Sei sulla pagina 1di 358

International Workshop

on

QUASICONFORMAL MAPPINGS AND THEIR APPLICATIONS

December 27, 2005 - January 01, 2006


Indian Institute of Technology Madras
Department of Mathematics

Proceedings of the

International Workshop
on
Quasiconformal Mappings And Their Applications
(IWQCMA05)
December 27, 2005 - January 01, 2006

Edited by
S. Ponnusamy
T. Sugawa
M. Vuorinen
Co-organized by

• Chennai Mathematical Institute, Chennai, India

• Institute of Mathematical Sciences, Chennai, India

Sponsored mainly by

• National Board for Higher Mathematics (DAE),


India

• Forum d’Analystes, Chennai, India

• National Science Foundation, USA

• The Abdus Salam International Center for Theoretical Physics,


Italy

• Commission on Development and Exchanges of the International


Mathematical Union

• Indian National Science Academy, India

• Council of Scientific and Industrial Research, India

• Department of Science and Technology, India


Preface

The Department of Mathematics, IIT Madras, Chennai hosted International


Workshop on Quasiconformal Mappings and their Applications (IWQCMA05)
December 27, 2005- January 1, 2006. This event was the first one in India on
this active research area which has its roots in geometric function theory and
which is closely connected with several topics of mathematical analysis.
The organizers gratefully acknowledge the financial support of
1. National Board for Higher Mathematics (DAE), India
2. National Science Foundation, USA
3. The Abdus Salam International Center for Theoretical Physics, Italy
4. Commission on Development and Exchanges of the International Mathe-
matical Union, Italy
5. Indian National Science Academy, India
6. Council of Scientific and Industrial Research, India
7. Department of Science and Technology, India
8. Forum d’Analystes, Chennai, India.
We also thank
1. Theivanai Ammal College for Women, Viluppuram, India
2. Canara Bank, IIT Madras Campus
3. State Bank of India, IIT Madras Campus
for their support. Many people have given us help, in particular the students of
Prof. S.Ponnusamy. Prior to the start of the workshop, preworkshop lectures
were given by Dr. Antti Rasila and Prof. Raimo Näkki from Finland. We thank
both of them. Also, we take this opportunity to thank Prof. R. Balasubramanian,
Prof. R. Parvatham, and Prof. C. S. Seshadri for their continued encouragement
and helpful advice.
The participants, who represented many different countries, received in most
cases financial support from their national funding organizations to cover their
expenses. ICTP’s generous support was useful in supporting mathematicians
from developing countries. Without the invaluable support from the aforemen-
tioned organizations, this conference would not have been possible in its present
form.
The main goal of the conference was to bring together internationally well-
known experts representing geometric function theory and some related topics.
They were requested to deliver a series of lectures for postgraduate students on
their respective areas. The audience consisted of mathematicians ranging from
graduate students to well-known experts from all the participating countries.
Conformal invariance and conformally invariant metrics have been important
research topics in geometric function theory during the past century. These
topics also were discussed or mentioned in several of the lectures. The organizing
committee was pleased to observe that the lectures were very well received and
lead to many lively discussions afterwards. We were also pleased to receive
positive response from the speakers to our request to contribute their lectures
for the proceedings. It is our hope that the publication of these proceedings will
make the results presented in this Workshop and also this research area and its
challenging open problems more widely known for a wide readership than what
is the case presently. The editorial work was carried out at IIT Madras, and the
www-pages
http://mat.iitm.ac.in/ samy/
http://www.cajpn.org/madras/
http://www.math.utu.fi/proceedings/madras
contain a copy of these proceedings.
Special thanks to Mr. Swadesh Kumar Sahoo, and Mrs. P. Vasundhra for
their help in the organization of the meeting. We take this opportunity to thank
Prof. Roger W. Barnard for his support in getting NSF grant for supporting US
participants, and for its partial support in bringing out this proceedings.

On behalf of the Organizing committee

S. Ponnusamy
IIT Madras
T. Sugawa
Hiroshima University
M. Vuorinen
University of Turku
Preface
Contents
Roger W. Barnard, Clint Richardson, 1
Alex Yu. Solynin
A note on a minimum area problem for
non-vanishing functions

Alan F. Beardon and David Minda 9


The hyperbolic metric and geometric function theory

Peter Hästö 57
Isometries of relative metrics

David A Herron 79
Uniform spaces and Gromov hyperbolicity

Ilkka Holopainen, and Pekka Pankka 117


p-Laplace operator, quasiregular mappings,
and Picard-type theorems

Henri Lindén 151


Hyperbolic-type metrics

Williams Ma and David Minda 165


Geometric properties of hyperbolic geodesics

Olli Martio 189


Quasiminimizers and potential theory

R. Michael Porter 207


History and recent developments in techniques for
numerical conformal mapping

Antti Rasila 239


Introduction to quasiconformal mappings in n-space

Toshiyuki Sugawa 261


The universal Teichmüller space and related topics

Matti Vuorinen 291


Metrics and quasiregular mappings

G. Brock Williams 327


Circle packing, quasiconformal mappings,
and applications

List of Participants 349


Proceedings of the International Workshop on Quasiconformal
Mappings and their Applications (IWQCMA05)

A Note on a Minimum Area Problem for Non-Vanishing


Functions

Roger W. Barnard, Clint Richardson, Alex Yu. Solynin

Abstract. We find the minimal area covered by the image of the unit disk
for nonvanishing univalent functions normalized by the conditions f (0) =
1, f ′ (0) = α. We discuss two different approaches, each of which contributes
to the complete solution of the problem. The first approach reduces the prob-
lem, via symmetrization, to the class of typically real functions, where we can
employ the well known integral representation to obtain the solution upon
prior knowledge about the extremal function. The second approach, requiring
smoothness assumptions, leads, via some variational formulas, to a boundary
value problem for analytic functions, which admits an explicit solution.

Keywords. Symmetrization, Minimal Area Problem.


2000 MSC. 30C70.

Contents

1. Introduction 1
2. Outline of Our Method 4
3. The Iceberg Problem 6
References 8

1. Introduction
 Z 
p p p
Let D = {z : |z| < 1} and A = f analytic in D : |f (z)| dA = ||f ||Ap < ∞ ,
D
the Bergman space of analytic functions in D.
Recently, Aharanov, Beneteau, Khavinson, and Shapiro [2] considered a gen-
eral minimization problem on Ap
inf{||f ||Ap : f ∈ Ap , ℓi (f ) = ci , i = 1, . . . , n}
where ℓi are bounded linear functionals on Ap , p > 1. They proved several general
results about this problem.
Version October 19, 2006.
Supported by NSF grant DMS-0412908.
2 Roger W. Barnard, Clint Richardson, Alex Yu. Solynin IWQCMA05

As we know, in recent years tremendous progress has been achieved in the


study of Bergman spaces. For a detailed account of this progress, we refer to the
recent monograph by Peter Duren and Alex Schuster, Bergman Spaces, [7].
Aharanov, Beneteau, Khavinson, and Shapiro [2] also mentioned that to ob-
tain a complete solution of a particular problem, one often needs additional
information which does not follow from their methods.
A particular example is the following open problem:
Z 
2 ′
inf |f | dA : f 6= 0 in D, f (0) = 1, f (0) = α
D
This is a “typical” extremal problem on the class of non-vanishing analytic
functions. The nonlinearity of the class is the obvious obstacle here.
But, we have a method which allows us to solve some problems similar to this
one.
Let
Nα = {f : f is univalent, and non-vanishing on D,
f (z) = 1 + a1 (f )z + . . . ,
normalized by a1 (f ) = α }
The area of the image f (D) is given by
Z ∞
X
′ 2
D(f ) = |f | dA = π n|an (f )|2 .
D n=1
Thus
D(f ) ≥ πα2 ,
with equality iff f (z) = 1 + αz.
Since this map f is in Nα , 0 < α ≤ 1, Koebe’s 1/4 Theorem implies Nα = ∅
for α > 4. So the nontrivial range is 1 < α < 4 .
For the non-trivial range, the minimal area problem for Nα is solved by
Theorem 1.1. For 1 < α < 4, let f ∈ Nα . Then
 √ 2  √  √ 
(1.1) D(f ) ≥παa2 a + a2 − 1 αa2 − 2 a2 − 1 a + a2 − 1

where a = a(α) is the solution to


" #
1 √ √  √ 
(1.2) = a2 1 − a2 − a(a + a2 − a)3 log (a + a2 − 1)4 /16a2 (a2 − 1) ,
α
which is unique in the interval 1 < α < ∞.
Equality in (1.1) holds iff f = fα defined by
Z z p
−β ξ 2 − a2 dz
fα (z) =  p 2  p √ 
−1 ξ + z
ξ2 + 1 a ξ 2 − 1 + ξ a2 − 1
A Note on a Minimum Area Problem for Non-Vanishing Functions 3

100
2
80
πα

60
AHΑL
40

Α(α)
20

1 2 3 4
Α

Figure 1. The graph of A(α).

Α=3

-4 4 8

-4

Figure 2. The extremal domain Dα = fα (D) for α = 3.

ia 1−z
√ 
with ξ = 2

z
and β = αa2 a + a2 + 1 .

For 0 < α < 4, let


A(α) = min D(f )
f ∈Nα

denote the minimal area covered by the images of functions in the class Nα . Note
A(α) is convex and increasing. This can be proven from the formulas, geometry,
and variational arguments. See Figure 1.
4 Roger W. Barnard, Clint Richardson, Alex Yu. Solynin IWQCMA05

2. Outline of Our Method


First consider the minimal area problem on Tα , the typically real nonvanishing
functions (not necessarily univalent). Use the linear structure of Tα and refor-
mulate to show uniqueness and get simple “sufficient conditions” for extremality
corresponding to linearized functions. This gives
Theorem 2.1. For 1 < α < 4, let f ∈ Tα . Then (1.1) holds with the same cases
of equality.
The technique of this proof was developed earlier in [1]. What is missing is
how to construct the extremal function!
Next. Assuming sufficient smoothness, we can apply a variant of Julia’s Vari-
ational Formula in [5]. This leads to boundary conditions for an extremal analytic
function. To obtain this “conditional” solution requires a priori smoothness.
Next, to achieve the “required smoothness,” we exploit geometric control of
the mapping radius and apply standard symmetrization techniques to obtain the
sufficient initial Jordan rectifiability as in [4].
Then we can apply earlier smoothing variations developed by Barnard and
Solynin in [5] giving “required smoothness.”
Thus the “conditional” proof becomes a true proof. We then verify that
the function recovered from the first step satisfies the sufficient conditions of
extremality which also leads then to a complete solution of the problem.
For a first step on Tα , we renormalize so that
f (0) = 1, f ′ (0) = α.
Subordination implies 0 < α ≤ 4. Since Tα is compact and convex, the mini-
mizer exists and is unique. The uniqueness follows by letting f1 and f2 be two
minimizers. Then Z
1
(2.1) D((f1 + f2 )/2) = |f1′ + f2′ |2 dσ
4 D
Z Z 
1 ′ 2 ′ 2
≤ |f1 | dσ + |f2 | dσ
2 D D
1
= (D(f1 ) + D(f2 )) ,
2
with equality iff f1′ ≡ f2′ .
We note here that the uniqueness obtained here is fortunate, since uniqueness
is in general not obtained when variational and approximation methods are used.
Reformulating the problem using the linearity of Tα , we use the following
lemma from [1, 3]
Lemma 2.2. For fα′ continuous on D, fα minimizes D(f ) on Tα iff fα minimizes
Z
L(f ) = ℜ fα′ (z)f ′ (z) dσ
D
on Tα .
A Note on a Minimum Area Problem for Non-Vanishing Functions 5

Proof. See Lemma 1 of [3].


Lemma 2.3. If fα′ is continuous on D, then
Z π
L(f ) = Kα (t)dµf (t),
0
where
2πα  it ′ it
Kα (t) = ℑ e f (e ) .
sin t
Proof. See [3].

Proof of Theorem 2.1 for Tα . For Dα = fα (D), first show (0, 1] ⊂ Dα by


considering
1 fα (τ z)
fe(z) = 1 − +
τ τ
e
for τ < 1 and compare D(f ) with D(fα ). Then fα (−1) = 0 since fα is not
identically 0 and fα ∈ Tα .
Thus with Lemmas 2.2 and 2.3, fα minimizes D(f ) on Tα iff fα minimizes
L(f ) under the constraints Z π
2 dµf = 1
0
Z π  
2 t 2
sec dµf =
0 2 α
Now we can use well known results to show fα is extremal iff Kα satisfies
 
2 t
Kα (t) = λ0 + λ1 sec ∀ t ∈ Supp (µfα )
2
 
2 t
Kα (t) ≥ λ0 + λ1 sec ∀t ∈/ Supp (µfα ),
2
where λ0 , λ1 are real constants.
Long computations, see [3], show our fα gives Kα that satisfies these condi-
tions!

Next we characterize the geometry of extremal domains for Nα .


Lemma 2.4. 1. ∀ α, 1 < α < 4, an extremal fα minimizing D(f ) exists in
Nα .
2. If fα is extremal, then fα (D) is bounded, starlike with respect to 1, and
circularly symmetric with respect to rays
ℓτ = {z = x + iy : y = 0, x ≥ τ, ∀ τ, 0 ≤ τ ≤ 1}.
3. The min area A(α) := D(fα ) for 1 < α < 4.

Proof. Apply circular and radial symmetrizations, then polarizations similar to


arguments in [5].
6 Roger W. Barnard, Clint Richardson, Alex Yu. Solynin IWQCMA05

f Α=3
α
L fr
4

1 -4 4 8
L nf
-4

Figure 3. The free (Lf r ) and non-free (Lnf ) portions of the boundary.

Now combine Theorem 2.1 and Lemma 2.2 to see that if fα is extremal in Nα ,
then since fα ∈ Tα , Theorem 2.1 implies Theorem 1.1.
Next we show how the extremal fα in Nα can be recovered from its boundary
values.
Lemma 2.5. Let fα be extremal for Nα . Then f ′ is continuous on D and |f ′ | ≡
β ≥ α ∀ z ∈ ℓf r . See Figure 2.

Proof. Apply the deep “two point variation techniques” from [5] twice giving
f ′ these properties on ℓf r . Then use the Julia-Wolff Theorem and boundary
behavior properties from Pommerenke [6], giving f ′ these properties everywhere.

Lemma 2.6. If fα is extremal, ϕ(z) = log(zfα′ (z)) maps as described in Figure 2,


with
i(1 − z)
q1 (z) = ϕ √
2 sin( 20 ) z
Z ξ
t2 − b2
ϕ2 (ξ) = ci √ dt + s.
0 (t − a ) 1 − t
2 2 2

Long computations are used to show monotonicity, then we use line integral
formulae to compute the area as in [4].

3. The Iceberg Problem


A related problem is known as the Iceberg Problem: Given a fixed volume
above the water, how deep can the iceberg go? See Figure 3.
This problem can be modeled by supposing a slice I is a continuum in C and
E = {II : cap I = 1, area [II ∩ UHP] = α}.
A Note on a Minimum Area Problem for Non-Vanishing Functions 7

−1
0 τ α = log |f ’(1)|

q
1 ϕ
2
UHP

Figure 4. The mapping ϕ(z) = log(zfα′ (z)).

h?

Figure 5. The Iceberg Problem.

We anticipate using similar arguments to those in this paper to find


h = min min{ℑz : z ∈ I }.
I ∈E
8 Roger W. Barnard, Clint Richardson, Alex Yu. Solynin IWQCMA05

References
[1] D. Aharonov, H. S. Shapiro, and A. Yu. Solynin, Minimal area problems for functions with
integral representation, J. Analyse Math., to appear.
[2] Dov Aharonov, Catherine Bénéteau, Dmitry Khavinson, and Harold Shapiro, Extremal
problems for nonvanishing functions in Bergman spaces, Selected topics in complex analysis,
Oper. Theory Adv. Appl., vol. 158, Birkhäuser, Basel, 2005, pp. 59–86.
[3] R. W. Barnard, C. Richardson, and A.Yu Solynin, A minimal area problem for non-
vanishing functions, Analysis and Algebra, accepted.
[4] R.W. Barnard, C. Richardson, and A.Yu Solynin, Concentration of area in half planes,
Proc. Amer. Math. Soc. 133 (2005), no. 7, 2091–99.
[5] R.W. Barnard and A.Yu Solynin, Local variations and minimal area problems, Indiana
Univ. Math. J. 53 (2004), no. 1, 135–167.
[6] Ch. Pommerenke, Boundary behavior of conformal maps, Springer-Verlag, 1992.
[7] Alex Schuster and Peter Duren, Bergman spaces of analytic function, Mathematical Surveys
and Monographs, no. 100, American Mathematical Society, Providence, RI, 2004.

Roger W. Barnard, Clint Richardson, Alex Yu. Solynin Address: Department of


Mathematics and Statistics, Texas Tech University, Lubbock, Texas 79409
Address: Department of Mathematics and Statistics, Stephen F. Austin University, Nacog-
doches, Texas 75962
Address: Department of Mathematics and Statistics, Texas Tech University, Lubbock, Texas
79409
E-mail: roger.w.barnard@ttu.edu
E-mail: crichardson@sfasu.edu
E-mail: alex.solynin@ttu.edu
Proceedings of the International Workshop on Quasiconformal
Mappings and their Applications (IWQCMA05)

The hyperbolic metric and geometric function theory

A.F. Beardon and D. Minda

Abstract. The goal is to present an introduction to the hyperbolic metric


and various forms of the Schwarz-Pick Lemma. As a consequence we obtain
a number of results in geometric function theory.

Keywords. hyperbolic metric, Schwarz-Pick Lemma, curvature, Ahlfors Lemma.

2000 MSC. Primary 30C99; Secondary 30F45, 47H09.

Contents

1. Introduction 10
2. The unit disk as the hyperbolic plane 11
3. The Schwarz-Pick Lemma 16
4. An extension of the Schwarz-Pick Lemma 19
5. Hyperbolic derivatives 21
6. The hyperbolic metric on simply connected regions 24
7. Examples of the hyperbolic metric 28
8. The Comparison Principle 33
9. Curvature and the Ahlfors Lemma 36
10. The hyperbolic metric on a hyperbolic region 41
11. Hyperbolic distortion 45
12. The hyperbolic metric on a doubly connected region 47
12.1. Hyperbolic metric on the punctured unit disk 47
12.2. Hyperbolic metric on an annulus 49
13. Rigidity theorems 51
14. Further reading 54
References 55

Version October 19, 2006.


The second author was supported by a Taft Faculty Fellowship and wishes to thank the Uni-
versity of Cambridge for its hospitality during his visit November 2004 - April, 2005.
10 Beardon and Minda IWQCMA05

1. Introduction

The authors are writing a book, The hyperbolic metric in complex analysis,
that will include all of the material in this article and much more. The ma-
terial presented here is a selection of topics from the book that relate to the
Schwarz-Pick Lemma. Our goal is to develop the main parts of geometric func-
tion theory by using the hyperbolic metric and other conformal metrics. This
paper is intended to be both an introduction to the hyperbolic metric and a con-
cise treatment of a few recent applications of the hyperbolic metric to geometric
function theory. There is no attempt to present a comprehensive presentation of
the material here; rather we present a selection of several topics and then offer
suggestions for further reading.

The first part of the paper (Sections 2-5) studies holomorphic self-maps of
the unit disk D by using the hyperbolic metric. The unit disk with the hyper-
bolic metric and hyperbolic distance is presented as a model of the hyperbolic
plane. Then Pick’s fundamental invariant formulation of the Schwarz Lemma is
presented. This is followed by various extensions of the Schwarz-Pick Lemma
for holomorphic self-maps of D, including a Schwarz-Pick Lemma for hyperbolic
derivatives. The second part of the paper (Sections 6-9) is concerned with the
investigation of holomorphic maps between simply connected proper subregions
of the complex plane C using the hyperbolic metric, as well as a study of neg-
atively curved metrics on simply connected regions. Here ‘negatively curved’
means metrics with curvature at most −1. The Riemann Mapping Theorem is
used to transfer the hyperbolic metric to any simply connected region that is
conformally equivalent to the unit disk. A version of the Schwarz-Pick Lemma is
valid for holomorphic maps between simply connected proper subregions of the
complex plane C. The hyperbolic metric is explicitly determined for a number
of special simply connected regions and estimates are provided for general sim-
ply connected regions. Then the important Ahlfors Lemma, which asserts the
maximality of the hyperbolic metric among the family of metrics with curvature
at most −1, is established; it provides a vast generalization of the Schwarz-Pick
Lemma. The representation of metrics with constant curvature −1 by bounded
holomorphic functions is briefly mentioned. The third part (Sections 10-13) deals
with holomorphic maps between hyperbolic regions; that is, regions whose com-
plement in the extended complex plane C∞ contains at least three points, and
negatively curved metrics on such regions. The Planar Uniformization Theorem
is utilized to transfer the hyperbolic metric from the unit disk to hyperbolic
regions. The Schwarz-Pick and Ahlfors Lemmas extend to this context. The
hyperbolic metric for punctured disks and annuli are explicitly calculated. A
new phenomenon, rigidity theorems, occurs for multiply connected regions; sev-
eral examples of rigidity theorems are presented. The final section offers some
suggestions for further reading on topics not included in this article.
The hyperbolic metric and geometric function theory 11

2. The unit disk as the hyperbolic plane


We assume that the reader knows that the most general conformal automor-
phism of the unit disk D onto itself is a Möbius map of the form
az + c̄
(2.1) z 7→ , a, c ∈ C, |a|2 − |c|2 = 1,
cz + ā
or of the equivalent form
z−a
(2.2) z 7→ eiθ , θ ∈ R, a ∈ D.
1 − āz
It is well known that these maps form a group A(D) under composition, and that
A(D) acts transitively on D (that is, for all z and w in D there is some g in A(D)
such that g(z) = w). Also, A(D, 0), the subgroup of conformal automorphisms
that fix the origin, is the set of rotations of the complex plane about the origin.
The hyperbolic plane is the unit disk D with the hyperbolic metric
2 |dz|
λD (z)|dz| = .
1 − |z|2
This metric induces a hyperbolic distance dD (z, w) between two points z and w
in D in the following way. We join z to w by a smooth curve γ in D, and define
the hyperbolic length ℓD (γ) of γ by
Z
ℓD (γ) = λD (z) |dz|.
γ
Finally, we set
dD (z, w) = inf ℓD (γ),
γ
where the infimum is taken over all smooth curves γ joining z to w in D.
It is immediate from the construction of dD that it satisfies the requirements
for a distance on D, namely
(a) dD (z, w) ≥ 0 with equality if and only if z = w;
(b) dD (z, w) = dD (w, z);
(c) for all u, v, w in D, dD (u, w) ≤ dD (u, v) + dD (v, w).
The hyperbolic area of a Borel measurable subset of D is
Z Z
aD (E) = λ2D (z)dxdy.
E

We need to identify the isometries of both the hyperbolic metric and the
hyperbolic distance. A holomorphic function f : D → D is an isometry of the
metric λD (z) |dz| if for all z in D,

(2.3) λD f (z) |f ′ (z)| = λD (z),
and it is an isometry of the distance dD if, for all z and w in D,

(2.4) dD f (z), f (w) = dD (z, w).
In fact, the two classes of isometries coincide, and each isometry is a Möbius
transformation of D onto itself.
12 Beardon and Minda IWQCMA05

Theorem 2.1. For any holomorphic map f : D → D the following are equivalent:
(a) f is a conformal automorphism of D;
(b) f is an isometry of the metric λD ;
(c) f is an isometry of the distance dD .

Proof. First, (a) implies (b). Indeed, if (a) holds, then f is of the form (2.1),
and a calculation shows that
|f ′ (z)| 1
= ,
1 − |f (z)| 2 1 − |z|2
so (b) holds. Next, (b) implies (a). Suppose that (b) holds; that is, f is an
isometry of the hyperbolic metric. Then for any conformal automorphism g of
D, h = g ◦ f is again an isometry of the hyperbolic metric. If we choose g so that
h(0) = g(f (0)) = 0, then
2|h′ (0)| = λD (h(0)|h′ (0)| = λD (0) = 2.
Thus, h is a holomorphic self-map of D that fixes the origin and |h′ (0)| = 1, so
Schwarz’s Lemma implies h ∈ A(D, 0). Then f = g −1 ◦ h is in A(D). We have
now shown that (a) and (b) are equivalent.
Second, we prove (a) and (c) are equivalent. If f ∈ A(D), then f is an isometry
of the metric λD . Hence, for any smooth curve γ in D,
Z Z
 
ℓD f ◦ γ = λD (w) |dw| = λD f (z) |f ′ (z)| |dz| = ℓD (γ).
f ◦γ γ

This implies that for all z, w ∈ D, dD (f (z), f (w)) ≤ dD (z, w). Because f ∈ A(D),
the same argument applies to f −1 , and hence we may conclude that f is a dD –
isometry. Finally, we show that (c) implies (a). Take any f : D → D that is
holomorphic and a dD –isometry. Choose any g of the form (2.1) that maps f (0)
to 0 and put h = g ◦ f . Then his holomorphic, a dD –isometry, and h(0) = 0.
Thus dD 0, h(z) = dD h(0), h(z) = dD (0, z). This implies that |h(z)| = |z| and
hence, that h(z) = eiθ z for some θ ∈ R. Thus h ∈ A(D, 0) and, as f = g −1 ◦ h,
f is also in A(D).

In summary, relative to the hyperbolic metric and the hyperbolic distance,


the group A(D) of conformal automorphism of the unit disk becomes a group of
isometries.
Theorem 2.2. The hyperbolic distance dD (z, w) in D is given by
1 + pD (z, w)
(2.5) dD (z, w) = log = 2 tanh−1 pD (z, w),
1 − pD (z, w)
where the pseudo-hyperbolic distance pD (z, w) is given by

z−w
(2.6)
pD (z, w) = .
1 − z w̄
The hyperbolic metric and geometric function theory 13

Proof. First, we prove that if −1 < x < y < 1 then


y−x
!
1 + 1−xy
(2.7) dD (x, y) = log y−x .
1 − 1−xy
Consider a smooth curve γ joining x to y in D, and write γ(t) = u(t) + iv(t),
where 0 ≤ t ≤ 1. Then
Z 1 Z 1
2 |γ ′ (t)| dt 2 u′ (t) dt
ℓD (γ) = ≥
0 1 − |γ(t)| 0 1 − u(t)
2 2

because |γ(t)|2 ≥ |u(t)|2 = u(t)2 and |γ ′ (t)| ≥ |u′ (t)| ≥ u′ (t). The second integral
can be evaluated directly and gives
  y−x
!
1+y 1−x 1 + 1−xy
ℓD (γ) ≥ log = log y−x .
1−y 1+x 1 − 1−xy
Because equality holds here when γ(t) = x + t(y − x), 0 ≤ t ≤ 1, we see that
(2.7) holds, so (2.5) is valid for −1 < x < y < 1.
Now we have to extend (2.5) to any pair of points z and w in D. Theorem
2.1 shows that each Euclidean rotation about the origin is a hyperbolic isometry
and this implies that, for all z, dD (0, z) = dD (0, |z|). Now take any z and w in
D, and let f (z) = (z − w)/(1 − z w̄). Then f is a conformal automorphism of D,
and so is a hyperbolic isometry. Thus
dD (z, w) = dD (w, z)

= dD f (w), f (z)

= dD 0, f (z)

= dD 0, |f (z)|

= dD 0, pD (z, w) ,
which, from (2.7) with x = 0 and y = pD (z, w), gives (2.5).

Note that (2.5) produces


1 + |z|
dD (0, z) = log , dD (0, z) = 2 tanh−1 |z|.
1 − |z|
Also,
dD (z, w) pD (z, w)
lim = λD (w) = 2 lim .
z→w |z − w| z→w |z − w|

A careful examination of the proof of (2.5) shows that if γ is a smooth curve


that joins x to y, where −1 < x < y < 1, then ℓD (γ) = dD (0, x) if and only if γ
is the simple arc from x to y along the real axis. As hyperbolic isometries map
circles into circles, map the unit circle onto itself, and preserve orthogonality, we
can now make the following definition.
14 Beardon and Minda IWQCMA05

Definition 2.3. Suppose that z and w are in D. Then the (hyperbolic) geodesic
through z and w is C ∩ D, where C is the unique Euclidean circle (or straight
line) that passes through z and w and is orthogonal to the unit circle ∂D. If γ is
any smooth curve joining z to w in D, then the hyperbolic length of γ is dD (z, w)
if and only if γ is the simple arc of C in D that joins z and w.

The unit disk D together with the hyperbolic metric is called the Poincaré
model of the hyperbolic plane. The “lines” in the hyperbolic plane are the hyper-
bolic geodesics and the angle between two intersecting lines is the Euclidean angle
between the Euclidean tangent lines at the point of intersection. The hyperbolic
plane satisfies all of the axioms for Euclidean geometry with the exception of the
Parallel Postulate. It is easy to see that if γ is a hyperbolic geodesic in D and
a ∈ D is a point not on γ, then there are infinitely many geodesics through a
that do not intersect γ and so are parallel to γ.
We shall now show that the hyperbolic distance dD is additive along geodesics.
By contrast, the pseudo-hyperbolic distance pD is never additive along geodesics.
Theorem 2.4. If u, v and w are three distinct points in D that lie, in this order,
along a geodesic, then dD (u, w) = dD (u, v) + dD (v, w). For any three distinct
points u, v and w in D, pD (u, w) < pD (u, v) + pD (v, w).

Proof. Suppose that u, v and w lie in this order, along a geodesic. Then there
is an isometry f that maps this geodesic to the real diameter (−1, 1) of D, with
f (v) = 0. Let x = f (u) and y = f (w), so that −1 < x < 0 < y < 1. It is
sufficient to show that dD (x, 0) + dD (0, y) = dD (x, y); this is a direct consequence
of (2.7).
It is easy to verify that pD a distance function on D, except possibly for the
verification of the triangle inequality. This holds because, for any distinct u, v
and w,
1
pD (u, w) = tanh dD (u, w)
2
1
≤ tanh [dD (u, v) + dD (v, w)]
2
tanh 21 dD (u, v) + tanh 12 dD (v, w)
=
1 + tanh 12 dD (u, v) tanh 21 dD (v, w)
1 1
< tanh dD (u, v) + tanh dD (v, w)
2 2
= pD (u, v) + pD (v, w).
This also shows that there is always a strict inequality in the triangle inequality
for pD for any three distinct points.

The following example illustrates how the hyperbolic distance compares with
the Euclidean distance in D.
The hyperbolic metric and geometric function theory 15

Example 2.5. The Poincaré model of the hyperbolic plane does not accurately
reflect all of the properties of the hyperbolic plane. For example, the hyperbolic
plane is homogeneous; this means that for any pair of points a and b in D there
is an isometry f with f (a) = b. Intuitively this means that the hyperbolic plane
looks the same at each point just as the Euclidean plane does. However, with
our Euclidean eyes, the origin seems to occupy a special place in the hyperbolic
plane. In fact, in the hyperbolic plane the origin is no more special than any
point a 6= 0.
Here is another way in which the Poincaré model deceives our Euclidean eyes.
Let x0 , x1 , x2 , . . . be the sequence 0, 21 , 34 , 87 , . . ., so that xn = (2n − 1)/2n , and
xn+1 is halfway between xn and 1 in the Euclidean sense. A computation using
(2.5) shows that dD (0, xn ) = log(2n+1 − 1). We conclude that dD (xn , xn+1 ) →
log 2 as n → ∞; thus the points xn are, for large n, essentially equally spaced
in the hyperbolic sense along the real diameter of D. Moreover, in any figure
representing the Poincaré model the points xn , for n ≥ 30, are indistinguishable
from the point 1 which does not lie in the hyperbolic plane. In brief, although the
hyperbolic plane contains arbitrarily large hyperbolic disks about the origin, our
Euclidean eyes can only see hyperbolic disks about the origin with a moderate
sized hyperbolic radius.
Let us comment now on the various formulae that are available for dD (z, w).
It is often tempting to use the pseudohyperbolic distance pD rather than the
hyperbolic distance dD (and many authors do) because the expression for pD is
algebraic whereas the expression for dD is not. However, this temptation should
be resisted. The distance pD is not additive along geodesics, and it does not
arise from a Riemannian metric. Usually, the solution is to use the following
functions of dD , for it is these that tend to arise naturally and more frequently
in hyperbolic trigonometry:
2 1 |z − w|2 1
(2.8) sinh dD (z, w) = = |z − w|2 λD (z)λD (w),
2 (1 − |z|) (1 − |w| )
2 2 4
and
1 |1 − z w̄|2 1
cosh2 dD (z, w) = = |1 − z w̄|2 λD (z)λD (w).
2 (1 − |z| )(1 − |w| )
2 2 4
These can be proved directly from (2.5), and together they give the familiar
formula
1 z−w
tanh dD (z, w) = = pD (z, w).
2 1 − z w̄
We investigate the topology defined on the unit disk by the hyperbolic dis-
tance. For this we study hyperbolic disks since they determine the topology. The
hyperbolic circle Cr given by {z ∈ D : dD (0, z) = r} is a Euclidean circle with
Euclidean center 0 and Euclidean radius tanh 12 r. Now let C be any hyperbolic
circle, say of hyperbolic radius r and hyperbolic center w. Then there is a hyper-
bolic isometry f with f (w) = 0, so that f (C) = Cr . As Cr is a Euclidean circle,
so is f −1 (Cr ), which is C. Conversely, suppose that C is a Euclidean circle in D.
16 Beardon and Minda IWQCMA05

Then there is a hyperbolic isometry f such that f (C) is a Euclidean circle with
center 0, so that f (C) = Cr for some r. Thus, as f is a hyperbolic isometry,
f −1 (Cr ) = C, is also a hyperbolic circle. This shows that the set of hyperbolic
circles coincides with the set of Euclidean circles in D. As the same is obviously
true for open disks (providing that the closed disks lie in D), we see that the
topology induced by the hyperbolic distance on D coincides with the Euclidean
topology on the unit disk.
Theorem 2.6. The topology induced by dD on D coincides with the Euclidean
topology. The space D with the distance dD is a complete metric space.

Proof. We have already proved the first statement. Suppose, then, that (zn )
is a Cauchy sequence with respect to the distance dD . Then (zn ) is a bounded
sequence with respect to dD and, as we have seen above, this means that the
(zn ) lie in a compact disk K that is contained in D. As λD ≥ 2 on D, we
see immediately from (2.8) that (zn ) is a Cauchy sequence with respect to the
Euclidean metric, so that zn → z ∗ , say, where z ∗ ∈ K ⊂ D. It is now clear that
dD (zn , z ∗ ) → 0 so that D with the distance dD is complete.

The Euclidean metric on D arises from the fact that D is embedded in the
larger space C and is not complete on D. By contrast, an important property of
the distance dD is that dD (0, |z|) → +∞ as |z| → 1; informally, the boundary ∂D
of D is ‘infinitely far away’ from each point in D. This is a consequence of the
fact that D equipped with the hyperbolic distance dD is a complete metric space
and is another reason why dD should be preferred to the Euclidean metric on D.
Exercises.
1. Verify that (2.1) and (2.2) determine the same subgroup of Möbius trans-
formations.
2. Suppose equality holds in the triangle inequality for the hyperbolic distance;
that is, suppose u, v, w in D and dD (u, w) = dD (u, v) + dD (v, w). Prove
that u, v and w lie on a hyperbolic geodesic in this order.
3. Verify that the hyperbolic disk DD (a, r) is the Euclidean disk with center
c and radius R, where

a 1 − tanh2 (r/2) (1 − |a|2 ) tanh(r/2)
c= and R = .
1 − |a|2 tanh2 (r/2) 1 − |a|2 tanh2 (r/2)
4. (a) Prove that the hyperbolic area of a hyperbolic disk of radius r is
4π sinh2 (r/2).
(b) Show that the hyperbolic length of a hyperbolic circle with radius r is
2π sinh r.

3. The Schwarz-Pick Lemma


We begin with a statement of the classical Schwarz Lemma.
The hyperbolic metric and geometric function theory 17

Theorem 3.1 (Schwarz’s Lemma). Suppose that f : D → D is holomorphic and


that f (0) = 0. Then either
(a) |f (z)| < |z| for every non-zero z in D, and |f ′ (0)| < 1, or
(b) for some real constant θ, f (z) = eiθ z and |f ′ (0)| = 1.

The Schwarz Lemma is proved by applying the Maximum Modulus Theorem


to the holomorphic function f (z)/z on the unit disk D. It says that if a holo-
morphic function f : D → D fixes 0 then either (a) f (z) is closer to 0 than z is,
or (b) f is a rotation of the plane about 0. Although both of these assertions are
true in the context of Euclidean geometry, they are only invariant under confor-
mal maps when they are interpreted in terms of hyperbolic geometry. Moreover,
as Pick observed in 1915, in this case the requirement that f has a fixed point
in D is redundant. We can now state Pick’s invariant formulation of Schwarz’s
Lemma [33].
Theorem 3.2 (The Schwarz-Pick Lemma). Suppose that f : D → D is holomor-
phic. Then either
(a) f is a hyperbolic contraction; that is, for all z and w in D,
 
(3.1) dD f (z), f (w) < dD (z, w), λD f (z) |f ′ (z)| < λD (z),
or
(b) f is a hyperbolic isometry; that is, f ∈ A(D) and for all z and w in D,
 
(3.2) dD f (z), f (w) = dD (z, w), λD f (z) |f ′ (z)| = λD (z)

Proof. By Theorem 2.1, f is an isometry if and only if one, and hence both,
of the conditions in (3.2) hold. Suppose now that f : D → D is holomorphic
but not an isometry. Select any two points z1 and z2 in D. Here is the intuitive
idea behind the proof. Because the hyperbolic plane is homogeneous, we may
assume without loss of generality that both z1 and f (z1 ) are at the origin. In
this special situation (3.1) follows directly from part (b) of Theorem 3.1. Now
we write out a formal argument. Let g and h be conformal  automorphisms (and
hence isometries) of D such that g(z1 ) = 0 and h f (z1 ) = 0. Let F = hf g −1 ;
then F is a holomorphic self-map of D that fixes 0. As g and h are isometries,
F is not an isometry  or else f would be too. Therefore, by Schwarz’s Lemma,

for all z, dD 0, F (z) < dD (0, z) and |F (0)| < 1. Thus, as F g = hf and g, h are
hyperbolic isometries,
 
dD f (z1 ), f (z2 ) = dD hf (z1 ), hf (z2 )

= dD F g(z1 ), F g(z2 )

= dD 0, F g(z2 )

< dD 0, g(z2 )

= dD g(z1 ), g(z2 )
= dD (z1 , z2 ).
18 Beardon and Minda IWQCMA05

This is the first inequality in (3.1). To obtain the second inequality, we apply
the Chain Rule to each side of F g = hf and obtain
|f ′ (z1 )|(1 − |z1 |2 )
|F ′ (0)| = < 1.
1 − |f (z1 )|2
This gives the second inequality in (3.1) at an arbitrary point z1 .

Often the Schwarz-Pick Lemma is stated in the following form: Every holo-
morphic self-map of the unit disk is a contraction relative to the hyperbolic
metric. That is, if f is a holomorphic self-map of D, then
 
(3.3) dD f (z), f (w) ≤ dD (z, w), λD f (z) |f ′ (z)| ≤ λD (z).
If equality holds in either inequality, then f is a conformal automorphism of D.
One should note that the two inequalities in (3.3) are equivalent. If the first
inequality holds, then
 dD (f (z), f (w)) |f (z) − f (w)| dD (z, w)
λD f (z) |f ′ (z)| = lim ≤ lim = λD (z).
w→z |f (z) − f (w)| |z − w| w→z |z − w|

On the other hand, if the second inequality holds, then integration over any path
γ in D gives ℓD (f ◦ γ) ≤ ℓD (γ). This implies the first inequality in (3.3).
Hyperbolic geometry had been used in complex analysis by Poincaré in his
proof of the Uniformization Theorem for Riemann surfaces. The work of Pick is
a milestone in geometric function theory, it shows that the hyperbolic metric, not
the Euclidean metric, is the natural metric for much of the subject. The definition
of the hyperbolic metric might seem arbitrary. In fact, up to multiplication by a
positive scalar it is the only metric on the unit disk that makes every holomorphic
self-map a contraction, or every conformal automorphism an isometry.
Theorem 3.3. For a metric ρ(z)|dz| on the unit disk the following are equivalent:
(a) For any holomorphic self-map of D and all z ∈ D, ρ(f (z))|f ′ (z)| ≤ ρ(z);
(b) For any f ∈ A(D) and all z ∈ D, ρ(f (z))|f ′ (z)| = ρ(z);
(c) ρ(z) = cλD for some c > 0.

Proof. (a)⇒(b) Suppose f ∈ A(D). Then the inequality in (a) holds for f . The
inequality in (a) also holds for f −1 ; this gives ρ(z) ≤ ρ(f (z))|f ′ (z)|. Hence, every
conformal automorphism of D is an isometry relative to ρ(z)|dz|.
(b)⇒(c) Define c > 0 by ρ(0) = cλD (0). Now, consider any a ∈ D. Let f be
a conformal automorphism of D with f (0) = a. Then because f is an isometry
relative to both ρ(z)|dz| and the hyperbolic metric,
ρ(a)|f ′ (0)| = ρ(0)
= cλD (0)
= cλD (a)|f ′ (0)|.
Hence, ρ(a) = cλD (a) for all a ∈ D.
(c)⇒(a) This is an immediate consequence of the Schwarz-Pick Lemma.
The hyperbolic metric and geometric function theory 19

Exercises.
1. Suppose f is a holomorphic self-map of the unit disk. Prove |f ′ (0)| ≤ 1.
Determine a necessary and sufficient condition for equality.
2. If a holomorphic self-map of the unit disk fixes two points, prove it is the
identity.
3. Let a and b be distinct points in D.
(a) Show that there exists a conformal automorphism f of D that inter-
changes a and b; that is, f (a) = b and f (b) = a.
(b) Suppose a holomorphic self-map f of D interchanges a and b; that is,
f (a) = b and f (b) = a. Prove f is a conformal automorphism with order
2, or f ◦ f is the identity.

4. An extension of the Schwarz-Pick Lemma


Recently, the authors [8] established a multi-point version of the Schwarz-Pick
Lemma that unified a number of known variations of the Schwarz and Schwarz-
Pick Lemmas and also has many new consequences. A selection of results from
[8] are presented in this and the next section; for more results of this type, consult
the original paper.
We begin with a brief discussion of Blaschke products. A function F : D → D
is a (finite) Blaschke product if it is holomorphic in D, continuous in D (the closed
unit disk), and |F (z)| = 1 when |z| = 1. If F is a Blaschke product then so are
the compositions g(F (z)) and F (g(z)) for any conformal automorphism g of D.
In addition, it is clear that any finite product of conformal automorphisms of D is
a Blaschke product. We shall now show that the converse is true. Suppose that
F is a Blaschke product. If F has no zeros in D then, by the Minimum Modulus
Theorem, F is a constant, which must be of modulus one. Now suppose that F
does have a zero in D. Then it can only have a finite number of zeros in D, say
a1 , . . . , ak (which need not be distinct), and
Y k  
z − am
F (z)
m=1
1 − ām z
is a Blaschke product with no zeros in D. This shows that F is a Blaschke
product if and only if it is a finite product of automorphisms of D. We say that
F is of degree k if this product has exactly k non-trivial factors.
We now discuss the complex pseudo-hyperbolic distance in D, and the hyper-
bolic equivalent of the usual Euclidean difference quotient of a function.
Definition 4.1. The complex pseudo-hyperbolic distance [z, w] between z and w
in D is given by
z−w
[z, w] = .
1 − wz
We recall that the pseudo-hyperbolic distance is |[z, w]|; see (2.6).
20 Beardon and Minda IWQCMA05

The complex pseudo-hyperbolic distance is an analog for the hyperbolic plane


D of the real directed distance x − y from y to x, for points on the real line R.
Definition 4.2. Suppose that f : D → D is holomorphic, and that z, w ∈ D
with z 6= w. The hyperbolic difference quotient f ∗ (z, w) is given by
[f (z), f (w)]
f ∗ (z, w) = . 
[z, w]

If we combine (2.6) with the Schwarz-Pick Lemma we see that



pD f (z1 ), f (z2 ) ≤ pD (z1 , z2 ),
and that equality holds for one pair z1 and z2 of distinct points if and only if f is
a conformal automorphism of D (in which case, equality holds for all z1 and z2 ).
It follows that if f : D → D is holomorphic, then either f is a hyperbolic isometry
and |f ∗ (z, w)| = 1 for all z and w, or f is not an isometry and |f ∗ (z, w)| < 1 for
all z and w.
We shall now discuss the hyperbolic difference quotient f ∗ (z, w). This is a
function of two variables but, unless we state explicitly to the contrary, we shall
regard it as a holomorphic function of the single variable z. Note that f ∗ (z, w)
is not holomorphic as a function of the second variable w. The basic properties
of f ∗ (z, w) are given in our next result.
Theorem 4.3. Suppose that f : D → D is holomorphic, and that w ∈ D.
(a) The function z 7→ f ∗ (z, w) is holomorphic in D.
(b) If f is not a conformal automorphism of D, then z 7→ f ∗ (z, w) is a holo-
morphic self-map D.
(c) The map z 7→ f ∗ (z, w) is a conformal automorphism of D if and only if f
is a Blaschke product of degree two.

Proof. Part (a) is obvious as w is a removable singularity of the function


! −1
∗ f (z) − f (w) z−w
f (z, w) = .
1 − f (w)f (z) 1 − wz
Now suppose that f is not a conformal automorphism of D. Then, as we have
seen above, |f ∗ (z, w)| < 1 and this proves (b).
To prove (c) we note first that there are conformal automorphisms g and h
(that depend on w) of D such that f ∗ (z, w) = g(f (z))/h(z) or, equivalently,
f (z) = g −1 f ∗ (z, w)h(z) . Clearly, if f ∗ (z, w) is an automorphism then f is
a Blaschke product of degree two. Conversely,
 suppose that f is a Blaschke
product of degree two. Then g f (z) is also a Blaschke product, say B, of
degree two and f ∗ (z, w) = B(z)/h(z). As f ∗ (z, w) is holomorphic in z, we see
that B(z) = h(z)h1 (z) for some automorphism h1 . Thus f ∗ (z, w) = h1 (z) as
required.

We shall now derive a three-point version of the Schwarz-Pick Lemma. Because


it involves three points rather than two points as in the Schwarz-Pick Lemma, the
The hyperbolic metric and geometric function theory 21

following theorem has extra flexibility and it includes all variations and extensions
of the Schwarz-Pick Lemma that are known to the authors. We stress, though,
that this theorem contains much more than simply the union of all such known
results. Although several Euclidean variations of the Schwarz-Pick Lemma are
known, in our view much greater clarity is obtained by a strict adherence to hy-
perbolic geometry. This and other stronger versions of the Schwarz-Pick Lemma
appear in [8].
Theorem 4.4 (Three-point Schwarz-Pick Lemma). Suppose that f is holomor-
phic self-map of D, but not an automorphism of D. Then, for any z, w and v in
D,

(4.1) dD f ∗ (z, v), f ∗ (w, v) ≤ dD (z, w).
Further, equality holds in (4.1) for some choice of z, w and v if and only if f is
a Blaschke product of degree two.

Proof. As f is holomorphic in D, but not an automorphism, Theorem 4.3(b)


shows that the left-hand side of (4.1) is defined. The inequality (4.1) now follows
by applying the Schwarz-Pick Lemma to the holomorphic self-map z 7→ f ∗ (z, v)
of D. The Schwarz-Pick Lemma also implies that equality holds in (4.1) if and
only if f ∗ (z, w) is a conformal automorphism of D and, by Theorem 4.3(c), this
is so if and only if f is a Blaschke product of degree two.

Theorem 4.4 is a genuine improvement of the Schwarz-Pick Lemma. Suppose,


for example that f : D → D is holomorphic, but not an automorphism, and that
f (0) = 0. Then the Schwarz-Pick Lemma tells us only that f (z)/z lies in the
hyperbolic plane D, and that |f ′ (0)| < 1. However, it we put w = 0 in (4.1),
and then let v → 0, we obtain the stronger conclusion that f (z)/z lies in the
hyperbolic disk with center f ′ (0) and hyperbolic radius dD (0, z).
Exercises.
1. If f (z) is a Blaschke product of degree k, prove that f ∗ (z, w) is a Blaschke
product of degree k − 1.
2. Verify the following Chain Rule for the ∗-operator: For all z and w in D,
and all holomorphic maps f and g of D into itself,

(f ◦ g)∗ (z, w) = f ∗ g(z), g(w) g ∗ (z, w).

5. Hyperbolic derivatives
Since the hyperbolic metric is the natural metric to study holomorphic self-
maps of the unit disk, one should also use derivatives that are compatible with
this metric. We begin with the definition of a hyperbolic derivative; just as
the Euclidean difference quotient leads to the usual Euclidean derivative, the
hyperbolic difference quotient results in the hyperbolic derivative.
22 Beardon and Minda IWQCMA05

Definition 5.1. Suppose that f : D → D is holomorphic, but not an isometry


of D. The hyperbolic derivative f h (w) of f at w in D is
[f (z), f (w)] (1 − |w|2 )f ′ (w)
f h (w) = lim = .
z→w [z, w] 1 − |f (w)|2
The hyperbolic distortion of f at w is
dD (f (z), f (w))
|f h (w)| = lim .
z→w dD (z, w)

By Theorem 4.3, |f h (z)| ≤ 1, and equality holds for some z if and only if
equality holds for all z, and then f is a conformal automorphism of D. Theorem
4.4 leads to the following upper bound on the magnitude of the hyperbolic dif-
ference quotient in terms of dD (z, w) and the derivative at any point v between
z and w.
Theorem 5.2. Suppose that f : D → D is holomorphic. Then, for all z and w
in D, and for all v on the closed geodesic arc joining z and w,
 
(5.1) dD 0, f ∗ (z, w) ≤ dD 0, f h (v) + dD (z, w).

Proof. First, it is clear that for any z and w, |f ∗ (z, w)| = |f ∗ (w, z)|. Thus
 
dD 0, f ∗ (z, w) = dD 0, f ∗ (w, z) .
Next, Theorem 4.4 (applied twice) gives
  
dD 0, f ∗ (z, w) ≤ dD 0, f ∗ (v, w) + dD f ∗ (v, w), f ∗ (z, w)

≤ dD 0, f ∗ (v, w) + dD (z, v)

= dD 0, f ∗ (w, v) + dD (z, v)

≤ dD 0, f ∗ (u, v) + dD (w, u) + dD (z, v).
We now let u → v, where v lies on the geodesic between z and w, and as
dD (z, v) + dD (v, w) = dD (z, w), we obtain (5.1).

Our next task is to transform (5.1) into a more transparent inequality about
f . This is the next result which we may interpret as a Hyperbolic Mean Value
Inequality, a result from [7].
Theorem 5.3 (Hyperbolic Mean Value Inequality). Suppose that f : D → D is
holomorphic. Then, for all z and w in D, and for all v on the closed geodesic
arc joining z and w,
  
(5.2) dD f (z), f (w) ≤ log cosh dD (z, w) + |f h (v)| sinh dD (z, w) .

This inequality is sharper than the Schwarz-Pick inequality for if we use


|f h (v)| ≤ 1 and the identity cosh t + sinh t = et , we recapture the Schwarz-
Pick inequality. It is known that equality holds in (5.2) if and only if f is a
Blaschke product of degree two and has a unique critical point c, such that ei-
ther c, z = v, w, or c, w = v, z, lie in this order along a geodesic. We refer
The hyperbolic metric and geometric function theory 23

the reader to [8] for a proof of this, and for the fact that a Blaschke product of
degree two has exactly one critical point in D.
Proof. First, we note that, for all u and v,
1
tanh dD (0, u) = |u|,
2
1 1
tanh dD (u, v) = pD (u, v) = tanh dD (0, [u, v]).
2 2

Next, using the definition of f (z, w), the inequality in Theorem 5.2, and the
addition formula for tanh(s + t), we have
1
tanh dD (f (z), f (w)) = pD (f (z), f (w))
2
= pD (z, w)|f ∗ (z, w)|
1 
= pD (z, w) tanh dD 0, f ∗ (z, w)
2
h1  1 i
h
≤ pD (z, w) tanh dD 0, f (v) + dD (z, w)
 2 2
pD (z, w) + |f h (v)|
= pD (z, w) .
1 + pD (z, w)|f h (v)|
Now the increasing function x 7→ tanh( 12 x) has inverse x 7→ log(1 + x)/(1 − x),
so we conclude that, with p = pD (z, w) and d = |f h (v)|,
   
1 + pd + p(d + p) 1 + p2 2p
dD (f (z), f (w)) ≤ log = log +d ,
1 + pd − p(d + p) 1 − p2 1 − p2
which is (5.2).
Next, we provide a Schwarz-Pick type of inequality for hyperbolic derivatives;
recall that the hyperbolic derivative is not holomorphic. This result is based
on the observation that if f : D → D is holomorphic, but not a conformal
automorphism of D, then f h (z) and f h (w) lie in D so that we can measure the
hyperbolic distance between these two hyperbolic derivatives.
Theorem 5.4. Suppose that f : D → D is holomorphic but not a conformal
automorphism of D. Then, for all z and w in D,
 
(5.3) dD f h (z), f h (w) ≤ 2dD (z, w) + dD f ∗ (z, w), f ∗ (w, z) .
Proof. Theorem 4.4 implies that for all z, w and v,

dD f ∗ (z, w), f ∗ (v, w) ≤ dD (z, v).
We let v → w and obtain

dD f ∗ (z, w), f h (w) ≤ dD (z, w)
and (by interchanging z and w),

dD f ∗ (w, z), f h (z) ≤ dD (z, w).
These last two inequalities and the triangle inequality yields (5.3).
24 Beardon and Minda IWQCMA05

It is easy to see that if f (0) = 0 then f ∗ (z, 0) = f ∗ (0, z) = f (z)/z. Thus we


have the following corollary originally established in [6].
Corollary 5.5. Suppose that f : D → D is holomorphic but not a conformal
automorphism of D, and that f (0) = 0. Then, for all z,

(5.4) dD f h (0), f h (z) ≤ 2dD (0, z),
and the constant 2 is best possible.
Example 5.6. The preceding corollary is sharp for f (z) = z 2 . Note that
f h (z) = 2z/(1 + |z|2 ) and dD (f h (z), f h (w)) = 2dD (z, w) whenever z, w lie on
the same hyperbolic geodesic through the origin. Thus, z 7→ f h (z) doubles all
hyperbolic distances along geodesics through the origin; this doubling is not
valid in general because in hyperbolic geometry there are no similarities except
isometries. Moreover, it is possible to verify that there is no finite K such that
dD (f h (z), f h (w)) ≤ KdD (z, w) for all z, w ∈ D, so z 7→ f h (z) does not even sat-
isfy a hyperbolic Lipschitz condition, so (5.4) is no longer valid when the origin
is replaced by an arbitrary point of the unit disk.

In spite of Example 5.6 a full-fledged result of Schwarz-Pick type is valid for


the hyperbolic distortion.
Corollary 5.7 (Schwarz-Pick Lemma for Hyperbolic Distortion). Suppose that
f : D → D is holomorphic but not a conformal automorphism of D. Then for all
z, w ∈ D, dD |f h (z)|, |f h (w)| ≤ 2dD (z, w).

Proof. Note that, from the proof of Theorem 5.4,


 
dD |f ∗ (z, w)|, |f h (w)| ≤ dD f ∗ (z, w), f h (w) ≤ dD (z, w),

and, similarly, dD |f ∗ (w, z)|, |f h (z)| ≤ dD (z, w). As |f ∗ (w, z)| = |f ∗ (z, w)|, the
desired inequality follows.
Exercises.
1. Verify the claims in Example 5.6.
2. Suppose that f : D → D is holomorphic but not a conformal automorphism
of D. Prove that for all conformal automorphisms S and T of D, and all z
and w in D,
|(S ◦ f ◦ T )∗ (z, w)| = |f ∗ (T (z), T (w))|.
In particular, deduce that the hyperbolic derivative is invariant in the sense
that
|(S ◦ f ◦ T )h (z)| = |f h (T (z))|.

6. The hyperbolic metric on simply connected regions


There are several equivalent definitions of what it means for a region in the
complex plane to be simply connected. A region Ω in C is simply connected if
and only if any one of the following (equivalent) conditions hold:
The hyperbolic metric and geometric function theory 25

(a) the set C∞ \Ω is connected;


(b) if f is holomorphic and never zero in Ω, then there is a single-valued holo-
morphic choice of log f in Ω;
(c) each closed curve in Ω can be continuously deformed within Ω to a point
of Ω.
A region in C∞ is simply connected if (a) or (c) holds. The regions D, C and
C∞ are all simply connected; an annulus is not.
Two subregions regions of C are conformally equivalent if there is a holomor-
phic bijection of one onto the other. This is an equivalence relation on the class
of subregions of C, and the fundamental result about simply connected regions
is the Riemann Mapping Theorem.
Theorem 6.1 (The Riemann Mapping Theorem). A subregion of C is confor-
mally equivalent to D if and only if Ω is a simply connected proper subregion of
C. Moreover, given a ∈ Ω there is a unique conformal mapping f : Ω → D such
that f (a) = 0 and f ′ (a) > 0.

The Riemann Mapping Theorem enables us to transfer the hyperbolic metric


from D to any simply connected proper subregion Ω of C.
Definition 6.2. Suppose that f is a conformal map of a simply connected plane
region Ω onto D. Then the hyperbolic metric λΩ (z)|dz| of Ω is defined by

(6.1) λΩ (z) = λD f (z) |f ′ (z)|.
The hyperbolic distance dΩ is the distance function on Ω derived from the hy-
perbolic metric.

We need to show λΩ is independent of the choice of the conformal map f that


is used in (6.1), for this will imply that λΩ is determined by Ω alone. Suppose,
then, that f is a conformal map of Ω onto D. Then the set of all conformal
maps of Ω onto D is given by h ◦ f , where h ranges over A(D). Any conformal
automorphism h of D is a hyperbolic isometry, so that for all w in D,

λD (w) = λD h(w) |h′ (w)|.
If we now let g = h ◦ f , w = f (z) and use the Chain Rule we find that
 
λD g(z) |g ′ (z)| = λD h(f (z) |h′ (f (z))||f ′ (z)|
= λD (f (z))|f ′ (z)|
so that λΩ as defined in (6.1) is independent of the choice of the conformal map
f.
Thus, Definition 6.2 converts every conformal map of a simply connected
proper subregion of C onto the unit disk into an isometry of the hyperbolic
metric. The hyperbolic distance dΩ on a simply connected proper subregion Ω of
C can be defined in two equivalent ways. First, one can pull-back the hyperbolic
distance on D to Ω by setting dΩ (z, w) = dD (f (z), f (w)) for any conformal map
f : Ω → D and verifying that this is independent of the choice of the conformal
26 Beardon and Minda IWQCMA05

mapping onto the unit disk. Alternatively, the hyperbolic length of a path γ in
Ω is Z
ℓΩ (γ) = λΩ (z)|dz|,
γ
and one can define
dΩ (z, w) = inf ℓΩ (γ),
where the infimum is taken over all piecewise smooth curves γ in Ω that join
z and w. These two definitions of the hyperbolic distance are equivalent. The
hyperbolic distance dΩ on Ω is complete. Moreover, a path γ in Ω connecting z
and w is a hyperbolic geodesic in Ω if and only if f ◦ γ is a hyperbolic geodesic
in D. Also, for any a ∈ Ω and r > 0, f (DΩ (a, r)) = DD (f (a), r).
In fact, the essence of Definition 6.2 is that the entire body of geometric
facts about the Poincaré model D of the hyperbolic plane transfers, without any
essential change, to an arbitrary simply connected proper subregion of C with
its own hyperbolic metric. If f : Ω → D is any conformal mapping, then f is an
isometry relative to the hyperbolic metrics and hyperbolic distances on Ω and
D. The next result is an immediate consequence of Definition 6.2 and we omit
its proof; it asserts that all conformal maps of simply connected proper regions
are isometries relative to the hyperbolic metrics and hyperbolic distances of the
regions.
Theorem 6.3 (Conformal Invariance). Suppose that Ω1 and Ω2 are simply con-
nected proper subregions of C, and that f is a conformal map of Ω1 onto Ω2 .
Then f is a hyperbolic isometry, so that for any z in Ω1 ,

(6.2) λΩ2 f (z) |f ′ (z)| = λΩ1 (z),
and for all z, w ∈ Ω1
dΩ2 (f (z), f (w)) = dΩ1 (z, w).

Note that if γ is a smooth curve in Ω1 , then (6.2) implies


ℓΩ2 (f ◦ γ) = ℓΩ1 (γ).
Theorem 6.3 implies that each element of A(Ω), the group of conformal auto-
morphisms of Ω, is a hyperbolic isometry.
Theorem 6.4 (Schwarz-Pick Lemma for Simply Connected Regions). Suppose
that Ω1 and Ω2 are simply connected proper subregions of C, and that f is a
holomorphic map of Ω1 into Ω2 . Then either
(a) f is a hyperbolic contraction; that is, for all z and w in Ω1 ,
 
dΩ2 f (z), f (w) < dΩ1 (z, w), λΩ2 f (z) |f ′ (z)| < λΩ1 (z),
or
(b) f is a hyperbolic isometry; that is, f is a conformal map of Ω1 onto Ω2 and
for all z and w in Ω1 ,
 
dΩ2 f (z), f (w) = dΩ1 (z, w), λΩ2 f (z) |f ′ (z)| = λΩ1 (z).
The hyperbolic metric and geometric function theory 27

Proof. Because of Theorem 6.3 we only need verify (a) when the holomorphic
map f : Ω1 → Ω2 is not a holomorphic bijection. Choose any point z0 in
Ω1 , and let w0 = f (z0 ). Next, construct a holomorphic bijection h of D, and a
holomorphic bijection g of Ω1 onto Ω2 ; these can be constructed so that h(0) = z0
and g(z0 ) = w0 = f (z0 ). Now let k = (gh)−1 f h. Then k is a holomorphic map
of D into itself and k(0) = 0. Moreover, k is not a conformal automorphism of
D or else f would be a holomorphic bijection. Thus |k ′ (0)| < 1 and, using the
Chain Rule, this gives |f ′ (z0 )| < |g ′ (z0 )|. With this,

λΩ2 f (z0 ) |f ′ (z0 )| < λΩ1 (z0 )
follows as (6.2) holds (with f replaced by g).
This establishes the second strict inequality in (a); the first strict inequality
for hyperbolic distances follows by integrating the strict inequality for hyperbolic
metrics.

This version of the Schwarz-Pick Lemma can be stated in the following equiv-
alent form. If f : Ω1 → Ω2 is holomorphic, then for all z and w in Ω1 ,
(6.3) dΩ2 (f (z), f (w)) ≤ dΩ1 (z, w),
and

(6.4) λΩ2 f (z) |f ′ (z)| ≤ λΩ1 (z).
Further, if either equality holds in (6.3) for a pair of distinct points or at one
point z in (6.4) , then f is a conformal bijection of Ω1 onto Ω2 .
Corollary 6.5 (Schwarz’s Lemma for Simply Connected Regions). Suppose Ω
is a simply connected proper subregion of Ω and a ∈ Ω. If f is a holomorphic
self-map of Ω that fixes a, then |f ′ (a)| ≤ 1 and equality holds if and only if
f ∈ A(Ω, a), the group of conformal automorphisms of Ω that fix a. Moreover,
f ′ (a) = 1 if and only if f is the identity.

Theorem 6.4 is the fundamental reason for the existence of many distortion
theorems in complex analysis. Consider the class of holomorphic maps of Ω1 into
Ω2 . Then any such map f will have to satisfy the universal constraints (6.3) and
(6.4) where the metrics λΩ1 and λΩ2 are uniquely determined (albeit implicitly)
by the regions Ω1 and Ω2 . Thus (6.3) and (6.4) are, in some sense, the generic
distortion theorems for holomorphic maps.
This is the appropriate place to point out that neither the complex plane C
nor the extended complex plane C∞ has a metric analogous to the hyperbolic
metric in the sense that the metric is invariant under the group of conformal
automorphisms. Recall that A(C) is the set of all maps z 7→ az + b, a, b ∈ C and
a 6= 0, and A(C∞ ) is the group M of Möbius transformations. The group A(C)
acts doubly transitively on C; that is, given two pairs z1 , z2 and w1 , w2 of distinct
points in C there is a conformal automorphism f of C with f (zj ) = wj , j = 1, 2.
Similarly, M acts triply transitively on C∞ . If there were a conformal metric
on either C or C∞ invariant under the full conformal automorphism group, then
28 Beardon and Minda IWQCMA05

the distance function induced from this metric would also be invariant under the
action of the full group of conformal automorphisms. The following result shows
that only trivial distance functions are invariant under A(C) or A(C∞ ).
Theorem 6.6. If d is a distance function on C or C∞ that is invariant under
the full group of conformal automorphisms, then there exists t > 0 such that
d(z, w) = 0 if z = w and d(z, w) = t otherwise.

Proof. Let d be a distance function on C that is invariant under A(C). Set t =


d(0, 1). Consider any distinct z, w ∈ C. Because A(C) acts doubly transitively
on C, there exists f ∈ A(C) with f (0) = z and f (1) = w. The invariance of d
under A(C) implies d(z, w) = d(f (0), f (1)) = d(0, 1) = t. The same argument
applies to C∞ .

The Euclidean metric |dz| on C is invariant under the proper subgroup of


A(C) given by z 7→ az + b, where |a| = 1 and b ∈ C. The spherical metric
2|dz|/(1 + |z|2 ) on C∞ is invariant under the group of rotations of C∞ , that is,
Möbius maps of the form
az − c̄
z 7→ , a, c ∈ C, |a|2 + |c|2 = 1,
cz + ā
or of the equivalent form
z−a
z 7→ eiθ , θ ∈ R, a ∈ C∞ .
1 + āz
The group of rotations of C∞ is a proper subgroup of M.
Exercises.
1. Suppose Ω is a simply connected proper subregion of C and a ∈ Ω. Let
F denote the family of all holomorphic functions f defined on D such that
f (D) ⊆ Ω and f (0) = a. Set M = sup{|f ′ (0)| : f ∈ F}. Prove M < +∞
and that |f ′ (0)| = M if and only if f is a conformal map of D onto Ω with
f (0) = a. Show M = 2/λΩ (a).
2. Suppose Ω is a simply connected proper subregion of C and a ∈ Ω. Let
G denote the family of all holomorphic functions f defined on Ω such that
f (Ω) ⊆ D. Set N = sup{|f ′ (a)| : f ∈ G}. Prove N < +∞ and that
|f ′ (a)| = N if and only if f is a conformal map of Ω onto D with f (a) = 0.
Show N = λΩ (a)/2.
3. Suppose Ω is a simply connected proper subregion of C and a ∈ Ω. Let
H(Ω, a) denote the family of all holomorphic self-maps of Ω that fix a.
Prove that {f ′ (a) : f ∈ H(Ω, a)} equals the closed unit disk.

7. Examples of the hyperbolic metric


We give examples of simply connected regions and their hyperbolic metrics.
These metrics are computed by using (6.2) in the following way: one finds an
explicit conformal map f from the region Ω1 whose metric is sought onto a region
The hyperbolic metric and geometric function theory 29

Ω2 whose metric is known. Then (6.2) enables one to find an explicit expression
for λΩ1 (z) for z in Ω1 . We omit almost all of the computations.
The simplest instance of the Riemann Mapping Theorem is the fact that
any disk or half-plane is Möbius equivalent to the unit disk. Because hyperbolic
circles (disks) in D are Euclidean circle (disks) in D, we deduce that an analogous
result holds for any disk or half-plane. Also, in any disk or half-plane hyperbolic
geodesics are arcs of circles orthogonal to the boundary; in the case of a half-
planes we allow half-lines orthogonal to the edge of the half-plane.
Example 7.1 (disk). As f (z) = (z − z0 )/R is a conformal map of the disk
D = {z : |z − z0 | < R} onto D, we find
2R |dz|
λD (z)|dz| = . 
R2 − |z − z0 |2

In particular,
2
λD (z0 ) = .
R
Example 7.2 (half-plane). Let H be the upper half-plane {x+iy : y > 0}. Then
g(H) = D, where g(z) = (z − i)/(z + i), so H = {x + iy : y > 0} has hyperbolic
metric
|dz| |dz|
λH (z)|dz| = = .
y Im z
Similarly, the hyperbolic metric of the right half-plane K = {x + iy : x > 0} is
|dz|/x. More generally, if H is any open half-plane, then
|dz|
λH (z)|dz| = ,
d(z, ∂H)
where d(z, ∂H) denotes the Euclidean distance from z to ∂H.
Theorem 7.3. If f : D → K is holomorphic and f (0) = 1, then
1 − |z| 1 + |z|
(7.1) ≤ Re f (z) ≤
1 + |z| 1 − |z|
and
2|z|
(7.2) |Im f (z)| ≤
1 − |z|2

Proof. This is an immediate consequence of the Schwarz-Pick Lemma after con-


verting the conclusion into weaker Euclidean terms. Fix z ∈ D and set
1 + |z|
r = dD (0, z) = 2 tanh−1 |z| = log .
1 − |z|
The Schwarz-Pick Lemma implies that f (z) lies in the closed hyperbolic disk
D̄K (1, r). The closed hyperbolic disk D̄K (1, r) has Euclidean center cosh r, Eu-
clidean radius sinh r and the bounding circle meets the real axis at e−r and er ;
30 Beardon and Minda IWQCMA05

see the exercises. Therefore, f (z) lies in the closed Euclidean square {z = x+iy :
e−r ≤ x ≤ er , |y| ≤ sinh r}. Since
1 − |z| 1 + |z|
e−r = and er = ,
1 + |z| 1 − |z|
(7.1) is established. Finally,
2|z|
sinh r =
1 − |z|2
demonstrates (7.2).
Theorem 7.4. Suppose that H is any disk or half-plane. Then for all z and w
in H,
1
sinh2 dH (z, w) = 41 |z − w|2 λH (z)λH (w).
2
Proof. It is easy to verify that for any Möbius map g we have
2
(7.3) g(z) − g(w) = (z − w)2 g ′ (z) g ′ (w).
Now take any Möbius map g that maps H onto D, and recall that g is an isometry
from H to D if both are given their hyperbolic metrics. Then, using (2.8) and
(7.3)
  1   ′
1
4
|z − w|2
λ H z λ H w = 4
|z − w|2
λ D g(z) λ D g(w) |g (z)| |g ′ (w)|
= 41 |g(z) − g(w)|2 λD (g(z))λD (g(w))
1
= sinh2 dD (g(z), g(w))
2
1 
= sinh2 dH z, w
2

There is another, less well known, version of the Schwarz-Pick Theorem avail-
able which is an immediate consequence of Theorem 7.4, and which we state in
a form that is valid for all disks and half-planes.
Theorem 7.5 (Modified Schwarz-Pick Lemma for Disks and Half-Planes). Sup-
pose that Hj is any disk or half-plane, j = 1, 2, and that f : H1 → H2 is
holomorphic. Then, for all z and w in H1 ,
|f (z) − f (w)|2 λH1 (z)λH1 (w)
≤  .
|z − w| 2
λH2 f (z) λH2 f (w)
Proof. By Theorem 7.4 and the Schwarz-Pick Lemma
1
1
4
|f (z) − f (w)|2 λH2 (f (z))λH2 (f (w)) = sinh2 dH2 (f (z), f (w))
2
1
≤ sinh2 dH1 (z, w)
2  
= 4 |z − w)|2 λH1 z λH1 w .
1
The hyperbolic metric and geometric function theory 31

Observe that if w → z in Theorem 7.5, then we obtain (6.4) in the special case
of disks and half-planes. We give an application of Theorem 7.5 to holomorphic
functions.
Example 7.6. Suppose that f is holomorphic in the open unit disk and that f
has positive real part. Then f maps D into K, and we have
|f (z) − f (w)|2 4 Re [f (z)] Re [f (w)]
≤ .
|z − w| 2 (1 − |z|2 )(1 − |w|2 )
This implies, for example, that if we also have f (0) = 1 then |f ′ (0)| ≤ 2.

Example 7.7 (slit plane). Since f (z) = z maps P = C\{x ∈ R : x ≤ 0} onto
K = {x + iy : x > 0}, the hyperbolic metric on P is
|dz|
λP (z) |dz| = √ √ .
2| z| Re [ z]
This gives
1 1
λP (z) = ≥ ,
2r cos(θ/2) 2|z|
where z = reiθ .
Example 7.8 (sector). Let S(α) = {z : 0 < arg(z) < απ}, where 0 < α ≤ 2.
Here, f (z) = z 1/α = exp α−1 log z is a conformal map of S(α) onto H, so S(α)
has hyperbolic metric
|z|1/α
λS(α) (z) |dz| = |dz|.
α|z| Im[z 1/α ]
Note that this formula for the hyperbolic metric agrees (as it must) with the
formula for λH in Example 7.2 (which is the case α = 1). The special case α = 2
is the preceding example.
Example 7.9 (doubly infinite strip). S = {x + iy : |y| < π/2} has hyperbolic
metric
|dz|
λS (z) |dz| = .
cos y
In this case we use the fact that ez maps S conformally onto K = {x + iy : x >
0}. Notice that λS (z) ≥ 1 with equality if and only if z lies on the real axis.
In particular, the hyperbolic distance between points on R is the same as the
Euclidean distance between the points.
Theorem 7.10. Let S = {z : |Im(z)| < π/2}. Then for any a ∈ R and any
holomorphic self-map f of S, |f ′ (a)| ≤ 1. Moreover, f ′ (a) = 1 if and only if
f (z) = z + c for some c ∈ R and f ′ (a) = −1 if and only if f (z) = −z + c for
some c ∈ R. In particular, for any interval [a, b] in R, the Euclidean length of
the image f ([a, b]) is at most b − a.
32 Beardon and Minda IWQCMA05

Proof. From Example 7.9 for z ∈ S


1
λS (z) = ≥1
cos y
and equality holds if and only if Im(z) = 0. This observation together with the
Schwarz-Pick Lemma gives
|f ′ (a)| ≤ λS (f (a))|f ′ (a)| ≤ λS (a) = 1
and equality implies f (a) ∈ R. In this case, f −f (a)+a is a holomorphic self-map
of S that fixes a and has derivative 1 at a, so it is the identity by the general form
of Schwarz’s Lemma. Thus, from Corollary 6.5 f ′ (a) = 1 implies f (z) = z + c
for some c ∈ R; the converse is trivial. If f ′ (a) = −1, then −f is a holomorphic
self-map of S with derivative 1 at a, and so f (z) = −z − c for some c ∈ R.

For a simply connected proper subregion Ω of C and a ∈ Ω, each hyperbolic


disk DΩ (a, r) = {z ∈ Ω : dΩ (a, z) < r} is simply connected and the closed
disk D̄Ω (a, r) is compact. When Ω is a disk or half-plane, hyperbolic disks are
Euclidean disks since any conformal map of the unit disk onto a disk or half-plane
is a Möbius transformation. Of course, this is no longer true when Ω is simply
connected and not a disk or half-plane. For particular types of simply connected
regions, more can be said about hyperbolic disks than just the fact that they are
simply connected.
Theorem 7.11. Suppose Ω is a convex hyperbolic region. Then for any a ∈ Ω
and all r > 0 the hyperbolic disc DΩ (a, r) is Euclidean convex.

Proof. Fix a ∈ Ω. Let h : D → Ω be a conformal mapping with h(0) = a. Since


h(DD (0, r)) = DΩ (a, r), it suffices to show that h maps each disc DD (0, r) =
D(0, tanh(r/2)) onto a convex set. Set R = tanh(r/2). Given b, c ∈ D(0, R)
we must show (1 − t)h(b) + th(c) lies in h(D(0, R)) for t ∈ I. Choose S so that
|b|, |c| < S < R and fix t ∈ I. The function
   cz 
bz
g(z) = (1 − t)h + th
S S
is holomorphic in D, g(0) = a and maps into Ω because Ω is convex. There-
fore, f = h−1 ◦ g is a holomorphic self-map of D that fixes the origin and so
f (D(0, R)) ⊆ D(0, R). Then (1 − t)h(b) + th(c) = g(S) = h(f (S)) lies in
h(D(0, S)) because f (S) ∈ D(0, S). Therefore, h(D(0, S)) = DΩ (a, r) is Eu-
clidean convex.

This result is effectively due to Study who proved that if f is convex univalent
in D, then for any Euclidean disk D contained in D, f (D) is Euclidean convex,
see [13]. The converse of Theorem 7.11 is elementary: If Ω is a simply connected
proper subregion of C and there exists a ∈ Ω such that every hyperbolic disk
DΩ (a, r) is Euclidean convex, then Ω is Euclidean convex since Ω = ∪{DΩ (a, r) :
r > 0}, an increasing union of Euclidean
√ convex sets. The radius of convexity
for a univalent function on D is 2 − 3; see [13]. This implies that if Ω is simply
The hyperbolic metric and geometric function theory 33

connected, then for each a ∈ Ω and 0 < r < (1/2) log 3 the hyperbolic disk
DΩ (a, r) is Euclidean convex.
In a general simply connected region hyperbolic geodesics are no longer arcs
of circles or segments of lines. It is possible to give a simple geometric property
of hyperbolic geodesics in Euclidean convex regions that characterize convex
regions, see [19] and [20].
Exercises.
1. Let K = {z = x + iy : x > 0}. For a > 0 and r > 0 verify that the
closed hyperbolic disk D̄K (1, r) is the Euclidean disk with Euclidean center
c = cosh r and Euclidean radius R = sinh r. This Euclidean disk meets the
real axis at e−r and er .
2. Suppose f : D → K is holomorphic. Prove that
(1 − |z|2 )|f ′ (z)| ≤ 2 Re f (z)
for all z ∈ D. When does equality hold?
3. Suppose f : K → D is holomorphic. Prove that
2|f ′ (z)|Re z ≤ 1 − |f (z)|2
for all z ∈ K. When does equality hold?
4. Suppose Ω is a simply connected proper subregion of C that is (Euclidean)
starlike with respect to a ∈ Ω. This means that for each z ∈ Ω the
Euclidean segment [a, z] is contained in Ω. For any r > 0 prove that the
hyperbolic disk DΩ (a, r) is starlike with respect to a.

8. The Comparison Principle


There is a powerful, and very general, Comparison Principle for hyperbolic
metrics, which we state here only for simply connected plane regions. This
Principle allows us to estimate the hyperbolic metric of a region in terms of other
hyperbolic metrics which are known, or which can be more easily estimated. In
general it is not possible to explicitly calculate the density of the hyperbolic
metric, so estimates are useful.
Theorem 8.1 (Comparison Principle). Suppose that Ω1 and Ω2 are simply con-
nected proper subregions of C. If Ω1 ⊆ Ω2 then λΩ2 ≤ λΩ1 on Ω1 . Further, if
λΩ1 (z) = λΩ2 (z) at any point z of Ω2 , then Ω1 = Ω2 and λΩ1 = λΩ2 .

Proof. Let f (z) = z be the inclusion map of Ω1 into Ω2 . Then the Schwarz-Pick
Lemma gives λΩ2 (z) ≤ λΩ1 (z). If equality holds at a point, then f is a conformal
bijection of Ω1 onto Ω2 , that is, Ω1 = Ω2 .

In other words, the Comparison Principle asserts that the hyperbolic metric
on a simply connected region decreases as the region increases. The hyperbolic
metric on the disk Dr = {z : |z| < r} is 2r|dz|/(r2 − |z|2 ) which decreases to zero
as r increases to +∞.
34 Beardon and Minda IWQCMA05

The Comparison Principle is used in the following way. Suppose that we want
to estimate the hyperbolic metric λΩ of a region Ω. We attempt to find regions
Ωj with known hyperbolic metrics (or metrics that can be easily estimated) such
that Ω1 ⊆ Ω ⊆ Ω2 ; then λΩ2 ≤ λΩ ≤ λΩ1 . The next result is probably the
simplest application of the Comparison Principle, and it gives an upper bound
of the hyperbolic metric λΩ of a region Ω in terms of the Euclidean distance
d(z, ∂Ω) = inf{|z − w| : w ∈ ∂Ω}
of z to the boundary of Ω. The geometric significance of this quantity is that
d(z, ∂Ω) is the radius of the largest open disk with center z that lies in Ω. Note,
however, that d(z, ∂Ω) (which is sometimes denoted by δΩ (z) in the literature)
is not conformally invariant. The metric
|dz| |dz|
=
d(z, ∂Ω) δΩ (z)
is called the quasihyperbolic metric on Ω. Example 7.2 shows that the quasihy-
perbolic metric for a half-plane is the hyperbolic metric.
Theorem 8.2. Suppose that Ω is a simply connected proper subregion of C.
Then for all z ∈ Ω
2
(8.1) λΩ (z) ≤ ,
d(z, ∂Ω)
and equality holds if and only if Ω is a disk with center z.

Proof. Take any z0 in Ω, and let R = d(z0 , ∂Ω) and D = {z : |z − z0 | < R}. As
D ⊆ Ω the Comparison Principle and Example 7.1 yield
2 2
λΩ (z0 ) ≤ λD (z0 ) = = ,
R d(z0 , ∂Ω)
which is (8.1). If λΩ (z0 ) = 2/d(z0 , ∂Ω) then λΩ (z0 ) = λD (z0 ) so, by the Compar-
ison Principle, Ω = D. The converse is trivial.

Theorem 8.2 gives an upper bound on the hyperbolic metric of Ω in terms of


the Euclidean quantity d(z, ∂Ω). It is usually more difficult to obtain a lower
bound on the hyperbolic metric. For convex regions it is easy to use geometric
methods to obtain a good lower bound on the hyperbolic metric.
Theorem 8.3. Suppose that Ω is convex proper subregion of C. Then for all
z∈Ω
1
(8.2) ≤ λΩ (z)
d(z, ∂Ω)
and equality holds if and only if Ω is a half-plane.

Proof. We suppose that Ω is convex. Take any z in Ω and let ζ be one of the
points on ∂Ω that is nearest to z. Let H be the supporting half-plane of Ω at ζ;
The hyperbolic metric and geometric function theory 35

thus Ω ⊆ H, and the Euclidean line that bounds H is orthogonal to the segment
from z to ζ. Thus, from the Comparison Principle, for any z ∈ Ω
1 1
λΩ (z) ≥ λH (z) = =
|z − ζ| d(z, ∂Ω)
which is (8.2). The equality statement follows from the Comparison Principle
and Example 7.2.

Theorems 8.2 and 8.3 show that the hyperbolic and quasihyperbolic metrics
are bi-Lipschitz equivalent on convex regions:
1 2
≤ λΩ (z) ≤ .
d(z, ∂Ω) d(z, ∂Ω)
Lower bounds for the hyperbolic metric in terms of the quasihyperbolic metric
are equivalent to covering theorems for univalent functions.
Theorem 8.4. Suppose that f is holomorphic and univalent in D, and that f (D)
is a convex region. Then f (D) contains the Euclidean disk with center f (0) and
radius |f ′ (0)|/2.

Proof. From Theorem 8.3 we have


2 = λD (0)

= λf (D) f (0) |f ′ (0)|
|f ′ (0)|
≥ .
d f (0), ∂f (D)
We deduce that 
d f (0), ∂f (D) ≥ |f ′ (0)|/2,
so that f (D) contains the Euclidean disk with center f (0) and radius |f ′ (0)|/2.

There is an analogous covering theorem for general univalent functions on the


unit disk, see [13].
Theorem 8.5 (Koebe 1/4–Theorem). Suppose that f is holomorphic and uni-
valent in D. Then the region f (D) contains the open Euclidean disk with center
f (0) and radius |f ′ (0)|/4.

The Koebe 1/4–Theorem gives a lower bound on the hyperbolic metric in


terms of the quasihyperbolic metric on a simply connected proper subregion of
C.
Theorem 8.6. Suppose that Ω is a simply connected proper subregion of C.
Then for all z ∈ Ω
1
(8.3) ≤ λΩ (z)
2d(z, ∂Ω)
and equality holds if and only if Ω is a slit-plane.
36 Beardon and Minda IWQCMA05

Proof. Fix z ∈ Ω and let f : D → Ω be a conformal map with f (0) = z. Koebe’s


1/4–Theorem implies d(z, ∂Ω) ≥ |f ′ (0)|/4. Now
2 = λΩ (f (0))|f ′ (0)|
≤ 4λΩ (z)d(z, ∂Ω)
which establishes (8.3). Sharpness follows from the sharp form of the Koebe
1/4–Theorem and Example 7.7.

Theorems 8.2 and 8.6 show that the hyperbolic and quasihyperbolic metrics
are bi-Lipschitz equivalent on simply connected regions:
1 2
(8.4) ≤ λΩ (z) ≤ .
2d(z, ∂Ω) d(z, ∂Ω)
Exercises.
1. (a) Suppose Ω is a simply connected proper subregion of C. Prove that
limz→ζ λΩ (z) = +∞ for each boundary point ζ of Ω that lies in C.
(b) Given an example of a simply connected proper subregion Ω of C that
has ∞ as a boundary point and λΩ (z) does not tend to infinity as z → ∞.
2. Suppose Ω is starlike with respect to the origin; that is, for each z ∈ Ω the
Euclidean segment [0, z] is contained in Ω. Use the Comparison Theorem
to prove that (8.3) holds; do not use Theorem 8.6.

9. Curvature and the Ahlfors Lemma


A conformal semimetric on a region Ω in C is ρ(z)|dz|, where ρ : Ω → [0, +∞)
is a continuous function and {z : ρ(z) = 0} is a discrete subset of Ω. A conformal
semimetric ρ(z)|dz| is a conformal metric if ρ(z) > 0 for all z ∈ Ω. The curvature
of a conformal semimetric ρ(z)|dz| can be defined at any point where ρ is positive
and of class C2 .
Definition 9.1. Suppose ρ(z)|dz| is a conformal metric on a region Ω. If a ∈ Ω,
ρ(a) > 0 and ρ(z) is of class C2 at a, then the Gaussian curvature of ρ(z)|dz| at
a is
△ log ρ(a)
Kρ (a) = − ,
ρ2 (a)
where △ is the usual (Euclidean) Laplacian,
∂2 ∂2
△= + .
∂x2 ∂y 2

Typically, we shall speak of the curvature of a conformal metric rather than


Gaussian curvature. In computing the Laplacian it is often convenient to use
∂2 ∂2
△=4 =4 ,
∂z∂ z̄ ∂ z̄∂z
The hyperbolic metric and geometric function theory 37

where the complex partial derivatives are defined by


 
∂ 1 ∂ ∂
∂= = −i ,
∂z 2 ∂x ∂y
 
¯ ∂ 1 ∂ ∂
∂= = +i .
∂ z̄ 2 ∂x ∂y
Alternatively the Laplacian expressed can be expressed in terms of polar coordi-
nates, namely
∂2 1 ∂ 1 ∂2
△= 2 + + 2 2.
∂r r ∂r r ∂θ
One reason semimetrics play an important role in complex analysis is that
they transform simply under holomorphic functions.
Definition 9.2. Suppose ρ(w)|dw| is a semimetric on a region Ω and f : ∆ → Ω
is a holomorphic function. The pull-back of ρ(w)|dw| by f is
(9.1) f ∗ (ρ(w)|dw|) = ρ(f (z))|f ′ (z)||dz|.

Since (ρ ◦ f )|f ′ | is a continuous nonnegative function defined on ∆, the pull-


back f ∗ (ρ(w)|dw|) of ρ(w)|dw| is a semimetric on ∆ provided f is nonconstant.
Sometimes we write simply f ∗ (ρ) to denote the pull-back. However, the nota-
tion (9.1) is precise and indicates that the formal substitution w = f (z) converts
ρ(w)|dw| into f ∗ (ρ(w)|dw|). The pull-back operation has several useful proper-
ties:
(f ◦ g)∗ (ρ(w)|dw|) = g ∗ (f ∗ (ρ(w)|dw|))
and
(f −1 )∗ = (f ∗ )−1 ,
when f is a conformal mapping. If f : Ω1 → Ω2 is a conformal mapping of
simply connected proper subregions of C, then the conclusion of Theorem 6.3 in
the pull-back notation is: f ∗ (λΩ2 ) = λΩ1 .
In the context of complex analysis, a fundamental property of the curvature
is its conformal invariance. More generally, curvature is invariant under the
pull-back operation.
Theorem 9.3. Suppose Ω and ∆ are regions in C, ρ(w)|dw| is a metric on Ω and
f : ∆ → Ω is a holomorphic function. Suppose a ∈ ∆, f ′ (a) 6= 0, ρ(f (a)) > 0
and ρ is of class C2 at f (a). Then Kf ∗ (ρ) (a) = Kρ (f (a)).

Proof. Recall that f ∗ (ρ(w)|dw|) = ρ(f (z))|f ′ (z)||dz|. Now,


log (ρ(f (z))|f ′ (z)|) = log ρ(f (z)) + log |f ′ (z)|
1 1
= log ρ(f (z)) + log f ′ (z) + log f ′ (z),
2 2
so that
∂ ∂ log ρ 1 f ′′ (z)
log (ρ(f (z))|f ′ (z)|) = (f (z))f ′ (z) + .
∂z ∂w 2 f ′ (z)
38 Beardon and Minda IWQCMA05

Then
∂2 ∂ 2 log ρ
log (ρ(f (z))|f ′ (z)|) = (f (z))f ′ (z)f ′ (z)
∂ z̄∂z ∂ w̄∂w
∂ 2 log ρ
= (f (z))|f ′ (z)|2
∂ w̄∂w
gives
△z [log (ρ(f (z))|f ′ (z)|)] = (△w log ρ) (f (z))|f ′ (z)|2 .
This is the transformation law for the Laplacian under a holomorphic function.
Consequently,
△z log(ρ(f (a))|f ′ (a)|)
Kf ∗ (ρ) (a) = −
ρ2 (f (a))|f ′ (a)|2
(△w log ρ) (f (a))|f ′ (a)|2
=−
ρ2 (f (a))|f ′ (a)|2
= Kρ (f (a)).

Theorem 9.4. The hyperbolic metric on a simply connected proper subregion of


C has constant curvature −1.

Proof. First, we establish the result for the unit disk. From
2 2
λD (z) = =
1 − |z|2 1 − z z̄
we obtain
∂2 2 ∂2
log =− log(1 − z z̄)
∂ z̄∂z 1 − z z̄ ∂ z̄∂z
∂ z̄
=
∂ z̄ 1 − z z̄
1
= .
(1 − z z̄)2
Consequently, KλD (z) = −1.
The general case of the hyperbolic metric on a simply connected proper sub-
region Ω of C follows from Theorem 9.3 since f ∗ (λD (w)|dw|) = λΩ (z)|dz| for any
conformal map f : Ω → D.

Ahlfors recognized that the Schwarz-Pick Lemma was a consequence of an


extremely important maximality property of the hyperbolic metric in D.
Theorem 9.5 (Maximality of the hyperbolic metric). Suppose ρ(z)|dz| is a C2
semimetric on a simply connected proper subregion Ω of C such that Kρ (z) ≤ −1
whenever ρ(z) > 0. Then ρ ≤ λΩ on Ω.
The hyperbolic metric and geometric function theory 39

Proof. First, we assume Ω = D. Given z0 in D choose any r satisfying |z0 | <


r < 1. The hyperbolic metric on the disk Dr = {z : |z| < r} is
2r
λr (z) = 2 .
r − |z|2
Consider the function
ρ(z)
v(z) =
λr (z)
which is defined on the disk Dr . Then v(z) ≥ 0 when |z| < r, and v(z) → 0 as
|z| → r, so that v attains its maximum at some point a in Dr . It suffices to show
that v(a) ≤ 1 for then v(z) ≤ 1 on Dr and we have
2r
ρ(z0 ) ≤ 2 .
r − |z0 |2
By letting r → 1 we find that ρ(z0 ) ≤ λD (z0 ).
We now show that v(a) ≤ 1. If ρ(a) = 0, then v(a) = 0 < 1. Otherwise,
ρ(a) > 0 and Kρ (a) ≤ −1. As a is a local maximum of v, it is also a local
maximum of log v so that
∂ 2 log v ∂ 2 log v
(a) ≤ 0, (a) ≤ 0.
∂x2 ∂y 2
We deduce that

0 ≥ △ log v (a)
 
= △ log ρ (a) − △ log λr (a)
= −Kρ (a)ρ(a)2 + Kλr (a)λr (a)2
(9.2) ≥ ρ(a)2 − λr (a)2 .
This implies that v(a) ≤ 1, and completes the proof in the special case Ω = D.
We now turn to the general case. Let h : D → Ω be a conformal mapping.
Then h∗ (ρ(w)|dw|) := τ (z)|dz| is a C2 semimetric on D such that Kτ (z) ≤ −1
whenever τ (z) > 0. Hence,
ρ(h(z))|h′ (z)| ≤ λD (z) = λΩ (h(z))|h′ (z)|,
and so ρ ≤ λΩ on Ω.

In fact, Ahlfors actually established a more general result (see [1] and [2]). The
stronger conclusion that either ρ < λΩ or else ρ = λΩ is valid but less elementary.
This sharp result was established by Heins [15]. Simpler proofs of the stronger
conclusion are due to Chen [12], Minda [28] and Royden [32].
The Schwarz-Pick Lemma is a special case of Theorem 9.5. If f : Ω1 → Ω2
is a nonconstant holomorphic function, then f ∗ (λΩ2 (w)|dw|) is a semimetric on
Ω1 with curvature −1 at each point where f ′ is nonvanishing, so is dominated
by the hyperbolic metric λΩ1 (z)|dz|, or equivalently, (6.4) holds. The equality
statement associated with (6.4) follows from the sharp version of Theorem 9.5.
40 Beardon and Minda IWQCMA05

Theorem 9.6. There does not exist a C2 semimetric ρ(z)|dz| on C such that
Kρ (z) ≤ −1 whenever ρ(z) > 0.
Proof. Suppose there existed a semimetric ρ(z)|dz| on C such that Kρ (z) ≤ −1
whenever ρ(z) > 0. Theorem 9.5 applied to the restriction of this metric to the
disk {z : |z| < r} gives
2r
(9.3) ρ(z) ≤ λr (z) = 2
r − |z|2
for |z| < r. If we fix z and let r → +∞, (9.3) gives ρ(z) = 0 for all z ∈ C. This
contradicts the fact that a semimetric vanishes only on a discrete set.
Corollary 9.7 (Liouville’s Theorem). A bounded entire function is constant.
Proof. Suppose f is a bounded entire function. There is no harm in assuming
that |f (z)| < 1 for all z ∈ C. If f were nonconstant, then f ∗ (λD (z)|dz|) would
be a semimetric on C with curvature at most −1, a contradiction.
Theorem 9.3 provides a method to produce metrics with constant curvature
−1. Loosely speaking, bounded holomorphic functions correspond to metrics
with curvature −1. If f : Ω → D is holomorphic and locally univalent (f ′ does
not vanish), then f ∗ (λD (z)|dz|) has curvature −1 on Ω. In fact, on a simply
connected proper subregion of C every metric with curvature −1 has this form;
see [36]. This reference also contains a stronger result that represents certain
semimetrics with curvature −1 at points where the semimetric is nonvanishing
by holomorphic (not necessarily locally univalent) maps of Ω into D.
Theorem 9.8 (Representation of Negatively Curved Metrics). Let ρ(z)|dz| be a
C3 conformal metric on a simply connected proper subregion Ω of C with constant
curvature −1. Then ρ(z)|dz| = f ∗ (λD (w)|dw|) for some locally univalent holo-
morphic function f : Ω → D. The function f is unique up to post-composition
with an isometry of the hyperbolic metric. Given a ∈ Ω the function f represent-
ing the metric is unique if f is normalized by f (a) = 0 and f ′ (a) > 0.
Moreover, ρ(z)|dz| = f ∗ (λD (w)|dw|) is complete if and only if f is a conformal
bijection; that is, the hyperbolic metric is the only conformal metric on Ω that
has curvature −1 and is complete.
Exercises.
1. Determine the curvature of the Euclidean metric |dz| and of the spherical
metric σ(z)|dz| = 2|dz|/(1 + |z|2 ).
2. Show that (1 + |z|2 )|dz| has negative curvature on C.
3. Determine the curvature of ex |dz| on C.
4. Determine the curvature of |dz|/|z| on C \ {0}.
5. Prove there does not exist a semimetric on C \ {0} with curvature at most
−1.
6. Prove the following extension of Liouville’s Theorem: If f is an entire
function and f (C) ⊆ Ω, where Ω is a simply connected proper subregion
of C, then f is constant.
The hyperbolic metric and geometric function theory 41

10. The hyperbolic metric on a hyperbolic region


In order to transfer the hyperbolic metric from the unit disk to nonsimply
connected regions, a substitute for the Riemann Mapping Theorem is needed. For
this reason we must understand holomorphic coverings. For the general theory
of topological covering spaces the reader should consult [22]. A holomorphic
function f : ∆ → Ω is called a covering if each point b ∈ Ω has an open
neighborhood V such that f −1 (V ) = ∪{Uα : α ∈ A} is a disjoint union of open
sets Uα such that f |Uα , the restriction of f to Uα , is a conformal map of Uα onto
V . Trivially, a conformal mapping f : ∆ → Ω is a holomorphic covering. If Ω is
simply connected, then the only holomorphic coverings f : ∆ → Ω are conformal
maps of a simply connected region ∆ onto Ω.
Example 10.1. The complex exponential function exp : C → C \ {0} is a
holomorphic covering. Consider any w ∈ C \ {0} and let θ = arg w be any
argument for w. Let V = C \ {−reiθ : r ≥ 0} be the complex
S plane slit from the
origin along the ray opposite from w. Then exp−1 (V ) = {Sn : n ∈ Z}, where
Sn = {z : θ − nπ < Im z < θ + nπ}. Note that exp maps each horizontal strip
Sn of width 2π conformally onto V .

A region Ω is called hyperbolic provided C∞ \ Ω contains at least three points.


The unit disk covers every hyperbolic plane region; that is, there is a holomorphic
covering h : D → Ω for any hyperbolic region Ω. As a consequence we demon-
strate that every hyperbolic region has a hyperbolic metric that is real-analytic
with constant curvature −1.
Theorem 10.2 (Planar Uniformization Theorem). Suppose Ω is a region in C.
There exists a holomorphic covering f : D → Ω if and only if Ω is a hyperbolic
region. Moreover, if a ∈ Ω, then there is a unique holomorphic universal covering
h : D → Ω with h(0) = a and h′ (0) > 0.

For a proof of the Planar Uniformization see [14] or [34]. If Ω is a simply


connected hyperbolic region, then any holomorphic universal covering h : D →
Ω is a conformal mapping. Therefore, the Riemann Mapping Theorem is a
consequence of the Planar Uniformization Theorem. When Ω is a nonsimply
connected hyperbolic region, then a holomorphic covering h : D → Ω is never
injective. In fact, for each a ∈ Ω, h−1 (a) is an infinite discrete subset of D. If
h : D → Ω is one holomorphic universal covering, then {h◦g : g ∈ A(D)} is the set
of all holomorphic universal coverings of D onto Ω. The Planar Uniformization
Theorem enables us to project the hyperbolic metric from the unit disk to any
hyperbolic region.
Theorem 10.3. Given a hyperbolic region Ω there is a unique metric λΩ (w)|dw|
on Ω such that h∗ (λΩ (w)|dw|) = λD (z)|dz| for any holomorphic universal cover-
ing h : D → Ω. The metric λΩ (w)|dw| is real-analytic with curvature −1.
Proof. We construct a metric with curvature −1 on any hyperbolic region. First
we define the metric locally. For a hyperbolic region Ω, let h : D → Ω be
42 Beardon and Minda IWQCMA05

a holomorphic covering. A metric is defined on Ω as follows. For any simply


connected subregion U of Ω let H = h−1 denote a branch of the inverse that
is holomorphic on U . Set λΩ (z) = λD (H(z))|H ′ (z)|. This defines a metric with
curvature −1 on U . In fact, this defines a metric on Ω. Suppose U1 and U2 are
simply connected subregions of Ω and U1 ∩ U2 is nonempty. Let Hj be a single-
valued holomorphic branch of h−1 defined on Uj . Then there is a g ∈ A(D) such
that H2 = g ◦ H1 locally on U1 ∩ U2 . Hence,
H2∗ (λD (z)|dz|) = (g ◦ H1 )∗ (λD (z)|dz|)
= H1∗ (g ∗ (λD (z)|dz|))
= H1∗ (λD (z)|dz|)
since each conformal automorphism of D is an isometry of the hyperbolic metric
λD (z)|dz|. Therefore, λΩ (z) is defined independently of the branch of h−1 that
is used and h∗ (λΩ ) = λD .
Moreover, this metric is independent of the covering. Suppose k : D → Ω is
another covering. Then k = h ◦ g for some g ∈ A(D), and so
k ∗ (λΩ ) = (h ◦ g)∗ (λΩ )
= g ∗ (h∗ (λΩ ))
= g ∗ (λD )
= λD .
That λΩ is real-analytic is clear from its construction. That the curvature is −1
follows from h∗ (λΩ ) = λD and Theorems 9.3 and 9.4.

The unique metric λΩ (w)|dw| on a hyperbolic region Ω given by Theorem 10.3


is called the hyperbolic metric on Ω. The hyperbolic distance on a hyperbolic
region is complete. The hyperbolic distance dΩ is defined by
dΩ (z, w) = inf ℓΩ (γ),
where the infimum is taken over all piecewise smooth paths γ in Ω that joining
z and w. Unlike the case of simply connected regions, a holomorphic covering
f : D → Ω onto a multiply connected hyperbolic region is not an isometry, but
only a local isometry. That is, each point a ∈ Ω has a neighborhood U such that
f |U is an isometry. In general, one can only assert that dΩ (f (z), f (w)) ≤ dD (z, w)
for z, w ∈ D. When Ω is multiply connected, then f is not injective, so there
exist distinct z, w ∈ D with f (z) = f (w). In this situation, dΩ (f (z), f (w)) = 0 <
dD (z, w).
In general, the hyperbolic metric is not just invariant under conformal map-
pings, but is invariant under holomorphic coverings.
Theorem 10.4 (Covering Invariance). If f : ∆ → Ω is a holomorphic covering
of hyperbolic regions, then f ∗ (λΩ (w)|dw|) = λ∆ (z)|dz|. In other words, every
holomorphic covering of hyperbolic regions is a local isometry.
The hyperbolic metric and geometric function theory 43

Proof. Let h : D → ∆ be a holomorphic covering. Then k = f ◦ h : D → Ω is


also a holomorphic covering, so
λD = k ∗ (λΩ )
= h∗ (f ∗ (λΩ )).
Thus, f ∗ (λΩ ) is a conformal metric on ∆ whose pull-back to the unit disk by a
covering projection is λD , so f ∗ (λΩ ) is the hyperbolic metric on ∆ by Theorem
10.3.

Theorem 10.4 implies that every h ∈ A(Ω) is an isometry of the hyperbolic


metric, and more generally, each holomorphic self-covering h of Ω is a local
isometry of the hyperbolic metric. A hyperbolic region can have self-coverings
that are not conformal automorphisms, see Section 12.
The maximality property of the hyperbolic metric given in Theorem 9.5 re-
mains valid for hyperbolic regions. As noted after the proof of Theorem 9.5 this
means that a version of the Schwarz-Pick Lemma holds for holomorphic maps
between hyperbolic regions. In order to establish a sharp result, we provide an
independent proof.
Theorem 10.5 (Schwarz-Pick Lemma - general version). Suppose ∆ and Ω are
hyperbolic regions and f : ∆ → Ω is holomorphic. Then for all z ∈ ∆,
(10.1) λΩ (f (z))|f ′ (z)| ≤ λ∆ (z).
If f : ∆ → Ω is a covering projection, then λΩ (f (z))|f ′ (z)| = λ∆ (z) for all
z ∈ ∆. If there exists a point in ∆ such that equality holds in (10.1), then f is
a covering.

Proof. Let k : D → ∆ and h : D → Ω be holomorphic coverings. The function


f ◦ k : D → Ω lifts relative to h to a holomorphic function F : D → D. Then
f ◦ k = h ◦ F and the Schwarz-Pick Lemma for the unit disk give
k ∗ (f ∗ (λΩ )) = (f ◦ k)∗ (λΩ )
= (h ◦ F )(λΩ )
= F ∗ (h∗ (λΩ ))
= F ∗ (λD )
≤ λD
= k ∗ (λ∆ ).
Because k is a surjective local homeomorphism, the inequality k ∗ (f ∗ (λΩ )) ≤
k ∗ (λ∆ ) gives the inequality (10.1). Because h and k are coverings, f is a covering
if and only if F is a covering. This observation establishes the sharpness.

We need to establish a result about holomorphic self-coverings of a hyperbolic


region that have a fixed point in order to obtain a good analog of Schwarz’s
Lemma for hyperbolic regions.
44 Beardon and Minda IWQCMA05

Theorem 10.6. A self-covering of a hyperbolic region that fixes a point is a


conformal automorphism. In particular, if a hyperbolic region Ω is not simply
connected and a ∈ Ω, then A(Ω, a) is isomorphic to the group of nth roots of
unity for some positive integer n.

Proof. Suppose Ω is a hyperbolic region, a ∈ Ω and f is a self-covering of Ω


that fixes a. We prove f is a conformal automorphism. The result is trivial
if Ω is simply connected since every covering of a simply connected region is
a homeomorphism. Let h : D → Ω be a holomorphic universal covering with
h(0) = a and h′ (0) > 0. Because a covering is surjective, it suffices to prove
f is injective. Let f˜ be the lift of f ◦ h relative to h that satisfies f˜(0) = 0.
Since h and f ◦ h are coverings, so is f˜. Because D is simply connected, f˜ is a
conformal automorphism of D. Then f˜(z) = eiθ z for some θ ∈ R. Because Ω
is not simply connected, the fiber h−1 (a) contains infinitely many points besides
0. As this fiber is a discrete subset of D, the nonzero elements of h−1 (a) have a
minimum positive modulus r; say h−1 (a) ∩ {z : |z| = r} = {aj : j = 1, . . . , m}.
From f˜(h−1 (a)) = h−1 (a), we conclude that f˜ induces a permutation of the set
{aj : j = 1, . . . , m}. Therefore, there exists n ≤ m! such that f˜n is the identity.
Then f n ◦h = h◦ f˜n = h and so f n is the identity. If n = 1, then f is the identity.
If n ≥ 2, then f n = I, the identity, implies f is a conformal automorphism of Ω
with inverse f n−1 .
This argument shows that if Ω is not simply connected, then there is a non-
negative integer m such that for all f ∈ A(Ω, a), f m is the identity. Therefore,
f ′ (a)m = 1, or f ′ (a) is an mth root of unity. Thus, f 7→ f ′ (a) defines a homo-
morphism of A(Ω, a) into the unit circle T and the image is a subgroup of the
mth roots of unity. Hence, A(Ω, a) is a finite group isomorphic to the group of
nth roots of unity for some positive integer n.
Example 10.7. Let AR = {z : 1/R < |z| < R}, where R > 1. The group
A(AR , 1) has order two; the only conformal automorphisms of AR that fix 1 are
the identity map and f (z) = 1/z.
Corollary 10.8 (Schwarz’s Lemma - General Version). Suppose Ω is a hyper-
bolic region, a ∈ Ω and f is a holomorphic self-map of Ω that fixes a. Then
|f ′ (a)| ≤ 1 and equality holds if and only if f ∈ A(Ω, a), the group of confor-
mal automorphisms of Ω that fix a. Moreover, f ′ (a) = 1 if and only if f is the
identity.

Proof. The Schwarz-Pick Lemma implies |f ′ (a)| ≤ 1 with equality if and only if
f is a self-covering of Ω that fixes a. Each f ∈ A(Ω, a) is a covering, so |f ′ (a)| = 1.
If f is a self-covering of Ω that fixes a, then f ∈ A(Ω, a) by Theorem 10.6. We
use the proof of Theorem 10.6 to verify that f ′ (a) = 1 implies f is the identity.
Let f˜ be the lift of f ◦ h as in the proof of Theorem 10.6. Then h ◦ f = f ◦ h
gives 1 = f ′ (a) = f˜′ (0), so f˜ is the identity. This implies f is the identity.

Picard established a vast generalization of Liouville’s Theorem.


The hyperbolic metric and geometric function theory 45

Theorem 10.9 (Picard’s Small Theorem). If an entire function omits two finite
complex values, then f is constant.

Proof. Suppose f is an entire function and f (C) ⊆ C\{a, b} := Ca,b , where a and
b are distinct complex numbers. We derive a contradiction if f were nonconstant.
The region Ca,b is hyperbolic; let λa,b (z)|dz| denote the hyperbolic metric on Ca,b .
If f were nonconstant, then f ∗ (λa,b (z)|dz|) would be a semi-metric on C with
curvature at most −1; this contradicts Theorem 9.6.
Exercises.
1. Verify that f (z) = exp(iz) is a covering of the upper half-plane H onto the
punctured disk D \ {0}.
2. Verify that for each nonzero integer n the function pn (z) = z n defines a
holomorphic covering of the punctured plane C \ {0} onto itself.
3. Verify that for each positive integer n the function pn (z) = z n defines a
holomorphic covering of the punctured disk D \ {0} onto itself.
4. Suppose Ω is a hyperbolic region and a ∈ Ω. Let F denote the family of
all holomorphic functions f : D → Ω such that f (0) = a and set M =
sup{|f ′ (0)| : f ∈ F}. Prove M is finite and for f ∈ F, |f ′ (0)| = M
if and only if f is a holomorphic covering of D onto Ω. Conclude that
M = 2/λΩ (a).
5. Suppose Ω is a hyperbolic region in C and a, b ∈ Ω are distinct points. If
f is a holomorphic self-map of Ω that fixes a and b, prove f is a conformal
automorphism of Ω with finite order. Give an example to show that f need
not be the identity when Ω is not simply connected.

11. Hyperbolic distortion


In Section 5 the hyperbolic distortion of a holomorphic self-map of the unit
disk was introduced. We now define an analogous concept for holomorphic maps
of hyperbolic regions.
Definition 11.1. Suppose Ω and ∆ are hyperbolic regions and f : ∆ → Ω is
holomorphic. The (local) hyperbolic distortion factor for f at z is

∆,Ω λΩ (f (z))|f ′ (z)| dΩ (f (z), f (w))


f (z) := = lim .
λ∆ (z) w→z d∆ (z, w)
If Ω = ∆, write f ∆ in place of f ∆,Ω .

The hyperbolic distortion factor defines a mapping of ∆ into the closed unit
disk by the Schwarz-Pick Lemma. If f is not a covering, then the hyperbolic
distortion factor gives a map of ∆ into the unit disk. There is a Schwarz-Pick
type of result for the hyperbolic distortion factor which extends Corollary 5.7 to
holomorphic maps between hyperbolic regions.
46 Beardon and Minda IWQCMA05

Theorem 11.2 (Schwarz-Pick Lemma for Hyperbolic Distortion). Suppose ∆


and Ω are hyperbolic regions and f : ∆ → Ω is holomorphic. If f is not a
covering, then
(11.1) dD (f ∆,Ω (z), f ∆,Ω (w)) ≤ 2d∆ (z, w)
for all z, w ∈ ∆.

Proof. Fix w ∈ Ω. Let h : D → ∆ and k : D → Ω be holomorphic coverings


with h(0) = w and k(0) = f (w). Then there is a lift of f to a self-map f˜ of D
such that k ◦ f˜ = f ◦ h. f˜ is not a conformal automorphism of D because f is
not a covering of ∆ onto Ω. We begin by showing that f˜D (z̃) = f ∆,Ω (h(z̃)) for
all z̃ in D. From k ◦ f˜ = f ◦ h and λD (z̃) = λ∆ (h(z̃))|h′ (z̃)| we obtain
λΩ (f (h(z̃)))|f ′ (h(z̃))|
f ∆,Ω (h(z̃)) =
λ∆ (h(z̃))
λΩ (k(f˜(z̃)))|k ′ (f˜(z̃))||f˜′ (z̃)|
=
λ∆ (h(z̃))|h′ (z̃)|
λD (f˜(z̃))|f˜′ (z̃)|
=
λD (z)
˜D
= f (z̃).
Now we establish (11.1). For z ∈ Ω there exists z̃ ∈ h−1 (z) with dD (0, z̃) =
dΩ (w, z). Then
dD (f ∆,Ω (z), f ∆,Ω (w)) = dD (f ∆,Ω (h(z̃)), f ∆,Ω (h(0)))
= dD (f˜D (z̃), f˜D (0))
≤ 2dD (z̃, 0)
= 2d∆ (z, w).

Corollary 11.3. Suppose ∆ and Ω are hyperbolic regions. Then for any holo-
morphic function f : ∆ → Ω,
f ∆,Ω (w) + tanh d∆ (z, w)
(11.2) f ∆,Ω (z) ≤ .
1 + f ∆,Ω (w) tanh d∆ (z, w)
for all z, w ∈ ∆.

Proof. Inequality (11.2) is trivial when f is a covering since both sides are
identically one, Thus, it suffices to establish the inequality when f is not a
covering of ∆ onto Ω. Then
dD (0, f ∆,Ω (z)) ≤ dD (0, f ∆,Ω (w)) + dD (f ∆,Ω (z), f ∆,Ω (w))
≤ dD (0, f ∆,Ω (w)) + 2d∆ (z, w)
The hyperbolic metric and geometric function theory 47

gives
 
∆,Ω 1 ∆,Ω
f (z) = tanh dD (0, f (z))
2
 
1 ∆,Ω
≤ tanh dD (0, f (w)) + d∆ (z, w)
2
f ∆,Ω (w) + tanh d∆ (z, w)
= .
1 + f ∆,Ω (w) tanh d∆ (z, w)

Exercises.
1. For a holomorphic function f : D → K, explicitly calculate f D,K (z).
2. Suppose Ω is a simply connected proper subregion of C and a ∈ Ω. Let
H(Ω, a) denote the set of holomorphic self-maps of Ω that fix a. Prove that
{f Ω (a) : f ∈ H(Ω, a)} is the closed unit interval [0, 1].

12. The hyperbolic metric on a doubly connected region


There is a simple conformal classification of doubly connected regions in C∞ .
If Ω is a doubly connected region in C∞ , then Ω is conformally equivalent to
exactly one of:
(a) C∗ = C \ {0},
(b) D∗ = D \ {0}, or
(c) an annulus A(r, R) = {z : r < |z| < R}, where 0 < r < R.
In the first case Ω itself is the extended plane C∞ punctured at two points and
so is not hyperbolic. In this section we calculate the hyperbolic metric for the
punctured unit disk D∗ and for the annulus AR = {z : 1/R < |z| < R}, where
R > 1.

12.1. Hyperbolic metric on the punctured unit disk. To determine the


hyperbolic metric on D∗ we make use of a holomorphic covering from H onto D∗
and Theorem 10.4. The function h(z) = exp(iz) is a holomorphic covering from
H onto D∗ . Therefore, the density of the hyperbolic metric on D∗ is
1
λD∗ (z) = .
|z| log(1/|z|)
For simply connected hyperbolic regions the only hyperbolic isometries of the
hyperbolic metric are conformal automorphisms. For multiply connected re-
gions there can be self-coverings that leave the hyperbolic metric invariant. For
instance, the maps z 7→ z n , n = 2, 3, . . ., are self-coverings of D∗ that leave
λD∗ (z)|dz| invariant. Up to composition with a rotation about the origin these
are the only self-coverings of D∗ that are not automorphisms.
Because each hyperbolic geodesic in D∗ is the image of a hyperbolic geodesic
in H under h, every radial segment [reiθ , Reiθ ], where 0 < r < R < 1, is part of
48 Beardon and Minda IWQCMA05

a hyperbolic geodesic. Since the density λD∗ is independent of θ, the hyperbolic


length of a geodesic segment σr,R = [reiθ , Reiθ ] is independent of θ. Direct
calculation gives
Z Z R
dt log R
ℓD∗ (σr,R ) = λD∗ (z)|dz| =
= log .
t log t log r
[r,R] r

As the formula shows, this length tends to infinity if either r → 0 or R → 1


which also follows from the completeness of the hyperbolic metric. The Euclidean
circle Cr = {z : |z| = r}, where 0 < r < 1, is not a hyperbolic geodesic; it has
hyperbolic length
Z
|dz| 2π
ℓD∗ (Cr ) = = .
|z|=r |z| log(1/|z|) log(1/r)
The hyperbolic length of Cr approaches 0 when r → 0 and ∞ when r → 1. The
hyperbolic area of the annulus A(r, R) = {z : r < |z| < R} ⊂ D∗ is
Z Z
1
aD∗ (A(r, R)) = 2 dx dy
A(r,R) |z| log |z|
2
Z R
dt
= 2π
t log2 t
r 
1 1
= 2π − .
log(1/R) log(1/r)
The hyperbolic area of A(r, R) tends to infinity when R → 1 and has the finite
limit 2π/ log(1/R) when r → 0.
There is a Euclidean surface in R3 that is isometric to {z : 0 < |z| < 1/e}
with the restriction of the hyperbolic metric on D∗ and makes it easy to see
these curious results about length and area in a neighborhood of the puncture.
If a tractrix is rotated about the y-axis and the resulting surface is given the
geometry induced from the Euclidean metric on R3 , then this surface has constant
curvature −1 and is isometric to {z : 0 < |z| < 1/e} with the restriction of the
hyperbolic metric on D∗ . This picture provides a simple isometric embedding of
a portion of D∗ into R3 . Radial segments correspond to rotated copies of the
tractrix and these have infinite Euclidean length. At the same time the surface
has finite Euclidean area.
Recall that for a simply connected region Ω, the hyperbolic density λΩ and
the quasihyperbolic density 1/δΩ are bi-Lipschitz equivalent; see (8.4). These
two metrics are not bi-Lipschitz equivalent on D∗ because the behavior of λD∗
near the unit circle differs from its behavior near the origin. For 1/2 < |z| < 1,
δD∗ (z) = 1 − |z| and so
lim δD∗ (z)λD∗ (z) = 1.
|z|→1

For 0 < |z| < 1/2, δ(z) = |z| and so


lim δD∗ (z)λD∗ (z) = 0.
|z|→1
The hyperbolic metric and geometric function theory 49

12.2. Hyperbolic metric on an annulus. Now we determine the hyperbolic


metric on an annulus by using a holomorphic covering from a strip onto an
annulus. In each conformal equivalence class of annuli we choose the unique
representative that is symmetric about the unit circle.
For 0 < r < R let A(r, R) = {z : r < |z| < R}. The number mod(A(r, R)) =
log(R/r) is called the modulus of A(r, R). Two annuli A(rj , Rj ), j = 1, 2, are
conformally (actually Möbius) equivalent if p
and only if R1 /r1 = R2 /r2 ; that is, if
and only if they have equal moduli. If S = (R/r), then AS = {z : 1/S < |z| <
S} is the unique annulus conformally equivalent to A(r, R) that is symmetric
about the unit circle. Here symmetry means AS is invariant under z 7→ 1/z̄,
reflection about the unit circle. Note that mod(AS ) = 2 log S.
The function k(z) = exp(z) is a holomorphic universal covering from the
vertical strip Sb = {z : |Im z| < b}, where b = log R, onto the annulus AR = {z :
1/R < |z| < R}. Therefore, the density of the hyperbolic metric of the annulus
AR is
π 1
λR (z) =  .
2 log R |z| cos π log |z|
2 log R

Example 12.1. We investigate the hyperbolic lengths of the Euclidean circles


Cr = {z : |z| = r} in AR . The hyperbolic length of Cr is
Z
π |dz| π2
ℓR (Cr ) =  =  .
|z|=r 2 log R |z| cos
π log |z| π log r
2 log R
(log R) cos 2 log R

The symmetry of AR about the unit circle is reflected by the fact that two
circles symmetric about the unit circle have the same hyperbolic length. Also,
the hyperbolic length of Cr increases from π 2 / log R = 2π 2 /mod AR to ∞ as r
increases from 1 to R. Hence, the hyperbolic lengths of the Euclidean circles Cr
in AR have a positive minimum hyperbolic length. The Euclidean circle C1 is a
hyperbolic geodesic; Cr is not a hyperbolic geodesic when r 6= 1.

If γn (t) = exp(2πint), then I(γn , 0) = n, where I(δ, 0) denotes the index or


winding number of a closed path δ about the origin, and
2π 2 |n|
ℓR (γn ) = .
mod A(R)
We now show that γn has minimal hyperbolic length among all closed paths in
AR that wind n times about the origin.
Theorem 12.2. Suppose γ is a piecewise smooth closed path in AR and I(γ, 0) =
n 6= 0. Then
2π 2 |n|
(12.1) ≤ ℓR (γ),
mod A(R)
where ℓR (γ) denotes the hyperbolic length of γ. Equality holds in (12.1) if and
only if γ is a monotonic parametrization of the unit circle traversed n times.
50 Beardon and Minda IWQCMA05

Proof. Suppose γ : [0, 1] → AR is a closed path with I(γ, 0) = n 6= 0. Then


Z
ℓR (γ) = λR (z) |dz|
γ
Z 1
π |γ ′ (t)| dt
=  
2 log R 0 |γ(t)| cos π log |γ(t)|
2 log R
Z 1
π |γ ′ (t)|

2 log R 0 |γ(t)|
since
 
π log |γ(t)|
(12.2) 0 < cos ≤1
2 log R
and equality holds if and only if |γ(t)| = 1 for t ∈ [0, 1]. Next,
Z 1 ′ Z
|γ (t)| 1 γ ′ (t)
(12.3) ≥ dt
0 |γ(t)| 0 γ(t)
Z
dz
=
z γ
= |2πiI(γ, 0)|
= 2π|n|.
Hence,
π 2 |n| 2π 2 |n|
ℓR (γ) ≥ = .
log R mod AR
It is straightforward to verify that if γn (t) = exp(2πint), t ∈ [0, 1], then equality
holds in (12.1). Conversely, suppose γ is a path for which equality holds. Then
equality holds in (12.2), so |γ(t)| = 1 for t ∈ [0, 1]. Let δ : [0, 1] → C be a
lift of γ relative to the covering exp : C → C∗ . From I(γ, 0) = n, we obtain
δ(1) − δ(0) = 2πni. The condition |γ(t)| = 1 implies δ(t) ∈ iR for t ∈ [0, 1].
The function h(t) = (δ(t) − δ(0))/2πi is real-valued, h(0) = 0 and h(1) = n.
Also, γ(t) = exp(2πih(t) + δ(0)). Equality must hold in (12.3) and this means
γ ′ (t)/γ(t) = 2πih′ (t) has constant argument. Hence, h′ (t) is either positive
or negative, so t 7→ exp(2πih(t) + δ(0) is a parametrization of the unit circle
starting at γ(0) = exp δ(0) and the unit circle is traversed either clockwise or
counterclockwise.
Exercises.
1. Consider the metric |dz|/|z| on C \ {0} := C∗ and let ℓC∗ (γ) denote the
length of a path γ in C∗ relative to this metric. If γ is a closed path in C∗ ,
prove
ℓC∗ (γ) ≥ 2π|I(γ, 0)|.
When does equality hold?
2. Suppose f is holomorphic on D and f (D) ⊆ D \ {0}. Prove that |f ′ (0)| ≤
2/e. Determine when equality holds.
The hyperbolic metric and geometric function theory 51

13. Rigidity theorems


If a holomorphic mapping of hyperbolic regions is not a covering, then strict
inequality holds in the general version of the Schwarz-Pick Lemma. It is often
possible to provide a quantitative version of this strict inequality that is inde-
pendent of the holomorphic mapping for multiply connected regions. We begin
by establishing a refinement of Schwarz’s Lemma.
Lemma 13.1. Suppose 0 6= a ∈ D and 0 < t < 1. If f is a holomorphic self-map
of D, f (0) = 0 and |f (a)| ≤ t|a|, then
t + |a|
(13.1) |f ′ (0)| ≤ < 1.
1 + t|a|
Proof. The Three-point Schwarz-Pick Lemma (Theorem 4.4) with z = 0 = v
and w = a gives
dD (f ′ (0), f (a)/a) = dD (f ∗ (0, 0), f ∗ (a, 0))
≤ dD (0, a).
Hence,
dD (0, |f ′ (0)|) = dD (0, f ′ (0))
≤ dD (0.f (a)/a) + dD (0, a)
≤ dD (0, t) + dD (0, −|a|)
= d(−|a|, t),
which is equivalent to (13.1).
Theorem 13.2. Suppose ∆ and Ω are hyperbolic regions with a ∈ ∆, b ∈ Ω
and ∆ is not simply connected. There is a constant α = α(a, ∆; b, Ω) ∈ [0, 1)
such that if f : ∆ → Ω is any holomorphic mapping with f (a) = b that is not a
covering, then
(13.2) λΩ (f (a))|f ′ (a)| ≤ αλ∆ (a);
or equivalently, f ∆,Ω (a) ≤ α. Moreover, for all z ∈ ∆
α + tanh d∆ (a, z)
(13.3) f ∆,Ω (z) ≤ < 1.
1 + α tanh d∆ (a, z)
Proof. Let h : D → ∆ and k : D → Ω be holomorphic coverings with h(0) = a
and k(0) = b. Because ∆ is not simply connected, the fiber h−1 (a) is a discrete
subset of D and contains infinitely many points in addition to 0. Let 0 < r =
min{|z| : z ∈ h−1 (a), z 6= 0} < 1. The set h−1 (a) ∩ {z : |z| = r} is finite, say
ãj , 1 ≤ j ≤ m. Next, {|z| : z ∈ k −1 (b)} is a discrete subset of [0, 1), so this set
contains finitely many values in the interval [0, r). Let s be the maximum value
of the finite set {|z| : z ∈ k −1 (b)} ∩ [0, r). Suppose f : ∆ → Ω is any holomorphic
mapping with f (a) = b and f is not a covering. Let f˜ be the unique lift of f ◦ h
relative to k that satisfies f˜(0) = 0. Then |f˜′ (0)| = f ∆,Ω (a). Because f is not
a covering, f˜ is not a rotation about the origin. From f ◦ h = k ◦ f˜ we deduce
52 Beardon and Minda IWQCMA05

that f˜ maps h−1 (a) into k −1 (b). In particular, |f˜(ã)| ≤ s = t|ã|, where ã = ã1
and t = s/r < 1. Lemma 13.1 gives
(s/r) + |ã| (s/r) + r
|f˜′ (0)| ≤ = = α < 1.
1 + (s/r)|ã| 1+s
Since |f˜′ (0)| = f ∆,Ω (a), this establishes (13.2). Inequality (13.3) follows imme-
diately from Corollary 11.3.

The pointwise result (13.2) is due to Minda [24] and was motivated by the
Aumann-Carathéodory Rigidity Theorem [4] which is the special case when
Ω = ∆ and a = b. The global result (13.3) is due to the authors [9]. The Aumann-
Carathéodory Rigidity Theorem asserts there is a constant α = α(a, Ω) ∈ [0, 1)
such that |f ′ (a)| ≤ α for all holomorphic self-maps of Ω that fix a and are not
conformal automorphisms. The exact value of the Aumann-Carathéodory rigid-
ity constant for an annulus was determined in [23]. The following extension of
the Aumann-Carathéodory Rigidity Theorem to a local result is due to the au-
thors [9]. The corollary is given in Euclidean terms and asserts that holomorphic
self-maps with a fixed point are locally strict Euclidean contractions if they are
not conformal automorphisms.
Corollary 13.3 (Aumann-Carathéodory Rigidity Theorem - Local Version).
Suppose Ω is a hyperbolic region, a ∈ Ω and Ω is not simply connected. There
is a constant β = β(a, Ω) ∈ [0, 1) and a neighborhood N of a such that if f is a
holomorphic self-map of Ω that fixes a and is not a conformal automorphism of
Ω, then |f ′ (z)| ≤ β for all z ∈ N .

Proof. From the theorem


λΩ (z) α + tanh dΩ (a, z)
|f ′ (z)| ≤ .
λΩ (f (z)) 1 + α tanh dΩ (a, z)
Set M (r) = max{λΩ (z) : dΩ (a, z) ≤ r} and m(r) = min{λΩ (z) : dΩ (a, z) ≤ r}.
Since f (DΩ (a, r)) ⊆ DΩ (a, r), we have
M (r) α + tanh dΩ (a, z)
|f ′ (z)| ≤ .
m(r) 1 + α tanh dΩ (a, z)
The right-hand side of the preceding equality is independent of f and tends to
α as r → 1. Therefore, for α < β < 1 there exists r > 0 such that
M (r) α + tanh dΩ (a, z)
≤β
m(r) 1 + α tanh dΩ (a, z)
for dΩ (a, z) ≤ r. Then |f ′ (a)| ≤ β holds in DΩ (a, r).

Our final topic is a rigidity theorem for holomorphic maps between annuli.
The original results of this type are due to Huber [17]. Marden, Richards and
Rodin [21] presented an extensive generalization of Huber’s work to holomorphic
self-maps of hyperbolic Riemann surfaces.
The hyperbolic metric and geometric function theory 53

Definition 13.4. Suppose 1 < R, S ≤ +∞ and f : AR → AS is a continuous


function. The degree of f is the integer deg f := I(f ◦ γ, 0), where γ(t) =
exp(2πit).
The positively oriented unit circle γ generates the fundamental group of both
AR and AS . Therefore, for any continuous map f : AR → AS , f ◦ γ ≈ γ n for
the unique integer n = deg f , where ≈ denotes free homotopy. Algebraically,
f induces a homomorphism f∗ from π(AR , a) ∼ = Z to π(AS , f (a)) ∼
= Z and the
image of 1 is an integer n; here π(AR , a) denotes the fundamental group of AR
with base point a. The reader should verify that
deg(f ◦ g) = (deg f ) (deg g).
Since the degree of the identity map is one, this implies that deg f = ±1 for any
homeomorphism f . If f is holomorphic, then
Z ′
1 f (z)
deg f = dz.
2πi γ f (z)
Suppose f, g : AR → AS are continuous functions. Then deg f = deg g if and
only if f and g are homotopic maps of AR into AS .
Example 13.5. For any integer n the holomorphic self-map pn (z) = z n of C∗
has degree n. Given annuli AR and AS with 1 < R, S < +∞, it is easy to
construct a continuous map of AR into AS with degree n; for example, the
function z 7→ (z/|z|)n has degree n. For R = S each conformal automorphism
has degree ±1. In fact, the rotations z 7→ eiθ z have degree 1 and the maps
z 7→ eiθ /z have degree −1. Constant self-maps of AR have degree 0. Can you find
a holomorphic self-map of AR with degree n 6= 0, ±1? Surprisingly, the answer
is negative! Holomorphic mappings of proper annuli are very rigid. The moduli
of the annuli provide sharp bounds for the possible degrees of a holomorphic
mapping of one annulus into another.
Theorem 13.6. If f : AR → AS is a holomorphic mapping, then
mod AS
(13.4) | deg f | ≤ .
mod AR
For n = deg f 6= 0 equality holds if and only if S = R|n| and f (z) = eiθ z n for
some θ ∈ R.
Proof. Let γ(t) = exp(2πit) for t ∈ [0, 1]. If n = deg f , then I(f ◦ γ, 0) = n, so
Theorem 12.2 gives
2π 2 |n|
≤ ℓS (f ◦ γ).
mod AS
Since holomorphic functions are distance decreasing relative to the hyperbolic
metric,
2π 2
ℓS (f ◦ γ) ≤ ℓR (γ) = .
mod AR
The preceding two inequalities imply (13.4). Suppose equality holds. Then f is a
covering of AR onto AS that maps the unit circle onto itself. By post-composing
54 Beardon and Minda IWQCMA05

f with a rotation about the origin, we may assume that f fixes 1. Equality
in (13.4) implies S = R|n| , where n = deg f . The covering f lifts relative to
pn (z) = z n to a holomorphic self-covering f˜ of AR that fixes 1. Theorem 10.6
implies that f˜ is the identity and so f (z) = z n .
Corollary 13.7 (Annulus Theorem). Suppose f is a holomorphic self-map of
AR . Then | deg f | ≤ 1 and equality holds if and only if f ∈ A(AR ).
Proof. The inequality follows immediately from Theorem 13.6. If f ∈ A(AR ),
then | deg f | = 1 since this holds for any homeomorphism. It remains to show
that if | deg f | = 1, then f ∈ A(AR ). Equality implies f maps the unit circle into
itself. By post-composing f with a rotation about the origin, we may assume
f fixes 1. By Theorem 10.6 if a self-covering of a hyperbolic region has a fixed
point, it is a conformal automorphism.
Note that if f is any holomorphic self-map of AR that is not a conformal
automorphism, then deg f = 0. A result analogous to Corollary 13.7 is not
valid for a punctured disk. For each integer n ≥ 0 the function z 7→ z n is a
holomoprhic self-map of D∗ with degree n.
Exercises.
1. Show that Theorem 13.2 is false when ∆ is simply connected. Hint: Sup-
pose ∆ = D and a = 0. For any number r ∈ [0, 1) show there exists a
holomorphic function f : D → Ω that is not a covering and f D,Ω (0) = r.
2. Suppose f is a holomorphic self-map of C \ {0}. Prove that deg f = 0 if
and only if f = exp ◦g for some holomorphic function g defined on C \ {0}.

14. Further reading


There are numerous topics involving the hyperbolic metric and geometric func-
tion theory that are not discussed in these notes. The subject is too extensive to
include even a reasonably complete bibliography. We mention selected books and
papers that the reader might find interesting. Anderson [3] gives an elementary
introduction to hyperbolic geometry in two dimensions. Krantz [18] provides an
elementary introduction to certain aspects of the hyperbolic metric in complex
analysis.
Ahlfors introduced the powerful method of ultrahyperbolic metrics [1]. A
discussion of this method and several applications to geometric function theory,
including a lower bound for the Bloch constant, can be found in [2]. Ahlfors’
method can be used to estimate various types of Bloch constants, see [25], [26],
[27]. In a long paper Heins [15] treates the general topic of conformal metrics on
Riemann surfaces. He gives a detailed treatment of SK-metrics, a generalization
of ultrahyperbolic metrics. Roughly speaking, SK-metrics are to metrics with
curvature −1 as subharmonic functions are to harmonic functions.
The circle of ideas surrounding the theorems of Picard, Landau and Schottky
and Montel’s normality criterion all involve three omitted values. Theorems of
The hyperbolic metric and geometric function theory 55

this type follow immediately from the existence of the hyperbolic metric on C∞
punctured at three points. Interestingly, only a metric with curvature at most −1
on a thrice punctured sphere is needed to establish these results. An elementary
construction of such a metric, based on earlier work of R. M. Robinson [31], is
given in [30].
Hejhal [16] obtained a Carathéodory kernel-type of theorem for coverings of
the unit disk onto hyperbolic regions. This result implies that the hyperbolic
metric depends continuously on the region.
The method of polarization was extended by Solynin to apply to the hyperbolic
metric, see [10]. It include earlier work of Weitsman [35] on symmetrization and
Minda [29] on a reflection principle for the hyperbolic metric.
For the role of hyperbolic geometry in the study of discrete groups of Möbius
transformations, see [5]. This reference includes a brief treatment of hyperbolic
trigonometry.

References
1. Ahlfors, L.V., An extension of Schwarz’s Lemma, Trans. Amer. Math. Soc., 43 (1938),
259-264.
2. Ahlfors, L.V., Conformal invariants, McGraw-Hill, 1973.
3. Anderson, J.W., Hyperbolic geometry, 2nd. ed., Springer Undergraduate Mathematics
Series, Springer, 2005.
4. Aumann, G. and Carathéodory, C., Ein Satz über die konforme Abbildung mehrfach
zusammenhängender Gebiete, Math. Ann., 109 (1934), 756-763.
5. Beardon, A.F., The geometry of discrete groups, Graduate Texts in Mathematics 91,
Springer-Verlag, New York 1983.
6. Beardon, A.F., The Schwarz-Pick Lemma for derivatives, Proc. Amer. Math. Soc., 125
(1997), 3255-3256.
7. Beardon, A.F. and Carne, T.K., A strengthening of the Schwarz-Pick Inequality, Amer.
Math. Monthly, 99 (1992), 216-217.
8. Beardon, A.F. and Minda, D., A multi-point Schwarz-Pick Lemma, J. d’Analyse Math.,
92 (2004), 81-104.
9. Beardon, A.F. and Minda, D., Holomorphic self-maps and contractions, submitted.
10. Brock, F. and Solynin, A.Yu., An approach to symmetrization via polarization, Trans.
Amer. Math. Soc., 352 (2000), 1759-1796.
11. Carathéodory, C., Theory of functions of a complex variable, Vol. II, Chelsea, 1960.
12. Chen, Huaihui, On the Bloch constant, Approximation, complex analysis and potential
theory (Montreal, QC 2000), NATO Sci. Ser. II Math. Phys. Chem., 37, Kluwer Acad.
Publ., Dordrecht, 2001.
13. Duren, P.L., Univalent functions, Springer-Verlag, New York 1983.
14. Goluzin, G.M., Geometric theory of functions of a complex variable, Amer. Math. Soc.,
1969.
15. Heins, M., On a class of conformal metrics, Nagoya Math. J., 21 (1962), 1-60.
16. Hejhal, D., Universal covering maps for variable regions, Math. Z., 137 (1974), 7-20.
17. Huber, H., Über analytische Abbildungen von Ringgebieten in Ringgebiete, Compositio
Math., 9 (1951), 161-168.
18. Krantz, S.G., Complex analysis: the geometric viewpoint, 2nd. ed., Carus Mathematical
Monograph 23, Mathematical Association of America, 2003.
56 Beardon and Minda IWQCMA05

19. Ma, W. and Minda, D., Geometric properties of hyperbolic geodesics, Proceedings of the
International Workshop on Quasiconformal Mappings and their Applications, pp ??.
20. Ma, W. and Minda, D., Euclidean properties of hyperbolic polar coordinates, submitted
21. Marden, A., Richards, I. and Rodin, B., Analytic self-mappings of Riemann surfaces, J.
Analyse Math., 18 (1967), 197-225.
22. Massey, W.S., Algebraic topology: an introduction, Springer-Verlag, 1977.
23. Minda, C.D., The Aumann-Carathéodory rigidity constant for doubly connected regions,
Kodai Math. J. 2 (1979), 420-426.
24. Minda, C.D., The hyperbolic metric and coverings of Riemann surfaces, Pacific J. Math.
84 (1979), 171-182.
25. Minda, C.D., Bloch constants, J. Analyse Math. 41 (1982), 54-84.
26. Minda, C.D., Lower bounds for the hyperbolic metric in convex regions, Rocky Mtn. J.
Math. 13 (1983), 61-69.
27. Minda, C.D., The hyperbolic metric and Bloch constants for spherically-convex regions,
Complex Variables Theory Appl. 5 (1986), 127-140.
28. Minda, C.D., The strong form of Ahlfors’ Lemma, Rocky Mtn. J. Math. 17 (1987), 457-461.
29. Minda, C.D., A reflection principle for the hyperbolic metric and applications to geometric
function theory, Complex Variables Theory Appl. 8 (1987), 129-144.
30. Minda, D. and Schober, G., Another elementary approach to the theorems of Landau,
Montel, Picard, and Schottky, Complex Variables Theory Appl. 2 (1983), 157-164.
31. Robinson, R.M, A generalization of Picard’s and related theorems, Duke Math. J. 5
(1939), 118-132.
32. Royden, H., The Ahlfors-Schwarz Lemma: the case of equality, J. Analyse Math. 46
(1986), 261-270.
33. Pick, G., Über eine Eigenschaft der konformen Abbildung kreisförmiger Bereiche, Math.
Ann. 77 (1915), 1-6.
34. Veech, W.A., A second course in complex analysis, W.A. Benjamin, 1967.
35. Weitsman, A., Symmetrization and the Poincaré metric, Ann. of Math.(2) 124 (1986),
159-169.
36. Yamada, A., Bounded analytic functions and metrics of constant curvature, Kodai Math.
J. 11 (1988), 317-324.

A.F. Beardon E-mail: A.F.Beardon@dpmms.cam.ac.uk


Address: Centre for Mathematical Sciences, University of Cambridge, Wilberforce Road,
Cambridge CB3 0WB, England

D. Minda E-mail: David.Minda@math.uc.edu


Address: Department of Mathematical Sciences, University of Cincinnati, Cincinnati, Ohio
45221-0025, USA
Proceedings of the International Workshop on Quasiconformal
Mappings and their Applications (IWQCMA05)

Isometries of relative metrics

Peter Hästö

Abstract. This survey contains a review of the basics of Möbius mappings


and some basic metrics used in relations with quasiconformal mappings, the
jG metric and the quasihyperbolic metric. The second and third sections are
devoted to the study of the isometries of these two metrics.

Keywords. Relative metrics, isometries, distance ratio metric, Seittenranta’s


metric.
2000 MSC. 30F45 (primary), 30C65 (secondary).

Contents

1. Overview 57
2. Möbius mappings 58
3. The jG metric 61
3.1. Isometries of j-type metrics 63
3.2. Other properties of j-type metrics 66
4. The quasihyperbolic metric 66
Additional notation 68
4.1. Isometries which are Möbius 68
4.2. Curvature of the quasihyperbolic metric 71
4.3. Isometries of the quasihyperbolic metric 73
References 76

1. Overview
These lecture notes consist of three parts: In the first part the basic theory
of Möbius mappings is reviewed. Particular emphasis will be given to concrete
calculations within the context of a single mapping in Euclidean space. Although
this presentation is perhaps not the most elegant one possible, it has the advan-
tage that it does a good job in preparing us for the isometry questions that come
up later. For a more detailed exposition of the basics of Möbius mappings see
e.g. [2, 31].
Supported by the Academy of Finland.
58 Peter Hästö IWQCMA05

The latter two parts deal with the problem of characterizing isometries of two
metrics which have turned out to be very important in the theory of quasicon-
formal mappings, namely, the jG and the quasihyperbolic metric. Specifically, in
the second part we deal with the jG metric — this part is based on joint work
with Z. Ibragimov and H. Lindén [18] in Computational Methods and Function
Theory. The third part reproduces parts of my recent manuscript [15], which
deals with isometries of the quasihyperbolic metric.
Characterizing isometries of a metric can in some sense be thought of as solving
a (system of) functional equation(s): we know that
df (G) (f (x), f (y)) = dG (x, y)
for all x, y ∈ G and we want to determine f . However, the fact that we have at
our disposal a continuum of functional equations implies that the methods used
to approach this problem are somewhat different than those usually found when
dealing with functional equations. Thus our methods will often be based on
some geometric considerations: we will employ geodesics (locally and globally),
intrinsic curvature, as well as limiting behavior of the metric in infinitesimal
regions more generally.
Many other properties of these and related metrics have also been studied. A
review of some of these results is presented in the chapter by H. Lindén in these
notes.

2. Möbius mappings
We denote by Rn = Rn ∪{∞} the one-point compactification of Rn , so its open
balls are the open balls of Rn , complements of closed balls in Rn and half-spaces.
If D ⊂ Rn we denote by ∂D and D its boundary and closure, respectively, all
with respect to Rn . By B n (x, r) and S n−1 (x, r) we denote the open ball centered
at x ∈ Rn with radius r > 0, and its boundary, respectively. For x ∈ D ( Rn we
denote δ(x) = d(x, ∂D) = min{|x − z| : z ∈ ∂D}. By [x, y], (x, y] we denote the
closed and half-open segment between x and y, respectively.
The (absolute) cross-ratio of four distinct points is defined by
|a − c| |b − d|
|a, b, c, d| = ,
|a − b| |c − d|
with the understanding that |∞−x|
|∞−y|
= 1 for all x, y ∈ Rn . A homeomorphism
f : Rn → Rn is a Möbius mapping if
|f (a), f (b), f (c), f (d)| = |a, b, c, d|
for every quadruple of distinct points a, b, c, d ∈ Rn . A mapping of a subdomain
of Rn is Möbius, if it is a restriction of a Möbius mapping defined on Rn .
Although the previous definition is very compact and brings out one important
aspect of Möbius mappings, it does not tell us what the behavior of a Möbius
mapping is in terms of geometry. However, it is not so difficult to get some
Isometries of relative metrics 59

results in this direction: Let us regard three points a, b, c as fixed and a fourth
point x as variable. Then the cross-ratio equation reads
|a − c| |b − x| |a′ − c′ | |b′ − x′ |
= ′ ,
|a − b| |c − x| |a − b′ | |c′ − x′ |
where a′ is the image of a under the Möbius mapping. We can rewrite this as
|b − x| |b′ − x′ |
=C ′ ,
|c − x| |c − x′ |
where C is a constant not depending on x. However, for fixed b, c and C > 0
the set
n |b − x| o
x ∈ Rn : =C
|c − x|
is a sphere. Thus the previous equation implies that the Möbius mapping maps
spheres to spheres. The converse of this statement is also true, see [4].
It is also possible to take a more constructive approach to Möbius mappings.
Let us first of all make the trivial observation that a mapping which preserves
Euclidean distances is Möbius. Second, we note that mappings preserving ra-
tios of Euclidean distances (so-called similarity mappings) are Möbius. These
mappings are:
• translations;
• reflections;
• rotations; and
• dilatations.
Are there any other Möbius mappings?
From the definition it is clear that the set of Möbius mappings is closed under
composition (in fact, the set is a group under composition). Thus we may employ
a very useful trick in trying to identify any other Möbius mappings, namely, we
normalize by mappings that we already know are Möbius. This means that we
consider the mapping g = s1 ◦ f ◦ s2 , where f is our original Möbius mapping and
s1 and s2 are similarities. Suppose first that f is such that f (∞) = ∞. Inserting
d = ∞ in the definition implies that
|a − c| |f (a) − f (c)|
= |a, b, c, ∞| = |f (a), f (b), f (c), ∞| = ,
|a − b| |f (a) − f (b)|
so f is a similarity. Otherwise there exists a finite point a such that f (a) = ∞.
By an auxiliary similarity we may assume that a = 0 (i.e. we choose s2 (x) = x+a
above). Similarly, f (∞) = b 6= ∞, and if we choose s1 (x) = x − b, then g is a
Möbius mapping which swaps 0 and ∞. Using this in the equation gives
|b| |g(c)|
= |∞, b, c, 0| = |0, g(b), g(c), ∞| = .
|c| |g(b)|
From this we see that |x| |g(x)| is a constant. By another similarity we may
assume that this constant equals 1, so that |g(x)| = |x|−1 for every x ∈ Rn . To
60 Peter Hästö IWQCMA05

get a grip of the non-radial action of g we use the equation inserting 0 and ∞ in
other places:
|b − d| |g(b) − g(d)|
= |∞, b, 0, d| = |0, g(b), ∞, g(d)| =
|d| |g(b)|
. Using the previous formula for |g(b)| this gives
|b − d|
(2.1) |g(b) − g(d)| = ,
|b| |d|
which is a central formula for calculating how a Möbius mapping affects distances.
We can rewrite (2.1) as
|g(b) − g(d)|2 |b − d|2
= .
|g(b)| |g(d)| |b| |d|
Using the cosine formula

c ,
|b − d|2 = |b|2 + |d|2 − 2|b| |d| cos b0d
c stands for the angle between the vectors b − 0 and d − 0, and similarly
where b0d
for |g(b)−g(d)|2 we see that g preserves angles at the origin and lines through the
origin. Thus, up to additional normalization by a reflection and/or a rotation,
we see that g(x) = x |x|−2 .
A Möbius mapping which swaps ∞ and with a point of Rn and which maps
every line through this point to itself is called an inversion. Note that every
inversion is an involution, i.e. it is its own inverse. The point which is mapped to
∞ is called the center of inversion. We have shown that every inversion equals
x 7→ x |x|−2 , up to similarity. In particular, every Möbius mapping can be written
as s ◦ i, where s is a similarity and i is an inversion or the identity.
Now that we have identified the Möbius mappings we can proceed to show
the following basic property: given two ordered triples of distinct points in Rn ,
(a, b, c) and (a′ , b′ , c′ ), there exists a Möbius map f with f (a) = a′ , f (b) = b′ and
f (c) = c′ . It is clearly sufficient to show this claim in the case when a′ , b′ and c′
are the vertices of an equilateral triangle. Let us first find a point x such that
|a − c| |b − x| |b − c| |a − x|
= = 1.
|a − b| |c − x| |a − b| |c − x|
The easiest way to see that such a point x exists is to use an inversion i with
center a. Then the equations to satisfy become
|i(b) − z| |i(b) − i(c)|
= = 1.
|i(c) − z| |i(c) − z|
We see that the first fraction describes a hyperplane which is the perpendicular
bisector of the segment [i(b), i(c)] and the second fraction the sphere with center
i(c) and radius |i(b) − i(c)|. Since these objects clearly intersect, we can find a
Isometries of relative metrics 61

suitable z, and then our x is given by i(z). Let ı̃ be an inversion with center x.
The choice of x implies that
|ı̃(a) − ı̃(c)| |ı̃(b) − ı̃(c)|
= = 1,
|ı̃(a) − ı̃(b)| |ı̃(a) − ı̃(b)|

so ı̃(a), ı̃(b), ı̃(c) are the vertices of an equilateral triangle. These can be
mapped to the given points (a′ , b′ , c′ ) by a similarity transform s, so our final
Möbius map is then s ◦ ı̃.

3. The jG metric
This section reproduces parts of the article [18] on isometries of some relative
metrics. The term “relative metric” implies that the metric is evaluated in
a proper subdomain of Rn relative to its boundary. More precisely, we want
the metric to blow up towards the boundary of the domain, i.e., we want the
boundary to be at infinity intrinsically.
Let D ( Rn be a domain containing the points x and y. The well-known
distance ratio metric is defined by
 
|x − y|
jD (x, y) = log 1 + ,
min{δ(x), δ(y)}
where δ(·) = dist( · , ∂D) denotes distance to the boundary. It was used, for in-
stance, by Gehring and Osgood [11] to characterize uniform domains (namely, in
such domains the jD metric is quasiconvex). Note that this metric has sometimes
been called simply “the relative metric”, and will be used in this meaning.
To see how these metrics fit into a larger framework we recall the concept of
an inner metric. Let d be a metric in D and γ be a path in D (i.e. a continuous
mapping from an interval I to D). The length (or, more explicitly, d-length) of
γ is defined as
k−1
X 
d(γ) = sup d γ(ti ), γ(ti+1 ) ,
i=1

where the supremum is taken over k and all increasing sequences (ti )ki=1 of points
in I. Then the inner or intrinsic metric of d is defined by
˜ y) = inf d(γ),
d(x,
γ

where the infimum is taken over all paths γ connecting x and y in D (note
that this need not be finite, unless D is rectifiably connected). It is clear that
˜ y) and that d(γ) = d(γ)
d(x, y) ≤ d(x, ˜ for any metric and path. The theory of
length-metrics, including in particular intrinsic metrics, is presented e.g. in [5, 6].
Suppose now that D ⊂ Rn and d is a metric in D. If
¯ = lim d(x, y)
d(x)
y→x |x − y|
62 Peter Hästö IWQCMA05

exists for all x ∈ D and is continuous, then we can express the inner metric of d
by Z
˜ ¯ |dz|,
d(x, y) = inf d(z)
γ γ
where |dz| represents integration with respect to d-arclength, and the infimum
is taken over rectifiable curves with end points x and y. In this case d˜ is called a
conformal metric. We easily see that the inner metric of jD is the quasihyperbolic
metric, Z
|dz|
j̃D (x, y) = kD (x, y) = inf .
γ γ d(z, ∂D)
Length-metrics are interesting from a geometric point of view, but for getting
explicit estimates they are often of little use. The role of point-distance functions,
like the jD metric, is that they share features with their inner metrics, but are
much more explicit.
In this paper we want to consider not only the jD metric, but all metrics
which resemble them in the very small and very large scale. The small scale
equivalence implies that the metrics have the same inner metrics, whereas the
large scale equivalence allows us to get a hold of the boundary behavior and thus
start unraveling the isometry story.
Remark 3.1. Note that jD is really families of metrics, namely for every domain
D we have one metric. We will continue to use this convention when talking about
this and other metrics in this paper.
Definition 3.2. We say that d is a j-type metric if the following three conditions
hold on every domain D ( Rn :
1. dD is a metric on D.
2. For each y ∈ D and for each sequence (xi ) with jD (xi , y) → 0 we have
dD (xi , y)
lim = 1.
i→∞ jD (xi , y)

3. For each y ∈ D and for each sequence (xi ) with jD (xi , y) → ∞ we have

lim dD (xi , y) − jD (xi , y) = 0.
i→∞

The fact that y can be any interior point in (2) means that being a j-type
metric is quite a strong condition; for instance, if d and f ◦ d are j-type metrics,
then f = id (Corollary 3.12).
It seems to be quite difficult to construct other natural metrics of j-type. The
main purpose of our more abstract treatment is to highlight the features that
are crucial, which in turn indicates that these techniques might be relevant also
for handling the isometries of the corresponding inner metrics.
In this part we characterize the isometries of j-type metrics. We start by
collecting some basic properties of these metrics. In Section 3.1 we solve the
Isometries of relative metrics 63

isometry problem for j-type metrics using a boundary rigidity result, and in
Section 3.2 we list some additional properties whose proofs can be found in [18].
The following result is more or less restatements of the definition. However,
it directly implies that we may restrict our focus very much without losing any
isometries.
Proposition 3.3. If d is a j-type, then
 
|x − y| dD (x, y) 1
lim dD (x, y) − log =0 and lim =
y→∂D\{∞} δ(y) y→x |x − y| δ(x)
for every x ∈ D. The inner metric of a j-type metrics is the quasihyperbolic
metric kD . In particular, every isometry of a j-type metric is an isometry of kD .

Every isometry of kD is a conformal mapping [29, Theorem 2.6]. Hence we


conclude:
Corollary 3.4. Every isometry of a j or δ-type metric is conformal. In partic-
ular, if n ≥ 3, then such an isometry is Möbius.

We plunge right into the main result of this section, a characterization of the
isometries of j-type metrics. In Section 3.2 we derive some miscellaneous results,
which give a clearer picture of j-type metrics.

3.1. Isometries of j-type metrics. The proof of the following theorem is


partly based on ideas from [17]. Incidentally, it is possible to give a much simpler
proof for the particular case of the jD metric itself, since in this case we can cancel
the logarithm and the 1+ terms. Thus f is a jD isometry if and only if
|x − y| |f (x) − f (y)|
= ,
min{δ(x), δ(y)} min{δ ′ (f (x)), δ ′ (f (y))}
where, as usual, δ ′ denotes the distance to the boundary in the image domain
f (D). The reader is challenged to find the very short argument which shows that
this implies that the isometry is a similarity.
In the general case of j-type metrics we have less information about the metric,
so we have to look at what happens at the boundary. In this case we can
nevertheless prove the following theorem, whose proof is reproduced from [18].
Theorem 3.5. Let d be a j-type metric, D ( Rn and f : D → Rn be a d-
isometry. Then either
1. f is a similarity, or
2. D = Rn \ {a} and, up to similarity, f is an inversion in a sphere centered
at a.

Proof. Denote D′ = f (D) and δ ′ (x) = d(x, ∂f (D)). Fix z ∈ ∂D \ {∞} and let
(zi ) be a sequence of points in D tending to z. We first assume
 that there exists
a subsequence, which we also denote by (zi ), such that f (zi ) converges to some
64 Peter Hästö IWQCMA05

point w1 ∈ Rn . Since d is a j-type metric we see, using Proposition 3.3 for the
third equality, that for every x ∈ D we have that

0 = lim dD′ (f (x), f (zi )) − dD (x, zi )
i→∞
 1   1 
= lim dD′ (f (x), f (zi )) − log ′ − lim dD (x, zi ) − log +
i→∞ δ (f (zi )) i→∞ δ(zi )
δ(zi )
+ lim log ′
i→∞ δ (f (zi ))
|f (x) − w1 | δ(zi )
= log + lim log ′ .
|x − z| i→∞ δ (f (zi ))
Taking exponentials gives
δ ′ (f (zi )) |f (x) − w1 |
(3.6) lim = < ∞.
i→∞ δ(zi ) |x − z|

Suppose now that (ẑi ) is a second sequence of points in D tending to z, but


that this time f (ẑi ) → w2 ∈ Rn \ {w1 }. Using x = ẑj for every j = 1, 2, . . . in
(3.6) gives
δ ′ (f (zi )) |f (ẑj ) − w1 |
lim = →∞
i→∞ δ(zi ) |ẑj − z|
as j → ∞, which is a contradiction. In other words, f (ẑi ) → w1 for every
sequence of points (ẑi ) → z, so we may extend f continuously to D by defining
f (z) = limi→∞ f (zi ). Therefore we conclude from (3.6), since the left-hand side
of this equation does not depend on x, that
|f (x) − f (z)| = hf (z)|x − z|
for some function hf : ∂D → (0, ∞). This means that for z, w ∈ ∂D we have
hf (z)|w − z| = |f (w) − f (z)| = hf (w)|w − z|,
so hf is in fact a constant. Therefore f acts as a similarity, say g, on the boundary.
We then extend f to all of Rn by setting f (x) = g(x) outside the original domain
of definition. Then it is clear that
(3.7) |f (x) − f (z)| = hf |x − z|
for every point x ∈ Rn , i.e. the sphere S n−1 (z, r) is mapped to S n−1 (f (z), hf r).
This clearly implies that the conformal mapping is Möbius, and a Möbius map-
ping satisfying (3.7) is a similarity.
We still have one assumption to consider. In the beginning of the proof we
assumed that we can find a boundary point z and a sequence (zi ) of points in
D tending to z such that f (zi ) tends to a finite limit. So we suppose now that
no such sequence can be found, i.e. that for every sequence (zi ) of points in D
tending to a boundary point z the sequence f (zi ) tends to ∞. As before we
Isometries of relative metrics 65

conclude that
 
0 = lim dD′ (f (x), f (zi )) − dD (x, zi )
i→∞
 
= lim jD′ (f (x), f (zi )) − jD (x, zi )
i→∞
 
|f (x) − f (zi )| δ(zi )
= lim log .
i→∞ min{δ ′ (f (zi )), δ ′ (f (x))} |x − z|
So it follows that
|f (x) − f (zi )|δ(zi )
lim = |x − z|.
i→∞ min{δ ′ (f (zi )), δ ′ (f (x))}

Since f (zi ) → ∞, we see that we can replace |f (x)−f (zi )| by |f (zi )| in the above
formula. Since the right-hand-side depends on x (which lies in an open set) we
see that the left-hand-side must do so, too, hence we have to choose the second
term in the minimum. Taking this into account we have
gf (z) = lim |f (zi )| δ(zi ) = |x − z| δ ′ (f (x)),
i→∞

where gf : ∂D → (0, ∞). Suppose that D has at least two finite boundary points,
and let a, b ∈ ∂D be such that the open segment (a, b) is contained in D. Now
if we first consider x (in the previous equation) to be the mid-point x of (a, b),
then we conclude that
gf (a) = |x − a| δ ′ (f (x)) = |x − b| δ ′ (f (x)) = gf (b).
But if we take some other point on the segment, then we get gf (a) 6= gf (b), a
contradiction. So only the case when D has a single boundary point remains to
consider. Then we have
lim |f (zi )| |zi − a| = |x − a| |f (x) − b|
i→∞

(for D = Rn \ {a} and D′ = Rn \ {b}) and we directly see that x 7→ f (x) + b − a


is an inversion, which concludes the proof.
Corollary 3.8. Let d be a similarity invariant j-type metric and let D ( Rn .
Then f : D → Rn is a d-isometry if and only if
1. f is a similarity, or
2. D = Rn \ {a} and, up to similarity, f is the inversion in a sphere centered
at a.

Proof. The previous proposition established that every d-isometry is of the given
kind. If f is a similarity, then it is an isometry by assumption. So it remains
(after normalization) to consider the case D = Rn \ {0}. In thiscase we see that
similarity invariance implies that dD (x, y) depends only on max |x| , |y| and the
|y| |x|
d On the other hand, an inversion in a sphere about the origin swaps
angle x0y.
d invariant, so we see that it is an isometry.
|x|/|y| and |y|/|x| and leaves x0y
66 Peter Hästö IWQCMA05

3.2. Other properties of j-type metrics. We said before that every j-type
metric has an upper bound in terms of the quasihyperbolic metric. Surprisingly,
it is also possible to get a universal lower bound by a metric, the so-called half-
apollonian metric [19]. For a domain D ( Rn this metric is defined by

|x − z|

ηD (x, y) = sup log .
z∈∂D |y − z|
The metric ηD is similarity invariant, and every Möbius mapping is bilipschitz.
The proofs of the following results can be found in [18].
Proposition 3.9. For every j-type metric d and every D ( Rn we have dD ≥ ηD .

Using the previous proposition and the quasihyperbolic upper bound we can
squeeze in j-type metrics to get the exact value on some subset of the domain:
Corollary 3.10. Let w ∈ D and z ∈ ∂D ∩ S n−1 (w, δ(w)). Then dD (x, y) =
jD (x, y) for every x, y ∈ [w, z).

The following is easily checked by a direct computation using the definition of


the j-metric, but also follows from Corollary 3.10 and the fact that line segments
are also geodesic rays for the ηD -metric ([19, Example 3.4]).
Corollary 3.11. Let w ∈ D and z ∈ ∂D ∩ S n−1 (w, δ(w)). Then [w, z) is a
geodesic ray for the j-type metric d, i.e. for every x, ξ, y ∈ [w, z) in this order we
have
dD (x, y) = dD (x, ξ) + dD (ξ, y).

In general, if we have a metric d and a subadditive function f : [0, ∞) → [0, ∞)


for which f (x) = 0 if and only if x = 0, then f ◦ d is also a metric. It turns out
that the conditions for begin a j-type metric are so rigid, that this transformation
is never possible in this context:
Corollary 3.12. Let d be a j-type metric and fD : [0, ∞) → [0, ∞) be a family
of arbitrary functions. If f ◦ d is a j-type metric, then fD = id for all relevant
D.

4. The quasihyperbolic metric


The remainder of this article is reproduced with minor modifications from
[15].
Let D ( R2 be an open set and denote δ(x) = d(x, ∂D), the distance to the
boundary. The quasihyperbolic metric in D is the conformal metric with the
density δ(x)−1 , in other words, the metric is given by
Z
ds(z)
kD (x, y) = inf ,
γ γ δ(z)
where the infimum is taken over paths γ connecting x and y in D and ds repre-
sents integration with respect to arc-length.
Isometries of relative metrics 67

The quasihyperbolic metric was first introduced in the seventies, and since
then it has found innumerable applications, especially in the theory of quasicon-
formal mappings, see, e.g. [11, 12, 22, 28, 29]; new connections are still being
made, for instance P. Jones and S. Smirnov [24] recently gave a criterion for re-
movability of a set in the domain of definition of a Sobolev space in terms of the
integrability of the quasihyperbolic metric, see also [25], and Z. Balogh and S.
Buckley [1] used the metric in a geometric characterization of Gromov hyperbolic
spaces.
Despite the prominence of the quasihyperbolic metric, there have been almost
no investigations of its geometry. Three exceptions are the papers by G. Martin
[28] and Martin and B. Osgood [29], the second of which was the main motivation
for the approach presented in this paper, and the thesis by H. Lindén [27]. Part of
the reason for this lack of geometrical investigations is probably that the density
of the quasihyperbolic metric is not differentiable in the entire domain, which
places the metric outside the standard framework of Riemanian metrics.
At least two modifications of the quasihyperbolic metric have been proposed
which do not suffer from this problem. J. Ferrand [10] suggested replacing the
density δ −1 by
|a − b|
σD (x) = sup .
a,b∈∂D |a − x| |b − x|
Note that δ(x)−1 ≤ σD (x) ≤ 2δ(x)−1 , so the Ferrand metric and the quasihyper-
bolic metric are bilipschitz equivalent. Moreover, the Ferrand metric is Möbius
invariant, whereas the quasihyperbolic metric is only Möbius quasi-invariant. A
second variant was proposed more recently by R. Kulkarni and U. Pinkall [26],
see also [23]. The K–P metric is defined by the density
n 2r o
µD (x) = inf : x ∈ B(z, r) ⊂ D .
(r − |x − z|)2
Equivalently, the infimum is taken over the hyperbolic densities of x in balls
contained in D. This density satisfies the same estimate as Ferrand’s density,
i.e. δ(x)−1 ≤ µD (x) ≤ 2δ(x)−1 , and the K–P metric is also Möbius invariant.
Although the Ferrand and K–P metrics are in some sense better behaved than
the quasihyperbolic metric, they suffer from the short-coming that it is very
difficult to get a grip even of the density, even in simple domains.
Despite this, D. Herron, Z. Ibragimov and D. Minda [21] recently managed to
solve the isometry problem of the K–P metric in most cases. By the isometry
problem of the metric d we mean characterizing mappings f : D → R2 with
dD (x, y) = df (D) (f (x), f (y))
for all x, y ∈ D. Notice that in some sense we are here dealing with two dif-
ferent metrics, due to the dependence on the domain. Hence the usual way of
approaching the isometry problem is by looking at some intrinsic features of the
metric which are then preserved under the isometry. Since irregularities (e.g.
cusps) in the domain often lead to more distinctive features, this implies that
the problem is often easier for more complicated domains.
68 Peter Hästö IWQCMA05

The work by Herron, Ibragimov and Minda [21] bears out this heuristic – they
were able to show that all isometries of the K–P metric are Möbius mappings
except in simply and doubly connected domains. Their proof is based on studying
the curvature of the metric. For the quasihyperbolic metric, formulae for the
curvature were worked out already in [29] (see Section 4.2, below), and were
used in that paper to prove that all the isometries of the disc are similarity
mappings. These will be our main tool in this paper. The other source of the
ideas used below are the papers [16, 17, 18, 19] on isometries of some other
similarity and Möbius invariant metrics.
There are three steps in characterizing quasihyperbolic isometries:
1. show that they are conformal;
2. show that they are Möbius; and
3. show that they are similarities.
The first step has been carried out by Martin and Osgood [29, Theorem 2.6] for
completely arbitrary domains, so there is no more work to do there. In Section 4.3
we will use the results from [29] on the curvature of the quasihyperbolic metric,
and some new ideas to prove that the conformal isometries are Möbius (second
step). For this we need to assume that the boundary of the domain is at least
C 3 -smooth. In Section 4.1 we will work on the third step – we show that Möbius
isometries are similarities provided the boundary is C 1 . In Section 4.2 we study
the Gaussian curvature of the quasihyperbolic metric, and the gradient of the
curvature.

Additional notation. We employ some additional conventions in this section:


We tacitly identify R2 with C, and speak about real and imaginary axes, etc.
We will often work with a mapping f : D → R2 . In such cases we will use a
prime to denote quantities on the image side, e.g. x′ = f (x), D′ = f (D) and
δ ′ (x) = d(x, ∂D′ ), and so on.

4.1. Isometries which are Möbius. Let D be a domain and ζ ∈ ∂D. We say
that ζ is circularly accessible, if there exists a disc B ⊂ D such that ζ ∈ ∂B.
Lemma 4.1. Let D ( R2 be a Jordan domain with circularly accessible bound-
ary, and let f : D → R2 be a quasihyperbolic isometry which is also Möbius.
Then, up to composition by similarity mappings, f is the identity or the inver-
sion in a circle centered at a boundary point.

Proof. Assume that f is not a similarity. Since f is a Möbius map, it is an


inversion, up to similarities, which are always isometries of the quasihyperbolic
metric. Thus it suffices to consider the case when f is an inversion in a unit
sphere. Let us denote the center of this sphere by w.
Suppose first that w 6∈ D and let ζ ∈ ∂D be the closest boundary point to w.
For simplicity we normalize the situation so that ζ lies on the positive real axis
and w = 0. Since ζ is circularly accessible, we find a disc B(z, r) ⊂ D which
contains ζ in its closure. Since ζ is the closest boundary point to w, we see that
Isometries of relative metrics 69

z has to lie on the positive real axis, as well. Let x and y be points satisfying
ζ < x < y ≤ ζ(ζ+2r)
ζ+r
. The right-hand inequality ensures that ζ is the closest
boundary point to [x, y], and that ζ ′ is the closest boundary point to [x′ , y ′ ].
Thus we find that
|x − ζ| |x′ − ζ ′ |
kD (x, y) = log and kD′ (x′ , y ′ ) = log ′ .
|y − ζ| |y − ζ ′ |
Since f is the inversion in the unit sphere, we have
|x − ζ|
|x′ − ζ ′ | = ,
|x| |ζ|
and similarly for y. Then the equation exp kD (x, y) = exp kD′ (x′ , y ′ ) gives us
|x − ζ| |x − ζ| |y| |ζ|
= ,
|y − ζ| |x| |ζ| |y − ζ|
i.e. |x| = |y|. This contradiction shows that w ∈ D. Since f maps D into R2 , it
is clear that w 6∈ D, so it follows that w is a boundary point.

We call D a C k domain, if ∂D is locally the graph of a C k function. Note that


if D is a C 1 domain, then certainly every boundary point is circularly accessible.
Proposition 4.2. Let D ( R2 be a C 1 domain, and let f : D → R2 be a quasi-
hyperbolic isometry which is also Möbius. If D is not a half-plane, then f is a
similarity.

Proof. We assume that f is not a similarity map. By the previous lemma we


see that there is no loss of generality in considering only the case when f is
the inversion centered at a boundary point. For simplicity of exposition, we
normalize so that the origin is this center.
Let ζ be a boundary point of D distinct from 0 and let u be the inward
pointing unit normal at ζ. For all sufficiently small t > 0, the point xt = ζ + tu
lies in D and its closest boundary point is ζ. For such s < t, we have
t
kD (xt , xs ) = log .
s

To estimate the distance of the image points, we use the inequality


 
′ ′ |x′ − y ′ |
jD′ (x , y ) = log 1 + ≤ kD′ (x′ , y ′ ),
min{δ ′ (x′ ), δ ′ (y ′ )}
which is always valid (since kD′ is the inner metric of jD′ , e.g. [12, Lemma 2.1]).
We also need the formula
|x − y|
|x′ − y ′ | =
|x| |y|
70 Peter Hästö IWQCMA05

for the length distortion of an inversion. Using these facts and the estimate
δ ′ (x′ ) ≤ |x′ − ζ ′ |, we derive the inequality
 |x′ − y ′ | 
kD′ (x′ , y ′ ) ≥ log 1 +
min{δ ′ (x′ ), δ ′ (y ′ )}
 |x − y|/(|x| |y|) 
≥ log 1 +
min{|x′ − ζ ′ |, |y ′ − ζ ′ |}
 |x − y| |ζ| 
= log 1 +
|x| |y| min{|x − ζ|/|x|, |y − ζ|/|y|}
 |x − y| |ζ| 
= log 1 + .
min{|y| |x − ζ|, |x| |y − ζ|}
Applying this inequality to the points xt and xs as defined before, we have
 (t − s) |ζ| 
kD′ (x′t , x′s ) ≥ log 1 + .
min{t |xs |, s |xt |}
Let us choose t = 2s. Since |x2s | and |xs | both tend to |ζ| as s → 0, we see that
the second term in the minimum is smaller. Since the inversion is supposed to be
an isometry, we can use the formula for kD (xt , xs ) from before with the previous
inequality to conclude that
2s  (2s − s) |ζ| 
log ≥ log 1 + .
s s |x2s |
Taking the exponential function gives |x2s | ≥ |ζ|. Since xs = ζ + su, this implies
that hζ − 0, ui ≥ 0 as s → 0, where h, i denotes the scalar product.
Applying the same argument, but starting with points on the image side, we
conclude that the opposite inequality is also valid. (There is actually a slight
asymmetry here: the domain D′ need not have circularly accessible boundary
at the origin. However, it is clear that this does not affect the argument so
far.) Thus it follows that hζ − 0, ui = 0 for all boundary points. But since the
boundary is assumed to be C 1 , this implies that the domain is a half-plane.

From [29, Theorem 2.8] we know that if f : D → R2 is a quasihyperbolic isom-


etry, then f is conformal in D. In dimensions three and higher every conformal
mapping is Möbius. It is easy to see that the proofs in this section work also in
the higher dimensional case. Therefore, we have proved the following result:
Corollary 4.3. Let D be a C 1 domains in Rn , n ≥ 3, which is not a half-space.
Then every quasihyperbolic isometry is a similarity mapping.
Example 4.4. Note that if we do not assume C 1 boundary, then there are some
further domains with non-trivial isometries: the punctured planes R2 \ {a} and
sector domains (i.e. domains whose boundary consists of two rays). In both
cases inversions centered at the distinguished boundary point (a or the vertex
of the sector). The previous proposition strongly suggests that these are all the
examples.
Isometries of relative metrics 71

4.2. Curvature of the quasihyperbolic metric. Let D be a domain in R2 .


We call a disc B ⊂ D maximal, if it is not contained in any other disc contained
in D. The set consisting of the centers of all maximal discs in D is called the
medial axis of D and denoted by MA(D). The medial axis and differentiability
properties of the distance-to-the-boundary function have been studied e.g. in
[7, 8, 9].
In a general domain the Gaussian curvature of the quasihyperbolic metric is
not defined, since the distance-to-the-boundary function is not C 2 . M. Heins
[20] considered this situation for a quite general class of metric, and defined the
notions of upper and lower curvature. Martin and Osgood worked with these
curvatures in the context of the quasihyperbolic metric, see [29, Section 3] for
details. However, if our domain is sufficiently regular (say C 2 ), and we are
considering points not on the medial axis, then the upper and lower curvature
agree, and define the curvature. In this case the curvature of kD is given by
KD (z) = −δ(z)2 △ log δ(z),
[20, (1.3)] or [29, (3.1)]. On the medial axis this formula does not make sense,
but the upper and lower curvatures still agree, and both equal −∞, by [29,
Corollary 3.12].
The next lemma is a specialization of Lemma 3.5, [29] to the case there the
upper and lower curvatures agree.
Lemma 4.5 (Lemma 3.5, [29]). Let G and G̃ be C 2 domains such that B(z, r) ⊂
G ∩ G̃ and ζ ∈ (∂G) ∩ (∂ G̃) ∩ (∂B(z, r)). If there is a neighborhood U of ζ such
that G ∩ U ⊂ G̃ ∩ U and d(z, ∂ G̃ \ U ) > d(z, ∂ G̃), then KG (z) ≤ KG̃ (z).

Using this lemma we can derive the following very plausible statement, which
says that the Gaussian curvature of the quasihyperbolic metric depends only on
the curvature of the boundary at the closest boundary point. We sill need some
more notation.
Let B be a disc with ζ ∈ (∂B) ∩ (∂D). Then we call B the osculating disc at
ζ if ∂B and ∂D have second order contact at ζ. Let D be at least a C 2 domain.
Then there exists an osculating disc at every boundary point ζ. If this disc has
radius r, then we define Rζ to be r if the disc lies in the direction of the interior of
D, and −r otherwise. Note that the function ζ 7→ 1/Rζ is C k−2 in a C k domain,
k ≥ 2.
Proposition 4.6. Let D ( R2 be a C 2 domain and z ∈ D \ MA(D) have closest
boundary point ζ ∈ ∂D. Then
Rζ 1
KD (z) = − =− .
Rζ − δ(z) 1 − δ(z)/Rζ
If z lies on the medial axis, then KD (z) = −∞.

Proof. The medial axis consists of points equidistant to two or more nearest
boundary points, and of centers of osculating circles. For the former, the claim
72 Peter Hästö IWQCMA05

that KD (z) = −∞ follows from [29, Corollary 3.12]. So we assume that z has a
unique nearest boundary point, ζ.
We suppose further that Rζ > 0, the other case begin similar. Let B(w, Rζ )
be the osculating disc at ζ. We define
w−ζ
Bt = B(w + Rζ
t, Rζ + t),
and note that ∂Bt contains ζ for all t > −Rζ . We have the formula
r r
KB(0,r) (x) = − =−
|x| r − d(x, ∂B(0, r))
for the curvature of the quasihyperbolic metric in a ball [29, Lemma 3.7], so we
can calculate KBt (z) explicitly.
Using the previous lemma with G = D and G̃ = Bt for t > 0 gives KD (z) ≤
KBt (z). If z is the center of B0 , then right-hand-side of this inequality tends
to −∞ as t → 0, which completes the proof of the claim regarding the medial
axis. So we assume that z is not the center of B0 , and then we can apply the
Lemma 4.5 with G = Bt for t < 0 (sufficiently close to 0) and G̃ = D to get
KBt (z) ≤ KD (z). Thus we have
KB−t (z) ≤ KD (z) ≤ KBt (z)
for small t > 0. Since KBt is continuous in t, we get KD (z) = KB0 (z) as we let
t → 0. The proof is completed by applying the aforementioned formula for the
curvature to the ball B0 = B(w, Rζ ).

Let f : D → R2 be a C 1 mapping. By ∇f we denote the gradient of f ,


˜ (z) we denote δ(z)∇f (z). The reason for
i.e. the vector (∂1 f, ∂2 f ), and by ∇f
multiplying by δ(x) is that
|x − y|
δ(y) = lim ,
x→y kD (x, y)

so that the ∇˜ operator is more natural in the setting where the quasihyperbolic
but not the Euclidean distance is preserved (see (4.9), below).
˜ D . For this need a mapping which
We next present an explicit formula for ∇K
associates to every point in D \ MA(D) its closest boundary point. We call this
mapping ζ = ζ(z).
Lemma 4.7. Let D ( R2 be a C 3 domain. Then
 
˜ D (z) = (KD (z) + 1) KD (z)∇δ(z) − (KD (z) + 1)∇Rζ(z)
∇K
for every z off the medial axis, where all differentiation is with respect to the
variable z.

Proof. We use the formula from Proposition 4.6. Thus


1 δ(z) KD (z)2
∇KD (z) = −∇ = KD (z)2 ∇ = (Rζ ∇δ(z) − δ(z)∇Rζ ),
1 − δ(z)/Rζ Rζ Rζ2
Isometries of relative metrics 73

where we understand ζ as a function of z. Note that Rζ and δ are C 1 , since D


is C 3 and we are not on the medial axis. From Proposition 4.6 we also get
δ(z) KD (z) + 1
= .
Rζ KD (z)
Thus we continue the equation by
 
˜ D (z) = KD (z)2 δ(z) ∇δ(z) − δ(z) ∇Rζ
∇K
Rζ Rζ

= (KD (z) + 1) KD (z)∇δ(z) − (KD (z) + 1)∇Rζ .

˜ is an intrinsic quantity of the quasihyperbolic metric.


We next show that |∇K|
Lemma 4.8. Let D be a C 3 domain. If f : D → R2 is a quasihyperbolic isometry,
˜ D (z)| = |∇K
then |∇K ˜ f (D) (f (z))| for every z ∈ D.

Proof. We know that f is conformal. For a unit vector u we find that




˜ D (z), u = lim KD (z + εu) − KD (z)
∇K
ε→0 kD (z + εu, z)
(4.9)
Kf (D) (f (z + εu)) − Kf (D) (f (z))
= lim .
ε→0 kf (D) (f (z + εu), f (z))
Next we note that f (z + εu) = f (z) + εf ′ (z)u + O(ε2 ). Here f ′ (z)u is understood
as complex multiplication. Let us define another unit vector ũ = |ff ′ (z)|
′ (z)
u. Then
we continue the previous equation by

˜
Kf (D) (f (z) + εf ′ (z)u) − Kf (D) (f (z))
∇KD (z), u = lim
ε→0 kf (D) (f (z) + εf ′ (z)u, f (z))
ε|f ′ (z)|h∇Kf (D) (f (z)), ũi
= lim
ε→0 ε|f ′ (z)|δ ′ (f (z))−1


= ∇K˜ f (D) (f (z)), ũ .
˜ D (z)| = |∇K
Since u was an arbitrary unit vector, we see that |∇K ˜ f (D) (f (z))|.

4.3. Isometries of the quasihyperbolic metric. We know that similarities


are always quasihyperbolic isometries, and we want to show that in most cases
these are the only ones. In view of the results in Section 4.1, it suffices for us to
show that a quasihyperbolic isometry is a Möbius mapping, so this will be what
we aim at in the proofs of this section.
A curve γ in D is a (quasihyperbolic) geodesic if
kD (x, y) = kD (x, z) + kD (z, y)
for all x, z, y ∈ γ in this order. It is clear from this definition that geodesics are
preserved by isometries. A geodesic ray is a geodesic which is isometric to R+ .
For every z ∈ D we easily find one geodesic ray, namely [z, ζ(z)), which also
happens to be a Euclidean line segment. The idea is to show that this geodesic
74 Peter Hästö IWQCMA05

is somehow special (from a quasihyperbolic point-of-view), so that it would map


to a geodesic ray of the same kind.
Lemma 4.10. Let D ( R2 be a C 2 domain with a boundary point ξ such that
1/Rξ = 0. Then every isometry f : D → R2 of the quasihyperbolic metric is
Möbius.
Proof. Let B ⊂ D be a non-maximal disc whose boundary contains ξ and let
z denote the center of B. By Proposition 4.6 we find that KD ≡ −1 on the
segment γ = [z, ξ). Thus Kf (D) ≡ −1 on γ ′ , so 1/Rζ′ ′ (z′ ) = 0 for every point z ′
on this curve. We consider two cases: either ζ ′ (z ′ ) is just a single point for all
z ′ ∈ γ ′ , or it sweeps out a non-degenerate subcurve of the boundary ∂D′ as z ′
varies over γ ′ . (There is no third possibility, since ζ ′ is a continuous function on
γ ′ .) In the single-point case we see that γ ′ has to be a line segment, since the
boundary does not have corners. In this case we find that

kD (x, y) = log |x−ξ| and kD′ (x′ , y ′ ) = log |x′ −ξ′ | ,

|y−ξ| |y −ξ |

where ξ is the closest boundary point to the every point on γ ′ . But this easily

implies that f is Möbius on γ. Since f is conformal it follows by uniqueness of


analytic extension that f is a Möbius mapping on all of D.
So we consider the second case, that ζ ′ (z ′ ) sweeps out a non-degenerate sub-
curve of the boundary ∂D′ . Since the curvature of the boundary at all these
points is zero, it follows that the piece of the boundary is a line segment, L′ .
Let U ′ ⊂ D′ be an open set such that (∂U ′ ) ∩ (∂D′ ) = L′ and the nearest
boundary point of every point in U ′ lies in L′ . Then the geometry of the quasi-
hyperbolic metric in U is the same as in a half-plane, in particular KD′ ≡ −1
on U ′ . Then KD ≡ −1 on U = f −1 (U ′ ), so it follows that (∂U ) ∩ (∂D) = L, for
some line segment L. So it follows that f |U is the restriction of a quasihyper-
bolic isometry of the half-plane. But these are only the Möbius mappings. Then
we again conclude from the uniqueness of analytic extension that f is a Möbius
mapping on all of D.
Let us call a domain strictly concave, if its complement is strictly convex.
Corollary 4.11. Let D ( R2 be a C 2 domain which is not a half-plane, strictly
convex or strictly concave. Then every quasihyperbolic isometry is a similarity
mapping.
Proof. Suppose that 1/Rζ 6= 0 for all boundary points. Since 1/Rζ is continuous
by assumption, this implies that it is either everywhere positive, or everywhere
negative. In these cases we have a strictly convex and strictly concave domain,
respectively, which was ruled out by assumption. So we find some point at which
1/Rζ = 0. Then it follows from Lemma 4.10 that the isometry is Möbius and
from Lemma 4.2 that it is a similarity.
So we are left with only two types of domains that we cannot handle: strictly
convex and strictly concave ones. As usual when working with isometries, the
Isometries of relative metrics 75

nicest domains turn out to be the most difficult. Unfortunately, we need to


assume more regularity of the boundary in order to take care of these cases.
Theorem 4.12. Let D ( R2 be a C 3 domain, which is not a half-plane. Then
every isometry f : D → R2 of the quasihyperbolic metric is a similarity mapping.

Proof. In view of Corollary 4.11, we may restrict ourselves to the case when
KD (z) 6= −1 for all z ∈ D. Let z ∈ D \ MA(D) and ζ be its nearest boundary
point. We note that ∇δ(z) and ∇Rζ are perpendicular – first of all, ∇δ(z) is
parallel to z − ζ; second, Rζ is a constant in the direction of z − ζ, since ζ is the
closest boundary point to all points on this line (near z).
If D is bounded, then it is clear that Rζ has a critical point. If D is unbounded,
then we note that 1/Rζ cannot have any other limit than 0 at ∞ (although a
limit need not exist, of course). Thus we see that Rζ has a critical point in the
unbounded case as well. Let ζ be a critical point of ξ 7→ Rξ and fix a point z ∈ D
with KD (z) 6= −∞ whose nearest boundary point is ζ. Of course, ∇Rζ = 0 at
the critical point ζ. Then it follows from Lemma 4.7 that
˜ D (z) = (KD (z) + 1)KD (z)∇δ(z).
∇K
Since the curvature is intrinsic to the metric, we have KD′ (z ′ ) = KD (z). Also,
˜ D′ (z ′ )| = |∇K
|∇K ˜ D (z)| by Lemma 4.8, so we have
 
(KD (z) + 1)KD (z)∇δ(z) = (KD (z) + 1) KD (z)∇δ ′ (z ′ ) − (KD (z) + 1)∇Rζ′ ′ (z′ ) .
We know that KD (z) 6= −1 and that ∇δ ′ (z ′ ) and ∇Rζ′ ′ (z′ ) are orthogonal. Thus
the previous equation simplifies to
2 2 2
KD (z)|∇δ(z)| = KD (z)|∇δ ′ (z ′ )| + (KD (z) + 1) ∇Rζ′ ′ (z′ ) .
Since |∇δ| = 1 off the medial axis for every domain, this equation implies that
∇Rζ ′ = 0.
So for our point z, ∇KD (z) and ∇KD′ (z ′ ) point to the nearest boundary
point of z and z ′ , respectively. Let γ = [z, ζ). Note that γ is a geodesic of the
quasihyperbolic metric. Also, ∇KD (z) and γ are parallel at z. Now γ is mapped
to some geodesic ray γ ′ , and since f is a conformal mapping, γ ′ is parallel to
∇KD′ (z ′ ) at z ′ . But [z ′ , ζ ′ ) is a geodesic parallel to ∇KD′ (z ′ ) at z ′ , and since
geodesics are unique (when the density is C 2 , i.e. except possibly on the medial
axis) we see that γ ′ = [z ′ , ζ ′ ).
So we have shown that f ([z, ζ)) = [z ′ , ζ ′ ). Moreover, we have

|x−ζ| ′ ′ |x′ −ζ ′ |
kD (x, y) = log |y−ζ| and kD′ (x , y ) = log |y′ −ζ ′ |
for x, y ∈ [z, ζ). Thus we see that f is just a similarity on [z, ζ). But f is a
conformal map, so this implies that f is a similarity in all of D.
Acknowledgment. I would like to thank Zair Ibragmov for several discussions
about the isometries of this and related metrics and Swadesh Sahoo for some
comments on this manuscript.
76 Peter Hästö IWQCMA05

References
[1] Z. Balogh and S. Buckley: Geometric characterizations of Gromov hyperbolicity, Invent.
Math. 153 (2003), no. 2, 261–301.
[2] A. Beardon: Geometry of Discrete Groups, Springer-Verlag, New York, 1983; corrected
reprint, 1995.
[3] : The Apollonian metric of a domain in Rn , pp. 91–108 in Quasiconformal mappings
and analysis (P. Duren, J. Heinonen, B. Osgood and B. Palka (eds.)), Springer-Verlag,
New York, 1998.
[4] A. Beardon and D. Minda: Sphere-preserving maps in inversive geometry, Proc. Amer.
Math. Soc. 130 (2002), 987–998.
[5] L. Blumenthal: Distance Geometry. A study of the development of abstract metrics. With
an introduction by Karl Menger, Univ. of Missouri Studies Vol. 13, No. 2, Univ. of Missouri,
Columbia, 1938.
[6] D. Burago, Yu. Burago and S. Ivanov: A course in metric geometry, Graduate Studies in
Mathematics, 33, Amer. Math. Soc., Providence, RI, 2001.
[7] L. Cafarelli and A. Friedman: The free boundary for elastic-plastic torsion problems,
Trans. Amer. Math. Soc. 252 (1979), 65–97.
[8] H. I. Choi, S. W. Choi and H. P. Moon: Mathematical theory of medial axis transform,
Pacific J. Math. 181 (1997), no. 1, 57–88.
[9] J. Damon: Smoothness and geometry of boundaries associated to skeletal structures.
I. Sufficient conditions for smoothness, Ann. Inst. Fourier (Grenoble) 53 (2003), no. 6,
1941–1985.
[10] J. Ferrand: A characterization of quasiconformal mappings by the behavior of a function of
three points, pp. 110–123 in Proceedings of the 13th Rolf Nevalinna Colloquium (Joensuu,
1987; I. Laine, S. Rickman and T. Sorvali (eds.)), Lecture Notes in Mathematics Vol. 1351,
Springer-Verlag, New York, 1988.
[11] F. Gehring and B. Osgood: Uniform domains and the quasihyperbolic metric, J. Anal.
Math. 36 (1979), 50–74.
[12] F. Gehring and B. Palka: Quasiconformally homogeneous domains, J. Anal. Math. 30
(1976), 172–199.
[13] P. Hästö: A new weighted metric: the relative metric II, J. Math. Anal. Appl. 301 (2005),
no. 2, 336–353.
[14] : Gromov hyperbolicity of the jG and ̃G metrics, Proc. Amer. Math. Soc. 134 (2006),
1137–1142.
[15] : Isometries of the quasihyperbolic metric, submitted.
[16] P. Hästö and Z. Ibragimov: Apollonian isometries of planar domains are Möbius mappings,
J. Geom. Anal. 15 (2005), no. 2, 229–237.
[17] : Apollonian isometries of regular domains are Möbius mappings, Ann. Acad. Sci.
Fenn. Math., to appear.
[18] P. Hästö, Z. Ibragimov and H. Lindén: Isometries of relative metrics, Comput. Methods
Funct. Theory 6 (2006), no. 1, 15–28.
[19] P. Hästö and H. Lindén: Isometries of the half-apollonian metric, Complex Var. Theory
Appl. 49 (2004), 405–415.
[20] M. Heins: On a class of conformal metrics, Nagoya Math. J. 21 (1962), 1–60.
[21] David Herron, Zair Ibragimov and David Minda: Geometry of the K-P metric, preprint
(2005).
[22] D. Herron and P. Koskela: Conformal capacity and the quasihyperbolic metric, Indiana
Univ. Math. J. 45 (1996), no. 2, 333–359.
[23] D. Herron, W. Ma and D. Minda: A Möbius invariant metric for regions on the Riemann
sphere, pp. 101–118 in Future Trends in Geometric Function Theory (RNC Workshop,
Jyväskylä 2003; D. Herron (ed.)), Rep. Univ. Jyväskylä Dept. Math. Stat. 92 (2003).
Isometries of relative metrics 77

[24] P. Jones and S. Smirnov: Removability theorems for Sobolev functions and QC maps,
Ark. Mat. 38 (2000), no. 2, 263–279.
[25] P. Koskela and T. Nieminen: Quasiconformal removability and the quasihyperbolic metric,
Indiana Univ. Math. J. 54 (2005), no. 1, 143–151.
[26] R. Kulkarni and U. Pinkall: A canonical metric for Möbius structures and its applications,
Math. Z. 216 (1994), 89–129.
[27] H. Lindén: Quasihyperbolic Geodesics and Uniformity in Elementary Domains, Ph.D.
Thesis, University of Helsinki, 2005.
[28] G. Martin: Quasiconformal and bilipschitz mappings, uniform domains and the quasihy-
perbolic metric, Trans. Amer. Math. Soc. 292 (1985), 169–191.
[29] G. Martin and B. Osgood: The quasihyperbolic metric and associated estimates on the
hyperbolic metric, J. Anal. Math. 47 (1986), 37–53.
[30] P. Seittenranta: Möbius-invariant metrics, Math. Proc. Cambridge Philos. Soc. 125
(1999), 511–533.
[31] M. Vuorinen: Conformal Geometry and Quasiregular Mappings, Lecture Notes in Math.,
Vol. 1319, Springer, Berlin, 1988.

Peter Hästö E-mail: peter.hasto@helsinki.fi


Address: Department of Mathematical Sciences, P.O. Box 3000, 90014 University of Oulu,
Finland
Proceedings of the International Workshop on Quasiconformal
Mappings and their Applications (IWQCMA05)

Uniform Spaces and Gromov Hyperbolicity

David A Herron

Abstract. This brief outline contains mostly definitions, background infor-


mation, and statements of theorems. Along with the title topics, we also
discuss the uniformization volume growth problem as well as certain capacity
and slice condition characterizations of uniformity.

Keywords. uniform spaces, Gromov hyperbolicity, quasihyperbolic metric,


volume growth, Ahlfors regular spaces, Loewner spaces, slice conditions.
2000 MSC. Primary: 30C65; Secondary: 53C23, 30F45.

Contents

1. Introduction 80
2. Metric Space Background 84
2.A. General Information 85
2.B. Abstract Domains 86
2.C. Maps and Gauges 87
2.D. Length and Geodesics 88
2.E. Connectivity Conditions 89
2.F. Doubling and Dimensions 90
2.G. Quasiconformal Deformations 92
2.H. Quasihyperbolic Distance and Geodesics 96
2.I. Modulus and Capacity 97
2.J. Ahlfors Regular and Loewner Spaces 99
2.K. Slice Conditions 101
3. Uniform Spaces 102
3.A. Euclidean Setting 102
3.B. Measure Metric Space Setting 102
3.C. Basic Information 103
4. Gromov Hyperbolicity 103
4.A. Thin Triangles Definition 104
Version October 19, 2006.
The author was supported by the NSF and the Charles Phelps Taft Research Center.
80 D.A.Herron IWQCMA05

4.B. Gromov Boundary 104


4.C. Connection with Uniform Spaces 105
5. Uniformization 105
5.A. Uniformization Problem 105
5.B. BHK Uniformization 105
5.C. Bounded Geometry and its Consequences 107
5.D. Lifts and Metric Doubling Measures 109
5.E. Volume Growth Problem 110
6. Characterizations of Uniform Spaces 111
6.A. Metric Characterizations 111
6.B. Gromov Boundary Characterizations 111
6.C. Characterizations using QC Maps 112
6.D. Capacity Conditions 112
6.E. LLC and Slice Conditions 113
References 113

1. Introduction
This survey is meant to supplement the talks I presented at the International
Workshop on Quasiconformal Mappings and their Applications and at the Interna-
tional Conference on Geometric Function Theory, Special Functions and their Appli-
cations. Primarily, I provide here basic background material including definitions,
terminology, and fundamental facts. I also list a few references, many of which
themselves contain additional references to this material. I have made no at-
tempt to render a complete list of references and apologize to all those whose
work I have neglected to mention. The reader is absolutely encouraged to consult
the many works referred to by the authors I do mention.
The goal of these notes is to provide the reader with a foundation enabling
them to understand the meaning and relevance of the recent work [BHK01],
[BHR01], [BKR98] of Bonk, Heinonen, Koskela and Rohde along with [Her04]
and [Her06]. I am delighted to thank Mario Bonk, Juha Heinonen and Pekka
Koskela for numerous helpful discussions and hours of blackboard sessions re-
garding these topics.
By now Euclidean uniform spaces (domains in Euclidean space in which points
can be joined by short twisted double cone arcs) are well recognized as being the
‘nice’ spaces for quasiconformal function theory as well as many other areas of
analysis (e.g., potential theory); see [Geh87], [Väi88] for Euclidean space and
[Gre01], [CGN00], [CT95] for the Carnot-Carathéodory setting. In [BHK01]
Bonk, Heinonen, and Koskela develop a uniformization theory which provides a
Uniformity and Hyperbolicity 81

two way correspondence between uniform spaces and Gromov hyperbolic spaces.
In particular, they prove the following fundamental result; see [BHK01, Theorem
1.1].
There is a one-to-one (conformal) correspondence between quasiisom-
etry classes of proper geodesic roughly starlike Gromov hyperbolic spaces
and quasisimilarity classes of bounded locally compact uniform spaces.
A simple, yet beautiful, example is the open unit disk in the plane. In terms of
its Euclidean geometry, each pair of points can be joined by a twisted double cone
which stays away from the boundary and is not much bigger than the distance
between the given points (it is a uniform space). On the other hand, the disk
also admits a non-euclidean geometry, in terms of its Poincaré hyperbolic metric,
and as such the disk is a Gromov hyperbolic space.
The Bonk, Heinonen, Koskela theory asserts that this phenomenon holds in a
very general setting. The complete proof of their result is presented in Chapters
2-5 of [BHK01] and beyond the scope of our discussion. However, there are two
basic results involved which are central to my workshop lectures: Fact 4.1 says
that every locally compact uniform space has a Gromov hyperbolic quasihyper-
bolization; Fact 5.1 says that every (proper geodesic) Gromov hyperbolic space
can be uniformized. I will describe what uniform spaces are, what their con-
nection is with Gromov hyperbolicity, and explain some of the ideas behind the
proofs. Time permitting, I will also look at the related question of when there
exists a uniformization with the property that the associated measure (see (2.7))
has regular volume growth. My conference lecture will focus on §6.D and §6.E.

For the remainder of this introduction, I advertise results from [Her04] and
[Her06] hoping to wet the reader’s appetite for this flavor of metric measure space
geometric function theory. See §2-§5 for precise definitions.
In [BHR01] and [BKR98] the authors investigate conformal deformations of
the unit ball in Euclidean space. The primary object of study in these notes
is the geometry of quasiconformal deformations of an abstract metric measure
space (Ω, d, µ). Following BHKR, we consider a metric-density ρ on Ω and Ωρ =
(Ω, dρ , µρ ) denotes the deformed space (see subsection 2.G). We are interested
in the situation when this new space Ωρ is uniform (see Section 3) and describe
this by calling such a ρ a uniformizing density. Every proper geodesic Gromov
hyperbolic space can be uniformized, and, there is a natural canonical proper
geodesic space associated with any locally compact abstract domain, namely, its
quasihyperbolization; see Facts 4.1 and 5.1. However, in general the associated
measure (see (2.7)) may fail to have Ahlfors regular volume growth. For example,
applying the BHK uniformization to the quasihyperbolized Euclidean unit ball
we obtain a new metric measure space which has exponential volume growth.
The theory developed in [BHK01] is exploited in [Her06] to extend some results
of [BHR01] to the setting of abstract metric measure spaces (Ω, d, µ). More
importantly, we establish the result given below which provides an answer to the
82 D.A.Herron IWQCMA05

question: When does an abstract domain admit a quasiconformal deformation


which is both uniformizing and such that the induced measure (2.7) satisfies the
natural volume growth estimate? That is, when is there a conformal uniformizing
density? In particular, the induced measure should be Ahlfors regular. Under
certain reasonable minimal hypotheses, this occurs precisely when the conformal
Assouad dimension of the space’s Gromov boundary is small enough. See §5.E
for a discussion of the proof of the following.
Theorem A. Let Ω be an abstract domain with bounded Q-geometry. Suppose Ω
admits a bounded uniformizing conformal density. Then Ω has a Gromov hyper-
bolic roughly starlike quasihyperbolization and the conformal Assouad dimension
of its Gromov boundary is strictly less than Q. The converse holds too, provided
we assume that the Gromov boundary of Ω is uniformly perfect.

The above result is quantitative: the asserted constants depend only on the data
associated with Ω and the density.
In what follows we consider metric measure spaces (Ω, d, µ) which satisfy the
following basic minimal hypotheses:
Ω is an abstract domain having bounded Q-geometry and a
Gromov hyperbolic roughly starlike quasihyperbolization.
Precise definitions are stated in subsections 2.B, 2.D, 2.H, 5.C; roughly, these
hypotheses ensure that Ω has ‘enough’ of the local properties enjoyed by domains
in Euclidean space. The data associated with these basic hypotheses consists of
six parameters: Q (the ‘dimension’), M , m, λ (the bounded geometry constants),
δ (the Gromov hyperbolicity constant) and κ (the rough starlike constant).
There are a number of auxiliary results (namely, Theorems B-F) needed for
the proof of Theorem A; all of these can be found in [Her04] or [Her06]. First
we have the so-called Gehring-Hayman Inequality (cf. [GH62]); it is an essential
tool for most of what follows. This was proved in [BKR98, Theorem 3.1] for
deformations of the Euclidean unit ball and in [HR93] for quasiconformal images
of uniform domains in Euclidean space; see also [BB03, Theorem 2.3], [BHK01,
Chpt. 5] and [HN94]. Our proof of the following (see [Her04, Theorem A]) utilizes
ideas from both [BKR98, Theorem 3.1] and [HR93, Theorem 1.1].
Theorem B. Let ρ be an Ahlfors Harnack density on a uniform Loewner met-
ric measure space (Ω, d, µ). Then there exists a constant Λ such that for all
quasihyperbolic geodesics [x, y]k with endpoints in Ω̄,
ℓρ ([x, y]k ) ≤ Λ dρ (x, y).

This result is quantitative: Λ depends only on the data associated with Ω.


Throughout this article the symbol Λ will stand for this Gehring-Hayman In-
equality constant.
Here is a simple, but useful, consequence of the Gehring-Hayman Inequality:
if there is an arc α joining some point w in Ω to some point ζ in ∂Ω with
Uniformity and Hyperbolicity 83

ℓρ (α) < ∞, then ℓρ (γ) < ∞ for every quasihyperbolic geodesic ray going to ζ. In
fact, there is even a ‘radial limit theorem’ [Her04, Theorem B] which says that
this is true for modQ -a.e. point of ∂Ω.
Next we communicate the primary tool employed in our proof of Theorem A.
It is based on a lifting procedure discussed in [BKR98, 2.7] and established for
the Euclidean unit ball as [BHR01, Proposition 1.25]. See (5.9) and (2.5) for the
definitions of ρν (the lift of ν) and δν,1/P (the quasimetric determined by ν). See
§5.D for a discussion of the proof of the following.
Theorem C. Assume the basic minimal hypotheses, that the Gromov boundary
of Ω is uniformly perfect, and that P < Q. Suppose ν is a P -dimensional metric
doubling measure on ∂G Ω. Then the lift ρ = ρν of ν is a doubling conformal
density on Ω and the natural map (∂ρ Ω, dρ ) → (∂G Ω, δν,1/P ) is bilipschitz.

The Bonk-Heinonen-Koskela uniformization theory is a crucial tool employed


in all our arguments and permits us to replace the space Ω with a bounded
uniform space Ωε where the geometry is more transparent; see Fact 5.1. A key
ingredient in our proof of Theorem A is the following generalization of [BHR01,
Proposition 2.11]. In particular, it asserts that a conformal density on a bounded
uniform space is uniformizing if and only if the associated measure (2.7) is a
doubling measure on the original space. (See §5.E for the precise definition of a
doubling conformal density.)
Theorem D. Assume the basic minimal hypotheses. Let Ωε be any BHK-
uniformization of Ω. Suppose ρ is a conformal density on Ω. Then the following
are quantitatively equivalent:
(a) ρ is doubling on Ω.
(b) Ωρ is bounded and uniform.
(c) Ωρ is bounded and Q-Loewner.
(d) Ωρ is bounded, Q-Loewner and Ahlfors Q-regular.
(e) the identity map Ωρ → Ωε is quasisymmetric.

Again, this result is quantitative: the asserted constants depend only on the data
associated with Ω and ρ, and the related data. Also, we point out that the proof
of (b) shows that the quasihyperbolic geodesics in Ω will be uniform arcs in Ωρ .
A crucial component of the proof of Theorem C is the following result which
permits us to estimate dρ (x) = distρ (x, ∂ρ Ω) in terms of ρ(x)d(x). More precisely,
it tells us that Ahlfors Harnack metric-densities are Koebe under the right condi-
tions. The lower bound is immediate via the Harnack inequality. To obtain any
upper bound, we at least need ∂ρ Ω 6= ∅. In fact, we require a condition which
ensures that Ω has a uniformly thick boundary as seen from each point. With
this in mind, we introduce the following notion: we say that (Ω, d, µ) satisfies a
Whitney ball modulus property if there exists a constant m > 0 such that
modQ (B̄(x; λ d(x)), ∂Ω; Ω) ≥ m for all x ∈ Ω.
84 D.A.Herron IWQCMA05

Theorem E. Let ρ be a Ahlfors Harnack metric-density on a uniform Loewner


abstract domain (Ω, d, µ). Suppose that Ω enjoys a Whitney ball modulus prop-
erty. Then there is a constant K such that for all x ∈ Ω,
K −1 ρ(x)d(x) ≤ dρ (x) ≤ Kρ(x)d(x);
the constant K depends only on the data associated with Ω.

An important consequence of Theorem E is that the quasihyperbolizations


of Ω and Ωρ are bilipschitz equivalent, and it follows that (Ωρ , kρ ) is a Gromov
hyperbolic space.
We mention that any uniform Loewner space with connected boundary satis-
fies a Whitney ball modulus property, provided it and its boundary are simulta-
neously bounded or unbounded. Similarly any bounded uniform Loewner space
with a finite number of non-degenerate boundary components will enjoy this
modulus property. Here is a sufficient condition for this property to hold which
allows for a totally disconnected boundary.
Theorem F. Let (Ω, d, µ) be a locally Loewner, uniform metric measure space.
Assume Ω and ∂Ω are either both bounded or both unbounded. Suppose that for
some p > 0, ∂Ω satisfies the Hausdorff p-content condition
p
H∞ (∂Ω ∩ B̄(ζ; r)) ≥ c rp for all 0 < r ≤ diam(∂Ω) and all ζ ∈ ∂Ω.
Then Ω enjoys a Whitney ball modulus property with a constant m which depends
only on c and the data associated with Ω.

In contrast to the Euclidean case, the converse to the above is false; see
[Her04, Example 3.2] which furnishes a space with an isolated boundary point
which nonetheless satisfies a Whitney ball modulus property.

Our notation is relatively standard and, for the most part, conforms with that
of [BHK01]. We write C = C(a, . . .) to indicate a constant C which depends only
on the parameters a, . . .; the notation A . B means there exists a finite constant
c with A ≤ cB, and A ≃ B means that both A . B and B . A hold. Typically
a, b, c, C, K, . . . will be constants that depend on various parameters, and we try
to make this as clear as possible often giving explicit values, however, at times
C will denote some constant whose value depends only on the data present but
may differ even on the same line of inequalities.

2. Metric Space Background


Naturally there are scores of references for metric space geometry. Here is a
brief list of some texts which I have found especially helpful: [BH99], [BBI01],
[Hei01], [Sem01], [Sem99], [DS97], and of course the references mentioned in these
works.
Uniformity and Hyperbolicity 85

2.A. General Information. In what follows (X, d) will always denote a generic
metric space possessing no additional presumed properties. For the record, this
means that d is a distance function; that is, d : X × X → R is positive semi-
definite, symmetric, and satisfies the triangle inequality. We often write the
distance between x and y as d(x, y) = |x − y|. The open ball (sphere) of radius r
centered at the point x is B(x; r) := {y : |x−y| < r} (S(x; r) := {y : |x−y| = r}).
When B = B(x; r) and λ > 0, λB := B(x; λ r). We say that X is a proper met-
ric space if it has the Heine-Borel property that every closed ball is compact (or
equivalently, the compact sets are exactly the closed and bounded sets).
In general, we work in the setting of a metric measure space (X, d, µ) with
X a non-complete locally complete (often locally compact) rectifiably connected
metric space and µ a Borel regular measure satisfying µ[B(x; r)] > 0 for each
ball.
Recall that every metric space can be isometrically embedded into a complete
metric space. We let X̄ denote the metric completion of a metric space X and
we call ∂X = X̄ \ X the metric boundary of X. Then d(x) = dist(x, ∂X) is the
distance from a point x ∈ X to the boundary ∂X of X; note that when ∂X is
closed in X̄, we have d(x) > 0 for all x ∈ X. For example, this holds when X
is locally compact. Of course, if X is complete to begin with, then ∂X = ∅ and
d(x) = ∞ for all x ∈ X. We call X locally complete provided d(x) > 0 for all
x ∈ X.
In a locally complete metric space we make extensive use of the notation
B(x) := B(x; d(x)).
In this setting, we call λB(x) = B(x; λ d(x)) a Whitney ball in X with associated
Whitney ball constant λ ∈ (0, 1).
It is convenient, at times, to consider quasimetric spaces (X, q). We call q a
quasimetric on X if q : X × X → R is symmetric and positive definite but only
satisfies
q(x, y) ≤ K (q(x, z) + q(y, z)) for all x, y, z ∈ X
in place of the triangle inequality. See [Hei01, 14.1], [Sem01] and [DS97].
Starting with a quasimetric q, there is a standard way to define a pseudometric
d with d ≤ q (cf. [BH99, 1.24, p.14]), but it may happen that d(x, y) = 0 for
some x 6= y. However, by first ‘snowflaking’ q and then applying this procedure
we can arrive at an honest distance function; see [Hei01, Proposition 14.5] or
[BH99, Proposition 3.21, p.435].
2.1. Fact. Let q be a quasimetric on X. There is an ε0 > 0 depending only on
the quasimetric constant K for q such that for all ε ∈ (0, ε0 ), the quasimetric
qε (x, y) = q(x, y)ε is bilipschitz equivalent to an honest distance function d on
X; in fact there is a constant L = L(ε, K) such that
L−1 qε (x, y) ≤ d(x, y) ≤ qε (x, y) for all x, y ∈ X.
86 D.A.Herron IWQCMA05

We remark that all the quasimetrics qε as defined above are QS equivalent to


each other.
Another useful notion, apparently introduced by Väisälä, is that of a meta-
metric m : X × X → R which is symmetric, non-negative, satisfies the triangle
inequality, but only
m(x, y) = 0 =⇒ x = y
and so possibly m(x, x) > 0. See [Väi05a, 4.2] for a treatment of metametric
spaces.
A metric space X is called uniformly perfect provided it has at least two
points and there is a constant ϑ ∈ (0, 1) such that for all balls B ⊂ X, B \
ϑB 6= ∅ provided X \ B 6= ∅. This concept, which involves three points, is
especially useful when dealing with quasisymmetric maps and also with doubling
measures (see §2.F). The property of being uniformly perfect is preserved by
quasisymmetric homeomorphisms, with the new constant depending only on the
original constant and the quasisymmetry data; in particular, one can ask whether
or not a conformal gauge is uniformly perfect (see §2.C).
It is a routine exercise to see that uniformly perfect locally compact spaces
contain quasisymmetrically embedded middle-third Cantor dusts. Using this
fact, together with a scaling argument and properties of quasisymmetric homeo-
morphisms (e.g. [Hei01, 11.10,11.11]), one can verify a version of the following.
For a simple more direct approach, which also provides the indicated explicit
constants, see [Her06, Lemma 4.2].
2.2. Fact. Suppose X is a uniformly perfect compact metric space. Then X
satisfies the p-dimensional Hausdorff measure density condition
rp
Hp [B(x; r)] ≥ for all 0 < r ≤ diam(X) and all x ∈ X,
6
where p = 1/ log2 (4/ϑ) and ϑ is the uniform perfectedness constant.

The above result can be used in conjunction with Theorem F to see that the
Whitney ball modulus property holds.

2.B. Abstract Domains. We call a metric measure space (Ω, d, µ) an abstract


domain if Ω is a non-complete locally complete rectifiably connected metric space
(and µ a Borel regular measure with dense support). An important example of
such a space is, of course, a proper subdomain of Euclidean space with either
Euclidean distance or the induced Euclidean length distance.
Unless explicitly indicated otherwise, the adjective locally means that the
modified property or condition holds in all Whitney-type balls λB(x) where
0 < λ < 1 is some fixed constant which we call the Whitney ball constant;
when there are several such local conditions in play, we always take λ to be the
minimum of all the associated Whitney ball constants.
Uniformity and Hyperbolicity 87

2.C. Maps and Gauges. An embedding f : X → Y from a metric space X


into a metric space Y is quasisymmetric, abbreviated QS, if there is a homeomor-
phism η : [0, ∞) → [0, ∞) (called a distortion function) such that for all triples
x, y, z ∈ X,
|x − y| ≤ t|x − z| =⇒ |f x − f y| ≤ η(t)|f x − f z|.
These mappings were studied by Tukia and Väisälä in [TV80]; see also [Hei01].
The bilipschitz maps form an important subclass of the quasisymmetric maps;
f : X → Y is bilipschitz if there is a constant L such that for all x, y ∈ X,
L−1 |x − y| ≤ |f x − f y| ≤ L|x − y|.
More generally, a map f : X → Y is an (L, C)-quasiisometry if L ≥ 1, C ≥ 0
and for all x, y ∈ X,
L−1 |x − y| − C ≤ |f x − f y| ≤ L|x − y| + C.
There seems to be no universal agreement regarding this terminology; some au-
thors use the adjective quasiisometry to mean what we have called bilipschitz,
and then a rough quasiisometry satisfies our definition of quasiisometry. So the
reader should beware! Of course a (1, 0)-quasiisometry is simply called an isom-
etry (onto its range).
Note that the above definitions also make sense for mappings of quasimetric
spaces.
Given a metric (or a quasimetric) on X, we can form the conformal gauge
G on X consisting of all metrics on X which are QS equivalent to the original
(quasi)metric. That is, G is the family of all metrics ∂ on X such that the identity
map (X, d) → (X, ∂) is QS. See [Hei01, Chapter 15] for more discussion of this
topic.
An embedding f : X → Y from a metric space X into a metric space Y is
called quasimöbius, abbreviated QM, if there is a homeomorphism ϑ : [0, ∞) →
[0, ∞) (called a distortion function) such that for all quadruples x, y, z, w of
distinct points in X,
|x, y, z, w| ≤ t =⇒ |f x, f y, f z, f w| ≤ ϑ(t)
where the absolute cross ratio is
|x − y||z − w|
|x, y, z, w| = .
|x − z||y − w|
These mappings were introduced and investigated by Väisälä in [Väi85]; see
also [Väi05a]. Every QS homeomorphism is QM; the converse holds in certain
special cases. Clearly Möbius transformations are QM maps in Euclidean space;
however, a Möbius transformation from the unit ball onto a half-space is not QS.
The QM maps are more flexible than the QS.
The QS and QM maps are defined by global conditions whereas QC (quasicon-
formal) maps only satisfy a local condition. I highly recommend Tyson’s recent
survey article [Tys03]. Väisälä’s notes [Väi71] are the classical reference for QC
maps in the Euclidean setting. These maps have been studied in the Heisenberg
88 D.A.Herron IWQCMA05

group setting and there is still much research underway there. Heinonen and
Koskela strongly advanced the theory in the general metric space setting; see
[HK95] and [HK98]. See Koskela’s notes [Kos07] for a ‘modern’ approach to QC
maps in the Euclidean setting.
There are three so-called definitions for QC maps: the metric definition, the
geometric definition, and the analytic definition. We present the first two. A
homeomorphism f : X → Y is (metrically) quasiconformal provided there is a
constant H < ∞ such that for all x ∈ X,
L(x, f, r)
lim sup H(x, f, r) ≤ H where H(x, f, r) = ,
rց0 l(x, f, r)
L(x, f, r) = sup{|f (y) − f (x)| : |x − y| ≤ r} ,
l(x, f, r) = inf{|f (y) − f (x)| : |x − y| ≥ r} .
A homeomorphism f : X → Y is (geometrically) quasiconformal provided there
is a constant K < ∞ such that for all curve families Γ in X,
K −1 mod(Γ) ≤ mod(f Γ) ≤ K mod(Γ).
Notice that unlike the metric definition, which makes sense for any pair of met-
ric spaces, the geometric definition requires measure metric spaces. These are
generally assumed to be Ahlfors Q-regular spaces (see §2.J) in which case mod(·)
denotes the Q-modulus.

2.D. Length and Geodesics. The length of a continuous path γ : [0, 1] → X


is defined in the usual way by
n
X
ℓ(γ) := sup |γ(ti ) − γ(ti−1 )| where 0 = t0 < t1 < · · · < tn = 1.
i=1

We call γ rectifiable when ℓ(γ) < ∞. We let Γ(x, y) = Γ(x, y; X) denote the
collection of all rectifiable paths joining x and y in X; in general we should
also indicate the metric in this notation, but it will always be understood from
context. Väisälä’s notes [Väi71, §1-§5] provide an excellent reference for studying
properties of curves, and the results are valid in the general metric space setting.
Each rectifiable path γ : [0, 1] → X has an associated arclength function s :
[0, 1] → [0, ℓ(γ)], given by s(t) = ℓ(γ[0, t]), which is of bounded variation. Given
a Borel measurable function ρ : X → [0, ∞], we define
Z Z 1
ρ ds := ρ(γ(t)) ds(t).
γ 0

An arc in a metric space X is the homeomorphic image of an interval I ⊂ R.


Given two points x and y on an arc α, we write α[x, y] to denote the subarc of
α joining x and y.
A geodesic in X is the image ϕ(I) of some isometric embedding ϕ : I → X
where I ⊂ R is an interval; we use the adjectives segment, ray, or line (re-
spectively) to indicate that I is bounded, semi-infinite, or all of R. When ϕ is
Uniformity and Hyperbolicity 89

L-bilipschitz we call ϕ(I) an L-quasigeodesic. More generally, if ϕ is (L, C)-


quasiisometric, then we call ϕ(I) an (L, C)-quasigeodesic. Thus γ is an L-
quasigeodesic precisely when
∀x, y ∈ γ : ℓ(γ[x, y]) ≤ L|x − y|;
classically, such curves in the plane R2 were called chord arc curves.
A metric space is geodesic if each pair of points can be joined by a geodesic
segment. We use the notation [x, y] to mean a (not necessarily unique) geodesic
segment joining points x, y; such geodesics always exist if our space is geodesic,
but may not be unique. (If there is some other distance function, such as k, then
we write [x, y]k to denote a k-geodesic joining x, y). We consider a given geodesic
[x, y] as being ordered from x to y (so we can use phrases such as the ‘first’ point
encountered). An unbounded metric space is roughly κ-starlike with respect to a
base point w if each point lies within distance κ of some geodesic ray emanating
from w.
The geodesic boundary ∂g X of an unbounded geodesic metric space X is the
set of equivalence classes of geodesic rays in X where two such rays are consid-
ered equivalent when they are at a finite Hausdorff distance from each other.
Equivalently, if α, β : [0, ∞) → X are geodesic rays in X, then α ≃ β if
supt |α(t) − β(t)| < ∞. The geodesic boundary of Rn is the sphere Sn−1 . The
geodesic boundary of hyperbolic n-space (Bn , h) is also the sphere Sn−1 .
Every metric space (X, d) admits a natural (or intrinsic) metric, the so-called
length distance given by
l(x, y) := inf{ℓ(γ) : γ a rectifiable curve joining x, y in Ω}.
A metric space (X, d) is a length space provided d(x, y) = l(x, y) for all points
x, y ∈ X; it is also common to call such a d an intrinsic distance function. Notice
that an l-geodesic [x, y]l is a shortest curve joining x and y.
The Hopf-Rinow Theorem (see [Gro99, p.9], [BBI01, p.51], [BH99, p.35]) says
that every locally compact length space is proper (and therefore geodesic). In
a general length space, when geodesics may not exist, one works with so-called
short arcs; see [Väi05a].
id
Since |x−y| ≤ ℓ(x, y) for all x, y, the identity map (X, l) → (X, d) is Lipschitz
continuous. It is important to know when this map will be a homeomorphism
(cf. [BHK01, Lemma A.4, p.92]). Notice that the identity map (X, d) → (X, l) is
uniformly locally Lipschitz when X is locally quasiconvex; see §2.E. More gen-
erally, one can show that the identity map (X, l) → (X, d) is a homeomorphism
precisely when X satisfies a weak notion of local quasiconvexity; see [BH07].

2.E. Connectivity Conditions. A metric space (X, d) is a-quasiconvex pro-


vided each pair of points can be joined by a path whose length is at most a
times the distance between its endpoints. A locally complete space X is locally
quasiconvex if there exists a constant a ≥ 1 such that for all z ∈ X, points
x, y ∈ λB(z) can be joined by a rectifiable arc α in X with ℓ(α) ≤ a|x − y|; we
90 D.A.Herron IWQCMA05

abbreviate this by the phrase ‘X is locally a-quasiconvex’. (Here it is understood


that there is some Whitney ball constant λ ∈ (0, 1) which may also depend on
other parameters).
A space (X, d) is c-linearly locally connected, or c-LLC, if c ≥ 1 and the
following conditions hold for all x ∈ X and all r > 0:
(LLC1 ) points in B(x; r) can be joined in B(x; c r)

and

(LLC2 ) points in X \ B̄(x; r) can be joined in X \ B̄(x; r/c).


Here the phrase ‘can be joined’ means ‘can be joined by a continuum’. We also
use the term LLC with respect to arcs in which case ‘can be joined’ means ‘can
be joined by a rectifiable arc’. Note that quasiconvexity implies LLC1 (even with
respect to arcs).
The generic example of a space which does not satisfy the LLC2 condition is
the interior of an infinite Euclidean cylinder such as Bn−1 × R ⊂ Rn . However,
for 2 ≤ k < n the regions Bn−k × Rk ⊂ Rn are easily seen to be 1-LLC2 . The
complement of a semi-infinite slab (e.g., Rn \ {(x1 , . . . , xn ) : x1 ≥ 0, |xn | ≤ 1})
fails to be LLC1 .
Ahlfors regular Loewner spaces are LLC; see [HK98, Theorem 3.13]. Uniform
domains also enjoy this property, but not necessarily uniform spaces. The LLC
condition was invented by Gehring who first used it to characterize quasidisks;
see [Geh82] and the references mentioned therein.

2.F. Doubling and Dimensions. The p-dimensional Hausdorff measure of a


set A ⊂ X is given by Hp (A) := limr→0 Hrp (A) where
X
Hrp (A) := inf{ diam(Bi )p : A ⊂ ∪Bi , Bi balls with diam(Bi ) ≤ r}.
p
The Hausdorff p-content of A is just H∞ (A). The Hausdorff dimension of A is
determined by
dimH (A) := inf {p > 0 : Hp (A) = 0} .
We also require the Assouad dimension of X which is given by
dimA (X) := inf{p : #S ≤ C(R/r)p for all S ⊂ X
with r ≤ |x − y| ≤ R for all x, y ∈ S}
where #S denotes the cardinality of the set S. See [Hei01, 10.15] or [Luu98,
3.2]. The spaces with finite Assouad dimension are precisely the doubling spaces
(which we discuss below in more detail). Finally, the conformal Assouad dimen-
sion of a metric space X is
c-dimA (X) := inf{dimA (X, d) : d ∈ G},
where G is the conformal gauge on X determined by the original metric; see
[Hei01, 15.8, p.125].
Uniformity and Hyperbolicity 91

A metric space (X, d) satisfies a (metric) doubling condition if there is a con-


stant N such that each ball in X of radius R can be covered by at most N balls of
radius R/2; these are precisely the spaces of finite Assouad dimension. A Borel
measure ν is a doubling measure on X if there is a constant D = Dν such that
ν[B(x; 2r)] ≤ D ν[B(x; r)] for all x ∈ X and all r > 0.
A Borel measure ν on X is p-homogeneous if there is a constant C = Cν such
that  p
ν[B(x; R)] R
≤C for all x ∈ X and all 0 < r ≤ R.
ν[B(x; r)] r
Obviously every homogeneous measure is doubling; the converse holds too with
C = D and p = log2 (D). Every Ahlfors Q-regular measure is Q-homogeneous.
The existence of a doubling measure is easily seen to imply a metric doubling
condition; the converse holds if our metric space is complete. Here is a precise
statement of this result, which is due to Vol’berg and Konyagin for compact
spaces, and Luukkainen and Saksman for complete spaces (see [Hei01, Theo-
rem 13.5]).
2.3. Fact. A complete doubling space X carries a p-homogeneous measure for
each p > dimA (X).

An especially important property of doubling measures is their exponential


decay on uniformly perfect spaces, which we record as follows; see [Hei01, (13.2)]
or [Sem99, Lemma B.4.7, p.420].
2.4. Fact. Let ν be a doubling measure on a uniformly perfect metric space.
There are constants C ≥ 1 and α > 0, depending only on the doubling constant
for ν and the uniformly perfect constant, such that for all balls B(z; r) ⊂ B(x; R),
ν[B(z; r)]  r α
≤C .
ν[B(x; R)] R
Now we discuss an interesting way to deform the geometry of a doubling space.
Let ν be a doubling measure on a metric space (X, d). For each α > 0 we define
δ = δν,α by
(2.5) δ(x, y) := ν[B(xy)]α , where B(xy) := B̄(x; |x − y|) ∪ B̄(y; |x − y|);
see [DS97, §16.2], [Sem99, (B.3.6)], [Hei01, 14.11]. This always defines a quasi-
metric on X, and, when X is uniformly perfect, the identity map (X, d) → (X, δ)
will be quasisymmetric and (X, δ, ν) will be Ahlfors (1/α)-regular. Moreover,
there is an α0 > 0 (depending only on the doubling constant for ν) such that
for all 0 < α < α0 , δν,α is bilipschitz equivalent to an honest distance function
on X (see Fact 2.1). In particular, if ν is p-homogeneous, then δν,1/p is already
bilipschitz equivalent to an honest distance function (e.g., if (X, d, ν) is Ahlfors
p-regular, then δν,1/p is bilipschitz equivalent to d).
In conjunction with the above chain of ideas, we declare ν to be a p-dimensional
metric doubling measure on X if ν is a doubling measure on X with the prop-
erty that δν,1/p is bilipschitz equivalent to a distance on X. For example, a
92 D.A.Herron IWQCMA05

p-homogeneous measure will be a p-dimensional metric doubling measure. We


summarize the above comments; see [Hei01, 14.11,14.14], [Sem99, B.3.7, B.4.6,
p.421], [DS97, 16.5,16.7,16.8].
2.6. Fact. Let ν be a p-dimensional metric doubling measure on a uniformly
perfect metric space (X, d). Define δ = δν,1/p as in (2.5). Then δ is a quasimetric
on X which is bilipschitz equivalent to a distance function on X, the identity map
(X, d) → (X, δ) is quasisymmetric, and (X, δ, ν) is an Ahlfors p-regular space.
All of the new parameters depend only on the original data for X and ν.

There is one final comment we wish to point out regarding the quasimetrics
δν,α . As above, suppose ν is a doubling measure on a metric space (X, d), and
suppose X has another metric, say, ∂ which is QS equivalent to d. Then by
using the doubling property of ν in conjunction with quasisymmetry we see that
ν[Bd (xy)] ≃ ν[B∂ (xy)] (where these sets are defined as above using balls centered
at x and y in the appropriate metrics); here the constant depends only on the
doubling constant and the quasisymmetry data. It therefore follows that the
quasimetric δd (defined as in (2.5) via Bd (xy)) is bilipschitz equivalent to δ∂
(defined via B∂ (xy)).
We note the important fact that quasisymmetric homeomorphisms preserve
these doubling conditions; cf. [Hei01, Theorem 10.18] or [DS97, Lemma 16.4].
In particular, the notions of doubling measure, the quasimetrics δν,α , and metric
doubling measures do not depend on the given distance function per se; they all
make sense for a conformal gauge.

2.G. Quasiconformal Deformations. Given an abstract domain (Ω, d, µ) and


a positive Borel measurable function ρ on Ω, we wish to define a new metric mea-
sure space Ωρ = (Ω, dρ , µρ ) which is a quasiconformal deformation of Ω. (Above
in §2.F we described another method for deforming the geometry of Ω which was
based on having a doubling measure. See Fact 2.6.)
We start by defining the ρ-length of a rectifiable curve γ via
Z
ℓρ (γ) := ρ ds
γ

and then the ρ-distance between two points x, y is


dρ (x, y) := inf{ℓρ (γ) : γ a rectifiable curve joining x, y in Ω};
see §2.D. The careful reader no doubt recognizes that, in general, dρ (x, y) could
be zero or even infinite; in order to ensure that dρ be an honest distance function,
we must require that 0 < dρ (x, y) < ∞ for all points x, y ∈ Ω. We designate
this by calling such ρ a metric-density on Ω. One way to guarantee this is to ask
that ρ be locally bounded away from zero and infinity. In practice, our densities
will always satisfy a Harnack inequality—see below—so this is never a problem
for us.
Uniformity and Hyperbolicity 93

The ρ-balls (etc.) are written as Bρ (x; r); these are the metric balls in Ωρ , so
Bρ (x; r) = {y ∈ Ω : dρ (x, y) < r}. We define a new measure µρ by
Z
(2.7) µρ (E) := ρQ dµ.
E

Here Q is usually the Hausdorff dimension of (Ω, d).


When Ωρ is non-complete (which will often be the case for us), we can form
∂ρ Ω = Ω̄ρ \ Ωρ and define dρ (x) = distρ (x, ∂ρ Ω). In this setting we also employ
the notation Bρ (x) = Bρ (x; dρ (x)); thus λBρ (x) is a Whitney ball in Ωρ .
We are especially interested in the metric-densities ρ for which Ωρ is a uniform
space, and we call such a ρ a uniformizing density (which implicitly includes
the hypothesis that Ωρ is non-complete). The Bonk-Heinonen-Koskela theory
produces uniformizing densities on proper geodesic Gromov hyperbolic spaces;
see Fact 4.1. Some other classes of metric-densities which we wish to single out
for attention include Harnack, Ahlfors, and Koebe densities; their definitions
follow below. We let Hρ , Aρ , Kρ denote the parameters associated with these
densities.
Before delving into the technical definitions, we wish to make a few com-
ments. The reader no doubt has encountered deformations of Euclidean domains
Ω ⊂ Rn by continuous densities ρ; in this setting Ωρ is a conformal deformation
of Ω, meaning that the identify map Ω → Ωρ is conformal (i.e., metrically 1-
quasiconformal). However, in our more general setting, even for the case ρ = 1
say, the identity map Ω → Ωρ = Ωl may fail to be quasiconformal (e.g., if Ω does
not satisfy some sort of local quasiconvexity condition). A similar phenomenon
holds for Borel metric-densities, even for domains Ω ⊂ Rn . Nonetheless, when
Ω is locally quasiconvex and ρ is a Harnack metric-density, Lemma 2.8 below
reveals that the identity map Ω → Ωρ is QC (and according to Proposition 2.9
even QS under the right circumstances). This is a good thing: we want Ωρ to be
a quasiconformal deformation of Ω.
With this in mind, we pronounce the following definitions. First, we declare ρ
to be a bounded density if the deformed space Ωρ is bounded, i.e., diamρ (Ω) < ∞.
Next, we call ρ a Harnack density provided it satisfies a uniform local Harnack
type inequality: for all points x in Ω,
1 ρ(y)
(H) ≤ ≤H for all y ∈ λB(x).
H ρ(x)
Here H = Hρ ≥ 1 and 0 < λ < 1 (generally λ will be small). Note that in contrast
to the situation in [BKR98, p.637], the validity of (H) for some 0 < λ < 1 need
not mean a similar set of inequalities will hold for λ = 1/2. The condition (H)
provides local control and permits the use of standard chaining type arguments;
e.g., see Lemma 2.13.
We call ρ an Ahlfors density if the associated metric measure space Ωρ =
(Ω, dρ , µρ ) is Ahlfors upper Q-regular (cf. §2.J); i.e., if µρ satisfies a global upper
94 D.A.Herron IWQCMA05

Ahlfors Q-regular volume growth estimate: there is a constant A = Aρ such that


for all points x in Ω,
(A) µρ [Bρ (x; r)] ≤ A rQ for all r > 0.
The positive real number Q is generally the Hausdorff dimension of our space;
it must agree with the number Q appearing in the definition of a Loewner space
(a notion also discussed in §2.J). The volume growth condition (A) ensures
that Ωρ satisfies an upper mass condition and so provides modulus estimates via
Facts 2.15, 2.16, 2.17.
Below (in §2.H) we discuss the density 1/d which determines the quasihyper-
bolic distance; of course this is a continuous Harnack density, but in general 1/d
does not satisfy the volume growth requirement (A).
We call ρ a Koebe density if Ωρ is non-complete and there is a constant K = Kρ
such that dρ (x) = distρ (x, ∂ρ Ω) enjoys the property
1 dρ (x)
(K) ≤ ≤K for all x ∈ Ω.
K ρ(x)d(x)
(Note that when ρ is a Harnack density, dρ (x) ≥ (λ/H)ρ(x)d(x) always holds,
and so it is the upper estimate which is needed.) For example, if ρ = |f ′ | where f
is a holomorphic homeomorphism defined in a subdomain Ω of the complex plane,
then a classical theorem in univalent function theory asserts that ρ is a Koebe
density with constant K = 4. As another example we note that Theorem E
asserts that any Harnack Ahlfors density on a uniform Loewner space (with suf-
ficiently ‘thick’ boundary) is a Koebe density; see [Her04, Theorem E]. We point
out that when ρ is a Koebe density on (Ω, d), the identity map (Ω, k) → (Ωρ , kρ )
is easily seen to be Kρ -bilipschitz; here (Ωρ , kρ ) denotes the quasihyperbolization
of Ωρ .
We employ the terminology conformal density for a metric-density which is
Harnack, Ahlfors, and Koebe. A basic example of a conformal density is ρ = |f ′ |
for any holomorphic homeomorphism |f ′ | defined in a subdomain of the complex
plane; we refer to [BKR98, Section 2] for other examples of conformal densities
on the Euclidean unit ball. The reader should be aware that the phrase ‘ρ is a
conformal density’ does not necessarily mean that the identity map Ω → Ωρ is
quasiconformal (unless Ω is locally quasiconvex).
Here are some especially useful estimates which also provide information con-
cerning the identity map Ω → Ωρ for certain densities. Roughly speaking, this
map is locally bilipschitz (therefore quasiconformal) for Harnack densities and
uniformly locally quasisymmetric for Harnack Koebe densities, provided Ω is lo-
cally quasiconvex. Proposition 2.9 gives a significant strengthening of this result.
2.8. Lemma. Let ρ : Ω → (0, ∞) be a Harnack density on a locally a-
quasiconvex abstract domain (Ω, d). Put η = λ/2a. Then for all z ∈ Ω,
1 dρ (x, y)
≤ ≤ aH for all points x 6= y in ηB(z);
H ρ(z)|x − y|
Uniformity and Hyperbolicity 95

in particular,
1 diamρ [ηB(z)]
≤ ≤ 2aH.
H ηρ(z)d(z)
If ρ is also a Koebe density, then for all 0 ≤ ϑ ≤ λ/2C and all x ∈ Ω,
C −1 ϑB(x) ⊂ ϑBρ (x) ⊂ CϑB(x),
where C = aHK. Here H = Hρ , K = Kρ and λ is the Whitney ball constant.

One immediate consequence of Lemma 2.8 is that the identity map Ω → Ωρ


is metrically quasiconformal with linear dilatation aH 2 . In addition, because of
the definition of the associated measure (see (2.7)), a straightforward calcula-
tion reveals that this identity map is geometrically quasiconformal with inner
dilatation H Q and outer dilatation (aH)Q . (Here we assume a Harnack density
on a locally quasiconvex Ω.) It is therefore natural to inquire about possible
quasisymmetry properties of this identity map.
Heinonen and Koskela proved that a quasiconformal map of bounded Ahlfors
regular spaces, with domain a Loewner space and a linearly locally connected
target space, is in fact quasisymmetric [HK98, Theorem 4.9]. The corollary to
the following analog of their result is used in the proof of Theorem D; note that
here our domain space is not assumed to be Ahlfors regular.
2.9. Proposition. Let Ω be a bounded locally quasiconvex Q-Loewner space.
Suppose ρ is a conformal density on Ω with Ωρ a bounded linearly locally con-
nected space. Then the identity map Ω → Ωρ satisfies the weak-quasisymmetry
condition
∀x, y, z ∈ Ω : |x − y| ≤ |x − z| =⇒ dρ (x, y) ≤ Ldρ (x, z)
for some constant L which depends only on the data associated with Ω, ρ, Ωρ ,
and the ratios r, q given in the proof.
2.10. Corollary. Let Ω be a bounded quasiconvex Q-Loewner space. Suppose
ρ is a conformal density on Ω and Ωρ is a bounded Q-Loewner space. Then the
identity map Ω → Ωρ is quasisymmetric with a distortion function which depends
only on the data associated with Ω, ρ, Ωρ , and the ratios r, q given in the proof
of Proposition 2.9.

As an exercise to help understand the various properties of these metric-


densities, the interested reader can provide a proof for the following [Her06,
Lemma 2.6].
2.11. Lemma. Suppose (Ω, d, µ) is a locally a-quasiconvex abstract domain. Let
τ be a positive Borel function on Ω which is locally bounded away from 0 and
∞. Put ∆ = Ωτ . If σ is a metric-density on ∆, then its pull-back ρ = σ τ is a
metric-density on Ω, Ωρ = ∆σ , and
(a) ρ and σ either are, or are not, both Ahlfors regular (with Aρ = Aσ ),
(b1) if σ, τ are both Koebe, then so is ρ with Kρ = Kσ Kτ ,
96 D.A.Herron IWQCMA05

(c1) if σ, τ are both Harnack, then so is ρ with Hρ = Hσ Hτ .


On the other hand, if ρ is a metric-density on Ω, then its push-forward σ = ρ τ −1
is a metric-density on ∆, ∆σ = Ωρ , (a) holds, and
(b2) if ρ, τ are Koebe, then so is σ with Kσ = Kρ Kτ ,
(c2) if ρ, τ are Harnack and τ is Koebe, then σ is Harnack with Hσ = Hρ Hτ .

2.H. Quasihyperbolic Distance and Geodesics. The quasihyperbolic dis-


tance in an abstract domain (Ω, d) is defined by
Z
ds
k(x, y) = kΩ (x, y) := inf ℓk (γ) = inf
γ d(z)

where the infimum is taken over all rectifiable curves γ which join x, y in Ω.
The quasihyperbolization of an abstract domain (Ω, d) is the metric space (Ω, k)
obtained by using quasihyperbolic distance. It is not hard to see that (Ω, k)
is complete, provided the identity map (Ω, ℓ) → (Ω, d) is a homeomorphism;
see [BHK01, Proposition 2.8]. Thus by the Hopf-Rinow theorem ([Gro99, p.9],
[BBI01, p.51], [BH99, p.35]), every locally compact abstract domain has a proper
(hence geodesic) quasihyperbolization.
We call the geodesics in (Ω, k) quasihyperbolic geodesics; see §2.D. Note that
when ρ is a Koebe density on Ω, the identity map (Ω, k) → (Ωρ , kρ ) is bilips-
chitz and we find that quasihyperbolic geodesics in Ω are quasihyperbolic quasi-
geodesics in Ωρ ; that is, a geodesic in (Ω, k) will be a quasigeodesic in (Ωρ , kρ )
(the quasihyperbolization of Ωρ ).
We remind the reader of the following basic estimates for quasihyperbolic
distance, first established by Gehring and Palka [GP76, Lemma 2.1]:
   
ℓ(x, y) |x − y| d(x)
k(x, y) ≥ log 1 + ≥ j(x, y) = log 1 + ≥ log .
d(x) ∧ d(y) d(x) ∧ d(y) d(y)
See also [BHK01, (2.3),(2.4)]. The first inequality above is a special case of the
more general (and easily proved) inequality,
 
ℓ(γ)
ℓk (γ) ≥ log 1 +
d(x) ∧ d(y)
which holds for any rectifiable curve γ with endpoints x, y.
An immediate consequence of the above inequalities is that the identity map
(Ω, k) → (Ω, d) is continuous; indeed,
Bk (x; R) ⊂ (eR − 1)B(x) for all x ∈ Ω and all R > 0,
where Bk (x; R) denotes the R-ball centered at x in (Ω, k). It is important to
know when this map will be a homeomorphism (which, according to [BHK01,
Lemma A.4, p.92], will be the case if and only if the identity map (Ω, ℓ) → (Ω, d)
is a homeomorphism). The following provides quantitative information concern-
ing this question; it is easy to verify via simple estimates for the quasihyperbolic
lengths of the ‘promised short arcs’.
Uniformity and Hyperbolicity 97

2.12. Lemma. Suppose that (Ω, d) is a locally a-quasiconvex abstract domain.


Then for all x ∈ Ω and all R > 0,
τ B(x) ⊂ Bk (x; R) provided 0 ≤ τ ≤ min{λ, R/[a(1 + R)]}.
As an exercise, the reader can check that for a domain in Rn , |x − y| ≤
[d(x) + d(y)]/2 =⇒ k(x, y) ≤ 2. Thus (2/3)B(x) ⊂ Bk (x; 2).
The Harnack inequality (H), as stated in §2.G, only requires that ρ be es-
sentially constant on Whitney type balls. We can do the usual chaining type
arguments to see that such a density ρ will satisfy a Harnack type inequality on
much bigger sets, of course with a change in the Harnack constant. Here is a
useful example of this phenomena.
2.13. Lemma. Let ρ be a Harnack density on an abstract domain (Ω, d). If
x, y ∈ Ω satisfy k(x, y) ≤ K, then 1/H1 ≤ ρ(y)/ρ(x) ≤ H1 , where H1 =
H1 (K, Hρ , λ).
We conclude this subsection with a covering lemma for quasihyperbolic geodesics.
2.14. Lemma. Suppose that (Ω, d) is a locally a-quasiconvex abstract domain.
Let γ be a quasihyperbolic geodesic segment or ray in Ω with endpoint x0 . Let
x0 , x1 , x2 , . . . be successive points along γ with k(xi , xi−1 ) = K ≤ log(1 + τ )
where τ = Pmin{λ, 1/2a}. Then the balls Bi = τ B(xi ) cover γ and have bounded
overlap: χBi ≤ Cχ∪Bi , where C = 1 + 4/K.
2.I. Modulus and Capacity. For p ≥ 1 we define the p-modulus of a family
Γ of curves in a metric measure space (X, d, µ) by
Z
modp Γ := inf ρp dµ,

where
R the infimum is taken over all Borel functions ρ : X → [0, ∞] satisfying
γ
ρ ds ≥ 1 for all locally rectifiable curves γ ∈ Γ. Then the p-modulus of a pair
of disjoint compact sets E, F ⊂ X is
modp (E, F ; X) := modp Γ(E, F ; X)
where Γ(E, F ; X) is the family of all curves joining the sets E, F in X. We also
let Γr (E, F ; X) be the subfamily of Γ(E, F ; X) consisting of the rectifiable paths
joining E, F .
An important property is that under fairly general circumstances, modp (E, F ; X)
agrees with the p-capacity of the pair E, F . There is extensive literature regard-
ing these “capacity equals modulus” results; for a start, see [HK98, Proposi-
tion 2.17].
For the reader’s convenience, we cite the following modulus estimates. First
we have the standard Long Curves Estimate; see [HK98, 3.15].
2.15. Fact. Let x ∈ X and suppose that the upper mass condition µ[B(x; R)] ≤
M Rp holds for some R > 0. Let Γ be a family of curves in B(x; R) and suppose
that each γ ∈ Γ has arclength ℓ(γ) ≥ L > 0. Then
modp Γ ≤ L−p µ[B(x; R)] ≤ M (R/L)p .
98 D.A.Herron IWQCMA05

Next we record the Spherical Ring Estimate; see [HK98, 3.14, p.17].
2.16. Fact. Let x ∈ X, 0 < 2r ≤ R, and suppose that the upper mass condition
µ[B(x; t)] ≤ M tp holds for all 0 < t < r + R. Then
modp (B̄(x; r), X \ B(x; R); X) ≤ C (log(R/r))1−p ,
where C = 2p+1 M/ log 2.

Finally, we require the following Basic Modulus Estimate; see [BKR98, Lemma
3.2].
2.17. Fact. Let (X, d, µ) be a metric measure space. Assume that ρ is a metric-
density on X whose associated measure (2.7) satisfies the Ahlfors volume growth
condition (A) at some point x ∈ E ⊂ X. Suppose that L > λ ≥ diamρ E,
and that Γ is some family of curves γ in X each having one endpoint in E and
satisfying ℓρ (γ) ≥ L. Then
modQ Γ ≤ C (log (1 + L/λ))1−Q ,
where C = 2Q+1 A/ log 2.
2.18. Corollary. Let (X, d, µ) be a metric measure space. Assume that for
some point x ∈ E ⊂ X, the upper mass condition µ[B(x; r)] ≤ M rQ holds for
all r > 0. Suppose that Γ is a family of curves γ in X each having one endpoint
in E and satisfying ℓ(γ) ≥ L > diam E. Then
modQ Γ ≤ C (log (1 + L/ diam E))1−Q ,
where C = 2Q+1 M/ log 2.

In Euclidean space Rn , the n-modulus is also called the conformal modulus and
simply denoted by mod(·). Below we state some well-known geometric estimates
for the conformal modulus mod(E, F ; Ω). Here and elsewhere in these notes,
∆(E, F ) := dist(E, F )/ min{diam(E), diam(F )}
is the relative distance between the pair E, F of nondegenerate disjoint continua.
2.19. Facts. Let E, F be disjoint compact sets in Rn .
(a) If E, F are separated by the spherical ring B(x; s) \ B̄(x; t), then
mod(E, F ; Rn ) ≤ ωn−1 (log(s/t))1−n .
(b) If E ∩ S(x; r) 6= ∅ 6= F ∩ S(x; r) for all t < r < s, then
mod(E, F ) ≥ σn log(s/t).
(c) If both E and F are connected, then
σn log(1 + 1/∆(E, F )) ≤ mod(E, F ; Rn ) ≤ Ωn (1 + 1/∆(E, F ))n .
(d) (Comparison Principle) If A, B, E, F ⊂ Ω with A, B also compacta, then
mod(E, F ; Ω) ≥ 3−n min{mod(E, A; Ω), mod(F, B; Ω), I},
where I = inf{mod(α, β; Ω) | α ∈ Γr (E, A; Ω), β ∈ Γr (F, B; Ω)}.
Uniformity and Hyperbolicity 99

(e) (Teichmüller Estimate) If E, F are both connected, then for all x, y ∈ E


and z, w ∈ F
 
n |x − z||y − w|
mod(E, F ; R ) ≥ τ
|x − y||z − w|
where τ (r) is the capacity of the Teichmüller ring
Rn \ {−1 ≤ x1 ≤ 0 or x1 ≥ r};
i.e, τ (r) = mod([−e1 , 0], [re1 , ∞]; Rn ).
(f) There exists λ = λ(n) ∈ [6, 5e(n−1)/2 ) such that when E, F are both con-
nected and ∆(E, F ) ≥ 1,
21−n ωn−1 [log(λ∆(E, F ))]1−n ≤ mod(E, F ; Rn ) ≤ ωn−1 [log(∆(E, F ))]1−n .
(g) (Carleman Inequality) For E ⊂ Ω,
mod(E, ∂Ω; Ω) ≥ nn−1 ωn−1 (log(|Ω|/|E|))1−n .
Here σn and ωn−1 , Ωn are the spherical cap constant and the measures of the
(n − 1)-sphere, n-ball respectively.

Most of these estimates can be found in [Väi71] or [Vuo88]. Lemma 2.5 in


[BH06] gives a precise formula for mod(E, F ; Rn ) in the case when E, F are
disjoint closed balls.

2.J. Ahlfors Regular and Loewner Spaces. A metric measure space (X, d, µ)
is Ahlfors Q-regular provided there exists a finite constant M = Mµ such that
for all x ∈ X and all 0 < r ≤ diam Ω,
M −1 rQ ≤ µ[B(x; r)] ≤ M rQ .
The positive real number Q will then be the Hausdorff dimension of (X, d),
and the Q-dimensional Hausdorff measure HQ on X will also satisfy the above
inequalities (possibly with a change in the constant M ). A metric space (X, d)
is Ahlfors Q-regular if (X, d, HQ ) is Ahlfors Q-regular. We use the adjectives
upper or lower to indicate that only one of these inequalities is in force, and—in
the abstract domain setting—add the adjective locally to mean that the required
inequality holds (or, inequalities hold) for Whitney balls (i.e., for radii 0 < r ≤
λd(x)).
There is an interesting result which gives upper estimates for the Assouad
dimension of subsets of Ahlfors regular spaces. See [BHR01, 3.12], [DS97, 5.8],
[Luu98, 5.2].
2.20. Fact. Suppose X is an Ahlfors Q-regular space and let M ⊂ X. Then
dimA M < Q if and only if M is porous in X; the constants depend only on each
other and the HQ -regularity constant.

The notion of a Loewner space was introduced by Heinonen and Koskela in


their study [HK98] of quasiconformal mappings of metric spaces; Heinonen’s
recent monograph [Hei01] renders an enlightening account of these ideas. A
100 D.A.Herron IWQCMA05

path-connected metric measure space (X, d, µ) is a Q-Loewner space, Q > 1,


provided the Loewner control function
ϕ(t) := inf{modQ (E, F ; X) : ∆(E, F ) ≤ t}
is strictly positive for all t > 0; here E, F are non-degenerate disjoint continua
in Ω and
∆(E, F ) := dist(E, F )/ min{diam(E), diam(F )}
is the relative size of the pair E, F . Note that we always have Q ≥ dimH (Ω) ≥ 1.
When (Ω, d, µ) is an n-Loewner space with Ω ⊂ Rn a domain and d, µ are
Euclidean distance and Lebesgue n-measure respectively, we simply call Ω a
Loewner domain. This is a generalization of Väisälä’s notion of a broad domain
(which he introduced in his analysis [Väi89, 2.15] of space domains QC equivalent
to a ball, and also used in his study [NV91, 3.8] of John disks), which in turn
is an analog of the quasiextremal distance domains first studied by Gehring and
Martio [GM85].
We call Ω ( Rn a ψ-QED domain if ψ : [0, ∞) → [0, ∞) is a homeomorphism
and for all disjoint continua E, F in Ω,
mod(E, F ; Ω) ≥ ψ(mod(E, F ; Rn )).
Clearly, ψ(t) ≤ t is a necessary restriction on such ψ. Also, every ψ-QED domain
is Loewner. The typical nonlinear functions ψ that arise in the literature have the
form ψp,M (t) = M −1 min{tp , t1/p } with p, M ≥ 1, a condition we call M -QEDp ,
or simply M -QED when p = 1.
The most important, and original, inequalities of this form are the M -QED
conditions corresponding to ψ(t) = t/M for some constant M ≥ 1. This
idea was introduced by Gehring and Martio who called such regions quasiex-
tremal distance domains. The terminology arises from the fact that the quantity
mod(E, F ; Ω)1/(1−n) is the extremal distance between E and F in Ω. When we
speak of a QED domain or a QED condition, we always mean an M -QED domain
or an M -QED condition for some M ≥ 1.
As in [HK96] we can consider the location of the continua E, F as well as
looking at special types of continua. In particular we can relax the ψ-QED
inequality by requiring it to hold only for all disjoint closed balls (or just closed
Whitney balls) to get the class ψ-QEDb (or ψ-QEDwb , respectively). Precise
definitions can be found in [BH06].
Every a-uniform domain in Rn is M -QED for some M = M (a, n); this follows
easily from Jones’ extension result for Sobolev spaces [Jon81, Theorem 1]. Also,
it is trivially true that
QED =⇒ ψ − QED =⇒ ψ − QEDb =⇒ ψ − QEDwb .
The converse of the middle implication fails; see[HK96, Example 4.1] and [BH06,
Example 4.2]. According to [BH06, Theorem 3.3], the last implication is re-
versible modulo a quantitative change in ψ. In addition, we always have
ψ − QED ⇐⇒ Loewner =⇒ QEDwb =⇒ ψ − QEDwb ⇐⇒ Loewnerwb
Uniformity and Hyperbolicity 101

where the last condition means that the Loewner condition is assumed only
for Whitney balls. Examples 4.2 and 4.3 in [BH06] illustrate that in general
the converses of the middle two implications fail to hold. The first equivalence
is established in [BH06, Theorem 1.3]. It remains open as to whether or not
Loewner domains (i.e. ψ-QED domains) are always QED.

2.K. Slice Conditions. There are various so-called slice conditions each de-
signed to handle their own specific problem. The ideas here are due to Buckley
et al. and his exposition [Buc03] is the place to begin reading about this topic.
He and his many co-authors have utilized an assortment of slice conditions to
investigate all kinds of different problems.
A non-empty bounded open set S ⊂ X is called a C-slice separating x, y
provided
∀ α ∈ Γ(x, y) : ℓ(α ∩ S) ≥ diam(S)/C
and
C −1 B(x) ∩ S = ∅ = S ∩ C −1 B(y) .
A set of C-slices for x, y ∈ X is a collection S of pairwise disjoint C-slices
separating x, y in X. One can show (see [BS03, (2.1)]) that the cardinality of
any such set S of C-slices separating x, y is always bounded by #S ≤ C 2 k(x, y).
We are interested in knowing when we can reverse this inequality. Since there
may be no C-slices separating x, y, we consider the quantity
dws (x, y) = dws (x, y; C) = dX
ws (x, y; C) := 1 + sup #S

where the supremum is taken over all S which are sets of C-slices in X separating
x, y, and #S denotes the cardinality of S.
We call (X, d) a weak C-slice space provided for all x, y ∈ X,
k(x, y) ≤ C dws (x, y; C),
Thus in these spaces dws (x, y) ≃ k(x, y), at least when k(x, y) ≥ 2. The weak
slice condition was introduced in [BO99, Section 5]; see also [BS03], [Buc03],
[Buc04]. When the weak C-slice space (X, d) is a domain Ω ( Rn , we call Ω a
weak C-slice domain.
The following rather technical lemma is quite useful for obtaining an up-
per bound for the cardinality of a set of slices; in weak slice spaces it provides
an upper bound for quasihyperbolic distances. It is the case α = 0 of [BS03,
Lemma 2.17].
2.21. Lemma. Let Γ be a 1-rectifiable subset of a rectifiably connected metric
space (X, d). Suppose ϕ : Γ → [ε, ∞) (with ε > 0) and S is a collection of
disjoint non-empty bounded subsets of X. Suppose also that there exist positive
constants b, c such that
(a) ∀S ∈ S : ℓ(S ∩ Γ) ≥ c diam(S) ,
(b) ∀S ∈ S , ∀z ∈ S ∩ Γ : ϕ(z) ≤ diam(S) ,
102 D.A.Herron IWQCMA05

(c) ∀t > 0 : ℓ(ϕ−1 (0, t]) ≤ b t .


Then the cardinality of S is at most #S ≤ 2(b/c) log2 (4ℓ(Γ)/cε).

3. Uniform Spaces
Roughly speaking, a space is uniform provided points in it can be joined by so-
called bounded turning twisted double cone arcs, i.e. paths which are not too long
and which stay away from the regions boundary. Uniform domains in Euclidean
space were first studied by John [Joh61] and Martio and Sarvas [MS79] who
proved injectivity and approximation results for them. They are well recognized
as being the ‘nice’ domains for quasiconformal function theory as well as many
other areas of geometric analysis (e.g., potential theory); see [Geh87] and [Väi88].
Every (bounded) Lipschitz domain is uniform, but generic uniform domains may
very well have fractal boundary. Recently, uniform subdomains of the Heisenberg
groups, as well as more general Carnot groups, have become a focus of study;
see [CT95], [CGN00], [Gre01].

3.A. Euclidean Setting. When our uniform space (see the definition given
below in §3.B) (Ω, d) is a domain Ω ⊂ Rn with Euclidean distance, we simply
call Ω a uniform domain. Every plane uniform domain is a quasicircle domain
(each of its boundary components is either a point or a quasicircle), and a finitely
connected plane domain is uniform if and only if it is a quasicircle domain. How-
ever, the plane punctured at the integers is not uniform. Such nice topological
information is not true for uniform domains in higher dimensions. For example,
a ball with a radius removed is uniform; this is not true when n = 2.
For domains in Rn we can consider uniformity both with respect to the Eu-
clidean distance and with respect to the induced length metric also. The latter
class of domains are usually called inner uniform; cf. [Väi98]. For example, a slit
disk in the plane is not uniform (with respect to Euclidean distance) but it is an
inner uniform domain. On the other hand, an infinite strip, or the inside of an
infinite cylinder in space, is not uniform nor inner uniform. The region between
two parallel planes is not uniform nor inner uniform. Every quasiball is uniform.

3.B. Measure Metric Space Setting. Following [BHK01], a uniform space is


an abstract domain (so, a non-complete, locally complete, rectifiably connected
metric space) (Ω, d) with the property that there is some constant a ≥ 1 such
that each pair of points can be joined by an a-uniform arc. A rectifiable arc γ
joining x, y in Ω is an a-uniform arc provided
ℓ(γ) ≤ a|x − y|
and
min{ℓ(γ[x, z]), ℓ(γ[y, z])} ≤ a d(z) for all z ∈ γ.
Here ℓ(γ) is the arclength of γ and γ[x, z] denotes the subarc of γ between x, z.
The second inequality above ensures that Ω contains the twisted double cone
∪{B(z; ℓ(z)/a) : z ∈ γ} where ℓ(z) denotes the left-hand-side of this inequality;
Uniformity and Hyperbolicity 103

the first inequality asserts that this twisted double cone is not too ‘crooked’.
Consequently, we call γ a double a-cone arc if it satisfies the second inequality
above (the phrases cigar arc and corkscrew are also used).

3.C. Basic Information. An important, and characteristic, property of uni-


form spaces is that quasihyperbolic geodesics are uniform arcs. (See [GO79,
Theorems 1,2] for domains in Euclidean space and [BHK01, Theorem 2.10] for
general metric spaces.) Slight alterations to the proof of [BHK01, Theorem 2.10]
yield the following generalization of this property.
3.1. Fact. In an a-uniform space, quasihyperbolic c-quasigeodesics are b-uniform
arcs where b = b(a, c).

In general, quasihyperbolic geodesics may not exist; see [Väi99, 3.5] for an
example due to P. Alestalo. However, one can still show that quasihyperbolically
short arcs are uniform arcs. One can prove that boundary points in a locally
compact uniform space can be joined by quasihyperbolic geodesics, and these
geodesics are still uniform arcs.
Another crucial piece of information is a characterization of uniformity due
to Gehring and Osgood [GO79, Theorems 1,2]; Bonk, Heinonen, and Koskela
[BHK01, Lemma 2.13] verified the necessity of this condition for the metric space
setting, while the Gehring-Osgood argument can be modified to establish the
sufficiency. Recalling the basic estimates for quasihyperbolic distance, we see that
uniform spaces are precisely those abstract domains in which the quasihyperbolic
distance is bilipschitz equivalent to the j distance. See also Theorem 6.1.
3.2. Fact. An abstract domain is a-uniform if and only if k(x, y) ≤ b j(x, y) for
all points x, y. The constants a and b depend only on each other.

We conclude this subsection with a useful fact regarding bounded uniform


spaces; see [Her06, Lemmas 2.12,2.13].
3.3. Lemma. Let ρ be a Harnack Koebe density on a bounded a-uniform space
(Ω, d). Suppose that Ωρ is bounded and that quasihyperbolic geodesics in Ω are
double a-cone arcs in Ωρ . Then for any positive constant C,
diamρ [CB(z)] ≃ dρ (z) for all z ∈ Ω,
where the constant depends only on C, Hρ , Kρ , a, λ and the quantity q given in
the proof.

4. Gromov Hyperbolicity
Good sources for information concerning Gromov hyperbolicity include [BHK01],
[BBI01], [BS00], [BH99], [Bon96] and especially the references mentioned in these
works. Väisälä has an especially nice treatment [Väi05a] of Gromov hyperbolicity
for spaces which are not assumed to be geodesic nor proper. Note however that
Bonk and Schramm have demonstrated that every Gromov δ-hyperbolic metric
104 D.A.Herron IWQCMA05

space can be isometrically embedded into some complete geodesic δ-hyperbolic


space; see [BS00, Theorem 4.1].
Hästö [Häs06] has an intriguing result giving a striking contrast between the
hyperbolicity of (Ω, j) versus that of (Ω, j̃) for Ω ( Rn : the latter space is
always Gromov hyperbolic whereas the former is Gromov hyperbolic precisely
when Ω has exactly one boundary point. This is quite surprising as these spaces
are bilipschitz equivalent (indeed, j̃ ≤ j ≤ 2j̃). It is known that for intrinsic
spaces, so also for geodesic spaces, Gromov hyperbolicity is preserved under
(L, C)-quasiisometries. In particular, Hästö’s result illustrates the failure of this
property in the non-intrinsic setting.

4.A. Thin Triangles Definition. A geodesic metric space is Gromov hyper-


bolic if its geodesic triangles are δ-thin for some δ > 0, which means that each
point on the edge of any geodesic triangle is within distance δ of some point on
one of the other two edges. That is, if [x, y] ∪ [y, z] ∪ [z, x] is a geodesic triangle,
then for all u ∈ [x, y], dist(u, [x, z] ∪ [y, z]) ≤ δ. (Recall that [x, y] denotes some
arbitrary, but fixed, geodesic joining x, y.)
There is a more general definition which applies to non-geodesic spaces. It is
based on the Gromov product
1
(x|y)w := (|x − w| + |y − w| − |x − y|) for points x, y, w in the space .
2
The Gromov product is useful even in geodesic spaces; it can be extended to the
Gromov boundary and then used to define a canonical conformal gauge there.
Roughly speaking, all simply connected manifolds with negative curvature are
Gromov hyperbolic; e.g., every CAT(κ) space with κ < 0. For a specific example,
consider any bounded strictly pseudoconvex domain Ω (with sufficiently smooth
boundary) in complex n-space together with any of the classical hyperbolic dis-
tances h; a result of Balogh and Bonk [BB00] asserts that (Ω, h) is a Gromov
hyperbolic space (with ∂G Ω = ∂Ω, the Euclidean boundary, and canonical con-
formal gauge determined by the Carnot-Carathéodory distance on ∂Ω).

4.B. Gromov Boundary. The Gromov boundary ∂G H of a proper geodesic


Gromov hyperbolic metric space (H, h) is defined as the set of equivalence classes
of geodesic rays, with two such rays being equivalent if they have finite Hausdorff
distance. That is, ∂G H is the geodesic boundary of H; see §2.D. An alterna-
tive description can be given in terms of (equivalent) sequences which converge at
infinity; in particular, this allows us to extend the Gromov product to the bound-
ary (cf. [Väi05a, 5.7] or [BH99, pp. 431-436]). This in turn yields a canonically
defined conformal gauge on the Gromov boundary generated by the quasimetrics
qw,ε (ξ, η) = exp[−ε(ξ|η)w ] for points ξ, η ∈ ∂G H.
For 0 < ε < ε(δ) = log 2/(4δ), each quasimetric qε is bilipschitz equivalent to an
honest metric on the boundary, and all these metric spaces are QS equivalent to
each other; in particular they all generate the same topology on ∂G H, and ∂G H
is compact. See Fact 2.1, [BH99, Proposition 3.21, pp.435-436], [BHK01, p.18].
Uniformity and Hyperbolicity 105

4.C. Connection with Uniform Spaces. Bonk, Heinonen and Koskela estab-
lished the following fundamental connection between uniform spaces and Gromov
hyperbolicity; see [BHK01, Proposition 2.8, Theorem 3.6]
4.1. Fact. The quasihyperbolization (Ω, k) of a locally compact a-uniform space
(Ω, d) is proper, geodesic and δ-hyperbolic where δ = δ(a) = 10000a8 . When
(Ω, d) is bounded, (Ω, k) is roughly κ-starlike with κ = 5000a8 .
˙ of
In fact they also prove that the Gromov boundary ∂G Ω ‘is’ the boundary ∂Ω
Ω in the one-point extension Ω̇ of Ω; see [BHK01, Proposition 3.12]. Moreover, in
the bounded case, the canonical gauge on ∂G Ω is naturally quasisymmetrically
equivalent to the conformal gauge determined by d on ∂Ω. See [Väi05b] for
similar results in the Banach space setting.

5. Uniformization
The celebrated Riemann Mapping Theorem asserts that every simply con-
nected proper subdomain of the plane can be mapped conformally onto the
unit disk, and hence supports a bounded conformal uniformizing metric-density,
namely, ρ = |f ′ | where f is the Riemann map. Koebe proved a similar result for
finitely connected plane domains: any one of these can always be conformally
mapped onto a circle domain (meaning that each boundary component is either
a point or a circle).
In space, every conformal map is (the restriction of) a Möbius transformation,
and thus the only space regions conformally equivalent to a ball are balls and
half-spaces. The problem of determining which space domains are QC equivalent
to a ball has been investigated for more than four decades by now (see [GV65]),
and the most significant result (that I know of) is Väisälä’s characterization in
[Väi89] describing the cylindrical domains (Ω = D × R ⊂ R3 ) which are QC
equivalent to B3 .

5.A. Uniformization Problem. Here we consider the metric space analog of


the Riemann Mapping Problem. We seek to characterize the abstract domains
which can be quasiconformally deformed into a uniform space. We ask the ques-
tion: What are necessary and sufficient conditions for an abstract domain to
support a conformal uniformizing metric-density? Theorem A provides an initial
answer.

5.B. BHK Uniformization. Bonk, Heinonen and Koskela developed a uni-


formization theory, which they call dampening, valid for proper geodesic Gromov
hyperbolic spaces. Their theory produces the following; see [BHK01, Proposition
4.5, Chapter 5]. (They established far more than we mention here:-)
5.1. Fact. Let (H, h) be an unbounded proper geodesic Gromov δ-hyperbolic
space. Fix a base point w ∈ H. For ε > 0 define ρε (x) = exp[−εh(x, w)] and let
Hε = (H, dε ) where dε = hρε . Then:
106 D.A.Herron IWQCMA05

(a) The geodesics in H are double a-cone arcs in Hε with a = a(ε, δ) = e1+8εδ .
(b) There is a constant ε0 = ε0 (δ) such that for all ε ∈ (0, ε0 ],
∀x, y ∈ H , ∀ geodesics [x, y] : ℓε [x, y] ≤ 20 dε (x, y);
here ℓε = ℓρε .

In fact, Hε is always bounded, and thus when ε ≤ ε0 we see that (H, h) has
been deformed, or dampened, (via the natural metric-density ρε ) to a bounded
20-uniform space Hε .
We briefly describe their theory in the special case which concerns us.
We consider a locally compact abstract domain (Ω, d) with the property that
the identity map (Ω, ℓ) → (Ω, d) is a homeomorphism (so the identity (Ω, k) →
(Ω, d) is also a homeomorphism) and such that its quasihyperbolization (Ω, k)
is a Gromov hyperbolic space. According to Fact 5.1, the space (Ω, k) admits a
uniformizing density of the form
ρε (x) := exp[−εk(x, w)];
here w ∈ Ω is a fixed base point and ε > 0 a sufficiently small parameter. More
precisely, when (Ω, k) is δ-hyperbolic and 0 < ε < ε(δ), the quasihyperbolic
geodesics in Ω are 20-uniform arcs in (Ω, dε ). (A careful check of BHK shows
that ε(δ) = [42(5 + 192δ + 1920δ 2 )]−1 ≤ (300 max{1, δ})−2 . :-) Here dε stands
for the distance function obtained by conformally deforming k via the metric-
density ρε . Since k was obtained from the original distance function d via the
quasihyperbolic density 1/d, Ωε = (Ω, dε ) is a conformal deformation of (Ω, d)
via the metric-density
πε (x) := ρε (x)/d(x) = d(x)−1 exp(−εk(x, w)) ;
again, πε will be a uniformizing density when 0 < ε < ε(δ) = (300 max{1, δ})−2 .
In order to determine when πε will be a Harnack or Koebe density, we need
the following information concerning ρε (see [BHK01, (4.4),(4.6),(4.17)]):
ρε (x)
(5.2) e−εk(x,y) ≤ ≤ eεk(x,y) ,
ρε (y)
ρε (x) ρε (x)
(5.3) ≤ dε (x) ≤ (2eεκ − 1) .
eε ε
The first set of inequalities (5.2) hold for all points x, y ∈ Ω and all ε > 0. They
guarantee that the identity map (Ω, k) → (Ω, dε ) is locally bilipschitz, so (Ω, dε )
is locally compact and rectifiably connected. On the other hand, Ωε = (Ω, dε ) is
non-complete, so we can form Ω̄ε , put ∂ε Ω = Ω̄ε \ Ω and let dε (x) = distε (x, ∂ε Ω).
(See [BHK01, pp.27-28]). We find that the leftmost inequality in (5.3) holds for
all x ∈ Ω and any ε > 0. However, in order to obtain the rightmost inequality
in (5.3), we must further require that (Ω, k) be roughly κ-starlike.
Now using (5.2), and obvious inequalities for 1/d, we deduce that
πε (y)
e−2R ≤ e−(ε+1)R ≤ ≤ e(ε+1)R ≤ e2R
πε (x)
Uniformity and Hyperbolicity 107

for all points x, y with k(x, y) ≤ R. According to Lemma 2.12, when (Ω, d) is
locally a-quasiconvex we have
τ B(x) ⊂ Bk (x; R) provided 0 < τ ≤ min{λ, R/[a(1 + R)]}.
Taking R = a (say) and τ = min{λ, 1/2a} we find that for all x ∈ Ω and all
y ∈ τ B(x),
πε (y)
e−2a ≤ e−(ε+1)a ≤ ≤ e(ε+1)a ≤ e2a ;
πε (x)
that is, πε is a Harnack density with constants H = e2a and τ which are inde-
pendent of ε.
Finally, since Ωε is non-complete, we can ask whether or not πε is a Koebe
density. Since πε (x) = ρε (x)/d(x), we see from (5.3) that πε will be a Koebe
density, with constant K = (2eεκ − 1)/ε (assuming εe ≤ 1), provided (Ω, k) is
roughly κ-starlike (and (Ω, d) uniformly locally quasiconvex).
These conditions describing when πε will be a Harnack or Koebe density do
not require that (Ω, k) be Gromov hyperbolic. We record the above information
for later reference; see also Lemma 2.8 and Corollary 5.8. Note too that (Ω, dε )
is bounded with diamε Ωε ≤ 2/ε.
5.4. Lemma. Let (Ω, d) be a locally a-quasiconvex abstract domain and fix a
base point w ∈ Ω. Then for any ε > 0,
πε (x) = ρε (x)/d(x) = d(x)−1 exp(−εk(x, w))
is a Harnack density with constant H = e2a . If in addition (Ω, k) is roughly κ-
starlike, then πε is also a Koebe density, with constant K = max{εe, (2eεκ −1)/ε}.
The above, in conjunction with Lemma 2.11 and Proposition 5.6(b), provides
a one-to-one correspondence between conformal metric-densities on Ω and the
same on Ωε . Here is a precise statement of this.
5.5. Corollary. Suppose (Ω, d, µ) is an abstract domain having bounded geom-
etry and a Gromov hyperbolic roughly starlike quasihyperbolization (Ω, k). Let
Ωε = (Ω, dε , µε ) be the deformation of Ω via the density πε defined just above.
If σ is a conformal density on Ωε , then its pull-back ρ = σ πε is a conformal
density on Ω, and conversely if ρ is a conformal density on Ω, then its push-
forward σ = ρ πε−1 is a conformal density on Ωε . In both cases Ωρ = (Ωε )σ , and
the metric-density parameters depend only on each other and the data associated
with Ω.
5.C. Bounded Geometry and its Consequences. Recall our definition that
an abstract domain (Ω, d, µ) is a locally complete, rectifiably connected, non-
complete metric measure space; these are our standing metric hypotheses. We
say that (Ω, d, µ) has bounded Q-geometry, Q > 1, provided it is both locally
upper Ahlfors Q-regular (see §2.J) and weakly locally Q-Loewner ; this latter
condition means that there exists a positive constant m such that for all x ∈ Ω,
and all non-degenerate disjoint continua E, F in λB(x),
∆(E, F ) ≤ 16 =⇒ mod Q (E, F ; Ω) ≥ m.
108 D.A.Herron IWQCMA05

Any space Ω with the property that Whitney balls λB(x) (with a fixed parame-
ter) are uniformly bilipschitz equivalent to Euclidean balls (in a fixed dimension)
is easily seen to have bounded geometry. Other examples include Riemann-
ian manifolds with Ricci bounded geometry as well as the exotic examples of
Bourdon-Pajot and Laakso for any Q > 1; see [BHK01, Exs.9.7, p.86] and the
references mentioned there. (Note that by Proposition 5.6, (Ω, d, µ) has bounded
geometry if and only if its quasihyperbolization satisfies the condition studied in
[BHK01, Chpt.9].)
The first part of bounded Q-geometry is a necessary condition for Ω to support
any conformal density (at least when Ω is locally quasiconvex). The second part
of bounded Q-geometry, the weak local Loewner criterion, can also be described
in terms of Poincaré inequalities as explained in [BHK01, Proposition 9.4] and
[HK98, §5]; it ensures that there are plenty of curves available (e.g., it gives local
quasiconvexity). However, to substantiate this existence of many curves requires
the use of certain modulus estimates (see Facts 2.15,2.16), and these estimates
in turn require an upper mass condition.
The Loewner part of bounded Q-geometry also performs an essential role in
two other places. First, it is a key player in the proof of Proposition 2.9, which
is the crucial ingredient in the proof of (d) implies (e) in Theorem D. Second,
for uniform Loewner spaces with appropriately ‘thick’ boundaries, the Koebe
condition for a metric-density follows from the Harnack and Ahlfors condition;
this fact is utilized in the proof of Theorem C.
Bounded geometry provides a number of essential properties for our underlying
space.
5.6. Proposition. Let (Ω, d, µ) be an abstract domain having bounded Q-
geometry (with constants M, m, λ). Then:
(a) µ is locally Ahlfors Q-regular.
(b) Ω is locally quasiconvex with constants which depend only on Q, M, m, λ.
(c) Ω is locally Q-Loewner with a control function ψ and parameters κ, ε0 which
depend only on the data Q, M, m, λ associated with Ω.

Here is a useful consequence of the above.


5.7. Corollary. Let ρ be a Harnack Koebe density on an abstract domain
(Ω, d, µ) having bounded Q-geometry. Then Ωρ is locally Ahlfors Q-regular and
locally Q-Loewner with parameters and a control function which depend only on
the data associated with ρ and Ω.

Note that if Ωρ above is also uniform, then it would be globally Loewner by


[BHK01, Theorem 6.4]. We record the following consequence of this observation.
5.8. Corollary. Let (Ω, d, µ) be an abstract domain having bounded Q-geometry
(with constants M, m) and a Gromov δ-hyperbolic roughly κ-starlike quasihy-
perbolization (Ω, k). Then any BHK-uniformization Ωε = (Ω, dε , µε ) of Ω is a
Uniformity and Hyperbolicity 109

bounded locally Ahlfors Q-regular locally Q-Loewner space (hence of bounded Q-


geometry), and even Q-Loewner when ε < ε(δ). Here the various parameters and
control functions depend only on δ, κ, Q, M, m, λ, ε.

5.D. Lifts and Metric Doubling Measures. Here we discuss the ideas be-
hind Theorem C and briefly outline the proof. Recall the notion of a metric
doubling measure discussed near the end of §2.F.
We define the lift ρν of ν via the formula
ν(Σx )1/P
(5.9) ρν (x) := ;
d(x)
here ν is a P -dimensional metric doubling measure on ∂G Ω and, for some fixed
base point w ∈ Ω, the shadow of a point x ∈ Ω is
Σx := {ζ ∈ ∂G Ω : k(x, [w, ζ)) ≤ R}.
It is not hard to see that there are constants R = R(δ, κ, ε) and C = C(δ, κ, ε)
such that
(5.10) Sx := ∂ε Ω ∩ 2Bε (x) ⊂ Σx ⊂ ∂ε Ω ∩ CBε (x);
of course we are using the natural identification of ∂G Ω with ∂ε Ω. Employing
(5.10), the doubling property of ν, and the fact that πε is Koebe, it is straight-
forward to verify that the push-forward of ρν (as defined in (5.9)) via the uni-
formizing density πε gives a density on Ωε which is bilipschitz equivalent to the
density defined via the formula
ν(Sx )1/P
ρ(x) := where Sx := ∂ε Ω ∩ 2Bε (x).
dε (x)
Theorem C now follows from Corollary 5.5 once we verify that ρ is a Borel
Harnack Ahlfors Koebe doubling metric-density on Ωε . (⌣)¨ The first two of
these are easy. The Koebe property follows from Theorem E once we know the
Ahlfors volume growth property. To see that Theorem E can be applied, we
first use Fact 2.2 along with Theorem F to see that the Whitney ball modulus
property holds.
To establish the Ahlfors property we need the doubling property. This in turn
requires the following ‘quasihyperbolic doubling’ result.
5.11. Proposition. Let (Ω, d, µ) be a locally a-quasiconvex locally Ahlfors Q-
regular abstract domain. Then its quasihyperbolization (Ω, k) is locally doubling
in the sense that if Σ ⊂ Ω is a set of points satisfying
0 < t ≤ k(x, y) ≤ T < ∞ for all x, y ∈ Σ, x 6= y,
then the cardinality of Σ is bounded by
#Σ ≤ 2M 2 8Q (2M 2 24Q )8aT /λt .
Here M is the local regularity constant and λ the Whitney ball constant.
110 D.A.Herron IWQCMA05

Here is an interesting application of Theorems C and D which provides a char-


acterization of doubling conformal densities in terms of metric doubling measures.
We say that a metric-density ρ on Ω is induced by a metric doubling measure ν
on ∂G Ω if there exists a constant C ≥ 1 such that
C −1 ρν (x) ≤ ρ(x) ≤ Cρν (x) for µ-a.e. x ∈ Ω.
5.12. Theorem. Assume the basic minimal hypotheses, that the Gromov bound-
ary of Ω is uniformly perfect, and P < Q. A metric-density ρ on Ω is a doubling
conformal density, with (∂ρ Ω, dρ ) Ahlfors P -regular, if and only if ρ is induced
by some P -dimensional metric doubling measure on ∂G Ω.

5.E. Volume Growth Problem. Here we briefly outline the proof of Theo-
rem A. Recall that this result provides our answer to the problem of deciding
when there exists a uniformizing conformal density (so, in particular, the asso-
ciated measure (2.7) should have Ahlfors regular volume growth). The necessity
in this result follows from Fact 4.1 along with Fact 2.20. The real work involved
is in establishing the sufficiency.
The first major step is to prove Theorem D. Following [BHR01], we say that
a metric-density ρ, on a uniform space (Ω, d, µ), is doubling provided µρ is a
doubling measure on (Ω, d); i.e., there exists a constant D = Dρ such that for all
x ∈ Ω,
(D) µρ [B(x; 2r)] ≤ D µρ [B(x; r)] for all r > 0.
When Ω is not uniform, the above doubling condition may fail to hold even if ρ
is ‘nice’; e.g., consider ρ = |f ′ | on Ω = {x + iy : |y| < 1} in the complex plane,
where f is a conformal homeomorphism of Ω onto the unit disk. To compensate
for this we employ the following definition: a conformal density ρ, on an abstract
domain (Ω, d, µ) with Gromov hyperbolic quasihyperbolization, is doubling if its
push-forward is doubling on some BHK-uniformization Ωε of Ω. (Since all such
spaces Ωε are QS equivalent, and doubling measures can be defined for conformal
gauges, there is no ambiguity here. See §2.C for a discussion of conformal gauges,
and also the very last paragraph of §2.F.)
Notice that the doubling condition (D) uses d-balls but µρ -measure and thus
interweaves the measure properties of the deformed space with the metric prop-
erties of the original (or uniformized) space.
Our proof of Theorem D requires the Gehring-Hayman Inequality (Theo-
rem B), Corollary 2.10, and the following two results. First we give a necessary
condition for a metric-density to be doubling.
5.13. Proposition. Let ρ be a Harnack density on an a-uniform locally Ahlfors
Q-regular space (Ω, d, µ) (with constants H, M, λ). Suppose that ρ is also doubling
on Ω (with constant D). Then quasihyperbolic geodesics in Ω are double c-cone
arcs in Ωρ where c = c(D, H, M, Q, a, λ). If in addition Ω is bounded, then so is
Ωρ and ∂Ω ⊂ ∂ρ Ω, with equality holding when Ω is also locally Q-Loewner.

Next we state a sufficient condition for a metric-density to be doubling.


Uniformity and Hyperbolicity 111

5.14. Proposition. Let ρ be a conformal density on an a-uniform locally


lower Ahlfors Q-regular space (Ω, d, µ) (with constants H, A, K, M, λ). Sup-
pose that quasihyperbolic geodesics in Ω are double c-cone arcs in Ωρ . Then
Ωρ is Ahlfors Q-regular with a parameter which depends only on c and the data
H, A, K, M, Q, a, λ associated with ρ and Ω. If in addition Ω and Ωρ are both
bounded, then ρ is doubling on Ω with a parameter which depends only on the
aforementioned data and the quantity q given in the proof of Lemma 3.3.

The second major step in the proof of Theorem A is to establish Theorem C.


Its proof is outlined above in §5.D. That done, we use Theorem C to obtain a
doubling conformal density, which by Theorem D is also bounded and uniformiz-
ing.

6. Characterizations of Uniform Spaces


Here we mention a number of characterizations for uniform spaces and uni-
form domains. In [Väi88] Väisälä provides a complete description of the various
possible twisted double cone conditions (which he calls length cigars, diameter
cigars, distance cigars, and Möbius cigars). The work [Mar80] of Martio should
also be mentioned.

6.A. Metric Characterizations. We have already mentioned that uniform


spaces are precisely the abstract domains in which the quasihyperbolic metric is
bilipschitz equivalent to the so-called j metric; see Fact 3.2.
It turns out that the following seemingly weaker quasihyperbolic metric con-
dition also characterizes uniform spaces. For uniform subdomains of Banach
spaces this result is due to Väisälä [Väi91, 6.16, 6.17]. Here we write
|x − y|
r(x, y) :=
d(x) ∧ d(y)
to denote the so-called relative distance between x, y. See [BH07] for the following
version.
6.1. Theorem. A locally quasiconvex abstract domain is uniform if and only if
there is a homeomorphism ϑ : [0, ∞) → [0, ∞) satisfying lim supt→∞ ϑ(t)/t < 1,
and such that for all points x, y, k(x, y) ≤ ϑ (r(x, y)). The uniformity constant
depends only on ϑ, and conversely in an a-uniform space, one can always take
ϑ(t) = b log(1 + t) with b = b(a).

6.B. Gromov Boundary Characterizations. We call Ω ( Rn a Gromov


domain if its quasihyperbolization (Ω, k) is Gromov hyperbolic. Bonk, Heinonen
and Koskela corroborated the following [BHK01, Proposition 7.12].
6.2. Fact. A Gromov domain in Euclidean space is uniform if and only if it is
linearly locally connected.

They utilized the above to establish the following [BHK01, Theorem 7.11].
112 D.A.Herron IWQCMA05

6.3. Fact. A (bounded) Gromov domain in Euclidean space is uniform if and


only if the canonical gauge on the Gromov boundary is quasisymmetrically equiv-
alent to the Euclidean gauge on the Euclidean boundary.

The careful reader will recognize that the Bonk-Heinonen-Koskela results were
established for regions on the sphere (i.e., using the spherical metric). Results
of Balogh and Buckley [BB06] are useful in this regard.
Väisälä has recently proven a Banach space analog of the above result; see
[Väi05b]. Along with providing a dimension free version of this result, he also
considers arbitrary domains (not just bounded) and replaces QS equivalence with
QM equivalence. In addition, he provides an example of a Gromov hyperbolic
domain which is LLC but not uniform.

6.C. Characterizations using QC Maps. It is evident that bilipschitz ho-


meomorphisms map uniform spaces to uniform spaces. This also holds true for
QS and QM maps of uniform domains in Euclidean space and in Banach spaces
[Väi99, Theorem 10.22], but not in the general metric space setting, and not for
QC maps. On the other hand, according to [BKR98, 2.4], the average derivative
of a quasiconformal map f : Bn → Ω ⊂ Rn is a conformal metric density on Bn
(a uniform n-Loewner n-regular space). Thus we can appeal to Theorem D and
read off a number of conditions which characterize when Ω will be uniform.

6.D. Capacity Conditions. It is known that given 0 < λ ≤ 1/2, there exists
a constant c = c(λ, n) > 0 such that
∀ x, y ∈ Ω : k(x, y) ≥ 2 =⇒ mod(λB̄(x), λB̄(y); D) ≥ c/k(x, y)n−1 ;
this is valid for any proper subdomain Ω of Rn . To prove it, one starts by us-
ing Lemma 2.14 to select an appropriate cover of any quasihyperbolic geodesic
joining x, y, and then a standard application of the Poincaré inequality applied
to adjacent balls leads to the asserted inequality. See the proof of [HK96, Theo-
rem 6.1].
Let C > 0 and 0 < λ ≤ 1/2. A proper subdomain Ω of Rn is a (C, λ)-k-cap
domain provided
∀ x, y ∈ D : k(x, y) ≥ 2 =⇒ mod(λB̄(x), λB̄(y); D) ≤ C/k(x, y)n−1 .
Thus in a k-cap domain D, we have mod(λB̄(x), λB̄(y); D) ≃ k(x, y)1−n for
points with k(x, y) ≥ 2, with constants of comparison dependent only on λ, n,
and the k-cap parameter.
This is the two-sided version of a condition introduced by Buckley in [Buc04] to
study quasiconformal images of domains which satisfy a quasihyperbolic bound-
ary condition. As explained on p.26 of that paper, a (C, λ)-k-cap condition
implies a (C ′ , λ′ )-k-cap condition for some C ′ = C ′ (C, λ, λ′ , n). We mainly con-
sider the case λ = 1/2, and refer to a (C, 1/2)-k-cap domain simply as a C-k-cap
domain.
Uniformity and Hyperbolicity 113

Every uniform domain in Rn is a k-cap domain, and the class of k-cap do-
mains is invariant under quasiconformal mappings (with a quantitative change
of parameter C). For proofs of these statements see [Buc04].
Recently we established the following characterization for uniform domains in
Euclidean space; see [BH06, Theorem 3.5].
6.4. Theorem. A proper subdomain of Rn is uniform if and only if it is both
QEDwb and a k-cap domain.

This result is quantitative.

6.E. LLC and Slice Conditions. By utilizing certain slice conditions, Balogh
and Buckley [BB03] established a number of geometric characterizations for Gro-
mov hyperbolic spaces. Here we mention the following new characterization of
uniform spaces; see [BH07]
6.5. Theorem. An abstract domain is uniform and LEC if and only if it is
quasiconvex, LLC2 with respect to arcs, and satisfies a weak slice condition.

These implications are quantitative.


We have modified the usual LLC conditions (see §2.E) by requiring that the
points in question be joinable by rectifiable arcs (rather than just by continua).
Every uniform domain in Rn is LLC. In fact every uniform space is quasiconvex
and thus LLC1 with respect to arcs. However, uniform spaces need not be LLC2 ;
e.g., an ‘asterik’ type space (the disjoint union of a point and a bunch of line
segments or rays, with its intrinsic length distance) may be uniform but not
LLC2 . We say that a locally complete metric space is locally externally connected,
abbreviated LEC, provided there is a constant c ≥ 1 such that the (LLC2 )
condition holds for all points x ∈ Ω and all r ∈ (0, d(x)/c).

References
[BB00] Z.M. Balogh and M. Bonk, Gromov hyperbolicty and the Kobayashi metric on strictly
pseudoconvex domains, Commet. Math. Helv. 75 (2000), no. 3, 504–533.
[BB03] Z.M. Balogh and S.M. Buckley, Geometric characterizations of Gromov hyperbolicty,
Invent. Math. 153 (2003), 261–301.
[BB06] , Sphericalization and flattening, Conform. Geom. Dyn, to appear (2006) .
[Bon96] M. Bonk, Quasi-geodesic segments and Gromov hyperbolic spaces, Geometriae Dedi-
cata 62 (1996), 281–298.
[BHK01] M. Bonk, J. Heinonen, and P. Koskela, Uniformizing Gromov hyperbolic spaces,
Astérisque. 270 (2001), 1–99.
[BHR01] M. Bonk, J. Heinonen, and S. Rohde, Doubling conformal densities, J. reine angew.
Math. 541 (2001), 117–141.
[BKR98] M. Bonk, P. Koskela, and S. Rohde, Conformal metrics on the unit ball in Euclidean
space, Proc. London Math. Soc. 77 (1998), no. 3, 635–664.
[BS00] M. Bonk and O. Schramm, Embeddings of Gromov hyperbolic spaces, Geom. Funct.
Anal. 10 (2000), 266–306.
[BH99] M.R. Bridson and A. Haefliger, Metric spaces of non-positive curvature, Springer-
Verlag, Berlin, 1999.
114 D.A.Herron IWQCMA05

[Buc03] S. Buckley, Slice conditions and their applications, Future Trends In Geometric Func-
tion Theory (Univ. Jyväskylä), vol. 92, Rep. Univ. Jyväskylä Dept. Math. Stat., 2003,
RNC Workshop held in Jyväskylä, June 15-18, 2003, pp. 63–76.
[Buc04] S.M. Buckley, Quasiconfomal images of Hölder domains, Ann. Acad. Sci. Fenn. Ser.
Math. 29 (2004), 21–42.
[BH06] S. Buckley and D.A. Herron, Uniform domains and capacity, Israel J. Math, to
appear (2006).
[BH07] , Uniform and weak slice spaces, in preparation (2007).
[BO99] S.M. Buckley and J. O’Shea, Weighted Trudinger-type inequalities, Indiana Univ.
Math. J. 48 (1999), 85–114.
[BS03] S.M. Buckley and A. Stanoyevitch, Distinguishing properties of weak slice conditions,
Conform. Geom. Dyn. 7 (2003), 49–75.
[BBI01] D. Burago, Y. Burago, and S. Ivanov, A course in metric geometry, American Math-
ematical Society, Providence, RI, 2001.
[CGN00] L. Capogna, N. Garofalo, and D-M Nhieu, Examples of uniform and NTA domains
in Carnot groups, Proceedings on Analysis and Geometry (Novosibirsk), Izdat. Ross.
Akad. Nauk Sib. Otd. Inst. Mat., 2000, (Novosibirsk Akad., 1999), pp. 103–121.
[CT95] L. Capogna and P. Tang, Uniform domains and quasiconformal mappings on the
Heisenberg group, Manuscripta Math. 86 (1995), no. 3, 267–281.
[DS97] G. David and S. Semmes, Fractured fractals and broken dreams:self-similar geometry
through metric and measure, Oxford Lecture Series in Mathematics and its Applica-
tions, vol. 7, Oxford University Press, Oxford, 1997.
[Geh82] F.W. Gehring, Characteristic properties of quasidisks, Les Presses de l’Université de
Montréal, Montréal, Quebec, 1982.
[Geh87] , Uniform domains and the ubiquitous quasidisk, Jahresber. Deutsch. Math.-
Verein 89 (1987), 88–103.
[GH62] F.W. Gehring and W.K. Hayman, An inequality in the theory of conformal mapping,
J. Math. Pures Appl. 41 (1962), no. 9, 353–361.
[GM85] F.W. Gehring and O. Martio, Quasiextremal distance domains and extension of qua-
siconformal mappings, J. Analyse Math. 45 (1985), 181–206.
[GO79] F.W. Gehring and B.G. Osgood, Uniform domains and the quasi-hyperbolic metric,
J. Analyse Math. 36 (1979), 50–74.
[GP76] F.W. Gehring and B.P. Palka, Quasiconformally homogeneous domains, J. Analyse
Math. 30 (1976), 172–199.
[GV65] F.W. Gehring and J. Väisälä, The coefficients of quasiconformality of domains in
space, Acta Math. 114 (1965), 1–70.
[Gre01] A.V. Greshnov, On uniform and NTA-domains on Carnot groups, Sibirsk. Mat. Zh.
42 (2001), no. 5, 1018–1035.
[Gro99] M. Gromov, Metric structures for Riemannian and non-Riemannian spaces, Metric
structures for Riemannian and non-Riemannian spaces, Progress in Mathematics,
vol. 152, Birkhäuser, Boston, 1999.
[Häs06] P. Hästö, Gromov hyperbolicity of the jG and j̃G metrics, Proc. Amer. Math. Soc.
134 (2006), 1137–1142.
[Hei01] J. Heinonen, Lectures on analysis on metric spaces, Springer-Verlag, New York, 2001.
[HK95] J. Heinonen and P. Koskela, Definitions of quasiconformality, Invent. Math. 120
(1995), 61–79.
[HK98] J. Heinonen and P. Koskela, Quasiconformal maps in metric spaces with controlled
geometry, Acta. Math. 181 (1998), 1–61.
[HN94] J. Heinonen and R. Nakki, Quasiconformal distortion on arcs, J. d’Analyse Math.
63 (1994), 19–53.
[HR93] J. Heinonen and S. Rohde, The Gehring-Hayman inequality for quasihyperbolic
geodesics, Math. Proc. Camb. Phil. Soc. 114 (1993), 393–405.
Uniformity and Hyperbolicity 115

[Her04] D.A. Herron, Conformal deformations of uniform Loewner spaces, Math. Proc.
Camb. Phil. Soc. 136 (2004), 325–360.
[Her06] , Quasiconformal deformations and volume growth, Proc. London Math. Soc.
92 (2006), 161–199.
[HK96] D.A. Herron and P. Koskela, Conformal capacity and the quasihyperbolic metric,
Indiana Univ. Math. J. 45 (1996), no. 2, 333–359.
[Joh61] F. John, Rotation and strain, Comm. Pure Appl. Math. 14 (1961), 391–413.
[Jon81] P.W. Jones, Quasiconformal mappings and extendability of functions in Sobolev
spaces, Acta Math. 147 (1981), 71–88.
[Kos07] P. Koskela, Lectures on quasiconformal mappings, Rep. Univ. Jyväskylä Dept. Math.
Stat., in preparation (2007).
[Luu98] J. Luukkainen, Assouad dimension: antifractal metrization, porous sets, and homo-
geneous measures, J. Korean Math. Soc. 35 (1998), no. 1, 23–76.
[Mar80] O. Martio, Definitions for uniform domains, Ann. Acad. Sci. Fenn. Ser. A I Math.
5 (1980), no. 1, 197–205.
[MS79] O. Martio and J. Sarvas, Injectivity theorems in plane and space, Ann. Acad. Sci.
Fenn. Ser. A I Math. 4 (1978/79), no. 2, 383–401.
[NV91] R. Näkki and J. Väisälä, John disks, Exposition. Math. 9 (1991), 3–43.
[Sem99] S. Semmes, Metric spaces and mappings seen at many scales (Appendix B), Metric
Structures in Riemannian and non-Riemannian spaces, Progress in Mathematics, vol.
152, Birkhäuser, Boston, 1999, pp. 401–518.
[Sem01] , Some novel types of fractal geometry, Oxford Mathematical Monographs,
Oxford University Press, Oxford, 2001.
[TV80] P. Tukia and J. Väisälä, Quasymmetric embeddings of metric spaces, Ann. Acad. Sci.
Fenn. Ser. A I Math. 5 (1980), 97–114.
[Tys03] J. Tyson, Quasiconformal maps on metric spaces: Questions and conjectures, Future
Trends in Geometric Function Theory (Jyväskylä, Finland), no. 92, Dept. Math.
Stat., Univ. Jyväskylä, 2003, RNC Workshop held in Jyväskylä, June 15-18, 2003,
pp. 249–262.
[Väi71] J. Väisälä, Lectures on n-dimensional quasiconformal mappings, Lecture Notes in
Math., No. 229 (Berlin), Springer-Verlag, 1971.
[Väi85] , Quasimöbius maps, J. Analyse Math. 44 (1984/85), 218–234.
[Väi88] , Uniform domains, Tôhoku Math. J. 40 (1988), 101–118.
[Väi89] , Quasiconformal maps of cylindrical domains, Acta Math. 162 (1989), 201–
225.
[Väi91] , Free quasiconformality in Banach spaces II, Ann. Acad. Sci. Fenn. Math.
16 (1991), 255–310.
[Väi98] , Relatively and inner uniform domains, Conformal Geometry and Dynamics
2 (1998), 56–88.
[Väi99] , The free quasiworld. Freely quasiconformal and related maps in Banach
spaces., Quasiconformal Geometry and Dynamics (Lublin, 1996) (Warsaw), vol. 48,
Inst. Math., Polish of Academy Sciences, Banach Center Publ., 1999, pp. 55–118.
[Väi05a] , Gromov hyperbolic spaces, Exposition. Math. 23 (2005), 187–231.
[Väi05b] , Hyperbolic and uniform domains in Banach spaces, Ann. Acad. Sci. Fenn.
Math. 30 (2005), 261–302.
[Vuo88] M. Vuorinen, Conformal geometry and quasiregular mappings, Lecture Notes in
Math., No. 1319 (Berlin), Springer Verlag, 1988.

David A Herron E-mail: david.herron@math.uc.edu


Address: Department of Mathematics, University of Cincinnati, OH 45221, USA
Proceedings of the International Workshop on Quasiconformal
Mappings and their Applications (IWQCMA05)

p-Laplace operator, quasiregular mappings, and


Picard-type theorems

Ilkka Holopainen and Pekka Pankka

Abstract. We describe the role of p-harmonic functions and forms in the


theory of quasiregular mappings.

Keywords. Quasiregular mapping, p-harmonic function, p-harmonic form,


conformal capacity.
2000 MSC. Primary 58J60, 30C65; Secondary 53C20, 31C12, 35J60.

Contents

1. Introduction 117
2. A-harmonic functions 120
3. Morphism property and its consequences 123
3.1. Sketch of the proof of Reshetnyak’s theorem 125
Further properties of f 127
4. Modulus and capacity inequalities 128
5. Liouville-type results for A-harmonic functions 130
6. Liouville-type results for quasiregular mappings 136
7. Picard-type theorems 137
8. Quasiregular mappings, p-harmonic forms, and de Rham cohomology 145
References 148

1. Introduction
In this survey we emphasize the importance of the p-Laplace operator as a
tool to prove basic properties of quasiregular mappings, as well as Liouville-
and Picard-type results for quasiregular mappings between given Riemannian
manifolds. Quasiregular mappings were introduced by Reshetnyak in the mid
sixties in a series of papers; see e.g. [36], [37], and [38]. An interest in studying
these mappings arises from a question about the existence of a geometric function
theory in real dimensions n ≥ 3 generalizing the theory of holomorphic functions
C → C.
118 Ilkka Holopainen and Pekka Pankka IWQCMA05

Definition 1.1. A continuous mapping f : U → Rn of a domain U ⊂ Rn is


called quasiregular (or a mapping of bounded distortion) if
1,n
(1) f ∈ Wloc (U ; Rn ), and
(2) there exists a constant K ≥ 1 such that
|f ′ (x)|n ≤ K Jf (x) for a.e. x ∈ U.

The condition (1) means that the coordinate functions of f belong to the
1,n
local Sobolev space Wloc (U ) consisting of locally n-integrable functions whose
distributional (first) partial derivatives are also locally n-integrable. In Condi-
tion (2) f ′ (x) denotes the formal derivative of f at x, i.e. the n × n matrix
Dj fi (x) defined by the partial derivatives of the coordinate functions fi of f .
Furthermore,
|f ′ (x)| = max|f ′ (x)h|
|h|=1

is the operator norm of f (x) and Jf (x) = det f ′ (x) is the Jacobian determinant

of f at x. They exist a.e. by (1). The smallest possible K in Condition (2) is


the outer dilatation KO (f ) of f . If f is quasiregular, then
Jf (x) ≤ K ′ ℓ(f ′ (x))n a.e.
for some constant K ′ ≥ 1, where
ℓ(f ′ (x)) = min |f ′ (x)h|.
|h|=1

The smallest possible K ′ is the inner dilatation KI (f ) of f . It is easy to


see by linear algebra that KO (f ) ≤ KI (f )n−1 and KI (f ) ≤ KO (f )n−1 . If
max{KO (f ), KI (f )} ≤ K, f is called K-quasiregular .
To motivate the above definition, let us consider a holomorphic function
f : U → C, where U ⊂ C is an open set. We write f as a mapping f =
(u, v) : U → R2 , U ⊂ R2 ,

f (x, y) = u(x, y), v(x, y) .
Then u and v are harmonic real-valued functions in U and they satisfy the
Cauchy-Riemann system of equations

D1 u = D2 v
D2 u = −D1 v,
where D1 = ∂/∂x, D2 = ∂/∂y. For every (x, y) ∈ U , the differential f ′ (x, y) : R2 →
R2 is a linear map whose matrix (with respect to the standard basis of the plane)
is    
D1 u D 2 u D1 u D 2 u
= .
D1 v D 2 v −D2 u D1 u
Hence
(1.1) |f ′ (x, y)|2 = det f ′ (x, y).
p-Laplace operator, quasiregular mappings, and Picard-type theorems 119

The first trial definition for mappings f : U → Rn of a domain U ⊂ Rn , sharing


some geometric and topological properties of holomorphic functions, could be
mappings satisfying a condition
(1.2) |f ′ (x)|n = Jf (x), x ∈ U.
However, it has turned out that, in dimensions n ≥ 3, a mapping f : U → Rn
1,n
belonging to the Sobolev space Wloc (U ; Rn ) and satisfying (1.2) for a.e. x ∈ U is
either constant or a restriction of a Möbius map. This is the so-called generalized
Liouville theorem due to Gehring [12] and Reshetnyak [38]; see also the thorough
discussion in [29].
Next candidate for the definition is obtained by replacing the equality (1.2)
by a weaker condition
(1.3) |f ′ (x)|n ≤ K Jf (x) a.e. x ∈ U,
where K ≥ 1 is a constant. Note that Jf (x) ≤ |f ′ (x)|n holds for a.e. x ∈ U . Now
there remains a question on the regularity assumption of such mapping f. Again
there is some rigidity in dimensions n ≥ 3. Indeed, if a mapping f satisfying
(1.3) is non-constant and smooth enough (more precisely, if f ∈ C k , with k = 2
for n ≥ 4 and k = 3 for n = 3), then f is a local homeomorphism. Furthermore,
it then follows from a theorem of Zorich that such mapping f : Rn → Rn is
necessarily a homeomorphism, for n ≥ 3; see [47]. We would also like a class of
maps satisfying (1.3), with fixed K, to be closed under local uniform convergence.
In order to obtain a rich enough class of mappings, it is thus necessary to weaken
the regularity assumption from C k -smoothness. See [15], [5], and [32] for recent
developments regarding smoothness and branching of quasiregular mappings.
The basic analytic and topological properties of quasiregular mappings are
listed in the following theorem by Reshetnyak; see [39], [41].
Theorem 1.2 (Reshetnyak’s theorem). Let U ⊂ Rn be a domain and let f : U →
Rn be quasiregular. Then
(1) f is differentiable a.e. and
(2) f is either constant or it is discrete, open, and sense-preserving.

Recall that a map g : X → Y between topological spaces X and Y is discrete


if the preimage g −1 (y) of every y ∈ Y is a discrete subset of X and that g is open
if gU is open for every open U ⊂ X. We also remark that a continuous discrete
and open map g : X → Y is called a branched covering.
To say that f : U → Rn is sense-preserving means that the local degree
µ(y, f, D) is positive for all domains D ⋐ U and for all y ∈ f D \ f ∂D. The
local degree is an integer that tells, roughly speaking, how many times f wraps
D around y. It can be defined, for example, by using cohomology groups with
compact support. For the basic properties of the local degree, we refer to [41,
Proposition I.4.4]; see also [11], [35], and [45]. For example, if f is differentiable
at x0 with Jf (x0 ) 6= 0, then µ(f (x0 ), f, D) = sign Jf (x0 ) for sufficiently small
120 Ilkka Holopainen and Pekka Pankka IWQCMA05

connected neighborhoods D of x0 . Another useful property is the following ho-


motopy invariance: If f and g are homotopic via a homotopy ht , h0 = f, h1 = g,
such that y ∈ ht D \ ht ∂D for every t ∈ [0, 1], then µ(y, f, D) = µ(y, g, D).

f
y

D
f ∂D
fD

The definition of quasiregular mappings extends easily to the case of continu-


ous mappings f : M → N, where M and N are connected oriented Riemannian
n-manifolds.
Definition 1.3. A continuous mapping f : M → N is quasiregular (or a mapping
1,n
of bounded distortion) if it belongs to the Sobolev space Wloc (M ; N ) and there
exists a constant K ≥ 1 such that
(1.4) |Tx f |n ≤ KJf (x) for a.e. x ∈ M.

Here again Tx f : Tx M → Tf (x) N is the formal differential (or the tangent


map) of f at x, |Tx f | is the operator norm of Tx f , and Jf (x) is the Jacobian
determinant of f at x uniquely defined by (f ∗ volN )x = J(x, f )(volM )x almost
everywhere. Note that Tx f can be defined for a.e. x by using partial derivatives
of local expressions of f at x. The geometric interpretation of (1.4) is that Tx f
maps balls of Tx M either to ellipsoids with controlled ratios of the semi-axes or
Tx f is the constant linear map.
Tx M
Tx f Tf (x) N

M N

We assume from now on that M and N are connected oriented Riemannian


n-manifolds.

2. A-harmonic functions
It is well-known that the composition u◦f of a holomorphic function f : U → C
and a harmonic function u : f U → R is a harmonic function in U . In other words,
holomorphic functions are harmonic morphisms. Quasiregular mappings have a
somewhat similar morphism property: If f : U → Rn is quasiregular and u is an
n-harmonic function in a neighborhood of f U , then u◦f is a so-called A-harmonic
p-Laplace operator, quasiregular mappings, and Picard-type theorems 121

function in U . In this section we introduce the notion of A-harmonic functions


and recall some of their basic properties that are relevant for this survey.
We denote by h·, ·i the Riemannian metric of M . Recall that the gradient of
a smooth function u : M → R is the vector field ∇u such that
h∇u(x), hi = du(x)h
for every x ∈ M and h ∈ Tx M.
The divergence of a smooth vector field V can be defined as a function
div V : M → R satisfying
LV ω = (div V ) ω,
where ω = volM is the (Riemannian) volume form and
(αt )∗ ω − ω
LV ω = lim
t→0 t
is the Lie derivative of ω with respect to V, and α is the flow of V.
V

αt (x)

x α = flow of V

We say that a vector field ∇u ∈ L1loc (M ) is a weak gradient of u ∈ L1loc (M ) if


Z Z
(2.1) h∇u, V i = − u div V
M M

for all vector fields V ∈ C0∞ (M ).


Conversely, a function div V ∈ L1loc (M ) is
a weak divergence of a (locally
R integrable) vector field V if (2.1) holds for all
u ∈ C0∞ (M ). Note that M div Y = 0 if Y is a smooth vector field in M with
compact support.
We define the Sobolev space W 1,p (M ) and its norm as
W 1,p (M ) = {u ∈ Lp (M ) : weak gradient ∇u ∈ Lp (M )}, 1 ≤ p < ∞,
kuk1,p = kukp + k|∇u|kp .

Let G ⊂ M be open. Suppose that for a.e. x ∈ G we are given a continuous


map
Ax : Tx M → Tx M
such that the map x 7→ Ax (X) is a measurable vector field whenever X is.
Suppose that there are constants 1 < p < ∞ and 0 < α ≤ β < ∞ such that
hAx (h), hi ≥ α|h|p
and
|Ax (h)| ≤ β|h|p−1
122 Ilkka Holopainen and Pekka Pankka IWQCMA05

for a.e. x ∈ G and for all h ∈ Tx M. In addition, we assume that for a.e. x ∈ G
hAx (h) − Ax (k), h − ki > 0
whenever h 6= k, and
Ax (λh) = λ|λ|p−2 Ax (h)
whenever λ ∈ R \ {0}.
1,p
A function u ∈ Wloc (G) is called a (weak) solution of the equation
(2.2) − div Ax (∇u) = 0
in G if Z
hAx (∇u), ∇ϕi = 0
G
for all ϕ ∈ C0∞ (G). Continuous solutions of (2.2) are called A-harmonic functions
(of type p). By the fundamental work of Serrin [43], every solution of (2.2) has
a continuous representative. In the special case Ax (h) = |h|p−2 h, A-harmonic
functions are called p-harmonic and, in particular, if p = 2, we obtain the usual
harmonic functions. The conformally invariant case p = n = the dimension of
M is important in the sequel. In this case p-harmonic functions are called, of
course, n-harmonic functions.
1,p
A function u ∈ Wloc (G) is a subsolution of (2.2) in G if
− div Ax (∇u) ≤ 0
weakly in G, that is Z
hAx (∇u), ∇ϕi ≤ 0
G
for all non-negative ϕ ∈ C0∞ (G). A function v is called supersolution of (2.2) if
−v is a subsolution. The proofs of the following two basic estimates are straight-
forward once the appropriate test function is found. Therefore we just give the
test function and leave the details to readers.
Lemma 2.1 (Caccioppoli inequality). Let u be a positive solution of (2.2) (for
a given fixed p) in G and let v = uq/p , where q ∈ R \ {0, p − 1}. Then
Z  p Z
p p β|q|
(2.3) η |∇v| ≤ v p |∇η|p
G α|q − p + 1| G

for every non-negative η ∈ C0∞ (G).

Proof. Write κ = q − p + 1 and use ϕ = uκ η p as a test function.


Remark 2.2. In fact, the estimate (2.3) holds for positive supersolutions if
q < p − 1, q 6= 0, and for positive subsolutions if q > p − 1.

The excluded case q = 0 above corresponds to the following logarithmic Cac-


cioppoli inequality.
p-Laplace operator, quasiregular mappings, and Picard-type theorems 123

Lemma 2.3 (Logarithmic Caccioppoli inequality). Let u be a positive superso-


lution of (2.2) (for a given fixed p) in G and let C ⊂ G be compact. Then
Z Z
(2.4) |∇ log u| ≤ c |∇η|p
p
C G
for all η ∈ C0∞ (G), with η|C ≥ 1, where c = c(p, β/α).

Proof. Choose ϕ = η p u1−p as a test function.

These two lemmas together with the Sobolev and Poincaré inequalities are
used in proving Harnack’s inequality for non-negative A-harmonic functions by
the familiar Moser iteration scheme. In the following |A| denotes the volume of
a measurable set A ⊂ M.
Theorem 2.4 (Harnack’s inequality). Let M be a complete Riemannian mani-
fold and suppose that there are positive constants R0 , C, and τ ≥ 1 such that a
volume doubling property
(2.5) |B(x, 2r)| ≤ C |B(x, r)|
holds for all x ∈ M and 0 < r ≤ R0 , and that M admits a weak (1, p)-Poincaré
inequality
 1/p
Z Z
(2.6) |v − vB | ≤ C r  |∇v|p 
B τB

for all balls B = B(x, r) ⊂ M, with τ B = B(x, τ r) and 0 < r ≤ R0 , and for all
functions v ∈ C ∞ (B). Then there is a constant c such that
(2.7) sup u ≤ c inf u
B(x,r) B(x,r)

whenever u is a non-negative A-harmonic function in a ball B(x, 2r), with 0 <


r ≤ R0 .

In particular, if the volume doubling condition (2.5) and the Poincaré inequal-
ity (2.6) hold globally, that is, without any bound on the radius r, we obtain a
global Harnack inequality. We refer to [18], [9], and [16] for proofs of the Harnack
inequality.

3. Morphism property and its consequences


The very first step in developing the theory of quasiregular mappings is to
prove, by direct computation, that quasiregular mappings have the morphism
property in a special case where the n-harmonic function is smooth enough.
Theorem 3.1. Let f : M → N be a quasiregular mapping (with a constant K)
and let u ∈ C 2 (N ) be n-harmonic. Then v = u ◦ f is A-harmonic (of type n) in
M, with
n
(3.1) Ax (h) = hGx h, hi 2 −1 Gx h,
124 Ilkka Holopainen and Pekka Pankka IWQCMA05

where Gx : Tx M → Tx M is given by
(
Jf (x)2/n Tx f −1 (Tx f −1 )T h, if Jf (x) exists and is positive,
Gx h =
h, otherwise.
The constants α and β for A depend only on n and K.

Proof. Let us first write the proof formally and then discuss the steps in more
detail. In the sequel ω stands for the volume forms in M and N. Let V ∈ C 1 (M )
be the vector field V = |∇u|n−2 ∇u. Since u is n-harmonic and C 2 -smooth, we
have div V = 0. By Cartan’s formula we obtain
d(V y ω) = d(V y ω) + V y (dω) = LV ω = (div V ) ω = 0
since dω = 0. Here Xy η is the contraction of a differential form η by a vector
field X. Thus, for instance, V y ω is the (n − 1)-form
V y ω(·, . . . , ·) = ω(V, ·, . . . , ·).
| {z } | {z }
n−1 n−1

Hence
a.e.
(3.2) df ∗ (V y ω) = f ∗ d(V y ω) = 0.
On the other hand, we have a.e. in M
(3.3) f ∗ (V y ω) = W y f ∗ ω = W y (Jf ω) = Jf W y ω,
where W is a vector field that will be specified later (roughly speaking, f∗ W =
V ). We obtain
(3.4) d(Jf W y ω) = 0,
or equivalently
(3.5) div(Jf W ) = 0
which can be written as
(3.6) div Ax (∇v) = 0,
where A is as in the claim.
Some explanations are in order. When writing
a.e.
f ∗ d(V y ω) = 0,
we mean that for a.e. x ∈ U and for all vectors v1 , v2 , . . . , vn ∈ Tx M
f ∗ d(V y ω)(v1 , v2 , . . . , vn ) = d(V y ω)(f∗ v1 , f∗ v2 , . . . , f∗ vn ) = 0,
where f∗ = f∗,x = Tx f is the tangent mapping of f at x. The equality on the
1,n
left-hand side of (3.2) holds in a weak sense since f ∈ Wloc (M ); see [39, p. 136].

This means that, for all n-forms η ∈ C0 (M ),
Z Z

(3.7) hf d(V y ω), ηi = hf ∗ (V y ω), δηi,
M M
p-Laplace operator, quasiregular mappings, and Picard-type theorems 125

where δ is the codifferential. Consequently, equations (3.4)–(3.6) are to be inter-


preted in weak sense. In particular, combining (3.2), (3.3), and (3.7) we get
Z Z Z

hJf W y ω, δηi = hf (V y ω), δηi = hf ∗ d(V y ω), ηi = 0
M M M

for all n-forms η ∈ C0∞ (M ), and so (3.4) holds in weak sense.


Let us next specify the vector field W. Let A = {x ∈ M : Jf (x) = det f∗,x 6= 0}.
Hence f∗,x is invertible for all x ∈ A, and W = f∗−1 V in A. In M \ A, either Jf (x)
does not exist, which can happen only in a set of measure zero, or Jf (x) ≤ 0.
Quasiregularity of f, more precisely the distortion condition (1.4), implies that
f∗,x = Tx f = 0 for almost every such x. Hence f∗,x = 0 for a.e. x ∈ M \ A.
Setting W = 0 in M \ A, we obtain

f ∗ (V y ω) = 0 = W y f ∗ ω

a.e. in M \ A. Hence f ∗ (V y ω) = W y f ∗ ω a.e. in M, and so (3.3) holds.

3.1. Sketch of the proof of Reshetnyak’s theorem. We shall use Theorem


3.1 to sketch the proof of Reshetnyak’s theorem in a way that uses analysis, in
particular, A-harmonic functions. First we recall some definitions concerning
p-capacity. If Ω ⊂ M is an open set and C ⊂ Ω is compact, then the p-capacity
of the pair (Ω, C) is
Z
(3.8) capp (Ω, C) = inf |∇ϕ|p ,
ϕ Ω

where the infimum is taken over all functions ϕ ∈ C0∞ (Ω), with ϕ|C ≥ 1. A
compact set C ⊂ M is of p-capacity zero, denoted by capp C = 0, if capp (Ω, C) =
0 for all open sets Ω ⊃ C. Finally, a closed set F is of p-capacity zero, denoted
by capp F = 0, if capp C = 0 for all compact sets C ⊂ F. It is a well-known fact
that a closed set F ⊂ Rn containing a continuum C cannot be of n-capacity zero.
This can be seen by taking an open ball B containing C and any test function
ϕ ∈ C0∞ (B), with ϕ|C = 1, and using a potential estimate
Z Z 
|∇ϕ| |∇ϕ|
|ϕ(x) − ϕ(y)| ≤ c dz + dz , x, y ∈ B,
B |x − z| B |y − z|
n−1 n−1

combined with a maximal function and covering arguments. Similarly, if C is


a continuum in a domain Ω and B is an open ball, with B̄ ⊂ Ω \ C, then
capn (C, B̄; Ω) > 0, where
Z
capn (C, B̄; Ω) = inf |∇ϕ|p > 0,
ϕ Ω

the infimum being taken over all functions ϕ ∈ C ∞ (Ω), with ϕ|C = 1 and
ϕ|B̄ = 0.
126 Ilkka Holopainen and Pekka Pankka IWQCMA05

f is light. Suppose then that U ⊂ Rn is a domain and that f : U → Rn is


a non-constant quasiregular mapping. We will show first that f is light which
means that, for all y ∈ Rn , the preimage f −1 (y) is totally disconnected, i.e. each
component of f −1 (y) is a point.
Fix y ∈ Rn and define u : Rn \ {y} → R by
1
u(x) = log .
|x − y|
Then u is C ∞ and n-harmonic in Rn \ {y} by a direct computation. By Theorem
3.1, v = u ◦ f,
1
v(x) = log ,
|f (x) − y|
is A-harmonic in an open non-empty set U \ f −1 (y) and v(x) → +∞ as x → z ∈
f −1 (y). We set v(z) = +∞ for z ∈ f −1 (y).
To show that f is light we use the logarithmic Caccioppoli inequality (2.4).
Suppose that C ⊂ f −1 (y) ∩ U is a continuum. Since f is non-constant and
continuous, there exists m > 1 such that the set Ω = {x ∈ U : v(x) > m} is an
open neighborhood of C and Ω̄ ⊂ U. We choose another neighborhood D of C
such that D̄ ⊂ Ω is compact. Now we observe that vi = min{v, i} is a positive
supersolution for all i > m. The logarithmic Caccioppoli inequality (2.4) then
implies that Z
|∇ log vi |n ≤ c capn (Ω, D̄) ≤ c < ∞
D
uniformly in i. Hence |∇ log v| ∈ Ln (D). Choose an open ball B such that
B̄ ⊂ D \ f −1 (y). We observed earlier that
capn (C, B̄; D) > 0
since C is a continuum. Let
MB = max log v.

Now the idea is to use
1 v
min{1, max{0, log }}
k MB
as a test function for capn (C, B̄; D) for every k ∈ N. We get a contradiction since
0 < capn (C, B̄; D) ≤ k −n k∇ log vkLp (D) → 0
as k → ∞. Thus f −1 (y) can not contain a continuum.

Differentiability a.e. Assume that f = (f1 , . . . , fn ) : U → Rn , U ⊂ Rn , is


quasiregular. Then coordinate functions fj are A-harmonic again by Theorem
3.1, since functions x = (x1 , . . . , xn ) 7→ xj are n-harmonic. Now there are at least
two ways to prove that f is differentiable almost everywhere. For instance, since
each fj is A-harmonic, one can show by employing reverse Hölder inequality
1,p
techniques that, in fact, f ∈ Wloc (U ), with some p > n. This then implies
that f is differentiable a.e. in U ; see e.g. [3]. Another way is to conclude
p-Laplace operator, quasiregular mappings, and Picard-type theorems 127

that f is monotone, i.e. each coordinate function fj is monotone, and therefore


1,n
differentiable a.e. since f ∈ Wloc (U ); see [41]. The monotonicity of fj holds
since A-harmonic functions obey the maximum principle.
1,n
f is sense-preserving. Here one first shows that conditions f ∈ Wloc (U ) and
Jf (x) ≥ 0 a.e. imply that f is weakly sense-preserving, i.e. µ(y, f, D) ≥ 0 for all
domains D ⋐ U and for all y ∈ f D \ f ∂D. This step employs approximation of
f by smooth mappings. Pick then a domain D ⋐ U and y ∈ f D \ f ∂D. Denote
by Y the y-component of Rn \ f ∂D and write V = D ∩ f −1 Y . Since f is light,
D \ f −1 (y) is non-empty. Thus we can find a point x0 ∈ f −1 (y) ∩ V. Next we
conclude that the set {x ∈ V : Jf (x) > 0} has positive measure. Otherwise, since
f is ACL and |f ′ (x)| = 0 a.e. in V , f would be constant in a ball centered at
x0 contradicting the fact that f is light. Thus there is a point x in V where f is
differentiable and Jf (x) > 0. Now a homotopy argument, using the differential
of f at z, and µ(y, f, D) ≥ 0 imply that f is sensepreserving.

f is discrete and open. This part of the proof is purely topological. A sense-
preserving light mapping is discrete and open by Titus and Young; see e.g. [41].

Further properties of f . Once Reshetnyak’s theorem is established it is pos-


sible to prove further properties for quasiregular mappings. We collect these
properties to the following theorems and refer to the books [39] and [41] for the
proofs.
Theorem 3.2. Let f : M → N be a non-constant quasiregular map. Then
1. |f E| = 0 if and only if |E| = 0.
2. |Bf | = 0, where Bf is the branch set of f, i.e. the set of all x ∈ M where
f does not define a local homeomorphism.
3. Jf (x) > 0 a.e.
4. The integral transformation formula
Z Z
(h ◦ f )(x)Jf (x)dm(x) = h(y)N (y, f, A)dm(y)
A N
holds for every measurable h : N → [0, +∞] and for every measurable A ⊂
M, where N (y, f, A) = card f −1 (y) ∩ A.
1,n 1,n
5. If u ∈ Wloc (N, R), then v = u ◦ f ∈ Wloc (M, R) and
∇v(x) = Tx f T ∇u(f (x)) a.e.
Furthermore, we have a generalization of the morphism property.
Theorem 3.3. Let f : M → N be quasiregular and let u : N → R be an A-
harmonic function (or a subsolution or a supersolution, respectively) of type n.
Then v = u ◦ f is f # A-harmonic (a subsolution or a supersolution, respectively),
where
(
Jf (x)Tx f −1 Af (x) (Tx f −1 )T h), if Jf (x) exists and is positive,
f # Ax (h) =
|h|n−2 h, otherwise.
128 Ilkka Holopainen and Pekka Pankka IWQCMA05

The ingredients of the proof of Theorem 3.3 include, for instance, the locality
of A-harmonicity, Theorem 3.2, and a method to ”push-forward” (test) functions;
see e.g. [16] and [41].

4. Modulus and capacity inequalities


Although the main emphasis of this survey is on the relation between quasireg-
ular mappings and p-harmonic functions, we want to introduce also the other
main tool in the theory of quasiregular mappings. Let 1 ≤ p < ∞ and let Γ
be a family of paths in M. We denote by F(Γ) the set of all Borel functions
̺ : M → [0, +∞] such that Z
̺ds ≥ 1
γ
for all locally rectifiable path γ ∈ Γ. We call the functions in F(Γ) admissible
for Γ. The p-modulus of Γ is defined by
Z
Mp (Γ) = inf ̺p dm.
̺∈F (Γ) M

There is a close connection between p-modulus and p-capacity. Indeed, suppose


that Ω ⊂ M is open and C ⊂ Ω is compact. Let Γ be the family of all paths in
Ω \ C connecting C and ∂Ω. Then
(4.1) capp (Ω, C) = Mp (Γ).
The inequality capp (Ω, C) ≥ Mp (Γ) follows easily since ̺ = |∇ϕ| is admissible
for Γ for each function ϕ as in (3.8). The other direction is harder and requires
an approximation argument; see [41, Proposition II.10.2].
If p = n = the dimension of M , we call Mn (Γ) the conformal modulus of Γ,
or simply the modulus of Γ. In that case Mn (Γ) is invariant under conformal
changes of the metric. In fact, Mn (Γ) can be interpreted as follows: Define a
new measurable Riemannian metric
hh·, ·ii = ̺2 h·, ·i.
Then, with respect to hh·, ·ii, the length of γ has a lower bound
Z
ℓhh·,·ii (γ) = ̺ds ≥ 1
γ

and the volume of M is given by


Z
Volhh·,·ii (M ) = ̺n dm.
M

Thus we are minimizing the volume of M under the constraint that paths in Γ
have length at least 1.
The importance of the conformal modulus for quasiregular mappings lies in
the following invariance properties; see [41, II.2.4, II.8.1]
p-Laplace operator, quasiregular mappings, and Picard-type theorems 129

Theorem 4.1. Let f : M → N be a non-constant quasiregular mapping. Let


A ⊂ M be a Borel set with N (f, A) := supy N (y, f, A) < ∞, and let Γ be a
family of paths in A. Then
(4.2) Mn (Γ) ≤ KO (f ) N (f, A) Mn (f Γ).
Theorem 4.2 (Poletsky’s inequality). Let f : M → N be a non-constant quasireg-
ular mapping and let Γ be a family of paths in M. Then
(4.3) Mn (f Γ) ≤ KI (f ) Mn (Γ).

The proof of (4.2) is based on the change of variable formula for integrals
(Theorem 3.2 3.) and on Fuglede’s theorem. The estimate (4.3) in the converse
direction is more useful than (4.2) but also much harder to prove; see [41, p.
39–50].
As an application of the use of p-modulus and p-capacity, we prove a Harnack’s
inequality for positive A-harmonic functions of type p > n − 1. Assume that
Ω ⊂ M is a domain, D ⋐ Ω another domain, and C ⊂ D is compact. For
p > n − 1, we set
λp (C, D) = inf Mp (Γ(E, F ; D)),
E,F
where E and F are continua joining C and Ω \ D, and Γ(E, F ; D) is the family
of all paths joining E and F in D.
Theorem 4.3 (Harnack’s inequality, p > n − 1). Let Ω, D, and C be as above.
Let u be a positive A-harmonic function in Ω of type p > n − 1. Then
 1/p
MC capp (Ω, D̄)
(4.4) log ≤ c0 ,
mC λp (C, D)
where
MC = max u(x), mC = min u(x),
x∈C x∈C
and c0 = c0 (p, β/α).

E Ω

γ C

F D

Proof. We may assume that MC > mC . Let ε > 0 be so small that MC − ε >
mC + ε. Then the sets {x : u(x) ≥ MC − ε} and {x : u(x) ≤ mC + ε} contain
continua E and F , respectively, that join C and Ω \ D. Write
log u − log(mC + ε)
w=
log(MC − ε) − log(mC + ε)
130 Ilkka Holopainen and Pekka Pankka IWQCMA05

and observe that w ≥ 1 in E and w ≤ 0 in F. Therefore |∇w| is admissible for


Γ(E, F ; D) and hence
Z
|∇w|p ≥ Mp (Γ(E, F ; D)) ≥ λp (C, D).
D

On the other hand,


Z
|∇ log u|p dm ≤ c(p, β/α) capp (Ω, D̄)
D

by the logarithmic Caccioppoli inequality (2.4), and


 
MC − ε
∇ log u = log ∇w.
mC + ε
Hence
 1/p
MC − ε capp (Ω, D̄)
log ≤ c0
mC + ε λp (C, D)
and (4.4) follows by letting ε → 0.

We can define λp (C, D) analogously for p ≤ n − 1, too. However, λp (C, D)


vanishes for p ≤ n − 1. Consequently, Theorem 4.3 is useful only for p > n − 1.
The idea of the proof is basically due to Granlund [13]. In the above form, (4.4)
appeared first time in [17]. In general, it is difficult to obtain an effective lower
bound for λp (C, D) together with an upper bound for capp (Ω, D̄). However, if
M = Rn and p = n, one obtains a global Harnack inequality by choosing C, D,
and Ω as concentric balls.

5. Liouville-type results for A-harmonic functions


We say that M is strong p-Liouville if M does not support non-constant
positive A-harmonic functions for any A of type p. We have already mentioned
that a global Harnack inequality
max u ≤ c min u
B(x,r) B(x,r)

holds for non-negative A-harmonic functions on B(x, 2r) with a (Harnack-)constant


c independent of x, r, and u if M is complete and admits a global volume doubling
condition and (1, p)-Poincaré’s inequality. It follows from the global Harnack in-
equality that such manifold M is strong p-Liouville.
Example 5.1. 1. Let M be complete with non-negative Ricci curvature. Then
it is well-known that M admits a global volume doubling property by the
Bishop-Gromov comparison theorem (see [2], [8]). Furthermore, Buser’s
isoperimetric inequality [6] implies that M also admits a (1, p)-Poincaré
inequality for every p ≥ 1. Hence M is strong p-Liouville.
p-Laplace operator, quasiregular mappings, and Picard-type theorems 131

2. Let Hn be the Heisenberg group. We write elements of Hn as (z, t), where


z = (z1 , . . . , zn ) ∈ Cn and t ∈ R. Furthermore, we assume that Hn is
equipped with a left-invariant Riemannian metric in which the vector fields
∂ ∂
Xj = + 2yj ,
∂xj ∂t
∂ ∂
Yj = − 2xj ,
∂yj ∂t

T = ,
∂t
j = 1, . . . , n, form an orthonormal frame. Harnack’s inequality for non-
negative A-harmonic functions on Hn was proved in [18] by using Jerison’s
version of Poincaré’s inequality. Jerison proved in [31] that (1,1)-Poincaré’s
inequality holds for the horizontal gradient
n
X
∇0 u = ((Xj u)Xj + (Yj u)Yj )
j=1

and for balls in so-called Carnot-Carathéodory metric. Since the Lp -norm


of the Riemannian gradient is larger than that of the horizonal gradient,
we have (1,1)-Poincaré’s inequality for the Riemannian gradient as well
if geodesic balls are replaced by Carnot-Carathéodory balls or Heisenberg
balls BH (r) = {(z, t) ∈ Hn : (|z|4 + t2 )1/4 < r} and their left-translations.

Classically, a Riemannian manifold M is called parabolic if it does not support


a positive Green’s function for the Laplace equation.
Definition 5.2. We say that a Riemannian manifold M is p-parabolic, with
1 < p < ∞, if
capp (M, C) = 0
for all compact sets C ⊂ M. Otherwise, we say that M is p-hyperbolic.
Example 5.3. 1. A compact Riemannian manifold is p-parabolic for all p ≥
1.
2. In the Euclidean space Rn we have precise formulas for p-capacities of balls:
(
c rn−p , if 1 ≤ p < n,
capp (Rn , B̄(r)) =
0, otherwise.
Hence Rn is p-parabolic if and only if p ≥ n.
3. If the Heisenberg group Hn is equipped with the left-invariant Riemannian
metric, we do not have precise formulas for capacities of balls. However,
for r ≥ 1,
capp (Hn , B̄H (r)) ≈ r2n+2−p
if 1 ≤ p < 2n + 2, and capp (Hn , B̄H (r)) = 0 if p ≥ 2n + 2. Hence Hn is
p-parabolic if and only if p ≥ 2n + 2.
132 Ilkka Holopainen and Pekka Pankka IWQCMA05

4. Any complete Riemannian manifold M with finite volume Vol(M ) < ∞


is p-parabolic for all p ≥ 1. This is easily
 seen by fixing a point o ∈ M
and taking a function ϕ ∈ C0∞ B(o, R) , with ϕ|B̄(o, r) = 1 and |∇ϕ| ≤
c/(R − r). We obtain an estimate

capp B(o, R), B̄(o, r) ≤ c Vol(M )/(R − r)p → 0
as R → ∞.
5. Let M n be a Cartan-Hadamard n-manifold, i.e. a complete, simply con-
nected Riemannian manifold of non-positive sectional curvatures and di-
mension n. If sectional curvatures have a negative upper bound KM ≤
−a2 < 0, then M is p-hyperbolic for all p ≥ 1. This follows since M n
satisfies an isoperimetric inequality
a
Vol(D) ≤ Area(∂D)
n−1
for all domains D ⋐ M, with smooth boundary; see [46], [7]. Another
proof uses the Laplace comparison and Green’s formula. If p > 1, then
v(x) = exp(−δd(x, o)) is a positive supersolution of the p-Laplace equation
for some δ = δ(n, p) > 0 (see [20]). Hence the p-hyperbolicity of M also
follows from the theorem below for p > 1.
Theorem 5.4. Let M be a Riemannian manifold and 1 < p < ∞. Then the
following conditions are equivalent:
1. M is p-parabolic.
2. Mp (Γ∞ ) = 0, where Γ∞ is the family of all paths γ : [0, ∞) → M such that
γ(t) → ∞ as t → ∞.
3. Every non-negative supersolution of
(5.1) − div Ax (∇u) = 0
on M is constant for all A of type p.
4. M does not support a positive Green’s function g(·, y) for (5.1) for any A
of type p and y ∈ M.

Here γ(t) → ∞ means that γ(t) eventually leaves any compact set. For the
proof of Theorem 5.4 as well as for the discussion below we refer to [17].
Let us explain what is Green’s function for (5.1). We define it first in a
”regular” domain Ω ⋐ M, where regular means that the Dirichlet problem for
A-harmonic equation is solvable with continuous boundary data. For this notion,
see [16]. We need a concept of A-capacity. Let C ⊂ Ω be compact, and assume
for simplicity that Ω \ C is regular. Thus there exists a unique A-harmonic
function in Ω \ C with continuous boundary values u = 0 on ∂Ω and u = 1 in C.
Call u the A-potential of (Ω, C). We define
Z
capA (Ω, C) = hAx (∇u), ∇ui.

Then
capA (Ω, C) ≈ capp (Ω, C)
p-Laplace operator, quasiregular mappings, and Picard-type theorems 133

and furthermore,
(5.2) capA (Ω1 , C1 ) ≥ capA (Ω2 , C2 )
if C2 ⊂ C1 and/or Ω1 ⊂ Ω2 . Note that this property is obvious for variational
capacities but capA is not necessary a variational capacity.
The definition of Green’s function, and in particular its uniqueness when p =
n, relies on the following observation.
Lemma 5.5. Let Ω ⋐ M be a domain and let C ⊂ Ω be compact such that Ω \ C
is regular. Let u be the A-potential of (Ω, C). Then, for every 0 ≤ a < b ≤ 1,
capA (Ω, C)
capA ({u > a}, {u ≥ b}) = .
(b − a)p−1
Definition 5.6. Suppose that Ω ⋐ M is a regular domain and let y ∈ Ω. A
function g = g(·, y) is called a Green’s function for (5.1) in Ω if
1. g is positive and A-harmonic in Ω \ {y},
2. limx→z g(x) = 0 for all z ∈ ∂Ω,
3.
lim g(x) = capA (Ω, {y})1/(1−p) ,
x→y
which we interpret to mean limx→y g(x) = ∞ if p ≤ n,
4. for all 0 ≤ a < b < capA (Ω, {y})1/(1−p) ,
capA ({g > a}, {g ≥ b}) = (b − a)1/(1−p) .
Theorem 5.7. Let Ω ⋐ M be a regular domain and y ∈ Ω. Then there exists a
Green’s function for (5.1) in Ω. Furthermore, it is unique at least if p ≥ n.

Monotonicity properties (5.2) of A-capacity and the so-called Loewner prop-


erty, i.e. capn C > 0 if C is a continuum, are crucial in proving the uniqueness
when p = n. Indeed, we can show that on sufficiently small spheres S(y, r)
|g(x, y) − capA (Ω, B̄(y, r))1/(1−n) | ≤ c, x ∈ S(y, r),
which then easily implies the uniqueness.
Next take an exhaustion of M by regular domains Ωi ⊂ Ωi+1 ⋐ M, M = ∪i Ωi .
We can construct an increasing sequence of Green’s functions gi (·, y) on Ωi . Then
the limit is either identically +∞ or
g(·, y) := lim gi (·, y)
i→∞

is a positive A-harmonic function on M \ {y}. In the latter case we call the limit
function g(·, y) a Green’s function for (5.1) on M.
We have the following list of Liouville-type properties of M (which may or
may not hold for M ):
(1) M is p-parabolic.
(2) Every non-negative A-harmonic function on M is constant for every A of
type p. (Strong p-Liouville.)
134 Ilkka Holopainen and Pekka Pankka IWQCMA05

(3) Every bounded A-harmonic function on M is constant for every A of type


p. (p-Liouville.)
(4) Every A-harmonic function u on M with ∇u ∈ Lp (M ) is constant for every
A of type p. (Dp -Liouville.)
We refer to [17] for the proof of the following general result, and to [18] and
[25] for studies concerning the converse directions.
Theorem 5.8.
(1) ⇒ (2) ⇒ (3) ⇒ (4).
Next we discuss the close connection between the volume growth and p-
parabolicity. Suppose
 that M is complete. Fix a point o ∈ M and write
V (t) = Vol B(o, t) .
Theorem 5.9. Let 1 < p < ∞ and suppose that
Z ∞ 1/(p−1)
t
dt = ∞,
V (t)
or Z ∞
dt

= ∞.
V (t)1/(p−1)
Then M is p-parabolic.
Proof. One can either construct a test function involving the integrals above, or
use a p-modulus estimate for separating (spherical) rings. More precisely, write
B(t) = B(o, t) and S(t) = S(o, t) = ∂B(o, t). For R > r > 0 and integers k ≥ 1,
we write ti = r + i(R − r)/k, i = 0, 1, . . . , k. Then, by a well-known property of
modulus,
k−1
1/(1−p) X 1/(1−p)
Mp Γ S(r), S(R); B̄(R) ≥ Mp Γ S(ti ), S(ti+1 ); B̄(ti+1 ) ;
i=0

see e.g. [41, II.1.5]. Here Γ S(r), S(R); B̄(R) is the family of all paths joining
S(r) and S(R) in B̄(R). For each i = 0, . . . , k − 1 we have an estimate

Mp Γ S(ti ), S(ti+1 ); B̄(ti+1 ) ≤ (V (ti+1 ) − V (ti )) (ti+1 − ti )−p .
Hence
(5.3)
k−1  1/(1−p)
1/(1−p) X V (ti+1 ) − V (ti )
Mp Γ S(r), S(R); B̄(R) ≥ (ti+1 − ti ).
i=0
ti+1 − ti
Thus the right-hand side of (5.3) tends to the integral
Z R
dt

r V (t)
1/(p−1)

as k → ∞. We obtain an estimate
Z R 1−p
 dt
Mp Γ S(r), S(R); B̄(R) ≤ ′
.
r V (t)
1/(p−1)
p-Laplace operator, quasiregular mappings, and Picard-type theorems 135

In particular, if Z ∞
dt
=∞
r V ′ (t)1/(p−1)
for some r > 0, then M is p-parabolic.

The converse is not true in general. That is, M can be p-parabolic even if
Z ∞ 1/(p−1)
t
dt < ∞
V (t)
or Z ∞
dt
< ∞;
V ′ (t)1/(p−1)
see [44].
It is interesting to study when the converse is true. We refer to [19] for the
proofs of the following two theorems.
Theorem 5.10. Suppose that M is complete and admits a global doubling prop-
erty and global (1, p)-Poincaré inequality for 1 < p < ∞. Then
Z ∞ 1/(p−1)
t
(5.4) M is p-hyperbolic if and only if dt < ∞.
V (t)

In some cases, we can estimate Green’s functions:


Theorem 5.11. Suppose that M is complete and has non-negative Ricci curva-
ture everywhere. Let 1 < p < ∞. Then
Z ∞ 1/(p−1)
t
M is p-hyperbolic if and only if dt < ∞.
V (t)
Furthermore, we have estimates for Green’s functions for (5.1)
Z ∞ 1/(p−1) Z ∞ 1/(p−1)
−1 t t
c dt ≤ g(x, o) ≤ c dt
2r V (t) 2r V (t)
for every x ∈ ∂M (r), where M (r) is the union of all unbounded components of
M \ B̄(o, r). The constant c depends only on n, p, α, and β.

Theorem 5.10 follows also from the following sharper result; see [21].
Theorem 5.12. Suppose that M is complete and that there exists a geodesic ray
γ : [0, ∞) → M such that for all t > 0,
|B(γ(t), 2s)| ≤ c|B(γ(t), s)|,
whenever 0 < s ≤ t/4, and that
 1/p
Z Z
 
|u − uBγ (t) |dm ≤ c  |∇u|p dm
Bγ (t) 2Bγ (t)
136 Ilkka Holopainen and Pekka Pankka IWQCMA05


for all u ∈ C ∞ 2Bγ (t) , where Bγ (t) = B(γ(t), t/8). Then M is p-hyperbolic if
Z ∞ 1/(p−1)
t
dt < ∞.
|B(γ(t), t/4)|

Theorem 5.12 can be applied to obtain the following.


Theorem 5.13. Let M be a complete Riemannian n-manifold whose Ricci curva-
ture is non-negative outside a compact set. Suppose that M has maximal volume
growth ( V (t) ≈ rn ). Then M is p-parabolic if and only if p ≥ n.

To our knowledge it is an open problem whether the equivalence (5.4) holds


for a complete Riemannian n-manifold whose Ricci curvature is non-negative
outside a compact set.

6. Liouville-type results for quasiregular mappings


Here we give applications of the above results on n-parabolicity and various
Liouville properties to the existence of non-constant quasiregular mappings be-
tween given Riemannian manifolds.
Let us start with the Gromov-Zorich ”global homeomorphism theorem” that
is a generalization of Zorich’s theorem we mentioned in the introduction; see [14],
[48].
Theorem 6.1. Suppose that M is n-parabolic, n = dim M ≥ 3, and that N is
simply connected. Let f : M → N be a locally homeomorphic quasiregular map.
Then f is injective and f M is n-parabolic.

Proof. We give here a very rough idea of the proof. First one observes that
f M is n-parabolic (see Theorem 6.2 below), and so N \ f M is of n-capacity
zero. Then one shows, again by using the n-parabolicity of M, that the set E
of all asymptotic limits of f is of zero capacity. Consequently, E is of Hausdorff
 zero. Recall that an asymptotic limit of f is a point y ∈ N such that
dimension
f γ(t) → y as t → ∞ for some path γ ∈ Γ∞ in M. Removing E ∪ (N \ F M )
from N has no effect on the simply connectivity for dimensions n ≥ 3. That is,
f M \ E remains simply connected. Thus one can extend uniquely any branch of
local inverses of f and obtain a homeomorphism g : f M \ E → g(f M \ E) such
that f ◦ g = id |(f M \ E). Finally, g can be extended to E to obtain the inverse
of f.

In [22] we generalized the global homeomorphism theorem for mappings of


finite distortion under mild conditions on the distortion. See also [49] for a
related result for locally quasiconformal mappings.
Theorem 6.2. If N is n-hyperbolic and M is n-parabolic, then every quasiregular
mapping f : M → N is constant.
p-Laplace operator, quasiregular mappings, and Picard-type theorems 137

Proof. Suppose that f : M → N is a non-constant quasiregular mapping. Then


f M ⊂ N is open. If f M 6= N, pick a point y ∈ ∂(N \ f M ) and let g =
g(·, y) be the Green’s function on N for the n-Laplacian. Then g ◦ f is a non-
constant positive A-harmonic function on M which gives a contradiction with
the n-parabolicity of M and Theorem 5.8. If f M = N, let u be a non-constant
positive supersolution on N for the n-Laplacian. Then u ◦ f is a non-constant
supersolution on M for some A of type n which is again a contradiction.
Example 6.3. 1. If N is a Cartan-Hadamard manifold, with KN ≤ −a2 < 0,
then every quasiregular mapping f : Rn → N is constant.
2. Let Hn be the Heisenberg group with a left-invariant Riemannian metric,
then every quasiregular mapping f : R2n+1 → Hn is constant.
Theorem 6.4. Suppose that M is strong n-Liouville while N is not. Then every
quasiregular map f : M → N is constant.

Proof. If N is not strong n-Liouville, then it is n-hyperbolic by Theorem 5.8.


Suppose that f : M → N is a non-constant quasiregular mapping. Then f M ⊂
N is open. If f M 6= N, choose a point y ∈ ∂(N \ f M ) and let g = g(·, y)
be the Green’s function for n-Laplacian on N . Then g ◦ f is a non-constant
positive A-harmonic function, with A of type n. This is a contradiction. If
f M = N , we choose a non-constant positive n-harmonic function u on N and
get a contradiction as above.
Theorem 6.5. Let N be a Cartan-Hadamard n-manifold, with −b2 ≤ K ≤
−a2 < 0, and let M be a complete Riemannian n-manifold admitting a global
doubling property and a global (1, n)-Poincaré inequality. Then every quasiregular
mapping f : M → N is constant.

Proof. By [20], N admits non-constant positive n-harmonic functions. Hence


N is not strong n-Liouville. On the other hand, the assumptions on M imply
that a global Harnack’s inequality for positive A-harmonic functions of type n
holds on M. Thus M is strong n-Liouville, and the claim follows from Theorem
6.4.
Theorem 6.6 (”One-point Picard”). Suppose that N is n-hyperbolic and M is
strong n-Liouville. Then every quasiregular mapping f : M → N \ {y}, with
y ∈ N, is constant.

Proof. Suppose that f : M → N \ {y} is a non-constant quasiregular mapping.


Then ∂(N \ f M ) 6= ∅. Choose a point z ∈ ∂(N \ f M ), and let g = g(·, z) be the
Green’s function on N for the n-Laplacian. Then g ◦ f is a non-constant positive
A-harmonic function for some A of type n leading to a contradiction.

7. Picard-type theorems
The classical big Picard theorem states that a holomorphic mapping of the
punctured unit disc {z ∈ C : 0 < |z| < 1} into the complex plane omitting two
138 Ilkka Holopainen and Pekka Pankka IWQCMA05

values has a meromorphic extension to the whole disc; see e.g. [1, Theorem 1-
14]. In [40] Rickman proved a counterpart of Picard’s theorem for quasiregular
mappings (Theorem 7.1) and its local version (Theorem 7.2) corresponding to
the big Picard theorem.
Theorem 7.1 ([40]). For each integer n ≥ 2 and each K ≥ 1 there exists
a positive integer q = q(n, K) such that if f : Rn → Rn \ {a1 , . . . , aq } is K-
quasiregular and a1 , . . . , aq are distinct points in Rn , then f is constant.
Theorem 7.2 ([40]). Let G = {x ∈ Rn : |x| > s} and let f : G → Rn \
{a1 , . . . , aq } be a K-quasiregular mapping, where a1 , . . . , aq are distinct points in
Rn and q = q(n, K) is the integer in Theorem 7.1. Then the limit lim|x|→∞ f (x)
exists.

In this section we consider corollaries and extensions of the big Picard theorem
for quasiregular mappings.
Although the short argument yielding Theorem 7.1 from Theorem 7.2 is well-
known, it seems that the following corollary employing the same argument has
gone unnoticed in the literature.
Corollary 7.3. Let K ≥ 1 and R > 0. Let f : Rn → Rn be a continuous mapping
omitting at least q = q(n, K) points, where q(n, K) is as in Theorem 7.1. Then
at least one of the following conditions fails:
(i) f |Rn \ B̄ n (R) is K-quasiregular,
(ii) f B n (r) is open for some r > R.
Proof. Suppose towards a contradiction that both conditions (i) and (ii) hold.
By Theorem 7.2, the mapping f has a limit at the infinity. Hence we may
extend f to a continuous mapping R̄n → R̄n . Moreover, f is K-quasiregular
in R̄n \ B̄ n (R). By composing f with a Möbius mapping if necessary, we may
assume that f (∞) = ∞. Since f is a non-constant quasiregular mapping on
R̄n \ B̄ n (R), f |R̄n \ B̄ n (R) is an open mapping. Hence f R̄n is open in R̄n , by (ii).
Since f R̄n is both open and closed, f R̄n = R̄n and f Rn = Rn . This contradicts
the assumption that f omits q points. The claim follows.
In [23] the authors consider quasiregular mappings of the punctured unit ball
into a Riemannian manifold N . We say that N has at least q ends, if there exists
a compact set C ⊂ N such that N \ C has at least q components which are not
relatively compact. Such a component of N \C is called an end of M with respect
to C. Let E be the set of ends of N , that is, E ∈ E is an end of N with respect to
some compact set C ⊂ N . We compactify N with respect to its ends as follows.
There is a natural partial order in E induced by inclusion. We call a maximal
totally ordered subset of E an asymptotic end of N . The set of asymptotic ends
of N is denoted by ∂N and N̂ = N ∪ ∂N . We endow N̂ with a topology such
that the inclusion N ⊂ N̂ is an embedding and for every e ∈ ∂N sets E ∈ e
form a neighborhood basis at e. The main result is the following version of the
big Picard theorem.
p-Laplace operator, quasiregular mappings, and Picard-type theorems 139

Theorem 7.4 ([23, Theorem 1.3]). For every K ≥ 1 there exists q = q(K, n)
such that every K-quasiregular mapping f : B n \{0} → N has a limit limx→0 f (x)
in N̂ if N has at least q ends.
In the spirit of Corollary 7.3 we formulate the following consequence Theorem
7.4.
Corollary 7.5. Given n ≥ 2 and K ≥ 1 there exists q̃ = q̃(n, K) such that the
following holds. Suppose that M is compact, {z1 , . . . , zk } ⊂ M , where 1 ≤ k < q̃,
and that N has at least q̃ ends. Let f : M \ {z1 , . . . , zk } → N be a continuous
mapping and let Ωi be a neighborhood of zi for every 1 ≤ i ≤ k. Then at least
one of the following conditions fails:
(i) f is K-quasiregular in Ωi \ {zi } for every i,
(ii) there exists a neighborhood Ω of M \ (Ω1 ∪ · · · ∪ Ωk ) such that f Ω is open.
Proof. Suppose that both conditions are satisfied. For every 1 ≤ i ≤ k we
fix a 2-bilipschitz chart ϕi : Ui → ϕi Ui at zi . We may assume that Ui ⊂ Ωi .
Every mapping f ◦ ϕ−1 n
i |ϕi (Ui \ zi ) is 2 K-quasiregular, and therefore it has a
limit at ϕi (zi ) by Theorem 7.4 if N has at least q(n, 2n K) ends. Hence f has
a limit at every point zi . We extend f to a continuous mapping fˆ: M → N̂ .
Denote M ′ = M \ {z1 , . . . , zk }. Since f is an open mapping, ∂f M ′ ∩ f M ′ = ∅.
Furthermore, since M is compact,
f M ′ ⊂ fˆM = fˆM.
Hence ∂f M ′ ⊂ fˆM \ f M ′ . Thus card(∂f M ′ ) ≤ card(fˆM \ f M ′ ) ≤ k and
N̂ = N = f M ′ = fˆM.
This is a contradiction, since
card(fˆM \ f M ′ ) ≤ k < q ≤ card(N̂ \ f M ′ ) = card(fˆM \ f M ′ ).

In [26] Holopainen and Rickman applied a method of Lewis ([33]) that relies on
Harnack’s inequality to prove the following general version of Picard’s theorem on
the number of omitted values of a quasiregular mapping. We say that a complete
Riemannian n-manifold M belongs to the class M(m, ϑ), where m : (0, 1) → N
and ϑ : (0, ∞) → (0, ∞) are given functions, if following two conditions hold:
(m) for each 0 < λ < 1 every ball of radius r in M can contain at most m(λ)
disjoint balls of radius λr, and
(ϑ) M admits a global Harnack’s inequality for non-negative A-harmonic func-
tions of type n with Harnack-constant ϑ(β/α), where α and β are the
constants of A.
Theorem 7.6 ([26]). Given n ≥ 2, K ≥ 1, m : (0, 1) → N, and ϑ : (0, ∞) →
(0, ∞) there exists q = q(n, K, m, ϑ) ≥ 2 such that the following holds. Suppose
that M belongs to the class M(m, ϑ) and that N has at least q ends. Then every
K-quasiregular mapping f : M → N is constant.
140 Ilkka Holopainen and Pekka Pankka IWQCMA05

Next we show that this theorem admits a local version. Suppose that M is
complete. We say that an asymptotic end e of M is of type E(m, ϑ) if there
exists E ∈ e such that
(Em) for each 0 < λ < 1 every ball of radius r in E can contain at most m(λ)
disjoint balls of radius λr, and
(Eϑ) E admits a uniform Harnack inequality for non-negative A-harmonic func-
tions of E of type n for balls B ⊂ E satisfying 4B ⊂ E. We also assume
that the Harnack constant ϑ depends only on β/α, where α and β are the
constants of A.
We also say that an asymptotic end e of M is p-parabolic (with p ≥ 1) if there
exists E ∈ e such that for every ε > 0 there exists E ′ ∈ e such that
Mp (Γ(E ′ , M \ E; M )) < ε.
Furthermore, we say that an asymptotic end e of M is locally C-quasiconvex
if for every E ∈ e there exists E ′ ∈ e, E ′ ⊂ E, such that each pair of points
x, y ∈ E ′ can be joint by a path in E ′ of length at most Cd(x, y), where d is the
Riemannian distance of M .
Theorem 7.7. Let n ≥ 2, K ≥ 1, m : (0, 1) → N, and ϑ : (0, ∞) → (0, ∞).
Then there exists q = q(n, K, m, ϑ) such that the following holds. Suppose that
M is complete and e is an n-parabolic locally C-quasiconvex asymptotic end of
M of type E(m, ϑ), and that N has at least q ends. Let E ∈ e and f : E → N be
a K-quasiregular mapping. Then f has a limit at e.
Corollary 7.8. Let n ≥ 2, K ≥ 1, m : (0, 1) → N, and ϑ : (0, ∞) → (0, ∞).
Then there exists q = q(n, K, m, ϑ) such that the following holds. Suppose that
a complete Riemannian n-manifold M has asymptotic ends {e1 , . . . , ek }, k < q,
of type E(m, ϑ) which are all n-parabolic and locally C-quasiconvex, and that N
has at least q ends. Let f : M → N be a continuous mapping and Ei ∈ ei for
every 1 ≤ i ≤ k. Then at least one of the following conditions fails:
(i) f is K-quasiregular in Ei for every i,
(ii) there exists a neighborhood Ω of M \ (E1 ∪ · · · ∪ Ek ) such that f Ω is open.

Proof. Suppose that both conditions hold. By Theorem 7.7, we may extend f
to a continuous mapping fˆ: M̂ → N̂ . Since M̂ is compact, we may follow the
proof of Corollary 7.5.

We need several lemmas in order to prove Theorem 7.7. Let us first recall
the definition of a Harnack function. Let M be a Riemannian manifold. A
continuous function u : M → R is called a Harnack function with constant θ if
M (h, x, r) := sup h ≤ θ inf h
B(x,r) B(x,r)

holds in each ball B(x, r) whenever the function h is nonnegative in B(x, 2r),
has the form h = ±u + a for some a ∈ R, and B̄(x, 2r) ⊂ M is compact. The
original version of Lewis’ lemma is stated for Harnack functions. It is well known
p-Laplace operator, quasiregular mappings, and Picard-type theorems 141

(see [16, 6.2]) that A-harmonic functions in the Euclidean setting are Harnack
functions with some θ depending only on n and on the constants p, α, and β of
A. In that case θ is called the Harnack constant of A.
Lemma 7.9. Let e be an n-parabolic locally C-quasiconvex asymptotic end of a
complete Riemannian n-manifold M . Suppose u : E → R, where E ∈ e, is a Har-
nack function with constant θ such that lim supx→e u(x) = ∞ and lim inf x→e u(x) <
0. Then for every C0 > 0 there exists a ball B = B(x0 , r0 ) ⊂ E such that
(1) B(x0 , 100Cr0 ) ⊂ E,
(2) u(x0 ) = 0, and
(3) maxB u ≥ C0 .

Proof. It is sufficient to modify the proof of [23, Lemma 2.1] as follows. Let
E ′ ∈ e be such that E ′ ⊂ E and E ′ is C-quasiconvex. Let F ′ ⊂ M be a compact
set such that E ′ is a component of M \ F ′ , fix o ∈ M , and let R0 > 0 be such
that F ′ ⊂ B̄(o, R0 /2).
Fix k ∈ N such that given r > R0 and x, y ∈ ∂B(o, r) ∩ E ′ there exists k balls
Bi = B(xi , r/1000), 1 ≤ i ≤ k, in E such that
(1) x ∈ B1 ,
(2) y ∈ Bk ,
(3) xi ∈ E ′ for every i, and
(4) Bi ∩ Bi+1 6= ∅ for every i ∈ {1, . . . , k − 1}.
Indeed, since Bi ∩B(o, R0 /2) = ∅, we have Bi ⊂ E, and since E ′ is C-quasiconvex,
we may choose any k > 2000C. We may now apply the proof of [23, Lemma 5]
almost verbatim.
Lemma 7.10. Let e be an n-parabolic locally C-quasiconvex asymptotic end of a
complete Riemannian n-manifold M and E ∈ e. If f : E → N is a quasiregular
mapping such that f E is n-hyperbolic, then f has a limit in N̂ at e.

Proof. Suppose that f E is n-hyperbolic. We may assume that E is C-quasiconvex.


If f has no limit at e, there exists a compact set F ⊂ N such that f E ′ ∩ F 6= ∅
for every E ′ ∈ e. Hence there exists a sequence (xk ) such that xk → e and
f (xk ) → z ∈ N as k → ∞. Let (yk ) be another sequence such that yk → e as
k → ∞. We show that the hyperbolicity of f E yields f (yk ) → z as k → ∞,
which is a contradiction.
For every k we fix a path αk : [0, 1] → E such that αk (0) = xk , αk (1) = yk ,
and ℓ(αk ) ≤ Cd(xk , yk ). Then
capn (E, |αk |) → 0
as k → ∞. By Poletsky’s inequality (4.3) and (4.1),
capn (f E, f |αk |) ≤ KI (f ) capn (E, |αk |)
for every k. Suppose that f (yk ) 6→ z. Then, by passing to a subsequence
if necessary, we may assume that d(f (yk ), z) ≥ δ > 0 for every k. Since
142 Ilkka Holopainen and Pekka Pankka IWQCMA05

d(f (αk (0)), f (αk (1))) ≥ δ/2 for large k, we have, by the n-hyperbolicity of f E,
that
capn (f E, f |αk |) ≥ ε > 0.
for every k. This is a contradiction.

The following lemma is a reformulation of [29, Lemma 19.3.2].


Lemma 7.11. Let E ⊂ M , let u : E → R be a non-constant Harnack function
with constant θ, and let α : [a, b] → E be a path. If ℓ(α) ≤ k dist(|α|, u−1 (0) ∪
M \ E), then u has a constant sign on |α|. Furthermore,
max u ≤ θk min u
|α| |α|

if u is positive on |α|, and


max u ≤ θ−k min u
|α| |α|

if u is negative on |α|.

Proof. Since |α| is connected, every non-vanishing function on |α| has constant
sign. We may assume without loss of generality that u is positive on |α|. Let
a = a0 < a1 < . . . < ak = b be a partition of [a, b] such that ℓ(α|[ai , ai+1 ]) =
ℓ(α)/k for every i = 0, 1, . . . , k − 1. For every i fix xi ∈ α([ai , ai+1 ]) such that
ℓ(α|[ai , xi ]) = ℓ(α|[xi , ai+1 ]). Then α([ai , ai+1 ]) ⊂ B̄(xi , ℓ(α)/(2k)). Further-
more, B(xi , ℓ(α)/k) ⊂ E and B(xi , ℓ(α)/k) ∩ u−1 (0) = ∅. Since α(ai+1 ) ∈
B̄(xi , ℓ(α)/(2k)) ∩ B̄(xi+1 , ℓ(α)/(2k)) for every i = 1, . . . , k − 1, a repeated use
of Harnack’s inequality yields max|α| u ≤ θk min|α| u.
Lemma 7.12 (Lewis’ lemma). Let M , e, E, and u be as in Theorem 7.7. Then
for every C0 > 0 there exists a ball B = B(x0 , r0 ) ⊂ E such that
(1) 6B ⊂ E,
(2) u(x0 ) = 0, and
(3) C0 ≤ max6B u ≤ θ6 maxB u.

Proof. Let C0 > 0 and B(x0 , R) be as in Lemma 7.9. Let Z = u−1 (0) and
ZR = Z ∩ B̄(x0 , 41R). For
S each x ∈ ZR we set rx = R − d(x, x0 )/41 and
Bx = B(x, rx ). Then F = x∈ZR B̄x is compact and x 7→ maxB̄x u is continuous.
Let a ∈ ZR be a point of maximum for this function. Thus
max u ≥ max u ≥ C0 .
B̄(a,ra ) B̄(x0 ,R)

As in [29, Lemma 19.4.1], we have that


5ra
dist(Z, B̄(a, 6ra ) \ F ) ≥ .
6
Let y0 ∈ B̄(a, 6ra ) be such that
u(y0 ) = max u ≥ C0 > 0.
B̄(a,6ra )
p-Laplace operator, quasiregular mappings, and Picard-type theorems 143

If y0 ∈ F , then, by the maximal property of ball B(a, ra ),


max u = u(y0 ) ≤ max u = max u ≤ θ6 max u.
B̄(a,6ra ) F B̄(a,ra ) B̄(a,ra )

If y0 6∈ F , let y1 ∈ F ∩B̄(a, 6ra ) be nearest to y0 in length metric. As B̄(a, ra ) ⊂ F


it follows that
d(y0 , y1 ) ≤ dist(y0 , B̄(a, ra )) ≤ 6ra − ra = 5ra .
Let α : [0, 1] → E be a path of minimal length such that α(0) = y0 and α(1) = y1 .
Then α[0, 1) ∩ F = ∅. Hence
5ra
dist(Z, |α|) ≥
.
6
Thus ℓ(α) ≤ Cd(y0 , y1 ) ≤ 6C dist(Z, |α|). By Lemma 7.11,
u(y0 ) ≤ max u ≤ θ6C min u ≤ θ6C u(y1 ) ≤ θ6C max u = θ6C max u.
|α| |α| F B̄(a,ra )

Lemma 7.13 ([24],[26]). Let N be an n-parabolic Riemannian manifold. Suppose


that C ⊂ N is compact such that N has q ends V1 , . . . , Vq with respect to C. Then
there exist n-harmonic functions vj , j = 2, . . . , q, and a positive constant κ such
that
(7.1) |vj | ≤ κ in C,
(7.2) |vj − vi | ≤ 2κ in V1 ,
(7.3) sup vj = ∞,
V1
(7.4) inf vj = −∞,
Vj

(7.5) vj is bounded in Vk for k 6= 1, j,


(7.6) if vj (x) > κ, then x ∈ V1 ,
(7.7) if vj (x) < −κ then x ∈ Vj .

Proof of Theorem 7.7. Suppose that a K-quasiregular mapping f : E → N


has no limit at e. By Lemma 7.10, N is n-parabolic. Let C ⊂ N be a compact
set such that N has q ends V1 , . . . , Vq with respect to C. For every j = 2, . . . , q
let us fix an n-harmonic function vj with properties (7.1) - (7.7) given in Lemma
7.13. For every j = 2, . . . , q we set uj = vj ◦f . Then functions uj are A-harmonic
in E. Next we show that
(7.8) lim sup uj (x) = +∞ and lim inf uj (x) = −∞,
x→e x→e

and hence they satisfy the assumptions of Lemma 7.9. This can be seen by
observing that the sets {x ∈ N : vj (x) > c} and {x ∈ N : vj (x) < −c} are non-
empty and open for every c > 0 and j = 2, . . . , q. By Lemma 7.10, f (E \ F )
intersects these sets for every compact F ⊂ M , and therefore (7.8) follows. By
144 Ilkka Holopainen and Pekka Pankka IWQCMA05

Lemma 7.12 there are sequences xi ∈ E and ri ∈ (0, ∞), i ∈ N, such that
u2 (xi ) = 0, B(xi , 3ri ) ⊂ E,
M (u2 , xi , 3ri ) ≤ θ6 M (u2 , xi , ri /2),
and M (u2 , xi , ri /2) → ∞ as i → ∞. Let us fix an index i such that M (u2 , xi , ri /2) ≥
4θκ, where θ > 1 is the Harnack constant of A and κ is the  constant in Lemma
7.13. We write x = xi and r = ri . By (7.6), f B(x, r/2) ∩ V1 6= ∅. Thus, by
(7.2), we have
(7.9) M (u2 , x, s) − 2κ ≤ M (uj , x, s) ≤ M (u2 , x, s) + 2κ
whenever s ≥ r/2. Next we conclude by using Harnack’s inequality that
(7.10) M (uj , x, r) ≤ (θ − 1)M (−uj , x, 2r)
for all j. Let us first show that uj (z) = 0 for some z ∈ B(x, r). Suppose on
the contrary, that uj > 0 in B(x, r). Then uj (y) ≤ θuj (x) for all y ∈ B(x, r/2)
by Harnack’s inequality. Since M (u2 , x, r/2) ≥ 4θκ, there exists y ∈ B(x, r/2)
such that uj (y) > 2θκ by (7.9). Thus uj (x) > 2κ, and so x ∈ V1 . By (7.2),
u2 (x) ≥ uj (x) − 2κ > 0 contradicting the assumption u2 (x) = 0. Therefore there
exists z ∈ B(x, r) such that uj (z) = 0. Thus inf B(x,r) uj ≤ 0. Inequality (7.10)
follows now from the calculation
 
M (uj , x, r) = sup uj = sup uj − inf uj + inf uj
B(x,r) B(x,r) B(x,2r) B(x,2r)
 
≤ θ inf uj − inf uj + inf uj
B(x,r) B(x,2r) B(x,2r)

= θ inf uj + (1 − θ) inf uj
B(x,r) B(x,2r)

≤ −(θ − 1) inf uj = (θ − 1) sup (−uj )


B(x,2r) B(x,2r)

= (θ − 1)M (−uj , x, 2r),


since uj − inf B(x,2r) uj ≥ 0 in B(x, 2r).
Inequalities (7.9) and (7.10), and the assumption M (u2 , x, r/2) ≥ 4θκ together
yield the inequality
(7.11) M (u2 , x, r) ≤ θM (−uj , x, 2r).
Indeed,
M (u2 , x, r) ≤ M (uj , x, r) + θ−1 M (u2 , x, r)
≤ (θ − 1)M (−uj , x, 2r) + θ−1 M (u2 , x, r),
which is equivalent to (7.11). We fix zj ∈ B̄(x, 2r) such that
(7.12) M (−uj , x, 2r) = −uj (zj ).
The well-known oscillation estimate (see e.g. [16, 6.6])
osc uj ≤ c(ρ/r)γ osc uj
B(y,ρ) B(y,r)
p-Laplace operator, quasiregular mappings, and Picard-type theorems 145

together with [24, Lemma 4.2] and (7.9) imply that


(7.13) osc uj ≤ c1 (ρ/r)γ M (u2 , x, 3r)
B(zj ,ρ)

for ρ ∈ (0, r). See [24, (5.5)] for details. Thus


max uj = osc uj + min uj
B̄(zj ,ρ) B(zj ,ρ) B̄(zj ,ρ)
γ
≤ c1 (ρ/r) M (u2 , x, 3r) + uj (zj )
≤ c1 (ρ/r)γ M (u2 , x, 3r) − θ−1 M (u2 , x, r),
by (7.13), (7.12), and (7.11). Since M (u2 , x, 3r) ≤ θ6 M (u2 , x, r), we obtain
c1 (ρ/r)γ M (u2 , x, 3r) ≤ (2θ)−1 M (u2 , x, r)
by choosing ρ = (2θ7 c1 )−1/γ r. Hence
max uj ≤ −(2θ)−1 M (u2 , x, r) ≤ −2κ.
B̄(zj ,ρ)

By (7.7), we conclude that f B(zj , ρ) ⊂ Vj and hence the balls B(zj , ρ) are
disjoint. Since B(zj , ρ) ⊂ B(x, 3r), there can be at most m(ρ/3r) of them.
Hence q has an upper bound that depends only on n, K, ϑ, and m.

8. Quasiregular mappings, p-harmonic forms, and de


Rham cohomology
The use of n-harmonic functions in studying Liouville-type theorems for quasireg-
ular mappings f : M → N is restricted to the case, where N is non-compact. The
reason for this restriction is simple: a compact Riemannian manifold does not
carry non-constant p-harmonic functions. Therefore, in the case of a compact
target manifold, we have to use p-harmonic forms. In this final section we discuss
briefly p-harmonic and A-harmonic forms and their connections to quasiregular
mappings. For detailed discussions on A-harmonic forms, see e.g. [27], [28], [29],
[30], and [42]. For the connection of A-harmonic forms to quasiregular mappings,
see e.g. [4], [29], and [34].

Vℓ The

Riemannian metric of M induces an inner product to the exterior bundle
T M for every ℓ ∈ {1, . . . , n}, see e.g. [29, 9.6] for details. We denote this
inner product by h·, ·i and the corresponding norm by | · |. As usual, sections of
V
the bundle ℓ T ∗ M are called ℓ-forms. The Lp -space of measurable ℓ-forms is
V
denoted by Lp ( ℓ M ) and the Lp -norm is defined by
Z 1/p
p
kξkp = |ξ| dx .
M
V
The local Lp -spaces of ℓ-forms are denoted by Lploc ( ℓ M ). The space of C ∞ -
V
smooth ℓ-forms on M is denoted by C ∞ ( ℓ M ), and the space of compactly
V
supported C ∞ -smooth ℓ-forms by C0∞ ( ℓ M ).
146 Ilkka Holopainen and Pekka Pankka IWQCMA05

V V
Let ℓ ∈ {1, . . . , n−1} and p > 1. Let A : ℓ T ∗ M → ℓ T ∗ M be a measurable
bundle map such that there exists positive constants a and b satisfying
(8.1) hA(ξ) − A(ζ), ξ − ζi ≥ a(|ξ| + |ζ|)p−2 |ξ − ζ|2 ,
(8.2) |A(ξ) − A(ζ)| ≤ b(|ξ| + |ζ|)p−2 |ξ − ζ|, and
(8.3) A(tξ) = t|t|p−2 A(ξ)
V
for all ξ, ζ ∈ ℓ Tx∗ M , t ∈ R, and for almost every x ∈ M . We also assume that
V
x 7→ Ax (ω) is a measurable ℓ-form for every measurable ℓ-form ω : M → ℓ T ∗ M .
We say that an ℓ-form ξ is A-harmonic (of type p) on M if ξ is a weakly closed
d,p Vℓ
continuous form in Wloc ( M ) and satisfies equality
δ(A(ξ)) = 0
weakly, that is, Z
hA(ξ), dϕi = 0
M
V d,p Vℓ
for all ϕ ∈ C0∞ ( ℓ−1 M ). Here Wloc ( M ) is the partial Sobolev space of
p Vℓ d,p Vℓ
ℓ-forms. A form ω ∈ Lloc ( M ) is in the space Wloc ( M ) if the distribu-
p Vℓ+1
tional exterior derivative dω exists and dω ∈ Lloc ( M ). The global space
d,p
Vℓ d,p Vℓ
W ( M ) is defined similarly. A form ω ∈ Wloc ( M ) is weakly closed if
d,p Vℓ−1
dω = 0 and weakly exact if ω = dτ for some τ ∈ Wloc ( M ).
Apart from minor differences between conditions (8.1)-(8.3) and the corre-
sponding conditions in Section 2, we can say that A-harmonic functions corre-
spond to A-harmonic weakly exact 1-forms.
Let f : M → N be a quasiregular mapping. Since f is almost everywhere
n/ℓ V
differentiable, we may define the pull-back f ∗ ξ of the form ξ ∈ Lloc ( ℓ N ) by
(f ∗ ξ)x = (Tx f )∗ ξf (x) .
n/ℓ V
By the quasiregularity of f , f ∗ ξ ∈ Lloc ( ℓ M ). Furthermore, d(f ∗ ξ) = f ∗ (dξ)
1,n/ℓ V 1,n/ℓ V 1,n/ℓ V
if ξ ∈ Wloc ( ℓ N ). Hence f ∗ ξ ∈ Wloc ( ℓ M ) for ξ ∈ Wloc ( ℓ M ). The
quasiregularity of f also yields that the pull-back f ∗ ξ of an (n/ℓ)-harmonic ℓ-form
V
is A-harmonic. Similarly to the case of A-harmonic functions, A : ℓ T ∗ M →
Vℓ ∗
T M is defined by
A(η) = hG∗ η, ηi(n/ℓ)−2 G∗ η,
where T
Gx = Jf (x)2/n (Tx f )−1 (Tx f )−1 a.e. .
Recently in [4] Bonk and Heinonen studied cohomology of quasiregularly el-
liptic manifolds using p-harmonic forms. A connected Riemannian manifold is
called K-quasiregularly elliptic if it receives a non-constant K-quasiregular map-
ping from Rn . The main result of [4] is the following theorem.
Theorem 8.1 ([4, Theorem 1.1]). Given n ≥ 2 and K ≥ 1 there exists a constant
C = C(n, K) > 1 such that dim H ∗ (N ) ≤ C for every K-quasiregularly elliptic
closed n-manifold N .
p-Laplace operator, quasiregular mappings, and Picard-type theorems 147

As the Picard-type theorem 7.6, also this theorem has a local counterpart.
Theorem 8.2 ([34, Theorem 2]). Given n ≥ 2 and K ≥ 1 there exists a constant
C ′ = C ′ (n, K) > 1 such that every K-quasiregular mapping f : B n \{0} → N has
a limit at origin if N is closed, connected, and oriented Riemannian n-manifold
with dim H ∗ (N ) ≥ C ′ .

We close this section with a sketch of the proof of Theorem 8.2. The following
theorem on exact A-harmonic forms is essential in the proof. For details, see
[34].
Theorem 8.3. Let n ≥ 3 and let η be a weakly exact A-harmonic ℓ-form, ℓ ∈
{2, . . . , n − 1}, on Rn \ B̄ n such that
Z
(8.4) |η|n/ℓ = ∞.
Rn \B̄ n (2)

Then there exists γ = γ(n, a, b) > 0 such that


Z
1
(8.5) lim inf γ |η|n/ℓ > 0.
r→∞ r B n (r)\B̄ n (2)

Here a and b are as in (8.1) and (8.2).

Sketch of the proof of Theorem 8.2. Let us first consider some exceptions.
For Riemannian surfaces the result is classical and follows from the uniformiza-
tion theorem and the measurable Riemann mapping theorem, see [34, Theorem
3]. For n ≥ 3 we may give a bound for the first cohomology using a well-known
result of Varopoulos on the fundamental group and n-hyperbolicity. For details,
see [34, Theorem 4]. Hence we may restrict our discussion to dimensions n ≥ 3
and to cohomology dimensions ℓ ≥ 2.
Let n ≥ 3 and 2 ≤ ℓ ≤ n − 1, and suppose that f : B n \ {0} → N does not
have a limit at the origin. Without changing the notation we precompose f with
a sense-preserving Möbius mapping σ such that σ(Rn \ B̄ n ) = B n \ {0}. Let us
now show that dim H ℓ (N ) is bounded from above by a constant depending only
on n and K. We fix p-harmonic ℓ-forms ξi generating H ℓ (N ), with p = n/ℓ. This
can be done by a result of Scott [42]. Furthermore, we may assume that forms
ξi are uniformly separated and uniformly bounded in Lp , that is, kξi − ξj kp ≥ 1
and kξi kp = 1 for every i and j.
A local version [34, Theorem 6] of the value distribution result of Mattila and
Rickman yields that
Z Z
∗ n/ℓ
(8.6) |f ξ| ∼ Jf
B n (r)\B n (2) B n (r)\B n (2)

for large radii r. Using Theorem 8.3 and a decomposition technique due to
Rickman, we find a radius R and a decomposition of the annulus B n (R) \ B n (2)
148 Ilkka Holopainen and Pekka Pankka IWQCMA05

into domains quasiconformally equivalent to B n in such a way that we have a


quasiregular embedding ψ : B n → Rn \ B̄ n (2) with properties
Z Z 1/4
(8.7) Jf & Jf
ψB n (1/2) B n (R)\B n (2)

and
Z Z
(8.8) Jf . Jf .
ψB n B n (R)\B n (2)

Combining (8.6) with (8.7) and (8.8), we have that forms ϕ∗ f ∗ ξi are uniformly
bounded in Lp (B n ) and uniformly separated in Lp (B n (1/2)). By compactness,
the number of forms is bounded by a constant depending on data.
Remark 8.4. The use of A-harmonic forms in the proof of Theorem 8.2 is very
similar to their use in the proof of Theorem 8.1. Also Theorem 8.3 corresponds
to a theorem of Bonk and Heinonen ([4, Theorem 1.11]).

References
[1] L. V. Ahlfors, Conformal invariants: topics in geometric function theory, McGraw-Hill
Series in Higher Mathematics, McGraw-Hill Book Co., New York, 1973.
[2] R. Bishop and R. Crittenden, Geometry of manifolds, Pure Appl. Math. 15, Academic
Press, New York, 1964.
[3] B, Bojarski and T. Iwaniec, Analytical foundations of the theory of quasiconformal map-
pings in Rn , Ann. Acad. Sci. Fenn. Ser. A I Math. 8 (1983), 257–324.
[4] M. Bonk and J. Heinonen, Quasiregular mappings and cohomology, Acta Math. 186
(2001), 219–238.
[5] M. Bonk and J. Heinonen, Smooth quasiregular mappings with branching, Publ. Math.
Inst. Hautes Études Sci. 100 (2004), 153–170.
[6] P. Buser, A note on the isoperimetric constant, Ann. Sci. École Norm. Sup. (4) 15 (1982),
213–230.
[7] I. Chavel, Riemannian geometry: A modern introduction, Cambridge University Press,
1993.
[8] J. Cheeger, M. Gromov, and M. Taylor, Finite propagation speed, kernel estimates for
functions of the Laplace operator, and the geometry of complete Riemannian manifolds,
J. Differential Geom. 17 (1982), 15–53.
[9] T. Coulhon, I. Holopainen, and L. Saloff-Coste, Harnack inequality and hyperbolicity for
subelliptic p-Laplacian with applications to Picard type theorems, Geom. Funct. Anal. 11
(2001), 1139– 1191.
[10] A. Eremenko and J. L. Lewis, Uniform limits of certain A-harmonic functions with
applications to quasiregular mappings, Ann. Acad. Sci. Fenn. Ser. A I Math. 16 (1991),
361–375.
[11] I. Fonseca and W. Gangbo, Degree theory in analysis and applications, Oxford Lecture
Series in Mathematics and its Applications, Clarendon Press, Oxford, 1995.
[12] F. W. Gehring, Rings and quasiconformal mappings in space, Trans. Amer. Math. Soc.
103 (1962), 353–393.
[13] S. Granlund, Harnack’s inequality in the borderline case, Ann. Acad. Sci. Fenn. Ser. A I
Math. 5 (1980), 159–163.
[14] M. Gromov, Metric structures for Riemannian and non-Riemannian spaces, Progress in
Mathematics 152, Birkhäuser Boston Inc., Boston, 1999; Structures métriques pour les
p-Laplace operator, quasiregular mappings, and Picard-type theorems 149

variétés riemanniennes. Lecture notes by P. Pansu et J. Lafontaine, Cedic Nathan, Paris,


1981.
[15] J. Heinonen, The branch set of a quasiregular mapping, in Proceedings of the Interna-
tional Congress of Mathematicians, Vol. II (Beijing, 2002), Higher Ed. Press, Beijing,
2002, 691–700.
[16] J. Heinonen, T. Kilpeläinen, and O. Martio, Nonlinear Potential Theory of Degenerate
Elliptic Equations, Oxford Mathematical Monographs, Clarendon Press, Oxford - New
York - Tokyo, 1993.
[17] I. Holopainen, Nonlinear potential theory and quasiregular mappings on Riemannian
manifolds, Ann. Acad. Sci. Fenn. Ser. A I Math. Diss. 74 (1990), 1–45.
[18] I. Holopainen, Positive solutions of quasilinear elliptic equations on Riemannian mani-
folds, Proc. London Math. Soc. (3) 65 (1992), 651–672.
[19] I. Holopainen, Volume growth, Green’s function, and parabolicity of ends, Duke Math. J.
97 (1999), 319-346.
[20] I. Holopainen, Asymptotic Dirichlet problem for the p-Laplacian on Cartan-Hadamard
manifolds, Proc. Amer. Math. Soc. 130 (2002), 3393–3400.
[21] I. Holopainen and P. Koskela, Volume growth and parabolicity, Proc. Amer. Math. Soc.
129 (2001), 3425–3435.
[22] I. Holopainen and P. Pankka, Mappings of finite distortion: Global homeomorphism the-
orem, Ann. Acad. Sci. Fenn. Math. 29 (2004), 59–80.
[23] I. Holopainen and P. Pankka, A big Picard type theorem for quasiregular mappings into
manifolds with many ends, Proc. Amer. Math. Soc. 130 (2005), 1143-1150.
[24] I. Holopainen and S. Rickman, A Picard type theorem for quasiregular mappings of Rn
into n-manifolds with many ends, Rev. Mat. Iberoamericana 8 (1992), 131-148.
[25] I. Holopainen and S. Rickman, Classification of Riemannian manifolds in nonlinear po-
tential theory, Potential Anal. 2 (1993), 37–66.
[26] I. Holopainen and S. Rickman, Ricci curvature, Harnack functions, and Picard type
theorems for quasiregular mappings in Analysis and topology, World Sci. Publishing,
River Edge, NJ, 1998, 315–326.
[27] T. Iwaniec, p-harmonic tensors and quasiregular mappings, Ann. of Math. 136 (1992),
589–624.
[28] T. Iwaniec and G. Martin, Quasiregular mappings in even dimensions, Acta Math. 170
(1993), 29–81.
[29] T. Iwaniec and G. Martin, Geometric function theory and non-linear analysis, Oxford
Mathematical Monographs, Oxford University Press, Oxford, 2001.
[30] T. Iwaniec, C. Scott, and B. Stroffolini, Nonlinear Hodge theory on manifolds with bound-
ary, Ann. Mat. Pura Appl. 177 (1999), 37–115.
[31] D. Jerison, The Poincaré inequality for vector fields satisfying Hörmander’s condition,
Duke Math. J. 53 (1986), 503–523.
[32] R. Kaufman, J. T. Tyson, and J.-M. Wu, Smooth quasiregular maps with branching in
Rn , Publ. Math. Inst. Hautes Études Sci. 101 (2005), 209–241.
[33] J. L. Lewis, Picard’s theorem and Rickman’s theorem by way of Harnack’s inequality,
Proc. Amer. Math. Soc. 122 (1994), 199–206.
[34] P. Pankka, Quasiregular mappings from a punctured ball into compact manifolds, Con-
form. Geom. Dyn. (to appear).
[35] T. Rado and P. V. Reichelderfer, Continuous transformations in analysis, Die
Grundlehren der mathematische Wissenschaften 75, Springer-Verlag, Berlin, 1955.
[36] Yu. G. Reshetnyak, Estimates of the modulus of continuity for certain mappings, Sibirsk.
Mat. Z. 7 (1966), 1106–1114.
[37] Yu. G. Reshetnyak, Spatial mappings with bounded distortion, Sibirsk. Mat. Z. 8 (1967),
629–659.
150 Ilkka Holopainen and Pekka Pankka IWQCMA05

[38] Yu. G. Reshetnyak, Liouville’s conformal mapping theorem under minimal regularity
hypotheses, Sibirsk. Mat. Z. 8 (1967), 835–840.
[39] Yu. G. Reshetnyak, Space mappings with bounded distortion, Translations of Mathemat-
ical Monographs 73, Amer. Math. Soc., Providence, RI, 1989.
[40] S. Rickman, On the number of omitted values of entire quasiregular mappings, J. Analyse
Math. 37 (1980), 100-117.
[41] S. Rickman, Quasiregular mappings, Ergebnisse der Mathematik und ihrer Grenzgebiete
26, Springer-Verlag, Berlin-Heidelberg-New York, 1993.
[42] C. Scott, Lp theory of differential forms on manifolds, Trans. Amer. Math. Soc. 347
(1995), 2075–2096.
[43] J. Serrin, Local behavior of solutions of quasi-linear equations, Acta Math. 111 (1964),
247–302.
[44] N. Varopoulos, Potential theory and diffusion on Riemannian manifolds in Conference
on Harmonic Analysis in Honor of Antoni Zygmund, Vol. 2, Wadsworth Math. Ser.,
Wadsworth, Belmont, Calif., 1983, 821–837.
[45] J. Väisälä, Discrete open mappings on manifolds, Ann. Acad. Sci. Fenn. Ser. A I Math.
392 (1966), 1–10.
[46] S. T. Yau, Isoperimetric constants and the first eigenvalue of a compact manifold, Ann.
Sci. École Norm. Sup. 8 (1975), 159–171.
[47] V. A. Zorich, The theorem of M. A. Lavrent’ev on quasiconformal mappings in space,
Mat. Sb. 74 (1967) 417–433.
[48] V.A. Zorich, Quasiconformal immersions of Riemannian manifolds and a Picard type
theorem, Functional Analysis and Its Appl. 34 (2000), 188-196.
[49] V.A. Zorich, Asymptotics of the admissible growth of the coefficient of quasiconformality
at infinity and injectivity of immersions of Riemannian manifolds, Publ. Inst. Math.
(Beograd) (N.S.) 75(89) (2004), 53–57.

Ilkka Holopainen Address: Department of Mathematics and Statistics, P.O. Box 68,
FIN-00014 University of Helsinki, Finland
E-mail: ilkka.holopainen@helsinki.fi

Pekka Pankka Address: Department of Mathematics and Statistics, P.O. Box 68,
FIN-00014 University of Helsinki, Finland
E-mail: pekka.pankka@helsinki.fi
Proceedings of the International Workshop on Quasiconformal
Mappings and their Applications (IWQCMA05)

Hyperbolic-type metrics

Henri Lindén

Abstract. The article is a status report on the contemporary research of


hyperbolic-type metrics, and considers progress in the study of the classes
of isometry- and bilipschitz mappings with respect to some of the presented
metrics. Also, the Gromov hyperbolicity question is discussed.

Keywords. Hyperbolic-type metric, intrinsic metric, isometry problem, bilipschitz-


mapping, Gromov hyperbolic space.
2000 MSC. Primary 30F45, Secondary 30C65,53C23.

Contents

1. Introduction 151
2. The metrics 152
3. Isometries and bilipschitz-mappings 157
4. Gromov hyperbolicity 160
References 163

1. Introduction
In geometric function theory there are many different distance functions around,
which — to a greater or lesser degree — resemble the classical hyperbolic metric.
Some of these are defined by geometric means, some by implicit formulas, and
many by integrating over certain weight functions.

What all these metrics have in common, is that they are defined in some
proper subdomain D ( Rn , and are strongly affected by the geometry of the do-
main boundary. Thus we should actually speak of families of metrics {dD }D(Rn ,
since the metric looks different in each domain, even though the defining formula
might be the same. In the literature, however, one usually abuses notation and
speaks only of “the metric d”, which we will do here also. The metrics typically
have negative curvature, ie. the geodesics, if they exist, avoid the boundary.
Most of the metrics described here also have an invariance property in the sense
that
(1.1) dD (x, y) = df (D) (f (x), f (y)),
152 H. Lindén IWQCMA05

for mappings f belonging to some fixed class, say similarities, Möbius transfor-
mations, or conformal mappings.
Many of the metrics, especially those with simple explicit formulas, have been
developed as tools for estimating other, more hard-to-handle metrics, such as the
quasihyperbolic metric, which is probably the one most commonly used metric
presented in this text. It has found applications in many branches of analysis,
and is a very natural generalization of the classical hyperbolic metric to any
domain D and dimension n ≥ 2. It has some flaws though, in most cases one
cannot compute it, and actually very little is known about the metric itself. The
difficulty of explicit computation is typical also for some other metrics, and for
this reason we have a lot of “similar” metrics around, which in many cases are
equivalent to each other; a handy feature, if one metric is suited for your study,
but the other is not. Here we will try to give a survey on some of these metrics.

2. The metrics
The classical starting point is the hyperbolic geometry developed by Poincaré
and Lobachevsky in the early 19:th century. Poincaré used the unit ball as
domain for his model, and Lobachevsky used the half space. These models
turned out to be equivalent in the sense that Möbius transformations between
them are isometries.
2.1. Definition. Let D ∈ {Hn , Bn }, and define a weight (or density) function
w : D → R by
1 2
w(z) = , for D = Hn and w(z) = , for D = Bn .
dist(z, ∂D) 1 − |z|2
Then the hyperbolic length ℓρ (γ) of a curve γ is defined by
Z
(2.2) ℓρ (γ) = ℓρ,D (γ) = w(z) |dz|,
γ

where |dz| denotes the length element. After this, the hyperbolic distance ρD is
defined for all x, y ∈ D by
Z
(2.3) ρD (x, y) = inf ℓρ,D (γ) = inf w(z) |dz|,
γ∈Γxy γ∈Γxy γ
where Γxy is the family of all rectifiable curves joining x and y within D.

The above method to define metrics is frequently used. In fact, to get a


completely new metric, the only thing that needs to be changed is the weight
function. After that, the length and the new distance function are defined as in
(2.2) and (2.3), respectively. The benefit of defining a metric d like this is that
it will automatically be intrinsic, in other words, it will be its own inner metric
ˆ This means that
d.
(2.4) dD (x, y) = dˆD (x, y) := inf ℓd,D (γ).
γ∈Γxy
Hyperbolic-type metrics 153

2.5. Geodesics. When a metric is defined in the way described above, one
might ask how to find the curve γ ∈ Γxy giving the desired infimum (which —
if it is found — is in fact a minimum). In general, this can be far from trivial,
even if such a curve exists. Curves minimizing the distance in this way are
called geodesics or geodesic segments. Another way of characterizing a geodesic,
is that it satisfies the triangle inequality with equality, ie. the curve γ ∈ Γxy is a
geodesic, if for all u, v, w ∈ |γ| properly ordered, we have

dD (u, w) = dD (u, v) + dD (v, w).

We denote by JdD [x, y] the geodesic segment between x and y in (D, d). This
segment may, however, not be unique, and no particular choice is made here. A
metric space in which geodesic segments exist between any two given points, is
called a geodesic metric space. If, in addition, the geodesic is unique, the space
is totally geodesic. Naturally a geodesic metric is always intrinsic.

2.6. Hyperbolic metric in G. It is also possible to define the hyperbolic


metric in a general simply connected subdomain G of the plane, since by the
Riemann mapping theorem there exists a conformal mapping f : G → f G = B2 .
Then the metric density is defined by

ρG (z) = ρB2 (f (z))|f ′ (z)|.

From the Schwarz lemma it follows that ρG is independent of the choice of f .


We then define the hyperbolic metric hG by (2.3) using the density ρG . This
definition automatically gives the hyperbolic metric the invariance property of
(1.1) for the class of conformal mappings. Note, that while in the classical cases
we use the traditional notation ρBn and ρHn for the hyperbolic metric, in general
domains we use hG . Also, note that when the dimension n ≥ 3, every conformal
mapping is a Möbius mapping, so it is not possible to extend the definition to
general simply connected domains like above. In fact, for n ≥ 3 the hyperbolic
metric is defined only in Bn and Hn .

The hyperbolic metric is well understood, and the geodesic flow is known. In
fact, in the classical models Bn and Hn the geodesics are known to be circular
arcs orthogonal to the boundary, and in other domains the geodesics simply are
induced by the conformal mapping. Moreover, for the classical cases there are
explicit formulas to calculate the value of the hyperbolic metrics in terms of
euclidean distances. For a comprehensive study on the classical cases, see the
book by Beardon [Be1]. The hyperbolic metric in an arbitrary domain has been
studied by F. Gehring, K. Hag and A. Beardon, see eg. the articles [Be3] and
[GeHa1].
154 H. Lindén IWQCMA05

Bn

n
y* H
y
x

y
x
x* x* y*

Figure 1: Hyperbolic geodesics in Bn and Hn .

One way to calculate the hyperbolic distance, is to use the absolute cross-ratio
defined by
|a − c||b − d|
|a, b, c, d| = , a, b, c, d ∈ Rn .
|a − b||c − d|
One can prove that, if C is the circle containing JρBn [x, y] or JρHn [x, y] and
{x∗ , y ∗ } = C ∩ ∂Bn or {x∗ , y ∗ } = C ∩ ∂Hn in the same order as in Figure 1, then
(2.7) ρBn (x, y) = log |x∗ , x, y, y ∗ | = ρHn (x, y).
Other explicit formulas have also been derived, see the book [Be1].

2.8. The Apollonian metric. The formula in (2.7) makes one wonder whether
a similar approach could be generalized to any domain D ( Rn . It turns out
that this is very much possible; the Apollonian distance in a domain D is defined
by
|z − x| |w − y|
(2.9) αD (x, y) = sup log ,
z,w∈∂D |z − y| |w − x|
for all x, y ∈ D. This is a metric, unless the boundary is the subset of a circle
or a line, in which case it is only a pseudo-metric, ie. the metric axiom d(x, y) =
0 ⇒ x = y need not hold.
Geometrically the Apollonian metric can be thought of in the following way:
an Apollonian circle (or sphere, when n ≥ 3) with respect to the pair (x, y), is a
set  

n |z − x|
Bx,y,q = z ∈ R =q .
|z − y|
Then the Apollonian metric is
αD (x, y) = log qx qy ,
where qx and qy are the ratios of the largest possible balls Bx,y,qx and By,x,qy still
contained in D.
Hyperbolic-type metrics 155

D
w
Bx
x

z y

By

Figure 2: The Apollonian balls approach.


The Apollonian metric is invariant in Möbius mappings in the sense of (1.1).
It is an easy exercise in geometry to show that in the case D = Hn the points z
and w are actually the points x∗ and y ∗ in (2.7), and thus ρHn = αHn .
The Apollonian metric has been studied in [GeHa2] and [Se], but especially
by P. Hästö and Z. Ibragimov in a series of articles, see e.g. [Hä1],[Hä2],[HäIb]
and [Ib].

The Apollonian metric is in a way a convenient construction with a clear


geometric interpretation, but as a shortcoming it has its lack of geodesics. In the
article [HäLi] some work is done to overcome this problem, by introducing the
half-Apollonian metric, defined by

|x − z|
(2.10)
ηD (x, y) = sup log ,
z∈∂D |y − z|
for all x, y ∈ D. The geometric intuition here is the same as for the Apollonian
metric, Indeed, instead of log qx qy we have
ηD (x, y) = log max{qx , qy }.
This metric is a only similarity invariant, but instead it has more geodesics than
the Apollonian metric. It is also bilipschitz equivalent to the Apollonian metric,
in fact
1
αD (x, y) ≤ ηD (x, y) ≤ αD (x, y).
2
2.11. The quasihyperbolic metric. The quasihyperbolic metric is perhaps
the most well-known and frequently used of the metrics considered here. It was
developed by F. Gehring and his collaborators in the 70’s. It is defined by the
method of 2.1 using
1
w(z) = , z∈D
dist(z, ∂D)
as weight function. It is immediate that for D = Hn the quasihyperbolic metric
coincides with the hyperbolic metric ρHn . The quasihyperbolic metric is invariant
under the class of similarity mappings.
156 H. Lindén IWQCMA05

The quasihyperbolic metric is well-behaved in many senses: the weight func-


tion is quite simple and it is a natural generalization of the hyperbolic metric.
Also, it is known to be geodesic for any domain D ( Rn [GeOs]. One of the
shortcomings of the metric is that in general the geodesics are not easy to deter-
mine. Besides the half-space Hn , the geodesics are known in the punctured space
Rn \ {z} and in the ball Bn , see [MaOs]. Recently the geodesics were determined
also for the punctured ball Bn \ {0}, and planar angular domains
Sϕ = {(r, θ) | 0 < θ < ϕ}, 0 < ϕ < 2π,
see [Li1].
2.12. Distance-ratio metrics. As the quasihyperbolic metric cannot be ex-
plicitly evaluated in the case of general domains, a typical way to overcome this
problem is to approximate it by another metric, often one of the distance-ratio
metrics or j-metrics. (Actually, by their construction also the Apollonian and
half-Apollonian metrics could be described as “distance-ratio metrics”). There
are two versions of these. The first, introduced by F. Gehring, is defined by
  
e |x − y| |x − y|
(2.13) jD (x, y) = log 1 + 1+ , x, y, ∈ D.
dist(x, ∂D) dist(y, ∂D)
The other one is defined by
 
|x − y|
(2.14) jD (x, y) = log 1 + . x, y, ∈ D.
dist(x, ∂D) ∧ dist(y, ∂D)
is a modification due to M. Vuorinen.

The two metrics have much in common, but also important differences, which
will be discussed further in Sections 2 and 3. Both are similarity invariant, and
can be used to estimate the quasihyperbolic metric. The metrics satisfy the
relation
jD (x, y) ≤ e
jD (x, y) ≤ 2 jD (x, y), x, y ∈ D.
The lower bound for the quasihyperbolic metric is given by the inequality
jD (x, y) ≤ kD (x, y)
proved in [GePa], which holds for points x, y in any proper subdomain D. The
upper bound holds for so called uniform domains, which is a wide class of domains
introduced in [MaSa].
2.15. Definition. A domain D ( Rn is called uniform or A-uniform, if there
exists a number A ≥ 1 such that the inequality
kD (x, y) ≤ A jD (x, y)
holds for all x, y ∈ D.

There are many definitions for uniform domains around, see eg. [Ge], so often
many “nice” domains can be shown to be uniform by other means, and so one
has access to the inequality in 2.15. However, typically very little can be said
about the constant A. These matters have been studied in [Li1].
Hyperbolic-type metrics 157

The j-metric defined in (2.14) has another important connection to the quasi-
hyperbolic metric. The quasihyperbolic metric is namely the inner metric of the
j-metric, in the sense of (2.4). In other words
kD (x, y) = inf ℓj,D (γ).
γ∈Γxy

Since the j-metric fails to be intrinsic, it cannot be geodesic either. In fact,


the j-metric has geodesics only in some special cases, see [HäIbLi, 3.7]. Very
little is known about the geodesic segments of the ej-metric, although it can be
conjectured that there is not much of them either.

3. Isometries and bilipschitz-mappings


As pointed out earlier, most of the hyperbolic-type metrics defined in this
article satisfy some kind of invariance property, that is, they satisfy the equality
(1.1) for some class of mappings f . Typically this invariance property follows
almost directly from the definition of the metric, for instance, it is easy to see
from the formulas (2.9) and (2.10) that the Apollonian metric is Möbius-invariant
and the half-Apollonian metric is similarity invariant. The interesting question
mostly regards the other implication. Is the class of “natural candidates” the
only mappings which give isometries in the metric in question? And what are the
“near-isometries”, that is, the bilipschitz mappings? There are still many open
ends regarding these questions, though some progress has been made recently.
3.1. Definition. Let D and D′ = f (D) be domains such that equipped with
distances dD and dD′ they are metric spaces. Then a continuous mapping f : D →
D′ is said to be L-bilipschitz in (or with respect to) the metric d if for all x, y ∈ D
we have
1
dD (x, y) ≤ dD′ (f (x), f (y)) ≤ L dD (x, y)
L
for some L ≥ 1. If the above inequality holds with L = 1, f is a d-isometry.
3.2. “One-point” and “two-point” metrics. In general, the hyperbolic-
type metrics can be divided into length-metrics, defined by means of integrating
a weight function, and point-distance metrics. The point-distance metrics may
again be classified by the number of boundary points used in their definition.
So for instance the j-metric and the half-Apollonian metric would be ‘one-point
metrics”, whereas the ej, and the Apollonian metrics are “two-point metrics”.
Actually also the length metrics can be characterized in the same way, by
looking at their weight function. Then the quasihyperbolic metric is a one-point
metric. An example of a two-point length metric is the so called Ferrand metric
n
σD , see [Fe1]. It is defined for a domain D ( R with card ∂D ≥ 2, using the
weight function
|a − b|
wD (x) = sup , x ∈ D \ {∞}.
a,b∈∂D |x − a||x − b|
158 H. Lindén IWQCMA05

This metric is Möbius invariant and coincides with the hyperbolic metric on Hn
and Bn . Moreover, it is bilipschitz equivalent to the quasihyperbolic metric by
the inequality
(3.3) kD (x, y) ≤ σD (x, y) ≤ 2 kD (x, y), x, y ∈ D.

Naturally one would expect the one-point point-distance metrics to be the


easiest ones to study. In fact, much can be said about these metrics when it
comes to the isometry question. The half-Apollonian metric has recently been
studied in [HäLi]. A point x ∈ D is called circularly accessible if there exists
a ball B ⊂ G such that x ∈ ∂B. If x is circularly accessible by two distinct
balls whose surfaces intersect at more than one point, it is called a corner point,
otherwise a regular point.
3.4. Theorem. Let D ( Rn be a domain which has at least n regular boundary
points which span a hyperplane. Then f : D → Rn is a homeomorphic η-isometry
if and only if it is a similarity mapping.

Furthermore, it was shown that Möbius mappings are in fact 2-bilipschitz


with respect to ηD .
For the j-metric, some results can be found in [HäIbLi], and in fact in a
slightly more general setting. The implications for the j-metric can be expressed
as follows;
3.5. Corollary. Let D ( Rn . Then f : D → Rn is a j-isometry if and only if
(1) f is a similarity, or
(2) D = Rn \ {a} and, up to similarity, f is the inversion in a sphere centered
at a.

Since ĵD = kD , it immediately follows that every isometry of the j-metric is


an isometry of the quasihyperbolic metric, of course in this case that does not
provide us with very much new information. However, a similar relation is true
for the Seittenranta metric δD defined in [Se] by
 
|x − y||a − b|
δD (x, y) = log 1 + sup , x, y ∈ D,
a,b∈∂D |a − x||b − y|

which is also studied in [HäIbLi]. Namely, here we have that δˆD = σD , so we


directly see that this is a Möbius invariant metric. In [Se] it is proved that at least
Euclidean bilipschitz mappings are bilipschitz with respect to δ. The converse is
not true, as can be shown by the counterexample
f : B2 \ {0} → B2 \ {0}, f (x) = |x| · x.
However, in [Se] it was shown that every bilipschitz δ-mapping is a quasicon-
formal mapping, and that every δ-isometry is conformal with respect to the
Euclidean metric (and thus Möbius for n ≥ 3). In [HäIbLi] it was shown that
also for n = 2 in fact the δ-isometries are exactly the Möbius mappings.
Hyperbolic-type metrics 159

For the j-metric there are still many open problems regarding the bilipschitz
question. It is well known (see [Vu]), that an Euclidean L-bilipschitz mapping is
L2 -bilipschitz with respect to the j (and k) metric.
For the Apollonian metric the isometry and bilipschitz questions have been
studied by several authors. The work was started by Beardon in [Be2], and con-
tinued by Gehring and Hag in [GeHa2] where they studied Apollonian bilipschitz
mappings. They proved the following theorem.
3.6. Theorem. Let D ( R2 be a quasidisk and f : D → D′ be an Apollonian
bilipschitz mapping.
(1) If D′ is a quasidisk, then f is quasiconformal in D and f = g|D , where
2 2
g : R → R is quasiconformal.
(2) If f is quasiconformal in D, then D′ is a quasidisk and f = g|D , where
2 2
g : R → R is quasiconformal.

In [Hä2] the above property (1) was generalized to hold also for n ≥ 3. In the
same article also a condition was introduced which determines when a Euclidean
bilipschitz mapping is also Apollonian bilipschitz. In the article [HäIb] it is
shown that for n = 2 the Apollonian isometries are exactly restrictions of Möbius
mappings.
For the quasihyperbolic metric the question regarding the isometries has long
been open. In [MaOs] it was shown that every kD -isometry is a conformal map-
ping. A similar proof gives the same result for Ferrand’s metric σD . However, in
[Hä3] it is shown that if the boundary of the domain is regular enough (C 3 , or C 2
unless the domain is either strictly convex or has strictly convex complement),
then the quasihyperbolic isometries are exactly the similarity mappings.
3.7. Conformal modulus. We conclude by introducing two new metrics which
are particularly interesting regarding the question of bilipschitz mappings. Let Γ
n
be a family of curves in R . By F(Γ) we denote the family of admissible functions,
n
that is, non-negative Borel-measurable functions ρ : R → R such that
Z
ρ ds ≥ 1
γ

for each locally rectifiable curve γ ∈ Γ. The n-modulus or the conformal modulus
of Γ is defined by Z
M(Γ) = Mn (Γ) = inf ρn dm,
ρ∈F (Γ) Rn
where m is the n-dimensional Lebesgue measure. It is a conformal invariant,
i.e. if f : G → G′ is a conformal mapping and Γ is a curve family in G, then
M(Γ) = M(f Γ).
n
For E, F, G ⊂ R we denote by ∆(E, F ; D) the family of all closed non-
n
constant curves joining E and F in D, that is, γ : [a, b] → R belongs to
∆(E, F ; D) if one of γ(a), γ(b) belongs to E and the other to F , and furthermore
γ(t) ∈ D for all a < t < b.
160 H. Lindén IWQCMA05

Now we will define two new conformal invariants in the following way. For
n
x, y ∈ D ( R λD is defined by

λD (x, y) = inf M ∆(Cx , Cy ; D) ,
Cx ,Cy

where Cz = γz [0, 1) and γz : [0, 1] → D is a curve such that z ∈ |γz | and γz (t) →
∂D when t → 1 and z = x, y. Correspondingly,

µD (x, y) = inf M ∆(Cxy , ∂D; D) ,
Cxy

where Cxy is such that Cxy = γ[0, 1] and γ is a curve with γ(0) = x and γ(1) = y.

It is not difficult to show that both quantities µD and λD are conformal invari-
ants, and that µD is a metric (often called the modulus metric) when cap ∂D > 0,
1/(1−n)
see [Gá]. λD is not a metric, but λ∗D = λD introduced in [Fe2] is, as long as
the boundary of the domain has more then two points.
One of the interesting feature regarding these metrics is that both are easily
seen — by their definitions — to be conformal invariants. Moreover, the following
can be shown (see [Vu, 10.19]);
3.8. Theorem. If f : D → D′ = f D is a quasiconformal mapping, then
(1) µD (x, y)/L ≤ µf D (f (x), f (y) ≤ L µD (x, y),
(2) λ∗D (x, y)/L1/(n−1) ≤ λ∗f D (f (x), f (y)) ≤ L1/(n−1) λ∗D (x, y)
hold for all x, y ∈ D, where L = max{KI (f ), KO (f )} is the maximal dilatation
of f .

It is not known if the class of bilipschitz mappings with respect to µ or λ∗


includes any other than quasiconformal mappings.

4. Gromov hyperbolicity
One way of telling “how hyperbolic” a metric in fact is, is to study whether
it satisfies hyperbolicity in the sense of M. Gromov. Classically such spaces
have been studied in the geodesic case, and then a space is said to be Gromov
δ-hyperbolic if for all triples of geodesics Jd [x, y], Jd [y, z] and Jd [x, z] we have
that
dist(w, Jd [y, z] ∪ Jd [z, x]) ≤ δ
for all w ∈ Jd [x, y], i.e. if all geodesic triangles are δ-thin.
4.1. The Gromov product. In non-geodesic spaces, however, we are con-
strained to use the definition involving the Gromov product. This can be defined
for two points x, y ∈ D with respect to a base point w by setting
1 
(x|y)w = d(x, w) + d(y, w) − d(x, y) .
2
Hyperbolic-type metrics 161

A space is then said to be Gromov δ-hyperbolic if it satisfies the inequality


(x|z)w ≥ (x|y)w ∧ (y|z)w − δ
for all x, y, z ∈ D and a base point w ∈ D. A space is said to be Gromov
hyperbolic if it is Gromov δ-hyperbolic for some δ. Sometimes one wants to use
the equivalent definition for Gromov hyperbolicity

(4.2) d(x, z) + d(y, w) ≤ d(x, w) + d(y, z) ∨ d(x, y) + d(z, w) + 2δ.

Recently the study of Gromov hyperbolicity has become quite popular, and
even hyperbolicity results on particular metrics in geometric function theory
have been developed by a number of authors. A systematic study of the different
metrics is made easier by the fact that Gromov hyperbolicity is preserved by
certain classes of mappings, so called rough isometries. We say that two metrics
d and d′ are roughly isometric if there exists a positive constant C such that
d(x, y) − C ≤ d′ (x, y) ≤ d(x, y) + C.
It is immediately clear from the definition (4.2) that roughly isometric metrics
are Gromov hyperbolic in the same domains. Moreover, we say that two metrics
are (A, C)-quasi-isometric if there is A ≥ 1, C ≥ 0 such that
A−1 d(x, y) − C ≤ d′ (x, y) ≤ A d(x, y) + C.
Also quasi-isometries (and thus bilipschitz mappings) are known to preserve Gro-
mov hyperbolicity, provided that the spaces are geodesic.
Naturally we would want the hyperbolic metric itself to be Gromov hyperbolic
also, and in fact it is, with constant δ = log 3, as is shown in [CoDePa]. One of
the more interesting and general results is one from the comprehensive study of
M. Bonk, J. Heinonen and P. Koskela [BoHeKo], where it is shown that for a
uniform domain D the space (D, kD ) is always Gromov hyperbolic.
For many of the other metrics Gromov hyperbolicity is easily proved or dis-
proved using the results from [Hä4]. Namely, it turns out that the e
j-metric is
Gromov hyperbolic in every proper subdomain of Rn , whereas the j-metric is
Gromov hyperbolic only in Rn \ {a}. Then, using inequalities
jD (x, y) − log 3 ≤ ηD (x, y) ≤ jD (x, y),
j̃D (x, y) − log 9 ≤ αD (x, y) ≤ j̃D (x, y),
and
αD (x, y) ≤ δD (x, y) ≤ αD (x, y) + log 3
we immediately get some results by rough isometry, that is, the results in Table
1 regarding the Apollonian, half-Apollonian and Seittenranta metrics. For prov-
ing Gromov hyperbolicity of the Ferrand metric one can use geodesity, Gromov
hyperbolicity of the quasihyperbolic metric, and the bilipschitz equivalence in
(3.3).
Finally, for the µ and λ∗ metrics positive results regarding Gromov hyperbol-
icity are shown in [Li2].
162 H. Lindén IWQCMA05

4.3. Theorem. The metric space (Bn , λ∗Bn ) is Gromov δ-hyperbolic, with Gro-
mov constant
 1−n
1  1 ωn−1  1−n
1 
δ ≤ 21 ωn−1
2
log 64
3
+ 4 log λ n ≤ 2 2
log 64
3
+ 4(log 2 + n − 1) ,
where ωn−1 denotes the (n − 1)-dimensional surface area of S n−1 and λn is the
Grötzsch constant. Also, any simply connected proper subdomain D ( R2 is
Gromov δ-hyperbolic with respect to the metric λ∗G , where
log 5462
δ≤ ≈ 1.3696.

4.4. Theorem. The metric space (Bn , µBn ) is Gromov δ-hyperbolic, with Gro-
mov constant
δ ≤ 2n−1 cn log 12,
where cn is the spherical cap inequality constant, see [Vu]. Especially, every
simply connected domain D ( R2 is Gromov hyperbolic with
2 log 12
δ≤ ≈ 1.5819.
π

4.5. Theorem. The metric space (Rn \ {z}, λ∗Rn \{z} ) is Gromov hyperbolic, with
1 1 
δ ≤ 2ωn−1
n−1
log 18λ2n ≤ 2ωn−1
n−1
log 72 + 2n − 2 .

As the below table indicates, the j-metric and the half-Apollonian metric are
the only metrics of the ones discussed here which fail to be Gromov hyperbolic
in most cases. These results indicate that these metrics are in a way “too easy”,
or have too little structure for satisfying Gromov hyperbolicity. On the other
hand, in other contexts that is one of their strongest features, as has been seen
in earlier sections.

Domain condition Proved where

kD D uniform [BoHeKo]
hD n = 2 all domains defined, n ≥ 3, D = Bn , Hn [CoDePa] and conf. invariance
αD All domains D ( Rn [Hä4] and rough isometry
ηD Only D = Rn \ {z}, δ = log 9 [Hä4],[HäLi]
jD Only D = Rn \ {z}, δ = log 9 [Hä4]
j̃D All domains D ( Rn [Hä4]
δD All domains D ( Rn [Hä4],[Se]
σD D uniform, for D = Bn δ = log 3 [Fe1],[BoHeKo]
λ∗D D = Bn , Rn∗ , n = 2 simply conn. domains [Li2]
µD D = Bn , n = 2 simply conn. domains [Li2]

Table 1: Gromov hyperbolicity of some metrics.


Hyperbolic-type metrics 163

References
[Be1] A. F. Beardon: The geometry of discrete groups. Graduate Texts in Mathe-
matics, Vol. 91, Springer-Verlag, Berlin-Heidelberg-New York, 1982.
[Be2] A. F. Beardon: The Apollonian metric of a domain in Rn . Quasiconformal
mappings and analysis (Ann Arbor, Michigan, 1995), Springer-Verlag, New
York, (1998), 91–108.
[Be3] A. F. Beardon: The hyperbolic metric in a rectangle II. Ann. Acad. Sci. Fenn.
Math., 28, (2003), 143–152.
[BoHeKo] M. Bonk, J. Heinonen and P. Koskela: Uniformizing Gromov hyperbolic
spaces. Astérisque 270, 2001, 1–99.
[CoDePa] M. Coornaert, T. Delzant and A. Papadopoulos: Géométrie et théorie
des groupes. Lecture Notes in Mathematics, Vol. 1441 Springer-Verlag, Berlin,
1990. (French, english summary).
[Fe1] J. Ferrand: A characterization of quasiconformal mappings by the behavior of
a function of three points. Proceedings of the 13th Rolf Nevanlinna Colloquium
(Joensuu, 1987; I. Laine, S. Rickman and T. Sorvali (eds.)), Lecture Notes in
Mathematics Vol. 1351, Springer-Verlag, New York, (1988), 110–123.
[Fe2] J. Ferrand: Conformal capacity and extremal metrics. Pacific J. Math. 180,
no. 1, (1997), 41–49.
[Gá] I. S. Gál: Conformally invariant metrics and uniform structures. Indag. Math.
22, (1960), 218–244.
[Ge] F. W. Gehring: Characteristic properties of quasidisks. Les Presses de
l’Universite de Montreal, Montreal, 1982.
[GeHa1] F. W. Gehring and K. Hag: A bound for hyperbolic distance in a quasidisk.
Computational methods and function theory (Nicosia, 1997), 233–240, Ser. Ap-
prox. Decompos. 11, World Sci. Publishing, River Edge, NJ, 1999.
[GeHa2] F. W. Gehring and K. Hag: The Apollonian metric and quasiconformal
mappings. In the tradition of Ahlfors and Bers (Stony Brook, NY, 1998), 143–
163, Contemp. Math. 256, Amer. Math. Soc., Providence, RI, 2000.
[GeOs] F. W. Gehring and B. G. Osgood: Uniform domains and the quasi-
hyperbolic metric. J. Anal. Math. 36 (1979), 50–74.
[GePa] F. W. Gehring and B. Palka: Quasiconformally homogeneous domains. J.
Anal. Math. 30 (1976), 172–199.
[Hä1] P. Hästö: The Apollonian metric: uniformity and quasiconvexity. Ann. Acad.
Sci. Fenn. Math., 28, (2003), 385–414.
[Hä2] P. Hästö: The Apollonian metric: limits of the approximation and bilipschitz
properties. Abstr. Appl. Anal., 20, (2003), 1141–1158.
[Hä3] P. Hästö: Isometries of the quasihyperbolic metric. In preparation (2005).
Available at http://www.helsinki.fi/˜hasto/pp/
[Hä4] P. Hästö: Gromov hyperbolicity of the jG and e jG metrics. Proc. Amer. Math.
Soc. 134, (2006), 1137–1142.
[HäIb] P. Hästö and Z. Ibragimov: Apollonian isometries of planar domains are
Möbius mappings. J. Geom. Anal. 15, no. 2, (2005), 229–237.
[HäIbLi] P. Hästö, Z. Ibragimov and H. Lindén: Isometries of relative metrics. In
preparation (2004). Available at http://www.helsinki.fi/˜hlinden/pp.html
[HäLi] P. Hästö and H. Lindén: Isometries of the half-Apollonian metric. Compl.
Var. Theory Appl. 49, no. 6 (2004), 405–415.
n
[Ib] Z. Ibragimov: On the Apollonian metric of domains in R Compl. Var. Theory
Appl. 48, no. 10, (2003), 837–855.
[Li1] H. Lindén: Quasihyperbolic geodesics and uniformity in elementary domains.
Ann. Acad. Sci. Fenn. Math. Diss. 146, (2005), 1–52.
164 H. Lindén IWQCMA05

[Li2] H. Lindén: Gromov hyperbolicity of certain invariant metrics. In preparation.


Available as preprint in Reports Dept. Math. Stat. Univ. Helsinki 409, (2005),
University of Helsinki.
[MaOs] G. Martin and B. Osgood: The quasihyperbolic metric and the associated
estimates on the hyperbolic metric. J. Anal. Math. 47 (1986), 37–53.
[MaSa] O. Martio and J. Sarvas: Injectivity theorems in plane and space. Ann.
Acad. Sci. Fenn. Ser. A I Math. 4, (1978/79), 383–401.
[Se] P. Seittenranta: Möbius-invariant metrics. Math. Proc. Camb. Phil. Soc.
125 (1999), 511–533.
[Vu] M. Vuorinen: Conformal geometry and quasiregular mappings. Lecture Notes
in Mathematics, Vol. 1319 Springer-Verlag, Berlin, 1988.

Henri Lindén E-mail: hlinden@iki.fi


Address: P.O.Box 68, 00014 University of Helsinki, FINLAND
Proceedings of the International Workshop on Quasiconformal
Mappings and their Applications (IWQCMA05)

Geometric properties of hyperbolic geodesics

W. Ma and D. Minda

Abstract. In the unit disk D hyperbolic geodesic rays emanating from the
origin and hyperbolic disks centered at the origin exhibit simple geometric
properties. The goal is to determine whether analogs of these geometric prop-
erties remain valid for hyperbolic geodesic rays and hyperbolic disks in a simply
connected region Ω. According to whether the simply connected region Ω is a
subset of the unit disk D, the complex plane C or the extended complex plane
(Riemann sphere) C∞ = C ∪ {∞}, the geometric properties are measured rel-
ative to the background geometry on Ω inherited as a subset of one of these
classical geometries, hyperbolic, Euclidean and spherical. In a simply con-
nected hyperbolic region Ω ⊂ C hyperbolic polar coordinates possess global
Euclidean properties similar to those of hyperbolic polar coordinates about
the origin in the unit disk if and only if the region is Euclidean convex. For
example, the Euclidean distance between travelers moving at unit hyperbolic
speed along distinct hyperbolic geodesic rays emanating from an arbitrary
common initial point is increasing if and only if the region is convex. A simple
consequence of this is the fact that the two ends of a hyperbolic geodesic in a
convex region cannot be too close. Exact analogs of this Euclidean separating
property of hyperbolic geodesic rays hold when Ω lies in either the hyperbolic
plane D or the spherical plane C∞ .

Keywords. hyperbolic metric, hyperbolic geodesics, hyperbolic disks, Eu-


clidean convexity, hyperbolic convexity, spherical convexity.
2000 MSC. Primary 30F45; Secondary 30C55.

Contents

1. Introduction 166
2. Hyperbolic polar coordinates in the unit disk 167
3. Hyperbolic polar coordinates in a disk or half-plane 169
4. Hyperbolic polar coordinates in simply connected regions 172
5. Euclidean convex univalent functions 173
6. Euclidean convex regions 175
7. Spherical geometry 177
8. Spherically convex univalent functions 179
9. Spherically convex regions 182
Version October 19, 2006.
The second author was supported by a Taft Faculty Fellowship.
166 Ma and Minda IWQCMA05

10. Hyperbolic geometry 183


11. Concluding remarks 185
References 186

1. Introduction
The results in this expository paper are adapted from [16] and [17] and concern
geometric properties of hyperbolic geodesics in a simply connected hyperbolic
region Ω and, to a lesser extent, geometric properties of hyperbolic disks. These
two references contain many results not mentioned here and as well as the details
that are not presented in this largely expository article. In particular, proofs not
given in this article can be found in these two references. There are three different
cases to consider according to whether the region Ω is a subset of the hyperbolic
plane D, the Euclidean plane C, or the spherical plane C∞ = C ∪ {∞}. Two
geometries on the region Ω will be considered. First, the intrinsic hyperbolic
geometry on Ω and, second, the geometry that Ω inherits as a subset of the
hyperbolic, Euclidean or spherical plane.
Here is a rough description of the types of behavior of hyperbolic geodesics
that we will consider. Fix a point w0 ∈ Ω. For θ ∈ R, let ρ(w0 , Ω) denote the
hyperbolic geodesic ray emanating from w0 that has unit Euclidean tangent eiθ
at w0 and let w0 (s, θ) be the hyperbolic arc length parametrization of this geo-
desic. Under what conditions does the point w0 (s, θ) move monotonically away
from w0 when s increases? Here motion away from w0 is measured relative to
the background distance. For example, if Ω lies in the Euclidean plane, this
means the Euclidean distance |w0 (s, θ) − w0 | should increase with s. The second
type of behavior we consider is whether the background distance between dis-
tinct geodesic rays increases as points move along these rays. In the Euclidean
case we inquire whether the Euclidean distance |w0 (s, θ1 ) − w0 (s, θ2 )| increases
with s when eiθ1 6= eiθ2 . Intuitively, one can think of two travelers departing
from w0 at the same time along different hyperbolic geodesic rays and traveling
at unit hyperbolic speed along the geodesics and asking whether the travelers
separate monotonically in the Euclidean sense. Finally, we investigate the shape
of hyperbolic circles relative to the background geometry. The main concern is
whether hyperbolic circles are convex curves relative to the background geome-
try. Hyperbolic rays emanating from a point w0 together with hyperbolic circles
centered at w0 form the coordinate grid for hyperbolic polar coordinates in Ω, so
our work can be interpreted as studying geometric properties of the hyperbolic
polar coordinate grid relative to the background geometry.
A descriptive outline of the paper follows. Hyperbolic polar coordinates in
the unit disk are defined in Section 2, while Section 3 extends hyperbolic polar
coordinates to any Euclidean disk or half-plane. Simple Euclidean properties of
the hyperbolic polar coordinate grid in any disk or half-plane are established as
the model for future investigations. Hyperbolic polar coordinates for a simply
Geometric properties of hyperbolic geodesics 167

connected region are introduced in Section 4. Loosely speaking hyperbolic polar


coordinates can be transferred from the unit disk to a simply connected region
Ω by using the Riemann Mapping Theorem; a conformal map f : D → Ω is a
hyperbolic isometry. Characterizations of Euclidean convex univalent functions
are discussed in Section 5. in Section 6 these characterizations are used to es-
tablish Euclidean properties of hyperbolic polar coordinates in Euclidean convex
regions and to show that these properties characterize Euclidean convex regions.
The remainder of the paper is devoted to analogs of these results in the spherical
and hyperbolic planes. The spherical plane is introduced in Section 7 along with
the notion of a spherically convex region. The results for regions in the spherical
plane parallels the Euclidean context. Spherically convex univalent functions are
presented in Section 8. The reader should note the number of parallels between
spherically convex univalent functions and Euclidean convex univalent functions.
The results for spherically convex univalent functions seem more involved than
those for Euclidean convex univalent functions; the more complicated nature of
formulas relating to spherically convex univalent functions is due to the fact that
the spherical metric has curvature 1 while the Euclidean metric has curvature
0. Nonzero curvature causes the appearance of extra terms. Applications of
some results for spherically convex functions to the behavior of the hyperbolic
coordinate grid in a spherically convex region are given in Section 9. Section 10
considers the behavior of the hyperbolic polar coordinate grid for hyperbolically
convex regions in the unit disk. Because of the strong similarity with the previ-
ous cases for Euclidean convexity and spherical convexity, we present a concise
discussion of the results. The reader should note that some theorems for hy-
perbolically convex univalent functions formally differ from those for spherically
convex univalent functions by certain sign changes; these alterations in sign are
due to the fact that the hyperbolic plane has curvature −1 while the spherical
plane has curvature 1. The brief final section directs the reader to some other
situations in function theory in which there are parallel results for the hyperbolic,
Euclidean and spherical planes.

2. Hyperbolic polar coordinates in the unit disk


We begin by recalling the unit disk as a model of the hyperbolic plane. The
hyperbolic metric on the unit disk D = {z : |z| < 1} is
2|dz|
λD (z)|dz| = .
1 − |z|2
The hyperbolic metric has curvature −1; that is,
△ log λD (z)
− = −1,
λ2Ω (z)
where z = x + iy and
∂2 ∂2 ∂2
∆= + = 4
∂x2 ∂y 2 ∂z∂ z̄
168 Ma and Minda IWQCMA05

denotes the usual Laplacian. For any piecewise smooth curve γ in D the hyper-
bolic length of γ is given by
Z
ℓD (γ) = λD (z)|dz|.
γ

The hyperbolic distance between z, w ∈ D is defined by


dD (z, w) = inf ℓD (γ),
where the infimum is taken over all piecewise smooth paths γ in D that join z
and w. In fact,

−1 a − b

dD (a, b) = 2 tanh .
1 − ¯ba
The group A(D) of conformal automorphisms of the unit disk is the set of holo-
morphic isometries of the hyperbolic metric and also of the hyperbolic distance.
A path γ joining z to w is called a hyperbolic geodesic arc if dD (z, w) = ℓD (γ).
The (hyperbolic) geodesic through z and w is C ∩ D, where C is the unique Eu-
clidean circle (or straight line) that passes through z and w and is orthogonal to
the unit circle ∂D. If γ is any piecewise smooth curve joining z to w in D, then
the hyperbolic length of γ is dD (z, w) if and only if γ is the arc of C in D that
joins z and w. A hyperbolic disk in the unit disk is DD (a, r) = {z : dD (a, z) < r},
where a ∈ D is the hyperbolic center and r > 0 is the hyperbolic radius. A hy-
perbolic disk in D is Euclidean disk with closure contained in D. In fact, DD (a, r)
is the Euclidean disk with center c and radius R, where

a 1 − tanh2 (r/2) (1 − |a|2 ) tanh(r/2)
c= and R = .
1 − |a|2 tanh2 (r/2) 1 − |a|2 tanh2 (r/2)
For more details about hyperbolic geometry on the unit disk the reader should
consult [1].
Hyperbolic polar coordinates on the unit disk relative to a specified pole or
center are defined as follows. Fix a point a in D, called the pole or center for
polar coordinates based at a. For θ in R let ρθ (a, D) = ρθ (a) denote the unique
hyperbolic geodesic ray emanating from a that is tangent to the Euclidean unit
vector eiθ at a. For θ = 0 the Euclidean unit tangent vector is 1 and ρ0 (a)
is called the horizontal hyperbolic geodesic emanating from a because the unit
tangent vector at a is horizontal. Of course, ρθ+2nπ (a) = ρθ (a) for all n in Z.
Let s 7→ za (s, θ), 0 ≤ s < +∞, be the hyperbolic arc length parametrization of
ρθ (a). This means
∂za (s, θ) eiΘ(s,θ)
(2.1) = ,
∂s λD (za (s, θ))
where eiΘ(s,θ) is a Euclidean unit tangent to ρθ (a) at the point za (s, θ). For fixed θ
the point za (s, θ) moves along the geodesic ray ρθ (a) with unit hyperbolic speed.
Two hyperbolic geodesic rays with distinct unit tangent vectors at a are disjoint
except for their common initial point and D = ∪{ρθ : 0 ≤ θ < 2π}. For each z in
D \ {a} there is a unique geodesic ray ρθ (a) with 0 ≤ θ < 2π that contains z, so
Geometric properties of hyperbolic geodesics 169

there exist unique s > 0 and θ in [0, 2π) with za (s, θ) = z. The hyperbolic polar
coordinates of the point z relative to the center or pole at a are the ordered pair
(s, θ), where za (s, θ) = z. The first coordinate, s = dD (a, z), is the hyperbolic
distance from a to z and the second polar coordinate, θ, is the angle between the
horizontal hyperbolic geodesic ray ρ0 (a) and the ray ρθ (a) that contains z at the
pole a. The hyperbolic circle with hyperbolic center a and hyperbolic radius s is
cD (a, s) = {z : dD (a, z) = s}. Note that each geodesic ray ρθ (a) is orthogonal to
every hyperbolic circle cΩ (a, s). Thus, the coordinate grid for hyperbolic polar
coordinates based at a consists of hyperbolic geodesics emanating from a and
hyperbolic circles centered at a. In terms of hyperbolic polar coordinates
λ2D (z)(dx2 + dy 2 ) = ds2 + sinh2 (s)dθ2 .
For a = 0, ρθ (0) is the radial segment [0, eiθ ) with hyperbolic arc length parametriza-
tion z0 (s, θ) = tanh(s/2)eiθ and
∂z0 (s, θ) eiθ 1 − |z0 (s, θ)|2
(2.2) = = z0 (s, θ).
∂s λD (z0 (s, θ)) 2|z0 (s, θ)|

Hyperbolic polar coordinates about the origin can be transported to any other
center in the unit disk by a hyperbolic isometry. Recall that each conformal
automorphism of D is an isometry of the hyperbolic metric and the hyperbolic
distance. For a ∈ D the Möbius transformation f (z) = (z + a)/(1 + āz) is a
conformal automorphism of D that sends the origin to a and f ′ (0) = (1−|a|2 ) > 0.
The fact that f ′ (0) > 0 insures that f (ρθ (0)) = ρθ (a) for all θ ∈ R and so
za (s, θ) = f (z0 (s, θ)) provides an explicit hyperbolic arc length parametrization
of ρa (θ):
tanh(s/2)eiθ + a
za (s, θ) = .
1 + ā tanh(s/2)eiθ

Trivially, the Euclidean distance from a = 0 to z0 (s, θ) is an increasing function


of s for each fixed θ and the Euclidean distance between z0 (s, θ1 ) and z0 (s, θ2 ) is
an increasing function of s when eiθ1 6= eiθ2 . It is plausible that these Euclidean
properties remain valid for any center a ∈ D. Rather than investigating these
assertions for the special case of the unit disk, we wait to consider the analogous
questions in any disk or half-plane. Also, hyperbolic circles centered at the origin
are Euclidean circles.

3. Hyperbolic polar coordinates in a disk or half-plane


We let ∆ denote any Euclidean disk or half-plane when it is not necessary
to distinguish between the cases; otherwise, we use D for a Euclidean disk and
H for a Euclidean half-plane. Given ∆ there is a Möbius transformation f that
maps ∆ onto the unit disk. Then the hyperbolic metric on ∆ is given by
λ∆ (z) = λD (f (z))|f ′ (z)|.
170 Ma and Minda IWQCMA05

This defines the hyperbolic density λ∆ independent of the Möbius map of ∆ onto
the unit disk. If D = {z : |z − a| < r}, then
2r|dz|
λD (z)|dz| = .
r2 − |z − a|2
If H is any half-plane, then
|dz|
λH (z)|dz| = ,
d(z, ∂H)
where d(z, ∂H) denotes the Euclidean distance from z to the boundary of H. In
particular, for the upper half-plane H = {z : Im (z) > 0},
|dz|
λH (z)|dz| = .
Im (z)
Because Möbius transformations map circles onto circles, hyperbolic geodesics
in a disk or half-plane are arcs of circles orthogonal to the boundary. Also,
hyperbolic disks are Euclidean disks with closure contained in the disk or half-
plane. Any Möbius map from ∆ onto D is an isometry from ∆ with the hyperbolic
metric to D with the hyperbolic metric. See [1] for details.
Hyperbolic polar coordinates are defined on ∆ analogous to the definition
for the unit disk. Fix a point w0 in ∆. For θ in R let ρθ (w0 , ∆) denote the
unique hyperbolic geodesic ray emanating from w0 that is tangent to eiθ at w0 .
ρ0 (w0 , ∆) is called the horizontal hyperbolic geodesic emanating from w0 since its
unit tangent vector at w0 is horizontal. When w0 and ∆ are fixed, we often write
ρθ in place of ρθ (w0 , ∆). Of course, ρθ+2nπ = ρθ for all n in Z. Let s 7→ w0 (s, θ),
0 ≤ s < +∞, be the hyperbolic arc length parametrization of ρθ . This means
∂w0 (s, θ) eiΘ(s,θ)
(3.1) = ,
∂s λ∆ (w0 (s, θ))
where eiΘ(s,θ) is a Euclidean unit tangent to ρθ at the point w0 (s, θ). Because
∆ = ∪{ρθ : 0 ≤ θ < 2π}, for each w in ∆ \ {w0 } there is a unique geodesic ray ρθ ,
0 ≤ θ < 2π, that contains w. Hence, there exist unique s > 0 and θ in [0, 2π) with
w0 (s, θ) = w. The hyperbolic polar coordinates for the point w relative to the
center or pole at w0 are (s, θ). The coordinate s = d∆ (w0 , w) is the hyperbolic
distance from w0 to w and θ is the angle between the horizontal hyperbolic
geodesic ray ρ0 and the ray ρθ at w0 . The hyperbolic circle with hyperbolic
center w0 and hyperbolic radius s is c∆ (w0 , s) = {w : d∆ (w0 , w) = s}. The
coordinate grid for hyperbolic polar coordinates consists of hyperbolic geodesics
emanating from w0 and hyperbolic circles centered at w0 . If f : D → ∆ is the
Möbius transformation with f (0) = w0 and f ′ (0) > 0, then w0 (s, θ) = f (z0 (s, θ)).
As we noted in the preceding section when a point in the unit disk moves
away from the origin along a hyperbolic geodesic, the Euclidean distance from
the origin increases and points along distinct geodesics separate monotonically
in the Euclidean sense. In fact these properties hold for any disk or half-plane
and for any center of hyperbolic polar coordinates.
Geometric properties of hyperbolic geodesics 171

Theorem 3.1. Let ∆ be any Euclidean disk or half-plane in C and w0 ∈ ∆.


(a) For each θ ∈ R, |w0 (s, θ) − w0 | is increasing for s ≥ 0.
(b) For eiθ2 6= eiθ1 , |w0 (s, θ1 ) − w0 (s, θ2 )| is an increasing function of s ≥ 0.

Proof. If f : D → ∆ is a Möbius mapping with f (0) = w0 and f ′ (0) > 0, then


w0 (s, θ) = f (z0 (s, θ)). Suppose
az + b
f (z) =,
cz + d
where ad − bc = 1. Because ∆ is a Euclidean disk or half-plane, ∞ does not
lie in ∆. Consequently, −d/c, the preimage of ∞, cannot lie in D; equivalently,
|c| ≤ |d|. Since ad − bc = 1, this implies d 6= 0. Also, w0 = f (0) = b/d.
(a) If D(s) = log |w0 (s, θ) − w0 | = log |f (z0 (s, θ)) − w0 |, then by using (2.2)
we obtain
f ′ (z0 (s, θ)) ∂z0 (s, θ)
D′ (s) = Re
f (z0 (s, θ)) − w0 ∂s
1 − |z0 (s, θ)| 2
z0 (s, θ)f ′ (z0 (s, θ))
(3.2) = Re .
2|z0 (s, θ)| f (z0 (s, θ)) − w0
From
1 z
f ′ (z) = and f (z) − w0 = ,
(cz + d)2 d(cz + d)
we obtain
zf ′ (z) d
= .
f (z) − w0 cz + d
Then for z ∈ D
zf ′ (z) dc̄z̄ + |d|2
(3.3) Re = Re >0
f (z) − w0 |cz + d|2
because |c| ≤ |d| and |z| < 1. Thus, (3.3) and (3.2) imply D(s) is increasing for
s ≥ 0, so |w0 (s, θ) − w0 | is increasing for s ≥ 0.
(b) We assume −π/2 ≤ θ1 = −θ < 0 < θ2 = θ ≤ π/2; the general case can be
reduced to this situation by performing a rotation. If
E(s) = log |w0 (s, θ) − w0 (s, −θ)| = log |f (z0 (s, θ)) − f (z0 (s, −θ))|,
then
′ f ′ (z0 (s, θ)) ∂z0∂s
(s,θ)
− f ′ (z0 (s, −θ)) ∂z0 (s,−θ)
∂s
E (s) = Re .
f (z0 (s, θ)) − f (z0 (s, −θ))
Because of (2.2) and |z0 (s, θ)| = |z0 (s, −θ)|, we obtain
 
′ 1 − |z0 (s, θ)|2 z0 (s, θ)f ′ (z0 (s, θ)) − z0 (s, −θ)f ′ (z0 (s, −θ))
(3.4) E (s) = Re .
2|z0 (s, θ)| f (z0 (s, θ)) − f (z0 (s, −θ))
Direct calculation produces
zf ′ (z) − ζf ′ (ζ) d2 − c2 ζz
= .
f (z) − f (ζ) (cz + d)(cζ + d)
172 Ma and Minda IWQCMA05

Set t = c/d. Then


zf ′ (z) − ζf ′ (ζ) 1 − t2 ζz
=
f (z) − f (ζ) (1 + tz)(1 + tζ)
 
1 1 − tζ 1 − tz
= + .
2 1 + tζ 1 + tz
Because (1 − w)/(1 + w) has positive real part for w ∈ D and |tζ|, |tz| < 1, we
conclude that for all z, ζ ∈ D
 ′ 
zf (z) − ζf ′ (ζ)
(3.5) Re > 0.
f (z) − f (ζ)
Hence, (3.4) and (3.5) imply E ′ (s) > 0 for s ≥ 0, so that |w0 (s, θ) − w0 (s, −θ)|
is an increasing function of s ≥ 0.

4. Hyperbolic polar coordinates in simply connected


regions
A region Ω in the complex plane C is hyperbolic if C \ Ω contains at least two
points. The hyperbolic metric on a hyperbolic region Ω is denoted by λΩ (w)|dw|
and is normalized to have curvature
△ log λΩ (w)
− = −1.
λ2Ω (w)
If f : D → Ω is any holomorphic universal covering projection, then the density
λΩ of the hyperbolic metric is determined from
2
(4.1) λΩ (f (z))|f ′ (z)| = .
1 − |z|2
For a, b in Ω the hyperbolic distance between these points is
Z
dΩ (a, b) = inf λΩ (w)|dw|,
δ
where the infimum is taken over all piecewise smooth paths δ in Ω joining a and
b. A path γ connecting a and b is a hyperbolic geodesic arc if
Z
dΩ (a, b) = λΩ (w)|dw|.
γ

A hyperbolic geodesic always exists, but need not be unique when Ω is multiply
connected. Given a ∈ Ω and r > 0, DΩ (a, r) = {z ∈ Ω : dΩ (a, z) < r} is the
hyperbolic disk with hyperbolic center a and hyperbolic radius r.
When Ω is simply connected, any conformal mapping f : D → Ω is an isometry
from the hyperbolic metric on D to the hyperbolic metric on Ω. In this case f
maps hyperbolic geodesics onto hyperbolic geodesics and hyperbolic disks onto
hyperbolic disks. If Ω is multiply connected, then a covering f is only a local
isometry, not an isometry.
Geometric properties of hyperbolic geodesics 173

Suppose Ω is a simply connected hyperbolic region, w0 ∈ Ω and f : D → Ω


is the unique conformal mapping with f (0) = w0 and f ′ (0) > 0. We can relate
hyperbolic polar coordinates on Ω with pole at w0 to those on D with pole at
the origin by using f . Tangent vectors for geodesic rays can be expressed in
terms of this conformal mapping. Because f is an isometry from the hyperbolic
metric on D to the hyperbolic metric on Ω and f ′ (0) > 0, w0 (s, θ) = f (z0 (s, θ))
is the hyperbolic arc length parametrization of ρθ (w0 , Ω) and the tangent vector
to ρθ (w0 , Ω) is
∂w0 (s, θ) f ′ (z0 (s, θ))eiθ
(4.2) = .
∂s λD (z0 (s, θ))
Thus,
∂w0 (0, θ) f ′ (0)eiθ
= ,
∂s 2
so that s 7→ w0 (s, θ) is parallel to eiθ at w0 . By making use of (4.1) we find
∂w0 (s, θ) f ′ (z0 (s, θ)) eiθ ei(ϕ(s,θ)+θ)
(4.3) = ′ = ,
∂s |f (z0 (s, θ))| λΩ (f (z0 (s, θ))) λΩ (f (z0 (s, θ)))
where eiϕ(s,θ) = |ff ′ (z

0 (s,θ))
(z0 (s,θ))|
. If arg f ′ (z) denotes the unique branch of the argument
of f ′ that vanishes at w0 , then ϕ(s, θ) = arg f ′ (z0 (s, θ)). From (3.1) and (4.3) we
obtain
(4.4) eiΘ(s,θ) = ei(ϕ(s,θ)+θ) .
In a similar manner, hyperbolic disks in Ω are the images of hyperbolic disks in D;
explicitly, if f : D → Ω is a conformal map with f (0) = w0 , then f (DD (0, r)) =
DΩ (w0 , r).

5. Euclidean convex univalent functions


Several characterizations of Euclidean convex univalent functions are needed
for our investigation of hyperbolic polar coordinates. We recall two classical
characterizations of Euclidean convex and starlike univalent functions. First, a
locally univalent holomorphic function f defined on D is a conformal map onto
a Euclidean convex region if and only if [2, p 42]
zf ′′ (z)
(5.1) 1 + Re ≥0
f ′ (z)
for z ∈ D. Second, if f (0) = w0 , then a holomorphic function f defined on D
maps D conformally onto a region starlike with respect to w0 if and only if [2, p
41]
zf ′ (z)
(5.2) Re ≥0
f (z) − w0
for z ∈ D.
174 Ma and Minda IWQCMA05

Theorem 5.1. Suppose f is holomorphic and locally univalent on D. f is Eu-


clidean convex univalent on D if and only if
zf ′ (z) − ζf ′ (ζ)
(5.3) Re >0
f (z) − f (ζ)
for all z, ζ in D.

Proof. We present the short proof. Suppose f is Euclidean convex univalent on


D. Then ([27] and [30])
 
2zf ′ (z) z+ζ
(5.4) Re − >0
f (z) − f (ζ) z − ζ
for z, ζ in D. If we interchange the roles of z and ζ in (5.4) and then add the
two inequalities, we obtain
zf ′ (z) − ζf ′ (ζ)
2 Re > 0,
f (z) − f (ζ)
which is equivalent to (5.3).
Conversely, suppose (5.3) holds for all z, ζ in D. Since
zf ′ (z) − ζf ′ (ζ) zf ′′ (z)
lim =1+ ′ ,
ζ→z f (z) − f (ζ) f (z)
we obtain (5.1). Hence, f is Euclidean convex univalent on D.
Theorem 5.2. If f is a normalized, f (0) = 0 and f ′ (0) = 1, Euclidean convex
univalent function on D and θ ∈ (0, π/2], then
 cos θ
2|z| sin θ iθ −iθ 2|z| sin θ 1 + |z|
(5.5) ≤ |f (e z) − f (e z)| ≤ .
1 + 2|z| cos θ + |z|2 1 − |z|2 1 − |z|
The lower bound is best possible for all θ ∈ (0, π/2] and the upper bound is sharp
for θ = π/2.

Proof. We sketch the idea of the proof. Fix θ in (0, π/2] and consider the
function
f (eiθ z) − f (e−iθ z) f (eiθ z) − f (e−iθ z)
g(z) = = .
eiθ − e−iθ 2i sin θ
From
zg ′ (z) eiθ zf ′ (eiθ z) − e−iθ zf ′ (e−iθ z)
= ,
g(z) f (eiθ z) − f (e−iθ z)
Theorem 5.1 implies that g is starlike with respect to the origin on D because
(5.2) holds with w0 = 0. If
X ∞
f (z) = z + an z n ,
n=2
then ∞
X sin nθ
g(z) = z + an z n .
n=2
sin θ
Geometric properties of hyperbolic geodesics 175

As f is convex univalent, |a2 | ≤ 1 [2]. Hence,



|g ′′ (0)| sin 2θ
= |a2 | ≤ 2 cos θ.
2 sin θ
Then [3] gives the inequalities in (5.5).
Corollary 5.3. If f is a normalized, f (0) = 0 and f ′ (0) = 1, Euclidean convex
univalent function on D, then for ϕ ∈ (0, π/2]
2r 2r
(5.6) ≤ |f (reiϕ ) − f (−reiϕ )| ≤ .
1+r 2 1 − r2
These bounds are sharp.
Example 5.4. If K(z) = z/(1 − z), then
eiθ z e−iθ z
K(eiθ z) − K(e−iθ z) = −
1 − eiθ z 1 − e−iθ z
(2i sin θ)z
= .
1 − (2 cos θ)z + z 2
This shows that the lower bound in (5.5) is sharp for K(z) when z = −r, r is in
(0, 1), for any θ in (0, π/2]. For θ = π/2 the upper bound in (5.5) is sharp for
the function K when z = ir, r in (0, 1). Also,
2r
K(r) − K(−r) =
1 − r2
and
2ir
K(ir) − K(−ir) = ,
1 + r2
so both bounds in (5.6) are sharp.

6. Euclidean convex regions


We establish various Euclidean properties for hyperbolic polar coordinates in
Euclidean convex regions; in fact, these Euclidean properties characterize con-
vex regions. Throughout this section we employ the notation of Section 4. In
particular, f will always denote a conformal map of D onto Ω with f (0) = w0
and f ′ (0) > 0. We show that for each fixed θ, the point w0 (s, θ) moves mono-
tonically away from w0 in the Euclidean sense. We give sharp upper and lower
bounds on |w0 (s, θ) − w0 | in terms of s and λΩ (w0 ). Also, in any convex re-
gion Ω distinct hyperbolic geodesic rays separate monotonically in the Euclidean
sense; this means that for eiθ2 6= eiθ1 , the distance |w0 (s, θ1 ) − w0 (s, θ2 )| is an
increasing function of s. We give sharp upper and lower bounds on the difference
|w0 (s, θ1 )−w0 (s, θ2 )|. These (and other) Euclidean properties of hyperbolic polar
coordinates characterize convex regions.
For example, a classical result of Study [29] implies that for every w0 ∈ Ω
each hyperbolic circle cΩ (w0 , s) is a Euclidean convex curve when Ω is convex.
The result of Study asserts that if f is a Euclidean convex univalent function,
176 Ma and Minda IWQCMA05

then f ({z : |z| < r}) is Euclidean convex for 0 < r < 1. Conversely, if every
hyperbolic circle is Euclidean convex, then Ω is an increasing union of Euclidean
convex regions and so is Euclidean convex.
Theorem 6.1. Let Ω be a simply connected hyperbolic region in C.
(a) If Ω is Euclidean convex and w0 ∈ Ω, then for each θ in R, |w0 (s, θ) − w0 | is
an increasing function of s and
1 − e−s es − 1
(6.1) ≤ |w0 (s, θ) − w0 | ≤ .
λΩ (w0 ) λΩ (w0 )
These bounds are best possible.
(b) Suppose that for every w0 in Ω and for each θ in R, |w0 (s, θ) − w0 | is an
increasing function of s. Then Ω is Euclidean convex.

The proof of Theorem 6.1 is given in [16].


Example 6.2. For the upper half-plane H, λH (w) = 1/Im(w). Then for w0 = i,
w0 (s, π/2) = i + i(es − 1), w0 (s, −π/2) = i − i(1 − e−s ) and 1/λH (i) = 1, so the
upper and lower bounds are best possible.
Theorem 6.3. Suppose Ω is a simply connected hyperbolic region in C.
(a) If Ω is Euclidean convex, w0 ∈ Ω and eiθ2 6= eiθ1 , then |w0 (s, θ1 ) − w0 (s, θ2 )|
is an increasing function of s ≥ 0 and
2 sin θ tanh s
(6.2) ≤ |w0 (s, θ1 ) − w0 (s, θ2 )|λΩ (a) ≤ 2es cos θ sin θ sinh s,
1 + cos θ tanh s
where θ = (θ2 − θ1 )/2.
(b) If for some w0 in Ω and all eiθ2 6= eiθ1 , |w0 (s, θ1 ) − w0 (s, θ2 )| is an increasing
function of s ≥ 0, then Ω is Euclidean convex.
Proof. We sketch the proof of (a). First, by translating Ω if necessary, we may
assume w0 = 0. Next, by rotating Ω about the origin if needed, we may assume
−π/2 ≤ θ1 = −θ < 0 < θ2 = θ ≤ π/2. Then
|w0 (s, θ) − w0 (s, −θ)| = |f (z(s, θ)) − f (z(s, −θ))|.
All of the quantities involved in the theorem are invariant under translation and
rotation. If
E(s) = log |w0 (s, θ) − w0 (s, −θ)| = log |f (z(s, θ)) − f (z(s, −θ))|,
then
′ f ′ (z(s, θ)) ∂z(s,θ)
∂s
− f ′ (z(s, −θ)) ∂z(s,−θ)
∂s
E (s) = Re .
f (z(s, θ)) − f (z(s, −θ))
Because of (2.2) and |z(s, θ)| = |z(s, −θ)|, we obtain
 
′ 1 − |z(s, θ)|2 z(s, θ)f ′ (z(s, θ)) − z(s, −θ)f ′ (z(s, −θ))
(6.3) E (s) = Re .
2|z(s, θ)| f (z(s, θ)) − f (z(s, −θ))
Suppose Ω is Euclidean convex. Then f is a Euclidean convex univalent function
and so (5.3) implies E ′ (s) > 0. Hence, |w0 (s, θ) − w0 (s, −θ)| is an increasing
function of s ≥ 0. Next, we establish (6.2). The function f /f ′ (0) is a normalized
Geometric properties of hyperbolic geodesics 177

Euclidean convex univalent function so (5.5) with r = tanh(s/2) gives the bounds
(6.2) for θ2 = θ and θ1 = −θ since f ′ (0) = 2/λΩ (a).
Corollary 6.4. Suppose Ω 6= C is a Euclidean convex region and w0 is a point
of Ω. Then |w0 (s, θ) − w0 (s, θ + π)| is increasing and
λΩ (a)
(6.4) tanh(s) ≤ |w0 (s, θ) − w0 (s, θ + π)| ≤ sinh(s).
2
Both bounds are sharp for a half-plane.

Proof. This is the special case of the theorem in which θ2 = θ + π and θ1 = θ.


It corresponds to two hyperbolic geodesic rays emanating from w0 in opposite
directions.

The lower bound in (6.2) has a simple geometric consequence. It gives


2 tan(θ/2)
lim |w0 (s, θ1 ) − w0 (s, θ2 )| ≥ .
s→+∞ λΩ (w0 )
For any Euclidean convex region this shows that the ‘ends’ of distinct hyperbolic
geodesic rays emanating from w0 cannot be too close. In particular, (6.4) implies
that the two ends of a single hyperbolic geodesic cannot be closer than 2/λΩ (w0 )
for any point w0 on the geodesic. This inequality is sharp for the upper half-
plane; we only consider the special case in which θ1 = 0 and θ2 = π. Consider
w0 = ib, where b > 0. Then λH (w0 ) = 1/b. For the hyperbolic geodesic γ through
ib that meets R in ±b,
2
lim |w0 (s, 0) − w0 (s, π)| = 2b = .
s→+∞ λH (w0 )

7. Spherical geometry
We discuss the geometry of the spherical plane C∞ with the chordal distance
χ, the spherical metric σ(z) |dz| and the induced spherical distance dσ .
The extended complex plane C∞ is sometimes called the Riemann sphere
because stereographic projection transforms C∞ into the unit sphere. Let S be
the unit sphere {x ∈ R3 : ||x|| = 1} in R3 , and let n = (0, 0, 1) be the ‘north
pole’. The stereographic projection ϕ of C∞ onto S is defined as follows. We
regard the complex plane C as a subset of R3 by identifying z = x + iy with the
point (x, y, 0). For z in C, the line through z = (x, y, 0) and n meets S at n and
at a second point ϕ(z). This defines ϕ on C, and we set ϕ(∞) = n. It is easy to
see that if z = x + iy ∈ C, then
 
2x 2y |z|2 − 1
ϕ(x + iy) = , , .
|z|2 + 1 |z|2 + 1 |z|2 + 1
Observe that ϕ(0) = (0, 0, −1), the south pole, and that ϕ(z) = z = (x, y, 0) if
and only if |z| = 1.
178 Ma and Minda IWQCMA05

The chordal distance χ is obtained by the following procedure. Use ϕ to


transfer points in C∞ to S; then measure the Euclidean distance in R3 between
the image points. Thus, χ is defined by
χ(z, w) = ||ϕ(z) − ϕ(w)||;
explicitly,
2|z − w| 2
χ(z, w) = p , χ(z, ∞) = p .
(1 + |z|2 )(1 + |w|2 ) (1 + |z|2 )
This interpretation of χ immediately shows that it is a distance function on C∞ .
Also, the metric space (C∞ , χ) is homeomorphic to S with the restriction of the
Euclidean metric; so (C∞ , χ) is compact and connected.
The spherical metric on C∞ is given by
2|dw|
σ(w)|dw| = ;
1 + |w|2
it has curvature
∆ log σ(w)
− = 1.
σ 2 (w)
The spherical distance on C∞ derived from this metric is


−1 z − w

dσ (z, w) = 2 tan ≤ π.
1 + w̄z
The chordal and spherical metrics are related to each other by the formula
 
1
χ(z, w) = 2 sin dσ (z, w) .
2
From 2θ/π ≤ sin θ ≤ θ when 0 ≤ θ ≤ π/2, we obtain
(2/π)dσ (z, w) ≤ χ(z, w) ≤ dσ (z, w),
so the two distances induce the same topology on C∞ . Note that
χ(z, w) dσ (z, w)
lim = σ(z) = lim .
w→z |z − w| w→z |z − w|

We present a complete description of the isometries of the spherical plane.


The orientation preserving conformal isometries of the spherical plane form a
group. All of the following groups are identical,
(1) the group of conformal isometries of the chordal distance;
(2) the group of conformal isometries of the spherical distance;
(3) the group of conformal isometries of the spherical metric;
(4) the group of Möbius maps of the form
az − c̄
z 7→ , |a|2 + |c|2 = 1.
cz + ā
The orientation-preserving isometries of R3 that fix the origin are the rotations
of R3 , and these are represented by the group SO(3) of 3 × 3 orthogonal matrices
with determinant one. The group SO(3) is conjugate to the group ϕ−1 SO(3)ϕ
Geometric properties of hyperbolic geodesics 179

which acts on C∞ ; the four identical groups above are equal to ϕ−1 SO(3)ϕ. For
this reason the isometries of the spherical plane are sometimes called rotations.
For antipodal z, w ∈ C∞ , that is, w = −1/z̄, any of the infinitely many great
circular arcs connecting z and w is a spherical geodesic. If z, w ∈ C∞ are not
antipodal, then the unique spherical geodesic arc is the shorter arc between z
and w of the unique great circle through z and w.
Just as one studies convex regions in the Euclidean plane it is natural to study
convex regions in the spherical plane. A simply connected region Ω on C∞ is
called spherically convex (relative to spherical geometry on C∞ ) if for each pair
of z, w ∈ Ω every spherical geodesic connecting z and w also lies in Ω. If Ω is
spherically convex and contains a pair of antipodal points, then Ω = C∞ . A
meromorphic and univalent function f defined on D is called spherically convex
if its image f (D) is a spherically convex subset of C∞ . A number of authors have
studied spherically convex functions; for example, [6], [8], [11], [15], [19], [21] and
[25].

8. Spherically convex univalent functions


In our discussion of Euclidean properties of hyperbolic geodesics, characteri-
zations of Euclidean convex functions played a crucial role. Therefore, it is not
surprising that characterizations of spherically convex functions play an impor-
tant role in investigating spherical properties of hyperbolic geodesics. One such
characterization obtained by Mejia and Minda [19] is
( )
zf ′′ (z) 2zf ′ (z)f (z)
(8.1) Re 1 + ′ − ≥0
f (z) 1 + |f (z)|2

for all z in D; also see [8]. Sometimes it is difficult to use (8.1) because it con-
tains the nonholomorphic term 2zf ′ (z)f (z)/(1 + |f (z)|2 ). One way to overcome
this difficulty is to establish two-variable characterizations for spherically convex
functions which are holomorphic in one of the two variables and are analogous
to Theorem 5.1
We now state two-variable characterizations for spherically convex functions
that will be applied to investigate properties of hyperbolic polar coordinates
on spherically convex regions and to derive other results for spherically convex
functions.
Theorem 8.1. Let f be meromorphic and locally univalent in D. Then f is
spherically convex if and only if
( )
2zf ′ (z) z+ζ 2zf ′ (z)f (ζ)
(8.2) Re − − >0
f (z) − f (ζ) z − ζ 1 + f (ζ)f (z)

for all z, ζ in D.
180 Ma and Minda IWQCMA05

Proof. Here we prove only the sufficiency. Observe that (8.2) is the spherical
analog of (5.4). Let
2zf ′ (z) z+ζ 2zf ′ (z)f (ζ)
(8.3) p(z, ζ) = − − .
f (z) − f (ζ) z − ζ 1 + f (ζ)f (z)
We show that if f satisfies the inequality (8.2), then (8.1) holds for all z ∈ D,
which characterizes spherically convex functions [19]. The assumption is that
Re {p(z, ζ)} > 0 for z, ζ ∈ D. Since
zf ′′ (z) 2zf ′ (z)f (z)
(8.4) p(z, z) = 1 + − ,
f ′ (z) 1 + |f (z)|2
f is spherically convex.
Corollary 8.2. Suppose f is meromorphic and locally univalent in D. Then f
is spherically convex if and only if
( )
zf ′ (z) − ζf ′ (ζ) zf ′ (z)f (ζ) + ζf ′ (ζ)f (z)
(8.5) Re − >0
f (z) − f (ζ) 1 + f (ζ)f (z)
for all z, ζ in D.

Proof. Note that (8.5) follows from (8.2) in the same manner that (5.3) was
derived from (5.4). Conversely, suppose (8.5) holds for all z, ζ in D. Set
zf ′ (z) − ζf ′ (ζ) zf ′ (z)f (ζ) + ζf ′ (ζ)f (z)
(8.6) q(z, ζ) = − .
f (z) − f (ζ) 1 + f (ζ)f (z)
Then Re {q(z, z)} > 0 for all z in D. As
zf ′′ (z) 2zf ′ (z)f (z)
q(z, z) = 1 + − ,
f ′ (z) 1 + |f (z)|2
the inequality (8.1) holds. Therefore, f is spherically convex.

As we pointed out earlier, we cannot easily derive properties of spherically


convex functions from (8.1) since it contains a nonholomorphic term. With
Theorem 8.1, this difficulty is overcome in some cases. If f is spherically convex,
then p(z, ζ) is holomorphic for in z ∈ D, has positive real part, and so satisfies

1 + |z|2 2|z|

(8.7) p(z, ζ) − 1 − |z|2 ≤ 1 − |z|2 .
The nonholomorphic function p(z, z) (see (8.4)) still satisfies the inequality (8.7),
which holds for the well known class consisting of holomorphic functions p(z) in
D with p(0) = 1 and Re {p(z)} > 0. Note that

1 + |z|2
p(z, z) − ≤ 2|z|
1 − |z| 1 − |z|2
2

also characterizes spherically convex functions and implies the inequality (8.1),
see [11].
Geometric properties of hyperbolic geodesics 181

This idea can be used to derive a number of results for spherically convex
functions. First, recall that a holomorphic and univalent function f in D with
f (0) = f ′ (0) − 1 = 0 is called starlike of order β ≥ 0 if Re {zf ′ (z)/f (z)} > β
in D. Using Theorem 8.1, we show that spherically convex functions are closely
related to starlike functions.
Theorem 8.3. If f (z) is spherically convex with f (0) = 0, then for every ζ ∈ D,
zζ f (z) − f (ζ)
Fζ (z) =  
f (ζ) (z − ζ) 1 + f (ζ)f (z)

is starlike of order 1/2.

Proof. Direct calculations yield


2zFζ (z)
− 1 = p(z, ζ).
Fζ (z)
Theorem 8.1 implies the result.

Since Re {F (z)/z} > 1/2 and F (z)2 /z is starlike if F is starlike of order 1/2
(see [26, p. 49]), we get the following results as corollaries of Theorem 8.3.
Corollary 8.4. If f (z) is spherically convex with f (0) = 0, then for every ζ ∈ D,
 
 ζ f (z) − f (ζ)  1
Re   >
 f (ζ) (z − ζ) 1 + f (ζ)f (z)  2

for all z in D.
Corollary 8.5. If f (z) is spherically convex with f (0) = 0, then for every ζ ∈ D,
Fζ2 (z) zζ 2 (f (z) − f (ζ))2
=  2
z f (ζ)2
(z − ζ) 1 + f (ζ)f (z)
2

is starlike in D.

Mejia and Pommerenke [21] obtained a number of results for spherically con-
vex functions by observing that f (z) is (Euclidean) convex when f (z) is spher-
ically convex and f (0) = 0. We now provide the sharp order of Euclidean
convexity for spherically convex functions that fix the origin.
Corollary 8.6. Let f (z) = αz + a2 z 2 + . . ., 0 < α < 1, be spherically convex.
Then for all z in D
   2
zf ′′ (z) α
Re 1 + ′ > √ .
f (z) 1 + 1 − α2
This result is best possible for each α.
182 Ma and Minda IWQCMA05

Example
 √8.7. For 0 < α ≤ 1, the spherical half-plane, or hemisphere, Ωα =
w : |w − 1 − α2 /α| < 1/α is spherically convex and
αz
kα (z) = √
1 − 1 − α2 z
maps D conformally onto Ωα . For the function kα ,
     2
zkα′′ (z) α
inf Re 1 + ′ :z∈D = √ .
kα (z) 1 + 1 − α2
Next, we give the sharp lower bound on Re {a2 f (z)} for normalized spherically
convex functions f (z) = αz+a2 z 2 +. . . . Similar results hold for Euclidean convex
functions [4] and hyperbolically convex functions [14].
Theorem 8.8. Let f (z) = αz + a2 z 2 + . . . , 0 < α ≤ 1, be spherically convex.
Then for all z in D

Re {a2 f (z)} ≥ 1 − α2 − 1 − α2 .
This result is best possible for all α.

It is easy to see that for the spherically


√ convex functions kα (z), the infimum
of Re {a2 kα (z)} over z ∈ D is 1 − α2 − 1 − α2 , so the lower bound is sharp for
all α ∈ (0, 1].

9. Spherically convex regions


Now, we establish certain properties of hyperbolic polar coordinates in spher-
ically convex regions. It is convenient to use the density of the hyperbolic metric
relative to the spherical metric; that is,
λΩ (w)|dw| 1
µΩ (w) = = (1 + |w|2 )λΩ (w).
σ(w)|dw| 2
Then λΩ (w)|dw| = µΩ (z)σ(w)|dw| and µΩ is invariant under all rotations of the
sphere.
Theorem 9.1. Suppose Ω is a hyperbolic region in C∞ .
(a) If Ω is spherically convex and w0 ∈ Ω, then dσ (w0 (s, θ), w0 ) is an increasing
function of s for all θ in R. Moreover, the sharp bounds
tanh(s/2) 1
p ≤ tan dσ (w(s, θ), w0 )
2
µΩ (w0 ) + tanh(s/2) µΩ (w0 ) − 1 2
tanh(s/2)
≤ p .
µΩ (w0 ) − tanh(s/2) µ2Ω (w0 ) − 1
hold.
(b) If dσ (w0 (s, θ), w0 ) is an increasing function of s for each w0 in Ω and all θ
in R, then Ω is spherically convex.

The proof of Theorem 9.1 is given in [17].


Geometric properties of hyperbolic geodesics 183

Example 9.2. Consider the function kα defined in Example 8.7. For w0 = 0,


µΩα (w0 ) = 1/α,
α tanh(s/2)
w0 (s, 0) = √
1 − tanh(s/2) 1 − α2

2
is the hyperbolic arc length parametrization of [0, 1+ α1−α ), and the upper bound
is equal to
α tanh(s/2) 1
√ = tan dσ (w0 (s, 0), 0).
1 − tanh(s/2) 1 − α2 2
This shows that the upper bound is sharp. Similarly,
−α tanh(s/2)
w0 (s, π) = √
1 + tanh(s/2) 1 − α2

2
is the hyperbolic arc length parametrization of ( −1+ α1−α , 0], and the lower bound
is equal to
α tanh(s/2) 1
√ = tan dσ (w0 (s, π), 0).
1 + tanh(s/2) 1 − α 2 2
Hence, the lower bound is also sharp.
Theorem 9.3. Suppose Ω is a hyperbolic region in C∞ .
(a) If Ω is spherically convex and w0 ∈ Ω, then dσ (w0 (s, θ1 ), w0 (s, θ2 )) is an in-
creasing function of s whenever eiθ2 6= eiθ1 .
(b) If there exists w0 ∈ Ω such that dσ (w0 (s, θ1 ), w0 (s, θ2 )) is an increasing func-
tion of s whenever eiθ2 6= eiθ1 , then Ω is spherically convex.

The reader can consult [17] for a proof of Thorem 9.3.


Geometrically, Theorem 9.3(a) indicates that in a spherically convex region
Ω, two hyperbolic geodesics starting off in different directions from a point w0 in
Ω will spread farther apart relative to the spherical distance.
If Ω is spherically convex, then so is every hyperbolic disk and conversely.
This follows from the analog of Study’s theorem for spherically convex functions;
see [19].

10. Hyperbolic geometry


In this section, we indicate similar monotonicity properties for hyperbolic
polar coordinates in hyperbolically convex regions. Because of the numerous
similarities with the Euclidean and spherical cases, we present even fewer details
in this situation. It is convenient to introduce the notation
λΩ (w)|dw| 1
νΩ (w) = = (1 − |w|2 )λΩ (w)
λD (w)|dw| 2
for the density of the hyperbolic metric of a region Ω ⊂ D relative to the back-
ground hyperbolic metric λD (w)|dw|.
A simply connected region Ω in D is called hyperbolically convex (relative to
the background hyperbolic geometry on D) if for all points z, w ∈ Ω the arc of
184 Ma and Minda IWQCMA05

the hyperbolic geodesic in D connecting z and w also lies in Ω. A holomorphic


and univalent function f defined on D with f (D) ⊂ D is called hyperbolically
convex if its image f (D) is a hyperbolically convex subset of D. Hyperbolically
convex functions have been studied by a number of authors [7], [8], [12], [13],
[14], [17], [20], [22]. The related concept of hyperbolically 1-convex functions
was investigated in [9].
There are known characterizations of hyperbolically convex functions. For
example, a holomorphic and locally univalent function f with f (D) ⊂ D is hy-
perbolically convex if and only if [12]
( )
zf ′′ (z) 2zf ′ (z)f (z)
(10.1) Re 1 + ′ + ≥0
f (z) 1 − |f (z)|2
for all z in D. Mejia and Pommerenke [22] (also see [14]) showed that a holomor-
phic and locally univalent function f with f (D) ⊂ D is hyperbolically convex if
and only if
( )
2zf ′ (z) z+ζ 2zf ′ (z)f (ζ)
(10.2) Re − + >0
f (z) − f (ζ) z − ζ 1 − f (ζ)f (z)
for all z, ζ in D. This is the hyperbolic analog of (8.2). Similar to the proof of
Corollary 8.2, we obtain the following characterization from (10.2).
Theorem 10.1. A holomorphic and locally univalent function f with f (D) ⊂ D
is hyperbolically convex if and only if
( )
zf ′ (z) − ζf ′ (ζ) zf ′ (z)f (ζ) + ζf ′ (ζ)f (z)
(10.3) Re + >0
f (z) − f (ζ) 1 − f (ζ)f (z)
for all z, ζ in D.

These two-point characterizations can be used to derive monotonicity proper-


ties of hyperbolic polar coordinates on hyperbolically convex regions in D.
Theorem 10.2. Let Ω ⊂ D.
(a) If Ω is hyperbolically convex and w0 ∈ Ω, then dD (w0 (s, θ), w0 ) is an increasing
function of s for all θ in R. Moreover, we have the following sharp bounds:
2 tanh(s/2)
q
νΩ (w0 )(1 + tanh(s/2)) + νΩ2 (w0 )(1 + tanh(s/2))2 − 4 tanh(s/2)
1
≤ tanh dD (w0 (s, θ), w0 )
2
2 tanh(s/2)
≤ q .
2 2
νΩ (w0 )(1 − tanh(s/2)) + νΩ (w0 )(1 − tanh(s/2)) + 4 tanh(s/2)
(b) If dD (w0 (s, θ), w0 ) is an increasing function of s for each w0 in Ω and all θ
in R, then Ω is hyperbolically convex.
Geometric properties of hyperbolic geodesics 185

Example 10.3. For 0 < α ≤ 1, the hyperbolic half-plane


 √ 
1 1 − α2

Hα = D \ w : w + ≤
α α
is hyperbolically convex and
2αz
Kα (z) = p
1−z+ (1 − z)2 + 4α2 z
maps D conformally onto Hα . When w0 = 0, νHα (0) = 1/α, w0 (s, 0) = Kα (tanh(s/2))
is the hyperbolic arc length parametrization of [0, 1), and the upper bound is
equal to
2α tanh(s/2) 1
p = tanh dD (w0 (s, 0), 0).
1 − tanh(s/2) + (1 − tanh(s/2)) + 4α tanh(s/2)
2 2 2
This shows that the upper bound is sharp. Similarly, √
w0 (s, π) = kα (− tanh(s/2))
−1− 1−α2
is the hyperbolic arc length parametrization of ( α
, 0], and the lower bound
is equal to
2α tanh(s/2) 1
p = tanh dD (w0 (s, π), 0).
1 + tanh(s/2) + (1 + tanh(s/2))2 − 4α2 tanh(s/2) 2
Thus, the lower bound is also sharp.

The proof of Theorem 10.4 below is analogous to the proof of Theorem 9.3;
the characterization (10.3) for hyperbolically convex functions is used in place of
(8.2).
Theorem 10.4. Suppose Ω ⊂ D.
(a) If Ω is hyperbolically convex and w0 ∈ Ω, then dD (w0 (s, θ1 ), w0 (s, θ2 )) is an
increasing function of s for all eiθ2 6= eiθ1 .
(b) If there exists w0 ∈ Ω such that dD (w0 (s, θ1 ), w0 (s, θ2 )) is an increasing func-
tion of s whenever eiθ2 6= eiθ1 , then Ω is hyperbolically convex.

Ω ⊂ D is hyperbolically convex if and only if every hyperbolic disk DΩ (w0 , r)


is hyperbolically convex as a subset of D. This is a direct consequence of the
analog of Study’s Theorem for hyperbolically convex functions; see [20] and [12].

11. Concluding remarks


Relative to the background geometry (hyperbolic, Euclidean, or spherical)
hyperbolic geodesics and hyperbolic disks have similar behavior in convex re-
gions. Moreover, there are numerous similarities between conformal maps of the
unit disk onto convex regions in each of the three geometries, although formu-
las for spherical or hyperbolic convexity can be more complicated than those
for Euclidean convexity because the spherical plane and the hyperbolic plane
have nonzero curvature. It is possible to obtain results for Euclidean convex-
ity as a limit of corresponding results for spherical convexity; for example, see
186 Ma and Minda IWQCMA05

Kim-Minda [6] for an illustration of the method. In the same manner Euclidean
results can be obtained as the limit of hyperbolic convexity results.
Holomorphic functions defined on the unit disk can be viewed as maps from D
to the Euclidean plane. Bounded holomorphic functions on the unit disk can be
regarded as maps from D to the hyperbolic plane, provided they are scaled to be
bounded by one. Finally, meromorphic functions on D can be considered as maps
into the spherical plane. Sometimes connections between classical results can be
made by adopting this geometric view. This paper showed the close connection
between Euclidean convexity, spherical convexity and hyperbolic convexity. Also,
by adopting this geometric viewpoint it is possible to recognize there should be
analogs of classical results for holomorphic functions for maps into the other two
geometries.
Some other function theory papers that relate to comparisons between hyper-
bolic, Euclidean and spherical geometry are [5], [23], [24], [25].

References
[1] A. F. Beardon and D. Minda, The hyperbolic metric and geometric function theory, Pro-
ceedings of the International Workshop on Quasiconformal Mappings and their Applica-
tions, Narosa Publishing House, India, (2006), 159-206.
[2] P. Duren, Univalent functions, Grundlehern Math. Wiss. 259, Springer, New York 1983.
[3] M. Finkelstein, Growth estimates for convex functions, Proc. Amer. Math. Soc. 18 (1967),
412-418.
[4] R. Fournier, J. Ma and S. Ruscheweyh, Convex univalent functions and omitted values,
Approximation Theory, 21 (1998), 225-241.
[5] S. Kim and D. Minda, A geometric approach to two-point comparisons for hyperbolic and
euclidean geometry on convex regions, J. Korean Math. Soc., 36 (1999), 1169-1180.
[6] S. Kim and D. Minda, The hyperbolic metric and spherically-convex regions, J. Math.
Kyoto Univ., 41 (2001), 285-302.
[7] S. Kim and T. Sugawa, Characterizations of hyperbolically convex regions, J. Math. Anal.
Appl. 309 (2005), 37-51.
[8] W. Ma, D. Mejia and D. Minda, Distortion theorems for hyperbolically and spherically
k-convex functions, Proc. of an International Conference on New Trends in Geometric
Function Theory and Application, R. Parvathan and S. Ponnusamy (editors), World Sci-
entific, Singapore, 1991, 46-54.
[9] W. Ma, D. Mejia and D. Minda, Hyperbolically 1-convex functions, Ann. Polon. Math.,
84 (2004), 185-202.
[10] W. Ma and D. Minda, Euclidean linear invariance and uniform local convexity, J. Austral.
Math. Soc. 52 (1992), 401-418.
[11] W. Ma and D. Minda, Spherical linear invariance and uniform local spherical convexity,
Current Topics in Geometric Function Theory, H. M. Srivastava and S. Owa (editors),
World Scientific, Singapore, 1993, 148-170.
[12] W. Ma and D. Minda, Hyperbolically convex functions, Ann. Polonici Math. 60 (1994),
81-100.
[13] W. Ma and D. Minda, Hyperbolic linear invariance and hyperbolic k-convexity, J. Aus-
tralian Math. Soc., 58 (1995), 73-93.
[14] W. Ma and D. Minda, Hyperbolically convex functions II, Ann. Polonici Math. 71 (1999),
273-285.
[15] W. Ma and D. Minda, Two-point distortion theorems for spherically-convex functions,
Rocky Mtn. J. Math., 30 (2000), 663-687.
Geometric properties of hyperbolic geodesics 187

[16] W. Ma and D. Minda, Euclidean properties of hyperbolic polar coordinates, Comput.


Methods Funct. Theory, to appear.
[17] W. Ma and D. Minda, Geometric properties of hyperbolic polar coordinates, submitted.
[18] D. Mejia and D. Minda, Hyperbolic geometry in k-convex regions, Pacific J. Math., 141
(1990), 333-354.
[19] D. Mejia and D. Minda, Hyperbolic geometry in spherically k-convex regions, Compu-
tational Methods and Function Theory, Proceedings, Valparaiso, Chili, S. Ruscheweyh,
E. B. Saff, L. C. Salinas and R. S. Varaga (editors), Lecture Notes in Mathematics, Vol.
1435, Springer-Verlag, New York, 1990, 117-129.
[20] D. Mejia and D. Minda, Hyperbolic geometry in hyperbolically k-convex regions, Rev.
Columbiana Math., 25 (1991), 123-142.
[21] D. Mejia and Ch. Pommerenke, On spherically convex functions, Michigan Math. J. 47
(2000), 163-172.
[22] D. Mejia and Ch. Pommerenke, On hyperbolically convex functions, J. Geom. Anal. 10
(2000), 365-378.
[23] D. Minda, Bloch constants, J. Analyse Math. 4 (1982), 54-84.
[24] D. Minda, Estimates for the hyperbolic metric, Kodai Math. J., 8 (1985), 249-258.
[25] D. Minda, Applications of hyperbolic convexity to euclidean and spherical convexity, J.
Analyse Math., 49 (1987), 90-105.
[26] R. Ruscheweyh, Convolutions in Geometric Function Theory, Les Presses de l’Universite
de Montréal, Montréal, 1982.
[27] T. Sheil-Small, On convex univalent functions, J. London Math. Soc. 1 (1969), 483-492.
[28] D. J. Struik, Lectures on Classical Differential Geometry, Addison-Wesley, Cambridge
1950.
[29] E. Study, Konforme Abbildung einfachzusammenhängerder Bereiche, Vorlesungen über
ausgewählte Gegenstande, Heft 2, Leipzig und Berlin 1913.
[30] T. J. Suffridge, Some remarks on convex maps of the unit disk, Duke Math. J. 37 (1970),
775-777.

W. Ma E-mail: wma@pct.edu
Address: School of Integrated Studies, Pennsylvania College of Technology, Williamsport, PA
17701, USA

D. Minda E-mail: David.Minda@math.uc.edu


Address: Department of Mathematical Sciences, University of Cincinnati, Cincinnati, Ohio
45221-0025, USA
Proceedings of the International Workshop on Quasiconformal
Mappings and their Applications (IWQCMA05)

Quasiminimizers and Potential Theory

Olli Martio

Abstract. Quasiminimizers are almost minimizers of variational integrals.


Although quasiminimizers do not form a sheaf and do not provide a unique
solution to the Dirichlet problem it is shown that they form an interesting basis
for a potential theory. Quasisuperminimizers and their Poisson modifications
are considered as well as their convergence properties. Special attention is
devoted to the theory on the real line.

Keywords. quasiminimizers, quasisuperminimizers, quasisuperharmonic func-


tions.
2000 MSC. Primary: 31C45; Secondary: 35J20, 35J60.

Contents

1. Introduction 189
2. Case n = 1 191
3. Properties of quasiminimizers 193
4. Quasisuperminimizers, Poisson modifications and regularity 194
5. More about n = 1 198
6. Quasisuperharmonic functions 199
7. Appendix 1 200
8. Appendix 2 202
References 205

1. Introduction
Quasiminimizers minimize a variational integral only up to a multiplicative
constant. More precisely, let Ω ⊂ Rn be an open set, K ≥ 1 and 1 ≤ p < ∞. In
the case of the p-Dirichlet integral, a function u belonging to the Sobolev space
1,p
Wloc (Ω) is a (p, K)-quasiminimizer or a K-quasiminimizer, if
Z Z
(1.1) p
|∇u| dx ≤ K |∇v|p dx
Ω′ Ω′

Version October 19, 2006.


190 O. Martio IWQCMA05

for all functions v ∈ W 1,p (Ω′ ) with v − u ∈ W01,p (Ω′ ) and for all open sets Ω′
with a compact closure in Ω. A 1-quasiminimizer, called a minimizer, is a weak
solution of the corresponding Euler equation
(1.2) div(|∇u|p−2 ∇u) = 0.
Clearly being a weak solution of (1.2) is a local property. However, being a K-
quasiminimizer is not a local property as one-dimensional examples easily show.
This indicates that the theory for quasiminimizers differs from the theory for
minimizers and that there are some unexpected difficulties.
Quasiminimizers have been previously used as tools in studying the regular-
ity of minimizers of variational integrals, see [GG1–2]. The advantage of this
approach is that it covers a wide range of applications and that it is based only
on the minimization of the variational integrals instead of the corresponding Eu-
ler equation. Hence regularity properties as Hölder continuity and Lp -estimates
are consequences of the quasiminimizing property. It is an important fact that
nonnegative quasiminimizers satisfy the Harnack inequality, see [DT].
Instead of using quasiminimizers as tools, the objective of these lectures is to
show that quasiminimizers have a fascinating theory themselves. In particular,
they form a basis for nonlinear potential theoretic model with interesting fea-
tures. From the potential theoretic point of view quasiminimizers have several
drawbacks: They do not provide unique solutions of the Dirichlet problem, they
do not obey the comparison principle, they do not form a sheaf and they do
not have a linear structure even when the corresponding Euler equation is lin-
ear. However, quasiminimizers form a wide and flexible class of functions in the
calculus of variations under very general circumstances. Observe that the quasi-
minimizing condition (1.1) applies not only to one particular variational integral
but the whole class of variational integrals at the same time. For example, if a
variational kernel F (x, ∇u) satisfies
(1.3) α|h|p ≤ F (x, h) ≤ β|h|p
for some 0 < α ≤ β < ∞, then the minimizers of
Z
F (x, ∇u) dx

are quasiminimizers of the p-Dirichlet integral


Z
(1.4) |∇u|p dx.

Hence the potential theory for quasiminimizers includes all minimizers of all
variational integrals similar to (1.4). The essential feature of the theory is the
control provided by the bounds in (1.3).
For example, the coordinate functions of a quasiconformal or, more generally,
quasiregular mapping are quasiminimizers of the n-Dirichlet integral
Z
|∇u|n dx
Quasiminimizers and Potential Theory 191

in all dimensions n = 2, 3, . . . .
Recently quasiminimizers have been considered in metric measure spaces. This
means that a metric space (X, d) is equipped with a Borel measure µ which
satisfies some standard assumptions like the doubling property. The Sobolev
space W 1,p is replaced by the so called Newtonian space N 1,p which for Rn and
the Lebesgue measure reduces to W 1,p . We do not consider metric spaces here
although most of the results hold in this case under appropriate conditions. For
this theory see [KM2].

2. Case n = 1
For n = 1 the definition (1.1) can be written in the following form: Let (a, b)
1,p
be an open interval in R and u ∈ Wloc (a, b). Then u is a (p, K)-quasiminimizer,
or K-quasiminimizer for short, if for all closed intervals [c, d] ⊂ (a, b)
Zd Zd
′ p
(2.1) |u | dx ≤ K |v ′ |p dx
c c

whenever u − v ∈ W01,p (c, d).


Now affine functions are minimizers, i.e. 1-quasiminimizers, for every p ≥ 1.
This fact can be easily deduced from the one dimensional version of (1.2) if p > 1.
For p = 1 this is trivial. Moreover, affine functions are the only minimizers for
p > 1. Thus choosing v(x) = α(x − c) + β where
α = (u(d) − u(c))/(d − c), β = u(c)
we see that u − v ∈ W01,p (c, d) and (2.1) yields
Zd
|u(d) − u(c)|p
(2.2) |u′ |p dx ≤ K ,
|d − c|p−1
c

see [GG2]. The inequality (2.2) gives another definition for a K-quasiminimizer
u: the function u is a locally absolutely continuous function in (a, b) that satisfies
(2.2) on each subinterval [c, d] of (a, b).
Observe that u ∈ W 1,p (c, d) in a bounded open interval (c, d) means that u is
absolutely continuous on [c, d] with
Zd
(2.3) |u′ |p dx < ∞.
c

If u ∈ W 1,p (c, d) and u − v ∈ W01,p (c, d), then v ∈ W 1,p (c, d) and v(c) =
1,p
u(c), v(d) = u(d). Functions u ∈ Wloc (a, b) are simply locally absolutely contin-
uous functions on (a, b) such that (2.3) holds in each subinterval [c, d] ⊂ (a, b).
We leave the following lemma as an exercise.
192 O. Martio IWQCMA05

Lemma 2.4. Suppose that u is a (p, K)-quasiminimizer in (a, b). Then u is a


monotone function. If p > 1 then u is either strictly monotone or constant.

The following lemma is more difficult to prove. It does not hold for p = 1.
Lemma 2.5. Let u be a (p, K)-quasiminimizer, p > 1, in an interval (a, b). If
b < ∞, then u has a continuous extension to b and (2.2) holds in all intervals
[c, d] ⊂ (a, b].

Proof. We may assume that u is increasing, b = 1, [0, 1] ⊂ (a, b] and u(0) = 0.


Fix 0 ≤ c < t < 1. Now
Zt Zt ! p1 Zt ! p1
u′ dx ≤ (t − c)(p−1)/p u′p dx ≤ (t − c)(p−1)/p u′p dx
c c 0
p−1 Zt   p−1 Zt
1 (t − c) p
′ 1 c p
≤ K p
p−1 u dx = K p 1− u′ dx
t p t
0 0

where we have used the Hölder inequality and (2.2). Next we choose c = 1 −
1
(2p K) 1−p . Then 0 < c < 1 and letting t ∈ (c, 1) we obtain
  p−1
c p p−1 1
1− < (1 − c) p = 1 .
t 2K p
The above inequalities yield
Zt Zc Zt Zc Zt
1
u′ dx = u′ dx + u′ dx ≤ u′ dx + u′ dx
2
0 0 c 0 0
and hence
Zt Zc
u(t) = u(t) − u(0) = u′ dx ≤ 2 u′ dx = 2u(c).
0 0
Since u is increasing, letting t → 1 we obtain
u(b) = u(1) = lim u(t) ≤ 2u(c) < ∞
t→1

and the last assertion of the lemma now follows from (2.2).
Lemma 2.5 shows that the natural domain of definition for a 1-dimensional
quasiminimizer is the closed interval [a, b].
Example 2.6. The function u(x) = xα , α > 1/2, is a (2, K)-quasiminimizer
in [0, ∞) for K = α2 /(2α − 1). This is a rather easy computation. Note that
u(x) = x1/2 is not a (2, K)-quasiminimizer in [0, ∞) (and in (0, ∞)) since u′ does
not belong to L2 (0, 1).

We will consider one dimensional quasiminimizers again in Chapter 5. The


one dimensional case was first studied in [GG2].
Quasiminimizers and Potential Theory 193

3. Properties of quasiminimizers
We start with a basic regularity property.
Theorem 3.1. Suppose that u is a (p, K)-quasiminimizer in Ω ⊂ Rn , p > 1. If
0 < r < R are such that the ball B(x, 2R) ⊂ Ω, then
osc(u, B(x, r)) ≤ C(r/R)α osc(u, B(x, R))
where C < ∞ and α ∈ (0, 1] depend on p, n and K only.

In particular Theorem 3.1 implies that u is locally Hölder continuous in Ω.


Another important property is the Harnack inequality.
Theorem 3.2. Let u be as in Theorem 3.1 and u ≥ 0. Then in each ball B(x, R)
such that B(x, 2R) ⊂ Ω
sup u ≤ C inf u
B(x,R) B(x,R)

where the constant C depends on p, n and K only.


1,p
Since quasiminimizers are functions in Wloc (Ω) only, Theorem 3.1 also means
that they can be made continuous after redefinition on a set of measure zero.
For p > n, and hence for all p > 1 for n = 1, Theorem 3.1 follows from
the Sobolev embedding lemma. Note that for p = 1 = n quasiminimizers are
continuous but they need not be Hölder continuous.
We do not prove Theorems 3.1 and 3.2 here. The proof for Theorem 3.2 in
the case n = 1 is relatively simple, see [GG2]. For the proof of Theorem 3.1 the
De Giorgi method can be used. The basic tool is the Sobolev type inequality
Z !1/t Z !1/p
− |u|t dx ≤ cr − |∇u|p dx
B(x,r) B(x,r)

for functions u ∈ W01,p (B(x, r)) where t > p. The main difficulty is to prove that
u is locally essentially bounded. For the proof see [GG1], [GG2] and [KS]. In the
paper [KS] metric measure spaces are considered and hence the regularity proof
of [KS] uses minimal assumptions.
In the general case n ≥ 2 the proof for Theorem 3.2 is rather complicated, see
[DT] and [KS]. The proof makes use of the Krylov–Safonov covering argument
[KSa]. Very recently it has turned out that the Moser method can be employed to
prove Theorems 3.1 and 3.2 for quasiminimizers even in metric measure spaces,
see [Ma].
In Potential Theory the Harnack inequality and Harnack’s principle are essen-
tially equivalent. From Theorem 3.1 and 3.2 it easily follows (p > 1): Suppose
that (ui ) is an increasing sequence of K-quasiminimizers in a domain Ω. If
lim ui (x0 ) < ∞ at some point x0 ∈ Ω, then lim ui is a K-quasiminimizer.
We will return to the proof of this fact in the next chapter and in Appendix 2.
194 O. Martio IWQCMA05

4. Quasisuperminimizers, Poisson modifications and


regularity
1,p
Let Ω be an open set in Rn . A function u ∈ Wloc (Ω) is called a (p, K)-
quasisuperminimizer, or a K-quasisuperminimizer, if
Z Z
(4.1) |∇u| dx ≤ K |∇v|p dx
p

Ω′ Ω′
1,p
holds for all open Ω′ ⊂⊂ Ω and all v ∈ Wloc (Ω) such that v ≥ u a.e. in Ω′ and
v − u ∈ W01,p (Ω′ ).
Remarks 4.2. (a) A 1-quasisuperminimizer is called a superminimizer.
(b) A superminimizer is a supersolution of the p-harmonic equation
∇ · (|∇u|p−2 ∇u) = 0,
i.e. u satisfies Z
|∇u|p−2 ∇u · ∇ϕ dx ≥ 0

for all non–negative ϕ ∈ C0∞ (Ω), see [HKM] for this theory. Observe that for
p = 2 every superharmonic (in the classical sense) function u is a superminimizer
1,2
provided that u belongs to Wloc (Ω), however, a superharmonic function need not
1,2
belong to Wloc (Ω). For n = 2 the classical example is u(x) = − log |x| which is
superharmonic in R2 but does not belong to W 1,2 (B(0, 1)). We return to this
problem in Chapter 6.
(c) The inequality (4.1) can be replaced by several other inequalities, for
example Z Z
p
|∇u| dx ≤ K |∇v|p dx,
Ω′ \E Ω′ \E

where E ⊂ Ω \ {u 6= v} is any measurable set. For the list of these conditions
see [B] and [KM2].

A function u is called a K-quasisubminimizer if −u is a K-quasisuperminimizer.


Note that if u is a K-quasisuperminimimizer, then αu and u + β are K-
quasisuperminimizers when α ≥ 0 and β ∈ R. However, the sum of two K-
quasisuperminimizers need not be a K-quasisuperminimizer even in the case
p = 2.
Lemma 4.3. Suppose that ui , i = 1, 2, are Ki -quasisuperminimizers in Ω. Then
min(u1 , u2 ) is a K-quasisuperminimizer in Ω with K = min(K1 K2 , K1 + K2 ).

Proof. We prove that u = min(u1 , u2 ) is a K-quasisuperminimizer with K ≤


K1 K2 ; this inequality is important in applications. The proof for K ≤ K1 + K2
is similar, see [KM2]. To this end let Ω′ ⊂⊂ Ω be an open set and v − u ∈
W01,p (Ω′ ), v = u in Ω \ Ω′ and v ≥ u. Now
Quasiminimizers and Potential Theory 195

Z Z Z
p p
|∇u| dx = |∇u1 | dx + |∇u2 |p dx
Ω′ {u1 ≤u2 }∩Ω′ {u1 >u2 }∩Ω′

and write w = max(min(u2 , v), u1 ). Then w ≥ u1 a.e. in Ω′ , w − u1 ∈ W01,p (Ω′ )


and w = u1 if u1 > u2 . Thus the quasisuperharmonicity of u1 , see Remark 4.2
(c), yields
Z Z
p
|∇u1 | dx ≤ K1 |∇w|p dx
{u1 ≤u2 }∩Ω′ {u1 ≤u2 }∩Ω′
Z Z
p
= K1 |∇v| dx + K1 |∇u2 |p dx.
{u1 ≤u2 }∩{v<u2 }∩Ω′ {u1 ≤u2 }∩{v≥u2 }∩Ω′

¿From these inequalities we obtain


Z Z
p
|∇u| dx ≤ K1 |∇v|p dx
Ω′ {u1 ≤u2 }∩{v<u2 }∩Ω′
Z Z
p
+ K1 |∇u2 | dx + |∇u2 |p dx
{u1 ≤u2 }∩{v≥u2 }∩Ω′ {u1 >u2 }∩Ω′
Z Z
p
≤ K1 |∇v| dx + K1 |∇u2 |p dx.
{u1 ≤u2 }∩{v<u2 }∩Ω′ {v≥u2 }∩Ω′

Note that Ω′ ∩ {u1 > u2 } ⊂ Ω′ ∩ {v ≥ u2 }. On the other hand max(u2 , v) − u2 ∈


W01,p (Ω′ ), max(u2 , v) ≥ u2 and max(u2 , v) − u2 = 0 in {v ≤ u2 } and hence the
quasisuperharmonicity of u2 implies
Z Z
p
|∇u2 | dx ≤ K2 |∇v|p dx.
{v≥u2 }∩Ω′ {v≥u2 }∩Ω′

This together with the previous inequality completes the proof.


The following corollary is important.
Corollary 4.4. Suppose that u is a K-quasisuperminimizer and h is a super-
minimizer in Ω. Then min(u, h) is a K-quasisuperminimizer in Ω.
Remark 4.5. Lemma 4.3 and Corollary 4.4 are the counterparts of a prop-
erty of classical superharmonic functions: If u1 and u2 are superharmonic, then
min(u1 , u2 ) is superharmonic.

Corollary 4.4 implies the necessity part of the following result. The other half
follows from Theorem 4.14 below.
1,p
Lemma 4.6. Suppose that u ∈ Wloc (Ω). Then u is a K-quasisuperminimizer,
p > 1, if and only if min(u, c) is a K-quasisuperminimizer for each c ∈ R.
196 O. Martio IWQCMA05

The Poisson modification is an important tool in Potential Theory. In the


classical case p = 2 this means the following: Let u be superharmonic in Ω
1,2
and u ∈ Wloc (Ω) (this assumption is not actually needed). If Ω′ ⊂⊂ Ω is an
open set let h be a minimizer (harmonic) in Ω′ with boundary values u, i.e.
u − h ∈ W01,2 (Ω′ ). The Poisson modification of u in Ω′ is defined as

′ h in Ω′ ,
(4.7) P (u, Ω ) =
u in Ω \ Ω′ .
Then P (u, Ω′ ) is a superharmonic function in Ω, P (u, Ω′ ) ≤ u and P (u, Ω′ ) is
harmonic in Ω′ . This theory works well for superminimizers for all p > 1, see
[HKM].
For quasisuperminimizers the above method does not work as above. In partic-
ular there exists a K-quasisuperminimizer (even a K-quasiminimizer) u such that
the function P (u, Ω′ ) in (4.7) is not a K ′ -quasisuperminimizer for any K ′ < ∞.
The example is one dimensional and requires some computation. However, there
are two replacements for P (u, Ω′ ).
The first Poisson modification is a modification of (4.7). Let u be a K-
quasisuperminimizer in Ω, p > 1, and Ω′ ⊂⊂ Ω an open set. Let h be the
minimizer with boundary values u in Ω′ , i.e. u − h ∈ W01,p (Ω′ ). Observe that
such a unique function h exists - this is a basic result in the theory of Sobolev
spaces, see e.g. [HKM]. Let

′ min(u, h) in Ω′ ,
(4.8) P1 (u, Ω ) =
u in Ω \ Ω′ .

Theorem 4.9. [KM2] The function P1 (u, Ω′ ) is K-quasisuperminimizer in Ω.

By the construction of P1 (u, Ω′ ), P1 (u, Ω′ ) ≤ u in Ω. Note also that if u is a


K-quasiminimizer, then P1 (u, Ω′ ) is also a K-quasiminimizer in Ω′ by Corollary
4.4.
Next we consider another possibility for a Poisson modification. For this we
need to consider obstacle problems. The obstacle method is the most important
method in the nonlinear potential theory. Let Ω ⊂ Rn be an open set and
u ∈ W 1,p (Ω). Write
Ku+ (Ω) = {v ∈ W 1,p (Ω) : v − u ∈ W01,p (Ω), v ≥ u a.e. in Ω}.
The class Ku− is defined similarly but v ≥ u is replaced by v ≤ u. The following
result is well-known.
Lemma 4.10. [HKM] The obstacle problem
Z
inf

|∇v|p dx, p > 1,
v∈Ku (Ω)


has a unique solution u ∈ Ku− (Ω). Moreover, u− is a subminimizer and contin-
uous if u is continuous.
Quasiminimizers and Potential Theory 197

A similar result holds for the class Ku+ (Ω) and the solution is a superminimizer.
Suppose now that u is a K-quasisuperminimizer in Ω and Ω′ ⊂⊂ Ω is open.
Let u− be the solution to the Ku− (Ω′ )-obstacle problem. Define
 −
′ u in Ω′ ,
(4.11) P2 (u, Ω ) =
u in Ω \ Ω′ .

Theorem 4.12. The function P2 (u, Ω′ ) is a K-quasisuperminimizer in Ω, a


subminimizer in Ω′ and a K-quasiminimizer in Ω′ .

The proof for Theorem 4.12 is in Appendix 1.


Superharmonic functions in the classical potential theory are lower semicon-
tinuous. It turns out that quasisuperminimizers can be defined pointwise and
the resulting function is lower semicontinuous.

Theorem 4.13. [KM2] Suppose that u is a K-quasisuperminimizer in Ω, p > 1.


Then the function u∗ : Ω → (−∞, ∞] defined by

u∗ (x) = lim ess inf u


r→0 B(x,r)

is lower semicontinuous and u∗ = u a.e. (u and u∗ are the same function in


1,n
Wloc (Ω)).

The proof for Theorem 4.13 is based on the De Giorgi method that is used to
prove a weak Harnack inequality
Z !1/σ
− uσ dx ≤ c ess inf u
B(x,3r)
B(x,r)

for a K-quasisuperharmonic function u ≥ 0. Here B(x, 5r) ⊂ Ω and c and σ > 0


depend only on n, p > 1 and K.
The following is Harnack’s principle for quasisuperminimizers.

Theorem 4.14. Suppose that (ui ) is an increasing sequence of K-quasisupermini-


mizers in Ω, p > 1. If u = lim ui is such that either

(i) u is locally bounded above or


1,p
(ii) u ∈ Wloc (Ω),

then u is a K-quasisuperminimizer.

The proof for Theorem 4.14 is somewhat tedious. It is presented in Appen-


dix 2.
198 O. Martio IWQCMA05

5. More about n = 1
We take a closer look at the case n = 1. Recall that a superminimizer is a
1-quasisuperminimizer. The next result is easy to prove, see [MS].
Lemma 5.1. Suppose that u : [a, b] → R is a superminimizer, p > 1. Then u is
a concave function.

¿From Lemma 5.1 it follows that a superminimizer is a Lipschitz function in


each interval [c, d] ⊂⊂ (a, b). It need not be a Lipschitz function in [a, b].
How regular are K-quasiminimizers and K-quasisuperminimizers? The an-
swer is not known for n ≥ 2 but for n = 1 some exact answers have been
obtained.
Let 1 < p < ∞ and let ω : [a, b] → [0, ∞) be a weight in L1 (a, b). Set
Z ! p1 Z !−1
Gp (ω) = sup − ω p dx − ω dx
I
I I

where the supremum is taken over all intervals I ⊂ [a, b]. If Gp (ω) < ∞, then ω
is said to belong to the Gp -class of Gehring.
For a non-constant quasiminimizer u in [a, b] set
Zd
(d − c)p−1
Kp (u) = sup |u′ |p dx
[c,d] |u(d) − u(c)|p
c

where the supremum is taken over all intervals [c, d] ⊂ [a, b]. In other words,
Kp (u) is the least constant in (2.2). The following lemma is immediate.
Lemma 5.2. Let u : [a, b] → R be absolutely continuous and non-constant with
u′ ≥ 0 a.e. Then u is Kp (u)-quasiminimizer with exponent p, p > 1, if and only
1
if u′ belongs to the Gp -class with Gp (u′ ) = Kp (u) p .

A. Korenovskii [K] has determined the optimal higher integrability bound p0 =


p0 (p, K) for a weight ω in the Gp -class. Hence Lemma 5.2 enables us to determi-
nate the optimal integrability bound for the derivative of a Kp -quasiminimizer.
Let γp,t : [p, ∞) → R, p > 1, t > 1, be the function
x−p x p
γp,t (x) = 1 − tp ( ),
x x−1
and let p1 (p, t) ∈ (p, ∞) be the unique solution of the equation γp,t (x) = 0. For
the properties of p1 (p, t) see [DS, Section 3].
Theorem 5.3. Suppose that u is a (p, K)-quasiminimizer, p ≥ 1, K ≥ 1, in
1
[a, b]. Then u′ ∈ Ls (a, b) for 1 ≤ s < p1 (p, K p ) if p > 1 and K > 1, u′ ∈ L1 (a, b)
if p = 1 and u′ ∈ L∞ (a, b) if p > 1 and K = 1. All these integrability conditions
are sharp.
Quasiminimizers and Potential Theory 199

Proof. Let first p > 1 and K > 1. We may assume that u is increasing. By
1
Lemma 5.2, Gp (u′ ) = Kp (u) p and from [K, Theorem 2] we conclude that u′ ∈
1
Ls (a, b) for 1 ≤ s < p1 (p, Gp (u′ )) = p1 (p, Kp (u) p ).
If p > 1 and K = 1, then u is affine and hence u′ ∈ L∞ (a, b). For p = 1, u′
trivially belongs to L1 (a, b).
1
To see that the bound α = p1 (p, K p ) is sharp for p > 1 and K > 1 it suffices
to consider the interval [0, 1]. The function
α α−1
u(x) = x α , x ∈ [0, 1],
α−1
1
has the derivative u′ (x) = x− α and a direct computation shows that u′ belongs
1
to the Gehring class with Gp (u′ ) = K p , see [DS, Proposition 2.3]. By Lemma 5.2
u is a K-quasiminimizer. On the other hand, u′ does not belong to Lα (0, 1). This
1
shows that the open ended upper bound p1 (p, K p ) is sharp.
For p = 1 the integrability of u′ cannot be improved since every increasing
absolutely continuous function u is a 1-quasiminimizer. The theorem follows.
Remark 5.4. For p = 2,
1
p1 (2, t) = 1 + t(t2 − 1)− 2 , t > 1,
and hence 1 1 1
p1 (2, K 2 ) = 1 + K 2 (K − 1)− 2 , K > 1.

In [MS] the inverse functions of one dimensional quasiminimizers are also


considered.

6. Quasisuperharmonic functions
In the nonlinear potential theory superharmonic functions can be defined in
many ways. The most natural definition uses the comparison principle, see (6.3)
below. Let p > 1 and let Ω be an open set in Rn . A function u : Ω → (−∞, ∞]
is said to be superharmonic, i.e. (p, 1)-superharmonic, if
(6.1) u is lower semicontinuous,
(6.2) u ≡
6 ∞ in any component of Ω,

(6.3) for each open set Ω′ ⊂⊂ Ω and each minimizer h ∈ C(Ω ), i.e. (p, 1)-
quasiminimizer, the inequality u ≥ h on ∂Ω′ implies u ≥ h in Ω′ .
For the theory of superharmonic functions in the nonlinear situation see [HKM].
Superharmonic functions can also be defined with the help of minimizers, see
[HKM, Theorem 7.10] and [HKM, Corollary 7.20]. For other definitions see [B].
Lemma 6.4. Suppose that u : Ω → (−∞, ∞] satisfies (6.1) and (6.2). Then
u is (p, 1)-superharmonic if and only if there is an increasing sequence (u∗i ) of
(p, 1)-quasisuperminimizers, i.e. superminimizers, with u = lim u∗i . Here u∗i is
the lower semicontinuous representative of a superminimizer ui .
200 O. Martio IWQCMA05

Note that given a superharmonic function u the sequence u∗i = min(u, i), i =
1, 2, . . ., is the required sequence.
In view of Lemma 6.4 the following definition for (p, K)-quasisuperharmonicity
is natural.
Definition 6.5. Let Ω ⊂ Rn be an open set, p > 1 and K ≥ 1. A function
u : Ω → (−∞, ∞] is said to be (p, K)-quasisuperharmonic if there is an increasing
sequence of K-quasisuperminimizers ui in Ω such that lim u∗i = u and u 6≡ ∞ in
each component of Ω.
Lemma 6.6. Suppose that u is a (p, K)-quasisuperharmonic function in Ω and
locally bounded above. Then u is a (p, K)-quasisuperminimizer.
Proof. By the definition for quasisuperharmonicity there is an increasing se-
quence of quasisuperminimizers u∗i : Ω → (−∞, ∞) such that lim u∗i = u. From
Theorem 4.14 it follows that u is a K-quasisuperminimizer as required.
Note that a (p, K)-quasisuperharmonic function is automatically lower semi-
continuous as a limit of an increasing sequence of lower semicontinuous functions.
Lemma 6.7. Let u be a K-quasisuperharmonic function in Ω and h a (contin-
uous) minimizer in Ω. Then min(u, h) is a K-quasisuperminimizer (and hence
K-quasisuperharmonic) in Ω.
Proof. Let u∗i be an increasing sequence of K-quasisuperminimizers in Ω such
that u∗i → u. Now min(u∗i , h) is lower semicontinuous and by Corollary 4.4,
min(u∗i , h) is a K-quasisuperminimizer. Since min(u∗i , h) ≤ h, it follows that
min(u, h) = lim min(u∗i , h) is a K-quasisuperharmonic function. By Lemma 6.6,
min(u, h) is a K-quasisuperminimizer.
Theorem 6.8. Suppose that u : Ω → (−∞, ∞] satisfies (6.1) and (6.2). Then u
is a K-quasisuperharmonic if and only if min(u, c) is a K-quasisuperminimizer
for each c ∈ R.
Proof. The only if part follows from Lemma 6.7. For the sufficiency choose c =
i, i = 1, 2, . . .. Then min(u, i) is a lower semicontinuous K-quasisuperminimizer
and it follows from Definition 6.5 that u is a K-quasisuperharmonic function.
The theory for K-quasisuperharmonic functions is still in its infancy. However,
the following result was proved in [KM2]: A set C ⊂ Rn is said to be (p, K)-polar,
if there is a neighborhood Ω of C and a (p, K)-quasisuperharmonic function u in
Ω such that u(x) = ∞ for each x ∈ C. Then C is a (p, K)-polar set if and only
if the p-capacity of C is zero. It has been previously known that a set C ⊂ Rn
is a (p, 1)-polar set if and only if the p-capacity of C is zero. Hence allowing the
freedom due to K ≥ 1 adds nothing new to the structure of polar sets.

7. Appendix 1
Proof for Theorem 4.12. That P2 (u, Ω′ ) is a subminimizer in Ω′ follows
from Lemma 4.9. In order to show that w = P2 (u, Ω′ ) is a K-quasisuperminimizer
Quasiminimizers and Potential Theory 201

in Ω let Ω′′ ⊂⊂ Ω be open and v a function such that v − w ∈ W01,p (Ω′′ ) and
v ≥ w in Ω′′ . We set v = w in Ω \ Ω′′ . For the K-quasisuperminimizing property
of w it suffices to show
Z Z
p
(a) |∇w| dx ≤ K |∇v|p dx.
Ω′′ ∩{w<v} Ω′′ ∩{w<v}

Hence we may assume that w < v in Ω′′ although Ω′′ ∩ {w < v} need not be
an open set. Write A = {x ∈ Ω : u(x) < v(x)}. Then A ⊂ Ω′′ because if
x ∈ A \ Ω′′ , then u(x) < v(x) = w(x) which is a contradiction since u ≥ w. The
quasisuperminimizing property of u yields
Z Z
(b) |∇u| dx ≤ K |∇v|p dx,
p

A A

see Remark 4.2 (c). The function min(u, v) satisfies w ≤ min(u, v) ≤ u in Ω and
min(u, v) can be continued as w to Ω′′ \ {w < u}. The resulting function is in
the right Sobolev space. Note also that min(u, v) = w outside Ω′′ ∩ Ω′ and that
min(u, v) and w coincide outside {w < u} ∩ Ω′′ in Ω. Since w is the solution
to the Ku− (Ω′ )-obstacle problem, w ≤ min(u, v) and min(u, v) has the correct
boundary values w in {w < u} ∩ Ω′′ , we obtain
Z Z
p
|∇w| dx ≤ |∇ min(u, v)|p dx
{w<u}∩Ω′′ {w<u}∩Ω
Z′′ Z
(c)
p
= |∇u| dx + |∇v|p dx.
{w<u}∩Ω′′ ∩{u<v} {w<u}∩Ω′′ ∩{u≥v}

Since
(Ω′′ ∩ {w = u}) ∪ ({w < u} ∩ Ω′′ ∩ {u < v}) ⊂ Ω′′ ∩ {u < v},
the inequalities (b) and (c) yield
Z Z Z
p
|∇w| dx = |∇u|p dx + |∇w|p dx
Ω′′ Ω′′ ∩{w=u} Ω′′ ∩{w<u}
Z Z
p
≤ |∇u| dx + |∇u|p dx
Ω′′ ∩{w=u} {w<u}∩Ω′′ ∩{u<v}
Z
+ |∇v|p dx
{w<u}∩Ω′′ ∩{u≥v}
Z Z
p
≤ |∇u| dx + |∇v|p dx
Ω′′ ∩{u<v} {w<u}∩Ω′′ ∩{u≥v}
Z Z Z
p p
≤ K |∇v| dx + |∇v| dx ≤ K |∇v|p dx.
Ω′′ ∩{u<v} {w<u}∩Ω′′ ∩{u≥v} Ω′′
202 O. Martio IWQCMA05

This is (a). We leave to an exercise to show that w is a K-quasiminimizer in Ω′ .


The proof is complete.

8. Appendix 2
Proof for Theorem 4.14. We show that the quasisuperminimizing property
is preserved under the increasing convergence if the limit is locally bounded above
1,p
or belongs to Wloc (Ω).
The proof of this theorem [KM2, Theorem 6.1] contains a gap which will be
settled here. The argument is quite similar as in [KM2]. The authors would like
to thank professor Fumi–Yuki Maeda for pointing out the error in the original
paper.
We consider the case (i) only. The case (ii) follows from (i) and from an easy
truncation argument, see [KM2, p. 477]. In the case (i) it follows from the De
Giorgi type upper bound
Z Z
p
|∇ui | dx ≤ c(R − ρ) −p
(ui − k)p dx,
B(x,ρ) B(x,R)

where
k < − sup{ess sup ui : i = 1, 2, . . .},
B(x,R)

0 < ρ < R and B(x, R) ⊂⊂ Ω, that the sequence (|∇ui |) is uniformly bounded
1,p
in Lp (Ω′ ) for every Ω′ ⊂⊂ Ω. This implies that u ∈ Wloc (Ω) and we may assume
p ′
that (|∇ui |) converges weakly to ∇u in L (Ω ).
Let C ⊂ Ω be a compact set and for t > 0 write

C(t) = {x ∈ Ω : dist(x, C) < t}.

Then C(t) ⊂⊂ Ω for 0 < t < dist(C, ∂Ω) = t0 .

Lemma 8.1. Let u and ui be as in Theorem 4.12. Then for almost every t ∈
(0, t0 ) we have
Z Z
p
lim sup |∇ui | dx ≤ c |∇u|p dx
i→∞
C(t) C(t)

where the constant c depends only on K and p.

Proof. Let 0 < t′ < t < t0 and choose a Lipschitz cut-off function η such that
0 ≤ η ≤ 1, η = 0 in Ω \ C(t) and η = 1 in C(t′ ). Let

wi = ui + η(u − ui ), i = 1, 2, . . . .
Quasiminimizers and Potential Theory 203

Then wi − ui ∈ W01,p (C(t)) and wi ≥ ui a.e. in C(t). Hence the quasisupermini-


mizing property of ui gives
Z Z Z
p p
|∇ui | dx ≤ |∇ui | dx ≤ K |∇wi |p dx
C(t′ ) C(t) C(t)
Z
≤ αK (1 − η)p |∇ui |p dx
C(t)
Z Z !
+ |∇η|p (u − ui )p dx + η p |∇u|p dx ,
C(t) C(t)

where α = 2p−1 . Adding the term


Z
αK |∇ui |p dx
C(t′ )

to the both sides and taking into account that η = 1 in C(t′ ) we obtain
Z Z
p
(1 + αK) |∇ui | dx ≤ αK |∇ui |p dx
C(t′ ) C(t)
Z Z
p p
+αK |∇η| (u − ui ) dx + αK |∇u|p dx.
C(t) C(t)

Set Ψ : (0, t0 ) → R, Z
Ψ(t) = lim sup |∇ui |p dx.
i→∞
C(t)

Now −ui belongs to the De Giorgi class (see [KM2, Lemma 5.1]), and hence Ψ
is a finite valued and increasing function of t. Hence the points of discontinuity
form a countable set. Let t, 0 < t < t0 , be a point of continuity of Ψ. Letting
i → ∞, we obtain from the previous inequality the estimate
Z

(1 + αK)Ψ(t ) ≤ αKΨ(t) + αK |∇u|p dx,
C(t)

because Z
|∇η|p (u − ui )p dx → 0
C(t)

as i → ∞ by the Lebesgue monotone convergence theorem. Since t is a point of


continuity of Ψ, we conclude that
Z
(1 + αK)Ψ(t) ≤ αKΨ(t) + αK |∇u|p dx,
C(t)
204 O. Martio IWQCMA05

or in other words Z
Ψ(t) ≤ αK |∇u|p dx.
C(t)

1,p
Proof of Theorem 4.14, case (i). As noted before u ∈ Wloc (Ω). Let Ω′ ⊂⊂ Ω
be open and v ∈ W (Ω ), v ≥ u almost everywhere and v − u ∈ W01,p (Ω′ ). By
1,p ′

[KM2, Lemma 6.2] it suffices to show that


Z Z
(a) |∇u| dx ≤ |∇v|p dx.
p

Ω′ Ω′

To this end let ε > 0 and choose open sets Ω′′ and Ω0 such that
Ω′ ⊂⊂ Ω′′ ⊂⊂ Ω0 ⊂⊂ Ω
and
Z
(b) |∇u|p dx < ε.
Ω0 \Ω′

Next choose a Lipschitz cut-off function η with the properties η = 1 in a neigh-


borhood of Ω′ , 0 ≤ η ≤ 1 and η = 0 on Ω \ Ω′′ . Set
wi = ui + η(v − ui ), i = 1, 2, . . . .
Then wi − ui ∈ W01,p (Ω′′ ) and wi ≥ ui . Thus
Z !1/p Z !1/p
|∇wi |p dx ≤ ((1 − η)|∇ui |p + η|∇v|)p dx
Ω′′ Ω′′
Z !1/p
+ |∇η|p (v − ui )p dx
Ω′′
= αi + βi .
The elementary inequality
(αi + βi )p ≤ αip + pβi (αi + βi )p−1
implies that
Z Z Z
(c) p p
|∇wi | dx ≤ (1 − η)|∇ui | dx + η|∇v|p dx + pβi (αi + βi )p−1 ,
Ω′′ Ω′′ Ω′′

where we also used the convexity of the function t 7→ tp . We estimate the terms
on the right-hand side separately.
Since η = 1 in a neighborhood of Ω′ , there is a compact set C ⊂ Ω′′ such that
C ∩ Ω′ = ∅ and Z Z
(1 − η)|∇ui | dx ≤ |∇ui |p dx.
p

Ω′′ C
Quasiminimizers and Potential Theory 205

We can choose C = Ω′′ \ Ω′ (t) for sufficiently small t > 0. Next choose t > 0
such that Z Z
p
lim sup |∇ui | dx ≤ c |∇u|p dx
i→∞
C(t) C(t)

and C(t) ⊂ Ω0 \ Ω′ .
This is possible by Lemma 8.1. We have
Z Z
lim sup (1 − η)|∇ui | dx ≤ lim sup |∇ui |p dx
p
t→∞ i→∞
Ω′′ C
Z Z
p
≤ lim sup |∇ui | dx ≤ c |∇u|p dx ≤ cε
i→∞
C(t) C(t)

where the last inequality follows from (b). Since ∇η = 0 in Ω′ and v = u in


Ω′′ \ Ω′ the Lebesgue monotone convergence theorem yields βi → 0 as i → ∞.
On the other hand the numbers αi remain bounded as i → ∞. Hence it follows
from (c) that
Z Z Z
p p
lim sup |∇wi | dx ≤ cε + η|∇v| dx ≤ cε + |∇v|p dx.
i→∞
Ω′′ Ω′′ Ω′′
Now ui is a K-quasisuperminimizer and hence for large i it follows that
Z Z Z Z
p
|∇ui | dx ≤ p p
|∇ui | dx ≤ K |∇wi | dx ≤ 2Kcε + K |∇v|p dx
Ω′ Ω′′ Ω′′ Ω′′
Z Z
≤ 2Kcε + K |∇v|p dx + K |∇v|p dx
Ω′ Ω′′ \Ω′
Z
≤ 3Kcε + K |∇v|p dx,
Ω′

where we used (b) and the fact that ∇u = ∇v in Ω′′ \ Ω′ . Since ε > 0 was
arbitrary and since
Z Z
|∇u| dx ≤ lim inf |∇ui |p dx
p
i→∞
Ω′ Ω′

by the weak convergence ∇ui → ∇u in Lp (Ω′ ), this completes the proof of (a)
and the proof for the case (i) is complete.

References
[B] Björn, A., Characterization of p-superharmonic functions on metric spaces, Studia
Math. 169 (1) (2005), 45-62.
[BB] Björn, A. and J. Björn, Boundary regularity for p-harmonic functions and solutions
of the obstacle problem, preprint, Linköping, 2004.
[BBS] Björn, A., J. Björn and N. Shanmugalingam, The Perron method for p-harmonic
functions, J. Differential Equations 195 (2003), 398-429.
206 O. Martio IWQCMA05

[BJ1] Björn, J., Poincaré inequalities for powers and products of admissible weights, Ann.
Acad. Sci. Fenn. Math. 26 (2001), 175-188.
[BJ2] Björn, J., Boundary continuity for quasiminimizers on metric spaces, Illinois J. Math.
46 (2002), 383-403.
[DS] D’Apuzzo, L. and C. Sbordone, Reverse Hölder inequalities. A sharp result, Rend.
Matem. Ser. VII, 10 (1990), 357-366.
[DT] Di Benedetto, E. and N.S. Trudinger, Harnack inequalities for quasi-minima of vari-
ational integrals, Ann. Inst. H. Poincáre Anal. Non Lineáire 1 (1984), 295-308.
[GG1] Giaquinta, M. and E. Giusti, On the regularity of the minima of variational integrals,
Acta Math. 148, (1982), 31-46.
[GG2] Giaquinta, M. and E. Giusti, Qasiminima, Ann. Inst. H. Poincáre Anal. Non Lineáire
1 (1984), 79-104.
[G] Giaquinta, M: Multiple Integrals in the Calculus of Variations and Nonlinear Elliptic
Systems, Ann. of Math. Studies 105, Princeton Univ. Press, 1983.
[HKM] Heinonen, J., T. Kilpeläinen and O. Martio: Nonlinear Potential Theory of Degener-
ate Elliptic Equations, Oxford University Press, 1993.
[KM1] Kinnunen, J. and O. Martio: Nonlinear potential theory on metric spaces, Illinois J.
Math. 46 (2002), 857-883.
[KM2] Kinnunen, J. and O. Martio, Potential theory of quasiminimizers, Ann. Acad. Sci.
Fenn. Math. 28 (2003), 459-490.
[KS1] Kinnunen, J. and N. Shanmugalingam, Regularity of quasi-minimizers on metric
spaces, Manuscripta Math. 105 (2001), 401-423.
[KS2] Kinnunen, J. and N. Shanmugalingam, Polar sets on metric spaces, Trans. Amer.
Math. Soc. 358.1 (2005), 11-37.
[KSa] Krylov, N.V. and M.V. Safalow, Certain properties of solutions of parabolic equations
with measurable coefficients (Russian), Izv. Akad. Nauk. SSSR 40 (1980), 161-175.
[K] Korenovskii, A.A., The exact continuation for a reverse Hölder inequality and Muck-
enhoupt’s conditions, Transl. from Matem. Zametki, 52(6) (1992), 32-44.
[Ma] Marola, N., Moser’s method for minimizers on metric measure spaces, Helsinki Uni-
versity of Technology, Institute of Mathematics, Research Reports A 478, 2004.
[MS] Martio, O. and C. Sbordone, Quasiminimizers in one dimension – Integrability of the
derivative, inverse function and obstacle problems, Ann. Mat. Pura Appl., to appear.
[Sh1] Shanmugalingam, N., Newtonian spaces: An extension of Sobolev spaces to metric
measure spaces, Revista Math. Iberoamericana 16 (2000), 243-279.
[Sh2] Shanmugalingam, N., Harmonic functions on metric spaces, Illinois J. Math. 45
(2001), 1021-1050.
[Sh3] Shanmugalingam, N., Some convergence results for p-harmonic functions on metric
measure spaces, Proc. London Math. Soc. 87 (2003), 226-246.

Olli Martio E-mail: olli.martio@helsinki.fi


Address: Department of Mathematics and Statistics, BOX 68, FI-00014 University of Helsinki,
FINLAND
Proceedings of the International Workshop on Quasiconformal
Mappings and their Applications (IWQCMA05)

History and Recent Developments in Techniques for


Numerical Conformal Mapping

R. Michael Porter

Abstract. A brief outline is given of some of the main historical develop-


ments in the theory and practice of conformal mappings. Originating with the
science of cartography, conformal mappings has given rise to many highly so-
phisticated methods. We emphasize the principles of mathematical discovery
involved in the development of numerical methods, through several examples.

Keywords. numerical conformal mapping, cartography, osculation, interpo-


lating polynomial method, mathematical discovery.
2000 MSC. 65-03 30C30 65-02 65S05.

Contents

1. Introduction 208
2. A Brief History of Mapmaking 208
3. The General Problem of Conformal Mappings 214
4. The Crowding Problem 216
5. Elementary Facts about Analytic Functions 217
6. Osculation Methods 218
6.1. Koebe’s method. 218
6.2. Graphical methods. 219
6.3. Grassmann’s method. 220
6.4. Sinh-log method 220
7. Schwarz-Christoffel Methods 222
8. Rapidly Converging Methods 223
8.1. Theodorsen’s Method. 223
8.2. Fornberg’s method. 225
9. Generalities on Conformal Mapping Methods 226
10. Interpolating Polynomial Method 226
10.1. Some properties of the half-click mapping. 228
10.2. Interpolating polynomial algorithm 229
Research supported by CONACyt grant 46936.
208 R. Michael Porter IWQCMA05

10.3. Numerical example 230


11. The Quest for Better Methods 233
11.1. Methods using derivatives 233
11.2. Method of simultaneous interpolation 233
11.3. Minimization approach 235
12. Combined Methods 235
13. Epilogue 236
References 237

1. Introduction
Although the present meeting is mainly concerned with the study of the the-
ory of quasiconformal mapping, our topic here is to understand some of the basic
principles of the theory of conformal mapping, with emphasis on the computa-
tional perspective.
The history of quasiconformal mappings is usually traced back to the early
1800’s with a solution by C. F. Gauss to a problem which will be briefly mentioned
at the end of Section 2, while conformal mapping goes back to the ideas of G.
Mercator in the 16th century. Since the early work had much to do with the
production of maps of the physical world, we will begin with a brief survey of
how this came about.

2. A Brief History of Mapmaking


From antiquity mankind has needed to map out his world: whether for con-
trolling dominated territories or to travel great distances. For the moment, by
map we will mean a representation of part of the earth’s surface on a flat paper.
Indeed, as one historian writes, mapmaking is older than the written word:
“The human activity of graphically translating one’s perception of his
world is now generally recognized as a universally acquired skill and one
that pre-dates virtually all other forms of written communication.”1

A map obviously must contain “known reference points” and somehow show
their relative distances and directions. This is the basis of the humor in the
following lines from “The Hunting of the Snark” by Lewis Carroll,

1
www.henry-davis.com/MAPS/AncientWebPages/100mono.html. As with many of the web
references cited here, the same information may be found on many sites and it is difficult to
pinpoint an original source.
Techniques for Numerical Conformal Mapping 209

Figure 1. “Map” lacking reference points

He had bought a large map2 representing the sea,


Without the least vestige of land:
And the crew were much pleased when they found it to be
A map they could all understand.

“What’s the good of Mercator’s North Poles and Equators,


Tropics, Zones, and Meridian Lines?”
So the Bellman would cry: and the crew would reply
“They are merely conventional signs!”

“Other maps are such shapes, with their islands and capes!
But we’ve got our brave Captain to thank”
(So the crew would protest) “that he’s bought us the best
A perfect and absolute blank!”

2
http://www.eq5.net/carrol/fit2.html

Figure 2. Left: Perspective view of terrain. Right: Topographi-


cal map of same terrain.
210 R. Michael Porter IWQCMA05

Figure 3. Right: Reconstruction of Catalhoyuk map, which is


said to have had the function of registering property ownership.

The picture of a terrain in Figure 2 (left) is likewise not a map. In the map of
Figure 2 (right), the reference points are “subjective,” involving loci of constant
height which need not correspond to any physically noticeable characteristics of
the terrain. Maps may incorporate our preconceptions or prejudices of what the
world looks like.
Of course, in modern mathematics the matter of reference points is resolved
by the precise notion “function,” for which it is postulated that every point of
the surface is made to correspond, albeit theoretically, to a unique point of the
map.
Let us take a quick trip though the history of maps. A nine-foot-long stone
map dating from 6200 B.C., which appears to be a plan of a town predating
Ankara, Turkey3 , is shown in Figure 3. It is said to have served to register
property rights, perhaps for tax purposes.
Moving now to larger-scale maps, we have a map of Africa produced in silk,
in 1389 by the Chinese map Great Ming Empire), measuring 17 square meters
(Figure 4). It is on display in South Africa and is said to be a copy of an earlier
stone one4 . However, it would be a mistake to say that the larger the scale, the
more recent the map. In 1999 there was discovered in Ireland a map dating from
3
John F. Brock, “The Oldest Cadastral Plan Ever Found: the Catalhoyuk town plan of
6200 B.C.,” http://www.mash.org.au/articles/articles2.htm
4
BBC News, http://news.bbc.co.uk/2/hi/africa/2446907.stm

Figure 4. Chinese map of Africa


Techniques for Numerical Conformal Mapping 211

Figure 5. Alleged ancient moon map

about 3000 B.C. which is claimed to be the oldest known map of the moon.5
Even older are the well-known drawings in from 14,000 B.C the Lascaux Caves,
France, discovered in 1940, with a drawing of the constellation Pleiades and of
the “Summer Triangle,” and a map of the constellation Orion from 30,000 B.C.
found engraved on a 4 cm. ivory tablet in Germany in 1979.6
At the time of this writing, it is being investigated whether or not the Chinese
admiral Zheng He discovered America and circled the globe 80 years before the
voyages of Christopher Columbus; part of the argument is based on a map from
the year 1418.7 At any rate, it is often stated incorrectly that in the epoch of
Columbus, people generally thought that the earth was flat. While the spherical
nature of the earth’s surface was known to the ancient Greeks, and its diameter
was measured by Hipparcus (190–120 A.C.), it is significant that no early map
was produced similar to that of Figure 6.
The principle here is that consciously or not, one tends to look for the simplest
possible answers to mathematical questions. Here it would be the topology which
is unnecessarily complicated. The possibility of closed curves on the earth’s
5
http://news.bbc.co.uk/1/hi/sci/tech/1205638.stm
6
http://news.bbc.co.uk/1/hi/sci/tech/871930.stm and http://news.bbc.co.uk/1/
hi/sci/tech/2679675.stm
7
The Sunday Statesman, New Delhi, 15 January, 2006; http://edition.cnn.com/2003
/SHOWBIZ/books/01/13/1421/index.html

Figure 6.
212 R. Michael Porter IWQCMA05

Figure 7. If there were an isometry φ from a neighborhood of


the North Pole to a planar region, then circles of radius r in the
two spaces would have different circumferences.

surface which are not contractible to a point would have many consequences,
among them a nonconstant curvature of the surface (at least according to our
intuitive notions of Euclidean space).
So let us return to the idea of a spherical earth. The oldest known maps of the
celestial sphere were not on flat surfaces such as paper, but rather on tortoise
shells. One reason for this may have been that, as Figure 7 shows, there can
be no isometry (that is, a distance-preserving map) from a spherical region to a
planar one.
This statement may sound disappointing, because it would be extremely useful
to be able to determine one’s position on a map using the distance one has
traveled from a previously determined point. However, navigators of the Middle
Ages realized that if distances cannot be preserved on planar maps of spherical
surfaces, then at least it would be useful to conserve angles. When the angles
between curves on the earth are equal to the corresponding angles on the map, the
map is said to be conformal (Figure 8). A navigator on the high seas, orienting

Figure 8. Conformal mapping of plane domains and of surfaces.


Techniques for Numerical Conformal Mapping 213

himself by the stars, has to go through some rather involved calculations to


determine the distance he has just traveled. However, the stars tell him quickly
the direction in which he is traveling, and hence the angle which his course is
making with lines of latitude and longitude. Thus a conformal (nonisometric)
map is not necessarily as impractical as it may first seem.
As is mentioned also in A. Rasila’s article, the idea of a conformal earth map
was developed by G. Mercator. The map is designed in two steps. The first
step is to project all points of the sphere except the poles to a cylinder in which
the sphere is inscribed (Figure 9). Then the horizontal lines in the image are
to be adjusted, placing the line corresponding to latitude θ at the height ϕ(θ).
Mercator did not know how to do this second step precisely; it was E. Wright in
1599 who found the theoretical solution,
Z θ

ϕ(θ) = .
0 cos θ

At that time no one knew how to evaluate this integral in elementary terms.
This led to the compilation of numerical tables for navigators, which was carried
out by successively summing small increments of the integrand. After some time,
people compared different tables and observed a curious coincidence: the integral
Z θ

0 cos θ

(1)-

(2)
?

Figure 9. Steps in Mercator projection


214 R. Michael Porter IWQCMA05

was apparently approximately equal to


θ π 
log tan + ,
2 4
a statement which became a conjecture and was finally proved in 1668 by J. Gre-
gory.
The point of this rather long digression is to illustrate that the interplay
between “pure” and “applied” mathematics has been with us for a long time.
Numerical evidence has long played an important role in the generation of con-
jectures and theorems; leaving aside questions arising from ancient geodesy, we
mention only the Prime Number Theorem conjectured by Gauss, and the location
of zeroes of the zeta function known as the Riemann Hypothesis.
Let us look again at Figure 8. The problem of finding a conformal correspon-
dence between an arbitrary region and a plane region is rather complicated, and
classically is called the problem of finding isothermic coordinates for a surface.
For real-analytic surfaces, this was solved by Gauss. The solution requires quasi-
conformal mappings as an intermediate step, and will not be explained in detail
here.

3. The General Problem of Conformal Mappings


Here we limit the discussion to plane domains. From the theory of functions
of a Complex Variable we have the following basic fact.
A correspondence f between plane domains is conformal if and only
if it can be represented locally as an analytic function of a complex
variable z = x + iy,
X∞
f (z) = ak (z − z0 )k .
0

Thus, once a conformal mapping is found from a given surface to a plane


region, all other conformal mappings are obtained from this one by composing
with an analytic function.
Applications of conformal mapping to physics, too numerous to mention here,
require finding conformal mappings between two given domains. For the simplest
situation, that of simply connected domains, the problem of finding a single
mapping is clarified by the following well-known result.
Theorem 1. RIEMANN MAPPING THEOREM: There exists a conformal map-
ping f from the unit disk D = {z : |z| < 1} to any simply connected proper
subdomain D of the complex plane (it is unique if suitably normalized).

Knowing that a conformal mapping exists is not the same as knowing how
to solve the numerical problem: given D, to calculate f to a given degree of
accuracy. In some simple cases the conformal mapping can be written as a
composition of elementary functions, such as Möbius transformations, powers
Techniques for Numerical Conformal Mapping 215

and roots, trigonometric functions, elliptic integrals, and the like. These basic
analytic formulas are also useful for reducing general problems to more accessible
ones (see Section 6.2 below). In practice, a mapping problem can be presented in
a variety of ways, partly because a domain can be presented in a variety of ways:
as set of points satisfying a certain condition (such as an inequality F (z) < c), or
a specific list of points (such as a set of pixels in an image). More commonly, a
domain is specified in terms of its boundary, which may be a condition satisfied
by its points (such as an equality F (z) = c), or a parametrization z = γ(t), or
simply a list of points along the boundary.
Furthermore, a map from one domain D1 to D2 is often described as a com-
position of a map from D to D1 composed with the inverse of a map from D to
D2 . Depending on the nature of the numerical description, it may be difficult or
inconvenient to calculate the inverse map.
The following result is also relevant to numerical work.
Theorem 2. (Carathéodory) Every conformal mapping between to Jordan re-
gions extends to a homeomorphism of the closures of the regions.

Because of the Cauchy Integral Formula


Z
f (z)
(1) f (z0 ) = dz, z0 ∈ D,
∂ D z − z0
which is valid for analytic functions in D continuous on the boundary, it is
sufficient to know the values of a conformal mapping on the boundary of D. This
observation reduces the complexity of the conformal mapping problem from 2

Given data:
(i) D (or ∂D)
(ii) z0 ∈ D

Figure 10. Illustration by G. Francis in [1]. Equipotential lines


of ideal fluid flow correspond to concentric circles under Riemann
mapping.
216 R. Michael Porter IWQCMA05

dimensions to 1, and this is why the vast majority of the numerical methods
which have been developed focus on calculating the boundary mapping f |∂ D . A
numerical scheme for evaluating (1), once the boundary mapping is known, can
be found in [12].
For this reason, we will consider the problem as defined in Figure 11 (except
for Section 6, where the inverse mapping will be sought). The closed curve
t 7→ γ(t) defines the boundary of D. A point z0 fixed inside of γ is to be equal
to f (0). We are interested in the boundary values f (eiθ ) for 0 ≤ θ < 2π. The
mapping θ 7→ eiθ of the interval to the unit circumference is so natural that we
use it in the statement of the problem: to find a function t = b(θ) so that
(2) f (eiθ ) = γ(b(θ)).

 
γ b eiθ


0 2π 0 2π
t θ
Figure 11. Elements of mapping problem (picture taken from [8])

4. The Crowding Problem


When f is a conformal mapping, the value of |f ′ | (or the density of the grid
points required) may vary greatly along different parts of the boundary. The
crowding factor is the ratio of the greatest to the least value of |f ′ |. For exam-
ple, consider conformal maps fa from D (or from a square, as in Figure 12) to
rectangles of height 1 and base a > 0, with fa (0) at the center of the rectangle.
Then the crowding factor grows exponentially with a as a → ∞.
This creates a numerical difficulty. Suppose we have calculated b(2πj/N ) for
j = 1, . . . , N , that is, for N equally spaced values of θ. If the crowding factor
Techniques for Numerical Conformal Mapping 217

Figure 12. Conformal maps between non-similar rectangles can-


not send vertices to vertices.
is very large, then many of these values will be bunched together, while other
points will be spread relatively far apart. This may cause an imperfect picture
of the behavior the mapping function (see for example the last picture in Figure
29 below). The crowding factor for the domain in Figure 11 is approximately
1000.

5. Elementary Facts about Analytic Functions


There are a great many theorems in Complex Analysis of the form “the func-
tion f is analytic if and only if . . . ”. A conformal mapping is the same thing as
a 1-to-1 analytic function, and this notion also admits many characterizations.
Theorem 3. An analytic and 1-to-1 function f : D → D is onto D when |f ′ (z0 )|
is maximal among the class of injective analytic functions D → D sending the
point z0 to 0.
Likewise there are many characterizations of boundary values on ∂ D = {|z| =
1} for conformal mappings. We list without detailed explanations, a few which
have been used as the basis for numerical methods.
Theorem 4. A continuous function γ : ∂ D → C is the boundary value of an
analytic function if:
1. [12] K[Re γ] = Im γ, where K is the conjugate boundary operator;
or if P
2. [8] its Fourier series γ(t) = ∞ −∞ bk e
kit
has bk = 0 for k < 0;
or if Rβ
3. [17],[22] |γ ′ (t)|1/2 S(γ(t), z0 ) + 0 |γ ′ (t)|1/2 A(γ(t), γ(s))|γ ′ (s)|1/2
= |γ ′ (t)|1/2 H(z0 , γ(t)),
where A and H are the Kerzman-Stein and the Cauchy kernels;
or if
4. etc., etc.
Generally speaking, we may say that each theoretical characterization of con-
formal mappings can lead to a numerical method or to a family of numerical
methods. At the risk of oversimplifying, we can say that most methods fall into
one of the following two classes.
• “Easy Methods”: D → D
• “Fast Methods”: D → D
218 R. Michael Porter IWQCMA05

× ×

1. 2.

×
×

3. 4.

Figure 13. One iteration of the Koebe square-root method

We give some illustrations below.

6. Osculation Methods
Most of the “Easy Methods” are classified as osculation methods, which consist
of first mapping D into D, and then mapping the image region to a larger
region inside of D, and so on. The desired approximation to f : D → D is the
composition of these mappings, and f itself is their limit.

6.1. Koebe’s method. The first osculation method ever created is based on
the proof given by P. Koebe in 1905 for Theorem 3, and which is found in many
Complex Variables texts, such as [2], [5]. This procedure is illustrated in Figure
13. The steps are prescribed as follows.
0. Suppose D ⊂ D

1. Find t0 with |γ(t0 )| < 1 minimal (“worst


boundary point”). Let a0 = γ(t0 )
w − a0
2. Move a0 to 0: w1 =
1 − a0 w

3. Take square root: w2 = w1

4. Move the image b0 of zero back to 0:


w2 − b 0
w=
1 − b 0 w2
The auxiliary mappings defined in steps 2, 3, and 4 combine to give a map
from D to itself which fixes the origin and has derivative greater than 1. Thus
Techniques for Numerical Conformal Mapping 219

the image under each successive iteration is “larger” than the previous one. Note
that if D = D, then step 2 cannot be applied.
The images of a domain under successive iterations of the Koebe method are
shown in Figure 14. The convergence to ∂D can be made slightly faster by using
the cube root instead of the square root in Step 3. If one uses the nth root and
lets n → ∞, one approaches the Koebe logarithm method, in which the logarithm
replaces the nth root and step 4 is modified accordingly.

6.2. Graphical methods. In the 1950’s, people were looking for ways to find a
mapping of a given domain to a domain reasonably close to D, in order to apply
then a fast method (such as Teodorsen’s method below) to the result. Today it
seems incredible that this was done by hand. For example, in [11] a method was
described by which the operator first fits manually the given domain D into a
disk from which a sector bounded by two arcs has been removed (Figure 15). A
conformal mapping from this slightly larger domain D1 ⊃ D to D can be written
in elementary terms,
 1/α
c + z1 c+z
(3) =k .
c − z1 c−z
as a composition of a power function with two Möbius transformations. These
Möbius transformations make the points 0, ∞ correspond to key points in the
original figure. An ingenious scheme was presented in which one places the
domain over specially designed “graph paper” showing circles passing through

Figure 14. Iterations of the Koebe method (left) and the loga-
rithmic Koebe method (right)
220 R. Michael Porter IWQCMA05

these two points and circles orthogonal to them. Then, much in the way that
one can magnify a picture by tracing a grid over it and then copying the part of
the picture in each small square in it to the corresponding square in a larger grid,
pieces of the boundary of the domain are copied to corresponding pieces within a
second graph paper. In this way the conformal mapping is approximated without
computing the elements of (3) numerically.

6.3. Grassmann’s method. In 1979, E. Grassmann [10] automated and re-


fined this idea of Albrecht-Heinhold, giving a much faster osculation method. The
first step is to detect automatically (that is, by a computer program) whether
an auxiliary mapping which opens a slit can be applied, and if not, applies other
alternatives. In the worst case, the Koebe method is used (Figure 16). The idea
behind this method is that for many domains, after a few iterations the domain is
“opened up” sufficiently so that the better auxiliary mappings will be applicable.

6.4. Sinh-log method. This method is based on the observation [19] that in
the logarithmic Koebe method, at a certain step during each iteration, the log-
arithmic image of the domain generally not only lies in the left half-plane, but
within some horizontal band; that is, the imaginary parts of all its points are
bounded above and below. An appropriate real-affine transformation takes this
half-band to the normalized half-band {−π/2 < Im z < π/2}, which in turn the
hyperbolic sine function sends to the whole left half-plane, as shown in Figure 17.
As a result, a much larger part of the half-plane is covered before the half-plane
is mapped back to the unit disk in the last step of the iterative procedure.

Figure 15. Manual graphical approach of Albrecht and Heinhold


Techniques for Numerical Conformal Mapping 221

Figure 16. Application of Grassmann’s method to a “daisy”


domain. Each figure shows the slit which is opened to form the
following one.

Thus the “sinh-log” method is defined by following steps 0, 1, and 2 of the


Koebe square-root algorithm, then (3) taking the logarithm of the resulting im-
age; (4) applying an affine transformation to a subdomain of the normalized
half-band; (5) applying sinh, and finally (6) a Möbius transformation of the left
half-plane to D so that the composition of all the maps mentioned fixes the
origin.
The sinh-log algorithm converges roughly as does Grassmann’s for many stan-
dard domains and for highly irregular domains it performs much better, while
being easier to program and more stable numerically. In particular, if one is
interested in calculating the composed mapping f and its inverse, it is easier to
save the data for a sequence of elementary maps of a single kind.

Figure 17. One iteration of the sinh-log method. The lower left
picture shows the half-strip containing the logarithmic image, as
well as the image of this under sinh
222 R. Michael Porter IWQCMA05

Figure 18. Iterations of the sinh-log method

It was mentioned that osculations are very “slow” methods. Their great ad-
vantage is that they apply to any domain bounded by a continuous (say piecewise
smooth) closed curve. In general, osculation methods have the following charac-
teristics.

Osculation Methods

COST OF ONE ITERATION: O(N )

RATE OF CONVERGENCE: very slow

GENERALITY: total

7. Schwarz-Christoffel Methods
The Schwarz-Christoffel methods are applicable to the particular case of a
domain with polygonal boundary. Although for numerical work every domain
can be considered in principle “polygonal,” if the number of vertices is extremely
large the advantage of Schwarz-Christoffel methods would be lost. The Schwarz-
Christoffel formula says that the Riemann mapping from the unit disk to a
polygonal domain is equal to the integral
Z z
dz
(4) f (z) = .
0 (z − z1 ) (z − z2 )1−α2 /π · · · (z − zn )1−αn /π
1−α 1 /π
Techniques for Numerical Conformal Mapping 223

α1
w3
w1
w2
Figure 19
For a given polygonal domain, the vertices wj = f (zj ) are known as well as
are the interior angles παj , but the prevertices zj are not known, so formula (4)
is useless until these values are determined. A strategy for a typical Schwarz-
Christoffel method would thus be
• guess approximate values for z1 , z2 , ...,
• evaluate the Schwarz-Christoffel integral,
• compare the results with w1 , w2 , ...,
• apply some type of correction
The difficulty is to make an initial guess sufficiently close for this to work.
A history and detailed explanation of various methods can be found in [6]. An
elegant solution to the Schwarz-Christoffel mapping question was invented by
T. A. Driscoll and S. A. Vavasis [7], which begins by triangulating the polygonal
domain in a special way, and then solving a set of equations for the cross ratios of
all rectangles formed by pairs of adjacent triangles. Not only is there no problem
to find an appropriate initial guess, but also the invariance of the cross ratio
makes it possible to avoid the crowding phenomenon: one can apply a Möbius
transformation of D to bring any part of the domain into focus.

8. Rapidly Converging Methods


It must be stressed that we will mention only a few of the very many methods
which have been developed. Details and history of many classical methods will
be found in [12]; for more methods see Section 13.

8.1. Theodorsen’s Method. This method, presented in 1931 for improving


the design of aircraft wings, applies only to starshaped domains (λz is in D
when z is in D and 0 < λ < 1). The region exterior to an airplane wing, in
which the physics of air movement takes place, is obviously not of this form, but
can be reduced to it by a suitable auxiliary transformation (Joukowski profile,
Figure 21).
So we assume from the start that the boundary ∂D is traced out by
(5) γ(t) = ρ(t)eit , 0 ≤ t ≤ 2π.
224 R. Michael Porter IWQCMA05

The idea behind the algorithm is that of condition 1 in Theorem 4. Consider


the real and imaginary parts u,v of the analytic function f (z) = u(z) + iv(z).
They are known to be harmonic conjugates. Now if we are given a function u0
defined only on ∂ D, it can be extended (by the Poisson integral) to a harmonic
function u in all of D. This function has a harmonic conjugate v defined in ∂ D.
The restriction v0 of v to ∂ D is called the conjugate boundary function of u0 , and
is written
(6) K[u0 ] = v0 .

Now we apply the following facts.


1. The conjugate boundary function can be calculated by a singular integral
Z 2π
it t−s
K[v0 ](e ) = v0 (eis ) cot ds.
0 2
2. If γ(b(θ)) = u0 (eiθ ) + iv0 (eiθ ) and if v0 = K[u0 ], then γ ◦ b defines the
boundary values of a conformal map.
Recall from Section 3 that we want t = b(θ) such that γ(b(θ)) defines the
boundary values of an analytic function according to (2). This is equivalent to
b(θ) − θ = K[log(ρ ◦ b)](θ).

The difference δ(θ) = b(θ) − θ satisfies


(7) δ(θ) = K[log(ρ(θ + δ(θ))].

Thus δ is a fixed point of a nonlinear operator, and Theodorsen’s method says


to construct a sequence of functions by iterating it,
δk+1 (θ) = K[ log(ρ(θ + δk (θ))) ].
Then δk → δ as k → ∞. The solution b is now given by b(θ) = θ + δ(θ).

Figure 20. Facsimile of Theodorsen’s Naca internal report


Techniques for Numerical Conformal Mapping 225

0.5

-1 -0.5 0.5 1

-0.5

-1

Figure 21. Joukowski profile and starshaped domain

Theodorsen’s Method

COST OF ONE INTERATION: O(N 2 )

RATE OF CONVERGENCE: linear

GENERALITY: starshaped domain;


close initial guess

8.2. Fornberg’s method. This method, published in 1980 in [8], is based on


criterion 2 of Theorem 4. We have as data for the problem γ : [0, 2π] → ∂D, a
periodic complex-valued function, together with a point z0 inside of D. We want
to find the boundary values γ ◦ b of the conformal mapping from D to D sending
0 to z0 . Like any map of the circle, it must have a Fourier series:

X
(8) γ(b(t)) = bk ekit .
k=−∞

From what
P we have said γ(b(θ)) is to give the values on ∂D of an analytic function
f (z) = ∞ a
k=0 k z k
, so by (2)

X ∞
X
(9) ak ekiθ = f (eiθ ) = γ(b(θ)) = bk ekiθ
k=0 k=−∞

and therefore 
0, k < 0,
bk =
ak , k ≥ 0.

This suggests the following procedure.


226 R. Michael Porter IWQCMA05

0. Guess initial values t1 , t2 , . . . , tN to be assigned as b(2jπ/N ) =


tj , i.e., we hope that f (e2jπi/N ) = γ(tj ) approximately.
1. For b determined this way, calculate the Fourier coefficients for
γ ◦ b:
b−N , b−(N −1) , . . . , b−1
2. Calculate the corresponding changes in bk which would go with
slightly different values of tj ,
t1 + ∆t1 , t2 + ∆t2 , . . . , tN + ∆tN
3. Solve a linear system for the differences
(∆t1 , ∆t2 , . . . , ∆tN )
to make the coefficients bk equal to zero (approximately).

Fornberg’s Method

COST OF ONE ITERATION: O(N log N )

RATE OF CONVERGENCE: linear

GENERALITY: requires close initial solution

9. Generalities on Conformal Mapping Methods


We have mentioned that there are a great many conformal mapping methods,
each based on a specific property of functions of a complex variable. Here we
have seen only a few. There are other better ones; for example, Wegmann’s
method [25] based on the Riemann-Hilbert problem, has the same O(N log N )
iteration cost and offers quadratic convergence. We suggest the following general
approach for discovering new such methods.
• Choose a characterization of analytic mapping functions
• Use it to measure in some sense how much a given function falls short of
this criterion
• Write an equation to describe this numerically
• Apply some numerical method to solve this equation
• . . . and hope that it works!

10. Interpolating Polynomial Method


We describe here the method presented in [18], together with some of the
ideas that led to its development, and some ideas for variations of the method.
We take as the first step in our heuristic procedure the following fact proved by
Wegmann [24].
Techniques for Numerical Conformal Mapping 227

P 2πk
i
e
2πk
N
i - P (e N )wk

Figure 22. Polynomial approximation of f

Theorem 5. For ∂D sufficiently smooth and for N sufficiently large, there is a


unique (suitably normalized) polynomial of degree N + 1 close to f which takes
the 2N -th roots of unity to points cyclicly ordered along ∂D. These polynomials
approach f as N → ∞.
Let us explain what is involved in this result. Recall that our problem is to
find wj ∈ ∂D such that e2πij/N ↔ wj .
Consider some points w1 , w2 , . . . , wN along ∂D. Let us investigate the hypoth-
2πk
esis that these indeed could be the values of f (e N i ) for a Riemann mapping f
from the unit disk to D.
Let us suppose that f is to be approximated by a polynomial P (z) of degree
2πi
N . Now, it is very easy to find a polynomial P (z) such that P (ej N ) = wj ,
j = 1, . . . , N . Thus the existence of such a polynomial tells us nothing about
whether (wj ) are the right points or not for f .
2πi
For N fixed, let us write ζj = ej N . Wegmann’s theorem says that if (wj )
are the right points for f at ζk , then the polynomial approximation P will not
only be right for f at ζj , but also at ζj+ 1 . For ease of reference, I will call
2
ζj+ 1 the “half-click” points. In Figure 23 we take D = D, z0 = 0, so that the
2
solution is f = identity (or any rotation). We have deliberately taken wk not to
be equally spaced along ∂D, and then calculated the image of all of ∂ D under
the corresponding interpolating polynomial. The images of the half-click points
are marked with an ×.
When wj are not the correct values for f (ζj ), then the polynomial approxima-
tion may be very far from ∂D at the half-click points. Further, the image P (∂ D)
may not be a simple curve.
This leads us to phrase the criterion for the conformal mapping: first we
discretize the problem to the N points ζj . Our complex-analysis criterion is
• The truncated power series of f of degree N maps all the
2N -th roots of unity close to ∂D.
The next step in the development of our algorithm is to “measure” how much
a given set of guessed points (wj ) fails to fulfill this criterion.
We choose N to be a power of 2 (to facilitate the numerical work). Given
~ = (wj ), there is a unique P = Pw~ , the interpolating polynomial for w,
w ~ such
228 R. Michael Porter IWQCMA05

that
(10) Pw~ is a polynomial of degree ≤ N ,
Pw~ (0) = 0,
2πi
Pw~ (e N j ) = wj for 0 ≤ j ≤ N − 1.
Since P is easy to calculate, so are its half-click values
 2πi 1 
(11) uj = Pw~ e N (j+ 2 ) .
The answer is given by a matrix multiplication,
(12) ~u = C w,
~
where
−1 i π 1
(13) cj = + cot( (j + )),
N N N 2
Cjk = cj−k ).
C is a circulant matrix, and C w
~ can be calculated via Fast Fourier Transform
(FFT) in O(N log N ) operations. This is also a key feature of Fornberg’s method.

10.1. Some properties of the half-click mapping. C is an orthogonal ma-


trix, and satisfies
C 2 = E = shift left by one index,
C = R[1/2] ,
E = R[1]
where
2πi
R[β] (w) = w[β] = (Pw~ (eiβ ej N ))j=0,...,N −1 .
By analogy one may call R a “β-click” of the W -values.
Thus we have uj = Pw~ (wj ), ~u = C(w),
~ C(~u) = E(w), ~ (C(~u))j = uj−1 .
With this we can measure the discrepancy of a given w~ from the criterion of
Wegmann’s theorem: define ρ = ρD to be the projection onto the nearest point

wk
uk
wk+1

Figure 23. Image of ∂ D under interpolating polynomial


Techniques for Numerical Conformal Mapping 229

u∗j
H
Y
C
ρ
?

ρ(u∗j )
wj+1 wj∗

~ 7→ ρ(C(w))
Figure 24. w ~

of ∂D = γ([0, 2π]). This is defined for points sufficiently near to ∂D. Then the
discrepancy can be described by the N -vector
(uj − ρ(uj ))j .

10.2. Interpolating polynomial algorithm. There are variations on this way


of describing the discrepancy. Begin with trial values
~ = (w0 , w1 , . . . , wN −1 ) ∈ (∂D)N .
w
Calculate
~ 7→ ρ(C(w)).
w ~
~ were the true solution, then ρ ◦ C(w)
If w ~ ∈ ∂D.
~ would be the same as C(w)
If we repeat the process, then the image of wj should go over to wj+1 , since
2
C = E. This suggests looking at ρ ◦ C ◦ ρ ◦ C and comparing it with the shift
E. We define the basic step of the algorithm as something very similar, and even
more convenient:
(14) Φ = ΦD = ρ ◦ C −1 ◦ ρ ◦ C
~ is a fixed point of Φ, it follows that ρ(uj ) = uj , so uj ∈ ∂D as required.
When (w)
The numerical method for solving Φ(w) ~ = w, ~ as it was in Theodorsen’s
method, is simply by iteration towards an attractive fixed point. When the orig-
~ ∗ is close enough to ∂D, the convergence to a fixed point of Φ is generally
inal w
linear.
To see more clearly why “ρC” appears twice in the definition of Φ, note that
the space of solutions normalized by f (0) = 0 can be identified with the circle
S 1 , being formed of
fβ (z) = f (eiβ z), 0 ≤ β < 2π.
We can think of w ~ = (w0 , w1 , . . . , wN −1 ) being shifted (or rotated) along ∂D
[β] [β] [β] [β]
to w ~ = (w0 , w1 , . . . , wN −1 ). The “ρC” algorithm approaches this space of
solutions, but not a particular solution.
The convergence characteristics of the Interpolating Polynomial method are
the same as for Fornberg’s. Recall that we have fixed N , and Φ depends upon
N . Therefore a solution of Φ(w)~ = w ~ is not a true solution of the mapping
problem, but rather a solution to the truncated problem for N -th degree. A true
230 R. Michael Porter IWQCMA05

algorithm for approximating the Riemann mapping to arbitrary accuracy would


have to involve a step for increasing the value of N at appropriate moments.

10.3. Numerical example. For the first example, we will calculate the same
Riemann mapping as did Fornberg in 1980. His domain Dα , shown in Figure 25,
is bounded by the curve
(15) ((Re w − .5)2 + (Im w − α)2 )(1 − (Re w − .5)2 − (Im w)2 ) − 1 = 0

1.
2.0

1.0
0.5

0.5

0 0.5 1.

-1.

Figure 25. Domain defined by equation (15) for α = 0.5, 1.0, 2.0

Iterating Φγ : Method of Fornberg:


α N FFTs k∆wk
~ ∞ k∆~ak∞ FFTs Accuracy of
Taylor coefs.
2.0 4 16 .48×10−1
2.0 8 4 .89×10−1
2.0 16 4 .16×10−1
2.0 32 4 .18×10−2
2.0 64 8 .91×10−5
2.0 128 0 —
1.5 128 24 .13×10−8 .70×10−9 96 .14×10−7
1.2 128 36 .45×10−8 .37×10−9 100 .33×10−5
1.2 256 0 — 50 .97×10−8
1.0 256 48 .68×10−7 .63×10−8 108 .17×10−5
1.0 512 0 — 50 .40×10−8
.9 512 44 .53×10−7 .82×10−8 116 .26×10−6
.9 1024 0 — 89 .98×10−10
.8 1024 80 .35×10−8 .33×10−9 155 .42×10−7
.8 2048 0 — 81 .55×10−11
.75 2048 112 .41×10−9 .37×10−10 143 .58×10−9
.72 2048 108 .40×10−8 .35×10−9 151 .40×10−8
.7 2048 108 .95×10−8 .81×10−9 120 .30×10−7
.7 4096 0 — 81 .31×10−11
.68 4096 148 .12×10−9 .97×10−11 151 .38×10−10
.66 4096 148 .53×10−9 .42×10−10 151 .13×10−9
.64 4096 136 .92×10−8 .69×10−9 159 .29×10−8
.62 4096 136 .33×10−7 .23×10−8 140 .19×10−7
.6 4096 144 .48×10−7 .33×10−8 148 .84×10−7
.6 8192 0 — 93 .23×10−10
.58 8192 168 .19×10−7 .12×10−8 167 .78×10−9
.56 8192 188 .14×10−7 .83×10−9 171 .90×10−9
.54 8192 180 .11×10−6 .63×10−8 171 .22×10−7
.54 16384 0 — 77 .28×10−10
.52 16384 180 .46×10−6 .23×10−7 178 .99×10−9
.5 16384 212 .25×10−6 .11×10−7 228 .15×10−7

Figure 26. Table of comparative behaviour of Fornberg’s method


and the Interpolating Polynomial method
Techniques for Numerical Conformal Mapping 231

1 1

0.5 0.5

-0.5 0.5 1 1.5 -0.5 0.5 1 1.5

-0.5 -0.5

-1 -1

True values of w
~ ~∗
Initial test values w
Figure 27.

with a variable parameter α. The high crowding factor for the domain D0.5
mentioned earlier is largely due to the fact that the base point z0 = 0 is near
the left-hand side of the domain. If one starts with evenly spaced points along
the boundary, then probably none of the “fast” methods will converge to the
solution. In fact, one has to be extremely close to the solution for convergence
to be possible. One way to get around this situation is to solve first the mapping
problem for D2.0 , which is nearly circular, and then project the solution points
to say, D1.5 , solve the problem there, and then project to another nearby Dα .
This was done in the same way for the Interpolating Polynomial method in [18],
and the fairly similar results make one wonder whether the two methods in some
sense may be based on essentially the same fundamental ideas.
Now we look at a simpler domain, a unit disk which is centered at α, 0 < α < 1.
The mapping function is
α−z
(16) f (z) = α − .
1 − αz

Figure 28. P (∂ D) for initial test values of Figure 27


232 R. Michael Porter IWQCMA05

Figure 29. Left: P (∂ D) after one application of the Interpolat-


ing Polynomial algorithm. Center: Result after second application.
Right: Imperfect view of domain due to discretization

For illustration we take α = 0.5 and N = 32. Supposing that we don’t know
the formula (16), we will naı̈vely guess that the wj are equally spaced; i.e., we
take wj = ζj as on the right of Figure 27. The resulting P (∂ D) turns out to be
the rather complicated curve shown in Figure 28, a very bad approximation of
the circumference of the disk.
After one iteration of Φ, a most of the wj have moved over to the left closer to
where they belong, and the image curve looks a bit more like ∂ D. At the second
iteration they cannot be distinguished visually from the true positions, and the
half-click images uj appear to lie exactly on ∂ D as well.
Of course, this is only an approximation, and to study its accuracy one may
graph the change in w∗~ from one iteration to the next. For this any convenient
norm will do; in Figure 30 we use log || ||∞ .
In Figure 31 we show the results for 0, 1, and 2 iterations of the method
for an ellipse and a square, starting in each case with equally spaced boundary
points. Note that a polynomial of degree 32 does not give a particularly good
approximation of the Riemann mapping for a square.

-0.5

-1

-1.5

-2

-2.5

-3

Figure 30. Logarithmic graph of amount w


~ is moved in each iteration
Techniques for Numerical Conformal Mapping 233

Figure 31. Interpolating polynomial algorithm applied to ellip-


tical and square domains.

11. The Quest for Better Methods


Once we have the basic idea of the solution of the mapping problem being a
fixed point for Φ, we can look for ways to calculate it more rapidly. Here we will
describe a few attempts.

11.1. Methods using derivatives. Given w ~ on ∂D, we can write the interpo-
lating polynomial P = Pw~ explicitly, and thus can calculate its derivative, giving
an N -vector
2πi
(17) wj′ = Pw~′ (ej N ) = (C ′ w)
~ j
for an appropriate matrix C ′ .
On the other hand, there are many other formulas involving f ′ . For example,
from (2) we have ieiθ f ′ (eiθ ) = γ ′ (b(θ))b′ (θ) so we can write

(18) ~b′ = i ζ~ 1 w ~ ′.
~γ ′

Many ideas present themselves for combining (17) and (18). For example, given
~ and ~γ ′ , and then can obtain ~b′ from which a new value of ~b can
~b one calculates w
be estimated. Alternatively, from an initial ~b written as a deviation ~b = ~b∗ + ∆~b
from the true solution ~b∗ , to obtain an equation for ∆w. ~ However, so far I have
not been able to create an algorithm which converges by using any such idea.

11.2. Method of simultaneous interpolation. Let w ~ ∗ ⊆ (∂D)N denote the


N
true solution of the mapping problem. Let w~ ⊆ (∂D) be an approximation. We
would like to devise an algorithm which moves w~ closer to w~ ∗ ; or in other words,
to calculate the difference ∆w~ approximately, and then add it to w ~ ∗.
~ to find w
N
Using the parameters ~t = (tj ) ∈ R , we see that within a neighborhood of each
234 R. Michael Porter IWQCMA05

γ(tj ) = wj the curve γ = ∂D can be approximated by a series, which we write


symbolically as
~γ ′′
(19) ~ + ~γ ′ ∆~t + ∆~t 2 + . . .
γ(~t + ∆~t) = w
2
In this formula we use coordinate-wise multiplication in a natural way, which
means a subscript “j” can be applied to all the letters. In particular, we want
to find the value of ∆~t for which (19) is equal to w ~ ∗.
To find it, write ~u = C w.
~ Then near ρ~u = γ(~s), ∂D is approximated by a
similar series
~τ ′′ ~ 2
(20) γ(~s + ∆~s) = w~ + ~γ ′ ∆ρ~u + ~τ ′ ∆~s + ∆s + ....
2
By the same token, there should be a value of ∆~s for which (20) is equal to
~u∗ = ρ~u∗ .
First we will examine the linear approximations obtained by truncating (19),
(20):
~∗ = w
w ~ + ~γ ′ ∆~t , ~u∗ = ρ~u + ~τ ′ ∆~s
which are connected by the relation
C(w ~ + ~γ ′ ∆~t ) = ρ~u + ~τ ′ ∆~s.
One solves this to find that
C(~γ ′ ∆~t )j ρuj − uj
(21) Im ′
= Im
τj τj′
since ∆sj ∈ R. The real-linear operator
C(~γ ′ ∆~t)
∆~t 7→ Im
τ′
from Rn to Rn can be seen to have a null vector d~ close to ~1 = (1, 1, . . . ). It
can be found by standard conjugate gradient methods, and ∆~t = d~ gives an
approximate solution of the mapping problem since ρ~u = ~u by (21).
Now we will use a quadratic approximation. The relation is
~γ ′′ 2 ~τ ′′
C( w~ + ~γ ′ ∆~t + ∆~t ) = ρ~u + ~τ ′ ∆~s + ∆~s 2 .
2 2
From what we already know concerning the linear approximation, we may
cancel several terms when this is expanded. We find thus a quadratic relation
between ∆~t and ∆~s. The linear system already gave an approximation for ∆~t,
which we substitute to solve for ∆~s.
The linear and quadratic versions of “simultaneous interpolation” method give
fairly good convergence, as shown in Figure 32. There are two disappointing
facts. One is that we have obtained slightly better convergence at a much higher
calculation cost. The other is that, according to these experiments, the quadratic
method seems no better than the linear one, in spite of costing a good deal more
work for each iteration.
Techniques for Numerical Conformal Mapping 235

-2.5

-5

-7.5

-10

-12.5

-15

Figure 32.

11.3. Minimization approach. We are looking for u∗j ∈ ∂D, so we try to


minimize the quantity
~ 22
||(I − ρ)C w||
~ ∈ (∂D)N . Here || · ||2 refers to the L2 norm. We calculate the Jacobian of
for w
the projection map at v near u. Because of the approximate relation
t − ρu
ρv = ρu + (u − ρu) i Im ,
u − ρu
we find that the Jacobian mapping is
 
1 (Re a)2 − Re a Im a
Jρ (v) =
|a| − Re a Im a (Im a)2
where a = u−ρu. Then the gradient of the real-valued function can be calculated
to be
||ϕ(~t)||22 = ||(I − ρ)C(w~ ∗ + ~γ ′~t)||22
which is given by
∇||ϕ||22 = 2JϕT ϕ
= 2 J~γT′ C t JI−ρ
T
ϕ.
Once one has the gradient, one can use it to find a value of ∆~t which minimizes the
function. Some experimentation has shown that this approach works, although
so far not very well.

12. Combined Methods


Recall that the “easy” methods are of general application but converge slowly,
and the “fast” methods only apply when one already has a good idea of where
the solution is. Thus in practice it is only logical to combine the two approaches.
Given a domain D, the first step is to map it to the interior of D, and then
apply an osculation method to obtain an image domain which is nearly circular.
Then one of the faster methods can be applied to this image domain.
Figure 33, taken from [19], explains this procedure, which can be automated
reasonably well. The initial domain is defined by several thousand points. The
236 R. Michael Porter IWQCMA05

(a) (b) ( )
i

Figure 33. Combined method. (a) Original domain. (b)


Nearly circular domain obtained by two iterations of an osculation
method. (c) After applying the Interpolation Polynomial method
to (b), the inverse of the osculation result is applied. Note the
effect of the crowding phenomenon and discretization.

image under the osculation mapping has the same number M of points (here
around 8000). The “fast” method is applied with a relatively small number N
of points (here 512). This is done because the cost O(N log N ) increases fairly
rapidly with N . Thus when the inverse of the osculation mapping is applied, we
only have N points to describe the domain. When the crowding phenomenon
is present, this may cause part of the figure to be badly represented. On the
other hand, if one is only interested in approximating the conformal mapping
near another part of the boundary (or in the interior), this may be a very useful
aspect of the method.

13. Epilogue
We suggest that the reader interested in knowing more about numerical con-
formal mapping consult the following.
The books [23] and [14] give detailed explanations of a great number of map-
pings with specific formulas. We mention also [3], a much older book, which
gives a general introduction to the theory of functions of a complex variable as
necessary to understand the topic of conformal mapping, as do [13], [16].
Reference [9], a half-century old text in German, is divided into two parts,
covering precisely what we have called the “easy” and “fast” methods for confor-
mal mapping. Bear in mind that the numerical examples were calculated without
computers!
Reference [12] gives a much more modern and very practical treatment, in-
cluding some of the methods we have described here (Koebe and Grassmann
osculation, Theodorsen’s method). In [15] one may find a great variety of other
conformal mapping methods. As to detailed treatiseson specific methods, we rec-
ommend the book [6] on the Schwarz-Christoffel method, and [20] which explains
the method of circle packings.
Techniques for Numerical Conformal Mapping 237

Finally we recommend the survey article [4], which gives a more recent per-
spective of the existing methods.

References
1. W. Abikoff, The uniformization theorem. Amer. Math. Monthly 88 (1981), 574–592.
2. L. Ahlfors, Complex Analysis: An introduction to the theory of analytic functions of one
complex variable, Third edition, International Series in Pure and Applied Mathematics,
McGraw-Hill Book Co., New York 1978.
3. L. Bieberbach, Conformal mapping, Chelsea, New York 1964.
4. T. K. DeLillo, The accuracy of numerical conformal mapping methods: a survey of exam-
ples and results, SIAM J. Numer. Anal. 31 (1994) 788–812.
5. John B. Conway, Functions of one complex variable, Second edition, Graduate Texts in
Mathematics 11, Springer-Verlag, New York-Berlin 1978.
6. T. A. Driscoll and L. N. Trefethen, Schwarz-Christoffel mapping, Cambridge Monographs
on Applied and Computational Mathematics, Cambridge University Press, Cambridge
2002.
7. T. A. Driscoll and S. A. Vavasis, Numerical conformal mapping using cross-ratios and
Delaunay triangulation, SIAM J. Sci. Comput. 19 (1998) 1783-1803.
8. B. A. Fornberg, A numerical method for conformal mapping of doubly connected regions,
SIAM J. Sci. Statist. Comput. 5 (1984) 771–783.
9. D. Gaier, Konstruktive Methoden der konformen Abbildung, Springer tracts in natural
philosophy, v. 3, Springer, Berlin 1964.
10. E. Grassmann, Numerical experiments with a method of successive approximation for
conformal mapping. Z. Angew. Math. Phys. 30 (1979) 873–884.
11. J. Heinhold, R. Albrecht, Zur Praxis der konformen Abbildung, Rend. Circ. Mat. Palermo
3 (1954) 130–148.
12. P. Henrici, Applied and computational complex analysis, Vol. 3, Pure and Applied Mathe-
matics (New York). A Wiley-Interscience Publication. John Wiley & Sons, Inc., New York
1986.
13. E. Hille, Analytic function theory, Chelsea Publishing Company, New York 1959.
14. H. Kober, Dictionary of conformal representations, Dover, New York 1957.
15. P. K. Kythe, Computational conformal mapping, Boston: Birkhüser, Boston 1998.
16. Z. Nehari, Conformal mapping, McGraw-Hill, New York 1952.
17. S. T. O’Donnell and V. Rokhlin, A fast algorithm for the numerical evaluation of conformal
mappings, SIAM J. Sci. Statist. Comput. 10 (1989) 475–487.
18. R. M. Porter, An interpolating polynomial method for numerical conformal mapping. SIAM
J. Sci. Comput. 23 (2001) 1027–1041.
19. R. M. Porter, An accelerated osculation method and its application to numerical conformal
mapping, Complex Var. Theory Appl., 48 (2003) 569–582.
20. K. Stephenson, Introduction to circle packing: The theory of discrete analytic functions,
Cambridge University Press, Cambridge, 2005.
21. L. N. Trefethen and T. A. Driscoll, A. Schwarz-Christoffel mapping in the computer era.
Proceedings of the International Congress of Mathematicians, Vol. III (Berlin, 1998), Doc.
Math. 1998, Extra Vol. III, 533–542.
22. M. R. Trummer, An efficient implementation of a conformal mapping method based on
the szegö kernel, SIAM J. Numer. Anal. 23 (1986) 853–872.
23. W. von Koppenfels, Praxis der konformen Abbildung, Springer-Verlag, Berlin-Göttingen-
Heidelberg 1959.
24. R. Wegmann, Discrete Riemann-Hilbert problems, interpolation of simply closed curves,
and numerical conformal mapping, J. Comput. Appl. Math. 23 (1988) 323–352.
238 R. Michael Porter IWQCMA05

25. R. Wegmann, Conformal mapping by the method of alternating projections, Numer. Math.
56 (1989) 291–307.

R. Michael Porter E-mail: mike@math.cinvestav.mx


Address: Department of Mathematics, Centro de Investigación y de Estudios Avanzados del
I.P.N., Apdo. Postal 14-740, 07000 México, D.F., Mexico
Proceedings of the International Workshop on Quasiconformal
Mappings and their Applications (IWQCMA05)

Introduction to quasiconformal mappings in n-space

Antti Rasila

Abstract. We give an introduction to quasiconformal mappings in the Eu-


clidean space Rn .

Keywords. quasiconformal mappings.

2000 MSC. 30C65.

Contents

1. Introduction: Mercator’s map 240


2. History and background 241
3. Preliminaries 242
ACLp functions 242
Conformal mappings 242
Möbius transformations 242
4. Modulus of a path family 244
Ring domains 248
Modulus in conformal mappings 250
Capacity of a condenser 251
Sets of zero capacity 252
Spherical symmetrizations 252
Canonical ring domains 253
Spherical metric 256
5. Quasiconformal mappings 256
Examples 257
6. An application of the modulus technique 258
References 259

Version October 19, 2006.


.
240 A. Rasila IWQCMA05

1. Introduction: Mercator’s map


Perhaps the greatest cartographer of the time, Gerardus Mercator (5 March
1512 – 2 Dec 1594) was born Gerhard Kremer of German parents in the town
of Rupelmonde near Antwerp. Like many other intellectuals of his time, he
Latinized his German name, which meant “merchant”, and changed it to the
name Mercator which means “world trader”. Mercator was a mapmaker, scholar,
and religious thinker. His interests ranged from mathematics to calligraphy and
the origin of the universe. Mercator studied mathematics in Louvain under the
supervision of mathematician and astronomer Gemma Frisius.

Figure 1: Gerardus Mercator (source: Wikipedia) and a World map using the
Mercator projection.

The Mercator map is defined by the formula



(x, y) = λ, log tan(π/4 + φ/2) ,
where φ is the latitude and λ is the longitude of the point on the sphere. Mercator
published the first map using this projection in 1569, a wall map of the world
on 18 separate sheets entitled: “New and more complete representation of the
terrestrial globe properly adapted for its use in navigation.” The projection did
not become popular until 30 years later (1599), when Edward Wright published
an explanation of it. An important property of the Mercator projection is that
it is conformal, i.e. the angles are preserved.

Figure 2: India and Finland in the Mercator projection.


Introduction to quasiconformal mappings in n-space 241

The Mercator projection is not without flaws, however. For example, from
the picture above one might conclude that India is approximately twice as large
as Finland. Actually, India’s land area is 3, 287, 590 km2 , almost ten times that
of Finland (338, 145 km2 ). This example also illustrates the reasons why we are
mainly interested in the local distortion of the geometry in this theory.

2. History and background


Conformal mappings play extremely important role in complex analysis, as
well as in many areas of physics and engineering. The class of conformal map-
pings turned out to be too restrictive for some problems. Quasiconformal map-
pings were introduced by H. Grötzsch provide more flexibility in 1928. Important
results were also obtained by O. Teichmüller and L. V. Ahlfors [1]. A compre-
hensive survey on quasiconformal mappings of the complex plane is [16]. See
also [15].
By the Riemann mapping theorem a simply-connected plane domain with
more than one boundary point can be mapped conformally onto the unit disk B2 .
On the other hand, Liouville’s theorem says that the only conformal mappings in
Rn , n ≥ 3, are the Möbius transformations. Hence the plane theory of conformal
mappings does not directly generalize to the higher dimensions.
Quasiconformal maps were first introduced in higher dimensions by M. A. Lav-
rent’ev in 1938. The systematic study of quasiconformal maps in Rn was begun
by F. W. Gehring [5] and J. Väisälä [20] in 1961. Since then the theory and
it’s generalizations have been actively studied [3, 4, 21, 23]. Generalizations
include quasiregular [18, 22, 19] and quasisymmetric mappings, and recently the
mappings of finite distortion [13] and the quasiconformal mappings in the metric
spaces [10, 11, 12].
Quasiconformal mappings in Rn are natural generalization of conformal func-
tions of one complex variable. Quasiconformal mappings are characterized by
the property that there exists a constant C ≥ 1 such that the infinitesimally
small spheres are mapped onto infinitesimally small ellipsoids with the ratio of
the larger “semiaxis” to the smaller one bounded from above by C.

l
L

Figure 3: Image of a small sphere.


242 A. Rasila IWQCMA05

For a comprehensive historical review of the theory of quasiconformal map-


pings in both plane and space settings, see [2]. A survey of the theory of qua-
siconformal mappings is given in [8] (see also [14]). This presentation is for the
most parts based on [7], [21] and [22].

3. Preliminaries
We shall follow standard notation and terminology adopted from [21], [22]
and [19]. For x ∈ Rn , n ≥ 2, and r > 0 let Bn (x, r) = {z ∈ Rn : |z − x| < r},
S n−1 (x, r) = ∂Bn (x, r), Bn (r) = Bn (0, r), S n−1 (r) = ∂Bn (r), Bn = Bn (1),
Hn = {x ∈ Rn : xn > 0}, Bn+ = Bn ∩ Hn , and S n−1 = ∂Bn . For t ∈ R
and a ∈ Rn \ {0}, P (a, t) = {x ∈ Rn : x · a = t} ∪ {∞}, is a hyperplane in
n
R = Rn ∪ {∞} perpendicular to the vector a and at distance t/|a| from the
origin. The surface area of S n−1 is denoted by ωn−1 and Ωn is the volume of Bn .
It is well known that ωn−1 = nΩn and that
π n/2
Ωn =
Γ(1 + n/2)
for n = 2, 3, . . ., where Γ is Euler’s gamma function. The standard coordinate
unit vectors are denoted by e1 , . . . , en . The k-dimensional Lebesgue measure
is denoted by mk . For k = n we omit the subscript and denote the Lebesgue
measure on Rn simply by m.
n
For nonempty subsets A and B of R , we let d(A) = sup{|x − y| : x, y ∈ A}
be the diameter of A, d(A, B) = inf{|x − y| : x ∈ A, y ∈ B} the distance between
the sets A and B, and in particular d(x, B) = d({x}, B).

ACLp functions. Let Q be a closed n-interval {x ∈ Rn : ai ≤ xi ≤ bi , i =


1, . . . , n}. A function f : Q → Rm is called ACL (absolutely continuous on lines)
if f is continuous and if f is absolutely continuous on almost every line segment in
Q parallel to one of the coordinate axes. Let U be an open set in Rn . A function
f : U → Rm is ACL if f |Q is ACL for every closed n-interval Q ⊂ U . Such a
function has partial derivatives Di f (x) a.e. in U , and they are Borel functions
[21, 26.4]. If p ≥ 1 and the partial derivatives of f are locally Lp -integrable, f is
said to be in ACLp or in ACLp (U ).

Conformal mappings. Let G, G′ be domains in Rn . A homeomorphism f : G →


G′ is called conformal if f is in C 1 (G), Jf (x) 6= 0 for all x ∈ G, and |f ′ (x)h| =
n
|f ′ (x)||h| for all x ∈ G and h ∈ Rn . If G, G′ are domains in R , a homeomorphism
f : G → G′ is conformal if its restriction to G \ {∞, f −1 (∞)} is conformal.
n n
Möbius transformations. A Möbius transformation is a mapping f : R → R
that is composed of a finite number of the following elementary transformations:
(1) Translation: f1 (x) = x + a.
(2) Stretching: f2 (x) = rx, r > 0.
(3) Rotation: f3 is linear and |f3 (x)| = |x| for all x ∈ Rn .
Introduction to quasiconformal mappings in n-space 243

(4) Reflection in plane P (a, t):


a
f4 (x) = x − 2(x · a − t) , f4 (∞) = ∞.
|a|2
(5) Inversion in a sphere S n−1 (a, r):
r2 (x − a)
f5 (x) = a + , f5 (a) = ∞, f5 (∞) = a.
|x − a|2
In fact every Möbius transformation can be expressed as a composition of a finite
number of reflections and inversions. It is easy to see that every elementary
transformation, and hence every Möbius transformation, is conformal.
Let a, b, c, d be distinct points in Rn . We define the absolute (cross) ratio by
|a − c| |b − d|
(3.1) |a, b, c, d| = .
|a − b| |c − d|
n
This definition can be extended for a, b, c, d ∈ R by taking limit.
An important property of Möbius transformations is that they preserve the
absolute ratios, i.e.
|f (a), f (b), f (c), f (d)| = |a, b, c, d|,
n n n n
if f : R → R is a Möbius transformation. In fact, a mapping f : R → R is a
Möbius transformation if and only if f preserves all absolute ratios.
Let a∗ = a/|a|2 for a ∈ Rn \ {0}, 0∗ = ∞ and ∞∗ = 0. Fix a ∈ Bn \ {0}. Let
σa (x) = a∗ + r2 (x − a∗ )∗ , r2 = |a|2 − 1
be an inversion in the sphere S n−1 (a∗ , r) orthogonal to S n−1 . Then σa (a) = 0,
σa (a∗ ) = ∞. Let pa denote the reflection in the (n − 1)-dimensional plane P (a, 0)
through the origin and orthogonal to a, and define a sense preserving Möbius
transformation by Ta = pa ◦ σa . Then Ta (Bn ) = Bn and Ta (a) = 0. For a = 0
we set Ta = id, i.e. the identity map.

n−1
S n−1
1 r S (a *,r)

0 a a*

Figure 4: Construction of the Möbius transformation Ta .


244 A. Rasila IWQCMA05

4. Modulus of a path family


A path in Rn is a continuous mapping γ : ∆ → Rn , where ∆ is a (possibly
unbounded) interval in R. The path γ is called closed or open according as ∆ is
compact or open. The locus |γ| of γ is the image set γ∆.
Let γ : [a, b] → Rn be a closed path. The length ℓ(γ) of the path γ is defined
by means of polygonal approximation (see [21], pages 1-8). The path γ is called
rectifiable if ℓ(γ) < ∞ and locally rectifiable if each closed subpath of γ is
rectifiable. If γ is a rectifiable path, then γ has a parameterization by means of arc
length, also called the normal representation of γ. The normal representation of
γ is denoted by γ 0 : [0, ℓ(γ)] → Rn . By making use of the normal representation,
one may define the integral over a locally rectifiable path γ.
Definition 4.1. Let Γ be a path family in Rn , n ≥ 2. Let F(Γ) be the set of all
Borel functions ρ : Rn → [0, ∞] such that
Z
ρ ds ≥ 1
γ

for every locally rectifiable path γ ∈ Γ. The functions in F(Γ) are called admis-
sible for Γ. For 1 < p < ∞ we define
Z
(4.2) Mp (Γ) = inf ρp dm
ρ∈F (Γ) Rn
and call Mp (Γ) the p-modulus of Γ. If F(Γ) = ∅, which is true only if Γ contains
constant paths, we set Mp (Γ) = ∞. The n-modulus or conformal modulus is
denoted by M(Γ).
Lemma 4.3. [21, 6.2] The p-modulus is an outer measure in the space of all path
families in Rn . That is,
(1) Mp (∅) = 0,
(2) If Γ1 ⊂ Γ2 then Mp (Γ1 ) ≤ Mp (Γ2 ), and
S  P 
(3) Mp j Γj ≤ j Mp Γj .
Proof. (1) Since the zero function is admissible for ∅, Mp (∅) = 0.
(2) If Γ1 ⊂ Γ2 then F(Γ2 ) ⊂ F(Γ1 ) and hence Mp (Γ1 ) ≤ Mp (Γ2 ).
(3) We may assume that Mp (Γj ) < ∞ for all j. Let ε > 0. Then we can
choose for each j a function ρj admissible for Γj such that
Z
ρpj dm ≤ Mp (Γj ) + 2−j ε.
Rn
Now let [
ρ = sup ρj , Γ= Γj .
j
j
Then ρ : Rn → [0, ∞] is a Borel function. Moreover, if γ ∈ Γ is locally rectifiable,
then γ ∈ Γj for some j, Z Z
ρ ds ≥ ρj ds ≥ 1,
γ γ
Introduction to quasiconformal mappings in n-space 245

and hence ρ is admissible for Γ. Now


Z Z X X
Mp (Γ) ≤ p
ρ dm ≤ ρpj dm ≤ Mp (Γj ) + ε.
Rn Rn j j

By letting ε → 0, the claim follows.


Let Γ1 and Γ2 be path families in Rn . We say that Γ2 is minorized by Γ1 and
write Γ1 < Γ2 if every γ ∈ Γ2 has a subpath in Γ1 .
Lemma 4.4. If Γ1 < Γ2 then Mp (Γ1 ) ≥ Mp (Γ2 ).

Proof. If Γ1 < Γ2 then obviously F(Γ1 ) ⊂ F(Γ2 ). Hence Mp (Γ1 ) ≥ Mp (Γ2 ).


Lemma 4.5. Let G be a Borel set in Rn , r > 0 and let Γ be the family of paths
in G such that ℓ(γ) ≥ r. Then Mp (Γ) ≤ m(G)r−p .

Proof. The claim follows immediately from (4.2) and the fact that the function
ρ = χG /r is admissible for Γ.
Lemma 4.6. Path family Γ has zero p-modulus if and only if there is an admis-
sible function ρ ∈ F(Γ) such that
Z Z
p
ρ dm < ∞ and ρ ds = ∞
Rn γ

for every locally rectifiable path γ ∈ Γ.

Proof. If ρ satisfies the above conditions, clearly ρ/k is admissible for Γ for all
k = 1, 2, . . . . Hence Z
Mp (Γ) ≤ k −p
ρp dm → 0
Rn
as k → ∞, and thus Mp (Γ) = 0.
Now let Mp (Γ) = 0 and choose a sequence of functions ρk ∈ F(Γ) such that
Z
ρpk dm < 4−k , k = 1, 2, . . . .
Rn
Define
X
∞ 1/p
ρ(x) = 2k ρpk (x) ,
k=1
and note that Z
ρp dm < ∞.
Rn
On the other hand,
Z Z
ρ ds ≥ 2k/p ρk ds ≥ 2k/p → ∞
γ γ

as k → ∞ for every locally rectifiable path γ ∈ Γ.


n
Corollary 4.7. Let Γ be a path family in R and denote by Γr the family of all
rectifiable paths in Γ. Then M(Γ) = M(Γr ).
246 A. Rasila IWQCMA05

The path families Γ1 , Γ2 , . . . are called separate if there exist disjoint Borel
sets Ei such that
Z
(4.8) χRn \Ei ds = 0
γ
for all locally rectifiable γ ∈ Γi , i = 1, 2, . . ..
Lemma 4.9. [19, Proposition II.1.5] Let Γ, Γ1 , Γ2 , . . . be a sequence of path fam-
ilies in Rn . Then
(1) If Γ1 , Γ2 , . . . are separate and Γ < Γj for all j = 1, 2, . . . , then
X
Mp (Γ) ≥ Mp (Γj ).
j
S
Equality holds if Γ = j Γj .
(2) If Γ1 , Γ2 , . . . are separate and Γj < Γ for all j = 1, 2, . . . , then
X
Mp (Γ)1/(1−p) ≥ Mp (Γj )1/(1−p) , p > 1.
j

Proof. (1) Let ρ be admissible for Γ, and let Ej be as in (4.8). Then for all
indices j the function ρj = χEj ρ is admissible for Γj . It follows that
X XZ p
XZ Z
p
Mp (Γj ) ≤ ρj dm = ρ dm ≤ ρp dm.
p j Rn j Ej Rn
S
(2) Let Ej be as in (4.8), and let E = j Ej . Then for all indices j the
P χEj ρ is admissible for Γj . Let (aj ) be a sequence such that aj ∈ [0, 1]
function
and j aj = 1. Let
X∞
ρ= a j χ Ej ρ j .
j=1
Next we show that ρ is admissible for Γ. Fix a locally rectifiable path γ ∈ Γ and
a subpath γj ∈ Γj for each j = 1, 2, . . . . Now
Z Z X  X Z
ρ ds = aj χEj ρj ds = aj χEj ρj ds
γ γ j j γ
X Z X
≥ aj χEj ρj ds ≥ aj = 1.
j γj j

Hence ρ is admissible for Γ and


Z Z
p
Mp (Γ) ≤ ρ dm = ρp dm
Rn E
XZ X p XZ
= ak χEk ρ dm = apj ρpj dm
j Ej k j Ej
Z X X Z
≤ apj ρpj dm ≤ apj ρpj dm.
Rn j j Rn
Introduction to quasiconformal mappings in n-space 247

By taking the infimum over all admissible ρj , we obtain


X p
(4.10) Mp (Γ) ≤ aj Mp (Γj ).
j

We may assume that Mp (Γ) > 0 (if that would not be the case, the left side of
the inequality is ∞ and there is nothing to prove). Hence by Lemma 4.4 we have
Mp (Γj ) ≥ Mp (Γ) > 0. Similarly, we may assume that Mp (Γj ) < ∞.
Let
1
tk = Pk , aj,k = Mp (Γj )1/(1−p) tk ,
j=1 Mp (Γj )1/(1−p)
Pk
for j = 1, . . . , k and k = 1, 2, . . . . Now j=1 aj,k = 1. We choose aj,k = 0 for
j ≥ k + 1, and by (4.10) we have
k
X X
k 1−p
Mp (Γ) ≤ tpk Mp (Γj )p/(1−p) Mp (Γj ) = Mp (Γj )1/(1−p) .
j=1 j=1

By letting k → ∞ the claim follows.


For E, F, G ⊂ Rn we denote by ∆(E, F ; G) the family of all nonconstant paths
joining E and F in G.
Lemma 4.11. [22, 5.22] Let p > 1 and let E, F be subsets of Hn . Then
1
Mp (∆(E, F ; Hn )) ≥ Mp (∆(E, F )).
2
11111111111111111111111111
00000000000000000000000000
00000000000000000000000000
11111111111111111111111111
00000000000000000000000000
11111111111111111111111111
00000000000000000000000000
11111111111111111111111111
00000000000000000000000000
11111111111111111111111111
F11111111111111111111111111
00000000000000000000000000
00000000000000000000000000
11111111111111111111111111
00000000000000000000000000
11111111111111111111111111
00000000000000000000000000
11111111111111111111111111
00000000000000000000000000
11111111111111111111111111
00000000000000000000000000
11111111111111111111111111
h
G
11111111111111111111111111
00000000000000000000000000
00000000000000000000000000
11111111111111111111111111
00000000000000000000000000
11111111111111111111111111
00000000000000000000000000
11111111111111111111111111
00000000000000000000000000
11111111111111111111111111
00000000000000000000000000
11111111111111111111111111
E11111111111111111111111111
00000000000000000000000000
00000000000000000000000000
11111111111111111111111111
00000000000000000000000000
11111111111111111111111111
00000000000000000000000000
11111111111111111111111111
00000000000000000000000000
11111111111111111111111111

Figure 5: Cylinder with bases E and F .


Example 4.12. Let E ⊂ {x ∈ Rn : xn = 0} be a Borel set, h > 0, F = E + hen .
We define a cylinder G with bases E, F by
G = {x ∈ Rn : (x1 , . . . , xn−1 , 0) ∈ E, 0 < xn < h}.
Then Mp (∆(E, F ; G)) = mn−1 (E)h1−p = m(G)h−p .
248 A. Rasila IWQCMA05

Proof. Choose ρ ∈ F(Γ) where Γ = ∆(E, F ; G)) and let γy be the vertical
segment from y ∈ E. Then γy ∈ Γ. We note that 1/p + (p − 1)/p = 1, and hence
by Hölder’s inequality
Z p  Z p−1  Z  Z
p p−1
1≤ ρ ds ≤ 1 ds ρ ds = h ρp ds.
γy γy γy γy

This holds for all y ∈ E and hence by the Fubini theorem


Z Z Z 
p p mn−1 (E)
ρ dm ≥ ρ ds dmn−1 ≥ .
Rn E γy hp−1
Since the above holds for any ρ ∈ F(Γ),
mn−1 (E)
Mp (Γ) ≥ .
hp−1
Next we choose ρ = 1/h inside G and ρ = 0 otherwise. Then ρ is admissible for
Γ and Z
mn−1 (E)
Mp (Γ) ≤ ρp dm = .
Rn hp−1

Remark 4.13. In Example 4.12 the modulus is invariant under similarity map-
pings if and only if p = n. This is the reason why the case p = n is so important
in the theory of quasiconformal mappings. Later in this section we will show
that M(Γ) is a conformal invariant.
n n
Ring domains. A domain G in R is called a ring, if R \ G has exactly two
components. If the components are E and F , we denote the ring by R(E, F ).
In general, it is difficult to calculate the modulus of a given path family. Next
two lemmas give us an important tool, letting us to obtain effective upper and
lower bounds for the modulus in many situations.

Figure 6: Spherical ring with 0 < a < b < ∞.


Introduction to quasiconformal mappings in n-space 249

n
Lemma 4.14. [21, 7.5] Let 0 < a < b < ∞, A = Bn (b) \ B (a) and

ΓA = ∆ S n−1 (a), S n−1 (b); A .
Then  b 1−n
M(ΓA ) = ωn−1 log .
a
Proof. Let ρ ∈ F(ΓA ). For each unit vector y ∈ S n−1 let γy : [a, b] → Rn the
radial line segment defined by γy (s) = sy. As in Example 4.12 by Hölder’s
inequality we obtain
Z n  Z b  Z b n−1
n n−1 1
1 ≤ ρ ds ≤ ρ(sy) s ds ds
γy a a s
 Z
b n−1 b
= log ρ(sy)n sn−1 ds.
a a
By integrating over y ∈ S n−1 , we have
 Z
b n−1
(4.15) ωn−1 ≤ log ρn dm.
a R n

Taking the infimum over all admissible ρ yields


 b n−1
ωn−1 ≤ log M(ΓA ).
a

Next we define ρ(x) = 1/ |x| log(b/a) for x ∈ A, and ρ(x) = 0 otherwise.
Clearly ρ is admissible for ΓA , and hence
Z  Z
n b −n b 1  b 1−n
M(ΓA ) ≤ ρ dm = ωn−1 log ds = ωn−1 log .
Rn a a s a

n
Lemma 4.16. [21, 7.8] Let x0 ∈ R and let Γ be the family of all nonconstant
paths through x0 . Then M(Γ) = 0.

Proof. If x0 = ∞, the claim follows immediately from Corollary 4.7.


If x0 6= ∞, we let
Γk = {γ ∈ Γ : |γ| ∩ S n−1 (x0 , 1/k) 6= ∅}.
We may assume that x0 = 0. Then for all R > 1/k
n 
Γk > ∆R , where ∆R = ∆ S n−1 (1/k), S n−1 (R); Bn (R) \ B (1/k) ,
and by Lemma 4.4 and Lemma 4.14 we have
 1−n
R
M(Γk ) ≤ M(∆R ) = ωn−1 log →0
1/k
S
as R → ∞, and thus M(Γk ) = 0. On the other hand, because Γ = k Γk we have
by Lemma 4.3 (3) X
M(Γ) ≤ M(Γk ) = 0.
k
250 A. Rasila IWQCMA05

n n
Modulus in conformal mappings. Let G ⊂ R and f : G → R be a continu-
ous function. Suppose that Γ is a family of paths in G. Then Γ′ = {f ◦γ : γ ∈ Γ}
is a family of paths in f (G). Γ′ is called the image of Γ under f .
Theorem 4.17. [21, 8.1] If f : G → f (G) is conformal, then M(f (Γ)) = M(Γ)
for all path families Γ in G.

Proof. By Lemma 4.16 we may assume that the paths of Γ, f (Γ) do not go
through ∞. Let ρ1 ∈ F(f (Γ)), and define

ρ(x) = ρ1 f (x) |f ′ (x)|
for x ∈ G and ρ(x) = 0 otherwise. Because f is a conformal mapping (see [21,
5.6]), Z Z Z
 ′
ρ ds = ρ1 f (x) |f (x)| |dx| = ρ1 ds ≥ 1
γ γ f ◦γ
for every locally rectifiable γ ∈ Γ. It follows that ρ ∈ F(Γ), and
Z Z Z Z
n n
 n
M(Γ) ≤ ρ dm = ρ1 f (x) |Jf (x)| dm = ρ1 dm = ρn1 dm
Rn G f (G) Rn

for all ρ1 ∈ F(f (Γ)), and thus M(Γ) ≤ M(f (Γ)). The inverse inequality follows
from the fact that f −1 is conformal.
Lemma 4.18. Let A ⊂ Hn , B ⊂ (∁Hn ), Γ = ∆(A, B), and let
Γ1 = ∆(A, ∂Hn ), Γ2 = ∆(B, ∂Hn ).
Then 
M(Γ) ≤ 2−n M(Γ1 ) + M(Γ2 ) .
In particular, the equality holds if A = g(B), where g is the reflection in Hn .

Proof. Let ρ1 ∈ F(Γ1 ) and ρ2 ∈ F(Γ2 ). We note that if γ ∈ Γ is a rectifiable


path, then γ has subpaths γ1 , γ2 such that γ1 ∈ Γ1 , γ2 ∈ Γ2 . Thus
Z Z
1 1
1≤ ρ1 ds + ρ2 ds.
2 γ1 2 γ2
We define ρ = ρ1 /2 + ρ2 /2. Now ρ is an admissible function for the curve family
Γ and hence Z
M(Γ) ≤ ρn dm.
Rn
We may assume that ρ1 (z) = 0 for z ∈ / Hn , and ρ2 (z) = 0 for z ∈ Hn . As
ρ = ρ1 /2 + ρ2 /2, we obtain
Z Z Z
n n
ρ dm = 2 −n
ρ1 dm + 2 −n
ρn2 dm
R n H n R \H
n n
Z Z 
= 2−n ρn1 dm + ρn2 dm .
Rn Rn
Introduction to quasiconformal mappings in n-space 251

It follows that 
M(Γ) ≤ 2−n M(Γ1 ) + M(Γ2 ) .
Next we consider the case A = g(B). Let ρ be an admissible function for the
path family Γ and denote ρ ◦ g by ρ̄. Now the function

ρ + ρ̄ on Hn ,
ρ̂ =
0 on ∁Hn ,
is admissible for the path family Γ1 . By the inequality (a + b)n ≤ 2n−1 (an + bn )
(for a, b ≥ 0) and the fact that M(Γ1 ) = M(Γ2 ) it follows that
Z Z
n 1
M(Γ1 ) ≤ ρ̂ dm = (ρ + ρ̄)n dm
Rn 2 Rn
Z Z
n−2 n n n−1
≤ 2 (ρ + ρ̄ )dm = 2 ρn dm.
Rn Rn
Hence, Z
n
M(Γ1 ) + M(Γ2 ) = 2M(Γ1 ) ≤ 2 ρn dm,
Rn
for any ρ admissible for the curve family Γ. By taking infimum over all admissible
ρ, the claim follows.

Capacity of a condenser. A condenser in Rn is a pair E = (A, C), where A


is open in Rn and C is a compact subset of A. The p-capacity of E is defined by
Z
(4.19) capp E = inf |∇u|p dm, 1 ≤ p < ∞,
u A

where the infimum is taken over all nonnegative functions u in ACLp (A) with
compact support in A and u|C ≥ 1. The n-capacity of E is called the conformal
capacity of E and denoted by capE.

Figure 7: Condenser E = (A, C).

Lemma 4.20. [22, 7.9] For all condensers (A, C) in Rn



(4.21) cap(A, C) = M ∆(C, ∂A; A) .
252 A. Rasila IWQCMA05

Sets of zero capacity. A compact set E in Rn is said to be of capacity zero,


denoted capE = 0, if there exists a bounded set A with E ⊂ A and cap(A, E) =
n n
0. A compact set E ⊂ R , E 6= R is said to be of capacity zero if E can
be mapped by a Möbius transformation onto a bounded set of capacity zero.
Otherwise E is said to be of positive capacity, and we write capE > 0.

n
Spherical symmetrizations. Let L be a ray from x0 to ∞ and E ⊂ R be
a compact set. We define spherical symmetrization of E in L as the set E ∗
satisfying the following conditions:
(1) x0 ∈ E ∗ if and only if x0 ∈ E,
(2) ∞ ∈ E ∗ if and only if ∞ ∈ E,
(3) For r ∈ (0, ∞) the set E ∗ ∩ S n−1 (x0 , r) is a closed spherical cap centered on
L with the same (n − 1)-dimensional Lebesgue measure as E ∩ S n−1 (x0 , r)
for E ∩ S n−1 (x0 , r) 6= ∅ and ∅ otherwise.
We note that E ∗ is always compact and connected if E is.

111
000
000
111
1111
0000
E
000
111
0000
1111 000
111
0000
1111
0000
1111 000
111
0000
1111 000
111
0000
1111
0000
1111 L
*
E
11111111111 L
00000000000
1111111
0000000 00000000000
11111111111
0000000
1111111 00000000000
11111111111
00000000000
11111111111
0000000
1111111 00000000000
11111111111
0000000
1111111
0000000
1111111 00000000000
11111111111
0000000
1111111
0000000
1111111
0000000
1111111

Figure 8: Spherical symmetrization.

Theorem 4.22. If E ∗ is the spherical symmetrization of E in a ray L, then


(1) m(E ∗ ) = m(E), and
(2) mn−1 (∂E ∗ ) ≤ mn−1 (∂E).

Proof. (Outline, [7, p.224]) By Fubini’s theorem


Z ∞ Z ∞
n−1

m(E ) = ∗
mn−1 (E ∩S (x0 , r))dr = mn−1 (E ∩S n−1 (x0 , r))dr = m(E),
0 0

which gives the first part.


To prove the second part, assume first that E is a polyhedron. Then for
r ∈ (0, ∞) the Brunn–Minkowski inequality yields
E ∗ (r) = {x : d(x, E ∗ ) ≤ r} ⊂ {x : d(x, E) ≤ r}∗ = E(r)∗ ,
Introduction to quasiconformal mappings in n-space 253

and hence
∗ m(E ∗ (r)) − m(E ∗ )
mn−1 (∂E ) ≤ lim sup
r→0 2r
m(E(r)) − m(E)
≤ lim sup = mn−1 (∂E).
r→0 2r
The result for the general domains is obtained by approximating the boundary
with polyhedrons.

C
1111111
0000000
0
0000000
1111111
0000000
1111111
C1
0000000
1111111 1111111
0000000
0000000
1111111 0000000
1111111
0000000
1111111
0000000
1111111
0000000
1111111 0000000
1111111
0000000
1111111
0000000
1111111 *
C0 C 1*
L0 0 L1 L0
11111111
00000000 0 111111111
000000000
000000000
111111111 L1
00000000
11111111 000000000
111111111
00000000
11111111
00000000
11111111 000000000
111111111

Figure 9: Spherical symmetrization of a ring.

Theorem 4.23. If R = R(C0 , C1 ) is a ring and if C0∗ and C1∗ are the sphrerical
symmetrizations of C0 and C1 in opposite rays L0 , L1 , then R∗ = R(C0∗ , C1∗ ) is a
ring with cap R∗ ≤ cap R.

Proof. (Idea, [7, p.225]) Let u be a locally lipschitz function that is admissible
for R. Choose u∗ such that {x : u∗ (x) ≤ t} = {x : u(x) ≤ t}∗ . Then u∗ is
admissible for R∗ and from Theorem 4.22 we obtain
Z Z
∗ n

cap(R ) ≤ |∇u | dm ≤ |∇u|n dm.
Rn Rn

By taking the infimum over all admissible u the claim follows.

Canonical ring domains. The complementary components of the Grötzsch


n
ring RG,n (s) in Rn are B and [se1 , ∞], s > 1, and those of the Teichmüller ring
RT,n (s) are [−e1 , 0] and [se1 , ∞], s > 0. We define two special functions γn (s),
s > 1 and τn (s), s > 0 by
 n 
γn (s) = M ∆(B , [se1 , ∞]) = γ(s),
τn (s) = M ∆([−e1 , 0], [se1 , ∞]) = τ (s),
respectively. The subscript n is omitted if there is no danger of confusion. We
shall refer to these functions as the Grötzsch capacity and the Teichmüller ca-
pacity.
254 A. Rasila IWQCMA05

Γ

se 1
B
n

−e1 0 se 1

Figure 10: Grötzsch ring RG,n (s) (left) and Teichmüller ring RT,n (s) (right).

Lemma 4.24. [22, 5.53] For all s > 1


γn (s) = 2n−1 τn (s2 − 1)
and that τn : (0, ∞) → (0, ∞) is a decreasing homeomorphism.

Proof. (Idea) Apply Lemma 4.18 and an auxiliary Möbius transformation.


Lemma 4.25. [22, 5.63(1)] Let s > 0. Then
τ (s) ≤ γ(1 + 2s) = 2n−1 τ (4s2 + 4s)

Proof. Let Γ = ∆(S n−1 (−e1 /2, 1/2), [se1 , ∞]). Then by Lemma 4.24
M(Γ) = γ(1 + 2s) = 2n−1 τ (4s2 + 4s).
By Lemma 4.4 τ (s) ≤ M(Γ).
Lemma 4.26. [3, (8.65),(8.62)] The following estimates hold for τn (t), t > 0:
 λ √ 
1−n n
√  1−n
τn (t) ≥ 2 ωn−1 log ( 1 + t + t) ,
2
and for γn (1/r), r ∈ (0, 1):
 √  
λn 1 + 1 − r2 1−n λn 1−n
γn (1/r) ≥ ωn−1 log ≥ ωn−1 log ,
2r r
where λn is the Grötzsch ring constant depending only on n.

The value of λn is known only for n = 2, namely λ2 = 4. For n ≥ 3 it is


known that 20.76(n−1) ≤ λn ≤ 2en−1 . For more information on λn , see [3, p.169].
Lemma 4.27. (see [9, 2.31]) Let 0 < r0 < 1. Then
 
M ∆(Bn (r), S n−1 ) ≥ γn (1/r) ≥ C(n, r0 )M ∆(Bn (r), S n−1 ) ,
for r0 > r > 0.

Proof. By Lemma 4.26,


 √  
λn 1 + 1 − r2 1−n λn 1−n
γn (1/r) ≥ ωn−1 log ≥ ωn−1 log .
2r r
Introduction to quasiconformal mappings in n-space 255

We note that  
λn log λn  1
log ≤ 1− log ,
r log r0 r
for 0 < r < r0 . Thus
 1 1−n
γn (1/r) ≥ C(n, r0 )ωn−1 log
r 
= C(n, r0 )M ∆(B (r), S n−1 ) ,
n

with
 1−n
log λn
C(n, r0 ) = 1− .
log r0
The second inequality follows immediately from the fact that the line segment
[0, r) is contained in the ball of radius r.
Remark 4.28. Note that C(n, r0 ) → 1 as r0 → 0 in Lemma 4.27.
Lemma 4.29. [22, 7.34] Let R = R(E, F ) be a ring in Rn , and let a, b ∈ E,
c, ∞ ∈ F be distinct points. Then
 
|a − c|
M(∆(E, F )) ≥ τ .
|a − b|
Equality holds for E = [−e1 , 0], a = 0, b = −e1 , F = [se1 , ∞), c = se1 , d = ∞.

It is not obvious from the definition how M(∆(E, F )), for nonempty E, F ∈
n
R , depends on the geometric setup and the structure of the sets E, F . The
following lemma gives a lower bound for M(∆(E, F )) in the terms of
d(E, F )/ min{d(E), d(F )}.
Lemma 4.30. [22, 7.38] Let E, F be disjoint continua in Rn with d(E), d(F ) > 0.
Then
M(∆(E, F )) ≥ τ (4s2 + 4s) ≥ cn log(1 + 1/s)
where s = d(E, F )/ min{d(E), d(F )} and cn > 0 is a constant depending only on
n.

This result can be improved to the following Lemma, which shows that M(∆(E, F ))
and s = d(E, F )/ min{d(E), d(F )} are simultaneously small or large, provided
that E, F are connected.
Lemma 4.31. [9, 2.30] For n ≥ 2 there are homeomorphisms h1 , h2 of the
positive real axis with the following property. If E, F are the components of the
n
complements of a nondegenerate ring domain in R , then
h1 (s) ≤ M(∆(E, F )) ≤ h2 (s),
where s = d(E, F )/ min{d(E), d(F )}.
256 A. Rasila IWQCMA05

n
Spherical metric. The stereographic projection π : R → S n ( 12 en+1 , 12 ) is de-
fined by
x − en+1
(4.32) π(x) = en+1 + , x ∈ Rn ; π(∞) = en+1 .
|x − en+1 |2
n
Stereographic projection is the restriction to R of the inversion in S n (en+1 , 1)
n+1
in R . Since π −1 = π, it follows that π maps the Riemann sphere S n ( 21 en+1 , 12 )
n n
onto R . The chordal metric q in R is defined by
n
(4.33) q(x, y) = |π(x) − π(y)|; x, y ∈ R .
Lemma 4.34. [22, 7.37] If R = R(E, F ) is a ring, then
 
1
(4.35) M(∆(E, F )) ≥ τ ,
q(E)q(F )
 
4q(E, F )
(4.36) M(∆(E, F )) ≥ τ .
q(E)q(F )

5. Quasiconformal mappings
A homeomorphism f : G → Rn , n ≥ 2, of a domain G in Rn is called qua-
siconformal if f is in ACLn , and there exists a constant K, 1 ≤ K < ∞ such
that
|f ′ (x)|n ≤ K|Jf (x)|, |f ′ (x)| = max |f ′ (x)h|,
|h|=1

a.e. in G, where f (x) is the formal derivative. The smallest K ≥ 1 for which
this inequality is true is called the outer dilatation of f and denoted by KO (f ).
If f is quasiconformal, then the smallest K ≥ 1 for which the inequality
|Jf (x)| ≤ Kl(f ′ (x))n , l(f ′ (x)) = min |f ′ (x)h|,
|h|=1

holds a.e. in G is called the inner dilatation of f and denoted by KI (f ). The


maximal dilatation of f is the number K(f ) = max{KI (f ), KO (f )}. If K(f ) ≤
K, f is said to be K-quasiconformal. It is well-known that
KI (f ) ≤ KOn−1 (f ), KO (f ) ≤ KIn−1 (f ),
and hence KI (f ) and KO (f ) are simultaneously finite.
Theorem 5.1. [21, 32.2,33.2] Let f : G → Rn be a quasiconformal mapping.
Then
(1) f is differentiable a.e.,
(2) f satisfies condition (N), i.e. if A ⊂ G and m(A) = 0, then m(f A) = 0.

The next lemma gives another definition of quasiconformality. This definition


is called the geometric definition, and it is very useful in applications. The proof
for equivalence of these definitions is given in [21].
Introduction to quasiconformal mappings in n-space 257

Lemma 5.2. A homeomorphism f : G → G′ is K-quasiconformal if and only if


M(Γ)/K ≤ M(f (Γ)) ≤ KM(Γ)
for every path family Γ in G.
We may also give geometric definitions for the inner and outer dilatations.
Again, we refer to [21] for the proofs for the equivalence of these definitions.
n
Let G, G′ be domains in R and f : G → G′ be a homeomorphism. Then
inner and outer dilatations of f are respectively
M (f (Γ)) M (Γ)
KI (f ) = sup , KO (f ) = sup ,
M (Γ) M (f (Γ))
where the suprema are taken over all path families Γ in G such that M (Γ) and
M (f (Γ)) are not simultaneously 0 or ∞. The maximal dilatation of f is
K(f ) = max{KI (f ), KO (f )}.
Theorem 5.3. [21, 13.2] Let f : G′ → G′′ , g : G → G′ be quasiconformal map-
pings. Then
(1) KI (f −1 ) = KO (f ),
(2) KO (f −1 ) = KI (f ),
(3) K(f −1 ) = K(f ),
(4) KI (f ◦ g) ≤ KI (f )KI (g),
(5) KO (f ◦ g) ≤ KO (f )KO (g),
(6) K(f ◦ g) ≤ K(f )K(g).
Examples. (see [21, pp.49-50]) (1) A homeomorphism f : G → f G satisfying
|x − y|/L ≤ |f (x) − f (y)| ≤ L|x − y|
for all x, y ∈ G is called L-bilipschitz. It is easy to see that L-bilipschitz maps
are L2(n−1) -quasiconformal.
(2) Let a 6= 0 be a real number, and let f (x) = |x|a−1 x. We can extend f to
n n
a homeomorphism f : R → R by defining f (0) = 0, f (∞) = ∞ for a > 0 and
f (0) = ∞, f (∞) = 0 for a < 0. Then f is quasiconformal with
KI (f ) = |a|, KO (f ) = |a|n−1 if |a| ≥ 1,
KI (f ) = |a| , KO (f ) = |a|−1
1−n
if |a| ≤ 1.
(3) Let (r, ϕ, z) be the cylindrical coordinates of a point x ∈ Rn , i.e. r ≥ 0,
0 ≤ ϕ ≤ 2π, z ∈ Rn−2 , and

 x1 = r cos ϕ,
x2 = r sin ϕ,

xj = zj−2 for 3 ≤ j ≤ n.
The domain Gα , defined by 0 < ϕ < α, is called a wedge of angle α, α ∈ (0, 2π).
Let 0 < α ≤ β < 2π. The folding f : Gα → Gβ , defined by
f (r, ϕ, z) = (r, βϕ/α, z),
is quasiconformal with KI (f ) = β/α, KO (f ) = (β/α)n−1 .
258 A. Rasila IWQCMA05

6. An application of the modulus technique


As an application, we give a bound for how close to a point α the values
attained by a quasiconformal mapping on a sequence of continua approaching
the boundary can be. The bound is given in the terms of the diameter of the
continua involved. In order to prove this result, we need the following lemmas.
This result is presented in [17, pp.638–639].
Lemma 6.1. Let w > 0 and t ∈ (0, min{w2 , 1/w}). Then
1 1 w 1
log < log < 2 log .
2 t t t

Proof. Since t < w2 , we have 1/ t < w/t. On the other hand, t < 1/w, or
w < 1/t, and hence w/t < 1/t2 . By taking logarithm the claim follows.
Lemma 6.2. Let C ⊂ Bn be connected and 0 < d(C) ≤ 1. Then m ≡
d(0, C)/d(C) < ∞ and if m > 0, then
1 n
M(Γ) ≥ τ (4m2 + 4m) ≥ 2−n τ (m); Γ = ∆(B (1/2), C; Bn ).
2
Proof. The second inequality holds by Lemma 4.25. To prove the first inequality,
n
we note that if C ∩ B (1/2) 6= ∅, then M(Γ) = ∞ and there is nothing to prove.
n
In what follows we may assume that C ∩ B (1/2) = ∅. Now the result follows
from the symmetry property of the modulus Lemma 4.11 and Lemma 4.30.
Theorem 6.3. Let f : Bn → Rn be a quasiconformal mapping or constant, α ∈
Rn and Cj a sequence of nondegenerate continua such that Cj → ∂Bn and |f (x)−
α| < Mj when x ∈ Cj , where Mj → 0 as j → ∞. If
  n−1
1 1
lim sup τ log = ∞,
j→∞ d(Cj ) Mj
then f ≡ α. In particular, if
 1−n  n−1
1 1
lim sup log log = ∞,
j→∞ d(Cj ) Mj
then f ≡ α.

Proof. Suppose that f is not constant. Let Γj = ∆(Bn (1/2), Cj ; Bn ). Then by


Lemma 6.2  
d(0, Cj )  1 
−n
M(Γj ) ≥ 2 τ ≥ 2−n τ .
d(Cj ) d(Cj )
n
Let w = d(f B (1/2), α) > 0. Now by Lemma 4.14
  1−n
w 1−n 1 1
M(f Γj ) ≤ ωn−1 log ≤ ωn−1 log ,
Mj 2 Mj
Introduction to quasiconformal mappings in n-space 259

whenever Mj < min{w2 , 1/w} by Lemma 6.1. Because M(Γj ) ≤ K(M(f Γj )), the
estimates above yield
  n−1
1 1
τ log ≤ 22n−1 Kωn−1 ,
d(Cj ) Mj
proving the first part of the claim.
The estimate (4.26) yields
 λ 
1−n n √ √  1−n
τ (t) ≥ 2 ωn−1 log 1+t+ t
2
where t = 1/d(Cj ). It follows that
 λ √   
n
√  1−n λ
n
√  1−n
log 1+t+ t ≥ log (1 + 2 t)
2 2
 λ 
n 2  1−n
= log 1+ p .
2 d(Cj )
We note that
 λ 1−n  1−n
n 2  λn 
log 1+ p ≥ 2 log p
2 d(Cj ) d(Cj )
whenever j is large enough. Let v = λn . Now by Lemma 6.1
  λ 1−n  1−n
n 1
2 log p ≥ 2 log ,
d(Cj ) d(Cj )
p
for d(Cj ) < min{v 2 , 1/v}. Hence
  n−1  1−n  n−1
1 1 2−2n 1 1
τ log ≤2 ωn−1 log log ,
d(Cj ) Mj d(Cj ) Mj
which gives the second part of the claim.

References
1. L. V. Ahlfors: Lectures on quasiconformal mappings, Van Nostrand Mathematical Stud-
ies, No. 10 D. Van Nostrand Co., Inc., Toronto, Ont.-New York-London, 1966.
2. C. Andreian–Cazacu: Foundations of quasiconformal mappings, Handbook of complex
analysis: geometric function theory, Vol. 2, edited by R. Kühnau, 687–753, Elsevier, Am-
sterdam, 2005.
3. G. D. Anderson, M. K. Vamanamurty and M. Vuorinen: Conformal invariants,
inequalities and quasiconformal mappings, Wiley-Interscience, 1997.
4. P. Caraman: n-dimensional quasiconformal (QCf) mappings, Editura Academiei
Române, Bucharest; Abacus Press, Newfoundland, N.J., 1974.
5. F. W. Gehring: Symmetrization of rings in space, Trans. Amer. Math. Soc. 101 (1961)
499–519.
6. F. W. Gehring: The Carathéodory convergence theorem for quasiconformal mappings in
the space, Ann. Acad. Sci. Fenn. Ser. A I 336/11, 1-21, 1963.
7. F. W. Gehring: Quasiconformal mappings, Complex analysis and its applications (Lec-
tures, Internat. Sem., Trieste, 1975), Vol. II, 213–268, IAEA, Vienna, 1976.
260 A. Rasila IWQCMA05

8. F. W. Gehring: Quasiconformal mappings in Euclidean spaces, Handbook of complex


analysis: geometric function theory, Vol. 2, edited by R. Kühnau, 1–29, Elsevier, Amster-
dam, 2005.
9. V. Heikkala: Inequalities for conformal capacity, modulus and conformal invariants,
Ann. Acad. Sci. Fenn. Math. Dissertationes 132 (2002), 1–62.
10. J. Heinonen: Lectures on analysis on metric spaces, Universitext, Springer-Verlag, New
York, 2001.
11. J. Heinonen and P. Koskela: Definitions of quasiconformality, Invent. Math. 120
(1995), no. 1, 61–79.
12. J. Heinonen and P. Koskela: Quasiconformal maps in metric spaces with controlled
geometry, Acta Math. 181 (1998), no. 1, 1–61.
13. T. Iwaniec and G. Martin: Geometric function theory and non-linear analysis, Oxford
Mathematical Monographs, The Clarendon Press, Oxford University Press, New York,
2001.
14. R. Kühnau (ed.): Handbook of complex analysis: geometric function theory, Vol. 1–2,
Amsterdam : North Holland/Elsevier, 2002, 2005.
15. O. Lehto: Univalent functions and Teichmller spaces, Graduate Texts in Mathematics,
109. Springer-Verlag, New York, 1987.
16. O. Lehto and K. I. Virtanen: Quasiconformal mappings in the plane, Second edition.
Translated from the German by K. W. Lucas. Die Grundlehren der mathematischen Wis-
senschaften, Band 126. Springer-Verlag, New York-Heidelberg, 1973.
17. A. Rasila: Multiplicity and boundary behavior of quasiregular maps, Math. Z. 250 (2005),
611–640.
18. Yu. G. Reshetnyak: Space mappings with bounded distortion, Translations of Mathe-
matical Monographs, 73, American Mathematical Society, Providence, RI, 1989.
19. S. Rickman: Quasiregular Mappings, Ergeb. Math. Grenzgeb. (3), Vol. 26, Springer-
Verlag, Berlin, 1993.
20. J. Väisälä: On quasiconformal mappings in space, Ann. Acad. Sci. Fenn. Ser. A I 298
(1961), 1–36.
21. J. Väisälä: Lectures on n-Dimensional Quasiconformal Mappings, Lecture Notes in
Math., Vol. 229, Springer-Verlag, Berlin, 1971.
22. M. Vuorinen: Conformal Geometry and Quasiregular Mappings, Lecture Notes in Math.,
Vol. 1319, Springer-Verlag, Berlin, 1988.
23. M. Vuorinen (ed.): Quasiconformal space mappings. A collection of surveys 1960–1990,
Lecture Notes in Mathematics, 1508, Springer-Verlag, Berlin, 1992.

Antti Rasila E-mail: antti.rasila@tkk.fi


Address:
Helsinki University of Technology, Institute of Mathematics,
P.O.Box 1100, FIN-02015 HUT, Finland
Proceedings of the International Workshop on Quasiconformal
Mappings and their Applications (IWQCMA05)

The universal Teichmüller space and related topics

Toshiyuki Sugawa

Abstract. In this survey, we give an expository account of the universal Teichmüller


space with emphasis on the connection with univalent functions. In the theory, the
Schwarzian derivative plays an important role. We also introduce many interesting results
involving Schwarzian derivatives and pre-Schwarzian derivatives, as well.

Keywords. universal Teichmüller space, univalent function, Schwarzian derivative, pre-


Schwarzian derivative.
2000 MSC. Primary: 30F60, Secondary: 30C55, 30C62.

Contents

1. Preliminary 262
1.1. Quasiconformal mappings 262
1.2. Hyperbolic Riemann surfaces 264
1.3. Quadratic differentials 264
1.4. Univalent functions 266
1.5. Grunsky inequality 266
1.6. Schwarzian derivative 268
2. The universal Teichmüller space 269
2.1. Definition 1: the quotient space of quasiconformal maps 269
2.2. Definition 2: quasisymmetric functions 269
2.3. Definition 3: marked quasidisks 270
2.4. Definition 4: Bers embedding 271
2.5. Equivalence of T1 through T4 271
3. Analytic properties of the Bers embedding 272
3.1. The Teichmüller space of a Riemann surface 272
3.2. Relationship with quasi-Teichmüller spaces 274
3.3. The Bers projection 275
3.4. Convexity 275
3.5. Teichmüller distance and other natural distances (metrics) 276
262 T. Sugawa IWQCMA05

4. Pre-Schwarzian models 277


4.1. The models T̂ (D) and T̂ (H) 277
4.2. The model T̂ (D∗ ) 278
4.3. Loci of typical subclasses of S 279
5. Univalence criteria 280
5.1. Univalence criteria due to Nehari and Ahlfors-Weill 280
5.2. Inner radius and outer radius 281
5.3. Pre-Schwarzian counterpart 282
5.4. Directions of further investigation 283
References 284

1. Preliminary
In this section, we prepare basic tools to understand the universal Teichmüller space.
The material is more or less standard, but for convenience, an expository account will
be given without proofs. The most convenient reference for overall topics is perhaps the
recently published handbook [61].

1.1. Quasiconformal mappings. A homeomorhism f of a plane domain D onto an-


other domain D′ is called a quasiconformal map if f has locally square integrable partial
derivatives (in the sense of distribution) and satisfies the inequality
|fz̄ | ≤ k|fz |
almost everywhere in D, where k is a constant with 0 ≤ k < 1,
fz = 21 (fx − ify ), fz̄ = 21 (fx + ify )
and
∂f ∂f
fx = , fy = .
∂x ∂y
It turns out that f preserves sets of (2-dimensional) Lebesgue measure zero and, in par-
ticular, fz 6= 0 a.e. Thus the quotient µ = fz̄ /fz is well defined as a Borel measurable
function on D and satisfies kµk∞ ≤ k < 1. This function is sometimes called the complex
dilatation of f and denoted by µf . More specifically, f is also called a K-quasiconformal
map, where K = (1 + k)/(1 − k). The minimal K = (1 + kµk∞ )/(1 − kµk∞ ) is called the
maximal dilatation of f and denoted by K(f ). It is known that a 1-quasiconformal map is
conformal (i.e., biholomorphic) and vice versa. The composition of a K1 -quasiconformal
map and a K2 -quasiconformal map is K1 K2 -quasiconformal map and the inverse map of
a K-quasiconformal map is also K-quasiconformal. In particular, K-quasiconformality is
The universal Teichmüller space and related topics 263

preserved under composition with conformal maps. Therefore, K-quasiconformality and,


hence, quasiconformality can be defined for homeomorphisms between Riemann surfaces.
In particular, we can argue quasiconformality of a homeomorphism of the Riemann sphere
b = C ∪ {∞}.
C
More precise information about compositions of quasiconformal maps will be needed
later. Let f : Ω → Ω′ and g : Ω → Ω′′ be quasiconformal maps. Then the complex
dilatation of g ◦ f −1 is given by

fz µg − µf
(1.1.1) (µg◦f −1 ◦ f ) = .
fz 1 − µf · µg

In particular, we obtain the following lemma.

Lemma 1.1.2. Let f : Ω → Ω′ and g : Ω → Ω′′ be quasiconformal maps. Then g ◦ f −1 is


conformal on Ω′ if and only if µf = µg a.e. in Ω.

Fundamental in the theory of quasiconformal maps is the following existence and


uniqueness theorem.

Theorem 1.1.3 (The measurable Riemann mapping theorem). For any measurable
function µ on C with kµk∞ < 1, there exists a unique quasiconformal map f : C → C
such that f (0) = 0, f (1) = 1 and fz̄ = µfz a.e. in C.

For the proof of the theorem and for more information about quasiconformal maps,
the reader should consult the books [3] and [69] as well as the paper [4] by Ahlfors and
Bers. See also the article “Beltrami Equation”, by Srebro and Yakubov, in [61, vol. 2] for
the recent development.
We denote by Belt(D) the open unit ball of the space L∞ (D) for a domain (or, more
generally, a measurable set) D. An element µ of Belt(D) is called a Beltrami coefficient
on D. For a Beltrami coefficient µ on C, the function f given in the measurable Riemann
mapping theorem will be denoted by f µ throughout the present survey.
Let µ be a Beltrami coefficient on the outside D∗ of the unit disk. We extend µ to
µ∗ ∈ Belt(C) by setting µ∗ (z) = µ(1/z̄) for z ∈ D. Let f be a quasiconformal auto-
morphism of C b fixing 1, −1, −i with µf = µ∗ . Since f (z) and 1/f (1/z̄) both have the
same complex dilatation µ∗ and satisfy the same normalization condition, they must be
equal by uniqueness part of the measurable Riemann mapping theorem. In particular,
|f (z)|2 = 1 for z ∈ ∂D, and consequently, f maps D∗ onto itself. We define wµ : Cb→C b
by wµ = f. Recall that wµ fixes 1, −1 and −i.
The following fact was observed by Ahlfors and Bers [4].
264 T. Sugawa IWQCMA05

Theorem 1.1.4. Let µt be a family of Beltrami coefficients on C holomorphically param-


eterized over a complex manifold X. Then the map t 7→ f µt (z) is holomorphic on X for a
fixed z ∈ C.

We do not explain the meaning of “holomorphically parameterized” here. It is, however,


sufficient practically to know that (tµ + ν)/(1 + tν̄µ) is a family of Beltrami coefficients
holomorphically parameterized over the unit disk |t| < 1, where µ, ν ∈ Belt(C).

1.2. Hyperbolic Riemann surfaces. A connected complex manifold of complex di-


mension one is called a Riemann surface. The Poincaré-Koebe uniformization theorem
tells us that every Riemann surface R admits an analytic universal covering projection
p of the unit disk D = {z ∈ C : |z| < 1} onto R except for the case when R is confor-
b the complex plane C, the punctured complex
mally equivalent to the Riemann sphere C,
plane C∗ = C \ {0} or a complex torus (a smooth elliptic curve). Those non-exceptional
Riemann surfaces are called hyperbolic.
The group of analytic automorphisms of R is denoted by Aut(R). The group of disk
automorphisms Aut(D) is identified with PSU(1, 1) and isomorphic to PSL(2, R) through
the Möbius transformation M : H = {z : Im z > 0} → D defined by M (z) = (z−i)/(z+i).
Thus Aut(D) inherits a structure of real Lie group. A subgroup Γ of Aut(D) is called
Fuchsian if Γ is discrete. It is known that Γ is discrete if and only if Γ acts on D properly
discontinuously. Note also that Γ is torsion-free if and only if Γ acts on D without fixed
points. The covering transformation group Γ = {γ ∈ Aut(D) : p ◦ γ = p} is a torsion-
free Fuchsian group and will be called the Fuchsian model of R. Conversely, for a given
torsion-free Fuchsian group Γ the quotient space D/Γ has natural complex structure so
that the projection D → D/Γ becomes an analytic universal covering. In this way, the
theory of hyperbolic Riemann surfaces can be translated into that of torsion-free Fuchsian
groups.
Since the Poincaré metric ρD (z)|dz| = |dz|/(1−|z|2 ) is invariant under the pull-back by
analytic automorphisms of D, it projects to a smooth metric, denoted by ρR = ρR (w)|dw|,
on the hyperbolic Riemann surface R via p. The metric ρR is called the hyperbolic metric
of R. Thus ρR is characterized by the relation ρD = p∗ (ρR ) = ρR (p(z))|p′ (z)||dz|.
Note that ρR has constant Gaussian curvature −4, in other words, ∆ log ρR = 4ρ2R .
The Schwarz-Pick lemma implies the contraction property f ∗ ρS ≤ ρR for a holomorphic
map f : R → S between hyperbolic Riemann surfaces R and S, where equality holds at
some (hence every) point in R iff f is a covering projection of R onto S.

1.3. Quadratic differentials. Let H(D) be the set of analytic functions on the unit
disk D and let n be a non-negative integer. For a Fuchsian group Γ, a function ϕ ∈ H(D)
is said to be automorphic for Γ (with weight −2n) if ϕ satisfies the functional equation
The universal Teichmüller space and related topics 265

(ϕ ◦ γ)(γ ′ )n = ϕ for every γ ∈ Γ, that is to say, ϕ(z)dz n is an invariant n-form for Γ. The
set of automorphic functions for Γ with weight −2n will be denoted by Hn (D, Γ).
An element ϕ of Hn (D, Γ) for a torsion-free Fuchsian group Γ projects to a holomorphic
n-form q = q(w)dwn on R = D/Γ so that p∗n q = ϕ(z)dz n , where p∗n q means the pull-back
q(p(w))(p′ (w))n of the n-form q by the canonical projection p : D → D/Γ.
We now define two norms for ϕ ∈ Hn (D, Γ) by
ZZ
kϕkAn (D,Γ) = |ϕ(z)|(1 − |z|2 )n−2 dxdy,
ω

kϕkBn (D,Γ) = sup |ϕ(z)|(1 − |z|2 )n ,


z∈D

where ω is a fundamental domain for Γ, that is, a subdomain of D such that ω ∩ γ(ω) = ∅
S
for every γ ∈ Γ with γ 6= id, γ∈Γ γ(ω̄) = D and ∂ω is of zero area. We denote by
An (D, Γ) and Bn (D, Γ) the subsets of Hn (D, Γ) consisting of ϕ with finite norm kϕkAn (D,Γ)
and kϕkBn (D,Γ) , respectively. It is easy to see that these become complex Banach spaces.
When Γ is the trivial group 1, we write An (D) and Bn (D) for An (D, 1) and Bn (D, 1),
respectively.
The definition of the spaces An (D) and Bn (D) can be extended for hyperbolic Riemann
surfaces R. Let Hn (R) denote the set of holomorphic n-forms on R and set
ZZ
kϕkAn (R) = |ϕ(w)|ρR (w)2−n dxdy,
R

kϕkBn (R) = sup |ϕ(w)|ρR (w)−n


w∈R

for ϕ = ϕ(w)dwn in Hn (R). Here, we should note that |ϕ(w)|ρR (w)−n does not depend on
the choice of the local coordinate w, in other words, |ϕ|ρ−n
R can be regarded as a function
on R.
The Banach spaces An (D, Γ) and An (D/Γ) (resp. Bn (D, Γ) and Bn (D/Γ)) are isomet-
rically isomorphic through the pull-back p∗n by the projection p : D → D/Γ. Also, the
following invariance property is convenient to note.

Lemma 1.3.1. Let R and S be hyperbolic Riemann surfaces and let p : R → S be a


conformal homeomorphism. Then the pullback operator p∗n : Bn (S) → Bn (R) is a linear
isometry, in other words,
kp∗n ϕkBn (R) = kϕkBn (S) , ϕ ∈ Bn (S).

In the theory of Teichmüller spaces, it is important to consider the spaces A2 and B2


as we shall see later. A 2-form q(w)dw2 is traditionally called a quadratic differential.
266 T. Sugawa IWQCMA05

1.4. Univalent functions. In connection with the universal Teichmüller space, the the-
ory of univalent functions is of particular importance. The best textbook in this direction
is [67] by O. Lehto.
We denote by S the set of analytic univalent functions f on the unit disk so normalized
that f (0) = 0 and f ′ (0) = 1. An analytic function f around the origin is said to be
strongly normalized if f (0) = f ′ (0) − 1 = f ′′ (0) = 0. Let S0 be the subset of S consisting
of strongly normalized functions. For f ∈ S , the function g = f /(1 + af ), where
a = f ′′ (0)/2, is strongly normalized but not necessarily analytic in D. It is thus natural
to consider the wider class

S˜0 = {f : meromorphic and univalent in D and strongly normalized}

than S0 .
The following meromorphic counterpart is also useful in the theory of univalent func-
tions. Let Σ be the set of meromorphic univalent functions F on the exterior D∗ = {ζ ∈
b : |ζ| > 1} of the unit disk so normalized that
C

b1 b2
(1.4.1) F (ζ) = ζ + b0 + + 2 + ...
ζ ζ

in |ζ| > 1.
For f ∈ S , the function F (ζ) = 1/f (1/ζ) belongs to Σ and satisfies the condition
0∈/ F (D∗ ), and vice versa. Let Σ′ denote the set of those functions F ∈ Σ that satisfy
0∈/ F (D∗ ). Moreover, b0 = 0 for a function F (ζ) = ζ + b0 + b1 /ζ + . . . in Σ if and only if
f ∈ S˜0 , where f (z) = 1/F (1/z). Hence, if we set

Σ0 = {F ∈ Σ : F (ζ) − ζ → 0 as ζ → ∞},

the correspondence f (z) 7→ F (ζ) = 1/f (1/ζ) gives bijections of S˜0 onto Σ0 and of S0
onto Σ′0 , where we define Σ′0 = Σ0 ∩ Σ′ .

1.5. Grunsky inequality. For a meromorphic function F near the point at infinity with
an expansion of the form (1.4.1), we take a single-valued branch of log((F (ζ)−F (ω))/(ζ −
ω)) in |ζ| > R and |ω| > R for sufficiently large R > 0 and expand it in the form

X∞ X ∞
F (ζ) − F (ω) bj,k
log =−
ζ −ω j=1 k=1
ζ j ωk

there. The coefficients bj,k are called the Grunsky coefficients of F. It is easy to see that
bj,k = bk,j and b1,k = bk for j, k ≥ 1, where bk is the coefficient in (1.4.1). The last relation
is deduced in the following way. If we write F (ζ) = ζ + b0 + G(ζ), then G(ζ) = O(|ζ|−1 )
The universal Teichmüller space and related topics 267

as ζ → ∞. Fix ω for a moment. Since


 
F (ζ) − F (ω) G(ζ) − G(ω) G(ζ) − G(ω)
log = log 1 + = + O(|ζ|−2 ),
ζ −ω ζ −ω ζ −ω
we obtain

X b1,k F (ζ) − F (ω)
= − lim ζ log
k=1
ωk ζ→∞ ζ −ω

G(ζ) − G(ω)
= − lim ζ
ζ→∞ ζ −ω
= G(ω),

from which the required relation follows.


The following theorem is greatly useful in the theory of Teichmüller spaces as well as
the theory of univalent functions. See [42], [29] or [86] for the proof and applications.

Theorem 1.5.1 (Grunsky). A meromorphic function F (ζ) with expansion of the form
(1.4.1) around ζ = ∞ is analytically continued to a univalent meromorphic function in
|ζ| > 1 if and only if the inequality
∞ 2
X∞ X X∞
|xj |2

(1.5.2) k bj,k xj ≤
j
k=1 j=1 j=1

holds for an arbitrary sequence of complex numbers x1 , x2 , . . . .

The inequality in (1.5.2) is known as the strong Grunsky inequality. Noting b1,k = bk ,
we take (x1 , x2 , x3 , . . . ) = (1, 0, 0, . . . ) to obtain

X
(1.5.3) k|bk |2 ≤ 1.
k=1

This inequality is known as Gronwall’s area theorem.


It is also known that inequality (1.5.2) can be replaced in the above theorem by the
(classical) Grunsky inequality:
∞ ∞
X X X ∞
|xj |2

(1.5.4) bj,k xj xk ≤ .
j
j=1 k=1 j=1


The symmetric matrix ( jkbj,k ) defines a linear operator on ℓ2 , where bj,k are the
Grunsky coefficient of a meromorphic function F (ζ) around ζ = ∞. This is sometimes
called the Grunsky operator and will be denoted by G[F ] in the following. The strong
268 T. Sugawa IWQCMA05

Grunsky inequality says that F ∈ Σ if and only if G[F ] is a bounded linear operator on
ℓ2 with operator norm ≤ 1. Here, the operator norm kG[F ]k of G[F ] is defined by

∞ X
2
X p
2
kG[F ]k = sup jkbj,k yj ,
kyk2 =1
k=1 j=1

P
where kyk2 = ( k |yk |2 )1/2 for y = (y1 , y2 , . . . ). Thus, F ∈ Σ ⇔ kG[F ]k ≤ 1. It is
known (cf. [86]) that F has a quasiconformal extension to C b if and only if kG[F ]k < 1.
See also the article “Univalent holomorphic functions with quasiconformal extensions”, by
Krushkal, in [61, vol. 2].

1.6. Schwarzian derivative. For a non-constant meromorphic function f on a domain,


we define Tf and Sf by
f ′′
Tf = = (log f ′ )′ ,
f′
 2
1′ f ′′′ 3 f′
Sf = (Tf ) − (Tf )2 = ′ − .
2 f 2 f

These are called the pre-Schwarzian derivative and the Schwarzian derivative of f, respec-
tively. Note that Tf is analytic at a finite point z0 if and only if f is analytic and injective
around z0 . Similarly, Sf is analytic at z0 if and only if f is meromorphic and injective
around z0 . The following two lemmas show usefulness of these operations.

Lemma 1.6.1. Let f be a non-constant meromorphic function on a domain D. The


pre-Schwarzian derivative of f vanishes on D if and only if f is (the restriction of )
a similarity. The Schwarzian derivative of f vanishes on D if and only if f is (the
restriction of ) a Möbius transformation.

Lemma 1.6.2. Let f and g be non-constant meromorphic functions for which the com-
position f ◦ g is defined. Then
Tf ◦g = (Tf ) ◦ g · g ′ + Tg = g1∗ (Tf ) + Tg ,
Sf ◦g = (Sf ) ◦ g · (g ′ )2 + Sg = g2∗ (Sf ) + Sg .

Combining these lemmas, we observe that SL◦f ◦M = M2∗ (Sf ) for Möbius transforma-
tions L and M. Thus the Schwarzian derivative behaves like a quadratic differential.
The universal Teichmüller space and related topics 269

2. The universal Teichmüller space

We have two choices to develop the theory of the (universal) Teichmüller space; the
unit disk model or the upper half-plane model. Although they can be translated into
each other, in principle, via the Möbius transformation z 7→ (z − i)/(z + i), both models
have their own advantage and thus can be chosen at will according to the purpose. In the
present survey, we will take the unit disk model to connect with the theory of univalent
functions in a direct way.

2.1. Definition 1: the quotient space of quasiconformal maps. We denote by


QC(D) the set of quasiconformal automorphisms of the unit disk D. As we will observe
later, every function in QC(D) extends to a unique homeomorphism of the closed unit
disk D. Thus, we may think that every f ∈ QC(D) is a self-homeomorphism of the closed
unit disk D. Two functions f and g in QC(D) are said to be Teichmüller equivalent and
T
denoted by f ∼g if there exists a disk automorphism L ∈ Aut(D) such that g = L ◦ f on
T
∂D. The quotient space QC(D)/∼ is a model of the universal Teichmüller space and will
be denoted by T1 for a moment. The equivalence class represented by f ∈ QC(D) will be
denoted by [f ] below.
Let f, g ∈ QC(D). The Teichmüller distance between p = [f ] and q = [g] is defined by

1
d1 (p, q) = inf log K(g1 ◦ f1−1 ).
f1 ∼f,g1 ∼g 2
T T

Recall here that K(f ) denotes the maximal dilatation of f. By a compactness property
of quasiconformal maps, one can check that d1 (p, q) is indeed a distance on T1 . In this
way, T1 becomes a metric space. It can also be shown that T1 is a complete metric
space with metric d1 by a normality property of the set of normalized K-quasiconformal
automorphisms of C (see [69]).

2.2. Definition 2: quasisymmetric functions. The notion of quasisymmetric func-


tions was created by Beurling and Ahlfors [18] for functions on the real line. We give
here a corresponding definition of quasisymmetric functions on the unit circle. A sense-
preserving homeomorphism h of the unit circle ∂D is called quasisymmetric if

1 |h(ei(s+t) ) − h(eis )| π
≤ ≤ M, s ∈ R, 0 < t <
M |h(eis ) − h(ei(s−t) )| 2

for a constant M ≥ 1. The set of all quasisymmetric functions on the unit circle will be
denoted by QS(∂D). The main result in [18] can be stated as follows.
270 T. Sugawa IWQCMA05

Theorem 2.2.1 (Beurling-Ahlfors). The restriction of a quasiconformal automorphism


of the unit disk to the unit circle is quasisymmetric. Conversely, a quasisymmetric func-
tion on the unit circle can be extended to a quasiconformal automorphism of the unit disk.

Two functions h1 and h2 on the unit circle are called Möbius equivalent if there exists
a disk automorphism L ∈ Aut(D) such that h2 = L ◦ h1 . Let T2 denote the quotient space
of QS(∂D) by the Möbius equivalence. By the above theorem of Beurling and Ahlfors,
one readily sees that T1 can be identified with T2 in a natural manner.
In order to get rid of taking quotient, we can define T2 as follows. A (sense-preserving)
homeomorphism h of ∂D is said to be normalized if h fixes the points 1, −1 and −i. Since
every Möbius equivalence class of quasisymmetric functions is represented by a unique
normalized one, one can identify T2 with the set of normalized quasisymmetric functions
on the unit circle.
See [37] for a modern treatment of quasisymmetric functions. The survey article “Uni-
versal Teichmüller space”, by Gardiner and Harvey, in [61, vol. 1] puts emphasis on the
connection with quasisymmetric functions.

2.3. Definition 3: marked quasidisks. A simply connected domain D in C b is called a


b If
quasidisk if D is the image of the unit disk under a quasiconformal automorphism of C.
D is the image under a K-quasiconformal automorphism, then D is called a K-quasidisk.
Many characteristic properties of quasidisks are known. See, for instance, [40].
Let D be a quasidisk (or a Jordan domain more generally) and x1 , x2 , x3 are positively
ordered (distinct) points on ∂D. The quadruple (D, x1 , x2 , x3 ) will be called a marked
quasidisk. By the Riemann mapping theorem and the Carathéodory extension theorem,
there exists a unique conformal homeomorphism g : H → D with g(0) = x1 , g(1) = x2
and g(∞) = x3 .
We denote by Q the set of all marked quasidisks in C. b Two marked quasidisks (D, xj )
and (D′ , x′j ) are said to be Möbius equivalent if D′ = L(D) and x′j = L(xj ), j = 1, 2, 3,
for some Möbius transformation L ∈ Möb = Aut(C). b We can define a pseudo-metric on
Q by
d((D, xj ), (D′ , x′j )) = kSf kB2 (D) ,
where f is a conformal homeomorphisms of D onto D′ with f (xj ) = x′j . It is easy to see
that d(D, D′ ) = 0 if and only if D and D′ are Möbius equivalent.
The set T3 of Möbius equivalence classes of all marked quasidisks constitutes another
model of the universal Teichmüller space and the above-defined pseudo-metric gives a
metric on T3 , which will be denoted by d3 , i.e.,
d3 (p, q) = inf d((D, xj ), (D′ , x′j ))
(D,xj )∈p,(D′ ,x′j )∈q
The universal Teichmüller space and related topics 271

for p, q ∈ T3 .
We can again take a suitable normalization to avoid the process of quotient and even
marking. For instance, we may say that a quasidisk D is normalized if its boundary
contains the points 0, 1 and ∞ in positive order along the boundary curve. If we denote
b then T3 can be identified with Q0 naturally,
by Q0 the set of normalized quasidisks in C,
and the restriction of the distance d on Q0 corresponds to the distance d3 on T3 .
In the above, the marking is important. For two simply connected hyperbolic domains
D1 and D2 , we set
d(D, D′ ) = inf kSf kB2 (D) .
f :D→D conformal

It is easy to see that d is a pseudo-distance. Lehto [67] posed a question whether or


not d(D, D′ ) = 0 implies that D and D′ are Möbius equivalent. Osgood and Stowe [82]
answered to this question in the negative (see also [19]).

2.4. Definition 4: Bers embedding. Let D be a hyperbolic domain in C. b We define


a subset T (D) of B2 (D) to consists of those holomorphic quadratic differentials ϕ(z)dz 2
on D such that ϕ = Sf for some univalent meromorphic function f on D which extends
to a quasiconformal automorphism of the Riemann sphere. Note that kSf kB2 (D) ≤ 12 for
every univalent meromorphic function f on D (see §5.2 and [10]). By Lemmas 1.6.1 and
1.6.2, for a Möbius transformation L, the pull-back L∗2 gives an isometric isomorphism of
B2 (L(D)) onto B2 (D). In particular, for a circle domain ∆, that is, the interior or the
exterior of a circle, or a half-plane, the space B2 (∆) is isomorphic, say, to B2 (D∗ ). The
space T4 = T (D∗ ) (or its equivalent) is a model of the universal Teichmüller space and
known as the Bers embedding of the universal Teichmüller space.
Ahlfors [2] showed the following.

Theorem 2.4.1. T (D∗ ) is a bounded, connected and open subset of B2 (D∗ ).

Thus T4 = T (D∗ ) inherits a complex structure and a metric from B2 (D∗ ). We denote by
d4 the distance, namely, d4 (ϕ, ψ) = kϕ − ψkB2 (D∗ ) for ϕ, ψ ∈ T4 . Since T (D∗ ) is bounded,
the distance d4 is not complete.

2.5. Equivalence of T1 through T4 . We see now that the above definitions of the
universal Teichmüller space are all equivalent. Firstly, consider the restriction map
QC(D) → QS(∂D) defined by f 7→ f |∂D . Then this map yields a bijection of T1 onto
T2 .
Secondly, we see the equivalence of T3 and T4 . For ϕ ∈ T4 = T (D∗ ), by definition,
there exists a quasiconformal map f of Cb fixing 0, 1, ∞ such that f is conformal on D∗
and satisfies Sf = ϕ. Then the image D = f (D∗ ) is a normalized quasidisk. Therefore, the
correspondence ϕ 7→ D gives a map T4 → T3 . We next show that this map is bijective.
272 T. Sugawa IWQCMA05

Suppose that a normalized quasidisk D is given. By definition, D = h(D∗ ) for some


quasiconformal map h of C b with h(1) = 0, h(−1) = 1 and h(−i) = ∞. Let µ = µh |D∗
b → C.
and set f = h ◦ (wµ )−1 : C b Then f is quasiconformal map of C,
b is conformal on
wµ (D∗ ) = D∗ and satisfies f (1) = 0, f (−1) = 1 and f (−i) = ∞. Therefore, ϕ = Sf
belongs to T4 = T (D∗ ). In this way, we obtain the map of T3 into T4 , which is obviously
the inverse map of the previously defined map of T4 to T3 . We have now concluded that
T3 and T4 are equivalent by those maps.
Finally, we connect T1 with T4 . Let h ∈ QC(D). We define µ ∈ Belt(C) by
(
µh on D,
µ=
0 on D∗

and define a quasiconformal map f : C b →C b by f = f µ , where f µ was defined in §1.1.


Since f is conformal in D∗ , the Schwarzian derivative Sf belongs to T4 = T (D∗ ). Note that
f ◦ h−1 is conformal in D by construction. Let h1 ∈ QC(D) be Teichmüller equivalent to
h and define f1 in the same way as above. We claim now that Sf1 = Sf . By assumption,
h1 = L ◦ h on ∂D for an L ∈ Aut(D). Define a map g : C b →C b by
(
f1 ◦ f −1 on Cb \ f (D),
g=
f1 ◦ h−1
1 ◦L◦h◦f
−1
on f (D).

It is clear that g is conformal on f (D) and f (D∗ ). Furthermore, since h−1


1 ◦ L ◦ h = id

on ∂D, the map g is continuous on C. b Since C = f (∂D) and g(C) = f1 (∂D) are both
b Since µg = 0 a.e., we conclude
quasicircles, it turns out that g is quasiconformal in C.
that g is conformal, hence, a Möbius map. Because of the relation f1 = g ◦ f on f (D∗ ),
Sf = Sf1 follows as required.
In this way, we obtain the mapping of T1 to T4 : [h] 7→ Sf |∗D . It is not difficult to see
that this mapping is bijective. This map is called the Bers embedding.

3. Analytic properties of the Bers embedding


3.1. The Teichmüller space of a Riemann surface. It is beyond the scope of the
present survey to develop the theory of Teichmüller spaces of Riemann surfaces in full
generality. Here, our focus will be on the Bers embedding of the Teichmüller space of a
Riemann surface. See [75], [45], [35], [36] for general properties of Teichmüller spaces. See
also [1], [113] for a differential geometric approach, [94] for an algebraic approach.
For simplicity, we assume a Riemann surface R to be hyperbolic, in other words,
there exists a torsion-free Fuchsian group Γ acting on D such that D/Γ is conformally
equivalent to R. Thus, we can identify R with D/Γ. We denote by p : D → D/Γ = R
The universal Teichmüller space and related topics 273

the canonical projection. Two quasiconformal maps fj : R → Sj , j = 1, 2, are called


Teichmüller equivalent if there exists a conformal homeomorphism g : S1 → S2 such that
f2−1 ◦ g ◦ f1 : R → R is homotopic to the identity relative to the ideal boundary. We omit
the explanation of the term “relative to the ideal boundary”. See the references given
above for details. Also, [33] gives several useful equivalent conditions for that.
The Teichmüller space Teich(R) of R is defined as the set of all the Teichmüller equiv-
alence classes of such quasiconformal maps of R onto another surface.
Suppose that f1 : R → S1 and f2 : R → S2 are quasiconformal maps. Let Γj be a
Fuchsian model of Sj acting on D and hj : D → D be a lift of fj , namely, pj ◦ hj = fj ◦ p,
where pj : D → D/Γj = Sj is the canonical projection. Then, it is known that f1 and f2
are Teichmüller equivalent if and only if h1 and h2 are Teichmüller equivalent in the sense
of §2.1. Note that hj ◦ γ ◦ h−1 −1
j ∈ Γj for each γ ∈ Γ, namely, hj Γhj = Γj .

Set
QC(D, Γ) = {h ∈ QC(D) : hΓh−1 is Fuchsian}
T
and denote by Teich(Γ) the quotient space QC(D, Γ)/∼. As we have seen, Teich(R) and
Teich(Γ) are canonically isomorphic through the universal covering projection p : D →
D/Γ = R. Also, Teich(Γ) is naturally contained in Teich(1) = T1 . In this sense, the
universal Teichmüller space T (D∗ ) contains all the Teichmüller space of an arbitrary
hyperbolic Riemann surface.
By using (1.1.1), the complex dilatation of f ∈ QC(D, Γ) is seen to be contained in

Belt(D, Γ) = {µ ∈ Belt(D) : (µ ◦ γ)γ ′ /γ ′ = µ ∀γ ∈ Γ}.

Furthermore, for h ∈ QC(D, Γ), let f be the function constructed in §2.5 and let γ ∈ Γ.
Since f and γ ◦ f ◦ γ −1 has the same complex dilatation, γ ◦ f ◦ γ −1 = L ◦ f for an
L ∈ Aut(C) b = Möb by Lemma 1.1.2. Lemma 1.6.2 now implies that γ ∗ (Sf ) = Sf .
2
Therefore, Sf is contained in the closed subspace B2 (D∗ , Γ) of B2 (D∗ ) defined in §1.3. As
in the previous section, we see that Sf depends only on the Teichmüller equivalence class
of h in QC(D, Γ) and the corresponding h 7→ Sf is one-to-one, we obtain an embedding
βΓ : Teich(D, Γ) → B2 (D∗ , Γ), which is called the Bers embedding of Teich(D, Γ). We set
T (D∗ , Γ) = βΓ (Teich(D, Γ)).
Bers [15] showed that T (D∗ , Γ) is a bounded domain in B2 (D∗ , Γ). It is obvious that
T (D∗ , Γ) is contained in T (D∗ ) by definition. Indeed, by using the Douady-Earle extension
[28], it can be seen that
T (D∗ , Γ) = T (D∗ ) ∩ B2 (D∗ , Γ)
and that T (D∗ , Γ) is contractible.
274 T. Sugawa IWQCMA05

3.2. Relationship with quasi-Teichmüller spaces. In view of the description of the


set T (D∗ , Γ), it may be natural to consider the following sets more generally. Let D be a
hyperbolic domain in C b and let G be a subgroup of Aut(D). Typically, G is a Kleinian
group and D is a connected component of its region of discontinuity. Then we set (cf. [98])

b s.t. ϕ = Sf and f is univalent in D},


S(D, G) = {ϕ ∈ B2 (D, G) : ∃f : D → C
b s.t. ϕ = Sf and f extends to a qc map of C},
T (D, G) = {ϕ ∈ B2 (D, G) : ∃f : D → C b

For a circle domain ∆ and a Fuchsian group Γ acting on ∆, the set S(∆, Γ) sometimes
called the quasi-Teichmüller space of Γ. (But, note that this terminology is not popular.)
Clearly, T (D, G) ⊂ S(D, G). It is easy to see that S(D∗ ) is closed while, as Ahlfors
showed, T (D∗ ) is open in B2 (D∗ ). The boundary of T (∆, Γ) in B2 (∆, Γ) is called the Bers
boundary and is important in relation with the deformation theory of Kleinian groups
(see [16]).

When G is the trivial group 1, we write S(D), T (D) for S(D, 1), T (D, 1), respectively.
Note that under the mapping f 7→ Sf , the sets S˜0 and Σ0 correspond to S(D) and
S(D∗ ), respectively, in one-to-one fashion. It is a challenging problem to characterize
those functions f in S whose Schwarzian derivatives lie on ∂T (D). See [7] and [43] for
some attempts.

In 1970’s, it had been a conjecture of Bers [16] that the closure of T (D∗ ) in B2 (D∗ )
is S(D∗ ). In 1978, Gehring [39] disproved it. Prior to it, Gehring [38] proved the weaker
assertion that the interior of S(D∗ ) in B2 (D∗ ) coincides with T (D∗ ). See [34] for a relevant
result. Thurston [110] proved the more striking result that S(D∗ ) even has an isolated
point in B2 (D∗ ) (see also [5]). After that, the Bers conjecture was reformulated in the form
that the closure of T (D∗ , Γ) is equal to S(D∗ , Γ) for a cofinite Fuchsian group Γ, that is, a
finitely generated Fuchsian group of the first kind. (This is nowadays generalized to the
Bers-Thurston density conjecture.) Shiga [95] proved a weaker version of it: the interior of
S(D∗ , Γ) in B2 (D∗ , Γ) coincides with T (D∗ , Γ) for a cofinite Γ. In the line of these studies,
the author showed that S(D∗ , Γ) \ T (D∗ ) 6= ∅ for a Fuchsian group Γ of the second kind
([99]) and that the interior of S(D∗ , Γ) in B2 (D∗ , Γ) coincides with T (D∗ , Γ) for a finitely
generated, purely hyperbolic Fuchsian group Γ of the second kind ([100]). Matsuzaki [71]
generalized the former to the case of a certain kind of infinitely generated Fuchsian groups
of the first kind. In recent years, a huge amount of progress has been made in the theory
of Kleinian groups, which enabled to prove the Bers-Thurston conjecture partially. See,
for instance, [20] and [80] for partial solutions to the conjecture.

We end the subsection with the following conjecture.


The universal Teichmüller space and related topics 275

Conjecture 3.2.1. The interior of quasi-Teichmüller space S(D∗ , Γ) in B2 (D∗ , Γ) is equal


to the Bers embedding T (D∗ , Γ) of the Teichmüller space of a Fuchsian group Γ acting on
D∗ .
Note that Zhuravlev (Žuravlev) [117] proved that T (D∗ , Γ) is the connected component
of the interior of S(D∗ , Γ) which contains the origin for an arbitrary Fuchsian group Γ (see
also [98]). Thus the conjecture is equivalent to connectedness of the interior of S(D∗ , Γ).

3.3. The Bers projection. Let D be a hyperbolic domain in C b and denote by E its
complement in C. b We define the map Φ : Belt(E) → B2 (D) by Φ(µ) = Sf µ | , where f µ
D

is defined as in §1.1 for µ which is extended to C by setting µ = 0 on D. Recall here that


Belt(E) is the open unit ball of the complex Banach space L∞ (E) with norm k · k∞ . It
is clear by definition that Φ(Belt(E)) = T (D). The map Φ : Belt(E) → T (D) is called
the (generalized) Bers projection. It is known that Φ : Belt(E) → B2 (D) is holomorphic
(cf. [98]) and that the Fréchet derivative d0 Φ : L∞ (E) → B2 (D) of Φ at the origin is
described by
ZZ
6 ν(ζ)
d0 Φ[ν](z) = − dξdη (ζ = ξ + iη)
E (ζ − z)
π 4

for ν ∈ L∞ (E). Bers [15] strengthened Ahlfors’ theorem (Theorem 2.4.1) to the following
form.
Theorem 3.3.1. The Bers projection Φ : Belt(D) → T (D∗ ) is a holomorphic split
submersion, in other words, the Fréchet derivative of Φ at every point exists and has a
(bounded) left inverse.

Indeed, Bers showed the above theorem for the projection Φ : Belt(D, Γ) → T (D∗ , Γ)
for an arbitrary Fuchsian group Γ. In particular, T (D∗ , Γ) is shown to be an open subset
of B2 (D∗ , Γ).

3.4. Convexity. Krushkal [57] proved that the Bers embedding T (D∗ ) of the universal
Teichmüller space is not starlike with respect to any point, and hence, not convex in
B2 (D∗ ). For non-starlikeness of general Teichmüller spaces, see Krushkal [60] and Toki
[111].
In spite of the above fact, the (Bers embededing of the) Teichmüller spaces enjoy many
kinds of convexity properties. We briefly list some of them in this subsection.
The most useful is perhaps the following “disk convexity” due to Zhuravlev [117], which
is shown as an application of the Grunsky inequality. A weaker version can be proved
also by the λ-lemma (see [102]).

Theorem 3.4.1 (Zhuravlev). Let Γ be a Fuchsian group acting on D∗ . Suppose that a


continuous map α : D → B2 (D∗ , Γ) is holomorphic in D and satisfies α(∂D) ⊂ S(D∗ ).
276 T. Sugawa IWQCMA05

Then α(D) ⊂ S(D∗ , Γ). Furthermore, if α(D)∩T (D∗ ) is non-empty, then α(D) ⊂ T (D∗ , Γ).

Outline of the proof. For each z ∈ D, there exists a unique Fz ∈ Σ0 such that SFz = α(z).
Let B(ℓ2 ) denote the complex Hilbert space consisting of bounded linear operators on ℓ2 .
Then the map β : D → B(ℓ2 ) defined by z 7→ G[Fz ] turns to be holomorphic. Then the
(generalized) maximum principle implies that
sup kβ(z)k = sup kβ(z)k ≤ 1
z∈D z∈∂D

and that either kβ(z)k < 1 for all z ∈ D or else kβ(z)k = 1 for all z ∈ D. Theorem
1.5.1 now yields that α(D) ⊂ S(D∗ ). If we assume that α(z0 ) ∈ T (D∗ ) for some point
z0 ∈ D in addition, then kβ(z0 )k < 1 and thus kβ(z)k < 1 for all z ∈ D. This means that
α(D) ⊂ T (D∗ ) ∩ B2 (D∗ , Γ) = T (D∗ , Γ).

We remark that the above argument is a variant of Lehto’s principle (see [13] or [67]).
A more sophisticated application of Grunsky inequality to Teichmüller spaces can be
found in [96].
Bers and Ehrenpreis [17] proved that finite dimensional Teichmüller spaces are holo-
morphically convex. Krushkal [58] strengthened it by showing that the Teichmüller space
of an arbitrary Riemann surface R is complex hyperconvex, that is to say, there exists a
negative plurisubharmonic function u(x) on Teich(R) such that u(x) → 0 when x tends to
∞. He proved it by pointing out that the function log tanh(d(x, y)) gives the Green func-
tion on Teich(R), where d(x, y) denotes the Teichmüller distance of Teich(R). Krushkal
[59] also proved that finite dimensional Teichmüller spaces are polynomially convex.

3.5. Teichmüller distance and other natural distances (metrics). In §2, we de-
fined two kinds of distances on the universal Teichmüller space; the Teichmüller distance
and the distance induced by the Bers embedding. These distances can be defined for the
Teichmüller space of an arbitrary Riemann surface. On the other hand, since Teichmüller
spaces have complex structure, it carries natural invariant distances for holomorphic maps
(see [46] or [52] as a general reference).
Let X be a complex (Banach) manifold. The Kobayashi pseudo-distance dK (x, y) is
defined as
N
X
inf dD (zj−1 , zj ),
j=1

where the infimum is taken over all finitely many holomorphic maps fj : D → X (j =
1, . . . , N ) which satisfy fj (zj ) = fj+1 (zj )(1 < j < N ), f1 (z0 ) = x, and fN (zN ) = y. Here,
The universal Teichmüller space and related topics 277

dD (z, w) denotes the Poincaré distance of D:



w−z
dD (z, w) = arctanh .
1 − z̄w

The following theorem was proved by Royden [91] for finite dimensional case and by
Gardiner (see [35] or [36]) for general case. (For a simple proof using the λ-lemma, see
[32].)

Theorem 3.5.1. The Kobayashi pseudo-distance of the Teichmüller space of a Riemann


surface is equal to the Teichmüller distance.

For other invariant metrics on Teichmüller spaces, see [75, Appendix 6].
Earle [31] proved that the Carathéodory (pseudo)distance of the Teichmüller space of
an arbitrary Fuchsian group is complete.
The Weil-Petersson metric is another important (Riemannian) metric on finite dimen-
sional Teichmüller spaces. Since the complex structure of the Teichmüller space of a
general Riemann surface is modelled on a complex Banach space which may not be re-
flexive, this metric cannot be defined on general Teichmüller spaces unless the structure
of the space is changed. However, some attempts were made to construct analogs of the
Weil-Petersson metric on the universal Teichmüller space, see [76], [77], [107], [108].

4. Pre-Schwarzian models
The Schwarzian derivative plays an important role in the definition of the Teichmüller
space. But, it is not easy to treat with Schwarzian derivative, in general, because of its
complicated form. Therefore, some attempts of replacing Schwarzian by pre-Schwarzian
have been made. See [116] and [6]. Though the pre-Schwarzian model is sometimes
called “poor man’s model” (cf. [43]) since it does not have much invariance, this model is
interesting in connection with geometric function theory.
When dealing with pre-Schwarzian derivative, the point at infinity plays a special role.
Therefore, we have to consider the case ∞ ∈ D separately.

4.1. The models T̂ (D) and T̂ (H). Let ∆ be a disk or a half-plane in C. Set

Ŝ(∆) = {Tf : f : ∆ → C is holomorphic and univalent}


and
b
T̂ (∆) = {Tf : f : ∆ → C is holomorphic and extends to a qc map of C}.

Here, Tf denotes the pre-Schwarzian derivative of f (see §1.6). By definition, T̂ (∆) ⊂


Ŝ(∆).
278 T. Sugawa IWQCMA05

We recall that the pre-Schwarzian derivative vanishes only when the function is affine.
Since each circle domain in C is similar (affinely equivalent) to either the unit disk D or
the half-plane H = {z ∈ C : Im z > 0}. Therefore, we may restrict ourselves on the two
cases ∆ = D and H. First let f ∈ S . By the well-known inequality (cf. [29])

(4.1.1) (1 − |z|2 )Tf (z) − 2z̄ ≤ 4,

we obtain kTf kB1 (D) ≤ 6. In particular, Ŝ(D) ⊂ B1 (D). Note also that the constant 6 is
sharp as the Koebe function K(z) = z/(1 − z)2 shows. It is easy to see that Ŝ(D) is closed
in B1 (D).

Let L(z) = (z − i)/(z + i). Note that kTL kB1 (H) = 4 and hence TL ∈ T̂ (H). Since
L∗1 : B1 (D) → B1 (H) is a linear isometry and Tf ◦L = L∗1 (Tf ) + TL , the space T̂ (H) is
contained in B1 (H) and it is isometrically equivalent to T̂ (D). In this sense, it is enough
to consider only T̂ (D).

We define the map π : B1 (D) → B2 (D) by π(ψ) = ψ ′ − ψ 2 /2. By definition, π(Ŝ(D)) =


S(D) and π(T̂ (D)) = T (D). Duren, Shapiro and Shields [30] noticed that this map is
continuous (see also §5.3).
Astala and Gehring [6] proved an analogous result to the case of Schwarzian derivative.

Theorem 4.1.2. The interior of Ŝ(D) in B1 (D) is equal to T̂ (D), while the closure of
T̂ (D) in B1 (D) is not equal to Ŝ(D). Moreover, ∂T (D) \ π(∂ T̂ (D)) is not empty.

Zhuravlev [116] revealed the following remarkable property of T̂ (D).

Theorem 4.1.3 (Zhuravlev). The space T̂ (D) decomposes into the uncountably many
connected components T̂0 and T̂ω , ω ∈ ∂D, where
T̂0 = {Tf ∈ T̂ (D) : f (D) is bounded } and T̂ω = {Tf ∈ T̂ (D) : f (z) → ∞ as z → ω}.

Moreover, {ψ ∈ B1 (D) : kψ − ψω kB1 (D) < 1} ⊂ T̂ω holds for each ω ∈ ∂D, where ψω (z) =
2ω̄/(1 − ω̄z) is the pre-Schwarzian derivative of the function z/(1 − ω̄z).

Note that the map π is not injective even in each connected component of T̂ (D).
Therefore, we should note that this model of the universal Teichmüller space has some
redundancy.

4.2. The model T̂ (D∗ ). There is some subtlety in consideration of the pre-Schwarzian
model of the universal Teichmüller space T̂ (D∗ ) on the exterior D∗ of the unit circle. The
The universal Teichmüller space and related topics 279

first thing to note is the fact that the Banach space B1 (D∗ ) is not the right space on which
T̂ (D∗ ) is modeled. We define
Ŝ(D∗ ) = {TF : F ∈ Σ}
and
b
T̂ (D∗ ) = {TF : F ∈ Σ extends to a quasiconformal map of C}.
If F (ζ) = ζ + b0 + b1 /ζ + b2 /ζ 2 + . . . , then TF (ζ) = 2b1 /ζ 3 + · · · = O(ζ −3 ) as ζ → ∞.
Therefore, the norm
(4.2.1) B(ψ) = sup (|ζ|2 − 1)|ζψ(ζ)|
ζ∈D∗

is more natural. Indeed, Becker’s univalence criterion [12] and Avhadiev’s inequality [8]
′′
ζF (ζ)
(4.2.2) (|ζ| − 1) ′
2 ≤6
F (ζ)
imply the following result.

Theorem 4.2.3. If a meromorphic function F (ζ) = ζ + b0 + b1 /ζ + . . . in |ζ| > 1 satisfies


B(TF ) ≤ 1, then F ∈ Σ. Conversely, every function F in Σ satisfies B(TF ) ≤ 6.

We set
B1′ (D∗ ) = {ψ ∈ B1 (D∗ ) : lim ζ 2 ψ(ζ) = 0}.
ζ→∞

Then, it is easy to see that B1′ (D∗ ) ∗


= {ψ : D → C holomorphic and B(ψ) < ∞}. The
above theorem now yields that Ŝ(D∗ ) is a bounded subset of B1′ (D∗ ).
We define π : B1′ (D∗ ) → B2 (D∗ ) as before by π(ψ) = ψ ′ − ψ 2 /2. Then π is continuous
[13, Lemma 6.1]. By definition, π(Ŝ(D∗ )) = S(D∗ ) and π(T̂ (D∗ )) = T (D∗ ). Since T (D∗ )
is an open set and T̂ (D∗ ) = π −1 (T (D∗ )), the set T̂ (D∗ ) is also open in B1′ (D∗ ). In this
way, we see that the space T̂ (D∗ ) is a complex Banach manifold modeled on B1′ (D∗ ). We
remark that π does not map B1 (D∗ ) into B2 (D∗ ).
The set T̂ (D∗ ) seems to be less investigated, but could be more useful. For instance,
the mapping F 7→ TF sends Σ0 to Ŝ(D∗ ) bijectively. Recall that the mapping F 7→ SF
sends Σ0 to S(D∗ ) bijectively. Therefore, the mapping π sends Ŝ(D∗ ) to S(D∗ ) bijectively.

4.3. Loci of typical subclasses of S . Since the differential operator Tf is closely


related with geometric function theory, many classical subclasses of univalent functions
correspond to sets with nice properties in Ŝ(D).
We recall several fundamental classes in univalent function theory. We denote by A the
set of analytic functions f in the unit disk D so normalized that f (0) = 0 and f ′ (0) = 1.
A function f ∈ A is called starlike (convex) if f is univalent and if f (D) is starlike with
280 T. Sugawa IWQCMA05

respect to the origin (convex). We denote by S ∗ and K the sets of starlike and convex
functions in A , respectively. A function f ∈ A is called close-to-convex if eiα f ′ /g ′ has
positive real part in D for a convex function g and for a real constant α. Denote by C the
set of close-to-convex functions in A . It is known that C ⊂ S (cf. [29]).
It is interesting to see how pre-Schwarzians of those functions are located in the space
Ŝ(D). The following result gives an answer to this question.

Theorem 4.3.1 ([27], [51]). {Tf : f ∈ K } and {Tf : f ∈ C } are both convex subsets of
Ŝ(D).

It may be natural to conjecture the following.

Conjecture 4.3.2 ([48]). The subset {Tf : f ∈ S ∗ } of Ŝ(D) is starlike with respect to
the origin.

Note that the vector operations in B1 (D) is translated to the Hornich operations in
the space of uniformly locally univalent functions (see, for example, [48]). Also, see Casey
[22] for relations between subclasses of S and (the closure) of T̂ (D).

5. Univalence criteria
As is well developed in Lehto’s textbook [67], univalence criteria are closely connected
with the universal Teichmüller space. The present section will be devoted to this topic.

5.1. Univalence criteria due to Nehari and Ahlfors-Weill. Nehari [78] proved the
following result, which is fundamental in the Teichmüller spaces.

Theorem 5.1.1. Every meromorphic univalent function f on the unit disk satisfies
the inequality kSf kB2 (D) ≤ 6. Conversely, if a meromorphic function f on the unit disk
satisfies the inequality kSf kB2 (D) ≤ 2, then f must be univalent.

The constants 6 and 2 are sharp since the Koebe function K(z) = z/(1 − z)2 satisfies
kSK kB2 (D) = 6 and since the function f (z) = ((1 + z)/(1 − z))iǫ , ǫ > 0, is never univalent
but kSf kB2 (D) = 2(1 + ε2 ) can approach 2 (Hille [44]). The former assertion was first
proved by Kraus [56] and reproved by Nehari. Therefore, it is called nowadays the Kraus-
Nehari theorem. The Kraus-Nehari theorem is a consequence of the Bieberbach theorem.
By the Möbius invariance of (1 − |z|2 )2 |Sf (z)|, it is enough to show the inequality only at
the origin, namely, |Sf (0)| ≤ 6 for f ∈ S . A straightforward computation gives Sf (0) =
6(a3 − a22 ) for f (z) = z + a2 z 2 + a3 z 3 + . . . . If we set F (ζ) = 1/f (1/ζ) = ζ + b0 + b1 /ζ + . . . ,
then b1 = a22 − a3 , and thus the inequality |b1 | ≤ 1 (see (1.5.3)) implies the required one.
The universal Teichmüller space and related topics 281

The class N = {f ∈ A : kSf kB2 (D) ≤ 2} is sometimes called the Nehari class. Gehring
and Pommerenke [41] showed that f ∈ N maps the unit disk conformally onto a Jordan
domain unless f (D) is Möbius equivalent to the parallel strip {z : |Im z| < π/4}. For
further development, see [24], [25] and [26].
In connection with Nehari’s theorem, Ahlfors and Weill established the following quasi-
conformal extension criterion. For ϕ ∈ B2 (D∗ ) with kϕkB2 (D∗ ) < 2, we set α[ϕ] ∈ Belt(D)
by α[ϕ](z) = −ρD (z)−2 ϕ(1/z̄)z̄ −4 /2. Note that the map α is the restriction of a bounded
linear operator which maps B2 (D∗ , Γ)2 = {ϕ ∈ B2 (D∗ , Γ) : kϕkB2 (D) < 2} into Belt(D∗ , Γ)
for every Fuchsian group Γ.

Theorem 5.1.2 (Ahlfors-Weill). The map α : B2 (D∗ )2 → Belt(D) is the local inverse
of the Bers projection Φ : Belt(D) → T (D∗ ), in other words, Φ(α[ϕ]) = ϕ for ϕ ∈ B2 (D∗ )
with kϕkB2 (D∗ ) < 2.

Corollary 5.1.3. The universal Teichmüller space T(D∗ ) contains the open ball centered
at the origin with radius 2 in B2 (D∗ ).

The map α : B2 (D∗ )2 → Belt(D) is sometimes called the Ahlfors-Weill section.

5.2. Inner radius and outer radius. Let D be a hyperbolic domain in C. b The inner
radius σI (D) and the outer radius σO (D) of univalence is defined respectively by
σI (D) = sup{σ ≥ 0 : kSf kB2 (D) ≤ σ ⇒ f is univalent in D},
b is univalent}.
σO (D) = sup{kSf kB2 (D) : f : D → C

We also define the number τ (D) ∈ [0, +∞] as kSp kB2 (D) , where p is a holomorphic universal
covering projection of D onto D. The quantity τ (D) is independent of the choice of p and
thus well defined. Note that τ (D) < ∞ if and only if ∂D is uniformly perfect (cf. [87] or
[103]).
Summarizing theorems of Ahlfors [2], Gehring [38], Nehari [78], we obtain the following.

Theorem 5.2.1. σI (∆) = 2, σO (∆) = 6, τ (∆) = 0 hold for a circle domain ∆. Let D be
a simply connected hyperbolic domain. Then σO (∆) ≤ 12 and τ (D) ≤ 6. Moreover, D is
a quasidisk if and only if σI (D) > 0.

b be univalent and set


The inequality σO (∆) ≤ 12 is shown as follows. Let f : D → C
Ω = f (D). Take a conformal map g : D∗ → D and set h = f ◦ g. Then, by Lemmas 1.3.1,
1.6.2 and the Kraus-Nehari theorem, we obtain
kSf kB2 (D) = kg2∗ (Sf )kB2 (D∗ ) = kSh − Sg kB2 (D∗ ) ≤ kSh kB2 (D∗ ) + kSg kB2 (D∗ ) ≤ 12.
282 T. Sugawa IWQCMA05

It is a remarkable fact due to Beardon and Gehring [10] that σO (D) ≤ 12 holds even for
an arbitrary hyperbolic domain D.
The inner and outer radii of univalence are better understood in the context of (quasi-)
Teichmüller space.

Lemma 5.2.2. Let g : D∗ → D be a conformal homeomorphism of D∗ onto a simply


connected hyperbolic domain D. Then {ϕ ∈ S(D∗ ) : kϕ−Sg k < σI (D)} is the maximal open
ball centered at Sg contained in T (D∗ ). On the other hand, σO (D) = max{kϕ − Sg kB2 (D∗ ) :
ϕ ∈ S(D∗ )}.

Lehto [65] proved the following relations.

Theorem 5.2.3. The relation σO (D) = τ (D) + 6 holds for a simply connected hyperbolic
domain D. Furthermore, 2 − τ (D) ≤ σI (R) ≤ min{2, 6 − τ (D)}.

As for the quantity τ (D), the following are known. For a convex domain D, we have
τ (D) ≤ 2. This result is repeatedly re-discovered by many mathematicians; [85], [90],
[112], [79], [65]. Suita [106] refined this result by showing the sharp inequality
(
2, 0 ≤ α ≤ 1/2,
τ (f (D)) ≤
8α(1 − α), 1/2 ≤ α ≤ 1

for a convex function f ∈ K of order α, namely, when Re (1 + zf ′′ (z)/f ′ (z)) > α.


It is known that τ (D) ≤ 6(K 2 − 1)/(K 2 + 1) for a K-quasidisk D (see [67]). See also
[68], [53], [23], [72], [9] for other classes of domains.
It is not easy to determine, or even to estimate from below, the value of σI (D), in
general. Known examples are sectors [62], triangles [64], the interiors and the exteriors
of regular polygons [21], [64], some other polygonal domains [73], [74], the exteriors of
hyperbolas [63].
For a general method of estimating σI (D) from below, see [66], [67] and [104]. See also
[105].

5.3. Pre-Schwarzian counterpart. One can define quantities similarly as in the pre-
vious section with respect to pre-Schwarzian derivative. We add the symbol ˆ to indicate
it. For instance,
σ̂I (D) = sup{σ ≥ 0 : kTf kB1 (D) ≤ σ ⇒ f is univalent in D}

for a hyperbolic domain D in C. In the case when D = D∗ , we adopt the norm B(ψ) :
σ̂I (D∗ ) = sup{σ ≥ 0 : B(TF ) ≤ σ ⇒ f is univalent in D∗ }.
The universal Teichmüller space and related topics 283


Duren, Shapiro and Shields [30] proved that σ̂I (D) ≥ 2( 5−2) = 0.472 · · · by observing
that kψ ′ kB2 (D) ≤ 4kψkB1 (D) and thus π(ψ) = ψ ′ − ψ 2 /2 is a continuous map of B1 (D) into
√ √
B2 (D). Note that Wirths [114] found the sharp constant C = (13 3 + 55 11)/64 =
3.20204 . . . for the estimate kψ ′ kB2 (D) ≤ CkψkB1 (D) . Nowadays, the best value for this
univalence criterion is known.
Theorem 5.3.1. σ̂I (∆) = 1 and σ̂O (∆) = 6 for ∆ = D, H and D∗ .

Becker [11], [12] showed that σ̂I (D) ≥ 1 and σ̂I (D∗ ) ≥ 1 and Becker-Pommerenke [14]
showed that equality hold for ∆ = D and that σ̂I (H) = 1. Pommerenke [88] showed the
sharpness for ∆ = D∗ .
By (4.1.1) and the fact that the Koebe function K satisfies kTK kB1 (D) = 6, we see that
σ̂O (D) = 6. σ̂O (H) = 6 can be seen by noting the relation
kψkB1 (H) = lim kψkB1 (∆r )
r→1−

for ψ ∈ B1 (H), where ∆r = {z : |z − i(1 + r2 )/(1 − r2 )| < 2r/(1 − r2 )}. The formula
σ̂O (D∗ ) = 6 follows from the fact that the inequality in (4.2.2) is sharp for each ζ.
For concrete estimates of τ̂ (D) for several geometric classes of domains, see [115], [101],
[81], [49], [50].
In spite of relative simplicity of the operation Tf , very little is known for quantities
σ̂I (D) and σ̂O (D). Stowe [97] gave non-trivial examples of domains D for which σ̂I (D) ≥ 1.

5.4. Directions of further investigation. The Bers embedding of Teichmüller spaces


is still mysterious. We know very little about the shape of it. Pictures of one-dimensional
Teichmüller spaces were recently given in [54] and [55]. Note that the first attempt towards
it was done by Porter [89] as early as in 1970’s.
It is an interesting and important problem to describe the intersection of T (∆) or T̂ (∆)
with a (complex) one-dimensional vector subspace of B2 (∆) or B1 (∆) for a circle domain.
Completely known examples are essentially, as far as the author knows, the linear hull of
1/(1 − z) in B1 (∆) [92] and the linear hull of z −2 in B2 (H) in [47], only.
The results presented above could be generalized to various directions. We end this
survey with remarks on possible ways to study furthermore.
In this section, we considered mainly the case when the domain is simply connected.
When the domain is multiply connected, the problem will become much more difficult.
See [83] and [84] for fundamental information.
We were concerned here with only pre-Schwarzian and Schwarzian derivatives. On the
other hand, several definitions of higher-order Schwarzian derivatives have been proposed
(e.g., [109], [93]). Thus, we may develop the theory for those higher-order Schwarzian
derivatives.
284 T. Sugawa IWQCMA05

Of course, we may consider domains in Cn or Rn but with great difficulty caused by


the lack of canonical metrics such as hyperbolic metric, the lack of Riemann mapping
theorem and so on. Note that Martio and Sarvas [70] gave some injectivity conditions
even in higher dimensions.

References
1. W. Abikoff, The Real Analytic Theory of Teichmüller Space, Lecture Notes in Mathematics, vol.
820, Springer-Verlag, 1976.
2. L. V. Ahlfors, Quasiconformal reflections, Acta Math. 109 (1963), 291–301.
3. , Lectures on Quasiconformal Mappings, van Nostrand, 1966.
4. L. V. Ahlfors and L. Bers, Riemann’s mapping theorem for variable metrics, Ann. of Math. (2) 72
(1960), 385–404.
5. K. Astala, Selfsimilar zippers, Holomorphic Functions and Moduli I (D. Drasin, C. J. Earle, F. W.
Gehring, I. Kra, and A. Marden, eds.), Springer-Verlag, 1988, pp. 61–73.
6. K. Astala and F. W. Gehring, Injectivity, the BM O norm and the universal Teichmüller space, J.
Analyse Math. 46 (1986), 16–57.
7. , Crickets, zippers, and the Bers universal Teichmüller space, Proc. Amer. Math. Soc. 110
(1990), 675–687.
8. F. G. Avhadiev, Conditions for the univalence of analytic functions (Russian), Izv. Vysš. Učebn.
Zaved. Matematika 1970 (1970), no. 11 (102), 3–13.
9. R. W. Barnard, L. Cole, K. Pearce, and G. B. Williams, The sharp bound for deformation of a disk
under a hyperbolically convex map, to appear in Proc. London Math. Soc.
10. A. F. Beardon and F. W. Gehring, Schwarzian derivatives, the Poincaré metric and the kernel
function, Comment. Math. Helv. 55 (1980), 50–64.
11. J. Becker, Löwnersche Differentialgleichung und quasikonform fortsetzbare schlichte Funktionen, J.
Reine Angew. Math. 255 (1972), 23–43.
12. , Löwnersche Differentialgleichung und Schlichtheitskriterien, Math. Ann. 202 (1973), 321–
335.
13. , Conformal mappings with quasiconformal extensions, Aspects of Contemporary Complex
Analysis, Proc. Conf. Durham, 1979, Academic Press, 1980, pp. 37–77.
14. J. Becker and Ch. Pommerenke, Schlichtheitskriterien und Jordangebiete, J. Reine Angew. Math.
354 (1984), 74–94.
15. L. Bers, A non-standard integral equation with applications to quasiconformal mappings, Acta Math.
116 (1966), 113–134.
16. , On boundaries of Teichmüller spaces and on Kleinian groups: I, Ann. of Math. (2) 91
(1970), 570–600.
17. L. Bers and L. Ehrenpreis, Holomorphic convexity of Teichmüller spaces, Bull. Amer. Math. Soc.
70 (1964), 761–764.
18. A. Beurling and L. V. Ahlfors, The boundary correspondence for quasiconformal mappings, Acta
Math. 96 (1956), 125–142.
19. V. Božin and V. Marković, Distance between domains in the sense of Lehto is not a metric, Ann.
Acad. Sci. Fenn. Math. 24 (1999), 3–10.
20. J. Brock and K. Bromberg, On the density of geometrically finite Kleinian groups, Acta Math. 192
(2004), 33–93.
The universal Teichmüller space and related topics 285

21. D. Calvis, The inner radius of univalence of normal circular triangles and regular polygons, Complex
Variables Theory Appl. 4 (1985), 295–304.
22. S. D. Casey, The inclusion of classical families in the closure of the universal Teichmüller space,
Michigan Math. J. 39 (1992), 189–199.
23. Y. M. Chiang, Some remarks on Lehto’s domain constant, Ann. Acad. Sci. Fenn. Ser. A I Math. 17
(1992), 285–293.
24. M. Chuaqui and B. Osgood, Sharp distortion theorems associated with the Schwarzian derivative, J.
London Math. Soc. 48 (1993), 289–298.
25. M. Chuaqui, B. Osgood, and Ch. Pommerenke, John domains, quasidisks, and the Nehari class, J.
Reine Angew. Math. 471 (1996), 77–114.
26. M. Chuaqui and Ch. Pommerenke, Characteristic properties of Nehari functions, Pacific J. Math.
188 (1999), 83–94.
27. J. A. Cima and J. A. Pfaltzgraff, A Banach space of locally univalent functions, Michigan Math. J.
17 (1970), 321–334.
28. A. Douady and C. J. Earle, Conformally natural extensions of homeomorphisms of the circle, Acta
Math. 157 (1986), 23–48.
29. P. L. Duren, Univalent Functions, Springer-Verlag, 1983.
30. P. L. Duren, H. S. Shapiro, and A. L. Shields, Singular measures and domains not of Smirnov type,
Duke Math. J. 33 (1966), 247–254.
31. C. J. Earle, On the Carathéodory metric in Teichmüller spaces, Discontinuous groups and Riemann
surfaces (Proc. Conf., Univ. Maryland, College Park, Md., 1973), Princeton Univ. Press, 1974,
pp. 99–103. Ann. of Math. Studies, No. 79.
32. C. J. Earle, I. Kra, and S. L. Krushkal’, Holomorphic motions and Teichmüller spaces, Trans. Amer.
Math. Soc. 343 (1994), 927–948.
33. C. J. Earle and C. McMullen, Quasiconformal isotopies, Holomorphic Functions and Moduli I
(D. Drasin, C. J. Earle, F. W. Gehring, I. Kra, and A. Marden, eds.), Springer-Verlag, 1988,
pp. 143–154.
34. B. Flinn, Jordan domains and the universal Teichmüller space, Trans. Amer. Math. Soc. 282 (1984),
603–610.
35. F. P. Gardiner, Teichmüller Theory and Quadratic Differentials, Wiley-Intersciences, New York,
1987.
36. F. P. Gardiner and N. Lakic, Quasiconformal Teichmüller Theory, Amer. Math. Soc., 2000.
37. F. P. Gardiner and D. Sullivan, Symmetric structures on a closed curve, Amer. J. Math. 114 (1992),
683–736.
38. F. W. Gehring, Univalent functions and the Schwarzian derivative, Comment. Math. Helv. 52
(1977), 561–572.
39. , Spirals and the universal Teichmüller space, Acta Math. 141 (1978), 99–113.
40. , Characteristic Properties of Quasidisks, Les Presses de l’Université de Montréal, 1982.
41. F. W. Gehring and Ch. Pommerenke, On the Nehari univalence criterion and quasicircles, Comment.
Math. Helv. 1984 (59), 226–242.
42. G. M. Goluzin, Geometric theory of functions of a complex variable, American Mathematical Society,
Providence, R.I., 1969, Translations of Mathematical Monographs, Vol. 26.
43. D. H. Hamilton, The closure of Teichmüller space, J. Analyse Math. 55 (1990), 40–50.
44. E. Hille, Remarks on a paper by Zeev Nehari, Bull. Amer. Math. Soc. 55 (1949), 552–553.
45. Y. Imayoshi and M. Taniguchi, An introduction to Teichmüller spaces, Springer-Tokyo, 1992.
286 T. Sugawa IWQCMA05

46. M. Jarnicki and P. Pflug, Invariant Distances and Metrics in Complex Analysis, Walter de Gruyter,
Berlin-New York, 1993.
47. C. I. Kalme, Remarks on a paper by Lipman Bers, Ann. of Math. (2) 91 (1970), 601–606.
48. Y. C. Kim, S. Ponnusamy, and T. Sugawa, Mapping properties of nonlinear integral operators and
pre-Schwarzian derivatives, J. Math. Anal. Appl. 299 (2004), 433–447.
49. Y. C. Kim and T. Sugawa, Growth and coefficient estimates for uniformly locally univalent functions
on the unit disk, Rocky Mountain J. Math. 32 (2002), 179–200.
50. , Norm estimates of the pre-schwarzian derivatives for certain classes of univalent functions,
Proc. Edinburgh Math. Soc. 49 (2006), 1–13.
51. Y. J. Kim and E. P. Merkes, On certain convex sets in the space of locally schlicht functions, Trans.
Amer. Math. Soc. 196 (1974), 217–224.
52. S. Kobayashi, Hyperbolic manifolds and holomorphic mappings, second ed., World Scientific Pub-
lishing Co. Pte. Ltd., Hackensack, NJ, 2005, An introduction.
53. W. Koepf, Close-to-convex functions, univalence criteria and quasiconformal extensions, Ann. Univ.
Marie Curie-Sklodowska 40 (1986), 97–103.
54. Y. Komori and T. Sugawa, Bers embedding of the Teichmüller space of a once-punctured torus,
Conform. Geom. Dyn. 8 (2004), 115–142.
55. Y. Komori, T. Sugawa, M. Wada, and Y. Yamashita, Drawing Bers embeddings of the Teichmüller
space of once-punctured tori, to appear in Exper. Math.
56. W. Kraus, Über den Zusamenhang einiger Charakteristiken eines einfach zusammenhängenden Bere-
iches mit der Kreisabbildung, Mitt. Math. Sem. Giessen 21 (1932), 1–28.
57. S. L. Krushkal, On the question of the structure of the universal Teichmüller space, Soviet Math.
Dokl. 38 (1989), 435–437.
58. , The Green function of Teichmüller spaces with applications, Bull. Amer. Math. Soc. (N.S.)
27 (1992), 143–147.
59. , Polynomial convexity of Teichmüller spaces, J. London Math. Soc. (2) 52 (1995), 147–156.
60. , Teichmüller spaces are not starlike, Ann. Acad. Sci. Fenn. Ser. A I Math. 20 (1995), 167–
173.
61. R. Kühnau (ed.), Handbook of Complex Analysis: Geometric Function Theory, 2 vols., Elsevier
Science B.V., Amsterdam, 2002, 2005.
62. M. Lehtinen, On the inner radius of univalency for non-circular domains, Ann. Acad. Sci. Fenn.
Ser. A I Math. 5 (1980), 45–47.
63. , Estimates of the inner radius of univalency of domains bounded by conic sections, Ann.
Acad. Sci. Fenn. Ser. A I Math. 10 (1985), 349–353.
64. , Angles and the inner radius of univalency, Ann. Acad. Sci. Fenn. Ser. A I Math. 11 (1986),
161–165.
65. O. Lehto, Domain constants associated with Schwarzian derivative, Comment. Math. Helv. 52
(1977), 603–610.
66. , Remarks on Nehari’s theorem about the Schwarzian derivative and schlicht functions, J.
Analyse Math. 36 (1979), 184–190.
67. , Univalent Functions and Teichmüller Spaces, Springer-Verlag, 1987.
68. O. Lehto and O. Tammi, Schwarzian derivative in domains of bounded boundary rotation, Ann.
Acad. Sci. Fenn. Ser. A I Math. 4 (1978/79), 253–257.
69. O. Lehto and K. I. Virtanen, Quasiconformal Mappings in the Plane, 2nd Ed., Springer-Verlag,
1973.
The universal Teichmüller space and related topics 287

70. O. Martio and J. Sarvas, Injectivity theorems in plane and space, Ann. Acad. Sci. Fenn. Ser. A I
Math. 4 (1978/79), 383–401.
71. K. Matsuzaki, Simply connected invariant domains of Kleinian groups not in the closures of Te-
ichmüller spaces, Complex Variables Theory Appl. 22 (1993), 92–100.
72. D. Mejı́a and Ch. Pommerenke, On spherically convex univalent functions, Michigan Math. J. 47
(2000), 163–172.
73. L. Miller-Van Wieren, Univalence criteria for classes of rectangles and equiangular hexagons, Ann.
Acad. Sci. Fenn. A I Math. 22 (1997), 407–424.
74. , On Nehari disks and the inner radius, Comment. Math. Helv. 76 (2001), 183–199.
75. S. Nag, The Complex Analytic Theory of Teichmüller Spaces, Wiley, New York, 1988.
76. , Diff(S 1 ), Teichmüller space and period matrices: canonical mappings via string theory,
Rev. Roumaine Math. Pures Appl. 39 (1994), 789–827.
77. S. Nag and D. Sullivan, Teichmüller theory and the universal period mapping via quantum calculus
and the H 1/2 space on the circle, Osaka J. Math. 32 (1995), 1–34.
78. Z. Nehari, The Schwarzian derivative and schlicht functions, Bull. Amer. Math. Soc. 55 (1949),
545–551.
79. , A property of convex conformal maps, J. Analyse Math. 30 (1976), 390–393.
80. K. Ohshika, Realising end invariants by limits of minimally parabolic, geometrically finite groups,
arXive:math. GT/0504546 (2005).
81. Y. Okuyama, The norm estimates of pre-Schwarzian derivatives of spiral-like functions, Complex
Variables Theory Appl. 42 (2000), 225–239.
82. B. Osgood and D. Stowe, The Schwarzian distance between domains: A question of O. Lehto, Ann.
Acad. Sci. Fenn. Ser. A I Math. 12 (1987), 313–318.
83. B. G. Osgood, Univalence criteria in multiply-connected domains, Trans. Amer. Math. Soc. 260
(1980), 459–473.
84. , Some properties of f ′′ /f ′ and the Poincaré metric, Indiana Univ. Math. J. 31 (1982),
449–461.
85. V. Paatero, Über die konforme Abbildung von Gebieten, deren Ränder von beschränker Drehung
sind, Ann. Acad. Sci. Fenn. A I Math.-Phys. XXXIII:9 (1931), 1–79.
86. Ch. Pommerenke, Univalent Functions, Vandenhoeck & Ruprecht, Göttingen, 1975.
87. Ch. Pommerenke, Uniformly perfect sets and the Poincaré metric, Arch. Math. 32 (1979), 192–199.
88. , On the Becker univalence criterion, Ann. Univ. Mariae Curie-Sklodowska Sect. A 36/37
(1982/83), 123–124 (1985).
89. R. M. Porter, Computation of a boundary point of Teichmüller space, Bol. Soc. Mat. Mexicana 24
(1979), 15–26.
90. M. S. Robertson, Univalent functions f (z) for which zf ′ (z) is spirallike, Michigan Math. J. 16
(1969), 97–101.
91. H. L. Royden, Automorphisms and isometries of Teichmüller space, Advances in the Theory of
Riemann Surfaces, Ann. of Math. Stud. 66, Princeton Univ. Press, 1971, pp. 369–383.
92. W. C. Royster, On the univalence of a certain integral, Michigan Math. J. 12 (1965), 385–387.
93. E. Schippers, Distortion theorems for higher order Schwarzian derivatives of univalent functions,
Proc. Amer. Math. Soc. 128 (2000), 3241–3249.
94. M. Seppälä and T. Sorvali, Geometry of Riemann Surfaces and Teichmüller Spaces, North-Holland,
Amsterdam, 1992.
288 T. Sugawa IWQCMA05

95. H. Shiga, Characterization of quasi-disks and Teichmüller spaces, Tôhoku Math. J. 37 (1985), 541–
552.
96. H. Shiga and H. Tanigawa, Grunsky’s inequality and its applications to Teichmüller spaces, Kodai
Math. J. 16 (1993), 361–378.
97. D. Stowe, Injectivity and the pre-Schwarzian derivative, Michigan Math. J. 45 (1998), 537–546.
98. T. Sugawa, The Bers projection and the λ-lemma, J. Math. Kyoto Univ. 32 (1992), 701–713.
99. , On the Bers conjecture for Fuchsian groups of the second kind, J. Math. Kyoto Univ. 32
(1992), 45–52.
100. , On the space of schlicht projective structures on compact Riemann surfaces with boundary,
J. Math. Kyoto Univ. 35 (1995), 697–732.
101. , On the norm of the pre-Schwarzian derivatives of strongly starlike functions, Ann. Univ.
Mariae Curie-Sklodowska, Sectio A 52 (1998), 149–157.
102. , Holomorphic motions and quasiconformal extensions, Ann. Univ. Mariae Curie-Sklodowska,
Sectio A 53 (1999), 239–252.
103. , Uniformly perfect sets: analytic and geometric aspects (Japanese), Sugaku 53 (2001), 387–
402, English translation in Sugaku Expo. 16 (2003), 225–242.
104. , A remark on the Ahlfors-Lehto univalence criterion, Ann. Acad. Sci. Fenn. A I Math. 27
(2002), 151–161.
105. , Inner radius of univalence for a strongly starlike domain, Monatsh. Math. (2003), 61–68.
106. N. Suita, Schwarzian derivatives of convex functions, J. Hokkaido Univ. Edu. (Sec. IIA) 46 (1996),
113–117.
107. L. A. Takhtajan and L.-P. Teo, Liouville action and Weil-Petersson metric on deformation spaces,
global Kleinian reciprocity and holography, Comm. Math. Phys. 239 (2003), 183–240.
108. , Weil-Petersson geometry of the universal Teichmüller space, Infinite dimensional algebras
and quantum integrable systems (Basel), Progr. Math., vol. 237, Birkhäuser, Basel, 2005, pp. 225–
233.
109. H. Tamanoi, Higher Schwarzian operators and combinatorics of the Schwarzian derivative, Math.
Ann. 305 (1995), 127–181.
110. W. P. Thurston, Zippers and schlicht functions, The Bieberbach conjecture, Proceedings of the
symposium on the occasion of the proof of the Bieberbach conjecture held at Purdue University, West
Lafayette, Ind., March 11–14, 1985 (Providence, RI) (D. Drasin, P. Duren, and A. Marden, eds.),
Mathematical Surveys and Monographs, vol. 21, American Mathematical Society, 1986, pp. 185–187.
111. M. Toki, On nonstarlikeness of Teichmüller spaces, Proc. Japan Acad. Ser. A Math. Sci. 69 (1993),
58–60.
112. S. Y. Trimble, A coefficient inequality for convex univalent functions, Proc. Amer. Math. Soc. 48
(1975), 266–267.
113. A. J. Tromba, Teichmüller Theory in Riemannian Geometry, Birkhäuser, Basel, 1992.
114. K.-J. Wirths, Über holomorphe Funktionen, die einer Wachstumsbeschränkung unterliegen, Arch.
Math. 30 (1978), 606–612.
115. S. Yamashita, Norm estimates for function starlike or convex of order alpha, Hokkaido Math. J. 28
(1999), 217–230.
116. I. V. Zhuravlev, Model of the universal Teichmüller space, Siberian Math. J. 27 (1986), 691–697.
117. I. V. Žuravlev, Univalent functions and Teichmüller spaces, Soviet Math. Dokl. 21 (1980), 252–255.
The universal Teichmüller space and related topics 289

Toshiyuki Sugawa Address: Department of Mathematics, Graduate School of Science, Hiroshima


University, Higashi-Hiroshima, 739-8526 Japan
E-mail: sugawa@math.sci.hiroshima-u.ac.jp
Proceedings of the International Workshop on Quasiconformal
Mappings and their Applications (IWQCMA05)

Metrics and quasiregular mappings

Matti Vuorinen

Abstract. This series of lectures intends to provide a gateway to some se-


lected topics of quasiconformal and quasiregular maps, in particular to the
main themes of [Vu1] and [Vu3]. Some of the basic notions and tools are
briefly reviewed. Several problems, exercises and open problems are given
throughout the text. At the end of the paper a short list of some generic
open problems is presented for metric spaces, which allow a great number of
variations in specific cases.

Keywords. quasiregular mappings, metric spaces.


2000 MSC. 30C65.

Contents

1. Introduction 291
2. Möbius transformations 295
3. Hyperbolic geometry 300
4. Quasihyperbolic geometry 305
5. Modulus and capacity 306
6. Conformal invariants 313
7. Distortion theory 319
8. Open problems 322
References 323

1. Introduction
The goal of these lectures is to provide an introduction to some of the main
properties of quasiconformal and quasiregular mappings. One of the central
themes here will be to study how these mappings deform distances and metrics
and therefore it is natural to study our mappings between metric spaces. In most
cases, the metrics will have some useful invariance or quasi-invariance properties
under a set Γ of transformations, called rigid motions. An important example is
the unit ball of Rn equipped with the hyperbolic metric in which case we may
take the set Γ to be the group of the Möbius self-mappings of the ball.
Version October 19, 2006.
292 M. Vuorinen IWQCMA05

The material is largely drawn from [Vu3] and [AVV2]. In order to give the
reader a chance to enter gradually this territory of mathematical research, prob-
lems of varying level are given, from easy exercises to research problems. Many
more can be found in [Vu3] and [AVV2] (the exercises in [AVV2] come with so-
lutions). Some research problems are collected at the end of the paper. Because
of limitations of space, most of the details/proofs are omitted with the general
reference to [V1] and [Vu3].
The idea of using invariance with respect to rigid motion to study function
theory is very old. In fact, it can be traced back to nineteenth century, in
particular, to the work of F. Klein. Perhaps the most natural notion of invariance
is conformal invariance under the group of conformal self-maps of a given simply-
connected domain. Several conformal invariants emerged from the studies of H.
Grötzsch, L. Ahlfors, and A. Beurling.
A pair (X, d) is called a metric space if X 6= ∅ and d : X ×X → [0, ∞) satisfies
the following four conditions


 (M1) d(x, y) ≥ 0 for all x, y ∈ X ,

(M2) d(x, y) = 0 iff x = y ,
(1.1)

 (M3) d(x, y) = d(y, x) for all x, y ∈ X ,
 (M4) d(x, y) ≤ d(x, z) + d(z, y) for all x, y, z ∈ X .

Let (X, d1 ) and (Y, d2 ) be metric spaces and let f : X → Y be a continuous


mapping. Then we say that f is uniformly continuous if there exists an increasing
continuous function ω : [0, ∞) → [0, ∞) with ω(0) = 0 and d2 (f (x), f (y)) ≤
ω(d1 (x, y)) for all x, y ∈ X . We call the function ω the modulus of continuity of
f . If there exist C, α > 0 such that ω(t) ≤ Ctα for all t > 0 , we say that f is
Hölder-continuous with Hölder exponent α . If α = 1 , we say that f is Lipschitz
with the Lipschitz constant C or simply C-Lipschitz. If f is a homeomorphism
and both f and f −1 are C-Lipschitz, then f is C-bilipschitz or C-quasiisometry
and if C = 1 we say that f is an isometry. These conditions are said to hold
locally, if they hold for each compact subset of X .
1.2. Exercise. If h : [0, ∞) → [0, ∞) is a function and h(t)/t is decreasing,
show that h(x + y) ≤ h(x) + h(y) for all x, y > 0. In particular, show that if
(X, d) is a metric space, then also (X, dα ), α ∈ (0, 1), is.
1.3. Exercise. Let f : [0, ∞) → [0, ∞) be Hölder-continuous with exponent
β > 1. Show that f (x) = f (0) for all x > 0 .
1.4. Example. Let f : Rn → Rn , f (x) = |x|α−1 x, f (0) = 0. Then f is Hölder-
continuous with exponent α .

In most examples below, the metric spaces will have some additional structure.
The metrics will often have some quasiinvariance properties. For instance, we
say that a pair of metric spaces (Xj , dj ), j = 1, 2, is quasiinvariant under a set
Γ of mappings f : (X1 , d1 ) → (X2 , d2 ) if there exists C ≥ 1 such that 1/C ≤
d2 (f (x), f (y))/d1 (x, y) ≤ C for all x, y ∈ X1 , x 6= y and all f ∈ Γ . In particular,
Metrics and quasiregular mappings 293

we will study metric spaces (X, d) where the group Γ of automorphisms of X


acts transitively (i.e. given x, y ∈ X there exists h ∈ Γ such that hx = y .) If
C = 1, then we say that d is invariant.
1.5. Examples.
Pn (1) The euclidean space Rn equipped with the usual metric
|x − y| = ( j=1 (xj − yj )2 )1/2 , Γ is the group of translation in Rn .
(2) The unit sphere S n = {z ∈ Rn+1 : |z| = 1} equipped with the metric of
n+1
R and Γ is the set of rotations of S n .
(3) Let G ⊂ Rn , G 6= Rn , for x, y ∈ G set
 
|x − y|
jG (x, y) = log 1 + .
min{d(x, ∂G), d(y, ∂G)}
Then one can prove that jG is a metric (this is a folklore result, see e.g. [S]). In
fact, there exists a constant C > 1 such that for the unit ball Bn of Rn
1/C ≤ ρBn (x, y)/jBn (x, y) ≤ C
for all x, y ∈ Bn , x 6= y . Here ρBn is the hyperbolic metric of Bn and it is
invariant under the group of Möbius self-mappings of Bn . For the definition of
ρBn see below or [Vu3, Section2].

A basic geometric object of a metric space (X, d) is the ball BX (z, r) = {x ∈


X : d(x, z) < r} . In order to study how balls and their boundary spheres are
deformed under homeomorphisms, we introduce a deformation measure Hf (x0 , r)
of a ball under a homeomorphism f : (X1 , d1 ) → (X2 , d2 ) at a point x ∈ X1
 
d2 (f (x), f (y))
Hf (x, r) = sup : d1 (z, x) = d1 (y, x) = r .
d2 (f (x), f (z))

r
Lr
x lr
f(x)

Figure 1. Hf (x, r) .

If f maps spheres centered at x onto spheres centered at f (x) , then Hf (x, r) =


1. For instance the above radial mapping x 7→ |x|α−1 x has the property Hf (0, r) =
1 for all r > 0 . Recall from Complex Analysis that a conformal map f : D1 →
D2 , Dj ⊂ C, j = 1, 2, satisfies limr→0 Hf (x, r) = 1 for all x ∈ D1 . We say that
294 M. Vuorinen IWQCMA05

a homeomorphism f : (X1 , d1 ) → (X2 , d2 ) is quasiconformal (with respect to


(d1 , d2 )), if there exists C ∈ [1, ∞) such that for all x ∈ X1
Hf (x) = lim sup Hf (x, r) ≤ C .
r→0

If f is L-bilipschitz, then f satisfies the above condition with C = L2 .


Let Gj ⊂ Rn , j = 1, 2, be domains and let f : G1 → G2 be a homeomorphism.
Suppose now that there exists a constant C ≥ 1 such that for all subdomains
D ⊂ G1 the mapping f |D : (D, jD ) → (f (D), jf (D) ) is C-Lipschitz. Fix x0 ∈ G1
and r ∈ (0, d(x0 , ∂G1 )/2) . If |x − x0 | = |y − x0 | = r and G = B n (x0 , 2r) \ {x0 } ,
then jG (x, y) ≤ log 3 and we obtain by the above C-Lipschitz-property


log |f (x) − f (x0 )| ≤ jf G (f (x), f (y)) ≤ CjG (x, y) ≤ C log 3 ,
|f (y) − f (x0 )|
and hence Hf (x0 ) ≤ 3C , where we used the triangle inequality (Lemma 3.21 (3)
below) and the fact that x0 ∈ ∂G . Now d1 and d2 are the usual metrics. Thus
we see that our map is quasiconformal.
In this argument the fact that x0 ∈ ∂G played a key role. For most of
the metrics that we will consider here, even one single boundary point will be
important. Most of the metrics will also be monotone with respect to the domain.
Thus, for instance, if G1 ⊂ G2 ⊂ Rn are domains, then jG1 (x, y) ≥ jG2 (x, y) for
all x, y ∈ G1 and for a fixed x0 ∈ G1 , jG1 (x0 , x) → ∞ as x → x1 ∈ ∂G . In
the above argument, it was assumed that f |D is Lipschitz continuous for all
subdomains D of G1 but we only used this property for subdomains of the form
B n (x, r) \ {x} , x ∈ G1 . In order to motivate this condition let us recall that a
conformal map is conformal also in every subdomain.
Here we have studied the metric j , mainly because it is very easy to define
and because it well represents the metrics we study here. There are now several
questions:
(a) Can we characterize the class of quasiisometries or isometries in the above
sense?
(b) Can we prove similar results for other metrics (and what are these met-
rics)?
(c) Can we say more for the case when the domains are ”nice” (for instance
quasidisks)?
Conformal invariants and conformally invariant metrics have been an impor-
tant topic in geometric function theory during the past century. One of the first
promoters of these ideas was F. Klein. In the context of quasiconformal map-
pings these ideas emerged as a result of the pioneering works of H. Grötzsch, O.
Teichmüller, L. Ahlfors and A. Beurling on quasiconformal maps in plane do-
mains [LV], [K]. Extension to higher dimensions is due to F.W. Gehring and J.
Väisälä [G] , [V1]. The case of metric measure spaces has been studied recently
by J. Heinonen, P. Koskela and many other people [H].
Metrics and quasiregular mappings 295

In the setup presented here, the aforementioned questions (a)-(c) were studied
already in [Vu1] and [Vu3]. But only very few answers are known, see [H1], [H2],
[HI]. These questions could also be investigated for some particular classes of do-
mains, which would bring a very wide spectrum of new questions into play. Some
examples of such classes of domains would be uniform domains and quasiconfor-
mal balls. As we will see, there still are numerous open problems in this area.
It is assumed that the reader is familiar with some basic facts and definitions of
the theory of quasiconformal and quasiregular maps [V1], [Vu3].

2. Möbius transformations
For x ∈ Rn and r > 0 let
B n (x, r) = { z ∈ Rn : |x − z| < r },
S n−1 (x, r) = { z ∈ Rn : |x − z| = r }
denote the ball and sphere, respectively, centered at x with radius r. The abbre-
viations B n (r) = B n (0, r), S n−1 (r) = S n−1 (0, r), Bn = B n (1), S n−1 = S n−1 (1)
will be used frequently. For t ∈ R and a ∈ Rn \ {0} we denote
P (a, t) = { x ∈ Rn : x · a = t } ∪ {∞}.
n
Then P (a, t) is a hyperplane in R = Rn ∪ {∞} perpendicular to the vector a,
at distance t/|a| from the origin.
2.1. Definition. Let D and D′ be domains in Rn and let f : D → D′ be a
homeomorphism. We call f conformal if (1) f ∈ C 1 , (2) Jf (x) 6= 0 for all
x ∈ D, and (3) |f ′ (x)h| = |f ′ (x)||h| for all x ∈ D and all h ∈ Rn . If D and
n
D′ are domains in R , we call a homeomorphism f : D → D′ conformal if the
restriction of f to D \ {∞, f −1 (∞)} is conformal.
2.2. Examples. Some basic examples of conformal mappings are the following
elementary transformations.
(1) A reflection in P (a, t):
a
f1 (x) = x − 2(x · a − t) , f1 (∞) = ∞ .
|a|2
(2) An inversion (reflection) in S n−1 (a, r):
r2 (x − a)
f2 (x) = a + , f2 (a) = ∞ , f2 (∞) = a .
|x − a|2
(3) A translation f3 (x) = x + a , a ∈ Rn , f3 (∞) = ∞.
(4) A stretching by a factor k > 0: f4 (x) = kx , f4 (∞) = ∞.
(5) An orthogonal mapping, i.e. a linear map f5 with
|f5 (x)| = |x| , f5 (∞) = ∞ .
296 M. Vuorinen IWQCMA05

2.3. Remarks. (1) The translation x 7→ x + a can be written as a composition


of reflections in P (a, 0) and P (a, 21 |a|2 ). The stretching x 7→ kx, k > 0, can be

written as a composition of inversions in S n−1 (0, 1) and S n−1 (0, k ). It can be
proved, that an orthogonal mapping can be composed of at most n+1 reflections
in planes (see [BE, p. 23, Theorem 3.1.3]).
(2) It is easy to show that f1 (f1 (x)) = x and f2 (f2 (x)) = x for all x ∈ Rn , i.e. f1
and f2 are involutions.
(3) It can also be shown that we have the difference formula
r2 |x − y|
|f2 (x) − f2 (y)| =
|x − a||y − a|
for all x, y ∈ Rn \ {a}.
(4) If an = 0, then one can show that f2 (Hn ) = Hn and that for all x, y ∈ Hn
|f2 (x) − f2 (y)|2 |x − y|2
= .
(f2 (x))n (f2 (y))n (x)n (y)n
(5) Reflections and inversions are sense-reversing. The composition of two sense-
reversing maps is sense-preserving.

n n
2.4. Definition. A homeomorphism f : R → R is called a Möbius transfor-
mation if f = g1 ◦ · · · ◦ gp where each gj is one of the elementary transformations
in 2.2(1)–(5) and p is a positive integer. Equivalently (see 2.3) f is a Möbius
transformation if f = h1 ◦ · · · ◦ hm where each hj is a reflection in a sphere or in a
n
hyperplane and m is a positive integer. If G ⊂ R the set of all (sense-preserving)
Möbius transformations mapping G onto itself is denoted by GM(G) (M(G)).
n
It will be convenient to identify R with the subset { x ∈ Rn : xn+1 = 0 }∪{∞}
n+1
of R . The identification is given by the embedding
(2.5) x 7→ x̃˜ = (x1 , . . . , xn , 0) ; x = (x1 , . . . , xn ) ∈ Rn .
We are now going to describe a natural two–step way of extending a Möbius
n
transformation of Rn to a Möbius transformation of Rn+1 . First, if f in GM(R )
is a reflection in P (a, t) or in S n−1 (a, r), let f˜˜ be a reflection in P (ã,
˜ t) or S n (ã,
˜ r),
n
respectively. Then if x ∈ R and y = f (x), by 2.2(1)–(2) we get
g
(2.6) f˜˜(x1 , . . . , xn , 0) = (y1 , . . . , yn , 0) = fg
(x) .
By (2.6) we may regard f˜˜ as an extension of f . Note that f˜˜ preserves the plane
xn+1 = 0 and each of the half–spaces xn+1 > 0 and xn+1 < 0. These facts follow
n
from the formulae 2.2(1)–(2). Second, if f is an arbitrary mapping in GM(R )
it has a representation f = f1 ◦ · · · ◦ fm where each fj is a reflection in a plane
or a sphere. Then f˜˜ = f˜˜1 ◦ · · · ◦ f˜˜m is the extension of f , and it preserves the
half–spaces xn+1 > 0, xn+1 < 0, and the plane xn+1 = 0. In conclusion, every
f in GM(R ) has an extension f˜˜ in GM(R ). It follows from [BE, p. 31,
n n+1
Metrics and quasiregular mappings 297

Theorem 3.2.4] that such an extension f˜˜ of f is unique. The mapping f˜˜ is called
˜ f˜˜,
the Poincaré extension of f . In the sequel we shall write x, f instead of x̃,
respectively.
Many properties of plane Möbius transformations hold for n–dimensional
n
Möbius transformations as well. The fundamental property that spheres of R
(which are spheres or planes in Rn , see [Vu1, Exercise 1.26, p.8]) are preserved
under Möbius transformations is proved in [BE, p. 28, Theorem 3.2.1].
n
2.7. Stereographic projection. The stereographic projection π: R →
S n ( 21 en+1 , 21 ) is defined by
x − en+1
(2.8) π(x) = en+1 + , x ∈ Rn ; π(∞) = en+1
|x − en+1 |2

n
Then π is the restriction to R of the inversion in S n (en+1 , 1). In fact, we can
identify π with this inversion. Because f −1 = f for every inversion f , it follows
n
that π maps the “Riemann sphere” S n ( 21 en+1 , 21 ) onto R .
n
The spherical (chordal) metric q in R is defined by
n
(2.9) q(x, y) = |π(x) − π(y)| ; x, y ∈ R ,
where π is the stereographic projection (2.8). From the definition (2.8) by cal-
culating we obtain

 |x − y|
 q(x, y) = p
 p ; x 6= ∞ 6= y,

 1 + |x|2 1 + |y|2
(2.10)

 1


 q(x, ∞) = p .
1 + |x|2

For x ∈ Rn \ {0} the antipodal (diametrically opposite) point x̃, is defined by


x
(2.11) x̃ = − 2
|x|
and we set ∞˜ = 0, 0̃ = ∞ . Then, by (2.10), q(x, x̃ ) = 1 and hence π(x),π(x̃ )
are indeed diametrically opposite points on the Riemann sphere.
n
2.12. Balls in the spherical metric. For x ∈ R and r ∈ (0, 1) we define the
spherical ball
n
(2.13) Q(x, r) = { z ∈ R : q(x, z) < r } .
Its boundary sphere is denoted by ∂Q(x, r). From the Pythagorean theorem it
follows that (cf. (2.11))
n √ 
(2.14) Q(x, r) = R \ Q x̃, 1 − r2 .
298 M. Vuorinen IWQCMA05

en+1

π(x)

π(y)

y 0

_
n

Figure 2. Formulae (2.8) and (2.9) visualized.

Figure 3. A cross-section of the Riemann sphere.

To gain insight into the geometry of spherical balls Q(x, r) it is convenient


to study the image πQ(x, r) under the stereographic projection π (see figure 3).
Indeed, by definition (2.9) we see that

(2.15) πQ(x, r) = B n+1 π(x), r ∩ S n ( 21 en+1 , 21 ).

Either by this formula or more directly by the definition of the spherical metric
(plus the fact that Möbius transformations preserve spheres) we see that in the
euclidean geometry, Q(x, r) is a point set of one of the following three kinds

(a) an open ball B n (u, s),


n
(b) the complement of B (v, t) in Rn ,
(c) a half–space of Rn .
Metrics and quasiregular mappings 299

n
Clearly, ∂Q(x, r) is either a sphere
√ or a hyperplane of R . Formula (2.14)
shows, in particular, that πQ(x, 1/ 2 ) is a half–sphere of the Riemann sphere
S n ( 21 en+1 , 21 ).
2.16. Absolute ratio. For an ordered quadruple a, b, c, d of distinct points in
n
R we define the absolute (cross) ratio by
q(a, c) q(b, d)
(2.17) | a, b, c, d | = .
q(a, b) q(c, d)
It follows from (2.10) that for distinct a, b, c, d in Rn
|a − c| |b − d|
| a, b, c, d | = .
|a − b| |c − d|
One of the most important properties of Möbius transformations is that they
preserve absolute ratios, i.e. if f ∈ GM, then
(2.18) | f (a), f (b), f (c), f (d) | = | a, b, c, d |
n
for all distinct a, b, c, d in R . As a matter of fact, the preservation of absolute
ratios is a characteristic property of Möbius transformations. It is proved in [BE,
n n
p. 72, Theorem 3.2.7] that a mapping f : R → R is a Möbius transformation
if and only if f preserves all absolute ratios.
2.19. Automorphisms in Bn . We shall give a canonical representation for
the maps in M(Bn ). Assume that f is in M(Bn ) and that f (a) = 0 for some
a ∈ Bn . We denote
a
(2.20) a∗ = 2 , a ∈ Rn \ {0}
|a|
and 0 = ∞, ∞ = 0. Fix a ∈ Bn \ {0}. Let
∗ ∗

(2.21) σa (x) = a∗ + r2 (x − a∗ )∗ , r2 = |a|−2 − 1


be an inversion in the sphere S n−1 (a∗ , r) orthogonal to S n−1 . Then σa (a) = 0,
σa (a∗ ) = ∞, σa (Bn ) = Bn .

Let pa denote the reflection in the (n − 1)–dimensional plane P (a, 0) through the
origin and orthogonal to a and define a sense–preserving Möbius transformation
by Ta = pa ◦ σa . Then, by (2.21), Ta Bn = Bn , Ta (a) = 0, and with ea = a/|a| we
have Ta (ea ) = ea , Ta (−ea ) = −ea . For a = 0 we set T0 = id, where id stands for
the identity map. The proof of the following fundamental fact can be found in
[A, p. 21], [BE, p. 40, Theorem 3.5.1].
We now define a spherical isometry tz in M(Rn ) which maps a given point
z ∈ Rn to 0 as follows. For z = 0 let tz = id and for z = ∞ let tz = p ◦ f , where
f is inversion in S n−1 and p is reflection in the (n − 1)–dimensional plane
p x1 = 0.
n n−1 2
For z ∈ R \ {0} let sz be inversion in S (−z/|z| , r), where r = 1 + |z|−2 .
According to [Vu1, (1.45)], the inversion sz is a spherical isometry and it is easy
to show that sz (z) = 0. Let pz be reflection in the plane P (z, 0). Defining
(2.22) tz = pz ◦ sz ,
300 M. Vuorinen IWQCMA05

s
S

1 r

0 a b

Figure 4. Inversion in S n−1 , b = a∗ .

we see that tz ∈ M(Rn ) is a spherical isometry with tz (z) = 0. Hence


√ 
tz (Q(z, r)) = Q(0, r) = B n r/ 1 − r2 ,
(2.23)
q(x, z)2
|tz (x)|2 =
1 − q(x, z)2
for all x, z ∈ Rn , r ∈ (0, 1).
2.24. Lemma. Let a ∈ Rn , r > 0, and let b ∈ Rn , u > 0, be such that
B n (a, r) = Q(b, u). If f is the inversion in S n−1 (a, r), then
f = t−1
b ◦ f1 ◦ tb ,

where tb √is the spherical isometry defined in (2.22) and f1 is the inversion in
S n−1 (u/ 1 − u2 ) = ∂Q(0, u).

3. Hyperbolic geometry
Hyperbolic geometry can be developed in the context of two spaces or, as they
are sometimes called, models. These two models of the hyperbolic space are the
unit ball Bn and the Poincaré half–space
Hn = Rn+ = { (x1 , . . . , xn ) ∈ Rn : xn > 0 } .
These two models can be equipped with a hyperbolic metric ρ that is unique
up to a multiplicative constant in either model. In either model the metric is
normalized (by giving the element of length of the metric) in such a way that for
all x, y ∈ Bn 
ρHn h(x), h(y) = ρBn (x, y)
whenever h ∈ GM and hBn = Hn . Therefore both models are conformally
compatible in the sense that the two metric spaces (Bn , ρ) and (Hn , ρ) can be
identified. This compatibility is very convenient in computations because we may
Metrics and quasiregular mappings 301

do a computation in that model in which it is easier, without loss of generality.


In what follows we shall use the symbols Rn+ and Hn interchangeably.

For A ⊂ Rn let A+ = { x ∈ A : xn > 0 }. We define a weight function


w : Rn+ → R+ = { x ∈ R : x > 0 } by
1
(3.1) w(x) = , x = (x1 , . . . , xn ) ∈ Rn+ .
xn
If γ : [0, 1) → Rn+ is a continuous mapping such that γ[0, 1) is a rectifiable curve
with length s = ℓ(γ), then γ has a normal representation γ 0 : [0, s) → Rn+
parametrized by arc length (see J. Väisälä [V, p. 5]). The hyperbolic length
of γ[0, 1) is defined by
Z s Z
0 ′ 0 |dx|
(3.2) ℓh (γ[0, 1)) = |(γ ) (t)| w(γ (t))dt = .
0 γ xn

If A ⊂ Rn+ is a (Lebesgue) measurable set we define the hyperbolic volume of A


by
Z
(3.3) mh (A) = w(x)n dm(x) ,
A

where m stands for the n–dimensional Lebesgue measure and w is as in (3.1). If


a, b ∈ Rn+ , then the hyperbolic distance between a and b is defined by
Z
|dx|
(3.4) ρ(a, b) = inf ℓh (α) = inf ,
α∈Γab α∈Γab α xn

where Γab stands for the collection of all rectifiable curves in Rn+ joining a and b.
Sometimes the more complete notation ρRn+ (a, b) or ρHn (a, b) will be employed.

Figure 5. Some geodesics of Hn = Rn+ .

The infimum in (3.4) is in fact attained: for given a, b ∈ Rn+ there exists a
circular arc L perpendicular to ∂Rn+ such that the closed subarc J[a, b] of L with
302 M. Vuorinen IWQCMA05

end points a and b satisfies


Z
|dx|
(3.5) ρ(a, b) = ℓh (J[a, b]) = .
xn
J[a,b]

If a and b are located on a normal of ∂Rn+ , then J[a, b] = [a, b] = { (1 − t)a + tb :


0 ≤ t ≤ 1 } (cf. [BE, p. 134]). Because of the (hyperbolic) length–minimizing
property (3.5), the arc J[a, b] will be called the geodesic segment joining a and b.
Knowing the geodesics, we calculate the hyperbolic distance in two special
cases. First, for r, s > 0 we obtain
Z r dt r

(3.6) ρ(ren , sen ) = = log .
s t s
Second, if ϕ ∈ (0, 12 π) we denote uϕ = (cos ϕ)e1 + (sin ϕ)en and calculate
Z Zπ/2
dα dα
(3.7) ρ(en , uϕ ) = = = log cot 12 ϕ .
sin α sin α
J[uϕ ,en ] ϕ

Figure 6. The points uϕ and en .

We shall often make use of the hyperbolic functions sh x = sinh x, ch x =


cosh x, th x = tanh x, cth x = coth x and their inverse functions. The above
formulae (3.6) and (3.7) are special cases of the general formula (see [BE, p. 35])
|x − y|2
(3.8) ch ρ(x, y) = 1 + , x, y ∈ Hn = Rn+ .
2xn yn
Note that by this formula the hyperbolic distance ρ(x, y) is completely deter-
mined once the euclidean distances xn = d(x, ∂Hn ), yn = d(y, ∂Hn ), and |x − y|
are known. In passing we note that if f2 ∈ GM(Hn ) is as defined in Remark
2.3(4), then ρ(x, y) = ρ(f2 (x), f2 (y)) for all x, y ∈ Hn . For another formulation
of (3.8) let z, w ∈ Hn , let L be an arc of a circle perpendicular to ∂Hn with
z, w ∈ L and let {z∗ , w∗ } = L ∩ ∂Hn , the points being labelled so that z∗ , z, w, w∗
occur in this order on L. Then (cf. [BE, p. 133, (7.26)])
(3.9) ρ(z, w) = log | z∗ , z, w, w∗ | .
Metrics and quasiregular mappings 303

Figure 7. The quadruple z∗ , z, w, w∗ .

Note that (3.6) is a special case of (3.9) when z∗ = 0 and w∗ = ∞ because


| 0, z, w, ∞ | = |w|/|z| for z, w ∈ Hn . The invariance of ρ is apparent by (3.9)
and (2.18): If f in GM(Hn ), then for all x, y ∈ Hn
(3.10) ρ(x, y) = ρ(f (x), f (y)) .

For a ∈ Hn and M > 0 the hyperbolic ball { x ∈ Hn : ρ(a, x) < M } is


denoted by D(a, M ). It is well known that D(a, M ) = B n (z, r) for some z and
r (this also follows from (3.10)! ). This fact together with the observation that
λten , (t/λ)en ∈ ∂D(ten , M ), λ = eM (cf. (3.6)), yields
 
 D(ten , M ) = B n (t ch M )en , t sh M ,
(3.11) B n (ten , rt) ⊂ D(ten , M ) ⊂ B n (ten , Rt) ,

r = 1 − e−M , R = eM − 1 .

Figure 8. The hyperbolic ball D(ten , M ) as a euclidean ball.

A counterpart of (3.8) for Bn is


 |x − y|2
(3.12) sh2 21 ρ(x, y) = , x, y ∈ Bn ,
(1 − |x| )(1 − |y| )
2 2

(cf. [BE, p. 40]). As in the case of Hn , we see by (3.12) that the hyperbolic
distance ρ(x, y) between x and y is completely determined by the euclidean
304 M. Vuorinen IWQCMA05

quantities |x − y|, d(x, ∂Bn ), d(y, ∂Bn ). Finally, we have also


(3.13) ρ(x, y) = log | x∗ , x, y, y∗ | ,
where x∗ , y∗ are defined as in (3.9): If L is the circle orthogonal to S n−1 with
x, y ∈ L, then {x∗ , y∗ } = L ∩ S n−1 , the points being labelled so that x∗ , x, y, y∗
occur in this order on L. It follows from (3.13) and (2.18) that
(3.14) ρ(x, y) = ρ(h(x), h(y))
for all x, y ∈ Bn whenever h is in GM(Bn ). Finally, in view of (2.18), (3.9), and
(3.13) we have
(3.15) ρBn (x, y) = ρHn (g(x), g(y)) , x, y ∈ Bn ,
whenever g is a Möbius transformation with gBn = Hn .
It is well known that the balls D(z, M ) of (Bn , ρ) are balls in the euclidean
geometry as well, i.e. D(z, M ) = B n (y, r) for some y ∈ Bn and r > 0. Making
use of this fact, we shall find y and r. Let Lz be a euclidean line through 0 and z
and {z1 , z2 } = Lz ∩ ∂D(z, M ), |z1 | ≤ |z2 |. We may assume that z 6= 0 since with
obvious changes the following argument works for z = 0 as well. Let e = z/|z|
and z1 = se, z2 = ue, u ∈ (0, 1), s ∈ (−u, u). Then it follows that
 1 + |z| 1 − s 
ρ(z1 , z) = log · =M ,
1 − |z| 1 + s
 1 + u 1 − |z| 
ρ(z2 , z) = log · =M .
1 − u 1 + |z|

Solving these for s and u and using the fact that



D(z, M ) = B n 1
(z
2 1
+ z2 ), 21 |u − s|
one obtains the following formulae:


 D(x, M ) = B n (y, r)

(3.16)

 x(1 − t2 ) (1 − |x|2 )t
 y= , r = , t = th 21 M ,
1 − |x| t
2 2 1 − |x| t
2 2

and
  

 B n x, a(1 − |x|) ⊂ D(x, M ) ⊂ B n x, A(1 − |x|) ,

(3.17)

 t(1 + |x|) t(1 + |x|)
 a= , A= , t = th 12 M .
1 + |x|t 1 − |x|t

We shall often need a special case of (3.16):


1
(3.18) D(0, M ) = B n (th M ) .
2
Metrics and quasiregular mappings 305

A standard application of formula (3.18) is the following observation. Let Tx be


in M(Bn ) as defined in 2.19 with Tx (x) = 0. Fix x, y ∈ Bn and z ∈ J[x, y] with
ρ(z, x) = ρ(z, y) = 12 ρ(x, y). Then Tz (x) = −Tz (y) and (3.18) yields

|Tx (y)| = th 12 ρ(x, y) ,
(3.19)
|Tz (x)| = th 14 ρ(x, y) .
For an open set D in Rn , D 6= Rn , define d(z) = d(z, ∂D) for z ∈ D and
 |x − y| 
(3.20) jD (x, y) = log 1 +
min{d(x), d(y)}
for x, y ∈ D. Then it is well-known that jD is a metric (see, e.g. [S]).
3.21. Lemma. The following inequalities
d(x)

(1) jD (x, y) ≥ log ,
d(y)
d(x)  |x − y| 

(2) jD (x, y) ≤ log + log 1 + ≤ 2 jD (x, y)
d(y) d(x)
|x − z|

(3) jD (x, y) ≥ log
|y − z|
hold for all x, y ∈ D, z ∈ ∂D .

In the next lemma we show that jD yields simple two–sided estimates for ρD
both when D = Bn and when D = Hn .
3.22. Lemma. (1) jBn (x, y) ≤ ρBn (x, y) ≤ 4 jBn (x, y) for x, y ∈ Bn .
(2) jHn (x, y) ≤ ρHn (x, y) ≤ 2 jHn (x, y) for x, y ∈ Hn .

4. Quasihyperbolic geometry
In an arbitrary proper subdomain D of Rn one can define a metric, the quasi-
hyperbolic metric of D, which shares some properties of the hyperbolic metric
of Bn or Hn . We shall now give the definition of the quasihyperbolic metric and
state without proof some of its basic properties which we require later on. The
quasihyperbolic metric has been systematically developed and applied by F. W.
Gehring and his collaborators.
Throughout this section D will denote a proper subdomain of Rn . In D we
define a weight function w : D → R+ by
1
(4.1) w(x) = ; x∈D.
d(x, ∂D)
Using this weight function one defines the quasihyperbolic length ℓq (γ) = ℓD q (γ)
of a rectifiable curve γ by a formula similar to (3.2). The quasihyperbolic distance
between x and y in D is defined by
Z
D
(4.2) kD (x, y) = inf ℓq (α) = inf w(x)|dx| ,
α∈Γxy α∈Γxy α
306 M. Vuorinen IWQCMA05

where Γxy is as in (3.4). It is clear that kD is a metric on D. It follows from (4.2)


that kD is invariant under translations, stretchings, and orthogonal mappings.
(As in (3.3) one can define the quasihyperbolic volume of a (Lebesgue) measurable
set A ⊂ D, but we shall not make use of this notion.) Given x, y ∈ D there exists
a geodesic segment JD [x, y] of the metric kD joining x and y (cf. [GO]). However,
very little is known about the structure of such geodesic segments JD [x, y] when
D is given. For some elementary domains, the geodesics were recently studied
by H. Lindén [L].
4.3. Remarks. Clearly, kHn = ρHn , and we see easily that ρBn ≤ 2 kBn ≤ 2 ρBn
(cf. (4.1)). Hence, the geodesics of (Hn , kHn ) are those of (Hn , ρHn ), but it is a
difficult task to find the geodesics of kD when D is given. The following monotone
property of kD is clear: if D and D′ are domains with D′ ⊂ D and x, y ∈ D′ ,
then kD′ (x, y) ≥ kD (x, y).
In order to find some estimates for kD (x, y) we shall employ, as in the case of
Hn and Bn , the metric jD defined in (3.20). The metric jD is indeed a natural
choice for such a comparison function since both kD and jD are invariant under
translations, stretchings and orthogonal mappings. A useful inequality is ([GP,
Lemma 2.1])
(4.4) kD (x, y) ≥ jD (x, y) ; x, y ∈ D .
In combination with 3.22, (4.4) yields
d(x)

(4.5) kD (x, y) ≥ log , d(z) = d(z, ∂D) .
d(y)
For easy reference we record Bernoulli’s inequality
(4.6) log(1 + as) ≤ a log(1 + s) ; a ≥ 1 , s > 0 .
4.7. Lemma. (1) If x ∈ D, y ∈ Bx = B n (x, d(x)), then
 |x − y| 
kD (x, y) ≤ log 1 + .
d(x) − |x − y|
(2) If s ∈ (0, 1) and |x − y| ≤ s d(x), then
1
kD (x, y) ≤ jD (x, y) .
1−s

5. Modulus and capacity


For the definition and basic properties of the modulus we refer the reader to
[V1]. The main sources for this section are [V1], [Vu3], [AVV2].
One of the main reasons why the modulus of a curve family is studied is
that we have a simple rule of transformation for the modulus of a curve family
under quasiconformal mappings. Further, we would like to use modulus as an
instrument so as to gain insight about ”the geometry”. Roughly speaking we
can say that the modulus of the family of all curves joining two connected non-
intersecting continua E, F ⊂ Rn behaves like min{d(E), d(F )}/d(E, F ) , where
Metrics and quasiregular mappings 307

d stands for the euclidean diameter. A long series of estimates is needed to


reach this conclusion and its variants and some of these estimates are given
in this and the following section. Some of these variants involve hyperbolic or
quasihyperbolic metric. In this fashion we step by step approach our goal, the
study of how quasiconformal mappings between metric spaces deform distances.
5.1. Lemma. The p–modulus Mp is an outer measure in the space of all curve
families in Rn . That is,
(1) Mp (∅) = 0 ,
(2) Γ1 ⊂ Γ2 implies Mp (Γ1 ) ≤ Mp (Γ2 ) ,
[∞  X ∞
(3) Mp Γi ≤ Mp (Γi ) .
i=1 i=1

Let Γ1 and Γ2 be curve families in Rn . We say that Γ2 is minorized by Γ1 and


write Γ2 > Γ1 if every γ ∈ Γ2 has a subcurve belonging to Γ1 .
5.2. Lemma. Γ1 < Γ2 implies Mp (Γ1 ) ≥ Mp (Γ2 ).
5.3. Remark. The family of all paths joining E and F in G is denoted by
n
∆(E, F ; G) see [Vu3, p.51]. If G = Rn or R we often denote ∆(E, F ; G) by
∆(E, F ). Curve families of this form are the most important S∞ for what follows.
The following subadditivity property is useful. If E = j=1 Ej and cE (F ) =
 P
Mp ∆(E, F ) = cF (E), then cF (E) ≤ cF (Ej ), see 5.1(3). More precisely if
n 
G ⊂ R is a domain and F ⊂ G is fixed, then cG F (E) = Mp ∆(E, F ; G) is an
outer measure defined for E ⊂ G. In a sense which will be made precise later
on, cE (F ) describes the mutual size and location of E and F . Assume now that
n
D is an open set in R and that F ⊂ D. It follows from 5.1(2) that
  
Mp ∆(F, ∂D; D \ F ) ≤ Mp ∆(F, ∂D; D) ≤ Mp ∆(F, ∂D) .
On the other hand, because ∆(F, ∂D; D) < ∆(F, ∂D) and ∆(F, ∂D; D \ F ) <
∆(F, ∂D; D) , 5.2 yields
  
(5.4) Mp ∆(F, ∂D) = Mp ∆(F, ∂D; D) = Mp ∆(F, ∂D; D \ F ) .
n
5.5. Lemma. Let D and D′ be domains in R and let f : D → D′ be a conformal
mapping. Then M(f Γ) = M(Γ) for each curve family Γ in D where f Γ = { f ◦ γ :
γ ∈ Γ }.
5.6. Lemma. Let p > 1 and let E and F be subsets of Rn+ . Then
 1 
Mp ∆(E, F ; Rn+ ) ≥ Mp ∆(E, F ) .
2
n
5.7. Corollary.
 Let E and F be sets in R with q( E, F ) ≥ a > 0. Then
M ∆(E, F ) ≤ c(n, a) < ∞.
5.8. Lemma. (1) Let 0 < a < b and let E, F be sets in Rn with
E ∩ S n−1 (t) 6= ∅ 6= F ∩ S n−1 (t)
308 M. Vuorinen IWQCMA05

for t ∈ (a, b). Then


 b
M ∆( E, F ; B n (b) \ B n (a) ) ≥ cn log .
a
Equality holds if E = (ae1 , be1 ), F = (−be1 , −ae1 ). Here cn > 0 depends only on
n (see [V1, (10.11),(10.4)]).
(2) Let 0 < a < b . Then

M ∆( S n−1 (a), S n−1 (a); B n (b) \ B n (a) ) = ωn−1 (log(b/a))1−n ,
where ωn−1 is the (n − 1)-dimensional surface area of S n−1 .
5.9. Corollary.
 If E and F are non–degenerate continua with 0 ∈ E ∩ F then
M ∆(E, F ) = ∞.
Proof. Apply 5.8 with a fixed b such that S n−1 (b) ∩ E 6= ∅ 6= S n−1 (b) ∩ F and
let a → 0.
5.10. Canonical ring domains. A domain (open, connected set) D in Rn is
called a ring domain or, briefly, a ring, if Rn \ D consists of two components
C0 and C1 . Sometimes we denote such a ring by R(C0 , C1 ). In our study two
canonical ring domains will be of particular importance. These are the Grötzsch
ring RG,n (s), s > 1, and the Teichmüller ring RT,n (t), t > 0, defined by

RG,n (s) = R(Bn , [se1 , ∞]), s > 1,
(5.11)
RT,n (t) = R([−e1 , 0], [te1 , ∞]), t > 0.
Sometimes we also use the bounded Grötzsch ring R(Rn \ Bn , [0, re1 ]) . An im-
portant conformal invariant associated with a ring is the modulus of the family
of curves joining its complementary components. In the case of Grötzsch ring
RG,n (s) and Teichmüller ring RT,n (t) the modulus is denoted by γn (s) and τn (t)
respectively. It is a well-known basic fact that γn : (1, ∞) → (0, ∞) is a decreas-
ing homeomorphism and that for all s > 1
(5.12) γn (s) = 2n−1 τn (s2 − 1) .

s t
Bn

se1 ∞ -e1 0 t e1 ∞

cap R (s) = M( s ) = (s) cap RT,n (t) = M( ) = n(t)


G,n n t

Figure 9. The Grötzsch and Teichmüller rings.


Metrics and quasiregular mappings 309

z w

-t t 1 f - r r 1
-1 -1

z w
i K '/2

-1 -t t 1 -K K -1 - r r 1
2 i K Log z + K
w = r sn( ,r), = π t

Figure 10. Conformal map of an annulus onto a disk minus a


symmetric slit.

z w
f g

-1 -t t 1 -1 - r r 1 -1 0 a 1

g f

Figure 11. Conformal map of an annulus onto a bounded


Grötzsch ring.

5.13. Elliptic integrals and γ2 (s). In the plane every ring domain can be
conformally mapped onto an annulus {z ∈ C : 1 < |z| < M } for some M . For
the Grötzsch ring this conformal mapping is given by the elliptic sn-function
[AVV2]. For more information on the involved special functions see [QV].
As shown in [LV, II.2]

(5.14) γ2 (s) = 2π/µ(1/s)


for s > 1 where
√ Z 1
π K( 1 − r2 )
µ(r) = , K(r) = [(1 − x2 )(1 − r2 x2 )]−1/2 dx
2 K(r) 0

for 0 < r < 1. The function K(r) is called a complete elliptic integral of the first
kind and its values can be found in tables.
310 M. Vuorinen IWQCMA05

5.5 5.5

5 5

4.5 4.5

4 4

3.5 3.5

3 3

2.5 K’ K 2.5
µ
2 2

1.5 1.5

1 1

0.5 0.5

0 0
0 0.5 1 0 0.5 1

Figure 12. The functions K(r) and µ(r) .

The modulus µ(r) satisfies the following three functional identities


 √ 
 µ(r)µ 1 − r 2 = 1 π2 ,

  4
1−r 1 2
(5.15) µ(r)µ 1+r = 2 π ,
  √ 

 µ(r) = 2µ 2 r .
1+r

From (5.15) one can derive several estimates for µ(r) [LV, p. 62]. By [LV, p. 62]
the following inequalities hold

1 1 + 3 1 − r2 4
(5.16) log < log < µ(r) < log
r r r
for 0 < r < 1. From (5.16) it follows that limr→0+ µ(r) = ∞ whence, by virtue of
the functional identities (5.15), limr→1− µ(r) = 0. Therefore, µ : (0, 1) → (0, ∞)
is a decreasing homeomorphism. For the sake of completeness we set µ(0) = ∞
and µ(1) = 0. By (5.14) and (5.15) we obtain
4 s − 1
(5.17) γ2 (s) = µ , s>1.
π s+1
5.18. Exercise. In the study of distortion theory of quasiconformal mappings
in Section 7 below the following special function will be useful
1
ϕK,n (r) = −1
γn (Kγn (1/r))
Metrics and quasiregular mappings 311

for 0 < r < 1, K > 0. (Note: [Vu1, Lemma 7.20] shows that γn is strictly
decreasing and hence that γn−1 exists.) Show that ϕAB,n (r) = ϕA,n (ϕB,n (r)) and
ϕ−1
A,n (r) = ϕ1/A,n (r) and that

ϕK,2 (r) = ϕK (r) = µ−1 K1 µ(r) .
Verify also that

(1) ϕ2 (r) = 21+rr
√ 2
(2) ϕK (r)2 + ϕ1/K 1 − r2 = 1 .
Exploiting (1) and (2) find ϕ1/2 (r). Show also that
 
1−r 1−ϕK (r)
(3) ϕ1/K 1+r = 1+ϕ K (r)
,
 √  √
2 ϕ (r)
(4) ϕK 21+rr = 1+ϕKK(r) .

Lemma 7.22 in [Vu3] yields the inequalities


(5.19) ωn−1 (log λn s)1−n ≤ γn (s) ≤ ωn−1 (log s)1−n ,
1−n
(5.20) ωn−1 (log(λ2n s)) ≤ τn (s − 1) ≤ ωn−1 (log s)1−n ,

for s > 1.
5.21. Theorem. The function gn (t) = (ωn−1 /γn (t))1/(n−1) −log t is an increasing
function on (1, ∞) with limt→∞ gn (t) = log λn where λn ∈ [4, 2en−1 ), λ2 = 4 , is
so-called Grötzsch ring constant.
5.22. Theorem. For s ∈ (1, ∞) and n ≥ 2
√ 1−n
(1) γn (s) ≤ ωn−1 µ(1/s)1−n < ωn−1 log(s + 3 s2 − 1 ) ,
s + 1 s − 1  s + 1
n−1 n−1 n−1
(2) 2 cn log ≤ γn (s) ≤ 2 cn µ <2 cn log 4 .
s−1 s+1 s−1

Moreover, if s ∈ (0, ∞) and a = 1 + 2(1 + 1 + s )/s, then

(3) cn log a ≤ τn (s) ≤ cn µ(1/a) < cn log(4a)


√ √
and (1 + 1/ s )2 ≤ a ≤ (1 + 2/ s )2 hold true. Furthermore, when n = 2, the
first inequality in (1), the second inequality in (2), and the second inequality in
(3) hold as identities.

5.23. Hyperbolic metric and capacity. As in Section 3 we let J[x, y] denote


the geodesic segment of the hyperbolic metric joining x to y, x, y ∈ Bn . It is
clear by conformal invariance that
cap(Bn , J[x, y]) = cap(Bn , Tx J[x, y])
312 M. Vuorinen IWQCMA05

5
G

3 F

2
g

f
1

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Figure 13. Bounds for γ3 .

where Tx is as defined in 2.19. We obtain by (3.19) and [Vu1, (7.25)]


 1  1 1−n
(5.24) cap(Bn , J[x, y]) = γn ≤ ωn−1 − log th ρ(x, y) .
th 12 ρ(x, y) 4
Next by (5.24), 5.22(2), and (5.16) we get
2n−1 cn ρ(x, y) ≤ cap(Bn , J[x, y]) ≤ 2n−1 cn µ(e−ρ(x,y) )
(5.25) < 2n−1 cn (ρ(x, y) + log 4) .
For large values of ρ(x, y) (5.25) is quite accurate. For small ρ(x, y) one obtains
better inequalities than (5.25) by combining 5.22(1) and (5.24).

It is left as an easy exercise for the reader to derive from (5.19) the following
inequality
(5.26) tα /λn ≤ γn−1 (Kγn (t)) ≤ λαn tα
for all t > 1 and K > 0, where α = K 1/(1−n) . From (5.26) it follows immediately
that
(5.27) rα λn −α ≤ ϕK,n (r) ≤ λn rα
holds for all K > 0 and r ∈ (0, 1). For K ≥ 1 this inequality can be refined if we
use Theorem 5.22 (1).
Metrics and quasiregular mappings 313

5.28. Theorem. For n ≥ 2, K ≥ 1, and 0 ≤ r ≤ 1


(1) ϕK (r) ≤ λ1−α
n rα , α = K 1/(1−n) ,
(2) ϕ1/K (r) ≥ λ1−β
n rβ , β = K 1/(n−1) .

A compact set E ⊂ Bn is said to be of capacity zero, denoted cap E = 0 , if


n
M(∆(E, S n−1 (2))) = 0 . A compact set E ⊂ R is said to be of capacity zero, if it
can be mapped by a Möbius transformation onto a set E1 ⊂ Bn of capacity zero.
Sets of capacity zero are very small: they have zero Hausdorff dimension, see
[Vu3, p.86]. For many purposes they are negligible. The next theorem provides
a convenient tool for estimating moduli of curve families in terms is geometric
quantities and a set function.
5.29. Theorem. For n ≥ 2 there exist positive numbers d1 , . . . , d4 and a set
n
function c(·) in R such that
n n n
(1) c(E) = c(hE) whenever h : R → R is a spherical isometry  and E ⊂ R .
n S
(2) c(∅) = 0, A ⊂ B ⊂ R implies c(A) ≤ c(B) and c ∞ j=1 Ej
P∞ n
≤ d1 j=1 c(Ej ) if Ej ⊂ R .
n
(3) If E ⊂ R is compact, then c(E) > 0 if and only if cap E > 0. Moreover
n
c(R ) ≤ d2 < ∞.
n
(4) c(E) ≥ d3 q(E)
 if E ⊂ R is connected and E 6= ∅. n
(5) M ∆(E, F ) ≥ d4 min{ c(E), c(F ) }, if E, F ⊂ R .
Furthermore, for n ≥ 2 and t ∈ (0, 1) there exists a positive number d5 such that
n
(6) M ∆(E, F ) ≤ d5 min{c(E), c(F )} whenever E, F ⊂ R and q(E, F ) ≥ t.

6. Conformal invariants
In the preceding sections
 we have studied some properties of the conformal
invariant M ∆(E, F ; G) . In this section we shall introduce two other conformal
invariants, the modulus metric µG (x, y) and its ”dual” quantity λG (x, y), where
n
G is a domain in R and x, y ∈ G. The modulus metric µG is functionally related
to the hyperbolic metric ρG if G = Bn , while in the general case µG reflects the
“capacitary geometry” of G in a delicate fashion. The dual quantity λG (x, y)
is also functionally related to ρG if G = Bn . As shown in [Vu3] for a wide
class of domains in Rn , the so–called QED–domains[GM], two–sided estimates
for λG (x, y) in terms of
|x − y|
rG (x, y) = .
min{ d(x, ∂G), d(y, ∂G) }
6.1. The conformal invariants λG and µG . If G is a proper subdomain of
n
R , then for x, y ∈ G with x 6= y we define

(6.2) λG (x, y) = inf M ∆(Cx , Cy ; G)
Cx ,Cy

where Cz = γz [0, 1) and γz : [0, 1) → G is a curve such that γz (0) = z and


γz (t) → ∂G when t → 1, z = x, y. It follows from 5.5 that λG is invariant under
314 M. Vuorinen IWQCMA05

conformal mappings of G. That is, λf G (f (x), f (y)) = λG (x, y), if f : G → f G is


conformal and x, y ∈ G are distinct.

Cy
G G
• •
y y
Cxy
• Cx • x
x
G (x,y) G(x,y)

Figure 14. The conformal invariants λG and µG .


n
6.3. Remark. If card(R \ G) = 1, then λG (x, y) ≡ ∞ by 5.9. Therefore λG is
n n
of interest only in case card(R \ G) ≥ 2. For card(R \ G) ≥ 2 and x, y ∈ G,
x 6= y, there are continua Cx and Cy as in (6.2) with C x ∩ C y = ∅ and thus
n
M ∆(Cx , Cy ; G) < ∞ by 5.7. Thus, if card(R \ G) ≥ 2, we may assume that
the infimum in (6.2) is taken over continua Cx and Cy with C x ∩ C y = ∅.
6.4. The extremal problems of Grötzsch and Teichmüller. The Grötzsch
and Teichmüller rings arise from extremal problems of the following type, which
were first posed for the case of the plane: Among all ring domains which separate
two given closed sets E1 and E2 , E1 ∩ E2 = ∅, find one whose module has the
greatest value.
Let E1 be a continuum and E2 consist of two points not separated by E1 . By
the conformal invariance of the modulus one may then suppose that E1 = S 1 and
E2 = {0, r}, 0 < r < 1. Then the extremal problem is solved by the bounded
Grötzsch ring R(R2 \ B 2 , [0, r]). In other words, cap(B 2 , E) ≥ γ2 (1/r), where
E ⊂ B 2 is any continuum joining the points 0 and r ∈ R. For details we refer
the reader to [LV, Ch. II].
The following function is the solution of the generalization of the Teichmüller
problem to Rn . For x ∈ Rn \ {0, e1 }, n ≥ 2, define
(6.5) p(x) = inf M(△(E, F )),
E,F
n
where the infimum is taken over all pairs of continua E and F in R with 0, e1 ∈
E, x, ∞ ∈ F . Teichmüller applied a symmetrization method to prove that for
n = 2,
p(x) ≥ p((1 + |x − e1 |)e1 )
with equality for x = (1 + |x − e1 |)e1 . For more details, see [HV] and [SoV].
n
For a proper subdomain G of R and for all x, y ∈ G define

(6.6) µG (x, y) = inf M ∆(Cxy , ∂G; G)
Cxy
Metrics and quasiregular mappings 315

2.5

2
F ∞
1.5

1
x
0.5

0 0 1

E
−0.5

−1
−1 0 1 2 3 4

Figure 15. The extremal problem of Teichmüller.

where the infimum is taken over all continua Cxy such that Cxy = γ[0, 1] and γ
is a curve with γ(0) = x and γ(1) = y. It is clear that µG is also a conformal
invariant in the same sense as λG . It is left as an easy exercise for the reader
to verify that µG is a metric if cap∂G > 0. [Hint: Apply 5.3 and 5.29.] If
cap∂G > 0, we call µG the modulus metric or conformal metric of G.
6.7. Remark. Let D be a subdomain of G. It follows from 5.3 and (5.4)
that µG (a, b) ≤ µD (a, b) for all a, b ∈ D and λG (a, b) ≥ λD (a, b) for all dis-
tinct a, b ∈ D. In what follows we are interested only in the non–trivial case
n
card(R \ G) ≥ 2. Moreover, by performing an auxiliary Möbius transformation,
n
we may and shall assume that ∞ ∈ R \ G throughout this section. Hence G
will have at least one finite boundary point.

In a general domain G, the values of λG (x, y) and µG (x, y) cannot be expressed


in terms of well–known simple functions. For G = Bn they can be given in terms
of ρ(x, y) and the capacity of the Teichmüller condenser.
6.8. Theorem. The following identities hold for all distinct x, y ∈ Bn :
!  
n−1 1 1
(1) µBn (x, y) = 2 τ =γ ,
sh2 21 ρ(x, y) th 21 ρ(x, y)
 
1 2 1
(2) λBn (x, y) = τ sh ρ(x, y) .
2 2
316 M. Vuorinen IWQCMA05

6.9. Remark. (1) In [Vu3, p. 193] it was stated as an open problem, whether
λD (x, y)1/(1−n) is a metric when D = Rn \ {0} and n = 2 . Subsequently the
problem was solved by A. Solynin [So] and J. Jenkins [J] for n = 2 . J. Ferrand
[F] proved that λD (x, y)1/(1−n) is a metric for all D ⊂ Rn , n ≥ 2.
(2) From 5.22(3) we obtain the following inequality for x, y ∈ Bn (exercise)
1 1  1
τ sh2 ρ(x, y) ≥ −cn log th ρ(x, y)
2 2   4
1 1
= 2cn arth e− 2 ρ(x,y) ≥ 2cn e− 2 ρ(x,y) .

Here the identities 2 ch2 A = 1 + ch 2A , sh 2A = 2 ch A sh A , and log th s =


−2 arth e−2s were applied. Recall that
1 |x − y|2
sh2 ρ(x, y) =
2 (1 − |x|2 )(1 − |y|2 )
by (3.12). Similarly, by 5.22(3) we obtain also
1 1  1 1  1 4
τ sh2 ρ(x, y) ≤ cn µ th2 ( ρ(x, y)) < cn log 2 1
2 2 2 4 2 th 4 ρ(x, y)
2
= cn log 1 .
th 4 ρ(x, y)
6.10. Lemma. Let G be a proper subdomain of Rn , x ∈ G, d(x) = d(x, ∂G),
Bx = B n (x, d(x)), let y ∈ Bx with y 6= x, and let r = |x − y|/d(x). Then the
following two inequalities hold:
1  r2  1
(1) λG (x, y) ≥ λBx (x, y) = τ > cn log
2 1−r 2 r
1  1 1−n
(2) µG (x, y) ≤ µBx (x, y) = γ ≤ ωn−1 log .
r r
6.11. Lemma. The inequality

p(x) ≥ max τ (|x|) , τ (|x − e1 |)
holds for all x ∈ Rn \ {0, e1 }. Equality holds if x = se1 and s < 0 or s > 1.

The following theorem summarizes some properties of p(x).


6.12. Theorem. For |x − e1 | ≤ |x|, x ∈ Rn \ {0, e1 }
(1) p(x) ≤ 2 τ (|x − e1 |) when |x + e1 | ≥ 2,
(2) p(x) ≤ 4 τ (|x − e1 |) when |x| ≥ 1,
(3) p(x) ≤ 2n+1 τ (|x − e1 |).

This result was improved by D. Betsakos [B] who proved the next theorem.
The sharp constant in Theorem 6.13 is not known for n > 2 , for n = 2, see [BV].
6.13. Theorem. For x ∈ Rn \ {0, e1 }
(6.14) p(x) ≤ 4τ (min{|x|, |x − e1 |}) .
Metrics and quasiregular mappings 317

For x ∈ Rn \{0} we denote by rx a similarity map with rx (0) = 0 and rx (x) = e1 .


Then it is easy to see that |rx (y) − e1 | = |x − y|/|x|. It follows immediately from
the definitions (6.2) and (6.5) that
(6.15) λRn \{0} (x, y) = min{ p(rx (y)) , p(ry (x)) } .

Next we deduce the following two–sided inequality for λRn \{0} (x, y).
6.16. Theorem. For distinct x, y ∈ Rn \ {0} the following inequality holds
 
1 ≤ λRn \{0} (x, y) τ |x − y|/ min{|x|, |y|} ≤ 4 .
6.17. Corollary. Let G be a proper subdomain of Rn , x and y distinct points
in G and m(x, y) = min{d(x), d(y)}. Then

λG (x, y) ≤ inf λRn \{z} (x, y) ≤ 4 τ |x − y|/m(x, y) .
z∈∂G

Proof. The first inequality follows from 6.7. For the second one fix z0 ∈ ∂G with
m(x, y) = d({x, y}, {z0 }). Applying 6.16 to Rn \ {z0 } yields the desired result.

We next show that 6.17 fails to be sharp for a Jordan domain G in Rn . For
t ∈ (0, 15 ) consider the family Gt = B n (−e1 , 1) ∪ B n (e1 , 1) ∪ B n (t) of Jordan
domains. Then by 6.17
λGt (−e1 , e1 ) ≤ 4 τ (2)
for all t ∈ (0, 51 ). But this is far from sharp because in fact

λGt (−e1 , e1 ) ≤ M ∆( [−2e1 , −e1 ] , [e1 , 2e1 ] ; Gt )
 1 1−n
≤ ωn−1 log −→ 0
t
as t → 0. However, for a wide class of domains, which we shall now consider,
the upper bound in 6.17 is essentially best possible.

−e1 e1

Figure 16. The family ∆( [−2e1 , −e1 ] , [e1 , 2e1 ] ; Gt ) .


318 M. Vuorinen IWQCMA05

n
6.18. QED domains. A closed set E in R is called a c–quasiextremal distance
set or c-QED exceptional set or c-QED set, c ∈ (0, 1], if for each pair of disjoint
n
continua F1 , F2 ⊂ R \ E
n  
(6.19) M ∆(F1 , F2 ; R \ E) ≥ c M ∆(F1 , F2 ) .
n n
If G is a domain in R such that R \ G is a c-QED set, then we call G a c-QED
domain.
6.20. Examples. (1) The unit ball Bn is a 12 –QED set by [GM1] or by the
above Lemma 5.6.
(2) If E is a compact set of capacity zero, then E is a 1–QED set. For instance
all isolated sets are 1–QED sets. The class of all 1–QED sets contains all closed
sets in Rn of vanishing (n − 1)–dimensional Hausdorff measure (see [V3], [GM1]).
(3) B2 \ [0, e1 ) is not a c-QED set for any c > 0.
6.21. Theorem. Let G be a c-QED domain in Rn . Then
λG (x, y) ≥ c τ (s2 + 2s) ≥ 21−n c τ (s)
where s = |x − y|/ min{d(x), d(y)}.

It should be noted that the lower bound of 6.21 is very close to that of 6.16;
in fact it differs only by a multiplicative constant.
In the next few theorems we shall give some estimates for the conformal
metric µG .
6.22. Lemma. Let G be a proper subdomain of Rn , s ∈ (0, 1), x, y ∈ G. If
kG (x, y) ≤ 2 log(1 + s), then
 1 
(1) µG (x, y) ≤ γ .
th(kG (x, y)/(1 − s))
Moreover, there exist positive numbers b1 and b2 depending only on n such that
(2) µG (x, y) ≤ b1 kG (x, y) + b2
for all x, y ∈ G.

It should be observed that (6.22(2)) is a generalization of the upper bound in


(5.25) to the case of an arbitrary domain. The lower bound in (5.25) will next
be generalized to the case of domains with connected boundary.
6.23. Lemma. Let G be a domain in Rn such that ∂G is connected. Then for
all a, b ∈ G, a 6= b,
(1) µG (a, b) ≥ τ (4m2 + 4m) ≥ cn jG (a, b)
where cn is the constant in 5.8 and m = min{d(a), d(b)}/|a − b|. If, in addition,
G is uniform, then
(2) µG (a, b) ≥ B kG (a, b)
for all a, b ∈ G.
Metrics and quasiregular mappings 319

7. Distortion theory
For the basic properties and definitions of K-quasiconformal and K-quasiregular
mappings we refer the reader to [Vu3] as well as to the other papers in this same
volume. See, in particular, Rasila’s paper. One of the key ideas is that under
a K-quasiconformal mapping, the modulus is changed at most by a constant
c ∈ [1/K, K] . The notions introduced in the previous chapters enable us to for-
mulate this basic property in a more concrete and geometric way, in terms of
metrics.
Theorem 7.1 and Corollary 7.2 are the key results of this paper, and the other
results in this section are just consequences. One should carefully observe that
although the transformation rule in Corollary 7.2 looks like a bilipschitz property;
the mappings need not be bilipschitz in the euclidean metric. This is because the
metric µG behaves in a non-linear fashion. In the euclidean metric quasiconformal
mappings are Hölder-continuous as the results below show.
7.1. Theorem. If f : G → Rn is a non–constant qr mapping, then
(1) µf G (f (a), f (b)) ≤ KI (f ) µG (a, b) ; a, b ∈ G .
In particular, f : (G, µG ) → (f G, µf G ) is Lipschitz continuous. If N (f, G) < ∞,
then
(2) λG (a, b) ≤ KO (f ) N (f, G) λf G (f (a), f (b))
for all a, b ∈ G with f (a) 6= f (b).
7.2. Corollary. If f : G → G′ = f G is a qc mapping, then

(1) µG (a, b) KO (f ) ≤ µf G (f (a), f (b)) ≤ KI (f ) µG (a, b) ,

(2) λG (a, b) KO (f ) ≤ λf G (f (a), f (b)) ≤ KI (f ) λG (a, b)
hold for all distinct a, b ∈ G.
7.3. Theorem. Let f : Bn → Rn be a non–constant K–qr mapping with f Bn ⊂
Bn and let α = KI (f )1/(1−n) . Then
  α
(1) th 21 ρ f (x), f (y) ≤ ϕK th 12 ρ(x, y) ≤ λ1−α n th 1
2
ρ(x, y) ,
 
(2) ρ f (x), f (y) ≤ KI (f ) ρ(x, y) + log 4 ,
hold for all x, y ∈ Bn , where λn is the Grötzsch ring constant.

7.4. Corollary. Let f : Bn → Bn be a K–qr mapping with f (0) = 0 and let


α = KI (f )1/(1−n) . Then
α
(1) |f (x)| ≤ ϕK,n (|x|) ≤ λ1−α
n |x| ≤ 2
1−1/K
K|x|1/K ,
a−1  1 + |x| KI (f )
(2) |f (x)| ≤ , a= 4 ,
a+1 1 − |x|
for all x ∈ Bn .
320 M. Vuorinen IWQCMA05

7.5. Example. Let g : B2 → B2 \ {0} = gB2 be the exponential function


z+1
g(z) = exp( z−1 ), z ∈ B2 . We shall show that g : (B2 , ρ) → (gB2 , kgB2 ) fails to
be uniformly continuous. To this end, let xj = (ej − 1)/(ej + 1), j = 1, 2, . . . .
Then it follows that ρ(0, xj ) = j and thus ρ(xj , xj+1 ) = 1. Let Y = B2 \ {0}.
Since g(xj ) = exp(−ej ) we get by (4.4) and (3.20)
 
kY g(xj ), g(xj+1 ) ≥ jY g(xj ), g(xj+1 )
  
= log 1 + exp ej+1 exp(−ej ) − exp(−ej+1 )
 
= log 1 + exp(ej+1 − ej ) − 1 = ej+1 − ej → ∞
as j → ∞. In conclusion, g : (B2 , ρ) → (Y, kY ) cannot be uniformly continuous,
because ρ(xj , xj+1 ) = 1.

7.6. Theorem. Let f : Bn → Rn be a non–constant qr mapping, let E ⊂


Rn \ f Bn be a non–degenerate continuum such that ∞ ∈ E, and let G = Rn \ E
be a domain.
(1) Then f : (Bn , ρ) → (G, jG ) is uniformly continuous.
(2) If G is uniform, then f : (Bn , ρ) → (G, kG ) is uniformly continuous.

7.7. Theorem. Suppose that f : G → Rn is a bounded qr mapping and that F


α
is a compact subset of G. Let α = KI (f )1/(1−n) and C = λ1−αn d(f G)/d(F, ∂G)
where λn is the Grötzsch constant. Then f satisfies the Hölder condition
(7.8) |f (x) − f (y)| ≤ C |x − y|α
for x ∈ F , y ∈ G.

7.9. Theorem. Let f : Bn → Rn be a non–constant qr mapping.


(1) If ϕ ∈ (0, 21 π) and f Bn ⊂ C(ϕ), then for all x ∈ Bn
 1 + |x| aϕ
|f (x)| ≤ |f (0)| 4aϕ
1 − |x|
where a depends only on n and KI (f ).
(2) If f Bn ⊂ { x ∈ Rn : x21 + · · · + x2n−1 < 1 }, then for all x, y ∈ Bn
|f (x)| ≤ |f (y)| + A KI (f ) (ρ(x, y) + log 4)
where A is a positive constant depending only on n.
7.10. Theorem. Let f : Bn → Bn be a qr mapping with N (f, Bn ) = N < ∞.
Then  β
th 41 ρ f (x), f (y) ≤ 2 th 14 ρ(x, y)
holds for all x, y ∈ Bn where β = 1/(N KO (f )). Furthermore, if f (0) = 0, then
for all x ∈ Bn
|f (x)|  |x| β
p ≤2 p .
1 + 1 − |f (x)|2 1 + 1 − |x|2
Metrics and quasiregular mappings 321

7.11. Exercise. Assume that f : Bn → Bn is K–qc with f (0) = 0 and f Bn =


Bn . Show that
 p 
|f (x)|2 ≤ min ϕ2K,n (|x|) , 1 − ϕ21/K,n 1 − |x|2 ,
 p 
|f (x)|2 ≥ max ϕ21/K,n (|x|) , 1 − ϕ2K,n 1 − |x|2 .

Note that in the case n = 2 we have ϕ2K,2 (r) = 1 − ϕ21/K,2 ( 1 − r2 ) for all K > 0
and 0 < r < 1, while the analogous relation fails to hold for n ≥ 3.

The next result is a famous theorem of A. Mori from 1956 [LV]. The theorem
has, however, one esthetic flaw: it is not sharp when K = 1 . It was conjectured
in the 1960’s that the constant 16 in the theorem could be replaced by 161−1/K
and also shown in [LV] that this would be sharp. In 1988 it was proven in [FeV]
that we can replace 16 by M (K) → 1 as K → 1 . Perhaps the latest paper
dealing with the problem of reducing the constant M (K) was written by S.-L.
Qiu [Q], but as far as we know it is still an open problem whether the constant
161−1/K could be achieved. Settling this problem would be remarkable progress,
since a lot of work has been done. For the spherical chordal metric this problem
was recently discussed by P. Hästö [H3].
7.12. Theorem. Let f : B2 → B2 be a K–qc mapping with f (0) = 0 and
f B2 = B2 . Then
|f (x) − f (y)| ≤ 16 |x − y|1/K
for all x, y ∈ B2 . Furthermore, the number 16 cannot be replaced by any smaller
absolute constant.

7.13. An open problem. For K ≥ 1, n ≥ 2, and r ∈ (0, 1) let


ϕ∗K,n (r) = ϕ∗K (r) = sup{ |f (x)| : f ∈ QC K (Bn ), f (0) = 0, |x| ≤ r }
where QC K (Bn ) = { f : Bn → f Bn | f is K–qc and f Bn ⊂ Bn }. As shown in
[LV, p. 64]
(7.14) ϕ∗K,2 (r) = ϕK,2 (r) ≤ 41−1/K r1/K
for each r ∈ (0, 1) and K ≥ 1. By 7.4(1)
(7.15) ϕ∗K,n (r) ≤ ϕK,n (r) ≤ λ1−α
n rα , α = K 1/(1−n) ,
for n ≥ 2, K ≥ 1, r ∈ (0, 1). A. V. Sychev [SY, p. 89] has conjectured that
(7.16) ϕ∗K,n (r) ≤ 41−α rα
for all n ≥ 2 and K ≥ 1. Because λ2 = 4, (7.16) agrees with (7.14) for n = 2.
In [AVV4] it is shown that ϕ∗K,n 6≡ ϕK,n for n ≥ 3. It follows from 7.10 and 7.11
that
(
ϕ∗ (r) ≤ 4 r1/K ,
(7.17)  K,n 2 √
ϕ∗K,n (r) ≤ 1 − ϕ21/K,n ( 1 − r2 ) .
322 M. Vuorinen IWQCMA05

From (7.17) and (7.15) it follows, as shown in [AVV], that


2
(7.18) ϕ∗K,n (r) ≤ 41−1/K r1/K
holds for all n ≥ 2, K ≥ 1, r ∈ (0, 1). Note that the right hand side of (7.18)
is bounded when K → ∞. Recall that λn → ∞ as n → ∞ and that λ1−α n ≤
21−1/K K. Note that Sychev’s conjecture (7.16) still remains open.
7.19. Another open problem. In [Vu4], the following problem was stated.
Let QC K (Rn ) = { f : Rn → Rn | f is K–qc and f (e1 ) = e1 }. Is it true that
(7.20) sup{|g(x)| : |x| = r, g ∈ QC K (Rn )} = sup{|f (−re1 )| : f ∈ QC K (Rn )} ,
when r > 0? For n = 2 the answer is in the affirmative by [LVV].

8. Open problems
Assume that G ⊂ Rn is a proper subdomain. For what follows, we will be
interested mainly in the cases when the domain is a member of some well-known
class of domains. Some examples are uniform domains, QED-domains, domains
with uniformly perfect (in the sense of Pommerenke [Su]) boundaries and qua-
siballs, i.e. domains G of the form G = f Bn for quasiconformal f : Rn → Rn .
We denote the class of domains with D . Let us consider collection of metrics
1/(1−n)
M = {αG , hG , jG , kG , λG , µG , q, | · |} where hG refers to the hyperbolic met-
ric when n = 2. Interesting categories of mappings, we denote them by C, are
Hölder, Lipschitz, isometries, quasiisometries and identity mappings.
The problems that we list below are just examples. There are a great many
variations, by letting the domain, mapping and metric independently vary over
the categories D , C , and M .
8.1. Convexity of balls and smoothness of spheres. Fix m ∈ M . Does
there exist constant T0 > 0 such that Dm (x, T ) = {z ∈ G : m(x, z) < T }, is
convex (in euclidean geometry) for all T ∈ (0, T0 )? Is ∂Gm (x, T ) smooth for
T < T0 ?
For instance, in the case m = kG both of these problems seem to be open. In
passing, we remark that it follows from (4.4) and Theorem 4.7 (2) that when
the radius tends to 0, quasihyperbolic balls become more and more round. The
quasihyperbolic metric is used as a tool for many applications, but very little
about the metric itself is known. See the theses [MA] and [L] and also Lindén’s
paper in this volume.
8.2. Lipschitz-constant of identity mapping. For x, y ∈ Bn , x 6= y , the
following inequality holds [Vu3, (2.27)]
ρBn (x, y) ρBn (x, y)
|x − y| ≤ 2 th < .
4 2
We may now regard this result as an inequality for the modulus of continuity of
id : (Bn , ρBn ) → (Bn , | · |) . Instead of considering the identity mapping we could
now take any mapping in our category of mappings and consider the problem
Metrics and quasiregular mappings 323

of estimating the modulus of continuity between any two metric spaces in our
category of metric spaces, see [Vu1], [S]. We list several particular cases of our
problem. For G = Rn \ {0} does there exist constants A or B such that for all
x, y ∈ G
q(x, y) ≤ AkG (x, y),
and
1/(1−n)
q(x, y) ≤ BλG (x, y) ?
For G = C \ {0, 1} does there exist constant C such that for all x, y ∈ G
q(x, y) ≤ ChG (x, y) ,

For G = Rn \ {0} does there exist a constant E such that for all x, y ∈ G
1/(1−n)
λG (x, y) ≤ EjG (x, y) ?
8.3. Characterization of isometries and quasiisometries. Given two met-
ric spaces in our category of spaces, does there exist a quasiisometry, mapping
the one space onto the other space? Again, we could consider, in place of quasi-
isometries, any other map in our category of maps.
1/(1−n)
What is the modulus of continuity of id : (G, µG ) → (G, λG )?
1/(1−n) 1/(1−n)
Is a quasiisometry f : (G, λG →
) (f G, λf G )
quasiconformal? J. Lelong-
Ferrand raised this question in [LF] and the question was answered in the nega-
tive in [FMV] . There it was also shown that the answer is affirmative under the
1/(1−n) 1/(1−n)
stronger requirement that f : (D, λD ) → (f D, λf D ) be uniformly contin-
uous for all subdomains D of G . However, it is not known what the isometries
are.
Are isometries f : (G, αG ) → (f G, αf G ) Möbius transformations? (see Beardon
[BE2], Hästö and Ibragimov [HI] and also Hästö’s paper in this volume).
8.4. Conformal invariants. The conformal invariant p(x) is relatively well-
known. See [HV] for further information. However, much less is known about
the invariants µG and λG . For domains whose boundaries are uniformly perfect
(in the sense of Pommerenke), there are some inequalities for µG in terms of jG ,
see [Vu2] and [JV]. Some results for λG when G = Bn \ {0} , were proved in [H]
and [BV]. But even the basic question of finding a formula for λB2 \{0} (x, y) is
open.

Some of these problems may be hard, some are very easy. Because of the
very general setup, it would require some effort even to single out the interesting
combinations of domains in D , mappings in C , and metrics in M .

References
[A] L. V. Ahlfors: Collected papers. Vol.1 and 2. Edited with the assistance of Rae
Michael Shortt. Contemporary Mathematicians. Birkhäuser, Boston, Mass., 1982.
xix+515 pp., xix+520 pp., ISBN: 3-7643-3076-7, ISBN: 3-7643-3075-9
324 M. Vuorinen IWQCMA05

[AVV1] G. D. Anderson, M. K. Vamanamurthy and M. Vuorinen: Sharp distortion


theorems for quasiconformal mappings. Trans. Amer. Math. Soc. 305 (1988), 95–111.
[AVV2] G. D. Anderson, M. K. Vamanamurthy, and M. Vuorinen: Conformal invari-
ants, inequalities and quasiconformal mappings. J. Wiley, 1997, 505 pp.
[BE1] A. F. Beardon: The geometry of discrete groups. Graduate Texts in Math. Vol. 91,
Springer-Verlag, Berlin–Heidelberg–New York, 1982.
[BE2] A. F. Beardon: The Apollonian metric of a domain in Rn . Quasiconformal mappings
and analysis (Ann Arbor, MI, 1995), 91–108, Springer, New York, 1998.
[B] D. Betsakos: On conformal capacity and Teichmüller’s modulus problem in space.
J. Anal. Math. 79 (1999), 201–214.
[BV] D. Betsakos and M. Vuorinen: Estimates for conformal capacity, Constr. Ap-
prox. 16 (2000), 589–602.
[FeV] R. Fehlmann and M. Vuorinen: Mori’s theorem for n-dimensional quasiconformal
mappings. Ann. Acad. Sci. Fenn. Ser. A I 13 (1988), 111–124.
[F] J. Ferrand: Conformal capacities and extremal metrics, Pacific J. Math. 180 (1997),
41–49.
[FMV] J. Ferrand, G. Martin, and M. Vuorinen: Lipschitz conditions in conformally
invariant metrics. J. Anal. Math. 56 (1991), 187–210.
[G] F.W. Gehring: Quasiconformal mappings in Euclidean spaces. Handbook of com-
plex analysis: geometric function theory. Vol. 2, ed. by R. Kühnau, 1–29, Elsevier,
Amsterdam, 2005.
[GM] F. W. Gehring and O. Martio: Quasiextremal distance domains and extension
of quasiconformal mappings. J. Anal. Math. 45 (1985), 181–206.
[GO] F. W. Gehring and B. G. Osgood: Uniform domains and the quasi–hyperbolic
metric. J. Anal. Math. 36 (1979), 50–74.
[GP] F. W. Gehring and B. P. Palka: Quasiconformally homogeneous domains. J.
Anal. Math. 30 (1976), 172–199.
[H1] P. A. Hästö: The Apollonian metric: uniformity and quasiconvexity. Ann. Acad.
Sci. Fenn. Math. 28 (2003), no. 2, 385–414.
[H2] P. A. Hästö: The Apollonian metric: quasi-isotropy and Seittenranta’s metric. Com-
put. Methods Funct. Theory 4 (2004), no. 2, 249–273.
[H3] P. A. Hästö: Distortion in the spherical metric under quasiconformal mappings.
Conform. Geom. Dyn. 7 (2003), 1–10 (electronic).
[HI] P. A. Hästö and Z. Ibragimov: Apollonian isometries of planar domains are
Möbius mappings. J. Geom. Anal. 15 (2005), no. 2, 229–237.
[He] V. Heikkala: Inequalities for conformal capacity, modulus, and conformal invariants.
Ann. Acad. Sci. Fenn. Math. Diss. No. 132 (2002), 62 pp.
[HV] V. Heikkala and M. Vuorinen: Teichmüller’s extremal ring problem.- Math. Z.
(to appear) and Preprint 352, April 2003, University of Helsinki, 20 pp.
[H] J. Heinonen: Lectures on Analysis on Metric Spaces. Springer, 2001.
[HB] D. A. Herron and S. M. Buckley: Uniform domains and capacity. Manuscript,
2005, 20pp.
[JV] P. Järvi and M. Vuorinen: Uniformly perfect sets and quasiregular mappings. J.
London Math. Soc. (2) (1996), 515–529.
[J] J. A. Jenkins: On metrics defined by modules, Pacific J. Math. 167 (1995), 289–292.
[K] R. Kühnau, ed.: Handbook of complex analysis: geometric function theory. Vol. 1.
and Vol. 2, Elsevier Science B.V., Amsterdam, 2002. xii+536 pp, ISBN 0-444-82845-1
and 2005, xiv+861 pp. ISBN 0-444-51547-X .
[LV] O. Lehto and K. I. Virtanen: Quasiconformal mappings in the plane. Die
Grundlehren der math. Wissenschaften Vol. 126, Second ed., Springer-Verlag, Berlin–
Heidelberg–New York, 1973.
Metrics and quasiregular mappings 325

[LVV] O. Lehto, K. I. Virtanen and J. Väisälä: Contributions to the distortion theory


of quasiconformal mappings. Ann. Acad. Sci. Fenn. Ser. A I No. 273 (1959) 14 pp.
[LF] J. Lelong-Ferrand: Invariants conformes globaux sur les varietes riemanniennes,
J. Differential Geom. 8 (1973), 487–510.
[L] H. Lindén: Quasihyperbolic geodesics and uniformity in elementary domains. Ann.
Acad.Sci. Fenn. Math. Diss. No 146, (2005), 50 pp.
[MA] G. Martin: Quasiconformal and bilipschitz mappings, uniform domains and the
hyperbolic metric. Trans. Amer. Math. Soc. 292 (1985), 169–192.
[Q] S.-L. Qiu: On Mori’s theorem in quasiconformal theory. A Chinese summary appears
in Acta Math. Sinica 40 (1997), no. 2, 319. Acta Math. Sinica (N.S.) 13 (1997), no. 1,
35–44.
[QV] S.-L. Qiu and M. Vuorinen: Special functions in geometric function theory. Hand-
book of complex analysis: geometric function theory. Vol. 2, ed. by R. Kühnau, 621–
659, Elsevier, Amsterdam, 2005.
[S] P. Seittenranta: Möbius-invariant metrics. Math. Proc. Cambridge Philos. Soc.
125, 1999, 511–533.
[So] A. Yu. Solynin: Moduli of doubly-connected domains and conformally invariant
metrics (in Russian), Zap. Nautsh. Semin. LOMI, tom 196 (1991), 122–131, Sankt
Peterburg “Nauka,” 1991.
[SoV] A. Yu. Solynin and M. Vuorinen: Extremal problems and symmetrization for
plane ring domains, Trans. Amer. Math. Soc. 348 (1996), 4095–4112.
[Su] T. Sugawa: Uniformly perfect sets: analytic and geometric aspects [translation of
Sūgaku 53 (2001), no. 4, 387–402]. Sugaku Expositions. Sugaku Expositions 16 (2003),
no. 2, 225–242.
[SY] A. V. Sychev: Moduli and n–dimensional quasiconformal mappings. (Russian). Iz-
dat. “Nauka”, Sibirsk. Otdelenie, Novosibirsk, 1983.
[T1] O. Teichmüller: Untersuchungen über konforme und quasikonforme Abbildung,
Deutsche Math. 3 (1938), 621–678.
[T2] O. Teichmüller: Gesammelte Abhandlungen, ed. by L. V. Ahlfors and F. W.
Gehring, Springer-Verlag, Berlin, 1982.
[V1] J. Väisälä: Lectures on n–dimensional quasiconformal mappings. Lecture Notes in
Math. Vol. 229, Springer-Verlag, Berlin–Heidelberg–New York, 1971.
[V2] J. Väisälä: Domains and maps. Quasiconformal space mappings, 119–131, Lecture
Notes in Math., 1508, Springer, Berlin, 1992.
[Vu1] M. Vuorinen: Conformal invariants and quasiregular mappings. J. Anal. Math. 45
(1985), 69–115.
[Vu2] M. Vuorinen: On quasiregular mappings and domains with a complete conformal
metric. Math. Z. 194 (1987) 459–470.
[Vu3] M. Vuorinen: Conformal geometry and quasiregular mappings. (Monograph, 208
pp.). Lecture Notes in Math. Vol. 1319, Springer-Verlag, 1988.
[Vu4] M. Vuorinen: Quadruples and spatial quasiconformal mappings. Math. Z. 205
(1990), no. 4, 617–628.
[Vu5] M. Vuorinen: Quasiconformal images of spheres. Mini-Conference on Quasiconfor-
mal Mappings, Sobolev Spaces and Special Functions, Kurashiki, Japan, 2003-01-08,
available at
http://www.cajpn.org/complex/conf02/kurashiki/vuorinen.pdf

Matti Vuorinen E-mail: vuorinen@utu.fi


Address: University of Turku
Proceedings of the International Workshop on Quasiconformal
Mappings and their Applications (IWQCMA05)

Circle Packings, Quasiconformal Mappings, and


Applications

G. Brock Williams

Abstract. We provide an overview of the connections between circle packings


and quasiconformal mappings, with particular attention to applications to
string theory and image recognition.

Keywords. Circle Packing, Quasiconformal Maps.

2000 MSC. 52C26, 30F60.

Contents

1. Introduction 328
2. Quasiconformal Maps 328
2.1. Analytic Definition of Quasiconformality 328
2.2. Geometric Definition of Quasiconformality 329
2.3. An Important Example 330
3. Conformal Welding 331
3.1. Quasisymmetries and Quasicircles 331
3.2. Conformal Welding Theorem 332
4. Circle Packing 333
4.1. Definitions and Examples 333
4.2. Packings and Maps 335
4.3. The Rodin-Sullivan Theorem 335
5. Applications 338
5.1. Image Recognition 338
5.2. Radnell-Schippers Quantum Field Theory 341
5.3. Circle Packing Measurable Riemann Mapping Theorem 342
References 343

Version June 10, 2006.


Supported by NSF Grant #0536665.
328 G. Brock Williams IWQCMA05

1. Introduction
The deep connections between the combinatorial and geometric properties of
circle packings and the analytic properties of the maps they induce have been the
subject of intense study in recent years. In 1985, William Thurston conjectured,
and Burt Rodin and Dennis Sullivan proved, that maps between circle packings
were nearly analytic [Thu85, RS87]. Since then the study of circle packings has
exploded to impact a great many other fields including conformal mapping [HS93,
HS96,Ste05], complex analysis [BS91,DS95a,Ste97,Ste02,Ste03] Teichmüller the-
ory [BS90, Bro96, Wil01b, BW02, Wil03, BS04b], brain mapping [Bea99, Kra99],
random walks [Ste96, Dub97, HS95, McC98, DW05], tilings [BS97, Rep98], mini-
mal surfaces and integrable systems [BS04a], numerical analysis [Moh93, CS99],
metric measure spaces [BK02] and much more.
The fundamental folk theorem of circle packing is that “packings desperately
want to be conformal.” They react to combinatorial or geometric changes in
precisely the same way as conformal maps. Maps between packings seem de-
termined to approximate conformal maps. There is, however, much to be said
about the relationships between circle packings and quasiconformal maps. It is
principally with these connections and the applications arising from them that
we will concern ourselves in this paper.
After some initial background on quasiconformal mappings in Section 2, we
describe the crucial concept of conformal welding in Section 3. We review the
fundamental concepts of circle packing in Section 4, and then describe three
applications of circle packings and quasiconformal maps in Section 5. Namely,
we discuss the use of packings in image recognition, in implementing Radnell-
Schippers quantum field theory, and in constructing quasiconformal maps.

2. Quasiconformal Maps
2.1. Analytic Definition of Quasiconformality. Quasiconformal mappings
form the heart of Teichmüller theory as developed in the 1950’s and 1960’s.
They are the natural generalization of analytic functions. For more detailed
explanations, a number of excellent resources are available, including [Ahl66,
LV73, Leh87, Nag88, IT92, GL00].
Definition 2.1. A homeomorphism f ∈ L2 is quasiconformal if
(2.1) ∂z f = µ∂z f
for some µ ∈ L∞ , ||µ||∞ < 1. Recall the complex partial derivatives are defined
by
1
∂z f = (∂x f + ∂y f )
2
1
∂z f = (∂x f − ∂y f ) .
2
Circle Packings, QC Maps, and Applications 329

Figure 1. A geometric measure of quasiconformality. The quo-


tient of the length of the dashed lines measures how close the curve
on the right is to being a circle.

Equation 2.1 is called the Beltrami equation and µ, a Beltrami differ-


ential. Notice that when µ ≡ 0, the Beltrami equation becomes ∂z f ≡ 0,
which when separated into real and imaginary parts is precisely the familiar
Cauchy-Riemann equations. Thus µ determines how “quasi” a quasiconformal
map really is. This measure of the “quasi-ness,” or distortion of a map is most
often expressed in terms of the dilatation
1 + ||µ||∞
K= ≥1
1 − ||µ||∞
of the map. A quasiconformal map f with dilatation K is called a K-quasiconformal
map; a 1-quasiconformal map is thus conformal.
The Beltrami differential µ corresponding to a quasiconformal map is often
called its complex dilatation. Notice, however, that the complex dilatation is
a complex function and actually measures the distortion of f at every point in its
domain. The dilatation, on the other hand, is a single real number and provides
a global bound on the distortion of f over the entire domain.

2.2. Geometric Definition of Quasiconformality. An equivalent measure


of the distortion of a quasiconformal map is provided by the dilatation quotient
supθ |f (z + reiθ ) − f (z)|
Df (z) = lim sup .
r→0+ inf θ |f (z + re ) − f (z)|

The dilatation quotient has a simple geometric interpretation. If we consider a


small circle of radius r about z in the domain, it will be mapped to some curve
about f (z) in the range. The dilatation quotient is then the ratio of the maximal
to the minimal distance from f (z) to this curve. See Figure 1.
Recall that conformal maps preserve angles; moreover, if f ′ (z) = reiθ 6= 0,
then
df = f ′ (z) dz = reiθ dz.
Thus infinitesimally, f acts geometrically like
z 7→ reiθ z + C
330 G. Brock Williams IWQCMA05

for some number C; that is, f acts like the composition of a scaling, rotation, and
translation. This not only explains the reason analytic maps with non-vanishing
derivative preserve angles, but also implies that they must map infinitesimal cir-
cles to infinitesimal circles. Consequently, the dilatation quotient of a conformal
map is identically 1.
It turns out that the dilatation of a quasiconformal map is nothing more than
the supremum of the dilatation quotient over the domain. Thus we have the
following equivalent definition of quasiconformality.
Definition 2.2. A homeomorphism f is K-quasiconformal if it is absolutely
continuous on lines and
Df (z) ≤ K
for all z in its domain.

2.3. An Important
  Example. If we think of the complex plane as R2 and
x
x + iy as , then it is natural to consider the effect of linear and affine
y
transformations. Suppose
    
a b x e
(2.2) f (x + iy) = + ,
c d y f
where ad − bc 6= 0.
A moment’s linear algebra shows f can be re-written as
 
ax + by + e
(2.3) f (x + iy) = .
cx + dy + f
Then
     
1 a −d 1 a+d
(2.4) ∂z f = − =
2 c b 2 c−b

     
1 a −d 1 a−d
∂z f = + = .
2 c b 2 c+b

Notice that ∂z f = 0 ifand only if a = d and c = −b, in which case, mul-


a b
tiplication by the matrix is equivalent to multiplication by the complex
c d
number a + ib.
In general, however, we will have
∂z f (a − d) + i(c + b)
µ= = ,
∂z f (a + d) + i(c − b)
and f will be quasiconformal.
Geometrically, f will map the basis vectors 1 and i to a + ic and b + id,
respectively, and then translate by e + if . It is easy to check that µ = 0 if and
Circle Packings, QC Maps, and Applications 331

only if the new basis vectors a + ic and b + id are perpendicular, and |µ| increases
toward 1 as the angle decreases toward 0.
Notice that affine maps have constant complex dilatation; conversely, if µ is
constant, it is a simple exercise to solve for the affine map whose dilatation is µ.
The importance of this example becomes apparent when we consider the infin-
itesimal behavior of any quasiconformal map. Just as we observed that confor-
mal maps act infinitesimally by rotation, scaling, and translation, quasiconformal
maps act infinitesimally as affine maps.

3. Conformal Welding
3.1. Quasisymmetries and Quasicircles. We continue our exploration of
quasiconformal maps with an investigation of their boundary values [BA56,LV73,
DE86, LP88, GL00]. Note that when maps extend continuously or smoothly to
the boundary, we will use same notation for the extended maps.
Definition 3.1. A homeomorphism ϕ : ∂D → ∂D is quasisymmetric or a
quasisymmetry if it is the boundary function of some quasiconformal map of
D onto itself.

As might be expected, quasisymmetries have a beautiful geometric character-


ization as well [BA56, LV73, Leh87, Krz87].
Definition 3.2. An orientation preserving homeomorphism ϕ : ∂D → ∂D is a
k - quasisymmetry if
1 |ϕ(I)|D
≤ ≤k
k |ϕ(J)|D
for any two adjacent intervals (subarcs) I and J of ∂D having equal length
|I|D = |J|D .

Essentially, this definition says quasisymmetries can’t map adjacent symmetric


intervals to extremely non-symmetric intervals.
Next, we temporarily leave quasisymmetries to consider the effect of quasicon-
formal maps on circles. However, as we will see, these quasicircles are intimately
connected to quasisymmetries.
Definition 3.3. A Jordan curve Γ is a K-quasicircle if it is the image of the
unit circle under a K-quasiconformal map of C onto itself.

As might be expected by now, quasicircles have both analytic and geometric


definitions [Ahl63, Ahl66].
Definition 3.4. A Jordan curve Γ is a quasicircle if there exists R > 1 so that
for all points x, y ∈ Γ
diam(Γx,y ) ≤ R|x − y|,
where Γx,y is the sub-arc of Γ connecting x and y which has the smaller diameter.
332 G. Brock Williams IWQCMA05

g
Figure 2. If Γ is a Jordan curve, then the Riemann Mapping
Theorem promises the existence of a conformal map f from the
inside of Γ to the inside of the unit disc D. Similarly, there exists
a conformal map g from the outside of Γ to the outside of the unit
disc.

f −1

Figure 3. Since f and g extend to the boundary, they induce a


homeomorphism ϕ = g ◦ f −1 : ∂D → ∂D.

Loosely speaking, this condition limits “pinching” – a quasicircle cannot visit a


point x, wander far away, and then return to a point very near x. Fred Gehring’s
monograph [Geh82] contains an extensive list of these and other characterizations
of quasicircles.

3.2. Conformal Welding Theorem. The intimate connection between qua-


sisymmetries and quasicircles is illustrated by the following two theorems [Pfl51,
LV73, Leh87, GL00].
Theorem 3.5. Suppose Γ is Jordan curve dividing the plane into complemen-
tary components Ω and Ω∗ . Let f : Ω → D and g : Ω∗ → D∗ be conformal
homeomorphisms, the existence of which are promised by the Riemann Mapping
Theorem. Then f and g extend to homeomorphisms of the boundary and
g ◦ f −1 : ∂D → ∂D
is a quasisymmetry if Γ is a quasicircle. See Figures 2 and 3.
Circle Packings, QC Maps, and Applications 333

The converse is also true. Given a quasisymmetry ϕ : ∂D → ∂D, we can  glue D


iθ iθ

and D together by attaching points e ∈ ∂D to their image points ϕ e ∈ ∂D∗ .
The result is a topological sphere. As D and D∗ struggle to fit together after the
welding, the “seam” between them will be pushed and pulled into a quasicircle.
Conformal Welding Theorem. Let ϕ : ∂D → ∂D be a quasisymmetry. Then
ϕ induces a conformal welding of D and D∗ . That is, there exist conformal
maps f : Ω → D and g : Ω∗ → D∗ of complementary Jordan domains in C with
boundary values satisfying
g ◦ f −1 (eiθ ) = ϕ(eiθ ).
Moreover, the Jordan curve Γ = f −1 (∂D) = g −1 (∂D∗ ) is unique up to Möbius
transformations.
For quasisymmetries defined on ∂D, it is customary to normalize our welding
maps so that f −1 (1) = g −1 (1) = ϕ(1) = 1, f (0) = 0, and g(∞) = ∞. With these
normalizations, the maps f and g and the curve Γ are unique.

4. Circle Packing
4.1. Definitions and Examples. Since William Thurston’s work in the mid-
1980’s, the connections between circle packings and analytic functions have been
widely studied. More detailed information is contained in the rapidly expanding
literature, including several recent survey articles [DS95b,Ste97,Ste02,Ste03] and
Ken Stephenson’s excellent new book [Ste05].
Definition 4.1. A CP-complex K is an abstract simplicial 2-complex such
that
1. K is simplicially equivalent to a triangulation of an (orientable) surface.
2. Every boundary vertex of K has an interior neighbor.
3. The collection of interior vertices is nonempty and edge-connected.
4. There is an upper bound on the degree of vertices in K.
The restrictions imposed by conditions 2 through 4 are extremely mild and
are met by most any reasonable triangulation.
Notice that a CP-complex is a purely combinatorial object. It possesses no
geometric structure until it is embedded in a surface by a circle packing. To
emphasize this fact, we will often refer to a CP-complex simply as an abstract
triangulation.
Definition 4.2. A circle packing is a configuration of circles with a specified
pattern of tangencies. In particular, if K is a CP-complex, then a circle packing
P for K is a configuration of circles such that
1. P contains a circle Cv for each vertex v in K,
2. Cv is externally tangent to Cu if [v, u] is an edge of K,
3. hCv , Cu , Cw i forms a positively oriented mutually tangent triple of circles if
hv, u, wi is a positively oriented face of K.
334 G. Brock Williams IWQCMA05

Figure 4. A finite circle packing (left). The underlying trian-


gulation can be recovered by connecting centers of tangent circles
with line segments (middle). The resulting collection of triangles
forms the carrier of the packing (right).

Figure 5. A portion of the “regular hex” packing. Notice that


every circle has the same radius.

A packing is called univalent if none of its circles overlap, that is, if no pair of
circles intersect in more than one point.

A univalent circle packing produces a geometric realization of its underlying


complex. Vertices can be embedded as centers of their corresponding circles,
and edges can be realized as geodesic segments joining centers of circles. The
collection of triangles embedded in this way is called the carrier of the packing,
written carr P . See Figure 4.
Example 4.3. William Thurston’s original interest in packings began with the
infinite “regular hex packing” in which every circle touches exactly 6 others. He
showed that the only univalent packing with this combinatorial pattern is the
one in which every circle has the same radius. (It remains an open question to
characterize the non-univalent ones.) See Figure 5.
Circle Packings, QC Maps, and Applications 335

Figure 6. A portion of the “ball bearing” packing. The carrier


has been drawn in to emphasize the lattice structure.

Example 4.4. Another useful infinite packing is the “ball bearing packing”
named by Tomasz Dubejko and Ken Stephenson [DS95b]. The underlying tri-
angulation is created from a lattice, and the original lattice structure is still
apparent in the resulting packing. Consequently, the carrier of the packing can
be decomposed into small squares. Moreover, there is a natural refinement of
the triangulation and carrier created by replacing each square with four copies
of the original. See Figure 6.

4.2. Packings and Maps. The connection between circle packings and func-
tion theory arises from the investigation of maps between the carriers of two
different packings for the same abstract complex. That is, suppose P and P e are
both Euclidean circle packings for the same underlying complex K. Then every
face in K is realized as both a Euclidean triangle T in carr P and a triangle Te
e It is easy now to construct an affine map between triangles T and Te.
in carr P.
If we translate one vertex of each to the origin, then the two edges meeting at
the origin form a basis for R2 and can be mapped one onto the other by a linear
map.
e by a piecewise
Thus the entire carrier of P can be mapped onto the carrier of P
affine map defined triangle by triangle. Notice that the individual triangle maps
agree on adjacent edges, so the complete map is continuous. Circle packing maps
constructed in this way are called discrete conformal maps. See Figure 7.

4.3. The Rodin-Sullivan Theorem. Recall from Section 2.3, that affine maps
are quasiconformal. The dilation on each triangle will be constant and depend
only by the difference between corresponding angles. If there are only finitely
many circles in the packings, the dilatation of a discrete conformal map will
be finite and depend only on the maximal difference in corresponding angles
between triangles in the two carriers.
336 G. Brock Williams IWQCMA05

Figure 7. Two circle packings with the same underlying trian-


gulation. The carrier for each is indicated and one pair of corre-
sponding triangles are shaded. Each triangle in the carrier on the
left can be mapped via an affine map to its corresponding triangle
in the carrier on the right.

At this point in our story, we come to Burt Rodin and Dennis Sullivan’s Ring
Lemma, the first connection between the analytic properties of discrete conformal
maps and the combinatorial properties of packings [RS87].
Ring Lemma. In a univalent packing, there is a lower bound Cn on the ratio of
the radius of any interior circle to the radius of any of its neighbors. This bound
depends only on the degree n (the number of neighbors) of the circle.

The sharp value of the bound Cn was determined by Dov Aharonov [Aha97].
Lemma 4.5. If {an } is the Fibonacci sequence, then
1
Cn = 2 .
an−2 + a2n−1 − 1
Cn
Moreover, converges to the square of the golden ratio.
Cn+1
The Ring Lemma thus connects a purely combinatorial property of the packing
(the degree) with a geometric property of the packing (the ratio of the radii of
adjacent circles). This geometric constraint on the circles implies angles in the
carrier must be bounded away from 0 and π. Hence there is a uniform bound
on the difference between corresponding angles in the carriers of two packings
with the same underlying triangulation. Consequently, the associated discrete
conformal map is quasiconformal with a bound on the dilatation determined
only the degree. In this way, a combinatorial property of the triangulation leads
directly to an analytic property of the associated discrete conformal maps.
In 1985, William Thurston conjectured the relationships between the combina-
torics, geometry, and mapping properties of packings run much deeper [Thu85].
Circle Packings, QC Maps, and Applications 337

Figure 8. A cross-shaped packing (left) which has been re-packed


in the unit disc (right). Since both packings share the same under-
lying triangulation, there is discrete conformal map between them
which approximates the classical Riemann map.

Suppose Ω ( C is a bounded simply connected region and p, q ∈ Ω, p 6= q. The


Riemann Mapping Theorem implies there is a unique conformal map f : Ω → D
with f (p) = 0 and f (q) > 0.
Now suppose Pn is a sequence of packings in Ω with mesh (radius of the largest
circle) decreasing to 0 and carr Pn → Ω as n → ∞. Let Kn be the underlying
triangulation of Pn . Paul Koebe [Koe36], E. M. Andreev [And70a, And70b], and
William Thurston [Thu] independently proved that any finite, simply connected
CP-complex (such as Kn ) can be realized by a packing in D which is “maximal”
in the sense that boundary circles are tangent to ∂D. This maximal, or Andreev,
packing is unique up to disc automorphisms.
Thus for each Pn ⊂ Ω, there is a maximal packing P en ⊂ D with the same
underlying triangulation Kn as Pn . Moreover, we can normalize P en so that
if Cp and Cq are the nearest circles in Pn to p and q, respectively, then the
corresponding circles C ep and C
eq in Pen are centered at 0 and on the positive real
axis, respectively.
Since Pn and P en share the same underlying triangulation, there is a discrete
conformal map
fn : carr Pn → carr Pen .
William Thurston conjectured that fn → f locally uniformly on Ω as n →
∞ [Thu85]. This was quickly proven by Burt Rodin and Dennis Sullivan [RS87].
See Figure 8.
Rodin-Sullivan Theorem. The discrete conformal maps described above con-
verge locally uniformly to the conformal map f : Ω → D with f (p) = 0 and
f (q) > 0.
Recall that if the degree of Kn is uniformly bounded for all n (Thurston’s
original conjecture was for packings with degree 6), then the Ring Lemma implies
338 G. Brock Williams IWQCMA05

each fn will be K-quasiconformal, with K independent of n. It remains to show


that the dilatation of fn must actually decrease to 1 as n → ∞. This follows
from the uniqueness of infinite packings.
Theorem 4.6. Every infinite, simply connected CP-complex has a packing in
either C or D. This packing is unique up to conformal automorphisms.

Various versions of Theorem 4.6 have been proven. Thurston’s original proof
was only for the regular hex packing of Example 4.3 and relied on deep results
from the theory of hyperbolic 3-manifolds [Thu]. Later improvements by Ken
Stephenson [Ste96], Alan Beardon and Ken Stephenson [BS90], Yves Colin de
Verdiére [dV89, dV91], Zheng-Xu He and Burt Rodin [HR93], and Zheng-Xu He
and Oded Schramm [HS96, HS98] utilized probabilistic techniques, variational
principles, the Perron method, or elementary topology.
The effect of Theorem 4.6 is to force the dilation of fn to decrease to 1 as
n → ∞. Consider a circle C “deep inside” Pn , that is, separated from ∂Ω by a
great many generations of other circles. If C is far enough from the boundary,
it can hardly tell if it is part of a finite packing, or the unique infinite one. The
same must be true for the corresponding circle C e in Pen ⊂ D. Thus triangles in
carr Pn and carr P en which are far from the boundary, must be nearly the same
(up to scaling, translation, and rotation). In particular, the corresponding angles
must be nearly the same, and the resulting affine map must be nearly conformal.
This is usually stated as the Packing Lemma [Ste96, Ste05].
Packing Lemma. Suppose Kn is a sequence of simply connected CP-complexes
with uniformly bounded degree and having univalent packings Pn in a bounded
simply connected domain Ω. If P en is any other sequence of univalent packings
for Kn , then the maximum difference between corresponding angles in carr Pn
and carr Pen goes to 0 locally uniformly as n → ∞.

Finally, recall that we assumed the mesh of Pn decreased to 0 as n → ∞;


thus on compact subsets of Ω, the number of generations of circles between
the compact subset and the boundary must go uniformly to infinity as n → ∞.
Consequently, the dilatation of fn will decrease to 1 uniformly on compact subsets
of Ω.

5. Applications
5.1. Image Recognition. In work with Ken Stephenson, we have applied cir-
cle packing techniques to two-dimensional image recognition problems. David
Mumford and Eitan Sharon have recently developed a technique for studying
two-dimensional shapes (Jordan curves) by means of the Weil-Peterson metric on
their associated welding homeomorphisms [MS04]. They restrict their attention
to smooth curves which then produce diffeomorphisms of ∂D. The Weil-Peterson
metric on these diffeomorphisms is invariant under Möbius transformations; thus
shapes which differ only by scaling or rotation are recognized as being the same.
Circle Packings, QC Maps, and Applications 339

(6.283,6.283)

(0.000,0.000)

Figure 9. A cross-shaped quasicircle (left) and the graph of the


resulting quasisymmetry, parametrized as a map from [0, 2π] onto
[0, 2π].

It is relatively easy to extend their program to shapes bounded by quasicircles


and to quasisymmetric maps on ∂D. By packing both the inside Ω and outside
Ω∗ of a quasicircle Γ, then repacking in D and D∗ , respectively, we can create
discrete analytic functions
fn :Ωn → D
gn :Ω∗n → D∗ ,
where Ωn → Ω and Ω∗n → Ω∗ .
It is much trickier to compare the boundary values of fn and gn since the
packings in Ω and Ω∗ don’t necessarily match up on the boundary. However, it
is possible with careful application of the geometry of quasicircles and a dash a
topology to create a map
ϕn : ∂D → ∂D
−1
which is essentially given by gn ◦fn . We then have the following theorem [Wil01a]:
Theorem 5.1. The mappings ϕn converge uniformly to the quasisymmetry ϕ
induced by the quasicircle Γ. Moreover, fn and gn converge locally uniformly to
the Riemann maps f : Ω → D and g : Ω → D∗ , respectively.

For example, consider the cross-shaped curve in Figure 9. Creating discrete


conformal maps as described above (Recall Figure 8), we can approximate the
corresponding quasisymmetry.
Repeating this procedure for a T-shaped curve and a hand-drawn cross in
Figures 10 and 11, the similarities and differences with the straight-sided cross
are easy to see.
A more difficult problem is to recover the shape given the map ϕ : ∂D →
∂D. The Conformal Welding Theorem guarantees that this is possible, but is
no help in actually computing the shape. Again, circle packing comes to the
340 G. Brock Williams IWQCMA05

(6.283,6.283)

(0.000,0.000)

Figure 10. A T-shaped quasicircle (left) and the graph of the


resulting quasisymmetry (right).

(6.283,6.283)

(0.000,0.000)

Figure 11. A hand-drawn cross (left) and the graph of the re-
sulting quasisymmetry (right). Compare with Figures 9 and 10.

Figure 12. A discrete welding for the map ϕ(eiθ ) = ei(θ+ 3 sin(3θ)) .
1

The circles corresponding to the “seam” in packing (left) are


shaded. The packing provides a realization on S 2 of the weld-
ing triangulations (middle). The edges along the “seam” form the
discrete welding curve (right).
Circle Packings, QC Maps, and Applications 341

rescue. Instead of welding D to D∗ , we will weld triangulations of discs. For


example, if ϕ is a homeomorphism from the boundary of an triangulation K to
the boundary of K∗ , we use ϕ to glue the triangulations together. After a few
minor refinements and adjustments, we attach every boundary edge e of K to its
image ϕ(e). This discrete welding then yields a triangulation K e of a sphere.
e
The welded triangulation K can be realized by a unique circle packing on S 2 .
The uniqueness of this packing is exactly analogous to the uniqueness of the
conformal structure on S 2 . The circles must push and pull against each other to
settle in locations compatible with the global pattern provided by K e in precisely
the same way that two welded discs settle in locations compatible with the global
conformal structure on S 2 . This circle packing provides a geometric realization
of the formerly purely combinatorial welding. In particular, the “seam” between
the original triangulations is realized as a polygonal Jordan curve, a discretized
version of the conformal welding curve.
Notice also that K e contains a copy of both K and K∗ . Thus we can define
discrete analytic functions from K and K∗ onto their copies in K. e This is, of
course, analogous to the existence of classical welding maps f and g onto com-
plementary regions of S 2 . Moreover, because of the way we used ϕ to weld K e
−1
together, a version of the welding condition g ◦ f = ϕ also holds.
In fact, the discrete version is more than just analogous to the classical case –
it converges to it as well. Welding finer and finer triangulations using the same
quasisymmetric map produces discrete welding curves that converge uniformly to
the classical conformal welding curve. Moreover, the discrete analytic functions
converge locally uniformly to the classical conformal welding maps [Wil04].
Discrete Conformal Welding Theorem. Given a quasisymmetric map ϕ :
∂D → ∂D, our construction produces discrete analytic functions {fn } and {gn }
converging locally uniformly to the conformal welding maps f and g induced by
ϕ. Moreover, the discrete conformal welding curves Γn converge uniformly to the
quasicircle Γ induced by ϕ.
5.2. Radnell-Schippers Quantum Field Theory. Recently David Radnell
and Eric Schippers [RS05] have developed a two-dimensional quantum field the-
ory based on conformal welding and rigged Teichmüller spaces. Very briefly, one
of fundamental ideas of string theory is that a one-dimensional closed string will
sweep out a surface, called its world sheet, as it travels through time. As a
string breaks apart and rejoins with itself, it alters the topology of the world
sheet. See Figure 13. Dennis Sullivan and Moira Chas have in this manner de-
scribed the topology of all world sheets in terms of the splitting and joining of
strings [Sul01].
While the topology of the world sheet captures the splitting and re-joining of a
string, it is the conformal structure of the world sheet that captures features such
as the relative size of the string and length of time between splittings and joinings.
Thus for many computations it is necessary to consider all possible conformal
structures on all possible surfaces. The Universal Teichmüller space contains
342 G. Brock Williams IWQCMA05

Figure 13. A depiction of a string traveling through time. As


the string breaks into two pieces and then rejoins, a topological
handle is created.

the Teichmüller spaces of all Riemann surfaces and as such has recently gained
the attention of physicists as a possible setting for string theory computations
[Pek94, Pek95].
Two common models for the Universal Teichmüller space are the space of
normalized quasicircles and the space of normalized quasisymmetries. The pro-
cess of conformal welding described in Section 3.2 provides the mechanism for
switching between the two models [Leh87, Krz95]. Our method of discrete con-
formal welding described above provides the means for actually computing this
correspondence as well [Wil01a, Wil04].
In the Radnell-Schippers model of quantum field theory, the ends of the world
sheets are parametrized (“rigged”) by quasisymmetric maps. The interaction
between two strings then corresponds to the welding of the two worldsheets via
the rigging [RS05]. These operations can be carried out using circle packings to
approximate the world sheets. The packable surfaces are dense [Bro86, Bro92,
Bro96, BS92, BS93, Wil03] in the moduli space of all surfaces, so nothing is lost
in this approach, while much is gained by the ability to actually compute the
new welded surface.

5.3. Circle Packing Measurable Riemann Mapping Theorem. Recall


that the distortion of a quasiconformal map f is described by its complex di-
latation µ, defined by the Beltrami equation
(5.1) ∂f = µ ∂f.
The classical Measurable Riemann Mapping Theorem asserts that given a Bel-
trami differential µ on a simply connected domain Ω ( C, there is a correspond-
ing quasiconformal map f µ from Ω to the unit disc D having µ as its complex
dilatation. If f µ is normalized to send two points p, q ∈ Ω, p 6= q, to 0 and
the positive real axis, respectively, then f µ is unique [LV73, Leh87, GL00].. The
original Riemann Mapping Theorem follows from the special case µ = 0.
Circle packings have been used previously by Zheng-Xu He [He90] to solve
Beltrami differential equations, but they appear indirectly. By applying our
discrete conformal welding technique, however, we can create quasiconformal
maps directly from their complex dilatation.
Circle Packings, QC Maps, and Applications 343

Given a Beltrami differential µ on a bounded simply connected region Ω ( C,


we pack Ω with a “ball bearing” packing. See Figure 6. The carrier divides Ω
into small squares. We approximate µ by a constant function on each square.
Recall from Section 2.3 that a map with constant dilatation is affine.
In work with Roger Barnard, we showed that the conformal structure on any
compact torus can be transformed into any other by cutting it open appropriately
and welding it back together [BW02]. However, the conformal structures of
compact tori can also be distorted by affine maps. Thus our work on welding
tori provides the mechanism for creating the effect of affine maps.
By refining our ball-bearing packing and performing a discrete conformal weld-
ing on each of the small squares in Ω, we can create a normalized discrete
quasiconformal map fn whose dilatation is approximately equal to µ on each
square [Wil].
Circle Packing Measureable Riemann Mapping Theorem. As the pack-
ings are refined, the discrete quasiconformal maps fn converge to the similarly
normalized quasiconformal map f µ : Ω → D with dilation µ.

References
[Aha97] Dov Aharonov, The sharp constant in the ring lemma, Complex Variables Theory
Appl. (1997), 27–31.
[Ahl63] Lars Ahlfors, Quasiconformal reflections, Acta Math. 109 (1963), 291–301.
[Ahl66] , Lectures on quasiconformal mappings, D. van Nostrand, Princeton, New
Jersey, 1966.
[And70a] E. M. Andreev, Convex polyhedra in Lobacevskii space, Mat. Sb. (N.S.) 10 (1970),
413–440 (English).
[And70b] , Convex polyhedra of finite volume in Lobacevskii space, Math. USSR Sbornik
12 (1970), 255–259 (English).
[BA56] A. Beurling and L. Ahlfors, The boundary correspondence under quasiconformal map-
pings, Acta Mathematica 96 (1956), 125–141.
[Bea99] Phillip L. Bowers and Monica K. Hurdal et al, Quasi-conformally flat mapping the
human cerebellum, Medical Image Computing and Computer-Assisted Intervention
- MICCAI ’99, Lecture Notes in Computer Science, vol. 1679, Springer-Verlag, 1999,
pp. 279–286.
[BK02] Mario Bonk and Bruce Kleiner, Quasisymmetric parametrizations of two-
dimensional metric spheres, Invent. Math. 150 (2002), no. 1, 127–183. MR
MR1930885 (2004k:53057)
[Bro86] Robert Brooks, Circle packings and co-compact extensions of Kleinian groups, In-
ventiones Mathematicae 86 (1986), 461–469.
[Bro92] , The continued fraction parameter in the deformation theory of classical
Schottky groups, Contemp. Math., vol. 136, Amer. Math. Soc., Providence, RI, 1992,
pp. 41–54.
[Bro96] , Some relations between graph theory and Riemann surfaces, Proceedings
of the Ashkelon Workshop on Complex Function Theory, Israel Math. Conf. Proc.,
vol. 11, Bar-Ilan Univ., Ramat Gan, 1996, pp. 61–73.
[BS90] Alan F. Beardon and Kenneth Stephenson, The uniformization theorem for circle
packings, Indiana Univ. Math. J. 39 (1990), 1383–1425.
[BS91] , Circle packings in different geometries, Tohoku Math. J. 43 (1991), 27–36.
344 G. Brock Williams IWQCMA05

[BS92] Philip L. Bowers and Kenneth Stephenson, The set of circle packing points in the
Teichmüller space of a surface of finite conformal type is dense, Math. Proc. Camb.
Phil. Soc. 111 (1992), 487–513.
[BS93] , Circle packings in surfaces of finite type: An in situ approach with applica-
tion to moduli, Topology 32 (1993), 157–183.
[BS97] , A regular pentagonal tiling of the plane, Conform. Geom. Dyn. 1 (1997),
58–68.
[BS04a] Alexander I. Bobenko and Boris A. Springborn, Variational principles for circle
patterns and Koebe’s theorem, Trans. Amer. Math. Soc. 356 (2004), no. 2, 659–689
(electronic). MR MR2022715
[BS04b] Philip L. Bowers and Kenneth Stephenson, Uniformizing dessins and Belyı̆ maps
via circle packing, Mem. Amer. Math. Soc. 170 (2004), no. 805, xii+97. MR
MR2053391 (2005a:30068)
[BW02] R.W. Barnard and G. Brock Williams, Combinatorial excursions in moduli space,
Pacific J. Math. 205 (2002), no. 1, 3–30.
[CS99] Charles Collins and Kennneth Stephenson, A circle packing algorithm, preprint.
[DE86] A. Douady and C. Earle, Conformally natural extension of homeomorphisms of the
circle, Acta Math. 157 (1986), 145–149.
[DS95a] Tomasz Dubejko and Kenneth Stephenson, The branched Schwarz lemma: a classical
result via circle packing, Mich. Math. J. 42 (1995), 211–234.
[DS95b] , Circle packing: Experiments in discrete analytic function theory, Experi-
ment. Math. 4 (1995), no. 4, 307–348.
[Dub97] Tomasz Dubejko, Recurrent random walks, Liouville’s theorem, and circle packings,
Math. Proc. Cambridge Philos. Soc. 121 (1997), no. 3, 531–546.
[dV89] Yves Colin de Verdière, Empilements de cercles: Convergence d’une methode de
point fixe, Forum Mathematicum 1 (1989), 395–402.
[dV91] , Une principe variationnel pour les empilements de cercles, Inventiones
Mathematicae 104 (1991), 655–669.
[DW05] David Dennis and G. Brock Williams, Layered circle packings, Int. Jour. Math. Math.
Sci. 15 (2005), 2429–2440.
[Geh82] Frederick W. Gehring, Characteristic properties of quasidisks, Les Presses De
L’Universiteé De Montréal, Montréal, 1982.
[GL00] Frederick P. Gardiner and Nikola Lakic, Quasiconformal Teichmüller theory, Math-
ematical Surveys and Monographs, vol. 76, American Mathematical Society, 2000.
[He90] Zheng-Xu He, Solving Beltrami equations by circle packing, Trans. Amer. Math. Soc.
322 (1990), 657–670.
[HR93] Zheng-Xu He and Burt Rodin, Convergence of circle packings of finite valence to
Riemann mappings, Comm. in Analysis and Geometry 1 (1993), 31–41.
[HS93] Zheng-Xu He and Oded Schramm, Fixed points, Koebe uniformization and circle
packings, Ann. of Math. 137 (1993), 369–406.
[HS95] , Hyperbolic and parabolic packings, Discrete & Computational Geom. 14
(1995), 123–149.
[HS96] , On the convergence of circle packings to the Riemann map, Invent. Math.
125 (1996), 285–305.
[HS98] , The C ∞ -convergence of hexagonal disk packings to the Riemann map, Acta
Mathematica 180 (1998), 219–245.
[IT92] Yoichi Imayoshi and Masahiko Taniguchi, An introduction to Teichmüller spaces,
Springer-Verlag, 1992.
[Koe36] Koebe, Kontaktprobleme der Konformen Abbildung, Ber. Sächs. Akad. Wiss. Leipzig,
Math.-Phys. Kl. 88 (1936), 141–164.
[Kra99] Steven G. Krantz, Conformal mappings, American Scientist 87 (1999), 436–445.
Circle Packings, QC Maps, and Applications 345

[Krz87] Jan G. Krzyż, Quasicircles and harmonic measure, Annales AcademiæScientiarum


FennicæSeries AI Mathematica 12 (1987), 19 – 24.
[Krz95] , On the notion of the universal Teichmüller space, Rev. Roumaine Math.
Pures Appl. 40 (1995), no. 2, 169–175.
[Leh87] O. Lehto, Univalent functions and Teichmüller spaces, Springer-Verlag, Berlin - Hei-
delberg - New York, 1987.
[LP88] A. Lecko and D. Partyka, An alternative proof of a result due to Douady and Earle,
Annales Universitatis Mariae Curie - Sklodowska - Lublin - Polonia 17 (1988), 59 –
68.
[LV73] O. Lehto and K.I. Virtanen, Quasiconformal mappings in the plane, second ed.,
Springer - Verlag, Berlin - Heidelberg - New York, 1973.
[McC98] Gareth McCaughan, A recurrence/transience result for circle packings, Proc. Amer.
Math. Soc. 126 (1998), 3647–3656.
[Moh93] Bojan Mohar, A polynomial time circle packing algorithm, Discrete Math. 117
(1993), 257–263.
[MS04] David Mumford and Eitan Sharon, 2D-Shape analysis using conformal mapping,
Proceedings IEEE Conference on Computer Vision and Pattern Recognition, June
2004.
[Nag88] Subhashis Nag, The complex analytic theory of Teichmüller spaces, Wiley, 1988.
[Pek94] Osmo Pekonen, The interface of Teichmüller theory and string theory, Rev.
Roumaine Math. Pures Appl. 38 (1994), no. 8, 829–854.
[Pek95] , Universal Teichmüller space in geometry and physics, Journal of Geometry
and Physics 15 (1995), 227–251.
[Pfl51] Albert Pfluger, Quasiconforme abbildungen und logarithmische kapazität, Ann. Inst.
Fourier (Grenoble) (1951), no. 2, 69–80.
[Rep98] Andrew Repp, Discrete Riemann maps and the parabolicity of tilings, Ph.D. thesis,
Virginia Polytechnic Institute and State University, May 1998.
[RS87] Burt Rodin and Dennis Sullivan, The convergence of circle packings to the Riemann
mapping, J. Differential Geometry 26 (1987), 349–360.
[RS05] David Radnell and Eric Schippers, Quasisymmetric sewing in rigged Teichmüller
space, preprint (2005).
[Ste96] Kenneth Stephenson, A probabilistic proof of Thurston’s conjecture on circle pack-
ings, Rend. Sem. Mat. Fis. Milano 66 (1996), 201–291.
[Ste97] , The approximation of conformal structures via circle packing, Computa-
tional methods and function theory 1997 (Nicosia) (N. Papamichael, S. Ruscheweyh,
and E.B. Saff, eds.), Ser. Approx. Decompos., World Scientific, 1997, pp. 551–582.
[Ste02] , Circle packing and discrete analytic function theory, Handbook of com-
plex analysis: geometric function theory, Vol. 1, North-Holland, Amsterdam, 2002,
pp. 333–370.
[Ste03] Kenneth Stephenson, Circle packing: a mathematical tale, Notices Amer. Math. Soc.
50 (2003), no. 11, 1376–1388.
[Ste05] Kenneth Stephenson, Introduction to circle packing: The theory of discrete analytic
functions, Cambridge University Press, 2005.
[Sul01] Dennis Sullivan, Strings, graphs, Riemann surfaces, Invited Address at the Graphs
and Patterns in Mathematics and Theoretical Physics Conference, June 2001.
[Thu] William Thurston, The geometry and topology of 3-manifolds, Princeton University
Notes, preprint.
[Thu85] , The finite Riemann mapping theorem, 1985, Invited talk, An International
Symposium at Purdue University on the occasion of the proof of the Bieberbach
conjecture, March 1985.
[Wil] G. Brock Williams, A circle packing measureable Riemann mapping theorem, to
appear in Proc. AMS.
346 G. Brock Williams IWQCMA05

[Wil01a] , Approximating quasisymmetries using circle packings, Discrete and Comput.


Geom. 25 (2001), no. 1, 103–124.
[Wil01b] , Earthquakes and circle packings, J. Anal. Math. (2001), no. 85, 371–396.
[Wil03] , Noncompact surfaces are packable, J. Anal. Math. 90 (2003).
[Wil04] , Discrete conformal welding, Indiana Univ. Math. J. 53 (2004), no. 3, 765–
804.

G. Brock Williams Address: Department of Mathematics, Texas Tech University,


Lubbock, Texas 79409
E-mail: williams@math.ttu.edu
URL: http://www.math.ttu.edu/∼williams
List of registered participants for the workshop
(excluding unregistered research scholars, Department of
Mathematics, IIT Madras)

Australia: Raimo Näkki


T-W. Ma Dept. of Mathematics and Statistics
School of Mathematics and Statistics P.O. Box 35 (MaD)
The University of Western Australia FIN-40014
Nedlands, W.A., 6009 University of Jyväskylä, Finland
Australia raimon@maths.jyu.fi
twma@maths.uwa.edu.au
Istvan Prause
Dept. of Mathematics and Statistics
P.O. Box 68, FI-00014
Finland: University of Helsinki, Finland
Peter A. Hästö istvan.prause@helsinki.fi
Department of Mathematical Sciences
P.O. Box 3000, FI-90014 Antti Rasila
University of Oulu, Finland Helsinki University of Technology
peter.hasto@helsinki.fi Institute of Mathematics
P.O. Box 1100, FIN-02015 HUT, Fin-
Ilkka Holopainen land
Dept. of Mathematics and Statistics arasila@iki.fi
P.O. Box 68, FI-00014
University of Helsinki, Finland Matti Vuorinen
iholopai@cc.helsinki.fi Department of Mathematics
University of Turku
Riku Klén FI-20014 Turku, Finland
Department of Mathematics vuorinen@utu.fi
University of Turku
FI-20014 Turku, Finland
riku.klen@utu.fi
India:
Henri Lindén Javid Ali
Dept. of Mathematics and Statistics Department of Mathematics
P.O. Box 68, FI-00014 Aligarh Muslim University
University of Helsinki, Finland Aligarh–202 002, India
hlinden@cc.helsinki.fi java81@rediffmail.com

Olli Martio Meena S. Atak


Dept. of Mathematics and Statistics Maharashtra Academy of Engineering
P.O. Box 68, FI-00014 Alandi–Pune
University of Helsinki, Finland Pin–412 105
olli.martio@helsinki.fi meena atak@yahoo.co.in
350 List of Participants IWQCMA05

A. Avudainayagam Sukhjit Singh Dhaliwal


Department of Mathematics Department of Mathematics
IIT Madras Sant Longowal Inst. of Engg. & Tech.
Chennai - 600 036 Longowal-148106 (Punjab)
avudai@iitm.ac.in sukhjit d@yahoo.com

R Balasubramanian Geetha
Department of Mathematics Department of Mathematics
The Institute of Mathematical Sciences Dheivanai Ammal College for Women
C.I.T. Campus, Taramani Villupuram – 605 602
Chennai – 600 113 Tamil Nadu
balu@imsc.res.in
K.R. Karthikeyan
S.S. Bhoosnurmath 24, Dhanalakshmi Ammal Street
Department of Mathematics Reddiar Garden, Kamaraj Nagar
Karnatak University Avadi, Chennai–600 071
Pavate Nagar, Dharwad – 580 003 kr karthikeyan1979@yahoo.com
Karnataka State
S. Karthikeyan
Bappaditya Bhowmik Department of Mathematics
Department of Mathematics Sona College of Technology
IIT Madras Sona Nagar
Chennai - 600 036 Thiagarajar Polytechnic College Road
ditya@iitm.ac.in Salem – 636 005, Tamil Nadu
mrsiva75@yahoo.co.in
Bhuvana
Hindustan Engineering College S.M. Khairnar
Padur, Kanchipuran Dist. Department of Mathematics
mabhuv@yahoo.com Anuradha Engineering College,
Chikhli–443201, Dist. Buldana, (M.S)
K. Chandrasekhran smkhairnar123@rediffmail.com
R-3 Sivakami Apartments
Somasundharam Street Kokila
Muthuvel Nagar, East Tambaram Department of Mathematics
Chennai-59 Dheivanai Ammal College for Women
kchandru1978@hotmail.com Villupuram – 605 602
Tamil Nadu
S.A. Choudum
Department of Mathematics V. Lakshmi
IIT Madras 52, Choolai high road
Chennai - 600 036 choolai, chennai – 600 112
sac@iitm.ac.in Tamil Nadu
lakshmipandian2005@yahoo.co.in
List of Participants 351

J. Lourthu Mary Tarakanta Nayak


Department of Mathematics Department of Mathematics
Madras Institute Technology IIT Guwahati, North Guwahati
Anna University Guwahati – 781039
Chennai – 600 044 tarakanta@iitg.ernet.in

N. Marikkannan Sanjay kumar Pant


Department of Applied Mathematics Deen Dayal Upadhyaya college(DU)
Sri Venkateswara College of Engg. Shivaji Marg, Karampura
Pennalur, Sripeumbudur – 602 105 New Delhi – 110 015
mari@svce.ac.in skpant@bol.net.in

A.K. Misra R. Parvatham


Department of Mathematics Ramanujan Institute (RIASM)
Berhampur University University of Madras
Bhanja Bihar Chennai – 600 005
Berhampur, Orissa parvatham@hotmail.com
akshayam2001@yahoo.co.in

Debasisha Mishra J. Patel


Department of Mathematics Department of Mathematics
Utkal University, Vani Vihar Utkal University, Vani Vihar
Bhubaneswar – 751004 Bhubaneswar–751004
kapa math@yahoo.com jpatelmath@sify.com

Sumit Mohanty Santosh Kumar Pattanayak


C/O Jagannath Patel Department of Mathematics
Department of Mathematics Chennai Mathematical Institute
Utkal University, Vani Vihar Plot H1, SIPCOT IT Park
Bhubaneswar – 751004 Padur PO, Siruseri – 603 103
sumit math2004@yahoo.co.in santosh@cmi.ac.in

S. Nanda S. Ponnusamy
North Orissa University Department of Mathematics
Sriram Chandra Vihar, Takatpur IIT Madras
Mayurbhanj Chennai - 600 036
Baripada – 757003 samy@iitm.ac.in
snanda@maths.iitkgp.ernet.in

U.H. Naik D.J. Prabhakaran


Department of Mathematics Dept. of Appl. Sci. & Humanities
Willingdon College M.I.T. Campus, Anna University
SANGLI – 416415 Chennai – 600 044
naikpawan@yahoo.com asirprabha@yahoo.com
352 List of Participants IWQCMA05

Z. Rebekal S. Selvaganesh
Dr. M.G.R. Deemed University 1/D Jawaharlal Street, Madupet
No. 122, Saniyasipuram Gudiyattam – 632602
2nd street, Medawalkam Vellore District, Tamil Nadu
Kilpauk, Chennai – 600 010 selva gym2000@yahoo.co.in

Rathiga C. S. Seshadri
Department of Mathematics Chennai Mathematical Institute
Aarupadai Veedu Inst. of Tech. Plot H1, SIPCOT IT Park
Paiyanoor Padur PO, Siruseri – 603 103
mabhuv@yahoo.com css@cmi.ac.in

Satyajit Roy R. Sivakumar


Department of Mathematics Department of Mathematics
IIT Madras Sona College of Technology
Chennai - 600 036 Sona Nagar
sjroy@iitm.ac.in Thiagarajar Polytechnic College Road
Salem – 636 005, Tamil Nadu
Pravati Sahoo mrsiva75@yahoo.co.in
Department of Mathematics
Mahila Maha Vidyalaya (MMV) S. Sivaprasad Kumar
Banaras Hindu University Department of Applied Mathematics
Varanasi – 221005 Delhi College of Engineering
pravatis@yahoo.co.in Bawana Road, New Delhi – 110 042
sivpk71@yahoo.com
Swadesh Kumar Sahoo
Department of Mathematics C.M. Subalakshmi
IIT Madras Dr. M.G.R. Deemed University
Chennai - 600 036 No. 23, Bashyam Reddy 2nd street
swadesh iitm@math.net Ooteri, Chennai – 600 012

N.D. Sangle K. Subathra


Annasaheb Dange College Department of Mathematics
of Engg. & Tech. Rajalakshmi Engg.College
Ashta, Dist. Sangli – 416 301 Rajalakshmi Nagar, Thaudalam
Maharastra k subathra@yahoo.co.in
navneet sangle@rediffmail.com

M. Suganthi
R. Sathya Priya Dept. of Math. and Computer Appl.
R.M.K. Engineering College PSG College of Technology
Kavarai Pattai Coimbatore – 641004
Thiruvalluvar – 601 206 msugan psgtech@yahoo.co.in
List of Participants 353

S. Sundar P. Veeramani
Department of Mathematics Department of Mathematics
IIT Madras IIT Madras
Chennai - 600 036 Chennai - 600 036
slnt@iitm.ac.in pvmani@iitm.ac.in

Jagmohan Tanti Murali K. Vemuri


Department of Mathematics Department of Mathematics
Chennai Mathematical Institute Chennai Mathematical Institute
Plot H1, SIPCOT IT Park Plot H1, SIPCOT IT Park
Padur PO, Siruseri – 603 103 Padur PO, Siruseri – 603 103
jagmohan@cmi.ac.in mkvemuri@cmi.ac.in

G. Thirupathi V. Vetrivel
Adhiyamaan College of Engineering Department of Mathematics
Hosur–635109,Tamil Nadu IIT Madras
gt venkat@rediffmail.com Chennai - 600 036
vetri@iitm.ac.in
E. Umamaheswari
Dr. M.G.D. Educational & Research
Inst. Iran:
No. 23, Venkatachala Mudali street R. Aghalary
Vepery – 600 007 Department of Mathematics
University of Urmia
Vanitha Urmia, Iran
Department of Mathematics raghalary@yahoo.com
Dheivanai Ammal College for Women
Villupuram – 605 602 Saeid Shams
Tamil Nadu Department of Mathematics
University of Urmia
Allu Vasudevarao Urmia, Iran
Department of Mathematics sa40shams@yahoo.com
IIT Madras
Chennai - 600 036
alluvasu@iitm.ac.in

P. Vasundhra
Department of Mathematics
IIT Madras
Chennai - 600 036
vasu2kk@yahoo.com
354 List of Participants IWQCMA05

Japan:
Toshiyuki Sugawa C.M. Pokhrel
Department of Mathematics Department of Mathematics
Graduate School of Science Nepal Engineering College
Hiroshima University G.P.O. Box 10210
1-3-1 Kagamiyama, Higashi-Hiroshima Kathmandu, Nepal
739-8526 JAPAN cmpokhrel@wlink.com.np
sugawa@math.sci.hiroshima-u.ac.jp

Hiroshi Yanagihara USA:


Department of Applied Science Om P. Ahuja
Yamaguchi University Department of Mathematics
Faculty of Engineering Kent State University
Tokiwadai Ube 755 – 8611 Ohio, U.S.A.
Japan oahuja@kent.edu
hiroshi@yamaguchi-u.ac.jp
Roger W. Barnard
Dept. of Mathematics and Statistics
Korea: Texas Tech University
Jae Ho Choi Lubbock, TX 79409, U.S.A
Department of Mathematics Education barnard@math.ttu.edu
Daegu National University of Educa-
tion Phillip Brown
1797-6 Daemyong 2 dong, Namgu Texas A&M University Galveston
Daegu 705-715, Korea PO Box 1675, Galveston
choijh@dnue.ac.kr Texas 77553 -1675, U.S.A
brownp@tamug.edu

Mexico: J. Lee Bumpus


R. Michael Porter K. Dept. of Mathematics and Statistics
Departamento de Matematicas Texas Tech University
CINVESTAV-I.P.N. Lubbock, TX 79409, U.S.A
Apdo. Postal 14-740 jlee wb@yahoo.com
07000 Mexico, D.F. MEXICO
mike@math.cinvestav.mx Atul Dixit
Dept. of Mathematics and Statistics
Texas Tech University
Nepal: Lubbock, TX 79409, U.S.A
Ajaya Singh atul.dixit@ttu.edu
Central Department of Mathematics
Tribhuvan University D. Freeman
Kirtipur, Kathmandu Department of Mathematical Sciences
Nepal University of Cincinnati
ajayas 2000@yahoo.com Cincinnati, OH 45221-0025, U.S.A
freemadd@email.uc.edu
List of Participants 355

David A. Herron Len Ruth


Graduate Program Director Department of Mathematical Sciences
Department of Mathematical Sciences University of Cincinnati
University of Cincinnati Cincinnati, OH 45221-0025, U.S.A
Cincinnati, Ohio 45221-0025, U.S.A len.ruth@sinclair.edu
david.a.herron@gmail.com
Alex Williams
Casey Hume Dept of Mathematics and Statistics
Dept. of Mathematics and Statistics Texas Tech University
Texas Tech University Lubbock, TX 79409
Lubbock, TX 79409, U.S.A alejandros son@yahoo.com
casey.hume@ttu.edu
Brock Williams
William Ma Dept. of Mathematics and Statistics
Department of Mathematics Texas Tech University
Pennsylvania College of Technology Lubbock, Texas 79409, U.S.A
22 Hillview Avenue, Williamsport, PA williams@math.ttu.edu
17701
U.S.A
wma@pct.edu

S.S. Miller
Department of Mathematics
State University of New York
Brockport, NY 14420, U.S.A
smiller@brockport.edu

David Minda
Department of Mathematical Sciences
University of Cincinnati
Cincinnati, OH 45221-0025, U.S.A
E-mail: David.Minda@math.uc.edu

Eric Murphy
U.S. Air Force
6611 Comet Circle
Apt no. 302
Springfield, VA 22150, U.S.A
eric.murphy@cox.net

Kent Pearce
Dept. of Mathematics and Statistics
Texas Tech University
Lubbock, TX 79409, U.S.A
kent.pearce@ttu.edu
356 List of Participants IWQCMA05

List of unregistered participants


(Research scholars, Department of Mathematics, IIT Madras)
• A. Anthony Eldred
• J. Anuradha
• A. Chandrashekaran
• R. Indhumathi
• Nachiketa Mishra
• V. Murugan
• Param Jeet
• H. Ramesh
• Jajati Keshori Sahoo
• V. Sankaraj
• E. Satyanarayana
• S. Suresh Kumar
• Manoj Yadav

Potrebbero piacerti anche