0% found this document useful (0 votes)
145 views137 pages

Advanced Complex Analysis Course

This document outlines the contents of a course on complex analysis on Riemann surfaces. It begins with an introduction that defines Riemann surfaces and provides examples, including the Riemann sphere, Euclidean plane, and hyperbolic plane. It also discusses the relations of the topic to fields like topology, geometry, algebra, number theory, and dynamics. The introduction then defines the key concepts of the course, including Riemann surfaces, the automorphism groups of different geometries on the Riemann sphere, and computing distances on surfaces.

Uploaded by

abdullah ghamdi
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
145 views137 pages

Advanced Complex Analysis Course

This document outlines the contents of a course on complex analysis on Riemann surfaces. It begins with an introduction that defines Riemann surfaces and provides examples, including the Riemann sphere, Euclidean plane, and hyperbolic plane. It also discusses the relations of the topic to fields like topology, geometry, algebra, number theory, and dynamics. The introduction then defines the key concepts of the course, including Riemann surfaces, the automorphism groups of different geometries on the Riemann sphere, and computing distances on surfaces.

Uploaded by

abdullah ghamdi
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 137

Complex Analysis on Riemann Surfaces

Math 213b — Harvard University


C. McMullen

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2 Maps between Riemann surfaces . . . . . . . . . . . . . . . . 9
3 Sheaves and analytic continuation . . . . . . . . . . . . . . . 22
4 Algebraic functions . . . . . . . . . . . . . . . . . . . . . . . . 29
5 Holomorphic and harmonic forms . . . . . . . . . . . . . . . . 40
6 Cohomology of sheaves . . . . . . . . . . . . . . . . . . . . . . 54
7 Cohomology on a Riemann surface . . . . . . . . . . . . . . . 64
8 Riemann-Roch . . . . . . . . . . . . . . . . . . . . . . . . . . 69
9 The Mittag–Leffler problems . . . . . . . . . . . . . . . . . . 76
10 Serre duality . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
11 Maps to projective space . . . . . . . . . . . . . . . . . . . . . 84
12 The canonical map . . . . . . . . . . . . . . . . . . . . . . . . 92
13 Line bundles . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
14 Curves and their Jacobians . . . . . . . . . . . . . . . . . . . 109
15 Hyperbolic geometry . . . . . . . . . . . . . . . . . . . . . . . 125
16 Uniformization . . . . . . . . . . . . . . . . . . . . . . . . . . 135

1 Introduction
Scope of relations to other fields.

1. Topology: Rgenus, manifolds. Algebraic topology, intersection form on


H 1 (X, Z), α ∧ β.

2. 3-manifolds. (a) Knot theory of singularities. (b) Isometries of H3 and


b (c) Deformations of M 3 and ∂M 3 .
Aut C.

3. 4-manifolds. (M, ω) studied by introducing J and then pseudo-holomorphic


curves.

4. Differential geometry: every Riemann surface carries a conformal met-


ric of constant curvature. Einstein metrics, uniformization in higher
dimensions. String theory.

1
5. Complex geometry: Sheaf theory; several complex variables; Hodge
theory.

6. Algebraic geometry: compact Riemann surfaces are the same as alge-


braic curves. Intrinsic point of view: x2 +y 2 = 1, x = 1, y 2 = x2 (x+1)
are all ‘the same’ curve. Moduli of curves. π1 (Mg ) is the mapping
class group.

7. Arithmetic geometry: Genus g ≥ 2 implies X(Q) is finite. Other


extreme: solutions of polynomials; C is an algebraically closed field.

8. Lie groups and homogeneous spaces. We can write X = H/Γ, and


M1 = H/ SL2 (Z). The Jacobian Jac(X) ∼ = Cg /Λ determines an ele-
ment of Hg / Sp2g (Z): arithmetic quotients of bounded domains.

9. Number theory. Automorphic forms P n2and theta functions. Let q =


exp(πiz) P
on H. Then f (q) = q is an automorphic form and
k
f (q) = n
an (k)q where an (k) is the number of ways to represent
n as a sum of k ordered squares.

10. Dynamics. Unimodal maps exceedingly rich, can be studied by com-


plexification: Mandelbrot set, Feigenbaum constant, etc. Billiards can
be studied via Riemann surfaces. The geodesic flow on a hyperbolic
surface is an excellent concrete example of a chaotic (ergodic, mixing)
dynamical system.

Definition and examples of Riemann surfaces. A Riemann surface is


a connected, Hausdorff topological space X equipped with an open covering
Ui and a collection of homeomorphisms fi : Ui → C such that there exist
analytic maps gij satisfying

fi = gij ◦ fj

on Uij = Ui ∩ Uj .
We put the definition into this form because it suggest that (gij ) is a
sort of 1-coboundary of the 0-chain (fi ). This equation will recur in the
definition of sheaf cohomology.
The Riemann sphere. The simplest compact Riemann surface is C b =
C∪∞ with charts U1 = C and U2 = C−{0}
b with f1 (z) = z and f2 (z) = 1/z.
b ∼
Alternatively, C = P is the space of lines in C2 :
1

P1 = (C2 − 0)/C∗ .

2
The isomorphism is given by [Z0 : Z1 ] 7→ z = Z1 /Z0 .
Note that we have many natural (inverse) charts C → P1 given by f (t) =
[ct + d, at + b]. The image omits a/c. Transitions between these charts are
given by Möbius transformations.
In fact the full automorphism group of Cb is the quotient of G = SL2 (C)
by ±I. This means every automorphism lifts to a linear map on C2 of deter-
minant 1, well-defined up to sign. One can also regard the full automorphism
group as PGL2 (C) = GL2 (C)/C∗ .
Other types of geometry on the Riemann sphere and its homogeneous
subdomains correspond to subgroups H ⊂ G as below.
What is a circle? Recall the classification of Hermitian forms, i.e. non-
degenerate bilinear forms on Cn satisfying z · w = w · z. Any such form
is equivalent, by an element of GLn (C), to the standard form of signature
(p, q), given by
z · z = |z1 |2 + · · · + |zp |2 − |zp+1 |2 − · · · − |zp+q |2
In the case of C2 , a form of signature (1, 1) (an indefinite form) has a light
cone of vectors with z · z = 0. The lines in this light cone give a circle on
P1 , and vice-versa.
The positive vectors in C2 pick out a distinguished component of the
complement of a circle. Consequently the space of oriented circles on C b is
isomorphic to C = SL2 (C)/ SU(1, 1).
The space C 0 = C/(Z/2) of unoriented circles can be thought of as the
complement of a ball B ⊂ RP3 , with the boundary of the ball identified
with Cb ∼
= S 2 , and the horizon as seen from p 6∈ B the circle attached to p.
We have π1 (C 0 ) ∼
= Z/2, and its double cover is the space of oriented circles.
Thus SL2 (C)/ SU(1, 1) is simply–connected.
The round sphere. The round metric on C b is given by 2|dz|/(1 + |z|2 )
and preserved exactly by
( ! )
a b
SU(2) = : |a| + |b| = 1 ∼
2 2
= S3.
−b a

Alternatively, if we represent points zi in P1 by unit vectors vi in C2 , then


the spherical distance satisfies:
cos2 (d(z1 , z2 )/2) = |hv1 , v2 i|2 .
√ √
For example, d(1, eiθ ) = θ, and if we set v1 = (1, 1)/ 2, v2 = (1, eiθ )/ 2,
then
|hv1 , v2 i|2 = |1 + eiθ |2 /4 = (1 + cos θ)/2 = cos2 (θ/2).

3
The factor of 1/2 is not unexpected: for real projective space, the map

S 1 ⊂ R2 − {0} → RP1 = S 1 /(±1)

doubles angles as well.


Alternatively, we can compute that in the round metric
Z r
2 ds
d(0, r) = 2
= 2 tan−1 (r),
0 1 + s

so r = tan(d/2). (This is the familiar transformation from calculus that


turns any rational function of sin(d) and cos(d) into the
√ integral of an ordi-
nary rational function.) Then v1 = (1, 0), v2 = (1, r)/ 1 + r2 , and hence
1 1
hv1 , v2 i2 = = = cos2 (d/2),
1+r 2 1 + tan2 (d/2)

using the identity 1 + tan2 θ = sec2 θ.


The Euclidean plane. The other two simply-connected Riemann surfaces
are C and H, with automorphism groups B = AN and SL2 (R). These are
simply their stabilizers in SL2 (C).
Note that B is the same as the stabilizer of ∞ ∈ C, b∼
b and hence C = G/B;
B is called a Borel subgroup because the associated homogeneous space is a
projective variety.
The hyperbolic plane. Note that SL2 (R) preserves not only the com-
b but also its orientation. (Note that z 7→ 1/z is not in
pactified real axis R,
SL2 (R.)
Miraculously, the full automorphism of H preserves the hyperbolic metric
|dz|/y. On the unit disk ∆ ∼ = H the automorphism group becomes
( ! )
a b
SU(1, 1) = : |a|2 − |b|2 = 1 ,
b a

and the metric becomes 2|dz|/(1 − |z|2 ).


As a 3-manifold, SU(1, 1) ∼= SL2 (R) is homeomorphic to a solid torus,
since it double covers the unit tangent bundle of ∆. Thus SL2 (R) retracts
onto SU(2) (as is well-known) but not onto SU (1, 1), since they have different
homotopy types. The reason is that the Gram–Schmidt process fails for
indefinite metrics, because a null vector cannot be normalized.
Hyperbolic distance in ∆. The unit disk ∆ can be interpreted as the
space of positive lines in C2 for the Hermitian form kZk2 = |Z0 |2 − |Z1 |2 .

4
As in the case of the sphere, we can use the underlying inner product to
compute distances in the hyperbolic metric on the disk. We find:

cosh2 (d(z1 , z2 )/2) = |hv1 , v2 i|2 .

Here it is useful to recall that t 7→ (cosh t, sinh t) sweeps out (one sheet of)
the unit ‘circle’ in R1,1 defined by x2 − y 2 = 1, just as (cos t, sin t) sweeps
out the unit circle in R2 . R
r
For example, d(0, r) = 0 2 dx/(1 − x2 ) = 2 tanh−1 (r), and hence

cosh2 (d(0, r)/2) = (1 − r2 )−1 ,

which agrees with the√inner product squared between the unit vectors v0 =
(1, 0) and vr = (1, r)/ 1 − r2 .
Aside: H as the moduli space of triangles. We can think of H as the
space of marked triangles (with ordered vertices) up to similarity (allowing
reversal of orientation). The triangle attached to τ ∈ H is the one with
vertices (0, 1, τ ).
The ‘modular group’ S3 acts by reflections, in the line x = 1/2 and the
circles of radius 1 centered at 0 and 1. It is generated by τ 7→ 1/τ and
τ 7→ 1 − τ . Note that the fixed loci of these maps are geodesics, and that
the special point where they come together corresponds to an equilateral
triangle, and that each line corresponds to isosceles triangles.
One can also consider barycentric subdivision. One of the 6 subdivided
triangles is given by replacing τ with the barycenter, i.e. β(τ ) = (1 + τ )/3.
As an exercise, one can show that S3 and β generate a dense subgroup
of Isom H. Thus any triangle can be approximated by one in Tn , the nth
barycentric subdivision of an equilateral triangle. A more subtle exercise is
to show that almost all of the triangles in Tn are long and thin.
Aside: the Schwarz lemma and dynamics. A fundamental property
of the hyperbolic metric: if f : H → H is holomorphic, then either f is an
isometry, or kf 0 (z)k < 1 for every z ∈ H, where the norm is measured in the
hyperbolic metric.
Sample use: if f (z) = z 2 + c and p is an attracting periodic point for f ,
then p attracts z = 0. (The immediate basin B of p is a bounded disk, by
the maximum principle; and this disk must contain a critical point of f n ,
where n is the period of p.)
Sample use: let p(z) be a polynomial. If Newton’s method works for the
point of inflection of p, then it works for almost every point on the sphere.
Idea of the proof: the critical points of Newton’s method N (z) = z −
p(z)/p0 (z) come exactly from the zeros of p(z)p00 (z).

5
Description of all Riemann surfaces by covering spaces. Once all
the universal covering spaces have been identified, we can easily see:

Theorem 1.1 Every Riemann surface X is isomorphic to


b C, C∗ , C/Λ or H/Γ,
C,

for some lattice Λ ⊂ C or torsion–free discrete subgroup Γ ⊂ SL2 (R).

The torsion–free condition insures that Γ acts freely on H.


Warning. It is not always true that the quotient of a planar domain by a
group acting freely with discrete orbits is a Hausdorff surface! (Consider the
acting (x, y) 7→ (2x, y/2) acting on R2 − {0, 0}.) However: if Γ is a discrete
group of isometries acting freely on a connected Riemannian manifold Y ,
then the action is properly discontinuous and Y covers the quotient manifold
Y /Γ.
Doubly-connected Riemann surfaces. Applying the result above we
can easily show that the Riemann surfaces with π1 (X) ∼ = Z are given by C∗ ,
∆∗ and
A(r) = {z : r < |z| < 1},
up to isomorphism.
In terms of covering spaces, we have C∗ ∼
= C/Z, ∆∗ ∼
= H/Z, and

A(r) ∼
= H/λZ ,

where λ > 1 and r = exp(−2π 2 / log λ).


The covering map f : H → A(r) can be factored as z 7→ log(z), z 7→ αz
and then z 7→ ez , where α = 2πi/ log λ. In other words, f (z) = z α with α
purely imaginary.
The fact that f (z) = z α = exp(iβ log z) can be guessed from the fact
that f should send R+ to the unit circle. For this map to have the right
deck group, we need

f (λ) = 1 = exp(iβ log λ) = exp(2πi),

which gives β = 2π/L, where L = log λ is the length of the closed hyperbolic
geodesic on A(r). Then r = f (−1) = exp(iβπi) = exp(−2π 2 /L). Thus
r → 0 as L → ∞.
Triply–connected Riemann surfaces, etc. Any finitely–connected pla-
nar surface is equivalent to C
b with a finite number of points and round disks
removed. In particular, such a surface depends on only finitely many moduli.

6
The double of a surface with g+1 circular boundaries is a closed Riemann
surface of genus g, with a real symmetry. In fact the space of such planar
regions is isomorphic to the space of Riemann surfaces defined over R with
the maximal number of real components.
Another canonical form for a planar region, say containing a neighbor-
hood of infinity, is the complement of finitely many disjoint horizontal seg-
ments. This canonical conformal representation can be found by maximizing
Re b1 , where f : U → C is given by f (z) = z + b1 /z + O(1/z 2 ). See e.g.
[Gol].
The triply–punctured sphere. The group Γ(2) ⊂ SL2 (Z); isometric for
hyperbolic metric; quotient is triply-punctured sphere. The map

λ:H→C
b − {0, 1, ∞}

can be interpreted as the elliptic modular function. That is, given τ ∈ H


we set Λ = Z ⊕ Zτ and form E = C/Λ; then E admits a degree two map
b branched over 0, 1, ∞ and λ(τ ). The ordering of the 4 branch points
to C
requires a marking of E[2], which explains that passage to Γ(2).
More constructions of Riemann surfaces.

1. Any connected, oriented Riemannian 2-manifold (X, g) has a unique


compatible complex structure, such that g is a conformal metric. (This
result requires isothermal coordinates or harmonic functions.)

2. Algebraic curves. A curve C ⊂ C2 defined by f (x, y) = 0 is smooth


df 6= 0 along C. Example: f (x, y) = y 2 − p(x) is smooth iff p has no
multiple roots. Non-example: y 2 = x3 , y 2 = x2 (x + 1). Double zeros
give nodes, triple zeros give cusps.

3. Polygons in C, C,
b H, glued together by isometries, naturally form
Riemannian surfaces. Polyhedra in R3 , such as the octahedron and
the icosahedron. Triangulated surfaces. (Relation to billiards.)

4. If X is a Riemann surface and π : Y → X is a covering map, with Y


connected, then there is a unique complex structure on Y such that π
is holomorphic and Y is a Riemann surface in its own right.

5. Finite branched coverings. Let X be a compact Riemann surface and


E ⊂ X a finite set, and let X ∗ = X − E and ρ : π1 (X ∗ ) → G is a map
to a finite group. Then the corresponding G-covering space Y ∗ → X ∗
completes to a give a holomorphic map π : Y → X, branched over E.

7
6. Finite quotients. If G ⊂ Aut(Y ) is a finite group of automorphisms,
then X = Y /G is a topological surface, and X carries a unique complex
structure such that π : Y → X = Y /G is holomorphic.
The Weierstrass ℘-function gives the map E 7→ E/G ∼ =C b where G =
Z/2 acts by z 7→ −z in the group law on an elliptic curve E.

The octagon, elaborated. To illustrate connections between these con-


structions more fully, we give 3 different descriptions of the same Riemann
surface.

1. Let Z be the regular octahedron, and B ⊂ Z its 6 vertices. Then Z has


the structure of a Riemann surface, and we can take the canonical 2-
fold covering X → Z branched over B. The result is a surface of genus
two that can be assembled out of 16 equilateral triangles, meeting 8
to a vertex.

2. Let X be the curve defined by y 2 = x(x4 − 1). This Riemann surface


regarded as the degree 2 cover of P1 branched over the 6 points
can be√
4
(0, ∞, 1).

3. Let X be a regular Euclidean octagon with opposite sides identified.

To see the first two surfaces are the same, we use the uniformization theorem
to see Z ∼
= C.b Now Z has symmetry group S4 , and the obvious rotation of
order 4 means we can take B equal to 0, ∞ and the 4th roots of unity. Then
the uniqueness of branched coverings shows the result is isomorphic to the
algebraic curve y 2 = x(x4 − 1).
This isomorphism reveals that Aut(X) has a subgroup G of order 48,
the ‘double cover’ of S4 , isomorphic to the preimage of S4 in SU(2). In fact
y 7→ −y gives the center Z/2 of√G. In particular, X has an automorphism of
order 8 given by (x, y) 7→ (ix, iy). (We will eventually prove, in the study
of hyperelliptic Riemann surfaces, that G = Aut(X).)
The regular octagon gives the same surface of genus two. Indeed, the
regular octagon P also has an automorphism of order 8, which gives rise
to an automorphism r : X → X of order 8 since the gluing instructions
are respected. Note that r has 6 fixed points: 1 coming from the center of
the octagon, 1 coming from the 8 vertices (which are all identified), and 4
coming from the edge midpoints (which are identified in pairs).
We can construct √ an explicit map of degree 2 from X = P/ ∼ to C b
4
branched over (0, 1, 1). More precisely, we will construct an isomorphism
of Y = X/hr4 i with C b such that X → Y ∼ =C b gives the desired branched

8
cover. Note that its six fixed points (the Weierstrass points) correspond to
the vertices and edge-midpoints of the octagon, and its center.
To describe Y , first observe that the quotient of the octagon by hr4 i
is a square Q. The edge identification on the square fold each edge at its
midpoint, identifying the two halves, and Y ∼ = Q/ ∼.
Topologically, the quotient of the boundary of the square by these iden-
tifications gives a plus sign. To make an isomorphism Y ∼ b map Q
= C,
conformally to complement in C b of the line segments from the 4th roots of
unity to the origin, sending the center of the square to infinity, and sending
the midpoint of each edge of the square to one of the 4th roots of unity.
This map folds the edges of the square and identifies them. Because of its
folding, this map respects the identifications on the sides of the square and
the octagon, and gives an explicit degree two map of the octagon surface to
b branched over the vertices of a octahedron.
C,

2 Maps between Riemann surfaces


The arrows in the category of Riemann surfaces are as important as the
objects themselves. In this section we discuss holomorphic maps f : X → Y
between Riemann surfaces. Particular types of maps with good features that
we will discuss are proper maps and branched coverings, especially regular
(Galois) branched coverings.
As we will explain, every compact Riemann surface X can be presented
as a branched covering of C,b and if X is defined over a number field then
we can arrange that the covering is branched over just 3 points.
Local analysis and multiplicity. Let f : X → Y be holomorphic, with
f (p) = q. Then we can find charts (U, p) ∼ = (∆, 0) and (V, q) ∼
= (∆, 0) so
that f goes over to an analytic map
F : (∆, 0) → (∆, 0).
Theorem 2.1 The charts can be chosen so that F (z) = 0 or F (z) = z d ,
d > 0.
Proof. First choose arbitrary charts, and assume F (z) 6= 0; then F (z) =
z d g(z) where g(0) 6= 0. The function g(z)1/d can therefore be defined for
|z| sufficiently small, yielding a factorization f (z) = h(z)d defined near 0,
where h0 (0) 6= 0. Since h is invertible near 0, it can be absorbed into the
choice of chart; then after suitably rescaling in domain and range, we obtain
F (z) = z d on ∆.

9
We let mult(f, p) = d when F (z) = z d , and mult(f, p) = ∞ when F (z) is
constant. Note that the multiplicity can be described topologically: in the
first case, f is locally d-to-1 near (but not at) p, while in the second case f
is locally constant.
By the description above, the set where mult(f, p) = ∞ is open and the
set where mult(f, p) is finite is open. Since X is connected, one of these
must be empty. This shows:

If f is not globally constant, then multp (f ) < ∞ for all p.

In particular, one f is constant on a nontrivial open set, it is globally con-


stant. (This is intuitively reasonable by analytic continuation.)

From now on in this section we assume f is not constant.

Critical points and branching. Let C(f ) = {p ∈ X : mult(f, p) > 1}


denote the discrete set of critical point of f . The branch locus B(f ) ⊂ Y
is its image, B(f ) = f (C(f )). The branch locus obeys the useful cocycle
formula:
B(f ◦ g) = B(f ) ∪ f (B(g)).
Here are some general properties of nonconstant analytic maps.

1. The map f is open and discrete. That is, when U is open so is f (U ),


and f −1 (q) is a discrete subset of X for all q ∈ Y .

2. If X is compact, then Y is compact and f (X) = Y , since f (X) is both


open and closed. Moreover the fibers of f are finite and C(f ) and
B(f ) are finite.

3. An analytic function f : X → C on a compact Riemann surface is


constant.

4. To prove C is algebraically closed, observe that a polynomial with no


zero would define an analytic map f : Cb →Y =C b − {0}, and Y is not
compact.

Proper maps. While compactness of X and Y are useful assumptions,


many results follow by imposing a compactness property on f . We say
f : X → Y is proper of f −1 sends compact sets to compact sets; that is,
whenever K ⊂ Y is compact, so is f −1 (K).
Assume f is analytic and proper. Then:

10
1. f is closed: i.e. E closed implies f (E) closed.

2. f is surjective. (Since f (X) is open and closed.)

3. If D ⊂ X is discrete, so is f (D).

4. In particular, the branch locus B(f ) ⊂ Y is discrete.

5. f −1 (q) is finite for all q ∈ Y . (Since f −1 (q) is compact and discrete.)

6. For any neighborhood U of f −1 (q) there exists a neighborhood V of q


whose preimage is contained in U . (Since f (X − U ) is closed and does
not contain q.)

7. If f is a proper local homeomorphism, then it is a covering map.


Indeed, a local homeomorphism is discrete, so given q ∈ Y we can
choose neighborhoods Ui of its preimages p1 , . . . , pn such that fi :
Ui → Vi is a homeomorphism, where fi = f |Ui . Let V be a disk
−1 (V ) is contained in
T
neighborhood of q such that V ⊂ f (Ui ) and f
−1
S
Ui . Then it is easy to check that f (V ) has a components Wi ⊂ Ui
and fi : Wi → V is a covering map.

8. For any q ∈ Y with f −1 (q) = {pS1 , . . . , pn }, there exists a disk V


containing q such that f −1 (V ) = Ui with Ui a disk, pi ∈ Ui and
fi = f |Ui satisfies fi : (Ui , pi ) → (V, q) is conjugate to z 7→ z di on ∆.
(Here one can use the Riemann mapping theorem to prove that if
g : ∆ → ∆ is given by g(z) = z m , and V ⊂ ∆ is a simply-connected
neighborhood of z = 0, then the pullback of g to V is also conjugate
to z 7→ z m .)
P
9. The function d(q) = f (p)=q m(f, p) is independent of q. It is called
the degree of f . In fact, the condition that d(q) is finite and constant
is equivalent to the condition that f is proper.

Finite branched coverings. Let f : X → Y be a proper map of degree


d = deg(f ). Let Y ∗ = Y − B(f ) and X ∗ = f −1 (Y ∗ ). Then:

f : X∗ → Y ∗

is a proper local homeomorphism, hence a covering map.

11
Conversely, one can easily show: given a discrete set B ⊂ Y and a finite
covering space f : X ∗ → Y ∗ = Y − B, there is a unique way to extend X ∗
and f to obtain a proper map

f : X → Y.

As a basic example, if Y = ∆, B = {0}, then Y ∗ = ∆∗ with π1 (Y ∗ ) ∼ = Z.


Thus a covering space is determined by its degree d. But fd (z) = z d : ∆∗ →
∆∗ gives an explicitly model for X ∗ /Y ∗ , and it clearly extends unique to a
proper map on ∆.
The general case follows from this example. Here is a related result on
filling in punctures. Let X ∗ = X − D be the complement of a discrete set
on a topological surface X. Suppose X ∗ is given the structure of a Riemann
surface. Then:

Theorem 2.2 The complex structure extends from X ∗ to X iff every point
x ∈ D has a neighborhood U such that U ∩ X ∗ ∼
= ∆∗ .

Here we use the removability of isolated singularities for bounded func-


tions. As an example, this result applies to the one-point compactification
of C; but not to the one-point compactification of C∗ . (For a fancy proof,
observe that if X = ∆ ∪ {∞} were a compact Riemann surface then it would
have to be isomorphic to C,
b and hence the complement of a point in C b would
be isomorphic to ∆, contradicting Liouville’s theorem.
Regular branched covers. We define the Galois group of a branched
covering (= proper) map f : X → Y by

Gal(X/Y ) = {g ∈ Aut(X) : f ◦ g = f }.

If | Gal(X/Y )| = deg(f ), we say X/Y is a regular (or Galois) branched cover.


Conversely, if G ⊂ Aut(X) is a finite group, the quotient Y = X/G can
be given the structure of a Riemann surface is a unique way such that the
quotient map
f : X → Y = X/G
is a proper holomorphic map. We have

C(f ) = {x ∈ X : |Gx | > 1},

where Gx denotes the stabilizer of x.


Examples.

12
1. Any analytic map between compact Riemann surfaces is proper.

2. Thus every rational map f : Cb →C b is proper, with B(f ) = {0, ∞}.


The map f (z) = z is regular, with G ∼
d
= Z/d generated by z 7→ ζd z.

3. The rational map f (z) = z + 1/z is regular with Galois group Z/2
generated by z 7→ 1/z and B(f ) = {−2, 2}. If we restrict this to a
map
f : C∗ → C,
we see the covering map from an infinite cylinder to the infinite pil-
lowcase.

4. The rational map f (z) = z n + 1/z n has the same branch locus and
Galois group the dihedral group D2n .

5. There are similar rational maps for A4 , S4 and A5 of degrees 12, 24


and 60. See Klein’s Lectures on the Icosahedron.

6. Any finite covering map between Riemann surfaces is proper.

7. For a complex torus E = C/Λ, we obtain a finite covering map fα :


E → E for every element of

End(E) = {α ∈ C∗ : αΛ ⊂ Λ}.

We have√ deg(fα ) = |α|2 . Note that many such α exist when e.g.
Λ = Z[ −d] is a ring (more generally, when Λ ⊗ Q is a quadratic
field.)

8. The Weierstrass ℘ function gives a regular proper map f : E → C


b of
degree 2 with |B(f )| = 4.

9. Let E = C/Z[i] and let G = Z/4 acting by rotations. Then E/G ∼ =Cb
and the natural map f : E/G → C has degree 4 and |B(f )| = 3.
b
Indeed, C(f ) consists of the 4 points represented by the center, vertex,
and edge midpoints of the square, and the two types of edge midpoints
are identified by the action of G.
The quotient can, if desired, be constructed as the composition of a
degree two map ℘ : E → C b branched over 0, ∞, ± − 1, with s(z) = z 2 ,
resulting in B(f ) = {0, 1, ∞} with f = s◦℘. This is a good illustration
of the principle
B(s ◦ ℘) = B(s) ∪ s(B(℘)).

13
Geometrically, E/G is the double of an isosceles right triangle. (It is
called the (2, 2, 4)–orbifold.)

10. An entire function is proper iff it is a polynomial.

11. Any proper map of the disk to itself is given by a Blaschke product,
d
Y z − ai
f (z) = exp(iθ) ·
1 − ai z
1

(Proof: f has have a least one zero. Taking a quotient with a Blaschke
product, the result is either a constant of modulus one or a proper map
g : ∆ → ∆ with no zeros. The latter is impossible.)

-2 -1 1 2

-2 2

Figure 1. Topology of a cubic polynomial.

12. A basic example of an irregular branched cover is given by f : C → C


the cubic polynomial
f (z) = z 3 − 3z.
(See Figure 1). This map has critical points C(f ) = {−1, 1} and
branch values B(f ) = {±2}; we have B e = f −1 (B) = {±1, ±2}, and
e → C−B is an irregular degree three cover. It corresponds to
f : C− B
the abbaab triple cover of a bouquet of circles with fundamental group
F2 = ha, bi.
We remark that every cubic polynomial p(z) is equivalent f (z) or
c(z) = z 3 , in the sense that it is equal to f or c after an affine change
of coordinates in domain and range.

13. The map f (z) = z 4 : H → C∗ is not proper, even though it is a local


homeomorphism. Indeed, the fibers of f have cardinality 2 for most
points in C∗ , but the fiber over z = 1 consists of the single point i with
multiplicity 1. Thus the ‘degree’ of f is not constant.

14
14. Let B ⊂ C b be a finite set with |B| = 2n, and let Y ∗ = C b − B.
∗ ∼
Then π1 (Y ) = F2n−1 , and there is a unique map to Z/2 which sends
each peripheral loop to 1. The resulting 2-fold covering space can
be completed to give a compact Riemann surface f : X → C b with
B(f ) = E, deg(f ) = 2 and the genus of X is n − 1. If B coincides
with the zero set of a polynomial p(x) of degree 2n, then X is nothing
more than the (completion of) the algebraic curve y 2 = p(x).

Infinite branched coverings. Another class of well–behaved maps be-


tween Riemann surfaces are the covering maps, possibly of infinite degree.
We would like to allow these maps to be branched as well. To this end, we
say f : X → Y is a branched covering if

1. Each q ∈ Y has a neighborhood V such that fi : Ui → V is proper,


when Ui are the components of f −1 (V ) and fi = f |Ui .

2. The branch locus B(f ) ⊂ Y is discrete.

(Some authors omit the second requirement).


The Galois group is defined as before. As before, we can set Y ∗ =
Y − B(f ), X ∗ = f −1 (Y ∗ ) and obtain a covering map

f : X ∗ → Y ∗.

This data uniquely determines (X, f ) up to isomorphism over Y . Conversely,


for any discrete set B ⊂ Y and any covering space p : X ∗ → Y ∗ as above,
with the property that whenever U ⊂ Y is a disk and |U ∩ B| ≤ 1, each
component of p−1 (U ∗ ) maps to U with finite degree, there is an associated
branched covering f : X → Y .
Regular branched coverings are also easily constructed: given any group
G ⊂ Aut(X) acting properly discontinuously on X, there exists a Riemann
surface Y ∼
= X/G and a regular branched covering

f : X → Y = X/G

with Galois group G.


Examples.

1. The map f : C → C∗ given by f (z) = eiz is a regular covering map,


with degree group G = 2πZ.

15
2. The map cos : C → C is a regular branched covering map, with
B(cos) = {±1}. This gives the universal covering map for the infi-
nite pillowcase. Its Galois group is 2πZ n Z/2.
The map cos(z) = (eiθ + 1/eiθ )/2 can be factored as C → C∗ → C,
where the first map is a covering map and the second has degree two.

3. Given a lattice Λ, the map ℘ : C → C


b is a regular branched covering,
with Galois group the dihedral group Λ n Z/2.

4. The map f (z) = tan(z) : C → f (C) ⊂ C


b is a covering map. In fact
2iz
tan(z) = g(e ) where g(z) = −i(z−1)/(z+1). Thus f (C) = C−{±i}.
b

The Riemann-Hurwitz formula. We now return our attention to com-


pact Riemann surfaces, or more generally Riemann surfaces of finite type.
The Euler characteristic of a compact Riemann surface X of genus g is
given by
χ(X) = 2 − 2g.
If we remove n points from X to obtain an open Riemann surface X ∗ , then

χ(X ∗ ) = 2 − 2g − n.

We refer to X ∗ as a Riemann surface of finite type.

Theorem 2.3 If f : X → Y is a branched covering map between Riemann


surfaces of finite type, then
X
χ(X) = dχ(Y ) − (mult(f, x) − 1).
x∈C(f )

Examples.

1. Let f : Cb →C b be a rational map of degree d, and let C be the number


of critical points of f , counted with multiplicities. Then 2 = 2d − C
and thus C = 2d − 2.

2. Let ℘ : E → Cb be the Weierstrass ℘-function. Then deg(℘) = 2, so all


its branched points are simple, and from

0 = 2 · 2 − |B(℘)|

we conclude that |B(℘)| = 4.

16
3. Let f : E → C b give the quotient of the square torus by a 90 degree
rotation, and let C denote the number of critical points of f counted
with multiplicities. Then

0=4·2−C

so C = 8. In fact, f has critical points of multiplicity 3 at the center


and vertices of a fundamental square, and simple critical points at the
edge midpoints.

4. If X is compact and f : X → X has degree d > 1, then χ(X) ≥ 0.


Thus the sphere and the torus are the only compact Riemann surfaces
admit self–maps of higher degree. Moreover, if X = E is a complex
torus, then f is simply a covering map.

5. Let f : X → Cb be a map of degree 2 branched over 2n points, and let


g be the genus of X. Then we have:

2 − 2g = 2 · 2 − 2n,

and thus g = n − 1 as previously discussed.

Branched covers in higher dimensions. It is a special feature of the


topology in dimension two that the branched cover of a surface is a surface,
and the quotient X/G is a surface when G preserved orientation.
A nice example in higher dimensions is furnished by the three spaces:

1. The quotient C2 /G where G acts by (x, y) 7→ (−x, −y);

2. The singular quadric z 2 = xy in C3 ; and

3. The 2-fold cover of C2 branched over the pair of lines x = 0 and y = 0.

These 3 spaces are isomorphic and they are all singular complex surfaces.
The topology of the isolated singularity is seen most clearly in the case of
(1): the link of the singularity is S 3 /G ∼ = RP3 . The reason this type of
example does not exist in complex dimension 1 is that S 1 /G ∼ = S 1 when G
is a finite group of rotations.
To see the isomorphic between (1) and (2), observe that (u, v, w) =
(x2 , y 2 , xy) are invariant under G, and w2 = uv.
As for (3), one should observe that the locus xy = 0 meets S 3 in the
Hopf link L ⊂ S 3 , and the 2-fold cover of S 3 branched over L is RP3 .

17
For a related example, y 2 = x3 meets S 3 in the trefoil knot, and y p = xq
meets S 3 in the (p, q) torus knot.
Belyi’s Theorem. The Riemann surfaces that arise as coverings of C b
branched over B = {0, 1, ∞} are especially interesting. In this case the
configuration B has no moduli, so X is determined by purely combinatorial
data.

Theorem 2.4 A compact Riemann surface X is defined over a number field


iff it can be presented as a branched cover of C,
b branched over just 3 points.

We first need to explain what it means for X to be defined over a number


field. For our purposes this means there exists a branched covering map
f : X → C b with B(f ) consisting of algebraic numbers. Alternatively, X
can be described as the completion of a curve in C2 defined by an equation
f (x, y) = 0 with algebraic coefficients.
Grothendieck wrote that this was the most striking theorem he had heard
since at age 10, in a concentration camp, he learned the definition of a circle
as the locus of points equidistant from a given center.
Proof. I. Suppose X is defined over a number field, and f : X → C b is given
with B(f ) algebraic. We will show there exists a polynomial p : C → C
b b such
that B(p) ⊂ {0, 1, ∞} and p(B(f )) ⊂ {0, 1, ∞}. This suffices, since

B(p ◦ f ) = B(p) ∪ p(B(f )).

We may assume ∞ 6∈ B(f ). Let deg(B) denote the maximal degree over
Q of the points of B, and let z ∈ B(f ) have degree deg(B). Then there
exists a polynomial p ∈ Q[z] of degree d such that p(z) = 0. Moreover
deg(B(p)) ≤ d − 1 since the critical values of B are the images of the zeros
of p0 (z). Thus
B 0 = B(p) ∪ p(B)
has fewer points of degree deg(B) over Q. Iterating this process, we can
reduce to the case where B ⊂ Q.
Now comes a second beautiful trick. Consider the polynomial p(z) =
Cz a (1 − z)b . This polynomial has critical points at 0, 1 and w = a/(a + b).
By choosing the value of C ∈ Q correctly, we can arrange that p(w) = 1
and hence B(p) ⊂ {0, 1, ∞}. Thus if x ∈ B is rational, we can first apply
a Möbius transform so 0 < x < 1; then, write x = a/(a + b), and choose p
as above so that B 0 = p(B) ∪ B(p) has fewer points outside {0, 1, ∞}. Thus
we can eventually eliminate all such points.

18
II. Let X be a Riemann surface presented as a branched covering f :
X →C b with B(f ) = {0, 1, ∞}. We need a nontrivial fact: there exists a
second g ∈ M(X) and a cofinite set X ∗ ⊂ X such that (f, g) : X ∗ → C2 is an
immersion with image the zero locus Z(p) of a polynomial p(x, y) ∈ C[x, y].
This polynomial has the property that the projection of (the normaliza-
tion of) Z(p) to the first coordinate — which is just f — is branched over
just 0, 1 and ∞. The set of all such polynomials of given degree is an alge-
braic subset W of the space of coefficients CN for some large N . This locus
W is defined by rational equations. Thus the component of W containing
our given p also contains a polynomial q with coefficients in a number field.
But Z(q) and Z(p) are isomorphic, since they have the same branch locus
and the same covering data under projection to the first coordinate. Thus
X∗ ∼= Z(p) ∼= Z(q) is defined over a number field.

Corollary 2.5 X is defined over a number field iff X can be built by gluing
together finitely many unit equilateral triangles.

Proof. If X can be built in this way, one can 2-color the barycentric
subdivision and hence present X as a branched cover of the double of a
30-60-90 triangle. The converse is clear, since the sphere can be regarded as
the double of an equilateral triangle.

Example: The square torus can be built by gluing together 8 isosceles


right triangles. But these can also be taken to be equilateral triangles, with
the same result! This is because the double of any two triangles is the same,
as a Riemann surface with 3 marked points.
Trees and Belyi polynomials. Let us say p : C → C is a Belyi polynomial
if p(0) = 0, p(1) = 1 and the critical values of p are contained in {0, 1}. Then
p−1 [0, 1] is a topological tree T ⊂ C. (This is because p−1 (C − T ) maps to
C − T by a covering map of degree deg(p).)
We adopt the convention that the inverse images of 0 or 1 coincide with
the vertices of T , and that the vertices of T at 0 and 1 are labeled. The
images of the vertices of the T then alternate between 0 and 1.
A tree of this type encodes a branched covering of C b of genus 0, which
by the uniformization theorem is simply a copy of C again. Thus:
b

Theorem 2.6 Any finite topological tree T ⊂ C with 2 vertices at 0 and 1


determines a unique Belyi polynomial.

Examples.

19
1. A star with d endpoints, and its center labeled zero, corresponds to
p(z) = z d .

2. A segment with 2 internal vertices marked 0 and 1 corresponds to


the cubic polynomial p(z) = 3z 2 − 2z 3 . To see this, we note that
p0 (z) = az(1 − z) since the two vertices are internal, and use the
normalization p(0) = 0 and p(1) = 1 to solve for a and then for p(z).

Describing compact Riemann surfaces and their meromorphic func-


tions. Let X be a compact Riemann surface, and let f : X → C
b be a mero-
morphic function of degree d ≥ 1. Then (X, f ) determines the following
data:

• A finite set B = B(f ) ⊂ C;


b and

• A monodromy map φ : π1 (C
b − B, y) → Sd .

To obtain the latter, choose y ∈ Y ∗ = C b − B and label the points in f −1 (y)


as x1 , . . . , xd . Then path lifting gives a natural action of π1 (Y ∗ , y) on the
fiber over y, and hence a representation φ : π1 (Y ∗ , y) → Sd . Explicitly,
ρ(γ)(i) = j if a lift of γ starting at xi terminates at xj .
Let X ∗ = f −1 (Y ∗ ). Then f : X ∗ → Y ∗ is a covering map, and clearly

π1 (X ∗ , xi ) ∼
= {γ ∈ π1 (Y ∗ , y) : ρ(γ)(i) = i}.

To be even more explicit, let us choose loops γ1 , . . . , γn based at y and en-


circling the points of B = (b1 , . . . , bn ) in order. Then π1 (Y ∗ , y) is generated
by these elements, with the single relation

γ1 · · · γn = e.

Thus φ is uniquely determined by the permutations σi = φ(γi ), which must


satisfy two properties:

• The product σ1 · · · σn = e in Sd ; and

• The group G = hσ1 , . . . , σn i is a transitive subgroup of Sd .

The conjugacy class of σi in Sd determines a partition pi of d, which is


independent of choices. This partition records how the sheets of X ∗ come
together over bi ; it can be alternatively defined by

pi = {mult(f, x) : f (x) = bi }.

20
Theorem 2.7 The pair (X, f ) is uniquely determined by the data (B, σ1 , . . . , σn )
above, and any data satisfying the product and transitivity conditions arises
from some (X, f ).

Proof. The covering space f : X ∗ → Y ∗ is determined, up to isomorphism


over Y , by the preimage of G ∩ Sd−1 in π1 (Y ∗ , y); filling in the branch
points, we obtain the desired compact Riemann surface X and meromorphic
function f .

Example: Belyi trees. For the case of Belyi polynomials, with B =


{0, 1, ∞}, we can take z = 1/2 as the basepoint of C b − B and use loops γ0
and γ1 around z = 0 and z = 1 as generators of the fundamental group.
The permutation representation, σi = ρ(γi ), is easily read off from a picture
of the tree formed by the preimage of [0, 1]. An example is shown in Figure
2.

0 1 0 1
a e
c
b d
f

Figure 2. This Belyi tree corresponds to σ0 = (abc)(df ), σ1 = (cde).

Moduli of Riemann surfaces. Note that if we fix the genus of X and


the degree of f , then the cardinality of B is also bounded. On the other
hand, one can show that X admits a meromorphic whose degree is bounded
by g + 1. Thus Riemann surfaces of genus g depend on only finitely many
parameters.
Hurwitz problem. Here is an unsolved problem. Let f : X → C b be a
branched covering of degree d with critical values B = (bi )n1 . Let pi be the
partition of d corresponding to the fiber over bi . What partitions (p1 , . . . , pn )
can be so realized?
This is really a problem in topology or group theory. We have to lift
each pi to an element gi ∈ Sd in the conjugacy class specified by pi , in such
a way that g1 · · · gn = e and hg1 , . . . , gn i acts transitively on (1, . . . , d).
Examples.

21
1. For n = 3 and d = 4, the partitions 2+2, 2+2, 3+1 cannot be realized;
since any 2 involutions in A4 commute, σ1 σ2 has order 1 or 2, not 3.

2. If σ1 , . . . , σn−1 lie in An , then so must σn . So for example, the parti-


tions of d = 5 given by 2 + 2 + 1, 3 + 1 + 1 and 2 + 1 + 1 + 1 cannot
coexist.
In fact, this An condition just insures that deg C(f ) is even — a nec-
essary condition, by the Riemann–Hurwitz formula. Note that the Euler
characteristic of X can be calculated from the partitions (pi ). For more
details, see [EKS].
Have we seen all compact Riemann surfaces? Yes! A basic result, the
Riemann Existence Theorem, to be proved later (for compact X), states:

Theorem 2.8 Every Riemann surface admits a nonconstant meromorphic


function. In fact, the elements of M(X) separate points and provide local
charts for X.

In more detail, this means for every p 6= q in X there is an f : X → C b


such that f (p) 6= f (q) and df (p) 6= 0. In classical terminology, this means f
also separates infinitely near points.

3 Sheaves and analytic continuation


Sheaves play an essential role in geometry, algebra and analysis. They allow
one to cleanly separate the local and global features of a construction, and
to move information between analytical and topological regimes.
For example, it is trivial to locally construct a meromorphic function on
a compact Riemann surface X. We will use sheaf theory to investigate the
global obstructions. It will turn out these obstructions are finite dimensional
and thus they can be surmounted.
Presheaves and sheaves. A presheaf of abelian groups on X is a functor
F(U ) from the category of open sets in X, with inclusions, to the category
of abelian groups, with homomorphisms. (This functor is covariant, i.e. it
reverses the directions of arrows.)
It is a sheaf if
(I) elements f ∈ F(U ) are determined by their restrictions to an
open cover Ui , and
(II) any collection fi ∈ F(Ui ) with fi = fj on Uij for all i, j
comes from an f ∈ F(U ).

22
These axioms say we have an exact sequence
Y Y
0 → F(U ) → F(Ui ) → F(Uij ).
S
Here U = Ui is an open cover, the index ij is considered to be ordered,
Uij = Ui ∩ Uj , and the maps are given by f 7→ f |Ui and (fi ) 7→ (fi |Uij ) −
(fj |Uij ).
Since F(U ) is simply the kernel of the consistency map, a sheaf is
uniquely determined once we know F(U ) for all U sufficiently small.
Cohomology. The exact sequence above hints at a cohomology theory,
and indeed it can be interpreted as:
d
0 → Z 0 (F) → C 0 (F) → C 1 (F),

where the chain groups are taken relative to the covering (Ui ). There are
no 0-coboundaries, so in fact

Z 0 (F) ∼
= H 0 (F) ∼
= F(U ).

Examples.

1. The sheaves C, C ∞ , O and M, of rings of continuous, smooth, holo-


morphic and meromorphic functions. The multiplicative group sheaves
O∗ and M∗ .

2. Applying (I) to the empty cover of the empty set, we see F(∅) = (0).
Thus the presheaf that assigns a fixed, nontrivial group G to every
open set is not a sheaf.

3. If G is a nontrivial abelian group, F(U ) = G for U nonempty and


F(∅) = (0) is a presheaf. But it is not a sheaf: for a pair of disjoint
nonempty open sets, we have G = F(U1 t U2 ) 6= G ⊕ G.

4. To rectify this, we define F(U ) to be the additive group of locally


constant maps f : U → G. This is now a sheaf; it is often denoted
simply by G. (E.g. R, C, Q, S 1 , Z.)

5. The presheaf F(U ) = O(U )/C(U ) is not a sheaf. For example, local
logarithms do not assemble. This can be thought of as the presheaf of
exact holomorphic 1-forms. The exactness condition is not local.

23
Pushforward. Sheaves naturally push forward under continuous maps, i.e.
we can define
π∗ (F)(U ) = F(π −1 (U )).
In particular, homeomorphisms between spaces give rise to isomorphisms of
sheaves.
We can also pull back a sheaf under a local homeomorphism. This means
we define (π ∗ (F))(U ) = F(U ) when U is small enough that π is a local
homeomorphism, and then extend using the sheaf axiom.
Structure. From a modern perspective, a Riemann surface can be defined
as a pair (X, OX ) consisting of a connected Hausdorff topological space and
a sheaf that is locally isomorphic to (C, OC ). Similarly definitions work for
smooth manifolds. For example: a homeomorphism f : X → Y between
Riemann surfaces is holomorphic iff it sends OX to OY .
Stalks. Let F be a presheaf.
The stalk Fx is the direct limit of F(U ) over the (directed) system of
open sets containing x ∈ X. It can be described directly as the disjoint
union of these groups modulo f1 ∼ f2 if they have a common restriction
near x. (Alternatively, Fx = (⊕F(U ))/N where N is generated by elements
of the form (f |U1 ) − (f |U2 ), f ∈ F(U1 ∪ U2 ).)
There is a natural map F(U ) → Fx for any neighborhood U of x. We
let fx denote the image of f ∈ F(U ) under this map. We have fx = 0 iff
there is a neighborhood V of x such that f |V = 0.
Example: The stalk as a local ring. For any point a ∈ X a Riemann
surface, we have Oa ∼ = C{{za }}, the ring of convergent power series, for
any local chart (uniformizer) za : U → C, za (a) = 0. Note that Oa is a
local ring, with maximal ideal ma = (za ) and residue field C = Oa /ma . The
natural map Oa → C is given by evaluation at a, and the associated discrete
valuation on Oa measure the order of vanishing of the germ of a function.

Theorem 3.1 Let F be a sheaf. Then f ∈ F(U ) is zero iff fx = 0 for all
x ∈ U.

Proof. If fx = 0 for each x ∈ U then there is a covering of U by open sets


Ui such that f |Ui = 0; then f = 0 by the sheaf axioms. The converse is
obvious.

Espace étalé. The espace étalé |F| of a presheaf is the disjoint union
of the stalks Fx , with a base for the topology given by sets of the form

24
[U, f ] = {fx : x ∈ U }. It comes equipped with a natural projection

p : |F| → X

which is a local homeomorphism. Every f ∈ F(U ) gives a continuous section


s : U → |F| by s(x) = fx .
Example. For a (discrete) group G, we have |G| = G × X with the product
topology.
We say F satisfies the identity theorem if whenever U is open and con-
nected, f, g ∈ F(U ) and fx = gx for some x ∈ U , then f = g. Examples: O,
M and G.

Theorem 3.2 Assume X is Hausdorff and locally connected, and F satis-


fies the identity theorem. Then |F| is Hausdorff.

Proof. We need to find disjoint neighborhoods centered at a pair of distinct


germs fx and gy . This is easy if x 6= y, since X is Hausdorff. Otherwise,
choose a connected open set U containing x = y, and small enough that
both fx and gx are represented by f, g ∈ F(U ). Then the identity theorem
implies the neighborhoods [U, f ] and [U, g] are disjoint, as required.

Structure of |O|. Since analytic functions satisfy the identity theorem,


|O| is Hausdorff. There is a unique complex structure on |O| such that
p : |O| → X is an analytic local homeomorphism.
There is also a natural map F : |O| → C given by F (fx ) = f (x). This
map is analytic. However |O| is never connected, since the stalks Ox are
uncountable.
Thus an analytic function f : U → C has three manifestations:

(i) as an element of an abelian group, f ∈ O(U );


(ii) as a complex number at each point, f (a) ∈ C; and
(iii) as the germ of an analytic function, fa ∈ Oa .

Values of functions. The construction of the function F on |O| appears


ad hoc. It can be made more understandable by recalling that the stalks Ox
are local rings with residue field kx = Ox /mx ∼ = C. Then F (fx ) gives the
value of fx in kx .
For kx to be canonically identified with C, one must regard O as a sheaf
of C-algebras; otherwise, complex conjugation might intervene.

25
Path lifting. We now recall some results from the theory of covering spaces.
Let p : (Y, y) → (X, x) be a local homeomorphism between Hausdorff
spaces, and let f : (I, 0) → (X, x) be a path. A lift of f is a path F : (I, 0) →
(Y, y), satisfying p ◦ F = f :

(Y, y)
v:
F vvv
v
vv
p
vv f 
(I, 0) / (X, x).

Theorem 3.3 (Uniqueness of lifts) A lift of f is unique if it exists.

Proof. If we have two liftings, F1 and F2 , the set of t ∈ I such that


F1 (t) = F2 (t) is open since p is a local homeomorphism, closed since X is
Hausdorff, and nonempty since F1 (0) = F2 (0) = x. By connectedness it is
the whole interval.

Now let fs (t) be a homotopy of paths parameterized by s ∈ [0, 1], such


that fs (0) = y0 and fs (1) = y1 are constant. Suppose for every s there is a
lift Fs of fs based at x0 . We then have:

Theorem 3.4 (Monodromy Theorem) The terminus Fs (1) = x1 is in-


dependent of s. Moreover Fs (t) is a lift of fs (t) as a function on I × I. In
particular, F0 (t) and F1 (t) are homotopic rel their endpoints.

Structure of |C k |. There are plenty of natural sheaves such that |F| is


not Hausdorff and path lifting is not unique.
A good example is provide by the sheaf C k of k-times continuously dif-
ferentiable functions on X = R. You can think of |C k | as the space of all
graphs of C k functions, with overlap graphs separated like the strands in a
knot projection. However graphs that coincide over an open set are glued
together. A point know which graph it is on, at least locally.
Now any C k function f : R → R defines a section s(x) = fx , and hence a
path in |C k |. Consider, for example, k = 0 and the sections s1 and s2 defined
by f1 (x) = x and f2 (x) = |x|. These agree for x > 0 but not for x ≤ 0.
Thus path–lifting is not unique, and the germs of x and |x| at x = 0 cannot
be separated by disjoint open sets in |C k |, so the space is not Hausdorff.
Note that we are at least guaranteed that when the lift of a path goes
wayward, the two germs at least have the same value at the point of depar-
ture (here x = 0).

26
In |C k | the first k derivatives must coincide at the point of departure. In
|C ∞ |,the graphs must make contact of infinite order. Finally in the space
of real–analytic functions, the graphs must coincide and the space becomes
Hausdorff.
Structure of Spec Z. There is a similar natural sheaf defined on the
spectrum of any commutative ring, e.g.

X = Spec Z = {0, 2, 3, 5, 7, 11, . . .}

with the Zariski topology. Here the primes p > 0 are closed points, but the
closure of the generic point 0 is the whole space. Over a prime p we have a
stalk
Op ∼= Z[1/q : gcd(p, q) = 1],
with residue field Fp , and over the generic point the stalk is Q. We can
define a map F : |O| → tFp as before; then an integer n ∈ O(X) = Z gives
a section n : X → tFp with n(p) = n mod p.
Completions; affine curves. In the analytic study of a Riemann surface,
all local rings are isomorphic. But from the algebraic point of view, this is
not the case. That is, if we start with X = Spec C[x, y]/p(x, y) a smooth
affine curve, with the Zariski topology and the algebraic structure sheaf,
then the field of fractions of Oa is K(X). The stalks remember too much.
To rectify this, it is standard to complete these local rings and consider
instead
lim Oa /mna ∼
= C[[za ]].
←−

This gives the ring of formal power series, which is now sufficiently localized
that it is the same ring at every point.
Analytic continuation. Let O be the sheaf of analytic functions on X =
C. Let γ : [0, 1] → C be path with a = γ(0) and b = γ(1). Let fa be the
germ of an analytic function at z = a, i.e. let f ∈ Oa .
We say fb ∈ Ob is obtained from fa by analytic continuation along γ
if there are analytic functions gi on balls Bi centered at γ(ti ), i = 0, . . . , n
such that:

1. 0 = t0 < t1 < · · · < tn = 1;

2. Bi meets Bi+1 , gi and gi+1 agree there, and γ[ti , ti+1 ] ⊂ Bi ∩ Bi+1 ;
and

3. g0 = fa and gn = fb .

27
Modern translation. Analytic continuation along γ is equivalent to path-
lifting to the space p : |O| → C, with the lift defined by γ e(t) = [gi ] ∈ Oγ(t)
for t ∈ [ti , ti+1 ]. Thus we have, from the previous results on path–lifting:

Corollary 3.5 An analytic continuation of fa along a given path γ is unique


if it exists.

This Corollary is also obvious from the definitions, by applying the iden-
tity theorem on the connected sets Bi ∩ Bi+1 and induction on i.

Corollary 3.6 (Monodromy theorem) If analytic continuation from a


to b is possible along a family of paths γs , s ∈ I, then the result fb is always
the same.

Maximal analytic continuation (spreads). To keep track of the poten-


tial multi-valuedness, we define an analytic continuation of fa ∈ Oa , a ∈ C,
to be a pointed Riemann surface (Y, b) endowed with an analytic local home-
omorphism p : (Y, b) → (C, a) and an F ∈ O(Y ) such that p∗ Fb = f Fa .
We say an analytic continuation (Y, b) is maximal (universal might be a
better terminology) if, for any other analytic continuation (Y 0 , b0 ), there is
an (analytic) local homeomorphism j : (Y 0 , b0 ) → (Y, b) making the natural
diagrams commute.
The map j is unique if it exists, by our assumption on basepoints and
uniqueness of path lifting. Thus any 2 maximal analytic continuations are
isomorphic.
However j need not be an embedding! For example, the 2-sheeted map
p : (C∗ , 1) → (C, 1) with F (z) = p(z) = z 2 gives an analytic continuation
of f (z) = z, even though the unique maximal analytic continuation is 1-
sheeted.

Theorem 3.7 Every germ of an analytic function fa has a unique maximal


analytic continuation, obtained by taking Y to be the connected component
of |O| containing fa , and restricting the natural maps F : |O| → C and
p : |O| → C to Y .

Proof. The map y 7→ p∗ (Fy ) gives a local homeomorphism of any analytic


continuation into |O|, which is then tautologically contained in the maximal
one described above.

28
Examples.
1. The maximal domain of f (z) = z n! is the unit disk.
P
P n
2. For f (z) = z , the maximal analytic continuation in C is p : C −
{1} → C with p(z) = z and F (z) = 1/(1 − z). Evidentally f (p(z)) =
1/(1 − z) = F (z).
P n+1
3. For f (z) = z /(n + 1) — which gives a branch of log(1/(1 − z))
on the unit disk — the maximal analytic continuation is given by the
universal covering map p : C → C − {1} with p(z) = 1 − ez and
F (z) = −z. Check: f (p(z)) = log(1/ez ) = −z.
p
4. Let f (z) = q(z) where q(z) = 4z 3 + az + b. Then its maximal
analytic continuation is given by Y = E − E[2] with E = C/Λ, with
p : Y → C given p(z) = ℘(z), and with F (z) = ℘0 (z).
Remark. One can replace the base C of analytic continuation with any
other Riemann surface X. Note however that the ‘maximal’ analytic con-
tinuation may become larger under an inclusion X ,→ X 0 . An example
follows.
Counterintuitive extensions. Let f : ∆ → C be an analytic function
such that Z(f ) ⊂ ∆ accumulates on every point of S 1 . Then f admits no
analytic continuation in the sense above.
However, if may be that (∆, f ) can be extended as an abstract function
on a Riemann surface. That is, there may be a proper inclusion j : ∆ ,→ X
and a map F : X → C that extends f . For this somewhat surprising
phenomenon to occur, it must certain be true that j(Z(f )) is a discrete
subsets of X.
To construct explicit examples, it is useful to know that there exist
Riemann mappings j : ∆ → U ⊂ C such that the radial limit of j(z) is
infinity at every point. (The Fatou set of J(λez ) for λ small provides an
explicit example of such a U .) Then one can construct a set B ⊂ U which is
discrete in C, such that j −1 (B) accumulates everywhere on S 1 . Now consider
an entire function F : C → C with zero set B, and let f (z) = F ◦ j(z) on
the disk. Then f has zeros accumulating everywhere on S 1 , but f extends
from ∆ to C under the embedding j.

4 Algebraic functions
We will develop two main results. Our main case of interest will be compact
Riemann surfaces, but we will formulate them in general.

29
Theorem 4.1 Let π : X → Y be a holomorphic proper map of degree d.
Then M(X)/M(Y ) is an algebraic field extension of degree d.

To see the degree is exactly d we need to know that M(X) separates the
points of X, i.e. we will appeal to the Riemann Existence Theorem.

Theorem 4.2 Let K/M(Y ) be a field extension of degree d. Then there is


a unique degree d branched covering π : X → Y such that K ∼
= M(X) over
M(Y ).
Moreover M(X)/M(Y ) is Galois iff X/Y is Galois, in which case there
is a natural isomorphism
Gal(M(X)/M(Y )) ∼
= Deck(X/Y ).
Putting these results together, we find for example that X 7→ M(X)
establishes an equivalence between (i) the category of finite-sheeted branched
coverings of C,
b and (ii) the category of finite field extensions of C(x).
Symmetric functions. The proof will implicitly use the idea of symmetric
functions. Recall that the ‘elementary symmetric functions’ si of f1 , . . . , fd
are related to the coefficients ci ∈ Z[f1 , . . . , fd ] of the polynomial
d
Y
P (T ) = (T − fi ) = T d + c1 T d−1 + · · · + cd (4.1)
1

(−1)i ci .
P P Q
by si = (Thus s1 = fi , s2 = i<j fi fj and sd = fi .) Clearly
these polynomials si are invariant under permutation of the fi , and in fact
the ring of invariant polynomials is given by:
Z[fi ]Sd = Z[si ];
see [La, §IV.6].
Separation of points. We also observe the following consequence of the
fact that M(X) separates points. This should be compared to the theorem
of the primitive element.

Proposition 4.3 Let A ⊂ X be a finite subset of a Riemann surface. Then


there exists a meromorphic function f : X → C
b such that f |A is injective.

Proof. Let V ⊂ M(X) be a finite-dimensional space of functions that


separate the points of A = {a1 , . . . , an }. Then for any i 6= j, the subspace
Vij ⊂SV where f (ai ) 6= f (aj ) has codimension at least one. It follows that
V 6= Vij , and hence a typical f ∈ V is injective on A.

30
Proof of Theorem 4.1. Let π : X ∗ → Y ∗ be the unbranched part of
the degree d branched covering π : X → Y , and consider f ∈ M(X). By
deleting more points if necessary, we can assume f ∈ O(X ∗ ). We will show
deg(f /M(Y )) ≤ d.
Given y ∈ Y ∗ , define a degree d polynomial in T by
Y X
P (T, y) = (T − f (x)) = ci (y)T d−i .
π(x)=y

Clearly P (T, y) is globally well-defined on Y ∗ .


We now check its local properties. Since π is a covering map over Y ∗ , we
can choose local sections qi : U → X ∗ such that P (T, y) = (T − f (qi (y)))
Q
on U . This shows that the coefficients of P (T, y) are analytic functions of
y. On the other hand, the definition of P (T, y) makes no reference to these
local sections, so we get a globally well-defined polynomial P ∈ O(Y ∗ )[T ]
satisfying P (f ) = 0.
Near any puncture of Y over which f takes finite values, the coefficients
ci (y) of P (T, y) are bounded, so they extend across the puncture. At any
puncture where f has a pole, we can choose a local coordinate z on Y and
n > 0 such that z n f has no poles. Then there is a monic polynomial Q(T )
of degree d, with holomorphic coefficients, such that Q(z n f ) = 0 locally,
constructed just as above. On the other hand, P (T ) = z −nd Q(z n T ), since
both sides are monic polynomials with the same zeros. Thus the coefficients
of P (T ) extend to these punctures as well, yielding meromorphic functions
on all of Y .
In other words, we have shown there is a degree d polynomial P ∈
M(Y )[T ] such that P (f ) = 0. It follows that deg(M(Y )/M(X)) is at most
d (e.g. by the theory of the primitive element).
To see the degree is exactly d, choose a function f ∈ M(Y ) which
separates the d points in a fiber over y0 ∈ Y ∗ . Then P (f ) = 0 for a
polynomial of degree d as above. If P factors Q(T )R(T ), then Q(f )R(f ) = 0
in M(Y ). Since X is connected, M(X) is a field and one of these terms
vanishes identically — say Q(f ) = 0. Then the values of f on the fiber
over y0 are zeros of the polynomial Qy0 (T ). Since deg Q < d, this is a
contradiction.

Basic point: pushforwards. The basic point in the preceding argument


is that if f ∈ M(X) and π : X → Y is proper, then the pushforward π∗ (f )
can be defined and lies in M(Y ). The pushfoward is just the sum of f over

31
the fibers of π, i.e. it is the ‘trace’, with multiplicities taken into account:
X
π∗ (f )(y) = mult(π, x)f (x).
π(x)=y

Then π∗ (f ) gives the first symmetric function of f , and all the symmetric
functions can be expressed in terms of π∗ (f k ), k = 1, 2, . . . , d.
From a field extension to a branched cover. We now turn to the proof
of Theorem 4.2. That is, starting from a finite extension field K/M(Y ) we
will construct a Riemann surface and a proper map π : X → Y , such that
M(X) ∼ = K over M(Y ). The pair (X, π) will be uniquely determined, up
to isomorphism over Y .
To see there must be something interesting in this construction, we can
ask: what will the branch locus B = B(π) will be? The field K must
implicitly know the answer to this question.
Our first approach to the construction of X will be related to analytic
continuation and sheaves.
Local algebraic functions. We first work locally. Let

P (T, z) = T d + c1 (z)T d−1 + · · · + cd (z)

be a monic polynomial with coefficients in M(Y ). Then for each a ∈ Y


outside the poles of the ci , we get two objects:

Pa (T ) ∈ Oa [T ] and P (T, a) ∈ C[T ].

We will see the behavior of the second influences the factorization of the
first.

Theorem 4.4 QdIf P (T, a) has simple zeros, then there exist fi ∈ Oa such
that P (T ) = 1 (T − fi ).

Proof. This is simply the statement that the zeros of P (T, z) vary holo-
morphically as its coefficients do. To make this precise, let w1 , . . . , wd be the
zeros of P (T, a), and let γ1 , . . . , γd be the boundaries of disjoint disks in C,
centered at these zeros. By continuity, S there is a connected neighborhood U
of a such that P has no zeros of γi . For z ∈ U we define

P 0 (t, z) dt
Z
1
Ni (z) =
2πi γi P (t, z)

32
and
tP 0 (t, z) dt
Z
1
fi (z) = ,
2πi γi P (t, z)
where P 0 = dP/dt. Then Ni (z) is the number of zeros of P (T, z) enclosed
by γi . But Ni (a) = 1 and Ni (z) is continuous, so Ni (z) = 1 for all z ∈ U .
It follows that fi (a) = wi and fi (z) is the unique zero of P (T, z) inside γi .
Since the loops γi enclosed Q disjoint disks, the values (f1 (z), . . . , fd (z)) are
distinct, and hence P (T ) = (T − fi ).

Resultant and discriminant. The preceding results shows the zeros of


P (T ) behave well when they are all simple. To analyze this condition, it is
useful to recall the resultant R(f, g).
Let K be a field, and let f, g ∈ K[x] be nonzero polynomials with
deg(f ) = d and deg(g) = e. Recall that K[x] is a PID and hence a UFD.
We wish to determine if f and g have a common factor, say h, of degree 1 or
more. In this case f = hf1 , g = hg1 and hence g1 f − f1 g = 0. Conversely,
if we can find nonzero r and s with deg(r) < e and deg(s) < d such that
rf + sg = 0, then f and g have a common factor.
The existence of such r and s is the same as a linear relation among the
elements (f, xf, . . . , xe−1 f, g, xg, . . . , xd−1 g), and hence it can be written as
a determinant R(f, g) which is simply a polynomial in the coefficients of f
and g. We have R(f, g) = 0 iff f and g have a common factor.
In general, gcd(f, g) = h can be found by the Euclidean division algo-
rithm, and (f, g) = (h) as an ideal in K[x]. We have R(f, g) = 0 iff (f, g) is
a proper ideal, rather than all of K[x].
The discriminant D(f ) = R(f, f 0 ) is nonzero iff f has simple zeros.
Example: the cubic discriminant. Let f (x) = x3 + ax + b, so f 0 (x) =
3x2 + a. Then D(f ) = R(f, f 0 ) is given by:
 
1 0 a b 0
 
0 1 0 a b 
 
det 3 0 a 0 0 = 4a3 + 27b2 .
 
 
0 3 0 a 0
 
0 0 3 0 a

The resultant R(f, g) and discriminant D(f ) are, from this perspective,
only
Q well-defined up to sign. The usual more precise definition, D(f ) =
(x − x )2 , gives D(f ) = −4a3 − 27b2 . With this sign, K[x]/f (x) is
i<j i j
Galois ⇐⇒ D(f ) is a square in K iff the Galois group of the splitting field

33
is Z/3 (rather than S3 ). In general the discriminant begin a square implies
the Galois group is contained
Q in An . p
This is because an even permutation
preserves the sign of i−j (xi − xj ) = D(f ).
Proof of Theorem 4.2. Let K be a degree d field extension of M(Y ). By
the theorem of the primitive element, there exists a monic irreducible poly-
nomial P (T ) with coefficients ci ∈ M(Y ) such that K ∼ = M(Y )[T ]/(P (T )).
By irreducibility, the discriminant D(P ) ∈ M(Y ) is not identically zero.
(Otherwise we could compute the gcd and P and P 0 to obtain a factor of
P .) Let Y ∗ ⊂ Y be the complement of the zeros and poles of D(P ), and of
the poles of the coefficients ci .
Let X ∗ ⊂ |OY ∗ | be the set of germs fa such that P (fa , a) = 0. By the
local analysis in Theorem 4.4, X ∗ is a degree d covering space of Y ∗ . The
tautological map f : X ∗ → C sending fa to fa (a) is analytic. We can com-
plete X ∗ to a branched covering π : X → Y , and extend f to a meromorphic
function on X using Riemann’s removable singularities theorem.
We must show that X is connected. But suppose instead X = X1 t X2 .
Then f |Xi satisfies a unique monic polynomial Q ∈ M(Y )[T ], and P =
Q1 Q2 . This contradicts irreducibility of P .
Thus M(X) is a field, and by construction, P (f ) = 0. Thus M(X)
contains a copy of K. But deg(π) = d, so deg(M(X)/M(Y )) ≤ d. Thus
M(X) ∼ = K.
It remains to consider the Galois group. Suppose X/Y is Galois as a
branched covering. Since Pa (T ) has distinct zeros for some a ∈ Y , the group
Deck(X/Y ) maps injectively into Gal(M(X)/M(Y )); indeed, only the iden-
tity stabilizes the function f . By degree considerations, this map is surjective
as well. Similarly, if M(X)/M(Y ) is Galois, then G = Gal(M(X)/M(Y ))
permutes the roots of the polynomial P (T ), and hence the germs in the fac-
torization of Pa (T ). These correspond to the sheets of X, so we get a map
G → Deck(X/Y ) which is an isomorphism again by degree considerations.

The Riemann surface X can be regarded as a completion of the maximal


analytic continuation of fa , for any germ fa ∈ Oa (Y ) satisfying P (fa , a) = 0
at a point where the discriminant of P (T ) is not zero.
Construction as a curve. An alternative to the construction of π : X →
Y is the following. Fix P (T ) ∈ M(Y )[T ], irreducible, and as before let Y ∗
be the locus where the discriminant is nonzero and where the coefficients of
P (T ) are holomorphic. Then define

X ∗ = {(x, y) : P (x, y) = 0} ⊂ C × Y ∗ .

34
Let (F, π) be projections of X ∗ to its two coordinates. Then π : X ∗ → Y ∗ is
a covering map, and F : X ∗ → C is an analytic function satisfying P (F ) = 0;
and the remainder of the construction carries through as before.
Examples: quadratic extensions. For any polynomial p(z) ∈ C[z] which
is not a square (i.e. which at least one root of odd order), we canp form the
Riemann surface π : X → C b corresponding to adjoining f (z) = p(z) to
K = C(z) = M(C). b When p has 2n or 2n − 1 simple zeros, the map π is
branched over 2n points (including infinity in the latter case), and hence X
has genus n−1. Note that the polynomial P (T ) = T 2 −p(z) has discriminant
Res(P, P 0 ) = −4p(z).
Remark: Compositums of fields. It is well-known in field theory that
if K1 , K2 are two finite extensions of k, then there exists an extension L
containing both – the compositum of K1 and K2 . An analogous fact holds
for Riemann surfaces.

Theorem 4.5 Let fi : Xi → C, b i = 1, 2 be a pair of Riemann surfaces


presented as branched coverings of C.
b Then there is a third Riemann surface
Y dominating them both.
More precisely, there is a Riemann surface Y and a pair of holomorphic
maps gi : Y → Xi such that f1 ◦ g1 = f2 ◦ g2 .

Proof. Let B = B(f1 ) ∪ B(f2 ). Then we can regard Xi as the branched


b determined by a subgroup of finite index Hi ⊂ π1 (C
covering of C b − B).
Now let Y be the branched covering determined by H1 ∩ H2 .

The field M(Y ) is a compositum of M(X1 ) and M(X2 ).


Note that Y is not quite canonical, because we must choose basepoints
to make the construction. However Y can be regarded as an irreducible
component of the fiber product

X1 ×Cb X2 = {(x1 , x2 ) ∈ X1 × X2 : f1 (x1 ) = f2 (x2 )},

which is canonical.
Puiseux series.

Avec les series de Puiseux,


Je marche comme sur des oeufs.
Il s’ensuit que je les fuis
Comme un poltron que je suis.
—A. Douady, 1996

35
(Variante: Et me refugie dans la nuit.)
Let Mp be the local field of a point p on a P
Riemann surface; it is isomor-
phic to the field of convergent Laurent series ∞ i
−n ai z in a local parameter
z ∈ Op with z(p) = 0.

Theorem 4.6 Every algebraic extension of Mp of degree d is of the form


K = Mp [ζ], where ζ d = z.

Proof. We may assume p = 0 in C. Let P (T ) ∈ Mp [T ] be the degree d


irreducible polynomial for a primitive element f ∈ K (so K = M0 (f )).)
Then we can find an r > 0 such that the coefficients ci of P (T ) are well-
defined on ∆(r), only have poles at z = 0, and the discriminant D(P )|Y
vanishes only at z = 0 as well.
Note that P (T ) remains irreducible as a polynomial in M(Y )[T ], Y =
∆(r) Thus it defines a degree d branched covering X → Y , branched only
over the origin z = 0. But there is also an obvious branched covering
∆(r1/d ) → Y given by ζ 7→ ζ d = z. Since π1 (Y ∗ ) ∼
= Z has a unique subgroup
∼ 1/d
of index d, we have X = ∆(r ) over Y , and thus M(X) = M(Y )[ζ]. It
follows that K = M0 [ζ].

Corollary 4.7 Any irreducible, degree d polynomial P (T ) = 0 with coeffi-


cients in Mp has a solution of the form

X
f (z) = ai z i/d .
−n

In particular, any polynomial is locally ‘solvable by radicals’, i.e. its


roots can be expressed in the form fi (z) = gi (z 1/d ), where gi (z) is analytic.
Equivalent, after the change of variables z = ζ d , the polynomial P (T ) factors
into linear terms.
Example. For P (T ) = T 3 + T 2 − z = 0, we have a solution

z 5z 3/2 231z 5/2


f (z) = z 1/2 − + − z2 + − ···
2 8 128
Why is this only of degree 1/2? Because the original polynomial has two
distinct roots when z = 0: it is reducible over M0 !
This calculation can be carried out recursively by setting ζ = z 1/2 and
observing that
f3 f6
 
p
3 2
f = ζ 1 − f /ζ = ζ 1 − 2 − 4 − · · · .
2ζ 8ζ

36
Note: it is a general fact that only the primes dividing d appear in the
denominators of the power series for (1 − x)1/d .

Riemann surfaces Number fields

K = M(C)
b = C(z) K=Q
A = C[z] A=Z
p∈C pZ prime ideal
uniformizer zp = (z − p) ∈ C[z] uniformizer p ∈ Z
order of vanishing of f (z) at p power of p dividing n ∈ Z
Ap = O p = { ∞ n Ap = Zp = { an pn }
P P
0 a n zp }
mp = zp Op mp = pZp
residue field k = Ap /mp = C residue field k = Ap /mp = Fp
value f (p), f ∈ C[z] value n mod p, n ∈ Z
L = M(X), π : X → C
b extension field L/Q
B = {f holomorphic on X0 = π −1 (C)} B = integral closure of A in L
P ∈ X0 : π(P ) = p prime P lying over p
BP · mp = meP ; e = ramification index
k0 = BP /mP = C k 0 = BP /mP = Fpf ; f = residue degree

Table 3. A brief dictionary.

Valuations on C. b How does M(X) know about the points of X? The


answer can be formulated in terms of valuations.
A discrete valuation on a field K is a surjective homomorphism v : K ∗ →
Z such that
v(f + g) ≥ min v(f ), v(g).
The homomorphism property means

v(f g) = v(f ) + v(g).

It is also convenient to define v(0) = +∞.


Examples.

37
1. For K = C(z) = M(C) b and p ∈ C, b there is a unique point valuation
vp such that vp (f ) = n > 0 iff f has a zero of order n at p. It is called
the point valuation vp (f ).

2. For K = Q and p a prime, there is a unique p-adic valuation vp such


that vp (pk m) = k when gcd(p, m) = 1.

3. We might try to define v10 (n) as the number of 0’s in the base 10
expansion of an integer n. This works well additively, but not multi-
plicatively: v10 (2 · 5) = 1 6= v10 (2)v10 (5).
This is an argument for work with ‘decimals’ in a prime base – the
most natural such base being b = 2.

Case of equality. Perhaps the most important property of a valuation is


that:
v(f + g) = min(v(f ), v(g)) provided v(f ) 6= v(g). (4.2)
For example, adding a constant cannot destroy a pole, while it always de-
stroys a zero. To see (4.2), just note that if v(f ) > v(g) then:

v(g) = v(f +g−f ) = min(v(f +g), v(f )) = v(f +g) ≥ min(v(f ), v(g)) = v(g).

(Here the first minimum cannot be v(f ) because v(f ) > v(g)).

Theorem 4.8 Every valuation on M(C)


b is a point valuation.

Proof. Let v : K ∗ → Z be a valuation. We claim v vanishes on the constant


subfield C ⊂ K. Indeed, if f ∈ K ∗ has arbitrarily large roots f 1/n , then
v(f 1/n ) = v(f )/n → 0 and thus v(f ) = 0.
Next suppose v(z − a) > 0 for some a. Then, v(z − a) > v(a − b) for all
b 6= a. Thus 4.2 implies v(z − b) = 0 for all b 6= 0. Since a rational function
is a product of linear terms and their reciprocals, this shows v is a multiple
of the point valuation at a. But v maps surjectively to Z, so we must have
v(z − a) = 1 and hence v = va .
Now suppose v(a) ≤ 0 for all a. Since v is surjective, v(z − a) < 0 for
some a. But then v(z − b) = v(z − a) for all b, by (4.2) again. It follows
that v(z − a) = −1 for all a and hence v = v∞ gives the order of vanishing
at infinity.

38
Valuations on Riemann surfaces. Using the Riemann existence theo-
rem, it can be shown that the valuations on M(X) correspond to the points
of X, for any compact Riemann surface X. This provides a natural way
to associate X canonically to the abstract field (C-algebra) M(X). By
pushing valuations forward, we can then associate to a finite field extension
M(Y ) ⊂ M(X) a canonical map f : X → Y , which turns out to be holo-
morphic. Of course we have already seen how to do this directly, by writing
M(X) = M(Y )[T ]/P (T ), looking at the discriminant of P (T ), etc.
The general theory of valuations on a field K is facilities by the following
observations:

1. The elements with v(x) ≥ 0 form an integral domain A ⊂ K.

2. The domain A is integrally closed. Indeed, if f ∈ K and f n +a1 f n−1 +


· · ·+an−1 f = an with ai ∈ A, and v(f ) < 0; then v(f n ) has the smallest
valuation in this sum, and thus v(f n ) = v(an ) < 0, a contradiction.

3. The elements with v(x) ≥ 1 form a prime ideal I ⊂ A.

Now in the case of a compact Riemann surface X → C, b a valuation v on


M(X) restricts to one on C(x), so it lies over a point evaluation vp for C;
b
and then the corresponding prime ideal must come from one of the points
on X lying over p.
Local rings and fields.
√ For a summary of connection, see Table 3. Ex-
ample: Let L = Q( D), where D ∈ Z is square-free. Then (forgetting the
prime 2) we have ramification over the primes p|D. At these primes we have
BP ∼= Zp [p1/2 ], just as for Puiseux series.
You might think there are ‘2 points’ lying above primes p which do not
divide D. In fact, for such primes, either:

(i) T 2 − D is irreducible mod p, and there is a unique P over p;


or
(ii) D is a square mod p, and there are two primes P1 and P2
over p.

In case (i) the prime p behaves like a circle (closed string?) rather than a
point, and P like a double cover of p. These cases are distinguished by the
Legendre symbol and each occurs half the time on average.
In general when P/p has ramification index e and residue degree f , it can
be thought of roughly as modeled on the map (z, w) 7→ (z e , wf ) of ∆ × S 1
to itself.

39
5 Holomorphic and harmonic forms
In this section we discuss differential forms on Riemann surfaces, from the
perspective of sheaves and stalks. A special role is played by the holomor-
phic 1-forms, and more generally the harmonic 1-forms, which can exist in
abundance even on a compact Riemann surface (where harmonic functions
do not).
Local rings again. Let p ∈ X be a point on a Riemann surface, and let
mp ⊂ Cp∞ be the ideal of smooth, complex-valued functions vanishing at p.
Let m2p be the ideal generated by products of element in mp .

Theorem 5.1 The ideal m2p consists exactly of the smooth functions all of
whose derivatives vanish at p.

Proof. (Cf. [Hel, p.10]). Clearly the derivatives of f g, f, g ∈ mp , all vanish.


For the converse, let us work locally at p = 0 ∈ C, and suppose f ∈ m0 .
Let f1 and f2 denote the partial derivatives of f with respect to x and y.
Then we have:
Z 1 Z 1 Z 1
d
f (x, y) = f (tx, ty) dt = x f1 (tx, ty) dt + y f2 (tx, ty) dt
0 dt 0 0
= xg1 (x, y) + yg2 (x, y)

where gi (0, 0) = fi (0, 0). Thus g1 , g2 ∈ mp if both derivatives of f vanish,


which implies f ∈ m2p .

(1)
Corollary 5.2 The vector space Tp = mp /m2p is isomorphic to C2 .

Comparison between C ∞ and O. Forster skirts this point (p.60) by


defining m2p to consist of the functions vanishing to order 2. Note that for
the local ring Op , it is much easier to analyze mp because it is principle —
mp = (zp ), m2p = (zp2 ), etc. For Cp∞ , on the other hand, m2p is not principal,
because the zero loci Z(f ) for smooth f ∈ m2p are not all the same!
(1)
Tangent and cotangent spaces. We refer to Tp as the complexified
cotangent space of the real surface X at p. The exterior differential of a
function can then be defined as the map

(df )p = [f (z) − f (p)] ∈ mp /m2p .

40
(1,0)
The subspace dOp is Tp = Tp∗ X ∼
= C, the complex (or holomorphic)
tangent space. Its complex conjugate is T (0,1) . In a local holomorphic
coordinate z = x + iy, we have:
Tp(1) = C · dx ⊕ C · dy = C · dz ⊕ C · dz = Tp(1,0) ⊕ Tp(0,1) .
The dual of Tp∗ is Tp , the complex tangent space to X at p.
Note that T X and T ∗ X are complex line bundles which naturally iso-
morphic to the 2-dimensional real vector bundles given by the tangent and
cotangent bundles. In other words, the complex structure on X turns these
real vector spaces into complex vector spaces, and these smooth vector bun-
dles into holomorphic line bundles.
As for T (1) X, it is isomorphic to T ∗ X ⊗R C.
The exterior algebra of forms. A smooth 1-form α on a Riemann sur-
face is a smooth section of the vector bundle T (1) (X). Each 1-form splits
naturally into its (1, 0) and (0, 1) parts, given locally by
α = f (z) dz + g(z) dz,
where f and g are smooth functions.
One we have exterior d on functions, we can form the deRham complex
by taking exterior products of forms, and extending d so d2 = 0 and the
Leibniz rule holds: X X
d( fi dαi ) = (dfi )dαi ,
and more generally
d(αβ) = (dα)β + (−1)deg α α) dβ.
The 2-forms on X are locally of the form
β = f (x, y)dz dz,
i.e. they are (1, 1) forms. Note that dx dy = (i/2)dz dz > 0, in the sense
that the integral of this form over any region is positive.
Stokes’ theorem gives that for any compact, smoothly bounded region
Ω ⊂ X, and any smooth 1-form α, we have
Z Z
dα = α.
Ω ∂Ω

For a general smooth manifold, we can now define the deRham cohomol-
ogy groups by
k
HDR (M ) = (closed k-forms) /d(arbitrary (k − 1) forms) .

41
Integrals of metrics. We will also meet metrics of the form ρ(z)|dz|, and
their associated area forms ρ2 = ρ(z)2 |dz|2 . These can also be integrated
along paths and regions. If γ : [0, 1] → X is a smooth path, then
Z Z 1
ρ= ρ(γ(t))|γ 0 (t)| dt,
γ 0

and if Ω ⊂ X is a region with coordinate z, then


Z Z Z
i
ρ2 = ρ(z)2 dx dy = ρ(z)2 dz ∧ dz.
Ω Ω 2 Ω

Note that a metric transforms by

f ∗ (ρ) = ρ(f (z)) |f 0 (z)| |dz|.

The d-bar operator. On a Riemann surface the exterior derivative has a


natural splitting d = ∂ + ∂.
On functions, this gives the splitting of df into its (1, 0) and (0, 1) parts:

df df
df = ∂f + ∂f = dz + dz.
dz dz
Here, if z = x + iy, we have
   
df 1 df df df 1 df df
= −i and = +i .
dz 2 dx dx dz 2 dx dx

Note that df /dz = 0 iff f is holomorphic (by the Cauchy–Riemann equa-


tions), in which case f 0 (z) = df /dz.
This splitting of df is related to the factor that a real-linear map T :
C → C can be written canonically in the form

T (z) = az + bz,

with a, b ∈ C. For T = Df , we have a = df /dz and b = df /dz. Intrinsically,


we have a real-linear map df : Tp X → C, and

df (v) = (∂f )(v) + (∂f )(v)

where these two pieces are complex–linear and conjugate–linear respectively.


Note that
d(f dz) = (∂f + ∂f ) dz = (∂f ) dz,

42
since dz ∧ dz = 0. Similarly, a 1-form α = f dz + g dz satisfies
 
df dg
dα = (∂f ) dz + (∂g) dz = − + dz dz.
dz dz

Holomorphic functions. A function f : X → C is holomorphic if ∂f = 0.


Equivalently, df = ∂f .
Holomorphic forms. A 1-form α on X is holomorphic if it has type (1, 0),
and the following equivalent conditions hold:
1. ∂α = 0.
2. dα = 0; i.e. α is closed.
3. Locally α = df with f holomorphic.
4. Locally α = α(z) dz with α(z) holomorphic.
The space of all such forms on X is denoted by Ω(X).
Examples: Ω(C)b = 0; Ω(C/Λ) = C · dz; dz/z ∈ Ω(C∗ ).
Cauchy’s theorem. Since a holomorphic form is closed, it integrates to
zero over any contractible loop in X. In particular we have Cauchy’s formula
in the plane: Z
f (z) dz = 0,
γ
if γ is contractible in the domain of f . Note that this really a theorem about
the 1-form f (z) dz.
Harmonic functions. A complex-valued function u : X → C is harmonic
if the following equivalent conditions hold:
1. ∆u = d2 u/dx2 + d2 u/dy 2 = 0 in local coordinates z = x + iy.
2. ∂∂u = 0.
3. ∂u is a holomorphic 1-form.
If u is real-valued, then we can add the condition:
Locally u = Re f for some holomorphic function f .
Indeed, f can be found by integrating ∂u to obtain a holomorphic function
with ∂f = ∂u; then ∂u = ∂f (since u is real), and hence du = d(f + f )
which implies u = 2 Re(f ) up to a constant.
Harmonic 1-forms. A 1-form ω on X is harmonic if the following equiv-
alent conditions hold:

43
1. Locally ω = du with u harmonic;
2. ∂ω = ∂ω = 0 (in particular, ω is closed);
3. There exist α, β ∈ Ω(X) such that ω = α + β.
Example: dx and dy are harmonic 1-forms on X = C/Λ.
In the last condition, α and β are simply the (1, 0) and (0, 1) parts of ω.
In particular, every holomorphic 1-form is harmonic. Denoting the space of
all holomorphic 1-forms by H1 (X), this gives:

Theorem 5.3 On any Riemann surface the space of harmonic forms splits
as:
H1 (X) = Ω(X) ⊕ Ω(X).

Harmonic conjugates. It is a classical fact that if u is a real–valued


harmonic function, then locally we can find a conjugate harmonic function
v such that u + iv is analytic. In fact, since ∂u is a holomorphic 1-form, we
can find an analytic function f such that
∂u = df,
and then, since u is real,
du = ∂u + ∂u = d(f + f ),
and so u = 2 Re(f ) and v = 2 Im(f ) will work, up to an additive constant.
Periods. On a compact Riemann surface, all these spaces are finite–
dimensional. This is a corollary of general facts about elliptic differential
operators such as the Laplacian. To see it explicitly in this case at hand, we
define the period map by associating to each 1-form ω the homomorphism
φ : π1 (X) → C
R
given by γ 7→ γ ω on π1 (X).

Theorem 5.4 If X is compact, then the period map


H1 (X) → Hom(π1 (X), C) ∼
= H 1 (X, C) ∼
= C2g
is injective.

Proof. Any form in the kernel can be expressed globally as ω = du where


u is harmonic — and hence constant — on the compact Riemann surface
X.

44
Corollary 5.5 If X has genus g, then dim Ω(X) ≤ g and dim H1 (X) ≤ 2g.

In fact equality holds, but to see this will require more work.
Example: the torus. For X = C/Λ we have Ω(X) = Cω with ω = dz,
and the period map sends π1 (X) ∼
= Z2 to Λ.
Example: the regular octagon. Here is concrete example of a compact
Riemann surface X and a holomorphic 1-form ω ∈ Ω(X). Namely let X =
Q/ ≡ where Q is a regular octagon in the plane, with vertices at the 8th
roots of unity, and ≡ identifies opposite edges. Since translations preserve
dz, the form ω = dz/ ≡ is well-defined on X. It has a zero of order two
at the single point p ∈ X coming from the vertex.R The edges of Q form a
system of generators for π1 (X), and the periods α ω are given by ζ i+1 − ζ i
where ζ is a primitive 8th root of unity.
Meromorphic forms and the residue theorem. A meromorphic 1-form
ω is a (1, 0)-form with meromorphic coefficients. That is, locally we have a
meromorphic function ω(z) such that
P ωn= ω(z) dz.
Suppose we locally have ω = an z dz. We then define the order of ω
at p by:
ordp ω = inf{n : an 6= 0},
and the residue by
Resp (ω) = a−1 .
The order of pole or zero of ω at a point p ∈ X is easily seen to be inde-
pendent of the choice of coordinate system. The invariance of the residue
follows from the following equivalent definition:
Z
Resp (ω) = ω,
γ

where γ is a small loop around p (inside of which the only pole of ω is at p).
On a compact Riemann surface, the degree of ω is defined by
X
deg(ω) = ordp (ω).
p∈X
P
Theorem 5.6 If X is compact, then p Resp (α) = 0.

By considering df /f = d log f , this gives another proof of:

45
Corollary 5.7 If f : X → C
b is a meromorphic function on a compact
Riemann surface, then X
ordp (f ) = 0.
p

In other words, f has the same number of zeros as poles (counted with mul-
tiplicity).
P
(The first proof was that f (x)=y mult(f, x) = deg(f ) by general consider-
ations of proper mappings.)
Since the ratio ω1 /ω2 of any two meromorphic 1-forms is a meromorphic
function (so long as ω2 6= 0), we have:

Theorem 5.8 The ‘degree’ of a meromorphic 1-form on a compact Rie-


mann surface — that is, the difference between the number of zeros and the
number of poles — is independent of the form.

Meromorphic forms on C. b Meromorphic 1-forms on C b have degree −2,


i.e. they have 2 more poles than zeros. To check this, note that dz has a
double order pole at z = ∞, since d(1/z) = −dz/z 2 . More generally, we
have:

Theorem 5.9 Let X be a compact Riemann surface of genus g. Then every


nonzero meromorphic form on X has degree 2g − 2.

Proof. This is an easy consequence of the Riemann–Hurwitz theorem. Let


f :X→C b be a meromorphic function of degree d > 0 whose branch locus
does not include ∞. Let ω = f ∗ (dz). Then over ∞, ω has a total of 2d
poles. Each critical point of f , on the other hand, creates a simple zero.
Thus
X
deg(ω) = (multp (f ) − 1) − 2 deg(f ) = −χ(X) = 2g − 2.
p∈X

Example: hyperelliptic Riemann surfaces. We will eventually show


that dim Ω(X) = g. Let us first check this for the important case of 2-fold
branched covers of the sphere.

46
Theorem 5.10 Let B ⊂ C be a finite set of cardinality 2n, and let p(T ) =
Q
B (T − b). Let π : X → C be the unique 2-fold covering of C branched over
b b
B. Then X has genus g = n − 1, and the forms
z i dz
ωi = p , i = 0, 1, . . . , n − 2
p(z)
form a basis for Ω(X). In particular dim Ω(X) = g.

Proof. To check this, we begin by investigating ω0 . Note that if z = w2 then


dz = 2w dw. Thus the pullback of dz to X has simplep zeros at the branch
points of π, which lie over the zeros of p(z). Also p(z), as a meromorphic
function on X, has simple zerospin the same locations. These zeros cancel
when we form the quotient dz/ p(z), and thus ω0 is holomorphic except
possibly over the two unbranched points p1 , p2 ∈ pX lying over z = ∞. But
now dz has a pole of order 2 at z = ∞, while 1/ p(z) has a zero of order
n at z = ∞. We conclude:
The form ω0 is holomorphic on X, with zeros of order n − 2 at
p1 and p2 and nowhere else.
See Table 4. In particular, the degree of a meromorphic 1-form on X is
2n − 4 = 2g − 2. Since z i has a pole of order i at z = ∞, it follows that the
g forms ωi on X are holomorphic for i ≤ n − 2.


zk dz 1/ p
B
e ≥0 +1 −1
∞ −k −2 +n

Table 4. Singularities on a hyperelliptic surface.

Note: periods and hyperelliptic integrals. When B = {r1 , . . . , r2n } ⊂


R, the periods of ω0 can be expressed in terms of the integrals
Z ri+1
dx
p ·
ri (x − r1 ) · · · (x − r2n )
We can also allow one point of B to become infinity. An example of a period
that can be determined explicitly comes from the square torus:
Z 1 √
dx 2 π Γ(5/4)
p = ·
0 x(1 − x2 ) Γ(3/4)

47
A more general type of period is:
Z
X
−3 dx dy dz
ζ(3) = n = ·
0<x<y<z<1 (1 − x)yz

See [KZ] for much more on periods.


Hodge theory on Riemann surfaces. The Laplacian and the notion
of harmonic functions and forms are actually elements of the differential
geometry of Riemannian manifolds. The part of discussion which survives
on a Riemann surface (without a chosen metric) is the Hodge star operator
on 1-forms, ω 7→ ∗ω; and the associated Hermitian inner product.
Hodge star. The Hodge ∗-operator is a real–linear map

∗ : Tp(1) X → Tp(1) X

for each p ∈ X, satisfying


∗2 = −1.
In local coordinates z = x + iy, it is given by

∗dx = dy and ∗ dy = −dx.

In other words, ∗ω ‘rotates’ ω by 90 degrees.


We adopt the convention (usual for complex manifolds) that ∗ is conju-
gate linear, meaning
∗(aω) = a(∗ω).
Thus an equivalent local definition of the Hodge star operator is

∗dz = i dz and ∗ dz = −i dz.

By applying the Hodge ∗ pointwise, we obtain an operator on 1-forms.


If we write a general 1-form as

ω = α + β,

where α, β are (1, 0) forms, then

∗ω = ∗(α + β) = i(α − β).

One can check that


ω1 ∧ ∗ω2 = ω2 ∧ ∗ω1 .
In other words, the product is Hermitian.

48
Positivity. It is useful to measure the size of a 1-form by the natural 2-form

|ω|2R = ω ∧ ∗ω.

This ‘real norm’ satisfies

|a dx + b dy|2R = (|a|2 + |b|2 ) dx dy.

This normalization is slightly unnatural for complex analysis, since

|dz|2R = 2 dx dy.

As an alternative, we have the ‘complex norm’, defined by the measure

|a dz + b dz|2C = (|a|2 + |b|2 ) |dz|2 .

The measure |dz|2 coincides with that induced by dx dy together with the
canonical orientation of X. With this identification, we have

|ω|2R = 2|ω|2C .

Both these (real) 2-forms are positive, in the sense that they are non-negative
multiples of the orientation form on X, or equivalently that they integrate
to non-negative quantities.
There is a natural real linear isomorphism between the real cotangent
(1,0)
space, Tp∗ , and the complex space Tp , sending dx to dz and dy to −i dz.
This map (whose inverse is α 7→ Re(α)) is an isometry from the real norm
to the complex norm.
Hodge norm and inner product. Now assume X is compact. We can
then define a Hermitian inner product on the space of smooth 1-forms by
Z
hω1 , ω2 i = ω1 ∧ ∗ω2 .
X

The associated L2 -norm is


Z
2
kωk = hω, ωi = |ω|2R .
X

Note that Ω(X) and Ω(X) are orthogonal with respect to the Hodge
inner product. On the space Ω(X) there is a second commonly used inner
product, Z
i
{α, β} = α ∧ β.
2 X

49
It satisfies Z
1
{α, α} = |α|2C = hα, αi.
X 2
Harmonic 1-forms. We can now give a more intrinsic definition of a
harmonic 1-form, that generalizes to arbitrary Riemannian manifolds.
From the formulas above, it is evident that d ∗ ω is a linear combination
of ∂ω and ∂ω, linear independent from dω, and thus

(dω = d ∗ ω = 0) ⇐⇒ (∂ω = ∂ω = 0) ⇐⇒ (ω is harmonic).

Now suppose X is compact and we have a deRham cohomology class [ω] ∈


H 1 (X). We wish to choose the representative ω to minimize kωk2 in its
cohomology class. Formally, this means ω must be perpendicular to the
exact forms, and thus
Z
hdf, ωi = df ∧ ∗ω = 0
X

for all smooth functions


R f . Integrating by parts, we see this is equivalent to
the condition that f d ∗ ω = 0 for all f , which just says that d ∗ ω = 0 —
the form ω is co-closed. This shows:

Theorem 5.11 A 1-form on a compact Riemann surface is harmonic iff it


is closed and (formally) norm-minimizing in its cohomology class.

The general theory of elliptic operators insures that this minimum is


actual achieved at a smooth form, and thus H1 (X) ∼ = H 1 (X, C). This is
the first instance of the Hodge theorem. We will soon prove it, for Riemann
surfaces, using sheaf cohomology.
The Laplacian. The Hodge star operator also provides any Riemann sur-
face with a natural Laplacian

∆ : (function on X) → (2-forms on X).

This is consistent with electromagnetism, where µ = ∆φ should be a mea-


sure giving the charge density for the static electric field potential φ. This
Laplacian is defined by
∆f = d ∗ df ;
it satisfies
d2 f d2 f
 
∆f = + dx dy
dx2 dy 2

50
in local coordinates z = x + iy. Integrating by parts, we find the basic
identity
Z Z Z Z
f ∆f = f (d ∗ df ) = − df ∧ ∗df = − kdf k2R .
X X X X

Note that in local coordinates we have


 2  2 !
2 df df
|df |R = + dx dy.
dx dy

Thus we obtain an intrinsic form of the familiar formula for compactly sup-
ported functions in Euclidean space,
Z Z
f ∆f = |∇f |2 .
Rn Rn

Complex Hessian. The Laplacian from functions to 2-forms can also be


thought of naturally as giving the complex Hessian of f : namely, we have

d2 f i
∆f = 2i∂∂f = 4 · dz dz.
dz dz 2

Curvature. We remark that, given a conformal metric ρ = ρ(z) |dz|, the


2-form R(ρ) = ∆ log ρ is naturally the Ricci curvature of ρ, and its ratio to
the volume form of ρ itself gives the Gaussian curvature:
∆ log ρ
K(ρ) = − ·
ρ2
For example: the curvature of the metric on C∗ given by

ρα = |z|α |dz|

is zero for all α ∈ R. Note that ρ−1 makes C∗ into a (complete) cylinder;
other values near −1 make it into a (incomplete) cone.
To see this metric is flat, just note that log |z| = Re log(z) is harmonic.
In fact, we can locally change coordinates so that ρα becomes |dz|: for
f (z) = z α+1 /(α + 1), we have

f ∗ (|dz|) = |f 0 (z)| |dz| = ρα .

For α = −1 we can take f (z) = log z.

51
Spectrum of the Laplacian. When we have a metric ρ on X, we can
divide by the volume form ρ2 to get a well-defined Laplace operator ∆ρ from
functions to functions. It is usually normalized so its spectrum is positive;
then, it is defined locally by:
∆f
∆ρ (f ) = − ·
ρ2
For example, the function f (y) = y α on H is an eigenfunction of the hyper-
bolic Laplacian (for the metric ρ = |dz|/y), since
α(1 − α)y α−2
∆ρ (y α ) = = α(1 − α)y α .
y −2
Note that the curvature is simple the ρ-Laplacian of log ρ:
K(ρ) = ∆ρ (log ρ).

1-forms and foliations. Geometrically, a (locally nonzero) 1-form de-


fines a foliation
R F tangent to Ker α, and a measure on transversals given
by m(τ ) = τ α. We have dα = 0 iff m is a transverse invariant measure,
meaning m(τ ) is unchanged if we slide it along the leaves of F. The trans-
verse measure on the orthogonal foliation, defined by ∗α, is also invariant
iff α is harmonic.
Example. The level sets of Re f (z) and Im f (z) give the foliations associ-
ated to the form α = du, u = Re f (z). The case f (z) = z + 1/z in particular
gives foliations by confocal ellipses and hyperboli, with foci ±1 coming from
the critical points of f .
Appendix: The Hodge star operator on Riemannian manifolds.
Let V be an n-dimensional real vector space with an inner product hv1 , v2 i.
Choose an orthonormal basis e1 , . . . , en for V . Then the wedge products
eI = ei1 ∧ · · · eik provide an orthonormal basis for ∧k V . The Hodge star
operator ∗ : ∧k V → ∧n−k V is the unique linear map satisfying
∗eI = eJ where eI ∧ ∗eI = e1 ∧ e2 ∧ · · · en .
Here J are the indices not occurring in I, ordered so the second equation
holds. More generally we have:
v ∧ ∗w = hv, wi e1 ∧ e2 ∧ · · · en ,
and thus v ∧ ∗w = w ∧ ∗v. Since v ∧ ∗v = (−1)k(n−k) ∗v ∧ v, we have
∗2 = (−1)k(n−k) on ∧k V . Equivalently, ∗2 = (−1)k when n is even, and
∗2 = 1 when n is odd.

52
Now let (M, g) be a compact Riemannian manifold. We canRthen try
to represent each cohomology class by a closed form minimizing M hα, αi.
Formally this minimization property implies:

dα = d ∗ α = 0, (5.1)

using the fact that a minimizer satisfies


Z Z Z
hdβ, αi = (dβ) ∧ ∗α = − β∧d∗α=0
M M M

for all smooth (k − 1)-forms β. Thus we call α harmonic if (5.1) holds, and
let Hk (M ) denote the space of all harmonic k-forms.

Theorem 5.12 (Hodge) There is a natural isomorphism Hk (M ) ∼ k (M ).


= HDR

Adjoints. The adjoint differential d∗ : E k (M ) → E k−1 (M ) is defined so


that:
hdα, βi = hα, d∗ βi,
R
where hα, βi = M α ∧ ∗β. It is given by

d∗ (α) = ± ∗ d ∗ α

for a suitable choice of sign, since:


Z Z Z
hdα, βi = dα ∧ ∗β = − α ∧ d ∗ β = ± α ∧ ∗(∗d ∗ β).

Since ∗2 = 1 on an odd-dimensional manifold, in that case we have d∗ =


− ∗ d∗. For a k-form β on an even dimensional manifold, we have instead:

d∗ β = (−1)k ∗ d ∗ β.

Here we have used the fact that ∗2 = (−1)k−1 on n − (k − 1) forms such as


d ∗ β.
Generalized Hodge theorem. Once the adjoint d∗ is in play, the argu-
ments of the Hodge theorem give a complete picture of all smooth k-forms
on M .

Theorem 5.13 The space of smooth k-forms has an orthogonal splitting:

E k (M ) = dE k−1 (M ) ⊕ Hk (M ) ⊕ d∗ E k+1 (M ).

53
The Laplacian. Once we have a metric we can combine d and d∗ to obtain
the Hodge Laplacian
∆ : E k (M ) → E k (M ),
defined by
∆α = (dd∗ + d∗ d)α.
Theorem 5.14 A form α on a compact manifold is harmonic iff ∆α = 0.

Proof. Clearly dα = d ∗ α = 0 implies ∆α = 0. Conversely if ∆α = 0 then:


Z
0 = h∆α, αi
ZM
= hdα, dαi + hd∗ α, d∗ αi
M

and so dα = d ∗ α = 0 as well.

Note that on functions these definitions give

∆f = d∗ df = − ∗ d ∗ df,
R
independent of the dimension of M . This satisfies hf, ∆f i ≥ 0, but differs
by
R a 00sign from
R 0the usual Euclidean Laplacian. (For example on S 1 we have
f f = − |f |2 ≤ 0 for the usual Laplacian.)
Back to Riemann surfaces. Now suppose M has even dimension n =
2k. The Hodge star on the middle-dimensional k-forms is then conformally
invariant. Thus it makes sense to talk about harmonic k-forms when only
a conformal structure is present. Similarly, on a 4-manifold one can talk
about self-dual 2-forms, satisfying ∗α = α. These play an important role in
Yang–Mills theory.

6 Cohomology of sheaves
In this section we introduce sheaf cohomology, an algebraic structure that
measures the global obstruction to solving equations for which local solutions
are available.
Maps of sheaves; exact sequences. A map between sheaves is always
specified at the level of open sets, by a family of compatible morphisms
F(U ) → G(U ). A map of sheaves induces maps Fp → Gp between stalks.
We say F → G is injective, surjective, an isomorphism, etc. iff Fp → Gp has
the same property for each point p.

54
We say a sequence of sheaves A → B → C is exact at B if the sequence
of groups
Ap → Bp → Cp
is exact, for every p.
The exponential sequence. As a prime example: on any Riemann surface
X, the sequence of sheaves

0 → Z → O → O∗ → 0

is exact. But it is only exact on the level of stalks! For every open set the
sequence
0 → Z(U ) → O(U ) → O∗ (U )
is exact, but the final arrow need not be surjective. (Consider f (z) = z ∈
O∗ (C∗ ); it cannot be written in the form f (z) = exp(g(z)) with g ∈ O(C∗ ).)
More generally, we have:

Theorem 6.1 The global section functor is left exact. That is, for any
short exact sequence of sheaves, 0 → A → B → C → 0, the sequence of
global sections
0 → A(U ) → B(U ) → C(U )
is also exact.

Proof. We have a commutative diagram

0 / A(U ) / B(U ) / C(U )

Q Q Q
0 / Ap / Bp / Cp

where the bottom row is exact and the vertical arrows are injective. Exact-
ness at A(U ) now follows easily: going down and across (from A(U )) is 1-1,
so going across and down must be too.
For exactness at B(U ), we must show that if b ∈ B(U ) maps to zero in
C(U ), then it is in the image of A(U ). This is true on the level of stalks: for
every p ∈ U there is a neighborhood Up of p and an ap ∈ A(Up ) such that
ap |Up 7→ b|Up . Now by the first part of the proof, A(Up ) → B(Up ) is 1 − 1,
so ap is unique. But then ap − aq = 0 on Upq , so these local solutions piece
together to give an element a ∈ A(U ) that maps to b.

55
Sheaf cohomology is the derived functor which measures the failure of
exactness to hold on the right.
Čech theory: the nerve of a covering. A precursor to sheaf cohomology
is Čech cohomology. The idea here is that any open covering U = (Ui ) of
X has an associated simplicial complex Σ(U) that is an approximation to
the topology of X. The simplices T in this complex are simply ordered finite
sequence of indices I such that I Ui 6= ∅.
This works especially well if we require that all the multiple intersections
are connected. Note that this is equivalent to requiring that Z(UI ) ∼ = Z
whenever UI 6= ∅. Then the ordinary simplicial cohomology H (Σ(U), Z)∗

turns out to agree with the sheaf cohomology H ∗ (U, Z).


Cochains, cocycles and coboundaries. When we have a general sheaf
F in play, we can modify Čech cohomology by allowing the ‘weights’ on
the simplex UI to take values in the group F(UI ). (For the usual Čech
cohomology, we would take F = Z.)
With these ‘weights on simplices’, we define the space of q-cochains by:
Y
C q (U, F) = F(UI ).
|I|=q+1

Here I ranges over ordered sets of indices (i0 , . . . , iq ), and

UI = Ui0 ∩ · · · ∩ Uiq .

Examples: a 0-cochain is the data fi ∈ F(Ui ); a 0-cochain is the data


gij ∈ F(Ui ∩ Uj ); etc.
Next we define a boundary operator

δ : C q (U, F) → C q+1 (U, F)

by setting δf = g where, for q = 0:

gij = fj − fi ;

for q = 1:
gijk = fjk − fik + fij ,
and more generally
q
X
gI = (−1)j fIj
0

where Ij = (i0 , i1 , . . . , ibj , . . . , iq+1 ). When two indices are eliminated, they
come with opposite sign, so δ 2 = 0.

56
The kernel of δ is the group of cocycles Z q (U, F), its image is the group
of coboundaries B q (U, F), and the qth cohomology group of F relative to
the covering U is defined by:

H q (U, F) = Z q (U, F)/B q (U, F).

Example: H 0 . A 0-cocycle (fi ) is a coboundary iff fj − fi = gij = 0 for


all i and j. By the sheaf axioms, this happens iff fi = f |Ui , and thus:

H 0 (U, F) = F(X).

Example: H 1 . A 1-cocycle gij satisfies gii = 0, gij = −gji and

gij + gjk = gik

on Uijk . It is a coboundary if it can be written in the form gij = fi − fj .


Example on S 1 . Let F = Z. Suppose we cover S 1 with n ≥ 3 intervals
U1 , . . . , Un such that Uij is an interval when i and j are consecutive, and
otherwise Uij isPempty. Then the space of 1-cocycles in simply Zn , but
gij = fi − fj iff gi,i+1 = 0. Thus H 1 (U, Z) = Z.
Refinement. Whenever V = (Vi ) is a finer covering than U = (Ui ), we can
choose a refinement map on indices such that Vi ⊂ Uρi . Once ρ is specified,
it determines maps F(UρI ) → F(VI ), and hence chain maps giving rise to a
homomorphism
H q (U, F) → H q (V, F),
which sends the chain (fI |UI ) to the chain (fρI |VI ).

Theorem 6.2 The refinement map H q (U, F) → H q (V, F) is independent


of ρ.

Definition. We define the cohomology of X with coefficients in F by:

H q (X, F) = lim H q (U, F),


−→

where the limit is taken over the system of all open coverings, directed by
refinement.

Theorem 6.3 The refinement map H 1 (U, F) → H 1 (V, F) is injective.

57
Proof. Suppose we are given coverings (Ui ) and (Vi ) with Vi ⊂ Uρi . Let
gij be a 1-cocycle for the covering (Ui ) that becomes trivial for (Vi ). That
means there exist fi ∈ F(Vi ) such that

gρi,ρj = fi − fj

on Vij .
Our goal is to find hk ∈ F(Uk ) so gkm = hk − hm . The first attempt is
to set
hik = fi
on Uk ∩ Vi , and hope that these patch together. But in fact we have

hik − hjk = fi − fj = gρi,ρj = gρi,k − gρj,k

on Uk ∩ Vi ∩ Vj . The clever part is the last expression, which follows from


the cocycle condition on gij . It suggests correcting our definition of hk as
follows:
hk = fi − gρi,k
on Uk ∩ Vi . Now this definition agrees on overlaps, so by the sheaf axioms
it gives a well–defined section hk ∈ F(Uk ).
The rest of the proof is now straightforward: by the cocycle condition
again, on Ukm ∩ Vi , we have

hk − hm = fi − gρi,k − fi + gρi,m = gkm

as desired.

Corollary 6.4 We have H 1 (X, F) = H 1 (X, U; F).


S

Intuition for q = 1. To get a nontrivial element of H 1 (X, U; Z), it suffices


to find a covering of X by connected open sets U1 , . . . , Un such that Ui meets
Ui+1 (and U1 meets Un ) and all other pairwise intersections are trivial. This
shows X is ‘coarsely’ a circle, and this property persists under refinements
because the Ui are connected.
Failure for q ≥ 2. Refinement is not injective for q ≥ 2 in general. For
example, let X be the 1-skeleton of a 3-simplex and let U = (U1 , . . . , U4 ) an
open covering obtained by thickening each of its triangular ‘faces’. Then the
nerve of U is a simplicial 2-sphere and hence H 2 (X, U; Z) = Z even though
H 2 (X; Z) = 0.

58
Note that U has a refinement V with no 3-fold intersections.
Leray coverings. An open set U is acyclic (for F) if H q (U, F) = 0 for all
q > 0. We say U is a Leray covering of (X, F) if UI is acyclic for all indices
I.

Theorem 6.5 (Leray) If U is a Leray covering, then

H q (X, F) ∼
= H q (U, F).

If just H 1 (Ui , F) = 0 for every i, U can still be used to compute H 1 .

(Note: Forster works only with Leray covers of ‘first order’, i.e. those in
the second part of the theorem.)
Vanishing theorem by dimension. Using the existence of fine coverings
where the (n + 2)-fold intersections are all empty, we have:

Theorem 6.6 For any n-dimensional space X and any sheaf F, H p (X, F) =
0 for all p > n.

Vanishing theorems for smooth functions, forms, etc. Let F be the


sheaf of C ∞ functions on a (paracompact) manifold X, or more generally a
sheaf of modules over C ∞ . We then have:

Theorem 6.7 The cohomology groups H q (X, F) = 0 for all q > 0.

Proof for q = 1. To indicate the argument, we will show H 1 (U, F) = 0


for any open covering U = (Ui ). We will use the fact that there exists a
partition of unity ρi ∈ C ∞ (X) subordinate to Ui : that is, a set of functions
withP Ki = supp ρi ⊂ Ui , such that Ki forms a locally finite covering of X
and ρi (x) = 1 for all x ∈ X.
Let gij ∈ Z 1 (U, F) be a 1-cocycle. Then gii = 0, gij = −gji and gij +
gjk = gik . Our goal is to write gij = fj − fi (or fi − fj ).
How will we ever get from gij , which is only defined on Uij , a function
fi define on all of Ui ? The central observation is that:

ρj gij , extended by 0, is smooth on Ui .

This is because ρj vanishes on Ui − Uj . More precisely, ρj gij is smooth


wherever gij is defined or ρj is zero. That is, it is smooth on

(Ui − Uj ) ∪ (Uij ) = Ui .

59
(Some care should be taken at the points of ∂Uj ∩ Ui .) Thus we can define:
X
fi = ρk gik ;
k

and then:
X X X
fi − fj = ρk (gik − gjk ) = ρk (gik + gkj ) = ρk gij = gij .
k k k

P
Proof for q = 2. Let fab = x ρx gabx . Then
X X
δ(f )abc = fbc − fac + fab = ρx (gbcx − gacx + gabx ) = ρx gabc = gabc .
x x

The exact cohomology sequence; deRham cohomology. We can now


explain how sheaf cohomology is used to capture global aspects of analytic
problems that can be solved locally.
Let E p denote the sheaf of smooth p-forms on a manifold X. Suppose
α ∈ E 1 (X) is closed; then locally α = df for f ∈ E 0 (X). When we can we
find a global primitive for α?
To solve this problem, let U = (Ui ) be an open covering of X by disks.
Then we can write α = dfi on Ui . On the overlaps, gij = fi − fj satisfies
dgij = 0, i.e. it is a constant function. Moreover we obviously have gij +gjk =
gik , i.e. gij is an element of Z 1 (U, C).
Now we may have chosen our fi wrong to fit together, since fi is not
uniquely determined by the condition dfi = αi ; we can always add a constant
function ci . But if we replace fi by fi + ci , then gij will change by the
coboundary ci − cj . Thus we can conclude:

α = df iff [gij ] = 0 in H 1 (X, C) .

The exact cohomology sequence. The conceptual theorem underlying


the preceding discussion is the following:

Theorem 6.8 Any short exact sequence of sheaves on X,

0 → A → B → C → 0,

60
gives rise to a long exact sequence

0 → H 0 (X, A) → H 0 (X, B) → H 0 (X, C) →


H 1 (X, A) → H 1 (X, B) → H 1 (X, C) →
H 2 (X, A) → H 2 (X, B) → H 2 (X, C) → · · ·

on the level of cohomology.

Note: for any open set U , we get an exact sequence

0 → A(U ) → B(U ) → C(U ), (6.1)

as can be checked using the sheaf axioms. Surjectivity of the maps Bx → Cx


implies that for any c ∈ C(X), there is an open covering (Ui ) and bi ∈ B(Ui )
mapping to c.
To obtain the connecting homomorphism

ξ : H 0 (X, C) → H 1 (X, A),

we use the exactness of (6.1) to write bi −bj as the image of aij . The resulting
cocycle [aij ] ∈ H 1 (X, A) is the image ξ(c).
Sheaves and deRham cohomology. Let Z p denote the sheaf of closed
p-forms. Recall that the deRham cohomology groups of X are given by:
p
HDR (X) = Z p (X)/dE p−1 (X).

The Poincaré Lemma asserts that


p
HDR (Rn ) = 0

for all p and n.


Chain homotopy. The proof is by induction on n. We have natural
maps π : Rn+1 → Rn and s : Rn → Rn+1 , such that π ◦ s = id. These
maps in fact give a homotopy equivalence — no surprise, since both spaces
are contractible. We wish to show they induce isomorphisms on deRham
cohomology, which amount so showing that the map on forms given by

φ = I − π ∗ ◦ s∗
∗ (Rn+1 ).
induces the identity on HDR

61
To this end there is a basic tool. Note that φ is a chain map — it
commutes with d. To show φ is trivial on cohomology, it suffices to construct
a chain homotopy K, which lowers the degree of forms by one, such that

φ = dK + Kd.

(Note that d(dK + Kd) = dKd = (dK + Kd)d). If [α] represents a coho-
mology class, then dα = 0, so

φ(α) = (dK + Kd)(α) = dK(α)

is cohomologous to zero, and we are done. Here are the details.


Proof. We can regard Rn+1 as a bundle over Rn with projection π and zero
section s. We wish to construct K such that

I − π ∗ ◦ s∗ = Kd + dK.

Let (t, x) denote coordinates on R × Rn , and let α denote a product of some


subset of dx1 , . . . , dxn . (So dα = 0). Then we define K(f (t, x)α) = 0, and

K(f (t, x) dt α) = F (t, x)α,


Rt
where F (t, x) = 0 f (t, x) dt.
The verification that K gives a chain homotopy is straightforward. For
example, dK +Kd should be the identity on a form like f dtα, since s∗ (dt) =
0. And indeed,

dK(f dt, α) = d(F α) = f dt α + (dx F )α

while
Kd(f dtα) = K((dx f ) dtα) = −(dx F )α.
The sign enters because we must move dt in front of dx f .
On a form like f α we should get (f (t, x) − f (0, x))α, and indeed K kills
this form while

Kd(f α) = K((df /f t) dtα + (dx f )α) = (f (t, x) − f (0, x))α.

62
The deRham theorem. Since every manifold looks locally like Rn , the
Poincaré Lemma implies we have an exact sequence of sheaves:
d
0 → Z p−1 → E p−1 → Z p → 0.
Now let p = 1. Then Z p−1 = C. By examining the associate long exact
sequence we find:
1 (X) ∼ H 1 (X, C).
Theorem 6.9 For any manifold X, we have HDR =
More generally, using all the terms in the exact sequence and all values
of p, we find:
Theorem 6.10 For any manifold X, we have
p
HDR (X) ∼
= H 1 (X, Z p−1 ) ∼
= H 2 (X, Z p−2 ) ∼
= · · · H p (X, Z 0 ) = H p (X, C).
A key point in this discussion is that:
The sheaf Z p , unlike E p , is not a module over C ∞ .
This is particularly clear for Z 0 = ∼ C, and it explains why the cohomology
p
of Z need not vanish.
Corollary 6.11 The deRham cohomology groups of homeomorphic smooth
manifolds are isomorphic.
(In fact one can replace ‘homeomorphic’ by ‘homotopy equivalent’ here.)
p
Finiteness. Now it is easy to prove that HDR (Rn ) = 0 for all p > 0. We
thus have, by taking a Leray covering:
Theorem 6.12 For any compact manifold, the cohomology groups H p (X, C)
are finite.

Periods revisited. Finally we mention the fundamental group:


Theorem 6.13 For any connected manifold X, we have
H 1 (X, C) ∼ 1
= HDR (X) ∼= Hom(π1 (X), C).
R
Proof. We have γ df = 0 for all closed loops γ, so the period map is
Rq
well-defined; if α has zero periods then f (q) = p α is also well-defined and
satisfies df = α. To prove surjectivity, take a (Leray) covering of X by
geodesicly convex sets, and observe that (i) every element of π1 (X) is repre-
sented by a 1-chain and (ii) every 1-boundary is a product of commutators.

63
Warning: exotic topological spaces . One can define the period map
H 1 (X, C) → Hom(π1 (X), C) directly, using γ : S 1 → X to obtain from ξ ∈
H 1 (X, C) a class φ(γ) ∈ H 1 (S 1 , C) ∼
= C. For more exotic topological spaces,
however, this map need not be an isomorphism: e.g. the ‘topologist’s sine
curve’ is a compact, connected space X with π1 (X) = 0 but H 1 (X, Z) = Z.

7 Cohomology on a Riemann surface


On a Riemann surface we have the notion of holomorphic functions and
forms. Thus in addition to the sheaves E p we have the important sheaves:

O — the sheaf of holomorphic functions; and


Ω — the sheaf of holomorphic 1-forms.

Let hi (F) = dim H i (X, F). We will show that a compact Riemann surface
satisfies:
h0 (O) = 1, h0 (Ω) = g,
h1 (O) = g, h1 (Ω) = 1.
The symmetry of this table is not accidental: it is rather the first instance
of Serre duality, which we will also prove.
We remark that a higher cohomology group, like H 1 (X, O), is a sort of
place holder, like paper money. It has no actual worth until it is exchanged
for something else. Serre duality makes such an exchange possible.
The Dolbeault Lemma. Just as the closed forms can be regarded as
the subsheaf Ker d ⊂ E p , the holomorphic functions can be regarded as the
subsheaf Ker ∂ ⊂ C ∞ = E 0 . So we must begin by studying the ∂ operator.
We begin by studying the equation df /dz = g ∈ L1 (C). An example is
given for each r > 0 by:
(
1/z if |z| > r,
fr (z) =
z/r2 if |z| ≤ r,

which satisfies gr = df /dz = (1/r2 )χB(0,r) (z). In particular


R
gr = π is
independent of r, which suggests the distributional equation:
d 1
= πδ0 .
dz z
Using this fundamental solution (and the fact that dx dy = (i/2)dz ∧ dz),
we obtain:

64
Theorem 7.1 For any g ∈ Cc∞ (C), a solution to the equation df /dz = g is
given by: Z
1 i g(w)
f (z) = g ∗ = dw ∧ dw.
πz 2π C z − w

Proof. It is enough to check df /dz at z = 0. Using that fact that df /dz =


1
(dg/dz) ∗ πz , we find:
  Z  
dg −1
Z
df g(z) dz
−2πi (0) = dz dz = d
dz C dz z C z
Z   Z
g(z) dz dz
= lim d = − lim g(z) = −2πi g(0).
r→0 C−B(0,r) z r→0 S 1 (0,r) z

The penultimate minus sign comes from the back that S 1 (0, r) acquires a
negative orientation as ∂(C − B(0, r)).

Theorem 7.2 For any g ∈ C ∞ (∆), there is an f ∈ C ∞ (∆) with df /dz = g.


P
Proof. Write g = gn where each gn is smooth and compactly supported
outside the disk Dn of radius 1 − 1/n. Solve dfn /dz = gn . Then fn is
holomorphic on Dn . Expanding it in power series, we can find holomorphic
functions
P hn on the disk such that |fn − hn | < 2−n n Dn . Then f =
(fn − hn ) = lim Fi converges uniformly, and F − Fi is holomorphic on Dn
for all i > n, so the convergence is also C ∞ .

Corollary 7.3 For any g ∈ C ∞ (∆), there is an f ∈ C ∞ (∆) with ∆f = g.

Proof. First solve dh/dz = g, then df /dz = h. Then ∆(f /4) = d2 f /dz dz =
g.

Remarks. The same results hold with ∆ replaced by C. It is easy to solve


the equation df /dz = g on C when g(z, z) is a polynomial: simply formally
integrate with respect to the z variable.
In higher dimensions, ω = ∂g should be thought of as a (0, 1)-form, and
this equation can be solved for g only when ∂ω = 0.
Functions with small ∂ derivative. Here is an easy consequence of the
Dolbeault lemma above. Suppose f (z) is smooth on the disk 2∆ of radius
2. Then on the unit disk ∆, we can write

f (z) = h(z) + g(z)

65
where h(z) is holomorphic and g is small on ∆ if ∂f is small on 2∆. Here
smallness can be measured, e.g. in the C k norm. Informally,

f = holomorphic + O(∂f ).

To see this, we simply solve ∂g = ∂ρf by convolution with 1/z, where ρ is


a smooth cutoff function = 1 on ∆ and = 0 outside 2∆.
Dolbeault cohomology. Let us define, for any Riemann surface X,

H 0,1 (X) = E 0,1 (X) ∂E 0 (X).




The preceding results show the sequence of sheaves:



0 → O → E 0 → E 0,1 → 0

is exact. Consequently we have:

Theorem 7.4 On any Riemann surface X, we have H 1 (X, O) ∼


= H 0,1 (X).

Thus the Dolbeault lemma can be reformulated as:

Theorem 7.5 The unit disk satisfies H 1 (∆, O) ∼


= H 0,1 (∆) = 0. The same
is true for the complex plane.

Corollary 7.6 We have H 1 (C,


b O) = 0.

Proof. Let U1 ∪ U2 be the usual covering by U1 = C and U2 = C b − {0}.


1
By the preceding result, H (Ui , O) = 0. Thus by Leray’s theorem, this
covering is sufficient for computing H 1 : H 1 (C,
b O) = H 1 (U, O). Suppose
g12 ∈ O(U12 ) = O(C ) is given. Then g12 (z) = ∞
∗ n
P
−∞ an z . Splitting this
Laurent series into its positive and negative parts, we obtain fi ∈ O(Ui )
such that g12 = f2 − f1 .

Similarly, we define

H 1,1 (X) = E 1,1 (X) ∂E 1,0 (X).




Theorem 7.7 On any Riemann surface X we have H 1 (X, Ω) ∼


= H 1,1 (X).

Proof. Consider the exact sequence of sheaves



0 → Ω → E 1,0 → E 1,1 → 0.

66
Corollary 7.8 The dimension of H 1 (X, Ω) is ≥ 1.

α on H 1,1 (X) has image C.


R
Proof. The map α 7→ X

We will later see that equality holds.


2
The Dolbeault isomorphism. Using the fact that ∂ = 0 one can defined
the Dolbeault cohomology groups for general complex manifolds, and prove
using sheaf theory the following variant of the deRham theorem:

Theorem 7.9 For any compact complex manifold X, we have H∂p,q (X) ∼
=
H q (X, Ωp ).

Here Ωq is the sheaf of holomorphic (q, 0)-forms.


Finiteness. Recall that we already know dim Ω(X) ≤ g, by period con-
siderations; in particular, Ω(X) is finite-dimensional. We remark that a
more robust proof of this finite-dimensionality can R be given by endowing
2 2
Ω(X) with a reasonable normal — e.g. kαk = X |α| — and then ob-
serving that the unit ball is compact. (The same proof applies to show
dim O(X) < ∞, without using the maximum principle. More generally the
space of holomorphic sections of a complex line bundle over a compact space
is finite-dimensional.)
A proof of the finiteness of dim H 1 (X, O), based on norms as in the
discussion, is given in Forster.
Serre duality, special case. We will approach finiteness instead through
elliptic regularity. We begin with a special case of Serre duality, which
allows one to relate H 1 and H 0 . The proof below is based on Weyl’s lemma,
which states that a distribution f is represented by a holomorphic function
if ∂f = 0.

Theorem 7.10 On any compact Riemann surface X, we have H 1 (X, O)∗ ∼


=
Ω(X). In particular, H 1 (X, O) is finite-dimensional.

We let ga = dim Ω(X), the arithmetic genus of X. We will eventually


see that ga = g = the topological genus.
Proof. By Dolbeault we have

H 1 (X, O) ∼
= H 0,1 (X) = E (0,1) (X)/∂E 0 (X).

We claim ∂E 0 (X) is closed in E (0,1) (X) in the C ∞ topology. If not, there


is a sequence fn ∈ E 0 (X) with ∂fn → ω in the C ∞ topology, such that

67
fn has no convergent subsequence in E 0 (X). Since ∂ annihilates constants,
we may normalize so that fn (p) = 0 for some p and all n. Since bounded
sets in E 0 (X) are compact, the latter condition implies for some k ≥ 0,
kfn kC k → ∞. From this we will obtain a contradiction.
Dividing through by the C k -norm, we can arrange that kfn kC k = 1 in
E 0 , and that ∂fn → 0 in the C ∞ topology. Solving the ∂-equation locally,
this means we can write locally write

fn = hn + gn

where hn is holomorphic and gn → 0 in C ∞ . But then hn is bounded, so


we can pass to a subsequence so hn converges smoothly. Then fn → F
smoothly, so ∂F = 0, so F is constant. Indeed, F = 0 since fn (p) = 0. On
the other hand, our normalization that kfn kCk = 1 implies kF kC k = 1, a
contradiction.
Since ∂E 0 (X) is closed, we have

H 1 (X, O)∗ = W ∼ = (∂E 0 (X))⊥ ⊂ (E 0,1 (X))∗ = D1,0 (X).


R
But any (1, 0)-current ω ∈ W satisfies ω ∧ (∂f ) = 0 for all smooth f , and
thus ∂ω = 0, which implies ω is holomorphic and thus W = Ω(X).

Corollary 7.11 We have natural isomorphisms H 1,0 (X) ∼


= Ω(X) and H 0,1 (X) ∼
=
Ω(X).

Proof. By definition, H 1,0 (X) consists of ∂-closed (1, 0) forms, which means
it is isomorphic to Ω(X).
As for H 0,1 (X), we have a natural map 0,1
R Ω(X) → H (X). On this
subspace, the pairing with Ω(X) given by α ∧ β is obviously perfect, since
(i/2) α ∧ α > 0. Thus Ω is isomorphic to H 0,1 (X).
R

Elliptic regularity. Here is a brief sketch of the proof of elliptic regularity.


We will show that if f is a distribution on Rn and ∆f = 0, then f is
represented by a smooth harmonic function.
The first point to observe is that morally, if f is a compactly supported
distribution and
∆f = u
with u ∈ C k , then f ∈ C k+2 . Here k is allowed to be negative. This can be
made precise in terms of Fourier transformsP and Sobolev spaces; it comes
from the fact that the symbol of ∆, namely ξi2 , has quadratic growth.

68
Now suppose we have a distribution with ∆f = 0. Let ρ be a C ∞ ,
compactly supported cutoff function. Then

∆(ρf ) = ρ∆f + u = u

where u involves only the first derivatives of f (and ρ). Thus if f is in C k ,


then u ∈ C k−1 and hence ρf is in C k+1 . With some finesse to get around
the fact that ρf is not quite the same as f , this shows f ∈ C ∞ .
For more details, see e.g. Rudin, Functional Analysis.
Taking stock. In preparation for Riemann-Roch, we now know H 1 (X, O)∗ ∼ =
Ω(X), ga = dim Ω(X) ≤ g, and dim H 1 (X, Ω) ≥ 1 because of the residue
map.
The space H 1 (X, Ω). We can mimic the proof that H 1 (X, O) = ∼ Ω(X)
to show
H 1,1 (X) ∼
= H 0 (X, O)∗ ∼
= C.
The main point is that

H 1,1 (X)∗ ∼
= (∂E 1,0 (X))⊥ ⊂ D0 (X),

and we have Z
f ∂α = 0

for all α ∈ E (1,0) (X) iff ∂f = 0 iff f is holomorphic.


Another argument will be given below.

8 Riemann-Roch
One of the most basic questions about a compact Riemann surface X is:
does there exist a nonconstant holomorphic map f : X → C? b But as we
have seen already in the discussion of the field M(X), it is desirable to ask
more: for example, do the meromorphic functions on X separate points?
Even better, does there exist a holomorphic embedding X → Pn for some
n?
To answer these questions we aim to determine the dimension of the
space of meromorphic functions with controlled zeros and poles. This is the
Riemann-Roch problem.
Divisors. Let X be a compact Riemann surface. The group of divisors
Div(X) is thePfree abelian group generated by the points of X. A divisor
is sum D = aP P , where aP ∈ Z and aP = 0 for all but finitely many
P ∈ X.

69
A divisor is effective if D ≥ 0, meaning aP ≥ 0 for all P . Any divisor
can be written uniquely as a difference of effective divisors, D = D+ − D− .
We write E ≥ D if E − D ≥ 0. P
The degree of a divisor, deg(D) = aP , defines a homomorphism deg :
Div(X) → Z. We let Div0 (X) denote the subgroup of divisors of degree
zero.
The sheaf of divisors. The quotient M∗ /O∗ is not officially a sheaf, but
it can be made into a sheaf in a standard way (take sections of the espace
étalé). A section of M∗ /O∗ over U is given by a covering Ui and elements
fi ∈ M∗ (Ui ) such that fi /fj ∈ O∗ (Uij ). Note that degP (fi ) gives the same
value for any i with P ∈ Ui . P
One can also define a sheaf by Div(U ) = { aP P } where the sum is
locally finite. We then have a natural isomorphism of sheaves,

M∗ /O∗ ∼
= Div,
P
send (fi ) to degP (fi ).
Principal divisors. The divisor of a global meromorphic function f ∈
M∗ (X) is defined by X
(f ) = ord(f, P ) · P.
P
The divisors that arise in this way are said to be principal. Note that:

(f g) = (f ) + (g) and deg((f )) = 0,

so f 7→ (f ) defines a homomorphism from

M∗ (X) → Div0 (X).

(Its cokernel is a beautiful complex torus, the Jacobian of X, to be discussed


later.)
Note that the topological degree of f : X → C b is given by deg((f )+ ),
which gives the number of zeros of f counted with multiplicities.
The map (f ) 7→ Div(f ) can be thought of as part of the long exact
sequence associated to

0 → O∗ → M∗ → M∗ /O∗ → 0,

which is short exact by the definition of the quotient.


Constrained zeros and poles We defined OD as the sheaf of complex
vector spaces given by meromorphic functions f such that (f ) + D ≥ 0. For

70
example, OnP (X) is the vector space of meromorphic functions on X with
poles of order at most n at p. In general OD keeps the poles of f from being
to big and forces some zeros at the points where aP < 0.
If D −E = (f ) is principal, we say D and E are linearly equivalent. Then
the map h 7→ hf gives an isomorphism of sheaves:

OD = OE+(f ) ∼
= OE ,

because
(hf ) + E = (h) + (f ) + E = (h) + D.

Constrained forms. We let M1 (X) denote the space of meromorphic 1-


forms on X. It is a 1-dimensional vector space over the field M(X). The
divisor K = (ω) of a nonzero meromorphic 1-form ω is defined just as for a
meromorphic function, in terms of the zeros and poles and ω.
Once we have such a canonical divisor, we get an isomorphism

OK ∼
=Ω

by h 7→ hω, because hω is holomorphic iff

(hω) = (h) + K ≥ 0

iff h ∈ OK .
Similarly, for any divisor D we let

ΩD (X) = {ω ∈ M1 (X) : (ω) + D ≥ 0}.

The sheaf ΩD is defined similarly, and it satisfies

ΩD ∼
= OK+D .

Global sections of this sheaf correspond to meromorphic forms with con-


strained zeros and poles.
Riemann-Roch Problem. The Riemann-Roch problem is to calculate or
estimate, for a given divisor D, the number

h0 (D) = dim H 0 (X, OD ) = dim OD (X).

Example: if X is a complex torus, we have h0 (nP ) = 0, 1, 1, 2, 3, . . . for


n = 0, 1, 2, 3, 4, . . .. This can be explained by the fact that we must have
ResP (f dz) = 0.

71
It is a general principle that Euler characteristics are more stable than
individual cohomology groups, and for any sheaf of complex vector spaces
F on X we define: X
χ(F) = (−1)q hq (F),
assuming these spaces are finite–dimensional. (Of course hq (F) = 0 for all
q > 2 by topological considerations.)
Example: Using the fact that h1 (O) = h0 (Ω) = ga , we have

χ(O) = 1 − ga .

We may now state:

Theorem 8.1 (Riemann-Roch, Euler characteristic version) For any


divisor D, we have
χ(OD ) = χ(O) + deg(D).

Equivalently, we have

h0 (OD ) − h1 (OD ) = deg D − ga + 1.

We will soon show that ga = g.


For the proof we will use:

Theorem 8.2 Let 0 → V1 → · · · P


→ Vn → 0 be an exact sequence of finite–
dimensional vector spaces. Then (−1)i dim(Vi ) = 0.

Proof. Let φi : Vi → Vi+1 . Then dim(Vi ) = dim Im φi + dim Ker φi , while


exactness gives dim Im φi = dim Ker φi+1 . Thus
X X
(−1)i dim Vi = (−1)i (dim Ker φi + dim Ker φi+1 ) = 0.

Corollary 8.3 If 0 → A → B → C → 0 is an exact sequence of sheaves,


and two of the three have well-defined Euler characteristics, then so does the
third, and we have:
χ(B) = χ(A) + χ(C).

Proof. Apply the previous result to the long exact sequence in cohomology.

72
Skyscrapers. The skyscraper sheaf CP is given by O(U ) = C if P ∈ U ,
and O(U ) = 0 otherwise. For any divisor D, we have the exact sequence:

0 → OD → OD+P → CP → 0, (8.1)

where the final map records the leading coefficient of the polar part of f at
P . It is easy to see (e.g. by taking fine enough coverings, without P in any
multiple intersections):

Theorem 8.4 We have H p (X, CP ) = 0 for all p > 0. In particular χ(CP ) =


h0 (CP ) = 1.

Theorem 8.5 The cohomology groups H p (X, OD ) are finite-dimensional


for p = 0, 1 and vanish for all p ≥ 2.

Proof. We have already seen the result is true for D = 0: the space
H 1 (X, O) ∼
= Ω(X)∗ is finite-dimensional, and using the Dolbeault sequence,
one can show H p (X, O) = 0 for all p ≥ 2. The result for general D then
follows induction, using the skyscraper sheaf.

Proof of Riemann-Roch. Since h1 (O) = dim Ω(X) = ga , the formula is


correct for the trivial divisor. Using (8.1) and the additivity of the Euler
characteristic for short exact sequences, we find:

χ(OD+P ) = χ(OD ) + χ(CP ) = χ(OD ) + 1,

which implies Riemann-Roch for an arbitrary divisor D.

Existence of meromorphic functions and forms. A more general form


of Serre duality will lead to a more useful formulation of Riemann-Roch, but
we can already deduce several useful consequences.

Theorem 8.6 Any compact Riemann surface admits a nonconstant map


f : X → P1 with deg(f ) ≤ ga + 1.

Proof. We have dim H 0 (X, OD ) ≥ deg D − ga + 1. Once deg D > ga this


gives dim H 0 (OD ) > 1 = dim C.

73
(Note: this bound on deg(f ) is not optimal. The optimal bound, coming
from Brill–Noether theory, is given by (g +3)/2 rounded down. We will later
see this bound explicitly for genus g ≤ 3.)

Corollary 8.7 Any surface with ga = 0 is isomorphic to P1 .

Proof. It admits a map to P1 of degree ga + 1 = 1.

Corollary 8.8 Any compact Riemann carries a nonzero meromorphic 1-


form.

Proof. Take ω = df .

Corollary 8.9 Canonical divisors K = (ω) exist, and satisfy OK ∼


= Ω.

The isomorphism is given by f 7→ f ω.


Arithmetic and topological genus; degree of canonical divisors;
residues.

Corollary 8.10 The degree of any canonical divisor is 2g − 2, where g is


the topological genus of X.

Proof. Apply Riemann-Hurwitz to compute deg(df ).

Corollary 8.11 The topological and arithmetic genus of X agree: we have


g = ga ; and dim H 1 (X, Ω) = 1.

Proof. Apply Riemann-Roch to a canonical divisor K: we get

h0 (K) − h1 (K) = ga − h1 (K) = 1 − ga + deg(K) = 2g − ga − 1.

Using the fact that h0 (K) = ga , we can solve for h1 (K):

h1 (K) = 2(ga − g) + 1.

Now we know ga − g ≤ 0 and h1 (K) ≥ 1, so we must have ga = g and


h1 (K) = 1.

74
Theorem 8.12 (Hodge theorem) On a compact Riemann surface every
1 (X, C) is represented by a harmonic 1-form. More precisely we
class in HDR
have
1
HDR (X) = Ω(X) ⊕ Ω(X) = H 1,0 (X) ⊕ H 0,1 (X) = H1 (X).

Proof. We already know the harmonic forms inject into deRham cohomol-
ogy, by considering their periods; since the topological and arithmetic genus
agree, they also surject.

The space of smooth 1-forms. On a compact Riemann surface, the full


Hodge theorem

E 1 (X) = dE 0 (X) ⊕ H1 (X) ⊕ d∗ E 2 (X)

becomes the statement:

E 1 (X) = (∂ + ∂)E 0 (X) ⊕ (Ω(X) ⊕ Ω(X)) ⊕ (∂ − ∂)E 0 (X).

We have now proved this statement. Indeed, it suffices to show ∂E 0 (X) ⊕


∂E 0 (X) spans the complement of the harmonic forms. But the isomorphism
H 0,1 (X) ∼
= Ω(X)∗ ∼ = Ω(X) gives E 0,1 ∼
= Ω(X) ⊕ ∂E 0 (X), and similarly for
E 0,1 .

Theorem
R 8.13 A smooth (1, 1) form α lies in the image of the Laplacian
iff X α = 0.

Proof. We must show α = ∂∂f . Since the residue map on H 1,1 (X) is
an isomorphism, we can write α = ∂β where β is a (1, 0)-form. Since
Ω(X) ∼= H 0,1 (X), we can find a holomorphic 1-form ω such that β − ω = ∂f
for some smooth function f . Then ∂f = β − ω, and hence ∂∂f = ∂β = α.

Remark: isothermal coordinates. The argument we have just given also


proves the Hodge theorem for any compact, oriented Riemannian 2-manifold
(X, g). To see this, however, we need to know that every Riemannian metric
is locally conformally flat; i.e. that one can introduce ‘isothermal coordi-
nates’ to make X into a Riemann surface with g a conformal metric.

75
9 The Mittag–Leffler problems
In this section we show that the residue theorem is the only obstruction to
the construction of a meromorphic 1-form with prescribed principal parts.
The residue map on H 1 (X, Ω) will also play a useful role in the discussion
of Serre duality.
We then solve the traditional Mittag-Leffler theorem for meromorphic
functions, and use it to prove the Riemann–Roch theorem for effective divi-
sors.
Mittag–Leffler for 1-forms. The Mittag-Leffler problem for 1-forms is to
construct a meromorphic 1-form ω ∈ M1 (X) with prescribed polar parts.
The input data is specified by meromorphic forms

βi ∈ M1 (Ui )

such that
αij = βi − βj ∈ Ω(Ui )
for all i. A solution to the Mittag-Leffler problem is given by a meromorphic
form ω with
ω = βi + γi , γi ∈ Ω(Ui ).
This just says that ω is a global meromorphic form whose principal parts
are those given by βi .
In terms of sheaves, we have

0 → Ω → M1 → M1 /Ω → 0,

where as before the quotient on the right must be converted to a sheaf


in the natural way. The Mittag-Leffler data (βi ) determines an element
of H 0 (M1 /Ω), and it has a solution if this element is in the image of
M1 )(X). Thus the obstruct to finding a solution is given by the Mittag–
Leffler coboundary
[αij ] ∈ H 1 (Ω).

The residue map. Now recall that the exact sequence


0 → Ω → E 1,0 → E 1,1 → 0

determines a natural isomorphism

H 1 (X, Ω) ∼
= H 1,1 (X) ∼
= C,

76
and that the residue map, defined by
Z
1
Res(ω) = ω,
2πi X

determines a natural isomorphism

H 1,1 (X) ∼
= C.

The terminology is justified by the following result:

Theorem 9.1 On a Mittag-Leffler coboundary, αij = βi − βj , we have


X
Res([αij ]) = Resp (βi ).
p

Here we choose any Ui such that p ∈ Ui to compute the residue.


Proof. We must find explicitly the form ω ∈ E 1,1 representing the cocycle
αij . To do this we first write

αij = ξi − ξj

for smooth (1, 0) forms ξi . Then we observe that, since αij is holomorphic,
the (1, 1)-forms ∂ξi fits together over patches to give a smooth (1, 1)-form
ω. Now we also have αij = βi − βj , and thus

ηi = βi − ξi = βj − ξj

on Uij . So these forms piece together to give a global form η such that

∂η = −∂ξi = −ω.

The only nuance is that η ∈ E 1,0 + M1 , i.e. it is locally the sum of a smooth
form and a form with poles. However, η is smooth on X ∗ , the result of
puncturing X at all the poles of the data (βi ). If we replace X ∗ by a surface
with boundary circles around each of these poles, then we get
Z Z Z X
2πi Res([αij ]) = ω≈ −∂η = η≈ 2πi Resp (βi ).
X X∗ −∂X ∗

Letting the boundary of X ∗ shrink to zero yields the Theorem.

77
Corollary 9.2 There exists a meromorphic 1-form with prescribed principal
parts iff the sum of its residues is equal to zero.

Corollary 9.3 Given any pair of distinct points p1 , p2 ∈ X, there is a mero-


morphic 1-form ω with simple poles of residues (−1)i at pi and no other
singularities.

This ‘elementary differential of the third kind’ is unique up to the addi-


tion of a global holomorphic differential. Example: the form dz/z works for
0, ∞ ∈ C.
b

Corollary 9.4 For any p ∈ X and n ≥ 2 there exists a meromorphic 1-


form ω with a pole of order n at p (but vanishing residue) and no other
singularities.

This is an ‘elementary differential of the second kind’.


Residues and currents. Underlying the calculation above (and the ∂
equation) is a simple calculation the distributional derivative on forms:
Z   Z
dz dz
∂ = = 2πi.
∆ z S z
1

Using distributions, we get a more direct proof of the calculation above.


Namely, we have an exact sequence


0 → Ω → D1,0 → D1,1 → 0,

and then if αij = βi −βj , the distributions ηi = ∂βi already agree on overlaps
and give a global (1, 1)-current η, satisfying, by the calculation above,
Z
1 X
Res(η) = η= Resp (βi ).
2πi X

Mittag-Leffler for functions. Given a finite set of points pi ∈ X, and


the Laurent tails
bn b1
fi (z) = n + · · · +
z z
of meromorphic functions fi in local coordinates near Pi , when can we find
a global meromorphic function f on X with the given principal parts?
In terms of sheaves, we are now studying the short exact sequence

0 → O → M → M/O → 0.

78
Given given fi ∈ M(Ui ), we want to determine when [fi − fj ] ∈ H 1 (X, O)
is a coboundary.
Now by the case of Serre duality we have already proven, H 1 (X, O) is
isomorphic to Ω(X)∗ , so there is a natural pairing between (δfi ) ∈ H 1 (X, O)
and ω ∈ Ω(X). In fact, by Theorem 9.1, this pairing is given by

hδfi , ωi = Res(ωfi ).

A class in H 1 (X, O) vanishes iff it pairs trivially with all ω ∈ Ω(X). Thus
we have:

Theorem 9.5 The Mittag-Leffler problem specified by (fi ) has a solution


iff X
Resp (fi ω) = 0
for every ω ∈ Ω(X).

Geometric Riemann–Roch. Using the solution to the Mittag–Leffler


problem, we get a geometric and natural version of the Riemann-Roch the-
orem for effective divisors. We will soon see that the same result holds
without the assumption that D ≥ 0.

Theorem 9.6 Suppose D ≥ 0. Then

h0 (D) = 1 − g + deg(D) + h0 (K − D).

Proof. Let T ∼ = Cn , n = deg D, be the vector space of all Laurent tails


associated to D. (More functorially, T is the space of global sections of
OD / O; the latter is a generalized skyscraper sheaf, with the same support
as D.) Taking into account the constant functions, we then have

H 0 (X, OD )/C ∼
= Ω(X)⊥ ⊂ T,

using the residue pairing above. But the map Ω(X) → T ∗ has a kernel,
namely ΩD (X), the space of holomorphic 1-forms that vanish to such higher
order that they automatically cancel all the principal parts in T . This shows

h0 (D) − 1 = dim Ω(X)⊥ = dim T − dim Ω(X) + dim ΩD (X)


= deg D − g + h0 (K − D),

which gives the formula above.

79
Remarks. The proof is quite effective; e.g. if we know the principal parts
of a basis for Ω(X) at P ∈ X, then we can compute explicitly the Laurent
tails of all meromorphic function in H 0 (X, OnP ).
The computation above can be summarized more functorially: for D ≥ 0
we have a short exact sequence of sheaves,
0 → O → OD → OD /O → 0,
which gives rise to a long exact sequence, using duality:
0 → C → H 0 (OD ) → H 0 (OD /O) → H 1 (O) ∼
= Ω(X)∗ → Ω−D (X)∗ → 0.
Note that H 0 (OD /O) is exactly the space of principal parts supported on
D; its dimension is deg D.
To carry out this computation we needed the inclusion OD ⊂ O, which
is equivalent to the condition that D ≥ 0. We remark that effective divisors
are ‘infinitely rare’ among all divisors.

10 Serre duality
The goal of this section is to present:

Theorem 10.1 (Serre duality) For any divisor D, we have a canonical


isomorphism
H 1 (X, O−D )∗ ∼
= ΩD (X).

The natural isomorphism comes from the pairing


H 0 (X, ΩD ) ⊗ H 1 (X, O−D ) → H 1 (X, Ω) ∼
= C.
This result allows us to eliminate h1 entirely from the statement of
Riemann-Roch, without the requirement that D is effective, and obtain:

Theorem 10.2 (Riemann-Roch, final version) For any divisor D on a


compact Riemann surface X, we have
h0 (D) = deg D − g + 1 + h0 (K − D).

Proof of Serre duality: analysis. A proof of Serre duality can be given


using the elliptic regularity, just as we did for the case Ω(X) ∼
= H 1 (X, O)∗ .
For this one should either consider forms that are locally (smooth) + (mero-
morphic), or consider smooth sections of the line bundle defined by a divisor,
which we will elaborate later.

80
Proof of Serre duality: dimension counts. We first show

Ω−D (X) → H 1 (X, OD )∗

is injective. For this, given ω ∈ Ω−D (X) choose a small neighborhood U0


centered at point P with U0 outside the support of D and (ω), on which we
have a local coordinate where ω = dz. Then set f0 = 1/z on U0 and f1 = 0
on Then set U1 = X − {P }. We then find δfi ∈ H 1 (X, OD )∗ , and

Res(fi ω) = 1,

so ω is nontrivial in the dual space.


This shows h0 (K − D) ≤ h1 (D), or h0 (K − D) − h1 (D) ≤ 0. Now we
sum the following two applications of Riemann-Roch:

h0 (K − D) − h1 (K − D) = 1 − g + deg(K − D)
h0 (D) − h1 (D) = 1 − g + deg(D)

to get
(h0 (D) − h1 (K − D)) + (h0 (K − D) − h1 (D)) = 0
(using the fact that deg(K) = 2g − 2). Since both terms on the left are ≤ 0,
they must vanish, and this gives Serre duality. (I am grateful to Levent

Alpoge for pointing out this last step in the argument.)


Remark on injectivity. It seems strange at first sight that, in our proof
that
Ω−D (X) ,→ H 1 (X, OD )∗ ,
almost no use was made of the divisor D. Why doesn’t the same argument
prove that Ω−E (X) ,→ H 1 (X, OD )∗ for any D and E?
The point is that we need the inclusion of Ω−D (X) into H 1 (X, OD ) to
be defined, and for this it is essential that ω pairs trivially with coboundaries
in the target. To get this fact we are really using D, so that the product
goes into H 1 (X, Ω) ∼
= C.
Proof of Serre duality: using just finiteness. Finally, here is an
argument that just uses finiteness of h1 (O). In particular it shows that
finiteness implies H 1 (X, O) ∼
= Ω(X)∗ . It is based loosely on Forster.
We begin with some useful qualitative dimension counts that follow im-
mediately from the fact that h0 (D) = 0 if deg(D) < 0, h1 (D) ≥ 0 and the
fact that Ω ∼= OK .

81
Theorem 10.3 For deg(D) > 0, we have

dim OD (X) ≥ deg(D) + O(1),


dim ΩD (X) ≥ deg(D) + O(1), and
1
dim H (X, O−D ) = deg(D) + O(1).

Pairings. Next we couple the product map

O−D ⊗ ΩD → Ω

together with the residue map Res : H 1 (X, Ω) → C to obtain a map

H 1 (X, O−D ) ⊗ ΩD (X) → H 1 (X, Ω) → C,

or equivalently a natural map

ΩD (X) → H 1 (X, O−D )∗ .

This map explicitly sends ω to the linear functional defined by

φ(ξ) = Res(ξω).

Using the long exact sequence associated to equation (8.1) we obtain:

Theorem 10.4 The inclusion OD → OD+P induces a surjection:

H 1 (X, OD ) → H 1 (X, OD+P ) → 0.

Corollary 10.5 We have a natural surjective map:

H 1 (X, OD ) → H 1 (X, OE ) → 0

for any E ≥ D.

Put differently, for E ≥ D we get a surjective map H 1 (X, O−E ) →


H 1 (X, O−D ) and thus an injective map on the level of duals. For organi-
zational convenience we take the direct limit over increasing divisors and
set
V (X) = lim H 1 (X, O−D )∗ .
D→+∞

Clearly ΩD (X) maps into V (X) for every D, so we get a map M1 (X) →
V (X).

82
Theorem 10.6 The natural map M1 (X) → V (X) is injective. Moreover
a meromorphic form ω maps into H 1 (X, O−D )∗ iff ω ∈ ΩD (X).

Proof. We have already shown injectivity in the previous proof. Now for
the second statement. If ω is in H 1 (X, O−D )∗ , then it must vanish on all
coboundaries for this group. Suppose however −(k+1) = ordP (ω) < −D(P )
for some P . Then k ≥ D(P ), so in the construction above we can arrange
that ξ = 0 in H 1 (X, O−D ). This is a contradiction. Thus ordP (ω) ≥ −D(P )
for all P , i.e. ω ∈ ΩD (X).

Completion of the proof of Serre duality. Note that both M1 (X)


and V (X) are vector spaces over the field of meromorphic functions M(X).
(Indeed the former vector space is one-dimensional, generated by any mero-
morphic 1-form.)
Given φ ∈ H 1 (X, O−D )∗ ⊂ V (X), we must show φ is represented by a
meromorphic 1-form ω. (By the preceding result, ω will automatically lie in
ΩD (X).)
The proof will be by a dimension count. We note that for n  0, we
have
dim H 1 (X, O−D−nP )∗ = n + O(1).
On the other hand, this space contains ΩD+nP (X) as well as OnP (X) · φ.
Both of these spaces have dimension bounded below by n+O(1). Thus for n
large enough, they meet in a nontrivial subspace. This means we can write
f φ = ω where f is a nonzero meromorphic function and ω is a meromorphic
form. But then φ = ω/f is also a meromorphic form, so we are done!

We can now round out the discussion by proving some results promised
above.

Theorem 10.7 For any divisor D we have

H 0 (X, OD ) ∼
= H 1 (X, Ω−D )∗ .

Proof. We have

H 0 (X, OD ) ∼
= H 0 (X, ΩD−K ) ∼
= H 1 (X, OK−D )∗ ∼
= H 1 (X, Ω−D )∗

83
Corollary 10.8 We have H 1 (X, OD ) = 0 as soon as deg(D) > deg(K) =
2g − 2.

Proof. Because then H 1 (X, OD )∗ ∼


= Ω−D (X) = 0.

Corollary 10.9 If deg(D) > 2g − 2, then h0 (D) = 1 − g + deg(D).

Corollary 10.10 We have H 1 (X, M) = H 1 (X, M1 ) = 0.

Proof. Any representative cocycle (fij ) for a class in H 1 (X, M) can be re-
garded as a class in H 1 (X, OD ) for some D of large degree. But H 1 (X, OD ) =
0 once deg(D) is sufficiently large. Thus (fij ) splits for the sheaf OD , and
hence for M.

Mittag-Leffler for 1-forms, revisited. Here is another proof of the


Mittag-Leffler theorem for 1-forms. Suppose we consider all possibleP prin-
cipal parts with poles of order at most ni > 0 at Pi , and letPD = ni Pi .
The dimension of the space of principal parts is then n = ni = deg D.
In addition, the principal part determines the solution to the Mittag-Leffler
problem up to adding a holomorphic 1-form. That is, the solutions lie in the
space ΩD (X), and the map to the principal parts has Ω(X) as its kernel.
Thus the dimension of the space of principal parts that have solutions
is:
k = dim ΩD (X) − dim Ω(X).
But by Riemann-Roch we have

dim ΩD (X) = h0 (K + D) = h0 (−D) + deg(K + D) − g + 1


= 2g − 2 + deg D − g − 1 = g + deg D − 1 = g + n − 1,

so the solvable data has dimension k = n − 1. Thus there is one condition on


the principal parts for solvability, and that condition is given by the residue
theorem.

11 Maps to projective space


In this section we explain the connection between the sheaves OD , linear
systems and maps to projective space. In the next section we relate this
classical theory to the modern concept of a holomorphic line bundle. Useful
references for this material include [GH, Ch. 1].

84
Historically, algebraic curves in projective space were studied from an
extrinsic point of view, i.e. as embedded subvarieties. From this perspec-
tive, it is not at all clear when the same intrinsic curve is being presented in
two different ways. The modern perspective is to fix the intrinsic curve X
— which we will consider as a Riemann surface — and then study its var-
ious projective manifestations. The embedding of X into Pn then becomes
visible as the family of effective divisors — the linear system — coming from
intersections of X with hyperplanes in Pn .
Projective space. Let V be an (n + 1)-dimensional vector space over C.
The space of lines (one-dimensional subspaces) in V forms the projective
space
PV = (V − {0})/C∗ .
It has the structure of a complex n-manifold. The subspaces S ⊂ V give
rise to planes PS ⊂ PV ; when S has codimension one, PS is a hyperplane.
The dual projective space PV ∗ parameterizes the hyperplanes in PV , via
the correspondence φ ∈ V ∗ 7→ Ker(φ) = S ⊂ V .
We can also form the quotient space W = V /S. Any line L in V that
is not entirely contained in S projects to a line in W . Thus we obtain a
natural map
π : (PV − PS) → P(V /S).
All the analytic automorphisms of projective space come from linear
automorphisms of the underlying vector space: that is,

Aut(PV ) = GL(V )/C∗ = PGL(V ).

We let Pn = PCn+1 with homogeneous coordinates [Z] = [Z0 : · · · : Zn ]. It


satisfies
Aut(Pn ) = PGLn+1 (C).
The case n = 1 gives the usual identification of automorphisms of Cb with
Möbius transformations.
The Hopf fibration. By considering the unit sphere in Cn+1 , we obtain the
Hopf fibration π : S 2n+1 → CPn with fibers S 1 . This shows that projective
space is compact. Moreover, for n = 1 the fibers of π are linked circles in
S 3 , and π generates π3 (S 2 ) ∼
= Z.
Metrics on projective space. The compact group SUn+1 also acts tran-
sitively on projective space and has a unique invariant metric, up to scale,
called the Fubini–Study metric. In this metric all linear subvarieties Pk ⊂ Pn
have the same (2k)–dimensional volume. In fact SUn+1 acts transitively on
these subvarieties.

85
More remarkably, the volume of any irreducible subvariety of Pn is deter-
mined by its dimension and its degree. In particular, all curves of degree d
in P2 have the same volume. This is a basic property of Kähler manifolds; it
comes from the fact that the symplectic form associated to the Fubini–Study
metric is closed.
Projective varieties. The zero set of a homogeneous polynomial f (Z)
defines an algebraic hypersurface V (f ) ⊂ Pn . A projective algebraic variety
is the locus V (f1 , . . . , fn ) obtained as an intersection of hypersurfaces.
Affine charts and cohomology. The locus An = Pn − V (Z0 ) is isomor-
phic to Cn with coordinates (z1 , . . . , zn ) = Zi /Z0 , while V (Z0 ) itself is a
hyperplane; thus we have

Pn ∼
= Cn ∪ Pn−1 .

By permuting the coordinates, we get a covering of Pn by n + 1 affine charts.


It is easy to see that the transition functions are algebraic. Thus Pn is a
complex manifold.
These affine charts shows that as a topological space, Pn is obtained by
adding a single cell in each even dimension 0, 1, . . . , 2n, and thus:

H 2k (Pn , Z) = Z for k = 0, 1, . . . , n;

moreover [Pk ] gives a canonical generator in dimension 2k. The remaining


cohomology groups are zero.
The degree of a subvariety. If Z k ⊂ Pn is a variety of dimension k, then
Z k is a topological cycle and its degree is defined so that:

[Z k ] = deg(Z) · [Pk ] ⊂ H 2k (Pn , Z).

Alternatively, deg(Z k ) = |Pn−k ∩ Z| is the number of intersections with a


generic linear subspace of complementary dimension. In particular, if Z =
V (f ) is a hypersurface and f is separable (a product of distinct irreducibles),
then
deg(Z) = deg(f ).

The space of hypersurfaces. The space of homogeneous polynomials of


degree d in Pn has dimension given by
 
d+n
dim Pd (Cn+1 ) = ·
n

86
This can be seen by inserting n markers into a list of d symbols, and turning
all the symbols up to the first marker into Z0 ’s, then the next stretch into
Z1 ’s, etc.
Thus the space of curves of degree d in P2 is itself a projective space
N
P , with N = d(d + 3)/2. E.g. there is a 5-dimensional space of conics, a
9-dimensional space of cubics and a 14-dimensional space of quartics. Note
that these spaces include reducible elements, e.g. the union of d lines is an
example of a (reducible) plane curve of degree d.
Meromorphic (rational) functions on Pn . The ratio

F (Z) = f1 (Z)/f0 (Z) = [f0 (Z) : f1 (Z)]

of two homogeneous polynomials of the same degree defines a meromorphic


‘function’
F : Pn 99K P1 .
Its values are undetermined on the subvariety V (f0 , f1 ), which has codimen-
sion two if f0 and f1 are relatively prime. Away from this subvariety, F (Z)
gives a holomorphic map to P1 . The field of all rational functions is

K(Pn ) ∼
= C(z1 , . . . , zn ).

Projectivization. Any ordinary polynomial p(zi ) has a unique homo-


geneous version P (Zi ) of the same maximal degree, such that p(zi ) =
P (1, z1 , . . . , zn ). Thus any affine variety V (p1 , . . . , pm ) has a natural com-
pletion V (P1 , . . . , Pm ) ⊂ Pm . This variety is smooth if it is a smooth sub-
manifold in each affine chart.
Examples: Curves in P2 .
1. The affine curve x2 − y 2 = 1 meets the line at infinity in two points
corresponding to its two asymptotes; its homogenization X 2 −Y 2 = Z 2
is smooth in every chart. In (y, z) coordinates it becomes 1 − y 2 = z 2 ,
which means the line at infinity (z = 0) in two points.

2. The affine curve defined by p(x, y) = y − x3 = 0 is smooth in C2 , but


its homogenization P (X, Y, Z) = Y Z 2 − X 3 defines the curve z 2 = x3
in the affine chart where Y 6= 0, which has a cusp.

The normalization of a singular curve. Every irreducible homogeneous


polynomial f on P2 determines a compact Riemann surface X together with
a generically injective map ν : X → V (f ). The Riemann surface X is called
the normalization of the (possibly singular) curve V (f ).

87
To construct X, use projection from a typical point P ∈ P2 − V (f ) to
obtain a surjective map π : V (f ) → P1 . After deleting a finite set from
domain and range, including all the singular points of V (f ), we obtain an
open Riemann surface X ∗ = V (f ) − C and a degree d covering map

π : X ∗ → (P1 − B).

As we have seen, there is a canonical way to complete X ∗ and π to a compact


Riemann surface X and a branched covering π : X → P1 . It is then easy to
see that the isomorphism

ν : X ∗ → V (f ) − C ⊂ P2

extends to a holomorphic map ν : X → P2 with image V (f ).


Examples: cusps and nodes. The cuspidal cubic y 2 = x3 is normalized
by ν : P1 → P2 given by ν(t) = (t2 , t3 ).
The nodal cubic y 2 = x2 (1−x) is also interesting to uniformize. The idea
is to project from the node at (x, y) = (0, 0), i.e. let us use the parameter
t = y/x or equivalently set y = tx. Then we find

t2 x2 = x2 (1 − x)

and so (x, y) = (1 − t2 , t(1 − t2 ) give a parameterization of this curve.


Maps to Pn . Now let X be a compact Riemann surface. Generalizing
the study of meromorphic functions f : X → P1 , we wish to describe all
holomorphic maps
φ : X → Pn .
There are at least two ways to approach the description of φ: using M(X),
and using Div(X).
Maps from meromorphic functions. Let S ⊂ M(X) be a linear sub-
space of meromorphic functions of dimension n + 1. We have a natural
map
φ : X → PS ∗
given by φ(x)(f ) = f (x). More concretely, if we choose a basis for S then
the map is given by
φ(x) = [f0 (x) : · · · : fn (x)].
The linear functional x 7→ f (x) in S ∗ makes sense at most points of X; the
only potential problems come from zeros and poles. To handle these, we
simply choose a local coordinate z near x with z(x) = 0, and define

φ(x)(f ) = (z p f (z))|z=0 ,

88
where p is chosen so all values are finite and so at least one value is nonzero.
Note that the map φ is nondegenerate: its image is contained in no
hyperplane.
Note also that if we replace S with f S, where f ∈ M∗ (X), then the map
φ remains the same. Thus we can always normalize so that 1 ∈ S.
Linear systems of divisors. We now turn to the approach via divisors.
Recall that divisors D, E are linearly equivalent if D − E = (f ) is principal;
then OD ∼ = OE .
The complete linear system |D| determined by D is the collection of
effective divisors E linearly equivalent to D. These are exactly the divisors
of the form E = (f ) + D, where f ∈ H 0 (X, OD ). The divisor only depends
on the line determined by f . We have a natural bijection

|D| ∼
= POD (X).

In particular, we have
dim |D| = h0 (D) − 1.
Technically |D| is a subset of the space of all divisors on X. Equivalent
divisors determine the same linear system, i.e.

|E| = |D| ⊂ Div+ ∼ (d) .


d (X) = X

We emphasize that all divisors in |D| are effective and of the same degree.
A general linear system is simply a subspace L ∼ = Pr ⊂ |D|. It corre-
sponds to a subspace S ⊂ H 0 (X, OD ).
Base locus and geometric linear systems. The base locus of a linear
system is the largest divisor B ≥ 0 such that E ≥ B for all E ∈ L. A linear
system is basepoint free if B = 0.
Let X ⊂ Pn be a smooth nondegenerate curve, and let H be a hyper-
plane. Then D = H ∩ X is an effective divisor called a hyperplane section.
Any two such divisors are linearly equivalent, because H1 − H2 = (f ) where
f : Pn 99K P1 is a rational map of degree one (a ratio of linear forms), and
f |X is a meromorphic function.
Thus the set of hyperplane sections determines a linear system L ⊂ |D|.
It is basepoint free because for any P ∈ X there is an H disjoint from P .
On the other hand, if we take only those hyperplanes passing through
P ∈ X, then we get a smaller linear system whose base locus is P itself.
Different perspectives on maps to projective space. The following
data, up to suitable equivalences, are in natural bijection:

89
1. Nondegenerate holomorphic maps φ : X → Pn of degree d; up to
Aut(Pn );

2. Subspaces S ⊂ M(X) of dimension (n + 1) with deg(f /g) = d for


generic f, g ∈ S; up to S 7→ f S;

3. Basepoint-free linear systems L ⊂ Div(X) of degree d and dimension


n.

Here are some of the relations between them.


1 =⇒ 2. Let S be spanned by the meromorphic function fi = Zi /Z0 ,
restricted to X. Note that f /g is the projection of φ(X) to P1 from a generic
Pn−2 , which will have degree d so long as the Pn−2 is disjoint from φ(X).
1 =⇒ 3. We let D = X ∩ H and let L ⊂ |D| be the collection of all
hyperplane sections.
2 =⇒ 1. We have seen that S canonically determines a map to PS ∗ ,
and clearly f S determines an equivalent map.
2 =⇒ 3. Let D be the smallest divisor such that D + (f ) is effective
for all (nonzero) f ∈ S. This is given, concretely, by D(x) = − inf S ordf (x).
Then we consider the linear system of divisors of the form E = (f ) + D such
that f ∈ S.
3 =⇒ 2. Picking any D ∈ L, we let S = {f : (f ) = E − D, E ∈ L}.
3 =⇒ 1. For each x ∈ X we get a hyperplane in L by considering the
set of E containing x. This gives a natural map X → Pn ∼ = L∗ .
Embeddings and smoothness for complete linear systems. We now
study the complete linear system |D|, and the associated map

φD : X → PH 0 (X, OD )∗

in more detail.
Recall that OD is a O-module. We say OD is generated by global sections
if for each x ∈ X we have a global section f ∈ H 0 (X, OD ) such that the
stalk OD,x , which is an Ox -module, is generated by f ; that is, OD,x = Ox ·f .

Theorem 11.1 The following are equivalent.

1. |D| is basepoint–free.

2. h0 (D − P ) = h0 (D) − 1 for all P ∈ X.

3. OD is generated by global sections.

4. For any x ∈ X there is an f ∈ OD (X) such that ordx (f ) = −D(x).

90
Example. Note that h0 (nP ) = 1 for n = 0 and = g for n = 2g − 1. Thus
(for large genus) there must be values of n such that h0 (nP ) = h0 (nP − P ).
In this case, D = nP is not globally generated.

Theorem 11.2 Let φ : X → Pn be a map of X to projective space with


the image not contained in a hyperplane. Then φ = π ◦ φD , where D is the
divisor of a hyperplane section, |D| is a base-point free linear system, and
where π : PN 99K Pn is projection from a linear subspace PS disjoint from
φD (X).

Proof. Since hyperplanes can be moved, the linear system |D| is base-point
free, and all hyperplane sections are linearly equivalent to D. Thus φ can
be regarded as the natural map from X to PS ∗ , where S is a subspace of
H 0 (X, OD ), so φ can be factored through the map φD to PH 0 (X, OD )∗ .

Theorem 11.3 Let |D| be base-point free. Then for any effective divisor
P = P1 + . . . + Pn on X, the linear system

P + |D − P | ⊂ |D|

consists exactly of the hyperplane sections E = φ−1 D (H) passing through


(P1 , . . . , Pn ). In particular, the dimension of the space of such hyperplanes
is
dim |D − P | = h0 (D − P ) − 1.

Proof. A hyperplane section E ∈ |D| passes through P iff E − P ≥ 0 iff


E = P + E 0 where E 0 ∈ |D − P |.

Theorem 11.4 If h0 (D − P − Q) = h0 (D) − 2 for any P, Q ∈ X, then |D|


provides a smooth embedding of X into projective space.

Proof. This condition says exactly that the set of hyperplanes passing
through φD (P ) and φD (Q) has dimension 2 less than the set of all hyper-
planes. Thus φD (P ) 6= φD (Q), so φD is 1 − 1. The condition on φ(D − 2P )
says that the set of hyperplanes containing φD (P ) and φ0D (P ) is also 2
dimensions less, and thus φD is an immersion. Thus φD is a smooth embed-
ding.

91
Theorem 11.5 The linear system |D| is base-point free if deg D ≥ 2g, and
gives an embedding into projective space if deg D ≥ 2g + 1.

Proof. Use the fact that if deg D > deg K = 2g − 2, then we have h0 (D) =
deg D − g + 1, which is linear in the degree.

Corollary 11.6 Every compact Riemann surface embeds in projective space.

Remark. By projecting we get an embedding of X into P3 and an immer-


sion into P2 .
Not every Riemann surface can be embedded into the plane! In fact a
smooth curve of degree d has genus g = (d − 1)(d − 2)/2, so for example
there are no curves of genus 2 embedded in P2 .
Examples of linear systems.

1. Genus 0. On P1 , we have OD ∼ = OE if d = deg(D) = deg(E). This


sheaf is usually referred to as O(d); it is well-defined up to isomor-
phism, |D| consists of all effective divisors of degree d. Note that
Od∞ (P1 ) is the d + 1-dimensional space of polynomials of degree d.
The corresponding map φD : P1 → Pd is given in affine coordinates by
φD (t) = (t, t2 , . . . , td ). Its image is the rational normal curve of degree
d. Particular cases are smooth conics and the twisted cubic.

2. Genus 1. On X = C/Λ with P = 0, the linear system |P | is not base-


point free, but |2P | is, and |3P | gives an embedding into the plane,
via the map z 7→ (℘(z), ℘0 (z)).
P
Recall that
P D= mi Pi is a principal divisor on X iff deg(D) = 0 and
e(D) = mi Pi in C/Λ is zero. Thus 3Q ∈ |3P | iff e(3(P − Q)) = 0.
This shows:

A smooth cubic curve X in P2 has 9 flex points, correspond-


ing to the points of order 3 in the group law on X.

12 The canonical map


Note that the embedding of a curve X of genus 1 into P2 by the linear series
|3P | breaks the symmetry group of the curve: since Aut(X) acts transitively,
the 9 flexes of Y = φ3P (X) are not intrinsically special.

92
For genus g ≥ 2 on the other hand there is a more natural embedding
— one which does not break the symmetries of X — given by the canonical
linear system |K|.
The linear system |K| corresponds to the map

φK : X → PΩ(X)∗ ∼
= Pg−1

given by φ(x) = [ω1 (x), . . . , ωg (x)], where ωi is a basis for Ω(X).

Theorem 12.1 (1) The linear system |K| is base-point free for g > 0.
(2) For g ≥ 1, either |K| gives an embedding of X into Pg−1 , or X is
hyperelliptic.

Proof. Let Pk ≥ 0 be an effective divisor of degree k on X. Then

h0 (K − Pk ) = 1 − g + (2g − 2) − k + h0 (Pk ).

We have 1 ≤ h0 (Pk ) for all k. If equality holds then we have

h0 (K − Pk ) = h0 (K) − k,

which is what we want to get a basepoint free linear system (k = 1) and an


embedding (k = 2). Thus |K| is basepoint free iff h0 (P1 ) = 1 iff g > 0, and
|K| gives an embedding iff h0 (P2 ) = 1 iff X is not hyperelliptic.

The hyperelliptic case. Let us analyze the canonical map in the hyper-
elliptic case. Then there is a degree two holomorphic map f : X → P1
branched over the zeros of a polynomial p(z) of degree 2g + 2. A basis for
the holomorphic 1-forms on X is given by

z i dz
ωi = p
p(z)

for i = 0, . . . , g − 1. That is, ωi (x) = f (x)i ω0 . It follows that the canonical


map φ : X → Pg−1 is given by

f (x) = [ωi (x)] = [ω0 (x)f (x)i ] = [f (x)i ].

In other words, the canonical map factors as φ = ψ ◦ f , where ψ : P1 → Pg−1


is the rational normal curve of degree g − 1.

93
Geometry of canonical divisors. With the canonical map in hand, it
is easy to visualize at once all the effective canonical divisors K on X, and
hence all the elements of Ω(X). There are just two cases:

1. The curve X ⊂ Pg−1 is embedded by the canonical map as a curve of


degree 2g − 2. Then the hyperplane sections K = X ∩ H are exactly
the canonical divisors, i.e. these are all the elements of the complete
linear system |K|.

2. The canonical maps factors through degree two map π : X → P1 .


Then the canonical divisors are the preimages K = π −1 (E) of divisors
of degree g − 1 on P1 .

In the second, hyperelliptic case, the image of the canonical map is the
rational normal curve Z ⊂ Pg−1 , which has degree (g − 1), and the canonical
map is the composition
π
φ : X → P1 ∼
= Z ⊂ Pg−1 .

Note that any set of g − 1 points on Z span a unique hyperplane H, so even


in the hyperelliptic case we can regard the canonical divisors as simple the
divisors of the form φ−1 (H).
Pn
Geometric picture for h0 (K − D). Let D = 1 Pi be an effective
divisor on X. The canonical curve proves a clear picture of the ‘hard term’
h0 (K−D) in Riemann–Roch, in terms of the space of hyperplanes containing
all these points Pi ; namely,

h0 (K − D) − 1 = dim |K − D|
= dim{H ⊂ Pg−1 : φ(Pi ) ∈ H, 1 ≤ i ≤ n}.

Special divisors. An effective divisor D = Pi is special if h0 (K −D) > 0,


P
i.e. if there is a holomorphic 1-form ω 6= 0 vanishing at the points (Pi ), or
equivalent if the points (Pi ) lie on a hyperplane in Pg−1 .
Of course divisors of degree ≤ g − 1 are all special. The first interesting
case is degree g. There exist plenty of such divisors – note that |H ∩φ(X)| =
2g − 2, so a given hyperplane determines many such divisors. On the other
hand, g typical points on φ(X) do not span a hyperplane, so these divisors
really are special.

Theorem 12.2 An effective divisor D of degree g is special if and only if


there is a nonconstant meromorphic function on X with (f ) + D ≥ 0.

94
Proof. By Riemann-Roch, we have

h0 (D) = 1 − g + deg D + h0 (K − D) = 1 + h0 (K − D) > 1

iff h0 (K − D) > 0 iff D is special.

Since special advisors do exist, we have:

Corollary 12.3 If g ≥ 2 then X admits a meromorphic function of degree


≤ g.

Example: genus 3. A curve of genus 3 either admits a map to P1 of


degree two, or it embeds as a curve X ⊂ P2 of degree 4. In the latter case,
projection to P1 from any P ∈ X gives a map of degree g = 3.
Explicit construction of meromorphic functions. The map f : X →
P1 determined by a special divisor D of degree g can be constructed geo-
metrically as follows. Let H be a hyperplane containing D, and let H ∩ X =
D + E. Then deg(E) = g − 2, so E lies in linear subspace J ∼ = Pg−3 . Then
projection of P g−1 1 1
from J to P gives a map f : X → P with D as one of
its fibers.
Weierstrass points. Now we focus on divisors of the form D = gP . We
say P is a Weierstrass point if gP is special.

Proposition 12.4 The following are equivalent:

1. P is a Weierstrass point.

2. There is a hyperplane H such that H ∩φ(X) has multiplicity g at φ(P ).

3. There is an ω ∈ Ω(X) vanishing to order g at P .

4. There is a meromorphic function f on X with a pole just at P and


with 1 ≤ deg(f ) ≤ g.

and otherwise holomorphic.


Example: there are no Weierstrass points on a Riemann surface of genus
1. The branch points of every hyperelliptic surface of genus g ≥ 2 are
Weierstrass points.
The map f can be constructed as follow: choose a subspace J ∼= Pg−3 in
P g−1 passing through the g − 2 points of H ∩ φ(X) other than P . Then let
f be the projection of X from J to P1 , normalized so H ∩ X maps to ∞.
Then f has a pole of order ≥ g at P and nowhere else.

95
The Wronskian. To have H ∩ φ(X) = D = gP , the hyperplane H should
contain not just P but the appropriate set of tangent directions at P , namely

(φ(P ), φ0 (P ), φ00 (P ), . . . , φ(g−1) (P )).

For these tangent (g − 1) tangent directions to span a (g − 2)-dimensional


plane H through φ(P ), there must be a linear relation among them; that is,
the Wronskian determinant W (P ) must vanish.
In terms of a basis for Ω and a local coordinate z at P , the Wronskian
is given by  j 
d ωi
W (z) = det ,
dz j
where j = 0, . . . , g − 1 and i = 1, . . . , g.
We can see directly the vanishing of the Wronskian is equivalent to gP
being special.

Theorem 12.5 The Wronskian vanishes at P iff there is a holomorphic


1-form ω 6= 0 with a zero at P of order at least g.

Proof. The determinant vanishes iff there is a linear combination of the


basis elements ωi whose derivatives through order (g − 1) vanish at P .

The quantity W = W (z) dz N turns out to be independent of the choice


of coordinate, where N = 1 + 2 + . . . + g = g(g + 1)/2. This value of N
arises because the jth derivative of a 1-form behaves like dz j+1 .
Thus W (z) is a section of ON K , so its number of zeros is deg N K =
N (2g − 2) = (g − 1)g(g + 1). This shows:

Theorem 12.6 Any Riemann surface of genus g has (g − 1)g(g + 1) Weier-


strass points, counted with multiplicity.

Weierstrass points of a hyperelliptic curve. These correspond to the


branch points of the hyperelliptic map π : X → P1 , since the projective
normal curve P1 → Pg−1 has no flexes.
Canonical curves of genus two. We will now describe more geometrically
the canonical curves of genus two, three and four.

Theorem 12.7 Any Riemann surface X of genus 2 is hyperelliptic, and any


degree two map of X to P1 agrees with the canonical map (up to Aut P1 ).

96
Proof. In genus 2, we have deg K = 2g − 2 = 2, so the canonical map
φ : X → Pg−1 = P1 already presents X as a hyperelliptic curve. If f :
X → P1 is another such map, with polar divisor P + Q, then we have
h0 (P + Q) = 2 = h0 (K − P − Q) + 2 − 2 + 1; thus there exists an ω with
zeros just at P, Q and therefore P + Q is a canonical divisor.

Corollary 12.8 The moduli space of curves of genus two is isomorphic to


the 3-dimensional space M0,6 of isomorphism classes of unordered 6-tuples
of points on P1 . Thus M2 is finitely covered by C3 − D, where D consists
of the hyperplanes xi = 0, xi = 1 and xi = xj .

Remark. Here is a topological fact, related to the fact that every curve
of genus two is hyperelliptic: if S has genus two, then the center of the
mapping-class group Mod(S) is Z/2, generated by any hyperelliptic involu-
tion. (In higher genus the center of Mod(S) is trivial.)
Canonical curves of genus 3. Let X be a curve of genus 3. Then
X is either hyperelliptic, or its canonical map realizes it as a smooth plane
quartic. We will later see that, conversely, any smooth quartic is a canonical
curve (we already know it has genus 3). This shows:

Theorem 12.9 The moduli space of curves of genus 3 is the union of the
5-dimensional space M0,8 and the 6-dimensional moduli space of smooth
quartics, (P14 − D)/ PGL3 (C).

Note: a smooth quartic curve that degenerates to a hyperelliptic one be-


comes a double conic. The eight hyperelliptic branch points can be thought
of as the intersection of this conic with an infinitely near quartic curve.
Weierstrass points in genus 3. The Weierstrass points on a smooth
plane quartic correspond to flexes; there are 2 · 3 · 4 = 24 of them in general.
At the flexes we have h0 (3P ) = 2. How can one go from a flex P to a
degree 3 branched covering f : X → P1 ? We can try projection fP from P ,
but in general the line L tangent to X at P will meet X in a fourth point
Q. Thus fP will have a double pole at P and a simple pole at Q.
Instead, we project from Q! Then the line L through P and Q has
multiplicity 3 at P , giving a triple order pole there.
The Fermat quartic. The Fermat quartic, defined by

X 4 + Y 4 + Z 4 = 0,

97
is somewhat similar; its symmetry group is PSL2 (Z/8), which has order 96;
the quotient is the (2, 3, 8) orbifold, of Euler characteristic −1/24, and this
curve is tiled by 12 octagons. In this case the full symmetry group is easily
visible: there is an S 3 coming from permutations of the coordinates,√and a
(Z/4)2 action coming from multiplication of X and Y by powers of −1.
On the Fermat curve, given in affine coordinates by x4 + y 4 = 1, there
are 12 flexes altogether, each of multiplicity 2. Of these, 8 lie in the affine
plane, and arise when one coordinate vanishes and the other is a 4th root of
unity.
These are examples of Weierstrass points of order two; a line tangent
to one of these flexes meets the curve in a unique point P . In this case,
projection from P itself really does give a rational map f : X → P1 with a
triple pole just at P .
Flexes of plane curves. In general, if C is defined by F (X, Y, Z) = 0,
then the flexes of C are the locus where both F and the Hessian H =
det(d2 F/dZi dZj ) of F vanish. For the Fermat curve, we have F (X, Y, Z) =
X 4 + Y 4 + Z 4 and H = 1728(XY Z)2 . On a smooth curve of degree d the
number of flexes is 3d(d − 2).
The Klein quartic. The most symmetric curve of genus 3 is the Klein
quartic, given by
X 3 Y + Y 3 Z + Z 3 X = 0.
Its symmetry group G ∼ = PSL2 (Z/7) has order 168 = 7 · 24 and gives as quo-
tient the (2, 3, 7)-orbifold. A subgroup of order 21 is easily visible in the form
above: we can cyclically permute the coordinates and also send (X, Y, Z) to
(ζ 2 X, ζY, ζ 4 Z) where ζ 7 = 1. More geometrically, the Klein quartic is tiled
by 24 regular 7-gons. It is easily shown that −1/42 = χ(S 2 (2, 3, 7)) is the
largest possible Euler characteristic for a hyperbolic orbifold, which implies:

Theorem 12.10 (Hurwitz) For any compact Riemann surface of genus g


we have | Aut(X)| ≤ 84(g − 1).

Equality is achieved in the case above.


Canonical curves of genus 4. Now let X ⊂ P3 be a canonical curve of
genus 4 (in the non-hyperelliptic case). Then X has degree 6.

Theorem 12.11 X is the intersection of an irreducible quadric and cubic


hypersurface in P3 .

Proof. The proof is by dimension counting again. There is a natural


linear map from Sym2 (Ω(X)) into H 0 (X, O2K ). Since dim Ω(X) = 4, the

98
first space has dimension 3+2

2 = 10, while the second has dimension 3g −
3 = 9 by Riemann-Roch. Thus there is a nontrivial quadratic equation
Q(ω1 , . . . , ω4 ) = 0 satisfied by the holomorphic 1-forms on X; equivalent X
lies on a quadric. The quadric is irreducible because X does not lie on a
hyperplane. Moreover, it is unique: if X lies on 2 quadrics, then it is a curve
of degree 4, not 6.
Carrying out a similar calculation for degree 3, we find dim Sym3 (Ω(X)) =
3+3
= 20 while dim H 0 (X, O3K ) = 5g − 5 = 15. Thus there is a 5-

3
dimensional space of cubic relations satisfied by the (ωi ). In this space, a
4-dimensional subspace is accounted for by the product of Q with an arbi-
trary linear equation. Thus there must be, in addition, an irreducible cubic
surface containing X.

We will later see that the converse also holds.


Dimension counts for linear systems. Here is another perspective on
the preceding proof. Intersection of surfaces of degree d in P3 with X gives
a birational map between projective spaces,

|dH| → |dK|.

Now note that in general a linear map φ : A → B between vector spaces


gives a birational map
Φ : PA 99K PB
which is projection from PC where C = Ker φ. Then dim A−dim B ≤ dim C
and thus
dim PC ≥ dim PA − dim PB − 1.
In the case at hand PC corresponds to the linear system Sd (X) of surfaces
of degree d containing X. Thus we get:

dim Sd (X) ≥ dim |dH| − dim |dK| − 1.

For d = 2 this gives

dim S2 (X) ≥ 9 − (3g − 4) − 1 = 9 − 8 − 1 = 0

which shows there is a unique quadric Q containing X. For d = 3 we get

dim S3 (X) ≥ 19 − (5g − 6) − 1 = 19 − 14 − 1 = 4.

Within S3 (X) we have Q + |H| which is 3-dimensional, and thus there must
be a cubic surface C not containing Q in S3 (X) as well.

99
This cubic is not unique, since we can move it in concert with Q + H.
The dimension of moduli space Mg : Riemann’s count. What is the
dimension of Mg ? We know the dimension is 0, 1 and 3 for genus g = 0, 1
and 2 (using 6 points on P1 for the last computation).
Here is Riemann’s heuristic. Take a large degree d  g, and consider
the bundle Fd → Mg whose fibers are meromorphic functions f : X → P1
of degree d. Now for a fixed X, we can describe f ∈ Fd (X) by first giving
its polar divisor D ≥ 0; then f is a typical element of H 0 (X, OD ). (The
parameters determining f are its principal parts on D.) Altogether with
find
dim Fd (X) = d + h0 (D) = 2d − g + 1.
On the other hand, f has b critical points, where
χ(X) = 2 − 2g = 2d − b,
so b = 2d + 2g − 2. Assuming the critical values are distinct, they can be
continuously deformed to determine new branched covers (X 0 , f 0 ). Thus the
dimension of the total space is given by
b = dim Fd = dim Fd (X) + dim Mg = 2d + 2g − 2 = 2d − g + 1 + dim Mg ,
and thus dim Mg = 3g − 3. This dimension is in fact correct.
On the other hand, the space of hyperelliptic Riemann surfaces clearly
satisfies
dim Hg = 2g + 2 − dim Aut P1 = 2g − 1,
since such a surface is branched over 2g + 2 points. Thus for g > 2 a typical
Riemann surface is not hyperelliptic.
Tangent space to Mg . As an alternative to Riemann’s count, we note
that the tangent space to the deformations of X is H 1 (X, Θ), where Θ ∼ =

Ω is the sheaf of holomorphic vector fields on X. By Serre duality and
Riemann-Roch, we have
dim H 1 (X, Θ) = dim H 1 (X, O−K ) = h0 (2K) = 4g − 4 − g + 1 = 3g − 3.
Serre duality also shows H 1 (X, Θ) is naturally dual to the space of holomor-
phic quadratic differentials Q(X).
Plane curves again. The space of homogeneous polynomials on Cn+1 of
degree d has dimension N = n+d n . Thus the space of plane curves of degree
d, up to automorphisms of P2 , has dimension
 
2+d
Nd = − 9.
d

100
We find 


−3 = dim Aut P1 for d = 2,

1 = dim M
1 for d = 3,
Nd =
6 = dim M3
 for d = 4,


12 < dim M = 15 for d = 5.
6

Thus most curves of genus 6 cannot be realized as plane curves. In a sense


made precise by the Theorem below, there is no way to simply parameterize
the moduli space of curves of high genus:

Theorem 12.12 (Harris-Mumford) For g sufficiently large, Mg is of


general type.

In fact g ≥ 24 will do.

13 Line bundles
Let X be a complex manifold. A line bundle π : L → X is a 1-dimensional
holomorphic vector bundle.
This means there exists a collection of trivializations of L over charts Ui
on X, say Li ∼ = Ui ×C. Viewing L in two different charts, we obtain clutching
data gij : Ui ∩ Uj → C∗ such that (x, yj ) ∈ Lj is equivalent to (x, yi ) ∈ Li iff
gij (x)yj = yi . This data satisfies the cocycle condition gij gjk = gik .
In terms of charts, a holomorphic section s : X → L is encoded by
holomorphic functions si = yi ◦ s(x) on Ui , such that

si (x) = gij (x)sj (x).

Examples: the trivial bundle X × C; the canonical bundle ∧n T∗ X. Here


the transition functions are gij = 1 for the trivial bundle and gij = 1/ det D(φi ◦
φ−1
j ) for the canonical bundle, with charts φi : Ui → C .
n

In detail, on a Riemann surface X, with local coordinates zi : Ui → C,


a section of the canonical bundle is locally given by ωi = si (z) dzi ; it must
satisfy si (z) dzi = sj (z) dzj , so si = (dzj /dzi )sj .
Tensor powers. From L we can form the line bundle L∗ = L−1 , and more
generally Ld , with transition functions gijd.

A line bundle is trivial if it admits a nowhere-vanishing holomorphic


section (which then provides an isomorphism between L and X × C). Such
a section exists iff there are si ∈ O∗ (Ui ) such that si /sj = gij , i.e. iff gij is
a coboundary.

101
Thus line bundles up to isomorphism over X are classified by the coho-
mology group H 1 (X, O∗ ).
Sections and divisors. Now consider a divisor D on a Riemann surface
X. Then we can locally find functions si ∈ M(Ui ) with (si ) = D. From
this data we construction a line bundle LD with transition functions gij =
si /sj . These transitions functions are chosen so that si is automatically a
meromorphic section of LD ; indeed, a holomorphic section if D is effective.

Theorem 13.1 The sheaf of holomorphic sections L of L = LD is isomor-


phic to OD .

Proof. Choose a meromorphic section s : X → L with (s) = D. (On a


compact Riemann surface, s is well-defined up to a constant multiple.) Then
a local section t : U → L is holomorphic if and only if the meromorphic
function f = t/s satisfies

(t) = (f s) = (f ) + D ≥ 0

on U , which is exactly the condition that f ∈ OD . Thus the map f 7→ f s


gives an isomorphism between OD (U ) and L(U ).

Line bundles on Riemann surfaces. Conversely, it can be shown that


every line bundle L on a Riemann surface admits a non-constant meromor-
phic section, and hence L = LD for some D. More precisely, if L is the sheaf
of sections of L one can show (see e.g. Forster, Ch. 29):

Theorem 13.2 The group H 1 (X, L) is finite-dimensional.

Corollary 13.3 Given any P ∈ X, there exists a meromorphic section


s : X → L with a pole of degree ≤ 1+h1 (L) at P and otherwise holomorphic.

Corollary 13.4 Every line bundle has the form L ∼


= LD for some divisor
D.

From the point of view of sheaf theory, we have

0 → O∗ → M∗ → Div → H 1 (X, O∗ ) → H 1 (X, M∗ ) → 0,

and since every line bundle is represented by a divisor, we find:

Corollary 13.5 The group H 1 (X, M∗ ) = 0.

102
Divisors and line bundles in higher dimensions. On complex man-
ifolds of higher dimension, we can similarly construct line bundles from
divisors. First, a divisor is simply an element of H 0 (M∗ /O ∗
P ); this means
it is locally a formal sum of analytic hypersurfaces, D = (fi ). Then the
associated transition functions are gij = fi /fj as before, and we find:

Theorem 13.6 Any divisor D on a complex manifold X determines a line


bundle LD → X and a meromorphic section s : X → L with (s) = D.

Failure of every line bundle to admit a nonzero section. However


in general not every line bundle arises in this way. For example, there exist
complex 2-tori M = C2 /Λ with no divisors but with plenty of line bundles
(coming from characters χ : π1 (M ) → S 1 ).
Degree. The degree of a line bundle, deg(L), is the degree of the divisor of
any meromorphic section.
In terms of cohomology, the degree is associated to the exponential se-
quence: we have

. . . H 1 (X, Z) → H 1 (X, O) → H 1 (X, O∗ ) → H 2 (X, Z) ∼


= Z.

This allows one to define the degree or first Chern class, c1 (L) ∈ H 2 (X, Z),
for a line bundle on any complex manifold.
Projective space. For projective space, we have H 1 (Pn , C) = 0, and in
fact H 1 (Pn , O) = H 0,1 (Pn ) = 0. (This follows from the Hodge theorem,
which implies that

H n (X, C) = ⊕p+q=n H p,q (X).

This is nontrivial for the case of Pn , n ≥ 2, because elements of H 0,1 are


represented by ∂-closed forms, rather than simply all (0, 1)-forms as in the
case of a Riemann surface).
It follows that line bundles on projective space are classified by their
degree: we have
0 → H 1 (Pn , O∗ ) → H 2 (Pn , Z) ∼
= Z.
We let O(d) denote the (sheaf of sections) of the unique line bundle of degree
d. It has the property that for any meromorphic section s ∈ H 0 (Pn , O(d)),
the divisor D = (s) represents d[H] ∈ H 2 (Pn , Z), where H ∼ = Pn−1 is a
hyperplane.
Example: the tautological bundle. Let Pn be the projective space of
Cn+1 with coordinates Z = (Z0 , . . . , Zn ). The tautological bundle τ → Pn

103
has, as its fiber over p = [Z], the line τp = C · Z ⊂ Cn+1 . Its total space is
Cn+1 with the origin blown up.
To describe τ in terms of transition functions, let Ui = (Zi 6= 0) ⊂ Pn .
Then we can use the coordinate Zi itself to trivialize τ |Ui ; in other words,
we can map τ |Ui to Ui × C the map

(Z0 , . . . , Zn ) 7→ ([Z0 : · · · : Zn ], Zi ).

(Here the origin must be blown up.) Then clearly the transition functions
are given simply by gij = Zi /Zj , since they satisfy

Zi = gij Zj .

Theorem 13.7 There is no nonzero holomorphic section of the tautological


bundle.

Proof. A section gives a map s : Pn → τ → Cn+1 which would have to be


constant because Pn is compact. But then the constant must be zero, since
this is the only point in Cn+1 that lies on every line through the origin.

As a typical meromorphic section, we can define s(p) to be the intersec-


tion of τp with the hyperplane Z0 = 1. In other words,

s([Z0 : Z1 : · · · : Zn ]) = (1, Z1 /Z0 , . . . , Zn /Z0 ).

Then si = Zi /Z0 . Notice that this section is nowhere vanishing (since Zi


has no zero on Ui ), but it has a pole along the divisor H0 = (Z0 ).
Thus we have τ ∼ = O(−1). Similarly, τ ∗ = O(1).
Homogeneous polynomials. Note that the coordinates Zi are sections
of O(1). Indeed, any element in V ∗ naturally determines a function on the
tautological bundle over PV , linear on the fibers, and hence a section of the
dual bundle. Similarly we have:

Theorem 13.8 The space of global sections of O(d) over Pn can be natu-
rally identified with the homogeneous polynomials of degree d on Cn+1 .

One can appeal to Hartog’s extension theorem to show all global sections
have this form.

Corollary 13.9 The hypersurfaces of degree d in projective space are ex-


actly the zeros of holomorphic sections of O(d).

104
The canonical bundle. To compute the canonical bundle of project space,
we use the coordinates zi = Zi /Z0 , i = 1, . . . , n to define a nonzero canonical
form
ω = dz1 · · · dzn
on U0 . To examine this form in U1 , we use the coordinates w1 = Z0 /Z1 ,
wi = Zi /Z1 , i > 1; then z1 = 1/w1 and zi = wi /w1 , i > 1, so we have
ω = d(1/w1 ) d(w2 /w1 ) · · · d(wn /w1 ) = −(dw1 · · · dwn )/w1n+1 .
Thus (ω) = (−n−1)H0 and thus the canonical bundle satisfies K ∼
= O(−n−
1) on Pn .
The adjunction formula.
Theorem 13.10 Let X ⊂ Y be a smooth hypersurface inside a complex
manifold. Then the canonical bundles satisfy
KX ∼
= (KY ⊗ LX )|X.
Proof. We have an exact sequence of vector bundles on X:
0 → TX → TY → TY / TX = N X → 0,
where N X is the normal bundle. Now (N X)∗ is the sub-bundle of T∗ Y |X
spanned by 1-forms that annihilate T X. If X is defined in charts Ui by
fi = 0, then gij = fi /fj defines LX . On the other hand, dfi is a nonzero
holomorphic section of (N X)∗ . The 1-forms dfi |X, however, do not fit to-
gether on overlaps to form a global section of (N X)∗ . Rather, on X we have
fj = 0 so
dfi = d(gij fj ) = gij dfj .
This shows (dfi ) gives a global, nonzero section of (N X)∗ ⊗ LX , and hence
this bundle is trivial on X.
On the other hand KY = KX ⊗ (N X)∗ , by taking duals and determi-
nants. Thus KX = KY ⊗ N X = KY ⊗ LX .

Smooth plane curves. Using the adjunction formula plus Riemann-Roch


we can obtain some interesting properties of smooth plane curves X ⊂ P2
of degree d.
Theorem 13.11 Let f : X → Pn be a holomorphic embedding. Then f is
given by a subspace of sections of the line bundle L → X, where L = f ∗ O(1).

Proof. The divisors of section of O(1) are hyperplanes.

105
Theorem 13.12 Every smooth plane curve of degree d has genus g = (d −
1)(d − 2)/2.

Proof. We have KX ∼ = KP2 ⊗ LX = O(d − 3). Any curve Z of degree d − 3


is the zero set of a section of O(d − 3) and hence restricts to the zero set of
a holomorphic 1-form on X. Thus we find 2g − 2 = d(d − 3).

Corollary 13.13 Every smooth quartic plane curve X is a canonical curve.

Proof. We have KX ∼ = O(d − 3) = O(1), which is the linear system that


gives the original embedding of X into P2 .

Next note that the genus g(X) = (d − 1)(d − 2)/2 coincides with the
dimension of the space of homogeneous polynomials on C3 of degree d − 3.
This shows:

Theorem 13.14 Every effective canonical divisor on X has the form K =


X ∩ Y , where Y is a curve of degree d − 3.

Explicit computation of Ω(X). If we unwind the proof the adjunction


formula, we get a useful algorithm for computing Ω(X) for a plane curve.
Let us work in affine coordinates (x, y) on C2 ⊂ P2 , and suppose X has
degree d and is defined by f (x, y) = 0. Let W (x, y) be a polynomial of
degree d − 3. We can then solve the equation

ω ∧ df = W (x, y) dx dy

to obtain 1-form ω on P2 . The pullback ω|T X gives a 1-form on X.


Note: ω is only well-defined up to adding a multiple of df ; but df |T X = 0,
so ω gives a well-defined 1-form on X.
Now the form df does not vanish along X, since X is smooth. Thus the
zeros of ω ∧ df coincide with the zeros of ω|T X, as well as with the zeros of
W (x, y). Since the curves defined by W and f meet in d(d − 3) = 2g − 2
points, the divisor (ω) must coincide with these points; hence ω|T X is a
holomorphic 1-form. By varying W (x, y) we get all such forms.
Examples. For the cubic curve y 2 = x3 −1 we have d−3 = 0, so W (x, y) = a
and we need to solve

ω ∧ (2y dy − 3x2 dx) = adx dy.

106
Clearly ω = dx/y does the job.
For the quartic curve x4 + y 4 = 1, we have W (x, y) = (ax + by + c), and
we need to solve

ω ∧ (4x3 dx + 4y 3 dy) = (ax + by + c) dx dy.

Clearly ω = dy/x3 works for W (x, y) = c, and then xω and yω give the
other 2 forms spanning Ω(X).
Gonality of plane curves. What is the minimum degree of a map of a
plane curve to P1 ?

Theorem 13.15 Any n + 1 distinct points in P2 impose independent con-


dition on curves of degree n.

Proof. Choose Y to be the union of n random lines through the first k ≤ n


points. Then Y is an example of a curve through the first k points that
does not pass through the k + 1st. This shows that adding the k + 1st point
imposes an additional condition on Y .

Theorem 13.16 A smooth curve X of degree d > 1 admits a nonconstant


map to P1 of degree d − 1, but none of degree d − 2.

Proof. For degree d − 1, simply projection from a point on X. For the


second assertion, suppose f : X → P1 has degree e ≤ d − 2. Let E ⊂ X be a
generic fiber of f . Then the d − 2 points E impose independent conditions
on the space of curves Y degree d − 3. Consequently

h0 (K − E) = g − deg E.

By Riemann-Roch we then have:

h0 (E) = 1 − g + deg(E) − h0 (K − E) = 1,

so |E| does not provide a map to P1 .

Hypersurfaces in products of projective spaces. Here are two further


instances of the adjunction theorem.

Theorem 13.17 Every smooth degree 6 intersection X of a quadric Q and


a cubic surface C is a canonical curve in P3 .

107
Proof. It can be shown that Q and C are smooth and transverse along X
(this is nontrivial). We then have KQ ∼
= KP3 ⊗ LQ and thus

= KP3 ⊗ LQ ⊗ LC ∼
= KQ ⊗ LC ∼
KX ∼ = O(−4 + 2 + 3) = O(1).

Theorem 13.18 Every smooth (d, e) curve on Q = P1 × P1 has genus g =


(d − 1)(e − 1).

Proof. It is easy to see that KX×Y = KX ⊗ KY . Thus KQ = O(−2, −2).


Therefore 2g − 2 = C · KC and KC = O(d − 2, e − 2), so 2g − 2 = d(e − 2) +
e(d − 2), which implies the result.

Canonical curves of genus 5. By a calculation with Riemann-Roch, one


can show that any canonical curve X ⊂ P4 is an intersection of 3 quadrics.
Conversely, a complete intersection of 3 quadrics is a canonical curve, be-
cause its canonical bundle is O(−5 + 2 + 2 + 2) = O(1).
In general, the canonical bundle of a complete intersection X of hyper-
surfaces of degrees di in Pn is given by

KX ∼
X
= O(−n − 1 + di ).

K3 surfaces. Manifolds with trivial canonical bundle are often interesting


— in higher dimensions they are called Calabi-Yau manifolds.
For Riemann surfaces, KX is trivial iff X is a complex torus. For 2-
dimensional manifolds, complex tori also have trivial canonical, but they
are not the only examples. Another class is provided by the K3 surfaces,
which by definition are simply-connected complex surfaces with KX trivial.
Example:

Theorem 13.19 Every smooth surface of degree 4 in X = P3 and of degree


(2, 2, 2) in X = P1 × P1 × P1 is a K3 surface.

Proof. Here KX = O(−3) or KX = O(−2, −2, −2), so the canonical bundle


is trivial. By the Lefschetz hyperplane theorem, a hypersurface in a simply-
connected complex 3-manifold is always itself simply-connected.

108
14 Curves and their Jacobians
We now turn to the important problem of classifying line bundles on a Rie-
mann surface X; equivalently, of classifying divisors modulo linear equiva-
lence.
The Jacobian. Recall that a holomorphic 1-form on X is the same thing
as a holomorphic map f : X → C well-defined up to translation in C. If
the periods of f happen to generate a discrete subgroup Λ of C, then we
can regard f as a map to C/Λ. However the periods are almost always
indiscrete. Nevertheless, we can put all the 1-forms together and obtain a
map to Cg /Λ.

Theorem 14.1 The natural map H1 (X, Z) → Ω(X)∗ has as its image a
lattice Λ ∼
= Z2g .

Proof. If not, the image lies in a real hyperplane defined, for some nonzero
ω ∈ Ω(X), by the equation Re α(ω) = 0. But then all the periods of Re ω
vanish, which implies the harmonic form Re ω = 0.

The Jacobian variety is the quotient space Jac(X) = Ω(X)∗ /H1 (X, Z),
the cycles embedded via periods.

R Q P ∈ X, there is a natural map φP :


Theorem 14.2 Given any basepoint
X → Jac(X) given by φP (Q) = P ω.

We will later show this map is an embedding, and thus Jac(X), roughly
speaking, makes X into a group.
Example: the pentagon. Let X be the hyperelliptic curve defined by
y 2 = x5 −1. Geometrically, X can be obtained by gluing together two regular
pentagons. Clearly X admits an automorphism T : X → X of order 5. Using
the pentagon picture, one can easily show there is a cycle C ∈ H1 (X, Z) such
that its five images T i (C) span H1 (X, Z) ∼
= Z4 . This means H1 (X, Z) = A·C
is a free, rank one A-module, where A = Z[T ]/(1 + T + T 2 + T 3 + T 4 ).
Now T ∗ acts on Ω(X). Since X/hT i has genus zero, T has no invariant
forms. Thus we can choose a basis (ω1 , ω2 ) for Ω(X) such that

T ∗ ωi = ζ i ωi

where ζi is a primitive 5th root


R of unity.
Let us scale these ωi so C ωi = 1. Let

π : H1 (X, Z) → Λ ⊂ C2

109
be the period map, defined by
Z Z 
π(B) = ω1 , ω2 .
B B

Since Z Z
ω= T ∗ ω,
T (B) B

we have !
ζ1 0
π(T B) = π(B).
0 ζ2
Since H1 (X, Z) = Z[T ] · C, we find that Λ ⊂ C2 is simply the image of the
ring Z[T ] under the ring homomorphism that sends T to (ζ1 , ζ2 ).
Since Λ is a lattice, we cannot have ζ2 = ζ 1 = ζ14 , nor can we have
ζ2 = ζ1 . Thus ζ2 = ζ12 or ζ13 . In the latter case we can interchange the
eigenforms to obtain the former case. This finally shows:

Theorem 14.3 The Jacobian of the hyperelliptic curve y 2 = x5 − 1 is iso-


morphic to C2 /Λ, where Λ is the ring Z[T ]/(1 + T + T 2 + T 3 + T 4 ) embedded
in C2 by
T 7→ (ζ, ζ 2 ),
and ζ is a primitive 5th root of unity.

This is an example of a Jacobian variety with complex multiplication.


The Picard group. Let Pic(X) denote the group of all line bundles on
X. Since every line bundle admits a meromorphic section, there is a natural
isomorphism between Pic(X) and Div(X)/(M∗ (X)), where M∗ (X) is the
group of nonzero meromorphic functions, mapping to principal divisors in
Pic(X). Under this isomorphism, the degree and a divisor and of a line
bundle agree.
The Abel-Jacobi map φ : Div0 (X) → Jac(X) is defined by
X  X Z Qi
φ Qi − Pi (ω) = ω.
Pi

Because of the choice of path from Pi to Qi , the resulting linear functional


is well-defined only modulo cycles on X.
For example: given any basepoint PR ∈ X, we obtain a natural map
Q
f : X → Jac(X) by f (Q) = φ(Q − P ) = P ω.
One of the most basic results regarding the Jacobian is:

110
Theorem 14.4 The map φ establishes an isomorphism between Jac(X) and
Pic0 (X) = Div0 (X)/(principal divisors).
The proof has two parts: Abel’s theorem, which asserts that Ker(φ) coin-
cides with the group of principal divisors, and the Jacobi inversion theorem,
which asserts that φ is surjective.
P
Theorem 14.5 (Abel’s theorem) A divisor D is principal iff D = Qi −
Pi and
X Z Qi
ω=0
Pi
for all ω ∈ Ω(X), for some choice of paths γi joining Pi to Qi .

Proof in one direction. Suppose D = (f ). We can assume after multiply-


ing f by a scalar, that none of its critical values are real. Let γ = f −1 ([0, ∞]).
Then we have Z ∞
X Z Qi Z
ω= ω= f∗ (ω) = 0
Pi γ 0

since f∗ (ω) = 0, being a holomorphic 1-form on C. b (To see this, suppose


d i i
locally f (z) = z . Then f∗ (z dz) = 0 unless z dz is invariant under rotation
by the dth roots of unity. This first happens when i = d − 1, in which case
z d−1 dz = (1/d)d(z d ), so the pushforward is proportional to dz.)

The curve in its Jacobian. Before proceeding to the proof of Abel’s


theorem, we derive some consequences.
Given P ∈ X, define
φP : X → Jac(X)
by φP (Q) = (Q − P ). Note that with respect to a basis ωi for Ω(X), the
derivative of φP (Q) in local coordinates is given by:
DφP (Q) = (ω1 (Q), . . . , ωg (Q)).
This shows:
Theorem 14.6 The canonical map X → PΩ(X)∗ is the Gauss map of φP .
Theorem 14.7 For genus g ≥ 1, the map φP : X → Jac(X) is a smooth
embedding.

Proof. If Q − P = (f ), then f : X → P1 has degree 1 so g = 0. Since |K|


is basepoint-free, there is a nonzero-holomorphic 1-form at every point, and
hence DφP 6= 0.

111
Theorem 14.8 (Jacobi) The map Div0 (X) → Jac(X) is surjective. In
fact, given (P1 , . . . , Pg ) ∈ X g , the map

φ : X g → Jac(X)

given by
X  Z Qi 
φ(Q1 , . . . , Qg ) = φ Qi − Pi = ωj
Pi
is surjective.

Proof. It suffices to show that det Dφ 6= 0 at some point, so the im-


age is open. To this end, just note that dφ/dQi = (ωj (Qi )), and thus
det Dφ(Q1 , . . . , Qg ) = 0 if and only if there is an ω vanishing simultane-
ously at all the Qi , i.e. iff (Qi ) lies on a hyperplane under the canonical
embedding. For generic Qi ’s this will not be the case, and hence det Dφ 6= 0
almost everywhere on X g .

Corollary 14.9 We have a natural isomorphism:

Pic0 (X) = Div0 (X)/(M∗ (X)) ∼


= Jac(X).

Bergman metric. We remark that the space Ω(X), and hence its dual,
carries a natural norm given by:
Z
2
kωk2 = |ω(z)|2 |dz|2 .
X

This induces a canonical metric on Jac(X), and hence on X itself.


This metric ultimately comes from the intersection pairing or symplectic
form on H1 (X, Z), satisfying ai · bj = δij .
Abel’s theorem, proof I: The ∂-equation. (Cf. Forster.) For the
converse, we proceed in two steps. First we will construct a smooth solution
to (f ) = D; then we will correct it to become holomorphic.
Smooth solutions. Let us say a smooth map f : X → C b satisfies (f ) = D
if near Pi (resp. Qi ) we have f (z) = zh(z) (resp. z −1 h(z)) where h is a
smooth function with values in C∗ , and if f has no other zeros or poles.
Note that for such an f , the distributional logarithmic derivative satisfies

∂f X 0
∂ log f = + (Qi − Pi0 ),
f

112
where Q0i and Pi0 are δ functions (in fact currents), locally given by ∂ log z,
and ∂f /f is smooth.
Inspired by the proof in one direction already given, we first construct
a smooth solution of (f ) = D which maps a disk neighborhood Ui of γi
diffeomorphicly to a neighborhood V of the interval [0, ∞].

Lemma 14.10 Given any arc γ joining P to Q on X, there exists a smooth


solution to (f ) = Q − P satisfying
Z Z Q
1 ∂f
∧ω = ω
2πi X f P

for all ω ∈ Ω(X), where the integral is taken along γ.

Proof. First suppose P and Q are close enough that they belong to a single
chart U , and γ is almost a straight line. Then we can choose the isomorphism
f :U →V ⊂C b so that f (P ) = ∞, f (Q) = 0 and f (γ) = [0, ∞] (altering γ
by a small homotopy rel endpoints). This f is already holomorphic on U ,
and it sends ∂U to a contractible loop in C∗ . Thus we can extend f to a
smooth function sending X − U into C∗ , which then satisfies (f ) = D.
Now note that z admits a single-valued logarithm on the region C−[0,
b ∞],
and thus log f (z) has a single-valued branch on Y = X − γ. Thus df /f =
d log f is an exact form on Y (and we only need the ∂ part when we wedge
with ω). However as one approaches γ from different sides, the two branches
of log f differ by 2πi. Applying Stokes’ theorem, we find:
Z Z
(∂ log f ) ∧ ω = 2πi ω.
X γ

To handle the case of well-separated P and Q, simply break γ up into


many small segments and take the product of the resulting f ’s.

Taking the product of the solutions for several pairs of points, and using
additivity of the logarithmic derivative, we obtain:

Corollary 14.11 Given P an arc γi joining Pi to Qi on X, there exists a


smooth solution to (f ) = Qi − Pi satisfying
Z Z
1 ∂f X
∧ω = ω
2πi X f γi

for all ω ∈ Ω(X).

113
From smooth to holomorphic. To complete the solution, it suffices to
find a smooth function g such that f eg is meromorphic. Equivalently, it
suffices to solve the equation ∂g = −∂f /f . (Note that ∂f /f is smooth even
at the zeros and poles of f , since near there f = z n h where h 6= 0.) Since
H 0,1 (X) ∼
= Ω(X)∗ , such a g exists iff
Z XZ Qi
1
(∂ log f ) ∧ ω ω=0
2πi Pi

for all ω ∈ Ω(X). This is exactly the hypothesis of Abel’s theorem.

The symplectic form on H 1 . For the second proof of Abel’s theorem, we


need to make the connection between the Jacobian and the symplectic form
on H 1 (X, C) more transparent. We now turn to some topological remarks
that will be useful in describing the Jacobian more explicitly, and also in
the second proof of Abel’s theorem.
First, recall that a surface of genus g admits a basis for H1 (X, Z) of the
form (ai , bi ), i = 1, . . . , g, such that ai · bi = 1 and all other products vanish.
In other words, the intersection pairing makes H1 (X, Z) into a unimodular
symplectic space.
This symplectic form on H1 (X, Z) gives rise to one on the space of peri-
ods, H 1 (X, C), defined by:
X
[α, β] = α(ai )β(bi ) − α(bi )β(ai ).

In matrix form, we can associate a vector of periods (αa , αb ) ∈ C2g to any


homomorphism α, and then we have
! ! !
0 I βa αa αb
[α, β] = (αa , αb ) = = αa · βb − αb · βa .
−I 0 βb βa βb

1 (X) and H 1 (X, C),


Theorem 14.12 Under the period isomorphism between HDR
we have Z
α ∧ β = [α, β].
X

Proof. Cut X along the (ai , bi ) curves to obtain a surface F with boundary,
on which we can write α = df . Then we have
Z Z Z
α ∧ β = (df ) ∧ β = f β.
X F ∂F

114
ai + b0i − a0i − bi , then
P
If we write ∂F =

f |a0i − f |ai = α(bi ) and f |b0i − f |bi = α(ai ).

On the other hand, β is the same along corresponding edges of ∂F . Therefore


we find Z X
α∧β = α(ai )β(bi ) − α(bi )β(ai ).

Poincaré duality. Note that for any cocycle N ∈ H 1 (X, Z), there is a dual
cycle C ∈ H1 (X, Z) such that
Z
[N, α] = α
C

for all α ∈ H 1 (X, C). In fact, since


X
[N, α] = N (ai )α(bi ) − N (bi )α(ai ),

we can simply take X


C= N (ai )bi − N (bi )ai .
Since the intersection form is unimodular, this construction gives a natural
isomorphism
H 1 (X, Z) ∼
= H1 (X, Z).

Recognizing holomorphic 1-forms. Our second remark uses the Rie-


mann surface structure. Namely, we note that a class α ∈ H 1 (X, C) can be
represented by a holomorphic 1-form iff

[α, ω] = 0 ∀ω ∈ Ω(X).

This is because H 1 (X, C) ∼


= Ω(X) ⊕ Ω(X).
Meaning of the norm. What is the meaning of kωk2 ? It is simply the
area of X under the map f : X → C defined locally by solving df = ω. This
map can be constructed globally on X once it is slit along the ai and bi
cycles.
For example, the area of a rectangle R in C with adjacent sides a and b,
positively oriented, is
i 
area(R) = Im(ab) = ab − ab .
2

115
And if X is a complex torus with a 1-form ω, with periods (a, b) = ω(a1 ), ω(b1 ),
then Z
|ω|2 = (i/2)[ω, ω] = (i/2)(ab − ba).
X

Abel’s theorem, proof II. We are now prepared to give a second proof
that a divisor with φ(D) = 0 is principal (cf. Lang, Algebraic Functions.) To
try to construct f with (f ) = D, we first construct a candidate for λ = df /f.

Theorem 14.13 For any divisor with deg D = 0, P there exists a meromor-
phic differential λ with only simple poles such that ResP (λ) · P = D.

Proof. By Mittag-Leffler for 1-forms, λ exists because the sum of its


residues is zero.
Alternatively, by Riemann-Roch, for any P, Q ∈ X we have dim H 0 (K +
P + Q) > dim H 0 (K). Thus there exists a meromorphic 1-form λ with a
simple pole at one of P or Q. By the residue theorem, λ has poles at both
points with opposite residues. Scaling λ proves the Theorem for Q − P , and
a general divisor of degree zero is a sum of divisors of this form.

Now if we could
R arrange that the periods of λ are all in the group 2πiZ,
then f (z) = exp λ would be a meromorphic function with (f ) = D.
(Compare Mittag-Leffler’s proof of Weierstrass’s theorem on functions
with prescribed zeros.)
Periods of λ. As we have just seen, D determines a meromorphic form λ
with simple poles and
X
Res(λ) = ResP (λ) · P = D.
P

However λ is not uniquely determined by this condition — it can be modified


by a form in Ω(X). Furthermore, the periods of λ are not well defined —
when we move a path across a pole of λ, it changes by an integral multiple
of 2πi.
What we do obtain, canonically from D, is a class

[λ] ∈ H 1 (X, C)/(2πiH 1 (X, Z) + Ω(X)),

where Ω(X) is embedded by its periods. And this class is zero iff there is a
meromorphic function such that (f ) = D.
Reciprocity. To make everything better defined, we now choose a sym-
plectic basis (ai , bi ) as before, and cut X open along this basis to obtain a

116
region F . We also arrange that the poles of λ lie inside F . Then the periods
of λ become well-defined, by integrating along these specific representatives
ai , bi of a basis for H1 (X, Z).
Now any ω ∈ Ω(X) can be written as ω = df on F . On the other hand,
ω ∧ λ = 0 in F , at least if we exclude small loops around the poles of λ.
Integration around these loops gives residues. Thus we find
Z Z X
0 = ω∧λ= f λ − 2πi Res(f Pi ) + Res(f Qi )
F ∂F
X Z Qi
= [ω, λ] − 2πi ω,
Pi
where the path of integration is taken in F .
Now suppose φ(D) = 0. Then the sum above vanishes modulo the
periods of ω. In other words, there exists a cycle C such that
Z
[ω, λ] = 2πi ω = 2πi[ω, N ]
C

for some N ∈ H 1 (X, C).But this means [ω, λ − 2πiN ] = 0 for all ω, and
hence λ − 2πiN is represented by a holomorphic 1-form. Modifying λ by
this form, we obtain [λ] = 2πiN ,R and thus the periods of the λ are integral
multiples of 2πi. Then f = exp( λ) satisfies (f ) = D.

Natural subvarieties of the Jacobian. Just as we can embed X into


Jac(X) (or more naturally into Pic1 (X)), we can map X (k) = X k /Sk to
Jac(X) (or into Pick (X)). By what we have just seen, the fibers of this map
consists of linearly equivalent divisors, and the dimensions of the fibers are
predicted by Riemann-Roch. In particular, we have:
Theorem 14.14 The fibers of the natural map X (k) → Pick (X) are projec-
tive spaces |D| corresponding to complete linear systems of degree k.
The images of these maps in Jac(X) (which are well-defined up to trans-
lation) are usually denoted Wk ; in particular, we obtain an important divi-
sors Wg−1 in this way, while Wg = Jac(X).
The Jacobian and X (g) . As a particular case, we note that the surjective
map
X (g) → Picg (X)
has, as its fibers, projective spaces corresponding to special divisors. For
example, when g = 2 the only special divisor class is |K|, and X (2) is
therefore a complex torus with one point blown up to a line (P1 ∼ = |K|).
Mordell’s Conjecture.

117
Theorem 14.15 Suppose X has genus g ≥ 2. Then, given any finitely
generated subgroup A ⊂ Jac(X), the set X ∩ A is finite.

This theorem is in fact equivalent to Mordell’s conjecture (Falting’s the-


orem), which states that X(K) is finite for any number field K.
The exponential sequence, the Picard group and the Jacobian. An
alternative description of the Jacobian is via the exponential sequence which
leads to the exact sequence

H 1 (X, Z) → H 1 (X, O) → H 1 (X, O∗ ) → H 2 (X, Z) → 0.

Under the isomorphisms H 1 (X, Z) ∼


= H1 (X, Z) by cup product, H 1 (X, O) ∼
=
Ω(X) by Serre duality, and H (X, Z) ∼
∗ 2
= Z by degree, we obtain the isomor-
phism

Pic0 (X) = Ker(H 1 (X, O) → H 2 (X, Z)) ∼


= Ω(X)∗ /H1 (X, Z) = Jac(X).

The norm map. We note that if f : X → Y is a nonconstant holomor-


phic map, we have a natural map on divisors and hence an induced map
Jac(X) → Jac(Y ) and even a pushforward map on line bundles, Pic(X) →
Pic(Y ). We also have a norm map N : M∗ (X) → M(Y )∗ , given by taking
the product over the fibers. Note that:

D − D0 = (g) =⇒ f (D) − f (D0 ) = (N (g)).

For this to be correct, the points of f (D) must be taken with multiplicity
according to the branching of f . In particular, deg f (P ) is not generally a
continuous function of P ∈ X.
Recalling that a divisor is represented by a cochain gi ∈ M∗ (Ui ) with
(δg)ij ∈ O∗ (Uij ), the pushforward of divisors can be defined by the local
norm map as well.
The same construction works for equidimensional maps between complex
manifolds of higher dimension. In this case f (D) = 0 if the image of f is a
lower-dimensional variety.
The Siegel upper half-space. The Siegel upper half-space is the space Hg
of symmetric complex g × g matrices P such that Im P is positive-definite.
The space Hg is the natural space for describing the g-dimensional com-
plex torus Jac(X), just as H is the natural space for describing the 1-
dimensional torus C/Λ.
To describe the Jacobian via Hg , we need to choose a symplectic basis
(ai , bi ) for H1 (X, Z).

118
Theorem 14.16 There exists a unique basis ωi of Ω(X) such that ωi (aj ) =
δij .

Proof. To see this we just need to show the map from Ω(X) into the space
of a-periods is injective (since both have dimension g. But suppose the
a-periods of ω vanish. Then the same is true for the (0, 1)-form ω, which
implies
Z Z X
2
|ω(z)| dz dz = ω ∧ ω = ω(ai )ω(bi ) − ω(bi )ω(ai ) = 0,

and thus ω = 0.

Definition. The period matrix of X with respect to the symplectic basis


(ai , bi ) is given by Z
τij = ωi = ωi (bj ).
bj

Theorem 14.17 The matrix τ is symmetric and Im τ is positive-definite.

Proof. To see symmetry we use the fact that for any i, j:


Z X
0 = ωi ∧ ωj = ωi (ak )ωj (bk ) − ωi (bk )ωj (ak ) = δij τjk − τik δjk = τji − τij .

Similarly, we have
Z Z
i
ω∧ω = |ω(z)|2 |dz|2 ≥ 0.
2
Thus, if we let
Z
i iX
Qij = ωi ∧ ω j = ωi (ak )ω j (bk ) − ωi (bk )ω j (ak )
2 2
i
= (τ ji − τij ) = Im τij ,
2
then Z X
X
Qij ci cj = |ci |2 |ωi |2 ,

and thus Im τ is positive-definite.

119
The symplectic group Sp2g (Z) now takes over for SL2 (Z), and two com-
plex tori are isomorphic as principally polarized abelian varieties if and only
if there are in the same orbit under Sp2g (Z).
With respect to the basis (b1 , . . . , bg , a1 , . . . , ag ) the symplectic group
acts on the full matrix of a and b periods by
! ! !
A B τ Aτ + B
= .
C D I Cτ + D

We must then change the choice of basis for Ω(X) to make the a-periods,
Cτ + D into the identity matrix; and we find

g(τ ) = (Aτ + B)(Cτ + D)−1 ,

which shows Sp2g (Z) acts on Hg by non-commutative fractional linear trans-


formations.
A non-Jacobian. We note that for g ≥ 2 there are period matrices which
do not arise from Jacobians. The simplest example comes from the rank 2
diagonal matrices !
t1 0
τij =
0 t2
with ti ∈ H (which, incidentally, show there is a copy of Hg in Hg ). It is
essential, here, that the matrix above is with respect to a symplectic base.
To see this matrix cannotRarise, suppose
R in fact τij = Jac(X). Note that
Per(ω1 ) = Z ⊕ Zt1 = Λ1 and a2 ω1 = b2 ω1 = 0. Thus by integrating ω1 we
get a map f : X → E = C/Λ1 such that f ∗ (dz) = ω1 . But then
Z Z Z
i
|f ∗ (dz)|2 = ω1 ∧ ω 1 = (i/2)(ω 1 (b1 ) − ω1 (b1 )) = Im(b1 ) = |dz|2 ,
X 2 X E

and hence deg(f ) = 1, which is impossible for a holomorphic map between


Riemann surfaces of different genera.
In fact A = C2 /Z2 ⊕τ Z2 can be realized as the Jacobian of a stable curve,
E1 ∨ E2 . And as a complex torus, A can be realized as a Jacobian — in fact,
in genus two, a complex torus is a Jacobian unless it splits symplecticly as a
product.
A dimension count shows that in genus g ≥ 4, there are principally
polarized Abelian varieties which cannot be realized by stable curves.
The infinitesimal Torelli theorem. A beautiful computation of Ahlfors
gives the explicit derivative of the map Mg → Ag provided by X 7→ Jac(X).

120
More precisely, the coderivative of this map sends the cotangent space at a
given periodic matrix (τij ) to the cotangent space to Mg at X, which is
naturally the space Q(X) of holomorphic quadratic differentials on X.
The map is given simply by
dτij = ωi ωj ∈ Q(X).
Now we have the following basic fact:
Theorem 14.18 (Max Noether) The natural map Ω(X)⊗Ω(X) → Q(X)
is surjective iff X is not hyperelliptic.
Corollary 14.19 (Local Torelli theorem) Away from the hyperelliptic
locus, the map Mg → Ag is an immersion.
Noether’s theorem can be rephrased as a property of the canonical curve
X ⊂ Pg−1 — the complete linear series of quadrics in Pg−1 restricts to a
complete linear series on X. In fact, any canonical curve is projectively
normal – see Griffiths and Harris.
Ahlfors’ formula: sketch of the proof. A new complex structure on
X is specified by a (small) Beltrami differential µ = µ(z)dz/dz. Then the
complex cotangent space is generated at each point, not by dz, but by
dz ∗ = (1 + µ) dz = dz + µ(z)dz.
In this way we obtain a new Riemann surface X ∗ . We let ωi denote the
basis for Ω(X) with ωi (aj ) = δij , and similarly for ωi∗ . We have
ωi∗ = ωi∗ (z)(dz + µ(z)dz).
Let k = sup |µ| which is assumed to be small. Since ωi −ωi∗ has vanishing
a-periods, this form has self-intersection number zero in cohomology. Thus
Z
i
0= (ωi − ωi∗ ) ∧ (ω i − ω ∗i ).
2 X
This yields Z Z
|ωi − ωi∗ |2 |dz|2 = |µ|2 |ωi∗ |2 |dz|2 ,
which implies
ωi∗ = ωi + O(k).
A simple computation also gives
Z Z
∗ ∗
τij − τij = ωi ∧ ωi = ωi ωj µ + O(k 2 ).
X X
Thus the variation in τij is given, in the limit, by pairing µ with ωi ωj .

121
Theta functions. The theory of θ-functions allows one to canonically
attach a divisor
Θ ⊂ A = Cg /(Zg ⊕ τ Zg )
to the principally polarized Abelian variety A determined by τ ∈ Hg . Using
the fact that Im τ  0, we define the entire θ-function θ : Cg → C by
X
θ(z) = exp(2πihn, zi) exp(πihn, τ ni).
n∈Zg

The zero-set of θ is Λ-invariant, and descends to a divisor Θ on A.


For example, when g = 1 and τ ∈ H, we obtain:
X
θ(z) = exp(2πinz) exp(πin2 τ ).
Z

Clearly θ(z + 1) = θ(z), and we have

θ(z + τ ) = θ(z) exp(−2πiz − πiτ ).

A useful notation for g = 1 is to set q = exp(πiτ ) and ζ = exp(2πiz);


then X 2
θ(ζ) = qn ζ n,
n∈Z

and we have
θ(q 2 η) = (qζ)−1 θ(ζ).

Zeros and poles. It turns out that θ has a unique zero on X = C/Z⊕τ Z =
C∗ /q 2Z . Indeed, we have
X 2 X 2
θ(−q) = q n +n (−1)n = q −1/4 q (n+1/2) (−1)n = 0,
2 2
since the terms q m and q (−m) , m ∈ Z + 1/2, occur with opposite signs.
A general meromorphic function on X can be expressed as a suitable
product of translates of θ functions. Indeed,
n
Y
fa (ζ) = θ(−qζ/ai )
1

has zeros at ai , and satisfies


n
Y
2 2 −n
fa (q ζ) = (q ζ) (−ai ).
1

122
Qn Qn
Thus if 1 ai = 1 bi , the function

fa (ζ)
F (ζ) =
fb (ζ)

is meromorphic on X, with zeros at (ai ) and poles at (bi ). The condition


that the products agree guarantees that the divisor of (F ) maps to zero
under the Abel-Jacobi map.
Theta functions for g ≥ 2. A convenient notation for θ functions in
higher genus is the following: we set

ζ = (ζ1 , . . . , ζg ) = (exp 2πizi ) ∈ (C∗ )g ,


q = qij = exp(πiτij ) ∈ exp(πiHg ),
n = (n1 , . . . , ng ) ∈ Zg ,
Y n
ζn = ζi i , and
i
n2 nn
Y
q = qiji j .
i,j

Then we set
2
X
θ(ζ) = qn ζ n,
n∈Zg

and find for all k ∈ Zg :


X 2 2
X 2 2
θ(q 2k ζ) = q n q 2nk ζ n = q −k ζ −k q (n+k) ζ n+k = q −k ζ −k θ(ζ).

Noting that ζ(zi + mi + τij kj ) = q 2k ζ, this shows that Θ = (θ) is a


P P
divisor on A = Cg /(Zg ⊕ τ Zg ).

Theorem 14.20 (Riemann) We have Θ = Wg−1 +κ for some κ ∈ Jac(X),


where Wg−1 ⊂ Jac(X) is the image of X g−1 under the Abel-Jacobi map.

In particular, for g = 2 the divisor Θ is isomorphic to X itself. Note


however that Θ makes sense even when A is not a Jacobian.
The Torelli theorem.

Theorem 14.21 Suppose (Jac(X), ω) and (Jac(Y ), ω 0 ) are isomorphic as


(principally) polarized complex tori (Abelian varieties). Then X is isomor-
phic to Y .

123
Sketch of the proof. Using the polarization, we can reconstruct the divisor
Wg−1 ⊂ Jac(X) up to translation, using θ-functions.
Now the tangent space at any point to Jac(X) is canonically identified
with Ω(X)∗ , and hence tangent hyperplanes in Jac(X) give hyperplanes in
PΩ(X), the ambient space for the canonical curve. (For convenience we will
assume X is not hyperelliptic.) In particular, the Gauss map on the smooth
points of Wg−1 defines a natural map
γ : Wg−1 → PΩ(X) ∼ = Pg−1 .
The composition of γ with the natural map C g−1 → Wg−1 simply sends g −1
points Pi on the canonical curve to the hyperplane H(Pi ) they (generically)
span.
Clearly this map is surjective. Since Wg−1 and Pg−1 have the same
dimension, γ is a local diffeomorphism at most points. The branch locus
corresponds to the hyperplanes that are tangent to the canonical curve X ⊂
Pg−1 . Thus from (Jac(X), ω) we can recover the collection of hyperplanes
X ∗ tangent to X ⊂ Pg−1 .
Finally one can show geometrically that X ∗ = Y ∗ implies X = Y . The
idea is that, to each point P ∈ X we have a (g − 3) dimensional family of
hyperplanes HP ⊂ X ∗ containing the tangent line TP (X). These hyper-
planes must all be tangent to Y as well; but the only reasonable way this
can happen is if TP (X) = TQ (Y ) for some Q on Y .
Now for genus g > 3 one can show a tangent line meets X in exactly one
point; thus Q is unique and we can define an isomorphism f : X → Y by
f (P ) = Q. (For example, in genus 4 the canonical curve is a sextic in P3 .
If a tangent line L ⊂ P3 were to meet 2P and 2Q, then the planes through
L would give a complementary linear series of degree 2, so X would be
hyperelliptic.) For g = 3 there are in general a finite number of ‘bitangents’
to the quartic curve X ⊂ P2 , and away from these points we can define f ,
then extend. (For details see Griffiths and Harris).

Explicit Torelli theorem in genus 2. In the case of genus two, the


Riemann surface X can be reconstructed directly from its period matrix
as the zero locus of the associated theta function. As remarked above, the
Gauss map then gives the canonical map from X to P1 . To give X as
an algebraic curve it suffices to determine the six branch points B of the
canonical map. But these are nothing more than the images under the
Gauss map of the six odd θ characteristics. In brief, we find:
Theorem 14.22 If g(X) = 2, then the 6 branch points of the canonical map
X → P1 correspond to the values of dθ at the 6 odd theta characteristics.

124
Spin structures in genus two. A spin structure on X is the choice of
a square-root of the canonical bundle. Equivalently it is a divisor class [D]
such that 2D ∼ K. The parity of a spin structure is defined to be the parity
of h0 (D).
The spin structures on X lie naturally in Picg−1 (X) and form a homo-
geneous space for Jac(X)[2] ∼ 2g
= F2 . The parity is, in fact, determined on
these points by the polarization of the Jacobian.
In the case of genus 2, it is easy to see that there are 6 odd spin structures,
one for each Weierstrass point Pi ∈ X. In fact the differences Pi − Pj of
Weierstrass points span Jac(X)[2].
To see this more clearly, first note P that 2Pi ∼ 2Pj ∼ K for any i and j.
ThusP if we take any (i ) ∈ F62 with i = 0 mod 2, and lift to integers such
that i = 0, then
X6
φ() = i Pi ∈ Jac(X)[2]
1

is well-defined. We also note that D = [5P 1 − 62 Pi ] is a principal divisor;


P
indeed, D = (y) where y 2 = f (x) and deg f = 5. P Thus φ(1, 1, 1, 1, 1, 1) = 0.
Taking the quotient of the space of (i ) with i = 0 mod 2 by subspace
∼ 4
h(1, 1, 1, 1, 1, 1)i, we obtain Jac(X)[2] = F2 .

15 Hyperbolic geometry
(The B-side.)
Elements of hyperbolic geometry in the plane. The hyperbolic metric
is given by ρ = |dz|/y in H and ρ = 2|dz|/(1 − |z|2 ) in ∆.
Thus d(i, iy) = log y in H, and d(0, x) = log(x + 1)/(x − 1) in ∆. Note
that (x + 1)/(x − 1) maps (−1, 1) to (0, ∞).
An important theorem for later use gives the hyperbolic distance from
the origin to a hyperbolic geodesic γ which is an arc of a circle of radius r:
1
sinh d(0, γ) = ·
r
To see this, let x be the Euclidean distance from 0 to γ. Then we have, by
algebra,
2x
sinh d(0, x) = · (15.1)
1 − x2
On the other hand, we have by a right triangle with sides 1, r and x + r.
Thus 1 + r2 = (x + r)2 which implies 2x/(1 − x2 ) = 1/r.

125
A more intrinsic statement of this theorem is that for any point p ∈ H
and geodesic γ ∈ H, we have

sinh d(p, γ) = cot(θ/2),

where θ is the visual angle subtended by γ as seen from p.


Area of triangles and polygons. The area of an ideal triangle is π. The
area of a triangle with interior angles (0, 0, α) is π − α. From these facts one
can see the area of a general triangle is given by the angle defect:

T (α, β, γ) = π − α − β − γ.

To see this, one extends the edges of T to rays reaching the vertices of an
ideal triangle I; then we have

T (α, β, γ) = I − T (π − α) − T (π − β) − T (π − γ)

which gives π − α − β − γ for the area.


Another formulation is that the area of a triangle is the sum of its exterior
angles minus 2π. In this form the formula generalizes to polygons.
Right quadrilaterals with an ideal vertex.
Theorem 15.1 Let Q be a quadrilateral with edges of lengths (a, b, ∞, ∞)
and interior angles (π/2, π/2, π/2, 0). Then we have

sinh(a) sinh(b) = 1.

Proof. First we make a remark in Euclidean geometry: let Q0 be an ideal


hyperbolic quadrilateral, centered at the origin, with sides coming from cir-
cles of Euclidean radii (r, R, r, R). Then rR = 1.
Indeed, from this picture we can construct a right triangle with right-
angle vertex 0, with hypotenuse of length r + R, and with altitude from the
right-angle vertex of 1. By basic Euclidean geometry of similar triangles, we
find rR = 12 = 1.
Now cut Q0 into 4 triangles of the type Q. Then we have sinh(a) = 1/r
and sinh(b) = 1/R, by (15.1). Therefore sinh(a) sinh(b) = 1.

Right hexagons. For a right-angled hexagon H, the excess angle is


6(π/2) − 2π = π, and thus area(H) = π.

Theorem 15.2 For any a, b, c > 0 there exists a unique right hexagon with
alternating sides of lengths (a, b, c).

126
Proof. Equivalently, we must show there exist disjoint geodesics α, β, γ in
H with a = d(α, β), b = d(α, γ) and c = d(β, γ). This can be proved by
continuity.
Normalize so that α is the imaginary axis, and choose any geodesic β at
distance a from α. Then draw the ‘parallel’ line L of constant distance b
from α, on the same side as β. This line is just a Euclidean ray in the upper
half-plane. For each point p ∈ Lp there is a unique geodesic γp tangent to
Lp at p, and consequently at distance b from α.
Now consider f (p) = d(γp , β). Then as p moves away from the juncture
of α and L, f (p) decreases from ∞ to 0, with strict monotonicity since γp ∪L
separates β from γq . Thus there is a unique p such that f (p) = c.

Doubling H along alternating edges, we obtain a pair of pants P . Thus


area(P ) = 2π.

Corollary 15.3 Given any triple of lengths a, b, c > 0, there exists a pair of
pants, unique up to isometry, with boundary components of lengths (a, b, c).

α
L

γ β

Figure 5. Right hexagons.

Pairs of pants decomposition.

Theorem 15.4 Any essential simple loop on a compact hyperbolic surface


X is freely homotopic to a unique simple geodesic. Any two disjoint simple
loops are homotopic to disjoint simple geodesics.

Corollary 15.5 Let X be a compact surface of genus g. Then X can be


cut along 3g − 3 simple geodesics into 2g − 2 pairs of pants. In particular,

127
we have
area(X) = 2π|χ(X)|.

Parallels of a geodesic. There is a nice parameterization of the geodesic


|z| = 1 in H: namely
δ(t) = tanh t + i sech t.
We have kδ 0 (t)k = 1 in the hyperbolic metric.
Now given a closed simple geodesic γ on X, let C(γ, r) be a parallel
curve at distance r from γ. Then we have:
L(C(γ), r) = L(γ) cosh(r).
Indeed, let γ can be covered by the imaginary axis iR+ in H. Then C(γ, r)
is a ray from 0 to ∞ which passes through δ(r). Thus the Euclidean slope
of C(γ, r) is the same as that of the vector (x, y) = (tanh t, sech t). Thus
projection along
p Euclidean horizontal lines from C(γ, r) to γ contracts by
a factor of y/ x2 + y 2 = sech(t). Therefore C(γ, r) is longer than γ by a
factor of cosh(t).
The collar lemma.
Theorem 15.6 Let α and β be disjoint simple geodesics on a compact Rie-
mann surface, of lengths a and b respectively. Define A and B by sinh(a/2) sinh(A) =
1 and sinh(b/2) sinh(B) = 1. Then the collars of widths A and B about α
and β are disjoint.
Proof. We can assume that α and β are part of a pants decomposition of X,
which reduces the result to the case where α and β are two cuffs of a pants
P . By the Schwarz lemma, we can assume the lengths of the boundaries of
P are (a, b, 0).
Now cut P along a simple loop γ that begins and ends at its ideal
boundary component, i.e. the cuff of length zero. Then the components of
P = γ are doubles of quadrilaterals with one ideal vertex and the remaining
angles π/2. The quadrilateral meeting α has finite sides of lengths a/2
and A satisfying sinh(a/2) sinh(A) = 1, and similarly for the quadrilateral
containing β. Thus these collars are disjoint.

Boundary of a collar.
Theorem 15.7 The length of each component of the boundary of the stan-
dard collar around α with L(α) = a satisfies
a2
L(C(α, r))2 = a2 cosh2 (r) = →4
1 + sinh−2 (a/2)

128
as a → 0. Thus the length of each component of the collar about α tends to
2 as L(α) → 0.

Proof. Apply the preceding formulas.


Check. The limiting case is the triply-punctured sphere, the double of
an ideal triangle T ⊂ H with vertices (−1, 1, ∞). The collars limit to the
horocycles given by the circles of radius 1 resting on ±1 together with the
horizontal line segment H at height 2 running from −1 + 2i to +1 + 2i. We
have L(H) = 1, so upon doubling we obtain a collar boundary of length 2.

Corollary 15.8 The collars about short geodesics on a compact hyperbolic


surface cover the thin part of the surface.

Proof. Suppose x ∈ X lies in the thin part — that is, suppose there is a
short essential loop δ through x. Then δ is homotopic to a closed geodesic γ,
which is necessarily simple. But since δ is short, we see by the result above
that δ must lie in the collar neighborhood of γ.

Thick-thin decomposition. There is a universal constant r > 0 such that


any compact Riemann surface of genus g can be covered by a collection of
O(g) balls B(xi , r) and O(g) standard collars about short geodesics.
Bers’ constant.
Theorem 15.9 There exists a constant Lg such that X admits a pants
decomposition with no cuff longer than Lg .

Theorem 15.10 We can take Lg = O(g), but there exist examples requiring

at least one curve of length > C g > 0.

A finite-to-one map to Mg .
Theorem 15.11 For each trivalent graph of G with b1 (G) = g, there is a
finite-to-one map
φG : ((0, Lg ] × S 1 )3g−3 → Mg ,
sending (ri , θi ) to the surface obtained by gluing together pants with cuffs of
lengths ri and twisting by θi , using pants and cuffs corresponding to vertices
and edges of G.
The union of the images of the maps φG is all of Mg .

Corollary 15.12 (Mumford) The function L : Mg → R sending X to


the length L(X) of its shortest geodesic is proper.

129
The Laplacian. Let M be a Riemannian manifold. The Laplace operator
∆ : C0∞ (M ) → C0∞ (M ) is defined so that
Z Z
2
|∇f | = f ∆f,
M M

both integrals taken with respect to the volume element on M .


For example, on R we find ∆f = −d2 f /dx2 by integrating by parts.
Similarly on Rn we obtain
X d2 f
∆f = − ·
dx2i

Note this is the negative of the ‘traditional’ Laplacian.


In terms of the Hodge star we can write
Z Z Z Z
h∇f, ∇f i = df ∧ ∗df = − f ∧ d ∗ df = f (− ∗ d ∗ df ) dV,

and therefore we have


∆f = − ∗ d ∗ df.
For example, on a Riemann surface with a conformal metric ρ(z)|dz|, we
have ∗dx = dy, ∗dy = −dx, and
 2
d2 f

−2 d f
∆f = −ρ (z) + 2 .
dx2 dy

As a particular case, for ρ = |dz|/y on H we see ∆y α = α(1 − α)y α , showing


that y α is an eigenfunction of the hyperbolic Laplacian.
The heat kernel. Let X be a compact hyperbolic surface. Enumerat-
ing the eigenvalues and eigenfunctions of the Laplacian, we obtain smooth
functions satisfying
∆φn = λn φn ,
λn ≥ 0. The heat kernel Kt (x, y) is defined by
X
Kt (x, y) = e−λn t φn (x)φn (y),

The heat kernel is the fundamental solution to the heat equation. That
is, for any smooth function f on X, the solution to the heat equation
dft
= −∆ft
dt

130
P
with initial data f0 = f is given by ft = Kt ∗ f . Indeed, if f (x) = an φn (x)
then X
ft (x) = Kt ∗ f = an e−λn t φn (x)
clearly solves the heat equation and has f0 (x) = f (x).
Note also that formally, convolution with Kt is the same as the operator
exp(−∆), which acts by exp(−λn ) on the λn -eigenspace.
Brownian motion. The heat kernel can also be interpreted using diffu-
sion; namely, Kt (x, ·) defines a probability measure on X that gives the
distribution of a Brownian particle xt satisfying x0 = x.
For example, on the real line, the heat kernel is given by
1
Kt (x) = √ exp(−x2 /(4t)).
4πt
Also we have Ks+t = Ks ∗ Kt , as befits a Markov process. R
To check this, note that Kt solves the heat equation, and that Kt = 1
for all t. Thus Kt ∗ f → f as t → 0, since Kt concentrates at the origin.
In terms of Brownian motion, the solution to the heat equation is given
by ft (x) = E(f0 (xt )), where xt is a random path with x0 = x.
The trace. The trace of the heat kernel is the function
Z X
Tr Kt = Kt (x, x) = e−λn t .
X

It is easy to see that the function Tr Kt determines the set of eigenvalues λn


and their multiplicities.
Length spectrum and eigenvalue spectrum.

Theorem 15.13 The length spectrum and genus of X determine the eigen-
values of the Laplacian on X.

Proof. The proof is based on the trace of the heat kernel. Let kt (x, y)
denote the heat kernel on the hyperbolic plane H; it satisfies kt (x, y) = kt (r)
where r = d(x, y). Then for X = H/Γ we have
X
Kt (x, y) = kt (x, γy),
Γ

where we regard Kt as an equivariant kernel on H.

131
Working more intrinsically on X, we can consider the set of pairs (x, δ)
where δ is a loop in π1 (X, x). Let `x (δ) denote the length of the geodesic
representative of δ based at x. Then we have:
X
Kt (x, x) = kt (`x (δ)).
δ

Let L(X) denote the space of nontrivial free homotopy classes of maps
γ : S 1 → X. For each γ ∈ L(X) we can build a covering space p : Xγ → X
corresponding to hγi ⊂ π1 (X).
The points of Xγ correspond naturally to pairs (x, δ) on X with δ freely
homotopic to γ. Indeed, given x0 in Xγ , there is a unique homotopy class
of loop δ 0 through x0 that is freely homotopic to γ, and we can set (x, δ) =
(p(x), p(δ 0 )). Conversely, given (x, δ), from the free homotopy of δ to γ
we obtain a natural homotopy class of path joining x to γ, which uniquely
determines the lift x0 of x to Xγ .
For x0 ∈ Xγ , let r(x0 ) denote the length of the unique geodesic through
x that is freely homotopic to γ. Then we have `x (δ) = r(x0 ). It follows that
0

Z Z XZ
Tr Kt = Kt (x, x) = kt (0) + r(x0 ).
X X L(X) Xγ
R
But X kt (0) = area(X)kt (0) depends only on the genus of X by Gauss-
Bonnet, and the remaining terms depend only on the geometry of Xγ . Since
the geometry of Xγ is determined by the length of γ, we see the length spec-
trum of X determines the trace of the heat kernel, and hence the spectrum
of the Laplacian on X.

Remark. Almost nothing was used about the heat kernel in the proof.
Indeed, the length spectrum of X determine the trace of any kernel K(x, y)
on X derived from a kernel k(x, y) on H such that k(x, y) depends only on
d(x, y).
Remark. In fact the genus is determined by the length spectrum.
Isospectral Riemann surfaces.
Theorem 15.14 There exist a pair of compact hyperbolic Riemann surfaces
X and Y , such that the length spectrum of X and Y agree (with multiplici-
ties), but X is not isomorphic to Y .

Isospectral subgroups. Here is a related problem in group theory. Let G


be a finite group, and let H1 , H2 be two subgroups of G. Suppose |H1 ∩C| =

132
|H2 ∩ C for every conjugacy class C in G. Then are H1 and H2 conjugate
in G?
The answer is no in general. A simple example can be given inside
the group G = S6 . Consider the following two subgroups inside A6 , each
isomorphic to (Z/2)2 :

H1 = he, (12)(34), (12)(56), (34)(56)i,


H2 = he, (12)(34), (13)(24), (14)(23)i.

Note that the second group actually sits inside A4 ; it is related to the sym-
metries of a tetrahedron.
Now conjugacy classes in Sn correspond to permutations of n, i.e. cycle
structures of permutations. Clearly |Hi ∩ C| = 3 for the cycle structure
(ab)(cd), and |Hi ∩ C| = 0 for other conjugacy classes (except that of the
identity). Thus H1 and H2 are isospectral. But they are not conjugate
(‘internally isomorphic’), because H1 has no fixed-points while H2 has two.
Construction of isospectral manifolds.
Theorem 15.15 (Sunada) Let X → Z be a finite regular covering of com-
pact Riemannian manifolds with deck group G. Let Yi = X/Hi , where H1
and H2 are isospectral subgroups of G. Then Y1 and Y2 are also isospectral.

Proof. For simplicity of notation we consider a single manifold Y = X/G


and assume the geodesics on Z are discrete, as is case for a negatively curved
manifold. Every closed geodesic on Y lies over a closed geodesic on Z.
Fixing a closed geodesic α on Y , we will show the set of lengths L of
geodesics on Y lying over α depends only on the numbers nC = H ∩ C for
conjugacy classes C in G.
For simplicity, assume α has length 1. Let α1 , . . . , αn denote the com-
ponents of the preimage of α on X. Let Si ⊂ G be the stabilizer of αi . The
subgroups Si fill out a single conjugacy class in G, and we have Si ∼ = Z/m
where nm = |G|. Each loop αi has length m.
Let k be the index of H ∩ Si in Si . Then k is the length of αi /H in Y .
Moreover, the number of components αj in the orbit H · αi is |H|/|H ∩ Si | =
|H||Si |/k, and of course all these components descend to a single loop on
X/H. Thus the number of times the integer k occurs in L is exactly
kAk
|L(k)| = ,
m|H|
where
Ak = |{i : [Si : Si ∩ H] = k}|.

133
Thus to determine L, it suffices to determine the integers Ak .
For example, let us compute A1 , the number of i such that we have
Si ⊂ H. Now H contains Si if and only if H contains a generator gi of Si .
We can choose the gi ’s to fill out a single conjugacy class C, since the groups
Si are all conjugate. Then the proportion of i’s satisfying Si ⊂ H is exactly
|H ∩ C|/|C|, and therefore

n|H ∩ C|
A1 = ·
|C|
An important point here: it can certainly happen that Si = Sj even
when i 6= j. For example if G is abelian, then all the groups Si are the
same. But the number of i such that Si is generated by a given element
g ∈ C is a constant, independent of g. Thus the proportion of Si generated
by an element of H is still |H ∩ C|/|C|.
Now for d|m, let Cd be the dth powers of the elements in C. Then the
subgroups of index d in the Si ’s are exactly the cyclic subgroups generated
by elements g ∈ Cd . Again, the correspondence is not exact, but constant-
to-one; the number of i such that hgi ⊂ Si is independent of g ∈ Cd . Thus
the proportion of Si ’s such that H ∩ Si contains a subgroup of index d is
exactly |H ∩ Cd |/|Cd |, which implies:
X n|H ∩ Cd |
Ak = ·
|Cd |
k|d

From these equations it is easy to compute Ak .

Cayley graphs. The spaces in Sunada’s construction do not have to be


Riemannian manifolds. For example, we can take Z to be a bouquet of
circles. Then X is the Cayley group of G, and Y1 and Y2 are coset graphs
on which G acts. The coset graphs Y1 and Y2 also have the same length
spectrum!
A small example. Let G = (Z/8)∗ n Z/8 be the affine group of A = Z/8,
i.e. the group of invertible maps f : A → A of the form f (x) = ax + b. Let

H1 = {x, 3x, 5x, 7x},


H2 = {x, 3x + 4, 5x + 4, 7x}.

Then the subgroups H1 and H2 are isospectral.


In both cases, the coset space Yi = G/Hi can be identified with Z/8;
that is, Z/8 × Hi = G.

134
To make associated graphs, Y1 and Y2 , we take hx + 1, 3x, 5xi as gener-
ators for G. Note that 3x and 5x have order 2. Then the coset graph Y1 is
an octagon, coming from the generator x + 1, with additional (unoriented,
colored) edges joining x to 3x and 5x. Similarly, Y2 is also an octagon, but
now the colored edges join x to the antipodes of 3x and 5x, namely 3x + 4
and 5x + 4.

Figure 6. Isospectral graphs

These graphs are isospectral. In counting the number of loops, it is


important to regard the graphs as covering spaces. For this it is best to
replace each colored edge which is not a loop by a pair of parallel edges with
opposite arrows. Each colored loop should be replaced by a single oriented
edge. Then the graphs become covering spaces of the bouquet of 3 circles,
and the number of loops of length n is the same for both graphs.
Not isometric. Using short geodesics, we can arrange Z such that one
can reconstruct the action of G on G/Hi from the intrinsic geometry of Yi .
Then Y1 and Y2 are isometric iff H1 and H2 are conjugate. So in this way
we obtain isospectral, but non-isometric, Riemann surfaces.

16 Uniformization
Theorem 16.1 Every simply connected Riemann surface is isomorphic to
b C or H.
C,

Poincaré series. We will now show that the uniformization theorem im-
plies the Riemann existence theorem; for example, it implies that every
compact Riemann surface can be presented as a branched cover of C.
b

135
For surfaces of genus 0 this is clear. For genus one, uniformization implies
X = C/Λ
b and then the construction of the Weierstrass ℘-function completes
the proof.
For surfaces of higher genus we have X = ∆/Γ. Here the idea is the
following: given a meromorphic form ω = ω(z) dz k on the unit disk, we
generate a Γ-invariant form by summation (just as one would do for the
℘-function):
X
Θ(ω) = γ ∗ ω.
Γ

More explicitly, this form is given by


X
Θ(ω) = ω(γ(z))(cz + d)−2k dz k ,
Γ

since γ 0 (z) = (cz + d)−2 when γ(z) = (az + b)/(cz + d) is normalized so that
ad − bc = 1.
There is no chance that this sum will converge when ω is a nonzero
function. However, when k = 2, |ω| is naturally an area form, and hence
Z Z
|Θ(ω)| ≤ |ω|.
X ∆

This shows
R the sum converges (and defines Ra meromorphic form on X) so
long as |ω| < ∞. In fact, it is enough that |ω| is finite outside a compact
set K ⊂ ∆.
A second problem is that Θ(ω) might be zero. Indeed, if ω = α − γ ∗ α,
then the sum will telescope and give zero. As in the case of the ℘ function,
we can insure the sum is nonzero by introducing a pole. That is, if a ∈ ∆
projects to p ∈ X, then
dz 2
 
Ωp = Θ
z−a
gives a meromorphic quadratic differential with a simple pole at p and else-
where holomorphic on X. It follows that for p 6= q,

fpq = Ωp /Ωq

defines a nonconstant meromorphic function on X; indeed, fpq (p) = ∞ and


fpq (q) = 0.
The multiplicities of these zeros and poles is not quite clear; for example,
Ωq might vanish at p, and then fpq has at least a double order pole at p.

136
But if p is close enough to q, then Ωq (p) 6= 0, and hence fpq provides a local
chart near p.
With a little more work we have established:

Corollary 16.2 Every Riemann surface carries a nonconstant meromor-


phic function, indeed, enough such functions to separate points.

Quasiconformal geometry. Here is another approach to the uniformiza-


tion theorem. For any µ on Cb with kµk∞ < 1, there exists a quasiconformal
map f : C → C with complex dilatation µ.
b b
Evidentally we can uniformize at least one Riemann surface Xg of genus
g, e.g. using a regular hyperbolic 4g-gon. Now take any other surface Y
of the same genus. By topology, there is a diffeomorphism f : Xg → Y .
Pulling back the complex structure to Xg and lifting to the universal cover,
we obtain by qc conjugacy a Fuchsian group uniformizing Y .

References
[EKS] A. Edmonds, R. Kulkarni, and R. E. Stong. Realizability of branched
coverings of surfaces. Trans. Amer. Math. Soc. 282(1984), 773–790.

[Gol] G. M. Goluzin. Geometric Theory of Functions of a Complex Vari-


able. Amer. Math. Soc, 1969.

[GH] P. Griffiths and J. Harris. Principles of Algebraic Geometry. Wiley


Interscience, 1978.

[Hel] S. Helgason. Differential Geometry, Lie Groups, and Symmetric


Spaces. Academic Press, 1978.

[KZ] M. Kontsevich and D. Zagier. Periods. In Mathematics Unlimited —


2001 and Beyond, pages 771–808. Springer-Verlag, 2001.

[La] S. Lang. Algebra. Addison-Wesley, 1984.

137

You might also like