1 Introduction

Let \(\mathbb {D}=\{z\in \mathbb {C}:\vert z\vert <1\}\) denote the open unit disk of the complex plane \(\mathbb {C}\) and \(H(\mathbb {D})\) denote the space of all analytic functions in \(\mathbb {D}\).

The Bloch space \(\mathcal {B}\) consists of those functions \(f\in H(\mathbb {D}) \) for which

$$\begin{aligned} \vert \vert f\vert \vert _{\mathcal {B}}=\vert f(0)\vert +\sup _{z\in \mathbb {D}}(1-\vert z\vert ^{2})\vert f'(z)\vert <\infty . \end{aligned}$$

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

$$\begin{aligned} \Vert f\Vert _{p}=\sup _{0\le r<1} M_p(r, f)<\infty , \end{aligned}$$

where

$$\begin{aligned} M_p(r, f)= & {} \left( \frac{1}{2\pi }\int _0^{2\pi }|f(re^{i\theta })|^p d\theta \right) ^{1/p}, \quad 0<p<\infty ,\\ M_{\infty }(r, f)= & {} \sup _{|z|=r}|f(z)|. \end{aligned}$$

For \(0<p<\infty \), the Dirichlet-type space \(D^{p}_{p-1}\) is the space of \(h\in H(\mathbb {D})\) such that

$$\begin{aligned} \Vert h\Vert _{D^{p}_{p-1}}^{p}=|h(0)|^{p}+ \int _{\mathbb {D}}|h'(z)|^{p}(1-|z|)^{p-1}dA(z)<\infty . \end{aligned}$$

When \(p=2\), the space \(D^{2}_{1}\) is just the Hardy space \(H^{2}\).

The Hardy–Littlewood space HL(p) consists of those function \(f\in H(\mathbb {D})\) for which

$$\begin{aligned} \Vert h\Vert ^{p}_{HL(p)}=\sum _{n=0}^{\infty }(n+1)^{p-2}|\widehat{h}(n)|^{p}<\infty . \end{aligned}$$

The space \(HL(\infty )\) consist of \(f(z)=\sum _{n=0}^{\infty }a_{n}z^{n}\in H(\mathbb {D})\) such that

$$\begin{aligned} \Vert f\Vert _{HL(\infty )}=\sup _{n\ge 0}(n+1)|a_{n}|<\infty . \end{aligned}$$

We also need the space \(H(\infty ,p)=\{h\in H(\mathbb {D}): \Vert h\Vert ^{p}_{H(\infty ,p)}=\int _{0}^{1}M_{\infty }^{p}(r,h)dr<\infty \}\).

It is well known that

$$\begin{aligned} D^{p}_{p-1}\subset H^{p}\subset HL(p),\quad 0<p\le 2, \end{aligned}$$
(1)
$$\begin{aligned} HL(p) \subset H^{p}\subset D^{p}_{p-1},\quad 2\le p<\infty , \end{aligned}$$
(2)

and

$$\begin{aligned} H^{p}\subset H(\infty ,p),\quad D^{p}_{p-1}\subset H(\infty ,p),\quad 0< p<\infty . \end{aligned}$$
(3)

The proofs of (1) and (2) can be found in [10, 11, 18], and the proof of (3) appears in [21, p. 127] and [14, Lemma 4].

Let \(1\le p<\infty \) and \(0<\alpha \le 1\), the mean Lipschitz space \(\Lambda ^p_\alpha \) consists of those functions \(f\in H(\mathbb {D})\) having a non-tangential limit almost everywhere such that \(\omega _p(t, f)=O(t^\alpha )\) as \(t\rightarrow 0\). Here \(\omega _p(\cdot , f)\) is the integral modulus of continuity of order p of the function \(f(e^{i\theta })\). It is known (see [10]) that \(\Lambda ^p_\alpha \) is a subset of \(H^p\) and

$$\begin{aligned} \Lambda ^p_\alpha =\left\{ f\in H(\mathbb {D}):M_p(r, f')=O\left( \frac{1}{(1-r)^{1-\alpha }}\right) , \quad \text{ as }\quad r\rightarrow 1\right\} . \end{aligned}$$

The space \(\Lambda ^p_\alpha \) is a Banach space with the norm \(\Vert \cdot \Vert _{\Lambda ^p_\alpha }\) given by

$$\begin{aligned} \Vert f\Vert _{\Lambda ^p_\alpha }=|f(0)|+\sup _{0\le r<1}(1-r)^{1-\alpha }M_p(r, f'). \end{aligned}$$

It is known (see e.g. [5, Theorem 2.5]) that

$$\begin{aligned} \Lambda ^{p}_{\frac{1}{p}}\subsetneq \mathcal {B}, \ \ 1<p<\infty . \end{aligned}$$

The Cesàro operator \(\mathcal {C}\) is defined in \(H(\mathbb {D})\) as follows:

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

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

The boundedness and compactness of the Cesàro operator \(\mathcal {C}\) and its generalizations on various spaces of analytic functions such as Hardy spaces, Bergman spaces, Dirichlet spaces, Bloch space, \(Q_{p}\) space, mixed norm space have been widely studied. See, e.g., [1, 2, 6, 9, 19, 23,24,25,26] and the references therein.

Recently, Galanopoulos et al. [12] introduced a Cesàro-like operator \(\mathcal {C}_\mu \) on \(H(\mathbb {D})\), which is a natural generalization of the classical Cesàro operator \(\mathcal {C}\). For a finite positive Borel measure \(\mu \) on the interval [0, 1), the Cesàro-like operator \(\mathcal {C}_\mu \) is defined on \(H(\mathbb {D})\) as follows:

$$\begin{aligned} \mathcal {C}_\mu (f)(z)=\sum ^\infty _{n=0}\left( \mu _n\sum ^n_{k=0}\widehat{f}(k)\right) z^n=\int _{0}^{1}\frac{f(tz)}{(1-tz)}d\mu (t), \quad z\in \mathbb {D}, \end{aligned}$$

where \(\mu _{n}\) stands for the moment of order n of \(\mu \), that is, \(\mu _{n}=\int _{0}^{1}t^{n}d\mu (t)\). They studied the operators \(\mathcal {C}_\mu \) acting on distinct spaces of analytic functions (e.g. Hardy space, Bergman space, Bloch space, etc.).

The Cesàro-like operator \(\mathcal {C}_\mu \) defined above has attracted the interest of many mathematicians. For instance, Jin and Tang [16] studied the boundedness (compactness) of \(\mathcal {C}_\mu \) from one Dirichlet-type space \(\mathcal {D}_{\alpha }\) into another one \(\mathcal {D}_{\beta }\). Bao, Sun and Wulan [3] studied the range of \(\mathcal {C}_\mu \) acting on \(H^{\infty }\). Blasco [4] investigated the operators \(\mathcal {C}_\mu \) induced by complex Borel measures on [0, 1). Galanopoulos et al. [13] studied the behaviour of the operators \(\mathcal {C}_\mu \) on the Dirichlet space and on the analytic Besov spaces. The operators \(\mathcal {C}_\mu \) associated to arbitrary complex Borel measures on \(\mathbb {D}\) the reader is referred to Galanopoulos et al. [15] and Zhou [27].

The following result was proved by Blasco [4, Theorem 3.17].

Theorem A

Let \(\mu \) be a finite positive Borel measure on [0, 1). Then \(\mathcal {C}_\mu : \mathcal {B}\rightarrow H^{2}\) is bounded if and only if

$$\begin{aligned} \sum _{n=0}^{\infty }\mu _{n}^{2}\log ^{2}(n+1)<\infty . \end{aligned}$$

It is natural to seek conditions on the measure \(\mu \) which makes the operator \(\mathcal {C}_\mu \) bounded (compact) from \(\mathcal {B}\) to \(H^{p}\) when \(p\ne 2\). The result for \(p\ne 2\) has not been given in the existing literature. In this paper, we extend the Bloch space \(\mathcal {B}\) to a class of spaces X and consider the target space \(H^{p}\) with any \(1\le p\le \infty \), where X is a Banach subspace of \(H(\mathbb {D})\) with \(\Lambda ^{s}_{\frac{1}{s}}\subset X\subset \mathcal {B}\) (\(1<s<\infty )\). We complete characterize the measure \(\mu \) for which \(\mathcal {C}_\mu \) is bounded (compact) from X to \(H^{p}\) \((1\le p\le \infty )\). It turns out that the boundedness and compactness from X to \(H^{p}\) are equivalent. This conclusion appears to be new even for \(p = 2\). At the end of the article, we also consider the case \(0<p<1\).

Our main results are included in the following.

Theorem 1.1

Suppose \(1\le p<\infty \), \(1<s<\infty \), and let \(\mu \) be a finite positive Borel measure on [0, 1). Let \(Y_{p}\in \{D^{p}_{p-1}, H^{p}, HL(p), H(\infty ,p)\}\) and let X be a Banach subspace of \(H(\mathbb {D})\) with \(\Lambda ^{s}_{\frac{1}{s}}\subset X\subset \mathcal {B}\). Then the following statements are equivalent.

  1. (1)

    The operator \(\mathcal {C}_\mu \) is bounded from X to \(Y_{p}\).

  2. (2)

    The operator \(\mathcal {C}_\mu \) is compact from X to \(Y_{p}\).

  3. (3)

    The measure \(\mu \) satisfies

$$\begin{aligned} \sum _{n=0}^{\infty }(n+1)^{p-2}\mu _{n}^{p}\log ^{p}(n+1)<\infty . \end{aligned}$$

Theorem 1.2

Let \(1<s<\infty \) and \(\mu \) be a finite positive Borel measure on [0, 1). Let X be a Banach subspace of \(H(\mathbb {D})\) with \(\Lambda ^{s}_{\frac{1}{s}}\subset X\subset \mathcal {B}\). Then the operator \(\mathcal {C}_\mu \) is bounded from X in \(H^{\infty }\) if and only if

$$\begin{aligned} \int _{0}^{1}\frac{\log \frac{e}{1-t}}{1-t}d\mu (t)<\infty . \end{aligned}$$

Throughout the paper, the measure \(\mu \) will be a positive finite Borel measure on the radius [0, 1), the letter C will denote an absolute constant whose value depends on the parameters indicated in the parenthesis, and may change from one occurrence to another. We will use the notation \(``P\lesssim Q"\) if there exists a constant \(C=C(\cdot ) \) such that \(`` P \le CQ"\), and \(`` P \gtrsim Q"\) is understood in an analogous manner. In particular, if \(``P\lesssim Q"\) and \( ``P \gtrsim Q"\), then we will write \(``P\asymp Q"\).

2 Proofs and some related results

We begin with some preliminary results which will be repeatedly used throughout the rest of the paper. The first lemma contains a characterization of \(L^{p}\)-integrability of power series with nonnegative coefficients. For a proof, see [20, Theorem 1].

Lemma 2.1

Let \(0<\beta ,p<\infty \), \(\{\lambda _{n}\}_{n=0}^{\infty }\) be a sequence of non-negative numbers. Then

$$\begin{aligned} \int _{0}^{1}(1-r)^{p\beta -1}\left( \sum _{n=0}^{\infty }\lambda _{n}r^{n}\right) ^{p}dr\asymp \sum _{n=0}^{\infty }2^{-np\beta }\left( \sum _{k\in I_{n}}\lambda _{k}\right) ^{p} \end{aligned}$$

where \(I_{0}=\{0\}\), \(I_{n}=[2^{n-1},2^{n})\cap \mathbb {N}\) for \(n\in \mathbb {N}\).

The following lemma is a consequence of Theorem 2.31 on page 192 of the classical monograph [28].

Lemma 2.2

  1. (a)

    The Taylor coefficients \(a_{n}\) of the function

    $$\begin{aligned} f(z)=\frac{1}{(1-z)^{\beta }}\log ^{\gamma }\frac{2}{1-z}, \quad \beta >0,\gamma \in \mathbb {R}, \quad z\in \mathbb {D} \end{aligned}$$

    have the property \(a_{n}\asymp n^{\beta -1}(\log (n+1))^{\gamma }\).

  2. (b)

    The Taylor coefficients \(a_{n}\) of the function

    $$\begin{aligned} f(z)=\log ^{\gamma }\frac{2}{1-z}, \quad \gamma >0,\quad z\in \mathbb {D} \end{aligned}$$

    have the property \(a_{n}\asymp n^{-1}(\log (n+1))^{\gamma -1}\).

We also need the following estimates (see, e.g. Proposition 1.4.10 in [22]).

Lemma 2.3

Let \(\alpha \) be any real number and \(z\in \mathbb {D}\). Then

$$\begin{aligned} \int ^{2\pi }_0\frac{d\theta }{|1-ze^{-i\theta }|^{\alpha }}\asymp {\left\{ \begin{array}{ll}1 &{} \hspace{5.0pt}\textrm{if} \quad \alpha <1,\\ \log \frac{2}{1-|z|^2} &{} \hspace{5.0pt}\textrm{if} \quad \alpha =1,\\ \frac{1}{(1-|z|^2)^{\alpha -1}} &{} \hspace{5.0pt}\textrm{if}\quad \alpha >1, \end{array}\right. } \end{aligned}$$

Proof of the implication \((1)\Rightarrow (3)\) Assume \(\mathcal {C}_\mu \) is bounded from X to \(Y_{p}\). Let \(f(z)=\log \frac{1}{1-z}\), then it is easy to check that \(f\in \Lambda ^{s}_{\frac{1}{s}}\subset X\). This implies that \(\mathcal {C}_\mu (f)\in Y_{p}\).

Case \(Y_{p}=H(\infty ,p)\). For \(0<r<1\), we have

$$\begin{aligned} M_{\infty }(r,\mathcal {C}_\mu (f))=\sup _{|z|=r}|\mathcal {C}_\mu (f)(z)|\ge \sum _{n=1}^{\infty } \mu _{n}\left( \sum _{k=1}^{n}\frac{1}{k}\right) r^{n}\gtrsim \sum _{n=1}^{\infty } \mu _{n}\log (n+1)r^{n}. \end{aligned}$$

Hence, by Lemma 2.1 we have

$$\begin{aligned} \Vert f\Vert ^{p}_{X}{} & {} \gtrsim \Vert \mathcal {C}_\mu (f)\Vert ^{p}_{Y_{p}}=\int _{0}^{1}M^{p}_{\infty }(r,\mathcal {C}_\mu (f))dr\\{} & {} \gtrsim \int _{0}^{1}\left( \sum _{n=1}^{\infty } \mu _{n}\log (n+1)r^{n}\right) ^{p}dr\\{} & {} \asymp \sum _{n=1}^{\infty } 2^{-n}\left( \sum _{k=2^{n-1}}^{2^{n}-1}\mu _{k}\log (k+1)\right) ^{p}\\{} & {} \gtrsim \sum _{n=1}^{\infty }2^{n(p-1)}\log ^{p}(2^{n}+1)\mu ^{p}_{2^{n}}\\{} & {} \gtrsim \sum _{k=1}^{\infty }(k+1)^{p-2}\mu ^{p}_{k}\log ^{p}(k+1). \end{aligned}$$

Case \(Y_{p}\ne H(\infty ,p)\). If \(p\ge 2\), then (2) and (3) show that \(Y_{p}\subset H(\infty ,p)\). This implies that

$$\begin{aligned} \Vert f\Vert ^{p}_{X}\gtrsim \Vert \mathcal {C}_\mu (f)\Vert ^{p}_{Y_{p}}\gtrsim \Vert \mathcal {C}_\mu (f)\Vert ^{p}_{H(\infty ,p)}. \end{aligned}$$

The desired result follows from the previous case.

If \(1\le p\le 2\), then (1) shows that \(Y_{p}\subset HL(p)\). Hence,

$$\begin{aligned} \Vert f\Vert ^{p}_{X}{} & {} \gtrsim \Vert \mathcal {C}_\mu (f)\Vert ^{p}_{Y_{p}}\gtrsim \Vert \mathcal {C}_\mu (f)\Vert ^{p}_{HL(p)}\\{} & {} = \sum _{n=1}^{\infty }(n+1)^{p-2}\mu _{n}^{p}\left| \sum _{k=1}^{n}\frac{1}{k}\right| ^{p}\\{} & {} \gtrsim \sum _{n=1}^{\infty }(n+1)^{p-2}\mu _{n}^{p}\log ^{p}(n+1). \end{aligned}$$

\(\square \)

Proof of the implication \((3)\Rightarrow (2)\) Let \(\{f_{k}\}_{k=1}^{\infty }\) be a bounded sequence in X which converges to 0 uniformly on every compact subset of \(\mathbb {D}\). Without loss of generality, we may assume that \(f_{k}(0)=0\) for all \(k\ge 1\) and \(\sup _{k\ge 1}\Vert f\Vert _{X}\le 1\).

Case \(1\le p\le 2\). Assume \(\sum _{n=1}^{\infty }(n+1)^{p-2}\mu _{n}^{p}\log ^{p}(n+1)<\infty \), then

$$\begin{aligned} \sum _{n=1}^{\infty }(n+1)^{p-2}\mu _{n}^{p}\log ^{p}(n+1){} & {} = \sum _{n=1}^{\infty }\left( \sum _{k=2^{n-1}}^{2^{n}-1}(k+1)^{p-2}\mu _{k}^{p}\log ^{p}(k+1)\right) \\{} & {} \gtrsim \sum _{n=1}^{\infty }2^{n(p-1)}\mu _{2^{n}}^{p}\log ^{p}(2^{n}+1)\\{} & {} \gtrsim \sum _{n=1}^{\infty }2^{-np}\left( \sum _{k=2^{n}}^{2^{n+1}-1}(k+1)^{1-\frac{1}{p}}\mu _{k}\log (k+1)\right) ^{p}. \end{aligned}$$

It follows that

$$\begin{aligned} \sum _{n=1}^{\infty }2^{-np}\left( \sum _{k=2^{n}}^{2^{n+1}-1}(k+1)^{1-\frac{1}{p}}\mu _{k}\log (k+1)\right) ^{p}<\infty . \end{aligned}$$

By Lemma 2.1 we have that

$$\begin{aligned}{} & {} \int _{0}^{1}(1-r)^{p-1}\left( \sum _{n=0}^{\infty }(n+1)^{1-\frac{1}{p}}\mu _{n}\log (n+1)r^{n}\right) ^{p}dr \\{} & {} \quad \asymp \sum _{n=0}^{\infty }2^{-np}\left( \sum _{k=2^{n}}^{2^{n+1}-1}(k+1)^{1-\frac{1}{p}}\mu _{k}\log (k+1)\right) ^{p}<\infty . \end{aligned}$$

Therefore, for any \(\varepsilon >0\) there exists a \(0<r_{0}<1\) such that

$$\begin{aligned} \int _{r_{0}}^{1}(1-r)^{p-1}\left( \sum _{n=0}^{\infty }(n+1)^{1-\frac{1}{p}}\mu _{k}\log (k+1)r^{n}\right) ^{p}dr<\varepsilon . \end{aligned}$$
(4)

By (1) and (3), we have that \(D_{p-1}^{p}\subset Y_{p}\). Hence, it is suffices to prove that

$$\begin{aligned} \lim _{k\rightarrow \infty }\Vert \mathcal {C}_\mu (f_{k})\Vert _{D_{p-1}^{p}}=0. \end{aligned}$$

It is clear that

$$\begin{aligned} \begin{aligned} \Vert \mathcal {C}_\mu (f_{k})\Vert ^{p}_{D_{p-1}^{p}}&=\int _{|z|\le r_{0}}|\mathcal {C}_\mu (f_{k})'(z)|^{p}(1-|z|)^{p-1}dA(z)\\&\quad +\, \int _{r_{0}<|z|<1}|\mathcal {C}_\mu (f_{k})'(z)|^{p}(1-|z|)^{p-1}dA(z)\\&:= J_{1,k}+J_{2,k}. \end{aligned} \end{aligned}$$

By the integral representation of \(\mathcal {C}_\mu \) we get

$$\begin{aligned} \mathcal {C}_\mu (f_{k})'(z)=\int _{0}^{1}\frac{tf'_{k}(tz)}{(1-tz)}d\mu (t) +\int _{0}^{1}\frac{tf_{k}(tz)}{(1-tz)^{2}}d\mu (t). \end{aligned}$$
(5)

Since \(\{f_{k}\}_{k=1}^{\infty }\) is converge to 0 uniformly on every compact subset of \(\mathbb {D}\), the Cauchy’s integral theorem implies that \(\{f'_{k}\}_{k=1}^{\infty }\) converges to 0 uniformly on every compact subset of \(\mathbb {D}\). Thus, for \(|z|\le r_{0}\), we have that

$$\begin{aligned} \begin{aligned} |\mathcal {C}_\mu (f_{k})'(z)|&\lesssim \int _{0}^{1}\frac{|f_{k}'(tz)|}{|1-tz|}+ \frac{|f_{k}(tz)|}{|1-tz|^{2}}d\mu (t)\\&\lesssim \sup _{|w|<r_{0}}\left( |f_{k}(w)|+|f_{k}'(w)|\right) \int _{0}^{1}\frac{1}{(1-tr_{0})^{2}}d\mu (t)\\&\lesssim \sup _{|w|<r_{0}}\left( |f_{k}(w)|+|f_{k}'(w)|\right) . \end{aligned} \end{aligned}$$

It follows that

$$\begin{aligned} J_{1,k} \rightarrow 0, \quad \text{ as }\quad k\rightarrow \infty . \end{aligned}$$

Next, we estimate \(J_{2,k}\).

Since \(X\subset \mathcal {B}\), for each \(k\ge 1\), we have

$$\begin{aligned} |f_{k}(z)|\lesssim \log \frac{e}{1-|z|}\quad \text{ and }\quad |f_{k}'(z)|\lesssim \frac{1}{1-|z|}\quad \text{ for } \text{ all }\quad z\in \mathbb {D}. \end{aligned}$$
(6)

By (5) and (6), Minkowski’s inequity, Lemma 2.3 we have that

$$\begin{aligned} \begin{aligned} M_{p}(r,\mathcal {C}_\mu (f_{k})')&= \left\{ \int _{0}^{2\pi }\left| \int _{0}^{1}\frac{tf_{k}'(tre^{i\theta })}{1-rte^{i\theta }}+\frac{tf_{k}(tre^{i\theta })}{(1-tre^{i\theta })^{2}}d\mu (t)\right| ^{p}d\theta \right\} ^{\frac{1}{p}}\\&\lesssim \left\{ \int _{0}^{2\pi }\left( \int _{0}^{1}\frac{1}{(1-tr)|1 -tre^{i\theta }|}d\mu (t)\right) ^{p}d\theta \right\} ^{\frac{1}{p}}\\&\quad +\, \left\{ \int _{0}^{2\pi }\left( \int _{0}^{1}\frac{\log \frac{e}{1-tr}}{|1-tre^{i\theta }|^{2}}d\mu (t)\right) ^{p}d\theta \right\} ^{\frac{1}{p}}\\&\lesssim \int _{0}^{1}\frac{1}{1-tr}\left( \int _{0}^{2\pi }\frac{d\theta }{|1-tre^{i\theta }|^{p}}\right) ^{\frac{1}{p}}d\mu (t)\\&\quad +\, \int _{0}^{1}\log \frac{e}{1-tr}\left( \int _{0}^{2\pi } \frac{d\theta }{|1-tre^{i\theta }|^{2p}}\right) ^{\frac{1}{p}}d\mu (t)\\&\lesssim \int _{0}^{1}F(t,r)d\mu (t), \end{aligned} \end{aligned}$$

where

$$\begin{aligned} F(t,r)=\left\{ \begin{array}{ll} \displaystyle {\frac{\log \frac{e}{1-tr}}{(1-tr)^{2-\frac{1}{p}}},} &{} \displaystyle {1<p\le 2};\\ \displaystyle {\frac{\log \frac{e}{1-tr}}{1-tr}, } &{} \displaystyle {p=1}. \end{array}\right. \end{aligned}$$

Lemma 2.2 yields that

$$\begin{aligned} M_{p}(r,\mathcal {C}_\mu (f_{k})')\lesssim \int _{0}^{1}F(t,r)d\mu (t) \asymp \sum _{n=0}^{\infty }(n+1)^{1-\frac{1}{p}}\mu _{n}\log (n+1)r^{n}. \end{aligned}$$

This together with (4) imply that

$$\begin{aligned} J_{2,k}&=\int _{r_{0}<|z|<1}|\mathcal {C}_\mu (f_{k})'(z)|^{p}(1-|z|)^{p-1}dA(z)\\&\lesssim \int _{r_{0}}^{1}(1-r)^{p-1}M^{p}_{p}(r,\mathcal {C}_\mu (f_{k})')dr\\&\lesssim \int _{r_{0}}^{1}(1-r)^{p-1}\left( \sum _{n=0}^{\infty }(n+1)^{1-\frac{1}{p}}\mu _{n}\log (n+1)r^{n}\right) ^{p}dr\\&\lesssim \varepsilon . \end{aligned}$$

Consequently,

$$\begin{aligned} \lim _{k\rightarrow \infty }\Vert \mathcal {C}_\mu (f_{k})\Vert _{D_{p-1}^{p}}=0. \end{aligned}$$

Case \(p>2\).

By (2) and (3) we see that \(HL(p)\subset Y_{p}\). To complete the proof, we have to prove that \(\lim _{k\rightarrow \infty }\Vert \mathcal {C}_\mu (f_{k})\Vert _{HL(p)}=0.\)

Since \(\sum _{n=1}^{\infty }(n+1)^{p-2}\mu _{n}^{p}\log ^{p}(n+1)<\infty \), for any \(\varepsilon >0\) there exists a positive integer N such that

$$\begin{aligned} \sum _{n=N+1}^{\infty }(n+1)^{p-2}\mu _{n}^{p}\log ^{p}(n+1)<\varepsilon . \end{aligned}$$
(7)

Set

$$\begin{aligned} f_{k}(z)=\sum _{j=1}^{\infty }\widehat{f_{k}}(j)z^{j},\quad z\in \mathbb {D}. \end{aligned}$$

Since \(\{f_{k}\}_{k=1}^{\infty }\) converges to 0 uniformly on every compact subsets of \(\mathbb {D}\), it follows that \(\widehat{f_{k}}(j)\longrightarrow 0\) as \(k\rightarrow \infty \) for every j. This yields that there exists \(K_{0}\in \mathbb {N}\) such that

$$\begin{aligned} \sum _{n=0}^{N}(n+1)^{p-2}\mu _{n}^{p} \left| \sum _{j=1}^{n}\widehat{f_{k}}(j)\right| ^{p}\lesssim \varepsilon \quad \text{ for } \text{ all } \ k>K_{0}. \end{aligned}$$
(8)

Note that \(\{f_{k}\}_{k=1}^{\infty }\subset X\subset \mathcal {B}\), then it follows from Corollary D in [17] we have that

$$\begin{aligned} \sup _{k\ge 1}\left| \sum _{j=1}^{n}\widehat{f_{k}}(j)\right| \lesssim \log (n+1) \sup _{k\ge 1} \Vert f_{k}\Vert _{\mathcal {B}}\lesssim \log (n+1). \end{aligned}$$
(9)

Hence, for \(k> K_{0}\), by (7)–(9) we have that

$$\begin{aligned} \begin{aligned} \Vert \mathcal {C}_\mu (f_{k})\Vert ^{p}_{HL(p)}&=\sum _{n=1}^{N}(n+1)^{p-2}\mu _{n}^{p}\left| \sum _{j=1}^{n}\widehat{f_{k}}(j)\right| ^{p}+ \sum _{n=N+1}^{\infty }(n+1)^{p-2}\mu _{n}^{p}\left| \sum _{j=1}^{n}\widehat{f_{k}}(j)\right| ^{p}\\&\lesssim \varepsilon +\sum _{n=N+1}^{\infty }(n+1)^{p-2}\mu _{n}^{p}\log ^{p}(n+1)\\&\lesssim \varepsilon +\varepsilon . \end{aligned} \end{aligned}$$

Therefore,

$$\begin{aligned} \lim _{k\rightarrow \infty } \Vert \mathcal {C}_\mu (f_{k})\Vert _{HL(p)}=0. \end{aligned}$$

The proof is complete. \(\square \)

Taking \(p = 1\) in Theorem 1.1, we can obtain the following corollary.

Corollary 2.4

Let \(1<s<\infty \), \(\mu \) be a finite positive Borel measure on [0, 1). Let X be a Banach subspace of \(H(\mathbb {D})\) with \(\Lambda ^{s}_{\frac{1}{s}}\subset X\subset \mathcal {B}\) and \(Y_{1}\in \{D, H^{1}, HL(1), H(\infty ,1)\}\). Then the following statements are equivalent.

  1. (1)

    The operator \(\mathcal {C}_\mu \) is bounded from X into \(Y_{1}\).

  2. (2)

    The operator \(\mathcal {C}_\mu \) is compact from X in to \(Y_{1}\).

  3. (3)

    The measure \(\mu \) satisfies

$$\begin{aligned} \int _{0}^{1}\log ^{2}\frac{e}{1-t}d\mu (t)<\infty . \end{aligned}$$

Carleson measures play a key role when we study the Cesàro-like operators. Recall that if \(\mu \) is a positive Borel measure on [0, 1), \(0\le \gamma <\infty \) and \(0< s < \infty \), then \(\mu \) is a \(\gamma \)-logarithmic s-Carleson measure if there exists a positive constant C such that

$$\begin{aligned} \mu ([t,1))\log ^{\gamma }\frac{e}{1-t} \le C (1-t)^{s}, \quad \text{ for } \text{ all }\quad 0\le t<1. \end{aligned}$$

In particular, \(\mu \) is an s-Carleson measure if \(\gamma =0\).

In the following, we provide a sufficient condition and a necessary condition in term of Carleson-type measure.

Corollary 2.5

Suppose \(1<p<\infty \), \(1<s<\infty \) and \(\gamma >1+\frac{1}{p}\), \(\mu \) is a positive Borel measure on [0, 1). Let \(Y_{p}\in \{D^{p}_{p-1}, H^{p}, HL(p), H(\infty ,p)\}\) and let X be a Banach subspace of \(H(\mathbb {D})\) with \(\Lambda ^{s}_{\frac{1}{s}}\subset X\subset \mathcal {B}\), then the following statements hold.

  1. (1)

    If \(\mu \) is a \(\gamma \)-logarithmic \(1-\frac{1}{p}\)-Carleson measure, then \(\mathcal {C}_\mu \) is bounded (equivalent to compact) from X to \( Y_{p}\).

  2. (2)

    If \(\mathcal {C}_\mu \) is bounded from X to \( Y_{p}\), then \(\mu \) is a 1-logarithmic \(1-\frac{1}{p}\)-Carleson measure.

Proof

  1. (1)

    Suppose \(\mu \) is a \(\gamma \)-logarithmic \(1-\frac{1}{p}\)-Carleson measure. Integrating by parts we have

    $$\begin{aligned} \begin{aligned} \mu _{n}&=\int _{0}^{1}t^{n}d\mu (t)=n \int _{0}^{1}t^{n-1}\mu ([t,1))dt\\&\lesssim n \int _{0}^{1}t^{n-1}(1-t)^{1-\frac{1}{p}}\log ^{-\gamma }\frac{e}{1-t}dt.\\ \end{aligned} \end{aligned}$$

    It is easy to estimate that

    $$\begin{aligned} n \int _{0}^{1}t^{n-1}(1-t)^{1-\frac{1}{p}}\log ^{-\gamma }\frac{e}{1-t}dt \asymp \frac{\log ^{-\gamma }(n+1)}{(n+1)^{1-\frac{1}{p}}}. \end{aligned}$$

    This shows that

    $$\begin{aligned} \mu _{n}= O\left( \frac{\log ^{-\gamma }(n+1)}{(n+1)^{1-\frac{1}{p}}}\right) . \end{aligned}$$

    Hence,

    $$\begin{aligned} \sum _{n=0}^{\infty }(n+1)^{p-2}\mu _{n}^{p}\log ^{p}(n+1)\lesssim \sum _{n=0}^{\infty }\frac{1}{(n+1)\log ^{p(\gamma -1)}(n+1)}\lesssim 1. \end{aligned}$$

    Now, Theorem 1.1 implies that \(\mathcal {C}_\mu \) is bounded (equivalent to compact) from X to \( Y_{p}\).

  2. (2)

    If \(\mathcal {C}_\mu \) is bounded from X to \( Y_{p}\), then for any \(N\ge 2\) we have

    $$\begin{aligned} \begin{aligned} 1&\gtrsim \sum _{n=0}^{\infty }(n+1)^{p-2}\mu _{n}^{p}\log ^{p}(n+1)\\&\gtrsim \sum _{n=0}^{N}(n+1)^{p-2}\mu _{n}^{p}\log ^{p}(n+1)\\&\gtrsim \mu ^{p}_{N} \sum _{n=0}^{N}(n+1)^{p-2}\log ^{p}(n+1)\\&\asymp \mu ^{p}_{N}N^{p-1} \log ^{p}(N+1). \end{aligned} \end{aligned}$$

    This yields that \(\mu \) is a 1-logarithmic \(1-\frac{1}{p}\)-Carleson measure.

\(\square \)

Proof of Theorem 1.2 Suppose \(\int _{0}^{1}\frac{\log \frac{e}{1-t}}{1-t}d\mu (t)<\infty \), then for any \(f\in X\) we have

$$\begin{aligned} \Vert \mathcal {C}_\mu (f)\Vert _{\infty }=\sup _{z\in \mathbb {D}}\left| \int _{0}^{1}\frac{f(tz)}{1-tz}d\mu (t)\right| \lesssim \sup _{z\in \mathbb {D}} \int _{0}^{1}\frac{\log \frac{e}{1-t|z|}}{1-t|z|}d\mu (t)\le \int _{0}^{1}\frac{\log \frac{e}{1-t}}{1-t}d\mu (t). \end{aligned}$$

This shows that \(\mathcal {C}_\mu \) is bounded from X to \(H^{\infty }\).

Conversely, assume that \(\mathcal {C}_\mu \) is bounded from X to \(H^{\infty }\). Take \(g(z)=\log \frac{e}{1-z}\in X\), then

$$\begin{aligned} \begin{aligned} \Vert g\Vert _{X}&\gtrsim \Vert \mathcal {C}_\mu (g)\Vert _{\infty }=\sup _{z\in \mathbb {D}}\left| \int _{0}^{1}\frac{\log \frac{e}{1-tz}}{1-tz}d\mu (t)\right| \\&\ge \sup _{x\in (0,1)}\int _{0}^{1}\frac{\log \frac{e}{1-tx}}{1-tx}d\mu (t). \end{aligned} \end{aligned}$$

By the Lebesgue control convergence theorem we have \(\int _{0}^{1}\frac{\log \frac{e}{1-t}}{1-t}d\mu (t)<\infty \). \(\square \)

Remark 2.6

In contrast with what happens in the case \(1\le p<\infty \), Theorem 1.2 is no longer valid for \(HL(\infty )\). In fact, \(\mathcal {C}_\mu \) is bounded from X to \(HL(\infty )\) if and only if \(\mu \) is a 1-logarithmic 1-Carleson measure. The proof is not difficult, we leave it to the interested reader. However, the condition \(\int _{0}^{1}\frac{\log \frac{e}{1-t}}{1-t}d\mu (t)<\infty \) implies that \(\mu \) is a 1-logarithmic 1-Carleson measure. But the reverse is not true, as the measure \(d\mu (t)=\log ^{-1}\frac{e}{1-t}dt\) shows. It is easy to check that \(\mu \) is a 1-logarithmic 1-Carleson measure but \(\int _{0}^{1}\frac{\log \frac{e}{1-t}}{1-t}d\mu (t)=\infty \).

Recall that the mixed norm space \(H(p,q,\alpha )\), \(0<p,q<\infty \), \(0<\alpha <\infty \), consists of those analytic functions on \(\mathbb {D}\) such that

$$\begin{aligned} \Vert f\Vert ^{q}_{p,q,\alpha }= \int _{0}^{1}(1-r)^{q\alpha -1}M^{q}_{p}(r,f)dr<\infty . \end{aligned}$$

Hardy and Littlewood proved (see [10, Theorem 5.11]) that

$$\begin{aligned} H^{p}\subset H\Bigg (q,p,\frac{1}{p}-\frac{1}{q}\Bigg ),\quad 0<p<q<\infty . \end{aligned}$$

In particular, for \(0<p<1\), \(H^{p}\subset H(1,p,\frac{1}{p}-1)\). In the following, we show that the range of \(\mathcal {C}_\mu \) acting on X is contained in \(H(1,p,\frac{1}{p}-1)\) for \(0<p<1\).

Theorem 2.7

Suppose \(1<s<\infty \), \(0<p<1\) and \(\mu \) is a finite positive Borel measure on [0, 1). Let X be a Banach subspace of \(H(\mathbb {D})\) with \(\Lambda ^{s}_{\frac{1}{s}}\subset X\subset \mathcal {B}\). Then \(\mathcal {C}_\mu (X)\subset H(1,p,\frac{1}{p}-1)\).

Proof

Let \(f\in X\subset \mathcal {B}\), then it is known that

$$\begin{aligned} |f(z)|\lesssim \Vert f\Vert _{\mathcal {B}}\log \frac{e}{1-|z|} \quad \text{ for } \text{ all }\ \ z\in \mathbb {D}. \end{aligned}$$

By Fubini theorem and Lemma 2.12.3 we have

$$\begin{aligned} \begin{aligned} \Vert \mathcal {C}_\mu (f)\Vert ^{p}_{1,p,\frac{1}{p}-1}&=\int _{0}^{1}(1-r)^{p(\frac{1}{p}-1)-1}M^{p}_{1}(r,\mathcal {C}_\mu (f))dr\\&=\int _{0}^{1}(1-r)^{-p}\left( \int _{0}^{2\pi }\left| \int _{0}^{1}\frac{f(tre^{i\theta })}{1-tre^{i\theta }}d\mu (t)\right| d\theta \right) ^{p}dr\\&\lesssim \Vert f\Vert ^{p}_{\mathcal {B}}\int _{0}^{1}(1-r)^{-p}\left( \int _{0}^{1}\log ^{2}\frac{e}{1-tr}d\mu (t)\right) ^{p}dr\\&\asymp \Vert f\Vert ^{p}_{\mathcal {B}}\int _{0}^{1}(1-r)^{-p} \left( \sum _{n=0}^{\infty }\frac{\log (n+1)}{n+1}\mu _{n}r^{n}\right) ^{p}dr\\&\asymp \Vert f\Vert ^{p}_{\mathcal {B}}\sum _{n=0}^{\infty }2^{-n(1-p)}\left( \sum _{k=2^{n}-1}^{2^{n+1}}(k+1)^{-1}\log (k+1)\mu _{k}\right) ^{p}\\&\lesssim \Vert f\Vert ^{p}_{\mathcal {B}}\sum _{n=0}^{\infty }2^{-n(1-p)} (n+1)^{p}\mu ^{p}_{2^{n}-1}\\&\lesssim \Vert f\Vert ^{p}_{\mathcal {B}}\sum _{n=0}^{\infty }2^{-n(1-p)} (n+1)^{p}\lesssim \Vert f\Vert ^{p}_{\mathcal {B}}\lesssim \Vert f\Vert ^{p}_{X}. \end{aligned} \end{aligned}$$

This yields that \(\mathcal {C}_\mu (X)\subset H\Bigg (1,p,\frac{1}{p}-1\Bigg )\). \(\square \)

We recall that \(f\in H(\mathbb {D})\) is a Cauchy transform if it admits a representation

$$\begin{aligned} f(z)=\int _{0}^{2\pi }\frac{d\lambda (\theta )}{1-e^{i\theta }z}, \ \ z\in \mathbb {D}, \end{aligned}$$

where \(\lambda \) is a finite complex Borel measure on \(\partial \mathbb {D}\). The space of all Cauchy transforms is denoted by \(\mathcal {K}\). 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 \(\Vert \cdot \Vert _{\infty }\)-norm. It turns out [8] that

$$\begin{aligned} \Vert f\Vert _{\mathcal {K}}=\sup \{|\langle f, g\rangle _{H^{2}}|: g\in \mathcal {A}, \Vert g\Vert _{H^{\infty }}\le 1 \}. \end{aligned}$$

It is known (see [7]) that \( H^{1}\subsetneq \mathcal {K}\subsetneq \bigcap _{0<p<1}H^{p}\). In view of Theorem 2.7, the following question arises naturally.

Question

Suppose \(1<s<\infty \), \(\mu \) is a finite positive Borel measure on [0, 1) and X is a Banach subspace of \(H(\mathbb {D})\) with \(\Lambda ^{s}_{\frac{1}{s}}\subset X\subset \mathcal {B}\). Is the operator \(\mathcal {C}_\mu \) bounded from X to \(\bigcap _{0<p<1}H^{p}\)?

We do not know the answer to this question. However, there exists a finite positive measure \(\mu \) on [0, 1) such that \(\mathcal {C}_\mu (X)\nsubseteq \mathcal {K}\). Actually, we have the following result.

Proposition 2.8

Suppose \(1<s<\infty \) and X is a Banach subspace of \(H(\mathbb {D})\) with \(\Lambda ^{s}_{\frac{1}{s}}\subset X\subset \mathcal {B}\). Then there exist \(f\in X\) and a finite positive measure \(\mu \) on [0, 1) such that \(\mathcal {C}_\mu (f)\notin \mathcal {K}\).

Proof

For \(1<\gamma <2\), let \(d\mu (t)=\left( (1-t)\left( \log \frac{e}{1-t}\right) ^{3-\gamma }\right) ^{-1}dt\). Then it is clear that \(\mu \) is a finite positive measure \(\mu \) on [0, 1). By Lemma 2.2 we have

$$\begin{aligned} \frac{1}{1-t}\left( \log \frac{e}{1-t}\right) ^{\gamma -3}\asymp \sum _{k=0}^{\infty }\left( \log (k+2)\right) ^{\gamma -3}t^{k}. \end{aligned}$$

For \(n\ge 2\), we have

$$\begin{aligned} \begin{aligned} \mu _{n}&=\int _{0}^{1}t^{n}d\mu (t)=\int _{0}^{1}t^{n}\left( (1-t) \left( \log \frac{e}{1-t}\right) ^{3-\gamma }\right) ^{-1}dt\\&\asymp \sum _{k=0}^{\infty }\left( \log (k+2)\right) ^{\gamma -3}\int _{0}^{1}t^{n+k}dt\\&=\sum _{k=0}^{n}\frac{\left( \log (k+2)\right) ^{\gamma -3}}{n+k+1}+\sum _{k=n+1}^{\infty } \frac{\left( \log (k+2)\right) ^{\gamma -3}}{n+k+1}\\&=S_{1}+S_{2}. \end{aligned} \end{aligned}$$

It is easy to see that

$$\begin{aligned} S_{1}=\sum _{k=0}^{n}\frac{\left( \log (k+2)\right) ^{\gamma -3}}{n+k+1}\asymp \frac{1}{n+1}\sum _{k=0}^{n} \left( \log (k+2)\right) ^{\gamma -3}\asymp \left( \log (n+2)\right) ^{\gamma -3}. \end{aligned}$$

Also, for every \(n\ge 2\) we have that

$$\begin{aligned} S_{2}= & {} \sum _{k=n+1}^{\infty }\frac{\left( \log (k+2)\right) ^{\gamma -3}}{n+k+1}\\= & {} \sum _{k=n+1}^{\infty }\frac{1}{(n+k+1)[\log (k+n+2)]^{3-\gamma }} \left( \frac{\log (k+n+2)}{\log (k+2)}\right) ^{3-\gamma }\\= & {} \sum _{k=n+1}^{\infty }\frac{1}{(n+k+1)[\log (k+n+2)]^{3-\gamma }} \left( 1+\frac{\log (1+\frac{n}{k+2})}{\log (k+2)}\right) ^{3-\gamma }\\&\asymp&\sum _{k=n+1}^{\infty }\frac{1}{(n+k+1)[\log (k+n+2)]^{3-\gamma }}\\&\asymp&\int _{n+1}^{\infty }\frac{1}{(n+x)[\log (n+x)]^{3-\gamma }}dx \\&\asymp&(\log (n+2))^{\gamma -2}. \end{aligned}$$

Hence, we have that

$$\begin{aligned} \mu _{n}\asymp \left( \log (n+2)\right) ^{\gamma -3}+ (\log (n+2))^{\gamma -2}\asymp (\log (n+2))^{\gamma -2}. \end{aligned}$$
(10)

Now, let \(f(z)=\log \frac{1}{1-z}\in X\) and \(g(z)=\sum _{n=0}^{\infty }\frac{z^{n}}{(n+1)(\log (n+2))^{\gamma }}\in \mathcal {A}\). To complete the proof, it is suffices to show that

$$\begin{aligned} |\langle \mathcal {C}_\mu (f), g\rangle _{H^{2}}|=\lim _{r\rightarrow 1^{-}}\sum _{n=1}^{\infty } \frac{\mu _{n}\left( \sum ^{n}_{k=1}\frac{1}{k}\right) }{(n+1)(\log (n+2))^{\gamma }}r^{2n}=\infty . \end{aligned}$$

In fact, it follows from (10) that

$$\begin{aligned} \sum _{n=1}^{\infty } \frac{\mu _{n}\left( \sum ^{n}_{k=1}\frac{1}{k}\right) }{(n+1)(\log (n+2))^{\gamma }}r^{2n}\asymp \sum _{n=1}^{\infty }\frac{r^{2n}}{(n+1)\log (n+2)}\rightarrow \infty , \ \ (r\rightarrow 1^{-}). \end{aligned}$$

The proof is complete. \(\square \)