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

Spectrum of a Composition Operator with Automorphic Symbol

Robert F. Allen1, Thong M. Le2, and Matthew A. Pons3 1Department of Mathematics, University of Wisconsin-La Crosse 2Department of Computer Science, University of California 3Department of Mathematics, North Central College [email protected], thmle@ucdavis@edu, [email protected]
Abstract.

We give a complete characterization of the spectrum of composition operators, induced by an automorphism of the open unit disk, acting on a family of Banach spaces of analytic functions that includes the Bloch space and BMOA. We show that for parabolic and hyperbolic automorphisms, the spectrum is the unit circle. For the case of elliptic automorphisms, the spectrum is either the unit circle or a finite cyclic subgroup of the unit circle.

Key words and phrases:
Composition operator, Spectrum, Automorphism
2010 Mathematics Subject Classification:
primary: 47B33, 47A10; secondary: 30H05.

1. Introduction

For an analytic self-map φ\displaystyle\varphi of the open unit disk 𝔻\displaystyle\mathbb{D} and a Banach space X\displaystyle X of functions analytic on 𝔻\displaystyle\mathbb{D}, we define the composition operator with symbol φ\displaystyle\varphi, denoted Cφ\displaystyle C_{\varphi}, by the rule Cφf=fφ\displaystyle C_{\varphi}f=f\circ\varphi for all fX\displaystyle f\in X. The study of composition operators began formally with Nordgren’s paper [12], where he explored properties of composition operators acting on the Hardy Hilbert space H2\displaystyle H^{2}. Since then, the study has proved to be an active area of research, most likely due to the fact that the study of such operators lies at the intersection of complex function theory and operator theory.

The spectrum of Cφ\displaystyle C_{\varphi} has been studied on many classical spaces of analytic functions, such as the Hardy spaces, Bergman spaces, weighted Hardy and Bergman spaces, Besov spaces, and the Dirichlet space. The interested reader is directed to [2] for general references.

The motivation of this paper was to determine the spectrum of a composition operator, induced by a disk automorphism, acting on the Bloch space. The Bloch space is the largest space of analytic functions on 𝔻\displaystyle\mathbb{D} that is Möbius invariant. This is one reason the Bloch space is a welcoming environment to study composition operators. The techniques developed apply to a larger class of spaces that includes the Bloch space.

The purpose of this paper is to determine the spectrum of Cφ\displaystyle C_{\varphi} acting on a family of Banach spaces, where φ\displaystyle\varphi is a disk automorphism. The spectrum will depend on the fixed point classification of the automorphisms of 𝔻\displaystyle\mathbb{D}. This is a standard approach to the study of composition operators induced by automorphisms. We show the spectrum of Cφ\displaystyle C_{\varphi}, acting on a particular family of Banach spaces, induced by a disk automorphism, must be a subset of the unit circle 𝔻\displaystyle\partial\mathbb{D}, and in some instances is the entire unit circle. Finally, we compare these results to particular examples of classical spaces.

2. Preliminaries

2.1. Automorphisms

The automorphisms of the open unit disk 𝔻={z:|z|<1}\displaystyle\mathbb{D}=\{z\in\mathbb{C}:|z|<1\} are precisely the analytic bijections on 𝔻\displaystyle\mathbb{D} which have the form

φ(z)=λaz1a¯z\displaystyle\varphi(z)=\lambda\frac{a-z}{1-\overline{a}z}

where λ\displaystyle\lambda is a unimodular constant and a\displaystyle a is a point in 𝔻\displaystyle\mathbb{D}. These automorphisms form a group under composition denoted by Aut(𝔻)\displaystyle\mathrm{Aut}(\mathbb{D}). Every element of Aut(𝔻)\displaystyle\mathrm{Aut}(\mathbb{D}) has two fixed points (counting multiplicity), and thus can be classified by the location of the fixed points:

elliptic:

one fixed point in 𝔻\displaystyle\mathbb{D} and one in the complement of 𝔻¯\displaystyle\overline{\mathbb{D}};

parabolic:

one fixed point on the unit circle 𝔻\displaystyle\partial\mathbb{D} (of multiplicity 2);

hyperbolic:

two distinct fixed points on 𝔻\displaystyle\partial\mathbb{D}.

Two disk automorphisms φ\displaystyle\varphi and ψ\displaystyle\psi are conformally equivalent if there exists a disk automorphism τ\displaystyle\tau for which ψ=τφτ1\displaystyle\psi=\tau\circ\varphi\circ\tau^{-1}. Many properties of automorphisms are preserved under conformal equivalence. The main advantage of conformal equivalence is in the placement of the fixed points. Every elliptic disk automorphism is conformally equivalent to one whose fixed point in 𝔻\displaystyle\mathbb{D} is the origin.

Lemma 2.1.

Let φ\displaystyle\varphi be an elliptic disk automorphism with fixed point a\displaystyle a in 𝔻\displaystyle\mathbb{D}. Then φ\displaystyle\varphi is conformally equivalent to ψ(z)=λz\displaystyle\psi(z)=\lambda z where λ=φ(a)\displaystyle\lambda=\varphi^{\prime}(a).

Proof.

Let τa\displaystyle\tau_{a} be the involution automorphism which interchanges 0 and a\displaystyle a, that is

τa(z)=az1a¯z.\displaystyle\tau_{a}(z)=\frac{a-z}{1-\overline{a}z}.

Define ψ=τaφτa1\displaystyle\psi=\tau_{a}\circ\varphi\circ\tau_{a}^{-1} on 𝔻\displaystyle\mathbb{D}. Since a\displaystyle a is a fixed point of φ\displaystyle\varphi, ψ\displaystyle\psi fixes the origin, and thus is a rotation. So there exists an unimodular constant λ\displaystyle\lambda such that ψ(z)=λz\displaystyle\psi(z)=\lambda z. To complete the proof, we will show λ=φ(a)\displaystyle\lambda=\varphi^{\prime}(a). Observe ψ(z)=λ\displaystyle\psi^{\prime}(z)=\lambda for all z𝔻\displaystyle z\in\mathbb{D}. In particular

λ=ψ(0)=τa(φ(τa(0)))φ(τa(0))τa(0)=φ(a)τa(a)τa(0)=φ(a).\displaystyle\lambda=\psi^{\prime}(0)=\tau_{a}^{\prime}(\varphi(\tau_{a}(0)))\varphi^{\prime}(\tau_{a}(0))\tau_{a}^{\prime}(0)=\varphi^{\prime}(a)\tau_{a}^{\prime}(a)\tau_{a}^{\prime}(0)=\varphi^{\prime}(a).

Thus φ\displaystyle\varphi is conformally equivalent to the rotation ψ(z)=φ(a)z\displaystyle\psi(z)=\varphi^{\prime}(a)z. ∎

Every parabolic disk automorphism is conformally equivalent to one whose fixed point (of multiplicity 2) is 1. The following Lemma is found as Exercise 2.3.5c of [2], and a complete proof can be found in [13].

Lemma 2.2.

[13, Lemma 4.1.2] Let φ\displaystyle\varphi be a parabolic disk automorphism. Then φ\displaystyle\varphi is conformally equivalent to either ψ1(z)=(1+i)z1z+i1\displaystyle\psi_{1}(z)=\displaystyle\frac{(1+i)z-1}{z+i-1} or ψ2(z)=(1i)z1zi1\displaystyle\psi_{2}(z)=\displaystyle\frac{(1-i)z-1}{z-i-1}.

Every hyperbolic disk automorphism is conformally equivalent to one whose fixed points in 𝔻\displaystyle\partial\mathbb{D} are ±1\displaystyle\pm 1.

Lemma 2.3.

[12, Theorem 6] Let φ\displaystyle\varphi be a hyperbolic disk automorphism. Then φ\displaystyle\varphi is conformally equivalent to ψ(z)=z+r1+rz\displaystyle\psi(z)=\displaystyle\frac{z+r}{1+rz} for some r(0,1)\displaystyle r\in(0,1).

2.2. The Space of Bounded Analytic Functions

The set of analytic functions on 𝔻\displaystyle\mathbb{D} is denoted by H(𝔻)\displaystyle H(\mathbb{D}). The space of bounded analytic functions on 𝔻\displaystyle\mathbb{D}, denoted H=H(𝔻)\displaystyle H^{\infty}=H^{\infty}(\mathbb{D}), is a Banach space under the norm

f=supz𝔻|f(z)|.\displaystyle\|f\|_{\infty}=\sup_{z\in\mathbb{D}}\,|f(z)|.

The bounded analytic functions on 𝔻\displaystyle\mathbb{D} is a rich space containing many interesting types of functions, such as polynomials and Blaschke products. In addition, the disk algebra 𝒜(𝔻)\displaystyle\mathcal{A}(\mathbb{D}), the set of analytic functions on 𝔻\displaystyle\mathbb{D} continuous to 𝔻\displaystyle\partial\mathbb{D}, is a closed subspace of H\displaystyle H^{\infty}.

The following two families of functions will be used in the next section. To prove these functions are in H\displaystyle H^{\infty}, we take a geometric approach using conformal mappings of the plane. To this effect, let \displaystyle\mathbb{H}_{\ell} and r\displaystyle\mathbb{H}_{r} denote the open left and right half planes respectively, i.e. ={Rez<0}\displaystyle\mathbb{H}_{\ell}=\{\mathrm{Re}\,z<0\} and r={Rez>0}\displaystyle\mathbb{H}_{r}=\{\mathrm{Re}\,z>0\}.

Lemma 2.4.

For s0\displaystyle s\geq 0, the function fs(z)=exp(s(z+1)z1)\displaystyle f_{s}(z)=\exp\left(\displaystyle\frac{s(z+1)}{z-1}\right) is in H\displaystyle H^{\infty}.

Proof.

If s=0\displaystyle s=0, then fs(z)=1\displaystyle f_{s}(z)=1. So, fs(z)\displaystyle f_{s}(z) is in H\displaystyle H^{\infty}. Now suppose s>0\displaystyle s>0. The function fs\displaystyle f_{s} is comprised of the functions

  1. (1)

    zz+1z1\displaystyle z\mapsto\frac{z+1}{z-1}; mapping 𝔻\displaystyle\mathbb{D} onto \displaystyle\mathbb{H}_{\ell},

  2. (2)

    zsz\displaystyle z\mapsto sz; mapping \displaystyle\mathbb{H}_{\ell} onto \displaystyle\mathbb{H}_{\ell},

  3. (3)

    zez\displaystyle z\mapsto e^{z}; mapping \displaystyle\mathbb{H}_{\ell} onto 𝔻{0}\displaystyle\mathbb{D}\setminus\{0\}.

z+1z1\displaystyle\frac{z+1}{z-1}sz\displaystyle szez\displaystyle e^{z}
Figure 1. Map fs(z)=exp(s(z+1)z1)\displaystyle f_{s}(z)=\exp\left(\displaystyle\frac{s(z+1)}{z-1}\right) for s>0\displaystyle s>0.

So fs\displaystyle f_{s} maps 𝔻\displaystyle\mathbb{D} into 𝔻\displaystyle\mathbb{D}, as depicted in Figure 1, and thus fs(z)\displaystyle f_{s}(z) is an element of H\displaystyle H^{\infty}. ∎

Lemma 2.5.

For real value t\displaystyle t, the function ft(z)=(1+z1z)it\displaystyle f_{t}(z)=\left(\displaystyle\frac{1+z}{1-z}\right)^{it} is in H\displaystyle H^{\infty}.

Proof.

For t=0\displaystyle t=0, ft\displaystyle f_{t} is identically 1, and thus is in H\displaystyle H^{\infty}. Now suppose t>0\displaystyle t>0. We will rewrite the function ft\displaystyle f_{t} as

ft(z)=exp(itlog1+z1z),\displaystyle f_{t}(z)=\exp\left(it\log\frac{1+z}{1-z}\right),

where log\displaystyle\log is the principle branch of the logarithm. Then ft\displaystyle f_{t} is comprised of the functions

  1. (1)

    z1+z1z\displaystyle z\mapsto\frac{1+z}{1-z}; mapping 𝔻\displaystyle\mathbb{D} onto r\displaystyle\mathbb{H}_{r},

  2. (2)

    zlogz\displaystyle z\mapsto\log z; mapping Hr\displaystyle H_{r} onto the horizontal strip Sh={0<Imz<2π}\displaystyle S_{h}=\{0<\mathrm{Im}\,z<2\pi\},

  3. (3)

    zitz\displaystyle z\mapsto itz; mapping Sh\displaystyle S_{h} onto the vertical strip Sv={2π<Rez<0}\displaystyle S_{v}=\{-2\pi<\mathrm{Re}\,z<0\},

  4. (4)

    zez\displaystyle z\mapsto e^{z}; mapping Sv\displaystyle S_{v} into A(e2π,1)={e2π<|z|<1}\displaystyle A(e^{-2\pi},1)=\{e^{-2\pi}<|z|<1\}.

1+z1z\displaystyle\frac{1+z}{1-z}logz\displaystyle\log zitz\displaystyle itzez\displaystyle e^{z}
Figure 2. Map ft(z)=exp(itlog1+z1z)\displaystyle f_{t}(z)=\exp\left(it\log\displaystyle\frac{1+z}{1-z}\right) for t>0\displaystyle t>0.

So ft\displaystyle f_{t} maps 𝔻\displaystyle\mathbb{D} into A(e2π,1)𝔻\displaystyle A(e^{-2\pi},1)\subseteq\mathbb{D}, as depicted in Figure 2. In the case of t<0\displaystyle t<0, the vertical strip Sv\displaystyle S_{v} becomes {0<Rez<2π}\displaystyle\{0<\mathrm{Re}\,z<2\pi\}. The map zez\displaystyle z\mapsto e^{z} takes Sv\displaystyle S_{v} into A(1,e2π)e2π𝔻\displaystyle A(1,e^{2\pi})\subseteq e^{2\pi}\mathbb{D}, as depicted in Figure 3.

1+z1z\displaystyle\frac{1+z}{1-z}logz\displaystyle\log zitz\displaystyle itzez\displaystyle e^{z}
Figure 3. Map ft(z)=exp(itlog1+z1z)\displaystyle f_{t}(z)=\exp\left(it\log\displaystyle\frac{1+z}{1-z}\right) for t<0\displaystyle t<0.

In either case, ft(z)\displaystyle f_{t}(z) is an element of H\displaystyle H^{\infty} since ft<e2π\displaystyle\|f_{t}\|_{\infty}<e^{2\pi} for all t\displaystyle t\in\mathbb{R}. ∎

These functions above, together with the monomials, play such a pivitol role in Section 3 that we denote the union of these functions by \displaystyle\mathcal{F}, i.e.

={fs:s0}{ft:t}{zk:k}.\displaystyle\mathcal{F}=\left\{f_{s}:s\geq 0\right\}\cup\left\{f_{t}:t\in\mathbb{R}\right\}\cup\left\{z^{k}:k\in\mathbb{N}\right\}.

2.3. Spectrum of Cφ\displaystyle C_{\varphi}

In this section we collect useful results regarding the spectrum of operators on Banach spaces. For a bounded linear operator T\displaystyle T on a Banach space X\displaystyle X, the spectrum of T\displaystyle T is given by

σ(T)={λ:TλI is not invertible}\displaystyle\sigma(T)=\{\lambda\in\mathbb{C}:T-\lambda I\text{ is not invertible}\}

where I\displaystyle I denotes the identity operator on X\displaystyle X. The spectrum is a nonempty, closed subset of \displaystyle\mathbb{C}. The spectral radius of T\displaystyle T is given by

ρ(T)=sup{|λ|:λσ(T)}.\displaystyle\rho(T)=\sup\,\{|\lambda|:\lambda\in\sigma(T)\}.

Due to the fact that the spectrum is closed, we have the spectrum of T\displaystyle T is contained in the closed disk centered at the origin of radius ρ(T)\displaystyle\rho(T).

Determining the spectrum of a particular composition operator can be difficult depending on the symbol of the operator and the space on which it is acting. However, the difficulties can be avoided if the operator is similar to a “simpler” operator. Linear operators S\displaystyle S and T\displaystyle T (not necessarily bounded) on a Banach space X\displaystyle X are similar if there exists a bounded linear operator U\displaystyle U on X\displaystyle X, having bounded inverse, such that T=USU1\displaystyle T=USU^{-1}. If S\displaystyle S and T\displaystyle T are both bounded operators, then similarity preserves the spectrum.

Theorem 2.6.

Let S\displaystyle S and T\displaystyle T be bounded operators on a Banach space X\displaystyle X. If S\displaystyle S and T\displaystyle T are similar, then σ(S)=σ(T)\displaystyle\sigma(S)=\sigma(T).

Proof.

Suppose S\displaystyle S and T\displaystyle T are similar operators on X\displaystyle X. By definition, there exists an invertible, bounded operator U\displaystyle U such that T=USU1\displaystyle T=USU^{-1}. Let λ\displaystyle\lambda\in\mathbb{C} and observe that,

TλI\displaystyle T-\lambda I =USU1λI\displaystyle=USU^{-1}-\lambda I
=USU1λUU1\displaystyle=USU^{-1}-\lambda UU^{-1}
=USU1U(λI)U1\displaystyle=USU^{-1}-U(\lambda I)U^{-1}
=U(SλI)U1.\displaystyle=U(S-\lambda I)U^{-1}.

Thus, we have that SλI\displaystyle S-\lambda I is not invertible if and only if TλI\displaystyle T-\lambda I is not invertible. Therefore σ(S)=σ(T)\displaystyle\sigma(S)=\sigma(T). ∎

3. Main Results

In this section, we determine the spectrum of Cφ\displaystyle C_{\varphi} for φ\displaystyle\varphi a disk automorphism acting on a particular family of Banach spaces of analytic functions. The spaces we consider will be denoted by 𝒳\displaystyle\mathcal{X} and have the following properties:

  1. (i)

    𝒳\displaystyle\mathcal{X} contains \displaystyle\mathcal{F},

  2. (ii)

    for all φAut(𝔻)\displaystyle\varphi\in\mathrm{Aut}(\mathbb{D}), Cφ\displaystyle C_{\varphi} is bounded on 𝒳\displaystyle\mathcal{X} and ρ(Cφ)=1\displaystyle\rho(C_{\varphi})=1.

The set of automorphisms of 𝔻\displaystyle\mathbb{D}, as seen previously, is a very nice subset of the analytic self-maps of 𝔻\displaystyle\mathbb{D}. By property (ii), every composition operator induced by a disk automorphism is bounded on 𝒳\displaystyle\mathcal{X}. In fact, every such composition operator is invertible. This result, that we prove below, can be viewed as a consequence of Theorem 1.6 of [2].

Proposition 3.1.

Let φ\displaystyle\varphi be a disk automorphism and Cφ\displaystyle C_{\varphi} the induced composition operator on 𝒳\displaystyle\mathcal{X}. Then Cφ\displaystyle C_{\varphi} is invertible with inverse Cφ1=Cφ1\displaystyle C^{-1}_{\varphi}=C_{\varphi^{-1}}.

Proof.

Since φAut(𝔻)\displaystyle\varphi\in\text{Aut}(\mathbb{D}), φ\displaystyle\varphi is invertible, and φ1\displaystyle\varphi^{-1} is an automorphism. The composition operator Cφ1\displaystyle C_{\varphi^{-1}} is bounded by property (ii) and

Cφ(Cφ1(f))=Cφ(fφ1)=fφ1φ=f\displaystyle C_{\varphi}\left(C_{\varphi^{-1}}(f)\right)=C_{\varphi}(f\circ\varphi^{-1})=f\circ\varphi^{-1}\circ\varphi=f
Cφ1(Cφ(f))=Cφ1(fφ)=fφφ1=f.\displaystyle C_{\varphi^{-1}}\left(C_{\varphi}(f)\right)=C_{\varphi^{-1}}(f\circ\varphi)=f\circ\varphi\circ\varphi^{-1}=f\ .

Therefore, Cφ\displaystyle C_{\varphi} is invertible with Cφ1=Cφ1\displaystyle C_{\varphi}^{-1}=C_{\varphi^{-1}}. ∎

Since the spectral radius of Cφ\displaystyle C_{\varphi} on 𝒳\displaystyle\mathcal{X} is 1 for φAut(𝔻)\displaystyle\varphi\in\mathrm{Aut}(\mathbb{D}), we see that the search for the spectrum can be restricted to subsets of 𝔻¯\displaystyle\overline{\mathbb{D}}. However, our search can be refined further to subsets of the unit circle.

Theorem 3.2.

Let φ\displaystyle\varphi be a a disk automorphism and Cφ\displaystyle C_{\varphi} the induced composition operator on 𝒳\displaystyle\mathcal{X}. Then σ(Cφ)𝔻\displaystyle\sigma(C_{\varphi})\subseteq\partial\mathbb{D}.

Proof.

By property (ii) of 𝒳\displaystyle\mathcal{X}, we have ρ(Cφ)=1\displaystyle\rho(C_{\varphi})=1. So, σ(Cφ)𝔻¯\displaystyle\sigma(C_{\varphi})\subseteq\overline{\mathbb{D}}. Since, by Proposition 3.1, Cφ\displaystyle C_{\varphi} is invertible with the inverse Cφ1=Cφ1\displaystyle C^{-1}_{\varphi}=C_{\varphi^{-1}}, then 0σ(Cφ)\displaystyle 0\notin\sigma(C_{\varphi}). So, the function f(z)=z1\displaystyle f(z)=z^{-1} is analytic in some neighborhood of σ(Cφ)\displaystyle\sigma(C_{\varphi}). By the Spectral Mapping Theorem (see Theorem 5.14 of [10]), we have σ(fCφ)=f(σ(Cφ))\displaystyle\sigma(f\circ C_{\varphi})=f(\sigma(C_{\varphi})), and so,

σ(Cφ1)=σ(Cφ1)=σ(Cφ)1={λ1:λσ(Cφ)}.\displaystyle\sigma(C_{\varphi^{-1}})=\sigma(C^{-1}_{\varphi})=\sigma(C_{\varphi})^{-1}=\{\lambda^{-1}:\lambda\in\sigma(C_{\varphi})\}\ .

Since φ1Aut(𝔻)\displaystyle\varphi^{-1}\in\mathrm{Aut}(\mathbb{D}), σ(Cφ1)𝔻¯.\displaystyle\sigma(C_{\varphi^{-1}})\subseteq\overline{\mathbb{D}}. Thus for λσ(Cφ)\displaystyle\lambda\in\sigma(C_{\varphi}), both λ\displaystyle\lambda and λ1\displaystyle\lambda^{-1} are in 𝔻¯\displaystyle\overline{\mathbb{D}}. This implies λ𝔻\displaystyle\lambda\in\partial\mathbb{D}. So σ(Cφ)𝔻\displaystyle\sigma(C_{\varphi})\subseteq\partial\mathbb{D}, as desired. ∎

Since the disk automorphisms are classified into three categories, according to fixed points, we will treat each type of automorphism separately. However, the strategy to determine σ(Cφ)\displaystyle\sigma(C_{\varphi}) is the same. For a disk automorphism φ\displaystyle\varphi, we have shown φ\displaystyle\varphi to be conformally equivalent to a particularly “nice” disk automorphism: in the elliptic case a disk automorphism that fixes 0, in the parabolic case a disk automorphism that fixes 1, and in the hyperbolic case a disk automorphism that fixes ±1\displaystyle\pm 1. In the next result, we show that conformally equivalent automorphisms induce similar composition operators on 𝒳\displaystyle\mathcal{X}. This result is not unique to the space 𝒳\displaystyle\mathcal{X}, but is true for any space for which automorphisms induce bounded composition operators (see pg. 250 of [2]).

Proposition 3.3.

Let φ\displaystyle\varphi and ψ\displaystyle\psi be conformally equivalent disk automorphisms. Then the induced composition operators Cφ\displaystyle C_{\varphi} and Cψ\displaystyle C_{\psi} on 𝒳\displaystyle\mathcal{X} are similar.

Proof.

Suppose φ\displaystyle\varphi and ψ\displaystyle\psi are conformally equivalent disk automorphisms. Then there exists a disk automorphism τ\displaystyle\tau such that ψ=τφτ1\displaystyle\psi=\tau\circ\varphi\circ\tau^{-1}. For f𝒳\displaystyle f\in\mathcal{X}, observe

Cψf=f(τφτ1)=((fτ)φ)τ1=(Cτ1CφCτ)f.\displaystyle C_{\psi}f=f\circ(\tau\circ\varphi\circ\tau^{-1})=((f\circ\tau)\circ\varphi)\circ\tau^{-1}=(C_{\tau^{-1}}C_{\varphi}C_{\tau})f.

Since Cτ1\displaystyle C_{\tau^{-1}} is bounded and invertible on 𝒳\displaystyle\mathcal{X} with Cτ11=Cτ\displaystyle C_{\tau^{-1}}^{-1}=C_{\tau}, then Cψ=Cτ1CφCτ11\displaystyle C_{\psi}=C_{\tau^{-1}}C_{\varphi}C_{\tau^{-1}}^{-1}. Therefore Cφ\displaystyle C_{\varphi} and Cψ\displaystyle C_{\psi} are similar. ∎

With Proposiition 3.3 and Lemmas 2.1, 2.2, and 2.3, it suffices to determine the spectrum of the “nice” disk automorphisms, since similarity of bounded operators preserves the spectrum.

Theorem 3.4.

Let φ\displaystyle\varphi be an elliptic disk automorphism with fixed point a\displaystyle a in 𝔻\displaystyle\mathbb{D}. Then the spectrum of Cφ\displaystyle C_{\varphi} acting on 𝒳\displaystyle\mathcal{X} is the closure of the positive powers of φ(a)\displaystyle\varphi^{\prime}(a). Moreover, this closure is a finite subgroup of the unit circle if φ(a)n=1\displaystyle\varphi^{\prime}(a)^{n}=1 for some natural number n\displaystyle n, and is the unit circle otherwise.

Proof.

By Lemma 2.1, φ\displaystyle\varphi is conformally equivalent to ψ(z)=λz\displaystyle\psi(z)=\lambda z where λ=φ(a)\displaystyle\lambda=\varphi^{\prime}(a). By Proposition 3.3, it suffices to show that σ(Cψ)\displaystyle\sigma(C_{\psi}) is the closure of the positive powers of λ\displaystyle\lambda. Let G=λ={λk:k}\displaystyle G=\langle\lambda\rangle=\{\lambda^{k}:k\in\mathbb{N}\}, which is a subset of 𝔻\displaystyle\partial\mathbb{D} since |λ|=1\displaystyle|\lambda|=1. For each k\displaystyle k\in\mathbb{N}, the function fk(z)=zk\displaystyle f_{k}(z)=z^{k} is in 𝒳\displaystyle\mathcal{X} by property (i), and we have (Cψfk)(z)=λkfk(z)\displaystyle(C_{\psi}f_{k})(z)=\lambda^{k}f_{k}(z). Thus λk\displaystyle\lambda^{k} is an eigenvalue of Cψ\displaystyle C_{\psi} corresponding to the eigenfunction fk\displaystyle f_{k}. So Gσ(Cψ)\displaystyle G\subseteq\sigma(C_{\psi}), and since the spectrum is closed, we have G¯σ(Cψ)=σ(Cφ)\displaystyle\overline{G}\subseteq\sigma(C_{\psi})=\sigma(C_{\varphi}). If the order of λ\displaystyle\lambda is infinite, then G\displaystyle G is dense in 𝔻\displaystyle\partial\mathbb{D}, and so G¯=𝔻\displaystyle\overline{G}=\partial\mathbb{D}.

Now suppose λ\displaystyle\lambda has order m<\displaystyle m<\infty. Then G={λk:k=1,,m}\displaystyle G=\{\lambda^{k}:k=1,\dots,m\}. So, G¯=G\displaystyle\overline{G}=G. We now wish to show σ(Cψ)G¯\displaystyle\sigma(C_{\psi})\subseteq\overline{G}. Since σ(Cψ)𝔻\displaystyle\sigma(C_{\psi})\subseteq\partial\mathbb{D} by Theorem 3.2, it suffices to show that if μ𝔻G¯\displaystyle\mu\in\partial\mathbb{D}\setminus\overline{G} then μσ(Cψ)\displaystyle\mu\notin\sigma(C_{\psi}). Suppose μ𝔻G¯\displaystyle\mu\in\partial\mathbb{D}\setminus\overline{G}.

Since μG¯\displaystyle\mu\notin\overline{G}, it clear that μG\displaystyle\mu\notin G and μm1\displaystyle\mu^{m}\neq 1. In order to show μσ(Cψ)\displaystyle\mu\notin\sigma(C_{\psi}), we will show that CψμI\displaystyle C_{\psi}-\mu I is invertible by proving that for every g𝒳\displaystyle g\in\mathcal{X}, there exists a unique f𝒳\displaystyle f\in\mathcal{X} such that fψμf=g\displaystyle f\circ\psi-\mu f=g.

Since the order of λ\displaystyle\lambda is m\displaystyle m, we have ψ(m)(z)=(ψψmtimes)(z)=λmz=z.\displaystyle\psi^{(m)}(z)=(\underbrace{\psi\circ\dots\circ\psi}_{m\mathrm{-times}})(z)=\lambda^{m}z=z. By repeated composition with ψ\displaystyle\psi, we obtain the system of linear equations:

fψμf\displaystyle f\circ\psi-\mu f =g\displaystyle=g
fψ(2)μ(fψ)\displaystyle f\circ\psi^{(2)}-\mu(f\circ\psi) =gψ\displaystyle=g\circ\psi
\displaystyle\;\,\vdots
fμ(fψ(m1))\displaystyle f-\mu(f\circ\psi^{(m-1)}) =gψ(m1).\displaystyle=g\circ\psi^{(m-1)}\ .

This system of linear equations can be expressed as the matrix equation Ax=b\displaystyle A\vec{x}=\vec{b} where

A=[μ10000μ100001100μ],x=[ffψfψ(m2)fψ(m1)], and b=[ggψgψ(m2)gψ(m1)].\displaystyle A=\left[\begin{matrix}-\mu&1&0&0&\cdots&0\\ 0&-\mu&1&0&\cdots&0\\ \vdots&0&\ddots&\ddots&&\vdots\\ \vdots&\vdots&\ddots&\ddots&\ddots&\vdots\\ 0&\vdots&&\ddots&\ddots&1\\ 1&0&\cdots&\cdots&0&-\mu\\ \end{matrix}\right],\vec{x}=\left[\begin{matrix}f\\ f\circ\psi\\ \vdots\\ \vdots\\ f\circ\psi^{(m-2)}\\ f\circ\psi^{(m-1)}\end{matrix}\right],\text{ and }\vec{b}=\left[\begin{matrix}g\\ g\circ\psi\\ \vdots\\ \vdots\\ g\circ\psi^{(m-2)}\\ g\circ\psi^{(m-1)}\end{matrix}\right].

The determinant of A\displaystyle A is (1)m(μm1)\displaystyle(-1)^{m}(\mu^{m}-1), which is not zero since μG\displaystyle\mu\notin G. Thus there is a unique solution for x\displaystyle\vec{x}. It gives us the unique solution f\displaystyle f, which is a finite linear combination of function in 𝒳\displaystyle\mathcal{X} of the form gψ(j1)\displaystyle g\circ\psi^{(j-1)} for j=1,,m\displaystyle j=1,\dots,m, and thus f\displaystyle f is in 𝒳\displaystyle\mathcal{X}. It follows that CψμI\displaystyle C_{\psi}-\mu I is invertible. So, μσ(Cψ)\displaystyle\mu\notin\sigma(C_{\psi}). Therefore, σ(Cφ)=σ(Cψ)G¯\displaystyle\sigma(C_{\varphi})=\sigma(C_{\psi})\subseteq\overline{G}. ∎

Theorem 3.5.

Let φ\displaystyle\varphi be a parabolic disk automorphism. Then the spectrum of Cφ\displaystyle C_{\varphi} acting on 𝒳\displaystyle\mathcal{X} is the unit circle.

Proof.

From Lemma 2.2, φ\displaystyle\varphi is conformally equivalent to either ψ1(z)=(1+i)z1z+i1\displaystyle\psi_{1}(z)=\displaystyle\frac{(1+i)z-1}{z+i-1} or ψ2(z)=(1i)z1zi1\displaystyle\psi_{2}(z)=\displaystyle\frac{(1-i)z-1}{z-i-1}. By Theorems 3.3 and 3.2 it suffices to show that 𝔻\displaystyle\partial\mathbb{D} is a subset of σ(Cψ1)\displaystyle\sigma(C_{\psi_{1}}) and σ(Cψ2)\displaystyle\sigma(C_{\psi_{2}}).

First suppose φ\displaystyle\varphi is conformally equivalent to ψ1\displaystyle\psi_{1}. Consider the function

fs(z)=exp(s(z+1)z1)\displaystyle f_{s}(z)=\exp\left(\displaystyle\frac{s(z+1)}{z-1}\right)

for s0\displaystyle s\geq 0. By property (i), fs\displaystyle f_{s} is in 𝒳\displaystyle\mathcal{X}. Observe

(Cψ1fs)(z)\displaystyle(C_{\psi_{1}}f_{s})(z) =fs(ψ1(z))=fs((1+i)z1z+i1)\displaystyle=f_{s}(\psi_{1}(z))=f_{s}\left(\frac{(1+i)z-1}{z+i-1}\right)
=exp(s((1+i)z1z+i1+1)(1+i)z1z+i11)=exp(s((1+i)z1+z+i1)(1+i)z1zi+1)\displaystyle=\exp\left(\frac{s\left(\frac{(1+i)z-1}{z+i-1}+1\right)}{\frac{(1+i)z-1}{z+i-1}-1}\right)=\exp\left(\frac{s((1+i)z-1+z+i-1)}{(1+i)z-1-z-i+1}\right)
=exp(s((2+i)z+i2)i(z1))=exp(s((12i)z+1+2i)z1)\displaystyle=\exp\left(\frac{s((2+i)z+i-2)}{i(z-1)}\right)=\exp\left(\frac{s((1-2i)z+1+2i)}{z-1}\right)
=exp(s(z+1)z12is)\displaystyle=\exp\left(\frac{s(z+1)}{z-1}-2is\right)
=ei(2s)fs(z).\displaystyle=e^{i(-2s)}f_{s}(z).

So, fs\displaystyle f_{s} is an eigenfunction of Cψ1\displaystyle C_{\psi_{1}} for s0\displaystyle s\geq 0. Then, 𝔻={ei(2s):s0}\displaystyle\partial\mathbb{D}=\{e^{i(-2s)}:s\geq 0\} is a subset of σ(Cψ1)\displaystyle\sigma(C_{\psi_{1}}). If φ\displaystyle\varphi is conformally equivalent to ψ2\displaystyle\psi_{2}, then by a similar calculation, we have

(Cψ2fs)(z)=e2isfs(z),\displaystyle(C_{\psi_{2}}f_{s})(z)=e^{2is}f_{s}(z),

and so 𝔻={e2is:s0}\displaystyle\partial\mathbb{D}=\{e^{2is}:s\geq 0\} is a subset of σ(Cψ2)\displaystyle\sigma(C_{\psi_{2}}). Therefore, σ(Cφ)=𝔻\displaystyle\sigma(C_{\varphi})=\partial\mathbb{D}, as desired. ∎

Theorem 3.6.

Let φ\displaystyle\varphi be a hyperbolic disk automorphism. Then the spectrum of Cφ\displaystyle C_{\varphi} acting on 𝒳\displaystyle\mathcal{X} is the unit circle.

Proof.

From Lemma 2.3, φ\displaystyle\varphi is conformally equivalent to ψ(z)=z+r1+rz\displaystyle\psi(z)=\displaystyle\frac{z+r}{1+rz} for some r(0,1)\displaystyle r\in(0,1). By Theorems 3.3 and 3.2 it suffices to show that 𝔻σ(Cψ)\displaystyle\partial\mathbb{D}\subseteq\sigma(C_{\psi}). Consider the function

ft(z)=(1+z1z)it\displaystyle f_{t}(z)=\left(\frac{1+z}{1-z}\right)^{it}

for t\displaystyle t\in\mathbb{R}. By property (i), ft\displaystyle f_{t} is in 𝒳\displaystyle\mathcal{X}. Observe

(Cψft)(z)\displaystyle(C_{\psi}f_{t})(z) =ft(ψ(z))=ft(z+r1+rz)\displaystyle=f_{t}(\psi(z))=f_{t}\left(\frac{z+r}{1+rz}\right)
=(1+z+r1+rz1z+r1+rz)it=(1+rz+z+r1+rzzr)it\displaystyle=\left(\frac{1+\frac{z+r}{1+rz}}{1-\frac{z+r}{1+rz}}\right)^{it}=\left(\frac{1+rz+z+r}{1+rz-z-r}\right)^{it}
=((r+1)z+(r+1)(r1)z(r1))it\displaystyle=\left(\frac{(r+1)z+(r+1)}{(r-1)z-(r-1)}\right)^{it}
=(r+1r1)itft(z).\displaystyle=\left(\frac{r+1}{r-1}\right)^{it}f_{t}(z).

So, ft\displaystyle f_{t} is an eigenfunction of Cψ\displaystyle C_{\psi} for t\displaystyle t real. Then 𝔻={(r+1r1)it:0<r<1,t}\displaystyle\partial\mathbb{D}=\left\{\left(\frac{r+1}{r-1}\right)^{it}:0<r<1,t\in\mathbb{R}\right\} is a subset of σ(Cψ)=σ(Cφ)\displaystyle\sigma(C_{\psi})=\sigma(C_{\varphi}). Therefore σ(Cφ)=𝔻\displaystyle\sigma(C_{\varphi})=\partial\mathbb{D}, as desired. ∎

4. Examples & Comparisons

In this section we first consider examples of spaces that satisfy the properties of 𝒳\displaystyle\mathcal{X}. For these spaces, our results characterize the spectrum of composition operators induced by disk automorphisms. Lastly, we consider spaces that do not satisfy the properties of 𝒳\displaystyle\mathcal{X} but for which the spectrum of composition operators induced by automorphisms is known. We will compare the spectra for those spaces with the characterization for 𝒳\displaystyle\mathcal{X}.

4.1. Examples

First, we will discuss examples of spaces that satisfy the properties of 𝒳\displaystyle\mathcal{X}.

4.1.1. Bounded analytic functions

The property (i) of 𝒳\displaystyle\mathcal{X} is satisfied by H\displaystyle H^{\infty} by Lemmas 2.4 and 2.5. In fact, on H\displaystyle H^{\infty}, any analytic self-map of 𝔻\displaystyle\mathbb{D} induces a bounded composition operator Cφ\displaystyle C_{\varphi} such that Cφ=1\displaystyle\|C_{\varphi}\|=1. Equality is achieved since H\displaystyle H^{\infty} contains the constant function 1. The spectral radius formula (see Theorem 5.15 of [10]) then implies that ρ(Cφ)=1\displaystyle\rho(C_{\varphi})=1. Thus, property (ii) is satisfied. This H\displaystyle H^{\infty} belongs to the family of Banach spaces of analytic functions 𝒳\displaystyle\mathcal{X}.

4.1.2. Bloch space

The Bloch space on 𝔻\displaystyle\mathbb{D}, denoted =(𝔻)\displaystyle\mathcal{B}=\mathcal{B}(\mathbb{D}), is the space of analytic functions on 𝔻\displaystyle\mathbb{D} such that βf=supz𝔻(1|z|2)|f(z)|<\displaystyle\beta_{f}=\displaystyle\sup_{z\in\mathbb{D}}\,(1-|z|^{2})|f^{\prime}(z)|<\infty. The quantity βf\displaystyle\beta_{f} is a semi-norm, called the Bloch semi-norm. The Bloch space is a Banach space under the norm

f=|f(0)|+βf.\displaystyle\|f\|_{\mathcal{B}}=|f(0)|+\beta_{f}.

It is well-known that \displaystyle\mathcal{B} is a Banach space of analytic functions that contains H\displaystyle H^{\infty}, and thus satisfies property (i) of 𝒳\displaystyle\mathcal{X}. In fact, every analytic self-map of 𝔻\displaystyle\mathbb{D} induces a bounded composition operator on \displaystyle\mathcal{B} (see pg. 126 [1]). Donaway, in his Ph.D. thesis, Corollary 3.9 of [3], proved the spectral radius of every composition operator induced by an analytic function on 𝔻\displaystyle\mathbb{D}, and in particular the disk automorphisms, is 1. So the Bloch space satisfies all the properties of 𝒳\displaystyle\mathcal{X}.

4.1.3. Analytic functions of bounded mean osciallation

The space of analytic functions on 𝔻\displaystyle\mathbb{D} with bounded mean oscillation on 𝔻\displaystyle\partial\mathbb{D}, denoted BMOA\displaystyle BMOA, is defined to be the set of functions in H(𝔻)\displaystyle H(\mathbb{D}) such that

f=supz𝔻fτaf(z)H2<,\displaystyle\|f\|_{*}=\sup_{z\in\mathbb{D}}\,\|f\circ\tau_{a}-f(z)\|_{H^{2}}<\infty,

where H2\displaystyle H^{2} is defined in Section 4.2.1. The space BMOA\displaystyle BMOA is a Banach space under the norm

fBMOA=|f(0)|+f.\displaystyle\|f\|_{BMOA}=|f(0)|+\|f\|_{*}.

It is well-known that BMOA\displaystyle BMOA is a Banach space of analytic functions, a subspace of the Bloch space, and contains H\displaystyle H^{\infty} as a subspace since fBMOA3f\displaystyle\|f\|_{BMOA}\leq 3\|f\|_{\infty}. Thus property (i) is satisfied by BMOA\displaystyle BMOA. The following result shows property (ii) is satisfied by BMOA\displaystyle BMOA also.

Theorem 4.1.

Let φ\displaystyle\varphi be an analytic self-map of 𝔻\displaystyle\mathbb{D}. Then Cφ\displaystyle C_{\varphi} acting on BMOA\displaystyle BMOA is bounded and ρ(Cφ)=1\displaystyle\rho(C_{\varphi})=1.

Proof.

As a result of the Littlewood subordination principle (see Theorem 1.7 of [4]), every analytic self-map φ\displaystyle\varphi of 𝔻\displaystyle\mathbb{D} induces a bounded composition operator on BMOA\displaystyle BMOA.

To compute the spectral radius of Cφ\displaystyle C_{\varphi} acting on BMOA\displaystyle BMOA, we first estimate the norm. By Corollary 2.2 of [9], there is a constant C>0\displaystyle C>0, independent of φ\displaystyle\varphi, such that

CφC(supa𝔻τφ(a)φτaH2+log21|φ(0)|2).\|C_{\varphi}\|\leq C\left(\sup_{a\in\mathbb{D}}\|\tau_{\varphi(a)}\circ\varphi\circ\tau_{a}\|_{H^{2}}+\log\frac{2}{1-|\varphi(0)|^{2}}\right). (4.1)

Since the function τφ(a)φτa\displaystyle\tau_{\varphi(a)}\circ\varphi\circ\tau_{a} is a composition of self-maps of the disk, the first term on the right is bounded above by 1. Also,

11|φ(0)|21+|φ(0)|1|φ(0)|21|φ(0)|\displaystyle\frac{1}{1-|\varphi(0)|^{2}}\leq\frac{1+|\varphi(0)|}{1-|\varphi(0)|}\leq\frac{2}{1-|\varphi(0)|}

and hence

log(21|φ(0)|2)log(41|φ(0)|)2log2log(1|φ(0)|).\displaystyle\log\left(\frac{2}{1-|\varphi(0)|^{2}}\right)\leq\log\left(\frac{4}{1-|\varphi(0)|}\right)\leq 2\log 2-\log\left(1-|\varphi(0)|\right).

Applying these estimates to Eq. (4.1), we have

CφC(1+2log2)Clog(1|φ(0)|).\displaystyle\|C_{\varphi}\|\leq C(1+2\log 2)-C\log(1-|\varphi(0)|).

This immediately implies that

CφnC(1+2log2)Clog(1|φn(0)|)\displaystyle\|C_{\varphi_{n}}\|\leq C(1+2\log 2)-C\log(1-|\varphi_{n}(0)|)

and it follows that ρ(Cφ)=1\displaystyle\rho(C_{\varphi})=1 for all bounded composition operators acting on BMOA\displaystyle BMOA by Theorem 3.7 of [3]. ∎

Thus BMOA\displaystyle BMOA satisfies all the properties of 𝒳\displaystyle\mathcal{X}.

4.2. Comparisons

We now investigate spaces that do not satisfy the properties of 𝒳\displaystyle\mathcal{X}. We compare the spectrum of induced composition operators on these spaces with those on 𝒳\displaystyle\mathcal{X}.

4.2.1. Hardy spaces

For 1p<\displaystyle 1\leq p<\infty, the Hardy space, denoted Hp=Hp(𝔻)\displaystyle H^{p}=H^{p}(\mathbb{D}), is the space of analytic functions on 𝔻\displaystyle\mathbb{D} such that

fHpp=sup0<r<1𝔻|f(reiθ)|pdθ2π<.\displaystyle\|f\|_{H^{p}}^{p}=\sup_{0<r<1}\,\int_{\mathbb{D}}|f(re^{i\theta})|^{p}\frac{d\theta}{2\pi}<\infty.

Under this norm, the Hardy spaces are Banach spaces and for p=2\displaystyle p=2 it is a Hilbert space.

It is well known that Hp\displaystyle H^{p} is a Banach space of analytic functions that contains H\displaystyle H^{\infty} as a subspace. For the cases of an elliptic or parabolic automorphism φ\displaystyle\varphi, it is the case that ρ(Cφ)=1\displaystyle\rho(C_{\varphi})=1 and the spectrum of Cφ\displaystyle C_{\varphi} on Hp\displaystyle H^{p} is the same as for Cφ\displaystyle C_{\varphi} acting on 𝒳\displaystyle\mathcal{X} (see Theorem 3.9 of [2]). However, it is not the case that the spectral radius is 1 for every composition operator induced by an automorphism. In fact, if φ\displaystyle\varphi is hyperbolic, then ρ(Cφ)=φ(a)1/p\displaystyle\rho(C_{\varphi})=\varphi^{\prime}(a)^{-1/p} where a\displaystyle a is the Denjoy-Wolff point of φ\displaystyle\varphi (see Theorem 3.9 of [2]). In this situation, φ(a)<1\displaystyle\varphi^{\prime}(a)<1 thus making ρ(Cφ)>1\displaystyle\rho(C_{\varphi})>1. In turn, the spectrum is the annulus φ(a)1/p|z|φ(a)1/p\displaystyle\varphi^{\prime}(a)^{1/p}\leq|z|\leq\varphi^{\prime}(a)^{-1/p} (see Theorem 4.9 of [8]).

4.2.2. Weighted Bergman spaces

For 1p<\displaystyle 1\leq p<\infty and α>1\displaystyle\alpha>-1, the standard weighted Bergman space, denoted Aαp=Aαp(𝔻)\displaystyle A_{\alpha}^{p}=A_{\alpha}^{p}(\mathbb{D}), is the space of analytic functions on 𝔻\displaystyle\mathbb{D} such that

fAαpp=𝔻(1|z|2)α|f(z)|p𝑑A(z)<,\displaystyle\|f\|_{A_{\alpha}^{p}}^{p}=\int_{\mathbb{D}}(1-|z|^{2})^{\alpha}|f(z)|^{p}\;dA(z)<\infty,

where dA(z)\displaystyle dA(z) is the normalized Lebesgue area measure on 𝔻\displaystyle\mathbb{D}. The weighted Bergman spaces are Banach spaces under the norm Aαp\displaystyle\|\cdot\|_{A_{\alpha}^{p}}.

It is well known that Aαp\displaystyle A_{\alpha}^{p} is a Banach space of analytic functions that contains H\displaystyle H^{\infty} as a subspace. For the cases of an elliptic or parabolic automorphism φ\displaystyle\varphi, it is the case that ρ(Cφ)=1\displaystyle\rho(C_{\varphi})=1 and the spectrum of Cφ\displaystyle C_{\varphi} on Aαp\displaystyle A_{\alpha}^{p} is the same as for Cφ\displaystyle C_{\varphi} acting on 𝒳\displaystyle\mathcal{X} (see Lemma 4.2 and Theorem 4.14 of [8]). However, as was the case for the Hardy spaces, it is not the case that the spectral radius is 1 for every composition operator induced by an automorphism. In fact, if φ\displaystyle\varphi is hyperbolic, then

ρ(Cφ)=max{1φ(a)s,1φ(b)s}\displaystyle\rho(C_{\varphi})=\max\left\{\frac{1}{\varphi^{\prime}(a)^{s}},\frac{1}{\varphi^{\prime}(b)^{s}}\right\}

where s=α+2p\displaystyle s=\frac{\alpha+2}{p}, a\displaystyle a is the Denjoy-Wolff point and b\displaystyle b is the other fixed point of φ\displaystyle\varphi (see Theorem 4.6 of [8]). In turn, the spectrum contains the annulus

min{1φ(a)s,1φ(b)s}|z|max{1φ(a)s,1φ(b)s}\displaystyle\min\left\{\frac{1}{\varphi^{\prime}(a)^{s}},\frac{1}{\varphi^{\prime}(b)^{s}}\right\}\leq|z|\leq\max\left\{\frac{1}{\varphi^{\prime}(a)^{s}},\frac{1}{\varphi^{\prime}(b)^{s}}\right\}

(see Corollary 4.7 of [8]).

4.2.3. Weighted Banach spaces

For 0<p<\displaystyle 0<p<\infty, the standard weighted Banach space on 𝔻\displaystyle\mathbb{D}, denoted Hp=Hp(𝔻)\displaystyle H^{\infty}_{p}=H^{\infty}_{p}(\mathbb{D}), is the space of analytic functions on 𝔻\displaystyle\mathbb{D} such that

fHp=supz𝔻(1|z|2)p|f(z)|<,\displaystyle\|f\|_{H^{\infty}_{p}}=\sup_{z\in\mathbb{D}}\,(1-|z|^{2})^{p}|f(z)|<\infty,

The weighted Banach spaces are, not surprising, Banach spaces under the norm Hp\displaystyle\|\cdot\|_{H^{\infty}_{p}}.

It is well known that Hp\displaystyle H^{\infty}_{p} is a Banach space of analytic functions that contain H\displaystyle H^{\infty} as a subspace. For the cases of an elliptic or parabolic automorphism φ\displaystyle\varphi, it is the case that ρ(Cφ)=1\displaystyle\rho(C_{\varphi})=1 and the spectrum of Cφ\displaystyle C_{\varphi} on Hp\displaystyle H^{\infty}_{p} is the same as for Cφ\displaystyle C_{\varphi} acting on 𝒳\displaystyle\mathcal{X} (see Lemma 4.2 and Theorem 4.14 of [8]). However, as was the case for the Hardy spaces, it is not the case that the spectral radius is 1 for every composition operator induced by an automorphism. In fact, if φ\displaystyle\varphi is hyperbolic, then

ρ(Cφ)=max{1φ(a)s,1φ(b)s}\displaystyle\rho(C_{\varphi})=\max\left\{\frac{1}{\varphi^{\prime}(a)^{s}},\frac{1}{\varphi^{\prime}(b)^{s}}\right\}

where a\displaystyle a is the Denjoy-Wolff point and b\displaystyle b is the other fixed point of φ\displaystyle\varphi (see Theorem 4.6 of [8]). In turn, the spectrum contains the annulus

min{1φ(a)p,1φ(b)p}|z|max{1φ(a)p,1φ(b)p}\displaystyle\min\left\{\frac{1}{\varphi^{\prime}(a)^{p}},\frac{1}{\varphi^{\prime}(b)^{p}}\right\}\leq|z|\leq\max\left\{\frac{1}{\varphi^{\prime}(a)^{p}},\frac{1}{\varphi^{\prime}(b)^{p}}\right\}

(see Corollary 4.7 of [8]).

4.2.4. Dirichlet space

The Dirichlet space on 𝔻\displaystyle\mathbb{D}, denoted 𝒟\displaystyle\mathcal{D}, is the space of analytic functions on 𝔻\displaystyle\mathbb{D} such that

𝔻|f(z)|2𝑑A(z)<\displaystyle\int_{\mathbb{D}}|f^{\prime}(z)|^{2}\;dA(z)<\infty

where dA\displaystyle dA denotes the normalized Lebesgue area measure on 𝔻\displaystyle\mathbb{D}. Under the norm

f𝒟2=|f(0)|2+𝔻|f(z)|2𝑑A(z)\displaystyle\|f\|_{\mathcal{D}}^{2}=|f(0)|^{2}+\int_{\mathbb{D}}|f^{\prime}(z)|^{2}\;dA(z)

the Dirichlet space has a Hilbert space structure. Although not every analytic self-map of 𝔻\displaystyle\mathbb{D} induce bounded composition operators on 𝒟\displaystyle\mathcal{D}, univalent maps, and thus the automorphisms, of 𝔻\displaystyle\mathbb{D} do.

Independently, Donaway (Corollary 3.11 of [3]) and Martín and Vukotić (Theorem 7 of [11]) showed that composition operators on 𝒟\displaystyle\mathcal{D} induced by univalent self-maps of 𝔻\displaystyle\mathbb{D}, and thus the automorphisms, have spectral radius 1. However, by direct calculation one can see that the functions in \displaystyle\mathcal{F} are not contained in the Dirichlet space; for the case of fs\displaystyle f_{s} this is shown in [14] (see pg. 455). Despite 𝒟\displaystyle\mathcal{D} not satisfying all the properties of 𝒳\displaystyle\mathcal{X}, the spectrum of automorphism induced composition operators on 𝒟\displaystyle\mathcal{D} are precisely the same as those on 𝒳\displaystyle\mathcal{X}.

To overcome the lack of eigenfunctions, the authors in [7] and [6] used two new approaches. In [7], the author produces approximate eigenfunctions and in [6] unitary similarity is the key tool.

Remark 4.2.

For all of the spaces discussed in Sections 4.1 and 4.2 (and those discussed in the next section), the spectrum of Cφ\displaystyle C_{\varphi} when φ\displaystyle\varphi is elliptic will be the same as that for Cφ\displaystyle C_{\varphi} acting on 𝒳\displaystyle\mathcal{X}. This is due to the fact that the eigenfunctions are the monomials, which are contained in all of these spaces.

5. Open Questions

We end this paper with open questions which were inspired while developing the examples and comparisons in Sections 4.1 and 4.2.

5.1. The little Bloch space

While the Bloch space contains the polynomials, they are not dense in \displaystyle\mathcal{B}. The closure of the polynomials in \displaystyle\|\cdot\|_{\mathcal{B}} is called the little Bloch space, denoted 0=0(𝔻)\displaystyle\mathcal{B}_{0}=\mathcal{B}_{0}(\mathbb{D}). More formally, the little Bloch space consists of the functions f\displaystyle f\in\mathcal{B} such that

lim|z|1(1|z|2)|f(z)|=0.\displaystyle\lim_{|z|\to 1}(1-|z|^{2})|f^{\prime}(z)|=0.

From Theorem 12 of [1], bounded composition operators on 0\displaystyle\mathcal{B}_{0} are induced exactly by functions in 0\displaystyle\mathcal{B}_{0}, which include the automorphisms. Donaway also proved the spectral radius of every bounded composition operator on 0\displaystyle\mathcal{B}_{0} is 1. Thus property (ii) is satisfied by 0\displaystyle\mathcal{B}_{0}. However, the following result shows that \displaystyle\mathcal{F} is not contained in 0\displaystyle\mathcal{B}_{0}, and thus property (i) of 𝒳\displaystyle\mathcal{X} is not satisfied.

Theorem 5.1.

The functions fs\displaystyle f_{s} and ft\displaystyle f_{t}, for s>0\displaystyle s>0 and t0\displaystyle t\neq 0, are not contained in the little Bloch space.

Proof.

Consider the function

ft(z)=exp(itlog1+z1z).\displaystyle f_{t}(z)=\exp\left(it\log\frac{1+z}{1-z}\right).

We show that this function is not in 0\displaystyle\mathcal{B}_{0} for t{0}.\displaystyle t\in\mathbb{R}\setminus\{0\}. Taking the derivative,

ft(z)=ft(z)(it1z1+z)2(1z)2\displaystyle f^{\prime}_{t}(z)=f_{t}(z)\left(it\frac{1-z}{1+z}\right)\frac{2}{(1-z)^{2}} =ft(z)2it(1z)(1+z).\displaystyle=f_{t}(z)\frac{2it}{(1-z)(1+z)}.

For t>0\displaystyle t>0, |ft(z)|e2π\displaystyle|f_{t}(z)|\geq e^{-2\pi} and, for t<0\displaystyle t<0, |ft(z)|1\displaystyle|f_{t}(z)|\geq 1. In either case, there is a constant C>0\displaystyle C>0 such that |ft(z)|C\displaystyle|f_{t}(z)|\geq C for all t{0}\displaystyle t\in\mathbb{R}\setminus\{0\} and all z𝔻\displaystyle z\in\mathbb{D}. Hence

|ft(z)|2C|t||z1||z+1|.\displaystyle|f^{\prime}_{t}(z)|\geq\frac{2C|t|}{|z-1||z+1|}.

To show that ft0\displaystyle f_{t}\not\in\mathcal{B}_{0}, we need to show that

lim|z|1(1|z|2)|ft(z)|0.\displaystyle\lim_{|z|\rightarrow 1}(1-|z|^{2})|f^{\prime}_{t}(z)|\neq 0.

To see this, first observe that

lim|z|1(1|z|2)|ft(z)|lim|z|1(1|z|2)2C|t||z1||z+1|\displaystyle\lim_{|z|\rightarrow 1}(1-|z|^{2})|f^{\prime}_{t}(z)|\geq\lim_{|z|\rightarrow 1}(1-|z|^{2})\frac{2C|t|}{|z-1||z+1|}

by our estimate from above. If we now take a radial path to 1, that is, we set z=r\displaystyle z=r and let r1\displaystyle r\uparrow 1, we have

limr1(1r2)2C|t|(1r)(1+r)=2C|t|>0\displaystyle\lim_{r\rightarrow 1^{-}}(1-r^{2})\frac{2C|t|}{(1-r)(1+r)}=2C|t|>0

when t0\displaystyle t\neq 0. Thus

lim|z|1(1|z|2)2Ct|z1||z+1|0\displaystyle\lim_{|z|\rightarrow 1}(1-|z|^{2})\frac{2Ct}{|z-1||z+1|}\neq 0

for t0\displaystyle t\neq 0, and hence ft\displaystyle f_{t} is not in 0\displaystyle\mathcal{B}_{0}.

Next consider the function

fs(z)=exp(s(z+1)z1).\displaystyle f_{s}(z)=\exp\left(\frac{s(z+1)}{z-1}\right).

We will show that this function is not in 0\displaystyle\mathcal{B}_{0} for s>0\displaystyle s>0. First observe that

fs(z)=exp(s(z+1)z1)(2s(z1)2)\displaystyle f^{\prime}_{s}(z)=\exp\left(\frac{s(z+1)}{z-1}\right)\left(\frac{-2s}{(z-1)^{2}}\right)

and thus we aim to show that

lim|z|1(1|z|2)|fs(z)|=lim|z|1(1|z|2)|exp(s(z+1)z1)|2s|1z|20.\displaystyle\lim_{|z|\rightarrow 1}(1-|z|^{2})|f^{\prime}_{s}(z)|=\lim_{|z|\rightarrow 1}(1-|z|^{2})\left|\exp\left(\frac{s(z+1)}{z-1}\right)\right|\frac{2s}{|1-z|^{2}}\neq 0.

Fix x0<0\displaystyle x_{0}<0 and consider the sequence {zn}\displaystyle\{z_{n}\} defined by

zn=x0+in+1x0+in1.\displaystyle z_{n}=\frac{x_{0}+in+1}{x_{0}+in-1}.

Since x0<0\displaystyle x_{0}<0, this sequence is contained in the unit disk and {zn}1\displaystyle\{z_{n}\}\rightarrow 1 as n.\displaystyle n\rightarrow\infty. To obtain our conclusion, we show

limn(1|zn|2)|exp(s(zn+1)zn1)|2s|1zn|20.\displaystyle\lim_{n\rightarrow\infty}(1-|z_{n}|^{2})\left|\exp\left(\frac{s(z_{n}+1)}{z_{n}-1}\right)\right|\frac{2s}{|1-z_{n}|^{2}}\neq 0.

First observe that the map ψ(z)=(z+1)/(z1)\displaystyle\psi(z)=(z+1)/(z-1) is its own inverse and hence ψ(zn)=x0+in\displaystyle\psi(z_{n})=x_{0}+in for each n\displaystyle n\in\mathbb{N}. Thus

|exp(s(zn+1)zn1)|=|exp(sx0+isn)|=esx0>0.\displaystyle\left|\exp\left(\frac{s(z_{n}+1)}{z_{n}-1}\right)\right|=|\exp(sx_{0}+isn)|=e^{sx_{0}}>0.

Substituting,

limn(1|zn|2)|exp(s(zn+1)zn1)|2s|1zn|2=limnesx0(1|zn|2)2s|1zn|2.\displaystyle\lim_{n\rightarrow\infty}(1-|z_{n}|^{2})\left|\exp\left(\frac{s(z_{n}+1)}{z_{n}-1}\right)\right|\frac{2s}{|1-z_{n}|^{2}}=\lim_{n\rightarrow\infty}e^{sx_{0}}(1-|z_{n}|^{2})\frac{2s}{|1-z_{n}|^{2}}.

Next,

1|zn|2=4x0(x01)2+n2\displaystyle 1-|z_{n}|^{2}=\frac{-4x_{0}}{(x_{0}-1)^{2}+n^{2}}

and

|1zn|2=4(x01)2+n2.\displaystyle|1-z_{n}|^{2}=\frac{4}{(x_{0}-1)^{2}+n^{2}}.

Thus

limnesx0(1|zn|2)2s|1zn|2\displaystyle\lim_{n\rightarrow\infty}e^{sx_{0}}(1-|z_{n}|^{2})\frac{2s}{|1-z_{n}|^{2}} =limnesx0(4x0(x01)2+n2)(s((x01)2+n2)2)\displaystyle=\lim_{n\rightarrow\infty}e^{sx_{0}}\left(\frac{-4x_{0}}{(x_{0}-1)^{2}+n^{2}}\right)\left(\frac{s((x_{0}-1)^{2}+n^{2})}{2}\right)
=limn(2sx0)esx0>0\displaystyle=\lim_{n\rightarrow\infty}(-2sx_{0})e^{sx_{0}}>0

and hence fs\displaystyle f_{s} is not in 0\displaystyle\mathcal{B}_{0} for s>0\displaystyle s>0. ∎

For the little Bloch space, we leave the reader with the following question.

Question 1.

For φ\displaystyle\varphi a parabolic or hyperbolic automorphism, what is the spectrum of Cφ\displaystyle C_{\varphi} on the little Bloch space?

5.2. Analytic functions of vanishing mean oscillation

Like the Bloch space, BMOA\displaystyle BMOA contains the polynomials, but they are not dense in BMOA\displaystyle BMOA. The closure of the polynomials in BMOA\displaystyle\|\cdot\|_{\mathrm{BMOA}} is denoted by VMOA\displaystyle VMOA. VMOA\displaystyle VMOA is the space of analytic functions with vanishing mean osciallation on 𝔻\displaystyle\partial\mathbb{D}, formally defined as the functions fBMOA\displaystyle f\in\mathrm{BMOA} such that

lim|a|1fτaf(a)H2=0.\displaystyle\lim_{|a|\to 1}\,\|f\circ\tau_{a}-f(a)\|_{H^{2}}=0.

By Corollary 4.2 of [9], Cφ\displaystyle C_{\varphi} is bounded on VMOA\displaystyle VMOA if and only if φVMOA\displaystyle\varphi\in VMOA. So every automorphism induces a bounded composition operator on VMOA\displaystyle VMOA. By the same argument as in Section 4.1.3, the spectral radius of Cφ\displaystyle C_{\varphi} induced by a disk automorphism is 1. Thus property (ii) of 𝒳\displaystyle\mathcal{X} is satisfied. Since VMOA\displaystyle VMOA is a subspace of the little Bloch space (see [5]), it follows that VMOA\displaystyle VMOA does not satisfy property (i), a corollary of Theorem 5.1.

Corollary 5.2.

The functions fs\displaystyle f_{s} and ft\displaystyle f_{t}, for s>0\displaystyle s>0 and t0\displaystyle t\neq 0, are not contained in VMOA\displaystyle VMOA.

For VMOA\displaystyle VMOA, we leave the reader with the following question.

Question 2.

For φ\displaystyle\varphi parabolic or hyperbolic automorphism, what is the spectrum of Cφ\displaystyle C_{\varphi} on VMOA\displaystyle VMOA?

Acknowledgements

The work of the second author was conducted while an undergraduate student at the University of Wisconsin-La Crosse and funded by the College of Science and Health Dean’s Distinguished Fellowship.

References

  • [1] J. Arazy, S. D. Fisher, and J. Peetre, Möbius invariant function spaces, J. Reine Angew. Math. 363 (1985), 110–145. MR 814017 (87f:30104)
  • [2] C. C. Cowen and B. D. MacCluer, Composition operators on spaces of analytic functions, Studies in Advanced Mathematics, CRC Press, Boca Raton, FL, 1995. MR 1397026 (97i:47056)
  • [3] R. L. Donaway, Norm and essential norm estimates of composition operators on Besov-type spaces, Ph.D. thesis, University of Virginia, 1999.
  • [4] P. L. Duren, Theory of Hp\displaystyle H^{p} spaces, Pure and Applied Mathematics, Vol. 38, Academic Press, New York-London, 1970. MR 0268655 (42 #3552)
  • [5] E. A. Gallardo-Gutiérrez, M. J. González, F. Pérez-González, Ch. Pommerenke, and J. Rättyä, Locally univalent functions, VMOA and the Dirichlet space, Proc. Lond. Math. Soc. (3) 106 (2013), no. 3, 565–588. MR 3048550
  • [6] E. A. Gallardo-Gutiérrez and A. Montes-Rodríguez, Adjoints of linear fractional composition operators on the Dirichlet space, Math. Ann. 327 (2003), no. 1, 117–134. MR 2005124 (2004h:47036)
  • [7] W. M. Higdon, Composition operators of the Dirichlet space, Ph.D. thesis, Michigan State University, 1997.
  • [8] O. Hyvärinen, M. Lindström, I. Nieminen, and E. Saukko, Spectra of weighted composition operators with automorphic symbols, J. Funct. Anal. 265 (2013), no. 8, 1749–1777. MR 3079234
  • [9] J. Laitila, Weighted composition operators on BMOA, Comput. Methods Funct. Theory 9 (2009), no. 1, 27–46. MR 2478261 (2010b:47059)
  • [10] B. D. MacCluer, Elementary functional analysis, Graduate Texts in Mathematics, vol. 253, Springer, New York, 2009. MR 2462971 (2010b:46001)
  • [11] M. J. Martín and D. Vukotić, Norms and spectral radii of composition operators acting on the Dirichlet space, J. Math. Anal. Appl. 304 (2005), no. 1, 22–32. MR 2124646 (2006j:47043)
  • [12] E. A. Nordgren, Composition operators, Canad. J. Math. 20 (1968), 442–449. MR 0223914 (36 #6961)
  • [13] M. A. Pons, Composition operators on Besov and Dirichlet type spaces, Ph.D. thesis, University of Virginia, 2007.
  • [14] by same author, 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 (2012d:47074)