This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

real symmetric, unitary, and complex symmetric weighted composition operators on Bergman spaces of polydisk

Pham Viet Hai School of Applied Mathematics and Informatics, Hanoi University of Science and Technology, Vien Toan ung dung va Tin hoc, Dai hoc Bach khoa Hanoi, 1 Dai Co Viet, Hanoi, Vietnam. hai.phamviet@hust.edu.vn
Abstract.

In this paper, we study weighted composition operators on Bergman spaces of analytic functions which are square integrable on polydisk. We develop the study in full generality, meaning that the corresponding weighted composition operators are not assumed to be bounded. The properties of weighted composition operators such as real symmetry, unitariness, complex symmetry, are characterized fully in simple algebraic terms, involving their symbols. As it turns out, a weighted composition operator having a symmetric structure must be bounded. We also obtain the interesting consequence that real symmetric weighted composition operators are complex symmetric corresponding an adapted and highly relevant conjugation.

Key words and phrases:
complex symmetry, weighted composition operators, Bergman spaces
2010 Mathematics Subject Classification:
47B33, 47B32

1. Introduction

1.1. Weighted composition operators

Our primary object of study is weighted composition operators that are defined as follows. Let 𝕆\mathbb{O} be a Banach space of analytic functions on some domain VdV\subseteq\mathbb{C}^{d}, where d1d\in\mathbb{Z}_{\geq 1}. Given functions g:VV,f:Vg:V\to V,f:V\to\mathbb{C} (often called the symbols), we consider the formal expression [f,g]~\widetilde{[f,g]} by setting

([f,g]~ξ)(z)=f(z)ξ(g(z)),where ξ()𝕆,zV.\displaystyle(\widetilde{[f,g]}\xi)(z)=f(z)\xi(g(z)),\quad\text{where $\xi(\cdot)\in\mathbb{O},z\in V$}.

The maximal weighted composition operator is defined as

dom(Wf,g,max)={ξ()𝕆:[f,g]~ξ()𝕆},\displaystyle\text{dom}(W_{f,g,\max})=\{\xi(\cdot)\in\mathbb{O}:\widetilde{[f,g]}\xi(\cdot)\in\mathbb{O}\},
Wf,g,maxξ()=[f,g]~ξ(),ξ()dom(Wf,gmax).\displaystyle W_{f,g,\max}\xi(\cdot)=\widetilde{[f,g]}\xi(\cdot),\quad\xi(\cdot)\in\text{dom}(W_{f,g\max}).

The operator Wf,gW_{f,g} is called a non-maximal weighted composition operator if the inclusion Wf,gWf,g,maxW_{f,g}\preceq W_{f,g,\max} holds. By the term bounded weighted composition operator, as it is used in the paper, is meant an operator Wf,gW_{f,g} which satisfies (i) dom(Wf,g)=𝕆\text{dom}(W_{f,g})=\mathbb{O} and (ii) there exists a constant R>0R>0 such that

Wf,ghRhh()𝕆.\displaystyle\|W_{f,g}h\|\leq R\|h\|\quad\forall h(\cdot)\in\mathbb{O}.

Note that the phrase "unbounded" is understood as "not necessarily bounded"; in other words, bounded operators belong to the unbounded class. When f()1f(\cdot)\equiv 1, writing CgC_{g} instead of W1,gW_{1,g} and it is named a composition operator.

The earliest reference on weighted composition operators appears to be the Banach-Stone theorem, which states that the only surjective isometries between Banach spaces of real-valued continuous functions are precisely of this type ([29]). Leeuw et al [9] and Forelli [10] obtained the similar results for Hardy spaces. Since then, weighted composition operators have become the subject matter of intensive and extensive research on various spaces and they have made important connections to the study of other operators ([6, 30]). What makes weighted composition operators worth to study is the fact that their properties can be characterized in simple algebraic terms.

When 𝕆\mathbb{O} is Hardy space of the unit disk, there is a large body of work on composition operators (see [8, 28]). It was proven long by Littlewood [23] that composition operators are always bounded. In contrast to the unweighted setting, the boundedness of weighted composition operators has not been well understood. There are many examples (see [8]) which show that weighted composition operators are not bounded. Other properties have been studied such as conditions for (weighted) composition operators to be real symmetric [7], be normal [24, 21], be unitary [21], or be invertible [15].

Considered on Fock spaces, weighted composition operators act very differently; for example, not all composition operators are bounded. Carswell, MacCluer and Schuster [4] showed that the only affine function g(z)=αz+βg(z)=\alpha z+\beta gives rise to the boundedness. Le [22] gave a complete characterization of bounded weighted composition operators on Fock spaces. Characterizations of compactness, isometry, normality, cohyponormality, and supercyclicity have also been produced, including those of Le [22], Hai and Khoi [17], and Carroll and Gilmore [3].

1.2. Complex symmetric operators

A complex symmetric operator is an unbounded linear operator X:dom(X)X:\text{dom}(X)\subseteq\mathcal{H}\to\mathcal{H} on a Hilbert space \mathcal{H} with the property that X=𝒞X𝒞X=\mathcal{C}X^{*}\mathcal{C}, where 𝒞\mathcal{C} is an isometric involution (in short, conjugation) on \mathcal{H}. To indicate the dependence on 𝒞\mathcal{C}, this case is often called 𝒞\mathcal{C}-selfadjoint.

The concept of complex symmetric operators is a natural generalization of complex symmetric matrices, and their general research was commenced by Garcia, Putinar [13, 14]. The class of complex symmetric operators is of great interest for several reasons. First, it has deep classical roots in various branches of mathematics, including matrix theory, function theory, projective geometry, etc. See [13, 14, 12] for details. Second, it is large enough to cover well-known operators such as normal operators, Volterra operators, Hankel operators, compressed Toeplitz operators. More interestingly, differential operators relevant in non-hermitian quantum mechanics obeying a parity and time symmetry, so called 𝒫𝒯\mathcal{PT}-symmetric, belong to this class (see [2]).

Putted forward by Garcia and Hammond [11], the problem of classifying complex symmetric weighted composition operators is of interest and it is being answered in several particular contexts. Jung et al [19] considered bounded weighted composition operators on Hardy spaces and characterized those which are complex symmetric with respect to the conjugation 𝒥f(z)=f(z¯)¯\mathcal{J}f(z)=\overline{f(\overline{z})}. The works [11, 19] make a motivation to study the problem in various spaces. We refer the reader to [25, 16] for Fock space and [31] for Hardy spaces in several variables and [18] for Lebesgue spaces.

1.3. Aim

In this paper, we study weighted composition operators on Bergman spaces of analytic functions which are square integrable on polydisk. We develop the study in full generality, meaning that the corresponding weighted composition operators are not assumed to be bounded. The properties of weighted composition operators such as real symmetry, unitariness, complex symmetry, are characterized fully in simple algebraic terms, involving their symbols. As it turns out, a weighted composition operator having a symmetric structure must be bounded. We also obtain the interesting consequence that real symmetric weighted composition operators are complex symmetric corresponding an adapted and highly relevant conjugation.

2. Preparation

2.1. Notations

Before getting closer to the content, we list notations and terminologies. The domain of an operator is denoted as dom()\text{dom}(\cdot). When dealing with unbounded operators, the symbol ABA\preceq B means that dom(A)dom(B)\text{dom}(A)\subseteq\text{dom}(B) and Ax=BxAx=Bx for xdom(A)x\in\text{dom}(A). Let 1,d={1,2,,d}\mathbb{Z}_{1,d}=\{1,2,\cdots,d\}. Always denote

(2.1) ωp(z)=p¯ppz1p¯z,p𝔻{0}.\displaystyle\omega_{p}(z)=\dfrac{\overline{p}}{p}\cdot\dfrac{p-z}{1-\overline{p}z},\quad p\in\mathbb{D}\setminus\{0\}.

Let 𝔻={z:|z|<1}\mathbb{D}=\{z\in\mathbb{C}:|z|<1\} and 𝕋={z:|z|=1}\mathbb{T}=\{z\in\mathbb{C}:|z|=1\}. For a fixed integer dd, the polydisk 𝔻d\mathbb{D}^{d} of d\mathbb{C}^{d} is the Cartesian product of dd copies of 𝔻\mathbb{D}. Bergman space (𝔻d)\mathcal{B}_{\ell}(\mathbb{D}^{d}) consists of all analytic functions ξ:𝔻d\xi:\mathbb{D}^{d}\to\mathbb{C} with

ξ2=𝔻d|ξ(z)|2j=1d(1+j)(1|zj|2)jdA(zj),\displaystyle\|\xi\|^{2}=\int\limits_{\mathbb{D}^{d}}|\xi(z)|^{2}\prod_{j=1}^{d}(1+\ell_{j})(1-|z_{j}|^{2})^{\ell_{j}}\,dA(z_{j}),

where dAdA denotes the normalized area measure on 𝔻\mathbb{D}. The reproducing kernel in (𝔻d)\mathcal{B}_{\ell}(\mathbb{D}^{d}) is given by

Kz(u)=j=1d1(1ujzj¯)j+2z,u𝔻d.\displaystyle K_{z}(u)=\prod_{j=1}^{d}\dfrac{1}{(1-u_{j}\overline{z_{j}})^{\ell_{j}+2}}\quad\forall z,u\in\mathbb{D}^{d}.

2.2. Elementary estimate

The following lemma may be well-known; but the proof is given, for a completeness of exposition.

Lemma 2.1.

Let 𝐜𝔻t\mathbf{c}\in\mathbb{D}^{t} and s1s\in\mathbb{Z}_{\geq 1}. Then there exists a constant Δ\Delta such that

𝔻s|h(𝐜,y)|2j=1s(1|yj|2)jdA(yj)Δ𝔻t×𝔻s|h(z)|2j=1t+s(1|zj|2)jdA(zj).\displaystyle\int\limits_{\mathbb{D}^{s}}|h(\mathbf{c},y)|^{2}\prod_{j=1}^{s}(1-|y_{j}|^{2})^{\ell_{j}}dA(y_{j})\leq\Delta\int\limits_{\mathbb{D}^{t}\times\mathbb{D}^{s}}|h(z)|^{2}\prod_{j=1}^{t+s}(1-|z_{j}|^{2})^{\ell_{j}}dA(z_{j}).

Consequently, if h()(𝔻t×𝔻s)h(\cdot)\in\mathcal{B}_{\ell}(\mathbb{D}^{t}\times\mathbb{D}^{s}), then the function h𝐜()(𝔻s)h_{\mathbf{c}}(\cdot)\in\mathcal{B}_{\ell}(\mathbb{D}^{s}), where h𝐜(y)=h(𝐜,y)h_{\mathbf{c}}(y)=h(\mathbf{c},y).

Proof.

Denote δj=(1|𝐜j|)/2\delta_{j}=(1-|\mathbf{c}_{j}|)/2 and D(𝐜,δ)t={z=(z1,z2,,zt)t:|zjcj|<δj1jt}D(\mathbf{c},\delta)^{t}=\{z=(z_{1},z_{2},\cdots,z_{t})\in\mathbb{C}^{t}:|z_{j}-c_{j}|<\delta_{j}\,\forall 1\leq j\leq t\}. Setting

W=𝔻t×𝔻s|h(z)|2j=1t+s(1|zj|2)jdA(zj),\displaystyle W=\int\limits_{\mathbb{D}^{t}\times\mathbb{D}^{s}}|h(z)|^{2}\prod_{j=1}^{t+s}(1-|z_{j}|^{2})^{\ell_{j}}dA(z_{j}),

by the Fubini theorem, we have

W=𝔻sj=1s(1|yj|2)jdA(yj)𝔻t|h(x,y)|2j=1t(1|xj|2)jdA(xj)\displaystyle W=\int\limits_{\mathbb{D}^{s}}\prod_{j=1}^{s}(1-|y_{j}|^{2})^{\ell_{j}}dA(y_{j})\int\limits_{\mathbb{D}^{t}}|h(x,y)|^{2}\prod_{j=1}^{t}(1-|x_{j}|^{2})^{\ell_{j}}dA(x_{j})
𝔻sj=1s(1|yj|2)jdA(yj)D(𝐜,δ)t|h(x,y)|2j=1t(1|xj|2)jdA(xj)\displaystyle\geq\int\limits_{\mathbb{D}^{s}}\prod_{j=1}^{s}(1-|y_{j}|^{2})^{\ell_{j}}dA(y_{j})\int\limits_{D(\mathbf{c},\delta)^{t}}|h(x,y)|^{2}\prod_{j=1}^{t}(1-|x_{j}|^{2})^{\ell_{j}}dA(x_{j})
j=1t(1(1+|𝐜j|2)2)j𝔻sj=1s(1|yj|2)jdA(yj)D(𝐜,δ)t|h(x,y)|2𝑑A(xj),\displaystyle\geq\prod_{j=1}^{t}\left(1-\left(\dfrac{1+|\mathbf{c}_{j}|}{2}\right)^{2}\right)^{\ell_{j}}\int\limits_{\mathbb{D}^{s}}\prod_{j=1}^{s}(1-|y_{j}|^{2})^{\ell_{j}}dA(y_{j})\int\limits_{D(\mathbf{c},\delta)^{t}}|h(x,y)|^{2}dA(x_{j}),

which implies, by the subharmonic property of the function x|h(x,y)|2x\mapsto|h(x,y)|^{2}, that

Wj=1t(1(1+|𝐜j|2)2)j𝔻sj=1s(1|yj|2)jdA(yj)πtj=1tδj|h(𝐜,y)|2\displaystyle W\geq\prod_{j=1}^{t}\left(1-\left(\dfrac{1+|\mathbf{c}_{j}|}{2}\right)^{2}\right)^{\ell_{j}}\int\limits_{\mathbb{D}^{s}}\prod_{j=1}^{s}(1-|y_{j}|^{2})^{\ell_{j}}dA(y_{j})\pi^{t}\prod_{j=1}^{t}\delta_{j}|h(\mathbf{c},y)|^{2}
=πtj=1tδj(1(1+|𝐜j|2)2)j𝔻s|h(𝐜,y)|2j=1s(1|yj|2)jdA(yj).\displaystyle=\pi^{t}\prod_{j=1}^{t}\delta_{j}\left(1-\left(\dfrac{1+|\mathbf{c}_{j}|}{2}\right)^{2}\right)^{\ell_{j}}\int\limits_{\mathbb{D}^{s}}|h(\mathbf{c},y)|^{2}\prod_{j=1}^{s}(1-|y_{j}|^{2})^{\ell_{j}}dA(y_{j}).

We recall a condition when a linear fractional function is a self-mapping of 𝔻\mathbb{D}.

Lemma 2.2 ([5]).

Let

ξ(z)=az+bcz+d,where a,b,c,d with adbc0.\displaystyle\xi(z)=\dfrac{az+b}{cz+d},\quad\text{where $a,b,c,d\in\mathbb{C}$ with $ad-bc\neq 0$}.

Then ξ()\xi(\cdot) is a self-mapping of 𝔻\mathbb{D} if and only if

|bd¯ac¯|+|adbc||d|2|c|2.\displaystyle|b\overline{d}-a\overline{c}|+|ad-bc|\leq|d|^{2}-|c|^{2}.

2.3. Algebraic observations

Lemma 2.3.

Let f:𝔻d,g:𝔻f:\mathbb{D}^{d}\to\mathbb{C},g:\mathbb{D}\to\mathbb{C} be analytic functions, where d1d\in\mathbb{Z}_{\geq 1}. Suppose that for every κ1,d\kappa\in\mathbb{Z}_{1,d}, the product g(uκ)f(u1,u2,,un)g(u_{\kappa})f(u_{1},u_{2},\cdots,u_{n}) is a function of variables u1,u2,,uκ1,uκ+1,,unu_{1},u_{2},\cdots,u_{\kappa-1},u_{\kappa+1},\cdots,u_{n}. Then for every κ1,d\kappa\in\mathbb{Z}_{1,d} the product f(u1,u2,,un)j=1κg(uj)f(u_{1},u_{2},\cdots,u_{n})\prod_{j=1}^{\kappa}g(u_{j}) is a function of variables uκ+1,uκ+2,,unu_{\kappa+1},u_{\kappa+2},\cdots,u_{n}.

Proof.

The lemma is proven by induction on κ\kappa and its proof is left to the reader. ∎

Lemma 2.4.

Let ψ:𝔻𝔻\psi:\mathbb{D}\to\mathbb{D} be an analytic function with the property that

(2.2) ψ(u)(1uψ(z)¯)2=ψ(z)¯(1ψ(u)z¯)2z,u𝔻.\displaystyle\psi^{\prime}(u)(1-u\overline{\psi(z)})^{2}=\overline{\psi^{\prime}(z)}(1-\psi(u)\overline{z})^{2}\quad\forall z,u\in\mathbb{D}.

Then

(2.3) ψ(z)=𝐚+𝐛z1𝐚¯z,\displaystyle\psi(z)=\mathbf{a}+\dfrac{\mathbf{b}z}{1-\overline{\mathbf{a}}z},

where coefficients satisfy

(2.4) 𝐚𝔻,𝐛,\displaystyle\mathbf{a}\in\mathbb{D},\mathbf{b}\in\mathbb{R},
(2.5) |𝐚(𝐛|𝐚|2+1)|+|𝐛|1|𝐚|2.\displaystyle|\mathbf{a}(\mathbf{b}-|\mathbf{a}|^{2}+1)|+|\mathbf{b}|\leq 1-|\mathbf{a}|^{2}.
Proof.

It is clear that the case when ψ()const\psi(\cdot)\equiv\text{const} verifies (2.2). Next is to consider the remaining case ψ()const\psi(\cdot)\not\equiv\text{const}. Let z𝔻z_{\star}\in\mathbb{D} such that ψ(z)0\psi(z_{\star})\neq 0. Letting z=zz=z_{\star} in (2.2), we can observe

ψ(u)(1ψ(u)z¯)2=ψ(z)¯(1uψ(z)¯)2.\displaystyle\dfrac{\psi^{\prime}(u)}{(1-\psi(u)\overline{z_{\star}})^{2}}=\dfrac{\overline{\psi^{\prime}(z_{\star})}}{(1-u\overline{\psi(z_{\star})})^{2}}.

After integrating with respect to the variable uu, ψ()\psi(\cdot) is of form

ψ(z)=Az+BCz+D,\displaystyle\psi(z)=\dfrac{Az+B}{Cz+D},

where A,B,C,DA,B,C,D are complex constants. There are two cases of the coefficient CC. If C=0C=0, then

ψ(z)=A^z+B^,where A^=AD and B^=BD.\displaystyle\psi(z)=\widehat{A}z+\widehat{B},\quad\text{where $\widehat{A}=\dfrac{A}{D}$ and $\widehat{B}=\dfrac{B}{D}$}.

Through setting ψ(z)=A^z+B^\psi(z)=\widehat{A}z+\widehat{B} in (2.2) and then equating coefficients, we get

A^,B^=0.\displaystyle\widehat{A}\in\mathbb{R},\quad\widehat{B}=0.

If C0C\neq 0, then

ψ(z)=G+Ez+F,where G=AC,E=BCADC2,F=DC.\displaystyle\psi(z)=G+\dfrac{E}{z+F},\quad\text{where $G=\dfrac{A}{C},E=\dfrac{BC-AD}{C^{2}},F=\dfrac{D}{C}$}.

It follows from ψ()const\psi(\cdot)\not\equiv\text{const}, that E0E\neq 0. By way of substituting this form of ψ()\psi(\cdot) back into (2.2) and then equating coefficients, we obtain

G¯2G2=E¯E,G¯G=F¯F,GF+E=EG¯E¯G.\displaystyle\dfrac{\overline{G}^{2}}{G^{2}}=\dfrac{\overline{E}}{E},\quad\dfrac{\overline{G}}{G}=\dfrac{\overline{F}}{F},\quad GF+E=-\dfrac{E\overline{G}}{\overline{E}G}.

Thus, (2.3)-(2.4) hold, where

𝐚=G+EF,𝐛=EF2;\displaystyle\mathbf{a}=G+\dfrac{E}{F},\quad\mathbf{b}=-\dfrac{E}{F^{2}};

meanwhile, (2.5) follows directly from Lemma 2.1. ∎

Lemma 2.5.

Let θ𝕋\theta\in\mathbb{T} and ψ:𝔻𝔻\psi:\mathbb{D}\to\mathbb{D} be an analytic function with the property that

(2.6) ψ(y)(1θyψ(x))2=ψ(x)(1θψ(y)x)2x,y𝔻.\displaystyle\psi^{\prime}(y)(1-\theta y\psi(x))^{2}=\psi^{\prime}(x)(1-\theta\psi(y)x)^{2}\quad\forall x,y\in\mathbb{D}.

Then

ψ(z)=α0+α1θz1α0θz,\displaystyle\psi(z)=\alpha_{0}+\dfrac{\alpha_{1}\theta z}{1-\alpha_{0}\theta z},

where coefficients satisfy

|α0+α0¯(α1α02)|+|α1|1|α0|2.\displaystyle|\alpha_{0}+\overline{\alpha_{0}}(\alpha_{1}-\alpha_{0}^{2})|+|\alpha_{1}|\leq 1-|\alpha_{0}|^{2}.
Proof.

It is clear that the case when ψ()const\psi(\cdot)\equiv\text{const} verifies (2.6). Next is to consider the remaining case ψ()const\psi(\cdot)\not\equiv\text{const}. Let z𝔻z_{\star}\in\mathbb{D} such that ψ(z)0\psi(z_{\star})\neq 0. Letting x=zx=z_{\star} in (2.6), we get

ψ(y)(1θψ(y)z)2=ψ(z)(1θyψ(z))2.\displaystyle\dfrac{\psi^{\prime}(y)}{(1-\theta\psi(y)z_{\star})^{2}}=\dfrac{\psi^{\prime}(z_{\star})}{(1-\theta y\psi(z_{\star}))^{2}}.

By way of integrating with respect to the variable yy, we observe ψ()\psi(\cdot) is of form

ψ(z)=Az+BCz+D,\displaystyle\psi(z)=\dfrac{Az+B}{Cz+D},

where A,B,C,DA,B,C,D are complex constants. There are two cases of CC. If C=0C=0, then

ψ(z)=A^z+B^,where A^=AD and B^=BD.\displaystyle\psi(z)=\widehat{A}z+\widehat{B},\quad\text{where $\widehat{A}=\dfrac{A}{D}$ and $\widehat{B}=\dfrac{B}{D}$}.

Through setting ψ(z)=A^z+B^\psi(z)=\widehat{A}z+\widehat{B} in (2.6) and then equating coefficients, we get B^=0\widehat{B}=0. Consider when C0C\neq 0 and then

ψ(z)=G+Ez+F,where G=AC,E=BCADC2,F=DC.\displaystyle\psi(z)=G+\dfrac{E}{z+F},\quad\text{where $G=\dfrac{A}{C},E=\dfrac{BC-AD}{C^{2}},F=\dfrac{D}{C}$}.

It follows from ψ()const\psi(\cdot)\not\equiv\text{const}, that E0E\neq 0. By way of substituting this form of ψ()\psi(\cdot) back into (2.6) and then equating coefficients, we obtain

{either G=F=0,E=θ¯,or E+GF=θ¯.\displaystyle\begin{cases}\text{either $G=F=0,E=\overline{\theta}$},\\ \text{or $E+GF=-\overline{\theta}$}.\end{cases}

Thus, the conclusion holds, where

α0=G+EF,α1=EF2θ¯.\displaystyle\alpha_{0}=G+\dfrac{E}{F},\quad\alpha_{1}=-\dfrac{E}{F^{2}}\overline{\theta}.

Lemma 2.6.

Let p𝔻{0}p\in\mathbb{D}\setminus\{0\} and ωp()\omega_{p}(\cdot) be the function given by (2.1). If the function ψ:𝔻𝔻\psi:\mathbb{D}\to\mathbb{D} satisfies

(2.7) ωp(y)ψ(u)[1uωp(ψ(y))]2=ωp(ψ(y))ψ(y)[1ωp(y)ψ(u)]2,\displaystyle\omega_{p}^{\prime}(y)\psi^{\prime}(u)[1-u\omega_{p}(\psi(y))]^{2}=\omega_{p}^{\prime}(\psi(y))\psi^{\prime}(y)[1-\omega_{p}(y)\psi(u)]^{2},

then either

(2.8) ψconst\displaystyle\psi\equiv\text{const}

or

(2.9) ψ(z)=G+Ez+F,\displaystyle\psi(z)=G+\dfrac{E}{z+F},

where coefficients satisfy

(2.10) pG|p|2=F|p|2+(GF+E)p¯,\displaystyle p-G|p|^{2}=-F|p|^{2}+(GF+E)\overline{p},
(2.11) E0,\displaystyle E\neq 0,
(2.12) |(GF+E)F¯G|+|E||F|21.\displaystyle|(GF+E)\overline{F}-G|+|E|\leq|F|^{2}-1.
Proof.

It is clear that the case when ψ()const\psi(\cdot)\equiv\text{const} verifies (2.7). Next is to consider the remaining case ψ()const\psi(\cdot)\not\equiv\text{const}. Since ωp(z)=p¯p|p|21(1p¯z)2\omega_{p}^{\prime}(z)=\frac{\overline{p}}{p}\cdot\frac{|p|^{2}-1}{(1-\overline{p}z)^{2}}, equation (2.7) is reduced to the following

(2.13) ψ(u)[1uωp(ψ(y))]2(1p¯y)2=ψ(y)[1ωp(y)ψ(u)]2(1p¯ψ(y))2\displaystyle\dfrac{\psi^{\prime}(u)[1-u\omega_{p}(\psi(y))]^{2}}{(1-\overline{p}y)^{2}}=\dfrac{\psi^{\prime}(y)[1-\omega_{p}(y)\psi(u)]^{2}}{(1-\overline{p}\psi(y))^{2}}

or equivalently to saying that

ψ(u)[1ωp(y)ψ(u)]2=ψ(y)(1p¯ψ(y))2(1p¯y)2[1uωp(ψ(y))]2.\displaystyle\dfrac{\psi^{\prime}(u)}{[1-\omega_{p}(y)\psi(u)]^{2}}=\dfrac{\psi^{\prime}(y)}{(1-\overline{p}\psi(y))^{2}}\cdot\dfrac{(1-\overline{p}y)^{2}}{[1-u\omega_{p}(\psi(y))]^{2}}.

Like the arguments in Lemmas 2.5 and 2.4, ψ()\psi(\cdot) is of form

(2.14) ψ(z)=Az+BCz+D,\displaystyle\psi(z)=\dfrac{Az+B}{Cz+D},

where A,B,C,DA,B,C,D are complex constants. Setting (2.1) in (2.13), we find

ψ(u)[p(1p¯ψ(y))p¯u(pψ(y))]2=ψ(y)[p(1p¯y)p¯(py)ψ(u)]2,\displaystyle\psi^{\prime}(u)[p(1-\overline{p}\psi(y))-\overline{p}u(p-\psi(y))]^{2}=\psi^{\prime}(y)[p(1-\overline{p}y)-\overline{p}(p-y)\psi(u)]^{2},

which implies, by (2.14), that

[(p¯A|p|2C)uy+(p¯B|p|2D)u+(pC|p|2A)y+pD|p|2B]2\displaystyle[(\overline{p}A-|p|^{2}C)uy+(\overline{p}B-|p|^{2}D)u+(pC-|p|^{2}A)y+pD-|p|^{2}B]^{2}
=[(p¯A|p|2C)uy+(pC|p|2A)u+(p¯B|p|2D)y+pD|p|2B]2.\displaystyle=[(\overline{p}A-|p|^{2}C)uy+(pC-|p|^{2}A)u+(\overline{p}B-|p|^{2}D)y+pD-|p|^{2}B]^{2}.

There are two cases.

- If

(p¯A|p|2C)uy+(p¯B|p|2D)u+(pC|p|2A)y+pD|p|2B\displaystyle(\overline{p}A-|p|^{2}C)uy+(\overline{p}B-|p|^{2}D)u+(pC-|p|^{2}A)y+pD-|p|^{2}B
=(p¯A|p|2C)uy+(pC|p|2A)u+(p¯B|p|2D)y+pD|p|2B,\displaystyle=(\overline{p}A-|p|^{2}C)uy+(pC-|p|^{2}A)u+(\overline{p}B-|p|^{2}D)y+pD-|p|^{2}B,

then after equating coefficients, we get p¯B|p|2D=pC|p|2A\overline{p}B-|p|^{2}D=pC-|p|^{2}A. It then implies (2.10), where

G=AC,F=DC,E=BCADC2.\displaystyle G=\dfrac{A}{C},\quad F=\dfrac{D}{C},\quad E=\dfrac{BC-AD}{C^{2}}.

Note that condition (2.12) follows directly from Lemma 2.2.

- If

(p¯A|p|2C)uy+(p¯B|p|2D)u+(pC|p|2A)y+pD|p|2B\displaystyle(\overline{p}A-|p|^{2}C)uy+(\overline{p}B-|p|^{2}D)u+(pC-|p|^{2}A)y+pD-|p|^{2}B
=(p¯A|p|2C)uy(pC|p|2A)u(p¯B|p|2D)ypD+|p|2B,\displaystyle=-(\overline{p}A-|p|^{2}C)uy-(pC-|p|^{2}A)u-(\overline{p}B-|p|^{2}D)y-pD+|p|^{2}B,

then after equating coefficients, we get

{p¯A|p|2C=0,p¯B|p|2D=(pC|p|2A),pD|p|2B=0,{A=p¯B,D=p¯B,C=p¯pB.\displaystyle\begin{cases}\overline{p}A-|p|^{2}C=0,\\ \overline{p}B-|p|^{2}D=-(pC-|p|^{2}A),\\ pD-|p|^{2}B=0,\end{cases}\Longrightarrow\begin{cases}A=-\overline{p}B,\\ D=\overline{p}B,\\ C=-\dfrac{\overline{p}}{p}B.\end{cases}

Thus, this case gives

ψ(z)=|p|2zpp¯z|p|2;\displaystyle\psi(z)=\dfrac{|p|^{2}z-p}{\overline{p}z-|p|^{2}};

but this is impossible as ψ\psi is not a self-mapping of 𝔻\mathbb{D} (see Lemma 2.2). ∎

2.4. Basic properties of Wf,gW_{f,g}

The closed graph of the maximal operator Wf,g,maxW_{f,g,\max} is left to the reader as its proof is similar to those used in [16].

Proposition 2.7.

The maximal operator Wf,g,maxW_{f,g,\max} is closed.

Consequently, we get a criterion for the boundedness of the maximal operator Wf,g,maxW_{f,g,\max}.

Corollary 2.8.

The maximal operator Wf,g,maxW_{f,g,\max} is bounded if and only if dom(Wf,g,max)=(𝔻d)\text{dom}(W_{f,g,\max})=\mathcal{B}_{\ell}(\mathbb{D}^{d}).

The following lemma will be used frequently to prove the main results as it shows that kernel functions always belong to the domain of Wf,g,maxW_{f,g,\max}^{*}.

Lemma 2.9.

Suppose that the operator Wf,gW_{f,g} is densely defined. Then equality Wf,gKz=f(z)¯Kg(z)W_{f,g}^{*}K_{z}=\overline{f(z)}K_{g(z)} holds for every z𝔻dz\in\mathbb{D}^{d}.

We take a while to focus on the very restrictive category of operators induced by the following functions

(2.15) φθ(z)=(θ1z11z1θ1¯,θ2z21z2θ2¯,,θdzd1zdθd¯)\displaystyle\varphi_{\theta}(z)=\left(\dfrac{\theta_{1}-z_{1}}{1-z_{1}\overline{\theta_{1}}},\dfrac{\theta_{2}-z_{2}}{1-z_{2}\overline{\theta_{2}}},\cdots,\dfrac{\theta_{d}-z_{d}}{1-z_{d}\overline{\theta_{d}}}\right)

and

(2.16) ψθ(z)=Kθ(z)/Kθ.\displaystyle\psi_{\theta}(z)=K_{\theta}(z)/\|K_{\theta}\|.
Proposition 2.10.

Let θ𝔻d\theta\in\mathbb{D}^{d}. Let 𝒜θ,max\mathcal{A}_{\theta,\max} be the maximal operator generated by [ψθ,φθ]~\widetilde{[\psi_{\theta},\varphi_{\theta}]}. Then the operator 𝒜θ,max\mathcal{A}_{\theta,\max} is unitary on (𝔻d)\mathcal{B}_{\ell}(\mathbb{D}^{d}).

Proof.

First, we show that dom(𝒜θ,max)=(𝔻d)\text{dom}(\mathcal{A}_{\theta,\max})=\mathcal{B}_{\ell}(\mathbb{D}^{d}) and 𝒜θ,maxh=h\|\mathcal{A}_{\theta,\max}h\|=\|h\| for every h(𝔻d)h\in\mathcal{B}_{\ell}(\mathbb{D}^{d}). Indeed, for h(𝔻d)h\in\mathcal{B}_{\ell}(\mathbb{D}^{d}), we consider

I=𝔻d|[ψθ,φθ]~h(z)|2j=1d(1+j)(1|zj|2)jdA(zj),\displaystyle I=\int\limits_{\mathbb{D}^{d}}|\widetilde{[\psi_{\theta},\varphi_{\theta}]}h(z)|^{2}\prod_{j=1}^{d}(1+\ell_{j})(1-|z_{j}|^{2})^{\ell_{j}}\,dA(z_{j}),

which implies, after doing the change of variables xj=θjzj1θj¯zjx_{j}=\frac{\theta_{j}-z_{j}}{1-\overline{\theta_{j}}z_{j}}, that

I=𝔻dj=1d|h(x)|2j=1d(1+j)(1|xj|2)jdA(xj),\displaystyle I=\int\limits_{\mathbb{D}^{d}}\prod_{j=1}^{d}|h(x)|^{2}\prod_{j=1}^{d}(1+\ell_{j})(1-|x_{j}|^{2})^{\ell_{j}}\,dA(x_{j}),

as desired. Next, by Lemma 2.9, we have

𝒜θ,max𝒜θ,maxKz(y)=ψθ(z)¯ψθ(y)Kφθ(z)(φθ(y))=Kz(y),\displaystyle\mathcal{A}_{\theta,\max}\mathcal{A}_{\theta,\max}^{*}K_{z}(y)=\overline{\psi_{\theta}(z)}\psi_{\theta}(y)K_{\varphi_{\theta}(z)}(\varphi_{\theta}(y))=K_{z}(y),

which implies, as the linear span of kernel functions is dense, that 𝒜θ,max𝒜θ,max=I\mathcal{A}_{\theta,\max}^{*}\mathcal{A}_{\theta,\max}=I. ∎

2.5. Conjugations

Let {U1,U2}\{U_{1},U_{2}\} be a partition of 1,d\mathbb{Z}_{1,d}, that is

(2.17) U1U2=1,d,U1U2=.\displaystyle U_{1}\cup U_{2}=\mathbb{Z}_{1,d},\quad U_{1}\cap U_{2}=\emptyset.

Let 𝐩=(𝐩1,𝐩2,,𝐩|U1|)𝔻|U1|{0}\mathbf{p}=(\mathbf{p}_{1},\mathbf{p}_{2},\cdots,\mathbf{p}_{|U_{1}|})\in\mathbb{D}^{|U_{1}|}\setminus\{0\} and 𝐪=(𝐪1,𝐪2.,𝐪|U2|)𝕋|U2|\mathbf{q}=(\mathbf{q}_{1},\mathbf{q}_{2}.\cdots,\mathbf{q}_{|U_{2}|})\in\mathbb{T}^{|U_{2}|}. Consider the operator

(2.18) 𝒞𝐩,𝐪h(z)=ϑ𝐩,U(z)h(ω(z)¯)¯,\displaystyle\mathcal{C}_{\mathbf{p},\mathbf{q}}h(z)=\vartheta_{\mathbf{p},U}(z)\overline{h(\overline{\omega(z)})},

where ϑ𝐩,U:𝔻d,ω=(ω1,ω2,,ωd):𝔻d𝔻d\vartheta_{\mathbf{p},U}:\mathbb{D}^{d}\to\mathbb{C},\omega=(\omega_{1},\omega_{2},\cdots,\omega_{d}):\mathbb{D}^{d}\to\mathbb{D}^{d} are given by

(2.19) ϑ𝐩,U(z)=jU1(1|𝐩j|2)1+j/2(1𝐩j¯zj)j+2,\displaystyle\vartheta_{\mathbf{p},U}(z)=\prod_{j\in U_{1}}\dfrac{(1-|\mathbf{p}_{j}|^{2})^{1+\ell_{j}/2}}{(1-\overline{\mathbf{p}_{j}}z_{j})^{\ell_{j}+2}},
(2.20) ωj(zj)={𝐩j¯𝐩j𝐩jzj1𝐩j¯zj,if jU1,𝐪jzj,if jU2.\displaystyle\omega_{j}(z_{j})=\begin{cases}\dfrac{\overline{\mathbf{p}_{j}}}{\mathbf{p}_{j}}\cdot\dfrac{\mathbf{p}_{j}-z_{j}}{1-\overline{\mathbf{p}_{j}}z_{j}},\quad\text{if $j\in U_{1}$},\\ \\ \mathbf{q}_{j}z_{j},\quad\text{if $j\in U_{2}$}.\end{cases}
Lemma 2.11.

The operator 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}} is a conjugation and moreover it satisfies 𝒞𝐩,𝐪Kz=ϑ𝐩,U(z)Kω(z)¯\mathcal{C}_{\mathbf{p},\mathbf{q}}K_{z}=\vartheta_{\mathbf{p},U}(z)K_{\overline{\omega(z)}}.

Proof.

For every h(𝔻d)h\in\mathcal{B}_{\ell}(\mathbb{D}^{d}), we consider the integral

I=𝔻d|ϑ𝐩,U(z)h(ω(z)¯)|2j=1d(1+j)(1|zj|2)jdA(zj),\displaystyle I=\int\limits_{\mathbb{D}^{d}}|\vartheta_{\mathbf{p},U}(z)h(\overline{\omega(z)})|^{2}\prod_{j=1}^{d}(1+\ell_{j})(1-|z_{j}|^{2})^{\ell_{j}}\,dA(z_{j}),

which implies, after doing the change of variables x=ω(z)x=\omega(z), that I=h2I=\|h\|^{2}. The equality shows that the operator 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}} is isometric. On one hand, we have ω(ω(z)¯)¯=z\overline{\omega(\overline{\omega(z)})}=z and on the other hand, we observe ϑ𝐩,U(z)ϑ𝐩,U(ω(z)¯)¯=1\vartheta_{\mathbf{p},U}(z)\overline{\vartheta_{\mathbf{p},U}(\overline{\omega(z)})}=1. Thus, the operator 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}} is involutive. ∎

3. Real symmetry

Recall that a linear operator QQ is called real symmetric if the equality Q=QQ=Q^{*} holds. In this section, we are concerned with how the function-theoretic properties of the symbols affect the real symmetry of the corresponding weighted composition operator, and vice versa. As a consequence, we show that a real symmetric weighted composition operator must be bounded.

We start the section by a lemma which focuses on symbol computation.

Lemma 3.1.

Let d1d\in\mathbb{Z}_{\geq 1} and 0d\ell\in\mathbb{Z}_{\geq 0}^{d}. Let f:𝔻d,g=(g1,g2,,gd):𝔻d𝔻df:\mathbb{D}^{d}\to\mathbb{C},g=(g_{1},g_{2},\cdots,g_{d}):\mathbb{D}^{d}\to\mathbb{D}^{d} be analytic functions. Suppose that

(3.1) f(z)¯Kg(z)(u)=f(u)Kz(g(u))z,u𝔻d.\displaystyle\overline{f(z)}K_{g(z)}(u)=f(u)K_{z}(g(u))\quad\forall z,u\in\mathbb{D}^{d}.

Then the following conclusions hold.

  1. (1)

    The functions f(),g()f(\cdot),g(\cdot) are of forms

    (3.2) f(z)=𝐜K𝐚(z),gκ(z)=𝐚κ+𝐛κzκ1𝐚κ¯zκκ1,d,\displaystyle f(z)=\mathbf{c}K_{\mathbf{a}}(z),\quad g_{\kappa}(z)=\mathbf{a}_{\kappa}+\dfrac{\mathbf{b}_{\kappa}z_{\kappa}}{1-\overline{\mathbf{a}_{\kappa}}z_{\kappa}}\quad\forall\kappa\in\mathbb{Z}_{1,d},

    where coefficients satisfy

    (3.3) 𝐜,𝐛=(𝐛1,𝐛2,,𝐛d)d,𝐚=(𝐚1,𝐚2,,𝐚d)𝔻d\displaystyle\mathbf{c}\in\mathbb{R},\quad\mathbf{b}=(\mathbf{b}_{1},\mathbf{b}_{2},\cdots,\mathbf{b}_{d})\in\mathbb{R}^{d},\quad\mathbf{a}=(\mathbf{a}_{1},\mathbf{a}_{2},\cdots,\mathbf{a}_{d})\in\mathbb{D}^{d}

    and

    (3.4) |𝐚κ(𝐛κ|𝐚κ|2+1)|+|𝐛κ|1|𝐚κ|2κ1,d.\displaystyle|\mathbf{a}_{\kappa}(\mathbf{b}_{\kappa}-|\mathbf{a}_{\kappa}|^{2}+1)|+|\mathbf{b}_{\kappa}|\leq 1-|\mathbf{a}_{\kappa}|^{2}\quad\forall\kappa\in\mathbb{Z}_{1,d}.
  2. (2)

    If the functions are given in item (1), then Wf,g,maxW_{f,g,\max} is bounded on (𝔻d)\mathcal{B}_{\ell}(\mathbb{D}^{d}).

Proof.

(1) Taking into account the form of kernel functions, (3.1) can be expressed in the following

(3.5) f(z)¯j=1d(1gj(u)zj¯)j+2=f(u)j=1d(1ujgj(z)¯)j+2.\displaystyle\overline{f(z)}\prod_{j=1}^{d}\left(1-g_{j}(u)\overline{z_{j}}\right)^{\ell_{j}+2}=f(u)\prod_{j=1}^{d}\left(1-u_{j}\overline{g_{j}(z)}\right)^{\ell_{j}+2}.

Take arbitrarily κ1,d\kappa\in\mathbb{Z}_{1,d}. Differentiating (3.5) with respect to the variable uκu_{\kappa} (i.e. taking derivative uκ\partial_{u_{\kappa}}), we obtain

f(z)¯t=1d(t+2)(1gt(u)zt¯)t+1zt¯uκ(gt(u))jt(1gj(u)zj¯)j+2\displaystyle-\overline{f(z)}\sum_{t=1}^{d}(\ell_{t}+2)\left(1-g_{t}(u)\overline{z_{t}}\right)^{\ell_{t}+1}\overline{z_{t}}\partial_{u_{\kappa}}(g_{t}(u))\prod_{j\neq t}\left(1-g_{j}(u)\overline{z_{j}}\right)^{\ell_{j}+2}
=uκ(f(u))j=1d(1ujgj(z)¯)j+2\displaystyle=\partial_{u_{\kappa}}(f(u))\prod_{j=1}^{d}\left(1-u_{j}\overline{g_{j}(z)}\right)^{\ell_{j}+2}
(3.6) f(u)(κ+2)(1uκgκ(z)¯)κ+1gκ(z)¯jκ(1ujgj(z)¯)j+2.\displaystyle-f(u)(\ell_{\kappa}+2)\left(1-u_{\kappa}\overline{g_{\kappa}(z)}\right)^{\ell_{\kappa}+1}\overline{g_{\kappa}(z)}\prod_{j\neq\kappa}\left(1-u_{j}\overline{g_{j}(z)}\right)^{\ell_{j}+2}.

Equation (3.6) divided by (3.5) is equal to

(3.7) t=1d(t+2)zt¯uκ(gt(u))1gt(u)zt¯=uκ(f(u))f(u)(κ+2)gκ(z)¯1uκgκ(z)¯.\displaystyle-\sum_{t=1}^{d}(\ell_{t}+2)\dfrac{\overline{z_{t}}\partial_{u_{\kappa}}(g_{t}(u))}{1-g_{t}(u)\overline{z_{t}}}=\dfrac{\partial_{u_{\kappa}}(f(u))}{f(u)}-(\ell_{\kappa}+2)\dfrac{\overline{g_{\kappa}(z)}}{1-u_{\kappa}\overline{g_{\kappa}(z)}}.

In particular with z=0z=0, we get

(3.8) uκ(f(u))f(u)=(κ+2)gκ(0)¯1uκgκ(0)¯.\displaystyle\dfrac{\partial_{u_{\kappa}}(f(u))}{f(u)}=(\ell_{\kappa}+2)\dfrac{\overline{g_{\kappa}(0)}}{1-u_{\kappa}\overline{g_{\kappa}(0)}}.

Setting

H(u)=(uκgκ(0)¯1)κ+2f(u),\displaystyle H(u)=(u_{\kappa}\overline{g_{\kappa}(0)}-1)^{\ell_{\kappa}+2}f(u),

then

uκ(f(u))=uκ(H(u))(uκgκ(0)¯1)κ+2(κ+2)gκ(0)¯H(u)(uκgκ(0)¯1)κ+3\displaystyle\partial_{u_{\kappa}}(f(u))=\dfrac{\partial_{u_{\kappa}}(H(u))}{(u_{\kappa}\overline{g_{\kappa}(0)}-1)^{\ell_{\kappa}+2}}-\dfrac{(\ell_{\kappa}+2)\overline{g_{\kappa}(0)}H(u)}{(u_{\kappa}\overline{g_{\kappa}(0)}-1)^{\ell_{\kappa}+3}}

and so (3.8) becomes uκ(H(u))=0\partial_{u_{\kappa}}(H(u))=0. For that reason, (uκgκ(0)¯1)κ+2f(u)(u_{\kappa}\overline{g_{\kappa}(0)}-1)^{\ell_{\kappa}+2}f(u) is a function of variables u1,u2,,uκ1,uκ+1,,udu_{1},u_{2},\cdots,u_{\kappa-1},u_{\kappa+1},\cdots,u_{d}. Since κ\kappa is arbitrary, by Lemma 2.3 the function f()f(\cdot) must be of form in (3.2). Next, we find the function g()g(\cdot) as follows.

Claim: gκ()g_{\kappa}(\cdot) is a function of one variable uκu_{\kappa}.

Indeed, after taking derivative zκ¯\partial_{\overline{z_{\kappa}}} in (3.7), the following is obtained

uκ(gκ(u))(1gκ(u)zκ¯)2=zκ(gκ(z))¯(1uκgκ(z)¯)2\displaystyle\dfrac{\partial_{u_{\kappa}}(g_{\kappa}(u))}{(1-g_{\kappa}(u)\overline{z_{\kappa}})^{2}}=\dfrac{\overline{\partial_{z_{\kappa}}(g_{\kappa}(z))}}{(1-u_{\kappa}\overline{g_{\kappa}(z)})^{2}}

or equivalently to saying

(3.9) uκ(gκ(u))(1uκgκ(z)¯)2=zκ(gκ(z))¯(1gκ(u)zκ¯)2.\displaystyle\partial_{u_{\kappa}}(g_{\kappa}(u))(1-u_{\kappa}\overline{g_{\kappa}(z)})^{2}=\overline{\partial_{z_{\kappa}}(g_{\kappa}(z))}(1-g_{\kappa}(u)\overline{z_{\kappa}})^{2}.

For every s1,d{κ}s\in\mathbb{Z}_{1,d}\setminus\{\kappa\}, we differentiate (3.9) with respect to the variable usu_{s}

(3.10) usuκ(gκ(u))(1uκgκ(z)¯)2=2zκ(gκ(z))¯(1gκ(u)zκ¯)zκ¯us(gκ(u)).\displaystyle\partial_{u_{s}}\circ\partial_{u_{\kappa}}(g_{\kappa}(u))(1-u_{\kappa}\overline{g_{\kappa}(z)})^{2}=-2\overline{\partial_{z_{\kappa}}(g_{\kappa}(z))}(1-g_{\kappa}(u)\overline{z_{\kappa}})\overline{z_{\kappa}}\partial_{u_{s}}(g_{\kappa}(u)).

If there is u𝔻du_{\star}\in\mathbb{D}^{d} with uκ(gκ(u))=0\partial_{u_{\kappa}}(g_{\kappa}(u_{\star}))=0, then (3.9) gives

zκ(gκ(z))=0z𝔻d,\displaystyle\partial_{z_{\kappa}}(g_{\kappa}(z))=0\quad\forall z\in\mathbb{D}^{d},

and hence by (3.10) we have

usuκ(gκ(u))=0u𝔻d;\displaystyle\partial_{u_{s}}\circ\partial_{u_{\kappa}}(g_{\kappa}(u))=0\quad\forall u\in\mathbb{D}^{d};

meaning that gκ()g_{\kappa}(\cdot) is a function of one variable uκu_{\kappa}. Now consider the case when uκgκ()0\partial_{u_{\kappa}}g_{\kappa}(\cdot)\not\equiv 0. Equation (3.10) divided by (3.9) is equal to

usuκ(gκ(u))uκ(gκ(u))=2zκ¯us(gκ(u))1gκ(u)zκ¯,\displaystyle\dfrac{\partial_{u_{s}}\circ\partial_{u_{\kappa}}(g_{\kappa}(u))}{\partial_{u_{\kappa}}(g_{\kappa}(u))}=-\dfrac{2\overline{z_{\kappa}}\partial_{u_{s}}(g_{\kappa}(u))}{1-g_{\kappa}(u)\overline{z_{\kappa}}},

which implies, by way of equating coefficients of zκ¯\overline{z_{\kappa}}, that

usuκ(gκ(u))=0=us(gκ(u)).\displaystyle\partial_{u_{s}}\circ\partial_{u_{\kappa}}(g_{\kappa}(u))=0=\partial_{u_{s}}(g_{\kappa}(u)).

The claim is proven.

Suppose that gκ(u)=ψ(uκ)g_{\kappa}(u)=\psi(u_{\kappa}) for some function ψ:𝔻\psi:\mathbb{D}\to\mathbb{C}. Thus, (3.9) is simplified to (2.2). By Lemma 2.4, the function gκ()g_{\kappa}(\cdot) is of the desired form.

(2) Since the function ff is bounded, it is enough to show that the composition operator Cg,maxC_{g,\max} is bounded. Before proving this, we denote

Ω={j1,d:bj=0},Δ={j1,d:bj0}.\displaystyle\Omega=\{j\in\mathbb{Z}_{1,d}:b_{j}=0\},\quad\Delta=\{j\in\mathbb{Z}_{1,d}:b_{j}\neq 0\}.

Suppose that n1,n2,,ntn_{1},n_{2},\cdots,n_{t} and m1,m2,,msm_{1},m_{2},\cdots,m_{s} are elements of Ω,Δ\Omega,\Delta, respectively; meaning

Ω={n1,n2,,nt},Δ={m1,m2,,ms}.\displaystyle\Omega=\{n_{1},n_{2},\cdots,n_{t}\},\quad\Delta=\{m_{1},m_{2},\cdots,m_{s}\}.

For each z=(z1,z2,,zd)dz=(z_{1},z_{2},\cdots,z_{d})\in\mathbb{C}^{d}, we express z=(zΩ,zΔ)z=(z_{\Omega},z_{\Delta}), where zΩ=(zn1,zn2,,znt)z_{\Omega}=(z_{n_{1}},z_{n_{2}},\cdots,z_{n_{t}}) and zΔ=(zm1,zm2,,zms)z_{\Delta}=(z_{m_{1}},z_{m_{2}},\cdots,z_{m_{s}}). Then g(z)=(𝐚Ω,gΔ(zΔ))g(z)=(\mathbf{a}_{\Omega},g_{\Delta}(z_{\Delta})) and

Cg,maxh(z)=h(𝐚Ω,gΔ(zΔ))=h𝐚Ω(gΔ(zΔ))=CgΔ,maxh𝐚Ω(zΔ).\displaystyle C_{g,\max}h(z)=h(\mathbf{a}_{\Omega},g_{\Delta}(z_{\Delta}))=h_{\mathbf{a}_{\Omega}}(g_{\Delta}(z_{\Delta}))=C_{g_{\Delta},\max}h_{\mathbf{a}_{\Omega}}(z_{\Delta}).

Note that the Jacobian determinant of gΔg_{\Delta} is

J=jΔ|bj|2|1aj¯zj|40,\displaystyle J=\prod_{j\in\Delta}\dfrac{|b_{j}|^{2}}{|1-\overline{a_{j}}z_{j}|^{4}}\neq 0,

so by [20, Theorem 10], the operator CgΔ,maxC_{g_{\Delta},\max} is bounded and together with Lemma 2.1, we get the boundedness of the operator Cg,maxC_{g,\max}. ∎

With all preparation in place we give one of the main results of this section. The following theorem provides a useful criteria to determine whether a maximal weighted composition operator is real symmetric or not.

Theorem 3.2.

Let d1d\in\mathbb{Z}_{\geq 1} and 0d\ell\in\mathbb{Z}_{\geq 0}^{d}. Let f:𝔻d,g:𝔻d𝔻df:\mathbb{D}^{d}\to\mathbb{C},g:\mathbb{D}^{d}\to\mathbb{D}^{d} be analytic functions. Then the operator Wf,g,maxW_{f,g,\max} is real symmetric on (𝔻d)\mathcal{B}_{\ell}(\mathbb{D}^{d}) if and only if the functions f(),g()f(\cdot),g(\cdot) are of forms in (3.2), where coefficients verify (3.3)-(3.4). In this case, the operator Wf,g,maxW_{f,g,\max} is bounded.

Proof.

Suppose that the operator Wf,g,maxW_{f,g,\max} is real symmetric on (𝔻d)\mathcal{B}_{\ell}(\mathbb{D}^{d}), which gives Wf,g,max=Wf,g,maxW_{f,g,\max}^{*}=W_{f,g,\max}. In particular, the following is obtained

(3.11) Wf,g,maxKz=Wf,g,maxKzz𝔻d.\displaystyle W_{f,g,\max}^{*}K_{z}=W_{f,g,\max}K_{z}\quad\forall z\in\mathbb{D}^{d}.

By Lemma 2.9, (3.11) becomes (3.1) and Lemma 3.1 gives the necessary condition.

For the sufficient condition, take f(),g()f(\cdot),g(\cdot) as in the statement of the theorem. Lemma 3.1 shows that the operator Wf,g,maxW_{f,g,\max} is bounded. By (3.2)-(3.4) and Lemma 2.9, the operator verifies (3.11) and so it must be real symmetric. ∎

The following result shows that the real symmetry cannot be separated from the maximal domain; in other words, a real symmetric weighted composition operator must be maximal.

Theorem 3.3.

Let d1d\in\mathbb{Z}_{\geq 1} and 0d\ell\in\mathbb{Z}_{\geq 0}^{d}. Let f:𝔻d,g:𝔻d𝔻df:\mathbb{D}^{d}\to\mathbb{C},g:\mathbb{D}^{d}\to\mathbb{D}^{d} be analytic functions. Then the operator Wf,gW_{f,g} is real symmetric on (𝔻d)\mathcal{B}_{\ell}(\mathbb{D}^{d}) if and only if it verifies two assertions.

  1. (1)

    (3.2)-(3.4) hold.

  2. (2)

    The operator Wf,gW_{f,g} is maximal; that is Wf,g=Wf,g,maxW_{f,g}=W_{f,g,\max}.

In this case, the operator Wf,gW_{f,g} is bounded.

Proof.

The implication ``"``\Longleftarrow" is proven in Theorem 3.2 and the remaing task is to prove the implication ``"``\Longrightarrow". Indeed, since Wf,gWf,g,maxW_{f,g}\preceq W_{f,g,\max}, by [26, Proposition 1.6], we have

Wf,g,maxWf,g=Wf,gWf,g,max.W_{f,g,\max}^{*}\preceq W_{f,g}^{*}=W_{f,g}\preceq W_{f,g,\max}.

Lemma 2.9 shows that Kzdom(Wf,g,max)K_{z}\in\text{dom}(W_{f,g,\max}^{*}), and so,

Wf,g,maxKz(u)=Wf,g,maxKz(u)z,u𝔻d.W_{f,g,\max}^{*}K_{z}(u)=W_{f,g,\max}K_{z}(u)\quad\forall z,u\in\mathbb{D}^{d}.

By Lemmas 2.9 and 3.1, conditions (3.2)-(3.4) hold, and hence, by Theorem 3.2, the operator Wf,g,max=Wf,g,maxW_{f,g,\max}=W_{f,g,\max}^{*}. Using this equality, item (2) is proven as follows

Wf,gWf,g,max=Wf,g,maxWf,g=Wf,g.W_{f,g}\preceq W_{f,g,\max}=W_{f,g,\max}^{*}\preceq W_{f,g}^{*}=W_{f,g}.

4. Unitary property

Recall that a bounded linear operator QQ is called unitary if the equality QQ=QQ=IQQ^{*}=Q^{*}Q=I holds. In this section, we describe all weighted composition operators that are unitary. The following lemma gives a partial characterization of the operator Wf,g,maxW_{f,g,\max} under the assumption that the symbol g()g(\cdot) fixes 0.

Lemma 4.1.

Let d1d\in\mathbb{Z}_{\geq 1} and 0d\ell\in\mathbb{Z}_{\geq 0}^{d}. Let f:𝔻d,g=(g1,g2,,gd):𝔻d𝔻df:\mathbb{D}^{d}\to\mathbb{C},g=(g_{1},g_{2},\cdots,g_{d}):\mathbb{D}^{d}\to\mathbb{D}^{d} be analytic functions. Suppose that

(4.1) f(z)¯f(u)Kg(z)(g(u))=Kz(u)z,u𝔻d.\displaystyle\overline{f(z)}f(u)K_{g(z)}(g(u))=K_{z}(u)\quad\forall z,u\in\mathbb{D}^{d}.

If g(0)=0g(0)=0, then

(4.2) f()𝐜,gκ(z)=𝐚κzϕ(κ),\displaystyle f(\cdot)\equiv\mathbf{c},\quad g_{\kappa}(z)=\mathbf{a}_{\kappa}z_{\phi(\kappa)},

where

(4.3) ϕ:1,d1,d is bijective and𝐜,𝐚κ𝕋.\displaystyle\text{$\phi:\mathbb{Z}_{1,d}\to\mathbb{Z}_{1,d}$ is bijective and}\quad\mathbf{c},\mathbf{a}_{\kappa}\in\mathbb{T}.
Proof.

Setting z=0z=0 in (4.1), we find f(0)¯f(u)=1\overline{f(0)}f(u)=1, which means f()𝐜f(\cdot)\equiv\mathbf{c}, where 𝐜𝕋\mathbf{c}\in\mathbb{T}. Consequently, taking into account the explicit form of kernel functions, (4.1) is reduced to the following

(4.4) j=1d(1gj(u)gj(z)¯)j+2=j=1d(1ujzj¯)j+2.\displaystyle\prod_{j=1}^{d}(1-g_{j}(u)\overline{g_{j}(z)})^{\ell_{j}+2}=\prod_{j=1}^{d}(1-u_{j}\overline{z_{j}})^{\ell_{j}+2}.

Let κ1,d\kappa\in\mathbb{Z}_{1,d}. Taking derivative uκ\partial_{u_{\kappa}}, (4.4) becomes

t=1d(t+2)(1gt(u)gt(z)¯)t+1uκ(gt(u))gt(z)¯jt(1gj(u)gj(z)¯)j+2\displaystyle-\sum_{t=1}^{d}(\ell_{t}+2)(1-g_{t}(u)\overline{g_{t}(z)})^{\ell_{t}+1}\partial_{u_{\kappa}}(g_{t}(u))\overline{g_{t}(z)}\prod_{j\neq t}(1-g_{j}(u)\overline{g_{j}(z)})^{\ell_{j}+2}
(4.5) =(κ+2)zκ¯(1uκzκ¯)κ+1jκ(1ujzj¯)j+2.\displaystyle=-(\ell_{\kappa}+2)\overline{z_{\kappa}}(1-u_{\kappa}\overline{z_{\kappa}})^{\ell_{\kappa}+1}\prod_{j\neq\kappa}(1-u_{j}\overline{z_{j}})^{\ell_{j}+2}.

Equation (4.5) divided by (4.4) is equal to

t=1d(t+2)uκ(gt(u))gt(z)¯1gt(u)gt(z)¯=(κ+2)zκ¯1uκzκ¯.\displaystyle\sum_{t=1}^{d}(\ell_{t}+2)\dfrac{\partial_{u_{\kappa}}(g_{t}(u))\overline{g_{t}(z)}}{1-g_{t}(u)\overline{g_{t}(z)}}=(\ell_{\kappa}+2)\dfrac{\overline{z_{\kappa}}}{1-u_{\kappa}\overline{z_{\kappa}}}.

In particular with u=0u=0, we get

(4.6) t=1d(t+2)uκ(gt(0))gt(z)¯=(κ+2)zκ¯κ1,d.\displaystyle\sum_{t=1}^{d}(\ell_{t}+2)\partial_{u_{\kappa}}(g_{t}(0))\overline{g_{t}(z)}=(\ell_{\kappa}+2)\overline{z_{\kappa}}\quad\forall\kappa\in\mathbb{Z}_{1,d}.

Let

𝔸=((1+2)u1(g1(0))¯(2+2)u1(g2(0))¯(d+2)u1(gd(0))¯(1+2)u2(g1(0))¯(2+2)u2(g2(0))¯(d+2)u2(gd(0))¯(1+2)ud(g1(0))¯(2+2)ud(g2(0))¯(d+2)ud(gd(0))¯)\displaystyle\mathbb{A}=\begin{pmatrix}(\ell_{1}+2)\overline{\partial_{u_{1}}(g_{1}(0))}&(\ell_{2}+2)\overline{\partial_{u_{1}}(g_{2}(0))}&\cdots&(\ell_{d}+2)\overline{\partial_{u_{1}}(g_{d}(0))}\\ &&&\\ (\ell_{1}+2)\overline{\partial_{u_{2}}(g_{1}(0))}&(\ell_{2}+2)\overline{\partial_{u_{2}}(g_{2}(0))}&\cdots&(\ell_{d}+2)\overline{\partial_{u_{2}}(g_{d}(0))}\\ \vdots&\vdots&\ddots&\vdots\\ (\ell_{1}+2)\overline{\partial_{u_{d}}(g_{1}(0))}&(\ell_{2}+2)\overline{\partial_{u_{d}}(g_{2}(0))}&\cdots&(\ell_{d}+2)\overline{\partial_{u_{d}}(g_{d}(0))}\end{pmatrix}

and 𝔹\mathbb{B} be the diagonal matrix given by

𝔹=diag(1+2,2+2,,d+2).\displaystyle\mathbb{B}=\text{diag}\left(\ell_{1}+2,\ell_{2}+2,\cdots,\ell_{d}+2\right).

Now equation (4.6) is rewritten as 𝔸g(z)=𝔹z\mathbb{A}g(z)=\mathbb{B}z, which gives g(z)=zg(z)=\mathbb{H}z, where =(hi,j)d×d=𝔸1𝔹\mathbb{H}=(h_{i,j})_{d\times d}=\mathbb{A}^{-1}\mathbb{B}.

Fix p1,dp\in\mathbb{Z}_{1,d}. Setting z=(0,,0,zp,0,,0)z=(0,\ldots,0,z_{p},0,\ldots,0) in (4.4), where zpz_{p} is the pp-th coordinate, we obtain

(4.7) j=1d(1gj(u)hj,pzp¯)j+2=(1upzp¯)p+2.\displaystyle\prod_{j=1}^{d}\left(1-g_{j}(u)\overline{h_{j,p}z_{p}}\right)^{\ell_{j}+2}=(1-u_{p}\overline{z_{p}})^{\ell_{p}+2}.

We continue with choosing u=(0,,0,up,0,,0)u=(0,\ldots,0,u_{p},0,\ldots,0), where upu_{p} is the pp-th coordinate, and then (4.7) is reduced to the following

j=1d(1|hj,p|2upzp¯)j+2=(1upzp¯)p+2.\displaystyle\prod_{j=1}^{d}(1-|h_{j,p}|^{2}u_{p}\overline{z_{p}})^{\ell_{j}+2}=(1-u_{p}\overline{z_{p}})^{\ell_{p}+2}.

Hence, since up,zp𝔻u_{p},z_{p}\in\mathbb{D} are arbitrary, coefficients hj,ph_{j,p} verify

(4.8) j=1d(1|hj,p|2x)j+2=(1x)p+2x𝔻.\displaystyle\prod_{j=1}^{d}(1-|h_{j,p}|^{2}x)^{\ell_{j}+2}=(1-x)^{\ell_{p}+2}\quad\forall x\in\mathbb{D}.

The highest power of the left-hand side is j=1d(j+2)\sum\limits_{j=1}^{d}(\ell_{j}+2); meanwhile of the right-hand side is p+2\ell_{p}+2. Thus, the coefficient of xj=1d(j+2)x^{\sum\limits_{j=1}^{d}(\ell_{j}+2)} in (4.8) must be 0; meaning that

j=1d|hj,p|2=0.\displaystyle\prod_{j=1}^{d}|h_{j,p}|^{2}=0.

Denote

Vp={j1,d:hj,p=0},Up={j1,d:hj,p0}.\displaystyle V_{p}=\{j\in\mathbb{Z}_{1,d}:h_{j,p}=0\},\quad U_{p}=\{j\in\mathbb{Z}_{1,d}:h_{j,p}\neq 0\}.

If Up=U_{p}=\emptyset, then hj,p=0h_{j,p}=0 for every j1,dj\in\mathbb{Z}_{1,d}; but this is impossible as the matrix \mathbb{H} is invertible. Now consider the situation when UpU_{p}\neq\emptyset. Setting u=wp:=(w1,,wp1,0,wp+1,,wd)u=w_{-p}:=(w_{1},\ldots,w_{p-1},0,w_{p+1},\ldots,w_{d}) in (4.7), where 0 is the pp-th coordinate, we get

j=1d(1gj(wp)hj,pzp¯)j+2=1.\displaystyle\prod_{j=1}^{d}\left(1-g_{j}(w_{-p})\overline{h_{j,p}z_{p}}\right)^{\ell_{j}+2}=1.

Through equating coefficients of zp¯\overline{z_{p}} and then using the fact that hj,p0h_{j,p}\neq 0 for jUpj\in U_{p}, we have

gj(wp)=0hj,t=0jUp,t1,d{p},\displaystyle g_{j}(w_{-p})=0\Longrightarrow h_{j,t}=0\quad\forall j\in U_{p},t\in\mathbb{Z}_{1,d}\setminus\{p\},

and so

(4.9) gj(u1,u2,,ud)=hj,pupjUp.\displaystyle g_{j}(u_{1},u_{2},\cdots,u_{d})=h_{j,p}u_{p}\quad\forall j\in U_{p}.

Note that equation (4.8) can be rewritten in the following form

jUp(1|hj,p|2x)j+2=(1x)p+2,\displaystyle\prod_{j\in U_{p}}(1-|h_{j,p}|^{2}x)^{\ell_{j}+2}=(1-x)^{\ell_{p}+2},

which implies, after equating coefficients of xp+2x^{\ell_{p}+2}, that

jUp|hj,p|j+2=1.\displaystyle\prod_{j\in U_{p}}|h_{j,p}|^{\ell_{j}+2}=1.

Since g(z)=zg(z)=\mathbb{H}z is a self-mapping of 𝔻d\mathbb{D}^{d}, we have

|hj,p|1jUp,p1,djUp|hj,p|j+21p1,d.\displaystyle|h_{j,p}|\leq 1\quad\forall j\in U_{p},p\in\mathbb{Z}_{1,d}\quad\Longrightarrow\quad\prod_{j\in U_{p}}|h_{j,p}|^{\ell_{j}+2}\leq 1\quad\forall p\in\mathbb{Z}_{1,d}.

Thus,

|hj,p|=1jUp,p1,d.\displaystyle|h_{j,p}|=1\quad\forall j\in U_{p},p\in\mathbb{Z}_{1,d}.

Claim: For every p1,dp\in\mathbb{Z}_{1,d}, UpU_{p} is a singleton set.

Since Up,p1,dU_{p},p\in\mathbb{Z}_{1,d} are subsets of 1,d\mathbb{Z}_{1,d}, it is enough to show that the family {Up:p1,d}\{U_{p}:p\in\mathbb{Z}_{1,d}\} consists of disjoint sets. Indeed, assume in contrary that there exist p,q1,dp,q\in\mathbb{Z}_{1,d} with pqp\neq q such that UpUqU_{p}\cap U_{q}\neq\emptyset. Let tUpUqt\in U_{p}\cap U_{q}. It follows from (4.9) that

gt(u)=ht,pup=ht,quq;\displaystyle g_{t}(u)=h_{t,p}u_{p}=h_{t,q}u_{q};

but this is impossible as |ht,p|=1=|ht,q||h_{t,p}|=1=|h_{t,q}| and pqp\neq q.

Let us define the map η:1,d1,d\eta:\mathbb{Z}_{1,d}\to\mathbb{Z}_{1,d} by setting η(p)=j\eta(p)=j, where j,p1,dj,p\in\mathbb{Z}_{1,d} satisfy (4.9). Then η\eta is bijective and (4.3) holds, where ϕ=η1\phi=\eta^{-1}. ∎

With the help of Lemma 4.1, we give a complete characterization of unitary weighted composition operators.

Theorem 4.2.

Let d1d\in\mathbb{Z}_{\geq 1} and 0d\ell\in\mathbb{Z}_{\geq 0}^{d}. Let f:𝔻d,g:𝔻d𝔻df:\mathbb{D}^{d}\to\mathbb{C},g:\mathbb{D}^{d}\to\mathbb{D}^{d} be analytic functions. Then the operator Wf,g,maxW_{f,g,\max} is unitary on (𝔻d)\mathcal{B}_{\ell}(\mathbb{D}^{d}) if and only if

(4.10) f(z)=𝐜j=1d(1|θj|2)1+j/2(1𝐚jθj¯zϕ(j))j+2,gκ(z)=𝐚κ(𝐚κ¯θκzϕ(κ))1𝐚κθκ¯zϕ(κ),\displaystyle f(z)=\mathbf{c}\prod_{j=1}^{d}\dfrac{(1-|\theta_{j}|^{2})^{1+\ell_{j}/2}}{(1-\mathbf{a}_{j}\overline{\theta_{j}}z_{\phi(j)})^{\ell_{j}+2}},\quad g_{\kappa}(z)=\dfrac{\mathbf{a}_{\kappa}(\overline{\mathbf{a}_{\kappa}}\theta_{\kappa}-z_{\phi(\kappa)})}{1-\mathbf{a}_{\kappa}\overline{\theta_{\kappa}}z_{\phi(\kappa)}},

where coefficients satisfy (4.3) and

(4.11) θ=(θ1,θ2,,θd)𝔻d.\displaystyle\theta=(\theta_{1},\theta_{2},\cdots,\theta_{d})\in\mathbb{D}^{d}.
Proof.

Note that the sufficient condition can be obtained by arguments similar to those used in Proposition 2.10. We prove the necessary condition as follows. Suppose that the operator Wf,g,maxW_{f,g,\max} is unitary on (𝔻d)\mathcal{B}_{\ell}(\mathbb{D}^{d}). Put θ=g(0)\theta=g(0) and define

F=fψθg,G=φθg.\displaystyle F=f\cdot\psi_{\theta}\circ g,\quad G=\varphi_{\theta}\circ g.

A direct computation gives [F,G]~=[f,g]~[ψθ,φθ]~\widetilde{[F,G]}=\widetilde{[f,g]}\widetilde{[\psi_{\theta},\varphi_{\theta}]} and so we get

WF,G,max=Wf,g,max𝒜θ,max.\displaystyle W_{F,G,\max}=W_{f,g,\max}\mathcal{A}_{\theta,\max}.

Using Proposition 2.10, we can show that the operator WF,G,maxW_{F,G,\max} is unitary with G(0)=φθ(g(0))=0G(0)=\varphi_{\theta}(g(0))=0. We have

Kz=WF,G,maxWF,G,maxKz=WF,G,max(F(z)¯KG(z))\displaystyle K_{z}=W_{F,G,\max}W_{F,G,\max}^{*}K_{z}=W_{F,G,\max}\left(\overline{F(z)}K_{G(z)}\right)
=F(z)¯FKG(z)G\displaystyle=\overline{F(z)}F\cdot K_{G(z)}\circ G
Kz(u)=F(z)¯F(u)KG(z)(G(u))z,u𝔻d.\displaystyle\Longrightarrow\quad K_{z}(u)=\overline{F(z)}F(u)K_{G(z)}(G(u))\quad\forall z,u\in\mathbb{D}^{d}.

The line above allows us to use Lemma 4.1 and the proof is complete. ∎

5. Complex symmetry

The section studies which the symbols give rise to weighted composition operators that are complex symmetric with respect to the conjugation 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}}. Such operators are called 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}}-symmetric. Like as the real symmetry, a 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}}-symmetric weighted composition operator must be bounded. As a byproduct, we obtain the interesting fact that real symmetric weighted composition operators are 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}}-symmetric corresponding an adapted and highly relevant selection of the parameters 𝐩,𝐪\mathbf{p},\mathbf{q}.

To study the necessary condition for a weighted composition operator to be 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}}-symmetric, we apply the symmetric condition to kernel functions. It turns out that the symbols generating such operators can be precisely computed.

Lemma 5.1.

Let d1,0dd\in\mathbb{Z}_{\geq 1},\ell\in\mathbb{Z}_{\geq 0}^{d}. Let U1,U2U_{1},U_{2} with condition (2.17) and 𝐩𝔻|U1|{0},𝐪𝕋|U2|\mathbf{p}\in\mathbb{D}^{|U_{1}|}\setminus\{0\},\mathbf{q}\in\mathbb{T}^{|U_{2}|}. Let 𝐫𝔻d\mathbf{r}\in\mathbb{D}^{d}, where

𝐫t={𝐩tif tU1,0if tU2.\displaystyle\mathbf{r}_{t}=\begin{cases}\mathbf{p}_{t}\quad\text{if $t\in U_{1}$},\\ 0\quad\text{if $t\in U_{2}$}.\end{cases}

Let f:𝔻d,g=(g1,g2,,gd):𝔻d𝔻df:\mathbb{D}^{d}\to\mathbb{C},g=(g_{1},g_{2},\cdots,g_{d}):\mathbb{D}^{d}\to\mathbb{D}^{d} be analytic functions. Suppose that

(5.1) ϑ𝐩,U(z)¯ϑ𝐩,U(g(ω(z)¯))f(ω(z)¯)Kω(g(ω(z)¯))¯(u)=f(u)Kz(g(u))z,u𝔻d.\displaystyle\overline{\vartheta_{\mathbf{p},U}(z)}\vartheta_{\mathbf{p},U}(g(\overline{\omega(z)}))f(\overline{\omega(z)})K_{\overline{\omega(g(\overline{\omega(z)}))}}(u)=f(u)K_{z}(g(u))\quad\forall z,u\in\mathbb{D}^{d}.

Then the following assertions hold.

  1. (1)

    The functions are of forms

    (5.2) f(u)=c~Kω(g(𝐫))¯(u),gκ(uκ)={Gκ+Eκuκ+Fκif κU1ακ+βκ𝐪κuκ1ακ𝐪κuκif κU2,\displaystyle f(u)=\widetilde{c}K_{\overline{\omega(g(\mathbf{r}))}}(u),\quad g_{\kappa}(u_{\kappa})=\begin{cases}G_{\kappa}+\dfrac{E_{\kappa}}{u_{\kappa}+F_{\kappa}}\quad\text{if $\kappa\in U_{1}$}\\ \\ \alpha_{\kappa}+\dfrac{\beta_{\kappa}\mathbf{q}_{\kappa}u_{\kappa}}{1-\alpha_{\kappa}\mathbf{q}_{\kappa}u_{\kappa}}\quad\text{if $\kappa\in U_{2}$},\end{cases}

    where coefficients satisfy

    (5.3) Gκ𝔻if κU1,Eκ=0,\displaystyle G_{\kappa}\in\mathbb{D}\quad\text{if $\kappa\in U_{1},E_{\kappa}=0$},
    (5.4) pGκ|pκ|2=Fκ|pκ|2+(GκFκ+Eκ)pκ¯if κU1,Eκ0,\displaystyle p-G_{\kappa}|p_{\kappa}|^{2}=-F_{\kappa}|p_{\kappa}|^{2}+(G_{\kappa}F_{\kappa}+E_{\kappa})\overline{p_{\kappa}}\quad\text{if $\kappa\in U_{1},E_{\kappa}\neq 0$},
    (5.5) |(GκFκ+Eκ)Fκ¯Gκ|+|Eκ||Fκ|21if κU1,Eκ0,\displaystyle|(G_{\kappa}F_{\kappa}+E_{\kappa})\overline{F_{\kappa}}-G_{\kappa}|+|E_{\kappa}|\leq|F_{\kappa}|^{2}-1\quad\text{if $\kappa\in U_{1},E_{\kappa}\neq 0$},
    (5.6) ακ𝔻if κU2,βκ=0,\displaystyle\alpha_{\kappa}\in\mathbb{D}\quad\text{if $\kappa\in U_{2},\beta_{\kappa}=0$},
    (5.7) |ακ+ακ¯(βκακ2)|+|βκ|1|ακ|2if κU2,βκ0.\displaystyle|\alpha_{\kappa}+\overline{\alpha_{\kappa}}(\beta_{\kappa}-\alpha_{\kappa}^{2})|+|\beta_{\kappa}|\leq 1-|\alpha_{\kappa}|^{2}\quad\text{if $\kappa\in U_{2},\beta_{\kappa}\neq 0$}.
  2. (2)

    If the functions are given in item (1), then Wf,g,maxW_{f,g,\max} is bounded on (𝔻d)\mathcal{B}_{\ell}(\mathbb{D}^{d}).

Proof.

(1) Setting z=ω(y)¯z=\overline{\omega(y)} in (5.1), we find

ϑ𝐩,U(ω(y)¯)¯ϑ𝐩,U(g(y))f(y)Kω(g(y))¯(u)=f(u)Kω(y)¯(g(u))y,u𝔻d.\displaystyle\overline{\vartheta_{\mathbf{p},U}(\overline{\omega(y)})}\vartheta_{\mathbf{p},U}(g(y))f(y)K_{\overline{\omega(g(y))}}(u)=f(u)K_{\overline{\omega(y)}}(g(u))\quad\forall y,u\in\mathbb{D}^{d}.

Consequently, taking into account the explicit form of kernel functions, we get

f(y)jU1(1|𝐩j|2)j+2j=1d[1gj(u)ωj(yj)]j+2\displaystyle f(y)\prod_{j\in U_{1}}(1-|\mathbf{p}_{j}|^{2})^{\ell_{j}+2}\prod_{j=1}^{d}[1-g_{j}(u)\omega_{j}(y_{j})]^{\ell_{j}+2}
(5.8) =f(u)jU1(1𝐩j¯gj(y))j+2(1𝐩jωj(yj))j+2j=1d[1ujωj(gj(y))]j+2.\displaystyle=f(u)\prod_{j\in U_{1}}(1-\overline{\mathbf{p}_{j}}g_{j}(y))^{\ell_{j}+2}(1-\mathbf{p}_{j}\omega_{j}(y_{j}))^{\ell_{j}+2}\prod_{j=1}^{d}[1-u_{j}\omega_{j}(g_{j}(y))]^{\ell_{j}+2}.

Let κ1,d\kappa\in\mathbb{Z}_{1,d}. After differentiating with respect to the variable uκu_{\kappa}, the equation above becomes

f(y)jU1(1|𝐩j|2)j+2t=1d(t+2)[1gt(u)ωt(yt)]t+1ωt(yt)uκ(gt(u))\displaystyle-f(y)\prod_{j\in U_{1}}(1-|\mathbf{p}_{j}|^{2})^{\ell_{j}+2}\sum_{t=1}^{d}(\ell_{t}+2)[1-g_{t}(u)\omega_{t}(y_{t})]^{\ell_{t}+1}\omega_{t}(y_{t})\partial_{u_{\kappa}}(g_{t}(u))
×jt[1gj(u)ωj(yj)]j+2\displaystyle\times\prod_{j\neq t}[1-g_{j}(u)\omega_{j}(y_{j})]^{\ell_{j}+2}
=uκ(f(u))jU1(1𝐩j¯gj(y))j+2(1𝐩jωj(yj))j+2j=1d[1ujωj(gj(y))]j+2\displaystyle=\partial_{u_{\kappa}}(f(u))\prod_{j\in U_{1}}(1-\overline{\mathbf{p}_{j}}g_{j}(y))^{\ell_{j}+2}(1-\mathbf{p}_{j}\omega_{j}(y_{j}))^{\ell_{j}+2}\prod_{j=1}^{d}[1-u_{j}\omega_{j}(g_{j}(y))]^{\ell_{j}+2}
f(u)jU1(1𝐩j¯gj(y))j+2(1𝐩jωj(yj))j+2\displaystyle-f(u)\prod_{j\in U_{1}}(1-\overline{\mathbf{p}_{j}}g_{j}(y))^{\ell_{j}+2}(1-\mathbf{p}_{j}\omega_{j}(y_{j}))^{\ell_{j}+2}
(5.9) ×(κ+2)[1uκωκ(gκ(y))]κ+1ωκ(gκ(y))jκ[1ujωj(gj(y))]j+2.\displaystyle\times(\ell_{\kappa}+2)[1-u_{\kappa}\omega_{\kappa}(g_{\kappa}(y))]^{\ell_{\kappa}+1}\omega_{\kappa}(g_{\kappa}(y))\prod_{j\neq\kappa}[1-u_{j}\omega_{j}(g_{j}(y))]^{\ell_{j}+2}.

Equation (5.9) divided by (5.8) is equal to

(5.10) t=1d(t+2)ωt(yt)uκ(gt(u))1gt(u)ωt(yt)=uκ(f(u))f(u)(κ+2)ωκ(gκ(y))1uκωκ(gκ(y)).\displaystyle-\sum_{t=1}^{d}(\ell_{t}+2)\dfrac{\omega_{t}(y_{t})\partial_{u_{\kappa}}(g_{t}(u))}{1-g_{t}(u)\omega_{t}(y_{t})}=\dfrac{\partial_{u_{\kappa}}(f(u))}{f(u)}-\dfrac{(\ell_{\kappa}+2)\omega_{\kappa}(g_{\kappa}(y))}{1-u_{\kappa}\omega_{\kappa}(g_{\kappa}(y))}.

Setting y=𝐫y=\mathbf{r}, we observe

uκ(f(u))f(u)=(κ+2)ωκ(gκ(𝐫))1uκωκ(gκ(𝐫)).\displaystyle\dfrac{\partial_{u_{\kappa}}(f(u))}{f(u)}=\dfrac{(\ell_{\kappa}+2)\omega_{\kappa}(g_{\kappa}(\mathbf{r}))}{1-u_{\kappa}\omega_{\kappa}(g_{\kappa}(\mathbf{r}))}.

Using the arguments similar to (3.8), the function f()f(\cdot) is of form as in (5.2). Setting ξ=ωκgκ\xi=\omega_{\kappa}\circ g_{\kappa}, (5.10) is rewritten as

(5.11) t=1d(t+2)ωt(yt)uκ(gt(u))1gt(u)ωt(yt)=uκ(f(u))f(u)(κ+2)ξ(y)1uκξ(y).\displaystyle-\sum_{t=1}^{d}(\ell_{t}+2)\dfrac{\omega_{t}(y_{t})\partial_{u_{\kappa}}(g_{t}(u))}{1-g_{t}(u)\omega_{t}(y_{t})}=\dfrac{\partial_{u_{\kappa}}(f(u))}{f(u)}-\dfrac{(\ell_{\kappa}+2)\xi(y)}{1-u_{\kappa}\xi(y)}.

We differentiate (5.11) with respect to the variable yκy_{\kappa}, where j{1,2}j\in\{1,2\}, and then

(5.12) ωκ(yκ)uκ(gκ(u))(1ωκ(yκ)gκ(u))2=yκ(ξ(y))(1uκξ(y))2\displaystyle\dfrac{\omega_{\kappa}^{\odot}(y_{\kappa})\partial_{u_{\kappa}}(g_{\kappa}(u))}{(1-\omega_{\kappa}(y_{\kappa})g_{\kappa}(u))^{2}}=\dfrac{\partial_{y_{\kappa}}(\xi(y))}{(1-u_{\kappa}\xi(y))^{2}}

or equivalently to saying that

(5.13) ωκ(yκ)uκ(gκ(u))(1uκξ(y))2=yκ(ξ(y))(1ωκ(yκ)gκ(u))2,\displaystyle\omega_{\kappa}^{\odot}(y_{\kappa})\partial_{u_{\kappa}}(g_{\kappa}(u))(1-u_{\kappa}\xi(y))^{2}=\partial_{y_{\kappa}}(\xi(y))(1-\omega_{\kappa}(y_{\kappa})g_{\kappa}(u))^{2},

where we denote

ωκ(yκ)={𝐩κ¯𝐩κ|𝐩κ|21(1𝐩κ¯zκ)2if κU1,𝐪κif κU2.\displaystyle\omega_{\kappa}^{\odot}(y_{\kappa})=\begin{cases}\dfrac{\overline{\mathbf{p}_{\kappa}}}{\mathbf{p}_{\kappa}}\cdot\dfrac{|\mathbf{p}_{\kappa}|^{2}-1}{(1-\overline{\mathbf{p}_{\kappa}}z_{\kappa})^{2}}\quad\text{if $\kappa\in U_{1}$},\\ \mathbf{q}_{\kappa}\quad\text{if $\kappa\in U_{2}$}.\end{cases}

Claim 1: gκ()g_{\kappa}(\cdot) is a function of one variable uκu_{\kappa}.

Let s1,d{κ}s\in\mathbb{Z}_{1,d}\setminus\{\kappa\}. Taking derivative us\partial_{u_{s}} on both sides of (5.13) gives

ωκ(yκ)usuκ(gκ(u))(1uκξ(y))2\displaystyle\omega_{\kappa}^{\odot}(y_{\kappa})\partial_{u_{s}}\circ\partial_{u_{\kappa}}(g_{\kappa}(u))(1-u_{\kappa}\xi(y))^{2}
(5.14) =yκ(ξ(y))2(1ωκ(yκ)gκ(u))ωκ(yκ)us(gκ(u)).\displaystyle=-\partial_{y_{\kappa}}(\xi(y))2(1-\omega_{\kappa}(y_{\kappa})g_{\kappa}(u))\omega_{\kappa}(y_{\kappa})\partial_{u_{s}}(g_{\kappa}(u)).

If there is u𝔻du_{\star}\in\mathbb{D}^{d} for which uκ(gκ(u))=0\partial_{u_{\kappa}}(g_{\kappa}(u_{\star}))=0, then (5.13) gives

yκ(ξ(y))=0y𝔻d,\displaystyle\partial_{y_{\kappa}}(\xi(y))=0\quad\forall y\in\mathbb{D}^{d},

and hence by (5.14),

usuκ(gκ(u))=0u𝔻d;\displaystyle\partial_{u_{s}}\circ\partial_{u_{\kappa}}(g_{\kappa}(u))=0\quad\forall u\in\mathbb{D}^{d};

meaning that gκ()g_{\kappa}(\cdot) is a function of one variable uκu_{\kappa}. Now consider the situation when uκgκ()0\partial_{u_{\kappa}}g_{\kappa}(\cdot)\not\equiv 0. Equation (5.14) divided by (5.13) is equal to the following

usuκ(gκ(u))uκ(gκ(u))=2ωκ(yκ)us(gκ(u))1ωκ(yκ)gκ(u),\displaystyle\dfrac{\partial_{u_{s}}\circ\partial_{u_{\kappa}}(g_{\kappa}(u))}{\partial_{u_{\kappa}}(g_{\kappa}(u))}=-\dfrac{2\omega_{\kappa}(y_{\kappa})\partial_{u_{s}}(g_{\kappa}(u))}{1-\omega_{\kappa}(y_{\kappa})g_{\kappa}(u)},

which implies, after equating coefficients of ωκ(yκ)\omega_{\kappa}(y_{\kappa}), that

usuκ(gκ(u))=0=us(gκ(u));\displaystyle\partial_{u_{s}}\circ\partial_{u_{\kappa}}(g_{\kappa}(u))=0=\partial_{u_{s}}(g_{\kappa}(u));

meaning that gκ()g_{\kappa}(\cdot) is a function of one variable uκu_{\kappa}.

Claim 2: ξ()\xi(\cdot) is a function of one variable yκy_{\kappa}, too.

Let t1,d{κ}t\in\mathbb{Z}_{1,d}\setminus\{\kappa\}. We proceed into taking derivative yt\partial_{y_{t}} on both sides of (5.12)

0=yt(ωκ(yκ)uκ(gκ(u))(1ωκ(yκ)gκ(u))2)=yt(yκ(ξ(y))(1uκξ(y))2)\displaystyle 0=\partial_{y_{t}}\left(\dfrac{\omega_{\kappa}^{\odot}(y_{\kappa})\partial_{u_{\kappa}}(g_{\kappa}(u))}{(1-\omega_{\kappa}(y_{\kappa})g_{\kappa}(u))^{2}}\right)=\partial_{y_{t}}\left(\dfrac{\partial_{y_{\kappa}}(\xi(y))}{(1-u_{\kappa}\xi(y))^{2}}\right)
0=ytyκ(ξ(y))(1uκξ(y))+2uκyκ(ξ(y))yt(ξ(y)).\displaystyle\Longrightarrow 0=\partial_{y_{t}}\circ\partial_{y_{\kappa}}(\xi(y))(1-u_{\kappa}\xi(y))+2u_{\kappa}\partial_{y_{\kappa}}(\xi(y))\partial_{y_{t}}(\xi(y)).

After equating coefficients of uκu_{\kappa}, we get

ytyκ(ξ(y))=0=yκ(ξ(y))yt(ξ(y));\displaystyle\partial_{y_{t}}\circ\partial_{y_{\kappa}}(\xi(y))=0=\partial_{y_{\kappa}}(\xi(y))\cdot\partial_{y_{t}}(\xi(y));

meaning that ξ()\xi(\cdot) is a function of one variable yκy_{\kappa}. Thus, we make use of Lemma 2.5 (when κU2\kappa\in U_{2}) and Lemma 2.6 (when κU1\kappa\in U_{1}).

(2) The part is the same as those of Lemma 3.1(2) but we give a proof, for a completeness of exposition. Since the function f()f(\cdot) is bounded, it is enough to show that the composition operator Cg,maxC_{g,\max} is bounded. Before proving this, we fix some symbols used. Denote

Ω={j1,d:gjconst},Δ={jU1:gjconst},Ξ={jU2:gjconst}.\displaystyle\Omega=\{j\in\mathbb{Z}_{1,d}:g_{j}\equiv\text{const}\},\quad\Delta=\left\{j\in U_{1}:g_{j}\not\equiv\text{const}\right\},\quad\Xi=\left\{j\in U_{2}:g_{j}\not\equiv\text{const}\right\}.

Suppose that n1,n2,,ntn_{1},n_{2},\cdots,n_{t} and m1,m2,,msm_{1},m_{2},\cdots,m_{s} and c1,c2,,crc_{1},c_{2},\cdots,c_{r} are elements of Ω,Δ,Ξ\Omega,\Delta,\Xi, respectively; meaning

Ω={n1,n2,,nt},Δ={m1,m2,,ms},Ξ={c1,c2,,cr}.\displaystyle\Omega=\{n_{1},n_{2},\cdots,n_{t}\},\quad\Delta=\{m_{1},m_{2},\cdots,m_{s}\},\quad\Xi=\{c_{1},c_{2},\cdots,c_{r}\}.

For each z=(z1,z2,,zd)dz=(z_{1},z_{2},\cdots,z_{d})\in\mathbb{C}^{d}, we express z=(zΩ,zΔ,zΞ)z=(z_{\Omega},z_{\Delta},z_{\Xi}), where zΩ=(zn1,zn2,,znt)z_{\Omega}=(z_{n_{1}},z_{n_{2}},\cdots,z_{n_{t}}) and zΔ=(zm1,zm2,,zms)z_{\Delta}=(z_{m_{1}},z_{m_{2}},\cdots,z_{m_{s}}) and zΞ=(zc1,zc2,,zcr)z_{\Xi}=(z_{c_{1}},z_{c_{2}},\cdots,z_{c_{r}}). Then g(z)=(𝐚Ω,gΔ(zΔ),gΞ(zΞ))g(z)=(\mathbf{a}_{\Omega},g_{\Delta}(z_{\Delta}),g_{\Xi}(z_{\Xi})) and

Cg,maxh(z)=h(𝐚Ω,gΔ(zΔ),gΞ(zΞ))=h𝐚Ω(gΔ(zΔ),gΞ(zΞ))=C(gΔ,gΞ),maxh𝐚Ω(zΔ,zΞ).\displaystyle C_{g,\max}h(z)=h(\mathbf{a}_{\Omega},g_{\Delta}(z_{\Delta}),g_{\Xi}(z_{\Xi}))=h_{\mathbf{a}_{\Omega}}(g_{\Delta}(z_{\Delta}),g_{\Xi}(z_{\Xi}))=C_{(g_{\Delta},g_{\Xi}),\max}h_{\mathbf{a}_{\Omega}}(z_{\Delta},z_{\Xi}).

Note that the Jacobian determinant of (gΔ,gΞ)(g_{\Delta},g_{\Xi}) is

J=jΔ|Ej(zj+Fj)2|2κΞ|βκ(1ακ𝐪κzκ)2|20,\displaystyle J=\prod_{j\in\Delta}\left|\dfrac{E_{j}}{(z_{j}+F_{j})^{2}}\right|^{2}\cdot\prod_{\kappa\in\Xi}\left|\dfrac{\beta_{\kappa}}{(1-\alpha_{\kappa}\mathbf{q}_{\kappa}z_{\kappa})^{2}}\right|^{2}\neq 0,

so by [20, Theorem 10], the operator CgΔ,maxC_{g_{\Delta},\max} is bounded and together with Lemma 2.1, we get the boundedness of the operator Cg,maxC_{g,\max}. ∎

Lemma 5.1 provides a necessary condition for a weighted composition operator to be 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}}-symmetric. It turns out that the assertion in Lemma 5.1 is also a sufficient condition.

Theorem 5.2.

Let d1,0dd\in\mathbb{Z}_{\geq 1},\ell\in\mathbb{Z}_{\geq 0}^{d}. Let U1,U2U_{1},U_{2} with condition (2.17) and 𝐩𝔻|U1|{0},𝐪𝕋|U2|\mathbf{p}\in\mathbb{D}^{|U_{1}|}\setminus\{0\},\mathbf{q}\in\mathbb{T}^{|U_{2}|}. Suppose that f:𝔻d,g=(g1,g2,,gd):𝔻d𝔻df:\mathbb{D}^{d}\to\mathbb{C},g=(g_{1},g_{2},\cdots,g_{d}):\mathbb{D}^{d}\to\mathbb{D}^{d} are analytic functions. Then the operator Wf,g,maxW_{f,g,\max} is 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}}-symmetric on (𝔻d)\mathcal{B}_{\ell}(\mathbb{D}^{d}) if and only if the functions f(),g()f(\cdot),g(\cdot) are of forms in (5.2), where coefficients verify (5.3)-(5.7). In this case, the operator Wf,g,maxW_{f,g,\max} is bounded.

Proof.

Suppose that the operator Wf,g,maxW_{f,g,\max} is 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}}-symmetric on (𝔻d)\mathcal{B}_{\ell}(\mathbb{D}^{d}), which gives

𝒞𝐩,𝐪Wf,g,max𝒞𝐩,𝐪=Wf,g,max.\displaystyle\mathcal{C}_{\mathbf{p},\mathbf{q}}W_{f,g,\max}^{*}\mathcal{C}_{\mathbf{p},\mathbf{q}}=W_{f,g,\max}.

In particular, the following is obtained

(5.15) 𝒞𝐩,𝐪Wf,g,max𝒞𝐩,𝐪Kz=Wf,g,maxKzz𝔻d.\displaystyle\mathcal{C}_{\mathbf{p},\mathbf{q}}W_{f,g,\max}^{*}\mathcal{C}_{\mathbf{p},\mathbf{q}}K_{z}=W_{f,g,\max}K_{z}\quad\forall z\in\mathbb{D}^{d}.

By Lemma 2.9, (5.15) becomes (5.1) and so we can make use of Lemma 5.1 to get the necessary condition.

For the sufficient condition, take f(),g()f(\cdot),g(\cdot) as in the statement of the theorem. Lemma 5.1 shows that the operator Wf,g,maxW_{f,g,\max} is bounded. By (5.2)-(5.7) and Lemma 2.9, the operator verifies (5.15) and so it must be 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}}-symmetric. ∎

Theorem 5.2 characterizes maximal weighted composition operators that are 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}}-symmetric. The following theorem proves that the maximal domain and boundedness are consequences of the 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}}-symmetry.

Theorem 5.3.

Let d1,0dd\in\mathbb{Z}_{\geq 1},\ell\in\mathbb{Z}_{\geq 0}^{d}. Let U1,U2U_{1},U_{2} with condition (2.17) and 𝐩𝔻|U1|{0},𝐪𝕋|U2|\mathbf{p}\in\mathbb{D}^{|U_{1}|}\setminus\{0\},\mathbf{q}\in\mathbb{T}^{|U_{2}|}. Suppose that f:𝔻d,g=(g1,g2,,gd):𝔻d𝔻df:\mathbb{D}^{d}\to\mathbb{C},g=(g_{1},g_{2},\cdots,g_{d}):\mathbb{D}^{d}\to\mathbb{D}^{d} are analytic functions. Then the operator Wf,gW_{f,g} is 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}}-symmetric on (𝔻d)\mathcal{B}_{\ell}(\mathbb{D}^{d}) if and only if it verifies two assertions.

  1. (1)

    (5.3)-(5.7) hold.

  2. (2)

    The operator Wf,gW_{f,g} is maximal; that is Wf,g=Wf,g,maxW_{f,g}=W_{f,g,\max}.

In this case, the operator Wf,gW_{f,g} is bounded.

Proof.

The implication ``"``\Longleftarrow" is proven in Theorem 5.2 and the remaining task is to prove the implication ``"``\Longrightarrow". Indeed, since Wf,gWf,g,maxW_{f,g}\preceq W_{f,g,\max}, by [26, Proposition 1.6], we have

𝒞𝐩,𝐪Wf,g,max𝒞𝐩,𝐪𝒞𝐩,𝐪Wf,g𝒞𝐩,𝐪=Wf,gWf,g,max.\mathcal{C}_{\mathbf{p},\mathbf{q}}W_{f,g,\max}^{*}\mathcal{C}_{\mathbf{p},\mathbf{q}}\preceq\mathcal{C}_{\mathbf{p},\mathbf{q}}W_{f,g}^{*}\mathcal{C}_{\mathbf{p},\mathbf{q}}=W_{f,g}\preceq W_{f,g,\max}.

Lemma 2.9 shows that Kzdom(Wf,g,max)K_{z}\in\text{dom}(W_{f,g,\max}^{*}), and so,

𝒞𝐩,𝐪Wf,g,max𝒞𝐩,𝐪Kz(u)=Wf,g,maxKz(u)z,u𝔻d.\mathcal{C}_{\mathbf{p},\mathbf{q}}W_{f,g,\max}^{*}\mathcal{C}_{\mathbf{p},\mathbf{q}}K_{z}(u)=W_{f,g,\max}K_{z}(u)\quad\forall z,u\in\mathbb{D}^{d}.

By Lemmas 2.9 and 5.1, conditions (5.2)-(5.7) hold, and hence, by Theorem 5.2, the operator Wf,g,maxW_{f,g,\max} is 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}}-symmetric. Using this, item (2) is proven as follows

Wf,gWf,g,max=𝒞𝐩,𝐪Wf,g,max𝒞𝐩,𝐪𝒞𝐩,𝐪Wf,g𝒞𝐩,𝐪=Wf,g.W_{f,g}\preceq W_{f,g,\max}=\mathcal{C}_{\mathbf{p},\mathbf{q}}W_{f,g,\max}^{*}\mathcal{C}_{\mathbf{p},\mathbf{q}}\preceq\mathcal{C}_{\mathbf{p},\mathbf{q}}W_{f,g}^{*}\mathcal{C}_{\mathbf{p},\mathbf{q}}=W_{f,g}.

Corollary 5.4.

Let d1d\in\mathbb{Z}_{\geq 1} and 0d\ell\in\mathbb{Z}_{\geq 0}^{d}. Let f:𝔻d,g=(g1,g2,,gd):𝔻d𝔻df:\mathbb{D}^{d}\to\mathbb{C},g=(g_{1},g_{2},\cdots,g_{d}):\mathbb{D}^{d}\to\mathbb{D}^{d} be analytic functions. If the operator Wf,gW_{f,g} is real symmetric, then it is complex symmetric.

Proof.

Suppose that the operator Wf,gW_{f,g} is real symmetric. By Theorem 3.3, the functions f(),g()f(\cdot),g(\cdot) satisfy (3.2)-(3.4). It then is 𝒞𝐩,𝐪\mathcal{C}_{\mathbf{p},\mathbf{q}}-symmetric, where U2=1,dU_{2}=\mathbb{Z}_{1,d} and

ακ={𝐚κif 𝐚κ0,0if 𝐚κ=0,𝐪κ={𝐚κ¯𝐚κif 𝐚κ0,1if 𝐚κ=0,βκ={𝐚κ𝐛κ𝐚κ¯if 𝐚κ0,𝐛κif 𝐚κ=0.\displaystyle\alpha_{\kappa}=\begin{cases}\mathbf{a}_{\kappa}\quad\text{if $\mathbf{a}_{\kappa}\neq 0$},\\ 0\quad\text{if $\mathbf{a}_{\kappa}=0$},\end{cases}\mathbf{q}_{\kappa}=\begin{cases}\dfrac{\overline{\mathbf{a}_{\kappa}}}{\mathbf{a}_{\kappa}}\quad\text{if $\mathbf{a}_{\kappa}\neq 0$},\\ 1\quad\text{if $\mathbf{a}_{\kappa}=0$},\end{cases}\beta_{\kappa}=\begin{cases}\dfrac{\mathbf{a}_{\kappa}\mathbf{b}_{\kappa}}{\overline{\mathbf{a}_{\kappa}}}\quad\text{if $\mathbf{a}_{\kappa}\neq 0$},\\ \mathbf{b}_{\kappa}\quad\text{if $\mathbf{a}_{\kappa}=0$}.\end{cases}

Acknowledgements

The paper was completed during a scientific stay of P.V. Hai at the Vietnam Institute for Advanced Study in Mathematics (VIASM). He would like to thank the VIASM for financial support and hospitality.

References

  • [1] Frédéric Bayart. Similarity to an isometry of a composition operator. Proc. Am. Math. Soc., 131(6):1789–1791, 2003.
  • [2] Carl M. Bender, Stefan Boettcher, and Peter N. Meisinger. 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric quantum mechanics. J. Math. Phys., 40(5):2201–2229, 1999.
  • [3] Tom Carroll and Clifford Gilmore. Weighted composition operators on Fock spaces and their dynamics. J. Math. Anal. Appl., 502(1):125234, 21, 2021.
  • [4] Brent J. Carswell, Barbara D. MacCluer, and Alex Schuster. Composition operators on the Fock space. Acta Sci. Math., 69(3-4):871–887, 2003.
  • [5] Manuel D. Contreras, Santiago Díaz-Madrigal, María J. Martín, and Dragan Vukotić. Holomorphic self-maps of the disk intertwining two linear fractional maps. In Topics in complex analysis and operator theory. Third winter school in complex analysis and operator theory, Valencia, Spain, February 2–5, 2010, pages 199–227. Providence, RI: American Mathematical Society (AMS), 2012.
  • [6] Carl C. Cowen. The commutant of an analytic Toeplitz operator. Trans. Am. Math. Soc., 239:1–31, 1978.
  • [7] Carl C. Cowen and Eungil Ko. Hermitian weighted composition operators on H2H^{2}. Trans. Am. Math. Soc., 362(11):5771–5801, 2010.
  • [8] Carl C. Cowen and Barbara D. MacCluer. Composition operators on spaces of analytic functions. Boca Raton, FL: CRC Press, 1995.
  • [9] Karel de Leeuw, Walter Rudin, and John Wermer. The isometries of some function spaces. Proc. Am. Math. Soc., 11:694–698, 1960.
  • [10] F. Forelli. The isometries of HpH^{p}. Can. J. Math., 16:721–728, 1964.
  • [11] Stephan Ramon Garcia and Christopher Hammond. Which weighted composition operators are complex symmetric? In Concrete operators, spectral theory, operators in harmonic analysis and approximation. 22nd international workshop in operator theory and its applications, IWOTA 11, Sevilla, Spain, July 3–9, 2011, pages 171–179. Basel: Birkhäuser/Springer, 2014.
  • [12] Stephan Ramon Garcia, Emil Prodan, and Mihai Putinar. Mathematical and physical aspects of complex symmetric operators. J. Phys. A, Math. Theor., 47(35):54, 2014. Id/No 353001.
  • [13] Stephan Ramon Garcia and Mihai Putinar. Complex symmetric operators and applications. Trans. Am. Math. Soc., 358(3):1285–1315, 2006.
  • [14] Stephan Ramon Garcia and Mihai Putinar. Complex symmetric operators and applications. II. Trans. Am. Math. Soc., 359(8):3913–3931, 2007.
  • [15] Gajath Gunatillake. Invertible weighted composition operators. J. Funct. Anal., 261(3):831–860, 2011.
  • [16] Pham Viet Hai. Unbounded weighted composition operators on Fock space. Potential Anal., 53(1):1–21, 2020.
  • [17] Pham Viet Hai and Le Hai Khoi. Boundedness and compactness of weighted composition operators on Fock spaces p()\mathcal{F}^{p}(\mathbb{C}). Acta Math. Vietnam., 41(3):531–537, 2016.
  • [18] M. R. Jabbarzadeh and M. Moradi. Complex symmetric weighted composition Lambert type operators on L2(Σ)L^{2}(\Sigma). Oper. Matrices, 12(1):271–285, 2018.
  • [19] Sungeun Jung, Yoenha Kim, Eungil Ko, and Ji Eun Lee. Complex symmetric weighted composition operators on H2(𝔻)H^{2}(\mathbb{D}). J. Funct. Anal., 267(2):323–351, 2014.
  • [20] Hyungwoon Koo, Michael Stessin, and Kehe Zhu. Composition operators on the polydisc induced by smooth symbols. J. Funct. Anal., 254(11):2911–2925, 2008.
  • [21] Trieu Le. Self-adjoint, unitary, and normal weighted composition operators in several variables. J. Math. Anal. Appl., 395(2):596–607, 2012.
  • [22] Trieu Le. Normal and isometric weighted composition operators on the Fock space. Bull. Lond. Math. Soc., 46(4):847–856, 2014.
  • [23] J. E. Littlewood. On inequalities in the theory of functions. Proc. Lond. Math. Soc. (2), 23:v–ix + 481–519, 1925.
  • [24] Rubén A. Martínez-Avendaño and Peter Rosenthal. An introduction to operators on the Hardy–Hilbert space, volume 237. New York, NY: Springer, 2007.
  • [25] Pham Viet Hai and Le Hai Khoi. Complex symmetry of weighted composition operators on the Fock space. J. Math. Anal. Appl., 433(2):1757–1771, 2016.
  • [26] K. Schmüdgen. Unbounded self-adjoint operators on Hilbert space. Graduate Texts in Mathematics, 265. Springer, Dordrecht, 2012.
  • [27] Joel H. Shapiro. The essential norm of a composition operator. Ann. Math. (2), 125:375–404, 1987.
  • [28] Joel H. Shapiro. Composition operators and classical function theory. New York: Springer-Verlag, 1993.
  • [29] R. K. Singh and J. S. Manhas. Composition operators on function spaces., volume 179. Amsterdam: North-Holland, 1993.
  • [30] Aristomenis G. Siskakis. Semigroups of composition operators on the Dirichlet space. Result. Math., 30(1-2):165–173, 1996.
  • [31] Maofa Wang and Kaikai Han. Complex symmetric weighted composition operators in several variables. J. Math. Anal. Appl., 474(2):961–987, 2019.