Conformal capacity of hedgehogs
Abstract.
We discuss problems concerning the conformal condenser capacity of “hedgehogs”, which are compact sets in the unit disk consisting of a central body that is typically a smaller disk , , and several spikes that are compact sets lying on radial intervals . The main questions we are concerned with are the following: (1) How does the conformal capacity of behave when the spikes , , move along the intervals toward the central body if their hyperbolic lengths are preserved during the motion? (2) How does the capacity depend on the distribution of angles between the spikes ? We prove several results related to these questions and discuss methods of applying symmetrization type transformations to study the capacity of hedgehogs. Several open problems, including problems on the capacity of hedgehogs in the three-dimensional hyperbolic space, will also be suggested.
Key words and phrases:
Conformal capacity, hyperbolic metric, hyperbolic transfinite diameter, potential function, hedgehogs, polarization, symmetrization, hyperbolic dispersion2010 Mathematics Subject Classification:
30C85, 31A15, 51M10In memoriam Jukka Sarvas (1944-2021).
Notation
-
- complex plane.
-
•
- open unit disk centered at .
-
•
- density of the hyperbolic metric in .
-
•
- hyperbolic length of .
-
•
- hyperbolic area of .
-
•
- pseudo-hyperbolic metric in .
-
•
- hyperbolic metric in .
-
•
and - closed and open Euclidean intervals with end points and .
-
•
and - closed and open hyperbolic intervals with end points and .
-
•
- conformal capacity of a compact set .
-
•
- logarithmic capacity of .
-
•
- modulus of the family of curves .
-
•
- reflection of a set with respect to the hyperbolic geodesic .
-
•
- polarization of a set with respect to the hyperbolic geodesic .
-
•
and - complete elliptic integrals of the first kind.
1. Introduction
The main theme discussed in this paper is the dependence of the condenser capacity on the geometric structure and characteristics of a compact set where is the unit disk in the complex plane . For brevity, we call the conformal capacity of or the capacity of and denote it by .
Due to the conformal invariance of the capacity it is natural to equip with the hyperbolic metric. Indeed, very recently it was shown in [30] that isoperimetric inequalities in hyperbolic metric yield simple upper and lower bounds for the capacity in the case when is a finite union of hyperbolic disks. In the subsequent work [29, 31, 32] these ideas were developed further, and it was also pointed out that similar ideas were also applied by F.W. Gehring [19] and R. Kühnau [26] fifty years earlier.
In most cases, we deal with compact sets consisting of a central body , which can be absent, and spikes , , that are closed intervals or any compact sets lying on radial intervals , where , . This type of compact sets have appeared in several research papers, for instance, in a recent paper [22] by J.-W. M. Van Ittersum, B. Ringeling, and W. Zudilin, where the term “hedgehog” was suggested for this shape of compact sets. Interestingly enough, estimates of the capacity and other characteristics of hedgehogs appeared to be useful in studies on the Mahler measure and Lehmer’s problem. Beside the above mentioned work of three authors, hedgehog structures appeared in the paper [34] by I. Pritsker and, earlier, the same hedgehog structure appeared in [39].
The hyperbolic metric in is defined by the element of length
Then the hyperbolic length of a compact subset of a rectifiable curve can be calculated as
Furthermore, the hyperbolic geodesics are circular arcs in that are orthogonal to the unit circle at their end points. The hyperbolic distance between points and in , that is equal to the hyperbolic length of the closed hyperbolic interval joining these points, is given by
(1.1) |
where is the pseudo-hyperbolic metric defined as
(1.2) |
Everywhere below, and stand, respectively, for the closed and open hyperbolic intervals with end points and . Similarly, notations and will be reserved for the closed and open Euclidean intervals with end points and . If and lie on the same diameter of then, of course, and .
When and , , the hyperbolic length of the interval and its Euclidean length are connected via the following formulas, which are often used in calculations:
(1.3) |
The hyperbolic area of a Borel measurable subset of is
(1.4) |
where stands for the -dimensional Lebesgue measure. In particular, the hyperbolic area of the disk , , is given by the following formula:
For the properties of geometric quantities defined above, we recommend the monograph of A. Beardon [9].
Our main focus in this paper will be on the quantity, which we call the conformal capacity, or just capacity that is related to the hyperbolic capacity studied in [19] and [46].
Definition 1.5.
Let be a compact set in . The conformal capacity of is defined as
(1.6) |
where the infimum is taken over all Lipschitz functions such that on the unit circle and for .
In terminology used in electrostatics, the conformal capacity is usually referred to as the capacity of the condenser with plates and and field . Therefore, many properties and theorems known for the capacity of a physical condenser can be applied to the conformal capacity as well. In Section 2, we collect some of these properties, which will be needed for the purposes of this work. Several available methods to prove the above mentioned properties also will be discussed in Section 2.
Explicit expressions for the conformal capacity of compact sets are available in a few cases only. Here are three examples, in which the conformal capacity is expressed in terms of both Euclidean and hyperbolic characteristics of the set.
Example 1.7.
The conformal capacity of the closure of the disk , with Euclidean radius , , and hyperbolic radius is given by
Example 1.8.
The conformal capacity of the interval , with hyperbolic length equal to , is
Here and below, and stand for the complete elliptic integrals of the first kind.
Example 1.9.
The conformal capacity of the interval , with hyperbolic length equal to , is
(1.10) |
As Examples 1.8 and 1.9 demonstrate, even for compact sets as simple as a hyperbolic interval, the conformal capacity can not be expressed in terms of elementary functions. Thus, estimates in terms of Euclidean characteristics of a set and numerical computations are important when working with this capacity.
This project originated with the following question raised by the third-listed author of this paper. This question arose in the course of recent work [29]-[32] and it was experimentally studied in [24].
Problem 1.11.
Suppose that and are such that the sets and have equal hyperbolic lengths. Is it true that conformal capacity of is greater than conformal capacity of ?
It appears that this question has many interesting ramifications. Thus, we decided to team up to discuss these questions, answer several of them demonstrating available technique and to point out a few remaining open questions. In the context of Problem 1.11, it is natural to consider compact sets lying on any finite number of radial intervals. This is how geometric shapes resembling animals with spikes, and therefore the term “hedgehog”, appeared in our study. Typically, the central body will be a disk , , or an emptyset and , , will be a collection of closed intervals attached to the central body. In this case, our compact sets look more like the sea creatures called “stylocidaris affinis”, that is shown in our Figure 1, than like hedgehogs as everyone knows them. But, because the term “hedgehog” was already applied in the context of our study in mathematical literature, we will stick with it in our paper.
![]() |
Our main results in Sections 3 and 4 deal with several extremal problems for the capacity of compact sets in the unit disk , where hedgehogs possessing certain symmetry properties play the role of the extremal configuration. Thus, in these sections we mainly work with compact sets in having components lying on a finite number of radial intervals. In Section 3, we first demonstrate our methods on simple cases, considered in Lemmas 3.1 and 3.4, when a compact set lies on the radial segment or on the diameter of . In particular, Lemma 3.1 provides an affirmative answer to the question stated in Problem 1.11. Then, in several theorems presented in Section 3, we extend our proofs to the case of compact sets lying on several radial intervals. In Section 4, we deal with several extremal problems on the conformal capacity for compact sets lying on a finite number of radial intervals evenly distributed over the unit disk.
As is well known, symmetrization type transformations (such as Steiner symmetrization, Schwarz symmetrization, Pólya circular symmetrization, Szegö radial symmetrization, polarization and other) provide a standard tool to estimate capacities and many other characteristics of sets. Most of the classical results on symmetrization can be found in the fundamental study by G. Pólya and G. Szegö [33]. More recent approaches to symmetrization were developed by A. Baernstein II [5], V. Dubinin [17], J. Sarvas [36] and also in the papers [37], [13], [40] and [41]. In Section 5, we will discuss hyperbolic counterparts of some of these transformations and how they can be applied in problems about conformal capacity.
Finally, in Section 6, we will mention possible generalizations of our results for conformal capacity in hyperbolic spaces of dimension .
2. Preliminary results on the conformal capacity
In this section, we recall properties of the conformal capacity needed for our work. We have already mentioned in the Introduction the connection of the conformal capacity with the condenser capacity. A condenser is a pair , where is a domain in the plane and is a compact subset of . The capacity of the condenser is defined by
(2.1) |
where the infimum is taken over all Lipschitz functions such that on and on . These functions will be called admissible for the condenser . We want to stress here that if , then the infimum in (2.1) can be taken over all admissible functions as above with an additional requirement that for all .
By Theorem 3.8 of Ziemer [50], the capacity of the condenser is equal to the modulus of the family of all curves in joining with . For the definition and the basic properties of the modulus of curve families, we refer to [23, Chapter II] and [21, Chapter 7].
The following invariance property of the conformal capacity, that we often use in our proofs below, is immediate from the well-known invariance property of the capacity of a condenser, see, for instance, [17, Theorem 1.12].
Proposition 2.2.
The conformal capacity is invariant under the Möbius self maps of and it is invariant under reflections with respect to hyperbolic geodesics. Thus, if is a compact subset of and is a Möbius automorphism or a reflection with respect to a hyperbolic geodesic, then
Similar to the “sets of measure zero” in measure theory, there are small sets that can be neglected when working with the conformal capacity. These sets are known in the literature as “polar sets” or “sets of zero logarithmic capacity”. For our purposes the second name is more appropriate and many other authors used it in a similar context. We recall that the logarithmic capacity of a set , not necessarily compact, is given by
where the supremum is taken over all Borel probability measures on whose support is a compact subset of . For the properties of the logarithmic capacity, we refer to Chapter 5 in T. Ransford book [35] and to the monographs [3], [27]. In Proposition 2.3 below we collect results identifying “sets of zero logarithmic capacity” as sets negligible for the value of the conformal capacity. For the proofs of these results we refer to H. Wallin’s paper [48].
Proposition 2.3.
The following hold:
-
(1)
If is compact, then if and only if .
-
(2)
Let , be compact sets in such that . Then if and only if .
-
(3)
If is such that , then .
We note that the set in part (2) and the set in part (3) of Proposition 2.3 are not necessarily compact. We stress here that the inverse statement for part (3) of this proposition is not true, in general. For example, if is the standard Cantor set, then its scaled version has zero hyperbolic length but positive logarithmic capacity, see, for example, [35, p.143]. A compact set of logarithmic capacity zero is of zero Hausdorff dimension [48].
Next, we will state a proposition about the existence of a function minimizing the Dirichlet integral in equation (2.1), and therefore in equation (1.6) as well. This proposition follows from classical potential theoretic results; see, for instance, [7, Theorem 1], [27, p.97].
Proposition 2.4.
Let be a compact set in . There is a unique function , called the potential function of , that minimizes the integral in (2.1); i.e. such that
Moreover, possesses the following properties:
-
(1)
is harmonic in and continuous on , except possibly for a subset of of zero logarithmic capacity.
-
(2)
for and for all , except possibly for a subset of of zero logarithmic capacity.
-
(3)
If every point of is regular for the Dirichlet problem in , then is continuous on and on .
Below are two examples of sets exceptional in the sense of parts (1) and (2) of Proposition 2.4.
Example 2.5.
Consider a compact set , where is an interval with length . Since
Wiener’s criterion (see [35, Theorem 5.4]) implies that the point is irregular for the Dirichlet problem. Therefore, the function is not a barrier at (see [35, Definition 4.1.4]). The latter implies that the limit does not exist. Since points , , are regular for the Dirichlet problem, is the only point in , where is not continuous.
To obtain a compact set such that has infinite number of discontinuities, we modify our previous example as follows. For , let . Thus, is obtained by translating and scaling the set . Let . Our previous argument can be applied to show that is not continuous at an infinite set of points , . Thus, has an infinite subset exceptional in the sense of part (1) of Proposition 2.4.
Example 2.6.
Let be a compact set of positive logarithmic capacity, which contains a nonempty subset , each point of which is isolated from other points of . Since is harmonic and bounded, every point is removable, which means that can be extended as a function harmonic at . Since is not constant, it follows that for every point . Thus, is an exceptional set, possibly infinite, as it was mentioned in part (2) of Proposition 2.4.
Next, we recall a subadditivity property of the conformal capacity, which we need in the following form.
Proposition 2.7.
Suppose that is the union of compact sets , , in . Then
(2.8) |
Moreover, if each has positive conformal capacity, then (2.8) holds with strict inequality.
Proof. Let , , be an admissible function for the condenser such that for . Then is an admissible function for the condenser such that for all . The latter inequality implies that for all , where the gradients exist. Therefore,
(2.9) |
Taking the infimum in this equation over all admissible functions , , with the properties mentioned above in this proof, we obtain the required subadditivity property (2.8).
If , then each of the condensers has the potential function that is a non-constant harmonic function in . Therefore, , almost everywhere in . Since is the potential function of it follows that (2.9) holds with . In this case, the strict inequality holds for all points in an annulus with such that . The latter implies that if then the third inequality in (2.9) is strict. Therefore, (2.8) holds with the sign of strict inequality in the case under consideration.
The proofs of our main theorems in Sections 3 and 4 rely on the polarization technique and on the geometric interpretation of the conformal capacity in terms of the hyperbolic transfinite diameter.
To define the polarization of compact sets with respect to a hyperbolic geodesic , we need the following terminology. To any hyperbolic geodesic , we can give an orientation by marking one of its complementary hyperbolic halfplanes and call it ; then the other complementary hyperbolic halfplane is given the name . Since every hyperbolic geodesic is an arc of a circle, we can define the classical symmetry transformation (called also inversion or reflection) with respect to . We note that the symmetry transformation with respect to a hyperbolic geodesic is a hyperbolic isometry on .
The polarization transformation of compact sets in can be defined as follows.
Definition 2.10.
Let be an oriented hyperbolic geodesic in . Let be the hyperbolic halfplanes determined by . Let be a compact set in and let denote the set symmetric to with respect to . The polarization of with respect to is defined by
(2.11) |
Equation (2.11) can be also written in the form
The polarization transformation was introduced by V. Wolontis in 1952, [49]. Wolontis’ work remained unnoticed until 1984, when V. N. Dubinin used this transformation to solve A. A. Gonchar’s problem on the capacity of a condenser with plates on a fixed straight line interval. The name “polarization” was also suggested by Dubinin [16]. The following proposition, describing the change of the conformal capacity under polarization, is an important ingredient of the proofs in the following sections, see [17, Theorem 3.4], [12, Theorem 2.8].
Proposition 2.12.
Let be a compact set in and be the polarization of with respect to an oriented hyperbolic geodesic . Then
(2.13) |
Furthermore, equality occurs in (2.13) if and only if coincides with up to reflection with respect to and up to a set of zero logarithmic capacity.
When working with hedgehog structures, the following particular case of Proposition 2.12 is useful.
Corollary 2.14.
One more useful characteristic of compact sets in the hyperbolic plane, the hyperbolic transfinite diameter, was introduced by M. Tsuji [46]. It is defined as follows (see [46, p.94] or [17, Section 1.4]).
Definition 2.15.
Let be a compact set in . The hyperbolic transfinite diameter of is defined as
(2.16) |
where stands for the pseudo-hyperbolic metric defined by (1.2) and the maximum is taken of all -tuples of points in .
The following relation was established in [46].
Proposition 2.17.
Let be a compact set in . Then
(2.18) |
Let be a compact set in . A map is called a hyperbolic contraction on if for every ,
where stands for the hyperbolic metric defined by (1.1). Furthermore, is called a strict hyperbolic contraction on if there is , , such that for every ,
Proposition 2.19.
Let be a compact set in . Let be a hyperbolic contraction. Then .
Moreover, if is a strict hyperbolic contraction on , then .
Remark 2.20.
We note here that the polarization transformation is not contracting, in general. For example, polarizing the set with respect to the diameter with its standard orientation, we obtain the polarized set . Then for every one-to-one map , there is a pair of points such that ; one can easily verify this inequality by considering each one of the possible maps .
To study the limit behavior of the conformal capacity , when some of the components of tend to the boundary of , we need a hyperbolic analog of the dispersion property of the Newtonian capacity discussed in [42]. Let be disjoint nonempty compact sets in , not necessarily connected, and let .
Definition 2.21.
By a hyperbolic dispersion of we mean a mapping satisfying the following properties:
-
(1)
For each , the restriction is a rigid hyperbolic motion of , which depends continuously on the parameter , such that for all .
-
(2)
If , then for each and , , the hyperbolic distances between the images and satisfy the following inequalities:
-
(3)
For each and , ,
Thus, hyperbolic dispersion of is a process moving the subsets farther and farther from each other, resembling the scattering of galaxies of our Universe.
We stress here that not every finite collection of compact sets admits hyperbolic dispersion. For example, the set , where and with and sufficiently small , cannot be hyperbolically dispersed in the sense of Definition 2.21. On the other hand, any union of two non-intersecting compact sets, each of which lies on a radial interval, can be hyperbolically dispersed.
The following useful result is a hyperbolic counterpart of Proposition 5 proved in [42].
Proposition 2.22.
Let be a hyperbolic dispersion of a compact set , as above. Then
(2.23) |
In the proof of Proposition 2.22, we will need the following elementary arithmetic result.
Lemma 2.24.
Let , , be such that . Then there are sequences of positive integers , , such that if , then and as .
Proof. Consider rational approximations of , ; i.e. consider sequences
where , are positive integers. Then consider the sequence
and the sequences
Clearly,
Also, we put
Then
The latter relation shows that for all sufficiently large. Therefore, we can remove a finite number of terms from the sequence and from the sequences and then re-enumerate these sequences to obtain sequences with the required properties.
Proof of Proposition 2.22. For and , we set , . Since the conformal capacity is invariant under hyperbolic motions, for all and all . This together with the subadditivity property of Proposition 2.19 implies that
(2.25) |
We assume without loss of generality that for all . Then we set and will use the sequences , , and defined as in the proof of Lemma 2.24 for our choice of .
Let , , be points in such that
(2.26) |
where the maximum is taken over all -tuples of points in . For , , , and , we set . Since is a hyperbolic motion on each , we have
(2.27) |
for all points , defined above and all .
For our choice of points, it follows from (2.27) and equation (2.16) of Definition 2.15 that
(2.28) |
where
(2.29) |
and
(2.30) |
where the product in (2.30) is taken over all pairs of points , such that , and .
Using equations (2.16), (2.26), and (2.29) and taking into account our choice of points , and the limit relation , we conclude that
(2.31) | ||||
Our assumption that when and and relations (1.1), (1.2), imply that for every there exists such that if then
This inequality together with (2.30) imply that
(2.32) |
Combining (2.28), (2.31), and (2.32), we obtain the following:
The latter inequality together with (2.18) implies that for all ,
(2.33) | ||||
Since can be chosen arbitrarily small, it follows from (2.33) that
(2.34) |
Remark 2.35.
We note here that the conformal capacity of a compact set is not monotone under hyperbolic dispersion, in general.
To give an example of such non-monotonicity, we consider a family of hedgehogs , , with consisting of a fixed central body , , , and single varying spike , which we define as follows.
We put , , where . Using polarization with respect to appropriate hyperbolic geodesics , we find that strictly decreases, when varies from to . At the same time and are constant for . Using these properties and the well-known convergence result, which is stated in Proposition 2.36 below, we conclude that there is a strictly increasing function such that , , and a function , , such that strictly decreases, strictly increases, while the hyperbolic length remains constant on .
Now, we put for and, for , we define as with and such that . The family of hedgehogs defines a dispersion of compact sets and such that strictly decreases on . Furthermore, using polarization with respect to appropriate geodesics, as we will demonstrate it later in the proof of Lemma 3.4, one can show that strictly increases on the interval .
For our proofs, we need two results on the sequences of compact sets in convergent in an appropriate sense.
Proposition 2.36 (see, [17, Theorem 1.11]).
Let , , be a sequence of compact sets in , such that for all , and let . Then
To state our next proposition, we recall that the Hausdorff distance between two compact sets in the plane is given by
The following convergence result follows from [4, Theorem 7].
Proposition 2.37.
For fixed , let , , be a sequence of compact sets on the diameter , each of which consists of a finite number of closed intervals such that the hyperbolic length of each of these intervals is . If the sequence converges in the Hausdorff metric to a compact set , then
3. Hedgehogs with geometric restrictions on the number of spikes
We start with the following monotonicity result, which, in particular, answers the question raised in Problem 1.11.
Lemma 3.1.
Suppose that and are fixed and varies in the interval . Let , , be such that . Let be a compact set consisting of a finite number of non-degenerate intervals and let . Then is a continuous function that strictly increases from to , when runs from to .
Proof. The continuity property of follows from Proposition 2.37. To prove the monotonicity of , we consider , such that and note that . Let be a hyperbolic geodesic that is orthogonal to the hyperbolic interval at its midpoint. We give an orientation to by marking its complementary hyperbolic halfplane with , see Figure 2, which illustrates the proof of this lemma. Notice that under our assumptions, and, since reflections with respect to hyperbolic geodesics preserve hyperbolic lengthes, the hyperbolic interval coincides with the reflection of with respect to . Therefore, the polarization of with respect to coincides with the set if and with the set otherwise, and the set is a non-degenerate interval and thus it has positive logarithmic capacity. Furthermore, since , it follows that the set differs from the reflection of with respect to by a set of positive logarithmic capacity. So, applying Proposition 2.12, we conclude that . Thus we proved that the function is strictly increasing. The assertion about the range of this function follows from the dispersion property of Proposition 2.22 and from the convergence property stated in Proposition 2.37.
Remark 3.2.
![]() |
Remark 3.3.
Next, we will use Lemma 3.1 to prove a lower bound for the conformal capacity of compact sets lying on the diameter of .
Lemma 3.4.
Let and let be defined as in (1.3). If is a compact set such that , then
(3.5) |
Equality occurs here if and only if coincides with some interval up to a set of zero logarithmic capacity.
Proof. (a) Suppose first that consists of intervals , . Let be the compact set obtained from by replacing the pair of intervals and with a single varying interval such that . It follows from the monotonicity property of Lemma 3.1 that . Applying this procedure of merging two intervals into a single interval times, we obtain the inequality (3.5) with the sign of strict inequality.
(b) If is a more general compact set, not the union of a finite number of intervals, we proceed as follows. Since the subset of isolated points of has zero logarithmic capacity we can remove it without changing the conformal capacity and the hyperbolic length of . Thus, we assume that does not have isolated points. Also, since both the conformal capacity and hyperbolic length are invariant under conformal automorphisms of , we may assume that .
We will use the following approximation argument. The set is an open subset of and therefore, in the case under consideration, it is a countably infinite union of open disjoint intervals , . We enumerate these intervals such that and has one of its end points at . Setting , , we obtain a sequence of compact sets , each consisting of a finite number of disjoint closed intervals on , such that for all and . So is approximated by a finite union of closed intervals. Therefore, and, by Proposition 2.36, .
Let be a closed interval on such that . Then, by part (a) of this proof,
(3.6) |
Furthermore, for all and . Thus, , by Proposition 2.36. Therefore, passing to the limit in (3.6), we obtain (3.5).
(c) Here we prove the equality statement. If coincides with some interval up to a set of zero logarithmic capacity then, by Proposition 2.3, and . This together with the conformal invariance property of the capacity implies that .
Suppose now that (3.5) holds with the sign of equality. Let and denote the infimum and supremum of the set of Lebesgue density points of . We may assume, without loss of generality, that .
If , then the set contains an open interval with . Let denote the polarization of with respect to a hyperbolic geodesic that is orthogonal to the hyperbolic interval at its midpoint and oriented such that . Notice, that under our assumptions, the set differs from by a set of positive one-dimensional Lebesgue measure, and therefore by a set of positive logarithmic capacity, and also it differs from the reflection of with respect to by a set of positive one-dimensional Lebesgue measure, and therefore by a set of positive logarithmic capacity. Hence, by Proposition 2.12,
(3.7) |
Since the polarization with respect to hyperbolic geodesics preserves hyperbolic length, we have . Therefore, it follows from the assumption and our proof in parts (a) and (b) above, that , which contradicts equation (3.7). Since the assumption that leads to a contradiction, we must have .
In the latter case, and, since , it follows from Proposition 2.3 that has zero logarithmic capacity, which completes the proof of the lemma.
Actually, our proof of Lemma 3.4 gives us a more general result, which we state as the following corollary.
Corollary 3.8.
For and , let be such that . Let be a hyperbolic geodesic passing through the point orthogonally to the diameter and oriented such that .
If is a compact subset in such that is a compact subset of and is a compact set such that , then
(3.9) |
Proof. Since the considered characteristics of compact sets are invariant under Möbius transformations, we may assume once more that . Notice that the polarization transformations used in the proof of Lemma 3.4 do not change the portion of the set , which lies in the halfspace . Therefore, our arguments used in the proof of Lemma 3.4 prove this corollary as well.
As concerns upper bounds for the conformal capacity of compact sets having fixed hyperbolic length, it is expected that there are no non-trivial upper bounds in this case. Below we present two examples which confirm these expectations.
Example 3.10.
It was shown by M. Tsuji [45] that the standard Cantor set has positive logarithmic capacity; more precisely, , see [35, p. 143]. Hence, by Proposition 2.3, the conformal capacity of the scaled Cantor set is positive.
For , let denote the image of the scaled Cantor set under the Möbius mapping , where . Since the hyperbolic length and conformal capacity are invariant under Möbius automorphisms of , we have and for all . A simple calculation shows that
This implies that the sequence of pseudo-hyperbolic distances strictly increases and as . Therefore, the sequence of hyperbolic distances also strictly increases and as . This implies that and are disjoint if and
(3.11) |
In our previous example, the hyperbolic diameters of sets tend to as . For compact sets with hyperbolic diameters bounded by some constant, say for compact sets on the interval such that , we have the strict inequality , which follows from the fact that contains a non-empty open interval that is a set of positive logarithmic capacity. In our next example, we show that for every and and are such that , , there is a compact set such that and .
Example 3.12.
First, we consider a condenser with the domain , where , , and a compact set , where , , .
Let denote the family of curves in joining the boundary circles of with the set and let denote the family of curves in the annular sector joining the arc with the set . It follows from Ziemer’s relation between the capacity of a condenser and the modulus of an appropriate family of curves (see [50, Theorem 3.8]) and from symmetry properties of the modulus of family of curves (see [1, Theorem 4] for an equivalent form of this symmetry property given in terms of the extremal length)that
(3.13) |
To find , we consider a function with
(3.14) |
where the parameter , , of the elliptic sine function is defined by the equation
(3.15) |
It is a well-known property of elliptic integrals [2, Theorem 5.13(1)] that if , then and the following expansion holds:
(3.16) |
From (3.15) and (3.16), we obtain that
(3.17) |
where when .
The function maps conformally onto the semidisk with . Furthermore, this function maps the arc onto an interval with . To find , we note that as and that converges to uniformly on compact subsets of as . Using these relations and equations (3.14), we find that and, therefore, we have the following asymptotic formula for the logarithmic capacity of the interval :
(3.18) |
where as .
Let denote the family of curves in joining the circle with the interval . Conformal invariance and symmetry properties of the modulus of family of curves imply that
(3.19) |
We have the following limit relation between the modulus of and the logarithmic capacity of :
(3.20) |
where as .
Finally, combining equations (3.13), (3.19) and (3.21), we obtain the following asymptotic formula for the capacity of the condenser :
(3.22) |
where when and are fixed and .
Let , , be such that
(3.23) |
Then there is a unique function mapping conformally onto such that . This function can be extended to a function continuous on and such that for all . Thus, maps onto a compact set .
The same conformal invariance and symmetry properties, which we used earlier in this example, together with equations (3.22) and (3.23) imply that
Notice that there is a constant such that for all . This implies that for every there is , , such that for every , , and all .
We fix , , and and consider a compact set such that and , Then, if is large enough, we have
The latter equation shows that for every , , and every , there is a compact set such that and . Since the hyperbolic length and conformal capacity are invariant under Möbius automorphisms of , compact sets with similar properties exist for every interval , .
Remark 3.24.
Our construction of a compact set in Example 3.12 is similar to the construction used in [38] to provide a counterexample for P.M. Tamrazov’s conjecture on the capacity of a condenser with plates of prescribed transfinite diameters. In turn, a counterexample used in [38] is based on the following result, that is example 5) in [28, Ch. II, §4]:
Let denote the orthogonal projection of the set introduced earlier onto the real axis. Then
Taking the limit in this equation as , we conclude that for every and every , there exists a compact set with Euclidean length such that
In particular, this answers a question raised in Problem 2 in [10] by showing that the supremum of the transfinite diameters of compact sets with Euclidean length , , is equal to .
As Examples 3.10 and 3.12 show, there are no upper bounds for the capacity expressed in terms of the hyperbolic length of a compact set , in general. In a particular case, when is connected, a non-trivial upper bound exists and is given in the following lemma.
Lemma 3.25.
Let be a Jordan arc having the hyperbolic length . Then
(3.26) |
Proof. The proof repeats the well-known proof for the logarithmic capacity, see, for example, [35, Theorem 5.3.2]. We consider a parametrization of by the hyperbolic arc-length. Then is contractive in the hyperbolic metric. Therefore, (3.26) follows from Proposition 2.19.
The polarization technique used in the proof of Lemma 3.1 can be applied in a more general situation as we demonstrate in our next theorem.
Theorem 3.27.
Consider distinct radial intervals , , of the unit disk . Suppose that , , is a compact set on such that and that is a hyperbolic interval having one end point at such that .
If each of the angles formed by the radial intervals and , , is greater than or equal to , then
(3.28) |
Equality occurs in (3.28) if and only if for each , coincides with up to a set of zero logarithmic capacity.
Proof. The proof is the same for all . Thus, we assume that . Rotating, if necessary, we may assume that . The diameter is a hyperbolic geodesic, which we orient such that . If , then the angles between the neighboring intervals are equal to and therefore the set can be represented as the union with . This shows that the sets , and satisfy the assumptions of Corollary 3.8. Therefore, by this corollary,
with the sign of equality if and only if coincides with up to a set of zero logarithmic capacity.
The same argument can be applied successively to the sets , , and to obtain the inequalities
Moreover, equality occurs in any one of these inequalities if and only if the corresponding sets and coincide up to a set of zero logarithmic capacity. Thus, the theorem is proved.
Remark 3.29.
If the angle between some radial intervals and is smaller than , then both our proofs, with polarization or with the contraction principle, fail even in the simplest case of two intervals and and when each of the sets and is a hyperbolic interval. However, the graphs of the results of numerical experiments performed by Dr. Mohamed Nasser, which are presented in Figure 3, suggest that the monotonicity property of the conformal capacity of two intervals remains in place for all angles. Therefore, we suggest the following.
Problem 3.30.
Given fixed , , and and varying , let with such that . Prove (or disprove) that strictly increases on the interval .
![]() |
As one can see from our proof of Theorem 3.27, the restriction on the number of radial intervals and restriction on angles between them is needed because otherwise polarization with respect to hyperbolic geodesics may destroy the radial structure of compact sets under consideration. Also, if at least one angle between radial intervals is , then the contraction principle of Proposition 2.19 can not be applied, in general. Still, under some additional assumptions on the hyperbolic lengths and angles, we have the following more general version.
Theorem 3.31.
Let , , be compact sets on the radial intervals , , such that
where stands for the minimal angle between the intervals . Then
(3.32) |
where is a hyperbolic interval having one end point at such that .
Equality occurs here if and only if for each , coincides with up to a set of zero logarithmic capacity.
Proof. (a) Let , and let be the hyperbolic geodesic orthogonal to the interval at and oriented such that belongs to the halfplane . An easy calculation shows that has its endpoints at the points . Consider sets , and . Let denote the closed hyperbolic interval with the initial point at such that and and let . Since the minimal angle between the intervals is and has its endpoints at , it follows that . Therefore, we can apply Corollary 3.8 to obtain the following:
(3.33) |
with the sign of equality if and only if coincides with up to a set of zero logarithmic capacity.
If , then and the inequality in (3.33) is equivalent to the inequality
(3.34) |
with the sign of equality if and only if coincides with up to a set of zero logarithmic capacity.
(b) If , then contains an open interval. In this case . Let , , denote the end point of the interval and let denote the midpoint of the hyperbolic interval . Notice that under our assumptions . Let be the hyperbolic geodesic orthogonal to the diameter at and oriented such that . Let and denote polarizations of the corresponding sets with respect to . Since and therefore , we have and . Applying Proposition 2.12 and using (3.33), we conclude that
(3.35) |
with the sign of strict inequality in the first inequality because differs from and from its reflection with respect to by an open interval and therefore by a set of positive logarithmic capacity.
Since and , using (3.35) and applying the same arguments as in part (a) of this proof to the set , we conclude that in the case the inequality (3.34) remains true with the same statement on the equality cases.
(c) Now, when (3.34) is proved in all cases, we can apply the iterative procedure as in the proof of Theorem 3.27 to conclude that the inequality (3.32) holds with the sign of equality if and only if, for each , coincides with up to a set of zero logarithmic capacity.
For compact sets and , lying on two orthogonal diameters of , we have the following result.
Theorem 3.36.
Let , be compact sets and let , , and be such that
Then
(3.37) |
where is such that .
Equality occurs in the first inequality if and only if coincides with and coincides with up to a set of zero logarithmic capacity. Equality occurs in the second inequality if and only if .
Proof. Let , , be such that the following holds:
Then, by Theorem 3.27,
(3.38) |
with the sign of equality if and only if the sets in the left and right sides of this inequality coincide up to a set of zero logarithmic capacity.
Suppose that , say . Then . Let denote polarization with respect to the geodesic that is orthogonal to the hyperbolic interval at its midpoint , . We assume here that is oriented such that . Under our assumptions, . Since the set has positive logarithmic capacity, it follows from Proposition 2.12 that
(3.39) |
The same polarization argument can be applied to show that, if , then
(3.40) |
Combining inequalities (3.38)–(3.40), we obtain the first inequality in (3.37) with the sign of equality if and only if coincides with and coincides with up to a set of zero logarithmic capacity.
To prove the second inequality in (3.37), we use the conformal mapping
from the doubly connected domain onto with
We note that conformal mappings preserve capacity of condensers and that the function is strictly increasing on the interval . Furthermore, it follows from formulas (1.3) that the sum of the hyperbolic lengths defined in the theorem is constant if and only if the following product is constant:
(3.41) |
where is constant.
Introducing new variables , and , we can express in terms of these variables as follows:
which we have to minimize under the constraint . Differentiating, we find that
Therefore, takes its minimal value when is as small as possible. By the classical arithmetic-geometric mean inequality , unless . Therefore, the minimal value of under the constraint occurs when . The latter implies that with the sign of equality if and only if . This proves the second inequality in (3.37).
Remark 3.42.
The inequality obtained in Theorem 3.36 is stronger, in general, than the inequality obtained by the classical Steiner symmetrization, which will be discussed in Section 5. To give an example, we consider two sets: and . Two Steiner symmetrizations, chosen appropriately, transform these sets to the sets and , respectively. The first inequality in (3.37) compares the conformal capacities of and with the conformal capacities of the sets
Numerical computation gives the following approximation and bounds for , ,
In our previous lemmas and theorems of this section, the extremal configurations were hedgehogs with spikes issuing from the central point. Similar results for compact sets with spikes emanating from a certain compact central body sitting in the disk , , as it is shown in Figure 4, also may be useful in applications.
Theorem 3.43.
Let be a compact set in the disk , . Let , , be compact sets on the radial intervals , , such that
where stands for the minimal angle between the intervals . Then
(3.44) |
where and is a hyperbolic interval having one end point at such that , .
Equality occurs in (3.44) if and only if for each , coincides with up to a set of zero logarithmic capacity.
![]() |
Proof. The proof of this theorem is essentially the same as the proof of Theorem 3.31. The only new thing we need is the following observation. Let , , be such that
and let be the middle point of the hyperbolic interval . Then every polarization performed with respect to the geodesic that is orthogonal to the radial interval and crosses it at some point such that , and such that , satisfies the following properties:
Therefore, applying polarizations successively and arguing as in the proof of Theorems 3.27 and 3.31, we obtain inequality (3.44) together with the statement on the equality in it.
Most of the results presented above in this section can be proved by two methods, either using the polarization technique or the contraction principle. Now, we give an example of a result, when polarization does not work but the contraction principle is easily applicable.
Theorem 3.45.
Let be a compact subset of and let be an upper semicontinuous function on . For , let be a compact set in such that the intersection is empty if and it is an interval with , , such that , if .
Then the conformal capacity is an increasing function on the interval .
Proof. Let , , be fixed and let . Consider the function defined as follows: if , then and . We claim that, if , then is a hyperbolic contraction on .
To prove this claim, we fix two points and consider the hyperbolic distance . Our claim will be proved if we show that this hyperbolic distance or, equivalently, the pseudo-hyperbolic distance
(3.46) |
is an increasing function of on the interval .
Rotating, if necessary, we may assume that
Then
where , are functions of .
Consider the function . After some algebra, we find that
(3.47) |
Since for fixed and , , , the hyperbolic distance between the points and is the same for all in the interval , it follows that
(3.48) |
Differentiating (3.47) and using formulas (3.48), we find:
(3.49) |
where is the following function:
It is clear that is an increasing function of on the interval . After simple calculation, we find that and therefore
This, together with (3.49), implies that and therefore the pseudo-hyperbolic distance in equation (3.46) decreases, when decreases from to . Therefore, our claim that is a hyperbolic contraction is proved. By Proposition 2.19, the latter implies that decreases when decreases from to , which proves the theorem.
4. Extremal properties of hedgehogs on evenly distributed radial intervals
In this section we consider problems with extremal configurations lying on radial intervals , . Since the intervals are evenly distributed in it is expected that extremal configurations possess rotational symmetry by angle .
First, we prove a theorem that generalizes the second inequality of Theorem 3.36 for sets lying on diameters.
Theorem 4.1.
Let , and let , , be such that
(4.2) |
and, for , let
Then
(4.3) |
Proof. To establish (4.3), we use the function , where , , which maps the sector conformally onto slit along the interval . Using symmetry properties of , we find that
One way to prove the monotonicity property of is to differentiate the function in (4.3) and check if its derivative is positive. Here, we demonstrate a different approach, which may be useful when an explicit expression for the derivative is not known.
The proof presented below is similar to the proof of the second inequality in (3.37) in Theorem 3.36. We use the function
to map the domain conformally onto the unit disk slit along the interval , where is defined as
Since possesses -fold rotational symmetry about , it follows from the symmetry principle for the module of family of curves, that
Therefore, to minimize under the constraint (4.2), we can minimize under the same constraint.
Using the variables , , constrained by the condition , we express as follows:
To minimize under the constraint , we introduce Lagrange’s function
Differentiating this function, we find
(4.5) |
where is defined as follows:
It follows from equation (4.5) that if is a critical point of the minimization problem under consideration, then . In this case
Since , , and there is only one critical point of the minimization problem under consideration and since , it follows that achieves its minimal value either at the point or at the point . In the latter case, we have
The inequality is equivalent to the following inequality:
Using binomial expansion, this inequality can be written as
Since for all and , the latter inequality holds true.
Thus, takes its minimal value when and therefore the inequality (4.4) is proved.
Remark 4.6.
The binomial inequalities similar to the one used in the proof of Theorem 4.1 are known to the experts. To prove that , we can argue as follows.
Let and . If , then . Suppose that for . Then
Now the required inequality follows by induction.
It was shown in the proof of Theorem 4.1 that there is only one critical point in the minimization problem considered in that theorem. Therefore, the following monotonicity result is also proved.
Corollary 4.7.
Since the conformal capacity is conformally invariant, the conformal capacity of an interval on remains constant when moves along the diameter so that its hyperbolic length is fixed. For intervals situated on equally distributed radial intervals the latter property is not true any more but, as our next theorem shows, if all these intervals have equal hyperbolic lengths and move synchronically, the conformal capacity of their union changes monotonically. Actually, the non-strict monotonicity is already established in Theorem 3.45. Thus, our intention here is to prove the strict monotonicity result and relate it with certain properties of relevant transcendental functions.
Theorem 4.8.
Let be a closed subinterval of with the initial point at and hyperbolic length . For , let . Then the conformal capacity is given by
(4.9) |
where
(4.10) |
Furthermore, strictly increases from with to with , when varies from to .
Proof. As in the proof of Theorem 4.1, we use the function , where , . Then maps the sector conformally onto slit along the interval . Furthermore, maps onto the interval with defined as in (4.10). Using symmetry properties of , we find that
As we have mentioned in the proof of Theorem 4.1, we know two approaches to prove the monotonicity property of the conformal capacity in that theorem. The same approaches can be used to prove the monotonicity statement of Theorem 4.8. Here, we demonstrate one more approach, which also may be useful when an explicit expression for the derivative is not known. We first note that is an analytic function of . This follows from equation (4.9). Since is not constant and analytic, it follows that is not constant on any subinterval of . Furthermore, it follows from Theorem 3.45 that is a non-decreasing function. Since it is non-decreasing and not constant on any interval, it is strictly increasing.
Our next theorem can be considered as a counterpart of the subadditivity property of the conformal capacity discussed in Proposition 2.7.
Theorem 4.11.
Let , , be compact sets on the interval having positive logarithmic capacities and such that every point of is regular for the Dirichlet problem in , . Then
(4.12) |
Equality occurs in the first inequality if and only if for each and , coincides with up to a set of zero logarithmic capacity.
Proof. The second inequality is just the subadditivity property of the conformal capacity stated in Proposition 2.7.
To prove the first inequality, we use the method of separation of components of a condenser in the style of Dubinin’s paper [15]. Let denote the potential function of the condenser . Since every point of is regular for the Dirichlet problem, it follows that is continuous on . Let and denote the functions obtained from , first by restricting , respectively, onto the sector or onto the sector and then extending this restriction by symmetry on the whole unit disk. Then each of the functions and is admissible for the condenser . Furthermore, each of these functions possesses -fold rotational symmetry about and is symmetric with respect to the real axis. Therefore, the following inequality holds:
(4.13) |
Summing up all the inequalities in (4.13), we obtain the first inequality in (4.12).
Furthermore, since every point of the sets , , is regular for the Dirichlet problem, it follows that or defined above in the proof is a potential function of if and only of for all . Therefore, if for some , then we have the strict inequality in (4.13) and in the first inequality in (4.12) as well.
Above we discussed results on the conformal capacity of compact sets lying on a fixed number of radial intervals. In our next theorem, we work with compact sets on radial intervals that are “densely spread” over in the sense that the angle between any two neighboring intervals is .
Theorem 4.14.
Let , , and let be a compact set of hyperbolic length . Let be such that , , . Consider compact sets , , and , where is such that and . Then
(4.15) |
Equality in the first inequality occurs if and only if .
Proof. The first inequality in (4.15), together with the statement on the equality cases, follows from Theorem 5 in [6]. Then the second inequality follows from the contraction principle stated in Theorem 3.45.
Polarization and the contraction principle are customarily applied when lower bounds for the conformal capacity are needed. In our next theorem, we present a result with an upper bound for this capacity.
Theorem 4.16.
For , let and let be a compact set of hyperbolic length , . Let . Let
Then
(4.17) |
Equality in the first inequality in (4.17) occurs if and only if , .
5. Symmetrization transformations in hyperbolic metric
In this section, we discuss possible counterparts of classical symmetrization-type transformations applied with respect to the hyperbolic metric. We note that Steiner, Schwarz and circular symmetrizations destroy hedgehog structures, in general, and therefore their applications to problems studied in previous sections of this paper are limited. We define symmetrizations using the following notations, which are convenient for our purposes: , , , . First, we mention results obtained with Steiner symmetrization.
Definition 5.1.
Let be a compact set. The Steiner symmetrization of with respect to the imaginary axis is defined to be the compact set
where stands for the one-dimensional Lebesgue measure.
Furthermore, the Steiner symmetrization of with respect to the line is defined to be the compact set .
For the properties and results obtained with the Steiner symmetrization, the interested reader may consult [33], [17], [5].
We note that if is a compact set in , then for all and therefore is well defined. As is well known, Steiner symmetrization does not increase the capacity of a condenser and therefore it does not increase the conformal capacity. Also, Steiner symmetrization preserves Euclidean area but, in general, it strictly decreases the hyperbolic area of that is defined by (1.4). Therefore, it is not an equimeasurable rearrangement with respect to the hyperbolic metric. Thus, to study problems on the hyperbolic plane, a version of Steiner symmetrization, which preserves the hyperbolic area and does not increase the conformal capacity, is needed. To define this symmetrization, we will use the function
which maps conformally onto the horizontal strip . We note that maps the hyperbolic geodesic onto the real axis and it maps the curves equidistant from , (which are circular arcs in joining the points and ), onto the horizontal lines , , in .
Definition 5.2.
Let be a compact set in . The hyperbolic Steiner symmetrization of with respect to the hyperbolic geodesic , centered at , is defined as
where stands for the Steiner symmetrization as in Definition 5.1.
Furthermore, the hyperbolic Steiner symmetrization of with respect to a hyperbolic geodesic , centered at the point , is defined as
where denotes the Möbius automorphism of , which maps onto the hyperbolic geodesic such that .
The hyperbolic Steiner symmetrization was introduced by A. Dinghas [14]. We note that this transformation symmetrizes sets along the hyperbolic equidistant lines and not along the hyperbolic geodesics. Previously it was used in several research papers, see, for instance, [25], [20]. In our next theorem, we collect properties of the hyperbolic Steiner symmetrization that are relevant to our study.
Theorem 5.3 (see [25]).
Let be the image of a compact set under the hyperbolic Steiner symmetrization with respect to a hyperbolic geodesic , centered at . Let be the hyperbolic geodesic orthogonal to . Then the following hold true:
-
(1)
If is a hyperbolic disk, then is a hyperbolic disk of the same hyperbolic area as having its center on .
-
(2)
is a compact set that is symmetric with respect to .
-
(3)
.
-
(4)
.
-
(5)
with the sign of equality if and only if coincides with up to a set of zero logarithmic capacity and up to a Möbius automorphism of that preserves .
Proof. We use the notation that we set in the definitions of Steiner and hyperbolic Steiner symmetrizations. We equip the strip with the hyperbolic metric induced by the hyperbolic metric in via the conformal mapping . That is, we have
So, trivially, is a hyperbolic isometry from to .
(1) Let be a hyperbolic disk in . Then is a hyperbolic disk in and is a hyperbolic disk in . It is easy to observe that hyperbolic disks in are horizontally convex (namely, their intersection with any horizontal line is either empty or a single horizontal rectilinear interval). It follows from the definition of Steiner symmetrization that is obtained by a horizontal rigid motion of and it is a hyperbolic disk in , symmetric with respect to the imaginary axis. Since both and are hyperbolic isometries and preserve symmetries, is a hyperbolic disk in , symmetric with respect to . Moreover, a horizontal motion in preserves the hyperbolic area. Therefore, .
(2) It is well known that Steiner symmetrization transforms compact sets to compact sets. So, if is compact, is compact, too. Moreover, is a set in , symmetric with respect to the imaginary axis. Hence, is a compact set in , symmetric with respect to the geodesic . It follows that is symmetric with respect to .
(3) Since is a hyperbolic isometry on , the set is a compact subset of the diameter having the same hyperbolic length as . Therefore, lies on the real axis and . The hyperbolic length (in ) on the real axis is proportional to the Euclidean length . Hence . Thus the equality in (3) follows at once.
(4) This is true because , , and Steiner symmetrization (with respect to the imaginary axis, in ) preserve hyperbolic areas.
(5) The conformal maps and preserve the capacity of condensers. So, to prove the inequality in (5), it suffices to show that for every compact subset of , we have . This is a well-known symmetrization theorem [17, Theorem 4.1]. The equality statement follows from [12, Theorem 1].
Remark 5.4.
Of course, any transformation bearing the name “symmetrization” must possess property (1) of Theorem 5.3; otherwise it is not a symmetrization. On the other side, it was shown by L. Karp and N. Peyerimhoff in [25] and by F. Guéritaud in [20] that even in the best case scenario the hyperbolic Steiner symmetrization changes hyperbolic triangles into sets that are not convex with respect to hyperbolic metric and therefore these sets are not hyperbolic triangles. In fact, the image of the hyperbolic triangle with vertices , , , and , , , is a proper subset of the hyperbolic isosceles triangle with vertices , , and .
In relation to Theorem 3.43 we suggest the following problem, where two hyperbolic Steiner symmetrizations provide some qualitative information about extremal configuration but these are not enough to give a complete solution of the problem. This problem is a counterpart of the problem for the logarithmic capacity in the Euclidean plane, which was solved in [8].
Problem 5.5.
Find the minimal conformal capacity among all compact sets with prescribed hyperbolic diameter and prescribed hyperbolic area . Describe possible extremal configurations.
The Schwarz symmetrization of in the plane with respect to a point replaces by the disk centered at of the same area. The hyperbolic analog of this symmetrization is the following.
Definition 5.6.
Let be a compact set in . Then its hyperbolic Schwarz symmetrization with respect to is the hyperbolic disk, we call it , centered at and such that .
The hyperbolic Schwarz symmetrization was first suggested by F. Gehring [19] and later used in [18] and [11]. In particular, the following result was proved.
Theorem 5.7 (see [19],[18],[11]).
If is the hyperbolic Schwarz symmetrization of a compact set , then
with the sign of equality if and only if coincides with up to a set of zero logarithmic capacity and up to a Möbius automorphism of .
Next, we will discuss the hyperbolic circular symmetrization of with respect to the hyperbolic ray , , . This is defined as the image of the interval under a Möbius automorphism of such that , .
Definition 5.8.
Let be a compact set in . Then its hyperbolic circular symmetrization with respect to the hyperbolic ray is a compact set such that: (a) if and only if , (b) for , if and only if , (c) if, for , , then is a closed circular arc on (which may degenerate to a point or may be the whole circle ) centered at such that .
Furthermore, the hyperbolic circular symmetrization of with respect to a hyperbolic geodesic ray is defined as
(5.9) |
where denotes the Möbius automorphism of , which maps onto the interval .
Remark 5.10.
We immediately admit here that the hyperbolic circular symmetrization with respect to is exactly the classical circular symmetrization with respect to the positive real axis. The reason for this is the fact that the hyperbolic density is constant on circles centered at . Thus, this transformation will not provide any new information that is not available via classical symmetrization methods. However, the variant in formula (5.9) can be used to obtain additional information that may be rather interesting.
It follows from Definition 5.8 that the hyperbolic area and conformal capacity do not depend on . For the Euclidean area of we have the following result.
Lemma 5.11.
Let be a compact set of positive area, and let denote the Euclidean area of , , , considered as a function of . If does not coincide with a hyperbolic disk centered at up to measure , then is strictly increasing in on the interval ; otherwise is constant for .
Proof. Let denote the circular symmetrization of with respect to , with defined as in Definition 5.8. Then , where
with chosen such that
The function strictly increases from to , when increases from to . Therefore, to prove monotonicity of a certain characteristic of as a function of , we can consider as a function of and treat as a function of . Thus, we will work with the function .
The Euclidean area can be found as follows:
(5.12) |
where is defined by the condition .
If , then is a non-degenerate proper arc on that is centered at . Using this symmetry and the monotonicity property of the function
we conclude that the inner integral in (5.12), that is the integral
is a strictly increasing function of , . Therefore, if differs by positive area from the disk centered at , which has the same area as , then the double integral in the right-hand side of (5.12) strictly increases on the interval . This proves that the area strictly increases on the interval . Of course, if coincides with a hyperbolic disk centered at up to measure , then is constant for .
Now we add two results in the spirit of our theorems on the conformal capacity of hedgehogs proved in previous sections. Figure 5 illustrates proofs of these results, which are given in Theorems 5.13 and 5.17 below. Examples of an admissible set and an extremal set of Theorem 5.13 are shown in parts (a) and (b) of Figure 5 and example of an admissible and extremal sets of Theorem 5.17 are shown in parts (c) and (d) of this figure.
Theorem 5.13.
For and , let and let be a compact set such that the length of its radial projection on the circle is . Then
(5.14) |
with the sign of equality if and only if coincides with up to a set of zero logarithmic capacity and up to a rotation about .
Proof. Consider a function . It can be easily shown that is a hyperbolic contraction from onto . Hence, by Proposition 2.19,
(5.15) |
As well known, the circular symmetrization decreases the conformal capacity, precisely, the following holds:
(5.16) |
with the sign of equality if and only if coincides with up to a set of zero logarithmic capacity and a rotation about . Combining (5.15) and (5.16), we obtain (5.14) together with the statement about cases of equality in it.
Theorem 5.17.
For , , and , let be a compact set such that and , where stands for the circular projection of onto the radius . Then
(5.18) |
with the sign of equality if and only if coincides with up to a set of zero logarithmic capacity and up to a rotation about .
Proof. Performing the circular symmetrization as in the proof of Theorem 5.13, we obtain the following inequality:
(5.19) |
with the sign of equality if and only if coincides with up to a set of zero logarithmic capacity and up to a rotation about .
If does not coincide with the interval up to a set of zero logarithmic capacity, then we can perform polarization transformations as in the proof of Lemma 3.4 to transform into the interval . Furthermore, our assumption that guarantees that these polarizations do not change . Hence, we obtain the inequality
(5.20) |
with the sign of equality if and only if coincides with up to a set of zero logarithmic capacity. Combining (5.19) and (5.20), we obtain (5.18) together with the statement about cases of equality in it.
The restriction in the previous theorem is due limitations of methods used in this paper. In relation with this theorem, we suggest two problems.
Problem 5.21.
Find over all compact sets satisfying the assumptions of Theorem 5.17 with .
Problem 5.22.
Let , , be a compact set such that its radial projection on the unit circle coincides with and its circular projection onto the radius coincides with the interval , as it is shown in Figure 6, and let . Is it true that ? If this is not true then identify the shape of minimizing the conformal capacity under the assumptions of this problem.
Remark 5.23.
Let denote the hyperbolic geodesic tangent to the circle . We assume that touches at the point , . Let denote the arc of with the endpoints , and let . It can be shown with the help of polarization that if minimizes the conformal capacity in Problem 5.22, then belongs to the compact set bounded by the arcs , and the circular arc .
![]() |
Now we turn to the hyperbolic version of the radial symmetrization which in the Euclidean setting was introduced by G. Szegö [44]. To define this radial symmetrization, we need the following notation. For , , and a compact set , define .
Definition 5.24.
Let be a compact set in such that , . Then the radial symmetrization of with respect to is the compact set with the property: for every , is a radial interval such that
(5.25) |
Equation(5.25) shows that Szegö’s radial symmetrization is a rearrangement that is equimeasurable with respect to the logarithmic metric. It found important applications to several problems in Complex Analysis and Potential Theory. However, because of the usage of the logarithmic metric, its applications to the hedgehog problems studied in this paper are rather limited. Indeed, certain compact sets on the interval having a big hyperbolic length are transformed by this symmetrization to the radial intervals with a very small hyperbolic length.
In the search for better estimates for the studied characteristics of compact sets in , we turned to the following version.
Definition 5.26.
The hyperbolic radial symmetrization of a compact set with respect to is defined to be a compact set starlike with respect to and such that
We recall that the compact sets in Theorems 3.27, 3.31, and 3.43 having the minimal conformal capacity among all sets admissible for these theorems can be obtained via the hyperbolic radial symmetrization defined above. Also, the graphs in Figure 3 of the results of numerical computations suggest that the hyperbolic radial symmetrization of two radial intervals reduces the conformal capacity of these intervals. Therefore, the following conjecture sounds plausible.
Problem 5.27.
Let be a compact set in . Prove (or disprove) that
We conclude this section with the following result, which describes how the hyperbolic area and hyperbolic diameter of a compact set behave under the hyperbolic radial symmetrization.
Lemma 5.28.
The hyperbolic area and the hyperbolic diameter do not increase under the hyperbolic radial symmetrization of .
Proof. First we deal with the hyperbolic area of a compact set in . Since every compact set in can be approximated by a finite union of polar rectangles, we may assume that is a finite union of sets of the form
where and . For such an , the hyperbolic radial symmetrization is again a union of sets of the same form, and moreover, each of the polar rectangles of is obtained by moving a rectangle of radially towards the origin, keeping its hyperbolic height fixed. More precisely, if (as defined above) is a rectangle of , then the corresponding rectangle of has the form
with and
(5.29) |
We express as a function of using (5.29) and find that
(5.30) |
Another elementary calculation gives
We consider the function
By straightforward differentiation of and taking into account (5.30), we find that is strictly increasing. This implies that . Summing up over the hyperbolic areas of all rectangles, we conclude that .
Next we turn to hyperbolic diameter, which we denote by . We will use the following basic fact of the hyperbolic geometry.
(a) Given a hyperbolic geodesic and a point , there is a unique hyperbolic geodesic orthogonal to . Let intersect at . Then the hyperbolic distance strictly increases as moves along from to .
To prove this result, we may assume that and , . Then and the monotonicity of for follows after simple calculations.
Furthermore, if and , , then intersects at the point such that .
We will also use the following result that can be checked by standard Calculus technique.
(b) Let , , and , . Then . Furthermore, the pseudo-hyperbolic distance , and therefore the hyperbolic distance , is a strictly increasing function of , and a strictly increasing function of , .
Now let be two points on such that . Note that and . We may assume that , and or , , . If , , then
and the required result is proved.
Next, we assume that with , . The set contains points and with , such that , . If , then it follows from the monotonicity property stated in (a) that and therefore in this case.
Now, we consider the case when , . In this case, the hyperbolic geodesic orthogonal to crosses at the point with . If , then
where the first inequality follows from the monotonicity property stated in (a).
Thus, we are left with the case , . We work with the points and defined above. If , then using twice the monotonicity property stated in (a) we obtain the required result:
If , , then
where the first and the third inequalities follow from the monotonicity property stated in (a) and the second inequality follows from the monotonicity property with respect to stated in (b). Now, the inequality is proved in all cases.
6. Hedgehog problems in
In this section we briefly mention how some of our results proved in the previous sections can be generalized for compact sets in the ball in . Here, , . The conformal capacity of a compact set is defined as
where stands for the three-dimensional Lebesgue measure and the infimum is taken over all Lipschitz functions such that on and on .
The conformal capacity of is just the conformal capacity of the condenser and therefore it is invariant under Möbius automorphisms of , see [19]. It is also known that the conformal capacity does not increase under polarization, see [5]. However, it is an open problem to prove that the conformal capacity is strictly decreasing under polarization unless the result of the polarization and the original compact set coincide up to reflection with respect to the plane of polarization and up to a set of zero logarithmic capacity; see, for instance, [43], where this question was discussed in the context of the Teichmüller’s problem in . This is why statements on the equality cases will be missing in all results, which we mention below.
Furthermore, the contraction principle, as it is stated in Proposition 2.19, is not available in the context of the conformal capacity in dimensions . This issue was discussed, for instance, in Section 5.6.2 in [5].
Below we list possible extensions of our results to and to spaces of higher dimension.
-
(1)
The non-strict monotonicity property in Lemma 3.1 for a finite number of intervals on the diameter remains true in any dimension .
- (2)
- (3)
- (4)
- (5)
Some of the proofs in this paper make essential use of methods only available in the planar case and therefore the corresponding result for dimensions remain open. Next we list a few of these open problems.
- (1)
-
(2)
The result stated in Theorem 3.45 easily follows from the contractions principle. But, as we have mentioned above, the contraction principle is not available to problems on conformal capacity in dimensions .
-
(3)
Theorem 4.8 can be proved by two methods: the first one uses the contraction principle and the second method uses explicit calculations. Both methods are not available in higher dimensions.
-
(4)
To prove Theorem 4.14 in the three-dimensional setting, we would need inequality (4.15) under the assumption that the central body is a ball , , and is the same as in Theorem 4.18. The proof of Theorem 4.14 is based on Baernstein’s -function method and the contraction principle. Both techniques are still waiting to be developed for the case of this type of problems in . Thus, the counterpart of Theorem 4.14 remains unproved in dimensions .
Acknowledgements. We dedicate this paper to the memory of Jukka Sarvas whose work on symmetrization is a standard reference in the potential-theoretic study of isoperimetric inequalities and symmetrization.
We are indebted to Prof. M. M.S. Nasser and Dr. Harri Hakula for kindly providing several numerical results for this paper. We also thank K. Zarvalis for his help with the figures.
We are grateful to the referee for many valuable suggestions.
References
- [1] L. V. Ahlfors, Lectures on quasiconformal mappings. Second edition. With supplemental chapters by C. J. Earle, I. Kra, M. Shishikura and J. H. Hubbard. University Lecture Series, 38. American Mathematical Society, Providence, RI, 2006. viii+162 pp.
- [2] G. D. Anderson, M. K. Vamanamurthy, and M. K. Vuorinen, Conformal invariants, inequalities, and quasiconformal maps. Canadian Mathematical Society Series of Monographs and Advanced Texts. John Wiley & Sons, Inc., New York, 1997.
- [3] D. H. Armitage and S. J. Gardiner, Classical potential theory. Springer Monographs in Mathematics. Springer-Verlag London, Ltd., London, 2001.
- [4] V. V. Aseev, Continuity of conformal capacity for condensers with uniformly perfect plates. (Russian) Sibirsk. Mat. Zh. 40 (1999), no. 2, 243-253; translation in Siberian Math. J. 40 (1999), no. 2, 205-213
- [5] A. Baernstein II, Symmetrization in analysis. With David Drasin and Richard S. Laugesen. With a foreword by Walter Hayman. New Mathematical Monographs, 36. Cambridge University Press, Cambridge, 2019. xviii+473 pp
- [6] A. Baernstein II and A. Yu. Solynin, Monotonicity and comparison results for conformal invariants. Rev. Mat. Iberoam. 29 (2013), no. 1, 91–113.
- [7] T. Bagby, The modulus of a plane condenser. J. Math. Mech. 17, 1967, 315–329.
- [8] R. W. Barnard, K. Pearce, and A. Yu. Solynin, An isoperimetric inequality for logarithmic capacity. Ann. Acad. Sci. Fenn. Math. 27 (2002), no. 2, 419–436.
- [9] A. F. Beardon, The geometry of discrete groups. Graduate texts in Math., Vol. 91, Springer-Verlag, New York, 1983.
- [10] D. Betsakos, Elliptic, hyperbolic, and condenser capacity; geometric estimates for elliptic capacity. J. Anal. Math. 96 (2005), 37–55.
- [11] D. Betsakos, Hyperbolic geometric versions of Schwarz’s lemma. Conform. Geom. Dyn. 17 (2013), 119–132.
- [12] D. Betsakos and S. Pouliasis, Equality cases for condenser capacity inequalities under symmetrization. Ann. Univ. Mariae Curie-Skłodowska Sect. A 66 (2012), no. 2, 1-24.
- [13] F. Brock and A. Yu. Solynin, An approach to symmetrization via polarization. Trans. Amer. Math. Soc. 352 (2000), no. 4, 1759–1796.
- [14] A. Dinghas, Minkowskische Summen und Integrale. Superadditive Mengenfunktionale. Isoperimetrische Ungleichungen, (German) Mémor. Sci. Math., Fasc. 149 Gauthier-Villars, Paris 1961 101 pp.
- [15] V. N. Dubinin, Change of harmonic measure in symmetrization. (Russian) Mat. Sb. (N.S.) 124(166) (1984), no. 2, 272–279.
- [16] V. N. Dubinin, Transformation of condensers in space. (Russian) Dokl. Akad. Nauk SSSR 296 (1987), no. 1, 18–20; translation in Soviet Math. Dokl. 36 (1988), no. 2, 217–219.
- [17] V. N. Dubinin, Condenser Capacities and Symmetrization in Geometric Function Theory, Birkhäuser, 2014.
- [18] A. Fryntov and J. Rossi, Hyperbolic symmetrization and an inequality of Dyn’kin. Entire functions in modern analysis (Tel-Aviv, 1997), 103–115, Israel Math. Conf. Proc., 15, Bar-Ilan Univ., Ramat Gan, 2001.
- [19] F. W. Gehring, Inequalities for condensers, conformal capacity, and extremal lengths. Michigan Math. J. 18 (1971), 1–20.
- [20] F. Guéritaud, A note on Steiner symmetrization of hyperbolic triangles, Elem. Math. 58 (2003), 21–25.
- [21] P. Hariri, R. Klén, and M. Vuorinen, Conformally Invariant Metrics and Quasiconformal Mappings, Springer Monographs in Mathematics, xix+502 pp, 2020.
- [22] J.-W. M. Van Ittersum, B. Ringeling, and W. Zudilin, Hedgehogs in Lehmer’s problem, Bull. Aust. Math. Soc. 105 ( 2022) , 236 - 242, doi:10.1017/S0004972721000654.
- [23] J. A. Jenkins, Univalent functions and conformal mapping. Reihe: Moderne Funktionentheorie Ergebnisse der Mathematik und ihrer Grenzgebiete, (N.F.), Heft 18. Springer-Verlag, Berlin-Göttingen-Heidelberg 1958 vi+169 pp.
- [24] E. M. Kalmoun, M. M.S. Nasser and M. Vuorinen, Numerical computation of preimage domains for a strip with rectilinear slits, arXiv:2204.00726.
- [25] L. Karp and N. Peyerimhoff, Extremal properties of the principal Dirichlet eigenvalue for regular polygons in the hyperbolic plane. Arch. Math. (Basel) 79 (2002), no. 3, 223–231.
- [26] R. Kühnau, Geometrie der konformen Abbildung auf der hyperbolischen und der elliptischen Ebene. VEB Deutscher Verlag der Wissenschaften, Berlin 1974.
- [27] N. S. Landkof, Foundations of modern potential theory. Translated from the Russian by A. P. Doohovskoy. Die Grundlehren der mathematischen Wissenschaften, Band 180. Springer-Verlag, New York-Heidelberg, 1972.
- [28] N. A. Lebedev, The area principle in the theory of univalent functions. Izdat. "Nauka”, Moscow, 1975. 336 pp.
- [29] M. M.S. Nasser and M. Vuorinen, Computation of conformal invariants.- Appl. Math. Comput., 389 (2021), 125617, arXiv:1908.04533
- [30] M. M.S. Nasser and M. Vuorinen, Isoperimetric properties of condenser capacity.- J. Math. Anal. Appl. 2021, 499, 125050, arXiv:2010.09704
- [31] M. M.S. Nasser, O. Rainio, and M. Vuorinen, Condenser capacity and hyperbolic diameter.- J. Math. Anal. Appl. 508(2022), 125870, arXiv:2011.06293
- [32] M. M.S. Nasser, O. Rainio, and M. Vuorinen, Condenser capacity and hyperbolic perimeter.- Comput. Math. Appl. 105(2022), 54–74. arXiv:2103.10237
- [33] G. Pólya and G. Szegö, Isoperimetric Inequalities in Mathematical Physics. Ann. Math. Studies 27, Princeton Univ. Press, 1952. MR0043486 (13:270d)
- [34] I. E. Pritsker, House of algebraic integers symmetric about the unit circle. J. Number Theory 236 (2022), 388–403.
- [35] Th. Ransford, Potential theory in the complex plane. London Mathematical Society Student Texts, 28. Cambridge University Press, Cambridge, 1995. x+232 pp.
- [36] J. Sarvas, Symmetrization of condensers in n-space. Ann. Acad. Sci. Fenn. Ser. A. I. 1972, no. 522, 44 pp.
- [37] A. Yu. Solynin, Polarization and functional inequalities. (Russian, with Russian summary) Algebra i Analiz 8 (1996), no. 6, 148–185; English transl., St. Petersburg Math. J. 8 (1997), no. 6, 1015–1038.
- [38] A. Yu. Solynin, Extremal configurations in some problems on capacity and harmonic measure. (Russian) Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI) 226 (1996), Anal. Teor. Chisel i Teor. Funktsii. 13, 170–195, 239; translation in J. Math. Sci. (New York) 89 (1998), no. 1, 1031–1049
- [39] A. Yu. Solynin, Harmonic measure of radial segments and symmetrization. (Russian) Mat. Sb. 189 (1998), no. 11, 121–138; translation in Sb. Math. 189 (1998), no. 11–12, 1701–1718.
- [40] A. Yu. Solynin, Continuous symmetrization via polarization. Algebra i Analiz 24 (2012), no. 1, 157–222; reprinted in St. Petersburg Math. J. 24 (2013), no. 1, 117–166.
- [41] A. Yu. Solynin, Exercises on the theme of continuous symmetrization. Comput. Methods Funct. Theory 20 (2020), no. 3-4, 465–509.
- [42] A. Yu. Solynin, Problems on the loss of heat: herd instinct versus individual feelings. Algebra i Analiz 33 (2021), no. 5, 1–50.
- [43] A. Yu. Solynin, Canonical embeddings of pair of arcs and related problems. Submitted (2022).
- [44] G. Szegö, On a certain kind of symmetrization and its applications. Ann. Mat. Pura Appl. (4) 40 (1955), 113–119.
- [45] M. Tsuji, On the capacity of general Cantor sets. Journal of the Mathematical Society of Japan, Vol. 5, No. 2, July, 1953.
- [46] M. Tsuji, Potential Theory in Modern Function Theory. Chelsea Publishing Co., New York, 1975.
- [47] M. Vuorinen, Conformal Geometry and Quasiregular Mappings, Lecture Notes in Mathematics, 1319, Springer-Verlag, 1988.
- [48] H. Wallin, Metrical characterization of conformal capacity zero. J. Math. Anal. Appl. 58 (1977), no. 2, 298–311.
- [49] V. Wolontis, Properties of conformal invariants. Amer. J. Math. 74 (1952), 587–606.
- [50] W. P. Ziemer, Extremal length and -capacity, Michigan Math. J. 16 (1969), 43–51.