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

footnotetext: File: bsv-arXive-09-09-2022.tex, printed: 2025-2-23, 10.34

Conformal capacity of hedgehogs

Dimitrios Betsakos, Alexander Solynin and Matti Vuorinen Department of Mathematics, Aristotle University of Thessaloniki, GR-54124 Thessaloniki, Greece [email protected] Department of Mathematics and Statistics, Texas Tech University, Box 41042, Lubbock, Texas 79409, USA [email protected] Department of Mathematics and Statistics, FI-20014 University of Turku, Finland [email protected]
Abstract.

We discuss problems concerning the conformal condenser capacity of “hedgehogs”, which are compact sets EE in the unit disk 𝔻={z:|z|<1}\mathbb{D}=\{z:\,|z|<1\} consisting of a central body E0E_{0} that is typically a smaller disk 𝔻¯r={z:|z|r}\overline{\mathbb{D}}_{r}=\{z:\,|z|\leq r\}, 0<r<10<r<1, and several spikes EkE_{k} that are compact sets lying on radial intervals I(αk)={teiαk: 0t<1}I(\alpha_{k})=\{te^{i\alpha_{k}}:\,0\leq t<1\}. The main questions we are concerned with are the following: (1) How does the conformal capacity cap(E){\rm cap}(E) of E=k=0nEkE=\cup_{k=0}^{n}E_{k} behave when the spikes EkE_{k}, k=1,,nk=1,\ldots,n, move along the intervals I(αk)I(\alpha_{k}) toward the central body if their hyperbolic lengths are preserved during the motion? (2) How does the capacity cap(E){\rm cap}(E) depend on the distribution of angles between the spikes EkE_{k}? 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 dispersion
2010 Mathematics Subject Classification:
30C85, 31A15, 51M10

In memoriam Jukka Sarvas (1944-2021).

Notation

  • \bullet

    \mathbb{C} - complex plane.

  • 𝔻={z:|z|<1}\mathbb{D}=\{z\in\mathbb{C}:\,|z|<1\} - open unit disk centered at z=0z=0.

  • λ𝔻(z)\lambda_{\mathbb{D}}(z) - density of the hyperbolic metric in 𝔻\mathbb{D}.

  • 𝔻(γ)\ell_{\mathbb{D}}(\gamma) - hyperbolic length of γ\gamma.

  • A𝔻(E)A_{\mathbb{D}}(E) - hyperbolic area of EE.

  • p𝔻(z1,z2)=|(z1z2)/(1z1z¯2)|p_{\mathbb{D}}(z_{1},z_{2})=\left|(z_{1}-z_{2})/(1-z_{1}\overline{z}_{2})\right| - pseudo-hyperbolic metric in 𝔻\mathbb{D}.

  • d𝔻(z1,z2)=log1+p𝔻(z1,z2)1p𝔻(z1,z2)d_{\mathbb{D}}(z_{1},z_{2})=\log\frac{1+p_{\mathbb{D}}(z_{1},z_{2})}{1-p_{\mathbb{D}}(z_{1},z_{2})} - hyperbolic metric in 𝔻\mathbb{D}.

  • [z2,z2][z_{2},z_{2}] and (z1,z2)(z_{1},z_{2}) - closed and open Euclidean intervals with end points z1z_{1} and z2z_{2}.

  • [z1,z2]h[z_{1},z_{2}]_{h} and (z1,z2)h(z_{1},z_{2})_{h} - closed and open hyperbolic intervals with end points z1z_{1} and z2z_{2}.

  • cap(E)=cap(𝔻,E){\rm cap}(E)={\rm cap}(\mathbb{D},E) - conformal capacity of a compact set E𝔻E\subset\mathbb{D}.

  • log.cap(E){\rm log.cap}(E) - logarithmic capacity of EE\subset\mathbb{C}.

  • 𝖬(Γ)\mathsf{M}(\Gamma) - modulus of the family of curves Γ\Gamma.

  • γ(E)\mathcal{R}_{\gamma}(E) - reflection of a set EE with respect to the hyperbolic geodesic γ\gamma.

  • 𝒫γ(E)\mathcal{P}_{\gamma}(E) - polarization of a set EE with respect to the hyperbolic geodesic γ\gamma.

  • 𝒦(k)\mathcal{K}(k) and 𝒦(k)=𝒦(1k2)\mathcal{K}^{\prime}(k)=\mathcal{K}(\sqrt{1-k^{2}}) - complete elliptic integrals of the first kind.

1. Introduction

The main theme discussed in this paper is the dependence of the condenser capacity cap(𝔻,E){\rm cap}(\mathbb{D},E) on the geometric structure and characteristics of a compact set E𝔻,E\subset\mathbb{D}, where 𝔻\mathbb{D} is the unit disk 𝔻={z:|z|<1}\mathbb{D}=\{z:\,|z|<1\} in the complex plane \mathbb{C}. For brevity, we call cap(𝔻,E){\rm cap}(\mathbb{D},E) the conformal capacity of EE or the capacity of EE and denote it by cap(E){\rm cap}(E).

Due to the conformal invariance of the capacity it is natural to equip 𝔻\mathbb{D} 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 EE 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 E=k=0nEkE=\cup_{k=0}^{n}E_{k} consisting of a central body E0E_{0}, which can be absent, and spikes EkE_{k}, k=1,,nk=1,\ldots,n, that are closed intervals or any compact sets lying on nn radial intervals I(αk)I(\alpha_{k}), where I(α)={teiα: 0t<1}I(\alpha)=\{te^{i\alpha}:\,0\leq t<1\}, I=I(0)I=I(0). 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 𝔻\mathbb{D} is defined by the element of length

λ𝔻(z)|dz|=2|dz|1|z|2.\lambda_{\mathbb{D}}(z)\,|dz|=\frac{2|dz|}{1-|z|^{2}}.

Then the hyperbolic length 𝔻(E)\ell_{\mathbb{D}}(E) of a compact subset EE of a rectifiable curve can be calculated as

𝔻(E)=E2|dz|1|z|2.\ell_{\mathbb{D}}(E)=\int_{E}\frac{2|dz|}{1-|z|^{2}}.

Furthermore, the hyperbolic geodesics are circular arcs in 𝔻\mathbb{D} that are orthogonal to the unit circle 𝕋=𝔻\mathbb{T}=\partial\mathbb{D} at their end points. The hyperbolic distance between points z1z_{1} and z2z_{2} in 𝔻\mathbb{D}, that is equal to the hyperbolic length 𝔻([z1,z2]h)\ell_{\mathbb{D}}([z_{1},z_{2}]_{h}) of the closed hyperbolic interval [z1,z2]h[z_{1},z_{2}]_{h} joining these points, is given by

(1.1) d𝔻(z1,z2)=log1+p𝔻(z1,z2)1p𝔻(z1,z2),d_{\mathbb{D}}(z_{1},z_{2})=\log\frac{1+p_{\mathbb{D}}(z_{1},z_{2})}{1-p_{\mathbb{D}}(z_{1},z_{2})},

where p𝔻(z1,z2)p_{\mathbb{D}}(z_{1},z_{2}) is the pseudo-hyperbolic metric defined as

(1.2) p𝔻(z1,z2)=|z1z21z1z¯2|.p_{\mathbb{D}}(z_{1},z_{2})=\left|\frac{z_{1}-z_{2}}{1-z_{1}\overline{z}_{2}}\right|.

Everywhere below, [z1,z2]h[z_{1},z_{2}]_{h} and (z1,z2)h(z_{1},z_{2})_{h} stand, respectively, for the closed and open hyperbolic intervals with end points z1z_{1} and z2z_{2}. Similarly, notations [z2,z2][z_{2},z_{2}] and (z1,z2)(z_{1},z_{2}) will be reserved for the closed and open Euclidean intervals with end points z1z_{1} and z2z_{2}. If z1z_{1} and z2z_{2} lie on the same diameter of 𝔻\mathbb{D} then, of course, [z1,z2]h=[z1,z2][z_{1},z_{2}]_{h}=[z_{1},z_{2}] and (z1,z2)h=(z1,z2)(z_{1},z_{2})_{h}=(z_{1},z_{2}).

When z1=0z_{1}=0 and z2=rz_{2}=r, 0<r<10<r<1, the hyperbolic length τ=τ(r)\uptau=\uptau(r) of the interval [0,r][0,r] and its Euclidean length r=r(τ)r=r(\uptau) are connected via the following formulas, which are often used in calculations:

(1.3) τ=log1+r1r,r=eτ1eτ+1.\uptau=\log\frac{1+r}{1-r},\quad\quad r=\frac{e^{\uptau}-1}{e^{\uptau}+1}.

The hyperbolic area of a Borel measurable subset EE of 𝔻\mathbb{D} is

(1.4) A𝔻(E)=Eλ𝔻2(z)𝑑m,A_{\mathbb{D}}(E)=\int_{E}\lambda_{\mathbb{D}}^{2}(z)\,dm,

where dmdm stands for the 22-dimensional Lebesgue measure. In particular, the hyperbolic area of the disk 𝔻r={z:|z|<r}\mathbb{D}_{r}=\{z:\,|z|<r\}, 0<r<10<r<1, is given by the following formula:

A𝔻(𝔻r)=4πr21r2=4πsinh2(τ/2).A_{\mathbb{D}}(\mathbb{D}_{r})=\frac{4\pi\,r^{2}}{1-r^{2}}=4\pi\sinh^{2}(\uptau/2).

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 EE be a compact set in 𝔻\mathbb{D}. The conformal capacity cap(E){\rm cap}(E) of EE is defined as

(1.6) cap(E)=inf𝔻|u|2𝑑m,{\rm cap}(E)=\inf\int_{\mathbb{D}}|\nabla u|^{2}\,dm,

where the infimum is taken over all Lipschitz functions uu such that u=0u=0 on the unit circle 𝕋=𝔻\mathbb{T}=\partial\mathbb{D} and u(z)1u(z)\geq 1 for zEz\in E.

In terminology used in electrostatics, the conformal capacity cap(E){\rm cap}(E) is usually referred to as the capacity of the condenser (𝔻,E)(\mathbb{D},E) with plates EE and 𝔻=¯𝔻\mathbb{D}^{*}=\overline{\mathbb{C}}\setminus\mathbb{D} and field 𝔻E\mathbb{D}\setminus E. 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 𝔻r\mathbb{D}_{r}, with Euclidean radius rr, 0<r<10<r<1, and hyperbolic radius τ>0,\uptau>0, is given by

cap(𝔻¯r)=2πlog(1/r)=2πlog((eτ+1)/(eτ1)).{\rm cap}(\overline{\mathbb{D}}_{r})=\frac{2\pi}{\log(1/r)}=\frac{2\pi}{\log((e^{\uptau}+1)/(e^{\uptau}-1))}.
Example 1.8.

The conformal capacity of the interval [r,r][-r,r], with hyperbolic length equal to 2τ=𝔻([r,r])=2log((1+r)/(1r))2\uptau=\ell_{\mathbb{D}}([-r,r])=2\log((1+r)/(1-r)), is

cap([r,r])=8𝒦(r2)𝒦(r2)=8𝒦((eτ1)2/(eτ+1)2)𝒦((eτ1)2/(eτ+1)2).{\rm cap}([-r,r])=8\frac{\mathcal{K}(r^{2})}{\mathcal{K}^{\prime}(r^{2})}=8\frac{\mathcal{K}((e^{\uptau}-1)^{2}/(e^{\uptau}+1)^{2})}{\mathcal{K}^{\prime}((e^{\uptau}-1)^{2}/(e^{\uptau}+1)^{2})}.

Here and below, 𝒦(k)\mathcal{K}(k) and 𝒦(k)=𝒦(1k2)\mathcal{K}^{\prime}(k)=\mathcal{K}(\sqrt{1-k^{2}}) stand for the complete elliptic integrals of the first kind.

Example 1.9.

The conformal capacity of the interval [0,r][0,r], with hyperbolic length equal to τ=𝔻([0,r])=log((1+r)/(1r))\uptau=\ell_{\mathbb{D}}([0,r])=\log((1+r)/(1-r)), is

(1.10) cap([0,r])=4𝒦(r)𝒦(r)=4𝒦((eτ1)/(eτ+1))𝒦((eτ1)/(eτ+1)).{\rm cap}([0,r])=4\frac{\mathcal{K}(r)}{\mathcal{K}^{\prime}(r)}=4\frac{\mathcal{K}((e^{\uptau}-1)/(e^{\uptau}+1))}{\mathcal{K}^{\prime}((e^{\uptau}-1)/(e^{\uptau}+1))}.

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 0<r<s<t<10<r<s<t<1 and 0<u<10<u<1 are such that the sets E=[0,r][s,t]E=[0,r]\cup[s,t] and E1=[0,u]E_{1}=[0,u] have equal hyperbolic lengths. Is it true that conformal capacity of EE is greater than conformal capacity of E1E_{1}?

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 E0E_{0} will be a disk 𝔻¯r\overline{\mathbb{D}}_{r}, 0<r<10<r<1, or an emptyset and EkE_{k}, 1km1\leq k\leq m, will be a collection of closed intervals attached to the central body. In this case, our compact sets EE 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.

Refer to caption
Figure 1. Hedgehog-like shapes: Live and geometrical.

Our main results in Sections 3 and 4 deal with several extremal problems for the capacity of compact sets in the unit disk 𝔻\mathbb{D}, where hedgehogs possessing certain symmetry properties play the role of the extremal configuration. Thus, in these sections we mainly work with compact sets in 𝔻\mathbb{D} 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 𝔻\mathbb{D}. 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 n3n\geq 3.

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 (D,E)(D,E), where DD is a domain in the plane and EE is a compact subset of DD. The capacity of the condenser (D,E)(D,E) is defined by

(2.1) cap(D,E)=infD|u|2𝑑m,{\rm cap}(D,E)=\inf\int_{D}|\nabla u|^{2}\,dm,

where the infimum is taken over all Lipschitz functions uu such that u0u\leq 0 on D\partial D and u1u\geq 1 on EE. These functions will be called admissible for the condenser (D,E)(D,E). We want to stress here that if D=𝔻D=\mathbb{D}, then the infimum in (2.1) can be taken over all admissible functions as above with an additional requirement that u(z)=0u(z)=0 for all z𝕋z\in\mathbb{T}.

By Theorem 3.8 of Ziemer [50], the capacity of the condenser (D,E)(D,E) is equal to the modulus 𝖬(Γ)\mathsf{M}(\Gamma) of the family Γ\Gamma of all curves in DED\setminus E joining EE with D\partial D. 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 𝔻\mathbb{D} and it is invariant under reflections with respect to hyperbolic geodesics. Thus, if EE is a compact subset of 𝔻\mathbb{D} and φ:𝔻𝔻\varphi:\mathbb{D}\to\mathbb{D} is a Möbius automorphism or a reflection with respect to a hyperbolic geodesic, then cap(φ(E))=cap(E).{\rm cap}(\varphi(E))={\rm cap}(E).

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 log.cap(E){\rm log.cap}(E) of a set EE\subset\mathbb{C}, not necessarily compact, is given by

log.cap(E)=supμeI(μ)withI(μ)=log|zw|dμ(z)𝑑μ(w),{\rm log.cap}(E)=\sup_{\mu}e^{I(\mu)}\quad{\mbox{with}}\quad I(\mu)=\iint\log|z-w|\,d\mu(z)d\mu(w),

where the supremum is taken over all Borel probability measures μ\mu on \mathbb{C} whose support is a compact subset of EE. 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. (1)

    If E𝔻E\subset\mathbb{D} is compact, then cap(E)=0{\rm cap}(E)=0 if and only if log.cap(E)=0{\rm log.cap}(E)=0.

  2. (2)

    Let E1E_{1}, E2E_{2} be compact sets in 𝔻\mathbb{D} such that E1E2E_{1}\subset E_{2}. Then cap(E1)=cap(E2){\rm cap}(E_{1})={\rm cap}(E_{2}) if and only if log.cap(E2E1)=0{\rm log.cap}(E_{2}\setminus E_{1})=0.

  3. (3)

    If E(1,1)E\subset(-1,1) is such that log.cap(E)=0{\rm log.cap}(E)=0, then 𝔻(E)=0\ell_{\mathbb{D}}(E)=0.

We note that the set E2E1E_{2}\setminus E_{1} in part (2) and the set EE 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 K[0,1]K\subset[0,1] is the standard Cantor set, then its scaled version K1/3={z: 3zK}K_{1/3}=\{z:\,3z\in K\} 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 EE be a compact set in 𝔻\mathbb{D}. There is a unique function uEu_{E}, called the potential function of EE, that minimizes the integral in (2.1); i.e. such that

cap(E)=𝔻|uE|2𝑑m.{\rm{cap}}(E)=\int_{\mathbb{D}}|\nabla u_{E}|^{2}\,dm.

Moreover, uEu_{E} possesses the following properties:

  1. (1)

    uEu_{E} is harmonic in 𝔻E\mathbb{D}\setminus E and continuous on 𝔻¯\overline{\mathbb{D}}, except possibly for a subset of EE of zero logarithmic capacity.

  2. (2)

    uE(z)=0u_{E}(z)=0 for z𝕋z\in\mathbb{T} and uE(z)=1u_{E}(z)=1 for all zEz\in E, except possibly for a subset of EE of zero logarithmic capacity.

  3. (3)

    If every point of E\partial E is regular for the Dirichlet problem in 𝔻E\mathbb{D}\setminus E, then uEu_{E} is continuous on 𝔻¯\overline{\mathbb{D}} and uE=1u_{E}=1 on EE.

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 Ee={0}(n=1In)E_{e}=\{0\}\cup\left(\cup_{n=1}^{\infty}I_{n}\right), where InI_{n} is an interval [en,en+en3][e^{-n},e^{-n}+e^{-n^{3}}] with length ln=en3l_{n}=e^{-n^{3}}. Since

n=1nlog(2/ln)=n=1nn3+log2<,\sum_{n=1}^{\infty}\frac{n}{\log(2/l_{n})}=\sum_{n=1}^{\infty}\frac{n}{n^{3}+\log 2}<\infty,

Wiener’s criterion (see [35, Theorem 5.4]) implies that the point z=0Eez=0\in E_{e} is irregular for the Dirichlet problem. Therefore, the function uEe(z)1u_{E_{e}}(z)-1 is not a barrier at z=0z=0 (see [35, Definition 4.1.4]). The latter implies that the limit limz0uEe(z)\lim_{z\to 0}u_{E_{e}}(z) does not exist. Since points zEez\in E_{e}, z0z\not=0, are regular for the Dirichlet problem, z=0z=0 is the only point in 𝔻\mathbb{D}, where uEeu_{E_{e}} is not continuous.

To obtain a compact set FeF_{e} such that uFeu_{F_{e}} has infinite number of discontinuities, we modify our previous example as follows. For nn\in\mathbb{N}, let Een={(z+1)/2n:zEe}E_{e}^{n}=\{(z+1)/2^{n}:\,z\in E_{e}\}. Thus, EenE_{e}^{n} is obtained by translating and scaling the set EeE_{e}. Let Fe={0}(n=1Een)F_{e}=\{0\}\cup\left(\cup_{n=1}^{\infty}E_{e}^{n}\right). Our previous argument can be applied to show that uFeu_{F_{e}} is not continuous at an infinite set of points zn=2nz_{n}=2^{-n}, nn\in\mathbb{N}. Thus, FeF_{e} has an infinite subset exceptional in the sense of part (1) of Proposition 2.4.

Example 2.6.

Let E𝔻E\subset\mathbb{D} be a compact set of positive logarithmic capacity, which contains a nonempty subset E0E_{0}, each point of which is isolated from other points of EE. Since uEu_{E} is harmonic and bounded, every point zE0z\in E_{0} is removable, which means that uEu_{E} can be extended as a function harmonic at zz. Since uEu_{E} is not constant, it follows that 0<uE(z)<10<u_{E}(z)<1 for every point zE0z\in E_{0}. Thus, E0EE_{0}\subset E 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 E=k=1nEkE=\cup_{k=1}^{n}E_{k} is the union of n2n\geq 2 compact sets EkE_{k}, k=1,,nk=1,\ldots,n, in 𝔻\mathbb{D}. Then

(2.8) cap(E)k=1ncap(Ek).{\rm{cap}}(E)\leq\sum_{k=1}^{n}{\rm{cap}}(E_{k}).

Moreover, if each EkE_{k} has positive conformal capacity, then (2.8) holds with strict inequality.

Proof. Let uku_{k}, k=1,,nk=1,\ldots,n, be an admissible function for the condenser (𝔻,Ek)(\mathbb{D},E_{k}) such that uk(z)=0u_{k}(z)=0 for z𝕋z\in\mathbb{T}. Then u(z)=max{u1(z),,un(z)}u(z)=\max\{u_{1}(z),\ldots,u_{n}(z)\} is an admissible function for the condenser (𝔻,E)(\mathbb{D},E) such that |u(z1)u(z2)|max1kn|uk(z1)uk(z2)||u(z_{1})-u(z_{2})|\leq\max_{1\leq k\leq n}|u_{k}(z_{1})-u_{k}(z_{2})| for all z1,z2𝔻z_{1},z_{2}\in\mathbb{D}. The latter inequality implies that |u(z)|max1kn|uk(z)||\nabla u(z)|\leq\max_{1\leq k\leq n}|\nabla u_{k}(z)| for all z𝔻z\in\mathbb{D}, where the gradients exist. Therefore,

(2.9) cap(E)𝔻|u|2𝑑m𝔻max1kn|uk|2dmk=1n𝔻|uk|2𝑑m.{\rm{cap}}(E)\leq\int_{\mathbb{D}}|\nabla u|^{2}\,dm\leq\int_{\mathbb{D}}\max_{1\leq k\leq n}|\nabla u_{k}|^{2}\,dm\leq\sum_{k=1}^{n}\int_{\mathbb{D}}|\nabla u_{k}|^{2}\,dm.

Taking the infimum in this equation over all admissible functions uku_{k}, k=1,,nk=1,\ldots,n, with the properties mentioned above in this proof, we obtain the required subadditivity property (2.8).

If cap(Ek)>0,k=1,2,,n{\rm cap}(E_{k})>0,\;k=1,2,\dots,n, then each of the condensers (𝔻,Ek)(\mathbb{D},E_{k}) has the potential function uEku_{E_{k}} that is a non-constant harmonic function in 𝔻Ek\mathbb{D}\setminus E_{k}. Therefore, |uEk|>0|\nabla u_{E_{k}}|>0, almost everywhere in 𝔻Ek\mathbb{D}\setminus E_{k}. Since uEku_{E_{k}} is the potential function of (𝔻,Ek)(\mathbb{D},E_{k}) it follows that (2.9) holds with uk=uEku_{k}=u_{E_{k}}. In this case, the strict inequality max1kn|uk(z)|<1kn|uk(z)|\max_{1\leq k\leq n}|\nabla u_{k}(z)|<\sum_{1\leq k\leq n}|\nabla u_{k}(z)| holds for all points zz in an annulus {z:ρ<|z|<1}\{z:\,\rho<|z|<1\} with 0<ρ<10<\rho<1 such that {z:ρ<|z|<1}𝔻1knEk\{z:\,\rho<|z|<1\}\subset\mathbb{D}\setminus\cup_{1\leq k\leq n}E_{k}. The latter implies that if uk=uEku_{k}=u_{E_{k}} then the third inequality in (2.9) is strict. Therefore, (2.8) holds with the sign of strict inequality in the case under consideration. \Box

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 γ\gamma, we need the following terminology. To any hyperbolic geodesic γ\gamma, we can give an orientation by marking one of its complementary hyperbolic halfplanes and call it H+H_{+}; then the other complementary hyperbolic halfplane is given the name HH_{-}. Since every hyperbolic geodesic γ\gamma is an arc of a circle, we can define the classical symmetry transformation (called also inversion or reflection) with respect to γ\gamma. We note that the symmetry transformation with respect to a hyperbolic geodesic γ\gamma is a hyperbolic isometry on 𝔻\mathbb{D}.

The polarization transformation of compact sets in 𝔻\mathbb{D} can be defined as follows.

Definition 2.10.

Let γ\gamma be an oriented hyperbolic geodesic in 𝔻\mathbb{D}. Let H+,HH_{+},H_{-} be the hyperbolic halfplanes determined by γ\gamma. Let EE be a compact set in 𝔻\mathbb{D} and let γ(E)\mathcal{R}_{\gamma}(E) denote the set symmetric to EE with respect to γ\gamma. The polarization 𝒫γ(E)\mathcal{P}_{\gamma}(E) of EE with respect to γ\gamma is defined by

(2.11) 𝒫γ(E)=((Eγ(E))H+¯)((Eγ(E))H).\mathcal{P}_{\gamma}(E)=((E\cup\mathcal{R}_{\gamma}(E))\cap\overline{H_{+}})\cup((E\cap\mathcal{R}_{\gamma}(E))\cap H_{-}).

Equation (2.11) can be also written in the form

𝒫γ(E)=((Eγ(E))H)((Eγ(E))H+).\mathcal{P}_{\gamma}(E)=((E\cup\mathcal{R}_{\gamma}(E))\setminus H_{-})\cup((E\cap\mathcal{R}_{\gamma}(E))\setminus H_{+}).

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 EE be a compact set in 𝔻\mathbb{D} and 𝒫γ(E)\mathcal{P}_{\gamma}(E) be the polarization of EE with respect to an oriented hyperbolic geodesic γ\gamma. Then

(2.13) cap(𝒫γ(E))cap(E).{\rm cap}(\mathcal{P}_{\gamma}(E))\leq{\rm cap}(E).

Furthermore, equality occurs in (2.13) if and only if 𝒫γ(E)\mathcal{P}_{\gamma}(E) coincides with EE up to reflection with respect to γ\gamma 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.

Under the assumptions of Proposition 2.12, let EE be the closure of the union of a finite or infinite number of non-overlapping closed intervals on the diameter (1,1)(-1,1). Then (2.13) holds with the sign of strict inequality unless 𝒫γ(E)\mathcal{P}_{\gamma}(E) coincides with EE up to reflection with respect to γ\gamma.

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 E𝔻E\subset\mathbb{D} be a compact set in 𝔻\mathbb{D}. The hyperbolic transfinite diameter of EE is defined as

(2.16) dh(E)=limnmax1j<kn(p𝔻(zj,zk))2/[n(n1)],{\rm d}_{h}(E)=\lim_{n\to\infty}\max\prod_{1\leq j<k\leq n}\left(p_{\mathbb{D}}(z_{j},z_{k})\right)^{2/[n(n-1)]},

where p𝔻(zj,zk)p_{\mathbb{D}}(z_{j},z_{k}) stands for the pseudo-hyperbolic metric defined by (1.2) and the maximum is taken of all nn-tuples of points z1,,znz_{1},\ldots,z_{n} in EE.

The following relation was established in [46].

Proposition 2.17.

Let EE be a compact set in 𝔻\mathbb{D}. Then

(2.18) cap(E)=[12πlogdh(E)]1.{\rm cap}(E)=\left[-\frac{1}{2\pi}\log{\rm d}_{h}(E)\right]^{-1}.

Let EE be a compact set in 𝔻\mathbb{D}. A map φ:E𝔻\varphi:E\to{\mathbb{D}} is called a hyperbolic contraction on EE if for every z1,z2Ez_{1},z_{2}\in E,

d𝔻(φ(z1),φ(z2))d𝔻(z1,z2),d_{\mathbb{D}}(\varphi(z_{1}),\varphi(z_{2}))\leq d_{\mathbb{D}}(z_{1},z_{2}),

where d𝔻(,)d_{\mathbb{D}}(\cdot,\cdot) stands for the hyperbolic metric defined by (1.1). Furthermore, φ:E𝔻\varphi:E\to{\mathbb{D}} is called a strict hyperbolic contraction on EE if there is kk, 0<k<10<k<1, such that for every z1,z2Ez_{1},z_{2}\in E,

d𝔻(φ(z1),φ(z2))kd𝔻(z1,z2).d_{\mathbb{D}}(\varphi(z_{1}),\varphi(z_{2}))\leq k\,d_{\mathbb{D}}(z_{1},z_{2}).

The following contraction principle is immediate from the Definition 2.15 and Proposition 2.17.

Proposition 2.19.

Let EE be a compact set in 𝔻\mathbb{D}. Let φ:E𝔻\varphi:E\to\mathbb{D} be a hyperbolic contraction. Then cap(φ(E))cap(E){\rm cap}(\varphi(E))\leq{\rm cap}(E).

Moreover, if φ\varphi is a strict hyperbolic contraction on EE, then cap(φ(E))<cap(E){\rm cap}(\varphi(E))<{\rm cap}(E).

Remark 2.20.

We note here that the polarization transformation is not contracting, in general. For example, polarizing the set E={±i/4,±(1i)/4}𝔻E=\{\pm i/4,\pm(1-i)/4\}\subset\mathbb{D} with respect to the diameter I=(1,1)I=(-1,1) with its standard orientation, we obtain the polarized set 𝒫I(E)={±i/4,(±1+i)/4}{\mathcal{P}}_{I}(E)=\{\pm i/4,(\pm 1+i)/4\}. Then for every one-to-one map φ:E𝒫I(E)\varphi:E\to{\mathcal{P}}_{I}(E), there is a pair of points z1,z2Ez_{1},z_{2}\in E such that d𝔻(φ(z2),φ(z1))>d𝔻(z2,z1)d_{\mathbb{D}}(\varphi(z_{2}),\varphi(z_{1}))>d_{\mathbb{D}}(z_{2},z_{1}); one can easily verify this inequality by considering each one of the possible maps φ\varphi.

To study the limit behavior of the conformal capacity cap(E){\rm{cap}}(E), when some of the components of EE tend to the boundary of 𝔻\mathbb{D}, we need a hyperbolic analog of the dispersion property of the Newtonian capacity discussed in [42]. Let E1,,EnE_{1},\ldots,E_{n} be disjoint nonempty compact sets in 𝔻\mathbb{D}, not necessarily connected, and let E=k=1nEkE=\cup_{k=1}^{n}E_{k}.

Definition 2.21.

By a hyperbolic dispersion of E=k=1nEkE=\cup_{k=1}^{n}E_{k} we mean a mapping φ:E×[0,)𝔻\varphi:E\times[0,\infty)\to\mathbb{D} satisfying the following properties:

  1. (1)

    For each kk, the restriction φ:Ek×[0,)𝔻\varphi:E_{k}\times[0,\infty)\to\mathbb{D} is a rigid hyperbolic motion of EkE_{k}, which depends continuously on the parameter t[0,)t\in[0,\infty), such that φ(x,0)=x\varphi(x,0)=x for all xEx\in E.

  2. (2)

    If 0t1<t20\leq t_{1}<t_{2}, then for each kk and jj, kjk\not=j, the hyperbolic distances between the images φ(Ek,t)\varphi(E_{k},t) and φ(Ej,t)\varphi(E_{j},t) satisfy the following inequalities:

    d𝔻(φ(Ek,t1),φ(Ej,t1))d𝔻(φ(Ek,t2),φ(Ej,t2)).d_{\mathbb{D}}\,(\varphi(E_{k},t_{1}),\varphi(E_{j},t_{1}))\leq d_{\mathbb{D}}\,(\varphi(E_{k},t_{2}),\varphi(E_{j},t_{2})).
  3. (3)

    For each kk and jj, kjk\not=j,

    d𝔻(φ(Ek,t),φ(Ej,t))as t.d_{\mathbb{D}}\,(\varphi(E_{k},t),\varphi(E_{j},t))\to\infty\quad{\mbox{as $t\to\infty$.}}

Thus, hyperbolic dispersion of EE is a process moving the subsets E1,,EnE_{1},\ldots,E_{n} 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 E=E1E2E=E_{1}\cup E_{2}, where E1={0}E_{1}=\{0\} and E2={z=reiθ:|θ|πε}E_{2}=\{z=re^{i\theta}:\,|\theta|\leq\pi-\varepsilon\} with 0<r<10<r<1 and sufficiently small ε>0\varepsilon>0, cannot be hyperbolically dispersed in the sense of Definition 2.21. On the other hand, any union E=E1E2E=E_{1}\cup E_{2} 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 φ:E×[0,)𝔻\varphi:E\times[0,\infty)\to\mathbb{D} be a hyperbolic dispersion of a compact set E=k=1nEkE=\cup_{k=1}^{n}E_{k}, as above. Then

(2.23) cap(φ(E,t))k=1ncap(Ek),as t.{\rm{cap}}(\varphi(E,t))\to\sum_{k=1}^{n}{\rm{cap}}(E_{k}),\quad{\mbox{as $t\to\infty$.}}

In the proof of Proposition 2.22, we will need the following elementary arithmetic result.

Lemma 2.24.

Let 0<αk<10<\alpha_{k}<1, k=1,,nk=1,\ldots,n, be such that k=1nαk=1\sum_{k=1}^{n}\alpha_{k}=1. Then there are kk sequences of positive integers mj,km_{j,k}, j=1,2,j=1,2,\ldots, such that if mj=k=1nmj,km_{j}=\sum_{k=1}^{n}m_{j,k}, then mjm_{j}\to\infty and mj,k/mjαkm_{j,k}/m_{j}\to\alpha_{k} as jj\to\infty.

Proof. Consider rational approximations of αk\alpha_{k}, k=1,,n1k=1,\ldots,n-1; i.e. consider (n1)(n-1) sequences

aj,kbj,kαk,k=1,,n1, as j,\frac{a_{j,k}}{b_{j,k}}\to\alpha_{k},\quad\ k=1,\ldots,n-1,\ {\mbox{ as $j\to\infty$,}}

where aj,ka_{j,k}, bj,kb_{j,k} are positive integers. Then consider the sequence

mj=jk=1n1bj,km_{j}=j\prod_{k=1}^{n-1}b_{j,k}\to\infty

and the sequences

mj,k=aj,kmjbj,k,k=1,,n1.m_{j,k}=\frac{a_{j,k}m_{j}}{b_{j,k}},\quad k=1,\ldots,n-1.

Clearly,

mj,kmj=aj,kbj,kαkas j.\frac{m_{j,k}}{m_{j}}=\frac{a_{j,k}}{b_{j,k}}\to\alpha_{k}\quad{\mbox{as $j\to\infty$.}}

Also, we put

mj,n=mjk=1n1mj,k.m_{j,n}=m_{j}-\sum_{k=1}^{n-1}m_{j,k}.

Then

mj,nmj=1k=1n1mj,kmj1k=1n1αk=αn>0as j.\frac{m_{j,n}}{m_{j}}=1-\sum_{k=1}^{n-1}\frac{m_{j,k}}{m_{j}}\to 1-\sum_{k=1}^{n-1}\alpha_{k}=\alpha_{n}>0\quad{\mbox{as $j\to\infty$.}}

The latter relation shows that mj,n>0m_{j,n}>0 for all jj sufficiently large. Therefore, we can remove a finite number of terms from the sequence mjm_{j} and from the sequences mj,km_{j,k} and then re-enumerate these sequences to obtain sequences with the required properties. \Box

Proof of Proposition 2.22. For k=1,2,,nk=1,2,\dots,n and t0t\geq 0, we set Et=φ(E,t)E^{t}=\varphi(E,t), Ekt=φ(Ek,t)E_{k}^{t}=\varphi(E_{k},t). Since the conformal capacity is invariant under hyperbolic motions, cap(Ekt)=cap(Ek){\rm{cap}}(E_{k}^{t})={\rm{cap}}(E_{k}) for all k=1,,nk=1,\ldots,n and all t0t\geq 0. This together with the subadditivity property of Proposition 2.19 implies that

(2.25) cap(Et)k=1ncap(Ekt)=k=1ncap(Ek).{\rm cap}(E^{t})\leq\sum_{k=1}^{n}{\rm cap}(E_{k}^{t})=\sum_{k=1}^{n}{\rm cap}(E_{k}).

We assume without loss of generality that cap(Ek)>0{\rm{cap}}(E_{k})>0 for all k=1,,nk=1,\ldots,n. Then we set αk=cap(Ek)/k=1ncap(Ek)\alpha_{k}={\rm{cap}}(E_{k})/\sum_{k=1}^{n}{\rm{cap}}(E_{k}) and will use the sequences mj,km_{j,k}, k=1,,nk=1,\ldots,n, and mjm_{j} defined as in the proof of Lemma 2.24 for our choice of αk\alpha_{k}.

Let zj,ksz_{j,k}^{s}, s=1,,mj,ks=1,\ldots,m_{j,k}, be points in EkE_{k} such that

(2.26) 1l<smj,kp𝔻(zj,kl,zj,ks)=max1l<smj,kp𝔻(zl,zs),\prod_{1\leq l<s\leq m_{j,k}}p_{\mathbb{D}}(z_{j,k}^{l},z_{j,k}^{s})=\max\prod_{1\leq l<s\leq m_{j,k}}p_{\mathbb{D}}(z^{l},z^{s}),

where the maximum is taken over all mj,km_{j,k}-tuples of points z1,z2,,zmj,kz^{1},z^{2},\ldots,z^{m_{j,k}} in EkE_{k}. For k=1,,nk=1,\ldots,n, j=1,2,j=1,2,\ldots, s=1,,mj,ks=1,\ldots,m_{j,k}, and t0t\geq 0, we set zj,ks,t=φ(zj,ks,t)z_{j,k}^{s,t}=\varphi(z_{j,k}^{s},t). Since φ\varphi is a hyperbolic motion on each EkE_{k}, we have

(2.27) p𝔻(zj,kl,t,zj,ks,t)=p𝔻(zj,kl,zj,ks),p_{\mathbb{D}}(z_{j,k}^{l,t},z_{j,k}^{s,t})=p_{\mathbb{D}}(z_{j,k}^{l},z_{j,k}^{s}),

for all points zj,klz_{j,k}^{l}, zj,ksz_{j,k}^{s} defined above and all t0t\geq 0.

For our choice of points, it follows from (2.27) and equation (2.16) of Definition 2.15 that

(2.28) dh(Et)lim supj[Πjtk=1nΠj,k]2/mj(mj1),d_{h}(E^{t})\geq\limsup_{j\to\infty}\left[\Pi_{j}^{t}\,\prod_{k=1}^{n}\Pi_{j,k}\right]^{2/m_{j}(m_{j}-1)},

where

(2.29) Πj,k=1l<smj,kp𝔻(zj,kl,zj,ks),\Pi_{j,k}=\prod_{1\leq l<s\leq m_{j,k}}p_{\mathbb{D}}(z_{j,k}^{l},z_{j,k}^{s}),

and

(2.30) Πjt=p𝔻(zj,k1l,t,zj,k2s,t),\Pi_{j}^{t}=\prod p_{\mathbb{D}}(z_{j,k_{1}}^{l,t},z_{j,k_{2}}^{s,t}),

where the product in (2.30) is taken over all pairs of points zj,k1l,tz_{j,k_{1}}^{l,t}, zj,k2s,tz_{j,k_{2}}^{s,t} such that 1lmj,k11\leq l\leq m_{j,k_{1}}, 1smj,k21\leq s\leq m_{j,k_{2}} and k1k2k_{1}\not=k_{2}.

Using equations (2.16), (2.26), and (2.29) and taking into account our choice of points zj,ksz_{j,k}^{s}, 1smj,k1\leq s\leq m_{j,k} and the limit relation limjmj,k/mj=αk\lim_{j\to\infty}m_{j,k}/m_{j}=\alpha_{k}, we conclude that

(2.31) limj(Πj,k)2/mj(mj1)\displaystyle\lim_{j\to\infty}\left(\Pi_{j,k}\right)^{2/m_{j}(m_{j}-1)} =limj(1l<smj,kp𝔻(zj,kl,zj,ks))2mj,k(mj,k1)mj,k(mj,k1)mj(mj1)\displaystyle=\lim_{j\to\infty}\left(\prod_{1\leq l<s\leq m_{j,k}}p_{\mathbb{D}}(z_{j,k}^{l},z_{j,k}^{s})\right)^{\frac{2}{m_{j,k}(m_{j,k}-1)}\cdot\frac{m_{j,k}(m_{j,k}-1)}{m_{j}(m_{j}-1)}}
=(dh(Ek))αk2.\displaystyle=\left(d_{h}(E_{k})\right)^{\alpha_{k}^{2}}.

Our assumption that d𝔻(Ek1t,Ek2t)d_{\mathbb{D}}(E_{k_{1}}^{t},E_{k_{2}}^{t})\to\infty when k1k2k_{1}\not=k_{2} and tt\to\infty and relations (1.1), (1.2), imply that for every ε>0\varepsilon>0 there exists tε>0t_{\varepsilon}>0 such that if k1k2k_{1}\not=k_{2} then

p𝔻(zj,k1l,t,zj,k2s,t)>1εfor all ttε.p_{\mathbb{D}}(z_{j,k_{1}}^{l,t},z_{j,k_{2}}^{s,t})>1-\varepsilon\quad{\mbox{for all $t\geq t_{\varepsilon}$.}}

This inequality together with (2.30) imply that

(2.32) (Πjt)2/mj(mj1)1εfor all ttε.\left(\Pi_{j}^{t}\right)^{2/m_{j}(m_{j}-1)}\geq 1-\varepsilon\quad{\mbox{for all $t\geq t_{\varepsilon}$.}}

Combining (2.28), (2.31), and (2.32), we obtain the following:

dh(Et)(1ε)k=1n(dh(Ek))αk2for all ttε.d_{h}(E^{t})\geq(1-\varepsilon)\,\prod_{k=1}^{n}\left(d_{h}(E_{k})\right)^{\alpha_{k}^{2}}\quad{\mbox{for all $t\geq t_{\varepsilon}$.}}

The latter inequality together with (2.18) implies that for all ttεt\geq t_{\varepsilon},

(2.33) cap(Et)\displaystyle{\rm{cap}}(E^{t}) =112πlogdh(Et)112πlog((1ε)k=1n(dh(Ek))αk2)\displaystyle=\frac{1}{-\frac{1}{2\pi}\log d_{h}(E^{t})}\geq\frac{1}{-\frac{1}{2\pi}\log\left((1-\varepsilon)\prod_{k=1}^{n}(d_{h}(E_{k}))^{\alpha_{k}^{2}}\right)}
=1k=1n(αk2/cap(Ek))12πlog(1ε)\displaystyle=\frac{1}{\sum_{k=1}^{n}(\alpha_{k}^{2}/{\rm{cap}}(E_{k}))-\frac{1}{2\pi}\log(1-\varepsilon)}
=k=1ncap(Ek)112πlog(1ε)k=1ncap(Ek).\displaystyle=\frac{\sum_{k=1}^{n}{\rm{cap}}(E_{k})}{1-\frac{1}{2\pi}\log(1-\varepsilon)\sum_{k=1}^{n}{\rm{cap}}(E_{k})}.

Since ε>0\varepsilon>0 can be chosen arbitrarily small, it follows from (2.33) that

(2.34) lim inftcap(Et)k=1ncap(Ek).\liminf_{t\to\infty}{\rm{cap}}(E^{t})\geq\sum_{k=1}^{n}{\rm{cap}}(E_{k}).

Finally, equations (2.25) and (2.34) imply (2.23). \Box

Remark 2.35.

We note here that the conformal capacity of a compact set E𝔻E\subset\mathbb{D} is not monotone under hyperbolic dispersion, in general.

To give an example of such non-monotonicity, we consider a family of hedgehogs E(t)E(t), t0t\geq 0, with E(t)E(t) consisting of a fixed central body Cr0(α)={r0eiθ:|θ|α}C_{r_{0}}(\alpha)=\{r_{0}e^{i\theta}:\,|\theta|\leq\alpha\}, 0<r0<10<r_{0}<1, 0<α<π0<\alpha<\pi, and single varying spike E(t)=[r1(t)ei(αt),r2(t)ei(αt)]E(t)=[r_{1}(t)e^{i(\alpha-t)},r_{2}(t)e^{i(\alpha-t)}], which we define as follows.

We put r1(0)=r1r_{1}(0)=r_{1}, r2(0)=r2r_{2}(0)=r_{2}, where r0<r1<r2<1r_{0}<r_{1}<r_{2}<1. Using polarization with respect to appropriate hyperbolic geodesics γ=(eiθ,eiθ)\gamma=(-e^{i\theta},e^{i\theta}), we find that cap(Cr0(α)[r1ei(αt),r2ei(αt)]){\rm cap}(C_{r_{0}}(\alpha)\cup[r_{1}e^{i(\alpha-t)},r_{2}e^{i(\alpha-t)}]) strictly decreases, when tt varies from 0 to α\alpha. At the same time d𝔻(Cr0(α),[r1ei(αt),r2ei(αt)])d_{\mathbb{D}}(C_{r_{0}}(\alpha),[r_{1}e^{i(\alpha-t)},r_{2}e^{i(\alpha-t)}]) and 𝔻([r1ei(αt),r2ei(αt)])\ell_{\mathbb{D}}([r_{1}e^{i(\alpha-t)},r_{2}e^{i(\alpha-t)}]) are constant for 0tα0\leq t\leq\alpha. 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 r1(t)r_{1}(t) such that r1(0)=r1r_{1}(0)=r_{1}, r1<r1(α)<1r_{1}<r_{1}(\alpha)<1, and a function r2(t)r_{2}(t), r1(t)<r2(t)<1r_{1}(t)<r_{2}(t)<1, such that cap(Cr0(α)[r1(t)ei(αt),r2(t)ei(αt)]){\rm cap}(C_{r_{0}}(\alpha)\cup[r_{1}(t)e^{i(\alpha-t)},r_{2}(t)e^{i(\alpha-t)}]) strictly decreases, d𝔻(Cr0(α),[r1(t)ei(αt),r2(t)ei(αt)])d_{\mathbb{D}}(C_{r_{0}}(\alpha),[r_{1}(t)e^{i(\alpha-t)},r_{2}(t)e^{i(\alpha-t)}]) strictly increases, while the hyperbolic length 𝔻([r1(t)ei(αt),r2(t)ei(αt)])\ell_{\mathbb{D}}([r_{1}(t)e^{i(\alpha-t)},r_{2}(t)e^{i(\alpha-t)}]) remains constant on 0tα0\leq t\leq\alpha.

Now, we put E(t)=Cr0(α)[r1(t)ei(αt),r2(t)ei(αt)]E(t)=C_{r_{0}}(\alpha)\cup[r_{1}(t)e^{i(\alpha-t)},r_{2}(t)e^{i(\alpha-t)}] for 0tα0\leq t\leq\alpha and, for tαt\geq\alpha, we define E(t)E(t) as Cr0(α)[r1(t),r2(t)]C_{r_{0}}(\alpha)\cup[r_{1}(t),r_{2}(t)] with r1(t)=(tα+αr1(α)/tr_{1}(t)=(t-\alpha+\alpha r_{1}(\alpha)/t and r2(t)r_{2}(t) such that 𝔻([r1(t),r2(t)])=𝔻([r1,r2])\ell_{\mathbb{D}}([r_{1}(t),r_{2}(t)])=\ell_{\mathbb{D}}([r_{1},r_{2}]). The family of hedgehogs E(t)E(t) defines a dispersion of compact sets Cr0(α)C_{r_{0}}(\alpha) and [r1eiα,r2eiα][r_{1}e^{i\alpha},r_{2}e^{i\alpha}] such that cap(E(t)){\rm cap}(E(t)) strictly decreases on 0tα0\leq t\leq\alpha. 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 cap(E(t)){\rm cap}(E(t)) strictly increases on the interval tαt\geq\alpha.

For our proofs, we need two results on the sequences of compact sets in 𝔻\mathbb{D} convergent in an appropriate sense.

Proposition 2.36 (see, [17, Theorem 1.11]).

Let EkE_{k}, k=1,2,k=1,2,\ldots, be a sequence of compact sets in 𝔻\mathbb{D}, such that Ek+1EkE_{k+1}\subset E_{k} for all k=1,2,k=1,2,\ldots, and let E=k=1EkE=\cap_{k=1}^{\infty}E_{k}. Then

cap(Ek)cap(E)as k.{\rm cap}(E_{k})\to{\rm cap}(E)\quad{\mbox{as $k\to\infty$.}}

To state our next proposition, we recall that the Hausdorff distance between two compact sets K,LK,L in the plane is given by

dH(K,L)=max{dist(x,K),dist(y,L):xL,yK}.d_{\rm H}(K,L)=\max\{{\rm dist}(x,K),{\rm dist}(y,L):x\in L,\;y\in K\}.

The following convergence result follows from [4, Theorem 7].

Proposition 2.37.

For fixed δ>0\delta>0, let EkE_{k}, k=1,2,k=1,2,\ldots, be a sequence of compact sets on the diameter (1,1)(-1,1), each of which consists of a finite number of closed intervals such that the hyperbolic length of each of these intervals is δ\geq\delta. If the sequence EkE_{k} converges in the Hausdorff metric to a compact set E𝔻E\subset\mathbb{D}, then

cap(Ek)cap(E)as k.{\rm cap}(E_{k})\to{\rm cap}(E)\quad{\mbox{as $k\to\infty$.}}

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 1<a<1-1<a<1 and τ>0\uptau>0 are fixed and bb varies in the interval [a,1)[a,1). Let c=c(b)c=c(b), b<c<1b<c<1, be such that 𝔻([b,c])=τ\ell_{\mathbb{D}}([b,c])=\uptau. Let E0(1,a]E_{0}\subset(-1,a] be a compact set consisting of a finite number of non-degenerate intervals and let E(b)=E0[b,c(b)]E(b)=E_{0}\cup[b,c(b)]. Then cap(E(b)){\rm cap}(E(b)) is a continuous function that strictly increases from cap(E0[a,c(a)]){\rm cap}(E_{0}\cup[a,c(a)]) to cap(E0)+cap([a,c(a)]){\rm cap}(E_{0})+{\rm cap}([a,c(a)]), when bb runs from aa to 11.

Proof. The continuity property of cap(E(b)){\rm cap}(E(b)) follows from Proposition 2.37. To prove the monotonicity of cap(E(b)){\rm cap}(E(b)), we consider b1b_{1}, b2b_{2} such that ab1<b2<1a\leq b_{1}<b_{2}<1 and note that c(b1)<c(b2)c(b_{1})<c(b_{2}). Let γ\gamma be a hyperbolic geodesic that is orthogonal to the hyperbolic interval [b1,c(b2)]h[b_{1},c(b_{2})]_{h} at its midpoint. We give an orientation to γ\gamma by marking its complementary hyperbolic halfplane H+H_{+} with b1H+b_{1}\in H_{+}, see Figure 2, which illustrates the proof of this lemma. Notice that under our assumptions, E0H+E_{0}\subset H_{+} and, since reflections with respect to hyperbolic geodesics preserve hyperbolic lengthes, the hyperbolic interval I1=[b1,c(b1)]hI_{1}=[b_{1},c(b_{1})]_{h} coincides with the reflection of I2=[b2,c(b2)]hI_{2}=[b_{2},c(b_{2})]_{h} with respect to γ\gamma. Therefore, the polarization 𝒫γ(E(b2))\mathcal{P}_{\gamma}(E(b_{2})) of E(b2)E(b_{2}) with respect to γ\gamma coincides with the set E(b1)E(b_{1}) if b1E0b_{1}\not\in E_{0} and with the set E(b1){c(b2)}E(b_{1})\cup\{c(b_{2})\} otherwise, and the set 𝒫γ(E(b2))E(b2)=(c(b1),c(b2))h\mathcal{P}_{\gamma}(E(b_{2}))\setminus E(b_{2})=(c(b_{1}),c(b_{2}))_{h} is a non-degenerate interval and thus it has positive logarithmic capacity. Furthermore, since E0H+E_{0}\subset H_{+}, it follows that the set 𝒫γ(E(b2))\mathcal{P}_{\gamma}(E(b_{2})) differs from the reflection of E(b2)E(b_{2}) with respect to γ\gamma by a set of positive logarithmic capacity. So, applying Proposition 2.12, we conclude that cap(E(b1))<cap(E(b2)){\rm cap}(E(b_{1}))<{\rm cap}(E(b_{2})). Thus we proved that the function cap(E(b)){\rm cap}(E(b)) 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. \Box

Remark 3.2.

The proof of Lemma 3.1 remains valid if E0E_{0} is any compact set in the hyperbolic halfplane H+H_{+} defined as in the proof above for the hyperbolic geodesic γ\gamma passing through the point aa such that the set E(b)=E0[b,c(b)]hE(b)=E_{0}\cup[b,c(b)]_{h} satisfies the assumptions of Proposition 2.37.

Refer to caption
Figure 2. Hedgehog with one moving interval.
Remark 3.3.

The non-strict monotonicity property in Lemma 3.1 also follows from the contraction principle of Proposition 2.19.

Next, we will use Lemma 3.1 to prove a lower bound for the conformal capacity of compact sets lying on the diameter of 𝔻\mathbb{D}.

Lemma 3.4.

Let τ>0\uptau>0 and let r(τ)r(\uptau) be defined as in (1.3). If E(1,1)E\subset(-1,1) is a compact set such that 𝔻(E)=τ\ell_{\mathbb{D}}(E)=\uptau, then

(3.5) cap([0,r(τ)])cap(E).{\rm cap}([0,r(\uptau)])\leq{\rm cap}(E).

Equality occurs here if and only if EE coincides with some interval [a,b](1,1)[a,b]\subset(-1,1) up to a set of zero logarithmic capacity.

Proof. (a) Suppose first that EE consists of n2n\geq 2 intervals [ak,bk][a_{k},b_{k}], 0=a1<b1<a2<b2<<an<bn<10=a_{1}<b_{1}<a_{2}<b_{2}<\cdots<a_{n}<b_{n}<1. Let E1E_{1} be the compact set obtained from EE by replacing the pair of intervals [an1,bn1][a_{n-1},b_{n-1}] and [an,bn][a_{n},b_{n}] with a single varying interval [an1,bn1][a_{n-1},b^{\prime}_{n-1}] such that 𝔻([an1,bn1])=𝔻([an1,bn1])+𝔻([an,bn])\ell_{\mathbb{D}}([a_{n-1},b^{\prime}_{n-1}])=\ell_{\mathbb{D}}([a_{n-1},b_{n-1}])+\ell_{\mathbb{D}}([a_{n},b_{n}]). It follows from the monotonicity property of Lemma 3.1 that cap(E1)<cap(E){\rm cap}(E_{1})<{\rm cap}(E). Applying this procedure of merging two intervals into a single interval n1n-1 times, we obtain the inequality (3.5) with the sign of strict inequality.

(b) If EE 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 EE has zero logarithmic capacity we can remove it without changing the conformal capacity and the hyperbolic length of EE. Thus, we assume that EE does not have isolated points. Also, since both the conformal capacity and hyperbolic length are invariant under conformal automorphisms of 𝔻\mathbb{D}, we may assume that min{Rez:zE}=0\min\{{\,\operatorname{Re}\,}z:\,z\in E\}=0.

We will use the following approximation argument. The set (1,1)E(-1,1)\setminus E is an open subset of (1,1)(-1,1) and therefore, in the case under consideration, it is a countably infinite union of open disjoint intervals IkI_{k}, k=1,2,k=1,2,\ldots. We enumerate these intervals such that I1=(1,0)I_{1}=(-1,0) and I2I_{2} has one of its end points at 11. Setting En=(1,1)k=1n+1IkE_{n}=(-1,1)\setminus\cup_{k=1}^{n+1}I_{k}, n=1,2,n=1,2,\dots, we obtain a sequence of compact sets EnE_{n}, each consisting of a finite number of disjoint closed intervals on (1,1)(-1,1), such that En+1EnE_{n+1}\subset E_{n} for all nn and E=n=1EnE=\cap_{n=1}^{\infty}E_{n}. So EE is approximated by a finite union of closed intervals. Therefore, limn𝔻(En)𝔻(E)\lim_{n\to\infty}\ell_{\mathbb{D}}(E_{n})\to\ell_{\mathbb{D}}(E) and, by Proposition 2.36, limncap(En)=cap(E)\lim_{n\to\infty}{\rm cap}(E_{n})={\rm cap}(E).

Let Fn=[0,an]F_{n}=[0,a_{n}] be a closed interval on [0,1)[0,1) such that 𝔻(Fn)=𝔻(En)\ell_{\mathbb{D}}(F_{n})=\ell_{\mathbb{D}}(E_{n}). Then, by part (a) of this proof,

(3.6) cap(Fn)<cap(En).{\rm cap}(F_{n})<{\rm cap}(E_{n}).

Furthermore, Fn+1FnF_{n+1}\subset F_{n} for all nn and n=1Fn=[0,r(τ)]\cap_{n=1}^{\infty}F_{n}=[0,r(\uptau)]. Thus, cap(Fn)cap([0,r(τ)]){\rm cap}(F_{n})\to{\rm cap}([0,r(\uptau)]), by Proposition 2.36. Therefore, passing to the limit in (3.6), we obtain (3.5).

(c) Here we prove the equality statement. If EE coincides with some interval [a,b][a,b] up to a set of zero logarithmic capacity then, by Proposition 2.3, 𝔻([a,b])=𝔻(E)=τ\ell_{\mathbb{D}}([a,b])=\ell_{\mathbb{D}}(E)=\uptau and cap([a,b])=cap(E){\rm cap}([a,b])={\rm cap}(E). This together with the conformal invariance property of the capacity implies that cap(E)=cap([0,r(τ)]){\rm cap}(E)={\rm cap}([0,r(\uptau)]).

Suppose now that (3.5) holds with the sign of equality. Let aa and bb denote the infimum and supremum of the set of Lebesgue density points of EE. We may assume, without loss of generality, that a=0a=0.

If r(τ)<br(\uptau)<b, then the set [0,r(τ)]E[0,r(\uptau)]\setminus E contains an open interval (c1,c2)(c_{1},c_{2}) with 0<c1<c2<r(τ)0<c_{1}<c_{2}<r(\uptau). Let 𝒫γ(E)\mathcal{P}_{\gamma}(E) denote the polarization of EE with respect to a hyperbolic geodesic γ\gamma that is orthogonal to the hyperbolic interval [(c1+c2)/2,b]h[(c_{1}+c_{2})/2,b]_{h} at its midpoint and oriented such that 0H+0\in H_{+}. Notice, that under our assumptions, the set 𝒫γ(E)\mathcal{P}_{\gamma}(E) differs from EE 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 EE with respect to γ\gamma 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) cap(𝒫γ(E))<cap(E).{\rm cap}(\mathcal{P}_{\gamma}(E))<{\rm cap}(E).

Since the polarization with respect to hyperbolic geodesics preserves hyperbolic length, we have l𝔻(𝒫γ(E))=τl_{\mathbb{D}}(\mathcal{P}_{\gamma}(E))=\uptau. Therefore, it follows from the assumption cap(E)=cap([0,r(τ)]){\rm cap}(E)={\rm cap}([0,r(\uptau)]) and our proof in parts (a) and (b) above, that cap(E)cap(Pγ(E){\rm cap}(E)\leq{\rm cap}(P_{\gamma}(E), which contradicts equation (3.7). Since the assumption that r(τ)<br(\uptau)<b leads to a contradiction, we must have b=r(τ)b=r(\uptau).

In the latter case, [0,r(τ)]E[0,r(\uptau)]\subset E and, since cap(E)=cap([0,r(τ)]){\rm cap}(E)={\rm cap}([0,r(\uptau)]), it follows from Proposition 2.3 that E[0,r(τ)]E\setminus[0,r(\uptau)] has zero logarithmic capacity, which completes the proof of the lemma. \Box

Actually, our proof of Lemma 3.4 gives us a more general result, which we state as the following corollary.

Corollary 3.8.

For τ>0\uptau>0 and 1<a<1-1<a<1, let ρ(a,τ)(a,1)\rho(a,\uptau)\in(a,1) be such that 𝔻([a,ρ(a,τ)])=τ\ell_{\mathbb{D}}([a,\rho(a,\uptau)])=\uptau. Let γ\gamma be a hyperbolic geodesic passing through the point aa orthogonally to the diameter (1,1)(-1,1) and oriented such that 1H+-1\in\partial H_{+}.

If E=E0E1E=E_{0}\cup E_{1} is a compact subset in 𝔻\mathbb{D} such that E0E_{0} is a compact subset of H+γH_{+}\cup\gamma and E1[a,1)E_{1}\subset[a,1) is a compact set such that 𝔻(E1)=τ\ell_{\mathbb{D}}(E_{1})=\uptau, then

(3.9) cap(E0[a,ρ(a,τ)])cap(E).{\rm cap}(E_{0}\cup[a,\rho(a,\uptau)])\leq{\rm cap}(E).

If E0E_{0} has positive logarithmic capacity, then equality occurs in (3.9) if and only if E1E_{1} coincides with the interval [a,ρ(a,τ)][a,\rho(a,\uptau)] up to a set of zero logarithmic capacity. Otherwise, equality occurs in (3.9) if and only if E1E_{1} coincides with some interval [b,c][a,1)[b,c]\subset[a,1) such that 𝔻([b,c])=τ\ell_{\mathbb{D}}([b,c])=\uptau up to a set of zero logarithmic capacity.

Proof. Since the considered characteristics of compact sets are invariant under Möbius transformations, we may assume once more that a=0a=0. Notice that the polarization transformations used in the proof of Lemma 3.4 do not change the portion E0E_{0} of the set EE, which lies in the halfspace H+={z𝔻:Rez0}H_{+}=\{z\in\mathbb{D}:\,{\,\operatorname{Re}\,}z\leq 0\}. Therefore, our arguments used in the proof of Lemma 3.4 prove this corollary as well. \Box

As concerns upper bounds for the conformal capacity of compact sets E(1,1)E\subset(-1,1) 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 KK has positive logarithmic capacity; more precisely, log.cap(K)1/9{\rm log.cap}(K)\geq 1/9, see [35, p. 143]. Hence, by Proposition 2.3, the conformal capacity κ=cap(K1/3)\kappa={\rm cap}(K_{1/3}) of the scaled Cantor set K1/3={z: 3zK}K_{1/3}=\{z:\,3z\in K\} is positive.

For nn\in\mathbb{N}, let K1/3n=φn(K1/3)K_{1/3}^{n}=\varphi_{n}(K_{1/3}) denote the image of the scaled Cantor set K1/3K_{1/3} under the Möbius mapping φn(z)=(z+rn)/(1+rnz)\varphi_{n}(z)=(z+r_{n})/(1+r_{n}z), where rn=11n!r_{n}=1-\frac{1}{n!}. Since the hyperbolic length and conformal capacity are invariant under Möbius automorphisms of 𝔻\mathbb{D}, we have 𝔻(K1/3n)=𝔻(K1/3)=0\ell_{\mathbb{D}}(K_{1/3}^{n})=\ell_{\mathbb{D}}(K_{1/3})=0 and cap(K1/3n)=cap(K1/3)=κ{\rm cap}(K_{1/3}^{n})={\rm cap}(K_{1/3})=\kappa for all nn\in\mathbb{N}. A simple calculation shows that

p𝔻(rn+1,rn)=nn+21n!,n.p_{\mathbb{D}}(r_{n+1},r_{n})=\frac{n}{n+2-\frac{1}{n!}},\quad n\in\mathbb{N}.

This implies that the sequence of pseudo-hyperbolic distances p𝔻(rn+1,rn)p_{\mathbb{D}}(r_{n+1},r_{n}) strictly increases and p𝔻(rn+1,rn)1p_{\mathbb{D}}(r_{n+1},r_{n})\to 1 as nn\to\infty. Therefore, the sequence of hyperbolic distances d𝔻(rn+1,rn)d_{\mathbb{D}}(r_{n+1},r_{n}) also strictly increases and d𝔻(rn+1,rn)d_{\mathbb{D}}(r_{n+1},r_{n})\to\infty as nn\to\infty. This implies that K1/3kK_{1/3}^{k} and K1/3lK_{1/3}^{l} are disjoint if klk\not=l and

(3.11) d𝔻(K1/3n+1,K1/3n) as n.d_{\mathbb{D}}(K_{1/3}^{n+1},K_{1/3}^{n})\to\infty\quad{\mbox{ as $n\to\infty$.}}

Given any C>0C>0, we fix jj\in\mathbb{N} such that jκ>Cj\kappa>C. Then, for any mm\in\mathbb{N}, we consider the compact sets K1/3m,j=s=mm+j1K1/3sK_{1/3}^{m,j}=\cup_{s=m}^{m+j-1}K_{1/3}^{s}. Using (3.11) and arguing as in the proof of the limit relation (2.23) of Proposition 2.22, we conclude that cap(K1/3m,j)jκ{\rm cap}(K_{1/3}^{m,j})\to j\kappa as mm\to\infty. The latter shows that for every constant C>0C>0 there are compact sets E(1,1)E\subset(-1,1) such that 𝔻(E)=0\ell_{\mathbb{D}}(E)=0 and cap(E)C{\rm cap}(E)\geq C.

In our previous example, the hyperbolic diameters of sets K1/3m,jK_{1/3}^{m,j} tend to \infty as mm\to\infty. For compact sets with hyperbolic diameters bounded by some constant, say for compact sets on the interval [a,b](1,1)[a,b]\subset(-1,1) such that 𝔻(E)<𝔻([a,b])\ell_{\mathbb{D}}(E)<\ell_{\mathbb{D}}([a,b]), we have the strict inequality cap(E)<cap([a,b]){\rm cap}(E)<{\rm cap}([a,b]), which follows from the fact that [a,b]E[a,b]\setminus E contains a non-empty open interval that is a set of positive logarithmic capacity. In our next example, we show that for every ε>0\varepsilon>0 and a,ba,b and τ\uptau are such that 1<a<b<1-1<a<b<1, 0<τ<𝔻([a,b])0<\uptau<\ell_{\mathbb{D}}([a,b]), there is a compact set E[a,b]E\subset[a,b] such that 𝔻(E)=τ\ell_{\mathbb{D}}(E)=\uptau and cap(E)>cap([a,b])ε{\rm cap}(E)>{\rm cap}([a,b])-\varepsilon.

Example 3.12.

First, we consider a condenser (A(ρ1,ρ),Kn(l))(A(\rho^{-1},\rho),K_{n}(l)) with the domain A(ρ1,ρ)A(\rho^{-1},\rho), where A(ρ1,ρ2)={z:ρ1<|z|<ρ2}A(\rho_{1},\rho_{2})=\{z:\,\rho_{1}<|z|<\rho_{2}\}, 0<ρ1<ρ20<\rho_{1}<\rho_{2}, and a compact set Kn(l)=j=1nKn,j(l)K_{n}(l)=\cup_{j=1}^{n}K_{n,j}(l), where Kn,j(l)={eiθ:|θ2π(j1)/n|l/n}K_{n,j}(l)=\{e^{i\theta}:\,|\theta-2\pi(j-1)/n|\leq l/n\}, j=1,,nj=1,\ldots,n, 0<l<π0<l<\pi.

Let Γ=Γ(ρ,n,l)\Gamma=\Gamma(\rho,n,l) denote the family of curves in A(ρ1,ρ)A(\rho^{-1},\rho) joining the boundary circles of A(ρ1,ρ)A(\rho^{-1},\rho) with the set Kn(l)K_{n}(l) and let Γ0=Γ0(ρ,n,l)\Gamma_{0}=\Gamma_{0}(\rho,n,l) denote the family of curves in the annular sector Sn(ρ)={z: 1<|z|<ρ,|argz|<π/n}S_{n}(\rho)=\{z:\,1<|z|<\rho,\ |\arg z|<\pi/n\} joining the arc {z=ρeiθ:|θ|π/n}\{z=\rho e^{i\theta}:\,|\theta|\leq\pi/n\} with the set Kn,1(l)K_{n,1}(l). 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) cap((A(ρ1,ρ),Kn(l)))=𝖬(Γ)=2n𝖬(Γ0).{\rm cap}((A(\rho^{-1},\rho),K_{n}(l)))=\mathsf{M}(\Gamma)=2n\mathsf{M}(\Gamma_{0}).

To find 𝖬(Γ0)\mathsf{M}(\Gamma_{0}), we consider a function φ(z)=φ2φ1(z)\varphi(z)=\varphi_{2}\circ\varphi_{1}(z) with

(3.14) φ1(z)=i(n𝒦(k)/π)logzandφ2(ζ)=sn(ζ,k),\varphi_{1}(z)=i(n\mathcal{K}(k)/\pi)\log z\quad{\mbox{and}}\quad\varphi_{2}(\zeta)=\mathrm{sn}(\zeta,k),

where the parameter kk, 0<k<10<k<1, of the elliptic sine function is defined by the equation

(3.15) 𝒦(k)𝒦(k)=2nlogρπ.\frac{\mathcal{K}^{\prime}(k)}{\mathcal{K}(k)}=\frac{2n\log\rho}{\pi}.

It is a well-known property of elliptic integrals [2, Theorem 5.13(1)] that if 𝒦(k)/𝒦(k)\mathcal{K}^{\prime}(k)/\mathcal{K}(k)\to\infty, then k0k\to 0 and the following expansion holds:

(3.16) 𝒦(k)𝒦(k)=2πlog4k+o(1).\frac{\mathcal{K}^{\prime}(k)}{\mathcal{K}(k)}=\frac{2}{\pi}\log\frac{4}{k}+o(1).

From (3.15) and (3.16), we obtain that

(3.17) log1k=nlogρlog4+o(1),\log\frac{1}{k}=n\log\rho-\log 4+o(1),

where o(1)0o(1)\to 0 when nn\to\infty.

The function w=φ(z)w=\varphi(z) maps Sn(ρ)S_{n}(\rho) conformally onto the semidisk 𝔻R+={w𝔻R:Imw>0}\mathbb{D}_{R}^{+}=\{w\in\mathbb{D}_{R}:\,{\,\operatorname{Im}\,}w>0\} with R=1/kR=1/\sqrt{k}. Furthermore, this function maps the arc Kn,1(l)K_{n,1}(l) onto an interval [cn(l),cn(l)][-c_{n}(l),c_{n}(l)] with 0<cn(l)<10<c_{n}(l)<1. To find cn(l)c_{n}(l), we note that 𝒦(k)π/2\mathcal{K}(k)\to\pi/2 as k0k\to 0 and that sn(ζ,k)\mathrm{sn}(\zeta,k) converges to sinζ\sin\zeta uniformly on compact subsets of \mathbb{C} as k0k\to 0. Using these relations and equations (3.14), we find that cn(l)=sin(l/2)+o(1)c_{n}(l)=\sin(l/2)+o(1) and, therefore, we have the following asymptotic formula for the logarithmic capacity of the interval [cn(l),cn(l)][-c_{n}(l),c_{n}(l)]:

(3.18) log.cap([cn(l),cn(l)])=(1/2)sin(l/2)+o(1),{\rm{log.cap}}([-c_{n}(l),c_{n}(l)])=(1/2)\sin(l/2)+o(1),

where o(1)0o(1)\to 0 as nn\to\infty.

Let Γ1=Γ1(ρ,n,l)\Gamma_{1}=\Gamma_{1}(\rho,n,l) denote the family of curves in 𝔻R\mathbb{D}_{R} joining the circle 𝕋R\mathbb{T}_{R} with the interval [cn(l),cn(l)][-c_{n}(l),c_{n}(l)]. Conformal invariance and symmetry properties of the modulus of family of curves imply that

(3.19) 𝖬(Γ0)=12𝖬(Γ1).\mathsf{M}(\Gamma_{0})=\frac{1}{2}\mathsf{M}(\Gamma_{1}).

We have the following limit relation between the modulus of Γ1\Gamma_{1} and the logarithmic capacity of [cn(l),cn(l)][-c_{n}(l),c_{n}(l)]:

(3.20) 12πlog(log.cap([cn(l),cn(l)]))=(𝖬(Γ1))112πlogR+o(1),-\frac{1}{2\pi}\log({\rm{log.cap}}([-c_{n}(l),c_{n}(l)]))=(\mathsf{M}(\Gamma_{1}))^{-1}-\frac{1}{2\pi}\log R+o(1),

where o(1)0o(1)\to 0 as RR\to\infty.

Using relations (3.18) and (3.20) with R=1/kR=1/\sqrt{k} and with kk as in the equation (3.17), we find that

(3.21) 𝖬(Γ1)=(12πlogsin(l/2)2+14π(nlogρlog4)+o(1))1.\mathsf{M}(\Gamma_{1})=\left(-\frac{1}{2\pi}\log\frac{\sin(l/2)}{2}+\frac{1}{4\pi}(n\log\rho-\log 4)+o(1)\right)^{-1}.

Finally, combining equations (3.13), (3.19) and (3.21), we obtain the following asymptotic formula for the capacity of the condenser (A(ρ1,ρ),Kn(l))(A(\rho^{-1},\rho),K_{n}(l)):

(3.22) cap((A(ρ1,ρ),Kn(l)))=4πlogρ+o(1),{\rm cap}((A(\rho^{-1},\rho),K_{n}(l)))=\frac{4\pi}{\log\rho}+o(1),

where o(1)0o(1)\to 0 when ρ\rho and ll are fixed and nn\to\infty.

Let s=s(ρ)s=s(\rho), 0<s<10<s<1, be such that

(3.23) cap([s,s])=cap((𝔻ρ,𝔻¯))=2πlogρ.{\rm cap}([-s,s])={\rm cap}((\mathbb{D}_{\rho},\overline{\mathbb{D}}))=\frac{2\pi}{\log\rho}.

Then there is a unique function ψ(z)\psi(z) mapping A(1,ρ)A(1,\rho) conformally onto 𝔻[s,s]\mathbb{D}\setminus[-s,s] such that ψ(ρ)=1\psi(\rho)=1. This function can be extended to a function continuous on A(1,ρ)¯\overline{A(1,\rho)} and such that ψ(z¯)=ψ(z)¯\psi(\overline{z})=\overline{\psi(z)} for all zA(1,ρ)¯z\in\overline{A(1,\rho)}. Thus, ψ(z)\psi(z) maps Kn(l)K_{n}(l) onto a compact set ψ(Kn(l))[s,s]\psi(K_{n}(l))\subset[-s,s].

The same conformal invariance and symmetry properties, which we used earlier in this example, together with equations (3.22) and (3.23) imply that

cap(ψ(Kn(l)))=12cap((A(ρ1,ρ),Kn(l)))=2πlogρ+o(1).{\rm cap}(\psi(K_{n}(l)))=\frac{1}{2}{\rm cap}((A(\rho^{-1},\rho),K_{n}(l)))=\frac{2\pi}{\log\rho}+o(1).

Notice that there is a constant C>0C>0 such that |ψ(eiθ)|C|\psi^{\prime}(e^{i\theta})|\leq C for all θ\theta\in\mathbb{R}. This implies that for every τ>0\uptau>0 there is l0l_{0}, 0<l0<2π0<l_{0}<2\pi, such that 𝔻(ψ(Kn(l)))τ\ell_{\mathbb{D}}(\psi(K_{n}(l)))\leq\uptau for every ll, 0<ll00<l\leq l_{0}, and all n2n\geq 2.

We fix ll, 0<ll00<l\leq l_{0}, and n2n\geq 2 and consider a compact set E[s,s]E\subset[-s,s] such that ψ(Kn(l))E\psi(K_{n}(l))\subset E and 𝔻(E)=τ\ell_{\mathbb{D}}(E)=\uptau, Then, if nn is large enough, we have

cap([s,s])>cap(E)cap(ψ(Kn(l)))=2πlogρ+o(1).{\rm cap}([-s,s])>{\rm cap}(E)\geq{\rm cap}(\psi(K_{n}(l)))=\frac{2\pi}{\log\rho}+o(1).

The latter equation shows that for every τ\uptau, 0<τ<𝔻([s,s])0<\uptau<\ell_{\mathbb{D}}([-s,s]), and every ε>0\varepsilon>0, there is a compact set E[s,s]E\subset[-s,s] such that 𝔻(E)=τ\ell_{\mathbb{D}}(E)=\uptau and cap(E)>cap([s,s])ε{\rm cap}(E)>{\rm cap}([-s,s])-\varepsilon. Since the hyperbolic length and conformal capacity are invariant under Möbius automorphisms of 𝔻\mathbb{D}, compact sets with similar properties exist for every interval [a,b][a,b], 1<a<b<1-1<a<b<1.

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 Kn+(l)[1,1]K_{n}^{+}(l)\subset[-1,1] denote the orthogonal projection of the set Kn(l)K_{n}(l) introduced earlier onto the real axis. Then

log.cap(Kn+(l))=(1/2)(sin(l/2))2/n.{\rm log.cap}(K_{n}^{+}(l))=(1/2)\left(\sin(l/2)\right)^{2/n}.

Taking the limit in this equation as nn\to\infty, we conclude that for every 0<s<20<s<2 and every ε>0\varepsilon>0, there exists a compact set E[1,1]E\subset[-1,1] with Euclidean length ss such that

log.cap(E)>log.cap([1,1])ε=1/2ε.{\rm log.cap}(E)>{\rm log.cap}([-1,1])-\varepsilon=1/2-\varepsilon.

In particular, this answers a question raised in Problem 2 in [10] by showing that the supremum of the transfinite diameters of compact sets E[1,1]E\subset[-1,1] with Euclidean length ss, 0<s<20<s<2, is equal to 1/21/2.

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 EE, in general. In a particular case, when EE is connected, a non-trivial upper bound exists and is given in the following lemma.

Lemma 3.25.

Let L𝔻L\subset\mathbb{D} be a Jordan arc having the hyperbolic length τ>0\uptau>0. Then

(3.26) cap(L)cap([0,r(τ)])=4𝒦((eτ1)/(eτ+1))𝒦((eτ1)/(eτ+1)).{\rm cap}(L)\leq{\rm cap}([0,r(\uptau)])=4\frac{\mathcal{K}((e^{\uptau}-1)/(e^{\uptau}+1))}{\mathcal{K}^{\prime}((e^{\uptau}-1)/(e^{\uptau}+1))}.

Proof. The proof repeats the well-known proof for the logarithmic capacity, see, for example, [35, Theorem 5.3.2]. We consider a parametrization T:[0,r(τ)]LT:\,[0,r(\uptau)]\to L of LL by the hyperbolic arc-length. Then TT is contractive in the hyperbolic metric. Therefore, (3.26) follows from Proposition 2.19. \Box

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 1n41\leq n\leq 4 distinct radial intervals Ik=I(αk)I_{k}=I(\alpha_{k}), k=1,,nk=1,\ldots,n, of the unit disk 𝔻\mathbb{D}. Suppose that EkIkE_{k}\subset I_{k}, k=1,,nk=1,\ldots,n, is a compact set on IkI_{k} such that 𝔻(Ek)=lk\ell_{\mathbb{D}}(E_{k})=l_{k} and that EkIkE_{k}^{*}\subset I_{k} is a hyperbolic interval having one end point at z=0z=0 such that 𝔻(Ek)=lk\ell_{\mathbb{D}}(E_{k}^{*})=l_{k}.

If each of the angles formed by the radial intervals IkI_{k} and IjI_{j}, kjk\not=j, is greater than or equal to π/2\pi/2, then

(3.28) cap(k=1nEk)cap(k=1nEk).{\rm cap}\left(\cup_{k=1}^{n}E_{k}\right)\geq{\rm cap}\left(\cup_{k=1}^{n}E_{k}^{*}\right).

Equality occurs in (3.28) if and only if for each kk, EkE_{k} coincides with EkE_{k}^{*} up to a set of zero logarithmic capacity.

Proof. The proof is the same for all nn. Thus, we assume that n=4n=4. Rotating, if necessary, we may assume that I1=[0,1)I_{1}=[0,1). The diameter (i,i)(-i,i) is a hyperbolic geodesic, which we orient such that 1/2H+1/2\in H_{+}. If n=4n=4, then the angles between the neighboring intervals are equal to π/2\pi/2 and therefore the set K=k=14EkK=\cup_{k=1}^{4}E_{k} can be represented as the union K=E0E1K=E_{0}\cup E_{1} with E0=k=24EkH+¯E_{0}=\cup_{k=2}^{4}E_{k}\subset\overline{H_{+}}. This shows that the sets E0E_{0}, E1E_{1} and K1=E1E2E3E4K_{1}=E_{1}^{*}\cup E_{2}\cup E_{3}\cup E_{4} satisfy the assumptions of Corollary 3.8. Therefore, by this corollary,

cap(K1)=cap(E0E1)cap(E0E1)=cap(K){\rm cap}(K_{1})={\rm cap}(E_{0}\cup E_{1}^{*})\leq{\rm cap}(E_{0}\cup E_{1})={\rm cap}(K)

with the sign of equality if and only if E1E_{1} coincides with E1E_{1}^{*} up to a set of zero logarithmic capacity.

The same argument can be applied successively to the sets K1K_{1}, K2=E1E2E3E4K_{2}=E_{1}^{*}\cup E_{2}^{*}\cup E_{3}\cup E_{4}, K3=E1E2E3E4K_{3}=E_{1}^{*}\cup E_{2}^{*}\cup E_{3}^{*}\cup E_{4} and K4=E1E2E3E4K_{4}=E_{1}^{*}\cup E_{2}^{*}\cup E_{3}^{*}\cup E_{4}^{*} to obtain the inequalities

cap(k=14Ek)=cap(K4)cap(K3)cap(K2)cap(K1)cap(K)=cap(k=14Ek).{\rm cap}\left(\cup_{k=1}^{4}E_{k}^{*}\right)={\rm cap}(K_{4})\leq{\rm cap}(K_{3})\leq{\rm cap}(K_{2})\leq{\rm cap}(K_{1})\leq{\rm cap}(K)={\rm cap}\left(\cup_{k=1}^{4}E_{k}\right).

Moreover, equality occurs in any one of these inequalities if and only if the corresponding sets EkE_{k} and EkE_{k}^{*} coincide up to a set of zero logarithmic capacity. Thus, the theorem is proved. \Box

Remark 3.29.

We want to stress once more that the inequality (3.28) also follows from the contraction principle of Proposition 2.19. Indeed, under the assumptions that all angles between radial intervals are π/2\geq\pi/2, there is always a contraction φ:k=1nEkk=1nEk\varphi:\cup_{k=1}^{n}E_{k}\to\cup_{k=1}^{n}E_{k}^{*}.

If the angle between some radial intervals IkI_{k} and IjI_{j} is smaller than π/2\pi/2, then both our proofs, with polarization or with the contraction principle, fail even in the simplest case of two intervals I1I_{1} and I2I_{2} and when each of the sets E1I1E_{1}\subset I_{1} and E2I2E_{2}\subset I_{2} 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 0<r<10<r<1, 0<s<0<s<\infty, and 0<α<π/20<\alpha<\pi/2 and varying 0t<10\leq t<1, let E(t)=[0,r]{τeiα:tτd(s,t)}E(t)=[0,r]\cup\{\uptau e^{i\alpha}:t\leq\uptau\leq d(s,t)\} with d(s,t)d(s,t) such that 𝔻([t,d(s,t)])=s\ell_{\mathbb{D}}([t,d(s,t)])=s. Prove (or disprove) that cap(E(t)){\rm cap}(E(t)) strictly increases on the interval 0t<10\leq t<1.

Refer to caption
Figure 3. Hyperbolic capacity of two intervals.

As one can see from our proof of Theorem 3.27, the restriction n4n\leq 4 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 <π/2<\pi/2, 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 EkE_{k}, k=1,,nk=1,\ldots,n, be compact sets on the radial intervals Ik=[0,eiβk)I_{k}=[0,e^{i\beta_{k}}), 0=β1<β2<<βn<βn+1=2π0=\beta_{1}<\beta_{2}<\ldots<\beta_{n}<\beta_{n+1}=2\pi, such that

𝔻(Ek)2logcotα2,k=1,,n,\ell_{\mathbb{D}}(E_{k})\geq 2\log\cot\frac{\alpha}{2},\quad k=1,\ldots,n,

where α\alpha stands for the minimal angle between the intervals IkI_{k}. Then

(3.32) cap(k=1nEk)cap(k=1nEk),{\rm cap}\left(\cup_{k=1}^{n}E_{k}\right)\geq{\rm cap}\left(\cup_{k=1}^{n}E_{k}^{*}\right),

where EkIkE_{k}^{*}\subset I_{k} is a hyperbolic interval having one end point at z=0z=0 such that 𝔻(Ek)=𝔻(Ek),k=1,,n\ell_{\mathbb{D}}(E_{k}^{*})=\ell_{\mathbb{D}}(E_{k}),k=1,\ldots,n.

Equality occurs here if and only if for each kk, EkE_{k} coincides with EkE_{k}^{*} up to a set of zero logarithmic capacity.

Proof. (a) Let τα=logcot(α/2)\uptau_{\alpha}=\log\cot(\alpha/2), rα=r(τα)=secαtanαr_{\alpha}=r(\uptau_{\alpha})=\sec\alpha-\tan\alpha and let γα\gamma_{\alpha} be the hyperbolic geodesic orthogonal to the interval (1,1)(-1,1) at z=rαz=r_{\alpha} and oriented such that 0 belongs to the halfplane H+=H+(γα)H_{+}=H_{+}(\gamma_{\alpha}). An easy calculation shows that γα\gamma_{\alpha} has its endpoints at the points e±iαe^{\pm i\alpha}. Consider sets E1+=E1H+¯E_{1}^{+}=E_{1}\cap\overline{H_{+}}, E1=E1H+E_{1}^{-}=E_{1}\setminus H_{+} and E0=E1+(k=2nEk)E_{0}=E_{1}^{+}\cup\left(\cup_{k=2}^{n}E_{k}\right). Let E~1\widetilde{E}_{1}^{-} denote the closed hyperbolic interval with the initial point at z=r(α)z=r(\alpha) such that E~1[rα,1)\widetilde{E}_{1}^{-}\subset[r_{\alpha},1) and 𝔻(E~1)=𝔻(E1)\ell_{\mathbb{D}}(\widetilde{E}_{1}^{-})=\ell_{\mathbb{D}}(E_{1}^{-}) and let E~1=E1+E~1\widetilde{E}_{1}=E_{1}^{+}\cup\widetilde{E}_{1}^{-}. Since the minimal angle between the intervals IkI_{k} is α\alpha and γα\gamma_{\alpha} has its endpoints at e±iαe^{\pm i\alpha}, it follows that E0H+¯E_{0}\subset\overline{H_{+}}. Therefore, we can apply Corollary 3.8 to obtain the following:

(3.33) cap(E0E~1)=cap(E~1(k=2nEk))cap(E){\rm cap}(E_{0}\cup\widetilde{E}_{1}^{-})={\rm cap}(\widetilde{E}_{1}\cup(\cup_{k=2}^{n}E_{k}))\leq{\rm cap}(E)

with the sign of equality if and only if E1E_{1}^{-} coincides with E~1\widetilde{E}_{1}^{-} up to a set of zero logarithmic capacity.

If E1+=[0,rα]E_{1}^{+}=[0,r_{\alpha}], then E1=E1+E~1E_{1}^{*}=E_{1}^{+}\cup\widetilde{E}_{1}^{-} and the inequality in (3.33) is equivalent to the inequality

(3.34) cap(E1(k=2nEk))cap(E){\rm cap}\left(E_{1}^{*}\cup\left(\cup_{k=2}^{n}E_{k}\right)\right)\leq{\rm cap}(E)

with the sign of equality if and only if E1E_{1}^{*} coincides with E1E_{1} up to a set of zero logarithmic capacity.

(b) If E1+[0,rα]E_{1}^{+}\not=[0,r_{\alpha}], then [0,rα]E1+[0,r_{\alpha}]\setminus E_{1}^{+} contains an open interval. In this case 𝔻(E~1)>𝔻([0,rα])\ell_{\mathbb{D}}(\widetilde{E}_{1}^{-})>\ell_{\mathbb{D}}([0,r_{\alpha}]). Let b1b_{1}, rα<b1<1r_{\alpha}<b_{1}<1, denote the end point of the interval E~1\widetilde{E}_{1}^{-} and let c1c_{1} denote the midpoint of the hyperbolic interval [0,b1]h[0,b_{1}]_{h}. Notice that under our assumptions rα<c1r_{\alpha}<c_{1}. Let γ1\gamma_{1} be the hyperbolic geodesic orthogonal to the diameter (1,1)(-1,1) at z=c1z=c_{1} and oriented such that 0H+(γ1)0\in H_{+}(\gamma_{1}). Let E^1=𝒫γ1(E~1)\widehat{E}_{1}=\mathcal{P}_{\gamma_{1}}(\widetilde{E}_{1}) and E^=𝒫γ1(E~1((k=2n(Ek)))\widehat{E}=\mathcal{P}_{\gamma_{1}}(\widetilde{E}_{1}\cup(\cup(_{k=2}^{n}(E_{k}))) denote polarizations of the corresponding sets with respect to γ1\gamma_{1}. Since rα<c1r_{\alpha}<c_{1} and therefore k=2nEkH+(γ1)\cup_{k=2}^{n}E_{k}\subset H_{+}(\gamma_{1}), we have E^1[0,rα]\widehat{E}_{1}\supset[0,r_{\alpha}] and E^=E^1(k=1nEk)\widehat{E}=\widehat{E}_{1}\cup(\cup_{k=1}^{n}E_{k}). Applying Proposition 2.12 and using (3.33), we conclude that

(3.35) cap(E^1(k=2nEk))<cap(E~1(k=2nEk))cap(E){\rm cap}(\widehat{E}_{1}\cup(\cup_{k=2}^{n}E_{k}))<{\rm cap}(\widetilde{E}_{1}\cup(\cup_{k=2}^{n}E_{k}))\leq{\rm cap}(E)

with the sign of strict inequality in the first inequality because E^1\widehat{E}_{1} differs from E~1\widetilde{E}_{1} and from its reflection with respect to γ1\gamma_{1} by an open interval and therefore by a set of positive logarithmic capacity.

Since 𝔻(E^1)=𝔻(E1)\ell_{\mathbb{D}}(\widehat{E}_{1})=\ell_{\mathbb{D}}(E_{1}) and E^1[0,rα]\widehat{E}_{1}\supset[0,r_{\alpha}], using (3.35) and applying the same arguments as in part (a) of this proof to the set E^=E^1(k=2n)\widehat{E}=\widehat{E}_{1}\cup(\cup_{k=2}^{n}), we conclude that in the case E1+[0,rα]E_{1}^{+}\not=[0,r_{\alpha}] 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 kk, EkE_{k} coincides with EkE_{k}^{*} up to a set of zero logarithmic capacity. \Box

For compact sets E1E_{1} and E2E_{2}, lying on two orthogonal diameters of 𝔻\mathbb{D}, we have the following result.

Theorem 3.36.

Let E1(1,1)E_{1}\subset(-1,1), E2(i,i)E_{2}\subset(-i,i) be compact sets and let rkr_{k}, 0<rk<10<r_{k}<1, and lk>0l_{k}>0 be such that

𝔻(Ek)=𝔻([rk,rk])=2lk,k=1,2.\ell_{\mathbb{D}}(E_{k})=\ell_{\mathbb{D}}([-r_{k},r_{k}])=2l_{k},\quad k=1,2.

Then

(3.37) cap(E1E2)cap([r1,r1][ir2,ir2])cap([r0,r0][ir0,ir0]),{\rm cap}(E_{1}\cup E_{2})\geq{\rm cap}([-r_{1},r_{1}]\cup[-ir_{2},ir_{2}])\geq{\rm cap}([-r_{0},r_{0}]\cup[-ir_{0},ir_{0}]),

where 0<r0<10<r_{0}<1 is such that 𝔻([r0,r0])=l1+l2\ell_{\mathbb{D}}([-r_{0},r_{0}])=l_{1}+l_{2}.

Equality occurs in the first inequality if and only if E1E_{1} coincides with [r1,r1][-r_{1},r_{1}] and E2E_{2} coincides with [ir2,ir2][-ir_{2},ir_{2}] up to a set of zero logarithmic capacity. Equality occurs in the second inequality if and only if l1=l2l_{1}=l_{2}.

Proof. Let 0rk±<10\leq r_{k}^{\pm}<1, k=1,2k=1,2, be such that the following holds:

𝔻([0,r1+])=𝔻(E1[0,1)),𝔻([0,r1])=𝔻(E1[0,1)),\ell_{\mathbb{D}}([0,r_{1}^{+}])=\ell_{\mathbb{D}}(E_{1}\cap[0,1)),\quad\ell_{\mathbb{D}}([0,-r_{1}^{-}])=\ell_{\mathbb{D}}(E_{1}\cap[0,-1)),
𝔻([0,r2+])=𝔻(E2[0,i)),𝔻([0,r2])=𝔻(E2[0,i)).\ell_{\mathbb{D}}([0,r_{2}^{+}])=\ell_{\mathbb{D}}(E_{2}\cap[0,i)),\quad\ell_{\mathbb{D}}([0,-r_{2}^{-}])=\ell_{\mathbb{D}}(E_{2}\cap[0,-i)).

Then, by Theorem 3.27,

(3.38) cap(E1E2)cap([0,r1+][0,r1][0,ir2+)[0,ir2]){\rm cap}(E_{1}\cup E_{2})\geq{\rm cap}([0,r_{1}^{+}]\cup[0,-r_{1}^{-}]\cup[0,ir_{2}^{+})\cup[0,-ir_{2}^{-}])

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 r1+r1r_{1}^{+}\not=r_{1}^{-}, say r1+>r1r_{1}^{+}>r_{1}^{-}. Then r1+>r1>r1r_{1}^{+}>r_{1}>r_{1}^{-}. Let 𝒫γ\mathcal{P}_{\gamma} denote polarization with respect to the geodesic γ\gamma that is orthogonal to the hyperbolic interval [r1,r1+][-r_{1},r_{1}^{+}] at its midpoint cc, 0<c<r10<c<r_{1}. We assume here that γ\gamma is oriented such that 0Hγ+0\in H_{\gamma}^{+}. Under our assumptions, 𝒫γ([r1,r1+][ir2,ir2+])=[r1,r1][ir2,ir2+]\mathcal{P}_{\gamma}([-r_{1}^{-},r_{1}^{+}]\cup[-ir_{2}^{-},ir_{2}^{+}])=[-r_{1},r_{1}]\cup[-ir_{2}^{-},ir_{2}^{+}]. Since the set [r1,r1+][r1,r1][-r_{1}^{-},r_{1}^{+}]\setminus[-r_{1},r_{1}] has positive logarithmic capacity, it follows from Proposition 2.12 that

(3.39) cap([r1,r1+][ir2,ir2+])>cap([r1,r1][ir2,ir2+]).{\rm cap}([-r_{1}^{-},r_{1}^{+}]\cup[-ir_{2}^{-},ir_{2}^{+}])>{\rm cap}([-r_{1},r_{1}]\cup[-ir_{2}^{-},ir_{2}^{+}]).

The same polarization argument can be applied to show that, if r2r2+r_{2}^{-}\not=r_{2}^{+}, then

(3.40) cap([r1,r1][ir2,ir2+])>cap([r1,r1][ir2,ir2]).{\rm cap}([-r_{1},r_{1}]\cup[-ir_{2}^{-},ir_{2}^{+}])>{\rm cap}([-r_{1},r_{1}]\cup[-ir_{2},ir_{2}]).

Combining inequalities (3.38)–(3.40), we obtain the first inequality in (3.37) with the sign of equality if and only if E1E_{1} coincides with [r1,r1][-r_{1},r_{1}] and E2E_{2} coincides with [ir2,ir2][-ir_{2},ir_{2}] up to a set of zero logarithmic capacity.

To prove the second inequality in (3.37), we use the conformal mapping

g(z)=(z2+r22)/(1+r22z2)g(z)=\sqrt{(z^{2}+r_{2}^{2})/(1+r_{2}^{2}z^{2})}

from the doubly connected domain 𝔻([r1,r1][ir2,ir2])\mathbb{D}\setminus([-r_{1},r_{1}]\cup[-ir_{2},ir_{2}]) onto 𝔻([r,r])\mathbb{D}\setminus([-r,r]) with

r=(r12+r22)/(1+r12r22).r=\sqrt{(r_{1}^{2}+r_{2}^{2})/(1+r_{1}^{2}r_{2}^{2})}.

We note that conformal mappings preserve capacity of condensers and that the function cap([r,r])=cap([r1,r1][ir2,ir2]){\rm cap}([-r,r])={\rm cap}([-r_{1},r_{1}]\cup[-ir_{2},ir_{2}]) is strictly increasing on the interval 0r<10\leq r<1. Furthermore, it follows from formulas (1.3) that the sum l1+l2l_{1}+l_{2} of the hyperbolic lengths defined in the theorem is constant if and only if the following product is constant:

(3.41) 1+r11r11+r21r2=C,\frac{1+r_{1}}{1-r_{1}}\cdot\frac{1+r_{2}}{1-r_{2}}=C,

where CC is constant.

Our goal now is to minimise the function F=F(r1,r2)F=F(r_{1},r_{2}), defined by

F=(r12+r22)/(1+r12r22),F=(r_{1}^{2}+r_{2}^{2})/(1+r_{1}^{2}r_{2}^{2}),

under the constraint (3.41).

Introducing new variables u=(1+r1)/(1r1)u=(1+r_{1})/(1-r_{1}), v=(1+r2)/(1r2)v=(1+r_{2})/(1-r_{2}) and w=u2+v2w=u^{2}+v^{2}, we can express FF in terms of these variables as follows:

F=(u2+v2)+(14C+C2)(u2+v2)+(1+4C+C2)=w+(14C+C2)w+(1+4C+C2),F=\frac{(u^{2}+v^{2})+(1-4C+C^{2})}{(u^{2}+v^{2})+(1+4C+C^{2})}=\frac{w+(1-4C+C^{2})}{w+(1+4C+C^{2})},

which we have to minimize FF under the constraint uv=Cuv=C. Differentiating, we find that

ddwF=4C(w+(1+4C+C2))2>0.\frac{d}{dw}F=\frac{4C}{(w+(1+4C+C^{2}))^{2}}>0.

Therefore, FF takes its minimal value when w=u2+y2w=u^{2}+y^{2} is as small as possible. By the classical arithmetic-geometric mean inequality u2+v2>2uv=2Cu^{2}+v^{2}>2uv=2C, unless u=vu=v. Therefore, the minimal value of FF under the constraint uv=Cuv=C occurs when u=vu=v. The latter implies that cap([r1,r1][ir2,ir2])cap([r0,r0][ir0,ir0]){\rm cap}([-r_{1},r_{1}]\cup[-ir_{2},ir_{2}])\geq{\rm cap}([-r_{0},r_{0}]\cup[-ir_{0},ir_{0}]) with the sign of equality if and only if r1=r2r_{1}=r_{2}. This proves the second inequality in (3.37). \Box

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: E1=[0,1/2][0,i/2]E_{1}=[0,1/2]\cup[0,i/2] and E2=[0,1/2][0,i/4]E_{2}=[0,1/2]\cup[0,i/4]. Two Steiner symmetrizations, chosen appropriately, transform these sets to the sets E1=[1/4,1/4][i/4,i/4]E_{1}^{*}=[-1/4,1/4]\cup[-i/4,i/4] and E2=[1/4,1/4][i/8,i/8]E_{2}^{*}=[-1/4,1/4]\cup[-i/8,i/8], respectively. The first inequality in (3.37) compares the conformal capacities of E1E_{1} and E2E_{2} with the conformal capacities of the sets

E1o=[(23),23][i(23),i(23)],E_{1}^{o}=[-(2-\sqrt{3}),2-\sqrt{3}]\cup[-i(2-\sqrt{3}),i(2-\sqrt{3})],
E2o=[(23),23][i(415),i(415)].E_{2}^{o}=[-(2-\sqrt{3}),2-\sqrt{3}]\cup[-i(4-\sqrt{15}),i(4-\sqrt{15})].

Numerical computation gives the following approximation and bounds for cap(Ek){\rm cap}(E_{k}), k=1,2k=1,2,

cap(E1)3.62589<cap(E1o)3.77702<cap(E1)4.28254,{\rm cap}(E_{1}^{*})\approx 3.62589<{\rm cap}(E_{1}^{o})\approx 3.77702<{\rm cap}(E_{1})\approx 4.28254,
cap(E2)3.19333<cap(E2o)3.29244<cap(E2)3.60548.{\rm cap}(E_{2}^{*})\approx 3.19333<{\rm cap}(E_{2}^{o})\approx 3.29244<{\rm cap}(E_{2})\approx 3.60548.

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 E0E_{0} sitting in the disk 𝔻¯r\overline{\mathbb{D}}_{r}, 0<r<10<r<1, as it is shown in Figure 4, also may be useful in applications.

Theorem 3.43.

Let E0E_{0} be a compact set in the disk 𝔻¯r\overline{\mathbb{D}}_{r}, 0<r<10<r<1. Let EkE_{k}, k=1,,nk=1,\ldots,n, be compact sets on the radial intervals Ik=[reiβk,eiβk]I_{k}=[re^{i\beta_{k}},e^{i\beta_{k}}], 0=β1<β2<<βn<βn+1=2π0=\beta_{1}<\beta_{2}<\ldots<\beta_{n}<\beta_{n+1}=2\pi, such that

𝔻(Ek)2log(cotα21r1+r),k=1,,n,\ell_{\mathbb{D}}(E_{k})\geq 2\log\left(\cot\frac{\alpha}{2}\frac{1-r}{1+r}\right),\quad k=1,\ldots,n,

where α\alpha stands for the minimal angle between the intervals IkI_{k}. Then

(3.44) cap(k=0nEk)cap(k=0nEk),{\rm cap}(\cup_{k=0}^{n}E_{k})\geq{\rm cap}(\cup_{k=0}^{n}E_{k}^{*}),

where E0=E0E_{0}^{*}=E_{0} and EkIkE_{k}^{*}\subset I_{k} is a hyperbolic interval having one end point at z=reiβkz=re^{i\beta_{k}} such that 𝔻(Ek)=𝔻(Ek)\ell_{\mathbb{D}}(E_{k}^{*})=\ell_{\mathbb{D}}(E_{k}), k=1,,nk=1,\ldots,n.

Equality occurs in (3.44) if and only if for each k=1,,nk=1,\ldots,n, EkE_{k} coincides with EkE_{k}^{*} up to a set of zero logarithmic capacity.

Refer to caption
Figure 4. Hedgehogs with central body E0E_{0} and spikes on five intervals.

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 ρ=ρ(r,α)\rho=\rho(r,\alpha), r<ρ<1r<\rho<1, be such that

𝔻([r,ρ])=2log(cotα21r1+r)\ell_{\mathbb{D}}([r,\rho])=2\log\left(\cot\frac{\alpha}{2}\frac{1-r}{1+r}\right)

and let cc be the middle point of the hyperbolic interval [r,ρ]h[r,\rho]_{h}. Then every polarization 𝒫γ\mathcal{P}_{\gamma} performed with respect to the geodesic γ\gamma that is orthogonal to the radial interval [0,eiβk)[0,e^{i\beta_{k}}) and crosses it at some point c1c_{1} such that c<|c1|<1c<|c_{1}|<1, and such that 0H+(γ)0\in H_{+}(\gamma), satisfies the following properties:

𝒫γ(Ek)[reiβk,eiβk)and𝒫γ(Ej)=Ejif jk.\mathcal{P}_{\gamma}(E_{k})\subset[re^{i\beta_{k}},e^{i\beta_{k}})\quad{\mbox{and}}\quad\mathcal{P}_{\gamma}(E_{j})=E_{j}\quad{\mbox{if $j\not=k$.}}

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. \Box

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 E0E_{0} be a compact subset of 𝕋=𝔻\mathbb{T}=\partial\mathbb{D} and let ρ(θ)0\rho(\theta)\geq 0 be an upper semicontinuous function on E0E_{0}. For 0<r<10<r<1, let E(r)E(r) be a compact set in 𝔻\mathbb{D} such that the intersection E(r)[0,eiθ]E(r)\cap[0,e^{i\theta}] is empty if eiθE0e^{i\theta}\not\in E_{0} and it is an interval [reiθ,seiθ][re^{i\theta},se^{i\theta}] with ss, rs<1r\leq s<1, such that 𝔻([r,s])=ρ(θ)\ell_{\mathbb{D}}([r,s])=\rho(\theta), if eiθE0e^{i\theta}\in E_{0}.

Then the conformal capacity cap(E(r)){\rm cap}(E(r)) is an increasing function on the interval 0r<10\leq r<1.

Proof. Let r0r_{0}, 0<r0<10<r_{0}<1, be fixed and let A~(r0)={z:r0|z|<1}\widetilde{A}(r_{0})=\{z:\,r_{0}\leq|z|<1\}. Consider the function φr:A~(r0)𝔻\varphi_{r}:\widetilde{A}(r_{0})\to\mathbb{D} defined as follows: if z=seiθA~(r0)z=se^{i\theta}\in\widetilde{A}(r_{0}), then argφr(z)=θ\arg\varphi_{r}(z)=\theta and d𝔻(reiθ,φr(z))=d𝔻(r0eiθ,z)d_{\mathbb{D}}(re^{i\theta},\varphi_{r}(z))=d_{\mathbb{D}}(r_{0}e^{i\theta},z). We claim that, if 0rr00\leq r\leq r_{0}, then φr\varphi_{r} is a hyperbolic contraction on A~(r0)\widetilde{A}(r_{0}).

To prove this claim, we fix two points z1,z2A~(r0)z_{1},z_{2}\in\widetilde{A}(r_{0}) and consider the hyperbolic distance d𝔻(φr(z1),φr(z2))d_{\mathbb{D}}(\varphi_{r}(z_{1}),\varphi_{r}(z_{2})). Our claim will be proved if we show that this hyperbolic distance or, equivalently, the pseudo-hyperbolic distance

(3.46) p𝔻(φr(z1),φr(z2))=|φr(z1)φr(z2)1φr(z1)φr(z2)¯|,p_{\mathbb{D}}(\varphi_{r}(z_{1}),\varphi_{r}(z_{2}))=\left|\frac{\varphi_{r}(z_{1})-\varphi_{r}(z_{2})}{1-\varphi_{r}(z_{1})\overline{\varphi_{r}(z_{2})}}\right|,

is an increasing function of rr on the interval 0rr00\leq r\leq r_{0}.

Rotating, if necessary, we may assume that

z1=a0eiα,z2=b0eiα,where r0a0,b0<10απ/2.z_{1}=a_{0}e^{i\alpha},\ \ z_{2}=b_{0}e^{-i\alpha},\quad{\mbox{where $r_{0}\leq a_{0},b_{0}<1$, $0\leq\alpha\leq\pi/2$.}}

Then

φr(z1)=aeiα,φr(z2)=beiα,\varphi_{r}(z_{1})=ae^{i\alpha},\quad\varphi_{r}(z_{2})=be^{-i\alpha},

where a=a(r)a=a(r), b=b(r)b=b(r) are functions of rr.

Consider the function F(r)=(p𝔻(aeiα,beiα))2F(r)=\left(p_{\mathbb{D}}(ae^{i\alpha},be^{-i\alpha})\right)^{2}. After some algebra, we find that

(3.47) F(r)=a2(r)2a(r)b(r)cosα+b2(r)(1a(r)b(r))2.F(r)=\frac{a^{2}(r)-2a(r)b(r)\cos\alpha+b^{2}(r)}{(1-a(r)b(r))^{2}}.

Since for fixed ss and θ\theta, r0s<1r_{0}\leq s<1, θ\theta\in\mathbb{R}, the hyperbolic distance between the points reiθre^{i\theta} and φr(seiθ)\varphi_{r}(se^{i\theta}) is the same for all rr in the interval 0rr00\leq r\leq r_{0}, it follows that

(3.48) dadr=1a21r2anddbdr=1b21r2.\frac{da}{dr}=\frac{1-a^{2}}{1-r^{2}}\quad{\mbox{and}}\quad\frac{db}{dr}=\frac{1-b^{2}}{1-r^{2}}.

Differentiating (3.47) and using formulas (3.48), we find:

(3.49) dFdr=2(1r2)(1ab)3Φ(a,b,α),\frac{dF}{dr}=\frac{2}{(1-r^{2})(1-ab)^{3}}\,\Phi(a,b,\alpha),

where Φ=Φ(a,b,α)\Phi=\Phi(a,b,\alpha) is the following function:

Φ=(a(1a2)(a(1b2)+b(1a2))cosα+b(1b2))(1ab)\Phi=(a(1-a^{2})-(a(1-b^{2})+b(1-a^{2}))\cos\alpha+b(1-b^{2}))(1-ab)
+(a22abcosα+b2)(a(1b2)+b(1a2)).+(a^{2}-2ab\cos\alpha+b^{2})(a(1-b^{2})+b(1-a^{2})).\ \ \ \

It is clear that Φ(a,b,α)\Phi(a,b,\alpha) is an increasing function of α\alpha on the interval 0απ/20\leq\alpha\leq\pi/2. After simple calculation, we find that Φ(a,b,0)=0\Phi(a,b,0)=0 and therefore

Φ(a,b,α)Φ(a,b,0)=0.\Phi(a,b,\alpha)\geq\Phi(a,b,0)=0.

This, together with (3.49), implies that dF(r)dr0\frac{dF(r)}{dr}\geq 0 and therefore the pseudo-hyperbolic distance in equation (3.46) decreases, when rr decreases from r0r_{0} to 0. Therefore, our claim that φr\varphi_{r} is a hyperbolic contraction is proved. By Proposition 2.19, the latter implies that cap(E(r)){\rm cap}(E(r)) decreases when rr decreases from r0r_{0} to 0, which proves the theorem. \Box

4. Extremal properties of hedgehogs on evenly distributed radial intervals

In this section we consider problems with extremal configurations lying on n2n\geq 2 radial intervals Ik={z=te2πi(k1)/n: 0t1}I^{*}_{k}=\{z=te^{2\pi i(k-1)/n}:\,0\leq t\leq 1\}, k=1,,nk=1,\ldots,n. Since the intervals IkI_{k} are evenly distributed in 𝔻\mathbb{D} it is expected that extremal configurations possess rotational symmetry by angle 2π/n2\pi/n.

First, we prove a theorem that generalizes the second inequality of Theorem 3.36 for sets lying on 2n42n\geq 4 diameters.

Theorem 4.1.

Let 0<rk<10<r_{k}<1, k=1,2k=1,2 and let rr, 0<r<10<r<1, be such that

(4.2) τ=𝔻([0,r])=12(𝔻([0,r1])+𝔻([0,r2]))\uptau=\ell_{\mathbb{D}}([0,r])=\frac{1}{2}(\ell_{\mathbb{D}}([0,r_{1}])+\ell_{\mathbb{D}}([0,r_{2}]))

and, for n2n\geq 2, let

En(r1,r2)=(k=0n1eπik/n[r1,r1])(k=0n1eπi(2k+1)/2n[r2,r2]).E_{n}(r_{1},r_{2})=(\cup_{k=0}^{n-1}e^{\pi ik/n}[-r_{1},r_{1}])\cup(\cup_{k=0}^{n-1}e^{\pi i(2k+1)/2n}[-r_{2},r_{2}]).

Then

(4.3) cap(En(r1,r2))=8n𝒦(κ)𝒦(κ),whereκ=r12n+r22n1+r12nr22n.{\rm cap}(E_{n}(r_{1},r_{2}))=8n\frac{\mathcal{K}(\kappa)}{\mathcal{K}^{\prime}(\kappa)},\quad{\mbox{where}}\quad\kappa=\frac{r_{1}^{2n}+r_{2}^{2n}}{1+r_{1}^{2n}r_{2}^{2n}}.

Furthermore, the following inequality holds:

(4.4) cap(En(r1,r2))cap(k=02n1eπik/2n[r,r]).{\rm cap}(E_{n}(r_{1},r_{2}))\geq{\rm cap}(\cup_{k=0}^{2n-1}e^{\pi ik/2n}[-r,r]).

Equality occurs in (4.4) if and only if r1=r2r_{1}=r_{2}.

Proof. To establish (4.3), we use the function φ=φ2φ1\varphi=\varphi_{2}\circ\varphi_{1}, where φ1(z)=z2n\varphi_{1}(z)=z^{2n}, φ2(z)=(z+r22n)/(1+r22nz)\varphi_{2}(z)=(z+r_{2}^{2n})/(1+r_{2}^{2n}z), which maps the sector S={z𝔻:|argz|<π/2n}S=\{z\in\mathbb{D}:\,|\arg z|<\pi/2n\} conformally onto 𝔻\mathbb{D} slit along the interval [1,0][-1,0]. Using symmetry properties of En(r1,r2)E_{n}(r_{1},r_{2}), we find that

cap(En(r1,r2))=2ncap([0,(r12n+r22n)/(1+r12nr22n)]).{\rm cap}(E_{n}(r_{1},r_{2}))=2n\,{\rm cap}([0,(r_{1}^{2n}+r_{2}^{2n})/(1+r_{1}^{2n}r_{2}^{2n})]).

This together with (1.10) gives (4.3).

One way to prove the monotonicity property of cap(En(r1,r2)){\rm cap}(E_{n}(r_{1},r_{2})) 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

g(z)=z2n+r22n1+r22nz2ng(z)=\frac{z^{2n}+r_{2}^{2n}}{1+r_{2}^{2n}z^{2n}}

to map the domain {z𝔻En(r1,r2):|argz|<π/2n}\{z\in\mathbb{D}\setminus E_{n}(r_{1},r_{2}):\,|\arg z|<\pi/2n\} conformally onto the unit disk slit along the interval (1,Fn](-1,F_{n}], where Fn=Fn(r1,r2)F_{n}=F_{n}(r_{1},r_{2}) is defined as

Fn=r12n+r22n1+r12nr22n.F_{n}=\frac{r_{1}^{2n}+r_{2}^{2n}}{1+r_{1}^{2n}r_{2}^{2n}}.

Since En(r1,r2)E_{n}(r_{1},r_{2}) possesses 2n2n-fold rotational symmetry about 0, it follows from the symmetry principle for the module of family of curves, that

cap(En(r1,r2))=2ncap([0,Fn(r1,r2]).{\rm cap}(E_{n}(r_{1},r_{2}))=2n\,{\rm cap}([0,F_{n}(r_{1},r_{2}]).

Therefore, to minimize cap(En(r1,r2)){\rm cap}(E_{n}(r_{1},r_{2})) under the constraint (4.2), we can minimize FnF_{n} under the same constraint.

Using the variables u=(1+r1)/(1r1)u=(1+r_{1})/(1-r_{1}), v=(1+r2)/(1r2)v=(1+r_{2})/(1-r_{2}), constrained by the condition uv=Cuv=C, we express FnF_{n} as follows:

Fn=(C1+(uv))2n+(C1+(vu))2n(C+1+(u+v))2n+(C+1(u+v))2n.F_{n}=\frac{(C-1+(u-v))^{2n}+(C-1+(v-u))^{2n}}{(C+1+(u+v))^{2n}+(C+1-(u+v))^{2n}}.

To minimize Fn(u,v)F_{n}(u,v) under the constraint uv=Cuv=C, we introduce Lagrange’s function

L=Fn(u,v)+λuv,λ.L=F_{n}(u,v)+\lambda uv,\quad\lambda\in\mathbb{R}.

Differentiating this function, we find

(4.5) Luλv=Lvλu=M(u,v,C),\frac{\partial L}{\partial u}-\lambda v=\frac{\partial L}{\partial v}-\lambda u=M(u,v,C),

where M=M(u,v,C)M=M(u,v,C) is defined as follows:

M=2n((C+1+(u+v))2n+(C+1(u+v))2n)2(M1N1M2N2),M=2n\,\left((C+1+(u+v))^{2n}+(C+1-(u+v))^{2n}\right)^{-2}\,(M_{1}N_{1}-M_{2}N_{2}),
M1=(C+1+(u+v))2n+(C+1(u+v))2n,{}M_{1}=(C+1+(u+v))^{2n}+(C+1-(u+v))^{2n},
M2=(C1+(uv))2n+(C1+(vu))2n,{}M_{2}=(C-1+(u-v))^{2n}+(C-1+(v-u))^{2n},
N1=(C1+(uv))2n1(C1+(vu))2n1,N_{1}=(C-1+(u-v))^{2n-1}-(C-1+(v-u))^{2n-1},
N2=(C+1+(u+v))2n1(C+1(u+v))2n1.N_{2}=(C+1+(u+v))^{2n-1}-(C+1-(u+v))^{2n-1}.

It follows from equation (4.5) that if (u,v)(u,v) is a critical point of the minimization problem under consideration, then u=v=Cu=v=\sqrt{C}. In this case

Fn(C,C)=2(C1)2n(C+1)4n+(C1)4n.F_{n}(\sqrt{C},\sqrt{C})=\frac{2(C-1)^{2n}}{(\sqrt{C}+1)^{4n}+(\sqrt{C}-1)^{4n}}.

Since u1u\geq 1, v1v\geq 1, uv=Cuv=C and there is only one critical point (u,v)=(C,C)(u,v)=(\sqrt{C},\sqrt{C}) of the minimization problem under consideration and since F(u,v)=F(v,u)F(u,v)=F(v,u), it follows that F(u,v)F(u,v) achieves its minimal value either at the point (u,v)=(C,C)(u,v)=(\sqrt{C},\sqrt{C}) or at the point (u,v)=(1,C)(u,v)=(1,C). In the latter case, we have

Fn(1,C)=(C1)2n(C+1)2n.F_{n}(1,C)=\frac{(C-1)^{2n}}{(C+1)^{2n}}.

The inequality Fn(C,C)<Fn(1,C)F_{n}(\sqrt{C},\sqrt{C})<F_{n}(1,C) is equivalent to the following inequality:

(C+1)4n+(C1)4n>2(C+1)2n.(\sqrt{C}+1)^{4n}+(\sqrt{C}-1)^{4n}>2(C+1)^{2n}.

Using binomial expansion, this inequality can be written as

k=02n(4n2k)C2nk>k=02n(2nk)C2nk.\sum_{k=0}^{2n}\binom{4n}{2k}C^{2n-k}>\sum_{k=0}^{2n}\binom{2n}{k}C^{2n-k}.

Since (2n2k)>(nk)\binom{2n}{2k}>\binom{n}{k} for all n2n\geq 2 and 1kn11\leq k\leq n-1, the latter inequality holds true.

Thus, Fn(u,v)F_{n}(u,v) takes its minimal value when u=v=Cu=v=\sqrt{C} and therefore the inequality (4.4) is proved. \Box

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 (2n2k)>(nk)\binom{2n}{2k}>\binom{n}{k}, we can argue as follows.

Let n2n\geq 2 and 0knk0\leq k\leq n-k. If k=0k=0, then (2n2k)=(nk)=1\binom{2n}{2k}=\binom{n}{k}=1. Suppose that (2n2k)(nk)\binom{2n}{2k}\geq\binom{n}{k} for 0knk10\leq k\leq n-k-1. Then

(2n2(k+1))=(2n2k)(2n2k)(2n2k1)(2k+1)(2k+2)(nk)2(nk)(2n2k1)2(k+1)(2k+1)\binom{2n}{2(k+1)}=\binom{2n}{2k}\frac{(2n-2k)(2n-2k-1)}{(2k+1)(2k+2)}\geq\binom{n}{k}\frac{2(n-k)(2n-2k-1)}{2(k+1)(2k+1)}
=(nk+1)2n2k12k1>(nk+1).=\binom{n}{k+1}\frac{2n-2k-1}{2k-1}>\binom{n}{k+1}.

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.

Under the assumptions of Theorem 4.1, suppose that r1=sr_{1}=s and r2=r2(s)r_{2}=r_{2}(s) is such that condition (4.2) holds. Then cap(En(s,r2(s))){\rm cap}(E_{n}(s,r_{2}(s))) strictly decreases from 8n𝒦(κ0)𝒦(κ0)8n\frac{\mathcal{K}(\kappa_{0})}{\mathcal{K}^{\prime}(\kappa_{0})} to 16n𝒦(κ1)𝒦(κ1)16n\frac{\mathcal{K}(\kappa_{1})}{\mathcal{K}^{\prime}(\kappa_{1})}, when ss runs from 0 to rr, where

κ0=(e2τ1e2τ+1)2n,κ1=(eτ1eτ+1)4n.\kappa_{0}=\left(\frac{e^{2\uptau}-1}{e^{2\uptau}+1}\right)^{2n},\quad\quad\kappa_{1}=\left(\frac{e^{\uptau}-1}{e^{\uptau}+1}\right)^{4n}.

Since the conformal capacity is conformally invariant, the conformal capacity of an interval EE on (1,1)(-1,1) remains constant when EE moves along the diameter (1,1)(-1,1) so that its hyperbolic length is fixed. For nn intervals situated on nn 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 Eτ(r)E_{\uptau}(r) be a closed subinterval of [0,1)[0,1) with the initial point at 0<r<10<r<1 and hyperbolic length τ>0\uptau>0. For nn\in\mathbb{N}, let Eτn(r)=k=1ne2πi(k1)/nEτ(r)E_{\uptau}^{n}(r)=\cup_{k=1}^{n}e^{2\pi i(k-1)/n}E_{\uptau}(r). Then the conformal capacity cap(Eτn(r)){\rm cap}(E_{\uptau}^{n}(r)) is given by

(4.9) cap(Eτn(r))=4n𝒦(κ)𝒦(κ),{\rm cap}(E_{\uptau}^{n}(r))=4n\frac{\mathcal{K}(\kappa)}{\mathcal{K}^{\prime}(\kappa)},

where

(4.10) κ=ρnrn1rnρn,ρ=eτ(1+r)(1r)eτ(1+r)+(1r).\kappa=\frac{\rho^{n}-r^{n}}{1-r^{n}\rho^{n}},\quad\quad\rho=\frac{e^{\uptau}(1+r)-(1-r)}{e^{\uptau}(1+r)+(1-r)}.

Furthermore, cap(Eτn(r)){\rm cap}(E_{\uptau}^{n}(r)) strictly increases from 4n𝒦(κ0)𝒦(κ0)4n\frac{\mathcal{K}(\kappa_{0})}{\mathcal{K}^{\prime}(\kappa_{0})} with κ0=(eτ1)n/(eτ+1)n\kappa_{0}=(e^{\uptau}-1)^{n}/(e^{\uptau}+1)^{n} to 4n𝒦(κ1)𝒦(κ1)4n\frac{\mathcal{K}(\kappa_{1})}{\mathcal{K}^{\prime}(\kappa_{1})} with κ1=(eτ1)/(eτ+1)\kappa_{1}=(e^{\uptau}-1)/(e^{\uptau}+1), when rr varies from 0 to 11.

Proof. As in the proof of Theorem 4.1, we use the function φ=φ2φ1\varphi=\varphi_{2}\circ\varphi_{1}, where φ1(z)=zn\varphi_{1}(z)=z^{n}, φ2(z)=(zrn)/(1rnz)\varphi_{2}(z)=(z-r^{n})/(1-r^{n}z). Then φ\varphi maps the sector S={z𝔻:|argz|<π/n}S=\{z\in\mathbb{D}:\,|\arg z|<\pi/n\} conformally onto 𝔻\mathbb{D} slit along the interval [1,0][-1,0]. Furthermore, φ\varphi maps Eτ(r)E_{\uptau}(r) onto the interval [0,ρ][0,\rho] with ρ\rho defined as in (4.10). Using symmetry properties of Eτn(r)E_{\uptau}^{n}(r), we find that

cap([rn,ρn])=cap([0,(ρnrn)/(1rnρn)]).{\rm cap}([r^{n},\rho^{n}])={\rm cap}([0,(\rho^{n}-r^{n})/(1-r^{n}\rho^{n})]).

This equation together with (1.10) gives (4.9).

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 cap(Eτn(r)){\rm cap}(E_{\uptau}^{n}(r)) is an analytic function of rr. This follows from equation (4.9). Since cap(Eτn(r)){\rm cap}(E_{\uptau}^{n}(r)) is not constant and analytic, it follows that cap(Eτn(r)){\rm cap}(E_{\uptau}^{n}(r)) is not constant on any subinterval of [0,1)[0,1). Furthermore, it follows from Theorem 3.45 that cap(Eτn(r)){\rm cap}(E_{\uptau}^{n}(r)) is a non-decreasing function. Since it is non-decreasing and not constant on any interval, it is strictly increasing. \Box

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 EkE_{k}, k=1,,nk=1,\ldots,n, be compact sets on the interval I=[0,1)I=[0,1) having positive logarithmic capacities and such that every point of EkE_{k} is regular for the Dirichlet problem in 𝔻Ek\mathbb{D}\setminus E_{k}, k=1,2,,nk=1,2,\dots,n. Then

(4.12) 1nk=1ncap(j=1ne2πi(j1)/nEk)cap(k=1ne2πi(k1)/nEk)<k=1ncap(Ek).\frac{1}{n}\sum_{k=1}^{n}{\rm cap}(\cup_{j=1}^{n}e^{2\pi i(j-1)/n}E_{k})\leq{\rm cap}(\cup_{k=1}^{n}e^{2\pi i(k-1)/n}E_{k})<\sum_{k=1}^{n}{\rm cap}(E_{k}).

Equality occurs in the first inequality if and only if for each kk and jj, EkE_{k} coincides with EjE_{j} 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 uu denote the potential function of the condenser (𝔻,k=1ne2πi(k1)/nEk)(\mathbb{D},\bigcup_{k=1}^{n}e^{2\pi i(k-1)/n}E_{k}). Since every point of EkE_{k} is regular for the Dirichlet problem, it follows that uu is continuous on 𝔻¯\overline{\mathbb{D}}. Let uku^{\prime}_{k} and uk′′u^{\prime\prime}_{k} denote the functions obtained from uu, first by restricting uu, respectively, onto the sector S1={z𝔻:π(2k3)/nargz2π(k1)/n}S_{1}=\{z\in\mathbb{D}:\,\pi(2k-3)/n\leq\arg z\leq 2\pi(k-1)/n\} or onto the sector S2={z𝔻: 2π(k1)/nargzπ(2k1)/n}S_{2}=\{z\in\mathbb{D}:\,2\pi(k-1)/n\leq\arg z\leq\pi(2k-1)/n\} and then extending this restriction by symmetry on the whole unit disk. Then each of the functions uku^{\prime}_{k} and uk′′u^{\prime\prime}_{k} is admissible for the condenser (𝔻,j=1ne2πi(j1)/nEk)(\mathbb{D},\cup_{j=1}^{n}e^{2\pi i(j-1)/n}E_{k}). Furthermore, each of these functions possesses nn-fold rotational symmetry about 0 and is symmetric with respect to the real axis. Therefore, the following inequality holds:

(4.13) 1ncap(j=1ne2πi(j1)/nEk)S1S2|uk|2𝑑m,k=1,,n.\frac{1}{n}{\rm cap}(\cup_{j=1}^{n}e^{2\pi i(j-1)/n}E_{k})\leq\int_{S_{1}\cup S_{2}}|\nabla u_{k}|^{2}\,dm,\quad k=1,\ldots,n.

Summing up all the inequalities in (4.13), we obtain the first inequality in (4.12).

Furthermore, since every point of the sets EjE_{j}, j=1,,nj=1,\ldots,n, is regular for the Dirichlet problem, it follows that uku^{\prime}_{k} or uk′′u^{\prime\prime}_{k} defined above in the proof is a potential function of (𝔻,j=1ne2πi(j1)/nEk)(\mathbb{D},\cup_{j=1}^{n}e^{2\pi i(j-1)/n}E_{k}) if and only of Ej=EkE_{j}=E_{k} for all j=1,,nj=1,\ldots,n. Therefore, if EjEkE_{j}\not=E_{k} for some jkj\not=k, then we have the strict inequality in (4.13) and in the first inequality in (4.12) as well. \Box

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 mn2m\geq n\geq 2 radial intervals that are “densely spread” over 𝔻\mathbb{D} in the sense that the angle between any two neighboring intervals is 2π/n\leq 2\pi/n.

Theorem 4.14.

Let E0=𝔻¯rE_{0}=\overline{\mathbb{D}}_{r}, 0<r<10<r<1, and let E1[r,1)E_{1}\subset[r,1) be a compact set of hyperbolic length τ>0\uptau>0. Let 0=α0=α1<α2<<αm<αm+1=2π0=\alpha_{0}=\alpha_{1}<\alpha_{2}<\ldots<\alpha_{m}<\alpha_{m+1}=2\pi be such that αk+1αk2π/n\alpha_{k+1}-\alpha_{k}\leq 2\pi/n, k=1,,mk=1,\ldots,m, mn2m\geq n\geq 2. Consider compact sets E=k=0meiαkE1E=\cup_{k=0}^{m}e^{i\alpha_{k}}E_{1}, E=k=0nei2π(k1)/nE1E^{*}=\cup_{k=0}^{n}e^{i2\pi(k-1)/n}E_{1}, and En(r,τ)=E0(k=1nei2π(k1)/n[r,r1])E_{n}(r,\uptau)=E_{0}\cup(\cup_{k=1}^{n}e^{i2\pi(k-1)/n}[r,r_{1}]), where r1r_{1} is such that r<r1<1r<r_{1}<1 and 𝔻([r,r1])=τ\ell_{\mathbb{D}}([r,r_{1}])=\uptau. Then

(4.15) cap(E)cap(E)cap(En(r,τ)).{\rm cap}(E)\geq{\rm cap}(E^{*})\geq{\rm cap}(E_{n}(r,\uptau)).

Equality in the first inequality occurs if and only if m=nm=n.

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. \Box

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 0ra<b<10\leq r\leq a<b<1, let E0=𝔻¯rE_{0}=\overline{\mathbb{D}}_{r} and let E1[a,b]E_{1}\subset[a,b] be a compact set of hyperbolic length τ>0\uptau>0, 0<τ<𝔻([a,b])0<\uptau<\ell_{\mathbb{D}}([a,b]). Let 0=α1<α2<<αn<αn+1=2π0=\alpha_{1}<\alpha_{2}<\ldots<\alpha_{n}<\alpha_{n+1}=2\pi. Let

E=E0(k=1neiαkE1),E=E0(k=1nei2π(k1)/nE1),Ea,b=E0(k=1nei2π(k1)/n[a,b]).E=E_{0}\cup(\cup_{k=1}^{n}e^{i\alpha_{k}}E_{1}),\quad E^{*}=E_{0}\cup(\cup_{k=1}^{n}e^{i2\pi(k-1)/n}E_{1}),\quad E^{a,b}=E_{0}\cup(\cup_{k=1}^{n}e^{i2\pi(k-1)/n}[a,b]).

Then

(4.17) cap(E)cap(E)<cap(Ea,b).{\rm cap}(E)\leq{\rm cap}(E^{*})<{\rm cap}(E^{a,b}).

Equality in the first inequality in (4.17) occurs if and only if αk=2π(k1)/n\alpha_{k}=2\pi(k-1)/n, k=1,,nk=1,\ldots,n.

Proof. The first inequality in (4.17) together with the statement on the cases of equality follows from Dubinin’s dissymmetrization results; see Theorem 4.14 in [17]. Then, since the conformal capacity is an increasing function of a set, the second inequality follows. \Box

Remark 4.18.

If E1E_{1} in Theorem 4.16 is an interval [c,d][a,b][c,d]\subset[a,b], then the upper bound in (4.17) can be replaced with cap(E0(k=1nei2π(k1)/n[s,b])){\rm cap}(E_{0}\cup(\cup_{k=1}^{n}e^{i2\pi(k-1)/n}[s,b])) with ss in (a,b)(a,b) such that 𝔻([s,b])=τ\ell_{\mathbb{D}}([s,b])=\uptau. This follows, for instance, from the monotonicity property stated in Theorem 3.45.

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: Lα={z:Im(eiαz)=0}L_{\alpha}=\{z:\,{\,\operatorname{Im}\,}(e^{-i\alpha}z)=0\}, La={z:Imz=a}L^{a}=\{z:\,{\,\operatorname{Im}\,}z=a\}, Cr={z:|z|=r}C_{r}=\{z:\,|z|=r\}, Rα={z=teiα:t0}R_{\alpha}=\{z=te^{i\alpha}:\,t\geq 0\}. First, we mention results obtained with Steiner symmetrization.

Definition 5.1.

Let EE\subset\mathbb{C} be a compact set. The Steiner symmetrization of EE with respect to the imaginary axis is defined to be the compact set

E={z=x+iy:ELy,|x|(1/2)(ELy)},E^{*}=\{z=x+iy:\,E\cap L^{y}\not=\emptyset,|x|\leq(1/2)\ell(E\cap L^{y})\},

where ()\ell(\cdot) stands for the one-dimensional Lebesgue measure.

Furthermore, the Steiner symmetrization of EE with respect to the line LαL_{\alpha} is defined to be the compact set Eα=ei(απ/2)(ei(π/2α)E)E_{\alpha}^{*}=e^{i(\alpha-\pi/2)}\left(e^{i(\pi/2-\alpha)}E\right)^{*}.

For the properties and results obtained with the Steiner symmetrization, the interested reader may consult [33], [17], [5].

We note that if EE is a compact set in 𝔻\mathbb{D}, then Eα𝔻E_{\alpha}^{*}\subset\mathbb{D} for all α\alpha\in\mathbb{R} and therefore cap(Eα){\rm cap}(E_{\alpha}^{*}) 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 A𝔻(E)A_{\mathbb{D}}(E) of EE 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

w=φ0(z)=log1+z1z,φ0(0)=0,w=\varphi_{0}(z)=\log\frac{1+z}{1-z},\quad\varphi_{0}(0)=0,

which maps 𝔻\mathbb{D} conformally onto the horizontal strip Π={w:|Imw|<π/2}\Pi=\{w:\,|{\,\operatorname{Im}\,}w|<\pi/2\}. We note that φ\varphi maps the hyperbolic geodesic (1,1)(-1,1) onto the real axis and it maps the curves equidistant from (1,1)(-1,1), (which are circular arcs in 𝔻\mathbb{D} joining the points 11 and 1-1), onto the horizontal lines {w:Imw=c}\{w:\,{\,\operatorname{Im}\,}w=c\}, 0<|c|<π/20<|c|<\pi/2, in Π\Pi.

Definition 5.2.

Let EE be a compact set in 𝔻\mathbb{D}. The hyperbolic Steiner symmetrization of EE with respect to the hyperbolic geodesic (i,i)(-i,i), centered at z=0z=0, is defined as

Eh=φ01((φ0(E))),E_{h}^{*}=\varphi_{0}^{-1}\left(\left(\varphi_{0}(E)\right)^{*}\right),

where (φ0(E))\left(\varphi_{0}(E)\right)^{*} stands for the Steiner symmetrization as in Definition 5.1.

Furthermore, the hyperbolic Steiner symmetrization of EE with respect to a hyperbolic geodesic γ\gamma, centered at the point aγa\in\gamma, is defined as

Eh(γ,a)=ψ1((ψ(E))h),E_{h}(\gamma,a)=\psi^{-1}\left(\left(\psi(E)\right)_{h}^{*}\right),

where ψ(z)\psi(z) denotes the Möbius automorphism of 𝔻\mathbb{D}, which maps γ\gamma onto the hyperbolic geodesic (i,i)(-i,i) such that ψ(a)=0\psi(a)=0.

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 Eh(γ,a)E_{h}(\gamma,a) be the image of a compact set E𝔻E\subset\mathbb{D} under the hyperbolic Steiner symmetrization with respect to a hyperbolic geodesic γ\gamma, centered at aγa\in\gamma. Let γa\gamma_{\perp}\ni a be the hyperbolic geodesic orthogonal to γ\gamma. Then the following hold true:

  1. (1)

    If EE is a hyperbolic disk, then Eh(γ,a)E_{h}(\gamma,a) is a hyperbolic disk of the same hyperbolic area as EE having its center on γ\gamma.

  2. (2)

    Eh(γ,a)E_{h}(\gamma,a) is a compact set that is symmetric with respect to γ\gamma.

  3. (3)

    𝔻(Eh(γ,a)γ)=𝔻(Eγ)\ell_{\mathbb{D}}(E_{h}(\gamma,a)\cap\gamma_{\perp})=\ell_{\mathbb{D}}(E\cap\gamma_{\perp}).

  4. (4)

    A𝔻(Eh(γ,a))=A𝔻(E)A_{\mathbb{D}}(E_{h}(\gamma,a))=A_{\mathbb{D}}(E).

  5. (5)

    cap(Eh(γ,a))cap(E){\rm cap}(E_{h}(\gamma,a))\leq{\rm cap}(E) with the sign of equality if and only if Eh(γ,a)E_{h}(\gamma,a) coincides with EE up to a set of zero logarithmic capacity and up to a Möbius automorphism of 𝔻\mathbb{D} that preserves γ\gamma_{\perp}.

Proof. We use the notation that we set in the definitions of Steiner and hyperbolic Steiner symmetrizations. We equip the strip Π\Pi with the hyperbolic metric λΠ(w)|dw|\lambda_{\Pi}(w)|dw| induced by the hyperbolic metric in 𝔻\mathbb{D} via the conformal mapping φ0:𝔻Π\varphi_{0}:\mathbb{D}\to\Pi. That is, we have

λΠ(w)|dw|=λ𝔻(z)|dz|,w=φ0(z),z𝔻,wΠ.\lambda_{\Pi}(w)|dw|=\lambda_{\mathbb{D}}(z)|dz|,\;\;\;w=\varphi_{0}(z),\;z\in\mathbb{D},\;w\in\Pi.

So, trivially, φ0\varphi_{0} is a hyperbolic isometry from 𝔻\mathbb{D} to Π\Pi.

(1) Let EE be a hyperbolic disk in 𝔻\mathbb{D}. Then ψ(E)\psi(E) is a hyperbolic disk in 𝔻\mathbb{D} and φ0ψ(E)\varphi_{0}\circ\psi(E) is a hyperbolic disk in Π\Pi. It is easy to observe that hyperbolic disks in Π\Pi 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 (φ0ψ(E))(\varphi_{0}\circ\psi(E))^{*} is obtained by a horizontal rigid motion of φ0ψ(E)\varphi_{0}\circ\psi(E) and it is a hyperbolic disk in Π\Pi, symmetric with respect to the imaginary axis. Since both ψ1\psi^{-1} and φ01\varphi_{0}^{-1} are hyperbolic isometries and preserve symmetries, Eh(γ,a)E_{h}(\gamma,a) is a hyperbolic disk in 𝔻\mathbb{D}, symmetric with respect to γ\gamma. Moreover, a horizontal motion in Π\Pi preserves the hyperbolic area. Therefore, A𝔻(Eh(γ,a))=A𝔻(E)A_{\mathbb{D}}(E_{h}(\gamma,a))=A_{\mathbb{D}}(E).

(2) It is well known that Steiner symmetrization transforms compact sets to compact sets. So, if E𝔻E\subset\mathbb{D} is compact, Eh(γ,a)E_{h}(\gamma,a) is compact, too. Moreover, (φ0(E))(\varphi_{0}(E))^{*} is a set in Π\Pi, symmetric with respect to the imaginary axis. Hence, EhE_{h}^{*} is a compact set in 𝔻\mathbb{D}, symmetric with respect to the geodesic (i,i)(-i,i). It follows that Eh(γ,a)E_{h}(\gamma,a) is symmetric with respect to γ\gamma.

(3) Since ψ\psi is a hyperbolic isometry on 𝔻\mathbb{D}, the set ψ(Eγ)\psi(E\cap\gamma_{\perp}) is a compact subset of the diameter (1,1)(-1,1) having the same hyperbolic length as EγE\cap\gamma_{\perp}. Therefore, φ0ψ(Eγ)\varphi_{0}\circ\psi(E\cap\gamma_{\perp}) lies on the real axis and Π(φ0ψ(Eγ))=𝔻(Eγ)\ell_{\Pi}(\varphi_{0}\circ\psi(E\cap\gamma_{\perp}))=\ell_{\mathbb{D}}(E\cap\gamma_{\perp}). The hyperbolic length (in Π\Pi) on the real axis is proportional to the Euclidean length \ell. Hence Π(φ0ψ(Eγ))=Π((φ0ψ(Eγ)))\ell_{\Pi}(\varphi_{0}\circ\psi(E\cap\gamma_{\perp}))=\ell_{\Pi}((\varphi_{0}\circ\psi(E\cap\gamma_{\perp}))^{*}). Thus the equality in (3) follows at once.

(4) This is true because φ0\varphi_{0}, ψ\psi, and Steiner symmetrization (with respect to the imaginary axis, in Π\Pi) preserve hyperbolic areas.

(5) The conformal maps φ0\varphi_{0} and ψ\psi preserve the capacity of condensers. So, to prove the inequality in (5), it suffices to show that for every compact subset FF of Π\Pi, we have cap(Π,F)cap(Π,F){\rm cap}(\Pi,F)\geq{\rm cap}(\Pi,F^{*}). This is a well-known symmetrization theorem [17, Theorem 4.1]. The equality statement follows from [12, Theorem 1]. \Box

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 ThT_{h}^{*} of the hyperbolic triangle TT with vertices z1=rz_{1}=-r, z2=rz_{2}=r, 0<r<10<r<1, and z3=is+1+s2eiβz_{3}=-is+\sqrt{1+s^{2}}e^{i\beta}, s>0s>0, arctans<β<π/2\arctan s<\beta<\pi/2, is a proper subset of the hyperbolic isosceles triangle T0T_{0} with vertices z1=rz_{1}=-r, z2=rz_{2}=r, and z3=(1+s2s)iz_{3}=(\sqrt{1+s^{2}}-s)i.

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 E𝔻E\subset\mathbb{D} with prescribed hyperbolic diameter d>0d>0 and prescribed hyperbolic area 0<A<4πsinh2(d/4)0<A<4\pi\sinh^{2}(d/4). Describe possible extremal configurations.

The Schwarz symmetrization of EE in the plane with respect to a point aa replaces EE by the disk centered at aa of the same area. The hyperbolic analog of this symmetrization is the following.

Definition 5.6.

Let EE be a compact set in 𝔻\mathbb{D}. Then its hyperbolic Schwarz symmetrization with respect to a𝔻a\in\mathbb{D} is the hyperbolic disk, we call it Ea#E_{a}^{\#}, centered at aa and such that A𝔻(Ea#)=A𝔻(E)A_{\mathbb{D}}(E_{a}^{\#})=A_{\mathbb{D}}(E).

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 Ea#E_{a}^{\#} is the hyperbolic Schwarz symmetrization of a compact set E𝔻E\subset\mathbb{D}, then

cap(Ea#)cap(E){\rm cap}(E_{a}^{\#})\leq{\rm cap}(E)

with the sign of equality if and only if Ea#E_{a}^{\#} coincides with EE up to a set of zero logarithmic capacity and up to a Möbius automorphism of 𝔻\mathbb{D}.

Next, we will discuss the hyperbolic circular symmetrization of EE with respect to the hyperbolic ray [a,eiα)h[a,e^{i\alpha})_{h}, a𝔻a\in\mathbb{D}, α\alpha\in\mathbb{R}. This is defined as the image of the interval [0,1)[0,1) under a Möbius automorphism φ\varphi of 𝔻\mathbb{D} such that φ(0)=a\varphi(0)=a, φ(1)=eiα\varphi(1)=e^{i\alpha}.

Definition 5.8.

Let EE be a compact set in 𝔻\mathbb{D}. Then its hyperbolic circular symmetrization with respect to the hyperbolic ray [0,1)h[0,1)_{h} is a compact set Eh𝔻E_{h}^{\circ}\subset\mathbb{D} such that: (a) 0Eh0\in E_{h}^{\circ} if and only if 0E0\in E, (b) for 0<r<10<r<1, EhCr=E_{h}^{\circ}\cap C_{r}=\emptyset if and only if ECr=E\cap C_{r}=\emptyset, (c) if, for 0<r<10<r<1, ECrE\cap C_{r}\not=\emptyset, then EhCrE_{h}^{\circ}\cap C_{r} is a closed circular arc on CrC_{r} (which may degenerate to a point or may be the whole circle CrC_{r}) centered at z=rz=r such that 𝔻(EhCr)=𝔻(ECr)\ell_{\mathbb{D}}(E_{h}^{\circ}\cap C_{r})=\ell_{\mathbb{D}}(E\cap C_{r}).

Furthermore, the hyperbolic circular symmetrization of EE with respect to a hyperbolic geodesic ray [a,eiα)h[a,e^{i\alpha})_{h} is defined as

(5.9) Eh(a,α)=φ1((φ(E))h),E_{h}^{\circ}(a,\alpha)=\varphi^{-1}\left(\left(\varphi(E)\right)_{h}^{\circ}\right),

where φ(z)\varphi(z) denotes the Möbius automorphism of 𝔻\mathbb{D}, which maps [a,eiα)h[a,e^{i\alpha})_{h} onto the interval [0,1)[0,1).

Remark 5.10.

We immediately admit here that the hyperbolic circular symmetrization with respect to [0,1)h[0,1)_{h} is exactly the classical circular symmetrization with respect to the positive real axis. The reason for this is the fact that the hyperbolic density λ𝔻(z)\lambda_{\mathbb{D}}(z) is constant on circles centered at 0. 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 cap(Eh(a,α)){\rm cap}(E_{h}^{\circ}(a,\alpha)) do not depend on α\alpha. For the Euclidean area of Eh(a,α)E_{h}^{\circ}(a,\alpha) we have the following result.

Lemma 5.11.

Let E𝔻E\subset\mathbb{D} be a compact set of positive area, and let A(α)A(\alpha) denote the Euclidean area of Eh(r,α)E_{h}^{\circ}(r,\alpha), 0<r<10<r<1, 0απ0\leq\alpha\leq\pi, considered as a function of α\alpha. If Eho(r,α)E_{h}^{o}(r,\alpha) does not coincide with a hyperbolic disk centered at rr up to measure 0, then A(α)A(\alpha) is strictly increasing in α\alpha on the interval 0απ0\leq\alpha\leq\pi; otherwise A(α)A(\alpha) is constant for 0απ0\leq\alpha\leq\pi.

Proof. Let E0E_{0} denote the circular symmetrization of φ(E)\varphi(E) with respect to [0,1)h[0,1)_{h}, with φ\varphi defined as in Definition 5.8. Then Eho(r,α)=ψ(E0)E_{h}^{o}(r,\alpha)=\psi(E_{0}), where

ψ(z)=(eiβz+r)/(1+eiβrz),\psi(z)=(e^{i\beta}z+r)/(1+e^{i\beta}rz),

with β=β(α)\beta=\beta(\alpha) chosen such that

eiα=ψ(1)=(eiβ+r)/(1+eiβr).e^{i\alpha}=\psi(1)=(e^{i\beta}+r)/(1+e^{i\beta}r).

The function β(α)\beta(\alpha) strictly increases from 0 to π\pi, when α\alpha increases from 0 to π\pi. Therefore, to prove monotonicity of a certain characteristic FF of Eho(r,α)E_{h}^{o}(r,\alpha) as a function of α\alpha, we can consider α=α(β)\alpha=\alpha(\beta) as a function of β\beta and treat FF as a function of β\beta. Thus, we will work with the function A1(β)=A(α(β))A_{1}(\beta)=A(\alpha(\beta)).

The Euclidean area A1(β)A_{1}(\beta) can be found as follows:

(5.12) A1(β)=E0|ψ(z)|2𝑑m=(1r2)201(θ(r)θ(r)dθ|1+rρei(θ+β)|4)ρ𝑑ρ,A_{1}(\beta)=\int_{E_{0}}|\psi^{\prime}(z)|^{2}\,dm=(1-r^{2})^{2}\int_{0}^{1}\left(\int_{-\theta(r)}^{\theta(r)}\frac{d\theta}{|1+r\rho e^{i(\theta+\beta)}|^{4}}\right)\,\rho\,d\rho,

where 0θ(r)π0\leq\theta(r)\leq\pi is defined by the condition γ(r)=E0Cr={reiθ:|θ|θ(r)}\gamma(r)=E_{0}\cap C_{r}=\{re^{i\theta}:\,|\theta|\leq\theta(r)\}.

If 0<θ(r)<π0<\theta(r)<\pi, then γ(r)\gamma(r) is a non-degenerate proper arc on CrC_{r} that is centered at z=rz=r. Using this symmetry and the monotonicity property of the function

g(t)=1+2rρcost+r2ρ2,0tπ,g(t)=1+2r\rho\cos t+r^{2}\rho^{2},\quad 0\leq t\leq\pi,

we conclude that the inner integral in (5.12), that is the integral

I(β)=θ(r)θ(r)dθ|1+rρei(θ+β)|4=θ(r)+βθ(r)+βdt(1+2rρcost+r2ρ2)4,I(\beta)=\int_{-\theta(r)}^{\theta(r)}\frac{d\theta}{|1+r\rho e^{i(\theta+\beta)}|^{4}}=\int_{-\theta(r)+\beta}^{\theta(r)+\beta}\frac{dt}{(1+2r\rho\cos t+r^{2}\rho^{2})^{4}},

is a strictly increasing function of β\beta, 0βπ0\leq\beta\leq\pi. Therefore, if E0E_{0} differs by positive area from the disk centered at 0, which has the same area as E0E_{0}, then the double integral in the right-hand side of (5.12) strictly increases on the interval 0βπ0\leq\beta\leq\pi. This proves that the area A1(β)=A(α)A_{1}(\beta)=A(\alpha) strictly increases on the interval 0απ0\leq\alpha\leq\pi. Of course, if Eho(r,α)E_{h}^{o}(r,\alpha) coincides with a hyperbolic disk centered at rr up to measure 0, then A(α)A(\alpha) is constant for 0απ0\leq\alpha\leq\pi. \Box

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 0<r<10<r<1 and 0απ0\leq\alpha\leq\pi, let Cr(α)={reiθ:|θ|α}C_{r}(\alpha)=\{re^{i\theta}:\,|\theta|\leq\alpha\} and let E{z:r|z|<1}E\subset\{z:\,r\leq|z|<1\} be a compact set such that the length of its radial projection Epr(r)E_{pr}(r) on the circle CrC_{r} is 2αr\geq 2\alpha r. Then

(5.14) cap(E)cap(Cr(α)){\rm cap}(E)\geq{\rm cap}(C_{r}(\alpha))

with the sign of equality if and only if EE coincides with Cr(α)C_{r}(\alpha) up to a set of zero logarithmic capacity and up to a rotation about 0.

Proof. Consider a function φ(z)=rz/|z|\varphi(z)=rz/|z|. It can be easily shown that φ\varphi is a hyperbolic contraction from EE onto Epr(r)E_{pr}(r). Hence, by Proposition 2.19,

(5.15) cap(E)cap(Epr(r)).{\rm cap}(E)\geq{\rm cap}(E_{pr}(r)).

As well known, the circular symmetrization decreases the conformal capacity, precisely, the following holds:

(5.16) cap(Epr(r))cap(Cr(α)){\rm cap}(E_{pr}(r))\geq{\rm cap}(C_{r}(\alpha))

with the sign of equality if and only if Epr(r)E_{pr}(r) coincides with Cr(α)C_{r}(\alpha) up to a set of zero logarithmic capacity and a rotation about 0. Combining (5.15) and (5.16), we obtain (5.14) together with the statement about cases of equality in it. \Box

Refer to caption
Figure 5. Admissible and extremal sets of Theorems 5.13 and 5.17.
Theorem 5.17.

For 0<r<10<r<1, 0απ0\leq\alpha\leq\pi, and ττ(r)=log((1+r)/(1r))\uptau\geq\uptau(r)=\log((1+r)/(1-r)), let E𝔻E\subset\mathbb{D} be a compact set such that 𝔻(ECr)=2αr/(1r2)\ell_{\mathbb{D}}(E\cap C_{r})=2\alpha r/(1-r^{2}) and 𝔻(Epro)=τ\ell_{\mathbb{D}}(E_{pr}^{o})=\uptau, where EproE_{pr}^{o} stands for the circular projection of EE onto the radius [0,1)[0,1). Then

(5.18) cap(E)cap(Cr(α)[0,r(τ)]),where r(τ)=(eτ1)/(eτ+1),{\rm cap}(E)\geq{\rm cap}(C_{r}(\alpha)\cup[0,r(\uptau)]),\quad\quad{\mbox{where $r(\uptau)=(e^{\uptau}-1)/(e^{\uptau}+1)$,}}

with the sign of equality if and only if EE coincides with Cr(α)[0,r(τ)]C_{r}(\alpha)\cup[0,r(\uptau)] up to a set of zero logarithmic capacity and up to a rotation about 0.

Proof. Performing the circular symmetrization as in the proof of Theorem 5.13, we obtain the following inequality:

(5.19) cap(E)cap(Cr(α)Epro){\rm cap}(E)\geq{\rm cap}(C_{r}(\alpha)\cup E_{pr}^{o})

with the sign of equality if and only if EE coincides with Cr(α)EproC_{r}(\alpha)\cup E_{pr}^{o} up to a set of zero logarithmic capacity and up to a rotation about 0.

If EproE_{pr}^{o} does not coincide with the interval [0,r(τ)][0,r(\uptau)] up to a set of zero logarithmic capacity, then we can perform polarization transformations as in the proof of Lemma 3.4 to transform EproE_{pr}^{o} into the interval [0,r(τ)][0,r(\uptau)]. Furthermore, our assumption that 𝔻(Epro)=τlog((1+r)/(1r))\ell_{\mathbb{D}}(E_{pr}^{o})=\uptau\geq\log((1+r)/(1-r)) guarantees that these polarizations do not change Cα(r)C_{\alpha}(r). Hence, we obtain the inequality

(5.20) cap(Cr(α)Epro)cap(Cr(α)[0,r(τ)]){\rm cap}(C_{r}(\alpha)\cup E_{pr}^{o})\geq{\rm cap}(C_{r}(\alpha)\cup[0,r(\uptau)])

with the sign of equality if and only if EproE_{pr}^{o} coincides with [0,r(τ)][0,r(\uptau)] 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. \Box

The restriction ττ(r)\uptau\geq\uptau(r) 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 minEcap(E)\min_{E}{\rm cap}(E) over all compact sets E𝔻E\subset\mathbb{D} satisfying the assumptions of Theorem 5.17 with 0<τ<τ(r)0<\uptau<\uptau(r).

Problem 5.22.

Let E{z:r1|z|r2}E\subset\{z:\,r_{1}\leq|z|\leq r_{2}\}, 0<r1<r2<10<r_{1}<r_{2}<1, be a compact set such that its radial projection on the unit circle 𝕋\mathbb{T} coincides with 𝕋\mathbb{T} and its circular projection onto the radius [0,1)[0,1) coincides with the interval [r1,r2][r_{1},r_{2}], as it is shown in Figure 6, and let E=𝔻¯r1[r1,r2])E^{*}=\overline{\mathbb{D}}_{r_{1}}\cup[r_{1},r_{2}]). Is it true that cap(E)cap(E){\rm cap}(E)\geq{\rm cap}(E^{*})? If this is not true then identify the shape of EE minimizing the conformal capacity under the assumptions of this problem.

Remark 5.23.

Let γ=γ(r1,r2)r2\gamma=\gamma(r_{1},r_{2})\ni r_{2} denote the hyperbolic geodesic tangent to the circle Cr1C_{r_{1}}. We assume that γ\gamma touches Cr1C_{r_{1}} at the point z1=r1eiθ0z_{1}=r_{1}e^{i\theta_{0}}, 0<θ0<π/20<\theta_{0}<\pi/2. Let α\alpha denote the arc of γ\gamma with the endpoints z1z_{1}, r2r_{2} and let α¯={z:z¯α}\overline{\alpha}=\{z:\,\bar{z}\in\alpha\}. It can be shown with the help of polarization that if EE^{*} minimizes the conformal capacity in Problem 5.22, then EE^{*} belongs to the compact set bounded by the arcs α\alpha, α¯\overline{\alpha} and the circular arc {z=r1eiθ:θ0θ2πθ0}\{z=r_{1}e^{i\theta}:\,\theta_{0}\leq\theta\leq 2\pi-\theta_{0}\}.

Refer to caption
Figure 6. An admissible set and conjectured extremal set of Problem 5.22.

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 0<r<10<r<1, α\alpha\in\mathbb{R}, and a compact set E𝔻E\subset\mathbb{D}, define E(r,α)=E[reiα,eiα)E(r,\alpha)=E\cap[re^{i\alpha},e^{i\alpha}).

Definition 5.24.

Let EE be a compact set in 𝔻\mathbb{D} such that 𝔻¯rE\overline{\mathbb{D}}_{r}\subset E, 0<r<10<r<1. Then the radial symmetrization of EE with respect to 0 is the compact set Erad𝔻E^{\rm rad}\subset\mathbb{D} with the property: for every α\alpha\in\mathbb{R}, Erad(r,α)E^{\rm rad}(r,\alpha) is a radial interval such that

(5.25) Erad(r,α)|dz||z|=E(r,α)|dz||z|.\int_{E^{\rm rad}(r,\alpha)}\frac{|dz|}{|z|}=\int_{E(r,\alpha)}\frac{|dz|}{|z|}.

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 [0,1][0,1] 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 𝔻\mathbb{D}, we turned to the following version.

Definition 5.26.

The hyperbolic radial symmetrization EhradE_{h}^{\rm rad} of a compact set E𝔻E\subset\mathbb{D} with respect to 0 is defined to be a compact set starlike with respect to 0 and such that

𝔻(Ehrad[0,eiα))=𝔻(E[0,eiα))for all α.\ell_{\mathbb{D}}(E_{h}^{\rm rad}\cap[0,e^{i\alpha}))=\ell_{\mathbb{D}}(E\cap[0,e^{i\alpha}))\quad{\mbox{for all $\alpha\in\mathbb{R}$.}}

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 EE be a compact set in 𝔻\mathbb{D}. Prove (or disprove) that

cap(Ehrad)cap(E).{\rm cap}(E_{h}^{\rm rad})\leq{\rm cap}(E).

We conclude this section with the following result, which describes how the hyperbolic area and hyperbolic diameter of a compact set EE 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 EE.

Proof. First we deal with the hyperbolic area Ah(E)A_{h}(E) of a compact set EE in 𝔻\mathbb{D}. Since every compact set in 𝔻\mathbb{D} can be approximated by a finite union of polar rectangles, we may assume that EE is a finite union of sets of the form

R={reit:r1rr2,t1tt2},R=\{re^{it}:r_{1}\leq r\leq r_{2},\;t_{1}\leq t\leq t_{2}\},

where 0r1<r2<10\leq r_{1}<r_{2}<1 and 0t1<t22π0\leq t_{1}<t_{2}\leq 2\pi. For such an EE, the hyperbolic radial symmetrization EhradE_{h}^{\rm rad} is again a union of sets of the same form, and moreover, each of the polar rectangles of EhradE_{h}^{\rm rad} is obtained by moving a rectangle of EE radially towards the origin, keeping its hyperbolic height τ\uptau fixed. More precisely, if RR (as defined above) is a rectangle of EE, then the corresponding rectangle of EhradE_{h}^{\rm rad} has the form

R^={reit:r^1rr^2,t1tt2},\widehat{R}=\{re^{it}:\hat{r}_{1}\leq r\leq\hat{r}_{2},\;t_{1}\leq t\leq t_{2}\},

with r^1r1\hat{r}_{1}\leq r_{1} and

(5.29) 𝔻([r^1,r^2])=𝔻([r1,r2])=τ.\ell_{\mathbb{D}}([\hat{r}_{1},\hat{r}_{2}])=\ell_{\mathbb{D}}([r_{1},r_{2}])=\uptau.

We express r2r_{2} as a function of r1r_{1} using (5.29) and find that

(5.30) r2=r2(r1)=(r1+r(τ))/(1+r(τ)r1),where r(τ) is defined in (1.3).r_{2}=r_{2}(r_{1})=(r_{1}+r(\uptau))/(1+r(\uptau)r_{1}),\quad{\mbox{where $r(\uptau)$ is defined in (\ref{Equation 1.1}).}}

Another elementary calculation gives

Ah(R)=r22r122(1r12)(1r22)(t2t1).A_{h}(R)=\frac{r_{2}^{2}-r_{1}^{2}}{2(1-r_{1}^{2})(1-r_{2}^{2})}(t_{2}-t_{1}).

We consider the function

g(r1)=r22r12(1r12)(1r22),r2=r2(r1),  0<r1<1.g(r_{1})=\frac{r_{2}^{2}-r_{1}^{2}}{(1-r_{1}^{2})(1-r_{2}^{2})},\;\;r_{2}=r_{2}(r_{1}),\;\;0<r_{1}<1.

By straightforward differentiation of g(r1)g(r_{1}) and taking into account (5.30), we find that g(r1)g(r_{1}) is strictly increasing. This implies that Ah(R)Ah(R^)A_{h}(R)\geq A_{h}(\widehat{R}). Summing up over the hyperbolic areas of all rectangles, we conclude that Ah(E)Ah(Ehr)A_{h}(E)\geq A_{h}(E_{h}^{r}).

Next we turn to hyperbolic diameter, which we denote by Diamh(E){\rm Diam}_{h}(E). We will use the following basic fact of the hyperbolic geometry.

(a) Given a hyperbolic geodesic γ\gamma and a point aγa\not\in\gamma, there is a unique hyperbolic geodesic γa\gamma_{\perp}\ni a orthogonal to γ\gamma. Let γ\gamma_{\perp} intersect γ\gamma at z=bz=b. Then the hyperbolic distance d𝔻(a,z)d_{\mathbb{D}}(a,z) strictly increases as zz moves along γ\gamma from bb to 𝕋\mathbb{T}.

To prove this result, we may assume that γ=(1,1)\gamma=(-1,1) and a=isa=is, 0s<10\leq s<1. Then γ=(i,i)\gamma_{\perp}=(-i,i) and the monotonicity of d𝔻(is,r)d_{\mathbb{D}}(is,r) for 0r<10\leq r<1 follows after simple calculations.

Furthermore, if γ=(1,1)\gamma=(-1,1) and Rea>0{\,\operatorname{Re}\,}a>0, Ima>0{\,\operatorname{Im}\,}a>0, then γ\gamma_{\perp} intersects γ\gamma at the point bb such that 0<b<Rea0<b<{\,\operatorname{Re}\,}a.

We will also use the following result that can be checked by standard Calculus technique.

(b) Let z1=reiθz_{1}=re^{i\theta}, 0<r<10<r<1, and z2=eiαz1z_{2}=e^{i\alpha}z_{1}, 0απ0\leq\alpha\leq\pi. Then p𝔻(z1,z2)=2rsin(α/2)12r2cosα+r4p_{\mathbb{D}}(z_{1},z_{2})=\frac{2r\sin(\alpha/2)}{\sqrt{1-2r^{2}\cos\alpha+r^{4}}}. Furthermore, the pseudo-hyperbolic distance p𝔻(z1,z2)p_{\mathbb{D}}(z_{1},z_{2}), and therefore the hyperbolic distance d𝔻(z1,z2)d_{\mathbb{D}}(z_{1},z_{2}), is a strictly increasing function of rr, 0r<10\leq r<1 and a strictly increasing function of α\alpha, 0απ0\leq\alpha\leq\pi.

Now let z1,z2z_{1},z_{2} be two points on EhradE_{h}^{\rm rad} such that d𝔻(z1,z2)=Diamh(Ehrad)d_{\mathbb{D}}(z_{1},z_{2})={\rm Diam}_{h}(E_{h}^{\rm rad}). Note that [0,z1]Ehrad[0,z_{1}]\subset E_{h}^{\rm rad} and [0,z2]Ehrad[0,z_{2}]\subset E_{h}^{\rm rad}. We may assume that z1=r1(0,1)z_{1}=r_{1}\in(0,1), and z2=0z_{2}=0 or z2=r2eitz_{2}=r_{2}e^{it}, 0<r2r10<r_{2}\leq r_{1}, 0tπ0\leq t\leq\pi. If z2=r2z_{2}=-r_{2}, 0r2r20\leq r_{2}\leq r_{2}, then

Diamh(Ehrad)=𝔻([r2,r1])=𝔻(E(1,1))Diamh(E){\rm Diam}_{h}(E_{h}^{\rm rad})=\ell_{\mathbb{D}}([-r_{2},r_{1}])=\ell_{\mathbb{D}}(E\cap(-1,1))\leq{\rm Diam}_{h}(E)

and the required result is proved.

Next, we assume that z2=r2eitz_{2}=r_{2}e^{it} with 0<r2r10<r_{2}\leq r_{1}, 0<t<π0<t<\pi. The set EE contains points ζ1=s1\zeta_{1}=s_{1} and ζ2=s2eit\zeta_{2}=s_{2}e^{it} with s1=max{|z|:zE[0,1)}s_{1}=\max\{|z|:\,z\in E\cap[0,1)\}, s2=max{|z|:zE[0,eit)}s_{2}=\max\{|z|:\,z\in E\cap[0,e^{it})\} such that s1r1s_{1}\geq r_{1}, s2r2s_{2}\geq r_{2}. If π/2t<π\pi/2\leq t<\pi, then it follows from the monotonicity property stated in (a) that d𝔻(ζ1,ζ2)d𝔻(z1,z2)d_{\mathbb{D}}(\zeta_{1},\zeta_{2})\geq d_{\mathbb{D}}(z_{1},z_{2}) and therefore Diamh(E)Diamh(Ehrad){\rm Diam}_{h}(E)\geq{\rm Diam}_{h}(E_{h}^{\rm rad}) in this case.

Now, we consider the case when 0<r2r10<r_{2}\leq r_{1}, 0<t<π/20<t<\pi/2. In this case, the hyperbolic geodesic γz1\gamma\ni z_{1} orthogonal to (eit,eit)(-e^{it},e^{it}) crosses (eit,eit)(-e^{it},e^{it}) at the point z=reitz^{*}=r^{*}e^{it} with 0<r<r10<r^{*}<r_{1}. If 0<r2r0<r_{2}\leq r^{*}, then

Diamh(Ehrad)=d𝔻(r1,r2eit)<d𝔻(r1,0)=𝔻(Ehrad[0,1))=𝔻(E[0,1))Diamh(E),{\rm Diam}_{h}(E_{h}^{\rm rad})=d_{\mathbb{D}}(r_{1},r_{2}e^{it})<d_{\mathbb{D}}(r_{1},0)=\ell_{\mathbb{D}}(E_{h}^{\rm rad}\cap[0,1))=\ell_{\mathbb{D}}(E\cap[0,1))\leq{\rm Diam}_{h}(E),

where the first inequality follows from the monotonicity property stated in (a).

Thus, we are left with the case r<r2r1r^{*}<r_{2}\leq r_{1}, 0<t<π/20<t<\pi/2. We work with the points ζ1=s1\zeta_{1}=s_{1} and ζ2=s2eit\zeta_{2}=s_{2}e^{it} defined above. If r<s2r1s1r^{*}<s_{2}\leq r_{1}\leq s_{1}, then using twice the monotonicity property stated in (a) we obtain the required result:

Diamh(Ehrad)=d𝔻(z1,z2)<d𝔻(z1,ζ2)d𝔻(ζ1,ζ2)Diamh(E).{\rm Diam}_{h}(E_{h}^{\rm rad})=d_{\mathbb{D}}(z_{1},z_{2})<d_{\mathbb{D}}(z_{1},\zeta_{2})\leq d_{\mathbb{D}}(\zeta_{1},\zeta_{2})\leq{\rm Diam}_{h}(E).

If r1<sksjr_{1}<s_{k}\leq s_{j}, kjk\not=j, then

Diamh(Ehrad)=d𝔻(z1,z2)d𝔻(z1,r1eit)d𝔻(sk,skeit)d𝔻(ζk,ζj)Diamh(E),{\rm Diam}_{h}(E_{h}^{\rm rad})=d_{\mathbb{D}}(z_{1},z_{2})\leq d_{\mathbb{D}}(z_{1},r_{1}e^{it})\leq d_{\mathbb{D}}(s_{k},s_{k}e^{it})\leq d_{\mathbb{D}}(\zeta_{k},\zeta_{j})\leq{\rm Diam}_{h}(E),

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 rr stated in (b). Now, the inequality Diamh(Ehrad)Diamh(E){\rm Diam}_{h}(E_{h}^{\rm rad})\leq{\rm Diam}_{h}(E) is proved in all cases. \Box

6. Hedgehog problems in 3{\mathbb{R}}^{3}

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 𝔹={x¯:|x¯|<1}\mathbb{B}=\{\overline{x}:\,|\overline{x}|<1\} in 3\mathbb{R}^{3}. Here, x¯=(x1,x2,x3)\overline{x}=(x_{1},x_{2},x_{3}), |x¯|=x12+x22+x32|\overline{x}|=\sqrt{x_{1}^{2}+x_{2}^{2}+x_{3}^{2}}. The conformal capacity of a compact set E𝔹E\subset\mathbb{B} is defined as

cap(E)=inf𝔹|u|3𝑑V,{\rm{cap}}(E)=\inf\int_{\mathbb{B}}|\nabla u|^{3}\,dV,

where dVdV stands for the three-dimensional Lebesgue measure and the infimum is taken over all Lipschitz functions uu such that u=0u=0 on 𝔻\partial\mathbb{D} and u=1u=1 on EE.

The conformal capacity of EE is just the conformal capacity of the condenser (𝔹,E)(\mathbb{B},E) and therefore it is invariant under Möbius automorphisms of 𝔹\mathbb{B}, 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 3\mathbb{R}^{3}. 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 n3n\geq 3. This issue was discussed, for instance, in Section 5.6.2 in [5].

Below we list possible extensions of our results to 3\mathbb{R}^{3} and to spaces of higher dimension.

  1. (1)

    The non-strict monotonicity property in Lemma 3.1 for a finite number of intervals on the diameter remains true in any dimension n3n\geq 3.

  2. (2)

    The inequality (3.5) of Lemma 3.4 holds in any dimension n3n\geq 3.

  3. (3)

    The inequality (3.28) of Theorem 3.27 holds true for a set EE lying on mm, 1m61\leq m\leq 6, distinct radial intervals IkI_{k} in 𝔹\mathbb{B} under the assumption that the angles between these intervals are greater than or equal to π/2\pi/2.

  4. (4)

    Theorems 3.31 and 3.43 also can be extended to the case of appropriate compact sets in 𝔹\mathbb{B}.

  5. (5)

    The result of Theorem 4.11 can be extended to the three-dimensional space. For this, under the assumptions of this theorem, we assume additionally that the unit disk 𝔻\mathbb{D} is embedded into 𝔹\mathbb{B} as follows: 𝔻=𝔹{(x1,x2,x3)3:x3=0}\mathbb{D}=\mathbb{B}\cap\{(x_{1},x_{2},x_{3})\in\mathbb{R}^{3}:\,x_{3}=0\}. Then, inequality (4.12) remain true for conformal capacities in 3\mathbb{R}^{3}.

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 n3n\geq 3 remain open. Next we list a few of these open problems.

  1. (1)

    Our proofs of Theorems 3.36 and 4.1 require computations related to conformal mapping of doubly connected domains, which is not available in higher dimensions.

  2. (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\geq 3.

  3. (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. (4)

    To prove Theorem 4.14 in the three-dimensional setting, we would need inequality (4.15) under the assumption that the central body E0E_{0} is a ball 𝔹¯r={x3:|x|r}\overline{\mathbb{B}}_{r}=\{x\in\mathbb{R}^{3}:\,|x|\leq r\}, 0r<10\leq r<1, and E1[r,1)E_{1}\subset[r,1) 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 3\mathbb{R}^{3}. Thus, the counterpart of Theorem 4.14 remains unproved in dimensions n3n\geq 3.

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 pp-capacity, Michigan Math. J. 16 (1969), 43–51.