On weighted compositions preserving the Carathéodory class
Abstract.
We characterize in various ways the weighted composition transformations which preserve the class 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, 47B331. Introduction
1.1. The class and its properties
Let denote the unit disk in the complex plane and the algebra of all functions analytic in . If and , we will say that is an analytic self-map of the disk. If, moreover, such satisfies , we will refer to it as a Schwarz-type function (as in the classical Schwarz lemma).
Denote by the Carathéodory class of all in with positive real part and normalized so that . An important example of a function in this class is the so-called half-plane mapping given by
This conformal map of the disk onto the right half-plane is extremal in many senses for the class . This is manifested, for example, by the growth theorem for the functions in the class:
The above estimate [11, Section 2.1] is a direct corollary of the Schwarz lemma and the elementary subordination principle since every function in is of the form , where is some Schwarz-type function. In the particular case when with , we will use the symbol to denote the functions ; that is, . In view of the Herglotz representation theorem [8, Chapter 1], the class equals , the closed convex hull of the collection in the topology of uniform convergence on compact subsets of (the compact-open topology). Every Schwarz-type function has radial limits almost everywhere on the unit circle with respect to the normalized arc length measure (see [7, Chapter 1]), hence so does every in .
1.2. On weighted composition transformations
Whenever is an analytic self-map of the disk, it is immediate that for every if and only if is a Schwarz-type function. Thus, it makes sense to consider the composition transformation on defined by the formula . 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 , given by , where . While there are many cases of such transformations in spaces of analytic functions, it turns out that such mapping can preserve the class only in the trivial case when .
One can consider the more general weighted composition transformations on , defined by a composition followed by a multiplication: , where , and , and ask whether such a transformation acts on or preserves in the sense that for every in . This will make sense more often since one of the two symbols and 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 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 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 for which whenever . 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 is of certain type, then the only possible in the trivial case: ; see Proposition 3. Alternatively, when is of special type then 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 and in terms of their qualitative behavior (image compactly contained in , 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 that preserves class has a unique fixed point whenever is not a rotation and this fixed point is obtained by applying the iterates of to an arbitrary function in , as one would expect. This is the content of Theorem 11. In the case when 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 -preserving weighted compositions
2.1. Initial observations
We begin by recording the obvious necessary conditions that must be satisfied by all admissible symbols and , i.e., by those which satisfy the initial assumptions , , and make the inclusion possible.
If for all in , it is clear that then ; this is immediately seen by choosing , which is a function in .
In addition to this, must be a Schwarz-type function since, after choosing , another function obviously in , we get
hence .
Thus, from now on we shall always work assuming these hypotheses: and is a Schwarz-type function.
It is quite easy to establish the lack of non-trivial pointwise multipliers of .
Proposition 1.
If preserves and (that is, for all in ) then .
Proof.
Indeed, since then , by the growth theorem for the functions in we have
for all in . Also, for any fixed we may choose to be a suitable rotation of the half-plane function for which
It follows that for all . Since , the maximum modulus principle implies that is identically constant, hence . ∎
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 on the set of all rotations of the half-plane function in order to know whether the transformation preserves . 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 . Note that for any function in , the function takes on the values only in and is a continuous function in the disk. Moreover, the argument of the product of two such functions and in , with values in , is still continuous and the formula
holds throughout . We will use this fact repeatedly.
Theorem 2.
Let be a Schwarz-type function, , and denote by the Schwarz-type function for which . Consider the argument function defined as above. Then the following conditions are equivalent:
-
(a)
.
-
(b)
.
-
(c)
The inequality
(1) holds for all in . In other words,
(2) -
(d)
The inequality
(3) holds for all in . Note also that
(4) where in the case when (recalling that ) the last equality should be understood as the limit: .
It should be noted that in the above result the inequalities in conditions (c) and (d) are both invariant under rotations of but not under the rotations in (or under the appropriate changes in ).
Proof.
We will show that (a) (b), (b) (c), and (c) (d).
(a) (b) . The implication (a) (b) is obvious so we only have to see that (b) (a). First of all, the image of a convex combination of functions in under is the same convex combination of their images and a convex combination of function in remains in . Hence, if we also have that as by assumption and the class is clearly convex.
Next, if uniformly on compact subsets of , then also in the same topology. Since is a compact family (in the classical terminology, meaning a closed set in the compact-open topology), we get .
To see that (b) (c), suppose that for all in . In other words, for all of modulus one and therefore also
for the Schwarz-type functions depending on each . This leads to the equation
which holds in the entire unit disk. Solving for , we get
The condition in is equivalent to
which amounts to the inequality
(5) |
Grouping the terms in (5) we obtain
for each in and for arbitrary with . For each point we can choose the argument of appropriately so as to get
Since this is valid at every point in the disk, the statement (1) follows.
To see that (c) (b), it suffices to observe that
and it is now easy to reverse the steps in the above proof.
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 in we write .
Recall that bounded analytic functions in have radial limits for almost every point on the unit circle with respect to the normalized Lebesgue arc length measure [7, Chapter 1]. An analytic function in the disk is called inner if for all in (equivalently, ) and also almost everywhere on . The following result generalizes our Proposition 1.
Proposition 3.
Let and let be inner. Then if and only if .
Proof.
The bounded functions and have radial limits almost everywhere on the circle. Thus, for almost every we may pass to the limit as in inequality (1) to conclude that almost everywhere on . Now it is an easy exercise to see that this together with implies . Just consider the bounded analytic function in whose boundary values on the circle have modulus one almost everywhere, hence , and note that ; it follows that , hence (that is, ). ∎
Here is the counterpart of this statement with assumptions on .
Proposition 4.
Let , where is an inner function. Then if and only if .
Proof.
After passing to the radial limits in (1) we get that almost everywhere on the unit circle.
If only on a set of measure zero on the circle, then Im almost everywhere on the circle. From the proof of the previous theorem we know that , which contradicts our initial assumption. Hence on a set of positive measure. By a classical theorem of Nevanlinna [7, Theorem 2.2], it follows that . ∎
It is easily seen from Theorem 2 that any admissible multiplication symbol can only carry a very small portion of the boundary of the unit disk to the imaginary axis.
Proposition 5.
Let , let and be two Schwarz-type functions, , and suppose that and preserves as before. Denote the radial limits of again by and let
Then .
Proof.
Assume the contrary: . After passing on to the radial limits in (1), we obtain
for almost all with . Specifically, holds at almost every point of (note that may not have radial limits at some subset of of total measure zero). Since the measure of is positive and , we must have either or on a set of positive measure in . The first case is excluded by the definition of and in the second case the Nevanlinna theorem implies that in , which is impossible in view of the assumption that . This shows that . ∎
In the context of (linear) weighted composition transformations the case in which is often important. However, in our context it should be noted that in this case we only obtain another rigidity situation. Namely, assuming that and choosing we get hence . The case of equality in the Schwarz lemma forces , hence , so our transformation 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 are possible when is compactly contained in or is contained in a sector, as the following results show.
Proposition 6.
Let and let be a Schwarz-type function such that . Then whenever the function satisfies
for all in , we have that .
Proof.
Follows from criterion (d) of Theorem 2 and the fact that the function is increasing in the interval . ∎
Example 1. An explicit example is , , and
a conformal map of the unit disk onto an angular sector with vertex at the origin. Condition (3) is clearly satisfied.
Proposition 7.
Let and . Write , . If is a Schwarz-type function such that
then .
Proof.
By assumption,
In view of condition (3) from Theorem 2 it suffices to check that
holds for all in . Equivalently,
must hold throughout . This will certainly be satisfied if
(6) |
in view of monotonicity of the sine function in and of in . But (6) is clearly equivalent to
This yields an elementary quadratic inequality in which is easily seen to be satisfied whenever
This proves the statement. ∎
We now formulate a counterpart of Proposition 6 with similar hypotheses on instead of which follows from our previous result.
Corollary 8.
Let , where is a Schwarz-type function. If and is a Schwarz-type function such that
then .
Proof.
Let . Then the function is clearly subordinated to the function
in the usual sense that . Thus, . It is plain that is the disk whose diameter has endpoints
hence its center and radius are respectively
Let us denote by the boundary of this disk. Let be the point of intersection of the circle with its tangent from the origin in the upper half-plane. By looking at the right triangle determined by the origin and the points and , we infer that
One argues similarly for the point of tangent in the lower half-plane and obtains that, for every in ,
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 we will denote by the standard lens map given by the formula
It is elementary that is a conformal map of the unit disk onto a lens-shaped region bounded by two circular arcs (symmetric with respect to the real axis) that intersect at the points forming an angle of opening at each of these points; see [13, p. 27].
The following simple geometric observation will be useful. The half-plane map is bijective between a lens-shaped region 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 , where is a lens map, and let be a Schwarz-type function. Then if and only if
where .
3.3. Composition symbols with radial limits of modulus one and/or angular derivatives
Recall that an analytic self-map of is said to have an angular derivative (in the restricted sense of Carathéodory [3, § 298-299]) at a point on the unit circle if it satisfies the following two conditions:
(a) the nontangential limit of at has modulus one;
(b) has a finite nontangential limit as .
The Julia-Carathéodory theorem (see [5] or [13]) states that has an angular derivative at if and only if
and, in this case, equals the above (unrestricted) lower limit (note that this limit is always strictly positive [13, p. 57]. Otherwise it is understood that .
Even though the angular derivative of need not exist anywhere on as a finite number, the function is well defined in this extended sense; being lower semicontinuous ([2], Lemma 2.5), it attains its minimum on (cf. also [5, Proposition 2.46]).
One ought to keep in mind that the function on as above, in general, does not coincide at all with the modulus of the boundary values of (if those exist). The most obvious example is the linear map onto a disk compactly contained in , which happens precisely when . 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 for every point 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 possesses even a mildly reasonable boundary behavior at a point on the unit circle then automatically cannot be “too good” at the same point.
Theorem 10.
Let , , and be as before and suppose that at some point on the unit circle the function has radial limit of modulus one. Then if the transformation preserves , the function cannot have a finite non-zero angular derivative at .
Proof.
Note that preserves if and only if the transformation preserves , where and , whenever . Hence, we may assume without loss of generality that .
Suppose that has a finite angular derivative at . Then the radial limit exists and . Taking the angular limit as in (1), we conclude that Im . Thus, either or .
Let us first consider the case . Since at the angular derivative of is neither nor , we know [12, p. 291] that it is actually univalent in some Stolz domain with vertex at :
for suitable and . Also, as is well known (cf. again [12, p. 291]), the function preserves angles between curves contained in that meet at . This shows that there exists a curve with and which is mapped by onto some non-horizontal segment
for an appropriate value of . (To see this, it suffices to look at the image under of the suitable Stolz domain mentioned earlier with vertex at , which will contain another Stolz domain with vertex at , and to select and so that the segment is contained in this new Stolz angle and is not contained in the real axis.) Keeping in mind that and is a Möbius transformation which maps the diameter to the positive semi-axis, we see that
Therefore, taking the limit as along in (3), we obtain , which is contrary to our construction of the segment . This completes the proof in the case when .
By (1), preserves if and only if with does, so we can argue as above in the case when to get a contradiction again. ∎
There are two ways in which the function 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 such that has a tangential contact with the unit circle. The price we pay for this is that is a dilated self-map of the disk (hence, in this example is compactly contained in ).
Example 2. For , let
Both are clearly Schwarz-type functions. Obviously, and is conformal at since . For a sufficiently large value of (which will be determined below) one can also check that our condition (1) is satisfied, hence . Indeed, it is immediate that
Checking our condition (1) in this case reduces to verifying that
holds for all in . (Note that as , both sides tend to zero but the strict inequality is maintained.) Since , it is clear that
while the right-hand side can be estimated from below as follows:
so it is only left to check that
for large enough and , which is clear. The inequality holds for all .
The natural question arises as to whether it is possible to have an example where both and 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 still preserves . The following example, illustrated by the figure below, gives an affirmative answer.
Example 3. Consider the planar domain
clearly symmetric with respect to both the real and imaginary axes. Let be a conformal map of onto which fixes the origin. Starting with the subdomain of in the upper half-plane and using the Schwarz reflection principle, one can also choose in such a fashion that it fixes the diameter and in the sense of a radial limit. Let . It can now easily be checked that our condition (1) is satisfied, hence .
Note, however, in relation to this “leaf-shaped” region that our mapping has boundary contact with the unit circle but does not have angular derivative at . The intuitive reason for this is that the corners at and 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 in polar coordinates:
Solving for , one obtains
and by elementary calculus one checks that for a fixed the integral
diverges and the conclusion follows by the Tsuji-Warschawski criterion [13, p. 72].

4. Fixed points of weighted composition transformations that preserve
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 be a weighted composition transformation such that where and are Schwarz-type functions, and is not a rotation. Then has a unique fixed point which is obtained by iterating applied to arbitrary in .
In the case when 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 applied to an arbitrary function in is a fixed point of the transformation. Define the iterations of in the usual way, being the identity function and , . Let . It is easy to see by induction that
for any integer . By our assumptions on , the origin is its only fixed point in . Since is not a disk automorphism, it follows that uniformly on compact subsets of and therefore also in the compact-open topology as . Thus, uniformly on compact subsets as . On the other hand,
so proving the uniform convergence on compact subsets of the infinite product is equivalent to proving the convergence on compact subsets of of the sums
For fixed, let . Let . Clearly, since is not a rotation. Applying the Schwarz lemma to , we get for any , and from here
whenever . Iterating this inequality, we get
for . Using the fact that is a Schwarz-type function, we obtain
in the disk . Thus, the series
converges uniformly on compact subsets of the disk and the infinite product is uniformly convergent on compact subsets to some function analytic in . Moreover, since is a compact class, . Combining both limits, we obtain
uniformly on compact subsets as for any . Now we can see that is a fixed point of the transformation. Applying to we have
Note that as constructed above does not depend on the initial choice of the function in . Now it is clear that this is the only fixed point of , because if satisfies , iterating the transformation we get
uniformly on compact subsets of .
It is only left to check our final comment in the statement of the theorem. Since is an inner function, by Proposition 3 we have so the equation for the fixed point: is the classical Shröder equation for the composition operator corresponding to the eigenvalue . By the first proposition from [13, Section 6.1], if were not identically constant, it would follow that . In view of the Schwarz lemma, this forces to be a rotation, contrary to our assumptions. ∎
The case when 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 be a weighted composition transformation such that , where , is a Schwarz-type functions, and is a rotation. Then the set of all fixed points of is as follows:
(a) all of , if ;
(b) the functions with -fold symmetry: , , whenever , where for some ;
(c) only the constant function one, if , where and for all .
Proof.
Since is an inner function, Proposition 3 forces , hence . Part (a) now follows trivially.
Parts (b) and (c) follow readily by comparing the Taylor series of both sides of the equality 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 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 , 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.