1 Introduction

Let \(\mathcal{{L}}(\mathcal{{H}})\) be the algebra of all bounded linear operators on a separable complex Hilbert space \(\mathcal{{H}}\). For \(T\in {\mathcal {L}}(\mathcal {H})\), \(T^{*}\) denotes the adjoint of T. An operator \(T\in {\mathcal {L}}(\mathcal {H})\) is said to be self-adjoint if \(T=T^{*}\), isometric if \(T^{*}T=I\), normal if \([T^{*},T]=0\), hyponormal if \([T^{*},T]\ge 0\), quasinormal if \([T^{*}T, T]=0\), and binormal if \([T^{*}T, TT^{*}]=0\), respectively, where \([R,S]:=RS-SR.\)

H-Toeplitz operators have been studied in various spaces. Recently, the authors in [13] studied the essential conditions for H-Toeplitz operators to become a co-isometry and a partial isometry, explored their invariant subspaces and kernels, and investigated their compactness and Fredholmness. In particular, they showed a nonzero H-Toeplitz operator cannot be a Fredholm operator on the Bergman space. Moreover, they considered the necessary and sufficient conditions for the commutativity of H-Toeplitz operators. In [25], the authors provided a characterization of the commutativity of H-Toeplitz operators with quasihomogeneous symbols on the Bergman space. In [22], the authors explored the characteristics of H-Toeplitz operators on the Bergman space and offered essential criteria for identifying both contractive and expansive operators. Additionally, the authors in [14] studied the slant Toeplitz operators on the Hardy space.

Basic properties of Toeplitz operators on the Hardy space and (weighted) Begman space can be found in [2, 7, 8, 18, 20, 28]. Recently, many authors have characterized the hyponormality of Toeplitz operators on the Bergman spaces and the weighted Bergman spaces (cf. [16, 17, 19, 21, 26, 27, 29]). The theory of Toeplitz operators is a vast and significant field that has made fundamental contributions to several problems in functional analysis and mathematical physics.

Several decades ago, researchers extensively studied contractive and expansive operators (cf. [3, 5, 6]). In particular, in [9], the authors investigated the problem of invariant subspaces for contractive operators. In [22], the authors studied the contractivity and expansivity of H-Toeplitz operators with analytic, co-analytic and harmonic symbols on the Bergman spaces.

In this paper, we study several classes of H-Toeplitz operators on the function spaces. In Sect. 2, we focus on the self-adjointness of H-Toeplitz operators on the Hardy space \(H^2\). Moreover, we consider complex symmetric H-Toeplitz operator on \(H^2.\) Furthermore, we investigate hyponormality, quasinormality, and binormality of H-Toeplitz operators. In particular, we show that for \(\varphi \in L^{\infty }\) the adjoint of H-Toeplitz operators is hyponormal. As an application of this, such an operator has a nontrivial invariant subspace. In Sect. 3, we will investigate the the algebraic properties of H-Toeplitz operators on the weighted Bergman spaces \(A^2_{\alpha }(\mathbb D)\). More concretely, we introduce the notion of H-Toeplitz operators on the weighted Bergman spaces, which combine the properties of both Toeplitz and Hankel operators. The importance of this notion is that it provides a unifying framework for a class of operators on the weighted Bergman spaces, which includes both Toeplitz and Hankel operators. Furthermore, we establish a convenient and explicit criterion for determining the contractivity and expansivity of H-Toeplitz operators.

2 H-Toeplitz operators on the Hardy spaces

Let \(\mathbb D\) be the open unit disk in the complex plane and let \(\mathbb T (\equiv \partial \mathbb D)\) be the unit circle. Let \(L^{\infty }(\mathbb T)\) denote the set of all essentially bounded measurable functions on \({\mathbb T}\). The Hilbert Hardy space \(H^2({\mathbb T})\) consists of all analytic functions f with the power series representation

$$\begin{aligned} f(z)=\sum _{n=0}^{\infty }a_nz^n\ \text{ where }\ \sum _{n=0}^{\infty }|a_n|^2<\infty . \end{aligned}$$

For a convenience, we denote \(L^{\infty }(\mathbb T)\) and \(H^{2}(\mathbb T)\) by \(L^{\infty }\) and \(H^2\), respectively. For any \(\varphi \in L^{\infty }\), the multiplication operator \(M_{\varphi }\) is defined by \(M_{\varphi }(f)=\varphi f\) for \(f\in H^2\), the Toeplitz operator \(T_{\varphi }:{H^2}\rightarrow {H^2}\) is defined by

$$\begin{aligned} T_{\varphi }{f}=P(\varphi f) \end{aligned}$$

for \(f\in {H^2}\) where P denotes the orthogonal projection of \(L^2\) onto \(H^2\), and the Hankel operator \(H_{\varphi }:{H^2}\rightarrow {H^2}\) is defined by

$$\begin{aligned} H_{\varphi }{f}=PM_{\varphi }Jf \end{aligned}$$

where \(J:{H^2}\rightarrow ({H^2})^{\perp }\) denotes the flip operator given by \(J(e_n)=e_{-n-1}\) for all \(n\ge 0\) where \(\{e_n\}_{n=-\infty }^{\infty }\) is an orthonormal basis for \(L^2\). Note that \(T_{\varphi }\) is bounded if and only if \(\varphi \in L^{\infty }\) and, in which case, \(\Vert T_{\varphi } \Vert =\Vert {\varphi }\Vert _{\infty }\).

Notation 2.1

Throughout this paper, a dilation operator K from \(H^2\) to \(L^2\) is denoted as \(K(e_{2n})=e_n\) and \(K(e_{2n+1})=e_{-n-1}\) for all \(n=0,1,2,\ldots \) where \(\{e_n\}_{n=-\infty }^{\infty }\) is an orthonormal basis for \(L^2\).

Let \({\mathbb N}\), \({\mathbb N}_0\), \({\mathbb Z}\), \({\mathbb R}\), and \({\mathbb C}\) be the set of positive integers, \(\text {non-negative integers}\), integers, real numbers, and complex numbers, respectively. A dilation operator K is bounded from \(H^2\) to \(L^2\) with \(\Vert K\Vert =1\) and its adjoint \(K^{*}\) from \(L^2\) to \(H^2\) is defined as

$$\begin{aligned} K^{*}(e_n)=e_{2n}~\text{ and }~K^{*}(e_{-n-1})=e_{2n+1} \end{aligned}$$

for all \(n=0,1,2,\ldots \). Thus \(K^{*}K=I\) on \(H^2\) and \(K^{*}K=I\) on \(L^2\). Indeed, since \(KK^{*}e_n=Ke_{2n}=e_n\) for each \(n\ge 0\), it follows that \(KK^{*}=I\) on \(H^2\). Moreover, since \(KK^{*}e_{-n-1}=Ke_{2n+1}=e_{-n-1}\) for each \(n\in {\mathbb N}\), we know that \(KK^{*}=I\) on \((H^2)^{\bot }\). Thus \(KK^{*}=I\) on \(L^2\). Hence K is unitary from \(H^2\) to \(L^2\).

The authors in [1] have introduced “H-Toeplitz operators" motivated by the Toeplitz, Hankel, and Slant Toeplitz operators.

Definition 2.2

For \(\varphi \in L^{\infty }\), an H-Toeplitz operator \(S_{\varphi }\) with the symbol \(\varphi \) on \(H^2\) is defined by

$$\begin{aligned} S_{\varphi }f=PM_{\varphi }Kf \end{aligned}$$

for each \(f\in H^2\) where P denotes the orthogonal projection of \(L^2\) onto \(H^2\).

In this case, \(\Vert S_{\varphi }\Vert =\Vert PM_{\varphi }K\Vert \le \Vert M_{\varphi }\Vert =\Vert \varphi \Vert _{\infty }.\) Note that if \(\{e_n\}_{n=0}^{\infty }\) denotes the orthonormal basis for \(H^2\), then

$$\begin{aligned} S_{\varphi }e_{2n}=PM_{\varphi }K{e_{2n}}=PM_{\varphi }e_n=T_{\varphi }e_n \end{aligned}$$

and

$$\begin{aligned} S_{\varphi }e_{2n+1}=PM_{\varphi }K{e_{2n+1}}=PM_{\varphi }e_{-n-1}=PM_{\varphi }Je_{n}=H_{\varphi }e_n \end{aligned}$$

for each \(n=0,1,2\ldots .\) Note that for \(\varphi \in L^{\infty }\), the adjoint of \(S_{\varphi }\) on \(H^2\) is given by

$$\begin{aligned} S_{\varphi }^{*}=K^{*}M_{\overline{\varphi }}. \end{aligned}$$

2.1 Basic properties of an H-Toeplitz operator

In this section, we consider the basic properties of an H-Toeplitz operator. We first study the self-adjointness of H-Toeplitz operators on \(H^2\).

Theorem 2.3

If \(\varphi (z)=\sum _{j=-\infty }^{\infty }a_je_j\) with respect to the orthonormal basis \(\mathcal {B}=\{e_n\}_{n=0}^{\infty }\) in \(L^{\infty }\) and \(S_{\varphi }\) is an H-Toeplitz operator on \(H^2\), then the matrices of \(S_{\varphi }\) and \(S_{\varphi }^{*}\) are represented as

$$\begin{aligned}{}[S_{\varphi }]_\mathcal {B}=\begin{pmatrix} a_0 &{}\quad a_1 &{}\quad a_{-1} &{}\quad a_2 &{}\quad a_{-2} &{}\quad a_3 &{}\quad a_{-3} &{}\quad \cdots \\ a_1 &{}\quad a_2 &{}\quad a_{0} &{}\quad a_{3} &{}\quad a_{-1} &{}\quad a_4 &{}\quad a_{-2} &{}\quad \cdots \\ a_2 &{}\quad a_3 &{}\quad a_{1} &{}\quad a_{4} &{}\quad a_{0} &{}\quad a_5 &{}\quad a_{-1} &{}\quad \cdots \\ a_3 &{}\quad a_4 &{}\quad a_{2} &{}\quad a_{5} &{}\quad a_{1} &{}\quad a_6 &{}\quad a_{0} &{}\quad \cdots \\ a_4 &{}\quad a_5 &{}\quad a_{3} &{}\quad a_{6} &{}\quad a_{2} &{}\quad a_7 &{}\quad a_{1} &{}\quad \cdots \\ a_5 &{}\quad a_6 &{}\quad a_{4} &{}\quad a_{7} &{}\quad a_{3} &{}\quad a_8 &{}\quad a_{2} &{}\quad \cdots \\ a_6 &{}\quad a_7 &{}\quad a_{5} &{}\quad a_{8} &{}\quad a_{4} &{}\quad a_9 &{}\quad a_{3} &{}\quad \cdots \\ \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \ddots \end{pmatrix} \end{aligned}$$
(1)

and

$$\begin{aligned}{}[S_{\varphi }^{*}]_\mathcal {B}=\begin{pmatrix} \overline{a_0} &{}\quad \overline{a_1} &{}\quad \overline{a_{2}} &{}\quad \overline{a_3} &{}\quad \overline{a_{4}} &{}\quad \overline{a_5} &{}\quad \overline{a_{6}} &{}\quad \cdots \\ \overline{a_1} &{}\quad \overline{a_2} &{}\quad \overline{a_{3}} &{}\quad \overline{a_4} &{}\quad \overline{a_{5}} &{}\quad \overline{a_6} &{}\quad \overline{a_{7}} &{}\quad \cdots \\ \overline{a_{-1}} &{}\quad \overline{a_0} &{}\quad \overline{a_{1}} &{}\quad \overline{a_2} &{}\quad \overline{a_{3}} &{}\quad \overline{a_4} &{}\quad \overline{a_{5}} &{}\quad \cdots \\ \overline{a_{2}} &{}\quad \overline{a_3} &{}\quad \overline{a_{4}} &{}\quad \overline{a_5} &{}\quad \overline{a_{6}} &{}\quad \overline{a_7} &{}\quad \overline{a_{8}} &{}\quad \cdots \\ \overline{a_{-2}} &{}\quad \overline{a_{-1}} &{}\quad \overline{a_{0}} &{}\quad \overline{a_1} &{}\quad \overline{a_{2}} &{}\quad \overline{a_3} &{}\quad \overline{a_{4}} &{}\quad \cdots \\ \overline{a_{3}} &{}\quad \overline{a_{4}} &{}\quad \overline{a_{5}} &{}\quad \overline{a_6} &{}\quad \overline{a_{7}} &{}\quad \overline{a_8} &{}\quad \overline{a_{9}} &{}\quad \cdots \\ \overline{a_{-3}} &{}\quad \overline{a_{-2}} &{}\quad \overline{a_{-1}} &{}\quad \overline{a_0} &{}\quad \overline{a_{1}} &{}\quad \overline{a_2} &{}\quad \overline{a_{3}} &{}\quad \cdots \\ \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \ddots \end{pmatrix}. \end{aligned}$$

Furthermore, \([S_{\varphi }]_\mathcal {B}\) is self-adjoint if and only if \([S_{\varphi }]_\mathcal {B}=0.\)

Proof

We know that \(S_{\varphi }\) is self-adjoint if and only if \(a_0,a_1,a_2,a_3,a_5,a_8,\ldots \) are real and \(a_2=\overline{a_{-1}}\), \(a_3=\overline{a_0}\), \(a_2=\overline{a_3}\), \(a_{-2}=\overline{a_4}\), \(a_3=\overline{a_5},\) \(a_{-3}=\overline{a_6}\), and so on (cf. [1, Page 151]).

On the other hand, the (ij) entry of the matrix \([S_{\varphi }]_\mathcal {B}\) is given by

$$\begin{aligned} a_{i,j}={\left\{ \begin{array}{ll} a_i~~~~~~~&{}\text{ if }~j=0\\ a_{i-n}~~~~&{}\text{ if }~j=2n\\ a_{i+n+1}~&{}\text{ if }~j=2n+1 \end{array}\right. } \end{aligned}$$

(see [1]). And the (ij) entry of the matrix \([S_{\varphi }^{*}]_\mathcal {B}\) is given by

$$\begin{aligned} \overline{a_{j,i}}={\left\{ \begin{array}{ll} \overline{a_j}~~~~~~~&{}\text{ if }~i=0\\ \overline{a_{j-n}}~~~~&{}\text{ if }~i=2n\\ \overline{a_{j+n+1}}~&{}\text{ if }~i=2n+1. \end{array}\right. } \end{aligned}$$
(2)

Thus \([S_{\varphi }]_\mathcal {B}=[S_{\varphi }^{*}]_\mathcal {B}\) if and only if \(a_{i,j}=\overline{a_{j,i}}\) for all ij. Hence \([S_{\varphi }]_\mathcal {B}\) is self-adjoint if and only if \(a_j=a_0\in {\mathbb R}\) for all \(j\in {\mathbb Z}\) if and only if \(a_j=0\) for all j since \([S_{\varphi }]_\mathcal {B}\) is bounded. \(\square \)

Proposition 2.4

Let \(\varphi \in L^{\infty }\) and \(S_{\varphi }\) be an H-Toeplitz operator on \(H^2\). Then \(S_{\varphi }\) is an isometry on \(H^2\) if and only if \(M_{\overline{\varphi }}PM_{\varphi }=I\) on \(L^2\). In particular, \(\varphi \) is not inner.

Proof

Since \(S_{\varphi }^{*}=K^{*}M_{\overline{\varphi }}\), we have \(S_{\varphi }^{*}S_{\varphi }=K^{*}M_{\overline{\varphi }}PM_{\varphi }K\). Then \(K^{*}M_{\overline{\varphi }}PM_{\varphi }K=I\) on \(H^2\). Hence \(M_{\overline{\varphi }}PM_{\varphi }=I\) on \(L^2\) since K is unitary from \(H^2\) to \(L^2.\) Thus \(S_{\varphi }\) is an isometry on \(H^2\) if and only if \(M_{\overline{\varphi }}PM_{\varphi }=I\) on \(L^2\).

If \(\varphi \) is inner, then \(M_{\overline{\varphi }}PM_{\varphi }-I=M_{\overline{\varphi }}M_{\varphi }-I=M_{|{\varphi }|^2}-I=0\). Thus \(S_{\varphi }\) is an isometry on \(H^2\). But, if \(\varphi \) is inner, then \(S_{\varphi }^{*}\) is an isometry on \(H^2\) (cf. [1]), and so \(S_{\varphi }^{*}\) is normal. Therefore, \(\varphi =0\) from [1], which is a contradiction. \(\square \)

Next, we study complex symmetric H-Toeplitz operator on \(H^2\). A conjugation on \(\mathcal H\) is an antilinear operator \({C}: \mathcal{H}\rightarrow \mathcal{H}\) which satisfies \({C}^{2}=I\) and \(\langle {C}x, {C}y \rangle =\langle y, x\rangle \) for all \(x,y\in \mathcal{H}\). If C is a conjugation on \(\mathcal {H}\), then there exists an orthonormal basis \(\{e_n\}_{n=0}^{\infty }\) for \(\mathcal{H}\) such that \({C}e_n=e_n\) for all n (see [10]). An operator \(T\in \mathcal{L(H)}\) is complex symmetric if there exists a conjugation C on \(\mathcal{H}\) such that \(T= {C}T^{*}{C}\). Complex symmetric operators have been widely studied by several mathematicians (see [10,11,12, 23, 24] for more details).

Proposition 2.5

For \(\varphi \in L^{\infty }\), let \(S_{\varphi }\) be an H-Toeplitz operator on \(H^2\) and C be a conjugation on \(L^2\) given by \(Cf(z)=\overline{f(\overline{z})}\) for \(f\in H^2\). Then \(S_{\varphi }\) is complex symmetric with the conjugation C if and only if \(T_{\overline{\varphi (\overline{z})}}e_{n}=K^{*}M_{\overline{\varphi (z)}}e_{2n}\) and \(H_{\overline{\varphi (\overline{z})}}e_n=K^{*}M_{\overline{\varphi (z)}}e_{2n+1}\) for \(n\in {\mathbb N}_0\).

Proof

Let C be a conjugation on \(L^2\) given by \(Cf(z)=\overline{f(\overline{z})}\) for \(f\in H^2\). Then \(Ce_n=e_n\) for \(n\ge 0\) and so \(CP=PC\) on \(L^2\) from [24]. Thus for \(n\ge 0,\)

$$\begin{aligned} CS_{\varphi }Ce_{2n}=CS_{\varphi }e_{2n}=CT_{\varphi }(e_n)=CP(\varphi e_n)=PC(\varphi e_n)=T_{\overline{\varphi (\overline{z})}}e_n \end{aligned}$$

and

$$\begin{aligned} CS_{\varphi }Ce_{2n+1}=CH_{\varphi }(e_n)=CPM_{\varphi }(Je_{n})=CP{\varphi }(Je_{n})=H_{\overline{\varphi (\overline{z})}}e_n \end{aligned}$$

hold. Since \(S_{\varphi }^{*}=K^{*}M_{\overline{\varphi }}\) for \(\varphi \in L^{\infty }\), we obtain that \(S_{\varphi }\) is complex symmetric with the conjugation C if and only if \(T_{\overline{\varphi (\overline{z})}}e_{n}=K^{*}M_{\overline{\varphi (z)}}e_{2n}\) and \(H_{\overline{\varphi (\overline{z})}}e_n=K^{*}M_{\overline{\varphi (z)}}e_{2n+1}\) for \(n\in {\mathbb N}_0\). \(\square \)

Theorem 2.6

For \(\varphi \in L^{\infty }\), let \(S_{\varphi }\) be an H-Toeplitz operator on \(H^2\). Assume that C is a conjugation on \(L^2\) given by \(Cf(z)=\overline{f(\overline{z})}\) for \(f\in H^2\) and \(C_{\mu ,\lambda }\) is a conjugation on \(L^2\) given by \(C_{\mu , \lambda }f(z)=\mu \overline{f(\lambda \overline{z})}\) for \(f\in H^2\) with \(|\lambda |=|\mu |=1\). Then the following statements are equivalent:

(i) \(S_{\varphi }\) is complex symmetric with the conjugation C.

(ii) \(S_{\varphi }\) is complex symmetric with the conjugation \(C_{\mu , \lambda }\).

(iii) \(\varphi =0.\)

Proof

(i) \(\Leftrightarrow \) (iii) Let \(\varphi (z)=\sum _{j=-\infty }^{\infty }a_je_j\) be with respect to the basis \(\mathcal {B}=\{e_n\}_{n=0}^{\infty }\). Since the matrix of \(S_{\varphi }\) is of the form (1), it follows that the matrix of \(CS_{\varphi }C\) is the followings:

$$\begin{aligned}{}[CS_{\varphi }C]_\mathcal {B}=\begin{pmatrix} \overline{a_0} &{}\quad \overline{a_1} &{}\quad \overline{a_{-1}} &{}\quad \overline{a_2} &{}\quad \overline{a_{-2}} &{}\quad \overline{a_3} &{}\quad \overline{a_{-3}} &{}\quad \cdots \\ \overline{a_1} &{}\quad \overline{a_2} &{}\quad \overline{a_{0}} &{}\quad \overline{a_{3}} &{}\quad \overline{a_{-1}} &{}\quad \overline{a_4} &{}\quad \overline{a_{-2}} &{}\quad \cdots \\ \overline{a_2} &{}\quad \overline{a_3} &{}\quad \overline{a_{1}} &{}\quad \overline{a_{4}} &{}\quad \overline{a_{0}} &{}\quad \overline{a_5} &{}\quad \overline{a_{-1}} &{}\quad \cdots \\ \overline{a_3} &{}\quad \overline{a_4} &{}\quad \overline{a_{2}} &{}\quad \overline{a_{5}} &{}\quad \overline{a_{1}} &{}\quad \overline{a_6} &{}\quad \overline{a_{0}} &{}\quad \cdots \\ \overline{a_4} &{}\quad \overline{a_5} &{}\quad \overline{a_{3}} &{}\quad \overline{a_{6}} &{}\quad \overline{a_{2}} &{}\quad \overline{a_7} &{}\quad \overline{a_{1}} &{}\quad \cdots \\ \overline{a_5} &{}\quad \overline{a_6} &{}\quad \overline{a_{4}} &{}\quad \overline{a_{7}} &{}\quad \overline{a_{3}} &{}\quad \overline{a_8} &{}\quad \overline{a_{2}} &{}\quad \cdots \\ \overline{a_6} &{}\quad \overline{a_7} &{}\quad \overline{a_{5}} &{}\quad \overline{a_{8}} &{}\quad \overline{a_{4}} &{}\quad \overline{a_9} &{}\quad \overline{a_{3}} &{}\quad \cdots \\ \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \ddots \end{pmatrix}. \end{aligned}$$

Then \([S_{\varphi }]_\mathcal {B}\) is complex symmetric with the conjugation C if and only if \(a_j=a_0\in {\mathbb C}\) for all \(j\in {\mathbb Z}\). Hence \(\varphi \) is of the form \(\varphi =\sum _{j=-\infty }^{\infty }\hat{\varphi }(0)e_j\) and so \(\varphi =0\) since \(\varphi \in L^{\infty }\).

(ii) \(\Leftrightarrow \) (iii) Let \(\varphi (z)=\sum _{j=-\infty }^{\infty }a_je_j\) be with respect to the basis \(\mathcal {B}=\{e_n\}_{n=0}^{\infty }\). It is known from [24] that \(C_{\mu ,\lambda }\) is unitarily equivalent to \(C_{1,\lambda }\). Since the matrix of \(S_{\varphi }\) is the form of (1), it follows that the matrix of \(C_{1,\lambda }S_{\varphi }C_{1,\lambda }\) is the followings:

$$\begin{aligned}{}[C_{1,\lambda }S_{\varphi }C_{1,\lambda }]_\mathcal {B}=\lambda I\begin{pmatrix} \overline{a_0} &{}\quad \overline{a_1} &{}\quad \overline{a_{-1}} &{}\quad \overline{a_2} &{}\quad \overline{a_{-2}} &{}\quad \overline{a_3} &{}\quad \overline{a_{-3}} &{}\quad \cdots \\ \overline{a_1} &{}\quad \overline{a_2} &{}\quad \overline{a_{0}} &{}\quad \overline{a_{3}} &{}\quad \overline{a_{-1}} &{}\quad \overline{a_4} &{}\quad \overline{a_{-2}} &{}\quad \cdots \\ \overline{a_2} &{}\quad \overline{a_3} &{}\quad \overline{a_{1}} &{}\quad \overline{a_{4}} &{}\quad \overline{a_{0}} &{}\quad \overline{a_5} &{}\quad \overline{a_{-1}} &{}\quad \cdots \\ \overline{a_3} &{}\quad \overline{a_4} &{}\quad \overline{a_{2}} &{}\quad \overline{a_{5}} &{}\quad \overline{a_{1}} &{}\quad \overline{a_6} &{}\quad \overline{a_{0}} &{}\quad \cdots \\ \overline{a_4} &{}\quad \overline{a_5} &{}\quad \overline{a_{3}} &{}\quad \overline{a_{6}} &{}\quad \overline{a_{2}} &{}\quad \overline{a_7} &{}\quad \overline{a_{1}} &{}\quad \cdots \\ \overline{a_5} &{}\quad \overline{a_6} &{}\quad \overline{a_{4}} &{}\quad \overline{a_{7}} &{}\quad \overline{a_{3}} &{}\quad \overline{a_8} &{}\quad \overline{a_{2}} &{}\quad \cdots \\ \overline{a_6} &{}\quad \overline{a_7} &{}\quad \overline{a_{5}} &{}\quad \overline{a_{8}} &{}\quad \overline{a_{4}} &{}\quad \overline{a_9} &{}\quad \overline{a_{3}} &{}\quad \cdots \\ \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \ddots \end{pmatrix}=\lambda [CS_{\varphi }C]_\mathcal {B}. \end{aligned}$$

Then \([S_{\varphi }]_\mathcal {B}\) is complex symmetric with the conjugation \(C_{1,\lambda }\) if and only if \(a_j=a_0\in {\mathbb C}\) for all \(j\in {\mathbb Z}\) and \(\lambda =1\). Hence, in this case, \(\varphi \) is of the form \(\varphi =\sum _{j=-\infty }^{\infty }\hat{\varphi }(0)e_j\) and so \(\varphi =0\) since \(\varphi \in L^{\infty }\). \(\square \)

Remark that if \(\varphi \in L^2\), then an unbounded H-Toeplitz operator is complex symmetric with the conjugation C if and only if \(\hat{\varphi }(j)=\hat{\varphi }(0)\in {\mathbb C}\) for all \(j\in {\mathbb Z}\). In previous theorem, if \(\hat{\varphi }(0)\not =0\), then \(\varphi =\sum _{j=-\infty }^{\infty }\hat{\varphi }(0)e_j\) does not belong to \(L^{\infty }\).

2.2 Hyponormal, quasinormal, and binormal H-Toeplitz operators

In this section, we study hyponormal, quasinormal, and binormal H-Toeplitz operators.

Lemma 2.7

For \(\varphi \in L^{\infty }\), let \(S_{\varphi }\) be an H-Toeplitz operator on \(H^2\). Then the following statements hold.

(i) \(S_{\varphi }S_{\varphi }^{*}=T_{|\varphi |^2}\), \(S_{\varphi }^{*}S_{\varphi }e_{2n}=K^{*}M_{\overline{\varphi }}T_{\varphi }e_n\), and \(S_{\varphi }^{*}S_{\varphi }e_{2n+1}=K^{*}M_{\overline{\varphi }}T_{\varphi }e_n\) hold for each \(n\in {\mathbb N}_0\).

(ii) \(S_{\varphi }^{*}\) is hyponormal if and only if the following equations hold.

$$\begin{aligned} {\left\{ \begin{array}{ll} T_{|\varphi |^2}e_{2n}\ge K^{*}M_{\overline{\varphi }}T_{\varphi }e_n\\ T_{|\varphi |^2}e_{2n+1}\ge K^{*}M_{\overline{\varphi }}H_{\varphi }e_n\end{array}\right. } \end{aligned}$$
(3)

for each \(n\in {\mathbb N}_0\). In particular, the equalities in (3) hold if and only if \(S_{\varphi }\) is normal.

Proof

(i) For \(\varphi \in L^{\infty }\), let \(S_{\varphi }\) be an H-Toeplitz operator on \(H^2\). Since \(S_{\varphi }^{*}=K^{*}M_{\overline{\varphi }}\), it follows that \(S_{\varphi }^{*}S_{\varphi }=K^{*}M_{\overline{\varphi }}PM_{\varphi }K\) and

$$\begin{aligned} S_{\varphi }S_{\varphi }^{*}=(PM_{\varphi }K)(K^{*}M_{\overline{\varphi }})=PM_{\varphi }M_{\overline{\varphi }}=PM_{|\varphi |^2}=T_{|\varphi |^2}. \end{aligned}$$

On the other hand, since \(S_{\varphi }^{*}S_{\varphi }=K^{*}M_{\overline{\varphi }}PM_{\varphi }K\), it follows that

$$\begin{aligned} S_{\varphi }^{*}S_{\varphi }e_{2n}= & {} K^{*}M_{\overline{\varphi }}PM_{\varphi }Ke_{2n}\\= & {} K^{*}M_{\overline{\varphi }}PM_{\varphi }e_{n}\\= & {} K^{*}M_{\overline{\varphi }}T_{\varphi }e_n \end{aligned}$$

and

$$\begin{aligned} S_{\varphi }^{*}S_{\varphi }e_{2n+1}= & {} K^{*}M_{\overline{\varphi }}PM_{\varphi }Ke_{2n+1}\\= & {} K^{*}M_{\overline{\varphi }}PM_{\varphi }e_{-n-1}\\= & {} K^{*}M_{\overline{\varphi }}H_{\varphi }e_n \end{aligned}$$

for each \(n\in {\mathbb N}_0\).

(ii) By (i), we obtain that \(S_{\varphi }^{*}\) is hyponormal if and only if for each n, it holds that

$$\begin{aligned} {\left\{ \begin{array}{ll} T_{|\varphi |^2}e_{2n}\ge K^{*}M_{\overline{\varphi }}T_{\varphi }e_n\\ T_{|\varphi |^2}e_{2n+1}\ge K^{*}M_{\overline{\varphi }}H_{\varphi }e_{n}.\end{array}\right. } \end{aligned}$$

In particular, we get that \(S_{\varphi }\) is normal if and only if

$$\begin{aligned} {\left\{ \begin{array}{ll} T_{|\varphi |^2}e_{2n}=K^{*}M_{\overline{\varphi }}T_{\varphi }e_n\\ T_{|\varphi |^2}e_{2n+1}=K^{*}M_{\overline{\varphi }}H_{\varphi }e_{n}\end{array}\right. } \end{aligned}$$

for each \(n\in {\mathbb N}_0\). \(\square \)

Using Lemma 2.7, we show that every H-Toeplitz operator on \(H^2\) is cohyponormal.

Theorem 2.8

For \(\varphi \in L^{\infty }\), let \(S_{\varphi }\) be an H-Toeplitz operator on \(H^2\). Then \(S_{\varphi }^{*}\) is hyponormal.

Proof. Set \(\varphi =\sum _{j=-\infty }^{\infty }\hat{\varphi }(j)e_j\in L^{\infty }\). Let \(S_{\varphi }\) be an H-Toeplitz operator on \(H^2\). Then \(S_{\varphi }^{*}\) is hyponormal if and only if

$$\begin{aligned} \Vert S_{\varphi }f\Vert ^2\le \Vert S_{\varphi }^{*}f\Vert ^2 \end{aligned}$$

for each \(f\in H^2\). Taking \(f=e_{2n}\) for each n, Lemma 2.7 implies that

$$\begin{aligned} \Vert S_{\varphi }^{*}e_{2n}\Vert ^2-\Vert S_{\varphi }e_{2n}\Vert ^2= & {} \Vert K^{*}M_{\varphi }^{*}e_{2n}\Vert ^2-\Vert PT_{\varphi }e_n\Vert ^2\\= & {} \Vert M_{\varphi }^{*}e_{2n}\Vert ^2-\Vert PT_{\varphi }e_n\Vert ^2\\= & {} \bigg \Vert \sum _{j=-\infty }^{\infty }\overline{\hat{\varphi }(j)}e_{2n-j}\bigg \Vert ^2-\bigg \Vert \sum _{j=-n}^{\infty }\hat{\varphi }(j)e_{n+j} \bigg \Vert ^2\\= & {} \sum _{j=-\infty }^{\infty }|{\hat{\varphi }(j)}|^2-\sum _{j=-n}^{\infty }|\hat{\varphi }(j)|^2\\= & {} \sum _{j=n-1}^{\infty }|\hat{\varphi }(-j)|^2\ge 0 \end{aligned}$$

since \(K^{*}\) is unitary. Put \(f(z)=e_{2n+1}\) for each n. Then Lemma 2.7 ensures that

$$\begin{aligned} \Vert S_{\varphi }^{*}e_{2n+1}\Vert ^2-\Vert S_{\varphi }e_{2n+1}\Vert ^2= & {} \Vert S_{\varphi }^{*}e_{2n+1}\Vert ^2-\Vert H_{\varphi }e_{n}\Vert ^2\\= & {} \Vert K^{*}M_{\varphi }^{*}e_{2n+1}\Vert ^2-\Vert PM_{\varphi }Je_{n}\Vert ^2\\= & {} \Vert M_{\varphi }^{*}e_{2n+1}\Vert ^2-\Vert PM_{\varphi }Je_{n}\Vert ^2\\= & {} \bigg \Vert \sum _{j=-\infty }^{\infty }\overline{\hat{\varphi }(j)}e_{2n+1-j}\bigg \Vert ^2-\bigg \Vert P(\sum _{j=-\infty }^{\infty }\hat{\varphi }(j)e_{j-n-1})\bigg \Vert ^2\\= & {} \sum _{j=-\infty }^{\infty }|{\hat{\varphi }(j)}|^2-\sum _{j=n+1}^{\infty }|\hat{\varphi }(j)|^2\\= & {} \sum _{j=-\infty }^{n}|\hat{\varphi }(j)|^2\ge 0. \end{aligned}$$

Hence we conclude that \(S_{\varphi }^{*}\) is hyponormal. \(\square \)

Theorem 2.9

Let \(\varphi \in L^{\infty }\) and \(S_{\varphi }\) be an H-Toeplitz operator on \(H^2\). Then the following statement hold.

(i) If \(\varphi \) is a nonzero constant function, then \(S_{\varphi }\) is not quasinormal, but its adjoint \(S_{\varphi }^{*}\) is quasinormal.

(ii) If \(\varphi =\lambda u\) for an inner function u and \(\lambda \in {\mathbb C}\), then \(S_{\varphi }^{*}\) is quasinormal.

Proof

(i) Let \(\varphi =\varphi _1+\overline{\varphi _2}\in L^{\infty }\) where \(\varphi _1,\varphi _2\in H^{\infty }\). Then \(S_{\varphi }\) is quasinormal if and only if, for each \(n\in {\mathbb N}_0\),

$$\begin{aligned} 0= & {} (S_{\varphi }^{*}S_{\varphi }S_{\varphi }-S_{\varphi }S_{\varphi }^{*}S_{\varphi })e_{2n}\\= & {} K^{*}M_{\overline{\varphi }}PM_{\varphi }KPM_{\varphi }Ke_{2n}-PM_{\varphi }KK^{*}M_{\overline{\varphi }}PM_{\varphi }Ke_{2n} \\= & {} K^{*}M_{\overline{\varphi }}PM_{\varphi }KPM_{\varphi }e_{n}-PM_{\varphi }M_{\overline{\varphi }}PM_{\varphi }e_{n} \\= & {} K^{*}M_{\overline{\varphi }}PM_{\varphi }KPM_{\varphi }e_{n}-PM_{|\varphi |^2}PM_{\varphi }e_{n} \end{aligned}$$
(4)

and

$$\begin{aligned} 0= & {} (S_{\varphi }^{*}S_{\varphi }S_{\varphi }-S_{\varphi }S_{\varphi }^{*}S_{\varphi })e_{2n+1}\\= & {} K^{*}M_{\overline{\varphi }}PM_{\varphi }KPM_{\varphi }Ke_{2n+1}-PM_{\varphi }KK^{*}M_{\overline{\varphi }}PM_{\varphi }Ke_{2n+1} \\= & {} K^{*}M_{\overline{\varphi }}PM_{\varphi }KPM_{\varphi }e_{-n-1}-PM_{\varphi }M_{\overline{\varphi }}PM_{\varphi }e_{-n-1} \\= & {} K^{*}M_{\overline{\varphi }}PM_{\varphi }KPM_{\varphi }e_{-n-1}-PM_{|\varphi |^2}PM_{\varphi }e_{-n-1}. \end{aligned}$$

If \(\varphi =c\) is nonzero constant and n is odd, then (4) becomes

$$\begin{aligned} K^{*}M_{\overline{\varphi }}PM_{\varphi }KPM_{\varphi }e_{n}-PM_{|\varphi |^2}PM_{\varphi }e_{n}= & {} K^{*}M_{\overline{\varphi }}PM_{\varphi }Kce_n-PM_{|c|^2}ce_n\\= & {} K^{*}M_{\overline{\varphi }}P(c^2e_{\frac{-n-1}{2}})-{|c|^2}ce_n=-{|c|^2}ce_n\\\ne & {} 0. \end{aligned}$$

Hence \(S_{\varphi }\) is not quasinormal.

On the other hand, \(S_{\varphi }^{*}\) is quasinormal if and only if \(S_{\varphi }S_{\varphi }^{*}S_{\varphi }^{*}-S_{\varphi }^{*}S_{\varphi }S_{\varphi }^{*}=0.\) Since \(S_{\varphi }S_{\varphi }^{*}=T_{|\varphi |^2}\), it follows that \(S_{\varphi }^*\) is quasinormal if and only if

$$\begin{aligned} T_{|\varphi |^2}S_{\varphi }^{*}=S_{\varphi }^{*}T_{|\varphi |^2}. \end{aligned}$$
(5)

If \(\varphi \) is a constant function, i.e. \(\varphi =c\), then

$$\begin{aligned} (T_{|\varphi |^2}S_{\varphi }^{*}-S_{\varphi }^{*}T_{|\varphi |^2})e_{2n}= & {} (T_{|c|^2}S_{c}^{*}-S_{c}^{*}T_{|c|^2})e_{2n}\\= & {} T_{|c|^2}K^{*}M_{\overline{c}}e_{2n}-K^{*}M_{\overline{c}}T_{|c|^2}e_{2n}\\= & {} P(\overline{c}|c|^2 K^{*}e_{2n})-\overline{c}|c|^2 K^{*}e_{2n}=0 \end{aligned}$$

and

$$\begin{aligned} (T_{|\varphi |^2}S_{\varphi }^{*}-S_{\varphi }^{*}T_{|\varphi |^2})e_{2n+1}= & {} (T_{|c|^2}S_{c}^{*}-S_{c}^{*}T_{|c|^2})e_{2n+1}\\= & {} T_{|c|^2}K^{*}M_{\overline{c}}e_{2n+1}-K^{*}M_{\overline{c}}T_{|c|^2}e_{2n+1}\\= & {} P(\overline{c}|c|^2 K^{*}e_{2n+1})-\overline{c}|c|^2 K^{*}e_{2n+1}=0 \end{aligned}$$

for each \(n\in {\mathbb N}_0\). Therefore, \(S_{\varphi }^{*}\) is quasinormal.

(ii) Since \(\varphi =\lambda u\) for an inner function u and \(\lambda \in {\mathbb C}\), it follows that

$$\begin{aligned} S_{\varphi }S_{\varphi }^{*}=(PM_{\varphi }K)(K^{*}M_{\overline{\varphi }})=PM_{\lambda u}M_{\overline{\lambda u}}=PM_{|\lambda |^2|u|^2}=T_{|\lambda |^2}=|\lambda |^2 I. \end{aligned}$$

Thus (5) holds. Hence \(S_{\varphi }^{*}\) is quasinormal. \(\square \)

We next consider the hyponormality and the binormality of \(S_{\varphi }\).

Proposition 2.10

For \(\varphi \in L^{\infty }\), let \(S_{\varphi }\) be an H-Toeplitz operator on \(H^2\). Then the following statements are equivalent.

(i) \(S_{\varphi }\) is normal.

(ii) \(S_{\varphi }\) is hyponormal.

(iii) \(\varphi =0\).

Proof

If \(\varphi =0\), then \(S_{\varphi }\) is normal, and hence hyponormal. If \(S_{\varphi }\) is hyponormal, the proof follows from [1]. \(\square \)

Theorem 2.11

Let \(\varphi \in L^{\infty }\)and \(S_{\varphi }\) be an H-Toeplitz operator on \(H^2\). Assume that one of the following statements hold.

(i) \(\varphi \) is a constant function.

(ii) \(\varphi =\lambda u\) for an inner function u and \(\lambda \in {\mathbb C}\).

(iii) \(\varphi =\lambda \overline{u}\) for an inner function u and \(\lambda \in {\mathbb C}\). Then \(S_{\varphi }\) is binormal.

Proof

Let \(\varphi \in L^{\infty }\). Then \(S_{\varphi }\) is binormal if and only if \(S_{\varphi }^{*}S_{\varphi }\) and \(S_{\varphi }S_{\varphi }^{*}\) commute. This is equivalent to \(S_{\varphi }^{*}S_{\varphi }\) and \(T_{|\varphi |^2}\) commute. Thus \(S_{\varphi }\) is binormal if and only if

$$\begin{aligned}{}[S_{\varphi }^{*}S_{\varphi },S_{\varphi }S_{\varphi }^{*}]=[S_{\varphi }^{*}S_{\varphi }, T_{|\varphi |^2}]=[K^{*}M_{\overline{\varphi }}PM_{\varphi }K,T_{|\varphi |^2}]=0. \end{aligned}$$
(6)

(i) If \(\varphi \) is a constant function, then (6) clearly holds.

(ii) If \(\varphi =\lambda u\) for an inner function u and \(\lambda \in {\mathbb C}\), then \(S_{\varphi }^{*}\) is quasinormal and so \(S_{\varphi }^{*}\) is binormal. Hence \(S_{\varphi }\) is binormal.

(iii) If \(\varphi =\lambda \overline{u}\) for an inner function u and \(\lambda \in {\mathbb C}\), then

$$\begin{aligned} S_{\varphi }S_{\varphi }^{*}=(PM_{\varphi }K)(K^{*}M_{\overline{\varphi }})=PM_{\lambda \overline{u}}M_{\overline{\lambda } u}=PM_{|\lambda |^2|u|^2}=T_{|\lambda |^2}=|\lambda |^2 I. \end{aligned}$$

Thus (6) clearly holds. Hence \(S_{\varphi }\) is binormal. \(\square \)

Example 2.12

If \(\varphi (z)=z^m\) for some m, then by Theorem 2.11, \(S_{z^m}\) is binormal and by Theorem 2.9, \(S_{z^m}\) is not quasinormal and \(S_{{z}^m}^{*}\) is quasinormal.

Example 2.13

Let \(\varphi (z)=\lambda \Big (\frac{z-\mu }{1-\overline{\mu }z}\Big )\) for \(\mu \in {\mathbb D}\) and \(\lambda \in {\mathbb C}\). Then \(S_{\varphi }\) is binormal from Theorem 2.11.

Corollary 2.14

For \(\varphi \in L^{\infty }\), let \(S_{\varphi }\) be an H-Toeplitz operator on \(H^2\). Assume that one of the following statements hold.

(i) \(\varphi \) is a constant function.

(ii) \(\varphi =\lambda u\) for an inner function u and \(\lambda \in {\mathbb C}\).

(iii) \(\varphi =\lambda \overline{u}\) for an inner function u and \(\lambda \in {\mathbb C}\).

Then \(S_{\varphi }^{*}\) has a nontrivial invariant subspace.

Proof

By Theorem 2.11, \(S_{\varphi }\) is binormal. Hence \(S_{\varphi }^{*}\) is binormal. Since \(S_{\varphi }^{*}\) is hyponormal by Theorem 2.8, we conclude that \(S_{\varphi }^{*}\) has a nontrivial invariant subspace from [4]. \(\square \)

3 H-Toeplitz operators on the weighted Bergman spaces

3.1 Preliminaries and auxiliary lemmas

For \(-1<\alpha <\infty \), the weighted Bergman spaces \(A_{\alpha }^2(\mathbb D)\) is the space of analytic functions in \(L^2(\mathbb D)\equiv L^2(\mathbb D, dA_{\alpha })\), where

$$\begin{aligned} dA_{\alpha }(z)=(\alpha +1)(1-|z|^2)^{\alpha }dA(z). \end{aligned}$$

The inner product on \(L^2(\mathbb D)\) is given by

$$\begin{aligned} \langle f, \ g \rangle _{\alpha }=\int _{\mathbb D}f(z)\overline{g(z)}dA_{\alpha }(z) \qquad (f, \ g \in L^2(\mathbb D, dA_{\alpha })). \end{aligned}$$

If \(\alpha =0\), then, \(A_{0}^2(\mathbb D)\) is the Bergman spaces. For \(n\in {\mathbb N}_0,\) let

$$\begin{aligned} e_n(z)=\sqrt{\frac{\Gamma (n+\alpha +2)}{\Gamma (n+1)\Gamma (\alpha +2)}}~z^n \ \ (z \in \mathbb D). \end{aligned}$$

Here, \(\Gamma (s)\) stands for the usual Gamma functions. It is easy to check that \(\{e_n\}_{n=0}^\infty \) be an orthonormal set in \(A_{\alpha }^2(\mathbb D)\) ([15]). Because the set of polynomials is dense in \(A_{\alpha }^2(\mathbb D)\), we conclude that \({e_n}\) forms an orthonormal basis for \(A_{\alpha }^2(\mathbb D)\). If \(f, g\in A_{\alpha }^2(\mathbb D)\) are functions of the form

$$\begin{aligned} f(z)=\sum _{n=0}^{\infty }a_nz^n \quad \hbox {and} \quad g(z)=\sum _{n=0}^{\infty }b_nz^n, \end{aligned}$$

then

$$\begin{aligned} \langle f, \ g \rangle _{\alpha }=\sum _{n=0}^{\infty }\frac{\Gamma (n+1)\Gamma (\alpha +2)}{\Gamma (n+\alpha +2)}a_n\overline{b}_n. \end{aligned}$$

The weighted harmonic Bergman spaces \(L^2_{\alpha }(\mathbb D)\) denote the space of all harmonic functions f on \(\mathbb D\) such that

$$\begin{aligned} |\!|f|\!|:=\Biggl ( \int _{\mathbb D}|f(z)|^2dA_{\alpha }(z)\Biggr )^{1/2}<\infty . \end{aligned}$$

The space \(L^2_{\alpha }(\mathbb D)\) is a closed subspace of \(L^2(\mathbb D)\) and therefore inherits the structure of a Hilbert space from \(L^2(\mathbb D)\). Let \(P_{ {harm}}\) denote the orthogonal projection from \(L^2(\mathbb D)\) onto \(L^2_{\alpha }(\mathbb D)\).

For \(\varphi \in L^{\infty }(\mathbb D),\) the multiplication operators \(M_{\varphi }\) on \(A_{\alpha }^2(\mathbb D)\) is defined by \(M_{\varphi }(f)=\varphi f\), and the Toeplitz operators \(T_{\varphi }\) on \(A_{\alpha }^2(\mathbb D)\) is defined by

$$\begin{aligned} T_{\varphi } (f) = P_{\alpha }({\varphi } f), \end{aligned}$$

where \(P_{\alpha }\) denotes the orthogonal projection of \(L^2 (\mathbb D)\) onto \(A_{\alpha }^2(\mathbb D)\) and \(f\in A_{\alpha }^2(\mathbb D)\). It is evident that those operators are bounded when \(\varphi \in L^{\infty }(\mathbb D)\). The Hankel operators \(H_{\varphi }\) on the \(A_{\alpha }^2(\mathbb D)\) is defined by

$$\begin{aligned} H_{\varphi } (f) = P_{\alpha }M_{\varphi }J(f), \end{aligned}$$

where the operators \(J:A_{\alpha }^2(\mathbb D) \rightarrow \overline{A_{\alpha }^2(\mathbb D)}\) is given by \(J(e_n(z))=\overline{e_{n+1}(z)}\) for all \(n\in {\mathbb N}_0\).

Now, we introduce the notion of H-Toeplitz operators on the weighted Bergman spaces and discuss their various familiar properties. First of all, we recall the well-known facts.

Lemma 3.1

[19] For any \(s, t\in {\mathbb N}_0\),

$$\begin{aligned} P_{\alpha }(\overline{z}^t z^s)={\left\{ \begin{array}{ll} \frac{\Gamma (s+1)\Gamma (s-t+\alpha +2)}{\Gamma (s+\alpha +2)\Gamma (s-t+1)} z^{s-t} \, \, \, \, \qquad &{}\text {if} \ s \ge t \\ 0 \qquad \qquad \qquad \qquad \qquad \, \, \, &{}\text {if} \ s < t. \end{array}\right. } \end{aligned}$$

In [13], the orthogonal projection from the space \(L^2(\mathbb D)\) onto the harmonic Bergman space is given. Using a similar method, the following results can be induced.

Lemma 3.2

In the weighted harmonic Bergman spaces \(L^2_{\alpha }(\mathbb D)\), for \(s, t\in {\mathbb N}_0\),

$$\begin{aligned} P_{ {harm}}(\overline{z}^t z^s)={\left\{ \begin{array}{ll} \frac{\Gamma (s-t+\alpha +2)\Gamma (s+1)}{\Gamma (s+\alpha +2)\Gamma (s-t+1)} z^{s-t} \qquad &{}\text { if} \ s \ge t \\ \frac{\Gamma (t-s+\alpha +2)\Gamma (t+1)}{\Gamma (t+\alpha +2)\Gamma (t-s+1)} \overline{z}^{t-s} \qquad \, &{}\text { if} \ s < t. \end{array}\right. } \end{aligned}$$

Proof

If \(s\ge t\), then

$$\begin{aligned} \begin{aligned} \langle P_{ {harm}}(\overline{z}^tz^s), z^k\rangle&=\langle \overline{z}^tz^s, z^k\rangle \\&={\left\{ \begin{array}{ll} \frac{\Gamma (\alpha +2)\Gamma (s+1)}{\Gamma (s+\alpha +2)} &{} \qquad \qquad \quad \text {if} \ k=s-t \\ 0\ \ {} &{}\qquad \qquad \quad \text {otherwise} \end{array}\right. }\\&={\left\{ \begin{array}{ll} \frac{\Gamma (k+\alpha +2)\Gamma (s+1)}{\Gamma (s+\alpha +2)\Gamma (k+1)}\langle z^k, z^k\rangle &{}\text {if} \ k=s-t \\ 0 &{}\text {otherwise} \end{array}\right. }\\&=\frac{\Gamma (s-t+\alpha +2)\Gamma (s+1)}{\Gamma (s+\alpha +2)\Gamma (s-t+1)}\langle z^{s-t}, z^k\rangle . \end{aligned} \end{aligned}$$

On the other hands, if \(s< t\), then

$$\begin{aligned} \langle P_{ {harm}}(\overline{z}^tz^s), \overline{z}^k\rangle= & {} \langle \overline{z}^tz^s, \overline{z}^k\rangle \\= & {} {\left\{ \begin{array}{ll} \frac{\Gamma (\alpha +2)\Gamma (t+1)}{\Gamma (t+\alpha +2)} &{}\qquad \qquad \quad \text {if} \ k=t-s \\ 0\ \ &{}\qquad \qquad \quad \text {otherwise} \end{array}\right. }\\= & {} {\left\{ \begin{array}{ll} \frac{\Gamma (k+\alpha +2)\Gamma (t+1)}{\Gamma (t+\alpha +2)\Gamma (k+1)}\langle \overline{z}^k, \overline{z}^k\rangle &{}\text {if} \ k=t-s \\ 0 &{}\text {otherwise} \end{array}\right. }\\= & {} \frac{\Gamma (t-s+\alpha +2)\Gamma (t+1)}{\Gamma (t+\alpha +2)\Gamma (t-s+1)}\langle \overline{z}^{t-s},\overline{z}^k\rangle . \end{aligned}$$

\(\square \)

Next, we find the matrix representations of Toeplitz operators \(T_{\varphi }\) and of Hankel operators \(H_{\varphi }\) with harmonic symbols \(\varphi \) on the weighted Bergman spaces. For the harmonic symbol \(\varphi (z)=\sum _{i=0}^{\infty }a_iz^i+\sum _{j=1}^{\infty }b_j\overline{z}^j\in L^{\infty }(\mathbb D)\), the \((m, n)^{th}\) entry of the matrix of \(T_{\varphi }\) with respect to orthonormal basis \({\mathcal B}=\{e_n\}_{n=0}^\infty \) of \(A_{\alpha }^2(\mathbb D)\) is given by

$$\begin{aligned} \begin{aligned} \langle T_{\varphi }e_n, e_m\rangle&= \langle P_{\alpha }({\varphi }e_n), e_m\rangle \\&=\sqrt{\frac{\Gamma (n+\alpha +2)}{\Gamma (n+1)\Gamma (\alpha +2)}}\sqrt{\frac{\Gamma (m+\alpha +2)}{\Gamma (m+1)\Gamma (\alpha +2)}}\Biggl \langle \Biggl (\sum _{i=0}^{\infty }a_iz^i+\sum _{j=1}^{\infty }b_j\overline{z}^j\Biggr )z^n,z^m\Biggr \rangle \\&=\sqrt{\frac{\Gamma (n+\alpha +2)}{\Gamma (n+1)\Gamma (\alpha +2)}}\sqrt{\frac{\Gamma (m+\alpha +2)}{\Gamma (m+1)\Gamma (\alpha +2)}} \Biggl (\sum _{i=0}^{\infty }a_i\langle z^{i+n},z^m\rangle +\sum _{j=1}^{\infty } b_j\langle {z}^n, z^{m+j}\rangle \Biggr ). \end{aligned} \end{aligned}$$

There are two cases to consider. If \(m\ge n\), then we have

$$\begin{aligned} \langle T_{\varphi }e_n, e_m\rangle= & {} \sqrt{\frac{\Gamma (n+\alpha +2)}{\Gamma (n+1)\Gamma (\alpha +2)}}\sqrt{\frac{\Gamma (m+\alpha +2)}{\Gamma (m+1)\Gamma (\alpha +2)}} \sum _{i=0}^{\infty }a_i\langle z^{i+n},z^m\rangle \\= & {} \sqrt{\frac{\Gamma (n+\alpha +2)}{\Gamma (n+1)\Gamma (\alpha +2)}}\sqrt{\frac{\Gamma (m+\alpha +2)}{\Gamma (m+1)\Gamma (\alpha +2)}} \frac{\Gamma (m+1)\Gamma (\alpha +2)}{\Gamma (m+\alpha +2)}a_{m-n}\\= & {} \sqrt{\frac{\Gamma (n+\alpha +2)\Gamma (m+1)}{\Gamma (n+1)\Gamma (m+\alpha +2)}}a_{m-n}. \end{aligned}$$

If \(m<n\), then we have

$$\begin{aligned} \begin{aligned} \langle T_{\varphi }e_n, e_m\rangle&=\sqrt{\frac{\Gamma (n+\alpha +2)}{\Gamma (n+1)\Gamma (\alpha +2)}}\sqrt{\frac{\Gamma (m+\alpha +2)}{\Gamma (m+1)\Gamma (\alpha +2)}} \sum _{j=1}^{\infty }b_j\langle z^{n},z^{m+j}\rangle \\&=\sqrt{\frac{\Gamma (n+\alpha +2)}{\Gamma (n+1)\Gamma (\alpha +2)}}\sqrt{\frac{\Gamma (m+\alpha +2)}{\Gamma (m+1)\Gamma (\alpha +2)}} \frac{\Gamma (n+1)\Gamma (\alpha +2)}{\Gamma (n+\alpha +2)}b_{n-m}\\&=\sqrt{\frac{\Gamma (m+\alpha +2)\Gamma (n+1)}{\Gamma (m+1)\Gamma (n+\alpha +2)}}b_{n-m}. \end{aligned} \end{aligned}$$

Thus, we have

$$\begin{aligned} \langle T_{\varphi }e_n, e_m\rangle ={\left\{ \begin{array}{ll}\sqrt{\frac{\Gamma (n+\alpha +2)\Gamma (m+1)}{\Gamma (n+1)\Gamma (m+\alpha +2)}}a_{m-n} \ \ \text {for} \ \ m\ge n\\ \sqrt{\frac{\Gamma (m+\alpha +2)\Gamma (n+1)}{\Gamma (m+1)\Gamma (n+\alpha +2)}}b_{n-m} \ \, \ \text {for} \ \ m<n, \end{array}\right. } \end{aligned}$$

where \(m, n\in {\mathbb N}_0\). Therefore, the matrix representation of \(T_{\varphi }\) is given by

$$\begin{aligned}{}[T_{\varphi }]_\mathcal {B}=\begin{pmatrix} a_0 &{}\quad \sqrt{\frac{1}{\alpha +2}}b_1 &{}\quad \sqrt{\frac{1\cdot 2}{(\alpha +2)(\alpha +3)}}b_2 &{}\quad \sqrt{\frac{1\cdot 2\cdot 3}{(\alpha +2)(\alpha +3)(\alpha +4)}}b_3 &{}\quad \cdots \\ \sqrt{\frac{1}{\alpha +2}}a_1 &{}\quad a_0 &{}\quad \sqrt{\frac{1\cdot 2}{\alpha +3}}b_1 &{}\quad \sqrt{\frac{1\cdot 2\cdot 3}{(\alpha +3)(\alpha +4)}}b_2 &{}\quad \cdots \\ \sqrt{\frac{1\cdot 2}{(\alpha +2)(\alpha +3)}}a_2 &{}\quad \sqrt{\frac{1\cdot 2}{\alpha +3}}a_1 &{}\quad a_0 &{}\quad \sqrt{\frac{3}{\alpha +4}}b_1 &{}\quad \cdots \\ \sqrt{\frac{1\cdot 2\cdot 3}{(\alpha +2)(\alpha +3)(\alpha +4)}}a_3 &{}\quad \sqrt{\frac{1\cdot 2\cdot 3}{(\alpha +3)(\alpha +4)}}a_2 &{}\quad \sqrt{\frac{3}{\alpha +4}}a_1 &{}\quad a_0 &{}\quad \cdots \\ \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \ddots \end{pmatrix} \end{aligned}$$

and the adjoint of the matrix representation of \(T_{\varphi }\) is given by

$$\begin{aligned}{}[T_{\varphi }^*]_\mathcal {B}=\begin{pmatrix} \overline{a_0} &{}\quad \sqrt{\frac{1}{\alpha +2}}\overline{a_1} &{}\quad \sqrt{\frac{1\cdot 2}{(\alpha +2)(\alpha +3)}}\overline{a_2} &{}\quad \sqrt{\frac{1\cdot 2\cdot 3}{(\alpha +2)(\alpha +3)(\alpha +4)}}\overline{a_3} &{}\quad \cdots \\ \sqrt{\frac{1}{\alpha +2}}\overline{b_1} &{}\quad \overline{a_0} &{}\quad \sqrt{\frac{1\cdot 2}{\alpha +3}}\overline{a_1} &{}\quad \sqrt{\frac{1\cdot 2\cdot 3}{(\alpha +3)(\alpha +4)}}\overline{a_2} &{}\quad \cdots \\ \sqrt{\frac{1\cdot 2}{(\alpha +2)(\alpha +3)}}\overline{b_2} &{}\quad \sqrt{\frac{1\cdot 2}{\alpha +3}}\overline{b_1} &{}\quad \overline{a_0} &{}\quad \sqrt{\frac{3}{\alpha +4}}\overline{a_1} &{}\quad \cdots \\ \sqrt{\frac{1\cdot 2\cdot 3}{(\alpha +2)(\alpha +3)(\alpha +4)}}\overline{b_3} &{}\quad \sqrt{\frac{1\cdot 2\cdot 3}{(\alpha +3)(\alpha +4)}}\overline{b_2} &{}\quad \sqrt{\frac{3}{\alpha +4}}\overline{b_1} &{}\quad \overline{a_0} &{}\quad \cdots \\ \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \ddots \end{pmatrix} \end{aligned}$$

and hence, we check that \(T_{\varphi }^*=T_{\overline{\varphi }}\).

Next, for the harmonic symbol \(\varphi (z)=\sum _{i=0}^{\infty }a_iz^i+\sum _{j=1}^{\infty }b_j\overline{z}^j\in L^{\infty }(\mathbb D)\), the \((m, n)^{th}\) entry of the matrix of \(H_{\varphi }\) with respect to orthonormal basis \({\mathcal B}=\{e_n\}_{n=0}^\infty \) of \(A_{\alpha }^2(\mathbb D)\) is given by

$$\begin{aligned} \begin{aligned} \langle H_{\varphi }e_n, e_m\rangle&= \langle P_{\alpha }M_{\varphi }Je_n, e_m\rangle \\&=\sqrt{\frac{\Gamma (n+\alpha +3)}{\Gamma (n+2)\Gamma (\alpha +2)}}\sqrt{\frac{\Gamma (m+\alpha +2)}{\Gamma (m+1)\Gamma (\alpha +2)}}\Biggl \langle \Biggl (\sum _{i=0}^{\infty }a_iz^i+\sum _{j=1}^{\infty }b_j\overline{z}^j\Biggr )\overline{z}^{n+1},z^m\Biggr \rangle \\&=\sqrt{\frac{\Gamma (n+\alpha +3)}{\Gamma (n+2)\Gamma (\alpha +2)}}\sqrt{\frac{\Gamma (m+\alpha +2)}{\Gamma (m+1)\Gamma (\alpha +2)}}\\&\qquad \bigg (\sum _{i=0}^{\infty }a_i\langle z^{i},z^{m+n+1}\rangle +\sum _{j=1}^{\infty } b_j\langle \overline{z}^j, z^{m+n+1}\rangle \bigg )\\&=\sqrt{\frac{\Gamma (n+\alpha +3)\Gamma (m+\alpha +2)}{\Gamma (n+2)\Gamma (m+1)}}\frac{\Gamma (m+n+2)}{\Gamma (m+n+\alpha +3)}a_{m+n+1} \end{aligned} \end{aligned}$$

for \(m, n\in {\mathbb N}_0\). Therefore, the matrix representation of \(H_{\varphi }\) is given by

$$\begin{aligned} {[}H_{\varphi }]_\mathcal {B}=\begin{pmatrix} \sqrt{\frac{1}{\alpha +2}}a_1 &{}\quad \sqrt{\frac{1\cdot 2}{(\alpha +2)(\alpha +3)}}a_2 &{}\quad \sqrt{\frac{1\cdot 2\cdot 3}{(\alpha +2)(\alpha +3)(\alpha +4)}}a_3 &{}\quad \cdots \\ {\frac{1\cdot 2}{\alpha +3}}a_2 &{}\quad \frac{3\sqrt{2}}{(\alpha +4)\sqrt{\alpha +3}}a_3 &{}\quad \frac{4\sqrt{6}}{(\alpha +5)\sqrt{(\alpha +3)(\alpha +4)}}a_4 &{}\quad \cdots \\ \frac{3\sqrt{2}}{(\alpha +4)\sqrt{\alpha +3}}a_3 &{}\quad \frac{12}{(\alpha +4)(\alpha +5)}a_4 &{}\quad \frac{20\sqrt{3}}{(\alpha +5)(\alpha +6)\sqrt{\alpha +4}}a_5 &{}\quad \cdots \\ \frac{4\sqrt{6}}{(\alpha +5)\sqrt{(\alpha +3)(\alpha +4)}}a_4 &{}\quad \frac{20\sqrt{3}}{(\alpha +5)(\alpha +6)\sqrt{\alpha +4}}a_5 &{}\quad \frac{4\cdot 5\cdot 6}{(\alpha +5)(\alpha +6)(\alpha +7)}a_6&{}\quad \cdots \\ \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \ddots \end{pmatrix}. \end{aligned}$$

Notation 3.3

For our convenience, we introduce the following notations:

$$\begin{aligned} \Lambda _{\alpha }(s)=\frac{\Gamma (s+1)\Gamma (\alpha +2)}{\Gamma (s+\alpha +2)} \quad \text {and} \quad \Lambda _{\alpha }(s,t)=\frac{\Gamma (s+1)^2\Gamma (s-t+\alpha +2)\Gamma (\alpha +2)}{\Gamma (s+\alpha +2)^2 \Gamma (s-t+1)}. \end{aligned}$$

Lemma 3.4

[19] For \(m\ge 0\), we have that

(i) \(\displaystyle {\mathinner {\Big |\!\Big |\overline{z}^m \sum _ {j=0}^{\infty } c_{j}z^{j}\Big |\!\Big |}^2 =\sum _{j=0}^{\infty } \Lambda _{\alpha }(j+m)|c_{j}|^2}\), and

(ii) \(\displaystyle {\mathinner {\Big |\!\Big |P_{\alpha }\bigg (\overline{z}^m \sum _ {j=0}^{\infty } c_{j}z^{j} \bigg )\Big |\!\Big |}^2}={\left\{ \begin{array}{ll} \displaystyle {\sum _{j=0}^{\infty } \Lambda _{\alpha }(j,m)|c_{j}|^2 \quad \text {if} \quad m \le j} \\ \displaystyle {\sum _{j=1}^{\infty } \Lambda _{\alpha }(j,m)|c_{j}|^2 \quad \text {if} \quad m > j}. \end{array}\right. }\)

Applying Lemmas 3.2 and 3.4, we obtain the following Remarks.

Remark 3.5

For \(m \ge 0\), we have

$$\begin{aligned} \displaystyle {|\!|P_{ {harm}}\bigg (\overline{z}^m \sum _ {j=0}^{\infty } c_{j}z^{j}\bigg )|\!|^2} =\displaystyle {\sum _{j=0}^{m} \Lambda _{\alpha }(m,j)|c_{j}|^2} + \sum _{j=m+1}^{\infty } \Lambda _{\alpha }(j,m)|c_{j}|^2. \end{aligned}$$

To define the notion of H-Toeplitz operators on \(A_{\alpha }^2(\mathbb D)\), we start by considering the operators \(K: A_{\alpha }^2(\mathbb D) \rightarrow L^2_{\alpha }(\mathbb D)\) defined by

$$\begin{aligned} K(e_{2n}(z)) = e_n(z) ~\text{ and } \,\,K(e_{2n+1}(z)) = \overline{e_{n+1}(z)} \end{aligned}$$
(7)

for all \(n\ge 0\) and \(z\in \mathbb D\). The operator K can be shown to be a bounded linear operator on \(A_{\alpha }^2(\mathbb D)\) with \(|\!|K|\!| = 1\). Furthermore, the adjoint operator \(K^*\) is given by

$$\begin{aligned} K^*(e_{n}(z)) = e_{2n}(z) \ \text {and} \ K^*(\overline{e_{n+1}(z)}) = e_{2n+1}(z) \quad \end{aligned}$$

for all \(n\ge 0\). From the definitions of the operators K and \(K^*\), we have that \(KK^*=I_{L^2_{\alpha }(\mathbb D)}\) and \(K^*K=I_{A_{\alpha }^2(\mathbb D)}\).

Remark 3.6

It follows from the definition of operator K, we have

$$\begin{aligned} K(z^{2n})=\frac{\sqrt{\Gamma (2n+1)\Gamma (n+\alpha +2)}}{\sqrt{\Gamma (n+1)\Gamma (2n+\alpha +2)}}z^n, K(z^{2n+1})=\frac{\sqrt{\Gamma (2n+2)\Gamma (n+\alpha +3)}}{\sqrt{\Gamma (n+2)\Gamma (2n+\alpha +3)}}\overline{z}^{n+1}, \end{aligned}$$

\(K^*(z^{n})=\frac{\sqrt{\Gamma (n+1)\Gamma (2n+\alpha +2)}}{\sqrt{\Gamma (2n+1)\Gamma (n+\alpha +2)}}z^{2n},\) and \( K^*(\overline{z}^{n+1})=\frac{\sqrt{\Gamma (n+2)\Gamma (2n+\alpha +3)}}{\sqrt{\Gamma (2n+2)\Gamma (n+\alpha +3)}}{z}^{2n+1}. \)

We next define H-Toeplitz operators on the weighted Bergman spaces \(A_{\alpha }^2(\mathbb D)\).

Definition 3.7

For \(\varphi \in L^{\infty }(\mathbb D)\), the H-Toeplitz operator \(B_{\varphi }\) on the weighted Bergman space is defined by \(B_{\varphi }(f) = P_{\alpha }M_{\varphi }K(f)\) for all \(f \in A_{\alpha }^2(\mathbb D)\) where K is defined as in (7).

We find the matrix representation of H-Toeplitz operators \(B_{\varphi }\) with harmonic symbol \(\varphi \) on the weighted Bergman spaces. If the harmonic symbol of the form \(\varphi (z)=\sum _{i=0}^{\infty }a_iz^i+\sum _{j=1}^{\infty }b_j\overline{z}^j\in L^{\infty }(\mathbb D)\), then

$$\begin{aligned} B_{\varphi }(e_{2n})=P_{\alpha }M_{\varphi }K(e_{2n})=P_{\alpha }M_{\varphi }(e_{n})=T_{\varphi }(e_n) \end{aligned}$$

and

$$\begin{aligned} B_{\varphi }(e_{2n+1})=P_{\alpha }M_{\varphi }K(e_{2n+1})=P_{\alpha }M_{\varphi }(\overline{e_{n+1}})=P_{\alpha }M_{\varphi }J(e_n)=H_{\varphi }(e_n) \end{aligned}$$

where \(\{e_n\}_{n=0}^\infty \) is an orthonormal set in \(A_{\alpha }^2(\mathbb D)\). Thus

$$\begin{aligned} \begin{aligned} \langle B_{\varphi }e_{2n}, e_m\rangle&=\langle T_{\varphi }e_{n}, e_m\rangle \\&={\left\{ \begin{array}{ll}\sqrt{\frac{\Gamma (n+\alpha +2)\Gamma (m+1)}{\Gamma (n+1)\Gamma (m+\alpha +2)}}a_{m-n} \quad \text {for} \ m\ge n\\ \sqrt{\frac{\Gamma (m+\alpha +2)\Gamma (n+1)}{\Gamma (m+1)\Gamma (n+\alpha +2)}}b_{n-m} \quad \, \text {for} \ m<n, \end{array}\right. } \end{aligned} \end{aligned}$$

and

$$\begin{aligned} \begin{aligned} \langle B_{\varphi }e_{2n+1}, e_m\rangle&=\langle H_{\varphi }e_{n}, e_m\rangle \\&=\sqrt{\frac{\Gamma (n+\alpha +3)\Gamma (m+\alpha +2)}{\Gamma (n+2)\Gamma (m+1)}}\frac{\Gamma (m+n+2)}{\Gamma (m+n+\alpha +3)}a_{m+n+1} \end{aligned} \end{aligned}$$

where \(m, n\in {\mathbb N}_0\). Thus \((m, n)^{th}\) entry of the matrix representation of \(B_{\varphi }\) with respect to orthonormal basis \(\mathcal {B}=\{e_n\}_{n=0}^\infty \) of \(A_{\alpha }^2(\mathbb D)\) is given by

$$\begin{aligned} {[}B_{\varphi }]_{\mathcal B}=\begin{pmatrix} a_0 &{}\quad \sqrt{\frac{1}{\alpha +2}}a_1 &{}\quad \sqrt{\frac{1}{\alpha +2}}b_1 &{}\quad \sqrt{\frac{1\cdot 2}{(\alpha +2)(\alpha +3)}}a_2 &{}\quad \cdots \\ \sqrt{\frac{1}{\alpha +2}}a_1 &{}\quad {\frac{1\cdot 2}{\alpha +3}}a_2 &{}\quad a_0 &{}\quad \frac{3\sqrt{2}}{(\alpha +4)\sqrt{\alpha +3}}a_3 &{}\quad \cdots \\ \sqrt{\frac{1\cdot 2}{(\alpha +2)(\alpha +3)}}a_2 &{}\quad \frac{3\sqrt{2}}{(\alpha +4)\sqrt{\alpha +3}}a_3 &{}\quad \sqrt{\frac{1\cdot 2}{\alpha +3}}a_1 &{}\quad \frac{12}{(\alpha +4)(\alpha +5)}a_4&{}\quad \cdots \\ \sqrt{\frac{1\cdot 2\cdot 3}{(\alpha +2)(\alpha +3)(\alpha +4)}}a_3 &{}\quad \frac{4\sqrt{6}}{(\alpha +5)\sqrt{(\alpha +3)(\alpha +4)}}a_4 &{}\quad \sqrt{\frac{1\cdot 2\cdot 3}{(\alpha +3)(\alpha +4)}}a_2 &{}\quad \frac{20\sqrt{3}}{(\alpha +5)(\alpha +6)\sqrt{\alpha +4}}a_5 &{}\quad \cdots \\ \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \ddots \end{pmatrix}. \end{aligned}$$

The following proposition presents some basic properties of H-Toeplitz operators on the weighted Bergman spaces (cf. [13]).

Proposition 3.8

For \(\varphi , \psi \in L^{\infty }(\mathbb D)\), the operator \(B_{\varphi }\) satisfies the following:

(i) \(B_{\varphi }\) is a bounded linear operators on \(A_{\alpha }^2(\mathbb D)\) with \(|\!|B_{\varphi }|\!|\le |\!|\varphi |\!|_{\infty }\).

(ii) For any scalars \(\alpha \) and \(\beta \), it holds \(B_{\alpha \varphi +\beta \psi }=\alpha B_{\varphi }+\beta B_{\psi }\).

(iii) The adjoint of the H-Toeplitz operators \(B_{\varphi }\) is given by \(B^*_{\varphi }=K^*P_{ {harm}}M_{\overline{\varphi }}\).

The following remark provides an important information regarding adjoint operators, showing the difference between the adjoint of Toeplitz operators and the adjoint of H-Toeplitz operators.

Remark 3.9

If fg are in \(L^{\infty }(\mathbb D)\), then, by the definition of Toeplitz operators, we have that

$$\begin{aligned} T^*_f = T_{\overline{f}} \quad \text {and} \quad T_{\overline{f}}T_g = T_{\overline{f}g} \ \text {if} \ f \ \text {or} \ g \ \text {is analytic.} \end{aligned}$$

However, for the case of the H-Toeplitz operators,

$$\begin{aligned} B^*_z(az)=K^*P_{ {harm}}M_{\overline{z}}(az)=K^*P_{ {harm}}({a\overline{z}}z)=K^*\left( \frac{\Gamma (2)\Gamma (\alpha +2)}{\Gamma (\alpha +3)}a\right) =\frac{a}{\alpha +2} \end{aligned}$$

and

$$\begin{aligned} B_{\overline{z}}(az)=P_{\alpha }M_{\overline{z}}K(az)=P_{\alpha }M_{\overline{z}}a\overline{z}=P_{\alpha }(a\overline{z}^2)=0. \end{aligned}$$

Therefore, \(B^*_z(az)\ne B_{\overline{z}}(az)\). It can be easily verified by computation that \(B_zB_z\ne B_{z^2}\).

Recall that a bounded linear operator T on a Hilbert space is called expansive if \(T^{*}T\ge I\), contractive if \(T^{*}T\le I\), and isometric if \(T^{*}T=I\), respectively. For \(k\in A_{\alpha }^2(\mathbb D)\), let \(k(z)=k_e(z)+k_o(z)\), where

$$\begin{aligned} k_e(z):= \sum _ {n=0}^{\infty } c_{2n}z^{2n} \quad \text {and} \quad k_o(z):= \sum _ {n=0}^{\infty } c_{2n+1}z^{2n+1}. \end{aligned}$$

3.2 H-Toeplitz operators with analytic symbols

In this subsection, we examine the characteristics of H-Toeplitz operators \(B_{\varphi }\) with analytic symbol functions \(\varphi \). First, we study the necessary condition for contractivity and expansivity of \(B_{\varphi }\) where \(\varphi (z)= \sum _{j=0}^{\infty }a_jz^j\) with \(a_j\in \mathbb C\) under a certain additional assumptions concerning the symbol \(\varphi \).

Theorem 3.10

Let \( \varphi (z)= \sum _{j=0}^{\infty }a_jz^j\) and \(a_j\in \mathbb C\).

(i) If \(B_{\varphi }\) is contractive, then

$$\begin{aligned} \sum _{j=0}^{\infty }|a_j|^2\le 1 \quad \text { and }\quad \sum _{j=s+1}^{\infty }\frac{\Lambda _{\alpha }(j,s+1)}{\Lambda _{\alpha }(s+1)}|a_{j}|^2 \le 1 \end{aligned}$$

for any \(s\in {\mathbb N}_0\).

(ii) If \(B_{\varphi }\) is expansive, then

$$\begin{aligned} \sum _{j=0}^{\infty }{\Lambda _{\alpha }(j)}|a_j|^2\ge 1 \quad \text { and }\quad \sum _{j=s+1}^{\infty }\frac{\Lambda _{\alpha }(j,s+1)}{\Lambda _{\alpha }(s+1)}|a_{j}|^2 \ge 1 \end{aligned}$$

for any \(s\in {\mathbb N}_0\).

Proof

For any \(k\in A_{\alpha }^2(\mathbb D)\), we have

$$\begin{aligned} \begin{aligned} B_{\varphi }k(z)&=P_{\alpha }M_{\varphi }K(k_e(z)+k_o(z))\\&=P_{\alpha }M_{\varphi }\sum _{n=0}^{\infty }\left( \frac{\sqrt{\Gamma (2n+1)\Gamma (n+\alpha +2)}}{\sqrt{\Gamma (n+1)\Gamma (2n+\alpha +2)}}c_{2n}z^n+\frac{\sqrt{\Gamma (2n+2)\Gamma (n+\alpha +3)}}{\sqrt{\Gamma (n+2)\Gamma (2n+\alpha +3)}}c_{2n+1}\overline{z}^{n+1}\right) \\&=\sum _{j=0}^{\infty }\sum _{n=0}^{\infty }\frac{\sqrt{\Gamma (2n+1)\Gamma (n+\alpha +2)}}{\sqrt{\Gamma (n+1)\Gamma (2n+\alpha +2)}}a_jc_{2n}z^{n+j}\\&\quad +\sum _{j=1}^{\infty }\sum _{n=0}^{j-1}\frac{\sqrt{\Gamma (2n+2)\Gamma (n+\alpha +3)}}{\sqrt{\Gamma (n+2)\Gamma (2n+\alpha +3)}}\cdot \frac{\Gamma (j+1)\Gamma (j-n+\alpha +1)}{\Gamma (j+\alpha +2)\Gamma (j-n)}a_jc_{2n+1}{z}^{j-n-1} \end{aligned} \nonumber \\ \end{aligned}$$
(8)

for any \(c_k\in \mathbb C \ (k=0,1,2,\ldots ).\) Then, from (8), the coefficient of \(z^m\) is

$$\begin{aligned} \begin{aligned}&\quad \sum _{n=0}^m\frac{\sqrt{\Gamma (2n+1)\Gamma (n+\alpha +2)}}{\sqrt{\Gamma (n+1)\Gamma (2n+\alpha +2)}}a_{m-n}c_{2n}\\&\quad +\sum _{n=0}^{\infty }\frac{\sqrt{\Gamma (2n+2)\Gamma (n+\alpha +3)}}{\sqrt{\Gamma (n+2)\Gamma (2n+\alpha +3)}}\cdot \frac{\Gamma (m+n+2)\Gamma (m+\alpha +2)}{\Gamma (m+n+\alpha +3)\Gamma (m+1)}a_{n+m+1}c_{2n+1}. \end{aligned} \end{aligned}$$

For a fixed \(\ell \in {\mathbb N}_0\), set \(c_{\ell }\ne 0\) and \(c_k=0\) for any \(k\ne \ell \). We consider the following two cases:

Case 1: If \(\ell =2s\) for any \(s\in {\mathbb N}_0\), then

$$\begin{aligned} B_{\varphi }k(z)=\sum _{j=0}^{\infty }\frac{\sqrt{\Gamma (2s+1)\Gamma (s+\alpha +2)}}{\sqrt{\Gamma (s+1)\Gamma (2s+\alpha +2)}}a_jc_{2s}z^{s+j}. \end{aligned}$$

If \(B_{\varphi }\) on \(A_{\alpha }^2(\mathbb D)\) is contractive, then

$$\begin{aligned} \sum _{j=0}^{\infty }\frac{{\Gamma (2s+1)\Gamma (s+\alpha +2)}}{{\Gamma (s+1)\Gamma (2s+\alpha +2)}}\Lambda _{\alpha }(s+j)|a_j|^2|c_{2s}|^2\le \Lambda _{\alpha }(2s)|c_{2s}|^2. \end{aligned}$$

Thus

$$\begin{aligned} {} \sum _{j=0}^{\infty }\Lambda _{\alpha }(s+j)|a_j|^2\le \frac{{\Gamma (s+1)\Gamma (2s+\alpha +2)}}{{\Gamma (2s+1)\Gamma (s+\alpha +2)}}\Lambda _{\alpha }(2s)=\Lambda _{\alpha }(s) \end{aligned}$$
(9)

for any \(s\in {\mathbb N}_0\). By a direct calculation, \( \frac{\Lambda _{\alpha }(s+j)}{\Lambda _{\alpha }(s)} \) is increasing for \(s\in {\mathbb N}_0\) and

$$\begin{aligned} \lim _{s\rightarrow \infty }\frac{\Lambda _{\alpha }(s+j)}{\Lambda _{\alpha }(s)}=1, \end{aligned}$$

and so (9) implies that

$$\begin{aligned} \sum _{j=0}^{\infty }|a_j|^2\le 1. \end{aligned}$$

Similarly, if \(B_{\varphi }\) on \(A_{\alpha }^2(\mathbb D)\) is expansive, then

$$\begin{aligned} {} \sum _{j=0}^{\infty }\Lambda _{\alpha }(s+j)|a_j|^2\ge \Lambda _{\alpha }(s) \end{aligned}$$
(10)

for any \(s\in {\mathbb N}_0\). By setting \(s=0\) in (10), we have the results.

Case 2: If \(\ell =2s+1\) for any \(s\in {\mathbb N}_0\), then

$$\begin{aligned} B_{\varphi }k(z)=\sum _{j=s+1}^{\infty }\frac{\sqrt{\Gamma (2s+2)\Gamma (s+\alpha +3)}}{\sqrt{\Gamma (s+2)\Gamma (2s+\alpha +3)}}\cdot \frac{\Gamma (j+1)\Gamma (j-s+\alpha +1)}{\Gamma (j+\alpha +2)\Gamma (j-s)}a_{j}c_{2s+1}z^{j-s-1}. \end{aligned}$$

If \(B_{\varphi }\) on \(A_{\alpha }^2(\mathbb D)\) is contractive, then

$$\begin{aligned} \sum _{j=s+1}^{\infty }\frac{{\Gamma (2s+2)\Gamma (s+\alpha +3)}}{{\Gamma (s+2)\Gamma (2s+\alpha +3)}}\cdot \frac{\Gamma (j+1)^2\Gamma (j-s+\alpha +1)^2}{\Gamma (j+\alpha +2)^2\Gamma (j-s)^2}\cdot \frac{\Lambda _{\alpha }(j-s-1)}{\Lambda _{\alpha }(2s+1)}|a_{j}|^2\le 1. \end{aligned}$$

Thus

$$\begin{aligned} \sum _{j=s+1}^{\infty }\frac{\Lambda _{\alpha }(j,s+1)}{\Lambda _{\alpha }(s+1)}|a_{j}|^2 \le 1 \end{aligned}$$

for any \(s\in {\mathbb N}_0\). Similarly, if \(B_{\varphi }\) on \(A_{\alpha }^2(\mathbb D)\) is expansive, then

$$\begin{aligned} \sum _{j=s+1}^{\infty }\frac{\Lambda _{\alpha }(j,s+1)}{\Lambda _{\alpha }(s+1)}|a_{j}|^2 \ge 1 \end{aligned}$$

for any \(s\in {\mathbb N}_0\). This completes the proof. \(\square \)

Example 3.11

Let \( \varphi (z)= \sum _{j=1}^{\infty }\frac{1}{j^{n/2}}z^j\) for any \(n\in \mathbb N\). Then,

$$\begin{aligned} \sum _{j=1}^{\infty }\frac{1}{j^n}=\zeta (n)>1, \end{aligned}$$

where \(\zeta (n)\) is the Riemann-zeta function for \(n\in \mathbb N\). Thus \(B_{\varphi }\) is not contractive from Theorem 3.10.

Example 3.12

Let \( \varphi (z)= \sum _{j=0}^{\infty }c^jz^j\) with any \(|c|< 1\). Then

$$\begin{aligned} \sum _{j=0}^{\infty }{|c|^{2j}}=\frac{1}{1-|c|^2}>1. \end{aligned}$$

Hence \(B_{\varphi }\) is not contractive from Theorem 3.10.

We give a description on the contractivity and the expansivity of H-Toeplitz operators in terms of the coefficients for the polynomial symbol \(\varphi \) of degree n on the Bergman spaces \(A_{0}^2(\mathbb D)\).

Corollary 3.13

Let \( \varphi (z)= \sum _{j=1}^{n}a_jz^j\) with any \(a_j\in \mathbb C\) and \(n\ge 1\). If \(B_{\varphi }\) is contractive on \(A_{0}^2(\mathbb D)\), then \( \sum _{j=1}^{n}|a_j|^2\le 1. \)

Proof

From the case 1 in the proof of Theorem 3.10, if \(B_{\varphi }\) is contractive, then

$$\begin{aligned} \sum _{j=1}^{n}|a_j|^2\le 1. \end{aligned}$$
(11)

Since \(\frac{\Lambda _0(j,s+1)}{\Lambda _0(s+1)}\) is increasing for \(j\le 2s\) and decreasing for \(j\ge 2s+1\), we have

$$\begin{aligned} \max _{j\ge s+1}\frac{\Lambda _0(j,s+1)}{\Lambda _0(s+1)}=\max \Bigl \{\frac{\Lambda _0(2s,s+1)}{\Lambda _0(s+1)}, \frac{\Lambda _0(2s+1,s+1)}{\Lambda _0(s+1)}\Bigr \}=\frac{s+2}{4(s+1)} \end{aligned}$$

for any \(s\in {\mathbb N}_0\), and

$$\begin{aligned} \max _{s\ge 0}\Bigl \{\frac{s+2}{4(s+1)}\Bigl \}=\frac{1}{2}. \end{aligned}$$

Thus, for any \(s\in {\mathbb N}_0\), the inequality given by

$$\begin{aligned} \sum _{j=s+1}^{n}\frac{\Lambda _0(j,s+1)}{\Lambda _0(s+1)}|a_{j}|^2 \le 1 \end{aligned}$$

implies that

$$\begin{aligned} \sum _{j=1}^{n}\frac{1}{2}|a_{j}|^2 \le 1. \end{aligned}$$
(12)

From (11) and (12), we have complete the proof. \(\square \)

Next, we consider the necessary and sufficient condition for the contractivity and the expansivity of \(B_{\varphi }\) with \(\varphi (z)= az^N\) for \(N\in \mathbb N\) and \(a\in \mathbb C\).

Theorem 3.14

For \( \varphi (z)= az^N\) with \(N\in \mathbb N\) and \(a\in \mathbb C\), \(B_{\varphi }\) is contractive if and only if \(|a|\le 1.\)

Proof

From the proof of Theorem 3.10, for any \(k\in A_{\alpha }^2(\mathbb D)\), we get that

$$\begin{aligned} \begin{aligned} B_{\varphi }k(z)&=P_{\alpha }M_{\varphi }K(k(z))\\&=\sum _{n=0}^{\infty }\frac{\sqrt{\Gamma (2n+1)\Gamma (n+\alpha +2)}}{\sqrt{\Gamma (n+1)\Gamma (2n+\alpha +2)}}ac_{2n}z^{n+N}\\&\quad +\sum _{n=0}^{N-1}\frac{\sqrt{\Gamma (2n+2)\Gamma (n+\alpha +3)}}{\sqrt{\Gamma (n+2)\Gamma (2n+\alpha +3)}}\cdot \frac{\Gamma (N+1)\Gamma (N-n+\alpha +1)}{\Gamma (N+\alpha +2)\Gamma (N-n)}ac_{2n+1}{z}^{N-n-1} \end{aligned} \end{aligned}$$

and

$$\begin{aligned} \begin{aligned} |\!|B_{\varphi }k(z)|\!|^2&=|a|^2\sum _{n=0}^{\infty }\frac{{\Gamma (2n+1)\Gamma (n+\alpha +2)}}{{\Gamma (n+1)\Gamma (2n+\alpha +2)}}\Lambda _{\alpha }(n+N)|c_{2n}|^2\\&\quad +|a|^2\sum _{n=0}^{N-1}\frac{{\Gamma (2n+2)\Gamma (n+\alpha +3)}}{{\Gamma (n+2)\Gamma (2n+\alpha +3)}}\Lambda _{\alpha }(N,n+1)|c_{2n+1}|^2. \end{aligned} \end{aligned}$$

Thus the contractivity of \(B_{\varphi }\) on \(A_{\alpha }^2(\mathbb D)\) is equivalent to

$$\begin{aligned} \begin{aligned}&|a|^2\sum _{n=0}^{\infty }\frac{{\Gamma (2n+1)\Gamma (n+\alpha +2)}}{{\Gamma (n+1)\Gamma (2n+\alpha +2)}}\Lambda _{\alpha }(n+N)|c_{2n}|^2\\&\quad +|a|^2\sum _{n=0}^{N-1}\frac{{\Gamma (2n+2)\Gamma (n+\alpha +3)}}{{\Gamma (n+2)\Gamma (2n+\alpha +3)}}\Lambda _{\alpha }(N,n+1)|c_{2n+1}|^2\\&\quad \le \sum _{j=0}^{\infty }\Lambda _{\alpha }(j)|c_j|^2. \end{aligned} \nonumber \\ \end{aligned}$$
(13)

There are two possibilities to consider. The first case is when \(c_{\ell }\ne 0\) for \(\ell \) is even, and \(c_{\ell }=0\) for \(\ell \) is odd, then by (13), we have

$$\begin{aligned} |a|^2\frac{{\Gamma (2n+1)\Gamma (n+\alpha +2)}}{{\Gamma (n+1)\Gamma (2n+\alpha +2)}}\Lambda _{\alpha }(n+N)|c_{2n}|^2\le \Lambda _{\alpha }(2n)|c_{2n}|^2, \end{aligned}$$

or equivalently,

$$\begin{aligned} |a|^2\le \frac{\Lambda _{\alpha }(n)}{\Lambda _{\alpha }(n+N)} \end{aligned}$$

for any \(n\in {\mathbb N}_0\). By a direct calculation, \(\frac{\Lambda _{\alpha }(n)}{\Lambda _{\alpha }(n+N)}\) is decreasing for n, and

$$\begin{aligned} |a|^2\le \min _{n\ge 0}\frac{\Lambda _{\alpha }(n)}{\Lambda _{\alpha }(n+N)}=\lim _{n\rightarrow \infty }\frac{\Lambda _{\alpha }(n)}{\Lambda _{\alpha }(n+N)}=1. \end{aligned}$$
(14)

The second case is when \(c_{\ell }\ne 0\) for \(\ell \) is odd, and \(c_{\ell }=0\) for \(\ell \) is even, then from (13), we have

$$\begin{aligned} |a|^2\frac{{\Gamma (2n+2)\Gamma (n+\alpha +3)}}{{\Gamma (n+2)\Gamma (2n+\alpha +3)}}\Lambda _{\alpha }(N,n+1)|c_{2n+1}|^2 \le \Lambda _{\alpha }(2n+1)|c_{2n+1}|^2, \end{aligned}$$

or equivalently,

$$\begin{aligned} |a|^2\le \frac{\Lambda _{\alpha }(n+1)}{\Lambda _{\alpha }(N,n+1)} \end{aligned}$$

for any \(0\le n\le N-1\). By a simple calculation,

$$\begin{aligned} \frac{\Lambda _{\alpha }(n+1)}{\Lambda _{\alpha }(N,n+1)}=\frac{\Gamma (N+\alpha +2)^2}{\Gamma (N+1)^2}\cdot \frac{\Gamma (N-n)}{\Gamma (N-n+\alpha +1)}\cdot \frac{\Gamma (n+2)}{\Gamma (n+\alpha +3)}>1, \nonumber \\ \end{aligned}$$
(15)

since \((N+j+1)^2>(N-n+j)(n+j+2)\) for any \(j\in \mathbb R\) and for all \(0\le n\le N-1\). From (14) and (15), \(B_{\varphi }\) is contractive if and only if \(|a|\le 1\). This completes the proof. \(\square \)

Corollary 3.15

If \( \varphi (z)= az^N\) with \(N\in \mathbb N\) and \(a\in \mathbb C\), then \(B_{\varphi }\) is a neither expansive nor isometric operator.

Proof

It follows from the proof of Theorem 3.14 that if \(B_{\varphi }\) is expansive, then

$$\begin{aligned} \begin{aligned}&|a|^2\sum _{n=0}^{\infty }\frac{{\Gamma (2n+1)\Gamma (n+\alpha +2)}}{{\Gamma (n+1)\Gamma (2n+\alpha +2)}}\Lambda _{\alpha }(n+N)|c_{2n}|^2\\&\qquad +|a|^2\sum _{n=0}^{N-1}\frac{{\Gamma (2n+2)\Gamma (n+\alpha +3)}}{{\Gamma (n+2)\Gamma (2n+\alpha +3)}}\Lambda _{\alpha }(N,n+1)|c_{2n+1}|^2\\&\quad \ge \sum _{j=0}^{\infty }\Lambda _{\alpha }(j)|c_j|^2. \end{aligned} \end{aligned}$$
(16)

If we substitute \(c_j=0\) for \(j\ne 2N+1\) in (16), then we obtain that

$$\begin{aligned} \Lambda _{\alpha }(2N+1)|c_{2N+1}|^2\le 0, \end{aligned}$$

which is a contradiction. \(\square \)

Corollary 3.16

Let \( \varphi (z)= az^N\) with \(N\in \mathbb N\) and \(a\in \mathbb C\). Then \(B_{\varphi }\) is not self-adjoint.

Proof

By the definition of the adjoint of \(B_{\varphi }\), we deduce that

$$\begin{aligned} \begin{aligned} B^*_{\varphi }k(z)&=K^*P_{ {harm}}M_{\overline{\varphi }}k(z)\\&=\overline{a}K^*\Biggl ( \sum _ {n=0}^{N-1} \frac{{\Gamma (N-n+\alpha +2)\Gamma (N+1)}}{{\Gamma (N-n+1)\Gamma (N+\alpha +2)}}c_{n}\overline{z}^{N-n}\\&\quad + \sum _ {n=N}^{\infty } \frac{{\Gamma (n-N+\alpha +2)\Gamma (n+1)}}{{\Gamma (n-N+1)\Gamma (n+\alpha +2)}}c_{n}{z}^{n-N}\Biggr )\\&= \overline{a}\sum _ {n=0}^{N-1} \frac{\Gamma (N+1)\sqrt{\Gamma (N-n+\alpha +2)\Gamma (2N-2n+\alpha +1)}}{\Gamma (N+\alpha +2)\sqrt{\Gamma (N-n+1)\Gamma (2N-2n)}}c_{n}{z}^{2N-2n-1}\\&\quad + \overline{a}\sum _ {n=N}^{\infty } \frac{\Gamma (n+1)\sqrt{\Gamma (n-N+\alpha +2)\Gamma (2n-2N+\alpha +2)}}{\Gamma (n+\alpha +2)\sqrt{\Gamma (n-N+1)\Gamma (2n-2N+1)}}c_{n}{z}^{2n-2N}. \end{aligned} \end{aligned}$$

Comparing constant terms in \(B_{\varphi }k(z)\) and \(B^*_{\varphi }k(z)\), they are

$$\begin{aligned} \frac{\sqrt{\Gamma (2N)\Gamma (N+1)}\Gamma (\alpha +2)}{\sqrt{\Gamma (2N+\alpha +1)\Gamma (N+\alpha +2)}}ac_{2N-1} \quad \text {and} \quad \frac{{\Gamma (N+1)\Gamma (\alpha +2)}}{{\Gamma (N+\alpha +2)}}\overline{a}c_{N}, \end{aligned}$$

respectively. As \(c_{2N-1}\) and \(c_{N}\) can be chosen arbitrarily, it follows that the constant terms in \(B_{\varphi }k(z)\) and \(B^*_{\varphi }k(z)\) are different, and hence \(B_{\varphi }\) is not self-adjoint. \(\square \)

Corollary 3.17

For \( \varphi (z)= az^N\) with \(N\in \mathbb N\) and \(a\in \mathbb C\), \(B_{\varphi }\) is not normal.

Proof

For any \(k \in A_{\alpha }^2(\mathbb D)\), the normality of \(B_{\varphi }\) is equivalent to \(B^*_{\varphi }B_{\varphi }k(z)=B_{\varphi }B_{\varphi }^*k(z)\) or \(|\!|B_{\varphi }k(z)|\!|^2=|\!|B^*_{\varphi }k(z)|\!|^2\). Using the proof of Theorem 3.14 and Corollary 3.16, we get

$$\begin{aligned} \begin{aligned} |\!|B_{\varphi }k(z)|\!|^2&=|a|^2\sum _{n=0}^{\infty }\frac{{\Gamma (2n+1)\Gamma (n+\alpha +2)}}{{\Gamma (n+1)\Gamma (2n+\alpha +2)}}\Lambda _{\alpha }(n+N)|c_{2n}|^2\\&\quad +|a|^2\sum _{n=0}^{N-1}\frac{{\Gamma (2n+2)\Gamma (n+\alpha +3)}}{{\Gamma (n+2)\Gamma (2n+\alpha +3)}}\Lambda _{\alpha }(N,n+1)|c_{2n+1}|^2 \end{aligned} \end{aligned}$$
(17)

and

$$\begin{aligned} |\!|B^*_{\varphi }k(z)|\!|^2=|a|^2 \sum _ {n=0}^{N-1} \frac{\Lambda _{\alpha }^2(N)}{\Lambda _{\alpha }(N-n)}|c_{n}|^2+ \sum _ {n=N}^{\infty } \frac{\Lambda _{\alpha }^2(n)}{\Lambda _{\alpha }(n-N)}|c_{n}|^2. \end{aligned}$$
(18)

If we substitute \(c_i=0\) for \(i\ne 2N+1\) in (17) and (18), then we obtain that \(|\!|B_{\varphi }k(z)|\!|^2=0\) and \(|\!|B^*_{\varphi }k(z)|\!|^2= \frac{\Lambda _{\alpha }^2(2N+1)}{\Lambda _{\alpha }(N+1)}|c_{2N+1}|^2\ne 0, \) which gives the results. \(\square \)

3.3 H-Toeplitz operators with coanalytic symbols

In this subsection, we examine the characteristics of H-Toeplitz operators \(B_{\varphi }\) with coanalytic symbol \(\varphi \). First, we examine the contractivity and the expansivity of \(B_{\varphi }\) where \(\varphi \) is of the form \(\varphi (z)= \sum _{j=1}^{\infty }b_j\overline{z}^j\) with \(b_j\in \mathbb C\).

Theorem 3.18

Let \( \varphi (z)= \sum _{j=1}^{\infty }b_j\overline{z}^j\) with \(b_j\in \mathbb C\). If \(B_{\varphi }\) is contractive, then

$$\begin{aligned} \sum _{j=1}^{s}\frac{1}{\Lambda _{\alpha }(s-j)}|b_j|^2\le \frac{1}{\Lambda _{\alpha }(s)} \end{aligned}$$

for any \(s\in \mathbb N\).

Proof

For any \(k\in A_{\alpha }^2(\mathbb D)\),

$$\begin{aligned} \begin{aligned} B_{\varphi }k(z)&=P_{\alpha }M_{\varphi }\left[ \sum _{n=0}^{\infty }\left( \frac{\sqrt{\Gamma (2n+1)\Gamma (n+\alpha +2)}}{\sqrt{\Gamma (n+1)\Gamma (2n+\alpha +2)}}c_{2n}z^n+\frac{\sqrt{\Gamma (2n+2)\Gamma (n+\alpha +3)}}{\sqrt{\Gamma (n+2)\Gamma (2n+\alpha +3)}}c_{2n+1}\overline{z}^{n+1}\right) \right] \\&=P_{\alpha }\left( \sum _{j=1}^{\infty }\sum _{n=0}^{\infty }\frac{\sqrt{\Gamma (2n+1)\Gamma (n+\alpha +2)}}{\sqrt{\Gamma (n+1)\Gamma (2n+\alpha +2)}}b_jc_{2n}z^n\overline{z}^j\right) \\&=\sum _{n=1}^{\infty }\sum _{j=1}^{n}\frac{\sqrt{\Gamma (2n+1)\Gamma (n+1)}}{\sqrt{\Gamma (n+\alpha +2)\Gamma (2n+\alpha +2)}}\cdot \frac{{\Gamma (n-j+\alpha +2)}}{{\Gamma (n-j+1)}}b_jc_{2n}z^{n-j}. \end{aligned} \nonumber \\ \end{aligned}$$
(19)

It follows from (19) that, the coefficient of \(z^m\) is

$$\begin{aligned} \sum _{n=m+1}^{\infty }\frac{\sqrt{\Gamma (2n+1)\Gamma (n+1)}}{\sqrt{\Gamma (n+\alpha +2)\Gamma (2n+\alpha +2)}}\cdot \frac{{\Gamma (m+\alpha +2)}}{{\Gamma (m+1)}}b_{n-m}c_{2n}. \end{aligned}$$

For some \(s\in \mathbb N\), we set \(c_{\ell }\ne 0\) if \(\ell =2s\) and \(c_{\ell }=0\) if \(\ell \ne 2s\). If \(B_{\varphi }\) on \(A_{\alpha }^2(\mathbb D)\) is contractive, then

$$\begin{aligned} \sum _{j=1}^{s}\frac{\Lambda _{\alpha }(2s)\Lambda _{\alpha }(s)}{\Lambda _{\alpha }(s-j)}|b_j|^2|c_{2s}|^2\le \Lambda _{\alpha }(2s)|c_{2s}|^2. \end{aligned}$$

Therefore,

$$\begin{aligned} \sum _{j=1}^{s}\frac{1}{\Lambda _{\alpha }(s-j)}|b_j|^2\le \frac{1}{\Lambda _{\alpha }(s)}. \end{aligned}$$

This completes the proof. \(\square \)

Corollary 3.19

For \( \varphi (z)= \sum _{j=1}^{\infty }b_j\overline{z}^i\) with \(b_j\in \mathbb C\), \(B_{\varphi }\) is not an expansive operator.

Proof

From the Eq. (19), if we substitute \(c_j=0\) for j is even, then we obtain that \(B_{\varphi }k(z)=0.\) Thus \(B_{\varphi }\) on \(A_{\alpha }^2(\mathbb D)\) is not an expansive operator. \(\square \)

Example 3.20

For \( \varphi (z)= \sqrt{\alpha +3}\overline{z}+ \sqrt{\alpha +2}\overline{z}^2\), we have

$$\begin{aligned} \sum _{j=1}^{2}\frac{1}{\Lambda _{\alpha }(2-j)}|b_j|^2=(\alpha +2)(\alpha +4) > \frac{(\alpha +2)(\alpha +3)}{2}=\frac{1}{\Lambda _{\alpha }(2)}. \end{aligned}$$

Hence by Theorem 3.18, \(B_{\varphi }\) is not contractive.

Next, we study the necessary and sufficient condition for the contractivity and the expansivity of \(B_{\varphi }\) with \(\varphi =b\overline{z}^N\) for \(N\in \mathbb N\) and \(b\in \mathbb C\).

Theorem 3.21

Let \( \varphi (z)= b\overline{z}^N\) with \(N\in \mathbb N\) and \(b\in \mathbb C\). Then \(B_{\varphi }\) is contractive if and only if \(|b|\le 1.\)

Proof

From the proof of Theorem 3.18, for any \(k\in A_{\alpha }^2(\mathbb D)\),

$$\begin{aligned} B_{\varphi }k(z)=\sum _{n=N}^{\infty }\frac{\sqrt{\Gamma (2n+1)\Gamma (n+1)}}{\sqrt{\Gamma (n+\alpha +2)\Gamma (2n+\alpha +2)}}\cdot \frac{{\Gamma (n-N+\alpha +2)}}{{\Gamma (n-N+1)}}bc_{2n}z^{n-N}. \end{aligned}$$

Thus

$$\begin{aligned} |\!|B_{\varphi }k(z)|\!|^2=|b|^2\sum _{n=N}^{\infty }\frac{\Lambda _{\alpha }(2n)\Lambda _{\alpha }(n)}{\Lambda _{\alpha }(n-N)}|c_{2n}|^2. \end{aligned}$$

Hence the contractivity of \(B_{\varphi }\) is equivalent to

$$\begin{aligned} |b|^2\sum _{n=N}^{\infty }\frac{\Lambda _{\alpha }(2n)\Lambda _{\alpha }(n)}{\Lambda _{\alpha }(n-N)}|c_{2n}|^2\le \sum _{n=0}^{\infty }\Lambda _{\alpha }(n)|c_n|^2. \end{aligned}$$

If we compare the terms involving \(|c_{2n}|^2\), then we have

$$\begin{aligned} \frac{\Lambda _{\alpha }(2n)\Lambda _{\alpha }(n)}{\Lambda _{\alpha }(n-N)}|b|^2|c_{2n}|^2\le \Lambda _{\alpha }(2n)|c_{2n}|^2, \end{aligned}$$

and so

$$\begin{aligned} |b|^2\le \frac{\Lambda _{\alpha }(n-N)}{\Lambda _{\alpha }(n)} \end{aligned}$$

for any \(n\ge N\). Since \(\frac{\Lambda _{\alpha }(n-N)}{\Lambda _{\alpha }(n)}\) is decreasing for \(n\ge N\), \(B_{\varphi }\) is contractive if and only if

$$\begin{aligned} |b|^2\le \min _{n\ge 0}\frac{\Lambda _{\alpha }(n-N)}{\Lambda _{\alpha }(n)}=\lim _{n\rightarrow \infty }\frac{\Lambda _{\alpha }(n-N)}{\Lambda _{\alpha }(n)}=1. \end{aligned}$$

This completes the proof. \(\square \)

Corollary 3.22

For \( \varphi (z)= b\overline{z}^N\) with \(N\in \mathbb N\) and \(b\in \mathbb C\), \(B_{\varphi }\) is neither expansive nor isometric.

Proof

Using the result as in the proof of Theorem 3.21, the expansivity of \(B_{\varphi }\) is equivalent to

$$\begin{aligned} |b|^2\sum _{n=N}^{\infty }\frac{\Lambda _{\alpha }(2n)\Lambda _{\alpha }(n)}{\Lambda _{\alpha }(n-N)}|c_{2n}|^2\ge \sum _{n=0}^{\infty }\Lambda _{\alpha }(n)|c_n|^2. \end{aligned}$$

If we substitute \(c_n= 0\) for \(n\ge 2N\), then we deduce that \( \sum _{n=0}^{2N-1}\Lambda _{\alpha }(n)|c_n|^2\le 0, \) which is a contradiction. \(\square \)

3.4 H-Toeplitz operators with harmonic symbols

Finally, we analyze the properties of H-Toeplitz operators \(B_{\varphi }\) that have harmonic symbols of the form \(\varphi (z)= \sum _{j=0}^{\infty }a_j{z}^j+\sum _{j=1}^{\infty }b_j{\overline{z}}^j\) with \(a_j, b_j \in \mathbb C\). Our focus is on determining the necessary and sufficient conditions for the contractivity and the expansivity of \(B_{\varphi }\).

Theorem 3.23

Let \( \varphi (z)= \sum _{j=0}^{\infty }a_j{z}^j+\sum _{j=1}^{\infty }b_j{\overline{z}}^j\) and \(a_j, b_j \in \mathbb C\).

(i) If \(B_{\varphi }\) is contractive, then

$$\begin{aligned} \sum _{j=0}^{\infty }\Lambda _{\alpha }(j)|a_{j}|^2\le 1, \quad \sum _{j=0}^{\infty }\frac{\Lambda _{\alpha }(s+j)}{\Lambda _{\alpha }(s)}|a_{j}|^2+\sum _{j=1}^{s}\frac{\Lambda _{\alpha }(s)}{\Lambda _{\alpha }(s-j)}|b_{j}|^2 \le 1 \end{aligned}$$

and

$$\begin{aligned} \sum _{j=s}^{\infty }\Lambda _{\alpha }(j,s)|a_{j}|^2 \le \Lambda _{\alpha }(s) \end{aligned}$$

for any \(s\in \mathbb N\).

(ii) If \(B_{\varphi }\) is expansive, then

$$\begin{aligned} \sum _{j=0}^{\infty }\Lambda _{\alpha }(j)|a_{j}|^2\ge 1, \quad \sum _{j=0}^{\infty }\frac{\Lambda _{\alpha }(s+j)}{\Lambda _{\alpha }(s)}|a_{j}|^2+\sum _{j=1}^{s}\frac{\Lambda _{\alpha }(s)}{\Lambda _{\alpha }(s-j)}|b_{j}|^2 \ge 1 \end{aligned}$$

and

$$\begin{aligned} \sum _{j=s}^{\infty }\Lambda _{\alpha }(j,s)|a_{j}|^2 \ge \Lambda _{\alpha }(s) \end{aligned}$$

for any \(s\in \mathbb N\).

Proof

By the similar arguments as in the proof of Theorems 3.10 and 3.18, for any \(k\in A_{\alpha }^2(\mathbb D)\),

$$\begin{aligned} \begin{aligned} B_{\varphi }k(z)&=\sum _{j=0}^{\infty }\sum _{n=0}^{\infty }\frac{\sqrt{\Gamma (2n+1)\Gamma (n+\alpha +2)}}{\sqrt{\Gamma (n+1)\Gamma (2n+\alpha +2)}}a_jc_{2n}z^{n+j}\\&\quad +\sum _{j=1}^{\infty }\sum _{n=0}^{j-1}\frac{\sqrt{\Gamma (2n+2)\Gamma (n+\alpha +3)}}{\sqrt{\Gamma (n+2)\Gamma (2n+\alpha +3)}}\cdot \frac{\Gamma (j+1)\Gamma (j-n+\alpha +1)}{\Gamma (j+\alpha +2)\Gamma (j-n)}a_jc_{2n+1}{z}^{j-n-1}\\&\quad +\sum _{n=1}^{\infty }\sum _{j=1}^{n}\frac{\sqrt{\Gamma (2n+1)\Gamma (n+1)}}{\sqrt{\Gamma (n+\alpha +2)\Gamma (2n+\alpha +2)}}\cdot \frac{{\Gamma (n-j+\alpha +2)}}{{\Gamma (n-j+1)}}b_jc_{2n}z^{n-j} \end{aligned} \end{aligned}$$

for any \(c_j\in \mathbb C \ (j=0,1,2,\ldots ).\) For some \(\ell \in {\mathbb N}_0\), set \(c_{\ell }\ne 0\) and \(c_j=0\) for any \(j\ne \ell \). Next, we examine the two cases below:

Case 1:  If \(\ell =0\), then \(B_{\varphi }k(z)=\sum _{j=0}^{\infty }a_jc_{0}z^{j}.\) Thus if \(B_{\varphi }\) on \(A_{\alpha }^2(\mathbb D)\) is contractive, then

$$\begin{aligned} \sum _{j=0}^{\infty }\Lambda _{\alpha }(j)|a_j|^2|c_{0}|^2 \le \Lambda _{\alpha }(0)|c_0|^2, \end{aligned}$$

or equivalently \(\sum _{j=0}^{\infty }\Lambda _{\alpha }(j)|a_j|^2\le 1\). Similarly, if \(B_{\varphi }\) on \(A_{\alpha }^2(\mathbb D)\) is expansive, then \(\sum _{j=0}^{\infty }\Lambda _{\alpha }(j)|a_j|^2\ge 1\).

Case 2:  If \(\ell =2s\) for any \(s\in \mathbb N\) and \(c_{2\,s}\ne 0\), then

$$\begin{aligned} B_{\varphi }k(z)=&\sum _{j=0}^{\infty }\frac{\sqrt{\Gamma (2s+1)\Gamma (s+\alpha +2)}}{\sqrt{\Gamma (s+1)\Gamma (2s+\alpha +2)}}a_jc_{2s}z^{s+j}\\&\quad +\sum _{j=1}^{s}\frac{\sqrt{\Gamma (2s+1)\Gamma (s+1)}}{\sqrt{\Gamma (s+\alpha +2)\Gamma (2s+\alpha +2)}}\cdot \frac{{\Gamma (s-j+\alpha +2)}}{{\Gamma (s-j+1)}}b_jc_{2s}z^{s-j}. \end{aligned}$$

If \(B_{\varphi }\) on \(A_{\alpha }^2(\mathbb D)\) is contractive, then

$$\begin{aligned}{} & {} \sum _{j=0}^{\infty }\frac{{\Gamma (2s+1)\Gamma (s+\alpha +2)}}{{\Gamma (s+1)\Gamma (2s+\alpha +2)}}\Lambda _{\alpha }(s+j)|a_j|^2|c_{2s}|^2\\{} & {} \quad +\sum _{j=1}^{s}\frac{\Lambda _{\alpha }(2s)\Lambda _{\alpha }(s)}{\Lambda _{\alpha }(s-j)}|b_j|^2|c_{2s}|^2\le \Lambda _{\alpha }(2s)|c_{2s}|^2 \end{aligned}$$

or equivalently

$$\begin{aligned} \sum _{j=0}^{\infty }\frac{\Lambda _{\alpha }(s+j)}{\Lambda _{\alpha }(s)}|a_{j}|^2+\sum _{j=1}^{s}\frac{\Lambda _{\alpha }(s)}{\Lambda _{\alpha }(s-j)}|b_{j}|^2 \le 1. \end{aligned}$$

Similarly, if \(B_{\varphi }\) on \(A_{\alpha }^2(\mathbb D)\) is expansive, then

$$\begin{aligned} \sum _{j=0}^{\infty }\frac{\Lambda _{\alpha }(s+j)}{\Lambda _{\alpha }(s)}|a_{j}|^2+\sum _{j=1}^{s}\frac{\Lambda _{\alpha }(s)}{\Lambda _{\alpha }(s-j)}|b_{j}|^2 \ge 1. \end{aligned}$$

Case 3:  If \(\ell =2s-1\) for any \(s\in \mathbb N\) and \(c_{2s-1}\ne 0\), then by the case 2 of Theorem 3.10, we have the results. This completes the proof. \(\square \)

The following results can be easily derived from Theorem 3.23.

Corollary 3.24

Let \( \varphi (z)= a_1z+b_1\overline{z}\) and \(a_1, b_1 \in \mathbb C\). If \(B_{\varphi }\) is contractive, then

$$\begin{aligned} |a_1|^2 \le \alpha +2, \quad \frac{2}{\alpha +3}|a_1|^2+\frac{1}{\alpha +2}|b_1|^2 \le 1 \end{aligned}$$

and

$$\begin{aligned} \frac{s+1}{s+\alpha +2}|a_1|^2+\frac{s}{s+\alpha +1}|b_1|^2 \le 1 \end{aligned}$$

for any \(s\ge 2\).

Example 3.25

For \( \varphi (z)= \frac{\sqrt{\alpha +2}}{\sqrt{2}}z+\frac{\sqrt{(\alpha +3)(3\alpha +7)}}{4}z^2-\frac{3\sqrt{(\alpha +1)(\alpha +2)}}{2\sqrt{2(\alpha +3)}}\overline{z}\), we have

$$\begin{aligned} \sum _{j=0}^{\infty }\Lambda _{\alpha }(j)|a_j|^2=\frac{1}{2}+\frac{3\alpha +7}{8(\alpha +2)}<1=\Lambda _{\alpha }(0), \\ \sum _{j=1}^{\infty }\Lambda _{\alpha }(j,1)|a_j|^2=\frac{1}{2(\alpha +2)}+\frac{3\alpha +7}{4(\alpha +2)(\alpha +3)} >\frac{1}{\alpha +2}=\Lambda _{\alpha }(1), \end{aligned}$$

and

$$\begin{aligned} \sum _{j=0}^{\infty }\frac{\Lambda _{\alpha }(1+j)}{\Lambda _{\alpha }(1)}|a_{j}|^2+\frac{\Lambda _{\alpha }(1)}{\Lambda _{\alpha }(0)}|b_{1}|^2=\frac{\alpha +2}{\alpha +3}+\frac{3(3\alpha +7)}{8(\alpha +4)}-\frac{9(\alpha +1)}{8(\alpha +3)} < 1. \end{aligned}$$

Hence, by the Theorem 3.23, \(B_{\varphi }\) is neither contractive nor expansive.