Download as pdf or txt
Download as pdf or txt
You are on page 1of 219

Riemann surfaces

Pablo Ares Gastesi


School of Mathematics, Tata Institute of Fundamental Research,
Bombay 400 005, India
pablo@math.tifr.res.in
ii
Acknowledgements
I would like rst of all to thank V. Srinivas for his continuous encouragement to
write this book.
I would also like to thank the people who attended my course, on which these notes
are based: Pralay Chatterjee, Preeti Raman and Vijaylaxmi Trivedi.
Finally I would like to thank R.R. Simha for many comments, all of them very
helpful.
Contents
Acknowledgements iv
Chapter 1. Riemann surfaces 1
1.1. Background 2
1.2. Topology of Compact Orientable Surfaces 12
1.3. Riemann Surfaces and Holomorphic Mappings 17
1.4. Dierential Forms 28
1.5. Sheaf Cohomology 42
Chapter 2. Compact Riemann surfaces 57
2.1. Divisors 58
2.2. Dolbeaults Lemma and Finiteness Results 62
2.3. The Riemann-Roch Theorem 68
2.4. Line bundles and Divisors 71
2.5. Serre Duality 81
2.6. Applications of the Riemann-Roch Theorem 95
2.7. Projective embeddings 99
2.8. Weierstrass Points and Hyperelliptic Surfaces 106
2.9. Jacobian Varieties of Riemann Surfaces 115
Chapter 3. Uniformization of Riemann surfaces 127
3.1. The Dirichlet Problem on Riemann surfaces 128
3.2. Uniformization of simply connected Riemann surfaces 141
3.3. Uniformization of Riemann surfaces and Kleinian groups 148
3.4. Hyperbolic Geometry, Fuchsian Groups and Hurwitzs Theorem 162
3.5. Moduli spaces 178
Exercises 187
vvi CONTENTS
Bibliography 201
Notation 203
Index 206
List of Figures 213
CHAPTER 1
Riemann surfaces
1.1 Background 2
1.2 Topology of Compact Surfaces 12
1.3 Riemann Surfaces and Holomorphic Mappings 17
1.4 Dierential Forms 28
1.5 Sheaf Cohomology 42
12 1. RIEMANN SURFACES
1.1. Background
This section is divided into three parts. In the rst one (1.1.1 to 1.1.13) we
review some results of Complex Analysis that we need in this book. Some of those
results will be extended later to the setting of Riemann surfaces. We also introduce
a few results from the theory of normal families of holomorphic functions, a topic
that might not be very familiar to some readers.
The second part (1.1.14 to 1.1.23) contains basic concepts and results of Topol-
ogy: fundamental groups, covering spaces and partitions of unity.
In the last part (1.1.24 and 1.1.25) we state a theorem of Schwartz on operators
on Banach spaces that we need in section 2.1.
References for the results in this section are given in 1.1.26.
Complex Analysis
1.1.1. Definition. Let be an open subset of the complex plane. A complex
valued function f : C is holomorphic at a point z
0
if the complex
derivative
f

(z
0
) = lim
zz
0
f(z) f(z
0
)
z z
0
exists. We say that f is holomorphic in if it is holomorphic at every point of .
1.1.2. For a complex valued function f with partial derivatives, f/x and
f/y, we dene the complex derivatives f
z
and f
z
by the expressions (see also 1.4.1),
f
z
=
f
z
=
1
2
_
f
x
i
f
y
_
, f
z
=
f
z
=
1
2
_
f
x
+ i
f
y
_
.
Informally speaking, the next result says that a function is holomorphic if it depends
on z but not on z.
Proposition. Let f : C be a function with continuous partial derivatives
at all points of . Then f is holomorphic if and only if f
z
= 0 (in ). If f is given
by f(z) = u(z) + iv(z), where u and v are real valued functions dened on , then1.1. BACKGROUND 3
f is holomorphic if and only if it satises the Cauchy-Riemann equations:
u
x
=
v
y
,
u
y
=
v
x
.
The functions u and v above are called the real and imaginary parts of f. Sometimes
we will use the notation Re(f) and Im(f) for u and v respectively.
1.1.3. We record for later use (2.2.4) how the operators

z
and

z
behave
under composition [18, pg. 31]. Assume f : C and g :

C are holomorphic
functions with f()

. Then
(g f)
z
= (g
z
f) (f
z
) + (g
z
f) (f
z
)
(g f)
z
= (w
z
f) (f
z
) + (g
z
f) (f
z
).
1.1.4. Proposition. A function f : C is holomorphic if and only if
for every point z
0
, there exists a neighbourhood U of z
0
, such that f can
be written as a convergent power series f(z) =

n=0
a
n
(z z
0
)
n
in U. If f is
holomorphic, then f

(z) =

n=1
a
n
n(z z
0
)
n1
.
Corollary. Holomorphic functions are smooth (of class C

) that is, they have


derivatives of all orders.
We actually have that holomorphic functions are of class C

; that is, they are


representable by power series (there exist functions with derivatives of all order
which are not given by power series).
The following result simplies the local expression of holomorphic functions.
1.1.5. Proposition. If f : C is a non-constant holomorphic function,
then for every point z
0
, there exists a positive integer n, and a neighbourhood
U of z
0
in , such that f(z) = (z z
0
)
n
g(z), where g : U C is a holomorphic
function satisfying g(z
0
) = 0.
The integer n is called the order of the zero of f(z) f(z
0
) at z
0
.4 1. RIEMANN SURFACES
1.1.6. Theorem (Identity Principle). Let f, g : C be two holomorphic
functions dened on a connected, open subset of C. Assume that there exists a
sequence {z
n
}
n
of points of , with z
n

n
z
0
, where z
0
, such that f(z
n
) = g(z
n
),
for all n. Then f(z) = g(z) for all z .
Remark. There are two important observations to be made about the Identity
Principle: rst of all, the set is assumed to be connected; otherwise, the result
will hold only in the connected component of containing the sequence {z
n
}
n
and
the limit point z
0
. Secondly, the limit point z
0
must belong to the set ; there exist
examples of distinct holomorphic functions that take the same values at each point
of a convergent sequence of points of an open set, where the limit point of such
sequence does not belong to the open set.
Recall that a subset A of a topological space X is called discrete if A does not have
accumulation points in X (A

= ).
Corollary. Let f : C be a non-constant holomorphic function. Then for
any complex number a, the set {z ; f(z) = a} is discrete in (it does not have
limit points in ).
1.1.7. Denote by D the unit disc, D = {z C; |z| < 1}.
Lemma (Schwarz). Let f : D C be a holomorphic function satisfying f(0) = 0
and |f(z)| 1. Then
(1) |f(z)| |z|;
(2) |f

(0)| 1.
Moreover, if equality holds in (1) for a point z = 0, or in (2), then f(z) = z, for a
complex number satisfying || = 1.
1.1.8. A point z
0
C is called an isolated singularity for a function f if f
is dened and holomorphic on U\{z
0
}, where U is a neighbourhood of z
0
. Isolated
singularities are of three types:
(1) removable if f can be extended to a holomorphic function in U;
(2) pole if there exists a positive integer n, such that (z z
0
)
n
f(z) can be
extended to a holomorphic function in U;
(3) essential singularity if none of the above conditions are satised.1.1. BACKGROUND 5
The following result describes the conditions under which a singularity is removable.
Theorem (Removable Singularity Theorem). An isolated singularity z
0
of the
function f is removable if and only if there exists a neighbourhood V of z
0
, such that
f is bounded in V \{z
0
}. Equivalently
lim
zz
0
(z z
0
)f(z) = 0
in V \{z
0
}.
Suppose z
0
is a pole of f and let n be the smallest (positive) integer such that
(z z
0
)
n
f(z) = g(z) is holomorphic on U. The function g(z) has a power series
expansion in U; dividing by (z z
0
)
n
we get that f has the following series expression
(called Laurent series), valid in U\{z
0
},
f(z) =
a
n
(z z
0
)
n
+ +
a
1
z z
0
+

j=0
a
j
(z z
0
)
j
,
with a
n
= 0. The point z
0
is called a pole of order n; the coecient a
1
is called
the residue of f at z
0
.
Definition. A function f is said to be meromorphic in a domain if there
exists a discrete subset A such that f is holomorphic in \A, and has a pole
at each point of A.
1.1.9. The following two results are classical and important properties of holo-
morphic functions that we will use frequently. In these results, we let denote a
connected open subset of C.
Theorem (Maximum Modulus Theorem). Let f : C be a holomorphic
function. Assume that there exists a point z
0
such that |f(z)| |f(z
0
)| for all
z . Then f is constant.
An easy consequence of this theorem is that if f is holomorphic on and K is a
compact subset of , then the maximum of |f| on K is given by the value of f at a
point in the boundary of K.
Theorem (Open Mapping Theorem). Let f : C be a non-constant holo-
morphic function. If U is an open set, then f(U) is an open set (in C).6 1. RIEMANN SURFACES
1.1.10. The next theorem tells us that the inverse of a bijective, holomorphic
mapping is also holomorphic. Observe that this result is not true in Real Analysis;
for example, the function f : R R given by f(x) = x
3
is bijective and innitely
dierentiable, but the inverse function does not have derivative at the point x = 0.
Theorem. Let f : C be a one-to-one holomorphic function. Then f

(z) =
0 for all z and f
1
: f() C is holomorphic.
1.1.11. The following local description of holomorphic functions (see [22, The-
orem 10.32, pg. 216]) is an easy consequence of 1.1.5. Its generalisation to Riemann
surfaces (1.3.10) will be a very useful tool.
Proposition. Let f : C be holomorphic. Let z
0
, and n the order of the
zero of f(z)f(z
0
) at z
0
. Then there exists a neighbourhood U of z
0
in , a positive
number r, and a holomorphic function g : U C, such that f(z) f(z
0
) = g(z)
n
in U. The function g satises |g(z)| < r and g

(z) = 0, for all z U.


1.1.12. A sequence of holomorphic functions {f
n
}
n
, dened on a domain (or
an open set) of C, is said to converge uniformly on compact subsets of to a
function f, if for every compact set K , the sequence {f
n
}
n
converges uniformly
to f in K; that is, for every compact subset K of and every > 0, there exists an
n
0
such that |f
n
(z) f(z)| < , for all n n
0
and all z K.
Theorem (Weierstrass). Let {f
n
}
n
be a sequence of holomorphic functions de-
ned on a domain of C. If {f
n
}
n
converges uniformly on compact subsets of
to a function f, then f is holomorphic. Moreover, the sequence of derivatives {f

n
}
n
converges uniformly on compact subsets of to f

.
1.1.13. Definition. A family A of holomorphic functions dened on a do-
main of C is called normal if for every sequence {f
n
}
n
of elements of A, there
exists a subsequence {f
n
j
}
j
that converges uniformly on compact subsets of .
Observe that in the above denition we do not require that the limit functions (of
subsequences of elements of A) belong the family A. For example, if = D and
A = {f
n
(z) = z/n; n Z}, then A is normal, but the limit function, f 0, is not1.1. BACKGROUND 7
in A.
Theorem (Montel). A family of holomorphic functions on a domain is nor-
mal if and only if it is uniformly bounded on compact subsets. That is, for every
K , compact, there exists a positive number M, such that |f(z)| M, for all
z K and all f A.
Since holomorphic functions are continuous, in the hypothesis of Montels theo-
rem we have that for each function f A there exists a constant M
f,K
such that
|f(z)| M
f,K
, for every z K. This constant depends on f and K. The important
point about Montels theorem is that the family A is normal if and only if we can
take a constant M for all functions on A.
Topology
1.1.14. Definition. Let n be a positive integer. A manifold M of dimension
n is a Hausdor topological space where every point has a neighbourhood homeo-
morphic to an open ball in R
n
. A manifold of dimension 2 is called a surface.
Some books require manifolds to be second countable (have a countable basis of
open sets); that is, there exists a countable collection of open sets, say A = {U
j
}

j=1
,
such that any open subset V of M can be written as a union of elements of A.
Riemann surfaces always satisfy this condition; see remark 4 in 1.3.3.
1.1.15. For the rest of this section all topological spaces (manifolds and sur-
faces in particular) are assumed to be connected, unless otherwise stated.
Definition. A surjective continuous map between two (connected) topological
spaces p : X Y is called a covering map if for every point q Y there is a
neighbourhood V Y , such that p
1
(V ) consists of a disjoint collection of open
subsets U
j
X, p
1
(V ) =

jJ
U
j
, and the restriction of p to each of these sets,
p|
U
j
: U
j
V , is a homeomorphism. The set V is called an evenly covered neigh-
bourhood of q.
By an abuse of notation we will say that X is a covering space of Y (although the8 1. RIEMANN SURFACES
map p is part of the denition). One can easily check that X is a manifold if and
only if Y is a manifold (exercise 7).
1.1.16. A path (or curve) on a topological space X is a continuous map
c : [a, b] X, from an compact interval of R to X. Although a path is actually a
mapping, sometimes in an abuse of notation we will identify a path with its image
{c(t); a t b}. We will usually consider paths dened on the interval [0, 1].
A space is said to be path connected if for any two points x
0
and x
1
, there exists
a path c satisfying c(0) = x
0
and c(1) = x
1
. We say that X is locally path
connected if every point has a path connected neighbourhood. It is easy to check
that a connected, locally path connected space is path connected (exercise 6). In
particular, connected manifolds are path connected, since open balls in Euclidean
space are path connected (the segment between two points in a ball is contained in
the ball).
The next proposition tells us that a path on a space can be lifted to a covering space;
we will see a more general result in 1.1.21.
Proposition. Let p : X Y be a covering map, and c : [0, 1] Y a path.
Then there exists a path c : [0, 1] X, such that p c = c. Moreover, if c(0) = y
0
,
and x
0
is a point in X with p(x
0
) = y
0
, then c is uniquely determined by requiring
that c(0) = x
0
.
1.1.17. For the rest of this section we will assume that all topological spaces
under consideration are path connected.
Definition. Let c
1
, c
2
: [0, 1] X be two paths on a topological space X with
the same initial and end points, c
1
(0) = c
2
(0) = p
0
and c
1
(1) = c
2
(1) = p
1
. We say
that c
1
and c
2
are homotopic if there exists a continuous map H : [0, 1][0, 1] X,
satisfying the following properties:
(1) H(t, 0) = c
1
(t), for all t [0, 1];
(2) H(t, 1) = c
2
(t), for all t [0, 1];
(3) H(0, s) = p
0
, for all s [0, 1].
(4) H(1, s) = p
1
, for all s [0, 1].1.1. BACKGROUND 9
We will write c
1
c
2
to denote that these two paths are homotopic. See gure 1.
c
1
cs
c
2
p
0
p
1
Figure 1. Homotopy between two paths.
1.1.18. A loop (based) at a point x
0
X is a path c with c(0) = c(1) = x
0
.
Let (X, x
0
) denote the set of loops at x
0
in X. It is not dicult to see that is
an equivalence relation in (X, x
0
).
Given two loops based at x
0
, say c
j
: [0, 1] X, j = 1, 2, we dene the composition
c
1
c
2
of c
1
and c
2
as the path given by the following equation:
(c
1
c
2
)(t) =
_

_
c
1
(2t), 0 t
1
2
,
c
2
(2t 1),
1
2
t 1.
Informally speaking, c
1
c
2
is the path that rst runs c
1
and then c
2
. It is not
dicult to see that the composition is preserved under homotopy; more precisely, if
c
1
c

1
and c
2
c

2
, then (c
1
c
2
) (c

1
c

2
). It turns out that the quotient space
(X, x
0
)/ with the composition is a group, called the fundamental group of X
at x
0
, written as
1
(X, x
0
). The identity is given by the class of the constant path
c
x
0
(t) = x
0
. For c
1
as above, let c
1
1
denote the path given by c
1
1
(t) = c
1
(1 t);
then we have that c
1
c
1
1
is homotopic to the constant path c
x
0
; that is, the class
of c
1
1
is the inverse of the class of c
1
(see [16, chapter 8] for details).
1.1.19. Given two points x
0
and x
1
in a path connected space, the fundamental
groups
1
(X, x
0
) and
1
(X, x
1
) are isomorphic; thus we will speak of the fundamental
group of a space, when reference to the base point is not important.
A space is said to be simply connected if its fundamental group is trivial. A
manifold has a simply connected covering, called the universal covering space [16,
pg. 397, Corollary 14.5.b]. Any two universal covering spaces are homeomorphic.
Let f : X Y be a continuous map, x
0
X and y
0
= f(x
0
). If c is a loop
based at x
0
, the composition f c is a loop based at y
0
. Moreover, if c
1
c
2
, then10 1. RIEMANN SURFACES
(f c
1
) (f c
2
). The mapping f

:
1
(X, x
0
)
1
(Y, y
0
), given by f

([c]) = [f c]
is a group homomorphism (we use square brackets for the classes of the loops in
the corresponding fundamental groups). If p : X Y is a covering map, the
homomorphism p

is injective ( exercise 12). Moreover, any subgroup of


1
(Y, y
0
)
is realised as the image of the fundamental group
1
(X, x
0
) of some covering space
X (see [13] for more details on covering spaces).
1.1.20. Definition. Let p : X Y be a covering map. A covering or deck
transformation is a homeomorphism f : X X, such that p f = p. The set of all
deck transformations will be denoted by Deck(X/Y ).
In the case of the universal covering (X simply connected), the groups Deck(X/Y )
and
1
(Y, y
0
) are isomorphic (for any point y
0
of Y ). Moreover, we can recover
Y from its universal covering X and the group Deck(X/Y ): we have that Y is
isomorphic to the quotient X/Deck(X/Y ) [9, pg. 25, exercise 5.1].
1.1.21. The following result generalises the lifting property of paths (1.1.16).
Proposition. Let p : X Y be the universal covering of Y . Let Z be a simply
connected topological space, and f : Z Y a continuous map. Then there exists a
continuous map

f : Z X, such that p

f = f.
1.1.22. Let p : X Y be a covering map between two surfaces. Choose
a point y
0
in Y , and an evenly covered neighbourhood V of y
0
. Then we have
that p
1
(V ) =
j
U
j
, where p : U
j
V is a homeomorphism (for all sets U
j
).
Fix j; shrinking V (and U
j
) if necessary, we can nd homeomorphisms : V
D, : U
j
D, where D denotes the unit disc in the complex plane, such that
( p )(z) = z. This leads us naturally to the concept of branched covering:
we simply require that (p(z)) = z
n
, for some positive integer n. This denition
is equivalent to require that every point x
0
X has a neighbourhood U, such that
p : U\{x
0
} V \{p(x
0
)} is a covering space (see [2, I.20B, pg. 39]). Branched
coverings share many of the properties of standard coverings. For example, if
p : X Y is a branched covering and c : [0, 1] Y is a path, then there exists
a lift of c to X; that is, a path c : [0, 1] X such that p c = c. To prove this1.1. BACKGROUND 11
observe that in the case of X and Y being the unit disc and p(z) = z
n
the statement
is clearly true; the general case can be reduced locally to this particular case.
1.1.23. Let U = {U
j
; j J} be a collection of open sets in a manifold X.
A partition of unity subordinate to U is a collection of continuous functions,

j
: X [0, +), j J, such that:
(1) supp(
j
) = {x X;
j
(x) = 0} U
j
;
(2) for every point x X, there exists a neighbourhood V such that V
supp(
j
) = for only nitely many j;
(3)

jJ

j
(x) = 1.
Observe that by (2) the expressions in (3) are actually nite sums.
By the theorem below we have that partitions of unity exist on connected manifolds.
That situation is enough for our applications; the reader interested on more general
results can nd them in a book on Topology (see the references at the end of this
section).
Theorem. Let M be a second countable, connected manifold, and U = {U
j
}
jJ
an open covering of M. Then there exists a partition of unity subordinate to U.
A Theorem of Functional Analysis
1.1.24. Let L be a vector space over the complex numbers, not necessarily of
nite dimension. A norm || || on L is a mapping || || : L [0, +), satisfying
the following conditions:
(1) ||v|| 0, for all v in L, and ||v|| = 0 if and only if v = 0;
(2) for v and w in L, ||v + w|| ||v|| +||w|| (triangle inequality);
(3) for v L and C, || v|| = || ||v||.
A norm induces a topology on L in a natural way; the basis of this topology is given
by the open balls
B(v, r) = {w L; ||v w|| < r},12 1. RIEMANN SURFACES
where v L and r is a positive number (the open sets in this topologies are the
unions of open balls). A Banach space is a complete (every Cauchy sequence
converges) normed vector space.
1.1.25. A linear mapping T : L M between Banach spaces is continuous if
and only if there exists a non-negative constant C, such that ||T(v)||
M
C ||V ||
L
(here || ||
M
and || ||
L
denote the norms in M and L respectively). We say that a
linear mapping is compact if there exists a neighbourhood U of 0 in L, such that
T(U) is relatively compact in M (the closure of T(U) is compact). In particular,
if T is compact and {v
n
}
n
is a sequence in L, with v
n
0, then there exists a
subsequence {v
n
j
}
j
, such that T(v
n
j
) converges to 0 in M.
The following result will be used in 2.2.10.
Theorem (L. Schwartz). Let T, S : L M be two continuous linear mappings
between Banach spaces. Assume that T is surjective and S compact. Then the space
(T S)(L) has nite codimension in M.
Recall that the codimension of a subspace N in M is the dimension of the quotient
space M/N.
1.1.26. Some references for Complex Analysis are [1], [5] and [20]. For the
Topology part see [13] and [16] (Indian version of [17]). Schwartzs theorem can be
found in [21] or [6].
1.2. Topology of Compact Orientable Surfaces
In this section we recall some facts of the Topology of compact orientable sur-
faces: genus, construction of homology groups and Euler-Poincare characteristic.
Throughout this section X will denote a connected, compact surface.
A good reference for the proofs of the results of this section is [13].
1.2.1. From the point of view of Topology compact orientable surfaces are
classied by a non-negative integer called the genus (genera in plural). Two surfaces
X and Y , of genera g and g

respectively, are homeomorphic if and only if g = g

.1.2. TOPOLOGY OF COMPACT ORIENTABLE SURFACES 13


A surface of genus 0 is homeomorphic to the 2-sphere,
S
2
= {(t
1
, t
2
, t
3
) R
3
; (t
1
)
2
+ (t
2
)
2
+ (t
3
)
2
= 1};
this surface is simply connected. If g > 0, X is homeomorphic to a sphere with g
handles attached (gure 2 for an example of a surface of genus 2). Its fundamental
group has a presentation given by 2g generators, a
1
, . . . , a
g
, b
1
, . . . , b
g
, and one single
relation

g
j=1
[a
j
, b
j
] = 1, where [a
j
, b
j
] is the commutator of a
j
and b
j
; that is
[a
j
, b
j
] = a
j
b
j
a
1
j
b
1
j
. If X has genus 1, its fundamental group is abelian; however,
if g 2 the group
1
(X, x
0
) is not abelian, we will prove this fact in 3.3.15.
a
1
b
1
a
2
b
2
Figure 2. A surface of genus 2 with generators of the fundamental group.
1.2.2. In the next paragraphs we dene the singular homology groups of com-
pact surfaces and state some results that we will need later in the text. We do
not provide proofs for those results; the reader can nd more details in [2] (for the
particular case of Riemann surfaces) or [14] (for the general case). Let denote
the triangle {(x, y) R
2
; 0 y x 1} in R
2
. A 2-simplex on X is a contin-
uous mapping f : X. A 1-simplex is a continuous mapping e : [0, 1] X.
A point, or a mapping from a point in R to X, is called a 0-simplex. The chain
groups C
n
(X), n = 0, 1, 2, consists of nite formal sums of the form

m
j=0
c
j
s
j
, where
c
j
Z, and c
j
are n-simplices (more precisely, C
n
(X) is the free abelian group on
n-simplices in X). For example, an element of C
0
(X) is a nite sum of the type
c
0
p
0
+ + c
n
p
n
, for some points p
j
of X and integers c
j
. The sets C
n
(X) are
abelian groups with formal addition as the group operation.
Like in the case of paths, simplices are mappings; thus two simplices might be dif-
ferent, although their images are the same: for example, if e is a 1-simplex, the
mapping e(t) = e(1 t) : [0, 1] X denes a simplex with the same image but a
dierent element on C
1
(X).14 1. RIEMANN SURFACES
1.2.3. The boundary operators between chain groups, : C
n
(X)
C
n1
(X), n = 1, 2, are dened as follows. In the case of n = 2, for a simplex
f : X, we set f = e
0
e
1
+ e
2
, where e
j
: [0, 1] X are the simplices given
by e
0
(t) = f(t, 0), e
1
(t) = f(1, 1 t) and e
2
(t) = f(1 t, 1 t). Thus the boundary
of f consists of the restriction of the mapping f to the boundary of in certain
order. For a 1-simplex e : [0, 1] X, we dene (e) = e(1) e(0). We extend to
C
2
(X) and C
1
(X) in the natural way to make it a group homomorphism. It is not
dicult to see that ( )(c) = 0 for c C
2
(X); one needs to check this equality
only on 2-simplices, say f : X,
((f)) = (e
1
e
2
+ e
3
) = e
1
(1) e
1
(0) e
2
(1) +e
2
(0) +e
3
(1) e
3
(0) =
= f(1, 0) f(0, 0) f(1, 0) +f(1, 1) +f(0, 0) f(1, 1) = 0.
1.2.4. Let Z
1
(X) be the kernel of : C
1
(X) C
0
(X), and B
1
(X) the image of
C
2
(X) in C
1
(X). The elements of these groups are called cycles and boundaries,
respectively. Since = 0, B
1
(X) is a (normal) subgroup of Z
1
(X); the quotient
group
H
1
(X, Z) = Z
1
(X)/B
1
(X)
is called the 1st homology group of X (with integer coecients). Observe that any
loop determines a 1-simplex in X without boundary, hence an element of H
1
(X, Z).
Thus we obtain a mapping (X, x
0
) H
1
(X, Z). We have that this mapping can be
used to compute the 1st homology group, as stated in Hurwitzs theorem: H
1
(X, Z)
is isomorphic to the abelianisation of the
1
(X, x
0
),
H
1
(X, Z) =
1
(X, x
0
)/[
1
(X, x
0
),
1
(X, x
0
)].
Recall that if G is a group, [G, G] denotes the subgroup generated by all commutators
of elements of G; that is, [G, G] is generated by all elements of the form aba
1
b
1
for a and b in G. It follows from Hurwitzs theorem that for a compact surface X
of genus g the group H
1
(X, Z) is a free abelian group of rank 2g:
H
1
(X, Z)

= Z
2g
= Z
2g factors
Z.1.2. TOPOLOGY OF COMPACT ORIENTABLE SURFACES 15
The homology classes of the paths in gure 2 are generators for this group. Observe
that in the case of genus 1 the fundamental group and the 1st homology group
coincide (observe that
1
(X, x
0
) is an abelian group).
1.2.5. The 2nd homology group is dened by
H
2
(X, Z) = ker ( : C
2
(X) C
1
(X)) .
For compact orientable surfaces we have H
2
(X, Z)

= Z.
The 0th homology group, H
0
(X, Z), is the image of C
1
(X) in C
0
(X) under
. This group is easy to compute: x a point x
0
X; since X is path connected,
for any point x
1
of X there exists a path c : [0, 1] X satisfying c(0) = x
0
and
c(1) = x
1
. Therefore (c) = x
1
x
0
. This means that the classes of these two points
in H
0
(X, Z) are equal; it is then easy to show that H
0
(X, Z)

= Z. In general, for
a non-connected surface Y the group H
0
(Y, Z) is a free abelian group of rank equal
to the number of connected components of Y .
1.2.6. As we have seen above the homology groups H
m
(X, Z), for m = 0, 1, 2, of
a compact surface X are free abelian groups. The rank of the nth group is called the
nth Betti number of X, denoted by b
n
(X). The Euler-Poincare characteristic
of X is dened as
(X) = b
2
(X) b
1
(X) + b
0
(X).
By the above computations we have that (X) = 2 2g.
1.2.7. A triangle T on X is a set homeomorphic to the standard triangle
dened above. The edges of T are the images of the edges of (under the given
homeomorphism). Similarly the images of the vertices of (the points (0, 0), (1, 0)
and (1, 1) in R
2
) are called the vertices of T. Observe that in this denition a trian-
gle T is a subset of X, and thus the homeomorphism between T and does not
play an important role as in the case of simplices or paths.
Definition. A triangulation of a compact surface X is a nite collection of
triangles, T = {T
1
, . . . , T
n
}, satisfying the following properties:
(1)

n
j=1
T
j
= X;
(2) if the intersection of two triangles is not empty, then it is equal to either a16 1. RIEMANN SURFACES
edge or a vertex of both triangles (see gure 3);
(3) for any vertex v, the union of all triangles containing v contains a neighbour-
hood of v.
Allowed in a triangulation Not allowed in a triangulation
Figure 3. Allowed and wrong triangulations.
T1
T2
v
Tn
Figure 4. Condition (3) in denition 1.2.7.
1.2.8. The following result gives us an alternative way of computing the Euler-
Poincare characteristic of X.
Proposition. Let T be a triangulation of a compact surface X with F triangles,
E edges and V vertices. Then
(X) = F E + V.
For an example, consider the triangulation of the 2-sphere given in gure 5. We
have F = 8, E = 12 and V = 6, so (S
2
) = 2.1.3. RIEMANN SURFACES AND HOLOMORPHIC MAPPINGS 17
v
5
v
4
v
6
v
1
v
2
v
3
Figure 5. A triangulation of the sphere.
1.3. Riemann Surfaces and Holomorphic
Mappings
After reviewing the necessary background in the last two sections we are ready to
introduce the concept of Riemann surface. In an informal way, a Riemann surface is
the most general topological space on which one can do Complex Analysis similar to
the analysis done in the complex plane [2]. In this section we give the formal deni-
tion of Riemann surface and some examples. We also dene holomorphic mappings
between Riemann surfaces and study some of their properties. In particular we are
interested on nding which properties of holomorphic mappings (or holomorphic
properties in general) are satised by mappings between surfaces.
Riemann surfaces
1.3.1. Informally speaking one can dene a Riemann surface as a collection
of open subsets of the complex plane glued by holomorphic functions. More pre-
cisely, let X be a connected surface; thus every point of X has a neighbourhood U
homeomorphic to an open set of R
2
, which we identify with C. One could dene a
Riemann surface then as a (topological) surface with a covering by open sets, homeo-
morphic to open sets in C, such that if U and V are two elements of the covering with
non-empty intersection, and : U (U) C and : V (V ) C the cor-
responding homeomorphisms, the mapping
1
is holomorphic. Although from
a working point of view this denition suces (see remark after denition 1.3.2),18 1. RIEMANN SURFACES
there are some technical problems with it. For example, the complex plane with the
identity function z z is a Riemann surface. If we now consider C with two open
sets, say U = C and V = C\{0}, and homeomorphisms given also by the identity on
U and V , we obtain a new Riemann surface; is this surface the same as the previous
one? Clearly, from a practical view point, both structures should give us the same
results; however, according to the above (informal) denition, these two surfaces
are dierent since the coverings are not equal. To avoid this type of problems one
denes Riemann surfaces via equivalence classes of coverings satisfying the above
conditions (complex atlases, to be more formal).
Definition. A complex atlas U on a connected surface X is a collection of
pairs U = {(U
j
, z
j
)}
jJ
, where:
(1) U
j
are open sets of X;
(2)

jJ
U
j
= X;
(3) z
j
: U
j
W
j
are homeomorphisms onto open subsets W
j
of the complex
plane;
(4) the functions z
j
satisfy the following compatibility condition: if U
j
U
k
= ,
then
z
k
z
1
j
: z
j
(U
j
U
k
) z
k
(U
j
U
k
)
is holomorphic.
Observe that the sets z
j
(U
j
U
k
) in (4) are open subsets of C, so one can talk
of z
k
z
1
j
being a holomorphic function.
The functions z
j
are called local coordinates and the pairs (U
j
, z
j
) coordinate
patches. The transition functions z
k
z
1
j
are known as changes of coordinates.
Definition. Two complex atlases U and V are equivalent or compatible if
their union is a complex atlas. A complex structure on X is an equivalence class
of complex atlases.
1.3.2. Definition. A Riemann surface is a connected surface with a com-
plex structure.1.3. RIEMANN SURFACES AND HOLOMORPHIC MAPPINGS 19
1.3.3. Remarks. 1. Let U be a complex atlas on X and let W denote the atlas
consisting on all local coordinates on X that are compatible with the coordinates of
U. Clearly W contains U, in the sense that any coordinate patch of U belongs to
W. Moreover, W is maximal in the sense that if V is another atlas on X containing
U, then V is contained in W. Clearly U and W are equivalent atlases, so they dene
the same complex structure on X. It follows from this observation that to dene a
Riemann surface structure on X we just need to give one complex atlas U and then
consider the unique maximal atlas (W in our notation) in the equivalence class of
U.
2. Observe that if U is an atlas on a surface X and (U, z) is a local coordinate of
U, then for any open set V with V U, we have that (V, z|
V
) is a local coordinate
on X, compatible with U. In other words, restrictions of local coordinates are local
coordinates; we will use this fact frequently.
3. The denition of complex structure generalises to higher dimensions in a natural
way; see 2.4.
4. By a theorem of Rado, Riemann surfaces have a countable basis of open sets.
We will not prove this result here; see, for example [8, pgs. 185-189].
Examples of Riemann surfaces
1.3.4. The complex plane C with the identity function z z is an example of
a Riemann surface. The maximal atlas for this structure consists of all pairs (U, f),
where U is any open subset of C and f is a one-to-one holomorphic function dened
on U (this is because f is an open mapping, whose inverse is holomorphic). Any open
subset of C with (the complex structure induced by) the restriction of the identity
function is also a Riemann surface. Two important examples of Riemann surfaces
of this type that will appear frequently are the unit disc D = {z C; |z| < 1} and
the upper half plane H = {z C; Im(z) > 0}.
1.3.5. A more interesting example is given by the one-point compactication of
the complex plane,

C = C{}. Recall that this space is constructed by adding one
point to the complex plane, called the point at innity, . The open sets of

C are20 1. RIEMANN SURFACES
the open subsets of C and the sets of the form {} (C\K), where K is a compact
subset of C (these last sets are the neighbourhoods of ). To dene a Riemann
surface structure on

C we consider two open sets, U
1
= C, and U
2
=

C\{0}. As
local coordinates we take the homeomorphisms z
1
: U
1
C, the identity function
z
1
(z) = z, and
z
2
: U
2
C, z
2
(z) =
_

_
1
z
, if z = ,
0, z = .
It is easy to see that the changes of coordinates are holomorphic functions: the
intersection of the two coordinate patches is given by U
1
U
2
= C\{0}, and the
function (z
2
z
1
1
)(z) = 1/z, is holomorphic in U
1
U
2
, since z = 0. This surface is
known as the Riemann sphere. It is not dicult to prove that

C is homeomorphic
to the 2-sphere (1.2.1); see exercise 15. We will show later that any Riemann surface
structure on

C is equivalent to the one we have just constructed (see 2.3.5, 2.6.8
and 3.2.8).
1.3.6. Our next example uses a standard technique for constructions of Rie-
mann surfaces, namely quotients (in a broad sense any Riemann surface is a
quotient; see the Uniformization theorem, 3.3.1). Let be a complex number
with positive imaginary part and G

the group of translations on C of the form


T

n,m
(z) = z + n + m, where n and m are integers. Denote by T

= C/G

the
quotient space; that is, T

consists of equivalence classes of points of C, where z and


w are equivalent if there exists a transformation T

n,m
of G

, such that T

n,m
(z) = w.
Let : C T

be the natural quotient map. We put the quotient topology on T

:
a set U T

is open if and only if


1
(U) is open in the complex plane. It is easy to
see that T

is a connected surface (exercise 31); moreover, T

is a compact surface.
To give a complex structure to T

, rst observe that there exists a positive number


r, such that d(0, n+m) > r, for all pairs of integers (n, m) = (0, 0). One can take,
for example, r =
1
2
min{1, ||}. If z
0
is a point in C, let U(z
0
) denote the disc centred
at z
0
and radius r. We have that U(z
0
) S(U(z
0
)) = , for any non-identity element
S of G. Thus the restriction of the quotient mapping |
U(z
0
)
: U(z
0
) (U(z
0
))
is a homeomorphism (since it is an injective, open mapping), so its inverse gives1.3. RIEMANN SURFACES AND HOLOMORPHIC MAPPINGS 21
us a local coordinate in a neighbourhood of (z
0
),
_
(U(z
0
)),
1
|
(U(z
0
))
_
. To
check the compatibility condition, let z
0
and z
1
be two points in the complex
plane with (z
0
) = (z
1
). This means that there is an element T

n,m
G

, sat-
isfying T

n,m
(z
0
) = z
1
. We have two local coordinates
_
(U(z
0
)),
1
|
(U(z
0
))
_
, and
_
(U(z
1
)),
1
|
(U(z
1
))
_
, dened in a neighbourhood of (z
0
). It is clear that the
disc U(z
0
) is mapped to the disc U(z
1
) by this transformation T

n,m
. It follows that
the change of coordinates is given by (|
U(z
1
)
)
1
|
U(z
0
)
= T

n,m
, which clearly is a
holomorphic function. The atlas U consisting on all pairs of the above form makes
T

a Riemann surface, called a torus. One can prove that T

is a compact surface
of genus 1. Abels theorem (2.9.11) gives us the converse statement: any compact
Riemann surface of genus 1 is of the form T

.
Any two tori are homeomorphic (1.2.1), but unlike in the case of the Riemann sphere,
there are many dierent Riemann surface structures on a torus (that is, not any two
complex structures on a torus are equivalent). (3.3).
1.3.7. In the above example the elements of G

(other than the identity) do


not have xed points in C, and thus we can use the projection to give a complex
structure on the quotient surface. A similar construction is possible for transfor-
mations with mild xed points, as we show next. Consider the unit disc D as a
Riemann surface with the structure induced by the identity function. Let R : D D
be the rotation of order 4 given by R(z) = iz, and denote by D
4
the quotient space
D/ < R >; that is, two points z and w on the unit disc are identied in D
4
if
z = R
n
(w), for some n = 0, 1, 2, 3 (R
4
is the identity map). It is not dicult to see
that D
4
is a manifold, homeomorphic to a disc; D
4
is obtained by taking the shaded
sector in gure 6 and identifying the sides l and l

; Let : D D
4
be the quotient
l
l

Figure 6. Branched covering.22 1. RIEMANN SURFACES


map; points on D
4
are of the form (z), where z is a point in the unit disc. We can
now get a coordinate on D
4
by the expression
4
((z)) = z
4
. We need to check that
this mapping is well dened: if (z) = (w), then z = i
n
w, where n = 0, 1, 2, 3, and
therefore z
4
= w
4
. Since
4
is globally dened on D
4
the compatibility condition
does not need to be checked. Thus we have that
4
induces a Riemann surface
structure on D
4
.
This examples shows how to give local coordinates to branched coverings (1.1.22;
the number 4 above is not important in the construction and can be substituted by
any positive integer).
Holomorphic mappings
1.3.8. The denition of a holomorphic function on an open set of the complex
plane (1.1.1) can be easily generalised to the context of Riemann surfaces.
Definition. A continuous mapping f : X Y between two Riemann surfaces
is called holomorphic if for every pair of local coordinates, (U, z) on X and (V, w)
on Y , with f(U) V = , the function
w f z
1
: z(U f
1
(V )) w(V )
is holomorphic.
If the target surface is the complex plane, Y = C, we say that f is a holomorphic
function on X.
A meromorphic function is a non-constant holomorphic mapping f : X

C from
a Riemann surface to the Riemann sphere; the points f
1
() are called the poles
of f. We will denote by O(X) and M(X) the sets of holomorphic and meromorphic
functions on the Riemann surface X, respectively.
Since the above denition is local some important properties of holomorphic
functions dened on C are satised by holomorphic mappings on Riemann surfaces.
For example, the Maximum Modulus Principle, the Open Mapping Theorem (1.1.9)1.3. RIEMANN SURFACES AND HOLOMORPHIC MAPPINGS 23
and the Identity Principle (1.1.6) hold in the setting of Riemann surfaces. Sim-
ilarly, a holomorphic, bijective mapping between two Riemann surfaces, called a
biholomorphism, has a holomorphic inverse (1.1.10).
1.3.9. Holomorphic functions of one variable are holomorphic mappings when
we consider open sets of the complex plane as Riemann surfaces (with the identity
as the coordinate function). One such example is given by a polynomial p : C C.
We can extend p to the Riemann sphere by dening p() = . To see that this
extended mapping is holomorphic assume that p is given by p(z) = a
n
z
n
+ +a
0
,
where a
j
are complex numbers, and a
n
satisfying a
n
= 0. It is clear that we need
to prove only that the extended function, also denoted by p, is continuous and
holomorphic at the point . Continuity follows from standard results of topology [4,
theorem 11.4, pg. 33]. To show that p is holomorphic at consider the local
coordinate z
2
dened in 1.3.5. The function g = z
2
p z
1
2
= 1/p(1/z) has the
following expression:
g(z) = (z
2
p z
1
2
)(z) =
z
n
a
n
+ + a
0
z
n
.
Since a
n
= 0 we have that g is holomorphic in a neighbourhood of 0, so p is a
holomorphic mapping on the Riemann sphere. Actually, p is a meromorphic function
whose only pole lies at the point . It also follows from the above expression that
the order of the pole at is n; we will use this fact in 1.3.14.
The mapping : C T

dened in 1.3.6 is holomorphic. This is easy to see


since local coordinates on T

are given precisely by the local inverses of . The


quotient mapping : D D
4
of 1.3.7 is also a holomorphic mapping; its expression
in local coordinates is given by z z
4
.
Degree of a holomorphic mapping
1.3.10. The following result is an direct consequence of 1.1.11.
Proposition. Let f : X Y be a non-constant holomorphic mapping between
two Riemann surfaces. Let p be a point of X, and q = f(p). Then there exist local
coordinates (U, z) and (V, w), on X and Y respectively, such that:24 1. RIEMANN SURFACES
(1) p U, z(p) = 0; q V , w(q) = 0;
(2) f(U) V ;
(3) the function w f z
1
is of the form
n
for some integer n (the integer
n depends on f and p, but it is independent of the choices of local coordinates p).
1.3.11. The integer n in the above proposition is called the ramication
number of f at p. The branching number (of f at p), denoted by b
p
(f), is
dened as b
p
(f) = n 1. Observe that if p

U, with p

= p, then b
p
(f) = 0. Thus
the set of points with positive branching number is discrete; in particular, if X is
a compact surface, it is a nite set. We also have that for any point q

V \{q},
the number of points in f
1
(q

) U is equal to n. If we count multiplicities, as


it is usually done in Complex Analysis, we have that every point in f(U) has n
preimages in U. It follows from this remark that the total number of preimages of
any point q
0
of Y is then given by

(b
p
0
(f) + 1), where the sum is taken over all
points p
0
X with f(p
0
) = q
0
. In the case of compact surfaces this sum is nite,
and independent of the point q
0
, as the next result shows.
Proposition. Let f : X Y be a non-constant holomorphic mapping between
compact Riemann surfaces. Then there exists a positive integer d, called the degree
of f, such that any point q Y has precisely d preimages counted with multiplicity.
In other words, for any q Y ,

f(p)=q
(b
p
(f) + 1) = d.
1.3.12. To prove the above proposition we need the following result.
Lemma. Let f : X Y be a non-constant holomorphic mapping between
two Riemann surfaces. Assume that X is compact. Then Y is compact and f is
surjective.
Proof. Since f is not constant we have that f(X) is open in Y (Open Mapping
Theorem, 1.1.9). On the other hand, since X is compact f(X) is compact in Y ,
and therefore closed. Since Y is connected and f(X) is not empty we have that
f(X) = Y .1.3. RIEMANN SURFACES AND HOLOMORPHIC MAPPINGS 25
1.3.13. Proof of proposition 1.3.11. Consider the decreasing sequence of
subsets of Y given by
Y
n
= {q Y ;

f(p)=q
(b
p
(f) + 1) n},
where n is a positive integer. Clearly Y
n
consists of the points of Y with at least n
preimages, counted with multiplicities. By the previous lemma we have that Y
1
= Y
(so at least one of these sets is not empty).
It follows from proposition 1.3.10 that Y
n
is an open subsets of Y , for all n. We
will show that Y
n
are closed sets as well. Fix n and let {y
k
}
k
be a sequence of points
in Y
n
, with limit point y
0
. Since the number of points of X with positive ramication
number is nite we can assume that f
1
(y
k
) consists of at least n distinct points,
say {x
1
k
, . . . , x
n
k
}. By the compactness of X, for each xed j = 1, . . . , n, there
exists a convergent subsequence of the sequence {x
j
k
}
k
. After renaming subindices
if necessary, we get n points, x
1
, . . . , x
n
, such that x
j
k
converges to x
j
(as k ).
Clearly f(x
j
) = y
0
by the continuity of f. If the points x
j
are distinct we have that
y
0
has at least n preimages; that is, y
0
belongs to Y
n
. Assume on the other hand
that some limit points coincide, say x
1
= x
2
; that is, the sequences {x
1
k
}
k
and {x
2
k
}
k
converge to the same limit point x
1
. Using again 1.3.10 we get that f should be at
least 2to-1 in a neighbourhood of x
1
(see remark after the proof). Thus b
x
1
(f) 1,
and this point counts as (at least) two preimages of y
0
. We obtain in this way that
y
0
has at least n preimages; that is, y
0
Y
n
, and therefore Y
n
is closed.
Since Y is connected, each Y
n
is either empty or equal to Y . Choose any point
y Y , and let d be the number of preimages of y. Then we have that y Y
d
but
y / Y
d+1
. It follows that Y
d
= Y . Since Y
n+1
Y
n
, the sets Y
n
are empty for n > d,
which completes the proof.
Remark. Suppose f : D D satises f(0) = 0. By 1.3.10 we can assume
f(z) = z
n
. If n = 1, then f is the identity, so there cannot exist two distinct
sequences of points in D, say {z
k
}
k
and {w
k
}
k
, converging to 0 and such that f(z
k
) =
f(w
k
). So we have that if such sequences exist then n 2. Similarly, if there are26 1. RIEMANN SURFACES
three sequences satisfying similar conditions then n 3. This should clarify the
proof above.
1.3.14. To give a concrete application of the above results we consider a poly-
nomial of degree n, say p(z) = a
n
z
n
+ + a
0
, where a
n
= 0. We have seen that p
can be extended to a holomorphic mapping on the Riemann sphere, whose only pole
of order n lies at the point . It follows that p has degree n; thus p has n zeroes,
counted with multiplicity: this is the Fundamental Theorem of Algebra. We also
see that the degree of a polynomial (in the sense of Algebra) is equal to the degree
when considered as a holomorphic mapping of the Riemann sphere.
The quotient mapping of 1.3.7 has degree 4. Although the surfaces in this
example are not compact, proposition 1.3.11 applies to a wider class of mappings
(proper mappings); see [8, pg. 29] (exercise 21).
Proposition. Let f :

C

C be a non-constant holomorphic mapping. Then
f is a rational function; i.e. there exist two polynomials p and q, with no common
zeroes, such that f(z) = p(z)/q(z). Moreover, the degree of f is the maximum of
the degrees of p and q.
Proof. We leave it as an exercise (exercise 22).
Corollary. The automorphisms (biholormophic self-mappings) of

C are the
Mobius transformations; that is, the mappings of the form T(z) =
a z + b
c z + d
, where
a, b, c, d are complex numbers satisfying ad bc = 0.
Proof. From the above proposition we have that if T is a biholomorphism
of

C then it must be a rational function of degree 1. The condition ad bc = 0 is
equivalent to the fact that the numerator and denominator of T do not have common
zeroes.
Triangulations and the Riemann-Hurwitz Formula
1.3.15. In the proof of the next proposition we will assume that compact Rie-
mann surfaces have non-constant meromorphic functions (a non trivial fact that will1.3. RIEMANN SURFACES AND HOLOMORPHIC MAPPINGS 27
be proved in 2.3.4).
Proposition. Compact Riemann surfaces are triangulable.
Proof. Let f : X

C be a non-constant meromorphic function and let S
denote the set of critical points of f,
S = {p X; b
p
(f) > 0}.
Since X is compact S is a nite set. Let S

be the set of critical values, S

= f(S).
If T is any triangulation of

C we can subdivide it to obtain another triangulation,
say T

, such that all points of S are vertices of T

(see gure 7), and at most one


vertex of any triangle of T

lies in S

. The triangles of T

lift via f to triangles in


z
0
z
0
Figure 7. Subdivision of triangles (z
0
is a critical value).
X, i.e. if T is a triangle of T

then f
1
(T) is a disjoint union of triangles in X. This
follows from the fact that f : X\f
1
(S) Y \S

is a covering map and the interiors


of the triangles are simply connected. Similarly, the (open) edges lift to edges. Thus
we can lift T

to a triangulation on X.
1.3.16. If f : X Y is a holomorphic mapping between compact surfaces of
degree d, for any point q in Y we have that the number of points in the set f
1
(q)
(counted without multiplicities) is given by d

f(p)=q
b
p
(f).
The total branching number of a holomorphic mapping f : X C, where
X is a compact surface, is dened by B(f) =

pX
b
p
(f). Observe that this sum
takes place over only nitely many points.
Proposition (Riemann-Hurwitz relation). Let f : X Y be a holomorphic
mapping of degree d between compact surfaces of genera g and g

respectively. Then
the following relation holds:28 1. RIEMANN SURFACES
2g 2 = d (2g

2) +B(f).
Proof. As in the previous proof we consider the sets S = {p X; b
p
(f) > 0} of
critical points and S

= f(S) Y of critical values of f. Let T be a triangulation of


Y such that all points of S

are vertices of it, and at most one vertex of any triangle


lies in S

. Let

T denote the triangulation of X obtained by lifting T via f
1
. We
have that any triangle of T lifts to d triangles on X, and similarly with the edges. So
if T has F triangles (faces), E edges and V vertices, then

T will have d F triangles
and d E edges. The vertices that do not belong to S

have d distinct preimages. If


y S

, from the observation before the proposition we see that f


1
(y) consists of
d

f(x)=y
b
x
(f) points. Thus f
1
(S

) consists of dn B(f) points, where n is the


number of points in S

. So

T has dV B(f) vertices. Applying the Euler-Poincare
characteristic formula to the triangulations T and

T we get F E + V = 2 2g

and dF dE + dV B(f) = 2 2g. Putting these two expressions together we


obtain the desired relation.
Corollary. With the hypothesis and notation of the above proposition we have
1. B(f) is an even integer;
2. g g

.
As an application of this corollary we see that there do not exist any non-constant
holomorphic mappings from the Riemann sphere to a surface of positive genus.
1.4. Dierential Forms
The denition of Riemann surface (1.3.2) can be easily modied to obtain other
type of structures on a (topological) manifold. For example, if we require changes
of coordinates to be smooth functions we have the concept of smooth manifolds
[3]. In particular Riemann surfaces are smooth manifolds (1.1.4) and thus we can
apply the results of Dierential Geometry to them. In this section we recall some
denitions and theorems about smooth manifolds and obtain some consequences
for the case of Riemann surfaces. We start with a working denition of forms,
following [7]; although this approach might look quite formal, it is short and good1.4. DIFFERENTIAL FORMS 29
enough to state the theorems we need. Nevertheless, in the third subsection we
explain a dierent point of view, as develop in [8], and its connections with ours as
well as the Dierential Geometry denitions. The complex structure enters in the
second part of this section, where we study harmonic and holomorphic forms. The
fourth part explains integration of forms on a surface; two important results, Stokes
(1.4.22) and Sards (1.4.23) theorems, are stated at the end of that part. We nish
this section with meromorphic forms, a natural generalisation of holomorphic forms;
the Residues theorem, a fundamental tool that we will use in the next chapters, is
obtained as an easy consequence of Stokes theorem.
Some of the results of Dierential Geometry require that manifolds have a count-
able basis of open sets; as remarked in 1.3.2, this is always true for Riemann surfaces.
Dierentiable functions and forms
1.4.1. Recall that by a smooth function we mean a function with derivatives
of all orders (class C

).
Definition. A continuous function f : X C dened on a Riemann surface
X is said to be smooth (or dierentiable) if for any coordinate patch (U, z), the
function (f z
1
) : z(U) C is smooth (as a function dened on an open set of R
2
and with values in R
2
).
We will denote by C(X) and S(X) the sets of continuous and dierentiable functions
on X, respectively. Similarly C(U) and S(U) denote the sets of continuous and
dierentiable functions on an open subset U of X (see also 1.5).
Using local coordinates (s, t) on the R
2
we dene two operators on smooth functions,
the partial derivatives, given by the following expressions:
f
x
(p) =
f
x
(p) =
(f z
1
)
s
(z(p)), and f
y
(p) =
f
y
(p) =
(f z
1
)
t
(z(p)),
where p U. From a Complex Analysis point of view it is more natural to consider
the operators f
z
and f
z
, dened as follows (see also 1.1.2):
f
z
=
f
z
=
1
2
_
f
x
i
f
y
_
, and f
z
=
f
z
=
1
2i
_
f
x
+ i
f
y
_
.30 1. RIEMANN SURFACES
Observe that, by the Cauchy-Riemann equations a function is holomorphic if and
only if f
z
= 0.
1.4.2. The most natural way of dening forms on a surface is by introducing
the tangent plane and then considering its dual. However that would take us into
a course on Dierential Geometry, which is outside the scope of this book. As
mentioned in the introduction to this section, we provide a practical approach,
and describe later in this section the relationship with other points of view.
A (smooth) 1-form on X is an assignment of two smooth complex valued
functions, f and g, to each local coordinate (U, z), such that the following invariance
property is satised. If (V, w = x + i y) is another local coordinate on X, with
U V = , and is given by

f and g on V , then
(1)

f = f
x
x
+ g
y
x
, g = f
x
y
+ g
y
y
,
on U V . We will write = fdx + gdy for a 1-form. The invariance property is
usually expressed by the equality f dx + g dy =

f d x + g d y. The space of 1-forms
on X will be denoted by S
1
(X); similarly we use S
1
(U) for the set of 1-forms on an
open subset U of X.
The sets S
1
(U) are free modules of rank 2 over the ring of C

complex valued
functions, with {dx, dy} being a basis of this module (that any form can be expressed
on U as fdx + gdy, where f and g are smooth functions). We dene two forms on
U by
dz = dx + idy and d z = dx idy.
It is easy to check that the pair {dz, d z} is also a basis of S
1
(U) (as a vector space
over R). Using dz and d z we can decompose S
1
(U) in two subspaces in a natural
way: S
1
(U) = S
(1,0)
(U) S
(0,1)
(U), where S
(1,0)
(U) (respectively S
(0,1)
(U)) is the
subspace generated by dz (respectively d z). Forms in S
(1,0)
(U) are called of type
(1, 0); in the local coordinate z they are of the form f(z) dz, where f is a smooth
function. Similarly the elements of S
(0,1)
(U) are called forms of type (0, 1).
The importance of this decomposition lies in the fact that it does not depend on
local coordinates: let (U, z) and (V, w) be a coordinate patches on X with V = U.1.4. DIFFERENTIAL FORMS 31
Assume that = f dz is an element of S
(1,0)
(U). Let =

fdw + gd w be the
expression of with respect to the coordinate w. We then have

f = f
z
w
, g = f
z
w
;
(this is similar to the expression (1) above; we have g = 0). Since the change of
coordinates in holomorphic, we have from Cauchy-Riemann equations that z/ w =
0, so g = 0. This shows that is also an element of S
(1,0)
(V ).
1.4.3. A (smooth) 2-form on X is an assignment of a smooth function f to
each local coordinate (U, z), satisfying the following invariance property: with the
above notation, if is given by the function

f on V , then

f = f
(x, y)
( x, y)
= f
_
x
x
y
y

x
y
y
x
_
.
We write = fdx dy or = fdxdy for a 2-form; we will denote the space of
2-forms by S
2
(X). We follow the convention dy dx = dxdy (dxdy = dy dx);
see below (1.4.5) for an explanation.
We will frequently use the word form to refer to 1 and 2 forms when it is clear from
the context the type of the form.
Exterior product and exterior derivative
1.4.4. The exterior product of two 1-forms,
j
= f
j
dx + g
j
dy, j = 1, 2, is
the 2-form dened by

1

2
= (f
1
g
2
f
2
g
1
) dx dy.
This operation : S
1
(X) S
1
(X) S
2
(X) satises (and it is characterised by)
the following properties:
(
1
+
2
)
3
= (
1

3
) + (
2

3
),
(
1
)
2
= (
1

2
),

2

1
=
1

2
,32 1. RIEMANN SURFACES
for forms
j
, j = 1, 2, 3 and C.
We have that the exterior product of dz and d z is given by
dz d z = (dx + idy) (dx idy) = 2i dx dy.
Using this equation, if
j
= u
j
dz + v
j
d z, j = 1, 2, are 1-forms, we have

1

2
= (u
1
v
2
u
2
v
1
) dz d z.
1.4.5. The exterior product of forms is a particular case of the following general
algebraic construction. Given a complex vector space V of dimension n let W be
a vector space of dimension d =
_
n
2
_
, with basis {w
1
, . . . , w
d
}; we dene a mapping
: V V W by
(v
j
, v
l
) w
i+jn
if 1 j < k n,
(v
l
, v
j
) w
i+jn
if 1 j < k n,
(v
i
, v
i
) 0 for i = 1, . . . , n,
and extend it linearly to V V . Denote the image of (v
j
, v
k
) by v
j
v
k
. The space
W is usually denoted by V
_
V or
_
2
V , and is called the 2nd exterior product of
V .
For a (xed) point x of X we denote by S
1
x
(X) the vector space generated by dx
and dy. Similarly we denote by S
2
x
(X) the vector space generated by dxdy. Then
we have that S
2
x
(X) = S
1
x
(X)
_
S
1
x
(X).
Since S
1
is a 2 dimensional space the space S
2
has dimension 1.
A similar denition for the mth exterior product is easy to obtain from this con-
struction. The dimension of
_
m
V is
_
n
m
_
for m n, and 0 if m > n; hence there
are no 3-forms on a surface. In general, one can dene forms of order at most n on
an n dimensional manifold. See [12, pg. 731] and [11, pg. 391] for an Algebraic
approach, or [25] for a Dierential Geometric point of view.
1.4.6. The exterior derivative d, is an operator on functions and forms de-
ned in the following way. For f S(X) we set
df = f
x
dx + f
y
dy;1.4. DIFFERENTIAL FORMS 33
for a 1-form = fdx + gdy we dene
d = f
y
dy dx + g
x
dx dy = (g
x
f
y
)dx dy;
and for S
2
(X) we put d = 0. The form df is called the dierential of the
mapping f. In the basis {dz, d z} the exterior derivative is given by:
df = f
z
dz + f
z
d z, d = (v
z
u
z
) dz d z,
where = udz + vd z.
One can easily check that df = 0 if and only if f is constant.
We can decompose the action of d on functions in its (1, 0) and (0, 1) parts,
: S(X) S
(1,0)
(X)

: S(X) S
(0,1)
(X)
f f
z
dz f f
z
d z.
It is clear that a function is holomorphic if and only if

f = 0. Since d
2
= d d = 0,
we have
2
=

2
=

+

= 0 (observe that the operators and

anti-commute).
1.4.7. The conjugation operator is an operator on the space of 1-forms, :
S
1
(X) S
1
(X) dened by:
(fdx + gdy) = gdx fdy,
or in complex notation,
(udz + vd z) = iudz + ivd z.
Observe that () = .
1.4.8. A 1-form is called closed if d = 0, and exact if = df for some
smooth function f. We say that is co-closed (co-exact) if is closed (exact,
respectively).
Since d
2
= 0 exact forms are closed. The next lemma is an important result that
shows that closed forms are locally exact (see the next paragraph for the meaning
of the word locally in this context). We will prove it in 1.5.24.
Lemma (Poincare Lemma). Let D C be an open disc, and a closed 1-form
on D. Then is exact.34 1. RIEMANN SURFACES
Holomorphic and Harmonic Forms
1.4.9. A 1-form is called holomorphic if it is given locally by = df,
where f is a holomorphic function. Locally means that for every point p of X there
exists a neighbourhood V , and a holomorphic function f : V C, such that = df
on V . The function f and the set V might depend on the point p. The space of
holomorphic forms is denoted by (X). In the next chapter we will show that if X
is a compact Riemann surface of genus g then (X) has (complex) dimension g.
1.4.10. Proposition. A form is holomorphic if and only if it can be written
as = u dz, where u is a holomorphic function.
Proof. Since f is holomorphic we have f
z
= 0, and therefore = df = f
z
dz.
To prove the converse statement consider a local coordinate (V, z) near a point p, and
let be given by = u dz, where u is a holomorphic function V . By shrinking V if
necessary we can assume that z(V ) is a disc on C. The function u z
1
: z(V ) C
is holomorphic and therefore it has a holomorphic primitive [20, chapter 10]; that
is, there exists a holomorphic function on z(V ), say g, such that g

= u z
1
. The
functionf = g z is holomorphic on V (f z
1
= g) and satises = df.
1.4.11. The Laplacian operator : S(X) S
2
(X) assigns a 2-form to a
smooth function f by the rule f = (f
xx
+ f
yy
) dx dy. It is easy to see that
= d d = 2 i

, where is the conjugation operator dened in 1.4.7. A function
f is called harmonic if f = 0. Harmonicity is a property that depends on the
Riemann surface structure of X; thus it might happen that a function dened on
a surface is harmonic for certain Riemann surface structure but not for a dierent
structure (see exercise 66 for an example). A 1-form is called harmonic if it can
be written locally as = df, where f is a harmonic function.
1.4.12. Proposition. A 1-form is harmonic if and only if it is closed and
co-closed.
Proof. If = df is harmonic then
d = d
2
f = 01.4. DIFFERENTIAL FORMS 35
and
d() = d df = f = 0,
so is closed and co-closed.
Conversely, if is closed then is locally exact (Poincare lemma, 1.4.8); that
is, = df for a smooth function f. If is co-closed we get d = f = 0, so the
function f is harmonic.
1.4.13. Proposition. A 1-form = udz + vd z is harmonic if and only if u
and v are holomorphic functions.
Proof. We have
d = (v
z
u
z
)dz d z
and
d() = d (iudz + ivd z) = i (v
z
+ u
z
) dz d z.
So is closed and co-closed (therefore harmonic, by the previous proposition) if
and only if u
z
= v
z
= 0. As we have remarked earlier, u
z
= 0 is equivalent to u
being holomorphic. On the other hand, v
z
= ( v)
z
, so we v
z
= 0 if and only if v is
holomorphic.
1.4.14. If f is a holomorphic function dened on the complex plane then its
real and imaginary parts (1.1.1) are harmonic functions. The next proposition is
the equivalent result for holomorphic 1-forms.
Proposition. A form is holomorphic if and only if it can be written as
= + i (), for some harmonic form .
Proof. We have seen (1.4.10) that if is holomorphic then it can be written
(locally) as = u dz, for some holomorphic function u. Let be dened by =
1
2
(u dz ud z). By the previous result is harmonic, and we have
i ( ) =
i
2
(i u dz + i ud z) =
1
2
(u dz + u d z) ,
so + i ( ) = .36 1. RIEMANN SURFACES
Conversely, if is harmonic we can write =
1
+
2
, where
1
and
2
are holo-
morphic forms. Then i() = i(i
1
+i
2
) =
1

2
; thus +i() is holomorphic.
A dierent approach to forms
1.4.15. In this subsection we explain the denition of forms as given in [8] and
its relation with the denitions in the previous subsections.
Given a point p of X, consider the collection of pairs (U, f), where U is a neigh-
bourhood of p and f : U C is a smooth function. Equivalently, take the union

S(U), as U varies over the neighbourhoods of p. Dene an equivalence relation

p
in this union by identifying (U, f) and (V, g) if there exists a neighbourhood W
of p, with W U V and such that f|
W
= g|
W
. Let S
p
=
_

pU
S(U)
_
/
p
denote the set of equivalence classes (see 1.5.15 for a general construction of this
type). Consider now two subspaces of S
p
: S
p
(1), consisting of all equivalence classes
of functions that vanish at p, and S
p
(2) containing the equivalence classes of func-
tions satisfying f(p) = f
x
(p) = f
y
(p) = 0. The quotient space T

p
X = S
p
(1)/S
p
(2)
is called the cotangent space of X at p. The union of all these spaces as p varies
over X is the cotangent bundle of X, T

X =

pX
T

p
X. There is a natural pro-
jection : T

X X given by (T

p
X) = x (in the language of 2.4, the cotangent
space is a smooth vector bundle over X of rank 2). If f is a function dened in a
neighbourhood of p the class of f f(p) in T

p
X is called the dierential of f at
p, written as d
p
f. For a local coordinate z = x + iy at p, its real and imaginary
parts are functions dened near p; let d
p
x and d
p
y be the corresponding elements
(dierentials) of T

p
X.
Claim. The dierentials {d
p
x, d
p
y} form a basis of T

p
X.
If d
p
f is an arbitrary element of this vector space we know from basic Calculus
that there exists a function g satisfying f = a(x x(p)) + b(y y(p)) + g, and
g S
p
(2). Thus d
p
f = a d
p
x + b d
p
y. We see from this formula that a =
f
x
(p) and
b =
f
y
(p). To prove the linear independence assume that r d
p
x + s d
p
y = 0. Then
r(x x(p)) + s(y y(p)) belongs to S
p
(2). Taking partial derivatives evaluated at
p, we see that r = s = 0.1.4. DIFFERENTIAL FORMS 37
An important result obtained in the proof of the above claim is the following ex-
pression for the dierential of a function:
d
p
f =
f
x
(p) d
p
x +
f
y
(p) d
p
y.
1.4.16. A1-form is a mapping fromX to its cotangent space, : X T

X,
such that = Id
X
. This means that (p) belongs to the cotangent space of
X at p. So we can write = f dx + g dy i.e. (p) = f(p) d
p
x + g(p) d
p
y. We
say that is smooth if f and g are smooth functions. The wedge product of two
1-forms is dened as in 1.4.4. By (T

p
)
2
X we mean T

p
X
_
T

p
X, the space consisting
of all exterior products of 1-forms at p. Since d
p
x d
p
y = d
p
y d
p
x, the space
(T

p
)
2
X has dimension 1. A 2 -form is a mapping : X (T

)
2
X, such that
(p) (T

p
)
2
X. We have that is locally given by = f dx dy. If f is smooth
we say that is a smooth 2-form.
Clearly the above approach agrees with our original denitions, at least locally.
From the usual Calculus formul of the derivative of the composition of two func-
tions one can get the invariance property (behaviour under changes of coordinates)
of 1 and 2-forms.
1.4.17. The relation between this denition and the usual one given in Die-
rential Geometry is as follows: the set S
p
has the structure of a vector space of
dimension 2 over R. Let T
p
X be the set of linear mappings L : S
p
R, satisfying
the Leibnitz rule L(f g)(p) = (Lf)(p) g(p) +f(p) (Lg)(p); T
p
X is called the tangent
space of X at p. The cotangent space T

p
X is simply the dual of T
p
X (from where
the notation comes).
Integration of forms
1.4.18. The invariance property of forms, which resembles the change of vari-
ables used in Calculus, allows us to dene integrals of forms on surfaces, as we show
in this subsection.38 1. RIEMANN SURFACES
A function f : [a, b] C is said to have right derivative at a point t if the
limit
(f

)
+
(t) = lim
h0
+
f(t + h) f(t)
h
exists. Similarly one denes the left derivative.
Definition. A curve : [a, b] H (continuous mapping) is called piecewise
smooth if there exist points a = t
0
< t
1
< < t
n
= b, such that is smooth on
each of the intervals (t
j
, t
j+1
) and it has right and left derivatives at the points t
j
,
whenever it is possible to compute them (for example, at a we can talk only of the
right derivative).
If c : [0, 1] X is a curve on a Riemann surface X, we say that c is piecewise
smooth if we can cover its image c([0, 1]) with coordinate patches {(U
j
, z
j
)}
jJ
, such
that z
j
c is piecewise smooth (for each j).
For the rest of this section we will assume that all curves are piecewise smooth,
unless otherwise stated.
1.4.19. Subdividing [0, 1] into smaller intervals if needed we can assume that
for each subinterval [t
j
, j
j+1
] there exists a local coordinate patch (U
j
, z
j
= x
j
+iy
j
),
such that c([t
j
, t
j+1
]) U
j
. Let be a 1-form on X, given by = f
j
dx + g
j
dy on
U
j
. We dene
_
c
=
n

j=1
_
t
j+1
t
j
_
f
j
(c(t))
(x
j
c)
t
(t)g
j
(c(t))
(y
j
c)
t
(t)
_
dt.
We leave to the reader to check that this denition is independent of the choices of
local coordinates or subdivision of the interval [0, 1].
1.4.20. See [25] for proofs of the next two results.
Proposition. Let be a closed 1-form on X, and c
j
: [0, 1] X, j = 1, 2,
two paths with the same end points. If c
1
is homologous to c
2
(the cycle c
1
c
2
is a
boundary) then
_
c
1
=
_
c
2
.1.4. DIFFERENTIAL FORMS 39
Corollary. If be closed 1-form, and c
1
and c
2
two homotopic paths on X
(with the same end points), then
_
c
1
=
_
c
2
.
1.4.21. The integration of a 2-form requires a little more work. First of all,
observe that one cannot speak of the value of a form at a point. However, since
changes of coordinates have non-zero Jacobians, it makes sense to say whether a
form vanishes or not at a given point. More precisely, if S
2
(X), we dene its
support by
supp() = {p X; (p) = 0}.
We have that supp() is a well dened set. Assume that for a certain form the set
supp() is compact and contained in a single coordinate patch, say (U, z = x + iy).
We dene the integral of by the expression
(2)
_
X
=
_
z(U)
(f z
1
) ds dt,
where (s, t) are coordinates on R
2
. As in the case of 1-forms, the invariance property
can be used to prove that this integral is well dened (its value does not depend on
the local coordinate z). More generally, if is a 2-form with compact support we
cover supp() by nitely many coordinates patches, say (U
j
, z
j
), j = 1, . . . , n. Let
{
1
, . . . ,
n
} be a partition of unity subordinate to this covering (1.1.23). We then
set
(3)
_
X
=
n

j=1
_
X
f
j
,
where is given by f
j
dx
j
dy
j
on U
j
. Observe that the forms
j
have compact
support contained in U
j
; their integrals are dened by the expression (2). One can
check that (3) is independent of the choices made to compute it.
1.4.22. One of the most important results on integration of forms is given by
the following theorem.
Theorem (Stokes). Let be a 1-form on a Riemann surface X. Let U be an
open subset of X with compact closure and smooth boundary U (that is, U consists
of a nite collection of smooth curves). Then
_
U
d =
_
U
.40 1. RIEMANN SURFACES
Corollary. Let X be a compact Riemann surface, and S
1
(X). Then
_
X
d = 0.
The integration on the boundary of U has to be done with certain orientation.
We will use Stokes theorem in two types of settings: in one case we consider two
disjoint domains U
1
and U
2
, with a common boundary, given by a smooth curve c,
U
1
= U
2
= c. Then
_
U
1
=
_
U
2
. In the other situation (2.2.1) the domain
of integration will be a disc, U = {z C; |z| < R} with the standard orientation of
C; its boundary, the circle of centre 0 and radius R, is oriented clockwise (opposite to
the standard orientation). For more details on Stokes theorem and its applications
see [25].
1.4.23. We will need two more theorems of Dierential Geometry. The rst
one tells us that the set of bad (critical) values of a smooth map is small in
measure. More precisely, let f : X R be a smooth proper map dened on a
(smooth) surface X. Recall that f is called proper if f
1
(K) is compact for any
K R compact. Let p be a point of X, and z = x + iy a local coordinate near p.
We say that p is a critical point of f is f
x
(p) = f
y
(p) = 0. Critical points are well
dened because changes of coordinates have non-vanishing derivatives. A critical
value is the image of a critical point under f.
Theorem (Sard). Let f : X R be a smooth proper mapping dened on a
smooth surface. Let C be the set of critical values of f. Then C has zero measure
in R.
1.4.24. The next result classies all smooth curves (1-dimensional mani-
folds) [15].
Theorem. Any compact, smooth 1-dimensional manifold is dieomorphic to a
circle.
Meromorphic forms
1.4.25. A 1-form is called meromorphic if it can be written locally as
= f(z) dz, where f is a meromorphic function. The set of poles of the functions1.4. DIFFERENTIAL FORMS 41
f are also known as the poles of . They are well dened (do not depend on the
function f chosen to represent the form) and form a discrete set of points of X. The
set of meromorphic forms on X will be denoted by M
1
(X). In a local coordinate
(U, z) the function f has a power series expansion of the form
f(p) =

j=N
a
j
(z(p))
j
, p U,
which we simplify as f =

j=N
a
j
z
j
. The coecient a
1
is called the residue of
the form at the point p, and we will denote it by res
p
().
Proposition. The residue of a meromorphic form is well dened.
Proof. We need to show that a
1
is independent of f and the coordinate z.
Assume then that z(p) = 0 as above. Let r and be positive numbers such that
D(0, r +), the disc of centre 0 and radius r + is contained in z(U). Since the poles
of form a discrete set we can further assume that the only singularity of the form
in z
1
(D(0, r + )) = {q U; |z(q)| < r + } lies at the point p. Consider now
the smooth curve c : [0, 1] U given by c() = z
1
(re
i
). One easily sees that
a
1
=
1
2i
_
c
.
Since this formula gives an way of computing a
1
in terms of the integral of over
a path we see that the residue does not depend of local coordinates (since path
integrals are coordinate-independent).
1.4.26. The following result is a consequence of Stokes Theorem.
Theorem (Residues Theorem). Let be a meromorphic form on a compact
surface X. Then

pX
res
p
() = 0.
Proof. First of all, since X is compact and the set of poles of is discrete we
have that the above sum is nite. So let us assume that these poles are at points
p
1
, . . . , p
n
. It is possible to nd a simply connected set with smooth boundary, say
U, containing all the poles, as follows: let c

be a smooth curve passing though the


poles (once by each pole) and take a small, smooth neighbourhood of c

, (which
can be done in local coordinates), as in gure 8. Let c denote the boundary of U.42 1. RIEMANN SURFACES
* * *
D
c
c

Figure 8. Proof of theorem 1.4.26.


We have that c is homotopic to a collection of small circles around the poles p
j
,
similar to the curves used in the proof of the previous result. Thus
_
c
=
n

j=1
res
p
j
.
On the other hand the form has no poles on X\U so it is holomorphic and therefore
closed. Applying Stokes Theorem (1.4.22) we get
0 =
_
X\U
d =
_
c
,
which completes the proof.
1.5. Sheaf Cohomology
The purpose of this section is to make the reader familiar with the language (and
basic techniques) of sheaf theory that will be used in the statements and proofs of
the results of the next chapter.
Sheaves of abelian groups
1.5.1. Definition. A presheaf of abelian groups F on a Riemann surface
(or topological space) X is an assignment of an abelian group, F(U), to each open
subset U of X, and a collection of group homomorphisms
U
V
: F(U) F(V ), for
every pair of open sets V and U with V U, satisfying the following two conditions:
(1)
U
U
is the identity homomorphism;
(2) if W V U are open sets then
V
W

U
V
=
U
W
.
The mappings
U
V
are called the restriction homomorphisms (of the presheaf).
If f F(U) and V U we will write f|
V
for
U
V
(f), or simply consider f as an1.5. SHEAF COHOMOLOGY 43
element of F(V ) if it is clear from the context. If the sets F(U) have extra structures,
for example they are vector spaces, one can dene presheaves of vector spaces, by
requiring the restriction homomorphisms to be linear mappings. Similarly one has
presheaves of rings, elds and other structures. In this book we will work only with
presheaves of abelian groups and vector spaces.
An example of presheaf is given by C, the set of continuous functions, with the
usual restriction of domains of functions: C(U) denotes the space of continuous
functions dened on the open set U. Here we are assuming that the functions take
values in C but one can consider the presheaf of real valued continuous functions
as well. Another examples of presheaves are given by S, O and M, consisting of
smooth, holomorphic and meromorphic functions respectively.
1.5.2. Definition. A presheaf F is called a sheaf if for every open set U of
X, and every collection of open sets {U
j
}
jJ
, such that

j
U
j
= U, the following
conditions are satised:
(1) if f and g are elements of F(U) such that f|
U
j
= g|
U
j
, for all j J, then
f = g;
(2) if f
j
F(U
j
), j J, and f
j
|
U
j
U
k
= f
k
|
U
j
U
k
, then there exists an element
f U with f|
U
j
= f
j
.
The examples of presheaves given above are actually sheaves (see exercise 36 for an
example of a presheaf that is not a sheaf). Another example of a sheaf is given by
O

, the set of nowhere vanishing holomorphic functions


O

(U) = {f : U C\{0}; f is holomorphic}.


Similarly we can dene M

, the sheaf of invertible meromorphic functions; that


is, M

(U) consists of the meromorphic functions on U which are not identically 0


on any component of U.
For the rest of this section F, G and H will denote sheaves of abelian groups.
Cohomology groups44 1. RIEMANN SURFACES
1.5.3. Let U = {U
j
}
jJ
be an open covering of a Riemann surface X. The
nth cochain group, n a non-negative integer, C
n
(U, F), is dened by
C
n
(U, F) =

(j
0
,...,jn)J
n+1
F(U
j
0
U
jn
).
Elements of C
0
(U, F) are of the form (f
j
)
jJ
, where f
j
F(U
j
); we will simplify this
notation by writing (f
j
). Similarly, (f
jk
) and (f
jkl
) will denote elements of C
1
(U, F)
and C
2
(U, F), respectively. The elements of C
n
(U, F) are called n-cochains.
1.5.4. The coboundary operators, : C
n
(U, F) C
n+1
(U, F), for n = 0, 1,
are dened in the following way:
1. if (f
j
) C
0
(U, F) then ((f
j
)) = (g
kj
), where g
jk
is given by g
jk
= f
k
f
j
, in
U
j
U
k
; (strictly speaking we mean g
jk
= f
k
|
U
j
U
k
f
j
|
U
j
U
k
; but see above for the
notation);
2. for (f
jk
) C
1
(U, F) we set ((f
jk
)) = (g
jkl
), with g
jkl
= f
kl
f
jl
+ f
jk
in
U
j
U
k
U
l
(similarly g
jkl
= f
kl
|
U
j
U
k
U
l
f
jl
|
U
j
U
k
U
L
+ f
jk
|
U
j
U
k
U
l
).
The operator satises
2
= = 0. To check this, let (f
j
) be a 0-cochain. Then
we have
(((f
j
))) = ((f
k
f
j
)) = (f
l
f
k
) (f
l
f
j
) + (f
k
f
j
) = 0.
1.5.5. The groups of cocycles and coboundaries, Z
1
(U, F) and B
1
(U, F),
are dened respectively as the kernel of : C
1
(U, F) C
2
(U, F) and the image of
: C
0
(U, F) C
1
(U, F). A 1-cochain (f
jk
) is a cocycle if it satises f
jl
= f
jk
+f
kl
in F(U
j
U
k
U
l
) (we can add the middle subindex). In particular a cocycle
satises f
jj
= 0 and f
jk
= f
kj
. A 1-cochain is a coboundary, or it splits, if
f
jk
= f
k
f
j
for some 0-cochain (f
j
). Observe that this is equivalent to the equality
f
jk
= g
j
g
k
where (g
k
) is the cocycle given by g
k
= f
k
. The group B
1
(U, F) is
a subgroup of Z
1
(U, F) because
2
= 0. Since cochain groups are abelian, it is a
normal subgroup, so the quotient is a group, called the 1st cohomology group of
X (with coecients in F with respect to the covering U),
H
1
(U, F) = Z
1
(U, F)/B
1
(U, F).1.5. SHEAF COHOMOLOGY 45
1.5.6. Our goal is to dene a group that does not depend of the covering U. To
do this we rst introduce an order in the collection of all open coverings of X: the
covering V = {V

}
A
is ner than the covering U = {U
j
}
jJ
, written as V < U,
if for every A there exists a j J, such that V

U
j
. Let : A J be a
mapping such that V

U
()
. Observe that might not be uniquely dened, but
we will show that our constructions are independent of the choice of this renement
mapping (1.5.7). Since V

is an open subset of U
()
the mapping composed with
the restriction homomorphisms induces a mapping between cochain groups in the
following way:

U
V
: C
1
(U, F) C
1
(V, F)
(f
jk
) (f
(a)(b)
|
Vav
b
).
It is clear that
U
V
preserves cocycles and coboundaries and thus it induces a mapping
between cohomology groups, which we will denote by
U
V
as well:

U
V
: H
1
(U, F) H
1
(V, F).
We next show that this mapping, at the cohomology level, does not depend of
the renement map .
1.5.7. Lemma. The mapping
U
V
: H
1
(U, F) H
1
(V, F) is independent of
.
Proof. Assume : A J is another renement map; that is, V

U
()
, for
all A. Dene two mappings by the expressions
K
1
: C
1
(U, F) C
0
(V, F)
(f
jk
) (f
()()
|
V
)
and
K
2
: C
2
(U, F) C
1
(V, F)
(f
jkl
)
_
(f
()()()
f
()()()
)|
VV

_
.46 1. RIEMANN SURFACES
We claim that K
2
K
1
=
U
V

U
V
. This is a simple computational exercise. Let
(f
jk
) C
1
(U, F); then
K
2
((f
jk
)) =K
2
(f
jk
f
jl
+ f
kl
) =
=f
()()
f
()()
+ f
()()
f
()()
+ f
()()
f
()()
,
and
(K
1
(f
jk
)) = (f
()()
) = f
()()
f
()()
.
If Z
1
(U, F) is a cocycle then K
1
() is a 0-cochain with respect to the covering
V, so (K
1
()) B
1
(V, F). On the other hand, if we write = (f
jk
) we have
K
2
(()) = f
()()
+ f
()()
f
()()
f
()()
= f
()()
f
()()
= 0.
Thus (
U
V

U
V
)() B
1
(V, F). But this implies that
U
V
and
U
V
induce the same
mappings in cohomology.
1.5.8. Lemma. The mapping
U
V
: H
1
(U, F) H
1
(V, F) is injective.
Proof. Assume that (f
jk
) is a cocycle (with respect to U) such that
U
V
((f
jk
))
splits. This means that there exists a cochain (h

), satisfying f
()()
= h

on
V

. Dene f
j
F(U
j
) by f
j
= f
j()
+ h

on F(U
j
V

). For A, on the
intersection U
j
V

we have
(h

+ f
j()
) (h

+ f
j()
) = h

+ f
j()
+ f
()j
= f
()()
+ f
()()
= 0,
because (f
jk
) is a cocycle. Since the sets U
j
V

cover U
j
as varies over A, we have
that f
j
is an element of F(U
j
), by property (2) in denition 1.5.2. On U
j
U
k
V

we get
f
j
f
k
= (f
jt()
+ h

) (f
kt()
h

) = f
jt()
+ f
t()k
= f
jk
,
and therefore, by property (1) of sheaves, we have that f
j
f
k
= f
jk
on U
j
U
k
;
that is, (f
jk
) splits.
1.5.9. The mappings
U
V
can be used to dene a cohomology group independent
of the coverings. First of all, if we have three coverings, U, V and W satisfying
W < V < U, it is easy to see that
V
W

U
V
=
U
W
. Consider now the disjoint union
of cohomology groups

H
1
(U, F), where U varies over all possible open coverings
of X. We dene an equivalence relation in this union by identifying H
1
(U, F)1.5. SHEAF COHOMOLOGY 47
and H
1
(V, F), , if
U
W
() =
V
W
() in H
1
(W, F), for some covering W
ner than both U and V. The quotient set
H
1
(X, F) :=
_
U
H
1
(U, F)/ ,
is called the 1st cohomology group of X (with coecients in the sheaf F). The
group structure on this set is dened by passing to ner coverings; if and are two
cohomology classes, choose H
1
(U, F) and H
1
(V, F), such that = [] and
= [] (where square brackets denote equivalence classes in H
1
(X, F)). Then we
dene + = [t
U
W
()+t
V
W
()], where W is a covering satisfying W < V and W < U.
Observe that such a covering always exists: if U = {U
j
}
jJ
and V = {V
a
}
aA
, then
W = {U
j
V
a
}
(j,a)JA
is ner than U and V. We leave as an exercise to show that this addition in H
1
(X, F)
is well dened. From the denition of cohomology groups and lemma 1.5.8 it follows
that for any covering U the natural mappings H
1
(U, F) H
1
(X, F) are injective.
1.5.10. The following result is obvious from these observations, but we state
it as a separate lemma because we will use it frequently.
Lemma. For a sheaf F of abelian groups on a Riemann surface X, the cohomol-
ogy group H
1
(X, F) vanishes if and only if for any open covering U of X the group
H
1
(U, F) vanishes.
1.5.11. Remark. Some authors consider cochain groups as in 1.5.3, but with
the condition that j
k
= j
l
if k = l. For example, if U consists of two open sets, say
U
1
and U
2
, and f is a 1-cochain in their sense, then f consists of f
12
and f
21
, while
with our denitions f will be given by f
11
, f
12
, f
21
and f
22
. But it can be proven
that both constructions give the same cohomology groups (if f is a cocycle then it
is determined by f
12
and f
21
since f
11
and f
22
must vanish), which are the objects
of our interest.
1.5.12. We have enough material to give an example of a cohomology group.
Proposition. For a Riemann surface X we have H
1
(X, S) = 0, where S is the
sheaf of smooth functions on X.48 1. RIEMANN SURFACES
Proof. Let U = {U
j
}
jJ
be an open covering of X and (f
jk
) an element of
H
1
(U, S). Let {
j
}
jJ
be a partition of unity subordinate to U. The functions
f
jk

k
belong to S(U
j
) (gure 9). Dene
h
j
=

k
f
jk

k
.
Then we have
h
j
h
k
=

l
f
jl

l
f
kl

l
=

l
(f
jl
f
kl
)
l
=

l
(f
jl
+ f
lk
)
l
=

l
f
jk

l
=f
jk

l
= f
jk
.
Observe that these sums are nite (at every point).
U
k
U
j
supp (
j
)
Figure 9. Proposition 1.5.12.
The above result also holds for the sheaves of 1 and 2 forms, S
1
and S
2
, as well as
for the sheaves of forms of type (1, 0) and (0, 1), S
(1,0)
and S
(0,1)
.
1.5.13. The following result is a useful tool to compute cohomology groups.
Theorem (Leray). Let U = {U
j
}
jJ
be an open covering of a Riemann surface
X and F a sheaf of abelian groups on X. Assume that H
1
(U
j
, F) = 0 for all j J;
then H
1
(X, F)

= H
1
(U, F).
Proof. Observe that the groups H
1
(U
j
, F) are the cohomology groups of the
sets U
j
considered as Riemann surfaces. Let V = {V
a
}
aA
be a covering ner than
U and : A J a renement mapping. We have proved that
U
V
: H
1
(U, F)
H
1
(V, F) is injective; we will show that it is surjective as well, and that will prove
the theorem. So given (f
ab
) in C
1
(V, F), we need to nd a cocycle (F
jk
) C
1
(U, F),
such that (F
t(a)t(b)
f
ab
) splits (with respect to the covering V).
For a xed j J the collection U
j
V = {U
j
V
a
}
aA
is an open covering of
U
j
. Since H
1
(U
j
, F) = 0, there exists a cochain (g
j
a
) C
0
(U
j
V, F), such that1.5. SHEAF COHOMOLOGY 49
f
ab
= g
j
a
g
j
b
(on U
j
V
a
V
b
). Set F
jk
= g
k
a
g
j
a
, on U
j
U
k
V
a
; by property (2)
in the denition of sheaf (1.5.2), (F
jk
) F(U
j
U
k
). To check that (F
jk
) is actually
a cocycle observe that
F
jk
+ F
kl
= (g
k
a
g
j
a
) + (g
l
a
g
k
a
) = g
l
a
g
j
a
= F
jl
,
and then use property (1) in (1.5.2). We claim that (F
t(a)t(b)
f
ab
) splits. Write
h
a
= g
t(a)
a
. Then we have
F
t(a)t(b)
f
ab
= g
t(b)
a
g
t(a)
a
g
t(b)
a
+ g
t(b)
b
= g
t(b)
b
g
t(a)
a
= h
b
h
a
,
which shows that F
t(a)t(b)
f
ab
splits.
A covering satisfying the hypothesis of the theorem is called a Leray covering.
1.5.14. Consider the subgroup of the group C
0
(U, F) given by
Z
0
(U, F) = ker
_
C
0
(U, F)

C
1
(U, F)
_
.
If a cochain (f
j
) satises ((f
j
)) = 0 then we have f
j
= f
k
on U
j
U
k
. From the
denition of a sheaf we have that there exists an element f F(X), such that
f|
U
j
= f
j
. Thus Z
0
(U, F) can be identied with F(X), for any covering U. The
B
0
(U, F) is the trivial group (it does not make sense to talk of 0-coboundaries);
thus we dene the 0th cohomology group of X (with coecients in F) by
H
0
(X, F) = F(X).
Sheaf associated to a presheaf
1.5.15. We now introduce a standard way of associating a sheaf to a given
presheaf. The reader familiar with continuation of analytic functions in the complex
plane will nd similarities between that construction and the one that follows. See
also 1.4.15. Let F be a presheaf of abelian groups on X. For a point p X, take
the disjoint union

U
F(U), where U varies over the open neighbourhoods of p.
We dene an equivalence relation
p
on this union as follows: let f F(U) and
g G(V ), then f
p
g if there exists a neighbourhood W of p, with W U V ,50 1. RIEMANN SURFACES
such that f|
W
= g|
W
. Let F
p
denote the quotient set, F
p
= (
U
F(U)) /
p
, called
the stalk of F at p. For f F(U) we will denote its class in F
p
by f
p
. Let
N(U, p) = {f
q
; q U}, and declare these sets, as U varies over all neighbourhoods
of p, a fundamental set of neighbourhoods of f
p
in |F| =
qX
F
q
. This construction
denes a topology in |F|. The mapping : |F| X, dened by (f
p
) = f(p) is
continuous (observe that f(p) is well dened, since all elements of F
p
agree on p).
For A X open, we set |F|(A) = {s : A |F|; s continuous and s = Id
A
}.
This makes |F| a sheaf on X.
Sheaf homomorphisms and sequences
1.5.16. Definition. Let F and G be sheaves of abelian groups on a Riemann
surface X. A sheaf homomorphism : F G is a collection of group homomor-
phisms,
U
: F(U) G(U), for each open set U of X, such that if V U, then
diagram
F(U)
u
//

U
V

G(U)

U
V

F(V )

V
//
F(V )
commutes. Here and denote the restriction homomorphisms of the sheaves F
and G respectively.
If it is clear from the context, we will drop the subindex in
U
and simply write
: F(U) G(U).
If F and G are sheaves of some other structure then the mappings
U
are required
to be compatible with that structure; for example, for sheaves of vector spaces the
mappings
U
should be linear mappings.
1.5.17. An example of a sheaf homomorphism is provided by the exterior de-
rivative, d : S S
1
, f df. The and

operators are also sheaf homomorphisms.
The kernel of

consists of the holomorphic functions (on U) and therefore it is a
sheaf. In general, if we dene
K(U) = ker
_

U
: F(U) G(U)
_
,1.5. SHEAF COHOMOLOGY 51
we have that K is a sheaf, called the kernel sheaf.
1.5.18. Another example of sheaf homomorphism that we will use in the next
chapter is given by the exponential map, exp : O(U) O

(U), exp(f) = e
2if
. The
kernel of this sheaf consists of the locally constant functions with integer values,
which we denote by Z as well: a continuous function f : U C belongs to Z(U) if
it takes integer values and every point of U has a neighbourhood where f is constant.
This is equivalent to require that f is constant on each connected component of U.
In particular, since Riemann surfaces are connected, Z(X) consists of the constant
functions with integer values, which is clearly isomorphic to Z. See also exercise 37.
1.5.19. Definition. Given two sheaf homomorphisms : F G and :
G H, we say that the sequence F

G

H is exact if the following two
conditions are satised (for every open set U of X):
(1)
U

U
= 0;
(2) if g G(U) satises
U
(g) = 0, then there exists an open covering {U
j
}
jJ
of U and elements f
j
F(U
j
), such that
U
j
(f
j
) = g|
U
j
.
The homomorphisms and induced group homomorphisms at the stalk level,
F
x
x
G
x
x
H
x
. The above denition is equivalent to requiring that, for each x in
X, this sequence is exact; that is, ker(
x
) = im(
x
)
1.5.20. Lemma. If 0 F

G

H is exact (where the rst homomorphism,
0 F is the trivial homomorphism) then for every open subset U of X, the sequence
0 F(U)
u
G(U)

U
H(U) is exact.
Proof. Step 1: : F(U) G(U) is injective. This fact follows immediately
from the denition of exact sequence. Assume that f F(U) is mapped to the 0
element by
U
. From (2) in 1.5.19 we have that there exists a covering {U
j
}
j
of U,
such that f|
U
j
= 0 (since the rst sheaf is trivial). But this implies that f = 0.
Step 2: (F(U)) ker(
U
). This is simply condition (1) in the denition of exact
sequence.
Step 3: ker(
U
) (F(U)). Let (g) = 0 for some g G(U). We have a
covering {U
j
}
j
of U, and elements f
j
F(U
j
), such that (f
j
) = g|
U
j
. But then
(f
j
f
k
) = g g = 0 on U
j
U
k
; since is injective by the rst step in this52 1. RIEMANN SURFACES
proof, we have that f
j
= f
k
. Thus there exists f F(U) such that f|
U
j
= f
j
, which
implies (f) = g.
The rst step in the above proof will be frequently used, so we write it as a separate
result for easier reference.
Corollary. If 0 F

G is exact, then for any open set U X, the mapping

U
: F(U)

G(U) is injective.
1.5.21. A sheaf homomorphism : F G induces mappings between
cohomology groups in a natural way. For the 0-th cohomology groups, since
H
0
(X, F) = F(X) and H
0
(X, G) = G(X), the mapping
X
: F(X) G(X) can be
considered as a mapping in cohomology,
0
=
X
: H
0
(X, F) H
0
(X, G).
To make a similar contruction with 1st cohomology groups, we consider a class
in H
1
(X, F). Choose a cochain to represent this class, = [(f
jk
)], in some
open covering U = {U
j
}
jJ
of X, with (f
jk
) Z
1
(U, F). It is easy to check that
(
U
j
U
k
(f
jk
)) is an element of Z
1
(U, G). Moreover, if (f
jk
) = (f
j
f
k
) is a coboundary
we have ((f
jk
)) = ((f
j
) (f
k
)) is also a coboundary. Thus we obtain a mapping

1
U
: H
1
(U, F) H
1
(U, G) dened by the expression
1
([(f
jk
)]) = [((f
jk
))]. Is it
easy to chek that if V is a covering of X ner that U, then
U
V

1
U
=
1
V

U
V
. Thus
we can dene mapping
1
: H
1
(X, F) H
1
(X, G) by
1
() = [((f
jk
))].
Assume now that 0 F

G

H 0 is an exact sequence. From the above
constructions we get two sequences
0 H
0
(X, F)

0
H
0
(X, G)

0
H
0
(X, H),
and
H
1
(X, F)

1
H
1
(X, G)

1
H
1
(X, H).
We want to have a homomorphism

: H
0
(X, H) H
1
(X, F) connecting them.
To do that, consider an element h of H
0
(X, H) = H(X). Since the sheaf sequence
0 F

X
G

X
H 0 is exact (lemma 1.5.20) and h is mapped to the 0 element
(the last homomorphism is trivial), we have a covering U = {U
j
}
jJ
, and elements
g
j
G(U
j
), such that (g
j
) = h|
U
j
. Clearly (g
j
g
k
) = 0, so there exist f
jk
1.5. SHEAF COHOMOLOGY 53
F(U
j
U
k
) with (f
jk
) = g
j
g
k
. It is not dicult to see that (f
jk
) is a cocycle:
(f
jk
+ f
kl
+ f
lj
) = (f
jk
) + (f
kl
) + (f
lj
) = g
j
g
k
+ g
k
g
l
+ g
l
g
j
= 0,
and therefore f
jk
+ f
kl
+ f
lj
= 0, since is injective (corollary 1.5.20). One can
check similarly that the cohomology class of (f
jk
) does not depend on the choices
made in its construction. We then set

(h) = f, where f = [(f


jk
)] H
1
(X, F).
1.5.22. Theorem. Assume 0 F

G

H 0 is an exact sequence.
Then the sequence
0 H
0
(X, F)

0
H
0
(X, G)

0
H
0
(X, H)

H
1
(X, F)

1
H
1
(X, G)

1
H
1
(X, H)
is exact.
Proof. Step 1. The rst terms of the sequence are
0 F(X)

0
G(X)

0
H(X),
which is exact by lemma 1.5.20.
Step 2: im(
0
) ker(

). Let h H(X) and g G(X) such that


0
(g) = h.
We have to show that

(h) = 0, in H
1
(X, F). In the construction of

we take
elements g
j
G(U
j
) (for some covering of X), such that (g
j
) = h|
U
j
. Clearly in
this situation we can take g
j
= g|
U
j
, and then we get (f
jk
) = g g = 0. Since is
injective we have f
jk
= 0, therefore

(h) = f = 0 (with the above notation).


Step 3: ker(

) im(
0
). Suppose

(h) = ([(f
jk
)]) splits; that is, [(f
jk
)] =
[(f
j
f
k
)], for some 0-cochain (f
j
). Let (g
j
) be the 0-cochain used in the construction
of

(h); then g
j
(f
j
) = g
k
(f
k
) on (U
j
U
k
). By the denition of sheaf, there
exists g G(X) satisfying g|
U
j
= g
j
(f
j
). Then (g)|
U
j
= (g
j
) ((f
j
)) =
(g
j
) = h|
U
j
, or, in other words, (g) = h.
Step 4: im(

) ker(
1
). If h H(X) satises

(h) = [(f
jk
)], then (f
jk
) =
g
j
g
k
, that is, ((f
jk
)) splits.
Step 5: ker(
1
) im(

). Assume (f
jk
) = g
j
g
k
. Then (g
j
) = (g
k
) +
((f
jk
)) = (g
k
), so there exists h H(X) with h|
U
j
= (g
j
). Clearly

(h) =
[(f
jk
)].54 1. RIEMANN SURFACES
Step 6: im(
1
) ker(
1
). This follows from the exactness of the sequence
F(U
j
U
k
)

G(U
j
U
k
)

H(U
j
U
k
).
Step 7: ker(
1
) im(
1
). Assume that (g
jk
) is a cocycle in X (with coecients in
the sheaf G), such that (g
jk
) = h
j
h
k
for some 0-cochain (with coecients in H).
Since G H 0 is exact, by shrinking the covering if necessary, we can further
assume that h
j
= (g
j
), for elements h
j
H(U
j
). Then
(g
jk
(g
j
g
k
)) = 0,
so there exists a 1-cochain (f
jk
) satisfying (f
jk
) = g
jk
g
j
+ g
k
. Clearly maps
(f
jk
+f
kl
+f
lj
) to 0, and therefore, by the injectivity of
U
j
U
k
U
l
, the cochain (f
jk
)
is a cocycle. Moreover, we have

1
([(f
jk
)]) = [((f
jk
))] = [(g
jk
g
j
+ g
k
)] = [(g
jk
)],
which completes the proof of the theorem.
1.5.23. Corollary. In the hypothesis of the above theorem, if H
1
(X, G) = 0,
then H
1
(X, F)

= H(X)/ (G(X)).
Proof. From the previous theorem we have that the following sequence is exact,
G(X)

H(X)

1
H
1
(X, F) 0.
So
1
is surjective, and the result follows.
The de Rham theorem
1.5.24. We have seen that the exterior derivative can be considered as a sheaf
homomorphism from S (smooth functions) to S
1
(1-forms). If a function f satises
df = 0, then it is locally constant. Thus C, the sheaf of locally constant functions
with values in the complex numbers, is the kernel of this sheaf homomorphism. Let
S
1
c
denote the sheaf of closed forms. We claim that the sequence
0 C S
d
S
1
c
01.5. SHEAF COHOMOLOGY 55
is exact, where the mapping C S is the natural inclusion. The only step that
requires some work is to show that d is locally surjective. In other words, we have
to prove that any closed form is locally exact. This is precisely the statement of
Poincares lemma (1.4.8), which we prove next.
Consider a closed form = fdx+gdy on the unit disc D. Since d = 0 we have
f
x
(x, y) = g
y
(x, y). Dene a function F : D C by the expression
F(x, y) =
_
1
0
(xf(tx, ty) + y g(tx, ty)) dt, (x, y) D.
For the reader used to path integration the function F can be written as
F(x, y) =
_
c
(f(z) + g(z)) dz,
where c is the path with image the straight line from 0 to (x, y); that is, c(t) =
(tx, ty) : [0, 1] D.
Clearly F is a smooth function; we will show that dF = . Observe rst that
f(tx, ty)
t
= x
f(tx, ty)
x
+ y
f(tx, ty)
y
;
and therefore, using the equality f
x
= g
y
we get
t
_
f(tx, ty)
t
_
= txf
x
(tx, ty) + ty g
x
(tx, ty).
If we now compute the derivative of F we see
F(x, y)
x
=
_
1
0

x
(xf(tx, ty) + y g(tx, ty)) dt =
=
_
1
0
(f(tx, ty) + txf
x
(tx, ty) + ty g
x
(tx, ty)) dt =
=
_
1
0
(f(tx, ty) + tf
t
(tx, ty)) dt =
_
1
0

t
(tf(tx, ty)) dt
=tf(tx, ty)|
t=1
t=0
= f(x, y).
A similar argument shows that F/y = g, so dF = , as claimed.
The group S
1
c
(X)/dS(X) is known as the de Rham group of X, and denoted
by H
1
dR
(X).
Theorem. The groups H
1
dR
(X) and H
1
(X, C) are isomorphic.56 1. RIEMANN SURFACES
Proof. It follows from the above computations, corollary 1.5.23 and the fact
that H
1
(X, S) = 0 (1.5.12).
Remark. It is possible to prove that in each class of H
1
dR
(X) there is a unique
harmonic form. This is the content of the Hodge theorem, which we will not show
here. See [7] for a classical approach, or [10] for a modern development.
The 2nd Cohomology Group
1.5.25. In 2.5 we will use 2nd cohomology groups as well. Since the con-
structions and proofs are similar to what we have done so far in this section we will
describe them only briey. The interested reader can ll in the details.
The coboundary homomorphism : C
2
(U, F) C
3
(U, F) is given by (f
ijk
) =
(g
ijkl
), where
g
ijkl
= f
jkl
f
ikl
+ f
ijl
f
ijk
on U
i
U
j
U
k
U
l
. The groups of cocycles and coboundaries are dened in a way
similar to the 1-cochains case. Lemma 1.5.7 and its proof generalise to this setting,
where the operator K
3
: C
3
(U, F) C
2
(V, F) is given by
K
3
(f
ijkl
) = f
()()()()
f
()()()()
+ f
()()()()
.
We leave it to the reader to nish the construction of H
2
(X, F).
1.5.26. The long exact sequence of 1.5.22 can be extended to include the second
cohomology groups.
Theorem. Assume 0 F

G

H 0 is an exact sequence. Then there
exists a group homomorphism

: H
1
(X, H) H
1
(X, F) such that the following
sequence is exact:
0 H
0
(X, F)

0
H
0
(X, G)

0
H
0
(X, H)

H
1
(X, F)

1
H
1
(X, G)

1

1
H
1
(X, H)

H
2
(X, F)

2
H
2
(X, G)

2
H
2
(X, H).
CHAPTER 2
Compact Riemann surfaces
2.1 Divisors 58
2.2 Dolbeaults Lemma and Finiteness Results 62
2.3 The Riemann-Roch Theorem 68
2.4 Line bundles and Divisors 71
2.5 Serre Duality 81
2.6 Applications of the Riemann-Roch Theorem 95
2.7 Projective embeddings 99
2.8 Weierstrass Points and Hyperelliptic Surfaces 106
2.9 Jacobian Varieties of Riemann Surfaces 115
5758 2. COMPACT RIEMANN SURFACES
In this chapter we study compact Riemann surfaces in some detail and obtain
several important results. We start (2.1) with the denition of divisor: loosely
speaking a divisor is a nite set of points of a surface to which certain integral
weights (numbers) are assigned. We are interested on nding functions and forms
whose zeroes and poles are determined by a divisor. This question is naturally
related (and in a some sense a generalisation) to the problem of nding non-constant
meromorphic functions on surfaces. The dimensions of the spaces of functions and
forms associated to a divisor are given by the Riemann-Roch theorem, one of the
most important results in the theory of compact Riemann surfaces. We prove a rst
version of this theorem in 2.3. A second version is proved in 2.6; for that we need
to introduce the concepts of line and vector bundles on Riemann surfaces (2.4) and
prove another important theorem, the Serre Duality theorem, which relates certain
spaces of functions and forms. We show the power of these result by giving a few
applications in the rest of the sections of the chapter. In particular we proved that
there exists only one Riemann surface structure on the Riemann sphere (2.6.8); a
generalisation of this result says that any surface can be considered as a subset of
a projective space (2.7). Using the Riemann-Roch and Serre Duality theorems we
construct in 2.9 certain higher dimensional manifolds, called Jacobian varieties,
which are a generalisation of tori (1.3.6), and a mapping from a Riemann surface to
its Jacobian variety, the Abel-Jacobi mapping. We will use this mapping to show
that any surface of genus 1 is indeed is a torus.
2.1. Divisors
This short section introduces the concept of divisor on a compact Riemann sur-
face X: a divisor D is simply a nite set of points of X with certain integral
weights. We associate two sheaves of vector spaces to any divisor, consisting of
meromorphic functions and forms, whose poles and zeroes have orders bounded by
D. The Riemann-Roch theorem, proved in a later section, relates the dimensions of
these two spaces.2.1. DIVISORS 59
2.1.1. Definition. A divisor D on a compact Riemann surface X is a formal
nite sum of the type
D =
m

j=1
n
j
p
j
,
where n
j
Z, m is a non-negative integer and p
j
are points of X.
The set of divisors on X, Div(X), is an abelian group where the operation is the
natural formal addition:
(n
1
p
1
+ + n
m
p
m
) + (r
1
q
1
+ + r
s
q
s
) = n
1
p
1
+ + n
m
p
m
+ r
1
q
1
+ + r
s
q
s
.
In another terminology, a divisor is an element of the free abelian group generated
by the points of X.
For a point p in X we will denote by D(p) the value of D at p, namely D(p) = n
j
if p = p
j
, for some j = 1, . . . , m and D(p) = 0 otherwise.
The zero divisor (m = 0) D = 0 is dened by D(p) = 0 for all p X. It is the
identity element of Div(X).
2.1.2. Definition. A divisor D is called eective, denoted by D 0, if
n
j
0, for all j = 1 . . . , m. We write D
1
D
2
, for divisors D
1
and D
2
, if D
1
D
2
is eective.
2.1.3. If f : X

C is a non-identically zero meromorphic function we can
assign to it a divisor div(f) by considering the order of f at the points of X. More
precisely, given a local coordinate z, dened on a neighbourhood of p, the function
f has a Laurent power series expansion of the form f(z) =

j=n
c
j
(z z(p))
j
,
with c
n
= 0. We set ord
p
(f) = n, and dene div(f) =

pX
ord
p
(f) p. Observe
that ord
p
(f) = 0 only when p is a zero or a pole of X, so this sum is nite (X is
compact). If p is a zero of f, then ord
p
(f) is the order of that zero. For the case of
a pole one has to be slightly careful; ord
p
(f) is a negative number, the opposite of
what is usually known as the order of a pole (1.1.8). To be more precise, consider
for example a polynomial p(z) = a(z z
1
) (z z
n
) dened on

C. Its divisor is
given by D = div(p) = z
1
+ +z
n
n (the sum here takes place in the group of
divisors, not in the complex plane!) So D() = n, although one usually speaks of
as a pole of order n. We hope this small matter of terminology does not confuse60 2. COMPACT RIEMANN SURFACES
the reader. We have that the divisor of a meromorphic function is the (formal) sum
of the zeroes and poles of the function, counted with multiplicities (and signs!).
We dene in a similar way the divisor of a meromorphic 1-form . If this form is
given locally by = fdz, we set ord
p
() = ord
p
(f), for p a point in the domain of
denition of the coordinate z. If t is another coordinate (dened in a neighbourhood
of p) and = gdt, then f = g
dt
dz
. Recall that
dt
dz
means
(tz
1
)
z
(z(q)), and therefore
dt
dz
= 0, since z and t are local coordinates. This shows that ord
p
(f) = ord
p
(g); or,
in other words, the order of at a point p, ord
p
(), is well dened. We can then
dene the divisor of by div() =

pX
ord
p
() p.
2.1.4. Definition. Two divisors D
1
and D
2
are linearly equivalent, D
1

D
2
, if there exists a meromorphic function f such that (f) = D
1
D
2
.
Consider on the Riemann sphere two divisors D
1
= z
1
and D
2
= z
2
, where z
1
and z
2
are points of

C. If z
1
= and z
2
= then the divisor of the function f(z) =
zz
1
zz
2
is equal to div(f) = z
1
z
2
, so D
1
D
2
. If z
1
= we take f(z) =
1
zz
2
, while we
consider the function f(z) = z z
1
if z
2
= . In any of these two last cases we
also have that div(f) = D
1
D
2
. Thus we have proved that D
1
and D
2
are linearly
equivalent. See the comments after the next denition for a more general statement.
If
1
and
2
are two meromorphic forms it is easy to see that the quotient
1
/
2
is a meromorphic function on X: suppose that
j
= f
j
dz on some local coordinate
z, where f
j
are meromorphic functions (j = 1, 2). Then we have

1

2
=
f
1
f
2
. If we have
another local coordinate, say t (whose domain of denition intersect that of z), then

j
= g
j
dt, and the relation f
j
= g
j
dt
dz
holds. We have
f
1
f
2
=
g
1
(dt/dz)
g
2
(dt/dz)
=
g
1
g
2
,
so
1
/
2
does not depend of the choice of local coordinates. This shows that the
divisors of any two meromorphic forms are linearly equivalent.
2.1.5. Definition. The degree of a divisor D =

m
j=1
n
j
p
j
is dened by the
integer
deg (D) =
m

j=1
n
j
.2.1. DIVISORS 61
It is easy to prove that two divisors on the Riemann sphere are linearly equivalent
if and only if they have the same degree (a generalisation of the above argument);
see exercise 42.
2.1.6. A meromorphic function has as many zeroes as poles, counted with mul-
tiplicities (1.3.11) and therefore deg (div(f)) = 0. It follows that linearly equivalent
divisors have the same degree. In particular, the divisors of two meromorphic forms
have the same degree (see 2.5.19).
Definition. A divisor is called principal if it is the divisor of a meromorphic
function, and canonical if it is the divisor of a meromorphic 1-form.
The divisor class group is the quotient of the group of divisors by the subgroup
of principal divisors, Div(X)/Div
P
(X). The canonical class is the class of any
canonical divisor in the divisor class group. Since the degree of a canonical divisor
does not depend of the form we have that the degree of the canonical class is well
dened; in 2.5.19 we will show that this degree is equal to 2g2, where g is the genus
of X. We will denote by K
X
(or K) both the canonical class and a representative
of it.
2.1.7. We associate a sheaf of meromorphic functions to a divisor D by setting
O(D)(U) = {f : U

C; f meromorphic and div(f) D on U or f 0},
where U is an open set of X. For the zero divisor, D = 0, we will simplify notation by
writing O (instead of O(0)). We have H
0
(X, O) is the set of holomorphic functions
on X; since X is compact, any holomoprhic function is constant, and thus we have
that H
0
(X, O) is isomorphic to C.
If D(x) 0, and div(f) D, then f can have a pole of order (as pole!) at most
D(x) at the point x. On the other hand, if D(x) < 0, f must have a zero of order
at least D(x) at x. For example, if z
0


C, and D = z
0
, the set O(D)(

C) consists of
the holomorphic functions on

C (i.e. the constant functions), and the meromorphic
functions with a simple pole at z
0
. These functions form a vector space of dimension
2; a basis is given, for example, by {f
1
1, f
2
(z) = 1/(z z
0
)}, when z
0
= , and62 2. COMPACT RIEMANN SURFACES
{f
1
1, f
2
(z) = z} if D = .
Proposition. If deg (D) < 0 then H
0
(X, O(D)) = 0.
Proof. Since div(f) has 0 degree and deg (D) is positive we cannot have
div(f) D.
We associate a sheaf of meromorphic forms on X to the divisor D by the expres-
sion
(D)(U) = { M
1
(U); div() D},
for U X open. We will simplify notation and write (U) when we consider the
zero divisor. The space H
0
(X, ) is the set of holomorphic 1-forms of X.
2.1.8. One can generalise denition 2.1.1 to non-compact surfaces in the fol-
lowing way. A divisor on a Riemann surface X is a map D : X Z such that for
any compact subset K of X, the set
{x X; D(x) = 0}
if nite. Clearly both denitions agree if X is compact.
2.2. Dolbeaults Lemma and Finiteness Results
The solutions of the Cauchy-Riemann equations, f
z
= 0, are the holomorphic
functions. In this section we study the inhomogeneous equation, f
z
= g, where g
is a smooth function, and show that it always has a solution. In the second part
of the section we use this fact to prove that the cohomology groups H
n
(X, O) are
nite dimensional vector spaces (for X compact and n = 0, 1).
Dolbeaults Lemma
2.2.1. We start with the case of g having compact support.
Theorem. Let g : C C be a smooth function with compact support. Then
there exists a smooth function f : C C, such that f
z
= g.
Proof. Dene f by the integral
f(z) =
1
2i
_
C
g(w)
w z
dwd w.2.2. DOLBEAULTS LEMMA AND FINITENESS RESULTS 63
We rst need to prove that f is well dened; that is, the integral is nite. Write
w = z + re
i
. Then
f(z) =
1
2i
_
C
g(z + re
i
)
re
i
r dr d =
1

_
C
g(z + re
i
) e
i
dr d.
In the above computation we have used the identity dw d w = 2i r dr d. The
function g has compact support so it is bounded and therefore the last integral in
the above expression is nite. This shows that f is well dened; moreover we can
interchange derivatives and integration when computing

f. Using this one easily
sees that
f
z
(z) =
1

_
C
g(z + re
i
)
z
e
i
dr d.
Write = re
i
; then
f
z
(z) =
1
2
lim
r0
_
||r
g(z + )
z
r

dr d.
Since the variables and z are interchangeable in the function g(z + ) we have
g
z
(z + )
1

=
g

(z + )
1

_
g(z + )

_
.
Substituting this identity in the previous integral we obtain
f
z
(z) =
1
2i
lim
r0
_
||0

_
g(z + )

_
d d

=
1
2i
lim
r0
_
||r
d
_
g(z + )

d
_
=lim
r0
_
||=r
g(z + )
2i
d.
In the last equality we have used Stokes theorem (1.4.22). The change of sign in the
integral is due to the orientation of the boundary of the integration domain when
applying Stokes theorem. If we put = re
i
in the last integral, we get
f
z
(z) =
1
2
lim
r0
_
=2
=0
g(z + re
i
) d.
The last integral is nothing but the average of the function g on the circle of centre
z and radius r. Since g is uniformly continuous these averages converge to the value
of g at the point z (see exercise 43 for hints). In other words
f
z
(z) = g(z).64 2. COMPACT RIEMANN SURFACES
2.2.2. By covering C with an increasing sequence of discs we can remove form
the previous result the condition of compactness of the the support of g.
Theorem (Dolbeault). Let g : C C be a smooth function. Then there exists
a smooth function f : C C with f
z
= g.
Proof. As remarked before the statement of the theorem the idea of the proof
is to cover C by an increasing sequence of compact sets, on which we can solve the
inhomogeneous Cauchy-Riemann equation; the limit of these compact solutions
will be the desired function.
We start by setting B
n
= {z C; |z| < n}. Since C is a normal space ([16,
Theorem 2.3 in pg. 198 and Urysohn Lemma in pg. 207]) there exist smooth
functions
n
, with compact support in B
n+1
, such that
n
1 on the disc {z
C; |z| < n+
1
2
}. (Observe that this disc contains the compact set B
n
). Let g
n
=
n
g.
These functions have compact support, and therefore we can nd smooth functions
h
n
, with support contained in B
n+1
, such that

h
n
= g
n
(in B
n+1
or C). The
functions h
n+1
h
n
are holomorphic on a neighbourhood of B
n
, so they have a
power series expansion that converges uniformly on B
n
. Truncating that power
series we get a polynomial p
n
, such that |(h
n+1
h
n
p
n
)(z)| < 2
n
, for all z in B
n
(this is a trivial case of Runges Theorem, [20, theprem 113.7, pg. 290]). Dene a
function f by the expression
f = h
n
+
_

m=n
(h
m+1
h
m
p
m
)
_
(p
1
+ + p
n1
) .
It is easy to see (exercise 44) that the function f is well dened. The right hand
side of the above equality is a uniformly convergent series of holomorphic functions,
so f is holomorphic and satises

f =

h
n
= g on X
n
.
Observe that solutions of the equation

f = g are not unique: if f is one solution
so will be f + h, where h is a holomorphic function (since

h = 0).
2.2.3. Corollary (of the proof). The above theorem holds for the case
of a disc; that is, when g is dened on {z C; |z| < R} (R > 0).
Actually the result holds for any open set of the complex plane, see [19].2.2. DOLBEAULTS LEMMA AND FINITENESS RESULTS 65
2.2.4. Suppose U is a domain of C biholomorphic to a disc; that is, there exists
a disc D
R
= {z C; |z| < R} and a biholomorphic mapping f : U D
R
. Then
the inhomogeneous Cauchy-Riemann equation can be solved on U for any smooth
function g : U C. The proof, which is a simple application of the formul 1.1.3,
is left as an exercise to the reader.
2.2.5. Corollary. The equation f = g has solution in C for any smooth
function g.
Proof. Let h and k be smooth functions such that h
z
= g and k
z
=

h. Then
f =
i
2
k satises f = g (use the identity = 2 i

)
2.2.6. Using Dolbeaults lemma we can compute the dimension of certain co-
homology groups.
Theorem. If X = C or X = {z C; |z| < R} then H
1
(X, O) = 0.
Proof. Let (f
jk
) Z
1
(U, O) for some covering U = {U
j
}
jJ
. Since holomorphic
functions are smooth and H
1
(X, S) = 0 (1.5.12) we have functions g
j
S(U
j
), such
that f
jk
= g
j
g
k
. This implies that

g
j
=

g
k
, so there exists a (smooth) function h
on X satisfying

g
j
= h|
U
j
. Let g be a solution of

g = h, and set f
j
= g
j
g. Then

f
j
= 0, i.e. f
j
is holomorphic, and f
jk
= f
j
f
k
. This shows that any arbitrary
cocycle splits and therefore H
1
(X, O) = 0.
Because of 2.2.4 the above theorem also holds if X is biholomorphic to a disc.
2.2.7. Corollary. For a Riemann surface X the following sequence is exact:
0 O S

S
(0,1)
0.
Proof. Since the kernel of the operator

is the set of holomorphic functions,
exactness at the rst step is satised. The question at the next step is local; we
need to show that any form of type (0, 1) is locally exact. Thus we can assume that
X is a disc in the complex plane. If is of type (0, 1) (in a disc), then = g d z,66 2. COMPACT RIEMANN SURFACES
for some smooth function g. Clearly the form is exact if and only if we can nd
a function f such that

f = g. But this is simply Dolbeaults lemma (2.2.3).
Finiteness Results
2.2.8. The sheaf O on a Riemann surface X has the structure of a vector space
(i.e. O(U) are vector spaces) and therefore the cohomology groups H
n
(X, O) are
vector spaces. In the next paragraphs we will show that these spaces are nite
dimensional, for n = 0, 1 and X a compact surface. The case of n = 0 is just a
simple consequence of the Maximum Modulus Principle (1.1.9 and 1.3.8).
Theorem. For a compact Riemann surface X, H
0
(X, O) = C.
2.2.9. The proof for H
1
(X, O) for the case when X is the Riemann sphere is
an easy consequence of basic results of Complex Analysis.
Theorem. H
1
(

C, O) = 0.
Proof. Consider the standard covering of

C given by U
1
= C and U
2
=

C\{0}
(1.3.5). Both sets U
1
and U
2
are biholomorphically equivalent to the complex plane,
so H
1
(U
j
, O) = 0, j = 1, 2; or, in other words, the covering {U
1
, U
2
} is a Leray
covering for the sheaf O. If we take a cocycle, (f
jk
)
j,k=1,2
(with coecients in the
sheaf O), since f
11
= f
22
= 0 and f
12
= f
21
, the only function we need to consider
is f
12
. Let f
12
(z) =

a
j
z
j
be a Laurent power series representation of this
function. The functions f
1
(z) =

j=0
a
j
z
j
and f
2
(z) =

j=1
a
j
z
j
are elements
of O(U
1
) and O(U
2
), respectively, satisfying f
12
= f
1
f
2
.
2.2.10. Theorem. If X is a compact Riemann surface then H
1
(X, O) is nite
dimensional.
Proof. Let D
r
denote the disc in the complex plane centred at 0 and radius
r, where r > 0. For every point p X there exists a coordinate patch (U, z), such
that z
p
(p) = 0 and z
p
(U) = D
2
. The collection {z
1
p
(D
1/2
)}, as p varies over all
the points of X and z is as above, forms an open covering of X and therefore there
exists a nite subcover W. To simplify notation we will write z
1
, . . . , z
n
for the
local coordinates of this subcover. Thus W = {W
j
= z
1
j
(D
1/2
)}
n
j=1
is our covering.2.2. DOLBEAULTS LEMMA AND FINITENESS RESULTS 67
The collections V = {V
j
= z
1
j
(D
1
)}
n
j=1
and U = {U
j
= z
1
j
(D
2
)}
n
j=1
are also open
coverings of X. Moreover, we have W < V < U. Observe that W
j
(respectively V
j
)
has compact closure contained in V
j
(respectively U
j
). Since the open sets of these
three coverings are biholomorphically equivalent to discs (in C) we have that W, V
and U are Leray coverings of X for the sheaf O.
Let (f
jk
) be an element of C
0
(U, O), and denote by g
jk
the restriction of f
jk
to
V
jk
. The functions g
jk
are bounded, since V
jk
has compact closure contained in U
jk
.
Assume that (g
jk
) splits, that is, g
jk
= g
j
g
k
. We claim that g
j
is bounded in V
j
.
To prove it consider a point p of the boundary of V
j
; then there exists an open set V
l
such that x V
l
. Let A be a neighbourhood of x with

A compact, contained in V
l
.
Then the functions g
l
and g
jl
are bounded in AV
j
V
l
, which is a neighbourhood
of x. Since c
j
= g
jl
+c
l
we have that c
j
is bounded in that neighbourhood. In other
words, every point of V
j
has a neighbourhood in V
j
where c
j
is bounded. By the
Maximum Modulus Principle (1.1.9) the function c
j
is bounded in V
j
.
The above argument shows that to compute H
1
(X, O) we can use bounded
cochains. The advantage of this is that the space of bounded cochains is a Banach
space. More precisely, if we consider = (f
j
) C
0
(V, O), where the functions f
j
s
are bounded, we can dene a norm by |||| =

n
j=1
||f
j
||

(||f
j
||

= sup{|f
j
(z)|; z
V
j
}). A similar denition gives us a norm in the space of 1-cochains. In what follows
we will assume all cochains to be bounded.
We now dene two mappings
: C
0
(W, O) C
1
(W, O) and : Z
1
(V, O) Z
1
(W, O)
as follows. The mapping is the coboundary homomorphism while is the mapping
induced by restriction (recall that W < V). Both mappings are clearly continuous
with respect to the topology induced by the norms introduced above. Moreover by
Montels theorem (1.1.13) we have that is a compact mapping.
Set:
: C
0
(W, O) Z
1
(V, O) Z
1
(WO)
(, ) () + ().68 2. COMPACT RIEMANN SURFACES
Since H
1
(X, O) = H
1
(W, O) = H
1
(V, O) the mapping is surjective. It follows
that is surjective as well.
The mapping
: C
0
(W, O) Z
1
(V, O) Z
1
(W, O)
(, ) ((, )) ((, )) = ().
is the dierence of a surjective mapping and a compact one. Its image is the space
of coboundaries, B
1
(, W, O). The quotient Z
1
(W, O)/B
1
(W, O) is nothing but
H
1
(X, O); by Schwartzs theorem (1.1.25) it has nite dimension.
2.2.11. Definition. If X is a compact Riemann surface, the dimension of
H
1
(X, O) is called the arithmetic genus of X.
2.3. The Riemann-Roch Theorem
Given a divisor D on a surface we have associated to it the shear O(D) consisting
of meromorphic functions whose zeroes and poles have orders bounded by D. If
this sheaf is not empty for certain divisors then we can easily prove the existence of
non-constant meromorphic functions on surfaces. It is thus of interest to know the
dimension of the cohomology groups of O(D). The Riemann-Roch theorem proves
that the dierence dim H
0
(X, O(D))dim H
1
(X, O(D)) is a constant that depends
only on the degree of D and the arithmetic genus of X. We prove this result here
and use it to show the existence of meromorphic functions on compact surfaces by a
proper choice of the divisor D. We also prove that any surface of arithmetic genus
0 must be biholomorphic to the Riemann sphere

C. In 2.6 we will show that the
arithmetic genus of a compact surface is equal to its topological genus; these two
facts show that the Riemann sphere admits only one Riemann surface structure (up
to biholomorphisms).
2.3.1. Let X be a compact surface x a point of it. The skyscraper sheaf C
p
is dened, on an open set U X, by
C
x
(U) =
_

_
C, if p / U,
0, otherwise,2.3. THE RIEMANN-ROCH THEOREM 69
with the obvious restriction mappings. Clearly H
0
(X, C
p
) = C
p
(X) = C. To
compute the rst cohomology group consider an open covering of X, say U = {U
j
}
j
.
Choose an index j
0
such that p U
j
0
, and dene a new covering of X, V = {V
j
}
j
by setting
V
j
=
_

_
U
j
0
, if j = j
0
,
U
j
\{p}, otherwise.
Since p does not belong to V
j
V
k
, for j = k, we have H
1
(V, C
p
) = 0. The natural
mapping H
1
(U, C
p
) H
1
(V, C
p
) is injective; therefore H
1
(U, C
p
) = 0. Since U is
an arbitrary covering of X we obtain H
1
(X, C
p
) = 0.
2.3.2. Let D be a divisor on X. Fix a point p X; we construct a sequence
(4) 0 O(D) O(D + p)

C
p
0
as follows. If f O(D), then (f) D D p, so O(D) is a subsheaf of
O(D + p). To dene , consider rst the case of a an open set U such that p / U;
then C
p
(U) = 0 so 0. On the other hand, if p U, let f be a meromorphic
function with f O(D+p)(U); choose a local coordinate z around x with z(p) = 0.
The function f will have a power series expansion of the form f(z) =

j=n1
a
j
z
j
near p. We set (f) = a
n1
, where D(p) = n (see the remark below). If C
p
(U) is
not trivial then the kernel of consists precisely of the functions that have a pole
of order at most n at p; that is, O(D)(U). On the other hand, if C
p
(U) = 0, then
x / U, so D|
U
= (D + p)|
U
(both divisors have the same values at all points of U)
and O(D)(U) = O(D + p)(U). Thus (4) is exact.
Remark. It can be easily checked that if w is another local coordinate on X
with w(p) = 0, and we write f(w) =

j=n1
b
j
w
j
, then b
n1
= a
n1
. Thus the
mapping above is well dened.
2.3.3. Theorem (Riemann-Roch). Let X be a compact Riemann surface and
D a divisor on X. Then the spaces H
0
(X, O(D)) and H
1
(X, O(D)) are nite di-
mensional and
dim H
0
(X, O(D)) dim H
1
(X, O(D)) = 1 g

+ deg (D),70 2. COMPACT RIEMANN SURFACES


where g

is the arithmetic genus of X, i.e. g

= dim H
1
(X, O).
Proof. Fix a point p X as above. The long exact sequence corresponding to
(4) is
0 H
0
(X, O(D)) H
0
(X, O(D + p)) C H
1
(X, O(D))
H
1
(X, O(D + p)) 0.
(5)
We can split (5) in two sequences given by
0 H
0
(X, O(D)) H
0
(X, O(D + p)) V 0
0 W H
1
(X, O(D)) H
1
(X, O(D + p)) 0,
where V is the image of the mapping H
0
(X, O(D +p)) C, and W = C/V . Both
the sequences are clearly exact because of the choice of V and W. Assume now that
the result is true for one of the two divisors, D or D + p. Since V and W have
nite dimension we have that two spaces in each of the above sequence are nite
dimensional, and thus all spaces must have nite dimension. From the exactness of
those sequences we see that
dim H
0
(X, O(D)) + dim V = dim H
0
(X, O(D + p))
dim H
1
(X, O(D + p)) + dim W = dim H
1
(X, O(D)).
These two equations together with the equality dim V + dim W = dim C = 1 give
dim H
0
(X, O(D)) dim H
1
(X, O(D)) =
dim H
0
(X, O(D + p)) dim H
1
(X, O(D + p)) 1
But deg (D) = deg (D + p) 1, so the above identity is equivalent to the following
dim H
0
(X, O(D)) dim H
1
(X, O(D)) deg (D) =
dim H
0
(X, O(D + p)) dim H
1
(X, O(D + p)) deg (D + p).
Since we are assuming that the theorem is true for one of the two divisors, one side
of the above equation is equal to 1 g

, and so is the other side.


To complete the proof we need three observations:
(1) if D is the zero divisor then dim H
0
(X, O) = dim C = 1 and dim H
1
(X, O) =
g

, so the theorem holds in this case.


(2) From the above arguments we have that if the theorem is true for a divisor2.4. LINE BUNDLES AND DIVISORS 71
E then it is also true for the divisors E +p and E p, where p is an arbitrary point
of X.
(3) Starting with the zero divisor we can obtain any divisor in a nite number
of steps by adding and subtracting points.
2.3.4. Corollary. If X is a compact Riemann surface, then there exists a
non-constant meromorphic function f : X

C.
Proof. Consider the divisor D = (g +1) p, for some point p of X. The constant
functions form a one-dimensional subspace of H
0
(X, O(D)). On the other hand,
from the Riemann-Roch theorem we obtain
dim H
0
(X, O(D)) 2,
so there must exist a non-constant function in O(D).
2.3.5. Corollary. If X is a compact surface of arithmetic genus 0 then X
is biholomorphic to

C.
Proof. From the previous result we have that if D = p then there exists a
non-constant meromorphic function f : X

C, with f H
0
(X, O(D)). Because
of the choice of D the function f can have at most a simple pole at p. On the other
hand, since f is not constant, it must have a pole at p (the only allowed singularity).
Thus f has degree 1 (see 1.3.11); that is, it is biholomorphic.
2.4. Line bundles and Divisors
The purpose of this section is to introduce the concept of vector bundles on
Riemann surfaces and prove some results that we need in later sections. Informally
speaking, a vector bundle is just a way of assigning a vector space to each point of
a surface such that, as we move holomorphically on the surface the corresponding
vector spaces change also holomorphically. The space of line bundles (vector
bundles of dimension 1) is called the Picard group of the surface; we show that this
group is isomorphic to H
1
(X, O

). Using vector bundles we give a new proof of


the existence of non-constant meromorphic functions on compact surfaces. We will
also show that there is a natural way of associating a line bundle to divisors on a72 2. COMPACT RIEMANN SURFACES
surface X. We will study the relationship between the sheaf O(D) and isomorphic
line bundles (we need those results for the next section).
Vector Bundles on Riemann surfaces
2.4.1. We begin this section by extending the notions of Riemann surfaces and
holomorphic mappings to the setting of several complex variables. Assume is an
open set of C
n
and f : C
m
is a continuous function. Write (z
1
, . . . , z
n
) for the
standard coordinates in C
n
, and f = (f
1
, . . . , f
m
). We say that f is holomorphic
if each f
j
is holomorphic on each variable z
k
, for j = 1, . . . , m and k = 1, . . . , n.
If each f
j
has partial derivatives, then f is holomorphic if and only if it satises
individual Cauchy-Riemann equations, i.e. f
j
/z
k
= 0.
Let Y be a 2n dimensional manifold; that is, every point of Y has a neigh-
bourhood homeomorphic to an open set of R
2n
. We can identify R
2n
with C
n
in a
standard way, (x
1
, x
2
, . . . , x
2n1
, x
2n
) (x
1
+ ix
2
, . . . , x
2n1
+ ix
2n
); it makes sense
then to talk about holomorphic changes of coordinates and complex structures, in
a way parallel to the denitions of 3. A complex manifold is a (topological)
manifold with a complex structure. It is natural to say that Y has complex dimen-
sion n (and real dimension 2n). From this approach, a Riemann surface is just a
complex manifold of (complex) dimension 1 (or in another terminology, a complex
curve). We leave it to the reader to dene holomorphic mappings between complex
manifolds and study their basic properties.
2.4.2. Definition. A holomorphic vector bundle of rank n on a Rie-
mann surface X is a complex manifold V , together with a surjective holomorphic
mapping : V X, such that for all p X, the bre V
p
=
1
(p) has the structure
of a complex vector space of (complex) dimension n, and is locally trivial in the
following sense:
(1) for every point p X there exists a neighbourhood U, and a biholomorphic2.4. LINE BUNDLES AND DIVISORS 73
mapping
U
:
1
(U) U C
n
, such that the following diagram is commutative:

1
(U)

U
//

##
F
F
F
F
F
F
F
F
F
U C
n
p
1
{{x
x
x
x
x
x
x
x
x
x
U
Here p
1
is the projection in the rst factor, p
1
(p, w) = p.
(2) Let p, U and
U
be as in (1). Dene a mapping
p
: V
p
C
n
by
U
(v) =

U
(p,
p
(v)). Then
p
is linear isomorphism of vector spaces.
2.4.3. A trivial example of a bundle is given by the product XC
n
X with
the projection on the rst coordinate. The isomorphism
X
is given by the identity
function. Any bundle is locally like the trivial bundle (condition (1) in the above
denition). So in some sense a bundle consists of many copies of the trivial bundle,
dened on open subsets of X and glued together by certain transition functions
(see 2.4.6 below) in the same way that a Riemann surface is a collection of open sets
of C related by holomorphic changes of coordinates.
2.4.4. Observe that the integer n in denition 2.4.2 is the same for all the
bres. The complex dimension of the manifold V is n + 1. A bundle of rank 1 is
called a line bundle.
By an abuse of notation some times we will say that V is a (vector) bundle on X
(although the projection and the mappings
U
are part of the denition).
Since our interest is in Complex Analysis, we have used holomorphic functions and
complex manifolds in 2.4.2. One can dene in a similar way continuous or smooth
bundles. All bundles in this book are assumed to be holomorphic.
2.4.5. The functions
U
are called local trivializations of the bundle. From
condition (1) we get that these functions are of the form
(3)
U
(p) = ((p),
U
(p)), p
1
(U),
where
U
:
1
(U) C
n
.
The Picard Group74 2. COMPACT RIEMANN SURFACES
2.4.6. There is a natural correspondence between bundles and certain cocycles
on X that we explain next for the case of line bundles, since this is the setting we
are interested in. Let L

X be a line bundle and U = {U
j
}
jJ
an open covering
of X; assume there are local trivializations
j
=
U
j
:
1
(U
j
) U
j
C dened
on the sets U
j
s. If U
j
U
k
= the mapping
k

1
j
, dened on (U
j
U
k
) C,
should be the identity in the rst variable (by (1) in 2.4.2 or (3) above) and a linear
isomorphism in the second; that is,

k

1
j
: (U
j
U
k
) C (U
j
U
k
) C
(p, v) (p, T
p
(v)).
Since T
p
is an isomorphism of 1-dimensional vector spaces it must be of the form
v
p
v, where
p
is a non-zero complex number. We then obtain holomorphic
functions, g
jk
: U
j
U
k
C

, dened by g
jk
(p) =
p
. These functions are called the
transition functions of L. This procedure associates to a line bundle L an element
(g
jk
) of the group of cochains C
0
(U, O

). For convenience we will use multiplicative


notation for this sheaf; in this way, a cocycle satises f
jk
f
kl
f
lj
1. With this
notation in mind, the cochain (g
jk
) is actually a cocycle, since
(x, w) = (
j

1
k

k

1
l

l

1
j
)(x, w) = (x, g
jk
(x) g
kl
(x) g
lj
(x)w).
So we have an assignment L (g
jk
) of a element of Z
1
(U, O

) to each line bundle


on X, which has a given trivialization on each U
j
U.
2.4.7. For the converse construction we start with a cocycle (g
jk
), and take
the disjoint union L
1
=

j
U
j
C. We dene an equivalence relation in L
1
as
follows: for (p, v) U
j
C and (q, w) U
k
C we set (p, v) (q, w) if p = q and
g
jk
(v) = w. The quotient space L = L
1
/ is a line bundle on X with transition
functions (g
jk
) (exercise 47).
2.4.8. These two constructions give us a correspondence between line bundles
and cocycles (with coecients in O

) on X. However one would like to actually


have a relation involving the cohomology group H
1
(X, O

) rather than groups of


cocycles. For that purpose we need to consider isomorphism classes of line bundles.
Definition. A holomorphic mapping F : V W between two vector bundles,2.4. LINE BUNDLES AND DIVISORS 75

1
: V X and
2
: W X on a Riemann surface X, is called a (bundle)
morphism if
2
F =
1
, and F|

1
1
(p)
:
1
1
(p)
1
2
(p) is a linear mapping, for
all p X. A morphism is an isomorphism if there exists a morphism F
1
: W V ,
such that F F
1
= Id
W
and F
1
F = Id
V
.
Two bundles are isomorphic if there exists an isomorphism between them.
2.4.9. Definition. A (holomorphic) vector bundle E

X of rank n is called
trivial if it is isomorphic to the bundle X C
n
p
1
X.
2.4.10. As explained above, our aim is to identify isomorphism classes of line
bundles on X with H
1
(X, O

). Since this last set is a group, the space of (isomor-


phism classes of) line bundles will get a group structure from such an identication.
However there exists an operation on line bundles that will turn out to be the right
one under the isomorphism we are looking for (namely, the bijection between the
set of line bundles and the cohomology group will be a group homomorphism). We
explain that operation next: let L and

L be two line bundles on X, and take an
open covering U = {U
j
}
jJ
, such that there are local trivializations on U
j
for both
bundles. Let (g
jk
) and ( g
jk
) denote the transition functions (with respect U) of L
and

L, respectively. The product (g
jk
g
jk
) is an element of Z
0
(U, O

) (recall that we
are using multiplicative notation for the sheaf O

) and thus it denes a line bundle


on X, called the tensor product of L and

L, denoted by L

L.
The identity element (in the set of line bundles) is given by the bundle correspond-
ing to the trivial cocycle; that is, h
jk
1. An easy computation shows that the
bundle constructed from this cocycle is just the trivial bundle X C
p
1
X (with
the notation of 2.4.7, the relation in L
1
is the trivial one).
2.4.11. The inverse element of L will be the line bundle dened by the cocycle
(1/g
jk
). Observe that these are well dened functions because g
jk
takes values in
C

. The bundle constructed from (1/g


jk
) is called the dual bundle of L, written
as L
1
.
Proposition. L L
1
is trivial.
Proof. The proof is clear from the correspondence between line bundles and
cocycles.76 2. COMPACT RIEMANN SURFACES
2.4.12. A right inverse of the projection : L X is a holomorphic way of
assigning to every point of X an element of L lying on the bre of that point.
Definition. A (holomorphic) section of a line bundle L

X is a holomorphic
mapping s : X L such that s = Id
X
.
If U = {U
j
} of X is a covering with local trivializations
j
and s a section of L,
we have (
j
s)(p) = (p, f
j
(p)), where f
j
: U
j
C are holomorphic functions. It
is easy to see that f
j
= g
jk
f
k
. Conversely, any set of functions f
j
satisfying these
identities denes a section of L X in the obvious way.
If L and

L are two line bundles, with sections s and s, given by (f
jk
) and (

f
jk
)
respectively; then f
j

f
j
induces a section of L

L, denoted by s s.
By an abuse of notation we will denote by L(U) the sheaf of holomorphic sections
of the bundle L dened on the open set U, and by L the space of global sections,
i.e. L(X).
2.4.13. Since the bres L
p
(for p a point of X) are vector spaces it makes sense
to talk of the zero vector of L
p
. Under a local trivialization
U
:
1
(U) UC the
zero vector (of L
p
) is mapped to (p, 0); in other words, the zero vector is
1
U
(p, 0).
It is easy to check that the image of this zero vector does not depend on the choice
of
U
.
Proposition. A line bundle L X is trivial if and only if it has a section
s : X L such that s(p) is never equal to the zero vector of L
x
, for any p X.
Proof. If s is a nowhere zero section of L X we have that the mapping
F : X C L, given by F(p, ) = s(p), is a bundle isomorphism.
Conversely, if F : X C L is an isomorphism, the section dened by s(p) =
F
1
(p, 1) is nowhere zero.
For the case of the trivial line bundle X C X, a section that never vanishes is
given by s(p) = (p, 1) (actually we can choose any non-zero complex number instead
of 1).
Remark. The above proposition generalises to the case of vector bundles of
rank greater than 1 in a natural way: a bundle of rank n, say V X, is trivial if2.4. LINE BUNDLES AND DIVISORS 77
and only if there exist n holomorphic sections, s
j
: X V , j = 1, . . . , n, such that
{s
1
(p), . . . , s
n
(p)} form a basis of V
p
, for all p X. In the case of line bundles this
condition is clearly the same as the previous result, since a vector in C is a basis if
and only if it is not the zero vector.
2.4.14. Let L and

L be two line bundles, F : L

L an isomorphism. Using
local trivializations we obtain functions f
j
: U
j
C

, satisfying g
jk
= (f
j
/f
k
) g
jk
.
This implies that the cohomology classes of the transition functions in H
1
(X, O

)
are equal, i.e. [(g
jk
)] = [( g
jk
)] H
1
(X, O

). So we have an identication between


this cohomology group and the Picard group Pic(X), the group of isomorphism
classes of line bundles on X.
Proposition. The mapping Pic(X) H
1
(X, O

), that sends the isomorphism


class of a line bundle to the cohomology class of the corresponding transition func-
tions, is a group isomorphism (with the tensor product as the operation on line
bundles).
Proof. The proof follows from the following two points, that are left as exercises
to the reader:
1. if L
1
and L
2
are two line bundles, isomorphic to L

1
and L

2
respectively, then
L
1
L
2
is isomorphic to L

1
L

2
;
2. let L be a line bundle, and U and V are two open coverings of X for which L has
local trivializations. Let (g
jk
) and ( g
jk
) be the elements of Z
1
(U, O

) and Z
1
(V, O)
constructed from these trivializations as in 2.4.6. Then (g
jk
) and (tildeg
jk
) represent
the same cohomology class in H
1
(X, O

).
Divisors and line bundles
2.4.15. The denition of section can be extended to include isolated singu-
larities as follows. Associated to a (holomorphic) section s there are holomorphic
functions f
j
: U
j
C, with respect to some covering {U
j
} of X. If these functions
are meromorphic we say that s is a meromorphic section of L. More formally, a
meromorphic section is a mapping s : X\A L, where A is a discrete subset of X,78 2. COMPACT RIEMANN SURFACES
such that
(1) s = Id
X\A
, and
(2) for every point p of A, there exists a local coordinate z, with z(a) = 0, and an
integer n, such that z
n
s extends to a holomorphic section of L in a neighbourhood
of p.
To dene the divisor of a meromorphic section, set ord
p
(s) = ord
p
(f
j
), if p U
j
,
the order of s at p. Since the transition functions are nowhere zero, the order is
well dened, i.e. it does not depend of the function f
j
. The divisor of s is given
by div(s) =

pX
ord
p
(s) p. As an example, consider the Riemann sphere with the
standard covering {U
1
, U
2
}. We dene a transition function by g
12
(z) = z, and two
functions f
j
: U
j
C by the expressions f
1
(z) = 1, f
2
() = 1/ (f
2
() = 0). The
function g
12
gives a line bundle L

C, and the pair (f
1
, f
2
) induces a section s of
L. The divisor of s consists of just one point, (s) = . Observe that, unlike the
case of meromorphic functions, the degree of this divisor is not 0.
2.4.16. As remarked in 2.4.12, we denote by L the vector space (sheaf) of
(global) holomorphic sections of a bundle L X over X.
Theorem. For a line bundle L

X on a compact Riemann surface X the
space H
1
(X, L) is nite dimensional.
Proof. The proof for the case of the trivial line bundle O (2.2.10) generalises
to any line bundle without any diculty (since all line bundles are locally trivial).
2.4.17. Proposition. Let L X be a line bundle on a compact surface.
Then there exists a meromorphic section of L which is not holomorphic.
Proof. Fix a point p X and let U be a neighbourhood of p, with a triv-
ialization mapping
U
of the bundle L. Assume further that there is a local co-
ordinate z on U satisfying z(p) = 0. For every positive integer k we dene a
meromorphic section s
k
of L on U\{p} by
U
(s
k
(q)) = (q, z(q)
k
). Let U denote
the covering of X consisting of the two open sets U
1
= U and U
2
= X\{p}. We
dene functions f
jk
, for j, k = 1, 2, on U
j
U
k
by f
(k)
12
= s
k
on U
1
U
2
= U\{p},
f
(k)
21
= f
(k)
12
and f
(k)
11
= f
(k)
22
= 0. This gives us an element f
(k)
of Z
1
(U, L). Since
H
1
(U, L) H
1
(X, L) is injective and H
1
(X, L) has nite dimension, say d, we have2.4. LINE BUNDLES AND DIVISORS 79
that
c
1
f
(1)
+ + c
d+1
f
(d+1)
B
1
(U, L),
for some complex numbers (c
1
, . . . , c
d+1
) = (0, . . . , 0). But this is equivalent to say
that there exist holomorphic sections s
j
of L on U
j
, j = 1, 2 satisfying
c
1
s
1
+ + c
d+1
s
d+1
= s
1
s
2
, on U\{p},
for sections s
1
and s
2
of L (on U\{p}). The section s
2
of L on U
2
= X\{p} extends
to a global meromorphic section of L over X; it is not dicult to see that this section
is not holomorphic.
2.4.18. Corollary. If X is a compact surface then there exists a non-
constant meromorphic function.
Proof. Let L X be a line bundle on X. Let p
1
and p
2
be two dierent
points of X; by the above proposition we have two sections of L, say s
1
and s
2
(with
p = p
j
in the above proof), such that each s
j
is holomorphic on X\{p
j
} and has a
pole at p
j
. We have that s
j
0, so there exists a function on X, say f, satisfying
s
1
= f s
2
. It is easy to see that f is non-constant meromorphic function, which will
in fact have a pole at p
1
and a zero at p
2
.
Remark. To complete the above proof we need to show that a (compact) sur-
face has line bundles. But on any Riemann surface there is a natural line bundle
called the canonical line bundle (2.5.12); see also the next subsection.
2.4.19. One can associate a line bundle L(D) to a divisor D in the following
form. Assume D =

n
j=1
c
j
p
j
. Choose coordinate patches (U
j
, z
j
) with z
j
(p
j
) = 0
and U
j
U
k
= , if j = k. Set U
n+1
= X\{p
1
, . . . , p
n
}. We dene meromorphic
functions on U
j
by f
j
= z
c
j
, for j = 1, . . . , n, and f
n+1
1, and construct L(D) as
the line bundle with transition functions given by the cocycle
g
jk
=
_

_
f
j
/f
k
, if U
j
U
k
= ,
1, otherwise.80 2. COMPACT RIEMANN SURFACES
The cochain (f
jk
) denes a section of L(D), called the canonical section and
denoted by s
D
. The divisor of s
D
is precisely D, as one can easily check.
The holomorphic sections of the bundle L(D) can be canonically identied with the
sheaf O(D) by the mapping
O(D)(U) L(D)(U)
f f s
D
.
Since (f) D, the divisor of the section f s
D
is eective; that is, it is a holomorphic
section of L(D). For the case of the zero divisor we have that that the sheaf O
of holomorphic functions on X is isomorphic to the sheaf of sections of the trivial
bundle XC. In the case of X being a compact surface this statement can be proved
directly: a section of the trivial bundle should be given by a function s : X XC
of the form s(p) = (p, f(p)), where f : X C is holomorphic. So f will be a
constant function, say f(p) =
s
, and the isomorphism between sections of X C
and O is given by s
s
.
It is clear from the above constructions that L(D)

is isomorphic to L(D).
2.4.20. Theorem. Any line bundle is isomorphic to the line bundle of a
divisor.
Proof. The theorem follows from 2.4.17. In fact, if s is a non-zero meromorphic
section of L, with div(s) = D, then the section s s
D
is a nowhere vanishing
section of L L(D). Here s
D
denotes the standard section of L(D). Thus
L

= L(D)

= L(D).
2.4.21. Theorem. If L X is a line bundle on a compact surface then
H
0
(X, L) is nite dimensional.
Proof. From the previous result we have that L is isomorphic to the line
bundle L(D) of a divisor D. But then (2.4.19) H
0
(X, L)

= H
0
(X, O(D)), so from
the Riemann-Roch theorem (2.3.3) we have that this space is nite dimensional.
2.4.22. Theorem. Two divisors D
1
and D
2
are linearly equivalent if and only
if the line bundles L(D
1
) and L(D
2
) are isomorphic.2.5. SERRE DUALITY 81
Proof. Let f : X

C be a meromorphic function with div(f) = D
2
D
1
. The
points of L(D
1
) are of the form (p, s
D
1
(p)), for X, in some local trivialization.
Then the mapping
(p, s
D
1
(p)) (p, s
D
2
(p))
denes an isomorphism between L(D
1
) and L(D
2
) (some care has to be taken for
the case of poles and zeroes of s
D
1
; we leave the details to the reader).
Conversely, if F : L(D
1
) L(D
2
) is an isomorphism, then f s
D
1
is a section
of L(D
2
), and thus there exists a meromorphic function f : X

C such that
F s
D
1
= f s
D
2
. Clearly div(f) = D
1
D
2
.
Remark. The above result can be restated as saying that the Picard group is
isomorphic to the divisor class group.
2.4.23. Definition. The degree of a line bundle L is denes as deg (D),
where D is a divisor such that L

= L(D).
By the above results, deg (L) is well dened.
2.5. Serre Duality
In 2.3 we have proved a rst version of the Riemann-Roch theorem, which relates
the dimensions of the spaces H
0
(X, O(D) and H
1
(X, O(D)). In this section we will
prove the Serre Duality theorem, which shows that H
1
(X, O(D)) is isomorphic to
H
0
(X, (D)). In this way we obtain a version of the Riemann-Roch theorem which
has only 0th cohomology groups. We will also see that there exists a line bundle
on a Riemann surface, called the canonical bundle, whose sections are nothing but
forms on the surface.
The proofs of this section are by R.R. Simha [24].
Some results on vector bundles on Riemann sur-
faces82 2. COMPACT RIEMANN SURFACES
2.5.1. Throughout this section V X will denote a vector bundle on a com-
pact Riemann surface. Recall that we also use V for the sheaf of holomorphic
sections of V ; that is, V (U) denotes the group of holomorphic sections of the bundle
on the open set U X.
Proposition. If X is a compact Riemann surface and V X a vector bundle
then H
1
(X, V ) is nite dimensional.
Proof. The proof is similar to the case of line bundles (2.2.10). We only need
the following result: U is an open set of X, biholomorphic to a disc D (in C),
and
U
:
1
(U) U C
n
is a local trivialization of the bundle V

X, then
the cohomology group H
1
(U, V ) is isomorphic to n copies of H
1
(D, O), that is,
H
1
(U, V )

=

n copies
H
1
(D, O), and hence it is trivial.
2.5.2. Corollary. There exists a non-identically zero meromorphic section
of V X.
2.5.3. Proposition. If V X is a vector bundle on X then there exists a
divisor D on X, a vector bundle V

and an exact sequence of the form:
0 L(D) V V

0.
The rank of V

satises rk(V

) = rk(V ) 1.
Proof. Let s : X V be a nonzero meromorphic section of V with divisor D.
The mapping L(D) V , given by s
D
(x) s(x), is an embedding of L(D) into
V . Identify L(D) with its image and dene V

= V/L(D).
2.5.4. Proposition. H
0
(X, V ) is nite dimensional.
Proof. From the sequence in the proposition above we get the following exact
sequence of cohomology groups
0 H
0
(X, L(D)) H
0
(X, V ) H
0
(L, V

).
Assume rk(V ) = 2. Then V

is a line bundle so H
0
(X, L(D)) and H
0
(X, V

) are -
nite dimensional, and hence so if H
0
(X, V ). The general cases follows from induction
on the rank of V .2.5. SERRE DUALITY 83
2.5.5. Corollary. Let L

C be a line bundle on the Riemann sphere. Then
the restriction of L to the complex plane, L|
C
C is trivial.
Proof. We will show that L|
C
is trivial by exhibiting a nowhere zero holomor-
phic section. From the previous result we have a section s :

C L. Consider its divi-
sor div(s) =

m
j=1
n
j
p
j
, and dene a meromorphic function f(z) =

p
j
C
(zp
j
)
n
j
.
Then fs is a section of L|
C
which never vanishes (the function f and the section s
have poles and zeros on C at the same point, but of opposite type, so they cancel).
2.5.6. Given a holomorphic mapping f : X Y between compact surfaces
and a line bundle V X, we can construct a bundle f

(V ) on Y as follows. First
of all, observe that to construct a bundle it suces to give its sections on open sets.
To see this remark, consider the following (easy) result.
Lemma. Let L

X be a line bundle, :
1
(A) AC a local trivialization.
Let x A and C. Then there exists a section of L on A such that s(x) = (x, ).
Proof. Dene s(y) =
1
(y, ).
This shows that there are local sections whose value at a point can be xed. So
there is always a nonzero local section. Thus, as we have claimed above, if we know
local sections, say {s
j
}, with respect to some open covering of X, we can recover
the transition functions {g
jk
} from the identity s
j
= g
jk
s
k
, and therefore the local
trivialisations. We dene the bundle f

(V ) on Y by setting f

(V )(A) = V (f
1
(A));
that is, the sections of f

(V ) on an open set are just the sections of V on the preimage


(under f) of the set V .
To make things a little more precise we consider the case of the function f : A B,
given by f(z) = z
n
= w, where A and B are two copies of the unit disc (we do
not use the notation D for the unit disc since we want to dierentiate between the
domain and the range of the function f). The bundle on A will be the trivial bundle,
denoted by O
A
. We have that f

(O
A
) = O(A), so the sections of f

(O
A
) are just
holomorphic functions on the variable z. Assume n = 2 for the sake of simplicity.
We claim that {1, z} is a basis of f

(O
A
)(B). This means that any section of f

(O
A
)
(on B) is of the form g
1
(w) 1 + g
2
(w) z, for holomorphic functions g
1
(w) and g
2
(w)84 2. COMPACT RIEMANN SURFACES
dened on B. Let h : A C be a holomorphic function (i.e. a section of the trivial
bundle O
A
) and write h(z) =

k=0
a
k
z
k
; then we have
h(z) =a
0
+ a
1
z + a
2
z
2
+ = (a
0
+ a
2
z
2
+ ) + z(a
1
+ a
3
z
2
+ ) =
=(a
0
+ a
2
w + ) + z(a
1
+ a
3
w + ) = g
1
(w) 1 +g
2
(w) z.
The rank of f

(O
A
) is equal to 2. In a more general setting of a holomorphic
function f : X Y and L X a line bundle we can nd a situation where
B Y is (biholomorphic to) a disc and f
1
(B) has more than one component. For
example, f
1
(B) = A
1

A
2
, where A
1
and A
2
are two discs in X. The function f
might even have dierent degree on the sets A
1
and A
2
, for example we could have
f(z) = z
2
and f(z) = z
3
respectively. To study this case, consider again the trivial
bundle over A
1

A
2
, say O
A
; let z
1
denote the section of f

(O
A
) that is equal to z
on A
1
and 0 on A
2
; dene z
2
similarly. Then one sees that a basis of the holomorphic
sections of f(O
A
) (on B) is given by the ve functions {1, z
1
, 1, z
2
, z
2
2
}. From these
two examples the reader should be able to construct f

(L) for the general case of a


line bundle L on X. The case of vector bundles of higher rank is also easy to obtain
from the above remarks.
We have that f

(V ) has rank equal to rk(f

(V )) = rk(V ) deg (f), where deg (f)


denotes the degree of the function f.
The denition of f

(V ) makes sense also in the case of f : X Y being a proper


mapping between two (arbitrary) Riemann surfaces, since in that case the degree of
f is well dened (exercise 21 and 1.3.14).
2.5.7. Proposition. For a Riemann surface X the group H
2
(X, S) = 0,
where S is the sheaf of smooth functions on X.
Proof. The proof is similar to the case of the 1st cohomology group but we
include it here for the sake of completeness. Let U = {U
j
}
j
be an open covering of X
and choose a cocycle (f
ijk
) Z
2
(U, S). Let {
j
}
j
be a partition of unity subordinate
to U. The functions
k
f
ijk
are smooth on U
i
U
j
. We set g
ij
=

k

k
f
ijk
; observe
that for any point of X this sum has only a nite number of nonzero terms. We
have (g
ij
) C
1
(U, F) and
g
ij
g
ik
+ g
jk
= f
ijk
,2.5. SERRE DUALITY 85
so (f
ijk
) splits and the result follows.
As in the case of the 1st cohomology group the above result is valid for the sheaves
of smooth forms and forms of type (1, 0) and (0, 1) as well.
2.5.8. Corollary. H
2
(C, O) = 0.
Proof. We have an exact sequence of sheaves on C,
0 O S

S
(0,1)
0,
from which we get the sequence
H
1
(C, S
(0,1)
) H
2
(C, O) H
2
(C, S).
Since the rst and third groups in this sequence are trivial we get that H
2
(C, O) = 0.
2.5.9. Now we return to the setting explained above, namely f : X Y is
a holomorphic mapping between compact surfaces (or a proper mapping between
Riemann surfaces), V X a bundle on X and f

(V ) the bundle on Y constructed


using f. Since f

(V )(A) = V (f
1
(A)), if U = {U
j
}
j
is a covering of Y , and we pull it
back to X via f
1
, f
1
(U) = {f
1
(U
j
)}
j
, we obtain a natural map H
n
(U, f

(V ))
H
n
(f
1
(U), V ), for n = 0, 1, 2. For example, for n = 0 we have
H
0
(Y, f

(V )) = f

(V )(Y ) = V (f
1
(Y )) = V (X) = H
0
(X, V ),
In more generality,
C
n
(U, V ) =

(f

(V )) (U
0
U
n
) =

V
_
f
1
(U
0
) f
1
(U
n
)
_
= C
n
(f
1
(U), V ).
Returning to the example of 2.5.6, f : A B given by f(z) = z
2
, we have that
the mapping f

(O(A))(B) O
A
(A) given by g
1
(w) 1 + g
2
(w) z h(z) is clearly
an isomorphism. A similar construction is possible for the other example, namely
f : A
1

A
2
B, f|
A
1
(z) = z
2
and f|
A
2
(z) = z
3
.
Theorem. The mappings H
n
(Y, f

(V )) H
n
(X, V ) are isomorphisms for
n = 0, 1, 2.
Before proving this theorem we need a technical result.86 2. COMPACT RIEMANN SURFACES
2.5.10. Proposition. If is an open disc of C then H
2
(, O) = 0.
Proof. Let (f
ijk
) be an element of Z
2
(U, O), for some covering U of . By 2.5.7
we can nd a smooth 1-chain, (g
ij
) C
1
(U, S), such that f
ijk
= g
jk
g
ik
+ g
ij
.
Since

f
ijk
= 0 we have

g
ij
=

g
ik
+

g
jk
, so (

g
ij
) Z
1
(U, S). But H
1
(, S) = 0
(1.5.12) so there exists a 0-chain of smooth functions, (h
i
) satisfying

g
ij
= h
i
h
j
.
Let
i
be smooth functions with

i
= h
i
. It is easy to check that the functions
f
ij
= g
ij

i
+
j
are holomorphic and (f
ij
) = (f
ijk
).
2.5.11. Proof of theorem 2.5.9. The case n = 0 was proved above. For
n = 1 or n = 2, we argue as follows. Choose a covering of Y , say U = {U
j
}
jJ
,
such that f

(V ) is trivial on U
j
and each U
j
is mapped to a disc in C (by some
local coordinate of Y ). Denote by f
1
(U) the open covering of X consisting of
{f
1
(U
j
)}
jJ
. By shrinking the sets U
j
we can assume that f
1
(U
j
) is a disjoint
union of coordinate patches, homeomorphic to discs in C, and such that the bundle
V is trivial on each component (of f
1
(U
j
)). From the previous proposition we have
that U and f
1
(U) are Leray coverings for f

(V ) and V respectively. It follows that


the natural mappings H
n
(U, f

(V )) H
n
(f
1
(U), V ) are bijective.
Corollary. If V is a vector bundle on a compact surface X, then the coho-
mology group H
2
(X, V ) is trivial.
The Canonical Line Bundle
2.5.12. On can construct in a natural way a line bundle on a surface X from
local coordinates as follows. Let {(U
j
, z
j
)}
j
be an open covering of X by coordinate
patches. On the intersections U
j
U
k
we have dz
j
= g
jk
dz
k
, where g
jk
= z
j
/z
k
.
From the chain rule of composition of functions we have that (g
jk
) is actually a
cocycle, so it denes a line bundle on X, denoted by K
X
and called the canonical
line bundle of the surface. Holomorphic (meromorphic) sections are just holomor-
phic (respectively meromorphic) 1-forms on X, as one can check by writing out the
section condition with respect to the above covering.
Recall that K denotes the canonical class of X; that is, the class of the divisor of2.5. SERRE DUALITY 87
a holomorphic form. We have also used K to denote a representative of that class.
So with this notation we have L(K) = K
X
.
2.5.13. Proposition. For any divisor D the sheaves O(D + K) and (D)
are isomorphic.
Proof. Let be a non-zero form and K = div(). The desired isomorphism is
given by f f.
2.5.14. If D
1
and D
2
are divisors it is easy to check that L(D
1
+D
2
) is isomor-
phic to L(D
1
) L(D
2
). Hence L(D + K)

= L(D) K
X
; we will use this notation
in this section.
The Riemann-Roch theorem revisited
2.5.15. We dene the characteristic of a bundle V X over a compact surface
as (V ) = dim H
0
(X, V )dim H
1
(X, V ) (recall that the second cohomology group
vanishes). In particular, for the trivial line bundle we have (O) = 1 g. In this
notation the Riemann-Roch theorem can be written as follows.
Theorem (Riemann-Roch). (L(D)) = deg (D) + (O) = deg (D) + 1 g.
Corollary. If X is a compact surface, p a point of X and V a bundle on X,
the group H
1
(X\{p}, V ) vanishes.
Proof. From the Riemann-Roch theorem (2.3.4) we have that there exists a
meromorphic function f : X

C whose only pole lies at the point p. Since f is
proper on X\{p} we have an isomorphism H
1
(X\{p}, V )

= H
1
(C, f

(V )). Thus we
need to consider only the case of X being the Riemann sphere and p = . If V is
a line bundle the result has been proved in 2.5.8. For the general case we use the
sequence of cohomology groups induced by the sequence 2.5.3 and use induction on
the rank of V as in the proof of 2.5.4.
2.5.16. For a vector bundle of rank n we have that the transition functions
g
jk
are linear isomorphisms of C
n
, so they can be naturally considered as mappings
from U
j
U
k
to GL(n, C), g
jk
: U
j
U
k
GL(n, C), where GL(n, C) is the group of
invertible square matrices of order n. The determinant of this matrix of functions is88 2. COMPACT RIEMANN SURFACES
nowhere zero, det(g
jk
) : U
j
U
k
C

; it is easy to show that (det(g


jk
)) is actually
a cocycle with values in the sheaf O

, so it denes a line bundle on X, called the


determinant line bundle of V , written as det(V ). We dene the degree of V ,
deg (V ), as the degree of its determinant line bundle. To compute a section of det(V )
we need to do it only locally, so assume that we have a trivial bundle of rank n over
the unit disk, DC
n
. Let {s
1
, . . . , s
n
} be sections of this bundle and {v
1
, . . . , v
n
} a
basis of C
n
. We can write each section as s
j
(p) = (p, s
j1
(p) v
1
+ + s
jn
(p) v
n
). The
n-th exterior product of C
n
is a one dimensional vector space with basis v
1
v
n
.
It is easy to check that the exterior product of the given sections is given by
(s
1
s
n
)(p) =(p, ( s
11
(p) v
1
+ + s
1n
(p) v
n
) ( s
n1
(p) v
1
+ + s
nn
(p) v
n
)) =
=(p, det( s
jk
)(p) (v
1
v
n
)).
From this we have that s
1
s
n
is a section of the determinant line bundle of
V . Or equivalently det(V ) is just the nth exterior power of V .
2.5.17. Proposition. If 0 V

V V

0 is an exact sequence of
bundles, then
(1) deg (V ) = deg (V

) + deg (V

), and
(2) (V ) = (V

) + (V

).
Proof. The rst part is an easy consequence of the fact that the determinant
bundle det(V ) is isomorphic to the bundle det(V

) det(V

). To prove (2) consider


the exact sequence
0 H
0
(X, V

) H
0
(X, V ) H
0
(X, V

) H
1
(X, V

) H
1
(X, V )
H
1
(X, V

) 0.
We have that the alternate sums of the dimensions of the spaces in this sequences
are equal, i.e.
dim H
0
(X, V

) + dim H
0
(X, V

) + dim H
1
(X, V ) =
dim H
0
(X, V ) + dim H
1
(X, V

) + dim H
1
(X, V

).
One can prove this from the fact that the if L : W W

is a linear map then


dim(ker(L)) + dim(im(L)) = dim W (for nite dimensional vector spaces). The
result follows from this equality.2.5. SERRE DUALITY 89
2.5.18. Corollary. (V ) = deg (V ) + rk(V ) (O).
Proof. The statement has been proved for line bundles. If rk(V ) = 2, in the
sequence
0 L V L

0
we have that L and L

are line bundles. Applying the previous proposition we get


(V ) = (L) + (L

) = deg (L) + deg (L

) + 2 (O) = deg (V ) + rk(V ) (O).


The general cases follows by induction on the rank of V .
2.5.19. Theorem. deg (K
X
) = 2g 2.
Proof. Let f : X

C be a non-constant meromorphic function of degree
d. By composing f with a Mobius transformation if necessary we can assume that
all its poles are simple; that is, b
p
(f) = 0 if f(p) = . The exterior derivative
of f, df = f
z
dz + f
z
d z = f
z
dz, is a meromorphic form on X, or equivalently a
meromorphic section of K
X
. If f has a zero of order n at a point p then f(z) = z
n
and df = nz
n1
dz = nz
bp(f)
dz. On the other hand, all poles of f are simple,
so f(z) =
1
z
at such points (this expression is an abuse of notation; it has to be
understood with respect to some local coordinate z) and therefore df =
1
z
2
dz has
a double pole at p. We get that the degree of df is equal to

bp(f)=0
b
p
(f) 2d =
B 2d, where d is the degree of f and B its total ramication number (1.3.16).
Hence deg (K
X
) = B 2d.
For the trivial line bundle O
X
we have (O
X
) = 1 g. Since H
1
(

C, O) = 0 (2.2.6),
(O
b
C
) = 1, and therefore, by the above corollary,
(O
X
) = (f

(O
X
)) = deg (f

(O
X
)) +d(O
b
C
) = deg (f

(O
X
)) +d.
If we can prove that deg (f

(O
X
)) = B/2 we will have (O
X
) =
B
2
+d. From the
Riemann-Roch theorem (2.5.15) we know that (O
X
) = 1 g, so B = 2(g 1 +d).
Thus
deg (K
X
) = B 2d = 2g 2.
So we need to show that deg (f

(O
X
)) = B/2. In order to simplify notation,
write Y for the Riemann sphere. We dene a function : f

(O
X
) O
Y
in the
following manner. If U Y is open we know that f

(O
X
)(U) = O
X
(f
1
(U)). For h90 2. COMPACT RIEMANN SURFACES
a holomorphic function on f
1
(U) (that is, h is an element of f

(O
X
)(U)), we have
that (h) O
Y
(U) should be a holomorphic function on U; if q U we set
(h)(q) =

pf
1
(q)
(b
p
(f) + 1) h(p).
Clearly (h) is holomorphic and a bundle morphism. Let L be the determinant line
bundle of f

(O
X
). We use to dene a morphism
: L L O
Y
as follows. First of all recall that f

(O
X
) is a bundle of rank d (the degree of f); let
(
1
, . . . ,
d
) and (
1
, . . . ,
d
) be elements of O
X
(f
1
(U)), for any suciently small
coordinate neighborhood U in Y . Then (
1

d
) (
1

d
) is a section
of L L; we denote it by and dene () = det((
i
) (
j
)) (that is, is
the determinant of the matrix whose (i, j)th entry is (
i
) (
j
).) To compute the
action of in local coordinates let q be a point of Y and assume rst that f
1
(q)
consists of one single point, say p (so b
p
(f) = d 1). We have, in a proper choice
of local coordinates, f(z) = z
d
= w. We have seen that {1, z, . . . , z
d1
} is a basis of
f

(O
U
) over W (f : U W, where U and W are two copies of the unit disc). Then
is given by det((z
i+j
))
i,j=o,...,n
. But
(z
i+j
) = z
i+j
(1 +
i+j
+ + (
i+j
)
d1
),
where is a (b
p
(f) + 1)-th root of unity. This is simply because the pre-images of
a point w in the unit disc under z z
m
are of the form z, z, . . . ,
m1
z, where
z
m
= w and is an m-th root of unity. Therefore
(z
i+j
) =
_

_
(b
p
(f) + 1) z
i+j
, if i + j = 0, b
p
(f) + 1
0, otherwise.
Hence det((z
i+j
)) is a constant multiple of z
(i+j)(d1)
= w
d1
and is simply
multiplication by a non-zero multiple of w
d1
. For example, if d = 3 and b
p
(f) = 2
we have w = z
3
and
det
_
(z
i+j
)
_
= det
_
_
_
_
_
3 0 0
0 0 3z
3
0 3z
3
0
_
_
_
_
_
= 27 z
6
= 27w
2
= 27w
d1
.2.5. SERRE DUALITY 91
If f
1
consists of more than one point the situation is a sum of computations like
that above, and again we have that we have that is multiplication by non-zero
multiple of w
P
bp(f)
, where the sum is taken over all p with f(p) = q.
If
1
is a holomorphic section of L L we have that
2
= (
1
) is a holomor-
phic section of O
Y
. In local coordinates we can write
2
(q) = w
nq
, and therefore
(
1
)(q) = w
P
bp(f)

1
(q) = w
nq
. From this we get
1
= w
nq
P
bp(f)
. The degree
of O
Y
is 0. The degree of L L is given by the sum of the numbers n
q

b
p
(f),
which gives
2 deg (L) = deg (L L) =

_
_
n
q

pf
1
(q)
b
p
(f)
_
_
= deg (O
Y
) B = B.
This completes the proof of the theorem.
2.5.20. Corollary (of the proof). The arithmetic genus g

= dim H
1
(X, O)
and topological genus of a compact Riemann surface are equal.
Proof. In the above proof we have shown that B = 2(g

1+d), where g

is the
arithmetic genus of X. If we apply the Riemann-Hurwitz formula to the function f
of the previous proof we get 2g 2 = 2d + B, where g is the topological genus of
X. It follows that g = g

.
2.5.21. Theorem (Riemann-Roch). If L X is a line bundle on a compact
surface X we have
dim H
0
(X, L) dim H
0
(X, K
X
L
1
) = deg (L) + 1 g.
Proof. It suces to show
(4) dim H
0
(X, L) dim H
0
(X, K
X
L
1
) deg (L) + 1 g.
If we prove this, then replacing L by K
X
L
1
we get that both sides of the inequality
change sings, so we must have an equality.
From the previous formulation of the Riemann-Roch theorem we have that
(5) dim H
0
(X, L) deg (L) + 1 g.92 2. COMPACT RIEMANN SURFACES
If deg (L) > deg (K
X
), then H
0
(X, K
X
L
1
) = 0 and thus 5 is simply 4. It follows
that to prove 4 we can assume that L = L(D) and the inequality holds for L(D+p
0
),
where p
0
is a point of X. It is clear that dim H
0
(X, L(D)) dim H
0
(X, L(D +
p
0
))+1; and similarly dim H
0
(X, K
X
L(D)) dim H
0
(X, K
X
L(Dp
0
))+1.
Thus
dim H
0
(X, L(D + p
0
)) dim H
0
(X, K
X
L(D p
0
))
dim H
0
(X, L(D)) dim H
0
(X, K
X
L(D)) + 2.
Since we are assuming that 4 holds for L(D + p
0
), then we have that the equation
will fail for L(D) if and only if
dim H
0
(X, L(D + p
0
)) = 1 + dim H
0
(X, L(D))
and
dim H
0
(X, K
X
L(D)) = 1 + dim H
0
(X, K
X
L(D p
0
)).
Let be in H
0
(X, L(D+p
0
)) but not in H
0
(X, L(D)); and let be in H
0
(X, K
X

L(D)) but not in H


0
(X, K
X
L(D p
0
)). Then = is a meromorphic
form with one single pole at p
0
. But this is not possible because of the Residue
theorem (1.4.26).
2.5.22. Corollary. dim H
1
(X, L) = dim H
0
(X, K
X
L
1
).
2.5.23. Corollary. dim H
0
(X, K
X
) = g and dim H
1
(X, K
X
) = 1.
The Serre Duality Theorem
2.5.24. Proposition. Given two distinct points p and q on a compact Rie-
mann surface X, there exists a meromorphic dierential on X whose only singular-
ities are simple poles at p and q, with residues 1 and 1 respectively.
Proof. Consider the line bundle L = K
X
L(p + q). Since deg (K
X
L
1
) =
2 < 0, we have that dim H
0
(X, L) = g + 1 (2.5.21, 2.1.7 and 2.4.19). The space
H
0
(X, K
X
) is a g-dimensional subspace of H
0
(X, K
X
L(p+q)), so we have a form,
say , in H
0
(X, K
X
L(p + q)) which is not holomorphic (it does not belong to
H
0
(X, K
X
)). Because of the form of the divisor p + q and the Residues theorem2.5. SERRE DUALITY 93
we have that must have at worst a simple pole at each of the points p and q.
Since the sum of the residues must be 0, res
p
() and res
q
() have opposite signs,
and non-zero (since is not holomorphic). A appropriate multiple of will be the
desired form.
2.5.25. Proposition. There exists a canonical isomorphism
Res : H
1
(X, K
X
) C.
Proof. Choose a point p X and a coordinate patch (U, z) near p, with
z(p) = 0. Let U = {U, X\{p}}. Taking residues at p denes a mapping from res
p
:
Z
1
(U, K
X
) C. This mapping is not identically 0; for example, res
p
(dz/z) = 1. If
C
0
(U, K
X
) we have res
p
() = 0. This is easy to see: will be given by a pair
(
1
,
2
) of holomorphic forms on U and X\{p} respectively. Since
1
is holomorphic,
res
p
(
1
) = 0. On the other hand, res
p
(
2
) = 0 by the Residues theorem. Thus we
obtain a mapping on cohomology, res
p
: H
1
(U, K
X
) C (by an abuse of notation
we use the same notation for both mappings). Since dim H
1
(U, K
X
) = 1 we have
that res
p
is actually an isomorphism.
To complete the proof of the proposition we need to show that the mapping res
p
is independent of the point p (this is the meaning of canonical in the statement of
the result). Let q then be another point of X. First of all observe that if we shrink
the set U in the covering U above the mapping res
p
does not change. So we can
choose coordinate patches (U, z) and (V, w), dened in neighbourhoods of p and q
respectively, and such that z(p) = 0 = w(q) and U V = . Let U be as above,
V = {V, X\{q}} and W = {U, X\{p, q}, V }. On H
1
(U, K
X
) we consider the
class of the form dz/z, dened on U\{p} (dz/z is a holomorphic form in U\{p});
we have seen that res
p
(dz/z) = 1; denote the class of this form by
p
. Similarly we
dene
q
as the class of dw/w in H
1
(V, K
X
). These two forms give the same class
in H
1
(W, K
X
). More precisely we claim that
p

q
= (

pq
), where

pq
is dened
as the class in H
1
(W, K
X
) of the chain given by by 0 in U V = ,
pq
on U\{p}
and
pq
on V \{q}, and
pq
is as in 2.5.24. The dierence
p

q
on H
1
(W, K
X
)
is given by 0, dz/z and dw/w on U V , U\{p} and V \{q} respectively. Since94 2. COMPACT RIEMANN SURFACES
the form
pq
has a simple pole with residue 1 at p we can write it as
1
+
dz
z
on the
local coordinate z, where
1
is a holomorphic form. Hence
p

pq
is given by
1
on U\{p}. Similarly
q

pq
is holomorphic on V \{q}. This shows that the class
of
p

q
splits in H
1
(W, K
X
) (the splitting is done by the class of

pq
).
If is a class in H
1
(X, K
X
) we can represent it by [
p
] or [
q
] in H
1
(W, K
X
),
where the complex number has to be the same since the classes [
p
] and [
q
]
are the same. If we now consider [
q
] as an element of H
1
(U, K
X
) we have that
res
p
(
p
) = ; similarly res
q
(
q
) = . Thus we obtain a canonical mapping
Res : H
1
(X, K
X
) C, by taking residues at any point of X.
If U is an open covering of X and L a line bundle, there is a natural mapping from
L (K
X
L
1
) to K
X
, which induces a paring
H
0
(X, ) Z
1
(U, K
X
L
1
) Z
1
(U, K
X
).
It is not dicult to see that this mapping induces a mapping on cohomology,
H
0
(X, ) H
1
(U, K
X
L
1
) H
1
(U, K
X
).
2.5.26. Theorem (Serre Duality). The natural mapping
: H
0
(X, L) H
1
(X, K
X
L
1
) H
1
(X, K
X
)
Res
C
is non-degenerate.
Proof. Given an element H
0
(X, L) we need to show that if () = 0 for
all H
1
(X, K
X
L
1
) then = 0. Fix p X and let (U, z) be a coordinate patch
with z(p) = 0. Consider the covering U = {U, X\{p}}. By the corollary in 2.5.15,
we have that U is a Leray covering for the sheaves (of holomorphic sections of the
line bundles) L, K
X
and K
X
L
1
. The forms z
n
dz, for n Z, can be considered
as elements of Z
1
(U, K
X
L
1
); let us use the same notation for the corresponding
classes in H
1
(U, K
X
L
1
). By computing the residues explicitly one can see that
if ( (z
n
dz)) = 0 for all n then all the coecients in the Taylor series of will
vanish; that is, = 0.
2.5.27. In the language of cohomology groups the above theorem takes the
following useful form.2.6. APPLICATIONS OF THE RIEMANN-ROCH THEOREM 95
Theorem (Serre Duality). Let X be a compact Riemann surface and D a divisor
on X. Then the group H
0
(X, (D)) is isomorphic to H
1
(X, O(D))

.
Proof. The above result shows that dim H
0
(X, L) dim H
1
(X, K
X
L
1
).
Combining this with 2.5.22 we actually have an equality. The theorem follows from
2.5.13.
2.6. Applications of the Riemann-Roch Theorem
In this section we obtain some simple consequences of the Riemann-Roch the-
orem. Perhaps the most striking fact shown here is that any surface of genus 0
must be (biholomorphically equivalent to) the Riemann sphere. We also re-write
the expression in the Riemann-Roch theorem using only 0-th cohomology groups,
namely H
0
(X, O(D)) and H
0
(X, (D)).
2.6.1. We have proved (2.5.20) that the topological genus g of a compact sur-
face X is equal to its arithmetic genus, i.e. g = dim H
1
(X, O). Using this fact
and the the Serre Duality formula we can rewrite the Riemann-Roch theorem in the
following form.
Theorem (Riemann-Roch). Let X be a compact surface of genus g and D a
divisor on X. Then
dim H
0
(X, O(D)) dim H
0
(X, (D)) = 1 g + deg (D).
The sheaves associated to the zero divisor are O and , that is, the sets of holomor-
phic functions and forms on the surface X. In this case the Riemann-Roch theorem
gives us the following expression:
dim H
0
(X, O) dim H
0
(X, ) = 1 g.
In other words, the space of holomorphic 1-forms on X has dimension equal to
the genus of the surface (since H
0
(X, O) is isomorphic to C and therefore a one-
dimensional space). In particular, there are no non-zero holomorphic forms in the
Riemann sphere, a result that can be also obtain from the fact that the degree of
the canonical class of

C is 2 (2.5.19).96 2. COMPACT RIEMANN SURFACES
2.6.2. Proposition. Let D be a divisor on a compact Riemann surface X.
Then
H
0
(X, O(D))

= H
1
(X, (D))

.
Proof. Let be a non-zero 1-form on X and K = () its divisor. We have seen
that there are sheaf isomorphisms (D)

= O(D + K) and O(D)

= (D K).
Therefore from the Serre Duality theorem we get
H
1
(X, (D))


= H
1
(X, O(D + K))


= H
0
(X, (D K))

= H
0
(X, O(D)).
2.6.3. The following theorem is an easy consequence of the previous result
and 2.1.7.
Theorem. Let X be a compact surface, D a divisor on X. Then
(1) if deg (D) > 2g 2, H
1
(X, O(D)) = 0;
(2) If deg (D) > 0, H
1
(X, (D)) = 0.
Corollary. If M denotes the sheaf of meromorphic functions on a compact
Riemann surface X, then H
1
(X, M) = 0.
Proof (of the corollary). If U = {U
a
}
aA
is a covering of X it is easy to
construct another covering, say V = {V
k
}
n
k=1
, with V < U, and renement mapping
: {1, . . . , n} A, such that V
k
is a compact subset of U
(k)
(use the compactness
of X). Let be an element of H
1
(X, M) and choose a cocycle (f
ab
) Z
1
(U, M)
to represent it. Thus f
ab
is a meromorphic function on U
a
U
b
. In particular, since
V
j
V
k
is a compact subset of U
(j)
U
(k)
, the function f
ab
has a nite number
of poles in V
j
V
k
. So we can nd an eective divisor D, with deg (D) > 2g 2,
and such that div(f
(j)(k)
) D on V
j
V
k
. By the previous theorem we have
that H
1
(X, O(D)) = 0, so there exist functions f
j
O(D)(V
j
), such that f
jk
=
f
j
f
k
on V
j
V
k
. In other words, (f
jk
) = 0 in H
1
(V, M). Since the mapping
from H
1
(U, M) to H
1
(V, M) is injective we have H
1
(U, M) = 0. It follows from
lemma 1.5.10 that H
1
(X, M) = 0.2.6. APPLICATIONS OF THE RIEMANN-ROCH THEOREM 97
2.6.4. Theorem. Let D be a divisor on a compact surface X of genus g and
p a point of X. Assume deg (D) > 2g 1. Then there exists a divisor E, linearly
equivalent to D, such that E(p) = 0.
Proof. Since deg (D p) = deg (D) 1 > 2g 2, by 2.6.3 the cohomology
groups H
1
(X, O(D)) and H
1
(X, O(D p)) vanish. Therefore dim H
0
(X, O(D)) =
dim H
0
(X, O(Dp)) + 1 and hence the inclusion mapping
H
0
(X, O(D p)) H
0
(X, O(D))
will not be surjective. The image of this mapping is the set of meromorphic functions
f with div(f) D and such that ord
p
(f) 1 D(p). Let f be a function in
H
0
(X, O(D))\H
0
(X, O(Dp)). Then ord
p
(f) = D(p), so we can write div(f) =
E D, where E is a divisor satisfying E(p) = 0. From this construction we have
that E is linearly equivalent to D.
2.6.5. Theorem. Let : L X a line bundle with deg (L) > 2g 1 on a
compact surface X of genus g and p a point of X. Then there exists a holomorphic
section s of L such that s(p) = 0.
Proof. The proof consists simply on rewriting the previous result in the lan-
guage of line bundles. Since any line bundle on X is isomorphic to the line bundle
of a divisor (2.4.20) we will assume that L

= L(D). By the previous result we have
a divisor E with E D and E(p) = 0. The canonical section s
E
of L(E) satises
(s
E
) = E; thus s
E
(p) = 0. A section of L can be obtained using the isomorphisms
L

= L(D)

= L(E) (these isomorphisms will take non-zero vectors to non-zero vec-
tors).
See [19] for a dierent proof of the above result.
2.6.6. Let X be a surface of the form X = T

; let : C T

denote the
canonical projection (1.3.6). The Weierstrass -function is dened by
(z) =
1
z
2
+

ZZ\{(0,0)}
_
1
(z )
2

1
z
2
_
,
where the sum is taken over all numbers = n+m, with (n, m) = (0, 0). Since is
invariant under the elements of G

( S
n,m
= , for all S
n,m
G

), the expression98 2. COMPACT RIEMANN SURFACES


f

((z)) = (z) denes a meromorphic function f

: T



C on the torus T

.
The only singularities of are double poles at the points of the form n + m; but
all these points are mapped to a single point of T

, say z
0
= (0). Thus the the
divisor D = 2 z
0
(in T

) satises dim H
0
(T

, O(D)) 2 (the constant functions are


obviously elements of this space).
If w and w
1
are local coordinates in T

, whose domains of denition have non-


empty intersection; we have seen that w
1
(p) = w(p) + n + m, for n, m Z (1.3.6).
This implies that dw denes a 1-form on T

without zeroes or poles. In other words,


K = 0, where K is the canonical class of T

. Hence O(D)

= (D) (2.5.13),
and since deg (D) > 0 we have H
0
(T

, (D)) = 0. Applying the Riemann-Roch


theorem we get
dim H
0
(T

, O(D)) = 1 g + deg (D) = 2.


The functions {1, f

} form a basis of this space.


Using similar techniques one can compute the dimensions of other spaces of forms
and functions on a surface (exercise 50).
2.6.7. Theorem. There does not exists a meromorphic function on T

with
only one simple pole.
Proof. Assume h : T



C is a meromorphic function whose only singularity
is a single pole. Without loss of generality (see remark 1 after the proof) we can
assume that the pole of h is at the point z
0
= (0). Then the divisor of h satises
div(h) z
0
. If D = 2z
0
, we have that h H
0
(X, O(D)). Thus we can write h
as a linear combination of the constant function 1 and f

: h = a +b f

. If b = 0 we
have that h is constant. On the other hand, if b = 0 then h will have a double pole
at z
0
. This contradiction proves the theorem.
Remarks. 1. If p
0
= (z
0
) and p
1
= (z
1
) are two points in T

the mapping
L : T

given by L((z)) = (z z
0
+z
1
) is clearly biholomorphic and satises
L(p
0
) = p
1
. Thus any two points on T

are equivalent.
2. Another way of proving the above theorem is as follows. Suppose that h : T



C
has one single pole as its only singularity; then h must have degree 1 and T

would2.7. PROJECTIVE EMBEDDINGS 99


have genus 0 (see 2.3.5). However we have included the above proof to show an
application of the Riemann-Roch theorem.
2.6.8. The following important result becomes now a straightforward conse-
quence of 2.3.5 and 2.5.20.
Theorem. A compact Riemann surface of genus 0 is biholomorphic equivalent
to the Riemann sphere.
2.7. Projective embeddings
We have seen that the Riemann sphere can be identied with the one-dimensional
complex projective space (exercise 17). In this section we generalise that result in
the following sense: we show that any compact Riemann sphere can be embedded
in a projective space (that is, X can be smoothly identied with a subset of a
projective space). The dimension of that projective space depends on whether the
given surface has a meromorphic function of degree 2 or not. Those surfaces with
such functions are in some sense dierent; they are called hyperelliptic surfaces and
we will study more properties of them in the next sections. The fact that a surface
can be considered as a subset of a projective space allows one to use many of the
results of Algebraic Geometry to obtain properties of Riemann surfaces; we will not
pursue such approach in this book, the interested reader can nd more material
in [10] or [19].
We start this section by recalling the construction of (complex) projective spaces.
Using the space of holomorphic forms on a surface X we dene a mapping, called
the canonical map, from X to certain projective space; we show that in most
situations this mapping is an embedding. In the remaining cases, by a small change
in the denition of the canonical map we can still produce an embedding of the
surface X in a projective space. We explain the relation of these embeddings with
certain spaces of forms on X. Finally, in the last part of this section we show another
way of obtaining embeddings by using sections of line bundles.
2.7.1. Let (C
n+1
)

denote the set C


n+1
\{(0, . . . , 0)}. Dene an equivalence
relation in this set, say , by identifying (z
0
, . . . , z
n
) and (w
0
, . . . , w
n
) if there exists100 2. COMPACT RIEMANN SURFACES
a non-zero complex number , such that z
j
= w
j
, for j = 0, . . . , n. The quotient
space (C
n+1
)

/ is called the n-dimensional (complex) projective space, written


as P
n
. We will use the notation [z
0
: : z
n
] for a point in P
n
. Observe that any
point in projective space has at least one coordinate not equal to 0. We put on P
n
the quotient topology induced by the natural mapping (C
n+1
)


P
n
: a set U of
P
n
is open if and only if
1
(U) is open in (C
n+1
)

(or equivalently, in C
n+1
). This
topology makes P
n
a manifold (of real dimension 2n). But as one might expect,
projective spaces are complex manifolds (of complex dimension n). To construct
an atlas we consider the covering {U
0
, . . . , U
n
}, where U
j
is the open set dened by
U
j
= {[z
0
: : z
n
]; z
j
= 0}; the coordinate functions are given by the mappings

j
: U
j
C
n
[z
0
: : z
n
]
_
z
0
z
j
, . . . ,
z
j1
z
j
,
z
j+1
z
j
, . . . ,
z
n
z
j
_
Changes of coordinates are holomorphic functions. For example
1

1
0
is the
function (w
1
, . . . , w
n
) (1/w
1
, w
2
/w
1
, . . . , w
n
/w
1
), dened for (w
1
, . . . , w
n
) C
n
,
with w
1
= 0.
2.7.2. The following proposition is an easy consequence of the Riemann-Roch
and Serre Duality theorems.
Proposition. Let X be a compact surface of genus g 1 and p a point of X.
Then there exists a holomorphic form on X that does not vanish at p.
Proof. Since the genus of X is positive there does not exist a meromorphic
function whose only singularity is a simple pole at p (see remark 2 in 2.6.7). Or
in terms of cohomology spaces, the dimension of H
0
(X, O(p)) is 1. The space of
holomorphic forms vanishing at p is H
0
(X, (p)). We can compute its dimension
using the Riemann-Roch theorem as follows:
dim H
0
(X, O(p)) dim H
0
(X, (p)) = 2 g.2.7. PROJECTIVE EMBEDDINGS 101
From this equation we get that dim H
0
(X, (p)) = g 1. Since the space of
holomorphic forms has dimension equal to g the result follows.
For the rest of this section we will assume that X is a compact surface of genus
g 2, unless otherwise stated.
2.7.3. Let f : X P
n
be a holomorphic function, and p a point of X. If
f(p) = [w
0
: : w
n
] we can assume without loss of generality, that w
0
= 0. Choose
a local coordinate z on X and the coordinate
0
described above, to obtain a map
F(x) = (F
1
(x), . . . , F
n
(x)) : U C
n
, where U is a neighbourhood of 0 (in C). The
dierential of F (or f) at p is dened as the linear mapping
d
p
F : C C
n
( F

1
(0), . . . , F

n
(0)).
Some times we will identify d
p
F with the vector (F

1
(0), . . . , F

n
(0)) to simplify no-
tation. The mapping f is called an embedding if it is injective and d
p
F is not
identically zero for any point p of X.
2.7.4. To dene an mapping of X into P
n
we consider a basis {
1
, . . . ,
g
}
of H
0
(X, ). Take a local coordinate z vanishing at p and write
j
= f
j
dz, for
holomorphic functions f
j
. We set i
1
: X P
g1
by i
1
(q) = [f
1
(q) : . . . : f
g
(q)], for
q in a neighbourhood of p. By 2.7.2 there exists a 1-form that does not vanish at
q, so i
1
(q) lies in P
g1
. It is also easy to see thati
1
does not depend on the choice
of local coordinate z (if we change coordinates, the values f
j
(q) are multiplied by a
non-zero complex number, give by the dereivative of the change of coordinates) and
so is well-dened. The mapping i
1
is called the canonical mapping. Our aim is
to nd under what conditions (on X) the mapping i
1
is an embedding.
For that purpose, let p be an arbitrary but xed point of X. We change the basis
{
1
, . . . ,
g
} (to adjust it to p) as follows: rst of all, by 2.7.3 we can assume
that f
1
(p) = 0; we then replace f
1
by
1
f
1
(p)
f
1
. By an abuse of notation we denote
this new function by f
1
as well. Substitute now f
j
by f
j
f
j
(p)f
1
, and again
abuse notation by using f
j
for the new functions (j = 2, . . . , g). Thus we have
that the basis {f
1
, . . . , f
g
} satises f
1
(p) = 0, and f
2
(p) = = f
g
(p) = 0. So102 2. COMPACT RIEMANN SURFACES
i
1
(p) = [1 : 0 : : 0]. We can use the local coordinate
0
near i
1
(p) to compute
d
p
(i
1
). We have that
(
0
i
1
)(z) =
_
(f
2
z
1
)(w)
(f
1
z
1
)(w)
, . . . ,
(f
g
z
1
)(w)
(f
1
z
1
)(w)
_
,
for w in a neighbourhood of 0 in C (recall that z is a local coordinate on X with
z(p) = 0). Thus di
1
(p) is given by
_
(f
2
z
1
)

(0)(f
1
z
1
)(0) (f
2
z
1
)(0)(f
1
z
1
)

(0)
(f
1
z
1
)
2
(0)
, . . .
. . . ,
(f
g
z
1
)

(0)(f
1
z
1
)(0) (f
g
z
1
)(0)(f
1
z
1
)

(0)
(f
1
z
1
)
2
(0)
_
=
=
_
(f
2
z
1
)

(0), . . . , (f
g
z
1
)

(0)
_
.
We see from this expression that d
p
(i
1
) gives us the order of the zeroes of
j
at the
point p (for j = 2, . . . , g).
2.7.5. We rst consider under what conditions the mapping i
1
is not one-to-
one. This means that there exists a point q in X, q = p, such that i
1
(q) = [1 : 0 :
: 0]; that is, if a holomorphic form vanishes at p, then it also vanishes at q. In
terms of cohomology groups this statement is given by the equality H
0
(X, (p)) =
H
0
(X, (pq)). From 2.7.2 we know that dim H
0
(X, (p)) = g1. If we apply
the Riemann-Roch theorem to the divisor p + q, we get
dim H
0
(X, O(p + q)) = dim H
0
(X, (p q)) + 3 g = g 1 + 3 g = 2.
The space of constant functions is a one dimensional subspace of H
0
(X, O(p + q)).
Hence there exists a non-constant meromorphic function f in O(p
q
)(X). Since the
genus g 2 the function f cannot have only one simple pole as its only singularity
(2.3.5 and 1.4.26). Thus f must have a pole at each of the points p and q and they
must be simple poles (because div(f) p q). We obtain that f is a function of
degree 2.
2.7.6. Definition. A compact Riemann surface is called hyperelliptic if it
has a meromorphic function of degree 2.2.7. PROJECTIVE EMBEDDINGS 103
2.7.7. The previous paragraphs show that if the canonical map i
1
fails to be
injective then X is hyperelliptic. We show next that if d
p
i
1
= (0, . . . , 0) then X
is also hyperelliptic. From the last line of 2.7.4 we have that if d
p
i
1
0 then any
holomorphic form that vanishes at p must do it with order at least 2. Equivalently,
H
0
(X, (p)) = H
0
(X, (2p)). We apply again the Riemann-Roch theorem, this
time to the divisor 2p:
dim H
0
(X, O(2p)) = dim H
0
(X, (2p)) + 3 g = 2;
which means again that X is hyperelliptic. We formulate this results in a single
statement.
Theorem. If X is a compact non-hyperelliptic surface then the canonical map-
ping i
1
: X P
g1
is an embedding.
2.7.8. To construct an embedding i
m
: X P
n
for the case of hyperelliptic
surfaces we consider the divisor mK, where m is a positive integer and K the
canonical class of X, or rather a representative of it, given by the divisor of a
holomorphic form, say div() = K. From exercise 50 we have that
dim H
0
(X, O(mK)) = (2m1)(g 1) = d > 0.
Let {f
1
, . . . , f
d
} be a basis of this space; dene i
m
: X P
d1
by i
m
(p) = [f
1
(p) :
: f
d
(p)]. We rst need to show that this mapping is well dened, i.e. the point
i
m
(p) is actually in P
d1
. Recall that there exists a holomorphic 1-form, say
p
that does not vanish at p. The space of holomorphic forms can be identied
with H
0
(X, O(K)) (2.5.13). Let f be the function in this space corresponding to

p
under the said isomorphism; then div(f) K and f(p) = 0. The function f
m
is in H
0
(X, O(mK)) and does not vanish at p. This shows that i
m
is well dened.
Without loss of generality we can assume that i
m
(p) = [1 : 0 : : 0]; that is
{f
2
, . . . , f
q
} are a basis of H
0
(X, O(mK p)). Assume that there exists a point q,
with q = p and i
m
(q) = i
m
(p). This means that any function in H
0
(X, O(mK))
which vanishes at p it will also vanish at q. Therefore
dim H
0
(X, O(mK p)) = dim H
0
(X, O(mK p q)).104 2. COMPACT RIEMANN SURFACES
It is not dicult to see (see exercise 50) that the space H
0
(X, O(mK p)) has
dimension d1. If we now apply the Riemann-Roch theorem to the divisor mKpq
we obtain
dim H
0
(X, O(mK p q)) = dim H
0
(X, (p +q mK)) +m(2g 2) 2 +1 g,
which implies that
dim H
0
(X, (p + q mK)) = 1.
The sheaves (p +q mK) and O(p +q (m1)K) are isomorphic (2.5.13); thus,
from 2.1.7 we have that the degree of p + q (m1)K must be non-negative. An
easy computation shows that this condition is given by (m1)(g1) 1. Therefore
if m 3 we have (m1)(g 1) > 1, since we are assuming that g 2, and in this
case i
m
will be one-to-one.
If di
m
(p) 0 we have that any function in H
0
(X, O(mK)) that vanishes at p it
does it with order at least 2. Equivalently,
H
0
(X, O(mK p)) = H
0
(X, O(mK 2p)).
Reasoning as above, with q replaced by p, we see that if m 3 then di
m
(p) does
not vanish.
2.7.9. We collect all the above results in the following theorem.
Theorem. Let X be a compact Riemann surface of genus g 2 and i
m
: X
P
n
the mth canonical map. If X is not hyperelliptic then i
1
is an embedding and
n = g 1. On the other hand, i
3
is always an embedding (in this case n = 5g 6).
The Riemann sphere

C is isomorphic to P
1
, and a torus can be embedded into P
2
.
Proof. The statement on the Riemann sphere was left to the reader (exercise
17) and the one regarding the torus is not dicult. We give some hints in exercise
54.
2.7.10. The above study in the case of hyperelliptic surfaces can be done in a
slightly dierent but equivalent way as we explain next. The reader can skip this
subsection without any problem.
Similar to the denition of 1-forms one has the concept of higher order forms: a
holomorphic m-form is an assignment of a holomorphic function f to each local2.7. PROJECTIVE EMBEDDINGS 105
coordinate z, such that the expression f(z) dz
m
is invariant. More precisely, if t is
another coordinate on X (with domain of denition not disjoint from that of z) and
is given by the function g in this coordinate ( is g(t) dt
m
), then f(z) = g(t) (
t
z
)
m
.
For example, if = h(z) dz is a holomorphic 1-form then
m
:= h
m
(z) dz
m
denes
an m-form, as the reader can easily check.
The space of m forms can be identied with the cohomology group H
0
(X, mK),
where K is the canonical class (or a representative of it). To see this choose a non-
zero holomorphic 1-form, say , and let K = div(). Let
m
(X) denote the space
of m-forms. Then the mapping H
0
(X, mK)
m
(X) given by h h
m
is an
isomorphism. The facts that i
m
is not injective, or d
p
i
m
0 can be translated to
results on higher order forms, similar to the statements regarding 1-forms and i
1
.
2.7.11. Holomorphic sections of (certain) line bundles produce embeddings of
Riemann surfaces into projective spaces as well. For a line bundle L on a compact
surface X we dene a mapping
L
: X P
n
, where n = dim H
0
(X, L) 1, as
follows. Choose a basis {s
0
, . . . , s
n
} of the space H
0
(X, L) of holomorphic sections
of L. For a point p of X let be a section of L, dened on a neighbourhood U of p,
which does not vanish at p. If the sections {s
j
}
n
j=0
do not have common zeroes the
expression [
s
0
(q)
(q)
: :
sn(q)
(q)
] denes a point of P
n
.
The bundle L is isomorphic to the line bundle L(D) of a divisor D and the space
H
0
(X, L) can be identied with H
0
(X, O(D)) (2.4.19). In the case of deg (D) >
2g 2 we have that
dim H
0
(X, O(D)) = deg (D) + 1 g,
so n = deg (D) g.
Theorem. If L is a line bundle of degree greater than 2g on a compact surface
X, the mapping
L
: X P
n
, for n = deg (L) g is an embedding.
Proof. Let q be an arbitrary point of X (the case q = p is also allowed; as a
matter of fact, we will use it below). Since the degree of the bundle L L(p) is
greater than 2g 1 we have that there exists a section s with s(q) = 0 (2.6.4). Let
s
p
denote the standard section of L(p) (so div(s
p
) = p). The expression s = s s
p
denes a section of L L(p) L(p). Since this last bundle is isomorphic to L we106 2. COMPACT RIEMANN SURFACES
can consider s as a section of L. If q = p the section s does not vanish at q; on the
other hand, if q = p we have that s has a simple zero at p.
Using these results, an argument similar to the above constructions nishes the proof
of the theorem. We left the details for the exercises (exercise 55).
2.7.12. Definition. A line bundle L on a surface X is called ample if the
mapping
L
dened above is an embedding, where L

= L
m
= L
m
L, for
some positive integer m. The line bundle L is called very ample if m = 1 (that is,
sections of L produce an embedding of X into projective space).
Using the above results it is easy to see when a line bundle is ample.
Proposition. A line bundle L is ample if and only if deg (L) > 0.
Proof. Assume that deg (L) > 0. For m positive integer we have
deg (L
m
) = mdeg (L) > 2g,
so the sections of L
m
give an embedding of X from the previous result (2.7.11).
On the other hand, if L is ample, and L

= L(D), then dim H
0
(X, O(mD)) > 0,
so deg (L
m
) = mdeg (L) = mdeg (D) must be positive.
Theorem 2.7.11 can be reformulated as saying that if L is a line bundle of degree
greater than 2g then L is very ample.
2.7.13. The following theorem is an easy consequence of the results of this
section.
Theorem. Let X be a compact Riemann surface of genus g 2. If X is not
hyperelliptic then the canonical bundle K
X
is very ample.
2.8. Weierstrass Points and Hyperelliptic Surfaces
We have seen that given an arbitrary point p on a compact Riemann surface X
of genus g there exists a meromorphic function whose only singularity is a pole at
p, of order at most g + 1 (2.3.4). It is therefore of interest to know whether there
are points on a surface that are poles of functions with order strictly less than g +1.
In this section we prove that such points, called Weierstrass points, always exist2.8. WEIERSTRASS POINTS AND HYPERELLIPTIC SURFACES 107
on a compact surface; we also study the relationship between Weierstrass points
and hyperelliptic surfaces. Throughout this section we will assume that all compact
surfaces have genus greater than 1, unless otherwise stated.
Weierstrass Points
2.8.1. We start with a theorem of Weierstrass that gives some information
about orders of poles of meromorphic functions on compact surfaces.
Theorem (Weierstrass Gap Theorem). Let X be a compact Riemann surface of
genus g 1 and p a point of X. Then there exist exactly g integers,
1 = n
1
< < n
g
< 2g,
such that there does not exist a meromorphic function on X whose only singularity
is a pole of order n
j
at p.
Proof. The genus 1 case has already been proved in 2.6.7. For the case of g 2,
rst of all observe that if there exists a meromorphic function on X, holomorphic
on X\{p} and with a pole of order k at p, then f belongs to H
0
(X, O(kp)) but not
to H
0
(X, O((k 1)p)). From the sequence
0 O((k 1)p) O(kp) C,
we obtain the following exact sequence:
0 H
0
(XO((k 1)p)) H
0
(X, O(kp)) C.
Thus dim H
0
(X, O(kp)) dim H
0
(X, O((k 1)p)) is either 1 or 0, depending on
whether a function f as above exists or not.
The Riemann-Roch theorem gives
dim H
0
(X, O(kp)) dim H
0
(X, O((k 1)p)) =
1 + dim H
1
(X, O(kp)) dim H
1
(X, O((k 1)p)).108 2. COMPACT RIEMANN SURFACES
From 2.6.3 we have that dim H
1
(X, O(kp)) = 0 if k > 2g 2. Therefore
2g1

k=1
_
dim H
0
(X, O(kp)) dim H
0
(X, O((k 1)p))
_
=
=
2g1

k=1
_
1 + dim H
1
(X, O(kp)) dim H
1
(X, O((k 1)p))
_
=
= 2g 1 dim H
1
(X, O) = g 1.
Thus there are exactly g 1 terms in the rst sum equal to 1 and g terms equal to
0. These last summands are the ones in the statement of the theorem.
2.8.2. A look into the above proof shows that there is nothing special about
choosing one point p, the important matter is the sequence of divisors p, 2p, 3p, . . ..
Thus Weierstrass theorem admits the following generalisation.
Theorem (Noether Gap Theorem). Let X be a compact Riemann surface of
genus g 2. Consider a sequence of divisors, D
0
= 0, D
1
= p
1
, D
2
= p
1
+ p
2
,. . . ,
D
2g
= p
1
+ + p
2g
on X. Then there exist precisely g integers,
1 = n
1
< . . . < n
g
< 2g,
such that
H
0
(X, O(D
j
)) = H
0
(X, O(D
j+1
))
if and only if j = n
k
for some k = 1, . . . , g.
2.8.3. Definition. A point p of a compact surface X is called a Weierstrass
point if there exists a meromorphic function f : X

C, such that f is holomorphic
on X\{p} and has a pole of order at most g at p.
From the Riemann-Roch theorem we have
dim H
0
(X, O(gp)) dim H
0
(X, (gp)) = g + 1 g = 1.
If p is a Weierstrass point then dim H
0
(X, O(gp)) 2 (the constant func-
tions form a one-dimensional subspace of H
0
(X, O(gp))), which implies that2.8. WEIERSTRASS POINTS AND HYPERELLIPTIC SURFACES 109
dim H
0
(X, (gp)) 1. Thus a point p is a Weierstrass point if and only if there
exists a holomorphic 1-form on X vanishing at p with order at least g.
2.8.4. In order to study Weierstrass points we need to recall some results from
Calculus. Let f
1
, . . . , f
m
be holomorphic functions dened on a domain (connected
open set) U of the complex plane. The Wronskian of f
1
, . . . , f
m
is dened by
W(f
1
, . . . , f
m
) = det
_
_
_
_
_
_
_
f
1
f
2
f
m
f

1
f

2
f

m

f
(m1)
1
f
(m1)
2
f
(m1)
m
_
_
_
_
_
_
_
,
where f
(j)
denotes the j-th complex derivative of f.
Lemma. The functions f
1
, . . . , f
m
are linearly dependent over C if and only if
W(f
1
, . . . , f
m
) 0.
Proof. One implication is clear: if a function, say f
j
depends linearly of the
other m 1 functions, then the j-th column in the above matrix will be a linear
combination of the other columns and the above determinant will be 0.
We will show the other implication for the case m = 2. If W(f
1
, f
2
) 0 then
f
1
f

2
f

1
f
2
0. If f

1
0 then either f
1
0 or f
2
is constant; in either case, f
1
and f
2
are linearly dependent. The case of f

2
0 is similar. So let us assume that
neither f

1
nor f

2
is identically 0. In the connected set V = {z U; f
2
(z) = 0}
we have that (f
1
(z)/f
2
(z))

= 0, so f
1
= f
2
, for some complex number . By the
Identity Principle the equality f
1
= f
2
holds on U.
We leave as an exercise to the reader (exercise 58) to show that if U is a connected,
open subset of C, and A is a discrete subset of U, then U\A is connected.
2.8.5. Let K
X
denote the canonical bundle of X and K
n
X
the tensor product
of K
X
with itself n times, K
n
X
= K
X

n
K
X
.
Theorem. Let n = g(g+1)/2. Then there exists a non-zero holomorphic section
W of K
n
X
such that the zeroes of W are precisely the Weierstrass points of X.
The proof of this theorem requires some more material, so we postpone it till a later
subsection (2.8.9).110 2. COMPACT RIEMANN SURFACES
The transition functions of the canonical bundle K
X
are given by dz/dw, where
z and w are two local coordinates on X. Hence the transition functions of the line
bundle K
n
X
are (dz/dw)
n
; a holomorphic section of this bundle then will be given
by a collection of holomorphic functions f(z) (in the local coordinate z) satisfying
f(z) = f(w)
_
dz
dw
_
n
.
Let {
1
, . . . ,
g
} be a basis of H
0
(X, ) and let p be a point of X. If z and w are
two local coordinates dened on neighbourhoods of p, we have
j
= f
j
dz = h
j
dw.
Hence f
j
= h
j
, for = dw/dz, from which it follows that
d
m
f
j
dz
m
=
m+1
_
d
m
h
j
dw
m
_
+
m1

k=0

k
_
d
k
h
j
dw
k
_
;
where
k
are holomorphic functions independent of j. For example,
df
j
dz
=
dh
j
dw
_
dw
dz
_
2
+ h
j
d
2
w
dz
2
,
and
d
2
f
j
dz
2
=
d
2
h
j
dw
2
_
dw
dz
_
3
+ 3
dh
j
dw
dw
dz
d
2
w
dz
2
+
dh
j
dw
d
3
w
dz
3
.
But then we see that
W(f
1
, . . . , f
g
) =det
_
d
m
f
j
dz
m
_
j=1,...,g
m=0,...,g1
= det
_

m+1
d
m
h
j
dz
m
_
j=1,...,g
m=0,...,g1
=
=
_
g1

m=0

m+1
_
W(h
1
, . . . , h
g
) =
n
W(h
1
, . . . , h
g
).
Thus the function W(f
1
, . . . , f
g
) gives a section of K
n
X
, for n = g (g +1)/2. We will
denote this section by W = W(
1
, . . . ,
g
; z)dz
n
.
2.8.6. We call the order of a zero p of W the multiplicity of p. We claim
that the zeroes of W do not depend of the choice of the basis {
1
, . . . ,
g
} of
H
0
(X, ). If {
1
, . . . ,
g
} is another basis then we have
j
=

g
k=1
c
jk

k
, Here
(c
jk
)
j,k=0,...,g
is the matrix of change of basis, and therefore det(d
jk
) = 0. This
implies W(
1
, . . . ,
g
; z)dz
n
= det(c
jk
) W(
1
, . . . ,
g
)dz
n
; so the zeroes of W are2.8. WEIERSTRASS POINTS AND HYPERELLIPTIC SURFACES 111
independent of the choice of basis, as claimed. From this argument it is also clear
that the multiplicity of a zero is independent of the chosen basis.
2.8.7. Proposition. The number of zeroes of W, counted with multiplicities,
is equal to (g 1)g(g + 1).
Proof. Since W is a holomorphic section of the line bundle K
n
X
it will have as
many zeroes as the degree of the bundle. This degree is equal to n times the degree
of the canonical bundle, that is
n(2g 2) =
g (g + 1)
2
(2g 2) = (g 1) g (g + 1).
2.8.8. Given a point p in X we say that a basis {
1
, . . . ,
g
} is adapted to
p if ord
p

1
< ord
p

2
< < ord
p

g
. We will denote ord
p

j
by m
j
. For a basis
adapted to p we dene
m(p) =
g

j=1
(m
j
j + 1).
Proposition. The multiplicity of a point is m(p).
Proof. Since the zeroes, and their multiplicity, do not depend of the basis of
holomorphic forms we can choose a basis adapted to a point p.
Although we are assuming that X has genus g 2 the construction of W makes
sense when g = 1; in that case the statement of the proposition holds clearly. The
case of higher genus follows by induction on g, as we prove next.
Observe that for holomorphic functions f, f
1
, . . . , f
n
we have
W(f f
1
, . . . , f f
n
) = f
n
W(f
1
, . . . , f
n
).
Assume that the proposition has been proved for g k. For the k +1 case we have
W(f
1
, . . . , f
k+1
) = f
k+1
W(1, f
2
/f
1
, . . . , f
k+1
/f
1
) =
=f
k+1
1
W((f
2
/f
1
)

, . . . , (f
k+1
/f
1
)

).
Since the basis chosen is adapted to p we have m
j
m
1
+ 1 > 0, for j 2. By the
induction hypothesis we see
ord
p
W(f
1
, . . . , f
k+1
) = (k + 1)m
1
+
k+1

j=1
(m
j
m
1
+ 1 +j 2),112 2. COMPACT RIEMANN SURFACES
which is the statement we wanted to show.
2.8.9. Proof of 2.8.5. As we have seen the section W(
1
, . . . ,
g
; z)dz
n
has
a zero of order m(p) =

g
j=1
(m
j
j +1) to p. We have that m(p) > 0 if and only if
m
g
> g 1 (since we are assuming the basis is adapted at p). But this is equivalent
to saying that p is a Weierstrass point of X.
Hyperelliptic Surfaces
2.8.10. Recall that a compact surface is called hyperelliptic if it has a mero-
morphic function of degree 2 (denition 2.7.6). Clearly the Riemann sphere and
the torus (1.3.6) are hyperelliptic. In the rst case one can take, for example, a
polynomial of degree 2. For the case of the torus the function f

, constructed from
the Weierstrass -function (2.6.6), has degree 2. We show in the next result that
surfaces of genus 2 are also hyperelliptic.
Proposition. A compact Riemann surface of genus 2 is hyperelliptic.
Proof. Let D be the divisor of a holomorphic form, so deg(D) = 2. By the
Riemann-Roch theorem we have
0 dim H
0
(X, O(D)) = 1 + dim H
0
(X, (D)).
Since dim H
0
(X, (D)) = 1 we have that H
0
(X, O(D)) = 2. The divisor D is
eective, so the constant functions form a 1-dimensional subspace of H
0
(X, O(D)).
Let f be then a non-constant function in H
0
(X, O(D)). Since X has genus 2 the
funtion f cannot have degree 1 (it would be an homeomorphism between X and the
Riemann sphere). Thus f must have degree 2 and X is hyperelliptic.
2.8.11. Proposition. If X is a hyperelliptic Riemann surface and f : X

C
is a function of degree 2 then the ramication points of f are precisely the Weierstrass
points of X.
Proof. First of all, by the Riemann-Hurwitz formula (1.3.16) the function
f has 2g + 2 ramication points. Choose one of those points, say p. We have
denoted by n
1
, . . . , n
g
the integers (gaps) between 1 and 2g given by the Weierstrass
Gap Theorem. If f(p) = then p is a pole of order 2. Otherwise, the function2.8. WEIERSTRASS POINTS AND HYPERELLIPTIC SURFACES 113
h(q) = 1/(f(q) f(p)) has a double pole at p. This means that there exists a
meromorphic function, say k, whose only singularity is a double pole at p (k = f or
k = h). The powers of this function, k, k
2
, . . . , k
g
gives us a sequence of functions
with singularities given by poles of orders 2, 4, . . . , 2g at p (and no other pole). Thus
m
j
= n
j
1 = 2j 2, so m
g
= 2g 2 > g 1 (since we are assuming that g > 1)
and p is a Weierstrass point of X. To complete the proof we need to show that these
are all the Weierstrass points of the surface.
From these computations we get that the multiplicity of each ramication points
of f is given by
g

j=1
(m
j
j + 1) =
g

j=1
(2j 1 j + 1) =
g

j=1
j =
g (g 1)
2
.
Since f has 2g + 2 ramication numbers, we obtain that these points contribute
(2g + 2)
g(g 1)
2
= (g + 1) g (g 1)
to the number of Weierstrass points (counted with multiplicities). But then,
from 2.8.5 and 2.8.7, we see that there cannot be more Weierstrass points.
2.8.12. The next result is a consequence of the proof of the previous proposi-
tion.
Proposition. There are exactly 2g +2 Weierstrass points on a compact hyper-
elliptic surface of genus g and each point has multiplicity g (g 1)/2.
It can be show that the converse of the above result also holds. See for exam-
ple [7, III.7.3, pg, 95].
2.8.13. Proposition. Let X be a hyperelliptic Riemann surface of genus
g 2, and let f and h be two meromorphic functions on X of degree 2. Then there
exists a Mobius transformation M such that h = M f.
Proof. It follows from the proof of 2.8.11 that if f
1
() = p
1
+ p
2
, (as a
divisor), and p
0
is a Weierstrass point of X, then p
1
+ p
2
2p
0
. Thus the divisor
h
1
() = q
1
+q
2
is linearly equivalent to p
1
+p
2
. Let k : X

C be a meromorphic
function with divisor p
1
+ p
2
q
1
q
2
. Multiplication by k is an isomorphism
between the spaces H
0
(X, O(p
1
+ p
2
)) and H
0
(X, O(q
1
+ q
2
)). With respect to the114 2. COMPACT RIEMANN SURFACES
basis {h, 1} and {f, 1}, this isomorphism will be given by a matrix
_
a b
c d
_
with non-
zero determinant, i.e. a d b c = 0. This gives us the following relations between
the functions f and h:
k h = a f + b k = c f + d.
Therefore h = Mf where M is the Mobius transformation M(z) = (az+b)/(cz+d).
2.8.14. Using a meromorphic function f of degree 2 on a hyperelliptic surface
X we can construct a biholomorphic mapping J : X X in the following way. If
p is a ramication point of f then we set J(p) = p. For a non-ramication point p
there exists a unique point p

in X, p = p

, such that f(p) = f(p

); we set J(p) = p

.
One sees easily that J is holomorphic and satises J
2
:= J J = Id
X
. We call J
the hyperelliptic involution. From this construction it seems that the mapping
J depends on the choice of the function f; however we will show (2.8.17) that this
is not the case. The quotient space Y = X/ < J >, with the quotient topology,
is a connected Hausdor space. Since J has only a nite number of xed points,
it is not dicult to see that Y is a Riemann surface; one can use the coordinates
of example 1.3.7 near the xed points of J. The natural projection : X Y
becomes holomorphic.
2.8.15. Proposition. A compact Riemann surface X of genus g 2 is
hyperelliptic if and only if it has an involution with 2g + 2 xed points.
Proof. We have seen one half of the proposition: if X is hyperelliptic the
involution J constructed above satises the conditions of the proposition.
Assume that h : X X is an involution with 2g +2 xed points. Let Y denote
the quotient space X/ < h > and : X Y the natural map. Then : X Y has
2g + 2 ramication points. Since has degree 2, by the Riemann-Hurwitz relation
we have that the genus of Y , say g

, satises
2g 2 = 2(2g

2) + 2g + 2 = 4g

+ 2g 2.
So g

= 0, i.e., Y is the Riemann sphere and : X



C is a meromorphic function
of degree 2. In other words, X is hyperelliptic.2.9. JACOBIAN VARIETIES OF RIEMANN SURFACES 115
2.8.16. Corollary (of the proof). If X is a hyperelliptic compact surface
of genus greater than 1 then the xed points of the hyperelliptic involution are the
Weierstrass points of X.
Proof. It follows from the proofs of the previous results and we leave it as an
exercise to the reader.
2.8.17. Corollary. Let X be a hyperelliptic compact surface of genus g 2.
Then the hyperelliptic involution J is the unique involution of X with 2g + 2 xed
points.
Proof. Let f : X

C be a degree 2 function and let J : X X be
the hyperelliptic involution constructed from f. We have that f J = f. Let
J
1
: X X be another involution with 2g + 2 xed points. Then f J
1
: X

C
is a function of degree 2, so by 2.8.13 there exist a Mobius transformation M such
that M f = f J
1
.
By 2.8.15 we have that the xed points of J
1
are the Weierstrass points of X.
For any of these points, say p, we have M(f(p)) = f(J
1
(p)) = f(p). Hence M has
2g +2 xed points, so it must be the identity (see corollary 3.3.5). This implies that
f J
1
= f; but this is precisely the condition that denes J, so J = J
1
.
2.9. Jacobian Varieties of Riemann Surfaces
In 1.3.6 we saw that a surface given by C/G

has genus 1. In this section we will


prove the converse statement, namely any surface of genus 1 is of the form C/G

.
We will also show that any compact surface (of positive genus) can be embedded
into a higher dimensional torus; that is, the quotient of C
n
by a nice group of
translations. This torus, known as the Jacobian variety of the surface, is related to
the Picard group, the group of equivalence classes of line bundles on the surface.
2.9.1. Throughout this section X will denote a compact Riemann surface of
positive genus. We know from Topology that X is homeomorphic to a sphere with
g handles attached (see gure 2 for a surface of genus 2). The homotopy classes
of the paths a
j
and b
j
, j = 1, . . . , g generate
1
(X, x
0
), and the corresponding
homology classes form a set of generators of H
1
(X, Z). Denote by D this collection116 2. COMPACT RIEMANN SURFACES
of paths. If we cut X along D (i.e. we consider X\{a
1
, . . . , a
g
, b
1
, . . . , b
g
}) we get
b
1
a
2
a
1
2
a
1
b
1
2
a
1
1
b
1
1
b
2
Figure 10. Polygon corresponding to a surface of genus 2.
a surface homeomorphic to a polygon with 4g sides. Identifying the sides in the
boundary of (see gure 10 for an example of a surface of genus 2) we recover X
(as a topological manifold). Let : X be the natural quotient map. Since
X\D is simply connected the mapping :

X\D, is a homeomorphism. All


these facts are well known and can be found, for example, in [13].
2.9.2. Let denote a smooth 1-form dened on a neighbourhood of D. We
set
A
j
() =
_
a
j
, B
j
() =
_
b
j
,
which are called the a- and b-periods of , respectively.
Let z denote the identity function on the complex plane. We can dene a smooth
form on as an expression of the type f(z) dz, where f is a smooth function on the
interior of , and with directional derivatives in the boundary points. The standard
result of Dierential Geometry apply to this setting (manifolds with boundary). If
is a smooth 1-form on X we can use the mapping : X to get a form on
as follows. For a point z
0
in

let w
0
= (z
0
). Since is a homeomorphism in

we
have that w = z
1
is a coordinate on X\D. The form will be then given by
= f(w) dw; it is easy to check that f(z) dz denes a form, on

(the lift of via


), say

, which can be extended to . The main property of

, as one can check


using local coordinates, is that if is a path in , then
_

=
_
()
. Details2.9. JACOBIAN VARIETIES OF RIEMANN SURFACES 117
can be found in a book on Dierential Geometry. Observe that this construction
is possible also if is dened only in a neighbourhood of the paths a
1
, . . . , b
g
(in
which case it yields a smmooth form in a neightborhood of ).
Let be a form on X; x a point z
0
in the interior of , and dene a function u
on by u(z) =
_
z
z
0

, where the integral is computed over a path in joining z


0
and z. Since the polygon is simply connected the function u is well dened (the
integral does not depend on the choice of path).
With this notation we have that the expression
_

makes sense.
Lemma.
_

=
g

j=1
_
A
j
()B
j
() A
j
()B
j
()
_
.
Proof. Let p be a point in a
j
and denote by p

the equivalent point in a


1
j
as
in gure 11 (that is, p and p

project to the same point in X). Consider a path


with end points p and p

, like in the gure. Clearly u(p) u(p

) =
_

. Since ()
is homologous to b
1
j
we have u(p) u(p

) = B
j
(). Similarly, for p in b
j
we get
u(p) u(p

) = A
j
(). By an abuse of notation we will use a
j
and b
j
for the paths
in X and the boundary curves of . Then
_

=
g

j=1
_
_
a
j
+
_
b
j
+
_
a
1
j
+
_
b
1
j
_
u

=
=
g

j=1
_
a
j
(u(p) u(p

))

+
g

j=1
_
b
j
(u(p) u(p

))

=
=
g

j=1
B
j
()
_
a
j

+
g

j=1
A
j
()
_
b
j

=
=
g

j=1
(A
j
()B
j
() A
j
()B
j
()) .
2.9.3. Proposition. Let X be a compact Riemann surface of genus g 1
and H
0
(X, ). Then
Im
g

j=1
A
j
()B
j
() < 0.118 2. COMPACT RIEMANN SURFACES
p
p

a
1
j
a
j

Figure 11. Proof of lemma 2.9.2.


Proof. Let u =
_
d

as above. By Stokes Theorem


_

u =
_

d(u

) =
_

=
_
X
.
If is given by = f dz (locally) we have
= fdz fd z = |f|
2
dz d z = 2i|f|
2
dx dy.
Hence
1
2i
_

u =
_
X
|f|
2
dxdy < 0.
Applying the previous lemma we get
0 >
1
2i
_

=
1
2i
g

j=1
(A
j
()B
j
() A
j
()B
j
()) =
=
1
2i
g

j=1
_
A
j
()B
j
() A
j
()B
j
()
_
=
1
2i
g

j=1
2 i Im
_
A
j
()B
j
()
_
.
2.9.4. Corollary. Let H
0
(X, ). If either
_
a
j
= 0, or Re
_
a
j
= Re
_
b
j
= 0,
for all j = 1, . . . , g, then 0.
Proof. The rst condition is simply A
j
() = 0, for j = 1, . . . , g. To see that
the second condition implies that 0 all one needs is the identity
Im(A
j
() B
j
()) = Re(A
j
()) Im(B
j
()) + Im(A
j
()) Re(B
j
()).2.9. JACOBIAN VARIETIES OF RIEMANN SURFACES 119
In both cases we get a contradiction with the previous theorem.
2.9.5. Corollary. Let {
1
, . . . ,
g
} be a basis of H
0
(X, ), then the matrix
A = (A
jk
) = (
_
a
j

k
)
j,k=1,...,g
is invertible.
Proof. Suppose c = (c
1
, . . . , c
g
) is a vector in C
g
satisfying Ac
t
= 0. Then

g
k=1
a
jk
c
k
= 0, for j = 1, . . . , g. In other words, the a-periods of the form =
c
1

1
+ + c
g

g
are all zero. From the previous corollary we get that 0, and
since the forms
1
, . . . ,
g
are linearly independent we have c
1
= = c
g
= 0.
2.9.6. We will say that a basis {
1
, . . . ,
g
} of H
0
(X, ) is normalised if A
is the identity matrix, i.e.
_
a
j

k
=
jk
. By Stokes theorem (1.4.22) the normalised
basis does not change if we change the paths in D by other loops in the same
homology classes. In particular we can assume that the paths (in D) avoid a xed
point, if needed. From now onwards we will assume that we are working with a
normalised base of holomorphic forms.
Theorem (Riemanns Bilinear Relations). Let X be a compact Riemann surface
of genus g 1. Let D be as above and {
j
}
g
j=1
a normalised basis of holomorphic
1-forms. Set B
jk
=
_
a
j

k
, j, k = 1, . . . , g, and B = (B
jk
). Then the matrix B is
symmetric and has positive denite imaginary part.
Proof. Set u
j
=
_
p
p
0

j
as above. By Stokes Theorem (1.4.22)
_

u
j

k
=
_
X

j

k
= 0.
On the other hand, by 2.9.2,
_

u
j

k
= B
j
(
k
) B
k
(
j
),
which proves that B is symmetric.
To prove the second statement, let c
1
, . . . , c
g
be real numbers, not all equal to 0,
and set =

g
j=1
c
j

j
. By 2.9.3 we have
Im
g

j=1
A
j
()B
j
() < 0.120 2. COMPACT RIEMANN SURFACES
Since A
j
() = c
j
are real numbers, we get
Im
g

j=1
A
j
()B
j
() = Im
g

j=1
g

k=1
c
j
c
k
B
j
(
k
) =
=
g

j=1
g

k=1
c
j
c
k
Im(B
jk
) < 0.
2.9.7. To give a more concrete feeling of the above theorem consider the case
of X given by X = C/G

(with the notation of 1.3.6). One can take to be the


polygon of vertices 0, 1, and 1 + in the complex plane. Then a = a
1
is given by
the path a(t) = t, 0 t 1, and b = b
1
by b(t) = t, 0 t 1. The dierential dz
is normalised since
_
1
0
dz = 1. The statement that B is symmetric is trivial, since
B is simply the complex number given by
_

0
dz = . We see that Im() is positive.
2.9.8. Given two distinct points of X, say p and q, we have seen that there
exists a meromorphic form on X whose only singularities are simple poles at p
and q, with residues 1 and 1 respectively (2.5.24). By adding a holomorphic form
to we can assume that the a-periods of this latter form are all 0 (by corollary 2.9.5
the a-periods of a holomorphic form can be arbitrarily chosen). We will denote this
normalised meromorphic form by
pq
.
Theorem (Reciprocity Theorem).
_
b
j

pq
= 2i
_
p
q

j
.
Proof. By 2.9.2 we have
_

u
j

k
=
g

k=1
A
k
(
j
)B
k
(
pq
) A
k
(
pq
)B
k
(
j
) = B
j
(
pq
) =
_
b
j

pq
.
Since the interior of is a simply connected set, by the Residue theorem we get
_

u
j

pq
= 2i(u
j
(p) u
j
(q)) = 2i
_
p
q

j
.2.9. JACOBIAN VARIETIES OF RIEMANN SURFACES 121
2.9.9. On C
g
consider the group of translations of the form z z +

,
where H
1
(X, Z) and

= (
_


1
, . . . ,
_


g
). The vectors

corresponding to
in D are given by (recall that the basis of H
0
(X, ) is assumed to be normalised)

a
1
= e
1
= (1, 0, . . . , 0), . . . ,
ag
= e
g
= (0, . . . , 0, 1),

b
1
= B
1
= (B
11
, . . . , B
1g
), . . .
bg
= B
g
= (B
g1
, . . . , B
gg
),
where B
jk
=
_
b
j

k
. For any other homology class, represented by a path , we
have

= n
1
e
1
+ + n
g
e
g
+ m
1
B
1
+ + n
g
B
g
, where n
j
and m
j
are integers.
By Riemanns Bilinear Relations the vectors {e
1
, . . . , e
g
, B
1
, . . . , B
g
} are linearly
independent over R, and is a discrete free abelian group of rank 2g ( is isomorphic
to Z
2g
). Some times we will identify the transformation z z +

with the vector

. The quotient space


J(X) := C
g
/
is called the Jacobian variety of X. It is the higher genus analogous of the case of
C/G

. It can be easily proved that J(X) is a complex manifold of dimension g (the


1-dimensional proof of 1.3.6 generalises without any diculty). Since the elements
of preserve the addition of points of C
n
we have that the Jacobian variety is an
abelian group.
2.9.10. One can dene a mapping, the Abel-Jacobi map A : X J(X), by
choosing a point, called the base point of the map, say x
0
X, and setting
A(x) =
__
x
x
0

1
, . . . ,
_
x
x
0

g
_
mod .
The notation above means that we compute the integrals over an arbitrary path
joining x
0
and x and then take the corresponding class in J(X) (this is the meaning
of mod in the above expression). If
1
and
2
are two such paths, then
1

2

H
1
(X, Z) so the point A(x) does not depend on the choice of path. We can extend
A to divisors by
A
_
r

j=1
n
j
x
j
_
=
r

j=1
n
j
A(x
j
),
which makes sense because J(X) is a group, as we have remarked above.
Let A
1
denote the Able-Jacobi mapping with another choice of base point. Then
we have that action of A
1
on points of X is given by the composition of the old122 2. COMPACT RIEMANN SURFACES
mapping A with a translation on J(X). The mapping induced by A
1
on divisors of
degree 0 is then equal to the mapping induced by A.
2.9.11. Theorem (Abel). With the above notation, a divisor D is principal
if and only if deg D = 0 and A(D) = 0 in J(X).
Proof. First of all, since meromorphic functions have as many zeroes as poles
(counted with multiplicity) the condition of deg D = 0 is natural.
Recall that if f is a meromorphic function dened on a neighbourhood of 0 in
the complex plane we can write f(z) = z
n
g(z), with g(z) = 0 (in a, perhaps smaller,
neighbourhood of 0). Then
f

(z)
f(z)
=
nz
n1
g(z) + z
n
g

(z)
z
n
g(z)
=
n
z
+
g

(z)
g(z)
;
so f

/f has a pole of order 1 at z = 0 with residue n.


Write D =

r
j=1
p
j

r
j=1
q
j
, where p
j
= q
k
, j, k = 1, . . . , r (although we may
have repetitions among the p
j
s or q
j
s). Assume f is a meromorphic function on X
with div(f) = D. Since the form df/f has poles at the points p
j
and q
j
it will be
of the form:
df
f
=
r

j=1

p
j
q
j
+
g

k=1
c
k

k
,
for some complex numbers c
k
.
If is a loop in X that avoids the points of D then
_

df/f 2iZ. This can


be easily seen by lifting to (the integral of df/f is log(f)).
Conversely, if is a dierential of the form =

j

p
j
q
j
+

k
c
k

k
, satisfying
that
_

is in 2iZ for all closed paths, then we can dene a function f on X by


f(x) = exp(
_
x
x
0
). It is easy to check that f is well dened, and div(f) = D.
Any form as above will have periods in 2iZ if and only if all its a- and b-
periods are in 2iZ. Thus we have reduced the statement of the theorem to the
following:
A divisor D =

r
j=1
p
j
q
j
is principal if and only deg D = 0 and there exist
complex numbers c
1
, . . . , c
g
, such that the form =

r
j=1

p
j
q
j
+

g
k=1
c
k

k
has a-
and b-periods in 2iZ.2.9. JACOBIAN VARIETIES OF RIEMANN SURFACES 123
By the normalisations made we have A
j
() = c
j
, and by the Reciprocity Theo-
rem
B
j
() =
r

k=1
2i
_
p
k
q
k

j
+
g

k=1
c
k
B
jk
.
Thus we see that
_

2iZ if and only if there exist integers, n


1
, . . . , n
g
, and
m
1
, . . . , m
g
, satisfying
_

_
2in
j
= c
j
,
2im
j
=

r
k=1
_
p
k
q
k

j
+

g
k=1
c
k
B
jk
.
Let denote the vector (
1
, . . . ,
g
), and
_
p
q
= (
_
p
q

1
, . . . ,
_
p
q

g
). We have that
the above two conditions are equivalent to
r

k=1
_
p
k
q
k
=
g

k=1
n
k
B
k
+
g

k=1
m
k
e
k
,
where e
k
and B
k
are as in 2.9.9. This equation is clearly equivalent to A(D) = 0
mod .
2.9.12. Theorem. If X is a compact Riemann surface of genus g 1 then
the Abel-Jacobi map A : X J(X) is an embedding.
Proof. We rst show that A is injective. This is an easy consequence of the
fact that g 1. If A(p) = A(q), for two distinct points p and q of X, then D = pq
would be a principal divisor. So we would have a meromorphic function f : X

C
with divisor div(f) = p q. This implies that X is homeomorphic to the Riemann
sphere

C (which have genus equal to 0), a contradiction.
The second point that we need to prove is that the derivative of A is not zero at
any point of X. Choose p X, and a local coordinate z near p, with z(p) = 0. The
elements of the normalised basis {
1
, . . . ,
g
} can be written as
j
= f
j
dz, where f
j
are holomorphic functions in a neighbourhood of p. Then the mapping A is given
by
A(q) =
__
q
p

1
, . . .
_
q
p

g
_
=
__
q
p
f
1
dz, . . . ,
_
q
p
f
g
dz
_
;
so its derivative is simply
dA(p) = (f
1
(0), . . . , f
g
(0)) .124 2. COMPACT RIEMANN SURFACES
If dA(p) = (0, . . . , 0) then
j
vanishes at p for all j = 1, . . . , g. But by 2.7.2 we
know that this cannot happen.
2.9.13. Corollary. If X is a compact Riemann surface of genus 1 then the
Abel-Jacobi mapping is biholomorphic.
Proof. J(X) is a complex manifold of (complex) dimension g, so when g = 1
we have that J(X) is a Riemann surface. The Abel-Jacobi mapping A thus become
a holomorphic mapping between two compact surfaces. Since A is not constant it
must be surjective (1.3.12); therefore it is a biholomorphic mapping (it is injective
by the previous result).
This corollary simply says that any surface of genus 1 is of the form C/G

.
2.9.14. In the case of higher genus, g 2, we cannot have a result like the
above corollary, since X has complex dimension 1 while J(X) has dimension g.
However, it is possible to construct a g-dimensional manifold from X, S
g
(X), and
a generalisation of the Abel-Jacobi mapping A : S
g
(X) J(X) that is surjective.
To do this consider the symmetric group S
g
of permutations of g letters. We have
that S
g
acts on X
g
= X X as follows: for S
g
we dene
X
g
(x
1
, . . . , x
g
)

(x
(1)
, . . . , x
(g)
) X
g
.
Let S
g
(X) = X
g
/S
g
denote the quotient space. One can consider S
g
(X) simply as
the set of eective divisors of degree g in X. From the above description it is easy
to put a complex structure to S
g
(X). We will not pursue further this point of view
here; the result we are interested is as follows.
Theorem (Jacobi Inversion Problem). The Abel-Jacobi map A : S
g
(X) J(X)
is surjective.
2.9.15. Before giving a proof of the above theorem we need a technical result.
If X has genus g 1, and D is an eective divisor of degree g, by the Riemann-Roch
theorem we have
dim H
0
(X, O(D)) = 1 + dim H
0
(X, (D)) 1.2.9. JACOBIAN VARIETIES OF RIEMANN SURFACES 125
We say that D is special if dim H
0
(X, O(D)) > 1. Observe that D is not special if
and only if the only holomorphic 1-form that vanishes at D is the zero form.
Lemma. Let D S
g
(X), where X is a compact Riemann surface of genus g 1.
Write D = p
1
+ +p
g
, and let U
j
, 1 j g, be a neighbourhood of p
j
in X. Then
there exists a divisor E = q
1
+ + q
g
, with q
j
U
j
, such that E is not special.
Moreover, one can choose E with q
j
= q
k
(for j = k).
Proof. Set D
0
= 0, and D
j
= p
1
+ + p
j
, for 1 j g. Using an induction
argument we will show that it is possible to nd a divisor E = q
1
+ + q
j
, with
q
k
U
k
, consisting of distinct points, and such that dim H
0
(X, (E
j
)) = g j.
When j = g we have the statement of the lemma.
If j = 0 there is nothing to prove. If j = 1, we have D
1
= p
1
, and this case
follows from the Riemann-Roch theorem:
dim H
0
(X, (p
1
)) = dim H
0
(X, O(p
1
)) 1 +g 1 = g 1.
Thus we can take q
1
= p
1
, E
1
= D
1
. Assume now that we have proven our statement
up to some j < g. Applying the Riemann-Roch Theorem for j + 1 we see
dim H
0
(X, (D
j+1
)) g j 1.
Let {
1
, . . . ,
gj
} be a basis of H
0
(X, (D
j
)). Since
j+1
is not identically 0,
there is a point q
j+1
, in U
j+1
, such that
j+1
(q
j+1
) = 0. We can vary q
j+1
is a small
neighbourhood of p
j+1
, so we can take q
j+1
= q
k
, for 1 k j. Thus
dim H
0
(X, (D
j+1
)) g j 1,
and the lemma is proved.
2.9.16. Proof of the Jacobi Inversion Problem. Choose a non-special
divisor D = p
1
+ + p
g
, with p
j
= p
k
(if j = k). For each point p
j
take a local
coordinate (U
j
, z
j
), with z
j
(p
j
) = 0. If {
1
, . . . ,
g
} is a normalised basis of H
0
(X, )
we write
j
= f
jk
dz
k
on U
k
. Then the Abel-Jacobi map (in U
1
U
g
) is given
by
(z
1
, . . . , z
g
) A(D) + (A
1
(z
1
, . . . , z
g
), . . . , A
g
(z
1
, . . . , z
g
)),126 2. COMPACT RIEMANN SURFACES
where
A
j
(z
1
, . . . , z
g
) =
g

k=1
_
z
k
0
f
jk
(z
k
)dz
k
.
Since (A
j
/z
k
) (0, . . . , 0) = f
jk
(0), we have that the Jacobian of A at the origin is
given by
_
_
_
_
_
f
11
(0) f
1g
(0)
.
.
.
.
.
.
f
g1
(0) f
gg
(0)
_
_
_
_
_
Because D is not special we have dim H
0
(X, (D)) = 0; it is easy to see that
this fact implies that the above matrix has rank g. Hence, by the Inverse Function
theorem, A : V
1
V
g
A(D) + U is a homeomorphism, where V
j
can be
considered as a neighbourhood of 0 in C, and U is a neighbourhood of (0, . . . , 0) in
C
g
.
Our goal is to show that A : S
g
(X) C
g
(the lift of A to the universal cover of
J(X)) is surjective. Let c = (c
1
. . . , c
g
) C
g
. Then there exists an integer n, such
that A(D) + (c/n) belongs to A(D) + U. Thus we have an eective divisor D

of
degree g, such that A(D

) = A(D) + (c/n). Or equivalently, n(A(D

) A(D)) = c.
Set D
1
= nD+nD

+gp
0
, where p
0
is the base point chosen to dene the map A.
By the Riemann-Roch theorem,
dim H
0
(X, O
D
1
) = 1 + dim H
0
(X,
D
1
) 1.
Let f : X

C be a meromorphic function with (f) D
1
. Write (f) = E D
1
.
Then E is an eective divisor of degree g. By Abels Theorem
0 = A((f)) = A(E) n(A(D

) A(D)) = A(E) c,
since A(p
0
) = 0. This completes the proof of the theorem.
CHAPTER 3
Uniformization of Riemann surfaces
3.1 The Dirichlet Problem on Riemann surfaces 128
3.2 Uniformization of simply connected Riemann surfaces 141
3.3 Uniformization of Riemann surfaces and Kleinian groups 148
3.4 Hyperbolic Geometry, Fuchsian Groups and Hurwitzs Theorem 162
3.5 Moduli of Riemann surfaces 178
127128 3. UNIFORMIZATION OF RIEMANN SURFACES
One of the most important results in the area of Riemann surfaces is the Uni-
formization theorem, which classies all simply connected surfaces up to biholomor-
phisms. In this chapter, after a technical section on the Dirichlet problem (solutions
of equations involving the Laplacian operator), we prove that theorem. It turns out
that there are very few simply connected surfaces: the Riemann sphere, the complex
plane and the unit disc. We use this result in 3.2 to give a general formulation of
the Uniformization theorem and obtain some consequences, like the classication of
all surfaces with abelian fundamental group. We will see that most surfaces have
the unit disc as their universal covering space, these surfaces are the object of our
study in 3.3 and 3.5; we cover some basic properties of the Riemaniann geometry,
automorphisms, Kleinian groups and the problem of moduli.
3.1. The Dirichlet Problem on Riemann surfaces
In this section we recall some result from Complex Analysis that some readers
might not be familiar with. More precisely, we solve the Dirichlet problem; that is,
to nd a harmonic function on a domain with given boundary values. This will be
used in the next section when we classify all simply connected Riemann surfaces.
Harmonic Functions and the Dirichlet Problem
3.1.1. Recall that a real-valued function u : U R, with continuous second
partial derivatives, is called harmonic if u =

2
u
x
2
+

2
u
y
2
= 0.
Lemma. Let U be an open subset of the complex plane and F : U C a
holomorphic function. Then Re(F) and Im(F), the real and imaginary parts of F,
are harmonic functions.
Proof. Write F = u +iv, where u and v are the real and imaginary parts of F
respectively. The Cauchy-Riemann equations says that u
x
= v
y
and u
y
= v
x
. So
we have
u = u
xx
+ u
yy
= (v
y
)
x
+ (v
x
)
y
= v
yx
v
xy
= 0,3.1. THE DIRICHLET PROBLEM ON RIEMANN SURFACES 129
since the second partial derivatives of u commute.
In 3.1.3 we will show a local converse of this result: a harmonic function is locally
the real part of a holomorphic function.
3.1.2. Let U be an open subset of C and f : U C a continuous function
dened on the boundary of U. The Dirichlet problem with data U and f consists
on nding a continuous function u : U R, harmonic on U and such that u = f
on U. As one might expect not every problem has a solution; we give one example
in 3.1.23. The next result shows that we can always nd a (unique) solution for the
Dirichlet problem when the domain U is a disc.
For a complex number z
0
and a positive real number r, we denote by D
r
(z
0
) the
open disc of radius r and centre z
0
, and by D
r
(z
0
) the closed disc. We will write D
r
for D
r
(0).
Theorem. Let R a positive number and f : D
R
R a continuous function.
Set
u(z) =
_

_
1
2
_
2
0
R
2
|z|
2
|Re
i
z|
2
f(Re
i
) d , for |z| < R,
f(z) , for |z| = R.
Then u solves the Dirichlet problem with data D
R
and f.
Proof. For z and complex numbers the function
P(z, ) =
||
2
|z|
2
| z|
2
is the real part of the function
F(z, ) =
+ z
z
,
which is holomorphic for z = . The expression for u in D
R
can be rewritten as
follows:
u(z) =
1
2
_
2
0
P(z, Re
i
)f(Re
i
) d = Re
_
1
2
_
2
0
F(z, Re
i
)f(Re
i
) d
_
=
= Re
_
1
2i
_
||=R
F(z, )f()
1

d
_
.
If z is in D
R
then F is holomorphic since |z| < || = R. Hence u is the real part of
a holomorphic function and therefore harmonic.130 3. UNIFORMIZATION OF RIEMANN SURFACES
Clearly u is continuous in D
R
. To complete the proof of the theorem we need
to show that u is continuous on the boundary of D
R
. Let G(z, ) = F(z, )/. We
have
1
2
_
2
0
P(z, Re
i
) d = Re
_
1
2i
_
||=R
+ z
z
1

d
_
=

||<R
res

G(z, ) = 1.
The function G is considered as a function of , where z is a xed point of D
R
. If
z = 0 then G(, 0) = 1/, so G has only one pole at = 0 with residue 1. On the
other hand, if z = 0 we have that G(, z) =
+z
(z)
; in this case, G has two poles, at
0 and z, with residues 1 and 2 respectively. We see that the sum of the residues
of G is equal to 1.
Let
0
be a point in D
R
, and > 0. Since f is continuous there exists a positive
number M, such that |f()| M, for all U. For z D
R
we have
u(z) u(
0
) = u(z) f(
0
) =
1
2
_
2
0
P(z, Re
i
)
_
f(Re
i
) f(
0
)
_
d.
By the continuity of f at
0
there exists a
0
> 0, such that |f() f(
0
)| < , if
U satises |
0
| <
0
. We partition the boundary of the disc D
R
into two
disjoint sets, A and B, where
A = { [0, 2]; |Re
i

0
| <
0
},
and B = [0, 2]\A. The set A consists of the angles that are close to the point
0
and B is its complement in the unit circle. We have
|u(z) f(
0
)|

1
2
_
A
P(z, )(f() f(
0
))d

+
+

1
2
_
B
P(z, )(f() f(
0
)) d

+
M

_
B
P(z, Re
i
) d.
The number in the above inequality comes from the fact that f() f(
0
) is small
for points in A and the total integral of P over the boundary of D
R
is equal
to 1. The bound of the second integral comes from the bound M of |f| and the
fact that P(z, ) > 0, for || > |z|. Let now z be in D
R
and close to
0
; that is,
|
0
z| <
0
/2. For B we have that
|Re
i
z| |Re
i

0
| |
0
z|

0
2
.3.1. THE DIRICHLET PROBLEM ON RIEMANN SURFACES 131
On the other hand,
R |z| = |
0
| |z| |
0
z| < .
Using these inequalities we have
P(z, Re
i
) =
R
2
|z|
2
|Re
i
z|
2

(R +|z|) (R |z|)
(
0
/2)
2

8R

2
0
.
We can now bound the above integral as follows:
M

_
B
P(z, Re
i
) d =
M

8R

2
0
2 =
16MR

2
0
.
For given the value of
0
is xed, so we can make small enough such that for we
have |u(z) u(
0
)| 2, for z as above. This shows that u is continuous at
0
.
The function P is called the Poisson kernel.
3.1.3. In the above proof we have shown the following result.
Corollary. If u is harmonic then u is locally the real part of a holomorphic
function.
3.1.4. Corollary. Let u : D
R
R be a harmonic function. Then u satises
u(0) =
1
2
_
2
0
u(re
i
)d,
for 0 < r < R.
Proof. Apply the above theorem on D
r
for the boundary values given by
u : D
r
R and observe that P(z, 0) = 1.
3.1.5. Corollary (Mean Value Property). Let u : D
R
R be a harmonic
function. Let z
0
be a point in D
R
and r > 0 a positive number such that D
r
(z
0
) is
contained in D
R
. Then
u(z
0
) =
1
2
_
2
0
u(z
0
+ re
i
) d,
for 0 < r < R.132 3. UNIFORMIZATION OF RIEMANN SURFACES
3.1.6. From corollary 3.1.3 one expects that harmonic functions share some of
the properties of holomorphic functions. In that sense one can consider the Mean
Value Property as the analogy of Cauchys Integral Formula. In the next result
we see that the Maximum Modulus Principle (1.1.9) is also satised by harmonic
functions.
Proposition (Maximum Modulus Principle). Let u : D
R
R be a harmonic
function. If there exists a point z
0
D
R
, such that u(z) u(z
0
) for all z D
R
,
then u is constant.
Proof. The set
E = {z D
R
; u(z) = u(z
0
)} = u
1
(u(z
0
))
is closed since u is a continuous function. Let z be an arbitrary point of E and r > 0
such D
r
(z) is contained in D
R
. From the Mean Value Property we get
u(z) =
1
2
_
2
0
u(z + re
i
) d
1
2
_
2
0
u(z
0
) d = u(z
0
) = u(z).
This implies that u(z +re
i
) = u(z
0
) for all in [0, 2]. Thus E is an open set. Since
D
R
is connected and E is not empty we have E = D
R
and therefore u is constant
on U.
A similar result with minimum instead of maximum can be obtained from the fact
that if u is a harmonic function then u is also harmonic; we leave the details for
the reader.
3.1.7. Corollary. If u : D
R
R is harmonic on D
R
and continuous on
D
R
then its maximum value is attained in the boundary of D
R
; that is, there exists
a point z
0
D
R
, such that u(z) u(z
0
), for all z D
R
.
Proof. Since D
R
is compact and u continuous there is a value u(z
1
) where u
attains its maximum. If z
1
is in D
R
there is nothing to prove. On the other hand, if
z
1
D
R
, from the previous corollary we have that u is constant in D
R
and therefore
in D
R
. In this case we can choose any point of D
R
as z
0
.
3.1.8. Corollary. If the Dirichlet Problem has a solution on a bounded
domain then the solution is unique.3.1. THE DIRICHLET PROBLEM ON RIEMANN SURFACES 133
Proof. Apply the Maximum Modulus Principle to u
1
u
2
and u
2
u
1
, where u
1
and u
2
are two solutions of the (same) Dirichlet problem. Here the condition of
the domain being bounded is necessary. Consider for example the Dirichlet problem
on the upper half plane with boundary values given by the identically 0 function (on
the real line). Then the contant function 0 and the function Im(z) are two distinct
solutions of this Dirichelt problem.
3.1.9. Using this result we can show that the Mean Value Property (3.1.5) is
also a sucient condition for harmonicity.
Proposition. Let u : U R be a continuous function on an open set U of the
complex plane. Assume u satises the Mean Value Property, namely
u(z
0
) =
1
2
_
2
0
u(z
0
+ re
i
) d,
for all z
0
U and all positive r such that the closed disc of centre z
0
and radius r,
is contained in U. Then u is harmonic.
Proof. Let v be the solution of the Dirichlet problem on D
r
(z
0
) with values
given by the function u. Observe that the proof of the Maximum Modulus Principle
uses only the Mean Value Property. Hence from 3.1.7 we have that v u has its
maximum on D
r
(z
0
); that is v(z) u(z) 0 for all z D
r
(z
0
). Applying the same
argument to the function u v we obtain that u(z) v(z) 0 and this completes
the proof.
3.1.10. Another similarity between harmonic and holomorphic functions is
given by the following result.
Corollary. Let u : D
R
R be a sequence of harmonic functions which con-
verges uniformly on compact subsets of D
R
to a (continuous) function u : D
R
R.
Then u is harmonic.
Proof. For z
0
in D
R
let r > 0 be such that D
r
(z
0
) D
R
. Then we have
u(z
0
) = lim
n
u
n
(z
0
) = lim
n
1
2
_
2
0
u
n
(z
0
+ re
i
)d =
1
2
_
2
0
u(z
0
+ re
i
)d.
For the last equality we have used that u
n
converges uniformly on compact sets to
u, so in particular on the circle of centre z
0
and radius r. It follows from 3.1.9 that
u is harmonic.134 3. UNIFORMIZATION OF RIEMANN SURFACES
3.1.11. The next lemma is needed to prove Harnacks inequality.
Lemma. Let z be a complex number with |z| = s and r a positive number
satisfying s < r. Then
r s
r + s

r
2
s
2
|re
i
z|
2

r + s
r s
,
for any real number .
Proof. To prove the left hand side inequality we use that
|re
i
z| |re
i
| +|z| = r + s.
The other inequality follows from
|re
i
z| |re
i
| |z| = r s.
These two inequalities, together with the expression r
2
s
2
= (r s)(r +s), prove
the result.
Let u : D
R
R
+
be a positive harmonic function and z D
R
a point with |z| = s.
Choose a positive real number r with s < r < R; then
u(z) =
1
2
_
2
0
r
2
s
2
|re
i
z|
2
u(re
i
) d
1
2
r + s
r s
_
2
0
u(re
i
) d =
r + s
r s
u(0).
Proposition (Harnacks inequality). Let u : D
R
R be a positive harmonic
function. Then, for all z
0
D
R
with |z| = s, and for all positive r such that
s < r < R, one has
r s
r + s
u(0) u(z)
r + s
r s
u(0).
Proof. The right hand side inequality was proved before the statement of the
proposition. The proof of the other inequality is similar.
3.1.12. The main application of the above inequality is the proof of the fol-
lowing theorem, which is similar to Montels theorem (1.1.13).
Theorem (Harnacks Principle). Let M be a real number and {u
n
}

n=1
be a
non-decreasing sequence of harmonic functions on D
R
satisfying u
n
M. Then the
sequence {u
n
} converges uniformly on compact subsets of D
R
to a harmonic function
u : D
R
R.3.1. THE DIRICHLET PROBLEM ON RIEMANN SURFACES 135
Proof. The pointwise convergence follows from the fact that u
n
(z) is a bounded,
non-decreasing sequence of real numbers (for xed z). Thus to complete the proof
we only need to show that the convergence is uniform on compact subsets of D
R
. Let
> 0 be given; then there exists an n
0
such that u
n
(0) u
m
(0) , for n
0
m n.
Choose a real number s such that 0 < s < r < R. We apply Harnacks inequality
to the non-negative function u
n
u
m
on the disc D
s
:
u
n
(z) u
m
(z)
r + s
r s
(u
n
(0) u
m
(0))
r + s
r s
.
Since can be made arbitrarily small we get that {u
n
}
n
converges uniformly on D
s
.
But any compact subset K of D
R
is contained in a disc of the form D
s
. Thus u is
harmonic (3.1.10).
3.1.13. The following more general result can be found in [1] (for our applica-
tions the previous version of Harnacks Principle is enough).
Theorem. Consider a sequence of functions u
n
(z), each dened and harmonic
in certain region
n
. Let be a region such that every point in has a neigh-
bourhood in all but a nite number of the
n
, and assume moreover that in this
neighbourhood u
n
(z) u
n+1
(z) for n suciently large. Then there are only two pos-
sibilities: either u
n
(z) tends uniformly to on every compact subset of , or u
n
(z)
converges uniformly on compact subsets of to a harmonic function u : R.
Subharmonic functions
3.1.14. Finding non-trivial harmonic functions on domains is not an easy prob-
lem. What we will do is to consider a more general class of functions, called sub-
harmonic functions, which are close enough to be harmonic; taking limits in this
class we obtain harmonic functions. The precise denition we need is as follows.
Definition. A continuous function u : U R on an open set U of the complex
plane is said to be subharmonic if for every harmonic function h : U R, and
every domain V U, the function u +h either is constant or has no maximum (on
V ).
Suppose V is a domain with compact closure V U. Let h : V R be a continuous136 3. UNIFORMIZATION OF RIEMANN SURFACES
function, harmonic on V . If u is subharmonic (on U) then the maximum of u + h
on V is attained in the boundary of V . The proof is similar to the case of harmonic
functions (3.1.7).
It is clear from the above denition that u is subharmonic if and only if it is locally
subharmonic; that is, every point of U has a neighbourhood where u is subharmonic.
3.1.15. Let D be a disc with D U and u : U R a subharmonic function.
Denote by P
D,u
the function that is equal to u on U\D and solves the Dirichlet
problem on D with boundary values given by u|
D
.
Proposition. A continuous function u : U R is subharmonic if and only if
u P
D,u
for every disc D whose closure is contained in U.
Proof. Assume rst that u is subharmonic on U. For any disc D with D U
we have that u P
D,u
is equal to 0 on D. Since u P
D,u
is continuous on D and
u is subharmonic, either u P
D,u
is identically 0 or it satises u P
D,u
0 on D.
To prove the converse let h : U R be a harmonic function and V U a
domain. Assume that u + h has a maximum value on V , say m
0
. Set
C = {z V ; u(z) + h(z) = m
0
}.
This set is a closed subset of V . Let z
0
be a point of C and D a disc of radius r
with centred at z
0
and such that D V . Then we have
m
0
= u(z
0
) + h(z
0
) P
D,u
(z
0
) + h(z
0
) =
1
2
_
2
0
_
u(z
0
+ re
i
) + h(z
0
+ re
i
)
_
d m
0
.
It follows that u(z
0
+re
i
) +h(z
0
+re
i
) = m
0
for all ; that is, C is an open subset
of V . Since V is connected we have that C = V and u + h is constant on V .
3.1.16. Corollary. Let u : U R be a continuous function. Then u is
subharmonic if and only if for every point z
0
U, and every positive number r such
that D
r
(z
0
) is contained in U, the following inequality holds:
u(z
0
)
1
2
_
2
0
u(z
0
+ re
i
) d.3.1. THE DIRICHLET PROBLEM ON RIEMANN SURFACES 137
Proof. If D = D
r
(z
0
) then the right hand side of the above inequality is simply
P
D,u
(z
0
).
Corollary (Maximum Modulus Principle). Subharmonic functions satisfy the
Maximum Modulus Principle.
Proof. The proof is similar to the case of harmonic functions.
3.1.17. Proposition. Let u, v : U R be subharmonic functions, c a posi-
tive real number, and D U a disc. Then the functions cu, u + v, max(u, v) and
P
D,u
are subharmonic (on U).
Proof. The fact that cu and u + v are subharmonic follows from the above
corollary.
To show that the maximum of two subharmonic functions is subharmonic consider
a point z
0
of U and assume that max(u, v)(z
0
) = u(z
0
); then we have
max(u, v)(z
0
) = u(z
0
)
1
2
_
2
0
u(z
0
+ re
i
)d
1
2
_
2
0
max(u, v)(z
0
+ re
i
)d.
It follows from the previous result that max(u, v) is subharmonic.
Consider now the function P
D,u
. Clearly this function is subharmonic on U\D (since
it is equal to u on this set) and on D (because it is harmonic). So we need to check
subharmonicity only at the points on the boundary of D. Let z
0
be one such point;
using the inequality u P
D,u
(3.1.15) we have
P
D,u
(z
0
) = u(z
0
)
1
2
_
2
0
u(z
0
+ re
i
) d
1
2
_
2
0
P
D,u
(z
0
+ re
i
) d.
We can now apply the previous corollary.
3.1.18. For functions of class C
2
(that is, with continuous partial derivatives
of second order) we have another characterisation of subharmonicity as follows.
Proposition. Let u : U R be a C
2
function. Then u is subharmonic if and
only if u 0 on U.
This result is taken some times as the denition of subharmonic functions. However a
function does not need to have partial derivatives in order to satisfy denition 3.1.14.138 3. UNIFORMIZATION OF RIEMANN SURFACES
3.1.19. Harmonic functions on Riemann surfaces were dened in 1.4.11. Since
a harmonic function u : U R, dened on a domain of the complex plane, is locally
the real part of a holomorphic function one sees that harmonicity is preserved under
changes of coordinates (recall that to compute the partial derivatives of a function on
Riemann surface we need to take local coordinates). However, to dene subharmonic
functions we need a little more of extra work. We begin with a denition.
Definition. A disc on a Riemann surface X is a domain D such that there
exists a local coordinate patch (U, z) with D U and z(D) is a closed disc on C.
Given a disc D on X, and a continuous function u : X R, we can dene P
D,u
:
X R as in 3.1.15.
Definition. A continuous function u : X R dened on a Riemann surface
X is called subharmonic if for every disc D of X, and every harmonic function on
D satisfying u h one has that u h of u < h (on D).
It is easy to show that for the case of X = C this denition is equivalent to 3.1.14.
Proposition 3.1.15 and corollary 3.1.16 extend to Riemann surfaces with similar
proofs. Therefore we see that talking of the Dirichlet problem and solutions of it on
Riemann surfaces makes sense.
In a more invariant way we have the following denition.
Definition. Let X be a Riemann surface, V X an open set of X and
u : V R a real-valued function dened on V . We say that u is harmonic
(respectively subharmonic) if for any local patch (U, z) with U V = , the
function
(u z
1
) : z(U V ) R
is harmonic (respectively, subharmonic).
Perrons method
3.1.20. The idea of Perrons method to nd harmonic functions consists of
taking a family of subharmonic functions that satisfy certain conditions and then
show that the pointwise supremum of such a family must be harmonic.3.1. THE DIRICHLET PROBLEM ON RIEMANN SURFACES 139
Theorem (Perrons method). Let U be a domain of the complex plane and
f : U R a bounded function. Denote by M the family of subharmonic functions
u : U R satisfying limsup
zz
0
u(z) f(z
0
), for all z U. Then the supremum
of the family M is a harmonic function.
Proof. First of all, observe that if |f(z)| K for z in U then v(z) K for
all z in U and all v in M (this is simply a consequence of the Maximum Modulus
Principle for subharmonic functions (3.1.16)).
It is easy to see that the family M has the following properties:
1. If u
1
and u
2
belong to M so does max(u
1
, u
2
).
2. If u M, and D is a disc contained in U then P
D,u
is in M.
Fix a point z
0
U and let D be a disc containing z
0
and satisfying D U.
Then there exists a sequence of functions {u
n
}
n
in M (the sequence may depend
on the point z
0
) such that limsup
n
u(z)n = u(z
0
). Let us dene functions v
n
by
v
n
= max(u
1
, . . . , u
n
). The sequence {v
n
}
n
is clearly non-decreasing and contained
in M (because of property 1 above). If we set w
n
= P
D,vn
we have that w
n
belongs
to M (property 2). Moreover the following inequalities hold:
u
n
(z
0
) v
n
(z
0
) w
n
(z
0
) u(z
0
).
So lim
n
w
n
(z
0
) = u
n
(z
0
). Let w be the limit of the sequence {w
n
}
n
. Then w is
harmonic on D by Harnacks principle and w u with w(z
0
) = u(z
0
).
Consider now another point of U, say z
1
, and let let u

n
be a sequence and D

a
disc similar to the ones considered above. We set

u

n
= max(u
n
, u

n
) and repeat the
above process to obtain a function w

satisfying w w

u, and w(z
1
) = w

(z
1
).
But then w(z
0
) = w

(z
0
) (since w(z
0
) w

(z
0
) u(z
0
) = w(z
0
)). So w w

on D

.
Thus w is harmonic on the domain U.
3.1.21. Lemma. Let U be a domain in the complex plane and z
0
a point of
U. Assume that there exists a continuous function : U [0, +) such that
(z
0
) = 0 and (z) > 0, for all z U\{z
0
}. If f : U R is a bounded function,
continuous at z
0
and M is as in theorem 3.1.20, then lim
zz
0
u(z) = f(z
0
), for
z U.
Proof. It suces to show that140 3. UNIFORMIZATION OF RIEMANN SURFACES
limsup
zz
0
u(z) f(z
0
) + , and liminf
zz
0
u(z) f(z
0
) ,
for > 0 arbitrary, z
0
U and z U.
Let W be a neighbourhood of z
0
such that |f(z)f(z
0
)| < for z W. Consider
the function
g(z) = f(z
0
) + +
(z)

0
_
K f(z
0
)
_
,
where
0
> 0 is the minimum of the harmonic function u on U\(W U). For z in
W we have g(z) f(z
0
) +, while for z not in W we see that g(z) K + > f(z).
By the Maximum principle we have that if v M then v < g. Thus u g, which
implies that limsup
zz
0
u(z) g(z
0
) f(z
0
) + .
The second inequality is proven in a similar way by using the function
h(z) = f(z
0
)
(z)

0
_
K + f(z
0
)
_
.
3.1.22. The function in the above lemma is called a barrier at z
0
. It is
clear that if every point of the boundary of U has a barrier then we can solve the
Dirichlet problem for that domain. One would like to have geometric conditions on
a domain so that we can easily see the existence of barriers at its boundary points.
An easy example is given by the upper half plane U = H = {z C; Im(z) > 0}.
Take any point, say z
0
= 0. Then (z) = Im(e
i/2
z) is a barrier at the origin. More
generally, let z
0
U and let z
1
denote a point not in U. Denote by [z
0
, z
1
] the
segment joining these two points and assume that [z
0
, z
1
] U = {z
0
}. Then the
function
(z) = Im
__
z z
0
z z
1
_
is a barrier at z
0
, for a proper choice of the square root.
3.1.23. We end this section with an example of a Dirichlet problem that has
no solution. Consider the open set U = D

= {z C : 0 < |z| < 1}, the punctured


unit disc, and the function f dened on U = S
1
{0} by f(0) = 1, f(z) = 0, for
|z| = 1. If u were a solution for the Dirichlet problem with this data, then u would
have its maximum (it must have a maximum since U is compact) at the boundary of3.2. UNIFORMIZATION OF SIMPLY CONNECTED RIEMANN SURFACES 141
the disc (0 is an interior point of U). But this would imply that u 0, contradicting
the fact that u(0) = 1.
3.2. Uniformization of simply connected Riemann
surfaces
We have seen that a compact, simply connected Riemann surface is biholomor-
phic to the Riemann sphere. By the Riemann mapping theorem we have that nay
simply connected open subset of the Riemann sphere is biholomorphic to either
the complex plane or the unit disc. In this section we show that these three sur-
faces are the only simply connected Riemann surfaces, up to biholomorphisms. The
proof assumes only a couple of results from Complex Analysis (that we state at the
beginning) and the theory of harmonic functions; it is based on a paper of R.R.
Simha [23].
3.2.1. Theorem (Koebe). Let A be the class of one-to-one holomorphic func-
tions dened on the unit disc f : D C and satisfying f(0) = 0, f

(0) = 1. Then
A is normal and compact in the topology of uniform convergence on compact subsets
of the disc.
3.2.2. Theorem (Riemann Mapping Theorem). If A is a simply connected
open subset of the complex plane, with C\A not empty, then A is biholomorphic to
the unit disc.
3.2.3. Lemma. Let h : R be a harmonic function dened on an open,
connected set of the complex plane. If there exists an open subset U of , such
that h|
U
is constant, then h is constant (on ).
Proof. Let p
0
be a xed point of U. For any point p of consider a path
: [0, 1] with (0) = p
1
, (1) = p. A harmonic function is locally the
real part of a holomorphic function (3.1.3); that is, for every point q of there
exists a neighbourhood V of q, and a holomorphic function f
V
: V C, such that
h = Re(f
V
) on V . Since the image of is compact we can nd connected open sets,
U
0
, . . . , U
n
, satisfying the following properties:
1. there exist holomorphic functions f
j
: U
j
C, such that h is the real part of142 3. UNIFORMIZATION OF RIEMANN SURFACES
f
j
on U
j
, for j = 0, . . . , n;
2. U
j
U
j+1
= , for j = 0, . . . , n 1.
Since U
0
U is not empty, and h is constant on U, we have that the real part of f
0
is constant on U
0
. But then f
0
must be constant. Similarly we get that f
1
must be
constant on U
1
; in particular, the real part of f
1
, which is equal to h|
U
1
, is constant.
By a nite number of steps we get that h|
Un
is constant and therefore h(p) = h(p
0
);
that is, h is constant on .
3.2.4. Lemma. Let f : C be a holomorphic function dened on a con-
nected open subset of the complex plane. Assume that f is a (branched) covering
map of degree d onto its image f(). Then
_

f d

f = d
_
f()
dz d z.
Proof. Assume rst that = f() = D and f is given by f(z) = z
n
. The proof
in this case is an easy calculation. If we write z = re
i
we have dzd z = 2irdrd
and therefore
_
D
df df =
_
D
nz
n1
dz d z
n1
d z = 2in
2
_
2
0
_
1
0
r
2n1
dr d = 2in.
On the other hand
_
D
dz d z = 2i
_
D
rdr d = 2i.
To prove the general case use the fact that f is a (branched) covering, and therefore
it behaves locally as the function z z
n
studied above.
Remark. The lemma, in a non-formal language, says that the area of f()
counted with multiplicity is equal to the true area of f() multiplied by the
degree of f.
3.2.5. We now prove that an annulus on a Riemann surface is always confor-
mally equivalent to a standard annulus on the complex plane.
Theorem (The Annulus Theorem). Let U be an open subset of R
2
containing
the closed annulus {z C; 1 |z| 2}. Suppose that there exists a Riemann3.2. UNIFORMIZATION OF SIMPLY CONNECTED RIEMANN SURFACES 143
surface structure on U such that the holomorphic functions (in that structure) are
smooth functions of R
2
. Then the open annulus
A = {z C; 1 < |z| < 2},
with the complex structure induced from U, is biholomorphic to a unique annulus
A
R
= {z C; 1 < |z| < R}
with the standard Riemann surface structure induced from C.
Proof. It is easy to see (use 3.1.22) that there exists a barrier at every point
of A so the Dirichlet problem has solution on A. For c a positive real number
let h
c
be the unique solution of the Dirichlet problem with boundary values 0 in
{z; |z| = 1} and c in {z; |z| = 2}. Observe that h
c
is linear on c: if c, d and are
positive numbers, then h
c+d
= h
c
+ h
d
and h
c
= h
c
. It is also easy to see that h
c
is a proper function.
By Sards theorem (1.4.23) the set of points where the function h does have
zero derivative has measure zero. Let t be a regular value (the image of a point
where h has non-zero derivative); then h
1
(t) is a collection of 1 dimensional closed
manifolds. By 1.4.24 these manifolds must be curves dieomorphic to circles. By the
h
1
(t)
Figure 12. h
1
(t).
maximum modulus theorem for harmonic functions we have that none of this circles
can enclosed a disc in A. Otherwise the maximum and minimum of h in that disc
will be achieved in the boundary, where h is constant (with value t), and thus h will
be constant in the whole disc; but then, h will be constant on A by 3.1.7. Similarly,
we cannot have two (disjoint) curves in h
1
(t) that bound an annulus inside A. So144 3. UNIFORMIZATION OF RIEMANN SURFACES
we see that h
1
(t) consists of a single curve, dieomorphic to S
1
, and homotopic to
the boundary curves of A, as in gure 12. We simplify notation and write C
t
for
this curve. We also have that if t < s then C
t
and C
s
bound a cylinder in A, with
C
t
closer to S
1
than C
s
(see the remark after the proof for a formal denition of
closer in this setting).
Consider the integral
c
=
_
Ct
dh
c
where 0 < t < c. From d dh
c
= h
c
we see
that d h
c
is a closed form and therefore
c
is independent of t (1.4.20). Since h
c
depends linearly on c so does
c
. In particular we have that either
c
= 0 for all
c > 0, or
c
+ as c +. If we write dh
c
= (h
c
/x)dx + (h
c
/y)dy,
we have (1.4.7) dh
c
dh
c
= ((h
c
/x)
2
+ (h
c
/y)
2
) dx dy. By the monotone
converge theorem we see that
0 <
_
A
dh
c
dh
c
= lim
0
_
hc(z)c
dh
c
dh
c
,
where we choose a regular value and such that 2 is is also a regular value. By
Stokes theorem this last integral is equal to
lim
0
_
C
c
C
d(h
c
(dh
c
)) = (c )
_
C
dh
c
()
_
C
dh
c
= c
c
.
This implies that
c
= 0. So there exists a unique value of c such that
c
= 2. Set
R = e
c
, and dene a holomorphic function on A by the expression
f(z) = exp
_
h
c
(z
0
) +
_
z
z
0
(dh
c
+ i(dh
c
))
_
,
where z
0
is an arbitrary (but xed) point of A. By our choice of c the periods of
the 1-form dh
c
+ i(dh
c
) are integer multiples of 2i, so f is well-dened. Observe
that |f| 1 as |z| 1, and |f| R when |z| 2. It follows that f : A A
R
is onto and proper. The surjectivity of f is a consequence of the fact that f is an
open mapping (it is holomorphic). To see that f is proper let K A
R
be a compact
set. We can assume that K = {z A
R
; r
1
|z| r
2
}, for 1 < r
1
< r
2
< R,
since any compact subset of A
R
is contained in one such annulus. Let > 0 be such
that 1 + < r
1
and r
2
< R . Then we have that there exists a > 0, such that
|f(z)| 1 +, for z with |z| < 1 +, and |f(z)| R, for z satisfying |z| > 2 .
Hence f
1
(K) is contained in the annulus {z A; 1 + |z| 2 }. Since
f
1
(K) is closed it follows that it is also compact, and therefore f is proper, as we3.2. UNIFORMIZATION OF SIMPLY CONNECTED RIEMANN SURFACES 145
claimed.
From 1.3.11 and exercise 21 we have that f is a (possibly branched) covering map,
of degree d 1. We need to show that d is precisely equal to 1. To see this we
use 3.2.4; rst of all, a simple computation gives
_
A
df d

f =lim
0
_
h(z)c
d(fd

f) = lim
0
i
_
hc(z)c
d(|f|
2
(dh
c
)) =
=2i(R
2
1) =
_
A
R
d d.
Here = dx dy is the standard area form in the plane. The above computation
simply shows that the area of f(A
R
), counted with multiplicity, is the same as
the area of A
R
. Thus f has to be one-to-one.
Remark. The formal way of saying that C
t
is closer to S
1
than C
s
is by saying
that C
t
lies in the annulus bounded by C
s
and S
1
.
3.2.6. Definition. A Riemann surface X is called planar is every smooth
closed 1-form on X with compact support is exact.
It is clear that any simply connected Riemann surface is planar: if is a form on
X, and p
0
is a xed point of X, the expression f(p) =
_
p
p
0
denes a function of
X such that df = . Here the integration is done on a path from p
0
to p; since X
is simply connected we have that the value of this integral does not depend of the
path. It is also clear that any open subset of a planar Riemann surface is planar.
Theorem. Let X be a Riemann surface, K a compact subset of X. Then there
exists a connected open subset U of X, with K U, and a compact Riemann surface
Y such that U is biholomorphic to an open subset of Y . Moreover, if X is planar
then Y can be chosen to be planar.
Proof. Without loss of generality we can assume that K is connected. Choose
a smooth function with compact support, : X R, such that (p) > 0 for all
p K. Let V =
1
((0, +)), and r = inf{(p); p K}. Observe that r > 0
because K is compact. We have that : V R
+
is proper. Let E be the set of
critical points of in V. By Sards theorem (1.4.23) (E) has zero measure in R
+
;
and since is proper on V , this set (E) is closed in R
+
. Therefore there exist two146 3. UNIFORMIZATION OF RIEMANN SURFACES
positive numbers, 0 < r
1
< r
2
< r, such that ([r
1
, r
2
]) E = . Let c be a point in
the interval (r
1
, r
2
), and U the connected component of
1
((c, +)) that contains
K. We will show that U satises the conditions in the statement of the theorem.
First of all the boundary of U is a collection of components of
1
(c). Since
is proper, and c is not a critical value of , it follows from 1.4.24 that U is a nite
collection of curves, {C
i
}
n
i=1
, where each curve C
i
is dieomorphic to the unit circle
S
1
. For each i = 1, . . . , n choose one such dieomorphism, : C
i
S
1
, and extend
it to a smooth function
i
: V
i
S
1
, where V
i
is a neighbourhood of C
i
. One can
easily check that the Jacobian of the mapping
g
i
= (
i
, ) : V
i
S
1
R
is never zero. Therefore there exists a neighbourhood T
i
of C
i
, T
i
V
i
, and a
positive number
i
, such that g
i
: T
i
S
1
(c
i
, c +
i
) is a dieomorphism. Fix
in (0, c); by the Annulus theorem (3.2.5) we have that there exists a biholomorphic
mapping h
i
: g
i
(S
1
(c , c + )) A
R
i
, where A
R
i
= {z C; 1 < |z| < R
i
}.
We can further assume that |h
i
| R
i
near g
1
i
(S
1
{c }) and |h
i
| 1 near
g
1
i
(S
1
{c + }). If that were not the case we only need to compose h
i
with the
mapping z R
i
/z, which interchanges the two components of the annulus A
R
i
.
We can thus use h
i
to attach (smoothly) the disc D
i
= {z C; |z| < R
i
} to U,
obtaining in this way a compact surface Y , that clearly contains a biholomorphic
copy of the set K.
To complete the proof of the theorem we need to show that Y can be chosen to be
planar when X is a planar surface. Let be a closed 1-form with compact support on
Y . Since D
i
is simply connected, closed forms are exact, and therefore there exists a
smooth function f
i
on D
i
, such that w|
D
i
= df
i
. Let a
i
be a positive number; if a
i
is
small enough we can choose a smooth function
i
, with compact support on D
i
, such
that
i
1 on the disc of radius R
i
a
i
(in D
i
), and

i
d(
i
f
i
) is a closed
1-form with compact support on X. Therefore we have that there exists a smooth
function g on X, with

= dg on U. Consider the form

d((1

i
)g).
We have now that

is a closed form with compact support in the disjoint union

i
D
i
; hence

i
, where

i
is a closed form with compact support on D
i
.3.2. UNIFORMIZATION OF SIMPLY CONNECTED RIEMANN SURFACES 147
Using again the fact that D
i
is simply connected we get functions F
i
such that

i
= dF
i
on D
i
, where F
i
are smooth and have compact support on D
i
. Therefore
the form is exact, as we wanted to show.
3.2.7. Theorem. Any planar connected Riemann surface is biholomorphic to
an open subset of

C.
Proof. If X is compact then we have that all forms have compact support.
The planarity condition implies that the space of holomorphic 1-forms have zero
dimension; that is, the surface has genus 0. We have already seen (2.3.5) that X
must be biholomorphically equivalent to the Riemann sphere

C.
Suppose now that X is not compact. Since X is metrizable [3] we can write
it as an increasing union of connected open subsets U
n
, with compact closure. By
theorem 3.2.6 we have that each U
n
is biholomorphic to an open subset of a planar
compact Riemann surface Y
n
, which by the above remarks should be biholomorphic
to

C. So we have a set of holomorphic, one-to-one (not necessarily surjective)
mappings f
n
: U
n
C. Choose a point p in U
1
, and a holomorphic chart (U, z)
around p, with z(p) = 0 and U U
1
. By replacing f
n
by a
n
f
n
+b
n
, where a
n
, b
n
C,
a
n
= 0, we can assume that, for all n:
(1) f
n
(p) = 0; (2) and f
n
(p) = dz(p).
Let K
n
be the set of one-to-one holomorphic functions on U
n
satisfying (1) and
(2). We have K
n
are non-empty sets, and by Koebes theorem (3.2.1) each K
n
is
compact. Hence the product K =

n
K
n
is compact.
The sets E
m
= {(g
1
, g
2
, . . .) K; g
m
|
Un
= g
n
, for n < m} are non-empty and
closed. Since E
m+1
E
m
, we have that the intersection of all the E
m
s is non-
empty. In other words, there exist holomorphic functions g
m
, dened on U
m
, such
that g
m+1
|
Um
= g
m
, on U
m
. So we have a holomorphic function g on X such that
g|
Um
= f
m
, for all m. Clearly g is on-to-one, and therefore it denes a biholomorphic
mapping from X onto an open subset of C.
3.2.8. Theorem (Poincare-Koebe Uniformization Theorem). Any simply con-
nected Riemann surface is biholomorphically equivalent to one (and only one) of the
following three surfaces: the Riemann sphere, the complex plane or the unit disc
(with their standard structures).148 3. UNIFORMIZATION OF RIEMANN SURFACES
Proof. If X is simply connected then it satises the planarity condition, and
therefore X is biholomorphically equivalent to

C or a simply connected open subset
of the complex plane. Using Riemanns Mapping theorem, we have that, in the
latter case, X is biholomorphic to either C or the unit disc D.
It is not dicult to show that the Riemann surfaces

C, C and D are not biholo-
morphic: rst of all, the Riemann sphere is compact but the complex plane and the
unit disc are not. The complex plane and the unit disc are not biholomorphic since
any holomorphic function f : C D must be constant, by Liouvilles theorem.
3.3. Uniformization of Riemann surfaces and
Kleinian groups
In this section we show that any Riemann surface can be written as a quotient

X/G, where

X is a simply connected surface (studied in the previous section). The
elements of the group G are Mobius transformations; we study some properties of
these groups, which are elements of a big class of groups known as Kleinian groups.
3.3.1. Let X be a Riemann surface and :

X X a universal covering.
From Topology (1.1.21) we have that X is homeomorphic to the quotient

X/G,
where G is the group of deck transformations of the covering. The elements of
G are homeomorphisms of

X, and non-identity transformations do not have xed
points in

X. In exercise 7 we asked the reader to show that

X is a manifold. But
we have more than that: there is a (unique) Riemann surface structure on

X such
that becomes a holomorphic mapping. We have left the proof of this fact to
the reader, but we include it here because of its importance. Let p be a point of
X and U an evenly covered neighbourhood of p. By shrinking U if necessary we
can assume that there is a local coordinate dened on it, say z : U z(U) C.
Write
1
(U) =

j
V
j
as a disjoint union of open sets, where |
V
j
: V
j
U is a
homeomorphism. The mapping w
j
= z |
V
j
: V
j
z(U) is a homeomorphism.
We take on

X the atlas consisting of all local coordinates of this form, {(V
j
, w
j
)}.
To show that

X is a Riemann surface we only need to check that the changes of
coordinates are holomorphic mappings. But this is clear since changes of coordinates3.3. UNIFORMIZATION OF RIEMANN SURFACES AND KLEINIAN GROUPS 149
on

X are equal to changes of coordinates on X. More precisely, if (

U, z) is another
local coordinate on X, with

U evenly covered, let (

V , w) be a local coordinate on

X
constructed as above. Assume

V V
j
= ; then we have
w
j
w
1
= z (|
V
j
) (|
e
V
)
1
z
1
= z z
1
,
which is holomorphic since it is a change of coordinates on the Riemann surface
X. Observe that we have taken restrictions of the mapping to sets where it is a
homeomorphism, so we can consider the inverse function ( in general will not have
a global inverse).
The expression of in the above coordinates (U, z) and (V
j
, w
1
) is given by
z w
1
1
= z (|
V
j
) (|
V
j
)
1
z
1
= Id : z(U) z(U),
which shows that is a holomorphic mapping.
The elements of G are homeomorphisms; moreover, they are biholomorphic map-
pings in the above Riemann surfaces structure. To prove this statement consider p
a point of

X and g G. Let p
1
= g(p) and denote by q the point q = (p) = (p
1
).
Choose a local coordinate (U, z) dened in a neighbourhood of q, with U evenly
covered, and let V
0
and V
1
the components of
1
(U) to which p and p
1
be-
long, respectively. We have then local coordinates around these points given by
(V
0
, w
0
= z (|
V
0
)) and (V
1
, w
1
= z (|
V
1
)). Observe that g(V
0
) = V
1
and
|
V
1
g = |
V
0
. To see then that g is holomorphic we need to compose it with these
local coordinates:
w
1
g w
1
0
= z (|
V
1
) g (|
V
0
)
1
z
1
= Id
z(U)
,
which proves our claim.
By the Uniformization theorem for simply connected surfaces (3.2.8) we have
that there exists a biholomorphic mapping f : Y

X, where Y is the Riemann
sphere, the complex plane or the unit disc. The mapping :

X X is a covering if
and only if f : Y X is a covering, so we can assume that

X is one of the three
mentioned surfaces. Putting all these facts together we obtain the general form of
the Uniformization theorem.150 3. UNIFORMIZATION OF RIEMANN SURFACES
Theorem (Uniformization theorem for Riemann surfaces). Let X be a Rie-
mann surface. Then X is biholomorphic to

X/G, where

X is the Riemann sphere,
the complex plane or the unit disc, and G is a group of biholomorphisms of

X,
isomorphic to the fundamental group of X.
Proof. We only need to show that X is biholomorphic to

X/G. From Topology
we have that there exists a homeomorphism between these two surfaces; the proof
that such mapping is actually holomorphic is similar to the above computations so
we leave it to the reader.
3.3.2. Let p
0
X and choose x
0


X satisfying (x
0
) = p
0
. Let U be an
evenly covered neighbourhood of x
0
and
1
=

j
V
j
as above. We have that

1
(p
0
) = {g(x
0
); g G}, and if g and h are distinct elements of G then g(x
0
) and
h(x
0
) belong to dierent V
j
s. Since these sets are disjoint we have that
1
(p
0
) is
a discrete subset of

X (it does not have accumulation points). More precisely, if
there is a sequence of transformations, say {g
n
}
n
with g
n
(x
0
) x
1
, then (x
1
) =
lim
n
(g
n
(x
0
)) = lim
n
(x
0
) = p
0
. The point x
1
belongs to one of the sets V
j
, say
V
j
1
. But then we will have that g
n
(x
0
) V
j
1
for n n
0
, a contradiction with the
fact that restricted to V
j
1
is a homeomorphism.
3.3.3. The group of automorphisms (biholomorphic self-mappings) of the Rie-
mann sphere, Aut(

C), is the group of Mobius transformations, as we have seen in


corollary 1.3.14. We can identied Aut(

C) with a group of matrices (or rather,


equivalence classes of matrices) as follows. Let GL(2, C) denote the group of square
matrices of order 2 with complex coecients and non-zero determinant (equivalently,
the group of invertible linear mappings of C
2
). We dene an equivalence relation
in this group by identifying M
1
and M
2
if there is a non-zero complex number, say
, such that M
2
= M
1
. The quotient space PGL(2, C) = GL(2, C)/ is known
as the projective general linear group. If we consider the subgroup SL(2, C) of
GL(2, C) of matrices of determinant equal to 1, and restrict the relation to com-
plex numbers with || = 1, we obtain a quotient group PSL(2, C) = SL(2, C)/ ,
known as the special projective linear group. Since any non-zero number has a3.3. UNIFORMIZATION OF RIEMANN SURFACES AND KLEINIAN GROUPS 151
square root in C it is not dicult to see that PGL(2, C) and PSL(2, C) are isomor-
phic groups. The identication between Aut(

C) and PSL(2, C) is given by


Aut(

C) PSL(2, C)
A(z) =
az + b
cz + d

_
_
a b
c d
_
_
,
where we use square brackets to denote equivalence classes of matrices in PSL(2, C).
It is easy to see that this mapping is a group homomorphism. From now onwards
we will freely interchange Mobius transformations with (classes of) matrices; for
example we will write the composition of two transformations as AB instead of the
more complicated notation A B.
3.3.4. Consider the map j : PSL(2, C) P
3
dened by
_
a b
c d
_
j
[a : b : c : d].
We can use j to put a topology on the group Aut(

C): a sequence of Mobius trans-


formation {A
n
}
n
converges to the transformation A if and only if j(A
n
) converges to
j(A). It is easy to see that this is equivalent to require that there exist elements of
PSL(2, C),
_
an bn
cn dn
_
and
_
a b
c d
_
, corresponding to A
n
and A respectively, such that
a
n
a, b
n
b and so on.
Although this is the most natural topology of Aut(

C) it does not behave nicely


with respect to the character of the transformations. For example, the sequence
of mappings {A
n
(z) = (1 +
1
n
) i z}
n
converges to A(z) = i z; the transformation A is
a rotation around the origin, it preserves the circles centred at that point, but the
mappings A
n
do not preserve any circle in the complex plane. Another example is
provided by the sequence of transformations {A
n
}
n
given by
A
n
=
_
(n + 1)/n 1
0 n/(n + 1)
_
.
Each of these mappings has two xed points in

C, z
n
=
n(n+1)
2n+1
and . The limit
of this sequence is A(z) = z + 1, which has only one xed point, namely (the
sequence of xed points z
n
converges to the point , so in the limit all xed points
collapse).152 3. UNIFORMIZATION OF RIEMANN SURFACES
3.3.5. The number of xed points can be used to classify Mobius transforma-
tions. We start with an easy lemma.
Lemma. A non-identity Mobius transformation has at least one and at most two
xed points in

C.
Proof. The xed points of the transformation A(z) =
az+b
cz+d
are given by the
solutions of the equation A(z) = z in the Riemann sphere. If c = 0 we have a second
degree equation, az + b = cz
2
+ d, which can have at most two distinct roots (and
it has at least one). On the other hand, if c = 0 we can write A as A(z) = z + ,
where = 0. If = 1 the transformation A xes only the point ; in the case of
= 1 the points and /(1 ) are xed by A.
Corollary. If a Mobius transformation has three xed points then it must be
the identity.
Assume A has only one xed point, say z
0
. If z
0
= , then A is of the form
A(z) = z + . Let S be the transformation S(z) =
1

z (since we are assuming that


A has only one xed point we have = 0); then SAS
1
is given by z z + 1. If
z
0
= , the transformation S
1
(z) =
1
zz
0
satises S
1
AS
1
1
(z) = z + 1.
If A has two xed points, say z
0
and z
1
, let S
2
(z) =
zz
0
zz
1
, where we substitute
a factor (numerator or denominator) by 1 if the corresponding xed point is the
point . It is easy to see that (S
2
AS
1
2
)(z) = z for some complex number , with
= 0, 1.
We can now give a classication of Mobius transformations.
Definition. Let A be a non-identity Mobius transformation. Then A is called
1. parabolic, if it is conjugate to z z + 1;
2. elliptic, if it is conjugate to z z, where || = 1 but = 1;
3. loxodromic, if it is conjugate to z z, where = 0, 1. If is real and
positive A is called hyperbolic.
The above classication can be given in terms of the trace of the transformation as
the following lemma shows. The proof is an easy exercise left to the reader.
Lemma. Let A(z) =
az+b
cz+d
be a Mobius transformation with ad bc = 1. Assume3.3. UNIFORMIZATION OF RIEMANN SURFACES AND KLEINIAN GROUPS 153
A is not the identity transformation. Then:
1. A is parabolic if and only if (a + d)
2
= 2;
2. A is elliptic if and only if (a + d)
2
< 4;
3. A is loxodromic if and only if (a + d)
2
does not belong to the interval [0, 4].
In particular A is hyperbolic if and only if (a + d)
2
> 4.
Observe that A has order 2 if and only if a + d = 0.
3.3.6. We can now look with more detail to some particular cases of the Uni-
formization theorem. We start with the easiest situation, when the universal cover-
ing is the sphere.
Proposition. If X is a Riemann surface whose universal covering space is
(biholomorphic to) the Riemann sphere then X is (biholomorphic to) the Riemann
sphere.
Proof. Non-identity covering transformations do not have xed points, but
any Mobius transformation has at least two xed points, so the covering group of

C X must be trivial.
3.3.7. The next case we consider is that of surfaces covered by the plane. For
a biholomorphic mapping A : C C we have that the point is a removable
singularity when we consider A as a mapping dened on the Riemann sphere (take
local coordinates and write A as a mapping from the punctured unit disc to itself).
If we extend A to

C we have that A is a Mobius transformation xing the point
so it must be of the form A(z) = z + , with = 0 (see also [4, theorem 11.4, pg.
33]). In other words,
Aut(C) = {g(z) = z + ; , C, = 0}.
Assume that X is a Riemann surface covered by C and let G be the group of
covering transformations. Since the elements of G, other than the identity mapping,
cannot have xed points, all transformations of G must then be of the formz z+.
If G is the trivial group then clearly X is the complex plane. Assume now that G
is cyclic; that is, it is of the form G = {A
n
(z) = z+n; n Z}. If we conjugate G by
an automorphism of C, say S, we obtain that C/G and C/SGS
1
are biholomorphic154 3. UNIFORMIZATION OF RIEMANN SURFACES
surfaces. Thus we can assume = 1. It is easy to see then that X is the punctured
plane, C

= {z C; z = 0}, and the covering mapping : C C

is given by the
exponential mapping, (z) = e
2iz
.
Suppose now that G has two generators; by a conjugation we can assume that
A(z) = z + 1 is an element of G. Let B(z) = z + be another element of G, not
in the subgroup generated by A. If = p/q is rational we can assume that p and q
are positive integers, with 0 < p < q and relatively prime. Let r and s be integers
such that r p + s q = 1. Then (A
s
B
r
)(z) = z + (1/q) and G will be cyclic. If is
real but not rational we can write = m+, for some integer n and a positive in
(0, 1). Since the pair {A(z) = z + 1, (A
m
B)(z) = z + } also generates G we can
assume that m = 0. For each positive integer n, there exists an integer p
n
, and a
non-rational number
n
in (0, 1), such that n = p
n
+
n
. Consider the elements C
n
of G given by C
n
= A
pn
B
n
; these transformations are of the form C
n
(z) = z +
n
.
We claim that the numbers
n
are all distinct: if
n
=
m
we will have
n
= n p
n
and thus m = p
m
+ n p
n
, which would imply that Q. Since all the
n
are
distinct we can get a subsequence, say {
n
j
}
j
, converging to some point of [0, 1]. In
such case the transformations C
1
n
j+1
C
n
j
are all distinct and converge to the identity.
But then (C
1
n
j+1
C
n
j
)(z) z for all z C, a contradiction with the denition of
covering space (see also 3.3.11). Hence we have that B(z) = z +, with not real;
we can assume that Im() > 0 (take B
1
if necessary), and obtain that G is of the
form G

, as in example 1.3.6, so X is a torus.


We claim that these three cases, the complex plane, the punctured plane and
tori, are all the possibilities of Riemann surfaces covered by C. To prove the claim,
let X be a Riemann surface of the form C/G. All the transformations of G are of
the form T

: z z + . Let r = min{||; T

G}. Observe that we take r to be


a minimum, not the inmum: if T
r
were not in G we could construct a sequence of
distinct elements of G converging to the identity, using a trick similar to the above
one (we leave the proof to the reader). By a conjugation we can assume that r = 1.
Let be such that T

G and || is minimum among the transformations in G


which are not of the form T
n
(z) = z +n, for n integer. If the group G is cyclic then
X should be either the complex plane or the punctured plane. On the other hand,3.3. UNIFORMIZATION OF RIEMANN SURFACES AND KLEINIAN GROUPS 155
if G is not cyclic the argument above applies and we see that cannot be a real
number. Thus G contains a subgroup of the form G

. We claim that G = G

. If
not, let T

be an element in G but not in G

. Since {1, } are linearly independent


over R we can write = r + s, where r and s are real numbers, but not integers.
Let m
1
and m
2
be two integers such that |r m
1
| 1/2 and |s m
2
| 1/2; the
number

= m
1
m
2
satises
|

| <
1
2
+
1
2
|| ||,
where the rst inequality is strict since is not a real number. But this contradicts
the choice of . These computations complete the proof of the following theorem.
Theorem. If X is a Riemann surface whose universal covering space is C, then
X is (biholomorphic to) C, C

or a torus.
It follows from this theorem and 3.3.6 that most surfaces are covered by the
unit disc. In particular any compact surface of genus greater than 1 has D as its
universal covering. We will see some applications of this fact in the next section (for
example, the Riemann-Hurwitz theorem 3.4.20).
3.3.8. The next two results, which are easy consequences of Schwarz lemma,
characterise the automorphisms of the unit disc.
Lemma. If f : D D is a biholomorphism of the unit disc with f(0) = 0. Then
f is a rotation around the origin; that is, f(z) = z, for some complex number of
absolute value 1.
Proof. This result is part of Schwarz lemma (1.1.7).
Proposition. The automorphisms of the unit disc D are the Mobius transfor-
mations of the form T
w,
(z) =
zw
1 wz
, where w D and || = 1.
Proof. We rst need to show that these transformations are automorphisms of
D. Let e
i
be a point of D (the boundary of the unit dist); then we have
|T
w,
(e
i
) = ||

e
i
w
1 we
i

= |e
i
|

e
i
w
e
i
w

= 1,
since the denominator of the last fraction is the complex conjugate of its numerator.
This computation only shows that T
w,
(S
1
) S
1
. Since T
1
w,
(z) = T
w,

we have156 3. UNIFORMIZATION OF RIEMANN SURFACES


that T
w,
(S
1
) = S
1
, and therefore T(D) is equal to either D of

C\D. But since the
image of the origin is given by T
w,
(0) = w, which is a point in D, we have that
T
w
(D) = D.
Let f : D D be an arbitrary automorphism of the unit disc. Write w
0
= f(0).
Then T
w
0
,1
f xes the origin, so it must be a rotation by the previous lemma; that
is (T
w
0
,1
f)(z) = z (|| = 1). A simple computation shows that
f(z) = T
1
w
0
,1
( z) = T
w,
(z).
3.3.9. From the Riemann Mapping theorem (or the Uniformization theorem)
we have that the upper half plane H is biholomorphic to the unit disc D; the Mobius
transformation T(z) =
zi
z+i
: H D gives one such identication. To see this observe
that for x real we have that |T(x)| is the ratio of the distance from x to i to the
distance from x to i and therefore |T(x)| = 1. By topological arguments we get
that T(R {}) = S
1
and T(H) must be either the unit disc or its exterior. Since
T(i) = 0 we have that T(H) = D. The advantage of using the upper half plane
over the unit disc is that many computations are easier. For example, the next
proposition shows that the automorphisms of H are just Mobius transformations
with real coecients and positive determinant, certainly simpler expressions than
those of elements of Aut(D).
Proposition. The automorphisms of H are the Mobius transformations of
the form A(z) =
az+b
cz+d
, where a, b, c, d are real numbers satisfying ad bc > 0 (or
equivalently ad bc = 1). The group Aut(H) acts transitively on H; that is, for
any two points w
0
and w
1
of H, there exists an element T Aut(H), such that
T(w
0
) = w
1
.
Proof. If A : H H is an automorphism of the upper half plane, then TAT
1
is an automorphism of the unit disc, where T(z) =
zi
z+i
. By 3.3.8 we have that TAT
1
is a Mobius transformation and therefore A is also a Mobius transformation. This
shows that Aut(H) is a group of Mobius transformations.
Let G denote the group of Mobius transformations of the form given in the state-
ment of the proposition; we want to show that G is the full group of automorphisms3.3. UNIFORMIZATION OF RIEMANN SURFACES AND KLEINIAN GROUPS 157
of H. If A G we have that
A(z) =
az + b
cz + d
=
az + b
cz + d
a z + b
c z + d
=
ac|z|
2
+ adz + bc z + bd
|cz + d|
2
;
so
Im(A(z)) =
(ad bc) Im(z)
|cz + d|
2
.
This shows that G is a subgroup of Aut(H) (if z has positive imaginary part so does
A(z)).
Let w
0
= x
0
+iy
0
be a point of the upper half plane. The transformation M(z) =
zx
0
y
0
satises M(w
0
) = i (since w
0
H we have y
0
> 0). We can write M as
M(z) =
z

y
0

x
0

y
0

y
0
,
so M belongs to G (in the above expression we have taken the positive square root of
y
0
, which is possible since y
0
is a positive real number). Therefore G acts transitively
on H (map w
0
to i and then i to w
1
), and consequently Aut(H) too.
If B is an element of Aut(H) xing the point i the transformation R = TBT
1
is
an automorphism of D that xes the origin. Hence R(z) =
2
z, for a complex
number of absolute value 1. The matrices corresponding to R and T are
R =
_
_
0
0

_
_
and T =
1

2i
_
_
1 i
1 i
_
_
,
respectively. If we write = cos() + i sin() an easy calculation shows that B is
given by
B(z) =
cos() z + sin()
sin() z + cos()
,
which belongs to G.
Let now C denote any automorphism of H; we have C(w
0
) = i for some point
w
0
in the upper half plane. Let M be as above. Then MC
1
xes the point i, so it
follows from the above computation that MC
1
G. Since M G we have that
Aut(H) = G.
In a way similar to the identication of Aut(

C) with PSL(2, C) we can give an iso-


morphism between Aut(H) and PSL(2, R), where this last group consists of equiv-
alence classes of matrices with real coecients and determinant 1. In this case158 3. UNIFORMIZATION OF RIEMANN SURFACES
we do not have that PSL(2, R) and PGL(2, R) are isomorphic, since a matrix in
GL(2, R) with negative determinant cannot be equivalent to a matrix with positive
determinant (negative numbers do not have square roots in R).
3.3.10. What elements of Aut(H) have xed points in H? First of all, if A
PSL(2, R) xes the point z
0


C, then A must also xed its conjugate z
0
, since the
coecients of A are real (we understand = ). If A is parabolic its xed point
must be in

R = R {}. If A is elliptic, with ad bc = 1, the solutions of the
equation A(z) = z are given by
z =
a d
_
(d a)
2
+ 4bc
2c
=
a d

d
2
+ a
2
2ad + 4bc
2c
=
=
a d
_
(a + d)
2
4
2c
.
Since c = 0 and 0 (a +d) < 4 the transformation A must have a xed point in H.
If A is loxodromic it must be hyperbolic and it is easy to see that its xed points
are both in

R. It follows from this computations that if X is a Riemann surface of
the form X = H/G then G does not have elliptic elements.
3.3.11. We next dene a general class of groups of Mobius transformations
and reformulate the Uniformization theorem.
Definition. A group of Mobius transformation G is said to act properly
discontinuously at a point z

C if there exists an open neighbourhood U of z,
such that the subgroup of G given by
{g G; g(U) U = }
is nite. We denote by (G) the (open) set of points of the Riemann sphere where
G acts properly discontinuously; this set is called the region of discontinuity of
G. The group G is called Kleinian if (G) = .
We can now rewrite the Uniformization theorem in terms of Kleinian groups.
Theorem (Uniformization theorem). Any (connected) Riemann surface X is
biholomorphic to a quotient of the form

X/G, where

X is the Riemann sphere, the
complex plane or the unit disc, and G is a Kleinian group satisfying

X (G).
Remark. Observe that in this above result we do not claim that

X is equal3.3. UNIFORMIZATION OF RIEMANN SURFACES AND KLEINIAN GROUPS 159
to the region of discontinuity (G) of G. There are cases when these two sets
are dierent. For example, if X is the punctured unit disc D

, then

X = H and
G = {z z +n; n Z}. But the region of discontinuity of G is the whole complex
plane and C/G is the punctured plane C

(the upper half plane covers the punctured


unit, the lower half plane the exterior of the unit disc and R covers S
1
).
3.3.12. Proposition. Kleinian groups are discrete.
Proof. By discrete we mean that G does not have accumulation points in
PSL(2, C), with the topology described in 3.3.4. Assume that G is not discrete;
then there exists a sequence of distinct elements of G, say {A
n
}
n
, such that A
n

A, where A is a Mobius transformation, not necessarily in G. The sequence of
Mobius transformations {B
n
= A
1
n+1
A
n
}
n
has innitely many distinct elements and
converges to the identity. But then B
n
(z) z, for all z in

C, and therefore G
cannot be Kleinian.
Let Gbe a Kleinian group, Aan elliptic transformation of G. By conjugating with an
element of PSL(2, C) if necessary we can assume that A is of the form A(z) = e
i
z.
It is easy to see that A has nite order if and only if is a rational number. Assume
now that / Q, and dene a mapping j :< A > S
1
, where < A >= {A
n
; n Z}
is the subgroup of G generated by A, by the expression j(A
n
) = e
i n
. Since A does
not have nite order we get that the image of j is an innite set of S
1
and therefore
it has an accumulation point, say e
i
0
. Let {n
j
}
j
be a sequence of integers such that
e
i n
j

e
i
0
. Then the transformations A
n
j
converge to z e
i
0
z, so G cannot be
Kleinian. We have proved the following result.
Proposition. If G is a Kleinian group and A is an elliptic element of G then
A has nite order.
3.3.13. Another interesting property of Kleinian groups is given in the follow-
ing proposition.
Proposition. A Kleinian group is either nite or innite countable.
Proof. Let z
0
be a point in the region of discontinuity of G and H the stabiliser
of z
0
in G; that is,
H = {g G; g(z
0
) = z
0
}.160 3. UNIFORMIZATION OF RIEMANN SURFACES
Since G acts properly discontinuously at z
0
we have that H is nite. Let G(z
0
)
denote the orbit of z
0
under the given group, G(z
0
) = {g(z
0
); g G}. We have that
G(z
0
) is a discrete set of

C and it must then countable (exercise 83). On the other
hand, it is easy to see that there is a bijection between G/H and G(z
0
), given by
[g] g(z
0
). It follows that G is either nite or innite countable.
3.3.14. Our next application of the Uniformization theorem is to determine
all surfaces with abelian fundamental group.
Lemma. Let A and B be two Mobius transformations, neither of them equal to
the identity. Assume that AB = BA. Then one and only one of the following cases
is satised:
1. if A is parabolic then B is also parabolic and they have the same xed points;
2. if A is not parabolic then B is not parabolic and either they have the same
xed points, or both transformations have order 2, and each of them interchanges
the xed points of the other.
Proof. First of all, the results of the lemma are invariant under conjugation,
so we can choose the xed points of the transformations in a way that computations
are easy. Observe that if A xes a set W pointwise (that is, A(w) = w for all w in
W), then B xes W as a set, B(W) = W, although B does not need to x each
point of W. Clearly this statement holds if we interchange A and B.
Assume rst that A is parabolic, say A(z) = z + 1 (remember that we are free
to conjugate A and B for our computations). Then B() = B(A()) = A(B())
so B() must be a xed point of A, which implies that B() = , and hence B
is of the form B(z) = z + . If B xes a point z
0
in C, from the above remark we
see that A must x the point z
0
, which cannot happen by hypothesis. Hence = 1
and B is a parabolic transformation with xed point .
Assume now that neither of the transformation is parabolic (by the above com-
putation, if one transformation is parabolic so is the other). Let A be of the form
A(z) = z. From the above computations we have that B({0, }) = {0, }, so
there are two possible cases:
1. B() = and B(0) = 0. Then A and B have the same xed points.
2. B() = 0 and B(0) = . In this case B(z) = /z. By a conjugation3.3. UNIFORMIZATION OF RIEMANN SURFACES AND KLEINIAN GROUPS 161
that does not change A we can assume that = 1. Then B xes 1. Since
A({1, 1}) = {1, 1} we see that A(z) = z (the possibility of A being the identity
does not occur by hypothesis), so A has order 2 and this completes the proof.
3.3.15. Theorem. Suppose X is a Riemann surface with abelian fundamental
group. The one (and only one) of the following cases occurs:
1. X is simply connected and X is

C, C or D;
2.
1
(X, x
0
)

= Z and X is C

, D

or A
r
= {z C; r < |z| < q}, for some real
number r (0, 1);
3.
1
(X, x
0
)

= Z Z and X is a torus.
Remark. In the above theorem all statements have to be understood up to
biholomorphisms.
Proof. The rst case is the Uniformization theorem for simply connected sur-
faces; the surfaces with universal covering space the complex plane have been studied
in 3.3.7. Thus we have only to study surfaces with abelian fundamental group and
the upper half plane as the universal covering space. Moreover, we can assume that
the fundamental group is not trivial.
If
1
(X, x
0
) is cyclic, then X = H/ <A>, where A is an element of PSL(2, R).
Since non-trivial deck transformations do not have xed points A will be either
parabolic or hyperbolic. In the rst case we can assume that A(z) = z + 1, after
a conjugation and taking inverses if necessary. Then X = D

and the covering


mapping is z exp(2iz). In the case of A hyperbolic we have A(z) = z, for some
number > 1; we get that X is an annulus, with covering mapping
z exp
_
2i
log z
log
_
,
where log is the principal branch of the logarithm, log(re
i
) = log r +i. The radius
of the annulus is given by r = exp(
2
2
log
) (0, 1).
To complete the proof of the theorem we need to show that there are no Riemann
surfaces with abelian fundamental group of rank greater than 1 and universal cover-
ing space the upper half plane. If one of the generators of G, say A, is parabolic, we
can assume that A(z) = z +1. By lemma 3.3.14 all elements of G are also parabolic,
and because they are automorphisms of H, they must be of the form z z + t,162 3. UNIFORMIZATION OF RIEMANN SURFACES
for t real. But when we studied surfaces covered by C we saw that in that case G
would not be discrete. If A is hyperbolic, we have that all elements of G are also
hyperbolic; a similar proof shows that this case cannot occur.
3.4. Hyperbolic Geometry, Fuchsian Groups and
Hurwitzs Theorem
In this section we will study some properties of groups of automorphisms of
the upper half plane. We show that there exists a natural metric on H, called
the hyperbolic metric, for which the elements of Aut(H) are isometries. It follows
from this that we can put a metric on compact surfaces (of genus greater than 1).
A somehow surprising result is that the area of a surface does not depend of the
Riemann surface structure. We will also prove that the group of automorphisms of
compact surfaces (covered by H) is nite.
3.4.1. If : [a, b] R
2
is a piecewise smooth curve (1.4.18) its length in
Euclidean geometry is given by the integral
||||
E
=
_
b
z
|

(t)| dt.
(Since is piecewise smooth the integral is nite.) This statement is usually for-
mulated by saying that the innitesimal length element (the length of the tangent
vector

) is |dz|. The distance d


E
between two points p
0
= (x
0
, y
0
) and p
1
= (x
1
, y
1
)
is the length of the segment joining them, that is:
d(p
0
, p
1
) = ||s||,
where
s(t) = (tx
0
+ (1 t)x
1
, ty
0
+ (1 t)y
1
), for 0 t 1.
One can check that the length of this curve is minimum among the lengths of the
curves joining p
1
and p
2
; that is,
d
E
(p
0
, p
1
) = ||s|| = inf{||||; (a) = p
1
, (b) = p
2
}.3.4. HYPERBOLIC GEOMETRY, FUCHSIAN GROUPS AND HURWITZS THEOREM 163
The segments are called geodesics. It can be easily veried that the distance between
any two points in a geodesic is given by the length of the piece of the geodesic joining
them.
3.4.2. By the above argument we see that to dene a metric on the upper half
plane it suces to give its innitesimal length element. We set this to be equal to
ds =
|dz|
Im(z)
. As explained in the case of the Euclidean metric, this simply means
that the length of a (piecewise smooth) curve : [a, b] H is given by the following
expression:
|||| =
_
b
a
|

(t)|
Im((t))
dt.
This integral is nite because the curve is assumed to be piecewise smooth. Similarly
one denes the distance d between two points z
0
and z
1
of H by
d(z
0
, z
1
) = inf{||||; (a) = z
0
, (b) = z
1
}.
We will use d for this new distance and d
E
for the (standard) Euclidean distance. ds
and d are called the hyperbolic metric and hyperbolic distance, respectively.
We need to prove that d is indeed a distance, but before that we will study some
properties of the metric ds and its relation with Mobius transformations.
3.4.3. Before proceeding further we recall some results from Complex Analysis.
Definition. Let z
j
, j = 1, . . . , 4 be four distinct points in

C; the cross ratio
of these points is dened by
(z
1
, z
2
; z
3
, z
4
) =
z
4
z
2
z
4
z
1
z
3
z
1
z
3
z
2
,
where we delete the corresponding terms (or we take limits) if one of the points is
the point .
Observe that in the case of one of the four points being there will be two terms
in the above expression with in them, one in the numerator and the other in the
denominator, so after removing those terms we are left with a well dened fraction.
Some authors change the order of the factors in the denition of cross ratio. However,
for the applications all denitions are equivalent. It can be easily proved that of the
possible 24 denitions of cross ratio (there are 24 permutations of four letters) there
are only 6 dierent values.164 3. UNIFORMIZATION OF RIEMANN SURFACES
If z
1
= , z
2
= 0 and z
3
= 1 then (z
1
, z
2
; z
3
, z
4
) = z
4
. More generally, an easy
computation show that S(z) = (z
1
, z
2
; z
3
, z) is the value (at z) of the unique Mobius
transformation S that takes z
1
, z
2
and z
3
to , 0 and 1, respectively. This remark
will be useful in the proof of the following result.
3.4.4. Lemma. Mobius transformations preserve cross ratios. More precisely,
if T is a Mobius transformation, and z
j
, j = 1, . . . , 4 four distinct points in the
Riemann sphere, then (T(z
1
), T(z
2
); T(z
3
), T(z
4
)) = (z
1
, z
2
; z
3
, z
4
).
Proof. The proof can be done with an easy (but long) direct calculation;
however, with the last remark in the above subsection we can get a short and elegant
proof as follows. Let S be the Mobius transformation that takes z
1
, z
2
and z
3
to ,
0 and 1, respectively (this S is given by S(z
4
) = (z
1
, z
2
; z
3
, z
4
)). Then ST
1
takes
T(z
1
), T(z
2
) and T(z
3
) to , 0 and 1 respectively and therefore we have
(T(z
1
), T(z
2
); T(z
3
), T(z
4
)) = ST
1
(T(z
4
)) = S(z
4
) = (z
1
, z
2
; z
3
, z
4
).
3.4.5. Lemma. For distinct points in the Riemann sphere lie on a line or
circle if and only if their cross ratio is real.
Proof. We will show rst that the image of

R = R {} under a Mobius
transformation is a line or a circle. If S(z) = (z
1
, z
2
; z
3
, z) is given by S(z) =
az+b
cz+d
,
then S(z) is real if and only if S(z) = S(z); that is,
az + b
cz + d
=
a z +

b
c z +

d
.
From this expression we obtain
(6) (a c ac)|z|
2
+ (a

bc)z + (b c ad) z + (b

bd)z = 0.
If a c a c = 0 then we must have a

bc = 0. Otherwise we get the following pair


of equations
a c = ac, a

d =

bc.
If a = 0 we have

d =

bc
a
, and thus adbc = a
b c
a
= a
bc
a
bc = 0, which is not possible.
On the other hand, if a = 0 we get

bc = 0, which again gives us ad bc = 0. So3.4. HYPERBOLIC GEOMETRY, FUCHSIAN GROUPS AND HURWITZS THEOREM 165
we see that a

bc = 0. Write a

bc = u + iv, z = x + i and b

d = r + is; then
equation (6) becomes
vx + uy + s = 0.
Since u and v cannot be simultaneously equal to 0 we get that this is the equation
of a line.
If a c ac = 0 equation (6) is equivalent to

z +
ad b c
ac a c

ad bc
ac a c

,
which the the equation of a circle.
To complete the proof of the lemma we argue as follows. If (z
1
, z
2
; z
3
, z
4
) is a real
number then z
j
lies in S
1
(

R), where S is the Mobius transformation that denes


the cross ratio (i.e. S(z) = (z
1
, z
2
; z
3
, z)). By the rst part of the proof we have that
S
1
(

R) is either a line or a circle.


Suppose now that (z
1
, z
2
; z
3
, z
4
) lie in a line or circle, say C. If we consider the
transformation S once more we have that S
1
(0), S
1
(1) and S
1
() are in C, so
S
1
(

R) = C, and therefore S(C) =



R. Thus S(z
4
) = (z
1
, z
2
; z
3
, z
4
) is a real number
(it cannot be by the denition of cross ratio).
Corollary. If C is the family of lines and circles in

C and A is a Mobius
transformation, then A(C) = C.
3.4.6. The Mobius transformation T(z) =
zi
z+i
: H D can be used to dene
a hyperbolic metric on the unit disc such that T becomes an isometry. This means
that if the metric on D is given by (z) |dz|, where is a positive function, then we
must have
(T(z)) |T

(z)| =
1
Im(z)
,
for z in H. If that is the case a simple use of the change of variables theorem for
integrals shows that |||| = ||T()|| for a piecewise smooth curve on H. It is easy to
see that is given by the expression
(z) =
2
1 |z|
2
.166 3. UNIFORMIZATION OF RIEMANN SURFACES
This allows us to switch between the upper half plane and the unit disc when we
prove results regarding the hyperbolic metric.
3.4.7. Proposition. Aut(H) acts by isometries with respect to the hyperbolic
metric: for any piecewise smooth curve : [a, b] H and any Mobius transforma-
tion A Aut(H), one has |||| = ||A()||.
Proof. Let A be given by A(z) =
az+b
cz+d
, where the coecients are real and
satisfy ad bc = 1. We have A

(z) =
1
(cz+d)
2
. In 3.3.8 we computed that
Im(A(z)) =
Im(z)
|cz + d|
2
.
Using these expressions we get
|A

(z)|
Im(A(z))
=
1
Im(z)
,
so
||A()|| =
_
b
a
|A

(t)| |

(t)|
Im(A((t))
dt =
_
b
a
|

(t)|
Im((t))
dt = ||||.
3.4.8. Theorem. d is a distance in H. The topology induced by it is the
standard Euclidean topology.
Proof. It is easy to see that d is symmetric, non-negative and satises the
triangle inequality. We will show that d(w
0
, w
1
) is strictly positive for w
0
= w
1
. We
will work in the unit disc, since the computation is easier in this case, and in the
process we will obtain a formula for the distance of a point in D to the origin that
will be useful later.
Let w
0
and w
1
be two distinct points in D. Using the Mobius transformation
M(z) =
z w
0
1 w
0
z
we can assume that w
0
= 0. By a rotation we can further assume that w
1
= t, where
t is a point in the open interval (0, 1). Consider a path : [0, 1] D joining 0 and
t. If we write (t) = x(t) + iy(t), we have that
|||| =
_
1
0
2|

(t)|
1 |(t)|
2
dt
_
1
0
2|x

(t)|
1 x(t)
2
dt
_ 2x(t)
1x(t)
2
0
dt = log
_
1 +t
1 t
_
.3.4. HYPERBOLIC GEOMETRY, FUCHSIAN GROUPS AND HURWITZS THEOREM 167
In the second inequality we have used that the function f(s) =
1
1s
2
is negative for
s < 0 and increasing for s 0. Observe that the hyperbolic length of the path
(s) = st, s [0, 1], is precisely the above expression, log
_
1+t
1t
_
; thus d(0, t) =
d(w
0
, w
1
) > 0. This completes the proof of the fact that d is a distance.
To prove that the topology induced by d is the standard topology of D we use
the above computations. We have that the hyperbolic disc D
h
(0, r), of centre 0 and
radius r > 0, is given by
D
h
(0, r) = {z D; |z| <
e
r
1
e
r
+ 1
};
that is, an Euclidean disc of centre 0 and dierent radius. This shows that the
neighbourhoods of 0 in the hyperbolic and Euclidean topologies are the same. Since
the group of Mobius transformations acts transitively by homeomorphisms in D
(3.3.8) we have that both topologies are the same.
As a corollary of the above computations we get that the distance from 0 to any
point w D is given by
d(0, w) = log
_
1 +|w|
1 |w|
_
.
In particular, as w approaches S
1
(in the Euclidean distance) we have that d(0, w)
goes to innity. This shows that the unit circle is at innity distance of any point
in the unit disc (apply the triangle inequality). Similarly the real axis is at innite
(hyperbolic) distance from points in the upper half plane. Thus the hyperbolic
metric is the natural one if we want to study properties of H (or D) on its own,
rather than considering it as a subset of the Riemann sphere.
3.4.9. A geodesic is a (smooth) curve that minimises the distance locally
between points in it (its image). More precisely, if is a geodesic dened on the
interval (a, b) and t
0
(a, b), then there is a neighbourhood U of (t
0
), such that the
distance between any two points in ((a, b)) U is given by the length of between
those two points. For example, if is a geodesic in the upper half plane, with the
same notation we have that if (t
j
) U, for j = 1, 2,
d((t
1
), (t
1
)) =
_
t
2
t
1
|

(t)|
Im((t))
dt.168 3. UNIFORMIZATION OF RIEMANN SURFACES
However a geodesic does not need to minimise distances globally. Consider the case
of the sphere S
2
where geodesics are given by great circles, that is, the intersection
of planes through the origin with S
2
. If p
1
and p
2
are two points in the sphere, not
diametrically opposed, then there are two geodesics joining them, one of which will
realize the distance between p
1
and p
2
while the other hand will have longer length.
We also have that there could be more than one geodesic between two points. In
the same example of the sphere, any two points diametrically opposed are joined
by innitely many dierent geodesics. And there are spaces where some points can-
not be joined by geodesics. The space R
2
\{(0, 0)} with the Euclidean metric is an
example; the points (1, 0) and (1, 0) are at distance 2, but there is no geodesic
between them realizing that distance. In the case of the hyperbolic metric we are in
the best possible situation: any two points can be joined by a unique geodesic that
realizes the distance between them.
3.4.10. Proposition. The hyperbolic geodesics of D are the circles and lines
perpendicular to S
1
. In the case of the upper half plane, the geodesics are the circles
and lines perpendicular to the real axis. Given any two points in D (or H) there
exists a unique geodesic between them; moreover, such geodesic realizes the distance
between any two points in its image.
Proof. In the proof of 3.4.8 we have obtained that the segment (0, 1) is a
geodesic in D. It is not dicult to see that the full diameter (1, 1) is a geodesic.
If C is a circle (or line) perpendicular to S
1
it is possible to nd an automorphism
of the unit disc, say A, such that A(C D) = (1, 1) (exercise 77). It follows from
this that any circle or line orthogonal to S
1
is a geodesic.
Conversely, if is a geodesic in the unit disc, consider two points w
j
= (t
j
), j = 0, 1,
close enough so that realizes the distance between them. By an automorphism of
D, say A, we can map w
0
to 0 and w
1
to a point in (0, 1). Since the automorphisms
of D are hyperbolic isometries we have that should be contained in the image of
(1, 1) under A
1
and thus it is a circle or line orthogonal to S
1
(to be more precise,
the image of (a, b) under is contained on a circle or line orthogonal to S
1
).3.4. HYPERBOLIC GEOMETRY, FUCHSIAN GROUPS AND HURWITZS THEOREM 169
To show that any two points of D lie in a unique geodesic it suces to consider
that case of one point being the origin and the other a point t in the interval (0, 1).
But again this is part of the proof of 3.4.8.
3.4.11. Theorem. The hyperbolic metric on D (H) is complete.
Proof. Let {z
n
}
n
be a Cauchy sequence in the unit disc with respect to the
hyperbolic metric. Given > 0 there exists an n
0
such that d(z
n
, z
m
) < , for
n, m n
0
. Therefore d(0, z
n
) d(0, z
n
0
) + d(z
n
0
, z
n
), and thus d(0, z
n
) is bounded.
The distance d(0, z
n
) is given by log(
1+|zn|
1|zn|
), so there exists a number 0 < R < 1,
such that |z
n
| R, for all n. Since the set {z; |z| R} is compact in the Euclidean
topology there exists a convergent subsequence. But the topologies induced by the
hyperbolic and Euclidean metrics are equivalent, so that subsequence converges (to
the same limit point) with respect to the hyperbolic metric. The triangle inequality
shows that the full sequence {z
n
}
n
converges in the hyperbolic metric.
3.4.12. The angle between two lines or circles in H meeting at a point z
0
is
dened as the angle formed by the tangent lines to the curves at z
0
. For simplicity
we say that two lines or circles meeting at a point of

R do it with angle equal to 0. A
triangle is the portion of H enclosed by three distinct geodesic that meet pairwise.
A triangle is called ideal if the geodesics meet in a point in the (extended) real axis.
The hyperbolic area of a region D of H is given by the integral
Area(D) =
_
D
1
y
2
dxdy.
Theorem (Gauss-Bonet for triangles). The hyperbolic area of a triangle with
angles , and is equal to ( + + ).
Proof. Consider rst the case of a triangle T with two angles equal to 0.
By using a Mobius transformation and the reection r(z) = z (which is also a
hyperbolic isometry and preserve angles, see exercise 78) we can assume that T is
as in gure 13. In this case we can compute the area directly as follows:
Area(T ) =
__
T
1
y
2
dxdy =
_
d
0
_

c
2
(xc)
2
1
y
2
dxdy =
=
_
d
0
1
_
c
2
(x c)
2
dx =
_

0
d = .170 3. UNIFORMIZATION OF RIEMANN SURFACES
If T has only one angle equal to 0 we can compute its area as the dierence of the
area of two triangle, each of them with two zero angles, as in the gure 14. The
general case follows in a similar way.
0 c d

T
Figure 13. Triangle with two zero angles.

T S
Figure 14. Gauss-Bonet.
3.4.13. We want now to apply some of these results to compact surfaces cov-
ered by the unit disc. We start with a denition.
Definition. A Kleinian group G is called Fuchsian if there exists a disc or
half plane H which is invariant under the elements of G, that is, g(H) = H for all
g G.
When talking about Fuchsian groups we will use the word disc to mean a disc or a
half plane.
A striking fact of Fuchsian groups is that discreteness and properly discontinuous
action are almost equivalent.
Proposition. Let G be a Fuchsian group with invariant disc H. The following
are equivalent:3.4. HYPERBOLIC GEOMETRY, FUCHSIAN GROUPS AND HURWITZS THEOREM 171
1. G acts discontinuously on H (i.e. H (G));
2. G is discrete.
Proof. We have seen that Kleinian groups are discrete, so we only need to
prove 2 1; we will show this by contradiction. By a conjugation we can assume
that D is the unit disc and G does not act properly discontinuously at the origin.
This means that for any neighbourhood U of 0, the set {g G; g(U) U = } is
innite (in particular Gis not nite, which is obvious since nite groups acts properly
discontinuously on the whole Riemann sphere). Let r
1
be a positive number and
dene U = D(0, r
1
), in the hyperbolic metric. Let z
1
and w
1
be points in U\{0},
such that z
1
= g
1
(w
1
), for some transformation g
1
G. Choose a positive number
r
2
satisfying r
2
< min{d(z
1
, 0), d(w
1
, 0)}. We can nd z
2
and w
2
in U\{0}, such
that z
2
= g
2
(w
2
), for g
2
G, and g
2
= g
1
. Continuing this process we nd sequences
of positive numbers {r
n
}
n
, points in the unit disc {z
n
} and distinct elements of G
{g
n
}
n
, such that:
(1). r
n
is a decreasing sequence converging to 0;
(2). d(0, z
n
) < r
n
;
(3). d(g
1
n
(z
n
), 0) < r
n
.
Since G acts by isometries on the hyperbolic distance we have
d(0, g
n
(0)) d(0, z
n
) + d(z
n
, g
n
(0)) = d(0, z
n
) + d(g
1
n
(z
n
), 0) < 2r
n
.
Let w
n
= g
n
(0). By proposition 3.3.8 we have that g
n
is of the form
g
n
(z) =
n
_
z + w
n
1 +w
n
z
_
,
where || = 1. Choose a subsequence {
n
j
}
j
with
n
j

j

0
. Since w
n
0 we have
that the transformations g
n
j
converge to the rotation R(z) =
0
z, and thus G is not
discrete.
3.4.14. The action of a Kleinian group on its region of discontinuity (or a part
of it) is better understood by taking a set that contains one element of each orbit.
For example, if G is the group of translations G = {T
n
(z) = z + n; n Z}, acting
on H, we have that every point of H can be mapped by an element of G to a point z,
with 0 Re(z) < 1. If we consider the vertical strip S = {z H; 0 Re(z) 1},172 3. UNIFORMIZATION OF RIEMANN SURFACES
we have that G identies the two vertical lines in the boundary of this strip. The
quotient space H/G is equivalent to S/G; it is easy to see (geometrically) that S/G
is a cylinder, which is clearly homeomorphic to the punctured disc. We have indeed
proved that H/G is the punctured disc; this discussion might help us to understand
why. The next denition generalises this situation to Fuchsian (or Kleinian) groups.
Definition. Let G be a Fuchsian group acting on H. A fundamental domain
of G for its action on H is a connected open set D satisfying the following conditions:
FD1: for every element g of G, not equal to the identity, g(D) D = ;
FD2: for every z in H there exists a transformation g of G, such that g(z) belongs
to D, the closure of D in H;
FD3: the boundary of D in H consists of a countable number of smooth curves,
called the sides of D. For every side s there exists another side, say s

, not necessarily
distinct from s, and an element g of G, such that g(s) = s

and (s

= s;
FD4: (local niteness) for every compact set K of H, the group
{g G; g(K) K = },
is nite.
3.4.15. The following result is needed to prove the existence of fundamental
domains.
Lemma. Let G be a non-cycle Fuchsian group with invariant disc . Then there
exists a point z
0
that is not xed by any non-trivial element of G.
Proof. Assume = D and that 0 is xed by some non-trivial element of G;
let H be the subgroup of G consisting of the elements that x 0. By lemma 3.3.8
all elements of H are rotations around the origin. If G = H by discreteness we
have that G must be cyclic. On the other hand, if H is a proper subgroup, since
G acts properly discontinuously at 0, we can nd a positive number r, such that
g(U) U = , for all g / H, where U is the disc of centre 0 and radius r. Any point
in U\{0} will satisfy the conditions of the lemma.
3.4.16. Let G be a Fuchsian group that leaves the upper half plane invariant
and choose p H satisfying the conditions of the above lemma. For g G, not3.4. HYPERBOLIC GEOMETRY, FUCHSIAN GROUPS AND HURWITZS THEOREM 173
equal to the identity, dene
H
g
= {z H; d(p, z) < d(g(p), z)},
where we use the hyperbolic metric to measure distances. Thus H
g
consists of the
points in H that are closer to p than to g(p). Geometrically one can obtain H
g
by
considering the hyperbolic geodesic segment that joins p and g(p), say L, and then
taking the geodesic L

orthogonal to L on its midpoint; H


g
will be the half plane
determined by L

containing p. The Dirichlet region D


p
(G) of G (relative to p) is
dened as the intersection of all such hyperplanes:
D
p
(G) =

g=Id
H
g
.
For example, if G is the group of translations G = {T
n
(z) = z + n; n Z}, and
p = (1/2) + i, the Dirichlet region D
p
(G) is precisely the (open) vertical strip we
considered in 3.4.14: D
p
(G) = {z H; 0 < Re(z) < 1}.
Proposition. The Dirichlet region is a fundamental domain for the action of
G on H.
Proof. To simplify notation we will write D for the Dirichlet region D
p
(G)
(relative to some point p).
Condition FD1. Let g = Id be an element of G and z a point of D. Since
z H
g
1 we have
d(g(p), g(z)) = d(p, z) < d(g
1
(p), z) = d(p, g(z)),
which implies that g(z) does not belong to D.
Condition FD2. Let z be a point in the upper half plane. By the discontinuous
action of G we have that there exists an element g G (not necessarily unique),
such that d(g(p), z) d(h(p), z) for all h G. If we write the elements of G as g h
we have
d(g
1
(z), p) = d(z, g(p)) d(z, (g h)(p)) = d(g
1
(z), h(p)).
This means that g
1
(z) belongs to the closure of H
h
for all h G and therefore to
the closure of D.
Condition FD3. Since G is countable it is clear that the boundary of D has at174 3. UNIFORMIZATION OF RIEMANN SURFACES
most countably many sides. Let z be a point in the relative interior of a side s. This
is equivalent to say that there exists a unique element g G, such that
d(z, p) < d(z, h(p)), h = g, Id and d(z, p) = d(z, g(p)).
Hence d(g
1
(z), p) = d(g
1
(z), g
1
(p)) = d(z, p), and for h = g
1
, Id we have
d(g
1
(z), h(p)) = d(z, (g h)(p)) > d(z, p) = d(g
1
(z), p).
Thus g
1
(z) belongs to the side s

with g
1
(s

) = s.
Condition FD4. Let K H be compact. Without loss of generality we can
assume that K is the closed disc centred at p and of hyperbolic radius r (any compact
subset of the upper half plane is contained in one such disc). We have that there are
only nitely many images of p (under transformations of G) in the disc of radius 2r
centred at p. From this it follows that if d(g
1
(p), p) > 2r then g(D) K = .
3.4.17. The following lemma is easy to prove; it shows that the Dirichlet region
is a good choice of fundamental domain.
Lemma. Let G be a Fuchsian group acting on H and D a Dirichlet region of G.
1. The quotient surface H/G is compact if and only if D is compact in H.
2. If D is compact then D has only nitely many sides.
3.4.18. Assume that X is a Riemann surface given by H/G; that is, X is
biholomorphic to the quotient surface H/G. We can put a metric on X by using
the natural projection : H X similar to the way we calculated the metric on D
from the mapping T : H D. Although does not need to be globally one-to-one,
it is so locally, and that is all we need. More precisely, for a point p
0
in X, let
U be an evenly covered neighbourhood of p
0
, and write
1
(U) =

j
V
j
for the
decomposition of the preimage of U in disjoint open sets of H, each homeomorphic
via to U. The functions z
j
((p)) = p, for p V
j
, serve as local coordinates on X.
Thus to dene a metric on X all we need to do is to nd expressions of the form

j
((p)) |dz
j
|, such that
j
=
k

dz
k
dz
j

. We then set
j
((p)) =
1
Im(p)
, for p V
j
. For
any other set V
k
as above, there exists an element g G, such that g(V
j
) = V
k
. The
value of
k
is given by
k
((q)) =
1
Im(q)
, and z
k
((q)) = q = g(p) = g(z
j
(p)), for3.4. HYPERBOLIC GEOMETRY, FUCHSIAN GROUPS AND HURWITZS THEOREM 175
q V
k
, so p = g
1
(q) belongs to V
j
. Under these circumstances we have that
1
Im(q)
=
1
Im(g(p))
=
|g

(p)|
Im(p)
=
1
Im(p)

dz
k
dz
j

We call this metric the hyperbolic metric of the surface. Observe that the metric
depends on the complex structure of X; however, the area of X does not, as the
following result shows.
Theorem (Gauss-Bonet for compact surfaces). If X is a compact surface of
genus g 2 then the hyperbolic area of X is equal to 2 (2g 2).
Proof. Let D be a Dirichlet region for G. By 3.4.17 we know that D has only
nitely many sides, so its boundary has zero area. On the other hand, by property
FD1 we have that : D X is one-to-one, so Area(D) = Area(X). Choose a point
p
0
in the interior of D and join it to the (nitely many) vertices of the boundary
of D by geodesics. This is possible, since D is a nite intersection of convex sets,
and thus it is convex. In this way we obtain a triangulation of D that projects to a
triangulation of X. Assume that the sides and vertices of D project to E sides and
V vertices on X. It is not dicult to see, using the Euler-Poincare formula, that
V E + 1 = 2 2g. On the other hand, by the Gauss-Bonet theorem for triangles
we have
Area(D) = 2E 2

(interior angles at the vertices) .


The term 2 comes from the sums of the angles at p
0
. Since the vertices of D
project to V points in X, we have that the sum of the interior angles is 2V , and
therefore
Area(X) = Area(D) = 2(2g 1 +V ) 2 = 2 (2g 2).
3.4.19. Our next goal is to study automorphisms of compact surfaces. Let
: H X be the universal covering of a compact surface of genus g 2, with
covering group G. If f : X X is an automorphism we can lift it to a biholo-
morphic mapping A : H H; in particular A is a Mobius transformation. The176 3. UNIFORMIZATION OF RIEMANN SURFACES
transformation A satises A = f; for any element g G we have
Ag A
1
= f g A
1
= f A
1
= f f
1
= ,
so there exists an element h G, such that Ag A
1
= h. In other words, A belongs
to N(G), the normaliser of G in Aut(H). (The group N(G) is the biggest subgroup
of Aut(H) on which G is normal; it consists of the element B Aut(H), such that
BGB
1
= G). The converse statement is also true; namely, if B N(G), then
the expression h((z)) = (B(z)) denes an automorphism, h, of X. Since the
elements of G will project to the identity mapping of X, we can identied Aut(X)
with N(G)/G. The following result guarantees that (under mild conditions on G)
the group N(G) is Fuchsian.
Proposition. If G is a torsion-free non-cyclic Fuchsian subgroup of PSL(2, R),
then N(G) is also Fuchsian.
Recall that a group is said to be torsion-free if there are no non-trivial elements of
nite order.
Proof. All we need to show is that N(G) is discrete (3.4.13). Assume that
there exists a sequence {h
n
} of distinct elements of N(G) converging to the identity.
For all g G we have that {h
n
gh
1
n
}
n
is a sequence of elements of G converging to
g. Since G is Fuchsian we get that h
n
gh
1
n
= g, for n > n
0
, for some positive integer
n
0
. By lemma 3.3.14 we get that h
n
and g must have the same xed points (since
we are assuming that G is torsion-free, the situation where h
n
and g have order 2
does not occur).
If all elements of G have the same xed points we would have that G consists of only
parabolic or hyperbolic transformations. In either case G would be cyclic, against
the hypothesis. Let us choose g
1
and g
2
in G with at least three distinct xed points,
say z
1
, z
2
and z
3
. Then every element h of N(G) will x z
j
, for j = 1, 2, 3. By 3.3.5
we get h = Id.
3.4.20. The automorphisms group of the Riemann sphere is the group of
Mobius transformations. For the case of a torus T

, any translation of the form


T(z) = z +c, with c C, induces an automorphism on T

. Thus in these two cases3.4. HYPERBOLIC GEOMETRY, FUCHSIAN GROUPS AND HURWITZS THEOREM 177
we have that the automorphisms group is not only innite, but it is not discrete
either. In the case of compact surfaces this cannot happen.
Theorem (Hurwitz). Let X be a compact Riemann surface of genus g 2.
Then Aut(X) has at most 84(g 1) elements.
Proof. If X = H/G, since G is torsion-free and non-cyclic we know that N(G)
is Fuchsian (3.4.19), and thus Y = H/N(G) is a Riemann surface. The covering
H Y clearly factors through X, so Y must be compact. The mapping q : X Y
has degree equal to the order of H = Aut(X), say n; thus H is a nite group
(remarks 1 and 2 below). To nd the bound on n we make a detailed study of the
Riemann-Hurwitz formula.
The set of points of an automorphism of X (other than the identity) are nite,
and since H is nite as well, we have that the set of points of X xed by some
non-trivial element of H is a nite set. Let p
1
, . . . , p
r
be a maximal set of inequiv-
alent xed points of non-trivial elements of H. That is, each p
j
is xed by some
automorphism of X not equal to the identity; and if j = k, we have that h(p
j
) = p
k
for all h H. Thus these points project under q to dierent points of Y . For each
j, let
j
be the order of the subgroup H
j
of H of automorphisms of X xing p
j
. We
have that there are n/
j
distinct points in X that project to the same point, q(p
j
)
of Y , and each such point is xed by a subgroup of H of order
j
. Thus we obtain
that the total branching number B of the mapping q is given by
B =
r

j=1
n

j
(
j
1) = n
r

j=1
_
1
1

j
_
.
Observe that
j
2, so 1(1/
j
) 1/2. The Riemann-Hurwitz formula in this
setting gives us
2g 2 = n
_
2 2 +
r

j=1
_
1
1

j
_
_
,
where is the genus of Y (and g ). If g = then n = 1 (recall that n is the
order of Aut(X), which we are trying to bound). In the case of g > we have the
following cases:
2. Then 2(g 1) 2n implies that n g 1.178 3. UNIFORMIZATION OF RIEMANN SURFACES
= 1. In this case we have a value
j
2, so the right hand side of the
Riemann-Hurwitz relation is, at least, equal to n/2, or equivalently, n
4(g 1).
= 0 and r 5. This cases gives n 4(g 1).
= 0 and r = 4. Since the right hand side of Riemann-Hurwitz relation must
be positive we get that at least one
j
3, and n 12(g 1).
= 0, r = 3. We can assume that 2
1

2
nu
3
. Then
3
> 3 and

2
3. There are several cases to study:
a. If
3
7 we get n 84(g 1), with equality in the case of
1
= 2,
2
= 3
and
3
= 7.
b.
3
= 6,
1
= 2. Then
2
4 and n 24(g 1).
c.
3
= 6,
1
3. Then n 12(g 1).
d.
3
= 5,
1
= 2. Then
2
4 and n 40(g 1).
e.
3
= 5,
1
3. Then n 15(g 1).
f.
3
= 4,
1
3. Then n 24(g 1).
Remarks. 1. If G is a Fuchsian group then H/G is a Riemann surface. 2. It
is easy to show that, in the situation of the above proof, there exists a point p X
which is not xed by any non-identity element of Aut(X). This shows that the order
of the covering X Y = X/Aut(X) is equal to the order of the group Aut(X).
3.5. Moduli spaces
3.5.1. So far in this book we have studied properties of a xed Riemann surface.
The problem of moduli spaces deals with the study of varying Riemann surface
structures on a xed topological surface. More precisely, two surfaces X and Y
are said to be conformally equivalent (or simply equivalent) if there exists a
biholomorphic mapping between them, f : X Y . Our goal is to know under what
conditions X and Y are equivalent. An example of this type of problem is given by
the Uniformization theorem (3.2.8); it classies all simply connected surfaces up to
biholomorphisms. The general problem is dicult and the study of it constitutes3.5. MODULI SPACES 179
a whole area of research on its own, with new mathematical tools. In this section
we will give a couple of examples of how this problem can be treated; the reader
interested on more results can nd a nice introductory text in [18].
3.5.2. Before we get to explain our examples we need to make a few general
remarks on the relation between conformally equivalent surfaces and their universal
coverings and covering groups. Let X and Y be two surfaces, with universal cov-
erings

X and

Y , and covering mappings
X
and
Y
respectively. Let

f :

X

Y
be the lift of f to the universal covering spaces; the following diagram will be then
commutative:

f
//

X
f
//
y
In particular we have that

f is a Mobius transformation (see 1.3.14, 3.3.3
and 3.3.9). The spaces

X and

Y are homeomorphic; we will then identify them
and consider

X as the universal covering space of both X and Y .
We can also give a more algebraic statement, in terms of the covering groups.
Let G
X
and G
Y
be the covering groups of X and Y respectively. Since we have
identied the universal covering spaces of X and Y we can consider these two groups
as subgroups of Aut(

X). Then the mapping

f will satisfy

fG
X

f
1
; that is, G
X
and
G
Y
are conjugate subgroups of Aut(

X). We will use this formulation of the problem
in our examples since it make many of the computations easier.
3.5.3. Our rst example consists on the study of equivalence classes of annuli.
Let r
1
, r
2
and r be real numbers satisfying r
1
< r
2
and 1 < r; we denote by
A(z
0
, r
1
, r
2
) the annulus {z C; r
1
< |zz
0
| < r
2
} and by A
r
the annulus A(0, 1, r).
Clearly A(z
0
, r
1
, r
2
) is equivalent to A
r
2
/r
1
by the transformation z
1
r
1
(z z
0
). So
every annulus is conformally equivalent to one of the form A
r
, which means that the
space of equivalence classes of annuli is contained in the interval (1, +). To fully
determine conformal equivalence of annuli we need to nd under what conditions
A
r
and A
s
(with r and s real numbers greater than 1) are equivalent. The universal
covering of A
r
is given by
r
: H A
r
, where
r
(z) = exp (2i log z/ log ), and r180 3. UNIFORMIZATION OF RIEMANN SURFACES
and are related by the expression r = exp (2
2
/ log ) (3.3.15). Here log is the
principal branch of the logarithm dened on C\[0, +). The covering group G
r
is
generated by the transformation g
r
(z) = z; that is, G
r
= {z
n
z; n Z}.
Let G
s
denote the covering group of the annulus A
s
, and let g
s
(z) = z be a
generator of G
s
. If f : A
r
A
s
is a biholomorphism, and

f : H H a lift to the
universal covering space, then

f will be an element of SL(2, C). Since

f conjugates
G
r
into G
s
we have that

fg
r

f
1
is equal to either g
s
or g
1
s
. The transformation
M(z) = 1/z is an automorphism of H that conjugates g
s
into g
1
s
, so we can
assume that

fg
r

f
1
= g
s
(otherwise we consider M

f, which is also a lift of f). The
xed points of g
r
and g
s
are 0 and , so

f({0, }) = {0, }. But since g
n
r
(z
0
)
for z
0
in H and n + we must have

f() = , and therefore

f(0) = 0. So

f
is of the form

f = k z, for k a positive real number. A simple computation shows
that, with this expression of

f, we have

fg
r

f
1
= g
r
, which mean that r = s. Thus
we have proved the following result.
Theorem. Any annulus A(z
0
, r
1
, r
2
) is conformally equivalent to one and only
one annulus of the form A
r
, for r > 1. More precisely we can take r = r
2
/r
1
.
Remark. See [22, pg. 291] for a purely analytic proof of the above theorem.
3.5.4. Consider now the case when X and Y are surfaces of genus 1. By the
Abel-Jacobi theorem (2.9.13) they must be of the form C/G

= T

, for some with


positive imaginary part. The classication of tori is given by the next result.
Theorem. Two tori T

and T

are conformally equivalent if and only if


(7) =
a + b
c + d
,
where a, b, c and d are integer numbers satisfying ad bc = 1 (that is, and are
related by an element of SL(2, Z)).
Proof. Let us denote by T

n,m
the transformation z z +n+m (and similarly
for ). Suppose

f is a mapping satisfying

fG


f
1
= G

. Since

f is an automorphism
of C it must be of the form

f(z) = z+. It is easy to check the following identities:

f T

1,0

f
1
(z) =z +

f T

0,1

f
1
(z) =z + .3.5. MODULI SPACES 181
We must then have = c +d and = a +b, for some integers a, b, c and d; that
is, and satisfy the relation 7. Since

f
1
G


f = G

, the transformation z
az+b
cz+d
must be invertible; that is, ad bc = 1. The imaginary part of
a+b
c+d
is equal to
adbc
Im()
, so we must have ad bc = 1. This proves one half of the theorem.
Assume now that and are related by an element of SL(2, Z), as in the theorem.
Let S be the Mobius transformation given by S(z) = cz + d. If c = 0 then a = d =
1, which implies that = so there is nothing to prove. Thus we can assume that
c = 0. It is not dicult to see that
ST

n,m
S
1
= T

nd+mb,nc+ma
;
this equation implies that SG

S
1
G

. Choosing (n, m) = (a, c) and (n, m) =


(b, d) we get that SG

S
1
contains the transformations T

1,0
and T

0,1
, and thus
SG

S
1
G

. So the transformation S conjugates G

into G

and therefore the


tori T

and T

are conformally equivalent.


3.5.5. From the above theorem we have that the space of equivalence classes of
tori, denoted by M
1
can be identied with H/SL(2, Z). To study this space we can
follow the techniques of 3.4. It is not dicult, for example, to nd a fundamental
domain for the action of SL(2, Z) on H. Let P be the open polygon bounded by the
geodesics:
L
1
={z H; Re(z) = 1/2} ,
L
2
={z H; Re(z) = 1/2} ,
L
2
={z H; |z| = 1} .
Claim. P is the Dirichlet region (for the action of SL(2, Z) on H) centred at the
point 2i.
Proof of the claim. We rst prove that L
1
is contained in the set of points
equidistant from 2i and 2i + 1; that is, L
1
D
A
for A(z) = z + 1, in the nota-
tion of 3.4.16. Observe that since d(z, 2i) = d(z, 2i + 1) we have d(R(z), R(2i)) =
d(R(z), R(2i + 1)) for R(z) =
1
2
+ iz. But then we have
d(R(z), 2i + 1) = d(R(z), R(2i)) = d(R(z), R(2i + 1)) = d(R(z), 2i).182 3. UNIFORMIZATION OF RIEMANN SURFACES
In other words, if d(z, 2i) = d(z, 2i + 1) then d(R(z), 2i) = d(R(z), 2i + 1). Since
R(z) = z if and only if Re(z) =
1
2
we have that L
1
D
A
(the geodesic D
A
is the
full vertical line containing L
1
).
Similarly one can prove that L
2
D
A
1 and L
3
D
B
, where B(z) = 1/z. These
three statements show that D
2i
is contained in P.
Assume now that D
2i
is a proper subset of P. Then there exists a point z
0
P
and a non-trivial element h SL(2, C), such that h(z
0
) P. Write h(z) =
az+b
cz+d
.
We have (see 3.3.9) Im(h(z
0
)) =
Im(z
0
)
|cz
0
+d|
2
. Write z
0
= x
0
+ iy
0
. Since z
0
is in P we
have
|z
0
|
2
= x
2
0
+ y
2
0
> 1 and
1
2
< x
0
<
1
2
.
Using these inequalities one can easily prove the following:
|cz
0
+ d|
2
= c |z
0
|
2
+ 2 c d x
0
+ d
2
> |c|
2
+|d|
2
|cd| = (|c| |d|)
2
+|cd|.
The last term in the above displayed formula is a positive integer (it cannot be 0
since ad bc = 0). Thus it is at least equal to 1, which implies that |cz
0
+ d|
2
> 1.
So we have that if z
0
and h(z
0
) are both in P then Im(h(z
0
)) < Im(z
0
). We can
apply he same argument to h(z
0
) and z
0
= h
1
(h(z
0
)), to get Im(z
0
) < Im(h(z
0
)).
This contradiction shows that P is indeed equal to D
2i
.
To have a picture of the space of (conformal equivalence classes) of tori we just need
to consider P, the closure of P in the hyperbolic plane, by the action of SL(2, Z)
(observe that the point of innity is not a part of P). The transformation A identies
L
2
and L
1
, while B xes the side L
3
(as a set, not necessarily pointwise). Thus we can
think of P as consisting of four sides, L
1
, L
2
, s = {z H; |z| = 1, 0 Re(z) <
1
2
}
and s

= {z H; |z| = 1,
1
2
< Re(z) 0}, with B(s) = s

(condition FD3 in
denition 3.4.14). Any torus will be conformally equivalent to one torus of the form
T

where belongs to P. The precise formulation is given in the next result.


Theorem. Any torus is conformally equivalent to one and only one torus T

with satisfying the following conditions:


1. || 1;
2.
1
2
< Re()
1
2
;
3. If || = 1 then Re() 0.3.5. MODULI SPACES 183
3.5.6. The boundary of P has three vertices: i,
1+i

3
2
and
1+i

3
2
(the point i
is the meeting point of the sides s and s

). The tori corresponding to these points


are special in the sense that we explain next.
An automorphism M : C C of the complex plane induces an automorphism on
a torus T

if and only if MG

M
1
= G

, where and are related by an element


of SL(2, Z). We observe that if M is of the form M(z) = z + , for a complex
number, then MG

M
1
is actually equal to G

. Moreover, if is not of the form


n + m (for n and m integers), then M induces a non-trivial automorphism of T

.
Thus we have that any torus has a group of automorphism with many elements
(see the remarks before Hurwitzs theorem in 3.4.20). The mapping M(z) = z
also conjugates G

into itself, so it will give another automorphism of T

. The xed
points of M are given by points z
0
of C satisfying M(z
0
) = z
0
+ n + m, for some
integers n and m. It is easy to see that there are only four possible points, up to
equivalence by elements of G

: 0, 1/2, /2 and (1 +)/2. So if f : T

denotes
the automorphism of the torus T

induced by M, we see that f has four xed points


on T

, corresponding to the above four points. In particular one can prove that f is
the hyperelliptic involution of T

(we use quotation marks since we have dened


hyperelliptic involutions only for surfaces of genus at least 2).
Suppose P corresponds to a torus with some automorphism dierent from
the ones in the previous paragraph. Then there exists a Mobius transformation
S(z) =
az+b
cz+d
, in SL(2, Z), such S() = . Solving this equation we get that should
be of the form
=
a d
2c
+ i
_
4 (a + d)
2
2c
.
We have that |a + d| < 2 (otherwise would be real) and c = 0 (since z z +
does not have xed points on P). Since || 1 we get
1ad
c
2
1. This inequality
gives us the following dierent options:
a = 0, c = 1, d = 0,
1
= i;
a = 1, c = 1, d = 0,
2
=
1 + i

3
2
;
a = 1, c = 1, d = 0,
3
=
1 + i

3
2
;184 3. UNIFORMIZATION OF RIEMANN SURFACES
Consider the torus T

1
, and let T
1
be the Mobius transformation T
1
(z) =
1
z(= iz).
We have T
4
1
= Id and T
1
T
i
n,m
T
1
1
= T
i
m,n
. Thus T
1
induces an automorphism of
order 4 on T

1
. The tori corresponding to
2
and
3
are conformally equivalent (by
the transformation z z 2), so we consider only one of them, say T

2
. Using
the identity
3
2
= 1 we see that T
2
(z) =
2
z satises T
2
T

2
n,m
T
1
2
= T

2
m.nm
. Since
the mapping (p, q) (q, p q) of Z
2
is invertible we have that T
2
induces an
automorphism on T

2
. It is easy to check that the order of that automorphism is 6.
We obtain that the tori corresponding to the vertices of P are precisely those with
some extra automorphisms.
3.5.7. Topologically all annuli are the same, that is, homeomorphic, and
similarly all tori. If one takes two annuli and identies the boundaries (glue
them by their boundaries) one gets a torus. This particular surface has an order 2
mapping interchanging the two annuli. Such mapping cannot be holomorphic (in
the torus) since it has many xed points, namely the two curves that formed the
boundaries of the two annuli. However, it is possible to show that this mapping is
anti-holomorphic (that is, its conjugate is holomorphic). More precisely, let = it
be a complex number with t > 1, and consider the symmetry (anti-holomorphic
mapping of order 2) of the complex plane given by : z z. We have
T

n,m

1
(z) = T

n,m
( z) = ( z + n + it) = z n it = T

n,m
(z).
Thus induces an automorphism of T

, say R. It is easy to check that R is anti-


holomorphic and has order 2. What are the xed points of R? If a point p in T

is xed by R, then there exist integers n and m such that z


0
= z
0
+ n + mti,
where z
0
is a point of C that projects to p under the natural quotient map. The
solutions of this equation are given by the lines L
n
= {z C; Re(z) =
n
2
}. Since L
n
is equivalent under elements of G

to L
n2n
we have only two set of solutions, the
imaginary axis and L
1
. These two lines project to two closed curves on T

(since
z z + ti belongs to G

), the ones corresponding to the boundaries of the annuli


above explained.
Consider now the transformation (z) = z; it is clear that induces an anti-
holomorphic involution on H. If S is an element of SL(2, C), then S = S, so3.5. MODULI SPACES 185
induces a mapping on M
1
, the space of tori. If is xed point of then there
exists an element T SL(2, C) such that () = T(). But then T T
1
will be
an anti-holomorphic mapping of H xing . So, without loss of generality we can
assume that () = . The solutions of this equation are given by H satisfying
Re() = 0; that is, the tori built by gluing two annuli.186 3. UNIFORMIZATION OF RIEMANN SURFACES187
Exercises
Exercise 1. Let f : C C be a holomorphic function. Assume that there exists
a constant M, such that Re(f(z)) M, for all z C. Show that f is constant.
Exercise 2. Prove Liouvilles theorem: a bounded holomorphic function dened
on the complex plane, f : C C, is constant.
Exercise 3. Prove the Fundamental Theorem of Algebra: any polynomial with
complex coecients has at least one root in C.
Exercise 4. Let f : D C be a holomorphic function dened on an open set
of the complex plane. Write f = u + iv (u and v are the real and imaginary parts
of f, respectively). Using the Cauchy-Riemann equations show that u and v are
harmonic functions.
Exercise 5. Prove by direct computation that if X is a simply connected surface
then H
1
(X, Z) is trivial.
Exercise 6. Show that any connected, locally path connected topological space is
path connected.
Exercise 7. Let p : X Y a covering. Show that X is a manifold if and only if Y
is a manifold.
Exercise 8. Let G be a group and H = [G, G] its commutator subgroup; that is,
H is generated by all elements of the form ghg
1
h
1
, for g and h in G. Prove that
H is a normal subgroup and G/H is abelian.
Exercise 9. Show that the punctured plane C

is homeomorphic to a cylinder.
Exercise 10. Prove the following version of Van Kampen Theorem:
let M be a topological space, U and V open subsets of M satisfying M = U V
and such that the intersection U V is path connected and non-empty. If U and V
are simply connected then M is simply connected.
Use this result to show that

C is simply connected.
Exercise 11. Consider the following subspace of R
2
: X = A B C, where:188
A = {(x, y); x 0, y = 1}, B = {(x, y); x 0, y = 1};
C = {(x, y); x < 0, y = 0}.
Dene a topology on X by putting the subspace topology on A\{(0, 1)}, B\{(0, 1)}
and C, and for the points (0, 1) (and (0, 1)) use the basis of neighbourhoods given
by N

= {(x, 1); 0 x < } {(x, 1); x < 0} (similarly for (0, 1)). Show
that X is locally homeomorphic to R but it is not a manifold.
Exercise 12. Let p : X Y be a covering mapping, where X and Y are man-
ifolds, and let p

:
1
(X, x
0
)
1
(Y, p(x
0
)) be the induced mapping between the
fundamental groups. Show that p

is injective.
Exercise 13. Find an example of a covering map p : X Y , and a continuous
function f : X X such that f p = f but f is not a homeomorphism (this shows
that the condition of homeomorphism cannot be removed from the denition on
1.1.20.
Exercise 14. Show that the result of exercise 7 is true if you substitute manifold
for Riemann surface.
Exercise 15. Can you nd a Riemann surface structure on C which is not compat-
ible with the usual structure, induced by (C, Id)?
Hint: C is homeomorphic to D.
Exercise 16. Prove that the stereographic projection P : S
2


C, given by
P(x
1
, x
2
, x
3
) =
_

_
x
1
+ix
2
1x
3
, if x
3
= 1,
, if x
3
= 1,
is a homeomorphism between the sphere S
2
of R
3
and the Riemann sphere

C.
Exercise 17. The projective line (or projective space of dimension 1) P
1
is dened as
follows: consider the equivalence relation in M = C
2
\{(0, 0)} given by (z
1
, z
2
)
(w
1
, w
2
) if there exists a non-zero complex number , such that z
1
= w
1
and
z
2
= w
2
. Set P
1
= M/ . Prove that

C is homeomorphic to P
1
.
Hint: denote by [z : w] the points P
1
(notation as in 2.7.1), consider the map from
P
1
to

C given by [z : w] z/w if w = 0 and [z : 0] .
Exercise 18. Let G be a group of Mobius transformations and D a connected
open subset of

C. Assume that the following conditions are satised:189
1. g(D) = D, for all g G.
2. For every point z in D, there exists a neighbourhood U of z, such that, if
g G satises g = ID, then g(U) U = .
Denote by D/G the quotient space of D by the action of G; that is, z
1
and z
2
are
equivalent if there exists g G, such that g(z
2
) = z
1
. Show that D/G carries a
unique Riemann surface structure such that the quotient mapping : D D/G is
holomorphic.
Exercise 19. Show that if f : X C is holomorphic and X is a compact Riemann
surface then f is constant.
Exercise 20. Show that the set of poles of a holomorphic function dened on a
Riemann surface is discrete.
Exercise 21. A mapping f : X Y between topological spaces is called proper
if f
1
(K) is compact in X, for any K compact subset of Y . Extend the denition
of degree and proposition 1.3.11 to the case of proper mappings (not necessarily
between compact surfaces).
Exercise 22. Prove proposition 1.3.14
Exercise 23. Let f : X Y be a holomorphic mapping between compact
Riemann surfaces of genera g and g

respectively. Show that if g = g

then f is a
covering map. Show that if g

= 1 then g = 1.
Exercise 24. Prove that the residue of a polynomial p(z) = a
0
+a
1
z + a
n
z
n
at
the point , when one considers p as a meromorphic mapping on

C, is equal to a
1
.
Exercise 25. Prove that a meromorphic function f :

C

C on the Riemann
sphere is rational, i.e. the quotient of two polynomials.
Exercise 26. Let R(z) =
p(z)
q(z)
be a rational function on the Riemann sphere, where
p and q are polynomials without common factors. Show that the point is a:
pole if deg (p) > deg (q);
zero if deg (p) < deg (q);
regular point if deg (p) > deg (q).
Exercise 27. Let f : X

C be a non-constant meromorphic function on a
Riemann surface. For a point p of X, let n denote the ramication index of f at190
that point. Dene
z(q) =
_

_
(f(q) f(p))
1/n
, if f(p) = ,
(f(q))
1/n
, if f(p) = ,
for q is a neighbourhood of p. Show that z is a local coordinate on X.
Exercise 28. Let X be a compact surface, p
1
, . . . , p
n
distinct points of X. Prove
that any non-constant holomorphic function f : X

= X\{p
1
, . . . , p
n
} C comes
arbitrarily close to any point z C. More precisely, show that for any z C, and
any > 0, there exists a point p X

such that |f(p) z| < .


Exercise 29. Let and be two complex numbers with positive imaginary parts.
Let T

and T

denote the corresponding tori (see 1.3.6). Show that if is a complex


number satisfying G

, then the function f : C C, given by f(z) = z,


induces a holomorphic mapping from F : T

. Prove that F is biholomorphic


if and only if G

= G

.
Exercise 30. Let f : X Y be a non-constant holomorphic mapping between
Riemann surfaces. Show that the mapping f

: O(Y ) O(X), dened by f

(g) =
g f, is a ring homomorphism.
Exercise 31. Let T

= C/G

be as in 1.3.6. Show that T

is a connected Hausdor
space. Moreover, show that T

is a topological manifold of dimension 2.


Exercise 32. Show that the operators and

satisfy = 0 and



= 0, by
direct computation.
Exercise 33. Let be a smooth (complex valued) 2-form on a surface X with
compact support. Let U = {U
i
}
n
i=1
and V = {V
j
}
m
j=1
be two nite open coverings
of the support of by coordinate patches. Let {f
n
}
n
i=1
and {g
j
}
m
j=1
be partitions of
unity subordinated to U and V respectively. Show that
n

i=1
_
X
f
i
=
m

j=1
_
X
g
j
.
Exercise 34. Let X =

X/G be a Riemann surface, where

X is the universal
covering space and G the group of deck transformations. Let p :

X X be the
covering map. Assume is a closed 1-form on X, and F a primitive of p

(), that
is, F :

X C satises dF = p

(). Show that for every g in G there exists a191


complex number a
g
, such that F F g
1
= a
g
.
Exercise 35. Use the Residues theorem to show that there does not exist a
function on C/G

with a just pole of order 1.


Exercise 36. Let Gbe a group with at least two elements, and X a topological space
with at least two non-empty disjoint open sets. Dene a presheaf F by F(U) = G
if U X is a non-empty open set, and F() the trivial group. The restriction
homomorphisms : F(U) F(V ) are given by the identity homomorphism if
V = and the trivial homomorphism if V = . Show that F is a presheaf but not
a sheaf. Compute the associated sheaf.
Hint: consider two disjoint open sets U
1
and U
2
, and let f F(U
1
U
2
) be given
by g
1
in U
1
and g
2
in U
2
.
Exercise 37. Show that if X is a simply connected surface, then H
1
(X, C) and
H
1
(X, Z) are trivial.
Exercise 38. Let C

= C\{0}. Consider the covering given by the open sets


U
1
= C

\R

and U
2
= C

\R
+
. Here R

and R
+
are the negative and positive real
axis, respectively.
1. Prove that U = {U
1
, U
2
} is a Leray covering for the sheaf of locally constant
functions with integer values Z.
2. Show that Z
1
(U, Z)

= Z(U
1
U
2
) Z(U
1
U
2
).
3. Show that Z(U
i
)

= Z, for i = 1, 2.
4. Prove that the boundary operator : C
0
(U, Z) C
1
(U, Z) is given by
(a
1
, a
2
) (a
2
a
1
, a
2
a
1
).
5. Conclude that H
1
(C

, Z)

= Z.
Exercise 39. Show that if X is a compact surface, the mapping H
1
(X, Z)
H
1
(X, C) induced by the inclusion Z C is injective. Here Z and C denote the
sheaves of locally constant functions with values in the integer and complex numbers
respectively.
Exercise 40. Show that the sheaf sequence:
0 C


0,
where (f) = df/f, is exact.
Exercise 41. Show that the mapping

of 1.5.15 is well dened.192


Exercise 42. Show that two divisors on the Riemann sphere are linearly equivalent
if and only if they have the same degree.
Exercise 43. Let g : C C be a smooth function with compact support. Show
that g is uniformly continuous. Use this to prove that if C, then the integral
_
2
0
g( + re
i
) d
converges to 2g() when r 0 (see 2.2.1).
Exercise 44. Use the inequality |(h
n+1
h
n
p
n
)(z)| < 2
n
to prove that the
function in 2.2.2 (pg. 64) is well dened.
Exercise 45. Let X be a compact Riemann surface and f : X

C a non-constant
meromorphic function of degree d. Show that the sheaf of meromorphic functions
on X is a nite algebraic extension eld of C(f) (the eld of rational functions on
f) as follows:
a) Let C be the set of critical points of f; that is, C consists of the points p X
such that d
p
(f) = 0. Set B = f(C) (critical values of f) and A = f
1
(C). Show
that f : X\A

C\B is a covering of degree d.
b) Let g : X

C be a non-constant meromorphic function. Let S be the set
of poles of g, S = g
1
(). If z

C\B\S and f
1
(z) = {p
1
, . . . , p
d
} (these points
are distinct), dene a
j
as the j-th elementary function on g(p
1
), . . . , g(p
d
). These
means that the a
j
, for j = 1, . . . , d satisfy the following equation:
(g(p))
d
+ a
1
(f(p)) (g(p))
d1
+ + a
d
(f(p)) 0,
for all p in X\A\f
1
(f(S)). Show that a
j
are holomorphic functions.
c) We want to show that a
j
extend to meromorphic functions on X. That will
prove that the equation displayed above holds for all points of X and will nish the
exercise. To show that, let a be a point in B f(S), choose a neighbourhood U of
a such that the poles of g on U lie in f
1
(a). Let w be a holomorphic function on
U, not identically 0 but satisfying w(a) = 0. Show that (w f)
n
g is holomorphic
for some non-negative integer n (on f
1
(U)). Let b
j
be the elementary symmetric
functions of (wf)
n
g on f
1
(z) = {p
1
, . . . , p
d
}, for z W\{a}, where W is an open
subset of U. Show that functions b
j
extend to holomorphic functions at a. Use the
relation a
j
= b
j
/w
jn
to show that a j is meromorphic on X.193
Exercise 46. Prove the Riemann-Hurwitz formula using the Riemann-Roch the-
orem.
Exercise 47. Prove the assertions of 2.4.7
Exercise 48. Let L be a line bundle on a Riemann surface X and L
1
the dual
bundle. Prove that for any point p X there is an isomorphism between (L
p
)

, the
dual space of L
p
, and (L
1
)
p
.
Exercise 49. Show that the mapping between the Picard group X and H
1
(X, O

)
of 2.4.14 is a group isomorphism.
Exercise 50. Let K be a canonical divisor on a compact surface X of genus g.
Show that the dimension of H
0
(X, O(m =, K)) is given as in the table 1 at the end
of the exercises.
Exercise 51. Let K be a canonical divisor on a compact surface X of genus g.
Show that
dim H
0
(X, O(mK p)) = dim H
0
(X, O(mK)) 1,
for any point p of X.
Exercise 52. Compute the dimension of the spaces of higher order dierentials.
Exercise 53. Prove that projective embeddings i
m
(see 2.7.4 and 2.7.8) are well
dened.
Exercise 54. Recall the Weierstrass function for the torus T

(that is, T

= C/G

where () > 0) is given by the expression (2.6.6):


(z) =
1
z
+

=0
_
1
(z )
2

1

2
_
.
Here is of the form = n + m with n and m integers.
a) Prove that is an even function ((z) = (z)).
b) Show that the derivative of is given by
(

)(z) =

1
(z )
3
,
where the sum is taken over all points of the form = n + m, including = 0.
Use this expression to show that

is an odd function (

(z) =

(z)).
c) Let f

: T

C be the mapping (of degree 2) induced by on the torus T

.
Show that the ramication points of f

are given by the points corresponding to 0,194


1/2, /2 and (1 +)/2.
d) Show that

vanishes (or it has a singularity if you do not take local coordi-


nates) precisely at the four points given above (and the corresponding points on C
congruent via the group G

).
e) Show that the mapping i : T

P
2
given by p [1 : (p) :

(p)] denes an
embedding of T

into the 2-dimensional projective space.


Exercise 55. Complete the proof of theorem 2.7.11 as follows:
a) To dene the map
L
: X P
n
in a neighbourhood of a point p we have used
a section that does not vanish at p. Show that if
1
is another section dened in
a neighbourhood of p, say U

, with
1
(p) = 0 then for every point q U U

we
have that
s
j
(q)
(q)
= h(q)
s
j
(q)

1
(q)
, where h is a non-zero holomorphic section. This shows
that
L
does not depend on the choice of as long as
L
(q) is in P
n
(see below).
b) Use 2.6.5 to prove that the sections s
j
do not have common zeroes and there-
fore
L
maps X into P
n
.
c) In the text we have shown that if p and q are two distinct points of X then
there exists a section of L which vanishes at p but it does not vanish at q. Use this
fact to prove that
L
is injective.
d) We have also shown that if p X then there exists a section of L with a
simple zero at p. Using this fact show that the dierential of
L
does not vanish.
This completes the proof of the fact that
L
: X P
n
is an embedding.
Exercise 56. Let z
1
, . . . , z
2g+2
be distinct points on the Riemann sphere. Prove
that the Riemann surface of the polynomial
w
2
= (z z
1
), . . . , (z z
2g+g
)
is a compact surface X of genus g. Let z denote the function on X, and p
j
the
preimage of z
j
(j = 1, . . . , 2g + g). Without loss of generality we can assume that
the points p
j
are not equal to or 0. Assume the divisor of z is given by div(z) =
q
3
+q
4
q
1
q
2
; prove that the divisor of w is given by div(w) = (p
1
+ +p
2g+2
)
(g + 1)q
1
(g + 1)q
2
. Use this to prove that the 1-forms
z
j
dz
w
, for j = 0, . . . , g 1
form a basis of the space of 1-forms on X. Basis for holomorphic dierentials on a
hyperelliptic surface.195
Exercise 57. Show that if f : X C is a meromorphic function of degree d on a
hyperelliptic surface of genus g, and d g, then d is even.
Exercise 58. Prove that if U is an open, connected subset of C, and A is a discrete
subset of U, then U\A is connected.
Exercise 59. Let 1 <
1
< <
g
2g be the complementary set of the numbers
n
1
, . . . , n
g
of Weierstrass theorem (2.8.1) in {1, . . . , 2g}. Call these numbers the
non-gaps at the point p (the n
j
s are called the gaps). Prove the following statement
regarding the non-gaps:
1. If
j
and
k
are two non-gaps with
j
+
k
2g, then
j
+
k
is also a
non-gap.
2. For each integer j with 0 < j < g one has
j
+
gj
2g.
3. If
1
= 2 then
j
= 2j and
j
+
g1
= 2g, for 0 < j < g.
4. If
1
> 2, then for some j with 0 < j < 2g one has
j
+
gj
> 2g.
5.

g1
j=1

j
g(g 1), and the equality occurs if and only if
1
= 2.
Exercise 60. Prove lemma 2.8.4 for the case of m 3.
Exercise 61. Let X be a hyperelliptic compact surface of genus g and f : X

C
a function of degree 2. Let j : X X denote the hyperelliptic involution. This
exercise shows an explicit basis of H
0
(X, ). Assume that f is not branched over
; that is that the xed points of j are not poles of f.
a) Choose a point p in X not xed by j; show that there exists a meromorphic
function u : X

C such that u(q) = u(j(q)) for q in a neighbourhood of p, and
that we can choose u to be holomorphic at p and j(p).
b) It follows from exercise 45 that there is an equation of the following form:
u
2
(p) + 2a
1
(f(p)) u(p) + a
2
(f(p)) = 0.
Write (u+a
1
)
2
= p/q, where p and q are polynomials on z. Then show that pq = p
1
q
2
1
,
where p
1
does not have multiple roots; moreover, p
1
(z) = c(z z
1
) (z z
2g+2
),
where c = is a complex number and z
j
= f(p
j
), for p
j
, j = 1, . . . , 2g + 2 the
Weierstrass points of X. Show that z
j
= z
k
if j = k.
c) Dene w = (u + a
1
)q/q
1
. Show that w(p
1
) = w(p
2
) for f
1
(z) = {p
1
, p
2
} in
an open set of X. Show that w = p
2
1
.196
d) Show that w(p) == w(j(p)) for p in an open set of X but w(p)
2
= w(j(p))
2
in that open set, and therefore on X by the identity principle. This implies that
w(p) = w(j(p)) on X. Show that w is a function of degree g + 1 on X.
e) Dene forms for j = 0, . . . , g 1 by

j
(p) =
f
j
(p) f

(p) dz
w(f(p))
,
where z is a local coordinate on X. Use the identity w = p
2
1
to show that
j
=
2f
j
dw/p

1
on f
1
(C) and thus
j
is holomorphic at those points.
f) To prove that
j
is holomorphic show that near a pole of f we have w =
c
1
f
g+1
O(1 +
1
z
) and therefore
j
= c
1
f
jg1
(1 +O(
1
z
)) is also holomorphic.
Exercise 62. Let f : X

C be a function of degree 2 on a hyperelliptic Riemann
surface X of genus g. Denote by p
1
, . . . , p
2g+2
the Weierstrass points of X. Show
that the images of these points under f are distinct; that is, f(p
j
) = f(p
k
), for
1 j < k 2g + 2.
Exercise 63. Prove Noether Gap Theorem (2.8.2).
Exercise 64. If D is a divisor on a surface X we say that O(D) is globally
generated if there exists a function f H
0
(X, O(D)) such that ord
p
(f) = D(p)
for every p in X. Equivalently, every function g dened on a neighbourhood of p with
div(g) D can be written as g = h f, where h is holomorphic on a neighbourhood
of p. Show that if D has degree greater than 2g 1 and X is compact of genus g
then O(D) is globally generated.
Exercise 65. Let X = C/G

, p
0
a point of X, D
n
the divisor np
0
. Show that
dim H
0
(X, O(D
n
)) =
_

_
0, n < 0,
1, n = 0,
n, n > 0.
Exercise 66. Consider the homeomorphism w : C D given by
w(z) =
z
1 +|z|
.
This mapping induces a Riemann surface structure on C, given by a single chart
(C, w). Denote by C
1
the complex plane with this structure. Let u : C R be197
dened by u(z) =
x
1+|z|
, where z = x + iy. Show that u is harmonic on C
1
, but it is
not when we consider the standard structure of the complex plane.
Exercise 67. Let A be an open subset of the complex plane, p a point in the
boundary of A. Suppose that there exists a disc D centred at a point q A such
that D A {p} and D A. Show that p is a regular point of A.
Hint: let c be the middle point of the segment joining p and q. Show that the
function (z) = log(r/2) log(|z c|) is a barrier at p.
Exercise 68. Let X be a Riemann surface,

X its universal covering space (iden-
tied with

C, C or H) and G the group of deck transformations (identied with a
group of Mobius transformations). Show that X is biholomorphic to the quotient

X/G.
Exercise 69. Let A(z) =
az+b
cz+d
be a Mobius transformation with ad bc = 1.
1. Show that A has order 2 (that is, A A = Id) if and only if a + d = 0.
2. Prove that A has only one xed point in

C if and only if |a + d| = 2.
Exercise 70. The Picard subgroup G of Aut(

C) is given by the transformations


of the form z
az+b
cz+d
, satisfying:
1. ad bc = 1.
2. a, b, c, d are in Z[i]; that is, their real and imaginary parts are integers.
Show that G is discrete but it does not act properly discontinuously at any point of
the Riemann sphere.
Exercise 71. Prove the second lemma in 3.3.5
Exercise 72. Show that |z w|
2
|1 wz|
2
| = |z|
2
+|w|
2
|z|
2
|w|
2
1. Use this
to show that if |z| = 1 and |w| < 1 then |
zw
1 wz
| < 1. Prove proposition 3.3.8
Exercise 73. Let G be a group of Mobius transformations of the form
T

: z z +, where is a complex number. Let r = inf{|}; T

G}. Show that


if there does not exist T

in G with r = || then G cannot be discrete.


Exercise 74. Compute the area of a hyperbolic pentagon with angles given as in
picture 13. More generally, compute the area of a hyperbolic pentagon (or convex
polyhedron).
Exercise 75. Show that if X = H/G is a compact surface of genus 2 then N(G)
strictly contains G.198
Exercise 76. Let z
i
, i = 1, . . . , 4 be four distinct point in the Riemann sphere.
Compute the cross ratios (z
(1)
, z
(2)
, : z
(3)
, z
(4)
) as varies over all 24 permutations
of four symbols. Show that there are only 6 distinct values among all these cross
ratios.
Exercise 77. Let C be a line or circle orthogonal to S
1
and A an automorphism
of the unit disc. Show that A(C) is a line or circle orthogonal to S
1
.
Exercise 78. Show that the mapping z z is an isometry in the hyperbolic
metric of H.
Exercise 79. Prove lemma 3.4.4 by direct computation.
Exercise 80. Prove lemma 3.4.17.
Exercise 81. Prove that the exponential sequence
0 Z C
exp
C

0
is exact.
Exercise 82. Show that the eld of meromorphic functions on a compact surface
is an algebraic eld of one variable. More precisely, let f : X

C be a non-constant
meromorphic function on a compact surface X; let d denote the degree of f. Let g be
another meromorphic function on X. Prove that there exist meromorphic functions
a
j
:

C

C, j = 1, . . . , d, such that
(g(p))
d
+ a
1
(f(p)) (g(p))
d1
+ + a
d
(f(p)) = 0,
for all p in X.
Exercise 83. Show that a discrete subset of the complex plane is either nite or
innite countable.
Hint: cover C by a countable increasing sequence of compact sets.199
g m dimension
0 m 0 1 2m
m > 0 0
1 0
g 1 m < 0 0
m = 0 1
m = 1 g
m > 1 (2g 1) (g 1)
Table 1.

Figure 15. Hyperbolic pentagon.200


Bibliography
[1] L. V. Ahlfors. Complex Analysis. Mc-Graw Hill, New York, USA, 1966.
[2] L. V. Ahlfors and L. Sario. Riemann surfaces. Princeton University Press, Princeton, New
Jersey, USA, 1960.
[3] W. M. Boothby. Introduction to Dierential Manifolds and Riemannian Geometry. Pure and
Applied Mathematics. Academic Press, New York, 1975.
[4] G. E. Bredon. Topology and Geometry, volume 139 of Graduate Texts In Mathematics.
Springer-Verlag, Berlin and New York, 1993.
[5] J. B. Conway. Functions of one complex variable, volume 11 of Graduate Texts In Mathematics.
Springer-Verlag, Berlin and New York, 1973.
[6] J. B. Conway. A Course in Functional Analysis, volume 96 of Graduate Texts In Mathematics.
Springer-Verlag, Berlin and New York, 1990.
[7] H. Farkas and I. Kra. Riemann Surfaces, volume 72 of Graduate Text in Mathematics.
Springer-Verlag, New York, Heidelberg and Berlin, 2nd edition, 1992.
[8] O. Forster. Lectures on Riemann Surfaces, volume 81 of Graduate Text in Mathematics.
Springer-Verlag, New York, Heidelberg and Berlin, 1981.
[9] M. J. Greenberg and J. R. Harper. Algebraic Topology, A First Course. The Benjamin Cum-
mings Publising Company, Reading, Massachusetts, U.S.A., 1981.
[10] P. Griths and J. Harris. Principles of Algebraic Geometry. Wiley Classics Library. John
Wiley & Sons, New York, USA, 1978. (First edition 1951).
[11] N. Jacobson. Basic Algebra, volume I. Hindustani Publishing Corporation (India), New Dehli,
India, 1974. Originally published by W. H. Freedman, 1974.
[12] S. Lang. Algebra. Addinson-Wesley Publishing Company, Reading, Massachusetts, U. S. A.,
3rd edition, 1993.
[13] W. S. Massey. Algebraic topology: an introduction. Harcourt, Brace & World, New York, 1967.
[14] W. S. Massey. Singular Homology Theory, volume 70 of Graduate Text in Mathematics.
Springer-Verlag, New York, 1980.
[15] J. Milnor. Topology from a Dierentiable Point of View. Universiry Press of Virginia, Char-
lottesville, U.S.A., 1965.
[16] J. R. Munkres. Topology: a rst course. Prentice-Hall of India, New Delhi, 11th indian edition,
1975.
201202
[17] J. R. Munkres. Topology: a rst course. Prentice-Hall, Englewood Clis, 1975.
[18] S. Nag. The Complex Analytic Theory of Teichm uller Spaces. John Wiley & Sons, 1988.
[19] R. Narasimhan. Compact Riemann Surfaces. Lectures in Mathematics, ETH Z urich.
Birkhauser, Basel-Boston-Berlin, 1992.
[20] W. Rudin. Real and Complex Analysis. McGraw-Hill, New York, USA, 1966.
[21] W. Rudin. Functional Analysis. McGraw-Hill, New York, USA, 1973.
[22] W. Rudin. Real and Complex Analysis. McGraw-Hill, New York, USA, international Edition
edition, 1987.
[23] R. R. Simha. The Uniformisation Ttheorem for planar Riemann surfaces. Arch. Math., 53:599
603, 1989.
[24] R.R. Simha. The Riemann-Roch theorem for compact Riemann surfaces. Ensign. Math. (2),
27:185196, 1981.
[25] M. Spivak. A Comprehensive Introduction to Dierential Geometry, volume 1. Publish or
Perish, Houston, USA, second edition, 1971.203
Notation
Term Page Meaning
Z - Integers
R - Real numbers
C - Complex numbers
C

- {z C; z = 0}
A\B - {p A; p / B}
Id
X
- Identity function on the set X
A B - disjoint union of sets A and B
U - boundary of the set U
D 4 unit disc, {z C; |z| < 1}
f
x
, f
y
, f
z
, f
z
2 partial derivatives
c
1
c
2
8 homotopic paths
(X, x
0
) 9 loops on the manifold X based at x
0

1
(X, x
0
) 9 fundamental group
Deck(X/Y ) 10 group of deck transformations (of a covering X Y )
supp(f) 11 support of a function
|| || 11 norm
S
2
13 2-sphere, {(x, y, z) R
3
; x
2
+ y
2
+ z
2
= 1}
C
n
(X) 13 chain group of X
: C
n
(X) C
n1
(X) 14 boundary operator between chain groups
H
n
(X, Z) 14 homology groups
(X) 15 Euler-Poincare characteristic
H 19 upper half plane, {z C; Im(z) > 0}

C 19 Riemann sphere or extended complex plane, C {}


T

21 torus
b
p
(f) 24 branching number of the function f at the point p
B(f) 27 total branching number of the function f,

pX
b
p
(f)
C(X) 29 sheaf of continuous functions on X204
S(X) 29 sheaf of smooth functions on X
S
1
(X) 30 sheaf of smooth 1-forms on X
S
(1,0)
(X) 30 sheaf of (1, 0) type forms on X
S
(0,1)
(X) 30 sheaf of (0, 1) type forms on X
S
2
(X) 31 sheaf of smooth 2-forms on X

1

2
31 exterior product of forms
d 32 exterior derivative
33 delta operator

33 delta bar operator


33 conjugation operator
T

X 36 cotangent space of a manifold X


M
1
(X) 40 meromorphic 1-forms on X
F 42 sheaf (or presheaf) on X
O(X) 43 sheaf of holomorphic functions on the surface X
M(X) 43 sheaf of meromorphic functions on the surface X
O

(X) 43 sheaf of non-zero holomorphic functions on the surface X


M

(X) 43 sheaf of meromorphic, non-identically zero functions


on the surface X
C
n
(U, F) 44 cochain group
Z
1
(U, F) 44 group of cocycles
B
1
(U, F) 44 group of coboundaries
H
1
(U, F) 44 cohomology group with respect to U
H
1
(X, F) 47 1st cohomology group
|F| 50 sheaf associated to a presheaf
: F G 50 sheaf homomorphism
K 51 kernel sheaf
S
1
c
54 sheaf of closed forms on X
H
1
dR
(X) 55 de Rham cohomology group
K(U) 44 kernel sheaf
D 59 divisor on a surface
D 0 59 eective divisor205
div(f) 59 divisor of a function
div() 60 divisor of a form
D
1
D
2
60 linearly equivalent divisors
deg (D) 60 degree of divisor D
Div
P
(X) 61 group of principal divisors of X
K
X
61 canonical class of X
O(D) 61 sheaf of meromorphic functions associated to a divisor
(D) 62 sheaf of meromorphic forms associated to a divisor
65, 128 Laplacian operator
C
x
68 skyscraper sheaf
L L

75 tensor product of bundles


L
1
75 dual bundle of L
Pic(X) 77 Picard group of the surface X
L(D) 79 line bundle associated to the divisor D
s
D
80 canonical section of the bundle L(D)
deg (L) 81 degree of the bundle L
K
X
86 canonical line bundle of the surface X
det(V ) 88 determinant line bundle of the bundle V
Res 93 Residue isomorphism between H
1
(X, K
X
) and C
97 Weiertrass -function
P
n
100 projective space
i
1
: X P
g1
101 canonical mapping
W(f
1
, . . . , f
m
) 109 Wronskian of the functions f
1
, . . . , f
m
A
j
(), B
j
() 116 periods of the form
J(X) 121 Jacobian variety of the surface X
PSL(2, C) 150 projective special linear group
(G) 158 region of discontinuity of the group G
(z
1
, z
2
; z
3
, z
4
) 163 cross ration of four complex numbers
D
p
(G) 173 Dirichlet region of G relative to p
N(G) 176 normalizer of G (in Aut(H))
A(z
0
, r
1
, r
2
) 179 annulus of centre z
0
and radii r
1
and r
2206 Index
Index
Numbers written in italic refer to the page where the corresponding entry is
described; numbers underlined refer to the denition; numbers in roman refer to the
pages where the entry is used.
Symbols
1-form . . . . . . . . . 30, 37
1-formintegration . . . . 38
1-forminvariance prop-
erty . . . . . . . . . 30
1-formmeromorphic . . . 40
1-formperiods . . . . . . 116
1-formsheaf of . . . . . . . 30
1-formtype (0, 1) . . . . . 30
1-formtype (1, 0) . . . . . 30
2-form . . . . . . . . . 31, 37
2-formintegration . . . . 39
2-forminvariance prop-
erty . . . . . . . . . 31
f
z
, f
z
. . . . . . . . . . . 2, 30
2-sphere . . . . . . . . . . . 13
2-sphereEuler-Poincare
characteristic . . 15
A
Abels theorem . . 122, 123
Abel-Jacobi map . . . . 121
ample line bundle . . . 106
annulus theorem . . . . 142
arithmetic genus . . . . . 68
automorphism 153, 155, 156
B
Banach space . . . . . . . 12
Banach spacecompact
mapping . . . . . . 12
barrier . . . . . . . . . . . 140
Betti number . . . . . . . 15
biholomorphism . . . . . 23
boundary operator . . . 14
branched covering . 10, 21
branching number . 24, 27
bundle . . . . . . . . . see
also vector bundle
C
canonical class . . . . . . 61
canonical classdegree . . 89
canonical divisor . . . . . 61
canonical line bun-
dle . . . . . . . 86, 106
canonical mapping . . 101
canonical section . . . . . 80
Cauchy-Riemann equa-
tions . . . . . . . . . 3
chain group . . . . . . . . 13
changes of coordinates . 18
closed form . . . . . . . . . 33
co-closed form . . . . . . . 33
co-exact form . . . . . . . 33
coboundary . . . . . 14, 44
coboundary operator . . 44
coboundarysplits . . . . . 44
cochain group . . . . . . . 44
cocycle . . . . . . . . . . . . 44
cohomology group . . . .
44, 47, 49, 56
cohomology groupexact
sequence . . . . 53, 56
cohomology groupmero-
morphic functions 96
cohomology groupof a
compact surface 66
cohomology groupof a
disc . . . . . . . 65, 86
cohomology groupof a
line bundle . . 78, 80
cohomology groupof a
vector bundle . . 82
cohomology groupof the
complex plane 65, 85
cohomology groupRie-
mann sphere . . . 66Index 207
cohomology groupsmooth
functions . . . . . 84
commutator subgroup . 14
compact mapping . . . . 12
compact surfacetriangu-
lation . . . . . . . . 27
complex atlas . . . . . . . 18
complex atlasequiva-
lence . . . . . . . . 18
complex atlasmaximal . 19
complex manifold . . . . 72
complex planeautomor-
phisms . . . . . . . 153
complex planecohomol-
ogy group . . . 65, 85
complex structure . . . . 18
conjugation . . . . . . . . 33
connecting homomor-
phism . . . . . . . . 52
continuous functionsheaf
of . . . . . . . . . . . 29
coordinate patch . . . . . 18
cotangent space . . 3637
countable basis of topol-
ogy . . . . . . . 7, 19
covering . . . . . . . . . . . . 7
covering map . . . . . . . . 7
covering transformation 10
coveringbranched . . . . 10
coveringuniversal . . . . . 9
cross ratio . . . . . 163, 164
curve . . . . . see also path
cycle . . . . . . . . . . . . . 14
cylinder . . . . . . . . . . 154
D
de Rhamcohomology . . 55
de Rhamgroup . . . . . . 55
de Rhamsequence . . . . 54
de Rhamtheorem . . . . 55
deck transformation . . 10
degree . . . . . . . . . 24, 60
degreeof a divisor . . . . 60
degreeof a line bundle . 81
determinant line bundle 88
dierential of a smooth
mapping . . . . . . 101
Dirichlet region . . 173, 174
divisor . . . . . . . . . 59, 62
divisor class group . . . 61
divisor of a meromorphic
form . . . . . . . . . 60
divisor of a meromorphic
function . . . . . . 59
divisorassociated line
bundle . . . . . . . 79
divisorassociated sheaf 61
divisorcanonical . . . . . 61
divisordegree . . . . . . . 60
divisoreective . . . . . . 59
divisorglobally gener-
ated . . . . . . . . . 196
divisorgroup of . . . . . . 59
divisorlinear equiva-
lence . . . . . . 60, 80
divisorof a meromorphic
form . . . . . . . . . 60
divisorof a meromorphic
function . . . . . . 59
divisorprincipal . . . . . . 61
divisorRiemann-Roch
theorem . . . . 69, 95
Dolbeaults lemma . . . 64
Dolbeaults lemmacom-
pact support case 62
dual bundle . . . . . . . . 75
E
eective divisor . . . . . . 59
elliptic Mobius transfor-
mation . . . 152, 158
embedding . 101, 104, 105
essential singularity . . . . 4
Euler-Poincare charac-
teristic . . . . . 15, 16
evenly covered neigh-
bourhood . . . . . . 7
exact form . . . . . . . . . 33
exact sequence . . . . . . 51
exact sequence in coho-
mology . . . . . 53, 56
exterior algebra . . . . . 32
exterior derivative . . . . 32
exterior product . . . . . 31
F
form . . . . . . . . . 30, 31, 37
formclosed . . . . . . . . . 33
formco-closed . . . . . . . 33208 Index
formco-exact . . . . . . . . 33
formconjugation . . . . . 33
formexact . . . . . . . . . . 33
formexterior derivative 32
formexterior product . . 31
formharmonic . . . . . . . 34
formholomorphic . . . . . 34
formmeromorphic . . . . 40
formperiods . . . . . . . 116
formresidue . . . . . . . . 41
formsupport . . . . . . . . 39
Fuchsian group . . . . . 170
fundamental domain . 172
fundamental do-
mainDirichlet re-
gion . . . . . 173, 174
fundamental group . 9, 13
Fundamental Theorem of
Algebra . . . . . . 26
G
Gap theorem . . . 107, 108
Gauss-Bonet theo-
rem . . . . . 169, 175
general linear group . 150
genus . . . . . . . . . . 12, 68
genusarithmetic . . . . . 68
genusarithmetic and
topological are
equal . . . . . . . . 91
genustopological . . . . . 12
geodesic . . . . . . . 167, 168
globally generated divi-
sor . . . . . . . . . . 196
group of divisors . . . . . 59
H
harmonic form . . . . . . 34
harmonic function 34, 128
harmonic functionbar-
rier . . . . . . . . . 140
harmonic functionHar-
nacks inequality 134
harmonic functionHar-
nacks principle . 134
harmonic functionMax-
imum Modulus
Principle . . . . . 132
harmonic function-
mean value prop-
erty . . . . . 131, 133
harmonic functionon a
Riemann surface 138
harmonic functionPer-
rons method . . 139
Harnacks inequality . 134
Harnacks principle . . 134
holomorphic form . . . . 34
holomorphic function 2, 22
holomorphic function-
local representa-
tion . . . . . . . . 3, 6
holomorphic func-
tionorder of a zero 3
holomorphic function-
power series . . . . 3
holomorphic functionsev-
eral variables . . 72
holomorphic functionuni-
form convergence 6
holomorphic mapping . 22
holomorphic mappingde-
gree . . . . . . . . . 24
holomorphic mappingex-
amples . . . . . . . 23
holomorphic mappinglo-
cal representation 23
holomorphic mappingRiemann-
Hurwitz relation 27
holomorphic section . . 76
holomorphic vector bun-
dle . . . . . . . . . . 72
homology group . . 14, 15
homology groupBetti
number . . . . . . 15
homotopy . . . . . . . . . . . 8
Hurwitz theorem . . . . 177
hyperbolic distance . . 163
hyperbolic distancecom-
pleteness . . . . . 169
hyperbolic distancein-
duced topology . 166
hyperbolic metric 163, 165
hyperbolic metricGauss-
Bonet theorem . 169
hyperbolic metric-
geodesic . . . . . . 168
hyperbolic metricisome-
tries . . . . . . . . . 166Index 209
hyperbolic Mobius trans-
formation . . . . . 152
hyperelliptic involution 114
hyperelliptic Riemann
surface . . . . . . .
102, 112, 113, 115
hyperelliptic Riemann
surfacehyperellip-
tic involution . . 114
I
Identity Principle . . 4, 23
integration of forms 3839
integration of 2-forms . 39
isolated singularity . . . . 4
isolated singularityessen-
tial singularity . . 4
isolated singularity-
pole . . . . . . . . 4, 5
isolated singularityre-
movable singular-
ity . . . . . . . . . 4, 5
isomorphic vector bun-
dles . . . . . . . . . 75
isomorphism of bundles 74
J
Jacobi inversion prob-
lem . . . . . . . . . 124
Jacobian variety . . . . 121
K
kernel sheaf . . . . . . . . 51
Kleinian group . . 158, 159
Kleinian groupfunda-
mental domain . 172
Kleinian groupregion of
discontinuity . . . 158
Koebe theorem . . . . . 141
L
Laplacian . . . . . . . . . . 34
Lerays theorem . . . . . 48
lift of a path . . . . . . . . . 8
line bundle . . . . . . 73, 88
line bundleample . . . . 106
line bundleassociated to a
divisor . . . . . . . 79
line bundlecanonical 86, 106
line bundlecanonical sec-
tion . . . . . . . . . 80
line bundlecohomology
group . . . . . . 78, 80
line bundledegree . . . . 81
line bundlemeromorphic
section . . . . . . . 78
line bundletrivial . . . . 83
line bundlevery ample 106
linearly equivalent divi-
sors . . . . . . . 60, 80
local coordinates . . . . . 18
local trivialization . . . . 73
locally path connected
space . . . . . . . . . 8
loop . . . . . . . . . . . . . . . 9
loxodromic Mobius
transformation . 152
M
manifold . . . . . . . . . . . . 7
manifoldcomplex . . . . . 72
manifoldsmooth . . . . . 28
manifoldsurface . . . . . . . 7
mappingdierential . . 101
maximal complex atlas 19
Maximum Modules The-
orem . . . . . . . . . 5
Maximum Modulus Prin-
ciple . . . . . . 23, 132
mean value prop-
erty . . . . . 131, 133
meromorphic form . . . 40
meromorphic formdivi-
sor . . . . . . . . . . 60
meromorphic function 5, 22
meromorphic functionco-
homology . . . . . 96
meromorphic functiondi-
visor . . . . . . . . 59
meromorphic functionex-
istence . . . . . 71, 79
meromorphic section 77, 78
meromorphic sectionexis-
tence . . . . . . . . 78
Montel theorem . . . . . . 7
morphism of bundles . . 74
multiplicity . . . . . . . . 110
Mobius transformation 26
Mobius transformation-
commuting . . . . 160210 Index
Mobius transformationel-
liptic . . . . 152, 158
Mobius transformation-
hyperbolic . . . . 152
Mobius transformation-
loxodromic . . . . 152
Mobius transformation-
parabolic . . . . . 152
N
Noether Gap theorem 108
normal family . . . . 6, 141
normalizerof a Fuchsian
group . . . . . . . . 176
normalizerof a subgroup 176
O
Open Mapping Theo-
rem . . . . . . . 5, 23
P
parabolic Mobius trans-
formation . . . . . 152
partition of unity . . . . 11
path . . . . . . . . . . . . . . . 8
path connected space . . 8
pathhomotopy . . . . . . . 8
pathlift . . . . . . . . . . . . . 8
pathloop . . . . . . . . . . . . 9
pathpiecewise smooth . 38
periods of a form . . . 116
Perrons method . . . . 139
Picard group . . . . . . . 77
piecewise smooth path 38
planar Riemann sur-
face . . . . . 145, 147
Poincare lemma . . 33, 54
Poincare-Koebe Uni-
formization theo-
rem . . . . . . . . . 147
Poisson kernel . . . 129, 131
pole . . . . . . . . . . 4, 5, 22
poleorder of . . . . . . . . . 5
poleresidue . . . . . . . . . . 5
power series . . . . . . . . . 3
presheaf . . . . . . . . . . . 42
presheafassociated sheaf 49
presheafrestriction homo-
morphisms . . . . 42
presheafstalk . . . . . . . 50
principal divisor . . . . . 61
projective general linear
group . . . . . . . . 150
projective space . . . . 100
projective special linear
group . . . . 150, 157
properly discontinuous
action . . . . . . . . 158
punctured plane . . . . 154
R
ramication number . . 24
rank of a vector bundle 72
rational function . . . . . 26
Reciprocity theorem . 120
renement mapping . . 45
region of discontinuity 158
removable singularity 4, 5
Removable Singularity
Theorem . . . . . . 5
residue . . . . . . . . . . 5, 41
Residues theorem . . . . 41
restriction homomor-
phisms of a
presheaf . . . . . . 42
Riemann mapping theo-
rem . . . . . . . . . 141
Riemann sphere . . 19, 71
Riemann sphereunique
complex structure 71
Riemann surface . . 18, 72
Riemann surfaceabelian
fundamental
group . . . . . . . . 161
Riemann surface-
branched covering 21
Riemann surfacecomplex
plane . . . . . . . . 19
Riemann surfacecomplex
structure . . . . . 18
Riemann surfacecount-
able basis of topol-
ogy . . . . . . . . . 19
Riemann surfacecylin-
der . . . . . . . . . . 154
Riemann surfaceexam-
ples . . . . . . . 1922
Riemann surfaceGauss-
Bonet theorem . 175Index 211
Riemann surfacehar-
monic function . 138
Riemann surfaceholo-
morphic mapping 22
Riemann surfaceHurwitz
theorem . . . . . . 177
Riemann surfacehyperel-
liptic . . . . . . . .
102, 112, 113, 115
Riemann surfacelocal co-
ordinates . . . . . 18
Riemann surfacepla-
nar . . . . . . 145, 147
Riemann surfacepunc-
tured plane . . . . 154
Riemann surfaceRiemann-
Roch theorem . . 87
Riemann surfaceSerre
Duality theo-
rem . . . . . . . 94, 95
Riemann surfacesubhar-
monic function . 138
Riemann surfacetorus 154
Riemann surfaceunit
disc . . . . . . . 4, 19
Riemann surfaceupper
half plane . . . . . 19
Riemanns Bilinear Rela-
tions . . . . . . . . 119
Riemann-Hurwitz rela-
tion . . . . . . . . . 27
Riemann-Roch theo-
rem . . . . 69, 87, 95
rotation . . . . . . . . . . 155
S
Sards theorem . . . . . . 40
Schwartzs theorem . . . 68
Schwarz lemma . . . 4, 155
section . . . . . . . . . . . . 76
sectioncanonical . . . . . 80
sectionholomorphic . . . 76
sectionmeromorphic . . 77
Serre Duality theo-
rem . . . . . . . 94, 95
sheaf . . . . . . . . . . . . . 43
sheafassociated to a divi-
sor . . . . . . . . . . 61
sheafassociated to a
presheaf . . . . . . 49
sheafcoboundary opera-
tor . . . . . . . . . . 44
sheafcochain group . . . 44
sheafcohomology group
44, 47, 49, 56
sheafconnecting homo-
morphism . . . . . 52
sheafde Rhamsequence 54
sheafexact sequence . . . 51
sheafhomomorphism . . 50
sheafkernel sheaf . . . . . 51
sheafLerays theorem . . 48
sheafmeromorphic func-
tionscohomology 96
sheafof continuous func-
tions . . . . . . . . 29
sheafof smooth func-
tions . . . 29, 47, 84
sheafof 1-forms . . . . . . 30
sheafof 2-forms . . . . . . 31
sheafskyscraper . . . . . . 68
sheafstalk . . . . . . . . . . 50
Shwartzs theorem . . . . 12
side derivative . . . . . . . 38
simplex . . . . . . . . . . . 13
simply connected space . 9
singularity . . . see also
isolated singularity
skyscraper sheaf . . . . . 68
smooth function . . . . . 29
smooth functiondieren-
tial . . . . . . . . 32, 33
smooth functionexterior
derivative . . . . . 32
smooth functionsheaf
of . . . . . . . . . 29, 47
smooth manifold . . 28, 40
smooth manifoldclassi-
cation . . . . . . . . 40
smooth mappingdieren-
tial . . . . . . . . . . 101
special linear group . . 150
stalk . . . . . . . . . . . . . 50
Stokes theorem . . . . . 39
subharmonic func-
tion . . . . . 135, 138
subharmonic functionon
a Riemann sur-
face . . . . . . . . . 138212 Index
subharmonic function-
Perrons method 139
support of a 2-form . . . 39
surface . . . . . . . . . . . . . 7
surfacefundamental
group . . . . . . . . 13
surfacehomology
group . . . . . . 14, 15
T
tangent space . . . . . . . 37
tensor product . . . . . . 75
torus . . . . . . 20, 124, 154
torusWeierstrass -
function . . . . . . 97
total branching number 27
transition functions . . . 74
triangulation . . . . 15, 27
trivial bundle . . . . 75, 76
U
Uniformization theo-
rem . 147, 150, 158
unit disc . . . . . . . . . . . . 4
unit discautomorphisms 155
universal covering space 9
upper half planeautomor-
phisms . . . . . . . 156
V
vector bundle . . . . . . . 72
vector bundle isomor-
phism . . . . . . . . 74
vector bundle morphism 74
vector bundledual . . . . 75
vector bundleisomorphic 75
vector bundleisomor-
phism . . . . . . . . 74
vector bundleline bundle 73
vector bundlelineample 106
vector bundlelinevery
ample . . . . . . . . 106
vector bundlelocal trivi-
alization . . . . . . 73
vector bundlemeromor-
phic section . . . 77
vector bundlemorphism 74
vector bundlePicard
group . . . . . . . . 77
vector bundlerank . . . . 72
vector bundlesection . . 76
vector bundletensor
product . . . . . . 75
vector bundletransition
functions . . . . . 74
vector bundletriv-
ial . . . . . 73, 75, 76
very ample line bundle 106
W
Weierstrass Gap theo-
rem . . . . . . . . . 107
Weierstrass point 108,
109, 112, 113, 115
Weierstrass theorem . . . 6
Weierstrass -function . 97
Wronskian . . . . . . . . 109
List of Figures
1 Homotopy between two paths. 9
2 A surface of genus 2 with generators of the fundamental group. 13
3 Allowed and wrong triangulations. 16
4 Condition (3) in denition 1.2.7. 16
5 A triangulation of the sphere. 17
6 Branched covering. 21
7 Subdivision of triangles (z
0
is a critical value). 27
8 Proof of theorem 1.4.26. 42
9 Proposition 1.5.12. 48
10 Polygon corresponding to a surface of genus 2. 116
11 Proof of lemma 2.9.2. 118
12 h
1
(t). 143
13 Triangle with two zero angles. 170
14 Gauss-Bonet. 170
15 Hyperbolic pentagon. 199
213

You might also like