Krantz Visual Complex Analysis
Krantz Visual Complex Analysis
Number Twenty-Three
Complex Analysis:
The Geometric Viewpoint
Second Edition
Steven G. Krantz
Washington University in St. Louis
Published by
THE MATHEMATICAL ASSOCIATION OF AMERICA
Committee on Publications
Gerald L. Alexanderson, Chair
Editorial Board
Kenneth A. Ross, Chair
Joseph Auslander
Harold P. Boas
Robert E. Greene
Roger Horn
Jeffrey Lagarias
Barbara Osofsky
The following Monographs have been published:
This book owes its existence to many people. I thank Don Albers,
Robert E. Greene, and Paul Halmos for convincing me to write it. I am
grateful to Marco Abate, Harold Boas, Ralph Boas, David Drasin, Paul
Halmos, Daowei Ma, David Minda, Marco Peloso, John Stapel, and
Jim Walker for reading various versions of the manuscript and making
valuable comments and suggestions. The Carus monograph committee
of the MAA helped me to find the right level and focus for the book.
Ralph Boas, the chairman of the Carus monograph committee,
played a special role in the development of this book. In addition to
shepherding the project along, he provided necessary prodding and ca-
joling at crucial stages to keep the project on track. Paul Halmos also
provided expert and sure counsel; he has been a good friend and mentor
for many years. To both men I express my sincere gratitude.
For the new edition, Ken Ross as Chair of the Carus Monographs
Editorial Board provided a strong and sure guiding hand. My friends
Harold Boas, Daowei Ma, David Minda, Jeff McNeal, Marco Peloso,
and Jim Walker offered valuable suggestions for the form of the new
edition. Robert Burckel, Robert E. Greene, and John P. D’Angelo were
especially generous in offering detailed commentary which led to deci-
sive improvements. Finally, the Carus Monograph Committee provided
yeoman service with many careful readings, sharp but constructive crit-
icisms, and helpful suggestions. Surely a better book has been the re-
sult.
ix
x Acknowledgments
—S.G.K.
Preface to the Second Edition
The warm reception with which the first edition of this book has been
received has been a source of both pride and pleasure. It is a special
privilege to have created a “for the record” version of Ahlfors’s seminal
ideas in the subject. And the geometric viewpoint continues to develop.
In the intervening decade, this author has learned a great deal more
about geometric analysis, and his view of the subject has developed and
broadened. It seems appropriate, therefore, to bring some new life to
these pages, and to set forth a fresh enunciation of the role of curvature
in basic complex function theory.
In this new edition, we explain how, in a natural and elementary
manner, the hyperbolic disc is a model for the non-Euclidean geometry
of Bolyai and Lobachevsky. Later on, we explain the Bergman kernel
and provide an introduction to the Bergman metric.
I have many friends and colleagues to thank for their incisive re-
marks and suggestions about the first edition of this book. I hope that I
do them justice in my efforts to implement a second edition. As always,
the Mathematical Association of America has been an exemplary pub-
lisher and has provided all possible support in the publication process.
I offer my humble thanks.
xi
Preface to the First Edition
Acknowledgments ix
Preface to the Second Edition xi
Preface to the First Edition xiii
xv
xvi Contents
Appendix 191
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
2. Expressing Curvature Intrinsically . . . . . . . . . . . . . . . . . . . . 191
3. Curvature Calculations on Planar Domains . . . . . . . . . . . . . 200
Symbols 205
References 209
Index 213
CHAPTER 0
Principal Ideas of Classical
Function Theory
D(P, r ) = {z ∈ C : |z − P| < r } ,
D(P, r ) = {z ∈ C : |z − P| ≤ r } ,
∂ D(P, r ) = {z ∈ C : |z − P| = r } .
We often use the lone symbol D to denote the unit disc D(0, 1). A
connected open set U ⊆ C is called a domain.
Complex analysis consists of the study of holomorphic functions.
Let F be a complex-valued continuously differentiable function (in the
sense of multivariable calculus) on a domain U in the complex plane.
We write F = u + iv to distinguish the real and imaginary parts of F.
Then F is said to be holomorphic, or analytic, if it satisfies the Cauchy–
Riemann equations:
∂u ∂v ∂u ∂v
= and =− .
∂x ∂y ∂y ∂x
that point it is routine to check that the function satisfies the Cauchy–
Riemann equations. The converse implication is a straightforward ex-
ercise. So, in the end, either definition is correct. From our perspective
the Cauchy–Riemann equations provide the most useful point of view.
This assertion will become more transparent as we develop the notion
of complex integration.
Figure 1.
Figure 2.
Figure 3.
D(P, r ) ≡ {z : |z − P| ≤ r }
1 1 1
= · .
ζ −z ζ − P 1 − ζz−P
−P
Thus
∞
1 1 z−P j
= · .
ζ −z ζ − P j=0 ζ − P
Substituting this power series expansion for the Cauchy kernel into the
Cauchy integral formula on D(P, r ) gives the desired power series ex-
pansion for the holomorphic function F.
It is a standard fact from the theory of power series that the zeros
of a function given by a power series expansion cannot accumulate in
the interior of the domain of that function. Thus we have:
This theorem once again bears out the dictum that holomorphic func-
tions are much like polynomials: The zero set of a polynomial a0 +
a1 z + a2 z 2 + · · · + an z n is discrete, indeed it is finite.
The Cauchy estimates on the derivatives of a holomorphic func-
tion in terms of the supremum of the function follow from direct esti-
mation of the formula (∗):
p(z) = (z − r1 ) · (z − r2 ) · · · (z − rk ).
10 Principal Ideas of Classical Function Theory
F (ζ ) 1
= + h(ζ ),
F(ζ ) ζ
where h is holomorphic near 0. Of course, h integrates to 0, by the
Cauchy integral theorem. And it is easily calculated that
1 1
dζ = 1,
2πi γ ζ
|F(P)| ≥ |F(z)|
For us, the main application of the maximum principle will be to the
classical Schwarz lemma.
ζ −a
φa (ζ ) =
1 − aζ
maps D(0, 1) to D(0, 1). [We shall use the phrase “Möbius transfor-
mation” specifically to mean a mapping of this form. Later on, we shall
consider compositions of rotations with Möbius transformations.] To
see this, notice that φa is certainly well defined and holomorphic in a
neighborhood of D(0, 1); moreover, φa (a) = 0. If we prove that φa
maps ∂ D(0, 1) to ∂ D(0, 1), then our claim will follow from the maxi-
mum principle. But for |ζ | = 1 we have that
ζ − a 1 ζ − a
|φa (ζ )| = = ·
1 − aζ ζ̄ 1 − aζ
ζ − a
= = 1,
ζ̄ − a
F(ζ ) = φa ◦ ρτ (ζ ).
F = φ−b ◦ ρτ .
and
1 − |w1 |2
|F (z 1 )| ≤ .
1 − |z 1 |2
If equality obtains in the first expression for some z 1 = z 2 or if equality
obtains in the second expression, then F must be a conformal self-map
of the disc.
Proof. Define
z + z1 z − w1
φ(z) = , ψ(z) = .
1 + z1 z 1 − w1 z
Then ψ ◦ F ◦ φ satisfies the hypotheses of Schwarz’s lemma. Therefore
(ψ ◦ F ◦ φ)(z) ≤ |z|, ∀z ∈ D.
Using the chain rule to write this out gives the second inequality.
The case of equality is analyzed as in Theorem 2.
16 Principal Ideas of Classical Function Theory
|F(z)| ≤ M
For the proof of this theorem, we shall need an important result from
real analysis. We begin with a little terminology.
for all F ∈ F.
Notice that the property of equicontinuity is stronger than uniform
continuity: not only is the choice of δ independent of the points z and w
in S (i.e., it depends only on ε), but it is independent of which F from
the family F we are considering. Next we define a companion notion.
|F(z)| ≤ M.
a positive distance between the two sets: that is, there exists a k > 0
such that if z ∈ D(P, R) and u ∈ c U , then |z − u| > k. Thus, for any
z ∈ D(P, R) and any F ∈ F we may apply the Cauchy estimates to F
on the disc D(z, k). We conclude that
M · 1!
|F (z)| ≤ ≡ C.
k1
It is now an easy exercise, using the fact that
z
F(z) − F(w) = F (ξ ) dξ,
w
to see that
F(z) − F(w) ≤ C|z − w|
for any z, w ∈ D(P, R). This estimate shows that the family F is
equicontinuous on D(P, R): given ε > 0, choose δ = ε/C.
Finally, if K is any compact subset of U , then K can be covered
by finitely many discs D(P, R) as above, and the equicontinuity on K
follows from the triangle inequality.
We may now apply the Ascoli/Arzelà theorem to F on K to
find a uniformly convergent subsequence {F jk } of any given sequence
{F j } ⊆ F. A standard diagonalization argument then yields a subse-
quence {F jkl } which converges uniformly on all compact subsets of U .
σ : U → V.
Normal Families and the Riemann Mapping Theorem 19
1
σ : ζ → (ζ − P)
2R
certainly maps P to 0, is holomorphic, and satisfies
σ (ζ ) < 1 (R + R) = 1.
2R
Obviously σ is one-to-one. So σ ∈ F and F is nonempty.
S TEP 2. The family F is a normal family. In fact, the elements of F
are holomorphic, and they are all bounded by 1. Montel’s theorem
thus applies, and the family is normal.
20 Principal Ideas of Classical Function Theory
S TEP 3. Let us define M = sup |σ (P) : σ ∈ F . We claim that M
is finite. In fact, let D(P, r ) be a closed disc about P that is con-
tained in U . Then the Cauchy estimates apply: for any σ ∈ F, we
have that
σ (P) ≤ 1! = 1 .
r r
S TEP 6. We claim that σ0 maps U onto the unit disc. It is in this step
that the topology of U comes in. One proves, by an advanced cal-
culus argument, that if F is any nonvanishing holomorphic function
on U , then log F is a well defined holomorphic function on U . In-
deed, one can define this logarithm by the formula
F (ζ )
log F(z) = dζ,
γz F(ζ )
1 + |β|
|ν (P)| = · M > M.
2|β|1/2
22 Principal Ideas of Classical Function Theory
It is plain that (i), (ii), (iii) are mutually exclusive and cover all possi-
bilities.
Isolated Singularities and the Theorems of Picard 23
is holomorphic on D(P, r ).
In the second case, we see that
Define
1
h(z) = .
F(z) − β
Then |h(z)| is bounded near P by the constant 1/δ. By the Riemann re-
movable singularities theorem, h can be continued analytically through
the point P. If h(P) = 0, then we have a contradiction since then F
could be continued analytically through P. But h(P) = 0 would imply
that F has a pole at P, and that is also false. The resulting contradiction
completes the proof.
It is sometimes useful to think about the point at ∞ just like any point
of the finite complex plane. The map
1
z →
z
allows us to pass back and forth between neighborhoods of 0 and
neighborhoods of ∞. In particular, a function F holomorphic on a set
{z ∈ C : |z| > R} is said to have a removable singularity, a pole, or an
essential singularity at ∞ if the function F (1/z) has, respectively, a
removable singularity, a pole, or an essential singularity at the origin.
We now record some striking properties about singularities at infinity.
Let Pk be the kth degree Taylor polynomial of F about zero. Then the
function
F(z) − Pk (z)
h(z) =
zk
is holomorphic on the entire plane and bounded at ∞. Thus, by Liou-
ville, h is constant. It follows that F is a polynomial.
F(z) = e z , then F takes all complex values except 0. The little the-
orem allows non-constant entire functions to omit one complex value.
It does not allow omission of two.
Both of Picard’s theorems are quite difficult to prove by classi-
cal techniques. The original proof involved the construction of a spe-
cial holomorphic function from the upper half plane to C \ {0, 1}. This
function, known as the elliptic modular function, is a powerful tool in
function theory. Here we briefly discuss the construction of the modular
function and its application to the proof of the little theorem.
Begin with a conformal mapping h of the right half of the re-
gion , shown in Figure 1, to the upper half plane, normalized so that
h(0) = 0, h(1) = 1, h(∞) = ∞. By Schwarz reflection [GRK], the
map h extends to a map from the entire region to C \ {0, 1}. By
more delicate reflection arguments, it is possible to continue the func-
tion analytically to a map from the entire upper half plane to C \ {0, 1}.
Denote the extended function by the symbol λ(z). Then λ is the elliptic
modular function.
Figure 1.
Isolated Singularities and the Theorems of Picard 27
0. Introductory Remarks
Differential geometry has developed into one of the most powerful
mathematical tools of modern mathematics. It has become an integral
part of the theories of differential equations, harmonic analysis, and
complex analysis, to name just a few examples.
In spite of their importance, the techniques of geometry have not
proliferated as much as they might have because of the complexity of
the language. The characterization of differential geometry as “that
portion of mathematics which is invariant under change of notation”
is unfortunately rather accurate.
The best way to learn new mathematics is in the context of what
is already familiar. In the milieu of one complex variable, the notions
of Riemannian metric, of geodesic, and of curvature become rather
simple. The standard geometric device of tensors and bundles is not
necessary. The result is that one can learn the flavor and some of the
methodology of differential geometry without being encumbered by its
notation and machinery.
In this chapter, therefore, we shall learn some of the basic ideas
of differential geometry, but only in the complex plane; we shall also
learn a few simple applications. In later chapters we shall learn about
curvature and more advanced applications.
29
30 Basic Notions of Differential Geometry
λ : X × X −→ R
1. λ(x, y) = λ(y, x)
2. λ(x, y) ≥ 0 and λ(x, y) = 0 iff x = y;
3. λ(x, y) ≤ λ(x, z) + γ (z, y).
The trouble with a metric defined in this generality is that it does not
interact well with calculus. What sort of interaction might we wish to
see?
Given two points P, Q ∈ X , one would like to consider the curve
of least length connecting P to Q. Any reasonable construction of such
a curve leads to a differential equation, and thus we require that our
metric lend itself to differentiation. Yet another consideration is curva-
ture: in the classical setting curvature is measured by the rate of change
of the normal vector field. The concepts of normal and rate of change
lead inexorably to differentiation. Thus we will now take a different
approach to the concept of “metric.”
Recall that in multivariable calculus we define the arc length of a
continuously differentiable curve γ : [a, b] → R2 by the formula
b
(γ ) = |γ̇ (t)| dt. (∗)
a
In point of fact, most calculus books give a heuristic which leads from
a reasonable notion of length to this formula. But, rigorously speaking,
(∗) must be considered to be the definition of arc length.
Notice that the definition (∗) hinges on the assumption that we
have previously defined |γ̇ (t)|, the length of the tangent vector γ̇ (t) to
Riemannian Metrics and the Concept of Length 31
ξ ρ,z ≡ ρ(z) · |ξ |,
Remark. Later in the book, we will consider metrics which are only
continuous (or even less smooth). We want a smooth metric now be-
cause we will differentiate it to calculate curvature.
For the record, the metrics we are considering here are a special
case of the type of differential metric called Hermitian. This terminol-
ogy need not concern us here. Classical analysts sometimes call these
metrics conformal metrics and write them in the form ρ(z)|dz|.
Technically speaking, our metric lives on the tangent bundle to the
domain . That is to say, the metric is a function of the variable (z, v),
where v is thought of as a tangent vector at the point z. This is just a
mathematical way of saying that the length of the vector v depends on
the point z at which it is positioned.
32 Basic Notions of Differential Geometry
ξ ρ,z ≡ ρ(z) · |ξ | = |ξ |.
1
ρ(z) =
1 − |z|2
This is the Poincaré metric, which has been used to gain deep insights
into complex analysis on the disc. It will receive our detailed attention
later on in this book. For now we do some elementary calculations with
the Poincaré metric.
For any ξ ∈ C we may calculate that
4
ξ ρ,(1/2+0i) = ρ(1/2 + 0i) · |ξ | = · |ξ |;
3
ξ ρ,0 = ρ(0) · |ξ | = 1 · |ξ | = |ξ |;
1
ξ ρ,(0+.9i) = ρ(.9i) · |ξ | = |ξ | = (5.2631578 . . . ) · |ξ |.
.19
The notion of the length of a vector varying with the base point is in
contradistinction to what we learn in calculus. In calculus, a vector has
direction and magnitude but not position. Now we declare that a vector
Riemannian Metrics and the Concept of Length 33
has position and the way that its magnitude is calculated depends on
that position. The next example shows how vectors with position arise
in practice.
1
η̇(t) ρ,η(t) = ρ(η(t)) · |η̇(t)| = .
1 − t2
We also have |µ̇(t)| = 1 for all t and
1
µ̇(t) ρ,µ(t) = ρ(µ(t)) · |µ̇(t)| = .
3/4 − t 2
(t)
(t)
Figure 1.
34 Basic Notions of Differential Geometry
γ : [a, b] →
Example 4. Let D ⊆ C be the unit disc and let ρ(z) = 1/(1 − |z|2 ),
the Poincaré metric on D. Fix > 0. Let us calculate the length of the
curve γ (t) = t, 0 ≤ t ≤ 1 − . Now
Riemannian Metrics and the Concept of Length 35
1−
ρ (γ ) = γ̇ (t) ρ,γ (t) dt
0
1− |γ̇ (t)|
= dt
0 1 − |γ (t)|2
1−
1
= dt
0 1 − t2
1 2−
= log .
2
lim ρ (γ ) = +∞.
→0+
This suggests that the boundary ∂ D is infinitely far from the origin, at
least along this particular path γ , in this the Poincaré metric.
Example 5. Equip the disc D with the Poincaré metric ρ(z) = 1/(1 −
|z|2 ). Fix > 0. Let us prove that, among all continuously differen-
tiable curves of the form
µ(t) = t + iw(t), 0 ≤ t ≤ 1 − ,
36 Basic Notions of Differential Geometry
that satisfy µ(0) = 0 and µ(1 − ) = 1 − + 0i, the one of least length
is γ (t) = t. Here w(t) is a continuously differentiable, real-valued
function.
In fact for any such candidate µ we have
1−
ρ (µ) = µ̇(t) ρ,µ(t) dt
0
1− 1
= |µ̇(t)| dt
0 1 − |µ(t)|2
1− 1
= · (1 + [w (t)]2 )1/2 dt.
0 1 − t 2 − [w(t)]2
However,
1 1
≥ and (1 + [w (t)]2 )1/2 ≥ 1.
1 − t2 − [w(t)] 2 1 − t2
We conclude that
1− 1
ρ (µ) ≥ dt = ρ (γ ).
0 1 − t2
This is the desired result.
Notice that, with only small modifications, this argument can also
be applied to piecewise continuously differentiable curves t + iw(t).
then it may cross itself. Of course we can eliminate the loops and
thereby create a shorter curve. If the resulting curve is still not the graph
of a function, then elementary comparisons show that it will be longer
than a curve of the form (∗) (see Figure 2). We may conclude that the
curve γ in the example is the shortest of all curves connecting 0 to
(1 − ) + 0i.
Riemannian Metrics and the Concept of Length 37
(1 - ) + 0i
0
Figure 2.
dρ (P, Q) = inf ρ (γ ) : γ ∈ C (P, Q) .
Check for yourself that the resulting notion of distance satisfies the
classical metric axioms listed at the outset of this section.
There is some subtlety connected with defining distance in this
fashion. If ρ(z) ≡ 1, the Euclidean metric, and if is the entire plane,
then dρ (P, Q) is the ordinary Euclidean distance from P to Q. But
if and P and Q are as shown in Figure 3, then there is no shortest
curve connecting P to Q. The distance from P to Q is suggested by
the dotted curve, but notice that this curve does not lie in . The crucial
issue here is whether the domain is complete in the metric, and we shall
have more to say about this point later.
We may use the language and notation of distance to summarize
the meaning of Examples 4 and 5 and the discussion following them: If
ρ(z) = 1/(1 − |z|2 ) is a metric on the disc D, P = 0, and Q = R + i0,
then dρ (P, Q) = (1/2) log((1 + R)/(1 − R)).
38 Basic Notions of Differential Geometry
Figure 3.
∂
f (z) = 0 on
∂z
Calculus in the Complex Domain 39
if and only if
1 ∂ ∂
+i u(z) + iv(z) = 0 on .
2 ∂x ∂y
Taking real and imaginary parts of this last equation leads to the pair of
equations
∂ ∂ ∂ ∂
u= v and v = − u.
∂x ∂y ∂x ∂y
But these are just the Cauchy-Riemann equations. We conclude that
∂
f = 0 on iff f is holomorphic on .
∂z
Further notice that
∂ ∂
z=1 z=0
∂z ∂z
∂ ∂
z=0 z = 1.
∂z ∂z
Thus ∂/∂z and ∂/∂z fit in a natural way into complex function theory.
Next observe that if f is holomorphic then
f (z + h) − f (z)
f (z) = lim
Ch→0 h
f (z + s) − f (z)
= lim
Rs→0 s
∂f
=
∂x
∂f
= −i
∂y
∂f
= .
∂z
Thus we see that the notion of complex derivative, introduced in Sec-
tion 0.1, is for holomorphic functions the same as ∂ f /∂z.
40 Basic Notions of Differential Geometry
∂2 ∂2
= +
∂x2 ∂y 2
may now be written as
∂ ∂ ∂ ∂
=4 =4 .
∂z ∂z ∂z ∂z
If we are going to use complex derivatives, then we need the usual
tools associated to derivatives. The linear properties (that is, the sum
and scalar multiplication rules) of ∂/∂z and ∂/∂z are obvious. The
product rule is a bit more tedious, but is safely left as an exercise. How-
ever, the chain rule is more subtle; it now takes the following form:
Proof. We will sketch the proof of the first identity and leave the sec-
ond as an exercise.
We have
∂ 1 ∂ ∂
( f ◦ g) = −i ( f ◦ g).
∂z 2 ∂x ∂y
We write g(z) = α(z) + iβ(z), with α and β real-valued functions, and
apply the usual calculus chain rule for ∂/∂ x and ∂/∂y. We obtain that
the last line equals
1 ∂ f ∂α ∂ f ∂β ∂ f ∂α ∂ f ∂β
+ −i −i . (∗)
2 ∂x ∂x ∂y ∂ x ∂ x ∂y ∂y ∂y
Calculus in the Complex Domain 41
Exercise.
∂2 a 2 (zz)a−1
log(1 + (zz)a ) = .
∂z∂z (1 + (zz)a )2
42 Basic Notions of Differential Geometry
( f ◦ g) = ( f ◦ g) · |g |2 .
3. Isometries
In any mathematical subject there are morphisms: functions which pre-
serve the relevant properties being studied. In linear algebra these are
invertible linear maps, in Euclidean geometry these are rigid motions,
and in Riemannian geometry these are “isometries.” We now define the
concept of isometry.
f : 1 → 2
Remark. The particular form that we use to define the pullback is mo-
tivated by the way that f induces mappings on tangent and cotangent
vectors, but this motivation is irrelevant for us here.
It should be noted that the pullback of any metric under a conju-
gate holomorphic f will be the zero metric. Thus we have designed our
definition of pullback so that holomorphic pullbacks will be the ones
of greatest interest. This assertion will be made substantive in Proposi-
tion 3 below.
Isometries 43
f : 1 → 2
f ∗ ρ2 (z) = ρ1 (z)
for all z ∈ 1 then f is called an isometry of the pair (1 , ρ1 ) with the
pair (2 , ρ2 ).
The differential definition of isometry (Definition 2) is very natu-
ral from the point of view of differential geometry, but it is not intuitive.
The next proposition relates the notion of isometry to more familiar
ideas.
f : 1 → 2
ρ1 (γ ) = ρ2 ( f ∗ γ ).
(c) Part (b) implies that the isometry f is one-to-one. Then f −1 is well
defined and f −1 is also an isometry.
By definition,
b
ρ2 ( f ∗ γ ) =
( f ∗ γ ) (t)
dt
ρ2 , f ∗ γ (t)
a
b
∂f
=
(γ (t)) · γ̇ (t)
dt.
∂z
a ρ2 , f ∗ γ (t)
since f is an isometry. Substituting this back into the formula for the
length of f ∗ γ gives
b
ρ2 ( f ∗ γ ) = γ̇ (t)ρ1 ,γ (t) dt = ρ1 (γ ).
a
hence
In other words,
1 − |a|2
=
|1 − az|2 − |z − a|2
1 − |a|2
=
1 − |z|2 − |a|2 + |a|2 |z|2
1
=
1 − |z|2
= ρ(z).
The Poincaré Metric 47
Q (P)
(Q)
P
Figure 1.
Next we have
since there is a rotation of the disc taking φ(Q) to |φ(Q)| + i0. Finally,
P−Q
|φ(Q)| = ,
1 − PQ
so that (∗) together with the special case treated in the the first sentence
gives the result.
These discs form a neighborhood basis for the origin in the Euclidean
topology. Thus we find that the two topologies are the same at the ori-
gin. Now the origin can be moved to any other point a ∈ D by the
Möbius transformation
The Poincaré Metric 49
z+a
z −→ .
1 + az
Since the Poincaré metric is invariant under Möbius transformations,
and since Möbius transformations take circles to circles (after all, they
are linear fractional), the two topologies are the same at every point.
One of the most striking facts about the Poincaré metric on the
disc is that it turns the disc into a complete metric space. How could this
be? The boundary is missing! The reason that the disc is complete in
the Poincaré metric is the same as the reason that the plane is complete
in the Euclidean metric—the boundary is infinitely far away. We now
prove this assertion.
Proposition 4. The unit disc D, when equipped with the Poincaré met-
ric, is a complete metric space.
dρ (0, p j ) ≤ M, all j.
e2M − 1
|pj| ≤ < 1.
e2M + 1
Thus our sequence is contained in a relatively compact subset of the
disc. A similar calculation yields that the sequence must in fact be
Cauchy in the Euclidean metric. Therefore it converges to a limit point
in the disc, as required for completeness.
50 Basic Notions of Differential Geometry
Figure 2.
Figure 3.
52 Basic Notions of Differential Geometry
~
Q
X
~
P
Figure 4.
z + z0
h(z) =
1 + z0z
The Poincaré Metric 53
then
h∗ρ
(0) = ρ
(0).
|h (0)|
ρ (h(0)) = ρ
(0)
or
1
(z 0 ) =
ρ ·ρ
(0) = ρ
(0) · ρ(z 0 ).
1 − |z 0 |2
Thus we have exhibited ρ
as the constant ρ
(0) times ρ.
Now that we know that the Poincaré metric is the right metric for
complex analysis on the disc, a natural next question is to determine
which other maps preserve the Poincaré metric.
Proof. First suppose that f (0) = 0. For R > 0 let C R be the set of
points in D that have Poincaré distance R from 0. Since the Poincaré
metric is invariant under rotations (after all, rotations are holomorphic
self-maps), it follows that C R is a Euclidean circle (however this circle
will have Euclidean radius (e2R −1)/(e2R +1), not R). Since f (0) = 0
and f preserves the metric, it follows that f (C R ) = C R . The fact that
f is distance-preserving then shows that f is one-to-one on each C R .
As a result, f is globally one-to-one.
Let P ∈ C R . Then
| f (P) − f (0)| | f (P)|
= = 1.
|P − 0| |P|
54 Basic Notions of Differential Geometry
g = ψ ◦ f ◦ φ.
f ∗ ρ(z) ≤ ρ(z).
ρ ( f ∗ γ ) ≤ ρ (γ ).
Proof. Now
1
f ∗ ρ(z) ≡ | f (z)|ρ( f (z)) = | f (z)| ·
1 − | f (z)|2
and
1
ρ(z) =
1 − |z|2
56 Basic Notions of Differential Geometry
Image of f
Figure 1.
The Schwarz Lemma 57
g (z 0 ) = (1 + ) f (z 0 ).
g ∗ ρ(z 0 ) ≤ ρ(z 0 ).
(1 + ) · f ∗ ρ(z 0 ) ≤ ρ(z 0 ).
Note that this inequality holds for any z 0 ∈ D. But now if γ : [a, b] →
D is any continuously differentiable curve, then we may conclude that
ρ ( f ∗ γ ) ≤ (1 + )−1 ρ (γ ).
The next result shows how to find the fixed point, and in effect
proves the contraction mapping fixed-point theorem in this special case.
B(P, R) = {z ∈ D : d(z, P) ≤ R}
58 Basic Notions of Differential Geometry
be the closed Poincaré metric ball with center P and Poincaré metric
radius R. Then the theorem tells us that
Observe that
∞
B(P, j) = D.
j=1
For God’s sake, I beseech you, give it up. Fear it no less than sensual
passions because it, too, may take all your time, and deprive you of your
health, peace of mind, and happiness.
János Bolyai did ultimately crack the problem (a couple of years in ad-
vance of Lobachevsky), producing an independent geometry in which
the parallel postulate fails. The elder Bolyai conveyed the discovery
to his friend Carl Friedrich Gauss (1777–1855). One can only imagine
János’s chagrin to read this reply from Gauss:
If I begin with the statement that I dare not praise such a work, you
will of course be startled for a moment: but I cannot do otherwise; to
praise it would amount to praising myself; for the entire content of the
work, the path which your son has taken, the results to which he is led,
coincide almost exactly with my own meditations which have occupied
my mind for from thirty to thirty-five years. On this account I find myself
surprised to the extreme.
My intention was, in regard to my own work, of which very little
up to the present has been published, not to allow it to become known
during my lifetime. Most people have not the insight to understand our
conclusions and I have encountered only a few who received with any
particular interest what I communicated to them. In order to understand
these things, one must first have a keen perception of what is needed,
and upon this point the majority are quite confused. On the other hand,
it was my plan to put all down on paper eventually, so that at least it
would not finally perish with me.
60 Basic Notions of Differential Geometry
These are the standard four axioms which give our Euclidean con-
ception of geometry. The fifth axiom, a topic of intense study for two
A Detour into Non-Euclidean Geometry 61
D
C
Figure 1.
P5 For each line and each point P that does not lie on there is a
unique line m through P such that m is parallel to .
Figure 2.
Figure 3.
(Figure 3). Then the endpoints of this arc will certainly have the form
−(1 − a 2 ) −2a (1 − a 2 ) −2a
, , ,
1 + a2 1 + a2 1 + a2 1 + a2
for some real a between −1 and 0. Then we see that our circular arc C
is just φia ().
Of course we take a point in our new geometry to be just a stan-
dard Euclidean point. Then it is straightforward to verify that the lines
satisfy axioms P1 and P2. Axiom P3 is not relevant for our discussion
of the parallel postulate since it relates to circles. Axiom P4 follows
from conformality.
The punch line, of course, is that Axiom P5 is not satisfied by the
lines in our new geometry. We repeat that two lines are parallel if they
do not intersect. Now fix the line as before, and consider the point
p = (0, 1 − ) not on that line. There is certainly a circular arc m of
the kind we have been discussing that (i) passes through p and (ii) is
disjoint from (Figure 4). Now any small rotation of m will also be
disjoint from and will intersect m in a point (Figure 5). Call that point
of intersection q. Then q is a point that is not on the line , yet we
have produced two distinct lines that pass through it and are disjoint
64 Basic Notions of Differential Geometry
Figure 4.
pq
Figure 5.
Figure 6.
CHAPTER 2
Curvature and Applications
− log ρ(z)
κU,ρ (z) = κ(z) ≡ . (∗)
ρ(z)2
67
68 Curvature and Applications
− log[ρ(h(z)) · |h
(z)|]
κU1 ,h ∗ ρ (z) ≡
[ρ(h(z)) · |h
(z)|]2
− log[ρ(h(z))] − [log |h
(z)|]
= .
[ρ(h(z)) · |h
(z)|]2
Remark. In fact the proof gives the slightly more general fact that if
U1 , U2 are domains and f : U1 → U2 is a holomorphic map (not nec-
essarily one-to-one or onto), then the following holds. If ρ is a metric
on U2 , then
κ f ∗ ρ (z) = κρ ( f (z))
It was Ahlfors [AHL1] who first realized that the Schwarz lemma
is really an inequality about curvature. In his annotations to his col-
lected works he modestly asserts that “This is an almost trivial fact
and anybody who sees the need could prove it at once.” However, he
goes on to say (most correctly) that this point of view has been a de-
cisive influence in modern function theory. It is a good place to begin
our understanding of curvature. Here is Ahlfors’s version of Schwarz’s
lemma:
f ∗ σ (z) ≤ ρ(z), ∀z ∈ D.
Proof following [MIS]. Let 0 < r < 1. On the disc D(0, r ) define the
metric
r
ρr (z) = .
r 2 − |z|2
Curvature and the Schwarz Lemma Revisited 71
f ∗σ
v= .
ρr
Observe that v is continuous and nonnegative on D(0, r ) and that
v → 0 when |z| → r (since f ∗ σ is bounded above on D(0, r ) ⊂ D
while ρr → ∞). It follows that v attains a maximum value M = M(r )
at some point τ ∈ D(0, r ). We will show that M ≤ 1, hence v ≤ 1 on
D(0, r ). Letting r → 1− then finishes the proof.
If f ∗ σ (τ ) = 0, then v ≡ 0. So we may suppose that f ∗ σ (τ ) > 0.
Therefore κ f ∗ σ is defined at τ . By hypothesis,
κ f ∗ σ ≤ −4.
This gives
f ∗ σ (τ )
≤1
ρr (τ )
or
M≤1
as desired.
f ∗ σ (0) ≤ ρ(0).
| f
(0)| · σ ( f (0)) ≤ ρ(0).
| f
(0)| ≤ 1.
2α
ραA (z) = √ .
A(α 2 − |z|2 )
κσ (z) ≤ −B < 0
for some positive constant B and for all z ∈ U . Then any holomorphic
function
f :C→U
must be constant.
f : D(0, α) → U.
Here D(0, α) is the Euclidean disc with center 0 and radius α, equipped
with the metric ραA (z) as in the last section. Theorem 4 of Section 1
yields, for any fixed z and α > |z|, that
√
∗ A
f σ (z) ≤ √ ραA (z).
B
74 Curvature and Applications
Letting α → +∞ yields
f ∗ σ (z) ≤ 0,
hence
f ∗ σ (z) = 0.
S = C \ {x + i0 : 0 ≤ x ≤ 1}.
z−1
c(z) =
z+1
takes the right half plane to the unit disc. So u(z) = c ◦ h(z) is a
bounded entire function. We conclude from Theorem 1 that u is con-
stant. Unravelling our construction, we have that f is constant.
The point of this easy example is that, far from being bounded,
an entire function need only omit a segment from its values in order
that it be forced to be constant. And a small modification of the proof
shows that the segment can be arbitrarily short. Picard considered the
question of how large a set can be omitted from the image of a non-
constant entire function.
Let us pursue the same line of inquiry rather modestly by asking
whether a non-constant entire function can omit one complex value.
The answer is “yes,” for f (z) = e z assumes all complex values except
zero. It also turns out (see Section 3.5) that it is impossible to construct
a metric on the plane less a point that has negative curvature bounded
away from zero.
The next step is to ask whether a non-constant entire function f
can omit two values. The striking answer, discovered by Picard, is “no.”
Because of Theorem 1, it suffices for us to prove the following:
κµ (z) ≤ −B < 0
∂2
=2 log 1 + [z · z]1/6 .
∂z∂z
Entire functions are of two types: there are polynomials and non-
polynomials (transcendental entire functions). Notice that a polyno-
mial has a pole at infinity. Conversely, any entire function with a pole
at infinity is a polynomial (see Section 0.4). So a transcendental entire
function cannot have a pole at infinity and, being unbounded (by Liou-
ville), cannot have a removable singularity at infinity. It therefore has
an essential singularity at infinity.
Now notice that, by the fundamental theorem of algebra, a poly-
nomial assumes all complex values. For a transcendental function, we
analyze its essential singularity at infinity by recalling the Casorati-
Weierstrass theorem (see Section 0.4): if 0 is an essential singularity for
a holomorphic function f on a punctured disc D
(0, ) ≡ D(0, ) \ {0}
then f assumes values on D
(0, ) which are dense in the complex
plane. One might therefore conjecture that the essential feature of Pi-
card’s theorem is not that the function being considered is entire, but
Normal Families and the Spherical Metric 79
| f (z)| ≤ M K , ∀z ∈ K , f ∈ F, (∗)
| f (z)| ≤ M, ∀z ∈ , f ∈ F,
Example 3. Let
U0 = C \ {x + i0 : 0 ≤ x ≤ 1}.
p(z)
Figure 1.
2x 2y 2
p(x, y) = , ,1 − .
1+x +y 1+x +y
2 2 2 2 1 + x 2 + y2
2
σ (z) =
1 + |z|2
Normal Families and the Spherical Metric 83
z−w
dσ (z, w) = 2 arctan . (∗)
1 + wz
Differentiation of the formula (∗) along a line in the plane gives rise
to the definition of σ . (Be reassured that σ can also be thought of as
the pullback of the surface Euclidean metric τ on the sphere by the
map p. However, a precise formulation of this assertion would take us
far afield and we omit it.) As an exercise, the reader may wish to derive
the formula for dσ (z, w) by imitating our derivation of the formula
for Poincaré distance (Proposition 2 of Section 1.4): first calculate the
special case z = 0, w = r + 0i, then use the invariance of the metric
under conformal maps of the Riemann sphere C (i.e., under suitable
linear fractional transformations).
Now suppose that in our study of normal families we study not
complex-valued holomorphic functions but meromorphic functions. A
meromorphic function is nothing other than a holomorphic function
taking values in the Riemann sphere Ĉ. In order to make this last point
clear, let m(z) be a meromorphic function on a domain U with a pole at
P ∈ U . Set I (z) = 1/z. Then I ◦ m is holomorphic in the usual sense
in a neighborhood of P. This is precisely the definition of the property
“holomorphic” for a Ĉ-valued function. Note that if the pole at P is of
order k then I ◦ m has a zero at P of order k, and conversely.
With this increased generality we can now give a more elegant
definition of “normal family.” This new definition captures the spirit of
the first one, but is strictly more general.
Definition 3
. Let U ⊆ C be a domain. Let F be a family of functions
f : U → C. Then F is called a normal family if every sequence of
elements in F has a subsequence that converges uniformly on compact
sets to a function from U to C. [Note that, in order to make sense of
this condition, one must use the provided metric on the range, which is
the Riemann sphere.]
84 Curvature and Applications
Exercises. The point of our new definition, and you should prove this
as an exercise, is that a compactly divergent sequence (according to
Definition 2 above) is actually normally convergent to the point ∞ ∈ Ĉ.
You should note that if we specialize down to holomorphic functions,
then Hurwitz’s theorem guarantees that any family which is normal
according to Definition 3
is also normal according to Definition 3.
As a second exercise, discuss Example 1 in light of this new defi-
nition.
{ f ∗ τ : f ∈ F}
2| f
(z)|
≤ MK .
1 + | f (z)|2
2| f
(z)|
≤ MK , ∀ f ∈ F, z ∈ K , (∗)
1 + | f (z)|2
Normal Families and the Spherical Metric 85
But this last is just M K times the Euclidean length of γ . Notice that
this estimate is independent of f ∈ F. By our definition of distance, it
follows that F is an equicontinuous family of functions when thought
of as mapping K equipped with the Euclidean metric to Ĉ equipped
with the spherical metric. A metric-space-valued version of the As-
coli/Arzelà theorem (really just the same as the one we discussed in
Section 0.3) now yields immediately that F is a normal family.
Now we treat the converse direction. Suppose that F is a normal
family. Define the expression (the spherical derivative)
2| f
(z)|
f # (z) = .
1 + | f (z)|2
The thrust of Marty’s theorem (and its proof) is that the condition of
being a normal family is really an equicontinuity condition on the fam-
ily, when measured in the correct metric. That metric becomes plain
when the functions are viewed as taking values in Ĉ.
In the next section we will derive two striking consequences of the
work done here.
σ (z) ≤ M · µ(z).
Therefore we have
We can now use this generalized Montel theorem to obtain the refine-
ment of Picard’s theorem which was promised at the end of Section 2.
This result is known as Picard’s great theorem.
0. Introductory Remarks
Refer to Section 0.3 for the statement and sketch of the proof of the
Riemann mapping theorem. The Riemann mapping function is the so-
lution to a certain extremal problem: to find a map of the given domain
U into the disc D which is one-to-one, maps a given point P to 0, and
has derivative of greatest possible modulus λ P at P. The existence of
the extremal function, which also turns out to be one-to-one, is estab-
lished by normal families arguments; the fact that the extremal function
is onto is established by an extra argument which is in fact the only step
of the proof where the topological hypotheses about U are used.
The point of the present discussion is to observe that the scheme
we just described can be applied even if U is not topologically equiv-
alent to the disc. Constantin Carathéodory’s brilliant insight was that
the quantity λ P can be used to construct a metric, now called the
Carathéodory metric. We maximize the derivative at P of maps φ of
U into D such that φ(P) = 0 but we no longer require the maps to
be injective. Of course the proof of the Riemann mapping theorem will
break down at the stage where we attempt to show that the limit map is
surjective; we will also be unable to prove that it is injective. All other
steps, including the existence of the extremal function, are correct and
give rise to a metric (as in Section 3.1 following) on the domain.
89
90 Some New Invariant Metrics
Definition 1. If P ∈ U , define
theorem. In that proof, it was necessary only to know that the extreme
value existed and was finite. Now we shall gain extra information by
comparing this value to other quantities.
ζ−P
φ : ζ −→
2R
satisfies φ ∈ (D, U ) P . Thus
1
FCU (P) ≥ |φ (P)| = > 0.
2R
h ∗ ρ2 (z) ≤ ρ1 (z), ∀z ∈ U1 .
92 Some New Invariant Metrics
ρ2 (h ∗ γ ) ≤ ρ1 (γ ).
The corollaries are proved just as they were proved for the Poincaré
metric in Chapter 1. We concentrate on proving the Proposition.
or
ρ1 (P) ≥ h ∗ ρ2 (P).
The Kobayashi Metric 93
Proof. First we calculate the metric at the origin. If φ ∈ (D, D)0 then,
by the Schwarz lemma, |φ (0)| ≤ 1. Therefore
FCD (0) ≤ 1.
φ(ζ ) = ζ
FCD (0) = 1.
Definition 1. If P ∈ U , define
|(φ ◦ ψ) (0)| ≤ 1
or
1
|φ (P)| ≤ .
|ψ (0)|
Taking the supremum over all φ gives
1
FCU (P) ≤ .
|ψ (0)|
Now taking the infimum over all ψ yields
φ R (ζ ) = P + Rζ
h ∗ ρ2 (z) ≤ ρ1 (z), ∀z ∈ U1 .
ρ2 (h ∗ γ ) ≤ ρ1 (γ ).
1
FKU2 (Q) ≤
|(h ◦ φ) (0)|
1
= .
|h (P)||φ (0)|
1
FKU2 (Q) ≤ · FKU1 (P)
|h (P)|
or
(h ∗ FK 2 )(P) ≤ FK 1 (P).
U U
h : C → C0
ζ → eζ
φ(ζ ) = ζ.
Then
1
FKD (0) ≤ = 1 = ρ(0).
|φ (0)|
It follows that
h:U → D
98 Some New Invariant Metrics
= (h ∗ ρ)(P);
= (h ∗ FKD )(P).
1
→ FKU (P).
|ψ j (0)|
h jk = φ jk ◦ ψ jk .
h = φ ◦ ψ .
ζ = h 0 (ζ )
= lim (φ ◦ ψ (ζ )) (∗∗)
→∞
= φ0 ◦ ψ0 (ζ ).
f : U −→ U
or
and that the metric does not vanish at P. Then f is a conformal map
of U onto U .
implies that
| f (P)| = 1.
Define
f1 = f
f2 = f ◦ f
···
f j
= f j−1
◦ f, j ≥ 2.
Then { f j } is a normal family. Since the numbers ( f j ) (P) all have unit
modulus (note that ( f j ) (P) = [ f (P)] j ), there is a subsequence { f }
j
j
such that ( f ) (P) → 1. (Exercise: if f (P) has argument which is
a rational multiple of π, then the assertion is clear; if it is an irrational
multiple of π, then the set of ( f j ) (P) forms a dense subset of the
circle.) Passing to another subsequence, which we still denote by { f },
j
m
∂
f˜kq (P)
∂z
Similarly,
g ◦ f (z) ≡ z.
Proof. We will prove the second set of inequalities. The proof of the
first inequalities is similar.
Now let P be a point in . Let r be a positive number that is
less than one third of the Euclidean distance of P to the boundary. Let
D(P, r ) ⊆ D(P, 2r ) be closed discs in . If z ∈ D(P, r ) then
D(P,2r )
ρK (z) ≥ ρ K (z).
D(P,2r ) 1 D(0,1)
ρK (z) = · ρK ((z − P)/[2r ]).
2r
And of course we know the Kobayashi metric on the unit disc explicitly.
D(0,1)
It equals the Poincaré metric. Hence ρ K ((z − P)/[2r ]) ≤ 4/3.
Putting our estimates together, we find that
4 1
ρ K (z) ≤ · .
3 2r
This is one half of our estimate.
For the other half, notice that since is bounded it is contained in
some large disc open D(0, R). Then, for z ∈ D(P, r ),
ρ K (z) ≥ ρ K
D(0,R)
(z).
D(0,1)
Now, by conformal mapping, this last is equal to [1/R]ρ K (z/[R]).
Thus it is at least equal to 1/R.
In summary, we have proved for points z ∈ D(P, r ) that
4 1
≥ ρ K (z) ≥ .
6r R
The compact set L ⊂ can be covered by finitely many discs of the
form D(P, r ). So, in the end, we obtain a uniform estimate
C3 ≤ ρ K (z) ≤ C4
Completeness of the Carathéodory and Kobayashi Metrics 105
Proof. Since the metrics are comparable, the induced balls are compa-
rable. The balls form a sub-basis for the topology. So the topologies are
comparable.
A bounded domain with infinitely many holes will not have C k bound-
ary, k ≥ 1, since any defining function will fail to be smooth at an
accumulation point of the holes. See [KRP2] for a thorough treatment
of defining functions.
Figure 1.
Completeness of the Carathéodory and Kobayashi Metrics 107
The curve
π π
µ(t) = (cos 2t)(cos t) + i(cos 2t)(sin t), − ≤t≤
4 4
Figure 2.
Figure 3.
108 Some New Invariant Metrics
Figure 4.
The functions
∂U P −→ ν P ∈ R2
and
∂U P −→ ν P ∈ R2
inf{|z − Q| : Q ∈ ∂U } = |z − P|.
Figure 5.
Proof. Define
T : ∂U × (−1, 1) −→ C
(Q, t) −→ Q + tν Q .
T : [0, 1] × (−1, 1) −→ C
(s, t) −→ γ (s) + tνγ (s) .]
-1
Figure 6.
110 Some New Invariant Metrics
z = Q + tν Q .
Figure 7.
Figure 8.
Figure 9.
Figure 10.
Completeness of the Carathéodory and Kobayashi Metrics 113
iP : U −→ D(C(P), r0 )
(r0 )2
ζ −→ C(P) +
ζ − C(P)
114 Some New Invariant Metrics
j P : D(C(P), r0 ) −→ D(0, 1)
ζ − C(P)
ζ −→
r0
is holomorphic. To estimate the Carathéodory metric at z, we use the
distance-decreasing property of the Carathéodory metric:
FCU (z) ≥ (j P ◦ i P )∗ FC
D(0,1)
(z)
(∗)
≡ |(j P ◦ i P ) (z)|FC
D(0,1)
(j P (i P (z))).
z P C(P)
r0
Figure 11.
Recall that, on the unit disc, the Carathéodory metric coincides with the
Poincaré metric. It follows from our calculation of the Poincaré metric
in Proposition 2 of Section 1.4 that
D(0,1) 1
FC (j P ◦ i P (z)) =
(δ/(r0 + δ))(2 − δ/(r0 + δ))
r0 + δ
≥ (∗ ∗ ∗)
2δ
r0 1
≥ · .
2 δ
In summary, using (∗), (∗∗), and (∗ ∗ ∗), we have
r0 r0 1 1
FCU (z) ≥ · · ≥ C0 · ,
(δ + r0 )2 2 δ δ
where C0 is a positive constant which depends only on r0 .
But this is precisely the estimate which enabled us, in Section 1.4,
to prove that the disc is complete in the Poincaré metric. We leave
it now as an exercise to provide the details which show that the
Carathéodory distance from any fixed interior point P0 ∈ U to a point
with Euclidean distance δ from the boundary has size C ·(1+| log 1/δ|)
and to conclude that U is complete in the Carathéodory metric.
116 Some New Invariant Metrics
Exercise. Let
U = D \ {0}
be the punctured disc. See Figure 12. This domain does not have
C 2 boundary. Use the Riemann removable singularities theorem and
Cauchy estimates to determine the behavior of the Carathéodory met-
ric near the boundary point 0 of U . Conclude that U is not complete in
the Carathéodory metric. What can you say about the Kobayashi met-
ric?
Figure 12.
Figure 13.
lim f (P − r ν P ) = .
r →0+
If α > 1, set
lim f (z) = .
α (P)z→P
(P)
Figure 14.
f : U1 −→ U2
The estimate
C
FCU (z) ≈ (∗)
dist(z, ∂U )
120 Some New Invariant Metrics
makes it a tedious but not difficult exercise to calculate that the regions
Mβ are comparable to the regions α (in this last formula, and in what
follows, “dist” means Euclidean distance). In point of fact, suppose
that z lies in some α ( p), some p ∈ ∂U j . Let us denote by τ p the
inward normal segment emanating from that boundary point p. Using
the estimate (∗), one can then estimate that distC (z, τ p ) ≤ C · α. For
the converse estimate, assume that z ∈ α ( p) and the same estimate
shows that distC (z, τ p ) ≥ C · α.
Thus we see that
iff
| f (P − tν P ) − Q| < .
f (z) ∈ B( f (P − tν P ), β).
But
We conclude that
| f (z) − Q| ≤ | f (z) − f (P − tν P )| + | f (P − tν P ) − Q|
≤ C + = C .
4. An Application of Completeness:
Automorphisms
This section is an introduction to a topic which is usually treated as
part of Riemann surface theory. (See [FK] for a nice treatment of the
Riemann surface point of view.) We shall not use any of this theory,
nor shall we make any reference to Riemann surfaces. Instead, we use
geometry.
ζ −a
ζ −→ µ ·
1 − aζ
where a, µ ∈ C, |a| < 1, and |µ| = 1.
the reflection
R ·r
σ : z −→ ,
z
An Application of Completeness: Automorphisms 123
Aut(U ) φ j −→ φ ∈ Aut(U )
α jk (z) = µ jk · z,
where the µ jk are unimodular constants. Since the unit circle is com-
pact, there is a subsequence—call it µm —which converges to a uni-
modular constant µ0 . But then the corresponding automorphisms
αm (z) = µm z
α jk (z) = σ (µ jk z),
αm (z) = σ (µm z)
β0 (ζ ) ≡ 1.
φ jk −→ φ0
normally.
Notice that the image of each φ j lies in U , hence the image of φ0
lies in the closure U of U . If φ0 is nonconstant then it satisfies the open
mapping principle. But
Aut(U ) φm −→ φ0 ∈ Aut(U ),
φ j (P) −→ w ∈ ∂U.
φ j (K ) ⊆ V.
See Figure 1.
K
(K)
j
Figure 1.
the metric ball B(P, R) contains K (see Proposition 1 in Section 3). Let
Q j = φ j (P). Since each φ j is distance-decreasing in the Carathéodory
metric, it follows that φ j (B(P, R)) ⊆ B(Q j , R). We claim that there is
a positive J such that whenever j ≥ J then B(Q j , R) ⊆ V . Assuming
the claim, we would then have
as required.
To prove the claim, recall from the proof of Theorem 6 in the last
section that (because the Carathéodory metric on U is complete) the
Euclidean radii of the metric balls B(Q j , R) must tend to 0. Choose
> 0 such that the Euclidean disc of center w and radius 2 lies in V .
We select J so large that when j > J , then both the Euclidean distance
of Q j to w is less than and the Euclidean radius of B(Q j , R) is less
than . The claim now follows from the triangle inequality.
φ j (P) −→ w ∈ ∂U,
for some w ∈ ∂U .
An Application of Completeness: Automorphisms 129
Let
γ : [0, 1] −→ U
Figure 2.
Let
K = {γ (t) : 0 ≤ t ≤ 1}.
: [0, 1] × [0, 1] −→ U ∩ V
such that
and
But then
(φ j )−1 ◦
has the property that φ P (P) = 0. Now if P and Q are arbitrary ele-
ments of D then (φ Q )−1 ◦ φ P is an automorphism of D that maps P
to Q. Thus the automorphism group of the disc does act transitively.
Finally, let
U = D(0, 1) \ d j.
j∈Z
Figure 3.
σ ≤ FKU .
or
1 σ (z)
≥ = σ (z).
|φ (0)| ρ(0)
Taking the infimum over all φ we conclude, for any ξ ∈ C, that
|ξ |
#ξ #µ,z = inf ≥ |ξ | · σ (z) = #ξ #σ,z ,
φ∈(U,D)z |φ (0)|
as desired.
Definition 2. A domain U ⊆ C is called hyperbolic if the Kobayashi
distance on U is actually a distance, that is, if
dKob (P, Q) > 0
κ ≤ −B < 0
f ∗ σ̃ ≤ ρ.
σ̃ ≤ FKU
or
C · σ ≤ FKU ,
√ √
where C = B/ 4 is a positive constant. Thus the Kobayashi dis-
tance is bounded from below by a positive constant times the σ -
distance. Since σ is assumed to be nondegenerate, so is FKU .
Corollary 3.1. Neither the plane C nor the punctured plane C \ {0}
possesses a metric of strictly negative curvature.
f n (z) = enz
Remark. One theme which comes through in this section is that the
only enemies are C and C \ {0}. As soon as a domain excludes at
least two points, then it has all relevant properties. In the theory of
Riemann surfaces it also holds that a domain is taut if and only if it is
hyperbolic (see [FK]). However, in several complex variables, matters
are no longer so simple. A detailed discussion of the relationship be-
tween tautness and hyperbolicity appears in [KOB2, p. 240]. See also
[KOB1].
0. Introductory Remarks
It is a remarkable fact—discovered by Stefan Bergman in 1927—that
a bounded domain in C or in Cn can be equipped with a “canoni-
cal” reproducing kernel. [Here we use the phrase “reproducing kernel”
to mean a function k(z, ζ ) of two variables—like the familiar Cauchy
kernel—with the property that integration against a holomorphic func-
tion f produces the value of f at z. In one complex variable such a
formula could take the form
f (z) = k(z, ζ ) f (ζ ) dζ
∂
or
f (z) = k(z, ζ ) f (ζ ) dξ dη
137
138 Introduction to the Bergman Theory
1. Bergman Basics
The Bergman theory centers on a special function space. Let ⊆ C be
a bounded domain. Now define the Bergman space
A2 () = { f holomorphic on : | f (z)|2 d A(z) < ∞}.
1. f A2 () ≥ 0;
2. f + g A2 () ≤ f A2 () + g A2 () ;
3. c f A2 () = |c| · f A2 () for any complex constant c.
Observe that properties (1) and (3) already imply (2). Also (2)
and (3) imply (1). But it is useful to have these attributes all laid out
explicitly. It is important to note that properties (1)–(4) characterize the
140 Introduction to the Bergman Theory
parts as
φ(ζ ) = φ1 (ζ ) + iφ2 (ζ ).
· + · =
= |φ (ζ )|2 .
∂τ ∂τ ∂τ ∂τ ∂τ ∂ζ
In the last step we have used the fact that ∂φ/∂τ = ∂φ/∂ζ = φ for a
holomorphic function φ.
In conclusion,
area (2 ) = 1 d A(z) = det Jac φ(ζ ) d A(ζ )
2 1
= |φ (ζ )|2 d A(ζ ),
1
Now we have
entiable)
φ (z) · K 2 (φ(z), φ(ζ )) · φ (ζ ) f (ζ ) d A(ζ )
1
= φ (z) · K 2 (φ(z), φ(φ −1 (ξ )))
2
2
× φ (φ −1 (ξ )) f (φ −1 (ξ ))
[φ −1 ] (ξ )
d A(ξ )
= φ (z) · K 2 (φ(z), ξ )
2
× φ (φ −1 (ξ )) f (φ −1 (ξ ))
1
× d A(ξ )
φ (φ −1 (ξ ))φ (φ −1 (ξ ))
= φ (z) · K 2 (φ(z), ξ )
2
−1 1
× f (φ (ξ )) · d A(ξ )
φ (φ −1 (ξ ))
≡ φ (z) · K 2 (φ(z), ξ ) · g(ξ ) d A(ξ ),
2
where g(ξ ) = f (φ −1 (ξ )) · 1
φ (φ −1 (ξ ))
.
It is easy to verify, using a change of variable, that g ∈ A2 (2 ). As
a result, the Bergman kernel K 2 for 2 will reproduce g. We conclude
that
φ (z) · K 2 (φ(z), φ(ζ )) · φ (ζ ) f (ζ ) d A(ζ )
1
The kernel φ (z)· K 2 (φ(z), φ(ζ ))·φ (ζ ) plainly satisfies the con-
jugate symmetric property (3) of the Bergman kernel for 1 . We also
note that k is holomorphic in the z-variable and conjugate holomorphic
in the ζ -variable. Properties (1) and (2) may be checked by a change
of variables just like the one we performed to prove the reproducing
property. We omit the details.
We may conclude, by the uniqueness of the Bergman kernel, that
This is the standard area form of the mean value property for holomor-
phic or harmonic functions.
144 Introduction to the Bergman Theory
or
1 1
= K (w, ζ ) d A(ζ ).
π π D
1
= K (w, 0)
π
for any w ∈ D.
Now, for a ∈ D fixed, consider the Möbius transformation
z−a
h(z) =
1 − az
that we studied in Section 1.4. We know that
1 − |a|2
h (z) = .
(1 − az)2
We may thus apply Theorem 2 of Section 4.2 with φ = h to find that
normalized to have unit length. We may confirm the first of these as-
sertions by noting that if j = k, then
z j , z k = z j z k d A(z)
D
1 2π
= r j ei jθ r k e−ikθ dθr dr
0 0
1 2π
= r j+k+1
dr ei( j−k)θ dθ
0 0
= 0.
We conclude that
√
π
z = √j
.
j +1
∞
( j + 1)z j ζ j
=
j=0
π
1 ∞
= · ( j + 1) · (zζ ) j .
π j=0
∞
d ∞
( j + 1)α j = α j+1
j=0
dα j=0
d 1
= α·
dα 1−α
1
= .
(1 − α)2
Applying this result to our expression for K (z, ζ ) yields that
1 1
K (z, ζ ) = · .
π (1 − z · ζ )2
This is consistent with the formula which we obtained by conformal
invariance in Subsection 3.1 for the Bergman kernel of the disc.
us now summarize the key ideas. Unlike the first two Bergman ker-
nel constructions, the present one will work for any domain with C 2
boundary.
First, the fundamental solution for the Laplacian in the plane is
the function
1
(ζ, z) = log |ζ − z|.
2π
∂2
K (z, ζ ) = 4 · G(ζ, z). ()
∂ζ ∂z
Calculation of the Bergman Kernel 149
Proof. Our proof will use a version of Stokes’s theorem written in the
notation of complex variables. It says that if u ∈ C 1 (), then
∂u
u(ζ ) dζ = 2i · dτ dη, (∗∗)
∂U U ∂ζ
1 1
G(ζ, z) = log(ζ − z) + log (ζ − z) + H (ζ, z). (†)
4π 4π
Here we think of the logarithm as a multivalued holomorphic function;
after we take a derivative, the ambiguity (which comes from an additive
multiple of 2πi) goes away.
Differentiating with respect to z (and using subscripts to denote
derivatives), we find that
1 −1
G z (ζ, z) = + Hz (ζ, z).
4π ζ − z
We may rearrange this formula to read
1
= −4π · G z (ζ, z) + 4π Hz (ζ, z).
ζ −z
We know that G, as a function of ζ , vanishes on ∂. Hence so does
G z . Let f ∈ C 2 () be holomorphic on . It follows that the Cauchy
formula
1 f (ζ )
f (z) = dζ
2πi ∂ ζ − z
can be rewritten as
2
f (z) = f (ζ )Hz (ζ, z) dζ
i ∂
150 Introduction to the Bergman Theory
or
−2i f (ζ )Hz (ζ, z) dζ = f (z).
∂
∂2
K (z, ζ ) = 4 · G(ζ, z).
∂ζ ∂z
That is the desired result.
as a glance at any classical complex analysis text will tell us (see, for
example, [COH] or [HIL]).
With formula () in mind, we can make life a bit easier by writing
1 1
G(ζ, z) = log(ζ − z) + log(ζ − z)
4π 4π
1 1
− log (1 − ζ z) − log 1 − ζ z .
4π 4π
Here we think of the expression on the right as the concatenation of
four multi-valued functions, in view of the ambiguity of the logarithm
function. This ambiguity is irrelevant for us because the derivative of
the Green’s function is still well defined (i.e., the derivative annihilates
additive constants).
Now we readily calculate that
∂G 1 −1 1 ζ
= · + ·
∂z 4π ζ − z 4π 1 − ζ z
and
∂2G 1 1
= · .
∂ζ ∂z 4π (1 − ζ z)2
In conclusion, we may apply Proposition 2 to see that
1 1
K (z, ζ ) = · .
π (1 − z · ζ )2
This result is consistent with that obtained in the first two calculations
(Subsections 4.3.1, 4.3.2).
1
= − 2 · φ j φ j · φ j φ j
j φjφ j j j
1
+ · φ j φ j
j φjφ j j
2
1 1
= |φ j |2 − ·
φ φ
j
.
j |φ j | j |φ j |
2 2 j
j j
About the Bergman Metric 153
(w)
∂z∂z z=(w)
∂2
= log (w) · K 2 ((w), (w)) (w)
∂w∂w
Here we have inserted terms with derivative zero, and have also applied
the chain rule. Now we have that this expression
∂2
= log K 1 (w, w)
∂w∂w
= ρ1 (w).
Therefore
1 1
K (z, z) = · .
π (1 − |z|2 )2
As a result,
∂ −2 2z
log K (z, z) = · (−z) = .
∂z 1 − |z|2 1 − |z|2
Therefore
∂2 2
log K (z, z) = .
∂z∂z (1 − |z|2 )2
In conclusion,
√
2
ρ D (z) = .
1 − |z|2
This is just the same (up to a constant multiple) as the Poincaré metric
which we calculated for the disc in Section 1.4.
z−w
ψ(z, w) = log
1 − z · w
ψ(z, w)) but that ψ is not of the form studied in the present book. There
is no contradiction because ψ is not obtained by integration against a
weight function ρ. [The metric defined here is sometimes termed the
“pseudohyperbolic metric.”]
for every z ∈ .
Application to Conformal Mapping 157
where γr and γθ are, respectively, the normal and tangential compo-
nents of γ . Clearly, if we construct a new curve γ by integrating the
vector field γθ , then the result is a curve that is shorter than γ .
Thus any circle in M is a length-minimizing geodesic. And the
converse is true as well.
Now let φ be a conformal self-map of A. Then, as a result of the
considerations in the last paragraph, φ will map circles in M (concen-
tric with the annulus) to circles in M (concentric with the annulus). But
more is true: any circle c that is centered at 0 has constant Bergman dis-
tance from any one given circle C in M. Let τc be the Bergman distance
from the arbitrary circle c to the fixed circle C. Then c will be mapped
by φ to another circle in A that has Bergman distance τc from φ(C). In
short, φ maps circles to circles. And the same remark applies to φ −1 .
Now the orthogonal trajectories to the family of circles centered
at P will of course be the radii of the annulus A. Therefore (by confor-
mality) these radii will get mapped to radii. Let R be the intersection
of the positive real axis with the annulus A. After composition with a
rotation, we may assume that φ maps R to R.
Application to Conformal Mapping 159
If M has just one circle in it, then that circle C must be fixed.
Since each circle c in the annulus (centered at P) has a distance τc from
C, then its image under φ will have the same distance from C. Thus
either c is fixed or it is sent to its image under inversion. By continuity,
whatever choice is valid for c will be valid for all other circles. Thus,
in this case, the map φ is either the identity or inversion.
Now suppose that M contains at least two circles. The only possi-
bilities for the action of φ on M ∩ R are preservation of the order of the
points or inversion of the order (as any other permutation of the points
is ruled out by elementary topology of a conformal mapping). By com-
posing φ with an inversion (z → R/z), we may assume that φ does not
act as an inversion on R. But of course φ must preserve M ∩ R. We
conclude that φ must be the identity.
What we have just proved is that, after we normalize φ so that it
maps the positive real axis to itself, then in fact φ must be the identity
(or an inversion). In other words, the original map φ must be a rotation
(or an inversion).
CHAPTER 5
A Glimpse of Several Complex Variables
{(z 1 , z 2 ) ∈ C2 : |z 1 |2 + |z 1 |2 < 1}
161
162 A Glimpse of Several Complex Variables
Which of these should serve as the model domain for the multivariable
Riemann mapping theorem? Before answering that question, perhaps
one should ask whether there is a biholomorphic equivalence between
the ball and the bidisc. [According to a result of Liouville (see [DFN]),
there are none but trivial conformal mappings in dimensions three and
higher. As a result, in several complex variables we consider biholo-
morphic mappings. These are mappings which are holomorphic, one-
to-one, and onto. The holomorphicity of the inverse is automatic. See
further discussion below in Section 1.] If there is such an equivalence,
then the other question need not be answered; if there is no equivalence,
then perhaps there is no Riemann mapping theorem.
As Poincaré discovered, there is in fact no equivalence between
the ball and the bidisc. In the last two decades, it has been discovered
that, generically, two domains in C2 are not biholomorphically equiv-
alent. Thus the entire question of the Riemann mapping theorem turns
into a rather complex subject in itself—that of classifying domains ac-
cording to biholomorphic equivalence. One of the main things that we
shall do in this chapter is to formulate carefully and to prove Poincaré’s
theorem concerning the inequivalence of the ball and the bidisc.
The other big development which occurred near the turn of the
twentieth century is Hartogs’s theorem about analytic continuation. Al-
though we cannot treat this topic in detail here, we should like to dis-
cuss it briefly in order to further convey the flavor of several complex
variables. So that we may keep matters as simple as possible, we shall
avoid formal definitions in this discussion.
A domain U in complex space is called a domain of holomorphy
if there is a holomorphic function defined on U which cannot be ana-
lytically continued to any open set which properly contains U . It turns
out (see [KR1]) that the Weierstrass theorem of classical function the-
ory can be used to show that any open set U in the plane is a domain of
holomorphy. Briefly, let U be such an open set. Let S ⊆ U be a set in U
such that (i) S has no accumulation point in the interior of U and (ii) S
Functions of Several Complex Variables 163
Figure 1.
1. Basic Concepts
We introduce here only the few definitions and pieces of notation which
we shall need for our immediate purposes. For a more comprehensive
introduction to several complex variables, see [KR1, KR2, KR4]. For
simplicity of notation, we confine attention to two complex variables.
We consider domains U ⊆ C2 , where
C2 ≡ C × C
Uw = {z 1 ∈ C : (z 1 , w) ∈ U }
Basic Concepts 165
and
U z = {z 2 ∈ C : (z, z 2 ) ∈ U }.
U1 = {(z 1 , z 2 ) : z 2 = −1}.
The function
g(z 1 , z 2 ) = z 1 (z 2 )2
is holomorphic on
B(P, r ) = {z ∈ C2 : |z 1 − P1 |2 + |z 2 − P2 |2 < r 2 }
and
be, respectively, the open ball and open bidisc of center P and radius r .
2
We also let B(P, r ) and D (P, r ) denote, respectively, the closed ball
and closed bidisc (defined by replacing the symbol < with the sym-
bol ≤).
A basic tool in complex analysis is the Cauchy integral formula
and its variants. We now derive one such.
2
Proposition 2. If f is holomorphic on a neighborhood of D (P, r )
(the closure of the bidisc), then for all (z 1 , z 2 ) ∈ D 2 (P, r ) we have
1 f (ζ1 , ζ2 )
f (z 1 , z 2 ) = dζ2 dζ1 .
(2πi)2 ∂ D(P1 ,r ) ∂ D(P2 ,r ) (ζ1 − z 1 )(ζ2 − z 2 )
Now apply the one variable Cauchy integral formula to the function
f (ζ1 , z 2 ) in the integrand in the second variable: the desired formula
follows.
Proof. Differentiate under the integral sign, just as in one complex vari-
able.
2
which converges absolutely and uniformly on D (P, r ). The coeffi-
cients are given by
j k
1 ∂ ∂
a j,k = f (P).
j!k! ∂z 1 ∂z 2
Proof. The proof follows familiar lines from one complex variable.
2
Since f is holomorphic on a neighborhood of D (P, r ), we may select
2
r > r so that f is holomorphic on a neighborhood of D (P, r ). We
write
1 f (ζ1 , ζ2 )
f (z 1 , z 2 ) = dζ2 dζ1 .
(2πi) ∂ D(P1 ,r ) ∂ D(P2 ,r ) (ζ1 − z 1 ) · (ζ2 − z 2 )
2
(∗)
However,
1 1
=
(ζ1 − z 1 ) (ζ1 − P1 ) − (z 1 − P1 )
1 1
= (∗∗)
(ζ1 − P1 ) 1 − ((z 1 − P1 )/(ζ1 − P1 ))
∞
(z 1 − P1 ) j
= .
(ζ − P1 ) j+1
j=0 1
Similarly,
1 ∞
(z 2 − P2 ) j
= . (∗ ∗ ∗)
(ζ2 − z 2 ) (ζ − P2 ) j+1
j=0 2
∞
∞
1
f (z 1 , z 2 ) =
j=0 k=0
(2πi)2 ∂ D(P1 ,r ) ∂ D(P2 ,r )
168 A Glimpse of Several Complex Variables
f (ζ1 , ζ2 )dζ2 dζ1
(ζ1 − P1 ) j+1 (ζ2 − P2 )k+1
× (z 1 − P1 ) j (z 2 − P2 )k
∞
≡ a j,k · (z 1 − P1 ) j (z 2 − P2 )k .
j,k=0
Here
1 f (ζ1 , ζ2 )dζ2 dζ1
a j,k = .
(2πi)2 ∂ D(P1 ,r ) ∂ D(P2 ,r ) (ζ1 − P1 ) j+1 (ζ2 − P2 )k+1
The familiar Cauchy theory of one complex variable now tells us that
j k
1 ∂ ∂
a j,k = f (P).
j!k! ∂z 1 ∂z 2
Proposition 4. Let
j
S∼ a j,k z 1 z 2k
F : U1 −→ U2
F = ( f 1 (z 1 , z 2 ), f 2 (z 1 , z 2 ))
converges to
k
∂ ∂
f,
∂z 1 ∂z 2
uniformly on compact sets.
Proof. Let
| f (z)| ≤ M < ∞, ∀ f ∈ F.
2
It follows that F is an equicontinuous family on D (P, r ). Therefore
the result follows from the Ascoli/Arzelà theorem.
f : U1 −→ U2
Remark. Notice that JacC f (z) is distinct from the real Jacobian,
JacR f , of calculus: the latter would be a 4×4 matrix which arises from
treating f as a function from a domain in R4 to a domain in R4 .
then
f (z) ≡ z.
The Automorphism Groups of the Ball and Bidisc 173
f (z 1 , z 2 ) = 0 + z + Am (z) + · · ·
f 1 (z 1 , z 2 ) = f
f 2 (z 1 , z 2 ) = f ◦ f
···
f j (z 1 , z 2 ) = f j−1
◦ f, ∀ j ≥ 2.
f 2 = z + 2Am + · · ·
f 3 = z + 3Am + · · ·
etc.
f : U −→ U
ρθ (z 1 , z 2 ) = (eiθ z 1 , eiθ z 2 ).
g = ρ−θ ◦ f −1 ◦ ρθ ◦ f.
Then
hence ρ±θ is its own Jacobian. Since ρ±θ is a multiple of the identity,
it commutes with all other 2 × 2 matrices. So
g(z) ≡ z.
f ◦ ρθ = ρθ ◦ f. (∗)
The Automorphism Groups of the Ball and Bidisc 175
a j,k = 0
z 1,k = (1 − 1/k)sgn(g11 ), (1 − 1/k)sgn(g12 ) ,
z 2,k = (1 − 1/k)sgn(g21 ), (1 − 1/k)sgn(g22 ) .
Now
and
Letting k → +∞ gives
|g11 | + |g12 | ≤ 1
(∗)
|g21 | + |g22 | ≤ 1.
and
Therefore
max{|g11 |, |g21 |} = 1
max{|g12 |, |g22 |} = 1. (∗∗)
The only way that (∗) and (∗∗) can both be true is if each column
of the matrix (gi j ) has one entry of modulus one and one entry 0. Say
The Automorphism Groups of the Ball and Bidisc 177
|g j,σ ( j) | = 1, j = 1, 2
and
g j,k = 0 if k = σ ( j).
as desired.
if and only if
2
z1 − a 2 1/2 2
+ (1 − |a| ) z 2 < 1
1 − az 1 − az
1 1
178 A Glimpse of Several Complex Variables
if and only if
if and only if
if and only if
if and only if
z ∈ B(0, 1).
M : C2 −→ C2
Proof. Since B(0, 1) is complete circular, the map g must be linear (by
Proposition 6). Now g must preserve the boundary of B(0, 1). [Why?—
Think about a linear map acting on a closed, convex set; it will preserve
the extreme points.] Therefore the function g takes Euclidean unit vec-
tors to Euclidean unit vectors. But this says that g is unitary.
The map
g = φ|w| ◦ α ◦ f
f = α −1 ◦ (φ|w| )−1 ◦ β.
The reader may verify that (φ|w| )−1 = φ−|w| . The proof is therefore
complete.
Exercise. The group Aut(D 2 (0, 1)) acts transitively on D 2 (0, 1). The
group Aut(B(0, 1)) acts transitively on B(0, 1).
The ball and the bidisc are the two most fundamental domains in
C2 . Both are contractible, both have transitive automorphism groups,
and both are domains of convergence for power series. It came there-
fore as a great surprise when Poincaré proved that there is no biholo-
morphic mapping
He proved this result by first showing that if such a φ existed, then the
induced map
Notice that we are integrating the lengths, in the metric, of the tangent
vectors to the curve. The Carathéodory length of a curve is defined
similarly. We comment, just as in Section 3.1, that FC and FK are inte-
grable because they are semicontinuous.
One of the basic properties which we shall prove is that holomor-
phic mappings decrease distance in the Carathéodory and Kobayashi
metrics. In several variables we express this assertion as follows.
f : U1 −→ U2
Then
and
U U
FK 1 (P, ξ ) ≥ FK 2 ( f (P), f ∗ (P)ξ ).
182 A Glimpse of Several Complex Variables
Proof. We give the proof for the Carathéodory metric. The proof for
the Kobayashi metric is similar.
Choose φ ∈ (D, U2 ) f (P) . Then φ ◦ f ∈ (D, U1 ) P . Hence
Exercise. Check that the Proposition implies that f decreases the in-
variant lengths of curves. That is, if γ is a continuously differentiable
curve in U1 and f ∗ γ ≡ f ◦ γ is the corresponding curve in U2 , then
show that
C ( f ∗ γ ) ≤
C (γ ) and
K ( f ∗ γ ) ≤
K (γ ).
and
i KP to i K
Q as claimed.
The proof for iCP is identical.
i0K (B) = B.
h(ζ ) ≡ φ(ζ ) · η.
Here “·” denotes the usual inner product of 2-vectors. We have that h
maps the disc to the disc and h(0) = 0. By the Schwarz lemma of one
184 A Glimpse of Several Complex Variables
variable,
|h (0)| ≤ 1.
|φ (0)| ≤ 1.
satisfies φ0 ∈ (B, D)0 and φ0 (ζ ) = ξ/|ξ | for any ζ . Thus φ0 (0) =
(1/|ξ |) · ξ is a positive multiple of ξ . Therefore
We conclude that
FKB (0, ξ ) = |ξ |,
hence that
i0K (B) = B.
i0K (D 2 ) = D 2 .
π1 (z 1 , z 2 ) = z 1 and π2 (z 1 , z 2 ) = z 2 .
Invariant Metrics and the Inequivalence of the Ball and the Bidisc 185
But the Schwarz lemma of one variable tells us easily that the last quan-
tity is just |η1 |. A similar argument shows that
2
FKD (0, η) ≥ |η2 |.
Therefore
i0K (D 2 ) ⊆ D 2 .
For the reverse inclusion, fix η as above and consider the function
ζ η1 ζ η2
φ(ζ ) = , , ζ ∈ D.
max{|η1 |, |η2 |} max{|η1 |, |η2 |}
0 (B) = B and i0 (D ) = D .
Exercise. Verify that iC C 2 2
φ : D 2 −→ B
g : D 2 −→ B
JacC g(0) : D 2 −→ B
Epilogue
In complex analysis, geometric methods provide both a natural lan-
guage for analyzing and recasting classical problems and also a rubric
for posing new problems. The interaction between the classical and the
modern techniques is both rich and rewarding.
Many facets of this symbiosis have yet to be explored. In particu-
lar, very little is known about explicitly calculating and estimating the
differential invariants described in the present monograph. It is hoped
that this book will spark some new interest in these matters.
Appendix on the Structure Equations
and Curvature
1. Introduction
Here we give a brief presentation of the connection between the cal-
culus notion of curvature (see [THO]) and the more abstract notion of
curvature which leads to the definition of κ in Chapter 2. It is a plea-
sure to acknowledge our debt to the clear and compelling exposition in
[ONE].
First, a word about notation. We use the language of differential
forms consistently in this appendix. On the one hand, classically trained
analysts are often uncomfortable with this language. On the other hand,
the best way to learn the language is to use it. And the context of cur-
vature calculations on plane domains may in fact be the simplest non-
trivial context in which differential forms can be profitably used. In any
event, this appendix would be terribly clumsy if we did not use forms,
so the decision is essentially automatic. All necessary background on
differential forms may be found in [RU1] or in [ONE].
191
192 Appendix on the Structure Equations and Curvature
p
U (u, v)−→(x1 (u, v), x2 (u, v), x3 (u, v)) ∈ M.
E 1 : M −→ R3
E 2 : M −→ R3
a E 1 (x1 , x2 , x3 ) + bE 2 (x1 , x2 , x3 ), a, b ∈ R.
δ1 = (1, 0, 0),
δ2 = (0, 1, 0),
δ3 = (0, 0, 1).
(Many calculus books call these vectors i, j, and k.) Then we may
write
Ei = ai, j (x1 , x2 , x3 )δ j , i = 1, 2, 3.
j
Expressing Curvature Intrinsically 193
The matrix
3
A ≡ ai, j i, j=1 ,
where the ai, j are functions of the space variables, is called the atti-
tude matrix of the frame (or basis) E 1 , E 2 , E 3 . Since A transforms one
orthonormal frame to another, A is an orthogonal matrix. Hence
A−1 = t A.
Dv f (P) = f ◦ φ(t)
,
dt t=0
Definition 2. If
α : M −→ R3
is a vector field on M,
S P (v) = −
v E 3 (P).
194 Appendix on the Structure Equations and Curvature
0 = Dv (E 3 · E 3 )
= (2 v E 3 ) · E 3
P
= −2S P (v) · E 3 (P).
θi E j (P) = δi j .
Expressing Curvature Intrinsically 195
β = β(E 1 , E 2 )θ1 ∧ θ2 .
0 = Dv (E i · E j )
= (
v E i ) · E j + E i · (
v E j )
= ωi, j (v) + ω j,i (v).
Thus
ωi, j = −ω j,i .
In particular,
ωi,i = 0.
196 Appendix on the Structure Equations and Curvature
We call the ωi, j the connection forms for M. We can now express the
shape operator in terms of these connection forms.
S P (v) = −
v E 3
3
= − (
v E 3 · E j )E j
j=1
3
= − ω3, j (v)E j
j=1
since ω3,3 = 0.
ω1,3 ∧ ω2,3 = κ θ1 ∧ θ2 .
S P (E 1 ) = ω1,3 (E 1 )E 1 + ω2,3 (E 1 )E 2
and
S P (E 2 ) = ω1,3 (E 2 )E 1 + ω2,3 (E 2 )E 2
Expressing Curvature Intrinsically 197
κ = det M P
= ω1,3 (E 1 )ω2,3 (E 2 ) − ω1,3 (E 2 )ω2,3 (E 1 )
= (ω1,3 ∧ ω2,3 )(E 1 , E 2 )
= λ.
Therefore
ω1,3 ∧ ω2,3 = λ θ1 ∧ θ2 = κ θ1 ∧ θ2 .
But
Ei = ai,k δk .
k
Therefore
v E i = (Dv ai,k )δk .
k
198 Appendix on the Structure Equations and Curvature
Then
ωi, j (v) ≡
v E i · E j
= (Dv ai,k )δk · a j,k δk
k k
= (Dv ai,k )a j,k
k
= dai,k (v)a j,k .
k
As a result,
ωi, j = a j,k dai,k .
k
Theorem 9. We have
dθi = ωi, j ∧ θ j , (1)
j
dωi, j = ωi,k ∧ ωk, j . (2)
k
hence
dθi = dai, j ∧ d x j .
j
Expressing Curvature Intrinsically 199
implies
dωi, j = − dai,k ∧ da j,k .
k
=− dai,m ∧ da j,m
m
= dωi, j .
In the antepenultimate line we have used the fact that A−1 = t A. The
result now follows.
Corollary. We have
dω1,2 = −κθ1 ∧ θ2 .
The Corollary has been our main goal in this subsection. It gives
an intrinsic way to calculate Gaussian curvature in the classical setting,
hence a way to define Gaussian curvature in more abstract settings. We
now proceed to develop this more abstract point of view.
3. Curvature Calculations on
Planar Domains
Let ⊆ C be a domain which is equipped with a metric ρ. Assume
for simplicity that ρ(z) > 0 at all points of . Define functions
(1, 0) (0, 1)
E1 ≡ and E2 ≡ .
ρ ρ
Curvature Calculations on Planar Domains 201
Then
θ1 = ρ d x and θ2 = ρ dy
are the dual covector fields. We define ωi, j according to the first struc-
ture equation:
dθ1 = ω1,2 ∧ θ2 ,
dθ2 = ω2,1 ∧ θ1 .
dω1,2 = −κθ1 ∧ θ2 .
One can check that these definitions are independent of the choice of
frame (or basis) E 1 , E 2 , but this is irrelevant for our purposes.
We conclude this Appendix by proving that the definition of cur-
vature which we just elicited from the structural equations coincides
with the one given in Section 2.1. First, by the way that we’ve defined
θ1 and θ2 , we have
dθ1 = dρ ∧ d x
= (ρx d x + ρ y dy) ∧ d x
= ρ y dy ∧ d x
ρy
= − d x ∧ ρ dy
ρ
ρy
= − d x ∧ θ2 .
ρ
Similarly,
dθ2 = dρ ∧ dy
= (ρx d x + ρ y dy) ∧ dy
= ρx d x ∧ dy
202 Appendix on the Structure Equations and Curvature
ρx
=− dy ∧ ρ d x
ρ
ρx
= − dy ∧ θ1 .
ρ
ρy
ω1,2 = − d x + τ dy
ρ
and
ρx
ω1,2 = −ω2,1 = − − dy + σ d x,
ρ
ρy ρx
ω1,2 = − dx + dy.
ρ ρ
Thus
∂ ρy ∂ ρx
dω1,2 = − dy ∧ d x + d x ∧ dy
∂y ρ ∂x ρ
ρ yy ρy ρy ρx x ρx ρx
= − + 2 dy ∧ d x + − 2 d x ∧ dy
ρ ρ ρ ρ
1
= 2 ρρ − (ρ y )2 − (ρx )2 d x ∧ dy
ρ
1
= 4 ρρ − (ρ y )2 − (ρx )2 θ1 ∧ θ2 .
ρ
1
κ=− (ρρ − |
ρ|2 ).
ρ4
Curvature Calculations on Planar Domains 203
Table of Symbols
Page
Symbol Number Meaning
Page
Symbol Number Meaning
Page
Symbol Number Meaning
Page
Symbol Number Meaning
129 a homotopy
R 14 the real numbers
r (P) 111 radius of curvature
ρ 31 a metric (weight)
ρ 70 the Poincaré metric
ρ(z) 106 defining function
ραA (z) 72 dilated, scaled Poincaré distance
metric
ρC 103, 104 Carathéodory metric
ρ
E 103, 104 Euclidean metric
ρK 103, 104 Kobayashi metric
ρ 152 Bergman metric
ρr (z) 70 the dilated Poincaré metric
ρτ 14 a rotation
ρ(z)|dz| 34 a conformal metric
S P (v) 193 shape operator or
Weingarten map
σ 122 the reflection map
σ (z) 69 the spherical metric
σ0 20 extremal function
a (z − P) j 2 power series expansion
j j
a jk (z 1 − P1 ) j (z 2 − P2 )k 166 power series expansion
T 109 tubular coordinate mapping
T P (M) 192 the tangent space to M at P
τ 84 induced Euclidean metric on sphere
τp 120 inward normal segment at p
θi 194 covector fields
U 2, 8 a domain
(U, D) P 93, 180 holomorphic functions f from
D to U such that f (0) = P
U0 81 the slit plane
Uw 164 slice of a domain in C2
Uz 165 slice of a domain in C2
W 108 tubular neighborhood
ξ
ρ,z 31 metric length of ξ
|ξ | 31 Euclidean length of ξ
References
209
210
213
214 Index
zeros
of a function given by a power
series, 8
of holomorphic functions, 31