1 Introduction

Let \({\mathbb {D}}\) be the open unit disk in the complex plane \({\mathbb {C}}\). Denote by \(H({\mathbb {D}})\) the space of functions analytic in \({\mathbb {D}}\). For \(f(z)=\sum _{n=0}^\infty a_nz^n\) in \(H({\mathbb {D}})\), the Cesàro operator \({\mathcal {C}}\) is defined by

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

See [7, 12, 14, 21, 23, 24] for the investigation of the Cesàro operator acting on some analytic function spaces.

Recently, P. Galanopoulos, D. Girela and N. Merchán [16] considered a Cesàro-like operator \({\mathcal {C}}_\mu \) on \(H({\mathbb {D}})\). For nonnegative integer n, let \(\mu _n\) be the moment of order n of a finite positive Borel measure \(\mu \) on [0, 1); that is,

$$\begin{aligned} \mu _n=\int _{[0, 1)} t^{n}d\mu (t). \end{aligned}$$

For \(f(z)=\sum _{n=0}^\infty a_nz^n\) belonging to \(H({\mathbb {D}})\), the Cesàro-like operator \({\mathcal {C}}_\mu \) is defined by

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

If \(d\mu (t)=dt\), then \({\mathcal {C}}_\mu ={\mathcal {C}}\). In [16, 19], the authors studied the action of \({\mathcal {C}}_\mu \) on distinct spaces of analytic functions.

We also need to recall some function spaces. For \(0<p<\infty \), \(H^p\) denotes the classical Hardy space [13] of those functions \(f\in H({\mathbb {D}})\) for which

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

where

$$\begin{aligned} M_p(r, f)= \left( \frac{1}{2\pi }\int _0^{2\pi }|f(re^{i\theta })|^p d\theta \right) ^{1/p}. \end{aligned}$$

As usual, denote by \(H^\infty \) the space of bounded analytic functions in \({\mathbb {D}}\). It is well known that \(H^\infty \) is a proper subset of the Bloch space \({\mathcal {B}}\) which consists of those functions \(f\in H({\mathbb {D}})\) satisfying

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

Denote by \(\text {Aut}({\mathbb {D}})\) the group of Möbius maps of \({\mathbb {D}}\), namely,

$$\begin{aligned} \text {Aut}({\mathbb {D}})=\{e^{i\theta }\sigma _a:\ \ a\in {\mathbb {D}}\ \text {and} \ \theta \ \ \text {is real}\}, \end{aligned}$$

where

$$\begin{aligned} \sigma _a(z)=\frac{a-z}{1-{\overline{a}}z}, \qquad z\in {\mathbb {D}}. \end{aligned}$$

In 1995 R. Aulaskari, J. Xiao and R. Zhao [2] introduced \({\mathcal {Q}}_p\) spaces. For \(0\le p<\infty \), a function f analytic in \({\mathbb {D}}\) belongs to \({\mathcal {Q}}_p\) if

$$\begin{aligned} \Vert f\Vert _{{\mathcal {Q}}_p}^2=\sup _{w\in {\mathbb {D}}} \int _{\mathbb {D}}|f'(z)|^2(1-|\sigma _w(z)|^2)^p dA(z)<\infty , \end{aligned}$$

where dA is the area measure on \({\mathbb {C}}\) normalized so that \(A({\mathbb {D}})=1\). \({\mathcal {Q}}_p\) spaces are Möbius invariant in the sense that

$$\begin{aligned} \Vert f\Vert _{{\mathcal {Q}}_p}=\Vert f\circ \phi \Vert _{{\mathcal {Q}}_p} \end{aligned}$$

for every \(f\in {\mathcal {Q}}_p\) and \(\phi \in \text {Aut}({\mathbb {D}})\). It was shown in [25] that \({\mathcal {Q}}_2\) coincides with the Bloch space \({\mathcal {B}}\). This result was extended in [1] by showing that \({\mathcal {Q}}_p={\mathcal {B}}\) for all \(1<p<\infty \). The space \({\mathcal {Q}}_1\) coincides with BMOA, the set of analytic functions in \({\mathbb {D}}\) with boundary values of bounded mean oscillation (see [5, 17]). The space \({\mathcal {Q}}_0\) is the Dirichlet space \({\mathcal {D}}\). For \(0<p<1\), the space \({\mathcal {Q}}_p\) is a proper subset of BMOA and has many interesting properties. See J. Xiao’s monographs [26, 27] for the theory of \({\mathcal {Q}}_p\) spaces.

For \(1\le p<\infty \) and \(0<\alpha \le 1\), the mean Lipschitz space \(\Lambda ^p_\alpha \) is the set of those functions \(f\in H({\mathbb {D}})\) with 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 well known (cf. [13, Chapter 5]) that \(\Lambda ^p_\alpha \) is a subset of \(H^p\) and \(\Lambda ^p_\alpha \) consists of those functions \(f\in H({\mathbb {D}})\) satisfying

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

Among these spaces, the spaces \(\Lambda ^p_{1/p}\) are of special interest. \(\Lambda ^p_{1/p}\) spaces increase with \(p\in (1, \infty )\) in the sense of inclusion and they are contained in BMOA (cf. [10]). By Theorem 1.4 in [4], \(\Lambda ^p_{1/p}\subseteq {\mathcal {Q}}_q\) when \(1\le p<2/(1-q)\) and \(0<q<1\). In particular, \(\Lambda ^2_{1/2}\subseteq {\mathcal {Q}}_q \subseteq {\mathcal {B}}\) for all \(0<q<\infty \).

Given an arc I of the unit circle \({\mathbb {T}}\) with arclength |I| (normalized such that \(|{\mathbb {T}}|=1\)), the Carleson box S(I) is given by

$$\begin{aligned} S(I)=\{r\zeta \in {\mathbb {D}}: 1-|I|<r<1, \ \zeta \in I\}. \end{aligned}$$

For \(0<s<\infty \), a positive Borel measure \(\nu \) on \({\mathbb {D}}\) is said to be an s-Carleson measure if

$$\begin{aligned} \sup _{I\subseteq {\mathbb {T}}}\frac{\nu (S(I))}{|I|^s}<\infty . \end{aligned}$$

If \(\nu \) is a 1-Carleson measure, we write that \(\nu \) is a Carleson measure characterizing \(H^p\subseteq L^p(d\nu )\) for \(0<p<\infty \) (cf. [13]). 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

$$\begin{aligned} {\tilde{\mu }}(E)=\mu (E \cap [0, 1)), \end{aligned}$$

for any Borel subset E of \({\mathbb {D}}\). Thus \(\mu \) is an s-Carleson measure on [0, 1) if there is a positive constant C such that

$$\begin{aligned} \mu ([t, 1)) \le C (1-t)^s \end{aligned}$$

for all \(t\in [0, 1)\). We refer to [8] for the investigation of this kind of measures associated with Hankel measures.

It is known that the Cesàro operator \({\mathcal {C}}\) is bounded on \(H^p\) for all \(0<p<\infty \) (cf. [21, 23, 24]) but this is not true on \(H^\infty \). In fact, N. Danikas and A. Siskakis [12] gave that \({\mathcal {C}}(H^\infty )\nsubseteq H^\infty \) but \({\mathcal {C}}(H^\infty )\subseteq BMOA\). Later M. Essén and J. Xiao [14] proved that \({\mathcal {C}}(H^\infty )\subsetneqq {\mathcal {Q}}_p\) for \(0<p<1\). Recently, the relation between \({\mathcal {C}}(H^\infty )\) and a class of Möbius invariant function spaces was considered in [7].

It is quite natural to study \({\mathcal {C}}_\mu (H^\infty )\). In [16] the authors characterized positive Borel measures \(\mu \) such that \({\mathcal {C}}_\mu (H^\infty )\subseteq H^\infty \) and proved that \({\mathcal {C}}_\mu (H^\infty )\subseteq {\mathcal {B}}\) if and only if \(\mu \) is a Carleson measure. Moreover, they showed that if \({\mathcal {C}}_\mu (H^\infty )\subseteq BMOA\), then \(\mu \) is a Carleson measure. In [16, p. 20], the authors asked whether or not \(\mu \) being a Carleson measure implies that \({\mathcal {C}}_\mu (H^\infty )\subseteq BMOA\). In this paper, by giving some descriptions of s-Carleson measures on [0, 1), for \(0<p<2\), we show that \({\mathcal {C}}_\mu (H^\infty )\subseteq {\mathcal {Q}}_p\) if and only if \(\mu \) is a Carleson measure, which gives an affirmative answer to their question. We also consider another Cesàro-like operator \({\mathcal {C}}_{\mu , s}\) and describe the embedding \({\mathcal {C}}_{\mu , s}(H^\infty )\subseteq X\) in terms of s-Carleson measures, where X is between \(\Lambda ^p_{1/p}\) and \({\mathcal {B}}\) for \(\max \{1, 1/s\}<p<\infty \).

Throughout this paper, the symbol \(A\thickapprox B\) means that \(A\lesssim B\lesssim A\). We say that \(A\lesssim B\) if there exists a positive constant C such that \(A\le CB\).

2 Positive Borel measures on [0, 1) as Carleson type measures

In this section, we give some characterizations of positive Borel measures on [0, 1) as Carleson type measures.

The following description of Carleson type measures (cf. [9] ) is well known.

Lemma A

Suppose \(s>0\), \(t>0\) and \(\mu \) is a positive Borel measure on \({\mathbb {D}}\). Then \(\mu \) is an s-Carleson measure if and only if

$$\begin{aligned} \sup _{a\in {\mathbb {D}}}\int _{{\mathbb {D}}} \frac{(1-|a|^2)^t}{|1-{\overline{a}}w|^{s+t}}d\mu (w)<\infty . \end{aligned}$$
(2.1)

For Carleson type measures on [0, 1), we can obtain some descriptions that are different from Lemma A. Now we give the first main result in this section.

Proposition 2.1

Suppose \(0<t<\infty \), \(0\le r<s<\infty \) and \(\mu \) is a finite positive Borel measure on [0, 1). Then the following conditions are equivalent:

  1. (i)

    \(\mu \) is an s-Carleson measure;

  2. (ii)
    $$\begin{aligned} \sup _{a\in {\mathbb {D}}}\int _{[0,1)}\frac{(1-|a|)^t}{(1-x)^{r}(1-|a|x)^{s+t-r}}d\mu (x)<\infty ; \end{aligned}$$
    (2.2)
  3. (iii)
    $$\begin{aligned} \sup _{a\in {\mathbb {D}}}\int _{[0,1)}\frac{(1-|a|)^t}{(1-x)^{r}|1-ax|^{s+t-r}}d\mu (x)<\infty . \end{aligned}$$
    (2.3)

Proof

\((i)\Rightarrow (ii)\). Let \(\mu \) be an s-Carleson measure. Fix \(a\in {\mathbb {D}}\) with \(|a|\le 1/2\). If \(r=0\), the desired result holds. For \(0<r<s\), using a well-known formula about the distribution function(cf. [15, p.20 ]), we get

$$\begin{aligned} \int _{[0,1)}&\frac{(1-|a|)^t}{(1-x)^{r}(1-|a|x)^{s+t-r}}d\mu (x)\nonumber \\ \thickapprox&\quad \int _{[0,1)}\left( \frac{1}{1-x}\right) ^rd\mu (x) \nonumber \\ \thickapprox&\quad r \int _0^\infty \lambda ^{r-1} \mu (\{x\in [0, 1): 1-\frac{1}{\lambda }<x \})d\lambda \nonumber \\ \lesssim&\quad \int _0^1 \lambda ^{r-1} \mu ([0, 1))d\lambda + \int _1^\infty \lambda ^{r-1} \mu ([1-\frac{1}{\lambda }, 1))d\lambda \nonumber \\ \lesssim&\quad 1+ \int _1^\infty \lambda ^{r-s-1} d\lambda \lesssim 1. \end{aligned}$$
(2.4)

Fix \(a\in {\mathbb {D}}\) with \(|a|>1/2\) and let

$$\begin{aligned} S_n(a)=\{x\in [0,1): 1-2^n(1-|a|)\le x<1\}, \ \ n=1, 2, \cdots . \end{aligned}$$

Let \(n_a\) be the minimal integer such that \(1-2^{n_a}(1-|a|)\le 0\). Then \(S_n(a)=[0, 1)\) when \(n\ge n_a\). If \( x\in S_1(a)\), then

$$\begin{aligned} 1-|a| \le 1-|a|x. \end{aligned}$$
(2.5)

Also, for \(2\le n\le n_a\) and \(x\in S_n(a)\backslash S_{n-1}(a)\), we have

$$\begin{aligned} 1-|a|x \ge |a|-x \ge |a|-(1-2^{n-1}(1-|a|))=(2^{n-1}-1)(1-|a|). \end{aligned}$$
(2.6)

We write

$$\begin{aligned} \int _{[0,1)}&\frac{(1-|a|)^t}{(1-x)^r(1-|a|x)^{s+t-r}}d\mu (x)\\ =&\int _{S_1(a)}\frac{(1-|a|)^t}{(1-x)^r(1-|a|x)^{s+t-r}}d\mu (x)\\&+\sum ^{n_a}_{n=2}\int _{S_n(a)\backslash S_{n-1}(a)}\frac{(1-|a|)^t}{(1-x)^r(1-|a|x)^{s+t-r}}d\mu (x)\\ =:&J_1(a)+J_2(a). \end{aligned}$$

If \(r=0\), bearing in mind (2.5), (2.6) and that \(\mu \) is an s-Carleson measure, it is easy to check that \(J_i(a)\lesssim 1\) for \(i=1, 2\). Now consider \(0<t<\infty \) and \(0< r<s<\infty \). Using (2.5) and some estimates similar to (2.4), we have

$$\begin{aligned} J_1(a) \lesssim (1-|a|)^{r-s}\int _{S_1(a)}\left( \frac{1}{1-x}\right) ^rd\mu (x) \lesssim 1. \end{aligned}$$

Note that (2.6) holds, \(0<t<\infty \), \(0< r<s<\infty \) and \(\mu \) is an s-Carleson measure. Then

$$\begin{aligned} J_2(a) \lesssim&\sum ^{n_a}_{n=2}\frac{(1-|a|)^{r-s}}{2^{n(s+t-r)}}\int _{S_n(a)\backslash S_{n-1}(a)}\left( \frac{1}{1-x}\right) ^rd\mu (x)\\ \lesssim&\sum ^{n_a}_{n=2}\frac{(1-|a|)^{r-s}}{2^{n(s+t-r)}}\int _0^\infty \lambda ^{r-1}\mu \Big (\Big \{x\in [1-2^n(1-|a|),1): 1-\frac{1}{\lambda }<x\Big \}\Big )d\lambda \\ \thickapprox&\sum ^{n_a}_{n=2}\frac{(1-|a|)^{r-s}}{2^{n(s+t-r)}} \bigg (\int _0^{\frac{1}{2^n(1-|a|)}}\lambda ^{r-1}\mu \big ([1-2^n(1-|a|),1)\big )d\lambda \\&+\int _{\frac{1}{2^n(1-|a|)}}^\infty \lambda ^{r-1} \mu \Big (\Big [1-\frac{1}{\lambda },1\Big )\Big )d\lambda \Bigg )\\ \lesssim&\sum ^{n_a}_{n=2}\frac{(1-|a|)^{r-s}}{2^{n(s+t-r)}} \Bigg (2^{ns}(1-|a|)^s\int _0^{\frac{1}{2^n(1-|a|)}}\lambda ^{r-1}d\lambda +\int _{\frac{1}{2^n(1-|a|)}}^\infty \lambda ^{r-1-s}d\lambda \Bigg )\\ \thickapprox&\sum ^{n_a}_{n=2} \frac{1}{2^{tn}}<\infty . \end{aligned}$$

Consequently,

$$\begin{aligned} \sup _{a\in {\mathbb {D}}}\int _{[0,1)}\frac{(1-|a|)^t}{(1-x)^r(1-|a|x)^{s+t-r}}d\mu (x)<\infty . \end{aligned}$$

The implication of \((ii)\Rightarrow (iii)\) is clear.

\((iii)\Rightarrow (i)\). For \(r\ge 0\), it is clear that

$$\begin{aligned} \int _{[0,1)}\frac{(1-|a|)^t}{(1-x)^{r}|1-ax|^{s+t-r}}d\mu (x)\ge \int _{[0,1)}\frac{(1-|a|)^t}{|1-ax|^{s+t}}d\mu (x) \end{aligned}$$

for all \(a\in {\mathbb {D}}\). Combining this with Lemma A, we see that if (2.3) holds, then \(\mu \) is an s-Carleson measure. \(\square \)

Remark 1

The condition \(0\le r<s<\infty \) in Proposition 2.1 can not be changed to \(r\ge s>0\). For example, let \(d\mu _1(x)=(1-x)^{s-1}dx\), \(x\in [0, 1)\). Then \(\mu _1\) is an s-Carleson measure but for \(r\ge s>0\),

$$\begin{aligned} \sup _{a\in {\mathbb {D}}}\int _{[0,1)}&\frac{(1-|a|)^t}{(1-x)^{r}|1-ax|^{s+t-r}}d\mu _1(x)\\ \ge&\int _0^1 (1-x)^{s-1-r}dx=+\infty . \end{aligned}$$

Remark 2

\(\mu \) supported on [0, 1) is essential in Proposition 2.1. For example, consider \(0<t<1\), \(0< r<s<1\) and \(s=r+t\). Set \(d\mu _2(w)=|f'(w)|^2(1-|w|^2)^sdA(w)\), \(w\in {\mathbb {D}}\), where \(f\in {\mathcal {Q}}_s\setminus {\mathcal {Q}}_t\). Note that for \(0<p<\infty \) and \(g\in H({\mathbb {D}})\), \(|g'(w)|^2(1-|w|^2)^pdA(w)\) is a p-Carleson measure if and only if \(g\in {\mathcal {Q}}_p\) (cf. [26]). Hence \(d\mu _2\) is an s-Carleson measure. But

$$\begin{aligned} \sup _{a\in {\mathbb {D}}}\int _{{\mathbb {D}}}&\frac{(1-|a|)^t}{(1-|w|)^{r}|1-a{\overline{w}}|^{s+t-r}}d\mu _2(w)\\&=\sup _{a\in {\mathbb {D}}}\int _{{\mathbb {D}}}|f'(w)|^2\frac{(1-|a|)^t(1-|w|)^{s-r}}{|1-a{\overline{w}}|^{s+t-r}}dA(w)\\&\thickapprox \sup _{a\in {\mathbb {D}}}\int _{{\mathbb {D}}}|f'(w)|^2(1-|\sigma _a(w)|^2)^tdA(w)=+\infty . \end{aligned}$$

Before giving the other characterization of Carleson type measures on [0, 1), we need to recall some results.

The following result is Lemma 1 in [20], which generalizes Lemma 3.1 in [18] from \(p=2\) to \(1<p<\infty \).

Lemma B

Let \(f\in H({\mathbb {D}})\) with \(f(z)=\sum ^\infty _{n=0}a_n z^n\). Suppose \(1<p<\infty \) and the sequence \(\{a_n\}\) is a decreasing sequence of nonnegative numbers. If X is a subspace of \(H({\mathbb {D}})\) with \(\Lambda ^p_{1/p}\subseteq X\subseteq {\mathcal {B}}\), then

$$\begin{aligned} f\in X \iff a_n=O\left( \frac{1}{n}\right) . \end{aligned}$$

We recall a characterization of s-Carleson measure \(\mu \) on [0, 1) as follows (cf. [6, Theorem 2.1] or [11, Proposition1]).

Proposition C

Let \(\mu \) be a finite positive Borel measure on [0, 1) and \(s>0\). Then \(\mu \) is an s-Carleson measure if and only if the sequence of moments \(\{\mu _n\}_{n=0}^\infty \) satisfies \(\sup _{n\ge 0} (1+n)^s \mu _n<\infty \).

The following characterization of functions with nonnegative Taylor coefficients in \({\mathcal {Q}}_p\) is Theorem 2.3 in [3].

Theorem D

Let \(0<p<\infty \) and let \(f(z)=\sum _{n=0}^\infty a_nz^n\) be an analytic function in \({\mathbb {D}}\) with \(a_n\ge 0\) for all n. Then \(f\in {\mathcal {Q}}_p\) if and only if

$$\begin{aligned} \sup _{0\le r<1} \sum _{n=0}^\infty \frac{(1-r)^p}{(n+1)^{p+1}}\left( \sum _{k=0}^n(k+1)a_{k+1}(n-k+1)^{p-1}r^{n-k}\right) ^2<\infty . \end{aligned}$$

We need the following well-known estimates (cf. [28, Lemma 3.10]).

Lemma E

Let \(\beta \) be any real number. Then

$$\begin{aligned} \int ^{2\pi }_0\frac{d\theta }{|1-ze^{-i\theta }|^{1+\beta }}\thickapprox {\left\{ \begin{array}{ll}1 &{} \text {if} \ \ \beta <0,\\ \log \frac{2}{1-|z|^2} &{} \text {if} \ \ \beta =0,\\ \frac{1}{(1-|z|^2)^\beta } &{} \text {if}\ \ \beta >0, \end{array}\right. } \end{aligned}$$

for all \(z\in {\mathbb {D}}\).

For \(0<s<\infty \) and a finite positive Borel measure \(\mu \) on [0, 1), set

$$\begin{aligned} f_{\mu , s}(z)=\sum _{n=0}^\infty \frac{\Gamma (n+s)}{\Gamma (s)n!} \mu _n z^n, \ \ z\in {\mathbb {D}}. \end{aligned}$$

Now we state the other main result in this section which is inspired by Lemma B and Proposition C.

Proposition 2.2

Suppose \(0<s<\infty \) and \(\mu \) is a finite positive Borel measure on [0, 1). Let \(1<p<\infty \) and let X be a subspace of \(H({\mathbb {D}})\) with \(\Lambda ^p_{1/p}\subseteq X\subseteq {\mathcal {B}}\). Then \(\mu \) is an s-Carleson measure if and only if \(f_{\mu , s}\in X\).

Proof

Let \(\mu \) be an s-Carleson measure. Clearly,

$$\begin{aligned} f_{\mu , s}(z)=\int _{[0, 1)} \frac{1}{(1-tz)^s}d\mu (t) \end{aligned}$$

for any \(z\in {\mathbb {D}}\). For \(p>1\), it follows from the Minkowski inequality and Lemma E that

$$\begin{aligned} M_p(r, f'_{\mu , s})\le&s \left( \frac{1}{2\pi }\int _0^{2\pi } \left( \int _{[0, 1)}\frac{1}{|1-tre^{i\theta }|^{s+1}}d\mu (t)\right) ^pd\theta \right) ^{1/p}\\ \le&s \int _{[0, 1)} \left( \frac{1}{2\pi }\int _0^{2\pi } \frac{1}{|1-tre^{i\theta }|^{(s+1)^p}}d\theta \right) ^{1/p} d\mu (t)\\ \lesssim&\int _{[0, 1)} \frac{1}{(1-tr)^{s+1-\frac{1}{p}}} d\mu (t) \end{aligned}$$

for all \(0<r<1\). Combining this with Proposition 2.1, we get \(f_{\mu , s}\in \Lambda ^p_{1/p}\) and hence \(f_{\mu , s}\in X\).

On the other hand, let \(f_{\mu , s}\in X\). Then \(f_{\mu , s}\in {\mathcal {Q}}_q\) with \(q>1\). By the Stirling formula,

$$\begin{aligned} \frac{\Gamma (n+s)}{\Gamma (s)n!}\thickapprox (n+1)^{s-1} \end{aligned}$$

for all nonnegative integers n. Consequently, by Theorem D we deduce

$$\begin{aligned} \infty> & {} \sum _{n=0}^\infty \frac{(1-r)^q}{(n+1)^{q+1}}\left( \sum _{k=0}^n(k+2)^{s}\mu _{k+1}(n-k+1)^{q-1}r^{n-k}\right) ^2\\ > rsim & {} \sum _{n=0}^\infty \frac{(1-r)^q}{(4n+1)^{q+1}}\left( \sum _{k=0}^{4n}(k+2)^{s}\mu _{k+1}(4n-k+1)^{q-1}r^{4n-k}\right) ^2\\ > rsim & {} \sum _{n=0}^\infty \frac{(1-r)^q}{(4n+1)^{q+1}}\left( \sum _{k=n}^{2n}(k+2)^{s}\int _r^1 t^{k+1}d\mu (t) (4n-k+1)^{q-1}r^{4n-k}\right) ^2\\ > rsim & {} \mu ^2([r, 1)) (1-r)^q\sum _{n=0}^\infty \frac{r^{8n+2}}{(4n+1)^{q+1}}\left( \sum _{k=n}^{2n}(k+2)^{s}(4n-k+1)^{q-1} \right) ^2\\ > rsim & {} \mu ^2([r, 1)) (1-r)^q \sum _{n=0}^\infty (4n+2)^{2s+q-1} r^{8n+2}\\&\thickapprox&\frac{\mu ^2([r, 1))}{(1-r)^{2s}} \end{aligned}$$

for all \(r\in [0, 1)\) which yields that \(\mu \) is an s-Carleson measure. The proof is complete. \(\square \)

3 \({\mathcal {Q}}_p\) spaces and the range of \({\mathcal {C}}_\mu \) acting on \(H^\infty \)

In this section, we characterize finite positive Borel measures \(\mu \) on [0, 1) such that \({\mathcal {C}}_\mu (H^\infty )\subseteq {\mathcal {Q}}_p\) for \(0<p<2\). Descriptions of Carleson measures in Proposition 2.1 play a key role in our proof.

The following lemma is from [22].

Lemma F

Suppose \(s>-1\), \(r>0\), \(t>0\) with \(r+t-s-2>0\). If r, \(t<2+s\), then

$$\begin{aligned} \int _{\mathbb {D}}\frac{(1-|z|^2)^s}{|1-{\overline{a}}z|^r|1-{\overline{b}}z|^t}dA(z)\lesssim \frac{1}{|1-{\overline{a}}b|^{r+t-s-2}} \end{aligned}$$

for all a, \(b\in {\mathbb {D}}\). If \(t<2+s<r\), then

$$\begin{aligned} \int _{\mathbb {D}}\frac{(1-|z|^2)^s}{|1-{\overline{a}}z|^r|1-{\overline{b}}z|^t}dA(z)\lesssim \frac{(1-|a|^2)^{2+s-r}}{|1-{\overline{a}}b|^{t}} \end{aligned}$$

for all a, \(b\in {\mathbb {D}}\).

We give our result as follows.

Theorem 3.1

Suppose \(0<p<2\) and \(\mu \) is a finite positive Borel measure on [0, 1). Then \({\mathcal {C}}_\mu (H^\infty )\subseteq {\mathcal {Q}}_p\) if and only if \(\mu \) is a Carleson measure.

Proof

Suppose \({\mathcal {C}}_\mu (H^\infty )\subseteq {\mathcal {Q}}_p\). Then \({\mathcal {C}}_\mu (H^\infty )\) is a subset of the Bloch space. By [16, Theorem 5], \(\mu \) is a Carleson measure.

Conversely, suppose \(\mu \) is a Carleson measure and \(f\in H^\infty \). Then f is also in the Bloch space \({\mathcal {B}}\). From Proposition 1 in [16],

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

Hence for any \(z\in {\mathbb {D}}\),

$$\begin{aligned} \Vert {\mathcal {C}}_\mu (f)\Vert _{{\mathcal {Q}}_p}&\lesssim \sup _{a\in {\mathbb {D}}} \left( \int _{{\mathbb {D}}}\left( \int _{[0,1)}\frac{|tf'(tz)|}{|1-tz|}d\mu (t)\right) ^2(1-|\sigma _a(z)|^2)^p dA(z)\right) ^{\frac{1}{2}} \nonumber \\&\quad +\sup _{a\in {\mathbb {D}}}\left( \int _{{\mathbb {D}}}\left( \int _{[0,1)}\frac{|tf(tz)|}{|1-tz|^2}d\mu (t)\right) ^2(1-|\sigma _a(z)|^2)^p dA(z)\right) ^{\frac{1}{2}} \nonumber \\&\lesssim \Vert f\Vert _{\mathcal {B}}\sup _{a\in {\mathbb {D}}} \left( \int _{{\mathbb {D}}}\left( \int _{[0,1)}\frac{1}{(1-|tz|)|1-tz|}d\mu (t)\right) ^2(1-|\sigma _a(z)|^2)^p dA(z)\right) ^{\frac{1}{2}} \nonumber \\&\quad +\Vert f\Vert _{H^\infty }\sup _{a\in {\mathbb {D}}}\left( \int _{{\mathbb {D}}}\left( \int _{[0,1)}\frac{1}{|1-tz|^2}d\mu (t)\right) ^2(1-|\sigma _a(z)|^2)^p dA(z)\right) ^{\frac{1}{2}}.\nonumber \\ \end{aligned}$$
(3.1)

Let c be a positive constant such that \(2c<\min \{2-p, p\}\). Then

$$\begin{aligned} (1-|tz|)^2\ge (1-t)^{2-2c} (1-|z|)^{2c} \end{aligned}$$
(3.2)

for all \(t\in [0, 1)\) and all \(z\in {\mathbb {D}}\). By the Minkowski inequality, (3.2), Lemma F and Proposition 2.1, we get

$$\begin{aligned}&\sup _{a\in {\mathbb {D}}} \left( \int _{{\mathbb {D}}}\left( \int _{[0,1)}\frac{1}{(1-|tz|)|1-tz|}d\mu (t)\right) ^2(1-|\sigma _a(z)|^2)^p dA(z)\right) ^{\frac{1}{2}} \nonumber \\&\quad \le \sup _{a\in {\mathbb {D}}} \int _{[0,1)} \left( \int _{{\mathbb {D}}} \frac{1}{(1-|tz|)^2|1-tz|^2} (1-|\sigma _a(z)|^2)^p dA(z) \right) ^{\frac{1}{2}} d\mu (t)\nonumber \\&\quad \lesssim \sup _{a\in {\mathbb {D}}}(1-|a|^2)^{\frac{p}{2}}\int _{[0,1)}\frac{1}{(1-t)^{1-c}}d\mu (t)\big (\int _{{\mathbb {D}}}\frac{(1-|z|^2)^{p-2c}}{|1-tz|^2|1-{\bar{a}}z|^{2p}} dA(z)\big )^{\frac{1}{2}} \nonumber \\&\quad \lesssim \sup _{a\in {\mathbb {D}}}\int _{[0,1)}\frac{(1-|a|^2)^{\frac{p}{2}}}{(1-t)^{1-c}|1-ta|^{\frac{p}{2}+c}}d\mu (t)<\infty . \end{aligned}$$
(3.3)

Similarly, it follows from Lemma F and Proposition 2.1 that

$$\begin{aligned} \sup _{a\in {\mathbb {D}}}&\left( \int _{{\mathbb {D}}}\left( \int _{[0,1)}\frac{1}{|1-tz|^2}d\mu (t)\right) ^2(1-|\sigma _a(z)|^2)^p dA(z)\right) ^{\frac{1}{2}} \nonumber \\&\quad \le \sup _{a\in {\mathbb {D}}} \int _{[0,1)} \left( \int _{{\mathbb {D}}} \frac{1}{|1-tz|^4} (1-|\sigma _a(z)|^2)^p dA(z)\right) ^{\frac{1}{2}} d\mu (t)\nonumber \\&\quad \lesssim \sup _{a\in {\mathbb {D}}}\int _{[0,1)}\frac{(1-|a|^2)^{\frac{p}{2}}}{(1-t^2)^{1-\frac{p}{2}}|1-at|^p}d\mu (t)<\infty . \end{aligned}$$
(3.4)

From (3.1), (3.3) and (3.4), we get that \({\mathcal {C}}_\mu (f)\in {\mathcal {Q}}_p\). The proof is complete. \(\square \)

Remark 3

Set \(d\mu _0(x)=dx\) on [0, 1). Then \(d\mu _0\) is a Carleson measure and \({\mathcal {C}}_{\mu _0}(1)(z)=\frac{1}{z}\log \frac{1}{1-z}\). Clearly, the function \({\mathcal {C}}_{\mu _0}(1)\) is not in the Dirichlet space. Thus Theorem 3.1 does not hold when \(p=0\).

Note that \({\mathcal {Q}}_p={\mathcal {B}}\) for any \(p>1\). Theorem 3.1 generalizes Theorem 5 in [16] from the Bloch space \({\mathcal {B}}\) to all \({\mathcal {Q}}_p\) spaces. For \(p=1\), Theorem 3.1 gives an answer to a question raised in [16, p. 20]. The proof given here highlights the role of Proposition 2.1. In the next section, we give a more general result where an alternative proof of Theorem 3.1 will be provided.

4 s-Carleson measures and the range of another Cesàro-like operator acting on \(H^\infty \)

It is also natural to consider how the characterization of s-Carleson measures in Proposition 2.2 can play a role in the investigation of the range of Cesàro-like operators acting on \(H^\infty \). We consider this topic by another kind of Cesàro-like operators.

Suppose \(0<s<\infty \) and \(\mu \) is a finite positive Borel measure on [0, 1). For \(f(z)=\sum _{n=0}^\infty a_nz^n\) in \(H({\mathbb {D}})\), we define

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

Clearly, \({\mathcal {C}}_{\mu , 1}\) is equal to \({\mathcal {C}}_{\mu }\).

Lemma 4.1

Suppose \(0<s<\infty \) and \(\mu \) is a finite positive Borel measure on [0, 1). Then

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

for \(f\in H({\mathbb {D}})\).

Proof

The proof follows from a simple calculation with power series. We omit it. \(\square \)

We have the following result.

Theorem 4.2

Suppose \(0<s<\infty \) and \(\mu \) is a finite positive Borel measure on [0, 1). Let \(\max \{1, \frac{1}{s}\}<p<\infty \) and let X be a subspace of \(H({\mathbb {D}})\) with \(\Lambda ^p_{1/p}\subseteq X\subseteq {\mathcal {B}}\). Then \({\mathcal {C}}_{\mu , s}(H^\infty )\subseteq X\) if and only if \(\mu \) is an s-Carleson measure.

Proof

Let \({\mathcal {C}}_{\mu , s}(H^\infty )\subseteq X\). Then \({\mathcal {C}}_{\mu , s}(1)\in X\); that is, \(f_{\mu , s}\in X\). It follows from Proposition 2.2 that \(\mu \) is an s-Carleson measure.

On the other hand, let \(\mu \) be an s-Carleson measure and \(f\in H^\infty \). By Lemma 4.1, we see

$$\begin{aligned} {\mathcal {C}}_{\mu , s}(f)'(z)=\int _{[0,1)}\frac{tf'(tz)}{(1-tz)^s}d\mu (t)+ \int _{[0,1)}\frac{stf(tz)}{(1-tz)^{s+1}}d\mu (t), \quad z\in {\mathbb {D}}. \end{aligned}$$

Then

$$\begin{aligned}&\sup _{0<r<1}(1-r)^{1-\frac{1}{p}}\left( \frac{1}{2\pi }\int _0^{2\pi }|{\mathcal {C}}_{\mu , s}(f)'(re^{i\theta })|^p d\theta \right) ^{\frac{1}{p}}\nonumber \\&\quad \lesssim \Vert f\Vert _{\mathcal {B}}\sup _{0<r<1}(1-r)^{1-\frac{1}{p}} \left( \frac{1}{2\pi }\int _0^{2\pi }\left( \int _{[0,1)} \frac{1}{|1-tre^{i\theta }|^{s}(1-tr)} d\mu (t)\right) ^p d\theta \right) ^{\frac{1}{p}} \nonumber \\&\qquad +\Vert f\Vert _{H^\infty }\sup _{0<r<1}(1-r)^{1-\frac{1}{p}} \left( \frac{1}{2\pi }\int _0^{2\pi }\left( \int _{[0,1)} \frac{1}{|1-tre^{i\theta }|^{s+1}} d\mu (t)\right) ^p d\theta \right) ^{\frac{1}{p}}. \nonumber \\ \end{aligned}$$
(4.1)

Note that \(ps>1\). By the Minkowski inequality, Lemma E and Lemma A, we deduce

$$\begin{aligned} \sup _{0<r<1}&(1-r)^{1-\frac{1}{p}} \left( \frac{1}{2\pi }\int _0^{2\pi }\left( \int _{[0,1)} \frac{1}{|1-tre^{i\theta }|^{s}(1-tr)} d\mu (t)\right) ^p d\theta \right) ^{\frac{1}{p}} \nonumber \\&\quad \le \sup _{0<r<1}(1-r)^{1-\frac{1}{p}} \int _{[0,1)} \left( \frac{1}{2\pi }\int _0^{2\pi } \frac{1}{|1-tre^{i\theta }|^{sp}(1-tr)^p} d\theta \right) ^{\frac{1}{p}} d\mu (t) \nonumber \\&\quad \lesssim \sup _{0<r<1}(1-r)^{1-\frac{1}{p}} \int _{[0,1)} \frac{1}{(1-tr)^{s+1-\frac{1}{p}}} d\mu (t)<\infty , \nonumber \\ \end{aligned}$$
(4.2)

and

$$\begin{aligned} \sup _{0<r<1}&(1-r)^{1-\frac{1}{p}} \left( \frac{1}{2\pi }\int _0^{2\pi }\left( \int _{[0,1)} \frac{1}{|1-tre^{i\theta }|^{s+1}} d\mu (t)\right) ^p d\theta \right) ^{\frac{1}{p}}\nonumber \\&\quad \lesssim \sup _{0<r<1}(1-r)^{1-\frac{1}{p}} \int _{[0,1)} \left( \frac{1}{2\pi }\int _0^{2\pi }\frac{1}{|1-tre^{i\theta }|^{(s+1)p} } d\theta \right) ^{\frac{1}{p}} d\mu (t) \nonumber \\&\quad \lesssim \sup _{0<r<1}(1-r)^{1-\frac{1}{p}} \int _{[0,1)} \frac{1}{(1-tr)^{s+1-\frac{1}{p}}} d\mu (t)<\infty . \end{aligned}$$
(4.3)

From (4.1), (4.2) and (4.3), \({\mathcal {C}}_{\mu , s}(f)\in \Lambda ^p_{1/p}\). Note that \(\Lambda ^p_{1/p}\subseteq X\). The desired result follows. \(\square \)