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

On weighted compositions preserving the Carathéodory class

Irina Arévalo Departamento de Matemáticas, Universidad Autónoma de Madrid, 28049 Madrid, Spain irina.arevalo@uam.es Rodrigo Hernández Universidad Adolfo Ibáñez, Facultad de Ingeniería y Ciencias, Av. Padre Hurtado 750, Viña del Mar, Chile rodrigo.hernandez@uai.cl María J. Martín University of Eastern Finland, Department of Physics and Mathematics, P.O. Box 111, 80101 Joensuu, Finland maria.martin@uef.fi  and  Dragan Vukotić Departamento de Matemáticas, Universidad Autónoma de Madrid, 28049 Madrid, Spain dragan.vukotic@uam.es
(Date: 17 February, 2017.)
Abstract.

We characterize in various ways the weighted composition transformations which preserve the class 𝒫{\mathcal{P}} of normalized analytic functions in the disk with positive real part. We analyze the meaning of the criteria obtained for various special cases of symbols and identify the fixed points of such transformations.

2010 Mathematics Subject Classification:
30C45, 47B33
Arévalo, Martín, and Vukotić are supported by MTM2015-65792-P from MINECO and FEDER/EU and partially by the Thematic Research Network MTM2015-69323-REDT, MINECO, Spain. Hernández and Martín are supported by FONDECYT 1150284, Chile. Martín is also supported by Academy of Finland Grant 268009.

1. Introduction

1.1. The class 𝒫{\mathcal{P}} and its properties

Let 𝔻{\mathbb{D}} denote the unit disk in the complex plane and (𝔻){\mathcal{H}}({\mathbb{D}}) the algebra of all functions analytic in 𝔻{\mathbb{D}}. If ϕ(𝔻)\phi\in{\mathcal{H}}({\mathbb{D}}) and ϕ(𝔻)𝔻\phi({\mathbb{D}})\subset{\mathbb{D}}, we will say that ϕ\phi is an analytic self-map of the disk. If, moreover, such ϕ\phi satisfies ϕ(0)=0\phi(0)=0, we will refer to it as a Schwarz-type function (as in the classical Schwarz lemma).

Denote by 𝒫{\mathcal{P}} the Carathéodory class of all ff in (𝔻){\mathcal{H}}({\mathbb{D}}) with positive real part and normalized so that f(0)=1f(0)=1. An important example of a function in this class is the so-called half-plane mapping \ell given by

(z)=1+z1z,z𝔻.\ell(z)=\frac{1+z}{1-z}\,,\qquad z\in{\mathbb{D}}\,.

This conformal map of the disk onto the right half-plane is extremal in many senses for the class 𝒫{\mathcal{P}}. This is manifested, for example, by the growth theorem for the functions in the class:

(|z|)=1|z|1+|z||f(z)|1+|z|1|z|=(|z|).\ell(-|z|)=\frac{1-|z|}{1+|z|}\leq|f(z)|\leq\frac{1+|z|}{1-|z|}=\ell(|z|)\,.

The above estimate [11, Section 2.1] is a direct corollary of the Schwarz lemma and the elementary subordination principle since every function ff in 𝒫{\mathcal{P}} is of the form ω\ell\circ\omega, where ω\omega is some Schwarz-type function. In the particular case when ω(z)=λz\omega(z)=\lambda z with |λ|=1|\lambda|=1, we will use the symbol λ\ell_{\lambda} to denote the functions ω\ell\circ\omega; that is, λ(z)=(1+λz)/(1λz)\ell_{\lambda}(z)=(1+\lambda z)/(1-\lambda z). In view of the Herglotz representation theorem [8, Chapter 1], the class 𝒫{\mathcal{P}} equals co ()¯\overline{\textrm{co\,}({\mathcal{L}})}, the closed convex hull of the collection ={λ:|λ|=1}{\mathcal{L}}=\{\ell_{\lambda}\,\colon\,|\lambda|=1\} in the topology of uniform convergence on compact subsets of 𝔻{\mathbb{D}} (the compact-open topology). Every Schwarz-type function has radial limits almost everywhere on the unit circle 𝕋{\mathbb{T}} with respect to the normalized arc length measure dm=dθ/(2π)dm=d\theta/(2\pi) (see [7, Chapter 1]), hence so does every ff in 𝒫{\mathcal{P}}.

1.2. On weighted composition transformations

Whenever ϕ\phi is an analytic self-map of the disk, it is immediate that fϕ𝒫f\circ\phi\in{\mathcal{P}} for every f𝒫f\in{\mathcal{P}} if and only if ϕ\phi is a Schwarz-type function. Thus, it makes sense to consider the composition transformation CϕC_{\phi} on 𝒫{\mathcal{P}} defined by the formula Cϕ(f)=fϕC_{\phi}(f)=f\circ\phi. For the theory of composition operators on Banach or Hilbert spaces of analytic functions, we refer the reader to [5] or [13].

It also seems reasonable to consider the multiplication transformations on 𝒫{\mathcal{P}}, given by MF(f)=FfM_{F}(f)=Ff, where F(𝔻)F\in{\mathcal{H}}({\mathbb{D}}). While there are many cases of such transformations in spaces of analytic functions, it turns out that such mapping can preserve the class 𝒫{\mathcal{P}} only in the trivial case when F1F\equiv 1.

One can consider the more general weighted composition transformations TF,ϕT_{F,\phi} on 𝒫{\mathcal{P}}, defined by a composition followed by a multiplication: TF,ϕ(f)=F(fϕ)T_{F,\phi}(f)=F(f\circ\phi), where FF, ϕ(𝔻)\phi\in{\mathcal{H}}({\mathbb{D}}) and ϕ(𝔻)𝔻\phi({\mathbb{D}})\subset{\mathbb{D}}, and ask whether such a transformation acts on 𝒫{\mathcal{P}} or preserves 𝒫{\mathcal{P}} in the sense that TF,ϕ(f)𝒫T_{F,\phi}(f)\in{\mathcal{P}} for every ff in 𝒫{\mathcal{P}}. This will make sense more often since one of the two symbols FF and ϕ\phi can compensate for the behavior of the other. Also, the nonlinearity of such transformations in our context makes the analysis non-trivial.

The linear weighted composition operators have been studied in a number of papers in the context of Banach spaces of analytic functions. The earliest references on this type of operators are [9] and [10]; among the numerous more recent references, we mention [1], [4], or [6]. Although the class 𝒫{\mathcal{P}} has no linear space structure, studying the question of when a weighted composition transformation preserves this class should be of interest as it would indicate the degree of rigidity when trying to produce new functions with positive real part from the given ones by composing and multiplying. Moreover, the knowledge of weighted composition transformations that preserve the class 𝒫{\mathcal{P}} could even have some applications to variational methods for solving non-linear extremal problems in geometric function theory.

1.3. Summary of main results

The question considered in the present paper is to characterize the ordered pairs of functions (F,ϕ)(F,\phi) for which F(fϕ)𝒫F(f\circ\phi)\in{\mathcal{P}} whenever f𝒫f\in{\mathcal{P}}. We solve this problem completely by exhibiting explicit conditions of geometric and analytic character which are equivalent to this property and are relatively easy to check in practice. The precise statement is formulated as Theorem 2.

Although this result is conclusive, the work does not end here by any means. In some situations the theorem allows us to deduce the following rigidity principle: if ϕ\phi is of certain type, then the only FF possible in the trivial case: F1F\equiv 1; see Proposition 3. Alternatively, when FF is of special type then ϕ\phi must be unique; cf. Proposition 4.

Another issue of interest is to construct non-trivial examples and to interpret the statement of Theorem 2 for some specific types of maps FF and ϕ\phi in terms of their qualitative behavior (image compactly contained in 𝔻{\mathbb{D}}, radial limits, angular derivatives, etc.). Roughly speaking, one could ask how much we can push the conditions for these maps to the limit and when one of the symbols is “good” in some sense, how “bad” can the other one be. Thus, various statements such as propositions 6 and 7 and Theorem 10 seem to require a more intricate analysis and may be as interesting as the characterization itself. Different relevant examples are also included.

We end the paper by showing that every transformation TF,ϕT_{F,\phi} that preserves class 𝒫{\mathcal{P}} has a unique fixed point whenever ϕ\phi is not a rotation and this fixed point is obtained by applying the iterates of TF,ϕT_{F,\phi} to an arbitrary function in 𝒫{\mathcal{P}}, as one would expect. This is the content of Theorem 11. In the case when ϕ\phi is a rotation, one can easily describe all fixed points. Even though the transformation is non-linear, here one can adapt the methods from linear functional analysis used in the study of composition operators.

2. Characterizations of the 𝒫{\mathcal{P}}-preserving weighted compositions

2.1. Initial observations

We begin by recording the obvious necessary conditions that must be satisfied by all admissible symbols FF and ϕ\phi, i.e., by those which satisfy the initial assumptions FF, ϕ(𝔻)\phi\in{\mathcal{H}}({\mathbb{D}}), ϕ(𝔻)𝔻\phi({\mathbb{D}})\subset{\mathbb{D}} and make the inclusion TF,ϕ(𝒫)𝒫T_{F,\phi}({\mathcal{P}})\subset{\mathcal{P}} possible.

\bullet If F(fϕ)𝒫F(f\circ\phi)\in{\mathcal{P}} for all ff in 𝒫{\mathcal{P}}, it is clear that then F𝒫F\in{\mathcal{P}}; this is immediately seen by choosing f1f\equiv 1, which is a function in 𝒫{\mathcal{P}}.

\bullet In addition to this, ϕ\phi must be a Schwarz-type function since, after choosing f(z)=1+zf(z)=1+z, another function obviously in 𝒫{\mathcal{P}}, we get

1=F(0)f(ϕ(0))=1+ϕ(0),1=F(0)f(\phi(0))=1+\phi(0)\,,

hence ϕ(0)=0\phi(0)=0.

Thus, from now on we shall always work assuming these hypotheses: F𝒫F\in{\mathcal{P}} and ϕ\phi is a Schwarz-type function.

It is quite easy to establish the lack of non-trivial pointwise multipliers of 𝒫{\mathcal{P}}.

Proposition 1.

If TF,ϕT_{F,\phi} preserves 𝒫{\mathcal{P}} and ϕ(z)=z\phi(z)=z (that is, Ff𝒫F\,f\in{\mathcal{P}} for all ff in 𝒫{\mathcal{P}}) then F1F\equiv 1.

Proof.

Indeed, since then Ff𝒫Ff\in{\mathcal{P}}, by the growth theorem for the functions in 𝒫{\mathcal{P}} we have

|F(z)||f(z)|1+|z|1|z||F(z)|\cdot|f(z)|\leq\frac{1+|z|}{1-|z|}

for all zz in 𝔻{\mathbb{D}}. Also, for any fixed zz we may choose ff to be a suitable rotation of the half-plane function for which

|f(z)|=1+|z|1|z|.|f(z)|=\frac{1+|z|}{1-|z|}\,.

It follows that |F(z)|1|F(z)|\leq 1 for all z𝔻z\in{\mathbb{D}}. Since F(0)=1F(0)=1, the maximum modulus principle implies that FF is identically constant, hence F1F\equiv 1. ∎

2.2. The main theorem

We now characterize in different ways all admissible pairs of symbols. Note that condition (b) below simply states that it suffices to test the action of TF,ϕT_{F,\phi} on the set ={λ:|λ|=1}{\mathcal{L}}=\{\ell_{\lambda}\,\colon\,|\lambda|=1\} of all rotations of the half-plane function in order to know whether the transformation preserves 𝒫{\mathcal{P}}. Condition (c) gives an effective analytic way of testing if a symbol is admissible or not while (d) provides conditions of geometric type. Each can be useful in its own way.

Throughout the paper, we shall consider the principal branch of the argument function with values in (π,π](-\pi,\pi]. Note that for any function ff in 𝒫{\mathcal{P}}, the function argf\arg f takes on the values only in (π/2,π/2)(-\pi/2,\pi/2) and is a continuous function in the disk. Moreover, the argument of the product of two such functions ff and gg in 𝒫{\mathcal{P}}, with values in (π,π)(-\pi,\pi), is still continuous and the formula

arg(fg)=argf+argg\mathrm{arg\,}(fg)=\mathrm{arg\,}f+\mathrm{arg\,}g

holds throughout 𝔻{\mathbb{D}}. We will use this fact repeatedly.

Theorem 2.

Let ϕ\phi be a Schwarz-type function, F𝒫F\in{\mathcal{P}}, and denote by ω\,\omega the Schwarz-type function for which F=ωF=\ell\circ\omega. Consider the argument function defined as above. Then the following conditions are equivalent:

  • (a)

    TF,ϕ(𝒫)𝒫T_{F,\phi}({\mathcal{P}})\subset{\mathcal{P}}.

  • (b)

    TF,ϕ()𝒫T_{F,\phi}({\mathcal{L}})\subset{\mathcal{P}}.

  • (c)

    The inequality

    (1) 4|ϕ(z)||Imω(z)|<(1|ω(z)|2)(1|ϕ(z)|2)4|\phi(z)|\cdot|\mathrm{Im\,}\omega(z)|<(1-|\omega(z)|^{2})(1-|\phi(z)|^{2})

    holds for all zz in 𝔻{\mathbb{D}}. In other words,

    (2) 2|ϕ(z)||ImF(z)ReF(z)|<1|ϕ(z)|2,forallz𝔻.2|\phi(z)|\cdot\left|\frac{\mathrm{Im\,}F(z)}{\mathrm{Re\,}F(z)}\right|<1-|\phi(z)|^{2}\,,\quad\mathrm{for\ all\ }z\in{\mathbb{D}}\,.
  • (d)

    The inequality

    (3) |argF(z)|<π2arcsin2|ϕ(z)|1+|ϕ(z)|2|\arg F(z)|<\frac{\pi}{2}-\arcsin\frac{2|\phi(z)|}{1+|\phi(z)|^{2}}

    holds for all zz in 𝔻{\mathbb{D}}. Note also that

    (4) π2arcsin2|ϕ(z)|1+|ϕ(z)|2=π2arctan2|ϕ(z)|1|ϕ(z)|2=arctan1|ϕ(z)|22|ϕ(z)|,\frac{\pi}{2}-\arcsin\frac{2|\phi(z)|}{1+|\phi(z)|^{2}}=\frac{\pi}{2}-\arctan\frac{2|\phi(z)|}{1-|\phi(z)|^{2}}=\arctan\frac{1-|\phi(z)|^{2}}{2|\phi(z)|}\,,

    where in the case when z=0z=0 (recalling that ϕ(0)=0\phi(0)=0) the last equality should be understood as the limit: arctan(+)=π2\arctan(+\infty)=\frac{\pi}{2}.

It should be noted that in the above result the inequalities in conditions (c) and (d) are both invariant under rotations of ϕ\phi but not under the rotations in ω\omega (or under the appropriate changes in FF).

Proof.

We will show that (a) \Leftrightarrow (b), (b) \Leftrightarrow (c), and (c) \Leftrightarrow (d).

(a) \Leftrightarrow (b) . The implication (a) \Rightarrow (b) is obvious so we only have to see that (b) \Rightarrow (a). First of all, the image of a convex combination of functions in 𝒫{\mathcal{P}} under TF,ϕT_{F,\phi} is the same convex combination of their images and a convex combination of function in 𝒫{\mathcal{P}} remains in 𝒫{\mathcal{P}}. Hence, if TF,ϕ()𝒫T_{F,\phi}({\mathcal{L}})\subset{\mathcal{P}} we also have that TF,ϕ(co ())=co TF,ϕ()𝒫T_{F,\phi}(\textrm{co\,}({\mathcal{L}}))=\textrm{co\,}T_{F,\phi}({\mathcal{L}})\subset{\mathcal{P}} as TF,ϕ()𝒫T_{F,\phi}({\mathcal{L}})\subset{\mathcal{P}} by assumption and the class 𝒫{\mathcal{P}} is clearly convex.

Next, if fnff_{n}\to f uniformly on compact subsets of 𝔻{\mathbb{D}}, then also F(fnϕ)F(fϕ)F(f_{n}\circ\phi)\to F(f\circ\phi) in the same topology. Since 𝒫{\mathcal{P}} is a compact family (in the classical terminology, meaning a closed set in the compact-open topology), we get TF,ϕ(𝒫)=TF,ϕ(co ()¯)=TF,ϕ(co ())¯𝒫T_{F,\phi}({\mathcal{P}})=T_{F,\phi}(\overline{\textrm{co\,}({\mathcal{L}})})=\overline{T_{F,\phi}(\textrm{co\,}({\mathcal{L}}))}\subset{\mathcal{P}}.

(b) \Leftrightarrow (c) . To verify that (1) is equivalent to (2), one easily checks that if F=ωF=\ell\circ\omega then

|ImF(z)ReF(z)|=2|Imω(z)|1|ω(z)|2.\left|\frac{\mathrm{Im\,}F(z)}{\mathrm{Re\,}F(z)}\right|=2\frac{|\mathrm{Im\,}\omega(z)|}{1-|\omega(z)|^{2}}\,.

To see that (b) \Rightarrow (c), suppose that F(fϕ)𝒫F(f\circ\phi)\in{\mathcal{P}} for all ff in {\mathcal{L}}. In other words, F(λϕ)𝒫F(\ell_{\lambda}\circ\phi)\in{\mathcal{P}} for all λ\lambda of modulus one and therefore also

F(λϕ)=1+ωλ1ωλF(\ell_{\lambda}\circ\phi)=\frac{1+\omega_{\lambda}}{1-\omega_{\lambda}}

for the Schwarz-type functions ωλ\omega_{\lambda} depending on each λ\lambda. This leads to the equation

1+ω1ω1+λϕ1λϕ=1+ωλ1ωλ\frac{1+\omega}{1-\omega}\frac{1+\lambda\phi}{1-\lambda\phi}=\frac{1+\omega_{\lambda}}{1-\omega_{\lambda}}

which holds in the entire unit disk. Solving for ωλ\omega_{\lambda}, we get

ωλ=λϕ+ω1+λϕω.\omega_{\lambda}=\frac{\lambda\phi+\omega}{1+\lambda\phi\omega}\,.

The condition |ωλ|<1|\omega_{\lambda}|<1 in 𝔻{\mathbb{D}} is equivalent to

|λϕ+ω|2<|1+λϕω|2|\lambda\phi+\omega|^{2}<|1+\lambda\phi\omega|^{2}

which amounts to the inequality

(5) |ϕ|2+|ω|2+2Re{λϕω¯}<1+|ϕω|2+2Re{λϕω}.|\phi|^{2}+|\omega|^{2}+2\mathrm{Re\,}\{\lambda\phi\overline{\omega}\}<1+|\phi\omega|^{2}+2\mathrm{Re\,}\{\lambda\phi\omega\}\,.

Grouping the terms in (5) we obtain

2Re{λϕ(z)(ω(z)¯ω(z))}<(1|ω(z)|2)(1|ϕ(z)|2)2\mathrm{Re\,}\{\lambda\phi(z)(\overline{\omega(z)}-\omega(z))\}<(1-|\omega(z)|^{2})(1-|\phi(z)|^{2})

for each zz in 𝔻{\mathbb{D}} and for arbitrary λ\lambda with |λ|=1|\lambda|=1. For each point zz we can choose the argument of λ\lambda appropriately so as to get

2Re{λϕ(z)(ω(z)¯ω(z))}=4|ϕ(z)||Imω(z)|.2\mathrm{Re\,}\{\lambda\phi(z)(\overline{\omega(z)}-\omega(z))\}=4|\phi(z)|\cdot|\mathrm{Im\,}\omega(z)|\,.

Since this is valid at every point zz in the disk, the statement (1) follows.

To see that (c) \Rightarrow (b), it suffices to observe that

2Re{λϕ(z)(ω(z)¯ω(z))}4|ϕ(z)||Imω(z)|2\mathrm{Re\,}\{\lambda\phi(z)(\overline{\omega(z)}-\omega(z))\}\leq 4|\phi(z)|\cdot|\mathrm{Im\,}\omega(z)|

and it is now easy to reverse the steps in the above proof.

(c) \Leftrightarrow (d) . Since F𝒫F\in{\mathcal{P}}, we know that |argF|<π/2|\arg F|<\pi/2. Thus, inequality (2) is clearly equivalent to

2|ϕ(z)||tan(argF(z))|<(1|ϕ(z)|2),z𝔻,2|\phi(z)|\cdot|\tan(\arg F(z))|<(1-|\phi(z)|^{2})\,,\qquad z\in{\mathbb{D}}\,,

which is the same as

|tan(argF(z))|<1|ϕ(z)|22|ϕ(z)|,z𝔻,|\tan(\arg F(z))|<\frac{1-|\phi(z)|^{2}}{2|\phi(z)|}\,,\qquad z\in{\mathbb{D}}\,,

understanding the right-hand side as ++\infty when ϕ(z)=0\phi(z)=0. The inverse tangent function is odd so this is the same as

|argF(z)|<arctan1|ϕ(z)|22|ϕ(z)|,z𝔻.|\arg F(z)|<\arctan\frac{1-|\phi(z)|^{2}}{2|\phi(z)|}\,,\qquad z\in{\mathbb{D}}\,.

Equalities (4) follow by elementary trigonometry, so the proof is complete. ∎

3. Some consequences and discussions

It should be stressed out that, even though our Theorem 2 gives different characterizations of all admissible pairs of symbols, in some special situations the information given by the theorem can be made more precise. Actually, in some situations it may not be obvious how many examples of admissible pairs can exist. The aim of this section is to explain what our main results amounts to in some important special situations.

3.1. Some rigidity principles

As is usual, for a bounded analytic function ϕ\phi in 𝔻{\mathbb{D}} we write ϕ=supz𝔻|ϕ(z)|=esssupζ𝕋|ϕ(ζ)|\|\phi\|_{\infty}=\sup_{z\in{\mathbb{D}}}|\phi(z)|=\operatorname{ess\,sup}_{\zeta\in{\mathbb{T}}}|\phi(\zeta)|.

Recall that bounded analytic functions in 𝔻{\mathbb{D}} have radial limits ϕ(ζ)=limr1ϕ(rζ)\phi(\zeta)=\lim_{r\to 1}\phi(r\zeta) for almost every point ζ\zeta on the unit circle 𝕋{\mathbb{T}} with respect to the normalized Lebesgue arc length measure [7, Chapter 1]. An analytic function ϕ\phi in the disk is called inner if |ϕ(z)|1|\phi(z)|\leq 1 for all zz in 𝔻{\mathbb{D}} (equivalently, ϕ1\|\phi\|_{\infty}\leq 1) and also |ϕ(ζ)|=1|\phi(\zeta)|=1 almost everywhere on 𝕋{\mathbb{T}}. The following result generalizes our Proposition 1.

Proposition 3.

Let F𝒫F\in{\mathcal{P}} and let ϕ\phi be inner. Then TF,ϕ(𝒫)𝒫T_{F,\phi}({\mathcal{P}})\subset{\mathcal{P}} if and only if  F1F\equiv 1.

Proof.

The bounded functions ϕ\phi and ω\omega have radial limits almost everywhere on the circle. Thus, for almost every ζ𝕋\zeta\in{\mathbb{T}} we may pass to the limit as zζz\to\zeta in inequality (1) to conclude that Imω(ζ)=0\mathrm{Im\,}\omega(\zeta)=0 almost everywhere on 𝕋{\mathbb{T}}. Now it is an easy exercise to see that this together with ω(0)=0\omega(0)=0 implies ω=0\omega=0. Just consider the bounded analytic function g=exp{iω}g=\exp\{i\omega\} in 𝔻{\mathbb{D}} whose boundary values on the circle have modulus one almost everywhere, hence g=1\|g\|_{\infty}=1, and note that g(0)=1g(0)=1; it follows that g1g\equiv 1, hence ω0\omega\equiv 0 (that is, F1F\equiv 1). ∎

Here is the counterpart of this statement with assumptions on ω\omega.

Proposition 4.

Let F=ωF=\ell\circ\omega, where ω\omega is an inner function. Then TF,ϕ(𝒫)𝒫T_{F,\phi}({\mathcal{P}})\subset{\mathcal{P}} if and only if ϕ0\phi\equiv 0.

Proof.

After passing to the radial limits in (1) we get that ϕImω=0\phi\mathrm{Im\,}\omega=0 almost everywhere on the unit circle.

If ϕ=0\phi=0 only on a set of measure zero on the circle, then Im ω=0\omega=0 almost everywhere on the circle. From the proof of the previous theorem we know that ω0\omega\equiv 0, which contradicts our initial assumption. Hence ϕ=0\phi=0 on a set of positive measure. By a classical theorem of Nevanlinna [7, Theorem 2.2], it follows that ϕ0\phi\equiv 0. ∎

It is easily seen from Theorem 2 that any admissible multiplication symbol FF can only carry a very small portion of the boundary of the unit disk to the imaginary axis.

Proposition 5.

Let F𝒫F\in{\mathcal{P}}, let ϕ\phi and ω\omega be two Schwarz-type functions, ϕ0\phi\not\equiv 0, and suppose that F=ωF=\ell\circ\omega and TF,ϕT_{F,\phi} preserves  𝒫{\mathcal{P}} as before. Denote the radial limits of FF again by FF and let

A={ζ𝕋:ReF(ζ)=0}={ζ𝕋:|ω(ζ)|=1,ω(ζ)1}.A=\{\zeta\in{\mathbb{T}}\,\colon\,\mathrm{Re\,}F(\zeta)=0\}=\{\zeta\in{\mathbb{T}}\,\colon\,|\omega(\zeta)|=1,\,\omega(\zeta)\neq 1\}\,.

Then m(A)=0m(A)=0.

Proof.

Assume the contrary: m(A)>0m(A)>0. After passing on to the radial limits in (1), we obtain

4|ϕ(ζ)||Imω(ζ)|(1|ω(ζ)|2)(1|ϕ(ζ)|2)4|\phi(\zeta)|\cdot|\mathrm{Im\,}\omega(\zeta)|\leq(1-|\omega(\zeta)|^{2})(1-|\phi(\zeta)|^{2})

for almost all ζ\zeta with |ζ|=1|\zeta|=1. Specifically, ϕImω=0\,\phi\mathrm{Im\,}\omega=0 holds at almost every point of AA (note that ϕ\phi may not have radial limits at some subset of AA of total measure zero). Since the measure of AA is positive and ϕ0\phi\not\equiv 0, we must have either ω=1\omega=1 or ω=1\omega=-1 on a set of positive measure in AA. The first case is excluded by the definition of AA and in the second case the Nevanlinna theorem implies that ω1\omega\equiv-1 in 𝔻{\mathbb{D}}, which is impossible in view of the assumption that ω(0)=0\omega(0)=0. This shows that m(A)=0m(A)=0. ∎

In the context of (linear) weighted composition transformations the case in which F=ϕF=\phi^{\prime} is often important. However, in our context it should be noted that in this case we only obtain another rigidity situation. Namely, assuming that Tϕ,ϕ(𝒫)𝒫T_{\phi^{\prime},\phi}({\mathcal{P}})\subset{\mathcal{P}} and choosing f1f\equiv 1 we get ϕ𝒫\phi^{\prime}\in{\mathcal{P}} hence ϕ(0)=1\phi^{\prime}(0)=1. The case of equality in the Schwarz lemma forces ϕ(z)=z\phi(z)=z, hence F=ϕ1F=\phi^{\prime}\equiv 1, so our transformation Tϕ,ϕT_{\phi^{\prime},\phi} reduces to the identity map.

3.2. Cases where one of the symbols has small range

Many non-trivial examples of weighted composition transformations that preserve class 𝒫{\mathcal{P}} are possible when ϕ(𝔻)\phi({\mathbb{D}}) is compactly contained in 𝔻{\mathbb{D}} or F(𝔻)F({\mathbb{D}}) is contained in a sector, as the following results show.

Proposition 6.

Let F𝒫F\in{\mathcal{P}} and let ϕ\phi be a Schwarz-type function such that ϕ=R<1\|\phi\|_{\infty}=R<1. Then whenever the function FF satisfies

|argF(z)|<π2arcsin2R1+R2|\arg F(z)|<\frac{\pi}{2}-\arcsin\frac{2R}{1+R^{2}}

for all zz in 𝔻{\mathbb{D}}, we have that TF,ϕ(𝒫)𝒫T_{F,\phi}({\mathcal{P}})\subset{\mathcal{P}}.

Proof.

Follows from criterion (d) of Theorem 2 and the fact that the function 2x/(1+x2)2x/(1+x^{2}) is increasing in the interval (0,1)(0,1). ∎

Example 1. An explicit example is ϕ(z)=Rz\phi(z)=Rz, 0<R<10<R<1, and

F(z)=(1+z1z)ε,0<ε<12πarcsin2R1+R2,F(z)=\left(\frac{1+z}{1-z}\right)^{\varepsilon}\,,\qquad 0<\varepsilon<1-\frac{2}{\pi}\arcsin\frac{2R}{1+R^{2}}\,,

a conformal map of the unit disk onto an angular sector with vertex at the origin. Condition (3) is clearly satisfied.

Proposition 7.

Let F𝒫F\in{\mathcal{P}} and K=supz𝔻|argF(z)|<π2K=\sup_{z\in{\mathbb{D}}}|\arg F(z)|<\frac{\pi}{2}. Write K=arcsin2R1+R2K=\arcsin\frac{2R}{1+R^{2}}, 0R<10\leq R<1. If  ϕ\phi is a Schwarz-type function such that

ϕ1R1+R\|\phi\|_{\infty}\leq\frac{1-R}{1+R}

then TF,ϕ(𝒫)𝒫T_{F,\phi}({\mathcal{P}})\subset{\mathcal{P}}.

Proof.

By assumption,

|argF(z)|K=arcsin2R1+R2,z𝔻.|\arg F(z)|\leq K=\arcsin\frac{2R}{1+R^{2}}\,,\qquad z\in{\mathbb{D}}\,.

In view of condition (3) from Theorem 2 it suffices to check that

arcsin2R1+R2<π2arcsin2|ϕ(z)|1+|ϕ(z)|2\arcsin\frac{2R}{1+R^{2}}<\frac{\pi}{2}-\arcsin\frac{2|\phi(z)|}{1+|\phi(z)|^{2}}

holds for all zz in 𝔻{\mathbb{D}}. Equivalently,

arcsin2|ϕ(z)|1+|ϕ(z)|2<π2arcsin2R1+R2\arcsin\frac{2|\phi(z)|}{1+|\phi(z)|^{2}}<\frac{\pi}{2}-\arcsin\frac{2R}{1+R^{2}}

must hold throughout 𝔻{\mathbb{D}}. This will certainly be satisfied if

(6) arcsin2ϕ1+ϕ2π2arcsin2R1+R2\arcsin\frac{2\|\phi\|_{\infty}}{1+\|\phi\|_{\infty}^{2}}\leq\frac{\pi}{2}-\arcsin\frac{2R}{1+R^{2}}

in view of monotonicity of the sine function in [0,π2)[0,\frac{\pi}{2}) and of u(x)=2x1+x2u(x)=\frac{2x}{1+x^{2}} in [0,1)[0,1). But (6) is clearly equivalent to

2ϕ1+ϕ2sin(π2arcsin2R1+R2)=cos(arcsin2R1+R2)=1(2R1+R2)2=1R21+R2.\frac{2\|\phi\|_{\infty}}{1+\|\phi\|_{\infty}^{2}}\leq\sin\left(\frac{\pi}{2}-\arcsin\frac{2R}{1+R^{2}}\right)=\cos\left(\arcsin\frac{2R}{1+R^{2}}\right)=\sqrt{1-\left(\frac{2R}{1+R^{2}}\right)^{2}}=\frac{1-R^{2}}{1+R^{2}}\,.

This yields an elementary quadratic inequality in ϕ\|\phi\|_{\infty} which is easily seen to be satisfied whenever

0ϕ1R1+R.0\leq\|\phi\|_{\infty}\leq\frac{1-R}{1+R}\,.

This proves the statement. ∎

We now formulate a counterpart of Proposition 6 with similar hypotheses on ω\omega instead of ϕ\phi which follows from our previous result.

Corollary 8.

Let F=ω𝒫F=\ell\circ\omega\in{\mathcal{P}}, where ω\omega is a Schwarz-type function. If ω<1\,\|\omega\|_{\infty}<1 and ϕ\phi is a Schwarz-type function such that

ϕ<1ω1+ω\|\phi\|_{\infty}<\frac{1-\|\omega\|_{\infty}}{1+\|\omega\|_{\infty}}

then TF,ϕ(𝒫)𝒫T_{F,\phi}({\mathcal{P}})\subset{\mathcal{P}}.

Proof.

Let R=ω<1R=\|\omega\|_{\infty}<1. Then the function FF is clearly subordinated to the function

R=1+Rz1Rz\ell_{R}=\frac{1+Rz}{1-Rz}

in the usual sense that F=R(ωR)F=\ell_{R}\circ\left(\frac{\omega}{R}\right). Thus, F(𝔻)R(𝔻)F({\mathbb{D}})\subset\ell_{R}({\mathbb{D}}). It is plain that R(𝔻)\ell_{R}({\mathbb{D}}) is the disk whose diameter has endpoints

R(1)=1R1+R,R(1)=1+R1R,\ell_{R}(-1)=\frac{1-R}{1+R}\,,\qquad\ell_{R}(1)=\frac{1+R}{1-R}\,,

hence its center and radius are respectively

C=1+R21R2,ρ=2R1R2.C=\frac{1+R^{2}}{1-R^{2}}\,,\qquad\rho=\frac{2R}{1-R^{2}}\,.

Let us denote by CR={z:|zC|=ρ}C_{R}=\{z\in{\mathbb{C}}\,\colon\,|z-C|=\rho\} the boundary of this disk. Let aa be the point of intersection of the circle CRC_{R} with its tangent from the origin in the upper half-plane. By looking at the right triangle determined by the origin and the points aa and CC, we infer that

arga=arcsinρC=arcsin2R1+R2.\arg a=\arcsin\frac{\rho}{C}=\arcsin\frac{2R}{1+R^{2}}\,.

One argues similarly for the point of tangent in the lower half-plane and obtains that, for every zz in 𝔻{\mathbb{D}},

|argF(z)|<arcsin2R1+R2.|\arg F(z)|<\arcsin\frac{2R}{1+R^{2}}\,.

The conclusion now follows from Proposition 7. ∎

Many interesting examples in geometric function theory are obtained from special types of conformal mappings such as the lens maps. In what follows, for 0<α<10<\alpha<1 we will denote by λα\lambda_{\alpha} the standard lens map given by the formula

λα(z)=(1α)(z)=α(z)1α(z)+1,z𝔻.\lambda_{\alpha}(z)=(\ell^{-1}\circ\ell^{\alpha})(z)=\frac{\ell^{\alpha}(z)-1}{\ell^{\alpha}(z)+1}\,,\qquad z\in{\mathbb{D}}\,.

It is elementary that λα\lambda_{\alpha} is a conformal map of the unit disk onto a lens-shaped region LαL_{\alpha} bounded by two circular arcs (symmetric with respect to the real axis) that intersect at the points ±1\pm 1 forming an angle of opening πα\pi\alpha at each of these points; see [13, p. 27].

The following simple geometric observation will be useful. The half-plane map \ell is bijective between a lens-shaped region LαL_{\alpha} and an angle with vertex at the origin and maps in a one-to-one fashion the largest disk contained in the lens-shaped region onto a disk tangent to the legs of the angle.

Our next result essentially shows that when the multiplication symbol is obtained by composing the half-plane map with a lens map, the statements of Proposition 6 and Proposition 7 can be unified into a single “if and only if” statement.

Proposition 9.

Let F=λαF=\ell\circ\lambda_{\alpha}, where λα\lambda_{\alpha} is a lens map, and let ϕ\phi be a Schwarz-type function. Then TF,ϕ(𝒫)𝒫T_{F,\phi}(\mathcal{P})\subset\mathcal{P} if and only if

ϕ1R1+R,\|\phi\|_{\infty}\leq\frac{1-R}{1+R}\,,

where 2R1+R2=sinαπ2\frac{2R}{1+R^{2}}=\sin{\frac{\alpha\pi}{2}}.

Proof.

Suppose first that ϕ1R1+R\|\phi\|_{\infty}\leq\frac{1-R}{1+R}. Then, since

supz𝔻|argF(z)|=απ2=K=arcsin2R1+R2,\sup_{z\in\mathbb{D}}|\arg\,F(z)|=\frac{\alpha\pi}{2}=K=\arcsin\frac{2R}{1+R^{2}}\,,

by Proposition 7 it follows that TF,ϕ(𝒫)𝒫T_{F,\phi}(\mathcal{P})\subset\mathcal{P}. Now, if TF,φ(𝒫)𝒫T_{F,\varphi}(\mathcal{P})\subset\mathcal{P}, by condition (3) of Theorem 2 we have

|argF(z)|+arcsin2|ϕ(z)|1+|ϕ(z)|2<π2.|\arg\,F(z)|+\arcsin\frac{2|\phi(z)|}{1+|\phi(z)|^{2}}<\frac{\pi}{2}\,.

Therefore, for almost every ζ𝕋\zeta\in\mathbb{T},

arcsin2|ϕ(ζ)|1+|ϕ(ζ)|2π2απ2=π2arcsin2R1+R2.\arcsin\frac{2|\phi(\zeta)|}{1+|\phi(\zeta)|^{2}}\leq\frac{\pi}{2}-\frac{\alpha\pi}{2}=\frac{\pi}{2}-\arcsin\frac{2R}{1+R^{2}}\,.

Then, as in the proof of Proposition 7,

2ϕ1+ϕ2sin(π2arcsin2R1+R2)=1R21+R2,\frac{2\|\phi\|_{\infty}}{1+\|\phi\|_{\infty}^{2}}\leq\sin\left(\frac{\pi}{2}-\arcsin\frac{2R}{1+R^{2}}\right)=\frac{1-R^{2}}{1+R^{2}},

and from here,

ϕ1R1+R.\|\phi\|_{\infty}\leq\frac{1-R}{1+R}\,.

3.3. Composition symbols with radial limits of modulus one and/or angular derivatives

Recall that an analytic self-map ϕ\phi of 𝔻{\mathbb{D}} is said to have an angular derivative ϕ(ζ)\phi^{\prime}(\zeta) (in the restricted sense of Carathéodory [3, § 298-299]) at a point ζ\zeta on the unit circle 𝕋{\mathbb{T}} if it satisfies the following two conditions:

(a) the nontangential limit of ϕ\phi at ζ\zeta has modulus one;

(b) ϕ(z)\phi^{\prime}(z) has a finite nontangential limit as zζz\to\zeta.

The Julia-Carathéodory theorem (see [5] or [13]) states that ϕ\phi has an angular derivative at ζ\zeta if and only if

lim infzζ1|ϕ(z)|1|z|<\liminf_{z\to\zeta}\frac{1-|\phi(z)|}{1-|z|}<\infty

and, in this case, |ϕ(ζ)||\phi^{\prime}(\zeta)| equals the above (unrestricted) lower limit (note that this limit is always strictly positive [13, p. 57]. Otherwise it is understood that |ϕ(ζ)|=|\phi^{\prime}(\zeta)|=\infty.

Even though the angular derivative of ϕ\phi need not exist anywhere on 𝕋{\mathbb{T}} as a finite number, the function |ϕ|:𝕋[0,]|\phi^{\prime}|:{\mathbb{T}}\to[0,\infty] is well defined in this extended sense; being lower semicontinuous ([2], Lemma 2.5), it attains its minimum on 𝕋{\mathbb{T}} (cf. also [5, Proposition 2.46]).

One ought to keep in mind that the function |ϕ||\phi^{\prime}| on 𝕋{\mathbb{T}} as above, in general, does not coincide at all with the modulus of the boundary values of ϕ\phi^{\prime} (if those exist). The most obvious example is the linear map ϕ(z)=az+b\phi(z)=az+b onto a disk compactly contained in 𝔻{\mathbb{D}}, which happens precisely when |a|+|b|<1|a|+|b|<1. Its usual derivative is constant everywhere, while the angular derivative does not exist at any point on the boundary; in this case, we interpret that |ϕ(ζ)|=|\phi^{\prime}(\zeta)|=\infty for every point ζ\zeta on the unit circle.

The concept of angular derivative is fundamental in the study of compactness of composition operators on Hardy and Bergman spaces, as well as in the iteration of analytic self-maps of the unit disk.

Our next result shows that if ϕ\phi possesses even a mildly reasonable boundary behavior at a point on the unit circle then ω\omega automatically cannot be “too good” at the same point.

Theorem 10.

Let FF, ω\omega, and ϕ\phi be as before and suppose that at some point ζ\zeta on the unit circle the function ϕ\phi has radial limit of modulus one. Then if the transformation TF,ϕT_{F,\phi} preserves 𝒫{\mathcal{P}}, the function ω\omega cannot have a finite non-zero angular derivative at ζ\zeta.

Proof.

Note that TF,ϕT_{F,\phi} preserves 𝒫{\mathcal{P}} if and only if the transformation TFλ,ϕλT_{F_{\lambda},\phi_{\lambda}} preserves 𝒫{\mathcal{P}}, where Fλ(z)=F(λz)F_{\lambda}(z)=F(\lambda z) and ϕλ(z)=ϕ(λz)\phi_{\lambda}(z)=\phi(\lambda z), whenever |λ|=1|\lambda|=1. Hence, we may assume without loss of generality that ζ=1\zeta=1.

Suppose that ω\omega has a finite angular derivative at ζ=1\zeta=1. Then the radial limit ω(1)\omega(1) exists and |ω(1)|=1|\omega(1)|=1. Taking the angular limit as z1z\to 1 in (1), we conclude that Im {ω(1)}=0\{\omega(1)\}=0. Thus, either ω(1)=1\omega(1)=-1 or ω(1)=1\omega(1)=1.

Let us first consider the case ω(1)=1\omega(1)=-1. Since at ζ=1\zeta=1 the angular derivative of ω\omega is neither 0 nor \infty, we know [12, p. 291] that it is actually univalent in some Stolz domain with vertex at z=1z=1:

Δ={z:|arg(1z)|<θ,r<|z|<1}\Delta=\left\{z\colon|\arg\,(1-z)|<\theta,\ r<|z|<1\right\}

for suitable r(0,1)r\in(0,1) and θ>0\theta>0. Also, as is well known (cf. again [12, p. 291]), the function ω\omega preserves angles between curves contained in Δ{1}{\Delta}\cup\{1\} that meet at z=1z=1. This shows that there exists a curve γ:[0,1]Δ{1}\gamma:[0,1]\to\Delta\cup\{1\} with γ(1)=1\gamma(1)=1 and which is mapped by ω\omega onto some non-horizontal segment

S={1+seiα0:0ss0},0<|α0|<π2,S=\{-1+se^{i\alpha_{0}}\colon 0\leq s\leq s_{0}\}\,,\quad 0<|\alpha_{0}|<\frac{\pi}{2}\,,

for an appropriate value of s0s_{0}. (To see this, it suffices to look at the image under ω\omega of the suitable Stolz domain mentioned earlier with vertex at 11, which will contain another Stolz domain with vertex at 1-1, and to select α0\alpha_{0} and s0s_{0} so that the segment SS is contained in this new Stolz angle and is not contained in the real axis.) Keeping in mind that F=ωF=\ell\circ\omega and \ell is a Möbius transformation which maps the diameter (1,1)(-1,1) to the positive semi-axis, we see that

argF(γ(t))=arg(ω(γ(t)))α0,ast1.\arg\,F(\gamma(t))=\arg\,\ell(\omega(\gamma(t)))\to\alpha_{0}\,,\quad\text{as}\quad t\to 1^{-}\,.

Therefore, taking the limit as z1z\to 1 along γ\gamma in (3), we obtain |α0|0|\alpha_{0}|\leq 0, which is contrary to our construction of the segment SS. This completes the proof in the case when ω(1)=1\omega(1)=-1.

By (1), TF,ϕT_{F,\phi} preserves 𝒫{\mathcal{P}} if and only if TG,ϕT_{G,\phi} with G=(1ω)/(1+ω)G=(1-\omega)/(1+\omega) does, so we can argue as above in the case when ω(1)=1\omega(1)=1 to get a contradiction again. ∎

There are two ways in which the function ω\omega can fail to have angular derivative: either it does not have a radial limit of modulus one or it does but the differential quotient fails to have a limit at the point in question. Here is an example of the first kind. It deals with the map ϕ\phi such that ϕ(𝔻)\phi({\mathbb{D}}) has a tangential contact with the unit circle. The price we pay for this is that ω\omega is a dilated self-map of the disk (hence, in this example ω(𝔻)\omega({\mathbb{D}}) is compactly contained in 𝔻{\mathbb{D}}).

Example 2. For K3/2K\geq 3/2, let

ϕ(z)=z(1+z)2,ω(z)=z(2z)2K.\phi(z)=\frac{z(1+z)}{2}\,,\qquad\omega(z)=\frac{z(2-z)}{2K}\,.

Both are clearly Schwarz-type functions. Obviously, ϕ(1)=1\phi(1)=1 and ϕ\phi is conformal at z=1z=1 since ϕ(1)0\phi^{\prime}(1)\neq 0. For a sufficiently large value of KK (which will be determined below) one can also check that our condition (1) is satisfied, hence TF,ϕ(𝒫)𝒫T_{F,\phi}({\mathcal{P}})\subset{\mathcal{P}}. Indeed, it is immediate that

Imω(z)=y(1x)K,z=x+iy.\mathrm{Im\,}\omega(z)=\frac{y(1-x)}{K}\,,\qquad z=x+iy\,.

Checking our condition (1) in this case reduces to verifying that

2|z(1+z)y|(1x)K<(1|z(1+z)2|2)(1|z(2z)2K|2)\frac{2\,\left|z(1+z)y\right|\,(1-x)}{K}<\left(1-\left|\frac{z(1+z)}{2}\right|^{2}\right)\left(1-\left|\frac{z(2-z)}{2K}\right|^{2}\right)

holds for all zz in 𝔻{\mathbb{D}}. (Note that as z1z\to 1, both sides tend to zero but the strict inequality is maintained.) Since x2+y2=|z|2<1x^{2}+y^{2}=|z|^{2}<1, it is clear that

2|z(1+z)y|(1x)K<4(1x)K\frac{2\,\left|z(1+z)y\right|\,(1-x)}{K}<\frac{4(1-x)}{K}

while the right-hand side can be estimated from below as follows:

(1|z(1+z)2|2)(1|z(2z)2K|2)\displaystyle\left(1-\left|\frac{z(1+z)}{2}\right|^{2}\right)\left(1-\left|\frac{z(2-z)}{2K}\right|^{2}\right) >\displaystyle> (1|1+z|24)(194K2)\displaystyle\left(1-\frac{|1+z|^{2}}{4}\right)\left(1-\frac{9}{4K^{2}}\right)
=\displaystyle= (1(1+x)2+y24)(194K2)\displaystyle\left(1-\frac{(1+x)^{2}+y^{2}}{4}\right)\left(1-\frac{9}{4K^{2}}\right)
\displaystyle\geq (12+2x4)(194K2)\displaystyle\left(1-\frac{2+2x}{4}\right)\left(1-\frac{9}{4K^{2}}\right)
=\displaystyle= 1x2(194K2)\displaystyle\frac{1-x}{2}\left(1-\frac{9}{4K^{2}}\right)

so it is only left to check that

4(1x)K<1x2(194K2)\frac{4(1-x)}{K}<\frac{1-x}{2}\left(1-\frac{9}{4K^{2}}\right)

for KK large enough and |x|<1|x|<1, which is clear. The inequality holds for all K>4+732K>4+\dfrac{\sqrt{73}}{2}.

The natural question arises as to whether it is possible to have an example where both ϕ\phi and ω\omega can have radial limits of modulus one at the same point (obviously, without having an angular derivative at the point in question) but the weighted composition TF,ϕT_{F,\phi} still preserves 𝒫{\mathcal{P}}. The following example, illustrated by the figure below, gives an affirmative answer.

Example 3. Consider the planar domain

Ω={x+iy: 4|y|x2+y2<(1x2y2)2},\Omega=\{x+iy\,\colon\,4|y|\sqrt{x^{2}+y^{2}}<(1-x^{2}-y^{2})^{2}\}\,,

clearly symmetric with respect to both the real and imaginary axes. Let ω\omega be a conformal map of 𝔻\,{\mathbb{D}} onto Ω\Omega which fixes the origin. Starting with the subdomain of Ω\Omega in the upper half-plane and using the Schwarz reflection principle, one can also choose ω\omega in such a fashion that it fixes the diameter (1,1)(-1,1) and ω(1)=1\omega(1)=1 in the sense of a radial limit. Let ϕ=ω\phi=\omega. It can now easily be checked that our condition (1) is satisfied, hence TF,ϕ(𝒫)𝒫T_{F,\phi}({\mathcal{P}})\subset{\mathcal{P}}.

Note, however, in relation to this “leaf-shaped” region that our mapping ω=ϕ\omega=\phi has boundary contact with the unit circle but does not have angular derivative at z=1z=1. The intuitive reason for this is that the corners at 1-1 and 11 are contained in lens-shaped regions and lens maps do not have angular derivatives. A rigorous proof of this fact can be given by using subordination. Alternatively, one can write the equation of the boundary of Ω\Omega in polar coordinates:

4r2|sinθ|=(1r2)2.4r^{2}|\sin\theta|=(1-r^{2})^{2}\,.

Solving for rr, one obtains

1r2=2(sin|θ|+sin2|θ|sin|θ|),1-r^{2}=2\left(\sqrt{\sin|\theta|+\sin^{2}|\theta|}-\sin|\theta|\right)\,,

and by elementary calculus one checks that for a fixed ε>0\varepsilon>0 the integral

0ε1r(θ)θ2𝑑θ\int_{0}^{\varepsilon}\frac{1-r(\theta)}{\theta^{2}}\,d\theta

diverges and the conclusion follows by the Tsuji-Warschawski criterion [13, p. 72].

Refer to caption
Figure 1. The boundary of the leaf-shaped region Ω\Omega.

4. Fixed points of weighted composition transformations that preserve 𝒫{\mathcal{P}}

Even though we are working in a non-linear context, it is possible to adapt the arguments typical for such situations; cf. [13, Sect. 6.1].

Theorem 11.

Let TF,ϕT_{F,\phi} be a weighted composition transformation such that TF,ϕ(𝒫)𝒫,T_{F,\phi}(\mathcal{P})\subset\mathcal{P}, where F=ω,F=\ell\circ\omega, ϕ\phi and ω\omega are Schwarz-type functions, and ϕ\phi is not a rotation. Then TF,ϕT_{F,\phi} has a unique fixed point which is obtained by iterating TF,ϕT_{F,\phi} applied to arbitrary ff in 𝒫{\mathcal{P}}.

In the case when ϕ\,\phi is inner but not a rotation, the unique fixed point is the constant function one.

Proof.

We first show that the limit of iterates of TF,ϕT_{F,\phi} applied to an arbitrary function ff in 𝒫{\mathcal{P}} is a fixed point of the transformation. Define the iterations of ϕ\phi in the usual way, ϕ0\phi_{0} being the identity function and ϕn+1=ϕnϕ\phi_{n+1}=\phi_{n}\circ\phi, n0n\geq 0. Let f𝒫f\in{\mathcal{P}}. It is easy to see by induction that

F(Fϕ)(Fϕn1)(fϕn)=TF,ϕnf𝒫F(F\circ\phi)\ldots(F\circ\phi_{n-1})(f\circ\phi_{n})=T^{n}_{F,\phi}f\in\mathcal{P}

for any integer n1n\geq 1. By our assumptions on ϕ\phi, the origin is its only fixed point in 𝔻{\mathbb{D}}. Since ϕ\phi is not a disk automorphism, it follows that ϕn0\phi_{n}\to 0 uniformly on compact subsets of 𝔻{\mathbb{D}} and therefore also ωϕn0\omega\circ\phi_{n}\to 0 in the compact-open topology as nn\to\infty. Thus, fϕn1f\circ\phi_{n}\to 1 uniformly on compact subsets as nn\to\infty. On the other hand,

k=0n1(Fϕk)=k=0n11+ωϕk1ωϕk,\prod_{k=0}^{n-1}(F\circ\phi_{k})=\prod_{k=0}^{n-1}\frac{1+\omega\circ\phi_{k}}{1-\omega\circ\phi_{k}}\,,

so proving the uniform convergence on compact subsets of the infinite product k=0(Fϕk)\prod_{k=0}^{\infty}(F\circ\phi_{k}) is equivalent to proving the convergence on compact subsets of 𝔻{\mathbb{D}} of the sums

k=0n1|11+ωϕk1ωϕk|=2k=0n1|ωϕk1ωϕk|.\sum_{k=0}^{n-1}\left|1-\frac{1+\omega\circ\phi_{k}}{1-\omega\circ\phi_{k}}\right|=2\sum_{k=0}^{n-1}\left|\frac{\omega\circ\phi_{k}}{1-\omega\circ\phi_{k}}\right|\,.

For r(0,1)r\in(0,1) fixed, let m(r)=max|z|r|ϕ(z)|m(r)=\max_{|z|\leq r}|\phi(z)|. Let δ=m(r)/r\delta=m(r)/r. Clearly, δ<1\delta<1 since ϕ\phi is not a rotation. Applying the Schwarz lemma to ϕ(rw)/m(r)\phi(rw)/m(r), we get |ϕ(rw)|/m(r)|w||\phi(rw)|/m(r)\leq|w| for any w𝔻w\in\mathbb{D}, and from here

|ϕ(z)|m(r)r|z|=δ|z||\phi(z)|\leq\frac{m(r)}{r}|z|=\delta|z|

whenever |z|r|z|\leq r. Iterating this inequality, we get

|ϕk(z)|δ|ϕk1(z)|δk|z||\phi_{k}(z)|\leq\delta|\phi_{k-1}(z)|\leq\ldots\leq\delta^{k}|z|

for |z|r|z|\leq r. Using the fact that ω\omega is a Schwarz-type function, we obtain

|11+ωϕk1ωϕk|=2|ωϕk1ωϕk|2|ωϕk|1|ωϕk|2|ϕk|1|ϕk|2δkr1δkr2r1rδk\left|1-\frac{1+\omega\circ\phi_{k}}{1-\omega\circ\phi_{k}}\right|=2\left|\frac{\omega\circ\phi_{k}}{1-\omega\circ\phi_{k}}\right|\leq 2\frac{|\omega\circ\phi_{k}|}{1-|\omega\circ\phi_{k}|}\leq 2\frac{|\phi_{k}|}{1-|\phi_{k}|}\leq 2\frac{\delta^{k}r}{1-\delta^{k}r}\leq\frac{2r}{1-r}\delta^{k}

in the disk {z:|z|r}\{z\,\colon\,|z|\leq r\}. Thus, the series

k=0|11+ωϕk1ωϕk|\sum_{k=0}^{\infty}\left|1-\frac{1+\omega\circ\phi_{k}}{1-\omega\circ\phi_{k}}\right|

converges uniformly on compact subsets of the disk and the infinite product k=0(Fϕk)\prod_{k=0}^{\infty}(F\circ\phi_{k}) is uniformly convergent on compact subsets to some function GG analytic in 𝔻{\mathbb{D}}. Moreover, since 𝒫\mathcal{P} is a compact class, G𝒫G\in\mathcal{P}. Combining both limits, we obtain

TF,ϕnf=F(Fϕ)(Fϕn1)(fϕn)GT^{n}_{F,\phi}f=F(F\circ\phi)\ldots(F\circ\phi_{n-1})(f\circ\phi_{n})\to G

uniformly on compact subsets as nn\to\infty for any f𝒫f\in\mathcal{P}. Now we can see that GG is a fixed point of the transformation. Applying TF,ϕT_{F,\phi} to GG we have

TF,ϕG\displaystyle T_{F,\phi}G =F(Gϕ)=F(limnk=0n1(Fϕk))ϕ=F(limnk=0n1(Fϕk+1))\displaystyle=F(G\circ\phi)=F\left(\lim_{n\to\infty}\prod_{k=0}^{n-1}(F\circ\phi_{k})\right)\circ\phi=F\left(\lim_{n\to\infty}\prod_{k=0}^{n-1}(F\circ\phi_{k+1})\right)
=F(limnk=1n(Fϕk))=limnk=0n(Fϕk)=G.\displaystyle=F\left(\lim_{n\to\infty}\prod_{k=1}^{n}(F\circ\phi_{k})\right)=\lim_{n\to\infty}\prod_{k=0}^{n}(F\circ\phi_{k})=G\,.

Note that GG as constructed above does not depend on the initial choice of the function ff in 𝒫{\mathcal{P}}. Now it is clear that this GG is the only fixed point of TF,ϕT_{F,\phi}, because if g𝒫g\in\mathcal{P} satisfies TF,ϕg=gT_{F,\phi}g=g, iterating the transformation we get

g=TF,ϕngGg=T^{n}_{F,\phi}g\to G

uniformly on compact subsets of 𝔻{\mathbb{D}}.

It is only left to check our final comment in the statement of the theorem. Since ϕ\phi is an inner function, by Proposition 3 we have F1F\equiv 1 so the equation for the fixed point: fϕ=ff\circ\phi=f is the classical Shröder equation for the composition operator CϕC_{\phi} corresponding to the eigenvalue λ=1\lambda=1. By the first proposition from [13, Section 6.1], if ff were not identically constant, it would follow that |ϕ(0)|=1|\phi^{\prime}(0)|=1. In view of the Schwarz lemma, this forces ϕ\phi to be a rotation, contrary to our assumptions. ∎

The case when ϕ\phi is a rotation leads to the well-known case of fixed points from the theory of composition operators, describing a trichotomy: the identity map, a rational rotation or an irrational rotation.

Proposition 12.

Let TF,ϕT_{F,\phi} be a weighted composition transformation such that TF,ϕ(𝒫)𝒫T_{F,\phi}(\mathcal{P})\subset\mathcal{P}, where F=ωF=\ell\circ\omega, ω\omega is a Schwarz-type functions, and ϕ\phi is a rotation. Then the set of all fixed points of TF,ϕT_{F,\phi} is as follows:

(a) all of 𝒫\,{\mathcal{P}}, if ϕ(z)z\phi(z)\equiv z;

(b) the functions with nn-fold symmetry: f(z)=g(zn)f(z)=g(z^{n}), g𝒫g\in{\mathcal{P}}, whenever ϕ(z)=λz\phi(z)=\lambda z, where λn=1\lambda^{n}=1 for some n>1n>1;

(c) only the constant function one, if ϕ(z)=λz\phi(z)=\lambda z, where |λ|=1|\lambda|=1 and λn1\lambda^{n}\neq 1 for all nn\in{\mathbb{N}}.

Proof.

Since ϕ\phi is an inner function, Proposition 3 forces F1F\equiv 1, hence TF,ϕf=fϕT_{F,\phi}f=f\circ\phi. Part (a) now follows trivially.

Parts (b) and (c) follow readily by comparing the Taylor series of both sides of the equality f(λz)=f(z)f(\lambda z)=f(z) in the disk. ∎

References

  • [1] R. Aron, M. Lindström, Spectra of weighted composition operators on weighted Banach spaces of analytic functions, Israel J. Math. 141 (2004), 263–-276.
  • [2] P. Bourdon, J.H. Shapiro, Riesz composition operators, Pacific J. Math. 181 (1997), 231–246.
  • [3] C. Carathéodory, Theory of Functions. II, Chelsea, New York 1960.
  • [4] M.D. Contreras, A.G. Hernández-Díaz, Weighted composition operators on Hardy spaces, J. Math. Anal. Appl. 263 (2001), no. 1, 224–233.
  • [5] C. Cowen, B. MacCluer, Composition Operators on Spaces of Analytic Functions, Studies in Advanced Mathematics, CRC Press, Boca Ratón 1995.
  • [6] Ž. Čučkovic, R. Zhao, Weighted composition operators on the Bergman space, J. London Math. Soc. (2) 70 (2004), no. 2, 499–-511.
  • [7] P.L. Duren, Theory of HpH^{p} Spaces, Pure and Applied Mathematics, Vol. 38, Second edition, Dover, Mineola, New York 2000.
  • [8] P.L. Duren, Univalent Functions, Springer-Verlag, New York 1983.
  • [9] H. Kamowitz, Compact operators of the form uCφuC_{\varphi}, Pacific J. Math. 80 (1979), no. 1, 205–211.
  • [10] A.K. Kitover, The spectrum of operators in ideal spaces (Russian), Investigations on linear operators and the theory of functions, VII, Zap. Naucn. Sem. Leningrad. Otdel Mat. Inst. Steklov (LOMI) 65 (1976), 196–198, 209–210.
  • [11] Ch. Pommerenke, Univalent Functions (with a chapter on quadratic differentials by G. Jensen), Vandenhoeck & Ruprecht, Göttingen 1975.
  • [12] Ch. Pommerenke, On the angular derivative and univalence, Anal. Math. 3 (1977), 291–297.
  • [13] J.H. Shapiro, Composition Operators and Classical Function Theory, Springer-Verlag, New York 1993.