Advanced Complex Analysis Course
Advanced Complex Analysis Course
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
5. Complex geometry: Sheaf theory; several complex variables; Hodge
theory.
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
3
The factor of 1/2 is not unexpected: for real projective space, the map
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:
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
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:
A(r) ∼
= H/λZ ,
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, ∞}
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.)
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.
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,
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:
10
1. f is closed: i.e. E closed implies f (E) closed.
3. If D ⊂ X is discrete, so is f (D).
f : X∗ → Y ∗
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.
Theorem 2.2 The complex structure extends from X ∗ to X iff every point
x ∈ D has a neighborhood U such that U ∩ X ∗ ∼
= ∆∗ .
Gal(X/Y ) = {g ∈ Aut(X) : f ◦ g = f }.
12
1. Any analytic map between compact Riemann surfaces is proper.
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 .
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.)
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.)
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
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).
f : X ∗ → Y ∗.
f : X → Y = X/G
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.
χ(X ∗ ) = 2 − 2g − n.
Examples.
0 = 2 · 2 − |B(℘)|
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
2 − 2g = 2 · 2 − 2n,
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.
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.
Examples.
19
1. A star with d endpoints, and its center labeled zero, corresponds to
p(z) = z d .
• A monodromy map φ : π1 (C
b − B, y) → Sd .
π1 (X ∗ , xi ) ∼
= {γ ∈ π1 (Y ∗ , y) : ρ(γ)(i) = i}.
γ1 · · · γn = e.
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 ).
0 1 0 1
a e
c
b d
f
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.
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.
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.
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.
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
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).
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.
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:
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:
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.
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.
(−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.
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
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
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 ).
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.
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.
which is canonical.
Puiseux series.
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.
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 .
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
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 ).
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.
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)).
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:
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.
(1)
Corollary 5.2 The vector space Tp = mp /m2p is isomorphic to C2 .
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
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
T (z) = az + bz,
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
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).
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.
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:
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.
√
zk dz 1/ p
B
e ≥0 +1 −1
∞ −k −2 +n
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
∗ : Tp(1) X → Tp(1) X
ω = α + β,
48
Positivity. It is useful to measure the size of a 1-form by the natural 2-form
|ω|2R = ω ∧ ∗ω.
|dz|2R = 2 dx dy.
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
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
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
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
d2 f i
∆f = 2i∂∂f = 4 · dz dz.
dz dz 2
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
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).
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)
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.
d∗ (α) = ± ∗ d ∗ α
d∗ β = (−1)k ∗ d ∗ β.
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.
and so dα = d ∗ α = 0 as well.
∆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.
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)∗
UI = Ui0 ∩ · · · ∩ Uiq .
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 0 (U, F) = F(X).
where the limit is taken over the system of all open coverings, directed by
refinement.
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
as desired.
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.
H q (X, F) ∼
= H q (U, F).
(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.
(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
0 → A → B → C → 0,
60
gives rise to a long exact sequence
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).
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
I − π ∗ ◦ s∗ = Kd + dK.
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
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.
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.
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,
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
The penultimate minus sign comes from the back that S 1 (0, r) acquires a
negative orientation as ∂(C − B(0, r)).
Proof. First solve dh/dz = g, then df /dz = h. Then ∆(f /4) = d2 f /dz dz =
g.
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 ).
Similarly, we define
66
Corollary 7.8 The dimension of H 1 (X, Ω) is ≥ 1.
Theorem 7.9 For any compact complex manifold X, we have H∂p,q (X) ∼
=
H q (X, Ωp ).
H 1 (X, O) ∼
= H 0,1 (X) = E (0,1) (X)/∂E 0 (X).
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
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
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
H 1,1 (X)∗ ∼
= (∂E 1,0 (X))⊥ ⊂ D0 (X),
and we have Z
f ∂α = 0
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:
0 → O∗ → M∗ → M∗ /O∗ → 0,
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.
OK ∼
=Ω
(hω) = (h) + K ≥ 0
iff h ∈ OK .
Similarly, for any divisor D we let
ΩD ∼
= OK+D .
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 .
Equivalently, we have
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):
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.
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.)
Proof. Take ω = df .
h1 (K) = 2(ga − g) + 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.
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.
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,
∂
0 → Ω → E 1,0 → E 1,1 → 0
H 1 (X, Ω) ∼
= H 1,1 (X) ∼
= C,
76
and that the residue map, defined by
Z
1
Res(ω) = ω,
2πi X
H 1,1 (X) ∼
= C.
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 ∗
77
Corollary 9.2 There exists a meromorphic 1-form with prescribed principal
parts iff the sum of its residues is equal to zero.
∂
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
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:
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
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:
80
Proof of Serre duality: dimension counts. We first show
Res(fi ω) = 1,
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
81
Theorem 10.3 For deg(D) > 0, we have
O−D ⊗ ΩD → Ω
φ(ξ) = Res(ξω).
H 1 (X, OD ) → H 1 (X, OE ) → 0
for any E ≥ 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).
We can now round out the discussion by proving some results promised
above.
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. 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.
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,
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 .
H 2k (Pn , Z) = Z for k = 0, 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
K(Pn ) ∼
= C(z1 , . . . , zn ).
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).
ν : X ∗ → V (f ) − C ⊂ P2
t2 x2 = x2 (1 − x)
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.
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 );
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 .
1. |D| is basepoint–free.
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.
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|
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.
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
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.
h0 (K − Pk ) = 1 − g + (2g − 2) − k + h0 (Pk ).
h0 (K − Pk ) = h0 (K) − k,
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)
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:
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 .
h0 (K − D) − 1 = dim |K − D|
= dim{H ⊂ Pg−1 : φ(Pi ) ∈ H, 1 ≤ i ≤ n}.
94
Proof. By Riemann-Roch, we have
1. P is a Weierstrass point.
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
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.
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).
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:
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.
|dH| → |dK|.
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
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
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.
(t) = (f s) = (f ) + D ≥ 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:
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
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 13.8 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.
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 .
105
Theorem 13.12 Every smooth plane curve of degree d has genus g = (d −
1)(d − 2)/2.
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:
ω ∧ df = W (x, y) dx 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
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 ?
h0 (K − E) = g − deg E.
h0 (E) = 1 − g + deg(E) − h0 (K − E) = 1,
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).
KX ∼
X
= O(−n − 1 + di ).
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.
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
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:
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 .
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.
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
∂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, ∞].
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
Taking the product of the solutions for several pairs of points, and using
additivity of the logarithmic derivative, we obtain:
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
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 =
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
[α, ω] = 0 ∀ω ∈ Ω(X).
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.
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
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.
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.
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.
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 ,
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
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
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
122
Qn Qn
Thus if 1 ai = 1 bi , the function
fa (ζ)
F (ζ) =
fb (ζ)
Then we set
2
X
θ(ζ) = qn ζ n,
n∈Zg
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).
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
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
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)
sinh(a) sinh(b) = 1.
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.
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
127
we have
area(X) = 2π|χ(X)|.
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. 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.
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 .
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
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
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
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 .
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 :
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.
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
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.
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
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
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:
References
[EKS] A. Edmonds, R. Kulkarni, and R. E. Stong. Realizability of branched
coverings of surfaces. Trans. Amer. Math. Soc. 282(1984), 773–790.
137