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

On the similarity of powers of operators with flag structure

Jianming Yang Department of MathematicsHebei Normal University 050016 Shijiazhuang, P.R.China curraddr nova_yang@petalmail.com  and  Kui Ji Department of MathematicsHebei Normal University 050016 Shijiazhuang, P.R.China curraddr jikui@hebtu.edu.cn
(Date: Dec.27th, 2023)
Abstract.

Let La2(𝔻)\mathrm{L}^{2}_{a}(\mathbb{D}) be the classical Bergman space and denote MhM_{h} for the multiplication operator by a function hh. Let BB be a finite Blaschke product with order nn. An open question proposed by R. G. Douglas is whether the operators MBM_{B} on La2(𝔻)\mathrm{L}^{2}_{a}(\mathbb{D}) similar to 1nMz\oplus_{1}^{n}M_{z} on 1nLa2(𝔻)\oplus_{1}^{n}\mathrm{L}^{2}_{a}(\mathbb{D})? The question was answered in the affirmative, not only for Bergman space but also for many other Hilbert spaces with reproducing kernels. Since the operator MzM_{z}^{*} is in Cowen-Douglas class B1(𝔻)B_{1}(\mathbb{D}) in many cases, Douglas’s question can be expressed as a version for operators in B1(𝔻)B_{1}(\mathbb{D}), and it is affirmative for many operators in B1(𝔻)B_{1}(\mathbb{D}). A natural question is how about Douglas’s question in the version for operators in Cowen-Douglas class Bn(𝔻)B_{n}(\mathbb{D}) (n>1n>1)? In this paper, we investigate a family of operators, which are in a norm dense subclass of Cowen-Douglas class B2(𝔻)B_{2}(\mathbb{D}), and give a negative answer. This indicates that Douglas’s question cannot be directly generalized to general Hilbert spaces with vector-valued analytical reproducing kernel.

2020 Mathematics Subject Classification:
Primary 47B13, 47B32 Secondary 32L05, 47B35

1. Introduction

Let 𝔻\mathbb{D} be the open unit disk in \mathbb{C} and 𝔻¯\overline{\mathbb{D}} be the closed unit disk. A function BB on 𝔻\mathbb{D} is said a finite Blaschke product with order nn if it has the following form

B(z)=eiθj=1nzaj1aj¯z,z𝔻,B(z)=e^{i\theta}\prod_{j=1}^{n}\frac{z-a_{j}}{1-\bar{a_{j}}z},\ z\in\mathbb{D},

for some θ[0,2π)\theta\in[0,2\pi) and some a1,,an𝔻a_{1},\dots,a_{n}\in\mathbb{D}. Let La2(𝔻)\mathrm{L}^{2}_{a}(\mathbb{D}) be the classical Bergman space on 𝔻\mathbb{D} and denote MhM_{h} for the multiplication operator on La2(𝔻)\mathrm{L}^{2}_{a}(\mathbb{D}) by a function hHol(𝔻)h\in\mathrm{Hol}(\mathbb{D}). A previously open question proposed by R. G. Douglas (Question 6 in [6]) is whether the operator MBM_{B} on La2(𝔻)\mathrm{L}^{2}_{a}(\mathbb{D}) similar to 1nMz\oplus_{1}^{n}M_{z} on 1nLa2(𝔻)\oplus_{1}^{n}\mathrm{L}^{2}_{a}(\mathbb{D}), where nn is the order of the finite Blaschke product BB?

The question was answered in the affirmative (see [17] or [9]). Later, the question was also answered in the affirmative on many other analytic function Hilbert spaces, such as weighted Bergman spaces Aα2\mathrm{A}^{2}_{\alpha} (see [18]), Sobolev disk algebra R(𝔻)\mathrm{R}(\mathbb{D}) (see [20] or [14]), and Dirichlet space 𝔇\mathfrak{D} (see [19]). Recently in [11], Hou and Jiang proved that the question still holds on the weighted Hardy space of polynomial growth, which covers the weighted Bergman space, the weighted Dirichlet space, and many weighted Hardy spaces defined without measures.

Douglas’s question originated in the study of reducing subspaces of analytic multiplication operators on Bergman space. Before the question was proposed, there had been many studies of the reducing subspaces of the multiplication operator by a finite Blaschke product (see [21], [12], and [10]). An application of Douglas’s question, together with the technology of strongly irreducible operator and K0\mathrm{K}_{0}-group, is the similarity of analytic Toeplitz operators (e.g. Theorem 1.1 in [18]), which is analogous to the result on Hardy space (see [3]).

Note that MBM_{B} is also equal to the analytic function calculus B(Mz)B(M_{z}), then we can describe Douglas’s question in the following general form: let TT be a certain bounded linear operator on a complex separable Hilbert space HH and suppose σ(T)𝔻¯\sigma(T)\subseteq\overline{\mathbb{D}}, then for any finite Blaschke product BB, is the operator B(T)B(T) similar to 1nT\oplus_{1}^{n}T on 1nH\oplus_{1}^{n}H, where nn is the order of BB?

Obviously, an operator TT satisfies Douglas’s question if and only if so does its adjoint TT^{*}. Since the adjoints MzM_{z}^{*} of the multiplication operators MzM_{z} by zz on many analytic function Hilbert spaces are in Cowen-Douglas class B1(𝔻)B_{1}(\mathbb{D}) proposed in [5], the works mentioned earlier correspond to Douglas’s question in the case that TT is in B1(𝔻)B_{1}(\mathbb{D}). In that way, how about Douglas’s question in the case that TT is in Cowen-Douglas class Bn(𝔻)(n>1)B_{n}(\mathbb{D})(n>1)?

For a long time, there has been a lack of sufficient understanding of Cowen-Douglas class Bn(Ω)B_{n}(\Omega) for the higher rank case. A new but important subclass FBn(Ω)FB_{n}(\Omega) has been introduced by G. Misra in [13]. All irreducible homogeneous operators in Bn(𝔻)B_{n}(\mathbb{D}) are in FBn(𝔻)FB_{n}(\mathbb{D}) (see [13]), and the class FBn(Ω)FB_{n}(\Omega) (or even its subclass CFBn(Ω)CFB_{n}(\Omega)) is norm dense in Bn(Ω)B_{n}(\Omega) (see [16]). G. Misra etc. proved the operator in FBn(Ω)FB_{n}(\Omega) possesses a flag structure. It is also proved that the flag structure is rigid, that is, the unitary equivalence class of the operator and the flag structure determine each other (see [13]).

In this paper, we investigate a family of operators in class FB2(𝔻)FB_{2}(\mathbb{D}) and give a negative answer to Douglas’s question.

Denote by Hol(𝔻)\mathrm{Hol}(\mathbb{D}) the set of all analytical functions on 𝔻\mathbb{D} and denote by Hol(𝔻¯)\mathrm{Hol}(\overline{\mathbb{D}}) the set of all analytical functions on 𝔻¯\overline{\mathbb{D}}. Let Hα2\mathrm{H}^{2}_{\alpha} be a weighted Hardy space with weight sequence α\alpha on 𝔻\mathbb{D} and denote Mα,hM_{\alpha,h} for the multiplication operator by hHol(𝔻)h\in\mathrm{Hol}(\mathbb{D}) on Hα2\mathrm{H}^{2}_{\alpha}. Let Hβ2\mathrm{H}^{2}_{\beta} be another weighted Hardy space with weight sequence β\beta on 𝔻\mathbb{D} and denote Mα,β,hM_{\alpha,\beta,h} for the multiplication operator by hHol(𝔻)h\in\mathrm{Hol}(\mathbb{D}) from Hα2\mathrm{H}^{2}_{\alpha} to Hβ2\mathrm{H}^{2}_{\beta}. These concepts will be introduced in detail later.

The following theorem is the main theorem of this paper.

Theorem 1.1.

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two Mo¨\ddot{o}bius invariant weighted Hardy spaces with property A for some integer n02n_{0}\geq 2. Suppose the weight sequences α\alpha and β\beta satisfy

supk0β(k)α(k)<andlimkα(k)kβ(k)=0.\sup_{k\geq 0}\frac{\beta(k)}{\alpha(k)}<\infty\quad\text{and}\quad\lim_{k\to\infty}\frac{\alpha(k)}{k\beta(k)}=0.

Denote TT be the operator

(Mα,zMα,β,h0Mβ,z)\begin{pmatrix}M_{\alpha,z}^{*}&M_{\alpha,\beta,h}^{*}\\ 0&M_{\beta,z}^{*}\\ \end{pmatrix}

on Hα2Hβ2\mathrm{H}^{2}_{\alpha}\oplus\mathrm{H}^{2}_{\beta}, where hHol(𝔻¯)h\in\mathrm{Hol}(\overline{\mathbb{D}}).

If for any finite Blaschke product BB, B(T)B(T) is similar to 1nT\oplus_{1}^{n}T, where nn is the order of BB, then h=0h=0.

In other words, under the hypothesis of Theorem 1.1, for any non-zero function hh in Hol(𝔻¯)\mathrm{Hol}(\overline{\mathbb{D}}), the operator

(Mα,zMα,β,h0Mβ,z)\begin{pmatrix}M_{\alpha,z}^{*}&M_{\alpha,\beta,h}^{*}\\ 0&M_{\beta,z}^{*}\end{pmatrix}

doesn’t satisfy Douglas’s question.

2. Preliminaries

2.1. Weighted Hardy space

In the present article, we denote the power series coefficients of an analytical function ff as (f^(k))k=0(\hat{f}(k))_{k=0}^{\infty}, i.e. f(z)=k=0f^(k)zkf(z)=\sum_{k=0}^{\infty}\hat{f}(k)z^{k}.

2.1.1.

Let α=(α(k))k=0\alpha=(\alpha(k))_{k=0}^{\infty} be a positive sequence, then the weighted Hardy space Hα2\mathrm{H}^{2}_{\alpha} with weight sequence α\alpha on 𝔻\mathbb{D} is defined as

{fHol(𝔻):k=0|f^(k)|2α(k)2<}.\left\{f\in\mathrm{Hol}(\mathbb{D})\colon\sum_{k=0}^{\infty}|\hat{f}(k)|^{2}\alpha(k)^{2}<\infty\right\}.

Hα2\mathrm{H}^{2}_{\alpha} has an inner product ,α\langle\cdot,\cdot\rangle_{\alpha} defined as

f,gα=k=0f^(k)g^(k)¯α(k)2,\langle f,g\rangle_{\alpha}=\sum_{k=0}^{\infty}\hat{f}(k)\overline{\hat{g}(k)}\alpha(k)^{2},

where f,gHα2f,g\in\mathrm{H}^{2}_{\alpha}. In the present article, we always assume the weight sequence α\alpha satisfying

(2.1) limkα(k+1)α(k)=1.\lim_{k\to\infty}\frac{\alpha(k+1)}{\alpha(k)}=1.

In this case, Hα2\mathrm{H}^{2}_{\alpha} is a Hilbert space and contains Hol(𝔻¯)\mathrm{Hol}(\overline{\mathbb{D}}). What’s more, if a complex sequence a=(a(k))k=0a=(a(k))_{k=0}^{\infty} satisfies k=0|a(k)|2α(k)2<\sum_{k=0}^{\infty}|a(k)|^{2}\alpha(k)^{2}<\infty and a function ff is defined as f(z)=k=0a(k)zkf(z)=\sum_{k=0}^{\infty}a(k)z^{k}, then fHol(𝔻)f\in\mathrm{Hol}(\mathbb{D}) and hence fHα2f\in\mathrm{H}^{2}_{\alpha}.

The norm of fHα2f\in\mathrm{H}^{2}_{\alpha} is fα=(k=0|f^(k)|2α(k)2)12\|f\|_{\alpha}=(\sum_{k=0}^{\infty}|\hat{f}(k)|^{2}\alpha(k)^{2})^{\frac{1}{2}}. Let ek(z)=zke_{k}(z)=z^{k}, z𝔻z\in\mathbb{D}, then {ek}k=0\{e_{k}\}_{k=0}^{\infty} forms an orthogonal basis of Hα2\mathrm{H}^{2}_{\alpha} (thus, Hα2\mathrm{H}^{2}_{\alpha} is separable). In addition, ekα=α(k)\|e_{k}\|_{\alpha}=\alpha(k) and f=k=0f^(k)ekf=\sum_{k=0}^{\infty}\hat{f}(k)e_{k} in the sense of norm α\|\cdot\|_{\alpha}.

It can be checked that for each ω𝔻\omega\in\mathbb{D}, the linear function δω:fHα2f(ω)\delta_{\omega}\colon f\in\mathrm{H}^{2}_{\alpha}\to f(\omega)\in\mathbb{C} is continuous. Then by Riesz representation Theorem, Hα2\mathrm{H}^{2}_{\alpha} has a reproducing kernel {kω}ω𝔻\{k_{\omega}\}_{\omega\in\mathbb{D}}, i.e. for each ω𝔻\omega\in\mathbb{D}, kωHα2k_{\omega}\in\mathrm{H}^{2}_{\alpha} and f,kωα=f(ω)\langle f,k_{\omega}\rangle_{\alpha}=f(\omega) whenever fHα2f\in\mathrm{H}^{2}_{\alpha}.

2.1.2.

Let Hα2\mathrm{H}^{2}_{\alpha} be a weighted Hardy space. The multiplier algebra Mult(Hα2)\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}) of Hα2\mathrm{H}^{2}_{\alpha} is defined as the set of all hHol(𝔻)h\in\mathrm{Hol}(\mathbb{D}) such that hfHα2hf\in\mathrm{H}^{2}_{\alpha} for any fHα2f\in\mathrm{H}^{2}_{\alpha}. For each hMult(Hα2)h\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}), one can define a multiplication operator Mα,h:fHα2hfHα2M_{\alpha,h}\colon f\in\mathrm{H}^{2}_{\alpha}\to hf\in\mathrm{H}^{2}_{\alpha}, which is bounded by closed graph Theorem. Sometimes Mα,hM_{\alpha,h} is also abbreviated as MhM_{h}. As well known, Mult(Hα2)H(𝔻)Hα2\mathrm{Mult}(\mathrm{H}^{2}_{\alpha})\subseteq\mathrm{H}^{\infty}(\mathbb{D})\cap\mathrm{H}^{2}_{\alpha}, where H(𝔻)\mathrm{H}^{\infty}(\mathbb{D}) is the set of all bounded analytic functions on 𝔻\mathbb{D}. The Mapping hMult(Hα2)Mα,h(Hα2)h\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha})\to M_{\alpha,h}\in\mathcal{L}(\mathrm{H}^{2}_{\alpha}) is an algebraic monomorphism with 1I1\to I, and for each hMult(Hα2)h\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}), Mα,hM_{\alpha,h} is invertible if and only if 1hMult(Hα2)\frac{1}{h}\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}).

As well known, the assumption (2.1) guarantees that the multiplication operator Mα,zM_{\alpha,z} is a bounded linear operator on Hα2\mathrm{H}^{2}_{\alpha} and Mα,z=supk0α(k+1)α(k)\|M_{\alpha,z}\|=\sup_{k\geq 0}\frac{\alpha(k+1)}{\alpha(k)}. Denote {Mα,z}\{M_{\alpha,z}\}^{\prime} for the commutant algebra of Mα,zM_{\alpha,z}, i.e. the set of all X(Hα2)X\in\mathcal{L}(\mathrm{H}^{2}_{\alpha}) such that XMα,z=Mα,zXXM_{\alpha,z}=M_{\alpha,z}X. Then it can be verified that XX is in {Mα,z}\{M_{\alpha,z}\}^{\prime} if and only if XX is equal to Mα,hM_{\alpha,h} for some hMult(Hα2)h\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}).

Furthermore, the assumption (2.1) also guarantees that Hol(𝔻¯)Mult(Hα2)\mathrm{Hol}(\overline{\mathbb{D}})\subseteq\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}) (see Lemma 3.4 in [7] or see [11]). Then it can be verified that the spectrum σ(Mα,z)\sigma(M_{\alpha,z}) of Mα,zM_{\alpha,z} is contained in 𝔻¯\overline{\mathbb{D}} and for any hHol(𝔻¯)h\in\mathrm{Hol}(\overline{\mathbb{D}}), the analytic function calculus h(Mα,z)h(M_{\alpha,z}) is Mα,hM_{\alpha,h}.

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two weighted Hardy spaces. The multiplier Mult(Hα2,Hβ2)\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}) from Hα2\mathrm{H}^{2}_{\alpha} to Hβ2\mathrm{H}^{2}_{\beta} is defined as the set of all hHol(𝔻)h\in\mathrm{Hol}(\mathbb{D}) such that hfHβ2hf\in\mathrm{H}^{2}_{\beta} for any fHα2f\in\mathrm{H}^{2}_{\alpha}. For each hMult(Hα2,Hβ2)h\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}), one can define a multiplication operator Mα,β,h:fHα2hfHβ2M_{\alpha,\beta,h}\colon f\in\mathrm{H}^{2}_{\alpha}\to hf\in\mathrm{H}^{2}_{\beta}, which is bounded by closed graph Theorem. Sometimes Mα,β,hM_{\alpha,\beta,h} is also abbreviated as MhM_{h}. Obviously, Mult(Hα2,Hβ2)Hβ2\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta})\subseteq\mathrm{H}^{2}_{\beta}.

In addition, it can also be verified that X(Hα2,Hβ2)X\in\mathcal{L}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}) satisfies XMα,z=Mβ,zXXM_{\alpha,z}=M_{\beta,z}X if and only if XX is equal to Mα,β,hM_{\alpha,\beta,h} for some hMult(Hα2,Hβ2)h\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}).

Remark 2.1.

A simple example of weighted Hardy space involved in the present article is the Hilbert space H(λ)\mathrm{H}_{(\lambda)}. For any λ\lambda\in\mathbb{R}, H(λ)\mathrm{H}_{(\lambda)} is defined as the weighted Hardy space Hαλ2\mathrm{H}^{2}_{\alpha_{\lambda}} with weight sequence αλ(k)=(k+1)λ\alpha_{\lambda}(k)=(k+1)^{\lambda}, k=0,1,k=0,1,\dots. This type of space contains many classical analytic function spaces on 𝔻\mathbb{D}. For example, the classical Hardy space H2(𝔻)\mathrm{H}^{2}(\mathbb{D}) (λ=0\lambda=0), the classical Bergman space La2(𝔻)\mathrm{L}^{2}_{a}(\mathbb{D}) (λ=12\lambda=-\frac{1}{2}) and the classical Dirichlet space 𝔇\mathfrak{D} (λ=12\lambda=\frac{1}{2}).

Remark 2.2.

A complicated example of weighted Hardy space involved in the present article is the weighted Hardy space of polynomial growth (which has been studied in [11] recently). A weighted Hardy space Hα2\mathrm{H}^{2}_{\alpha} (usually assume α(0)=1\alpha(0)=1) is called of polynomial growth if the weight sequence α\alpha satisfies

supk(k+1)|α(k)α(k1)1|<.\sup_{k\in\mathbb{N}}(k+1)|\frac{\alpha(k)}{\alpha(k-1)}-1|<\infty.

The weighted Hardy spaces of polynomial growth cover the weighted Bergman space, the weighted Dirichlet space, and many weighted Hardy spaces defined without measures.

2.2. Cowen-Douglas operators

Recall some basic concepts of the Cowen-Doug-las operator class Bn(Ω)B_{n}(\Omega), which was introduced in [5], and the subclass FB2(Ω)FB_{2}(\Omega) of B2(Ω)B_{2}(\Omega), which was introduced in [13].

Let HH be a complex separable Hilbert space and (H)\mathcal{L}(H) denote the collection of bounded linear operators on HH. For Ω\Omega a connected open subset of \mathbb{C} and nn a positive integer, let Bn(Ω)B_{n}(\Omega) denote the operators TT in (H)\mathcal{L}(H) which satisfy:

(a) Ωσ(T)={ω:Tωnot invertible}\Omega\subseteq\sigma(T)=\left\{\omega\in\mathbb{C}\colon T-\omega\ \text{not invertible}\right\};

(b) Ran(Tω)=H\mathrm{Ran\,}(T-\omega)=H for ωΩ\omega\in\Omega;

(c) ωΩKer(Tω)=H\vee_{\omega\in\Omega}\,\mathrm{Ker\,}(T-\omega)=H; and

(d) dimKer(Tω)=n\dim\mathrm{Ker\,}(T-\omega)=n for ωΩ\omega\in\Omega.

As well known, the adjoint Mα,zM_{\alpha,z}^{*} of the multiplication operator Mα,zM_{\alpha,z} acting on the weighted Hardy space Hα2\mathrm{H}^{2}_{\alpha} is in class B1(𝔻)B_{1}(\mathbb{D}).

The operator class FB2(Ω)FB_{2}(\Omega) is the set of all bounded linear operators TT of the form

(T0S0T1),\begin{pmatrix}T_{0}&S\\ 0&T_{1}\\ \end{pmatrix},

where T0T_{0} and T1T_{1} are in class B1(Ω)B_{1}(\Omega) and the operator SS is a non-zero intertwiner between them, i.e. T0S=ST1T_{0}S=ST_{1}.

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two weighted Hardy spaces. Throughout this paper, we will denote Tα,β,hT_{\alpha,\beta,h} as the operator

(Mα,zMα,β,h0Mβ,z)\begin{pmatrix}M_{\alpha,z}^{*}&M_{\alpha,\beta,h}^{*}\\ 0&M_{\beta,z}^{*}\\ \end{pmatrix}

on Hα2Hβ2\mathrm{H}^{2}_{\alpha}\oplus\mathrm{H}^{2}_{\beta} for a function hh in Mult(Hα2,Hβ2)\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}). One can see that Tα,β,hT_{\alpha,\beta,h} is in class FB2(𝔻)FB_{2}(\mathbb{D}) whenever hh is a non-zero function.

3. Two propositions

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two weighted Hardy spaces. Let T=Tα,β,hT=T_{\alpha,\beta,h}, where hh is a non-zero function in Mult(Hα2,Hβ2)\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}). To investigate whether for any finite Blaschke product BB, B(T)B(T) is similar to 1nT\oplus_{1}^{n}T, where nn is the order of BB, we will treat BB in two cases: the case of B(z)=eiθza1a¯zB(z)=e^{i\theta}\frac{z-a}{1-\bar{a}z} and the case of B(z)=znB(z)=z^{n}. In 4, we deal with the first case and the main point is Proposition 4.3. In 5, we deal with the second case and the main points are Proposition 5.6 and Proposition 5.10. Before that, in 3, we will first prove the following two propositions: Proposition 3.1 and Proposition 3.4.

In this paper, we write TT~T\sim\tilde{T} for two bounded linear operators TT and T~\tilde{T} if TT is similar to T~\tilde{T} (i.e. there exists an invertible operator XX such that TX=XT~TX=X\tilde{T}).

Let T0T_{0} and T1T_{1} be two bounded linear operators on Hilbert spaces H0H_{0} and H1H_{1} respectively. Denote σT0,T1(X):=T0XXT1\sigma_{T_{0},T_{1}}(X):=T_{0}X-XT_{1} for any X(H1,H0)X\in\mathcal{L}(H_{1},H_{0}). Then a linear operator σT0,T1:(H1,H0)(H1,H0)\sigma_{T_{0},T_{1}}\colon\mathcal{L}(H_{1},H_{0})\to\mathcal{L}(H_{1},H_{0}) can be defined. Let σT\sigma_{T} be the operator σT,T\sigma_{T,T}.

From Lemma 2.18 in [13] and Theorem 2.19 in [13], a useful conclusion can be obtained that for any TT in class B1(Ω)B_{1}(\Omega), KerσTRanσT={0}\mathrm{Ker\,}\sigma_{T}\cap\mathrm{Ran\,}\sigma_{T}=\left\{0\right\}. But we need to slightly generalize this.

Proposition 3.1.

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two weighted Hardy spaces. If the weight sequences α\alpha and β\beta satisfy condition

(3.1) limkα(k)kβ(k)=0,\lim_{k\to\infty}\frac{\alpha(k)}{k\beta(k)}=0,

then KerσMα,z,Mβ,zRanσMα,z,Mβ,z={0}\mathrm{Ker\,}\sigma_{M_{\alpha,z}^{*},M_{\beta,z}^{*}}\cap\mathrm{Ran\,}\sigma_{M_{\alpha,z}^{*},M_{\beta,z}^{*}}=\left\{0\right\}.

Proof.

Clearly, we just need to explain that for any Y(Hα2,Hβ2)Y\in\mathcal{L}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}), the equations YMα,z=Mβ,zYYM_{\alpha,z}=M_{\beta,z}Y and Y=XMα,zMβ,zXY=XM_{\alpha,z}-M_{\beta,z}X, for some X(Hα2,Hβ2)X\in\mathcal{L}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}), imply Y=0Y=0.

The equation YMα,z=Mβ,zYYM_{\alpha,z}=M_{\beta,z}Y is equivalent to Y=Mα,β,hY=M_{\alpha,\beta,h} for some hMult(Hα2,Hβ2)h\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}). Then we have Y=XMα,zMβ,zXY=XM_{\alpha,z}-M_{\beta,z}X, XMα,z=Mβ,zX+Mα,β,hXM_{\alpha,z}=M_{\beta,z}X+M_{\alpha,\beta,h}. Hence, Xzk=zXzk1+hzkXz^{k}=zXz^{k-1}+hz^{k}, k=1,2,k=1,2,\dots. Let g:=Xz0Hβ2g:=Xz^{0}\in\mathrm{H}^{2}_{\beta}. Then by induction, we have that Xzk=zkg+kzk1hXz^{k}=z^{k}g+kz^{k-1}h, k=1,2,k=1,2,\dots.

Let Xzk=j=0xjkzjXz^{k}=\sum_{j=0}^{\infty}x_{jk}z^{j}, k=0,1,k=0,1,\dots. Then

(3.2) Xzk,zl=j=0xjkzj,zl=j=0xjkzj,zl=xlkβ(l)2,\langle Xz^{k},z^{l}\rangle=\langle\sum_{j=0}^{\infty}x_{jk}z^{j},z^{l}\rangle=\sum_{j=0}^{\infty}x_{jk}\langle z^{j},z^{l}\rangle=x_{lk}\beta(l)^{2},

where k,l=0,1,k,l=0,1,\dots. On the other hand,

zkg(z)=j=0g^(j)zj+k=j=0g^(jk)zj,z^{k}g(z)=\sum_{j=0}^{\infty}\hat{g}(j)z^{j+k}=\sum_{j=0}^{\infty}\hat{g}(j-k)z^{j},
zkh(z)=j=0h^(j)zj+k=j=0h^(jk)zj,z^{k}h(z)=\sum_{j=0}^{\infty}\hat{h}(j)z^{j+k}=\sum_{j=0}^{\infty}\hat{h}(j-k)z^{j},

where k=0,1,k=0,1,\dots, and g^(j)=0=h^(j)\hat{g}(j)=0=\hat{h}(j) whenever j<0j<0. Hence,

(3.3) Xzk,zl=zkg+kzk1h,zl=zkg,zl+kzk1h,zl=j=0g^(jk)zj,zl+kj=0h^(jk+1)zj,zl=g^(lk)β(l)2+kh^(lk+1)β(l)2,\begin{split}\langle Xz^{k},z^{l}\rangle&=\langle z^{k}g+kz^{k-1}h,z^{l}\rangle\\ &=\langle z^{k}g,z^{l}\rangle+k\langle z^{k-1}h,z^{l}\rangle\\ &=\langle\sum_{j=0}^{\infty}\hat{g}(j-k)z^{j},z^{l}\rangle+k\langle\sum_{j=0}^{\infty}\hat{h}(j-k+1)z^{j},z^{l}\rangle\\ &=\hat{g}(l-k)\beta(l)^{2}+k\hat{h}(l-k+1)\beta(l)^{2},\end{split}

where k=1,2,k=1,2,\dots and l=0,1,l=0,1,\dots. Note that Xz0=gXz^{0}=g, equation (3.3) still holds for k=0k=0. From (3.2) and (3.3), it follows that

xlk=g^(lk)+kh^(lk+1),x_{lk}=\hat{g}(l-k)+k\hat{h}(l-k+1),

where k,l=0,1,k,l=0,1,\dots.

Let t=0,1,t=0,1,\dots, k=1,2,k=1,2,\dots, and l=k+t1l=k+t-1. Then xk+t1,k=g^(t1)+kh^(t)x_{k+t-1,k}=\hat{g}(t-1)+k\hat{h}(t). Hence,

(3.4) xk+t1,kk=g^(t1)k+h^(t)h^(t),k,fort=0,1,.\frac{x_{k+t-1,k}}{k}=\frac{\hat{g}(t-1)}{k}+\hat{h}(t)\to\hat{h}(t),\ k\to\infty,\ \text{for}\ t=0,1,\dots.

From condition (3.1) and assumption (2.1), we can see that

α(k)kβ(k+t1)0,k,fort=0,1,.\frac{\alpha(k)}{k\beta(k+t-1)}\to 0,\ k\to\infty,\ \text{for}\ t=0,1,\dots.

From (3.2), it follows that

|xk+t1,k|=|Xzk,zk+t1|β(k+t1)2Xzkαzk+t1ββ(k+t1)2=α(k)β(k+t1)X.\begin{split}|x_{k+t-1,k}|&=\frac{|\langle Xz^{k},z^{k+t-1}\rangle|}{\beta(k+t-1)^{2}}\\ &\leq\frac{\|X\|\|z^{k}\|_{\alpha}\|z^{k+t-1}\|_{\beta}}{\beta(k+t-1)^{2}}\\ &=\frac{\alpha(k)}{\beta(k+t-1)}\|X\|.\end{split}

Hence,

(3.5) |xk+t1,k|kα(k)kβ(k+t1)X0,k,fort=0,1,.\frac{|x_{k+t-1,k}|}{k}\leq\frac{\alpha(k)}{k\beta(k+t-1)}\|X\|\to 0,\ k\to\infty,\ \text{for}\ t=0,1,\dots.

From (3.4) and (3.5), it can be observed that h^(t)=0\hat{h}(t)=0 when t=0,1,t=0,1,\dots. This implies h=0h=0 and consequently Y=Mα,β,h=0Y=M_{\alpha,\beta,h}=0. ∎

Remark 3.2.

Let λ,μ\lambda,\mu\in\mathbb{R}. The weight sequences of the Hilbert spaces H(λ)=Hαλ2\mathrm{H}_{(\lambda)}=\mathrm{H}^{2}_{\alpha_{\lambda}} and H(μ)=Hαμ2\mathrm{H}_{(\mu)}=\mathrm{H}^{2}_{\alpha_{\mu}} are αλ(k)=(k+1)λ\alpha_{\lambda}(k)=(k+1)^{\lambda} and αμ(k)=(k+1)μ\alpha_{\mu}(k)=(k+1)^{\mu}, respectively. Then the weight sequences α=αλ\alpha=\alpha_{\lambda} and β=αμ\beta=\alpha_{\mu} satisfy the condition (3.1) if and only if λμ<1\lambda-\mu<1.

The following proposition, which can be used to deal with two similar operators in FB2(Ω)FB_{2}(\Omega), is Proposition 3.3 in [13].

Lemma 3.3.

If XX is an invertible operator intertwining two operators TT and T~\tilde{T} in FB2(Ω)FB_{2}(\Omega), i.e. XT=T~XXT=\tilde{T}X, then XX is upper triangular:

X=(X11X120X22),X=\begin{pmatrix}X_{11}&X_{12}\\ 0&X_{22}\\ \end{pmatrix},

and the operators X11X_{11} and X22X_{22} are invertible.

Proposition 3.4.

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two weighted Hardy spaces, where the weight sequences α\alpha and β\beta satisfy condition (3.1). If the operators Tα,β,hT_{\alpha,\beta,h} and Tα,β,h~T_{\alpha,\beta,\tilde{h}} are similar, where hh and h~\tilde{h} are two non-zero functions in Mult(Hα2,Hβ2)\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}), then there exist h1Mult(Hα2)h_{1}\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}) and h2Mult(Hβ2)h_{2}\in\mathrm{Mult}(\mathrm{H}^{2}_{\beta}) such that 1h1Mult(Hα2)\frac{1}{h_{1}}\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}), 1h2Mult(Hβ2)\frac{1}{h_{2}}\in\mathrm{Mult}(\mathrm{H}^{2}_{\beta}), and h=h~h1h2h=\tilde{h}h_{1}h_{2}.

Proof.

Since T:=Tα,β,hT~:=Tα,β,h~T:=T_{\alpha,\beta,h}\sim\tilde{T}:=T_{\alpha,\beta,\tilde{h}}, there exists an invertible operator XX such that XT=T~XXT=\tilde{T}X. According to Lemma 3.3 and TT and T~\tilde{T} are in FB2(𝔻)FB_{2}(\mathbb{D}), it follows that XX is upper triangular:

X=(X11X120X22),X=\begin{pmatrix}X_{11}&X_{12}\\ 0&X_{22}\\ \end{pmatrix},

and the operators X11X_{11} and X22X_{22} are invertible.

From XT=T~XXT=\tilde{T}X, we have

X11Mα,z=Mα,zX11,X_{11}M_{\alpha,z}^{*}=M_{\alpha,z}^{*}X_{11},
X22Mβ,z=Mβ,zX22,X_{22}M_{\beta,z}^{*}=M_{\beta,z}^{*}X_{22},
X11Mα,β,h+X12Mβ,z=Mα,zX12+Mα,β,h~X22.X_{11}M_{\alpha,\beta,h}^{*}+X_{12}M_{\beta,z}^{*}=M_{\alpha,z}^{*}X_{12}+M_{\alpha,\beta,\tilde{h}}^{*}X_{22}.

The first two equations show that X11=Mα,g1X_{11}=M_{\alpha,g_{1}}^{*} and X22=Mβ,g2X_{22}=M_{\beta,g_{2}}^{*} for some g1Mult(Hα2)g_{1}\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}) and g2Mult(Hβ2)g_{2}\in\mathrm{Mult}(\mathrm{H}^{2}_{\beta}). The third equation shows that

X11Mα,β,hMα,β,h~X22=σMα,z,Mβ,z(X12).X_{11}M_{\alpha,\beta,h}^{*}-M_{\alpha,\beta,\tilde{h}}^{*}X_{22}=\sigma_{M_{\alpha,z}^{*},M_{\beta,z}^{*}}(X_{12}).

Since Mα,g1=X11M_{\alpha,g_{1}}^{*}=X_{11} and Mβ,g2=X22M_{\beta,g_{2}}^{*}=X_{22} are invertible, we have 1g1Mult(Hα2)\frac{1}{g_{1}}\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}) and 1g2Mult(Hβ2)\frac{1}{g_{2}}\in\mathrm{Mult}(\mathrm{H}^{2}_{\beta}). What’s more, we can directly calculate

σMα,z,Mβ,z(X11Mα,β,hMα,β,h~X22)=0.\sigma_{M_{\alpha,z}^{*},M_{\beta,z}^{*}}(X_{11}M_{\alpha,\beta,h}^{*}-M_{\alpha,\beta,\tilde{h}}^{*}X_{22})=0.

Therefore, X11Mα,β,hMα,β,h~X22X_{11}M_{\alpha,\beta,h}^{*}-M_{\alpha,\beta,\tilde{h}}^{*}X_{22} is in RanσMα,z,Mβ,zKerσMα,z,Mβ,z\mathrm{Ran\,}\sigma_{M_{\alpha,z}^{*},M_{\beta,z}^{*}}\cap\mathrm{Ker\,}\sigma_{M_{\alpha,z}^{*},M_{\beta,z}^{*}}. According to Proposition 3.1, this implies X11Mα,β,h=Mα,β,h~X22X_{11}M_{\alpha,\beta,h}^{*}=M_{\alpha,\beta,\tilde{h}}^{*}X_{22} and consequently Mα,β,hMα,g1=Mβ,g2Mα,β,h~M_{\alpha,\beta,h}M_{\alpha,g_{1}}=M_{\beta,g_{2}}M_{\alpha,\beta,\tilde{h}}. Thus, hg1=g2h~hg_{1}=g_{2}\tilde{h}.

Finally, take h1=1g1h_{1}=\frac{1}{g_{1}} and h2=g2h_{2}=g_{2}, and then we complete the proof. ∎

Remark 3.5.

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two weighted Hardy spaces. Let hh and h~\tilde{h} be in Mult(Hα2,Hβ2)\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}). If there exist h1Mult(Hα2)h_{1}\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}) and h2Mult(Hβ2)h_{2}\in\mathrm{Mult}(\mathrm{H}^{2}_{\beta}) such that 1h1Mult(Hα2)\frac{1}{h_{1}}\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}), 1h2Mult(Hβ2)\frac{1}{h_{2}}\in\mathrm{Mult}(\mathrm{H}^{2}_{\beta}), and h=h~h1h2h=\tilde{h}h_{1}h_{2}, then Tα,β,hT_{\alpha,\beta,h} and Tα,β,h~T_{\alpha,\beta,\tilde{h}} are similar.

Proof.

To see this, take the invertible operator

X=((Mα,h11)00Mβ,h2),X=\begin{pmatrix}(M_{\alpha,h_{1}}^{-1})^{*}&0\\ 0&M_{\beta,h_{2}}^{*}\\ \end{pmatrix},

then, by simple calculation, XTα,β,hX1=Tα,β,h~XT_{\alpha,\beta,h}X^{-1}=T_{\alpha,\beta,\tilde{h}}, and this implies that Tα,β,hT_{\alpha,\beta,h} and Tα,β,h~T_{\alpha,\beta,\tilde{h}} are similar. ∎

4. The case of B(z)=eiθza1a¯zB(z)=e^{i\theta}\frac{z-a}{1-\bar{a}z}

Recall a basic concept of weakly homogeneous operators, which was introduced by Clark and Misra [2]. Denote by Aut(𝔻)\mathrm{Aut\,}(\mathbb{D}) the analytic automorphism group of 𝔻\mathbb{D}, which is the set of all analytic bijection from 𝔻\mathbb{D} to itself. As well know, a function ϕ\phi is in Aut(𝔻)\mathrm{Aut\,}(\mathbb{D}) if and only if it has the following form: ϕ(z)=eiθza1a¯z,z𝔻,\phi(z)=e^{i\theta}\frac{z-a}{1-\bar{a}z},\ z\in\mathbb{D}, for some θ[0,2π)\theta\in[0,2\pi) and some a𝔻a\in\mathbb{D}. A bounded linear operator TT on a complex separable Hilbert space HH is called weakly homogeneous if σ(T)𝔻¯\sigma(T)\subseteq\overline{\mathbb{D}} and for any ϕAut(𝔻)\phi\in\mathrm{Aut\,}(\mathbb{D}), ϕ(T)\phi(T) is similar to TT.

We say a weighted Hardy space Hα2\mathrm{H}^{2}_{\alpha} be Mo¨\ddot{o}bius invariant if for each ϕAut(𝔻)\phi\in\mathrm{Aut\,}(\mathbb{D}), fϕHα2f\circ\phi\in\mathrm{H}^{2}_{\alpha} whenever fHα2f\in\mathrm{H}^{2}_{\alpha}. Let Hα2\mathrm{H}^{2}_{\alpha} be a Mo¨\ddot{o}bius invariant weighted Hardy space. Then for each ϕAut(𝔻)\phi\in\mathrm{Aut\,}(\mathbb{D}), one can define a composition operator Cα,ϕ:fHα2fϕHα2C_{\alpha,\phi}\colon f\in\mathrm{H}^{2}_{\alpha}\to f\circ\phi\in\mathrm{H}^{2}_{\alpha}, which is bounded by closed graph Theorem. What’s more, Cα,ϕC_{\alpha,\phi} is invertible and Cα,ϕ1=Cα,ϕ1C_{\alpha,\phi}^{-1}=C_{\alpha,\phi^{-1}}. By simple calculation, ϕ(Mα,z)Cα,ϕ=Mα,ϕCα,ϕ=Cα,ϕMα,z\phi(M_{\alpha,z})C_{\alpha,\phi}=M_{\alpha,\phi}C_{\alpha,\phi}=C_{\alpha,\phi}M_{\alpha,z}. Thus, Mα,zM_{\alpha,z} is a weakly homogeneous operator.

Remark 4.1.

It is known that for each λ\lambda\in\mathbb{R}, the Hilbert space H(λ)\mathrm{H}_{(\lambda)} is Mo¨\ddot{o}bius invariant (see [22], [4]).

Remark 4.2.

Theorem 3.1 in [11] shows that each weighted Hardy space of polynomial growth is Mo¨\ddot{o}bius invariant.

The main idea of the next proposition originates from Theorem 3.16 in [8]. Here we use Proposition 3.4 to improve it in part.

Proposition 4.3.

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two Mo¨\ddot{o}bius invariant weighted Hardy spaces. Suppose the weight sequences α\alpha and β\beta satisfy condition (3.1). If the operator Tα,β,hT_{\alpha,\beta,h} is weakly homogeneous, where h𝒞(𝔻¯)Mult(Hα2,Hβ2)h\in\mathcal{C}(\overline{\mathbb{D}})\cap\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}), then hh is zero everywhere on 𝔻¯\overline{\mathbb{D}} or hh is non-zero everywhere on 𝔻¯\overline{\mathbb{D}}.

Proof.

If hh is a zero function on 𝔻\mathbb{D}, then hh is a zero function on 𝔻¯\overline{\mathbb{D}} since h𝒞(𝔻¯)h\in\mathcal{C}(\overline{\mathbb{D}}). So we suppose that hh is a non-zero function on 𝔻\mathbb{D}.

Let ϕAut(𝔻)\phi\in\mathrm{Aut\,}(\mathbb{D}) and let ϕ~\tilde{\phi} be the function defined by ϕ~(z)=ϕ(z¯)¯\tilde{\phi}(z)=\overline{\phi(\overline{z})}. Then ϕ~\tilde{\phi} is also in Aut(𝔻)\mathrm{Aut\,}(\mathbb{D}). Since T:=Tα,β,hT:=T_{\alpha,\beta,h} is weakly homogeneous, ϕ~(T)T\tilde{\phi}(T)\sim T. From Mα,zMα,β,h=Mα,β,hMβ,zM_{\alpha,z}^{*}M_{\alpha,\beta,h}^{*}=M_{\alpha,\beta,h}^{*}M_{\beta,z}^{*}, it can be calculated that

ϕ~(T)=(ϕ~(Mα,z)ϕ~(Mα,z)Mα,β,h0ϕ~(Mβ,z))=(Mα,ϕMα,β,hϕ0Mβ,ϕ).\tilde{\phi}(T)=\begin{pmatrix}\tilde{\phi}(M_{\alpha,z}^{*})&\tilde{\phi}^{\prime}(M_{\alpha,z}^{*})M_{\alpha,\beta,h}^{*}\\ 0&\tilde{\phi}(M_{\beta,z}^{*})\\ \end{pmatrix}\\ =\begin{pmatrix}M_{\alpha,\phi}^{*}&M_{\alpha,\beta,h\phi^{\prime}}^{*}\\ 0&M_{\beta,\phi}^{*}\\ \end{pmatrix}.

Mo¨\ddot{o}bius invariance guarantees that the composition operators Cα,ϕC_{\alpha,\phi} and Cβ,ϕC_{\beta,\phi} are invertible and that Cα,ϕ1=Cα,ϕ1C_{\alpha,\phi}^{-1}=C_{\alpha,\phi^{-1}} and Cβ,ϕ1=Cβ,ϕ1C_{\beta,\phi}^{-1}=C_{\beta,\phi^{-1}}. Then, by simple calculation,

(Cα,ϕ00Cβ,ϕ)ϕ~(T)(Cα,ϕ00Cβ,ϕ)1=(Mα,zMα,β,(hϕ1)(ϕϕ1)0Mβ,z).\begin{pmatrix}C_{\alpha,\phi}^{*}&0\\ 0&C_{\beta,\phi}^{*}\\ \end{pmatrix}\tilde{\phi}(T)\begin{pmatrix}C_{\alpha,\phi}^{*}&0\\ 0&C_{\beta,\phi}^{*}\\ \end{pmatrix}^{-1}\\ =\begin{pmatrix}M_{\alpha,z}^{*}&M_{\alpha,\beta,(h\circ\phi^{-1})(\phi^{\prime}\circ\phi^{-1})}^{*}\\ 0&M_{\beta,z}^{*}\\ \end{pmatrix}.

So

T=Tα,β,hϕ~(T)Tα,β,(hϕ1)(ϕϕ1).T=T_{\alpha,\beta,h}\sim\tilde{\phi}(T)\sim T_{\alpha,\beta,(h\circ\phi^{-1})(\phi^{\prime}\circ\phi^{-1})}.

Since ϕ\phi is a bijection from 𝔻\mathbb{D} to itself and ϕ(z)=eiθ1|a|2(1a¯z)2\phi^{\prime}(z)=e^{i\theta}\frac{1-|a|^{2}}{(1-\bar{a}z)^{2}} is non-zero everywhere on 𝔻\mathbb{D}, then (hϕ1)(ϕϕ1)(h\circ\phi^{-1})(\phi^{\prime}\circ\phi^{-1}) is a non-zero function on 𝔻\mathbb{D}. Consequently, according to Proposition 3.4, there exists h1=h1,ϕMult(Hα2)h_{1}=h_{1,\phi}\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}) and h2=h2,ϕMult(Hβ2)h_{2}=h_{2,\phi}\in\mathrm{Mult}(\mathrm{H}^{2}_{\beta}) such that 1h1Mult(Hα2)\frac{1}{h_{1}}\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}), 1h2Mult(Hβ2)\frac{1}{h_{2}}\in\mathrm{Mult}(\mathrm{H}^{2}_{\beta}), and h=(hϕ1)(ϕϕ1)h1h2h=(h\circ\phi^{-1})(\phi^{\prime}\circ\phi^{-1})h_{1}h_{2}. Thus,

(4.1) h(ϕ(z))=h(z)ϕ(z)h1(ϕ(z))h2(ϕ(z)),z𝔻.h(\phi(z))=h(z)\phi^{\prime}(z)h_{1}(\phi(z))h_{2}(\phi(z)),\ z\in\mathbb{D}.

Then, similar to the last part of the proof of Theorem 3.16 in [8], we can obtain that hh is non-zero everywhere on 𝔻¯\overline{\mathbb{D}}. ∎

Remark 4.4.

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two Mo¨\ddot{o}bius invariant weighted Hardy spaces. Let hh be in Hol(𝔻¯)Mult(Hα2,Hβ2)\mathrm{Hol}(\overline{\mathbb{D}})\cap\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}). If hh is zero everywhere on 𝔻¯\overline{\mathbb{D}} or hh is non-zero everywhere on 𝔻¯\overline{\mathbb{D}}, then T=Tα,β,hT=T_{\alpha,\beta,h} is weakly homogeneous.

Proof.

It is trivial if hh is zero everywhere on 𝔻¯\overline{\mathbb{D}}. So we suppose that hh is non-zero everywhere on 𝔻¯\overline{\mathbb{D}}. Then 1hHol(𝔻¯)\frac{1}{h}\in\mathrm{Hol}(\overline{\mathbb{D}}). Let ϕAut(𝔻)\phi\in\mathrm{Aut\,}(\mathbb{D}) and let ϕ~\tilde{\phi} be the function defined by ϕ~(z)=ϕ(z¯)¯\tilde{\phi}(z)=\overline{\phi(\overline{z})}. From the proof of Proposition 4.3, we see that ϕ~(T)Tα,β,(hϕ1)(ϕϕ1)\tilde{\phi}(T)\sim T_{\alpha,\beta,(h\circ\phi^{-1})(\phi^{\prime}\circ\phi^{-1})}. It is not hard to see that the function h~:=(hϕ1)(ϕϕ1)\tilde{h}:=(h\circ\phi^{-1})(\phi^{\prime}\circ\phi^{-1}) is in Hol(𝔻¯)\mathrm{Hol}(\overline{\mathbb{D}}) and non-zero everywhere on 𝔻¯\overline{\mathbb{D}}. Then 1h~Hol(𝔻¯)\frac{1}{\tilde{h}}\in\mathrm{Hol}(\overline{\mathbb{D}}). Since h=h~1h~hh=\tilde{h}\frac{1}{\tilde{h}}h and Hol(𝔻¯)\mathrm{Hol}(\overline{\mathbb{D}}) is contained in both Mult(Hα2)\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}) and Mult(Hβ2)\mathrm{Mult}(\mathrm{H}^{2}_{\beta}), then according to Remark 3.5, T=Tα,β,hTα,β,(hϕ1)(ϕϕ1)T=T_{\alpha,\beta,h}\sim T_{\alpha,\beta,(h\circ\phi^{-1})(\phi^{\prime}\circ\phi^{-1})}. Thus, ϕ~(T)T\tilde{\phi}(T)\sim T. Since ϕ~\tilde{\phi} is arbitrary in Aut(𝔻)\mathrm{Aut\,}(\mathbb{D}), we obtain that TT is weakly homogeneous. ∎

5. The case of B(z)=znB(z)=z^{n}

5.1. Power of a special operator

Let Hα2\mathrm{H}^{2}_{\alpha} be a weighted Hardy space and nn be a positive integer. Denote

Hα,j2:={zj+kn}k=0Hα2,j=0,1,,n1.\mathrm{H}^{2}_{\alpha,j}:=\vee\left\{z^{j+kn}\right\}_{k=0}^{\infty}\subseteq\mathrm{H}^{2}_{\alpha},\ j=0,1,\dots,n-1.

Then, it can be verified that

Hα2=Hα,02Hα,12Hα,n12\mathrm{H}^{2}_{\alpha}=\mathrm{H}^{2}_{\alpha,0}\dotplus\mathrm{H}^{2}_{\alpha,1}\dotplus\cdots\dotplus\mathrm{H}^{2}_{\alpha,n-1}

is a Hilbert direct sum, and {zj+kn}k=0\left\{z^{j+kn}\right\}_{k=0}^{\infty} is an orthogonal basis of Hα,j2\mathrm{H}^{2}_{\alpha,j}. What’s more, Hα,j2\mathrm{H}^{2}_{\alpha,j} is an invariant subspace of the multiplication operator Mα,zn=Mα,znM_{\alpha,z^{n}}=M_{\alpha,z}^{n}.

We say that a weighted Hardy space Hα2\mathrm{H}^{2}_{\alpha} has property A for a positive integer nn, if there exists c1=c1(n)>0c_{1}=c_{1}(n)>0 and c2=c2(n)>0c_{2}=c_{2}(n)>0 such that

α(k)c1α(j+kn)\alpha(k)\leq c_{1}\alpha(j+kn)

and

α(j+kn)c2α(k)\alpha(j+kn)\leq c_{2}\alpha(k)

for any j=0,1,,n1j=0,1,\dots,n-1 and k=0,1,k=0,1,\dots.

Inspired by [1], we give the following lemma.

Lemma 5.1.

Let Hα2\mathrm{H}^{2}_{\alpha} be a weighted Hardy space with property A for a positive integer nn. Then for each j=0,1,,n1j=0,1,\dots,n-1, the map

Xα,j:f=k=0f^(k)zkHα2k=0f^(k)zj+knHα,j2X_{\alpha,j}\colon f=\sum_{k=0}^{\infty}\hat{f}(k)z^{k}\in\mathrm{H}^{2}_{\alpha}\to\sum_{k=0}^{\infty}\hat{f}(k)z^{j+kn}\in\mathrm{H}^{2}_{\alpha,j}

is an invertible bounded linear operator. Moreover, for each j=0,1,,n1j=0,1,\dots,n-1, Mα,zn|Hα,j2M_{\alpha,z^{n}}|_{\mathrm{H}^{2}_{\alpha,j}} Xα,jX_{\alpha,j} == Xα,jX_{\alpha,j} Mα,zM_{\alpha,z}. Thus, Mα,znM_{\alpha,z^{n}} is similar to 1nMα,z\oplus_{1}^{n}M_{\alpha,z}.

Proof.

Let f=k=0f^(k)zkHα2f=\sum_{k=0}^{\infty}\hat{f}(k)z^{k}\in\mathrm{H}^{2}_{\alpha}, then

k=0f^(k)zj+kn2=k=0|f^(k)|2α(j+kn)2c22k=0|f^(k)|2α(k)2=c22f2,\begin{split}\|\sum_{k=0}^{\infty}\hat{f}(k)z^{j+kn}\|^{2}&=\sum_{k=0}^{\infty}|\hat{f}(k)|^{2}\alpha(j+kn)^{2}\\ &\leq c_{2}^{2}\sum_{k=0}^{\infty}|\hat{f}(k)|^{2}\alpha(k)^{2}=c_{2}^{2}\|f\|^{2},\end{split}

which implies g=k=0f^(k)zj+knHα,j2g=\sum_{k=0}^{\infty}\hat{f}(k)z^{j+kn}\in\mathrm{H}^{2}_{\alpha,j} and gc2f\|g\|\leq c_{2}\|f\|. Thus, it can be seen that Xα,jX_{\alpha,j} is not only well defined but also a bounded linear operator. Obviously, Xα,jX_{\alpha,j} is an injection.

Let g=k=0g~(k)zj+knHα,j2g=\sum_{k=0}^{\infty}\tilde{g}(k)z^{j+kn}\in\mathrm{H}^{2}_{\alpha,j}, then

k=0g~(k)zk2=k=0|g~(k)|2α(k)2c12k=0|g~(k)|2α(j+kn)2=c12g2,\begin{split}\|\sum_{k=0}^{\infty}\tilde{g}(k)z^{k}\|^{2}&=\sum_{k=0}^{\infty}|\tilde{g}(k)|^{2}\alpha(k)^{2}\\ &\leq c_{1}^{2}\sum_{k=0}^{\infty}|\tilde{g}(k)|^{2}\alpha(j+kn)^{2}=c_{1}^{2}\|g\|^{2},\end{split}

which implies that f=k=0g~(k)zkHα2f=\sum_{k=0}^{\infty}\tilde{g}(k)z^{k}\in\mathrm{H}^{2}_{\alpha} and g=Xα,jfg=X_{\alpha,j}f. Thus, Xα,jX_{\alpha,j} is a surjection. Therefore, Xα,jX_{\alpha,j} an invertible operator.

The remaining conclusions can be easily checked. ∎

Remark 5.2.

Form the assumption (2.1), we can see that a weighted Hardy space Hα2\mathrm{H}^{2}_{\alpha} has property A for a positive integer nn if and only if there exists c1=c1(n)>0c_{1}=c_{1}(n)>0 and c2=c2(n)>0c_{2}=c_{2}(n)>0 such that

c1α(n1+kn)α(k)c2c_{1}\leq\frac{\alpha(n-1+kn)}{\alpha(k)}\leq c_{2}

for any k=0,1,k=0,1,\dots.

Remark 5.3.

For each λ\lambda\in\mathbb{R}, the weighted Hardy spaces Hαλ2=H(λ)\mathrm{H}^{2}_{\alpha_{\lambda}}=\mathrm{H}_{(\lambda)} has property A for any positive integer nn. In fact, this follows Remark 5.2 since

αλ(n1+kn)αλ(k)=nλ\frac{\alpha_{\lambda}(n-1+kn)}{\alpha_{\lambda}(k)}=n^{\lambda}

for any k=0,1,k=0,1,\dots.

Remark 5.4.

Each weighted Hardy space Hα2\mathrm{H}^{2}_{\alpha} of polynomial growth has property A for any positive integer nn.

Proof.

The weighted Hardy space Hα2\mathrm{H}^{2}_{\alpha} is of polynomial growth means that there exists NN\in\mathbb{N} such that for each kk\in\mathbb{N},

k+1k+N+1α(k)α(k1)k+N+1k+1\frac{k+1}{k+N+1}\leq\frac{\alpha(k)}{\alpha(k-1)}\leq\frac{k+N+1}{k+1}

(see [11]). Then for each k=0,1,k=0,1,\dots, whenever n1+kn>kn-1+kn>k,

j=k+1n1+knj+1j+N+1α(n1+kn)α(k)=j=k+1n1+knα(j)α(j1)j=k+1n1+knj+N+1j+1.\prod_{j=k+1}^{n-1+kn}\frac{j+1}{j+N+1}\leq\frac{\alpha(n-1+kn)}{\alpha(k)}=\prod_{j=k+1}^{n-1+kn}\frac{\alpha(j)}{\alpha(j-1)}\leq\prod_{j=k+1}^{n-1+kn}\frac{j+N+1}{j+1}.

Form

j=k+1n1+knj+N+1j+1=j=1Nn(k+1)+jk+1+jj=1Nn=nN,\prod_{j=k+1}^{n-1+kn}\frac{j+N+1}{j+1}=\prod_{j=1}^{N}\frac{n(k+1)+j}{k+1+j}\leq\prod_{j=1}^{N}n=n^{N},

we get

1nNα(n1+kn)α(k)nN.\frac{1}{n^{N}}\leq\frac{\alpha(n-1+kn)}{\alpha(k)}\leq n^{N}.

The last inequality holds for each k=0,1,k=0,1,\dots. Thus, this remark follows Remark 5.2. ∎

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two weighted Hardy spaces. It is not hard to check that the following conditions are equivalent:

(a) Hα2Hβ2\mathrm{H}^{2}_{\alpha}\subseteq\mathrm{H}^{2}_{\beta};

(b) 1Mult(Hα2,Hβ2)1\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta});

(c) supk0β(k)α(k)<\sup_{k\geq 0}\frac{\beta(k)}{\alpha(k)}<\infty.

If one of the conditions above is true, the multiplication operator Mα,β,1M_{\alpha,\beta,1} is the inclusion mapping ι:Hα2Hβ2\iota\colon\mathrm{H}^{2}_{\alpha}\to\mathrm{H}^{2}_{\beta}. In addition, Hol(𝔻¯)Mult(Hα2)Mult(Hα2,Hβ2)\mathrm{Hol}(\overline{\mathbb{D}})\subseteq\mathrm{Mult}(\mathrm{H}^{2}_{\alpha})\subseteq\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}).

Remark 5.5.

Let λ,μ\lambda,\mu\in\mathbb{R}. The weighted Hardy spaces Hαλ2=H(λ)\mathrm{H}^{2}_{\alpha_{\lambda}}=\mathrm{H}_{(\lambda)} and Hαμ2=H(μ)\mathrm{H}^{2}_{\alpha_{\mu}}=\mathrm{H}_{(\mu)} satisfy H(λ)H(μ)\mathrm{H}_{(\lambda)}\subseteq\mathrm{H}_{(\mu)} if and only if λμ0\lambda-\mu\geq 0.

Proposition 5.6.

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two weighted Hardy spaces with property A for a positive integer nn. Suppose Hα2Hβ2\mathrm{H}^{2}_{\alpha}\subseteq\mathrm{H}^{2}_{\beta}. Denote T=Tα,β,1T=T_{\alpha,\beta,1}, then TnT^{n} is similar to T(j=1n1T~)T\oplus(\oplus_{j=1}^{n-1}\tilde{T}), where T~=Tα,β,z\tilde{T}=T_{\alpha,\beta,z}.

Proof.

It is enough to prove (T)nT(j=1n1T~)(T^{*})^{n}\sim T^{*}\oplus(\oplus_{j=1}^{n-1}\tilde{T}^{*}), where

T=(Mα,z0Mα,β,1Mβ,z),T~=(Mα,z0Mα,β,zMβ,z).T^{*}=\begin{pmatrix}M_{\alpha,z}&0\\ M_{\alpha,\beta,1}&M_{\beta,z}\\ \end{pmatrix},\ \tilde{T}^{*}=\begin{pmatrix}M_{\alpha,z}&0\\ M_{\alpha,\beta,z}&M_{\beta,z}\\ \end{pmatrix}.

From Mα,β,1Mα,z=Mβ,zMα,β,1M_{\alpha,\beta,1}M_{\alpha,z}=M_{\beta,z}M_{\alpha,\beta,1}, it can be calculated that

(T)n=(Mα,zn0nMα,β,zn1Mβ,zn).(T^{*})^{n}=\begin{pmatrix}M_{\alpha,z^{n}}&0\\ nM_{\alpha,\beta,z^{n-1}}&M_{\beta,z^{n}}\end{pmatrix}.

Since

(I001nI)(T)n(I001nI)1=(Mα,zn0Mα,β,zn1Mβ,zn)=:A,\begin{pmatrix}I&0\\ 0&\frac{1}{n}I\\ \end{pmatrix}(T^{*})^{n}\begin{pmatrix}I&0\\ 0&\frac{1}{n}I\\ \end{pmatrix}^{-1}=\begin{pmatrix}M_{\alpha,z^{n}}&0\\ M_{\alpha,\beta,z^{n-1}}&M_{\beta,z^{n}}\\ \end{pmatrix}=:A,

it followed that (T)nA(T^{*})^{n}\sim A.

Let

K0\displaystyle K_{0} =Hα,02Hβ,n12,\displaystyle=\mathrm{H}^{2}_{\alpha,0}\oplus\mathrm{H}^{2}_{\beta,n-1},
K1\displaystyle K_{1} =Hα,12Hβ,02,\displaystyle=\mathrm{H}^{2}_{\alpha,1}\oplus\mathrm{H}^{2}_{\beta,0},
\displaystyle\cdots
Kn1\displaystyle K_{n-1} =Hα,n12Hβ,n22,\displaystyle=\mathrm{H}^{2}_{\alpha,n-1}\oplus\mathrm{H}^{2}_{\beta,n-2},

then

Hα2Hβ2=K0K1Kn1\mathrm{H}^{2}_{\alpha}\oplus\mathrm{H}^{2}_{\beta}=K_{0}\dotplus K_{1}\dotplus\cdots\dotplus K_{n-1}

is Hilbert direct sum. One can check that Hα,j2\mathrm{H}^{2}_{\alpha,j} and Hβ,j2\mathrm{H}^{2}_{\beta,j} are invariant subspace of Mα,znM_{\alpha,z^{n}} and Mβ,znM_{\beta,z^{n}}, respectively, and that

Mα,β,zn1(Hα,02)\displaystyle M_{\alpha,\beta,z^{n-1}}(\mathrm{H}^{2}_{\alpha,0}) Hβ,n12,\displaystyle\subseteq\mathrm{H}^{2}_{\beta,n-1},
Mα,β,zn1(Hα,12)\displaystyle M_{\alpha,\beta,z^{n-1}}(\mathrm{H}^{2}_{\alpha,1}) Hβ,02,\displaystyle\subseteq\mathrm{H}^{2}_{\beta,0},
\displaystyle\cdots
Mα,β,zn1(Hα,n12)\displaystyle M_{\alpha,\beta,z^{n-1}}(\mathrm{H}^{2}_{\alpha,n-1}) Hβ,n22.\displaystyle\subseteq\mathrm{H}^{2}_{\beta,n-2}.

Hence, K0,K1,,Kn1K_{0},K_{1},\dots,K_{n-1} are invariant subspace of AA.

To complete the proof, it is enough to prove that

A|K0T,A|K1T~,,A|Kn1T~.A|_{K_{0}}\sim T^{*},\;A|_{K_{1}}\sim\tilde{T}^{*},\;\dots,\;A|_{K_{n-1}}\sim\tilde{T}^{*}.

This implies that (T)nAT(j=1n1T~)(T^{*})^{n}\sim A\sim T^{*}\oplus(\oplus_{j=1}^{n-1}\tilde{T}^{*}).

It is not hard to see that

A|K0\displaystyle A|_{K_{0}} =(Mα,zn|Hα,020PHβ,n12Mα,β,zn1PHα,02Mβ,zn|Hβ,n12),\displaystyle=\begin{pmatrix}M_{\alpha,z^{n}}|_{\mathrm{H}^{2}_{\alpha,0}}&0\\ P_{\mathrm{H}^{2}_{\beta,n-1}}M_{\alpha,\beta,z^{n-1}}P_{\mathrm{H}^{2}_{\alpha,0}}&M_{\beta,z^{n}}|_{\mathrm{H}^{2}_{\beta,n-1}}\\ \end{pmatrix},
A|K1\displaystyle A|_{K_{1}} =(Mα,zn|Hα,120PHβ,02Mα,β,zn1PHα,12Mβ,zn|Hβ,02),\displaystyle=\begin{pmatrix}M_{\alpha,z^{n}}|_{\mathrm{H}^{2}_{\alpha,1}}&0\\ P_{\mathrm{H}^{2}_{\beta,0}}M_{\alpha,\beta,z^{n-1}}P_{\mathrm{H}^{2}_{\alpha,1}}&M_{\beta,z^{n}}|_{\mathrm{H}^{2}_{\beta,0}}\\ \end{pmatrix},
\displaystyle\cdots
A|Kn1\displaystyle A|_{K_{n-1}} =(Mα,zn|Hα,n120PHβ,n22Mα,β,zn1PHα,n12Mβ,zn|Hβ,n22).\displaystyle=\begin{pmatrix}M_{\alpha,z^{n}}|_{\mathrm{H}^{2}_{\alpha,n-1}}&0\\ P_{\mathrm{H}^{2}_{\beta,n-2}}M_{\alpha,\beta,z^{n-1}}P_{\mathrm{H}^{2}_{\alpha,n-1}}&M_{\beta,z^{n}}|_{\mathrm{H}^{2}_{\beta,n-2}}\\ \end{pmatrix}.

Lemma 5.1 shows that for any j=0,1,,n1j=0,1,\dots,n-1, the maps Xα,j:Hα2X_{\alpha,j}\colon\mathrm{H}^{2}_{\alpha}Hα,j2\to\mathrm{H}^{2}_{\alpha,j} and Xβ,j:Hβ2X_{\beta,j}\colon\mathrm{H}^{2}_{\beta}Hβ,j2\to\mathrm{H}^{2}_{\beta,j} are invertible operators. Take the invertible operators

X0=(Xα,000Xβ,n1),X1=(Xα,100Xβ,0),,Xn1=(Xα,n100Xβ,n2).X_{0}=\begin{pmatrix}X_{\alpha,0}&0\\ 0&X_{\beta,n-1}\\ \end{pmatrix},\,X_{1}=\begin{pmatrix}X_{\alpha,1}&0\\ 0&X_{\beta,0}\\ \end{pmatrix},\,\dots,\,X_{n-1}=\begin{pmatrix}X_{\alpha,n-1}&0\\ 0&X_{\beta,n-2}\\ \end{pmatrix}.

By calculation we obtain that Mα,zn|Hα,j2Xα,j=Xα,jMα,zM_{\alpha,z^{n}}|_{\mathrm{H}^{2}_{\alpha,j}}X_{\alpha,j}=X_{\alpha,j}M_{\alpha,z}, Mβ,zn|Hβ,j2Xβ,j=Xβ,jMβ,zM_{\beta,z^{n}}|_{\mathrm{H}^{2}_{\beta,j}}X_{\beta,j}=X_{\beta,j}M_{\beta,z} and that

PHβ,n12Mα,β,zn1PHα,02Xα,0\displaystyle P_{\mathrm{H}^{2}_{\beta,n-1}}M_{\alpha,\beta,z^{n-1}}P_{\mathrm{H}^{2}_{\alpha,0}}X_{\alpha,0} =Xβ,n1Mα,β,1,\displaystyle=X_{\beta,n-1}M_{\alpha,\beta,1},
PHβ,02Mα,β,zn1PHα,12Xα,1\displaystyle P_{\mathrm{H}^{2}_{\beta,0}}M_{\alpha,\beta,z^{n-1}}P_{\mathrm{H}^{2}_{\alpha,1}}X_{\alpha,1} =Xβ,0Mα,β,z,\displaystyle=X_{\beta,0}M_{\alpha,\beta,z},
\displaystyle\cdots
PHβ,n22Mα,β,zn1PHα,n12Xα,n1\displaystyle P_{\mathrm{H}^{2}_{\beta,n-2}}M_{\alpha,\beta,z^{n-1}}P_{\mathrm{H}^{2}_{\alpha,n-1}}X_{\alpha,n-1} =Xβ,n2Mα,β,z.\displaystyle=X_{\beta,n-2}M_{\alpha,\beta,z}.

The above equation indicates that

A|K0X0=X0T,A|K1X1=X1T~,,A|Kn1Xn1=Xn1T~,A|_{K_{0}}X_{0}=X_{0}T^{*},\ A|_{K_{1}}X_{1}=X_{1}\tilde{T}^{*},\ \dots,\ A|_{K_{n-1}}X_{n-1}=X_{n-1}\tilde{T}^{*},

and hence

A|K0T,A|K1T~,,A|Kn1T~.A|_{K_{0}}\sim T^{*},\;A|_{K_{1}}\sim\tilde{T}^{*},\;\dots,\;A|_{K_{n-1}}\sim\tilde{T}^{*}.

5.2. Dissimilarity

In the remaining part of this section, we need some conclusions related to strongly irreducible operators (‘strongly irreducible’ is also abbreviated as ‘(SI)’).

Denote K0()\mathrm{K}_{0}(\mathcal{B}) for the K0\mathrm{K}_{0}-group of a Banach algebra \mathcal{B}. For a bounded linear operator AA on a Hilbert space HH, denote {A}\left\{A\right\}^{\prime} for the commutant algebra of AA and A(n)A^{(n)} for the nn copies nnA\oplus_{n}^{n}A of AA. Then the following theorem is Theorem 1 in [15].

Theorem 5.7.

Let T=A1(n1)A2(n2)Ak(nk)T=A_{1}^{(n_{1})}\oplus A_{2}^{(n_{2})}\oplus\cdots\oplus A_{k}^{(n_{k})}, where A1,A2,,AkA_{1},A_{2},\dots,A_{k} are strongly irreducible Cowen-Douglas operators, AiA_{i} and AjA_{j} are not similar whenever iji\neq j, and n1,n2,,nkn_{1},n_{2},\dots,n_{k} are positive integers. Then K0({T})k\mathrm{K}_{0}(\left\{T\right\}^{\prime})\cong\mathbb{Z}^{k}.

The following proposition, which is Proposition 2.22 in [13], gives a characterization of strong irreducibility in FB2(Ω)FB_{2}(\Omega).

Lemma 5.8.

An operator

T=(T0S0T1)T=\begin{pmatrix}T_{0}&S\\ 0&T_{1}\end{pmatrix}

in FB2(Ω)FB_{2}(\Omega) is strongly irreducible if and only if SRanσT0,T1S\notin\mathrm{Ran\,}\sigma_{T_{0},T_{1}}.

Lemma 5.9.

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two weighted Hardy spaces, where the weight sequences α\alpha and β\beta satisfy condition (3.1). Then

(a) for any non-zero function hh in Mult(Hα2,Hβ2)\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}), the operator Tα,β,hT_{\alpha,\beta,h} is strongly irreducible.

(b) for any non-zero functions hh and h~\tilde{h} in Mult(Hα2,Hβ2)\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}) such that hh and h~\tilde{h} have different zero points on 𝔻\mathbb{D}, the operators Tα,β,hT_{\alpha,\beta,h} and Tα,β,h~T_{\alpha,\beta,\tilde{h}} are not similar.

Proof.

(a) Since σMα,z,Mβ,z(Mα,β,h)=0\sigma_{M_{\alpha,z}^{*},M_{\beta,z}^{*}}(M_{\alpha,\beta,h}^{*})=0, it follows that Mα,β,hRanσMα,z,Mβ,zM_{\alpha,\beta,h}^{*}\notin\mathrm{Ran\,}\sigma_{M_{\alpha,z}^{*},M_{\beta,z}^{*}} according to Proposition 3.1. Then Tα,β,hT_{\alpha,\beta,h} is strongly irreducible by Lemma 5.8.

(b) Clearly, this is a corollary of Proposition 3.4. ∎

Proposition 5.10.

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two weighted Hardy spaces with property A for a positive integer n2n\geq 2. Suppose Hα2Hβ2\mathrm{H}^{2}_{\alpha}\subseteq\mathrm{H}^{2}_{\beta} and suppose the weight sequences α\alpha and β\beta satisfy condition (3.1). Denote T=Tα,β,1T=T_{\alpha,\beta,1}, then TnT^{n} is not similar to 1nT\oplus_{1}^{n}T.

Proof.

Proposition 5.6 shows that TnT(j=1n1T~)=T(1)T~(n1)T^{n}\sim T\oplus(\oplus_{j=1}^{n-1}\tilde{T})=T^{(1)}\oplus\tilde{T}^{(n-1)}, where T~=Tα,β,z\tilde{T}=T_{\alpha,\beta,z}. Lemma 5.9 shows that TT and T~\tilde{T} are strongly irreducible Cowen-Douglas operators and TT~T\nsim\tilde{T}. Assume TnT^{n} is similar to 1nT=T(n)\oplus_{1}^{n}T=T^{(n)}, then

K0({T(1)T~(n1)})K0({Tn})K0({T(n)}).\mathrm{K}_{0}(\left\{T^{(1)}\oplus\tilde{T}^{(n-1)}\right\}^{\prime})\cong\mathrm{K}_{0}(\left\{T^{n}\right\}^{\prime})\cong\mathrm{K}_{0}(\left\{T^{(n)}\right\}^{\prime}).

But from Theorem 5.7, it follows that

K0({T(1)T~(n1)})2,K0({T(n)}).\mathrm{K}_{0}(\left\{T^{(1)}\oplus\tilde{T}^{(n-1)}\right\}^{\prime})\cong\mathbb{Z}^{2},\ \ \mathrm{K}_{0}(\left\{T^{(n)}\right\}^{\prime})\cong\mathbb{Z}.

So we get a contradiction since 2\mathbb{Z}^{2}\ncong\mathbb{Z}. ∎

6. Main results

In this section, we complete the proof of the main theorem and give some concrete examples.

Proof of Theorem 1.1.

Suppose h0h\neq 0. Since the functions in Aut(𝔻)\mathrm{Aut\,}(\mathbb{D}) are finite Blaschke products with order 11, TT is weakly homogeneous. Since supk0β(k)α(k)<\sup_{k\geq 0}\frac{\beta(k)}{\alpha(k)}<\infty, i.e. Hα2Hβ2\mathrm{H}^{2}_{\alpha}\subseteq\mathrm{H}^{2}_{\beta}, Hol(𝔻¯)Mult(Hα2)Mult(Hα2,Hβ2)\mathrm{Hol}(\overline{\mathbb{D}})\subseteq\mathrm{Mult}(\mathrm{H}^{2}_{\alpha})\subseteq\mathrm{Mult}(\mathrm{H}^{2}_{\alpha},\mathrm{H}^{2}_{\beta}). Hence, according to Proposition 4.3, hh is non-zero everywhere on 𝔻¯\overline{\mathbb{D}}, which implies 1hHol(𝔻¯)\frac{1}{h}\in\mathrm{Hol}(\overline{\mathbb{D}}). Thus, h,1hMult(Hα2)h,\frac{1}{h}\in\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}). From Remark 3.5, we have T=Tα,β,hTα,β,1=:T^T=T_{\alpha,\beta,h}\sim T_{\alpha,\beta,1}=:\hat{T}. Since B(z)=zn0,z𝔻B(z)=z^{n_{0}},z\in\mathbb{D} is a finite Blaschke product with order n0n_{0}, it follows that Tn0=B(T)1n0TT^{n_{0}}=B(T)\sim\oplus_{1}^{n_{0}}T and hence T^n01n0T^\hat{T}^{n_{0}}\sim\oplus_{1}^{n_{0}}\hat{T}. But this contradicts the conclusion of Proposition 5.10 that T^n01n0T^\hat{T}^{n_{0}}\nsim\oplus_{1}^{n_{0}}\hat{T}. ∎

Corollary 6.1.

Let λμ[0,1)\lambda-\mu\in[0,1). Let

T=(MzMh0Mz),T=\begin{pmatrix}M_{z}^{*}&M_{h}^{*}\\ 0&M_{z}^{*}\\ \end{pmatrix},

on H(λ)H(μ)\mathrm{H}_{(\lambda)}\oplus\mathrm{H}_{(\mu)}, where hHol(𝔻¯)h\in\mathrm{Hol}(\overline{\mathbb{D}}). If for any finite Blaschke product BB, B(T)B(T) is similar to 1nT\oplus_{1}^{n}T, where nn is the order of BB, then h=0h=0.

Proof.

This corollary now follows Remarks 3.2, 4.1, 5.3, 5.5 and Theorem 1.1. ∎

Corollary 6.2.

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two weighted Hardy spaces of polynomial growth. Suppose the weight sequences α\alpha and β\beta satisfy

supk0β(k)α(k)<andlimkα(k)kβ(k)=0.\sup_{k\geq 0}\frac{\beta(k)}{\alpha(k)}<\infty\quad\text{and}\quad\lim_{k\to\infty}\frac{\alpha(k)}{k\beta(k)}=0.

Denote T=Tα,β,hT=T_{\alpha,\beta,h}, where hHol(𝔻¯)h\in\mathrm{Hol}(\overline{\mathbb{D}}). If for any finite Blaschke product BB, B(T)B(T) is similar to 1nT\oplus_{1}^{n}T, where nn is the order of BB, then h=0h=0.

Proof.

This corollary now follows Remark 4.2, 5.4 and Theorem 1.1. ∎

For many weighted Hardy space Hα2\mathrm{H}^{2}_{\alpha}, the multiplier algebra Mult(Hα2)\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}) has better properties, such as Mult(Hα2)=H(𝔻)\mathrm{Mult}(\mathrm{H}^{2}_{\alpha})=\mathrm{H}^{\infty}(\mathbb{D}). In Theorem 1.1, if a higher condition is required for multiplier algebra of weighted Hardy space, a stronger conclusion can be obtained.

Theorem 6.3.

Let Hα2\mathrm{H}^{2}_{\alpha} and Hβ2\mathrm{H}^{2}_{\beta} be two Mo¨\ddot{o}bius invariant weighted Hardy spaces with property A for some integer n02n_{0}\geq 2. Assume 𝒞(𝔻¯)Hol(𝔻)Mult(Hα2)\mathcal{C}(\overline{\mathbb{D}})\cap\mathrm{Hol}(\mathbb{D})\subseteq\mathrm{Mult}(\mathrm{H}^{2}_{\alpha}). Suppose the weight sequences α\alpha and β\beta satisfy

supk0β(k)α(k)<andlimkα(k)kβ(k)=0.\sup_{k\geq 0}\frac{\beta(k)}{\alpha(k)}<\infty\quad\text{and}\quad\lim_{k\to\infty}\frac{\alpha(k)}{k\beta(k)}=0.

Denote T=Tα,β,hT=T_{\alpha,\beta,h}, where h𝒞(𝔻¯)Hol(𝔻)h\in\mathcal{C}(\overline{\mathbb{D}})\cap\mathrm{Hol}(\mathbb{D}). If for any finite Blaschke product BB, B(T)B(T) is similar to 1nT\oplus_{1}^{n}T, where nn is the order of BB, then h=0h=0.

Proof.

Similar to the proof of Theorem 1.1. ∎

Corollary 6.4.

Let H2(𝔻)\mathrm{H}^{2}(\mathbb{D}) be the classical Hardy space and La2(𝔻)\mathrm{L}^{2}_{a}(\mathbb{D}) be the classical Bergman space. Let

T=(MzMh0Mz),T=\begin{pmatrix}M_{z}^{*}&M_{h}^{*}\\ 0&M_{z}^{*}\\ \end{pmatrix},

on H2(𝔻)La2(𝔻)\mathrm{H}^{2}(\mathbb{D})\oplus\mathrm{L}^{2}_{a}(\mathbb{D}), where h𝒞(𝔻¯)Hol(𝔻)h\in\mathcal{C}(\overline{\mathbb{D}})\cap\mathrm{Hol}(\mathbb{D}). If for any finite Blaschke product BB, B(T)B(T) is similar to 1nT\oplus_{1}^{n}T, where nn is the order of BB, then h=0h=0.

Proof.

As will known, Mult(H2(𝔻))\mathrm{Mult}(\mathrm{H}^{2}(\mathbb{D})) and Mult(La2(𝔻))\mathrm{Mult}(\mathrm{L}^{2}_{a}(\mathbb{D})) are both H(𝔻)H^{\infty}(\mathbb{D}). Note that H2(𝔻)\mathrm{H}^{2}(\mathbb{D}) is H(0)\mathrm{H}_{(0)} and La2(𝔻)\mathrm{L}^{2}_{a}(\mathbb{D}) is H(12)\mathrm{H}_{(-\frac{1}{2})}. Then this corollary now follows Remarks 3.2, 4.1, 5.3, 5.5 and Theorem 6.3. ∎

References

  • [1] M. F. Ahmadi and K. Hedayatian, On similarity of powers of shift operators, Turkish Journal of Mathematics 36 (2012), 596–600.
  • [2] D. N. Clark and G. Misra, On homogeneous contractions and unitary representations of su(1,1), Journal of Operator Theory 30 (1993), no. 1, 109–122.
  • [3] C. C. Cowen, On equivalence of toeplitz operators, Journal of Operator Theory 7 (1982), no. 1, 167–172.
  • [4] C. C. Cowen and B. D. MacCluer, Composition operators on spaces of analytic functions, Studies in Advanced Mathematics, CRC Press, Boca Raton, FL, 1995.
  • [5] M. J. Cowen and R. G. Douglas, Complex geometry and operator theory, Acta Mathematica 141 (1978), no. 1, 187–261.
  • [6] R. G. Douglas, Operator theory and complex geometry, Extracta Mathematicae 24 (2009), no. 2, 135–165.
  • [7] E. Fricain, J. Mashreghi, and D. Seco, Cyclicity in reproducing kernel hilbert spaces of analytic functions, Computational Methods and Function Theory 14 (2013), 665 – 680.
  • [8] S. Ghara, The orbit of a bounded operator under the möbius group modulo similarity equivalence, Israel Journal of Mathematics (2018), 1–41.
  • [9] K. Guo and H. Huang, On multiplication operators on the bergman space: Similarity, unitary equivalence and reducing subspaces, Journal of Operator Theory 65 (2011), no. 2, 355–378.
  • [10] K. Guo, S. Sun, D. Zheng, and C. Zhong, Multiplication operators on the bergman space via the hardy space of the bidisk, Journal für die reine und angewandte Mathematik 2009 (2009), no. 628, 129–168.
  • [11] B. Hou and C. Jiang, Analytic automorphism group and similar representation of analytic functions, 2023, arXiv:2209.03852.
  • [12] J. Hu, S. Sun, X. Xu, and D. Yu, Reducing subspace of analytic toeplitz operators on the bergman space, Integral Equations and Operator Theory 49 (2004), no. 3, 387–395.
  • [13] K. Ji, C. Jiang, D. K. Keshari, and G. Misra, Rigidity of the flag structure for a class of cowen–douglas operators, Journal of Functional Analysis 272 (2017), no. 7, 2899–2932.
  • [14] K. Ji and R. Shi, Similarity of multiplication operators on the sobolev disk algebra, Acta Mathematica Sinica, English Series 29 (2013), 789–800.
  • [15] C. Jiang, X. Guo, and K. Ji, K-group and similarity classification of operators, Journal of Functional Analysis 225 (2005), no. 1, 167–192.
  • [16] C. Jiang, K. Ji, and D. K. Keshari, Geometric similarity invariants of cowen-douglas operators., J. Noncommut. Geom. 17 (2023), no. 2, 573–608.
  • [17] C. Jiang and Y. Li, The commutant and similarity invariant of analytic toeplitz operators on bergman space, Science in China Series A: Mathematics 50 (2007), 651–664.
  • [18] C. Jiang and D. Zheng, Similarity of analytic toeplitz operators on the bergman spaces, Journal of Functional Analysis 258 (2010), no. 9, 2961–2982.
  • [19] Y. Li and P. Ma, On similarity and commutant of a class of multiplication operators on the dirichlet space, Banach Journal of Mathematical Analysis 17 (2022), no. 1, 9.
  • [20] Z. Wang, R. Zhao, and Y. Jin, Finite blaschke product and the multiplication operators on sobolev disk algebra, Science in China Series A: Mathematics 52 (2009), 142–146.
  • [21] K. Zhu, Reducing subspaces for a class of multiplication operators, Journal of the London Mathematical Society 62 (2000), no. 2, 553–568.
  • [22] N. Zorboska, Composition operators on SaS_{a} spaces, Indiana University Mathematics Journal 39 (1990), no. 3, 847–857.