Download as pdf or txt
Download as pdf or txt
You are on page 1of 352

Vladimir N.

Dubinin

Condenser Capacities
and Symmetrization
in Geometric
Function Theory
Vladimir N. Dubinin

Condenser Capacities and


Symmetrization in
Geometric Function Theory

Translated from the Russian by Nikolai G. Kruzhilin


Vladimir N. Dubinin
Institute of Applied Mathematics
Far Eastern Federal University
Vladivostok, Russia

ISBN 978-3-0348-0842-2 ISBN 978-3-0348-0843-9 (eBook)


DOI 10.1007/978-3-0348-0843-9
Springer Basel Heidelberg New York Dordrecht London
Library of Congress Control Number: 2014947338

Mathematics Subject Classification (2010): 30A10, 30C25, 30C45, 30C55, 30C70, 30C75, 30C85, 31A15,
31A99, 31C15

© Springer Basel 2014


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the
material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now
known or hereafter developed. Exempted from this legal reservation are brief excerpts in connection with
reviews or scholarly analysis or material supplied specifically for the purpose of being entered and executed
on a computer system, for exclusive use by the purchaser of the work. Duplication of this publication or parts
thereof is permitted only under the provisions of the Copyright Law of the Publisher’s location, in its current
version, and permission for use must always be obtained from Springer. Permissions for use may be obtained
through RightsLink at the Copyright Clearance Center. Violations are liable to prosecution under the
respective Copyright Law.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
While the advice and information in this book are believed to be true and accurate at the date of publication,
neither the authors nor the editors nor the publisher can accept any legal responsibility for any errors or
omissions that may be made. The publisher makes no warranty, express or implied, with respect to the
material contained herein.

Printed on acid-free paper

Springer Basel is part of Springer Science+Business Media (www.birkhauser-science.com)


To my parents
Contents

Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix
1 Conformal Capacity
1.1 Lipschitz functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Condensers with two plates on the Riemann sphere . . . . . . . . . . 6
1.3 Generalized condensers . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.4 Geometric interpretation of capacity . . . . . . . . . . . . . . . . . . 22
2 The Asymptotic Behaviour of the Capacity as some Plates
of the Condenser Degenerate into Points
2.1 The Green, Robin, and Neumann functions . . . . . . . . . . . . . . 25
2.2 The case when one plate is nondegenerate . . . . . . . . . . . . . . . 33
2.3 Condensers all of whose plates are degenerating . . . . . . . . . . . . 41
2.4 Reduced moduli . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3 Special Transformations
3.1 Contractions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.2 Polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.3 The Gonchar transformation . . . . . . . . . . . . . . . . . . . . . . 71
3.4 Linear and radial transformations . . . . . . . . . . . . . . . . . . . . 75
3.5 Averaging transformations . . . . . . . . . . . . . . . . . . . . . . . . 79
4 Symmetrization
4.1 Symmetrization along straight lines or circles . . . . . . . . . . . . . 89
4.2 Composites of symmetrizations and conformal mappings . . . . . . . 100
4.3 Separating and averaging symmetrizations . . . . . . . . . . . . . . . 103
4.4 Dissymmetrization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5 Metric Properties of Sets and Condensers
5.1 Finite sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
5.2 Projections of sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
5.3 Subsets of a circle and a line interval . . . . . . . . . . . . . . . . . . 137
5.4 Polygons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
5.5 Ring domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

vii
viii Contents

6 Extremal Decomposition Problems


6.1 Fixed poles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
6.2 Free poles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
6.3 Möbius invariants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
7 Univalent Functions
7.1 Classical inequalities, with some additions and comments . . . . . . 202
7.2 Covering theorems for radial segments . . . . . . . . . . . . . . . . . 217
7.3 Distortion theorems related to n-symmetry . . . . . . . . . . . . . . 225
7.4 Variational principles of conformal mappings . . . . . . . . . . . . . 238
7.5 Behaviour of level curves . . . . . . . . . . . . . . . . . . . . . . . . . 250
7.6 Inequalities involving the Schwarzian derivative . . . . . . . . . . . . 259
8 Multivalent Functions
8.1 Majorization principles . . . . . . . . . . . . . . . . . . . . . . . . . . 277
8.2 Covering theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
8.3 Distortion theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
8.4 p-valent functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300
Appendix
A.1 Dirichlet’s principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
A.2 A uniqueness theorem for contractions . . . . . . . . . . . . . . . . . 306
A.3 On separating transformations of sets and condensers . . . . . . . . . 310
A.4 Invariance of the reduced modulus under geometric
transformations of domains . . . . . . . . . . . . . . . . . . . . . . . 314
A.5 Quadratic differentials . . . . . . . . . . . . . . . . . . . . . . . . . . 324
A.6 Unsolved problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
List of Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
Preface

The theory of functions of a complex variable is rich with manifestations of sym-


metry. This fact probably owes itself to the symmetry of the centre of the theory,
namely conformality. If the conditions of a problem involve some symmetry then
a similar symmetry is to be expected from its solutions. This is yet another confir-
mation of the well-known Curie principle: symmetric premises lead to symmetric
conclusions. Let us also recall that Weyl emphasized the importance of investigat-
ing the symmetry (automorphism) groups of an object under study. On the other
hand there exists a deep connection between the phenomenon of symmetry, con-
servation laws, and extremal problems. This connection is confirmed by numerous
examples in natural sciences and thus must also be reflected in function theory,
whose subject matter is close to mathematical physics. This explains the role of
the method of symmetrization, the method implying the investigation and sys-
tematic use of such transformations as symmetrization and dissymmetrization to
solve extremal problems. Here we mean by symmetrization a procedure reducing
the asymmetry of the objects and resulting in the appearance of some elements
of symmetry in them. Dissymmetrization is the (reverse) process of the successive
removal of elements of symmetry.
Let us briefly outline the common features inherent in the method of sym-
metrization. Let U be an abstract set and F(u) be a real-valued function on U.
Assume that we investigate the following extremal problem:

F(u) → inf, u ∈ U. (0.1)

The method of symmetrization hinges on the idea that any extremal configuration
is determined by some kind of symmetry, so that we shall seek a solution among
the objects possessing some elements of that symmetry. A priori the symmetry of
the extremal object is unknown in general. The task of symmetrization is to find a
proper subset U∗ ∈ U comprising the elements u possessing some sort of symmetry
and to define a map Sym : U → U∗ such that

F(u)  F(Sym u), u ∈ U. (0.2)

If such a map exists problem (0.1) reduces to

F(u) → inf, u ∈ U∗ .

ix
x Preface

The map Sym is called symmetrization while inequality (0.2) is called a sym-
metrization principle. The method of symmetrization consists of the symmetriza-
tion principles corresponding to various maps Sym. Certain generalizations are
also possible, such as passing from one extremal problem to another in the process
of symmetrization or an analysis of the reverse procedure, dissymmetrization. It
is not unusual to apply several types of symmetrization in solving one particular
problem. As a rule, Sym is a geometric transformation while F(u) is a compli-
cated analytic characteristic. Estimating less accessible quantities (capacities, in-
ner radii) in terms of more accessible ones (length, area) is one of the main features
of the method of symmetrization. The idea of symmetrization has a centuries-old
history: it goes back to ancient Greeks. Today it finds applications in various ar-
eas of natural sciences. The pioneering research of symmetrization in mathematics
was carried out by Steiner and some time later by Hardy, Littlewood, Pólya and
Szegő. The development of the method of symmetrization in geometric theory of
functions of a complex variable has been considerably influenced by the articles
by Pólya and Szegő [PS], Hayman [H], Jenkins [J], Markus [M], Mituyk [Mit1],
Baernstein [Bae1]. As applied to this theory, the method of symmetrization devel-
ops in the following interrelated directions: constructing new transformations Sym
satisfying the symmetrization principle (0.2); finding new functions F(u) satisfying
(0.2) and investigating the properties of these functions related to certain classes
of conformal maps. In the present book the main object of study is the conformal
capacity cap C of a condenser C:

u = C, F(u) = cap C,

while Sym is some kind of geometric transformation of the condenser C such as


the Steiner symmetrization St or the circular symmetrization Cr. Hence the main
symmetrization principles will have the form

cap C  cap Sym C.

The limiting cases of these principles yield inequalities for the logarithmic capac-
ities of sets, the inner radii of domains, etc. Our choice of the conformal capacity
for F is motivated solely by applications to the theory of analytic functions. This
capacity is defined as the infimum of the Dirichlet integrals of real-valued func-
tions with prescribed boundary values. Moreover, the conformal invariance of the
Dirichlet integral provides opportunities for various types of symmetrization in the
plane. The role of the Dirichlet integral in many classical extremal problems in
analysis, geometry, and mathematical physics is well known. One can also demon-
strate the connection of this functional with the basic notions of such efficient
methods of geometric function theory as the area method, contour integration,
Grötzsch strips (see [Gr], [Gol]) and extremal metrics [Ahl2], [J], [Oht1].
The material offered here comprises primarily the research carried out in the
solution of some specific classical and modern problems in function theory. For
Preface xi

instance, Mityuk suggested proving the segment covering theorem ([Gol], Ch. IV,
§ 6) using symmetrization, which eventually led to the discovery of ‘separating
transformations’ of condensers. The Szegő problem which contains the segment
covering theorem, was also proved on the way, as well as a number of other re-
sults. Their proofs required the introduction of generalized condensers. At present,
separating transformations are part of the capacitive approach to various prob-
lems in function theory, which has a large intersection with the method of extremal
metrics and essentially coopts Grötzsch’s method of strips. Following the ideas of
Grötzsch and Teichmüller about applications of the reduced moduli of degener-
ating doubly-connected domains, we also consider condensers with two and more
plates some of which degenerate into points. In this way the notion of the reduced
modulus of a domain with respect to a system of boundary points, which is of
independent interest. Another line of research owes its appearance to Gonchar’s
problem about the condenser of minimum capacity. When plates of this condenser
lie on a straight line segment this problem was solved by Tamrazov [T1] by ‘mixing
signed measures’. Zorich and Gutlyanskii pointed out to the author that while the
result of [T1] is important, the method of solution is cumbersome and wondered if
other approaches to the solution of this problem could be developed. The answer
to their question and some other questions is contained in the simple observation
that the transformation of a real-valued function given by

max[u(z), u(z)], Im z > 0,
P u(z) =
min[u(z), u(z)], Im z  0,

preserves the Dirichlet integral [D2]. If the function u is admissible for a condenser
C then the function P u is admissible for the ‘polarized’ condenser P C with smaller
capacity. Clearly, this remark not only pertains to the Dirichlet integral but also
to quite general functionals depending on the values of a function and its first
partial derivatives [D3]. In the limit, repeated applications of polarization give
various types of symmetrization. In its essence this approach goes back to general
ideas about rearrangements [HaLiP] (see also Wolontis’ paper [W], p. 598). The
above considerations together with the independent research of other authors re-
sulted in the solution of many problems in function theory and partial differential
equations and inspired a new wave of research on polarization of functions and
condensers from various standpoints (see [Bae2], [Sol2], [Oht2], and many other
works). Finally, a totally new transformation, ‘dissymmetrization’ of functions
and condensers, appeared as the author was working on the solution of Gonchar’s
problem of harmonic measure. Just like symmetrization, dissymmetrization does
not increase the condenser capacity. Though this one has a possibility, so far quite
rare, to estimate the capacity of a condenser from above in terms of the capacity of
a symmetric condenser. (Answering numerous questions about the term ‘dissym-
metrization’ I note that the words ‘dissymmetrization’ and ‘desymmetrization’ are
equally applicable. I use the former term following some classical developments in
crystallography [Sh].)
xii Preface

The book does not seek to expound all questions relating to applications
of condenser capacity and symmetrization to geometric function theory. We have
deliberately minimized overlappings with the famous monographs [J] and [H]. On
the other hand, an attempt to encompass the state of the art would be too hard
a job for the author. In particular, one cannot imagine symmetrization methods
without the research of Baernstein related to polarization and the so-called *-
function [Bae1]–[Bae4]. In our view, another very important research direction
follows from Solynin’s proof of the theorem describing the behaviour of Green’s
function under polarization [Sol2]. Unfortunately, we could not touch upon inter-
esting works by Avkhadiev, Betsakos, Weitsman, Pruss, Fryntov and many others.
In comparison with the Russian edition, the list of literature was slightly extended;
in particular, several recent publications of the author were added, in which he
developed the methods presented in this book. However, we make no pretension
to completeness: given the current development of the Internet we do not think
this necessary.
The book is based on the survey article [D4], lectures on conformal invariants
and symmetrization regularly delivered at Far Eastern State University, and papers
by the author and his students E.G. Prilepkina (E.G. Akhmedzyanova), L.V. Ko-
valev, E.V. Kostyuchenko, V.Yu. Kim, D.A. Kirillova, and N.V. Èirikh. It is note-
worthy that the author has been influenced by lectures of G.K. Antonyuk and
I.P. Mityuk and by the research of Mityuk’s students B.E. Levitsky, V.A. Shlyk.
Yu.V. Chernykh and A.Yu. Solynin as well as by such scientific schools as
A.A. Gonchar’s seminar at Steklov Institute of Mathematics, G.V. Kuzmina’s
seminar at the St.-Petersburg Branch of Steklov Institute of Mathematics, and the
seminar organized by O. Martio and M. Vuorinen at the University of Helsinki.
The author expresses special thanks to E.V. Strizheva (E.V. Sysoeva) for type-
setting the manuscript and to D.B. Karp who produced all figures in this book.
The author appreciates the informal approach of N.G. Kruzhilin to the transla-
tion of the book, which contributed to some improvements of the text and to the
correction of misprints.
Chapter 1

Conformal Capacity

1.1 Lipschitz functions


Throughout, R will be the set of real numbers, C the plane of complex numbers
z = x + iy, C = C ∪ {∞}, and U (z0 , r) = {z : |z − z0 | < r} if z0 is a finite point;
we set U (∞, r) = {z : |z| > 1/r}.
For E ⊂ C let Lip(E) denote the class of Lipschitz functions v : E → R. This
means that for each v ∈ Lip(E) there exists a constant K such that

|v(z) − v(z  )|  K|z − z  |

for any finite points z, z  ∈ E In the case when ∞ ∈ E we also assume that v is
continuous at this point. We shall use the notation Lip(C) = Lip.
It can readily be verified that a Lipschitz function on a set E is continuous
on E, and if u and v are bounded Lipschitz functions on E ⊂ C, then uv also
belongs to Lip(E).
Theorem 1.1. Let E be a compact subset of C and let v : E → R be a function
such that each point in E has a neighbourhood U such that v ∈ Lip(U ∩ E). Then
v belongs to the class Lip(E).

Proof. Assuming the converse we can find two sequences of points {zn }∞
n=1 and
{z  }∞
n=1 , z n 
= z 
n , such that z n ∈ E\{∞}, z 
n ∈ E\{∞}, and

|v(zn ) − v(zn )|
→∞ as n → ∞. (1.1)
|zn − zn |

We can assume that zn → z0 and zn → z0 , where z0 and z0 are some points
in E. If z0 = z0 , then (1.1) contradicts the condition that v is bounded in some
neighbourhoods of z0 and z0 . On the other hand, if z0 = z0 , then (1.1) contradicts
the inclusion v ∈ Lip (U ∩ E), where U is a neighbourhood of z0 . 

© Springer Basel 2014 1


V.N. Dubinin, Condenser Capacities and Symmetrization in Geometric Function
Theory, DOI 10.1007/978-3-0348-0843-9_1
2 Chapter 1. Conformal Capacity

Theorem 1.2. If v is a continuously differentiable function in an open subset G


of C, then it is a Lipschitz function on compact subsets of G ( for each compact
subset E of G v ∈ Lip(E) ).
Proof. In view of Theorem 1.1, it is sufficient to verify that for each point z0 ∈ G
there exists an open disc U such that z0 ∈ U ⊂ G and v ∈ Lip(U ). Let U ,
z0 ⊂ U ⊂ G, be an arbitrary open disc such that the absolute values of the first
partial derivatives of v are bounded by some constant K in U . By the mean value
theorem, for any points z, z  ∈ U we have
 
 ∂v ∂v ∗ 
|v(z) − v(z  )| =  (z ∗ ) x + (z ) y 
∂x ∂y
 2  2
∂υ ∗ ∂υ ∗  √
 (z ) + (z ) ( x)2 + ( y)2  2K|z − z  |,
∂x ∂y

where z ∗ ∈ [z, z  ] ⊂ U u z − z  = x + i y. 
Theorem 1.3. Let Ek , k = 1, . . . , n, be closed subsets of C and assume that E =
n
Ek is a convex set. If v ∈ Lip(Ek ), k = 1, . . . , n, then v ∈ Lip(E).
k=1

Proof. If E contains z = ∞, then this point lies in some Ek , so v is continuous at z.


Now let z and z  be two finite points in E. We look at the line segment γ := [z, z  ]
directed in the natural way from z to z  and construct a finite sequence of points
z1 , . . . , zm in γ, m  n + 1, such that z1 = z, zm = z  , the point zi+1 goes after zi
and both zi+1 and zi lie in the same set in the system {Ek }nk=1 , i = 1, . . . , m − 1.
For example, this construction can be made as follows. Let Ek1 be a set containing
z1 . Then we set z2 := sup(γ ∩Ek1 ). As Ek1 is closed, thus z2 ∈ Ek1 . Now if z2 = z  ,
then there exists a set Ek2 containing z2 and distinct from Ek1 , for otherwise z2

n
belongs to the set γ\ Ek , which is open in γ. Then some neighbourhood (in
k=1
k=k1

n
γ) of z2 is not covered by the union Ek , in contradiction with the assumption
k=1
that E is convex. Now we set z3 = sup(γ ∩ Ek2 ), and so on. Let K be a common
Lipschitz constant for the sets Ek , k = 1, . . . , n. By construction,


m−1
m−1
|υ(z) − υ(z  )|  |υ(zi ) − υ(zi+1 )|  K |zi − zi+1 | = K|z − z  |.
i=1 i=1

Thus υ ∈ Lip(E). 
Throughout, by the truncation of a function υ : Ω → R = R ∪ {∞} we shall
mean the special case of truncation

trun υ(z) := max(min(υ(z), 1), 0), z ∈ Ω.


1.1. Lipschitz functions 3

Corollary 1.1. If u and v are functions in Lip, then max(u, v), min(u, v), and
trun v are also functions in this class. Furthermore, if u = v on some straight
line or circle α, which partitions C into two domains with closures C− and C+ ,
then the function equal to max(u, v) in C+ and to min(u, v) in C− belongs to the
class Lip.
Lemma 1.1. Let ζ = ϕ(z) be a function which maps a neighbourhood of a compact
set E ⊂ Cz conformally and univalently onto a neighbourhood of a set G ⊂ Cζ
so that ϕ(E) ⊂ G. Let v(ζ) ∈ Lip(G) and assume that E is convex. Then the
composite function v(ϕ(z)) belongs to Lip(E).

Proof. Let K be a Lipschitz constant for v(ζ) and assume that for some C > 0 we
have |ϕ (z)|  C for all z ∈ E. Then

|v(ϕ(z)) − v(ϕ(z  ))|  K|ϕ(z) − ϕ(z  )|  KC|z − z  |

for any points z, z  ∈ E. 


Theorem 1.4. Assume that a conformal univalent mapping ζ = ϕ(z) takes an open
set B ⊂ Cz to a set D ⊂ Cζ and assume that the boundaries of B and D consist
of finitely many piecewise analytic curves. Let H be the set of boundary points
of B at which the boundary fails to be analytic or which are taken to points at
which the boundary of D fails to be analytic. Let v(ζ) be a function in the class
Lip(D) taking equal values at the boundary points of D corresponding to the same
point in B. If E is a compact subset of B \ H such that every point z0 ∈ E has a
neighbourhood U (z0 , r) such that either U (z0 , r) ∩ B is connected or U (z0 , r) ⊂ B,
then v(ϕ(z)) belongs to Lip(E).

Proof. Let z0 be a point in E which also lies in B. Then it has a neighbour-


hood U (z0 , r) such that U (z0 , r) ⊂ B. By Lemma 1.1 the composite function
v(ϕ(z)) belongs to Lip(U (z0 , r)). Now let z0 ∈ E ∩ ∂B and assume that for some
neighbourhood U (z0 , r) the intersection U (z0 , r) ∩ B is connected. We extend
ζ = ϕ(z) to an analytic function in a neighbourhood of z0 keeping the same no-
tation. Assume that ζ = ψ(w) maps some disc U (0, 1 + ε), ε > 0, conformally
and univalently onto a neighbourhood of ϕ(z0 ) so that the points lying on the real
axis are taken to boundary points of D and points above the real axis are taken
to D. Applying Lemma 1.1 to the composite function v(ψ(w)) we conclude that
v(ψ(w)) ∈ Lip({w : |w|  1, Imw  0}). By Theorem 1.3, extending v(ψ(w)) into
the disc U (0, 1) by the formula

v(ψ(w)) = v(ψ(w)), w ∈ U (0, 1),




we obtain a function in Lip U (0, 1) . Using Lemma 1.1 we see that the corre-
sponding extensions of v(ζ) into a neighbourhood of ϕ(z0 ) and of v(ϕ(z)) into a
neighbourhood of z0 are Lipschitz in some neighbourhoods of these points. Hence
for some r > 0 we have v(ϕ(z)) ∈ Lip(U (z0 , r ) ∩ E). Finally, let z0 ∈ E ∩ ∂B
4 Chapter 1. Conformal Capacity

and assume that the set U (z0 , r) ∩ B is disconnected and that U (z0 , r) ⊂ B. Re-
peating the above reasoning for each connected
component and taking account of
Theorem 1.3 we obtain v(ϕ(z)) ∈ Lip U (z0 , r) . Using Theorem 1.1 we complete
the proof. 
We see from these theorems that the classes of Lipschitz functions are fairly
large and functions in these classes can be extended in the natural way. They also
have some properties of differentiable functions. For example, if v(x, y) := v(z) ∈
Lip(E) and an interval a  y  b of a vertical line x = const lies in E, then for
fixed x the function v(x, y) is absolutely continuous in y on [a, b]. Hence the partial
derivative ∂v/∂y exists almost everywhere on [a, b] and, moreover,
b
∂v(x, y)
v(x, b) − v(x, a) = dy. (1.2)
∂y
a

We leave out the proofs of the next two results because they are formally
the same as the corresponding proofs found in many textbooks on calculus or dif-
ferential equations. The formulae in these statements are called Green’s formulae.
Theorem 1.5 is a consequence of Fubini’s theorem and Newton–Leinbniz formula
(1.2). The proof of Theorem 1.6 reduces to the use of Theorem 1.5: verifying the
necessary conditions is a simple exercise.
Theorem 1.5. Let D be a bounded domain in the complex plane C with positively
oriented boundary γ, which is formed by finitely many analytic curves. Let u and
v be functions in Lip(D). Then
   
∂v ∂u
udx + vdy = − dxdy.
∂x ∂y
γ D

Theorem 1.6. Let D ⊂ C be a domain bounded by finitely many analytic curves


γ ⊂ C. Let u ∈ Lip(D) and let v be function with continuous second partial
derivatives in D. Then
    2   
∂v ∂ v ∂2v ∂u ∂v ∂u ∂v
u ds = − u + + + dxdy,
∂n ∂x2 ∂y 2 ∂x ∂x ∂y ∂y
γ D

where ∂/∂n denotes differentiation along the inward normal.


Functions in the class Lip are absolutely continuous on all the straight lines
parallel to the coordinate axes. They have bounded partial derivatives almost
everywhere in C, which are square integrable on each bounded measurable subset
B of C. In particular, for each function v ∈ Lip and every set B of this type the
integral
   2  2 
∂v ∂v
I(v, B) := + dxdy,
B ∂x ∂y
1.1. Lipschitz functions 5

is well defined; it is called the Dirichlet integral. The integral I(v, B) is treated as
Lebesgue integral and is usually taken over a Borel set B. If B is an unbounded
Borel set, then the Dirichlet integral I(v, B) is defined as an improper integral.
If B ⊂ C is an open set and v ∈ Lip(E) for each compact subset E of B, then
I(v, B) is also defined as an improper integral; more precisely,

I(v, B) = lim I(v, Bn ),


n→∞

where {Bn }∞
n=1 is an exhaustion of B, that is, a sequence of open sets Bn with


compact closures B n such that B n ⊂ Bn+1 ⊂ B, n = 1, 2, . . . , Bn = B.
n=1

Theorem 1.7. Let B be an open subset of C, z0 ∈ B, B0 = B\{z0 }, and let v : B0 →


[0, 1] be a Lipschitz function on compact subsets of B0 such that I(v, B0 ) < +∞.
Then for any positive numbers r and ε there exist functions vj : B → [0, 1], j =
0, 1, that are Lipschitz on compact subsets of B and have the following properties :
v0 = v1 = v outside U (z0 , r); vj = j in a neighbourhood of z0 , and |I(vj , B0 ) −
I(v, B0 )| < ε, j = 0, 1.

Proof. For fixed r > 0 and some r1 and r2 , r2 < r1 < r, we set

[log |r1 /(z − z0 )|]/ log(r1 /r2 ) for z0 = ∞,
u(z) =
[ log |r1 z|]/ log(r1 /r2 ) for z0 = ∞.

Direct calculations show that

I (u, U (z0 , r1 )\U (z0 , r2 )) = 2π/ log(r1 /r2 ).

In view of Theorems 1.1–1.3, the functions

v0 = min(v, 1 − trun u) and v1 = max(v, trun u),

which are defined at z0 by continuity, belong to Lip(E) for each compact subset
E of B and have the following properties: 0  vj  1 on B, j = 0, 1; v0 = v1 = v
outside U (z0 , r1 ); v0 = 0 and v1 = 1 in a neighbourhood of z0 . Now if r1 and r2
are positive numbers such that

I(v, U (z0 , r1 )\{z0 }) < ε/2 and 2π/ log(r1 /r2 ) < ε/2,

then

|I(vj , B0 ) − I(v, B0 )|
= |I(vj , {z ∈ B0 : vj (z) = v(z)}) − I(v, {z ∈ B0 : vj (z) = v(z)})|
 I(u, U (z0 , r1 )\U (z0 , r2 )) + I(v, U (z0 , r1 )\{z0 }) < . 
6 Chapter 1. Conformal Capacity

Exercises 1.1
(1) Show that it is essential in Theorem 1.1 that E is a compact set.
(2) Give an example of a function v : G → R that is continuously differentiable
on an open set G ⊂ C, but not Lipschitz in G.
(3) Can we drop the assumption in Theorem 1.3 that the sets Ek , k = 1, . . . , n
are closed?
(4) Show that the convexity assumption is essential in Theorem 1.3
(5) Prove Corollary 1.1.
(6) Prove Theorem 1.6.
(7) Prove that any domain B ⊂ C can be exhausted by a sequence of domains
each of which is bounded by finitely many analytic curves.
(8) Show that if a function v is Lipschitz on compact subsets of an open set B,
then the value of the Dirichlet integral I(v, B) is independent of the way in
which B is exhausted by a sequence of sets {Bn }∞ n=1 .
(9) Express the Dirichlet integral in the polar coordinates.
(10) Calculate the integral I (u, U (z0 , r1 )\U (z0 , r2 )) in the proof of Theorem 1.7.
(11) Under the assumptions of Theorem 1.4 verify that the Dirichlet integral is
conformally invariant, that is, I (v(ϕ(z)), B) = I(v, D).
(12) Let f = u+iv be a function mapping a domain G conformally and univalently
onto a domain f (G). Show that the area of f (G) is equal to the Dirichlet
integral I(u, G).

1.2 Condensers with two plates on the Riemann sphere


By a condenser on the Riemann sphere C we mean an arbitrary ordered pair
C = (E0 , E1 ) of disjoint closed nonempty subsets E0 and E1 of C. We call the
open set G = C\(E0 ∪ E1 ) the field of the condenser C, and we call the sets E0
and E1 the plates of C. Some authors also use the notation C = {G, E0 , E1 }. The
capacity of the condenser C is the quantity

cap C = inf {I(v, C) : v ∈ Lip, v = j on Ej , j = 0, 1} .

Since for a function v in the class Lip its truncation trun v also belongs to this
class (Corollary 1.1) and I(trun v, C)  I(v, C), it follows that

cap C = inf {I(v, C) : v ∈ L (C)} ,

where L (C) = {v : v ∈ Lip, 0  v  1 on C, v = j on Ej , j = 0, 1}. We can


show that if we replace the class Lip in the definition of the capacity by the set
of continuous functions in C belonging to the well-known class ACL2 (G\ {∞}) or
by the set of continuous functions in C that are infinitely smooth in G\ {∞} and
1.2. Condensers with two plates on the Riemann sphere 7

have a finite Dirichlet integral, or by subclasses of these classes consisting of the


functions vanishing in a neighbourhood of E0 and equal to 1 in a neighbourhood
of E1 , then the value of the capacity remains the same (cf. [GResh], § 5.1). Note
also that if v = j on Ej , then 1 − v = j on E1−j , j = 0, 1 and I(1 − v, C) = I(v, C).
Hence
cap (E0 , E1 ) = cap (E1 , E0 ).

We shall say that a condenser C2 = (E20 , E21 ) contains a condenser C1 =


(E10 , E11 ) if E2j ⊃ E1j , j = 0, 1. Usually this is expressed as C2 ⊃ C1 . The
following important property of monotonicity is an immediate consequence of the
definition of capacity.
Theorem 1.8. If C2 ⊃ C1 , then

cap C2  cap C1 .

We can show that if C2 ⊃ C1 and the fields of the condensers C2 and C1


are distinct nondegenerate doubly connected domains, then cap C2 > cap C1 (see
Theorem 1.14).
Theorem 1.9. The addition of finitely many points to a plate of a condenser does
not change does not change its capacity.

Proof. We can assume that we add a point z0 to the plate E0 and that this point
lies in the field of the condenser C = (E0 , E1 ). We set C0 = (E0 ∪ {z0 }, E1 ). Let
v be a function in the class L (C) such that I(v, C) < +∞. Let ε be a positive
number and r be a positive number such that the neighbourhood U (z0 , r) lies in
C\(E0 ∪ E1 ). Then the function v0 in Theorem 1.7 lies in the class L (C0 ), and
we have
|I(v0 , C) − I(v, C)| < .
Thus
cap C0  I(v0 , C)  I(v, C) + ε.
Hence
cap C0  cap C.
The reverse inequalities are consequences of the monotonicity property of capacity.


Theorem 1.10. If Ck = (Ek0 , Ek1 ), k = 1, 2, are condensers, then

cap (E10 ∪ E20 , E11 ∩ E21 ) + cap (E10 ∩ E20 , E11 ∪ E21 )  cap C1 + cap C2

(if some plates have empty intersection then the capacity of the corresponding
‘condenser’ is set equal to zero).
8 Chapter 1. Conformal Capacity

Proof. This follows from the equality

I(v1 , C) + I(v2 , C) = I(min(v1 , v2 ), C) + I(max(v1 , v2 ), C),

where v1 and v2 are arbitrary functions in the classes L (C1 ) and L (C2 ), re-
spectively, and the functions min(v1 , v2 ) and max(v1 , v2 ) belong to Lip by Corol-
lary 1.1. 
We call a sequence {Cn }∞ n=1 of condensers Cn = (En0 , En1 )∞an exhaustion
of a condenser C = (E0 , E1 ) if Enj ⊃ En+1j , n = 1, 2 . . . , and n=1 Enj = Ej ,
j = 0, 1.
It is natural to call the following property of condenser capacity continuity.
Theorem 1.11. If {Cn }∞
n=1 is an exhaustion of a condenser C, then

cap C = lim cap Cn .


n→∞

Proof. By Theorem 1.8 the sequence cap Cn is decreasing and bounded below by
the quantity cap C. Hence it has a limit

lim cap Cn = inf cap Cn  cap C.


n→∞ n

Assume that the last inequality is strict. Then there exists a function v in the class
L (C) such that
inf cap Cn > I(v, C). (1.3)
n

For k = 2, 3, . . . the sets {z : v(z) < 1/k} and {z : v(z) > 1 − 1/k} are open and
contain the closed sets E0 and E1 , respectively. Then for each k > 1 there exists
a condenser Cnk = (Enk 0 , Enk 1 ) such that

Enk 0 ⊂ {z : v(z) < 1/k} and Enk 1 ⊂ {z : v(z) > 1 − 1/k}.

Hence
     2
kv − 1 kv − 1 k
cap Cnk  I trun , C I , C = I(v, C),
k−2 k−2 k−2

which is in contradiction with (1.3) if k is sufficiently large. 


The key property for applications is the conformal invariance of capacity. Let
C = {G, E0 , E1 } be a condenser in Cz and C  = {G,
 E0 , E
1 } be a condenser in Cw ;
let w = f (z) be a function mapping the field G conformally and univalently onto
the field G so that the plates of C correspond to the plates of C:  this means that
if z ∈ G and the spherical distance ρ(z, Ej ) → 0 then ρ(f (z), E j ) → 0, j = 0, 1.

In this case we call C the image of the condenser C and use the notation f (C).
Obviously, f −1 (f (C)) = C.
1.2. Condensers with two plates on the Riemann sphere 9

Theorem 1.12. The equality


cap C = cap f (C)
holds.
Proof. Let Cn = {Gn , En0 , En1 }∞ n=1 be an exhaustion of the condenser C by a
sequence of condensers with plates bounded by finitely many analytic curves.
Then for each n the condenser f (Cn ) also has this property and furthermore,
f (Cn ) ⊃ f (C). Let v ∈ L (Cn ). By Theorem 1.4 the function v(f −1 (w)) which is
extended to Cw \f (Gn ) by continuity and takes there the values 0 and 1 belongs
to the class L (f (Cn )). As the Dirichlet integral is conformally invariant, we have
I(v, Cz ) = I(v(f −1 (w)), Cw )  cap f (Cn )  cap f (C).
Taking the minimum we obtain
cap Cn  cap f (C).
It remains to use the fact that the condenser capacity is continuous and then
repeat the above arguments for the inverse mapping z = f −1 (w). 
We say that a condenser C = {G, E0 , E1 } is admissible if there exists a
continuous real function ω(z) in C that is equal to 0 on E0 , to 1 on E1 and
is harmonic in G. We call ω(z) the potential function of the condenser C. It is
known that if each boundary point of the field G of the condenser C is an end-
point of some arc lying outside G, then C is admissible [R]. Note that although
the potential function does not necessarily belong to the class Lip, the function
ωt (z) = trun((ω(z) − t/(1 − 2t)), 0 < t < 1/2, belongs to this class and
cap C  I(ωt , G)  (1 − 2t)−2 I(ω, G) → I(ω, G), t → 0.
Moreover, the following result holds.
Theorem 1.13. Let C = {G, E0 , E1 } be a condenser in C and let v be a continuous
function in C which vanishes on E0 , is equal to 1 on E1 , and is Lipschitz in some
neighbourhood of every point in G with the possible exception of finitely many
points. Then
I(v, G)  cap C. (1.4)
In (1.4) equality holds if and only if C is an admissible condenser and v is its
potential function.
Proof. Inequality (1.4) can be deduced by the techniques developed in the proofs of
the previous results. The assertion about equality sign follows from Dirichlet’s prin-
ciple and is based on the properties of harmonic functions. For the proofs of various
modifications of Dirichlet’s principle the reader can consult monographs on func-
tion theory, the theory of potential, or partial differential equations [C], [Sobol], [H]
(see also Appendix A1). 
10 Chapter 1. Conformal Capacity

Now we introduce the notions of


6 5 a regular condenser and its decompo-
sitions. Note that the field of an ar-
H4 bitrary condenser consists of countably
4 many disjoint domains. We say that the
3
H condenser is regular if each of the do-
C 1 mains making up its field has boundary
points on both plates of the condenser.
Let C = {G, E0 , E1 } be a regular admis-
H3 sible condenser and ω(z) be the poten-
C tial function. Then it is easy to see that
the field G consists of finitely many do-
Figure 1.1: Grötzsch’s lemmas. mains, G = {z : 0 < ω(z) < 1} and
Ej = {z : ω(z) = j}, j = 0, 1. Let
Ck = {{z : tk < ω(z) < tk+1 }, {z :
ω(z)  tk }, {z : ω(z)  tk+1 }}, k = 1, . . . , n, where 0 = t1 < t2 < · · · < tn+1 = 1.
We call the ordered system of condensers {C1 , . . . , Cn } the decomposition of C
corresponding to the partition {tk }n+1
k=1 . The next result is a far-reaching general-
ization of the well-known Grötzsch lemma (cf. [J], Theorem 2.6 and [PS]).

Theorem 1.14. If condensers C = {G, E0 , E1 } and Ck = {Gk , Ek0 , Ek1 }, k =


1, . . . , n, satisfy the following conditions :

Gk ∩ Gl = ∅, k = l, k, l = 1, . . . , n,

n
Ej ⊂ Ekj , j = 0, 1,
k=1
then
 −1

n
−1
cap C  (cap Ck ) . (1.5)
k=1

In addition, if C and the Ck , k = 1, . . . , n, are regular admissible condensers, then


equality in (1.5) holds if and only if the system {C1 , . . . , Cn } is a decomposition
of the condenser C up to ordering.

Proof. Let vk be a function in the class L (Ck ), k = 1, . . . , n. Then the function


n 
n
v= dk vk , where 0  dk  1 and dk = 1, belongs to the class L (C). Using
k=1 k=1
the definition of capacity and the first assumption of the theorem we obtain in
succession
n
n
cap C  I(v, G) = I(v, Gk ) = d2k I(vk , Gk ).
k=1 k=1
1.2. Condensers with two plates on the Riemann sphere 11

Taking the lower bounds over the vk we obtain



n
cap C  d2k cap Ck ,
k=1
n
which yields inequality (1.5) for dk = (1/cap Ck )/ k=1 (1/cap Ck ).
In Figure 1.1 we give the simplest case of two Grötzsch lemmas. The parts
related to Theorem 1.14 are the two shaded ring domains G1 and G2 bounded by
the Jordan curves γ1 , γ2 , γ3 , and γ4 (n = 2). The field G of the condenser C is
bounded by γ1 and γ4 . The plate E1 = E11 is the closure of the interior of the
curve γ1 ; E0 = E20 is the closure of the exterior of γ4 ; E10 is the closure of the
exterior of γ2 and E21 is the closure of the interior of γ3 .
Now assume that C and the Ck , k = 1, . . . , n, are regular admissible con-
densers and we have equality in (1.5). Let ω and ωk be the potential functions of the
n
condensers C and Ck , respectively, and let vk = ωk . Then ω ≡ dk ωk by Dirich-
k=1

n
let’s principle (Exercise 1.2(4)). Hence G = Gk and ω(z) = dk ωk (z) + tk , z ∈
k=1
Gk , where the tk are some constants, 0  tk < 1, k = 1, . . . , n. Now we change
n
the numbering of the Gk . Since G = Gk , at least one of these sets (which we
k=1
denote by G1 ) has a boundary point on E0 . Then t1 = 0. Assume that d1 < 1
and let z1 be a smooth point of the curve {z : ω(z) = d1 }. Then z1 is a boundary
point of some Gk distinct from G1 , for instance, of G2 , so that t2 = d1 . Assume
that d1 + d2 < 1 and let z2 be a smooth point of the curve {z : ω(z) = d1 + d2 }. It
is a boundary point of some Gk , for instance, of G3 = G2 , G1 . Then t3 = d1 + d2
and so on, until we obtain d1 + · · · + dn = 1.
Next we set G  k = {z : tk < ω(z) < tk+1 }, k = 1, . . . , n (tn+1 = 1). By defini-
 k ⊃ Gk . In view of the equality G =  Gk , we have G
n
tion G  k = Gk , k = 1, . . . , n.
k=1
We see that the system {Ck }nk=1 with the new numbering is a decomposition of
the condenser C.
Conversely, let {Ck }nk=1 be the decomposition of the condenser C correspond-
ing to a partition {tk }n+1
k=1 and let ω be the potential function of C. Then the
function ωk = (ω − tk )/(tk+1 − tk ) coincides in the field Gk with the potential
function of the condenser Ck in the field Gk . Bearing in mind that y = 1/x is a
convex function we find
 n −1

−1
(cap C) = (tk+1 − tk )I(ω, Gk )/(tk+1 − tk )
k=1

n
n
n
 (tk+1 − tk )2 /I(ω, Gk ) = (I(ωk , Gk ))−1 = (cap Ck )−1 .
k=1 k=1 k=1

In combination with (1.5) this gives us the required equality. 


12 Chapter 1. Conformal Capacity

Exercises 1.2
(1) Show that the class L (C) is nonempty.
(2) Show that the capacity of a condenser does not change if in place of the condi-
tions ‘v = j on Ej ’ we assume in the definition that v = j in a neighbourhood
of the set Ej , j = 0, 1.
(3) Show that any condenser has an exhaustion by condensers with plates bound-
ed by finitely many analytic Jordan curves.
(4) Prove Theorem 1.13.
(5) Let C = {G, E0 , E1 } be an admissible condenser and v : C → R a continuous
function in C which is continuously differentiable on an open subset B of
G. Assume that there exists a sequence of functions vj ∈ L (C) such that
∇vj = 0 a.e. on Cz \B, j = 1, 2, . . . ; vj → v on Cz as j → ∞, and ∇vj ⇒ ∇v
on compact subsets of B as j → ∞. Show that if I(v, B) = cap C, then v is
the potential function of the condenser C.
(6) Let G be a doubly connected domain on the Riemann sphere and let E1 and
E2 be the connected components of the complement of G. Assume that a
function f maps G conformally and univalently onto an annulus r1 < |z| < r2 ;
0 < r1 < r2 < ∞. Then the quantity

1 r2
MG = log
2π r1

is called the modulus of G separating the boundary components of G [J]).


Show that
M G = 1/ cap(E1 , E2 ).

(7) Find the capacity of the condenser ({z = x+ iy : |x|  1, y = 0}, {z = x+ iy :


|x|  t, y = 0}), t > 1.
(8) Bearing in mind the interpretation of condenser capacity in electrostatics
give a physical interpretation of Theorem 1.14 (see [PS]).

1.3 Generalized condensers


Let B be an open subset of C. A generalized condenser in B is a triple C =
(B, E, ), where E = {Ek }nk=1 is a system of closed pairwise disjoint nonempty
sets Ek ⊂ B, k = 1, . . . , n, and = {δk }nk=1 is a tuple of real numbers δk ,

n
k = 1, . . . , n, n  2. The set B \ Ek , which is open in B, is called the field of
k=1
the condenser C, the set Ek are the plates of the condenser, and the δk are the levels
of the potential or briefly potentials of the plates Ek , k = 1, . . . , n. The capacity
1.3. Generalized condensers 13


n
cap C of C is defined as the infimum of the Dirichlet integrals I(v, B \ Ek )
k=1
over all the admissible functions v, that is, real functions v which are continuous
in B, Lipschitz on compact subsets of B and equal to δk on Ek , k = 1, . . . , n.
n
If (∂B) \ Ek is empty or consists of finitely many analytic curves and there
k=1

n
exists a continuous function u in B which is harmonic in B \ Ek , equal to δk
k=1

n
on Ek , k = 1, . . . , n, and satisfies ∂u/∂n = 0 on (∂B) \ Ek , then we call u
k=1
the potential function of C and the condenser C is said to be admissible. Using
conformal maps we can extend these definitions to a broader class of condensers.
Under the assumptions of Theorem A1 the potential function exists and

cap C = I(u, B).

In what follows, speaking about condensers we shall suppress the adjective ‘gen-
eralized’. Obviously,

C = (E0 , E1 ) ≡ (C, {E0 , E1 }, {0, 1}).

When B is a finitely connected domain in C we shall also consider condensers


of the form C = (B, E, ), defined as above, but with B replaced by [B], the com-
pactification of the domain B by means of Carathéodory prime ends; its boundary
∂[B] is the set of prime ends [Gol]. In this case we mean by a neighbourhood an
arbitrary open subset of [B], and plates in E are closed subsets of [B]. It is impor-
tant for applications to distinguish between different attainable boundary points
of B having the same support. For simplicity, in this and the following sections we
mostly give statements for condensers endowed with the topology induced from
C. For condensers in [B] we have similar results, but their statements (and the
main of their proofs) coincide in appearance with the corresponding results for
condensers in B, so we do not present them. We can explain the difference be-
tween the two types of condenser to the reader not acquainted with the theory
of prime ends using the following simple example (see Figure 1.2). Let B be the
unit disc |z| < 1 cut along the radius (0, 1), E1 be the arc of the circle |z| = 1
going from i to −1, and E2 be the interval [1/2, 1]. Let f be a conformal univalent
mapping of B onto the disc U := {w : |w| < 1} and assume that E1 corresponds to
an arc α of the circle |w| = 1 under the mapping f 1 , the upper side of the interval
[1/2, 1] corresponds to an arc β and its lower side to an arc γ. Then the condenser
C = (B, {E1 , E2 }, {0, 1}) in B corresponds to the condenser (U, {α, β ∪ γ}, {0, 1})
with the same capacity. On the other hand, if E 2 is the upper side of the cut made

along [1/2, 1], then the condenser C = (B, {E1 , E2 }, {0, 1}) in [B] corresponds to

(U, {α, β}, {0, 1}), but cap C = cap C.
1 so that for a point z in B approaching E1 , the point f (z) approaches α.
14 Chapter 1. Conformal Capacity

j

F3
f4
F 2
2 3
3


Figure 1.2.

In a similar way to condensers with two plates, the definition of conformal


capacity can be modified as follows. However, showing that the two definitions are
equivalent requires some work in general.

Lemma 1.2. If the class of admissible functions is reduced to the subclass of func-
tions equal to δk in a neighbourhood 2 of the plate Ek , k = 1, . . . , n, then the
capacity of a condenser C = (B, E, ) does not change.

Proof. Making a conformal mapping we can assume that one plate of the condenser
C contains the point at infinity. Without loss of generality we also assume that

−1 = δ1 < δ2 < · · · < δn = 1.

Let cap0 C be the infimum of the Dirichlet integrals over all the admissible func-
tions equal to δk in a neighbourhood of Ek , k = 1, . . . , n. Since we have narrowed
down the class, we have
cap C  cap0 C.

Now let v be an arbitrary admissible function of the condenser C and assume that

n
I(v, B \ Ek ) < ∞. For each k we look at neighbourhoods Uk and Uk satisfying
k=1
the following conditions: each Uk (Uk ) is the intersection of B with an open subset

of C bounded by finitely many closed analytic Jordan curves; Ek ⊂ Uk ⊂ U k ⊂ Uk ;
the oscillation of v on Uk does not exceed (min(δ2 − δ1 , δn − δn−1 ))/2; U k ∩ U l = ∅,
k = l, k, l = 1, . . . , n. Let ω be some fixed continuous function in B which is

n−1
Lipschitz on compact subsets of B, vanishes on B \ Uk , is equal to δk on
k=2
 
n−1
U k, k = 2, . . . , n−1, and satisfies I(ω, B\ Ek ) < ∞. Then such neighbourhoods
k=2
and such a function ω exist is easy to establish (for n = 2, ω ≡ 0). For sufficiently

2 By a neighbourhood in B we mean the intersection of B with an open subset of C.


1.3. Generalized condensers 15

small ε > 0 we define the following function on B:




⎪ 1 + ω(z) for v(z)  ω(z) + 1 − ε,





⎪ v(z) − 2εω(z) − ε

⎪ for ω(z) + ε  v(z)  ω(z) + 1 − ε,

⎪ 1 − 2ε



vε (z) = ω for ω(z) − ε < v(z) < ω(z) + ε,



⎪ v(z) − 2εω(z) + ε

⎪ for ω(z) − 1 + ε  v(z)  ω(z) − ε,

⎪ 1 − 2ε






⎩ −1 + ω(z) for v(z)  ω(z) − 1 + ε,

where v(z) = max(min(v(z), 1), −1), z ∈ B. It readily follows from the theorems
in § 1.1 that vε (z) is continuous in B and Lipschitz on compact subsets of B. Now
we have vε (z) = −1 on the set {z ∈ U1 : v(z) < ω(z) − 1 + ε}, which is open
in B, and vε (z) = 1 on the open set {z ∈ Un : v(z) > ω(z) + 1 − ε}. In the
intersection of the open sets Uk and {z : ω(z) − ε < v(z) < ω(z) + ε} we have
vl (z) = δk , k = 2, . . . , n − 1. Thus


n
I(vε , B \ Ek )  cap0 C.
k=1

Passing to the limit as ε → 0 we obtain


n 
n
I(v, B \ Ek )  I(v, B \ Ek )  cap0 C.
k=1 k=1

Taking the lower bound over the functions v we complete the proof. 

A simple but important example of a condenser which is different from


the examples in § 1.2 is a quadrilateral with two marked opposite sides. By a
quadrilateral (Q; a, b, c, d) we mean a simply connected domain Q on the Riemann
sphere C with four distinct boundary points (prime ends) a, b, c, d, arranged in
accordance with the positive direction of the boundary of Q. These points are
called vertices of the quadrangle, while the pieces of the boundary of Q lying be-
tween neighbouring vertices (but not containing the vertices themselves) are called
sides and denoted by (a, b), (b, c), (c, d), and (d, a). The capacity of the condenser
(G, {(a, b), (c, d)}, {0, 1}) coincides with the conformal modulus of the quadrilat-
eral (Q; a, b, c, d) with respect to the sides (a, b) and (c, d) (see Exercise 1.3(1)).
Now we go over to the properties of condenser capacity, which are mainly
responsible for the importance of this concept in geometric function theory. We
start with several types of monotonicity.
16 Chapter 1. Conformal Capacity

Theorem 1.15. If two condensers C1 = (B1 , {E1k }nk=1 , {δk }nk=1 ) and C2 = (B2 ,
{E2k }m
k=1 , {δk }k=1 ), where n  m, satisfy B1 ⊂ B2 and E1k ⊂ E2k , k = 1, . . . , n,
m

then
cap C1  cap C2 .

Proof. Let v be an admissible function for C2 . Then its restriction v1 to B 1 is


admissible for the condenser C1 , and we have

cap C1  I(v1 , B1 )  I(v, B2 ).

It remains to take the lower bound over all possible admissible functions v. 

We agree to say that a set B2 is


obtained by extending a set B1 across
F3 F4 a part of its boundary γ ⊂ ∂B1 if
C B1 ⊂ B2 and (∂B1 ) ∩ B2 lies in γ. It
follows from the above theorem that if
B is extended across parts of ∂B not
lying on the plates Ek , k = 1, . . . , n,
then the capacity of the condenser C =
(B, {Ek }nk=1 , ) does not decrease. If B
Figure 1.3: Examples when the capacity is extended across plates lying on ∂B,
of a condenser decreases. we have the reverse inequality.
Theorem 1.16. Let Ci = (Bi , {Eik }nk=1 , ), i = 1, 2, and assume that B2 is ob-
tained by extending B1 across the parts E1k ∩ ∂B1 of the boundary, k = 1, . . . , n,
so that each connected component of the set B 2 \ B 1 has common boundary points
with at most one plate of C1 . Assume that E2k ⊂ E1k ∪ (B 2 \ B 1 ), k = 1, . . . , n.
Then
cap C1  cap C2 .

Proof. To any admissible function v1 for the condenser C1 which is equal to δk


in a neighbourhood of E1k , k = 1, . . . , n, we assign the function v2 constructed
as follows. We set v2 = v1 on B 1 . If a connected component of B 2 \ B 1 has
common boundary points with the plate E1k , then we set v2 = δk in this connected
component. We set v2 = 0 at the other points in B 2 . It is easy to see that v2 is an
admissible function for C2 and

I(v1 , B1 ) = I(v2 , B2 )  cap C2 .

It remains to take the lower bound over all possible functions v1 . 

From the conformal invariance of the Dirichlet integral and the boundary
properties of univalent functions we can deduce various types of conformal in-
variance of the capacity of a condenser C = (B, {Ek }nk=1 , ). In practice, the
1.3. Generalized condensers 17


n
preservation of capacity by conformal univalent mappings of the set B \ ( Ek )
k=1
is easy to establish in each particular case. So here we do not present these re-
sults, but just point out that if f is a conformal univalent function in B, defined
by means of boundary correspondence on the boundary of B, then we use the
notation
f (C) = (f (B), {f (Ek )}nk=1 , ).
Now we discuss in greater detail the composition principles for condensers.
Theorem 1.17. Let Bi , i = 1, . . . , m, be pairwise disjoint open subsets of an
open set B ⊂ C, and assume that condensers Ci = (Bi , {Eij }nj=1 i
, {δij }nj=1
i
),
i = 1, . . . , m, and C = (B, {Ek }k=1 , {δk }k=1 ) satisfy the following condition: each
n n

plate Eij , 1  j  ni , of any condenser Ci , 1  i  m, lies in the union of the


plates of C having the same potential as Eij . Then


m
cap Ci  cap C.
i=1

Proof. Let v be an admissible function for C. Then the restrictions vi of this


function to the sets B i are admissible for the corresponding condensers Ci , i =
1, . . . , m. Hence

m
m
I(v, B)  I(vi , Bi )  cap Ci .
i=1 i=1

Taking the infimum we complete the proof. 


Note a useful consequence of Theorem 1.17. Sometimes, the reciprocal of the
capacity
1
|C| :=
cap C
is called the modulus of the condenser .

m
Let λi , 0  λi  1, i = 1, . . . , m, be real numbers, λi = 1. Bearing in
i=1
mind that y = 1/x is a convex function we obtain
 −1  −1

m
m
m
−1 −1
|C|  |Ci | = λi (λi |Ci |)  λ2i |Ci |.
i=1 i=1 i=1

Thus under the assumptions of Theorem 1.17 we have



m
|C|  λ2i |Ci |
i=1

(cf. [Sol3], Theorem 1.5).


18 Chapter 1. Conformal Capacity

Theorem 1.18. Let Ci = (Bi , {Eij }nj=1


i
, {δij }nj=1
i
), i = 1, . . . , m, and

C = (B, {Ek }nk=1 , {δk }nk=1 )

be condensers satisfying the following conditions:



m
1) B = Bi;
i=1
2) for i = i each point in the set B i ∩B i (i, i = 1, . . . , m), also belongs to some
plates of the condensers Ci and Ci whose potentials in these condensers are
equal ;
3) each plate Ek of the condenser C lies in the union of the plates of the con-
densers Ci whose potentials are equal to that of Ek .
Then

m
cap Ci  cap C.
i=1

Proof. Let vi be an admissible function for Ci which is equal to δij in a neigh-


bourhood of the plate Eij , j = 1, . . . , ni , i = 1, . . . , m. We define a function v(z)
on B to be equal to vi (z) for z belonging to B i , i = 1, . . . , m. By condition 2) the
function v is well defined and by conditions 1)–3) and the definitions of the vi it
is admissible for the condenser C. Hence

m
I(vi , Bi ) = I(v, B)  cap C.
i=1

It remains to take the greatest lower bounds over all possible functions vi , i =
1, . . . , m. The proof is complete. 

We illustrate the composition principles by the following simple examples.

F34 F44
3 F44
3
C4
C3 C4 
C3 F34 ? F43
2 F33
2 F33
F43

Figure 1.4.

In the left-hand figure we give the case of Theorem 1.17. Here B is the union
of the sets B1 and B2 (or the interior of the closure of this union), E1 = E11 ∪ E21
1.3. Generalized condensers 19

and E2 = E12 ∪ E22 . By Theorem 1.17,


cap(B, {E1 , E2 }, {0, 1})
 cap(B1 , {E11 , E12 }, {0, 1}) + cap(B2 , {E21 , E22 }, {0, 1}).
In the right-hand figure, defining the set B as before and using Theorem 1.18 we
obtain the inequality
cap(B, {E11 , E22 }, {0, 1})
 cap(B1 , {E11 , E12 }, {0, δ}) + cap(B2 , {E21 , E22 }, {δ, 1})
for each real δ (cf. [Ahl2], p. 55, Figs. 4-2).

m
Theorem 1.19. For any real λi such that 0  λi  1, i = 1, . . . , m, and λi = 1,
i=1


m
cap C  λi cap Ci ,
i=1

where
Ci = (B, {Ek }nk=1 , {δki }nk=1 ), i = 1, . . . , m,
and   n 

m
C= B, {Ek }nk=1 , λi δki .
i=1 k=1

Proof. Let vi be admissible functions for the condensers Ci , i = 1, . . . , m. Then



m
v := λi vi is an admissible function for C. By the definition of condenser capac-
i=1
ity, bearing in mind that y = x2 is a convex function we obtain in succession
   m
2

m 
 
cap C  I(v, B) = ∇ λi vi  dxdy  λi |∇vi |2 dxdy.
 
B i=1 i=1 B

Taking the infimum we obtain the required inequality for capacities. 


In conclusion we present an analogue of the symmetry principle used in the
method of extremal lengths (cf. [Oht1], Theorem 2.47). For an open set B ⊂ C let
B + = {z ∈ B : Im z > 0}, and for a closed set E let E + = {z ∈ E : Im z  0}. A
set A is said to be mirror-symmetric relative to the real axis if A = {z : z ∈ A}.
We say that a condenser C = (B, E, ) is mirror-symmetric relative to the real
axis if the set B and the plates in E have this symmetry.
Theorem 1.20. If a condenser C = (B, {Ek }nk=1 , {δk }nk=1 ) is mirror-symmetric
relative to the real axis, then
cap C = 2 cap C + ,
where C + = (B + , {Ek+ }nk=1 , {δk }nk=1 ).
20 Chapter 1. Conformal Capacity

Proof. Standard arguments show that it is sufficient to look at the case when the

n
sets B and Ek are bounded by finitely many Jordan curves. Making a conformal
k=1

n
mapping if necessary we can assume that the boundary of G := B\ Ek is formed
k=1
by finitely many analytic curves. In this case let u be the potential function of the
condenser C. By uniqueness, u(z) ≡ u(z), that is, u is mirror-symmetric relative
to the real axis. (In particular, ∂u/∂n = 0 at points in (∂B + ) ∩ G lying on the
real axis.) Hence u is the potential function of the condenser C + , and therefore
cap C = I(u, B) = 2I(u, B + ) = 2 cap C + . 

It is clear how to state the symmetry principle for an arbitrary straight line
(distinct from the real axis) or a circle.

Exercises 1.3
(1) Let w = f (z) be a function mapping conformally and univalently the quadri-
lateral (Q; a, b, c, d) onto a rectangle 0 < u < 1, 0 < v < M (w = u + iv)
and taking the vertices a, b, c, d to 0, 1, 1 + iM, iM , respectively. Then the
quantity
M = M (Q; a, b, c, d)
is called the (conformal) modulus of the quadrilateral (Q; a, b, c, d) with re-
spect to the sides (b, c) and (d, a) (or the modulus of the family of curves
lying in Q and joining the sets (b, c) and (d, a) [J]).
Show that
M (Q; a, b, c, d) = cap(Q, {(b, c), (d, a)}, {0, 1})
= 1/ cap(Q, {(a, b), (c, d)}, {0, 1}).

(2) A polygonal quadrilateral is a quadrilateral with straight line intervals as


sides. Show that a polygonal quadrilateral mirror-symmetric relative to a
diagonal has modulus 1.
(3) Let (Q; a, b, c, d) be a polygonal quadrilateral with interior angle at a at most
π, and let a be a point on the side (a, b) such that the straight-line interval
(d, a ) lies in Q. Verify the inequality
M (Q; a, b, c, d) < M (Q ; a , b, c, d),
where (Q ; a , b, c, d) also denotes a polygonal quadrilateral.
(4) In the notation of Exercise (3) let Q be a convex domain and let a = a be
a point in the closed set bounded by the side (a, b) and the extensions of
the sides (d, a) and (b, c) over the vertices a and b, respectively to the point
of their intersection (which can be at infinity). Show that the inequality in
Exercise (3) also holds in this case.
1.3. Generalized condensers 21

(5) Verify that Theorem 1.17 yields the following ‘Grötzsch principle’ (cf. [Gol],
Ch. IV, § 6). Let Qi , i = 1, . . . , n, be pairwise disjoint quadrilaterals in an
annulus r < |z| < R, such that two opposite sides of each quadrilateral lie on
the circles |z| = r and |z| = R. Then the moduli M Qi of these quadrilateral
with respect to the distinguished sides satisfy

n
M Qi  2π/ log(R/r).
i=1

Describe all the equality cases in this relation.


(6) Show that Theorem 1.18 contains a result due to Grötzsch: for any pairwise
disjoint doubly connected domains Gi , i = 1, . . . , n, lying in an annulus
G = {z : r < |z| < R} and separating its boundary components,


n
1
M Gi  M G = log(R/r)
i=1

(see Exercise 1.2(5)).


(7) Let C = (B, {Ek }nk=1 , {δk }nk=1 ) and Ci = (Bi , {Eij }nj=1
i
, {δij }nj=1
i
), i = 1, . . .,
m, be condensers satisfying the following conditions:
Bi ∩ Bi = ∅, i = i , i, i = 1, . . . , m,

m
E1 ⊂ Ei1 ,
i=1
(∂Bi \ Ei1 ) ∩ B = ∅, δi1 = δ1 , i = 1, . . . , m,
for each k = 2, . . . , m there exists a plate Eij , 1  j  ni , 1  i  m, such
that Ek ⊂ Eij , δk = δij . Prove the inequality


m
cap C  cap Ci .
i=1

(8) Let C = (Cz , {Ek }nk=1 , ) be a condenser mirror-symmetric relative to the


 = (Cw , {E
circle |z| = 1, and let C k }n , ) be a condenser mirror-symmetric
k=1
relative to |w| = 1. Assume that there exists a univalent analytic function
w = f (z) in the disc U = {z : |z| < 1} such that |f (z)| < 1 for z ∈ U and
f (Ek ∩ U ) ⊂ E k , k = 1, . . . , n. Prove the inequality


cap C  cap C.
22 Chapter 1. Conformal Capacity

1.4 Geometric interpretation of capacity


Geometric definitions of the capacity of a condenser with two plates, including the
limiting case as one of the plates contracts to the point at infinity, are well known
in potential theory. The reader can find in the literature fairly complete proofs
that different approaches are equivalent, so we content ourselves with definitions
and the statements of the main results.
Assume that a condenser C = (E0 , E1 ) has a finite field and its plates E0
and E1 are infinite sets. For an arbitrary ordered quadruple z1 , z2 , z3 , z4 on the
Riemann sphere C, where z1 = z3 and z2 = z4 , we set

|z1 , z2 , z3 , z4 | = |ϕ(z1 )/ϕ(z3 )|,

where ϕ is a Möbius transformation of the sphere C such that ϕ(z2 ) = 0 and


ϕ(z4 ) = ∞. Let a1 , . . . , an be some points in the plate E0 and b1 , . . . , bn some
points in E1 . We set
 −1
n
Wn = inf log |ai , bj , bi , aj |. (1.6)
2
1i<jn

The sequence {Wn } is nondecreasing. Indeed, assume that some points a1 , . . .,


an ∈ E0 and b1 , . . . , bn ∈ E1 deliver the infimum in (1.6). Then for each fixed l,
1  l  n,    
n n−1
Wn  log |al , bj , bl , aj | + Wn−1 .
2 2
1jn
j=l

Summing these relations over l we obtain


     
n n n−1
n Wn  2 Wn + n Wn−1 ,
2 2 2

and therefore Wn  Wn−1 . We denote the (finite or infinite) limit of the sequence
{Wn } by mdC.
Theorem 1.21. The following equality holds:

cap C = .
mdC
The proof of Theorem 1.21 was given by Bagby [B]. The following important
special case of this result was considered by Tsuji. Let E be an infinite closed
subset of the unit disc |z| < 1 and let C = (E, {z : |z|  1}). We set
⎧ ⎫
⎨   2/[n(n−1)]
 ak − al ⎬
dn,h (E) = max   .
⎩ak ,al ∈E  1 − ak al ⎭
1k<ln
1.4. Geometric interpretation of capacity 23

As above, we can verify that {dn,h (E)}∞


n=2 is a nonincreasing sequence, so that
there exists a limit
lim dn,h (E) = dh (E),
n→∞

which is called the hyperbolic transfinite diameter of the set E. The following
theorem is a consequence of Tsuji’s results [Ts].
Theorem 1.22. The following equality holds:
 −1
1
cap C = − log dh (E) .

The quantity dh (E) is also called the hyperbolic capacity of E; it is denoted by

caph E = dh (E).

Now let E be a bounded infinite closed set in the plane C. If r > 0 is small
enough, then we can look at the condenser

C(r) = (U (∞, r), E).

By Theorem 1.14
1 r
|C(r )|  |C(r)| + log  for any r < r.
2π r
Hence the quantity
1
|C(r)| + log r

increases as r tends to zero and the limit
 
1
M := lim |C(r)| + log r
r→0 2π
exists. Thus, for M = ∞ we have the asymptotic formula
   2  2 
2π 2π 1
cap C(r) = − −M +o , r → 0.
log r log r log r

The quantity
cap E = exp{−2πM }
is called the logarithmic capacity of the set E. If M = ∞, then cap E = 0. This
capacity has the following geometric interpretation. Let
⎧ ⎫2/[n(n−1)]
⎨ ⎬
dn (E) = max |ak − al | .
⎩ak ,al ∈E ⎭
1k<ln
24 Chapter 1. Conformal Capacity

It can easily be shown that {dn (E)}∞


n=2 is a nonincreasing sequence, so the limit

lim dn (E) = d(E)


n→∞

which is called the transfinite diameter of E, exists.


Theorem 1.23. (Fekete–Szegő) The following equality holds:

cap E = d(E).

The proof of Theorem 1.23 can be found in monographs on function the-


ory and potential theory ([Gol], Ch. VII, § 3). Logarithmic capacity also has a
probabilistic interpretation (see, for instance, [Doob] or [L]).

Exercises 1.4
(1) (The Teichmüller problem.) Assume that the field of a condenser C = (E0 ,E1 )
is a doubly connected domain, −1, 0 ∈ E0 , and a, ∞ ∈ E1 . Show that

cap C  cap([−1, 0], {z : Im z = 0, Re z  |a|}).

(2) (The Grötzsch problem.) Assume that the field of a condenser C = (E0 , E1 )
is a doubly connected domain, E0 ⊃ {z : |z|  1}, and E1  a, ∞. Show that

cap C  cap({z : |z|  1}, {z : Im z = 0, Re z  |a|}).


Chapter 2

The Asymptotic Behaviour of the


Capacity as some Plates of the
Condenser Degenerate into Points

2.1 The Green, Robin, and Neumann functions


In this section we have collected some definitions and properties of the functions
mentioned in the title. For a domain B on the Riemann sphere C and a point
z0 ∈ B, z0 = ∞ assume that there exists a function gB (z, z0 ) which is continuous
in C, harmonic in B apart from z0 , vanishes outside B, and has the following
property: the function
gB (z, z0 ) + log |z − z0 |
is harmonic in a neighbourhood of z0 . Then gB (z, z0 ) is called the (classical) Green
function of the domain B with pole at z0 . For z0 = ∞ we define the function
gB (z, ∞) in a similar way, with the exception that gB (z, ∞) − log |z| must be
harmonic in a neighbourhood of the point at infinity. A domain B which carries
a classical Green function will be said to be admissible. It is known that if each
boundary point of B is the end-point of some Jordan arc lying in the exterior of
B, then B is an admissible domain [R]. For an arbitrary domain B in C its Green
function is defined as the limit

gB (z, z0 ) = lim gBn (z, z0 ), z ∈ C,


n→∞

where {Bn }∞n=1 is an exhaustion of B by domains with classical Green functions,




that is, Bn ⊂ Bn+1 , z0 ∈ B1 and Bn = B. We can show that the Green
n=1
function is independent of the exhaustion of B and is equal to the classical Green
function in the case when the latter exists in B. If the above limit is equal to infinity
at some (and then each) point z, we say that B has no Green function [St]. It is

© Springer Basel 2014 25


V.N. Dubinin, Condenser Capacities and Symmetrization in Geometric Function
Theory, DOI 10.1007/978-3-0348-0843-9_2
26 Chapter 2. The Asymptotic Behaviour of the Capacity

easy to see that the Green function gB (z, z0 ) increases with the growth of B; it is
positive in B \ {z0 } and has the property of symmetry

gB (z1 , z2 ) = gB (z2 , z1 ) for z1 , z2 ∈ B.

Furthermore, the Green function is invariant under conformal univalent map-


pings f :
gf (B) (f (z), f (z0 )) = gB (z, z0 ), z ∈ B \ {z0 }.
In many cases this enables us to find an analytic expression for the Green function.
Indeed, it is seen by direct verification that the Green function of the disc U =
{z : |z| < 1} with pole at the origin has the form

gU (z, 0) = − log |z|.

Hence the Green function of the same disc U with pole at an arbitrary point
z0 ∈ U is  
 1 − z0z 
gU (z, z0 ) = log  .
z − z0 
An expression for the Green function of a doubly connected domain looks much
more complicated. For instance, the Green function of the annulus K(R) = {z :
1 < |z| < R}, R = ∞, with pole at a positive point z0 has the following represen-
tation [Henr]:
 
log r  1 − z0 reiθ 
gK(R) (re , z0 ) = −
iθ 
log z0 + log  iθ 
log R re − z0 

R−n z0n − z0−n n
− n − R−n
(r − r−n ) cos nθ.
n=1
n R

For an arbitrary open set B in the plane C we shall mean by the Green
function gB (z, z0 ) the Green function of the connected component of B containing
the point z0 . If B has the Green function with pole at some point z0 ∈ B, then
it has the Green function with pole at an arbitrary point in the component of B
containing z0 [St].
The inner radius of an open set B with respect to a point z0 ∈ B is the
quantity

r(B, z0 ) = exp{ lim [gB (z, z0 ) + log |z − z0 |]} for z0 = ∞,


z→z0

r(B, ∞) = exp{ lim [gB (z, ∞) − log |z|]}.


z→∞

If B has no Green function, then we set r(B, z0 ) = +∞. It is easy to see that the
inner radius of a simply connected domain of hyperbolic type with respect to a
finite point coincides with the conformal radius of the domain [Gol]. Let E be a
2.1. The Green, Robin, and Neumann functions 27

closed bounded subset of C and B the connected component of the complement


C \ E which contains the point z = ∞. Then the quantity

γ = log r(B, ∞) = log r(C \ E, ∞)

is called the Robin constant of B [Gol], [St] and cap E = e−γ is called the loga-
rithmic capacity of E. It is known that the logarithmic capacity cap E is equal to
the Chebyshev constant τ (E) and to the transfinite diameter d(E) of E:

cap E = τ (E) = d(E) = 1/r(B, ∞) (2.1)

(see [Gol]). Since the Green function is a conformal invariant, if a function f maps
the domain B conformally and univalently onto a domain f (B), then

r(B, z0 )|f  (z0 )| = r(f (B), f (z0 )) (2.2)

for each z0 ∈ B. In Chapter 8 we discuss the behaviour of the inner radius under
maps which are not univalent.
Now we proceed to define the Robin function in the case when B ⊂ C is a
domain bounded by finitely many analytic Jordan curves. Let Γ be a nonempty
closed subset of ∂B consisting of finitely many nonsingular Jordan arcs and let z0
be a finite point in B \ Γ. Let gB (z, z0 , Γ) be a continuous real-valued function on
B \{z0 } which extends to a harmonic function in a neighbourhood of B \(Γ∪{z0 })
and satisfies the following conditions:

gB (z, z0 , Γ) = 0 for z ∈ Γ;

gB (z, z0 , Γ) = 0 for z ∈ ∂B \ (Γ ∪ {z0 });
∂n
gB (z, z0 , Γ) + log |z − z0 | is a bounded harmonic function in a neighbourhood of
z0 (∂/∂n denotes differentiation along the inward normal to ∂B).
If z0 = ∞, then gB (z, ∞, Γ) is defined in a similar way, with the only ex-
ception that gB (z, ∞, Γ) − log |z| must be harmonic in a neighbourhood of the
point at infinity. If z0 ∈ B, then gB (z, z0 , Γ) is called the Robin function of the
domain B with pole at z0 [Dur]. We shall use this term for the functions defined
for z0 ∈ B \ Γ. If Γ = ∂B, then the Robin function is equal to the Green function
gB (z, z0 ) in B. It is easy to verify that under the above assumptions the function
gB (z, z0 , Γ) is uniquely defined.
Now let B be an arbitrary finitely connected domain in C without isolated
boundary points. By the boundary ∂B of B we shall mean the set of prime ends.
If there can be no confusion, we shall use the same letter for the support of an
attainable boundary point and the point itself (see the explanations in § 1.3); recall
that a boundary point ζ of a domain G is called an attainable boundary point if it
can be connected with some interior point of G by a continuous curve l : z = z(t),
a  t  b, so that the whole curve apart from the end-point ζ = z(b) lies in G. We
28 Chapter 2. The Asymptotic Behaviour of the Capacity

define a neighbourhood U (ζ, r, B) of an attainable boundary point as a connected


component of the intersection B ∩ U (ζ, r) (where U (ζ, r) := {z : |z − ζ| < r}) that
contains an open arc of the curve l with end-point at z(b). Two attainable boundary
points ζ1 and ζ2 are said to be different if they have disjoint neighbourhoods
U (ζk , r, B), k = 1, 2; note that the corresponding points in the plane (the supports
of ζ1 and ζ2 ) may coincide. We also say that a subset of the domain B adjoins an
attainable boundary point ζ if it contains a neighbourhood U (ζ, r, B) of this point.
A small neighbourhood of an interior point ζ ∈ B is an ordinary disc U (ζ, r) ⊂ B.
The notation z → ζ means a net of points in which the subsequent points belong
to the same neighbourhood of the point ζ as the previous ones.
Assume that a function f maps the domain B conformally and univalently
onto a bounded domain f (B) in C whose boundary consists of finitely many
analytic Jordan curves. A point z0 is called an admissible point with exponent
α, 0 < α  2 for B, if either z0 ∈ B or z0 is an attainable boundary point of B
and the following expansion holds in some neighbourhood of this point:

f (z) = f (z0 ) + (z − z0 )1/α (A + o(1)), z → z0 , (2.3)

where A is a nonzero constant. For the point z0 = ∞ we replace the local parameter
z − z0 in the expansion (2.3) by 1/z. If z0 ∈ B, z0 = ∞, then (2.3) holds a fortiori
with α = 1 and A = f  (z0 ). It is clear that the expansion (2.3) and the value of α
are independent of the choice of the function f . Thus, the fact that z0 is admissible
only depends on the domain B, but not on f . In particular, if the boundary of B
in a neighbourhood of z0 = ∞ is formed by two closed analytic arcs with common
end-point z0 , which form an interior angle απ > 0, then z0 is an admissible point
with exponent α by the Lichtenstein–Warschawski theorem ([Henr], p. 359). Let
B and f be as above, let Γ be a nonempty closed subset of the boundary ∂B
which consists of finitely many nonsingular connected components, and let z0 be
an admissible point with exponent α for B which does not belong to Γ. Then we
define the Robin function by

gB (z, z0 , Γ) = αgf (B) (f (z), f (z0 ), f (Γ)). (2.4)

This definition does not depend on the choice of f because a conformal univalent
map of a domain with real-analytic boundary onto a similar domain preserves the
property of being harmonic, boundary conditions, and the form of the expansion
in a neighbourhood of a pole. It is easy to see that similarly to the behaviour
of the Green function, gB (z, z0 , Γ) + log |z − z0 | is harmonic in a neighbourhood
of a finite admissible point z0 , and for z0 = ∞ the function gB (z, z0 , Γ) − log |z|
is harmonic. In some cases the Robin function can be expressed in terms of the
Green functions of appropriate domains. For example, if B = {z : Im z > 0} and
Γ = {z = x + iy : x  0, y = 0}, then thanks to symmetry relative to the real axis,
the Green function with pole at z0 < 0 of the domain C \ Γ is equal to the Robin
function gB (z, z0 , Γ).
2.1. The Green, Robin, and Neumann functions 29

It is also easy to verify that for γ = {z : |z| = 1}

gK(R) (z, z0 , γ) = gK(R2 ) (z, R2 /z 0 ) + gK(R2 ) (z, z0 ).

On the other hand, let G(z; R) be the Grötzsch function, which maps the annulus
K(R), R < ∞, conformally and univalently onto the exterior |w| > 1 of a disc
which is cut along the positive real half-axis, from some point P (R) to infinity, so
that G(R; R) = P (R). Then for positive z0
 
 1 − G(z; R)G(z0 ; R) 

gK(R) (z, z0 , γ) = log  .
G(z; R) − G(z0 ; R) 

By the Robin radius of a domain B with respect to a point z0 and a set Γ,


for z0 = ∞ we mean the quantity

r(B, Γ, z0 ) = exp{ lim [gB (z, z0 , Γ) + log |z − z0 |]};


z→z0

we also set
r(B, Γ, ∞) = exp{ lim [gB (z, ∞, Γ) − log |z|]}.
z→∞

If z0 ∈ B and Γ = ∂B, then r(B, Γ, z0 ) = r(B, z0 ). In the case when B is simply


connected, Γ is a boundary arc, and z0 ∈ (∂B)\Γ the quantity r(B, Γ, z0 ) occurs in
the literature under various names and denoted in various ways. In the general case
this quantity is not very well understood. If f is a conformal univalent mapping
of the domain B onto a domain with smooth boundary, which has the expansion
(2.3) in a neighbourhood of an admissible point z0 , then

r(B, Γ, z0 )|A|α = rα (f (B), f (Γ), f (z0 )). (2.5)

As previously, calculating Robin radii can be reduced to finding the inner radii or
the Green functions, similarly to how it is done with Robin functions. Note also
that if B is a circular sector {z ∈ U : ϕ1 < arg z < ϕ2 } and Γ = {z : |z| = 1, ϕ1 
arg z  ϕ2 }, then gB (z, 0, Γ) = − log |z| so that

r(B, Γ, 0) = r(B, 0)

for whatever ϕ1 and ϕ2 can be. If we replace here the disc with a regular polygon
with centre at the origin, the circular sector with sector cut out from the polygon
by the rays arg z = ϕk , k = 1, 2, and Γ with the corresponding piece of the
boundary of the polygon, then the above equality holds if and only is the rays in
question pass through vertices of the polygon.
Several ways to define the Neumann function can be found in the literature
(see, for example, [Henr] and [Dur]). We require a new definition of this function,
which covers the case when its poles are attainable boundary points. Again, let
B ⊂ C be a domain bounded by finitely many analytic Jordan curves and let z0
and z ∗ be different points in B. By the Neumann function of B with poles at z0
30 Chapter 2. The Asymptotic Behaviour of the Capacity

and z ∗ we shall mean the real-valued function n(z) = nB (z, z0 , z ∗ ) satisfying the
following conditions:
n(z) is harmonic in B \ {z0 , z ∗ };
∂n(z)
= 0 for z ∈ (∂B) \ {z0 , z ∗ };
∂n
n(z) + log |z − z0 | is a bounded harmonic function in a neighbourhood of the
point z0 ;
in a neighbourhood of z ∗
σ(z0 )
n(z) = log |z − z ∗ | + o(1), z → z ∗ , z ∈ B,
σ(z ∗ )
where σ(z) = 1 for z ∈ ∂B and σ(z) = 2 for z ∈ B; if z0 = ∞ (z ∗ = ∞), then
the local parameter z − z0 (z − z ∗ ) must be replaced by 1/z. That such a function
exists and is unique easily follows from well-known results in potential theory (see,
for instance, [Henr], p. 271). If B = C (∂B = ∅) then we set
nC (z, z0 , z ∗ ) = log |(z − z ∗ )(z ∗ − z0 )/(z − z0 )|
for z ∗ = ∞ and
nC (z, z0 , ∞) = − log |z − z0 |.
Now let B be an arbitrary finitely connected domain in C without isolated bound-
ary points and let f be a function mapping B conformally and univalently onto a
domain f (B) bounded by finitely many analytic curves. Let z0 be an admissible
point in B with exponent α and z ∗ be another admissible point in B with exponent
α∗ . Then we define the Neumann function of B by
σ(z0 ) ∗
nB (z, z0 , z ∗ ) = αnf (B) (f (z), f (z0 ), f (z ∗ )) − ∗
log |A∗ |α , (2.6)
σ(z )
where A∗ is the constant in the expansion (2.3) of f in a neighbourhood of z ∗ ,
σ(z) is the exponent of the admissible point z if z ∈ ∂B and σ(z) = 2 for z ∈ B.
It follows from (2.6) that if z0 is a finite point, then nB (z, z0 , z ∗ ) + log |z − z0 |
is a bounded harmonic function in a neighbourhood of z0 , while (assuming now
that z ∗ = ∞) the function nB (z, z0 , z ∗ ) − (σ(z0 )/σ(z ∗ )) log |z − z ∗ | is harmonic
in a neighbourhood of z ∗ . The Neumann function of a half-plane can easily be
constructed using symmetry arguments. Indeed, let B = {z : Im z > 0}, z0 ∈
B, z ∗ ∈ B, z0 = z ∗ . The function
 
 (z − z ∗ )(z − z ∗ )(z ∗ − z0 )(z ∗ − z 0 ) 
log  

(z − z0 )(z − z 0 )2Im z ∗
is mirror-symmetric relative to the real axis (= ∂B), hence it coincides with the
Neumann function nB (z, z0 , z ∗ ). Similarly, for z0 ∈ ∂B we obtain
|(z − z ∗ )(z − z ∗ )|1/2 |z ∗ − z0 |
nB (z, z0 , z ∗ ) = log .
|z − z0 ||2Im z ∗ |1/2
2.1. The Green, Robin, and Neumann functions 31

The Neumann function of a disc can be expressed in a similar way. If B is an


arbitrary simply connected domain, then we can use formula (2.6).
In conclusion we define the quantity N (B, z0 , z ∗ ), which is an analogue of
the Robin radius in certain respects, by setting
N (B, z0 , z ∗ ) = exp{ lim [nB (z, z0 , z ∗ ) + log |z − z0 |]} for z0 = ∞,
z→z0

and
N (B, ∞, z ∗ ) = exp{ lim [nB (z, ∞, z ∗ ) − log |z|]}.
z→∞

Exercises 2.1
(1) Find the Green function of the upper half-plane with pole at an arbitrary
point in this half-plane.
(2) Find the Green function of the domain {z : Im z > 0, |z| > 1/2} with pole at
z0 = i.
(3) Show that if a domain B1 lies in a domain B2 , then
r(B1 , z0 )  r(B2 , z0 ), z 0 ∈ B1 ,
and if B1 has the classical Green function, then r(B1 , z0 ) = r(B2 , z0 ) only in
the case when B1 = B2 .
(4) Prove the property of continuity: for each exhaustion {Bn }∞ n=1 of a domain
B and any z0 ∈ B1 ,
r(B, z0 ) = lim r(Bn , z0 ).
n→∞

(5) Show that the inner radius r(B, z) of a domain is a continuous function of
z as long as z is a finite point in B. Is the inner radius also continuous at
z = ∞?
(6) Prove Fekete’s theorem: for any closed bounded set E and each polynomial
P (z) = z n + · · · , 
d(P −1 (E)) = n d(E).
(7) Show that if a domain B has infinite inner radius with respect to some point,
then ∂B only consists of degenerate boundary components.
(8) Find the inner radius
a) of the disc |z| < R with respect to a point z0 , |z0 | < R,
b) of the strip |Im z| < l with respect to a point z0 , Im z0 = 0,
c) of the plane Cz cut along the n rays arg z n = θ, |z n |  l, with respect
to the origin,
d) of a square with side a with respect to its centre of symmetry,
e) of the annulus with respect to an arbitrary point z0 in K(R).
(9) Prove that cap E = cap ∂E.
32 Chapter 2. The Asymptotic Behaviour of the Capacity

(10) Find the logarithmic capacity


a) of a line segment of length l,
b) of the disc |z|  R,
c) of the arc | arg z|  θ < π, |z| = 1,
d) of a square with side a,
e) of an ellipse with semiaxes a and b.
(11) Prove the equality
σ(z1 )gB (z1 , z2 , Γ) = σ(z2 )gB (z2 , z1 , Γ).
Consider the following cases separately: both points lie in B; both points lie
on the boundary of B; one point lies in B and the other lies on the boundary.
(12) Assuming that the Robin function gB (z, z0 , Γ) exists, let zk , k = 1, . . . , n,
be different points in B \ Γ which are distinct from z0 , and let B(ρ) =

n
B\ U (zk , ρ). Show that
k=1

gB(ρ) (z, z0 , Γ) ⇒ gB (z, z0 , Γ)



n
on compact subsets of B \ {zk }.
k=1
(13) Let E be a set formed by finitely many closed subarcs of the unit circle
|z| = 1. Prove the formula
r(U, E, 0) = (cap E)−2 .
(14) (Robinson [Rob]) Show that if the set E in Exercise (13) is mirror-symmetric
relative to the real axis, then

cap E = 2 cap F ,
where F is the projection of E onto this axis.
(15) Verify that the Neumann function of the exterior of the unit disc has the
following form:
 
 (z − z ∗ )(1 − z ∗ z)(z ∗ − z0 )(1 − z 0 z ∗ ) 

log   , |z0 | > 1, |z ∗ | > 1, z0 , z ∗ = ∞.
(z − z0 )(1 − z 0 z)(1 − |z ∗ |2 ) 
What is the form of the Neumann function for z ∗ = ∞? for |z0 | = 1?
(16) Find N (B, −1, 1) for B = {z : |Im z| < π/2}.
(17) Let B be an open set containing the origin and let Bn = {z : z n ∈ B}. Prove
the equality
r(B, 0) = (r(Bn , 0))n .
(18) Verify that the quantity
r(B, Γ1 , z)/r(B, Γ2 , z), z ∈ B (where Γ1 = ∅, Γ2 = ∅),
is a conformal invariant.
2.2. The case when one plate is nondegenerate 33

2.2 The case when one plate is nondegenerate


We say that a parametric family of domains D(z0 , r), 0 < r < r0 , is asymptotically
circular if the inclusions

U (z0 , s(r)) ⊂ D(z0 , r) ⊂ U (z0 , t(r))

hold for some positive functions s(r) and t(r), 0 < r < r0 , such that s(r) ∼ t(r) ∼
r as r → 0. An element D(z0 , r) of an asymptotically circular family is called an
almost disc with centre z0 and radius r, and we shall call the intersection of an
almost disc with a half-plane whose boundary contains z0 an almost half-disc. Let
B = C be an open subset of C, Z = {zk }nk=1 a set of distinct points in B, and =
{δk }nk=0 a tuple of real numbers such that δ0 = 0. Also let Ψ = {ψk }nk=1 , where
ψk = ψk (r) ≡ μk rνk and μk , νk , k = 1, . . . , n, are arbitrary positive numbers. Let
D(zk , ψk (r)), 0 < r < r0 , k = 1, . . . , n, be asymptotically circular families of
domains. For sufficiently small r > 0 the condenser

C(r; B, ∂B, Z, , Ψ) = (B, {Ek }nk=0 , ),

where E0 = ∂B, Ek = D(zk , ψk (r)), k = 1, . . . , n, is well defined. We consider


the asymptotic behaviour of the capacity of this condenser as r → 0, that is,
when one plate E0 is fixed, while the other plates Ek degenerate into points
zk , k = 1, . . . , n, respectively. We shall show that the capacity of the condensers
C(r; B, ∂B, Z, , Ψ) is
 2  2 
1 1 1
Q +R +o , r → 0, (2.7)
log r log r log r

with real constants Q and R independent of r. First of all we must verify that the
asymptotic expression (2.7) is independent of the choice of almost discs.
Lemma 2.1. Assume that the almost discs D(zk , ψk (r)) in the definition of the
condenser C(r; B, ∂B, Z, , Ψ) satisfy the inclusions

U (zk , sk (ψk (r))) ⊂ D(zk , ψk (r)) ⊂ U (zk , tk (ψk (r))), 0 < r < r0 , (2.8)

for some positive continuous functions sk (τ ) and tk (τ ), 0 < τ < τ0 , with sk (τ ) ∼


tk (τ ) ∼ τ as τ → 0, k = 1, . . . , n, and assume that the capacity of the condenser
C(r; B, ∂B, Z, , Ψ) has the asymptotic expression (2.7). Then for any other fami-
lies of almost discs D(zk , ψk (r)) the capacity of the condenser C(r; B, ∂B, Z, , Ψ)
has the representation (2.7) with the same constants Q and R.
 B, ∂B, Z, , Ψ) denote the condenser defined in a similar way
Proof. Let C(r;
to C(r; B, ∂B, Z, , Ψ) but with the almost discs D(zk , ψk (r)) replaced by their
neighbourhoods U (zk , ψk (r)), k = 1, . . . , n. It is easy to see that the functions

r∗ (r) = min ψk−1 (sk (ψk (r))) and r∗ (r) = max ψk−1 (tk (ψk (r)))
1kn 1kn
34 Chapter 2. The Asymptotic Behaviour of the Capacity

are continuous on some interval 0 < r < r0 and satisfy r∗ (r) ∼ r∗ (r) ∼ r as r → 0.
By continuity, for each sufficiently small r > 0 there exist r and r such that
r∗ (r ) = r = r∗ (r ). From (2.8) we obtain

D(zk , ψk (r )) ⊂ U (zk , ψk (r)) ⊂ D(zk , ψk (r )), k = 1, . . . , n.

By Theorem 1.15,
 B, ∂B, Z, , Ψ)  cap C(r ; B, ∂B, Z, , Ψ).
cap C(r ; B, ∂B, Z, , Ψ)  cap C(r;

Adding the terms


 2
1 1
−Q −R
log r log r
to all the parts of this inequality and bearing in mind that r ∼ r ∼ r , after
simple calculations we arrive at the formula
 2  2 
 B, ∂B, Z, , Ψ) = Q 1 1 1
cap C(r; + +R +o , r → 0.
log r log r log r
(2.9)
Now let D(zk , ψk (r)), k = 1, . . . , n, be arbitrary almost discs. Using (2.8)
again, we obtain

U (zk , ψk (r∗ (r))) ⊂ D(zk , ψk (r)) ⊂ U (zk , ψk (r∗ (r))), k = 1, . . . , n,

and by Theorem 1.15


 ∗ (r); B, ∂B, Z, , Ψ)  cap C(r; B, ∂B, Z, , Ψ)
cap C(r
 ∗ (r); B, ∂B, Z, , Ψ).
 cap C(r

Repeating the above argument, but taking now (2.9) into account we see that the
condenser C(r; B, ∂B, Z, , Ψ) has capacity (2.7), which completes the proof. 
Theorem 2.1. Let B be an open subset of C having the Green function and let
Z, , and Ψ be sets as above. Then
n  2
δk2 1
cap C(r; B, ∂B, Z, , Ψ) = − 2π + R(B, ∂B, Z, , Ψ)
νk log r log r
k=1
 2 
1
+o , r → 0, (2.10)
log r

where
 $
r(B, zk ) δk δl
n n n
δk2
R(B, ∂B, Z, , Ψ) = −2π log + gB (zk , zl ) .
νk2 μk νk νl
k=1 k=1 l=1
l=k
2.2. The case when one plate is nondegenerate 35

Proof. Assume that each connected component of B is bounded by finitely many


analytic curves and δk = 0, k = 1, . . . , n. We set gk (z) = gB (z, zk ) and Rk =
r(B, zk ), k = 1, . . . , n, and for each fixed r, 0 < r < r0 , we look at the auxiliary
function n n
g(z) = δk βlk gl (z), (2.11)
k=1 l=1
where 
−(log ψk )−1 (log ψl )−1 gk (zl ), l = k,
βlk = −1 −1
−(log ψk ) [1 + (log ψk ) log Rk ], l = k,
ψk = μk rνk , l, k = 1, . . . , n. The function g(z) vanishes in C\B, is continuous in B\
n n
{zk } and is harmonic in B \ {zk }. We set D(zk , ψk ) := {z ∈ U (zk , ψk (2r)) :
k=1 k=1
g(z)/δk > 1}, k = 1, . . . , n. For sufficiently small r the boundary of D(zk , ψk )
coincides with {z ∈ U (zk , ψk (2r)) : g(z) = δk }. For points z in this set we have

n
n
n
δk = δk βlk gl (z) + δj βlj gl (z)
l=1 j=1 l=1
j=k

= δk −(log ψk )−1 [1 + (log ψk )−1 log Rk ][− log |z − zk | + log Rk + o(1)]


n $
− (log ψk )−1 (log ψl )−1 gk (zl )[gl (zk ) + o(1)]
l=1
l=k

n 
+ δj −(log ψj )−1 (log ψk )−1 gj (zk )[− log |z − zk | + log Rk + o(1)]
j=1
j=k
$
− (log ψj )−1 gj (zk ) + o((log r)−1 ) , r → 0,

provided that zk = ∞. Hence


δk = (δk + o(1))(log ψk )−1 log |z − zk |, r → 0,
and
log |z − zk | ∼ log ψk , r → 0.
In view of the last relation, in the above formula we have
%
δk = δk (log ψk )−1 log |z − zk | + [(log ψk )−1 log Rk ](1 + o(1))
&
− (log ψk )−1 log Rk + o((log r)−1 )
n % −1
+ j=1 δj (log ψj ) gj (zk )(1 + o(1))
j=k
&
− (log ψj )−1 gj (zk ) + o((log r)−1 )
% &
= δk (log ψk )−1 log |z − zk | + o((log r)−1 ) .
36 Chapter 2. The Asymptotic Behaviour of the Capacity

Therefore,
|z − zk | ∼ ψk , r → 0.
Thus the family of domains D(zk , ψk (r)), 0 < r < r0 , is asymptotically circular
and its elements are almost discs with radii ψk and centres zk . In a similar way,
the family is asymptotically circular for zk = ∞. Let C(r; B, ∂B, Z, , Ψ) be a
condenser whose degenerating plates are equal to the sets D(zk , ψk ) introduced
above, k = 1, . . . , n. By Dirichlet’s principle (Theorem A1)
 
n
cap C(r; B, ∂B, Z, , Ψ) = I g, B \ D(zk , ψk ) .
k=1

Let ρ > 0 be sufficiently small so that the sets U (zk , ρ) lie in the D(zk , ψk ), k =
1, . . . , n, respectively. Using Green’s function and Gauss’ theorem we obtain in
succession
   

n n
∂g n
∂g
I g, B \ D(zk , ψk ) = − δk ds = − δk ds
∂n ∂n
k=1 k=1∂D(z ,ψ ) k=1 ∂U(zk ,ρ)
k k


n
n
= 2π δk δl βkl + o(1), ρ → 0.
k=1 l=1

Here we differentiate along the outward normals to the boundaries of D(zk , ψk )


and U (zk , ρ), k = 1, . . . , n. Substituting here the values of βkl and bearing in mind
that the capacity of the condenser is independent of ρ we obtain
 n
cap C(r; B, ∂B, Z, , Ψ) = 2π − δk2 (log ψk )−1 [1 + (log ψk )−1 log Rk ]
k=1

n
n $
− δk δl (log ψk )−1 (log ψl )−1 gl (zk ) . (2.12)
k=1 l=1
l=k

Equality (2.12) in combination with Lemma 2.1 yields the required formula (2.10).
Now let B be an open set with the (not necessarily classical) Green function.
We can assume that B consists of finitely many ( n) domains. Let {Bs }∞ s=1
be an exhaustion of B by sets Bs bounded by analytic curves. It is known that
gBs (z, zk ) ⇒ gk (z) as s → ∞ uniformly on compact subsets of B \ zk , so that
r(Bs , zk ) → Rk as s → ∞. Let g(z) and D(zk , ψk ), k = 1, . . . , n, be the function
and the domain defined above, and let g s (z) be the function constructed from
the gBs (z, zk ) similarly to g(z). In view of uniform convergence, using standard
arguments we find that
 

n
lim I g s , Bs \ D(zk , ψk ) = cap C(r; B, ∂B, Z, , Ψ).
s→∞
k=1
2.2. The case when one plate is nondegenerate 37

On the other hand, we see as before that the integral under the limit sign is
equal to

n 
n 
∂g s ∂g s
− (δk + αks (z)) ds = − δk ds + αs
∂n ∂n
k=1∂D(z ,ψ ) k=1 ∂U(zk ,ρ)
k k
 n
= 2π − δk2 (log ψk )−1 [1 + (log ψk )−1 log r(Bs , zk )]
k=1
n n $
− δk δl (log ψk )−1 (log ψl )−1 gBs (zk , zl ) + αs ,
k=1 l=1
l=k

where the αks (z) and αs are infinitesimal quantities as s → ∞. Passing to the
limit as s → ∞, we again obtain (2.12) and therefore also (2.10).
Now let δk = 0 for k = 1, . . . , m and δk = 0 for k = m + 1, . . . , n, 1  m 
n − 1. We set

n
B(ρ) = B \ U (zk , ψk (ρ)), Z  = {zk }m
k=1 ,  = {δk }m
k=0 , Ψ = {ψk }m
k=1 .
k=m+1

Since the capacity is monotone, for r  ρ we have

cap C(r; B, ∂B, Z  ,  , Ψ )  cap C(r; B, ∂B, Z, , Ψ)

 cap C(r; B(ρ), ∂B(ρ), Z  ,  , Ψ ).


Adding the same expression to all the parts of this inequality and taking account
of what we said above, we obtain
 2   2  2 
1 1 1
o  Φ(r; B, Z, , Ψ)  λ(ρ) +o , (2.13)
log r log r log r

where

Φ(r; B, Z, , Ψ)

n    2
δk2 1 1
= cap C(r; B, ∂B, Z, , Ψ) + 2π − R(B, ∂B, Z, , Ψ) ,
νk log r log r
k=1

λ(ρ) = R(B(ρ), ∂B(ρ), Z ,  , Ψ ) − R(B, ∂B, Z, , Ψ) → 0, as ρ → 0.




It follows from (2.13) that the limit

lim Φ(r; B, Z, , Ψ)(log r)2 = 0


r→0

exists. This means that formula (2.10) holds in the general case. 
38 Chapter 2. The Asymptotic Behaviour of the Capacity

 Now we look at the asymptotic behaviour


of the capacity of a generalized condenser in
{l ){l * B when one of its plates does not degenerate,
C while some other plates shrink to interior or
)* ? 4 boundary points of B.
Let B be a finitely connected domain
without isolated boundary points in the Rie-
)* ? 4 mann sphere C, Γ be a nonempty closed subset
of the boundary ∂B which consists of finitely
 many nondegenerate connected components,
)* ? 3 and Z = {zk }nk=1 be a set of different admis-
sible points in B with zk ∈ / Γ, k = 1, . . . , n.
Let = {δk }nk=0 be a tuple of real num-
Figure 2.1. bers with δ0 = 0, and let Ψ = {ψk }nk=1 ,
where ψk = ψk (r) ≡ μk rνk and μk , νk , k =
1, . . . , n, are arbitrary positive numbers. Let
D(zk , ψk (r)), 0 < r < r0 , k = 1, . . . , n, be asymptotically circular families of
domains. Then for sufficiently small r > 0 we have a well-defined condenser

C(r; B, Γ, Z, , Ψ) = (B, {Ek }nk=0 , ),

where this time we have E0 = Γ, Ek = D(zk , ψk (r)) for zk ∈ B and if zk is an


attainable boundary point of B, then Ek is set to be the closure in [B]3 of the
connected component of the intersection B ∩ D(zk , ψk (r)) which adjoins the point
zk , k = 1, . . . , n. It can easily be seen that the whole proof of Lemma 2.1 carries
over to condensers of the form C(r; B, Γ, Z, , Ψ), so we can replace the condenser
C(r; B, ∂B, Z, , Ψ) by C(r; B, Γ, Z, , Ψ) in the statement of the lemma.
Theorem 2.2. Let B be a domain, let Γ ⊂ ∂B, and let Z, , and Ψ be sets as
above. Then

cap C(r; B, Γ, Z, , Ψ)
 2  
n σ(zk )δ 2 1 1 2
= −π k
+ R(B, Γ, Z, , Ψ) +o , r → 0,
k=1 νk log r log r log r
where

R(B, Γ, Z, , Ψ)
 $
n σ(zk )δk2 r(B, Γ, zk ) n n σ(zk )δk δl
= −π log + g (z ,
B k l z , Γ) ,
k=1 νk2 μk k=1 l=1
l=k νk νl

σ(zk ) is the exponent of the admissible point zk if zk ∈ ∂B and σ(zk ) = 2 if zk


lies in B, k = 1, . . . , n.
3 see § 1.3. If the intersection B ∩ D(z , ψ (r)) is connected and bounded by a Jordan curve, then
k k
this is just the usual closure in C.
2.2. The case when one plate is nondegenerate 39

Proof. As the capacity is a conformal invariant, in view of (2.3), (2.4), and (2.5),
it is sufficient to consider the case when B is bounded by finitely many analytic
Jordan curves, all points in Z are finite, and for each boundary point zk in Z
there exists a disc U (zk , rk ) such that the intersection U (zk , rk ) ∩ ∂B is a line
interval. In this case, following the proof of Theorem 2.1, we introduce the aux-
iliary function g(z) by formula (2.11), where δk = 0, gk (z) = gB (z, zk , Γ), and

n
Rk = r(B, Γ, zk ), k = 1, . . . , n. The function g(z) is continuous on B \ {zk },
k=1

n
harmonic in B \ (Γ ∪ {zk }), and satisfies the following conditions: g(z) = 0
k=1

n
on Γ and ∂g/∂n = 0 for z ∈ (∂B) \ (Γ ∪ {zk }). We set D(zk , ψk ) := {z ∈
k=1
B ∩ U (zk , ψk (2r)) : g(z)/δk > 1}, k = 1, . . . , n. Repeating the corresponding
part of the proof of Theorem 2.1 we see that for zk ∈ B the family of domains
D(zk , ψk (r)), 0 < r < r0 , is asymptotically circular, with elements that are al-
most discs with centres zk and radii ψk , and if zk ∈ ∂B, then for sufficiently
small r each set D(zk , ψk (r)) is an almost half-disc. Let C(r; B, Γ, Z, , Ψ) be a
condenser with degenerating plates equal to the sets D(zk , ψk ) introduced above
k = 1, . . . , n. Again, following the proof of Theorem 2.1, we obtain
  
n n
∂g
cap C(r; B, Γ, Z, , Ψ) = I g, B \ D(zk , ψk ) = − δk ds
∂n
k=1 k=1B∩∂D(z ,ψ )
k k


n  n n
∂g
=− δk ds = π σ(zk )δk δl βkl + o(1), ρ → 0.
∂n
k=1 B∩∂U(zk ,ρ) k=1 l=1

Hence

n
cap C(r; B, Γ, Z, , Ψ) = π − σ(zk )δk2 (log ψk )−1 [1 + (log ψk )−1 log Rk ]
k=1
n
n $
− σ(zk )δk δl (log ψk )−1 (log ψl )−1 gl (zk ) ,
k=1 l=1,l=k

which yields the required formula for δk = 0, k = 1, . . . , n.


Now let δk = 0 for k = 1, . . . , m and δk = 0 for k = m + 1, . . . , n, where
1  m  n − 1. In addition to the notations B(ρ), Z  ,  , and Ψ used in the proof
of Theorem 2.1, we set Γρ := (∂B(ρ)) \ ∂B. By the monotonicity of the capacity

cap C(r; B, Γ, Z  ,  , Ψ )  cap C(r; B, Γ, Z, , Ψ)


 cap C(r; B(ρ), Γ ∪ Γρ , Z  ,  , Ψ )

for r  ρ. The end of the proof is essentially the same as in Theorem 2.1. 
40 Chapter 2. The Asymptotic Behaviour of the Capacity

Asymptotic formulae for the capacity of a condenser at least two plates of


which are not degenerating are consequences of the formulae above. This can be
seen in the following example.
Theorem 2.3. Let B be a finitely connected domain in C, {Ek }m k=1 be a system
of pairwise disjoint closed sets Ek ⊂ B, k = 1, . . . , m, and {δk }m
k=1 be a tuple

m
of real numbers. Assume that the boundary of Ek consists of finitely many
k=1

m
analytic curves and that the components of (∂B) \ Ek are also analytic curves.
k=1
Let u denote the potential function of the condenser (B, {Ek }m k=1 , {δk }k=1 ); for
m

m
k = m + 1, . . . , n let zk be points in B  = B \ Ek and let D(zk , ψk (r)), 0 <
k=1
r < r0 , be asymptotically circular families of simply connected Jordan domains
which are almost discs with centres zk and radii ψk (r) = μk rνk , where μk and νk ,
k = m + 1, . . . , n, are positive numbers. Then

cap(B, {Ek }nk=1 , {δk }nk=1 ) = cap(B, {Ek }m k=1 , {δk }k=1 )
m


n  n
(δk − u(zk ))2 (δk − u(zk ))2 r(B  , zk )
− 2π − 2π log
νk log r νk2 μk
k=m+1 k=m+1
n n $ 2
(δk − u(zk ))(δl − u(zl )) 1
+ gB  (zk , zl )
νk νl log r
k=m+1 l=m+1
l=k
 2 
1
+o , r → 0,
log r

where Ek = D(zk , ψk (r)) and the δk are some real numbers, k = m + 1, . . . , n.


Proof. Let us introduce some notation:
ω is the potential function of the condenser (B, {Ek }nk=1 , {δk }nk=1 ),
g is the potential function of the condenser (B  , {Ek }n+1  n+1
k=m+1 , {δk }k=m+1 ),
where En+1 = ∂B  , δk = δk − u(zk ), k = m + 1, . . . , n, δn+1 = 0;
n
we also set B  = B \ Ek .
k=1
From Green’s formula we obtain

∂ω
I(ω − g − u, B  ) = I(ω, B  ) + I(g + u, B  ) + 2 (g + u) ds
∂n
∂B 

   ∂ω
= I(ω, B ) + I(g, B ) + I(u, B ) + 2 ω ds + O(r)
∂n
∂B 
  
= −I(ω, B ) + I(g, B ) + I(u, B ) + O(r).
2.3. Condensers all of whose plates are degenerating 41

On the other hand



 ∂(ω − g − u)
I(ω − g − u, B ) = − (ω − g − u) ds = O(r).
∂B  ∂n
Thus
I(ω, B  ) = I(u, B  ) + I(g, B  ) + O(r).
It remains to use Theorem 2.1. 

Exercises 2.2
(1) Let B be a bounded domain in the plane C containing the origin and let
E(r), 0 < r < r0 , be a parametric family of continua such that 0 ∈ E(r) ⊂
B, 0 < r < r0 , which have logarithmic capacity cap E(r) = r. Find an
asymptotic formula for the capacity of the condenser (E(r), C \ B) as r → 0.
(2) How (2.10) will change if we replace the almost discs in the definition of the
condenser C(r; B, ∂B, Z, , Ψ) D(zk , ψk (r)), 0 < r < r0 , with the families
of line intervals [zk , zk + eiϕk ψk (r)], 0 < r < r0 , where the ϕk are some real
numbers, k = 1, . . . , n?
(3) Verify that the sets D(zk , ψk (r)) in the proof of Theorem 2.2 are either almost
discs (if zk ∈ B) or almost half-discs (if zk ∈ ∂B), k = 1, . . . , n.
(4) Give a detailed proof of Theorem 2.2 in the case when some of the δk vanish.

2.3 Condensers all of whose plates are degenerating


Let B be a finitely connected domain without isolated boundary points in C,
Z = {zk }nk=1 a set of different admissible points in B; = {δk }nk=1 a tuple of
real numbers, and Ψ = {ψk }nk=1 , where ψk = ψk (r) ≡ μk rνk and μk and νk , k =
1, . . . , n, are some positive numbers. Let D(zk , ψk (r)), 0 < r < r0 , k = 1, . . . , n,
be asymptotically circular families of domains. For small r > 0 we look at the
condenser
C(r; B, ∅, Z, , Ψ) := (B, {Ek }nk=1 , ),
where Ek = D(zk , ψk (r)) for zk ∈ B, while if zk is an attainable boundary point
of B, then the plate Ek is the closure in [B] of the connected component of
the intersection B ∩ D(zk , ψk (r)) which adjoins zk , k = 1, . . . , n. Similarly to
Lemma 2.1 we can prove the following result.
Lemma 2.2. Assume that some of the almost discs D(zk , ψk (r)) in the definition of
the condenser C(r; B, ∅, Z, , Ψ) satisfy the inclusions (2.8) for some continuous
positive functions sk (τ ) and tk (τ ), 0 < τ < τ0 , such that sk (τ ) ∼ tk (τ ) ∼ τ as
τ → 0, k = 1, . . . , n, and that the capacity C(r; B, ∅, Z, , Ψ) displays the asymp-
totic behaviour (2.7). Then for any other families of almost discs D(zk , ψk (r)) the
capacities of the condensers C(r; B, ∅, Z, , Ψ) have the representation (2.7) with
the same constants Q and R.
42 Chapter 2. The Asymptotic Behaviour of the Capacity

Theorem 2.4. Let B, Z, Ψ and = {δk }nk=1 , n 2, be the domainand sets defined
n 
n
above, but suppose that B = C and δn = δ := σ(zk )δk /νk / σ(zk )/νk .
k=1 k=1
Then the following asymptotic formula holds:

cap C(r; B, ∅, Z, , Ψ)
 2  2 

n
σ(zk )(δk − δ)2 1 1
= −π + R(B, ∅, Z, , Ψ) +o , r → 0,
νk log r log r log r
k=1

where

n
σ(zk )(δk − δ)2 N (B, zk , zn )
R(B, ∅, Z, , Ψ) = −π log
νk2 μk
k=1
n−1
n−1 $
σ(zk )(δk − δ)(δl − δ)
+ nB (zk , zl , zn ) ,
νk νl
k=1 l=1
l=k

N (B, zn , zn ) := 1.
Proof. Since the capacity C(r; B, ∅, Z, , Ψ) does not change after the replace-
ment of δk by δk − δ, k = 1, . . . , n, we can set δ = 0. As the capacity is a
conformal invariant, in view of (2.3) and (2.6), it can easily be verified that
it is sufficient to consider the case when B is a domain bounded by finitely
many analytic Jordan curves, Z consists of finite points, zn = 0, and for each
boundary point zk in Z there exists a disc U (zk , rk ) such that the intersection
U (zk , rk ) ∩ ∂B is a straight line interval. Now we follow the proof of Theorem 2.1
and introduce the auxiliary function g(z) defined by (2.11), where δk = 0, gk (z) =
nB (z, zk , 0), Rk = N (B, zk , 0), k = 1, . . . , n − 1, and the sum is taken from 1 to

n 
n
n − 1. Then g(z) is harmonic in B \ {zk } and ∂g/∂n = 0 on (∂B) \ {zk }. We
% k=1 & k=1
set D(zk , ψk ) := z ∈ B ∩ U (zk , ψk (2r)) : g(z)/δk > 1 , k = 1, . . . , n. Since in a
neighbourhood of the poles zk , k = 1, . . . , n−1, the functions gk (z) have expansion
of the same form as the corresponding expansions of the analogous functions in
the proof of Theorem 2.1, repeating the arguments in that proof we see that for
zk ∈ B the family of domains D(zk , ψk (r)), 0 < r < r0 , consists of almost discs,
while if zk ∈ ∂B, then for small r each set D(zk , ψk (r)) is an almost half-disc. Now
we examine the structure of D(zn , ψn (r)) for small r. In a neighbourhood of z = 0


n−1
n−1  
σ(zl )
δn = δk βlk log |z| + o(1)
σ(zn )
k=1 l=1

n−1
σ(zk )δk
= − (log ψk )−1 [1 + (log ψk )−1 log Rk ](log |z| + o(1))
σ(zn )
k=1
2.3. Condensers all of whose plates are degenerating 43


n−1
n−1  
σ(zl )
− δk (log ψk )−1 (log ψl )−1 gk (zl ) log |z| + o(1)
σ(zn )
k=1 l=1
l=k
n−1
−σ(zk )δk (log ψk )−1  log Rk

−1
= (log ψn ) (log |z|) 1+
σ(zn ) (log ψn )−1 νk log r
k=1
  n−1 
1 n−1
−σ(zl )δk (log ψk )−1 −1
+o + gk (zl ) (log ψl )
log r σ(zn ) (log ψn )−1
k=1 l=1
l=k
 n−1
νn σ(zk )δk ν /ν
Rk μnk n
−1
= (log ψn ) (log |z|) δn − log
σ(zn ) log r νk2 μk
k=1
  
n−1
n−1 σ(zl )δk 1
+ gk (zl ) + o .
νk νl log r
k=1 l=1
l=k

The map z → az, a > 0, does not change the capacity of C(r; B, ∅, Z, , Ψ), while
other quantities are transformed as follows: Rk → Rk a1+σ(zk )/σ(zn ) , gk (zl ) →
σ(zk )
gk (zl ) + σ(z n)
log a, k, l = 1, . . . , n − 1, δk → δk , μk → μk a, νk → νk , σ(zk ) →
σ(zk ), k = 1, . . . , n. Hence after applying this map we obtain the same asymptotic
formula as in Theorem 2.4; furthermore, we can pick a so that the expression in
square brackets in the above formula vanishes:


n−1
σ(zk )δk Rk μnk
ν /νn n−1
n−1 σ(zl )δk
2 log + gk (zl ) = 0. (2.14)
νk μk νk νl
k=1 k=1 l=1
l=k

In this case we obtain


|z| ∼ ψn (r), r → 0.
Thus the sets D(zn , ψn (r)) are also almost (half-)discs. Let C(r; B, ∅, Z, , Ψ) be a
condenser whose degenerating plates are equal to the sets D(zk , ψk ), k = 1, . . . , n.
Repeating the above proofs and taking (2.14) into account we can calculate the
capacity of this condenser:


n−1
n−1
n−1
n−1
σ(zl )
cap C(r; B, ∅, Z, , Ψ) = π σ(zk )δk δl βkl − πσ(zn )δn δk βlk
σ(zn )
k=1 l=1 k=1 l=1

n−1
n−1
=π σ(zk )δk δl βkl
k=1 l=1
n−1
σ(zk )δk ' (
− πσ(zn )δn −(log ψk )−1 (1 + (log ψk )−1 log Rk )
σ(zn )
k=1
44 Chapter 2. The Asymptotic Behaviour of the Capacity



n−1
n−1
σ(zl )
− δk gk (zl )(log ψk )−1 (log ψl )−1
σ(zn )
k=1 l=1
l=k


n−1
σ(zk )δk2
= −π
νk log r
k=1
n−1  2
σ(zk )δ 2 Rk σ(zk )δk δl
n−1 n−1
1
−π k
log + gl (zk )
νk2 μk νk νl log r
k=1 k=1 l=1
l=k
  2
δn δn 1
+ πσ(zn )δn − + 2 (log μn )
νn log r νn log r
n−1  2 
1 σ(zk )δk Rk μnk n σ(zl )
ν /ν n−1 n−1
1
+ log + δk gk (zl )
σ(zn ) νk2 μk νk νl log r
k=1 k=1 l=1
l=k
 2 
1
+o
log r
 2  2 

n
σ(zk )δk2 1 1
= −π + R(B, ∅, Z, , Ψ) +o , r → 0.
νk log r log r log r
k=1

Using Lemma 2.2 we arrive at the formula in Theorem 2.4 for arbitrary sets
D(zk , ψk ) and numbers δk = 0, k = 1, . . . , n.
Now let δk = 0 for k = 1, . . . , m, and δk = 0 for k = m + 1, . . . , n, where
1  m  n − 2. If we construct the function g(z) as previously, but take the
sum from m + 1 to n − 1, then g(z) coincides with the potential function of
a condenser C(r) = C(r; B, ∅, {zk }nk=m+1 , {δk }nk=m+1 , {ψk }nk=m+1 ) with certain
specially chosen plates. If R > 0 is sufficiently small, then the sets U (zk , R), k =
1, . . . , n, are pairwise disjoint. We take the auxiliary functions
  
 z − zk 
hk (z) = Mk (r) log   / log R , z ∈ U (zk , R) \ U (zk , ψk ),
ψk  ψk

where Mk (r) = max{|g(z)| : z ∈ U (zk , R)}, k = 1, . . . , m. Note that Mk (r) =

O((log r)−1 ), r → 0. For small r the function



⎪ 
m

⎨ g(z), z∈/ U (zk , R),
k=1
h(z) =

⎪ max{min[g(z), hk (z)], −hk (z)}, z ∈ B ∩ U (zk , R) \ U (zk , ψk ),

0, z ∈ B ∩ U (zk , ψk ), 1  k  m

is admissible for the condenser C  (r) = (B, {Ek }nk=1 , ), where Ek = B∩U (zk , ψk )
for k = 1, . . . , m, and the plates with larger indices are as in the condenser C(r).
2.3. Condensers all of whose plates are degenerating 45

Then we have
   

n 
n

cap C(r)  cap C (r)  I h, B \ Ek I g, B \ Ek
k=1 k=m+1

m
+ I(hk , U (zk , R) \ U (zk , ψk )) = cap C(r) + o((log r)−2 ), r → 0.
k=1

In combination with what we have already proved and with Lemma 2.2 this yields
the required asymptotic formula. 
Assume that B ⊂ C is a domain bounded by analytic curves and the set
Z lies in B. Then the asymptotic formula in Theorem 2.4 can be expressed in a
symmetric form in terms of the classical Neumann function N (z, ζ) [C], [Henr].
Recall that by definition N (z, ζ) is the function with the following properties:
1) N (z, ζ) is harmonic in B apart from the point z = ζ;
2) the sum N (z, ζ) + log |z − ζ| is also harmonic at z = ζ;
3) ∂N
) (z, ζ)/∂n = 2π/L on the boundary ∂B, where L is the total length of ∂B;
4) N (z, ζ)dsz = 0.
∂B
We set N (zk , zk ) = lim (N (z, zk ) + log |z − zk |), k = 1, . . . , n. Then
z→zk

nB (z, zk , zn ) = N (z, zk ) − N (z, zn ) − N (zn , zk ) + N (zn , zn ),


log N (B, zk , zn ) = N (zk , zk ) − N (zk , zn ) − N (zn , zk ) + N (zn , zn ).
Setting δ = 0 for simplicity, after straightforward calculations we obtain a sym-
metric expression:
 n 
δ2 1
n n
δk δl
R(B, ∅, Z, , Ψ) = −2π k
log + N (zk , zl ) .
νk2 μk νk νl
k=1 k=1 l=1

In particular, this suggests that the condition δn = δ in Theorem 2.4 may be


excessive.
Theorem 2.5. Let Z = {zk }nk=1 be a set of distinct points on the sphere C, =
{δk }nk=1 be a tuple of real numbers, and let Ψ = {ψk }nk=1 , where ψk = ψk (r) ≡
μk rνk and μk and νk , k = 1, . . . , n, are positive numbers. Then
cap C(r; C, ∅, Z, , Ψ) (2.15)
 2  2 

n
(δk − δ) 2
1 1
= −2π + R(C, ∅, Z, , Ψ) +o , r → 0,
νk log r log r log r
k=1

where  *

n
n
δ= δk /νk 1/νk ,
k=1 k=1
46 Chapter 2. The Asymptotic Behaviour of the Capacity

and
R(C, ∅, Z, , Ψ)
 n 
(δk − δ)2
n
n
 (δk − δ)(δl − δ)
= 2π log μk + log |zk − zl | .
νk2 νk νl
k=1 k=1 l=1
l=k

A prime on a summation sign means that if zk = ∞ (zl = ∞), then the corre-
sponding term is set equal to zero.
Proof. We can assume that δ = 0 and δn = 0. Now if zk = ∞, k = 1, . . . , n,
then we repeat the proof of Theorem 2.4 replacing the functions nB (z, zk , 0) with
nC (z, zk , 0) = log |zzk /(z − zk )|, k = 1, . . . , n − 1 (zn = 0), which gives us
(2.15). In the case when one of the points (for instance, zm ) is at infinity, we
perform a translation z → z + c if necessary to make the other points distinct
from zero. The map z → 1/z takes the condenser C(r; C, ∅, Z, , Ψ) to a condenser
C(r; C, ∅, Z  , , Ψ ) with the same capacity, whose plates are the closures of almost
discs with centres at the finite points zk = 1/zk and with radii μm = μm and
μk = μk /|zk2 | for k = 1, . . . , n, k = m, and whose parameters are Z  = {zk }nk=1
and Ψ = {μk rνk }nk=1 . The capacity of this condenser satisfies (2.15). After simple
calculations we see that
R(C, ∅, Z  , , Ψ ) = R(C, ∅, Z, , Ψ). 

Exercises 2.3
(1) Prove Lemma 2.2.
(2) Write out formulae showing how a linear fractional transformation of the
domain B changes the function nB (z, z0 , z ∗ ) and quantity N (B, z0 , z ∗ ).
(3) Show that if B is an open set whose complement has harmonic measure zero,
then
cap C(r; B, ∅, Z, , Ψ) = cap C(r; C, ∅, Z, , Ψ).
(4) Does Theorem 2.4 hold for δn = δ?
(5) Let B be a simply connected domain on the Riemann sphere C with n differ-
ent marked admissible boundary points zk with exponents αk , k = 1, . . . , n,
n  2; let Z = {zk }nk=1 , let = {δk }nk=1 be a set of real numbers not all of
which are equal, and let Ψ = {μk rνk }nk=1 , where the μk and νk are positive
numbers, k = 1, . . . , n. Take a condenser C(r; B, ∅, Z, , Ψ) as in the begin-
ning of § 2.3 and consider a conformal univalent mapping f of B onto the
upper half-plane Im w > 0. Show that
cap C(r; B, ∅, Z, , Ψ)
 2  2 

n
αk (δk − δ)2 1 1
= −π +R +o , r → 0,
νk log r log r log r
k=1
2.4. Reduced moduli 47

 

n 
n
where δ = αk δk /νk /( αk /νk ),
k=1 k=1


n
α2k (δk − δ)2 1/α
R=π log(μk k |Ak |)
νk2
k=1

n n
 αk αl (δk − δ)(δl − δ)
+ log |f (zk ) − f (zl )| ,
νk νl
k=1 l=1
l=k

and the constants Ak can be determined from the conditions

f (z) − f (zk ) = (z − zk )1/αk (Ak + o(1)), z → zk , k = 1, . . . , n;

for zk = ∞ (or f (zk ) = ∞) the local parameter z − zk (f (z) − f (zk ), respec-


tively) must be replaced by 1/z (by 1/f (z)).

2.4 Reduced moduli


Below we present the concept of reduced modulus, which is a natural generalization
of the classical reduced modulus of a simply connected domain B with respect to
one of its points [Tei] to the case of an arbitrary open set and a system of points
zk , k = 1, . . . , n, n  2. For greater transparency we consider the cases when all
of these points are interior and when some of them lie on the boundary separately.
Let B = C be an open subset of C, Z = {zk }nk=1 be a set of different points
n
in B, = {δk }nk=0 be a tuple of real numbers, δ0 = 0, and δk2 = 0; and
k=1
let Ψ = {ψk (r)}nk=1 , where ψk (r) = μk rνk and μk and uk are positive numbers,
k = 1, . . . , n, n  1. We set
 −1

n
ν= δk2 /νk .
k=1

We look at asymptotically circular families of domains D(zk , ψk (r)), 0 < r <


r0 , k = 1, . . . , n, and the corresponding condenser C(r; B, ∂B, Z, , Ψ) (see § 2.2).
By definition, the modulus of this condenser is the reciprocal value of the capacity:

|C(r; B, ∂B, Z, , Ψ)| = (cap C(r; B, ∂B, Z, , Ψ))−1 .

By the reduced modulus of B with respect to its boundary ∂B and the sets Z, , Ψ
we shall mean the limit
+ ν ,
M (B, ∂B, Z, , Ψ) = lim |C(r; B, ∂B, Z, , Ψ)| + log r
r→0 2π
48 Chapter 2. The Asymptotic Behaviour of the Capacity

if it exists and is finite. For δk = 1, k = 1, . . . , n, it follows from Theorem 1.14


that for all sufficiently small r < r
 n
−1

r −1
|C(r ; B, ∂B, Z, , Ψ)|  |C(r; B, ∂B, Z, , Ψ)| + 2π νk log  ,
r
k=1

Hence the quantity in curly brackets increases as r approaches zero, so it has a


limit, which can be finite or equal to +∞. In the last case, as well as for B = C,
we set M (B, ∂B, Z, , Ψ) = +∞ by definition. It follows from Theorem 2.1 that
if B has the Green function, then the reduced modulus exists, is independent of
the choice of the almost discs D(zk , ψk (r)), k = 1, . . . , n, and is equal to
 n 
ν 2 δk2 r(B, zk ) δk δl
n n
M (B, ∂B, Z, , Ψ) = log + gB (zk , zl ) . (2.16)
2π νk2 μk νk νl
k=1 k=1 l=1
l=k

In the special case when B is a simply connected domain of hyperbolic type in C


which contains a point z0 , we obtain the classical definition of the reduced modulus
(see [J], § 2.6):
1
M (B, ∂B, {z0 }, {0, 1}, {r}) = log r(B, z0 ).

Let B = C, let Z, Ψ be the sets defined above, and let = {δk }nk=1 be a
system of real numbers not all of which are equal. Then the reduced modulus of
the Riemann sphere with respect to the sets Z, , Ψ is defined by analogy:
+ ν ,
M (Z, , Ψ) := M (C, ∅, Z, , Ψ) = lim |C(r; C, ∅, Z, , Ψ)| + log r .
r→0 2π

n
If we also have δk /νk = 0, then
k=1
 n 
ν2 δ2
n
n
 δk δl
M (Z, , Ψ) = − k
log μk − log |zk − zl | (2.17)
2π νk2 νk νl
k=1 k=1 l=1
l=k

by Theorem 2.5. Throughout, unless otherwise stated, for a fixed set = {δk }nk=0
with δ0 = 0 we set = {δk }nk=1 ; the other way round we can recover the system
= {δk }nk=0 with δ0 = 0 from = {δk }nk=1 .
From properties of generalized condensers we can deduce the corresponding
properties of reduced moduli. For example, Theorem 1.16 yields

M (B1 , ∂B1 , Z, , Ψ)  M (B2 , ∂B2 , Z, , Ψ)  M (Z, , Ψ) (2.18)

for all B1 ⊂ B2 ⊂ C, Z, and Ψ.


2.4. Reduced moduli 49

Theorem 2.6. If all Z lies in the intersection of open sets B1 and B2 , then for any
Ψ and = {0, 1, . . . , 1}

M (B1 ∩ B2 , ∂(B1 ∩ B2 ), Z, , Ψ) + M (B1 ∪ B2 , ∂(B1 ∪ B2 ), Z, , Ψ)


 M (B1 , ∂B1 , Z, , Ψ) + M (B2 , ∂B2 , Z, , Ψ).

Proof. We set C(r; B) = C(r; B, ∂B, Z, , Ψ). By Theorem 1.10

cap C(r; B1 ∩ B2 ) + cap(r; B1 ∪ B2 )  cap(r; B1 ) + cap(r; B2 ).

Furthermore, since the capacity is monotonic, it follows that

cap C(r; B1 ∪ B2 )  cap C(r; Bk )  cap C(r; B1 ∩ B2 ), k = 1, 2.

Making elementary transformations, from these inequalities we obtain

|C(r; B1 ∩ B2 )| + |C(r; B1 ∪ B2 )|  |C(r; B1 )| + |C(r; B2 )|.

This yields the inequality in Theorem 2.6. 


Note that if B  ⊂ B and B\B  consists of the connected components of B
not containing points in Z, then for any and Ψ

M (B  , ∂B  , Z, , Ψ) = M (B, ∂B, Z, , Ψ).

We say that an open set B is regular with respect to a point set Z if each
connected component of B contains at least one point of Z. We also say that B
is admissible if each of its connected components has the classical Green function.
Now we define a decomposition of an open set similar to a decomposition of a
condenser, which was defined before Theorem 1.14. Let B be an admissible regular
set with respect to Z = {zl }m l=1 and A = {αl }l=1 be a system of positive numbers.
m

We also set

m
g(z) = αl gB (z, zl ), B(t) = {z : g(z) > t}, B  = B(tn+1 ),
l=1

Ck = B(tn+2−k )), C\B(tn+1−k ) , k = 1, . . . , n,

where 0 < t1 < t2 < · · · < tn+1 < ∞ and g(zl ) := +∞, l = 1, . . . , m. Then the
ordered tuple of sets {B  , C1 , . . . , Cn } is called the decomposition of B with respect
to Z and A which corresponds to the partition {tk }n+1 k=1 .

Theorem 2.7. If the point set Z = {zl }m l=1 , the open sets B and B , and the
condensers Ck = {Gk , Ek0 , Ek1 }, k = 1, . . . , n, satisfy the following conditions:
(i) zl ∈ B  ⊂ B, l = 1, . . . , m, Gk ⊂ B\B  , k = 1, . . . , n,
(ii) Gk ∩ Gl = ∅, k =  l, k, l = 1, . . . , n,
n 
n
(iii) zl ∈ Ek0 , l = 1, . . . , m, C\B ⊂ Ek1 ,
k=1 k=1
50 Chapter 2. The Asymptotic Behaviour of the Capacity

then for = {0, 1, . . . , 1} and each system of functions Ψ = {μl rνl }m


l=1 ,


n
M (B, ∂B, Z, , Ψ)  M (B  , ∂B  , Z, , Ψ) + |Ck |. (2.19)
k=1

In addition, if the condensers Ck , k = 1, . . . , n, are admissible and regular and the


sets B and B  are admissible and regular with respect to Z, then equality holds in
(2.19) for a fixed set of functions Ψ = {μl rνl }m l=1 if and only if the system of sets
{B  , C1 , . . . , Cn } coincides with the decomposition of B with respect to the points
in Z and numbers in A = {νl−1 }m l=1 up to the ordering.

Proof. Assume that the condenser C(r; B) is defined as in the proof of Theo-
rem 2.6, and let C(r; B  ) be the analogous condenser for the set B  . Let ν =
m 
 −1 −1
νl . In view of (i)–(iii), Theorem 1.14 yields
l=1

ν ν
n
|C(r; B)| + log r  |C(r; B  )| + log r + |Ck |.
2π 2π
k=1

Passing to the limit as r → 0 we obtain (2.19) (see Figure 1.1).


Now let {B  , C1 , . . . , Cn } be the decomposition of B with respect to Z and
the system of numbers A = {νl−1 }m l=1 which corresponds to some partition {tk }k=1
n+1

and let t > tn+1 . The potential function of the condenser (B(t), C\B) coincides
with 1 − g(z)/t in the field of the condenser. Hence the system of condensers
{(B(t), C\B  ), C1 , . . . , Cn } is a decomposition of the condenser (B(t), C\B). It is
easy to see that if t is large enough, then the part of the curve g(z) = t lying in a
 
neighbourhood of zl is an ‘almost circle’ with centre zl and radius μl rνl (1/(μl rνl )

for zl = ∞), where the μl , l = 1, . . . , m, are some positive numbers and r = e−t .
By Theorem 1.14

ν ν n
|(B(t), C\B)| + log r = |(B(t), C\B  )| + log r + |Ck |;
2π 2π
k=1

letting r → 0 (t → +∞), this yields

 
n
 
M (B, ∂B, Z, , {μl rνl }m
l=1 ) = M (B , ∂B , Z, , {μl r }l=1 ) +
νl m
|Ck |.
k=1

In combination with (2.16) this gives us equality in (2.19).


It remains to show that equality in (2.19) is only possible when {B  , C1 ,
. . . , Cn } is a decomposition of B with respect to Z and A.
2.4. Reduced moduli 51


m
We set g  (z) = νl−1 gB  (z, zl ), B  (t) = {z : g  (z) > t}, where t is sufficiently
l=1
large. If equality holds in (2.19), then it follows from what we have proved that

n
M (B, ∂B, Z, , Ψ) = M (B  , ∂B  , Z, , Ψ) + |Ck |
k=1

n
= M (B  (t), ∂B  (t), Z, , Ψ) + |(B  (t), C\B  )| + |Ck |
k=1

 M (B  (t), ∂B  (t), Z, , Ψ) + |(B  (t), C\B)|


 M (B, ∂B, Z, , Ψ).
This is only possible when

n
|(B  (t), C\B  )| + |Ck | = |(B  (t), C\B)|.
k=1

Again, Theorem 1.14 shows that the system of condensers


{(B  (t), C\B  ), C1 , . . . , Cn }
coincides with the decomposition of the condenser (B  (t), C\B) up to the ordering.
Let ω(z) be the potential function of this condenser and z  be an arbitrary bound-
ary point of B  . Then we have the identity ω(z)/ω(z ) ≡ 1 − g  (z)/t on B  \B  (t).

m
Hence, in particular, g  (z) extends to B\ {zl } as a harmonic function, which
l=1
we also denote by g  (z). From the uniqueness theorem for harmonic functions we
easily deduce

m
g  (z) + t(1/ω(z  ) − 1) ≡ g(z) := νl−1 gB (z, zl ).
l=1

Thus g(z) ≡ t((1 − ω(z))/ω(z )), z ∈ B\B (t), and {B  , C1 , . . . , Cn } is a decom-


 

position of B with respect to Z and A up to the ordering. 


In the theorems below we establish connections between the reduced moduli
of the Riemann sphere and its subsets.
Theorem 2.8. Let Bk , k = 1, . . . , n , be disjoint open subsets of the sphere C,
and zkl ∈ Bk be distinct points in these sets, l = 1, . . . , mk , let Zk = {zkl }m k
l=1 ,
k = 1, . . . , n. Assume that the sets k , k = {δkl }l=1 , and Ψk = {μkl r }l=1
mk νkl mk
k 2
m  k
n m
satisfy the following conditions: δkl = 0, k = 1, . . . , n, and δkl /νkl = 0.
l=1 k=1 l=1
Then
m 2  2

n k
n
mk
2 2
δkl /νkl M (Bk , ∂Bk , Zk , k , Ψk )  δkl /νkl M (Z, , Ψ),
k=1 l=1 k=1 l=1
52 Chapter 2. The Asymptotic Behaviour of the Capacity

where Z = {zkl : 1  k  n, 1  l  mk }, = {δkl : 1  k  n, 1  l  mk },


and Ψ = {μkl rνkl : 1  k  n, 1  l  mk }.
Proof. We set
m −1  −1
k
n
mk
2 2
νk = δkl /νkl , k = 1, . . . , n, ν= δkl /νkl .
l=1 k=1 l=1

The reduced moduli exist under the assumptions of Theorem 2.8; it follows from
their definition that

|C(r; Bk , ∂Bk , Zk , k , Ψk )|−1


   2
2π 2π
=− − M (Bk , ∂Bk , Zk , k , Ψk )
νk log r νk log r
 2 
1
+o , r → 0, k = 1, . . . , n,
log r
   2
2π 2π
|C(r; C, ∅, Z, , Ψ)|−1 = − − M (Z, , Ψ)
ν log r ν log r
 2 
1
+o , r → 0.
log r

On the other hand, from Theorem 1.18, since the capacity is monotonic, we obtain


n
|C(r; Bk , ∂Bk , Zk , k , Ψk )|−1  |C(r; C, ∅, Z, , Ψ)|−1 .
k=1

It remains to sum the above relations and, after some simple transformations, take
the limit as r → 0. 
Theorem 2.9. If the sets Z = {zl }nl=1 , , = {δl }nl=1 , and Ψ = {μl rνl }nl=1 satisfy

n 
n
the conditions zl = ∞, l = 1, . . . , m, δl /νl = 0, and δl2 = 0, then for each
l=1 l=1
real number a,
M (B(a), ∂B(a), Z, , Ψ) = M (Z, , Ψ),

n
where B(a) = { z : (δl /νl ) log |z − zl | = a }.
l=1

Proof. Fix a real number a. Let Bl denote the connected component of B(a)
containing zl . Applying the generalized maximum principle to the function


n
n
a− (δl /νl ) log |z − zl | − (δl /νl )gBl (z, zl )
l=1 l=1
2.4. Reduced moduli 53

in each domain Bl , we obtain



n
n
a− (δl /νl ) log |z − zl | ≡ (δl /νl )gBl (z, zl ) in B(a).
l=1 l=1

Adding (δk /νk ) log |z − zk | to both sides and passing to the limit as z → zk , we
find that
n
δl δk
n
δl
a− log |zk − zl | = log r(Bk , zk ) + gB (zk , zl ).
νl νk νl l
l=1 l=1
l=k l=k

Hence

n
δk2
n
n
δk δl
− 2 log μk − log |zk − zl |
νk νk νl
k=1 k=1 l=1
l=k

n
δk2 r(Bk , zk ) δk δl
n n
= log + gB (zk , zl ).
νk2 μk νk νl l
k=1 k=1 l=1
l=k

It remains to use formulae (2.16) and (2.17). 


Now let B be a finitely connected domain without isolated boundary points
on the Riemann sphere C; let Γ be a nonempty closed subset of the boundary
∂B which consists of finitely many nondegenerate connected components; let Z =
{zk }nk=1 be a set of different admissible points in B, zk ∈/ Γ, k = 1, . . . , n, let
n
= {δk }k=0 be a tuple of real numbers with δ0 = 0,
n
δk2 = 0, and let Ψ =
k=1
{ψk (r)}nk=1 , where ψk (r) = μk rνk and μk and νk , k = 1, . . . , n, are positive
numbers. Then we set
 −1
1 n
2
ν= σ(zk )δk /νk .
2 k=1

In the case when all the points zk lie in B, k = 1, . . . , n, this quantity is just
the previously introduced ν. Now we look at the asymptotically circular families
D(zk , ψk (r)), 0 < r < r0 , k = 1, . . . , n, and the condenser C(r; B, Γ, Z, , Ψ)
from § 2.2. The reduced modulus of the domain B with respect to Γ and the sets
Z, , Ψ is defined by
+ ν ,
M (B, Γ, Z, , Ψ) = lim |C(r; B, Γ, Z, , Ψ)| + log r . (2.20)
r→0 2π
By Theorem 2.2 the reduced modulus exists and
M (B, Γ, Z, , Ψ)
 n 
ν 2 σ(zk )δk2 r(B, Γ, zk ) σ(zk )δk δl
n n
= log + gB (zk , zl , Γ) .
4π νk2 μk νk νl
k=1 k=1 l=1
l=k
54 Chapter 2. The Asymptotic Behaviour of the Capacity

Let B and Z, Ψ be the domain and sets defined above. We assume that
B = C, Γ = ∅, and let = {δk }nk=1 be a tuple of real numbers not all of which are

n
equal such that σ(zk )δk /νk = 0. Then the domain B has a reduced modulus
k=1
with respect to the sets Z, , Ψ, which is defined by equality (2.20) with Γ = ∅.
Theorem 2.4 yields
M (B, ∅, Z, , Ψ)
 n 
ν 2 σ(zk )δk2 N (B, zk , zn ) σ(zk )δk δl
n−1 n−1
= log + nB (zk , zl , zn ) ,
4π νk2 μk νk νl
k=1 k=1 l=1
l=k

where it is assumed that δn = 0 and N (B, zn , zn ) = 1.


Theorem 2.10. In the conditions of the definition (2.20), if B2 is a domain obtained

n
by extending B1 across the part of its boundary (∂B1 \ Γ1 ) \ U (zk , ε) for some
k=1
ε > 0, then
M (B1 , Γ1 , Z, , Ψ)  M (B2 , Γ1 , Z, , Ψ).
If the domain B2 is obtained by extending B1 across the part Γ1 of its boundary
and a closed set Γ2 lies in (∂B2 ) \ ((∂B1 ) \ Γ1 ), then the reverse inequality holds:
M (B1 , Γ1 , Z, , Ψ)  M (B2 , Γ2 , Z, , Ψ).
Proof. This follows from the definition of the reduced modulus and Theorems 1.15
and 1.16. 
We can give a description of all equality cases in Theorem 2.10 using Ap-
pendix A4. The behaviour of the reduced modulus under a conformal univalent
mapping is described in the next theorem.
Theorem 2.11. Let B, Γ, Z, and Ψ be as in the definition of the reduced modulus
and assume that a function w = f (z) maps B conformally and univalently onto
a domain f (B) so that Γ is taken to a set f (Γ) by the boundary correspondence,
and the points zk in Z are mapped to some points f (zk ) forming a set W which
are admissible for f (B), k = 1, . . . , n. Assume that
|f (z) − f (zk )| ∼ ck |z − zk |σ(f (zk ))/σ(zk ) , z → zk , (2.21)
at the points zk lying on the boundary, where ck > 0, k = 1, . . . , n; here if zk =
∞ (f (zk ) = ∞), then z − zk (f (z) − f (zk )) must be replaced with 1/z (with
1/f (z)). Then

M (B, Γ, Z, , Ψ) = M (f (B), f (Γ), W, , Ψ),
where Ψ = {
μk rνk }nk=1 with μk = μk |f  (zk )| and νk = νk if zk is an interior point
σ(f (zk ))/σ(zk )
of B and with μk = ck μk and νk = νk σ(f (zk ))/σ(zk ) if zk ∈ ∂B, 1 
k  n.
2.4. Reduced moduli 55

Proof. In view of the geometric meaning of the absolute value of the deriva-
tive, the map f takes an almost disc D(zk , μk rνk ), zk ∈ B, to an almost disc
k rνk ). By (2.21), f takes the connected component of the intersection
D(f (zk ), μ
B ∩ D(zk , μk rνk ) of B and the almost disc which adjoins the point zk ∈ ∂B
to the connected component of f (B) ∩ D(f (zk ), μ k rνk ) adjoining f (zk ), where
k
k r ) is some almost disc in the corresponding family. By the confor-
D(f (zk ), μ ν

mal invariance of the Dirichlet integral



|C(r; B, Γ, Z, , Ψ)| = |C(r; f (B), f (Γ), W, , Ψ)|.

It remains to observe that


 n −1  n −1
1 2 1 2
ν= σ(zk )δk /νk = ν := σ(f (zk ))δk /
νk . 
2 2
k=1 k=1

One immediate consequence of Theorem 1.20 is a symmetry principle.


Theorem 2.12. Let B and Γ ⊂ ∂B be a domain and a closed set which are mirror-
symmetric relative to the real axis and let Z = {zk }m k=1 , = {δk }k=0 , and Ψ =
m

{μk r }k=1 be as in the definition of M (B, Γ, Z, , Ψ). Assume that the points zk ,
νk m

k = 1, . . . , n, where n  m, have nonnegative imaginary parts, {zk }m


k=1 = {z k }k=1 ,
m

and if zk and zl are symmetric points in Z, then δk = δl , μk = μl , and νk = νl .


Assume that B + = {z ∈ B : Im z > 0} is also a domain, and let Γ+ = {z ∈ Γ :
Im z  0}, Z + = {zk }nk=1 , + = {δk }nk=0 , and Ψ+ = {μk rνk }nk=1 . Then

M (B + , Γ+ , Z + , + , Ψ+ ) = 2M (B, Γ, Z, , Ψ).

Combining the symmetry principle with conformal mappings we arrive at the


following method for calculating the reduced moduli.
Theorem 2.13. Let B be a a finitely connected domain without isolated boundary
points on the Riemann sphere C; let Γ be a (maybe empty) closed subset of the
boundary ∂B consisting of finitely many nondegenerate connected components such
that (∂B) \ Γ is nonempty and lies in one boundary component of B. Let Z =
{zk }nk=1 be a set consisting of different points z1 , . . . , zs in B and admissible points
zs+1 , . . . , zn in (∂B) \ Γ (one of these two subsets can be empty, but not both), and
let and Ψ be sets in the definition (2.20) of the reduced modulus. Assume that
a function w = f (z) maps B conformally and univalently into the upper half-
plane Im w > 0 so that L := f ((∂B) \ Γ) lies on the real axis and (2.21) holds with
σ(f (zk )) = 1, k = s+1, . . . , n. If the domain D := f (B)∪L∪{w : w ∈ f (B)} = C,
then
n+s  2 
ν2 δk δk δl
n+s
k )
r(D, w
M (B, Γ, Z, , Ψ) = log + k , w
gD (w l ) .
π νk k
μ νk νl
k=1 k,l=1
k=l
56 Chapter 2. The Asymptotic Behaviour of the Capacity


s 
n
On the other hand, if D = C (Γ = ∅) and 2δk /νk + σ(zk )δk /νk = 0, then
k=1 k=s+1
 n+s  2 
ν2 δk
n+s
 δk δl
M (B, ∅, Z, , Ψ) = − k −
log μ log |w
k − w
l | .
π νk νk νl
k=1 k,l=1
k=l

The following notation is used in both formulae:

k = f (zk ),
w δk = δk , k = μk |f  (zk )|,
μ νk = νk , k = 1, . . . , s;
δk = δk , μ
1/σ(zk )
k = f (zk ),
w k = ck μk , νk = νk /σ(zk ), k = s + 1, . . . , n;
k = f (zk−n ), δk = δk−n , μ
w 
k = μk−n |f (zk−n )|,
n+s −1

νk = νk−n , k = n + 1, . . . , n + s; ν = 2
δk /
νk .
k=1

Proof. This follows by applying in turns Theorem 2.11, Theorem 2.12, and formu-
lae (2.10), (2.15), and (2.20). 
In conclusion we discuss the limiting case of the composition principle for
condensers presented in § 1.3.
Theorem 2.14. Let B, Γ, and Z = {zk }nk=1 , = {δk }nk=0 , and Ψ = {μk rνk }nk=1 be
the sets and systems of points, numbers, and functions from the definition of the re-
duced modulus M := M (B, Γ, Z, , Ψ). Let Bi ⊂ B, i = 1, . . . , m, be pairwise dis-
joint domains and let Γi , Zi = {zij }nj=1
i
, i , i = {δij }nj=1
i
, and Ψi = {μij rνij }nj=1
i

be as in the definition of the reduced moduli Mi := M (Bi , Γi , Zi , i , Ψi ), i =


1, . . . , m. Assume that the following conditions are fulfilled: each point zij in Zi
coincides with some point zk ∈ Z with k = k(i, j)4 ; for any i and j and for
k = k(i, j), δij = δk , μij = μk , and νij = νk ; Γi ⊂ Γ, i = 1, . . . , m. Also let

n
m
ni
σ(zk )δk2 /νk = 2
σ(zij )δij /νij . (2.22)
k=1 i=1 j=1

Then ⎛ ⎞2
 2

n
m ni
M σ(zk )δk2 /νk  Mi ⎝ 2
σ(zij )δij /νij ⎠ .
k=1 i=1 j=1

Proof. By Theorem 1.17



m
|C(r; B, Γ, Z, , Ψ)|−1  |C(r; Bi , Γi , Zi , i , Ψi )|−1 . (2.23)
i=1
4 When zij is an attainable boundary point of Bi , we mean that its support coincides with zk or
with the support of zk .
2.4. Reduced moduli 57

It follows from the definition of the reduced modulus that

|C(r; B, Γ, Z, , Ψ)|−1
   2  2 
2π 2π 1
=− −M +o , r → 0,
ν log r ν log r log r
|C(r; Bi , Γi , Zi , i , Ψi )|−1
   2  2 
2π 2π 1
=− − Mi +o , r → 0,
νi log r νi log r log r

where
 −1 ⎛ ⎞−1
1 1
n ni
ν= σ(zk )δk2 /νk and νi = ⎝ 2
σ(zij )δij /νij ⎠ , i = 1, . . . , m.
2 2 j=1
k=1

Taking (2.22) and (2.23) into account we obtain the required inequality. 

Exercises 2.4
(1) Write out the relation between the logarithmic capacity of a bounded closed
set E and the reduced modulus of the open set C \ E with respect to the
point at infinity.
(2) For a domain D ⊂ C with Green’s function gD (z, ζ) := − log |z − ζ| + h(z, ζ)
verify the formula
1
M (D, ∂D, {z1 , z2 }, {0, −1, 1}, {r, r}) = {log |z1 − z2 |2 + H(D, z1 , z2 )},

where z1 , z2 ∈ D ∩ C and H(D, z, ζ) = h(z, z) + h(ζ, ζ) − 2h(z, ζ).
(3) Prove the following asymptotic formula using the above notation:

H(D, z0 − ρeiϕ , z0 + ρeiϕ )


= −4πρ2 {K(z0 , z 0 ) − Re[e2ϕi l(z0 , z0 )]} + o(ρ2 ), ρ → 0,

where K(z, ζ) and L(z, ζ) = 1/[π(z − ζ)2 ] − l(z, ζ) are kernels well known in
function theory:

2 ∂ 2 g(z, ζ) 2 ∂ 2 g(z, ζ)
K(z, ζ) = − , L(z, ζ) = − .
π ∂z ∂ζ π ∂z ∂ζ

(4) Prove Renggli’s inequality:

(cap(E1 ∪ E2 ))(cap(E1 ∩ E2 ))  (cap E1 )(cap E2 ).


58 Chapter 2. The Asymptotic Behaviour of the Capacity

(5) Using (2.17) and the symmetry principle for harmonic functions calculate
the reduced modulus M (B, ∂B, Z, , Ψ) in the case when B is a disc or a
half-plane.
(6) Describe all the equality cases in (2.18).
(7) Prove (2.17) by setting B = {z : |z| < R} in formula (2.16) and passing
to the limit as R → ∞. Verify that the limit transitions as r → 0 (in the
definition of reduced modulus) and as R → ∞ are interchangeable.
(8) Prove the following result. Let Z1 = {z1l }ml=1 , Z2 = {z2l }l=1 be fixed tuples
1 m2

of different finite points, let 1 = {δ1l }l=1 and 2 = {δ2l }m


m1 2
l=1 be tuples of
1 −1
m
1’s, and let Ψ1 = {rν1l }m
l=1 and Ψ2 = {r
1
}l=1 , where νjl > 0 and
ν2l m2
ν1l =
l=1
2
m
−1 −1
ν2l := ν . Then for any open disjoint sets B1 and B2 in the plane C
l=1
such that zjl ∈ Bj , j = 1, 2, l = 1, . . . , mj ,

M (B1 , ∂B1 , Z1 , 1 , Ψ1 ) + M (B2 , ∂B2 , Z2 , 2 , Ψ2 )


m11 m12 −1 −1

2 |z1k − z2l |ν1k ν2l


ν k=1 l=1
 log 1 −1 −1 1 −1 −1
π |z1k − z1l |ν1k ν1l |z2k − z2l |ν2k ν2l
1k<lm1 1k<lm2

(for mj = 1 the corresponding product is set equal to 1). Describe all the
equality cases in this inequality. As a consequence, deduce the results on
the product of the conformal radii of two nonoverlapping domains due to
Lavrent’ev and Kufarev. Is it also possible to prove Goluzin’s corresponding
results on three nonoverlapping domains in this way (see [Leb], Introduc-
tion, § 2)?
(9) Let B be a finitely connected domain in the sphere C, z0 a finite point in it,
and Γ some marked nondegenerate boundary component of this domain. It
is known that there exists a unique conformal univalent mapping w = f (z)
of B onto the disc |w| < 1 cut along several pieces of radii going out of the
centre which takes the continuum Γ to the circle |w| = 1 and satisfies the
conditions f (z0 ) = 0 and f  (z0 ) > 0 ([J], Theorem 5.5). Using Dirichlet’s
principle with free boundary [HurC] verify that

1
M (B, Γ, {z0 }, {0, 1}, {r}) = − log f  (z0 ).

(10) A strip (digon) is a simply connected domain of hyperbolic type with two
marked different attainable points, which are called vertices (Figure 2.2).
We consider only strips P with vertices z1 and z2 which have the following
property. If a function w = f (z) maps P conformally and univalently onto
the straight strip {w : 0 < Re w < 1} so that f (z1 ) = −i∞ and f (z2 ) = +i∞,
2.4. Reduced moduli 59

then in a neighbourhood of either of the zk

(−1)k+1 i
f (z) = log(z − zk ) + ck + o(1), z → zk , (2.24)
ϕk π

where ϕk > 0 and ck are constants, k = 1, 2 (if zk = ∞, then the local


parameter z − zk must be replaced by 1/z). Clearly, ϕk π is equal to the
interior angle of the domain P at zk , k = 1, 2. Following Kuz’mina and
Emel’yanov [Kuz2], we mean by the reduced modulus of P the quantity

M (P ) = Im(c2 − c1 ).

Show that

M (P ) = 4M (P, ∅, {z1 , z2 }, {−1, 1}, {rϕ1 , rϕ2 }).

{5 {5
{3 (4  {3 (5 
(3  (3 
{4
{ 3 (3  (4  {4
{4

Figure 2.2: Left to right: a strip, a half-strip, and a trilateral.

(11) A half-strip is a simply connected domain of hyperbolic type P ⊂ C with


three different marked attainable boundary points z1 , z2 , and z3 , following
one another in accordance with the positive orientation of the boundary of
P ; one of them (for instance, z1 ) is called the vertex. We only consider half-
strips P whose vertex z1 has the following property. If a function w = f (z)
maps P conformally and univalently onto a straight half-strip {w : 0 <
Re w < 1, Im w < 0} so that f (z1 ) = −i∞, f (z2 ) = 1, and f (z3 ) = 0, then
the expansion (2.24) with k = 1 holds in a neighbourhood of z1 . Following
Solynin [Sol3], we define the reduced modulus of the half-strip P (called a
trilateral in [Sol3]) as
M (P ) = −Im c1 .
Prove the equality

M (P ) = M (P, Γ, {z1 }, {0, 1}, {rϕ1 }) = M (P, Γ, {z1 }, {0, 1}, {r}),

where Γ is the part of the boundary of P lying between the points z2 and z3
(that is, the ‘side’ opposite to z1 ).
(12) State and prove an analogue of Theorem 2.12 in which the real axis is replaced
by an arbitrary straight line or a circle.
60 Chapter 2. The Asymptotic Behaviour of the Capacity

(13) Let B = {z : |z| < 1} be the unit disc and Γ a closed subset of its boundary.
Show that
1
M (B, Γ, {0}, {0, 1}, {r}) = − log cap Γ,
π
where the reduced modulus is defined by (2.20) with no extra conditions on Γ.
(14) Let B be a triangle bounded by straight-line segments with end-points z1 =
0, z2 = 1, and z3 , where Im z3 > 0, and let Z = {zk }3k=1 . Let αk π > 0, k =
1, 2, 3, denote the angles at the corresponding vertices of B. Let = {δk }3k=1
3
and Ψ = {μk rνk }3k=1 be sets satisfying the relation αk δk /νk = 0. Verify
k=1
the equality
 3
−2 3
1 2 2
M (B, ∅, Z, , Ψ) = − αk δk2 /νk log (μk αk B(α1 , α2 ))αk δk /νk ,
π
k=1 k=1

where B(α1 , α2 ) is the beta function.


(15) State Theorem 2.14 in the following special cases:
a) B is a simply connected domain, Γ = ∂B, the set Z consists of a
single point z0 ∈ B, and the domains Bi are half-strips as in Exercise 11,
with vertex at z0 and opposite sides lying on Γ;
b) the domains B and Bi , i = 1, . . . , m, are strips with the same vertices
(see Exercise 10);
c) the domains B and Bi , i = 1, . . . , m, are half-strips with a common
vertex and opposite sides Γi ⊂ Γ, i = 1, . . . , m.
(16) Using Theorem 1.18 formulate and prove a composition principle for reduced
moduli ensuring the reverse inequality to the one in Theorem 2.14. State a
special case of this principle when B is a strip and B1 and B2 are disjoint
half-strips in B whose vertices coincide with the vertices of B.
Chapter 3

Special Transformations

3.1 Contractions
Let E be a compact subset of the plane C. A map T : E → C is called a contrac-
tion if
|T (z) − T (ζ)|  |z − ζ|
for all z, ζ ∈ E
The following result yields many estimates for the logarithmic capacity of
compact sets.
Theorem 3.1. If T is a contraction of a compact set E onto T (E), then

cap E  cap T (E).

Proof. This immediately follows from the geometric interpretation of the logarith-
mic capacity (Theorem 1.23). 
The equality case in Theorem 3.1 is considered in Appendix A2. As conse-
quences of Theorem 3.1, we present the so-called ‘1/4-estimates’ for the logarithmic
capacity of a compact set.
Theorem 3.2. Let E be a compact subset of C.
a) If E is connected and has diameter d, then

cap E  d/4.

b) If E is a rectifiable curve of length l, then

cap E  l/4.

c) If E lies on a straight line and has Lebesgue measure m there, then

cap E  m/4.

© Springer Basel 2014 61


V.N. Dubinin, Condenser Capacities and Symmetrization in Geometric Function
Theory, DOI 10.1007/978-3-0348-0843-9_3
62 Chapter 3. Special Transformations

d) If the orthogonal projection of E onto some straight line has Lebesgue mea-
sure m, then
cap E  m/4.

e) If E is a subset of the unit half-circle |z| = 1, Imz  0 with angular mea-


sure A, then
cap E  sin(a/4).

Proof. a) After a rotation and a parallel translation we can assume that 0, d ∈ E.


Let T : C → R denote the orthogonal projection onto the real axis. It is easy to
see that T is a contraction. By Theorem 3.1

cap E  cap T (E)  cap ([0, d]) = d/4.

b) Let T : [0, l] → E be the natural parametrization of a curve E. Then T is


a contraction, and again, by Theorem 3.1

l/4 = cap ([0, l])  cap E.

c) Let E be a compact subset of the real axis. It is sufficient to observe


that if T (x) is the linear Lebesgue measure of the set E ∩ (−∞, x], then the map
T : R → R is a contraction and T (E) = [0, m].
d) This follows from the application of Theorem 3.1 to the orthogonal pro-
jection T and from part c) (cf. [Gol], Ch. VII, § 2, Theorem 2).
e) The proof is similar to c). 

Let C = (E0 , E1 ) be a condenser with plates in the open plane C. We call a


map T : E0 ∪ E1 → C a contraction of C if the following conditions hold: for all
z, ζ ∈ E0 or z, ζ ∈ E1
|T (z) − T (ζ)|  |z − ζ|;

while for any z ∈ E0 and ζ ∈ E1

|T (z) − T (ζ)|  |z − ζ|.

Theorem 3.3. If T is a contraction of a condenser (E0 , E1 ), then

cap (E0 , E1 )  cap (T E0 , T E1 ).

Proof. This follows from the geometric interpretation of the condenser capacity
(Theorem 1.21). 
3.2. Polarization 63

Exercises 3.1
(1) Using Theorem 3.2 a), prove the Koebe–Bieberbach 1/4-theorem in the class
S (see [Gol], Ch. II, § 4).
(2) Give details of the proof of part e) of Theorem 3.2.
(3) Show that in part e) of Theorem 3.2 the half-circle can be replaced with the
circle |z| = 1.
(4) Let m(θ, E) be the linear measure of the intersection of a bounded closed set
E with the ray arg z = θ and let
LRE = {reiθ : 0  r  m(θ, E), 0  θ  2π}.
Show that if E lies on several rays with initial point at the origin which form
angles with opening at least π/2, then
cap E  cap LRE. (3.1)

(5) Show that if E lies on the rays arg z = ±θ1 and arg z = ±θ2 , where 0  θ1 
θ2 − π/2  π/2, or if E is mirror-symmetric relative to the real axis, then
(3.1) holds.
(6) Let θ1 , θ2 , . . . , θn be real numbers and F be a compact subset of the nonneg-
ative half-axis. Show that inequality (3.1) holds for the set

n
E= {reiθk : r ∈ F }.
k=1

(7) (Gehring [Geh]) Using Theorem 1.21 verify Vuorinen’s conjecture on plane
condensers:
cap (E0 , E1 )  cap (E0 , E1∗ ),
where E0 , E1 ⊂ {z : Imz  0} and E1∗ = {z : z ∈ E1 }.

3.2 Polarization
Let α be a fixed oriented straight line or a circle in C, and let C− +
α and Cα be the

closures in C of the two domains bounded by α (the set Cα lies ‘to the left’ of α).
We shall use the following notation:
z ∗ is the point symmetric to z ∈ C relative to α;
A∗ = {z : z ∗ ∈ A}, A− = A ∩ C− α, A+ = A ∩ C+ α, A ⊂ C;
− ∗ − ∗ +
Pα A = (A ∪ A ) ∪ (A ∩ A ) , Pα+ A = (A ∪ A∗ )+ ∪ (A ∩ A∗ )− .
By a polarization of a set A with respect to α we mean the transformation of A
into Pα− A or Pα+ A.
Now let C = (E0 , E1 ) be a condenser in C. We define the polarization of
C with respect to α as the transformation of this condenser into the condenser
Pα C = (Pα− E0 , Pα+ E1 ).
64 Chapter 3. Special Transformations

 ,  ,
Q , F3
F3


F2 Q  F2

Figure 3.1.

Theorem 3.4 ([W, D2]). For any condenser C

cap C  cap Pα C. (3.2)

In addition, if C is an admissible condenser with connected field, then equality in


(3.2) holds if and only if C either coincides with the condenser Pα C or they are
symmetric relative to α.
Proof. It is sufficient to look at the case when α is a straight line. Let v(z) be a
function in the class L(C) and v ∗ (z) := v(z ∗ ), z ∈ C. We set

Δ1 = {z : v(z)  v ∗ (z)}+ ∪ {z : v(z)  v ∗ (z)}− ,

and

Δ2 = {z : v(z) > v ∗ (z)}− ∪ {z : v(z) < v ∗ (z)}+ .

Let us introduce the auxiliary function


 
min(v(z), v ∗ (z)), z ∈ C−
α, v(z), z ∈ Δ1 ,
Pα v(z) = = (3.3)
max(v(z), v ∗ (z)), z ∈ C+
α , v ∗ (z), z ∈ Δ2 .

By Corollary 1.1, Pα v(z) belongs to the class Lip; furthermore, 0  Pα v  1 in C,


Pα v = 0 on the set Pα− E0 , and Pα v = 1 on Pα+ E1 . Hence

I(Pα v, C)  cap Pα C.

Note that Pα v = v almost everywhere on Δ1 and ∇Pα v = ∇v ∗ on 2 . Bearing


in mind that the Dirichlet integral is invariant under Möbius transformations
3.2. Polarization 65

we obtain

I(Pα v, C) = I(Pα v, 1 ) + I(Pα v, 2 ) = I(v, 1 ) + I(v ∗ , ∗2 )


= I(v, 1 ) + I(v, 2 ) = I(v, C).

Thus
I(v, C)  cap Pα C,
and (3.2) follows since v can be arbitrary.
Now let C be an admissible condenser with connected field and assume that
we have equality in (3.2). Let ω(z) be the potential function of the condenser C and
{C1 , C2 , C3 } be the decomposition of C corresponding to a partition {0, t, 1 − t, 1},
where 0 < t < 1/2. We can apply the above arguments to v = trun((ω −t)/(1−2t))
and the condenser C2 ; then we see that

I(v, C) = I(Pα v, C).

On the other hand it follows from Theorem 1.14 and the above that
3
3

|C| = |Ck |  |Pα Ck |  |Pα C| = |C|.
k=1 k=1

Hence
cap Pα C2 = cap C2 = I(v, C),
so that
cap Pα C2 = I(Pα v, C).
From Theorem 1.13 we conclude that Pα v is the potential function of the condenser
Pα C2 . By the uniqueness theorem for harmonic functions, in each connected com-
ponent of the field G2 of Pα C2 we either have Pα v ≡ v or Pα v ≡ v ∗ . Hence in
each connected component of G2 either Pα ω ≡ ω or Pα ω ≡ ω ∗ . Since C is a regu-
lar condenser, we see from the properties of harmonic functions that one of these
identities holds in the field of C, and therefore the condenser Pα C either coincides
with C or is symmetric to it relative to α. 
Inequality (3.2) also holds when the condenser C is ‘defined’ on a set B ⊂ C
which is mirror-symmetric relative to the curve α. Namely, let

C = (B, {E0 , E1 }, {0, 1}) and Pα C = (B, {Pα− E0 , Pα+ E1 }, {0, 1})

(Figures 3.2 and 3.3).


Theorem 3.5. If the open set B ⊂ C is mirror-symmetric relative to a straight line
or a circle α, then (3.2) holds. Furthermore, if here α is a straight line, B a strip
bounded by lines orthogonal or parallel to α, and C an admissible condenser with
connected field, then equality holds in (3.2) if and only if C either coincides with
the condenser Pα C or is symmetric to it relative to α.
66 Chapter 3. Special Transformations


F2 F3 Q  F2 Q , F3

Figure 3.2.

F2 F3 Q  F2 Q , F3

Figure 3.3.

Proof. The special case of Theorem 3.5 when B is a strip is important for us; this
case reduces to Theorem 3.4 by means of the following procedure. Let α be the
imaginary axis and B = {z : 0 < Im z < π}. Let A  be the union of the set A
with its reflection in the real axis. From Theorem 1.20 and since the capacity is a
conformal invariant, we obtain
0 , E
2 cap C = cap({z : |Im z| < π}, {E 1 }, {0, 1})
= cap(exp E 0 , exp E
1 ).

The polarization of the condenser C with respect to α corresponds to the polariza-


tion of the condenser (exp E 0 , exp E
1 ) with respect to the anticlockwise oriented
circle |z| = 1. Thus (3.2) holds. In addition, if C is an admissible condenser,
B \ (E0 ∪ E1 ) is a connected set, and (∂B) \ (E0 ∪ E1 ) = ∅, then (exp E0 , exp E1 ) is
an admissible condenser with connected field. So Theorem 3.4 gives us conditions
for equality. In the case when (∂B) \ (E0 ∪ E1 ) = ∅ the condenser (exp E 0 , exp E
1 )
falls into two, each having a connected field. Again, the equality conditions follow
from Theorem 3.4. The case when the line α is parallel to the sides of B (for in-
stance, α : Im z = π/2, while B is as above) can be treated in a similar way, except
that now the polarization of C with respect to α corresponds to the polarization
of the condenser (exp E 0 , exp E1 ) with respect to the imaginary axis. For an ar-
bitrary mirror-symmetric set B the proof of (3.2) (leaving aside the discussion
of conditions for equality) repeats in fact the corresponding part of the proof of
Theorem 3.4, where the class L(C) is set to consist of the admissible functions for
the condenser C. 
Now let B be an arbitrary open set in C, let Z = {zk }m k=1 , m  1, be an
arbitrary system of different points in B, = {0, 1, . . . , 1}, and Ψ = {ψk (r)}m
k=1 ,
3.2. Polarization 67

where ψk (r) = μk rνk with some positive numbers μk and νk , k = 1, . . . , m. Let


α be an oriented straight line. We set zk (α) = Pα− zk if the system Z contains
no point that is symmetric to zk relative to α or such a point (zl say) exists but
ψk (r)  ψl (r) for sufficiently small r; otherwise let zk (α) = Pα+ zk .
Theorem 3.6. The following inequality holds:

M (B, ∂B, Z, , Ψ)  M (Pα− B, ∂Pα− B, {zk (α)}m


k=1 , , Ψ). (3.4)

In addition, if B is an admissible domain, then equality is attained in (3.4) if and


only if either B = Pα− B and zk = Pα− zk for k = 1, . . . , m, or B = Pα+ B and
zk = Pα+ zk for k = 1, . . . , m.
Proof. Let  

m
C(r) := U (zk , ψk (r)), C\B
k=1

and assume that r > 0 is sufficiently small. It is easy to see that the condenser
Pα C(r) has the following form:
 m  

− −
Pα C(r) = Pα U (zk , ψk (r)) , C\Pα B
k=1
 

m
= U (zk (α), ψk (r)), C\Pα− B .
k=1

Now inequality (3.4) follows from Theorem 3.4 for the condensers C(r) and Pα C(r)
and from the definition of the reduced modulus.
If equality holds in (3.4), then we look at the decomposition {B  , C} of B
with respect to the sets Z and {νk−1 }mk=1 that corresponds to a partition {0, t},
where t is sufficiently large. Using Theorem 2.7, the above result, Theorem 3.4,
and Theorem 2.7 again we obtain successively

M (B, ∂B, Z, , Ψ) = M (B  , ∂B  , Z, , Ψ) + |C|


 M (Pα− B  , ∂Pα− B  , {zk (α)}m
k=1 , , Ψ) + |Pα C|
 M (Pα− B, ∂Pα− B, {zk (α)}m
k=1 , , Ψ) = M (B, ∂B, Z, , Ψ).

Thus we have equality signs in all the intermediate relations. In particular, |C| =
|Pα C| and it follows from Theorem 3.4 that the condensers C and Pα C coincide
up to a reflection in the line α. Since t can be arbitrary, the same can be said about
the relation of the domain B and the points zk , k = 1, . . . , m, to the domain Pα− B
and the points Pα− zk , k = 1, . . . , m. 
Corollary 3.1. For an arbitrary open set B ⊂ C and any z0 ∈ B

r(B, z0 )  r(Pα− B, Pα− z0 ).


68 Chapter 3. Special Transformations

In addition, if B is an admissible domain, then equality holds here if and only if


B and z0 coincide with the set Pα− B and the point Pα− z0 up to a reflection in the
line α.

Corollary 3.2. For each compact set E ⊂ C and each straight line α

cap E  cap Pα+ E. (3.5)

If C \ E is an admissible domain, then equality in (3.5) holds only for E = Pα+ E


or E = Pα− E.

Let H be the upper half-plane Im z > 0 or strip 0 < Im z < 1. Fix real x0
and let α = {z : Re z = x0 } be a straight line oriented in the increasing direction
of Im z.

Theorem 3.7. Let B be a finitely connected domain in C without isolated boundary


points such that the set (H \ B) ∪ (B \ H) is bounded. Assume that the difference
H \B lies in the half-plane Re z < x0 and B \H lies in Re z > x0 . Let Γ be a closed
subset of ∂B consisting of finitely many nondegenerate connected components and
such that Γ ⊃ {∂[(H \ B) ∪ (B \ H)]} \ ∂H and if z ∈ Γ ∩ ∂H, Re z  x0 , then
−z + 2x0 ∈ Γ ∪ (H \ B). Let Z = {zk }nk=1 be a set of different admissible points in
B \ Γ, = {δk }nk=0 with δ0 = 0 and δk > 0, k = 1, . . . , n, and let Ψ = {r, . . . , r}.
Then
M (B, Γ, Z, , Ψ)  M (B, Γ, Z  , , Ψ),

where Z  = {zk }nk=1 is the set of points defined as follows: for fixed k, if the set Z
contains a point zl ∈ Z which is symmetric to zk ∈ Z relative to the line α, then
zk = zk in the case when Re zk  x0 and δk  δl or Re zk  x0 and δk  δl , and
zk = zl in the case when Re zk  x0 and δk < δl or Re zk  x0 and δk > δl ; on
the other hand, if the symmetric point −z k + 2x0 does not belong to Z, then by
definition zk = zk if Re zk  x0 and zk = −zk + 2x0 if Re zk  x0 .

Proof. Let g(z) be an auxiliary function defined by (2.11), where gk (z) =


gB (z, zk , Γ), Rk = r(B, Γ, zk ), and ψk = r, k = 1, . . . , n, and assume that r
is sufficiently small. We set Dk (zk , r) = {z ∈ B ∩ U (zk , 2r) : g(z)/δk > 1},
k = 1, . . . , n. As we mentioned in the proof of Theorem 2.2, for zk ∈ B the domains
D(zk , r), 0 < r < r0 , make up an asymptotically circular family; its elements are
almost discs with centre zk and radius r, while if zk ∈ ∂B, then the set D(zk , r)
is an almost half-disc. Let C(r; B, Γ, Z, , Ψ) be a condenser with degenerating
plates D(zk , r), k = 1, . . . , n. We set D(zk , r) = Pα+ D(zk , r) if Re zk  x0 and
D(zk , r) = Pα− D(zk , r) otherwise. We see from (2.11) that if z ∈ B and z ∈ / zk ,
1  k  n, then g(z) → 0 as r → 0. Bearing this in mind and following the
proof of Theorem 3.4 we can show that the function Pα g(z) defined by (3.3) (with
g(z) = v(z)) is admissible for the condenser C(r; B, Γ, Z  , , Ψ) with degenerating
3.2. Polarization 69

plates D(zk , r), k = 1, . . . , n. Moreover,


 

n
cap C(r; B, Γ, Z, , Ψ) = I g, B \ D(zk , r)
k=1
 

n
=I Pα g, B \ D(zk , r)  cap C(r; B, Γ, Z  , , Ψ).
k=1

It remains to use the definition (2.20). 


Let α be the same straight line as above. Repeating the proof of Theorem
3.7 we verify the following result.
Theorem 3.8. Let B be a finitely connected domain without isolated boundary
points in C which is mirror-symmetric relative to α; let Γ be a closed subset
of ∂B consisting of finitely many nondegenerate connected components; let Z =
{zk }nk=1 be some set of different admissible points of B not lying in Γ, and let
Ψ = {r, . . . , r}. If = {δk }nk=0 with δ0 = 0 and δk > 0, k = 1, . . . , n, then

M (B, Γ, Z, , Ψ)  M (B, Pα− Γ, Z  , , Ψ), (3.6)

where the set Z  was defined in Theorem 3.7. Inequality (3.6) also holds for an
arbitrary tuple = {δk }nk=0 with δ0 = 0 and δk = 0, k = 1, . . . , n, provided that
Γ is mirror-symmetric in α and Z  is as above, with one exception: if the point
−zk + 2x0 symmetric to zk does not belong to Z and δk < 0, then by definition
zk = zk for Re zk  x0 and zk = −z k + 2x0 for Re zk  x0 .
Using the methods presented above the reader can easily describe the be-
haviour under polarization of the reduced moduli in other examples.

Exercises 3.2
(1) Show that Pα− A = (Pα+ A)∗ and Pα+ (C \ A) = C \ Pα− A.
(2) Is it true that after polarization with respect to a line α a set E with sym-
metry axis parallel to α shifts in the direction orthogonal to α?
(3) Let (E0 , E1 ) be a condenser in C. Check that the pair (Pα− E0 , Pα+ E1 ) is
indeed a condenser.
(4) Give an example of a domain G such that Pα+ G is not a domain.
+
(5) Show that if G is a domain, then (Pα+ G) ∩ Cα is a connected set.
(6) In the proof of Theorem 3.4 verify that Pα v ∈ L(Pα C).
(7) In the proof of Theorem 3.4 verify that ∇Pα v = ∇v a.e. on 1 and ∇Pα v =
∇v ∗ on 2 .
(8) Examine the equality case in Theorem 3.5 for arbitrary admissible con-
densers C.
70 Chapter 3. Special Transformations

(9) Let δ ∈ (0, 2) be a fixed number, E a compact subset of the imaginary axis
and F a compact set in C such that Pα− F = F for each straight line α given
by the equation Re z = t with t ∈ (−1, −δ/2) and oriented in the increasing
direction of Im z. Assume that the condenser

C1 (α) = (F, E(α)),

where E(α) = {x + iy : x ∈ [−1, α] ∪ [α + δ, 1], iy ∈ E}, is well defined. Show


that cap C1 (α) is a strictly increasing function on the interval (−1, −δ/2).
(10) In the notation of the previous exercise let F = {z : Im z  d}, where
d > sup E, B = {z : Im z > c}, where c < inf E, and

C2 (α) = (B, {F, E(α)}, {0, 1}).

Show that cap C2 (α) is a strictly increasing function on (−1, −δ/2).



(11) Let E(α) be as above and let E(α) be the set symmetric to E(α) relative to
the straight line Im z = c. Show that the capacity of the condenser

C3 (α) = (E(α), E(α))

is strictly increasing on (−1, −δ/2) (see the left-hand plot in Figure 3.4).

 D6 )*

 D7 ) *
F)*
D5 ) * 

3 , 3 G

Figure 3.4.


n
(12) Let C4 (ν) = (F, E(ν)), where F = Fk with Fk = {reiϕ : ϕ ∈ [−γk , γk ],
k=1
m
r ∈ Ik } for k = 1, . . . , n and E(ν) = El (ν) with El (ν) = {z = reiϕ : ϕ ∈
l=1
[ν − λl , ν + λl ], r ∈ Jl } for l = 1, . . . , m, 0 < γk < π and 0 < λl < π. Let
Ik and Jl be subintervals of (0, +∞) such that Ik ∩ Jl = ∅, k = 1, . . . , n,
l = 1, . . . , m. Show that cap C4 (ν) is a strictly decreasing and cap(F ∪ E(ν))
a strictly increasing function on (0, π) (he right-hand plot in Figure 3.4).
(13) State and prove the analogues of Theorems 3.6–3.8 in the case when α is a
circle and the analogues of Theorems 3.7 and 3.8 for an arbitrary set Ψ =
{μk rνk }nk=1 and Γ = ∅.
3.3. The Gonchar transformation 71

3.3 The Gonchar transformation


We look at a condenser C = (E0 , E1 ) in C with plates Ej = Tj ×Kj := {z = x+iy :
x ∈ Tj , y ∈ Kj }, j = 0, 1, where T0 and T1 are disjoint nonempty closed subsets
of the interval [−1, 1] of the real axis and K0 and K1 are bounded nonempty closed
subsets of the imaginary axis. In 1969 Gonchar made the following conjecture:

cap C  cap GC (3.7)

for the condenser GC := (G− E0 , G+ E1 ) with plates G− E0 = [−1, −1 + mT0 ] × K0


and G+ E1 = [1 − mT1 , 1] × K1 , where mTj is the linear Lebesgue measure of
Tj , j = 0, 1. Using polarization of condensers we can answer this question in the
affirmative.

H F 2

F2
F3 H, F 3
3 3 3 3
U2 U3 H U 2 H, U 3

Figure 3.5: The Gonchar transformation.

Before we give the proof, we present an elementary lemma.


Lemma 3.1. Let A and B be subsets of the real axis which are formed from finitely
many disjoint (maybe degenerate) closed intervals. Then

λ(A ∪ B) + λ(A ∩ B) = λ(A) + λ(B),

where λ(T ) is the number of intervals forming the set T .

Proof. The equality is obvious for λ(A) + λ(B) = 2. Now assume that it holds for
any sets with the total number of intervals equal to s  2, and let A and B be sets
such that λ(A) + λ(B) = s + 1. Then one of these sets, for instance, A, contains at
least 2 intervals. Let A be the set obtained by removing an interval, say I, from
A, so that A = A ∪ I, and let t be the number of intervals forming I ∩ B. It is
easy to see that λ(A ∪ B) = λ(A ∪ B) − t + 1 and λ(A ∩ B) = λ(A ∩ B) + t. From
the inductive assumption we obtain

λ(A ∪ B) + λ(A ∩ B) = λ(A ∪ B) + λ(A ∩ B) + 1


= λ(A ) + λ(B) + 1 = λ(A) + λ(B). 
72 Chapter 3. Special Transformations

Now we turn to the main result.


Theorem 3.9 ([D2] and [D3]). Let C and GC be the condensers defined above. Then
inequality (3.7) holds. In addition, if C is an admissible condenser, then equality
holds in (3.7) if and only if C either coincides with GC, or they are symmetric
relative to the imaginary axis.
Proof. We start with the case when each of the sets T0 and T1 is a union of
finitely many disjoint nondegenerate intervals. Assume that T0 consists of k closed
intervals and T1 consists of l intervals; for definiteness assume that sup T0 < sup T1 .
Let x1 be the left-hand end-point of the leftmost interval in T1 and x2 be the left-
hand end-point of the rightmost interval in T1 . Polarizing the condenser C with
respect to the line α = {z : Re z = (x1 + x2 )/2} codirectional with the imaginary
axis we obtain a condenser Pα C = (T 20 × K0 , T
21 × K1 ), where

T0 = Pα− T0 and T1 = Pα+ T1 .

By Theorem 3.4,
cap C  cap Pα C.
The plates of Pα C have the following properties:
a) T0 ∪ T1 ⊂ [−1, 1];
b) mTj = mTj , j = 0, 1;
c) the set T0 consists of k pairwise disjoint (maybe degenerate) closed intervals,
and T1 consists of l intervals;
d) x1 is an isolated point (degenerate interval) in T1 .
Properties a) and d) are obvious, and b) follows from the chain of equalities

mT0 = (mT0 + mT0∗ )/2 = (m(T0 ∪ T0∗ ) + m(T0 ∩ T0∗ ))/2


20 .
= m(T0 ∪ T0∗ )− + m(T0 ∩ T0∗ )+ = mT

In the proof of c) we use Lemma 3.1:


20 ) = λ((T0 ∪ T ∗ )− ∪ (T0 ∩ T ∗ )+ )
λ(T 0 0
= λ((T0 ∪ T0∗ )− ) + λ((T0 ∩ T0∗ )+ ) − λ((T0 ∪ T0∗ )− ∩ (T0 ∩ T0∗ )+ )
= λ(T0− ∪ (T0∗ )− ) + λ(T0− ∩ (T0∗ )− ) − λ(T0− ∩ T0+ )
= λ(T0− ) + λ((T0∗ )− ) − λ(T0− ∩ T0+ )
= λ(T0− ) + λ(T0+ ) − λ(T0− ∩ T0+ ) = λ(T0− ∪ T0+ ) = λ(T0 )

(the case j = 1 is similar). Thus the plates have indeed properties a)–d).
Removing all the isolated points from T 20 and T21 we obtain sets T 1 and T 1 ,
0 1
respectively. Since the capacity is monotonic, we have

cap Pα C  cap(T01 × K0 , T11 × K1 ).


3.3. The Gonchar transformation 73

The sets Tj1 , j = 0, 1, are formed by finitely many disjoint nondegenerate intervals,
and we have λ(T01 )  k and λ(T11 )  l − 1 (see properties c) and d)). Furthermore,

T01 ∪ T11 ⊂ [−1, 1] and mTj1 = mTj , j = 0, 1,

(properties a) and b)).


Now we repeat our arguments replacing C = (T0 × K0 , T1 × K1 ) with the
condenser (T01 × K0 , T11 × K1 ) and so on, until after several (say, r) steps we obtain
a condenser (T0r × K0 , T1r × K1 ) such that

λ(T0r )  k, λ(T1r ) = 1, T0r ∪ T1r ⊂ [−1, 1], mTjr = mTj , j = 0, 1,

and
cap(T01 × K0 , T11 × K1 )  cap(T0r × K0 , T1r × K1 ).
Making similar transformations of the condenser (T0r × K0 , T1r × K1 ) which
involve the right end-points of intervals in T0r we finally obtain a condenser (T0s ×
K0 , T1s × K1 ) such that

λ(T0s ) = λ(T1s ) = 1, T0s ∪ T1s ⊂ [−1, 1], mTjs = mTj , j = 0, 1,

and
cap(T0r × K0 , T1r × K1 )  cap(T0s × K0 , T1s × K1 ).
Now let x be the left end-point of T1s and x be the right end-point of T0s . Making
polarizations, first with respect to the line α1 = {z : Re z = (x + 1)/2} codirec-
tional with the imaginary axis and then with respect to the (codirectional) line
α2 = {z : Re z = (−1 + x )/2}, we obtain

Pα2 Pα1 (T0s × K0 , T1s × K1 ) = GC.

Using Theorem 3.4 again, in view of the above inequalities, we conclude that

cap C  cap(T0s × K0 , T1s × K1 )  cap GC,

where equality is only possible in the case described in the statement of Theo-
rem 3.9.
Now if we know from the outset that sup T0 > sup T1 , then performing the
polarization of the condenser C with respect to the line Re z = (inf T0 + sup T1 )/2
and maybe also shifting the result parallel to the real axis we reduce this case to
the one we have already considered.
For an arbitrary condenser C = (T0 × K0 , T1 × K1 ) the proof goes as follows.
Let v be a function in L(C) that vanishes on some open set U0 ⊃ T0 × K0 and is
equal to 1 on another open set U1 ⊃ T1 × K1 . Let τm be the partition of [−1, 1]
into equal intervals of length 2/m. Let Sjm denote the union of all the intervals in
τm that intersect Tj . If m is sufficiently large, then S0m ∩ S1m = ∅ and we have the
inclusions
Tj × Kj ⊂ Sjm × Kj ⊂ Uj , j = 0, 1.
74 Chapter 3. Special Transformations

Hence v ∈ L((S0m × K0 , S1m × K1 )). By the definition of the capacity and by what
we have proved,
I(v, C)  cap(S0m × K0 , S1m × K1 )
 cap G(S0m × K0 , S1m × K1 )  cap GC.
It remains to take the infimum over all functions v of the form described above.
Now let C be an admissible condenser such that equality holds in (3.7). Let
Hj = [aj , bj ] be the smallest interval containing Tj , j = 0, 1. If Hj \ Tj = ∅ for
some j (for instance, j = 1), then, first, we can assume that b0 < b1 and, second,
there exists a point x ∈ / T1 such that a1 < x and b0 < x < b1 . We perform
polarization Pα with respect to the line α = {z : Re z = (a1 + x)/2} codirectional
with the imaginary axis. Clearly, Pα C = C and Pα C is distinct from the condenser
symmetric to C relative to α. By Theorem 3.4 and by what we have proved,
cap C > cap Pα C  cap GPα C = cap GC,
which contradicts the assumed equality. Now if Hj = Tj , j = 0, 1, then we see from
the assumption of equality and from what we have proved that C must coincide
with the condenser GC or its reflection in the imaginary axis. 

Exercises 3.3
(1) Let C = (E0 , E1 ), where Ej = {reiϕ : ϕ ∈ Tj , r ∈ Kj }, j = 0, 1, and
T0 and T1 are disjoint nonempty closed subsets of an interval [0, ψ], where
0 < ψ  2π, and K0 and K1 are closed nonempty bounded subsets of the
interval 0 < r < ∞. Let G1 C = ({reiϕ : ϕ ∈ [ψ − mT0 , ψ], r ∈ K0 }, {reiϕ :
ϕ ∈ [0, mT1 ], r ∈ K1 }). For ψ  π show that
cap C  cap G1 C.

(2) Does the same inequality hold for ψ = 3π/2?


(3) Assume that the plates of a condenser C = (E0 , E1 ) lie in an interval [a−1 , a]
with a > 1, and let G2 C = ([a−1 , a−1 exp(lm E0 )], [a exp(−lm E1 ), a]), where

dr
lm e = , e ⊂ [a−1 , a].
e r
Prove the inequality
cap C  cap G2 C.
(4) (Tamrazov [T2]) For a condenser C = (E0 , E1 ) with plates on [−1, 1] con-
sider the condenser T C = (T − E0 , T + E1 ) with plates T − E0 = [−1, d0 ] and
T + E1 = [d1 , 1], where d0 and d1 are some numbers such that cap T − E0 =
cap E0 and cap T + E1 = cap E1 . Does the inequality
cap C  cap T C
hold?
3.4. Linear and radial transformations 75

3.4 Linear and radial transformations


The transformation below is a version of the extension of the Gonchar transfor-
mation to the case when [−1, 1] is replaced with an infinite interval. Interestingly
enough, in this case plates need not be Cartesian products: they can vary arbi-
trarily in the direction of the imaginary axis.
Now we turn to the statements. Let S be the horizontal strip 0 < Im z < 1.
We look at a condenser C = (S, {E0 , E1 }, {0, 1}) such that the set {Re z : z ∈ E0 }
is bounded above and {Re z : z ∈ E1 } is bounded below. The linear transformation
in the strip S takes C to the condenser LC = (S, {L− E0 , L+ E1 }, {0, 1}) with the
following plates:
L− E0 = {x + iy : x  lim [−c + m(E0 ∩ {z : Rez  −c, Imz = y})], 0  y  1},
c→+∞
+
L E1 = {x + iy : x  lim [c − m(E1 ∩ {z : Rez  c, Imz = y})], 0  y  1}.
c→+∞

F2 F3

M F 2 M, F 3

Figure 3.6.

The functions of c under the limit signs are monotonic. If the first limit is
equal to −∞, this means that the plate L− E0 contains the attainable boundary
point −∞ of the strip S, while if the second limit is +∞, then L+ E1 contains +∞.
If for some c > c0 the plates Ej of C contain the half-strips {z ∈ S : (−1)j+1 Re z 
c}, j = 0, 1, respectively, then starting from this c the functions under the limit
sign are constants, so the limit signs can be suppressed.
Sometimes we shall identify the condenser LC with the quadrilateral
(Q; a, b, c, d) such that Q is the interior of the field of LC, the vertices a and
b lie on the plate L− E0 , the vertices c and d lie on L+ E1 , Im a = Im d = 1, and
Im b = Im c = 0. Obviously,
cap LC = 1/M (Q; a, b, c, d).
Theorem 3.10 (cf. [M]). The following inequality holds:
cap C  cap LC. (3.8)
In addition, if C is an admissible condenser with connected field, then equality
holds in (3.8) if and only if C = LC.
76 Chapter 3. Special Transformations

Proof. We can assume that −∞ ∈ E0 and +∞ ∈ E1 . Otherwise cap LC = 0 and


the result is obvious. We partition the sets (−∞, +∞) and [0, 1] into equal intervals
{ xμ }μ and { yν }m ν=1 , respectively, and look at the corresponding decomposition
{ xμ × yν }μ,ν of the strip S. Similarly to the argument at the end of the proof of
Theorem 3.9, we see that to prove (3.8) it is sufficient to take a condenser C such
that each of its plates is the union of finitely many closed rectangles in the above
decomposition and an appropriate half-strip {z ∈ S : (−1)j+1 Re z  c}, where
c > 0 is sufficiently large. Within the strip Pν = {x + iy : −∞ < x < +∞, y ∈
yν } the condenser C has a similar structure to the condenser in Theorem 3.9.
The difference between them is the presence of unbounded half-strips and the
inequality c = 1. Now we look at the intersections of the plates of C with P1 .
Following the proof of Theorem 3.9, after performing finitely many polarizations
of C with respect to straight lines parallel to the imaginary axis and removing
some parts of the plates we obtain a condenser C1 = (S, {E01 , E11 }, {0, 1}) such
that
E0k ∩ Pν = (L− E0 ) ∩ Pν and E1k ∩ Pν = (L+ E1 ) ∩ Pν (3.9)
for k = ν = 1. Next we apply the same procedure ‘in the strip P2 ’ to the condenser
C1 , which gives us a condenser C2 = (S, {E02 , E12 }, {0, 1}), satisfying (3.9) for k =
ν = 2. Note that the polarizations of condensers with respect to lines codirectional
with the imaginary axis do not change the sets E01 ∩ P1 and E11 ∩ P1 . Hence (3.9)
holds for k = 2 and ν = 1, 2. Now we look at the strip P3 and so on, until
we obtain a condenser Cm = (S, {E0m , E1m }, {0, 1}) satisfying (3.9) for k = m
and ν = 1, 2, . . . , m. In other words, Cm = LC. Using Theorem 3.5 and the
monotonicity of the capacity we obtain

cap C  cap C1  cap C2  · · ·  cap Cm = cap LC.

Now let C be an admissible condenser such that S \ (E0 ∪ E1 ) is connected and


C = LC. We can assume that E1 = L+ E1 . Then there exist points z1 ∈ E1 and
z2 ∈
/ E1 such that 0  Im z1 = Im z2  1 and Re z1 < Re z2 . By Theorem 3.5 and
the above
cap C > cap Pα C  cap LPα C = cap LC,
where α is the line Re z = Re(z1 + z2 )/2 codirectional with the imaginary axis.
Thus we have no equality in (3.8). 
The linear transformation in an arbitrary (straight) strip S  is defined as the
composite Lϕ = ϕ−1 ◦ L ◦ ϕ, where ϕ is a linear map transforming S  into S. By
Theorem 3.10
cap C  cap Lϕ C
for each ϕ.
In 1964, Marcus defined the radial transformation which takes an open set
B, 0 ∈ B ⊂ C, to the starlike set

M B = {reiθ : 0  r < M (θ, B), 0  θ  2π},


3.4. Linear and radial transformations 77

where
⎛ ⎞

⎜ dr ⎟
⎠ , E(ρ, θ, B) := {r : re ∈ B, ρ  r < ∞}.

M (θ, B) = ρ exp ⎝
r
E(ρ,θ,B)
(3.10)
Here M (θ, B) is independent of ρ such that {z : |z|  ρ} ⊂ B.

2 2 2

C NC UOB C
3

B? < 3< 3
5 5 5

Figure 3.7: Left to right: the Marcus radial transformation and the
Szegő symmetrization.

In a similar way we define the radial transformation of a bounded closed set E:

M E = {reiθ : 0  r  M (θ, E), 0  θ  2π},

where this time we set


⎛ ⎞

⎜ dr ⎟
M (θ, E) = lim ρ exp ⎝ ⎠, ρ > 0. (3.11)
ρ→0 r
E(ρ,θ,E)

Now let C = (E0 , E1 ) be a condenser such that E0 contains z = 0 and E1 contains


the point at infinity. Let

M C = (M E0 , C \ M (C \ E1 )).

Theorem 3.11 ([M], [Mit3]). The following inequality holds:

cap C  cap M C.

In addition, if C is an admissible condenser with connected field, then equality


holds here if and only if C = M C.
Proof. Conceptually, this is similar to the proofs of Theorems 3.9 and 3.10. In place
of polarizations with respect to straight lines codirectional with the imaginary axis,
now we must use polarizations with respect to anticlockwise-oriented circles with
78 Chapter 3. Special Transformations

centre at the origin. Note also that a suitable branch r0 (z) of (1/(2π)) log z maps
the z-plane cut along the positive real half-axis conformally and univalently onto
the strip S. The radial transformation in the z-plane corresponds to the linear
transformation in the strip S:

M C = r0−1 (Lr0 (C)).

This observation suggests the sequence of polarizations which we must perform.



In addition, making a reference to Theorem 2.7, similarly to the proof of
Theorem 3.6 we arrive at the following results.
Corollary 3.3. For any open set B, 0 ∈ B ⊂ C,

r(B, 0)  r(M B, 0).

Moreover, if B is an admissible domain, then equality holds here if and only if


B = M B.
Corollary 3.4. For any compact set E ⊂ C

cap E  cap M E.

If C \ E is an admissible domain, then equality holds here if and only if E = M E.


The radial transformations of sets and condensers have the following general-
ization, which was proposed by Mityuk [Mit3]. Let rψ (z), where ψ ∈ (−π/2, π/2),
be the regular branch of the function ζ = (1/(2π cos ψeiψ )) log z mapping the
plane C cut along the spiral θ cos ψ − log r sin ψ = 0 (z = reiθ ) conformally and
univalently onto the strip S. We can define rψ (z) on the sides of the cut be means
of boundary correspondence. By the result of the spiral transformation (or ‘spi-
ral symmetrization’ in Mityuk’s terminology) of a bounded closed set E we mean
the set
−1
Mψ E = rψ (L− rψ (E)),
which is ψ-spiral-shaped set with respect to z = 0.
An open set B, 0 ∈ B ⊂ C, is transformed by the formula
−1
Mψ B = rψ (S \ L+ rψ (C \ B)).

Finally, the spiral transformation takes a condenser C = (E0 , E1 ), 0 ∈ E0 , ∞ ∈ E1


to the condenser
−1
Mψ C = rψ (Lrψ (C)).
If ψ = 0, then M0 = M is the radial transformation. For −π/2 < ψ < π/2 we
have the inequalities

r(B, 0)  r(Mψ B, 0), cap C  cap Mψ C.


3.5. Averaging transformations 79

The first inequality follows from the second, and the second can be proved using
either equimeasurable rearrangements of functions or polarizations.
Let m(θ, E) be the linear measure of the intersection of a compact set E ⊂ C
with the ray arg z = θ, 0  θ  2π. (Note that m(θ, E)  M (θ, E).) The linear
radial transformation with respect to the point z = 0 is the transition from E to
the set
LRE = {reiθ : 0  r  m(θ, E), 0  θ  2π}.
In some special cases Theorem 3.1 on contractions yields the inequality

cap E  cap LRE (3.12)

(see Exercises 3.1). Whether this inequality holds for arbitrary sets E is still an
open question.
In § 4.3 we present radial transformations of sets which are constructed using
the linear measure m(θ, E), but are distinct from LRE.

Exercises 3.4
(1) Show that the assumption in Theorems 3.10 and 3.11 that the condenser C
has a connected field is excessive.
(2) Prove Theorem 3.11 using polarizations with respect to circles.
(3) Define the radial transformation M of condensers C lying in a sector 0 <
arg z < ϕ (ϕ < 2π). Using Theorem 3.10 and the function (1/(2π)) log z
prove Theorem 3.11 for this transformation.
(4) Prove Corollaries 3.3 and 3.4.
(5) Prove that the area of a measurable set B is no less than the area of M B.

3.5 Averaging transformations


For P = {z : 0 < Re z < 1} we look at a family of condensers {Ck }nk=1 lying in the
vertical strip P and having the form Ck = (P, {Ek0 , Ek1 }, {0, 1}), with the plates

Ek0 = {x + iy : 0  x  1, y  yk0 (x)}


and
Ek1 = {x + iy : 0  x  1, y  yk1 (x)},
where the ykj (x) are some functions on [0, 1] such that −∞  yk0 (x) < yk1 (x) 
+∞, k = 1, . . . , n. Let A = {αk }nk=1 be a tuple of positive numbers such that
n
αk = 1. Then the linear averaging transformation in the strip P takes the
k=1
family {Ck }nk=1 to the condenser

LA {Ck }nk=1 = (P, {LA {Ek0 }nk=1 , LA {Ek1 }nk=1 }, {0, 1})
80 Chapter 3. Special Transformations

D5

D4

2 3 2 3
D3
3

MB gDl h5l?3 < B ? < 3< 3


5 5 5

Figure 3.8.

such that
 

n
LA {Ek0 }nk=1 = x + iy : 0  x  1, y  αk yk0 (x)
k=1
and  

n
LA {Ek1 }nk=1 = x + iy : 0  x  1, y  αk yk1 (x) ,
k=1

where y can take infinite values.


Theorem 3.12 (cf. [M]). The following inequality holds:


n
αk cap Ck  cap LA {Ck }nk=1 . (3.13)
k=1

In addition, if all the condensers Ck , k = 1, . . . , n, are admissible, then equality


in (3.13) holds if and only if each of them can be obtained from LA {Ck }nk=1 by a
translation parallel to the imaginary axis.
We preface the proof of Theorem 3.12 with the following result.
Lemma 3.2. Assume that a condenser C = (P, {E0 , E1 }, {0, 1}) has plates defined
by the relations E0 = {x + iy : 0  x  1, y  y0 (x)} and E1 = {x + iy : 0  x 
1, y  y1 (x)}, where y0 (x) ≡ −∞ and y1 (x) ≡ +∞; let u be the potential function
of C. Then each level curve γt := {z ∈ P : u(z) = t}, 0 < t < 1, intersects each
line Re z = x, 0  x  1, in a unique point.
Proof. Under the assumptions of the lemma, u coincides with the potential func-
tion of the corresponding quadrilateral, so it is well defined. We can give a standard
proof of Lemma 3.2 using methods in function theory (cf. [Gol], Ch. IV, § 5). How-
ever, an approach based on capacities is also of interest. Fix some t, 0 < t < 1. The
3.5. Averaging transformations 81

analytic curve γt decomposes P into open sets P0 and P1 (let Ej ⊂ Pj , j = 0, 1). We


look at the condensers C0 = (P0 , {E0 , P 1 }, {0, t}) and C1 = (P1 , {P 0 , E1 }, {t, 1}).
If Lemma 3.2 fails, then Lϕ C0 = C0 and Lϕ C1 = C1 , where ϕ(z) = i(1 − z). By
Theorem 3.10

cap C = cap C0 + cap C1 > cap Lϕ C0 + cap Lϕ C1 ,

and this contradicts Theorem 1.18, which claims that

cap Lϕ C0 + cap Lϕ C1  cap C. 

Proof of Theorem 3.12. We can assume that each condenser Ck has a potential
function uk = uk (z) ≡ uk (x, y), k = 1, . . . , n. We define functions yk = yk (x, t),
y ∗ = y ∗ (x, t), and u∗ = u∗ (z) ≡ u∗ (x, y) by the relations

uk (x, yk (x, t)) = t, 0  x  1, 0 < t < 1, k = 1, . . . , n,


n
y ∗ (x, t) = αk yk (x, t), 0  x  1, 0 < t < 1,
k=1
u∗ (x, y ∗ (x, t)) = t, 0  x  1, 0 < t < 1,

0, (x, y) ∈ LA {Ek0 }nk=1 ,
u∗ (x, y) =
1, (x, y) ∈ LA {Ek1 }nk=1 .

By Lemma 3.2 the harmonic function ∂uk /∂y is nonnegative on Qk := {z ∈


P : 0 < uk (z) < 1}, so that ∂uk /∂y > 0 in Qk . The function yk (x, t) is well
defined, so in view of the implicit function theorem, we have
5 5
∂uk ∂yk ∂yk ∂uk ∂yk
=− and =1
∂x ∂x ∂t ∂y ∂t
in Qk . In a similar way
5 5
∂u∗ ∂y ∗ ∂y ∗ ∂u∗ ∂y ∗
=− and =1
∂x ∂x ∂t ∂y ∂t
on the set Q∗ := {z ∈ P : 0 < u∗ (z) < 1}. The function u∗ (x, y) is continuous in
P , Lipschitz continuous on compact subsets of Q∗ , vanishes on LA {Ek0 }nk=1 and
is equal to 1 on LA {Ek1 }nk=1 . By the Dirichlet principle

cap LA {Ck }nk=1  I(u∗ , Q∗ ).



n
On the other hand, setting yj∗ (x) = αk ykj (x), j = 0, 1, we obtain
k=1

1 (x) 2 

1 y 2 
∗ ∗ ∂u∗ ∂u∗
I(u , Q ) = dx + dy
∂x ∂y
0 y0∗ (x)
82 Chapter 3. Special Transformations

⎡
2 ⎤
∂y ∗
1 ⎢ ∂x1
⎢ 1 ⎥ ⎥
= dx ⎢ + ⎥ dt
⎣ ∂y ∗ ∂y ∗ ⎦
0 0
∂t ∂t
⎡  2 ⎤
n ∂yk
1 1 ⎢ αk ⎥
⎢ ∂x 1 ⎥
= dx ⎢ k=1 + ⎥ dt
⎣ n ∂y k n ∂y k ⎦
0 0 α k αk
k=1 ∂t k=1 ∂t
⎡  2 ⎤
∂y k
1 1 ⎢
αk ⎥
n αk n
⎢ ∂x ⎥
 dx ⎢ + ⎥ dt
⎣ ∂yk ∂yk ⎦
0 0 k=1 k=1
∂t ∂t

n 1  (x)
yk1 2  2 
∂uk ∂uk
= αk dx + dy
∂x ∂y
k=1 0 yk0 (x)

n
n
= αk I(uk , Qk ) = αk cap Ck .
k=1 k=1
  
Here we have used Cauchy’s inequality ( ab)2  ( a2 )( b2 ) twice. We have
equality sign in the last inequality only for

∂yk ∂yl ∂yk ∂yl


= and = , k, l = 1, . . . , n.
∂t ∂t ∂x ∂x
Hence equality in (3.13) only holds for yk = yl + const, k, l = 1, . . . , n. 

Now let {Bk }nk=1 be a family of domains in C which are starlike with respect
to the origin, so that if a point z lies in such a domain, than the whole of the line
segment [0, z] also does. Let A = {αk }nk=1 be a tuple of positive numbers such
n
that αk = 1. By the result of the radial averaging transformation of the family
k=1
{Bk }nk=1 we mean the domain
 n

RA {Bk }nk=1 = re iθ
:0r< [m(θ, Bk )] αk
, 0  θ  2π ,
k=1

where m(θ, Bk ) is the linear measure of the intersection of Bk with the ray arg z =
θ. The radial averaging transformation of a family {Ek }nk=1 of starlike closed sets
Ek with respect to the origin is defined in a similar way (with strict inequality
sign replaced with nonstrict one). Let {Ck }nk=1 be a family of condensers Ck =
(Ek0 , Ek1 ) such that Ek0 and C \ Ek1 are starlike sets in C with respect to z = 0.
3.5. Averaging transformations 83

Then the radial averaging transformation assigns to {Ck }nk=1 the condenser


RA {Ck }nk=1 = RA {Ek0 }nk=1 , C \ RA {C \ Ek1 }nk=1 .

Theorem 3.13 (Marcus [M]). 5 In the above notation,



n
αk cap Ck  cap RA {Ck }nk=1 . (3.14)
k=1

In addition, if both plates in each condenser Ck , k = 1, . . . , n, are distinct from a


point, then equality in (3.14) holds if and only if the condensers Ck , k = 1, . . . , n,
coincide with RA {Ck }nk=1 up to dilations with centre z = 0.
Proof. If one of the plates in some condenser is a point, the both sides of (3.14)
vanish. If not, we let r denote the single-valued branch of the function (i/(2π)) log z
mapping the plane C cut along the positive real half-axis onto the strip P and
defined on the sides of the cut by means of boundary correspondence. Let uk =
ωk ◦ r−1 , where ωk is the potential function of the condenser Ck , k = 1, . . . , n. Let
Qk , Q∗ , and u∗ be the sets and function in P constructed from the functions uk ,
k = 1, . . . , n, as in the proof of Theorem 3.12, and let ω ∗ = u∗ ◦ r (similarly to
Lemma 3.2 we can show that the level curves of the ωk are ‘starlike’, meaning here
that they are boundaries of starlike domains). Repeating the proof of Theorem 3.12
we obtain

n
n
n
I(ω ∗ , C) = I(u∗ , Q∗ )  αk I(uk , Qk ) = αk I(ωk , C) = αk cap Ck .
k=1 k=1 k=1

The function ω ∗ is continuous in C, Lipschitz-continuous on compact subsets of


r−1 (Q∗ ), vanishes on RA {Ek0 }nk=1 and is equal to 1 on C \ RA {C \ Ek1 }nk=1 . By
the Dirichlet principle
I(ω ∗ , C)  cap RA {Ck }nk=1 .
The proof of (3.14) is complete. If we have equality in (3.14), then we must also
have equality in the preceding relation. Hence uk = ul ◦ ϕkl , where ϕkl is a trans-
lation parallel to the imaginary axis, or equivalently, ωk = ωl ◦ ψkl , where ψkl is a
dilation with centre at z = 0. 
Theorem 3.14 (Marcus [M]). Let {Bk }nk=1 be a family of starlike domains in C
with respect to the origin. Then for any tuple of positive numbers A = {αk }nk=1
n
such that αk = 1,
k=1

n
(r(Bk , 0))αk  r(RA {Bk }nk=1 , 0). (3.15)
k=1
5 Conditions for equality in Theorems 3.13 and 3.14 were found by Mityuk and Shlyk [Mit3].
84 Chapter 3. Special Transformations

If all the Bk are distinct from C, then equality holds in (3.15) is and only if all
these domains are equal up to dilations with centre at z = 0.
Proof. We can assume that the domains Bk are distinct from the whole plane and
look at the condensers
Ck (r) = (U (0, r), C \ Bk ), k = 1, . . . , n,
and
RA {Ck (r)}nk=1 = (U (0, r), C \ RA {Bk }nk=1 ).
By formula (2.10)
 2
2π 1
cap Ck (r) = − − 2π log r(Bk , 0)
log r log r
 2 
1
+o , r → 0, k = 1, . . . , n,
log r
 2
2π 1
cap RA {Ck (r)}k=1 = −
n
− 2π log r(RA {Bk }k=1 , 0)
n
log r log r
 2 
1
+o , r → 0.
log r

We deduce (3.15) from Theorem 3.13. Now assume that we have equality in (3.15).
Let {Bk , Ck } denote the decomposition of Bk with respect to the ‘tuples’ {0}
and {1}, k = 1, . . . , n corresponding to the partition {0, t}. Since |Ck | = t/(2π),
0 < t < ∞, k = 1, . . . , n, using Theorem 3.13 again we obtain

n
t 1 1
αk |Ck | = =   = |RA {Ck }nk=1 |. (3.16)
2π n
cap RA {Ck }nk=1
k=1 αk cap Ck
k=1

Now from Theorem 2.7, inequality (3.15) for the family {Bk }nk=1 , inequality (3.16),
and Theorem 2.7 again we obtain in succession

n
αk
n
αk
n

log r(Bk , 0) = log r(Bk , 0) + αk |Ck |
2π 2π
k=1 k=1 k=1
1
 log r(RA {Bk }nk=1 , 0) + |RA {Ck }nk=1 |

1
 log r(RA {Bk }nk=1 , 0).

Hence we have equality in (3.16), and by Theorem 3.13 the condensers Ck can
be obtained from RA {Ck }nk=1 by dilations with centre at the origin, k = 1, . . . , n.
Hence the domains Bk coincide up to dilations. It is easy to see that this condition
is also sufficient for equality in (3.15). 
3.5. Averaging transformations 85

Corollary 3.5. If Ek , k = 1, . . . , n, are nondegenerate continua which are starlike


with respect to z = 0, then
n
(cap Ek )αk  cap RA {Ek }nk=1
k=1


n
for any tuple of positive numbers A = {αk }nk=1 such that αk = 1. Equality holds
k=1
here if and only if the Ek are equal to RA {Ek }nk=1 up to dilations with centre at
the origin.
Proof. This follows from (2.1) and the previous result. 
In conclusion we note that if {Ck }nk=1 is a family of condensers

Ck = (Ek0 , Ek1 ), where 0 ∈ Ek0 and ∞ ∈ Ek1 ,

then the family {M Ck }nk=1 satisfies the assumptions of Theorem 3.13. In view of
Theorem 3.11,

n
n
αk cap Ck  αk cap M Ck  cap RA {M Ck }nk=1 .
k=1 k=1

D4

D5 D3

SB gN Dl h5l?3 < B ? < 3< 3


5 5 5 .

Figure 3.9.

Usually, the condenser RA {M Ck }nk=1 is also called the result of the radial
averaging transformation of the family of condensers {Ck }nk=1 (see Figure 3.9). In
a similar way, making only slight modifications, we can deal with the composite
of the linear transformation and linear averaging transformation.
86 Chapter 3. Special Transformations

Exercises 3.5
(1) Concerning the definition of the linear averaging transformation, can we say
that arbitrary functions ykj (x), with yk0 (x) < yk1 (x) for 0  x  1, define
some condensers Ck , k = 1, . . . , n?
(2) Show that the sets LA {Ekj }nk=1 , j = 0, 1, are closed in [P].
(3) Let {Qk }nk=1 be a family of quadrilaterals lying in a sector 0 < arg z < ϕ,
ϕ < 2π, such that two opposite sides of each quadrilateral lie on the sides of
the sector and the other two are given by equations r = rk1 (θ) and r = rk2 (θ)
(0 < rk1 (θ) < rk2 (θ) < ∞), 0  θ  ϕ, k = 1, . . . , n. Let A = {αk }nk=1 be a
n
tuple of positive numbers such that αk = 1. We let RA {Qk }nk=1 denote
k=1
a quadrilateral of the form described above, but with two sides given by
1
n 1n
r = αk
rk1 (θ) and r = αk
rk2 (θ), 0  θ  ϕ, and let M (·) denote the
k=1 k=1
modulus of a quadrilateral with respect to the sides defined by equations.
Show that
n
αk M (Qk )  M (RA {Qk }nk=1 ).
k=1

Describe all the equality cases in this relation.


(4) Let {Qk }nk=1 be a family of quadrilaterals in the annulus ρ < |z| < R such
that two opposite sides of each quadrilateral lie on the circles |z| = ρ and
|z| = R, and the other two sides are defined by equations θ = θk1 (r) and
θ = θk2 (r) (θk1 (r) < θk2 (r)), ρ  r  R, k = 1, . . . , n. Let A = {αk }nk=1
be as in Exercise 3, let ΘA {Qk }nk=1 denote a quadrilateral of the same form,
n 
n
with two sides given by θ = αk θk1 (r) and θ = αk θk2 (r), ρ  r  R,
k=1 k=1
and let M (·) be the modulus of a quadrilateral with respect to the sides
defined by equations. Prove the inequality


n
αk M (Qk )  M (ΘA {Qk }nk=1 )
k=1

and examine the equality cases in it.


(5) Let {Ck }nk=1 be a family of condensers Ck = (Ek0 , Ek1 ) in C such that the
sets Ek0 and C \ Ek1 are ψ-spiral-shaped (−π/2 < ψ < π/2) relative to z = 0
(which means that together with any point they also contain an arc of some
spiral θ cos ψ − log r sin ψ = c, where 0  c < 2π cos ψ, which has the end-
points at this point and the origin). Following Mityuk we define the result of
the spiral averaging transformation of the family of condensers as follows:
−1
Mψ {Ck }nk=1 = rψ (LA {
rψ (Ck )}nk=1 ),
3.5. Averaging transformations 87

where rψ (z) = irψ (z) + 1 and the function rψ (z) was defined at the end of
§ 3.4. Show that Mψ {Ck }nk=1 is a condenser. Following the proof of Theorem
3.13 show that
n
αk cap Ck  cap Mψ {Ck }nk=1 .
k=1

(6) Let {Bk }nk=1 be a family of domains in a sector 0 < arg z < ϕ, ϕ < 2π, each of
which is bounded by two intervals of the sides of the sector with initial point
at the origin and a curve Γk defined by an equation r = rk (θ) > 0, 0  θ  ϕ,
n
k = 1, . . . , n. For a tuple of positive numbers A = {αk }nk=1 , αk = 1, let
k=1
RA {Bk }nk=1 be a domain of a similar form bounded by two line intervals and
1
n
a curve Γ : r = rkαk (θ), 0  θ  ϕ. Prove the inequality
k=1

n
(r(Bk , Γk , 0))αk  r(RA {Bk }nk=1 , Γ, 0)
k=1

and show that equality holds here only when the domains Bk are equal to
RA {Bk }nk=1 up to a dilation with centre z = 0.
(7) Using the previous exercise show that, in the class of rectilinear triangles B
with fixed area and fixed acute angle at the vertex z = 0, the maximum
Robin radius r(B, Γ, 0) is attained only for an isosceles triangle with equal
sides at the vertex z = 0. Here Γ is the opposite side to z = 0.
(8) Let {Ck }nk=1 be a family of condensers in the horizontal strip S which satisfy
the assumptions required for the definition of the linear transformation L
(§ 3.4). Construct the linear averaging transformation of {Ck }nk=1 as a com-
posite of the linear transformation L, rotations about the origin, and the
linear averaging transformation LA , where A = {αk }nk=1 is a tuple of posi-
n
tive numbers such that αk = 1. Write out an inequality connecting the
k=1
capacities of the Ck and the resulting condenser. Describe all the equality
cases in it. Using this result, prove the inequality for the radial averaging
transformation RA presented before Exercises 3.5.
Chapter 4

Symmetrization

The result of the symmetrization of a set or condenser A is usually denoted in


the literature by an asterisk A∗ . To distinguish between different symmetrization
methods we shall use the notation St A, Cr A, RA, and so on. If Sym denotes
symmetrization of open (or closed) subsets of C, then by the symmetrization of a
condenser C = (E0 , E1 ) we mean the transformation taking it to the condenser

Sym C = (SymE0 , C \ Sym(C \ E1 )). (4.1)

The transformations below are linked to classical coordinate systems and in a


certain sense are limiting cases of polarization.

4.1 Symmetrization along straight lines or circles


We set P = P (a, b) := {x + iy : a < x < b}, −∞  a < b  +∞ and let l(x)
denote the vertical line Rez = x, −∞ < x < +∞. Let B be a (relatively) open
subset of P . By the Steiner symmetrization of B with respect to the real axis we
mean the transformation of this set into the mirror-symmetric set StB defined by

St B = {x + iy : B ∩ l(x) = ∅, 2|y| < m(B ∩ l(x))},

where m(·) is linear Lebesgue measure. For a closed set E ⊂ P its Steiner sym-
metrization is

St E = {x + iy : E ∩ l(x) = ∅, 2|y|  m(E ∩ l(x))}.

The reader can verify that St B is an open and St E is a closed subset of P .


Moreover, if B is a domain, then St B is also a domain (cf. [H], p. 115). The
Steiner symmetrization in the set ϕ(P ) obtained from P by a motion ϕ is defined
as the composite transformation ϕ ◦ St ◦ϕ−1 .

© Springer Basel 2014 89


V.N. Dubinin, Condenser Capacities and Symmetrization in Geometric Function
Theory, DOI 10.1007/978-3-0348-0843-9_4
90 Chapter 4. Symmetrization

F3 Q o Uv)Q o F3 *

F2 Uv F2

F3
Q o Uv)Q o F3 *

Figure 4.1: The Steiner symmetrization.

Now let C = (P, {E0 , E1 }, {0, 1}) be an arbitrary condenser such that E1
contains the points at infinity. Then we set

St C = (P, {St E0 , P \ St(P \ E1 )}, {0, 1}).

The following symmetrization principle goes back to Pólya and Szegő [PS].
Theorem 4.1. The following inequality holds:

cap C  cap St C. (4.2)

In addition, if C is an admissible condenser with connected field, then equality


holds in (4.2) if and only if this condenser coincides with St C up to a translation
along the imaginary axis.
Proof. We partition [a, b] and (−∞, +∞) into equal subintervals { xμ }μ and
{ yν }ν , respectively, and look at the corresponding decomposition = { xμ ×
yν }μ,ν of the strip P . Assume that the plate E0 of C consists of finitely many
closed rectangles from this decomposition, while E1 is the union of finitely many
rectangles in , two strips (or half-planes) of the form {z ∈ P : Im z  c1 } or
{z ∈ P : Im z  c2 } tiled by rectangles in , and a half-plane Re z  c3 (if
a = −∞) or Re z  c4 (if b = +∞; both the half-planes must also be tiled by
rectangles in ). Let Pμ := {x + iy ∈ P : x ∈ xμ }, μ = 1, . . . , n, be the strips
corresponding to the partition {xμ }μ and intersecting the field of C. To each strip
Pμ we assign a fixed line lμ = l(xμ ), where xμ is an interior point in the interval
xμ , μ = 1, . . . , n. Now we look at the strip P1 . Assume that, apart from two
rays, the intersection l1 ∩ E1 also contains finite intervals. Let z1 be the upper
end-point of the top interval in l1 ∩ E1 and z2 be the initial point of the downward
ray in l1 ∩ E1 . Making polarization of the condenser C with respect to the line
|z − z1 | = |z − z2 | co-directed with the real axis, and then removing the degenerate
4.1. Symmetrization along straight lines or circles 91

rectangles (points and line intervals) from the plates of the resulting condenser
we obtain a condenser C 1 = (P, {E01 , E11 }, {0, 1}) with the following properties.
Its plate E01 consists of finitely many rectangles; the other plate E11 consists of
rectangles, half-strips and perhaps also half-planes of the form indicated above,
the number of rectangles in these plates is no greater than the corresponding
number for C, and the number of rectangles in E11 which lie in P 1 is at least one
less than the number of rectangles in E1 which lie in P 1 . In addition,
m(E0τ ∩ lμ ) = m(E0 ∩ lμ ) and m((Pμ \ E1τ ) ∩ lμ ) = m((Pμ \ E1 ) ∩ lμ ), (4.3)
for τ = 1 and μ = 1, . . . , n. We described similar structures in detail when we
were discussing Gonchar transformations (see the proof of Theorem 3.9). Now we
repeat the above argument for C replaced by C 1 and so on, until in finitely many
(say, r) steps we obtain a condenser C r = (P, {E0r , E1r }, {0, 1}) such that (4.3)
holds for τ = r. The number of rectangles in the plates of C r is no greater than
the corresponding number for C 1 , and the intersection l1 ∩ E1r is the union of two
rays. Assuming that l1 ∩ E0r does not reduce to a single line interval, let z1r be
the lower end-point of the top interval in l1 ∩ E0r and z2r be the lower end-point
of the bottom interval in l1 ∩ E0r . Making polarization of C r with respect to the
line |z − z1r | = |z − z2r | co-directed with the real axis, removing the degenerate
components of the plates and then making similar polarization takes us in finitely
many steps to a condenser C s = (P, {E0s , E1s }, {0, 1}) satisfying (4.3) for τ = s
and such that the intersection l1 ∩ E0s is a line interval. Let z1s be the mid-point
of the line segment l1 ∩ E0s and z2s the mid-point of l1 ∩ (P1 \ E1s ). Polarizing the
condenser C s with respect to the line |z − z1s | = |z − z2s | co-directed with the real
axis, and removing the degenerate components of the plates we obtain a condenser
C1 whose plates intersect l1 in sets having a common centre of symmetry: one of
these intersections is a compact line segment and the other consists of two disjoint
rays. If l1 ∩ E0r = ∅, then we set C1 = C r . It follows from the construction of C1
that
St C1 = St C.
Since the capacity is monotonic, from Theorem 3.5 we obtain
cap C  cap C1 .
Next we perform in succession similar transformations of the strips P2 , . . . , Pn .
Note that the polarization with respect to the straight line assigned to the strip
P2 and co-directed with the real axis does not affect the symmetry properties of
the parts of the plates which lie in P1 , and so on. At the last step we obtain a
condenser Cn = (P, {E0 (n), E1 (n)}, {0, 1}) with the following property: for each
straight line lμ the intersection lμ ∩E1 (n) consists of two rays, and the intersection
lμ ∩ E0 (n) is empty or equal to a closed interval with the same centre of symmetry
as lμ ∩ E1 (n), μ = 1, . . . , n, and furthermore,
St Cn = St C,
cap C1  cap Cn .
92 Chapter 4. Symmetrization

Let ζ1 be the centre of symmetry of l1 ∩ E1 (n) and ζ2 the centre of symmetry of


l2 ∩ E1 (n). Polarizing the condenser Cn with respect to the line |z − ζ1 | = |z − ζ2 |
we obtain a condenser Cn+1 such that the parts of its plates which lie in P1 and P2
have the same axis of symmetry. Performing the same operation with the strips P2
and P3 we obtain a condenser Cn+2 such that the parts of its plates which lie in P1 ,
P2 , and P3 have the same axis of symmetry and so on, until we obtain a condenser
C2n−1 with axis of symmetry, which coincides with St C up to a translation along
the imaginary axis; furthermore,

cap C  cap Cn  cap C2n−1 = cap St C.

For an arbitrary condenser C = (P, {E0 , E1 }, {0, 1}) with E1  ±i∞, in-
equality (4.2) can be proved as follows. Let v be an admissible function for C,
which is equal to 0 in a neighbourhood U0 of the plate E0 and to 1 in a neighbour-
hood U1 of the plate E1 . Let Sj be the union of all the rectangles in the decom-
position that intersect Ej , j = 0, 1. If is sufficiently fine, then S0 ∩ S1 = ∅
and
Ej ⊂ Sj ⊂ Uj , j = 0, 1.
Hence v is an admissible function for the condenser (P, {S0 , S1 }, {0, 1}). By what
we have already proved,

I(v, P )  cap(P, {S0 , S1 }, {0, 1})


 cap St(P, {S0 , S1 }, {0, 1})  cap St C.

Using Lemma 1.2 we obtain (4.2).


Now let C be an admissible condenser with connected field, and assume that
we have equality sign in (4.2) but C does not coincides with the condenser St C
up to a translation along the imaginary axis. Then at least one of the following
results holds:
a) there exists a point x ∈ (a, b) such that E0 ∩ l(x) = ∅ and this set is not an
interval;
b) there exists a point x ∈ (a, b) such that (P \ E1 ) ∩ l(x) = ∅ and this set is
not an interval;
c) for some x ∈ (a, b) the intersection E0 ∩l(x) is a closed interval with centre of
symmetry z1 , the intersection (P \ E1 ) ∩ l(x) is a closed interval with centre
of symmetry z2 , and z1 = z2 ;
d) for some x1 , x2 ∈ (a, b), x1 = x2 , the sets (P \ E1 ) ∩ l(xk ) are open intervals
with centres of symmetry zk , k = 1, 2, such that Im z1 = Im z2 .
In all cases a)–d) we can easily indicate a straight line α such that the polarization
of C with respect to α in accordance with Theorem 3.5 strictly reduces its capacity.
For example, in cases c) and d) we can take for α the straight line of points
equidistant from z1 and z2 , which is codirectional with the real axis. In cases a)
4.1. Symmetrization along straight lines or circles 93

and b) such a line is chosen similarly to the proofs of Theorems 3.9 and 3.10. As
a result, we obtain a contradiction:
cap C > cap Pα C  cap St(Pα C) = cap St C. 
Corollary 4.1. Let B be an open subset of the plane C possessing a Green function;
let Z = {zk }nk=1 be a tuple of points in B with pairwise distinct values of the real
parts Re zk ; let = {δk }nk=0 be a set of real numbers such that δ0 = 0 and δk > 0
for k = 1, . . . , n. Finally, let Ψ = {ψk (r)}nk=1 , where ψk (r) = μk rνk , for some
positive μk and νk , k = 1, . . . , n, n  1. Then
M (B, ∂B, Z, , Ψ)  M (St B, ∂(St B), {Re zk }nk=1 , , Ψ).
In addition, if B is an admissible domain, then equality sign holds if and only if
B = {z : z − ic ∈ St B} and Im zk = c, k = 1, . . . , n, where c is an arbitrary real
constant.
Proof. We set  = {0, 1, . . . , 1} and Ψ = {μk rνk /δk }nk=1 . From (2.16) we conclude
that
 n 2 *  n 2

M (B, ∂B, Z, , Ψ) = δk /νk δk2 /νk M (B, ∂B, Z,  , Ψ )
k=1 k=1

and
M (St B, ∂(St B), {Re zk }nk=1 , , Ψ)
 n 2 *  n 2

= δk /νk δk /νk M (St B, ∂(St B), {Re zk }nk=1 ,  , Ψ ).
2

k=1 k=1

The proof of the inequality


M (B, ∂B, Z,  , Ψ )  M (St B, ∂(St B), {Re zk }nk=0 ,  , Ψ ),
and the description of the equality cases are virtually the same as in the proof of
Theorem 3.6, where we must take the Steiner symmetrization St in place of the po-
larization Pα− and the points zk (α) must be replaced with Re zk , k = 1, . . . , n. 
Corollary 4.2. For each compact subset E of C
cap E  cap St E, (4.4)
and for compact subsets E of the disc U
caph E  caph St E. (4.5)
In addition, if C \ E (U \ E) is an admissible domain, than equality in (4.4) holds
if and only if E coincides with St E up to a translation along the imaginary axis,
while equality in (4.5) holds only if E = St E.
The verification of this result can be left to the reader.
94 Chapter 4. Symmetrization

F3 L o Et)L o F3 *

F2
Et F2

Figure 4.2: Circular symmetrization.

The most popular method for symmetrization in the plane is Pólya’s circular
symmetrization. We consider this transformation for condensers in an annulus K,
where K = K(ρ, R) := {z : ρ < |z| < R}, 0  ρ < R  ∞. If ρ = 0 (R = ∞),
then we assume that the point 0 (∞, respectively) belongs to K. Let γ(r) denote
the circle |z| = r for 0 < r < ∞, let γ(0) := 0 and γ(∞) := ∞. Let B be an open
subset of K. Circular symmetrization with respect to the positive real half-axis
assigns to a set B the ‘circularly symmetric’ set

Cr B = {reiθ : B ∩ γ(r) = ∅, 2|θ|r < m(B ∩ γ(r))} ∪ {−r : γ(r) ⊂ B},

where, as before, m(·) is Lebesgue linear measure and m(∞) := ∞. In a similar


way we define symmetrization of a closed subset E of K:

Cr E = {reiθ : E ∩ γ(r) = ∅, 2|θ|r  m(E ∩ γ(r))}.

By definition, circular symmetrization with respect to an arbitrary ray L is


the composite transformation ϕ−1 ◦ Cr ◦ϕ, where ϕ is the motion of C taking the
ray L to the positive real half-axis.
For a condenser C = (K, {E0 , E1 }, {0, 1}) we set

Cr C = (K, {Cr E0 , K \ Cr(K \ E1 )}, {0, 1}).

Theorem 4.2 ([P]). 6 In the above notation,

cap C  cap Cr C.

In addition, if C is an admissible condenser with connected field, then equality


holds if and only if this condenser coincides with Cr C up to a rotation about the
origin.
6 Statements on equality signs in Theorems 4.1 and 4.2 were first obtained by Jenkins [J] and

also Ohtsuka [Oht1]. A more general case was considered by Shlyk [Shl].
4.1. Symmetrization along straight lines or circles 95

Proof. This is similar to the case of the Steiner symmetrization, but is in fact
slightly simpler because the plates of the condenser are ‘on equal footing’. In place
of the rectangles forming the plates of the condenser C in the proof of Theorem 4.1
we must take circular rectangles constructed with the use of the polar coordinate
grid and must perform polarizations with respect to straight lines through the
origin (Figure 4.3). 

4 4
3 3
F3 F2

D ? )F2 < F3 * Q 3 D Q 4 Q 3 D

Figure 4.3.

Corollary 4.3. Let B be an open subset of C possessing a Green function; let


Z = {zk }nk=1 be a system of point in B with pairwise distinct absolute values;
let = {δk }nk=0 be a tuple of real numbers such that δ0 = 0 and δk > 0 for
k = 1, . . . , n, and let Ψ = {ψk (r)}nk=1 , where ψk (r) = μk rνk for some positive μk
and νk , k =, . . . , n, n  1. Then

M (B, ∂B, Z, , Ψ)  M (Cr B, ∂(Cr B), {|zk |}nk=1 , , Ψ).

In addition, if B is an admissible domain, then equality holds here if and only


if B = {z : ze−iψ ∈ Cr B} and all the points zk = 0 have the same argument
arg zk = ψ, 1  k  n, where ψ is some real number.
Corollary 4.4. For each compact set E ⊂ C

cap E  cap Cr E,

and if C \ E is an admissible domain, then equality holds here if and only if E


coincides with Cr E up to a rotation about the origin.
Now we introduce some further notation: Ω = Ω(a) := {z : 0 < arg z < a}
with 0 < a < 2π, Ω(2π) = C, L(θ) is the ray arg z = θ; for an open or closed
subset A of Ω we set
⎛ ⎞

⎜1 dr ⎟
R(A ∩ L(θ)) = exp ⎝ ⎠.
2 r
A∩L(θ)
96 Chapter 4. Symmetrization

TF

Figure 4.4: Symmetrization with respect to a circle.

By symmetrization of a (relatively) open subset B of Ω with respect to the circle


|z| = 1 we shall mean the transformation of B into the set
RB = {reiθ : B ∩ L(θ) = ∅, R−1 (B ∩ L(θ)) < r < R(B ∩ L(θ))},
which is mirror-symmetric relative to |z| = 1. Symmetrization of a closed subset
E of Ω is similarly defined, except that strict inequality signs must be replaced
by nonstrict ones. Let r(z) be a regular branch of ζ = (i/(2π)) log z that takes
the plane C cut along the positive real half-axis to the strip 0 < Re ζ < 1 and is
defined on the sides of the cut by means of boundary correspondence. Then it is
easy to see that
RA = r−1 (St r(A)) (4.6)
for each subset A of Ω. Here St is the Steiner symmetrization in the strip 0 
Re ζ  a/(2π). Let C = (Ω, {E0 , E1 }, {0, 1}) be a condenser whose plate E1
contains the origin and the points at infinity. By definition we set
RC = (Ω, {RE0 , Ω \ R(Ω \ E1 )}, {0, 1}).
Theorem 4.3. The capacity of the condenser RC does not exceed the capacity of
C. Furthermore, in the case of admissible condensers with connected field cap C =
cap RC if and only if the condensers C and RC coincide up to a dilation with
centre at z = 0.
Proof. For a = 2π this follows from the representation (4.6) and Theorem 4.1. If
a = 2π (Ω = C), then the proof is similar to that of Theorem 4.1, but in place of
the rectangles forming the plates of C we must take circular rectangles constructed
with the use of the polar coordinate grid and perform polarization with respect to
circles with centre at the origin. 
Corollary 4.5. If B is an admissible domain in C \ {0}, then
r(B, z0 )  r(RB, z0 /|z0 |)|z0 |
4.1. Symmetrization along straight lines or circles 97

for each point z0 ∈ B. Equality holds here only for B = {z : z/t ∈ RB} and
|z0 | = t, where t is an arbitrary positive number.
Theorem 4.4. For an arbitrary open set B ⊂ C, B = C, which contains 0 and ∞,

M (B, ∂B, {0, ∞}, {0, 1, (−1)k}, {r, r})  M (G, ∂G, {0, ∞}, {0, 1, (−1)k}, {r, r}),
(4.7)
k = 1, 2, where G = C \ R(C \ B).
Proof. We can assume that B is admissible. For k = 2, (4.7) can be proved in the
standard way. In fact, by Theorem 4.3,

cap C(r; B, ∂B, {0, ∞}, {0, 1, 1}, {r, r})


= cap(C, {C \ B, U (0, r) ∪ U (∞, r)}, {0, 1})
 cap(C, {R(C \ B), U (0, r) ∪ U (∞, r)}, {0, 1})
= cap C(r; G, ∂G, {0, ∞}, {0, 1, 1}, {r, r}).

It remains to use the definition of a reduced modulus.


For k = 1 we require formula (2.16), which shows that

8πM (B, ∂B, {0, ∞}, {0, 1, ±1}, {r, r}) = log[r(B, 0)r(B, ∞)] ± 2gB (0, ∞). (4.8)

The function u(z) = gB (z, 0) − gB (z, ∞) is harmonic on the set B \ {0, ∞}, and
we have u(z) → +∞ as z → 0 and u(z) → −∞ as z → ∞. The Green function of
B1 := {z : u(z) > 0} with pole at z = 0 coincides with u(z) in B1 . Hence

log r(B1 , 0) = lim (u(z) + log |z|) = log r(B, 0) − gB (0, ∞).
z→0

In a similar way, in B2 := {z : u(z) < 0} the function −u(z) coincides with the
Green function of B2 with pole at z = ∞. Hence

log r(B2 , ∞) = lim (−u(z) − log |z|) = log r(B, ∞) − gB (0, ∞).
z→∞

Adding together the above relations, in view of (4.8) and what we have already
proved, we obtain

log[r(B, 0)r(B, ∞)] − 2gB (0, ∞)


= log[r(B1 , 0)r(B2 , ∞)] = log[r(B1 ∪ B2 , 0)r(B1 ∪ B2 , ∞)] + 2gB1 ∪B2 (0, ∞)
 0)r(G,
 log[r(G,  ∞)] + 2g  (0, ∞),
G

where G  = C \ R(C \ (B1 ∪ B2 )). Since B1 ∩ B2 = ∅, for each θ the ray L(θ)
intersects the set C \ (B1 ∪ B2 ). Hence, by the definition of symmetrization with
respect to a circle the set R(C \ (B1 ∪ B2 )) contains the circle |z| = 1. Hence the
connected components of G  containing z = 0 and z = ∞ are disjoint and

gG (0, ∞) = 0.
98 Chapter 4. Symmetrization

Thus,

 0)r(G,
log[r(B, 0)r(B, ∞)] − 2gB (0, ∞)  log[r(G,  ∞)] − 2g  (0, ∞)
G
 log[r(G, 0)r(G, ∞)] − 2gG (0, ∞).

The last inequality follows from (2.18). Taking (4.8) into account again, we obtain
(4.7) for k = 1. 

Exercises 4.1
(1) Show that if B is an open subset of P (a, b) and E is a closed subset of P (a, b),
then St B is open in P (a, b) while St E is closed in P (a, b).
(2) Show that the Steiner symmetrization transforms a domain into a simply
connected one.
(3) Show that the Steiner and the circular symmetrization in C do not increase
the length of the boundary of a simply connected domain, and the area of
the domain remains the same.
(4) Show that if a domain B in C is starlike with respect to z = 0, then St B is
also starlike.
(5) Show that the Steiner symmetrization preserves the convexity of a domain.
(6) Let B be a finitely connected domain in P (a, b) without isolated boundary
point and Γ be a closed subset of the boundary ∂B consisting of finitely
many nondegenerate connected components. Let γ := (∂B) \ Γ ⊂ ∂P (a, b)
and z0 ∈ γ. Prove the inequalities

r(B, Γ, z0 )  r(St B, ∂(St B) \ St γ, Re z0 ).

(7) (Krzyż [Kr]) Prove the inequality

gB (z, ζ)  gSt B (Re z, Re ζ)

for simply connected domains B ⊂ C.


(8) Prove Corollary 4.2.
(9) Let D be an admissible domain in C which is mirror symmetric relative to a
straight line l∗ and let z1 , z2 ∈ l∗ be a pair of points in D which are symmetric
relative to l∗ . Let l be the straight line through z1 and z2 . Show that

H(D, z1 , z2 )  H(Stl D, z1 , z2 ),

where H(D, z1 , z2 ) was defined in Exercise 2.4(2) and Stl denotes the Steiner
symmetrization with respect to the line l, that is, the composite ϕ−1 ◦ St ◦ϕ,
where ϕ is a motion in C taking l to the real axis.
4.1. Symmetrization along straight lines or circles 99

(10) Let B be a finitely connected subdomain of the annulus K(ρ, R) which has
no isolated boundary points and one of whose boundary components is a
circle γ := {z : |z| = ρ}, let Z = {zk }nk=1 be a system of distinct points in

n
B ∪ γ, = {δk }nk=0 a tuple of real numbers such that δ0 = 0 and δk2 = 0;
k=1
and let Ψ = {μk rνk }nk=1 , where the μk and νk , k = 1, . . . , n, are positive
numbers. Assume that each circle |z| = r, ρ  r < R, contains at most two
points in Z, and if zk and zl have the same modulus, then δk δl < 0. Prove
the inequality
M (B, (∂B) \ γ, Z, , Ψ)  M (K(ρ, R), {z : |z| = R}, Z  , , Ψ),
where Z  = {zk }nk=1 with zk = |zk | for δk > 0 and zk = −|zk | for δk < 0,
k = 1, . . . , n.
(11) Let E be a compact subset of C and B an open set containing E. Show that
for sufficiently large t > 0
Crt B ⊃ St E,
where Crt is the circular symmetrization with respect to the half-axis {x+iy :
x  −t, y = 0}.
(12) For fixed v > 0 let ϕv (z) := z/v − i be a linear transformation taking the
circles |z − iv| = v to |z| = 1, and let ϕ−1
v be the inverse transformation.
Show that if v > 0 is sufficiently large, then for the sets E and B in the
previous exercise
Rv B ⊃ St E,
where Rv = ϕ−1 v ◦ R ◦ ϕv .
(13) Let E1 , E2 , and E3 be closed pairwise disjoint subsets of P (a, b), ∞ < a <
b < +∞, such that −i∞ ∈ E1 and i∞ ∈ E3 . Let E 1 and E 3 denote the
connected components of the set P (a, b) \ St(P (a, b) \ (E1 ∪ E3 )) such that
−i∞ ∈ E 1 and +i∞ ∈ E 3 . Show that

cap(P (a, b), {E1 , E2 , E3 }, {−1, 0, 1})


 cap(P (a, b), {E 1 , St E2 , E
3 }, {−1, 0, 1}).

(14) Let z0 be a fixed point in the plane C and B an open set in C. The Schwarz
symmetrization with respect to a point z0 assigns to B the open disc Sh B
with centre at z0 and area equal to that of B. For a closed set E ⊂ C it
assigns to E the closed disc Sh E with centre z0 and the same area as E. Let
C = (E0 , E1 ) be the condenser in C such that E1  {∞}. Using Theorem
4.1 prove the following symmetrization principle:
cap C  cap ShC.
In addition, if C is an admissible condenser, then equality holds here if and
only if C coincides with the condenser Sh C up to a motion.
100 Chapter 4. Symmetrization

(15) Let P be a rectangle in a strip x1 < x < x2 such that two opposite sides of
P lie on the lines x = x1 and x = x2 , and let M P be the modulus of P with
respect to these sides. We shall say that symmetrization transforms P into
the rectangle P ∗ = {z : x1 < x < x2 , 0 < y < h}, where (x2 − x1 )h is the
area of P . Prove the following inequality:

M P  M P ∗.

(16) Prove Grötzsch’s principle (Exercise 1.3(5)) using symmetrization in Exer-


cise 15.

4.2 Composites of symmetrizations and


conformal mappings
Let f be a linear fractional automorphism of the complex sphere C, and let Sym
be one of the symmetrization methods in C discussed above. Then we can define
the symmetric condenser f −1 (Sym f (C)), which we shall naturally refer to as
the result of the generalized symmetrization of C. The following inequality is a
consequence of the conformal invariance of the capacity (Theorem 1.12) and the
symmetrization principles (Theorems 4.1, 4.2 and 4.3):

cap C  cap f −1 (Sym f (C)).

Non-Möbius maps f for which this inequality holds are also of interest. For exam-
ple, a regular branch of the logarithm brings about symmetrization with respect
to a circle (4.6). We shall consider other examples of such symmetrizations. To do
this we need the following simple result.
Lemma 4.1. For an arbitrary condenser C = (B, {Ek }nk=1 , ) and a closed subset
F of the real axis
cap C  cap(B \ F, {Ek }nk=1 , ).
Here equality sign holds if the condenser C is mirror-symmetric relative to the real
axis.
Proof. The inequality in question follows from the monotonicity of the capacity
(Theorem 1.15). In addition, if C is mirror-symmetric relative to the real axis,
then the condenser (B \ F, {Ek }nk=1 , ) also has this symmetry. By Theorem 1.20

cap C = 2 cap C + = 2 cap(B \ F, {Ek }nk=1 , )+ = cap(B \ F, {Ek }nk=1 , ),

where C + = (B \ F, {Ek }nk=1 , )+ = (B + , {Ek+ }nk=1 , ). 



Let g(z) denote the single-valued branch of the function ζ = z + z 2 − 1 that
maps the sphere C cut along the interval [−1, 1] onto the exterior of the unit disc.
4.2. Composites of symmetrizations and conformal mappings 101

We define g(z) on [−1, 1] using the correspondence between attainable boundary


points ([Gol], Ch. II, § 3). This function takes the family of confocal ellipses

|z + 1| + |z − 1| = ρ + 1/ρ, 1 < ρ < ∞,

to the family of concentric circles |ζ| = ρ, 1 < ρ < ∞, and the interval [−1, 1] is
taken to the circle |ζ| = 1. We emphasize that each point in (−1, 1) corresponds
to two points on the circle |ζ| = 1, which are symmetric relative to the real axis.
By the result of elliptic symmetrization of an open or closed set A in C we
shall mean the set
El A := g −1 (Cr g(A)),
where Cr is circular symmetrization in |ζ|  1.
We can show that elliptic symmetrization transforms open and closed subsets
of C into open and closed subsets, respectively. The symmetrization of a condenser
is defined by formula (4.1) with Sym = El.
Theorem 4.5. If C = (E0 , E1 ) is a condenser, then

cap C  cap El C.

Proof. Let D = {ζ : |ζ| > 1}. Using Lemma 4.1, the conformal invariance of
capacity, and Theorem 4.2 we see that

cap C  cap(C \ [−1, 1], {E0 , E1 }, {0, 1}) = cap(D, {g(E0 ), g(E1 )}, {0, 1})
 cap(D, {Cr g(E0 ), D \ Cr(D \ g(E1 ))}, {0, 1})
= cap(C \ [−1, 1], {g −1(Cr g(E0 )), g −1 (D \ Cr(D \ g(E1 )))}, {0, 1})
= cap(C, {El E0 , C \ g −1 (Cr g(C \ E1 ))}, {0, 1}) = cap El C. 

Now let B be a domain in C and z0 be a finite point in B which does not


lie in the interval [−1, 1]. From Theorem 4.5 and (2.10) we obtain the following
symmetrization principle for domains:

r(B, z0 )|(z02 − 1)−1/2 sinh(Re cosh−1 z0 )|  r(El B, cosh(Re cosh−1 z0 )).

In a similar way, for z0 = ±1 we have

r(B, z0 )  r(El B, 1).

Let h(z) denote the regular branch of ζ = cos−1 z that takes the sphere C
cut along the rays {z : arg z 2 = 0, |z|  1} onto the strip {ζ : −π < Re ζ < 0}. We
define this function on the sides of the cuts in the natural way. This map takes
the family of confocal hyperbolae

|z + 1| − |z − 1| = 2 cos α, 0 < α < π,


102 Chapter 4. Symmetrization

to the family of vertical lines Re ζ = −α, 0 < α < π, and it takes the boundary
rays to the straight lines Re ζ = −π and Re ζ = 0. Note that each point in
{z : arg z 2 = 0, |z| > 1} corresponds to two points on the line Re ζ = −π or
Re ζ = 0, which are symmetric relative to the real axis.
Let A be an open or closed set in the plane C. We call the following set the
result of hyperbolic symmetrization of A:

Hyp A = h−1 (St h(A)),

where St is the Steiner symmetrization in the strip {ζ : −π  Re ζ  0}. For


an arbitrary condenser C = (E0 , E1 ) with E1  ∞ its hyperbolic symmetrization
is defined by (4.1). The following result is proved similarly to the case of elliptic
symmetrization.
Theorem 4.6. If C = (E0 , E1 ) is a condenser such that ∞ ∈ E1 , then

cap C  cap Hyp C.

Using asymptotic formula (2.10) we obtain the following principle of hyper-


bolic symmetrization for domains:

r(B, z0 )|(1 − z02 )−1/2 sin(Re arccos z0 )|  r(Hyp B, cos(Re arccos z0 )),

where B is a domain in the plane C, z0 is a point in B lying apart from the rays
{z : arg z 2 = 0, |z|  1}, and arccos z denotes the branch of the inverse cosine
described above.

Let p(z) be the regular branch of ζ = z which maps the plane C cut along
the real negative half-axis onto the right half-plane Re ζ > 0. This mapping takes
the family of parabolas

|z| = Re (2c − z), 0 < c < ∞,



to the family of vertical lines Re ζ = c, 0 < c < ∞, and the boundary cor-
respondence takes the negative half-axis to the imaginary axis, so that points
on the negative half-axis correspond to pairs of points in the ζ-plane which are
mirror-symmetric relative to the real axis.
For an open or closed subset A of C we define its parabolic symmetrization by

Par A = p−1 (St p(A)).

This can be regarded as symmetrization along the above family of parabolas.


Symmetrization of condensers C = (E0 , E1 ) such that E1  ∞ is defined by (4.1).
Theorem 4.7. For each condenser C = (E0 , E1 ) such that ∞ ∈ E1 ,

cap C  cap Par C.


4.3. Separating and averaging symmetrizations 103

Proof. This is similar to the proof of Theorem 4.5, although we must use Theorem
4.1 in place of Theorem 4.2. 

Let B be a domain in the plane C and z0 a point in B lying apart from the
negative real half-axis. From Theorem 4.7 and (2.10) we obtain
 √
r(B, z0 )Re z0 /|z0 |  r(Par B, (Re z0 )2 ).

These examples are far from exhausting the method for constructing sym-
metrization transformations which we have sketched
) above. In particular, for each
quadratic differential Q(z)dz 2 the map ζ = i Q1/2 (z)dz in combination with
Steiner symmetrization St result in some symmetrization along the trajectories
of this quadratic differential. We need not formulate the details of each of these
transformations: it is sufficient that we know the principles of their construction.

4.3 Separating and averaging symmetrizations


Using the principles of composition and symmetry and combining concrete trans-
formations from Chapter 3 bring about various symmetrization methods for sets
and condensers. We shall consider several of these methods, whose efficiency has
been verified in practice. Let Dk , k = 1, . . . , n (n  1), be simply connected
disjoint domains in C bounded by finitely many piecewise analytic curves, and
let {pk }nk=1 be a family of functions ζ = pk (z) mapping the domains Dk con-
formally and univalently onto the right half-plane Re ζ > 0 and defined on the
boundary ∂Dk by means of boundary correspondence. For an arbitrary set A ⊂ C
let A(k) be the union of pk (A ∩ Dk ) with its reflection in the imaginary axis. Let
C = (B, {Ei }m i=1 , {δi }i=1 ) be a condenser satisfying the following condition: each
m

Dk , k = 1, . . . , n, intersects at least two plates of C. By the result of the sepa-


rating transformation of the condenser C with respect to the family of functions
{pk }nk=1 we mean the family {Ck }nk=1 of condensers
(k)
Ck = (B (k) , {Ekj }m
j=1 , {δkj }j=1 )
k mk

mirror-symmetric relative to the imaginary axis, where the set {Ekj }m k


j=1 consists
of the nonempty intersections of the form Ei ∩ Dk , 1  i  m (mk  m), and
δkj = δi if Ekj ⊂ Ei , j = 1, . . . , mk , k = 1, . . . , n.
Theorem 4.8. In the above notation

1
n
cap C  cap Ck . (4.9)
2
k=1

Proof. The open sets B ∩ Dk , k = 1, . . . , n, lie in B and are pairwise disjoint. The
k := (B ∩ Dk , {Ekj }mk , {δkj }mk ), k = 1, . . . , n and C satisfy the
condensers C j=1 j=1
104 Chapter 4. Symmetrization

k each plate Ekj lies in some plate Ei of


following condition: in each condenser C
the condenser C which has the same potential δi = δkj . By Theorem 1.17

n
cap C  k .
cap C
k=1

In view of conformal invariance,


k = cap(pk (B ∩ Dk ), {pk (Ekj )}mk , {δkj }mk ),
cap C j=1 j=1

k = 1, . . . , n. It remains to use the symmetry principle (Theorem 1.20), which


states that
1
cap(pk (B ∩ Dk ), {pk (Ekj )}m
j=1 , {δkj }j=1 ) =
k mk
cap Ck ,
2
k = 1, . . . , n. 

q3 D3
E3
D

q4
D4
E4

Figure 4.5: The separating transformation.

Now we look closer at one important special case of separating transforma-


tion, when C lies in the extended plane C, (B = C) and consists of just two plates.
Then the condition on C = (E0 , E1 ) required for the separating transformation
takes the following form:
Ej ∩ Dk = ∅, j = 0, 1, k = 1, . . . , n,
and after the separating transformation we obtain a system of condensers {Ck }nk=1 ,
(k) (k)
where Ck = (E0 , E1 ), k = 1, . . . , n. We have inequality (4.9) again. Moreover,
the following result holds (see also Theorem A3).
4.3. Separating and averaging symmetrizations 105

Theorem 4.9. If C = {E0 , E1 } is a regular admissible condenser, then equality



n
holds in (4.9) if and only if the field of the condenser satisfies G ⊂ Dk and
k=1
the potential function of C is mirror-symmetric relative to each analytic arc in
(∂Dk ) ∩ G7 , 1  k  n, in a neighbourhood of this arc. In particular, if C is
an admissible condenser with connected field, and some intersection (∂Dk ) ∩ G,
1  k  n, contains a piece of a straight line or circle γ and if equality holds in
(4.9), then C is mirror-symmetric relative to γ.
Proof. Assume that we have equality sign in (4.9). Let ω be the potential function
of the condenser C. Then

n
cap C = I(ω, G)  I(ω, G ∩ Dk ). (4.10)
k=1

Let us introduce the auxiliary functions



ω(p−1
k (ζ)), Re ζ  0,
ωk (ζ) =
ω(p−1
k (−ζ)), Re ζ  0, k = 1, . . . , n.

It is easy to see that for each k the function ωk is continuous in C, vanishes on


(k) (k)
the set E0 and is equal to 1 on E1 . It follows from Theorems 1.3 and 1.4 that
the function ωk is Lipschitz in a neighbourhood of each finite point in the field
G(k) of Ck , with the possible exception of finitely many points on the imaginary
axis which correspond to points where the boundary of Dk fails to be analytic and
also of the point corresponding to z = ∞ under the map pk . Since the Dirichlet
integral is invariant under Möbius transformations, taking account of Theorem
1.13 we obtain in turns
1 1
I(ω, G ∩ Dk ) = I(ωk , G(k) )  cap Ck , k = 1, . . . , n. (4.11)
2 2
It follows from (4.10), (4.11) and our assumption that we have equality signs in all
these relations. Equality in (4.10) yields I(ω, C\ ∪nk=1 Dk ) = 0. Now, if a point in
the field G lies in C\ ∪nk=1 Dk , then the function ω is constant in the corresponding
connected component of G, which contradicts the assumption that C is a regular
condenser. Thus
n
G⊂ Dk . (4.12)
k=1

Now let α be an analytic arc in (∂Dk ) ∩ G. Then equality in (4.11) and Theorem
1.13 ensure that ωk is a harmonic function in G(k) . By the uniqueness theorem, ωk
coincides with the analytic extension of ω ◦ p−1
k in a neighbourhood of the interval
pk (α). However, ωk is mirror-symmetric relative to pk (α). Hence the function ω
7 That is, the potential function takes equal values at points symmetric relative to this arc.
106 Chapter 4. Symmetrization

is mirror-symmetric relative to α in a neighbourhood of α. In particular, if the


field G is connected and α is an interval of a straight line or circle γ, then by the
uniqueness theorem ω is mirror-symmetric relative to γ, so that the condenser C
is also mirror-symmetric relative to γ.
Conversely, assume that (4.12) holds and the potential function ω of C has
the symmetry required in the theorem. Then for each k, 1  k  n, and any
analytic arc α ⊂ (∂Dk ) ∩ G the analytic extension of ω ◦ p−1 k is mirror-symmetric
relative to the interval pk (α) and therefore coincides with ωk in a neighbourhood of
this segment. Hence ωk is a harmonic function in a neighbourhood of each point in
G(k) lying on the imaginary axis. In view of the definition of ωk , it is the potential
function of Ck . Note that if (∂Dk ) ∩ G = ∅, then ωk is the potential function of
the condenser Ck by definition. In either case we have (4.11), and in combination
with the obvious equality in (4.10) this yields the equality in (4.9). 

Let Dk and pk be the domains and functions defined above and assume that
there exists a point z0 which is the support of a unique boundary element of each
Dk and such that the functions pk satisfy the asymptotic equalities

|pk (z) − pk (z0 )| ∼ dk |z − z0 |1/βk , z → z0 , k = 1, . . . , n,


n
where the dk and βk , k = 1, . . . , n, are positive quantities and βk = 2 (if z0 = ∞
k=1
(pk (z0 ) = ∞), then z − z0 (pk (z) − pk (z0 )) must be replaced by 1/z (1/pk (z))).
We call this family of functions {pk }nk=1 an admissible family for a separating
transformation of open sets with respect to the point z0 . Let B be an arbitrary
open subset of C containing z0 . Then we call the family of sets {B (k) }nk=1 (which
are mirror-symmetric relative to the imaginary axis) the result of the separating
transformation of B with respect to the family of functions {pk }nk=1 .
Theorem 4.10. The following inequality holds:

n  βk2 /2
r(B (k) , pk (z0 ))
r(B, z0 )  . (4.13)
dk
k=1

In addition, if B is an admissible domain, then equality sign holds in (4.13) if and


only if for each analytic subarc of the set (∂Dk )∩B, 1  k  n, the Green function
of B with pole at z0 is mirror-symmetric relative to this arc in some neighbourhood
of it. In particular, if (∂Dk ) ∩ B, 1  k  n, contains a piece of a straight line or
circle γ and equality holds in (4.13), then B is mirror-symmetric relative to γ.

Proof. We can assume that B is an admissible set. If some intersection (C\B) ∩


Dk is empty, then r(B (k) , pk (z0 )) = ∞ and (4.13) holds trivially. Otherwise we
can apply the separating transformation with respect to the family of functions
{pk }nk=1 to the condenser C(r) := C(r; B, ∂B, {z0 }, {1}, {r}). Let {Ck (r)}nk=1 be
4.3. Separating and averaging symmetrizations 107

the result of this transformation. Then, since y = 1/x is a convex function, we


have  n −1
βk −1 <

n
βk2
|C(r)|  |Ck (r)| βk  |Ck (r)|
2 2
k=1 k=1

by Theorem 4.8. In view of (2.10) and the asymptotic equalities for the pk ,

1 r(B, z0 )
|C(r)| = log + o(1), r→0
2π r
and
1 r(B (k) , pk (z0 ))
|Ck (r)|  log + o(1), r → 0,
2π dk r1/βk

k = 1, . . . , n. Summing these relations we arrive at (4.13).


Now let B be an admissible domain and assume that equality sign holds in
(4.13). We look at the decomposition {B  , C} of the domain B corresponding to
the partition {0, t} with respect to the ‘tuples’ {z0 } and {1}, where t is sufficiently
large. As above, we obtain
n
βk2
|C|  |Ck |, (4.14)
2
k=1

where the system {Ck }nk=1 is obtained from C by the separating transformation
with respect to the family of functions {pk }nk=1 . Using Theorem 2.7, the above
inequality, and (4.13) we conclude that

1 1
log r(B, z0 ) = log r(B  , z0 ) + |C|
2π 2π
1 βk2 r((B  )(k) , pk (z0 )) βk2
n n
 log + |Ck |
2π 2 dk 2
k=1 k=1

1 n
βk2 r(B (k)
, pk (z0 ))
 log .
2π 2 dk
k=1

Hence we also have equality signs in the intermediate relations. In particular,


equality in (4.14) and Theorem 4.9 ensure the symmetry of the Green function
gB (z, z0 ) and the domain B claimed in Theorem 4.10.
Conversely, assume that gB (z, z0 ) has this symmetry. Then we set up the
auxiliary functions

gB (p−1
k (ζ), z0 )/βk , Re ζ  0,
gk (ζ) = −1
gB (pk (−ζ), z0 )/βk , Re ζ  0, k = 1, . . . , n.

For each k, 1  k  n, and any analytic arc α ⊂ (∂Dk ) ∩ B, the analytic extension
of the function gB (p−1
k (ζ), z0 )/βk is mirror-symmetric relative to the interval pk (α)
108 Chapter 4. Symmetrization

and therefore coincides with gk in a neighbourhood of this interval. Hence the


function gk is harmonic on Bk \ pk (z0 ). Since we know its asymptotic behaviour
in a neighbourhood of pk (z0 ), we conclude that it is the Green function of B (k) .
Hence
 $
r(B (k) , pk (z0 )) = exp lim (gk (ζ) + log |ζ − pk (z0 )|)
ζ→pk (z0 )
 $
= dk exp lim βk−1 (gB (z, z0 ) + log |z − z0 |) = dk [r(B, z0 )]1/βk .
z→z0

This yields equality sign in (4.13). 


Conditions ensuring equality in (4.13) in the general case are presented in
Theorem A4.
Now let ζ = pk (z), z ∈ Dk , k = 1, . . . , n, be elements of an admissible family
of functions for a separating transformation of sets with respect to z = ∞, and
n
let |pk (z)| ∼ d−1
k |z|
1/βk
as z → ∞, z ∈ Dk , where dk , βk > 0 and βk = 2.
k=1
Let E be an infinite bounded closed subset of C such that E ∩ Dk = ∅, k =
1, . . . , n. As above, E (k) is the union of the set pk (E ∩ Dk ) and its reflection in the
imaginary axis. We can naturally call the family of sets {E (k) }nk=1 the result of the
separating transformation of E with respect to the family of functions {pk }nk=1 .
From Theorem 4.10 we deduce the following result.
Corollary 4.6. The following inequality holds:
n βk2 /2
cap E  dk cap E (k)
k=1

where the assertion about equality sign is the same as in Theorem 4.10, provided
that B is the connected component of C \ E containing the point at infinity.
In 1955, Szegő defined an ‘averaging’ symmetrization, taking at arbitrary
starlike set to an n-fold rotationally symmetric one [S2]. Subsequently, Marcus,
Mityuk, and other authors generalized this definition in various directions. First
we state a general result. Let {αk }nk=1 be a set of positive numbers such that
n
αk = 1; let {βk }nk=1 be a set of real numbers such that βk = ±1, k = 1, . . . , n;
k=1
and let {γk }nk=1 , n  2, be an arbitrary set of real numbers. For an open set
B ⊂ C, where 0 ∈ B, let
 n

VB = reiθ : 0  r < [M (βk θ + γk , B]αk , 0  θ  2π ,
k=1

where M (·, B) is defined by (3.10). For a closed bounded set E let


 n

VE = reiθ : 0  r  [M (βk θ + γk , E)]αk , 0  θ  2π ,
k=1
4.3. Separating and averaging symmetrizations 109

where M (·, E) is defined in (3.11). For a condenser C = (E0 , E1 ) with the first
plate E0  0 and the second plate E1  ∞ we set

V C = (V E0 , C \ V (C \ E1 )).

Theorem 4.11. The following inequalities hold :

r(B, 0)  r(V B, 0), (4.15)


cap E  cap V E, (4.16)
cap C  cap V C. (4.17)

If B is an admissible domain, then equality holds in (4.15) if and only if B is


starlike with respect to the origin and invariant under the transformations λ−1 k ◦
λl , k, l = 1, . . . , n, where λk (reiθ ) := reiβk (θ−γk ) , k = 1, . . . , n. For a compact set
E such that C \ E is an admissible domain, equality holds in (4.16) also only if E
is starlike and invariant under these transformations. Finally, if C = (E0 , E1 ) is
an admissible condenser with connected field, then equality sign in (4.17) holds if
and only if both E0 and C \ E1 are starlike with respect to the origin and invariant
under the transformations λ−1 k ◦ λl , k, l = 1, . . . , n.

Proof. Using the radial transformation (Corollary 3.3) we obtain the inequalities

r(B, 0) = r(λk (B), 0)  r(M λk (B), 0), k = 1, . . . , n,

and if B is an admissible domain, then for each k equality holds if and only if

λk (B) = M λk (B).

In particular, in the case of equality B is a starlike domain with respect to z = 0,


distinct from the whole plane. By Theorem 3.14
n
(r(M λk (B), 0)αk  r(RA {M λk (B)}nk=1 , 0) = r(V B, 0),
k=1

where A = {αk }nk=1 and RA {M λk (B)}nk=1 = V B. For M λk (B) = C equality


holds only when the domains M λk (B) coincide with V B up to a dilation with
centre at z = 0. The above relations yield (4.15). If we have equality sign in
(4.15), then we also have equality in the preceding inequalities. Hence the domain
B is starlike, and the sets λk (B), k = 1, . . . , n, coincide with V B up to a dilation
with centre at the origin. As B = C, such dilations must have coefficient 1, so
λ−1
k (λl (B)) = B, k, l = 1, . . . , n. Conversely, under the above assumptions we
have equality in (4.15). The proof of inequalities (4.16) and (4.17) is similar, but
we must use Corollaries 3.4 and 3.5 in the first case and Theorems 3.11 and 3.13
in the second in place of Corollary 3.3 and Theorem 3.14. 
110 Chapter 4. Symmetrization

If we set βk = 1 and γk = 2πk/n, k = 1, . . . , n, in the V -transformation


above, then its result is n-fold rotationally symmetric, that is, the sets V B, V E
and the plates of the condenser V C are invariant under the rotation by 2π/n about
the origin. This V -transformation is called the Szegő–Marcus symmetrization. We
denote it by SMA (see Figure 3.7). By Theorem 4.11, for V = SMA we have
inequalities (4.15)–(4.17) [M], with equality signs only when B = SMA B, E =
SMA E, and C = SMA C [Mit3]. The limiting case as n → ∞ of the symmetrization
SMA is of interest. For simplicity, let B ⊂ C, 0 ∈ B, be an open set bounded by
analytic curves and in the definition of the symmetrization SMA let A = {αk }nk=1
with αk = 1/n, k = 1, . . . , n. Then the set SMA B is bounded by a piecewise
smooth curve r = r(θ), 0  θ  2π. For each θ
 
1
n
2πk
log r(θ) = log M θ + ,B
n n
k=1
2π 2π
1 1
→ log M (θ + ψ, B)dψ = log M (ψ, B)dψ, n → ∞.
2π 2π
0 0

Consequently,
⎡ ⎤
2π 2π
1 1
r(θ) → exp ⎣ log M (ψ, B)dψ ⎦  M (ψ, B)dψ, n → ∞.
2π 2π
0 0

Let R(B) be the radius of the disc having the same area as B. Then
⎡ ⎤1/2
2π 2π
1 1 ⎣
M (ψ, B)dψ  2π M 2 (ψ, B)dψ ⎦ R
2π 2π
0 0

(see Exercise 3.4(5)). Hence (4.15) is slightly stronger then the inequality
r(B, 0)  r(Sh B, 0), (4.18)
which follows from Schwarz’ symmetrization principle (Exercise 4.1(14)).
Setting n = 2, α1 = α2 = 1/2, β1 = 1, β2 = −1 and γ1 = γ2 = 0 in
the V -transformation, we obtain a version of Aharonov–Kirwan symmetrization
(V = AK), which, given an open or closed set G, assigns to G a set AK G which
is starlike with respect to the origin and mirror-symmetric relative to the real
axis [AhKir]. By Theorem 4.11 Aharonov–Kirwan symmetrization does not reduce
the inner radius of the domain with respect to the origin and does not increase
the logarithmic and conformal capacities. All these quantities remain invariant if
and only if the symmetrized set remains the same.
For an arbitrary bounded closed set E let
+  ,
AKL E = reiθ : 0  r  m(θ, E)m(−θ, E), 0  θ  2π .
4.3. Separating and averaging symmetrizations 111

The set AKL E is different from AK E in that the logarithmic metric is replaced
with the linear metric m(θ, E). It is closed, starlike with respect to z = 0, and
mirror-symmetric relative to the real axis.
Theorem 4.12. If E is a compact set lying on rays arg z = ±θ1 and arg z = ±θ2 ,
where 0  θ1  θ2 − π/2  π/2, then
cap E  cap AKL E.
Proof. Let {E (k) }2k=1
be obtained by the separating transformation of the set E
with respect to the family of functions {pk }2k=1 , where p1 (z) = −iz, D1 = {z :
Im z > 0} and p2 (z) = iz, D2 = {z : Im z < 0}. By Corollary 4.6
 =
cap E  cap E (1) cap E (2) = cap p2 (E (1) ) cap p1 (E (2) ).
A contraction and an averaging transformation yield
= =
cap p2 (E (1) ) cap p1 (E (2) )  cap LRp2 (E (1) ) cap LRp1 (E (2) )
 cap RA {LRp2(E (1) ), LRp1 (E (2) )}
(see Exercise 3.1(5) and Corollary 3.5, where A = {1/2, 1/2}). It remains to ob-
serve that
RA {LRp2 (E (1) ), LRp1 (E (2) )} = AKL E. 
Theorem 4.13. If E is a compact set lying on the rays arg z = 2πk/n, k = 1, . . . , n,
then
cap E  cap SML E,
 $
1n
where SML E denotes the ‘star’ z : arg z n = 0, 0  |z n |  m(2πk/n, E) .
k=1

Proof. Let {E (j) }2n


j=1 be the set obtained from E by the separating transformation
with respect to the family of functions {pj }2n j=1 , pj (z) = −iz , z ∈ Dj := {z :
n

π(j − 1)/n < arg z < πj/n}, j = 1, . . . , 2n. From Corollary 4.6 we obtain
2n n
2 2
cap E  (cap E (j) )1/2n = (cap E (2k−1) )1/n .
j=1 k=1

By Fekete’s theorem
1/n
cap E (2k−1) = cap E2k−1 ,
where E2k−1 = {z : −iz n ∈ E (2k−1) }, k = 1, . . . , n (Exercise 2.1(6)). A contrac-
tion followed by an averaging transformation yield
n n
(cap E2k−1 )1/n  (cap LR E2k−1 )1/n
k=1 k=1
 cap RA {LR E2k−1 }nk=1 = cap SML E,
A = {1/n, . . . , 1/n}. 
112 Chapter 4. Symmetrization

In Theorems 4.12 and 4.13 we deal with linear radial transformations distinct
from LR and, moreover, applied to some special sets E. However, they provide
more information in these special cases than the transformations AK, M , and
SMA . For example, if E is the union of the line interval [0, 1 −√i] and some subset
of the ray {reiπ/4 : ε  r < ∞}, ε > 0,√with linear measure 2, then AKL E is
formed by two line segments of length 2, AK E = {0}, and the Marcus radial
transformation produces the single line interval M E = [0, 1 − i]. In this case the
inequality in Theorem 4.2 is stronger than (4.16) (V = AK) or the inequality in
Corollary 3.4. The description of the equality cases in Theorems 4.12 and 4.13 is
obtained with the help of the previous theorems and Appendices A2 and A3.
Replacing the radial transformations in Theo-
rem 4.11 by transformations along logarithmic spirals
we obtain the Mityuk spiral averaging symmetriza-
tion Mnψ [Mit3]. More precisely, let C = (E0 , E1 )
be the condenser with plates E0  0 and E1  ∞;
let ψ be a fixed number in the interval (−π/2, π/2),
A = {αk }nk=1 be a tuple of positive numbers satis-

n
fying αk = 1, and let {λk }nk=1 be a family of
k=1
rotations about the origin: λk (reiθ ) = rei(θ−2πk/n) ,
O7>6 F k = 1, . . . , n. Then by definition

Figure 4.6: The result of Mnψ C = Mψ {Mψ λk (C)}nk=1 .


the spiral averaging sym- The condenser Mn C is invariant under rotations in
ψ
metrization of the set E. the family {λk }nk=1 and
cap Mnψ C  cap C
(see Exercise 3.5(5)).

Exercises 4.3
(1) Show that elements of the family {Ck }nk=1 involved in the definition of the
separating transformation are indeed condensers.
(2) Let
Dk = {z : |arg(z − z0 ) − 2πk/n| < π/n},

ζ = pk (z) ≡ [(z − z0 )/a]n , z ∈ Dk , Re ζ > 0, a > 0,
d(ζ) = (1 − ζ)/(1 + ζ), p(ζ) = z0 + ζ 2/n .
Let C be a condenser whose plates intersect each sector Dk , k = 1, . . . , n, and
let {Ck }nk=1 be the result of the separating transformation of C with respect
to the family of functions {pk }nk=1 . Then the condensers

Ck∗ := p(d−1 (Cr d(Ck ))), k = 1, . . . , n,


4.4. Dissymmetrization 113

are n-fold rotationally symmetric relative to the point z0 . Note that although
the mapping p(ζ) is not single-valued, the condenser Ck∗ is uniquely defined
because its inverse image is symmetric in the origin. Prove the inequality
1
n
cap C  cap Ck∗ .
n
k=1

(3) Let us generalize the V -transformation defined before Theorem 4.11. Namely,
let {αk }nk=1 and {γk }nk=1 be sets defined above, let {βk }nk=1 be a system of
nonzero integers and B ⊂ C an open set such that 0 ∈ B. Set
 n

V B = re : 0  r <
iθ α /|β |
[M (βk θ + γk , B)] k k , 0  θ  2π .
k=1

Prove the inequality



n
r(B, 0)  r(V B, 0)1/b , where b = αk /|βk |.
k=1

Hint: using Exercise 2.1(17) make appropriate modifications in the proof of


Theorem 4.11.
(4) Show that (4.18) remains valid if the disc Sh B is replaced with the disc
whose area is equal to that of the star of B (the set of points in B that can
be joined to the origin by a line segment lying in B entirely).

4.4 Dissymmetrization
In many problems in geometric function theory the extremal object has the sym-
metry group of a regular polygon, and if, for instance, this object is a solution
of some maximum problem, then the object providing the minimum in the same
problem has a single symmetry axis. A characteristic feature of dissymmetrization
is as follows: by contrast to the known symmetrization transformation, dissym-
metrization enables us to obtain ‘reverse’ estimates. Originally, dissymmetrization
was designed to solve Gonchar’s problem of harmonic measure. Later it became
clear that this transformation is interesting on its own [D4].
Let z0 be a fixed finite point and let L∗k , k = 1, . . . , n, where n  2, be fixed
rays in C, going out of z0 at equal angles. Let Φ be the group of symmetries of
C formed by the composites of the reflections in the rays L∗k , k = 1, . . . , n, and in
the bisectors of the angles formed by these rays. (Obviously, we need not mention
bisectors for odd n.) In other words, Φ is the ‘dihedral group’. Throughout this
section symmetry means Φ-invariance. We say that a set A ⊂ C is symmetric
(or Φ-symmetric) if ϕ(A) = A for any isometry ϕ ∈ Φ. We say that a condenser
C = (B, E, ) is symmetric if the set B and the plates of C are symmetric. Finally,
a real function v on a symmetric set Ω is said to be symmetric if v(z) ≡ v(ϕ(z)) for
114 Chapter 4. Symmetrization

any ϕ ∈ Φ. We call a system of closed sectors with vertices at z0 a decomposition


of the sphere C if no two sectors have common interior points and their union is C.
Let {Pk }Nk=1 be a symmetric decomposition of C, that is, {ϕ(Pk )}k=1 = {Pk }k=1
N N

for any isometry ϕ ∈ Φ. We call a set of rotations {λk }k=1 of the form λk (z) =
N

z0 + (z − z0 )eiθk , k = 1, . . . , N , a dissymmetrization of a symmetric decomposition


{Pk }Nk=1 if the set of images {Sk }k=1 , Sk = λk (Pk ), k = 1, . . . , N , also is a
N

decomposition of C and ) for each nonempty intersection Sk ∩ Sl , k, l = 1, . . . , N ,


there exists an isometry ϕ ∈ Φ such that ϕ(λ−1 −1
k (Sk ∩ Sl )) = λl (Sk ∩ Sl ).
Let A be an arbitrary subset of the sphere let C, C = (B, {Ek }nk=1 , ) be a
symmetric condenser, and v be a symmetric function defined on a symmetric set
Ω. Then we set
N
Dis A = λk (A ∩ Pk ),
k=1
Dis C = (Dis B, {Dis Ek }nk=1 , ),
Dis v(z) = v(λ−1
k (z)), z ∈ Sk ∩ Dis Ω, k = 1, . . . , N.

As v is symmetric, condition ) ensures that the function Dis v is well defined


on Dis Ω. We say that the dissymmetrization {λk }N k=1 takes a set A (a condenser
C, a function v) to the set Dis A (condenser Dis C, function Dis v). Using a more
geometric language, dissymmetrization is a rearrangement of sectors after which an
arbitrary symmetric function v is transformed into a single-valued function Dis v.
Now we state two lemmas on the existence of nontrivial dissymmetrizations.
Lemma 4.2. Let Lk , k = 1, . . . , n, be different rays in C going out of a point z0 , and
let ψ be the smallest positive angle between the Lk . Then there exist a symmetric
decomposition {Pk }N k=1 , N  n, and a dissymmetrization {λk }k=1 such that for
N

each k = 1, . . . , n the sector Pk has a bisector Lk , the opening angle of Pk is equal
to ψ, and Dis L∗k = Lk .

M4 M5 Tl
Ql b
l l )bl * ? m )bm *
bm
Qm Tm
M4
M3

B M3
Fku B

M5

Figure 4.7: Dissymmetrization.


4.4. Dissymmetrization 115

Proof. We can assume that z0 = 0 and L∗k = {z : arg z = 2πk/n}, k = 1, . . . , n.


We shall call a set of the form {z : |arg z n | = θ}, 0  θ  π, a symmetric pencil
of rays. For k = 1, . . . , n let Lk = {z : arg z = αk }, let Qk be the closed sector
with opening angle ψ such that its bisector is equal to the ray L∗k , let δk (z) =

m
z exp(i(αk − 2πk/n)) and Rk = δk (Qk ). We shall use the notation Am = Qk
k=1

m
and Bm = Rk . Note that the boundary rays of An form a symmetric pencil
k=1
of rays. If Bn = C, then the construction is complete. Otherwise C \ Bn consists
of n1 open sectors (0 < n1 < n) and C \ An consists of n open sectors with
equal opening angles and the same total magnitude. Let Qn+1 be the closure of
one of the latter sectors and R be the largest sector adjacent to Bn (that is,
meeting Bn only at boundary points). By definition, the rotation δn+1 takes Qn+1
to a sector Rn+1 ⊂ R adjacent to Bn . Then the number of sectors in C \ Bn+1
remains the same (= n1 ), and the number of sectors C \ An+1 gets reduced by
one. After repeating this procedure n − n1 times, we obtain n1 sectors in both
C \ A2n−n1 and C \ B2n−n1 . Let ψ1 be the opening angle of the smallest sector
in the last set. Then we draw 2n1 different closed plane sectors Qk adjacent to
A2n−n1 , k = 2n − n1 + 1, . . . , 2n + n1 , with opening angle ψ1 /2 and we associate
to each Qk the corresponding rotation δk which maps Qk to a sector Rk adjacent
to B2n−n1 so that the common side of Qk and A2n−n1 goes to the common side of
Rk and B2n−n1 , but on the other hand Rk has no common interior points with the
other sectors Rk , k  = k. Next we perform the analogous construction for A2n+n1
and B2n+n1 in place of An and Bn , respectively. Since the number of sectors in
the complement of BN decreases strictly on each step, for some N  the system of

sectors {Rk }N
k=1 forms a decomposition of C.

We see from this construction that the decomposition {Rk }N k=1 and the sys-
N
tem of rotations {δk }k=1 satisfy condition ) (in appropriate notation). How-

ever, generally speaking, {Qk }N k=1 is not symmetric. Let {Pk }k=1 be an arbi-
N
N
trary symmetric decomposition produced from {Qk }k=1 by subdividing the Qk
with k = n + 1, . . . , N  into smaller sectors (we set Pk = Qk for k = 1, . . . , n).
Then each sector Pk , 1  k  N , lies in some Qk , 1  k   N  , and we set
λk (z) ≡ δk (z). It is easy to see that the symmetric decomposition {Pk }Nk=1 and
the dissymmetrization {λk }N k=1 are as required. 

Remark 4.1. In the hypotheses of Lemma 4.2 there also exist a symmetric de-
composition {Pk }N
  N ∗
k=1 and a dissymmetrization {λk }k=1 such that each ray Lk is

the common boundary of two sectors in {Pk }k=1 and Dis Lk = Lk , k = 1, . . . , n.
N ∗

Indeed, starting from the decomposition {Pk }N k=1 constructed in Lemma 4.2 we
 
construct another decomposition {Pk }k=1 by replacing the first n sectors by their
N

halves. We set the maps associated with these halves to be equal to the rotations
associated with the whole sectors Pk , k = 1, . . . , n.
116 Chapter 4. Symmetrization

Lemma 4.3 (Haliste [Hal]). Let E ∗ be a symmetric subset of the circle |z − z0 | = 1


consisting of n closed disjoint arcs, and let E be a subset of |z − z0 | = 1 also con-
sisting of n closed arcs with disjoint interiors whose total length is equal to that of
E ∗ . Then there exist a symmetric decomposition {Pk }N K=1 and a dissymmetriza-

tion {λk }Nk=1 such that Dis E = E.
Proof. We can assume that z0 = 0, L∗k = {z : arg z = 2πk/n}, k = 1, . . . , n, and
E ∗ = {z : |z| = 1, |arg z n |  l/2}, 0 < l < 2π. Let P be the union of closed
infinite sectors ‘based’ on the arcs in E ∗ and S the similar union of sectors ‘based’
on E. Let f be a function which has the form f (z) = eiθk z 2π/l (with real θk )
in each sector in P and takes it to a sector bounded by two neighbouring rays
L∗k and L∗k+1 (where we set L∗n+1 = L∗1 ) so that the image of the whole of P
is C. Let g be a function defined in a similar way for the set S (here we must
take some suitable rays Lk in place of the L∗k ). Let {Pk }N
  N
k=1 and {λk }k=1 be
the decomposition and dissymmetrization in Remark 4.1. The system of sectors
 −1   , makes up a decomposition of P , and the
{Pk }N
k=1 , Pk = f (Pk ), k = 1, . . . , N
 −1   , makes up a
system {Sk }Nk=1 , Sk = λk (Pk ), λk = g (λk (f (z))), k = 1, . . . , N
decomposition of S. Note that the Sk and λk satisfy condition ) and


N
E= λk (E ∗ ∩ Pk ).
k=1

Replacing E and E by {z : |z| = 1} \ E and {z : |z| = 1} \ E ∗ , respectively, and
repeating the above construction we obtain a decomposition of the whole of C and
a dissymmetrization such that Dis E ∗ = E. 
Theorem 4.14. For any Φ-symmetric condenser C = (B, E, )
cap C  cap Dis C. (4.19)
In addition, if C is an admissible condenser with connected field, then equality in
(4.19) holds if and only if C coincides with Dis C up to a rotation about z0 .
Proof. First we show that the capacity of a symmetric condenser is the infimum
of the Dirichlet integrals over all symmetric admissible functions. For a proof it is
sufficient that to each admissible function v for C we assign a symmetric admissible
function v ∗ with Dirichlet integral not exceeding the Dirichlet integral of v. The
straight lines containing the rays L∗k or the bisectors of the sectors formed by
these rays subdivide C into 2n closed sectors Dj , j = 1, . . . , 2n. Assume that an
admissible function v for the condenser C = (B, E, ) satisfies
I(v, B ∩ Dj  ) = min I(v, B ∩ Dj ), 1  j   2n.
1j2n

Then we construct a symmetric function v ∗ from v as follows:


v ∗ (z) = v(z), z ∈ B ∩ Dj  ,
∗ ∗
v (z) = v (ϕ(z)), for any z ∈ B and ϕ ∈ Φ.
4.4. Dissymmetrization 117

Using the symmetry of C and Theorem 1.3 we conclude that v ∗ is admissible for
C. Since the Dirichlet integral is Φ-invariant, we have
I(v ∗ , B) = 2nI(v ∗ , B ∩ Dj  ) = 2nI(v, B ∩ Dj  )  I(v, B).
Now let v ∗ be a symmetric admissible function for C = (B, {Ek }m k=1 , {δk }k=1 ).
m

Then the function Dis v is continuous on Dis B, Lipschitz on compact subsets of
Dis B (see Theorem 1.3), and equal to δk on Dis Ek , k = 1, . . . , m. In view of the
conformal invariance of the Dirichlet integral and the properties of dissymmetriza-
tion,

N
N
I(v ∗ , B) = I(v ∗ , B ∩ Pk ) = I(Dis v ∗ , (Dis B) ∩ Sk )
k=1 k=1

= I(Dis v , Dis B)  cap Dis C.
Taking the infimum over all possible v ∗ we obtain inequality (4.19).
Assume that we have equality sign in (4.19), and let C be an admissible

m
condenser with connected field G := B \ Ek . The potential function u of C is
k=1
Φ-symmetric because it is unique and for each ϕ ∈ Φ the function u ◦ ϕ is also
a potential function of C. The function Dis u is continuous in Dis B, Lipschitz on
compact subsets of the field of Dis C, and equal to δk on Dis Ek , k = 1, . . . , m. By
the Dirichlet principle, in view of our calculations,
cap C = I(u, B) = I(Dis u, Dis B)  cap Dis C = cap C.
Hence Dis u is the potential function of the condenser Dis C. We take k  such that
 denote the connected component of the field
G ∩ Pk = ∅, 1  k   N , and let G
of Dis C containing the set λk (G ∩ Pk ). On this set
Dis u = u ◦ λ−1
k .

By the uniqueness theorem for harmonic functions this equality holds in the whole
 and therefore G
of G,  = λk (G). Now,

I(Dis u, Dis B) = I(u, G) = I(u ◦ λ−1 


k , λk (G)) = I(Dis u, G).


 is the field of the condenser Dis C, and so Dis C = λk (C).


Hence G 
Theorem 4.15. Assume that the reduced modulus M (B, Γ, Z, , Ψ) exists for the
sets B and Γ and systems Z = {zk }m k=1 , = {δk }k=0 , and Ψ (any version of
m

the reduced modulus can be used here, including the cases when Γ = ∂B, B = C,
or Γ = ∅ : then must be taken in place of ). Assume that the sets B, Γ, and

m
{zk } are Φ-symmetric and if ϕ(zk ) = zl for some zk and zl and for ϕ ∈ Φ,
k=1
then necessarily δk = δl . Then
M (B, Γ, Z, , Ψ)  M (Dis B, Dis Γ, {Dis zk }m
k=1 , , Ψ),

provided that the reduced modulus on the right-hand side exists.


118 Chapter 4. Symmetrization

Proof. For all versions of the definition of the reduced modulus the proof is the
same. For definiteness we shall consider the reduced modulus of an open set B
with respect to its boundary ∂B and the systems Z = {zk }m k=1 , = {δk }k=0 ,
m
 2
m
where δ0 = 0 and δk = 0, and Ψ = {μk rνk }m
k=1 . By definition
k=1
+ ν ,
M (B, ∂B, Z, , Ψ) = lim |C(r; B, ∂B, Z, , Ψ)| + log r ,
r→0 2π
m −1
 2
where ν = δk /νk and the condenser C = (r; B, ∂B, Z, , Ψ) was defined
k=1
in the beginning of § 2.2. In view of Lemma 2.1, we can assume that this condenser
is Φ-symmetric. Hence we can define the condenser Dis C(r; B, ∂B, Z, , Ψ), which
is actually the condenser C (r; Dis B, ∂(Dis B), {Dis zk }mk=1 , , Ψ). As above,

M (Dis B, ∂(Dis B), {Dis zk }m


k=1 , , Ψ)
+ ν ,
= lim |C(r, Dis B, ∂(Dis B), {Dis zk }mk=1 , , Ψ)| + log r .
r→0 2π
It remains to use Theorem 4.14. 

Now we take a closer look at an important special case.


Theorem 4.16. Let z ∗ be the point z0 (from the definition of dissymmetrization)
or ∞, and let B ⊂ C, z ∗ ∈ B, be an open Φ-symmetric set. Then

r(B, z ∗ )  r(Dis B, z ∗ ). (4.20)

In addition, if B is an admissible domain, then equality holds if and only if B


coincides with Dis B up to a rotation about z0 .

Proof. Inequality (4.20) follows from Theorem 4.15 and (2.16). The description
of the equality cases is based on Theorem 4.14; it repeats the proof of the corre-
sponding result for polarization (Theorem 3.4) with obvious modification and with
the help of the following simple fact: the Green function of a symmetric domain
is symmetric relative to the group Φ. 
Corollary 4.7. For an infinite bounded closed symmetric set E

cap E  cap Dis E

with the same equality cases as in Theorem 4.16, provided that B is the connected
component of C \ E containing the point at infinity.
In conclusion we present the solution of Gonchar’s problem of the harmonic
measure of radial line segments. Let ω(z, E, D) be the harmonic measure of the
set E with respect to the domain D calculated at the point z [St].
4.4. Dissymmetrization 119

F F

Figure 4.8: The generalized Gonchar problem.

Theorem 4.17. Let U be the unit disc |z| < 1, ρ and θk , k = 1, . . . , n, be real
numbers such that 0 < ρ < 1 and 0  θ1 < θ2 < · · · < θn < 2π, and let

n 
n
E= {r exp(iθk ) : ρ  r  1}, E∗ = {r exp(2πi(k − 1)/n) : ρ  r  1}.
k=1 k=1

Then
ω(0, E, U \ E)  ω(0, E ∗ , U \ E ∗ ).
Here equality sign holds only when the domain U \ E coincides with U \ E ∗ up to
a rotation about the origin.
Proof. Let us introduce some notation: Γ = {z : |z| = 1}, f is a function mapping
the domain U \ E conformally and univalently onto the disc U so that f (0) = 0
and defined on the boundary of U \ E by means of boundary correspondence; f ∗
is a similar function for U \ E ∗ . Assume that

ω(0, E, U \ E)  ω(0, E ∗ , U \ E ∗ ).

In view of the conformal invariance of the harmonic measure and the fact that
the harmonic measure at z = 0 of a subset of Γ with respect to U is equal to the
linear measure of this subset divided by 2π, we see that

mf (Γ)  mf ∗ (Γ).

By Lemma 4.3 there exists a dissymmetrization Dis taking f ∗ (Γ) to a subset of


Γ containing f (Γ) (z0 = 0). Then from Corollary 4.7 and the monotonicity of the
logarithmic capacity we obtain

cap f ∗ (Γ)  cap Dis f ∗ (Γ)  cap f (Γ).

On the other hand, using Exercise 2.4(13) and equalities (2.2) and (2.5) we find
1 r(U \ E, 0)
(cap f (Γ))2 = = = r(U \ E, 0).
r(U, f (Γ), 0) r(U \ E, Γ, 0)
120 Chapter 4. Symmetrization

In the last equality we have used the representation gU \E (z, 0, Γ) = − log |z|. In a
similar way
(cap f ∗ (Γ))2 = r(U \ E ∗ , 0).
The domain U \ E can be regarded as the result of the dissymmetrization of the
domain U \ E ∗ (see Lemma 4.2). In view of (4.20),
r(U \ E, 0)  r(U \ E ∗ , 0),
and in combination with the previous relations this yields
cap f (Γ) = cap f ∗ (Γ).
Thus equality holds in all the intermediate relations. In particular, ω(0, E, U \E) =
ω(0, E ∗ , U \ E ∗ ) and U \ E coincides with the domain U \ E ∗ up to a rotation
about the origin. 
It is essential for the above proof that the set E consists of n line segments
attached to the circle |z| = 1. Baernstein conjectured that Theorem 4.17 also holds
when the interval ρ  r  1 in the definition of E is replaced by an arbitrary
compact subset of (0, 1] (Fig. 23). So far this conjecture has only been proved in
several special cases (see also Problem 12 in Appendix A6).

Exercises 4.4
(1) Let B be a symmetric open and E a symmetric closed subset of C. Show
that Dis B is open and Dis E is closed.
(2) Is it true that a dissymmetrization takes a symmetric domain to a domain?
(3) (Haliste [Hal]) Construct the dissymmetrization in Lemma 4.3 without using
Lemma 4.2.
(4) (Solynin [Sol1]) Let Dn∗ be a regular n-gon and Dm an arbitrary m-gon with
m  n, both circumscribed about the circles |z| = 1. Show that there exist
a symmetric decomposition {Pk }N k=1 and a dissymmetrization {λk }k=1 such
N

that Dis Dn∗ ⊂ Dm (z0 = 0). Prove the inequality




n22/n tan πn Γ 1 − n1
r(Dm , 0) 
,
π 1/2 Γ 12 − n1
where Γ(·) is the Euler gamma function.
(5) (Solynin [Sol1]) Let Dn∗ be as above and Dm be a convex m-gon with m 
n which has the same perimeter as Dn∗ . Let Φ be the symmetry group of
Dn∗ . Show that the complement C \ Dn∗ has a Φ-symmetric decomposition

{Pk }Nk=1 into closed infinite sectors having vertices at vertices of Dn and

closed half-strips with finite sides lying on sides of Dn and there exists a
dissymmetrization, a system of motions {λk }N k=1 (which are not necessarily
rotations) such that the set of images {Sk }N
k=1 , Sk = λk (Pk ), k = 1, . . . , N ,
is a decomposition of C \ Dm and condition ) holds.
4.4. Dissymmetrization 121

(6) Show that a dissymmetrization of the cross section of a homogeneous isotropic


elastic tube does not reduce the torsional rigidity of the tube and does not
increase the principal frequency of a homogeneous, uniformly stretched mem-
brane (see [PS], §§ 1.26, 1.27 for definitions).
(7) For arbitrary real θk , k = 1, 2, 3, 4, and ρk , 0 < ρk < 1, k = 1, 2, 3, 4,
ρ1 = ρ3 , ρ2 = ρ4 , set
4
 
n
E= {r exp(iθk ) : ρk  r}, E∗ = {r exp(πi(k − 1)/2) : ρk  r}.
k=1 k=1

Does the inequality

r(U (0, R) \ E, 0)  r(U (0, R) \ E ∗ , 0)

hold for each R > 0?


Chapter 5

Metric Properties of Sets


and Condensers

5.1 Finite sets


Let zk , k = 1, . . . , n, be some points on the unit circle Γ = {z : |z| = 1} and let
zk∗ = exp(2πi(k − 1/n), k = 1, . . . , n, be points positioned symmetrically on the
circle. The following classical inequality goes back to Pólya and Schur:
n n n n
|zk − zl |  |zk∗ − zl∗ | = nn . (5.1)
k=1 l=1 k=1 l=1
l=k l=k

In potential-theoretic terms (5.1) this means that the mutual energy


n
n
log |zk − zl |−1
k=1 l=1
l=k

is smallest when the points zk are positioned symmetrically. In this section we


consider some extensions and supplements to (5.1).
Theorem 5.1. Let θk , k = 1, . . . , 2n, be real numbers such that 0  θ1 < θ2 <
· · · < θ2n < 2π (n  2); and let r1 , . . . , rm be different positive numbers (m  1).
Let {zk }2mn
k=1 denote the set of different points in C with arguments θk and absolute
values rl , 1  k  2n, 1  l  m. Let {δk }2mn k=1 be a sequence of real numbers
satisfying the following condition: δk = δl if |zk | = |zl |, 1  k, l  2mn. Then

2mn 2mn    δk δl


 zk − zl δk δl 2mn 2mn  zk (η) − zl (η) 
     , (5.2)
 R 2 − z k zl   R2 − z (η)z (η) 
k=1 l=1 k=1 l=1 k l
l=k l=k

© Springer Basel 2014 123


V.N. Dubinin, Condenser Capacities and Symmetrization in Geometric Function
Theory, DOI 10.1007/978-3-0348-0843-9_5
124 Chapter 5. Metric Properties of Sets and Condensers

where R is an arbitrary quantity larger than max rl , and the different, sym-
1lm
metrically positioned points zk (η) are defined by the relations |zk (η)| = |zk | and

n
|arg zkn (η)| = η/2, where η = (θ2k − θ2k−1 ), k = 1, . . . , 2mn.
k=1

Proof. For fixed R we look at the reduced modulus M (B, ∂B, Z, , Ψ), where
B = {z : |z| < R}, Z = {zk }2mn 2mn
k=1 , = {δk }k=1 , and Ψ = {r, . . . , r}. By (2.16),

M (B, ∂B, Z, , Ψ)
⎧ ⎫

⎨2mn  2 ⎪
1 2
R − |zk | 2 2mn
2mn
 R − z k zl ⎬
= 2mn 2 ⎪ δk log2
+ δk δl log   .
 ⎩ k=1 R R(zk − zl ) ⎪

2π 2
δk k=1 l=1
l=k
k=1

This relation also holds if we replace the set Z by Z ∗ = {zk (η)}2mn


k=1 . Thus, we
must show that

M (B, ∂B, Z, , Ψ)  M (B, ∂B, Z ∗ , , Ψ).

By Lemma 4.3 there exists a dissymmetrization Dis such that

Dis zk (η) = zk , k = 1, . . . , 2mn.

It remains to use Theorem 4.15. 

g{l h3:
l?3 g{l )*h3:
l?3

Figure 5.1.

Inequality (5.2) contains (5.1). Indeed, taking m = 1, r1 = 1, and δ1 = 1 in


(5.2) we multiply both sides of (5.2) by R4n(2n−1) and pass to the limit as R → ∞.
Then we obtain
2n 2n 2n 2n
|zk − zl |  |zk (η) − zl (η)|. (5.3)
k=1 l=1 k=1 l=1
l=k l=k
5.1. Finite sets 125

Now let zk , k = 1, . . . , n, be some different points on the circle Γ. We can assume


that 0  arg z1 < arg z2 < · · · < arg zn < 2π for some values of the argument.
For sufficiently small η we pick arbitrary points zk on Γ, k = 1, . . . , 2n, such that
0  arg z1 < arg z2 < · · · < arg z2n
 
< 2π, the closed arc z2k−1 
z2k of Γ contains the
 
point zk and |z2k−1 − z2k | = η/n, k = 1, . . . , n. Now we derive (5.1) from (5.3) by
writing (5.3) for the points zk , k = 1, . . . , 2n, dividing both sides by η n and taking
the limit as η → 0.
In a similar way, from Theorem 5.1 we obtain the following result. Let θk ,
k = 1, . . . , n, be numbers such that 0  θ1 < θ2 < · · · < θn < 2π, and let r1 , . . . , rm
be some distinct positive numbers. Let {zk }mn k=1 and {δk }k=1 be sets of points and
mn

numbers defined as in Theorem 5.1, but for just n quantities θk instead of the 2n
quantities taken previously. Then
mn mn mn mn
|zk − zl |δk δl  |zk∗ − zl∗ |δk δl , (5.4)
k=1 l=1 k=1 l=1
l=k l=k

where the zk∗ have the same moduli as the zk : |zk∗ | = |zk |, k = 1, . . . , mn, but
have arguments θk∗ = 2π(k − 1)/n, k = 1, . . . , n. Taking particular values of rk
and δk in (5.4) we obtain various elementary inequalities. For example, we can set
r1 = 1 − ρ, r2 = 1 + ρ (ρ > 0), δ1 = −1, δ2 = 1 (m = 2) and passing to the limit
as ρ → 0, after simple calculations, obtain the relation


n
n
1 n
1
n
n(n2 − 1)
2
 ∗ ∗ 2
= , (5.5)
|zk − zl | |zk − zl | 12
k=1 l=1 k=1 l=1
l=k l=k

which holds for any different points zk on the circle Γ, k = 1, . . . , n,.


We can give no nontrivial lower estimate for the left-hand side of (5.1) with-
out imposing additional constraints on the zk , k = 1, . . . , n. In this connection the
following result is of interest.

Theorem 5.2. Let {zk }N k=1 be a system of different points in C \ {0}, and let Dj ,
j = 1, . . . , n, be open pairwise nonoverlapping infinite sectors with vertex at the
origin such that each point in {zk }N k=1 lies on the common boundary of two different
sectors in the set {Dj }nj=1 , and the boundary of each Dj contains a point in this
system. Let pj (z) = cj z π/θj for z ∈ Dj , where |cj | = 1 and θj is the opening angle
of Dj ; then pj maps Dj conformally and univalently onto the right half-plane,
j = 1, . . . , n. Let {δk }N
k=1 be a system of real numbers such that

 2 ⎛ ⎞2

N n
π ⎝
δk = δk ⎠ . (5.6)
2θ j
k=1 j=1 zk ∈Dj
126 Chapter 5. Metric Properties of Sets and Condensers

Then
N N
|zk − zl |δk δl (5.7)
k=1 l=1
l=k
n
   δk 
2

π 2
π −1 δk δl
 |zk | θj |pj (zk ) − pj (zl )| 2 .
θj
j=1 zk ∈Dj zk ∈Dj zl ∈Dj
l=k

Equality holds in (5.7), for example, for a symmetric system of points {zk }N k=1 of
the form {rs exp(2πi(k − 1)/n) : k = 1, . . . , n, s = 1, . . . , m} with different positive
rs , s = 1, . . . , m, the sectors Dj∗ = {z : 2π(j − 1)/n < arg z < 2πj/n}, j =
1, . . . , n, and numbers {δk }N
k=1 such that δk = δl if |zk | = |zl |, 1  k, l  N .

Proof. We can assume that all the δk , k = 1, . . . , N , are distinct from zero. For
some sufficiently large R and sufficiently small r, 0 < r < R, we can define the
condenser C := C(r; B, ∂B, Z, , Ψ), where B = {z : |z| < R}, Z = {zk }N k=1 ,
= {δk }N
k=0 , where δ 0 = 0, and Ψ = {r, . . . , r}. Let {C } n
j j=1 be the result of the
separating transformation of C with respect to the family of functions {pj }nj=1 .
By Theorem 4.8
1
n
cap C  cap Cj . (5.8)
2 j=1
Taking the symmetric system of points {zk }N k=1 and numbers {δk }k=1 indicated at
N

the end of the statement of Theorem 5.2 we can assume that C is mirror-symmetric
relative to the rays arg z n = 0. Then we conclude from the symmetry principle
and the choice of the sectors Dj∗ , j = 1, . . . , n, that equality sign holds in (5.8).
Now asymptotic formula (2.10) yields
N
N
δk2 R2 − |zk |2
cap C = − 2π − 2π δk2 log
log r R
k=1

k=1
   2  2 
N N
 R 2 − z k zl  1 1
+ δk δl log   +o , r → 0,
R(zk − zl )  log r log r
k=1 l=1
l=k
 2π 2π
δk2
R θj − |zk | θj
cap Cj = − 2π − 2πlog π π δk2
log r π −1
zk ∈Dj zk ∈Dj R θj |zk | θj
θj
 2π 
 R θj − p (z )p (z ) 
 j k j l 
+ δk δl log π 
 θj 
zk ∈Dj zl ∈Dj R (pj (zk ) − pj (zl ))
l=k
 2  2 
1 1
× +o , r → 0,
log r log r
5.1. Finite sets 127

j = 1, . . . , n. Substituting this into (5.8) we obtain


  δk2   2 
N
R2 − |zk |2
N N
 R − z k zl δk δl
 
R  R(zk − zl ) 
k=1 k=1 l=1
l=k
⎧⎡ ⎤ ⎫
⎪  2π
2
2π  δk  2π  δk2δl ⎪

⎨⎢  R θj − |z | θj  2 ⎥   ⎪

n
  ⎥  R θj − pj (zk )pj (zl ) 
 ⎢  π
k
 ⎦  
⎪⎣  θj π π
−1   π  ⎪
j=1 ⎪
⎩ z k ∈D j R |zk | θj  θ
zk ∈Dj zl ∈Dj R j (pj (zk ) − pj (zl ))
 ⎪

θj l=k

where equality holds for symmetrically positioned points {zk }N k=1 , domains
{Dj∗ }nj=1 , and numbers {δk }Nk=1 as above. In view of (5.6), we can cancel out R
N
raised to the power ( δk )2 on both sides, and then taking the limit as R → ∞
k=1
we obtain the required inequality (5.7) with the same equality case as before. 

Now we present two particular cases of (5.7). If each ray arg z = 2π(k − 1)/n
contains a single point zk , k = 1, . . . , n, then setting Dk = Dk∗ , k = 1, . . . , n, and
δk = 1, k = 1, . . . , n, we easily verify condition (5.6). From (5.7) we obtain
n n n n n
n n n
|zk − zl |  |zk | 2 −1 (|zk | 2 + |zk+1 | 2 ), (zn+1 = z1 ).
2
k=1 l=1 k=1
l=k

Replacing the arithmetic mean by the geometric mean we finally find


n n n
|zk − zl |  nn |zk |n−1 .
k=1 l=1 k=1
l=k

Note that condition (5.6) holds if we have δk = 0 for all j = 1, . . . , n. In
zk ∈Dj
particular, let zk , k = 1, . . . , 2n, be points on the circle Γ with 0  arg z1 <
arg z2 < · · · < arg zn < 2π for some values of their arguments. Setting δk = (−1)k
and Dk = {z : arg zk < arg z < arg zk+1 } in (5.7), k = 1, . . . , 2n (here arg z2n+1 =
arg z1 + 2π), we obtain
2n 2n
k+l
2n
π n 2n
|zk − zl |(−1)   .
2θj 2
k=1 l=1 k=1
l=k

(Here we have again used an inequality between classical mean values.)


Sometimes we can use conformal mappings to obtain inequalities for finite
subsets of arbitrary curves. We see this in the following simple example.
128 Chapter 5. Metric Properties of Sets and Condensers

Theorem 5.3. Let {zk }nk=1 be a subset of the ellipse

|z + 1| + |z − 1| = ρ + 1/ρ, ρ>1

which is mirror-symmetric relative to the real axis. Then


 n  n n  2n n
=
2 n −n2 ρ −1
|zk − 1| |zk − zl |  n 2 .
ρn
k=1 k=1 l=1
l=k

Proof. We set Z = {z1 , . . . , zn , ∞}, = {1, . . . , 1, −n}, and


=   $
Ψ= 2
|z1 − 1|/ρ r, . . . , 2
|zn − 1|/ρ r, 2r .

Then the condenser C(r; C, ∅, Z, , Ψ) is√mirror-symmetric relative to the real axis.


Let f be the regular branch of w = z + z 2 − 1 that maps C \ [−1, 1] conformally
and univalently onto the exterior |w| > 1 of the unit disc. By Lemma 4.1 and the
conformal invariance of capacity

cap C(r; C, ∅, Z, , Ψ) = cap C(r; C \ [−1, 1], ∅, Z, , Ψ)


= cap f (C(r; C \ [−1, 1], ∅, Z, , Ψ)) = cap C(r; {z : |z| > 1}, ∅, W, , Ψ),

where W = {f (z1 ), . . . , f (zn ), ∞} and Ψ = {r, . . . , r}. Let C ∗ be the condenser


mirror-symmetric relative to the circle |w| = 1 and coinciding with C(r; {z :
|z| > 1}, ∅, W, , Ψ) in |w| > 1. Obviously, C ∗ = C(r; C, ∅, {0, w1 , . . . , w2n , ∞},
{−n, 1, . . . , 1, −n}, {r, μ1 r, . . . , μ2n r, r}), where wk = 1/f (zk ) and μk = 1/ρ2 for
k = 1, . . . , n, while wk = f (zk−n ) and μk = 1 for k = n + 1, . . . , 2n. By the
symmetry principle

cap C ∗ = 2 cap C(r; {z : |z| > 1}, ∅, W, , Ψ).

Thus,
1
cap C(r; C, ∅, Z, , Ψ) = cap C ∗ .
2
By Theorem 2.5 the first condenser has capacity
⎡ ⎤
 n   2
2 = n n
n + n2 ⎢ 2n 2 ⎥ 1
−2π + 2π log ⎣ n |zk − 1| |zk − zl |⎦
log r ρ log r
k=1 k=1 l=1
l=k
 2 
1
+o , r → 0,
log r
5.1. Finite sets 129

and the condenser C ∗ has capacity


⎡ ⎤
2
  2n 2n  2
2(n + n ) ⎢ 1 ⎥ 1
−2π + 2π ⎣log |wk − wl |⎦
log r ρ2n log r
k=1 l=1
l=k
 2 
1
+o , r → 0.
log r

Hence  
2
n = n n 2n 2n
2 n
|zk2 − 1| |zk − zl | = |wk − wl |1/2 .
k=1 k=1 l=1 k=1 l=1
l=k l=k

It remains to use (5.4) and calculate the right-hand side of this inequality for
symmetrically positioned points (with r1 = 1/ρ, r2 = ρ, and δk = 1, k =
1, . . . , 2n). 

We find the right-hand side of (5.4) and other inequalities of this type using
elementary calculations. At the same time, an approach based on capacities is
also of interest. In fact, let wk∗ = (1/ρ) exp(2πi(k − 1)/n) for k = 1, . . . , n, wk∗ =
ρ exp(2πi(k − 1)/n) for k = n + 1, . . . , 2n, and D = {z : 0 < arg z < 2π/n}. By
symmetry considerations

cap C(r; C, ∅, {0, w1∗ , . . . , w2n



, ∞}, {−n, 1, . . . , 1, −n}, {r, . . . , r})
= n cap C(r; D, ∅, {0, w1∗ , w2∗ , wn+1
∗ ∗
, wn+2 , ∞}, {−n, 1, . . . , 1, −n}, {r, . . . , r})
= n cap C(r; C, ∅, {0, 1/ρn, ρn , ∞}, {−n, 1, 1, −n}, {rn, (n/ρn−1 )r, nρn−1 r, rn }).

From the asymptotic formula in Theorem 2.5 we conclude that


2n 2n  2n
ρ2n − 1
|wk∗ − wl∗ | =n 2n
.
ρn
k=1 l=1
l=k

Now let zk , k = 1, . . . , n, be points in the interval (−1, 1). We apply Theorem


5.3 to the points ζk on the ellipse such that Re ζk = Re ζn+k = zk , k = 1, . . . , n;
then the limiting procedure as ρ → 1 yields
 n  n n
= 2
1 − zk2 |zk − zl |  nn 2n−n .
k=1 k=1 l=1
l=k

Conversely, if we let ρ approach infinity, then the ellipse in Theorem 5.3 is


transformed into a circle, and the inequality in the theorem gives us (5.1).
130 Chapter 5. Metric Properties of Sets and Condensers

Exercises 5.1
(1) Show that equality holds in (5.1) only when zk = zk∗ eiθ , k = 1, . . . , n, where
θ is an arbitrary real number.
(2) Let {zk }nk=1 be a set of different points on the circle Γ and {zk∗ }nk=1 a set
of symmetrically positioned points on Γ. Let E be a closed subset of (0, 1],

n
g(z, ζ) the Green function of the domain U \ {zk r : r ∈ E}, and g ∗ (z, ζ)
k=1

n
the Green function of U \ {zk∗r, r ∈ E}. For 0 < r < 1 prove the inequality
k=1

2π2π 2π2π
g(reiϕ , reiθ )dϕdθ  g ∗ (reiϕ , reiθ )dϕdθ.
0 0 0 0

(3) Derive a corollary to inequality (5.7) by taking the points

zk = (1 − ρ) exp(i(2π(k − 1)/n)) and zn+k = (1 + ρ) exp(i(2π(k − 1)/n))

and numbers δk = −1 and δn+k = 1, k = 1, . . . , n, and passing to the limit


as ρ → 0.
(4) Generalize Theorem 5.2 by allowing points zk in the interior of the sectors
Dj , 1  k  N , 1  j  n.

5.2 Projections of sets


The orthogonal projection onto the real axis is a contraction mapping. It follows
from the geometric interpretation of the logarithmic capacity of compact sets
E ⊂ C that for their projections P E we have

cap E  cap P E.

Moreover,
 = (m(P E))/4,
cap P E  cap E
 = [−m(P E)/2, m(P E)/2] and m(·) is Lebesgue linear measure. Thus,
where E

cap E  (m(P E))/4 (5.9)

(Theorem 3.2 (d)). These results are also consequences of the properties of the
Steiner symmetrization (Corollary 4.2). In application to hyperbolic capacity,
Steiner symmetrization with respect to the real axis and then with respect to
the imaginary axis yields

caph E  caph P E  caph E (5.10)
5.2. Projections of sets 131

for each compact subset E of the disc U . In a similar way, Theorem 4.2 shows
that the circular projection onto the positive real half-axis does not increase the
logarithmic or hyperbolic capacity. Theorems 4.5–4.7 show that the projections
in elliptic or parabolic systems of coordinates do not increase the capacity of a
compact set either. The situation with the radial projection is more complicated.
For example, if E lies in a small neighbourhood of the origin, then its radial
projection onto the unit circle Γ = {z : |z| = 1} has a larger logarithmic capacity
than E, and conversely, if E ⊃ U , then cap E is no smaller than that of the
projection of E onto Γ.
Theorem 5.4. Let E be a compact subset of the plane C, 0 ∈ E, and let F =
{z/|z| : z ∈ E} be the radial projection of E onto the circle Γ. Then

(cap E) cap{z : 1/z ∈ E} exp(−2gC\E (0, ∞))  (cap F )4  (sin(l/4))4 ,

where l is the length of the projection F .

Proof. We set B = C \ E, G = C \ RE, and D = C \ F . Using Theorem 4.4,


the monotonic behaviour of the reduced modulus, and Corollary 4.3 we obtain in
turns

M (B, ∂B, {0, ∞}, {1, 1}, {r, r})  M (G, ∂G, {0, ∞}, {1, 1}, {r, r})
 M (D, ∂D, {0, ∞}, {1, 1}, {r, r})  M (Cr D, (∂ Cr D), {0, ∞}, {1, 1}, {r, r}).

By formula (2.16)

M (B, ∂B, {0, ∞}, {1, 1}, {r, r})


1
= {− log[(cap E) cap{z : 1/z ∈ E} + 2gB (0, ∞)}

and similar relations hold for other reduced moduli. For the third modulus in the
chain we take the projection F for E, and for the last we take an arc of Γ with
length l. It remains to note that if F is an arbitrary subset of Γ and D = C \ F ,
then
gD (z, 0) + log |z| = gD (z, ∞).
Hence
− log cap E = log r(D, 0) = log r(D, ∞) = gD (0, ∞),
and therefore
1
M (D, ∂D, {0, ∞}, {1, 1}, {r, r}) = − log cap F. 

By the projection PQ (z) of a point z in C onto a closed set Q ⊂ C we mean the


point closest to z in Q. The projection of a set E onto Q is PQ E = {PQ (z) : z ∈ E}.
132 Chapter 5. Metric Properties of Sets and Condensers

Theorem 5.5. Let E be a bounded closed set in the plane C and Q a closed infinite
sector in the plane with opening angle απ, 0 < α  1. Let p and l denote the linear
measures of the intersections of the projection PQ E with the sides of Q. Then

cap E  pl(4 − 2α)α/2−1 (2α)−α/2 . (5.11)
Equality sign is attained for the set E lying on the boundary of the sector Q and
formed by two line intervals with the same length and end-point at the vertex of
the sector.
Proof. We can assume that Q = {z : |arg z|  απ/2}. It is known that the
projection onto a closed convex set is a contraction mapping. By Theorem 3.1
cap E  cap PQ E  cap ((PQ E) ∩ (∂Q)) .
Using AKL-symmetrization we obtain
cap ((PQ E) ∩ (∂Q))  cap AKL((PQ E) ∩ (∂Q)),
and the set AKL((PQ E) ∩ (∂Q)) consists of two equal closed
√ subintervals of the
sides of ∂Q which go out of the origin and have length pl (see Theorem 4.12).
We finish the proof by calculating the logarithmic capacity of this set, which can
easily be performed by means of a well-known conformal mapping. 
Note that taking a suitable sector Q with opening angle π (α = 1), we obtain
from (5.11) the classical result (5.9). We can describe all the equality cases in
(5.11) with the help of Appendices A2 and A3.
Now we define the projection of a compact
4 F set E onto the symmetric ‘star’ formed by the n
rays arg z n = 0, n  1. To do this we subdivide
Qo F C into n sectors Dk , where Dk = {z : 2π(k −
3 1)/n − π/n < arg z < 2π(k − 1)/n + π/n},
k = 1, . . . , n, and in each sector take the family
Γk = {γk (r) : 0  r < ∞}√ of curves √ γk (r)
Qo F
defined by √ the equations Re z n = rn , z ∈
Dk , where√ 1 = 1, k = 1, . . . , n (for n = 1 the
branch of z is taken in D1 and is defined on
F ∂D1 by means of boundary correspondence).
5
We can easily see that for n = 1 the family Γ1
consists of parabolas, for n = 2, Γ1 and Γ2 are
Figure 5.2. families of parallel straight lines, and for n = 4
the families Γk , 1  k  4, consist of branches
of hyperbolae. By the projection of the compact set E onto the rays arg z n = 0 we
shall mean the set

n
Pn E = {r exp(i(2π(k − 1)/n)) : γk (r) ∩ E = ∅}.
k=1
5.2. Projections of sets 133

By Theorem 4.13 the logarithmic capacity of the projection satisfies


>
? n
?1
cap Pn E  cap SML E = @ n
m(2π(k − 1)/n, Pn E),
4
k=1

where m(θ, F ) is the linear measure of the intersection of F and the ray arg z = θ.
A weaker inequality, with m(θ, F ) replaced with M (θ, F ), is obtained from Szegő–
Marcus symmetrization SMA with A = {1/n, . . . , 1/n}. In the case when Pn E
consists of n line segments with end-point at the origin these two estimates coincide
and give a solution to the Fekete problem [S2]. For example, this is the case when
E is a continuum. Assuming that E ∩ Dk = ∅, k = 1, . . . , n, n  2, let mk be
the largest value of r such that the intersection of the continuum E with γk (r) is
nonempty. Then it immediately follows from Corollary 4.6 and (5.9) that
>
? n
n n
?1
= @
2 2
(k) 2/n (k) 2/n
cap E  (cap E )  (cap P E ) n
mk , (5.12)
4
k=1 k=1 k=1

where {E (k) }nk=1 is the result of the separating transformation√ of the set E with
respect to the √ family of functions {pk }nk=1 , where pk (z) = z n for z ∈ Dk , k =
1, . . . , n (here 1 = 1). For arbitrary sets we have the following result.
Theorem 5.6. The following inequality holds:
>
? n
?1
cap E  @ n
m(2π(k − 1)/n, Pn E).
4
k=1

Proof. Let E (k) be the sets defined above and



Ek = {z : arg z n = 0, | z n | ∈ P E (k) }, k = 1, . . . , n.

Similarly to (5.12) and in view of Exercise 3.1(6),


n n n
2
cap E  (cap P E (k) )2/n = (cap Ek )1/n  (cap LREk )1/n .
k=1 k=1 k=1

It remains to use Corollary 3.5:


>
?
n
?1 n
(cap LREk ) 1/n
 cap RA {LREk }nk=1 = @
n
m(2π(k − 1)/n, Pn E),
4
k=1 k=1

where A = {1/n, . . . , 1/n}. 


134 Chapter 5. Metric Properties of Sets and Condensers

Theorem 5.7. For an arbitrary closed subset E of the unit disc U ,


 √ 
π K( 1 − τ 2 )
caph E  exp − , (5.13)
2n K(τ )

where K(·) is the completeelliptic integral of the first kind, τ is defined by the
1n
equation τ −1/2 − τ 1/2 = n [L−1
k (E) − Lk (E)], 0  τ < 1, and Lk (E) =
k=1
) n
n 2 −1
2r dr, k = 1, . . . , n. Equality in (5.13) holds for the symmetric
{r:γk (r)∩E=∅}
sets
En∗ (t) = {z : |z|  t, arg z n = 0}, 0  t < 1.
Proof. We can assume that Lk (E) = 0, k = 1, . . . , n. We look at the following
condensers:
C = (E, C \ U ); {Ck }nk=1 is the result of the separating transformation
√ of C with
respect to the family √ of functions {p } n
k k=1 , where p k (z) = z n for z ∈ D ,
k
k = 1, . . . , n (here 1 = 1);
S Ck is obtained by applying Steiner symmetrization with respect to the real axis
to the condenser Ck and then using Steiner symmetrization with respect to
the imaginary axis;

Ck = ({ζ : |Re ζ|  Lk (E), Im ζ = 0}, Cζ \ Uζ );
Ck = ({w : Re w = 0, |Im w|  L−1 k (E) − Lk (E)}, [−2, 2]);
C ∗ = ({w : Re w = 0, |Im w|  t−n/2 − tn/2 }, [−2, 2]), where tn = τ is as in
Theorem 5.7;
throughout, the index k ranges between 1 and n.
Using Theorems 4.8 and 4.1 in turns we obtain

1 1
n n
cap C  cap Ck  cap S Ck .
2 2
k=1 k=1

Since the capacity is monotonic and conformally invariant (with respect to the
mapping w = iζ + i/ζ), we have
cap S Ck  cap Ck = cap Ck , k = 1, . . . , n.
Finally we use the Marcus radial averaging transformation (Theorem 3.13):

n
cap Ck  n cap C ∗ .
k=1

Note that if E = En∗ (t) and therefore C = Cn∗ := (En∗ (t), C \ U ), then all the above
inequalities turn to equalities. Summing these relations we obtain
cap C  (n/2) cap C ∗ = cap Cn∗ .
5.2. Projections of sets 135

Hence
caph E  caph En∗ (t).
The easiest way to finding the hyperbolic capacity of a symmetric set is through
the equality
caph En∗ (t) = (caph E1∗ (tn ))1/n ,
which reduces the calculations to the well-known mapping of a circular annulus
to the Grötzsch ring. 

It is easy to see that Theorem 5.7 can be treated as a covering theorem.


Namely, for each closed subset E of the disc U we have
>
? n
?
@ [L−1 −1/2
− τ 1/2 ,
k (E) − Lk (E)]  τ
n
(5.14)
k=1

where τ is the solution of the equation


 √ 
π K( 1 − τ 2 )
caph E = exp − . (5.15)
2n K(τ )

Corollary 5.1. Under the assumptions of Theorem 5.7


>
?
? n

@
n
Lk (E)  τ,
k=1

where τ is the solution of equation (5.15). Here equality is attained at the sets
En∗ (t), 0  t < 1.

1
n
Proof. We can assume that Lk (E) = 0. We set xk = log[L−1
k (E) − Lk (E)] for
k=1
−1/2
k = 1, . . . , n, and x0 = log(τ − τ 1/2 ). The function
 A B$
1 x  2x
Ψ(x) = log e + e +4
2

is convex and increasing on the real axis. Thus, in view of (5.14),


 
1 1
n n
Ψ(x0 )  Ψ xk  Ψ(xk ).
n n
k=1 k=1

Returning to the original variables we obtain the required inequality. 


136 Chapter 5. Metric Properties of Sets and Condensers

Corollary 5.2. If E ⊂ U is a continuum joining a point z ∈ U to the radius


[0, −z/|z|], then
|z|  τ,
where τ is the solution of equation (5.15) with n = 1. If zk , where k = 1, . . . , n, n 
2, are points in E ⊂ U lying on n respective rays going at equal angles out of z = 0,
then >
? n
?
@
n
|zk |  τ 1/n ,
k=1

where τ satisfies (5.15). Here equality holds for continua E formed by n straight
line intervals of equal length going out of the origin z = 0 at equal angles and
joining the origin to the points zk , k = 1, . . . , n.

Proof. We can assume that the points zk , k = 1, . . . , n, lie on the rays arg z n = 0,
or, in the case of the first assertion we set z1 = z. It remains to observe that
|zk |n/2  Lk (E), k = 1, . . . , n. 

After a limiting procedure inequalities involving hyperbolic capacity yield


the corresponding inequalities for the logarithmic capacity cap E of the compact
set E.

Exercises 5.2
(1) Describe the equality cases in (5.9) and (5.12).
(2) Prove the inequality
>
?
?1 n
cap E  @
n 2/n
Lk (E).
4
k=1

For n = 2 this result goes back to Pólya.


(3) Let P (z) = z m + · · · be a polynomial of degree m  1, and let E(P ) = {z :
P (z) ∈ [0, 1]}. Show that
>
? n
?
@
n
Lk (E(P ))  21−n/m
k=1

and verify that equality holds in the case when m/n is an integer, for

1
P (z) = [Tm/n (22n/m−1 z n − 1) + 1] = z m + · · · ,
2
where Tm/n (ζ) is the Chebyshev polynomial.
5.3. Subsets of a circle and a line interval 137

5.3 Subsets of a circle and a line interval


Let Γ = {z : |z| = 1} and let E be a closed subset of Γ. By the length of E
we shall mean its Lebesgue linear measure. It is easy to see that the transfinite
diameter or, equivalently, the logarithmic capacity of a length-l arc of the circle
Γ is sin(l/4). Using the principle of circular symmetrization (Corollary 4.4) we
deduce the following result.
Theorem 5.8 (Beurling). The least possible value of the logarithmic capacity of a
closed subset of the circle Γ which has length l is sin(l/4); it is attained at arcs of Γ.
This result was already known to experts in the 1930s. What is the upper
bound for the logarithmic capacity? Note that by Fekete’s theorem ([Gol], Ch.
VII, § 1) the set En∗ := {z ∈ Γ : |arg z n |  l/2}, which also has length l, has
logarithmic capacity
cap En∗ = (sin(l/4))1/n → 1 = cap Γ, n → ∞.
Thus, if we impose no additional constraints on a set E ⊂ Γ, then we only have
the trivial upper bound. There was a conjecture that a set consisting of n or fewer
arcs cannot have a greater capacity than the set En∗ having the same length. An
affirmative answer here was given with the help of dissymmetrization [Hal].
Theorem 5.9 (Haliste). Let E be the union of n closed arcs of the unit circle Γ
with total length l. Then
cap E  cap En∗ = (sin(l/4))1/n .
Here equality holds if and only if the set E coincides with En∗ up to a rotation
about the origin.
Proof. This follows from Lemma 4.3 and Corollary 4.7. 
Note that since the analytic capacity A(E), introduced by Ahlfors [Ahl1],
satisfies the inequality A(E)  cap E, under the assumptions of Haliste’s theorem
A(E)  (sin(l/4))1/n .
Theorem 5.9 shows that a larger capacity is associated with a symmetric set which
is ‘as loosely scattered on the circle Γ as possible’. In this connection we can ask
whether for ‘sufficiently scattered’ sets we can have a better lower bound than
Beurling’s.
Theorem 5.10. Let E be a closed subset of the unit circle Γ and {σk }nk=1 , n 
1, be a decomposition of Γ into closed arcs σk without common interior points:
 n
k=1 σk = Γ. Then
n
2
cap E  [sin(lk /(2βk ))]βk /2 , (5.16)
k=1
where lk is the linear measure of the intersection of E with the arc σk , and βk π is
the length of σk , k = 1, . . . , n.
138 Chapter 5. Metric Properties of Sets and Condensers

Proof. Let Dk denote the open infinite sector with vertex at the origin that is
bounded by the rays passing through the end-points of the arc σk , let βk π be
the opening angle of Dk , and let pk (z) = αk z 1/βk , |αk | = 1, be the function
mapping Dk conformally and univalently onto the right half-plane, k = 1, . . . , n.
We can assume that E ∩ Dk = ∅, k = 1, . . . , n. Then we can consider the system
{E (k) }nk=1 obtained by the separating transformation of E with respect to the
family of functions {pk }nk=1 . By Corollary 4.6
n
2
cap E  (cap E (k) )βk /2 .
k=1

Beurling’s theorem completes the proof:

cap E (k)  sin(lk /(2βk )), k = 1, . . . , n. 

Now we look at subsets of the interval [−1, 1]. For a closed set e ⊂ [−1, 1] we
denote 
dx
μ(e) = √ .
1 − x2
e

Theorem 5.11. Let E be a closed subset of the interval [−1, 1], and let {ek }nk=1 ,
n  1, be a decomposition of [−1, 1] into closed intervals ek with disjoint interiors:

n
ek = [−1, 1]. Then
k=1

n  2(μ(ek ))2 /π2


1 πμ(ek ∩ E)
cap E  sin .
2 2μ(ek )
k=1

Proof. We consider the subset F of the circle Γ which is mirror-symmetric relative


to the real axis and whose orthogonal projection onto this axis is equal to E. The
decomposition {ek }nk=1 induces a decomposition {σk }2nk=1 of Γ: it consists of the 2n
closed arcs of the circle whose orthogonal projections onto the real axis coincide
with the intervals ek , 1  k  n. (The points ±1 are end-points of some of these
arcs.) Note that if for some k and l the projection of σk is el , then linear Lebesgue
measure m(σk ) of the arc σk is equal to μ(el ) and m(σk ∩ F ) = μ(el ∩ E). For the
set F and decomposition {σk }2n k=1 inequality (5.16) takes the following form:

2n  (m(σk ))2 /(2π2 ) n  (μ(ek ))2 /π2


πm(σk ∩ F ) πμ(ek ∩ E)
cap F  sin = sin .
2m(σk ) 2μ(ek )
k=1 k=1

It remains to use Robinson’s equality



cap F = 2 cap E

(see Exercise 2.1(14)). 


5.3. Subsets of a circle and a line interval 139

In the case when E is a union of finitely many intervals, Theorem 5.11 can
be improved as follows.

n
Theorem 5.12. Let E = [ak , bk ], where −1 = a1 < b1 < a2 < b2 < · · · < an <
k=1
bn = 1, n  2. Then


n   $ θ(δk )−θ(δk−1 ) 2
1 1 π(θ(bk ) − θ(δk )) π(θ(ak ) − θ(δk )) π
capE  max cos − cos ,
2 2 θ(δk−1 ) − θ(δk ) θ(δk−1 ) − θ(δk )
k=1

where θ(x) = cos−1 x and the maximum is taken over all the δk such that δ0 =
−1, δn = 1, bk < δk < ak+1 , k = 1, . . . , n − 1.
Proof. We look at the subset F of the circle Γ which is mirror-symmetric relative
to the real axis and whose orthogonal projection onto the axis equal to E. For
the δk in the hypotheses of Theorem 5.12 we set θk = θ(δn−k+1 ), k = 1, . . . , n +
1, and θk = 2π − θ2n−k+2 for k = n + 2, . . . , 2n + 1. The system of functions
{pk }2n k=1 , where ζ = pk (z) ≡ −i(e
−iθk π/(θk+1 −θk )
z) for θk < arg z < θk+1 , k =
1, . . . , 2n, and for suitably chosen branches of power functions, is admissible for
a separating transformation of domains with respect to z = ∞. By Robinson’s
equality and Corollary 4.6
 2n
2
/(2π 2 )
2 cap E = cap F  (cap F (k) )(θk+1 −θk ) , (5.17)
k=1

where F (k) is the union of the sets pk ({z ∈ F : θk  arg z  θk+1 }) and its
reflection in the imaginary axis, k = 1, . . . , 2n. Each of these sets is one or two
arcs of Γ which are symmetric relative to the imaginary axis. The orthogonal
projection of F (k) onto the imaginary axis coincides with the orthogonal projection
of F (2n−k+1) onto the same axis; the length of this projection is
π(θ(bn−k+1 ) − θ(δn−k+1 )) π(θ(an−k+1 ) − θ(δn−k+1 ))
lk := cos − cos ,
θ(δn−k ) − θ(δn−k+1 ) θ(δn−k ) − θ(δn−k+1 )
k = 1, . . . , n. Taking account of Robinson’s equality again we obtain

cap F (k) = cap F (2n−k+1) = lk /2, k = 1, . . . , n.
Substituting these values in (5.17) and making the change of index n − k + 1 → k
we obtain the required estimate. 

n
Theorem 5.13. If E = [ak , bk ], −1 = a1 < b1 < a2 < b2 < · · · < an < bn = 1,
k=1
n  2, then
  1/(n−1)
1
n−1
1 −1 −1
cap E  cos (cos ak+1 − cos bk ) .
2 2
k=1
140 Chapter 5. Metric Properties of Sets and Condensers

Proof. Let θ(x) := cos−1 x and let F ⊂ Γ be the set defined as in the proof of
Theorem 5.12. It consists of 2(n − 1) arcs of the circle Γ, with total length

n−1
n−1
l := 2θ(an ) + 2(θ(ak ) − θ(bk )) + 2(π − θ(b1 )) = 2π + 2 (θ(ak+1 ) − θ(bk )).
k=2 k=1

Using successively Robinson’s equality and Theorem 5.9 we obtain


1 1 1
cap E = (cap F )2  (cap E2(n−1)

)2 = (sin(l/4))1/(n−1) . 
2 2 2
Although the proofs in this section are short, this does not mean the prob-
lems under consideration are trivial. For instance, arguments based on the values
of extremal capacities meet here with significant difficulties. For logarithmic ca-
pacity we have the simple formula cap En∗ = (sin(l/4))1/n . However, already for
hyperbolic capacity the calculations of the quantity caph E ∗ (or equivalently, of
the capacity of a condenser whose plates are arcs of concentric circles) is quite
cumbersome. It is even more difficult to find δ(E ∗ ) and δh (E ∗ ) (see Exercises
5.3(3) and 5.3(4)). It is perhaps for this reason that the problem of finding upper
bounds for logarithmic capacity, which we mentioned above, remained open for a
long period of time (Theorem 5.9).

Exercises 5.3
(1) In Theorem 5.10 let σk = {z ∈ Γ : |arg z −2πk/n|  π/n}, k = 1, . . . , n. Show
that equality in (5.16) holds only for even n and for E formed by n/2 equal
arcs with midpoints either at the points exp(i(π/n+ 4πk/n)), k = 1, . . . , n/2,
or at the points exp(i(−π/n + 4πk/n)), k = 1, . . . , n/2.
(2) Describe all the equality cases in the inequality in Theorem 5.12.
(3) For 0 < r1 < r2 < 1 let E be a union of n closed arcs of the circle |z| = r2 with
total length l and let E ∗ = {z : |z| = r2 , |arg z n |  l/(2r2 )} be the symmetric
set with the same length. By the Robin capacity of E with respect to the
domain Ω(E) = {z : |z| > r1 , z ∈ E} we mean the quantity
δ(E) = (r(Ω(E), E, ∞))−1 .
Prove the inequality
δ(E)  δ(E ∗ ).
(4) In the notations of the previous exercise the quantity
δh (E) = exp(−2π(cap C(E))−1 ),
C(E) = (Ω(E) ∩ U, {E, Γ}, {0, 1}), is called the hyperbolic Robin capacity of
E with respect to Ω(E) ∩ U . Prove that
δh (E)  δh (E ∗ ).
5.4. Polygons 141

5.4 Polygons
In § 1.3 we defined a quadrilateral (Q; a, b, c, d) and its modulus M (Q; a, b, c, d). In
particular, we mentioned there that
M (Q; a, b, c, d) = cap(Q, {(b, c), (d, a)}, {0, 1}).
If the sides of a quadrilateral are straight line segments, then we call it a polyg-
onal quadrilateral. The simplest properties of the moduli of quadrilaterals are
consequences of the monotonicity of the capacity of generalized condensers and
composition principles (see, for instance, Exercises 1.3 (1–6)). Some deeper prop-
erties follow from the polarization and symmetrization of condensers. For example,
polarization with respect to the bisector of the infinite sector in Figure 3.2 does
not reduce the modulus of a polygonal quadrilateral with respect to its sides lying
on the sides of the sector. The same holds for sectors with zero opening angle,
that is, strips.
Theorem 5.14. Let (Q; a, b, c, d) be a quadrilateral in the strip 0 < Re z < 1 such
that the sides (a, b) and (c, d) are intervals of the straight lines Re z = 1 and
Re z = 0, respectively. Then
M (Q; a, b, c, d)  M (St Q; a , b , c , d ),
where Re a = Re b = 1, Re c = Re d = 0, Im a = −Im b = Im(a − b)/2, and
Im c = −Im d = Im(c − d)/2.
Proof. We introduce some notation:
u is the potential function of the condenser (Q; {(b, c), (d, a)}, {0, 1});
γ = {z ∈ Q : u(z) = 1/2};
Q1 = {z ∈ Q : 0 < u(z) < 1/2}; (St Q)+ = {z ∈ St Q : Im z > 0};
Q2 = {z ∈ Q : 1/2 < u(z) < 1}; (St Q)− = {z ∈ St Q : Im z < 0}.
Then we have the following chain of relations
M (Q; a, b, c, d) = cap(Q; {(b, c), (d, a)}, {0, 1})
= cap(Q1 , {(b, c), γ}, {0, 1/2}) + cap(Q2 , {γ, (d, a)}, {1/2, 1})
= cap(Q, {γ, (b, c) ∪ (d, a)}, {1/2, 1})
 cap(St Q, {[0, 1], (b, c ) ∪ (d , a )}, {1/2, 1})
= cap((St Q)+ , {(b , c ), [0, 1]}, {1, 1/2})
+ cap((St Q)− , {[0, 1], (d , a )}, {1/2, 1})
= cap(St Q, {b , c ), (d , a ), {0, 1})
= M (St Q; a , b , c , d ).
Here we have used Theorem 4.1 and the symmetry principle. 
142 Chapter 5. Metric Properties of Sets and Condensers

Uv Q Uv

Figure 5.3.

In Figure 5.3 we depict the result of the successive application of the Steiner
symmetrization with respect to the real axis, polarization relative to the line α,
and the Steiner symmetrization again to a polygonal quadrilateral. At each step
the modulus of the quadrilateral with respect to the sides lying in the strip remains
the same or decreases.
In the same strip 0 < Re z < 1 we take a domain G containing the interval
(−1, 1) and having the following property: if z0 ∈ G, then G contains also the ray
{z : Re z = Re z0 , Im z  Im z0 }. Let 1 + iα and iβ (0 < α  ∞, 0 < β  ∞)
be two different marked attainable boundary points of G, and let γ be the part of
the boundary of G confined between 1 + iα and iβ.
Let γ(t) be the straight line y = 2tx − t, depending to the parameter t, δ1 <
t < δ2 , where δ1 = inf{t : γ(t) ∩ γ = ∅} and δ2 = sup{t : γ(t) ∩ γ = ∅}, and let Q(t)
be the part of G lying over γ(t). We shall study
3 , j the behaviour of the function

g(t) = M (Q(t); 1 + it, 1 + iα, iβ, −it),
j

R)u* δ1 < t < δ2 , or in other words, we shall examine


how the conformal modulus of the quadrilateral
ju u changes as one of its sides rotates about the
3>4 3 point (1/2, 0).
2
3 , ju Theorem 5.15. The function g(t) is strictly
convex on the interval (δ1 , δ2 ). In addition, if
{1 − z : z ∈ G, Re z  1/2} ⊂ {z ∈ G : Re z 
1/2}, then g(t) is strictly decreasing on (δ1 , 0)
and g(t)  g(−t) for each t ∈ (δ1 , 0). Here
Figure 5.4: Variation of the mod- equality sign holds if and only if both G and
ulus of a quadrilateral with ro- the set γ are mirror-symmetric relative to the
tating side. straight line Re z = 1/2.
5.4. Polygons 143

Proof. We set P = {z : 0 < Re z < 1}; Et0 = {x + iy : 0  x  1, y  2tx − t};


and let E1 be the set of points in the closed strip P which lie over γ (including γ
itself).
The function g(t) has the following representation:

g(t) = cap(Q(t), {γ, [−it, 1 + it]}, {1, 0}) = cap(P, {Et0 , E1 }, {0, 1}).

Applying Theorem 3.12 to arbitrary t1 , t2 ∈ (δ1 , δ2 ) and positive numbers α1 , α2


such that α1 + α2 = 1 we conclude that

α1 g(t1 ) + α2 g(t2 ) > g(α1 t1 + α2 t2 ),

which means that g(t) is convex on (δ1 , δ2 ).


Now let σ be the straight line Re z = 1/2, which is oppositely directed from
the imaginary axis, and assume that δ1 < t < 0. By Theorem 3.5

g(t) = cap(P, {Et0 , E1 }, {0, 1})  cap Pσ (P, {Et0 , E1 }, {0, 1})
= cap(P, {E−t0 , E1 }, {0, 1}) = g(−t),

where equality holds only when G and γ are mirror-symmetric relative to σ.


That g(t) is monotone follows from the above and its convexity:
(t2 + t1 )g(t1 ) − (t2 − t1 )g(−t1 )
g(t2 ) <  g(t1 )
2t1
for δ1 < t1 < t2 < 0. 
The general properties of condensers and of the transformations considered in
Chapters 3 and 4 are very useful for isoperimetric inequalities. Here we consider
the Pólya–Szegő problem of the maximum inner radius of an n-gon with pre-
scribed area. If M3 is a triangle with area S, then performing repeatedly Steiner
symmetrization with respect to its heights, we obtain in the limit an equilateral
triangle M3∗ with the same area S. It follows from the symmetrization principle
in Corollary 4.1 that the inner radius of the equilateral triangle M3∗ with respect
to its centre is the maximum inner radius possible. We can show by similar argu-
ments that a square has the largest inner radius among all the quadrangles M4
with fixed area S. Pólya and Szegő mentioned that “to prove (or disprove) the
analogous theorems for regular polygons with more than four sides is a challenging
task” ([PS], § 7.4).
Theorem 5.16. [Sol4] If Mn is an n-gon with area S and z0 ∈ Mn , then


1 C
2/n Γ
1 − n n π
r(Mn , z0 )  2 1 1 S tan , (5.18)
Γ 2−n π n

where Γ(·) is the Euler gamma function. Equality is attained in (5.18) only for a
regular n-gon with centre at z0 .
144 Chapter 5. Metric Properties of Sets and Condensers

Proof. We partition the star of Mn with respect to z0 into m  n triangles Dk



m
with vertex z0 . It is sufficient to prove (5.18) in the case when S = Sk , where
k=1
Sk is the area of the triangle Dk , k = 1, . . . , m. Let 2αk π be the angle of Dk at the
vertex z0 , and let pk be the function mapping Dk conformally and univalently onto
the right half-plane Re ζ > 0 and taking the vertices of the triangle to 0, −i, ∞,
(pk (z0 ) = 0). By Theorem 4.10


m
m
log r(Mn , z0 )  2α2k log(r(C \ γ, 0)/dk ) = 2α2k log(4/dk ), (5.19)
k=1 k=1

where γ = {ζ : Re ζ = 0, Im ζ  −1}, and the constants dk are defined by the


asymptotic equalities

|pk (z)| ∼ dk |z − z0 |1/(2αk ) , z → z0 , k = 1, . . . , m.

No Do

2
{2
El
j

Figure 5.5.

We have equality sign in (5.19) if and only if Mn is mirror-symmetric relative


to all the sides of the triangles Dk which have an end-point at z0 . By (2.5)

2αk log(4/dk ) = log r(Dk , Dk ∩ (∂Mn ), z0 ), k = 1, . . . , m. (5.20)



Let Bk = {z : |arg z| < αk π : 0 < Re z < Sk cot αk π}, let Γk be the
side of Bk opposite to the vertex z = 0, let Bk+ = Bk ∩ {z : Im z > 0} and
Γ+k = Γk ∩ {z : Im z  0}, k = 1, . . . , m. From Exercise 3.5(7) we conclude that

4αk Sk cot αk π
r(Dk , Dk ∩ (∂Mn ), z0 )  r(Bk , Γk , 0) = r(Bk+ , Γ+
k , 0) = , (5.21)
αk B(αk , 1/2)
5.4. Polygons 145

k = 1, . . . , m, where B(·, ·) is the Euler integral of the first kind. We calculate the
Robin radius r(Bk+ , Γ+ k , 0) by formula (2.5), where f is the well-known conformal
mapping of the triangle Bk+ onto the upper half-plane taking the vertices of Bk+
to 0, 1, and ∞ (f (0) = 0). Equality holds in (5.21) if and only if Dk is an isosceles
triangle with apex z0 . Summing (5.19)–(5.21) we obtain
m  √ 
Sk cot αk π
2
log r(Mn , z0 )  αk log 4 + αk log =: F ({αk }, {Sk }),
αk B(αk , 1/2)
k=1

with equality holding only for the regular n-gon (n = m) with centre at z0 .
Now we look at the following extremal problem:
F ({αk }, {Sk }) → sup (5.22)
under the conditions

m
m
αk = 1, Sk = S (5.23)
k=1 k=1
and constraints
0 < αk < 1/2, 0 < Sk < S. (5.24)
As a point (αk , Sk ) approaches the boundary of the rectangle {(αk , Sk ) :
0  αk  1/2, 0  Sk  S}, the corresponding term in the sum F ({αk }, {Sk })
has a nonpositive upper limit. Hence the function F ({αk }, {Sk }) takes its maxi-
mum under conditions (5.23) in the interior of the parallelepiped described by the
constraints (5.24).
Now we write out the Lagrange function of the problem (5.22), (5.23):

m
m
Φ({αk }, {Sk }, λ1 , λ2 ) = F ({αk }, {Sk }) + λ1 αk − (2λ2 )−1 Sk ,
k=1 k=1

and the system of equations for stationary points:


αk (2Sk )−1 − (2λ2 )−1 = 0, (5.25)
 
1 1
2αk log 4 + log(Sk cot αk π) − log αk Γ(αk ) + log Γ αk + (5.26)
2 2
 
1 1
− αk ψ(αk ) + αk ψ αk + − παk sin−1 2παk − log π − 1 + λ1 = 0,
2 2
where ψ(z) is the logarithmic derivative of the gamma function. By (5.25)
Sk = αk λ2 . (5.27)
Substituting (5.27) into (5.26) and dropping the subscript of αk for brevity we
obtain the equation
1
R(α) = 1 + log π − λ1 , (5.28)
2
146 Chapter 5. Metric Properties of Sets and Condensers

where
1 1
R(α) = 2α log 2 − log(tan απ) − απ tan απ − log α
2  2
1
− απ cot 2απ − log Γ(α) + log Γ α +
2
 
1 1
− αψ(α) + αψ α + + log λ2 .
2 2

We claim that for fixed λ1 and λ2 equation (5.28) has one solution or none.
For a proof we calculate the derivative R (α):
       
 1 1 1
R (α) = 4log2 − π cotαπ − cot α + π + απ cot α + π − cotαπ
2 2 2
   
1 1 1
− α−1 − 2ψ(α) + 2ψ α + − αψ  (α) + αψ  α + .
2 2 2
Using the formula ψ(z) = ψ(1 − z) − π cot πz we transform this expression as
follows:
   
 −1 1 1
2R (α) = 8 log 2 − α + 2ψ + α − 2ψ(1 − α) + 2ψ −α
2 2
   
1 1
− αψ  (α) + αψ  + α − αψ  − α + αψ  (1 − α) − 2ψ(α).
2 2
After some calculations we obtain

R (α) → 0 as α → +0. (5.29)

Next we find R (α):


    
 −2   1   1
2R (α) = α − 3 ψ (α) − ψ + α − ψ (1 − α) + ψ −α
2 2
     
  1  1 
− α ψ (α) − ψ +α −ψ − α + ψ (1 − α) .
2 2
Using the expansions of ψ  (z) and ψ  (z) in power series we find

  −2
1

2R (α) = − 3 (k + 1 + α)−2 − k + + α − (k + 1 − α)−2
2
k=0
 −2  ∞
  −3
1 1
+ k+ −α + 2α (k + 1 + α)−3 − k + + α
2 2
k=0
 −3 
1
+(k + 1 − α)−3 − k + − α . (5.30)
2
5.4. Polygons 147

If 0 < α < 1/2, then the second sum in (5.30) is negative, while the first is
decreasing and equal to 0 for α = 0. Thus R (α) < 0. Hence it follows from the
limit relation (5.29) that R(α) is monotonic for 0 < α < 1/2. This means that if the
system (5.23), (5.25), (5.26) is solvable, then all the quantities α1 , α2 , . . . , αm are
equal. Using this observation in combination with (5.23), (5.27), and (5.28) we find
the unique solution of (5.25), (5.26) under the constraints (5.24): α1 = α2 = · · · =
αm = 1/m, S1 = S2 = · · · = Sm = S/m, λ2 = (2S)−1 , λ1 = 1 + 12 log π − R(1/m).
Substituting these values of α1 , α2 , . . . , αm and S1 , S2 , . . . , Sm in F ({αk }, {Sk })
and using the formula for the inner radius of a regular n-gon Mn∗ ([PS], p.330) we
obtain

r(Mn , z0 )  r(Mm , z0 )  r(Mn∗ , z0 ),
which coincides with (5.18). If equality holds in (5.18), then Mn is a regular poly-
gon. 
The reader can find other examples of isoperimetric inequalities for polygons
derived by symmetrization in [Sol1] (see also Exercises 4.4(4) and 4.4(5)).

Exercises 5.4
1. (Vuorinen [DV]) Assume that 0 < r < 1, Re b < 1 + r, Im b > 0, |b − 1| > r,
and let 1 + reiϕ0 be the point of tangency between the circle |z − 1| = r and a
straight line through b that is closer to 1+r. Show that the conformal modulus
M (Q; 1 + reiϕ , b, 0, 1) of a polygonal quadrilateral is strictly increasing on
[0, ϕ0 ].
2. Establish the following asymptotic formula for polygonal quadrilaterals:

M (Q(h); t, t + i, i, heiϕ ) = t + (t sin ϕ − cos ϕ)h/2 + o(h), h → 0.

Here t > 0 and ϕ ∈ (0, π/2) are some fixed numbers.


3. (Vuorinen [DV]) Show that the modulus M (Q(y); 1 + iα, 1 + iy, iγ, iδ) of a
polygonal quadrilateral with γ > δ is a decreasing convex function on the
interval (α, ∞).
4. For real numbers ϕ, r1 , and r2 such that 0 < ϕ < π/2 and 1/ cos ϕ < r1  r2 ,
prove that the conformal modulus

q2 (r) = M (Q2 (r), re−iϕ , r2 e−iϕ , r1 eiϕ , (2 cos ϕ − 1/r)−1 eiϕ )

of a polygonal quadrilateral has the following properties:


1) q2 (r) is decreasing on the interval ((2 cos ϕ − 1/r1 )−1 , 1/ cos ϕ),
2) q2 (r) > q2 ((2 cos ϕ − 1/r)−1 ) for r ∈ ((2 cos ϕ − 1/r1 )−1 , 1/ cos ϕ),
3) the function q2 (1/ρ) is convex on (1/r2 , 2 cos ϕ − 1/r1 ).
This result characterizes the behaviour of the modulus of a quadrilateral as
one of its sides rotates about the point z = 1.
148 Chapter 5. Metric Properties of Sets and Condensers

5.5 Ring domains


We mean by a ring domain an arbitrary doubly connected domain G in the Rie-
mann sphere C. Let M G be the modulus of G with respect to the family of curves
separating its boundary components, and let E0 and E1 be the components of
the complement to G. We associate the condenser (E0 , E1 ) with G. It is known
that M G = |(E0 , E1 )| = (cap(E0 , E1 ))−1 (Exercise 1.2(6)). The most common
‘canonical’ domains are (circular) annuli and Grötzsch and Teichmüller rings. The
last two rings are obtained from the domains G(P ) = {z : |z| > 1} \ [−∞, −P ]
with P > 1 and T(P ) = C \ ([−∞, −P ] ∪ [0, 1]) with P > 0, respectively, by
motions of C.
The following results are well known.
Lemma 5.1 (Rengel). Among the ring domains G with fixed finite area in the
logarithmic metric (2π|z|)−1 |dz| which separate the points 0 and ∞ only the annuli
bounded by concentric circles with centre at the origin have the maximum modulus.
Lemma 5.2 (Grötzsch). Among the ring domains such that one component of their
complement contains the unit disc and the other contains the points −P and ∞,
only the ring G(P ) has the maximum modulus.
Lemma 5.3 (Teichmüller). Among the ring domains separating the pair of points
{0, 1} from the pair {∞, −P } only the ring T(P ) has the maximum modulus.

Rengel’s lemma follows from the Szegő–Marcus symmetrization principle and


Grötzsch’s and Teichmüller’s lemmas follow from the principle of circular sym-
metrization (see also Exercises 1.4 (1 and 2)).
We shall require analogues of these lemmas for arbitrary condensers (E0 , E1 ),
when the field G = C \ (E0 ∪ E1 ) is not necessarily a ring domain. Throughout
§ 5.5 we use the notation

Dk = {reiθ : 0 < r < ∞, |θ − 2πk/n| < π/n}, k = 1, . . . , n, n  1.

In particular, if n = 1 then D1 = C. We say that a connected component G of


the set G ∩ Dk , 1  k  n, n  1, separates the points 0 and ∞ in Dk if the
complement Dk \ G is the union of two disjoint closed sets E and F such that
one of them contains 0, the other contains ∞, and furthermore E ∩ ∂G ⊂ E0
and F ∩ ∂G ⊂ E1 . An analogue of Rengel’s lemma, giving no description of the
equality case, can be stated in terms of condensers as follows.
Lemma 5.4. If the field of a condenser (E0 , E1 ) has a component which separates
0 and ∞ in C and has area in the logarithmic metric not exceeding some quantity
S, 0 < S < ∞, then
|(E0 , E1 )|  M {z : r < |z| < R}, (5.31)
where r and R are arbitrary real numbers such that S = (1/2π) log(R/r).
5.5. Ring domains 149

Proof. This easily follows from the Marcus–Szegő averaging symmetrization prin-
ciple (4.17). 
Now we go over to generalizing Grötzch’s and Teichmüller’s lemmas.
Throughout, t is a complex parameter with Re t > 0; w = dt (z) = (z − t)/(z + t)
is a linear fractional function, and d−1 t (w) is its inverse function. For arbitrary
complex numbers a and b and real numbers r and R such that Re a > 0, Re b > 0,
a = b, 0  r < R  ∞, r < |b|  R, and r  |a| < R let G(t, r, a, b, R) denote
the doubly connected domain {z : r < |z| < R , z ∈ [−a , a ] ∪ [−R, −b ] ∪ [b , R]},
where the real numbers a , b , r , and R are defined by d|t| (a ) = −|dt (a)|,
d|t| (b ) = |dt (b)|, r = |d−1  −1 i(π−ϕ)
|t| (−e )|, and R = |d|t| (e

)|; here θ is the an-
gle at which the interval [−ir, ir] is seen from the point t and ϕ is the an-
gle at which [−iR, iR] is seen from t. Note that if t is positive, then r = r
and R = R; in addition, if a and b are positive and satisfy a  t  b, then
G(t, r, a, b, R) = {z : r < |z| < R, z ∈ [−a, a] ∪ [−R, −b] ∪ [b, R]}. Now if a = r
and R = ∞ or r = 0 and b = R, then G(t, r, a, b, R) can be obtained from the
Grötzsch ring G(P ) by a Möbius map while if r = 0 and R = ∞, then it is can be
obtained from the Teichmüller ring T(P ) by a similar map.
We say that a set E joins a point a to a set F if some connected component
of E contains both a and some point b ∈ F.
Lemma 5.5. Let a, b, r and R be as above and let (E0 , E1 ) be a condenser satisfying
the following conditions: the plate E0 contains the disc |z| < r and joins the point
a to −a, and the plate E1 contains the set |z| > R and joins b to −b. Then

|(E0 , E1 )|  M G(t, r, a, b, R) (5.32)

for each t with Re t > 0. In addition, if the field G of the condenser (E0 , E1 ) is a
ring domain, then equality in (5.32) holds only when G coincides with G(t,r,a,b,R)
up to a shift along the imaginary axis.
Proof. Let (E01 , E11 ) be the image of the condenser (E0 , E1 ) under the mapping
w = dt (z). The plate E01 contains the set K0 = dt ({z : |z| < r}) and connects
dt (a) with the point symmetric to it relative to the circle |w| = 1, and the plate
E11 contains K1 = dt ({z : |z| > R}) and connects dt (b) to the point symmetric
to it relative to the same circle. Let θ and ϕ be the angles at which the arcs
K 0 ∩ {w : |w| = 1} and K 1 ∩ {w : |w| = 1} are seen from w = 0, respectively.
Now we apply circular symmetrization with respect to the negative real half-axis
to the condenser (E01 , E11 ). Then we obtain a condenser (E02 , E12 ) such that

cap(E0 , E1 ) = cap(E01 , E11 )  cap(E02 , E12 ).

The plate E02 of (E02 , E12 ) connects the point −|dt (a)| with the point symmetric to
it relative to |w| = 1, and E12 connects |dt (b)| with the analogous point. Evaluated
at w = 0, the harmonic measure of the arc K 0 ∩ {w : |w| = 1} in the disc |w| < 1
is equal to θ /2π; on the other hand it coincides with the harmonic measure,
150 Chapter 5. Metric Properties of Sets and Condensers

evaluated at the point z = t, of the interval [−ir, ir] in the half-plane Re z > 0
(see, for instance, [C]). The latter is equal to θ/π, so θ /2 = θ. In a similar way
ϕ /2 = π −ϕ. After the circular symmetrization the centres of the boundary circles
of K0 and K1 occur on the real axis, and the images of the arcs K 0 ∩ {w : |w| = 1}
and K 1 ∩ {w : |w| = 1} are seen from w = 0 at the same angles θ and ϕ as
before. Returning to the variable z = d−1|t| (w) we see that the inverse image of the
2 2
condenser (E0 , E1 ) contains a condenser with the field G(t, r, a, b, R). Since the
capacity is monotonic, we have (5.32).
Now assume that equality sign holds in (5.32). Then it follows from Theorem
4.2 that the points a, −a, b, −b, and t lie on the same straight line, and if r > 0
or R < ∞, then they lie on the real axis. Furthermore, G must coincide with the
ring domain G(t, r, a, b, R) up to a translation along the imaginary axis. 
It is easy to see that Lemma 5.5 contains Grötzsch’s and Teichmüller’s lem-
mas, which correspond to some special choice of the parameters. Now we turn to
another kind of generalization.
For any R, r, and b such that Re b > 0 and 0 < r < |b|  R  ∞ we look at
the doubly connected domain
H1 (r, b, R) = {z : r < |z| < R, z ∈ [−R, −b] ∪ [b , R]},
where b = d−1 r (|dr (b)|). In a similar way, for any r, R and a such that 0  r 
|a| < R < ∞, Re a > 0, we construct the ring domain H2 (r, a, R) = {z : r < |z| <
R, z ∈ [−a , a ]}, where a = d−1
r (−|dr (a)|).

Lemma 5.6. Assume that 0 < r < |b|  R  ∞ and Re b > 0. If the complement
of the plate E0 of a condenser (E0 , E1 ) contains no pair of points z1 , z2 , such that
z1 z2 = r2 and the plate E1 contains the set |z| > R and joins b to −b, then
|(E0 , E1 )|  M H1 (r, b, R). (5.33)
In addition, if the field G of (E0 , E1 ) is a ring domain, then equality holds in
(5.33) only when G coincides with H1 (r, b, R).
Proof. Let (E01 , E11 ) be the image of the condenser (E0 , E1 ) under the mapping
w = dr (z). The set Cw \E01 contains no pair of points w1 , w2 , such that w1 = −w2 :
indeed, if dr (z1 ) = −dr (z2 ) for some z1 , z2 ∈ Cz \ E0 , then we readily see that
z1 z2 = r2 , while Cz \ E0 contains no such points by assumption. Performing
the circular symmetrization Cr− with respect to the negative half-axis we find
that the plate E02 of the condenser (E02 , E12 ) := Cr− (E01 , E11 ) contains the half-
plane Re w < 0, while E12 connects |dr (b)| with −1/|dr (b)| and contains the set
dr ({z : |z| > R}). Returning to the variable z = d−1 r (w) we conclude that the
inverse image of (E02 , E12 ) contains a condenser with the field H1 (r, b, R). As a
result,
cap(E0 , E1 ) = cap(E01 , E11 )  cap(E02 , E12 )  (M H1 (r, b, R))−1 .
The equality part is verified in a similar way to Lemma 5.5. 
5.5. Ring domains 151

Combining our generalizations of Rengel’s, Grötzsch’s and Teichmüller’s lem-


mas we present the following result for ring domains, which goes back to He’s
principle of partial symmetrization [D4], p. 42.
Theorem 5.17. Let G be a doubly connected domain in C, and let E0 and E1 be
the connected components of its complement. Assume that for each k = 1, . . . , n
(n  2) the domain G satisfies at least one of the following conditions:
(1) There exists a connected component of the intersection G∩D k which separates
the points 0 and ∞ in Dk . Assume that the area of this component in the
logarithmic metric does not exceed some quantity Sk , 0 < Sk < ∞.
(2) For some rk , ak , bk , and Rk , where 0  rk < Rk  ∞, ak , bk ∈ Dk ∩ {z :
rk  |z|  Rk }, one component of the complement of G (for instance, E0 )
contains the sector {z : |z| < rk } ∩ Dk and connects ak with the boundary of
Dk , while the other component (E1 ) contains the sector {z : |z| > Rk } ∩ Dk
and connects the point bk with the boundary of Dk .
(3) For some rk , bk , and Rk , where 0 < rk < Rk  ∞ and bk ∈ Dk ∩ {z :
rk < |z|  Rk }, one component of the complement of G (for instance, E1 )
contains the sector {z : |z| > Rk } ∩ Dk and connects the point bk with the
boundary of Dk . The set G ∪ E1 contains no pair of points z1 , z2 such that
z1 z2 = rk2 ei4kπ/n .
(4) For some rk , ak , and Rk , where 0  rk < Rk < ∞ and ak ∈ Dk ∩ {z :
rk  |z| < Rk }, one component of the complement of G, for instance, E0 ,
contains the sector {z : |z| < rk } ∩ Dk and connects the point ak with the
boundary of Dk . The set G ∪ E0 contains no pair of points z1 , z2 such that
z1 z2 = Rk2 ei4kπ/n .
Then  −1
1
n
MG  (M Gk )−1 , (5.34)
n
k=1

where the ring domains Gk , k = 1, . . . , n, are defined as follows: if the domain G


satisfies condition (1) for some k, then Gk = {z : rk < |z| < Rk } with arbitrary
rk and Rk such that Sk = (1/2πn) log(Rk /rk ); if G satisfies condition (2), then
√ n/2   n/2
Gk = {z : z n ∈ G(t, rk , ank , bnk , Rk )}, where t is an arbitrary complex

number with Re t > 0 and 1 = 1; if G satisfies condition (3), then Gk has
√ n/2  n/2
the form {z : z n ∈ H1 (rk , bnk , Rk )}, and in the case of G satisfying (4),
√ n/2  n/2
Gk = {z : z n ∈ H2 (rk , ank , Rk )}.
Equality holds in (5.34) only when G is an n-fold rotationally symmetric
ring domain equal to some Gk , 1  k  n, where in the case of condition (2) the
n/2 n/2
parameter t is real and satisfies the inequality |ak |  t  |bk |.
Proof. Let C(k) = (E0 (k), E1 (k)) be the result of the separating transformation
of the condenser C = (E0 , E1 ) with field G with respect to the family of functions
152 Chapter 5. Metric Properties of Sets and Condensers


ζ = pk (z) ≡ z n (z ∈ Dk , Re ζ > 0). By Theorem 4.8

1
n
cap C  cap C(k).
2
k=1

Each of the condensers C(k), k = 1, . . . , n, satisfies at least one of the following


conditions:
(1 ) A connected component of the field of the condenser C(k) separates the
points 0 and ∞ in Cζ . The area of this component
√ in the logarithmic metric
has the bound n2 Sk /2 (the mapping ζ = z n has increased the logarith-
mic area n2 /4-fold, after which the area was doubled because the separating
transformation has added the ‘half’ of the condenser lying in the left half-
plane Re ζ < 0).
(2 ) One plate of C(k) contains the disc {ζ : |ζ| < rk } and connects the point
n/2
 n  n/2
a with − ank and the other contains the set {ζ : |ζ| > Rk } and connects
 nk  n
bk with − bk .
(3 ) One plate of the condenser C(k) contains the set {ζ : |ζ| > Rk } and
n/2
 
connects the point bnk with − bnk . The complement of the other plate
contains no pairs of points ζ1 , ζ2 such that ζ1 ζ2 = rkn .
(4 ) One plate of the condenser C(k) contains the disc {ζ : |ζ| < rk } and
n/2
 
connects the point ank with − ank . The complement of the other plate
contains no pairs of points ζ1 , ζ2 such that ζ1 ζ2 = Rkn .
If any of conditions (1 ), (2 ), (3 ) hold, then we use Lemma 5.4, 5.5, or 5.6. Case
(4 ) reduces to (3 ) by means of the map η = 1/ζ. Then we obtain the inequalities
cap C(k)  (M G(k))−1 , k = 1, . . . , n,
where the G(k) are the following ring domains: in the first case G(k) has the
form {ζ : r(k) < |ζ| < R(k)} with arbitrary r(k) and R(k) such that n2 Sk /2 =
n/2   n/2
(1/2π) log(R(k)/r(k)); in the second case G(k) = G(t, rk , ank , bnk , Rk ), in
n/2  n n/2
the third case G(k) = H1 (rk , bk , Rk ), and finally, in the fourth case G(k) =
n/2  n/2
H2 (rk , ank , Rk ). The ring domains G(k), 1  k  n, are mirror-symmetric

relative to the coordinate axes. Applying the (multivalued) mapping z = n ζ 2 to
G(k) we obtain an n-fold rotationally symmetric ring domain Gk such that each of
the n symmetrically positioned parts forming Gk corresponds univalently to half
the ring G(k). Finally, we obtain
2
(M G(k))−1 = (M Gk )−1 .
n
The proof of (5.34) is complete. By Theorem 4.9, if equality holds in (5.34), then
G must be mirror-symmetric relative to the straight lines containing the rays
5.5. Ring domains 153

{z : arg z = π(2k − 1)/n}, k = 1, . . . , n. Hence the fields of the condensers C(k)


are also ring domains and it remains to cite the uniqueness in Lemmas 5.1, 5.5,
and 5.6. 

Vjg‚5.hqnf‚u{oogvtke‚Itqv|uej‚tkpi Vjg‚5.hqnf‚u{oogvtke‚Vgkejow
nngt‚tkpi

Figure 5.6.

In some special cases the modulus of G(t, r, a, b, R) can easily be calculated


with the help of elliptic integrals. The following result is a consequence of case (2)
in Theorem 5.17 for rk = 0 and Rk = ∞, k = 1, . . . , n.
Corollary 5.3. If G is a doubly connected domain and for each k = 1, . . . , n (n 
2) some component of the complement of G connects a point ak ∈ Dk with the
boundary of Dk , and the other component of the complement connects a point
bk ∈ Dk with the boundary of Dk , then
⎧ ⎫−1
1 ⎨ ⎬
n
K(qk )
MG   , (5.35)
2⎩ K 1 − q2 ⎭
k=1 k

 n  n
where qk = d−1 −1
|t| (−|dt ak |)/d|t| (|dt b |), t is an arbitrary complex number such
 n  n k
that Re t > 0, Re ak > 0, Re bk > 0, and K(·) is the complete elliptic integral
n/2
of the first kind. In particular, if ak = rk ei2kπ/n , bk = Rk ei2kπ/n , and rk  t 
n/2 
Rk , then qk = rkn /Rkn . Equality holds in (5.35) only in the case of an n-fold
rotationally symmetric Teichmüller ring

C \ ({z : 0  z n  rn } ∪ {z : z n  Rn })

with 0 < r < R < ∞, ak = rei2kπ/n , bk = Rei2kπ/n , k = 1, . . . , n, and


rn/2  t  Rn/2 .
154 Chapter 5. Metric Properties of Sets and Condensers

Exercises 5.5
(1) Show that
1 
M G(P ) = K 1 − P −2 /K(P −1 ),
4
where K(·) is the complete elliptic integral of the first kind.
(2) Prove the equality
M T(P 2 − 1) = 2M G(P ).

(3) Prove Carleman’s lemma: an annulus bounded by concentric circle has the
maximum modulus among the ring domains in the plane whose boundary
components enclose regions with prescribed area;
(4) Using Carleman’s lemma prove Grötzsch’s assertion cited in Exercise 1.3(6).
(5) Prove Lemma 5.4.
(6) Let G be a ring domain in C and assume that points rk exp(2πik/n) and
points Rk exp(2πik/n), where 0 < rk < Rk < ∞, k = 1, . . . , n (n  2), lie in
different components of the complement of G. Show that
>
? n
1  ? rk
MG  K 2
1 − τ /K(τ ), where τ =@ ,
2n Rk
k=1

and prove that here equality holds only if G is an n-fold rotationally sym-
metric Teichmüller ring.
(7) Show that f (θ) = M [C \ ([0, 1] ∪ {z : arg z = θ, |z|  1})] is an increasing
function on (0, π).
(8) For any positive s and t find the modulus of the ring domain


n
C\ [{z : arg z n = 0, |z|  s} ∪ {z : arg z n = π, |z|  t}]
k=1

(9) Deduce Grötzsch’s lemma from Teichmüller’s lemma.


(10) Establish Teichmüller’s lemma using Grötzsch’s lemma.
(11) Assume that a component of the complement of a ring domain G contains a
ray with initial point at the origin and that the smallest closed sector with
vertex at the origin that contains the other component of the complement
of G has opening angle α, 0 < α < 2π. Find a sharp upper bound for the
modulus M G and describe the rings delivering this bound.
Chapter 6

Extremal Decomposition Problems

By extremal decomposition problems we mean problems of finding upper bounds


for sums of the form α1 M1 + α2 M2 + · · · + αn Mn , where the αk are fixed positive
numbers and the Mk are the moduli or reduced moduli of nonoverlapping domains
Bk satisfying some conditions, k = 1, . . . , n. An extremal system of such domains
forms an ‘extremal decomposition’. The simplest case is when we consider two
nonoverlapping domains, B1 and B2 . Let 0 ∈ B1 and ∞ ∈ B2 . By Grötzsch’s
lemma

M (B1 , ∂B1 , {0}, {1}, {r}) + M (B2 , ∂B2 , {∞}, {1}, {r})  0, (6.1)

with equality sign holding only for B1 = {z : |z| < R} and B2 = {z : |z| > R},
where R is an arbitrary positive number. The following equivalent form of (6.1)
is more common in the literature:

r(B1 , a1 )r(B2 , a2 )  |a1 − a2 |2 , B1 ∩ B2 = ∅, aj ∈ Bj , j = 1, 2. (6.1 )

This result, which was established by Lavrent’ev, triggered the investigations of a


variety of extremal decomposition problems, which now have a long history [Kuz1],
[Kuz2]. We limit ourselves to just a few problems for the reduced moduli of open
sets Bk with respect to (interior or boundary) points ak . We state the results in
terms of inner or Robin radii, or as inequalities for quadratic forms with coeffi-
cients depending on the Green and Robin functions gBk (z, ak ) and gBk (z, ak , Γk ).
Provisionally, we divide the material into three parts: when the points ak are
fixed, when they have some freedom, and when the problem reduces to estimates
for Möbius invariants. In the first two cases it is natural to say that we have ex-
tremal problems with fixed and free poles, respectively, since the ak are poles of
the associated quadratic differential or poles of the Green (Robin) functions of the
sets under consideration.

© Springer Basel 2014 155


V.N. Dubinin, Condenser Capacities and Symmetrization in Geometric Function
Theory, DOI 10.1007/978-3-0348-0843-9_6
156 Chapter 6. Extremal Decomposition Problems

6.1 Fixed poles


First we discuss consequences of the monotonicity of the capacity and of compo-
sition principles for condensers.
Theorem 6.1 ([Neh]). Let Bk ⊂ C be pairwise nonoverlapping domains and let
ak ∈ Bk , k = 1, . . . , n (n  2). Then for any real numbers δk , k = 1, . . . , n, such

n
that δk = 0,
k=1
n n n
2 
(r(Bk , ak ))δk
 |ak − al |−δk δl .
k=1 k=1 l=1
l=k


n
Proof. We set B = Bk , Z = {ak }nk=1 , = {0, δ1 , . . . , δn }, = {δ1 , . . . , δn },
k=1
and Ψ = {r, . . . , r}. In this case inequality (2.18) takes the following form:

M (B, ∂B, Z, , Ψ)  M (Z, , Ψ). (6.2)

It remains to use (2.16) and (2.17). 


Theorem 6.2 ([Kuf]). If B1 and B2 are nonoverlapping subdomains of the disc
|z| < 1, then the following sharp estimate holds for any points ak ∈ Bk , k = 1, 2 :

r(B1 , a1 )r(B2 , a2 )  |a2 − a1 |2 [1 − |(a2 − a1 )/(1 − a1 a2 )|2 ]. (6.3)

Here equality sign holds if and only if B1 ∪ B2 = {z : |z|  1} and the common
boundary of B1 and B2 has the equation

|(z − a1 )/(1 − a1 z)| = |(z − a2 )/(1 − a2 z)|, |z|  1.

Proof. We can assume that ak = 0, k = 1, 2, and add the domains B3 = {z :


1/z ∈ B1 } and B4 = {z : 1/z ∈ B2 } symmetric to B1 and B2 , respectively. Let

4
B= Bk , Z = {a1 , a2 , 1/a1 , 1/a2 }, = {0, 1, −1, −1, 1}, = {1, −1, −1, 1},
k=1
and Ψ = {r, r, r, r}. Again, (2.18) yields inequality (6.2) for the new choice of the
parameters. Using formulae (2.16) and (2.17), we arrive at the required inequality
after simple calculations. The easiest way to see that (6.3) is sharp is to cite
Theorem 2.9. 
In solving extremal problems for decompositions into families of subsets with
‘spacers’ we can sometimes use Theorem 2.7. For example, it has the following
immediate consequence.
Theorem 6.3. Let ak , k = 1, . . . , n, be different fixed points in the unit disc U

n
and let δk , k = 1, . . . , n, be fixed positive numbers such that δk = 1. Then for
k=1
6.1. Fixed poles 157

any pairwise nonoverlapping domains Bk in U , k = 1, . . . , n, such that ak ∈ Bk


n
and for any condenser C such that one of its plates contains Bk and the other
k=1
contains the exterior of U
 δk2  
n
r(Bk , ak )
n n
 1 − ak al δk δl
e2π|C|    . (6.4)
r(U, ak )  ak − al 
k=1 k=1 l=1
l=k

In addition, if the domains Bk are admissible and C is an admissible regular



n
condenser, then equality holds in (6.4) if and only if the pair { Bk , C} is a
k=1
decomposition of the disc U with respect to the points ak and numbers δk .
Note that if equality holds in (6.4) and ak → al , then necessarily |C| → +∞.
Hence for small |C| Theorem 6.3 does not give us a sharp estimate. In this case
the extremal configuration is not the union of n disjoint domains Bk and the field
of C any longer. The direct calculations of the modulus of the condenser C for
which the extremal configuration breaks down into disjoint domains take us to the
following result.
Corollary 6.1. Let B1 and B2 , B1 ∩ B2 = ∅, be fixed subdomains of U and let
ak ∈ Bk , k = 1, 2, be fixed points. If there exists a condenser with one plate
containing B1 ∪ B2 and the other (‘spacer’) containing C\U such that the modulus
of this condenser is at least (1/2π) log(1/ρ), where ρ + 1/ρ = 2|1 − a1 a2 |/|a1 − a2 |,
0 < ρ < 1, then the following sharp bound holds:
1
r(B1 , a1 )r(B2 , a2 )  (1 − ρ4 )2 |a1 − a2 |2 .
16
For admissible domains B1 and B2 equality holds here if and only if the are bounded
by the curve
|(z − a1 )(z − a2 )| = ρ2 |(1 − a1 z)(1 − a2 z)|.
Recall that if B is an admissible domain with Green’s function gB (z, ζ), then
log r(B, a) = hB (a, a), where hB (z, ζ) = gB (z, ζ) + log |z − ζ| is the regular part
of the Green function. Thus, for instance, the maximum problem for the product
1
n
(r(Bk , ak ))αk is equivalent to the maximum problem for the sum
k=1


n
αk hBk (ak , ak ).
k=1

We would like to have at our disposal a quantity taking account of the regular terms
in the expansion of the Green function, so we look at the symmetric difference

H(B, z, ζ) = hB (z, z) + hB (ζ, ζ) − 2hB (z, ζ).


158 Chapter 6. Extremal Decomposition Problems

We can show that for z0 ∈ B and any real ϕ

lim H(B, z0 − ρeiϕ , z0 + ρeiϕ )/ρ2 = −4πK(B, z0 , ϕ),


ρ→0

where we set
K(B, z0 , ϕ) = K(z0 , z 0 ) − Re [e2ϕi l(z0 , z0 )]
for the kernels K(z, ζ) and l(z, ζ) defined in Exercise 2.4(3).
The quantity K(B, z0 , ϕ) is nonnegative and nonincreasing in B. Thus we
obtain the natural problem to find the minimum of the sum


n
αk K(Bk , ak , ϕk )
k=1

over arbitrary pairwise nonoverlapping domains Bk , when certain constraints are


imposed on the parameters αk , ak , and ϕk , k = 1, . . . , n.
Theorem 6.4 ([A]). For any pairwise nonoverlapping domains Bk , points ak , ak ∈
Bk , positive numbers αk , and real numbers ϕk , where k = 1, . . . , n, n  2,

n √
n
αk αl ei(ϕk +ϕl )
n
1
αk K(Bk , ak , ϕk )  Re . (6.5)
π (ak − al )2
k=1 k=1 l=1
l=k

Proof. We can assume that each Bk has a classical Green function. For suffi-
ciently small r > 0 and ρ > 0 we take the tuples Z = {a1 − ρeiϕ1 , . . . , an −
√ √ √ √
ρeiϕn , a1 + ρeiϕ1 , . . . , an + ρeiϕn }, = {− α1 , . . . , − αn , α1 , . . . , αn }, and
Ψ = {r, . . . , r}. By Theorem 2.8


n
√ √
α2k M (Bk , ∂Bk , {ak − ρeiϕk , ak + ρeiϕk }, {0, − αk , αk }, {r, r})
k=1
 2

n
 αk M (Z, , Ψ).
k=1

Hence

n
αk [2 log(2ρ) + H(Bk , ak − ρeiϕk , ak + ρeiϕk )]
k=1
  n √

n
n
αk αl ei(ϕk +ϕl )
 αk 2 log(2ρ) − 4Re ρ2 + o(ρ2 ), ρ → 0,
(ak − al )2
k=1 k=1 l=1
l=k

which yields the required inequality. 


6.1. Fixed poles 159

Corollary 6.2. Under the assumptions of Theorem 6.4, for any complex numbers
βk , k = 1, . . . , n,
 
 
 n
1 βk βl 
n n n
 2
 βk l(ak , ak ) +   βk β k K(ak , ak ),
 π (ak − al )2 
k=1 k=1 l=1  k=1
l=k

where K(z, ak ) and l(z, ak ) are the Bergman kernels of the domain Bk of the first
and second kind, respectively, with respect to the class of regular functions in Bk
with integrable square of the absolute value, k = 1, . . . , n.
Proof. This follows from Theorem 6.4, in which we set αk = |βk |2 and ϕk =
arg βk + θ, k = 1, . . . , n, and use that θ can be arbitrary. 
Below we present general results indicating that extremal decomposition
problems for reduced moduli with respect to interior points are connected with
problems of decomposition into strips and half-strips. The reduced moduli of strips
and half-strips (M (P )) were defined in Exercises 2.4(10) and 2.4(11). By definition
a trajectory of a strip P is the inverse image of an arbitrary line Re w = u, 0 < u <
1, with respect to the mapping f (see Exercise 2.4(10)). A trajectory of a half-strip
P is the inverse image of an arbitrary half-line Re w = u, Im w  0, 0 < u < 1, with
respect to the corresponding mapping f . In particular, a trajectory of a half-strip
contains a boundary element on the ‘side’ opposite to the vertex of the half-strip.
For a strip (k = 1, 2) or a half-strip (k = 1) the quantities ϕk in the expansion
(2.24) are equal to the exponents of the admissible points zk in P .
Now we take several distinct points {ak }nk=1 on the sphere C and a finite
set of pairwise nonoverlapping strips and half-strips {P } such that the supports
of their vertices are points in the set {ak }nk=1 and each ak is the support of some
vertex. Let akl , l = 1, . . . , mk , be the vertices of domains in {P } which have the
support ak , k = 1, . . . , n. Let ϕkl be the exponent of the admissible point akl and
let Pkl be the domain with vertex akl , l = 1, . . . , mk , k = 1, . . . , n. Clearly, each
strip is numbered twice in this way and each half-strip is numbered only once. We
have the following result.
Theorem 6.5. Let {ak }nk=1 and {P } be the systems of points and domains defined
above, let B be an open subset of C containing the points ak , k = 1, . . . , n, and
let {xk }nk=1 be a set of real numbers. Assume that the following conditions are
fulfilled :
k
m
1) ϕkl = 2, k = 1, . . . , n;
l=1
2) for each strip Pkl in {P }, |xk ϕkl | = |xk ϕk l |, where ϕkl and ϕk l are the
exponents of the vertices of Pkl ;
3) if Pkl is a half-strip or a strip with vertices akl and ak l , and if xk xk > 0,
where 1  l, l  mk and 1  k, k   n, then no trajectory of the domain Pkl
lies in B.
160 Chapter 6. Extremal Decomposition Problems

Then
 

n
n
n
n
π
mk
x2k log r(B, ak )+ xk xl gB (ak , al )  x2k 2
λkl ϕkl M (Pkl ) , (6.6)
2
k=1 k=1 l=1 k=1 l=1
l=k

where λkl = 1/2 if Pkl is a strip and λkl = 1 if Pkl is a half-strip.


Proof. We can assume that each connected component of B is bounded by finitely
many smooth curves and contains al least one of the points ak , 1  k  n.
Moreover, it is sufficient to prove Theorem 6.5 in the case when the domains in {P }
have piecewise smooth boundaries and the numbers xk , k = 1, . . . , n, are distinct
from zero. For l = 1, . . . , mk and k = 1, . . . , n let w = fkl (z) be the function in the
definition of the domain Pkl ∈ {P } such that fkl (akl ) = −i∞, which is defined
on the boundary of Pkl by means of the boundary correspondence, We introduce
some further notation:
Ek = U (ak , r1/|xk | ), k = 1, . . . , n;
C(r) = (B, {∂B, E1 , . . . , En }, {0, sign x1 , . . . , sign xn });
ω is the potential function of the condenser C(r);
γ = {z ∈ B : ω(z) = 0};
Ckl (r) = (Pkl , {γ ∩ P kl , Ek ∩ P kl , Ek ∩ P kl }, {0, sign xk , sign xk }), provided
that Pkl is a strip with vertices akl and ak l ;
Ckl (r) = (Pkl , {γ ∩ P kl , Ek ∩ P kl }, {0, sign xk }), provided that Pkl is a half-
strip with vertex akl .
Then it follows from the composition principle (Theorem 1.17) that

cap C(r) = cap(C, {γ, E1 , . . . , En }, {0, sign x1 , . . . , sign xn })


n mk
 λkl cap Ckl (r).
k=1 l=1

Let
kl (r) = (Pkl , {γ ∩ P kl , (Ek ∪ Ek ) ∩ P kl , {0, 1}) if Pkl is a strip
C
and
kl (r) = (Pkl , {γ ∩ P kl , Ek ∩ P kl }, {0, 1}) if Pkl is a half-strip.
C

We can easily see that


kl (r) = cap fkl (C
cap Ckl (r) = cap C kl (r)), k = 1, . . . , n, l = 1, . . . , mk .

Finally, if Pkl is a strip, then let SCkl (r) be the result of the Steiner symmetrization
of the condenser fkl (C kl (r)) with respect to the real axis, while if Pkl is a half-strip,
6.1. Fixed poles 161

then let SCkl (r) be obtained by the Steiner symmetrization from the condenser
in the strip 0  Re w  1 which is mirror-symmetric relative to the real axis
kl (r)) in the half-strip {w : 0  Re w  1, Imw  0. By
and coincides with fkl (C
Theorem 4.1, for strips we have
kl (r))  cap SCkl (r),
cap fkl (C
while for half-strips, in view of the symmetry principle,

kl (r))  1 cap SCkl (r).


cap fkl (C
2
Finally,
1
nk m
cap C(r)  cap SCkl (r).
2
k=1 l=1
If Pkl is a strip with vertices akl and ak l and if xk xk < 0, then each trajectory of
this strip meets γ because ω is continuous. Otherwise each trajectory of Pkl meets
γ by assumption 3) of Theorem 6.5. Hence the plate of the condenser SCkl (r) on
which the potential function is equal to 0 contains the interval [0, 1]. The second
plate of this condenser is the union of two ‘almost straight’ half-strips, which ‘go’
to infinity as r → 0. Citing the expansion (2.24) and the second assumption of
Theorem 6.5 we obtain
2
cap SCkl (r) 
1
(− log r) + λkl M (Pkl ) + o(1)
|xk ϕkl |π
  $
2|xk ϕkl |π λkl M (Pkl )|xk ϕkl |π 1
= 1− +o , r → 0,
− log r − log r log r
l = 1, . . . , mk , k = 1, . . . , n. Summing the above relations and taking condition 1)
into account we obtain
n  
−1
cap C(r)  2π |xk |
log r
k=1
 2  2 
n mk
1 1
2 2 2
− λkl M (Pkl )xk ϕkl π +o , r → 0.
log r log r
k=1 l=1

Comparing this with the asymptotic behaviour of the capacity of the condenser
C(r) expressed by (2.10) we obtain the required estimate. 
Theorem 6.6. Under the assumptions of Theorem 6.5 let B be a union of finitely
many pairwise nonoverlapping finitely connected domains without isolated bound-
ary points, and let Γ be a nonempty closed subset of ∂B which consists of finitely
many nondegenerate connected components, and assume that
 k
n m
Pkl ⊂ C \ ((∂B) \ Γ).
k=1 l=1
162 Chapter 6. Extremal Decomposition Problems

Then

n
n
n
x2k log r(Bk , Γk , ak ) + xk xl gBk (al , ak , Γk )
k=1 k=1 l=1
l=k
 

n
π
mk
 x2k λkl ϕ2kl M (Pkl ) ,
2
k=1 l=1

where Bk is the connected component of B containing ak , Γk = Γ ∩ (∂Bk ), and


the Robin function gBk (z, ak , Γk ) is set equal to zero outside Bk , k = 1, . . . , n.
Proof. This is similar to the proof of Theorem 6.5 apart from some details. For
example, the set ∂B in the definition of C(r) must be replaced with Γ. The
composition principle must now be applied to the condenser (C \ ((∂B) \ Γ),
{γ, E1 , . . . , En }, {0, sign x1 , . . . , sign xn }), whose capacity is equal to that of C(r).
On the other hand, this capacity is equal to the sum of the capacities of certain
condensers C(r; Bk , Γk , Zk , k , Ψk ), where Zk = {al : al ∈ Bk } and k and Ψk
are the tuples of numbers in the set {0, sign x1 , . . . , sign xn } and of functions in
{r1/|x1 | , . . . , r1/|xn | } which correspond to the points al in the domain Bk , k =
1, . . . , n. The asymptotic formula for such capacities is given by Theorem 2.2. 
Now we extend Theorem 6.6 to the case of boundary points ak . We shall
content ourselves with one domain to avoid excessive complexity. Let B be a
finitely connected domain without isolated boundary points in the Riemann sphere
C. We consider a system {ak }nk=1 of distinct admissible points in B and a finite
system {P } of pairwise nonoverlapping strips and half-strips in C which have the
following property: for any sufficiently small r > 0 the neighbourhood U (z, r, P )
of each vertex z of any domain P in {P } lies in the neighbourhood U (ak , r, B)
of some ak , 1  k  n. The other way round, each neighbourhood U (ak , r, B)
contains the neighbourhood U (z, r, P ) of a vertex z of some domain P ∈ {P }. It
particular, the vertices z and the corresponding attainable points ak have the same
supports. (For the definition of the neighbourhoods in question see § 2.1.). Let akl ,
l = 1, . . . , mk , denote the vertices of domains P ∈ {P } such that U (ak , r, B) with
small r contains neighbourhoods of these points, and let ϕkl be the exponent of
the vertex akl , l = 1 . . . , mk , k = 1, . . . , n.
Theorem 6.7. Let B and the systems {ak }nk=1 and {P } be as defined above, let Γ be
a nonempty closed subset of the boundary ∂B formed by finitely many nondegen-
erate connected components, and let {xk }nk=1 be a system of real numbers. Assume
that ak ∈/ Γ, k = 1, . . . , n, the domains P in {P } are disjoint from (∂B) \ Γ, and
the following conditions hold :
k
m
1) ϕkl = σ(ak ), k = 1, . . . , n;
l=1
2) if ϕkl and ϕk l are the exponents of the vertices of a strip Pkl in {P }, then
|xk ϕkl | = |xk ϕk l |;
6.1. Fixed poles 163

3) if Pkl is a half-strip or if Pkl is a strip with vertices akl and ak l , 1  l, l 


mk , 1  k, k   n, and if xk xk > 0, then no trajectory of Pkl lies in B.
Then

n
n
n
x2k σ(ak ) log r(B, Γ, ak ) + xk xl σ(ak )gB (ak , al , Γ)
k=1 k=1 l=1
l=k
 

n
mk
 x2k π λkl ϕ2kl M (Pkl ) ,
k=1 l=1

where the λkl are as in Theorem 6.5.


Proof. This repeats the proof of Theorem 6.5, provided that we take account of
our comments on Theorem 6.6. Yet another difference is that the Ek must be
replaced with the sets U (ak , r1/|xk | , B), k = 1, . . . , n. 
Theorems 6.5–6.7 can be combined into a single statement concerning es-
timates for quadratic forms. However, the detailization of particular conditions
would make such a statement cumbersome and therefore inconvenient in applica-
tions. In addition, note that our approach also gives estimates for quadratic forms
with coefficients depending on the Neumann functions.
In Theorems 6.5–6.7 the classical nonoverlapping assumption for domains
in extremal decomposition problems is either replaced with the ‘noncovering’ as-
sumption for trajectories or dropped altogether. For example, if we take two finite
points a1 and a2 (n = 2), a single strip P = C \ [a1 , a2 ] with vertices at a1 and a2 ,
and the numbers x1 = x2 = 1 in Theorem 6.5, then inequality (6.6) shows that
Lavrent’ev’s result (6.1 ) also holds in the case when B1 ∩B2 = ∅, but B = B1 ∪B2
contains no circle arc with end-points a1 and a2 . Moreover, under these constraints

r(B, a1 )r(B, a2 ) exp(2gB (a1 , a2 ))  |a1 − a2 |2 .

Now let {γ} be the set of circles passing through a1 and a2 . Let B1∗ be the set of
points in B1 (recall that a1 ∈ B1 ) that can be connected with a1 by a circle arc
γ ∈ {γ} lying entirely in B1 and not containing a2 . In a similar way, let B2∗ be the
set of points in B2 (recall that a2 ∈ B2 ) that can be connected with a2 by a circle
arc γ ∈ {γ} lying in B2 and not containing a1 . Using Theorem 6.6 for a suitable
decomposition of the strip P = C \ [a1 , a2 ] into half-strips we arrive at inequality
(6.1 ) again, albeit under the weaker constraint

B1∗ ∩ B2∗ = ∅.

Our next result uses some notions from Appendix A5.


Corollary 6.3. Let G be a domain in C bounded by finitely many simple closed
analytic curves (maybe none at all). Let Q(z)dz 2 be a positive quadratic differen-
tial on G which is regular everywhere apart from m  0 simple poles (the set of
164 Chapter 6. Extremal Decomposition Problems

simple poles can be empty) and n  1 second-order poles a1 , . . . , an such that in a


neighbourhood of the latter, in terms of a local parameter taking ak to z = 0,
 2 
xk
Q(z)dz = − 2 + · · · dz 2 ,
2
z
where xk = 0 is a real number, k = 1, . . . , n. Assume that G is the inner closure of
the union of the circle domains G1 , . . . , Gn corresponding to the poles a1 , . . . , an .
Let B ⊂ G be an open set which contains the points ak , k = 1, . . . , n, but no
orthogonal trajectory of the quadratic differential Q(z)dz 2 with limit end-points ak
and ak such that xk , xk > 0, 1  k, k   n. Then

n
n
n
n
x2k log r(B, ak ) + xk xl gB (ak , al )  x2k log r(Gk , ak ). (6.7)
k=1 k=1 l=1 k=1
l=k


n
Equality holds in (6.7) for B = Gk and arbitrary xk , k = 1, . . . , n.
k=1

Proof. Let Φ be the union of the trajectories of the quadratic differential Q(z)dz 2
that have a limit end-point at a zero or a simple pole of Q(z)dz 2 . Let z0 ∈ Φ and
z
ζ = F (z) = Q1/2 (z)dz.
z0

Then the analytic function F is locally univalent in the domain G apart from the
zeros and poles of Q(z)dz 2 . It maps it onto a Riemann surface R over the ζ-plane.
In Gk \ ak the function F has the following form:
F(z) = ±i|xk | log fk (z), (6.8)
where fk maps Gk conformally and univalently onto the unit disc |w| < 1 and
fk (ak ) = 0, k = 1, . . . , n. Let E be the union of all the orthogonal trajectories of
the quadratic differential Q(z)dz 2 connecting double poles ak with zeros or poles
on the boundary of the corresponding domain Gk , k = 1, . . . , n. Let E denote the
F -image of the set E. From the local structure of trajectories of the quadratic
differential we see that each connected component of R \ E covers a strip of the
form a < Re ζ < b or a half-strip of the form a < Re ζ < b, Im ζ > 0 (Im ζ < 0)
with multiplicity 1. Thus, G \ E is a union of several strips and half-strips Pkl ,
l = 1, . . . , mk , k = 1, . . . , n, which make up a set {P } (see the construction
before Theorem 6.5). Assumptions 1)-3) of Theorem 6.5 hold automatically. To
calculate the right-hand side of (6.6) in this case, note that if Pkl is a strip with
vertices akl and ak l and ϕkl and ϕk l are the corresponding exponents, then the
representation (6.8) shows that
1 1
M (Pkl ) = − log |fk (ak )| − log |fk  (ak )|.
ϕkl π ϕk l π
6.1. Fixed poles 165

If Pkl is a half-strip with vertex akl , then


1
M (Pkl ) = − log |fk (ak )|.
ϕkl π

Hence  

n
π
mk
1
n mk
x2k 2
λkl ϕkl M (Pkl ) = Ψkl ,
2 2
k=1 l=1 k=1 l=1

where
 1
' 2 (
− 2 xk ϕkl log |fk (ak )| + x2k ϕk l log |fk  (ak )| if Pkl is a strip,
Ψkl =
− x2k ϕkl log |fk (ak )| if Pkl is a half-strip.

The last sum has the form

1
nk m n n
− x2k ϕkl log |fk (ak )| = − x2k log |fk (ak )| = x2k log r(Gk , ak ). 
2
k=1 l=1 k=1 k=1

By contrast with well-known results in extremal decomposition, Corollary 6.3


is in fact a ‘covering theorem’. We also see that for some quadratic differentials
we meet with the phenomenon of domain break-up: while a connected component
of B can contain several points ak , in the extremal case each component Gk of B
contains only one point, ak . We shall explain this using the quadratic differential

z 4 + 6z 2 + 1 2
Q(z)dz 2 = − dz .
z 2 (z 2 − 1)2

It is regular on the whole of C apart from the second-order poles at 0, ∞, and


±1. We can easily analyze the structure of its trajectories and see that it has
no orthogonal trajectories with one end-point at 0 and the other at ∞. Let Bk ,
k = 1, . . . , 4, be pairwise nonoverlapping domains (with the possible exception of
B1 and B2 ) which contain the points z1 = 0, z2 = ∞, z3 = 1, and z4 = −1,
respectively. Then from Corollary 6.3 we obtain
= √ √
r(B, 0)r(B,
 ∞)r(B3 , 1)r(B4 , −1) exp(g  (0, ∞))  8( 2 − 1)2 2 ,
B

where B  = B1 ∪ B 2 .
If B is the union of n pairwise nonoverlapping simply connected domains Bk
such that ak ∈ Bk , k = 1, . . . , n, but B contains no first-order poles, then the
assumptions of Corollary 6.3 certainly hold and (6.7) takes the following form:
n n
2 2
(r(Bk , ak ))xk  (r(Gk , ak ))xk . (6.9)
k=1 k=1
166 Chapter 6. Extremal Decomposition Problems

Jenkins deduced this from the ‘general coefficient theorem’ [J]. It is known that in
several special cases the right-hand side of (6.7) (or (6.9)) can be found explicitly.
This makes it possible to refine the corresponding estimates due to Lavrent’ev,
Goluzin, Kufarev and Fales, Kolbina and other authors. As an example, we con-
sider here Goluzin’s result on three simply connected nonoverlapping domains in
C [Gol], Ch. IV, § 4. Let Bk , k = 1, 2, 3,, be some simply connected domains in
3
C containing points ak ∈ Bk , k = 1, 2, 3,, and let B = Bk . Taking account of
k=1
linear fractional automorphisms of C, we can assume that ak = exp(2πi(k − 1)/3),
k = 1, 2, 3. We set G = C and
−9z
Q(z)dz 2 = dz 2
(z 3 − 1)2
in Corollary 6.3. The domains Gk , k = 1, 2, 3, are sectors bounded by the rays
arg z 3 = π. If (6.7) holds with |xk | = 1, k = 1, 2, 3, then
 3 
r(B, ak ) exp{2x1 x2 gB (a1 , a2 )
k=1 (6.10)
 3
4
+ 2x2 x3 gB (a2 , a3 ) + 2x3 x1 gB (a3 , a1 )} 
3

3
with equality for B = Gk .
k=1
Next we set B 2k−1 to be the connected component of {z ∈ B : 2π(k − 1)/3 <
arg z < 2πk/3} with boundary point ak , k = 1, 2, 3; and we set B 2k−2 to be
the connected component of {z ∈ B : 2π(k − 2)/3 < arg z < 2π(k − 1)/3} with
boundary point ak , k = 2, 3, 4 (a4 = a1 ). Suppose that B 1 ∩ B2 = B3 ∩ B4 =
 
B5 ∩ B6 = ∅. Then the assumptions of Corollary 6.3 with xk = 1, k = 1, 2, 3,
hold for the set B and we obtain (6.10). If only one intersection is empty, for
1 ∩ B
instance, B 2 = ∅, then (6.10) holds again, this time for x1 = x2 = 1 and
x3 = −1. Recall that Goluzin proved (6.10) for x1 = x2 = x3 = 0, provided that
the domains Bk , k = 1, 2, 3, are pairwise disjoint.
In Theorem 6.5 we can take the strip domains of a quadratic differentials
equal to a full square for the strips in the set {P }. The next result refines Theo-
rem 6.1.
Corollary 6.4. Let ak , k = 1, . . . , n, be different finite points in C and let δk and
xk , k = 1, . . . , n, be real numbers such that δ1 +δ2 +· · ·+δn = 0 and |δk | = |xk | = 0,
k = 1, . . . , n (n  2). Assume that an open set B ⊂ C contains the ak , k = 1, . . . , n,
and does not contain trajectories of the quadratic differential
 n 2
δk
Q(z)dz 2 = dz 2
z − ak
k=1
6.1. Fixed poles 167

with limit end-points ak and ak such that xk xk > 0, 1  k, k   n. Then

n
n
n
n
n
x2k log r(B, ak ) + xk xl gB (ak , al )  − δk δl log |ak − al |.
k=1 k=1 l=1 k=1 l=1
l=k l=k

Equality sign holds here if xk = δk , k = 1, . . . , n and


 
n
B= z: δk log |z − ak | = c ,
k=1

where c is an arbitrary real constant.


Proof. This is similar to the proof of Corollary 6.3, except that
 n
F (z) = [−Q(z)]1/2 dz = i δk log |z − ak | + c.
k=1

The equality cases are described with the help of Theorem 2.9. 
Note that the numbers δk corresponding to the two end-points of a trajectory
of a strip domain of the quadratic differential Q(z)dz 2 have opposite signs. Hence
if xk = δk , k = 1, . . . , n, then the assumptions of Corollary 6.4 certainly hold.
In particular, if B is the union of pairwise nonoverlapping domains Bk , ak ∈ Bk ,
k = 1, . . . , n, then Corollary 6.4 yields Nehari’s inequality (Theorem 6.1).
In conclusion we present several results for the Robin radii of nonoverlapping
domains. Throughout what follows we only consider finitely connected domains
B ⊂ C without isolated boundary points and nonempty closed subsets Γ of ∂B
consisting of finitely many nondegenerate connected components.
Theorem 6.8. The inequality
' (−1
r(B1 , Γ1 , a1 )r(B2 , Γ2 , a2 )  |a2 − a1 |2 1 − |(a2 − a1 )/(1 − a1 a2 )|2 (6.11)

holds for any domains Bk , sets Γk , and points ak satisfying

B1 ∩ B2 = ∅, ak ∈ Bk ⊂ U, (∂Bk ) ∩ U ⊂ Γk ⊂ ∂Bk , k = 1, 2,

For arbitrary fixed points a1 and a2 in U equality is attained in (6.11) if B1 ∪ B2 =


U and the common boundary of B1 and B2 coincides with Γ1 = Γ2 and is a non-
Euclidean straight line orthogonal to the non-Euclidean straight line joining a1 and
a2 which passes through the midpoint of the line interval between these points.
Proof. For definiteness assume that ak = 0, k = 1, 2, the set (∂B1 ) \ Γ1 is empty,
but γ := (∂B2 ) \ Γ2 is nonempty. We look at the condensers
C1 = C(r; B1 , ∂B1 , {a1 }, {−1}, {r}),
C1∗ = C(r; B1∗ , ∂B1∗ , {1/a1 }, {−1}, {r/|a1|2 }),
168 Chapter 6. Extremal Decomposition Problems

C2 = C(r; B2 , Γ2 , {a2 }, {1}, {r}),


C2∗ = C(r; B2∗ , ∂B2∗ , {a2 , 1/a2 }, {1, 1}, {r, r/|a2|2 }),
C3 = C(r; C, ∅, {a1 , a2 , 1/a1 , 1/a2 }, {−1, 1, −1, 1}, {r, r, r/|a1|2 , r/|a2 |2 }),
where B1∗ = {z : 1/z ∈ B1 } and B2∗ = B2 ∪ γ ∪ {z : 1/z ∈ B2 }. Since the capacity
is a conformal invariant, it follows from the symmetry principle that
cap C1 = cap C1∗ , cap C2∗ = 2 cap C2 .

Extending the domains B1 , B1∗ , and B2∗ if necessary so that B1 ∪ B1∗ ∪ B2∗ =
C, but the domains remain disjoint, and using the monotonicity of the capacity
and the composition principle (Theorem 1.18) we obtain
cap C1 + cap C1∗ + cap C2∗  cap C3 .
Thus
2 cap C1 + 2 cap C2  cap C3 .
Hence we obtain (6.11) by letting r → 0 and taking account of the asymptotic
formulae for the capacities of degenerating condensers (Theorem 2.1 and 2.2),
which yield
 2  2 
2π 1 1
capC1 = − − 2πlogr(B1 ,a1 ) +o , r → 0,
logr logr logr
 2  2 
2π 1 1
capC2 = − − 2πlogr(B2 ,Γ2 ,a2 ) +o , r → 0,
logr logr logr
 2  2 
8π |a1 − a2 |2 |1 − a1 a2 |2 1 1
capC3 = − − 4πlog +o , r → 0.
logr ||a1 |2 − 1|||a2 |2 − 1| logr logr

The equality cases in (6.11) are straightforward. In a similar way we consider the
cases when ak = 0 for some point, or both the sets (∂Bk ) \ Γk , k = 1, 2, are empty
or both of them are nonempty. 
Let K be the annulus t < |z| < 1 and assume that a function T maps K
conformally and univalently onto a Teichmüller ring, the plane cut along the union
of some interval [T (1), −1], T (1) < −1, and the nonnegative half of the real axis,
from 0 to infinity. Clearly, T is mirror-symmetric relative to the real axis and
T (t) = ∞.
Theorem 6.9. For any domains Bk , sets Γk , and points ak such that
B1 ∩ B2 = ∅, ak ∈ Bk ⊂ K, (∂Bk ) ∩ K ⊂ Γk ⊂ ∂Bk , k = 1, 2,
the following inequality holds:
|T (|a2 |) − T (−|a1 |)|2
r(B1 , Γ1 , a1 )r(B2 , Γ2 , a2 )  . (6.12)
|T  (|a2 |)T  (−|a1 |)|
6.1. Fixed poles 169

Equality holds in (6.12), for instance, if a1 < 0, a2 > 0, B1 ∪ B2 = K, and the


common boundary of the domains B1 and B2 coincides with Γ1 = Γ2 = {z ∈ K :
|T (z) − T (a1 )| = |T (z) − T (a2 )|}.
Proof. We look at the condensers C1 = C(r; B1 , Γ1 , {a1 }, {−1}, {r/|T (−|a1 |)|})
and C2 = C(r; B2 , Γ2 , {a2 }, {1}, {r/|T (|a2 |)|}). By Theorem 2.2

cap C1 + cap C2
 2
4π 1
=− − 2π{log[|T  (−|a1 |)T  (|a2 |)|r(B1 , Γ1 , a1 )r(B2 , Γ2 , a2 )]}
log r log r
 2 
1
+o , r → 0.
log r

On the other hand, by Theorem 4.2

cap C1 + cap C2  4 cap C  4 cap Cr C,

where C = (K, {D(a2 , r/|T  (|a2 |)|), D(a1 , r/|T  (−|a1 |)|)}, {0, 1}). Since the capac-
ity is a conformal invariant, it follows from (2.15) that

4 cap Cr C = cap C(r; C, ∅, {T (−|a1|), T (|a2 |)}, {−1, 1}, {r, r})
 2  2 
4π 1 1
=− − 4π{log |T (|a2 |) − T (−|a1 |)|} +o , r → 0.
log r log r log r

Adding together the above relations we obtain (6.12).


Now assume that a1 < 0 and a2 > 0, and let the domains Bk and sets Γk ,
k = 1, 2 be as defined at the end of the statement of Theorem 6.9. Taking suitable
smooth almost discs in the definition of the condensers C1 and C2 we obtain
equalities for capacities in all the above relations. This yields equality in (6.12).

Note that the equality case in (6.12) described in Theorem 6.9 is far from
being unique. Apparently, we shall describe all the equality cases if apart from the
function T we also consider its composites with linear fractional automorphisms
of the complex sphere preserving the real axis. If the sets (∂Bk ) \ Γk , k = 1, 2, lie
on the same boundary circle of K, inequality (6.12) can be improved as follows.
Theorem 6.10. For any domains Bk ⊂ K, sets Γk ⊂ ∂Bk , and points ak ∈ Bk ,
k = 1, 2, such that B1 ∩ B2 = ∅, and {z ∈ ∂Bk : t  |z| < 1} ⊂ Γk the following
inequality holds:

r(B1 , Γ1 , a1 )r(B2 , Γ2 , a2 )
16|T (|a2|) − T (−|a1 |)|2 (6.13)
   4
|T  (|a2 |)T  (−|a1 |)| 4 T (|a2 |)/T (−|a1|) + 4 T (−|a1 |)/T (|a2 |)
170 Chapter 6. Extremal Decomposition Problems

(the positive value of root functions are taken throughout). Equality holds in (6.13)
for a1 < 0, a2 > 0, and B1 ∪ B2 = K, when the common boundary of the domains
B1 and B2 is given by the equation
 
−T (|a2 |)|T (z) − T (−|a1 |)| = −T (−|a1|)|T (z) − T (|a2 |)|, z ∈ K,
and the sets Γ1 and Γ2 are formed by this boundary and some arcs of the circle
|z| = t.
Proof. In the main this repeats the proof of the previous theorem. At the end of the
proof the condenser C(r; C, ∅, {T (−|a1|), T (|a2 |)}, {−1, 1}, {r, r}) must be replaced
with C(r; C \ Γ, Γ, {T (−|a1|), T (|a2 |)}, {−1, 1}, {r, r}), where Γ is the nonnegative
real half-axis. By the formula in Theorem 2.2, this condenser has capacity

− − 2π
log r
  
 −T (|a |) + −T (−|a |) 
 2 1 
× log |16T (|a2 |)T (−|a1 |)| − 2 log    
 −T (|a2 |) − −T (−|a1 |) 
 2  2 
1 1
× +o , r → 0,
log r log r

which yields (6.13). The equality case is also treated similarly to Theorem 6.9,
with the same replacement of the condenser. 
The right-hand side of (6.13) is strictly smaller than the right-hand side of
(6.12). This can immediately be verified, but it is simpler to compare the capacities
of the condensers under consideration, or more precisely, the second terms in the
expansions for these capacities.

Exercises 6.1
(1) Prove (6.1) using only the fact that the capacity of a condenser is monotonic
and verify that this inequality is equivalent to (6.1 ). Investigate the equality
cases in these relations.
(2) Deduce (6.3) from (6.1 ) using only conformal mappings.
(3) Prove the Duren–Schiffer inequality [DurSch]

n
n
n
αl αk log |al − ak | + α2k log r(Bk , ak )  s2 log R,
k=1 l=1 k=1
l=k


n
where the αk , k = 1, . . . , n, are arbitrary real numbers, s = αk , Bk ,
k=1
k = 1, . . . , n, are pairwise nonoverlapping simply connected subdomains of
the complement of some domain Ω, ∞ ∈ Ω, and cap(C \ Ω) = R.
6.1. Fixed poles 171

(4) Verify that Theorem 6.3 is indeed a consequence of Theorem 2.7.


(5) Recover the proof of Corollary 4.2.
(6) Under the assumptions of Theorem 6.4, for sufficiently smooth domains Bk
√ √
look at the reduced moduli M (Bk , ∅, ak − ρeiϕk , {ak + ρeiϕk }, {− αk , αk },
{r, r}), k = 1, . . . , n. Replacing the moduli in the proof of Theorem 6.4 with
these moduli and using Theorem 2.14 (in place of Theorem 2.8) deduce an
inequality for the classical Neumann functions N (z, ak ). Compare the result
with (6.5) in the case of simply connected domains Bk , k = 1, . . . , n.
(7) Following the proof of Theorem 6.5 and our comments give full proofs of
Theorems 6.6 and 6.7.
(8) Under the assumptions of Corollary 6.3 let all the xk be positive numbers
and let B be the union of domains Bk , ak ∈ Bk , k = 1, . . . , n, which have
classical Green functions. Using Theorem 2.7 and the proof of Theorem 6.5
show that equality sign can only hold in (6.9) when the level curves of the
functions gBk (z, ak ) are the closures of trajectories of the quadratic differ-
ential Q(z)dz 2 . In particular, if the Bk , k = 1, . . . , n, are simply connected
domains, then equality in (6.9) holds if and only if Bk = Gk , k = 1, . . . , n.
(9) Under the assumptions of Corollary 6.4 let xk = δk , k = 1, . . . , n, let B
be the union of domains Bk , which have classical Green functions, and let
ak ∈ Bk , k = 1, . . . , n. Show that
 equality sign holds in the$ inequality in
n
this corollary if and only if B = z : δk log |z − ak | = c , where c is an
k=1
arbitrary real constant.
(10) (Kolbina [Kol]) Show that for any different finite points a1 and a2 , any dif-
ferent positive numbers α and β, and any nonoverlapping simply connected
domains B1 , B2 ⊂ C such that ak ∈ Bk , k = 1, 2,
√ √ 2√αβ
4α+β αα β β  α − β 
(r(B1 , a1 )) (r(B2 , a2 )) 
α β
√ √ |a1 − a2 |α+β .
|α − β|α+β  α + β 

Hint: for α < β consider the quadratic differential

(a2 − a1 )[z − a1 − α(a2 − a1 )/(α − β)] 2


− dz
(z − a1 )2 (z − a2 )2

and use Corollary 6.3.


(11) For arbitrary pairwise nonoverlapping domains B1 , B2 , and B3 in C which
contain points 0, −i, and i, respectively, and for any α, 0 < α < 2, prove the
inequality
2 2
+6 α2 2 2
(r(B1 , 0))α r(B2 , −i)r(B3 , i)  2α α (2 − α)−(2−α) /2
(2 + α)−(2+α) /2 .
(6.14)
172 Chapter 6. Extremal Decomposition Problems

Equality is attained for the circle domains of the quadratic differential

z 2 (α2 − 4) + α2 2
Q(z)dz 2 = − dz .
z 2 (1 + z 2 )2

(12) Using Corollary 6.3 prove the inequality


2 2 2 2
(r(B1 , 0)r(B4 , ∞))α r(B2 , −i)r(B3 , i)  4α2α |1 − α|−(1−α) (1 + α)−(1+α)
(6.15)
for any positive number α and any pairwise nonoverlapping domains Bk ⊂ C,
k = 1, 2, 3, 4, which contain the points 0, −i, i, ∞, respectively, (for α = 1 the
right-hand side of (6.15) is equal to 1/4). Verify that if α  1 and B1 ∩B4 = ∅,
B2 ∩ B3 = ∅ or α  1 and B2 ∩ B3 = ∅, B1 ∩ B4 = ∅, then inequality (6.15)
also holds.
Hint: look at the quadratic differential

α2 z 4 + (2α2 − 4)z 2 + α2 2
Q(z)dz 2 = − dz .
z 2 (1 + z 2 )2

(13) Deduce (6.11) from Theorem 6.6.

6.2 Free poles


Let E be some fixed subset of the Riemann sphere C, and let αk , k = 1, . . . , n
(n  2), be fixed positive numbers. We shall consider the following problem:
n
(r(Bk , ak ))αk → sup;
k=1
ak ∈ E ∩ Bk , k = 1, . . . , n, Bk ∩ Bl = ∅, k = l, k, l = 1, . . . , n.

If E is an arbitrary compact set and α1 = α2 = 1 (n = 2), then it follows from


Lavrent’ev’s inequality (6.1 ) that the required supremum is the square of the
diameter of E. Note that the precise value of the supremum of the above product
over arbitrary fixed points ak and pairwise nonoverlapping domains Bk is only
known for small values of n. Nevertheless, in some particular cases this problem
with free poles ak ∈ E can be solved for any n  2. First we look at the simplest
example.
Theorem 6.11. For any different points ak on the circle |z| = 1, k = 1, . . . , n
(n  2), and any pairwise nonoverlapping domains Bk ⊂ C such that ak ∈ Bk ,
k = 1, . . . , n,
n  n
4
r(Bk , ak )  . (6.16)
n
k=1
6.2. Free poles 173

In addition, if the Bk , k = 1, . . . , n, are admissible domains, then equality holds in


(6.16) if and only if ak = exp(i(θ + 2πk/n)) and Bk = {z : |arg z − θ − 2πk/n| <
π/n}, k = 1, . . . , n, where θ is an arbitrary real constant.
Proof. We use the notation θk = arg ak , k = 1, . . . , n, θ0 = θn − 2π, θn+1 = 2π,
ϕk = θk+1 − θk , k = 0, 1, . . . , n. We can assume that 0 = θ1 < θ2 < · · · < θn < 2π.
We also set

Dk = {reiθ : 0 < r < ∞, θk < θ < θk+1 },


ζ = pk (z) ≡ −i(ze−iθk )π/ϕk , z ∈ Dk , Re ζ > 0, k = 0, 1, . . . , n.

Each pair {pk−1 , pk } makes up an admissible family of functions for a separating


(1) (2)
transformation of open sets with respect to the point ak . Let {Bk , Bk } be the
(1) (1)
result of this transformation of Bk , k = 1, . . . , n, and let Bn+1 = B1 . From

Theorem 4.10 and (6.1 ) we conclude that
n n =
(1) (2)
r(Bk , ak )  (ϕk−1 ϕk /π 2 )r(Bk , i)r(Bk , −i)
k=1 k=1
n =  n (6.17)
ϕk (2) (1) 4
= r(Bk , −i)r(Bk+1 , i)  ,
π n
k=1

with the corresponding assertion about equality sign. 


Now we add a domain B0 containing the origin to the
domains B1 , . . . , Bn in Theorem 6.11 and look at the C4
following problem:
b4
n
r(Bk , ak ) → sup;
k=0 C2 b3
a0 = 0, |ak | = 1, k = 1, . . . , n, C5
ak ∈ Bk , k = 0, 1, . . . , n, b5 C3
Bk ∩ Bl = ∅, k = l, k, l = 0, 1, . . . , n.

It is considerably more difficult than the first prob- Figure 6.1: The extremal
lem. In fact, we may conjecture that for some domains configuration in Theorem
Bk , k = 1, . . . , n, positioned without symmetry the 6.12 for n = 3.
product of their inner radii is smaller than for some
symmetrically positioned domains. On the other hand, dissymmetrization strictly
increases the inner radius of the domain B0 . Nevertheless, the following result
holds.
Theorem 6.12. For any different points ak , k = 0, 1, . . . , n (n  2), a0 = 0,
|ak | = 1 for k = 1, . . . , n, and any pairwise nonoverlapping domains Bk ⊂ C,
174 Chapter 6. Extremal Decomposition Problems

ak ∈ Bk , k = 0, 1, . . . , n,
n  2
4n+1/n nn n−1
r(Bk , ak )  2 . (6.18)
(n − 1)n+1/n n+1
k=0

In addition, if the domains Bk , k = 0, 1, . . . , n, are admissible, then equality holds


in (6.18) if and only if the Bk and ak are the circle domains and poles of a
quadratic differential

(eiθ z)n (n2 − 1) + 1 2


Q(z)dz 2 = − dz ,
z 2 ((eiθ z)n − 1)2

where θ is some real constant.


(1) (2)
Proof. Let θk , ϕk , Dk , pk (z), Bk , Bk be as in the proof of Theorem 6.11. Then
the family of functions {pk }nk=1 is admissible for the separating transformation of
(k)
open sets with respect to z = 0. Let {B0 }nk=1 be obtained from the domain B0
by this transformation. Then by Theorem 4.10
 ϕ 2
n A B1 k
(k) 2 π
r(B0 , a0 )  r(B0 , 0) . (6.19)
k=1

For each k we look at the quadratic differential

−ζ 2 (4 − (ϕk /π)2 ) + (ϕk /π)2 2


Qk (ζ)dζ 2 = − dζ .
ζ 2 (1 + ζ 2 )2

It is regular on Cζ apart from the second-order poles at ζ = 0, −i, i; and in a


neighbourhood of these points we have the expansions

Qk (ζ) = −(ϕk /π)2 /ζ 2 + . . . , Qk (ζ) = −1/(ζ + i)2 + · · · ,


Qk (ζ) = −1/(ζ − i)2 + · · · .
(k) (2) (1)
Now, the domains B0 , Bk , Bk+1 are (generally speaking, multiply connected)
pairwise nonoverlapping domains in Cζ . Their union contains only finitely many
closures of orthogonal trajectories of Qk (ζ)dζ 2 since such a trajectory either has
an end-point at a zero of Qk (ζ)dζ 2 or both its end-points lie at different poles of
Qk (ζ)dζ 2 (ϕk < 2π). Corollary 6.3, taking account of Exercise 6.1(11), yields
ϕ 2 A B(ϕk /π)2
k (k) (2) (1)
r(B0 , 0) r(Bk , −i)r(Bk+1 , i)  Φ(ϕk /π), (6.20)
π
where
2
+6 u2 +2 2 2
Φ(u) = 2u u (2 − u)−(2−u) /2
(2 + u)−(2+u) /2
, 0 < u < 2.
6.2. Free poles 175

Now we look at the extremal problem


n
n
Φ(uk ) → sup; uk = 2. (6.21)
k=1 k=1

The necessary conditions here have the following form:


n
Φ (uk )/Φ(uk ) = −λ/ Φ(uk ), k = 1, . . . , n. (6.22)
k=1

We claim that all the uk are equal. It is easy to verify that the function F (u) =
Φ (u)/Φ(u) is strictly decreasing on some interval (0, u0 ] and increasing on [u0 , 2),
u0 > 1. We introduce the auxiliary function F(u) = F (u) − F (2 − u), 0 < u  1.
Then simple calculations show that F(u) is positive on (0, 1). If one of the uk is
greater than 1 (for instance, uk > 1), then for the other uk we have uk  2−uk <
1 < u0 , and therefore
F (uk )  F (2 − uk ) > F (2 − (2 − uk )) = F (uk ).
This contradicts (6.22). Hence uk ∈ (0, 1] for each k. Since F (u) is monotone on
(0, u0 ], it follows by (6.22) that uk = 2/n, k = 1, . . . , n, at the supposed extremum
point. We readily see that the point (2/n, . . . , 2/n) solves indeed problem (6.21).
In combination with (6.17), (6.19), and (6.20) this gives us the required inequality.
Now let Bk , k = 0, 1, . . . , n, be admissible domains and suppose that equality
holds in (6.18). Then necessarily ϕk = 2π/n, k = 1, . . . , n, and the theorem follows
by applying Exercise 6.1(8) to the quadratic differential Q(z)dz 2 with arbitrary
constant θ in the statement. We can easily find the value of the expression on
the right-hand side of (6.9) if we use Exercise 6.1(11) and the following fact: the
differential Q(z)dz 2 can be obtained from Qk (ζ)dζ 2 (ϕk = 2π/n) by the change
of variable ζ = i(eiθ z)n/2 . 
The statement of the following problem is borrowed from Bakhtina’s papers
(see [Bakh]):
n
r(Bk , ak ) → sup;
k=0
a0 = 0, |ak | = 1, k = 1, . . . , n, ak ∈ Bk , k = 0, 1, . . . , n,
Bk ∩ Bl = ∅, k = l, k, l = 0, 1, . . . , n, Bk = {z : 1/z ∈ Bk }, k = 1, . . . , n.
Theorem 6.13. Let ak , k = 1, . . . , n (n  2), be arbitrary different points in the
circle |z| = 1, let a0 = 0, and let Bk , k = 0, 1, . . . , n, be pairwise nonoverlapping
domains in C such that ak ∈ Bk , k = 0, 1, . . . , n, B0 ⊂ U , and the Bk with
k = 1, . . . , n, are mirror-symmetric relative to the circle |z| = 1. Then

n
 √  2
22n+1/n n− 2
r(Bk , ak )  2 √ .
(n − 2)n/2+1/n n + 2
k=0
176 Chapter 6. Extremal Decomposition Problems

In addition, if the Bk , k = 0, 1, . . . , n, are admissible domains, then equality holds


here if and only if the Bk and ak are the circle domains and poles of the quadratic
differential
(eiθ z)2n + (eiθ z)n (2n2 − 2) + 1 2
Q(z)dz 2 = − dz , (6.23)
z 2 ((eiθ z)n − 1)2
where θ is an real constant.
This is an immediate consequence of the following result.
Theorem 6.14. For any different points ak on the circle |z| = 1, k = 1, . . . , n
(n  2), for a0 = 0 and an+1 = ∞, and for any pairwise nonoverlapping domains
Bk ⊂ C such that ak ∈ Bk , k = 0, 1, . . . , n + 1,

 √  2
 n 2n+1/n
2 n− 2
r(B0 , a0 )r(Bn+1 , an+1 ) r(Bk , ak )  √ .
k=1
(n2 − 2)n/2+1/n n+ 2

In addition, if the Bk , k = 0, 1, . . . , n + 1, are admissible domains, them equality


holds here if and only if the Bk are the circle domains and the ak are the poles of
the quadratic differential (6.23), where θ is an arbitrary real constant.
Proof. Throughout the proof we use the notations from the proofs of Theorems
6.11 and 6.12. The family of functions {pk }nk=1 is admissible for the separation
(k)
transformation of open sets with respect to z = ∞. Let {Bn+1 }nk=1 be obtained
from the domain Bn+1 by this transformation. Then by Theorem 4.10
 ϕ 2
n A B1 k
(k) 2 π
r(Bn+1 , an+1 )  r(Bn+1 , ∞) . (6.24)
k=1

For each k we look at the quadratic differential

ζ 4 (ϕk /π)2 /2 − ζ 2 (4 − (ϕk /π)2 ) + (ϕk /π)2 /2 2


Rk (ζ)dζ 2 = − dζ .
ζ 2 (1 + ζ 2 )2

It is regular in Cζ apart from the second-order poles at ζ = 0, ∞, −i, i; in a


neighbourhood of these poles we have the expansions

Rk (ζ) = −[(ϕk /π)2 /2]/ζ 2 + · · · , Rk (ζ) = −[(ϕk /π)2 /2]/ζ 2 + · · · ,

Rk (ζ) = −1/(ζ + i)2 + · · · , Rk (ζ) = −1/(ζ − i)2 + · · · .


(l)
In accordance with Exercise 6.1(12), for the circle domains Gk of Rk (ζ)dζ 2 we
have
A B(ϕk /π)2 /2 √
(0) (3) (1) (2)
r(Gk , 0)r(Gk , ∞) r(Gk , −i)r(Gk , i) = (π/ϕk )2 Ψ(ϕk /π 2),
6.2. Free poles 177

where √
2 2 2
Ψ(v) = 8v 2v +2
|1 − v|−(1−v) (1 + v)−(1+v) , 0<v 2
(Ψ(1) = 1/2). Furthermore,

ϕ 2 A B1
 ϕ 2
k  
k (k) (k) 2 π (2) (1) ϕk
r(B0 , 0)r(Bn+1 , ∞) r(Bk , −i)r(Bk+1 , i) Ψ √ .
π π 2
(6.25)
Now we look at the extremal problem
n
n

Ψ(vk ) → sup; vk = 2.
k=1 k=1

Let H(v) = Ψ (v)/Ψ(v). Elementary calculations show √ that this is a decreasing


function on (0, v0 ] and an increasing function on [v
√ 2], where 0, 85 < v0 < 1.
0 ,
Bearing
√ also in mind that H(0, 564) > 0 and H(√ 2) < 0, we see that H(v) −
H( 2 − v) is positive on the interval 0√< v < √2/2. Similarly to the proof of
Theorem 6.12 we verify that the point ( 2/n, . . . , 2/n) is the unique solution of
the extremal problem. In combination with (6.17), (6.19), (6.24), and (6.25) this
yields the inequality in Theorem 6.14. The equality case, similarly to the proof of
Theorem 6.12, follows from Exercise 6.1(8). 

Theorem 6.15. For any points ak and bk such that C4


argak = arg bk , |ak | = ρ, |bk | = R, k = 1, . . . , n (n  c4
2), 0 < ρ < R < ∞, and any pairwise nonoverlapping b4
open sets Bk ⊂ C such that ak , bk ∈ Bk , k = 1, . . . , n, c5
b5 b3 c 3
n
r(Bk , ak )r(Bk , bk ) exp(−2gBk (ak , bk )) C5 C3
k=1
 √ 2n
4 ρR(Rn/2 − ρn/2 ) Figure 6.2: Illustration to
 , (6.26)
n(Rn/2 + ρn/2 ) Theorem 6.15.

In (6.26) equality sign holds for the domains Bk∗ = {z : 2πk/n < arg z < 2π(k +
1)/n} and points a∗k = ρ exp(i(π/n + 2πk/n)) and b∗k = R exp(i(π/n + 2πk/n)),
k = 1, . . . , n.
(1) (2)
Proof. Let θk , ϕk , Bk , and Bk be as in the proof of Theorem 6.11 k = 1, . . . , n.
Using Theorems 4.8 and 1.18 we obtain


n
cap C(r; Bk , ∂Bk , {ak , bk }, {1, −1}, {r, r})
k=1
n    π $
1 (1) (1)
π
 ϕ
cap C r; Bk , ∂Bk , iρ , iR
k ϕ k , {1, −1},
2
k=1
178 Chapter 6. Extremal Decomposition Problems

 $   $
π ϕπ −1 π ϕπ −1 (2) (2)
π π
ρ k r, R k r + cap C r; Bk , ∂Bk , −iρ ϕk , −iR ϕk ,
ϕk ϕk
 $
π ϕπ −1 π ϕπ −1
{1, −1}, ρ k r, R k r
ϕk ϕk
n A
1 (1) (1)
= cap C r; Bk+1 , ∂Bk+1 ,
2
k=1
 π π
$  $
π ϕπ −1 π ϕπ −1
ϕ
iρ , iR
k ϕ k , {−1, 1}, ρ k r, R k r
ϕk ϕk
  π π
$
(2) (2)
+ cap C r; Bk , ∂Bk , −iρ ϕk , −iR ϕk , {1, −1},
 $
π ϕπ −1 π ϕπ −1
ρ k r, R k r
ϕk ϕk
  $
1
n π π π π
 cap C r; C, ∅, −iR ϕk , −iρ ϕk , iρ ϕk , iR ϕk ,
2
k=1
 $
π ϕπ −1 π ϕπ −1 π ϕπ −1 π ϕπ −1
{−1, 1, −1, 1}, R k r, ρ k r, ρ k r, R k r .
ϕk ϕk ϕk ϕk

Replacing the capacities of condensers with their asymptotic expansions in Theo-


rems 2.1 and 2.5 we obtain
n
log r(Bk , ak )r(Bk , bk ) exp(−2gBk (ak , bk ))
k=1

ϕk √
ρ π/ϕk √
ρ n/2

n
2π 4 ρR 1 − R 4 ρR 1 − R
2 log
ρ π/ϕk  2n log
ρ n/2 .
k=1 1+ R n 1+ R

In the last relation we have used that log[x(1 − τ 1/x )/(1 + τ 1/x )] is a concave
function for 0 < x < 1, (0 < τ < 1). For the domains Bk∗ and points a∗k , b∗k ,
k = 1, . . . , n, we have equality signs in all the above inequalities. 

Taking the union of disjoint domains Bk and Bk for Bk in Theorem 6.15 we
obtain the following result.
Corollary 6.5 ([E]). Under the assumptions of Theorem 6.15, if Bk = Bk ∪ Bk ,
k = 1, . . . , n, where Bk and Bk are simply connected domains such that Bk ∩ Bk =
∅, and if ak ∈ Bk and bk ∈ Bk , k = 1, . . . , n, then the following sharp estimate
holds:
n  √ 2n
4 ρR(Rn/2 − ρn/2 )
r(Bk , ak )r(Bk , bk )  .
k=1
n(Rn/2 + ρn/2 )

In fact the condition that Bk and Bk are simply connected is excessive here.
6.2. Free poles 179

Corollary 6.6. For any points ak on the circle |z| = ρ, 0 < ρ < 1, and any pairwise
nonoverlapping domains Bk such that ak ∈ Bk ⊂ U , k = 1, . . . , n (n  2)
n  n
4ρ(1 − ρn )
r(Bk , ak )  .
n(1 + ρn )
k=1

Equality holds here for the points ρ exp(i(π/n + 2πk/n)) and domains {z ∈
U : 2πk/n < arg z < 2π(k + 1)/n}, k = 1, . . . , n.

Proof. This follows by applying Theorem 6.15, to the points ak and 1/ak and to
the sets Bk ∪ {z : 1/z ∈ Bk }, k = 1, . . . , n. 
Theorem 6.16. Let B be an open set with Green’s function which contains some
points ak , k = 1, . . . , 2n, lying on the circle |z| = 1 and arranged counterclockwise.
Then  2n   2n 2n 

r(B, ak ) exp (−1) gB (ak , al )  (2/n)2n .
k+l

k=1 k=1 l=1


l=k

Equality is attained, for example, for B ∗ = {z ∈ C \ {0} : arg z 2n = 0} and


a∗k = exp(i(π/2n + π(k − 1)/n)), k = 1, . . . , 2n.

Proof. We set

Z = {ak }2n 2n
k=1 , = {δk }k=0 , δ0 = 0, δk = (−1) , k = 1, . . . , 2n,
k

= {δk }2n
k=1 , Ψ = {r, . . . , r}.

Since the reduced moduli behave monotonically (see (2.18)), we have

M (B, ∂B, Z, , Ψ)  M (Z, , Ψ). (6.27)

By (2.16),
 2n 2n
2n

1
M (B, ∂B, Z, , Ψ) = log r(B, ak ) + k+l
(−1) gB (ak , al ) ,
2π(2n)2
k=1 k=1 l=1
l=k

while from (2.17) we obtain


 n n 
1
M (Z, , Ψ) = − (−1)k+l log |ak − al | .
2π(2n)2
k=1 l=1
l=k

It remains to use the inequality deduced before Theorem 5.3. The equality cases
are straightforward. 
180 Chapter 6. Extremal Decomposition Problems

Theorem 6.17. Let B1 and B2 be disjoint open subsets of the complex sphere C
such that B1 contains the origin and B2 contains some points ak such that arg ak =
1n 
2πk/n, k = 1, . . . , n; let  ak  = R. Then
k=1
   
2
n
n
n
(r(B1 , 0))n r(B2 , ak ) exp gB2 (ak , al )  n−n Rn+1 . (6.28)
k=1 k=1 l=1
l=k

Equality
 sign holds in (6.28) for the sets B1∗ and√B2∗ bounded by the curves z =
n
R/2 + it, −∞ < t < +∞ and the points a∗k = n R exp(2πik/n), k = 1, . . . , n.
Proof. We set Z = {0, a1 , . . . , an }, = {0, −n, 1, . . . , 1}, = {−n, 1, . . . , 1}, and
Ψ = {r, . . . , r}. Using the fact that the reduced modulus is monotonic again, we
obtain (6.27) with B = B1 ∪ B2 . Now formulae (2.16) and (2.17) take us to the
inequality

n
n
n
n2 log r(B1 , 0) + log r(B2 , ak ) + gB2 (ak , al )
k=1 k=1 l=1
l=k

n
n
n
 2n log |ak | − log |ak − al |.
k=1 k=1 l=1
l=k

It remains to use the second special case of inequality (5.7):



n
n
n
log |ak − al |  log nn + (n − 1) log |ak |.
k=1 l=1 k=1
l=k

The equality case is straightforward to verify. 


Note that B2∗ is a union of n simply connected pairwise nonoverlapping do-
mains. In the extremal case the left-hand side of (6.28) has the following form:
n
2
(r(B1∗ , 0))n r(B2∗ , a∗k ).
k=1

Thus, (6.28) generalizes the corresponding inequality in the classical extremal


decomposition problem with free poles to domains which are allowed to overlap.
Other examples when extremal decomposition problems with free poles on
circles or rays are solved can be found in recent papers by Kuz’mina, Emel’yanov,
and Bakhtin. The first two authors develop the extremal-metric approach [Kuz1],
while Bakhtin uses variational methods and the separating transformation of do-
mains [Bakh]. In some cases our results enable us to solve problems with free
poles lying in subsets different from rays or circles. For example, using the lin-
6.2. Free poles 181

ear fractional transformation z = (1 − iζ)/(1 + iζ), which takes the real axis to
|z| = 1, we derive the following result from Theorem 6.11: for arbitrary points ak
on the real axis and pairwise nonoverlapping domains Bk such that ak ∈ Bk ⊂ Cζ ,
k = 1, . . . , n, n  2, the sharp bound
n  n
r(Bk , ak ) 2

1 + a2k n
k=1

holds. Setting here λak with any positive λ in place of ak , we easily obtain a sharp
bound for the product of inner radii of the nonoverlapping domains in the case
when the points ak lie on a fixed interval. The following is yet another result in
this direction.
Theorem 6.18. Let ak , k = 1, . . . , n, n  1, be points in the interval (−1, 1); let
Bk be pairwise nonoverlapping domains such that ak ∈ Bk ⊂ C, k = 1, . . . , n,, and
n
assume that the complement C\ Bk contains a continuum joining −1 to 1.
k=1
Then  n
n
r(Bk , ak ) 2
  .
k=1
1 − a2k n

Equality sign holds here for ak = cos[π(2k − 1)/(2n)], k = 1, . . . , n, and for the
domains Bk bounded by the curves {z = (ζ + 1/ζ)/2 : ζ 2n ∈ [0, 1]}.

Proof. To simplify the calculations (and skip writing out formulae for the capaci-
ties of condensers again) we use inequality (6.16). We can assume that the comple-

n
ment C \ Bk contains a Jordan curve γ joining −1 to 1. Let f1 and f2 be the
k=1
branches of the inverse of the Zhukovskii (Joukowsky) transform z = (ζ + 1/ζ)/2
in the simply connected domain C \γ. Since the domains fj (C\γ), j = 1, 2, are dis-
joint, the domains fj (Bk ), j = 1, 2, k = 1, . . . , n, are also pairwise nonoverlapping
and =
r(fj (Bk ), fj (ak )) = r(Bk , ak )/ 1 − a2k , j = 1, 2, k = 1, . . . , n.

Using (6.16) we obtain


n 2 n  2n
(r(Bk , ak ))2 4
= r(fj (Bk ), fj (ak ))  .
1 − a2k j=1
2n
k=1 k=1

The assertion on the equality case is a consequence of the equalities in (6.16). 

Note that if the Bk , k = 1, . . . , n, are simply connected pairwise nonover-


lapping domains which do not contain the points −1 and 1, then they satisfy the
assumptions of Theorem 6.18.
182 Chapter 6. Extremal Decomposition Problems

Corollary 6.7. For arbitrary different points ak on the positive real half-axis, k =
1, . . . , n, n  1, and for any pairwise nonoverlapping domains Bk such that ak ∈

n
Bk ⊂ C, k = 1, . . . , n, and the complement C \ Bk contains a continuum
k=1
joining 0 to ∞,
n  n
r(Bk , ak ) 2
√  .
(1 + ak ) ak n
k=1

Equality is attained here for ak = cot2 [π(2k − 1)/(4n)], k = 1, . . . , n, and the


domains Bk bounded by the curves {z = −(ζ + 1)2 (ζ − 1)−2 : ζ 2n ∈ [0, 1]}.
Proof. This follows from Theorem 6.18 with the help of the linear fractional map
z = (ζ − 1)/(ζ + 1), which takes the positive real half-axis to the interval (−1, 1).

Now we consider the extremal decomposition prob-
V )b3 < s* lem with free poles for products of the Robin radii
C3 3 r(Bk , Γk , ak ) in place of the inner radii r(Bk , ak ).
Theorem 6.19. Let ak , k = 1, . . . , n, n  2, be points
on the circle |z| = ρ, 0 < ρ < 1; let Bk ⊂ U be sim-
ply connected pairwise nonoverlapping domains such that
ak ∈ Bk , k = 1, . . . , n,, and let Γk ⊂ ∂Bk , k = 1, . . . , n,
be closed sets such that each set (∂Bk )\Γk is either empty
or equal to an arc of the circle |z| = 1. Then
Figure 6.3.  n
n n
4ρ 1 + ρn
r(Bk , Γk , ak )  r(Bk∗ , Γ∗k , a∗k ) = ,
n 1 − ρn
k=1 k=1
(6.29)
where
a∗k = ρ exp(2πik/n), Bk∗ = {z ∈ U : | arg z − 2πk/n| < π/n},

and Γ∗k = {z ∈ B k : z n ∈ [−1, 0]}, k = 1, . . . , n.
Proof. We can assume that the domains Bk , k = 1, . . . , n, are bounded by Jordan
curves. For sufficiently
small r > 0 we look at the condensers Ck∗ (r) = (Bk∗ , {Γ∗k ,
U (a∗k , r)}, {0, 1} , k = 1, . . . , n (see Figure 6.3). By Theorem 2.2 their capacity
has the expression
 2  2 
∗ 2π ∗ ∗ ∗ 1 1
cap Ck (r) = − − 2π [log r(Bk , Γk , ak )] +o , r → 0.
log r log r log r


n
Let u∗ be a continuous function in U \ U (a∗k , r) which is mirror-symmetric
k=1
relative to the rays arg z 2n = 0, harmonic in Bk∗ \U (a∗k , r) apart from the ray
6.2. Free poles 183

arg z = arg a∗k , equal to 1 on the line intervals [0, (ρ − r)a∗k /|a∗k |], to zero on
the intervals [(ρ + r)a∗k /|a∗k |, a∗k /|a∗k | ] and the circle |z| = 1, and satisfies the
following conditions: ∂u∗ /∂n = 0 on ∂U (a∗k , r) and at the interior points of the
Γ∗k , k = 1, . . . , n. Then it is easy to see that

I(u∗ , Bk∗ \U (a∗k , r) = 4/ cap Ck∗ (r), k = 1, . . . , n.

Let v = Dis u∗ be obtained by a dissymmetrization of u∗ taking the points a∗k to


ak , k = 1, . . . , n. Then

n 
n
n

I(u∗ , U \ U (a∗k , r)) = I(v, U \ U (ak , r))  I(v, Bk \ U (ak , r) .
k=1 k=1 k=1

We map Bk conformally and univalently onto the disc Uw by a function gk such


that gk (ak ) = 0 and gk (ak ) > 0. Let rk = rk (r) be a real function such that
U (0, rk (r)) ⊃ gk (U (ak , r)), rk (r) ∼ gk (ak )r as r → 0, and vk (w) = v(gk−1 (w)) is
defined in the closure Ωk of the annulus Ωk := {w : rk (r) < |w| < 1} (maybe
after extending it to the outer circle by continuity). Since the Dirichlet integral is
conformally invariant, it follows from Theorem 4.2 that

I(v, Bk \ U (ak , r)) = I(vk , Ωk )


 cap(Ωk , {{w : vk (w) = 0}, {w : vk (w) = 1}}, {0, 1})
 cap Cr(Ωk , {{w : vk (w) = 0}, {w : vk (w) = 1}}, {0, 1})
 cap(Ωk , {[rk (r), 1] ∪ γk , [−1, −rk (r)]}, {0, 1}) = 4/ cap Ck (r),

where γk is the closed arc of the circle |w| = 1 with midpoint at z = 1 and
length equal to the length of gk ((∂Bk ) \ Γk ) (for some k this arc can degenerate
into a point) and Ck (r) = (Uw , {gk (Γk ), U (0, rk (r))}, {0, 1}), k = 1, . . . , n. Using
Theorem 2.2 again we obtain
 2
2π 1
cap Ck (r) = − − 2π[log r(Uw , gk (Γk ), 0)]
log rk (r) log rk (r)
 2 
1
+o , r → 0, k = 1, . . . , n.
log r

Summing the above relations while bearing in mind that

r(Bk , Γk , ak ) gk (ak ) = r(Uw , gk (Γk ), 0), k = 1, . . . , n,

we obtain the inequality in (6.29). To verify the equality on the right-hand side
of (6.29) we map the sector Bk∗ conformally and univalently onto the ζ-plane cut
along the rays [−∞, 0] and [1/4, +∞] of the real axis by means of the mapping
zn
ζ = ζ(z) = .
(1 + z n )2
184 Chapter 6. Extremal Decomposition Problems

Then

r(Bk∗ , Γ∗k , a∗k )|ζ  (a∗k )| = r(Cζ \[−∞, 0], ρn(1 + ρn )−2 ) = 4ρn (1 + ρn )−2 ,

which yields the required equality. 


Our method of the proof of Theorem 6.19 allows us to extend it to pairwise
nonoverlapping subdomains of an annulus.
Theorem 6.20. The following equalities hold for arbitrary different points ak , k =
1, . . . , n (n  2), on the circle |z| = 1 and any pairwise nonoverlapping domains
Bk ⊂ C such that ak ∈ Bk , k = 1, . . . , n :

n
n(n2 − 1)
K(Bk , ak , arg ak + π/2)  , (6.30)
12π
k=1

n
n(n2 + 2)
K(Bk , ak , arg ak )  . (6.31)
24π
k=1

If ak = exp(i(θ + 2πk/n)) and Bk = {z : | arg z − θ − 2πk/n| < π/n}, k = 1, . . . , n,


then equality sign holds in (6.30) and (6.31). Here θ is an arbitrary real number.
Proof. Setting ϕk = arg ak + π/2 in Theorem 6.4 we obtain

n
1 −ak al
n n
K(Bk , ak , arg ak + π/2)  Re
π (ak − al )2
k=1 k=1 l=1
l=k

1
n n
1 n(n2 − 1)
=  .
π |ak − al |2 12π
k=1 l=1
l=k

We used (5.5) at the final step.


Inequality (6.31) follows from Theorem 6.15, in which we must set ρ = 1 − r
and R = 1 + r and, after simple transformations of (6.26), take the limit as r → 0.
It is easier to calculate the quantities K(Bk , ak , ·) in the equality case with
the use of the well-known formula for transformations of the Bergman kernels ([C],
§§ 2.44 and 2.45 in the Appendix). 

Exercises 6.2
(1) For k = 1, . . . , n (n  2) let zk = eiθk , where θ1 < θ2 < · · · < θn < θn+1 =
θ1 + 2π, let Pk = {z : θk < arg z < θk+1 }, let xk = ±1, and set xn+1 = x1 .
Let fk denote the regular branch of

1 (ze−iθk )π/(θk+1 −θk ) − 1


w= log ,
πi (ze−iθk )π/(θk+1 −θk ) + 1
6.2. Free poles 185

which maps Pk conformally and univalently onto the strip 0 < Re w < 1.
Consider an open set B containing the points zk , k = 1, . . . , n, but not
containing the curves Re fk (z) = u, 0 < u < 1, in the case when xk xk+1 > 0,
1  k  n. Prove the inequality


n
n
n
log r(B, zk ) + xk xl gB (zk , zl )
k=1 k=1 l=1
l=k

n
 log[2(θk+1 − θk )/π]  n log(4/n).
k=1

Verify that for any choice of xk equality holds for zk = exp(i(θ + 2πk/n)),

n
k = 1, . . . , n, and B = {z : |arg z−θ−2πk/n| < π/n}, where θ is arbitrary.
k=1
(2) (Kuz’mina [Kuz3]) Under the assumptions of Theorem 6.12 find the least
upper bound for the product
n
α
(r(B0 , a0 )) r(Bk , ak )
k=1

for arbitrary α, 0 < α < 1.


Hint: repeat the proof of Theorem 6.12 (α = 1) using the quadratic differen-
tial
(n2 − α)z n + α 2
Q(z)dz 2 = − 2 n dz .
z (z − 1)2

(3) Show that if α > n, then the product of powers of the inner radii in the
previous exercise has no upper bound.
(4) (Kovalev [K2]) Show that the condition B0 ⊂ U is excessive in Theorem 6.13.
(5) Find the least upper bound for the product

n−1
r(Bk , ak )
(r(B1 , −1)r(Bn , 1))1/4 
k=2
1 − a2k

over various points ak ∈ (−1, 1), k = 2, . . . , n − 1, and pairwise nonover-


lapping domains Bk , k = 1, . . . , n, such that ak ∈ Bk for k = 2, . . . , n − 1,
−1 ∈ B1 and 1 ∈ Bn .
(6) Following the proof of Theorem 6.19 establish the following result. Let ak ,
k = 1, . . . , n, n  2, be points on the circle |z| = ρ, 0 < t < ρ < 1; let Bk ,
ak ∈ Bk ⊂ K(t) = {z : t < |z| < 1}, k = 1, . . . , n, be simply connected
pairwise nonoverlapping domains and let Γk ⊂ ∂Bk , k = 1, . . . , n, be closed
186 Chapter 6. Extremal Decomposition Problems

sets such that for each k the difference (∂Bk ) \ Γk is empty or equal to an
arc of the circle |z| = 1, k = 1, . . . , n. Then
n n  n
4 |1 + Tn (ρn )|
r(Bk , Γk , ak )  r(Bk∗ , Γ∗k , a∗k ) = ,
nρn−1 |Tn (ρn )|
k=1 k=1

where a∗k = ρ exp(2πik/n), Bk∗ = {z ∈ K(t) : | arg z − 2πk/n| < π/n}, Γ∗k =
{z ∈ ∂Bk∗ : |z| = 1}, k = 1, . . . , n, and the function Tn is defined similarly to
the function T before Theorem 6.9 but involving tn in place of t.
(7) For fixed t and ρ, 0 < t < ρ < 1, find the maximum of the product
n
r(Bk , ak )
k=1

for various different points ak , |ak | = ρ, and pairwise nonoverlapping sub-


domains Bk of the annulus t < |z| < 1 such that ak ∈ Bk , k = 1, . . . , n.
Describe all the extremal configurations of points and domains.
(8) Prove the inequality

n π 1
(1 + a2k )2 K Bk , ak ,  n(n2 + 2)
2 6π
k=1

for arbitrary points ak on the real axis and pairwise nonoverlapping domains
Bk ⊂ C such that ak ∈ Bk , k = 1, . . . , n, n  2.
(9) Prove the equality a∗1 + a∗2 + · · · + a∗n = 0. for the extremal points a∗k in the
problem
n
r(Bk , ak ) → sup;
k=1
ak ∈ Bk ⊂ U, k = 1, . . . , n, Bk ∩ Bl = ∅, k = l, k, l = 1, . . . , n.

(10) Consider the following extremal problem ‘with free poles’:


n
r(Bk , Γk , ak )(1 − |ak |2 ) → sup, (6.32)
k=1

where the supremum is taken over all possible points ak , domains Bk ⊂ U ,


and sets Γk satisfying ak ∈ Bk , (∂Bk ) ∩ U ⊂ Γk ⊂ ∂Bk , and Bk ∩ Bl = ∅ for
k = l, k, l = 1, . . . , n, n  2. Show that if n = 2 then the supremum in (6.32)
cannot be attained and is equal to 16.
(11) Show that for fixed n  3 the product on the left-hand side of (6.32) is
bounded above. In addition, if the supremum in (6.32) is attained at some
points a∗k such that |a∗k | < 1, k = 1, . . . , n, then a∗1 + a∗2 + · · · + a∗n = 0.
6.3. Möbius invariants 187

6.3 Möbius invariants


Let B be an open subset of the Riemann sphere C which has a Green function
and let A = {ak }nk=1 be an ordered set of different points in B. Then by (2.16)

M (B, A) ≡ M (B, ∂B, A, {1, . . . , 1}, {r, . . . , r})


 n 
1 n n
= log r(B, ak ) + gB (ak , al ) .
2πn2
k=1 k=1 l=1,
l=k

We look at the following Möbius invariants

exp{2πn2 M (B, A)}


I(B, A) = 1 ,
|ak − al |2/(n−1)
1k<ln

exp{2πn2 M (B, A)}


J(B, A) = , an+1 = a1 , n  2,
1n

|ak − ak+1 |
k=1

where the dash on a product sign means that if one of the points ak is equal
to ∞, then the corresponding factor is set equal to 1. For n = 2 and n = 3
these two quantities coincide. Their invariance under the group of linear fractional
automorphisms of the sphere means that

I(B, {ak }nk=1 ) = I(ϕ(B), {ϕ(ak )}nk=1 )


and
J(B, {ak }nk=1 ) = J(ϕ(B), {ϕ(ak )}nk=1 )

for each Möbius transformation ϕ : C → C. Note that the Green function is a


conformal invariant, and it is straightforward that the quantity

r(B, a1 )r(B, a2 )
|a1 − a2 |2

is invariant under Möbius transformations. The other invariants are obtained by


multiplying fractions of this form and some factors involving Green’s functions
and cross ratios. In particular, for n  4 J(B, A) is different from I(B, A) by a
product of certain powers of the absolute values of the cross ratios of points in A.
In the case when B is a union of pairwise nonoverlapping domains Bk such that
ak ∈ Bk , k = 1, . . . , n, the following notation is used:
 n 

I(B, A) = In ≡ r(Bk , ak ) |ak − al |2/(1−n) .
k=1 1k<ln
188 Chapter 6. Extremal Decomposition Problems

If two or more domains in the set {Bk }nk=1 intersect, then


 

n
I Bk , A > In .
k=1

An upper estimate for the Möbius invariant In is equivalent to an extremal problem


with free poles:
n
r(Bk , ak ) → sup;
k=1

n n
ak ∈ Bk ⊂ C, Bk ∩ Bl = ∅, k = l, k, l = 1, . . . , n, |ak − al | = 1 (n  2).
k=1 l=1
l=k

Lavrent’ev’s inequality (6.1 ) ensures the sharp bound

I2  1.

By a theorem of Goluzin ([Gol], Ch. IV, § 5),

I3  3−9/2 43

provided that the Bk , k = 1, 2, 3, are simply connected domains (see our com-
ment on inequality (6.10)). Under the same assumption, Kuz’mina (see [Kuz2])
established the following result:

I4  4−8/3 32 .

One of the extremal configurations of domains Bk and points ak in this problem


is given in Figure 6.1.
Finding the least upper bound for the invariant In with n  5 is an exciting
complicated problem. So far, it has only been solved for n = 5, under the additional
assumption that the set {ak }5k=1 is mirror-symmetric relative to a circle or a
straight line (Theorem 6.23).
We start with a general result.
Theorem 6.21. For any open set G ⊂ C

sup I(B, {ak }nk=1 ) = sup I(B, {ak }nk=1 ),


ak ∈B⊂G ak ∈B⊂C
k=1,...,n k=1,...,n

sup J(B, {ak }nk=1 ) = sup J(B, {ak }nk=1 ).


ak ∈B⊂G ak ∈B⊂C
k=1,...,n k=1,...,n
6.3. Möbius invariants 189

Proof. We can assume without loss of generality that G contains the origin. Let
α be the left-hand side of the first relation and β be its right-hand side. By the
definition of the supremum α  β. Let β  < β and assume that for some open set
B  ⊂ C and an n-point subset A = {ak }nk=1 of B  we have

β  < I(B  , A ).

Making a linear fractional transformation we can assume that B  does not contain
the point at infinity. Then starting with some index j0 , each set Gj := {z : z/j ∈
G} contains all the points ak , k = 1, . . . , n. The sequence of sets
{B  ∩Gj }jj0 is an
  
exhaustion of B , that is, B ∩ Gj ⊂ B ∩ Gj+1 for j  j0 , and (B  ∩ Gj ) = B  .
jj0
Hence I(B  ∩ Gj , A ) → I(B  , A ) as j → ∞. Again, from the definition of the
supremum we obtain

α  I(B  ∩ Gj , A ) = I ({z ∈ G : zj ∈ B  }, {ak /j}nk=1 ) , j  j0 .

Hence
β  < α  β.
As β  < β can be arbitrary, we arrive at the required equality α = β. The case of
the invariant J(B, A) is similar. 
Thus, in investigating extremal decompositions with free poles which are
defined in terms of the quantities I(B, A) and J(B, A) we can consider sets B
positioned arbitrarily in C.
Lemma 6.1. Let B be an open set intersecting both upper and lower half-planes
and assume that in an ordered set A = {ak }nk=1 , where ak ∈ B, the number of
points lying in the upper half-plane is equal to the number of points lying in the
lower half-plane. Let B + denote the union of the part of B lying in the half-plane
{z : Imz  0} and its reflection in the real axis, and let B − denote the union of
B ∩ {z : Imz  0} and its reflection in the same axis. Let A+ = {a+ k }k=1 be the
n

subset consisting of the points in A lying in the upper half-plane {z : Imz  0}


and the points in {z : Imz < 0} symmetric to them if there are any, and let
A− = {a− k }k=1 be the subset consisting of the points in A lying in the lower half-
n

plane {z : Imz  0} and the ones in {z : Imz > 0} symmetric to them. Then

I 2 (B, A)  I(B + , A+ )I(B − , A− ).

Proof. The symmetry principle (Theorem 2.12) and composition principle (Theo-
rem 2.14) yield
2M (B, A)  M (B + , A+ ) + M (B − , A− ). (6.33)
Now let ak and al be two finite points in A, 1  k, l  n. If, for instance,
+
Im ak  0, and Im al > 0, then there exist points a+
s = ak and at = al such that

+ + + + +
|ak − al |2 = |a+ + +
s − at ||as − at | = |as − at ||ai − aj |,
190 Chapter 6. Extremal Decomposition Problems

for some i and j, j = t, 1  s, t  n, 1  i, j  n. If Im ak  0 and Im al < 0, then



in a similar way we see that there exist points a− s = ak and at = al such that

− − −
|ak − al |2 = |a−
s − at ||ai − aj |

for some i and j, j = t.


If Imak = Imal = 0, then
+ − −
|ak − al |2 = |a+
s − at ||ai − aj |

for some s, t, i, and j, t = j.


Finally, if Imak > 0 and Imal < 0, then it is easy to see that
+ − −
|ak − al |2  4Imak (−Imal ) = |a+
s − at ||ai − aj |

for some s, t, i, and j. Summarizing, we conclude that


 2

|ak − al |2/(n−1)
1k<ln
  
 |a+
k − a+
l |
2/(n−1)
|a−
k − a−
l |
2/(n−1)
.
1k<ln 1k<ln

Comparing this with (6.33) we arrive at the required inequality. 


Now we discuss several concrete upper bounds for the invariants I(B, A) and
J(B, A) in special cases when we can use the approach based on capacities and
symmetrization. First of all, note that we can deduce the following result from
Theorem 6.5. If an open set B contains no circle arcs connecting two fixed points
a1 and a2 and A = {a1 , a2 }, then

I(B, A) = J(B, A)  1.

Equality holds here for the sets B = {z : |(z − a1 )/(z − a2 )| = c}, where c is
an arbitrary positive number (see our comment after Theorem 6.7). In the case
when B is a union of two disjoint domains this result coincides with Lavrent’ev’s
theorem. Similar observations about classical estimates can be made for three or
more points in A, but their statements are too cumbersome. We shall go into cases
when the conditions imposed on the points in A and the set B are more or less
transparent. We agree to say that an open set B separates ak ∈ A from al ∈ A if
these points belong to different connected components of B.
Theorem 6.22. Let A = {ak }4k=1 be a set of distinct points in the Riemann sphere
C, and assume that a set B separates the points a1 and a2 from the points a3 and
a4 . Then
I(B, A)  |(a1 , a3 , a4 , a2 )(a1 , a4 , a3 , a2 )|4/3 , (6.34)
6.3. Möbius invariants 191

where (·, ·, ·, ·) is the cross ratio of a quadruple of points. Equality holds in (6.34)
in the case when B is bounded by the curves described by

|(z − a1 )(z − a2 )| = |(z − a3 )(z − a4 )|. (6.35)

Proof. We look at the quadratic differential


 2
2 1 1 1 1
Q(z)dz = + − − dz 2 .
z − a1 z − a2 z − a3 z − a4

Let γ be a trajectory lying in a strip domain of this differential. As a point z goes


along γ, the imaginary part of the function

z
F (z) = [−Q(z)]1/2 dz
z0

varies from −∞ to +∞ (or from +∞ to −∞). Hence the two end-points of γ cor-
respond to different points ak ; moreover, the points a1 and a2 (a3 and a4 ) cannot
be end-points of the same γ. Hence the trajectories of the quadratic differential
Q(z)dz 2 , with a possible exception of finitely many of them, do not lie in B. By
Corollary 6.4,
 4

r(B, ak ) exp {2 (gB (a1 , a2 ) + gB (a3 , a4 ))}
k=1
|a1 − a3 |2 |a1 − a4 |2 |a2 − a3 |2 |a2 − a4 |2
 .
|a1 − a2 |2 |a3 − a4 |2

Hence

(|a1 − a3 ||a1 − a4 ||a2 − a3 ||a2 − a4 |)4/3


I(B, A)  8/3
(|a1 − a2 ||a3 − a4 |)
 4/3  4/3
|a1 − a4 ||a3 − a2 | |a1 − a3 ||a2 − a4 |
= .
|a1 − a2 ||a3 − a4 | |a1 − a2 ||a3 − a4 |

Now the set B bounded by the curves (6.35) is open, and the points a1 and a2 lie
in different connected components of B from a3 and a4 . It is easy to see that
 
 (z − a1 )(z − a2 ) 
gB (z, a1 ) + gB (z, a2 ) = − log  ,
(z − a3 )(z − a4 ) 
 
 (z − a1 )(z − a2 ) 
gB (z, a3 ) + gB (z, a4 ) = log  .
(z − a3 )(z − a4 ) 
192 Chapter 6. Extremal Decomposition Problems

Therefore,
 
 (a1 − a3 )(a1 − a4 ) 

log r(B, a1 ) + gB (a1 , a2 ) = log  ,
a1 − a2 
 
 (a2 − a3 )(a2 − a4 ) 

log r(B, a2 ) + gB (a2 , a1 ) = log  ,
a1 − a2 

and similar equalities with a1 replaced a3 and a2 by a4 also hold. Consequently,


 
 (a1 − a3 )2 (a1 − a4 )2 (a2 − a3 )2 (a2 − a4 )2 
2
2πn M (B, A) = log   ,
(a1 − a2 )2 (a3 − a4 )2 

which yields equality in (6.34). 


Corollary 6.8. Let A = {ak }4k=1 be a set of points positioned anticlockwise on the
circle |z| = 1 and let B be a set separating a1 and a3 from a2 and a4 . Then

I(B, A)  4−4/3 , J(B, A)  1/4.

In either case equality is attained for ak = exp(i(θ + πk/2)), k = 1, . . . , 4 and for


B = {z ∈ C \ {0} : arg z 4 = 4θ + π}, where θ is an arbitrary real constant.
Proof. In view of symmetry and conformal invariance, we can assume that a1 = 1,
a2 = eiϕ , a3 = −1, and a4 = −i. Using Theorem 6.22 (with another numbering of
the points) we arrive at the inequality
 4/3
sin ϕ
I(B, A)  2−4/3  4−4/3
sin ϕ + 1

and the corresponding result on equality signs.


Next, similar arguments give us the second inequality:

I(B, A)(|a1 − a3 ||a1 − a2 ||a1 − a4 ||a3 − a2 ||a3 − a4 ||a2 − a4 |)2/3


J(B, A) =
|a1 − a3 ||a3 − a2 ||a2 − a4 ||a4 − a1 |
|a1 − a3 ||a1 − a4 ||a3 − a2 ||a2 − a4 | sin ϕ 1
 = 
|a1 − a2 |2 |a3 − a4 |2 2(sin ϕ + 1) 4

(here we have interchanged the indices of a2 and a3 ). 


In view of Schur’s inequality (5.1),

|ak − al |2/3  44/3 .


1k<l4

Hence the first inequality in Corollary 6.8 is an improvement over the result of
Theorem 6.11 for n = 4.
6.3. Möbius invariants 193

Theorem 6.23. 8 For arbitrary different points ak , k = 1, . . . , 5, two of which are


symmetric relative to the straight line or circle passing through the other three and
for any pairwise nonoverlapping domains Bk such that ak ∈ Bk ⊂ C, k = 1, . . . , 5,

I5 = I5 (a1 , . . . , a5 , B1 , . . . , B5 )  3−3/4 411/3 5−25/6 .

In addition, if the Bk are admissible domains, k = 1, . . . , n, then equality sign


holds here if and only if the points ak and domains Bk coincide with the points
0, 1, e2πi/3 , e−2πi/3 , ∞ and the circle domains Bk∗ of the quadratic differential

z 6 + 7z 3 + 1 2
Q(z)dz 2 = − dz
z 2 (z 3 − 1)2

up to a linear fractional transformation.


Proof. We can assume that a1 = 0, a5 = ∞, and the points a2 , a3 , and a4 lie on
the unit circle |z| = 1 and moreover, a2 = 1. Assume also that π/2  arg a3 <
arg a4  3π/2. Then we set ϕ1 = arg a3 , ϕ2 = arg a4 − ϕ1 , and ϕ3 = 2π − ϕ1 − ϕ2
and repeat the first part of the proof of Theorem 6.14 replacing the term (ϕk /π)2
with 2(ϕk /π)2 in the definition of the quadratic differential Rk (ζ)dζ 2 . Then we
obtain the inequality
3
I52  Ψ(ϕk /π), (6.36)
k=1

2 2 2
2
where Ψ(v) = sin(πv/2) v 2v +2 |1−v|−(1−v) (1+v)−(1+v) for 0 < v  3/2, Ψ(0) = 0,
and Ψ(1) = 1/8. Now we conclude from Corollary 6.3 that (6.36) also holds under
the following constraints:

a1 = 0, a2 = |a3 | = |a4 | = 1,
π/2  arg a3  π  arg a4  3π/2,
a5 = ∞, ak ∈ Bk ⊂ C, k = 1, . . . , 5, (6.37)
Bk ∩ Bl = ∅, k = l, k, l = 2, 3, 4, B5 = {z : 1/z ∈ B1 },
B1 ∩ Bk ∩ {z : |z| < 1} = B5 ∩ Bk ∩ {z : |z| > 1} = ∅, k = 2, 3, 4.

Thus the proof reduces to solving the extremal problem

X (u, v) := Ψ(u)Ψ(2 − u − v)Ψ(v) → max;


(6.38)
(u, v) ∈ Ω := {(u, v) : 1/2  u  3/2, 1/2  v  3/2, u + v  2}.

For convenience we set

Ω∗ = Ω ∪ {(u, v) : u  1, v  1, 1  u + v}
8 This result was independently proved by Kuz’mina ([Kuz2], no. 5, p. 26) and this author.
194 Chapter 6. Extremal Decomposition Problems

and look at the extremal problem

X (u, v) → max; (u, v) ∈ Ω∗ . (6.39)

By Weierstrass’ theorem, (6.39) has at least one solution, which either lies on the
boundary of Ω∗ or in its interior. Depending on this we split our further analysis
into two steps.
a) Analyzing X (u, v) on the boundary of Ω∗ .
On the part of the boundary of Ω∗ where u + v = 2 the function X (u, v)
vanishes. On the parts of the boundary of Ω∗ where either u = 1, or v = 1, or
u + v = 1 X (u, v) takes values in the range of the function

g(v) = Ψ(1)Ψ(v)Ψ(1 − v), 0  v  1.

The function g(v) takes equal values at points symmetric relative to the line
d2
v = 1/2. We see from simple calculations that dv 2 {log g(v)} < 0 on the inter-

val (0, 1/2]. Hence 1/2 is a maximum point of log g(v) on the interval (0, 1) and
at the same time a maximum point of X (u, v) on the corresponding parts of the
boundary of Ω∗ :
X (u, v)  g(1/2) = Ψ2 (1/2)Ψ(1).
Finally, it remains to look at the part of the boundary of Ω∗ where either u = 1/2
or v = 1/2. In this case it suffices to analyze the maximum/minimum properties
of the function

h(v) = Ψ(1/2)Ψ(v)Ψ(3/2 − v), 0  v  1/2.


3 2
3 {log h(v)} > 0 on (0, 1/2). Hence dv 2 {log h(v)}
d d
It can easily bee seen that dv
2 2
2 {log h(v)} < dv 2 {log h(0.35)} < 0 for v ∈
d d
is increasing on this interval and dv
(0, 0.35). Consequently, dv {log h(v)} is decreasing on (0, 0.35) and dv
d d
{logh(v)} 
dv {logh(0.35)} > 0 for v ∈ (0, 0.35]. On the remaining interval (0.35, 0.5) we have
d

 
d 2 π cos(πv/2)
{log h(v)} = − + [4v log v + 2(1 − v) log(1 − v)
dv v 2 sin(πv/2)
 −1    
3 3 3
+2(1 + v) log(1 + v)] − 2 −v −4 − v log −v
2 2 2
       
1 1 5 5 π cos(πv)
+2 − v log −v +2 − v log −v − .
2 2 2 2 2 1 + sin(πv)
d
Taking lower bounds for each separate term we obtain dv {log h(v)} > 0 for 0.35 <
v < 0.5. Hence log h(v) is an increasing function on (0, 1/2) and it follows that
h(v)  h(1/2) = Ψ2 (1/2)Ψ(1) for v ∈ [0, 1/2]. Summing the inequalities on the
boundary of Ω∗ we obtain

X (u, v)  Ψ2 (1/2)Ψ(1), (u, v) ∈ ∂Ω∗ . (6.40)


6.3. Möbius invariants 195

b) Analyzing X (u, v) in the interior of Ω∗ .


If (u∗ , v ∗ ) is a solution of (6.39) and (u∗ , v ∗ ) lies in the interior of Ω∗ , then the
point (u∗ , v ∗ , w∗ ) with w∗ = 2 − u∗ − v ∗ is a local conditional maximum point of
Ψ(u)Ψ(v)Ψ(w) under the constraint u + v + w = 2. From the necessary conditions
for an extremum we obtain
F (u∗ ) = F (v ∗ ) = F (w∗ ), (6.41)
where F (t) = d
dt {log Ψ(t)} , 0 < t < 2. Finding separate estimates for the second
derivative  π 3 cos(πt/2) 
d2 2
{F (t)} = 2 −
dt2 t3 (1 − t2 ) 2 sin3 (πt/2)
on (0, 2/3] and [2/3, 1) we see that this derivative is positive on (0, 1). Hence the
first derivative
d 2 t2 π 2 1
{F (t)} = − 2 + 2 log +
dt t 2
|1 − t | 2 sin2 (πt/2)
has a unique zero on (0, 1). It is also easy to see that the first derivative is positive
on [1, 2), so dF (t)/dt has a unique zero on (0, 2). Hence F (t) takes each value at
two points in (0, 2) at most and we can conclude that if (6.41) holds, then at least
two numbers among u∗ , v ∗ , w∗ coincide. Thus, the solution (u∗ , v ∗ ) of problem
(6.39) lies on one of the following three line intervals: u = v, 1/2  v  1;
u = 2 − 2v, 1/2  v  1; u = (2 − v)/2, 0  v  1. In each case (6.39) reduces to
finding the maximum of the function
f (v) := Ψ2 (v)Ψ(2 − 2v),
on [1/2, 1]. Now we look at the auxiliary function
1 d % &
F(v) = log(Ψ2 (v)Ψ(2 − 2v)) , v ∈ [1/2, 1).
2 dv
It is easy to see that sign df (v)/dv = sign F (v), v ∈ [1/2, 1). Making simple
calculations we obtain
d2
{F (v)} = π 3 (−2 cos4 (πv/2) − 2 cos2 (πv/2) + 1) sin−3 (πv)
dv 2
+ 4v −3 (1 − v 2 )−1 − 16(2 − 2v)−3 (1 − (2 − 2v)2 )−1 .
We can readily show that this function is negative on [1/2, 1), so F(v) is concave
on this interval. As F(2/3) = 0 and dF(2/3)/dv < 0, it follows that F (v) 
F (2/3) + (dF(2/3)/dv)(v − 2/3) < 0 for v > 2/3. Now if the interval (1/2, 2/3)
contains no zeros of F (v), then
     $  
1 2 2
max f (v) = max f ,f , f (1) = f = 3−3/2 422/3 5−25/3 , (6.42)
[1/2,1] 2 3 3
and 2/3 is the unique maximum point. If this interval contains a unique zero
of F (v), then f (v) has a local minimum at this points and (6.42) also holds.
196 Chapter 6. Extremal Decomposition Problems

It remains to observe that F(v) cannot have two zeros or more on (1/2, 2/3)
since it is concave on this interval. Comparing (6.40) and (6.42) we conclude that
(2/3, 2/3) ∈ Ω ⊂ Ω∗ is the unique solution of problems (6.39) and (6.38). From
relations (6.36) and (6.42) we obtain
I5  3−3/4 411/3 5−25/6 .
Direct calculations show that we have equality sign here in the case of the extremal
configuration indicated in the statement of Theorem 6.23. In a similar way to the
proof of Theorem 6.12 we show that this extremal configuration is unique.
It remains to discuss the case when
a1 = 0, a2 = 1, a3 = eiϕ1 , a4 = e−iϕ2 , 0 < ϕk < π/2, k = 1, 2, a5 = ∞.
We set
ζ = p1 (z) ≡ −iz, Im z > 0; ζ = p2 (z) ≡ iz, Im z < 0.
The pair {p1 , p2 } is an admissible family of functions for the separating transfor-
(1) (2)
mation of open sets with respect to the points a1 , a2 , and a5 . Let {Bk , Bk } be
the result of such a transformation of the domain Bk , k = 1, 2, 5. Then (4.13)
yields
(1) (1) (2) (2)
r2 (Bk , ak )  r(Bk , ak )r(Bk , ak ),
(j)
ak = pj (ak ), j = 1, 2, k = 1, 2, 5.
Now we set
(1) (1)
a3 = −ia3 , a4 = −ia3 ,
(2) (2)
a3 = ia4 , a4 = ia4 ,
(1) (1)
B3 = {ζ : iζ ∈ B3 }, B4 = {ζ : −iζ ∈ B3 },
(2) (2)
B3 = {ζ : iζ ∈ B4 }, B4 = {ζ : −iζ ∈ B4 }.
In view of the previous inequality,
1
5
(j) (j)
2 r(Bk , ak )
I52  (j) (j)
k=1
(j) (j) (j) (j)
.
j=1 |(a2 − a3 )(a3 − a4 )(a4 − a2 )|1/2
The linear fractional transformation
((−1)j+1 iζ + 1) tan(ϕj /2) + i((−1)j+1 iζ − 1)
z=
((−1)j+1 iζ + 1) tan(ϕj /2) − i((−1)j+1 iζ − 1)

(j) (j) (j) (j)
takes the system of sets a1 , . . . , a5 , B1 , . . . , B5 to a system of sets satisfying
(6.37), j = 1, 2. For such configurations we have (6.36) and it remains to repeat
the previous proof. 
6.3. Möbius invariants 197

Corollary 6.9. Let A = {ak }5k=1 be a set of distinct points in C, and let B be a
union of pairwise nonoverlapping domains Bk , ak ∈ Bk , k = 1, . . . , 5, such that
B4 and B5 lie to opposite sides of the straight line or circle passing through a1 , a2 ,
and a3 . Then

I(B, A) = I5 (a1 , . . . , a5 , B1 , . . . , B5 )  3−3/4 411/3 5−25/6 (6.43)

with the same equality cases as in Theorem 6.23.


Proof. We can assume that a1 , a2 , and a3 lie on the real line and that Ima4 > 0
and Ima5 < 0. Then in the notation of Lemma 6.1 we obtain

I 2 (B, A)  I(B + , {a1 , a2 , a3 , a4 , a4 })I(B − , {a1 , a2 , a3 , a5 , a5 }). (6.44)

In view of the assumptions of Corollary 6.9, the connected components of B +


containing a1 , a2 , a3 , a4 , and a4 are pairwise disjoint and the connected components
of B − containing a1 , a2 , a3 , a5 , and a5 are disjoint, too. It remains to use Theorem
6.23, which states that each factor on the right-hand side of (6.44) is no greater
than the right-hand side of (6.43) and provides the corresponding result on equality
signs. 
Theorem 6.24. Let A = {ak }6k=1 be a set of points positioned anticlockwise on the
circle |z| = 1, let a7 = a1 , and assume that a set B separates ak from ak+1 , k =
1, . . . , 6. Then
 6
2
J(B, A)  . (6.45)
3
Equality in (6.45) holds for A = {exp(i(θ + πk/3))}6k=1 and B = {z ∈ C \ {0} :
arg z 6 = 6θ + π}, where θ is an arbitrary real constant.
Proof. Since J(B, A) is a Möbius invariant we can assume that a1 = 1, a2 = eiϕ1 ,
a3 = e2πi/3 , a4 = e2πi/3+iϕ2 , a5 = e4πi/3 , and a6 = e4πi/3+iϕ3 , where the ϕj are
some quantities in the interval (0, 2π/3), j = 1, 2, 3. Then
6 3
ϕj π ϕj
|ak − ak+1 | = 26 sin sin − . (6.46)
j=1
2 3 2
k=1

We look at the condenser C(r) = C(r; B, ∂B, A, {1, . . . , 1}, {r, . . . , r}). By Theo-
rem 2.1
   2  2 
−1 2 1 1
cap C(r) = 12π − (12π) M (B, A) +o , r → 0.
log r log r log r

Let us introduce some notation:


Dj = {z : 2π(j − 1)/3 < arg z < 2πj/3};
Cj = (B ∩ Dj , {(∂B) ∩ Dj , E ∩ Dj }, {0, 1});
198 Chapter 6. Extremal Decomposition Problems

fj is the branch of the function w = z 3/2 taking Dj to the upper half-plane;


Cj∗ is the condenser mirror-symmetric relative to the real axis which ‘coin-
cides’ with fj (Cj ) in Im > 0;
Bj∗ is the union of fj (B ∩ Dj ) and its reflection in the real axis, j = 1, 2, 3.
Then by Theorem 4.8
3
1
cap C(r)  cap Cj∗ .
2 j=1

Using Theorem 2.1 again we obtain


 
−1
cap Cj∗ = 8π
log r
  2
3 ∗ ∗ 3ϕj i/2 −3ϕj i/2 1
+ 2π 4 log − 32πM (Bj , ∂Bj , {1, e , −1, e })
2 log r
 2 
1
+o , r → 0, j = 1, 2, 3.
log r

Summing these relations and letting r → 0 we arrive at the inequality

exp{2π62 M (B, A)}


 6 3 + , 12
2
 exp 2π42 M (Bj∗ , ∂Bj∗ , {1, e3ϕj i/2 , −1, e−3ϕj i/2 }) .
3 j=1

By the hypotheses of Theorem 6.24 the set Bj∗ separates ±1 from the points
e±3ϕj i/2 , j = 1, 2, 3. Using the second inequality in Corollary 6.8 we obtain
 6 3   2  2  12
2 2 1 3ϕj i/2   3ϕj i/2 
exp{2π6 M (B, A)}  1 − e  1 + e 
3 j=1 4
 6 3
2 3ϕj
= sin .
3 j=1 2

In view of (6.46),
 6 3 3ϕ
1 sin 2 j
J(B, A)  ϕj
π ϕj . (6.47)
3 j=1 sin 2 sin 3 − 2

The maximum value of the product on the right-hand side of (6.47) is the cube of
the maximum value of
sin 3x
f (x) :=
sin x sin (π/3 − x)
6.3. Möbius invariants 199


on (0, π/3). We set f to be equal to 2 3 at 0 and π/3 by continuity. Since f is
symmetric in x = π/6, it is sufficient to consider it on [0, π/6]. It is easy to see
that

(log f ) = 3 cot 3x − cot x + cot (π/3 − x)


and
−9 1 1
(log f ) = + + .
sin2 3x sin2 x sin2 (π/3 − x)

We claim that the second derivative is negative on [0, π/6]. This is equivalent
to the inequality

(24 − 16 sin2 x)(sin2 (π/3 − x) + sin2 x) > 9.

On [0, π/6] the first expression in parentheses is no smaller than 20, and the
second is  1/2. Thus (log f ) < 0, so that (log f ) is decreasing on [0, π/6]. Since
(log f ) = 0 at π/6, it follows that (log f )  0 on [0, π/6] and f is increasing on
[0, π/6]. It remains to observe that f (π/6) = 22 ; in combination with (6.47) this
yields (6.45). The equality cases are straightforward to verify. 

Exercises 6.3
(1) Show that the invariant I(B, A) is equal to J(B, A) times the product of
some powers of the absolute values of the cross ratios of points in the set A.
(2) (Fedorov [F]) Which quantity is larger,
n−2
max In (1, ωn−1 , . . . , ωn−1 , ∞, B1 , . . . , Bn )
or
max In (1, ωn , . . . , ωnn−1 , B1 , . . . , Bn )?

Here the maximum is taken over all possible pairwise nonoverlapping domains
Bk and ωk = exp(2πi/k).
(3) (Kuz’mina [Kuz2]) Prove that

I4  4−8/3 32

and describe all the equality cases.


Hint: use the proof of Theorem 6.12.
(4) For what real values of s and t is the quantity
exp{32πM (B, {a1, a2 , a3 , a4 })}
2 2 2
|(a1 − a2 )(a3 − a4 )| 3 −s |(a1 − a4 )(a2 − a3 )| 3 −t |(a1 − a3 )(a2 − a4 )| 3 +s+t
a Möbius invariant?
Chapter 7

Univalent Functions

If f is a univalent meromorphic function in the disc U = {z : |z| < 1}, f (0) = ∞,


then
|f  (0)| = r(f (U ), f (0))
by the definition of the conformal radius. For this reason many results in the
previous section can also be formulated for univalent functions sharing no values.
In this chapter we consider other applications of the capacity-based approach and
symmetrization methods to the theory of univalent functions.
We introduce our notation:
M is the class of univalent meromorphic functions in the disc U ;
M0 is the subclass of M consisting of functions f normalized by the condition
f (0) = 0;
M0 is the subclass of M0 consisting of functions normalized by the condition
f  (0) = 1;
R is the subclass of holomorphic functions in M;
R0 is the subclass of holomorphic functions in M0 ;
S is the class of holomorphic univalent functions f in U such that f (0) =
f  (0) − 1 = 0;
SB and SB0 are the subclasses of M and M0 , respectively, consisting of
functions such that |f (z)| < 1 for z ∈ U ;
Σ is the class of functions f (z) = z + a0 + a−1 /z + · · · that are meromorphic
and univalent in the domain |z| > 1;
M(R) is the class of meromorphic univalent functions w = f (z) in the annulus
K = K(R) = {z : 1 < |z| < R}, R < ∞, such that the set f (K) of
values of f (z) in K lies in |w| > 1 and f takes the circle |z| = 1 to
|w| = 1. Each function f in M(R) has an analytic extension into the
annulus 1/R < |z| < R by the Riemann–Schwarz symmetry principle.
We shall denote this extension by the same letter f .
M(R, M ) is the subclass of M(R) consisting of functions such that |f (z)| <
M for z ∈ K (R < M  ∞).

© Springer Basel 2014 201


V.N. Dubinin, Condenser Capacities and Symmetrization in Geometric Function
Theory, DOI 10.1007/978-3-0348-0843-9_7
202 Chapter 7. Univalent Functions

7.1 Classical inequalities, with some additions


and comments
In this section we consider elementary inequalities of the type of growth or distor-
tion theorems, which follow in a unified fashion from the solution of the Teichmüller
problem (Lemma 5.3). We also comment on and supplement some classical results.
For fixed numbers rk , k = 1, 2, 3, 4, −1  r1 < r2 < r3 < r4  1 we set
K(r1 , r2 , r3 , r4 ) = U \ {[−1, r1 ] ∪ [r2 , r3 ] ∪ [r4 , 1]}, and let (·, ·, ·, ·) denote the cross-
ratio of four points.
Theorem 7.1. Let f be a function in M, and let w2 and w3 be arbitrary differ-
ent points on the curve f ([r2 , r3 ]); let w1 and w4 be different points in another
component of the complement of the f -image of the ring domain K(r1 , r2 , r3 , r4 ).
Then

|(w3 , w1 , w2 , w4 )|  −(k(r3 ), k(r1 ), k(r2 ), k(r4 )), (7.1)


|(w3 , w2 , w1 , w4 )|  (k(r3 ), k(r2 ), k(r1 ), k(r4 )), (7.2)

where k(z) := z(1 + z)−2 is the Koebe function. In both (7.1) and (7.2) equality
holds only for the composites of k(z) with arbitrary linear fractional maps and for
the points wk equal to the images of the corresponding points rk or of the points
r5−k , k = 1, 2, 3, 4.
Proof. First we look at (7.1). Let g(w) denote the linear fractional map of Cw into
itself that takes the points w1 , w2 and w3 to ∞, −1, and 0, respectively. By the
invariance of the cross ratio

(w3 , w1 , w2 , w4 ) = (0, ∞, −1, g(w4 )) = −1/g(w4 ).

Using the conformal invariance of the modulus of a doubly connected domain and
Lemma 5.3 we obtain

M (K(r1 , r2 , r3 , r4 )) = M (g(f (K(r1 , r2 , r3 , r4 ))))  M (T− (P )),

where P = |g(w4 )| and T− (P ) = C \ {[−1, 0] ∪ [P, ∞]}. On the other hand, if


m(w) is the linear fractional map taking k(r1 ), k(r2 ), and k(r3 ) to ∞, −1, and 0,
respectively, then in a similar way we have

(k(r3 ), k(r1 ), k(r2 ), k(r4 )) = (0, ∞, −1, m(k(r4 ))) = −1/m(k(r4 )),
M (K(r1 , r2 , r3 , r4 )) = M (T− (Q)),

where Q = m(k(r4 )). By the monotonicity of the modulus, P  Q, which is


equivalent to (7.1). Now assume that equality holds in (7.1). Then necessarily
P = Q and the functions g ◦ f and m ◦ k map K(r1 , r2 , r3 , r4 ) onto the same ring
domain T− (P ). Since f is analytic in U , it follows that either g(f (r2 )) = −1 or
7.1. Classical inequalities, with some additions and comments 203

g(f (r2 )) = 0. In the first case g ◦ f = m ◦ k and therefore f ≡ g −1 ◦ m ◦ k, which


proves the assertion on equality for f (rk ) = wk , k = 1, 2, 3, 4. In the second case the
composite function h◦g ◦f , where h(w) = (P w+P )/(w−P ), maps K(r1 , r2 , r3 , r4 )
onto T− (P ) and takes r2 to −1, which again yields f ≡ g −1 ◦ h−1 ◦ m ◦ k with
f (r1 ) = w4 , f (r2 ) = w3 , f (r3 ) = w2 , and f (r4 ) = w1 .
Inequality (7.2) becomes an elementary consequence of (7.1) once we have
observed that (w3 , w2 , w1 , w4 ) = 1 − (w3 , w1 , w2 , w4 ). 

Note that other cross ratios of four points reduce to the cases considered in
(7.1) and (7.2). In what follows we limit ourselves to inequality (7.1) and write it
out as
 
 (w2 − w3 )(w4 − w1 )  (r3 − r2 )(1 − r2 r3 )(r4 − r1 )(1 − r1 r4 )
  (7.1 )
 (w2 − w1 )(w4 − w3 )   (r2 − r1 )(1 − r1 r2 )(r4 − r3 )(1 − r3 r4 ) .

Let f be a function in the class M and assume that we take its derivative at a
point distinct from a pole. Taking w2 = f (r2 ) and making the limiting procedure
as r3 → r2 , from (7.1 ) we deduce
 
 (w2 − w1 )(w4 − w2 )(r4 − r1 )(1 − r22 )(1 − r1 r4 ) 
 
|f (r2 )|   , (7.3)
(w4 − w1 )(r2 − r1 )(r4 − r2 )(1 − r1 r2 )(1 − r2 r4 ) 

−1  r1 < r2 < r4  1. In a similar way, letting r1 → r2 we obtain


 
 (w3 − w2 )(w4 − w2 )(r4 − r3 )(1 − r22 )(1 − r3 r4 ) 
 
|f (r2 )|   , (7.4)
(w4 − w3 )(r3 − r2 )(r4 − r2 )(1 − r2 r3 )(1 − r2 r4 ) 

where −1 < r2 < r3 < r4  1. Finally, taking the limit as r1 → r2 and r4 → r3


yields

(1 − r22 )(1 − r32 )


|f  (r2 )f  (r3 )|  |f (r2 ) − f (r3 )|2 , −1 < r2 < r3 < 1. (7.5)
(r3 − r2 )2 (1 − r2 r3 )2

In each case (7.3)–(7.5) equality is attained for an appropriate choice of the points
wk and the composite of the Koebe function and some linear fractional map.
If we set r2 = 0, w2 = 0, and wk = f (rk ) in (7.3) and (7.4), k = 1, 3, 4, then
we obtain the following estimates for functions f in M0 :
 
 f (r4 ) − f (r1 )  (r1 − r4 )(1 − r1 r4 )
 
 f (r4 )f (r1 )   r1 r4
, −1 < r1 < 0 < r4 < 1,
 
 f (r4 ) − f (r3 )  (r4 − r3 )(1 − r4 r3 )
 
 f (r4 )f (r3 )   r4 r3
, 0 < r3 < r4 < 1.

Here equality holds for the functions of the form g(z) = z/(1 − cz + z 2 ), where
c ∈ C is arbitrary. For r1 = −r4 the first inequality is equivalent to a result due
204 Chapter 7. Univalent Functions

to Grötzsch ([J], p. 149). Taking a linear fractional map of the disc |z| < 1 onto
itself we can easily see that (7.5) is equivalent to Fan’s inequality

|f  (z1 )f  (z2 )| (1 − |z1 |2 )(1 − |z2 |2 )


 ,
|f (z1 ) − f (z2 )|2 |(1 − z1 z 2 )(z1 − z2 )|2

where f ∈ M and the points zk satisfying |zk | < 1, k = 1, 2, are not poles of
f [Fan].
Now we set r1 = −1, r2 = 0, w1 = ∞, w2 = 0, and w4 = f (r4 ), 0 < r4 < 1,
in (7.3). Then we see that for each f in the class S,
r4
|f (r4 )|  .
(1 + r4 )2

Applying this estimate to e−iθ f (eiθ z) ∈ S and taking z ∈ U such that θ = arg z,
|z| = r4 , we arrive at the classical inequality

|z|
|f (z)|  , |z| < 1, (7.6)
(1 + |z|)2

where equality holds for the function k(z) and z ∈ [0, 1). In a similar way, from
(7.4) with r2 = 0, r4 = 1, w2 = 0, w3 = f (r3 ), and w4 = ∞, for functions in S we
obtain
|z|
|f (z)|  , |z| < 1, (7.7)
(1 − |z|)2
where equality holding for the Koebe function at every point z ∈ (−1, 0].
For r1 = 0, r4 = 1, w1 = 0, w2 = f (r2 ), and w4 = ∞ inequality (7.3) gives
the classical estimate for the absolute value of the logarithmic derivative in the
class R0 :   
 f (z)  1 + |z|
 
 f (z)   |z|(1 − |z|) , 0 < |z| < 1.

Equality sign holds for the function k(z) and z ∈ (−1, 0). In the same class, setting
r3 = 0, r4 = 1, w2 = f (r2 ), w3 = 0, and w4 = ∞ in (7.4) we obtain a sharp lower
bound for the absolute value of the logarithmic derivative:
  
 f (z)  1 − |z|
 
 f (z)   |z|(1 + |z|) , 0 < |z| < 1,

which is attained for z ∈ (0, 1) and the function k(z). In combination with (7.6)
and (7.7) the last two inequalities give us a sharp estimate for the absolute value
of the derivative of a function in the class S:
1 − |z| 1 + |z|
 |f  (z)|  .
(1 + |z|)3 (1 − |z|)3
7.1. Classical inequalities, with some additions and comments 205

Under the assumption that f  (r2 ) = 1 inequality (7.5) with r2 = 0 and


w2 = 0, yields Löwner’s estimate
|f  (z)| 1 − |z|2
 , |z| < 1,
|f (z)|2 |z|2
for f ∈ M0 (see also [J], pp. 142–144).
Let f ∈ M0 . Applying (7.4) to f (−z) and setting r4 = 0, w4 = f (−r4 ) = 0,
w3 = f (−r3 ), and w2 = f (−r2 ), after changing the notation (−r3 ) → r2 , (−r2 ) →
r1 we obtain the following result:
      
 f (r1 )   f (r1 ) f (r1 ) r2 (1 − r1 )2 
  
 f (r2 )   f (r2 ) f (r2 ) − 1 r1 (r1 − r2 )(1 − r1 r2 )  , (7.8)

where r1 and r2 are points distinct from the poles and 0 < r2 < r1 < 1. The
same inequality for −1 < r2 < 0 < r1 < 1 can be deduced by applying (7.4) with
r3 = 0, w3 = f (−r3 ) = 0, w4 = f (−r4 ), and w2 = f (−r2 ), to the function f (−z)
and then changing the notation (−r2 ) → r1 , (−r4 ) → r2 . Here equality is attained
for the composite of the Koebe function and h(w) = aw/(w − b), where a, b = 0
are arbitrary, provided that b = k(rj ), j = 1, 2.
An upper bound follows by applying (7.3) with r1 = 0, w1 = f (r1 ) = 0,
w2 = f (r2 ), and w4 = f (r4 ) to f ∈ M0 and by changing the notation r2 → r1 ,
r4 → r2 in the final relation, which yields
      
 f (r1 )   f (r1 ) f (r1 ) r2 (1 − r1 )2 
  − 1  , 0 < r1 < r2 < 1.
 f (r2 )   f (r2 ) f (r2 ) r1 (r2 − r1 )(1 − r1 r2 ) 
Here equality holds for the extremal functions in (7.8).
Now we give estimates for the ratio |f  (r1 )/f (r2 )| which are independent of
the values of the function f. To do this we need some auxiliary results, which are
of interest, too. If f ∈ R and −1 < r1 < r2 < 1, where f (r2 ) = 0, then
 
(1 + r1 )(1 − r2 )2  f  (r1 )  (1 − r1 )(1 + r2 )2

  .
(1 − r1 )|r1 − r2 |(1 − r1 r2 ) f (r1 ) − f (r2 )  (1 + r1 )|r1 − r2 |(1 − r1 r2 )
(7.9)
For −1 < r2 < r1 < 1 the inequality signs in (7.9) must be reversed. On the right-
hand side of (7.9) equality holds for the composite of the Koebe function and an
arbitrary linear function and on the left-hand side equality holds for the composite
of z(1 − z)−2 and a linear function. The right-hand inequality is a consequence of
(7.3) with r1 = −1 and w1 = ∞, and the left-hand one follows from (7.4) with
r4 = 1 and w4 = ∞. Now if f ∈ R0 and 0 < r2 < r1 < 1, then
 
(r1 − r2 )(1 − r1 r2 )  f (r1 ) − f (r2 )  (r1 − r2 )(1 − r1 r2 )
   . (7.10)
r2 (1 + r1 )2 f (r2 ) r2 (1 − r1 )2
The left-hand side of (7.10) follows from (7.1 ), where we must set r1 = −1, r2 = 0,
w1 = ∞, w2 = 0, w3 = f (r3 ), and w4 = f (r4 ) and introduce the corresponding
206 Chapter 7. Univalent Functions

changes in the notation. The right-hand side is also a consequence of (7.1 ) with
r1 = 0, r4 = 1, w4 = ∞, w1 = 0, w3 = f (r3 ), and w2 = f (r2 ). From (7.9) and
(7.10) we deduce the following estimates for functions in R0 :
  
(1 − r1 )(1 + r2 )2  f (r1 )  (1 + r1 )(1 − r2 )2

  , 0 < r2 < r1 < 1. (7.11)
r2 (1 + r1 )3 f (r2 )  r2 (1 − r1 )3

If −1 < r1 < r2 < 0, then the inequality signs in (7.11) must be reversed. On the
right-hand side of (7.11) equality is attained for the functions az(1 − z)−2 and on
the left-hand side equality holds for az(1 + z)−2 with an arbitrary a = 0.
Turning to the class of bounded univalent functions, the following Pick func-
tion is extremal in many results:

p(z; λ) = −k −1 (λk(−z)), 0 < λ  1,

where k(z) = z(1+z)−2 . Let f ∈ SB, let ϕ be a linear fractional map of the disc U
onto itself, and h be a univalent meromorphic function in U . Then the composite
h ◦ f ◦ ϕ is also univalent and meromorphic in U , and therefore Theorem 7.1 can
be used. In particular, setting ϕ(z) ≡ z and h(w) = w(1 + w)−2 we arrive at the
following result.
Theorem 7.2. Let f ∈ SB and assume that −1  r1 < r2 < r3 < r4  1. Let
w2 and w3 be arbitrary different points on the curve f ([r2 , r3 ]) and let w1 and
w4 be different points in the disc |w| < 1 which lie in another component of the
complement of the f -image of the ring domain U \ {[−1, r1 ] ∪ [r2 , r3 ] ∪ [r4 , 1]} (if
such points exist). Then
 
 (w3 − w2 )(1 − w3 w2 )(w4 − w1 )(1 − w4 w1 ) 
 
 (w2 − w1 )(1 − w2 w1 )(w4 − w3 )(1 − w4 w3 ) 
  (7.12)
 (r3 − r2 )(1 − r3 r2 )(r4 − r1 )(1 − r4 r1 ) 
  .
(r2 − r1 )(1 − r2 r1 )(r4 − r3 )(1 − r4 r3 ) 

In (7.12) equality holds for the Pick functions p(z; λ), 0 < λ  1, and points
wk = p(rk ; λ), k = 1, 2, 3, 4.
Proof. From inequality (7.1) we obtain

|h(w3 ), h(w1 ), h(w2 ), h(w4 )|  −(k(r3 ), k(r1 ), k(r2 ), k(r4 )),

with h as above. Writing out explicitly the cross ratios we arrive at (7.12). The
equality case is straightforward. 
Similarly to the derivation of (7.3)–(7.5) from (7.1 ), for functions f ∈ SB
we deduce the following relations from (7.12):
 
 (w2 − w1 )(1 − w2 w1 )(w4 − w2 )(1 − w4 w2 )(1 − r22 )(r4 − r1 )(1 − r4 r1 ) 
 
|f (r2 )|   ,
(1 − w22 )(w4 − w1 )(1 − w4 w1 )(r2 − r1 )(1 − r2 r1 )(r4 − r2 )(1 − r4 r2 ) 
7.1. Classical inequalities, with some additions and comments 207

where w2 = f (r2 ) and −1  r1 < r2 < r4  1;


 
 (w3 − w2 )(1 − w3 w2 )(w4 − w2 )(1 − w4 w2 )(1 − r22 )(r4 − r3 )(1 − r4 r3 ) 
 
|f (r2 )|   ,
(1 − w22 )(w4 − w3 )(1 − w4 w3 )(r3 − r2 )(1 − r3 r2 )(r4 − r2 )(1 − r4 r2 ) 
where again we take w2 = f (r2 ) and −1  r2 < r3 < r4  1; finally,
 
 (f (r2 ) − f (r3 ))2 (1 − f (r2 )f (r3 ))2 (1 − r22 )(1 − r32 ) 
|f  (r2 )f  (r3 )|   
(1 − f 2 (r2 ))(1 − f 2 (r3 ))(r2 − r3 )2 (1 − r2 r3 )2 
for −1 < r2 < r3 < 1. Up to a linear fractional mapping of the disc U onto itself
the last inequality is equivalent to Nehari’s bound in [Neh], p. 259. In particular,
if r2 = 0, f (0) = 0, and r3 = r, then
 2 
 f (r)(1 − r2 ) 
  
|f (0)f (r)|   .
(1 − f 2 (r))r2 
Assume in addition that the angular limit f (1) = 1 exists, so that the angular
derivative f  (1) is well defined [Pom]. Then letting r → 1, from the above we
obtain 
|f  (1)|  1/|f  (0)|. (7.13)
Now we set
1−z ζ−z
Bz (ζ) = · .
1 − z 1 − ζz
Applying (7.13) to the function f = Bw ◦ ψ ◦ Bz−1 we arrive at the estimate

1 (1 − |z|2 )|1 − ψ(z)|4


|ψ  (z)|  · , z ∈ U,
β 2 |1 − z|4 (1 − |ψ(z)|2 )
for any conformal mapping ψ of U into itself such that ψ(1) = 1 and ψ  (1) = β.
Many other special cases of inequality (7.12) are obtained in a similar way to how
we have deduced inequalities for unbounded functions.
Let G be a domain in Cz mirror-symmetric relative to the real axis L and
such that (Cz \ G) ∪ L is a connected set. Let w = l(z) denote a function mapping
G conformally and univalently into Cw so that the points on the real axis and the
boundary points are taken to the real axis. By the Riemann–Schwarz symmetry
principle the function w = l(z) takes points symmetric relative to the real axis to
points symmetric relative to the real axis.
The two-point estimate (7.5) for univalent functions in the disc can also be
extended to functions in G. Moreover, we have the following result.
Theorem 7.3. If f is a univalent meromorphic function in the domain G, then for
arbitrary different points z1 and z2 in G∩L distinct from the poles of the functions
f and l,
|f  (z1 )f  (z2 )| |l (z1 )l (z2 )|
 , (7.14)
|f (z1 ) − f (z2 )|2 |l(z1 ) − l(z2 )|2
208 Chapter 7. Univalent Functions

while a reverse inequality holds for arbitrary different points z1 and z2 in G which
are distinct from the poles of f and l and symmetric relative to the real axis L :
|f  (z1 )f  (z2 )| |l (z1 )l (z2 )|
2
 . (7.15)
|f (z1 ) − f (z2 )| |l(z1 ) − l(z2 )|2
In (7.14) and (7.15) equality is attained for the composite functions of the form
s ◦ l, where s is an arbitrary conformal automorphism of the Riemann sphere Cw .
Proof. The expression on the left-hand sides of (7.14) and (7.15) is a Möbius
invariant. Hence we have indeed equalities for the composites s ◦ l. Now we turn
to (7.14). Here we can assume that z1 < z2 . Let x > 0 be sufficiently small
so that the points z1 = z1 + x and z2 = z2 + x also lie in G and the set
G = G \ ([−∞, z1 ] ∪ [z1 , z2 ] ∪ [z2 , +∞]) is a doubly connected domain. We set
(f (z1 ) − f (z1 ))(f (z) − f (z2 )) (l(z1 ) − l(z1 ))(l(z) − l(z2 ))
f(z) = and 
l(z) = .
(f (z1 ) − f (z2 ))(f (z) − f (z1 )) (l(z1 ) − l(z2 ))(l(z) − l(z1 ))
Comparing the moduli of doubly connected domains with the help of Lemma 5.3,
as in the proof of Theorem 7.1, we obtain
M G = M f(G )  M G(f ), M G = M G(l),
  )|, +∞]} and G(l) = Cw \{[−1, 0]∪[|
where G(f ) = Cw \{[−1, 0]∪[|f(z l(z2 )|, +∞]}.
2
Hence
|f(z2 )|  |
l(z2 )|.
Dividing both sides by ( x)2 and passing to the limit as x → 0 we obtain (7.14).
To prove (7.15) we compare the reduced moduli of domains after applying
the mappings f and l to these domains. In view of (2.16), (2.18), and (2.17), we
obtain in succession
M (G, ∂G, {z1 , z2 }, {0, 1, −1}, {r, r})
= M (f (G), ∂f (G), {f (z1 ), f (z2 )}, {0, 1, −1}, {r|f (z1 )|, r|f  (z2 )|})
 M ({f (z1 ), f (z2 )}, {1, −1}, {r|f (z1 )|, r|f  (z2 )|})
1
= {− log |f  (z1 )f  (z2 )| + 2 log |f (z1 ) − f (z2 )|} .

In view of Theorem 2.9, for the function l we have equalities in this chain of
relations. 
We say that a finitely connected domain B ⊂ Cz and a closed subset γ of its
boundary ∂B are admissible if the boundary ∂B has no isolated points and the
nonempty set γ consists of finitely many nondegenerate connected components.
By analogy with the case of the unit disc and boundary circle, for admissible B
and γ we consider the conformal invariant
δ(ζ, z; B, γ) := exp{−gB (ζ, z, γ)},
7.1. Classical inequalities, with some additions and comments 209

where ζ, z ∈ B. Let w = f (z) be a function mapping the domain B conformally and


univalently onto an admissible domain G ⊂ Cw and let Γ ⊂ ∂G be an admissible
set. For each points z ∈ B we set
δ(f (ζ), f (z); G, Γ) r(B, γ, z)
|Df (z)| := lim = |f  (z)|.
ζ→z δ(ζ, z; B, γ) r(G, Γ, f (z))

This definition also extends to boundary points z ∈ (∂B) \ γ (f (z) ∈ (∂B) \ Γ) if


f  (z) is understood, for instance, as the angular derivative. In the special case of
B = {z : |z| < 1}, γ = ∂B and G = {w : |w| < 1}, Γ = ∂G we have
 
 ζ −z 
δ(ζ, z; B, γ) =  
1 − zζ 

and a similar formula holds for δ(f (ζ), f (z); G, Γ), so that

1 − |z|2
|Df (z)| = |f  (z)|.
1 − |f (z)|2
The next result covers many known two-point distortion theorems.
Theorem 7.4. Let B and G be admissible domains in the Riemann spheres Cz and
Cw , respectively, and let γ ⊂ ∂B and Γ ⊂ ∂G be admissible sets. Let f be a function
mapping B onto G conformally and univalently so that f (γ) ⊂ Γ, where f (γ) is
treated in the sense of boundary correspondence. Then for any points z1 , z2 ∈ B
and any real numbers t1 and t2
 2t1 t2
t21 t22 δ(z1 , z2 ; B, γ)
|Df (z1 )| |Df (z2 )|  .
δ(f (z1 ), f (z2 ); G, Γ)

On the other hand, if f is a conformal mapping such that f (B) ⊂ G and f ((∂B) \
γ) ⊂ (∂G) \ Γ, then
 2t1 t2
t21 t22 δ(z1 , z2 ; B, γ)
|Df (z1 )| |Df (z2 )|  .
δ(f (z1 ), f (z2 ); G, Γ)
for any points z1 , z2 ∈ B and real numbers t1 , t2 .
Proof. Set Z = {z1 , z2 }, = {t1 , t2 }, Ψ = {r, r}, W = {f (z1 ), f (z2 )}, and
 = {|f  (z1 )|r, |f  (z2 )r|}. Let f (γ) ⊂ Γ. In view of Theorems 2.11 and 2.10 (see
Ψ
also Theorems 1.15 and 1.16),

M (B, γ, Z, , Ψ) = M (f (B), f (γ), W, , Ψ)
  M (G, Γ, W, , Ψ).
 M (G, f (γ), W, , Ψ) 

Using the formula for the reduced modulus (2.20) we obtain the first inequality in
Theorem 7.4. The proof of the second is based on the same results. 
210 Chapter 7. Univalent Functions

It follows from the results in Appendix A4 that in the first inequality in


Theorem 7.4 equality sign holds only for f (B) = G, f (γ) = Γ, and the set G ∩
∂f (B) consisting of finitely many piecewise smooth curves such that the normal
derivative of the function t1 gG (w, f (z1 ), Γ) + t2 gG (w, f (z2 ), Γ) vanishes at their
interior points. In the second case equality holds if and only if f (B) = G and
t1 gG (w, f (z1 ), Γ) + t2 gG (w, f (z2 ), Γ) = 0 on G ∩ f (γ).
Note that for t1 = 1, t2 = 0 Theorem 7.4 turns into a result of the type of
majorization principle: if f (γ) ⊂ Γ, then for each point z in B

|Df (z)|  1,

while for f ((∂B) \ γ) ⊂ (∂G) \ Γ we obtain

|Df (z)|  1,

for each z ∈ B. In particular, for γ = ∂B and Γ = ∂G the second inequality


is known as the monotonicity of the inner radius of a domain. If γ = ∂B and
Γ = ∂G, then both inequalities extend to boundary points z ∈ (∂B) \ γ, provided
the derivative f  has the corresponding limit.
Let f be a function in M. Using various types of symmetrization from Chap-
ter 4, we can slightly refine two-point distortion theorems by taking account of
certain sets covered by f (U ). For example, let

μf (z1 , z2 ) = t−1 (1 − t)−1 dt,
σ

where σ = {t ∈ (0, 1) : f (z1 ) + t(f (z2 ) − f (z1 )) ∈ f (U )}, f (zk ) = ∞, k = 1, 2.


Theorem 7.5. If f is a function in the class M and points z1 , z2 in the disc U are
distinct from the poles of f , then

cosh4 (μf (z1 , z2 )/4)|f  (z1 )f  (z2 )| |1 − z1 z 2 |2


 . (7.16)
|f (z1 ) − f (z2 )|2 (1 − |z1 |2 )(1 − |z2 |2 )|z1 − z2 |2
Here equality is attained for f (z) = z + 1/z and points z1 = −z2 = −r, 0 < r < 1.
Proof. We look at the domain D := Cw \{w = f (z1 )+t(f (z2 )−f (z1 )) : (1−τ )/2 
t  (1 + τ )/2}, where τ = tanh(μf (z1 , z2 )/4). In the notations of Exercise 2.4(2)

H(f (U ), f (z1 ), f (z2 ))  H(D, f (z1 ), f (z2 )). (7.17)

In fact, let ϕ(w) = (w − f (z1 ))/(w − f (z2 )). Taking account of Exercise 2.4(2)
and formula (2.16) we conclude that (7.17) is equivalent to the following relation
between reduced moduli:

M (ϕ(f (U )), ∂ϕ(f (U )), {0, ∞},{0, −1, 1}, {r, r})
 M (ϕ(D), ∂ϕ(D), {0, ∞},{0, −1, 1}, {r, r}).
7.1. Classical inequalities, with some additions and comments 211

The last inequality follows from Theorem 4.4. Elementary calculations show that
for any z1 , z2 ∈ U distinct from the poles of f ,

(1 − |z1 |2 )(1 − |z2 |2 )|z1 − z2 |2 |f  (z1 )f  (z2 )|


H(f (U ), f (z1 ), f (z2 )) = log
|1 − z1 z 2 |2 |f (z1 ) − f (z2 )|2

and H(D, f (z1 ), f (z2 )) = 2 log(1 − τ 2 ).


Thus (7.16) is a consequence of (7.17). The case of equality sign is straight-
forward. 
Combining (7.16) with Fan’s inequality we obtain the sharp estimate

|1 − z1 z 2 |
μf (z1 , z2 )  4cosh−1  .
(1 − |z1 |2 )(1 − |z2 |2 )

Here equality sign is attained in the same case as in Theorem 7.5.


In addition to the well-known inequality (7.13) we present an estimate for the
absolute value of the derivative |f  (1)| from the other side. We cannot have such an
estimate without additional constraints on the behaviour of f in a neighbourhood
of z = 1.
Theorem 7.6. Let f be a function in the class SB which extends to the arc γ =
{z : |z| = 1, | arg z| < θ/2} by continuity so that |f (z)| = 1 on γ. Then
2
|f  (1)|  (1/|f  (0)|)1/(2 sin (θ/4))
.

Here equality is attained for f (z) ≡ cz, where |c| = 1.


Proof. We can assume that 0 < θ < 2π. By Corollary 6.3 the product
2
r(B1 , 0)[r(B2 , 1)]4 sin (θ/4)
r(B3 , ∞)

takes its maximum among all the triples of pairwise disjoint simply connected
domains B1 , B2 , B3 in C such that 0 ∈ B1 , 1 ∈ B2 and ∞ ∈ B3 at the circular
domains G1 , G2 , and G3 of the quadratic differential

(z − eiθ/2 )(z − e−iθ/2 ) 2


Q(z)dz 2 = − dz
z 2 (z − 1)2

(trajectories of this quadratic differential correspond to trajectories of the one in


in Exercise 6.1(11) under the map z  = i(1 − z)/(1 + z)). The domain G1 lies
in the disc |z| < 1 and is symmetric to G3 relative to |z| = 1, while the arc γ
partitions G2 into two parts symmetric relative to |z| = 1. We set B1 = f (G1 ),
B3 = {z : 1/z ∈ B1 }, and set B2 to be the union of f (G2 ∩ {z : |z|  1}) and the
reflection of this set in the circle |z| = 1. Since the circular domains are extremal,
2 2
r(B1 , 0)[r(B2 , 1)]2 sin (θ/4)
 r(G1 , 0)[r(G2 , 1)]2 sin (θ/4)
. (7.18)
212 Chapter 7. Univalent Functions

On the other hand

r(B1 , 0) = r(G1 , 0)|f  (0)|, r(B2 , 1)  r(G2 , 1)|f  (1)|. (7.19)

The first relation in (7.19) requires no comment. As regards the second, we observe
that, by the symmetry principle, the analytic extension of f across the arc γ
maps G2 onto the domain B2 . We complete the proof by combining (7.18) and
(7.19). 

In conclusion we consider an inequality for bounded functions which takes


account of boundary distortion.
Theorem 7.7. Let f be a function in the class SB0 which takes the arc γ = {z :
|z| = 1, | arg z| < α}, 0 < α < π, to some arc f (γ) with length 2β of the circle
|w| = 1 by means of boundary correspondence. Then for any r1 and r2 , −1  r1 <
0 < r2  1,

(|f (r2 )| + |f (r1 )|)(1 + |f (r2 )f (r1 )|) (r2 − r1 )(1 − r2 r1 )


 .
|f (r2 )|(|f (r1 )|2 + 2|f (r1 )| cos β + 1) r2 (r12 − 2r1 cos α + 1)

Equality holds for the Pick functions p(z; λ), 0 < λ  1, and arbitrary r1 , r2 , α
such that −1  r1 < 0 < r2  1 and 0 < α < π.

Proof. We introduce the condensers

C1 = (Cζ , {[0, k(r2 )], {ζ : Im ζ = 0, Re ζ  k(r1 ) or Re ζ  k(eiα )}}, {0, 1});


C2 = (Uz , {[0, r2 ], [−1, r1 ] ∪ [(∂Uz ) \ γ]}, {0, 1});
C3 = (f (Uz ), {f ([0, r2 ]), f ([−1, r1 ]) ∪ [(∂f (Uz )) \ f (γ)]}, {0, 1});
C4 = (Uw , {[0, |f (r2 )|], [−1, −|f (r1 )|] ∪ [(∂Uw ) \ {w : |w| = 1,
| arg w| < β}]}, {0, 1});
C5 = (Cω , {[0, k(|f (r2 )|)], {ω : Im ω = 0, Re ω  k(−|f (r1 )|) or
Re ω  k(eiβ )}}, {0, 1}).

Since the capacity is a conformal invariant and the condensers C1 and C5 are
mirror-symmetric relative to the real axis, it follows that

cap C1 = cap C2 = cap C3 and cap C4 = cap C5 .

Since the capacity is monotonic, from Theorem 4.2 we obtain

cap C3  cap Cr C3  cap C4 .

Thus
cap C1  cap C5 .
7.1. Classical inequalities, with some additions and comments 213

On the other hand we can compare the capacities of these condensers using the
linear fractional maps

ζ(k(r2 ) − k(r1 ))
ξ=
(ζ − k(r1 ))k(r2 )
and
ω(k(|f (r2 )|) − k(−|f (r1 )|))
η= .
(ω − k(−|f (r1 )|))k(|f (r2 )|)

Hence
k(eiα )(k(r2 ) − k(r1 )) k(eiβ )(k(|f (r2 )|) − k(−|f (r1 )|))
 .
(k(eiα ) − k(r1 ))k(r2 ) (k(eiβ ) − k(−|f (r1 )|))k(|f (r2 )|)
Direct calculations yield the inequality in Theorem 7.7. The equality case is
straightforward. 

Corollary 7.1. If f (z) = c1 z + c2 z 2 + · · · is a function satisfying the assumptions


of Theorem 7.7, then
c2
Re  2(cos α − |c1 | cos β)
c1
with equality for the Pick functions.

Proof. It is sufficient to use Theorem 7.7 for −r1 = r2 = r and take the limit as
r → 0. 

Note that in terms of functions in SB0 the classical Pick inequality has the
following form:
|c2 |  2(1 − |c1 |)|c1 |.

Exercises 7.1
(1) Let d(z1 , z2 ) = tanh−1 |(z1 −z2 )/(1−z 1 z2 )| be the hyperbolic distance between
the points z1 and z2 in the disc U . Using Theorem 7.3 show that if f is a
function in M and points zk ∈ U, k = 1, 2, are distinct from poles of f , then
the following sharp inequalities hold:

1 
|f (z1 ) − f (z2 )|  sinh(2d(z1 , z2 )) |f  (z1 )f  (z2 )|(1 − |z1 |2 )(1 − |z2 |2 ),
2 
|f (z1 ) − f (z2 )|  tanh d(z1 , z2 ) |f  (z1 )f  (z2 )|(1 − |z1 |2 )(1 − |z2 |2 ).

Verify that the first relation is equivalent to Fan’s inequality. Give exam-
ples of functions f and points z1 , z2 for which equality signs hold in these
inequalities.
214 Chapter 7. Univalent Functions

(2) Let f be a function in the class SB that takes a closed arc γ of the circle
|z| = 1 with length 2α to an arc f (γ) = {w : |w| = 1, | arg w|  β}, 0 < β <
π by means of boundary correspondence and assume that the images f (zk )
of two points zk ∈ U, k = 1, 2, lie on the real axis. Prove the inequality

(f 2 (z2 ) − 2f (z2 ) cos β + 1)(1 − f (z1 ))2


(f 2 (z1 ) − 2f (z1 ) cos β + 1)(1 − f (z2 ))2
(t2 + 2t2 cos α + 1)(1 + t1 )2
 22 ,
(t1 + 2t1 cos α + 1)(1 + t2 )2

where t1 = max(|z1 |, |z2 |) and t2 = min(|z1 |, |z2 |). Equality holds here for the
Pick functions p(z; λ), 0 < λ  1 and for −1 < z1 < z2  0.
(3) The Grötzsch function w = G(z) ≡ G(z; R) maps the annulus K(R) = {z :
1 < |z| < R}, R < ∞ conformally and univalently onto the exterior of the
disc |w| > 1 cut from some point P (R) to ∞ along the real positive half-axis,
where G(R; R) = P (R). Verify that
  
i
G(z; R) = τ sn2 log(zR) + 1 K(τ ); τ ,
π

where τ = τ (R) = 1/P (R) is the solution of the equation


π 
log R = K 1 − τ 2 /K(τ ),
2
K(τ ) is the complete elliptic integral of the first kind with modulus τ and
sn(·; τ ) is the Jacobi elliptic function. What is the image of K(R) under the
Teichmüller function

T (z) ≡ T (z; R) := (G(z; R) + 1/G(z; R))/4 − 1/2 ?

(4) Let f be a meromorphic univalent function in the annulus K(R) and assume
that 1  |rk |  R, k = 1, 2, 3, 4, r1 < r2 < r3 < r4 , r1 r2 > 0, and r3 r4 > 0.
Assume that the f -image of the ring domain K(r1 , r2 , r3 , r4 ) := K(R) \
{[−R, r1 ] ∪ [r2 , r3 ] ∪ [r4 , R]} separates some points w2 and w3 from points w1
and w4 , where w2 and w3 lie in the connected component of the complement
of f (K(r1 , r2 , r3 , r4 )) containing the limit values of f (z) as z → r2 , z → r3 .
Following the proof of Theorem 7.1 prove the inequality

|(w3 , w1 , w2 , w4 )|  |(T (r3 ), T (r1 ), T (r2 ), T (r4 ))|. (7.20)

Show that equality holds in (7.20) only for the composite functions f =
ϕ(T (eiθ z)) and the points wk that are the f -images of the points e−iθ rk or
of the points e−iθ r5−k , k = 1, 2, 3, 4. Here ϕ is an arbitrary linear fractional
mapping and θ is an arbitrary real number in the case when r1 = −R,
r2 = −1, r3 = 1, and r4 = R; time; otherwise θ = 0.
7.1. Classical inequalities, with some additions and comments 215

(5) Prove the following result by applying (7.20) to the function f (z/R) in the
annulus K(R2 ). Let f be a function in M(R) and assume that 1/R  |rk | 
R, k = 1, 2, 3, 4, r1 < r2 < r3 < r4 , r1 r2 > 0, and r3 r4 > 0,. Assume that the
f -image of the ring domain {1/R < |z| < R} \ {[−R, r1 ] ∪ [r2 , r3 ] ∪ [r4 , R]}
separates some points w2 and w3 from points w1 and w4 . Then

|(w3 , w1 , w2 , w4 )|  |(G(r3 ), G(r1 ), G(r2 ), G(r4 ))|. (7.21)

In (7.21) equality holds for the composite functions ϕ ◦ G, where ϕ is a


conformal automorphism of the exterior |w| > 1 of the unit disc, and for
the points wk that are the images of the points rk or of the points r5−k ,
k = 1, 2, 3, 4, under this map.
(6) For holomorphic functions f in the class M(R) prove the inequality

|G(|z|)|  |f (z)|  |G(−|z|)|, z ∈ K(R).

(7) Using inequality (7.21) show that for each holomorphic function f ∈ M(R)
      
 G (|z|)   f  (z)   G (−|z|) 
   
 G(|z|)   f (z)   G(−|z|)  ,
|G (|z|)|  |f  (z)|  |G (−|z|)|, z ∈ K(R).

(8) Let Fρ be the class of holomorphic univalent functions f in the annulus


Rρ := {z : ρ < |z| < 1} such that f (Rρ ) ⊂ {w : 0 < |w| < 1} and
f ({z : |z| = 1}) = {w : |w| = 1}. In view of the analytic extension of
f across |z| = 1 by the symmetry principle we see that in fact the class
Fρ is the subclass of holomorphic functions in M(1/ρ). Prove the following
estimates for the absolute values of f in Fρ :

|G(−|z|; 1/ρ)|  |f (z)|  |G(|z|; 1/ρ)|, z ∈ Rρ .

(9) For functions f in Fρ deduce Duren’s inequality

|f  (z)|  |G (|z|; 1/ρ)|, z ∈ Rρ .

(10) Verify that


|f  (z)|  |G (−|z|; 1/ρ)|

for functions f in Fρ and points z ∈ Rρ such that |G(−|z|; 1/ρ)|  2 − 1.
(11) Using polarization of condensers and then using the scheme of the proof of
Theorem 7.1 prove the following result. Let f be a univalent meromorphic
function in the annulus K(R) and let z1 , z2 be different points in this annulus.
Let w3 and w4 be points in the connected component of the complement of
f (K(R)) which contains the limit values of f (z) as z → 1. Then

|(f (z1 ), w3 , f (z2 ), w4 )|  |(T (−|z1 |), T (1), T (|z2 |), T (−1))|.
216 Chapter 7. Univalent Functions

If w3 and w4 lie in the other connected component of the complement of


f (K(R)), then

|(f (z1 ), w3 , f (z2 ), w4 )|  |(T (−|z1 |), T (R), T (|z2 |), T (−R))|.

In either case equality holds if and only if the points z1 and z2 lie on the
same straight line through the origin, to opposite sides of the origin, and
the points wj are w3 = f (z2 /|z2 |) and w4 = f (z1 /|z1 |) in the first case
and w3 = f (Rz2 /|z2 |) and w4 = f (Rz1 /|z1 |) in the second, while f is the
composite of the rotation ζ = −z|z1 |/z1 , the function T , and an arbitrary
linear fractional automorphism of the Riemann sphere.
(12) Let f be a function in the class SB0 which takes a set α formed by finitely
many open arcs of the circle |z| = 1 to the circle |w| = 1 by means of
boundary correspondence. Deduce from Theorem 7.4 that

|f  (0)| cap{w : |w| = 1, w ∈ f (α)}  cap{z : |z| = 1, z ∈ α}.

(13) ([Pom], p. 217]) Using Theorem 7.4 again, deduce that if f ∈ SB0 and f
takes a closed subset γ of the circle |z| = 1 to a closed subset Γ of the circle
|w| = 1 by means of boundary correspondence, then

|f  (0)| cap Γ  cap γ.

(14) Let f ∈ SB0 be a function such that f (γ) ⊂ Γ, where γ is a closed subset of
the circle |z| = 1 distinct from the whole circle and Γ is a similar subset of
|w| = 1. For arbitrary points z1 , z2 in the disc |z| < 1 and arbitrary real t1
and t2 prove the inequality
2  t2
r2 (Cz \ γ, zk )(1 − |f (zk )|2 )|f  (zk )| k
k=1
r2 (Cw \ Γ, f (zk ))(1 − |zk |2 )
⎧ A B ⎫2t1 t2
⎨ exp gC \Γ (f (z1 ), f (z2 )) + gC \Γ (f (z1 ), 1/f (z2 )) ⎬
w w
 A B .
⎩ exp gCz \γ (z1 , z2 ) + gCz \γ (z1 , 1/z2 ) ⎭

(15) Let f be a holomorphic univalent function in the annulus Ω = {z : ρ <


|z| < 1} such that the image f (Ω) lies in the disc |w| < 1 and |f (z)| = 1 for
|z| = 1. Let β be a closed subset of the circle |z| = 1 consisting of finitely
many nondegenerate arcs. Prove that for any points z1 and z2 in the set
Ω ∪ {z : |z| = 1, z ∈ β},
2
r(Ω, γ, zk )|f  (zk )|
 exp{2 [gΩ (z1 , z2 , γ) − gUw (f (z1 ), f (z2 ), f (β))]},
r(Uw , f (β), f (zk ))
k=1

where γ = β ∪ {z : |z| = ρ}, Uw = {w : |w| < 1}.


7.2. Covering theorems for radial segments 217

7.2 Covering theorems for radial segments


Let f be a function in the class M0 . Setting G = C, m  2, n = 1, a1 = 0,
r(G1 , 0) = 1, and B = f (U ) in Corollary 6.3 we easily conclude that f (U ) contains
at least one orthogonal trajectory of the quadratic differential Q(z)dz 2 with an
end-point at a simple pole. In what follows we consider more delicate questions,
relating to covering straight-line segments with end-point at the origin. For f ∈ M0
let Λf (θ) be the distance from the origin to the closest boundary point of the set
f (U ) which lies on the ray arg w = θ, 0 < |w| < ∞. If no such point exists for
some θ, then we set Λf (θ) = +∞. In 1922, Szegő set the problem of finding the
lower bounds for the mean values of the Λf (θ + 2πk/n), k = 1, . . . , n, for functions
f in S, arbitrary θ, and fixed n  2. If n = 1, then we need not take the mean
value and the answer is given by the classical Koebe–Bieberbach theorem:

1
Λf (θ)  , f ∈ S. (7.22)
4

Equality in (7.22) holds only for the function z(1+e−iθ z)−2 which maps the disc U
conformally and univalently onto the plane cut along the ray arg w = θ, |w|  1/4.
Theorem 7.8 ([S1]). Let f be a function in the class M0 and for some real θ let
{w = r exp(i(θ + π(k − 1))), 0  r  ∞} ⊂ f (U ), k = 1, 2. Then
 −1 C
1 −1 −1 1
(Λ (θ) + Λf (θ + π))  . (7.23)
2 f 4

Equality holds in (7.23) if and only if f (z) ≡ z((e−iθ z)2 + c(e−iθ z) + 1)−1 , where
c is an arbitrary real coefficient with |c|  2.

Proof. We can assume that θ = 0. Then the points wk = Λf (π(k − 1)) exp(iπ(k −
1)), k = 1, 2, do not lie in f (U ). Let B be the image of the disc U under 1/f . Since
B is a simply connected domain, its complement E = Cw \ B is a continuum. By
(2.1)
cap E = 1/r(B, ∞) = 1.
The points 1/wk , k = 1, 2, belong to E. They lie on the real axis, to opposite sides
of the origin. By Theorem 3.2(a) or Corollary 4.2,

1 = cap E  |1/w1 − 1/w2 |/4 = (1/|w1 | + 1/|w2 |)/4,

which coincides with (7.23). If equality holds, E must be the interval of the real
axis joining 1/w1 to 1/w2 . Hence

1/f (z) ≡ z + 1/z + c, z ∈ U,

where c is a real number with |c|  2. 


218 Chapter 7. Univalent Functions

Theorem 7.8 refines inequality (7.22). Note also that Szegő’s result can be
treated in terms of inner radii as follows. If B is an open subset of C containing
the origin and its complement contains a continuum connecting points a < 0 and
b > 0, then
r(B, 0)  4/(1/b − 1/a). (7.24)
Equality holds in (7.24) only for B = C \ [{z : Re z  a, Im z = 0} ∪ {z : Re z 
b, Im z = 0}].
Theorem 7.9. For each function f in the class M0 and any real θ
>
? n   C
? 2πk n 1
@n
Λf θ +  , (7.25)
n 4
k=1

with a possible exception of the case when n = 1 and the domain f (U ) covers
the whole of the ray {−ρeiθ : 0  ρ  ∞}. Equality in (7.25) holds only for
the function w = z[1 + (e−iθ z)n ]−2/n , which maps the disc U conformally
 and
univalently onto the plane Cw cut along the rays arg wn = nθ, |w|  n 1/4.
Proof. We can assume √that θ = 0 and n  2. We set Dk = {w : | arg w − 2πk/n| <
π/n}, ζ = pk (w) := wn , and (w ∈ Dk , pk (e2πik/n ) = 1), k = 1, . . . , n, and let
{B (k) }nk=1 be the sets obtained by the separating transformation of the domain
f (U ) with respect to the family of functions {pk }nk=1 . By Theorem 4.10
n
2
1 = r(f (U ), 0)  [r(B (k) , 0)]2/n . (7.26)
k=1

The complement of f (U ) is connected and contains the points Λf (2πk/n)e2πik/n .


Hence the complement of each set B (k) contains a continuum connecting the points
±Λf (2πk/n). In view of (7.24), r(B (k) , 0)  2Λf (2πk/n), k = 1, . . . , n. In
n/2 n/2

combination with (7.26) this yields inequality (7.25).


Now if equality holds in (7.25) then it also holds in the preceding relations. In
particular, equality in (7.24) ensures that the sets B (k) are planes cut along the rays
arg ζ 2 = 0, |ζ|  Λf (2πk/n), k = 1, . . . , n. Taking account of equality in (7.26)
n/2

we conclude that all the Λf (2πk/n), k= 1, . . . , n, coincide and f (U ) is the plane
cut along the rays arg wn = 0, |w|  n 1/4. Hence f (z) = z(1 + z n )−2/n . 
Corollary 7.2 ([Gol], Ch. IV, § 6, Theorem 2). If f ∈ S, then the f -image of
the disc U covers N straight-line segments with end-point w = 0 which form equal
angles (= 2π/N ) with one another and whose total length is arbitrarily close to N .
Proof. Assume the converse. Then there exists ε > 0 such that for each θ, 0  θ 
2π, we have
 
1
N

1−ε> Λf θ + k .
N N
k=1
7.2. Covering theorems for radial segments 219

Now we set in turns θ = 2π/mN, 2 ·2π/mN, . . . , m·2π/mN = 2π/N . Multiplying


the resulting relations and using the well-known inequality between the quadratic
and geometric means we obtain
  
1
m N
2π 2π
(1 − ε) >
m
Λf j+ k
j=1
N mN N
k=1
> >
m ??N   ?mN
?  
@ 2π 2π 2π
 N
Λf j+ k = @ N
Λf k .
j=1
mN N mN
k=1 k=1

Thus, >
?mN  
? 2π
1−ε @
mN
Λf k ,
mN
k=1

which contradicts (7.25) (with θ = 0 and n = mN ) for sufficiently large m. 


Corollary 7.3 ([Gol, Ch. IV, § 6, Theorem 3]). If f ∈ S, then the star of the f -image
of the disc U has area at least π.
Proof. Let f be a function in the class S. Then the function g(z) = f (ρz)/ρ,
0 < ρ < 1, also belongs to S and takes U to a domain bounded by a closed
analytic Jordan curve. We use inequality (7.25) (with θ = 0) and apply the well-
known inequality between the geometric and arithmetic means to w = g(z), which
gives us > >
C ? n   ?  
1 ? 2π ?1 n 2π
 @ k @
n n
Λg Λ2g k .
4 n n n
k=1 k=1

Hence
 2  
1 2 2π
n
1 n 2π
π  Λg k .
4 2 n n
k=1
Passing to the limit as n → ∞ we obtain the required result for the function
w = g(z). We complete the proof by the limiting procedure as ρ → 1. 
Theorem 7.10. For each function f in S such that |f (z)| < M in U , (1 < M < ∞)
and for each real number θ
n
Λf (θ + 2πk/n) 1
2/n
 .
(1 + (Λf (θ + 2πk/n)/M )n ) 4
k=1

Equality holds only for the function in S mapping the disc U onto the domain
|w| < M cut along the line segments
 −2 $

w : wn = ρeiθn , 1 + 1 − M −n  ρ  Mn .
220 Chapter 7. Univalent Functions

Proof. On the level of ideas, this repeats the proof of the previous theorem, where
the parameters of the separating transformation of the domain f (U ) must be
replaced with

Dk = {w : | arg w − 2πk/n| < π/n, 0 < |w| < M },



ζ = pk (w) = wn [1 + (w/M )n ]−2 (Re ζ > 0) , k = 1, . . . , n. 

Corollary 7.4. For any θ the following sharp estimate holds for functions f in the
class S such that |f (z)| < M for z ∈ U :
>
? n −2/n
? √
@n
Λf (θ + 2πk/n)  1 + 1 − M −n .
k=1

Equality holds only for the function in Theorem 7.10.



Proof. Set x∗ = log(1+ 1 − M −n )−2/n and xk = log Λf (θ+2πk/n), k = 1, . . . , n.
The function
n A B
F (x) = − log ex (1 + (ex /M )n )−2/n
2
is decreasing and convex on (−∞, log M ]. Taking Theorem 7.10 into account we
obtain
 n 
A B 1
n
1
∗ n −n/2
F (x ) = log (1 + r )r  F (xk )  F xk .
n n
k=1 k=1

Hence
1
n
xk  x∗ . 
n
k=1

Next we set
' (−1/n 
f (z; θ, c, n) = z 1 + (c−n − 2)(e−iθ z)n + (e−iθ z)2n , n
1/4  c  ∞.

It is easy to see that the function w = f (z; θ, c, n) belongs to M0 and maps the disc
U onto the plane w cut along the rays arg wn = θn, |w|  c and arg wn = π + θn,
|w|  (4 − c−n )−1/n .
Theorem 7.11. If f ∈ M0 , then for any θ and n  2,
 −2  n −2
1 n/2 1 n/2
n
2π π 2π
Λf (θ + k) + Λf (θ + + k)  4. (7.27)
n n n n n
k=1 k=1

Here equality sign holds only forthe functions f (z) = f (z; θ, c, n), where c is an
arbitrary real number such that n 1/4  c  ∞.
7.2. Covering theorems for radial segments 221

Proof. We set aj = Λf (θ + πj/n), j = 1, . . . , 2n. If aj = ∞ for some j, then (7.27)


follows from (7.25). In fact, for odd j we have
 −2  −2  −2
1 n/2 1 n/2 1 n/2
n n n n
a2k + a2k−1 = a2k  a−1
2k  4.
n n n
k=1 k=1 k=1 k=1


Here equality is attained only for f (z; θ, n 1/4, n). The case of even j is similar.
Now assume that aj < ∞, j = 1, . . . , 2n. Then using (7.25) again we see that

 2
1 n/2 
n n
n
c= a2k > 1/4.
n
k=1

Let Dj denote the image of the sector {z : | arg z − θ − πj/n| < αj π/2, |z|  1}
j = 1, . . . , 2n, under the map w = f (z; θ, c, n), where αj = 1/n + (−1)j+1 (1/n) −
2α), απn = cos−1 (1 − c−n /2), 0 < α < 1/n (note that f (ei(θ+απ) ; θ, c, n) = ∞)
j = 1, . . . , 2n. Let ζ = pj (w),, be the functions defined by
 $  1 
' ( 1 ' (− 1 −1
ζ= (−1)j (e−iθ z)n αj n + (−1)j (e−iθ z)n αj n 1 αj n = 1 ,

w = f (z; θ, c, n), w ∈ Dj , j = 1, . . . , 2n.

Making the separating transformation of the domain f (U ) with respect to the


family of functions {pj }2n
j=1 , similarly to the proof of Theorem 7.9 we arrive at the
inequality

 α22
π α1
n 2 n

2p1 (a2k−1 ei(θ+ n ) ) 2p2 (a2k e i(θ+ n )
)  1. (7.28)
k=1 k=1

It remains to verify the relations


 n
 
π(2−m)
 1/n
i θ+ n
p2−m a2k−m e
k=1
⎛ 2/n  ⎞ (7.29)
1 n/2
n π(2−m)
i θ+
 p2−m ⎝ a2k−m e n ⎠,
n
k=1

m = 0, 1. It fact, it follows from (7.29) with m = 0 that

n  α22  α22 n
2π 2π
i(θ+ n ) i(θ+ n )
2p2 (a2k e )  2p2 (ce ) =1
k=1
222 Chapter 7. Univalent Functions

and relations (7.28) and (7.29) with m = 1 yield


⎛ ⎛ 2/n ⎞⎞α21 n
n 2

n
π α1 1 π
2p1 (a2k−1 ei(θ+ n ) )  ⎝2p1 ⎝ ei(θ+ n ) ⎠⎠
n/2
1 a2k−1 .
n
k=1 k=1

Hence  2/n
1 n/2
n
π
−i(θ+ n )
a2k−1  p−1
1 (1/2)e = (4 − c−n )−1/n ,
n
k=1
which is equivalent to (7.27). Equality sign holds here only when we have equality
in (7.28), that is, when
 A B−1 
−1 −iθ 1/α2n iθ 2/α2n
f (z) ≡ p2n (e z) 1 + (e z) = f (z; θ, c, n), z ∈ U.

Note that for n 1/4  c  ∞ we have equality sign in (7.27) for the functions
f (z; θ, c, n).
Now we prove (7.29). Let m = 0 (for m = 1 the proof is similar). The
function ω  = X(ω) ≡ − log p2 (ω 2/n ei(θ+2π/n) ) maps the domain B = {ω : ω =
wn/2 , wei(θ+2π/n) ∈ D2 } (1n/2 = 1) conformally and univalently onto the strip
|Im ω  | < π/2. The equation of the ‘upper’ boundary curve of B can be written as
A Bn/2
ω = ω(t) = e−i(θ+2π/n) f (t1/n ei(θ+2π/n+απ) ; θ, c, n) , 0  t  1.

We set ϕ = arg ω, and ψ = απn. After simple transformations we obtain


cos 2ϕ = (Re ω 2 )/|ω 2 | = (1 − t)(t2 − 2t cos 2ψ + 1)−1/2 cos ψ. Hence ((Im ω)2 )t =
(|ω 2 | sin2 ϕ)t > 0. Thus Im ω(t) is an increasing function on (0, 1), so that B
contains its own right translations. We fix some points ω1 and ω2 , 0 < ω1 <
ω2 < ∞. Then the functions ω = Ψi (z) ≡ X −1 (log((1 − z)/(1 + z)) + X(ωi )) − ωi ,
i = 1, 2, map the disc U conformally and univalently onto some domains Bi , where
B1 ⊂ B2 and B1 = B2 by the above. By the Lindelöf principle Ψ1 (0) < Ψ2 (0).
Hence X  (ω1 ) < X  (ω2 ). This means that X(ω) is a convex function on (0, +∞)
and inequality (7.29) with m = 0 follows from the well-known properties of convex
functions. 
We can regard the next inequality as a supplement to (7.25) and (7.27).
 n −2
 n/2
Theorem 7.12. If f ∈ M0 and n1 Λf (θ + 2πk/n)  2 for some θ and
k=1
n  2, then
 n −2  −1
1 n/2
n
2π π 2π
Λf (θ + k) + Λf (θ + + k)  4. (7.30)
n n n n
k=1 k=1

Equality sign
 holds onlyfor f (z) = f (z; θ, c, n), where c is an arbitrary constant
such that n 1/4  c  n 1/2.
7.2. Covering theorems for radial segments 223

Proof. We stick to the notation used in the proof of Theorem 7.11. As previously,

we can assume that aj < ∞, j = 1, . . . , 2n. We look at the function ω  = X(ω) ≡
− log p1 (e i(θ+π/n) ω 
e ), which maps the domain B = {ω : e ω+i(θ+π/n)
∈ D1 } onto
the strip |Im ω  | < π/2. The ‘lower’ boundary curve of B  has the equation ω =
 
ω(t) = log f (t1/n ei(θ+απ) ; θ, c, n)−i(θ+π/n), 0  t  1. Since n 1/4 < c  n 1/2,
it follows that 1/(2n)  α < 1/n. Simple calculations show that for α = 1/(2n)
the function cos Im(nω(t)) = (t − 1)(t2 − 2t cos 2ψ + 1)−1/2 cos ψ is decreasing on
(0, 1) and −π < Im(nω(t)) < 0 on this interval. Hence Im ω(t) is also decreasing
on (0, 1) and the domain B  contains its own right translations. Repeating the
corresponding part of the proof of Theorem 7.11, we see that the function X(ω) 

is convex on the real axis (if α = 1/(2n), then X(ω) is a linear function). Taking
(7.28) and (7.29) into account, in view of convexity, we obtain
   α21 n
n α22 n
1/n
1 2p1 a2k−1 ei(θ+π/n) 2p2 cei(θ+2π/n) .
k=1

Hence n
1/n
a2k−1  p−1
1 (1/2)e
−i(θ+π/n)
= (4 − c−n )−1/n ,
k=1

which
 is equivalent
 to (7.30). Equality holds here only for w = f (z; θ, c, n), where
n
1/4  c  n 1/2 by the hypotheses of the theorem. 
We introduce some further notation:
G(z; R) is the Grötzsch extremal function (see § 2.1 and Exercise 7.1(3)),

Pn = G(Rn ; Rn ), dn ≡ dn (c) = (cPn − 1)/(c − Pn ), dn (∞) = Pn ,



1 − dn G(z n ; Rn )
G(z; c, n, R) = n , Pn < c  ∞,
G(z n ; Rn ) − dn
 √
n
G(z; Pn , n, R) ≡ G(z; n, R) = n G(z n ; Rn ) ( 1 = 1).

It is easy to see that w = G(z; c, n, R), where Pn  c  ∞, is a function


in M(R) which maps the annulus K onto the domain |w| > 1 cut along the rays
arg wn = 0, |wn |  c; arg wn = π, |wn |  dn . Here and below we set 0−1 = ∞ and
∞−1 = 0.
For an arbitrary function f in M(R) let Xf (θ) denote the distance from
the origin to the closest boundary point of f (K) lying on the ray arg w = θ,
1 < |w| < ∞. If no such point exists for some θ, then we set Xf (θ) = +∞.
Theorem 7.13. If f ∈ M(R), then for each θ
>
? n  
? 2π 2π 
@
n n/2
Xf (θ + k) − Xf
−n/2
(θ + k)  (Pn − 1)/ Pn
n n
k=1
224 Chapter 7. Univalent Functions

with a possible exception of the case when n = 1 and the whole of the ray arg w =
θ + π, 1 < |w|  ∞ lies in the domain f (K). Equality holds if and only if
f (z) ≡ eiθ G(αz; n, R), where α is an arbitrary constant such that |α| = 1.
Proof. For θ = 0 the inequality in Theorem 7.13 is an immediate consequence of
inequality (5.14) for the continuum E = {z ∈ U : 1/z ∈ f (K)}. The uniqueness of
the extremal function (for fixed α) follows from the conditions ensuring equality
sign in the inequalities in the proof of Theorem 5.7, in the special case when E is
as above. 
In a similar way, from Corollary 5.2 we obtain the inequality
n

Xf (θ + k)  Pn , (7.31)
n
k=1

which holds for any θ and f ∈ M(R), with the possible exception of the case when
n = 1 and the ray arg w = θ + π, 1 < |w|  ∞, lies in f (K) entirely. In particular,
for holomorphic functions in M(R) we have

Xf (θ)  P (R)
and
1 
(Xf (θ) + Xf (θ + π))  P (R) + P 2 (R) − 1,
2
for each real θ (P (R) = P1 ).

Exercises 7.2
(1) Establish (7.22) using the principle of circular symmetrization.
(2) (Emel’yanov [E]) Prove (7.25) using Theorem 6.5. Hint: see Figure 7.1.

g )V *
g )2*

Figure 7.1.
7.3. Distortion theorems related to n-symmetry 225

(3) (Bermant; see [Gol], Ch. IV, § 6) Show that the product of the areas of the
star of the image of the disc U under a map in the class S and the inverse
image of this star is at least π 2 .
(4) Let f be a function in M(R). Show that for each real θ the quantities
   
1 n 1 n
n n
2π π 2π
A= Xf θ + k and B = Xf θ + + k
n n n n n
k=1 k=1

satisfy
(A + B)−1 + (A−1 + B −1 )−1  Pn
with the possible exception of the case when n = 1 and one of the rays
arg w2 = 2θ, 1 < |w|  ∞, lies in f (K) entirely. Equality sign holds here if
and only if
f (z) ≡ eiθ G(αz; c, n, R),
where α and c are arbitrary constants satisfying |α| = 1 and Pn  c  ∞.
(5) Assume that for some f ∈ S the f -image of the disc U intersects each line
Re w = u, u  0, in a set with linear measure at most π/2. Using Steiner
symmetrization and circular symmetrization prove that

Λf (0) > π/(2 π 2 − 1).

(6) Let f be a function in the class S. Show that if 0 < θ  π, then for some
u  (cos θ)/(4 sin θ/2) the line interval lu = {w = u + iv : |v|  (cos θ/2)/2}
lies in the f -image of the disc U . The only exception is when f maps the
disc U conformally and univalently onto the plane Cw cut along the ray
{w = u + iv : u  (cos θ)/(4 sin θ/2), v ± (cos θ/2)/2}.
Hint: Use in succession the polarizations with respect to the lines

v = ((cos θ/2)/2)(1 − 2−n ), n = 0, 1, 2, . . . ,

which are oppositely directed relative to the real axis (Corollary 3.1).

7.3 Distortion theorems related to n-symmetry


Here we understand n-symmetry as mirror symmetry relative to all the rays arg(z−
z0 ) = θ + 2πk/n, k = 1, . . . , n, for some z0 and θ. We considered the case n = 1
or (equivalently) n = 2 in § 7.1. In particular, one of the extremal functions in the
class M turned out to be the Koebe function k(z) = z(1 + z)−2 , which possesses
2-symmetry. For an arbitrary n its analogue has the form kn (z) := (k(z n ))1/n .
The function kn (z) maps the disc U conformally and univalently onto the w-plane
cut along the rays arg wn = 0, |wn |  1/4. Now we set Dk := {reiθ : 0 < r <
∞, |θ − 2πk/n| < π/n}, k = 1, . . . , n.
226 Chapter 7. Univalent Functions

Theorem 7.14. Let f be a function in the class M and let zk , k = 1, . . . , n (n  2),


be any points such that |zk | = r, 0 < r < 1, and f (zk ) ∈ Dk , k = 1, . . . , n. Then
>  n 
?  f 2 −1 (z )f  (z ) 
? n
 k  1 − rn
@
n

k
 . (7.32)
 Re f 2 (z )  r(1 + rn )
n
k=1 k

Here equality holds for f (z) = ckn (z) and zk = rei2kπ/n , k = 1, . . . , n, where c is
an arbitrary positive coefficient.

Proof. We look at the condensers


 n 
 
n
C∗ = [0, rei2kπ/n ], [(r + r)ei2kπ/n , ei2kπ/n ] ∪ {z : z ∈
/ U}
k=1 k=1
and  

n 
n
C= [0, zk ], [zk + zk r/r, zk /r] ∪ {z : z ∈
/ U} ,
k=1 k=1

where the positive number r is taken sufficiently small so that the points f (zk +
zk r/r) also lie in the sectors Dk . We can obtain the condenser C from C ∗ using
dissymmetrization from Lemma 4.2, so their capacities satisfy

cap C ∗  cap C. (7.33)

Let f (C) be the f -image of C. Then by Corollary 5.3


n
K(q )
cap f (C)  2  k , (7.34)
k=1 K 1 − qk2
√ n
where qk = 1/d−1
√ 1 (|dt ( f (zk + zk r/r))|) for dτ (z) := (z − τ )/(z + τ ) and t =
f n (zk ), Re t > 0. For the complete elliptic integral K(τ ) we have the following
asymptotic expansions ([Ach], § 32):

4
K(τ ) = log √ + o(1), τ → 1,
1 − τ2
 π 
K 1 − τ 2 = + o( 1 − τ 2 ), τ → 1,
2
Thus
K(q ) 2 4
 k = log  + o(1), k = 1, . . . , n.
K 2
1 − qk π 1 − qk2

Here and below o(1) denotes an infinitesimal quantity as r → 0.


7.3. Distortion theorems related to n-symmetry 227

On the other hand the condenser C ∗ is conformally equivalent to


({z : 0  z n  1}, {z : z n  Rn }) ,
where
 2/n
rn/2 + r−n/2
R= , R > 1,
(r + r)n/2 + (r + r)−n/2
and its capacity is

K R−n 4n 4
cap C ∗ = 2n
√ = log √ + o(1).
K 1−R −n π 1 − R−n
In view of (7.33) and (7.34),
1
n
(1 − qk2 )
k=1
1 (1 + o(1)). (7.35)
(1 − R−n )n
We look at the expressions on the right-hand side of (7.35) separately:
n r(1 − r2n + o(1))
1 − R−n = ,
r(r + r)n (rn + r−n + 2)
√ n (7.36)
1 − |dt ( f (zk + zk r/r))|
qk = √ ,
1 + |dt ( f n (zk + zk r/r))|
√ n
and for t = f (zk ),
√ n √
 f (zk + zk r/r) − f n (zk )
dt ( f (zk + zk r/r)) = √
n
√ .
f n (zk + zk r/r) + f n (zk )
Since
  n zk r 
f n (zk + zk r/r) = f n (zk ) + f (zk )f n/2−1 (zk ) + o( r),
2 r
it follows that
 n/2−1 
f (zk )f  (zk ) 
1− qk2 
= n r  √ + o( r), r → 0. (7.37)
Re f n (z )  k

Substituting (7.36) and (7.37) into (7.35) and letting r approach zero we obtain
the required inequality (7.32). The equality case is straightforward. 
Note that if arg f (zk ) = 2πk/n, k = 1, . . . , n, then the left-hand side of
(7.32) has the following form:
>
? n
?
@
n
|f  (zk )/f (zk )|.
k=1
228 Chapter 7. Univalent Functions

Corollary 7.5. If f ∈ M, then the following sharp bound holds for each fixed r,
0 < r < 1, and any different points z1 , z2 , z3 on the circle |z| = r,
 
1 3 
 f (zk )
 √  3
 3 1 − r3
k=1
 .
|(f (z1 ) − f (z2 ))(f (z2 ) − f (z3 ))(f (z3 ) − f (z1 ))| 9 r(1 + r3 )
Here equality sign holds for f (z) = k3 (z) and the points zk = r exp(2πi(k −
1)/3), k = 1, 2, 3. (If some zk is a pole of f , then the derivative f  (zk ) is treated
as the limit lim ((z −zk )f (z))−1 , and the factors containing f (zk ) on the left-hand
z→zk
side of the inequality are dropped).
Proof. For f ∈ M let Φ be the conformal automorphism of the w-plane normalized
by the conditions Φ(f (zk )) = exp(2πi(k − 1)/3), k = 1, 2, 3. Applying Theorem
7.14 to F (z) = Φ(f (z)) we obtain
 3 
 
 
 F (zk )  [(1 − r3 )/(r(1 + r3 ))]3 .

 
k=1

It remains to calculate the product of the derivatives F  (zk ). The simplest way is to
use the fact that the expression on the left-hand side of the inequality in Corollary
7.5 is invariant under linear fractional maps. The equality case is straightforward.

The next two statements trace back to Goluzin’s results in [Gol], Ch.IV, § 2.
Here we limit ourselves to sharp estimates such that the corresponding extremal
functions possess n-symmetry. Both inequalities follow from the monotonicity of
reduced moduli.
Theorem 7.15. The following sharp estimate holds for f ∈ Σ and any points zk ,
k = 1, . . . , n, n  1, on the circle |z| = ρ with ρ > 1 :
 n  n n  
   f (zk ) − f (zl )  n n
      ρ−2n2
 f (zk )  zk − zl  |1 − z k zl |. (7.38)
 
k=1 k=1 l=1 k=1 l=1
l=k

In (7.38) equality is attained for f (z) = z(1+z −n)2/n and zk = ρ exp(2πik/n), k =


1, . . . , n, or alternatively, for the zk = ρ exp(πi/n + 2πik/n), k = 1, . . . , n.
Proof. We set

B = f ({z : |z| > 1}), W = {f (z1 ), . . . , f (zn ), ∞},


= {1, . . . , 1, −n} and Ψ = {r, . . . , r, r}.

Then the first inequality in Theorem 2.10 yields

M (B, ∅, W, , Ψ)  M (W, , Ψ). (7.39)


7.3. Distortion theorems related to n-symmetry 229

By (2.17) the reduced modulus on the right-hand side of (7.39) is equal to


 n n 
1
− log |f (zk ) − f (zl )| .
2π(n + n2 )2
k=1 l=1
l=k

The simplest way to calculate the second reduced modulus in (7.39) is to use The-
orem 2.13 with some obvious modifications while taking account of the symmetry
relative to the circle |z| = 1, in place of the symmetry relative to the real axis:
 n
1
n
n
 2
M (B, ∅, W, , Ψ) = log |zk f (zk )| − log |zk − zl |
4π(n + n2 )2
k=1 k=1 l=1
l=k
   

n
n
1 1 
n
n
 1 
− 
log  −  − 2 
log zk −  .
zk zl zl
k=1 l=1 k=1 l=1
l=k

Substituting the expressions for the reduced moduli into (7.39) we obtain (7.38).
The equality case is straightforward, but we can also observe that equality in (7.39)
holds thanks to the symmetry of the extremal functions and the system of points
zk , k = 1, . . . , n. 

Theorem 7.16. If f ∈ Σ and z1 , . . . , z2n are points on some circle |z| = ρ with
ρ > 1, then
   
 2n  2n 2n
 f (zk ) − f (zl ) (−1)
k+l
2n 2n
     k+l+1
 f (zk )  zk − zl   |1 − z k zl |(−1) . (7.40)
 
k=1 k=1 l=1 k=1 l=1
l=k

In (7.40) equality holds for f (z) = z(1 + z −2n )1/n and zk = ρ exp(πi/2n + πik/n),
k = 1, . . . , 2n.

Proof. Let B = f ({z : |z| > 1}), W = {f (z1 ), f (z2 ), . . . , f (z2n )}, = {0, −1, 1,
. . . , −1, 1}, = {−1, 1, . . . , −1, 1}, Ψ = {r, r, . . . , r}. Then by (2.18),

M (B, ∂B, W, , Ψ)  M (W, , Ψ).

As above, we calculate the reduced modulus of the Riemann sphere by formula


(2.17):
 2n 2n 
1
M (W, , Ψ) = − (−1)k+l
log |f (zk ) − f (zl )| .
2π(2n)2
k=1 l=1
l=k
230 Chapter 7. Univalent Functions

We can find the second reduced modulus using (2.16):

M (B, ∂B, W, , Ψ) = M (D, ∂D, Z, , Ψ) 


 2n
2n
2n  
1 |zk |2 − 1  1 − z k zl 
= log + (−1)k+l 
log   ,
2π(2n)2 1/|f  (zk )| zk − zl 
k=1 k=1 l=1
l=k

where

D = {z : |z| > 1}, Z = {z1 , . . . , z2n },


and
 = {r/|f  (z1 )|, r/|f  (z2 )|, . . . , r/|f  (z2n )|}.
Ψ

It remains to add up the above relations. The equality case can be verified directly
or using symmetry arguments. 
Comparing (7.38) for n = 2 with (7.40) for n = 1 we arrive at Goluzin’s
theorem on the distortion of chords under mappings of the domain |z| > 1 by
functions in Σ:  
 f (z1 ) − f (z2 ) 
 1− 1 ,
 z1 − z2  ρ2
where z1 and z2 are any points on the circle |z| = ρ, ρ > 1 ([Gol], Ch. IV, § 2).
Theorem 7.17. Let f be a function mapping the disc U conformally and univalently
onto a bounded domain D with transfinite diameter d(D)  1. Let zk be points in
U such that arg f (zk ) = 2π(k − 1)/n, k = 1, 2, . . . , n. Then
1
n 1
n
|zk − zl |
n k=1 l=1
l=k
nn |f  (zk )||f (zk )|n−1  1n 1 n .
k=1 |1 − zk z l |
k=1 l=1

Here equality holds for w = [3z n /(4−2z n )]1/n and zk = (1/2)1/n exp(2πi(k−1)/n),
k = 1, 2, . . . , n.
Proof. We introduce some notation: Z = {zk }nk=1 , = {δk }nk=0 , Ψ = {μk r}nk=1 ,
where μk = δk = 1 for k = 1, . . . , n and δ0 = 0; F (z) = 1/f (z), W = {F (zk )}nk=1 .
By formula (2.16)
 n  
1 n n
 1 − zk z l 
2
log(1 − |zk | ) + log   = M (U, ∂U, Z, , Ψ)
2πn2 zk − zl 
k=1 k=1 l=1
l=k

1
n
=− log(|f  (zk )||f (zk )|−2 ) + M (F (U ), ∂F (U ), W, , Ψ).
2πn2
k=1
7.3. Distortion theorems related to n-symmetry 231

Let D0 be the connected component of the image of the set Cw \ D under the map
ζ = 1/w which contains the origin. By the hypotheses r(D0 , 0)  1. Let g(ζ, ζ  )
denote the Green function of the domain F (U ). Again, using formula (2.16) in
combination with (6.28) we obtain

M (F (U ), ∂F (U ), W, , Ψ)
 n 
1
n
n
= log r(F (U ), F (zk )) + g(F (zk ), F (zl ))
2πn2
k=1 k=1 l=1
l=k
 
1
n
n
n
 n2 log r(D0 , 0) + log r(F (U ), F (zk )) + g(F (zk ), F (zl ))
2πn2
k=1 k=1 l=1
l=k
 n
  n

1 1
 log n−n |F (zk )| n+1
= log n−n |f (zk )| −n−1
.
2πn2 2πn2
k=1 k=1

Summing the above relations we arrive at the required inequality. The equality
case is straightforward. 
 √
The n-symmetric Grötzsch function G(z; n, R) = n G(z n ; Rn ) ( n 1 = 1),
which maps the annulus K conformally and univalently onto the domain |w| > 1
cut along the rays arg wn = 0, Pn  |wn |  ∞, where Pn = G(Rn ; Rn ), is one of
the extremal functions in the class M(R).
Theorem 7.18. If f ∈ M(R), then for arbitrary real θ and n  2,
 
 n

 f  (exp(i(θ + 2πk/n)))  |G (exp(iπ/n); n, R)|n .

k=1

Proof. We can assume that θ = π/n. We set zk = exp(i(π/n + 2πk/n)), k =


1, . . . , n, and look at the family of subdomains Dk = {z : R−1 < |z| < R, | arg z −
π/n − 2πk/n| < π/n}, k = 1, . . . , n, of the annulus R−1 < |z| < R. The function f
takes them to the family of pairwise nonoverlapping domains f (Dk ), k = 1, . . . , n,
so that the points f (zk ) lie on the circle |w| = 1. The Grötzsch function G(z; n, R)
maps the Dk onto n disjoint sectors with vertex at the origin and equal opening
angles so that G(zk ; n, R) = zk , k = 1, . . . , n. By Theorem 6.11
n n
r(f (Dk ), f (zk ))  r(G(Dk ; n, R), G(zk ; n, R)) = r(G(Dn ; n, R), zn )n .
k=1 k=1

It remains to observe that for each function f in the class M(R) we have

r(f (Dk ), f (zk )) = |f  (zk )|r(Dk , zk ), k = 1, . . . , n. 


232 Chapter 7. Univalent Functions

Theorem 7.19. If f is a function in the class M(R), then for n  1, 0 < ϕ < π/n,
and 1  ρ < R,
 2n 
  2n 2n
  k+l
 f  (zk ) |f (zk ) − f (zl )|(−1)
 
k=1 k=1 l=1
l=k (7.41)
 2n
 G (ρn einϕ ; Rn ) 

  n−1  ,
nρ (G(ρ e ; R ) − G(ρ e
n inϕ n n −inϕ ; R )) 
n

where

z2j−1 = ρ exp(i(−ϕ + 2πj/n)) and z2j = ρ exp(i(ϕ + 2πj/n)), j = 1, . . . , n.

In (7.41) equality holds for the functions G(z; m, R) with m such that the system
of points {zk }2n
k=1 is mirror-symmetric relative to the rays arg z
m
= 0.

Proof. We set K ∗ = {z : 1/R < |z| < R}, Z = {zk }2n 2n


k=1 , W = {f (zk )}k=1 ,
k 2n νk 2n
= {(−1) }k=1 , and Ψ = {μk r }k=1 , where μk = νk = 1 for k = 1, . . . , 2n.
Using in turns formula (2.16), the monotonicity of reduced modulus, and formula
(2.17) for the reduced modulus of the Riemann sphere we obtain

M (K ∗ , ∂K ∗ , Z, , Ψ)
2n
∗ ∗ 1
= M (f (K ), ∂f (K ), W, , Ψ) − log |f  (zk )|
8πn2
k=1
2n

1
 M (W, , Ψ) − log |f  (zk )|
8πn2
k=1
 2n
2n 2n

1

= − (−1)k+l
log |f (zk ) − f (zl )| − log |f (zk )| .
8πn2
k=1 l=1 k=1
l=k

If m is an integer such that the set {zk }2n k=1 is mirror-symmetric relative to the
rays arg z m = 0, then the set {G(zk ; m, R)}2n
k=1 is mirror-symmetric relative to the
2n
rays arg wm = 0. Hence the function (−1)k log |w − G(zk ; m, R)| vanishes on
k=1
the boundary of the domain G(K ∗ ; m, R). From Theorem 2.9 we conclude that for
the function G(z; m, R) with such m we have equalities in the above relations. It
remains to show that the quantity exp(−8πn2 M (K ∗ , ∂K ∗ , Z, , Ψ)) is equal to the
right-hand side of (7.41). We shall use the behaviour of the reduced modulus under
conformal mappings. Let Kn∗ = {ζ : 1/Rn < |ζ| < Rn }, Zn = {ρn e−inϕ , ρn einϕ },
n = {0, −1, 1}, and Ψn = {r/nρn−1 , r/nρn−1 }. Then using (2.16) and bearing
7.3. Distortion theorems related to n-symmetry 233

in mind the symmetry of H(ζ) := G(ζ; Rn ) we obtain

1
M (K ∗ , ∂K ∗ , Z, , Ψ) = M (Kn∗ , ∂Kn∗ , Zn , n , Ψn )
 n 
1 r(Kn∗ , ρn einϕ ) n −inϕ n inϕ
= 2 log − 2gKn∗ (ρ e ,ρ e )
8πn 1/nρn−1
 
1 r(H(Kn∗ ), H(ρn einϕ )) n −inϕ
= 2 log − 2gH(Kn∗ ) (H(ρ e n inϕ
), H(ρ e ))
8πn 1/nρn−1
1
− log |H  (ρn einϕ )|
4πn
1 1
= M ({H(ρn e−inϕ ), H(ρn einϕ )}, n , Ψn ) − log |H  (ρn einϕ )|
n 4πn
1 ' (
= −2 log(1/nρn−1 ) + 2 log |H(ρn e−inϕ ) − H(ρn einϕ )|
8πn
1
− log |H  (ρn einϕ )|. 
4πn

We preface the proof of the next theorem by two lemmas on the calculation
of the reduced moduli of symmetric configurations.
Lemma 7.1. Let K ∗ = {z : 1/R < |z| < R}, Z = {1, e2πi/n , . . . , e2πi(n−1)/n },
= {0, 1, 1, . . . , 1}, and Ψ = {r, r, . . . , r}. Then
 ∞

1 2/k
∗ ∗
M (K , ∂K , Z, , Ψ) = log(R n/2
/n) − .
2πn 1 + R2nk
k=1

Proof. By the definition of a reduced modulus


 
1
M (K ∗ , ∂K ∗ , Z, , Ψ) = lim |C(r; K ∗ , ∂K ∗ , Z, , Ψ)| + log r ,
r→0 2πn

where we can assume that the condenser C(r; K ∗ , ∂K ∗ , Z, , Ψ) is mirror-sym-


metric relative to the circle |z| = 1 and the rays arg z = πk/n, k = 0, . . . , 2n − 1.
Let ω(z) denote the potential function of the condenser; it also has this symmetry.
We look at the map w = ((zeπi/n )n/2 + (zeπi/n )−n/2 )/2, which takes the domain
G := {z ∈ K : | arg z| < π/n} to D ∩ {w : Im w > 0}, where D is the interior of
the ellipse with foci at ±1 and semiaxes (Rn/2 + R−n/2 )/2 and (Rn/2 − R−n/2 )/2.
Let z = z(w) be the inverse map and let

ω(z(w)), Im w  0,
 (w) =
ω (7.42)
ω(z(w)), Im w  0.

In view of the symmetry principle for harmonic functions and Dirichlet’s principle,
234 Chapter 7. Univalent Functions

we find that
  −1
|C(r; K ∗ , ∂K ∗ , Z, , Ψ)| = 2n |∇ω|2 dxdy
G
  −1
1
= n ω|2 dudv
|∇ = |C(r; D, ∂D, {0}, {0, 1}, {(n/2)r})|.
n
D

Recalling again the definition of the reduced modulus we obtain


M (D, ∂D, {0}, {0, 1}, {(n/2)r})
 
1
lim |C(r; D, ∂D, {0}, {0, 1}, {(n/2)r})| + log r .
r→0 2π
Thus
1
M (K ∗ , ∂K ∗ , Z, , Ψ) =M (D, ∂D, {0}, {0, 1}, {(n/2)r})
n
1
= log[r(D, 0)/(n/2)].
2πn
It remains to use the formula for the inner radius of the domain D (see the tables
at the end of [PS]). 
Lemma 7.2. Let Q = Cz \ {z = x + i0 : |x|  1/a or |x|  a}, Z = {i, −i},
= {0, 1, 1}, and Ψ = {μr, μr}, where a > 1 and μ > 0. Then
1
M (Q, ∂Q, Z, , Ψ) = log[(a + 1/a)/μ].

Proof. Similarly to the proof of the previous theorem, we look at the condenser
C(r; Q, ∂Q, Z, , Ψ), which is mirror-symmetric relative to the real axis and the
circle |z| = 1; let ω(z) denote its potential function. Let z(w) be the branch of the
inverse of the Zhukovskii (Joukowsky) transform w = (z + 1/z)/2 which maps the
upper half-plane Im w > 0 onto {z : |z| > 1, Im z > 0} =: G. Defining ω  (w) as in
(7.42) we obtain
  −1    −1
|C(r; Q, ∂Q, Z, , Ψ)| = 4 |∇ω|2 dxdy = 2 ω |2 dudv
|∇
G Cw
1
= |C(r; D, ∂D, {0}, {0, 1}, {μr})|,
2
where this time D is the w-plane cut along the rays {w = u + iv : v = 0, |u| 
(a + 1/a)/2}. Hence
1
M (Q, ∂Q, Z, , Ψ) = M (D, ∂D, {0}, {0, 1}, {μr})
2
1 1
= log[r(D, 0)/μ] = log[(a + 1/a)/μ]. 
4π 4π
7.3. Distortion theorems related to n-symmetry 235

The quantity Xf (θ), f ∈ M(R), was defined at the end of § 7.2.


Theorem 7.20. For functions f in the class M(R),
∞ 
n
|f  (zk )| 2n 2n/k
 n2 /2 exp , (7.43)
Xf (αk )π/βk + Xf (αk )−π/βk R 1 + R2nk
k=1 k=1

where zk = exp(2πik/n), n  2, αk = (ϕk+1 + ϕk )/2, and βk = ϕk+1 − ϕk , where


ϕk = arg f (zk ) for k = 1, . . . , n and ϕn+1 = ϕ1 + 2π, and the continuous branch of
the argument is chosen so that ϕ1 < ϕ2 < · · · < ϕn . In (7.43) equality is attained
for the function w = G(eiπ/n z; n, R).
Proof. We take the family of domains

{Dk }nk=1 , Dk = {w : ϕk < arg w < ϕk+1 }, k = 1, . . . , n,

and the family of functions {pk }nk=1 , ζ = pk (w) ≡ i(e−iϕk w)π/βk , which maps the
domains Dk onto the right half-plane Re ζ > 0, respectively. Let {Ck }nk=1 be the
result of the separating transformation of the condenser

C(r; f (K ∗ ), ∂f (K ∗ ), W, , Ψ)

with respect to the family of functions {pk }nk=1 , where

W = {f (z1 ), . . . , f (zn )}, = {0, 1, . . . , 1},  = {|f  (z1 )|r, . . . , |f  (zn )|r}.
Ψ

Since the capacity is a conformal invariant, from Theorem 4.8 we obtain



cap C(r; K ∗ , ∂K ∗ , Z, , Ψ) = cap C(r; f (K ∗ ), ∂f (K ∗ ), W, , Ψ)
1
n
 cap Ck , Z = {z1 , . . . , zn }, Ψ = {r, . . . , r}.
2
k=1

Next we set
Ck∗ = s−1 (R(s(Ck ))), k = 1, . . . , n,
where s(ζ) = (ζ − i)/(ζ + i) and R denotes symmetrization with respect to the
unit circle. Finally, we set
k = C(r; Qk , ∂Qk , {i, −i}, {0, 1, 1}, {μkr, μk r}),
C

where

Qk = Cζ \ {ζ = ξ + i0 : |ξ|  Xf (αk )−π/βk or |ξ|  Xf (αk )π/βk },


and
π 
μk = |f (zk )f  (zk+1 )|, k = 1, . . . , n.
βk
236 Chapter 7. Univalent Functions

Since the capacity is monotonic, it follows from Theorem 4.3 that


k ,
cap Ck  cap Ck∗  cap C k = 1, . . . , n.

Summing these relations we obtain

1
n
cap C(r; K ∗ , ∂K ∗ , Z, , Ψ)  k .
cap C
2
k=1

Hence

n
2 ∗ ∗
n M (K , ∂K , Z, , Ψ)  2M (Qk , ∂Qk , {i, −i}, {0, 1, 1}, {μkr, μk r}).
k=1

We complete the proof of inequality (7.43) using Lemmas 7.1 and 7.2 and the
classical inequality
1
n n
1/n 2π
βk  βk = .
n n
k=1 k=1
iπ/n
For f (z) = G(e z; n, R) we have equality in all the above relations, and therefore
we also have equality in (7.43). 

Exercises 7.3
(1) From (7.32) deduce a lower bound for the absolute value of the logarithmic
derivative in the class R0 .
(2) Deduce Fan’s inequalities (see § 7.1) using Theorem 7.14.
(3) (Schiffer [Sch]) For a function f (z) = z + a0 + a−1 /z + a−2 /z 2 + · · · in
the class Σ prove the sharp bound |a−2 |  2/3. Here equality holds for
f (z) = z(1 + eiθ z −3 )2/3 + c, where c is a complex and θ a real constant.
Hint: Use Corollary 7.5.
(4) Let K(m/M ) be the class of holomorphic univalent functions which do not
vanish in the annulus m < |z| < M (m > 0) and have the following property:
for m < r1 < r2 < M the image of the circle |z| = r1 lies inside the image
of |z| = r2 (see [Gol], the appendix). Let f ∗ denote the function in the class
K(m/M ) such that f ∗ (m) = 1 and f ∗ maps m < |z| < M conformally and
univalently onto the w-plane cut along the line segments arg wn = 0, 0 
|w|  1, and rays arg wn = 0, λ  |wn |  +∞, where λ = λ(n, m, M ).
Let f be a function in K(m/M ) and let ak exp(2πik/n) and bk exp(2πik/n),
(0 < ak < bk ) be points in different components of the complement of the
f -image of the annulus m < |z| < M , k = 1, . . . , n. Prove the inequality
>
? n
? n/2
(ak + bk )2
n/2
(λ1/2 + 1)2
@
n
 .
k=1
(ak bk )n/2 λ1/2
7.3. Distortion theorems related to n-symmetry 237

Here
√ we have equality sign for the function f ∗ and points exp(2πik/n) and
n
λ exp(2πik/n), k = 1, . . . , n. Deduce a distortion theorem in the class
K(m/M ) from this result.
(5) Verify that the n-symmetric Teichmüller function
 √
n
T (z; n, R) := n T (z n ; Rn ) + 1, z ∈ K ( 1 = 1),

maps the annulus K conformally and univalently onto the plane cut along the
line segments arg wn = 0, 0  |w|  1 and rays arg wn = 0, T (Rn , Rn ) + 1 
|wn |  ∞. Here T (z; R) is the function in Exercise 7.1(3).
(6) Let f be a univalent meromorphic function in the annulus K, let r be a real
number such that 1 < r < R, and zk , k = 1, . . . , n (n  2), be arbitrary
points such that |zk | = r and | arg f (zk ) − 2πk/n| < π/n for some value of
the argument, k = 1, . . . , n. Prove the inequality
>
? n  
?  f (zk )n/2−1 f  (zk )  T  (r; n, R)
@  
 Re f (zk )n/2   T (r; n, R) .
n

k=1

Equality holds for functions of the form f (z) = cT (z; n, R) and for the points
zk = r exp(2πik/n), k = 1, . . . , n, where c is an arbitrary positive constant.
(7) Prove the following inequality for f ∈ M(R) and points zk such that |zk | = r,
1 < r < R, and arg f (zk )/f (z1 ) = 2π(k − 1)/n, k = 1, . . . , n:
>
? n  
?  f  (zk )  G (r; n, R)
@  
 f (zk )   G(r; n, R) .
n

k=1

Equality is attained for the function f (z) = eiψ G(e−iθ z; n, R) and points
zk = r exp(i(θ + 2πk/n)), k = 1, . . . , n, where θ and ψ are real constants.
(8) For f ∈ M(R) prove the following sharp estimate for triples of points z1 , z2 , z3
on the circle |z| = 1:
 3 
1  
 f (zk ) √   3
 3 G (r; 3, R)
k=1
 ,
|(f (z1 ) − f (z2 ))(f (z2 ) − f (z3 ))(f (z3 ) − f (z1 ))| 9 G(r; 3, R)

where the equality holds for f (z) = eiψ G(e−iθ z; 3, R) and the points zk =
exp(i(θ + 2πk/3)), k = 1, 2, 3, where θ and ψ are real constants.
238 Chapter 7. Univalent Functions

7.4 Variational principles of conformal mappings


The principles mentioned in the title of this section are important for function the-
ory and continuum mechanics. Here we discuss qualitative variational principles,
which enable us to control the variations of conformal mappings under changes
of the boundaries of the source domains. For example, a classical qualitative vari-
ational principle says that if a function w = f (z) maps a domain D such that
∞ ∈ D ⊂ z := {z : |z| > 1} conformally and univalently onto the exterior w of
the unit disc and f (∞) = ∞, then at the point at infinity we have |f  (∞)|  1; at
any point z ∈ (∂D) ∩ (∂ z ) we have |f  (z)|  1 (if the derivative is defined), and
for each ρ > 1 the level curve |f (z)| = ρ lies in the set |z|  ρ. In other words, as z
is deformed into D, the absolute values of the derivatives at infinity and at fixed
boundary points decrease and level curves extend. Similar principles hold for func-
tions performing conformal mappings onto other types of canonical domain: the
half-plane Hw := {w : Im w > 0} and the strip Sw := {w : 0 < Im w < 1} [LaSha],
§ 61 (by contrast with [LaSha], we compare the function f with the identity map-
ping, which in involves no loss of generality). These principles can be treated as
properties of complex potentials of stationary plane-parallel flows. They follow, for
instance, from Schwarz’ lemma or the maximum principle for harmonic functions.
We are interested in variational principles that can be deduced from potential
theory and symmetrization in a unified fashion. For a fixed canonical domain Kz
we analyze the properties of a function f : D → Kw in their dependence on the
deformation
(D \ Kz ) ∪ (Kz \ D).

If D ⊂ Kz , then we speak about an inner deformation of the domain Kz and


if D ⊃ Kz , then we speak about an outer deformation. We shall focus on the
variation of the absolute value of the derivative of f , that is, on estimates for
dilations. Apart from the canonical domains above we shall also look at the annulus
Kw (R) = {w : 1 < |w| < R} and the polygonal quadrilateral Qw (a1 , a2 , a3 , a4 )
in Cw whose vertices at the points a1 , a2 , a3 , a4 are arranged in accordance with
the positive orientation of its boundary. Now we introduce the following function
classes, which match the variational principles under consideration:
V1 is the class of functions w = f (z) mapping a simply connected domain
D = D(f ), ∞ ∈ D ⊂ Cz conformally and univalently onto the exterior of the unit
disc w and having an expansion of the following form at infinity:

f (z) = αz + O(1), z → ∞;

V2 is the class of functions w = f (z) mapping a domain D = D(f ) such that


the union (Hz \ D) ∪ (D \ Hz ) is a bounded set, conformally and univalently onto
the half-plane Hw and having an expansion of the form

f (z) = z + O(1), z ∈ D, z → ∞;
7.4. Variational principles of conformal mappings 239

V3 is the class of functions w = f (z) mapping a domain D = D(f ) such that


the union (Sz \ D) ∪ (D \ Sz ) is a bounded set, onto the strip Sw and satisfying

lim f (z) = +∞, lim f (z) = −∞.


z→+∞ z→−∞
z∈D z∈D

V4 (l) is the class of functions w = f (z) satisfying the following conditions:


the domain of f is a quadrilateral D = D(f ) with vertices at points a1 , a2 , a3 ,
a4 , arranged in accordance with the positive orientation of the boundary of D and
such that the side (a1 , a2 ) lies on a line segment [l, l + i] with l > 0 and the side
(a3 , a4 ) lies on [i, 0]; the function w = f (z) maps D conformally and univalently
onto a polygonal quadrilateral Q(L, L + i, i, 0), L = L(f ) > 0, so that f (a1 ) = L,
f (a2 ) = L + i, f (a3 ) = i, and f (a4 ) = 0;
V5 is the class of functions w = f (z) which are defined in a doubly connected
domain D = D(f ) separating the disc |z|  1 from infinity and map this domain
conformally and univalently onto an annulus Kw (R) with R = R(f ) so that the
outer boundary component of D corresponds to the circle |w| = R.
We point out constraints imposed on functions in these classes in the case
of inner deformations of the canonical domain. For instance, if of functions f in
V1 we request additionally to satisfy D(f ) ⊂ z , then Schwarz’ lemma yields
|α|  1, where the equality |α| = 1 is only possible for f (z) = eiϕ z with some real
ϕ. Thus for |α| > 1 the subclass of f ∈ V1 , f  (∞) = α, such that D(f ) ⊂ z is
empty; and if |α| = 1, then this subclass contains only the rotations of the domain
z about the origin. Further, if f (z) = z + c/z + · · · is a function in V2 and
D(f ) ⊂ Hz , then c  0 with c = 0 only for f (z) ≡ z. Comparing the conformal
moduli of quadrilaterals and rings we also obtain the following results: for inner
deformations in the class V3 (that is, for f satisfying D(f ) ⊂ Sz ) we necessarily
have
lim (f (z) − z) − lim (f (z) − z)  0;
z→+∞ z→−∞
z∈D z∈D

from the conditions f ∈ V4 (l) and D(f ) ⊂ Qz (l, l + i, i, 0) it follows that l  L;


and if f ∈ V5 and D(f ) ⊂ Kz (R), then necessarily R(f )  R.
Now we discuss a general variational result for quadratic forms with coeffi-
cients depending on Robin capacities and Robin functions. Here we limit ourselves
to inner deformations of domains: the corresponding results for outer deformations
will only be different in notation. Let D and Kz be finitely connected domains
without isolated boundary points, let D ⊂ Kz ⊂ Cz , and let Kw be a copy of
Kz on the Riemann sphere Cw . Let Γ and Γ be closed subsets of ∂D and ∂Kz ,
respectively, which consist of finitely many nondegenerate connected components;
let {zk }nk=1 be a set of different admissible points in D with exponent 1 which

n
lie in D \ (Γ ∪ Γ ); let {δk }nk=1 be a tuple of real numbers such that δk2 = 0.
k=1
Assume that a function w = f (z) mapping D conformally and univalently onto
the domain Kw exists and the derivatives f  (zk ) are well defined, k = 1, . . . , n. Let
240 Chapter 7. Univalent Functions

f (Γ) be the image of Γ in the sense of boundary correspondence. By conformal


invariance

n
n
σ(zk )δk δl [gD (zk , zl , Γ) − gKz (zk , zl , Γ )]
k=1 l=1
(7.44)
n
n

= σ(zk )δk δl [gKw (f (zk ), f (zl ), f (Γ)) − gKz (zk , zl , Γ )] ,
k=1 l=1

where for k = l the difference in square brackets is defined by the formula


gϕ(B) (ϕ(z), ϕ(z), ϕ(Γ)) − gB  (z, z, Γ)
= lim [gϕ(B) (ϕ(ζ), ϕ(z), ϕ(Γ)) − gB  (ζ, z, Γ )];
ζ→z

here ϕ can be an arbitrary map (for example, the identity map). Inequalities for
the left-hand side of (7.44) can be treated as expressing some monotonicity of the
quadratic form, while inequalities for the right-hand side are qualitative variational
principles.
Thus the proof of some variational principles reduces to using methods in
potential theory. The function

n
u(z) = δk gB (z, zk , Γ)
k=1

can be important in the analysis of equality cases; it is called the potential function
of the domain B, the set Γ, and the tuples {zk }nk=1 and {δk }nk=1 .
Theorem 7.21. Let D, Kz , Kw , Γ, Γ , {zk }nk=1 , {δk }nk=1 , and f be the domains,
sets, tuples, and function defined above. Then the following results hold:
1) if D = Kz and Γ ⊂ Γ then

n
n
σ(zk )δk δl [gKw (f (zk ), f (zl ), f (Γ)) − gKz (zk , zl , Γ )]  0; (7.45)
k=1 l=1

the reverse inequality



n
n
σ(zk )δk δl [gKw (f (zk ), f (zl ), f (Γ)) − gKz (zk , zl , Γ )]  0 (7.46)
k=1 l=1

holds for Γ ⊃ Γ , with equality attained only for Γ = Γ ;


 
n
2) if (∂D) ∩ Kz ) ∩ Γ ∪ ( {zk }) = ∅ and Γ = Γ , then (7.45) also holds,
k=1
with equality sign holding if and only if D = Kz and (∂D) ∩ Kz consists of
finitely many piecewise smooth curves such that the potential function u(z)
of the sets Kz , Γ and tuples {zk }nk=1 and {δk }nk=1 satisfies ∂u/∂n = 0 at the
interior points of these curves;
7.4. Variational principles of conformal mappings 241

3) if (∂D) ∩ Kz ⊂ Γ and Γ = (∂Kz )\((∂D)\Γ), then inequality (7.46) holds


with equality sign if and only if D = Kz and the potential function of the sets
Kz , Γ and tuples {zk }nk=1 , {δk }nk=1 vanishes on Γ ∩ Kz .
Proof. All the three assertions of the theorem follow in a unified fashion from the
monotonicity of the capacity (Theorems 1.15 and 1.16), Theorem 2.2, and the
uniqueness results in Appendix A4. In fact, in case 1) we see from Theorem 1.15
that
cap C(r; Kz , Γ , Z, , Ψ)  cap C(r; D, Γ, Z, , Ψ),
where Z = {zk }nk=1 , = {δk }nk=0 , δ0 = 0, Ψ = {r, . . . , r}, and r > 0 is sufficiently
small. Using Theorem 2.2 we obtain

n
n
n
σ(zk )δk2 log r(Kz , Γ , zk ) + σ(zk )δk δl gKz (zk , zl , Γ )
k=1 k=1 l=1
l=k

n n n
 σ(zk )δk2 log r(D, Γ, zk ) + σ(zk )δk δl gD (zk , zl , Γ),
k=1 k=1 l=1
l=k

and in view of (7.44), this is equivalent to (7.45).


Now if equality sign hold in (7.45), then the corresponding reduced moduli
are equal:
M (Kz , Γ , Z, , Ψ) = M (D, Γ, Z, , Ψ)
and the sets Γ and Γ coincide by Theorem A5. Cases 2) and 3) are similar, but
we must use other monotonicity results for capacities and other conditions for the
equality of reduced moduli. 
Taking a particular canonical domain for Kz in Theorem 7.21 we obtain
variational principles for the corresponding function classes. For example, let f ∈
V1 , D(f ) ⊂ z , and let Γ be a subset of the unit circle |z| = 1 consisting of finitely
many arcs such that Γ ⊂ ∂D(f ). Part 2) of Theorem 7.21 yields a distortion
theorem in which the absolute values |f  (zk )/f (zk )| find estimates in terms of the
inner radii and Green’s functions of the domains Cz \Γ and Cw \f (Γ). In particular,
if in addition we have n = 1, z1 = ∞, and δ1 = 1, then inequality (7.45) (with
Γ = Γ ) coincides with the Komatu–Pommerenke inequality

|f  (∞)| cap Γ  cap f (Γ).

When the arc Γ shrinks to a boundary point z at which the derivative f  (z) exists,
the above inequality yields

|f  (∞)|  |f  (z)|,

which refines the classical principle mentioned in the beginning of § 7.4. This prin-
ciple coincides with the boundary Schwarz lemma.
242 Chapter 7. Univalent Functions

Now let f ∈ V2 and D(f ) ⊂ Hz . Then f can be regarded as the complex


potential of an infinitely deep stream over a flat bottom, past an obstacle Hz \D(f ),
with hydrodynamic normalization which means that the flow is unperturbed at
infinity.
Corollary 7.6. If f ∈ V2 , D(f ) ⊂ Hz , and Γ is a closed bounded subset of ∂D(f ),
Γ ∩ Hz \ D(f ) = ∅, then
cap f (Γ)  cap Γ.
Proof. This follows from Theorem 7.21 (case 2, n = 1, z1 = ∞), once we have
observed that the Robin function with pole at infinity of the upper half-plane (with
respect to Γ (or f (Γ))) is the restriction of the Green function of Cz \Γ (Cw \f (Γ))
with pole at z = ∞ to this half-plane. The above inequality contains the classical
bound f  (z)  1, which holds for all points z on the boundary of D(f ) apart from
the obstacle Hz \D(f ) [LaSha], § 61. 
Corollary 7.7. Let f ∈ V2 , D(f ) ⊂ Hz and let β1 and β2 , β1  β2 . be real numbers.
If [β1 , β2 ] ∩ Hz \ D(f ) = ∅, then for any point z on the real axis apart from the
set [β1 , β2 ] ∪ (Hz \D(f )),
f  (z)Φ(z; β1 , β2 )  Φ(f (z); f (β1 ), f (β2 )),
  −4
where Φ(ζ; s1 , s2 ) = (ζ − s1 )(ζ − s2 ) |ζ − s1 | + |ζ − s2 | and the positive
values of the root function are taken. On the other hand, if the interval [β1 , β2 ]
contains all the points in Hz \D(f ) lying on the real axis, then the reverse inequality
holds for each point z on the real axis lying outside this interval.
Proof. Assume that the first condition in the statement holds. Then we set D =
D(f ), Kz = Hz , Kw = Hw and Γ = Γ = [β1 , β2 ] in Theorem 7.21 and take
{z, ∞} to be the set of points and {−1, 1} to be the number tuple. By the second
assertion of the theorem we obtain (7.45). To calculate the Robin functions on
the left-hand side of (7.45) we use the conformal invariance of these functions. A
suitably chosen branch of the root function
C
z − s2
ζ= (s1 < s2 )
z − s1
maps the half-plane Hz conformally and univalently onto the coordinate sector
Ω = {ζ : Re ζ > 0, Im ζ > 0} so that the interval [s1 , s2 ] is taken to the ray
 = {ζ : Re ζ = 0, Im ζ  0}. By symmetry
Γ
 
 
 ζ + ζ 

gΩ (ζ, ζ , Γ) = log  
ζ − ζ 
for positive ζ  . Setting first sk = βk , k = 1, 2, and then sk = f (βk ), k = 1, 2,
we easily find all the parameters of the left-hand side of (7.45). The proof of the
second part of Corollary 7.7 proceeds in a similar way, but uses the third assertion
of Theorem 7.21. 
7.4. Variational principles of conformal mappings 243

We can prove similar corollaries for other function classes defined at the
beginning of this section. Some of these results can be stated equivalently in other
function classes thanks to the existence of an appropriate conformal mapping. For
example, for functions in V3 , that is, for the complex potentials of flows in curved
channels ([LaSha], §§ 48 and 62) we have the following result.
Corollary 7.8. Let f ∈ V3 and D(f ) ⊂ Sz . If an interval [β1 , β2 ] lies in the set
(∂D(f )) ∩ (∂Sz ) and [β1 , β2 ] ∩ Sz \ D(f ) = ∅, then for each finite point z lying on
the boundary ∂Sz apart from the set [β1 , β2 ] ∪ (Sz \ D(f )),

f  (z)eπ(f (z)−z) Φ(eπz ; eπβ1 , eπβ2 )  Φ(eπf (z) ; eπf (β1 ) , eπf (β2 ) ).

On the other hand, if [β1 , β2 ] contains all the points in Sz \ D(f ) that lie on the
boundary ∂Sz , then the reverse inequality holds for each finite point z ∈ ∂Sz lying
apart from [β1 , β2 ].

For appropriately chosen parameters Theorem 7.21 can be treated as a dis-


tortion theorem in some classes of univalent functions. For example, the following
result holds.
Corollary 7.9. (Nehari [Neh]) If f is a function in the class SB and zk , k =
1, . . . , n, are points in U , then
  δk δl   
n n n
 zk − zl  n n  1 − f (z )f (z ) δk δl
 2
   k l 
|f (zk )| δk
  f (zk ) − f (zl )   
 1 − zk z l 
k=1 k=1 l=1 k=1 l=1
l=k

for any real δk , k = 1, . . . , n.

Proof. We set h(w) = f −1 (w), wk = f (zk ), D = f (U ), Kz = Kw = U , Γ = ∂D,


and Γ = ∂U . Applying assertion 3) of Theorem 7.21 to h and the points wk , k =
1, . . . , n, we obtain


n
n
n
n
n
− δk2 log |h (wk )| + δk δl gU (h(wk ), h(wl ))  δk δl gU (wk , wl ).
k=1 k=1 l=1 k=1 l=1
l=k l=k

It remains to observe that


 
 1 − ζz 
gU (z, ζ) = log  . 
z−ζ 

Theorem 7.21 can be supplemented with a result on the case Γ = ∅, where


we shall use Neumann functions in place of Robin functions. However, there is also
another approach to deducing estimates of this type. We consider it in a special
case and state the result as a distortion theorem.
244 Chapter 7. Univalent Functions

Theorem 7.22. Let f be a function in SB such that f and its derivative are also
defined at some boundary points zk and |zk | = |f (zk )| = 1, k = 1, . . . , l. Let zk
with k = l + 1, . . . , n be points in the disc U . If δk , k = 1, . . . , n, are real numbers
such that
l n
δk + 2 δk = 0,
k=1 k=l+1

then
 l
 n
2 2
|f  (zk )| δk
|f  (zk )|2δk
k=1 k=l+1
  θks δk δs   2δk δs
n n
 zk − zs  n n  1−z z 
   k s 
  f (zk ) − f (zs )    ,
 1 − f (zk )f (zs ) 
k=1 s=1 k=1 s=l+1
s=k

where θks = 2 if both k  l + 1 and s  l + 1 and θks = 1 otherwise.



n
Proof. We can assume that δk2 = 0 and zk = 0 f (zk ) = 0, k = 1, . . . , n.
k=1
Note also that f  (zk ) = 0, k = 1, . . . , l: this is easy to see, for instance, by
applying the boundary Schwarz lemma to the composite of f and a suitable Möbius
automorphism of the unit disc U . We look at mirror-symmetric condensers of the
form

C(r) = (Cz , {E(z1 , r), . . . , E(zn , r),


2
E(1/zl+1 , r/|zl+1 |), . . . , E(1/zn , r/|zn2 |)},
{δ1 , . . . , δn , δl+1 , . . . , δn })

and

C(r) = (Cw , {E(f (z1 ), r|f  (z1 )|), . . . , E(f (zn ), r|f  (zn )|), E(1/f (zl+1 ),
r|f  (zl+1 )/f 2 (zl+1 )|), . . . , E(1/f (zn ), r|f  (zn )/f 2 (zn )|)},
{δ1 , . . . , δn , δl+1 , . . . , δn }).

(Here E(z, r) denotes the closure of the almost disc D(z, r) in § 2.2.)
It easily follows from the geometric interpretation of the derivative that we
can take almost discs such that the assumptions of Exercise 1.3(8) are satisfied.
Then we have

cap C(r)  cap C(r)
for sufficiently small r > 0. It remains to use asymptotic formula (2.15). 
The set of boundary points zk in Theorem 7.22 can be empty. Then in the

l 1
l 2
statement of the theorem we set l = 0, δk = 0, |f  (zk )|δk = 1, and θks = 2,
k=1 k=1
7.4. Variational principles of conformal mappings 245

k, s = 1, . . . , n. The proof remains formally the same, but the notation is simpler
because the corresponding plates of condensers are not involved.
By the classical variational principles, in the case of an inner deformation of
a domain the dilation decreases at boundary points fixed by the deformation. It
is natural to conjecture that, for example, at points more distant from the defor-
mation locus the changes in dilation will be less profound. The comparison of the
dilations at different points for a fixed deformation can sometimes be interpreted
as the comparison of deformations moving away from a fixed point. Moreover,
comparing deformations in their impact on the dilation and other characteristics
of a function is a question of independent interest. Here we discuss some of these
questions for the function classes introduced above.
Theorem 7.23. Let f be a function in the class V2 or V3 , and let Kz = Hz in
the first case and Kz = Sz in the second, let D = D(f ), and let β be a real
number. Also assume that the interior part of the deformation Kz \D lies in the
half-plane Re z < β and its exterior D\Kz lies in the half-plane Re z > β. Let
Γ be a closed subset of ∂D consisting of finitely many nondegenerate connected
components with the following property: Γ ⊃ {∂[(Kz \D) ∪ (D\Kz )]}\∂Kz , and if
z ∈ Γ ∩ ∂Kz , Re z  β, then −z + 2β ∈ Γ ∪ (Kz \ D). Let {zk }nk=1 be a set of
different admissible points in D\Γ and {δk }nk=1 a tuple of positive numbers. Then


n
σ(zk )δk δl [gKw (f (zk ), f (zl ), f (Γ)) − gKw (f (zk ), f (zl ), f (Γ))]  0, 9 (7.47)
k,l=1

where the points zk , k = 1, . . . , n, are defined as follows: if a point zk in the set
{zk }nk=1 has a symmetric point zl ∈ {zk }nk=1 relative to the line Re z = β, then
zk = zk in the case when Re zk  β and δk  δl or Re zk  β and δk  δl , and
zk = zl in the case when Re zk  β and δk < δl or Re zk  β and δk > δl ; on the
other hand, if the symmetric point −zk + 2β does not belong to the set {zk }nk=1 ,
then zk = zk for Re zk  β and zk = −zk + 2β for Re zk  β.

Proof. We set Z = {zk }nk=1 , Z  = {zk }nk=1 , = {δk }nk=0 , δ0 = 0, and Ψ =


{r, . . . , r}. By Theorem 3.7

M (D, Γ, Z, , Ψ)  M (D, Γ, Z  , , Ψ),

and taking account of formula (2.20) for the reduced moduli we obtain


n
n
σ(zk )δk δl [gD (zk , zl , Γ) − gD (zk , zl , Γ)]  0.
k=1 l=1

By the conformal invariance this is equivalent to (7.47). 


9 If k = l, then we treat the difference of Robin functions as the corresponding limit.
246 Chapter 7. Univalent Functions

Following the methods presented in Appendix A4 we can show that if the


deformation is nontrivial then equality can be attained in (7.47) only for zk = zk ,
k = 1, . . . , n. We explain (7.47) by looking at several simple examples.
Corollary 7.10. If f ∈ V2 , D(f ) ⊂ Hz , and Hz \D(f ) lies in the half-plane Re z <
σ, then for any points z1 and z2 on the real axis such that σ < z1 < z2 ,

f  (z1 ) f  (z2 )
 .
f (z1 ) − f (σ) f (z2 ) − f (σ)

Proof. In Theorem 7.23 we set β = (z1 + z2 )/2, and let Γ be the part of the
boundary ∂D(f ) lying between the points ∞ and σ in the sense of the positive
orientation of the boundary, let n = 1, let {z1 } be the set of points and let δ1 = 1.
Then z2 = z1 and inequality (7.47) gives us

gHw (f (z2 ), f (z2 ), [−∞, f (σ)]) − gHw (f (z1 ), f (z1 ), [−∞, f (σ)])  0,

where [−∞, f (σ)] is an interval of the real axis. A direct calculation of the Robin
capacities yields the required relation. 
Corollary 7.11. Let f ∈ V2 , D(f ) ⊂ Hz and assume that Hz \D(f ) lies in the left
half-plane Re z < 0. Then for any positive z1 , z2 , and y, where z1 < z2 ,

Im f (z1 + iy)  Im f (z2 + iy),

f  (z1 )  f  (z2 ).

Proof. In Theorem 7.23 we set β = (z1 + z2 )/2, Γ = ∂D(f ), and take the set
of points {z1 + iy, z ∗ }, where z ∗ = (z1 + z2 )/2 + it, and the number tuple {δ, 1}
(t > 0, δ > 0). From (7.47) we see that

δgHw (f (z1 + iy), f (z ∗ ), f (Γ)) + δ 2 log r(D(f ), Γ, z1 + iy)


 δgHw (f (z2 + iy), f (z ∗ ), f (Γ)) + δ 2 log r(D(f ), Γ, z2 + iy).

Dividing by δ and passing to the limit as δ → 0 we obtain

gHw (f (z1 + iy), f (z ∗), f (Γ))  gHw (f (z2 + iy), f (z ∗ ), f (Γ))

of equivalently,
   
 f (z + iy) − f (z ∗ )   f (z + iy) − f (z ∗ ) 
 1   2 
  .
 f (z1 + iy) − f (z ∗ )   f (z2 + iy) − f (z ∗ ) 

Hence, taking account of the asymptotic expansion of f in a neighbourhood of


infinity we obtain the first inequality. The second follows from the first, which
completes the proof. 
7.4. Variational principles of conformal mappings 247

Thus under the assumptions of Corollary 7.11 (and in terms of the complex
potential) whatever shape the obstacle Hz \D(f ) may have, stream curves go down
monotonically to the right of the imaginary axis and the velocity of the flow is
nondecreasing along the real axis (Figure 7.2). Furthermore, using Hopf’s lemma
and the proof of Corollary 7.11 we can show that the velocity is monotonically
increasing on the positive half-axis.

2 {3 {4

Figure 7.2.

Using circular and Steiner symmetrization we obtain similar variational prin-


ciples for functions in the classes V1 −V3 and V5 . Now we give another kind of exam-
ple. If f ∈ V4 (l), then D(f ) can be interpreted, for instance, as a filtration domain,
f /L(f ) will be the complex potential of the filtration flow and 1/L(f ) the flow
discharge. We look at the condenser C(f ) = (D(f ), {(a4 , a1 ), (a2 , a3 )}, {0, 1}).
As the condenser capacity is monotonic, if the side (a1 , a2 ) coincides with a line
interval (l, l + i) and (a3 , a4 ) coincides with (i, 0), then for D(f ) ⊂ Qz (l, l + i, i, 0)
we have L(f )  l, while for D(f ) ⊃ Qz (l, l + i, i, 0) we have L(f )  l. In the next
theorem we show that inner deformations of a quadrilateral affect the quantity
L(f ) stronger than outer deformations of the same form do.
Theorem 7.24. If f ∈ V4 (l), D(f ) ⊂ Π := {z : 0  Re z  l}, and the intervals
(l, l + i) and (i, 0) are two sides of the quadrilateral D(f ), while the other two sides
are parallel translations of each other in the direction of the imaginary axis, then
l  L(f ).
Now let fk ∈ V4 (l), D(fk ) ⊂ Π, k = 1, 2, and assume that each of the quadri-
laterals D(f1 ) and D(f2 ) has sides (l, l + i) and (i, 0). Assume that the other two
sides of D(f1 ) are curves γ1 and γ2 such that 0 ∈ γ1 and l ∈ γ1 and the sides
of D(f2 ) are obtained from those of D(f1 ) by reflecting γ1 in the real axis and
reflecting γ2 in the line Im z = 1. Then l  (L(f1 ) + L(f2 ))/2.
Finally, if f ∈ V4 (l), D(f ) ⊂ Π, the intervals (l, l + i) and (i, 0) are sides of
the quadrilateral D(f ) and the other two sides are centrally symmetric (namely,
(a4 , a1 ) is symmetric in the point l/2 and (a2 , a3 ) is symmetric in l/2 + i), then
l  L(f ).
248 Chapter 7. Univalent Functions

Proof. The first assertion follows from Theorem 5.14. The second follows from
Theorem 3.12, where we must set n = 2 and α1 = α2 = 1/2. Finally, the third
assertion is obtained by using in turns the composition principle (Figure 1.4, the
left-hand part) and Theorem 3.12 again. In applying the composition principle we
decompose C(f ) by the line Re z = l/2 into two condensers of a similar form. 

Close results for functions in V1 and V5 can be established with the use of
the radial averaging transformation from § 3.5.

Exercises 7.4
(1) Supplement Theorem 7.21 by considering the case Γ = ∅. In this case one
assumes that
n
σ(zk )βk = 0
k=1

and takes Neumann functions in place of Robin functions.


(2) Let f be a function in SB0 such that f and its derivative are also defined and
continuous at some boundary points zk , k = 1, . . . , n, which are positioned
n-symmetrically on the unit circle. Let |f (zk )| = 1, k = 1, . . . , n.
Prove the inequality
 
 n 
  2
 f  (zk )  |1/f  (0)|n /2  1.
 
k=1

(3) (cf. [AnVas]) Prove that if f is a mapping in the class SB0 which is conformal
at boundary points z1 , . . . , zn and f (zk ) = zk , k = 1, . . . , n, then for any real

n
numbers αk , k = 1, . . . , n, such that αk = 1,
k=1

n
2
|f  (0)|  |f  (zk )|−2αk .
k=1

(4) Establish the following analogue of Nehari’s inequality for functions on an


annulus. Let f be a function in the class M(R, M ), γ be a nonempty closed
subset of ∂K(R) consisting of finitely many nondegenerate Jordan arcs, and
let Γ be a subset of ∂K(M ) with the same structure. Let zk , k = 1, . . . , n,
be points in K(R) \ γ, and for |zk | = R assume in addition that M =
∞, |f (zk )| = M , and f is conformal at the zk , 1  k  n. Let f (γ) ⊂ Γ,
where f (γ) is defined by means of boundary correspondence. Then for any
7.4. Variational principles of conformal mappings 249

real numbers δk , k = 1, . . . , n,

n
n
n
σ(zk )δk2 
log(r(K(R), γ, zk )|f (zk )|) + σ(zk )δk δl gK(R) (zk , zl , γ)
k=1 k=1 l=1
l=k

n
 σ(zk )δk2 log r(K(M ), Γ, f (zk )) (7.48)
k=1
n
n
+ σ(zk )δk δl gK(M) (f (zk ), f (zl ), Γ).
k=1 l=1
l=k


n
For δk2 = 0 equality in (7.48) holds if and only if f (K(R)) = K(M ),
k=1
f (γ) = Γ, and the set K(M ) ∩ ∂f (K(R)) consists of finitely many piecewise
smooth curves such that the normal derivative of the function

n
δk gK(M) (w, f (zk ), Γ) (7.49)
k=1

vanishes at the interior points of these curves. If f ((∂K(R))\γ) ⊂ (∂K(M ))\


Γ, then the reverse inequality to (7.48) holds for any real δk k = 1, . . . , n.
n
Furthermore, for δk2 = 0 equality holds if and only if f (K(R)) = K(M ),
k=1
f (γ) ∩ ∂K(M ) = Γ, and the function in (7.49) vanishes on the set K(M ) ∩
f (γ).
(5) Let f be a function in the class M(R, M ), M = ∞, which is also defined at
a boundary point z0 with |z0 | = R, let |f (z0 )| = M and assume that f is
conformal at z0 . Show that for each closed subset γ of the unit circle |z| = 1
consisting of finitely many nondegenerate Jordan arcs

r(K(M ), f (γ), f (z0 ))


|f  (z0 )|  ,
r(K(R), γ, z0 )

where equality sign holds in the case when f maps the annulus K(R) onto
the annulus K(M ) cut along some interval [−M, δ(M, R)], f (1) = 1, z0 = R,
and the set γ is mirror-symmetric relative to the real axis. In the simplest
case of γ = {z : |z| = 1} (so that f (γ) = {w : |w| = 1}) the above inequality
can be written as
r(K(M 2 ), M )
|f  (z0 )|  , |z0 | = R, |f (z0 )| = M.
r(K(R2 ), R)

For functions in the class SB0and points z such that |z0 | = |f (z0 )| = 1,
deduce the bound |f  (z0 )|  1/ |f  (0)| from this inequality.
250 Chapter 7. Univalent Functions

(6) Let γ be a subset of the circle |z| = 1 consisting of finitely many arcs, let
Z = {zk }nk=1 be a tuple of points zk ∈ U , k = 1, . . . , n, and = {δk }nk=1 a

n
tuple of real numbers δk , k = 1, . . . , n, such that δk2 = 0. For fixed γ, Z,
k=1
and set
2n
2n
2n
H(γ, Z, ) = δk2 log r(C \ γ, zk ) + δk δl gC\γ (zk , zl ),
k=1 k=1 l=1
l=k

where δn+k = δk and zn+k = 1/zk , k = 1, . . . , n. Similarly to the proof of


Theorem 7.22 prove the following generalization of the well-known Löwner
lemma to univalent functions. Let f be a function in the class SB and assume
that, as z approaches a set γ which consists of finitely many open subarcs of
the circle |z| = 1, we have |f (z)| → 1. Then for any set of points Z and tuple
of numbers


n  
  zk f  (zk ) 
2 δk2 log    H(f (γ), W, ) − H(γ, Z, ),
f (zk ) 
k=1

where W = {f (zk )}nk=1 , f (γ) is the limit set of f (z) as z → γ, and the dash
over a summation sign means that in the case of zk = 0 (f (zk ) = 0) the
corresponding factor is set equal to 1.

7.5 Behaviour of level curves


Methods based on symmetrization of condensers reveal some simple properties of
level curves of univalent functions in the classes introduced above. The reader can
establish some of these properties on his own, by making Exercises 7.5. We shall
consider only one general trick based on dissymmetrization, which enables us, in
particular, to extend the results in § 7.2 to coverings of line segments by the images
of subsets of a disc or annulus. The statement of the corresponding problem about
the behaviour of level curves goes back to Bazilevich. Let Λf (r, θ), where f ∈ M0
and 0 < r  1, denote the distance from the origin to the closest boundary point
of the set f ({z : |z| < r}) that lies on the ray arg w = θ. If there exist no such
point for fixed r and θ, then we set Λf (r, θ) = +∞ (Λf (1, θ) is equal to Λf (θ)
introduced in § 7.2). Let L(r, f ) be the f -image of the circle |z| = r, 0 < r < 1.
The curve L(r, f ) is usually called a level curve of f . Thus, if the level curve L(r, f )
intersects the ray arg w = θ, then Λf (r, θ) is the distance from the origin to the
closest point of intersection of L(r, f ) and this ray.
The statement below reduces estimates for the Λf (r, θ) to estimates for the
Λf (1, θ). Let Φ = Φ(x1 , . . . , xn , y1 , . . . , ym ) be a real function which is nondecreas-
ing in the variables yl and homogeneous on the set xk > 0, yl > 0, k = 1, . . . , n,
7.5. Behaviour of level curves 251

l = 1, . . . , m; more precisely,

Φ(αx1 , . . . , αxn , αy1 , . . . , αym ) = αn Φ(x1 , . . . , xn , y1 , . . . , ym )

for all possible α > 0. Let N be a subclass of functions in M0 which satisfies
the following condition: if f ∈ N, then for each conformal homeomorphism g
from U into itself with g(0) = 0 and g  (0) > 0 the function f (g(z))/g  (0) also
belongs to the class N. For fixed ϕk and ψl , where k = 1, . . . , n and l = 1, . . . , m,
and for 0 < r  1, we consider the functional on the class N given by A(f ) ≡
A(f ; r, ϕ1 , . . . , ϕn , ψ1 , . . . , ψm ) = Φ(Λf (r, ϕ1 ), . . . , Λf (r, ϕn ), Λf (1, ψ1 ),
. . . , Λf (1, ψm )). Then the following result holds.
Lemma 7.3. If
A(f ; 1, ϕ1 , . . . , ϕn , ψ1 , . . . , ψm )  1 (7.50)
in the class N, then also

A(f ; r, ϕ1 , . . . , ϕn , ψ1 , . . . , ψm )  4rn (1 + rn )−2 (7.51)

for any r, 0 < r < 1 and f ∈ N. If equality sign holds in (7.51), then fD(z) =
f (eiβ hn (e−iβ z))/hn (0) is an extremal function in (7.50), were β is a real number
and the function z = hn (ζ) with hn (0) = 0 and hn (0) > 0 maps the disc |ζ| < 1
conformally and univalently onto the disc |z| < 1 cut along the segments of radii
[r exp(2πik/n), exp(2πik/n)], k = 1, . . . , n.
Proof. Fix r with 0 < r < 1 and f ∈ N. Let z = g(ζ), g(0) = 0, g  (0) > 0, be
a function mapping the disc |ζ| < 1 conformally and univalently onto the disc
|z| < 1 cut along the segments of radii [r exp(iθk ), exp(iθk )], where

θk = arg f −1 (Λf (r, ϕk ) exp(iϕk )), k = 1, . . . , n.

(some of the θk can coincide). Then the composite function

F (ζ) = f (g(ζ))/g  (0)

belongs to the class N, and we have

ΛF (1, ϕk ) = Λf (r, ϕk )/g  (0), k = 1, . . . , n,


ΛF (1, ψl )  Λf (1, ψl )/g  (0), l = 1, . . . , m.

In succession, from (7.50) and the properties of Φ we obtain

1  A(F ; 1, ϕ1 , . . . , ϕn , ψ1 , . . . , ψm )
= Φ(ΛF (1, ϕ1 ), . . . , ΛF (1, ϕn ), ΛF (1, ψ1 ), . . . , ΛF (1, ψm ))
 Φ(Λf (r, ϕ1 )/g  (0), . . . , Λf (r, ϕn )/g  (0), Λf (1, ψ1 )/g  (0), . . . , Λf (1, ψm )/g  (0))
= (g  (0))−n Φ(Λf (r, ϕ1 ), . . . , Λf (r, ϕn ), Λf (1, ψ1 ), . . . , Λf (1, ψm ))
= (g  (0))−n A(f ; r, ϕ1 , . . . , ϕn , ψ1 , . . . , ψm ).
252 Chapter 7. Univalent Functions

It remains to use Theorem 4.16, which states that



g  (0)  hn (0) = 4r(1 + rn )−2/n
n

with equality sign only for g(ζ) = eiβ hn (e−iβ ζ), where β is a real number. 

Theorem 7.25. For each function f in M0 and any θ and r, where 0 < r < 1,
>
?
? n
@
n
Λf (r, θ + 2πk/n)  r(1 + rn )−2/n , (7.52)
k=1

with a possible exception of the case when n = 1 and the domain f (U ) covers the
whole of the ray {−ρeiθ : 0  ρ  ∞}. Equality sign holds in (7.52) only for the
function f (z; θ, n) := z[1 + (e−iθ z)n ]−2/n .

Proof. For n = 1 let N be the subclass of functions in M0 not covering the whole
of the ray {−ρeiθ : 0  ρ  ∞}, and for n > 1 let N = M0 . In either case we
1n
set Φ = 4 xk and ϕk = θ + 2πk/n, k = 1, . . . , n. Note that for r = 1 (7.52)
k=1
coincides with the already established inequality (7.25). Thus we are under the
assumptions of Lemma 7.3, which states that (7.52) holds for any r, 0 < r < 1.
Here equality sign is only possible for

f (eiβ hn (e−iβ ζ))/h (0) ≡ f (ζ; θ, n).

As f is conformal in U , we have

f (z) ≡ hn (0)f (eiβ h−1


n (e
−iβ
z); θ, n) ≡ f (z; θ, n).

A direct verification shows that, indeed, for f (z; θ, n) we have equality sign in
(7.52). 

Corollary 7.12. If f ∈ S, 0 < r < 1, and if zk , k = 1, . . . , n, are points in the disc


U such that |zk | = r and arg(f (zk )/f (z1 )) = 2π(k − 1)/n, k = 1, . . . , n, then
>
?
? n
@
n
|f  (zk )|  (1 − rn )/(1 + rn )1+2/n .
k=1

Here equality holds for f (z) = f (z; θ, n) and the points zk = r exp(i(θ + 2πk/n)),
k = 1, . . . , n.

This is an immediate consequence of Theorems 7.14 and 7.25. On the other


hand, both this result and Theorem 7.25 are consequences of Theorem 7.14.
Let f (z; θ, c, n) be the function from Theorem 7.11.
7.5. Behaviour of level curves 253

Theorem 7.26. If f ∈ M0 , then for any θ, r with 0 < r < 1, and n  2
 n  −2  −2
n 
1 n2

n
2π 1 π 2π
Λf r, θ + k + 2
Λf r, θ + + k  2 rn + r−n .
n n n n n
k=1 k=1
(7.53)
Equality sign
 holds only for f (z) = f (z; θ, c, n), where c is an arbitrary real number
satisfying n 1/4  c  ∞.
Proof. We set N = M0 , ϕj = θ + πj/n for j = 1, . . . , 2n, and
⎧ −2  n −2 ⎫−2
⎨ 1 n n
1 n ⎬
Φ = 16 2
x2k + 2
x2k−1 .
⎩ n n ⎭
k=1 k=1

Theorem 7.11 (r = 1) and Lemma 7.3 yield inequality (7.53) for 0 < r < 1.
Equality sign in (7.53)is only possible if f (z) = h2n (0)f (eiβ h−1
2n (e
−iβ
z); θ, c , n)
for some c such that n 1/4  c  ∞. Here β and h2n (ζ) are as in the statement
of
 Lemma 7.3. Since f is conformal in U , we have f (z) = f (z; θ, c , n),  where
n
1/4  c  ∞. A direct verification shows that for each c such that n 1/4 
c  ∞ and any r such that 0 < r  1 we have equality sign in (7.53) for the
functions f (z; θ, c, n). 
Theorem 7.27. If f ∈ M0 , then for any θ, any r such that 0 < r < 1, and any
n2
 n  −2  −2
n 
1 n2
n
2π 1 π 2π
Λf r, θ + k + Λf2 1, θ + + k  rn + 2 + r−n .
n n n n n
k=1 k=1

Equality sign
 holds only for f (z) = f (z; θ, c, n), where c is an arbitrary real number
satisfying n 1/4  c  ∞.
Proof. For r = 1 the inequality in Theorem 7.27 coincides with (7.27). We set
N = M0 , ϕk = θ + 2πk/n for k = 1, . . . , n, ψl = θ + π/n + 2πl/n for l = 1, . . . , n,
and ⎧ −2  n −2 ⎫−1
⎨ 1 n
1 ⎬
n/2 n/2
Φ=4 xk + yl
⎩ n n ⎭
k=1 l=1

and use Lemma 7.3 again. The equality case is considered in a similar way to the
theorems above. 
The above approach can easily be extended to bounded functions and to
functions in an annulus. For example, let f ∈ M(R) and let Xf (r, θ) be the distance
form the origin to the closest boundary point of the set f ({z : 1 < |z| < r}) lying
on the ray arg w = θ, 1 < |w| < ∞. If no such point exists for some r and θ, then
we set Xf (r, θ) = +∞.
254 Chapter 7. Univalent Functions

Yg )s< 2*
3 s S

Figure 7.3.

Theorem 7.28. If f ∈ M(R), then for any θ and r such that 1 < r < R,
>
? n  n 
? 2π n
−2 2π 
@n 2
Xf (r, θ + k) − Xf (r, θ + k)  (G(rn ; Rn ) − 1)/ G(rn ; Rn ),
n n
k=1

with the possible exception of the case when n = 1 and the whole of the ray arg w =
θ + π, 1 < |w|  ∞ lies in the domain f (K). Equality sign holds if and only if
f (z) ≡ eiθ G(αz; n, R), where α is an arbitrary number such that |α| = 1.
Proof. Let f ∈ M(R) and 1 < r < R. Let θk denote the argument of the point
f −1 (Xf (r, θ + 2πk/n) exp(i(θ + 2πk/n))), k = 1, . . . , n. Let z = hn (ζ), (hn (1) =
exp(iθn )) be a function mapping some annulus 1 < |ζ| < Qn (r) conformally and

n
univalently onto the domain G = K \ [r exp(iθk ), R exp(iθk )] and z = gn (ζ)
k=1
(gn (1) = 1) be a similar function mapping an annulus 1 < |ζ| < Rn (r) onto G∗ =

n
K \ [r exp(2πik/n), R exp(2πik/n)]. We can regard the condenser corresponding
k=1
to G as the result of dissymmetrization in Lemma 4.2 applied to the condenser
with field G∗ . By the dissymmetrization principle (Theorem 4.14)
Qn (r)  Rn (r).
The composite function F (ζ) = f (hn (ζ)) belongs to the class M(Qn (r)), and we
have
XF (Qn (r), θ + 2πk/n) = Xf (r, θ + 2πk/n), k = 1, . . . , n.
By Theorem 7.13
>
? n  
? 2π 2π
@n n/2
Xf (r, θ + k) − Xf
−n/2
(r, θ + k)
n n
k=1

 (G(Qnn (r); Qnn (r)) − 1) / G(Qnn (r); Qnn (r))
 
 (G(Rnn (r); Rnn (r)) − 1) / G(Rnn (r); Rnn (r)) = (G(rn ; Rn ) − 1) / G(rn ; Rn ).
7.5. Behaviour of level curves 255

Here equality sign is only possible for

F (ζ) ≡ eiθ G(αζ; n, Qn (r)), |α| = 1,


Qn (r) = Rn (r), and hn (ζ) ≡ eiθn gn (ζ).

Taking account of normalization we arrive at the identity

F (ζ) ≡ eiθ G(ζ; n, Rn (r)), 1 < |ζ| < Rn (r).

Hence

f (z) ≡ eiθ G(gn−1 (e−iθn z); n, Rn (r)) ≡ eiθ G(e−iθn z; n, R), 1 < |z| < R.

On the other hand, if f (z) ≡ eiθ G(αz; n, R), where |α| = 1, then

αhn (ζ) ≡ gn (ζ), Qn (r) = Rn (r), F (ζ) ≡ eiθ G(gn (ζ); n, R) ≡ eiθ G(ζ; n, Rn (r))

and we have equality sign in all the above relations. 


In a similar way we can prove a result which for r = 1 corresponds to the in-
equality in Exercise 7.2(4). From Theorem 7.28, similarly to the proof of Corollary
5.1 we obtain the following statement.
Corollary 7.13. Let f ∈ M(R). Then for any θ and r such that 1 < r < R,
>
? n
? 
@
n
Xf (r; θ + 2πk/n)  n G(rn ; Rn ),
k=1

with the possible exception of the case when n = 1 and the whole of the ray arg w =
θ + π, 1 < |w|  ∞ lies in the domain f (K). Here equality holds only for the
functions f (z) = eiθ G(αz; n, R), where α is an arbitrary number such that |α| = 1.
In what follows, for more transparent statements we limit ourselves to func-
tions in S. Extremal functions will have the form

−2/n
f (z; θ, n) = z 1 + (e−iθ z)n .

Let wk (r, θ, f ) be an arbitrary point of intersection of the level curve L(r, f ) and
the ray arg w = θ + 2π(k − 1)/n, k = 1, . . . , n.
Theorem 7.29. If f is a function in S, then for any θ, r, and ρ such that 0 < r < 1
and 0 < ρ < 1,
n    
1  wk (r, θ, f )   r(1 + ρn )2/n 
 log  log . (7.54)
n wk (ρ, θ, f )   ρ(1 + rn )2/n 
k=1

Here equality sign holds for f (z) = f (z; θ, n) and arbitrary r and ρ.
256 Chapter 7. Univalent Functions

Proof. We can assume that 0 < r < ρ < 1. We set γk = [wk (r, θ, f ), wk (ρ, θ, f )];
then
n   n     
  |dw|
n
 f (z) 
log wk (r, θ, f )  =   
 wk (ρ, θ, f )  |w|  f (z)  |dz|.
k=1 k=1 γ k=1f −1 (γ )
k k

For each k, 1  k  n, we can easily construct a piecewise continuous function


z = zk (t) on [r, ρ] such that

zk (t) ∈ f −1 (γk ), |zk (t)| = t, r  t  ρ.

Then we have
    ρ   
 f (z)   
  |dz|   f (zk (t))  dt, k = 1, . . . , n.
 f (z)   f (zk (t)) 
f −1 (γk ) r

By Theorem 7.14

n  
ρ  ρ
1  f (zk (t))  1 − tn ρ(1 + rn )2/n
  dt  dt = log
n  f (zk (t))  t(1 + tn ) r(1 + ρn )2/n
r k=1 r

(see also Exercise 7.3(1)). Combining the above relations we obtain (7.54). The
equality case is straightforward. 
Theorem 7.30. If f ∈ S, then for any θ, r and ρ such that 0 < r < 1 and 0 < ρ < 1,
 
1  
n
 r ρ .
|wk (r, θ, f ) − wk (ρ, θ, f )|   n 2/n
− n 2/n 
n (1 + r ) (1 + ρ )
k=1

Here equality sign is attained for f (z) = f (z; θ, n).


Proof. As in the proof of Theorem 7.29, we take 0 < r < ρ < 1 and γk =
[wk (r, θ, f ), wk (ρ, θ, f )]. In this case
 
1 1 1
n n n
|wk (r, θ, f ) − wk (ρ, θ, f )| = |dw|  |f  (z)||dz|.
n n n
k=1 k=1 γ k=1f −1 (γ )
k k

Now we follow the proof of Theorem 7.29, but use Corollary 7.12 in place of
Theorem 7.14. 
7.5. Behaviour of level curves 257

Exercises 7.5
(1) By a level curve of a function f in the class V1 or V5 (see § 7.4) authors usually
mean a closed analytic curve Cρ (f ) defined by the equation |f (z)| = ρ, where
ρ > 1 in the first case and 1 < ρ < R in the second. Let z = p(w; α) with
0 < α  1 denote the Pick function given by the equation
z αw
= , w ∈ w ,
(1 − z)2 (1 − w)2
which maps the domain w onto the domain z cut along the interval
[p(−1; α), −1]. Using circular symmetrization show that if f ∈ V1 , f  (∞) =
α > 0, and D(f ) ⊂ z , then for each ρ > 1 the level curve Cρ (f ) lies in the
annulus
p(ρ; α)  |z|  −p(−ρ; α).
For ρ > 1 the curve Cρ (f ) can only meet the boundary of this annulus when
p(e−iϕ f (zeiϕ ); α) ≡ z for some real ϕ.
(2) For fixed R1 and R2 , 1 < R1  R2 < ∞, let z = r(w; R1 , R2 ) denote the
function mapping the annulus Kw (R1 ) conformally and univalently onto the
annulus Kz (R2 ) cut along the interval [−R2 , r(−R1 ; R1 , R2 )] of the real axis.
Let f ∈ V5 and let D(f ) ⊂ Kz (R), where R(f )  R. Prove the following
result: for 1 < ρ < R(f ) the level curve Cρ (f ) lies in the set
|z|  −r(−ρ; R(f ), R). (7.55)
In addition, if one of the boundary components of D(f ) is equal to the circle
|z| = 1, then the level curve Cρ (f ) also lies in the disc
|z|  r(ρ; R(f ), R). (7.56)
In the cases under consideration a curve Cρ (f ) with 1 < ρ < R(f ) can touch
the boundary of the corresponding

set (7.55) or (7.56) only when for some
real ϕ1 and ϕ2 we have r eiϕ2 f (zeiϕ1 ); R(f ), R ≡ z.
(3) If f is a function in the class V2 or V3 , then a level curve Cv (f ) of f is defined
as the curve given by the equation Im f (z) = v, with v > 0 in the first case
and 0 < v < 1 in the second. Show that if f ∈ V2 , f (z) = z + c/z + o(1/z)
as z → ∞, (z ∈ D(f )), and D(f ) ⊂ Hz , then for each v > 0 the level curve
Cv (f ) lies in the strip 
v  Im z  v 2 + 2c.
For v > 0 the level curve Cv (f ) can touch the bottom of the strip √ only if
f (z) ≡ z (c = 0). It can touch the top of the strip
√ only if f (z) = z 2 + 2c
with c  0 (and then the point of tangency is i 2c).
(4) Let f ∈ V3 , D(f ) ⊂ Sz , and let
lim (f (z) − z) − lim (f (z) − z) = a.
z→+∞ z→−∞
z∈D z∈D
258 Chapter 7. Univalent Functions

Using Steiner symmetrization prove a result due to Goryainov: for 0 < v < 1
the level set Cv (f ) lies in the half-plane
Im z  Im s(iv; a).
2 −1 −aπ/4
Here z = s(w; a) ≡ π cosh (e cosh(πw/2)) is a function which maps
the strip Sw conformally and univalently onto the strip Sz cut along the
interval [0, s(0, a)] of the imaginary axis. Show that if we assume in addition
that Sz \ D(f ) ∩ {z : Im z = 1} = ∅, then the level curve Cv (f ) also lies in
the half-plane
Im z  v.
(5) For f ∈ M0 such that for z ∈ U we have |f (z)|  M with 1 < M < ∞ and
for arbitrary θ and r such that 0 < r < 1 prove the inequality
n
Λf (r, θ + 2πk/n)
 rn (1 + rn )−2 .
(1 + (Λf (r, θ + 2πk/n)/M )n )2/n
k=1

Here equality holds only for the function in M0 taking the disc U to the disc

√ −2
|w| < M cut along the line segments {w : wn = ρeiθn , 1 + 1 − M −n 
ρ  M n }.
(6) For arbitrary θ and r such that 0 < r  1 prove the following sharp estimate
for functions in the previous exercise:
>
? n −2/n
? √ 
@
n n
Λf (r, θ + 2πk/n)  4r 1 + rn + (1 + rn )2 − 4rn M −n .
k=1

Equality is attained only for the same function as in Exercise 7.5(5).


(7) Let f be a function in M0 . Show that for 0 < r < 1 and for θ such that
{ρ exp(i(θ + π(k − 1))) : 0  ρ  ∞} ⊂ f ({z : |z| < r}), k = 1, 2, we have
 −1
1 −1 −1
(Λ (r, θ) + Λf (r, θ + π))  r(1 + r2 )−1 .
2 f
Here equality holds if and only if f (z) ≡ z((e−iθ z)2 + c(e−iθ z) + 1)−1 , where
c is an arbitrary real constant with absolute value |c|  (1 + r2 )r−1 .
(8) Let L(r, f ), 1 < r < R, be the image of the circle |z| = r under a map
f ∈ M(R) (that is, L is a level curve) and let wk (r, θ, f ), k = 1, . . . , n, be
the points defined before Theorem 7.29. If f is a holomorphic function in the
class M(R), then prove the following inequality for arbitrary θ, r, and ρ such
that 1  r < R and 1  ρ < R:
n    
1  wk (r, θ, f )   G(r; n, R) 
n log wk (ρ, θ, f )   log G(ρ; n, R)  .
k=1

Equality sign holds for f (z) = eiθ G(z; n, R) and arbitrary r and ρ.
7.6. Inequalities involving the Schwarzian derivative 259

7.6 Inequalities involving the Schwarzian derivative


Let f be a holomorphic function in a domain B ⊂ C such that f  (z) = 0 for z ∈ B.
Then the expression
   2
f  (z) 1 f  (z)
Sf (z) = −
f  (z) 2 f  (z)

is well defined in B; it is called the Schwarzian derivative or Schwarzian of f . If f


has a simple pole at z, then we define the Schwarzian derivative at this point by

Sf (z) = S1/f (z).

Finally, if f is holomorphic and univalent in a neighbourhood of infinity, then the


authors usually set
Sf (∞) = lim z 4 Sϕ (z),
z→0
10
where where ϕ(z) = f (1/z). Thus we have defined the Schwarzian derivative of a
meromorphic locally univalent function f in any domain B ⊂ C. A straightforward
calculation shows that
Sf ◦g = (Sf ◦ g)(g  )2 + Sg .
The Möbius transformations f can be characterized by the condition Sf = 0.
Using asymptotic formulae for the capacities of condensers with shrinking
plates and then drawing the ‘accumulation points’ of the plates closer together we
obtain estimates for some combinations of the coefficients of univalent functions.
Note that a plate shrinking to a point yields an estimate for the absolute value of
the derivative f  (z); two mutually approaching and shrinking plates with opposite
signs (a dipole) yield an estimate for the Schwarzian Sf (z); and a simplest four-
plate multipole (or three-plate one, with signs of the plates chosen appropriately)
leads to an estimate for Sf (z) − Sf2 (z). We explain this using the following simple
example. Let f ∈ M be a function with the following expansion in a neighbourhood
of the origin:
f (z) = c0 + c1 z + c2 z 2 + c3 z 3 + · · · .
Since the capacity is conformally invariant and monotonic (see Theorem 1.15), for
sufficiently small ρ, r, and arbitrary ϕ we have

cap C(r; U, ∂U, {ρeiϕ , −ρeiϕ }, {0, 1, −1}, {r, r})


= cap C r; f (U ), ∂f (U ), {f (ρeiϕ ), f (−ρeiϕ )},



{0, 1, −1}, {|f (ρeiϕ )|r, |f  (−ρeiϕ )|r}
 cap C(r; C, ∅, {f (ρeiϕ), f (−ρeiϕ )}, {1, −1}, {|f (ρeiϕ )|r, |f  (−ρeiϕ )|r}).
10 See another definition of Sf (∞) after Theorem 7.36.
260 Chapter 7. Univalent Functions

Using asymptotic formulae (2.10) and (2.15) as r → 0 we obtain

4ρ2 (1 − ρ2 )2 |f (ρeiϕ ) − f (−ρeiϕ )|2


 .
(1 + ρ2 )2 |f  (ρeiϕ )f  (−ρeiϕ )|
Taking account of the expansion of f in a neighbourhood of the origin and letting
ρ → 0, after simple calculations we find
 
2iϕ c3 c22 1
1  Re e − 2 = Re e2iϕ Sf (0).
c1 c1 6
As ϕ can be arbitrary, we finally arrive at Kraus’ inequality

|Sf (0)|  6.

Interestingly enough, taking the condenser

C(r; U, ∅, {ρeiϕ , −ρeiϕ }, {0, 1, −1}, {r, r})

in place of C(r; U, ∂U, {ρeiϕ , −ρeiϕ }, {0, 1, −1}, {r, r}) we obtain in a similar way
a reverse inequality for the capacity but the same inequality for the Schwarzian.
Theorem 7.31 (Aharonov [Ah]). If f ∈ M, then

|Sf (0) − Sf2 (0)|  60.

Equality is attained at the Koebe function k(z) = z(1 + z)−2 .


Proof. Since the Schwarzian is a linear fractional invariant, we can assume that


f ∈ S; let f (z) = z + cn z n . It follows from the monotonicity of the reduced
n=2
modulus that for any points w1 , . . . , wn in f (U ) and any real numbers δ1 , . . . , δn ,

n
such that δk = 0,
k=1

n
n
δk δl h(wk , wl )  0, (7.57)
k=1 l=1

where h(w, ζ) = gf (U) (w, ζ) + log |w − ζ|. Now let

w1 = 0, w2 = f (ρeiϕ ), w3 = f (−ρeiϕ ); δ1 = 2, δ2 = δ3 = −1;

then

h(w1 , w1 ) = 0;
  
8
h(w2 , w2 ) = Re 2c2 ρeiϕ + (3c3 − 2c22 )ρ2 e2iϕ + 4c4 − 6c2 c3 + c32 ρ3 e3iϕ
3
  $
9 ρ4
+ 5c5 − 8c2 c4 − c23 + 12c22 c3 − 4c42 ρ4 e4iϕ − ρ2 − + o(ρ4 );
2 2
7.6. Inequalities involving the Schwarzian derivative 261

  
8
h(w3 , w3 ) = Re −2c2 ρeiϕ + (3c3 − 2c22 )ρ2 e2iϕ − 4c4 − 6c2 c3 + c32 ρ3 e3iϕ
3
  $
9 ρ4
+ 5c5 − 8c2 c4 − c23 + 12c22 c3 − 4c42 ρ4 e4iϕ − ρ2 − + o(ρ4 );
2 2
    
c2 c3
h(w1 , w2 ) = Re c2 ρeiϕ + c3 − 2 ρ2 e2iϕ + c4 − c2 c3 + 2 ρ3 e3iϕ
2 3
 2 4
 $
c c
+ c5 − c2 c4 − 3 + c22 c3 − 2 ρ4 e4iϕ + o(ρ4 );
2 4
  2
  
c2 2 2iϕ c32
h(w1 , w3 ) = Re −c2 ρe + c3 −

ρ e − c4 − c2 c 3 + ρ2 e3iϕ
2 3
  $
c23 2 c42 4 4iϕ
+ c5 − c 2 c 4 − + c2 c3 − ρ e + o(ρ4 );
2 4
   $
2 2iϕ c23 4 4iϕ ρ4
h(w2 , w3 ) = Re c3 ρ e + c5 − ρ e + ρ2 − + o(ρ4 ),
2 2
and inequality (7.57) takes the form
% &
Re (4c5 − 8c2 c4 − 6c23 + 16c22 c3 − 6c42 )e4iϕ  2.
Hence |2c5 − 4c2 c4 − 3c23 + 8c22 c3 − 3c42 |  1. It is easy to verify that the left-
hand side of this inequality is equal to |Sf (0) − Sf2 (0)|/60. The equality case is
straightforward. 
Similar inequalities also hold for bounded univalent functions (Exercises
7.6(1), (2)).
Now we look at an example of a two-point distortion theorem involving
Schwarzian derivatives.
Theorem 7.32 (Alenitsyn [A]). If f is a function in the class M and if f (±r) = ∞,
where 0 < r < 1, then
 
    2
Sf (r) + Sf (−r) − 12 f (r)f (−r) + 3   48r , (7.58)
 (f (r) − f (−r)) 2 r 2  (1 − r4 )2
   $
 12  f  (r)f  (−r) 3(6r2 − r4 − 1)
Sf (r) + Sf (−r) +  − 12Re  .
 2
(1 + r )2  (f (r) − f (−r)) 2 r2 (1 − r2 )2
(7.59)
Equalities in (7.58) and (7.59) are attained at the composites of the Koebe function
with linear fractional automorphisms of the sphere.
Proof. For sufficiently small ρ > 0 set
z1 = r + ρeiϕ , z2 = r − ρeiϕ , z3 = −r + ρeiθ , z4 = −r − ρeiθ ,
W = {f (zk )}4k=1 , = {(−1)k+1 }4k=1 ,
Ψ = {μk τ }4k=1 , μk = 1, k = 1, 2, 3, 4.
262 Chapter 7. Univalent Functions

Since the reduced modulus is monotonic, it follows that

M (f (U ), ∂f (U ), W, , Ψ)  M (W, , Ψ).

Expressing the left-hand side in terms of inner radii and Green’s functions (2.16)
and using formula (2.17) for the reduced modulus of the complex sphere we obtain


n
n
(−1)k+l h(zk , zl )  0, (7.60)
k=1 l=1

where
 
 f (zk ) − f (zl ) 

h(zk , zl ) = log (1 − zk z l ) k = l,
zk − zl ,
h(zk , zk ) = log |(1 − |zk |2 )f  (zk )|.

We write out the expansions of f in a neighbourhood of r and −r:





f (z) = ak (z − r)k and f (z) = bk (z + r)k ,
k=0 k=0

which gives us
  
2a2 eiϕ r(eiϕ + e−iϕ )
h(z1 , z1 ) = log |a1 (1 − r2 )| + Re ρ −
a1 1 − r2
  2 
2iϕ
2 3a3 e 2a22 e2iϕ 1 r2 eiϕ + e−iϕ
+ρ − − − + o(ρ2 ),
a1 a21 1 − r2 2 1 − r2
  
2 2a2 eiϕ r(eiϕ + e−iϕ )
h(z2 , z2 ) = log |a1 (1 − r )| + Re ρ − +
a1 1 − r2
  2 
2iϕ
2 3a3 e 2a22 e2iϕ 1 r2 eiϕ + e−iϕ
+ρ − − − + o(ρ2 ),
a1 a21 1 − r2 2 1 − r2
  
2b2 eiθ r(eiθ + e−iθ )
h(z3 , z3 ) = log |b1 (1 − r2 )| + Re ρ +
b1 1 − r2
  2 
2iθ
2 3b3 e 2b22 e2iθ 1 r2 eiθ + e−iθ
+ρ − − − + o(ρ2 ),
b1 b21 1 − r2 2 1 − r2
  
2 2b2 eiθ r(eiθ + e−iθ )
h(z4 , z4 ) = log |b1 (1 − r )| + Re ρ − −
b1 1 − r2
   
2iθ 2 2iθ 2 iθ −iθ 2
3b 3 e 2b e 1 r e + e
+ρ2 − 22 − − + o(ρ2 ),
b1 b1 1 − r2 2 1 − r2
7.6. Inequalities involving the Schwarzian derivative 263

 
eiϕ − e−iϕ 1
h(z1 , z2 ) = log |a1 (1 − r2 )| + Re −ρr + ρ 2
1 − r2 1 − r2
 2 
a3 e2iϕ r2 eiϕ − e−iϕ
+ − + o(ρ2 ),
a1 2 1 − r2
 
eiθ − e−iθ 1
h(z3 , z4 ) = log |b1 (1 − r2 )| + Re ρr 2
+ ρ 2
1−r 1 − r2
  2

b3 e2iθ r2 eiθ − e−iθ
+ − + o(ρ2 ),
b1 2 1 − r2
 
(1 + r2 )|a0 − b0 | a1 eiϕ − b1 eiθ eiϕ − eiθ
h(z1 , z3 ) = log + Re ρ −
2r a0 − b 0 2r
−iθ
  2iϕ 2iθ iθ 2
r(e − e )

a2 e − b 2 e (e − e )

+ + ρ2 +
1 + r2 a0 − b 0 8r2
  2  2

1 a1 eiϕ − b1 eiθ r2 eiϕ − e−iθ ei(ϕ−θ)
− − − + o(ρ2 ),
2 a0 − b 0 2 1 + r2 1 + r2
 
(1 + r2 )|a0 − b0 | a1 eiϕ + b1 eiθ eiϕ + eiθ
h(z1 , z4 ) = log + Re ρ −
2r a0 − b 0 2r
−iθ
  2iϕ 2iθ iθ 2

r(e + e ) a2 e − b 2 e iϕ
(e + e )
+ + ρ2 +
1 + r2 a0 − b 0 8r2
 2  2 
1 a1 eiϕ + b1 eiθ r2 eiϕ + e−iθ ei(ϕ−θ)
− − + + o(ρ2 ),
2 a0 − b 0 2 1 + r2 1 + r2
 
(1 + r2 )|a0 − b0 | a1 eiϕ + b1 aiθ eiϕ + eiθ
h(z2 , z3 ) = log + Re ρ − +
2r a0 − b 0 2r
 
r(eiϕ + e−iθ ) a 2 e 2iϕ
− b 2 e 2iθ
(e iϕ
+ e iθ 2
)
− + ρ2 +
1 + r2 a0 − b 0 8r2
  2   2

1 a1 eiϕ + b1 eiθ r2 eiϕ + e−iθ ei(ϕ−θ)
− − + + o(ρ2 ),
2 a0 − b 0 2 1 + r2 1 + r2
 
(1 + r2 )|a0 − b0 | −a1 eiϕ + b1 eiθ eiϕ − eiθ
h(z2 , z4 ) = log + Re ρ +
2r a0 − b 0 2r
−iθ
  2iϕ 2iθ iθ 2
r(e − e )

a2 e − b 2 e (e − e )

− 2
+ ρ2 +
1+r a0 − b 0 8r2
  2  2

1 −a1 eiϕ + b1 eiθ r2 eiϕ − e−iθ ei(ϕ−θ)
− − − + o(ρ2 ).
2 a0 − b 0 2 1 + r2 1 + r2
264 Chapter 7. Univalent Functions

In the limit as ρ → 0 inequality (7.60) takes the following form:



a1 b 1 a3 a1 − a22
Re 2ei(ϕ+θ) + e2iϕ
(a0 − b0 ) 2 a21
2
$
2iθ b3 b1 − b2 ei(ϕ+θ) 2ei(ϕ−θ) 2
+e − −  .
b21 2r2 (1 + r2 )2 (1 − r2 )2

Now we set ϕ = θ + π and, as θ can be arbitrary, we obtain (7.58), while setting


ϕ = −θ + π we obtain (7.59). 
Following the proof of Theorem 7.32 we easily deduce an estimate for a linear
combination of Schwarzian derivatives at one or several points in the disc U and
similar estimates in multiply connected domains.
From theorems on extremal decompositions (Chapter 6) we can derive some
results for meromorphic and univalent functions sharing no values. For exam-
ple, let f be a function in M. Elementary calculations show that the quantity
K(f (U ), f (z), ϕ) introduced before Theorem 6.4 has the representation

K(f (U ), f (z), ϕ) (7.61)


  $
1 1 2iϕ Sf (z)
= − Re e , z ∈ U, f (z) = ∞
π (1 − |z|2 )2 |f  (z)|2 6(f  (z))2

(see, for instance, equalities A2.12, A2.44, A2.45, and A2.47 in [C]). Hence we can
deduce the following result from Theorem 6.4: if fk with fk (0) = ∞, k = 1, . . . , n,
n  2, are meromorphic univalent functions in the disc U which map U onto
pairwise nonoverlapping domain, then for arbitrary positive numbers αk and real
numbers ϕk , k = 1, . . . , n, we have

n   $ n √
n
1 2iϕk Sfk (0) αk αl ei(ϕk +ϕl )
αk  2
− Re e  2
 Re . (7.62)
|fk (0)| 6(fk (0)) (fk (0) − fl (0))2
k=1 k=1 l=1
l=k

Theorem 7.33. For any meromorphic univalent functions fk in the disc U satisfy-
ing |fk (0)| = 1, k = 1, . . . , n, n  2, and mapping U onto pairwise nonoverlapping
domains

n
1 n(n2 + 2) fk2 (0)Sfk (0)
n

 −  Re
|fk (0)|2 24 6(fk (0))2
k=1 k=1
(7.63)
n
1 n(n2 − 1)
− + .
|fk (0)|2 12
k=1

Both equalities in (7.63) are attained for the functions fk∗ (z) = exp(2πi(k − 1)/n)
× [(1 + z)/(1 − z)]2/n , k = 1, . . . , n, where the branch of the root function taking
1 to 1 is selected.
7.6. Inequalities involving the Schwarzian derivative 265

Proof. Setting αk = 1 and eiϕk = ifk (0), k = 1, . . . , n, in (7.62) we obtain


n  $
n
n
1 fk2 (0)Sfk (0) −fk (0)fl (0)
+ Re  Re
|fk (0)|2 6(fk (0))2 (fk (0) − fl (0))2
k=1 k=1 l=1
l=k

n n
1 n(n2 − 1)
=  .
|fk (0) − fl (0)|2 12
k=1 l=1
l=k

Here we have used (5.5). Thus we have proved the right-hand inequality in (7.63).
As regards the left-hand inequality in (7.63), it is a consequence of (6.31) with
Bk = fk (U ) and ak = fk (0), k = 1, . . . , n, and the above formula (7.61). The
equality cases are verified by direct calculations. 
The next result supplements the classical estimate for the Schwarzian of a
bounded function (Exercise 7.6(1)) by taking account of the distortion of boundary
arcs lying on the circle ∂U .
Theorem 7.34. If f is a function in the class SB0 and f (z) → γ  ⊂ {w : |w| = 1}
as z approaches a boundary arc γ ⊂ {z : |z| = 1}, then
   
1  
 Sf (0) + e−i2t sin2 σ − e−i2t (f  (0))2 sin2 σ   2 cos σ − |f  (0)|2 cos σ ,
3 2 2 2 2
(7.64)
where σ is the length of γ, eit is the midpoint of this arc, σ  is the length of γ  ,

and eit is the midpoint of γ  . Equality is attained for the Pick function l(z; λ) =

k −1 (λ
k(z)) and the arcs {eiθ : |θ| < σ/2}, where the numbers σ and λ satisfying
0 < σ < 2π and sin2 (σ/4)  λ  1 can be arbitrary (and  k(z) = z(1 − z)−2 is the
Koebe function).

Proof. Applying (7.64) to the function e−it f (eit z) we see that we can content
ourselves with the case t = t = 0. In Exercise 7.4(6) let n = 2, z1 = reiϕ ,
z2 = −reiϕ , δ1 = 1, and δ2 = −1, where r > 0 is sufficiently small and ϕ is an
arbitrary real number. Setting
f (z) = α1 z + α2 z 2 + α3 z 3 + · · · ,
we arrive at the following representation for the left-hand side of the inequality in
that exercise:
  
2α3 3α2
4Re − 22 ei2ϕ r2 + o(r2 ), r → 0.
α1 2α1
To calculate the right-hand side we look at the function
2w sin(σ/4)
ζ = h(w) =  ,
1−w+ (1 − weiσ/2 )(1 − we−iσ/2 )
which maps the exterior B of the arc {w = eiθ : |θ|  σ/2} conformally and
univalently onto the disc |ζ| < 1 (here we take the branch of the root function
266 Chapter 7. Univalent Functions

equal to 1 at w = 0). Using well-known power expansions of elementary functions


we carry out the following calculations:
1 − |h(f (±reiϕ ))|2
log r(B, f (±reiϕ )) = log
|h (f (±reiϕ ))|

σ σ σ
= − log sin − |α1 |2 r2 sin2 − Re ±2α1 reiϕ cos2
4 4 4
  
 
$
σ σ σ σ
+r2 ei2ϕ 2α2 cos2 + α21 cos2 3 cos − 2 cos2 + o(r2 ),
4 4 2 4
r(B, f (±reiϕ ))
log r(B, 1/f (±reiϕ )) = log = log r(B, f (±reiϕ ))
|f 2 (±reiϕ )|
  $
α2 α3 α2
− 2 log(|α1 |r) − 2Re ± reiϕ + r2 ei2ϕ − 22 + o(r2 ),
α1 α1 2α1

gB (f (reiϕ ), f (−reiϕ )) = gB (1/f (reiϕ ), 1/f (−reiϕ ))


 
 1 − h(f (reiϕ ))h(f (−reiϕ )) 
 
= log  
 h(f (reiϕ )) − h(f (−reiϕ )) 
 
 σ  σ
= − log(2r) − log α1 sin  + |α1 |2 r2 sin2
4 4
  
$
α3 σ σ σ
− Re r2 ei2ϕ + 2α2 cos2 + α21 cos2 cos + o(r2 ),
α1 4 4 2
 
 1 − h(f (±reiϕ ))h(1/f (±reiϕ )) 
 
gB (f (±re ), 1/f (±reiϕ )) = log 


 h(f (±reiϕ )) − h(1/f (±reiϕ )) 

σ 2 2 2 σ

σ
= − log sin + |α1 | r cos + Re ∓α1 reiϕ ∓ α1 re−iϕ cos2
4 4 4
 
   

σ σ σ 1 1 σ
+ r2 ei2ϕ −α2 cos2 − α21 cos2 cos − + sin4
4 4 2 2 2 4
 
   
$
σ σ σ 1 σ
+r2 e−i2ϕ −α2 cos2 − α1 2 cos2 cos − cos4 + o(r2 ),
4 4 2 2 4
 
 1 − h(f (±reiϕ ))h(1/f (∓reiϕ )) 
 
gB (f (±re ), 1/f (∓re )) = log 
iϕ iϕ 
 h(f (±reiϕ )) − h(1/f (∓reiϕ )) 

σ 2 2 2 σ

σ
= − log sin − |α1 | r cos + Re ∓α1 reiϕ ± α1 re−iϕ cos2
4 4 4
 
  

σ σ 1 1 σ
+ r2 ei2ϕ −α2 cos2 − α21 cos2 − + sin4
4 4 2 2 4
 
   
$
2 −i2ϕ 2 σ 2 2 σ σ 1 4 σ
+r e −α2 cos − α1 cos cos − cos + o(r2 ).
4 4 2 2 4
7.6. Inequalities involving the Schwarzian derivative 267

Hence
H(f (γ), {f (reiϕ ), f (−reiϕ )}, {1, −1})
= log r(B, f (reiϕ )) + log r(B, f (−reiϕ )) + log r(B, 1/f (reiϕ ))
A
+ log r(B, 1/f (−reiϕ )) + 2 gB (f (reiϕ ), 1/f (reiϕ )) + gB (f (−reiϕ ),

1/f (−reiϕ )) − gB (f (−reiϕ ), 1/f (reiϕ )) − gB (f (reiϕ ),


B
1/f (−reiϕ )) − 2gB (f (reiϕ ), f (−reiϕ ))
 2  
2 2 σ α2 2 2 σ

= 4 log 2 + 8|α1 | r cos + 4Re + α1 sin e i2ϕ
r2 + o(r2 ), r → 0.
2 2α21 2
Setting f (z) ≡ z we obtain
H(γ, {reiϕ , −reiϕ }, {1, −1})
σ A σ B
= 4 log 2 + 8r2 cos + 4Re sin2 ei2ϕ r2 + o(r2 ), r → 0.
2 2
It remains to substitute the expressions obtained in the inequality in Exercise
7.4(6); bearing in mind that ϕ can be arbitrary, this yields (7.64). The equality
case is verified directly. 
Theorem 7.35. Let f and p be functions in M0 taking real values on the inter-
val (−1, 1). Assume that the Steiner symmetrization of f (U ) with respect to the
imaginary axis lies in p(U ). Then
6 + Sf (0) 6 + Sp (0)
 .
(f  (0))2 (p (0))2
Proof. Using the notation from Exercises 2.4(2), (3) and 4.1(9), for sufficiently
small ρ > 0 we obtain
H(f (U ), −iρ, iρ)  H(Stl f (U ), −iρ, iρ)  H(p(U ), −iρ, iρ),
where l is the imaginary axis and the last inequality follows because the reduced
modulus is monotonic. Passing to the limit as ρ → 0 we obtain
K(f (U ), 0, π/2)  K(p(U ), 0, π/2);
in combination with the representation (7.61) this gives us the required relation.

Corollary 7.14. If w = f (z) = z + a2 z 2 + a3 z 3 + · · · is a holomorphic univalent
function in the disc U and an , n = 2, 3, . . ., are some real numbers, then
a3 − a22  π 2 /(6d2f ) − 1, where df = sup mes{w ∈ f (U ) : Im w = v}.
v

Equality sign is attained for the function f (z) = 2i log[(1 − iz)/(1 + iz)], which
maps U conformally and univalently onto the strip |Re w| < π/2.
268 Chapter 7. Univalent Functions

Let E be a nondegenerate continuum and let


a−1
f (z) = a1 z + a0 + + ···
z
be a function mapping the exterior D := {z : |z| > 1} of the unit disc conformally
and univalently onto the connected component of Cw \ E containing the point at
infinity. Let
a∗
f ∗ (z) = a∗1 z + a∗0 + −1 + · · ·
z
denote the function mapping D conformally and univalently onto Cw \ E ∗ , where
E ∗ is the set obtained from E by Steiner symmetrization with respect to the real
axis. Then the Pólya–Szegő inequality

|a1 |  |a∗1 | (7.65)

is equivalent to the following inequality for logarithmic capacities:

cap E  cap E ∗

(see Corollary 4.2). It is natural to ask about the behaviour of other coefficients in
the expansion of f under Steiner symmetrization. There are no interesting relations
between |a0 | and |a∗0 |. In fact, a translation of E parallel to the imaginary axis
does not change Re a0 , Re a∗0 and Im a∗0 = 0, while Im a0 can be made arbitrary.
On the other hand we can give examples when Re a0 = Re a∗0 , and then translating
E along the real axis we can achieve the inequality |Re a0 | < |Re a∗0 |. Further, if
E is a line segment making an acute angle with the real axis, then |a−1 | > |a∗−1 |.
For E = {w : |w| = 1} we obtain the reverse inequality |a−1 | = 0 < 1/2 = |a∗−1 |.
Theorem 7.36. If functions f and f ∗ are as above, then

|a1 |2 − Re a1 a−1  |a∗1 |2 − Re a∗1 a∗−1 . (7.66)

This inequality is equivalent to the relation


1 1 Sf (∞) 1 1 Sf ∗ (∞)
+ Re   ∗ + Re ∗  ,
|f  (∞)|2 6 (f (∞)) 2 |f (∞)| 2 6 (f (∞))2

where the derivative and Schwarzian at infinity are calculated with respect to the
local parameters, that is, f  (∞) is equal to the derivative of the function 1/f (1/z)
at the origin and Sf (∞) = S1/f (1/z) (0).
We can interpret (7.66) as follows in terms of capacities. Let E be a mirror-
symmetric set with respect to the real axis, and for H := {w : Im w > 0} let
H \ E be a simply connected domain. Let g denote the function mapping H \ E
conformally and univalently onto H so that

lim [g(w) − w] = 0.
w→∞
7.6. Inequalities involving the Schwarzian derivative 269

Then the limit


hcap(E ∩ H) = lim w[g(w) − w]
w→∞
is called the ‘half-plane’ capacity of E ∩ H with respect to the point at infinity [L],
p. 69. If f is as above, then g has the following expansion in a neighbourhood of
infinity:
a2 − a1 a−1
g(w) = w + 1 + ··· ,
w
where a1 and a−1 are real numbers. Thus
hcap(E ∩ H) = |a1 |2 − Re a1 a−1
and (7.66) transforms into
hcap(E ∩ H)  hcap(E ∗ ∩ H).
Proof of Theorem 7.36. For fixed v > 0 let ϕv (w) := w/v − i be a linear transfor-
mation taking the circle |w − iv| = v to the unit circle |w| = 1, and let ϕ−1
v be
the inverse map. We regard the symmetrization of the closed bounded set E with
respect to the circle |w − iv| = v as the composite map
Rv E = ϕ−1
v (R ϕv (E)),

where R is symmetrization with respect to the unit circle. For an open set B
containing iv and ∞ let
Sv B = Cw \ Rv (Cw \ B).
Let {Bn }∞ be an exhaustion of f (D) by simply connected domains Bn such
n=1


that ∞ ∈ Bn , B n ⊂ Bn+1 , n = 1, 2, . . ., and Bn = f (D). Using Theorem 4.4
n=1
(k = 1) and taking (2.16) into account we conclude that for sufficiently large v > 0
log[r(Bn , iv)r(Bn , ∞)] − 2gBn (iv, ∞)
= log[r(ϕv (Bn ), 0)r(ϕv (Bn ), ∞)] − 2gϕv (Bn ) (0, ∞)
' (
 log r(C \ R(C \ ϕv (Bn )), 0)r(C \ R(C \ ϕv (Bn )), ∞)
− 2gC\R(C\ϕv (Bn )) (0, ∞)
= log[r(Sv Bn , iv)r(Sv Bn , ∞) − 2gSv Bn (iv, ∞).
Finally,
log[r(Bn , iv)r(Bn , ∞)] − 2gBn (iv, ∞)
(7.67)
 log[r(Sv Bn , iv)r(Sv Bn , ∞)] − 2gSv Bn (iv, ∞).
Since the reduced modulus is monotonic, for two domains G1 and G2 with Green’s
functions such that G1 ⊂ G2 and for different points ζ and w in the domain G1 ,
we have
log[r(G1 , ζ)r(G1 , w)] − 2gG1 (ζ, w)  log[r(G2 , ζ)r(G2 , w)] − 2gG2 (ζ, w). (7.68)
270 Chapter 7. Univalent Functions

By Exercise 4.1(12), for sufficiently large v

Sv Bn ⊂ Cw \ (Cw \ f (D))∗ ⊂ f ∗ (D).

Thus it follows from (7.67) and (7.68) that

log[r(Bn , iv)r(Bn , ∞)] − 2gBn (iv, ∞)


(7.69)
 log[r(f ∗ (D), iv)r(f ∗ (D), ∞)] − 2gf ∗ (D) (iv, ∞).

Now we let
an−1
+ ···
fn (z) = an1 z + an0 +
z
denote a function mapping D conformally and univalently onto Bn , and let

w an an
hn (w) = n − 0n − −1 + · · ·
a1 a1 w

be the expansion on the inverse mapping in a neighbourhood of the point at


infinity. Then for sufficiently large v

r(Bn , ∞)|an1 | = r(D, ∞) = 1,


r(Bn , iv)|hn (iv)| = r(D, hn (iv)) = |hn (iv)|2 − 1,
gBn (iv, ∞) = gD (hn (iv), ∞) = log |hn (iv)|.

Hence
|hn (iv)|2 − 1
r(Bn , iv)r(Bn , ∞)e−2gBn (iv,∞) =
|an1 hn (iv)||hn (iv)|2
 −2
 iv 
1 −  n + O(1)
  
a1 n 2 1 1
=   
 = 1 − (|a 1 | − Re a n n
a
1 −1 ) 2
+ o , v → +∞.
n n
1 − a1 a−1 + o 1  v v2
 v2 v2 

Repeating the above calculations for f ∗ in place of fn , we conclude from (7.69)


that
|an1 |2 − Re an1 an−1  |a∗1 |2 − Re a∗1 a∗−1 .
Taking the limit as n → ∞ we obtain (7.66). 
Let f be the function defined above and

a∗−1 (θ)
fθ∗ (z) = a∗1 (θ)z + a∗0 (θ) + + ···
z
be a function mapping the domain D conformally and univalently onto the com-
plement of the set obtained by the Steiner symmetrization of the continuum E
7.6. Inequalities involving the Schwarzian derivative 271

with respect to the straight line w = teiθ , −∞ < t < +∞ (f0∗ = f ∗ ). Applying
Theorem 7.36 to e−iθ f and e−iθ f0∗ we arrive at the following conclusion:

|a1 |2 − Re a1 a−1 e−2iθ  |a∗1 (θ)|2 − Re a∗1 (θ)a∗−1 (θ)e−2iθ .

In particular, for θ = π/2 (that is, when Steiner symmetrization is performed with
respect to the imaginary axis) we obtain

|a1 |2 + Re a1 a−1  |a∗1 (π/2)|2 + Re a∗1 (π/2)a∗−1 (π/2). (7.66 )

Note that the classical inequality (7.65) holds for any Steiner symmetrization:

|a1 |  |a∗1 (θ)|, 0  θ  2π.

Concerning the question of whether the left-hand sides of (7.66) and (7.66 ) ac-
tually decrease after Steiner symmetrization with respect to an arbitrary straight
line, we make the following observation. The function f (z) = (z + 1/z) exp(i(π/2 +
ϕ)) = a1 (ϕ)z + a0 (ϕ) + a−1 (ϕ)/z + · · · maps D conformally and univalently onto
the exterior of the interval [−2 exp(i(π/2 + ϕ)), 2 exp(i(π/2 + ϕ))], and the func-

tion fπ/2 (z) = i(z + 1/z) cos ϕ maps D onto the exterior of [−2i cos ϕ, 2i cos ϕ].
For these functions we have

|a1 (ϕ)|2 − Re a1 (ϕ)a−1 (ϕ) = |a∗1 (π/2)|2 − Re a∗1 (π/2)a∗−1 (π/2), 0  ϕ  2π,
|a1 (ϕ)|2 + Re a1 (ϕ)a−1 (ϕ) > |a∗1 (π/2)|2 + Re a∗1 (π/2)a∗−1 (π/2), ϕ = πn, n ∈ Z.

We can show that for any holomorphic univalent function f in D with real coeffi-
cients we have

|a1 |2 − Re a1 a−1  |a∗1 (π/2)|2 − Re a∗1 (π/2)a∗−1 (π/2).

We can complement Theorem 7.36 with the following result.


Theorem 7.37. Let f(z) =  a1 z + 
a0 + 
a−1 /z + · · · be a function mapping D
conformally and univalently onto the exterior of a continuum E  ⊂ E ∗ . Then

|a∗1 |2 − Re a∗1 a∗−1  |


a1 |2 − Re 
a1 
a−1 . (7.70)

Proof. This follows from inequality (7.68), in which we must set G1 = f ∗ (D),
G2 = f(D), w = ∞, and ζ = iv, where v is sufficiently large. Next we perform cal-
culations for the functions f ∗ and f by repeating the end of the proof of Theorem
7.36 for fn . As a result, we obtain (7.70). 

As applications of Theorems 7.36 and 7.37, we state some covering results


for the class Σ.
272 Chapter 7. Univalent Functions

Corollary 7.15. Let f (z) = z + a0 + a−1 /z + · · · be a function in Σ and let w0 be


a point in E = Cw \ f (D). Then
m4f (w0 , ϕ) + 16Rf4 (w0 )
 1 + Re e−2iϕ a−1
8m2f (w0 , ϕ)

for each real number ϕ, where Rf (w0 )  0 is the radius of the largest disc with
centre at w0 which lies in E and mf (w0 , ϕ) is the linear Lebesgue measure of the
intersection of E with the line {w = w0 + teiϕ : t ∈ R}. Equality sign is attained
for functions f (z) = w0 + eiϕ λ−1 h−1 (λh(e−iϕ z)), where h(ζ) = ζ + 1/ζ and λ > 1
is arbitrary.
Proof. We start with the case w0 = 0, ϕ = π/2. For λ = [m2f (0, π/2)/(4Rf2 (0)) +
1]/[mf (0, π/2)/Rf (0)] the function
 
iRf (0) 1
f(z) := iRf (0)h−1 (λh(z)) = iλRf (0)z + λ− + ···
z λ
maps the domain D conformally and univalently onto the complement of the set
E := {w : |w|  Rf (0)} ∪ {w : Re w = 0, |Im w|  mf (0, π/2)/2}. Under
 ⊂ E ∗ (E = Cw \ f (D)). Hence
the assumptions of Corollary 7.15 we have E
inequalities (7.66) and (7.70) yield
m4f (0, π/2) + 16Rf4 (0)
1 − Re a−1  Rf2 (0)(2λ2 − 1) = .
8m2f (0, π/2)

To treat arbitrary w0 and ϕ we must apply this argument to the function

a0 − e−2iϕ a−1 /z + · · · .
ei(π/2−ϕ) [f (ei(ϕ−π/2) z) − w0 ] = z + 

The conditions for equality are straightforward. 


In particular, the above estimate contains the equalities

m2f (w0 , (arg a−1 )/2)/8 − 1  |a−1 |  1 − m2f (w0 , (arg a−1 + π)/2)/8.

The right-hand inequality improves a well-known consequence of the area theorem,


the bound |a−1 |  1 (cf. [Gol], Ch. II, § 4). Both inequalities supplement the
classical bound
mf (w0 , ϕ)  4, for any ϕ,
which is a consequence of (7.65). Namely,

mf (w0 , (arg a−1 )/2)  8(1 + |a−1 |)  4,

mf (w0 , (arg a−1 + π)/2)  8(1 − |a−1 |).

Here equalities hold for |a−1 | = 1 and f (z) = z + w0 + e2iϕ /z. Finding sharp
estimates for fixed values of |a−1 | other than 1 would be of interest.
7.6. Inequalities involving the Schwarzian derivative 273

Corollary 7.16. Assume that for some α, β, and γ a function f (z) = z + a0 +


a−1 /z + · · · in Σ satisfies the inequality

m ((Cw \ f (D)) ∩ l(u))  α for any u, β  u  γ.

Then
Re a−1  1 − c2 (1 − k 2 )/2,
where c and k are the real constants defined by the condition
1  2
ζ − k2
c dζ = (γ − β)/2 − iα/2, c > 0, 0 < k < 1.
ζ2 − 1
0

Equality sign holds for f in Σ which maps the domain D conformally and uni-
valently onto the exterior of the rectangle with two sides lying on the lines u =
β, u = γ, where γ − β < 4, and with a suitable height α.
Proof. We can assume that β = −γ. The function
ζ 
ζ 2 − k2
F (ζ) = c dζ + iα/2 = cζ + c0 − [c(1 − k 2 )/2]/ζ + · · ·
ζ2 − 1
0

maps the upper half-plane Im ζ > 0 conformally and univalently onto the quadri-
lateral {w : Im w > 0} \ E, where this time we set E
 = {w : |Re w|  γ, |Im w| 
α/2}. By the Riemann–Schwarz symmetry principle the function
   
1 1 c 1 1
f(z) := F (z + ) = z + c0 − c − k2 + ···
2 z 2 2 z
 Under the assumptions
maps D conformally and univalently onto the exterior of E.
 ∗
of Corollary 7.16, E ⊂ (Cw \ f (D)) . It remains to use inequalities (7.66) and
(7.70). 

Exercises 7.6
(1) (Nehari) For functions f (z) = c1 z + · · · in the class SB0 prove the inequality

|Sf (0)|  6(1 − |c1 |2 ).

Is this bound sharp?


(2) Using Theorem 7.22 deduce the sharp bound

|Sf (0) − Sf2 (0)|  60(1 − |c1 |4 − 2|c2 |2 )

for functions f (z) = c1 z + c2 z 2 + · · · in SB0 . Give an example of a function


with equality sign in this relation.
274 Chapter 7. Univalent Functions

(3) (Singh [Sin]) In the conditions of Exercise 7.1(15) let β coincide with the
circle |z| = 1. For an arbitrary point z in the annulus Ω prove the inequality
 
1  |f  (z)|2
 Sf (z) + πl(z, z)  πK(z, z) − ,
6  (1 − |f (z)|2 )2
where K(·, ·) and l(·, ·) are the Bergman kernels of the first and second kind
in Ω with respect to the class of holomorphic functions in Ω with square
integrable absolute value.
(4) (Alenitsyn [A]) Let f and g be functions mapping the disc U meromorphi-
cally and univalently onto nonoverlapping domains and satisfying Im f (0) =
Im g(0), f  (0) > 0, g  (0) > 0. Prove that
f  (0)g  (0) 1
 1 − |Sf (0) + Sg (0)|.
(f (0) − g(0))2 12
Equality holds for f (z) = (1 + z)/(1 − z) and g(z) = (z − 1)/(z + 1).
(5) For all functions f in the class M(R) prove the following inequalities:
 n    
n2 − 1  Gn (zk ) 2
n−  
2  f  (zk ) 
k=1
 n  2 
n 2
Re[zk (Sf (zk ) − SGn (zk ))] n2 + 2  Gn (zk ) 
   f  (zk )  − n ,
|f  (zk )|2 4
k=1 k=1

where Gn (z) = G(z; n, R) and zk = exp(i(π/n + 2π(k − 1)/n)), k = 1, . . . , n.


(6) For functions f in M set

νf (z0 , ϕ) = t−2 dt,
æ

where

æ = {t ∈ R : f (z0 ) + teiϕ ∈ f (U )}, z0 ∈ U, f (z0 ) = ∞.

For each real ϕ prove the inequality


4
νf2 (z0 , ϕ)  8(1 − |z0 |2 )−2 |f  (z0 )|−2 − Re[e2iϕ Sf (z0 )(f  (z0 ))−2 ].
3
Equality sign holds for f (z) = z(1 + (e−iϕ z)2 )−1 and z0 = 0.
(7) Let f ∈ M be a function with an expansion f (z) = z + a2 z 2 + a3 z 3 + · · · at
the origin. For each real number ϕ prove that

Re[e2iϕ (a3 − a22 )]  1 − νf2 (0, ϕ)/8.

Equality is attained for the same function as in the previous exercise.


7.6. Inequalities involving the Schwarzian derivative 275

(8) Let f be a function in SB0 which extends analytically to an open subarc


of the circle |z| = 1 containing z = 1 and takes it to some arc of the circle
|w| = 1. Prove the inequality

Re[2Sf (0) + Sf (1)]  12(1 − |f  (0)|2 ).

(9) Let E1 and E2 be nondegenerate continua which are mirror-symmetric rela-


tive to the real axis R and satisfy (E1 \ E2 ) ∪ (E2 \ E1 ) ⊂ R. Let

fk (z) = a1 (k)z + a0 (k) + a−1 (k)/z + · · ·

be a function mapping the domain |z| > 1 conformally and univalently onto
the connected component of C \ Ek containing the point at infinity, k = 1, 2.
Prove the equality

|a1 (1)|2 − Re a1 (1)a−1 (1) = |a1 (2)|2 − Re a1 (2)a−1 (2).


Chapter 8

Multivalent Functions

Applications of symmetrization to the analysis of the properties of multivalent


functions were exposed in the monographs [H], [J], [Mit4]. Roughly speaking, there
exist two lines of such applications, different in their approach to the multiplicity
of the covering. The first approach consists in using majorization principles (which
concern the behaviour of the inner radius of a domain, the capacity of a condenser,
and so on under regular mappings) and then symmetrizing the plane images and
using the monotonicity of the quantity under consideration. The second approach
is based on the symmetrization of sets and condensers lying on the Riemann surface
of the inverse mapping. In §§ 8.2–8.4 we give examples of such applications; most of
them are different from the examples in [H], [J], [Mit4]. The reader can find new ap-
plications of symmetrization of condensers to multivalent functions in [D5]–[D10]

8.1 Majorization principles


One of the most important majorization principles in function theory is Lindelöf’s
principle stated below.
Theorem 8.1. Let B ⊂ Cz and G ⊂ Cw be domains possessing Green’s function
and let f : B → G be a nonconstant meromorphic function in B. Then for each
fixed finite point w0 in f (B),

gG (f (z), w0 )  pk gB (z, zk ), z ∈ B, (8.1)
k0

where z0 , z1 , . . . are the zeros of f − w0 and p0 , p1 , . . . are their multiplicities,


respectively.
In addition,  if B is an admissible domain, then equality sign in (8.1) at some
point z ∈ B \ {zk } implies equality sign at all the points in B.
k0

© Springer Basel 2014 277


V.N. Dubinin, Condenser Capacities and Symmetrization in Geometric Function
Theory, DOI 10.1007/978-3-0348-0843-9_8
278 Chapter 8. Multivalent Functions

Proof. We can assume that the function f − w0 has only


 finite zeros. Suppose
B is admissible and fix some n  0. In the domain B \ {zk } we consider the
k0
auxiliary function


n
hn (z) = gG (f (z), w0 ) − pk gB (z, zk ).
k=0

In view of the expansion of f in a neighbourhood of zk :

f (z) = w0 + cpk (z − zk )pk + · · · , cpk = 0

and of the representation for the Green function in a neighbourhood of its pole,
we conclude that the points zk , k n, are removable singular points of hn . Thus
hn is a harmonic function in B \ {zk }, tends to +∞ as z → zk with k > n
k>n
(if there are zk with k > n), and 
is nonnegative on the boundary of B. By the
maximum principle hn  0 in B \ {zk }. If f − w0 has finitely many zeros, then
k>n
taking n + 1 equal to their numberwe obtain hn  0 in B, so that (8.1) holds.
Here equality at some point in B \ {zk } implies equality throughout B. If there
k0
are infinitely many zeros, then we use Harnack’s theorem and see that the series

gG (f (z), w0 ) − pk gB (z, zk )
k0

is uniformly convergent on compact subsets of B. Hence its sum is a nonnegative


harmonic function in B. By the maximum principle, if this function vanishes at
some point in B then it vanishes identically.
Now let B be a domain with Green’s function and let {Bn }∞ n=1 e ban ex-
haustion of B by domains Bn with classical Green functions. For fixed l we choose
n(l) such that Bn(l) contains the zeros z0 , z1 , . . . , zl . Then by the above


l
gG (f (z), w0 )  pk gBn (z, zk )
k=0

for z ∈ Bn and n  n(l). Passing to the limit as n → ∞ in the last inequality we


obtain

l
gG (f (z), w0 )  pk gB (z, zk )
k=0

for all z ∈ B. Since this holds for each l, inequality (8.1) holds. 
We say that a function f brings about a complete n-covering of a domain
G by a domain B if the following conditions are satisfied: f (B) = G; as z ∈ B
8.1. Majorization principles 279

approaches the boundary of B, all the limit values of f (z) lie on the boundary
of G; each point w ∈ G has precisely n inverse images (with multiplicities) in
B. Note that if the first two conditions are satisfied and some point w0 ∈ G
has precisely n inverse images, then f brings about a complete n-covering of the
domain G by B. Indeed, by the properties of analytic functions each point w ∈ G
has a neighbourhood U (w) with the following properties: for each w ∈ U (w) the
number of its inverse images f −1 (w ) is the same. Since every point G can be
joined to w0 by a curve lying in G and this curve can be covered by a finite system
of such neighbourhoods, each point in G has precisely n inverse images.
Let B be a domain in Cz , z0 be a finite point in B, w0 be a point in the
Cw -plane, and p  1 an integer. Then we let Mpw0 (B, z0 ) denote the class of
meromorphic functions f in B that have an expansion
f (z) = w0 + ap (z − z0 )p + · · · , ap = 0,
in a neighbourhood of z0 .
Theorem 8.2 (Mityuk [Mit1]). If f ∈ Mpw0 (B, z0 ), then

r(f (B), w0 )  |ap |rp (B, z0 ) exp pk gB (z0 , zk ), (8.2)
k1

where the zk , k  1, are the zeros of f − w0 distinct from z0 and the pk are
their multiplicities. In addition, if B is an admissible domain and the equation
f (z) − w0 = 0 has n roots with multiplicities taken into account, then equality
in (8.2) holds if and only if f (B) is an admissible domain and f brings about a
complete n-covering of the domain f (B) by B.
Proof. We can assume that B possesses a Green function. If the domain f (B) does
not have a Green function, then strict inequality holds in (8.2), Otherwise, using
Lindelöf’s principle we obtain

gf (B) (f (z), w0 )  pgB (z, z0 ) + pk gB (z, zk ).
k1

Hence
log r(f (B), w0 ) − log |f (z) − w0 |

 p log r(B, z0 ) − p log |z − z0 | + pk gB (z, zk ) + o(1), z → z0 .
k1

Adding p log |z − z0 | to both sides of this inequality and passing to the limit as
z → z0 we arrive at (8.2).
Let B be an admissible domain and assume that the equation f (z) − w0 = 0
has n roots with multiplicities. Consider the following harmonic function in B:

h(z) = gf (B) (f (z), w0 ) − pgB (z, z0 ) − pk gB (z, zk ).
k1
280 Chapter 8. Multivalent Functions

At z0 we have

h(z0 ) = log r(f (B), w0 ) − log [|ap |rp (B, z0 )] − pk gB (z0 , zk ).
k1

We conclude from the maximum principle that equality in (8.2) holds if and only
if h ≡ 0 in B. Now assume that we have equality sign in (8.2). Then it follows
from the definition of h that if w → ∂f (B) then gf (B) (w, w0 ) → 0, that is, f (B) is
an admissible domain. Furthermore, since h ≡ 0, it follows that f (z) → ∂f (B) as
z → ∂B. Hence the function f brings about a complete n-covering of the domain
f (B) by B. Conversely, if f brings about an n-covering and f (B) is an admissible
domain, then


m
m
h(z) = gf (B) (f (z), w0 ) − pgB (z, z0 ) − pk gB (z, zk ), p+ pk = n.
k=1 k=1

Under these assumptions, h(z) → 0 as z → ∂B. Hence h ≡ 0 and equality holds


in (8.2). 

Corollary 8.1 (Hayman [H], p. 124). Let f (z) = a0 +a1 z +· · · be a regular function
in the disc U taking values in a domain G. Then

|a1 |  r(G, a0 ).

Equality is attained for functions f mapping U onto G conformally and univa-


lently.

A conformal univalent mapping preserves the Green function and the inner
radius with respect to a point is multiplied by the modulus of the derivative of the
mapping at this point. This ensures the behaviour of the reduced modulus under
univalent mappings that is described by (2.16). If we have a multivalent mapping
and all elements of the tuple have the same sign, then the relation between the
reduced moduli of the image and inverse image follows again from (2.16), Theorem
8.2, and Lindelöf’s principle.

Theorem 8.3. Let B and G be domains in the Riemann spheres Cz and Cw , re-
spectively, which have Green’s functions and let f be a meromorphic function in
B taking values in G. Let wl , l = 1, 2, . . . , m, be different points in f (B) and let
zkl for k = 1, 2, . . . , nl and l = 1, 2, . . . , m be points in B such that expansions of
the form
f (z) − wl = akl (z − zkl )pkl + o((z − zkl )pkl ), z → zkl ,

in which akl = 0 and pkl > 0, k = 1, 2, . . . , nl , l = 1, 2, . . . , m, hold at these points


(if zkl = ∞ (wl = ∞) then z − zkl (f (z) − wl ) must be replaced with 1/z(1/f (z))).
8.1. Majorization principles 281


nl
Assume that the sum pkl = p is independent of l = 1, 2, . . . , m. Also set
k=1

W = {wl }m
l=1 , w = {δl }m
l=1 , Ψw = {μl rνl }m
l=1 ,

Z = {zkl : k = 1, . . . , nl , l = 1, . . . , m}, z = {δkl : k = 1, . . . , nl , l = 1, . . . , m},


δkl = δl , k = 1, 2, . . . , nl , l = 1, 2, . . . , m,
1/pkl νl /pkl
Ψz = {(μl /|akl |) r : k = 1, . . . , nl , l = 1, . . . , m}.

Then for any positive δl , l = 1, 2, . . . , m,


pM (B, ∂B, Z, z , Ψz )  M (G, ∂G, W, w , Ψw ). (8.3)

In addition, if B and G are admissible domains and f brings about a complete


p-covering of G by B, then equality sign holds in (8.3).
Proof. In view of (2.16), inequality (8.3) is equivalent to the relation
m 

m
m
m nl
2
p δl log r(G, wl ) + δl δk gG (wl , wk )  δl2 pjl log |ajl |
l=1 l=1 k=1 l=1 j=1
k=l
(8.4)

m
nl
m
m
nl
nk
+ δl2 p2jl log r(B, zjl ) + δl pjl δk pik gB (zjl , zik ),
l=1 j=1 l=1 k=1 j=1 i=1
(i=j for k=l)

which holds for any positive δl , l = 1, . . . , m. For fixed l and j such that 1  l  m
and 1  j  nl , Mityuk’s theorem yields

nl
log r(G, wl )  log |ajl | + pjl log r(B, zjl ) + pil gB (zil , zjl ).
i=1
i=j

We multiply this by pjl and take the sum for j from 1 to nl , after which we
multiply the result by δl2 and takes the sum for l from 1 to m. Then we obtain

m
m
nl
p δl2 log r(G, wl )  δl2 pjl log |ajl |
l=1 l=1 j=1

m
nl
m
nl
nl (8.5)
+ δl2 p2jl log r(B, zjl ) + δl2 pjl pil gB (zil , zjl ).
l=1 j=1 l=1 j=1 i=1
i=j

In a similar way, for fixed l and j such that 1  l  m and 1  j  nl , from


Lindelöf’s principle we obtain

nk
gG (wl , wk )  pik gB (zjl , zik ).
i=1
282 Chapter 8. Multivalent Functions

We multiply this by pjl and take the sum for j from 1 to nl , after which we
multiply the result by δl δk and take the sum for l and k from 1 to m, k = l. Then
we obtain

m
m
m
m
nl
nk
p δl δk gG (wl , wk )  pjl δl pik δk gB (zjl , zik ).
l=1 k=1 l=1 k=1 j=1 i=1
k=l k=l

In view of (8.5), this inequality yields (8.4) and therefore also (8.3). The equality
case is a consequence of the equality cases in Theorem 8.2 and Lindelöf’s principle
(see the end of the proof of Theorem 8.2). 
The question of the validity of (8.3) for arbitrary (not necessarily positive)
real δl arises in a natural way. Of course, then we must use the multiplicity of the
covering; for instance, we can limit ourselves to p-valent functions. A meromorphic
function f in a domain B is said to be p-valent in B if it takes each value at p
points at most.
Theorem 8.4. Under the assumptions of Theorem 8.3, if f is a p-valent meromor-
phic function in B, then inequality (8.3) holds for any real δl , l = 1, . . . , m, with
the corresponding assertion about equality sign.
Proof. We can assume that all the δl are distinct from 0, that μl = νl = 1, l =
1, . . . , m, and that B is bounded by finitely many analytic Jordan curves. We set
W = {wl }m l=1 , w = {δl }l=1 , Ψw = {tl (r)}l=1 , tl (r) ≡ r for l = 1, . . . , m, and νw =
m m
m
( δl2 )−1 . For sufficiently small r > 0, the condenser C(r; G, ∂G, W, w , Ψw ) is
l=1
well defined in Cw . By Theorem 2.1 the reduced modulus
νw
M (G, ∂G, W, w , Ψw ) = lim |C(r; G, ∂G, W, w , Ψw )| + log r (8.6)
r→0 2π
is also well defined. Corresponding to the condenser C(r; G, ∂G, W, w , Ψw ) there
is a condenser C(r; B, ∂B, Z, z , Ψz ) in Cz , in which the connected component of
the inverse image of D(wl , r) with respect to f lying in a neighbourhood of zkl is
taken for the plate Ekl , k = 1, . . . , nl , l = 1, . . . , m. It is easy to see that the sets
Ekl are almost discs. In view of Theorem 2.1,
νz
M (B, ∂B, Z, z , Ψz ) = lim |C(r; B, ∂B, Z, z , Ψz )| + log r , (8.7)
r→0 2π
where m n −1  −1
l
m
2
νz = δkl pkl = δl2 /p = νw /p.
l=1 k=1 l=1

Now let u(z) be the potential function of C(r; B, ∂B, Z, z , Ψz ), which


 means that
u(z) is a continuous real-valued function in Cz , harmonic in B\ Ekl , equal to 0
k,l
8.1. Majorization principles 283

on Cz \B, and equal to δkl on Ekl . We look at the function


1
v(w) = u(z)
p
f (z)=w

on f (B). For each w its value v(w) is a sum of p terms at most. The function
v(w) is continuous in f (B), vanishes on the boundary of f (B), is equal to δl on
D(wl , r), and is Lipschitz in a neighbourhood of each finite point in f (B), with a
possible exception of finitely many points. As y = x2 is a convex function and the
Dirichlet integral is conformally invariant, from Dirichlet’s principle we obtain
 
−1 2 1
|C(r; G, ∂G, W, w , Ψw )|  |∇v| dσw  |∇u|2 dσz .
p
f (B) B

Taking account of Dirichlet’s principle again, we conclude that


|C(r; G, ∂G, W, w , Ψw )|  p|C(r; B, ∂B, Z, z , Ψz )|. (8.8)
By (8.6)–(8.8) we have inequality (8.3).
Now assume that G has a classical Green function and f brings about a com-
plete p-covering of the domain G by B. Then the condenser C(r;G,∂G,W, w ,Ψw )
has a potential function ω(w). The composite ω(f (z)) is the potential function of
the condenser C(r; B, ∂B, Z, z , Ψz ), and since the Dirichlet integral is confor-
mally invariant and the covering has multiplicity p, it follows that
|C(r; G, ∂G, W, w , Ψw )| = p|C(r; B, ∂B, Z, z , Ψz )|.
In view of (8.6) and (8.7), this ensures equality sign in (8.3). 
If the δl have different signs, then the condition that f is p-valent is essential
in Theorem 8.4. We see this already in the following example, when p = 1. Let B
be the plane Cz cut along the rays [−∞, 0] and [1, +∞] of the real axis and let
G = Cw \[1, +∞]. Let w = f (z) denote the branch of a power function in B which
is defined by the relations w = r3/2 ei3ϕ/2 and z = reiϕ , −π < ϕ < π. The function
w = f (z) is two-valent, but each of the points w1 = e3πi/8 and w2 = e−3πi/8 has a
unique inverse image counting multiplicities, the point z1 = eiπ/4 or z2 = e−iπ/4 ,
respectively. Direct calculations show that
 

log r(G, w1 )r(G, w2 ) − 2gG (w1 , w2 ) = 2 log 2 sin .
8
At the same time,
 
3
log r(B, z1 )r(B, z2 ) − 2gB (z1 , z2 , ) + log |f  (z1 )f  (z2 )| = 2 log √ ,
2
in contradiction with (8.3).
Now we present a result on the behaviour of the reduced modulus of the
Riemann sphere.
284 Chapter 8. Multivalent Functions

Theorem 8.5. Let W = {wl }m l=1 , w = {δl }l=1 , and Ψw = {μl r }l=1 be tuples
m νl m

m  2
m
such that (δl /νl ) = 0 and δl = 0. Let w = R(z) be a rational function of
l=1 l=1
degree p  1, and assume that the following representation hold in neighbourhoods
of the points zkl , the inverse images of the wl under the mapping w = R(z) :

R(z) − wl = akl (z − zkl )pkl + o((z − zkl )pkl ), z → zkl ,


akl = 0, pkl > 0, k = 1, . . . , nl , l = 1, . . . , m,


nl

where pkl = p, l = 1, . . . , m if zkl = ∞ (or wl = ∞), then z − zkl (R(z) − wl ,


k=1
respectively) must be replaced with 1/z (with 1/R(z)) . Then

p M (Z, z , Ψz ) = M (W, w , Ψw ),

where Z = {zkl : k = 1, . . . , nl , l = 1, . . . , m}, z = {δkl : k = 1, . . . , nl , l =


1, . . . , m} with δkl = δl for k = 1, . . . , nl and l = 1, . . . , m, and Ψz = {μkl rνkl : k =
1, . . . , nl , l = 1, . . . , m} with μkl = (μl /|akl |)1/pkl and νkl = νl /pkl , k = 1, . . . , nl ,
l = 1, . . . , m.
Proof. First of all note that


m
nl
m
nl
m
2
δkl = 0, (δkl /νkl ) = p (δl /νl ) = 0,
l=1 k=1 l=1 k=1 l=1

so the reduced modulus M (Z, z , Ψz ) exists. Let C(r)  be the inverse image of the
condenser C(r; C, ∅, W, w , Ψw ) under the mapping w = R(z). It is easy to see
that the plates of the condenser C(r)  are almost discs with centres zkl and radii
νkl
μkl r , k = 1, . . . , nl , l = 1, . . . , m. Since w = R(z) is a complete p-covering of
the sphere Cw , it follows that

p|C(r)| = |C(r; C, ∅, W, w , Ψw )|.

Hence
 m n −1 
1 l

p 2
(δkl /νkl ) 
log r + |C(r)|

l=1 k=1
m −1
1
= (δl2 /νl ) log r + |C(r; C, ∅, W, w , Ψw )|

l=1

and we obtain the required equality by the limiting procedure as r → 0. 


Now we look at the behaviour of the capacity of a condenser under multivalent
mappings. Let B be a subdomain of C and C = (B, {E0 , E1 }, {0, 1}) be a condenser
8.1. Majorization principles 285

such that the boundary of B lies in the union of the plates of C. Let f be a
meromorphic function in B such that f (E0 ) ∩ f (E1 ) ∩ ∂f (B) = ∅. (Here f (Ek ) is
the limit set of f (z) as z → Ek , k = 0, 1.) Then the condenser

f (C) := (f (B), {f (E0 ), f (E1 ) ∩ ∂f (B)}, {0, 1})

is well defined, and the following majorization principle holds (cf. [Kl], Theorem 1).
Theorem 8.6. If the above assumptions are satisfied and if f (E0 ∩ ∂B) ⊂ ∂f (B),
then
cap C  cap f (C). (8.9)
If also E1 ⊂ ∂B and f is univalent in B, then equality holds in (8.9).
Proof. Let v be an admissible function for C which is equal to 0 in a neighbourhood
of E0 and to 1 in a neighbourhood of E1 and satisfies 0  v(z)  1 for z ∈ B. We
look at the following function on f (B):


⎨ 0, w ∈ f (E0 ),
V (w) = 1, w ∈ f (E1 ) ∩ ∂f (B),


min{v(z) : f (z) = w}, w ∈ f (B).

Note that f is nonconstant under the assumptions of the theorem, so it takes each
value in f (B) on a finite or countable infinite set of points in B, which accumulate
to the boundary of B. Hence the minimum in the definition of V is taken over
finitely many values of v. It is easy to see that V is continuous in f (B) and
Lipschitz in a neighbourhood of each point in f (B), with a possible exception of
finitely many points w such that f (z) = w and f  (z) = 0 (see the theorems in
§ 1.1). Now we complete f (C) to a condenser in Cw in the natural way and use
Theorem 1.13, which shows that

I(v, B)  I(V, f (B))  cap f (C).

Using Lemma 1.2 we arrive at the required inequality (8.9). The conditions for
equality in (8.9) are sufficient because the Dirichlet integral is conformally invari-
ant. 
Corollary 8.2. Let f be a regular function in U satisfying |f (z)| < 1 for z ∈ U .
Then for each compact set γ ⊂ U ,

caph γ  caph f (γ).

Corollary 8.3. [Kub] Let f be a regular function in the annulus K = {z : r <


|z| < 1} (r > 0) such that |f (z)| < 1 for z ∈ K and |f (z)| = 1 for |z| = 1. Then

caph {w : |w| < 1, w ∈ f (K)}  r.

Equality holds when f is univalent.


286 Chapter 8. Multivalent Functions

Now let Rp (K) denote the class of regular functions w = f (z) in K = K(R) =
{z : 1 < |z| < R} satisfying the following conditions: the set f (K) of values of
f (z) in K lies in the domain |w| > 1; f takes the circle |z| = 1 to |w| = 1; a circuit
in the positive direction along the circle |z| = 1 corresponds to precisely p circuits
along the circle |w| = 1: 
d arg f (z) = 2πp.
|z|=1

Theorem 8.7 (Mityuk [Mit3]). If f is a function in the class Rp (K), then


cap C  p cap f (C), (8.10)
where C = ({z : |z|  1}, {z : |z|  R}) and f (C) = ({w : |w|  1}, Cw \ (f (K) ∪
{w : |w|  1})). Equality holds in (8.10) if and only if f (C) is an admissible
condenser and f brings about a p-fold covering of f (K) by the domain K.
Proof. Let u be the potential function of C and α := {z : |z| = 1}. Using Theorems
1.13 and 1.6 we obtain

∂u
cap C = I(u, K) = dsz , (8.11)
∂nz
α

where ∂/∂nz denotes the inward normal derivative on the boundary of K. Now
let {Cn }∞n=1 be an exhaustion of f (C) by condensers Cn = ({w : |w|  1}, En )
whose plates En are bounded by finitely many analytic Jordan curves, and let ωn
be the potential function of Cn , n = 1, 2, . . .. By Harnack’s theorem the sequence
of functions ωn , n = 1, 2, . . ., converges to a harmonic function ω uniformly on
compact subsets of αf ∪ f (K), where αf := {w : |w| = 1}, and we have ω = 0 on
the circle αf and the limit values of ω(w) as w → ∂f (K) do not exceed 1. In view
of Theorems 1.11, 1.13, and 1.6, we conclude that
cap f (C) = lim cap Cn = lim I(ωn , f (K) \ En )
n→∞ n→∞
 
∂ωn ∂ω (8.12)
= lim dsw = dsw ,
n→∞ ∂nw ∂nw
αf αf

where ∂/∂nw denotes the inward normal derivative on the boundary of f (K). We
look at the auxiliary function
v(z) = ω(f (z)) − u(z),
which is harmonic in the annulus K and differentiable on α ∪ K, vanishes on the
circle α and has nonpositive limit boundary values as |z| → R. By the maximum
principle v  0 in K. Hence
 
∂ω(f (z)) ∂u
dsz − dsz  0. (8.13)
∂nz ∂nz
α α
8.1. Majorization principles 287

Since  
∂ω(f (z)) ∂ω
dsz = p dsw ,
∂nz ∂nw
α αf

relations (8.11)–(8.13) yield (8.10).


Now, if equality holds in (8.10), then, as there is equality in (8.13), we have
v ≡ 0 in K. Hence f (z) → ∂f (K) as z → ∂K. Bearing also in mind that a circuit
along α corresponds to p circuits along αf we see that f brings about a complete
p-covering of the domain f (K) by K. Moreover, as v ≡ 0, it follows that f (C)
is an admissible condenser. Conversely, if f (C) is an admissible condenser and f
brings about such a covering, then the composite ω(f (z)) is the potential function
of the condenser C and
 
∂ω(f (z)) ∂ω
cap C = dsz = p dsw = p cap f (C). 
∂nz ∂nw
α αf

Using the scheme of the above proof, Mityuk proved a more general result
than Theorem 8.7 [Mit2]. Note that the proofs in [Kub], [Mit3] and [Kl] use the
properties of the function v, while the proof of Theorem 8.6 is based on another
approach. We can show that if we modify this approach similarly to the proof of
Theorem 8.4 then we also arrive at Theorem 8.7.

Exercises 8.1
(1) (Mityuk [Mit3]) For regular functions f in the class Mp0 (B, z0 ) prove the
inequality
sup |f (z)|  |ap |rp (B, z0 ).
B

Can equality hold here? Consider the cases of simply connected and nonsim-
ply connected domain B separately.
(2) (Mityuk [Mit3]) Let f (z) = a0 + an z n + an+1 z n+1 + · · · be a regular function
in C such that the f -image of the disc U (R) = {z : |z| < R} has inner radius
r(f (U (R)), a0 ) = o(Rn ) as R → ∞. Show that f ≡ a0 .
(3) (Mityuk [Mit3]) Let f be a regular function in a domain B which satisfies
one of the following conditions: f ≡ const in B or r(f (B), f (z)) < ∞ for
z ∈ B; assume that the series

gB (z0 , zk ),
k1

constructed from the sequence {zk }k0 of zeros of f − a, is divergent. Prove


that f ≡ a in B.
288 Chapter 8. Multivalent Functions

(4) Let B and G be finitely connected domains without isolated boundary points
and let γ ⊂ ∂B and Γ ⊂ ∂G be subsets consisting of finitely many nonde-
generate connected components. Let f be a regular function in B such that
f (B) ⊂ G and f ((∂B) \ γ) ⊂ (∂G) \ Γ (so that for each sequence of points
ζn ∈ B approaching the boundary (∂B) \ γ the corresponding sequence of
points f (ζn ) → (∂G) \ Γ). Let w0 be a point in f (B) and let {zk }k0 be
points in B such that f (zk ) = w0 ; let pk be the order of the zero zk of the
function f − w0 . Prove the inequality

gG (f (z), w0 , Γ)  pk gB (z, zk , γ)
k0

for z ∈ B. Show that if equality holds at some point z ∈ B \ {zk }, then it
k0
holds throughout B.
(5) Under the assumptions of the previous exercise assume that

f (z) = w0 + cn (z − z0 )n + · · · , cn = 0,

in a neighbourhood of z0 (n = p0 ). Prove the inequality


 

r(G, Γ, w0 )  |cn |r (B, γ, z0 ) exp
n
pk gB (z0 , zk , γ) .
k1

(6) In Exercise 8.1(4) let z0 ∈ B, w0 = f (z0 ), let zk , k = 1, 2, . . . , be the zeros


of f − w0 distinct from z0 and pk be the order of the zero at zk , k = 1, 2, . . ..
Show that  

|Df (z0 )|  exp − pk gB (z0 , zk , γ)  1,
k1

where the ‘derivative’ |Df (z0 )| was defined before Theorem 7.4.
(7) Describe all the equality cases in Theorems 8.3 and 8.4.
(8) Establish an analogue of Theorem 8.4 for the reduced moduli

M (B, γ, Z, z , Ψz ) and M (G, Γ, W, w , Ψw ),

where γ = ∂B, Γ = ∂G.


(9) Find necessary and sufficient conditions for equality in (8.9).
8.2. Covering theorems 289

8.2 Covering theorems


Combinations of majorization and symmetrization principles produce various cov-
ering theorems for multivalent functions. Here we give just a few examples of such
applications.
Theorem 8.8. Let f ∈ Mpw0 (B, z0 ), and let Lf < ∞ be the linear measure of the
set of numbers R such that the whole of the circle |w| = R lies in the domain
f (B). Then
 

−p
|ap |  4(Lf + |w0 |)(r(B, z0 )) exp − pk gB (z0 , zk ) , (8.14)
k1

where the zk , k  1, are zeros of f − w0 distinct from z0 and the pk are their
multiplicities. If B is an admissible domain and the total number of zeros (with
multiplicities) of f (z) − w0 is n, then equality holds in (8.14) if and only if f
brings about a complete n-covering of the Cw -plane cut along the ray |w|  Lf ,
arg w = arg w0 + π if w0 = 0 or along the ray |w|  Lf , arg w = θ with an
arbitrary real θ if w0 = 0 by the domain B.
Proof. Assume that the domain f (U ) possesses a Green function. Then by Corol-
lary 4.3
r(f (B), w0 )  r(Cr f (B), |w0 |) = r(G, 0),
where G = {w : w + |w0 | ∈ Cr f (B)}. Corollary 3.3 yields

r(G, 0)  r(M G, 0).

It is easy to see that

M (π, G)  m(π, G)  Lf + |w0 |.

Hence the domain M G lies in G∗ := Cw \ {w : |w|  Lf + |w0 |, arg w = π}. Since


the inner radius is monotonic, it follows that

r(M G, 0)  r(G∗ , 0) = 4(Lf + |w0 |).

Summing the above relations we obtain

r(f (B), w0 )  4(Lf + |w0 |). (8.15)

Inequality (8.14) is a consequence of (8.2) and (8.15). If f (U ) possesses no Green


function, then we apply the above calculations to admissible subdomains of f (U )
and arrive at the relation ∞  4(Lf + |w0 |) < ∞, a contradiction.
Now let B be an admissible domain and assume that the equation f (z)−w0 =
0 has n roots counting multiplicities. If equality holds in (8.14), then it also holds
in (8.2) and (8.15). From equality in (8.2) we conclude that f (B) is an admissible
290 Chapter 8. Multivalent Functions

domain and f brings about a complete n-covering of f (B) by B. Corollary 4.3


says that the domain f (B) and point w0 are only different from Cr f (U ) and |w0 |
by a possible rotation, and by Corollary 3.3, Cr f (U ) is obtained from M G by
a translation by −|w0 |. Finally, if equality holds in (8.15) then M G = G∗ by
Exercise 2.1(3). Thus, f (B) coincides with the domain indicated in the statement
of Theorem 8.8. Conversely, if f (B) has this form and f performs a complete n-
covering of f (B) by B, then equality sign holds in (8.2), (8.15) and therefore also
in (8.14). 
Replacing circular symmetrization Cr in the proof of Theorem 8.8 by elliptic
symmetrization El we obtain an estimate for the absolute value |ap | in terms of the
linear measure of the set of L > |w0 − a|/2 such that the ellipse |w − w0 |+ |w − a| =
2L lies in f (B), where a is an arbitrary fixed point in C.
In what follows we drop the exponential in applications of (8.2) for simplicity
or set p = 1; on the other hand, in the case of equality we shall content ourselves
with indicating the extremal function.
Theorem 8.9 (Mityuk [Mit4]). If f ∈ Mpw0 (B, z0 ) and the domain f (B) intersects
each line Re w = const by a system of intervals with total length at most l, then
|ap |  2l/(πrp (B, z0 )). (8.16)
Equality is attained in (8.16) in the case when B is an admissible domain and f
performs a complete p-covering of the strip |Im(w − w0 )| < l/2 by B.
Proof. Since the inner radius is monotonic, using Corollary 4.1 we obtain in suc-
cession
r(f (B), w0 )  r(St f (B), Re w0 )  r(G∗ , Re w0 ) = 2l/π,
where this time G∗ = {w : |Im w| < l/2}. In combination with Theorem 8.2 this
yields the required inequality. In the case when f performs a complete p-covering
by an admissible domain B of a strip |Im (w − w0 )| < l/2, equality sign holds in
(8.2) and in the above inequalities. 
Using symmetrization we can show that (8.16) also holds in the case when for
each real u the complement Cw \f (B) contains a sequence of points u+i(v(u)+nl),
n = 0, ±1, ±2, . . . (see [H] and [Mit4]). For B = U and z0 = w0 = 0 the extremal
function in Theorem 8.9 has the form
l 1 + zp
f (z) = log .
π 1 − zp
In addition, if p = 1, then the assumptions of Theorem 8.9 can be relaxed consid-
erably by assuming that the Riemann surface of the inverse function f −1 contains
no vertical line intervals with length greater than l [D1].
Throughout what follows, in the case when f (∞) = ∞, we shall mean
by f  (∞) the quantity lim (f (z)/z)−1 , while for f (0) = ∞ we set f  (0) =
z→∞
lim (zf (z))−1 .
z→0
8.2. Covering theorems 291

Theorem 8.10. If f is a meromorphic function on the set B = Cz \ {z = eiθ : |θ| 


m/2}, 0 < m  2π, which is normalized by the conditions f (0) = 0, f (∞) = ∞,
and |f  (0)f  (∞)|  1, then the projection of the set E = Cw \ f (B) onto the unit
circle |z| = 1 has linear measure at most m. Equality is attained for f (z) ≡ z.

Proof. Assume the converse holds. Then from Theorem 8.3, using symmetrization
with respect to the circle |z| = 1, the principle of circular symmetrization (Corol-
lary 4.3), and the monotonicity of the moduli of domains we obtain in succession

1
M (B, ∂B, {0, ∞}, {1, 1}, {r, r}) + log |f  (0)f  (∞)|

 M (f (B), ∂f (B), {0, ∞}, {1, 1}, {r, r})
 M (Cz \ RE, ∂RE, {0, ∞}, {1, 1}, {r, r})
 M (Cz \ Cr RE, ∂ Cr RE, {0, ∞}, {1, 1}, {r, r})
< M (B, ∂B, {0, ∞}, {1, 1}, {r, r}).

We have obtained a contradiction, which proves Theorem 8.10. 

For an open set B containing z = 0 we denote


  −1
1 dr
S(θ, B) = − , E(ρ, θ, B) = {r : reiθ ∈ B, ρ  r < ∞},
ρ r2
E(ρ,θ,B)

where ρ > 0 is sufficiently small so that {z : |z|  ρ} ⊂ B. It is easy to see that


S(θ, B) is independent of ρ. Furthermore,

S(θ, B)  M (θ, B)  m(θ, B),

where m(θ, B) is the linear measure of the intersection of B with the ray arg z = θ
and M (θ, B) is as defined in (3.10). For the proof of the left-hand inequality we set
Tρ = [0, 1/ρe−iθ ] \ B  , where B  is the image of E(ρ, θ, B) under the map w = 1/z.
Then S(θ, B) = lim (m(−θ, Tρ ))−1  lim (M (−θ, Tρ ))−1 = M (θ, B).
ρ→0 ρ→0

Theorem 8.11. If f is a meromorphic function in U such that f (0) = 0 and


f  (0) = 1, then for each real θ and each positive integer n
n
1
S(θ + 2πk/n, f (U ))  .
4
k=1

Equality sign holds here for the function w = z[1 + (e−iθ z)n ]−2/n , which maps the
disc U conformally
 and univalently onto the plane Cw cut along the rays arg wn =
θn, |w|  1/4.
n
292 Chapter 8. Multivalent Functions

Proof. The function g(z) = 1/f (z) is meromorphic in U , and g(z) = 1/z + a0 +
a1 z + · · · in a neighbourhood of the origin. We set B = f (U ) and θk = θ +
2πk/n, k = 1, . . . , n. Then Corollary 8.1 and Theorem 5.6 yields in succession
n
1
1  [cap(C \ g(U ))]n  m(−θk , C \ g(U ))
4
k=1
n
   n
1 1 dr 1
= − = S −1 (θk , B).
4 ρ r2 4
k=1 E(ρ,θk ,B) k=1

The case of equality sign is straightforward. 


Theorem 8.12 (Marcus). If f (z) = z + a2 z 2 + · · · is a regular function in the unit
disc U , then for each real θ,
n   n  
−1 2πk −1 π 2πk
M θ+ , f (U ) + M θ+ + , f (U )  4.
n n n
k=1 k=1

Equality is attained for the functions of the form



f (z; θ, c, n) := z[1 + (c−n − 2)(e−iθ z)n + (e−iθ z)2n ]−1/n , n
1/4  c  ∞,

which map U conformally and univalently onto the plane Cw cut along the rays
arg wn = θn, |w|  c and arg wn = π + θn, |w|  (4 − c−n )−1/n .
Proof. By Theorem 8.11
> C
?
? n
2πk 1
@
n
M (θ + , f (U )) =: c 
n
.
n 4
k=1

If we assume that the inequality in the theorem fails, then


>
? n
? π 2πk
@
n
M (θ + + , f (U )) < (4 − c−n )−1/n .
n n
k=1

Hence the set S MA f (U ) with parameters A = {1/n, . . . , 1/n} is a proper subset


of the domain f (U ; θ, c, n). Using in succession Corollary 8.1, inequality (4.15) (for
V = S MA ), and the fact that the inner radius is monotonic we obtain

1  r(f (U ), 0)  r(S MA f (U ), 0) < r(f (U ; θ, c, n), 0) = 1.

This contradiction proves the inequality in Theorem 8.12. The equality case is
straightforward. 
In the notation introduced before Theorem 7.13 we have the following result.
8.2. Covering theorems 293

Theorem 8.13. Let f ∈ Rp (K), let ϕk , k = 1, . . . , n, be arbitrary real numbers


such that ϕ1 < ϕ2 < · · · < ϕn < ϕn+1 = ϕ1 + 2π, and let ak , k = 1, . . . , n, be
points on the outer boundary of f (K) such that arg ak = (ϕk+1 − ϕk )/2. Then
n

|ak | n(ϕk+1 −ϕk )  G(Rpn ; Rpn ). (8.17)
k=1

Equality holds in (8.17) for f (z) = αG(βz p ; n, Rp ) and


 √
n
ak = α exp(2πik/n) n G(Rpn ; Rpn ) ( 1 = 1), k = 1, . . . , n,

where |α| = |β| = 1.


Proof. By Theorem 8.7
cap C  p cap f (C).
Let {Ck }nk=1be the condensers obtained by the separating transformation of the
condenser f (C) with respect to the family of functions

ζ = pk (w) ≡ −i(e−iϕk w)π/(ϕk+1 −ϕk ) ,

ϕk < arg w < ϕk+1 , pk (ak ) = |ak |π/(ϕk+1 −ϕk ) > 0, k = 1, . . . , n. Then by Theo-
rem 4.8
1
n
cap f (C)  cap Ck .
2
k=1
Using Lemma 5.5, for each condenser Ck we obtain

cap Ck  cap Ck∗ ,


where Ck∗ = ({ζ : |ζ|  1}, {ζ : Im ζ = 0, |Re ζ|  pk (ak )}). Finally, by Theorem
3.13
1
n
cap Ck∗  cap RA {Ck∗ }nk=1 ,
n
k=1

where A = {1/n, . . . , 1/n} and


 n

π
RA {Ck∗ }nk=1 = {ζ : |ζ|  1}, {ζ : Im ζ = 0, |Re ζ|  |ak | n(ϕk+1 −ϕk ) } .
k=1

Summing the above relations yields


pn
cap C  cap RA {Ck∗ }nk=1 . (8.18)
2
Taking account of Theorems 8.7 and 4.9 we conclude that equality in (8.18) holds
for f (z) = G(z p ; n, Rp ) and the points ak equal to the nth roots of G(Rpn ; Rpn ).
Since the modulus of ring domains behaves strictly monotonically, we have (8.17).
The equality case is obvious. 
294 Chapter 8. Multivalent Functions

Exercises 8.2
(1) (Marcus) For a regular function f (z) = a1 z + a2 z 2 + · · · in U prove the
estimate   2π 
1
|a1 |  exp log M (θ, f (U ))dθ .
2π 0
Equality holds for f (z) = a1 z.
(2) Let f be a regular function in the disc U such that f (0) = 0 and f  (0) = 1, and
let Sf be a subset of the Riemann surface of the inverse function consisting of
points lying on this surface together with the straight-line segments joining
them to the f -image of the origin and not passing through the ramification
points. Show that if f (z) ≡ z, then for any real numbers ϕk , k = 1, . . . , n,
n that ϕ1 < ϕ2 < · · · < ϕn < ϕn+1 = ϕ1 + 2π, and any positive αk with
such
k=1 αk = 1 there exists ϕ such that
n
mαk (ϕ + ϕk , Sf ) > (r∗ )−1 ,
k=1

where π(r∗ )2 is the area of the inverse image of the star Sf with respect to f .
(3) In the notations of Theorem 8.11 show that if f is a meromorphic function
in U such that f (0) = 0 and f  (0) = 1, then for each real θ,
' −1 (2
S (θ, f (U )) + S −1 (θ + π, f (U ))
' (2
+ S −1 (θ + (π/2), f (U )) + S −1 (θ + (3π/2), f (U ))  16.

Equality sign holds for w = z[1 + c(e−iθ z)2 + (e−iθ z)4 ]−1/2 , where c is an
arbitrary constant such that |c|  2.
(4) (Marcus) Let f (z) = a1 z + a2 z 2 + · · · be a regular function in U such that
M (θ, f (U ))  M for 0  θ  2π and some constant M < ∞. Let Rn (θ) =
1
n
M (θ + 2πk/n, f (U )). For arbitrary θ such that 0  θ  2π prove the
k=1
inequality
' −1 (' (
Rn (θ) + Rn−1 (θ + π/n) 1 + Rn (θ)Rn (θ + π/n)/M 2n  4/|a1 |n .
Indicate the extremal function.
(5) (Aharonov–Kirwan) A regular function f in the disc U is called a Bieber-
bach–Eilenberg function if f (0) = 0 and f (z1 )f (z2 ) = 1 for all z1 , z2 ∈ U .
Prove the following inequality for such functions:
M (θ, f (U ))M (−θ, f (U ))  1, 0  θ  2π.

(6) (Marcus) For a Bieberbach–Eilenberg function f (z) = a1 z + a2 z 2 + · · · show


that
[Rn−1 (0) + Rn−1 (π/n)][1 + Rn (0)Rn (π/n)]  4/|a1 |n .
8.3. Distortion theorems 295

8.3 Distortion theorems


In this section we look at examples of applications of the approach based on capac-
ities and symmetrization in the following directions: estimates for the distortion in
a neighbourhood of the boundary, distortion estimates taking account of the cover-
ing, and distortion theorems for function sharing no values. The reader interested
in other applications of symmetrization to distortion theorems for multivalent
function can consult the monographs [H], [J], and [Mit4].
Let B denote the class of regular functions f in the disc U such that |f (z)| < 1
for z ∈ U . If f is a function in B which has an angular limit f (z) with |f (z)| = 1
at some boundary point z, |z| = 1, then the angular derivative f  (z) exists by the
Julia–Wolf lemma. When it is finite, the analytic function f  (ζ) has an angular
limit as ζ → z, which is equal to f  (z) ([Pom], pp. 79–83).
Theorem 8.14. Let f be a function in the class B, and let zk , k = 1, . . . , n, n  2,
be points on the circle |z| = 1 at which f has distinct angular limits f (zk ), k =
1, . . . , n, which are symmetrically positioned on the circle |w| = 1 (so that f n (zk )
is independent of k = 1, . . . , n). Then
 
 n 
 
 f  (zk )  1.
 
k=1

Equality is attained for f (z) ≡ αz with |α| = 1.

Proof. We can assume that f n (zk ) = 1, k = 1, . . . , n. We look at the continua



n 
n
γ = [0, rzk ] and γ ∗ = [0, r exp(2πik/n)], 0 < r < 1. Using Corollary 8.2
k=1 k=1
and Theorem 5.7, we obtain in succession

⎡ ⎤
π K 1 − τ12
caph γ  caph f (γ)  exp ⎣− ⎦,
2n K(τ1 )

where K(·) is the complete elliptic integral of the first kind,


 n
1/n
−1/2 1/2
τ1 − τ1 = [L−1
k (r) − Lk (r)] ,
k=1

Lk (r) = Re(f (rzk ))n/2 , k = 1, . . . , n, 0  τ1 < 1.

On the other hand Theorem 4.14 and the same Theorem 5.7 yield
⎡  ⎤
π K 1 − τ22
caph γ  caph γ ∗ = exp ⎣− ⎦,
2n K(τ2 )
296 Chapter 8. Multivalent Functions

−1/2 1/2
where τ2 − τ2 = r−n/2 − rn/2 . Consequently,
 n 1/n
(L−1
k (r) − Lk (r))  r−n/2 − rn/2 .
k=1

Hence
 n 1/n  1/n
  n
 
(f (zk ))n/2 − (f (rzk ))n/2   (1 − Lk (r))
k=1 k=1
n
 (1 − r) + o((1 − r)), r → 1.
2
It remains to divide the first and last expressions in this inequality by 1 − r and
to pass to the limit as r → 1. 
Note that in fact we have obtained an estimate for the modulus of the ‘radial
derivative’, provided that the ‘radial limit’ exists.
For a meromorphic function f in U and points z1 , z2 ∈ U such that f (z1 ) =
f (z2 ), f (zk ) = ∞, k = 1, 2, let αf (z1 , z2 ) denote the linear measure of θ such that
0  θ  2π and for the circle going through f (z1 ) and f (z2 ) and making at f (z1 )
angle θ with the direction of the real axis, its arc with end-points f (z1 ) and f (z2 )
does not lie in f (U ) entirely.
Theorem 8.15. If f is a meromorphic function in the disc U , then for any points
z1 , z2 ∈ U which are distinct from the poles of f and satisfy f (z1 ) = f (z2 ),
sin4 (αf (z1 , z2 )/4)|f  (z1 )f  (z2 )| |z1 − z2 |2
2
 .
|f (z1 ) − f (z2 )| (1 − |z1 | )(1 − |z2 |2 )|1 − z 1 z2 |2
2

Equality is attained for f (z) = z + 1/z and any points z1 and z2 in U not lying
on the real axis and satisfying z1 = z 2 .
Proof. We look at the meromorphic function F (z) = (f (z) − f (z1 ))/(f (z) − f (z2 ))
in U . Since the Green function is conformally invariant, by Theorem 5.4 (with
E = C \ F (U )) we obtain
log[r(f (U ), f (z1 ))r(f (U ), f (z2 ))] + 2gf (U ) (f (z1 ), f (z2 )) (8.19)
2
= log[r(F (U ), 0)r(F (U ), ∞)] + 2gF (U ) (0, ∞) + log |f (z1 ) − f (z2 )|
 log[|f (z1 ) − f (z2 )|2 sin−4 (αf (z1 , z2 )/4).
By Theorem 8.3 (or Theorems 8.1 and 8.2) the left-hand side of (8.19) is no less
than
log[|f  (z1 )f  (z2 )|r(U, z1 )r(U, z2 )] + 2gU (z1 , z2 )
= log[|f  (z1 )f  (z2 )|(1 − |z1 |2 )(1 − |z2 |2 )] + 2 log[|1 − z 1 z2 |/|z1 − z2 |].
The equality case is straightforward. 
8.3. Distortion theorems 297

Corollary 8.4. If f (z) = z + a1 z 2 + · · · is a regular function in the disc U , then


the following sharp estimate holds at any point z ∈ U distinct from a zero of f :
  

4 αf (0, z)  f (z) 
 |z|2
sin 
 f 2 (z)  1 − |z|2 .
4

The next two theorems extend some results due to Goluzin ([Gol], Ch. 4, § 5)
and Alenitsyn ([Gol], Appendix, § 2.1) to multivalent functions.
Theorem 8.16. The following sharp estimate holds for functions fk ∈ Mpwkk (U, 0),
k = 1, 2, 3, which map the disc U onto pairwise nonoverlapping domains:
 3 
 
 (pk )  64
 fk (0)/pk !  √ |(w1 − w2 )(w2 − w3 )(w3 − w1 )|. (8.20)
  81 3
k=1

Equality holds only for the functions of the form


 2/n
1 + βk z pk
fk (z) = αk ,
1 − βk z pk
where |αk | = |βk | = 1, k = 1, 2, 3, and also for the functions obtained from these
by means of an arbitrary linear fractional transformation (the same for all k).
Proof. By Theorem 8.2,
 
 (pk ) 
fk (0)/pk !  r(fk (U ), wk ), (8.21)

for k = 1, 2, 3 and equality holds in (8.21) if and only if fk (U ) is an admissible


domain and fk brings about a complete pk -covering of fk (U ) by the disc U , k =
1, 2, 3. Let Φ denote the Möbius transformation of Cw taking the points wk to
ak = exp(2πik/3), k = 1, 2, 3, respectively. Since I3 is invariant (see § 6.3) it
follows that
1
3 1
3 1
3
r(fk (U ), wk ) r(Bk , ak ) r(Bk , ak )
k=1
= k=1
= k=1
√ .
|(w1 − w2 )(w2 − w3 )(w3 − w1 )| |(a1 − a2 )(a2 − a3 )(a3 − a1 )| 3 3
To complete the proof of (8.20) we use Theorem 6.11, which states that
3
r(Bk , ak )  (4/3)3 . (8.22)
k=1

Now if equality sign holds in (8.20), then equality holds in (8.21) and (8.22).
Hence each function Φ ◦ fk brings about a complete pk -covering of the sector
| arg w − 2πk/3| < π/3 by the disc U , and

Φ ◦ fk (z) = ak + ck z pk + · · · , ck = 0, k = 1, 2, 3,
298 Chapter 8. Multivalent Functions

in a neighbourhood of the origin. Hence we easily obtain the representation for


the fk , k = 1, 2, 3, given by Theorem 8.16. Conversely, if the fk have this rep-
resentation, then by Theorems 8.2 and 6.11 equality holds in (8.21) and (8.22),
respectively. Hence equality also holds in (8.20). 

Theorem 8.17. If functions fk ∈ Mpwkk (U, 0), k = 1, . . . , n (n  2), map the disc
U onto pairwise nonoverlapping domains, then for any real constants δk , k =

n
1, . . . , n, such that δk = 0,
k=1

n  δk2 n n
 (pk ) 
fk (0)/pk !  |wk − wl |−δk δl .
k=1 k=1 l=1
l=k

Proof. This follows from inequality (8.21) and Theorem 6.1. 

We can significantly relax the condition that the domains fk (U ) in Theorems


8.16 and 8.17 do not overlap by taking account of the remark to Corollary 6.3
and 6.4.

Exercises 8.3
Let f be a function in the class B such that the radial limit lim f (rz) =: f (z)
r→1
exists at a point z, |z| = 1, and |f (z)| = 1. We define the ‘derivative’ of |f | at z by

1 − |f (rz)|
∂f (z) = lim .
r→1 1−r

Note that if f has an angular limit f (z) at z, then it is equal to the radial limit
and furthermore |f  (z)|  ∂f (z). On the other hand by the Schwarz–Pick lemma

1 − |f (rz)|2
|f  (rz)|  .
1 − r2
Thus, under these assumptions

|f  (z)| = lim |f  (rz)|  ∂f (z),


r→1

so that ∂f (z) = |f  (z)|. If 0 < |f  (z)| < ∞ then we say that f is conformal at z.

(1) For f ∈ B such that if it has radial limits f (zk ) at two different points z1 , z2
on the circle |z| = 1 and |f (zk )| = 1, k = 1, 2 prove the inequality

|∂f (z1 )∂f (z2 )|  |f (z1 ) − f (z2 )|2 /|z1 − z2 |2 .


8.3. Distortion theorems 299

(2) Let f ∈ B and assume that it has radial limits f (zk ) at three different points
zk on the circle |z| = 1 and |f (zk )| = 1, k = 1, 2, 3. Prove the inequality
3
1
∂f (zk )  √ |(f (z1 ) − f (z2 ))(f (z2 ) − f (z3 ))(f (z3 ) − f (z1 ))|.
k=1
3 3
Equality holds for any conformal automorphism of the disc U .
(3) For a map f in B which is conformal at boundary points z1 , . . . , zn and
satisfies f (zk ) = zk , k = 1, . . . , n, f (0) = 0, f  (0) = 0, prove that either
|f  (zk )|  2 at some point zk , 1  k  n, or
n
2
|f  (0)|  |2 − f  (zk )|2αk
k=1


n
for arbitrary real αk , k = 1, . . . , n, such that αk = 1.
k=1
Hint: the function w = g(z) = z 2 /f (z) maps the domain {z ∈ U : |z 2 /f (z)| <
1} conformally and univalently onto the disc |w| < 1.
(4) Let f be a function in B which extends analytically into a neighbourhood of
z = 1 and takes some open arc of the circle |z| = 1 containing this point to
an arc of the circle |w| = 1. Show that
Re Sf (1)  0.

(5) On the basis of Theorem 8.3 and Exercise 2.4(8) state and prove a gen-
eral distortion theorem for two multivalent functions in the disc which share
no values. As an example of its application prove the following result. Let
z0 , z1 , z2 be different points in the unit disc U , and let w0 , w1 , w2 be arbitrary
different points in Cw . Let fj , j = 0, 1, be meromorphic functions in U that
map this disc onto disjoint domains so that f0 (z0 ) = w0 , f1 (z1 ) = w1 , and
f1 (z2 ) = w2 . Then the following sharp estimate holds:
|z1 − z2 |2 |(w0 − w1 )(w0 − w2 )|4
|f0 (z0 )4 f1 (z1 )f1 (z2 )|  .
(1 − |z0 | ) (1 − |z1 |2 )(1 − |z2 |2 )|1 − z 1 z2 |2 |w1 − w2 |2
2 4

Equality is attained for univalent functions w = fj (z) defined as the com-


posites w = w(ζ), ζ = ζj (z), j = 0, 1, where
 2
w − w0 w2 − w1 2ζ 1 − z0z
= , τ (ζ0−2 (z) − 1) = eiθ ,
w − w1 w2 − w0 ζ +1 z − z0
   2
1 2 z(τ + σz 1 ) − τ z1 − σ
− τ ζ1 (z) + τ = ,
τ στ + z1 − z(1 + στ z 1 )
θ is an arbitrary real number, σ = ((z2 − z1 )/(1 − z 1 z2 ))|(1 − z 1 z2 )/(z2 − z1 )|,
and τ can be found from the equation 2τ /(1 + τ 2 ) = |z2 − z1 |/|1 − z1 z2 |, 0 <
τ < 1.
300 Chapter 8. Multivalent Functions

8.4 p-valent functions


Recall that a meromorphic function f in a domain B ⊂ C is said to be p-valent
in B if it takes each value at p points in B at most. The next statement supple-
ments Theorem 8.15 in the case when the covering brought about by f has finite
multiplicity.
Theorem 8.18. Let f be a meromorphic p-valent function in the disc U , and let
w1 and w2 be some different finite points in f (U ), each of which has p inverse
images in U with respect to f . Then the following inequality holds for the sets of
points {zk1 }nk=1
1
and {zk2 }nk=1
2
taken to w1 and w2 , respectively:
2 nl  pkl
4p −2p  (pkl ) 
sin (αf (z11 , z12 )/4)|w1 − w2 | f (zkl )/pkl !
l=1 k=1
 2
 2 2 nl nk  
nl
 zjl − zik pjl pik

2
(1 − |zkl |2 )−pkl   ,
 1 − z jl zik 
l=1 k=1 l=1 k=1 j=1 i=1
(i=j for k=l)

where pkl is the multiplicity of the zero zkl of the function f − wl , k = 1, . . . , nl ,



n1 n2
l = 1, 2, so that pk1 = pk2 = p. Equality holds, for example, for f (z) =
k=1 k=1
z p +1/z p and the points zk1 = r exp(πi/(2p)+2πik/p) and zk2 = r exp(−πi/(2p)+
2πik/p, 0 < r < 1, k = 1, . . . , p.

Proof. Repeating the first part of the proof of Theorem 8.15 we arrive at (8.19)
with f (z1 ) = w1 , f (z2 ) = w2 , z1 = z11 , and z2 = z12 . An estimate for the left-hand
side of this inequality can be found with the use of Theorem 8.4, by taking (2.16)
into account. In fact,

log[r(f (U ), w1 )r(f (U ), w2 )] + 2gf (U) (w1 , w2 )


= 8πM (f (U ), ∂f (U ), {w1 , w2 }, {0, 1, 1}, {r, r})  8πpM (U, ∂U, Z, z , Ψz )
 2 n 2
1
l nl
= pkl log |akl | + p2kl log r(U, zkl )
p
l=1 k=1 l=1 k=1
2 2

nl n k

+ pjl pik gU (zjl , zik ) ,


l=1 k=1 j=1 i=1
(i=j for k=l)

where Z = {zkl : k = 1, . . . , nl , l = 1, 2}, z = {0, 1, 1, . . . , 1}, and Ψz = { |pkl !/


f (pkl ) (zkl )|1/pkl r1/pkl , where k = 1, . . . , nl , l = 1, 2}. We complete the proof of the
inequality in Theorem 8.18 by simple calculations of the inner radii and values
of the Green function of U . For f and the points zkj indicated at the end of the
statement of the theorem the above inequalities turn to equalities. 
8.4. p-valent functions 301

Many applications of symmetrization to properties of p-valent functions and


also circumferentially mean p-valent or a really mean p-valent functions can be
found in [H], [J], and [Mit4]. So we limit ourselves to describing an approach
distinct from the one in [H], [J], and [Mit4] by looking at an example of the proof
of a covering theorem for radial segments.
Throughout what follows, by a Riemann surface we shall mean a surface
pasted together from finitely many plane domains, with naturally defined projec-
tions and neighbourhoods of points in the surface [HurC]. We call a Jordan curve
γ on a Riemann surface a line interval if the projection map takes γ one-to-one
onto some line interval in the plane. The length of γ is set to be the length of the
corresponding projection. In the next statement we generalize Corollary 7.13 to
p-valent functions. For simplicity we limit ourselves to regular mappings and do
not claim that the extremal function is unique.
Theorem 8.19. Let f be a function in Rp (K) which is p-valent in the annulus
1  |z| < R and let Lf (r, ϕ), where 1 < r  R, denote the supremum of the
lengths of the line intervals lying over the ray arg w = ϕ on the Riemann surface
of the inverse function of w = f (z), 1 < |z| < r, and having an end-point over the
circle |w| = 1. Then for any θ, n  1, and 1 < r  R,
n
(Lf (r, θ + 2πk/n) + 1)  G(rnp , Rnp ). (8.23)
k=1

Equality is attained for f (z) = eiθ G(αz p ; n, Rp ) with |α| = 1.


Proof. We can assume that θ = 0 and r < R. Let us introduce some notation:
α = {z : |z| = 1}; λk = Lf (r, 2πk/n) + 1; Gk = {w : |w| > 1, | arg w − 2πk/n| <
π/n} \ {t exp(2πik/n) : 1  t  λk }; G k is obtained from the domain Gk by
adding the attainable boundary points; z = Fk (w) is the function mapping Gk
conformally and univalently onto the rectangle {z : 0 < Re z < ak , |Im z| < 1}
with some ak > 0 so that the arc |w| = 1, 2πk/n − π/n < arg w < 2πk/n
is taken to (−i, ak − i), and the arc |w| = 1, 2πk/n < arg w < 2πk/n + π/n
is taken to (ak + i, i), k = 1, . . . , n. We call curves in G k which are given by
equations Re Fk (w) = x, 0  x  1, trajectories. Let Rf (ρ), r  ρ < R, be the
Riemann surface such that f maps the annulus 1 < |z| < ρ one-to-one onto this
surface (we use the same letter f for maps to the w-plane and to this Riemann
surface). For suitable ρ the surface Rf (ρ) is bounded by two analytic curves and
has finitely many ramification points. For each ray arg w = 2πk/n, k = 1, . . . , n,
there are at least p different line intervals lying in Rf (r) over this ray and joining
a point over |w| = 1 to a point W (j) on the other component of the boundary
of Rf (r), j = 1, . . . , np. When a ramification point lies over such a ray, these
intervals can intersect, so the total number of intervals can be larger than np, but
there are precisely np points W (j) . We set θj = arg f −1 (W (j) ), j = 1, . . . , np,

np
and let Ω denote the ring domain {z : 1 < |z| < ρ} \ [r exp(iθj ), ρ exp(iθj )].
j=1
302 Chapter 8. Multivalent Functions

For each fixed k we draw all the trajectories Re Fk (w) = xkl , l = 1, . . . , nk , in


G k which pass through the projections of ramification points in Rf (ρ) or are
tangent to the projection of the boundary of f (Ω); we also draw the trajectories
Re Fk (w) = 0, and the trajectories through the points at which pr ∂f (Ω) fails to
be analytic; assume that 0 = xk1 < xk2 < · · · < xknk = ak . Note that for no k and
l does the domain f (Ω) on Rf (ρ) contain a curve over a trajectory Re Fk (w) = x
with xkl < x < xkl+1 , l = 1, . . . , nk − 1, that connects two points over |w| = 1
in Rf (ρ): otherwise this curve γ would decompose the doubly connected domain
f (Ω) into two parts, and the boundary of each part would contain points in the
same boundary component f ((∂Ω)\α), which is disjoint from γ. For example, such
points lie over the ray arg w = 2πk/n and can be attained in the following way.
Let Wk be a point in γ whose projection lies on the ray arg w = 2πk/n. Moving
along the surface Rf (ρ) over this ray, away from Wk and towards the origin we
must reach a boundary point in the component f ((∂Ω) \ α) (by the definition of
Lf (r, 2πk/n) and W (j) ). On the other hand, moving away from Wk in the opposite
direction we also reach the boundary component f ((∂Ω) \ α) because the point
w = ∞ is not covered by the projection of Rf (ρ).
We conclude from this that the different curves on Rf (ρ) lying over the
trajectories Re Fk (w) = xkl , l = 1, . . . , nk , k = 1, . . . , n, subdivide f (Ω) into
finitely many domains so that for any k and l, 1  k  n, 1  l  nk − 1, we can
find 2p domains in this decomposition each of which covers univalently a strip-like
domain bounded by the trajectories Re Fk (w) = xkl and Re Fk (w) = xkl+1 and
has boundary points over the circle |w| = 1. Let these domains be denoted by
D+ − +
klj and Dklj , j = 1, . . . , p (the boundary of Dklj contains a curve lying over the
arc |w| = 1, 2πk/n < arg w < 2πk/n + π/n, while the boundary of D− klj contains
a curve lying over the arc |w| = 1, 2πk/n − π/n < arg w < 2πk/n). We shall
identify these domains with their projections.
Now we look at the following condensers:
C(ρ) = (Ω, {α, (∂Ω) \ α}, {0, 1});
f (C(ρ)) is the image of C(ρ) on the surface Rf (ρ);
±
Cklj is the ‘restriction’ of f (C(ρ)) to the set D±
klj ;

klj = Fk (C − ), while for j = p + 1, . . . , 2p C


for j = 1, . . . , p, we set C klj will
klj
+
be the reflection of the condenser Fk (Cklj−p ) in the real axis for (in the strips
xkl  x  xkl+1 we augment the plates of the condensers in the natural way). Since
capacity is a conformal invariant, from the composition principle (see Theorem
1.17) we obtain

cap C(ρ) = cap f (C(ρ))



n n p
k −1 2p
k −1
n n
 +
cap Cklj −
+ cap Cklj = klj .
cap C
k=1 l=1 j=1 k=1 l=1 j=1
8.4. p-valent functions 303

As f is p-valent, from Theorem 3.12 with A = (1/2p, . . . , 1/2p) we find that


n 2p
k −1 k −1
n
klj 
cap C klj }2p
2p cap LA {C j=1
l=1 j=1 l=1
k −1
n
 2p (xkl+1 − xk ) = 2pak ,
l=1

k = 1, . . . , n. By the symmetry principle the symmetric condenser


Ck∗ := (Cw , {{w : |w|  1}, {w : arg wn = 0, |w|  λk }}, {0, 1})
has capacity
cap Ck∗ = 2nak , k = 1, . . . , n.
Now Theorem 3.13 on the radial averaging transformation yields

n

cap Ck∗  n cap C,
k=1

where
  n
 
=
C Cw , {w : |w|  1}, {w : arg w = 0, |w| 
n 1/n
λk } , {0, 1} .
k=1

Summing the above relations we obtain



n
cap C(ρ)  
2pak  p cap C.
k=1

Hence

cap C(R)  p cap C.

On the other hand let C (R) be the condenser constructed similarly to C(R) but
with the number θj replaced with 2πj/(np), j = 1, . . . , np. By Theorem 4.14 on
dissymmetrization
cap C(R)  cap C ∗ (R),
and by the symmetry of h(z) := G(z p ; n, Rp ) we have
cap C ∗ (R) = p cap h(C ∗ (R)),
where
+  ,
h(C ∗ (R)) = Cw , {w : |w|  1},{w : argwn = 0, |w|  n G(rnp ;Rnp )} ,{0,1} .

Thus
  cap h(C ∗ (R)).
cap C
The required inequality follows because the capacity is strictly monotonic. The
equality case is a consequence of our definition of the extremal function. 
304 Chapter 8. Multivalent Functions

Exercises 8.4
(1) Let f , wl , zkl , and pkl , k = 1, . . . , nl . l = 1, 2, be the function, points and
numbers introduced in Theorem 8.18. Assume that |f (z)| < R for z ∈ U .
Prove the inequality
 p 2 nl  pkl
|R2 − w 1 w2 |2  (pkl ) 
f (z kl )/p kl ! 
(R2 − |w1 |2 )(R2 − |w2 |2 )|w1 − w2 |2
l=1 k=1
 2 n  2 2 n n   k+l
l l k
 zjl − zik (−1) pjl pik

2
2 −pkl
(1 − |zkl | )   .
 1 − z jl zik 
l=1 k=1 j=1 i=1
l=1 k=1
(i=j for k=l)

(2) For a rational function R(z) of degree n prove the equality


⎡ ⎤
n n n
* n n
  ⎢ ⎥
|R (ak )R (bk )| = ⎣ |ak − al ||bk − bl |⎦ |ak − bl |2 ,
k=1 k=1 l=1 k=1 l=1
l=k

where the ak , k = 1, . . . , n, are the roots of the equation R(z) = 0 and the
bk , k = 1, . . . , n, are the roots of the equation R(z) = 1.
(3) Following the proof of Theorem 7.5 generalize it to meromorphic p-valent
functions f .
(4) Passing from f to f (zR)/R, 1/R < |z| < 1, in (8.23) and letting R → ∞
deduce the following result. Let w = f (z) = z p + · · · be a regular p-valent
function in the disc |z| < 1, and let Λ  f (r, ϕ) be the supremum of the lengths
of the intervals on the Riemann surface of the inverse function of w = f (z),
|z| < r, which lie over the ray arg w = ϕ and have an end-point over w = 0.
Then for any θ, n  1 and 0 < r  1
n
 f (r, θ + 2πk/n)  rp (1 + rnp )−2/n .
Λ
k=1

Equality is attained for the functions f (z) = z p [1 + (e−iθ z p )n ]−2/n , which


cover the w-plane cut along the rays arg wn = θn, |wn |  1/4, with multi-
plicity p.
(5) We say that functions fk , k = 1, . . . , n (n  2) are jointly p-valent in the
disc U if in totality they take each value at p points in U at most, counting
multiplicities. Find the supremum of the product
n
|ck |
k=1

over all the sets of regular functions w = fk (z) = ak + ck z p + · · · in U , which


are jointly p-valent and satisfy |ak | = 1, k = 1, . . . , n .
Appendix

A1. Dirichlet’s principle


In what follows we consider a special case of Dirichlet’s principle in the form
required for the investigation of generalized condensers in Chapters 1, 2, and 3.
Theorem A1. Let B be an open set in C, let E = {Ek }nk=1 be a system of pairwise
disjoint nonempty closed sets Ek ⊂ B, k = 1, . . . , n, and let = {δk }nk=1 be
an n-tuple of real numbers, n  2. Assume that each connected component of

n
the set G := B \ Ek is a finitely connected domain without isolated boundary
k=1

n
points and that (∂G) ∩ ( Ek ) consists of finitely many nondegenerate connected
k=1
components. Then the condenser C = (B, E, ) has a potential function u and
I(u, G) = cap C.
Proof. The existence of a potential function (which is a solution to a Dirichlet
problem) follows from relevant theorems presented in well-known monographs on
function theory and potential theory (see, for instance, [GarM] or [R]). We can
assume that B is a domain and the plates in the set E lie at the boundary of B
(G = B). Bearing in mind that the Dirichlet integral is conformally invariant we
shall also assume that the boundary of B consists of finitely many analytic Jordan
curves. In these conditions the potential function u is an admissible function for
the condenser C in the sense of the definition in § 1.3. Hence
I(u, B)  cap C.

For an arbitrary admissible function v we have


  
∂(v − u) ∂u ∂(v − u) ∂u
I(v, B) = I(u, B) + I(v − u, B) + 2 + dxdy.
∂x ∂x ∂y ∂y
B


n
Assume now that the boundary of B lies in the union of the plates Ek , and v
k=1
is equal to δk in a neighbourhood of the plate Ek , k = 1, . . . , n. Then v ∈ Lip(B)

© Springer Basel 2014 305


V.N. Dubinin, Condenser Capacities and Symmetrization in Geometric Function
Theory, DOI 10.1007/978-3-0348-0843-9
306 Appendix

and the function u has continuous second derivatives in B. Using Green’s formula
(Theorem 1.6) we obtain
   
∂(v − u) ∂u ∂(v − u) ∂u ∂u
+ dxdy = (v − u) ds. (A1)
∂x ∂x ∂y ∂y ∂n
B ∂B

Since u and v have the same boundary values, the integral on the right-hand side
of (A1) vanishes and we arrive at the inequality
I(v, B)  I(u, B).
In view of the freedom which we have in the choice of v, we conclude from Lemma
1.2 that
cap C  I(u, B).

n
If (∂B) \ Ek is nonempty, then we take an exhaustion {Bn }∞ n=1 of B by do-
k=1
mains Bn each of which is bounded by finitely many analytic Jordan curves. An
admissible function v for the condenser C belongs to the class Lip(Bn ), and u is
harmonic in Bn . Hence we have (A1) for Bn in place of B, n = 1, 2, . . .. Note that

n
if ζ is an end-point of a connected component of the set Ek , then |∇u(z)| =
k=1
O(|z − ζ|−1/2 ) as z → ζ in a neighbourhood of ζ (see, for instance, [GarM], p. 455);
on the other hand u is harmonic at the other boundary points of B. Hence

∂u
lim (v − u) ds = 0
n→∞ ∂n
∂Bn

and we arrive at the inequality I(v, B)  I(u, B) again, which yields cap C 
I(u, B). 

A2. A uniqueness theorem for contractions


Recall that a map T of a compact set E ⊂ C into C is called a contraction if for
any points z, ζ ∈ E
|T (z) − T (ζ)|  |z − ζ|.
By Theorem 3.1,
cap E  cap T (E). (A2)
To describe the equality cases in (A2) we require some concepts and notations
used in potential theory [R], as well as several auxiliary results.
By definition, the potential pμ : C → [−∞, ∞) of a finite Borel measure μ
with compact support supp μ in C is the function

pμ (z) = log |z − ζ|dμ(ζ), z ∈ C.
A2. A uniqueness theorem for contractions 307

It is known that pμ (z) is a subharmonic function in C, which is harmonic in


C \ (supp μ) and satisfies the asymptotic equality

pμ (z) = μ(C) log |z| + O(|z|−1 ), z → ∞.

The energy of a finite Borel measure μ with compact support on C is the


following quantity:
 
I(μ) := log |z − ζ|dμ(z)dμ(ζ) = pμ (z)dμ(z).

A probability Borel measure ν on a compact set E is called the equilibrium


measure of E if
I(ν) = sup I(μ),

where the supremum is taken over all the probability Borel measures on E.
A set E ⊂ C is said to be polar if I(μ) = −∞ for each nontrivial finite Borel
measure μ with compact support in E. The expression ‘nearly everywhere’ means
‘everywhere away from a Borel polar subset’.
If E is a compact set, then it has an equilibrium measure ν ([R], p. 58), and
if E is not polar, then the equilibrium measure ν of E is unique and supp ν lies in
∂e E, the outer boundary of E ([R], p. 75). The logarithmic capacity of the compact
set E is the quantity
cap E = exp I(ν)

(here exp(−∞) = 0).


We denote the connected component of the complement C\ E of the compact
set E that contains the point at infinity by CE (∂CE = ∂e E).

Lemma A1. If E1 and E2 are nonpolar compact sets such that E1 ⊂ E2 and the
domain CE1 contains a nonpolar subset of E2 , then cap E2 > cap E1 .

Proof. Let cap E2 = cap E1 . Then the equilibrium measure for E1 coincides with
the equilibrium measure for E2 ; it will be denoted by ν. We look at the function
pν (z) in CE1 . Since supp ν ⊂ E1 , the function pν (z) is harmonic in CE1 . By
Frostman’s theorem pν (z)  I(ν) on C and pν (z) = I(ν) on E2 , with a possible
exception of a polar subset of E2 . By assumption CE1 contains a nonpolar subset of
E2 , so there exists a point ζ ∈ E2 ∩CE1 such that pν (ζ) = I(ν). Then pν (z) = I(ν)
in CE1 \ {∞} by the minimum principle, in contradiction with the representation
of the potential in a neighbourhood of infinity. 

Lemma A2. Let E be a nonpolar compact set and ν its equilibrium measure. Then
each point z in (∂e E) \ supp ν has a neighbourhood U (z) such that U (z) ∩ E is a
polar set.
308 Appendix

Proof. If z ∈ supp ν, then it has a neighbourhood U (z) such that ν(U (z) ∩ E) = 0.
Hence the support of ν is a subset of (∂e E) \ U (z) and cap E = cap((∂e E) \ U (z)).
If U (z) ∩ E is nonpolar, then we apply Lemma A1 to the compact sets E and
(∂e E) \ U (z), which yields cap E > cap((∂e E) \ U (z)). Thus U (z) ∩ E is polar. 

Lemma A3. Let E and ν be as in Lemma A2, and let z and ζ be points on the outer
boundary of E such that the straight line interval (z, ζ) lies in the complement of
CE. Then z and ζ lie in the support of ν.

Proof. Assume the contrary: for example, assume that z ∈ supp ν. Then by Lemma
A2 there exists a neighbourhood U (z) such that U (z) ∩ E is a polar set. In partic-
ular, U (z) ∩ ∂CE is polar. Then U (z) \ ∂CE is connected (see, for instance, [R],
p. 68). Picking points z1 ∈ CE and z2 ∈ (z, ζ) which lie in U (z) we connect
these by a continuous curve γ lying in U (z) \ ∂CE. By assumption z2 lies in the
complement of CE, which contradicts the connectedness of γ. 

Lemma A4. Let T (z) be a contraction which maps a compact set E onto a compact
set T (E), and let cap E = cap T (E) > 0. Then |T (z) − T (ζ)| = |z − ζ| for any
points z and ζ in the support of the equilibrium measure of E.

Proof. Let ν denote the equilibrium measure of T (E). By Theorem A.4.4 in [R]
there exists a Borel probability measure μ on E such that μ(T −1 ) = ν and for
each continuous function φ : T (E) → R,
 
φ(w)dν(w) = φ(T (z))dμ(z). (A3)
T (E) E

For fixed ω ∈ T (E) the function f (w) := log |w − ω| is upper semicontinuous and
bounded above on T (E). Furthermore, by Frostman’s theorem ([R], p. 59)

f (w)dν(w) = pν (ω)  I(ν) > −∞.
T (E)

Approximating f (w) by a monotone sequence of continuous functions and using


B. Levi’s theorem, from (A3) we conclude that

pν (ω) = log |T (z) − ω|dμ(z). (A4)
E

In its turn the function pν (ω) is upper semicontinuous and bounded above on
T (E), and we have 
pν (ω)dν(ω) = I(ν) > −∞.
T (E)
A2. A uniqueness theorem for contractions 309

Repeating the above argument for this function while taking equality (A4) into
account we obtain
  
I(ν) = pν (T (ζ))dμ(ζ) = log |T (z) − T (ζ)|dμ(z)dμ(ζ). (A5)
E E E

Now if μ is the equilibrium measure of E, then we use the equality cap E =


cap T (E), relation (A5), and the assumption that T is a contraction in succession,
which yields
 
I(μ) = I(ν)  log |z − ζ|dμ(z)dμ(ζ) = I(μ)  I(
μ).
E E

Since the equilibrium measure is unique, μ is an equilibrium measure of the com-


pact set E.
Now assume that for some points z0 and ζ0 in the support of μ we have the
strict inequality |T (z0 ) − T (ζ0 )| < |z0 − ζ0 |. As T is continuous, we can easily show
that there exist ε, 0 < ε < 1, and some neighbourhoods U (z0 ) and U (ζ0 ) such
that
|T (z) − T (ζ)| < ε|z − ζ|
for any z ∈ E ∩ U (z0 ) and ζ ∈ E ∩ U (ζ0 ). By (A5), since T is a contraction, it
follows that
 
I(ν)  log εdμ(z)dμ(ζ) + I(μ) = μ(U (z0 ))μ(U (ζ0 )) log ε + I(ν) < I(ν).
U (z0 ) U (ζ0 )

This contradiction completes the proof of Lemma A4. 


Theorem A2. For a compact set E and a contraction T : E → C equality in (A2)
holds if and only if T is an isometry nearly everywhere on E.
Proof. We can assume that cap E = 0. Let ν denote the equilibrium measure of
E. Let cap E = cap T (E) and let {Ui }, i = 1, 2, . . . be a countable system of open
sets forming a basis for the topology in C. By Lemma A2, each z ∈ ∂CE \ supp ν
has aneighbourhood Ui (z) in this basis such that Ui (z) ∩ E is polar. We set
B = (Ui (z) ∩ E), where the union is taken over all z ∈ ∂CE \ supp ν. Then B
z
is a polar Borel set. We claim that T is an isometry on E1 = E \ B. Let z and
ζ be points in E1 . If z ∈ ∂CE or ζ ∈ ∂CE, then we set z1 := z or ζ1 := ζ by
definition (then z1 , ζ1 ∈ supp ν). Now let z ∈ C \ CE and let (z1 , z2 ) denote the
largest interval on the line zζ that contains z and lies in C \ CE, so that z lies
between z1 and ζ. In a similar way, if ζ ∈ C \ CE, then let (ζ1 , ζ2 ) be the largest
interval on the line zζ that contains ζ and lies in C \ CE, so that ζ lies between z
and ζ1 . By Lemma A3, z1 and ζ1 lie in supp ν. In any case the points z and ζ lie
on a (closed) line interval [z1 , ζ1 ] with end-points z1 , ζ1 ∈ supp ν. By Lemma A4,
310 Appendix

|T (z1 )− T (ζ1 )| = |z1 − ζ1 |. Since T is a contraction, it easily follows that the points
T (z) and T (ζ) lie on the interval [T (z1 ), T (ζ1 )] and |T (z) − T (ζ)| = |z − ζ|.
Conversely, assume that there exists a set E1 ⊂ E such that the Borel set
E \ E1 is polar and T is an isometry on E1 . Taking the closure of E1 we can
assume that E1 is compact. Since the equilibrium measure ν of the compact set
E is concentrated on E1 , so that ν(E \ E1 ) = 0 ([R], p. 56), it follows that
cap E = cap E1 . Hence cap E = cap T (E1 )  cap T (E). In view of the reverse
inequality, cap E = cap T (E). 

A3. On separating transformations of sets


and condensers
In this appendix we consider only condensers C = (E0 , E1 ) such that the set D :=
C \ (E0 ∪ E1 ) has a nonpolar boundary. Note that if some connected component of
D has a polar boundary, then D coincides with this connected component and ∂D
is totally disconnected. In each connected component of D we define the potential
function ω of the condenser C to be equal to the solution of the ‘generalized’
Dirichlet problem [R] with boundary function h(ζ) = j, ζ ∈ Ej , j = 0, 1. The
function ω is harmonic in D, satisfies the inequality 0  ω  1, vanishes at the
regular boundary points of E0 , and is equal to 1 at the regular boundary points
of E1 . The irregular boundary points of D form a polar set. By the field G of
C we shall mean the union of the connected components of D whose boundary
contains regular points from both plates (here we diverge from the definition in
§ 1.2, where the field of a condenser is defined as the whole of D). It is known
that cap C = I(ω, G) (see, for instance, [Shl]). Let {Ck }nk=1 be the result of the
separating transformation of C with respect to the family of functions {pk }nk=1 .
(Throughout what follows we use the definitions and notation from § 4.3.) The
next result supplements Theorem 4.9 for an arbitrary (not necessarily admissible)
condenser C.
Theorem A3. In (4.9) equality is attained if and only if the field G of C satisfies

n
the condition G ⊂ Dk and the potential function of C is mirror-symmetric
k=1
relative to each analytic arc in (∂Dk ) ∩ G, 1  k  n, in a neighbourhood of this
arc. In particular, if C is a condenser with connected field and some intersection
(∂Dk ) ∩ G, where 1  k  n, contains a piece of a straight line or circle γ and if
equality holds in (4.9) then the condenser C is mirror-symmetric relative to γ up
to a polar set.

Proof. Let ω be the potential function of C. Then


n
cap C = I(ω, G)  I(ω, G ∩ Dk ). (A6)
k=1
A3. On separating transformations of sets and condensers 311

Let {Cj }∞ be a sequence of regular condensers approximating C, let ω j be the


j=1
 
potential function of Cj and Gj be its field. It is known that the functions ωj con-
verge uniformly to ω on compact subsets of G and lim I( j ) = I(ω, G) [Shl].
ωj , G
j→∞
We let {C jk }n denote the result of the separating transformation of C j with
k=1
respect to the family of functions {pk }k=1 . In follows from the definition of the
n

separating transformation that C jk ⊃ Ck , k = 1, . . . , n, j = 1, 2, . . . .


Let us introduce several auxiliary functions:
 
ω(p−1
k (ζ)), Re ζ  0, ωj (p−1
k (ζ)), Re ζ  0,
ωk (ζ) = −1

ω jk (ζ) = −1
ω(pk (−ζ)), Re ζ  0, ωj (pk (−ζ)), Re ζ  0.
As in the proof of Theorem 4.9, we arrive at the inequality

I(  j ∩ Dk ) = 1 I(
ωj , G  (k) )  1 cap C
ωjk , G jk  1 cap Ck .
j
2 2 2
Taking the limit as j → ∞ we obtain
1
I(ω, G ∩ Dk )  cap Ck . (A7)
2
Now (4.9) follows from (A6).
Assume that equality sign holds in (4.9). Then it follows from equality in (A6)
n 
n
that I(ω, C\ Dk ) = 0. If some point in the field G lies in the set C\ Dk , then
k=1 k=1
ω is constant in the corresponding connected component of G. This contradicts the

n
definition of G. Thus G ⊂ Dk . Since the Dirichlet integral is invariant under
k=1
Möbius transformations, we have I(ω, G ∩ Dk ) = 12 I(ωk , G(k) ), and we conclude
from equality in (A7) that
I(ωk , G(k) ) = cap Ck .
Using Theorem 1.13 we readily see that ωk (ζ) is a harmonic function in the
field G(k) .
Now let α be an analytic arc in (∂Dk ) ∩ G. By the uniqueness theorem,
in a neighbourhood of the line interval pk (α) ωk coincides with the analytic
extension of ω(p−1
k (ζ)). However, ωk is mirror-symmetric relative to pk (α). Hence
the function ω is mirror-symmetric relative to α in some neighbourhood of this
arc. In particular, if G is connected and α is a piece of a straight line or circle γ,
then by the uniqueness theorem ω is mirror-symmetric relative to γ and therefore
the condenser C is also mirror-symmetric relative to this curve up to a polar set.
Conversely, if the potential function of a condenser enjoys the symmetry

n
mentioned in the statement and G ⊂ Dk , then repeating the sufficiency proof
k=1
for the equality case in Theorem 4.9 almost word-for-word we obtain equality in
(4.9). 
312 Appendix

Now we look at a domain B with nonpolar boundary in C and let gB (z, z0 )


be the ‘generalized’ Green function with pole at z0 for this domain [R], [St]. Then
the inner radius r(B, z0 ) of B with respect to z0 is defined by the expansions

gB (z, z0 ) = − log |z − z0 | + log r(B, z0 ) + o(1), z → z0 , for z0 = ∞


and
gB (z, ∞) = log |z| + log r(B, ∞) + o(1), z → ∞.

Let {pk }nk=1 be an admissible family of functions for separating transformations


of open sets with respect to z0 and let {B (k) }nk=1 be the result of the separating
transformation of B with respect to this family. We define Bk as the connected
component of B (k) containing ζk := pk (z0 ), k = 1, . . . , n. By Theorem 4.10

n  βk2 /2
r(Bk , ζk )
r(B, z0 )  . (A8)
dk
k=1

The following result complements the description of the equality case in this the-
orem.

Theorem A4. Equality in (A8) holds if and only if the generalized Green function
of B with pole at z0 is symmetric relative to each analytic arc in (∂Dk ) ∩ B,
1  k  n, in a neighbourhood of this arc. In particular, if (∂Dk ) ∩ B, 1  k  n
contains a piece of a straight line or circle γ and equality holds in (A8), then B is
symmetric relative to γ up to a polar set.

Proof. Let λ > 0 be a number such that B(λ) = {z : gB (z, z0 ) > λ} is a domain
and γλ = {z : gB (z, z0 ) = λ} is a closed analytic curve. We look at the condenser
Cλ = (B(λ), C \ B). The function ωλ (z) = gB (z, z0 )/λ is the potential function of
Cλ , so cap Cλ = I(ωλ , B \ B(λ)). Using the Green function we obtain
 
∂ωλ 1 ∂g 2π
cap Cλ = − ωλ ds = − ds = .
∂n λ ∂n λ
γλ γλ

Hence λ = 2π/ cap Cλ .


It is clear that the Green function of B(λ) is equal to gB (z, z0 )−λ, so it follows
from the definition of the inner radius that log r(B(λ), z0 ) = log r(B, z0 )−λ. Hence

log r(B, z0 ) = log r(B(λ), z0 ) + 2π/ cap Cλ . (A9)

Assume that equality holds in (A8), that is,


n
β2 r(Bk , ζk )
k
log r(B, z0 ) = log .
2 dk
k=1
A3. On separating transformations of sets and condensers 313

Then in view of (A9),

2π β2
n
r(Bk , ζk )
k
log r(B(λ), z0 ) + = log .
cap Cλ 2 dk
k=1

Using inequality (A8) we obtain


n
β2 r(B(λ)k , ζk ) 2π β2 n
r(Bk , ζk )
k
log +  k
log , (A10)
2 dk cap Cλ 2 dk
k=1 k=1

where B(λ)k is the connected component of B (k) (λ) containing the point ζk , k =
1, . . . , n, and {B (k) (λ)}nk=1 is the result of the separating transformation of B(λ)
with respect to the family of functions {pk }nk=1 . Let Cλk denote the condenser
(B(λ)k , C \ Bk ). Then


log r(B(λ)k , ζk ) +  log r(Bk , ζk ), k = 1, . . . , n,
cap Cλk

by Theorem 2.7. After elementary transformations we obtain


n
n  
βk2 βk2 2π
log r(B(λ)k , ζk )  log r(Bk , ζk ) − .
2 2 cap Cλk
k=1 k=1

Taking (A10) into account yields


n
β2 k
(cap Cλk )−1  (cap Cλ )−1 .
2
k=1

λk }n be the result of the separating transformation of Cλ . Then by


Let {C k=1
(4.9)
1 βk
n n
cap Cλ  λk = 1
cap C
2
cap Cλk . (A11)
2 2 2 βk
k=1 k=1

Using Jensen’s inequality for the concave function f (x) = x−1 and recalling that
n
βk = 2 we obtain
k=1


n
β2 k
(cap Cλk )−1  (cap Cλ )−1 .
2
k=1

In view of the reverse inequality, we see that equality holds in (A11). In this case,
by Theorem A3 the function ωλ possesses symmetry indicated in the statement of
Theorem A4.
314 Appendix

Conversely, assume that gB (z, z0 ) possesses this symmetry. Let us introduce


the auxiliary functions

gB (p−1
k (ζ), z0 )/βk , Re ζ  0,
gk (ζ) = −1
k = 1, . . . , n.
gB (pk (−ζ), z0 )/βk , Re ζ  0,

For each k, 1  k  n, and any analytic arc α ⊂ (∂Dk ) ∩ B the analytic extension
of gB (p−1
k (ζ), z0 )/βk is symmetric relative to the interval pk (α) and therefore co-
incides with gk (ζ) in a neighbourhood of this interval. Hence gk (ζ) is harmonic in
Bk \ ζk . In view of its asymptotic behaviour in a neighbourhood of ζk , we conclude
that it is the Green function of Bk . Hence
 $
r(Bk , ζk ) = exp lim (gk (ζ) + log |ζ − ζk |)
ζ→ζk
 $
= dk exp lim βk−1 (gB (z, z0 ) + log |z − z0 |) = dk [r(B, z0 )]1/βk .
z→z0

This yields equality in (A8). 

A4. Invariance of the reduced modulus under


geometric transformations of domains
We shall expound on a general approach to uniqueness results for reduced moduli.
As applications we shall find necessary and sufficient conditions for the preserva-
tion of the reduced modulus in the situation when a domain or some distinguished
part of its boundary extend or are transformed by means of polarization or dis-
symmetrization.
In the notations of § 2.4 and under the assumption that the reduced modulus
(2.20) exists we look at the function


n
δk
u(z) = gB (z, zk , Γ),
νk
k=1

which we shall call the potential function for the domain B, set Γ, and tuples
Z, , and Ψ. In some neighbourhoods of the points zk

δk
u(z) = − log |z − zk | + ak + o(1), z → zk , z ∈ B, (A12)
νk
where
δk δl
n
ak = log r(B, Γ, zk ) + gB (zk , zl , Γ), k = 1, . . . , n.
νk νl
l=1
l=k
A4. Invariance of the reduced modulus 315

If zk = ∞, then some obvious modifications must be made in (A12). In our case u =


0 on (∂B)\Γ, with a possible exception of finitely many points. (If the boundary
∂B is smooth, we can easily see this, for example, by applying Hopf’s lemma to |u|

n
at some points in ∂B.) Let Br denote the difference B\ E(zk , ψk (r), B), where
k=1
E(zk , ψk (r), B), k = 1, . . . , n, is the closure in [B] of the connected component of
B ∩ U (zk , ψk (r)) adjoining the point zk .
Lemma A5. The following asymptotic formula holds:
 n 

n
σ(zk )δk2 σ(zk )δk  δk 1
I(u, Br ) = −π(log r) +π log + ak + o(1),
νk νk νk μk
k=1 k=1
r → 0.
Proof. In view of the conformal invariance of the Dirichlet integral and Green’s
formula,

n 
∂u
I(u, Br ) = − u ds,
∂n
k=1γ (r)
k

where γk (r) = ∂E(zk , ψk (r), B) ∩ B, k = 1, . . . , n. Taking (A12) into account we


obtain
    
∂u δk δk
u ds = − log ψk (r) + ak + o(1) − + o(1) dϕ
∂n νk νk
γk (r) γk (r)
 
δk2 δk
= πσ(zk ) 2 (log μk + νk log r) − ak + o(1), k = 1, . . . , n,
νk νk
which gives us the required result. 
We call a real-valued function v an admissible function for the domain B, set

n
Γ, and tuples Z, , Ψ if it is continuous in B\ {zk }, vanishes in Γ, satisfies the
k=1
Lipschitz condition on compact subsets of B and has the following expansions in
a neighbourhood of the zk :
δk
v(z) = − log |z − zk | + bk + o(1), z → zk , z ∈ B, (A13)
νk
where the bk , k = 1, . . . , n, are real constants.
Lemma A6. If the potential function u with the expansion (A12) and an admis-
sible functions v with the expansion (A13) satisfy v − u ≡ const in the domain

n
B\ {zk }, then there exists a positive number ε such that for sufficiently small
k=1
r > 0,

n
σ(zk )δk
ε < I(v, Br ) − I(u, Br ) − 2π (bk − ak ).
νk
k=1
316 Appendix

Proof. We pick a sufficiently small positive r0 so that I(v − u, Br0 ) = 0. Then for
all r  r0


n 
∂u
2ε := I(v − u, Br0 )  I(v − u, Br ) = I(v, Br ) + I(u, Br ) + 2 v ds
∂n
k=1γ (r)
k


n 
∂u
= I(v, Br ) − I(u, Br ) + 2 (v − u) ds,
∂n
k=1γ (r)
k

where the γk (r) k = 1, . . . , n, are as in the proof of Lemma A5. In view of (A12),
(A13), and the definition of the potential function u,
   
∂u δk
(v − u) ds = (bk − ak + o(1)) − + o(1) dϕ
∂n νk
γk (r) γk (r)

σ(zk )δk
= −π (bk − ak ) + o(1), r → 0,
νk

for each k = 1, . . . , n. 

Lemma A7. Let B, Γ, and Z = {zk }nk=1 , , Ψ, and also B  , Γ , and Z  = {zk }nk=1 ,
, Ψ be domains, sets, and tuples as in the definition of the reduced modulus such
that if zk ∈ B, then zk ∈ B  , and if zk ∈ ∂B, then zk ∈ ∂B  and σ(zk ) =
σ(zk ), 1  k  n. Let u and u be the potential functions for B, Γ, Z, , Ψ
and B  , Γ , Z  , , Ψ, respectively. Assume that there exists a function v  which is
admissible for B  , Γ , Z  , , Ψ and satisfies bk = ak for k = 1, . . . , n, (here the bk
and ak , k = 1, . . . , n, are the constants in the expansions (A13) for v  and (A12)
for u), and
I(v  , Br )  I(u, Br ) (A14)
for all sufficiently small r > 0. Then

M (B, Γ, Z, , Ψ) = M (B  , Γ , Z  , , Ψ) (A15)


n
if and only if v  ≡ u in the domain B  \ {zk }.
k=1


n
Proof. Assume that (A15) holds, but v  ≡ u in B  \ {zk }. Then v  − u ≡ const
k=1
in this domain because v  and u coincide on Γ . By Lemma A6, for some ε > 0
and sufficiently small r


n
σ(z  )δk
ε < I(v  , Br ) − I(u , Br ) − 2π k
(bk − ak ).
νk
k=1
A4. Invariance of the reduced modulus 317

Applying Lemma A5 to the integrals I(u, Br )  I(v  , Br ) and I(u , Br ) and taking
account of the equalities bk = ak and σ(zk ) = σ(zk ), k = 1, . . . , n, we conclude
that
n
σ(zk )δk
n
σ(zk )δk 
ak < ak .
νk νk
k=1 k=1

Hence formula (2.20) for the reduced modulus yields the inequality

M (B, Γ, Z, , Ψ) < M (B  , Γ , Z  , , Ψ),

which contradicts (A15).



n
Conversely, assume that v  ≡ u in the domain B  \ {zk }. Then ak = ak ,
k=1
k = 1, . . . , n, which yields (A15) in view of formula (2.20). 
Remark A1. Lemmas A5–A7 hold also in the case when zk = ∞ for some k: then
in the corresponding expansion (A12) or (A13)the quantity log |z − zk |, 1  k  n
must be replaced with − log |z|. These lemmas also hold when Γ = ∂B for some
domain B with the classical Green function: the proofs remain the same.
Remark A2. In Lemma A6 and therefore also in Lemma A7 we can replace Lips-
chitz continuity on compact subsets of B in the definition of an admissible func-
tion v by the analogous requirement in a subdomain B0 ⊂ B containing the sets
E(zk , ψk (r), B) with small r > 0 and the additional condition v = 0 in B\B0 . The
proofs remain the same.
Now let B = C. As above, we introduce the potential function for the Rie-
mann sphere and the tuples Z, , and Ψ:


n
 δk
u(z) = − log |z − zk |,
νk
k=1

and keep our former notation for this function. In some neighbourhood of the
points zk we have the expansions (A12), in which


n
δl
ak = − log |zk − zl |, k = 1, . . . , n.
νl
l=1
l=k

It is clear how (A12) must be modified in the case when zk = ∞ for some k,
1  k  n. An admissible function v for the complex sphere and the tuples Z, ,
and Ψ is introduced similarly to the case when B = C and the expansion (A13)
holds as before.
Lemma A8. Let Z = {zk }nk=1 , , Ψ and Z  = {zk }nk=1 , , Ψ be tuples as
in the definition (2.17) of the reduced modulus of the Riemann sphere, and let
318 Appendix


n
δk /νk = 0. Let u and u be the potential functions for Z, , Ψ and Z  , , Ψ,
k=1
respectively. Assume that there exists a function v  which is admissible for Z  , ,
Ψ and has the following properties: bk = ak for k = 1, . . . , n and

I(v  , Cr )  I(u, Cr )

for sufficiently small r > 0 (here the bk and ak , k = 1, . . . , n, are the constants in
the expansions (A13) for v  and (A12) for u).Then the equality

M (Z, , Ψ) = M (Z  , , Ψ)

n
holds if and only if v  ≡ u + const in the domain C\ {zk }.
k=1

We leave out the proof of these results since it is perfectly similar to the
proof of Lemma A7 (including the proofs of the analogues of Lemmas A5 and
A6). The main difference consists in the use of formula (2.17) in place of (2.20)
for the reduced modulus.
We also consider the case when B is a simply connected domain of hyperbolic
type and Γ = ∅. Using the notations in Theorem 2.13 we define the potential
function as follows:
δk
n+s
u(z) = − log |f (z) − w
k |.
νk
k=1

Note that u is defined up to a constant depending on f . An admissible function v is


defined similarly to (A13), with a single difference: there is no condition that v = 0
on Γ since Γ = ∅. Repeating the proof of Lemmas A5–A7 and taking Theorem
2.13 into account we arrive at the following result.
Lemma A9. Suppose that the assumptions of Theorem 2.13 hold for a domain B
and tuples Z = {zk }nk=1 , , Ψ and for a domain B  and tuples Z  = {zk }nk=1 ,
, Ψ (Γ = ∅) and furthermore, suppose that if zk ∈ B, then zk ∈ B  , but if
zk ∈ ∂B, then zk ∈ ∂B  and σ(zk ) = σ(zk ), 1  k  n. Let u and u be the
potential functions for B, Z, , Ψ and B  , Z  , , Ψ, respectively, and assume
that there exists an admissible function v  for B  , Z  , , Ψ such that bk = ak for
k = 1, . . . , n (here the bk and ak , k = 1, . . . , n, are the constants in the expansions
of the functions v  and u, respectively) and

I(v  , Br )  I(u, Br )

for sufficiently small r > 0. Then the equality

M (B, ∅, Z, , Ψ) = M (B  , ∅, Z  , , Ψ)

n
holds if and only if v  ≡ u + const in B  \ {zk }.
k=1
A4. Invariance of the reduced modulus 319

Remark A3. Under the assumptions of Lemma A7 (or Lemma A9), if the reduced
moduli in question are equal, then

I(v  , Br ) = I(u, Br ) + o(1), r → 0.

Indeed, we can see as in the proof of Lemma A6 that when the reduced
moduli are equal, we have

I(u , Br )  I(v  , Br ).

It remains to use the inequality I(v  , Br )  I(u, Br ) and Lemma A5.
Now we turn to uniqueness and start with the simplest property of the mono-
tonicity of the reduced modulus under certain modifications of the domain B and
the set Γ. The case Γ = ∂B was considered by Kovalev [K3] in full generality. So
in the three theorems below we limit ourselves to the case when B ⊂ C is a finitely
connected domain without isolated boundary points and Γ is a nonempty closed
subset of ∂B consisting of finitely many nondegenerate connected components. In
the case of a simply connected domain B it is possible that Γ = ∅. We shall also
assume that the tuples Z, , Ψ satisfy the conditions for the existence of the
relevant reduced moduli of the domain B as well as the tuples corresponding to
the extension B  of B do (Theorems A6 and A7).
Theorem A5. If Γ ⊃ Γ , then

M (B, Γ, Z, , Ψ) = M (B, Γ , Z, , Ψ)

if and only if Γ = Γ .

Proof. Sufficiency is obvious. Now let the reduced moduli be equal and let u be
the potential function for the domain B, set Γ, and tuples Z, , and Ψ. As Γ ⊃ Γ ,
u is also an admissible function for B, Γ , Z, , Ψ, and the assumptions of Lemma
A7 are satisfied. Hence u coincides with the potential function for B, Γ , Z, , Ψ
and therefore Γ = Γ . 

Theorem A6. Let B  be a domain obtained by extending the domain B across a part
Γ of its boundary, let Γ = (∂B  )\((∂B)\Γ), and let u be the potential function
for the sets B  , Γ and tuples Z, , Ψ. Then

M (B, Γ, Z, , Ψ) = M (B  , Γ , Z, , Ψ) (A16)

if and only if B = B  and u = 0 on Γ ∩ B  .

Proof. Let u be the potential function for B, Γ, Z, , Ψ. We look at the function



 u(z), z ∈ B,
v (z) =
0, z ∈ B  \B.
320 Appendix

Assume that (A16) holds. Then by Lemma A7 and in view of Remark A2, v  ≡ u

n
in the domain B  \ {zk }. Hence it follows, in particular, that B  \B is empty.
k=1
Indeed, if z0 ∈ B  \B, then z0 is a limit point of B  , but not of B. Hence some
a neighbourhood of z0 contains a point in B  whose small neighbourhood lies in
B  \B. Then u ≡ 0 by the uniqueness theorem, in contradiction with the definition
of the reduced modulus. Thus B  \B = ∅ and so B  = B. By continuity we also
n
have u ≡ u in B  \ {zk }. Hence u = 0 on Γ ∩ B  . Conversely, let B = B 
k=1
and let u = 0 on Γ ∩ B  . We can assume that B  is an analytic Jordan domain.
The extension of u − u to the points zk lying in B is a harmonic function in
B. Let z be a point on ∂B. If z ∈ B  , then in view of the definition of this
domain, z ∈ Γ and therefore u (z) − u(z) = 0. If z ∈ B  , then necessarily z ∈ ∂B 
and two cases are possible: either z ∈ Γ or z ∈ Γ . In the first case it follows
from the definition of Γ that z ∈ Γ and u (z) − u(z) = 0 again. In the second case
z ∈ (∂B)\Γ and then ∂(u (z)− u(z))/∂n = 0 everywhere with a possible exception
of finitely many points. From the generalized modulus principle and Hopf’s lemma
we conclude that u ≡ u in B. In view of (2.20), Theorem 2.2, and equality (A12),
this yields (A16). 

Theorem A7. If B  is a domain obtained by extending the domain B across the



n
part ((∂B)\Γ)\ {zk } of its boundary so that the inner angles with vertices at
k=1
the boundary points zk , 1  k  n, keep their magnitudes, then

M (B, Γ, Z, , Ψ) = M (B  , Γ, Z, , Ψ) (Π17)

if and only if B = B  and ((∂B)\Γ)∩B  consists of finitely many piecewise smooth


curves such that the potential function u of the sets B  , Γ and tuples Z, , Ψ
satisfies ∂u /∂n = 0 at the interior points of these curves.

Proof. Let u and u be the potential functions for B, Γ, Z, , Ψ and B  , Γ,


Z, , Ψ and let ω be the restriction of u to B defined by continuity at the
n
boundary points. Suppose (A17) holds. Then ω ≡ u + const in B\ {zk } by
k=1
Lemma A7 or Lemma A9. Moreover, B = B  . In fact, as in the proof of the
previous theorem, we can verify that otherwise B  \B contains interior points. By
Remark A3, I(u , B  \B) = 0, and therefore u ≡ const, in contradiction with
the definition of the potential function. Now let z0 ∈ γ := ((∂B)\Γ) ∩ B  be an
arbitrary point distinct from the branch points of level curves of u , and let v
be the harmonic function conjugate to u in a neighbourhood of z0 . Since this
function is also conjugate to u in B, the set γ is a curve on which v = const.
Hence γ is a smooth curve in a neighbourhood of z0 and ∂u /∂n = 0 on this curve.
Conversely, let B = B  and let ∂u /∂n = 0 at the smooth points of γ. We can
assume that B  is an analytic Jordan domain. We take the harmonic function u −u
A4. Invariance of the reduced modulus 321

in B. By definition u − u = 0 on Γ. If z ∈ (∂B)\Γ and z ∈ ∂B  , then, again by


definition, ∂(u (z) − u(z))/∂n = 0 apart from finitely many points. If z ∈ (∂B)\Γ
and z ∈ ∂B  , then since B = B  , we have z ∈ γ and ∂(u (z) − u(z))/∂n = 0
at the smooth points. By the maximum principle and in view of Hopf’s lemma
u ≡ u + const in B, and therefore (A17) holds. 

Following § 3.2, by the polarization of a set A ⊂ C with respect to the real


axis we mean the set

P A = {z : (Im z  0) ∧ [(z ∈ A) ∨ (z ∈ A)]}


∪ {z : (Im z  0) ∧ (z ∈ A) ∧ (z ∈ A)} .

Theorem A8. Assume that B and P B are domains with classical Green functions,
let Z = {zk }nk=1 be a system of different points in B, = {δk }nk=1 be a tuple of
positive numbers, and let Ψ = {ψk (r)}nk=1 with ψk (r) ≡ μk rνk , where μk and νk ,
k = 1, . . . , n, are arbitrary positive numbers. Let Z  = {zk }nk=1 be the system of
points defined as follows: zk = P zk if Z does not contain the point symmetric to
zk relative to the real axis; and if there exists such a point (say, zl ), then zk = P zk
in the case when δk νl > δl νk , zk = P zk in the case when δk νl < δl νk , and zk = zk
and zl = zl in the case when δk νl = δl νk , 1  k  n. Then

M (B, ∂B, Z, , Ψ) = M (P B, ∂P B, Z  , , Ψ) (Π18)

if and only if u(z) ≡ u (z) or u(z) ≡ u (z), where u and u are the potential
functions for the sets B, ∂B and the tuples Z, , and Ψ and for the sets P B,
∂P B and the tuples Z  , , and Ψ, respectively.

Proof. We set u and u to be equal to zero outside the original domain of definition.
Note that in our case, for instance, u(z) has the following form:


n
δk
u(z) = gB (z, zk ),
νk
k=1

where the gB (z, zk ) are the Green functions of B with poles at the points zk ,
respectively, k = 1, . . . , n. We set

 min(u(z), u(z)), Im z  0,
v (z) =
max(u(z), u(z)), Im z  0.

The function v  is admissible for the sets P B, ∂P B and tuples Z  , , and Ψ, and
we have
I(v  , (P B)r ) = I(u, Br ).
The other assumptions of Lemma A7 are also fulfilled by the definition of the set
Z  (Γ = ∂B, Γ = ∂P B). As we noted above, Lemma A7 holds also for such Γ
322 Appendix


n
and therefore we have (A18) if and only if v  ≡ u in P B\ {zk }. We conclude
k=1
from the definition of v  and the uniqueness theorem for harmonic functions that
either u(z) ≡ u (z) or u(z) ≡ u (z). 

Theorem A9. Let G be a mirror-symmetric domain relative to the real axis and
let E be a closed set such that E ⊂ G, the sets G\E and P (G\E) are finitely
connected domains with complements consisting of nondegenerate components, and
the sets Γ := E ∩ ∂(G\E) and Γ := {z : z ∈ P E} ∩ ∂P (G\E) consist of finitely
many nondegenerate connected components. Let Z = {zk }nk=1 be a set of different
admissible points in G\E, = {δk }nk=1 be a tuple of positive numbers, let Ψ =
{μk rνk }nk=1 , where μk > 0 and νk > 0, and let Z  = {zk }nk=1 be the set defined as
in Theorem A8. Then

M (G\E, Γ, Z, , Ψ) = M (P (G\E), Γ , Z  , , Ψ)

if and only if u(z) ≡ u (z) or u(z) ≡ u (z), where u and u are the potential
functions for G\E, Γ, Z, , Ψ and P (G\E), Γ , Z  , , Ψ, respectively.

Proof. This can be proved by reproducing formally the proof of Theorem A8.
The difference is that now the potential function u is a linear combination of
functions of the form gG\E (z, zk , Γ), 1  k  n. Again, we use Lemma A7, and
use significantly the mirror symmetry of G to prove that the assumptions of that
lemma about the boundary points zk ∈ Z are satisfied. 

Theorem A10. Let B ⊂ C be a finitely connected domain without isolated boundary


points that is mirror-symmetric relative to the real axis. Let Γ be a nonempty closed
subset of ∂B consisting of finitely many nondegenerate connected components,
which is also mirror-symmetric relative to the real axis. (If B is simply connected,
then it is possible that Γ = ∅.) Let Z = {zk }nk=1 be a set of different admissible
points in B\Γ, = {δk }nk=1 be a tuple of nonzero real numbers, and let Ψ =
{ψk (r)}nk=1 , where ψk (r) ≡ μk rνk , with μk > 0 and νk > 0, k = 1, . . . , n,. Let
Z  = {zk }nk=1 be the set of points defined as follows: for fixed k, 1  k  n, if Z
does not contain the point symmetric to zk relative to the real axis, then zk = P zk
for δk > 0 and zk = P zk for δk < 0. But if Z contains such a point (say zl ), then
zk = P zk in the case when δk νl > δl νk , zk = P zk in the case when δk νl < δl νk ,
and zk = zk and zl = zl in the case when δk νl = δl νk . Then

M (B, Γ, Z, , Ψ) = M (B, Γ, Z  , , Ψ)

if and only if zk = zk for all k = 1, . . . , n or zk = zk for all k = 1, . . . , n. (If Γ = ∅,


n n
then by definition δk /νk = 0 if B = C and σ(zk )δk /νk = 0 if B = C; recall
k=1 k=1
that σ(zk ) is the exponent of the admissible point zk for zk ∈ ∂B and σ(zk ) = 2
for zk ∈ B.)
A4. Invariance of the reduced modulus 323

Proof. This is similar to the proof of Theorem A8. For Γ = ∅ we use Lemma A7.
If Γ = ∅ and B = C, then we use Lemma A8, and if Γ = ∅ and B = C then we
use Lemma A9. 

Uniqueness theorems for circular and Steiner symmetrization can be reduced


to uniqueness theorems for polarization, so we do not discuss these questions in
detail and turn to the next transformation, dissymmetrization. Using the notation
in § 4.4 we can state the following result.
Theorem A11. Let B be a domain symmetric with respect to the group Φ, and
assume that one of the following conditions hold :
a) B and Dis B are finitely connected domains without isolated boundary points;
then assume that Γ and Dis Γ are nonempty closed subsets of ∂B and ∂DisB,
respectively, which consist of nondegenerate connected components and that
Γ is Φ-symmetric;
b) B and Dis B are domains in C with classical Green functions and Γ = ∂B;
c) B = C and Γ = ∅;
d) B and Dis B are simply connected domains of hyperbolic type and Γ = ∅.
Let Z = {zk }nk=1 , = {δk }nk=1 , and Ψ = {ψk (r)}nk=1 be tuples of admissible
points, real numbers, and functions such that zk ∈ / ∂B ∩ ∂Pj for k = 1, . . . , n and
j = 1, . . . , N , ϕ(Z) = Z for each isometry ϕ ∈ Φ and moreover, if ϕ(zk ) = zl ,
then δk = δl and ψk (r) ≡ ψl (r), 1  k, l  n. In addition, assume that in each case
(a)–(d) all the necessary conditions for the existence of the corresponding reduced
modulus in § 2.4 are fulfilled. Then

M (B, Γ, Z, , Ψ) = M (DisB, DisΓ, {Diszk }nk=1 , , Ψ)

if and only if u(z) ≡ u (z0 + (z − z0 )eiθ ) + const, where u and u are the potential
functions for B, Γ, Z, , Ψ and DisB, DisΓ, {Diszk }nk=1 , , Ψ, respectively, and θ
is a real number.

Proof. We look at the potential functions u and u for B, Γ, Z, , Ψ and DisB,


DisΓ, {Diszk }nk=1 , , Ψ, respectively. It is easy to show that u is Φ-symmetric,
that is, u(z) ≡ u(ϕ(z)) for any ϕ ∈ Φ. Hence there exists a uniquely defined
function Dis u (see § 4.4) that is admissible for DisB, DisΓ, {Diszk }nk=1 , , Ψ, and
satisfies
I(Dis u, (DisB)r ) = I(u, Br ).
Using Lemma A7 in cases (a) and (b), Lemma A8 in case (c), and Lemma A9 in
case (d) we conclude that Dis u ≡ u + const, and therefore u ◦ λ−1 
k ≡ u + const
for some k, 1  k  N, (in cases (a) and (b) const = 0). 

The above proofs and a part of the proof of Theorem A6 show that we can
generalize Theorem A8 and case (b) of Theorem A11 by replacing the domain
P B (DisB) in their statements by a sufficiently ‘regular’ domain B  such that
324 Appendix

P B ⊂ B  (DisB ⊂ B  ). Then in case (b) of Theorem A11, to the necessary


and sufficient conditions for the equality of moduli we must add that DisB =
B  and u = 0 on the set (∂DisB) ∩ B  , where u is the potential function for
B  , ∂B  , {Diszk }nk=1 , , Ψ. In the case of Theorem A8 the conditions for equality
keep their form.

A5. Quadratic differentials


The contents of this appendix is borrowed from Jenkins’ book [J]; we leave out
proofs and give the simplified version for plane domains. Let G be a domain in
Cz . By a quadratic differential in G we mean the expression

Q(z)dz 2 , (Π19)

where Q(z) is a meromorphic function in G. If D ⊂ Cw is conformally equivalent


to a domain G and ϕ : D → G is a univalent conformal mapping, then we say that
the quadratic differential (A19) generates the quadratic differential


Q(w)dw 2
= Q(ϕ(w))(ϕ (w))2 dw2

in D by means of ϕ. A finite point z0 in G is called a zero or a pole of order n of


the differential (A19) if it is a zero or a pole, respectively, of the function Q(z).
In the case of the point at infinity, by definition a zero (a pole) of the differential
(A19) is a zero (a pole) of the quadratic differential

Q(1/z)(1/z 4)dz 2

at z = 0. For example, the quadratic differential

dz 2
Q(z)dz 2 =
z 2 (z − 1)

has a second-order pole at z = 0 and first-order poles at z = 1 and z = ∞. We


denote the set of zeros and simple poles of the quadratic differential (A19) by H1
and the set of its poles of order  2 by H2 . A maximal regular curve z(t) on which

Q(z)dz 2 ≡ Q(z(t))(z  (t))2 dt2 > 0

(Q(z)dz 2 < 0) if called a trajectory (an orthogonal trajectory, respectively) of the


quadratic differential (A19). It is clear that if Q(z)dz 2 generates a quadratic differ-

ential Q(w)dw 2
in D, then trajectories of Q(z)dz 2 are the ϕ-images of trajectories

of Q(w)dw .2

Theorem A12. (cf. [J], Ch. III). Let Q(z)dz 2 be a quadratic differential in a domain
G and z0 ∈ G. Then there exists a function z = ϕ(w) mapping the disc |w| < 1
A5. Quadratic differentials 325

conformally and univalently onto a neighbourhood V of z0 such that ϕ(0) = z0


and the quadratic differential Q(z)dz 2 on V generates by means of ϕ a quadratic

differential Q(w)dw 2
in |w| < 1, whose trajectories are defined by the following
equations:
a) Im w = c if z0 ∈ G \ (H1 ∪ H2 ),
b) Im w(2+n)/2 = c if z0 is a zero of order n,
c) Im w1/2 = c if z0 is a first-order pole,
d) Im (a+ib) log w = c if z0 is a second-order pole:

Q(w)dw 2
= (a + ib)2 w−2 dw2 ,
e) Im w(2−n)/2 = c if z0 is a pole of order n  3;
here a, b, and c are real constants.

Theorem A12 gives us a qualitative picture of the trajectories in a neigh-


bourhood of a point z0 ∈ G.
Now let G be a domain bounded by finitely many simple closed analytic
curves (maybe none at all). A quadratic differential in G (that is, in a neighbour-
hood of G is said to be positive if the boundary of G contains no poles of the
differential and is formed by the closures of some trajectories. We agree to treat
any quadratic differential on the Riemann sphere as positive. By an F -set with
respect to Q(z)dz 2 we means a subset K of G such that each trajectory of Q(z)dz 2
which intersects K lies in K entirely. The inner closure of E is defined as the
interior of the closure of E; it is denoted by E.D By a ring domain, a circle domain,
a strip domain, or an end domain for Q(z)dz 2 we mean a maximal connected open
F -set with the following properties.
A ring domain for Q(z)dz 2 contains no points in H1 ∪ H2 and is filled by
trajectories of Q(z)dz 2 which are closed Jordan curves. For ) a suitable choice of
a purely imaginary constant τ the function w = exp{τ Q(z)1/2 dz} takes this
domain conformally onto an annulus r1 < |w| < r2 .
A circle domain U for Q(z)dz 2 contains a unique double pole z0 of Q(z)dz 2
and U \ {z0 } is filled by trajectories of Q(z)dz 2 , which are closed Jordan curves,
each separating z0 from the boundary of U. )For a suitable choice of a purely
imaginary constant τ the function w = exp{τ Q(z)1/2 dz}, set equal to zero at
z0 , maps U conformally onto a disc |w| < R.
A strip domain P for Q(z)dz 2 contains no points in H1 ∪ H2 and is filled by
trajectories of the differential Q(z)dz 2 , each of which has the limit end-point in one
direction at some point z1 ∈ H2 and the limit end-point in the other )direction at
a point z2 ∈ H2 (not necessarily distinct from z1 ). The function ζ = Q(z)1/2 dz
maps the domain P conformally onto a strip a < Im ζ < b (where a and b, a < b,
are finite real numbers).
An end domain for Q(z)dz 2 contains no points in H1 ∪ H2 and is filled by
trajectories of Q(z)dz 2 , each of which has the limit end-points in both directions
) at
the same point z0 ∈ H2 . This domain is mapped conformally by ζ = Q(z)1/2 dz
326 Appendix

onto the upper or lower half-plane of the ζ-plane (depending on the choice of the
branch of the root function).

Theorem A13. Let Q(z)dz 2 be a positive quadratic differential in G ⊂ Cz and


assume that the configuration is neither one of the following ones nor can be
obtained from these by conformal equivalence: G = Cz and Q(z)dz 2 = dz 2 ;
G = Cz and Q(z)dz 2 = teiα dz 2 /z 2 where t > 0 and α is real. Let Φ denote the
union of all the trajectories of Q(z)dz 2 that have a limit end-point at some point
in H1 . Then G \ Φ consists of finitely many ring domains, circle domains, strip
domains, and end domains.
Each second-order pole of the differential Q(z)dz 2 has a neighbourhood lying
in a circle domain or in the inner closure of a union of finitely many strip domains,
and each pole of order n  3 has a neighbourhood lying in the inner closure of the
union of n − 2 end domains and finitely many strip domains (maybe none at all).

D of Φ is not necessarily empty [J]. The following result


The inner closure Φ
holds.

Theorem A14. If a quadratic differential Q(z)dz 2 on Cz has at most three different


D is empty.
poles, then for Φ as in the previous theorem the set Φ

For example we look at the quadratic differential on the sphere Cz in Exercise


6.1(10):
(z − α/(α − β)) 2
Q(z)dz 2 = − dz ,
z 2 (z − 1)2

where α and β are fixed positive numbers such that α < β. This differential has
a first-order zero at z = α/(α − β) < 0, a simple pole at infinity, and second-
order poles at z = 0 and z = 1. The set Φ D is empty and Cz \ Φ consists of two
circle domains. The function z = ϕ(w) := wn , where n > 1 is an integer, maps
the sector D := {w : −π/n < arg w < π/n} conformally and univalently onto
the half-plane Cw cut along the real negative half-axis. The quadratic differential
Q(z)dz 2 generates the quadratic differential

 2 n2 (wn − α/(α − β)) 2


Q(w)dw =− dw
w2 (wn − 1)2

in D by means of this function. In fact, this differential is defined in the whole of


Cw (cf. the description of an extremal configuration in Theorem 6.12). The picture
of the trajectories of this differential is clear from the structure of the trajectories
of Q(z)dz 2 (Figure 6.1). This connection between quadratic differentials also sim-
plifies the calculation of the products of powers of inner radii of circle domains in
the corresponding extremal problem.
A6. Unsolved problems 327

A6. Unsolved problems


We present below several unsolved problems in geometric function theory and
potential theory which are related to the subject of this book in one way or another.

1. Let LRE be the set obtained by the linear radial transformation of a compact
set E (see Exercise 3.1(4)). Is it true that

cap E  cap LRE ?

For some special sets E the contraction in § 3.1 answers this question in the
affirmative (Exercises 3.1(4–6)).
2. For fixed δ ∈ (0, 2) and disjoint nonempty compact subsets E0 and E1 of the
imaginary axis set

C5 (α) = (E0 (α), E1 (α)),


Ej (α) = {x + iy : x ∈ [−1, α] ∪ [α + δ, 1], iy ∈ Ej }, j = 0, 1,

where α ∈ (−1, −δ/2) (see the left-hand image in Figure 3.4). Show that the
function cap C5 (α) is strictly increasing on (−1, −δ/2). In the case when the
plates of the condenser C5 (α) are the reflections of each other in a straight
line (so that C5 (α) looks like C3 (α) in Figure 3.4) the solution follows from
the properties of polarization (Exercise 3.2(11)).
3. (Karp [DKarp]) Let ε ∈ (0, 1) and δ ∈ (0, 1 − ε) be real numbers and let E
be a compact subset of the imaginary axis. We set

C6 (α) = ({x + iy : x ∈ [−1, −α − δ] ∪ [−α, −ε], iy ∈ E},


{x + iy : x ∈ [ , α] ∪ [α + δ, 1], iy ∈ E}),

α ∈ (ε, 1 − δ). It follows from Theorem 3.4 that the capacity cap C6 (α) is
strictly increasing on (ε, (1 + ε − δ)/2). Is it also increasing on the rest of the
interval, [(1 + ε − δ)/2, 1 − δ)? This problem can be generalized for condensers
with asymmetric plates in the natural way.
4. (Solynin [Sol5]) Let f1 (x) and f2 (x) be continuous positive functions on an
interval a < x < b, where −∞  a < b  +∞; let f0 (x) := (f1 (x) + f2 (x))/2
and let Bk = {z : a < Re z < b, |Im z| < fk (x)}, k = 0, 1, 2. For each
x ∈ (a, b) show that

r(B0 , x)  r(B1 , x)r(B2 , x).

Equality sign holds if and only if f1 (x) ≡ f2 (x) on (a, b).


5. (Solynin [Sol5]) Let dn (E), 2  n  ∞, be the nth diameter of the compact
set E ([Kuz2], no. 3, § 4.7).
328 Appendix

(a) Find the supremum of the nth Jung radius over all the connected com-
pact sets with nth diameter equal to 1; that is, find the quantity
r(n) = max min R.
dn (E)=1 U (z,R)⊃E

b) Prove that the function r(n) increases from 3/3 to 2 for n increasing
from 2 to ∞.
The case n = 2 in (a) corresponds to Jung’s classical theorem, and for n = ∞
this problem was solved by Solynin.
6. (Krzyż [Sol5]) Let B be a simply connected domain in the plane C such that
the intersection of B with any horizontal line has linear measure at most
2a > 0, and the intersection of B with any vertical line has linear measure
at most 2b > 0. Let Π(a, b) := {z : |Re z| < a, |Im z| < b}. Prove that
r(B, z)  r(Π(a, b), 0) (A20)
for all points z ∈ B, with equality holding in (A20) if and only if B is a
translation of the rectangle Π(a, b) and z is the centre of B.
Solynin has conjectured that (A20) must hold for any domains B ⊂ C
(not necessarily simply connected).
7. (Solynin [Sol5]) Let B be a domain in C satisfying the following condition:
the intersection with B of no horizontal line contains a closed interval with
length 2a, and the intersection with B of no vertical line contains a closed
interval with length 2b. Prove that the densities of hyperbolic metrics ([Gol],
Ch. VIII, § 2) satisfy the inequality
λ(B, z)  λ(Π(a, b), 0) (A21)
for all z ∈ B. Equality holds in (A21) only in the case indicated in the
previous problem.
8. Does the inequality in Exercise 4.1(9) hold in the same conditions, apart from
he assumption that the domain D is symmetric relative to the line l∗ ?
9. (Solynin [Sol5]) Let γk , k = 1, . . . , n (n > 2), be rectifiable Jordan arcs with
fixed length l > 0 going out of the origin. Prove the inequality

n
cap γk  cap E ∗ ,
k=1

where E ∗ = {z : arg z n = 0, 0  |z|  l}. In addition, for γk ⊂ U, l < 1,


prove the inequality
n
caph γk  caph E ∗ .
k=1
In the case when the γk , k = 1, . . . , n, are line intervals, both inequalities
follow from the dissymmetrization principle.
A6. Unsolved problems 329

10. For a set of points a = (a1 , . . . , an ) on the unit circle Γ = {z : |z| = 1}, n  1,
and a compact set E ⊂ (0, 1] let


n
B(a, E) = U \ {z = tak : t ∈ E}.
k=1

For arbitrary sets of points a and   show that


a and arbitrary sets E and E

r(B(a, E) ∩ B(
a, E),  0),
 0)  r(B(a∗ , E) ∩ B(a∗ , E),

where a∗ is the set of nth roots of unity and a∗ consists of all the nth roots
of −1. For E = E  the above inequality follows from Theorem 4.16. For n = 1
the affirmative answer follows by polarization. This author knows the proof
in the case when n = 2 and the sets E and E  are closed intervals containing 1.
11. (Haliste [Hal]) In the notation of the previous problem, let ga (z) be the Green
function of the domain B(a, E) with pole at z = 0. For an arbitrary r ∈ [0, 1]
prove the inequality

2π 2π
ga (re )dθ 

ga∗ (reiθ )dθ.
0 0

12. (Baernstein [Bae4]) In the notation of Problem 10 let ωa (z) be the value of
the harmonic measure of the circle Γ with respect to the domain B(a, E) at
the point z. Show that for each increasing convex function Φ and 0  r  1

2π 2π
Φ(ωa (reiθ ))dθ  Φ(ωa∗ (reiθ ))dθ.
0 0

Baernstein proved this inequality for n  3. For n > 3 the question is open
even for E = [ρ, 1]. It is also not known if the weaker inequality ωa (0) 
ωa∗ (0) holds for arbitrary E = [ρ, 1] and n > 3.
13. Let Pn E be the projection of a compact set E onto the rays arg z n = 0 (see
the definition before Theorem 5.6). Is it true that

cap E  cap Pn E ?

The answer is affirmative for n = 2 (§ 3.1) and in the case when E is mirror
symmetric relative to the rays arg z n = π.
14. In the conditions of Exercise 5.3(1) (which is a special case of Theorem 5.10)
find the sharp lower bound for the logarithmic capacity in the case of odd n.
15. Assume that a component of the complement of a doubly connected domain
G ⊂ C contains points tk a∗k , k = 1, . . . , n and the other component contains
330 Appendix

points sk  a∗k , k = 1, . . . , n, such that tk and sk are positive numbers, the



ak are the different nth roots of unity, and the  a∗k are the nth roots of −1,
k = 1, . . . , n, n  2. Find the sharp upper bound for the modulus M G
in terms of tk and sk , if only in the case when tk = t and sk = s for all
k = 1, . . . , n.
16. Consider the product
n
rα (B0 , a0 ) r(Bk , ak ),
k=1

where B0 , B1 , . . . , Bn (n  2) are pairwise disjoint domains in C, a0 =


0, |ak | = 1, k = 1, . . . , n, and 0 < α  n. Show that it attains its maximum
at a configuration of domains Bk and points ak possessing rotational n-
symmetry. The proof is due to this author for α = 1 and to Kuz’mina [Kuz3]
for 0 < α < 1. Subsequently, Kovalev [K1] solved this problem under the
additional assumption that √ the angles between neighbouring line segments
[0, ak ] do not exceed 2π/ α.
17. Find the maximum value of the product
n
r(Bk , ak )
k=1

over various systems of nonoverlapping domains Bk ⊂ C and points ak ∈


Bk ∩ [−1, 1], k = 1, . . . , n, n > 4.
18. (Kuz’mina [Kuz2]) Show that the upper bound for the invariant I5 which was
established in Theorem 6.23 holds for arbitrary (not necessarily symmetric)
systems of points {ak }5k=1 .
19. In the notation of § 7.2, is it true that
  −1 C
1 −1
n
2πk n 1
Λf θ+  , f ∈ S, n > 3?
n n 4
k=1

For n = 2 this was proved by Szegő and for n = 3 by Reich and Schiffer.
20. For f ∈ M0 and n  2 is it true that
n   n  
2πk π 2πk
Λ−1
f θ + + Λ−1
f θ + +  4?
n n n
k=1 k=1

This inequality holds in the case when Λf (θ +2πk/n) = c  n 2, k = 1, . . . , n
(Theorem 7.12) and in the case when f maps the disc U onto a starlike
domain with respect to the origin (see Theorem 8.12). Theorem 7.11 yields
an analogous inequality, but for another kind of average.
A6. Unsolved problems 331

21. Find a sharp lower bound for Λf (0) in Exercise 7.2(5). The conjectural ex-
tremal function maps the disc U onto the union of the half-strip {w = u + iv :
u  0, |v| < π/4} and the half-plane Re w > 0, cut along a ray on the real
axis.
22. Let f (z) = z + a0 + a−1 /z + · · · be a function in the class Σ. For fixed a−1
and ϕ find the sharp upper estimate for the linear measure of the intersection
of C \ f ({z : |z| > 1}) and the straight line {w = teiϕ : t ∈ R}, 0  ϕ  2π
(see the comment on Corollary 7.15).
23. (Aseev [As]) Let (E0 , E1 ) be a condenser in R3 = {(x, y, z)}∪{∞} with plates
E0 and E1 which are linked closed curves in R3 (similarly to links of a chain).
Show that the conformal capacity of such a condenser is no smaller than that
of the condenser (E0∗ , E1∗ ), where E0∗ is the x-axis and E1∗ is a circle in the
yOz-plane with centre at the origin. (By ))) definition the conformal capacity
in R3 is the infimum of the integrals |∇v|3 dxdydz over all sufficiently
R3
smooth functions v vanishing in a neighbourhood of E0 and equal to 1 in a
neighbourhood of E1 .)
The last problem lies outside the framework of our book. On the other hand,
our approach and many of the results in Chapters 1–6 can be extended to higher-
dimensional spaces in the natural way (see, for instance, [D3] and [DPr]). The
same holds for some of the problems above. By contrast with other problems of
this kind, Aseev’s problem is essentially spatial and is of particular interest for
this reason.
Bibliography

[A] Yu.E. Alenitsyn, Univalent functions without common values in a multiply


connected domain, Tr. Mat. Inst. Steklov. 94 (1968), 4–18 (in Russian).
[Ach] N.I. Akhiezer, Elements of the theory of elliptic functions, Nauka, Moscow
(1970) (in Russian).
[Ah] D. Aharonov, Sequence of necessary conditions for univalence of a mero-
morphic function, Duke Math. J. 38 (1971), 595–598.
[AhKir] D. Aharonov and W.E. Kirwan, A method of symmetrization and appli-
cations, Trans. Amer. Math. Soc. 163 (1972), 369–377.
[Ahl1] L.V. Ahlfors, Bounded analytic functions, Duke Math. J. 14 (1947), 1–11.
[Ahl2] L.V. Ahlfors, Conformal invariants. Topics in geometric function theory,
McGraw-Hill Book Co., New York, (1973).
[AnVas] J.M. Anderson and A. Vasil’ev, Lower Schwarz-Pick estimates and angu-
lar derivatives, Ann. Acad. Sci. Fenn. Math. 33 (2008), 101–110.
[As] V.V. Aseev, Private communication (1988).
[B] T. Bagby, The modulus of a plane condenser, J. Math. and Mech. 17 (1967),
no. 4, 315–329.
[Bae1] A. Baernstein, II, Integral means, univalent functions and circular sym-
metrization, Acta math. 133 (1974), no. 3-4, 139–169.
[Bae2] A. Baernstein, II, A unified approach to symmetrization, Partial Differen-
tial Equations of Elliptic Type, Symposia Math. 35, A. Alvino et al., eds.,
Cambridge Univ. Press, Cambridge (1994), 47–91.
[Bae3] A. Baernstein, II, The ∗ -function in complex analysis, Handbook of com-
plex analysis: Geometric function theory 1, Elsevier, Amsterdam (2002),
229–271.
[Bae4] A. Baernstein, II, On the harmonic measure of slit domains, Complex Vari-
ables 9 (1987), 131–142.
[Bakh] A.K. Bakhtin, G.P. Bakhtina, and Yu.B. Zelinskii, Topologico-algebraic
structures and geometric methods in complex analysis, Institute of Math-
ematics, National Academy of Sciences, Kiev (2008).
[C] R. Courant, Dirichlet principle, conformal mapping, and minimal surfaces,
Interscience, New York (1950).

© Springer Basel 2014 333


V.N. Dubinin, Condenser Capacities and Symmetrization in Geometric Function
Theory, DOI 10.1007/978-3-0348-0843-9
334 Bibliography

[D1] V.N. Dubinin, Coverings of vertical segments under a conformal mapping,


Mat. Zametki 28, no. 1 (1980), 25–32; English transl. in Math. Notes 28
(1981), 476–480.
[D2] V.N. Dubinin, Transformation of functions and Dirichlet’s principle, Mat.
Zametki 38 (1985), no. 1, 49–55; English transl. in Math. Notes 38 (1985),
539–542.
[D3] V.N. Dubinin, Capacities and geometric transformations of subsets in n-
space, Geometric and Functional Analysis 3 (1993), no. 4, 342–369.
[D4] V.N. Dubinin, Symmetrization in the geometric theory of functions of
a complex variable, Uspekhi Mat. Nauk, 49 (1994), no. 1, 3–76; English
transl. in Russ. Math. Surveys, 49 (1994), no. 1, 1–79.
[D5] V.N. Dubinin, On the finite-increment theorem for complex polynomials,
Mat. Zametki, 88 (2010) no. 5, 673–682; English transl. in Math. Notes 88
(2010) no. 5, 647–654.
[D6] V.N. Dubinin, Boundary values of the Schwarzian derivative of a regular
function, Mat. Sb. 202 (2011) no. 5, 29–44; English transl. in Sb. Math.
202 (2011) no. 5, 649–663;
[D7] V.N. Dubinin, A new version of circular symmetrization with applications
to p-valent functions, Mat. Sb., 203 (2012) no. 7, 79–94; English transl. in
Sb. Math. 203 (2012) no. 7, 996–1011.
[D8] V.N. Dubinin, Markov-type inequality and a lower bound for the moduli
of critical values of polynomials, Dokl. Ross. Akad. Nauk 451 (2013) no. 5,
1–3; English transl. in Dokl. Math. 88 (2013) no. 1, 449–450.
[D9] V.N. Dubinin, Symmetrization of condensers and inequalities for functions
multivalent in a disk, Mat. Zametki, 94 (2013) no. 6, 846–856; English
transl. in Math. Notes 94 (2013) no. 6, 876–884.
[D10] V.N. Dubinin, On the Jenkins circles covering theorem for functions holo-
morphic in a disk, Zap. Nauchn. Semin. St. Peterburg. Otdel. mat. Inst.
Steklov. 418 (2013), 60–73 (in Russian).
[DKarp] V.N. Dubinin and D. Karp, Capacities of certain plane condensers and
sets under simple geometric transformations. Complex Variables and El-
liptic Equations 53 (2008), 607–622.
[Doob] J.L. Doob, Classical potential theory and its probabilistic counterpart,
Springer-Verlag, Berlin, 2001.
[DV] V.N. Dubinin and M. Vuorinen, On conformal moduli of polygonal quadri-
laterals, Israel J. Math. 171 (2009), 111–125.
[DPr] V.N. Dubinin and E.G. Prilepkina, Extremal decomposition of spatial do-
mains, Zap. Nauchn. Semin. St. Peterburg. Otdel. mat. Inst. Steklov. 254
(1998), 95–107; English transl. in J. Math. Sci. (New York) 105 (2001),
no. 4, 2180–2189.
Bibliography 335

[Dur] Duren P., Robin capacity, Computational methods and Function Theory
(CMFT’97). N. Papamichael, St. Ruscheweyh, E.B. Saff (eds.), World sci-
entific Publishing Co. (1999), 177–190.
[DurSch] P.L. Duren and M.M. Schiffer, Conformal mappings onto nonoverlapping
regions, in: Complex analysis, Birkhäuser Verlag, Basel–Boston, MA 1988,
27–39.
[E] E.G. Emel’yanov, On a connection between two problems on extremal de-
compositions, Zap. Nauchn. Sem. Leningr. Otd. Mat. Inst. Steklov. 169
(1987), 91–98; English transl. in J. Soviet Math. 52 (1990), 3063–3068.
[F] S.I. Fedorov, Maximum of a conformal invariant in the problem of nonover-
lapping domains, Zap. Nauchn. Sem. Leningr. Otd. Mat. Inst. Steklov. 112
(1981), 172–183; English transl. in J. Soviet Math. 25 (1984), 1093–1101.
[Fan] K. Fan, Distortion of univalent functions, J. Math. Anal. Appl. 66 (1978),
626–631.
[Geh] F.W. Gehring, A private letter to M. Vuorinen (1988).
[Gol] G.M. Goluzin, Geometric theory of functions of a complex variables, Nauka,
Moscow, 1966; English transl. Amer. Math. Soc., Prividence (RI), 1969.
[Gr] G. Grötzsch, Über einige Extremalprobleme der konformen Abbildung,
Ber. Verh. Sächs. Akad. Wiss. Leipzig, Math.-phys. Kl 80 (1928), 367–376.
[GResh] V.M. Gol’dshtein and Yu.G. Reshentyak, Introduction to the theory
of functions with generalized derivatives, and quasiconformal mappings
Nauka, Moscow, 1983.
[GarM] J.B. Garnett and D.E. Marshall, Harmonic measure, Cambridge Univ.
Press, Cambridge, 2005.
[H] W.K. Hayman, Multivalent functions, Cambridge Univ. Press. 2nd ed.
Cambridge, 1994.
[Hal] K. Haliste, On an extremal configuration for capacity, Arkiv for Mat. 27
(1989), no. 1, 97–104.
[Henr] P. Henrici, Applied and computational complex analysis, 3, Wiley-Inter-
science, New York, 1986.
[HurC] A. Hurwitz and R. Courant, Vorlesungen über allgemeine Funktionenthe-
orie und elliptische Funktionen, Springer-Verlag, Berlin–New York 1964
[HaLiP] G.G. Hardy, J.E. Littlewood, G. Pólya, Inequalities, Cambridge Univer-
sity Press, Cambridge 1988.
[J] J.A. Jenkins, Univalent functions and conformal mapping, Springer-Verlag,
Berlin–Göttingen–New York 1958, 1962.
[K1] L.V. Kovalev, On the problem of extremal decomposition with free poles
on a circle, Dal’nevostochnyi Mat. Sb. 2 (1996), 96–98 (in Russian).
[K2] L.V. Kovalev, On the inner radii of symmetric nonoverlapping domains,
Russian Mathematics (Iz. VUZ.) 44 (2000), 77–78.
336 Bibliography

[K3] L.V. Kovalev, Monotonicity of the generalized reduced module, Zap.


Nauchn. Sem. St. Peterburg. Otd. Mat. Inst. Steklov. 276 (2001), 219–236;
English transl. in J. Math. Sci. (New York) 118 (2003) no. 1, 4861–4870.
[Kl] H. Kloke, Some inequalities for the capacity of plane condensers, Results
in Math. 9 (1986), no. 1-2, 82–94.
[Kol] L.I. Kolbina, Some extremal problems in conformal mapping, Dokl. Akad.
Nauk SSSR 84 (1952), 865–868 (in Russian).
[Kr] J. Krzyż, Circular symmetrization and Green’s function, Bull. Acad.
Polon. Sci. Ser. Sci. Math. Astr. Phys. 7 (1959), 327–330.
[Kub] T. Kubo, Hyperbolic transfinite diameter and some theorems on analytic
functions in an annulus, J. Math. Soc. Jap. 10 (1958), no. 4, 348–364.
[Kuf] P.P. Kufarev, On conformal mappings of complementary domains, Dokl.
Akad. Nauk SSSR 73 (1950), 881–884 (in Russian).
[Kuz1] G.V. Kuz’mina, Moduli of families of curves and quadratic differentials,
Tr. Mat. Inst. Steklova 139 (1980), 1–240; English transl. in Proc. Steklov
Inst. Math. 139 (1982) 1–231.
[Kuz2] G.V. Kuz’mina, Methods of geometric function theory, I, II, Algebra i
Analiz 9 (1997), no. 3, 41–103; no. 5, 1–50; English transl. in St. Petersbg.
Math. J. 9 (1997), no. 3, 455–507; no. 5, 889–930.
[Kuz3] G.V. Kuz’mina, The method of extremal metric in extremal decomposition
problems with free parameters, Zap. Nauchn. Sem. St. Petersburg. Otd.
Mat. Inst. Steklov. 302 (2003), 52–67; English transl. in J. Math. Sci. (New
York) 129 (2005), no. 3, 3843–3851.
[L] G.F. Lawler, Conformally invariant processes in the plane, Amer. Math.
Soc. Math. Sur. and Mon. 114, 2005.
[Leb] N.A. Lebedev, The area principle in the theory of univalent functions,
Nauka, Moscow, 1975 (in Russian).
[LaSha] M.A. Lavrent’ev and B.V. Shabat, Methods of the theory of functions of
a complex variable, Nauka, Moscow 1973 (in Russian).
[M] M. Marcus, Radial averaging of domains, estimates for Dirichlet integrals
and applications, J. Anal. Math. 27 (1974), 47–78.
[Mit1] I.P. Mityuk, The symmetrization principle for multiply connected domains
and some of its applications, Ukrain. Math. Zh. 17 (1965), no. 4, 46–54.
[Mit2] I.P. Mityuk, Some properties of functions regular in a multiply connected
region, Dokl. Akad. Nauk SSSR 164 (1965), no. 3, 495–498; English transl.
in Sov. Math. Dokl. 6 (1965), 1252–1255.
[Mit3] I.P. Mityuk, Symmetrization methods and their application to geometric
function theory. An introduction to symmetrization methods, Kuban’ Uni-
versity Publishing House, Krasnodar 1980 (in Russian).
[Mit4] I.P. Mityuk, Application if symmetrization methods to geometric function
theory Kuban’ University Publishing House, Krasnodar 1985 (in Russian).
Bibliography 337

[Neh] Z. Nehari, Some inequalities in the theory of functions, Trans. Amer. Math.
Soc. 75 (1953), no. 2, 256–286.
[Oht1] M. Ohtsuka, Dirichlet problem, extremal length and prime ends, Van Nos-
trand, New York, 1970.
[Oht2] M. Ohtsuka, Extremal length and precise functions, Math. Sci. and Appl.
19, Tokio, 2003.
[Pom] Ch. Pommerenke, Boundary behaviour of conformal maps, Springer-
Verlag, Berlin, 1992.
[P] G. Pólya, Sur la symetrisation circulaire, C.R. Acad. Sci. Paris 230 (1950),
25–27.
[PS] G. Pólya and G. Szegő, Isoperimetric inequalities in mathematical physics,
Princeton University Press, Princeton (NJ) 1951.
[R] T. Ransford, Potential theory in the complex plane, Cambridge Univ.
Press, Cambridge, 1995.
[Rob] R.M. Robinson, On the transfinite diameters of some related sets, Math.
Z. 108 (1969), 377–380.
[Sch] M. Schiffer, Sur un problème d’extrémum de la représentation conforme,
Bull. Soc. Math. France 66 (1938), 48–55.
[Sin] V. Singh, Grunsky inequalities and coefficients of bounded schlicht func-
tions, Ann. Acad. Sci. Fenn. Math. 310 (1962) 1–21.
[Sobol] S.L. Sobolev, Some applications of functional analysis in mathematical
physics, Leningrad State University, Leningrad 1950; English transl. Amer.
Math. Soc., Providence (RI) 1991.
[S1] G. Szegő, Jahresber. Deutsch. Math. Ver. 32 (1923), 45.
[S2] G. Szegő, On a certain kind of symmetrization and its applications, Ann.
Mat. Pura ed Appl. Ser. 4, 40 (1955), 113–119.
[Sol1] A.Yu. Solynin, Isoperimetric inequalities for polygons and dissymmetriza-
tion, Algebra i Analiz 4 (1992), no. 2, 210–234; English transl. in St. Pe-
tersbg. Math. J. 4 (1993), no. 2, 377–396.
[Sol2] A. Yu. Solynin, Functional inequalities via polarization, Algebra i Analiz 8
(1996), no. 6, 148–185; English transl. in St. Petersb. Math. J. 8 (1997),
no. 6, 1015–1038.
[Sol3] A.Yu. Solynin, Modules and extremal metric problems, Algebra i Analiz
11 (1999), no. 1, 3–86; English transl. in St. Petersbg. Math. J. 11 (2000)
no. 1 1–65.
[Sol4] A.Yu. Solynin, Solution of the Polya–Szegő isoperimetric problem, Zap.
Nauchn. Sem. Leningr. Otd. Mat. Inst. Steklov. 168 (1988), 140–153. En-
glish transl. in J. Soviet Math. 53 (1991), 311–320.
[Sol5] A.Yu. Solynin, Private communication (2008).
338 Bibliography

[St] S. Stoı̈lov, Leçons sur les principes topologiques de la théorie des fonctions
analytiques Gauthier-Villars, Paris 1956.
[T1] P.M. Tamrazov, Capacities of condensers. The method of mixing signed
measures, Mat. Sbornik 115 (1981), no. 1, 40–73; English transl. in Math.
USSR. Sb. 43 (1982) 33–62.
[T2] P.M. Tamrazov, Two open problems connected with capacities, Lecture
Notes in Math. 1165 (1985), 292–293.
[Tei] O. Teichmüller, Untersuchungen über konforme und quasikonforme Abbil-
dung, Deutsche Math. 3 (1938), 621–678.
[Ts] M. Tsuji, Potential theory in modern function theory, Tokyo, 1959.
[Sh] A.V. Shubnikov, The problem of dissymmetry of material objects, Pub-
lishing House of the USSR Academy of Sciences, Moscow, 1961.
[Shl] V,A. Shlyk, A uniqueness theorem for the symmetrization of arbitrary
condensers, Sibirsk. Mat. J. 23 (1982), no. 2, 165–175; English transl. in
Sib. Math. J. 23 (1982), 267–276.
[Vas] A. Vasil’ev, Moduli of families of curves for conformal and quasiconformal
mappings, Lecture Notes in Math. 1788, Springer, 2002.
[W] V. Wolontis, Properties of conformal invariants, Amer. J. Math. 74 (1952),
no. 3, 587–606.
Index

admissible point, 28 principle, 9, 305


exponent of, 28 dissymmetrization, 114
almost domain
disc, 33 admissible, 25
half-disc, 33 circle, 325
asymptotically circular family of end, 325
domains, 33 ring, 325
strip, 325
capacity
‘half-plane’, 269
hyperbolic, 23 function
logarithmic, 23, 27 admissible, 13, 315
monotonicity of, 7, 15, 16 Bieberbach–Eilenberg, 294
composition principle, 17, 18, 56 Grötzsch, 29, 214
Green’s, 25, 26, 312
condenser, 6
Lipschitz, 1
admissible, 9
Neumann, 29, 45
decomposition of, 10
exhaustion of, 8 p-valent, 282
field of, 6, 12, 310 potential, 9, 240, 310, 314, 317,
generalized, 12 318
plate of, 6, 12 Robin, 27
potential function of, 13
regular, 10 Green’s formula, 4
condenser capacity, 6, 13
contraction, 61, 62
half-strip, 59
decomposition harmonic measure, 118, 329
of sets, 49
deformation, 238 lemma
Dirichlet Grötzsch’s, 10, 21, 50, 148
integral, 5 Teichmüller’s, 148

© Springer Basel 2014 339


V.N. Dubinin, Condenser Capacities and Symmetrization in Geometric Function
Theory, DOI 10.1007/978-3-0348-0843-9
340 Index

majorization principle, 277, 279, Schwarzian (derivative), 259


281, 282, 285, 286 set
Möbius invariants, 187, 188 admissible, 49
modulus regular, 49
of condenser, 17, 47 strip, 58
of doubly connected domain, 12, symmetrization
148 Aharonov–Kirwan, 110
of half-strip, 59 circular, 94
of quadrilateral, 20 elliptic, 101
of strip, 59 hyperbolic, 102
reduced, 47, 48, 53 Mityuk, 112
monotonicity parabolic, 102
of capacity, 15, 16 Steiner, 89
of inner radius, 31 Szegő–Marcus, 110
of reduced modulus, 48 with respect to circle, 96
symmetrization principle, x
neighbourhood of an attainable symmetry principle
point, 28 for condensers, 20
for reduced moduli, 55
polarization
of condensers, 63 theorem
of sets, 63 Fekete’s, 31
potential Robinson’s, 32
level of, 12 trajectory
of plate, 12 of half-strip, 159
projection of sets, 131 of strip, 159
onto rays, 132
transformation
Gonchar, 71
quadratic differential, 324 linear, 75, 76
orthogonal trajectory of, 324 linear averaging, 79
trajectory of, 324 linear radial, 79
quadrilateral, 15, 20 radial, 76, 77
radial averaging, 82, 83, 85
radius separating
inner, 26 of condensers, 103
Robin, 29 of sets, 106
ring spiral, 78
Grötzsch, 148 spiral averaging, 86
Teichmüller, 148 truncation, 2
List of Symbols

AK The Aharonov–Kirwan symmetrization, 110


AKL The symmetrization transformation obtained from AK by replacing the
logarithmic metric with the linear metric, 110
[B] The compactification of the domain B, 13
(B, E, ) A generalized condenser, 12
C The complex plane, 1
C The Riemann sphere, 1
|C| The modulus of the condenser C, 17, 47
cap C The capacity of the condenser C, 6, 13
cap E The logarithmic capacity of the set E, 23, 27
caph E The hyperbolic capacity of the set E, 23
C(r; B, Γ, Z, , Ψ) A condenser with fixed plate Γ ⊂ ∂B, 33, 38
C(r; B, ∅, Z, , Ψ) A condenser with all plates shrinking to points, 41
Cr B The circular symmetrization of the set B, 94
Dis The dissymmetrization, 114
D(z0 , r) An almost disc, 33
(E0 , E1 ) The condenser in C with plates E0 and E1 , 6
El The elliptic symmetrization, 101
gB (z, z0 ) The Green function of the set B with pole at z0 , 25, 26, 312
gB (z, z0 , Γ) The Robin function, 27
GC The Gonchar transformation of the condenser C, 71
G(P ) The Grötzsch ring, 148

© Springer Basel 2014 341


V.N. Dubinin, Condenser Capacities and Symmetrization in Geometric Function
Theory, DOI 10.1007/978-3-0348-0843-9
342 List of Symbols

G(z; R), G(z; c, n, R) The Grötzsch extremal functions, 29, 223


H(B, z, ζ) The symmetric difference, 157
Hyp The hyperbolic symmetrization, 102
I(B, A) A Möbius invariant, 187
In A Möbius invariant, 187
I(v, B) The Dirichlet integral of the function v over the set B, 4
J(B, A) A Möbius invariant, 187
K(·) The complete elliptic integral of the first kind, 134
K(B, z0 , ϕ) The ‘symmetric derivative’, 158, 264
K(R) An annulus, 26, 201, 286
LA {Ck }nk=1 The linear averaging transformation of the family of condensers
{Ck }nk=1 , 79
LC The linear transformation of the condenser C, 75
Λf (θ), Xf (θ), Λf (r, θ), Xf (r, θ) The distances from the origin to certain
sets, 217, 223, 250, 253
Lip(E) The class of Lipschitz functions on the set E, 1
LRE The linear radial transformation of the set E, 79
M (B, Γ, Z, , Ψ) A reduced modulus, 47, 53
M (B, ∅, , Ψ) A reduced modulus, 54
MB The Marcus radial transformation of the set B, 76
MC The Marcus radial transformation of the condenser C, 77
mE The linear Lebesgue measure of the set E, 71
MG The modulus of the ring (doubly connected domain) G, 12, 148
M, M0 , M0 , R, R0 , S, SB, SB0 Classes of univalent functions, 201
M (P ) A reduced modulus, 59
Mψ B The Mityuk spiral transformation of the set B, 78
Mψ C The Mityuk spiral transformation of the condenser C, 78
Mnψ The Mityuk spiral averaging symmetrization, 112
M (Q; a, b, c, d) The conformal modulus of a quadrilateral, 20
List of Symbols 343

M (θ, B) The quantity defined in (3.10), 77


M (θ, E) The quantity defined in (3.11), 77
m(θ, E) The linear measure of the intersection of the set E with the ray
arg z = θ, 79
Mpw0 (B, z0 ), Rp (K) Classes of multivalent functions, 279, 286

M (Z, , Ψ) := M (C, ∅, Z, , Ψ) A reduced modulus, 48


N (B, z0 , z ∗ ) The Neumann radius, 31
nB (z, z0 , z ∗ ) A Neumann function, 30
N (z, ζ) A Neumann function, 45
Pα− A (Pα+ A) The result of a polarization of the set A with respect to α, 63
Pα C The polarized condenser C, 63
Par The parabolic symmetrization, 102
Pn E The projection of the set E onto the rays arg z n = 0, 132
PQ E The projection of the set E onto the closed set Q, 131
(Q; a, b, c, d) A quadrilateral, 15
Q(z)dz 2 A quadratic differential, 324
R The real axis, 1
RA {Bk }nk=1 The radial averaging transformation of the family of sets
{Bk }nk=1 , 82
RA {Ck }nk=1 The radial averaging transformation of the family of condensers
{Ck }nk=1 , 83, 85
RB The symmetrization of the set B with respect to the unit circle, 96
r(B, Γ, z0 ) The Robin radius, 29
r(B, z0 ) The inner radius of the set B with respect to the point z0 , 26
Sf (z) The Schwarzian (derivative), 259
SML The symmetrization transformation obtained from SMA by replacing the
logarithmic metric with the linear metric, 111
SMA The Szegő–Marcus symmetrization, 110
St B The Steiner symmetrization of the set B, 89
344 List of Symbols

σ(z) The exponent of the admissible point z for z ∈ ∂B; σ(z) = 2 for
z ∈ B, 30
T(P ) The Teichmüller ring, 148
trun υ The truncation of the function υ, 2
U (∞, r) The exterior of the disc U (0, 1/r), 1
U (z0 , r) The open disc with centre at z0 = ∞ and radius r, 1
U = U (0, 1) The unit disc, 13, 201

You might also like