1 Introduction
Let ${\mathbb D} $ and $\partial {\mathbb D}$ be the open unit disc and the unit circle in the complex plane, respectively. Denote by $H({\mathbb D})$ the space of all analytic functions on ${\mathbb D}$ . For any interval $I\subset \partial {\mathbb D}$ , let
be the Carleson box based on I, where $|I|$ is the normalized arc length of I. Let $0<\alpha <\infty $ and $\mu $ be a positive Borel measure on ${\mathbb D}$ . We say that $\mu $ is an $\alpha $ -Carleson measure (see [Reference Pau and Zhao19]) if
$\mu $ is the classical Carleson measure when $\alpha =1$ .
For $0<p<\infty $ , the Hardy space $H^p$ is the set of all $f\in H(\mathbb {D})$ with
Let $H^\infty $ denote the space of all bounded analytic functions with the supremum norm $\|f\|_{H^{\infty }}=\sup _{z\in {\mathbb D}}|f(z)|.$ For $0<p<\infty $ , the weighted Dirichlet space $\mathcal {D}_p$ is the space consisting of all $f\in H({\mathbb D})$ such that
where $dA(z)={1\over \pi }dxdy$ is the normalized Lebesgue area measure in ${\mathbb D}$ . Clearly, $\mathcal {D}_1$ is the classical Hardy space $H^2$ , and $\mathcal {D}_0$ is the classical Dirichlet space $\mathcal {D}$ .
Let $K:\lbrack 0,\infty )\to \lbrack 0,\infty )$ be a nondecreasing function and not identically zero. The space $Q_K$ consists of those functions $f\in H({\mathbb D})$ such that (see, e.g., [Reference Essén and Wulan7, Reference Wulan and Zhu25])
where $g(z,a)=-\log |\varphi _{a}(z)|$ is the Green function at a and $\varphi _a(z)={a-z \over 1-\bar {a}z}$ is the Möbius map that interchanges points $0$ and a. When $K(t)=t^p, 0<p<\infty $ , $Q_K$ is the $Q_p$ space (see, e.g., [Reference Xiao27]). For the case $p=1$ , $Q_1$ coincides with $BMOA$ , the class of holomorphic functions of bounded mean oscillation on ${\mathbb D}$ (see [Reference Girela10]), meanwhile for $p>1$ , $Q_p$ is the Bloch space.
Let $0\leq \lambda ,p \leq 1$ . Galanopoulos, Merch $\acute {{\textrm {a}}}$ n, and Siskakis [Reference Galanopoulos, Merchán and Siskakis8] defined the Dirichlet–Morrey space $\mathcal {D}_p^{\lambda }$ , which consists of all functions $f\in \mathcal {D}_p$ such that
By a simple calculation, we see that
It is easy to check that $\mathcal {D}_1^{\lambda }=\mathcal {L}^{2,\lambda }$ , the Morrey space (see [Reference Li, Liu and Lou13, Reference Liu and Lou14, Reference Wu and Xie23, Reference Wulan and Zhou24]). Moreover, $\mathcal {D}_p^{1}=Q_p,\mathcal {D}_p^{0}=\mathcal {D}_p$ , and
Let X be a Banach space of analytic functions on ${\mathbb D} $ and $ M(X)$ be the algebra of pointwise multipliers of $X $ , that is,
In this paper, we investigate the corona problem for $M(\mathcal {D}^\lambda _p) $ when $0<\lambda ,p<1$ . This problem can be stated as follows.
The $M(\mathcal {D}^\lambda _p) $ corona problem. Let $0<\lambda ,p<1$ , and let $g_1,g_2,\ldots ,g_n\in M(\mathcal {D}_p^{\lambda })$ . What are the sufficient and necessary conditions on these functions such that there exist $ f_1,f_2,\ldots ,f_n\in M(\mathcal {D}_p^{\lambda })$ such that
The corona theorem holds for $H^\infty $ by a famous result of Carleson (see [Reference Carleson6]). In [Reference Nicolau and Xiao15], Nicolau and Xiao gave the corona theorem on spaces $H^\infty \cap Q_p$ , which has been extended to $Q_p$ space by Xiao in [Reference Xiao26]. In addition, the same results on $ M(Q_p)$ and $H^\infty \cap Q_K$ were established by Pau in [Reference Pau16, Reference Pau17], respectively. In [Reference Li and Wulan12], Li and Wulan also showed that the corona theorem holds for $M(Q_K)$ . See [Reference Bao, Lou, Qian and Wulan4, Reference Treil21, Reference Trent22] and the references therein for more results of the corona theorem.
In this paper, we first show that the corona theorem holds for $M(\mathcal {D}^\lambda _p) $ when ${0<\lambda},{p<1}$ , that is, the unit disc is dense in the maximal ideal space of $M(\mathcal {D}^\lambda _p) $ . This can be reformulated in the following way.
Theorem 1 Let $0<\lambda ,p<1$ , and let $g_1,g_2,\ldots ,g_n\in M(\mathcal {D}_p^{\lambda })$ with
Then there exist $ f_1,f_2,\ldots ,f_n\in M(\mathcal {D}_p^{\lambda })$ such that
Banjade and Trent showed that the Wolff theorem holds for the multipliers of the Dirichlet space and weighted Dirichlet spaces in [Reference Banjade and Trent2, Reference Banjade and Trent3], respectively. Banjade also gave the Wolff theorem on spaces $H^\infty \cap Q_p$ and $M(Q_p)$ in [Reference Banjade1]. In [Reference Li and Wulan12], Li and Wulan showed that the Wolff theorem holds for the space $M(Q_K)$ . Motivated by these results, we investigate the Wolff theorem on the space $M(\mathcal {D}_p^{\lambda })$ . The second main result can be reformulated in the following way.
Theorem 2 Let $0<\lambda ,p<1$ , and let $g,g_1,g_2,\ldots ,g_n\in M(\mathcal {D}_p^{\lambda })$ with
Then there exist $ f_1,f_2,\ldots ,f_n\in M(\mathcal {D}_p^{\lambda })$ such that
The paper is organized as follows. In Section 2, we give some auxiliary properties, which will be used in the proofs of Theorems 1 and 2. In Section 3, we give a detail proof for Theorem 1. In Section 4, we give a proof for Theorem 2.
In this paper, we write $F\approx G$ if $F\lesssim G \lesssim F$ , where $G \lesssim F$ means that there exists a nonnegative constant C such that $ G\le CF$ .
2 Some auxiliary properties
In this section, we state some lemmas that are useful for the proof of the corona theorem on spaces $ M(\mathcal {D}^\lambda _p)$ .
Lemma 1 [Reference Galanopoulos, Merchán and Siskakis8, Propositions 2.1 and 2.2]
Let $0<\lambda ,p<1$ , $f\in H({\mathbb D})$ . Then the following statements hold:
-
(a) f belongs to $\mathcal {D}_p^{\lambda }$ if and only if
$$ \begin{align*}\sup_{I\subset \partial{\mathbb D}} {1\over |I|^{p\lambda }}\int_{S(I)}|f'(z)|^2(1-|z|^2)^pdA(z) <\infty.\end{align*} $$ -
(b) If $f\in \mathcal {D}_p^{\lambda }$ , then
$$ \begin{align*}|f(z)|\lesssim{\|f\|_{\mathcal{D}_p^{\lambda}} \over (1-|z|^2)^{{p\over 2}(1-\lambda)}}.\end{align*} $$
Lemma 2 [Reference Zhu28, Lemma 3.10]
Let $z\in {\mathbb D}$ , $\alpha>-1$ , and $\beta>0$ . Then
Lemma 3 [Reference Pau16, Lemma 2.2]
Let $0<p<1$ . If $ |f(z)|^2(1-|z|^2)^pdA(z)$ is a p-Carleson measure, then $ |f(z)|dA(z)$ is a Carleson measure.
Lemma 4 Let $0<\lambda ,p<1$ such that
has sense and it defines an analytic function. If $ |f(z)|^2(1-|z|^2)^pdA(z)$ is a $p\lambda $ -Carleson measure, then $ |T_f(z)|^2(1-|z|^2)^pdA(z)$ is also a $p\lambda $ -Carleson measure.
Proof For the Carleson box $S(I)$ , it is obvious that
From Lemma 7.2.2 in [Reference Xiao27], we have
Applying the Cauchy–Schwarz inequality, we obtain that
Therefore, $ |f(z)|dA(z)$ is a $1+{p\lambda \over 2}-{p\over 2} $ -Carleson measure. This deduces that
So $ |T_f(z)|^2(1-|z|^2)^pdA(z)$ is a $p\lambda $ -Carleson measure. The proof is complete.▪
Let $f\in H({\mathbb D})$ . The Volterra integral operator $J_f$ , which was first introduced by Pommerenke in [Reference Pommerenke20], is defined as
Its related operator $I_f$ is defined by
It is clear that $M_f(g)(z)=J_f(g)(z)+I_f(g)(z)+f(0)g(0)$ , where $M_f(g)(z)=f(z)g(z)$ , called the multiplication operator.
Lemma 5 Let $0<\lambda ,p<1$ . Then $f\in M(\mathcal {D}_p^{\lambda })$ if and only if $f\in H^\infty $ and the measure $d\mu (z)=|f'(z)|^2|g(z)|^2(1-|z|^2)^pdA(z)$ is a $p\lambda $ -Carleson measure for any $g\in \mathcal {D}_p^{\lambda }$ .
Proof Suppose first that $f\in M(\mathcal {D}_p^{\lambda })$ . For any $g\in \mathcal {D}_p^{\lambda }$ , we have $M_f(g)\in \mathcal {D}_p^{\lambda }$ . Corollary 3.1 in [Reference Galanopoulos, Merchán and Siskakis8] gives that $f\in H^\infty $ , which implies that $I_f(g)\in \mathcal {D}_p^{\lambda }$ . Therefore, $J_f(g)\in \mathcal {D}_p^{\lambda }$ , that is, the measure $d\mu (z)=|f'(z)|^2|g(z)|^2(1-|z|^2)^pdA(z)$ is a $p\lambda $ -Carleson measure.
Conversely, suppose that $f\in H^\infty $ and the measure $\mu $ is a $p\lambda $ -Carleson measure. For any $g\in \mathcal {D}_p^{\lambda }$ , using the norm of $\mathcal {D}_p^{\lambda }$ , we can easily obtain the desired result.▪
Let $0<\lambda ,p<1$ , $f\in L^2(\partial {\mathbb D})$ . We say that $f\in \mathcal {D}_p^{\lambda }(\partial {\mathbb D}) $ if
A function of $H^2$ belongs to $\mathcal {D}_p^{\lambda }$ if and only if its boundary values belong to $ \mathcal {D}_p^{\lambda }(\partial {\mathbb D}) $ (see [Reference Galanopoulos, Merchán and Siskakis8, Theorem 2.1]).
Lemma 6 [Reference Bao and Pau5, Lemma 2.2]
Let $0<p<1$ and $f\in L^2(\partial {\mathbb D})$ . Let $F\in \mathcal {C}^1({\mathbb D})$ such that $\lim _{r\to 1^-}F(re^{i\theta })=f(e^{i\theta })$ for almost everywhere $ e^{i\theta }\in \partial {\mathbb D}$ . For any interval $I\subset \partial {\mathbb D}$ ,
Using Lemma 6, we get the following result.
Corollary 1 Let $0<\lambda ,p<1$ and $f\in L^2(\partial {\mathbb D})$ . Let $F\in \mathcal {C}^1({\mathbb D})$ such that $\lim _{r\to 1^-}F(re^{i\theta })=f(e^{i\theta })$ for almost everywhere $ e^{i\theta }\in \partial {\mathbb D}$ . If $ |\nabla F(z)|^2(1-|z|^2)^pdA(z)$ is a $p\lambda $ -Carleson measure, then $ f\in \mathcal {D}_p^{\lambda }(\partial {\mathbb D})$ .
For $z=x+i y$ , define
Let f be $\mathcal {C}^{1}$ and bounded on ${\mathbb D}$ . Then the $\bar {\partial }$ -equation $\bar {\partial } g=f$ has a standard solution
on ${\mathbb D}$ . In this paper, Jones’ solution of the $\bar {\partial }$ -equation is suitable for our purpose, i.e., the following lemma (see [Reference Jones11, Reference Xiao26]).
Lemma 7 Let $ \mu $ be a Carleson measure on ${\mathbb D}$ . If for $z \in {\mathbb D} \cup \partial {\mathbb D}$ and $\varsigma \in {\mathbb D}$ ,
then
satisfies $\bar {\partial }G=\mu $ . Moreover, if $z \in \partial {\mathbb D}$ , then the above integral converges absolutely and obeys
and hence $G\in L^\infty (\partial {\mathbb D})$ and $\|G\|_{L^\infty (\partial {\mathbb D})}\le C\|\mu \|_1.$
Thus, if $|g(z)|dA(z)$ is a Carleson measure, then the equation $\bar {\partial }G=g $ has a solution $G\in L^\infty (\partial {\mathbb D})$ .
3 Proof of Theorem 1
Proof of Theorem 1
Let $g_1,g_2,\ldots ,g_n\in M(\mathcal {D}_p^{\lambda })$ such that (1) holds. Without loss of generality, we suppose that $\|g_j\|_{H^\infty }\le 1,j=1,2,\ldots ,n$ . Consider the equation
and its nonanalytic solutions
For $ 1\le k,j\le n$ , we assume that the $\bar {\partial }$ -equation
has solution $b_{j,k}$ such that $b_{j,k} f \in \mathcal {D}_p^{\lambda }(\partial {\mathbb D})$ for every $f\in \mathcal {D}_p^{\lambda }$ . By a simple calculation, we see that
satisfy $\sum _{j=1}^ng_jf_j=\sum _{j=1}^ng_j\varPhi _j=1$ .▪
Since
we have $f_j\in H({\mathbb D})$ . By (1), we have that (see also [Reference Pau16])
To obtain that $ f_j\in M(\mathcal {D}_p^{\lambda })$ , we need to show that $ ff_j\in \mathcal {D}_p^{\lambda }$ for any $ f\in \mathcal {D}_p^{\lambda }$ . Applying Lemmas 1 and 5, we get
Since $|\varPhi _j(z)|\le \delta ^{-1}$ , we obtain
Therefore,
which implies that $\varPhi _j f \in \mathcal {D}_p^{\lambda }$ . Since $ g_k\in M(\mathcal {D}_p^{\lambda })$ , $b_{j,k}f\in \mathcal {D}_p^{\lambda }, 1\le j,k\le n$ , we obtain that
This implies that
To finish the proof, we only need to show that there exists a $ b_{j,k}$ such that $\bar {\partial }b_{j,k}=\varPhi _j\bar {\partial }\varPhi _k$ , where $b_{j,k} f \in \mathcal {D}_p^{\lambda }(\partial {\mathbb D})$ for every $f\in \mathcal {D}_p^{\lambda }$ . Consider the $\bar {\partial }$ -equation $ \bar {\partial }u=G$ , where $G=\varPhi _j\bar {\partial }\varPhi _k$ . Applying (1), we get that
Since $g_k\in M(\mathcal {D}_p^{\lambda })\subset Q_p $ (see Corollary 3.1 in [Reference Galanopoulos, Merchán and Siskakis8]), then $ |g^{\prime }_k(z)|^2(1-|z|^2)^pdA(z)$ is a p-Carleson measure. So, $|G(z)|^2(1-|z|^2)^pdA(z) $ is also a p-Carleson measure. Using Lemma 3, we have that $|G(z)|dA(z) $ is a Carleson measure. So, by Lemma 7, we get Jones’ solution u of the $\bar {\partial } $ problem $\bar {\partial }u=G$ . Moreover, u has the boundary values in $L^\infty (\partial {\mathbb D})$ . Let
From the proof of Theorem 3.1 of [Reference Nicolau and Xiao15], we have that v has the same boundary values as $zu$ and
Since $|G(z)|dA(z) $ is a Carleson measure, then $ v\in L^\infty ({\mathbb D})$ from [Reference Jones11].
Now, we show that $vf\in \mathcal {D}_p^{\lambda }(\partial {\mathbb D})$ for every $f\in \mathcal {D}_p^{\lambda }$ . For this purpose, we need to prove that for any $f\in \mathcal {D}_p^{\lambda }$ ,
is a $p\lambda $ -Carleson measure by Corollary 1.
Since $f\in \mathcal {D}_p^{\lambda }$ and $ v\in L^\infty ({\mathbb D})$ , we see that $ |v(z)|^2|\nabla f(z)|^2(1-|z|^2)^pdA(z)$ is a $p\lambda $ -Carleson measure. Then it remains to verify that $ |\nabla v(z)|^2| f(z)|^2(1-|z|^2)^pdA(z)$ is also a $p\lambda $ -Carleson measure. By inequalities (3) and (4), we obtain that
where
We write $ fT_{|g^{\prime }_k|}=T_{f|g^{\prime }_k|}+(fT_{|g^{\prime }_k|}-T_{f|g^{\prime }_k|} )$ . Since $g_k\in M(\mathcal {D}_p^{\lambda })$ , by Lemma 5, we get that $ |f(z)|^2|g^{\prime }_k(z)|^2(1-|z|^2)^pdA(z)$ is a $p\lambda $ -Carleson measure. Lemma 4 yields that $|T_{f|g^{\prime }_k|}|^2(1-|z|^2)^pdA(z)$ is also a $p\lambda $ -Carleson measure. Applying the identity
the Cauchy–Schwarz inequality, and $g_k\in M(\mathcal {D}_p^{\lambda })\subset Q_p\subset \mathcal {D}_p^{\lambda } $ (see Corollary 3.1 in [Reference Galanopoulos, Merchán and Siskakis8]), we obtain that
Using Proposition 4.27 of [Reference Wulan and Zhu25] and following the ideas of the proof of Lemma 1 in [Reference Liu and Lou14], we obtain that
The Cauchy–Schwarz inequality and Lemma 2 yield that
Since
applying Lemma 2 and Lemma D in [Reference Pau and Peláez18] and changing the variable $z=\varphi _w(u)$ , we have
This implies that $ |\nabla (vf)(z)|^2(1-|z|^2)^pdA(z)$ is a $p\lambda $ -Carleson measure. Then Corollary 1 deduces that $vf\in \mathcal {D}_p^{\lambda }(\partial {\mathbb D})$ .▪
Let $g=(g_1,g_2,\ldots ,g_n)\in H({\mathbb D})^n$ and $f=(f_1,f_2,\ldots ,f_n)\in H({\mathbb D})^n$ . Consider the operator $M_g$ defined by
Similarly to the proof of Theorem 1.2 in [Reference Pau16] (using the ideas of Theorem 3.1 of [Reference Xiao26]), by Theorem 1, we obtain the following result. We omit the details of the proof.
Theorem 3 Let $0<\lambda ,p<1$ and $g\in H({\mathbb D})^n$ . The following statements are equivalent:
-
(a) $M_g: M(\mathcal {D}_p^{\lambda })\times \cdots \times M(\mathcal {D}_p^{\lambda })\to M(\mathcal {D}_p^{\lambda })$ is surjective;
-
(b) $M_g: \mathcal {D}_p^{\lambda }\times \cdots \times \mathcal {D}_p^{\lambda }\to \mathcal {D}_p^{\lambda }$ is surjective; and
-
(c) $g_1,\ldots ,g_n\in M(\mathcal {D}_p^{\lambda })$ and (1) holds.
4 Proof of Theorem 2
In this section, we give a detail proof for Theorem 2, i.e., we show that the Wolff theorem holds for the algebra of pointwise multipliers of the Dirichlet–Morrey space $\mathcal {D}_p^{\lambda } $ .
Proof of Theorem 2
First, considering the trivial case $g\equiv 0$ , we can choose $f_j\equiv 0, j=1,2,\ldots ,n$ . Then the desired result is obtained. Next, we consider the nontrivial case. Let $g,g_1,g_2,\ldots ,g_n\in M(\mathcal {D}_p^{\lambda })$ satisfying (2). Without loss of generality, we suppose that $g,g_1,g_2,\ldots ,g_n$ are the analytic function on $\bar {{\mathbb D}}$ and satisfying $\|g\|_{H^\infty }\le 1$ , $\|g_j\|_{H^\infty }\le 1,j=1,2,\ldots ,n$ (see [Reference Garnett9, Proof of Theorem 2.3, p. 319]). Let
It is obvious that $\Psi _j $ , satisfying $|\Psi _j(z)|\le 1$ , is the $C^\infty $ function on $\bar {{\mathbb D}}$ such that $ \sum _{j=1}^ng_j\Psi _j=g$ . For $ 1\le k,j\le n$ , we assume that the following equation
has solution $b_{j,k}\in M(\mathcal {D}_p^{\lambda })$ and $\|b_{j,k}\|_{H^\infty }\le 1$ . We write
Then we get that
In addition,
Therefore, $f_j\in H({\mathbb D})$ . Moreover, $\|f_j\|_{H^\infty } \leq 1+2n$ , and hence $f_j\in H^\infty $ .▪
Since $ { \partial \overline {g_j}\over \partial \bar {z}}=\overline {g^{\prime }_j}$ , and
we see that
Similarly,
Therefore, by the assumption, we get that
Now, we show that $ f_j\in M(\mathcal {D}_p^{\lambda })$ . To do this, we need to show that $ ff_j\in \mathcal {D}_p^{\lambda }$ for any $ f\in \mathcal {D}_p^{\lambda }$ . First of all, we are going to prove that $fg^2\Psi _j \in \mathcal {D}_p^{\lambda }$ . Since $g\in H^\infty $ and $|\Psi _j(z) |\leq 1$ , we get that
Since $g\in M(\mathcal {D}_p^{\lambda }) $ , applying Lemma 5, we get
Using the fact that $|g\nabla \Psi _j|^2\lesssim |g'|^2+ \sum _{k=1}^n|g^{\prime }_k|^2 $ and Lemma 5 again, we obtain that
So, using the fact that
we get that $fg^2\Psi _j \in \mathcal {D}_p^{\lambda }$ .
Now, we are going to prove that the other part of $ f_jf$ belongs to $ \mathcal {D}_p^{\lambda } $ . Since $ g_k\in M(\mathcal {D}_p^{\lambda })$ , $b_{j,k}f\in \mathcal {D}_p^{\lambda }, 1\le j,k\le n$ , we obtain that
Hence, $f_jf \in \mathcal {D}_p^{\lambda },$ that is, $ f_j\in M(\mathcal {D}_p^{\lambda }),j=1,2,\ldots ,n$ .
Now, it remains to show that we can find a solution of $\bar {\partial }b_{j,k}=g\Psi _j\bar {\partial }\Psi _k$ satisfying $ b_{j,k}\in M(\mathcal {D}_p^{\lambda })$ . To do this, it is sufficient to deal with the equation $ \bar {\partial }u=G$ , where $G=g\Psi _j\bar {\partial }\Psi _k$ . It is easy to see that
Since $ M(\mathcal {D}_p^{\lambda })\subset Q_p$ (see Corollary 3.1 in [Reference Galanopoulos, Merchán and Siskakis8]) and $ g_k\in M(\mathcal {D}_p^{\lambda })$ , we obtain that $|G(z)|^2(1-|z|^2)^pdA(z) $ is a p-Carleson measure. Lemma 3 yields that $|G(z)|dA(z) $ is a Carleson measure. Then we can obtain a solution $v\in M(\mathcal {D}_p^{\lambda })$ by a similar method to Theorem 1.