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

Compact Differences of Composition Operators
on Weighted Dirichlet Spaces

Robert F. Allen1, Katherine C. Heller2, and Matthew A. Pons2 1Department of Mathematics and Statistics, University of Wisconsin-La Crosse 2Department of Mathematics, North Central College rallen@@uwlax.edu, kheller@noctrl.edu, mapson@noctrl.edu
Abstract.

Here we consider when the difference of two composition operators is compact on the weighted Dirichlet spaces 𝒟α\displaystyle\mathcal{D}_{\alpha}. Specifically we study differences of composition operators on the Dirichlet space 𝒟\displaystyle\mathcal{D} and S2\displaystyle S^{2}, the space of analytic functions whose first derivative is in H2\displaystyle H^{2}, and then use Calderón’s complex interpolation to extend the results to the general weighted Dirichlet spaces. As a corollary we consider composition operators induced by linear fractional self-maps of the disk.

Key words and phrases:
Composition operator; Compact difference; Weighted Dirichlet space; Complex interpolation.
2010 Mathematics Subject Classification:
primary: 47B33; secondary: 46E20, 47B32

1. Introduction

For an analytic self-map φ\displaystyle\varphi of the unit disk 𝔻\displaystyle\mathbb{D} and a Banach space 𝒴\displaystyle\mathcal{Y} of functions analytic on the unit disk, we define the composition operator Cφ\displaystyle C_{\varphi} with symbol φ\displaystyle\varphi by the rule Cφf=fφ\displaystyle C_{\varphi}f=f\circ\varphi for all f𝒴\displaystyle f\in\mathcal{Y}. The study of these types of operators began formally with Nordgren’s paper [15] where he explored properties of composition operators acting on the Hardy Hilbert space H2\displaystyle H^{2}. Over the past fifty years, the study has proved to be a lively source of inquiry, most likely due to the fact that the study of such operators lies at the intersection of complex function theory and operator theory. With this perspective, the goal of such an investigation seeks to relate the operator properties of Cφ\displaystyle C_{\varphi} to the analytic and geometric properties of the symbol function φ\displaystyle\varphi. In this note we will focus on the property of compactness.

Recall that an operator T\displaystyle T acting on a Banach space 𝒴\displaystyle\mathcal{Y} is compact if it takes the unit ball in 𝒴\displaystyle\mathcal{Y} (which is not compact in the infinite dimensional setting) into a set with compact closure. For composition operators, compactness is generally classified by how the symbol function behaves near the boundary of the disk. For instance, it is well known that a composition operator Cφ\displaystyle C_{\varphi} is compact on H(𝔻)\displaystyle H^{\infty}(\mathbb{D}) if and only if φ<1\displaystyle\|\varphi\|_{\infty}<1. This condition is sufficient on many other spaces, but is often not necessary, meaning that the symbol can have some contact with the boundary and still induce a compact composition operator. This phenomenon has been studied in depth on the Hardy and weighted Bergman spaces; in [12] the authors supply the intuitive message for the case of the Hardy spaces: “Cφ\displaystyle C_{\varphi} will be compact on Hp\displaystyle H^{p} if and only if φ\displaystyle\varphi squeezes the unit disc rather sharply into itself”. They then make sense of this intuitive notion using the finite angular derivative of the symbol φ\displaystyle\varphi.

Shapiro and Taylor were the first to observe the role that the angular derivative plays in the study of the compactness problem. In [21] they showed that the symbol of a compact composition operator Cφ\displaystyle C_{\varphi} on H2\displaystyle H^{2} cannot have finite angular derivative at any point of 𝔻\displaystyle\partial\mathbb{D}. This result was extended to the Bergman space A2\displaystyle A^{2} by Boyd in [4]. In [12], the authors use Carleson measure techniques to show that nonexistence of the angular derivative of φ\displaystyle\varphi is also a sufficient condition for Cφ\displaystyle C_{\varphi} to be compact on A2\displaystyle A^{2}, however this is not the case on H2\displaystyle H^{2} and the authors provide an example demonstrating such a φ\displaystyle\varphi. Shapiro later characterized the compact composition operators on H2\displaystyle H^{2} in terms of the Nevanlinna counting function and the essential norm of the operator in [19].

In contrast with this, for Cφ\displaystyle C_{\varphi} acting on the space S2\displaystyle S^{2}, the situation is much simpler. First, if Cφ\displaystyle C_{\varphi} is bounded on S2\displaystyle S^{2}, then φ\displaystyle\varphi must have finite angular derivative at any point in 𝔻\displaystyle\partial\mathbb{D} which is mapped to 𝔻\displaystyle\partial\mathbb{D}; see [7] Theorem 4.13. For compactness, it turns out that Cφ\displaystyle C_{\varphi} is compact on S2\displaystyle S^{2} if and only if φ<1\displaystyle\|\varphi\|_{\infty}<1; see, for example, [18]. Thus we see a drastic (and interesting) shift in behavior among spaces that are closely related to each other.

Here we are interested in determining when the difference of two composition operators is compact. In [11] MacCluer investigated this on the Hardy space to understand the topological structure of the collection of compact composition operators within the set of all composition operators. More recently, Moorhouse considered this on a broader range of spaces in [14] and further considered the role of the second order data of the symbol φ\displaystyle\varphi; Bourdon also considered this same question on the Hardy space in [3]. Here we aim to extend some of those results.

In the next section we gather the necessary prerequisites. In Section 3 we work primarily on the Dirichlet space and S2\displaystyle S^{2}. Our techniques mimic those of MacCluer and Moorhouse, but required a change in perspective due to the behavior of the reproducing kernels in our spaces of interest. To overcome this, we instead use the kernels for evaluation of the first derivative. Finally, we appeal to Calderón’s method of complex interpolation to provide an extension to the general weighted Dirichlet spaces. This work is, in part, an invitation for other researchers to employ these newer techniques to the study of composition operators.

2. Preliminaries

2.1. Spaces of analytic functions

We let 𝔻\displaystyle\mathbb{D} denote the open unit disk in the complex plane, 𝔻={z:|z|<1}\displaystyle\mathbb{D}=\left\{z\in\mathbb{C}:|z|<1\right\}, and let H(𝔻)\displaystyle H(\mathbb{D}) be the space of functions analytic on 𝔻\displaystyle\mathbb{D}. The following classical spaces of analytic functions have received much attention in the study of composition operators. The Hardy space is defined by

H2(𝔻)={f in H(𝔻):fH22=limr102π|f(reiθ)|2dθ2π<}\displaystyle H^{2}(\mathbb{D})=\left\{f\textup{ in }H(\mathbb{D}):\|f\|_{H^{2}}^{2}=\lim_{r\rightarrow 1^{-}}\int_{0}^{2\pi}|f(re^{i\theta})|^{2}\,\frac{d\theta}{2\pi}<\infty\right\}

where dθ\displaystyle d\theta is the Lebesgue arc-length measure on the unit circle. For β>1\displaystyle\beta>-1, the standard weighted Bergman space is defined by

Aβ2(𝔻)={f in H(𝔻):fAβ22=𝔻|f(z)|2(1|z|2)β𝑑A<}\displaystyle A_{\beta}^{2}(\mathbb{D})=\left\{f\textup{ in }H(\mathbb{D}):\|f\|_{A_{\beta}^{2}}^{2}=\int_{\mathbb{D}}|f(z)|^{2}(1-|z|^{2})^{\beta}\,dA<\infty\right\}

where dA\displaystyle dA is the Lebesgue area measure normalized so that A(𝔻)=1\displaystyle A(\mathbb{D})=1; the Dirichlet space is given by

𝒟(𝔻)={f in H(𝔻):f𝒟2=|f(0)|2+𝔻|f(z)|2𝑑A<}.\displaystyle\mathcal{D}(\mathbb{D})=\left\{f\textup{ in }H(\mathbb{D}):\|f\|_{\mathcal{D}}^{2}=|f(0)|^{2}+\int_{\mathbb{D}}|f^{\prime}(z)|^{2}\,dA<\infty\right\}.

Recall that a reproducing kernel Hilbert space \displaystyle\mathcal{H} with inner product ,\displaystyle\langle\cdot,\cdot\rangle_{\mathcal{H}} has the property that for each w𝔻\displaystyle w\in\mathbb{D}, there is a unique function Kw\displaystyle K_{w}\in\mathcal{H} such that

f(w)=f,Kw.\displaystyle f(w)=\langle f,K_{w}\rangle_{\mathcal{H}}.

For the Hardy and weighted Bergman spaces, the kernels have a similar form:

Kw(z)=11w¯z\displaystyle K_{w}(z)=\frac{1}{1-\overline{w}z}

on H2\displaystyle H^{2} and

Kw(z)=1(1w¯z)β+2\displaystyle K_{w}(z)=\frac{1}{(1-\overline{w}z)^{\beta+2}}

on Aβ2\displaystyle A_{\beta}^{2} with β>1\displaystyle\beta>-1. On the Dirichlet space the kernel takes on a more complicated form,

Kw(z)=1+log11w¯z,\displaystyle K_{w}(z)=1+\log\frac{1}{1-\overline{w}z},

where logz\displaystyle\log z denotes the principal branch of the logarithm.

Though often convenient from the computational point of view, presenting the norms for these spaces in terms of integrals obscures the relationship between the spaces, though it is somewhat revealed in the representations of the reproducing kernels. To make the relationship more explicit, we can consider the spaces with a series norm, (equal to the norm given above for the Hardy and Dirichlet spaces, but only equivalent to the Bergman norm):

H2(𝔻)={f(z)=n=0anzn in H(𝔻):n=0|an|2<};\displaystyle H^{2}(\mathbb{D})=\left\{f(z)=\sum_{n=0}^{\infty}a_{n}z^{n}\textup{ in }H(\mathbb{D}):\sum_{n=0}^{\infty}|a_{n}|^{2}<\infty\right\};
Aβ2(𝔻)={f(z)=n=0anzn in H(𝔻):|a0|2+n=1|an|2nβ+1<};\displaystyle A_{\beta}^{2}(\mathbb{D})=\left\{f(z)=\sum_{n=0}^{\infty}a_{n}z^{n}\textup{ in }H(\mathbb{D}):|a_{0}|^{2}+\sum_{n=1}^{\infty}\frac{|a_{n}|^{2}}{n^{\beta+1}}<\infty\right\};
𝒟(𝔻)={f(z)=n=0anzn in H(𝔻):|a0|2+n=1n|an|2<}.\displaystyle\mathcal{D}(\mathbb{D})=\left\{f(z)=\sum_{n=0}^{\infty}a_{n}z^{n}\textup{ in }H(\mathbb{D}):|a_{0}|^{2}+\sum_{n=1}^{\infty}n|a_{n}|^{2}<\infty\right\}.

With these characterizations, we see the obvious containment relationship 𝒟H2A2,\displaystyle\mathcal{D}\subset H^{2}\subset A^{2}, but more importantly it is apparent that there are other spaces that deserve consideration. One particular space that has received more attention as of late is S2\displaystyle S^{2} which can be defined with a series norm or an equal integral norm,

S2(𝔻)\displaystyle S^{2}(\mathbb{D}) ={f in H(𝔻):fS22=|f(0)|2+fH22<}\displaystyle=\left\{f\textup{ in }H(\mathbb{D}):\|f\|_{S^{2}}^{2}=|f(0)|^{2}+\|f^{\prime}\|_{H^{2}}^{2}<\infty\right\}
={f(z)=n=0anzn in H(𝔻):fS22=|a0|2+n=1n2|an|2<}.\displaystyle=\left\{f(z)=\sum_{n=0}^{\infty}a_{n}z^{n}\textup{ in }H(\mathbb{D}):\|f\|_{S^{2}}^{2}=|a_{0}|^{2}+\sum_{n=1}^{\infty}n^{2}|a_{n}|^{2}<\infty\right\}.

While this space is also a reproducing kernel Hilbert space, one of the first difficulties encountered in this setting is that there is not a “nice” closed form for the reproducing kernel functions with respect to this norm. The reason for this is the fact that, on S2\displaystyle S^{2}, the kernel for evaluation at w\displaystyle w takes the form

Kw(z)=1+n=1(w¯z)nn2,\displaystyle K_{w}(z)=1+\sum_{n=1}^{\infty}\frac{(\overline{w}z)^{n}}{n^{2}},

however we cannot identify this sum as an elementary function; for more on this, see [9]. We will discuss how to overcome this obstacle shortly.

In general, for α1\displaystyle\alpha\geq-1 we define the weighted Dirichlet space

𝒟α(𝔻)={f(z)=n=0anzn in H(𝔻):f𝒟α2=|a0|2+n=1n1α|an|2<}.\displaystyle\mathcal{D}_{\alpha}(\mathbb{D})=\left\{f(z)=\sum_{n=0}^{\infty}a_{n}z^{n}\textup{ in }H(\mathbb{D}):\|f\|_{\mathcal{D}_{\alpha}}^{2}=|a_{0}|^{2}+\sum_{n=1}^{\infty}n^{1-\alpha}|a_{n}|^{2}<\infty\right\}.

These are all Hilbert spaces and we see that H2=𝒟1\displaystyle H^{2}=\mathcal{D}_{1} with equal norm; for β>1\displaystyle\beta>-1, the standard weighted Bergman space Aβ2=𝒟β+2\displaystyle A_{\beta}^{2}=\mathcal{D}_{\beta+2} with an equivalent norm. Also, 𝒟=𝒟0\displaystyle\mathcal{D}=\mathcal{D}_{0} and S2=D1\displaystyle S^{2}=D_{-1} with equal norm. Moreover, if 1α<β<\displaystyle-1\leq\alpha<\beta<\infty, 𝒟α𝒟β\displaystyle\mathcal{D}_{\alpha}\subset\mathcal{D}_{\beta} with continuous inclusion and the analytic polynomials are dense in 𝒟α\displaystyle\mathcal{D}_{\alpha}.

As is the case with S2=D1\displaystyle S^{2}=D_{-1}, there are no nice closed forms for the reproducing kernels for 𝒟α\displaystyle\mathcal{D}_{\alpha} when 1<α<0\displaystyle-1<\alpha<0 which is a drawback since these kernels are quite useful in the study of composition operators. In particular, if \displaystyle\mathcal{H} is a functional Hilbert space of functions on the disk and Cφ\displaystyle C_{\varphi} is bounded on \displaystyle\mathcal{H}, then for w𝔻\displaystyle w\in\mathbb{D}, we have

CφKw=Kφ(w).\displaystyle C_{\varphi}^{*}K_{w}=K_{\varphi(w)}.

To overcome this drawback, in many instances we will consider the linear functional for evaluation of the first derivative at a point in the disk; for a reference see [7] Theorem 2.16. If =𝒟α\displaystyle\mathcal{H}=\mathcal{D}_{\alpha} for α1\displaystyle\alpha\geq-1, these functionals are bounded, and thus the Riesz Representation Theorem ([6] Theorem I.3.4) guarantees the existence of a function, denoted Kw(1)\displaystyle K_{w}^{(1)} for w𝔻\displaystyle w\in\mathbb{D}, such that

f(w)=f,Kw(1).\displaystyle f^{\prime}(w)=\langle f,K_{w}^{(1)}\rangle_{\mathcal{H}}.

As with the point evaluation kernels, these kernels behave predictably under the action of the adjoint of a composition operator and it is easy to see that

CφKw(1)=φ(w)¯Kφ(w)(1).\displaystyle C_{\varphi}^{*}K_{w}^{(1)}=\overline{\varphi^{\prime}(w)}K_{\varphi(w)}^{(1)}.

In particular, we will employ these in the spaces 𝒟\displaystyle\mathcal{D} and S2\displaystyle S^{2}. On 𝒟\displaystyle\mathcal{D}, we find that

Kw(1)(z)=z1w¯zandKw(1)𝒟2=1(1|w|2)2,\displaystyle K_{w}^{(1)}(z)=\frac{z}{1-\overline{w}z}\hskip 14.45377pt\textup{and}\hskip 14.45377pt\|K_{w}^{(1)}\|_{\mathcal{D}}^{2}=\frac{1}{(1-|w|^{2})^{2}},

whereas on S2\displaystyle S^{2} we have

Kw(1)(z)=1w¯log11w¯zandKw(1)S22=11|w|2.\displaystyle K_{w}^{(1)}(z)=\frac{1}{\overline{w}}\log\frac{1}{1-\overline{w}z}\hskip 14.45377pt\textup{and}\hskip 14.45377pt\|K_{w}^{(1)}\|_{S^{2}}^{2}=\frac{1}{1-|w|^{2}}.

2.2. Julia Carathéodory Theorem

For an analytic self-map φ\displaystyle\varphi of the unit disk, the angular derivative plays a key role in determining compactness of composition operators on many of the spaces in question. For ζ𝔻\displaystyle\zeta\in\partial\mathbb{D} and M>1\displaystyle M>1, a nontangential approach region at ζ\displaystyle\zeta is defined by

Γ(ζ,M)={z𝔻:|zζ|<M(1|z|)}\displaystyle\Gamma(\zeta,M)=\{z\in\mathbb{D}:|z-\zeta|<M(1-|z|)\}

and a function f\displaystyle f has a nontangential limit at ζ\displaystyle\zeta if limzζf(z)\displaystyle\lim_{z\rightarrow\zeta}f(z) exists in each nontangential region Γ(ζ,M)\displaystyle\Gamma(\zeta,M). When a nontangential limit exists, we denote it by limzζf(z).\displaystyle\angle\lim_{z\rightarrow\zeta}f(z). Furthermore, if φ\displaystyle\varphi is a self-map of the disk and ζ𝔻\displaystyle\zeta\in\partial\mathbb{D}, then φ\displaystyle\varphi has finite angular derivative at ζ\displaystyle\zeta if there exists η𝔻\displaystyle\eta\in\partial\mathbb{D} such that

φ(ζ):=limzζηφ(z)ζz\displaystyle\varphi^{\prime}(\zeta):=\angle\lim_{z\rightarrow\zeta}\frac{\eta-\varphi(z)}{\zeta-z}

exists as a (finite) complex value. One obvious implication of the existence of a finite angular derivative for φ\displaystyle\varphi at ζ\displaystyle\zeta is that φ\displaystyle\varphi has nontangential limit of modulus 1 at ζ.\displaystyle\zeta. The Julia-Carathéodory Theorem provides several other implications; for a reference see [7] Theorem 2.44.

Theorem 2.1 (Julia-Carathéodory Theorem).

For an analytic self-map φ\displaystyle\varphi of 𝔻\displaystyle\mathbb{D} and ζ𝔻\displaystyle\zeta\in\partial\mathbb{D}, the following are equivalent:

  1. (a)

    d(ζ):=lim infzζ(1|φ(z)|)/(1|z|)<\displaystyle d(\zeta):=\liminf_{z\rightarrow\zeta}(1-|\varphi(z)|)/(1-|z|)<\infty, where the limit is taken as zζ\displaystyle z\rightarrow\zeta unrestrictedly in 𝔻\displaystyle\mathbb{D};

  2. (b)

    φ\displaystyle\varphi has finite angular derivative φ(ζ)\displaystyle\varphi^{\prime}(\zeta) at ζ\displaystyle\zeta;

  3. (c)

    both φ\displaystyle\varphi and φ\displaystyle\varphi^{\prime} have finite nontangential limits at ζ\displaystyle\zeta, with |η|=1\displaystyle|\eta|=1 where η=limr1φ(rζ).\displaystyle\eta=\lim_{r\rightarrow 1}\varphi(r\zeta).

Moreover, when these conditions hold, we have limzζφ(z)=φ(ζ)=ζ¯ηd(ζ)\displaystyle\angle\lim_{z\rightarrow\zeta}\varphi^{\prime}(z)=\varphi^{\prime}(\zeta)=\overline{\zeta}\eta d(\zeta), i.e. d(ζ)=|φ(ζ)|>0\displaystyle d(\zeta)=|\varphi^{\prime}(\zeta)|>0, and d(ζ)=limzζ(1|φ(z)|)/(1|z|).\displaystyle d(\zeta)=\angle\lim_{z\rightarrow\zeta}(1-|\varphi(z)|)/(1-|z|).

In characterizing compact differences, we will be interested in maps which have similar behavior on the boundary of 𝔻\displaystyle\mathbb{D}. If φ\displaystyle\varphi and ψ\displaystyle\psi are two self-maps of the unit disk both with finite angular derivative at ζ𝔻\displaystyle\zeta\in\partial\mathbb{D}, then we say that the maps have the same first order data at ζ\displaystyle\zeta if φ(ζ)=ψ(ζ)\displaystyle\varphi(\zeta)=\psi(\zeta) (as radial limits) and φ(ζ)=ψ(ζ)\displaystyle\varphi^{\prime}(\zeta)=\psi^{\prime}(\zeta). If in addition φ\displaystyle\varphi and ψ\displaystyle\psi are twice differentiable at ζ\displaystyle\zeta (meaning that if we consider φ\displaystyle\varphi and ψ\displaystyle\psi as functions on 𝔻{ζ}\displaystyle\mathbb{D}\cup\{\zeta\}, then they are twice continuously differentiable) with φ′′(ζ)=ψ′′(ζ)\displaystyle\varphi^{\prime\prime}(\zeta)=\psi^{\prime\prime}(\zeta), then we say that φ\displaystyle\varphi and ψ\displaystyle\psi have the same second order data at ζ\displaystyle\zeta.

2.3. Linear fractional self-maps of the disk

Recall that a linear fractional map has the form φ(z)=(az+b)/(cz+d)\displaystyle\varphi(z)=(az+b)/(cz+d); the condition that adbc0\displaystyle ad-bc\neq 0 is necessary and sufficient for such a φ\displaystyle\varphi to be a univalent, nonconstant mapping of the Riemann sphere onto itself. Our focus here is on linear fractional self-maps of the disk and we point the reader to Chapter 0 of [20] for more information. With this narrowed focus we may assume that d0\displaystyle d\neq 0, in which case we can represent φ\displaystyle\varphi in the form φ(z)=(az+b)/(cz+1)\displaystyle\varphi(z)=(az+b)/(cz+1).

It is easy to see that every linear fractional self-map φ\displaystyle\varphi of 𝔻\displaystyle\mathbb{D} will induce a bounded composition operator on the weighted Dirichlet spaces under consideration. This is due to the fact that the map φ\displaystyle\varphi^{\prime} is continuous, and hence bounded, on 𝔻¯\displaystyle\overline{\mathbb{D}}. Of particular interest here is the role of second order data. The following statement seems to be known but we were unable to find a proof in the literature.

Lemma 2.2.

If φ\displaystyle\varphi and ψ\displaystyle\psi are linear fractional self-maps of 𝔻\displaystyle\mathbb{D} with the same second order data at a point in 𝔻\displaystyle\partial\mathbb{D}, then φ=ψ\displaystyle\varphi=\psi.

Proof.

Assume that φ(z)=(az+b)/(cz+1)\displaystyle\varphi(z)=(az+b)/(cz+1) and ψ(z)=(αz+β)/(γz+1)\displaystyle\psi(z)=(\alpha z+\beta)/(\gamma z+1) are nonconstant linear fractional self-maps of 𝔻\displaystyle\mathbb{D} with the same second order data. By composing with rotations, we may assume without loss of generality that φ(1)=ψ(1)=1\displaystyle\varphi(1)=\psi(1)=1. Thus we have the following assumptions:

  1. (i)

    φ(1)=a+bc+1=α+βγ+1=ψ(1)=1\displaystyle\displaystyle\varphi(1)=\frac{a+b}{c+1}=\frac{\alpha+\beta}{\gamma+1}=\psi(1)=1;

  2. (ii)

    φ(1)=abc(c+1)2=αβγ(γ+1)2=ψ(1)\displaystyle\displaystyle\varphi^{\prime}(1)=\frac{a-bc}{(c+1)^{2}}=\frac{\alpha-\beta\gamma}{(\gamma+1)^{2}}=\psi^{\prime}(1);

  3. (iii)

    φ′′(1)=2c(abc)(c+1)3=2γ(αβγ)(γ+1)3=ψ′′(1)\displaystyle\displaystyle\varphi^{\prime\prime}(1)=\frac{-2c(a-bc)}{(c+1)^{3}}=\frac{-2\gamma(\alpha-\beta\gamma)}{(\gamma+1)^{3}}=\psi^{\prime\prime}(1).

First note that c,γ1\displaystyle c,\gamma\neq-1 since each of the above quantities exist as finite complex values. Now, considering (ii) and (iii), it immediately follows that

cc+1=γγ+1\displaystyle\frac{c}{c+1}=\frac{\gamma}{\gamma+1}

and hence c=γ\displaystyle c=\gamma. Substituting this into (i), we have a+b=α+β\displaystyle a+b=\alpha+\beta or aα=βb\displaystyle a-\alpha=\beta-b. Moreover, this same substitution in (ii) implies that

aα=bcβc=c(bβ)\displaystyle a-\alpha=bc-\beta c=c(b-\beta)

or

βb=c(bβ).\displaystyle\beta-b=c(b-\beta).

Thus it must be the case that b=β\displaystyle b=\beta, since c1\displaystyle c\neq-1. It is then immediate that a=α\displaystyle a=\alpha and hence φ=ψ\displaystyle\varphi=\psi. ∎

2.4. Calderón’s complex interpolation

Let (X0,0)\displaystyle(X_{0},\|\cdot\|_{0}) and (X1,1)\displaystyle(X_{1},\|\cdot\|_{1}) be a compatible pair of Banach spaces in the sense of Calderón (see [5]). Both X0\displaystyle X_{0} and X1\displaystyle X_{1} may be continuously embedded in the complex topological vector space X0+X1\displaystyle X_{0}+X_{1} when equipped with the norm

xX0+X1=inf{y0+z1:x=y+z,yX0,zX1}.\displaystyle\|x\|_{X_{0}+X_{1}}=\inf\left\{\|y\|_{0}+\|z\|_{1}:x=y+z,y\in X_{0},z\in X_{1}\right\}.

In addition, the space X0X1\displaystyle X_{0}\cap X_{1}, with norm

xX0X1=max{x0,x1},\displaystyle\|x\|_{X_{0}\cap X_{1}}=\max\left\{\|x\|_{0},\,\|x\|_{1}\right\},

maps continuously into X0\displaystyle X_{0} and X1\displaystyle X_{1}. In this note we will further assume that the space X0X1\displaystyle X_{0}\cap X_{1} is dense in both X0\displaystyle X_{0} and X1\displaystyle X_{1} and define the interpolation algebra [X0,X1]\displaystyle\mathcal{I}[X_{0},X_{1}] to be the set of all linear operators T:X0X1X0X1\displaystyle T:X_{0}\cap X_{1}\rightarrow X_{0}\cap X_{1} that are both 0-continuous and 1-continuous. The interpolation algebra defined above first appeared in the Lp\displaystyle L^{p}-space setting in [2]; for properties and applications to the study of spectra, see [1], [10], [16], and [17].

For a Banach space 𝒴\displaystyle\mathcal{Y}, we let (𝒴)\displaystyle\mathcal{B}(\mathcal{Y}) denote the set of all bounded operators on 𝒴\displaystyle\mathcal{Y}. By continuity any operator T[X0,X1]\displaystyle T\in\mathcal{I}[X_{0},X_{1}] induces a unique operator Ti(Xi)\displaystyle T_{i}\in\mathcal{B}(X_{i}), i=0,1\displaystyle i=0,1. For t(0,1)\displaystyle t\in(0,1), let Xt=[X0,X1]t\displaystyle X_{t}=[X_{0},X_{1}]_{t} be the interpolation space obtained via Calderón’s method of complex interpolation; it follows then that X0X1\displaystyle X_{0}\cap X_{1} is dense in Xt\displaystyle X_{t} and T\displaystyle T also induces a unique operator Tt(Xt)\displaystyle T_{t}\in\mathcal{B}(X_{t}) satisfying

Tt(Xt)T0(X0)1tT1(X1)t,t(0,1).\displaystyle\|T_{t}\|_{\mathcal{B}(X_{t})}\leq\|T_{0}\|_{\mathcal{B}(X_{0})}^{1-t}\|T_{1}\|_{\mathcal{B}(X_{1})}^{t},\hskip 7.22743ptt\in(0,1).

To apply interpolation techniques to our study, we first verify that the weighted Dirichlet spaces can be interpreted as interpolation spaces. One can see this by considering these spaces as weighted 2\displaystyle\ell^{2}-spaces or by considering the techniques developed in [13]; a direct proof of this nature can be found in [16].

Proposition 2.3.

Suppose 1<α<γ<β<\displaystyle-1<\alpha<\gamma<\beta<\infty. If t(0,1)\displaystyle t\in(0,1) with γ=(1t)α+tβ\displaystyle\gamma=(1-t)\alpha+t\beta, then [𝒟α,𝒟β]t=𝒟γ\displaystyle[\mathcal{D}_{\alpha},\mathcal{D}_{\beta}]_{t}=\mathcal{D}_{\gamma} with the series norm given above.

One appealing property of working with composition operators on these spaces in the interpolation setting is the nested behavior mentioned earlier. Specifically, 𝒟α𝒟β\displaystyle\mathcal{D}_{\alpha}\subset\mathcal{D}_{\beta}, for 1<α<β\displaystyle-1<\alpha<\beta, with continuous inclusion, so 𝒟α=𝒟α𝒟β\displaystyle\mathcal{D}_{\alpha}=\mathcal{D}_{\alpha}\cap\mathcal{D}_{\beta}. Combining this with the fact that the analytic polynomials are dense in each weighted Dirichlet space implies that 𝒟α𝒟β\displaystyle\mathcal{D}_{\alpha}\cap\mathcal{D}_{\beta} is dense in 𝒟α\displaystyle\mathcal{D}_{\alpha} and 𝒟β\displaystyle\mathcal{D}_{\beta}. Moreover, MacCluer and Shapiro showed in [12] that boundedness of Cφ\displaystyle C_{\varphi} on 𝒟α\displaystyle\mathcal{D}_{\alpha} implies boundedness on 𝒟β\displaystyle\mathcal{D}_{\beta} when 1<α<β.\displaystyle-1<\alpha<\beta. Thus, for our purposes, it suffices to know that Cφ\displaystyle C_{\varphi} is bounded on the single endpoint space 𝒟α\displaystyle\mathcal{D}_{\alpha}. Furthermore, the fact that we are defining our operators on a dense subset of each space implies that for t(0,1)\displaystyle t\in(0,1) the interpolated operator satisfies (Cφ)t=Cφ\displaystyle(C_{\varphi})_{t}=C_{\varphi}.

When using interpolated operators, the goal is to determine properties of the operator on an interpolation space Xt\displaystyle X_{t} based on properties of the operator on the endpoint spaces, or to extrapolate properties from one interpolation space to the other interpolation spaces and/or the endpoint spaces. The result that we will make use of here is Cwikel’s compactness result which extrapolates compactness on one interpolation space to the other interpolation spaces, but not necessarily to the endpoint spaces.

Theorem 2.4 ([8] Theorem 2.1).

If T0\displaystyle T_{0} and T1\displaystyle T_{1} are bounded and Tt\displaystyle T_{t} is compact for some t(0,1)\displaystyle t\in(0,1), then Tx\displaystyle T_{x} is compact for all x(0,1)\displaystyle x\in(0,1).

3. Compact Differences of Composition Operators

As mentioned previously, our goal is to extend the work of Moorhouse [14]. There the author characterized when two composition operators have compact difference on the weighted Bergman spaces (α>1\displaystyle\alpha>1 in our scale of weighted Dirichlet spaces) and provided partial results for weighted Dirichlet spaces in the range 0<α1\displaystyle 0<\alpha\leq 1. Here we extend some of those results to the entire range of weighted Dirichlet spaces. We first work on the Dirichlet space by modifying techniques used in the Hardy and weighted Bergman space setting. We then discuss extending this to the space S2\displaystyle S^{2}. Finally, we discuss the compact difference problem on an arbitrary weighted Dirichlet space where we apply Calderón’s complex interpolation.

Lemma 3.1.

Let φ\displaystyle\varphi and ψ\displaystyle\psi be analytic self-maps of 𝔻\displaystyle\mathbb{D} such that Cφ\displaystyle C_{\varphi} and Cψ\displaystyle C_{\psi} are bounded on 𝒟\displaystyle\mathcal{D}. Further assume that φ\displaystyle\varphi and ψ\displaystyle\psi have finite angular derivative at some point ζ𝔻\displaystyle\zeta\in\partial\mathbb{D}. If CφCψ\displaystyle C_{\varphi}-C_{\psi} is compact on 𝒟\displaystyle\mathcal{D}, then φ\displaystyle\varphi and ψ\displaystyle\psi have the same first order data at ζ\displaystyle\zeta.

Proof.

First note that it suffices to prove the statement in the case when ζ=1\displaystyle\zeta=1 since the rotations of the disk give rise to unitary operators on 𝒟\displaystyle\mathcal{D}. By the Julia-Carathéodory Theorem it follows that |φ(1)|=|ψ(1)|=1\displaystyle|\varphi(1)|=|\psi(1)|=1 and by hypothesis φ(1)=s<\displaystyle\varphi^{\prime}(1)=s<\infty and ψ(1)=t<\displaystyle\psi^{\prime}(1)=t<\infty. For the sake of notation we set φ(1)=η\displaystyle\varphi(1)=\eta. Now, we will show that if φ\displaystyle\varphi and ψ\displaystyle\psi do not have the same first order data then CφCψ\displaystyle C_{\varphi}-C_{\psi} is not compact by showing that the essential norm of CφCψ\displaystyle C_{\varphi}-C_{\psi} is bounded away from 0; in particular we will show that

CφCψe21.\displaystyle\|C_{\varphi}-C_{\psi}\|_{e}^{2}\geq 1.

To this end, we first obtain lower estimates on the norm of (CφCψ)\displaystyle(C_{\varphi}-C_{\psi})^{*} acting on the normalized reproducing kernels for evaluation of the first derivative; for w𝔻\displaystyle w\in\mathbb{D} we have

(CφCψ)(Kw(1)Kw(1))2=CφKw(1)CψKw(1)2Kw(1)2\displaystyle\left\|(C_{\varphi}-C_{\psi})^{*}\left(\frac{K_{w}^{(1)}}{\|K_{w}^{(1)}\|}\right)\right\|^{2}=\frac{\|C_{\varphi}^{*}K_{w}^{(1)}-C_{\psi}^{*}K_{w}^{(1)}\|^{2}}{\|K_{w}^{(1)}\|^{2}}

and

CφKw(1)CψKw(1)2Kw(1)2=CφKw(1)2Kw(1)2+CψKw(1)2Kw(1)22ReCφKw(1)Kw(1),CψKw(1)Kw(1).\displaystyle\frac{\|C_{\varphi}^{*}K_{w}^{(1)}-C_{\psi}^{*}K_{w}^{(1)}\|^{2}}{\|K_{w}^{(1)}\|^{2}}=\frac{\|C_{\varphi}^{*}K_{w}^{(1)}\|^{2}}{\|K_{w}^{(1)}\|^{2}}+\frac{\|C_{\psi}^{*}K_{w}^{(1)}\|^{2}}{\|K_{w}^{(1)}\|^{2}}-2\textup{Re}\left\langle\frac{C_{\varphi}^{*}K_{w}^{(1)}}{\|K_{w}^{(1)}\|},\frac{C_{\psi}^{*}K_{w}^{(1)}}{\|K_{w}^{(1)}\|}\right\rangle.

Considering the action of the adjoint and the formulas for these kernels and their norms in 𝒟\displaystyle\mathcal{D}, we see that previous line is equal to

|φ(w)|2(1|w|2)2(1|φ(w)|2)2+|ψ(w)|2(1|w|2)2(1|ψ(w)|2)22Re(φ(w)¯ψ(w)(1|w|2)2(1φ(w)¯ψ(w))2).\displaystyle\frac{|\varphi^{\prime}(w)|^{2}(1-|w|^{2})^{2}}{(1-|\varphi(w)|^{2})^{2}}+\frac{|\psi^{\prime}(w)|^{2}(1-|w|^{2})^{2}}{(1-|\psi(w)|^{2})^{2}}-2\textup{Re}\left(\frac{\overline{\varphi^{\prime}(w)}\psi^{\prime}(w)(1-|w|^{2})^{2}}{(1-\overline{\varphi(w)}\psi(w))^{2}}\right).

By the Julia-Carathéodory Theorem

limw11|w|21|φ(w)|2=1sandlimw1φ(w)=φ(1)=s,\displaystyle\angle\lim_{w\rightarrow 1}\frac{1-|w|^{2}}{1-|\varphi(w)|^{2}}=\frac{1}{s}\hskip 14.45377pt\textup{and}\hskip 14.45377pt\angle\lim_{w\rightarrow 1}\varphi^{\prime}(w)=\varphi^{\prime}(1)=s,

from which it follows that

limw1|φ(w)|2(1|w|2)2(1|φ(w)|2)2=1.\displaystyle\angle\lim_{w\rightarrow 1}\frac{|\varphi^{\prime}(w)|^{2}(1-|w|^{2})^{2}}{(1-|\varphi(w)|^{2})^{2}}=1.

Next we consider two cases. If ψ(1)η\displaystyle\psi(1)\neq\eta (as a radial limit), then we can find a sequence {rn}\displaystyle\{r_{n}\} increasing to 1 such that limnψ(rn)η\displaystyle\lim_{n\rightarrow\infty}\psi(r_{n})\neq\eta. Then limn(1φ(rn)¯ψ(rn))0\displaystyle\lim_{n\rightarrow\infty}(1-\overline{\varphi(r_{n})}\psi(r_{n}))\neq 0 and the fact that s,t<\displaystyle s,t<\infty guarantee us that

limn2Re(φ(rn)¯ψ(rn)(1|rn|2)2(1φ(rn)¯ψ(rn))2)=0.\displaystyle\lim_{n\rightarrow\infty}2\textup{Re}\left(\frac{\overline{\varphi^{\prime}(r_{n})}\psi^{\prime}(r_{n})(1-|r_{n}|^{2})^{2}}{(1-\overline{\varphi(r_{n})}\psi(r_{n}))^{2}}\right)=0.

This shows that in this case

(CφCψ)(Kw(1)Kw(1))1.\left\|(C_{\varphi}-C_{\psi})^{*}\left(\frac{K_{w}^{(1)}}{\|K_{w}^{(1)}\|}\right)\right\|\geq 1. (3.1)

For the second case suppose that ψ(1)=η\displaystyle\psi(1)=\eta (as a radial limit) but st\displaystyle s\neq t. Observe that

(1φ(w)¯ψ(w))2φ(w)¯ψ(w)(1|w|2)2=1φ(w)¯ψ(w)(1φ(w)¯ψ(w)1|w|2)2\displaystyle\frac{(1-\overline{\varphi(w)}\psi(w))^{2}}{\overline{\varphi^{\prime}(w)}\psi^{\prime}(w)(1-|w|^{2})^{2}}=\frac{1}{\overline{\varphi^{\prime}(w)}\psi^{\prime}(w)}\left(\frac{1-\overline{\varphi(w)}\psi(w)}{1-|w|^{2}}\right)^{2}

which equals

1φ(w)¯ψ(w)(1|φ(w)|21|w|2+φ(w)¯(1w)1|w|2(ηψ(w)1wηφ(w)1w))2.\displaystyle\frac{1}{\overline{\varphi^{\prime}(w)}\psi^{\prime}(w)}\left(\frac{1-|\varphi(w)|^{2}}{1-|w|^{2}}+\frac{\overline{\varphi(w)}(1-w)}{1-|w|^{2}}\left(\frac{\eta-\psi(w)}{1-w}-\frac{\eta-\varphi(w)}{1-w}\right)\right)^{2}.

For M>1\displaystyle M>1, consider the boundary of a nontangential approach region

γM={w𝔻:|1w|1|w|2=M}.\displaystyle\gamma_{M}=\left\{w\in\mathbb{D}:\frac{|1-w|}{1-|w|^{2}}=M\right\}.

As w1\displaystyle w\rightarrow 1 along γM\displaystyle\gamma_{M}, the Julia-Carathéodory Theorem guarantees us that

|φ(w)¯(1w)1|w|2(ηψ(w)1wηφ(w)1w)|M|ts|.\displaystyle\left|\frac{\overline{\varphi(w)}(1-w)}{1-|w|^{2}}\left(\frac{\eta-\psi(w)}{1-w}-\frac{\eta-\varphi(w)}{1-w}\right)\right|\rightarrow M|t-s|.

Thus for N>0\displaystyle N>0, by choosing M\displaystyle M sufficiently large, we may find a sequence {wn}\displaystyle\{w_{n}\} approaching 1 along γM\displaystyle\gamma_{M} such that for n\displaystyle n large enough it follows that

|(1φ(wn)¯ψ(wn))2φ(wn)¯ψ(wn)(1|wn|2)2|>N.\displaystyle\left|\frac{(1-\overline{\varphi(w_{n})}\psi(w_{n}))^{2}}{\overline{\varphi^{\prime}(w_{n})}\psi^{\prime}(w_{n})(1-|w_{n}|^{2})^{2}}\right|>N.

Equivalently, for 0<ε<1\displaystyle 0<\varepsilon<1, we may find a sequence {wn}\displaystyle\{w_{n}\} converging to 1 nontangentially such that

2Re(φ(wn)¯ψ(wn)(1|wn|2)2(1φ(wn)¯ψ(wn))2)|2φ(wn)¯ψ(wn)(1|wn|2)2(1φ(wn)¯ψ(wn))2|<ε\displaystyle 2\textup{Re}\left(\frac{\overline{\varphi^{\prime}(w_{n})}\psi^{\prime}(w_{n})(1-|w_{n}|^{2})^{2}}{(1-\overline{\varphi(w_{n})}\psi(w_{n}))^{2}}\right)\leq\left|\frac{2\overline{\varphi^{\prime}(w_{n})}\psi^{\prime}(w_{n})(1-|w_{n}|^{2})^{2}}{(1-\overline{\varphi(w_{n})}\psi(w_{n}))^{2}}\right|<\varepsilon

for n\displaystyle n sufficiently large. Thus we see that

(CφCψ)(Kw(1)Kw(1))1ε\displaystyle\left\|(C_{\varphi}-C_{\psi})^{*}\left(\frac{K_{w}^{(1)}}{\|K_{w}^{(1)}\|}\right)\right\|\geq 1-\varepsilon

for all ε\displaystyle\varepsilon with 0<ε<1\displaystyle 0<\varepsilon<1, and hence in this case the estimate from  (3.1) also holds.

To show that CφCψe21\displaystyle\|C_{\varphi}-C_{\psi}\|_{e}^{2}\geq 1, recall that the normalized kernel functions Kw(1)/Kw(1)\displaystyle K_{w}^{(1)}/\|K_{w}^{(1)}\| converge weakly to zero as |w|1\displaystyle|w|\rightarrow 1 (see [7] Proposition 7.13). If we then consider any compact operator Q\displaystyle Q, it must then be the case that

Q(Kw(1)Kw(1))0\displaystyle\left\|Q^{*}\left(\frac{K_{w}^{(1)}}{\|K_{w}^{(1)}\|}\right)\right\|\rightarrow 0

as |w|1\displaystyle|w|\rightarrow 1. Combining this with the estimate

CφCψQ(CφCψ)(Kw(1)Kw(1))Q(Kw(1)Kw(1))\displaystyle\|C_{\varphi}-C_{\psi}-Q\|\geq\left\|(C_{\varphi}-C_{\psi})^{*}\left(\frac{K_{w}^{(1)}}{\|K_{w}^{(1)}\|}\right)\right\|-\left\|Q^{*}\left(\frac{K_{w}^{(1)}}{\|K_{w}^{(1)}\|}\right)\right\|

it follows that

CφCψe2=inf{CφCψQ2:Q compact}1\displaystyle\|C_{\varphi}-C_{\psi}\|_{e}^{2}=\inf\{\|C_{\varphi}-C_{\psi}-Q\|^{2}:Q\textup{ compact}\}\geq 1

completing the proof.∎

Our next result is a Dirichlet space analog of [14] Theorem 4. The statement there concerns weighted Dirichlet spaces 𝒟α\displaystyle\mathcal{D}_{\alpha} with α>0\displaystyle\alpha>0. Our statement takes a slightly different form in that the derivatives of φ\displaystyle\varphi and ψ\displaystyle\psi appear; this is due to the fact that we are using the kernels for evaluation of the first derivative.

Theorem 3.2.

Let φ\displaystyle\varphi and ψ\displaystyle\psi be analytic self-maps of 𝔻\displaystyle\mathbb{D} such that Cφ\displaystyle C_{\varphi} and Cψ\displaystyle C_{\psi} are bounded on 𝒟\displaystyle\mathcal{D} and define

ρ(z)=|φ(z)ψ(z)1φ(z)¯ψ(z)|\displaystyle\rho(z)=\left|\frac{\varphi(z)-\psi(z)}{1-\overline{\varphi(z)}\psi(z)}\right|

for z𝔻\displaystyle z\in\mathbb{D}. If CφCψ\displaystyle C_{\varphi}-C_{\psi} is compact on 𝒟\displaystyle\mathcal{D}, then

lim|z|1ρ(z)(|φ(z)|(1|z|2)1|φ(z)|2+|ψ(z)(1|z|2)1|ψ(z)|2)=0.\displaystyle\lim_{|z|\rightarrow 1}\rho(z)\left(\frac{|\varphi^{\prime}(z)|(1-|z|^{2})}{1-|\varphi(z)|^{2}}+\frac{|\psi^{\prime}(z)(1-|z|^{2})}{1-|\psi(z)|^{2}}\right)=0.

Before giving the proof, notice that the quantity ρ(z)\displaystyle\rho(z) is simply the pseudo-hyperbolic distance between φ(z)\displaystyle\varphi(z) and ψ(z)\displaystyle\psi(z) and thus we have the well known equality

1ρ2(z)=(1|φ(z)|2)(1|ψ(z)|2)|1φ(z)¯ψ(z)|2.\displaystyle 1-\rho^{2}(z)=\frac{(1-|\varphi(z)|^{2})(1-|\psi(z)|^{2})}{|1-\overline{\varphi(z)}\psi(z)|^{2}}.
Proof.

We again argue by contrapositive and assume that

lim|z|1ρ(z)(|φ(z)|(1|z|2)1|φ(z)|2+|ψ(z)(1|z|2)1|ψ(z)|2)0.\lim_{|z|\rightarrow 1}\rho(z)\left(\frac{|\varphi^{\prime}(z)|(1-|z|^{2})}{1-|\varphi(z)|^{2}}+\frac{|\psi^{\prime}(z)(1-|z|^{2})}{1-|\psi(z)|^{2}}\right)\neq 0. (3.2)

To show that CφCψ\displaystyle C_{\varphi}-C_{\psi} is not compact, we will show that there is a sequence {zn}\displaystyle\{z_{n}\} in 𝔻\displaystyle\mathbb{D} with |zn|1\displaystyle|z_{n}|\rightarrow 1 such that

(CφCψ)(Kzn(1)Kzn(1))↛0.\displaystyle\left\|(C_{\varphi}-C_{\psi})^{*}\left(\frac{K_{z_{n}}^{(1)}}{\|K_{z_{n}}^{(1)}\|}\right)\right\|\not\rightarrow 0.

Since Kw(1)/Kw(1)0\displaystyle K_{w}^{(1)}/\|K_{w}^{(1)}\|\rightarrow 0 weakly as |w|1\displaystyle|w|\rightarrow 1, we will conclude that (CφCψ)\displaystyle(C_{\varphi}-C_{\psi})^{*}, and hence CφCψ\displaystyle C_{\varphi}-C_{\psi}, is not compact.

As in the proof of Lemma 3.1, for w𝔻\displaystyle w\in\mathbb{D} we have

(CφCψ)(Kw(1)Kw(1))2=CφKw(1)CψKw(1)2Kw(1)2\displaystyle\left\|(C_{\varphi}-C_{\psi})^{*}\left(\frac{K_{w}^{(1)}}{\|K_{w}^{(1)}\|}\right)\right\|^{2}=\frac{\|C_{\varphi}^{*}K_{w}^{(1)}-C_{\psi}^{*}K_{w}^{(1)}\|^{2}}{\|K_{w}^{(1)}\|^{2}}

which is greater than or equal to

CφKw(1)2Kw(1)2+CψKw(1)2Kw(1)22|CφKw(1)Kw(1),CψKw(1)Kw(1)|.\frac{\|C_{\varphi}^{*}K_{w}^{(1)}\|^{2}}{\|K_{w}^{(1)}\|^{2}}+\frac{\|C_{\psi}^{*}K_{w}^{(1)}\|^{2}}{\|K_{w}^{(1)}\|^{2}}-2\left|\left\langle\frac{C_{\varphi}^{*}K_{w}^{(1)}}{\|K_{w}^{(1)}\|},\frac{C_{\psi}^{*}K_{w}^{(1)}}{\|K_{w}^{(1)}\|}\right\rangle\right|. (3.3)

Manipulating the third term here,

|CφKw(1)Kw(1),CψKw(1)Kw(1)|=|φ(w)||ψ(w)|(1|w|2)2|1φ(w)¯ψ(w)|2\displaystyle\left|\left\langle\frac{C_{\varphi}^{*}K_{w}^{(1)}}{\|K_{w}^{(1)}\|},\frac{C_{\psi}^{*}K_{w}^{(1)}}{\|K_{w}^{(1)}\|}\right\rangle\right|=\frac{|\varphi^{\prime}(w)||\psi^{\prime}(w)|(1-|w|^{2})^{2}}{|1-\overline{\varphi(w)}\psi(w)|^{2}}

which is equal to

(1|φ(w)|2)(1|ψ(w)|2)|1φ(w)¯ψ(w)|2|φ(w)|(1|w|2)1|φ(w)|2|ψ(w)|(1|w|2)1|ψ(w)|2\displaystyle\frac{(1-|\varphi(w)|^{2})(1-|\psi(w)|^{2})}{|1-\overline{\varphi(w)}\psi(w)|^{2}}\frac{|\varphi^{\prime}(w)|(1-|w|^{2})}{1-|\varphi(w)|^{2}}\frac{|\psi^{\prime}(w)|(1-|w|^{2})}{1-|\psi(w)|^{2}}

or, more simply,

(1ρ2(w))CφKw(1)Kw(1)CψKw(1)Kw(1).\displaystyle(1-\rho^{2}(w))\frac{\|C_{\varphi}^{*}K_{w}^{(1)}\|}{\|K_{w}^{(1)}\|}\frac{\|C_{\psi}^{*}K_{w}^{(1)}\|}{\|K_{w}^{(1)}\|}.

Substituting into the expression in  (3.3) and factoring we see that

CφKw(1)CψKw(1)2Kw(1)2(CφKw(1)CψKw(1)Kw(1))2+2ρ2(w)CφKw(1)Kw(1)CψKw(1)Kw(1).\displaystyle\frac{\|C_{\varphi}^{*}K_{w}^{(1)}-C_{\psi}^{*}K_{w}^{(1)}\|^{2}}{\|K_{w}^{(1)}\|^{2}}\geq\left(\frac{\|C_{\varphi}^{*}K_{w}^{(1)}\|-\|C_{\psi}^{*}K_{w}^{(1)}\|}{\|K_{w}^{(1)}\|}\right)^{2}+2\rho^{2}(w)\frac{\|C_{\varphi}^{*}K_{w}^{(1)}\|}{\|K_{w}^{(1)}\|}\frac{\|C_{\psi}^{*}K_{w}^{(1)}\|}{\|K_{w}^{(1)}\|}.

As we are assuming that the limit in Eqn.  (3.2) is not 0, it must be the case that there is a sequence {zn}\displaystyle\{z_{n}\} in 𝔻\displaystyle\mathbb{D} with |zn|1\displaystyle|z_{n}|\rightarrow 1 such that either

ρ(zn)(|φ(zn)|(1|zn|2)1|φ(zn)|2)=ρ(zn)CφKzn(1)Kzn(1):=an\displaystyle\rho(z_{n})\left(\frac{|\varphi^{\prime}(z_{n})|(1-|z_{n}|^{2})}{1-|\varphi(z_{n})|^{2}}\right)=\rho(z_{n})\frac{\|C_{\varphi}^{*}K_{z_{n}}^{(1)}\|}{\|K_{z_{n}}^{(1)}\|}:=a_{n}

or

ρ(zn)(|ψ(zn)|(1|zn|2)1|ψ(zn)|2)=ρ(zn)CψKzn(1)Kzn(1):=bn\displaystyle\rho(z_{n})\left(\frac{|\psi^{\prime}(z_{n})|(1-|z_{n}|^{2})}{1-|\psi(z_{n})|^{2}}\right)=\rho(z_{n})\frac{\|C_{\psi}^{*}K_{z_{n}}^{(1)}\|}{\|K_{z_{n}}^{(1)}\|}:=b_{n}

does not converge to 0. Since both the sequences {an}\displaystyle\{a_{n}\} and {bn}\displaystyle\{b_{n}\} are bounded (by the boundedness of Cφ\displaystyle C_{\varphi} and Cψ\displaystyle C_{\psi} together with the fact that ρ(z)1\displaystyle\rho(z)\leq 1 for all z𝔻\displaystyle z\in\mathbb{D}), we may, by passing to a subsequence if necessary, assume that anα\displaystyle a_{n}\rightarrow\alpha and bnβ\displaystyle b_{n}\rightarrow\beta with either α0\displaystyle\alpha\neq 0 or β0\displaystyle\beta\neq 0. By symmetry we may assume that α0\displaystyle\alpha\neq 0. By passing to a further subsequence if necessary, we may also assume that ρ(zn)p\displaystyle\rho(z_{n})\rightarrow p.

First we note that p>0\displaystyle p>0. Indeed, if p=0\displaystyle p=0, then it must be the case that

limn|φ(zn)|(1|zn|2)1|φ(zn)|2=limnCφKzn(1)Kzn(1)=,\displaystyle\lim_{n\rightarrow\infty}\frac{|\varphi^{\prime}(z_{n})|(1-|z_{n}|^{2})}{1-|\varphi(z_{n})|^{2}}=\lim_{n\rightarrow\infty}\frac{\|C_{\varphi}^{*}K_{z_{n}}^{(1)}\|}{\|K_{z_{n}}^{(1)}\|}=\infty,

but this contradicts the fact that Cφ\displaystyle C_{\varphi} is bounded. Thus we may (by passing to another subsequence if necessary) assume that

limn|φ(zn)|(1|zn|2)1|φ(zn)|2=limnCφKzn(1)Kzn(1)=αp.\displaystyle\lim_{n\rightarrow\infty}\frac{|\varphi^{\prime}(z_{n})|(1-|z_{n}|^{2})}{1-|\varphi(z_{n})|^{2}}=\lim_{n\rightarrow\infty}\frac{\|C_{\varphi}^{*}K_{z_{n}}^{(1)}\|}{\|K_{z_{n}}^{(1)}\|}=\frac{\alpha}{p}.

Similarly, the boundedness of Cψ\displaystyle C_{\psi} implies that we may assume

limn|ψ(zn)|(1|zn|2)1|ψ(zn)|2=limnCψKzn(1)Kzn(1)=βp.\displaystyle\lim_{n\rightarrow\infty}\frac{|\psi^{\prime}(z_{n})|(1-|z_{n}|^{2})}{1-|\psi(z_{n})|^{2}}=\lim_{n\rightarrow\infty}\frac{\|C_{\psi}^{*}K_{z_{n}}^{(1)}\|}{\|K_{z_{n}}^{(1)}\|}=\frac{\beta}{p}.

Now, if αβ\displaystyle\alpha\neq\beta, then

limn(CφKzn(1)CψKzn(1)Kzn(1))2=(αpβp)20.\displaystyle\lim_{n\rightarrow\infty}\left(\frac{\|C_{\varphi}^{*}K_{z_{n}}^{(1)}\|-\|C_{\psi}^{*}K_{z_{n}}^{(1)}\|}{\|K_{z_{n}}^{(1)}\|}\right)^{2}=\left(\frac{\alpha}{p}-\frac{\beta}{p}\right)^{2}\neq 0.

On the other hand, if α=β0\displaystyle\alpha=\beta\neq 0, then

limn2ρ2(zn)CφKzn(1)Kzn(1)CψKzn(1)Kzn(1)=2αβ0.\displaystyle\lim_{n\rightarrow\infty}2\rho^{2}(z_{n})\frac{\|C_{\varphi}^{*}K_{z_{n}}^{(1)}\|}{\|K_{z_{n}}^{(1)}\|}\frac{\|C_{\psi}^{*}K_{z_{n}}^{(1)}\|}{\|K_{z_{n}}^{(1)}\|}=2\alpha\beta\neq 0.

In either case,

(CφCψ)(Kzn(1)Kzn(1))↛0\displaystyle\left\|(C_{\varphi}-C_{\psi})^{*}\left(\frac{K_{z_{n}}^{(1)}}{\|K_{z_{n}}^{(1)}\|}\right)\right\|\not\rightarrow 0

as desired.∎

In the case of S2\displaystyle S^{2}, the proofs are nearly identical to those just given for 𝒟\displaystyle\mathcal{D} except for the fact that the kernel functions take a slightly different form in S2\displaystyle S^{2}; for a reference we point the reader to [9]. Notice also that the hypotheses of Lemma 3.3 are slightly altered. This is due to the fact that on S2\displaystyle S^{2}, the boundedness of Cφ\displaystyle C_{\varphi} implies that φ\displaystyle\varphi has finite angular derivative at any point ζ𝔻\displaystyle\zeta\in\partial\mathbb{D} with |φ(ζ)|=1\displaystyle|\varphi(\zeta)|=1 ([7] Theorem 4.13).

Lemma 3.3.

Let φ\displaystyle\varphi and ψ\displaystyle\psi be analytic self-maps of 𝔻\displaystyle\mathbb{D} such that Cφ\displaystyle C_{\varphi} and Cψ\displaystyle C_{\psi} are bounded on S2\displaystyle S^{2}. Further assume that there exists ζ𝔻\displaystyle\zeta\in\partial\mathbb{D} such that |φ(ζ)|=|ψ(ζ)|=1\displaystyle|\varphi(\zeta)|=|\psi(\zeta)|=1. If CφCψ\displaystyle C_{\varphi}-C_{\psi} is compact on S2\displaystyle S^{2}, then φ\displaystyle\varphi and ψ\displaystyle\psi have the same first order data at ζ\displaystyle\zeta.

The following theorem should be compared to the result of Theorem 3.2 for the Dirichlet space and [14] Theorem 4. Here we see the square of the modulus of the derivative appearing. Again, this difference in form is due to the use of the kernel for evaluation of the first derivative in S2\displaystyle S^{2}.

Theorem 3.4.

Let φ\displaystyle\varphi and ψ\displaystyle\psi be analytic self-maps of 𝔻\displaystyle\mathbb{D} such that Cφ\displaystyle C_{\varphi} and Cψ\displaystyle C_{\psi} are bounded on S2\displaystyle S^{2}. If CφCψ\displaystyle C_{\varphi}-C_{\psi} is compact on S2\displaystyle S^{2}, then

lim|z|1ρ(z)(|φ(z)|2(1|z|2)1|φ(z)|2+|ψ(z)|2(1|z|2)1|ψ(z)|2)=0.\displaystyle\lim_{|z|\rightarrow 1}\rho(z)\left(\frac{|\varphi^{\prime}(z)|^{2}(1-|z|^{2})}{1-|\varphi(z)|^{2}}+\frac{|\psi^{\prime}(z)|^{2}(1-|z|^{2})}{1-|\psi(z)|^{2}}\right)=0.

We close with our main theorem and an interesting corollary for one particular class of maps.

Theorem 3.5.

Let φ\displaystyle\varphi and ψ\displaystyle\psi be analytic self-maps of 𝔻\displaystyle\mathbb{D}. Let F(φ)\displaystyle F(\varphi) be the set of points ζ𝔻\displaystyle\zeta\in\partial\mathbb{D} with |φ(ζ)|=1\displaystyle|\varphi(\zeta)|=1 and similarly for F(ψ)\displaystyle F(\psi). Further suppose that F(φ)=F(ψ):=F\displaystyle F(\varphi)=F(\psi):=F.

  1. (a)

    Suppose Cφ\displaystyle C_{\varphi} and Cψ\displaystyle C_{\psi} are bounded on S2=𝒟1\displaystyle S^{2}=\mathcal{D}_{-1}. If φ\displaystyle\varphi and ψ\displaystyle\psi have second order data at each point ζF\displaystyle\zeta\in F and CφCψ\displaystyle C_{\varphi}-C_{\psi} is compact on S2\displaystyle S^{2}, then φ\displaystyle\varphi and ψ\displaystyle\psi have the same second order data at each point ζF\displaystyle\zeta\in F.

  2. (b)

    Let γ>1\displaystyle\gamma>-1 and suppose Cφ\displaystyle C_{\varphi} and Cψ\displaystyle C_{\psi} are bounded on 𝒟α\displaystyle\mathcal{D}_{\alpha} for some α\displaystyle\alpha with 1α<γ\displaystyle-1\leq\alpha<\gamma. If φ\displaystyle\varphi and ψ\displaystyle\psi have finite angular derivative and second order data at each point ζF\displaystyle\zeta\in F and CφCψ\displaystyle C_{\varphi}-C_{\psi} is compact on 𝒟γ\displaystyle\mathcal{D}_{\gamma}, then φ\displaystyle\varphi and ψ\displaystyle\psi have the same second order data at each point ζF\displaystyle\zeta\in F.

Proof.

For (a), we know that φ\displaystyle\varphi and ψ\displaystyle\psi have the same first order data at each ζF\displaystyle\zeta\in F by Lemma 3.3. If φ′′(ζ)ψ′′(ζ)\displaystyle\varphi^{\prime\prime}(\zeta)\neq\psi^{\prime\prime}(\zeta), then by [14] Proposition 1 there is a sequence {zn}\displaystyle\{z_{n}\} in 𝔻\displaystyle\mathbb{D} with znζ\displaystyle z_{n}\rightarrow\zeta, i.e. |zn|1\displaystyle|z_{n}|\rightarrow 1,such that ρ(zn)↛0\displaystyle\rho(z_{n})\not\rightarrow 0 and

1|zn|21|φ(zn)|2↛0.\displaystyle\frac{1-|z_{n}|^{2}}{1-|\varphi(z_{n})|^{2}}\not\rightarrow 0.

Since φ(ζ)0\displaystyle\varphi^{\prime}(\zeta)\neq 0, this contradicts Theorem 3.4 as we are assuming that CφCψ\displaystyle C_{\varphi}-C_{\psi} is compact on S2\displaystyle S^{2}, and hence φ′′(ζ)=ψ′′(ζ).\displaystyle\varphi^{\prime\prime}(\zeta)=\psi^{\prime\prime}(\zeta).

For (b), it suffices to prove the result for γ\displaystyle\gamma with 1<γ1\displaystyle-1<\gamma\leq 1; the other cases follow from [14] Theorem 6 (we will actually appeal to this result to obtain our conclusion momentarily). Choose β\displaystyle\beta and δ\displaystyle\delta such that 1α<γ1<δ<β\displaystyle-1\leq\alpha<\gamma\leq 1<\delta<\beta. Then it must be the case that 𝒟γ\displaystyle\mathcal{D}_{\gamma} and 𝒟δ\displaystyle\mathcal{D}_{\delta} are interpolation spaces for the pair [𝒟α,𝒟β]\displaystyle[\mathcal{D}_{\alpha},\mathcal{D}_{\beta}]. Furthermore, by Theorem 2.4, we know that CφCψ\displaystyle C_{\varphi}-C_{\psi} is compact on 𝒟δ\displaystyle\mathcal{D}_{\delta}. Appealing to [7] Theorem 9.16 we see that φ\displaystyle\varphi and ψ\displaystyle\psi must have the same first order data at each ζF\displaystyle\zeta\in F; [14] Theorem 6 guarantees us that φ\displaystyle\varphi and ψ\displaystyle\psi have the same second order data at each ζF\displaystyle\zeta\in F. ∎

As in [14], we can extend the previous result to a stronger result for the class of linear fractional symbols. We first have one final lemma which is also a known result on a variety of spaces; we include references where appropriate and a proof for the spaces not explicitly referenced. Recall again that any nonconstant linear fractional self-map of 𝔻\displaystyle\mathbb{D} is univalent and induces a bounded composition operator on each weighted Dirichlet space.

Lemma 3.6.

Let γ1\displaystyle\gamma\geq-1 and let φ\displaystyle\varphi be a linear fractional self-map of 𝔻\displaystyle\mathbb{D}. Then Cφ\displaystyle C_{\varphi} is compact on 𝒟γ\displaystyle\mathcal{D}_{\gamma} if and only if φ<1\displaystyle\|\varphi\|_{\infty}<1.

Proof.

If 1γ<0\displaystyle-1\leq\gamma<0, then the statement follows from [18] Theorem 2.1 and the remarks following the theorem. For γ=1\displaystyle\gamma=1, the Hardy space case, see [20] (pages 23, 29 - 31); the case for γ>1\displaystyle\gamma>1 follows similarly.

Finally, when 0γ<1\displaystyle 0\leq\gamma<1, choose β>1\displaystyle\beta>1. The spaces 𝒟γ\displaystyle\mathcal{D}_{\gamma} and 𝒟1=H2\displaystyle\mathcal{D}_{1}=H^{2} are interpolation spaces for the pair [S2,𝒟β]\displaystyle[S^{2},\mathcal{D}_{\beta}]; recall S2=𝒟1\displaystyle S^{2}=\mathcal{D}_{-1}. Now, if Cφ\displaystyle C_{\varphi} is compact on 𝒟γ\displaystyle\mathcal{D}_{\gamma}, it follows from Theorem 2.4 that Cφ\displaystyle C_{\varphi} is compact on H2\displaystyle H^{2}. Then φ<1\displaystyle\|\varphi\|_{\infty}<1 since φ\displaystyle\varphi is linear fractional. Conversely, suppose that φ<1\displaystyle\|\varphi\|_{\infty}<1. Then Cφ\displaystyle C_{\varphi} is compact on H2\displaystyle H^{2} and hence on 𝒟γ\displaystyle\mathcal{D}_{\gamma} by Theorem 2.4. ∎

Corollary 3.7.

Let φ\displaystyle\varphi and ψ\displaystyle\psi be linear fractional self-maps of 𝔻\displaystyle\mathbb{D} and let γ1\displaystyle\gamma\geq-1. Then CφCψ\displaystyle C_{\varphi}-C_{\psi} is compact on 𝒟γ\displaystyle\mathcal{D}_{\gamma} if and only φ=ψ\displaystyle\varphi=\psi or both Cφ\displaystyle C_{\varphi} and Cψ\displaystyle C_{\psi} are compact on 𝒟γ\displaystyle\mathcal{D}_{\gamma}.

Proof.

If φ=ψ\displaystyle\varphi=\psi or both Cφ\displaystyle C_{\varphi} and Cψ\displaystyle C_{\psi} are compact on 𝒟γ\displaystyle\mathcal{D}_{\gamma}, then it is clear that CφCψ\displaystyle C_{\varphi}-C_{\psi} is compact on 𝒟γ\displaystyle\mathcal{D}_{\gamma}. Conversely, suppose that CφCψ\displaystyle C_{\varphi}-C_{\psi} is compact on 𝒟γ\displaystyle\mathcal{D}_{\gamma}. If either Cφ\displaystyle C_{\varphi} or Cψ\displaystyle C_{\psi} is compact on 𝒟γ\displaystyle\mathcal{D}_{\gamma}, then both Cφ\displaystyle C_{\varphi} and Cψ\displaystyle C_{\psi} are compact on 𝒟γ\displaystyle\mathcal{D}_{\gamma} by the fact that the compact operators form a linear subspace within the set of all bounded operators on 𝒟γ\displaystyle\mathcal{D}_{\gamma}. Thus we may further assume that neither Cφ\displaystyle C_{\varphi} nor Cψ\displaystyle C_{\psi} is compact on 𝒟γ\displaystyle\mathcal{D}_{\gamma}. It follows now by Lemma 3.6 that φ=1\displaystyle\|\varphi\|_{\infty}=1 and ψ=1\displaystyle\|\psi\|_{\infty}=1, i.e. φ\displaystyle\varphi and ψ\displaystyle\psi have contact with the boundary of 𝔻\displaystyle\mathbb{D}.

We now consider a few cases. For γ>1\displaystyle\gamma>1, see [14] Corollary 2. Next assume that γ=1\displaystyle\gamma=-1. Then Lemma 3.3 implies that φ\displaystyle\varphi and ψ\displaystyle\psi have the same first order data. Applying Theorem 3.5(a) we see that φ\displaystyle\varphi and ψ\displaystyle\psi must also have the same second order data and hence φ=ψ\displaystyle\varphi=\psi by Lemma 2.2. A similar argument holds for γ=0\displaystyle\gamma=0 but we will use interpolation.

Next consider 1<γ1\displaystyle-1<\gamma\leq 1. We again use an interpolation scheme and choose δ\displaystyle\delta and β\displaystyle\beta with 1<γ1<δ<β\displaystyle-1<\gamma\leq 1<\delta<\beta. Here we have 𝒟γ\displaystyle\mathcal{D}_{\gamma} and 𝒟δ\displaystyle\mathcal{D}_{\delta} as interpolation spaces for the pair [S2,𝒟β]\displaystyle[S^{2},\mathcal{D}_{\beta}]. Since CφCψ\displaystyle C_{\varphi}-C_{\psi} is compact on 𝒟γ\displaystyle\mathcal{D}_{\gamma}, it follows by Theorem 2.4 that CφCψ\displaystyle C_{\varphi}-C_{\psi} is compact on 𝒟δ\displaystyle\mathcal{D}_{\delta} and thus φ\displaystyle\varphi and ψ\displaystyle\psi have the same first order data by [7] Theorem 9.16. Theorem 3.5(b) then implies that φ\displaystyle\varphi and ψ\displaystyle\psi have the same second order data and hence φ=ψ\displaystyle\varphi=\psi by Lemma 2.2. ∎

Acknowledgments

Part of this work is taken from the second author’s Ph.D. dissertation written at the University of Virginia under the direction of Professor Barbara D. MacCluer.

References

  • [1] Bruce A. Barnes, Continuity properties of the spectrum of operators on Lebesgue spaces, Proc. Amer. Math. Soc. 106 (1989), no. 2, 415–421. MR 969515
  • [2] by same author, Interpolation of spectrum of bounded operators on Lebesgue spaces, Proceedings of the Seventh Great Plains Operator Theory Seminar (Lawrence, KS, 1987), vol. 20, 1990, pp. 359–378. MR 1065835
  • [3] Paul S. Bourdon, Components of linear-fractional composition operators, J. Math. Anal. Appl. 279 (2003), no. 1, 228–245. MR 1970503
  • [4] David M. Boyd, Composition operators on the Bergman space, Colloq. Math. 34 (1975/76), no. 1, 127–136. MR 407644
  • [5] Alberto P. Calderón, Intermediate spaces and interpolation, the complex method, Studia Math. 24 (1964), 113–190. MR 167830
  • [6] John B. Conway, A course in functional analysis, second ed., Graduate Texts in Mathematics, vol. 96, Springer-Verlag, New York, 1990. MR 1070713
  • [7] Carl C. Cowen and Barbara D. MacCluer, Composition operators on spaces of analytic functions, Studies in Advanced Mathematics, CRC Press, Boca Raton, FL, 1995. MR 1397026
  • [8] Michael Cwikel, Real and complex interpolation and extrapolation of compact operators, Duke Math. J. 65 (1992), no. 2, 333–343. MR 1150590
  • [9] Katherine C. Heller, Composition Operators on S2(D), ProQuest LLC, Ann Arbor, MI, 2010, Thesis (Ph.D.)–University of Virginia. MR 2844199
  • [10] Domingo A. Herrero and Karen Saxe Webb, Spectral continuity in complex interpolation, Math. Balkanica (N.S.) 3 (1989), no. 3-4, 325–336. MR 1048054
  • [11] Barbara D. MacCluer, Components in the space of composition operators, Integral Equations Operator Theory 12 (1989), no. 5, 725–738. MR 1009027
  • [12] Barbara D. MacCluer and Joel H. Shapiro, Angular derivatives and compact composition operators on the Hardy and Bergman spaces, Canad. J. Math. 38 (1986), no. 4, 878–906. MR 854144
  • [13] John E. McCarthy, Geometric interpolation between Hilbert spaces, Ark. Mat. 30 (1992), no. 2, 321–330. MR 1289759
  • [14] Jennifer Moorhouse, Compact differences of composition operators, J. Funct. Anal. 219 (2005), no. 1, 70–92. MR 2108359
  • [15] Eric A. Nordgren, Composition operators, Canadian J. Math. 20 (1968), 442–449. MR 223914
  • [16] Matthew A. Pons, The spectrum of a composition operator and Calderón’s complex interpolation, Topics in operator theory. Volume 1. Operators, matrices and analytic functions, Oper. Theory Adv. Appl., vol. 202, Birkhäuser Verlag, Basel, 2010, pp. 451–467. MR 2723292
  • [17] Karen Saxe, Compactness-like operator properties preserved by complex interpolation, Ark. Mat. 35 (1997), no. 2, 353–362. MR 1478785
  • [18] Joel H. Shapiro, Compact composition operators on spaces of boundary-regular holomorphic functions, Proc. Amer. Math. Soc. 100 (1987), no. 1, 49–57. MR 883400
  • [19] by same author, The essential norm of a composition operator, Ann. of Math. (2) 125 (1987), no. 2, 375–404. MR 881273
  • [20] by same author, Composition operators and classical function theory, Universitext: Tracts in Mathematics, Springer-Verlag, New York, 1993. MR 1237406
  • [21] Joel H. Shapiro and Peter D. Taylor, Compact, nuclear, and Hilbert-Schmidt composition operators on H2\displaystyle H^{2}, Indiana Univ. Math. J. 23 (1973/74), 471–496. MR 326472