0% found this document useful (0 votes)
244 views331 pages

Algebraic Geometry

The document outlines lecture notes on algebraic geometry, beginning with introductory examples of algebraic curves in the plane. It then discusses how points on a conic curve can be constructed through five generic points using linear algebra and properties of the evaluation map. The chapter also explains how Pascal's mystic hexagon theorem allows constructing additional points on a smooth conic through initial generic points using only a straightedge.

Uploaded by

alin444444
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
244 views331 pages

Algebraic Geometry

The document outlines lecture notes on algebraic geometry, beginning with introductory examples of algebraic curves in the plane. It then discusses how points on a conic curve can be constructed through five generic points using linear algebra and properties of the evaluation map. The chapter also explains how Pascal's mystic hexagon theorem allows constructing additional points on a smooth conic through initial generic points using only a straightedge.

Uploaded by

alin444444
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 331

Lecture Notes

Algebraic Geometry III/IV

Matt Kerr

Contents

Part 1.

Introduction and Motivation

Chapter 1.

Two theorems on conics in the plane

Chapter 2.

Riemann surfaces and algebraic curves

21

Chapter 3.

The normalization theorem

39

Chapter 4.

Lines, conics, and duality

49

Part 2.

General denitions and results

63

Chapter 5.

Complex manifolds and algebraic varieties

65

Chapter 6.

More on projective algebraic varieties

75

Chapter 7.

Smooth varieties as complex manifolds

87

Chapter 8.

The connectedness of algebraic curves

97

Chapter 9.

Hilbert's nullstellensatz

107

Chapter 10.

Local analytic factorization of polynomials

113

Chapter 11.

Proof of the normalization theorem

121

Chapter 12.

Intersections of curves

131

Chapter 13.

Meromorphic 1-forms on a Riemann surface

143

Chapter 14.

The genus formula

155

Chapter 15.

Some applications of Bzout

167

Part 3.

Cubic curves

173

Chapter 16.

The singular cubic

175

Chapter 17.

Putting a nonsingular cubic in standard form

187

CONTENTS

Chapter 18.

Canonical normalization of the Weierstrass cubic

195

Chapter 19.

Group law on the nonsingular cubic

207

Chapter 20.

Abel's theorem for elliptic curves

219

Chapter 21.

The Poncelet problem

225

Chapter 22.

Periods of families of elliptic curves

241

Chapter 23.

Counting

255

Part 4.

Fp -rational

points on elliptic curves

Curves of higher genus

263

Chapter 24.

The algebraicity of global analytic objects

265

Chapter 25.

The Riemann-Roch Theorem

273

Chapter 26.

Applications of Riemann-Roch, I: special Riemann


surfaces

Chapter 27.

281

Applications of Riemann-Roch, II: general Riemann


surfaces

291

Chapter 28.

Abel's Theorem, part I

303

Chapter 29.

Abel's Theorem, part II

311

Appendix: genera of singular curves

325

Part 1

Introduction and Motivation

CHAPTER 1

Two theorems on conics in the plane

The primary subject of this course is algebraic curves, the simplest


example of which is the solution set of a two-variable polynomial equa-

tion in the plane. What plane?


this course, mainly

C2 ,

R2 , C2 , F2p , Q2 . . .?

For the purposes of

with excursions into the others.

Remark 1.0.1. A perennial point of confusion is whether to call

the complex plane.

This is henceforth forbidden! It is the complex (ane) line, and a real

C
= R2 . (A complex plane will mean something
2
2
over C, so C will be the complex ane plane, P the

(ane) plane via


2-dimensional

complex projective plane, and so forth.

We'll worry about ane vs.

H := {x + iy | x, y R, y >
plane in C; that terminology is

projective in the next chapter.) Note that

0} C

will denote the upper-half

unavoidable.

The objects which shall concern us, then, will be 1-dimensional over

C (complex algebraic curves), hence 2-dimensional over R (Riemann


surfaces). Our approach will be quite intuitive and visual for the rst
few chapters, to get an idea of what algebraic geometry is before settling
into a more measured approach. My feeling has always been that you
need motivation for introducing formalism, in this case for layering lots
of algebra onto geometry. In this chapter that motivation might consist
of the subtle gaps that open as we try to prove some famous results on
conics from (mostly) linear algebra.
7

1. TWO THEOREMS ON CONICS IN THE PLANE

Example 1.0.2. (a) Consider the set of rational points on the Fer-

mat quartic (degree 4) curve:

{(x, y) Q2 | x4 + y 4 = 1} Q2 .

(1.0.1)

By a case of Fermat's last theorem, (1.0.1) is the empty set, since



4
a 4
+ dc = 1 means (ad)4 + (bc)4 = (bd)4 with ad, bc, bd Z and
b

bd 6= 0.
(b) Next we look at the Fermat cubic (degree 3) curve

{(x, y) C2 | x3 + y 3 = 1} C2 .

(1.0.2)

We will see that (1.0.2) has the structure of a complex


on the left-hand side in

1-torus,

shown

1+

The right-hand side represents the quotient


by the lattice
phism?
any

(for some

It holds topologically (convince yourself of this visually) for

H,

where

:= Z h1, i

C/ of the complex line


H). What is the isomor-

but complex analytically only for the values

a b
c d

!
SL2 (Z) and 0 :=

a0 +b
c0 +d

1+ 3
. We call (1.0.2) an elliptic
2

curve: it isn't an ellipse in any sense, but originally arose in connection


with the arc-length of one.
(c) Finally, take the real degree 2 (a.k.a. quadric or conic) curve
(1.0.3)

{(x, y) R2 | x2 + y 2 = 1} R2 ,

which is of course a circle:

First we shall pursue conics to get a preliminary feel for the interplay
between algebra and geometry. These aren't too hard to visualize: you
know what the real solution sets look like (ellipses, hyperbolas, pairs

1.1. ALGEBRAIC CURVES IN

R2

of lines, etc.) and the complex solutions are all spheres once we add
points at innity.

1.1. Algebraic curves in


Let

P2n

R2

n in x and y . (If
n
write P .) In an exercise

denote the real polynomials of degree

there is no possibility for confusion, I'll just


below, you are asked to prove that

P2n is a real vector space of dimension


n+2
2

Example 1.1.1. (a) A basis for

dim(P2 ) = 6.
3
(b) For P , the
dimension is 10).

P2

is given by

{1, x, y, xy, x2 , y 2 },

so

basis is

{1, x, y, xy, x2 , y 2 , x3 , x2 y, xy 2 , y 3 }

(and the

For a conguration of distinct points

S = {p1 , . . . , pm } R2 ,
dene a linear  evaluation map

evSn : P n Rm
via

f (p1 )
.
.
.

f 7

f (pm )
Recall from linear algebra the

Rank + Nullity theorem:

dim(image(evS )) + dim(ker(evS )) = dim(P n ).


|
{z
}
|
{z
}

(1.1.1)

rank

nullity

Definition 1.1.2. The conguration of points

evSn

is surjective (onto).

Now if

evSn

is surjective, its rank is

space of polynomials vanishing on

S.

m;

while its kernel is just the

By (1.1.1) we have:

n polynomials vanishing

n+2
dimension
m.
2

Proposition 1.1.3. The space of degree

on a general conguration of

1or

S is called n-general 1

points has

just general or generic if the context is understood

10

1. TWO THEOREMS ON CONICS IN THE PLANE

What does all this have to do with algebraic curves? Well, given a
polynomial in

Pn ,

I can look at its solution set in

2
we have the assignment

P n \P n1

{degree n

R2 .

More precisely,

real ane algebraic curves}

given by

f 7 Cf := {(x, y) R2 | f (x, y) = 0}.


Notice that

f ker(evS ) Cf S.

(1.1.2)

Proposition 1.1.4. Through ve (2-)general points

in the real plane, there exists a unique conic


Proof.

Q.

6 5 = 1.

OK, OK, so what the proof means is:


of degree

A, B, C, D, E

by Prop.

1.5, the space

polynomials vanishing on a general conguration of

2+2
2

5=1. So given two degree-2 polynomials


f, g P 2 vanishing at all 5 points in S , we must have g = a f (for
some a R). But f = 0 and a f = 0 dene the same curve Q; and
by (1.1.2) Q contains A, B, C, D, E .
There are two issues with this. First, in order for f to dene a conic,
we need to know that the 1-dimensional space of solutions doesn't lie in
P 1 (inside P 2 ). But if a linear (degree 1) polynomial h vanishes at all 5
2
2
2
points, so does h ; and then since h, h P are not linearly dependent,
evS2 doesn't have maximal rank, contradicting the genericity condition
on S . (At this point, Proposition 1.6 is completely proved as stated.)

points has dimension

The second problem is more interesting: what exactly does it mean

A, B, C, D, E to be 2-generic? If you think about our little proof, the


existence of Q has nothing to do with this genericity, but the uniqueness has everything to do with it. Moreover, the statement that  Q is
unique if S is generic is somewhat circular without knowing (beyond
Denition 5) what the genericity condition is: namely, that no 4 of
for

the points are collinear. It's best to wait until we're doing projective
geometry to prove that, and so we will.

2notation: A\B

denotes set-theoretic exclusion, sometimes also written

Ane just means that the curves are in

R2 .

A B.

1.2. BLAISE PASCAL AND THE MYSTIC HEXAGON

11

Remark 1.1.5. More generally, the abstract ethos surrounding the

word general in algebraic geometry is that you are working in the


complement of nitely many algebraic conditions. The algebraic condition we are avoiding in this section is the vanishing of all

mm

n
minors in a matrix representing evS .

1.2. Blaise Pascal and the mystic hexagon


Similarly, given eight points

A, B, C, D, E, P1 , P2 , P3
in (3-)general position, the vector space of cubic polynomials vanishing
at all of them has dimension

3+2
2


8 = 2.

V , and write (xA yA ), . . . , (xP3 , yP3 ) for the coor3


points. V lies in P , which consists of elements of the

Call this vector space


dinates of given
form

a1 + a2 x + a3 y + + a9 xy 2 + a10 y 3 = fa (x, y).


Asking a polynomial of this form to vanish at our

points yields the

equations

0 =

a1 + a2 xA + a3 yA + + a9 xA yA2 + a10 yA3

.
.
.

i.e.

0 = a1 + a2 xP3 + a3 yP3 + + a9 xP3 yP2 3 + a10 yP3 3

linear constraints on the

10

variables

{aj }

expressing what it

fa to lie in V .
Now let A, B, C, D, E be (2-)general and Q the unique conic through

means for

them:

B
Q
C
A

D
E

In fact, we shall make the stronger assumption that no three of

Q is smooth (an ellipse or hyperbola, not a pair of


We would like to construct (arbitrarily many) points on Q using

are collinear, so that


lines).

A, B, C, D, E

12

1. TWO THEOREMS ON CONICS IN THE PLANE

only a straightedge. Start by drawing secant lines connecting adjacent


points:

B
Q
C
A

D
BC

E
AB

CD
DE

P1

AB DE =: P1 .3 Next, draw any line ` through


through B, C, D, or E , and set ` CD =: P2 .

where we have labelled

that does not pass

C
A

D
BC

E
AB
l

CD

DE

P
1

P
2

(Note that our choice of

constructing.) Next, draw

3Of

q Q we
P1 P2 BC =: P3 .

will determine the point

P1 P2 ,

and label

end up

course, the lines may be parallel. There are two ways to x this: either make

A, B, C, D, E

more general so that none of the lines we intersect are parallel; or

work projectively. Since this chapter is entirely motivational we won't worry about
that level of detail..

1.2. BLAISE PASCAL AND THE MYSTIC HEXAGON

13

C
A

D
BC

E
AB
l

CD

DE
P P
1 2

P
1

Finally, we draw

P
3

P
2

EP3

and set

q := EP3 `.

C
A

D
q

BC

AB
l

CD

EP3

DE

P
1

P
3

P
2

Proposition 1.2.1.

q Q.

To prove this, we shall require:

Lemma 1.2.2.

when

A, B, C, D, E

A, B, C, D, E, P1 , P2 , P3

is in (3-)general position,

are general in the strong sense assumed above.

We won't prove the lemma here. (It is a special case of something


called the Cayley-Bacharach Theorem, which will be easy to prove once
we know a little about residues in algebraic geometry.)

fAB (x, y) = 0 for the (linear)


AB , fQ (x, y) = 0 for the conic Q, and so forth.

Proof. [of Prop.

equation of the line

1.8] We write

Consider the three cubic polynomials

f1 (x, y) := fQ (x, y) fP1 P2 (x, y),


f2 (x, y) := fAB (x, y) fCD (x, y) fEP3 (x, y),
f3 (x, y) := f` (x, y) fBC (x, y) fDE (x, y).

14

1. TWO THEOREMS ON CONICS IN THE PLANE

Their vanishing sets are

C1 := Q P1 P2 ,
C2 := AB CD EP3 ,
C3 := ` BC DE,
each of which contains the set

S := {A, B, C, D, E, P1 , P2 , P3 }.
f1 , f2 , f3 belong to V := ker(evS3 ). Since the dimen2, f1 , f2 , f3 cannot be linearly independent and we have a

So by (1.1.2),
sion of

is

4
nontrivial relation

f1 + f2 + f3 = 0

(1.2.1)
with real coecients.
Suppose

=0

in (1.2.1). Then

f2 = f3 ,

so that

f2

and

f3

are

proportional hence cut out the same curve:

AB CD EP3 = ` BC DE.
` = AB, CD or EP3 , which implies ` contains A, B, C, D,
E  a contradiction to our choice of `!
So 6= 0, and we may rewrite (1.2.1) as

But this means


or

f1 = f2 f3 .

Now since ` and EP3 both contain q , f2 and f3 both vanish at q . By


(1.2.2), f1 also is zero at q , so one of its factors has to be. Therefore q
is contained in Q or P1 P2 .
Suppose q P1 P2 . Then the lines P1 P2 , EP3 , and ` collapse to
the same line (look at the last diagram) and so in particular E `,
again in contradiction to our choise of `.
We conclude that q Q.

(1.2.2)

In the construction described pictorially above,

Proposition 1.9  ultimately just the point where

have proved the beautiful statement:

4not

meets

Q.

Since

` was essentially free, A, B, C, D, E, and q can be thought


6 distinct but otherwise arbitrary points of Q. Consequently, we

our choice of
of as

was  in light of

all of

, ,

are zero

1.3. PONCELET'S PORISM

15

Theorem 1.2.3. [B. Pascal, 1639] Intercepts of opposite edges

of a hexagon inscribed in a conic, lie on a line.


When we get to the notion of duality in projective geometry, Theorem 1.10 will dualize for free to:
Corollary 1.2.4. The (three) lines joining opposite vertices of a

hexagon circumscribed about a conic, meet in a single point.

1.3. Poncelet's Porism


According to my dictionary, a porism is a proposition that uncovers
the possibility of nding such conditions as to make a specic problem

5 The result of Poncelet I'll describe


6
here just had a whole book devoted to it, and in the late 1970's P.

capable of innumerable solutions.

Griths (my Ph.D. advisor) and J. Harris devoted two nice articles to
it.

If your local pub had an ellipse-shaped pool table it would even

have practical applications.

C and D be two conics in R2 . They are the vanishing sets


of some fC , fD P2 . For simplicity, assume they are ellipses, with D
contained in the interior of C .
So let

Problem: Does there exist a closed polygon (self-intersections are


OK) inscribed in

and circumscribed about

D?

Solution: Sometimes. But existence of one such polygon implies


that there are innitely many.

This is clearly a porism.

We shall call such polygons as in the

Problem circuminscribed when a specic pair

C ,D

is understood.

The precise statement is:


Theorem 1.3.1. [Poncelet, 1822] If the pair

circuminscribed polygon, then for any point on


scribed

n-sided

polygon with one vertex on

C, D has an n-sided
there is a circumin-

C.

Another way of putting this is that circuminscribed polygons can


be rotated continuously: there are beautiful pictures of this at http://

5rather than a complication of the swine


6L. Flatto, Poncelet's Theorem, AMS,

u
2009.

It's on reserve in the library and

aimed at advanced undergraduates, i.e. you. Highly recommended.

16

1. TWO THEOREMS ON CONICS IN THE PLANE

enriques.mathematik.uni-mainz.de/intgeo/poncelet.html.

While this

picture

x0

START

etc.

L0

Poncelet
Construction

D
x2

L1

x1

is to those ones as the math building (at virtually any university) is


to Durham cathedral, it allows us to characterize Theorem 1.12 in one
more way:
(1.3.1)

If

xn = x0

for any

x0 ,

then

xn = x 0

for every

x0 .

We will skirt around projective geometry in explaining the idea


here, but can't avoid
solutions to

fC = 0

C.

Henceforth,

and

fD = 0

and

 that is,

shall denote all complex

C, D C2 .

Topologi-

cally these are real surfaces (in fact, spheres with one or two missing
points), and are complex-analytically isomorphic to

or

C ,

but it

won't hurt to draw them schematically as real curves on a sheet of


paper. You can think of this as the real solutions standing in for the
complex ones. In general, when we want to see the topology of a complex algebraic curve, we'll draw a surface; when we want to see how
dierent curves intersect or how they lie in space, we'll draw a curve.
Consider the set of pairs

E := {(x, L) C D | x L},
where

is the set of lines tangent to

D.

An involution is a map

which, applied twice, gives the identity. Here are two involutions on

E:

1.3. PONCELET'S PORISM

17

1(x,L):=(x,L)

(1) =id.
x

and

2(x,L):=(x,L)

(2)

2
x

=id.

The composition

:= 2 1
takes

1
2
(x, L) 7
(x0 , L) 7
(x0 , L0 ) =: (x1 , L1 )

(1 , 2 and

are all maps from

to

E ).

More generally,

(xi , Li ) =: (xi+1 , Li+1 )


denes the

ith

iteration of the Poncelet construction. The construction

starting from some

(x, L)

closes if and only if

n (x, L) = (x, L)

(1.3.2)

for some

n.

C and D aren't tangent


 that is, C and D are in general

Thinking of complex points, and assuming


anywhere and don't meet at innity

position in some sense  they meet in exactly four points. This is a


rst taste of Bezout's theorem, which we will prove carefully later. Now
let

x C:

tangent to

x
/ C D, there are exactly two lines containing x and
D; if x C D, there is only one. So we nd that the

if

projection

E
C
(x, L) 7 x

is a two-sheeted branched covering with four branch points:

18

1. TWO THEOREMS ON CONICS IN THE PLANE

(x,L)

(x,L)

E is an elliptic curve, hence isomorphic to


that : E E is a translation

This turns out to mean that

C/lattice.

One deduces

C/ C/
u 7 u + ,
(x, L) is some point u0 . Whether or not n (u0 )
do with whether n , and nothing to do with

and our starting point

u0

has everything to

the choice of

u0 ,

and so (1.3.1) follows.

We will see a more in-depth treatment of this after studying elliptic


curves, including an algebraic criterion for deciding when
a given

C, D,

n =

id. for

and an analysis of elliptical billiards. For now, here are

some examples of Poncelet in action:

Example 1.3.2. (a)

C = {x2 + y 2 = 1}, D = { xa2 +

y2
1a2

n = 4.

C
L0

x1

x0=x4 =(a, 1a2 )

D
L1

x2

(b)

L3

x3

L2

4x
C = {x2 + y 2 = 1}, D = { (1+T
+
)2

4y 2
(1T )2

= 1}, n = 3.

= 1},

EXERCISES

19

1T
2

1+T
2

Exercises
(1)

P2n

n+2
. [For a more challenging
2

is a vector space of dimension

problem, try it with polynomials in

variables instead of two 

obviously with a dierent answer.]


(2) Consider the two conics

r2 }.

C = {x2 + y 2 = 1}

and

The corresponding Poncelet elliptic curve

D = {x2 + y 2 =
E is singular (see

Denition 2.9), which means that the problem degenerates and is


solvable by hand using secondary school maths. For which
between

and

1)

does the

th

n,

(say,

iterate of the Poncelet construction

(starting at an arbitrary point on


(Hint: for each

C)

return to the starting point?

there are nitely many; just nd how many, and

the equation they must satisfy.)

CHAPTER 2

Riemann surfaces and algebraic curves

In this chapter we will dene (complex) algebraic curves (represented by  C ),

1 complex

1-manifolds

(represented by  M ), and Rie-

mann surfaces, and start to consider under what additional hypotheses


they are equivalent concepts.

2.1. Algebraic curves


Definition 2.1.1. Let

gree

m in x, y (the  2

S2m

denote homogeneous polynomials of de-

2 These are polynomials

stands for  2 variables).

of the form

fm (x, y) =

cjk xj y k ,

j,k0
j+k=m

that is, each term has total degree


More generally,
mials in

Skm

m.

is the space of

S2m is a subset of P2m .


degree-m homogeneous polynoClearly

variables  that is, linear combinations

ci1 ,...,ik xi11 xikk

i1 ,...,ik 0
i1 ++ik =m

m. Elements f Skm have the propfm (x1 , x2 , . . . , xk ) = m fm (x1 , x2 , . . . , xk ). (See exercise

of monomials with total degree


erty that
3 below.)

Given a real ane algebraic curve of degree

C := {0 = f (x, y) = fd (x, y) + fd1 (x, y) + + f0 } R2


1in

the ane resp. projective plane. Later we will dene algebraic curves in higher

dimensional projective spaces and products thereof, but the most intrinsic denition of an algebraic curve as a

1-dimensional reduced scheme (some authors require

this to be irreducible and over an algebraically closed eld as well) is probably


something to learn only once you have a rst course in algebraic geometry under
your belt.

2the

eld of denition, from which the

cjk

is a vector space over that eld.


21

are taken, will depend on context;

S2m

22

2. RIEMANN SURFACES AND ALGEBRAIC CURVES

with

fd

not identically zero, we would like to count its intersections

with a real line given parametrically by

L : t 7 (t, t)
(where

are real constants). These are just the solutions of

0 = fd (, )td + fd1 (, )td1 + f0 .

(2.1.1)

Naively, we would like to get

points:

#{C L} = d ??

(2.1.2)
Some issues arise . . .

Problem
(a)

Solution

is not algebraically closed!

(b)

solutions at innity!

(c)

multiple roots!

(d)

might contain

pass to

C2

add a line at innity to

C2

count intersections with multiplicity

L!

Uh-oh

Each Problem is an obstruction to (2.1.2), and the object of each


Solution is to remove the obstruction.
In a little more depth, (a) says that in spite of the fact that the

fi (, ) R, roots of (2.1.1) can be non-real. So we had better consider


C and L as complex algebraic curves  take x, y C in the denition
of C and t C in the denition of L. Their dimensions over R then,
of course, double, and C L now contains the points corresponding to
non-real roots of (2.1.1).
1
d
Next, if we plug t = s
into (2.1.1) and multiply by s , then it
becomes

0 = f0 sd + f1 (, )sd1 + + fd (, ).

(2.1.3)

 to (2.1.1) is a solution at 0 to (2.1.3), which exists if


and only if fd (, ) = 0. For this to be counted in C L, we must add a
2
2 3
line L to C to get P , the complex projective plane (which we shall
discuss in a moment). This adds a point to L at  corresponding to
s = 0, yielding a P1 (projective line), and adds points at innity to C

A solution at

(yielding a compact projective curve).

3called CP2

or

P2C

in some books

2.1. ALGEBRAIC CURVES

We know that to get

23

d solutions out of a degree d polynomial equa-

tion you have to count a twice repeated root as two solutions. So to

d intersection points you will certainly have to count intersections


of C with L however many times the corresponding root of (2.1.1) is
repeated. The curious case is an m-times repeated root at innity: via
(2.1.3), this corresponds to fd (, ) = = fdm+1 (, ) = 0. In that
case, (2.1.1) is only in fact of degree d m. One does have to worry
get

about these degenerate cases, but they will look completely natural in
projective coordinates.
Finally, if all

f` (, ) = 0,

then

C L

and we are in trouble

 there is no way around (d). We shall have to demand that plane


curves intersect properly (in points only), disallowing this possibility,
in order to make any statement about the number of intersection points.

Example 2.1.2. Given a quintic curve

as in the following picture

L
8

(5,5)

the number of intersection points in

(i,i)
(1,1)
R2

is only

2.

But the number of

complex intersection points, counting multiplicities and intersections


at innity, is

So .

innity to

5.

how does one go about adding a line (resp.


(resp.

L)?

First, visualize

as

point) at

24

2. RIEMANN SURFACES AND ALGEBRAIC CURVES

i
1

and think of all arrows as going o to the same point. Adding this point

C {},

resulting in a sphere

gives the 1-point compactication

1
circle

unit

0
This is an informal way of thinking of

P1

Definition 2.1.3. Projective space

through the origin

in the following:

Pn

is the set of complex lines

Cn+1 .4 More precisely,

Pn := D

(Cn+1 \{0})
(z0 ,z1 ,...,zn )(z0 ,z1 ,...,zn )
C

consists of nonzero vectors in

Cn+1 ,

modulo the equivalence relation

equating all vectors lying on a complex line.

Elements are written

[z0 : z1 : : zn ].
For

n=1

this yields the projective line

phism

P1 ,

which has the isomor-

P1
C {}
[z0 : z1 ] 7 zz01 =: z
4as will be proved in Chapter 5, one should really think of Pn
(but we haven't dened these yet).

denotes

(0, 0, . . . , 0).

as a complex manifold

2.1. ALGEBRAIC CURVES

25

given by taking slope (of the line represented by


is well-dened as

[z0 : z1 ] 7

z1
.
z0

[z0 : z1 ]).

This map

The honest topological

is

picture of

z1
z0

i
1

circle

unit

0
while the schematic real picture is

[0:1]

Next, setting

0
n=2

[1:0]
=

we have the projective plane

P2 ,

and the iso-

morphism

P2

if z0 6=0

[z0 : z1 : z2 ]

(2.1.4)

if z0 =0

expresses how

P2

1
(C C)

 P
z1 z2
,
C2
z0 z0

[z1 : z2 ] P1

adds a line at innity (the

P1 )

to

C2 .

For

P2 ,

the

(rather bad, but standard) schematic picture is

[0:0:1]

L , i.e.
z0=0

z1=0

all 3
P1s
[0:1:0]

[1:0:0]

z2=0

While I'm not going to try to represent

real dimensions on paper,

here is a mostly honest topological depiction of the

3 P1 's:

26

2. RIEMANN SURFACES AND ALGEBRAIC CURVES

z0=0
z1=0

z2=0
Remark 2.1.4. Equation (2.1.4) relates ane coordinates (on

C2 )

P2 ). Instead of the {zi }, I will frequently



X Y
2
use [Z : X : Y ] for a point in P and (asuming Z 6= 0)
,
=: (x, y)
Z Z
2
for the corresponding point in C .
2
Also, a warning is in order: [0 : 0 : 0] is not a point in P . With
and projective coordinates (on

homogeneous coordinates, some entry must be nonzero.


Returning to our degree

{f (x, y) = 0} C

described above?

C=
P2 as

(and now complex) algebraic curve

, what happens to it as we compactify

to

To treat this, we rst need to introduce the main

object of study of this course.


Since
equation

[Z : X : Y ] = [Z : X : Y ], in order for a
F (Z, X, Y ) = 0 to make sense projectively (i.e.

polynomial
in

P2 ),

we

must have
(2.1.5)

F (Z, X, Y ) = 0 = F (Z, X, Y ) = 0 ( C ).

This condition is guaranteed by homogeneity of


in Denition 2.1).
homogeneity of

(cf.

(In fact, as we shall see later it is equivalent to

F .)
C P2
F S3d .

Definition 2.1.5. A projective algebraic curve

d,

the property

is the zero set of a homogeneous polynomial

of degree

Here, then, is a general procedure for going between ane and


projective curves:

f (x, y) = 0 7 Z f

(2.1.6)

X Y
,
Z Z

corresponds to taking the projective closure


curve

CC

. Conversely, if the given


=0

C P2

of a given ane

is already projective (dened

2.1. ALGEBRAIC CURVES

by

F = 0),

27

then

F (Z, X, Y ) = 0 7 F (1, x, y) = 0

(2.1.7)
restricts

to the ane curve

C C2 .

Given an ane curve, taking

closure then restricting gets you back to where you started.


Example 2.1.6. Starting from the homogeneous cubic polynomial

F (Z, X, Y ) = ZXY +3Z 2 Y +4Y 3 , the anization is f (x, y) = F (1, x, y) =


xy + 3y + 4y 3 . Conversely, if we start from f (x, y) = x3 y y 2 + 2x, the
X Y
4
3
2 2
3
projectivization is F (Z, X, Y ) = Z f ( , ) = X Y Y Z + 2XZ .
Z Z
Take

F, C

deg Fi ),

Q
F = Fi (so that deg F =
set of Fi , we have C = Ci .

to be as in Denition 2.5. If

then writing

Ci

for the zero

Definition 2.1.7. We say that

no proper (deg

1)

is irreducible if and only if

has

homogeneous factors.

Now let's consider our intersection problem (2.1.2) once more, in the
complex projective setting. Referring to the discussion up to Example

m-fold

intersection at innity, then the degree

of the polynomial in (2.1.1) is

d m. The Fundamental Theorem of


has d m complex roots counted with

2.2, if

and

have an

Algebra then says that (2.1.1)

multiplicity, and we dene these to be the intersection multiplicities for

and

in

C2

as indicated in our discussion. We have proved a baby

version of Bezout's theorem:


Proposition 2.1.8. Let

L P2

be a (projective) line in

P2 ,

i.e. an

algebraic curve of degree one. A projective algebraic curve of degree


in

not containing

L,

meets

in

points counted with multiplicity.

In proving this result we did a tiny bit of complex analysis on

L, so

were implicitly using its structure as a complex 1-manifold. In general it


is quite useful to be able to do analytic computations on curves, but not
all irreducible algebraic curves are complex manifolds (at least, without
doing something to them called normalization). The obstructions are
called singularities and will be explored in greater depth later.

For

now, we will just give a denition and a few examples.


Definition 2.1.9. A singularity or singular point of an ane al-

gebraic (plane) curve

f (x, y) = 0

is a point in

C2

where

f,

f
, and
x

28

2. RIEMANN SURFACES AND ALGEBRAIC CURVES

f
f
f
are all zero  that is, a point on the curve where
and
vany
x
y
ish. A singularity of a projective algebraic curve

F (Z, X, Y ) = 0

is a

F
F F
,
, and
are all zero. A curve with one or more
point where F ,
X Y
Z
singular points is called singular ; a curve with none is called smooth.
Example 2.1.10. Here are some local real (schematic) pictures of

plane curve singularities:

cusp

An example of a cusp is the point


on

x3 = y 2 );

[1 : 0 : 0] on X 3 Y 2 Z = 0 (or (0, 0)

XY = 0 has
at [1 : 0 : 0]:

the curve

or normal crossing)

ordinary
triple point

ordinary
double point

an ODP (ordinary double point,

X=0

Y=0

Now, this is just two

P1 's

(namely,

X=0

and

Y = 0)

touching at one

point. A more (though not completely) topologcially honest picture


of this is:

X=0

Y=0

which makes it apparent that an ODP is actually a bi-conical singularity.

2.2. COMPLEX

1-MANIFOLDS

29

Before we pass to the analytic side of our story, there are 2 more
facts about homogeneous polynomials worth a quick mention.

First,

d
d
the map S3 P2 given by F (Z, X, Y ) 7 F (1, x, y) is an isomorphism,
d
d
so dim(S3 ) = dim(P2 ) in particular. Second is the Euler formula
N
X

(2.1.8)

i=0

Zi

F
=dF
Zi

for

d
F SN
+1

which will be used in later chapters (cf. Chapter 6 for a proof ).

2.2. Complex

1-manifolds

Recall from basic point-set topology that a topological space is a set

together with a collection

{UI }I

of open sets containing

X,

the

empty set, and all unions and all nite intersections of its members.
(Here

is some typically huge index set. A base for the topology of

is a sub-collection of the

unions, and

{UI }I

which generates it under taking

is said to be second countable if it has a countable base.)

is called Hausdor if points can be separated: i.e. given

there exist disjoint open sets

and

containing

and

q,

respectively.

In topology, a homeomorphism is a continuous, 1-to-1, open

5 map.

U 3 p that are homeomorphic


n
to R (or equivalently, an open ball in R )  these are called open
neighborhoods of p. If these always exist, we say X is locally homeon
morphic to R . A second countable, Hausdor topological space that
n
is locally homeomorphic to R , is called a real n-manifold.
In the case n = 2, we are going to layer complex analyticity onto
Given a point

p X,

and

we like open sets

this construction:

Definition 2.2.1. A complex

1-manifold

consists of

(i) a connected Hausdor topological space

{U } of M (this is a nite set of


{UI }, such that U = M ); and

(ii) an open cover


from amongst the

5a

map

is open

M;
open sets taken

is open (resp. continuous) if the image (resp. preimage) of any open set

30

2. RIEMANN SURFACES AND ALGEBRAIC CURVES

(iii) mappings

z : U C

that are homeomorphisms onto their

image, such that the transition functions

:= z z1 : z (U ) z (U )
are biholomorphic (i.e., analytic isomorphisms).

The

{z }, { }

are called local coordinates, and the

ing) functions ; the entire collection


atlas. The functions

are key:

transition (or patchis called an analytic

is an complex analytic manifold

because they are complex analytic. If in (iii) we replace

by

Rn

and

require the transition functions to be smooth (i.e., have continuous partial derivatives of all orders), then
dierentiable) real

n-manifold

would have been a smooth (or

instead.

If we think of the (complex analytic) transition functions in Denition 2.11 as maps from

R2

to

R2 ,

then

( |{z}
x ,
y
) = ( u(x, y) , v(x, y) )
|{z}
| {z }
| {z }
real
part

is smooth and

u, v

imag.
part

real
part

imag.
part

satisfy the Cauchy-Riemann equations. These may

be expressed in terms of the Jacobian matrix

ux uy
vx vy
6if U

and

V .

ux vx
vx ux

!
,
U := U U .
goes from V

are distinct open sets in our open cover, we write

Leter we will write


to

for

z (U )

and

for

z (U ),

so that

1-MANIFOLDS

2.2. COMPLEX

which consequently has positive determinant:

is

Example 2.2.2.

of complex

u2x + vx2

(obviously

biholomorphic. Therefore
7
orientable as a real 2-manifold.

cannot equal zero since


orientation, so

31

is

0)

preserves

C, C , H, P1 , and C/ are (the simplest) examples

1-manifolds.

For the rst three, producing an analytic atlas

is trivial (since you only need one

U ),

and we will do this below for

the latter two.

Now assume
nite subcover.

is compact, that is, every open cover has a -

(In fact, since a complex

1-manifold

always admits

a metric, compactness is equivalent to every sequence of points hav-

an = n has no limit in C but


[0 : 1] in P1 , which is compact.)

ing a convergent subsequence. Clearly

an = [1 : n] =

[ n1

Then viewed over

2-manifold.

: 1] does limit to
R, M is an orientable,

compact, connected, smooth

By a theorem in topology, this means that

morphic to a sphere with

handles, and we say

is homeo-

has genus

g:

etc.
g=0

g=2

g=1

g 0 occur for complex manifolds; we'll show this


g = 0, 1. To do complex analysis on M , you can use

It is a fact that all


in a moment for

the local coordinates, but for some purposes it is also convenient to cut

7in

into a simply connected region, e.g.


fact, a matrix of the form

A
B

B
A


is a rotation times a dilation, hence

preserves angles  that is to say, under the assumption of the CR-equations,

is

conformal. So a complex 1-manifold is essentially a dierentiable real 2-manifold


with conformal transition functions.

32

2. RIEMANN SURFACES AND ALGEBRAIC CURVES

g=1

g=2

Extrapolating from this, one sees that if you begin with a sphere with

handles, the cut-open version is a polygon with

4g

edges identied

in pairs. (The black points on the right-hand side are

all

identied.)

8
Using this we can do a quick computation of the Euler characteristic
of

M:
M := faces edges + vertices = 1 2g + 1 = 2 2g.

(2.2.1)

g = 0)

Example 2.2.3. (

nates

[X : Y ].

Let

Consider the open

[0:1]=" "

M := P1 with homogeneous coordicover {U0 , U1 } of M given by:

U0

[1:0}="0"

U01

M
U1
That is,

U0 = M \{[0 : 1]} and U1 = M \{[1 : 0]}.

8traditionally
lation of

M,

For local coordinates,

one would use the numbers of faces, edges, and vertices in a triangu-

but using a polygonal decomposition like this is also OK

2.3. RIEMANN SURFACES

33

we take

z0 :

U0
C
Y
[X : Y ] 7 X

z1 :

U1
C
.
[X : Y ] 7 X
Y

and

Writing

C C.

U01 := U0 U1 ,

z0 (U01 ) = C C and z1 (U01 ) =


(which goes from z0 (U01 ) to z1 (U01 )

we have

The transition function

by denition) is then

10 : C C
.
u 7 u1

g = 1) Let 1 , 2 C be linearly independent


over R. Then := Z h1 , 2 i = Z1 + Z2 is a lattice, and we set
M := C/. (This means that z, z 0 C give the same point in M if
0
and only if z z .) We endow M with local coordinates on the
Example 2.2.4. (

neighborhoods shown

covering by coordinate
neighborhoods

),

basically by using the coordinate on

(before quotienting by

nd that the transition functions

are all either the idenitity or

translation by some

C
ij

Topologically,

is a torus (cf. the

and

g=1

pictures above), which is evident from performing the identications


on the sides of a fundamental region for

C/

as shown.

2.3. Riemann surfaces


Traditionally, a Riemann surface

is a compact complex

1-manifold

obtained as the existence domain of an algebraic (typically multivalued) function over

P1 .

That isn't the denition I'll use here, but I do

want to explain the concept.

34

2. RIEMANN SURFACES AND ALGEBRAIC CURVES

For example, given distinct complex numbers


tion

i , the algebraic func-

v
u2g+2
uY
F(z) := t
(z i )
i=1

on

can be made single-valued on a complex manifold

(of genus

g,

P1 , cut identical
j = 1, . . . , g + 1. Then

it turns out) constructed as follows: On two copies of


nonintersecting slits from
glue the two copies of
which

2j1

to

2j

for

together on these slits, forming a set on

F becomes single-valued;

nally, endow this set with an analytic

atlas to get a complex manifold M . This manifold has a distinguished

1
morphism M P presenting it as a nite branched cover of the
projective line. We won't do this explicitly here  especially endowing
it with an analytic atlas, since that is really a special case of normalizing
an algebraic curve (cf.

3.1).9

Instead, let's visualize what a couple of existence domains for


algebraic functions look like, starting with the Riemann surface of
1

(w =)z 3

over the unit disk. This is some object tting (as  {z

= w3 })

into the following picture:


2 i
3

{z=w3}

3
4 i

e3
wdisk
3 2
1

zdisk
To construct it, think about following

z3

around the disk once coun-

terclockwise: when you reach your starting point the function has become

2i
3

times the branch of

more, you get

9it

4i
3

1
3;

z3

you started with; going around once

and one more time gets you back to your original

makes a very instructive exercise though, and the next example gives a hint on

how to do it

2.3. RIEMANN SURFACES

branch.

35

So taking three unit disks, slitting them along the positive

reals, and gluing them as indicated

III

II

"z1/3"

"e 3 z1/3"

"e 3 z1/3"

we get the parking lot

{z=w3}

zdisk
(The green segments are glued but I can't draw in 4 dimensions.) An
easier way to visualize this Riemann surface is this: it's just the
disk. The diculty is in seeing the

w-disk

over the

w-

z -disk.

Next, let's construct an existence domain for

p
(z a)(z b)(z c)

F(z) =
over

P1 .

In a neighborhood of

surface of

(z z0 )

1
2

z0 = a, b, c

this looks like the Riemann

over a disk, which is the same as the construction

we just did except with 2 unit disks instead of 3. Indeed, going once
around

a, b,

or

takes

F 7 F;

and furthermore, because the degree

of the polynomial under the square root is odd, going once around

does the same thing. Since

is equivalent to

going around two points at once gives no change. So taking two

P1 's

36

2. RIEMANN SURFACES AND ALGEBRAIC CURVES

and cutting and pasting them as indicated, we end up with a manifold


of genus

on which

becomes well-dened:

"open" the
cuts and
join

"+ f "

" f "
8

8
c

a
(In the picture,

and

are called

1-cycles;

there just there to make

the topology clear.) The same construction works if we replace


by

p
(z a)(z b)(z c)(z d),

with

replacing

F(z)

In fact, by a deep result (on existence of nonconstant meromorphic


functions on complex 1-manifolds) any compact complex

1-manifold

is an existence domain of the sort we have just discussed: they are


equivalent objects in the end. The following is motivated by this, and
the desire to keep things simple:
Definition 2.3.1. A Riemann surface is a compact complex

1-

manifold.

Exercises
(1) Take projective closures of
and

L := {x = 0}

in

P2

C := {y 2 = (x 1)(x 2)(x 3)(x 4)}

(nd associated homogeneous equations),

and determine all intersections and their multiplicities (give the


projective coordinates of the points).

What is the sum of multi-

plicities?

Z03 + Z13 + Z23 = Z0 Z1 Z2 .


degree 3  is a scalar, not a

(2) Find the ane equation associated to


(This equation is homogeneous of
coordinate).

EXERCISES

(3) Let

37

F be a polynomial in 3 variables. Prove that F (z0 , z1 , z2 ) =


= F (z0 , z1 , z2 ) = 0 ( C ) forces F to be homo-

geneous (of some degree). [You will have to assume the following

g (in (z0 , z1 , z2 )), with vanish{f = 0} {g = 0}, then f divides a power of g . This

result: given two polynomials


ing locus

and

is called Study's lemma and will be proved later.]


challenging, show the converse. If you like, do both.

If this is too

CHAPTER 3

The normalization theorem

We state (but do not yet prove) the promised relationship between


algebraic curves and Riemann surfaces, and explain how to work it out
directly for conics. To state the general relationship, however, we need
the notion of meromorphic functions on a Riemann surface, so we will
rst dene and prove a few results about those.

3.1. Meromorphic functions on a Riemann surface

M
{(U , z )}
Let

be a Riemann surface (Denition 2.15) with analytic atlas


(Denition 2.11), and write

analytic chart

: V U ( M )

V := z (U ) C.

is simply dened to be the (com-

position) inverse of of the local coordinate

z 1

The local

z .

(I just don't like writing

since in some settings this is easy to confuse with


Definition 3.1.1. A meromorphic (resp.

1
.)
z

holomorphic) function

f K(M ) (resp. O(M )) is a collection of continuous maps f : U


P1 such that
the {f } agree on overlaps (viz., f = f on U ), and
f is a meromorphic (resp. holomorphic) function, in the
sense of complex analysis, for all .
1

U
f

39

40

3. THE NORMALIZATION THEOREM

Remark 3.1.2. (a) One really works with functions of the coordi-

z ,

nate

i.e. the function

f =: g

(mapping

V P1 ),

and then

the compatibility condition reads

g = g .

(3.1.1)
(b)

K(M )

is a eld, since you can multiply, add, and invert (addi-

tively and multiplicatively) meromorphic functions.


Proposition 3.1.3. [Liouville's Theorem]

O(M )
=C

compact

(constant functions).

f O(M ) = f (M ) (P1 \{}) =


C; while on the other, M compact and f continuous = f (M ) is
compact. Applying absolute value gives a compact subset |f (M )|
R0 . This has a maximum element, whch is assumed at some point
p M , and this p lies in some U . Hence, the absolute value of
the holomorphic function g = f attains a maximum on V (at
(p)), and by the maximum modulus principle, g (and thus f ) is
some constant c C.
Let U be any open set of the atlas meeting U . Since f = f = c
on U , and U has accumulation points, f = c on U . One continues
this argument now for any open set meeting U or U , and so forth.
By connectedness of M , this shows f = c on all open sets of the atlas,
hence on all of M .

Proof. On the one hand,

Definition 3.1.4. Let

any

p M, f

f K(M )

be a meromorphic function. For

is locally of the form

z m h(z)

(3.1.2)

m Z, z

p (i.e. z(p) = 0), and


h(z) a local holomorphic function of z with h(0) 6= 0.1 We say that the
order p (f ) of f at p is m.

with

a local coordinate vanishing at

With this bit of language it is easy to compute the meromorphic


function eld for Riemann surfaces of genus
Theorem 3.1.5. (a)

1To
then

be absolutely precise, if

K(P1 )
= C(z) (z
z

and

V.

1.

an indeterminate).

is a local coordinate on

is a holomorphic function on

understood.

U 3 p,

with

V = z(U ),

I'll frequently assume things like this to be

3.1. MEROMORPHIC FUNCTIONS ON A RIEMANN SURFACE

41

:= {m1 1 + m2 2 | mi Z} (1 , 2 C linearly
independent over R) for a lattice, K(C/)
= C(, 0 ) where (u) is
the Weierstrass -function for .
(b) Writing

Proof. (a) Referring to Example 2.13, write

for the two local coordinates.

I am really going

z = z0 and w = z1
to use z as a global

P1 ; the statement we want to prove is that meromorphic


1
functions on P are precisely the rational functions of z .
In one direction, this is easy: if P , Q are polynomials in z (with
P (z)
Q 6= 0), clearly Q(z)
is the restriction to U0 of a meromorphic function

coordinate on

on

P1

(on

U1 ,

it is

1
P(w
)
1 ).
Q( w
)

Conversely, are all meromorphic functions rational?

K(P1 ), p (f ) < 0

Given

2
at nitely many points zi (=

p), and we shall for simP


plicity assume none of these is the point . Let Pi (z) =
zi (f )k<0 ik (z
k
zi ) (sum is over k ) be the principal part of the Laurent expansion of
P
f at zi , and consider G(z) = Pi (z). Then f G O(P1 ) is constant
by Liouville; and since G is rational, we're done.
(b) Next, f K(C/) if and only if f is a doubly-periodic meromorphic function on C: that is, f (u) = f (u + m1 1 + m2 2 ) for all
m1 , m2 Z (also known as an elliptic function). We will see later that
these are generated (rationally) by

X
1
(u) := 2 +
u

1
1

(u )2 2

6=0

and its derivative.


Definition 3.1.6. A morphism (or holomorphic map)

of Riemann surfaces

3 is a collection

{U })

maps (agreeing on the

|z {F 1 (Ui )U }

F : U M

of continuous

such that the composition

is holomorphic for all

tion works more generally for complex

, i.

M M

zi F

(Note that this deni-

1-manifolds

 compactness is

inessential.)

2otherwise

1
1
f have an accumutaion point =
f
identically 0. (Also, note that I am identifying points by the value of the coordinate
z on P1 . If M were not P1 , I would write pi instead of zi .)
compactness

3write {U , z } and {U
i , zi }

4this composition renders z

zeroes of

for the atlases.

i as a function of

this sense)

(and is a local snapshot of

in

42

3. THE NORMALIZATION THEOREM

M
F
Ui

zi

F(U)

Vi

F (U ) U
i

U
z

the composition

(this can ramify!)

p (U )M and q (Ui )M with F (p) = q ,


zi (q) = 0, as shown in the above gure. Assuming

Now suppose we have

z (p) = 0 and
F is nonconstant, then after normalizing the local coordinates,5 we
have z
i (z ) = (z ) for some (unique) Z>0 . One says that f has
ramication index at p (over q ). If this index is > 1, we say that f
is branched over q (or ramies at p).

= 3, we have already seen this picture in 2.3.


In general, for a holomorphic map of Riemann surfaces : X Y ,
1
for all but nitely many y Y the number |
(y)| is the same, and
this is called the degree of the mapping . (This will be explained in
greater depth in a later chapter.) The branch points of are just the
remaining points of Y . Usually we will just draw a schematic picture
Remark 3.1.7. For

like

Y
and it is understood that the picture is really as in
around the point on the base
cover

moves you between branches of the

X.

Proposition 3.1.8. Let

erally, a complex

1-manifold).

be a Riemann surface (or, more genThe holomorphic maps

cluding the constant map sending all points to


meromorphic functions

5see

2.3  so that going

Exercise 4 below

K(M ).

{},

M P1 ,

ex-

are simply the

3.2. RIEMANN SURFACES PARAMETRIZE ALGEBRAIC CURVES

Proof. Again refer to Example 2.13: given a morphism

P1

43

F :M

z0 F is holomorphic on the
, while z1 F = z0 F1 is
6
of the preimage of 0. Hence, F

(Denition 3.6), by denition

complement of the preimage of


holomorphic on the complement
is meromorphic and

{F }

denes a meromorphic function (Denition

3.1). The converse is even more tautological!

Later we will discuss morphisms (holomorphic maps) of complex


manifolds of any dimension. The following is a special case:

[Z0 : Z1 : : Zn ] for (projective) coordin


nates on P . A map from a Riemann surface M to P is called holomorphic if and only if all compositions [Zi : Zj ] are holomorphic
1
as maps to P on the open subsets of M where they are well-dened.
Definition 3.1.9. Write

(M )

Remark 3.1.10. If the image

does not live in a coordinate

hyperplane, this is the same as saying that composing

coordinate on

and

Zj

with each ane

gives a meromorphic function.

To see this, rst write

Zi

Mij

for the subsets of

where (under

are not both zero; these are the open subsets in the last

7
denition. By Prop. 3.1.8, the conditions of Defn. 3.1.9 mean that
Zj
are meromorphic functions on the Mij . We need to show that the
Zi
Z
zj = Z0j extend to meromorphic functions on all of M . First, M is

p
/ M0j (i.e. Zj and Z0
vanish at p), we have a neighborhood U containing p where some other
Zi does not vanish, so that U Mij , Mi0 . Now, on U M0j we can
 1
Zj
ZZ0i
write zj =
as a product of functions which are meromorphic
Zi
on all of U , hence showing that zj extends as desired.
covered by the open sets

{Zi 6= 0}.

Hence, for

3.2. Riemann surfaces parametrize algebraic curves


Here is the Normalization Theorem. We will prove part (A) in this
course.

C P2 ,
: M P2

Theorem 3.2.1. (A) Given an irreducible algebraic curve

there exists a Riemann surface


with

6The

holomorphicity of

and a holomorphic map

as its image which is 1-to-1 on

1 (C\sing(C)).

1
z0 F guarantees, in particular, that
poles and not essential singularities.

7Some, but not all, of these M

ij may be empty if

has only

(M ) is contained in a coordinate

hyperplane. We have excluded this degenerate possibility for simplicity.

44

3. THE NORMALIZATION THEOREM

(B) Given a Riemann surface

M,

there exists a holomorphic map

: M P2 such that
(M ) is an irreducible algebraic curve with sing((M )) consisting
of ordinary double points (or empty), and

is 1-to-1 o the preimage of these ordinary double points.

In this sense, irreducible smooth projective algebraic plane curves


are equivalent to, and are isomorphically parametrized by, Riemann
surfaces.

" "

If a curve

is not smooth, then the normalization desingularizes it

(and we shall see this quite explicitly later on). In either case, we say
that

is the normalization of

C.

Let's look briey at the meaning of (B), which we will not prove
in this course.

For a given Riemann surface (i.e.

compact complex

M , it guarantees a holomorphic map to P2 , with image


(M ) = C =projective closure of {f (x, y) = 0}. Changing coordinates
2
on P if necessary, we may assume that C does not pass through [0 :
0 : 1]. So it makes sense to consider the composition
1-manifold)

2 {[0:0:1]}

[Z:X:Y]

projection

which exhibits

[Z:X]

as a branched cover of

P1

existence domain of the algebraic function

 or more precisely, as the

g(x)

obtained by solving

f (x, g(x)) = 0.
So Theorem 3.10(B) contains the statement that every complex 1manifold is an existence domain in the sense of

2.3.

3.3. STEREOGRAPHIC PROJECTION

45

We should also note that any Riemann surface admits a holomorphic embedding

: M , P3 ,

an even nicer result than part (B) above!

3.3. Stereographic projection


As a plausibility check on Theorem 3.10(A), we'd like a recipe for
normalizing conics  i.e.

degree-2 (conic) curves

C P2 .

Given a

p C , and any line line ` through p, by Proposition 2.8 ` either


meets C in two points with mutliplicity 1 or in 1 point with multiplicity

point

2. Put dierently, we have either

` C = {p, q}
or

` C = 2p,

i.e.

` = Tp C

is the tangent line to

at

p.

(We will give a systematic treatment of tangent lines below.)


versely, given

and any other point

through them (and it doesn't meet

on

C,

Con-

there is a unique line

anywhere else).

There are two ways to think of why this gives a parametrization of

C . One possibility is to
p to project C onto it:

take a xed line (


=

P1 )

and use lines through

C
p

tan
gen
t

line

This is where the term stereographic projection comes from.


But this auxiliary projective line is superuous, because the family

p already gives a P1 . (Indeed this is close to the original


1
1
denition of what P is.) We can parametrize this P by the slope of
Z Z
the line with respect to suitable coordinates (usually (x, y) = ( 1 , 2 )).
Z0 Z0
of lines through

The upshot is that we get a 1-1 correspondence between lines through

p and points on C , so that we are in the situation of 3.2 with M


= P1 .
Example 3.3.1. Suppose we wish to nd a parametrization

P1 C

{X 2 + Y 2 = Z 2 } P2 , which in ane coordinates is


x2 + y 2 = 1. We choose a point on C , say p = (1, 0), and draw lines
y = (x1) through p. (The slope here is , and this should be viewed
1
2
2
as a choice of coordinate on P .) Substituting into x + y = 1 and
of the conic

46

3. THE NORMALIZATION THEOREM

solving for

in terms of

we have

x2 + 2 (x 1)2 = 1
=

(x 1){(1 + 2 )x + (1 2 )} = 0.

=
Ignoring the solution

x=

(2 + 1)x2 22 x + (2 1) = 0

x=1

2 1
,
2 + 1

p), we have

2 1
2

1
=
.
2 + 1
2 + 1

(which corresponds to


y=

Hence, we nd


() =

2 1 2
,
2 + 1 2 + 1


.

One can also do stereographic projection to construct normalizations of singular cubic curves:

p
P

The idea here is to consider lines through the singular point


any such

` already meets C

point (by Proposition 2.8).

twice, it will only hit

p;

since

in one additional

You'll work an example in the exercises

below. This will not work for a smooth cubic.

Exercises

m 7 (x(m), y(m)) (hence an isomorphism


P C ) of the smooth conic curve C that is the projective closure
2
2
of 3x y = 1. (You may work in ane coordinates.)
Show that for any RS M and meromorphic function (0 6=) f
P
K(M ), one has pM p (f ) = 0. [Hint: Use the residue theorem

(1) Give a parametrization

(2)

from complex analysis.

Cut open the RS as in Chapter 2, and

df
along the boundary.]
integrate
f
(3) Convince yourself that the order
on a RS

p (f )

of a meromorphic function

(Denition 3.1.4) is independent of the choice of local

coordinate.
(4) Prove the following, which was claimed in Denition 3.1.6: Given

M, M 0

Riemann surfaces with a holomorphic map

f : M M0

EXERCISES

47

f (p) = q ). Then there exist (U, z) on M and (V, w) on M 0

satisfying z(p) = 0 = w(q), such that w = z (for some N) is


the local form taken by f near p. [Here for example  (U, z) means
an open disk U M with local coordinate z : U C.]
1
2
2
Find a parametrization P C of the singular cubic Y Z X Z +
X 3 = 0 in P2 . (C has an ordinary double point p at [Z : X : Y ] =
[1 : 0 : 0]. Check that this point is indeed a singularity of C .) To
do this, convert to ane coordinates, substitute in y = mx, and

(and

(5)

solve for the other intersection point's coordinates as a function of

m.

Two points will go to

p = (0, 0).

Picture:

p
p

What are

and

Change coordinates on

(fractional linear

0 and are sent


: P1 C sending

transformation) so that in your new coordinate,

p. Your parametrization should read now


z 7 (x(z), y(z)) with 0, 7 p. This will
to

exercise.

be used in a later

CHAPTER 4

Lines, conics, and duality

To complete our introduction to algebro-geometric concepts on the


level of curves, in this chapter we'll study projective transformations,
tangent lines, and dual curves. Our convention will be to write

Z =

Z0

Z1 for column vectors in C3 (written with respect to the standard


Z2
2
basis e), and [Z] = [Z0 : Z1 : Z2 ] for the corresponding point in P .

4.1. The classication of complex conics


The story begins in even lower degree, with lines  i.e. degree
algebraic curves. These are subsets of

of the form

L = { t z = 0},

(4.1.1)
where

is a nonzero vector in

C3 .

Note that for

By stereographic projection (cf.


isomorphically parametrized by
Theorem.

3.3),

C , L = L .

lines and smooth conics are

in the sense of the Normalization

(For lines, the projection is done through a point not on

the line; for conics, one chooses any point on the conic.)

However,

not all conics are smooth, and so we will need to classify conics up to

1 The two key non-smooth examples to keep in

projective equivalence.

mind are the pair of lines

{XY = 0} = {X = 0} {Y = 0}
and the double line

{X 2 = 0}.
1There

is a somewhat subtle point here. For smooth curves in general, projective

equivalence is ner (equates fewer curves) than isomorphism as Riemann surfaces.


However, you have to consider curves of degree at least

5 to see this discrepancy.

As

far as conics are concerned, we like projective equivalence simply because it gives
a uniform and algebraic treatment of singular and smooth curves.
49

50

4. LINES, CONICS, AND DUALITY

The rst has two irreducible components (and is hence reducible ), while
the second has one component of multiplicity two (and is said to be

non-reduced ).

To dene projective equivalence, we introduce the projective general


linear group

GL(n, C)
.
h id. | C i

(We have A B B = A for some C .) Consider the action


2
of P GL(3, C) on P by

a00 a01 a02

T a10 a11 a12 [Z0 : Z1 : Z2 ] =


a20 a21 a22
"
#
a00 Z0 + a01 Z1
a10 Z0 + a11 Z1
a20 Z0 + a21 Z1
:
:
+a02 Z2
+a12 Z2
+a22 Z2
P GL(n, C) :=

or in more compact notation

T (A)[Z] = [A Z].
(We are, consistently with the notation mentioned at the beginning of
the chapter, letting the matrix

A act on Z

viewed as a column vector. )

This action is well-dened:

Z to 0 (recall [0] is not a point in P2 );


Z = Y , then T (A)[Z] = [AZ] = [A Y ] = [AY ] = [AY ];
A = B , then T (A)[Z] = [BZ] = [BZ] = T (B)[Z].

it sends no nonzero
if
if

Definition 4.1.1. The transformations

T (A) : P2 P2 , A

P GL(3, C), are the projective linear transformations (or projectivities )


2
of P .

2roughly

speaking, reduced means all of its irreducible components are of mul-

tiplcity one. So while

{XY = 0}

is reduced, something like

{XY 3 = 0}

is not.

Obviously this is going a bit beyond the notion of an algebraic curve as a solution
set, since it incorporates multiplicity. To really formalize what such an object is,
we would have to work with scheme theory or algebraic cycles, which I do not want
to do. So unless otherwise stated, in this course an algebraic curve is assumed to
be reduced.

3the

dot indicates matrix multiplication. This will often be omitted.

4.1. THE CLASSIFICATION OF COMPLEX CONICS

Remark 4.1.2. The analogue of projectivities on

P1

51

are simply the

fractional linear transformations:

a b
c d

!!
[Z0 : Z1 ] = [aZ0 + bZ1 : cZ0 + dZ1 ].

So writing  z  for the point

[1 : z],
!!
a b
c + dz
(z) =
.
a + bz
c d

You probably know from complex analysis that such transformations


preserve the cross-ratio of 4 points. Furthermore, they are the only automorphisms of

P1

(invertible morphisms from

P1 P1 )

as a complex

1-manifold.
C = {F (Z) =
0} of degree d, points in T (A)C are of the form T (A)Z for Z C . These
How do projectivities aect algebraic curves? For a curve

are precisely the solutions of the equation


{F T (A1 )() = 0} (= T (A)C)

(4.1.2)
since then

F (T (A1 )T (A)Z) = F (T (A1 A)Z) = F (Z) = 0.


4

Since (4.1.2) just substitutes linear forms for


for

C,

Z0 , . . . , Zn in the equation

we nd:

Proposition 4.1.3. The images of (smooth resp.

braic curves of degree

singular) alge-

under projectivities, are again (smooth resp.

singular) algebraic curves of degree

d.

So lines are carried to lines, conics to conics, and so on.


eral, if

T (A)C = C

for some

A,

then the curves

C, C

In gen-

are said to be

projectively equivalent.
Proof. To see why smoothness is preserved, write

[] = T (A)[Z]); and suppose (for a contradicF


F
Z C we have Z
(Z) 6= 0 but
(T (A)Z) = 0
0
i

(where we have in mind


tion) that for some
(i).

4i.e.

F () = F (T (A1 ))

homogeneous polynomials of degree

in

Z0 , . . . , Z n

52

4. LINES, CONICS, AND DUALITY

If

F = F T (A1 ),

chain rule

then

F = F T (A),

and by the (multivariable)

X T (A)i F X F
F
=
=
,
ai0
Z0
Z0 i
i
i
i

so that

0 6=

X F
F
(Z) =
(T (A)Z) = 0,
ai0
Z0
i
i


a contradiction.

Next we want to get formulas for the eect of projectivities on lines


and conics.

Given

( A1 )Z ,

so that

t t

L = { t Z = 0},

0 = t A1 Z =

T (A)L = L( t A1 ) .

(4.1.3)
Since

(4.1.2) gives

GL(3, C) acts transitively on C3 , this implies the (relatively triv-

ial)

Proposition 4.1.4. All lines in

P2

are projectively equivalent.

Let

Q = {0 = aZ02 + bZ12 + cZ22 + dZ0 Z1 + eZ0 Z2 + f Z1 Z2 }


be an arbitrary conic. We can rewrite the equation

0=

Z0 Z1 Z2

d
2

d
2
e
2

in terms of a (unique) symmetric

f
2

Z0

t
Z1 =: ZBZ
c
Z2
e
2
f
2

5 matrix

B.

(The expression

called a symetric bilinear form.) Given such a


in

A Z

for

Z,

QB ,

ZBZ

is

(4.1.2) substitutes

yielding

0 = t (A1 Z)BA1 Z = t Z( t A1 BA1 )Z


so that

T (A)QB = Q( t A1 BA1 ) .
M , the transformation B 7 t M BM =:
B, B 0 are cogredient over C. All nonzero

Given an invertible complex matrix

B0

is called a cogredience, and

5i.e.

the transpose

equals

4.2. TANGENT LINES

/C to

1 0 0

0 1 0 ,
0 0 0

symmetric matrices are cogredient

1 0 0

0 0 0 ,
0 0 0

53

one of the form

1 0 0

0 1 0 .
0 0 1

or

We conclude:

P2

Proposition 4.1.5. All conics in

are projectively equivalent to

one of

{X 2 = 0}, {X 2 + Y 2 = 0},

or

{X 2 + Y 2 + Z 2 = 0}.

X 2 + Y 2 = (X + 1Y )(X 1Y ) is a pair of lines,


projectively equivalent to XY = 0.

Notice that
and so

Corollary 4.1.6. (i) All smooth

7 conics are projectively equiva-

lent.
(ii)

QB

is smooth

Proof. (i) Since

det B 6= 0.
{X 2 + Y 2 + Z 2 = 0}

is the only smooth option

in Prop. 4.1.5, by Prop. 4.1.3 all smooth conics must be equivalent to


this hence to each other.

B 7 t M BM multiplies determinant by (det M )2 ,


which is always nonzero (as M GL(3, C)); so projectivities preserve
non-zero-ness of det B .

(ii) Cogredience

4.2. Tangent lines


Let

C = {F (Z0 , Z1 , Z2 ) = 0}

be a projective algebraic curve, and

suppose

R (, ) C
t 7 f (t) = [Z0 (t) : Z1 (t) : Z2 (t)]
is a dierentiable path segment in

C.

Then,

2
X
dZi
F
0 = (F f ) (0) =
(f (0))
(0)
Zi
dt
i=0
0

6This

is Sylvester's theorem. For an easy proof of the real version, see pp. 161-162

of my linear algebra book.

(To get rid of the  1 entries, hence arrive at the

simpler complex version, just multiply the relevant basis vectors by

7or

equivalently, irreducible

1.)

54

4. LINES, CONICS, AND DUALITY

0
(0)
Z
0



0
F
F
F
= Z
(f
(0))
(f
(0))
(f
(0))
Z

1 (0)
Z1
Z2
0
Z20 (0)

= t F (f (0)) f 0 (0),
LF (f (0)) contains
C through f (0).

and so the line


paths in

all tangent vectors

f 0 (0)

to all such

F (f (0)) = 0,
C to be smooth

There is one catch: if the gradient vector


does not dene a line at all. So we must ask

then it
at

f (0)

for this computation to work.

Tp C to a curve C = {F =
p = [p0 : p1 : p2 ] C is LF (p) .

Definition 4.2.1. The tangent line

0} P

at a smooth point

f(, )

p=f(0)

f (f(0))

TpC =L

C
The next proposition makes the intuitively obvious statement that projectivities respect tangent lines:

L
T (A)C

Proposition 4.2.2. If

is the tangent line to

is the tangent line to


at

at

p,

then

T (A)L

T (A)p.

Proof. We must show

T (A)LF (p) = L(F T (A1 ))(T (A)p) .


Writing

F = F T (A)1 ,

this is equivalent to

T (A)L(F T (A))(p) = L F (T (A)p)


hence to

L t A1 (F T (A))(p) = L F (T (A)p)
or

(F T (A))(p) t A F (T (A)p)

(4.2.1)
where

means up to multiplication by

C .

As you may wish to check

by writing everything out, equality of both sides of (4.2.1) is just an


expression of the chain rule.

4.3. THE DUAL PROJECTIVE PLANE

55

Now, the tangent line to a line (at any point) is the line itself; for
conics the story is less trivial. First we write

F (Z) = t ZBZ,
B

symmetric, and compute the gradient: writing

dard

e0 , e1 , e2

for the stan-

0
0
1

basis vectors 0 , 1 , 0 ,
1
0
0

t
F/Z0
ZBe0 + t e0 BZ

F (Z) = F/Z1 = t ZBe1 + t e1 BZ


t
F/Z2
ZBe2 + t e2 BZ

t
e0 BZ

= 2 t e1 BZ = 2BZ.
t
e2 BZ
t

ZBei = t ( t ei t BZ) = t ( t ei BZ) = t ei BZ


t
symmetric and ei BZ is 11, i.e. a scalar.)
(Here

Proposition 4.2.3. The tangent line to

uses the fact that

QB

at

[p] QB

is

is

LBp .8

4.3. The dual projective plane

V /C, a basis e = {ei } of V and an


T : V V with matrix [T ]e =: M .

Suppose we have a vector space


invertible linear transformation
Recall that the dual of

is the vector space

V := Hom(V, C)
of linear functionals (f

one has the tautological pairing

h,i
V V C

(4.3.1)
given by

: V C);

hf, vi = f (v).

We have a dual basis


e = {ei } ( ei , ej = ij ),

with respect to which one has

hf, vi =

([f ]e )[v]e
{z }

matrix multiplication
Finally, there is a dual transformation

T : V V

dened by


T f, T v = hf, vi
8here we are treating p as a column vector (which is consistent with earlier notation)

56

4. LINES, CONICS, AND DUALITY

with matrix

[T ]e = t M 1 .
This gives a more conceptual way to look at the story of lines in
above: put

C
= C

V = C3 , [T ]e = A, f V , = [f ]e

P2

and so on. Of course

as vector spaces, but we want to keep them conceptually

separate.
The crucial point is to projectivize

and

V :

writing

3 \{0}
C3 \{0}
C
2

P =
, P =
,
C
C
2.
P2 correspond to points [] P
2

we see that lines in

In fact, since the

notion of duality is dened by (4.3.1), it is symmetric:


points in

correspond to lines in

2.
P

V = V ,

and so

This entire correspondence is

invariant under projectivities provided one operates simultaneously on

P2

with

T (A)

and

2
P

T ( t A1 ).

with

A bit more formally, then:

2 is the space of lines


P

Definition 4.3.1. The dual projective plane

in

P2 .
Now write

p0

p = p1 ,
p2

= 1
2

for column vectors. Though this is maybe a little awkward, here is how
I want to standardize notation:

p = [p] = [p0 : p1 : p2 ] P2
2,
= [ t ] = [0 : 1 : 2 ] P
in other words, points in
points in

2
P

P2

are thought of as column vectors and

as row vectors. As above the line

L P2

is dened by

the equation

Z =0

and we say that its dual

solve
= .
L

for

Z0

Z = Z1 ,
Z2

Moreover,

denes a line

2
Lp P

via

W p=0

solve for

W =

W0 W1 W2



4.3. THE DUAL PROJECTIVE PLANE

and we write

57

p = Lp .

What about the dual of a conguration of


(a) a point

on a line

L()

(important in Poncelet)?

L, L through a point p?
points p, q on a line L?

(b) a pair of lines


(c) a pair of

For the rst one, the equation

t
expresses  p

Lp ).

L ;

p=0

but from the above it also expresses

p
L

(i.e.

Repeating this reasoning, we have

(
a) a line

p through

a point

;
L

a pair of points
(b)
(
c) a pair of lines

L on a line p;
L,
.
p, q through a point L

Here is something more interesting to dualize, which is left as an exercise for you.

L, L
L) take

Theorem 4.3.2. [Pappus of Alexandria, c. 300 AD] Let

P2

s = L L. On L (resp.
(3)
(resp. q , q , q ) dierent from s, and set
:= p(i) q (j) p(j) q (i) (where {i, j, k} = {1, 2, 3}).

be two distinct lines, and write


(1)

(2)

(3)

p ,p ,p
(k)
(for k = 1, 2, 3) r
(1)
(2)
(3)
Then r , r , r
are
distinct

(1)

(2)

collinear.

p(3)
p(2)

r(1)

(2)

p(1)

q(3)

(3)

(2)

(1)

58

4. LINES, CONICS, AND DUALITY

p(1) q (2) p(3) q (1) p(2) q (3) is a hexagon inscribed in the


(i)
conic LL, and the {r } are the intercepts of its opposite edges. After
Proof. In fact,

changing by a projectivity (which preserves the gure), this conic is

XY = 0, which is obviously the limit of the smooth conic XY Z 2 = 0


as 0. Since Pascal's theorem implies collinearity of the hexagon
edge intercepts for all 6= 0, this remains true at = 0.

4.4. Dual conics and polar lines

2
C P

Definition 4.4.1. The dual

C = {F = 0} P is the set of
2 | p C}.
= {TpC P
That is, C

of a smooth algebraic curve

(dual points of ) tangent lines to

C.

This is consistent with our denition for lines. For higher degree
curves, however, the dual is not one point: consider the duality map

2
DC : P2 P
sending

p 7 [ t F (p)].
C = DC (C).
= TpC , then T C = p.

Proposition 4.4.2. (a)

(b) If

is smooth at

Proof. (a) For

p C , Tp C = LF (p) = TpC = [ t F (p)].

So

this is practically a tautology.


(b) Here we jump into a little deep water. It suces to show that
for any path

() : (, ) C

through

TpC = [ t F (p)] = DC (p),

d t
(0) p = 0.
dt

(4.4.1)

C is the image of C
(, ) C through p. So

Since

by

DC , t (t) = (DC q)(t)

for some

q() :

the left-hand side of (4.4.1) becomes

d
(DC q)(0) p =
dt
(4.4.2)

q00 (0) q10 (0) q20 (0)

9i.e., [Z : Z : Z ] 7
0
1
2

2F
Z02
2F
Z1 Z0
2F
Z2 Z0

F
Z0 (Z0 , Z1 , Z2 )

2F
Z0 Z1
2F
Z12
2F
Z2 Z1

2F
Z0 Z2
2F
Z1 Z2
2F
Z22

F
Z1 (Z0 , Z1 , Z2 )



p0

p1
.


p2

p

F
Z2 (Z0 , Z1 , Z2 )

4.4. DUAL CONICS AND POLAR LINES

The matrix in the middle is the Hessian of


determinant if

DC (p).

is nonsingular at

59

p and has nonvanishing

at

We will return to the Hessian

later in this course.


Now, since each

F
is homogeneous (of degree
Zi

the Euler formula ((2.1.8), with

(d 1).

p)

set equal to

q00 (0) q10 (0) q20 (0)

(d 1).F (q(0))
which is indeed zero by the beginning of

d1, if d = deg(C)),
collapses (4.4.2) to

F
(p)
Z0
F
(p)
Z1
F
(p)
Z2

dq
(0),
dt
4.2.

Remark 4.4.3. One can show that the dual of a smooth algebraic

curve of degree

d2

d is an algebraic curve of degree d(d 1);

moreover, for

this dual is singular, so the duality map cannot be reversed as

dened.

QB . By the computation
F (Z) = ZBZ ), F (Z) = 2 ZB t ZB and so


DQB (QB ) = [ t ZB | [Z] C]

We consider the dual of the conic


(for

in

4.2

= {[ t ZB] | t ZBZ = 0}.


Making the substitition

= t ZB Z = t B 1 ,

this becomes

2 | t W B 1 B t B 1 W = 0}
B = {[ t ] P
Q
2 | t B 1 = 0}
= { P
where we have used the fact that

is symmetric. This gives part (i)

of:
Proposition 4.4.4. (i)

B = QB1 ,
Q

the dual of a smooth conic is a smooth

= Q . (In particular,
Q
B
B
conic, since det B 6= 0 =

and

det B 1 6= 0.)
(ii) Given
tangent to

p P2 \QB ,

there exist exactly two lines through

and

QB .

Proof. (ii) By Proposition 4.4.2(b), this is dual to the statement:

if the line

2
p P

is not tangent to

B,
Q

then it meets

B
Q

in exactly

points. This last statement then follows from Proposition 2.1.8.

2


60

4. LINES, CONICS, AND DUALITY

QB be a smooth conic and p be a point not


on QB , with Tq QB and Tr QB the two tangent lines to QB containing p.
2
(Here q, r QB .) Then the polar line L(p,QB ) P of p with respect to
QB is the line through q and r.
Definition 4.4.5. Let

p
Tq QB

Tr QB

r
L

p P2 \QB , with polar line L = L(r,QB ) (


2 (of the dual point L
with respect
P2 ). Then the polar line L(L,
Q
B) P
to the dual conic) is p
(the dual line of p). In a picture, where dual
Proposition 4.4.6. Let

objects are the same color:


p
Tq QB

L
Tr QB

q
q

Tq QB

r
Tr QB

(a,b,c)
the denition of the dual curve, and Proposition 4.4.2(b). 

2
2
2
Example 4.4.7. Q = {4Z0 +Z1 +Z2 = 0} B =
1
.
1
Let p = [1 : 2 : 2] (
/ Q) so that the polar line L is given by

Z0

 4

0= 1 2 2
1
Z1 = 4Z + 2X + 2Y,
1
Z2
Proof. This is an immediate consequence of the rules

, c),
(a, b

EXERCISES

i.e.

L = L

where

61

= [4 : 2 : 2].

On the dual

2
P

side:

= ; Q

has matrix

41

B 1 =

1
p is the line 0 = W0 + 2W1 +
hence equation 0 =
is
2W2 . On the other hand, the polar line of with respect to Q

1
W0

 4

0 = 4 2 2
1
W1 = W0 + 2W1 + 2W2 ,
W2
1
14 Z02

agreeing with

+ Z12

+ Z22 ; and

p.

It is instructive to think about what happens to Pascal and Poncelet


under duality. While the dual of Poncelet is again just Poncelet (but in

2 ), we do nd that if polygons inscribed in C and circumscribed about


P
2 ) inscribed in
D close up after n sides, then so do the polygons (in P
and circumscribed about C .
D
The dual of Pascal, on the other hand, does give a dierent statement:

Proposition 4.4.8. The 3 lines through opposite vertices of a hexagon

circumscribed about a conic, pass through a single point.

Proof is basically the same as for Proposition 4.4.6; I'll let you work
it out.

Exercises
(1) Given a conguration of

4 points p, q, r, s P2 in general position,

i.e. no three of them collinear, show there exists a unique projec-

p 7 [1 : 0 : 0], q 7 [0 : 1 : 0], r 7 [0 : 0 : 1], s 7 [1 :


1 : 1]. [Hint: work with vectors p, q, r, s C3 . You only have to
send p (resp. q, r, s) to a multiple of e0 (resp. e1 , e2 , e0 + e1 + e2 ).]
tivity sending

(2) (a) Give a direct proof of Pappus's theorem.

[Hint: use the last

exercise to rst simplify the coordinates of several of the points.]


(b) State a dual version of Pappus's theorem, and draw a gure.
[Note: it would be better to state the dual version in
of Pappus in

2
P

and then dualize that.]

P2 :

think rst

62

4. LINES, CONICS, AND DUALITY

(3) Show that the equation of the polar line of


has equation

pBZ = 0.

with respect to

QB

Use this to give another (short) proof of

Proposition 4.4.6.
(4) Prove that all automorphisms of

P1

(as a complex manifold) are

fractional linear transformations. Deduce that


[Use material from

3.1.]

Aut(P1 )
= P GL2 (C).

Part 2

General denitions and results

CHAPTER 5

Complex manifolds and algebraic varieties

In Chapters 2-3 we introduced Riemann surfaces and plane algebraic curves, and stated the Normalization Theorem which produces
a strong relation between them. Here we will introduce the arbitrarydimensional generalizations of these objects. While it is true that an
algebraic variety of dimension

n-manifold, the converse


complex 2-manifolds.
plex

n has a desingularization1 which is a comis false: already there are non-algebraic

However, it is true that any global analytic object (functions or differential forms, for example) on a projective algebraic variety viewed
as a complex manifold, is a algebraic. This is Serre's GAGA (global
analytic

global algebraic) principle. For example, global meromor-

phic functions in this context turn out to be nothing but restrictions


to the algebraic variety of rational functions on the ambient projective
space (elements of

C(z1 , . . . , zn )).

5.1. Complex

n-manifolds

These are the generalization of complex

1-manifolds (or of Riemann

surfaces, if we assume compactness) to higher dimension. Once again,


we begin with a Hausdor topological space
and write

U = U U .

with open cover

This is made into a complex

the additional data of an analytic atlas on

X:

{U }

n-manifold

by

that is, a collection of

2
holomorphic coordinates
'

z : U V Cn ,

1normalization

is no longer the correct term (refers to a weaker process which still

produces a singular object)

2As

usual, an underline means a vector or tuple of some kind; in this case,

(z1 , . . . , zn ).
65

z =

66

5. COMPLEX MANIFOLDS AND ALGEBRAIC VARIETIES

or homeomorphisms between

and an open set in

Cn ,

such that the

transition functions

:= z z 1 : V V
are biholomorphic.

X
U

C
V

(z 1,...,z n)space
V := z (U )

Here

and

(z1 ,...,z n)space

V := z (U )

are open subsets of

Cn ,

and we need to explain what biholomorphic means. First, a function

f : Cn C is holomorphic if and only if it looks locally (about each


P
i1 i2
I
I
in
point) like f (z) =
I=(i1 ,...,in ) aI z where z means z1 z2 zn and
aI C are constants. A holomorphic map Cn Cm is just m of these:
(z1 , . . . zn ) = z 7 (f1 (z), . . . , fm (z)). (Since these denitions are local,
n
they immediately have meaning when C etc. are replaced by open
sets.) Finally, biholomorphic simply indicates a bijective map (
is bijective by construction) which is holomorphic in each direction.
To generalize the morphisms of Riemann surfaces introduced in

3.1,

we can dene a morphism

of complex manifolds.

need not be of the same dimension; for X we keep the


'
m
above notation and for Y write Z : U V C . A morphism F
Here

F : X Y

and

is then a collection of continuous functions


the

U )

F : U Y

(agreeing on

such that each composition


Z F z 1 : z F1 (U F (U )) Z (U F (U ))

Pn

5.2.

AS A COMPLEX MANIFOLD

yields a holomorphic map (from a subset of

V Cm ).

If

n = m = 1

67

V Cn

to a subset of

then this reproduces Denition 3.1.6.

Moreover, compositions of morphisms are morphisms.


Basic examples of complex manifolds include (besides Riemann surfaces when

n = 1)

Cartesian products of Riemann surfaces, complex

n-tori


Cn /Z 1 , , 2n
1 , . . . , 2n
projective n-space

(where

Pn :=

are linearly independent over

in

Cn
= R2n ),

and

Cn+1 \{0}
.
h(0 , . . . , n ) (0 , . . . , n ) C i

Demonstrating that

Pn

is a complex

n-manifold

(as we do in the next

section) immediately gives meaning to a morphism of complex manifolds from a Riemann surface to

Pn .

This notion is equivalent to (but

more intrinsic than) Denition 3.1.9, as we shall see.

Pn

5.2.

as a complex manifold

Pn is covered by the open sets Ui := {i 6= 0}, with local coordinates


!

bi
n
0
=
,..., ,...,
: Ui Cn .
zi = (zi1 , . . . , zin ) :=
i
i
i
(Here  i replaces  ,

Vi = Cn ,

and

c
()

means to omit that term.) We

need to check that the transition functions

ji : Vij Vji
ji tells us how to write the zj = (zj1 , . . . , zjn )
functions of the zi = (zi1 , . . . , zin ) in such a way that
!
!
0
bi
j
n
0
i
bj
n
,..., ,..., ,...,
is sent to
,..., ,..., ,...,
i
i
i
i
j
j
j
j

are holomorphic. Now,


as

(where for convenience we assume


subset where

zij 6= 0.

j > i).

Now,

Vij Cn

is simply the

So the correct transition function is

ji (zi1 , . . . , zin ) = (zj1 (zi1 , . . . , zin ), . . . , zjn (zi1 , . . . , zin ))


3it

might be a good idea to glance back at the picture there (for intuition purposes)

68

5. COMPLEX MANIFOLDS AND ALGEBRAIC VARIETIES

where

zik /zij ,
zjk (zi1 , . . . , zin ) =
zi,k1 /zij ,

1/zij,

(5.2.1)

For

z1 =

0
1

k i, k > j
for i + 1 < k j .
for k = i + 1
for

z0 = 10 and
1
are the two local coordinates, while (5.2.1) becomes z1 (z0 ) =
,
z0
P1 , zi

reduces to

zi (i = 0, 1).

More precisely,

so that we recover Example 2.2.3. Here is a schematic picture of the

P1 :

1 / 0

0 /

local coordinates on

0
For

1 2
,
0 0

z0 = (z01 , z02 ) =


z2 = (z21 , z22 ) = 02 , 21 , with

and

, we have

e.g.

0 2
,
1 1

z1 = (z11 , z12 ) =
,


01
20 (z01 , z02 ) = z102 , zz02
.

Again, the local coordinates can be visualized as follows:

[0:0:1]

1 / 2

0 / 2

/
2

1 / 0

0 /

[0:1:0]

[1:0:0]
So, for instance, the coordinates
plement

U1

2/ 1
1

z1 =

0 2
,
1 1

are dened on the com-

of the vertical line, and both vanish at

[0 : 1 : 0].

Remark 5.2.1. Whenever you have a local holomorphic coordinate

'

zi : Ui Vi Cn , the inverse mapping i = zi 1 : Vi


Ui X (or just Vi , X ) is called a local analytic chart. In case
X = Pn , i : Cn , Pn is given by
(system) like

i (zi1 , . . . , zin ) = [zi1 : : zii : 1 : zi,i+1 : : zin ]


Cn , (Cn+1 \{0})  Pn .
P1 and P2 :

, and one can visualize this as a map from


Here are pictures of the image of

for

5.2.

0=1

C2

Pn

AS A COMPLEX MANIFOLD

69

U0

U0
0
1
0=1

The next statement says that the notion of holomorphic map from
a Riemann surface to projective space (Defn.

3.1.9) is just a special

case of morphism of complex manifolds. It is enough to consider (as


we do) the situation where the image is not contained in a coordinate
hyperplane, since these are just smaller-dimensional projective spaces
(included into

Pn

by morphisms).

M be a complex 1-manifold, and consider


F : M Pn with F (M ) not contained in any

Proposition 5.2.2. Let

a continuous mapping

{Zi = 0}. The following statements are then equivalent:


(i) F is a morphism of complex manifolds;
(ii) each composition [i F : j F ] is a morphism of complex
1
manifolds to P (on the open subset of M where it is well-dened);

(iii) each i F gives a meromorphic function on M .


j
Ui := {Zi 6= 0} Pn as above. For each
{i, j} (where i 6= j ), the projections ij : Ui Uj  P1 dened by
[0 : : n ] 7 [i : j ] are morphisms of complex manifolds. So if
F
M Pn is one, then ij F = [i F : j F ] is one too, showing
(i) = (ii). Next, (ii) = (iii) is Remark 3.1.10. Finally, if all
j
the
F give meromorphic functions on all of M , then in particular
i


j
0
\
i
n
zi F = i F, . . . , i F , . . . , i F, . . . , i F is holomorphic on
Proof. First, write

F 1 (Uj )

(or a suitable covering of it by coordinate neighborhoods).

These give the local holomorphic representations of


morphism, proving

(iii) = (i).

We will rene Proposition 5.2.2 in Chapter

required for a


7

below.

Whilst we are dwelling on the subject of projective space, I would


like to mention (just for

P2 )

a trick for drawing the real solution sets

70

5. COMPLEX MANIFOLDS AND ALGEBRAIC VARIETIES

of homogeneous equations on the page: barycentric coordinates. First


draw 3 points

A(0) , A(1) , A(2)

on a piece of paper:

0
Think of these as vectors

[0 : 1 : 2 ]

is. Now, plot

A(i) R2 ;
as

2
X

(5.2.2)

it doesn't matter where the origin

P2
i=0

j=0 j

A(i) .

To draw an algebraic curve, simply nd all the solutions


with

i R,

and use (5.2.2) to plot them.

Example 5.2.3. (i) The line

completes to

[0 : 1 : 2 ]

2 = 0 .

y=

(ii) The conic

xy = 1

y =

R) projectively
[1 : x : ] in this way gives

(assume

Plotting the points

[1: x :]

completes to

1 2 = (0 )2 ,

and we get

5.3. AFFINE AND PROJECTIVE ALGEBRAIC VARIETIES

71

for its real barycentric plot.


(iii) The cubic curve

0 )(1 0 )

y 2 = x(x + 1)(x 1)

becomes

(2 )2 0 = 1 (1 +

with picture

In fact, this is the precise meaning of the schematic real 1-dimensional


pictures of complex algebraic curves we have been drawing and will
continue to draw  we are plotting the real solutions in barycentric
coordinates.

5.3. Ane and projective algebraic varieties


We are going to approach this from a slightly more algebraic angle
than, take the common solution of a bunch of polynomial equations.
Start with the commutative ring

Sn := C[z1 , . . . , zn ]

of polynomials in

variables.
Let

J Sn

be an ideal. The ane variety associated to

V (J) := {z = (z1 , . . . , zn ) | f (z) = 0 f J} Cn ,

is

72

5. COMPLEX MANIFOLDS AND ALGEBRAIC VARIETIES

which is the vanishing locus of all polynomials in

J.

By a result in

Sn is nitely
generated, that is, of the form (f1 , . . . , fk ); consequently V (J) is simply
of the form f1 (z) = = fk (z) = 0. However, working in terms
algebra known as Hilbert's basis theorem, any ideal in

of ideals does have a payo, in the form of the famous theorem on


zeroes or nullstellensatz :

g Sn
to J .

Theorem 5.3.1. [D. Hilbert, 1893] If

on

V (J),
If

does,

then for some

m N,g

belongs

J = (f ), then this just says if g


then f divides some power of g .

vanishes identically

vanishes (in

Cn )

wherever

Sn+1 = C[Z0 , . . . , Zn ].
d
Its underlying additive group can be viewed as the direct sum d Sn+1 ,
d
where Sn+1 denotes homogeneous polynomials of degree d in n + 1
variables. Hence, any polynomial G can be written uniquely as a nite
P
sum of homogeneous terms
d Gd .
Next we consider the projective case, writing

Definition 5.3.2. An ideal

I Sn+1

is homogeneous if and only

if the condition

GI

Gd I (d)

is satised.
The projective variety associated to a homogeneous ideal

I Sn+1 is4

V (I) := {[Z] = [Z0 : : Zn ] | F (Z) = 0 F I} Pn .


A version of the Nullstellensatz suited to this case, which is an immediate consequence of Theorem 3.1, is:
Corollary 5.3.3. Given a homogeneous polynomial

on all of

V (I),

some power of

Remark 5.3.4. If

F1 , . . . , F k

belongs to

vanishing

I.

are homogeneous polynomials (of var-

ious degrees), then

I := (F1 , . . . , Fk ) is a homogeneous ideal;


(I) = {F1 (Z) = = Fk (Z) = 0}.
(ii) V
(i)

4technically,

one should keep track of multiplicities of irreducible components,

rather than just dening


detail.

and

V (I)

as a set.

For the most part we will suppress this

5.3. AFFINE AND PROJECTIVE ALGEBRAIC VARIETIES

73

As in the case of curves, we want to be able to go between the ane


and projective settings. To restrict a projective variety to the ane
world, start with the ring (or algebra) homomorphism

Sn+1 Sn
induced by

F (Z0 , Z1 , . . . , Zn ) 7 F (1, z1 , . . . , zn ).
If we write

for the image of a homogeneous ideal

under this map,

then

V (I ) = V (I) Cn .
To go the other way, recall the space
of degree at most

d.

Pnd = dj=1 Snj

of polynomials

We have a homomorphism of abelian groups (or

vector spaces)

d
d
Pnd
Sn+1

which is dened by

f (z) 7 (Z0 )d f (Z/Z0 ) .


J Sn an ideal. Given generators {fi }ki=1 (fi of degree di )
for J , so that J = (f1 , . . . , fk ), set J := (d1 (f1 ), . . . , dk (fk )). Then we
Now, take

have

Cn = V (J).
V (J)
So if

n = 3,

then

V (J)

is adding stu in the  P

at innity

{Z0 = 0}

to complete your ane variety to a projective one, as suggested by the


picture:

hyperplane {Z 0=0} at

in which the black points get added in the process of completing the
curve.

74

5. COMPLEX MANIFOLDS AND ALGEBRAIC VARIETIES

Example 5.3.5. A key example of ane or projective varieties are

the hypersurfaces cut out of

5
d
Sn+1
and

Cn

or

Pn

by a single equation. Let

X = V (F ) = {F (Z) = 0} Pn
the corresponding projective hypersurface of degree

d.

(Like algebraic

curves, these are called linear, quadric, cubic, quartic, quintic, etc.

d = 1, 2, 3, 4, 5, . . ..) We will dene dimension rigorously


below, but X is (n 1)-dimensional (and thus of codimension 1).
Now since Sn is a unique factorization domain, we can factor
according as

F =

k
Y

Fimi ,

i=1
uniquely (up to order), where each

Fi

is prime (irreducible). We can

then write unambiguously

X=

k
X

mi Xi ,

i=1
and say

is reduced when all

mi = 1,

and irreducible when

k = 1.

Exercises
(1) Show that each projection

ij : Ui Uj  P1

described in the proof

of Prop. 5.2.2 is a morphism of complex manifolds. [This is a quick


one.]
(2) Sketch the real solutions of [the projective closure of]

ai )}

5here

in

, if

a1 < a2 < < a2g+2

we really mean

V (F ).

V ((F )),

{y 2 =

Q2g+2
i=1

(x

are real numbers.

the variety of the ideal

(F ),

but we shorten this to

CHAPTER 6

More on projective algebraic varieties

We warm up with two examples we can get our hands on immediately: linear varieties and quadric hypersurfaces.

Then we launch

into what it means for an algebraic variety to be singular resp. smooth


at a point, and in the latter case introduce its tangent space at that
point (which is a linear variety).

This leads to a careful denition

of dimension for algebraic varieties. As a sort of appendix I'll give


a long-overdue introduction to plane curve singularities (which were
glossed over in Chapter 2).

6.1. Linear subvarieties of

Pn

We start by generalizing the projectivities of Chapter 4. Recall


that the projective general linear group is dened as the quotient of
invertible matrices by the scalar action:

P GL(n + 1, C) :=

GL(n + 1, C)
.

+
*
0


..
.
C

This group acts on projective space by the rule

P GL(n + 1, C) Pn Pn
(M, [Z]) 7 [M.Z] =: T (M )[Z].
That is, for each
of

M P GL(n + 1, C), T (M )

gives an automorphism

as a complex manifold. In fact, the projectivities

T (M )

give all

automorphisms (generalizing the last exercise of Chapter 4), but we


won't prove this here.
75

76

6. MORE ON PROJECTIVE ALGEBRAIC VARIETIES

A system of

linear equations

`10 0 + + `1n n = 0
.
.
.

(6.1.1)

`k0 0 + + `kn n = 0

denes a linear subspace

V Pn .

Recalling that the rank of a matrix

is its number of linearly independent row (or equivalently, column)


vectors, the matrix

`10
.
.
.

..

.
.
.

`k0
has rank(L)

=: r k .

codim(V

`1n

=: L

`kn

Dening

) := r (equivalently, dim(V ) = n r),

we have the

Proposition 6.1.1. (i) All projective linear subvarieties of the

same (co)dimension are projectively equivalent.


(ii) A linear subvariety of

nr

Pn

of codimension

is isomorphic to

as a complex manifold.

L and V as above, note that if the equations (6.1.1)


are not independent (i.e. k > r ), then without changing V or r , we can
eliminate equations (reducing k ) until they are (and k = r , i.e. L has
maximal rank). Assume this has been done, so that reordering Zi 's if
Proof. Given

necessary,

det

`10
.
.
.

..

`k0

`1,k1
.
.
.

6= 0.

`k,k1

M be the (n + 1) (n + 1) matrix whose rst k rows are given by


L (a k (n + 1) matrix) and last n k + 1 rows by (0 , Ink+1 ), where
0 denotes a (n k + 1) k matrix of zeroes and Im always means an
m m identity matrix.
n
Consider the automorphism T (M ) of P . By denition of V , [] V
if and only if (matrix multiplication by) L kills . So one should view
T (M ) as taking V to the subspace V0 = {0 = = k1 = 0},
Let

6.2. QUADRIC HYPERSURFACES

which proves (i) since


evidently a

Pnk

77

was arbitrary. This also proves (ii) since

[k : : n ]).

(with homogeneous coordinates

A linear subvariety of codimension

V0

is

is called a hyperplane.

6.2. Quadric hypersurfaces


Recall that a (projective) hypersurface is a subvariety

F (Z) = 0.
2), so that X

out by a single homogeneous equation


in the case where

2
Sn+1
(degree

X Pn

cut

We are interested
is a quadric. The

polynomial can be written

F (Z) = t ZBZ =

Z0

Zn

b00
.
.
.

..

.
.
.

bn0
with

b0n

Z0
.
.
.

bnn

Zn

symmetric. Under a linear change of projective coordinates

Z0
.
.
.

=M

Y0
.
.
.

(M GL(n + 1, C)),

Yn

Zn
we nd

F (Z) =

Y0

Yn

M BM

Y0
.
.
.

=: G(Y ),

Yn
t

where (as in Chapter 4)

M BM

is said to be cogredient to

Lemma 6.2.1. [Sylvester's Theorem

complex
matrices

(n + 1) (n + 1)

matrix

/C]

Any given symmetric

is cogredient to exactly one of the

Mk =

B.

..

1
0
..

0
where

is the number of

1's.

Corollary 6.2.2. A given quadric hypersurface in

Pn

is projec-

tively equivalent to (or transformable by a linear change of coordinates

78

6. MORE ON PROJECTIVE ALGEBRAIC VARIETIES

into) exactly one of the quadrics

Qk =

( k1
X

)
Yj2 = 0

(k = 1, . . . , n + 1).

j=0

Q1 = {Y02 = 0} is a double hyperplane, Q2 = {Y02 +Y12 =


union of two hyperplanes (each
= Pn1 ), and Qk for k 3 is

Note that

0}

is a

irreducible (equation does not factor). You'll investigate these a tiny


bit further in one of the exercises.

6.3. Singularities, tangent planes, and dimension


We'll need the Euler formula mentioned in

2.1,

so let's prove it

rst:
Lemma 6.3.1. [Euler's formula]

d
=
F Sn+1

Pn

i=0

F
Zi Z
=
i

d.F
(F =)Z0d0 Zndn ,
P
P
P

di = d. We have i Zi Z
(Z0d0 Zndn ) = i Zi Zdii (Z0d0 Zndn ) =
i
P

( i di )Z0d0 Zndn = dZ0d0 Zndn .
Proof. It suces to check this on monomials

Now, the denition of smoothness for hypersurfaces is similar to


what we have learned for curves; the general case of varieties cut out
by more than one equation is trickier.

So we'll start, then, with an

ane hypersurface

V = V (f ) Cn ,
and a point

pV.

f
V is smooth at p z
(p) 6= 0 for some
j
j {1, . . . , n}. Otherwise, p is a singular point (or singularity ) of V .
(ii) If V is smooth at all of its points, V is smooth. Otherwise, V is
Definition 6.3.2. (i)

singular.
(iii) If

V is
(

smooth at

p,

dene the tangent plane

(z1 (p) + 1 , . . . , zn (p) + n ) |

Tp V :=

n
X
i=1

(Here the

)
f
i
(p) = 0 Cn .
zi

i C.)

So one can think of


coordinates

{i }.

Tp V

as a copy of

Cn1

with origin at

and

It's also worth noting the formal correspondence

6.3. SINGULARITIES, TANGENT PLANES, AND DIMENSION

79

Tp V ) and dierential operators

i i zi . This is not misleading at all, and in fact the

between tangent vectors (points in


at

p,

namely

intrinsic construction of tangent planes (for complex or more generally


dierentiable manifolds) uses local dierential operators.
As for singularities, i.e.

fzn (p) = 0,

f (p) = fz1 (p) = =

points where

we saw examples of those in Example 2.1.10 for curves.

y 2 = x3 x2

(0, 0) since both partials


of y x + x . On the other hand, y = x x is smooth because

this equation together with 0 =


(y 2 x3 + x) = 3x2 + 1 and
x

0 = y
(y 2 x3 + x) = 2y admit no common solution. This is easy to
1
1
see: the points ( , 0) and ( , 0) where both partials vanish, do not
3
3
Here is one more:

is singular at

lie on the curve.


Next, consider a projective hypersurface

V = V (F ) Pn ,
where

is homogeneous and

P V.

F
V is smooth at P Z
(P ) 6= 0 for some
j
j {0, . . . , n}. Otherwise, P is a singular point (or singularity ) of V .
(ii) If V is smooth at all of its points, V is smooth. Otherwise, V is
Definition 6.3.3. (i)

singular.

V
(

(ii) If

TP V :=

is smooth at

P,

dene the tangent plane (


=

[Z0 (P ) + 0 : . . . : Zn (P ) + n ] |

n
X

i=0

Pn1 )
)

F
(P ) = 0
Zi

Pn .

Now a priori, the denition of a singular point is one at which

F (P ) = FZ0 (P ) = = FZn (P ) = 0; but by the Euler


X
F
(6.3.1)
Zi (q)
(P ) = deg(F ) F (P )
Z
i
i
and so it is in fact enough to check

formula,

FZ0 (P ) = = FZn (P ) = 0.

In fact,

(6.3.1) also implies (for the projective case only!) the simplication

(
(6.3.2)

[0 : : n ] |

TP V =

X
i

F
(P ) = 0
i
Zi

Note that (6.3.2) is really just the solution set of


Chapter 4 (but now in

rather than

).

F (P ) = 0,

as in

80

6. MORE ON PROJECTIVE ALGEBRAIC VARIETIES

As you might expect, the notions of tangent plane in ane and


projective cases agree, in the sense that  at a point on an ane
hypersurface  the tangent plane of the projective completion is the
completion of the tangent plane:

Tp V (F (1, z1 , . . . , zn )) = T[1:p] V (F )Cn , where


[1 : z1 (p) : : zn (p)].

Proposition 6.3.4.

(P =)[1 : p]

means

Proof. Given

and

Q = [1 : q],

q = (z1 (q), . . . , zn (q)) Cn .

Writing

f (z) = F (1, z)

we want to show

q Tp V (f ) Q TP V (F ).

(6.3.3)

The left-hand (ane) condition is, writing

zi (q) = zi (p) + i

in Deni-

tion 6.3.2(iii),

n
X
f
(zi (q) zi (p))
(p) = 0.
zi
i=1
This is really

n
X

(Zi (Q) Zi (P ))

i=1

F
(P ) = 0,
Zi

which by Euler becomes

n
X

Zi (Q)

i=1
Since

F (P ) = 0,

F
F
(P ) deg(F ) F (P ) + 1
(P ) = 0.
Zi
Z0

we get

X
F
F
1
(P ) +
Zi (Q)
(P ) = 0,
Z0
Z
i
i=1
which is exactly the right-hand (projective) condition of (6.3.3).

Now let's have a look at singularities and smoothness in the general


projective case. The denition is complicated, but after this chapter
we won't use it much. Let

V = V (F1 , . . . , Fk ) Pn ,
and

be a point on

V.

6.3. SINGULARITIES, TANGENT PLANES, AND DIMENSION

Definition 6.3.5. (i)

neighborhood

1
such that

W Pn

of

V
p

is smooth at

81

if and only if there exists a

and sub-index set{i1 , . . . , ic }

{1, . . . , k}

V W = V (Fi1 , . . . , Fic ) W , and


Fi1

F
(p) Zin1 (p)
Z0

.
.
..
.
.
(b) rank
.
= c.
.
.

(a)

Fic
(p)
Z0
has codimension

V
(ii) If V

We say

Fic
(p)
Zn
(or dimension

n c) at p.
is smooth at each point p V , then V is smooth
c

(otherwise,

is singular ).
(iii) If

has the same (co)dimension at each smooth point

p V,

V is equidimensional. If moreover that codimension is c, we just


2
say V is a variety of codimension c (dimension n c).
n
(iv) The tangent plane Tp V P to V at a smooth point p is the
solution set of L.p = 0, where

F1
F1
(p)
(p) Z
Z0
n

.
.
..
.
.
L=
.
.
.
.

then

Fk
(p)
Z0

Fk
(p)
Zn

Condition 6.3.5(i)(a) says that locally about

= Fic (Z) = 0,

p, once you set Fi1 (Z) =

the remaining equations are redundant; and roughly

speaking, the condition (b) on rank says that no more (none of the

F i` )

are redundant. In the terminology of

6.1, Tp V

is a linear subva-

riety, and it follows from condition 6.3.5(i)(b) that its codimension is

c.
V

That is, we have really just dened the (co)dimension of a variety


at a smooth point

p,

to be the (co)dimension of

Tp V

 something

we already knew how to dene.


Finally, the neighborhood
set containing

in the denition is an analytic open

(such as a ball), but the denition would also work

if we only permitted algebraic open sets dened by complements of


(other) subvarieties,

3 known as Zariski open sets. In general, if you

1the matrices in this denition assume a particular representative (Z (p), . . . , Z (p))


0
n
in

Cn+1

of the projective coordinate

[Z0 (p) : : Zn (p)].

It doesn't matter which

one you take, as long as you are consistent.

2note that the dimension of V


a

m-dimensional variety must

3this

at a singular point is not dened, so the denition of


be one that is of dimension

is exactly how Example 6.3.6(iii) is started below

m at all smooth

points.

82

6. MORE ON PROJECTIVE ALGEBRAIC VARIETIES

want to view an algebraic variety as a complex analytic space (or manifold, if it is smooth), then you must use analytic open sets; on the
other hand, the Zariski open sets introduce a dierent topology on
or

, which is coarser but has the advantage of being algebraic (and

is still Hausdor ). We need both. In brief, when we study varieties analytically, we use the analytic topology; when we want to make heavy
use of the correspondence between varieties and ideals in commutative
rings, we use the Zariski topology.

V be the ane variety {z12 + z22 + z32 = 0}


P 2
derivatives zi (
j zj ) = 2zi all vanish at p = (0, 0, 0)

Example 6.3.6. (i) Let

C3 .

The partial

and so

is singular there:

Z1 Z3 = 0
3
(ii) Now for a nasty one. Let V P be dened by
Z2 Z3 = 0
and take p = [1 : 0 : 0 : 0], q = [1 : 0 : 0 : 1], r = [1 : 1 : 0 : 0]:

)
,

r, Z1 6= 0 and so having set Z1 Z3 = 0 (i.e. Z3 = 0), the


second equation Z2 Z3 = 0 is redundant. So the relevant matrix from



6.3.5(i)(b) is
Z0 (Z1 Z3 ) Z1 (Z1 Z3 ) Z2 (Z1 Z3 ) Z3 (Z1 Z3 ) =
r




|
=
,
which
has
rank
1
proving
that
0 Z3 0 Z1 r
0 0 0 1
V has codimension 1 (dimension 2) at r. Locally about q , Z3 6= 0 and
so the equations are eectively Z1 = 0 and Z2 = 0, neither of which is
Locally about

6.4. SINGULARITIES OF PLANE CURVES

83

0 Z3 0 Z1
0 0 Z3 Z2

redundant. The matrix in 6.3.5(i)(b) is now

!


=

q

0 1 0 0
,
0 0 1 0
mension 2 at q .

which does have rank


So

conrming that

is not equidimensional.

Finally, at

equation is redundant but the matrix 6.3.5(i)(b) is


meaning

is singular at

has codi-

neither
!

0 0 0 0
0 0 0 0

p.

(iii) Finally, consider the variety


tions

C P3

dened by the three equa-

Z0 Z3 Z1 Z2 = 0 (I)
Z12 Z0 Z2 = 0
(II)

2
Z2 Z1 Z3 = 0 (III)

Ui := {Zi 6= 0} of P3 . In U0 , we can divide by Z0 ,


Z12
2
so that (II) becomes Z2 =
. Together with (I), this gives Z2 =
Z0
Z2 Z2 = Z1 Z2 ZZ01 = Z0 Z3 ZZ01 = Z1 Z3 . Consquently, (III) is redundant on
U0 . Now, for any point in U0 C , you can check that the matrix in
6.3.5(i)(b) has rank 2, showing that C is smooth of dimension 1 at all
of those points. To nish, and show that C is a 1-dimensional smooth
variety, carry out a similar analysis in each of U1 , U2 , and U3 (exercise).
and the covering

1 is called a curve,
d 3 a d-fold. So-called

Somewhat unsurprisingly, a variety of dimension


of dimension

a surface, and of dimension

Calabi-Yau threefolds play a central role in mathematical string theory.

6.4. Singularities of plane curves


Consider a curve

C = {F (Z) = 0} P2
F S3d (of degree 3 in Z0 , Z1 , Z2 ).
A point p C is a singularity if and only if FZ0 (p) = FZ1 (p) = FZ2 (p) =
0, and (moving C by a projectivity if necessary) we may assume that
p = [1 : 0 : 0]. To locally analyze C at p, we can pass to ane coordiZ
Z
nates x = 1 , y = 2 and replace F by
Z0
Z0
dened by a homogeneous polynomial

f (x, y) = F (1, x, y) =

d
X
m=k

fm (x, y),

84

6. MORE ON PROJECTIVE ALGEBRAIC VARIETIES

fm S2m for each m, and fk 6= 0. Now 0 = fx (p) = fy (p) = f (p)


(p = (0, 0)) translates to 0 = f1 = f0 , so that k 2. We say that p is
a k -tuple point of C , or a singularity of order k .
where

So far, we have not dened tangent planes at singular points. Indeed, this can really only be done for curves in general.

0 should be, we think of [x : y] as


1
homogenous coordinates on the P of lines through (0, 0) = p. The
lowest-order homogeneous term fk of f denes a 0-dimensional vari1
ety p (C) := {fk (x, y) = 0} in this P . For each [x0 : y0 ] p (C),
what the tangent lines to

one should think of

To decide

at

y0
as the slope of a line tangent to some local
x0

irreducible component

4 of the curve

at

p.

C at a
of p (C).

Definition 6.4.1. The tangent lines to a curve

are the lines through

corresponding to points

singularity

fk (x, y) = 0 has k solutions counted with multplicity. If these


are all distinct, i.e. if p (C) is reduced, then we say p is an ordinary
k -tuple point. The most geometric way to think of this is that C has
k distinct tangent lines at p.
Any line through a k -tuple point p other than one of C 's tangent
lines there, meets C with multiplicity k at p: if L is given parametrically
by t 7 (at, bt) (fk (a, b) 6= 0) then the intersection multiplicity is comk
puted as in Chapter 2 by taking the order of f (at, bt) = t fk (a, b) +
at t = 0.
Now,

ab xa y b it
can be useful (for various purposes) to plot the nitely many (a, b) with
ab 6= 0. If you do this when (0, 0) is a k -tuple point and f has degree
n, then these lie in the shaded region
Remark 6.4.2. Given a polynomial

4this

f (x, y) =

will be made precise when we do local normalization

(a,b)Z20

6.4. SINGULARITIES OF PLANE CURVES

85

d
k

This may be useful for one of the exercises below.


There is more to singularities, it turns out, than their order or even
the tangent line conguration reected by

p (C).

A local analytic clas-

5
sication of so-called simple singularities of curves has been carried
out.

For the purposes of this classication, if

are two curves

p are considered equivalent if


0
2
there are small neighborhoods U, U of (0, 0) in C and a biholomor'
0
0
phism U U carrying p to p and C to C . The dierent classes of
through

p = (0, 0),

C, C 0

their singularities at

simple singularities carry A-D-E labels, which reect a relation to


other classications in mathematics (simple Lie algebras/Dynkin diagrams, rational surface singularities, etc.)
The results are that double points are all equivalent to one of

An : x2 + y n+1 = 0

(n 1),

and (simple) triple points to one of

Dn : y(x2 + y n2 ) = 0

(n 4),

E6 : x3 + y 4 = 0,
E7 : x(x2 + y 3 ) = 0,
E8 : x3 + y 5 = 0.
The ODP's (ordinary double points: two distinct tangents) are all of
type

A1 , as all An2

have only one tangent; amongst the latter, cusps

are the singularities of type

5cf.

A2 .6

OTP's (ordinary triple points) are all

[Barth, Hulek, Peters, and van de Ven, Compact complex surfaces, Springer,

2004] for the denition. Simple singularities encompass all double (2-tuple) points
and some triple (3-tuple) points, and nothing of higher order.

6refer

back to Chapter 2 for a few pictures (cusps, ODP, OTP)

86

6. MORE ON PROJECTIVE ALGEBRAIC VARIETIES

D4 ; we note that the tangent lines to y(x2 + y 2 ) = 0 have slopes


0, i, i. All Dn5 have two distinct tangents (one with mutliplicity 2)
and the E6,7,8 each have one tangent (of multiplicity 3).

of type

Exercises

Pn dened by a symmetric
bilinear form B is smooth if and only if det(B) 6= 0.
(ii) Corollary 6.2.2 associates a number k to each projective quadric
n
hypersurface in P . Show that any two are projectively equivalent
if and only if they have the same value of k . [This is easy.]
2
2
3
Show that [the closure in P of] y = 4x + ax + b is smooth unless
a3 + 27b2 = 0.
4
4
4
Find the tangent plane to the complex surface 2x +y +z 4xyz =
0 (in C3 ) at the point p = (1, 1, 1).
Finish the proof in Example 6.3.6(iii) that C is a smooth curve.
2
What form does a degree k projective algebraic curve (in P ) take
if it has a singularity of order k ?
2
2 2
2
3
2
Analyze the singularity of C = {(x + y ) + 3x y y = 0} C

(1) (i) Prove that a quadric hypersurface in

(2)

(3)

(4)
(5)

(6)

at the origin. (What is its order, and type?)


(7) For which values of

are the algebraic curves

singular (in (a) and (b) below)?

F (X, Y, Z) = 0

in

Attempt a sketch of each of

the singular curves, saying where the singularities are located and
what type they are.

F (X, Y, Z) = X 3 + Y 3 + Z 3 + (X + Y + Z)3 ,
3
3
3
(b) F (X, Y, Z) = X + Y + Z + 3XY Z .

(a)

CHAPTER 7

Smooth varieties as complex manifolds

This Chapter starts the long slog toward a proof of part (A) of the
Normalization Theorem 3.2.1. After introducing a bit of the theory of
several complex variables, we'll use the holomorphic implicit function
theorem to put a complex manifold structure on any smooth irreducible
(ane or projective) algebraic variety:
Theorem 7.0.3. A smooth irreducible algebraic curve

C Pn

is

M
: M , P with C

a Riemann surface. (More precisely, there exists a Riemann surface


and an injective morphism of complex manifolds

as its image.)
This is, of course, the smooth case of Thm. 3.2.1(A). As for going
the other way, from Riemann surfaces to algebraic curves, here is a
statement which is dierent in character from 3.2.1(B):

M with n + 1 linearly indef0 , . . . , fn K(M ), yields an algebraic

Theorem 7.0.4. A Riemann surface

pendent meromorphic functions


curve in

Pn

not contained in any proper linear subvariety.

We won't prove this in full  just the existence of a morphism

M Pn

of complex manifolds which is nondegenerate, i.e.

image is not contained in any

n1

whose

. Proving that the image is described

by algebraic equations (hence yields an algebraic curve) is harder.

7.1. Background from several complex variables

On (or C{z1 , . . . , zn }) denote the ring of convergent power series


I aI z in z1 , . . . , zn , or equivalently, holomorphic functions dened on
n
some neighborhood of 0 C (cf. 5.1).
Let

Proposition 7.1.1. [W. Osgood, 1900] Let

be a function on

0C

which is holomorphic in each zi as the


f
other {zj }j6=i are held xed; that is, z = 0 (i). Then f is in fact a
i
an open neighborhood of

holomorphic function (and so gives an element of


87

On ).

88

7. SMOOTH VARIETIES AS COMPLEX MANIFOLDS

Proof. We will only give the proof for

phic in

z2 ,

we have

f (z1 , z2 ) =

n = 2.

Since

is holomor-

f (z1 , 2 )
d2 ;
2 z1

2 1

z1 , this

using the holomorphicity in

(2 1)2

=
(2 1)2

f (1 , 2 )
d1 d2
(1 z1 )(2 z2 )

f (1 , 2 )d1 d2


1
1 2 1 z11

z2
2

.

Now using the power-series expansion

1
1

zi
i

X  zi k
k0

whose uniform convergence allows us to swap integration and summation, we nd

X 

f (z1 , z2 ) =

k1 ,k2 0

(2 1)2

f (1 , 2 )d1 d2
1k1 +1 2k2 +1

z1k1 z2k2 .


In order to put a complex manifold structure on a smooth variety,


we will need a way to parametrize zero-loci of holomorphic functions.
This is given by the holomorphic implicit function theorem which here
I will just state and prove in the two variable case.
Proposition 7.1.2. Let

Then there exists

w O1

f O2

f (0, 0) = 0,

such that in a neighborhood of

f (z1 , z2 ) = 0
The upshot of this is that
on

with

z2

f
(0, 0)
z1

6= 0.
(0, 0) in C2 ,

z1 = w(z2 ).

gives a local holomorphic coordinate

{f (z1 , z2 ) = 0}.
Proof. We will assume the

just check that the

implicit function theorem, and

it yields is holomorphic:

0=
=

f (w(z2 ), z2 )
z2

f
f
w
f
w
(w(z2 ), z2 ) +
(w(z2 ), z2 )
+
(w(z2 ), z2 )
.
z2
z1
z2 z1
z2

7.2. SMOOTH NORMALIZATION

Now since

f
z1

f O2 ,

locally. So we nd that

f
z2

w
z2

= 0; moreover, by
= 0, so that w O1 .

89

assumption

f
z1

6= 0


Here is a visual explanation of why the nonvanishing condition on

f /z1

matters:

z2

z2

z1

f (0)=0
z1

f (0)=0
z1

In the left-hand picture, you can write

z1

z1 as a function of z2 (as desired);

on the right-hand side, you cannot.

7.2. Smooth normalization


The more general statement which implies Theorem 7.0.3 is:
Theorem 7.2.1. Given

n-manifold X ;
a system of open neighborhoods {W X} covering Y (with local
holomorphic coordinates z = (z1 , . . . , zn ));
holomorphic functions f1 , . . . , f` O(W ) (for each ) such
1
that Y W = V ({fj }j=1,...,` ) W ; and (also for each )

a closed connected subset

of a compact complex

rank {fj /zk }


= ` [the Jacobian condition ].
j = 1, . . . , `

k = 1, . . . , n
Then Y is an (n `)-dimensional compact complex (n `)-manifold.
We say that

is a codimension-` complex submanifold of

X.

In

fact, Theorem 7.2.1 immediately gives:


Corollary 7.2.2. Any smooth irreducible projective algebraic va-

riety

Y Pn

of dimension

1this condition makes Y


the vanishing locus

is a compact complex

into an analytic subvariety of

f1 = = f` = 0,

d-manifold.

X ; here V (f1 , . . . , f` ) means

just as in the algebraic setting.

90

7. SMOOTH VARIETIES AS COMPLEX MANIFOLDS

Proof. Put

X = Pn and ` = nd.

That

is smooth of dimension

(Defn. 6.3.5) implies the Jacobian condition required in Thm. 7.2.1.


Now we prove the Theorem.

Proof. Rening the covering if necessary, we can arrange to have


det

(7.2.1)

Write  z  for
I

z I , z II

(z1 , . . . , z` )

fj
zk

6= 0.
1j,k`

and  z

II

 for

(z,`+1 , . . . , zn ).

(So

z =

A schematic picture:

z
z

By the condition (7.2.1), and the general holomorphic implicit function


theorem, we have holomorphic functions
subsets of

Cn`

to

C` )

{w }

(mapping from open

such that



Y W = z I = w z II


for each . Hence, the z
give local coordinates
II
which constitute an open cover of Y .
Consider the transition functions for X

on the

{Y W },

'

z , z
I

II

: z (W ) z (W )

 




7 I z , z
, II z , z
=: z I , z II
I

II

corresponding to change of coordinates on


describing change of coordinates on

Y : z

II

This is 1-to-1 because

7 II w

II

W .

Clearly the functions

Y W are then


, z
=: z II .
z
II

II

is, and holomorphic because

So we have the data of an analytic atlas on

Y.

II

and

are.

7.3. NONDEGENERATE MORPHISMS

91

7.3. Nondegenerate morphisms


The statement related to Theorem 7.0.4 which we shall prove is:
Proposition 7.3.1. Given a Riemann surface

M,

the following

data are equivalent:


(a)

n+1

linearly independent meromorphic functions

fi K(M );

(b) a nondegenerate holomorphic map (morphism of complex manifolds)

: M Pn .

We will need the notion of a meromorphic function on a complex


manifold of any dimension.
Definition 7.3.2. A meromorphic function

plex manifold

X)

{(U , g , h )}
X;

is a collection

{U } is an open cover of
g , h O(U ) (they are
g h = g h on U .
g
2
We write  F =
 on U .
h

F K(X)

(on a com-

such that

holomorphic functions); and

dim(X) = 1, this coincides with the earlier


Denition 3.1.1 (via g/h); by Prop. 3.1.8 meromorphic functions on
1
Riemann surfaces yield morphisms X P . But this does not generalize: if dim(X) > 1, a meromorphic function on X need not even
1
yield a well-dened mapping X P .
Remark 7.3.3. For

X = C2 with complex coordinates x, y .


Then F := x/y (one U = X ; g = x, h = y ) denes a meromorphic
1
function, which is not well-dened (as a mapping to P ) at (0, 0).
Example 7.3.4. Consider

=2

=1
= 12
x

???
2the

third condition says that

g
h

are dened! (see Remark 7.3.3)

g
h on overlaps  at least, where the quotients

92

7. SMOOTH VARIETIES AS COMPLEX MANIFOLDS

The notion of blowing up in algebraic geometry is motivated (in


part) by the desire to remove such indeterminacies. In the last example,

X = C2

P1 of lines
(mapping down onto X ) on
through the origin, yielding a new space X
the idea would be to replace the origin in

by the

which the meromorphic function becomes well-dened as a morphism.


Example 7.3.5. The meromorphic functions on

Pn

and its smooth

subvarieties (viewed as complex manifolds) are the rational functions


P (Z)
Z
d
F = Q(Z)
for P, Q Sn+1 . For instance, the ane coordinates zi = i
Z0
Z
are meromorphic functions (and more generally, zji = i is one).
Zj
Here is how to see at least that rational functions are meromorphic in the sense of Denition 7.3.2. (That meromorphic functions are
rational is more nontrivial.) In

Uj = {Zj 6= 0},

set

gj (zj ) := P (zj0 , . . . , 1 , . . . , zjn ) = P (Z/Zj ) =


j th
entry

hj (Zj ) := Q(zj0 , . . . , 1 , . . . , zjn ) =


j th
entry

then

gj hi =

1
P (Z)
Zjd

1
Q(Z);
Zjd

1 1
P (Z)Q(Z) = gi hj .
Zjd Zid

f : C X from a
Riemann surface to a complex manifold, and let F K(X) be given
by {(g , h , U )}. Assume that f (C)|U {h = 0} is a nite point
1
set, and put W := f
(U ), G := g f , H := h f . Then

f F := {(G , H , W )} (or rather, G/H ) belongs to K(C).


Example 7.3.6. Consider a holomorphic map

The last two examples will be used in the proof of Prop.

7.3.1,

which we now give:

n + 1 meromorphic
points in M which cause a

Proof. The rst issue is how we get from

functions to a morphism to

. The set of

problem is

:= {q M | fi (q) = 0 for

all

i} {q M | fi (q) = for

Dene

f : (M \) Pn

some

i}

7.3. NONDEGENERATE MORPHISMS

93

by

p 7 [f0 (p) : : fn (p)].


q let z be a local holomorphic coordinate with z(q) = 0, then
(f )
write fi (z) = z q i hi (z) (where hi are local holomorphic functions not
vanishing at q ), and put := mini{0,...,n} {q (fi )}. For z 6= 0,


f (z) = z f0 (z) : : z fn (z) ;

Near

none of the entries in this blows up locally, and at least one does not
vanish at

z = 0

(i.e.

at

q ).

Hence,

extends to all of

evident that this extension is still holomorphic as a map


Next, given a morphism

f : M Pn ,

M , and
1
to P .

it is

we want to product an

(n + 1)-tuple of meromorphic functions. Referring to Examples 7.3.5

(for zi ) and 7.3.6 (for f ), simply take fi := f zi and you're done.


Finally, to see that f is degenerate i the {fi } are linearly depenn+1
dent, consider the correspondence between nonzero vectors v C
n
n1
(up to scale) and hyperplanes in P , by taking Pv
to be the projectin+1 v
cation of (C
) . Degeneracy of f occurs i f (M ) Pvn1 for some
v , which is to say (f0 (p), . . . , fn (p)) v for all p M . But this just
P
reads
vi fi (p) = 0 (p), which is a nontrivial linear relation.


We give two examples of nondegenerate projective embeddings of


Riemann surfaces (the rst is actually a series of examples). For these
cases we actually give algebraic equations for the image.

Example 7.3.7. The so-called rational canonical curves are the

images of the nondegenerate morphisms

f : P1 , Pn
given, for each

n N,

by

[Z0 : Z1 ] 7 [Z0n : Z0n1 Z1 : : Z1n ].


(In ane terms, one can think of this as

z 7 [1 : z : . . . : z n ],

with

7 [0 : : 0 : 1].)
Let's see what this looks like for the rst few values of

map.

for

n = 1, f

sends

[Z0 : Z1 ] 7 [Z0 : Z1 ]

n:

and so is just the identity

94

7. SMOOTH VARIETIES AS COMPLEX MANIFOLDS

for n = 2, we have [Z0 : Z1 ] 7 [Z02 : Z0 Z1 : Z12 ]. If we write


[Y0 : Y1 : Y2 ] for the homogeneous coordinates on P2 , then the image is
2
2
the conic {Y1 Y0 Y2 = 0} P .
for n = 3, [Z0 : Z1 ] 7 [Z03 : Z02 Z1 : Z0 Z12 : Z13 ](= [Y0 : Y1 : Y2 : Y3 ])
has image

V := V (Y0 Y3 Y1 Y2 , Y12 Y0 Y2 , Y22 Y1 Y3 ) P3 .


By exercise 4 from Chapter 6 you know that

Example 7.3.8. Let

1-torus.

is smooth.

a lattice) be a complex

We want to demonstrate that there is a (nondegenerate) mor-

phism from
present

M = C/ ( C

to

P2

with a cubic curve as image. Note that this will

as the normalization of such a cubic curve:

Some of the steps will be exercises.


First, there exists a unique meromorphic function

K(C)

satis-

fying

(u + ) = (u) for every and u C


(u) = u2 + h(u), where h K(C) is holomorphic
borhood of 0, has all its poles in \{0}, and h(0) = 0.
Existence is an exercise. Uniqueness is easy: if

in a neigh-

were another such

Q = ( u ) (Q u ) has no pole at 0 and is periodic, hence has no poles in either. But the only possible zeroes
were in , and so Q is entire. By compactness of a fundamental
region for , any -periodic entire function is bounded hence (by Liouville) constant. Since Q is zero at 0, this constant is zero and
= Q.
function,

EXERCISES

95

is an even function
some a, b C. In each

In the exercises below, you will also show that


((u)

= (u))

and

(0 )2 = 43 + a + b

for

case, you get equality by showing the right-hand side minus the lefthand side has no poles and is zero at some point (as in the uniqueness
argument just described). The upshot is that

f : C/ P2
dened by

u 7 [1 : (u) : 0 (u)]

for

u 6= 0

and

0 7 [0 : 0 : 1]
parametrizes (or normalizes)

C = {Z0 Z22 = 4Z13 + aZ1 Z02 + bZ03 },

smooth cubic with the ane equation

y 2 = 4x3 + ax + b.
f (M ) C ,
manifold, and f as

What we have said so far only gives that


the smooth curve

M C,

as a complex

but viewing
a morphism

the open mapping theorem from complex analysis says the

image is open; while on the other hand the image of a compact set by
a continuous map is compact (hence closed in
and closed in

C,

and thus

C ).

So

f (M )

is open

f (M ) = C .

Exercises

f : P1 Pn has the following


property: the image of any collection of k ( n + 1) distinct points
{w1 , . . . , wk } P1 is in general position (spans a Pk1 in Pn ). [Hint:
Vandermonde determinant.] Then, taking k = 2, explain why this
shows f is injective.
Turning to the case n = 3 in Example 7.3.7, (a) actually prove that
V = Image(f ) and (b) that you cannot throw out any of the three
equations dening V .
2
5
2
Show that the map P P given by [Z0 : Z1 : Z2 ] 7 [(Z0 ) :
Z0 Z1 : Z0 Z2 : (Z1 )2 : Z1 Z2 : (Z2 )2 ] is (a) well-dened and (b)

(1) Show that the rational canonical map

(2)

(3)

holomorphic (i.e.

a morphism of complex manifolds), then (c)

write (polynomial) equations expressing the image as an algebraic


variety. (For (c) you can just write the equations and not prove it.)

96

7. SMOOTH VARIETIES AS COMPLEX MANIFOLDS

(4) Let

1 , 2 C be two complex numbers which are R-linearly inde-

pendent, and let

= Z1 + Z2 = {n1 1 + n2 2 | n1 , n2 Z}
be the lattice in

that they generate.

(a) Show that the series

1
(u) = 2 +
u

1
1
2
2
(u )

6= 0
is absolutely and uniformly convergent on any compact subset of
the complex

u-plane which does not contain any of the points of .

[Hint: any compact subset is contained inside one of the following

|u| K |u |  (). Break the sum into terms


with || 2K , and || > 2K , and use (essentially) the Weierstrass

form:

M-test.]
(b) Verify the pole condition in Example 7.3.8: that all poles are
on

and in a neighborhood

holomorphic and

h(0) = 0.

of

0, (u) = u2 + h(u)

with

[Hint: what do you know about an

absolutely and uniformly convergent series of analytic functions?]


(c) Show that

is a doubly-periodic function; that is, show that

(u + ) = (u)
[Hint:

for every

uC

and every

From (a), you can calculate the derivative

ferentiating each term of the series dening

(u + ) = (u),

(u).

0 (u)

by dif-

First prove

then integrate.]

(5) Now forget the explicit formula for

(u) just given,

and retain just

K(C) is -periodic with all poles , and


2
locally of the form (u) = u
+ h(u) with h holomorphic (on
some U C containing a fundamental domain) and h(0) = 0.
0
Prove that (a) (u) = (u) [ = h even = h odd] and (b)
(0 (u))2 = 4((u))3 + a(u) + b for some a, b C. [See hint given

these facts: that

in the Example.]

CHAPTER 8

The connectedness of algebraic curves

The main theorem of this chapter will be that the smooth part

C\sing(C)

of an irreducible algebraic curve

(and then, of course, so is

C ).

C P2

For example, in Exercise 5 of Chapter

3, you showed that the complement of the ODP

p = [1 : 0 : 0]

in the

{Y Z X Z + X = 0}, viewed as a complex

isomorphic to C  which is certainly connected.

singular cubic curve


1-manifold, is

is path-connected

Just so that there is no confusion, we should say what the situation


is for reducible curves right away and why the result does not generalize.
For plane projective algebraic curves with more than one irreducible
component, say

C = Ci ,

the components

Ci

must intersect (this will

be one consequence of Bezout's theorem later), making


But the complement of the singularities in

connected.

will not be connected, as

these will include all of the intersection points.


We begin by introducing a new, somewhat technically involved, tool
for dealing with singularities, intersections, and projections of curves.

8.1. Resultants and discriminants


Let

be a unique factorization domain (UFD), where we recall

that this is a commutative domain in which each element has a unique


factorization into irreducibles, up to reordering and multiplication by
units. In a UFD, amongst other things, the notion of a greatest com-

2 has meaning. By the Gauss lemma,

mon divisor

In practice we will always take

to be

or

C[x].

D[y]

is also a UFD.

(Note that

C[x]

is a

C[x, y] is not.)
m
m1
Consider f (y) = a0 y + a1 y
+ + am , g(y) = b0 y n + b1 y n1 +
+ bn elements of D[y] with a0 , b0 6= 0.
PID, but

1we will show that the set sing(C) of singular points is always nite
2recall that these are well-dened up to units (invertible elements); for

C[x] or C[x, y] the units are C , hence the notion of monic gcd
well-dened).
97

example in

(which is completely

98

8. THE CONNECTEDNESS OF ALGEBRAIC CURVES

Definition 8.1.1. The resultant

the element of

3 of

and

g,

given by the determinant of the

R(f, g), is
(n + m) (n + m)
written

4
Sylvester matrix

M(f,g)

am

0 a0 a1

.. . .
..
.
.
.

0 0

:=
b0 b1

0 b0 b1

.. . .
..
.
.
.
0 0

am

Now writing

a0

a1

a0

bn

bn

R(f, g) = 0
h)

Proof. The gcd (say,

..

.
.
.

..

..

..

bn1

for the eld of fractions of

Proposition 8.1.2.

am1

b0

D,

am

.
0

.
.

0
bn

we have the

gcdK[y] (f, g)

6= 1.5

is nontrivial if and only if

F g = Gf

(8.1.1)

F = A0 y m1 + + Am1 and G = B0 y n1 + + Bn1 in


D[y]. Indeed, if h 6= 1 then put F = f /h and G = g/h. Conversely,
since deg F < deg f and deg G < deg g , and both sides of (8.1.1) factor
into the same irreducibles, f and g have a common factor of degree
> 0.
for some

In turn, (8.1.1) is equivalent to

(8.1.2)

a0 B0 = b0 A0
a1 B0 + a0 B1 = b1 A0 + b0 A1
.
.
.

am Bn1 = bn Am1
n1
{Ai }m1
i=0 , {Bj }j=0 D. To get from
m+n1
coecients of y
, y m+n2 , . . . , 1.

being satised for some


(8.1.2), just take

(8.1.1) to

3also called eliminant, since y is eliminated


4the line in the matrix is just an organizational device  it has no meaning
5two further equivalent conditions: (i) deg (gcd
(f, g)) > 0; and, noting
y

K[y]

is a PID, so that the ideal

D[y]

(f, g)K[y] = (gcdK[y] (f, g)),

(ii)

that

(f, g)K[y] 6= (1)K[y] .

8.1. RESULTANTS AND DISCRIMINANTS

99

Now notice that (8.1.2) can be rephrased in matrix multiplication


terms: there exist

{Ai }, {Bj }

such that

B0

.
.

n1
t
M(f,g) .
A0

.
.

Am1
In other words, we have shown
i.e.

det(M(f,g) ) = 0.
Definition 8.1.3.

= 0.

h 6= 1 is the same as ker(t M(f,g) ) 6= {0},




D(f ) := R(f, f 0 ) is the discriminant

of

f.

Here

f
.
denotes the formal derivative
y
Example 8.1.4. If

f C[y],

then

D(f ) C

is a number, and the

criterion
(8.1.3)

D(f )

vanishes

has a multiple root

follows immediately from Prop. 8.1.2. For the ane curve

z 2 = 4y 3 + ay + b
to be singular, we need two of the roots of the right-hand side to coincide. That is, by (8.1.3), we need


4 0 a



4 0

3
2
0 = R(4y + ay + b, 12y + a) = 12 0 a

12 0



12










a


0 a
b
a b

which after a bit of row-reduction

4
0
0
0
0


0 a
b
0

4 0
a
b

0 2a 3b 0 = 16(4a3 + 12 9b2 )

0 0 2a 3b

0 12
0
a
= 64(a3 + 27b2 ).

100

8. THE CONNECTEDNESS OF ALGEBRAIC CURVES

This recovers the result from Exercise 2 of Chapter 6.

Example 8.1.5. If

f C[x, y],

then

D(f ) C[x]

is a polynomial

and from Prop. 8.1.2 we have:

D(f )

(8.1.4)

vanishes at

The collection of

D(f ),
curve

x0 's

x0 f (x0 , y)

has a multiple root in

y.

where this happens, that is, the set of roots of

is called the discriminant locus for the projection of the ane

{f (x, y) = 0}
y

onto the

x-line:
f=0

x
discriminant locus
Proposition 8.1.6. An irreducible (reduced) algebraic curve

0} P

{F =

has (if any) nitely many singularities.

f (x, y) = F (1, x, y) has multiple


roots in y for x in the discriminant locus = {(D(f ))(x) = 0} C.
We may assume f has positive degree in y , since otherwise V (f ) is just
Proof. The ane polynomial

a vertical line.
Since

is irreducible in

C[x, y]

cal vanishing of

would imply that

reduced. So

D(f )
D(f ) is a

nontrivial polynomial, and

(8.1.5)

#{x C | y

such that

It is easy to argue directly that were

},
6or

then

(x )

would divide

y , the identiV (F ), was non-

of positive degree in

V (f ),

hence

is nite:

f (x, y) = fy (x, y) = 0} <


V (f ) to contain a vertical line{x =

(contradicting irreducibility). So by

you can wait for Study's lemma in the next Chapter

8.2. MONODROMY AND CONNECTEDNESS

101

(8.1.5) and Prop. 2.1.8, in fact

#{p C2 | f (p) = fy (p) = 0} < .


V (f ). The only possible

additional singularities of V (F ) are the (nitely many) points where it


meets the line at .


The set in brackets includes all singularities of

8.2. Monodromy and connectedness


Let

be a region, that is, an open connected subset.

holomorphic function,
when

be a small disk about a point

f O().

on which one is given a

We are interested in the question of

extends to a holomorphic function on all of

doesn't always happen, take

Let

=C

and

To see why this

a small disk about

z = 1:

then f =
only extends to a holomorphic function on C . Even worse,
z
f = log(z) becomes multivalued on C and so (as a holomorphic
function) only extends to

C\R0 .

To give a condition which will ensure the existence of a well-dened


holomorphic extension, we need the concept of analytic continuation.

from p to q to be the image of a continuous


function P : [0, 1] with P(0) = p and P(1) = q . (Here we are
allowed to pick q = p.) An analytic continuation of f along

Dene a path

,f

an
aly

tic

co
n

tin

ua

tio

of

C
i

consists of

i ,fi

102

8. THE CONNECTEDNESS OF ALGEBRAIC CURVES

into segments {i }N
i=0 ,
a covering of by disks i i (with 0 = ), and
functions fi O(i ) (with f0 = f ) satisfying fi fi+1
i i+1 .

a partition of

If we continue

along two dierent paths from

fN O(N )

p to q

on

and compare the

q)
in each case, these need not agree. In the above example of f = log(z)
on a disk about p = {z = 1}, we can analytically continue f along

any path in C . However, if we take q = p so that the path is closed,

then we do not have fN (p) = f(0) (p): they dier by 2 1 times the
winding number of the path about z = 0, hence the multivaluedness

results, i.e. the last function

(in the neighborhood of

referred to above. This problem only occurs, however, for non-simplyconnected regions:
Proposition 8.2.1. [Riemann Monodromy Principle] Given

a region

which is simply connected, i.e.

1 () = {0}.

f O() can be
starting at p . Then

be a small disk, and assume that

continued along any path

f O()

extending

Let

analytically
there exists

f.

We will frequently use this together with the


Proposition 8.2.2. [Heredity principle] Given

f O()

F (x, y) O(C2 ),

with

F (x, f (x)) = 0.

(8.2.1)

Then the analytic continuation of

along any path

must satisfy

(8.2.1).
fi in the analytic continuation are holomorphic, so is each F (x, fi (x)) (on i ). But F (x, f (x)) 0 on =
0 by assumption, and since f = f0 f1 on 0 1 , we have
F (x, f1 (x)) 0 on 0 1 and therefore (by basic complex analysis) on all of 1 . Simply iterate this argument for i = 1, . . . , N .

Proof. Since

and each

Now given an ane algebraic curve


degree

n,

it is convenient to write

polynomial in
(8.2.2)

over

C = {f0 (x0 , y) = 0}

with

f0

of

as the vanishing locus of a monic

C[x]:

f (x, y) = y n + a1 (x)y n1 + + an (x) = 0.

8.2. MONODROMY AND CONNECTEDNESS

This is acheived by performing a change of variable

x0 = x + y

103

and

f (x, y) := f0 (x0 , y) = f0 (x + y, y), which has coecient of


y depending polynomially on ; choose so that this coecient is 1.
n
(The main point is that in f0 (x0 , y), the y term may be zero, and we
writing

want to remedy that.)


Having put the equation of

in this form, we write

C
C
(x, y) 7 x

for the projection of the curve to the

x-axis.

Writing

D := {D(f )(x) =

0} for the discriminant locus of this projection, by (8.1.4) we have that


1
for x C\D , the bre
(x) consists of n distinct points. For some
xed disk C\D , label these points {y1 (x), . . . , yn (x)}. Notice that
R(f, f
) = D(f )(x) 6= 0 implies that f
6= 0 on {f = 0} 1 (), so
y
y
that the holomorphic IFT (Prop. 7.1.2) gives yi (x) O(). The point
here is that the roots of (8.2.2) in y are algebraic  hence multivalued
 functions of x, but we can take well-dened holomorphic branches
of them over . As we shall see, the multivaluedness will intertwine
them outside .
Label the points of D = {p1 , . . . , pK }, and let be the path in
1
P consisting of segments connecting to p1 , p1 to p2 , and so on up
1
to pK . Then the region := (P \) C is simply connected. By
Propositions 8.2.1-2, the {yi (x)} extend to functions in O() which
still satisfy

f (x, yi (x)) = 0.

(8.2.3)

to "i "

= P \
1

= points of D

Analytically continued through

in

C\D,

the

(8.2.3) by the heredity principle, but may swap.

yi

continue to satisfy

104

8. THE CONNECTEDNESS OF ALGEBRAIC CURVES

Example 8.2.3.

through

1
3
3

cyclically
x.

f (x, y) = y 3 x, D = {0}, = R0 . Passing

2 1
3
permutes y1 (x) =
x, y2 (x) = e 3 3 x, y3 (x) =
7

This swapping (or permutation) of the

yi (x) gives rise to an equiv-

alence relation  :

yi (x) yj (x) if one may be analytically continued


into the other in C\D . An equivalence class is just all the {y } which
are equivalent to a given yi in this sense.
Proposition 8.2.4. For any equivalence class

ordering if necessary) by

of

formed (re-

y1 (x), . . . , ym (x),
m
Y

(8.2.4)

(y y (x))

=1

C[x, y].

belongs to

Put dierently:
functions on

C\D,

while the

{y (x)}m
=1

are multivalued algebraic

the elementary symmetric polynomials in them are

not multivalued; in fact, they are polynomials!


Corollary 8.2.5.

irreducible

= C\ 1 (D) is connected ( =

connected).
Proof. [assuming Prop. 8.2.4] If

f C[x, y]

doesn't factor, then

by the Proposition there can be only one equivalence class:

{1, . . . , n}.

So the complete set of branches

{yi (x)}

E =

is acted on tran-

D, and one can therefore draw a path on


two points.


sitively by monodromy about

C\ (D)

connecting any

We now prove Proposition 8.2.4, using some theorems from complex


analysis.

In particular, recall that Rouch's theorem asserts that for

f, g O(R) on a simply connected region9


satisfying |f | > |g| on a path , f + g and f have the same number
of zeroes (counted with multiplicity) inside .

two holomorphic functions

7the

transformations of an algebraic structure arising from its transport around

loops (in this case, loops in

about points of

D)

are what is meant by the word

monodromy in general. So the Riemann monodromy principle is really a statement


about the absence of monodromy.

8we use to index E (i.e. 1, . . . , m) and i to index {1, . . . , n}


9the main point is that R should contain the interior of

8.2. MONODROMY AND CONNECTEDNESS

105

Proof. The product (8.2.4) is clearly well-dened on

monodromy about

simply swaps its factors; hence it is

C\D, since
in O(C\D).

Write

Ym

(8.2.5)

=1

(y y (x)) =

m
X

emj (y1 (x), . . . ym (x))y j

j=0

emj (y1 (x), . . . , ym (x)) =: emj (x) denotes the elementary symmetric polynomials in the {y }. Again, because these are not changed
under monodromy, we have emj (x) O(C\D). Observe that given
D with neighborhood N (a small disk about ), the polynomials
aj (x) from (8.2.2) satisfy

where

x N = |aj (x)| M (j)


for some

M N.

Fixing

x0 N ,

F(y) = y n ,
{yi (x0 )}

so that the

put

aj = aj (x0 )

and

G(y) = y n + a1 y n1 + + an ,

are the roots of

G.

On

= {|y| = M + 1} C,

we have

|F G| = |a1 y n1 + + an | M ((M + 1)n + + 1)


= (M + 1)n 1 < (M + 1)n = |F|.
F and G have the same number
n zeroes (at y = 0!), we nd that

By Rouch,

F=y

has

|yj (x0 )| < M + 1


Consequently the

for all

of zeroes inside

j = 1, . . . , n

ek (x) O(C\D)

and

since

x0 N .

are bounded on

N (C\D) =

N \{}, and so by the Riemann removable singularity theorem extend


across {}. Doing this for each D , we conclude that ek (x) O(C).
j
So the coecients of the y 's in (8.2.5) are entire functions of x. To
prove that they are polynomials in x, we shall have to consider their
behavior about x = . If we work in the local coordinates x
= x1 ,
y = xy about [0 : 1 : 0] in P2 , then the polynomial (8.2.2) dening C
becomes

10

x f
10here
to

1 y
,
x x

 
 
1
1
n1
n
= y + xa1
y
+ + x an
,
x
x
n

we are essentially taking the projective completion of

U1 P2 .

and restricting that

106

8. THE CONNECTEDNESS OF ALGEBRAIC CURVES

with roots
(8.2.6)

 
1
.
yi (
x) = xyi
x

N be a small neighborhood of x = 0 and N


:= N \{
x = 0}. By
n
(8.2.6), the monodromy of the {
yi }i=1 about x = 0 stabilizes the subset {
y }m
y1 (
x), . . . , ym (
x)) = xk ek ( x1 ) are well-dened
=1 , so that the ek (

holomorphic functions on N . Since deg(aj (x)) j , the x


j aj ( x1 ) are
polynomials in x
hence bounded on N . Using Rouch as above, the
m

ek ({
y (
x)}=1 ) are also bounded on N
, and thus extend to holomorphic functions on N .
1
In other words, ek (x) = ek ( ) has a pole at x = of order at most
x

k . Since ek (x) was also holomorphic on C, we have ek K(P1 ). Now


P (x)
K(P1 )
= C(P1 ) and so ek (x) = Q(x) where P, Q are polynomials; since
its only pole is at , Q is a constant. Therefore each ek C[x], and
with (8.2.5) we see that (8.2.4) is a polynomial in C[x, y].

Let

Exercises
(1) Are the real points

11 of a smooth algebraic curve

P2

necessarily

connected?
(2) For what values of

11i.e.

a, b

does

x4 + ax + b

points on the curve which can be written

have a multiple root?

[X0 : X1 : X2 ],

with all

Xi R

CHAPTER 9

Hilbert's nullstellensatz

In something of an algebraic detour, we will now prove Theorem


5.3.1 for ane hypersurfaces. In the general case, we shall also state
(but not prove) a more precise theorem which lays out the correspondence between ane algebraic varieties and ideals in commutative
rings.

(This is part of the foundation for scheme theory, which you

can explore further in the books by R. Hartshorne and E. Kunz.)

9.1. Resultants (bis)

8.1 let D be a
UFD with fraction eld K ; and for f = a0 Y + a1 Y
+ + an and
g = b0 Y m + b1 Y m1 + + bm polynomials in D[Y ], dene R(f, g) :=
det M(f,g) . (In case D is itself a polynomial ring, we will sometimes write
RY (f, g) to make it clear that Y is the variable being eliminated.)
We will need some more results on resultants. As in

R(f, g) = Gf +F g
deg G < deg g , deg F < deg f .
Proposition 9.1.1.

Proof. If

R(f, g) = 0,

n1

for some

F, G D[Y ] with

then we are done by (8.1.1).

Otherwise,

write
(9.1.1)

Y m1 f = a0 Y n+m1 +a1 Y n+m2 +


Y m2 f =
a0 Y n+m2 +

+an Y m1

+an Y m2

.
.
.

f
n1

Y
g
Y n2 g

=
a0 Y n
+

n+m1
n+m2
n1
= b0 Y
+b1 Y
+ +bm Y
=
b0 Y n+m2 +

+bm Y n2

+an

.
.
.

b0 Y m

Working in

K[Y ],

divide by

a0

107

+bm .

m equations, then use elm columns (to the right of

in the rst

ementary row operations to kill the rst

108

9. HILBERT'S NULLSTELLENSATZ

Y j 's

 =) apart from the

along the diagonal.

Then, divide out the

equations by the leading coecients to the right of  =, normalizing


those leading coecients to
row-reduction on

M(f,g)

1.

That is, we have essentially carried out

in the context of the linear system, under the

assumption that its determinant is nonzero.


The new system takes the form

?? = Y n+m1
?? =
Y n+m2

(9.1.2)

.
.
.

..

?? =
where each ?? is a

K -linear

combination of the entries to the left of

 = in (9.1.1). In particular, the last row of (9.1.2) is

G0 f + F0 g = 1
deg G0 m 1, deg F0 n 1. Now
multiply by the product of all the elements of D we divided the equations by, clearing denominators of G0 , F0 to give elements G, F of D[Y ].
But this product is in fact one denition of the determinant of M(f,g) ,
namely R(f,g) . The result follows.

where

G0 , F0 K[Y ]

satisfy

We should mention the formula for the resultant of two polynomials

y,

whose irreducible factors are all linear (or constant) in

although we

will probably neither use nor prove it:

f and g
j (Y yj )

Proposition 9.1.2. If

a0 i (Y xi ), g = b0
Q
n
am
0 b0
i,j (xi yj ).

f =
R(f, g) =

decompose into linear factors


(for

xi , yj D),

then

9.2. Study's lemma


We continue to assume that

n.

Given

at

:

D,

is a UFD with

f D[Y ]

of degree

we have the ring homomorphism given by evaluation

D[Y ]
D
.
G(Y ) 7 G()

Proposition 9.2.1. (i) If

(Y ) | f (Y ).
(ii) f has at most n

f ()(= (f )) = 0,

then

roots in

D.

i.e.

is a root of

f,

9.2. STUDY'S LEMMA

109

Proof. (i) By the division algorithm,

f = q.(Y ) + r

(9.2.1)
where

deg r < deg(Y ) = 1,

i.e.

r D.

Applying

to (9.2.1), we

have

0 = f () = q().0 + r
and thus

r = 0,

so that

(Y r)

divides

f.

(ii) Follows from (i) (and the fact that


have at most

n = deg(f )

D[Y ]

is a UFD) since

can

linear factors.

Now we will specialize to the case

D = C[X];

more generally, the

results of this section will hold with any algebraically closed eld replacing
Let

C and (X1 , . . . , Xn1 ) replacing X .


F D[Y ] = C[X, Y ] = S2 .
V (F ) = C2 ,
of S2 .

Proposition 9.2.2. If

then

F =0

as an element

Proof. Suppose

of

D[Y ], F

C.

Since

vanishes on all of

C2 ,

By Prop. 9.2.1(ii), viewed as an element

has a nite number of roots in

be constants in
that

F 6= 0.

i.e.

C[X].

Some of these may

C such
F (X, )(= (F )) 6= 0 in

is an innite eld, there exists

is not one of these roots, and then

F (X, ) itself has nitely many roots,


so there exists C such that F (, ) 6= 0. Hence, F is not identically
2
zero on C .

C[X].

Again by Prop. 9.2.1(ii),

Proposition 9.2.3. [Study's Lemma] Given

irreducible and

V (f ) V (g).

Then

divides

f, g S2 ,

with

g.
f be irreV (fi ) V (f )

Remark 9.2.4. Suppose we drop the requirement that

ducible, so that

f =

fimi (fi irreducible in

S2 ).

for each i, and by the Proposition fi |g for each


P
f |g mi , i.e. f divides a power of g .

Then

i.

This implies that

f |0 is trivial, we take g 6= 0. By Prop. 9.2.2, we have


2
which implies V (f ) 6= C hence f 6= 0. We may assume

Proof. Since

V (g) =
6 C,
that f
/ C (since a constant divides anything), and furthermore
degY (f ) 6= 0 (otherwise just swap X and Y ). Writing
f = a0 (X)Y n + a1 (X)Y n1 + + an (X)
/ C[X]

that

110

9. HILBERT'S NULLSTELLENSATZ

> 0 and a0 (X) 6= 0),


g
/ C[X].
(n

claim:1 we can assume that

I make the

claim,

f and g are of degree > 0 in Y , so by


Prop. 9.1.1 (with D = C[X]), RY (f, g) = F g + Gf C[X] for
degY F < degY f , degY G < degY g . Given any C\V (a0 ), since
C is algebraically closed there exists a root C of f (, Y ). From
V (f ) V (g) we see that (, ) V (f (, Y )) V (f (, Y )) C, so
that f (, Y ) and g(, Y ) have a common root for every C\V (a0 ).
It follows that a0 RY (f, g) C[X] evaluates to zero at every C,
hence is zero. As a0 6= 0, we nd RY (f, g) = 0 in C[X]; and then
by Prop. 8.1.2, degY (gcdS (f, g)) > 0. (Alternately, F g = (G)f
2
= f, g have a divisor of nonzero degree in Y .) But f is irreAssuming the

ducible, so divides any nonzero non-unit dividing it; we conclude that

f | gcdS2 (f, g) | g .
g C[X]\{0}. Then there exists
C\V (g.a0 ). Viewed as a function on C2 , g is constant in Y , so
g(, ) 6= 0 C. But since a0 () 6= 0, degY (f (, Y )) > 0; and
then (as C is algebraically closed) C such that f (, ) = 0. By
assumption, V (f ) V (g) and so g(, ) = 0, a contradiction.

To prove the

claim, suppose

9.3. The nullstellensatz


The proof of Study immediately generalizes to

Cn .

This yields a

version of Hilbert's nullstellensatz for hypersurfaces:

V (f ) = V (g) for f, g Sn and . . .

(i) f, g are irreducible, then f = g ( C )


N
M
(ii) f, g are not irreducible, then M, N N such that f |g , g|f .
Equivalently, f and g have the same irreducible factors.
Corollary 9.3.1. If

Proof. (i) Study

= f |g

and

g|f ;

(ii) is by Remark 9.2.4.

The point of this is that, modulo issues with powers, there is a

bijection between hypersurfaces and principal ideals (i.e. polynomials


up to multiplication by constants) in

Sn

which reverses inclusion. That

f and g are reduced (all irreducible factors


1), (f ) (g) f |g V (f ) V (g).

is, provided
multplicity

1at

this point, of course, we can't just swap

and

Y

occur with

9.3. THE NULLSTELLENSATZ

111

To get a more general perspective on this, we introduce a few new

X Cn ,

ideas. First, given a subset

we dene the ideal of

by

I(X) := {f Sn | f (z) = 0 z S}.


For example, if
Lemma.

I(V (f )) = (f ) by Study's
it is of the form V (J) for

is reduced, we clearly have

A subset

X C

is algebraic if

J Sn . (Indeed, this is just an ane algebraic variety.)


statement V (I(X)) = X is true (almost a tautology) for algebraic

some ideal
The

subsets.
Given any ideal

J Sn ,

we let rad(J) denote the radical of

which is the ideal comprising all elements of

J.

belongs to
Finally,

Sn

some power of which

A radical ideal is an ideal which equals its own radical.

Sn /J is a domain ( J
ideals in Sn ), and maximal Sn /J

is prime

the monoid of

J,

Theorem 9.3.2. Let

is irreducible in
is a eld.

be an algebraically closed eld.

(i) Every maximal ideal

m Sn

is of the form

(Z1 1 , . . . , Zn

Cn );
(ii) Let J ( Sn be an ideal: then V (J) 6= ;
(iii) For any ideal J Sn , I(V (J)) = rad(J).

n ) = m

(here

Corollary 9.3.3. The correspondence

ideals

subsets

{J Sn }

{X Cn }


V

induces bijections

{radical ideals} {algebraic subsets}


.

{prime ideals} {irred. alg. subsets}


This is clear from the Theorem; to see the last correspondence,
notice that

V (J1 J2 ) = V (J1 ) V (J2 ).

One can push the relation between commutative algebra and ane
algebraic geometry much further.

For example, the ring of regular

functions on an irreducible ane variety

V = V (P) (P

is dened by

C[V ] := Sn /P,

a prime ideal)

112

9. HILBERT'S NULLSTELLENSATZ

and it is easy to see that this embeds (say, for

O(V ).
O(V ) given

smooth) in

Sn to
by restricting polynomial functions to , and so Sn /P is its image.)
C[V ] is sometimes also called the coordinate ring of V . Furthermore,
, the eld of
if V is the ane part of a smooth projective variety V
) is isomorphic to the fraction eld C(V )
meromorphic functions K(V
(or V ).
of C[V ]. Usually C(V ) is called the function eld of V
(The idea is that

is the kernel of the map from

There is even a way to recover varieties from their coordinate rings;


this is the Spec operation. Very roughly speaking, the ane story is
this: any nitely generated commutative domain

may be presented

C[z1 , . . . , zN ]/I (where I C[z1 , . . . , zN ] is an ideal), and then you


N
take V (I) C . This gives one realization of Spec(A); of course,
there are many ways of writing A in this form (dierent N , dierent
I , etc.). From the standpoint of scheme theory, Spec(A) is something

as

intrinsic, an ane scheme which exists in the absence of any particular


embedding in an ane space

CN .

The best resources on this are the

book by E. Kunz and the classic text by R. Hartshorne.

Exercises

X Cn , V (I(X)) = X; (ii)
J1 , J2 C[Z1 , . . . , Zn ], V (J1 J2 ) = V (J1 )

(1) Prove: (i) that for any algebraic subset


that for any two ideals

V (J2 ).

CHAPTER 10

Local analytic factorization of polynomials

Recall the idea of normalization for an irreducible algebraic curve

C P2 :

there should exist a Riemann surface

phically to

P
nonsingular C

with

as its image.

mapping holomor-

In Chapter 7 we did this for

by using the holomorphic implicit function theorem to

put a complex manifold structure on


sisted, for each

p C,

itself.

This essentially con-

in exhibiting a neighborhood

and a (bi)holomorphic parametrization of

U C.

Np C

Np P2

of

by some open set

(The holomorphicity of the transition functions was then a

consequence.)
Now suppose

has an ordinary double point (ODP) at

 recall

that this is a singularity with 2 distinct tangent lines. Denoting disjoint


union by  q, one has

Np C '
that is,

U1 q U2
;
0U1 0U2

U1 , U2 ( C) glued together
C , U1 and U2 must be detached:

locally looks like two disks

one point. In order to normalize

113

at

114

10. LOCAL ANALYTIC FACTORIZATION OF POLYNOMIALS

q2

C
q

q2

q1

(irreducible)

p
p

U2

U1

p
(reducible)

topological picture

schematic picture
Our overarching goal is to produce

and

as in this gure. Geomet-

rically it seems clear that the local analytic curve


even though the global curve

is not.

Np C

is reducible,

The rst step, then, will be

to nd an appropriate formalism (in terms of 2-variable power series)


for working with

Np C ,

which one might call analytic localization.

In this setting, the local equation can be uniquely factored. This will
allow us (in the next Chapter) to carry out local normalization  that
is, put local coordinates on the irreducible components of

Np C .

Fi-

nally, we will patch these parametrizations together with those of open


subsets of

C\sing(C)

to obtain

C .

p. For con2
venience, replace C for the moment by its anization in C . From
9.3, we have the coordinate ring R = C[C], and to any point p C
corresponds a maximal ideal in m R (consisting of polynomials vanishing at p). Inverting all primes not contained in m, or localizing R at
m, replaces polynomial functions by rational functions with poles anywhere but p, which roughly corresponds to replacing C by C minus any
set of points not including p. This is quite dierent from intersecting
C with an analytic ball at p, and will not produce a local factorization
of a globally irreducible C . Instead of rational functions, we need conThere are algebraic approaches to localization of

at

vergent power series. The closest construction in algebra is something

10.1. ANALYTIC LOCALIZATION

called completion (or Henselian localization).

115

If you are curious (we

won't get into this), a good reference is the book by D. Eisenbud.

10.1. Analytic localization

{f (x, y) = 0} C2
passing through p = (0, 0). The dening polynomial f C[x, y] is,
trivially, a convergent power series; so we may consider how f factors
in O2 = C{x, y} (cf. 7.1). In fact, for purposes of examining the intersection of C with a small neighborhood of the origin, we will show that
f may be replaced by an element of C{x}[y] ( O2 ) in a particularly
It will suce to think of

as an ane curve

nice form:

Definition 10.1.1. The subset

W C{x}[y] of Weierstrass

poly-

nomials comprises elements of the form

y d + a1 (x)y d1 + + ad1 (x)y + ad (x)


where each

aj (x) C{x}

(d Z0 )

satises

aj (0) = 0.
Lemma 10.1.2. Let

, > 0
(a)

f O2

with

f / 0

y -axis.

Then

such that:

f 6= 0

on (i){|x|

< , |y| = }

and (ii){x

(b) the number of roots (counted with


with

on the

|y| < ,

is constant in

for

= 0, 0 < |y| < };


multiplicity) of f (x, y) in y

|x| < .

f (0, y) are isolated: otherwise they would


have a limit point, forcing f to be identically zero. We may therefore
choose  so that f (0, y) 6= 0 for 0 < |y| . To get (a)(i) from this, just
use continuity and choose suciently small. The number of roots in
Proof. The zeroes of

1technically,

a submonoid  you can multiply (but not add) elements, and it has

the identity element

1;

the notions of irreducible element and uniqueness of

factorization still have meaning.


10.2.2) and has

Since

is inside a UFD (see proof of Thm.

as its sole unit, it does indeed have unique factorization in a very

strong sense. (See the discussion after the proof of Thm. 10.2.2.)

2in

S is some subset of

/ 0 on S  as  f is
not vanish on S  (i.e. f

general, if

the domain of denition of a function

should read  f

not identically zero on

as  f does

is zero at no point of

S ,

S)

and  f

f,

6= 0

on

one

S

 two very dierent

meanings. Henceforth the symbols will be used with no further explanation.

116

10. LOCAL ANALYTIC FACTORIZATION OF POLYNOMIALS

(b) is computed by

2 1
x

which is continuous in

|y|=

fy (x, y)
dy Z,
f (x, y)


and therefore constant.

Lemma 10.1.3. For

as in Lemma 10.1.2, let

{y (x)}=1,...,d

be the

3
roots described in (b). Denote the elementary symmetric polynomials
in them by

ej (x) (=

1 <<j

y1 (x) yj (x)).

Then

w := y d + e1 (x)y d1 + + ed (x)
is a Weierstrass polynomial.
Proof. Note that for each

Clearly then the

ej (x)

, y (0) = 0

from Lemma 10.1.2(a)(i).

are well-dened and satisfy

ej (0) = 0;

we must

{|x| < }. First we have


X
fy (x, y)
1

yk
dy =
(y (x))k =: k (x),
2 1 |y|= f (x, y)

show that they are holomorphic on

y (x) of the argument is (y (x))k Resy (x) ( ffy ) =


(y (x))k ordy (x) (f (x, )). Here the Newton symmetric polynomials
k (x) span the same vector space over C as the ej (x) (a general fact
from algebra). From the integral expression, the k are evidently holomorphic, and therefore so are the ej .

since the residue at each

Let

U := O2 O2

denote the units, which are just the invertible

convergent power series, or equivalently the convergent power series


with nonzero constant term. (That is, given

g O2 , g U

1
g

O2 .)
f

Lemma 10.1.4. For

such that

uw = f,

Proof. Write

Qd

=1 (y

y (x)),

and

u :=

f
w

uU
|y| }.

as above, there exists a unique

and this holds on all of

V := {|x| <

O(V \{w = 0}).

For xed

and

x, w(x, y) =

ej (x) as coecients.
|x| < ), w(x, y) and f (x, y) have

as mutliplying this out gives the

Consequently, for each xed

3These

(with

{|x| < }  in particular, one should expect


0. So the y (x) are really only well-dened
of the disk {|x| < } (e.g., deleting the positive

may well be multivalued on

them to be permuted as

goes about

on some simply-connected subset


real numbers gives a slit disk).

10.2. UNIQUENESS OF LOCAL FACTORIZATION

y ). Therefore u 6= 0 on V , and u(x, y)


in y . Now, for any given y0 with |y0 | < ,

1
u(x, y)
u(x, y0 ) =
dy .
2 1 |y|= y y0

the same roots (in

x)

holomorphic

117

is (for each

u(x, y) is holomorphic on a neighborhood of |y| = , this formula


shows u
(x, y0 ) is holomorphic in x. By Osgood's lemma, we have u
O(V ). Since u 6= 0, it has nonzero constant term u(0, 0), and is thus a
unit. Uniqueness is clear since u
w = f and uw = f = (
u u)w = f
= u u = 0.


Since

10.2. Uniqueness of local factorization


The uniqueness of

in the last Lemma was trivial. A slightly less

trivial uniqueness question would be: can we write

as a product of a

unit and a Weierstrass polynomial in two dierent ways  i.e., with a


dierent

and

u?

We cannot:

Lemma 10.2.1. Given

decomposition

f = wu

f O2

(with

f / 0

in Lemma 10.1.4 (i.e., into

y -axis), the
w W and u U)

on the

is unique.

u(0, 0) 6= 0, shrinking , (hence V )


if necessary, we have u 6= 0 on V . Thus if f = wu, the zeroes of f and
Q
w are the same. This forces w = (y y (x)) = y d + e1 (x)y d1 + +
ed (x), which makes w (hence u) unique.

Proof. Since any unit

has

Making use of the last two lemmas, we now show that


uniquely (up to units) into irreducibles

fi O2 .

polynomial dening an irreducible algebraic


then the local piece

{fi = 0}.

CV

f O2 factors

f began its life as a


2
curve C = {f = 0} C ,
If

breaks (uniquely) into irreducible components

Provided there is more than one of them, the

fi

are no longer

polynomials, for that would contradict (global) irreducibility of


Theorem 10.2.2.

O2

is a UFD.

Proof. We must demonstrate that

f1 f`

f O2 factors into irreducibles

uniquely up to order and units.

O1 = C{x} is a UFD: given g O1 , we have a


(f )
decomposition g(x) = x 0
h(x), where h is a unit (convergent

First, note that


unique

C.

118

10. LOCAL ANALYTIC FACTORIZATION OF POLYNOMIALS

h(0) 6= 0) and 0 (f ) Z. The irreducibles in this


case are just the factors of x.
By the Gauss lemma, it follows that C{x}[y] is a UFD.
P
a b
Next, suppose that f (x, y) =
a,b ab x y O2 vanishes identically
P
b
on the y -axis; that is, 0 f (0, y) =
b 0b y . It follows that all 0b = 0

for all b, so that f = x f0 where > 0 and f0 (0, y)


/ 0. We must prove
unique factorization for f0 .
Let f O2 with f (0, y) /
0. Lemmas 10.1.4 and 10.2.1 give
f = uw uniquely. Since w belongs to the UFD C{x}[y], we have a
unique decomposition w = h1 h` into irreducibles hj C{x}[y].
Clearly also hj (0, y) /
0, and so Lemma 10.1.4 applied to each hj
gives uniquely hj = uj wj , with each wj a Weierstrass polynomial irreducible in C{x}[y] (since hj is). This yields w = (u1 w1 ) (u` w` ) =
(u1 u` )w1 w` =: uw , and by Lemma 10.2.1 u must be 1. So far
we have f = uw1 w` .
We do not know yet whether wj is irreducible in O2 . If wj =
0 00
v v (v 0 , v 00 O2 ), then wj (0, y)
/ 0 = the same thing for v 0 , v 00 .
0
0 0
00
00 00
Lemma 10.1.4 applies to yield v = u w and v = u w , so that wj =
(u0 u00 )(w0 w00 ); applying Lemma 10.2.1 yet again gives u0 u00 = 1 =
wj = w0 w00 . But w0 , w00 W C{x}[y], contradicting irreducibility of
wj in C{x}[y].
To see uniqueness, write factorizations f = f1 f` = g1 gk into
irreducibles in O2 ; we may assume f (0, y)
/ 0. Then Lemma 10.1.4
gives fj = uj wj and gi = u
i wi with wj , wi irreducible Weierstrass polynomials. We then have (u1 u` )(w1 w` ) = (
u1 uk )(w1 wk ), so
that by Lemma 10.2.1 u1 u` = u
1 uk and w1 w` = w1 wk .
By uniqueness of factorization in C{x}[y] (and Lemma 10.2.1), the
{wj } and {wi } are the same (up to reordering), and ` = k .

power series with

Note the key statement that comes out of this proof: given
with

f (0, y)
/ 0,

we have

f = uw1 w` ,

(10.2.1)
where

uU

and

wi

are Weierstrass polynomials which are irreducible

(as Weierstrass polynomials, as elements of


of

O2 ).

f O2

C{x}[y],

and as elements

Moreover, this decomposition is completely unique, up to re-

ordering of the

wi .

Finally  this also comes out of the proof  if

EXERCISES

119

was a Weierstrass polynomial, then u = 1 in (10.2.1), and degy (f ) =


P`
i=1 degy (wi ). This will be useful in working the following problems.

Exercises

f (x, y) = x3 x2 + y 2 is (a) irreducible in C[x, y]


reducible in C{x}[y].
3
2
Show g(x, y) = x y is irreducible in C{x}[y].

(1) Show

(2)

and (b)

CHAPTER 11

Proof of the normalization theorem

The purpose of this chapter is twofold: to nd a method for explicitly parametrizing neighborhoods of singular points on algebraic curves;
and, using this, to completely prove part (A) of Theorem 3.2.1. In fact,
we shall prove a stronger result which contains a uniqueness statement:
Theorem 11.0.3. Let

with

S =

C P2

be an irreducible algebraic curve,

sing(C) its set of singular points. Then there exists a Rie-

mann surface

and morphism (of complex manifolds)

: C P2

such that

=C
(C)
1
(b) #{ (S)} <
1 (S)) (C\S) =: C is injective (hence an isomorphism).
(c) : (C\
(a)

) is called the normalization of C , and is unique in the


(C,

=
0 , 0 ) is another, then there exists a morphism : C
sense that if (C
C 0 such that = 0 .
The pair

9.3) between ideals I


C[x, y], varieties V = V (I), and rings C[V ] = C[x,y]
, normalization
I
means taking the integral closure of C[V ] in C(V ). Taking Spec of
with a morphism to V . This
the result produces an ane variety V
We remark that in the correspondence (cf.

procedure may be carried out for projective varieties by patching ane


ones together, and if this is done for curves (V
just

= C ),

then

is really

constructed algebraically. While this is beyond the scope of our

course, it's instructive to look at an example.


Example 11.0.4. If we take

V = {x3 x2 + y 2 = 0} C2 ,

then

the coordinate ring

C[x, y]
x2 + y 2 )
is not integrally closed in its fraction eld C(V ).
2 + (x 1) = 0, while irreducible in C[V ][], is
C[V ] =

(x3

121

That is, the equation


solved by

y
, as
x

122

y2
x2

11. PROOF OF THE NORMALIZATION THEOREM

x2 x3
x2

cubic curve

= 1x
V is

C(V ).

in

A schematic picture of the irreducible

(0,0)

and

(1,0)

y
can be viewed as separating the branches of
x

point

at the singular

(0, 0).

In an exercise above you were asked to carry out the (local) analytic
approach, proving that
in

C{x}[y].

x3 x2 +y 2 is irreducible in C[x, y] but reducible

Here is another such example.

y 4 + x3 x2 (= 0), which
C{x}[y], into the product of

Example 11.0.5. Consider the equation

is irreducible in

C[x, y]

but reducible in

Weierstrass polynomials

(y 2 x 1 x)(y 2 + x 1 x).

x 1 x is
ishing at x = 0.

Here

regarded as a convergent power series (in


The local picture (near

(0, 0))

C{x})

van-

described by this fac-

torization is of two parking lots (topologically, these are just disks)


attached at their centers:

g1

g2

We need a procedure that gives the indicated holomorphic parametrizations of these two branches.

11.1. Overview
Informally, here is the main idea of the proof of Theorem 11.0.3.
Given an irreducible algebraic curve

with singular point

p,

we may

11.2. IRREDUCIBLE LOCAL NORMALIZATION

123

P GL(3, C)-transformation (i.e. a projectivity) of P2 to move


p 7 [1 : 0 : 0]. By another linear transformation of coordinates (cf.
8.2), we can put the ane equation in the form
use a

(11.1.1)

f = y n + a1 (x)y n1 + + an (x) (= 0) ,

aj (x) C[x].

Now we want to normalize a neighborhood of the singularity

(0, 0).

D(f )(x) is not idenm`


m1
tically zero in C[x]. Hence the local factorization f = f1 f`
into irreducibles in C{x}[y] will have no repeated factors (all mi = 1).
Writing = {|x| < } and W = {|y| < } for suciently
small ,  > 0, this corresponds to the decomposition of C W into
C1 C` , where each Ci is homeomorphic to a disk and the union

Since

is irreducible in

C[x][y],

its discriminant

attaches them only at their centres.


More precisely, writing

f = uw1 w`
10.2, the Ci

{wi = 0} of irreducible Weierstrass



polynomials. If we can write down 1-to-1 holomorphic maps
i :
is some other disk related to ) with image i ()
= Ci , and
C2 (

repeat this procedure over all singular points, then the normalization C
as in

are the zero-loci

C = C\sing(C), we have a covering


1
by holomorphic parametrizations = z (from 7.2). Composing the
i with the z whenever Ci U in nonempty, yields holomorphic tran
sition functions. Thinking of C as an abstract complex 1-manifold,

these transition functions indicate how to attach each Ci to C to


. To obtain C (topologically) from
yield a new complex 1-manifold C

this, you simply reattach the centers of the Ci .

can be constructed as follows. On

The rst step indicated in his outline, which we do not yet know
how to do, was the construction of the

{i }.

We shall now do this.

11.2. Irreducible local normalization

w = y k + b1 (x)y k1 + + bk (x) be a Weierstrass polynomial,


irreducible in C{x}[y]. Unless k = 1, the discriminant (D(w))(x) has a
zero at x = 0. Since D(w) is not identically zero, this zero is isolated,
and we can take small enough that x = 0 is its only zero on =
{|x| < }.
Let

124

11. PROOF OF THE NORMALIZATION THEOREM

Q
w = k=1 (y y (x)) which is valid in
the sense of 8.2, but not in C{x}[y]. Namely, the {y (x)} are multival1
ued on , but become well-dened on minus a slit. (Another, more
algebraic, way to think of this factorization, if 0 < |x0 | < , is as taking
place in C{xx0 }[y].) The multivaluedness is manifested as follows: by
Now, there is a factorization

the heredity principle, going once counterclockwise around the origin

, permutes the roots of w by y (x) 7 y () (x) where Sk (=the


symmetric group on k elements). This permutation must be transitive,
i.e. a k -cycle: otherwise, it splits into a product of (smaller) cycles,
each of which gives rise to an irreducible proper factor of w in C{x}[y],

in

in contradiction to its irreducibility.


Here, then, is how to parametrize the set
Proposition 11.2.1. Let

pick any

{1, . . . , k}.

w W

{w = 0} W :

be irreducible of degree

k,

and

1
k

:= {t C| |t| < },

Then writing

C2
g:
t 7 (tk , y (tk ))
2 with image the local analytic curve

is well-dened and injective,

C := {(x, y) | w(x, y) = 0, |x| < , |y| < },


and gives a biholomorphism (of complex 1-manifolds)

\{0}
C \{(0, 0)}.
Remark 11.2.2. Here

C \{(0, 0)}

is a complex 1-manifold by the

holomorphic implicit function theorem as in

7.2,

neighborhoods with local holomorphic coordinate

and is covered by

x.

One can regard

the last biholomorphism as giving the transition function between


and (more generally) any open set in

(0 , t)

with holomorphic coordinate

x.
Proof. (of Prop. 11.2.1) Recall that

y (x)

is well-dened on the

:= \{x R0 }. Analytic continuation of y (x) once


counterclockwise around x = 0 yields y () (x); going around once more
k
gives y 2 () (x), and so on. Since is a k -cycle, () = and going
k
around zero k times returns us to y (x). But t does precisely this when

slit disk

1except at 0, since all y (0) = 0

2The meaning of  y (tk ) will be

dened in proof.

11.3. FINISHING LOCAL NORMALIZATION

goes around

function on

once, and so

y (tk )

125

extends to a well-dened analytic

= k1
(j) into pie-slices

j=0
(j) := {0 < |t| < and j arg(t) j+1 }. On the interior of each

2
2
j
j+1
slice (that is, where
< arg(t) < 2 ), we can dene a holomorphic
2
k
k
function by y j () (t ), since t 7 t maps this interior (isomorphically)

to
where y j () (x) is dened. Extending these functions continu (j) , they patch together (in fact, analytically continue into
ously to
. This
one another) to yield a single holomorphic function y
(tk ) on
by the removable
is bounded exactly as in 8.2, and so extends to O()
A bit more carefully, we subdivide
1
k

singularity theorem.
Let

k := e

. If

(tk1 , y (tk )) = (tk2 , y (tk2 ))

then

t2 = (k )` t1
for some

` Z,

and

y ` () (tk1 ) = y (tk1 ).
{y } are all distinct away from 0, the last equation is impos`
sible unless k|`, which implies (k ) = 1 so that t1 = t2 . This proves
that g is injective.

Since is transitive, g maps surjectively onto C . It gives, Since the

nally, a holomorphic map of Riemann surfaces on the complement of

) and x (for open


0 since in local coordinates t (on
C \{(0, 0)}) we have x = g(t)'s x-coordinate = tk .

subsets covering

11.3. Finishing local normalization


Referring back to

8.1,

for each of the irreducible factors

wj

we now apply Proposition 11.2.1. This yields normalizations

j C
gj :
j
of the irreducible components of the local analytic curve

C W :

of

126

11. PROOF OF THE NORMALIZATION THEOREM

g3

C3

C2

C1

|x|<
Each restriction

=

gj :
(Cj \{(0, 0)}) , C
j
is biholomorphic with respect to local coordinates on

Cj \{(0, 0)}.

covering of

In fact, it takes the form

j and an open

t 7 tk (= x) as

indicated at the end of the last proof. These may be regarded as the
glueing maps that will attach each
holes in

to

thereby plugging the

, which is what we do next.

Before that, we just note that one should carry out the construction
of
of

gj 's
C.

p = (0, 0),

as we have done near

at all the other singular points

11.4. Global normalization (patching)


Suppose for the moment
that

C = C\{(0, 0)}.

(0, 0)

is the only singular point of

Then we put

1
2 ` ,
C := C

g1

where

1
C
g1

g2

g`

means

1
C q
,
1 )
g1 (p) p (p
1
2
C

g1

g2

means

1 q
2
C q
,
1 , q
2 )
g1 (p) p , g2 (q) q (p
and so forth.

C,

so

11.5. EXAMPLES OF LOCAL NORMALIZATION

127

If there are more singularities, then repeat this patching at each

S = sing(C).
P2 with image C , set
get a map : C

c , for c C
(c) :=
.
gj (c), for c
j

point in
To

These two prescriptions are compatible with the patching.


To see that

is compact: given an open cover

{U }

of

C ,

pick

C 0

U(q) containing each q (S). The complement


of these
is isomorphic to a closed subset of C , since is bijective away
in C
1
from
(S). Now a closed subset of C is a closed subset in P2 , P2
0
is compact, and a closed subset of a compact set is compact. So C
0 } has a nite subcover {Ui C 0 }. The {Ui }
is compact and {U C
.
together with the {U(q) } then furnish a nite subcover of C
one

We have now proved all but the uniqueness part of 11.0.3 and it is
time to backtrack and get explicit.

11.5. Examples of local normalization


Example 11.5.1. Assuming

gcd(k, a) = 1,

y k xa
is irreducible in

11.2.1. The

y1 (x) =

C{x}[y],

and we shall apply the procedure of Prop.

(multivalued) roots of

y k xa = 0

If we plug

are

xa , y2 (x) = k xa , . . . , yk (x) = (k )k1 k xa ;

they are well dened on the slit disk

in

into

y1 (x)

{0 < |x| < , arg(x) (0, 2)}.

and analytically continue, we get

y1 (tk ) = ta .
Hence by denition

g(t) = (tk , ta ).
We should check that the image of
the statement that

(tk )a = (ta )k .

lies in

y k xa = 0:

this is just

128

11. PROOF OF THE NORMALIZATION THEOREM

Example 11.5.2. Here is a more complicated example where there

is more than one

gj

(as in

5.3):

f = y 8 + y 4 x6 + x3 x 2 y 4 + x5 x2 .
Viewed in

C{x}[y], this is not a Weierstrass polynomial (the coecient


1 x of y is not zero at x = 0), so we should expect a nontrivial unit
u in (10.2.1). Indeed,
2

f = (y 4 x3 + 1)(y 4 + x3 x2 )

= (y 4 x3 + 1)(y 2 x 1 x)(y 2 + x 1 x),


|
{z
}|
{z
}|
{z
}
u

where

w1

w2

u(0, 0) 6= 0.

is a unit because

w1 , w2 are irreducible Weierstrass polynomials and so we apply


11.2.1 (with k = 2) to normalize their zero-sets.
p
Beginning with w1 , the roots are y11 (x) =
x 1 x and y12 (x) =
p
x 1 x, which are swapped as x goes around 0. So y11 (t2 ) is ob2
tained by substituting t for x and analytically continuing: informally,
p

t2 1 t2 = t 4 1 t2 . So

4
g1 (t) = (t2 , t 1 t2 ).
p
p
For w2 , the roots are y21 (x) = i
x 1 x and y22 (x) = i x 1 x;
Now

and this yields

4
g2 (t) = (t2 , it 1 t2 ).

w2 = 0: one need
p

4
2
2
x(t) 1 x(t) = (it 1 t ) + t2 1 t2 = 0.
Let's check this parametrizes

only write

(y(t))2 +

11.6. Uniqueness
Begin with two normalizations:

CO
?

C \ 1 (S)

PO 2 o
?

C \S

CO 0
?

C 0 \ ( 0 )1 (S)

6
S U
X Z ] _ a d f i k

with

, 0

holomorphic maps of complex manifolds. Essentially what

we have to show is that neighborhoods of the points of


and

1 (S)

(in

C )

( 0 )1 (S) (in C 0 ) are isomorphic in a way which is compatible with

11.6. UNIQUENESS

and

0.

129

Put together with the bottom dotted arrow these isomor-

phisms will yield the desired map

: C C 0

of Riemann surfaces

making the diagram

C ?
?

??
??
??


/ 0
C
0 ~~~
~~
~ ~

C
U P2 be a small
open set containing it. For simplicity assume p = [1 : 0 : 0] and choose
coordinates so that U {|x| < , |y| < } and C U is given by the

vanishing set of a Weierstrass polynomial. Write U := U \{p}.


1
Now pick q
(p); by continuity of , 1 (U ) is open in C . So
there exists an open set V , which we may assume to be connected, with
q V 1 (U ). Since (V \{q}) C U must then be connected,

and C U is homeomorphic to a disjoint union of punctured disks,


maps V \{q} into one of these punctured disks. Consequently, V is
3
mapped into only one (local) irreducible component W of C U . This

To start, let

pS C

be a singular point and

yields the following diagram:

T
d b a _ ] \ Z X
f
+
i g

=
/ W o
{|t|
V OO
g
O O
mm
OX
mmm
m
O O
m

mm
O O
mmm t7tk

'

vmmm
 prx
/ ({|x| < } Cx )
C2

(x,y)

in which

prx

composition
on

< k } Ct

and

are morphisms of complex manifolds, so that their

X is evidently a holomorphic (obviously bounded) function

V.
The composition

tion on

V.

is also evidently a bounded, well-dened func-

By the holomorphic IFT (and holomorphicity of

holomorphic on

V \{q};

X),

it is

hence by the removable singularity theorem,

T O(V ). It is also clear that T(0) = 0. So by the open mapping


theorem, T maps V onto a neighborhood N of 0 in Ct (which we may
assume is a disk). Shrinking U (and thus W ) if necessary, we may
conclude that
 the restriction of to a neighborhood of q  maps
V onto W . From the diagram, this
is just g T.
3remember

that these components are homeomorphic to disks; take out

is where the punctured disks came from

and that

130

11. PROOF OF THE NORMALIZATION THEOREM

is 1-to-1 o 1 (S), no neighborhood of any other point q0


1 (S) can be sent to W . Repeating the argument above by varying
q , sets up a 1-to-1 correspondence between  V 's (i.e. neighborhoods
1
of points in
(S) in C ) and  W 's (irreducible local components of
C at points of S ). We can play the same game for the normalization
C 0 , and nd that for a unique q 0 ( 0 )1 (S) we have a neighborhood
V 0 and an isomorphism T0 : V 0 N whose composition with g gives

0 : V 0 W .
0 0
The piece of carrying (V, q) to (V , q ) is now dened simply by
(T0 )1 T. This is automatically holomorphic, and its composition with
0 is g T = as desired.
Since

Exercises

f (x, y) = y 4 (x + 1)7
4
6
7
of g(x, y) = y x + x

(1) Locally normalize the zero-set of


(2) Locally normalize the zero-set

(1, 0).
at (0, 0).

at

CHAPTER 12

Intersections of curves

Now we come to the applications of normalization, which will occupy this chapter and Chapter 14. You may recall that in Chapter 2
we studied intersections of an plane algebraic curve

with a (projec-

L C were each assigned a multiplicity by


restricting the equation of C under a parametrization of L, and looking
tive) line

L.

The points of

at the multplicities of the roots of the resulting one-variable polynomial. With this denition, the multiplicities added up to the degree of
the curve (cf. Prop. 2.1.8).
If we had tried to replace

by an arbitrary curve

at that point,

we would have run into the problem of no longer knowing how to locally
parametrize

near the intersection points. Now that we can do this

(Prop. 11.2.1), we can pull the dening equation of

back under the

parametrization and look at its order of vanishing at the intersection


point. This leads to the general denition of intersection multiplicity,
and with this in hand that we can nally state (and prove!) Bezout's
theorem in general. In its proof the intersection divisor will make an
appearance, so we begin with a short bit on divisors.

12.1. Divisors on a Riemann surface


Let

be a Riemann surface. The group of divisors on

free abelian group on points of

(
Div(M )

:=

X
nite

is the

M,


)


mi [pi ] mi Z , pi M .

The uncountably many symbols

[pi ]

are the generators of this (very

big) abelian group. Associated to a divisor

D=

a degree

deg(D) :=
131

mi .

mi [pi ] Div(M )

is

132

12. INTERSECTIONS OF CURVES

The resulting group homomorphism


Div(M )

(12.1.1)

deg

Z.

is called the degree map.


The divisor of a (nontrivial) meromorphic function

(f ) :=

p (f ).[p]

is given by

Div(M ),

pM
where

p (f )

is the order of

at

(Defn. 3.1.4). Note that the sum is

actually nite (as required by the denition of divisor) since at all but

M , p (f ) = 0. Now K(M ) is a multiplicative


Sending f (f ) yields a homomorphism

nitely many points of


abelian group.

()

K(M ) Div(M )

(12.1.2)

of abelian groups, as you will show in an exercise below, which takes


multplication to addition:

(f g) = (f ) + (g), (f 1 ) = (f ).

With these denitions, the composition of (12.1.2) with (12.1.1)


takes

0.

to

pM

p (f ), which by Exercise 3.2 is zero.

That is,

deg () =

Note that one can dene meromorphic functions and divisors more

generally on complex 1-manifolds, but it is only in the compact case


(Riemann surfaces) that the divisors of meromorphic functions are always of degree

0.

Example 12.1.1. On

is

the easiest meromorphic function around

Z1
Z0

K(P1 ) . Writing simply 0, for the points [1 : 0], [0 : 1],


divisor is (z) = [0] [], obviously of degree 0.

z=

its

P1 ,

12.2. Intersection multiplicities


For a polynomial in one variable

f (x)

with

f (0) = 0, deg(f )

is

the exponent of the highest degree term, while the order of vanishing
ord0 (f )

:= 0 (f )

is the exponent of the term of lowest degree. Order

(unlike degree) also makes sense for power series in 1 variable.


How does all this generalize to two variables? First, a polynomial

F (x, y) can be written as a sum of homogeneous terms. If this is


F = Fk +Fk+1 + +Fd1 +Fd , then deg(F ) := d (highest homogeneous
degree) while ord(0,0) (F ) := k (lowest homogeneous degree). From 6.4,
k is also the order of singularity of the curve C = {F = 0} at (0, 0), i.e.
the number of tangent lines to C counted with multiplicity. When we

12.2. INTERSECTION MULTIPLICITIES

133

don't want to refer to the polynomial, we will write ord(0,0) C ; remember


this is

when

is smooth at

crossing) there, and so on.

(0, 0), 2

when

has an ODP (normal

Finally, ord(0,0) also makes sense for

2-

variable power series.


Now suppose

V = {f (x, y) = 0}, W = {h(x, y) = 0}

are reduced

ane algebraic curves that intersect properly  i.e. have no common

V W
p V W,

irreducible components. Then


is itself reduced. For

has no repeated components, so

(f h))(p) = fx (p)h(p) + hx (p)f (p) = fx (p).0 + hx (p).0 = 0


x

and similarly

( y
(f h))(p) = 0.

Therefore

V W

sing(V

W ),

and

Prop. 8.1.6 yields


(12.2.1)

#{V W } #{sing(V W )} < .


V

Definition 12.2.1. Assume

tinct), and let

p V W.

Let

Writing the local decomposition of

W are irreducible (and disU C2 be a neighborhood of p.


V into irreducibles (uniquely)
and

V U = V1 + + Vk ,
with local normalizations (again, essentially unique)

gi : Vi ( C2 )
ti 7 (xi (t), yi (t)),
we dene the (local) intersection multiplicity at

(V W )p :=

k
X

ord0 (h(gi (t))).

i=1
The (global) intersection number is then dened by

(V W ) :=

(V W )p ,

pV W
in which the sum is nite by (12.2.1).

Remark 12.2.2. (a) If either

or

is smooth, the intersection

number is actually the degree of a divisor,

V W :=

X
pV W

(V W )p [p].

134

12. INTERSECTIONS OF CURVES

This is because we can regard the smooth one (say,


surface and then

V W

V W

Div(W ).

as a formal sum of points of

W)

as a Riemann

Alternatively, you can think of

, known as a zero-cycle

1 on

P2 .

The degree is dened in the same way as for divisors.

h gi appearing in Defn. 12.2.1 will frequently

be written gi (h)  that is, we are pulling the function h back by the

local normalization gi .
(b) The composition

The local intersection multiplicities are well-dened essentially by


the uniqueness of local normalizations. They also have some reasonable
properties:
Proposition 12.2.3.

(V W )p = (W V )p .

Proposition 12.2.4.

(V W )p

precisely when none of

at

V 's

ordp (V

tangents at

) ordp (W ),

with equality

coincide with the tangents of

p.

We will postpone proof of these results in

12.4 5, since the proofs

get a bit technical.


Example 12.2.5. Here are two pictures of smooth curves meeting

at a point

p:

distinct tangents

(I)

(II)

ordp W = 1 because the curves are smooth. But


in the rst case, (V W ) = 2, while in the second (which has distinct
tangents) (V W ) = 1.
In each case, ordp V

1zero

refers to the fact that we are taking a formal sum of zero-dimensional sub-

varieties (i.e. points) in

P2

12.2. INTERSECTION MULTIPLICITIES

Example 12.2.6. Let

a, b, m, n N

135

with

gcd(n, a) = gcd(m, b) = 1.
Then by Prop. 12.2.4, we should have


{y n = xa } {y m = xb } (0,0) min(n, a) min(m, b).
Let's check this by actually computing the left-hand side. The normalg
n
a
n a
ization of {y = x } is just t 7 (t , t ) by Example 11.5.1. Writing

h = y m xb ,

we have

g (h) = g (y m xb ) = (ta )m (tn )b = tam tbn


and the order of this at

(0, 0)

is the least of

{y n = xa } {y m = xb }


(0,0)

am

and

bn:

= min(am, bn).

This clearly satises the inequality, and it is easy to cook up an example


where equality doesn't hold:
becomes

with

n = 3 , a = 4, m = 2, b = 9

it

8 6.

P
(V W )p to the more general setting where V = mj Vj
P
W = nk Wk with {Vj } and {Wk } irreducibles, we simply put
X
(V W )p :=
mj nk (Vj Wk ).

To extend
and

j,k
Remark 12.2.7. Here are two other approaches to local intersection

multiplicity which give the same numbers.


(a) The commutative algebra approach makes use of localization.

C(x, y) denotes the fraction eld of C[x, y]. Let p = (a, b)


C . The local ring at p, denoted Op , is the subset of C(x, y) consisting
G1
of rational functions
(here G1 , G2 C[x, y]) with G2 (p) 6= 0. You
G2
can easily check that this is a ring, and it obviously contains C[x, y].
It has a unique maximal ideal mp consisting of functions which vanish
at p.
Now let V = {f = 0}, W = {h = 0} be as above, and assume
p V W . Writing (f, h)p for the ideal in Op generated by f and h,
Recall that

we dene

(V W )p := dimC (Op /(f, h)p )


2for

instance, the polynomial

y n xa

has order given by the smallest of

and

a.

136

12. INTERSECTIONS OF CURVES

by viewing the quotient

vector space. (Note that from

this denition,

Op /(f, h)p as a
invariance of (V W )p

under projectivities is imme-

diately clear.)

As a simple example, we know that the intersection

multiplicity at

p = (0, 0)

of

{x = 0}

{y 2 x = 0} should be 2.
basis 1, y . See Chapter 4 of [L.

and

The quotient vector space, indeed, has

Flatto, Poncelet's Theorem] for more on this approach.


(b) For an approach via resultants, it is convenient to work with

= {H = 0}, P =
V = {F = 0},W
(in homogeneous coordinates [Z : X : Y ] on
[P0 : P1 : P2 ] V W
P2 ). Assume that [0 : 0 : 1] neither belongs to (i) C D, nor (ii) any
line joining points of C D , nor (iii) any line tangent to C or D at a
point of C D . Then we may dene
homogeneous polynomials. Write

)P := ord[P :P ] (RY (F, H)).


(V W
0 1
C[Z, X][Y ] and RY (F, H),
Z and X ; it is in fact homogeneous and of degree deg(F ) deg(H). Its order at [P0 : P1 ] is just the
highest power of (P0 X P1 Z) dividing it.

Here we are thinking of

F, H

as elements of

3
which eliminates Y , is a polynomial in

Justifying this denition takes a bit of work, but it leads immediately to a proof of Bezout since the intersection multiplicities have to
add up to

deg RY (F, H) = deg V deg W

by construction. This is the

point of view taken in [F. Kirwan, Complex Algebraic Curves].

12.3. Bezout's theorem


We rst do a quick recap of Prop. 2.1.8:
Proposition 12.3.1. Let

curve,

L (
= P1 ) P2

a line

C = {F (Z, X, Y ) = 0} P2 be a degree
not contained in C . Then (L C) = d.

Proof. By a change of coordinates, we may assume

and

[0 : 1 : 0]
/ C.

L = {Y = 0}

Then by the Fundamental Theorem of Algebra,

k
Y
F (Z, X, 0) =
(X i Z)di ,
i=1

Pk

di = d since F
C L = {[1 : i : 0]}ki=1 .

where

3as

i=1

has is homogeneous of degree

d.

Hence

usual you can think about this resultant in terms of a projection onto the

(or rather,

[Z : X]-)

axis.

x-

12.3. BEZOUT'S THEOREM

137

Passing to ane coordiantes (f =


(x i )di ) and locally normalgi
izing L at (i , 0) by t 7 i + t, we have

We conclude that

(L C)(i ,0) := ord0 (gi f ) = di .


P
(L C) = di = d.

C, E P2 be properly
(C E) = deg C deg E .

Theorem 12.3.2. [E. Bezout, 1779] Let

intersecting projective algebraic curves. Then

k = deg E , and choose lines


L1 , . . . , Lk avoiding the points of C E . Write E = {H(Z, X, Y ) = 0},
Lj = {j (Z, X, Y ) = 0}. Then by Propositions 12.2.3 and 12.3.1,
Proof. Assume

is irreducible. Let

(C Lj ) = (Lj C) = deg C,
and

(C

(kj=1 Lj ))

k
X
=
(C Lj ) = deg C deg E.
j=1

Now by Example 7.3.5, the quotient of two homogeneous polynomials of the same degree gives a meromorphic function on projective
space.

is of degree

and each

:=

is of degree 1, so we may dene

H
K(P2 ).
1 k

= C ) for the normalization, we have


: C P2 (with (C)

We can compute the divisor of this


Example 7.3.6 K(C).

Writing
by

meromorphic function if we notice that locally about each point of

CE
[resp.

[resp.

Lj ].

C (Lj )],

[resp.

1
] gives a dening equation for

So by Defn. 12.2.1,

( ) =

pCE

q ( )[q]

qC(Lj )

ordp (

)[p]

pCE

p ( )[p] +

ordq (

qC(Lj )

(C E)p [p]

pCE

1
)[q]

(C (Lj ))q [q].

qC(Lj )

But as divisors of meromorphic functions on Riemann surfaces have


degree

0,
0 = deg(( )) =

X
pCE

(C E)p

(C (Lj ))q

qC(Lj )

138

12. INTERSECTIONS OF CURVES

= (C E) deg C deg E.
C

is reducible, break it into irreducible components and

sum the results!

Finally, if

Remark 12.3.3. In terms of zero-cycles (cf.

Bzout is saying that

C E

has degree

Remark 12.2.2(a)),

deg C deg E .

12.4. Proof of Prop. 12.2.3


We now show the symmetry of intersection numbers. Write

V =

{f = 0}, W = {h = 0}, p V W . For simplicity assume that


p = (0, 0), V and W are irreducible, and the dening (polynomial)
equations are in the form

f = y m + B1 (x)y m1 + + Bm (x) ,

h = y n + b1 (x)y n1 + + bn (x).

We decompose these according to (10.2.1): viz.,

f = u1 v1 vr ,

h = u2 w1 ws

vj , wk are irreducible Weierstrass polynomials. For the roots


(j)
of vj [resp. wk ] on a slit disk {|x| < , x
/ R>0 } we shall write y (x)
(k)
( = 1, . . . , mj ) [resp. z (x) ( = 1, . . . , nk )]. On the non-slit x-disk
where the

these become multivalued, and we will assume that counterclockwise


analytic continuation sends y 7 y+1 to keep the numbering simple.
(j)
(k)
As in 11.2 the y
(tmj ) [resp. z (tnk )] are well-dened on a small

t-disk {|t| < 0 },

and we have

(j)

0 mj
y+0 (tmj ) = y(j) ((m
t) )
j

mth
j root of unity mj . (This changes the branch you
(j)
m
start at when arg(t) = 0.) Write gj (t) := (t j , y
(tmj )) and Gk (t) :=
(k)
(tnk , z (tnk )) for the parametrizations of {vj = 0} and {wk = 0}.
for some primitive

We then have the key identity

(12.4.1)

mj
Y



(j)
wk tnk mj , y+0 (tnk mj )

0 =1

4Warning:

you cannot write

(j)
of y
, since this assumes
composition with the

0 mj
(m
t) = (mj )0 mj tmj = tmj
j

inside the argument

(j)
y is well-dened on an entire disk (whereas only its

mth
j -power

map is!).

12.5. PROOF OF PROP. 12.2.4

139

mj nk n
o
Y
Y
(j)
(k)
y+0 (tnk mj ) z+0 (tnk mj )
=
0 =1 0 =1
nk
Y

(12.4.2)



(k)
vj tnk mj , z+0 (tnk mj ) ,

0 =1
which uses the factorization of each Weierstrass polynomial into a product (for each xed
of a disk by

x)

2/mj

of linear factors. Bearing in mind that rotation

does not change the order of a function at

0,

we

compute

ord0 ((12.4.1))

= nk

mj
X

ord0


(j)
wk (tmj , y+0 (tmj ))

0 =1


= nk mj ord0 wk (tmj , y(j) (tmj ))

Dividing this by

mj nk

= nk mj ord0 (gj wk ).
Pr Ps

and applying

k=1 gives

j=1

!
X

ord0

gj

wk

ord0 (gj h)

= (V W )p .

Similarly
ord0 ((12.4.2))
and dividing out

mj nk

= nk mj ord0 (Gk vj ),
(W V )p .

and summing yields

Q.E.D.

12.5. Proof of prop. 12.2.4


With the same notation as in the last section, we also write out the
irreducible Weierstrass polynomials

(j)

(j)

vj = y mj + amj 1 (x)y mj 1 + + a0 (x).

have

(j)

a0 (x) is the product of the multivalued


P
ord(0,0) vj mj ,
j ord(0,0) vj = ord(0,0) f , and

Note that

ord(0,0) (vj (x, y))

(j)

ord0 (a0 (x)) =

1
=
ord0
mj

mj
Y

roots

(j)

y (x).

1
(j) m
ord0 (a0 (t j ))
mj
!

(j)

y+0 (tmj )

0 =1

= ord0 (
y(j) (tmj )).

We

140

12. INTERSECTIONS OF CURVES

Therefore

(V W )p =

r
X

ord0


h(tmj , y(j) (tmj ))

j=1

r
X

(ord(0,0) h) min

m
ord0 (t j ), ord0 (
y(j) (tmj ))

j=1

(ord(0,0) h)

r
X

min

m
ord0 (t j ), ord(0,0) (vj (x, y))

j=1

= ord(0,0) h

r
X

ord(0,0) vj

j=1

= ord(0,0) h ord(0,0) f
= ordp V ordp W ,
Q.E.D.

Exercises

M be a Riemann surface. Show that the divisor map () :


K(M ) Div(M ) is a homomorphism of (abelian) groups. [Hint:

(1) Let

use local coordinates.]

V = {yx =
0} and W = {y x = 0}. (This will depend on C.)
2
Let C P be an algebraic curve of degree n > 1 and L a (pron
+ 1 singular points of C . (Note: bc
jective) line containing
2

(2) Compute the intersection multiplicity

(3)

(V W )(0,0)

for

is the greatest integer function, which takes the greatest integer


less than a given real number.)

Use Bezout's theorem to prove

C L hence cannot be irreducible. [Hint: prove rst that the


intersection multiplicity of L and C at each singular point through
which L passes, is at least 2.]
2
Let C P be an algebraic curve of degree 4 with 4 singular points.
Using Bezout's theorem and Prop. 12.2.4, prove that C cannot be
irreducible. [Hint: use the Hint from (3) together with a conic Q
through the following 5 points: the 4 singularities of C plus one
more point of C .]
2
A degree d algebraic curve C P can be taken to go through
(d+1)(d+2)
1 distinct points. (This is just because dim(S3d ) =
any
2
that

(4)

(5)

(d+1)(d+2)
.)
2

Prove that if all of these points are taken to lie in a

single curve

of degree

e<

d
2

+ 1,

then

is reducible.

EXERCISES

141

(V W )(0,0) for V = {y 2 x3 =

0} and (a) W = {x3 x2 +y 2 = 0} or (b) W = {(y x)3 4 2xy =


0}.

(6) Compute the intersection multiplicity

CHAPTER 13

Meromorphic 1-forms on a Riemann surface

In the next chapter we will see one more application of the normalization business, via intersection numbers: the degree-genus formula.
As more will be needed for its proof, presently we make a detour to
dene and study dierential forms (with poles) on manifolds  how to
patch them together via local coordinates, how to pull them back under
a morphism, and so forth. Like meromorphic functions, 1-forms have
an associated divisor. In contrast to the function case, the degree of
this divisor is not zero: it tells you the genus of the Riemann surface,
via the so-called Poincar-Hopf theorem.

This result will be key to

proving the Riemann-Hurwitz and genus formulae.

13.1. Dierential 1-forms


These are the expressions you integrate over paths in calculus and
complex analysis. For example, on

R2

= F (x, y)dx + G(x, y)dy


is a

1-form.

Given a dierentiable map

: R2 R2
given by

(u, v) 7 (x(u, v), y(u, v)),


the pullback of

by

is

(13.1.1)

:=
=

F (x(u, v), y(u, v)) d(x(u, v)) + G(x(u, v), y(u, v))d(y(u, v))


y
x
(u, v) + G(x(u, v), y(u, v)) u
(u, v) du
F (x(u, v), y(u, v)) u


+ F (x(u, v), y(u, v)) x
(u, v) + G(x(u, v), y(u, v)) y
(u, v) dv.
v
v

A  0-form is just a function

f (x, y),

and

f := f = f (x(u, v), y(u, v))


143

144

13. MEROMORPHIC 1-FORMS ON A RIEMANN SURFACE

. (13.1.1) is simply the analogue


with  . This is exactly what you are

is nothing but precomposing with


for

1-forms

of precomposition

doing when you change variables in an integral.


We want to generalize
to complex

1-manifolds),

1-forms from R2 to real 2-manifolds (and then


which seems to call for a bit of motivation.

2-manifold, f : M R a dieren3
tiable function, and p M a point. If M R , then the notion of
taking partial derivatives of f at p in directions tangent to M  makes
immediate sense  you just precompose f with a (dierentiable) path
in M having a given tangent at p, and dierentiate with respect to the
Let

be a dierentiable real

variable parametrizing this path.


In abstract dierential topology, one has no embedding in
the dierentiability of
tions

R3 .

Rather,

is arranged by requiring the transition func-

relative to local coordinates on an open cover, to be smooth:

(x ,y )

(x ,y )

(This was discussed at the beginning of

2.2.)

One then denes the

tangent spaces

Tp M := vector

space of linear dierential operators (at

=R


+


,
x p y p

and tangent bundle

T M := pM Tp M.

p)

13.1. DIFFERENTIAL 1-FORMS

145

1 (p) = Tp M . A
1
global section of T M , that is, is a smooth map : M T M with
= idM , is called a vector eld on M . (Typically one writes ~v ,
with the understanding that ~
v (p) Tp M .)

One has a projection map

: TM M

with

Now integration is dual to dierentiation, so dierentials are dual

, a dual basis (for the dual vector space)


x y

to tangent vectors. For


is

dx , dy :

we write


 

dx
= 1 , dy
=0,
x
x
 
 

dx
= 0 , dy
= 1.
y
y
The co tangent spaces are then

D
E
Tp M
= R dx |p , dy |p .
Global sections of the cotangent bundle
the dierential

1-forms

on

M.

T M = pM Tp M

are then

In local coordinates a dierential

1-form

looks like:

= F (x , y )dx + G (x , y )dy .

(13.1.2)

M given locally by {g : V R} must







= g |V = g |V ,

Just as a function on

g |V

{ }

the

satisfy

are subject to compatibility conditions



|V = |V .

Now since
and

(hence each

is smooth, smoothness of

(i.e. of

in (13.1.2)) is preserved under pullback, and it makes sense to

dene

A1R (M ) :=
=

1-forms on M
{F , G } innitely dierentiable.

smooth, real-valued
collections

For a complex

{ }

with

1-manifold,

of smooth real 2-manifold (the

which we recall from is a special kind

are conformal), the labels on the

diagram change:

1to

dene smoothness one has to put a manifold structure on

do here.

TM,

which I won't

146

13. MEROMORPHIC 1-FORMS ON A RIEMANN SURFACE

C
V

Omitting subscript
indicate

R C ,

's

for the moment, and writing a subscript

to

one has

*

+
+


,
,
TC,p M = C
=C
x p y p
z p z p
D
E
D
E

TC,p
M C = C dx|p , dy|p
z |p ,
= C dz|p , d





where
,
, dz := dx +
:=

1
:=
+
1
z
2 x
y
z
2 x
y

1dy , d
z := dx 1dy . (This makes dz( z ) = 1, dz( z ) = 0,

) = 0, d
z ( z ) = 1 so that the bases are dual.) A smooth section
d
z ( z

of the complexied cotangent bundle TC M thus looks locally like


*

F (x, y)dz + G(x, y)d


z

= (F + G)dx + 1(F G)dy,


for

and

G smooth (innitely dierentiable) complex-valued functions.

The 1-forms we are after are substantially more restricted:


Definition 13.1.1. A holomorphic [resp.

1 (M )

[resp.

= f (z )dz ,

K1 (M )]2
with

meromorphic]

1-form

is a collection of expressions

f : V C holomorphic

[resp. meromorphic],

satisfying
(13.1.3)

2recall

the notation



|V = |V

, .

K(M )

for meromorphic functions; this is short for

one can think of such functions as meromorphic 0-forms.

K0 (M ),

as

13.1. DIFFERENTIAL 1-FORMS

147

Explicitly, (13.1.3) says that

f (z )dz = f ( (z ))d( (z ))
= f ( (z ))0 (z )dz ,
and is thus equivalent to

f (z ) = f ( (z ))0 (z ).

(13.1.4)

1 , 2 K1 (M ), we can consider their quotient as a mero


morphic function 1 K(M ). This is because in local coordinates, one
2
f (z )dz
(z )
can cancel the dz 's  viz.,
= fg (z
 and the compatibility
g (z )dz
)
Given

condition (13.1.4) implies that such quotients do patch together (the

0 (z )

factors cancel). Conversely, a meromorphic function times a

meromorphic

1-form

gives a new meromorphic

1-form.

M = P1 , let 1 = be arbitrary and 2 = dz .


Z
1
Here z = 1 on P as usual, and dz looks as if it should be not just
Z0
Z
meromorphic but holomorphic. But in the coordinate at  w = 0 ,
Z1
. So dz in fact has a pole of order 2 at [0 : 1].
dz becomes d( w1 ) = dw
w2

Now consider F (z) := 1 =


K(P1 )(
= C(z) by Thm. 3.1.5(a));
2
dz
we have then = F (z)dz . Therefore



P (z)
1
1
K (P ) =
dz P, Q C[z] of the same degree .
Q(z)
Example 13.1.2. On

M = C/ a complex 1-torus, write u for the


coordinate on C. Since each transition function sends u 7 u +
0
(for some ), their derivatives are all identically 1. Hence, du
gives a well-dened global holomorphic 1-form on M (i.e. belongs to
1 (C/)).
So take 1 = arbitrary, 2 = du. The same argument as above,
Example 13.1.3. For

using Thm. 3.1.5(b), gives

K1 (C/)
= {f (u)du | f = -periodic

Example 13.1.4. Let

can represent

df :=

f K(M )

meromorphic function on

C} .

be a meromorphic function. We

as a collection of maps

f : V P1 .

The

1-forms

df
dz are then compatible (via pullback) with the transition
dz

148

13. MEROMORPHIC 1-FORMS ON A RIEMANN SURFACE

functions, as in (13.1.3); hence, they patch together to give a global


meromorphic 1-form

df K1 (M ).

We will refer to this as the dieren-

f.

tial of

K1 (M ) be given by a collection {f (z )dz }; we would like


dene its order p () at a point p U M . We simply set
Let

to

p () := z (p) (f );
if this is negative

p U

has a pole

at

p.

As a well-denedness check, suppose

also. Then (using (13.1.4))


p (f ) = p f 0 (z ) = p (f )
must have nonvanishing derivative
point. If has a pole at p U , then its residue is

1
Resp () := Resz (p) (f ) =
f (z )dz
2 1 C (p)

since, as a biholomorphism,
every

where

C (p)

is a small circle (in

V )

about

z (p).

at

The well-denedness

check boils down to change of variable in the integral.


Let

= {f (z )dz } K1 (M )

be a form, and

= M

be a

3
smooth real closed curve. Then we dene

:=

f (z )dz ,

where we observe that 1-forms have been set up so that the right-hand
side is independent of choices of local coordinates and the partition of

into local pieces. The following can be viewed as a version of either

Stokes's theorem or Cauchy's theorem.

Proposition 13.1.5. Let

smooth boundary

3A real curve

be a closed region with piecewise

= .

means something 1-dimensional over

of a closed path on the Riemann surface; and

R (not C), so you should think


U are the segments from which

the path is pieced together.

4the

technical term here is

2-chain,

though we won't get into this here

13.1. DIFFERENTIAL 1-FORMS

149

M
Assume that the meromorphic form

containing

is holomorphic on some open set

Then

= 0.

= , but assume that is only


containing (so that may contain poles

Proposition 13.1.6. Again let

holomorphic on an open set


of

).

p()<0

M
(a) Then we have the residue formula

2 1

Resp ().

p
p () < 0

(b) In general for

K1 (M ),

pM

Resp () = 0.

0 be a sum of circular
paths about those p where has poles. Let 0 be the complement
in M of the union of disks containing these {p}, with 0 = 0 . Apply
Prop. 13.1.5 to the pair 0 , 0 .
Applying the residue formula to the case = M , = gives
(b).

Proof. For the residue formula (a), take

Corollary 13.1.7. Consider a nonconstant meromorphic function

f K(M ).

Then

150

13. MEROMORPHIC 1-FORMS ON A RIEMANN SURFACE

(a)

pM

p (f ) = 0,

i.e. the number of zeroes (counted with mult-

plicity) equals to number of poles (counted with multplicity); and


(b)

#{f 1 ()}

(counted with multplicity) is independent of

Proof. (a) is Prop. 13.1.6(b) applied to

f ,

P1 .

df
. Replacing
f

by

and noting that the number of poles doesn't change, by (a) the

number of zeroes can't change either, giving (b).

f , deg(f ), is dened to be the


Thinking of f as a covering map from

Definition 13.1.8. The degree of

number in Cor.

M P

13.1.7(b).

deg(f )

can be visualized as the number of branches (or

5
sheets).
Remark 13.1.9. We have said nothing about

when

is not a

boundary:

M
Indeed, there is nothing we can say yet  this is the study of periods,
which depend on the complex analytic structure of

M.

We will be able

to compute some periods of holomorphic forms on algebraic curves later


in the course.

13.2. Poincar-Hopf theorem


The usual statement of this theorem is that the sum of indices of

any

6 vector eld

~v

on a compact oriented smooth manifold

M ; we'll only worry


M is 2. In that case, the

is equal

to the Euler characteristic

about the case where

the real dimension of

index Indp (~
v ) of

~v

at

p M is the number of counterclockwise rotations done by (the head


of ) ~
v as one goes once counterclockwise on a small circle about p. It
can only be nonzero if ~
v (p) = 0.
5you may wish to refer back to Remark 3.1.7
6technical point: ~v should have only nitely many

zeroes

13.2. POINCAR-HOPF THEOREM

151

I'll give a heuristic proof of the italicized statement, which is probably more illuminating than a formal one. Subdivide a given compact
smooth oriented real 2-manifold

into triangles:

etc.

Then put one marked point on each edge, vertex, and face of the triangulation:

Next draw the following vector eld on each triangle:

These match up to give a global vector eld on


of this

M.

Evidently the index

~v is 1 at the marked points on the edges, and +1 at the marked

152

13. MEROMORPHIC 1-FORMS ON A RIEMANN SURFACE

points on the faces and vertices. Hence,

(13.2.1)

Indp (~
v)

= #F #E + #V = M = 2 2g

pM
where
on

is the genus of

M.

That (13.2.1) holds for any vector eld

is the version of the theorem proved by Poincar. It still holds

if we allow

~v

to have singularities at a nite set of points

(i.e. it is just a section over


indices of

~v

at the

pi

M \{p1 , . . . , pn }),

{p1 , . . . , pn }

provided one adds the

to the sum.

In fact, (13.2.1) even holds if

A1R (M \{p1 , . . . , pn }).

~v

~v

is replaced by a smooth

1-form

M , i.e. a
nonvanishing section of Sym (T M ), to smoothly identify T M with
T M . The corresponding notion of index, if (in local coordinates at p)
takes the form F dx + Gdy , is
 

1
G
(13.2.2)
Indp :=
d arctan
,
2
F
The idea is to use a metric on

and once again the sum in (13.2.1) must be over all zeroes of

and the

{pi }.
1-manifold, and write K1 (M )
be locally in the form f.dx + g.dy where f, g are complex-valued. To
get in the above setting, we may of course view M as a smooth real
2-manifold, and take the real part of :
Now let

be a compact complex

loc

:= <() = <(f )dx + <(g)dy.


= p (). Of course, in a local
7
holomorphic coordinate z about p with z(p) = 0, we have


loc
z dz = r cos() + 1 sin() (dx + 1dy)



= r cos() 1 sin() dx+r sin() + 1 cos() dy.
Let

be a zero or pole of

and put

So locally we have for the real part

cos()dx + sin()dy,
r
and thus by (13.2.2)

1
Indp () =
2
7up

d[] = = p ()

to multiplication by a locally nonvanishing holomorphic function (which will

not aect index)

EXERCISES

153

p () = 2g 2.

p
We have arrived at the following corollary of (13.2.1), which will henceforth be the meaning of Poincar-Hopf  for us:
Theorem 13.2.1. Let

1-form

K1 (M )

on a Riemann surface of genus

(#
|

of zeroes

#
{z

be a nonvanishing meromorphic

g.

Then

of poles) of

counted with multiplicity

= 2g 2.

Remark 13.2.2. Just as for meromorphic functions we can consider

the divisor

() :=

p ()[p]

pM
of a meromorphic 1-form. In this context, the Theorem says that

deg(()) = 2g 2.
Exercises

dx
y

1 (E). (We can


E
talk about holomorphic 1-forms on a smooth algebraic curve now,

(1) Let

E = {y 4x 4x = 0}, =

because they are Riemann surfaces by the smooth normalization


Prop. 7.1.) Consider the complex analytic automorphism

(x, y) 7 (x, iy), and apply this to the 1-form: com


pute the pullback A ().
(a) In Example 13.1.2, dz denes a meromorphic dierential 1-form
1
1
1
on P . Compute its divisor (dz). Explain why (P ) = {0}. (b)
What is the divisor of du on C/, from Example 13.1.3? Explain
why it is the unique holomorphic 1-form on C/ up to scale.
2
2
Practice with pullbacks: for the map : R R that sends
(x, y) 7 (u(x, y), v(x, y)) := (x2 3xy, y 3 +x), compute where
= udv + vdu. Write it in the form f (x, y)dx + g(x, y)dy .
E

(2)

(3)

A: E

sending

(4) Continuing from Exercise 5 of Chapter 3, compute the pullback of

dx
under
y

y(z)

: P1 C .

[Hint: simply plug in your nal

x(z)

and

from that exercise. After simplication, your answer should

be very simple indeed.]

CHAPTER 14

The genus formula

We are ready to prove two formulas for the genus of a Riemann


surface (RS) which are especially useful in algebraic geometry.

For

the rst result (the Riemann-Hurwitz formula), the RS will arise as a


nite branched cover of another RS whose genus is known. The proof
makes essential use of Poincar-Hopf and a ramication divisor which
we introduce below.

For the second result, which is an application

of the rst (and of the intersection theory from Chap.

12), the RS

will arise as the normalization of an irreducible algebraic curve in

P2

with only ordinary double point (ODP) singularities. (In fact, the 4H
people will, in their reading material, learn how to deal with worse
singularities later.)

This is a very concrete payo for the preceding

hard work: now we can compute the genus of [the desingularization of]
a projective algebraic curve!

14.1. Order and multiplicity for maps of Riemann surfaces


Consider a nonconstant morphism
faces with

f (p) = q .

f : M M0

of Riemann sur-

In Exercise 4 of Chapter 3, the following was

established: there exist

f (U ) V , and
local holomorphic coordinates z : U C and w : V C with
z(p) = 0 = w(q),
neighborhoods

w f = z

U 3 p, V 3 q

with

N. More informally, in

these local coordinates f takes the form (w =)f (z) = z . We write


p (f ) := . This is the ramication index, and f ramies at p precisely
when it exceeds 1.
0
For any q M , consider the sum
X
d(q) :=
p (f ).
such that

for some unique

pM with f (p)=q
155

156

14. THE GENUS FORMULA

p of index d lies over q , then over a nearby point


q0 , p is replaced by d points with ramication index 1. This is by virtue

of the local form w = z , as is the fact that the ramication points are
isolated hence nite in number (M is compact!). Evidently then, d(q)
is constant in q ; we will call this constant d N the degree deg(f ) of
1
the morphism f .
If a ramication point

Here is a more gentried way to dene this. We can think of a point

q M0

as a divisor

[q] Div(M ),

and pull it back to a divisor on

2
by the formula

f 1 ([q]) :=

p (f )[p]

Div(M ).

f (p)=q
We then put (for any

q M 0,

it doesn't matter)

X

deg(f ) := deg f 1 ([q]) =
p (f ) .
f (p)=q

f : M M 0 , nally, is the ramication


X
Rf :=
(p (f ) 1) [p] Div(M ).

Associated to

divisor

pM
By the above remarks, the sum is clearly nite.

14.2. Riemann-Hurwitz formula


Again take

d := deg(f ),

f : M M0

to be a nonconstant morphism, write

and put

r := deg(Rf ).
In the following

resp.

Theorem 14.2.1.

g0

will refer to the genus of

resp.

M 0.

r = 2 {g + d dg 0 1}.

Remark 14.2.2. Some alternative ways to write this result are:

g = (g 0 1)d + 2r + 1
(ii) M = deg(f )M 0 deg(Rf )
(i)

These better represent the way you want to think of it: as a formula

1when f

function and take the degree of


can think of it as a morphism of

0.

P1 , you can think of it as a meromorphic


its divisor, deg((f )), which is always 0. Or, you
Riemann surfaces and take deg(f ), which is never

is a nonconstant map from

to

So that extra parenthesis matters!

2This

extends linearly to dene

f 1 (D)

for any

D Div(M ).

14.2. RIEMANN-HURWITZ FORMULA

for the genus (or Euler characteristic) of


and data about how

Proof. For

z, w

as in

14.1

sits over

p M
so that

with a
f
a

M,

157

if you know that of

M0

M 0.

= p (f ),

we choose local coordinates

z 7 z (= w).

We shall need to assume the existence of a nonzero meromorphic

1-form K1 (M 0 ).

This is obvious if

tion of an algebraic curve in

M0

arises as the normaliza-

, as you can just pull back any non-

constant meromorphic function (say,

Z1 /Z0 )

and take its dierential.

Every Riemann surface arises in this way, but to see that you need
the Riemann-Roch theorem. We proceed with the proof modulo this
detail.
Locally writing

= g(w)dw,

we have

loc

f = g(z a )d(z a ) = a.g(z a )z a1 dz ,


hence

p (f ) = a.0 (g) + (a 1) = p (f ).f (p) () + (p (f ) 1).


In Div(M ) we have therefore

(f ) :=

p (f )[p] =

p (f ).f (p) ()[p] +

pM

|
=

q ()

qM 0

(p (f ) 1) [p]

{z

Rf

p (f )[p] + Rf

f (p)=q

q ()f 1 ([q]) + Rf

qM 0

!
= f 1

q ()[q]

+ Rf

=f
() Div(M 0 ).

1
Now f K (M ),

(()) + Rf ,

where

so Poincar-Hopf on

tells us that

2g 2 = deg ((f ))
which by the computation just done


= deg f 1 (()) + deg Rf

158

14. THE GENUS FORMULA

q ()

p (f ) + r

f (p)=q

{z

deg(f )

X
= deg(f )
q () + r.
| {z }
deg(())

Applying Poincar-Hopf once more (but on

M 0 ),

we get that this

= d(2g 0 2) + r.
So we have shown

2 2g = d(2 2g 0 ) r,

which is the version of R-H

stated in Remark 14.2.2(ii).


We turn to some examples.

C = {y 2 =

Q2m

2
i=1 (xi )} C , and let M be
P2 . The original curve
the normalization of its projective closure C
Example 14.2.3. Let

had a projection map to the

x-axis ((x, y) 7 x),

and this extends to

f : M P1 =: M 0 ,
as depicted below:

ramification points

Clearly

g 0 = 0, d = 2,

and

r=
since

p (f ) 1 = 1

(p (f ) 1) = 2m

at each of the ramication points. So by Remark

14.2.2(i)

g = (0 1).2 +

2m
+ 1 = m 1.
2

M = M 0 = C/ be a complex 1-torus; as
usual = {m1 1 + m2 2 | m1 , m2 Z}, where 1 , 2 C are inde
pendent over R. Now assume for some C . Then we have
Example 14.2.4. Let

a complex multiplication map

M M 0

14.3. THE GENUS OF A PROJECTIVE ALGEBRAIC CURVE

159

z 7 z
which has

Rf = 0.

You will treat this setting in an exercise below.

14.3. The genus of a projective algebraic curve

C = {F (Z, X, Y ) = 0} P2 be an irreducible algebraic curve


of degree d with S = sing(C) its set of singular points. We assume that
these are all ordinary double points, and that there are exactly |S| =
of these; write S = {p1 , . . . , p }. Of course, = 0 S =
C is smooth.
 C its normalization, we shall deduce from
Denoting by : C
Let

Theorem 14.2.1 the formula:


Theorem 14.3.1.

has genus

g=

(d 1)(d 2)
.
2

To get a feel for this before launching into the proof, for

smooth

we have

d = 1 = g = 0,
d = 2 = g = 0,
d = 3 = g = 1,
d = 4 = g = 3,
and so on. For degree 3 with one ODP, we get

g=

(3 1)(3 2)
1 = 0,
2

as we found using stereographic projection. Indeed, we know how to


parametrize all three genus

cases (smooth

d = 1, 2;

singular

d = 3)

by a Riemann sphere.
The rest of this section is devoted to the proof. Begin by choosing
coordinates on

P2

so that

L C consists of d distinct points,


none of the tangents to C at its ODP's are vertical (i.e.
form X = aZ ), and
C does not contain [0 : 0 : 1].
The latter requirement allows us to project from
map

C P1 =: M 0

[0 : 0 : 1]:

of the

that is, the

160

14. THE GENUS FORMULA

given by

[Z : X : Y ] 7 [Z : X],
roughly speaking the projection of
Writing

M := C ,

to the

x-axis,

is well-dened.

the main idea of the proof of to apply Riemann-

Hurwitz to the composition

f = x : M M 0.

In a picture, where

VT refers to a point with vertical tangent:

C
x
VT

VT
ODP
1

P
Now for

M 0 = P1 , g 0 = 0

so that Thm. 14.2.1 gives

rf = 2(genus(M ) + deg(x) 1) = 2(g + d 1).

(14.3.1)

In particular, the degree of the map

is

because the projection is

done along vertical lines, all but nitely many such lines meet
points by Bezout, and

in

is 1-to-1 o nitely many such points. So we

see that if we can compute the degree of the ramication divisor

Rf

then we are done.


To do this, let

E := {FY = 0}
where

FY

is the partial derivative. Obviously

by Bzout,
(14.3.2)

(E C) = (d 1)d.

deg(E) = d 1,

and so

14.3. THE GENUS OF A PROJECTIVE ALGEBRAIC CURVE

P0
Denoting by
p the sum over points where
P
and by
j=1 the sum over ODP's, we have
X0

(E C) =

(E C)p +

161

has a vertical tangent,

(E C)pj .

j=1

We will show

Rf =

(14.3.3)

X0

(E C)p [
p]

p
where

p = 1 (p) C .

(Recall that by our choice of coordinates, a

point with vertical tangent cannot be a singular point, and so has a

unique preimage point under normalization.) Taking degrees of both


sides of (14.3.3) gives

rf =

(14.3.4)

X0

(E C)p = (E C)

(E C)pj .

j=1

Further, we will deduce that

(E C)pj = 2 (j);

(14.3.5)

together with (14.3.2) and (14.3.3), this yields

rf = d(d 1) 2.
Now put this together with (14.3.1) to get

2g + 2(d 1) = d(d 1) 2,
2g = (d 2)(d 1) 2,
and divide the last line by

to get Theorem 14.3.1. It remains only to

check (14.3.3) and (14.3.5).

p C E.
By assumption, p
FX (p) 6= 0. By the
holomorphic implicit function theorem, we can parametrize C locally
by writing x = X/Z as an implicit function of y = Y /Z , viz.
If

has a VT at

p, then F (p) = FY (p) = 0;

this implies

3
is a smooth point, so that

0 = F (1, x(y), y).

3if F (p) = 0
X
at

then

FZ (p) = 0

too by the Euler formula, contradicting smoothness

162

14. THE GENUS FORMULA

Now, dierentiating gives

0=

d
F (1, x(y), y) = FX (1, x(y), y) x0 (y) + FY (1, x(y), y).
dy

For the two functions on the right-hand side to sum to zero, they must
have the same order to

y(p):

ordy(p) FY

(1, x(y), y) = ordy(p) x0 (y),

in other words

(E C)p = {ordy(p) x(y) 1}


= {p (x) 1}
= {p(f ) 1}.
As the only ramication points of

Rf :=

are (

(q (f ) 1)[q] =

of ) vertical tangent points,

X0

(E C)p [
p]

qC
as claimed.

Finally, to see (14.3.5), assume for simplicity (for some

(0, 0).

j ) pj =

The local ane equation about an ODP is of the form

F (1, x, y) = ax2 + 2bxy + cy 2 + {higher-order

terms}.

To nd the tangent lines, recall that one solves

0 = ax2 + 2bxy + cy 2 =

x y

!
a b
c d
| {z }

x
y

in

P1

(for their slopes). That the solution

consists of two distinct

pj is an ODP) = Q is smooth = det B 6= 0 =


ac b2 6= 0. That there is no vertical tangent = [x : y] = [0 : 1] is
not a solution = c 6= 0. Consider the partial

points (as

FY (1, x, y) = 2bx + 2cy + {higher-order


whose vanishing denes
about

pj

E;

evidently

terms}

can be locally parametrized

by

b
y = y(x) = x + {higher-order
c

terms}.

14.4. BEYOND STEREOGRAPHIC PROJECTION

To compute its intersection number against


equation of

C,

163

we pull the dening

back along this parametrization and take the order at

0:

(E C)(0,0) = ord0 (F (1, x, y(x)))



= ord0 ax2 + 2bx y(x) + c(y(x))2 + {higher-order terms}


ac b2 2
= ord0
x + {higher-order terms}
c
= 2,
Q.E.D.

14.4. Beyond stereographic projection


The genus formula is very nice, but needs to pass a smell test: if it

C P2 has genus zero normalization, then we should


parametrize C by the unique genus zero Riemann surface

says that a curve


be able to

P1 .

We know that this can be done for a smooth conic and a nodal

cubic (i.e. a cubic with one ODP); the rst new case predicted by the
formula is that of an irreducible

4 quartic curve (d

= 4)

with 3 ODP's

( = 3):
g=
Let's give this a try.
another curve

2.

If

(4 1)(4 2)
3 = 0.
2
Write {pi }i=0,1,2 for the

D passes through one of these:

then by 12.2.4,

is a line, then it cannot pass through all 3

have

ODP's, and suppose

pi ,

(C D)pi

as then we would

2
X
4 = deg C deg L = (C L)
(C L)pi 6,
i=0

a contradiction. So the ODP's are not collinear, and by a similar argu-

C , then no three of p0 , p1 , p2 , p3
are collinear. We may therefore move C (and the pi ) by a projectivity
2
of P , to have p0 = [1 : 0 : 0], p1 = [0 : 1 : 0], p2 = [0 : 0 : 1],
p3 = [1 : 1 : 1]. (We'll do so for this abstract analysis but not for the

ment if

p3

is any xed smooth point of

concrete example that follows.)


The general conic in

P2

is of the form

aXY + bY Z + cXZ + dX 2 + eY 2 + f Z 2 = 0.
4We

have to say

is irreducible explicitly, because the union of a smooth cubic

and a general line is a quartic with

5the 6

gets replaced by a

ODP's.

in the inequality above

164

14. THE GENUS FORMULA

By substitution, we nd that the general conic through the above four
points is of the form

Q[a:b] = {aXY + bY Z (a + b)XZ = 0}.


[a : b] P1 .
The zero-cycle (cf. Remark 12.2.2(a)) Q[a:b] C has degree 8 by
Bzout, and is of the form 2[p0 ] + 2[p1 ] + 2[p2 ] + [p3 ] + more. This
more can only be one more point q[a:b] with multiplicity one, since
what is already written has degree 7 (and by construction, one doesn't
have negative intersection numbers). Naturally, q could be one of the
pi : if it is p3 , then this would say that Q is tangent to C there. Dene
This is a 1-parameter family parametrized by

a map

: P1 C
by

[a : b] 7 q[a:b] (:= Q[a:b] C {2[p0 ] + 2[p1 ] + 2[p2 ] + [p3 ]}).


P1 P2 (I
that is onto

In fact, this is a morphism of complex manifolds from


won't prove this carefully). Also, since

is irreducible,

essentially follows from the open mapping theorem and compactness of

P1 .
is 1-to-1 o the singular points of C . Take q C
distinct from the pi ; since no three of the pi are collinear, no four of
q, p0 , p1 , p2 , p3 are collinear, so there exists a unique conic Q through
all ve. (The uniqueness when q = p3 then essentially follows from
continuity of .)
We claim that

Example 14.4.1. So what does such a normalization look like?

Take the very concrete quartic curve

C = {X 2 Z 2 + Y 2 Z 2 + 2X 2 Y 2 = 0}.
Irreducibility can be checked by putting the polynomial in ane form

y 2 (1+2x2 )+x2 and showing it doesn't factor into terms of lower degree
in y . I will let you check that the only singularities are p0 = [1 : 0 : 0],

p1 = [0 : 1 : 0], p2 = [0 : 0 : 1]; pick p3 := [i : 1 : 1] (i = 1). The


general conic through these 4 points is

Q[:] := {XZ + Y Z = i( + )XY }.

EXERCISES

Substituting this into

times the equation of

165

gives

0 = (i( + )XY Y Z)2 + 2 Y 2 Z 2 + 22 X 2 Y 2 =




2 + 2 2
2
2
2
= ( + )Y (Z iX) Z i
X ,
2 + 2

X
in which the last factor gives us the x =
-coordinate of the point
Z
q[:] . The y -coordinate is obtained by substituting into the equation
of Q, and we nd ([ : ]) =

i(2 + 2 2 )(2 + 2 2 ) : (2 + 2 )(2 + 2 2 )

: (2 + 2 )(2 + 2 2 ) .
Or, in ane coordinates (t


(t) = i

in particular),

1 + t2
1 + t2
,
i
t2 + 2t 1 t2 2t 1


.

Exercises

C , C 0 compact RS with g =
genus(C), g 0 = genus(C 0 ), f : C C 0 nonconstant holomorphic
0
map of degree d. Show that for any d 1, g g . (The covering

(1) Recall the setup of Riemann-Hurwitz:

surface has at least as many handles.)


(2) Let

z=

Z1
(where
Z0

[Z0 : Z1 ]

are the homogeneous coordinates) be

P1 . If a holomorphic map f : P1 P1
n
n1
takes the form f (z) = z + a1 z
+ . . . + an , then
(a) What is deg(f )?
(b) What can you say about the ramication divisor Rf ? (at least,
the canonical coordinate on

what is its degree?)


(c) Use Riemann-Hurwitz to check your answers.

= {m1 + n2 | m, n Z} is a lattice in
C. (In particular, 1 and 2 are independent over R.) Suppose

that C satises . [Remark: if


/ Z this places a
strong condition on ; we say , or C , has complex multiplication
(or CM ). ] The multiplication by induces a holomorphic map
: C C , i.e. an automorphism of the RS C .
(a) Show that the ramication divisor R Div(C) for this map is

(3) Let

C = C/

where

zero.
(b) Prove that the degree of
image lattice

equals the index

[ : ]

of the

166

14. THE GENUS FORMULA

e of the irreducible curve C


C
x2 + x2 y 2 + y 2 = 0 in P2 . (First

(4) Find the genus of the normalization


given by taking the closure of

convert to homogeneous coordinates and check for singularities.


Then apply the genus formula. This is very similar to something
above...)
(5) This problem complements (3) above, but you won't use anything

f : C C of
f : C C (i.e.

from this chapter in doing it. A (holomorphic) map


Riemann surfaces is nothing but an analytic map
an entire function) such that for all

()
i.e.

z1 z2

denedness

, z C,

f(z + ) f(z) ,

= f(z1 ) f(z2 ) mod (this is just the wellcondition for f ). Show that such a map is necessarily

mod

ane, i.e. of the form

f(z) = z + .
[If

other than

works, then we are of course in the CM

case described above. So a non-CM complex 1-torus, which is the

z 7 nz + , n Z,
().

generic case, has endomorphisms of the form


and that's all.] Hint: consider

f0 (z),

and use

CHAPTER 15

Some applications of Bzout

We have already put Bzout's theorem to use in proving the genus


formula (and in several interesting exercises at the end of Chapter
12). Now we shall use it to prove a general result on curves through
congurations of points, which in particular will yield a short (and
rigorous) proof of Pascal's theorem from Chapter 1.

We shall also

deduce some results on cubics will will come in handy in studying the
group law on elliptic curves.
Throughout this Chapter we shall use the following dictionary:

P2

algebraic curve

dening equation

degree

C
D
E

(homogeneous polynomial)

F S3d
G S3d
H S3e

d
d
e

Recall the theorem we are wanting to apply:

Bzout.

C E

is

0-dimensional (consists of points) = (C E) =

de.
Part of the content of the (equivalent) contrapositive statement is:

tuozB. The number of points

|C E|

exceeds

de = E

have a common component.

From Chapter 9, we have:

Study.

irreducible and

Putting tuozB and

BS.

E C = H

divides

Study together gives:

irreducible and

|C E| > de = H | F .

We'll make use of this statement below.


167

F.

and

168

15. SOME APPLICATIONS OF BZOUT

15.1. Cayley-Bacharach theorem


Let

p1 , . . . , p n P2

be distinct points, and dene

S d (p1 , . . . , pn ) :=

in

[Z : X : Y ]

vanishing at

E is irreducible
pa+1 , . . . , pn
/ E . Then

Lemma 15.1.1. Suppose

some

a > ed,

while

d)
p1 , . . . , p n .

homogeneous polynomials (of degree

and

p1 , . . . , p a E

for

S d (p1 , . . . , pn ) = H S de (pa+1 , . . . , pn ).
Proof. The inclusion of the RHS into the LHS is easy, since it is

just saying that the product of a polynomial vanishing at the last


points by a polynomial vanishing at the rst

na

points, vanishes at all

of them. So we turn to the reverse inclusion.

S d (p1 , . . . , pn ) is nonzero, take a nonzero element F ; this


denes a degree d curve C containing p1 , . . . , pn . Clearly p1 , . . . , pa
C E , so |C E| > ed, and by BS, H | F . We can therefore write
F = F0 H with F0 S de . Since F = 0 but H 6= 0 at pa+1 , . . . , pn , F0
de
must vanish at these points. It follows that F0 S
(pa+1 , . . . , pn ) as
desired.

Assuming

Theorem 15.1.2. Let

and assume exactly

d(d e)
d e.

remaining
of degree

ed

|C D| = d2 with d > e,
C D lie on E . Then the

be irreducible,

of the points of

points lie on a (not necessarily irreducible!) curve

[A : B : C] E\{(C D) E}, and set =


F (A, B, C), = G(A, B, C). Dene P := G + F S d ; this vanishes on C D and at [A : B : C]. Label (C D) E =: {p1 , . . . , ped },
[A : B : C] =: ped+1 , and (C D)\{(C D) E} = {ped+2 , . . . , pd2 +1 };
2
set a := ed + 1 and n = d + 1.
d
Since a > ed, Lemma 15.1.1 tells us that S (p1 , . . . , pd2 +1 ) = H
S de (ped+2 , . . . , pd2 +1 ). But then, since P S d (p1 , . . . , pd2 +1 ), we have
P = HP0 for some P0 S de (ped+2 , . . . , pd2 +1 ). This P0 denes the
required curve.

Proof. Let

Here is the nice application to Pascal:

1it

is enough to check, in applying this, that

points) lie on

can't lie on

E.

E.

ed of the points (not exactly ed of the


ed of these points simply

This is because by Bzout, more than

15.2. INTERSECTIONS OF CUBICS

169

Corollary 15.1.3. The (three) intercepts of opposite sides of a

hexagon inscribed in a conic are collinear.


Proof. Referring to the picture

p5
L2
p6

L1
L6

L3

L4 p 2 L5

p1

p4

p3
q3

q2

C := L1 L3 L5 , D = L2 L4 L6 , and E = Q. Clearly this


means d = 3 and e = 2, and we do indeed see that de = 6 points of
C D = {p1 , . . . , p6 } {q1 , q2 , q3 } lie on E . So the last three points of
C D, which are the intercepts, lie on a curve of degree d e = 1 by
the Theorem.

we put

Remark 15.1.4. If one wanted instead to plug the technical gap in

the proof of Pascal suggested in Chapter 1, part of what one needs is the
statement: if

p1 , . . . , p8 P2 are distinct and in general position

in the

sense that no 4 are collinear and no 7 conconic (lying on an irreducible


conic), then

dim S 3 (p1 , . . . , p8 ) = 2.

This is proved in Reid's book.

15.2. Intersections of cubics


The results of

15.1

dealt with the case where all intersections of

curves have multplicity one (the transversal case), since we required

|C D| = d2 = (C E).
assuming E is smooth and

To deal with the general case, at least


irreducible (so that we may view it as a

Riemann surface), write

C E :=

(E C)p [p]

Div(E).

pEC
If

is irreducible but singular, with an unique ODP

nition gives a divisor

p
/ E C.

C E Div(E)

p,

the same de-

(on the normalization) provided

170

15. SOME APPLICATIONS OF BZOUT

Theorem 15.2.1. Let

(If

C, D, E

E irreducible.
p
/ E C, E D.) Writing

be distinct cubics, with

is singular, assume moreover that

by Bzout
9
X

DE =

[qi ]

Div(E)

i=1

where the

qi

need not be distinct, and assuming


8
X

C E =

[qi ] + [q] Div(E),

i=1

we have

q = q9 .

In the intersection multiplicity one case, the Theorem gives immediately:


Corollary 15.2.2. Let

D E = {q1 , . . . , q9 }

C, D, E

be distinct cubics,

(distinct points) and

then it passes through

irreducible. If

passes through

q1 , . . . , q 8 ,

q9 .
E

Actually this is true without assuming


doesn't share any components with

or

Proof. (of Theorem) First assume

C ),

irreducible (provided

meromorphic function on

P2 .

but we won't prove that.

is smooth. Recall that the

quotient of two homogeneous polynomials  say,

D = {G = 0}

F/G

By Example 7.3.6, since

in points, we may pull this back to

 yields a

intersects

E:


F
f := K(E) .
G E
Suppose (for a contradiction) that

D = {G = 0},

the divisor of

q 6= q9 .

Since

C = {F = 0}

and

is evidently

(f ) = C E D E = [q] [q9 ] Div(E).


This says that

has one zero (at

q)

and one pole (at

holomorphic map of Riemann surfaces

1.

That is,

EP ,f

q9 );

hence, as a

has mapping degree

is 1-to-1; and since (using the open mapping theorem) its

image must be open and closed (and

P1 is connected), f

is surjective. So

f gives an isomorphism E
= P . Trouble is, this is total rubbish. Since
E is a smooth cubic, its genus is 1 by the genus formula, whereas the
1
genus of P is zero. So they can't be isomorphic for purely topological
1

EXERCISES

171

reasons. This contradication tells us that, indeed, our assumption

q9

q 6=

was wrong, and so they are equal.

E is singular with ODP


: E P2 (of E ) to
P1 . As before,
from E

To extend this argument to the case where

F
p, rst pull back G
along the normalization

. We regard f as a map
f K(E)
assuming q 6= q9 leads to deg(f ) = 1. However, a dierent objection
to  deg(f ) = 1 will be required as there is no topological obstruction:

indeed, E
= P1 by the genus formula (a nodal cubic has genus zero
normalization). So argue as follows: since p

/ C, D, we nd that
F
2
K(P ) is well-dened at p, so its pullback via cannot separate
G
mapping
the two branches of E there. That is, at the two points of E
to p
(under ), f will take the same value. But then, the mapping
degree of f cannot be 1.
The other possibility is that p
is a cusp.2 We may assume p = [1 :
0 : 0] and the equation is of the form x3 = y 2 . Again we need to show
F
, if nonconstant, cannot have mapping degree 1. Let
that f =
G
F (1,x,y)
F (1,0,0)
1
p = (
p), and write R(x, y) := G(1,x,y)
G(1,0,0)
. Then f (t) f (0) =
( R)(t) = R(t2 , t3 ), and
obtain

deg(f ) = deg f 1 ([f (0)])


ord0 (R(t2 , t3 ))
ord(0,0) (R(x, y)) min{ord0 (t2 , t3 )}
{z
}
|
{z
} |
=2

2.

Exercises

F = 0, G = 0, H = 0),
of respective degrees d, d, e with 3 e d. Suppose C and D
2
intersect in d distinct points, and assume that E is smooth (hence
irreducible). Show that if E passes through ed 1 of these, it
passes through ed of them. (Imitate the argument from the proof

(1) Let

C , D,

and

be as above (dened by

of Theorem 15.2.1.)

2see

the paragraph immediately preceding

16.1

below.

Part 3

Cubic curves

CHAPTER 16

The singular cubic

Recall that a singular cubic curve

D P2

ographic projection through its singular point

is normalized via stere-

p;

that is, we get a nor-

malization morphism

: P1 P2
with image

D.

In particular, all singular cubics have normalization of

genus zero. Moreover, they are all projectively equivalent to one of two
examples.

The nodal cubic. A node is just an ordinary double point. Let

D = {Y 2 Z = X 2 (Z X)};

the ane equation is

y 2 = x2 (1 x)

and a

schematic picture is

where I have denoted points with real coordinates in blue and points
with only

x-coordinate

real in red. (How these sit inside the full set

of complex points will be pictured below; the dotted stu will connect
up.) By Exercise 5 of Chapter 3, this is parametrized by

: P1 /{0, } D
1D

is for degenerate
175

176

16. THE SINGULAR CUBIC


t 7
The  P

/{0, }

4t
t(1 + t)
,
2
(1 t)
(1 t)3


=: (x(t), y(t)).

means the Riemann sphere with the top and bottom

2
points identied.

The cuspidal cubic. Take


2

y =x

D = {Y 2 Z = X 3 },

ane equation

, schematic picture

p
where I have only drawn real points. To do stereographic projection
through the cusp

= x =

(0, 0),

write

y = 1t x

and substitute to get

1 2
x
t2

= x3

1
. Hence we get a normalization
t2

: P1 D
dened by



1 1
t 7 1 : 2 : 3 .
t t
One of our overarching themes in the next few chapters will be the
study of algebro-geometrically dened group laws on cubics.

In this

chapter, we focus on the above two singular examples, as the smooth


case is more dicult.
be equivalent (via

For the nodal cubic, the law will turn out to

C = P1 \{0, }; while in
1
addition on C = P \{}. In both

to multiplication on

the cuspidal case, it identies with

cases, these sets are the preimages under normalization of the smooth
points of

2Note

D,

which is where the group laws will be dened.

that one can homogenize the formula for

4T1 T0 (T0 T1 ) : T1 T0 (T0 + T1 )],


[0 : 0 : 1].

and then it is

[T0 : T1 ] 7 [(T0 T1 )3 :
clear that (1) = ([1 : 1]) =

by

16. THE SINGULAR CUBIC

177

In the course of studying such laws as well as addition theorems on


these curves, we will pull back rational functions on

P2

(quotients of

homogeneous polynomials of equal degree, or equivalently elements of

C(x, y))

to get meromorphic functions on

P1

(the normalization of our

curve). So in illustrating the simplicity of the group law, hence

Principle 1: Singularities make curves of a given degree more trivial and easier to study,

we will be seeing a concrete example of the following

C P2 an irreducible algebraic
is of the
: C P2 , every f K(C)

Principle 2: Given
normalization

curve with
form

F ,

F C(x, y).
In other words, writing
of

and

gC0 (x, y)

C[C0 ] :=

C0 := C (P2 \{Z = 0})

for the ane part

for its dening equation, if we dene

C[x, y]
,
(gC0 )

C(C) := ( fraction
]
eld of C[C0)
F C(x, y)

F
,
=
F
/ on C

then Principle 2 says that

K(C)
analytic
Since

was projective,

C(C).
algebraic

is compact, and that turns out to be of

fundamental importance: e.g.,

C[C0 ]

is only a subring of, rather than

equal to, the ring of holomorphic functions on (the desingularization


of )

C0 .
Before continuing on, we should address one point:

why should

be an ODP

the only possible singularities of an irreducible cubic


or cusp, and why must it have only one?

First of all, if it had two

L through those two points.


Both intersection multiplicities (of C with L at these two points) would
have to be 2, and so (C L) 4 in violation of Bzout. (See what
a useful theorem this is?) So C can only have one singular point,
and as its equation is of degree 3 that point can only be of order 2
or 3. If it is of order 3, then by the result of Chapter 6 Exercise 5,
singular points, then we could take a line

178

16. THE SINGULAR CUBIC

is a union of

lines, contradicting irreducibility. Finally, the local

equation about a non -ordinary double point of

3. An explicit
local analytic transformation puts this in the form (
x) + (
y )3 = 0. So
2
n
it is a cusp. Alternately, anything which looks like x + y = 0 has
intersection multiplicity n with the line x = 0, again violating Bzout
(in the context of our cubic curve) if n > 3.
form

x + f3 (x, y) = 0,

with

f3

can only be of the

homogeneous of degree

16.1. Warm-up: Functions on a nonsingular conic


Our smooth conic is named
a map

C P

C.

Any

F K(C)

can be viewed as

. Composing this with the stereographically produced

1 =
normalization : P C , yields

=
P1 / C F / < P1 ,

F
that is, a meromorphic function on

t := T1 /T0 ), F

P1 .

Since

K(P1 ) = C(t)

(here

must be of the form

G(T0 , T1 )
g(t)
=
h(t)
H(T0 , T1 )
where

g, h, G, H

same degree.

are polynomials and

are homogeneous of the

By the fundamental theorem of algebra, we can write

this as

for some

G, H

, i , j C.

Q
(T i T0 )mi
Qi 1
,
nj
j (T1 j T0 )

T0N

As

deg G = deg H = N +

mi

nj = 0,

the expression simplies to

Q
(t i )mi
Q
(= ( F )(t)).
(t j )nj
Note that
(16.1.1)

( F )() 6= 0,

mi =

nj .

16.2. Functions on a nonsingular cubic (nodal case)


Let

F : D P1

be

the restriction to
(16.2.1)

of a rational function on

6 0,
=
p D.

which is well-dened and


at the singular point

P2

16.2. FUNCTIONS ON A NONSINGULAR CUBIC (NODAL CASE)

P1 D

Since the normalization

0, 7 p

sends

179

but is otherwise

1-to-1, we get

(P1 /{0, })

/
:

P1

F
with

F (0) = F () C

. Henceforth we shall, by abuse of notation,

refer to this composition as

Thinking of

F.

as a meromorphic function on

P1 ,

(16.1.1) applies

and we get

Q
(t i )mi
F (t) = Q
(t j )nj
Furthermore,

with

mi =

nj .

Q mi

= F () = F (0) = Q inj
j

so that

(16.2.2)
relating the

z -coordinates

imi =

j j ,

of the zeroes and poles of

F.

Now introduce the multivalued function

u :=
1
on

dt
= log(t)
t

, which takes well-dened values in

(16.2.2) in terms of

u:

C/2 1Z.

We can restate

viz.,

p (F ) u(p) 0

mod

2 1Z.

pD
This leads to Abel's theorem for the singular cubic :

P, Z Div(D\
p)

Proposition 16.2.1. Given

eective divisors of

the same degree,

dz
0
z

mod

P = poles
Z = zeroes
3recall

2 1Z

)
of some

as in (16.2.1).

divisors are formal sums of points on a complex manifold with integer coef-

cients. A divisor is eective if none of those coecients are negative.

180

16. THE SINGULAR CUBIC

P
P
P = nj [j ] and Z = mi [i ] are of the same degree
P
P
P
(d =
nj =
mi ), then we may write Z P = dk=1 ([zk ] [pk ])
Z
P zk
by some choice of paths. Also, in the statement
:=
and
pk
P
Explicitly, if

poles and zeroes are as usual meant with multiplicity.

This is a

rst baby case of a general statement for algebraic curves (Abel's


theorem) connecting integrals of

1-forms

to the question of when a

divisor is the divisor of a meromorphic function.

16.3. Group law on the nodal cubic


4

Fix a normalization


=
: P1 /{0, } D
t 7 (x(t), y(t))
1 7 (1) =: e.
p, q D be arbitrary nonsingular points, and Lpq be the line
through p and q . (If they are the same, then take L to be the tangent line Tp D .) By Bzout, (Lpq D) = 3 and so Lpq meets D in a third
point which we call p q . More precisely, everything is counted with
multiplicity (p q need not be distinct from p or q ) so we really mean
Let

[p q] := Lpq D [p] [q].


Now let

L0

be the line through

pq

and

(or

Te D

if they coincide),

and put

[p + q] := L0 D [p q] [e].
That is,

p+q

is the extra intersection point of this line with

guaranteed by Bzout. Here's a useful picture of the construction:

4when

we need to use homogemeous coordinates,

(t) = [Z(t) : X(t) : Y (t)]

16.4. ADDITION THEOREMS FOR THE NODAL CUBIC

181

p+q

p
e

0,

t(p*q)=1/(t(p)t(q))

t(q)

1=t(e)

t(p)
t(p+q)=t(p)t(q)

0
p*q

Now writing

fL

for the equation of a line


fLpq
F :=
fL0
satises (16.2.1).
back along

F (t) =
But since

L,

observe that

: D P1
D

In terms of the

t-coordinate

on

P1 ,

i.e.

pulling

we must have:

(t t(p))(t t(q))(t t(p q))


(t t(p))(t t(q))
=
.
(t t(p + q))(t 1)(t t(p q))
(t t(p + q))(t 1)

F (0) = F (), by (16.2.2)


Y
Y
{locations of zeroes} =
{locations

of poles}

= t(p) t(q) = t(p + q) t(e) = t(p + q).


|{z}
1

C = P1 \{0, } with
the one just dened on D\
p. Alternately, taking log gives

u(p) + u(q) u(p + q) mod 2 1Z,

identifying addition on D\
p with addition in C/2 1Z. This may
This identies the group law (multiplication) on

be rewritten

(16.3.1)

t(p)

dt
+
t

t(q)

dt

t(p+q)

dt
t

mod

2 1Z.

16.4. Addition theorems for the nodal cubic


Let's unwind the equivalence of group laws in the nodal cubic
example from the beginning of the chapter. Noting that

[0 : 0 : 1],

here is a picture of how the group law works:

e := (1) =

182

16. THE SINGULAR CUBIC

t.

x=cons

p+q
p

p*q
In particular, the
the

y -coordinates

x-coordinates of p + q
are of each other.

and

pq

are the same, while

Just to clarify the topology of the situation, here is what the projection of the normalization of

onto the

x-axis

"schematic" picture

looks like:

topological picture

0
8

xvalues

Px

1
8

1
tvalues

(real axis in bold)

It is a 2-sheeted cover with 2 branch points, with the closed path


indicating the equator (or unit circle

|t| = 1)

on the upper

(i.e.

).
D
Now we get to work. Start by inverting the equivalence

t(p + q):
(t1 ) + (t2 ) = (t1 t2 ).
| {z } | {z } | {z }
p

p+q

t(p)t(q) =

16.4. ADDITION THEOREMS FOR THE NODAL CUBIC

Since

p, q ,

and

pq

183

are collinear by construction,

Z(p) Z(q) Z(p q)

0 = det X(p) X(q) X(p q) .


Y (p) Y (q) Y (p q)
Assuming none of them is

e, Z(p)Z(q)Z(p q)) 6= 0

and we get the

1st addition theorem:

1
1
1

0 = det x(t1 ) x(t2 ) x(t1 t2 ) .


y(t1 ) y(t2 ) y(t1 t2 )

x(t1 t2 ) from x(t1 )


p
y(t) = x(t) 1 x(t).

This allows you to compute


equation of
Next,

dt
, while
t

D


to write

and

x(t2 ),

using the

= d(x(t))
= [use Exercise 4 from Chap.
y(t)
D

dx
= xdx1x ; so (16.3.1) may be expressed
y

dx
y

13]

x(q)
x(p+q)
dx
dx
dx

.
x 1x
x(e)(=) x 1 x
x(e) x 1 x
x(e)

dx

dt
(Note that 2 1 =
=
. Going modulo its integer multi|t|=1 t
y
ples, which is what   means here, is necessary not to have the equation's correctness depend upon the choice of paths from to x(p), to
x(q), and to x(p + q).) Solving

1
1
1

det
x(p)
x(q)
x(p + q)
=0
p
p
p
x(p) 1 x(p) x(q) 1 x(q) x(p + q) 1 x(p + q)
for

x(p)

x(p + q)

yields

x(p)x(q)
x(p + q) = p
2 .
p
1 x(p) + 1 x(q)
Forgetting the association with

p, q , p + q D

we get the

2nd addition theorem:


x1

x2
x1x2
dx
dx
dx
( 1x1 + 1x2 )2

x 1x x 1x
x 1x

Note that

dx
x 1x

= log

1x1
1x+1

mod

2 1Z.

by explicit computation of the

184

16. THE SINGULAR CUBIC

integral.

(One way to view this function is

log(t) (= u)

viewed as a

multivalued function of x.) So we have discovered a functional equation




1x1
for log
, which is ugly to check by hand.
1x+1
One aspect of the game we have just played here is:

start with

a natural choice of dierential 1-form on the curve (if possible, one


which is smooth away from any singularities of the curve). In the above,

dx
| . You can think of this as a multivalued 1-form on the
y D
x-axis, and then D is the existence domain of the 1-form over P1x .
this was

Then you integrate this 1-form, which gives a transcendental function


which is multivalued even on

(you have to go to its universal cover

to make it well-dened), and try to produce a functional equation for


it (as a function of

x).

In the last section we'll summarize this story

for a couple of other curves.

16.5. Other addition theorems (conic, cuspidal cubic)

C = {y 2 + x2 = 1}, parametrized


2t t2 1

t 7 2
,
t + 1 t2 + 1

Consider the example

as in

3.3.

by

P1

via

We compute


2dt
dx
= 2
= 2d(arctan(t)),


y C
t +1

dx
dx
=
= d(arcsin(x)).

y C
1 x2
1
On the universal cover of Px \{1} let = arcsin(x) (starting at x =
0 t = 0 = 0).5 Its role is similar to that of u = log t
above, as the integral of our chosen dierential 1-form on the curve;
takes well-dened values in C/2Z. Writing x(1 ) =: x1 , x(2 ) =: x2 ,
x(1 + 2 ) =: x12 , the standard trigonometry relations give
q
q
x12 = x1 1 x22 + x2 1 x21 .

5Note: t 7

2t
t2 +1 = x is a degree-2 map (from C to the x-axis) with ramication
points t = 1 over x = 1. On the complements of these points, we have a 2-to-1

map C C . The universal cover of C is C, and so we have maps C Ct Cx


2 tan 2
sending to t to x. So our setup encodes the relation
= x(t()) = x() =
(tan 2 )2 +1

sin().

EXERCISES

185

The second addition formula for the conic then reads

x1

dx

+
1 x2

x2

dx

1 x2

which is a functional equation for

x12

arcsin.

dx

1 x2

mod

2Z,

More simply put, it is just

the inverse of the trigonometric identity.


Next, look back to the cuspidal example from the beginning of the
chapter. We have

1
t2
1
t3


d
dx
d(x(t))
=
=

y D
y(t)


= 2dt,

while


dx
dx
= 3.

y D x2
dx
(Note that this time, the integral of
| is just 2t
y D
valued on D .) Clearly if t12 = t1 + t2 , then
t1
t2
t12
dt +
dt =
dt

0
1
(=x(t1 ))
t2
1

dx
x

3
2

1
(=x(t2 ))
t2
2

dx
x

3
2

and is not multi-

1
(=x(t1 +t2 ))
t2
12

dx
3

x2

1
So we get a functional equation for , which is unfortunately rather
x
stupid: it says

1
1
  21 +   12 = 
1
t21

1
t22

1
1
(t1 +t2 )2

 12 .

In an exercise below, you will show a less trivial addition theorem for
the cuspidal cubic, to the eect that

P, Q, R (D\
p)

are collinear

t(P ) + t(Q) + t(R) = 0.

Exercises

D = {Y 2 Z = X 3 } P2 and
1
1
1
normalize it as above, with : P D given by t 7 [1 : 2 : 3 ] =
t
t
[Z : X : Y ]. (The singular point is p = [1 : 0 : 0].) Prove directly
1
that the group law given by addition on (P \{})
= C (namely,
t1 , t2 7 t1 + t2 ) corresponds to the following process on (D\{
p}):
0
take the line L through (t1 ) and (t2 ), then a line L through
the third intersection point (t1 ) (t2 ) (of L with D ) and the
neutral point [0 : 0 : 1], and nally locate the third intersection

(1) Consider the cuspidal cubic curve

186

16. THE SINGULAR CUBIC

point of this

L0

with

to get  (t1 ) + (t2 ) (also as above, for the

P, Q, R (D\{
p})
t(P ) + t(Q) + t(R) = 0. (Here P, Q, R

nodal cubic). Do this simply by showing that

are collinear if and only if

are distinct.) Hint: use the determinant of the matrix

a3 a 1

3
b b 1 ,
c3 c 1

and rewrite

[1 :

1
t2

1
]
t3

= [t3 : t : 1].

CHAPTER 17

Putting a nonsingular cubic in standard form

An irreducible algebraic curve


genus of its normalization

is

E P2

is an elliptic curve if the

(topologically it looks like a donut).

By the genus formula, all smooth cubic curves are elliptic. In the next
two chapters we will show not only that such a curve is isomorphic to

C/

for some lattice

but will get a description of

which shows

E . This is important, since for two dierent lattices


= Z h, i and 0 = Z h0 , 0 i, the complex 1-tori C/ and C/0

its dependence on

need not be isomorphic as Riemann surfaces. (More precisely, they are

[0 : 0 ] by an integral
1
projectivity, i.e. a transformation of P induced by A P SL2 (Z).)
Even more signicant is how we do this: by putting E in Weierstrass
isomorphic if and only if

[ : ]

is carried to

form, integrating a holomorphic form on it to get a map to a complex


torus, and showing that the Weierstrass
invert this map.

To put

-function

and its derivative

in this form, a choice of ex is required.

What is that?

17.1. Flexes

C = {F (Z, X, Y ) = 0} P2 be an irreducible algebraic curve of


degree d 3. One way of thinking of the tangent line at a nonsingular
point p C is as the unique line satisfying (C Tp C) 2.
Let

Definition 17.1.1. A smooth point

pC

is called a ex if the

intersection multiplicity

(C Tp C) 3.

TpC

C
C , and can

(see 4.4). Since C

Intuitively these are the inection points of

be seen to

correspond to cusps of the dual curve C

has nitely

187

188

17. PUTTING A NONSINGULAR CUBIC IN STANDARD FORM

many singularities, this gives one proof that there are nitely many
exes; we will however take a dierent approach.
Denoting partial derivatives by subscript, e.g.

Hessian of

FZX :=

2F
, the
ZX

is the polynomial matrix

FZZ FZX FZY

HessF = FXZ FXX FXY .


FY Z FY X FY Y

Its determinant

H := det(HessF )
3(d 2). Call HC :=
associated to C .

is clearly a homogeneous polynomial of degree

{H(Z, X, Y ) = 0} P

Lemma 17.1.2. Let

the Hessian curve

pC

be a smooth point. Then

p is a ex

p HC .
Proof. Since intersection numbers are invariant under projectiv-

p = [1 : 0 : 0], Tp C = {Y = 0}. In ane


coordinates, writing f (x, y) := F (1, x, y), this means that the curve
{f (x, y) = 0} C2 contains (0, 0) and is tangent to {y = 0}. So
f (0, 0) = 0 and (fx (0, 0), fy (0, 0)) = (, 0) where 6= 0, so that
ities, we may assume

f (x, y) = y + (ax2 + 2bxy + cy 2 ) + higher-order


Parametrizing

Tp C

by

t 7 (t, 0),

terms.

we have

(C Tp C)p = ord0 (f (t, 0)) = ord0 (at2 + h.o.t.),


which is

(yielding a ex) if and only if

Now the above form of

a = 0.

implies

F (Z, X, Y ) = Y Z d1 + (aX 2 + 2bXY + cY 2 )Z d2 +


so that

0
0 (d 1)

HessF (1, 0, 0) =
0
2a
2b
.
(d 1) 2b
2c

Taking the determinant,

H(p) = det(HessF (p)) = 2(d 1)2 2 a.


This is clearly zero (i.e.

p HC )

if and only if

a = 0.

17.1. FLEXES

Now Bezout guarantees intersections of

189

HC .

and

If

is singular

then these might all be at singular points, so that there might be no


exes (though this isn't typical: see the exercises). On the other hand,
if

is smooth then by Lemma 17.1.2 we do have exes. Rening this

observation:

Proposition 17.1.3. On a nonsingular curve

there exists at least one and at most

3d(d 2)

of degree

d 3,

exes.

Proof. By Bezout,

(C HC )p = (C HC ) = deg(C) deg(HC ) = d 3(d 2).

pCHC
So the number of points in

C HC

is between

and

3d(d 2),

all

points are smooth points, and we apply Lemma 17.1.2.

HessF is just the multivariable derivative


2
2
(Jacobian matrix) of DC : P P (4.4), the intersections of C and
HC may be viewed as degeneracies of the map DC |C : C  C . This is
referred to above.
what gives rise to the cusps in C
Remark 17.1.4. Since

Definition 17.1.5. The multiplicity of a ex

be

pC

is dened to

(C HC )p .
Now take

C = E

to be a smooth elliptic curve (d

= 3).

Then

in the proof of Lemma 17.1.2, the precise form of the homogeneous


polynomial is
(17.1.1)

F (Z, X, Y ) = Y Z 2 +(aX 2 +2bXY +cY 2 )Z+X 3 +X 2 Y +XY 2 +Y 3 .


Assume

a=0

[1 : 0 : 0]. (Note that must


in order that Y not divide F  which would make E

so that we have a ex at

then be nonzero,

reducible hence singular.) Then a short computation gives

2y
2by
2 + 2bx

HessF (1, x, y) =
2by
6x + 2y
2x + 2y + 2b .
2 + 2bx 2x + 2y + 2b 6y + 2x + 2c
Pull this back to

Tp E = {y = 0}

by making the substitution

0
0 2 + 2bt

HessF (1, t, 0) =
0
6t
0
;
2 + 2bt 0 2t + 2c

190

17. PUTTING A NONSINGULAR CUBIC IN STANDARD FORM

this has determinant

H(1, t, 0) = (2 + 2bt)2 6t,


and since

, 6= 0
(Tp E HE )p = ord0 (H(1, t, 0)) = 1.

So

HE

is smooth at

and

Tp E

is not its tangent line.

But then it

E transversely (since they have distinct tangent lines), so


(E HE )p = 1. This computation is valid at any ex of E (after a

intersects
that

projective change of coordinates, of course), and so proves:


Proposition 17.1.6. Any smooth cubic has

exes, each of mul-

tiplicity one.

deg(HE ) = 3(d 2) = 3, Bezout gives us 9 intersection points of HE and E , counted with multiplicity; and we have
demonstrated that the multiplicities are all 1.

Proof. Since

17.2. Weierstrass form


Consider an arbitrary smooth cubic curve

E = {F (Z, X, Y ) = 0} P2 .
In this section we will show that there exists a choice of projective
coordinates putting

uniquely into a convenient form. (Alternately,

you can view this as the existence of a projectivity putting

into this

form, in the same coordinates.)

E has a ex, and rst of all we can choose coordinates so


that this is at [0 : 0 : 1] =: O with TO E = {Z = 0}. To get the general
equation of such a cuve: take (17.1.1), set a = 0 (for a ex), swap Z
and Y , and (without loss of generality since 6= 0) normalize to 1;
We know

this gives

F (Z, X, Y ) = ZY 2 + (2bXZ + cZ 2 )Y + X 3 + X 2 Z + XZ 2 + Z 3 ,
with ane form

f (x, y) := F (1, x, y) = y 2 + yf2 (x) + f3 (x).


Now the discriminant

Dy (f (x, y)) = Ry (y 2 + yf2 (x) + f3 (x), 2y + f2 (x))

17.2. WEIERSTRASS FORM

191

1 f2
f3
1 f2 f3

= det 2 f2 0 = det
f2 2f3
2
f2
0 2 f2

= f22 + 4f3 = (2bx + c)2 + 4(x3 + x2 + x + )


3 since 6= 0. Roots of (Dy (f ))(x)
correspond to vertical lines x = x0 which are tangent to (the ane part
of ) E at some point. Bezout tells us that the intersection number there
can only be 2, since deg(E) = 3 and {X = x0 Z} already meets E at O .

is a polynomial in

of degree

Such rst order tangencies mean the roots each have multiplicity one.
Therefore
at

has three vertical tangents (apart from

L = {Z = 0}),

p1 , p2 , p3 .
Lemma 17.2.1. The

{pi }3i=1

are collinear.

Proof. In the picture

T E

p3
p1
q

Lp p 2
1

p2

p to be the third intersection point of Lp1 p2 and E , and q the third


intersection point of LOp with E . Consider the cubic curves C1 = E ,
C2 = LOp1 + LOp2 + LOp3 , and C3 = TO E + 2Lp1 p2 . We have
dene

C1 C2 = 3O + 2p1 + 2p22 + p + q
and

C1 C3 = 3O + 2p1 + 2p2 + 2p.


15.2, the ratio of the homogeneous polynomials dening
C2 and C3 gives a degree 1 map E P1 (which is impossible) if p 6= q .
So p = q , and LOp is tangent to E at p. It follows that p is p1 , p2 , or
Arguing as in

192

p3 .

17. PUTTING A NONSINGULAR CUBIC IN STANDARD FORM

p1 doesn't
p1 , p2 , p3 Lp1 p2 .

The rst two are impossible since the tangent to

through

p2

p = p3 .

and vice versa; so

Hence

pass

O to Lp1 p2 (
= P1 ) presents E as
1
a 2 : 1 cover of P branched over p1 , p2 , p3 , and the image TO E Lp1 p2 of
O. Furthermore Lp1 p2 , LOp1 , TO E form a triangle, and so we can choose
0
0
0
0
new projective coordinates X , Y , Z in order that Lp1 p2 = {Y = 0},
LOp1 = {X 0 = 0}, and TO E = {Z 0 = 0}. For simplicity I'll drop the
primes and just write X, Y, Z for this new coordinate system. The
Now stereographic projection from

following picture summarizes what we know:

TOE

LOp
E

p1

Y = 0 ) p1
p2 (resp. p3 ).

where (on

X
at
Z

p2

Lp p

1 2

is at

X
Z

= 0.

Write

(resp.

2 )

for the value of

We would like an equation to correspond to this picture. Now, in


the new coordinate system, the equation of

is still of the form

F (Z, X, Y ) = ZY 2 + (2bXZ + cZ 2 )Y + X 3 + X 2 Z + XZ 2 + Z 3 ,
[0 : 0 : 1] with tangent line Z = 0.
But now (referring to the picture) also [1 : 0 : 0] E , which implies
= 0. Moreover, FY (= 2Y Z + 2bXZ + cZ 2 ) = 0 at p1 = [1 : 0 : 0],
p2 = [1 : 1 : 0], and p3 = [1 : 2 : 0] since the tangents are vertical
there. This yields c = 0, then 2b1 = 2b2 = 0. As the {pi } are distinct
(so i 6= 0), we have , and
because we still have a ex at

F (Z, X, Y ) = Y 2 Z + X(X 2 + XZ + Z 2 )
= Y 2 Z + X(X 1 Z)(X 2 Z).

EXERCISES

193

Now dene new coordinates by the projective transformation

r
X=

4
1 + 2
X0 +
Z0 , Y = iY0 , Z = Z0 ,

which makes the equation

r
F (Z0 , X0 , Y0 ) = F

4
1 + 2
X0 +
Z0 , iY0 , Z0

= Y02 Z0 + 4X03 g2 X0 Z02 g3 Z03 .


Dropping the subscript

0's and taking the ane equation,

we have put

in Weierstrass form :
Proposition 17.2.2. (a) Any smooth cubic

E P2

is projectively

1
equivalent to a curve with ane equation of the form

y 2 = 4x3 g2 x g3 .
(b) This form is unique (given
4

( g2 , g3 )

where

; in particular,

j :=
is an invariant of

E ) up to a change of the form (g2 , g3 ) 7


g23
C
g23 27g32

E.

Proof. We have just seen (a).

equation

To see (b), write the projective

Y Z = 4X g2 XZ g3 Z 3 .

It is not dicult to see that

any linear transformation preserving the form of this equation (up to


rescaling) has the form

3
Z . Taking
2 0

gives exactly the

is unchanged by this

X = X0 , Y = Y0 , Z =
claimed eect on (g2 , g3 ), and j

:=

transformation.

Exercises
(1) Show that the cubic curve

P2

C = {0 = X 3 +Y 3 XY (X +Y +Z)}

has one (ODP) singular point and exactly three collinear exes.

[Hint: start by computing the Hessian, then nd the Hessian curve
and determine its intersections with

1note

that the vanishing of the

roots sum to zero

x2

C .]

term of the right-hand side indicates that its

CHAPTER 18

Canonical normalization of the Weierstrass cubic

This chapter will focus on the precise relationship between Weierstrassform elliptic curves and complex 1-tori (or equivalently, 2-lattices in

C).

E a period lattice
E . These will ulti-

We will begin by associating to a Weierstrass cubic

E ,

and to a (full) lattice

a Weierstrass cubic

mately be shown to be bijections of sets and mutual inverses. The key

-function and its derivative


Abel map u : E C/E . This

step is the inversion of the Weierstrass

P2 )

(embedding a 1-torus in

by the

map is closely related to the elliptic integral

dx
p
,
4x3 g2 x g3

a variant of which will be studied in the exercises.

18.1. Holomorphic forms on an elliptic curve


Let

be a Weierstrass cubic, viz., the projective closure of

f (x, y) := y 2 Q(x) = 0
in

P2 ,

where

Q(x) = (x e1 )(x e2 )(x e3 ) ,


e1 + e2 + e3 = 0.

dx
Claim 18.1.1. :=
1 (E) is nowhere vanishing.
y
E

Remark 18.1.2. This statement perhaps requires clarication. You

dx
may interpret
y

in either of two equivalent ways:


E
dx
2
(a) any algebraic dierential form (such as
) on C extends to a meroy
2
morphic form on P , and you can think of |E as shorthand for pullback

: E , P2 just to write dx
);
y
Y
X
(b) alternatively, writing x =
and y =
exhibits x and y as meroZ
Z
2
morphic functions on P (and hence, via pullback, on E ), and Example
d( x|E )
13.1.4 tells us that
is a meromorphic 1-form.
y|E
1
Either way, we have K (E); and part of the content of the Claim is
to

(rather than introducing

195

196

18. CANONICAL NORMALIZATION OF THE WEIERSTRASS CUBIC

that

is actually holomorphic:

vanishing statement says that

p () 0 for all p E . The


actually p () = 0 for all p.

Proof. Look at the ane part

E\O.

Wherever

nowhere

fy 6= 0,

so that

dx
gives a local coordinate,
y

is holomorphic and nonvanishing. We


E
precisely at the three points {(ei , 0)}i=1,2,3 ,

f = 0 and fy = 0
0
where fx = Q (ei ) 6= 0 so that y is a

dx
0
0 = df = 2ydy Q (x)dx so that y
have

local coordinate. On

we have


dy
2 0 ,
Q (x) E
which is evidently nonvanishing and holomorphic in a neighborhood of
each

(ei , 0).
O = [0 : 0 : 1]? By PoincarpE p () = 2g 2 = 0, so that if p () = 0 for all
be no contribution from O either.


What about the (ex) point at innity

g = 1 =
p E\O, there can

Hopf,

1 (E)
= C hi.
multiple of .

Corollary 18.1.3.

1-form

on

is a

That is, every holomorphic

0 1 (E), the discussion preceding Example


0
K(E). But since is nowhere vanishing, 0 is
13.1.2 tells us

actually a holomorphic function. Now use Liouville's theorem (O(E)


=
C).

Proof. For any

Amongst the standard topological invariants of a 1-manifold

its rst homology group. An ad hoc denition is

(
H1 (M, Z) := (

free abelian group generated by


closed piecewise-C

paths on

subgroup generated by
boundaries of nitely triangulable regions

or simply cycles modulo boundaries. From the picture

),

is

18.1. HOLOMORPHIC FORMS ON AN ELLIPTIC CURVE

197

it isn't hard to convince yourself that

H1 (E, Z)
= Z h, i .
That is, for any closed
(with boundary

E , there exists a closed set E

path

such that

= m + n + .
The integers

m, n

are uniquely determined by

+m

+n

=m

+n

One then has

by Cauchy's theorem (Prop. 13.1.5). The values of the integrals

, and we dene

, C.

over cycles are called the periods of

E := Z

the period lattice

This furnishes an invariant of the complex structure on

E which, unlike

the topological invariant, actually distinguishes elliptic curves which


are non-isomorphic as complex manifolds (or algebraic curves).
Remark 18.1.4. Given a lattice of the form

we have a Weierstrass

Z h1 , 2 i =: C,

P -map
P

C/ P2
u 7 [1 : (u) : 0 (u)]
whose image (by Exercise 5 of Chapter 7) is a Weierstrass cubic! Dene

E := P(C/),
which we henceforth consider to be the range of the map

P.

Obviously

it is of interest to nd out whether all Weierstrass cubics arise in this


fashion (as

1algebraic

E 's).

geometers call this a transcendental invariant

198

18. CANONICAL NORMALIZATION OF THE WEIERSTRASS CUBIC

P is injective. Its composi1


tion with (the x-coordinate projection) x : E P has degree 2 since
has a unique pole on C/ (at 0), which is a double pole. But since
Before moving on we should note that

mapping degrees of Riemann surfaces multiply under composition, and


the degree of

itself is

2,

that of

P : C/ E

must be

1.

18.2. The Abel map


Let

E = {y 2 = 4x3 g2 x g3 } P2
|
{z
}
Q(x)

be a Weierstrass cubic with

dx
y

1 (E).

Integrating this gives

E
a (holomorphic) map of Riemann surfaces

u : E C/E
p
p 7

O
where the integration is over any

map is well-dened: if

0 , 00

path from

to

p.

This Abel

are two such paths, then their dierence

is closed and so

0 00 = + m + n.
Integrating, we have

= m

00

E .

+n

E has rank 2 (something


2
we haven't yet addressed). Otherwise, C/E is isomorphic to C or C ,
both of which are noncompact. Since E is compact and u is continuous,
its image must be a compact submanifold of C/E . If the latter is
noncompact then the image is therefore a point, meaning u is constant,
clearly false since 6= 0.
Remark 18.2.1. It is now easy to see that

A baby version of Abel's theorem for elliptic curves

3 is then:

Theorem 18.2.2. The Abel map is injective.

=
2

E = Z (rank 1) = C/E = C/2iZ C (by


3The real theorem, which we will deal with later,
divisors.

taking

exp).

has to be phrased in terms of

18.2. THE ABEL MAP

u(p) u(q)

Proof. (Sketch) Suppose

E;

then

mod

for

p 6= q

points of

= u(p) u(q) E .

199

Modifying the path from

to

get

m + n

by

(for some

m, n Z),

we

= 0.
p
Dirichlet's existence theorem (which we won't prove) guarantees the

0 K1 (E) with only simple poles, only at p and q , with


Resp (0 ) = Resq (0 ) = 1. This is true for any two (distinct) points p
and q , and has nothing to do with our assumption (that u(p) = u(q)).

existence of

Now referring to the picture

we have

H1 (E\({p} {q}), Z)
= Z h, , i

(18.2.1)

where

0 = 2i.

Next, normalize

0 ,

putting



0

:= 0
,

which has the same residues as

0 .

Observe that

= 2i,

while

= 0.

Cutting open the above gure along


domain

4In

and

yields the fundamental

4
(the yellow region):

order to accomodate the path from

by an integer factor

M.

to

p,

it may be necessary to dilate

(You can think of this as the fundamental domain of a

200

18. CANONICAL NORMALIZATION OF THE WEIERSTRASS CUBIC

pat

q
O
On the interior of

F, U :=

gives a holomorphic function which is

continuous on the boundary. Now

= U(p) U(q)

0=
p
which by the Residue theorem

Noting that

(resp.

  to  ), this

1
=
2i

where

6= 0.

U .
F

) is the change in U from  

Hence,

1
=
2i

1
=
2i

 

to   (resp.


,

= 0.

By (18.2.1), any closed path on

n + m + ` ; and
` = 2i`. Consequently,

of the form

so

E\({p} {q}) is, up to boundaries,


the integral of over such a path is


F := exp
O

{p} {q}. Let


z (resp. w) be a local coordinate about p (resp. q) with z(p) = 0
dz
(resp. w(q) = 0). We know that the leading term of at p is
, and
z
dw
at q is
. This makes F locally at p (resp. q ) the product of a
w
dz
dw
log z
= z (resp. e w = w1 ), so that
holomorphic function by e z = e

is a well-dened function on

by

which is holomorphic o

E .) This doesn't
M + M M M .

nite unbranched covering of


replacing

really aect the proof, except for

18.3. ABEL INVERTS WEIERSTRASS

is meromorphic on

201

with divisor

(F ) = [p] [q].
Therefore

deg(F ) = 1,

making

F : E P1

an isomorphism, which is

impossible.
We conclude from this contradiction that

p and q

cannot have been

distinct.

Abel's theorem is usually paired with something called Jacobi inversion, the baby version of which is:
Proposition 18.2.3. The Abel map

is surjective (and thus an

isomorphism).
Proof. This is trivial:

is a closed mapping (since continuous),

and an open mapping (since holomorphic and not constant). The image
is therefore open and closed in

C/E ;

since the latter is connected,

we're done.

Essentially all of the foregoing (with the exception of Remark 18.1.4)


works for any nonsingular cubic. There is a unique holomorphic

1-form

up to scale; it vanishes nowhere; and integrating it from a base point


gives an isomorphism from the cubic to a complex 1-torus. This follows
from the last 2 sections by applying the projective transformation of
Chapter 17 to put the cubic in Weierstrass form (which has just been
slightly more convenient for writing down

).

For the next section,

however, the Weierstrass form will be crucial.

18.3. Abel inverts Weierstrass


We now make the Big Claim that
(18.3.1)

a Weierstrass cubic is always the image

(cf. Remark 18.1.4), and we have


(18.3.2)

u P = idC/

and
(18.3.3)

P u = idE .

of a Weierstrass

P -map

202

18. CANONICAL NORMALIZATION OF THE WEIERSTRASS CUBIC

First we study the case where

is (by assumption) the image of a

P -map.
= Z h1 , 2 i C

Proposition 18.3.1. Let

be a lattice.

The

composition

(
(
=)
=)
C/ E C/E
u
P
is the identity.
Proof. Obviously part of the claim is that


 



E :=
H1 (E , Z)
= .

(18.3.4)

For

E ,

(x, y)

not a lot is lost by working in ane

there is only

at

coordinates, since

and we know that corresponds to

u0

on the

complex 1-tori. (Note also that  u is used both as the Abel map and
as the coordinate on
Since

C; which one will be clear from the context.)


P(u) = ((u), 0 (u)),
!

dx
0 (u)du
d((u))

P = P
=
= du.
=
y E
0 (u)
0 (u)
P is an isomorphism
H1 (C/, Z)

Moreover, that
image of some

means any cycle

on

is the

so that

=
gives a bijection between

P(u0 )

u(P(u0 )) =
O

P =

=
E

dx
=
y

P (
)

u0 C/,

and

hence (18.3.1). So then taking

P(u0 )

P =

=
P(0)

du

u0

du = u0
0

18.3. ABEL INVERTS WEIERSTRASS

203

proves the Proposition.

Now let

be any Weierstrass cubic.

Proposition 18.3.2. The composition

(
(
=)
=)
E C/E EE P2
P
u
is the identity. In fact,

E = EE

(18.3.5)

exactly as subsets of

P2 .
(x, y) E :

Proof. Here is what the composition looks like, where

(x, y) 7

dx
P
p
7
Q(x)

dx
p
Q(x)

, 0

dx
p
Q(x)

is determined by y . We must show that the right-hand side


recovers (x, y), or equivalently that the inverse (x(u), y(u)) : C/E
E of the Abel map u identies with ((u), 0 (u)).
Let's start with x, and compare the elliptic functions x(u) and P(u)
on C/E . First I claim that both have double poles at u = 0: you
1
already know that (u) = 2 + higher-order terms. For x, it suces to
u
check this on E , using

X
x = K(E)
Z E

where the

{Z=0}=TOE

X=0

. . . which is easy:

O (x) = (E {X = 0})O (E {Z = 0})O = 1 3 = 2.

!!
,

204

18. CANONICAL NORMALIZATION OF THE WEIERSTRASS CUBIC

Now

x(u) =

A
u2

+ h.o.t.,

with

A=

x dw 2

Q(w)
2

lim x(u) u = lim x u(x) = lim


u0
x
x
1/ x


lim

!2
p
1/ Q(x)
1/2x

3
2

4x3
= 1.
x Q(x)

= lim

Dene an involution

:EE
by

(x, y) 7 (x, y);


p E,
(p)
p
p
dx
dx
dx
u((p)) =

=
=
= u(p),
y
O=(O) y
O
O y

this xes

O.

For

and so

y(u) = y(u).

x(u) = x(u) ,

x(u) and (u) are both even -periodic


1
functions locally of the form 2 + h.o.t., and so their dierence has no
u
poles and must (by Liouville) be constant: x = c.
x dx
du
Next, dierentiating u =
gives
= y1 , or
dx
y
All told, we now have that

x0 (u) =

dx
= y(u);
du

and then

d
(c) = x0 (u) 0 (u) = y(u) 0 (u).
du
All that is left is to check that c = 0.
The xed points of the involution u 7 u are the 2-torsion
i.e. those u C/E with 2u 0
0=

u3

O
since we must have
(by

u)

u u

u1

P)

E . These are, of course, the images


in E , since u = u. They also
points of (x, y) 7 (x, y) in EE , since

mod

of the xed points of

must map (by

points,

to the xed

EXERCISES

((u), 0 (u)) = ((u), 0 (u)).

205

Writing

y 2 = 4x3 g2 x g3 = 4(x e1 )(x e2 )(x e3 )


for the equation of

E,
O

e2

e1

x (2-torsion
Similarly, if

e3

points)

= e1 , e2 , e3 , .

EE = {y 2 = 4x3 g2 x g3 = 4(x e1 )(x e2 )(x e3 )}

then

(2-torsion

points)

= e1 , e2 , e3 , ;

and clearly

e1 + e2 + e3 = e1 + e2 + e3 = 0.
Since

(u) = x(u) + c,
(u1 ) + (u2 ) + (u3 ) = x(u1 ) + x(u2 ) + x(u3 ) + 3c

which becomes

0 = 0 + 3c
so

c = 0.
We conclude that

(u(x, y)) = x

and

0 (u(x, y)) = y .

Exercises
(1) [Adapted from Silverman-Tate.]
Let

0 < ,

and consider the ellipse

dened by

y2
x2
+
= 1.
2 2
E is given by the
p
1 k 2 sin2 d

Show that the arc-length of

4
0

integral

206

18. CANONICAL NORMALIZATION OF THE WEIERSTRASS CUBIC

for an appropriate choice of constant

depending on

and

Then prove that this is equal to

4
0

1 k 2 t2
p
dt.
(1 t2 )(1 k 2 t2 )

hence the problem of determining arc-length comes down to evaluating the integral

on the elliptic curve

Remark:

1 k 2 t2
dt
u
0
u2 = (1 t2 )(1 k 2 t2 ).

this curve obviously isn't cubic but is still of genus 1

because of the singularity at innity. You can think of this as what


you get if instead of writing an elliptic curve as
branching at (i.e.

2:1

over

P1

with

order 2 ramication over) 3 nite points plus

innity, you take the branching to be over 4 nite points. So it's


very close to a Weierstrass cubic, even though it's quartic.

The integral in this problem is of course related to the Abel map/abelian


integral above, and is meant to demonstrate why such integrals
(and hence these curves) are called elliptic.

CHAPTER 19

Group law on the nonsingular cubic

It is now high time for the smooth version of Chapter 16: group
laws and addition theorems for elliptic curves. We start by introducing
an algebro-geometrically dened binary operation on the points of a
cubic curve, and prove it coincides with addition on

C/E

under the

Abel isomorphism. This gives one proof that the operation denes an
abelian group, and we give another more natural one as well.

From

the fact that it makes Abel's map into a homomorphism we will then
derive functional equations for elliptic integrals.

19.1. Denition of the group law


Let

E P2

be a nonsingular cubic, which we shall not require to

be in Weierstrass form, and x a ex

Step 1: Draw the line

L1

O.

Let

and

q:

through

and

be points of

E.

O
p

L1
p*q

q
By Bezout, there is a third intersection point, which we shall denote

p q,

so that

L1 E = [p] + [q] + [p q].


Note that the three points need not be distinct  any two, or all three,
may coincide.

This has the usual interpretation: double-intersection


207

208

19. GROUP LAW ON THE NONSINGULAR CUBIC

means

L1

is tangent to

at that point; triple-intersection means you

have a ex.

Step 2: Draw the line

L2

through

and

p q:

O
p

L2
p*q

p+q

The third intersection point is denoted

p + q,

with the same interpre-

tations as above.
In the special case where

is in Weierstrass form,

{X = x0 Z}. So the map sending p q


: E E taking (x, y) 7 (x, y):
line

to

p+q

L2

is a vertical

is just the involution

E
j
e1

e2

e3

We have therefore constructed a binary operation

EE E
(p, q) 7 p + q
on the (set consisting of the) points of

E.

It is not yet clear that this

denes a group. It is clear that it is commutative, so that if it denes


a group, then that group is abelian.

19.2. RELATION TO THE GROUP STRUCTURE ON THE 1-TORUS

209

19.2. Relation to the group structure on the 1-torus


Let

T : P2 P2

be the projective transformation putting

into

O0 := [0 : 0 : 1]). Denote the


0
binary operation dened on points of E := T (E) (via the method just
0
0
described, with O replacing O ) by + . Since projectivities preserve
0
lines, intersection multiplicities, and so forth, it is clear that T (p) +
T (q) = T (p+q). So to show that  + denes a group law for arbitrary
E , it suces to check this for Weierstrass cubics.
Hence we may assume E is in Weierstrass form (and O = [0 : 0 : 1]).

Weierstrass form (and taking

to

Take the Abel map

u : E C/E
to be as in Chapter 18, with inverse

E 

denes a group law on

our operation on

C/E .

P.

We know that addition mod

The next result implies not only that

E denes a group law, it says that u is an isomorphism

of groups.

Theorem 19.2.1. The Abel map respects binary operations. That

is, the equivalent formulas

P(u1 ) + P(u2 ) = P(u1 + u2 )

(19.2.1)

u(p) + u(q) u(p + q)

(19.2.2)

mod

hold.

Proof. We will prove (19.2.2), in a manner reminiscent of the

proof of Theorem 18.2.1.


polynomial dening

Li ,

Writing

FLi

for the degree-1 homogeneous

consider the meromorphic function


FL2
f :=
K(E) .
FL1 E
Reading its divisor o from the intersection points of the

{Li }

(f ) = [p + q] + [p q] + [O] ([p] + [q] + [p q])


= [p + q] [p] [q] + [O].
So for its pullback

P f = f P

to

C/E ,

(P f ) = [u(p + q)] [u(p)] [u(q)] + [0].

and

E,

210

19. GROUP LAW ON THE NONSINGULAR CUBIC

Cut open
region

C/E

and put

U :=

F:

du

on the resulting fundamental

u(p+q)
u(q)

u(p)

Write

U(p)

for

U(u(p))

(and so on) for simplicity. By the residue theo-

rem,

U(p + q) U(p) U(q) (+0)

1
d(P f )
=
U
2i F
P f





1
1

dlog(P f )
du +
dlog(P f )
du.
=
2i
2i

Since both terms in braces are integers, the whole thing belongs to


du = E .

du,

U(p), U(q), U(p + q) are lifts to C of u(p), u(q), u(p + q),


modulo E we see that u(p + q) u(p) u(q) 0.

Since

going

We will generalize this argument in the next chapter to get Abel's


theorem for

E.

Remark 19.2.2. For an arbitrary smooth cubic

to scale) a unique

(C),

u : C C/(C,) .

C , one still has (up

which gives rise to an Abel isomorphism

(19.2.2) still holds in this setting by more or less

the same proof; this avoids passing through Weierstrass form.

19.3. A more algebro-geometric approach


Returning to the setup of

19.1,

let us suppose that the coecients

of the homogeneous cubic polynomial dening


eld

k C.

extension of

of

We shall say

(e.g.

itself, or

belong to some sub-

K C

is a eld

then we can consider the

K -points

is dened over

C),

E
k.

If

E
E(K) := {[Z : X : Y ] E | Z, X, Y K}.
Proposition 19.3.1.

quently a subgroup of

E.

E(K)

is closed under  +, and is conse-

19.3. A MORE ALGEBRO-GEOMETRIC APPROACH

211

p, q E(K)

then the line L1 = Lpq is dened over K ,

1 =
and so can be parametrized P Lpq over K . (This means that the
Proof. If

formulas expressing
ecients in

[Z : X : Y ] as functions of [T0 : T1 ] P1 involve co-

K .)

So the pullback of the homogeneous polynomial denP3


3j j
is dened over K . Now this can be written
j=0 j Z0 Z1 =

ing E
Q3
i=1 (Z1

i Z0 ), assuming without loss of generality that there are no


Z0 factors; and what we know is that the j K . This is no guarantee
that the i K . But in this case we know that two of them  say,
1 , 2  correspond to p, q and so must belong to K . Consequently
3 K as well, and its image point p q is also dened over K . Repeat
the argument for L2 and the claim follows.

Remark 19.3.2. (a) In light of the above denition, the correct

notation for the set of complex points of

E,

which we have heretofore

denoted simply  E , is

E(C).
(b) As a C-module, E(C)
= C/E has rank 1, but as an abelian
group (i.e.  Z-module), its rank is innite. (Consider a bunch of Qlinearly independent complex numbers modulo E  there is no bound
on the size of the Q-vector space you can generate in this fashionl.) On
the other hand, for the subgroup E(Q), a famous theorem of Mordell
(1922) asserts that the rank (as an abelian group) is always nite. In
problem (6) below, you will show computationally that the rank of

E(Q)

in one example is at least

Now the Abel map

1.

u is non-algebraic (i.e., transcendental); it should

be seen as providing a link between the complex algebraic and the complex analytic. Such maps, which include multivariate abelian functions,
modular and automorphic forms, are very important in arithmetic algebraic geometry.

While they do not preserve the eld of denition,

they have nonetheless been essential to the study of things like class
eld theory, the proof of Fermat's last theorem, and so on.
light of the subgroup structures

E(K) E(C),

Still, in

it is a bit sloppy to

prove that  + is a group law in a manner that only works over

C.

So we will now give the fully algebraic approach to this proof, by


checking
(19.3.1)

O + p = p ( p E),

212

19. GROUP LAW ON THE NONSINGULAR CUBIC

(p q) + (p + q) = O ( p, q E),

(19.3.2)

+ is

(19.3.3)

Notice that (19.3.2) implies

p O = p,

and

associative.

(p O) + (p + O) = O,

and thus that

so we indeed have inverses. (19.3.1) and (19.3.3) are the

other two group axioms.


To verify (19.3.1), we have the pictorial depiction of the two-step
process for adding

and

p:

O*p

O*p

L2

L1
E

O+p

Step II

Step I

L1 = LOp and L2 = LO,Op = L1 !! Consequently the third


intersection point O + p of L2 with E , is none other than p. (19.3.2)
in which

isn't much harder; again, a picture:

O =(p*q)*(p+q)

L2
p*q

L1
q
p+q
L1 = Lpq, p+q . Since the line through O and p q intersects
(and denes) p + q , it follows that L1 's third intersection point

where

in

19.3. A MORE ALGEBRO-GEOMETRIC APPROACH

(p q) (p + q)

with

is just

O.

213

Then

L2 := LO,(pq)(p+q) = LO,O = TO E,
and since
again

is a ex ((TO E

E)O = 3),

the third intersection point is

O.

Finally we come to the associativity issue (19.3.3).

Here I won't

break the two steps up into two pictures. Instead, here is a depiction
of

(p + q) + r,

the addition of

where the blue lines compute

and the green ones

to the result:

L1

p+q

L1

r
p

(p+q)*r
p*q

q
p+q

(p+q)+r
L2

L2

...

p + (q + r) looks
p):
l1
l1
r
p

and here is what

for adding result to

like (blue lines for

q + r,

p*(q+r)

q*r

l2
q
p+(q+r)

q+r
l2
We have to show

(p + q) + r = p + (q + r),

or equivalently

(p + q) r = p (q + r).

green

214

19. GROUP LAW ON THE NONSINGULAR CUBIC

Now look at the three cubics

E , C = L1 `2 L01 , and D = `1 L2 `01 ,

with intersections

E C = [O] + [p] + [q] + [r] + [p q] + [p + q] + [q r] + [q + r] + [(p + q) r]


and

E D = [O]+[p]+[q]+[r]+[pq]+[p+q]+[q r]+[q +r]+[p(q +r)].


Argue la
and

15.2:

the ratio of the homogeneous polynomials dening

induces a meromorphic function on

with divisor

E C E D = [(p + q) r] [p (q + r)],
leading as usual to a contradiction unless these two points are the same.

19.4. Addition theorems


Now assume

E = {y 2 = Q(x)}

(with

Q(x) = 4x3 g2 x g3 )

is in

Weierstrass form; we would like to unwind the statements (19.2.1) and


(19.2.2) that

and

u are group homomorphisms (hence isomorphisms),

to produce something more computationally explicit.


We do this rst for the Weierstrass map.

((u1 ), (u1 )), q = P(u2 ) = ((u2 ), (u2 )),

p = P(u1 ) =
pq

Writing

we have

= (p+q) = (P(u1 )+P(u2 )) = (P(u1 +u2 )) = ((u1 +u2 ), 0 (u1 +u2 ))


= ((u1 + u2 ), 0 (u1 + u2 )).
Now

p, q ,

and

pq

are collinear by construction  they lie on

in the group law process for

E.

L1

We may express this in projective

coordinates by saying that

1
1
1

0 = det (u1 ) (u2 ) (u1 + u2 ) .


0 (u1 ) 0 (u2 ) 0 (u1 + u2 )
This is the rst addition theorem, and is the analogue for bi-periodic
functions of the standard trigonometric angle-addition formulas.
really does express

It

(u1 +u2 ) in terms of (u1 ) and (u2 ), since () =

p
Q(()).
Let's actually compute the group law on

ax + b for L1

E.

Start by writing

and substituting this into the equation of

y=

to intersect

19.4. ADDITION THEOREMS

215

them. This gives

0 = 4x3 g2 x g3 (ax + b)2 = 4(x x(p))(x x(q))(x x(p + q))


since

L1

and

meet in

p, q, p q .

(Note that

x(p + q) = x(p q).)1

From expanding these two expressions and comparing coecients of

x2 ,

one nds that

a2 = 4(x(p) + x(q) + x(p + q));


and since

is the slope of

L1

it is obvious that

a=

y(q) y(p)
.
x(q) x(p)

Therefore we have


2
1 y(q) y(p)
x(p) x(q).
x(p + q) =
4 x(q) x(p)
p dx
x(p) dx
x(q) dx

Now u(p) =
=
, similarly u(q) =
O y

Q(x)
Q(x)
x(p+q) dx

u(p + q) =
. Re-expressing

and

Q(x)

u(p) + u(q) u(p + q)

mod

using all these formulas yields the second addition theorem :


(
)

x2
1 Q(x2 ) Q(x1 ) 2 x2 x1
x1
4
x x

dx
p
+
Q(x)

dx
p

Q(x) E

dx
p
Q(x)

dx

which is a nontrivial functional equation for the elliptic integral


.

Q(x)

The problems below (with the exception of the last one) take place on a

E := {y 2 = 4x3 g2 xg3 } P2 in Weierstrass form,


dx
with base point O = [0 : 0 : 1], holomorphic form =
| 1 (E),
y E
p
and Abel map u : E C/E , u(p) =
(recall this is an isoO
0
morphism), with inverse P(u) = [1 : (u) : (u)]. I have written

nonsingular cubic

1the

x(p) = (u1 ), y(p) = 0 (u1 ); x(q) = (u2 ),


y(q) = (u2 ); x(p + q) = (u1 + u2 ), y(p + q) = 0 (u1 + u2 ).
dictionary we have in mind is:

216

19. GROUP LAW ON THE NONSINGULAR CUBIC

everything in ane form, which you can convert to projective coordinates if needed.

Exercises

C/E correspond to the x3


intercepts {(ei , 0)}i=1 and the point O .
Assume g2 , g3 Q. Given p, q E(Q) (i.e. the x, y coordinates are
rational), give another proof that p + q E(Q), using the addition

(1) Show that the 2-torsion points on

(2)

theorems.

g2 = 4, g3 = 0.
Consider the complex analytic automorphism A : E E sending
(x, y) 7 (x, iy). In Ch. 13 Exercise (1), you showed that A =
i .

(a) Find A u (i.e. compute u A).


(b) Prove that iE = E . (In fact, E is a square lattice  so

(3) For this and the following three problems take

this is a very special elliptic curve!)


(4) Special case of the 2nd addition theorem (or rather, what we did
in

19.4

above doesn't exactly work in the case we'll do here, so

(2u) in terms of
(u), for E as in exercise (3), i.e. with equation y 2 = 4x3 + 4x.
0
0
[Hint: write (u) and then the slope a of E at ((u), (u)), in
terms of (u). Write y = ax + b for the line tangent to E at this
3
2
point. Then factor 4x + 4x (ax + b) into linear factors (what are
you'll have to work it out from scratch): write

the roots?), multiply out both expressions, and compare coecients


of

x2 .]

(5) Continuing the last problem, show that


point.

(1, 2 2)

is a

4-torsion

Use the CM recalled in exercise (3) to get three more

4-torsion

points. Can you use the group law to nd them all? (If

not, why?)
(6) This problem also depends on (4). Consider a point

of

with

x-coordinate x0 = 2a q , where the fraction is written in


lowest terms, a is an odd natural number and p and q are odd
integers. Show that P is of innite order (in the group). [Hint:
write (x0 , y0 ) for this point, and let (x1 , y1 ) := 2(x0 , y0 ) under the
group law; if x0 = (u), then x1 = (2u). So rewriting your

rational

EXERCISES

formula from (4) as a formula for

in terms of

x0

and simplifying,

a. Then suppose
the starting point was an N -torsion point for some N and produce
show that

x1

x1

217

is of the same form, but with larger

a contradiction via the pigeonhole principle.]


(7) This problem takes place on an arbitrary nonsingular cubic

P , with base point O = choice of ex in E , holomorphic form


1 (E) with periods generating a lattice E , and Abel map
p

=
u : E C/E , u(p) = O .
Prove that the 3-torsion points on C/ correspond (under u) to
the exes on E .

CHAPTER 20

Abel's theorem for elliptic curves

Given a divisor

D=

ni [pi ] on an elliptic curve E , we can formally

compute the sum in the group law, ending up with a single point on

E.

It seems of interest to ask if anything special is true if this point

is the origin

O.

In fact, assuming

is true precisely if

ni = 0 ,

it will turn out that this

is the divisor of a meromorphic function on the

curve. We begin by describing the statement of Abel's theorem for a


curve of arbitrary genus (which does not have a group law), to place
the statement for genus one in a broader context. Then we prove the
genus-one case, introducing theta functions along the way.

20.1. The Jacobian of an algebraic curve


Let

be a Riemann surface of genus

g.

We will need to accept

some facts in order to state Abel's theorem for

M.

(These will be

returned to in later chapters, along with the proof of Abel.) It turns


out that the space of holomorphic
abelian group of
has rank

2g .

1-cycles

1-forms

has dimension

modulo boundaries (cf.

18.1

g,

whilst the

for denitions)

In terms of bases,

H1 (M, Z)
= Z h1 , . . . , 2g i ,
1 (M )
= C h1 , . . . , g i .
Remark 20.1.1. A visual explanation of the statement about ho-

mology groups may be the best one:

g+2

g+1

2g
g

Chapter 3 of Griths's Introduction to Algebraic Curves gives one


219

220

20. ABEL'S THEOREM FOR ELLIPTIC CURVES

approach to computing the holomorphic forms of

M,

provided one be-

lieves that any Riemann surface is the normalization of an algebraic


curve

in

P2

with only ordinary double point (if any) singularities.

(This statement relies on the existence of nonconstant meromorphic


functions on M , which is nontrivial.)
Since the genus g of M is
(d1)(d2)
(d = deg(C), =# of ODPs), it is enough to show
2

24.1) and furthermore


2
that holomorphic pullbacks of rational 1-forms from P span a space of

d1
dimension
. This is done in Griths, and will be discussed a
2
little more in 24.2.
Just to get an idea of how this works, suppose C = {F (Z, X, Y ) =
0} is smooth of degree d, and recall S3m denotes degree-m homogeneous

m+2
polynomials in 3 variables, with dimension
. If G is a homoge2
neous polynomial of degree n, write g(x, y) = G(1, x, y) (and similarly
f (x, y) = F (1, x, y)). Then the meromorphic 1-form on P2 which in
gdx
ane coordinates takes the form
, restricts to a holomorphic 1-form
fy
1
on C precisely if n = d 3. (This is equivalent to saying deg(g)

 (d1)(d2)
d1
=
=
d 3.) Hence,2 1 (C) has dimension (d3)+2
.
2
2
2
that all meromorphic

Anyhow, let

1-forms

are rational (cf.

j H1 (M, Z)

period vector

be a basis element; associated to it is a

j :=

j 1
.
.
.

j
Together these form a

g 2g

Cg .

period matrix

R-linearly independent columns. (This isn't obvious, and will be addressed in 24.2.)
g
Hence their columns generate (over Z) a 2g -lattice M C (
= R2g ).
Recall that if V is a vector space (say, over C) then the dual space

is the space of linear functions V


:= Hom(V, C).
Definition 20.1.2. The Jacobian of

with

is the abelian group

J(M ) :=
1the

(1 (M ))
,
image {H1 (M, Z)}

computation in Griths proving this is ugly but straightforward; Poincar

residues facilitate a conceptual and essentially 1-line proof (but at the cost of more
sophisticated machinery).

2putting

o to

24.2

that this formula encompasses all rational holomorphic forms

20.2. THE ABEL-JACOBI MAP

221

where the denominator means the linear functions on

1 (M )

obtained

1 (M ) over 1-cycles. Evaluation of linear functions


basis {1 , . . . , g } induces an isomorphism

by integrating
against the

J(M )
that is, the Jacobian is a complex

Cg
;
M

g -torus.

Lemma 20.1.3. Any morphism of complex manifolds

Cg /M

: P1

is constant.

Cg /M has g independent holomorphic 1-forms: du1 , . . . , dug


g

(where u1 , . . . , ug are just the coordinates on C ). Since (dui )


1 (P1 ) and 1 (P1 ) = {0}, we have
Proof.

0 = (dui ) = d( ui )
locally

ui = ui
i = 1, . . . , g .

which implies
each

(well-dened only locally) is constant for

20.2. The Abel-Jacobi map

(f ), for some nontrivial mero


morphic function f on M ? Since deg((f )) = 0 for any f K(M ) , it
is clear that D must be of degree 0  i.e. in the kernel of
When is

Div(M ) of the form

deg : Div(M ) Z
X
X
ni [pi ] 7
ni .
So consider a divisor

in
Div

(M ) := ker(deg).

We may write

D=

([qj ] [rj ]) =

j j

|
where   means topological boundary and
to

{z

=:

r
j qj

qj .
Definition 20.2.1. The Abel-Jacobi map

AJ :

r
q

Div

(M ) J(M )

}
is a

path from

rj

222

20. ABEL'S THEOREM FOR ELLIPTIC CURVES

sends

D (= )

to

viewed as a functional on

qj

rj

1 (M ).

The rst question that arises is whether this is even well-dened,


which in this case means independent of the choice of 1-chain (sum

of paths)

To check this, let

meaning that

= D = 0 .

is a 1-cycle hence represents a

Consequently,

belongs to the denominator of

AJ

=
0

( 0 ) = 0,
class in H1 (M, Z).

Then

J(M ).

It's even easier to check that

is a homomorphism (of abelian groups), which is left to you.


Now suppose

D = (f ),

and consider the family of divisors

Dt := f 1 (t) Div(M ),
parametrized by

t P1 .

Then

D = D0 D ,

and the composition

AJ

P1 Div0 (M ) J(M )
sending

t 7 D0 Dt 7 AJ(D0 Dt )
is constant by Lemma 20.1.3, and zero at

t = 0.

Thus

AJ(D) = 0, and

we observe that

AJ

0
factors through Pic (M )

:=

(M )
(K(M ) )
Div

in a well-dened fashion. (The denominator means divisors of meromorphic functions, and the statement is simply that
Pic

(M )

is called the Picard group of

AJ

kills these.)

M.

Theorem 20.2.2. [Abel's Theorem]

AJ :

Pic

(M ) J(M )

is

an isomorphism.

Leaving aside the surjectivity part of this, the meaning of the welldenedness + injectivity of this map is that for

(for

D = (f )

some f K(M ) )

D Div0 (M ),

AJ(D) 0

mod

M ,

20.3. DIRECT PROOF OF ABEL FOR GENUS ONE

223

completely answering the question we asked at the outset. (Note that


the forward implication [

] is just well-denedness, which is com-

pletely proved. What is nontrivial is the injectivity/backward implication, since you actually have to nd some

having

as its divisor.)

1 case,
1
i.e. for M = E (the normalization of ) an elliptic curve. Let (E)
P
0
be nonzero, and consider D Div (E). We can write D =
ni [pi ]
P
with
ni = 0, and
X

X

X pi
X
AJ
ni [pi ] = AJ
ni ([pi ] [O]) =
ni
=
ni u(pi )
Example 20.2.3. We consider what this means in the genus-

O
where

C/E ,

is the Abel map. Here the right-hand sum is taking place in

and we see right away that

AJ

X

ni [pi ] = 0

ni u(pi ) 0.
E

By Abel's theorem (on the left) and the fact that

(C/E , +)

u : (E, +)

is a group-isomorphism (on the right), we have that

(20.2.1)
for

ni [pi ] = (f )

some f K(E)

ni pi = O

in the group law on

E(C).

As above, the forward implication has been proved.

20.3. Direct proof of Abel for genus one


In this section we will deduce a result equivalent to the backward
implication in (20.2.1), recasting it as an existence theorem for elliptic
functions. For simplicity take

= Z h1, i, H

+1
2

(upper half-plane):

1+

P
mj = 0 and
mj uj 0 mod .
P
Then, writing D :=
mj [uj ] Div(C/), there exists g K(C/)
such that (g) = D . (You may think of g as a -periodic meromorphic
function on C.)
Theorem 20.3.1. Suppose

224

20. ABEL'S THEOREM FOR ELLIPTIC CURVES

Proof. Introduce the theta function (on

(u) :=

C)

2 +2nu}

ei{n

nZ

-periodic, it has the properties


(a) (u) = (u)
(b) (u + 1) = (u) [cf. Exercises]
2i( 2 +u)
(c) (u + ) = e
(u). To check this, write ( + u)
X
X
2
2
=
ei{n +2nu+2n } =
ei{(n+1) +2(n+1)u 2u}

While it is not

nZ

nZ

m = n + 1,
X
2
ei(m +2mu)
= ei 2iu

which becomes, reindexing by

mZ
as required.
(d)

has a simple (order

1)

zero at

+1
and no-where else in the
2

fundamental domain bounded by vertices

0, 1, , 1 + .

Now consider

m
Y 
+1 j
;
f (u) :=
u uj +
2
j
clearly

f (u + 1) = f (u)

by property (b); but also (using property (c))


 !mj
u uj + +1
+

2

u uj + +1
2

Y
mj
1
=
e2i( + 2 +uuj )

f (u + ) Y
=
f (u)
j

j
1

= e2i( + 2 +u) mj e2i mj uj .


P
P
mj = 0 and
mj uj = M + N ,

By asssumption,
pression equals

2iN

so the last ex-

. The function

g(u) := e2iN u f (u)


will therefore satisfy
and the denition of

(g) =

g(u + ) = g(u) = g(u + 1).


f together with property (d)

So it is

-periodic,

makes it clear that

mj [uj ].

Exercises
(1) Verify property (b) for the theta function above (20.3).

CHAPTER 21

The Poncelet problem

First let's recall the most elementary statement of the porism


from Chapter 1.

One starts with two conics

CR , D R

in

R2 ,

which

for simplicity we can take to be two ellipses cut out by polynomials

fC , fD P 2

with real coecients:

CR

DR

1.3 whether there exists a closed polygon inscribed in CR


circumscribed about DR . The result stated there, Theorem 1.3.1,

We asked in
and

said that if there is one then there is an innite family. Our goal in
this chapter is not just to esh out the sketch of proof given there of
this porism, but to actually provide a way of deciding for which pairs
there does exist a circuminscribed polygon.
A slight reformulation of the theorem is this: starting from some
point
hits

x0

CR

on

CR ,

draw a line segment tangent to

DR ,

continue until it

again. Begin again at this new point, by drawing the other line

segment through it and tangent to

CR

DR :

x0
etc.

L0

DR
x2

L1

x1

Iterating this construction, we may ask whether it ever closes up 


i.e. returns to its starting point. (We will not care whether the path
crosses itself.) What we will show is that the answer is independent of
the choice of starting point

x0 .
225

226

21. THE PONCELET PROBLEM

21.1. Proof of Theorem 1.3.1


This Theorem has nothing to do with
and

fD

and

being ellipses,

fC

being real polynomials, and so forth  it makes sense more

generally for pairs of conics in the complex projective plane


that is the context in which we view it for the proof.

P2 ,

and

Namely, let

C = {FC (Z, X, Y ) = 0}, D = {FD (Z, X, Y ) = 0} be the conics cut


2
out by homogeneous degree-2 polynomials FC , FD S . If the latter
have coecients in R (not essential for what follows), then the real
points C(R), D(R) make sense and then CR , DR above are just their
intersections with ane space. Now, these ane real points need not
meet (as in the above picture), but by Bezout

and

must meet

in four points counted with multiplicity. We will carry out our proof

under the assumption that the multiplicities are all one, i.e.
meet transversely and so

and

|C D| = 4.

Consider the incidence correspondence

E := {(x, L) | x L} C D
2 is the dual curve consisting of lines tangent to D (at any
P
D
point). In 1.3 we dened pictorially two involutions 1 : E E and
meets C in two points (counted
2 : E E . The idea is that each L D
with multiplicity), and swapping those points gives 1 ; whereas each
x C is in two lines tangent to D (counted with multiplicity), and
swapping those lines gives 2 . Composing involutions gives := 2 1 ,

where

which is no longer an involution and is the complex geometry analogue


of the iteration described just above.

(x0 , L0 ) E ,

If we pick a starting point

then we are interested in whether

n (x0 , L0 ) = (x0 , L0 )
for some

n N.

The projection

: E C(
= P1 )
(x, L) 7 x
has

21.1. PROOF OF THEOREM 1.3.1

mapping degree

2:

there exist two lines

through a general point

227

L, L0

tangent to

xC

C
D
x

4
of

ramication points (each of order two): namely, the points

xed by the involution

In particular, the ramication points of

C D,

identify with the points of

since through each of these there is a unique tangent to

(rather than two):

C
D

L
By the Riemann-Hurwitz formula (for ), E = dC r = 224 = 0.
This implies E is elliptic, and so has an Abel map u mapping it isomor1
phically to a 1-torus C/ (where depends on E hence ultimately on
C and D).
1to

dene the Abel map you also have to choose a holomorphic

aects the scaling of the lattice but not its isomorphism class

1-form

on

E;

this

228

21. THE PONCELET PROBLEM

We could have carried out this same computation using


(sends
points

(x, L) 7 L), whose ramication points


of 1 and hence identify with bitangents:

(in

E)

:E D

are the xed

C
D

L
There are four of these since

and

are conics in

2
P

hence have

= 4.
|C D|
Now consider an arbitrary involution
nate on

of

u. Any automorphism
u
7 au + b by Ch. 14 Exercise

is denoted

takes the form

C/, where the coordiof C/ (in particular I )


(5), and squaring this

gives

u 7 au + b 7 a(au + b) + b = a2 u + b(a + 1).


If this is to be the identity, we must either have (i)
or (ii)

a = 1

and

bC

a=1

and

b /2,

arbitrary. Case (i) has no xed points as it is

a translation by a 2-torsion point.


By abuse of notation

C/.

Since

and

2 we will think of , , as automorphisms of


1 2

are involutions of

C/

with xed points, they

belong to case (ii):

1 (u) b1 u ,

2 (u) b2 u

(mod ).

Therefore

(u) = 2 (1 (u)) b2 (b1 u) = u + (b2 b1 ),


| {z }
=:

i.e.

is a translation on

C/.

u0 for the image of (x0 , L0 ) under the Abel map. Clearly


(x0 , L0 ) = (x0 , L0 ) i n (u0 ) u0 (mod ). But n (u0 ) = u0 + n ,
which u0 i n 0, i.e. n . We conclude that the Poncelet
Write

2strictly speaking one should write u u1


1
ing to

on

for the involution of

C/ correspond-

21.2. EXPLICIT SOLUTION OF THE PONCELET PROBLEM

construction (starting from


and only if
on

is

n-torsion

(x0 , L0 ))

closes up at the

relative to the lattice. Since

this has nothing to do with the choice of

229

nth iteration if
depends only

(x0 , L0 ).

Q.E.D.

21.2. Explicit solution of the Poncelet problem


The exes are the preferred choices of origin for the group law on

E,
it turns out that the best choice for O is one of the xed points of 2
(the four (x, L) with x C D ). Writing C D = {p1 , p2 , p3 , p }, we
set O := (p , L ) E .
a cubic plane curve. On the incidence-correspondence elliptic curve

Le
p2
p1
e

p3

C
D

(e, Le ) is the rst point in the Poncelet iteration, i.e. (O);


clearly = u((e, Le )), with u : E C/ the usual Abel isomorphism.
n
The question of whether is the identity can be restated in terms of
the (unique) group law on E with origin O :
Here

(21.2.1)

Is

(e, Le )

an

N -torsion

point?

The approach we take to its solution in this section is work of Cayley


as presented in the nice expository article [P. Griths and J. Harris, On

Cayley's explicit solution to Poncelet's porism, L'Enseignement Math.


24 (1978), 31-40.].

A family of conics. Consider the collection of conics depending


on

t P1 :


Dt := p P2 | tFC (p) + FD (p) = 0 ,

where

D = C

and

D0 = D.

Each

Dt

passes through

p1 , p2 , p3 , p .

230

21. THE PONCELET PROBLEM

Recall that the equation of a conic may always be written

p.M.p = 0 ,

a symmetric 3x3 matrix;

det M = 0. Write MC , MD for the


matrices corresponding to C, D , so that tMC + MD corresponds to Dt .
Those t for which Dt is singular, are then just the ti in
the conic is singular if and only if

det (tMC + MD ) = (t t1 )(t t2 )(t t3 ).

(21.2.2)

There are three singular conics through the

{pi }i=1,2,3, :
t3

t2

p3

p1

p1

p2

p3

p2

p3

Dt

Dt

For any

t P1 ,

p2

t1

p1

Dt

let

`t := tangent

line (through

`t C =: p + qt

p )

(denes

to

Dt

qt ).

From the pictures above, we see that

qti = pi

(i = 1, 2, 3),

q = p (double

intersection),

q0 = e.
So stereographic projection through
ization)

p gives

an isomorphism (normal-

P1 C
sending

t 7 qt ,
in particular

ti 7 pi (i = 1, 2, 3),
7 p ,

21.2. EXPLICIT SOLUTION OF THE PONCELET PROBLEM

231

0 7 e.
This makes

existence domain (cf.

3 of

P1 branched at t1 , t2 , t3 ,  i.e. the


p
2.3) of (21.2.2), which is to say the Riemann

a double-cover

surface

(21.2.3)


s2 = det(tMC + MD ) =: E.

(e, Le ) on E corresponds to a point over t = 0 on E ; call this


. (Moreover, O E corresponds to [0 : 0 : 1] =: O E , as it should.)

The point

Summarizing everything in a picture:

(e,L e )
(p ,L )

P1

t1,t2,t3,t 4

p 1,p 2,p 3,p4 = C D

=(0,s 0)

P1

t gives local
(t 1,0)

holo. coord.
with t( )=0

(t 2,0)

(t 3,0)

Our main question (21.2.1) becomes:


Is
Now

t1 + t2 + t3

N -torsion

E?

on

may not be zero and we are lacking a factor of

4,

so

is not quite in Weierstrass form. But it is easy to see that we have

a normalization

P : C/ E
given by

Clearly this sends

P


ti 0 (u)
u 7 (u) +
,
.
1
3
2 2
0 7 O; dene u0 C/ to be the

point sent to

The question is now:


Is

3remember, E

u0 N -torsion

was already (via

on

C/?

a double-cover of

branched over

p1 , p2 , p3 , p

232

21. THE PONCELET PROBLEM

Normal elliptic curves and a multiple addition theorem.


Put

uj := u0 + j ,

where

j C.

Abel's theorem implies

Proposition 21.2.1. There exists a

tion

only if

with order-N pole at

-periodic

meromorphic func-

and simple zeroes at

u1 , . . . , u N ,

if and

u1 + + uN 0 (mod ).

What we are really after here is the vector space

of meromorphic

E with at worst an order-N pole at O (and no other poles).


There are N 1 degrees of freedom coming from pushing around the
P
{uj } (while keeping
uj 0) and one degree of freedom from multiplying the function by a constant. So dim V = N ; let {f1 , f2 , . . . , fN }
(with f1 constant) be a basis, and dene
functions on

1
N : E PN
[w1 ::wN ]
by

([1 : t : s] =:) z 7 [f1 (z) : : fN (z)].


Definition 21.2.2. The image of

mal elliptic curve of degree

f1 , f2 , f3
For

to be

N.

N ,

denoted

(Note that

E3 is

EN ,

is called a nor-

essentially

 take

1, t, s.)

uj 0,

there exists a function

on

EN

with zeroes at

N (P(uj )), and order N pole at N (O). Now the hyperplane at inN
nity {w1 = 0} P 1 intersects EN only at N (O) (with multiplicity
N ). If written as the pullback to EN of a rational function, it follows
that F has denominator w1 ; the numerator must
then also be a ho
H(w)
1
. It follows that the
mogeneous linear form H SN , i.e. F =
w1
N (P(uj ))

all lie on

{H = 0} PN 1 ,

EN

and so

coords. of
N (P(uj ))
(21.2.4)

z }| {
0 = det[ fi (P (uj )) ].
|{z}
=:Fi

N (P(uj ))
H(w)
function
w1

Conversely if this is satised then the

{H = 0}; one writes down


PN
divisor
j=1 [uj ] N [0], and

the

lie on a hyperplane
and computes its

EN

concludes (by Abel) that

uj 0.

21.2. EXPLICIT SOLUTION OF THE PONCELET PROBLEM

We can push this computation further. Expand

233

Fi (u0 + j ) =

(N 1)

Fi (u0 ) + Fi0 (u0 )j + +

Fi
(u0 ) N 1

+ N
j ( ),
(N 1)! j

then apply multilinearity of the determinant to expand the RHS of


(21.2.4):

h
i
Y

(j1)
0 = const. (k ` )det Fi
(u0 ) +

terms of higher

homog. degree

k>`
Dividing by

uj u0 ),

{j }

in the

` ) and taking the limit as all j 0 (i.e.

k>` (k

all

this becomes

0 = det

(21.2.5)

(j1)
Fi
(u0 )

i = 1, . . . , N
j = 1, . . . , N

The determinant on the RHS of (21.2.5) is called the Wronskian of


P
N P . Notice that in the limit N
j=1 uj 0 becomes N u0 0; so
this last condition is equivalent to (21.2.5)!
Example 21.2.3. Here is what the above calculation (using mul-

tilinearity of the determinant) looks like for


degree higher than


F + F0
1
1 1

F2 + 1 F20

F
1
=
F2

N = 2,

{j }:

0
F1 + 2 F1 F1 + 1 F10
=
F2 + 2 F20 F2 + 1 F20



(2 1 )F10

= (2 1 )
0

(2 1 )F2

ignoring terms of

in the

(2
(2

1 )F10
1 )F20


F1 F10
.
F2 F20

Using the chain rule and again multilinearity of  det, one nds
that the vanishing of the Wronskian is independent of the choice of
local coordinate on

E.

So we can replace

by

(and hence

by

f ),

which yields our multiple addition theorem:

u0 is N -torsion in C/ (and
th step) if and only if
the N
h
i
(j1)
det fi
(0) i = 1, . . . , N = 0.

Theorem 21.2.4.

tion closes up at
(21.2.6)

the Poncelet itera-

j = 1, . . . , N

4note:

higher homogeneous degree in the

{j }

means higher than

k>` (k ` )

234

21. THE PONCELET PROBLEM

(j1)

(0) probably requires explanation: rst, we are viewing f locally as a function of t (rather than
st derivative is (total derivof [1 : s : t] =: z on E ), and the (j 1)
ative) with respect to t. The  0 just means t is set to 0 at the end;
this is because we are evaluating at (i.e. u0 ), which has coordinates
[(s, t) =] (s0 , 0).
Remark 21.2.5. The meaning of

Application in the case

fi

odd. Obviously we can't compute

the Wronskian (21.2.6) unless we know the


Take

N = 2m + 1.

Then for

fi .

f1 , . . . , fm+1 ; fm+2 , . . . , f2m

we may

choose

1, t, . . . , tm ; s, st, . . . , stm1 .
These have order of pole at

0, 2, . . . , 2m; 3, 5, . . . , 2m + 1.

dj1 ti1
The determinant in (21.2.6) is then (using that
dtj1

= 0

unless

j = i)

Writing


1
0
0


.
..
.

.
.

0
m!
0


m+1 s
d


dtm+1

. .
.
.
.
.. . . ..
.


m+1
m1 )


d dt(st

m+1

p
s = s(t) = det(tMC + MD )

A0 = s0 ),





.
.

.


0

.
2m
d s

dt2m


.
.

.


d2m (stm1 )

dt2m

..

..

= A0 + A1 t + A2 t2 +

(here

this becomes a nonzero constant times


Am+1 A2m

..
.
..
.
.
.
.

A
Am+1
2

(21.2.7)

(2m + 1)-gon
(21.2.7) vanishes.

We conclude that there is a circuminscribed


a family of such) for the pair

C, D

i

Example 21.2.6. We work out the case

determinant

(21.2.7)

is just

A2 ,

N = 3,

i.e

(and hence

m = 1.

The

so we can get a Poncelet triangle

21.3. ELLIPTIC BILLARDS

A2 = 0.

Writing

Ti =

235

1
, calculate
ti

v
u 3
u Y
p
s = det(tMC + MD ) = t (t ti )
i=1

r

3 
Y
t
Ti2 2
Ti
1 =C
t
=C
1 t
t
2
8
i
i=1
i=1
3
Y

A2
1X 2 1
=
T + (T1 T2 + T2 T3 + T1 T3 ) .
C
8 i=1 i
4

=
If

T1 = 1,

solving a quadratic equation we nd

A2 = 0

equation of

(1 + T )2
(1 T )2
, T3 =
for some T
4
4



4
4
2
E reads s = (t1) t
t
.
(1 + T )2
(1 T )2
T2 =

If we take

MD =

4
(1+T )2

4
(1T )2

1 0 0

MC = 0 1 0
0 0 1

0 ,
1

corresponding to


C=
then

= 1


4y 2
4x2
+
=1 ,
(1 + T )2 (1 T )2



D = x2 + y 2 = 1 ,

and indeed


det (tMC + MD ) = t

4
(1 + T )2


t

4
(1 T )2


(1 t) .

This recovers Example 1.3.2(b) from the beginning of the course! It's
easy to draw one triangle, but seems quite nontrivial that you get one
independent of the starting point.

21.3. Elliptic billards


Returning to the real world, let

CR R2
2

be an ellipse with foci

F1

F2 . (CR consists of all points in R , the sum of whose distances from


F1 and F2 is a xed constant.) We imagine that CR is the boundary of
a pool table (frictionless, of course!). A billiard trajectory for CR is a
sequence of pairs (xi , Li )i0 with xi , xi+1 C Li and where Li1 , Li
make equal angles with Txi CR
and

236

21. THE PONCELET PROBLEM

xi

Li1

Li
x i+1

x i1

 i.e. one has equality of angles of incidence and reection.


If

DR

is another conic (ellipse or hyperbola) then a (real) Poncelet

(CR , DR ) is a sequence of pairs (xi , Li )i0


Li tangent to DR .

trajectory for

C Li

and

Theorem 21.3.1. [L. Flatto, 2003] (a) Assume

5
with

CR .

with

DR

xi , xi+1

is confocal

Then the (real) Poncelet trajectories are billiard trajectories

with respect to

CR .

(b) Conversely, any billiard trajectory for

F1 or F2 and not along the minor


CR and some DR confocal with CR .

CR

not passing through

axis, is a Poncelet trajectory for

We will prove only (a); Flatto does (b) in Appendix E of his book
[L. Flatto, Poncelet's Theorem, AMS, 2009] (which is, by the way,
written for undergraduates).

At any rate, the two proofs are very

similar.
Remark 21.3.2. It's worth pointing out right away that given

(x0 , L0 ) (L0

not containing

F1

F2 and not the


with CR and tangent
or

minor axis), there

DR confocal
to L0 . If L0 passes
between F1 and F2 , DR is a hyperbola; otherwise, it's an ellipse. One
p
det(tMC + MD ) obtains infordetermines this DR , and then from
mation (as in 21.2) on whether the Poncelet trajectory closes up. By

is a unique conic

the Theorem, this is also the billiard trajectory! You'll use this to do
a computation in Exercise (2) below. But I should emphasize that if
you change

(x0 , L0 )

change the choice of

5that

is,

F1

and

F2

(i.e. the choice of billiard trajectory), you have to

DR

accordingly.

are the foci of

DR .

If

DR is a hyperbola, this just means that


F1 and F2 must remain constant.

the dierence of distances from its points to

21.3. ELLIPTIC BILLARDS

237

Proof. (of (a)): We begin with a general principle, for a conic

QR

with foci

F1 , F2 .

Given

p0 QR ,

let

L := Tp0 QR
F2

p0

F1

F2

QR
and denote by

F20

the reection of

|pq|

for the segment and

F2

in

L.

p, q write pq
:= |F1 p0 | + |p0 F2 | and

Given points

for its length. Set

note that by denition of ellipse,

|F1 q| + |qF2 | =

(q QR ).

p L\{p0 }, |F1 p| + |pF20 | = |F1 p| + |pF2 | visibly exceeds ,


0
that taking p = p0 minimizes |F1 p| + |pF2 |. It follows that

If

F1 p0 p0 F20 = F1 F20 .

(21.3.1)
Now let

meaning

CR , DR be confocal  assume that DR is an ellipse.

Apply-

ing the principle that (21.3.1) holds for the above construction, leads
to a picture

F2
1+1

p
2
2

F
1

TpC R

A
F1

F
2

F2

CR
DR
in which the solid black lines are part of a Poncelet iteration and we
must show

1 = 2

(so that it is a billiard trajectory).

reection in

238

21. THE PONCELET PROBLEM

Tp CR

(dotted black) is denoted by one prime, reection in solid black

lines by two primes.

F1 A + AF2 = F1 B + BF2 ,

By denition of ellipse,

which implies

|F100 F2 | = |F1 F200 |.


F100 pF2 and F1 pF200 are rotations
of each other (through p), so that + 21 = + 22 ( = 1 = 2 ).
It is obvious from the picture that 1 + 1 = 2 + 2 , and so we indeed
conclude that 1 = 2 .

From there it is clear that the triangles

Exercises
(1) Consider the pair of conics

C, D

from Exercise (2) of Chapter 1

once more  but in the following form: write

1 0 0

MC = 0 1 0 ,
0 0 1

1 0 0

MD = 0 1 0
0 0 r2

and use these to dene quadratic forms by e.g.

QC (X, Y, Z) =

So

QC = 0

X Y

1
0
0
X


Z 0 1 0 Y = X2 + Y 2 Z2
0 0 1
Z

denes

and

QD = 0

Working in homogeneous coordinates

D
[V : T : U ],
denes

P2 .

as conics in

dene an elliptic

curve by

U 2 V = det (T MC + V MD ) .
In ane coordinates, this is

u =

U
.
V

u2 = det(t MC + MD ),

where

t=

This is the general prescription for the elliptic curve

arising in the Poncelet construction, exactly as above.


have to show is that in the present situation, with
as given, the elliptic curve is singular.

MC

T
,
V

All you
and

MD

It's practically a one-line

problem, and you may use the ane setup.

But now you are in

a position to see why the curve being singular should make the
Poncelet problem easier, and even to see why (from the abstract
perspective) the solution involved trig functions.

EXERCISES

(2) Let

a>b>0

239

and


CR =


x2 y 2
+ 2 =1 ;
a2
b

( a2 b2 , 0). We plan to shoot our pool ball vertically

along the line (L0 =) {x = c}, where 0 < c < a (and c 6=


a2 b2 ).
For what value of c does the resulting billiard trajectory yield a
triangle? [Hint: the conics confocal with CR are all of the form
n 2
o
y2
x
+ b2 = 1 . Also: while straightforward, this is not a 1a2
it has foci

line computation!]

CHAPTER 22

Periods of families of elliptic curves


2
E E P are simply elements of the
, where , are 1-cycles generating

The periods of an elliptic curve


period lattice

E = Z

1 (E) is some canonically chosen generator. (That


dened over Q, then should be the restriction of a rational
2
dierential 1-form on P dened over Q.) If E is taken to vary with
1
respect to a parameter t P , the periods give interesting multivalH1 (E, Z)
is, if E is

and

ued transcendental functions (e.g. hypergeometric functions) which are


related to modular forms.
In this Chapter we explore (via examples) two dierent approaches
to computing period functions of this sort  the Euler integral
method and the Picard-Fuchs method.

The rst of these is just a

way of computing the integral using Laurent polynomials; the second


derives a homogeneous linear ordinary dierential equation satised by
the periods, which yields a recurrence relation for their power-series
coecients. Actually, both methods yield power series at rst but one
can sometimes recognize what functions they are the power series of.
This may sound like complex function theory, but in fact the power
series coecients (esp. when related to modular forms) can have arithmetic meaning, as we shall see in the next chapter; in the context of
mirror symmetry (one of several interfaces between algebraic geometry
and string theory), power series derived from periods are related to
counting curves on threefolds.
Sections

22.1

and

22.3

will have a bit of overlap with

22.1. Holomorphic 1-forms on a smooth cubic

F S3

18.1.
P2

dene a smooth curve E = {F (Z0 , Z1 , Z2 )


(31)(32)
= 1, so that E is elliptic.
genus formula g =
2

= 0};

F = Z0 Z1 Z2 t(Z03 + Z13 + Z23 ),


2i
P1 \{0, 13 , 33 , 3 } (3 = e 3 ).

for any

Let

Example 22.1.1.

32

241

by the

242

22. PERIODS OF FAMILIES OF ELLIPTIC CURVES

For the ane forms of the equation we shall use the following notation:

v
u

L1
y

L0
2

P
x

L2
x=

Z1
,
Z0

y=

Z2
, and equation is
Z0

P2 \L2 , the coordinates are u =


g(u, v) := Z13 F (Z0 , Z1 , Z2 );

Z0
,
Z2

v=

Z1
, and equation is
Z2

the third neighborhood is left to you;

P2 \L0 ,
f (x, y) :=
on

the coordinates are

1
F (Z0 , Z1 , Z2 );
Z03

on

P2 \L0 L2 , we have u =


1 x
3
f (x, y) = y g y , y .
on

1
,
y

v =

x
;
y

y =

1
,
u

x =

u
; and
v

on E by




dx
du
dv
dy
=
=
=
=
fy E\(L0 VT)
fx E\(L0 HT)
gv E\(L2 HT)
gu E\(L2 VT)

Now dene a form

where

the notation

intersects

E\(L0 VT) means E minus those points where


L0 or has a vertical tangent line (similarly, HT

means horizontal tangent);

equality of any two dierentials above is meant in the sense of


where both are dened;

for example: on

dx
fy

E , f = 0 = 0 = df = fx dx + fy dy =
= 0 and) fx , fy 6= 0;

dy
where (f
fx

the   means that the third neighborhood stu is left to


you.

fx and fy
do not simultaneously vanish (E is smooth!), {E\(L0 VT)}{E\(L0
HT)} is all of E\L0 . So the 6 dierent domains of denition glue to give


(E\L0 ) (E\L1 ) (E\L2 ), which is all of E . Morover, dx
fy
Now consider the domains of the rst two expressions: since

etc.

are all holomorphic where they are dened.

1 (E).

E\(L0 VT)
We conclude that

22.2. PERIOD OF A FAMILY OF CUBIC CURVES (EULER INTEGRAL METHOD)


243

deg(()) = 2g 2 = 2 2 = 0, and so 's lack


0
1
of poles implies it has no zeroes either. Any other (E) has
0

O(E), and then by Liouville 0 is a constant multiple of . So

1 (E) has dimension 1, and spans it.


By Poincar-Hopf,

22.2. Period of a family of cubic curves (Euler integral


method)

Et

Now consider the Hesse family

of elliptic curves, already given

in Example 22.1.1, with ane form

f (x, y) = xy t(x3 + y 3 + 1) = 0,
(For the four values of

t C.

excluded in the example,

not an elliptic curve. I won't write

ft

Et

is singular hence

because the subscript is reserved

here for partial derivatives.) An alternate form of the equation, valid


on

C C ,

is

 3

x + y3 + 1
1t
= 0,
xy
|
{z
}
=:(x,y)

where

C[x, x1 , y, y 1 ].
holomorphic 1-forms

belongs to the ring of Laurent polynomials

From the last section, we have the family of


dx
1 (Et ).
t :=

fy Et
We can obtain a family of 1-cycles by noticing that
is empty for

|t| <

1
, since
3

|(x, y)| < 3

for

x, y

{|x| = |y| = 1} Et
in the unit circle.

Indeed,

t := {|x| = 1, |y| 1} Et
has this empty set as its boundary

t ; and so we would like to compute

the period

P (t) :=

t
t

as a function of
rank

2,

t,

on the open disk

1
. Now since
3

|t| <

there is a complementary 1-cycle

period

Q(t) :=

t .
t

on

Et ,

H1 (Et , Z)

has

and cooresponding

244

22. PERIODS OF FAMILIES OF ELLIPTIC CURVES

Noticing from the homogeneous form of the equation that

E0 = {Z0 Z1 Z2 =

0} is a union of 3 lines (
= P1 ), we can easily visualize what happens to
Et , t , and t as t tends to zero:

Et

E0

Z1 =0

Z0 =0

undergoes
degeneration

(thin black cycles are


pinched to points)

Z2 =0

t 0) to dx
on Z2 = 0, and 0 passes
x
through the poles of this form while 0 traverses the unit circle around
them, we infer that Q(t) as t 0 but P (t) 2i.
We now compute P (t) more precisely, by rst noting that the area
From the fact that

tends (as

integral

|x|=|y|=1
(since

dx dy
=
f (x, y)

dx df
fy f

df = fx dx + fy dy and dx dx = 0)



df (x, y)
1
=

dx
fy (x, y)
|x|=1
|y|=1 f (x, y)

(where inside the parentheses

is a xed constant).

Now thinking

f (x, y) = 0 for |t| small (and x xed with |x| = 1),


y + ay + b = 0 where a = xt is big and b = x3 + 1 is not. This

about the equation


we have

means that two of the roots are big and one is small  in particular,
there is exactly one solution

y(x) with modulus less than 1.

Therefore,

by Cauchy's residue theorem, the integral above


2i

=
|x|=1

= 2i

1
fy (x, y(x))

dx

dx
= 2i
fy

|x| = 1
y = y(x)

So

1
P (t) =
2i

|x|=|y|=1

1
=
2i

dx dy
1
=
f (x, y)
2i
X

|x|=|y|=1 n0

t .
t

|x|=|y|=1

dx
x

dy
y

1 t(x, y)

tn n dlogx dlogy

22.2. PERIOD OF A FAMILY OF CUBIC CURVES (EULER INTEGRAL METHOD)


245

where dlogx

dx
and
x

means simply the

nth

power of

(x, y).

Using

Cauchy residue twice this

= 2i

tn n (0, 0)

n0
in which

n (0, 0) =: [n ]0 is the constant term of n = (x2 y 1 +x1 y 2 +

x1 y 1 )n .
We can make this more explicit. Given a product

(x2 y 1 + x1 y 2 + x1 y 1 ) (x2 y 1 + x1 y 2 + x1 y 1 )
|
{z
}
n times

each contribution to the constant term comes from exponents sum-

(0, 0) (i.e. multiplying to x0 y 0 ). But the only combinations of (2, 1), (1, 2), (1, 1) summming to (0, 0) are: m(2, 1) +
m(1, 2) + m(1, 1). Hence, the only possibility for a nozero constant term is to have n = 3m (i.e. 3|n), and the number of ways to

ming to

choose

is

2 1

x y from m factors
x1 y 2 from m factors

1 1
x y from m factors

3m
m,m,m

:=

(3m)!
. That is,
m!m!m!

P (t) = 2i

3m

m0

(3m)!

= 2i 2 F1
(m!)3

where by denition (writing

1 2
, ; 1; (3t)3
3 3

(a)m := a(a + 1) (a + m 1)

for the

Pochhammer symbol)

2 F1 (a, b; c; u)

:= 1 +

X (a)m (b)m
um
(c)
m!
m
m1

is the Gauss hypergeometric function.


Notice that

P (t)

is really a function of

the symmetry in the family

Et :
Et E3 t

(x, y) 7 (3 x, y).

(3t3 ) = u.

This reects

246

22. PERIODS OF FAMILIES OF ELLIPTIC CURVES

The Gauss hypergeometric function satises a well-known ODE. In this

P0 (u) = P (t) and Q0 (u) = Q(t))




d2
2
d
u(1 u) 2 + (1 2u)
P0 (u) = 0.
du
du 9

case (writing

Q0 (the
log u
P (u)
other period), which turns out to have a term of the form
2i 0
It turns out that the ODE satised by

reecting the fact that following

P0

must be satised by

around

t = 0

yields

t + 3t

in

H1 (Et , Z).
22.3. Cohomology of an elliptic curve

C, with basis {ek }nk=1 .


2
The second tensor power of V , written V V , is the n -dimensional
P
vector space consisting of nite sums
i vi wi (vi , wi V ) subject
to bilinearity (e.g., on the left (v + w) u = v u + w u); it
V2
n
has basis {ek e` }k,`=1 . The second exterior power
V consists of
P
nite sums
vi wi satisfying bilinearity and also v w = w v (so
that v v = 0); it may be viewed as a quotient- or sub-space of V V ,

n
and has basis {ek e` }1k<`n hence dimension
. In particular, if
2
V2
dim V = 2, then dim( V ) = 1; this is essentially the only case we
Let

be a nite-dimensional vector space over

shall use.
Dualizing the homology groups

H1 (E, Z) =
18.1,

from

Z hclosed
Z hboundaries

paths on

Ei

of regions in

Ei

we dene cohomology groups (with complex coecients)

by

H 1 (E, C) := Hom (H1 (E, Z), C) (


= C2 ).
Write

A0 (E)

for

functions and

A1 (E)

f dx + gdy , where
A (E). These are objects

latter locally of the form


have the

2-forms

C 1-forms on E , the
z = x + iy ; and nally we

for

locally of the form

i
i
z = Gd
z dz)
Gdx dy (= Gdy dx = Gdz d
2
2
with

smooth, which you may think of as a eld of innitesimal area

elements.
bundle

1the

V2

In more sophisticated terms, they are C


sections of the
V
2

T E = pE
Tp E . (Refer to 13.1 for notation.)

homology class represented by a 1-cycle

is written

[]

22.3. COHOMOLOGY OF AN ELLIPTIC CURVE

247

The various degrees of forms are connected by exterior dierentiation

d : A0 (E) A1 (E)
sending

F 7 dF := Fx dx + Fy dy
F
F
=
dz +
d
z,
z
z
and

d : A1 (E) A2 (E)
f dx + gdy 7 df dx + dg dy
= (gx fy )dx dy.
The

1st

de Rham cohomology group of


1
HdR
(E, C) :=

ker(d) A1 (E)
=
d(A0 (E))

the class represented by a

1-form

Lemma 22.3.1. The map

[] 7 {[] 7

is then dened by
closed
exact

is written

C forms
;
C forms

[].

1
: HdR
(E, C) H 1 (E, C)

given by

is well-dened, and an isomorphism.

is closed ( = 0) and

= d = = 0. If

(d = 0) then

=
=

Proof. First we check well-denedness: if

exact (

= d ), then ([]) = 0 since


is a boundary ( = ) and is closed

d = 0, so that ([]) is dened on the

level of homology classes.

(The middle equality in both cases  swapping

and

 is Stokes's

theorem, a generalization of the fundamental theorem of calculus for


dierential forms.)

is injective, assume ([]) = 0, and let p be a point of


q
E . Then F(q) = p denes a C function F on E . (The reason F

isn't multivalued is that two paths dier by a cycle , and


=0

by the assumption.) Now = dF by the fundamental theorem of


calculus, and so [] = 0.
Finally, write , for a basis of H1 (E, Z). Using the identication

=
1
H (E, C) C2 which evaluates a functional against this basis,
! a nice

way to think about the map is as sending [] 7


C2 .

To see that

2 [] 7 

integrating

means the complex-linear functional on homology classes given by


over a representative 1-cycle

248

22. PERIODS OF FAMILIES OF ELLIPTIC CURVES

E C/E identies with du. Rescaling

(by a complex constant)


so that
= 1, the 2-vector becomes

!
!

du
1
([du]) =
=
(where we may assume H), and we
du

Moreover, the Abel map

have the standard picture

C/
coord. u

Noting that

that

([d
u]) =

is surjective since

1(=u)

!
!

d
u
du

=
=
du
d
u

!
!
1
1
2
,
span C .

!
, we conclude

Remark 22.3.2. (a) Note that meromorphic 1-forms on a Riemann

d{f dz} = df dz =
f
we have used
= 0,
z

surface are always closed since locally (f mero.)

f
dz
z

dz

dz dz = dz dz = 0. (Here
which expresses the holomorphicity of f o its poles.) Using the

formula as above (i.e. 7 { 7


}), we can dene a map

and

same

ker(Res) K1 (E)
H 1 (E, C)
d(K(E))
which also turns out to be an isomorphism.

(Here

ker(Res)

consists

of forms with no residues  in particular, with no simple poles. This


doesn't mean they're holomorphic though!)

1-form cannot be d of a smooth function G or meromorphic function f . (Locally the integral of is a


holomorphic function, so in either case G or f would actually have to
be holomorphic, hence by Liouville constant, making zero.) So we
(b) A nonzero holomorphic

have a commuting diagram of injective homomorphisms


(22.3.1)


1 (E)
t


/ ker(d)A1 (E)

exact
NNN
NNN
NNN

NNN =
N'



ker(Res)K1 (E)
/ H 1 (E, C)

exact

22.4. DIFFERENTIATING COHOMOLOGY CLASSES

.In what follows

will all just be denoted

249

, which you should read

take the period vector associated to this 1-form.

22.4. Dierentiating cohomology classes


Given a family

t)

many

{Et }tP1

of elliptic curves (smooth but for nitely

t 1 (Et ),
!
P (t)
(t ) =:
.
Q(t)

with holomorphic forms

write

This assumes a choice (this unfortunately only works locally in


basis

t , t

for

H1 (Et , Z),

so that

P (t) =

t , Q(t) =

t .

t)

of

We can

dierentiate this period vector to obtain

P 0 (t)
Q0 (t)

!
,

t (considering the isomorphisms in (22.3.1)) is of some1


1
1
thing in ker(d) A (Et ) or ker(Res) K (E) (but not (Et )). We
0
will use the latter, and we write t for a family of residue-free mero!
0
P
(t)
0
. The point is that by
morphic 1-forms satisfying (t ) =
Q0 (t)
which for each

dierentiating families of cohomology classes you get a new family of


cohomology classes.

Example 22.4.1. Consider the Legendre family

Et P2

given by

the projective closure of

y 2 = x(x 1)(x t),


with holomorphic 1-form family


dx
t =
1 (Et ).
y Et
We have

(t ) =
where

t , t

P (t)
Q(t)

!
=

dx

t x(x1)(xt)

,
t t
=
dx

t t
t

x(x1)(xt)

are the 1-cycles exhibited in the schmatic picture

250

22. PERIODS OF FAMILIES OF ELLIPTIC CURVES

Et
1

"2 sheets over P

t
or the topological picture

" x(x1)(xt) "

"+ x(x1)(xt) "


t
t
 which shows the two sheets (each is a

be

with slits from

to

and

2.3).
From the latter picture, it is clear that for t small we may take t to
stationary on its sheet as t moves, and the two pieces of t on the

to

P1

being glued together to give

dierent sheets not to change either.

Et

(cf.

Therefore we may dierentiate

the above integrals under the integral sign (by

P 0 (t)
Q0 (t)


t
=

d
) to obtain
dt

1
dx
2

(xt)

x(x1)(xt)
1
dx
2

t (xt)

x(x1)(xt)

22.5. THE PICARD-FUCHS EQUATION

and

00

P (t)
Q00 (t)
obviously the rst is


t
=

1/2 dx
|
xt y Et

x(x1)(xt)

3
dx
4

t (xt)2

4 dx
2

(xt)

251

x(x1)(xt)

and the second

3/4 dx
|
(xt)2 y Et

, and

so we have

t0


dx
=
,
(x t) y Et
1/2

These both belong to

(t, 0)

t00


dx
=
.
(x t)2 y Et

ker(Res) K1 (Et ),

3/4

since their only poles are at

(orders 2 and 4 resp.) and the sum of the residues of a meromor-

phic form must always be zero.

(t ), (t0 ), and (t00 ) must be linearly dependent in C2 !


[t ], [t0 ], and [t00 ] are linearly dependent in K1 (Et ) modulo

Of course,
Therefore,

d(K(Et )),

i.e. as cohomology classes.

22.5. The Picard-Fuchs equation


Since what is being dierentiated in the last section is really cohomology classes (via the identication with

C2 ),

it makes sense to

write

Dt [t ] = [t0 ] ,

Dt2 [t ] = [t00 ].

With this notation, the linear dependence observation above implies


an ODE of the form


A(t)Dt2 + B(t)Dt + C(t) () = 0
|
{z
}

(22.5.1)

=:DPF

satised by

[t ]

(as a varying cohomology class) hence by

P (t)

and

Q(t)!
However, to nd

A, B,

and

C,

we have to compute. We start by

dierentiating a meromorphic function





2y
4ydx
2dy
d
=
+
,
(x t)2 Et
(x t)3 Et (x t)2 Et
|
{z
}
K(E)

which using

y 2 = x(x 1)(x t) = dy = 3x 2(1+t)x+t


dx
2y



4y 2
3x2 2(1 + t)x + t

=
+
dx

(x t)3 y
(x t)2 y
Et

becomes

252

22. PERIODS OF FAMILIES OF ELLIPTIC CURVES

and using

y 2 = x(x 1)(x t)

again


x2 + (2 2t)x + t
=
dx
(x t)2 y
Et

(x t)2 2tx + t2 + (2 2t)x + t dx
=
(x t)2
y Et

(2 4t)x + t2 + t dx
= t +
(x t)2
y
Et

= = t + 4(1

2t)t0

+ 4t(1 t)t00 .

d of a meromorphic function, hence (has both


1
trivial in H (Et , C). We conclude that (dividing

So this last expression is

0 and) is
by 4 to simplify)

its periods
through

DPF = t(t 1)Dt2 + (2t 1)Dt +


kills

[t ].

1
4

From ODE theory, the associated indicial equation is





C(t)
B(t)
t r + lim
= r2
r(r 1) + lim
t0 A(t)
t0 A(t)
which has a double root, implying one holomorphic solution (unique
up to scale) and one logarithmic solution near

t = 0.

22.6. Computation of a period (Picard-Fuchs method)


First let's compute its limit

t = lim
t0
t0
t
| {z }

lim

dx
p
.
x(x 1)(x t)

P (t)

P1 ) above, this



dx
dx

=
= 2i Res0
x x1
x x1

Referring to the picture of

(on the slit

1
= 2i
= 2,
1
P
P (t) must be the holomorphic solution. Write P (t) = 2 an tn ,
a0 = 1, and apply DPF :
X
0 = DPF
an tn

X
1
=
t(t 1)(n + 2)(n + 1)an+2 + (2t 1)(n + 1)an+1 + an tn
4
n0

and so

22.6. COMPUTATION OF A PERIOD (PICARD-FUCHS METHOD)

where we have shifted indices after dierentiating.


with like powers of

X
n0

t,

253

Collecting terms

this


X
1
an (n + 1)an+1 tn +
[2(n + 1)an+1 (n + 1)(n + 2)an+2 ] tn+1
4
n0
X

(n + 1)(n + 2)an+2 tn+2 .

n0
Shifting indices once more, we have

X 1

n0

an (n + 1)an+1 + 2nan n(n + 1)an+1 + n(n 1)an tn


"
#
2
X 
1
=
n+
an (n + 1)2 an+1 tn .
2
n0

Since this power series is zero, we get a recurrence relation for the
coecients of

P (t):

an+1 =

n + 12
n+1

2
an ,

so that

an





1/2 3/2
1/2 3/2 (1/2 n + 1) 2
1/2 + n 1 2
= a0
=

|{z} 1 2
n
n!
=1

2

1/2

=
and

P (t) = 2

X 1/22
n

n0
Again, the situation as

t0

tn .

looks like

0
t

t
In the next chapter the formula for
rational points on cubics over

Fp .

P (t)

will be connected to counting

254

22. PERIODS OF FAMILIES OF ELLIPTIC CURVES

Exercises
(1) Check that 2 F1

1 2
, ; 1; (3t)3
3 3

m0

t3m

(3m)!
by writing out
(m!)3

the Pochhammer symbols.


(2) Show that the curves

Et := {Z0 Z1 W0 W1 t(Z1 Z0 )2 (W1 W0 )2 =

0} P1Z0 :Z1 P1W0 :W1 are in fact elliptic (except at those nitely
many t  which ones?  for which they are singular). You could
1
do this by projecting to the rst P and using Riemann-Hurwitz to
compute the genus. (To use R-H in this way, rst nd all vertical
tangents  places on the curve where the partials with respect to

W0

and

W1

vanish.)

(3) Writing the family of curves from the last exercise in ane form,

xy t(x 1)2 (y 1)2 = 0 (or 1 t(x, y) = 0), dene a family


of loops t H1 (Et , Z) for small t, and a family of holomorphic
1
forms t (Et ), exactly as in the text. Compute the period

P (t) := t t as a power series, using the computation done above


as a model.

in Remark 22.3.2 is well-dened and an isomorphism.


P
log t
Find Q(t) in the Legendre example by plugging
P (t)+ n1 bn tn
2

(4) Check that


(5)

into the Picard-Fuchs equation.

CHAPTER 23

Counting

Fp -rational

points on elliptic curves

In this nal chapter on elliptic curves, we take a brief dip into something much more arithmetic, counting the number (mod
in

P (Fp )

p) of solutions

to the equations for the Hesse and Legendre cubics from the

last chapter. These cubics still depend on

t,

which is taken to be an

integer now (rather than a complex number) so that we can reduce

p.

modulo

In a truly bizarre twist, the number of points (over

Fp )

in

each case is given by nearly the same power series as the holomorphic
period on the corresponding complex family of elliptic curves from Ch.
22. We briey explain one abstract way, due to Y. Manin, to understand this connection between arithmetic and transcendental algebraic

geometry.

23.1. Sum formulas

Fp the eld with p elements (i.e. Z/pZ


viewed as a ring). Equality in Fp will generally be denoted by  =, not
 . (We will use the latter for counting points mod p.)
Let

be an odd prime, and

(p)

Lemma 23.1.1. For

k Z,
(

xk =

xFp (or Fp )

in

0, p 1 - k
1, p 1 | k

Fp .
y Fp , the assignment x 7 yx
groups Fp Fp . Therefore
X
X
X
yk
xk =
(xy)k =
xk .

Proof. Given

phism of additive
(23.1.1)

xFp

1The

xFp

yields an isomor-

xFp

interested reader is urged to study the theory of modular forms(!) and look

at papers by Ahlgren, Ono, Papanikolas et al in particular. Modular forms are, on


the one hand, essentially given by periods on families of (self-products of ) elliptic
curves; on the other hand, they are frequently constructed via point-counts over
nite elds.
255

256

23. COUNTING

Now,

Fp

Fp -RATIONAL

p 1,

is a cyclic (multiplicative) group of order

(Fp )k = {1}
Provided

POINTS ON ELLIPTIC CURVES

p 1 - k,

then, there exists

and so

p 1 | k.

y Fp

with

y k 6= 1.

By (23.1.1),

we have

0 = (y k 1)

xk

xFp
which implies (dividing by

y 1)
X
0=
xk .
xFp

p 1 | k , then xk = 1 for all x Fp


X
X
xk =
1 = p 1 = 1.

On the other hand, if

and so

xFp

xFp


Lemma 23.1.2. For

x Fp ,

p1
X

(
xk =

k=1
Proof. If

0, x 6= 1
.
1, x = 1

x = 1 then the sum is 1| + {z


+ 1} = p1 = 1.

If

x 6= 1

p1 times

then

(1 x)
| {z }
6=0

p1
X

x =

k=1

p1
X

k=1

= xx

p
X

xk

k=2
p

= x(1 xp1 )
which (since

Fp

is cyclic of order

p 1)
= x(1 1) = 0


Lemma 23.1.3. Let

Fp .

F2p
Proof. Since

over, if

Fp

Then

p1
2

1,

(p)

p1
2

and

1.
(p)
p1

p 1, ( 2 )2 = p1 = 1. Morep1

cannot have a 2 = 1 (Fp would then

is cyclic of order

is a generator then we

23.2.

Fp -POINTS

ON THE LEGENDRE ELLIPTIC CURVE

p1
, a contradiction). Hence
2

have order

x 7 x

p1
2

257

yields a surjective

homomorphism of multiplicative groups

Fp  {+1, 1},

1
|F | = p1
.
2 p
2
p1
1. As 2 elements of

whose kernel necessarily has order


square, then
exhaust the

p1
2

= p1 =
kernel of and

23.2.

Fp -points

Now if

Fp

= 2

is a

are squares, these

the non-square elements go to

1.

on the Legendre elliptic curve

Consider once again the Legendre family of cubics

Et = {Y 2 Z = X(X Z)(X tZ)},


t Z.
Et (Fp ) Et (Fp ),

p we can look at the


solutions
i.e. with X, Y, Z in Fp resp. its algebraic
.
closure; there is a clear analogy to Et (Q) Et (Q)
We are going to compute the number of points |Et (Fp )| modulo p,
i.e. in Fp . (Computing the number in Z is a much harder problem.)
but this time with

After reducing mod

First we claim that

|Et (Fp )| 1 +

(23.2.1)

(p)

o
Xn
p1
1 + [x(x 1)(x t)] 2 .
xFp

[0 : 0 : 1] at ; the rest


of the curve is described by y = x(x 1)(x t). By Lemma 23.1.3, the
quantity in curly brackets yields (mod p) 2 if x(x1)(xt) is a square,
1 if x(x 1)(x t) = 0, and 0 if x(x 1)(x t) is not a square. This
2
exactly counts pairs (x, y) Fp solving the ane equation, conrming
The leading  1 on the RHS counts the point

(23.2.1).

P
Fp ) xFp 1 = 0, so the RHS of
X p1
p1
p1
1+
x 2 (x 1) 2 (x t) 2

Now (summing in

(23.2.1) is

xFp

p1
2  p1 
X

p1

2  p1 

X
p1
p1
p1
2
2
= 1+
x 2
x 2 ` (t)`
xk (1) 2 k

`
k
xFp
`=0
k=0
p1
p1





2
2

p1
p1
X
X
X
p1
p1
`
`
k
k
2
2
2
= 1+
x
x (t)
x (1)
.

`
k
xF
`=0
k=0
X

258

23. COUNTING

Fp -RATIONAL

The sum here can be rewritten

POINTS ON ELLIPTIC CURVES

xFp

xp1 G(x), where G(x) =

P p1
2

m= p1
2

a0 , and so the above


p1

p1


2 
2 
X
X

p1/2
p1/2
p1
= 1
x` (t)`
xk (1) 2 k

`
k
`=0
k=0

By Lemma 23.1.1, this is just

p1
2

= 1

(t) (1)

p1
`
2

= 1 (1)

p1

2 
X
p1/2 2

`=0

= 1 + (1)

2

p1/2

`=0
p+1
2

p+1
2

t`

X 1/2
t`
`
`0

=: P (t).
For the last step, we use the denition (in

1/2

1/2

:=

Fp )

3/2 (1/2 ` + 1)
,
`!

which is evidently

1p1+2
2

p
2

to be

1
0 when p > ` > p1
(since
p+1
+1 =
2
2
2
p1 
p1
2
equals
for 0 `
(and is dened
2
`

= 0), and
` p). We conclude:

for

P (t) counts (mod p) the Fp -rational points of

Proposition 23.2.1.

Et .
Notice that
(23.2.2)

P (t)
23.3.

is a mod

p

version of the period

Fp -rational

For this section, take

P (t)

from

22.6!

points on the Hesse cubic

to be an odd prime with

p 1

mod

3.

I can't resist doing the same exercise for the other main example
from the last chapter, namely

Et = {XY Z = t(X 3 + Y 3 + Z 3 )}
where again we assume

t Z.

This has ane form

xy = t(x3 + y 3 + 1)

am x m .

23.3.

Fp -RATIONAL

POINTS ON THE HESSE CUBIC

259

and toric form

1 = t(x2 y 1 + x1 y 2 + x1 y 1 ),
|
{z
}
=:(x,y)

where the Laurent polynomial

(x, y)

is dened for

(x, y) (Fp )2 .

Using the toric form and Lemma 23.1.2, it is easy to compute the

Fp -points
Et := Et (Fp ) (Fp )2 .
Namely, in

Fp

(i.e. mod

|Et |

p)
X

p1
X

tk ((x, y))k ,

(x,y)(Fp )2 k=1
the point being (besides the Lemma) that
exactly those

(x, y)

on

Et .

t(x, y)

is

(in

Fp )

for

Switching the order of summation this

becomes

(23.3.1)

p1
X

tk

(x, y)k .

(x,y)(Fp )2

k=1
Now by Lemma 23.1.1

xi y j =

xi

xFp

(x,y)(Fp )2
For

yj =

yFp

1, p 1 | i, j
.
0, otherwise

k [1, p 2],
(
((x, y))k = []0 +

Our assumption on

x, y
p1

terms with powers of


not both divisible by

implies that

3 - p 1,

)
.

and so

((x, y))p1 = [p1 ]0 + x2(p1) y (p1) + x(p1) y 2(p1) + x(p1) y (p1)


(
)
terms with powers of x, y
+
not both divisible by p 1
(in particular, there are no

xp1

or

y p1

x(p1) , y (p1) , xp1 y (p1) , x(p1) y p1 ,

terms). So (23.3.1) becomes

p1
X

tk [k ]0 tp1 3.

k=1
Recall from

22.2

that

[k ]0 =

3m
m,m,m

if

k = 3m

(and

if

3 - k ).

260

23. COUNTING

Fp -RATIONAL

POINTS ON ELLIPTIC CURVES

On the other hand, looking along the coordinate axes

Z=0

X = 0, Y = 0,

we get (only) the points

[1 : 1 : 0] , [0 : 1 : 1] , [1 : 0 : 1]
in

Et (Fp ).

and so may assume


only

in

Z=0

For example, along

Y = 1;

(on

Et )

we must have

then the equation is

X, Y 6= 0

X + 1 = 0.

This has

X = 1 as solution: otherwise we would have an element of order


Fp , so 6 | p 1, contradicting our assumption on p.

We conclude that

p1

|Et (Fp )| 3(1 t


(p)

p1
bX
3 c

m=1

again very reminiscent of

P (t)


3m
t3m ,
m, m, m

from before!

23.4. Deep reasons for (23.2.2)


The issue is this:
counting the point at

in

23.2,

why on earth does

|Et (Fp )| 1

(not

appear to solve the Picard-Fuchs equation


p+1

t(t 1)Dt2 + (2t 1)Dt +


where P (t) is the solution

2
() = 0 ? Indeed, P (t) 1 = (1)2 P (t),
from 22.6!! The two computations were

1
4

quite elementary, after all, so maybe there's an elementary explanation


for their equivalence?
Nope.

This is dealt with in [Clemens, A scrapbook of complex

curve theory, pp.65-69] and I'll just give a hint of the avor here. It involves an algebro-geometric version of the Lefschetz trace formula (the
formula from topology for the number of xedmm points of a mapping),
the Riemann-Roch theorem, Serre duality, and abstract sheaf theory.
Not elementary, but we can give a brief summary (with no claim to
total accuracy).
Consider
points, and

Et over Fp , t Fp . Then writing FP for number


f robp for the map [Z : X : Y ] 7 [Z p : X p : Y p ],


|Et (Fp )| = FP f robp : Et (Fp ) Et (Fp ) .

This should make sense to you because as an automorphism of

pth -power

(Frobenius) map xes exactly the elements of

Lefschetz-type theorem, it turns out that this

n
o
= 1 trace f rob |H 1 (Et /Fp ,O)

Fp .

of xed

Fp ,

the

By the

23.4. DEEP REASONS FOR (23.2.2)

261

H 1 is Cech cohomology computed with respect to the Zariski

topology, O is the sheaf of regular functions, and f rob is the action


by pullback (under f rob) on cohomology classes. It is a 1-dimensional
where the

vector space, with generator represented by a certain rational function

h with two simple poles, at q = [1 : 0 : 0] and some other point


p Et (Fp ). More precisely, H 1 (Et /Fp , O) is isomorphic to the space of
rational functions on Et with poles allowed only at p and q modulo the
subspace of rational functions with poles allowed at either p or q (not
both). [That this space is 1-dimensional in the more familiar complex
case is an exercise below.] You should also note that pulling back by

f robp

stabilizes the vector space we have just described, since (as

are taken to be in

f robp .

Et (Fp )

rather than

Et (Fp )) p

and

p, q

are xed under

So the displayed expression at least makes sense.

P
h = y1 + `0 b` y ` about
q ,2 and also expand a generator t 1 (Et /Fp ) (regular dierentials) as
P
[1 + k1 ak (t)y k ]dy . Recall also from the complex case, that residues
h

Next, we expand

in formal power series

of meromorphic functions require, and depend on, a choice of local


coordinate; while residues of meromorphic 1-forms are invariant (i.e.
require no such choice, as they can already be integrated around a loop
without appending a  dz ).
take residue by computing

So for functions

Resq (F );

if

with a pole at

q,

we

has no other pole, then (as

residues sum to zero) the residue has to be zero.

f rob above, we have f rob h(=


h f rob) = y1p + `0 b` y `p = h + f + g . (The last equality, in which f
has only a pole at q and g has only a pole at p, is by 1-dimensionality
1
of H (Et , O) and the explicit description we gave of it. In that vector

space, this reads f rob [h] = [h].) Moreover, Resq (ht ) = 1 while

Now, writing

for the trace of

= Resq (ht )+Resq (f t )+Resq (gt ) = Resq ((f rob h)t ) = ap1 (t),
with the last equality obtained by multiplying out the explicit expressions for

f rob h

and

t .

So we end up with

|Et (Fp )| = 1 ap1 (t),

2note

that

gives a local coordinate about

[1 : 0 : 0]

on

Et ; x

does not.

262

23. COUNTING

Fp -RATIONAL

and (like the periods of


tion because

[t ]

does.

POINTS ON ELLIPTIC CURVES

t ) ap1 (t) must satisfy the Picard-Fuchs equaAgain, the regular solution of DPF () = 0 is

unique up to scale, and from there we are essentially done.

Exercises
(1) Check that

Et (Fp )
2

E P
p, q E(C)

(2) Let

is closed under

f robp ,

for

t Fp .

be a smooth cubic over the complex numbers, and


two distinct points.

meromorphic functions on

Let

be the vector space of

with poles only at

and

q,

with sub-

Wp and Wq (the meromorphic functions with poles only at


p and q respectively). Using Abel's theorem, prove that the dimension of V /(Wp + Wq ) is one. [Hint: you will also need to use
1
the fact that (E) has no zeroes, and that the residues of
F V given by Resp (F ) and Resq (F ) must sum to zero (cf.
spaces

Prop. 13.1.6(b)).]

Part 4

Curves of higher genus

CHAPTER 24

The algebraicity of global analytic objects

To kick o the last part of this course, on curves of higher genus,


this Chapter will demonstrate two approaches to the following result:

meromorphic (or holomorphic) functions and forms on normalizations


of algebraic curves, all arise as pullbacks of functions and forms on
projective space constructed from rational functions (quotients of homogeneous polynomials) and their dierentials.

This is a special in-

stance of Serre's GAGA principle (global analytic is global algebraic


in the projective setting), and is proved (in

24.1)

using techniques

from Chapter 8 together with the primitive element theorem.

For holomorphic forms, we would like a more precise result (already


hinted at in Remark 20.1.1) on how to think of the holomorphic forms
on a normalization rationally. It is important at this point to recall
part (B) of the Normalization Theorem 3.2.1, which says that every

Riemann surface can be obtained as the normalization of an algebraic


curve in

P2 ,

even one with only ordinary double point (ODP) singu-

larities. So in the course of analyzing curves with ODP's in

24.2

we

will actually have proved (cf. Prop. 24.2.1(c)) that for any Riemann
surface

of genus

g , dim(1 (M )) = g .

Featuring prominently in this

section is the space of homogeneous polynomials vanishing at the set


of ODP's, which will play a key role in the proof of Riemann-Roch in
the next Chapter.

24.1. Chow's theorem for algebraic curves


Let

C P2

be an irreducible projective algebraic curve of degree

d; applying a projective transformation if necessary we have [0 : 0 :


1]
/ C . Start by normalizing C ; that is, express it as the image of
e (P2 \{[0 : 0 : 1]}). One may evidently produce
a morphism : C
1What I haven't proved here, is that a global analytic subvariety of projective space
is an algebraic subvariety; but I will use this at a couple of points in later chapters
(apologies).
265

266

24. THE ALGEBRAICITY OF GLOBAL ANALYTIC OBJECTS

e, by pulling back the


C
Y
2
x= X
, y =
on P ; and
Z
Z

meromorphic functions on the Riemann surface


rational functions

C(x, y)

under

C(x, y) consists of all


in X, Y, Z of the same degree.)
the eld

(Recall

quotients of homogeneous polynomials

C(x, y)C for the


subring of rational functions whose polar set does not contain C , dene
e by
the eld of rational functions on C
More precisely, writing

e := C(x, y)C .
C(C)
Next we consider the projection


: P2 \{[0 : 0 : 1]}  P1
[Z : X : Y ] 7 [Z : X],
e as a

:= with the normalization presents C


d-sheeted2 branched cover of P1 . Write B ( P1 ) for the branch locus,
1
and for a path containing B with P \ simply connected (cf. 8.2).
whose composition

We have inclusions

e K(C)

C(x) C(C)
(24.1.1)

rat'l

mero.

fcns.

fcns.

where the rst is obtained by noting

C(x) C(x, y)C

and

C(x) =

C(x) C(x, y)C .


Now, one might initially speculate that the right-hand inclusion of
(24.1.1) is proper when

has singularities such as ODP's, since: (a)

of a function would seem to

e, (b) at rst glance the pullback


p, q C
have the same value at p and q , and (c)

e
C

ought to be able to take dierent values

an ODP has 2 preimage points

meromorphic functions on

at distinct points, right? The weak link in this chain of reasoning is


(b), as you can see from the following

C = {Y 2 Z = X 2 (X Z)}
[1 : 0 : 0]. The pullback of xy

Example 24.1.1.

at its ODP

values

and

has tangent lies


to

e
C

Y =

therefore takes

(resp.) at the 2 points lying over the ODP.

The point is that rational functions are not well-dened at all points
of

P2 , and this can be used to our advantage to get more

functions on

singular curves. So it becomes plausible that the right-hand inclusion

2the

mapping degree

deg(
) = d (= deg(C))

by Bzout

24.1. CHOW'S THEOREM FOR ALGEBRAIC CURVES

267

of (24.1.1) is an equality, and that is exactly what we shall prove in the


rest of the section.
To that end, let
and denote by

e
K(C)

be a nonzero meromorphic function,

the set of poles of

Writing

0 = f (x, y) = y d + a1 (x)y d1 + + ad (x)

(24.1.2)

for the ane equation of

C,

we have as in

8.2

distinct solutions

{yj (x)}dj=1 to

f (x, ) = 0 over P \, which are interchanged as one


passes through \B. Moreover, by irreducibility of C (hence f ), (24.1.2)

is the minimal polynomial of y , proving that


e :
[C(C)
C(x)] d.

(24.1.3)

x P1 \(
(P )), one can think of (x, yj (x)) as belonge with
ing to C
(x, yj (x)) = x. Consider the elementary symmetric

polynomials (i = 0, . . . , d, with e0 = 1)
For each

ei (x) := ei ((x, y1 (x)), . . . , (x, yd (x))) ,


P1 \(B
(P )). As in 8.2,
the fact that they are bounded away from
(P ) guarantees (by Rie1
mann) their extension to holomorphic functions on P \
(P ). Further,
1
if x0
(P ) has neighborhood x0 P , then for k N suciently
e. By the
large,
:=
((x x0 )k ) is holomorphic in
1 (x0 ) C
which are well-dened and holomorphic on

8.2), ei (x) extends holomorphically across x0 ;


x0 )ik ei (x), ei (x) extends meromorphically
this argument at all points of
(P ), we nd that

same argument (from

but since ei (x) = (x


across

x0 .

Repeating

ei K(P1 );
and by Theorem 3.1.5(a)
Next observe that for

0 =

K(P1 )
= C(x).
1
any x P \(
(P ))

d
Y

and

j {1, . . . , d},

((x, yj (x)) (x, yi (x)))

i=1

= (x, yj (x))d e1 (x)(x, yj (x))d1 +e2 (x)(x, yj (x))d2 +(1)d ed (x) ;

3as

in

8.2

we may change projective coordinates if necessary to put the equation

in this form

268

24. THE ALGEBRAICITY OF GLOBAL ANALYTIC OBJECTS

that is, for a dense subset of points

0 =

d
X

e, (p)
pC

satises the equation

(1)i ei (
(p)) (p)di .

i=0
Therefore the meromorphic function

itself satises

d
X
0 =
(1)i (
ei ) di ,

(24.1.4)

i=0
with coecients in

C(x).

Finally, recall the primitive element theorem, which says that an


algebraic eld extension (of degree
ment (of degree

e).

e)

is generated by a single ele-

(A transcendental eld extension, likewise, has

a transcendental element, which satises no algebraic equation.) Were

e :
[K(C)
C(x)] > d, there would thus
e was arbitrary, (24.1.4)
but as K(C)

be an element of degree

> d;

shows this is not so. hence

e :
[K(C)
C(x)] d.
Putting this together with (24.1.1) and (24.1.3), we see that

e = C(C),
e
K(C)
proving the
Theorem 24.1.2. Every meromorphic function on the normaliza-

tion of an irreducible projective algebraic curve is rational, i.e.

the

pullback of a ratio of homogeneous polynomials.


Corollary 24.1.3. Every meromorphic 1-form on a normalization

is rational (i.e.

f dg

f, g

where

are rational).

e , and let 0 K1 (C)


e
(dx) =: K1 (C)
0
e. Then belongs to K(C)
e ,
be any other meromorphic 1-form on C

hence is rational by Theorem 24.1.2.



Proof. Consider (say)

24.2. Cohomology of a Riemann surface


Let

be a Riemann surface of genus

the 1st homology group


rank

H (M, Z) =

g.

Recall from

20.1

that

closed loops
boundaries is an abelian group of

2g , and dene M 's 1st cohomology group

to be the

vector space of complex-linear functionals

H 1 (M, C) := Hom(H1 (M, Z), C).

2g -dimensional

24.2. COHOMOLOGY OF A RIEMANN SURFACE

269

22.3 (for elliptic curves) we have the de Rham cohomology

Exactly as in

groups
d

1
(M )
HdR

:=

To any closed

ker{A1 (M ) A2 (M )}
d
image{A0 (M )

1-form

A1 (M )}

we may assign the

C 1-forms
.
exact C 1-forms

functional 7

closed

on

loops. By the rst 2 paragraphs of the proof of Lemma 22.3.1 (which


work for any

M ),

this induces a well-dened injective map

1
HdR
(M ) , H 1 (M, C).

(24.2.1)

Surjectivity also holds but will require a little more work than for elliptic curves.
Writing

1 (M ) for the space of anti-holomorphic

forms (the com-

plex conjugates of holomorphic ones), we can embed

1
1 (M ) 1 (M ) , HdR
(M )

(24.2.2)
via

(, ) 7 [ + ].
d(1 (M )) = 0 = d(1 (M ))
injectivity, suppose + = df ,

The map (24.2.1) is well-dened because


(cf.

Remark 22.3.2(a)).

f A0 (M ).

To prove

Then

d(f ) = f d + df = ( + ) =
|{z}
=0

looks locally like a function times dzdz(= 0). Now breaking


M up into triangular regions i with local holomorphic coordinates

zi = xi + 1yi ,

X
X
=
gi dzi gi dzi = 2 1
|gi |2 dxi dyi .
M
i
i
| i
{z
}
since

R0

Since each integral

= 0 gi 0, we have

= 0 0.
M

But using Stokes's theorem and

=
M

M = ,

d(f ) =
M

f = 0
M

270

24. THE ALGEBRAICITY OF GLOBAL ANALYTIC OBJECTS

df = = f
= 0 = f O(M ) = f
z
constant (by Liouville) = = 0. So (24.2.1) is injective.
d3
By now you are quite familiar with the fact that dim(S3
) =

(d1)(d2)
(d3)+2
2
=
. If S is a set of points in P , then the homo2
2
geneous polynomials of degree d 3 vanishing on each of these points
are subject to (possibly dependent) linear conditions. Denoting the
d3
space of such polynomials by S3
(S), we therefore have
which implies

0.

So

(d 1)(d 2)
.
2

dim(S3d3 (S))

(24.2.3)

: M P2 is injective o a
algebraic curve C = {F (Z, X, Y ) = 0}

Now assume
image an

nite point set, with


of degree

d (F S3d ),

having only ODP singularities (as in part (B) of the Normalization


Theorem). Write

for the collection of these ODP's, and note

e.
M =C

By (24.2.3) and the genus formula,

dim(S3d3 (S)) g.
4 we have a map

By Remark 20.1.1,

S3d3 (S) 1 (M )
given by

G 7

g dx
fy

g(x, y) = G(1, x, y) etc. This is necessarily injective: were gdx


to
fy
vanish on C , we would have G 0 on C ; since F is irreducible then
F |G by Study, which is impossible (unless G is trivial) as deg(G) =
d 3 < d = deg(F ).
where

All told, we have a sequence of injective maps of complex vector


spaces

1
(M ) , H 1 (M, C).
S3d3 (S) S3d3 (S) , 1 (M ) 1 (M ) , HdR
Notice that the left-hand side has dimension
side has dimension exactly

2g .

2g

and the right-hand

All the injections are therefore isomor-

phisms and we conclude:


Proposition 24.2.1. For a Riemann surface

malizing an algebraic curve of degree

4see

(of genus

g)

nor-

with ordinary double points

Griths's book Introduction to algebraic curves for more details

24.2. COHOMOLOGY OF A RIEMANN SURFACE

271

S = {p1 , . . . , p },

the holomorphic forms are all pullbacks of rational


gdx
forms of the form f (as described above). Moreover, we have:
y
1
1
(a) [de Rham theorem] H (M, C)
(M );
= HdR
1
1

(b) [Hodge decomposition] HdR (M ) = (M ) 1 (M ); and


(d1)(d2)
1
= dim(S3d3 (S)).
(c) dim (M ) = g =
2
We also get an application to the period matrices

20.1.

1 , . . . , 2g

Recall that if

a basis for

(M ),

is a basis for

then


1 1

.
.
.

..

..


1 g

H1 (M, Z)

2g 1

.
.
.

2g

Proposition 24.2.2. Viewed as vectors in

of

are

R-linearly

described in

and

1 , . . . , g

R2g (
= Cg ),

the columns

independent.

Proof. Suppose otherwise, i.e. that there exists a nonzero vector

a R2g

satisfying

0 = a ;
then we have also (by complex conjugating)

0 = a.
That is,

and so the rank of

b C2g

a=0

!
is less than

2g .

But then there is a nonzero

such that

which means explicitly for each

!
= t0 ,
that

g
g
X
X
(
bi i +
bg+1 i ) = 0.

j i=1

| {z }
=:

|i=1 {z

=:

(, ) 1 (M )1 (M ) goes to zero in Hom(H1 (M, Z), C) =


H 1 (M, C). By our sequence of injections above, = = 0. But since
b 6= 0, this contradicts linear independence of 1 , . . . , g in 1 (M ). 

We nd then that

272

24. THE ALGEBRAICITY OF GLOBAL ANALYTIC OBJECTS

Exercises
(1) Write a basis for the holomorphic 1-forms on the (smooth) curve

C P2 with
dim(1 (C))?

ane equation

1 + x6 + y 6 xy 5 = 0.

What is

CHAPTER 25

The Riemann-Roch Theorem

As you know, there are no nonconstant holomorphic functions on a

p but
1
no poles anywhere else? Then you still get nothing, unless M is P (in
1
which case there is (zz(p)) ). This is because for g = genus(M ) 1,
there is a nonzero holomorphic form which doesn't vanish at p. For
P
any meromorphic function f on M , we know that
qM Resq (f ) = 0;
so if f has a simple pole at p, then Resp (f ) 6= 0 and f must have

Riemann surface

M.

What if we allow a simple pole at one point

another pole to cancel this term.

p (but still no
other poles)? Then the answer is more complex; if g = 0 or 1 there are
nonconstant such functions (e.g. the Weierstrass -function), while if
g 2 it can depend on the point p. In general, the vector spaces of
functions f K(M ) with a single pole (at p) with p (f ) k has
dimension max{1, k g + 1}. You are guaranteed to get something
nonconstant as soon as k g + 1 2.
What if we are prepared to allow a double pole at

In the 1850's, Riemann proved a more general inequality which replaces

(and

k)

by multiple points and orders; a decade later, his

student Roch turned this into an exact equality (Theorem 25.2.3 below) incorporating another term related to meromorphic 1-forms.
encompasses the equality

dim( (M )) = g

It

and gives a powerful tool

for studying embeddings of Riemann surfaces into higher dimensional


projective spaces, among other things.

Its statement is in terms of

spaces of functions and forms related to divisors, and we will start


(25.1) by dening these spaces precisely.
You may prefer this shorter introduction to the topic from a lecture
by Lefschetz: Well, a Riemann surface is a certain kind of Hausdor
space. You know what a Hausdor space is, don't you? Its also compact, ok. I guess it is also a manifold. Surely you know what a manifold

273

274

25. THE RIEMANN-ROCH THEOREM

is. Now let me tell you one nontrivial theorem, the Riemann-Roch Theorem.

25.1. Eective divisors and rational equivalence

P
mp [p] and E = pM np [p]
divisors on M . (Of course, only nitely many mp and np are nonzero.)
If for all p mp np , then we write D E .
Let

D=

be a Riemann surface,

Definition 25.1.1.

pM

D Div(M )

Example 25.1.2. The divisor

is eective

()

D 0.

of a holomorphic

1-form

is

eective. (Why?)
We can use this idea to put constraints on meromorphic functions

D = 3[q] 2[r],

f K(M ) with
divisor (f ) =
pM p (f )[p]. Then imposing the inequality (f )+D 0
forces q (f ) + 3 0 and r (f ) 2 0; that is, f is allowed a pole of
order no worse than 3 at q , and must have a zero of order at least
2 at r. Likewise, if K1 (M ) then () D means has a zero of
order at least 3 at q , and is allowed a pole of order no worse than 2 at
r. The next denition formalizes this and denes the quantities which
and forms. For instance, suppose

and

the Riemann-Roch theorem will relate.

D Div(M ),

Definition 25.1.3. For any

L(D) := {f K(M ) | (f ) + D 0} {0},


I(D) := { K1 (M ) | () D} {0}.
(The  {0} just means that the zero-function is included, so as to
produce a vector space.) Set

`(D) := dim L(D) ,

i(D) := dim I(D).

The next step is to dene an equivalence relation on divisors which


is ubiquitous in algebraic geometry.
Definition 25.1.4. Divisors

alent i there exists

f K(M )

1from A Beautiful Mind by S. Nasar


2by Chow's theorem all meromorphic

D, E Div(M )
with

are rationally equiv-

(f ) = D E ;

we write

rat

D E.

functions are rational, hence the terminology

rational equivalence (sometimes also called linear equivalence).

25.1. EFFECTIVE DIVISORS AND RATIONAL EQUIVALENCE

Proposition 25.1.5. If

rat

D E,

275

then

deg(D) = deg(E);
(ii) L(D)
= L(E);
(iii) I(D)
= I(E); and
(iv) `(D) = `(E) and i(D) = i(E).
(i)

Furthermore,

rat

respects the abelian group structure of

Div(M ).

D E = (f ). Now Exercise
(i). Given g L(D),

Proof. By assumption

deg((f )) = 0,

which yields

3.2 says that

(f g) + E = (f ) + (g) + E = (g) + D 0 ;
so

g 7 f g

denes a map

L(D) L(E),

and

h 7

h
denes an inverse
f

map. This gives (ii), and (iii) is done in the same way. (iv) obviously
follows from (ii)-(iii). The last statement about
that

rat

is essentially just

(D + (f )) + (E + (g)) = (D + E + (f g)).

Remark 25.1.6. In fact, the Picard group


P ic(M )

of

20.1

is just

the group of equivalence classes

Div(M )
rat

So Proposition 25.1.5(i,iv) can be thought of as saying that

dene functions from

P ic(M )

to

deg, `,and

Z.

K Div(M ) is just the


divisor of any meromorphic 1-form K (M ). Since any two such are
Definition 25.1.7. A canonical divisor

rationally equivalent (easy exercise), there is a single canonical divisor

class

[K] P ic(M ).

The next (basic) result is sometimes called Brill-Noether reciprocity:


Proposition 25.1.8. Let

D Div(M )

be arbitrary, and

canonical divisor. Then

I(D)
= L(K D),
and so

i(D) = `(K D).

K = (); if (f ) + K D 0, then (f ) = (f ) + K
D K + K = D. So f 7 f maps L(K D) I(D), and 7
gives an inverse.

Proof. Let

276

25. THE RIEMANN-ROCH THEOREM

25.2. Proof and statement

d
projective algebraic curve with ODP singularities S = {p1 , . . . , p }.

e
Let M := C
P2 ((M ) = C ) be its normalization, with 1 (pi ) =
{qi , ri }, and dene a divisor
Throughout this section we take

to be an irreducible degree

X
E := (S) =
[qi ] + [ri ] Div(M )
1

i=1
of degree

2 .

H P2

Given any line

( H  for hyperplane), write

H := 1 (H C) Div(M )
for the intersection divisor (of degree

d).3
m N,

Lemma 25.2.1. For all suciently large

`(mH E) md 2 g + 1
and

i(mH E) = 0,
where

g=

(d1)(d2)
2

is the genus of

M.

R S31 and F S3d for the dening


H and C (resp.). Consider the map

Proof. Write

polynomials of

homogeneous

S3m (S) L(mH E)




G

G 7
.
Rm
By Study's lemma,

and so

G
Rm


0

G|C 0

F |G,

ker = F S3md ( S3m (S)).

Therefore, taking dimensions of

L(mH E) im(),

3If H

passes through an ODP

pi ,

then

qi and ri

will both show up in

multiplicities determined by the local intersection multiplicities of


local analytic components of

at

pi .

(See Def. 12.2.1  )

H,

with

with the two

25.2. PROOF AND STATEMENT

277

we nd


`(mH E) dim(im()) = dim

S3m (S)
F S3md (S)

= dim S3m (S) dim S3md .


By Prop. 24.2.1(c), this

(m + 1)(m + 2)
(m d + 1)(m d + 2)

2
2
d(d 3)
= dm
2
(d 1)(d 2)
+1
= dm
2
= dm 2 g + 1.


Lemma 25.2.2. Let

D Div(M ), p M .

Then

0 `(D + [p]) `(D) (i(D + [p]) i(D)) 1.


Proof. First note that

L(D) L(D+[p]) = `(D+[p])`(D)

0.
P
D = qM nq [q],
f K(M ) satisfying

Next, writing
function

(f ) + D + [p] 0

(25.2.1)
If

f, g

an element of

and

so that

f g L(D).

So

:= limxp

f (x)
, we have
g(x)

g) np

`(D + [p]) `(D) 1,

and we conclude

0 `(D + [p]) `(D) 1.

(25.2.2)
Similarly, writing

for a canonical divisor,

0 `(K D) `(K D [p]) 1


or equivalently
(25.2.3)

is a

p (f ) = (np + 1).

are two such functions, then setting


ordp (f

L(D + [p])\L(D)

0 i(D) i(D + [p]) 1.

Altogether,

0 `(D + [p]) `(D) + i(D) i(D + [p]) 2

278

25. THE RIEMANN-ROCH THEOREM

and we just have to show that  2 is impossible.

f satises (25.2.1),
K (M ) satises

Suppose (for a contradiction) that


equivalent to  1 in (25.2.2), and

() D

which is

and

p () = np ,

which is equivalent to  1 in (25.2.3). Then

(f ) = (f ) + () [p]
with

p (f ) = p (f ) + p () = 1.
But the sum of residues of a meromorphic form must be zero (Prop.
13.1.6(b)), so

having a single simple pole (and no other poles) is

absurd.

Theorem 25.2.3. [Riemann-Roch] Let

of genus

g, D

a divisor on

M.

be a Riemann surface

Then

`(D) i(D) = deg(D) g + 1.


Proof. By part (B) of the Normalization Theorem 3.2.1, we can

assume we are in the situation described in the beginning of the section,


with

e.
M =C

By Lemma 25.2.1, there exists

m0 Z

such that

m m0 =

`(mH E) i(mH E) md 2 g + 1.
Now for any two lines
lines in

H1 , H2 ,

we have

rat

H1 H2 ;

so if

H1 , . . . , Hm

are

then by Proposition 25.1.5(iv)

`(H1 + + Hm E) i(H1 + + Hm E) md 2 g + 1.
Taking
points

m large enough and lines through (a) all points of S and (b) all
Pm
in D , we can ensure that
i=1 Hi E D is eective, so that
H1 + + Hm E = D + [P1 ] + + [Pk ]

where

k = md 2 deg(D)

(and the

Pj

are points of

M ).

Therefore

we have

` D+

k
X
j=1

!
[Pj ]

k
X
i D+
[Pj ]
j=1

!
k + deg(D) g + 1.

25.2. PROOF AND STATEMENT

279

Repeatedly applying the right-hand inequality of Lemma 25.2.2 gives

k
X
k + `(D) i(D) ` D +
[Pj ]

!
i D+

j=1

k
X

!
[Pj ]

j=1

and we conclude that

`(D) i(D) deg(D) g + 1.

(25.2.4)

Next we show the reverse inequality. Plugging

K D

into (25.2.4),

we have

`(K D) i(K D) deg(K D) g + 1


which becomes (using Brill-Noether reciprocity)

i(D) `(D) 2g 2 deg(D) g + 1 = (deg(D) g + 1),


so that

`(D) i(D) deg(D) g + 1.



Amongst the easy corollaries of this theorem are the Riemann in-

equality

`(D) deg(D) g + 1,
and (by putting

D=0

in the theorem) the formula

dim 1 (M ) = g.
Here is one more simple application:
Proposition 25.2.4. Up to isomorphism,

surface of genus

P1

is the only Riemann

0.

0; then rst of all the above corollary


1
of Riemann-Roch says that dim (M ) = 0. If we take (for some
p M ) D = [p], then I(D) 1 (M ) = {0} = i(D) = 0. So by
Proof. Suppose

has genus

Riemann-Roch,

`(D) = deg(D) g + 1 = 1 0 + 1 = 2.
Now

L(D)

consists of functions with a simple pole allowed at

(and

L(D); and since


dim L(D) = 2 there is also a nonconstant function f L(D), which by

no other poles). The constant function

belongs to

Liouville must have the allowed simple pole. Therefore the mapping

280

25. THE RIEMANN-ROCH THEOREM

degree of

f : M P1

is (cf.

14.1)

deg(f ) = deg(f 1 ([])) = deg([p]) = 1;


that is,

is an isomorphism.

Exercises
(1) Check that any two canonical divisors on a Riemann surface are
rationally equivalent.

D Div(M ), g = genus(M ). Prove that if deg D > 2g 2,


then i(D) = 0. Likewise show that if deg D < 0, then `(D) = 0.
Let M be a genus g Riemann surface, and p M . Using RiemannRoch, nd the smallest value of k for which there must exist f
K(M ) having a pole at p of order no worse than k (i.e. p (f )
k ), and no other poles.
1
Let M have genus g 2. (a) Prove that M has a morphism to P
of degree g + 1. [Hint: use Exercise (3)] (b) Prove that M has
1
a morphism to P of degree g . [Hint: let p M , and look at
i((g 2)[p]). This is a bit harder than (a).]
Assume D > 0. By Exercise (2), if g 1 then i(D) = 0, and
Riemann-Roch becomes `(D) = deg(D) for g = 1 and deg(D) + 1
for g = 0. Prove this directly (a) for M
= P1 and (b) for M
= C/
(1-torus).

(2) Let

(3)

(4)

(5)

CHAPTER 26

Applications of Riemann-Roch, I: special Riemann


surfaces

We now focus our attention on Riemann surfaces with a degree-two


mapping to

P1 ,

starting with the case of genus 1. (The higher genus

cases can be viewed as a generalization of elliptic curves, though there


is no group law.)

The rst section begins with some general claims

which will be more thoroughly investigated in the next Chapter.

26.1. Curves of genus 1


For the proof of Riemann-Roch (Theorem 25.2.3), we needed to use
the (unproved) Normalization Theorem 3.2.1(B). It's actually possible
to argue the other way, from Riemann-Roch to the existence of plane
projective immersions (with ODP singularities) for arbitrary Riemann
surfaces.
When can we do better? The degree-genus formula tells you that
(d1)(d2)
,
only curves of genera 0, 1, 3, 6, 10, 15, . . . (numbers expressible as
2
d N) can ever be embedded as smooth curves in P2 . There is no reason to believe, from this or from the Normalization Theorem, that an

arbitrary Riemann surface of one of these genera can be so embedded. In fact, it isn't true once you get to genera

6, 10, 15, . . ..

That it

works for genus 1 and genus 3 (almost; see next Chapter) is a bit of
a miracle!
So: if you buy that any genus 1 Riemann surface is a complex 1torus and any torus can be Weierstrassed into
isn't surprising.

P2 , the following result

On the other hand, it shows that Riemann-Roch is

powerful and gives us a hint of how we might prove similar results in


higher genus (e.g.,

and

Theorem 26.1.1. Let

3)
M

later.

be a Riemann surface of genus one. There

exists an injective morphism of complex manifolds


image

(M )

a smooth algebraic curve of degree


281

3.

: M , P2

with

282
26. APPLICATIONS OF RIEMANN-ROCH, I: SPECIAL RIEMANN SURFACES

p M,

Proof. Given

we know that

i(2[p]) = 0 = i([p])

by Exer-

cise 25.2, so that Riemann-Roch yields

`(2[p]) = deg(2[p]) g + 1 = 2 1 + 1 = 2,
`([p]) = deg([p]) g + 1 = 1.
In terms of the spaces of meromorphic functions, this says that

L(2[p]) ( L([p]) = L(0) = O(M ),


where

dim L([p]) = 1 means L([p]) consists of constant (or equivalently,

holomorphic) functions. Therefore, we have an element

x L(2[p])\L([p]),
i.e. a meromorphic function with a double pole at

Regard

as a morphism

M P1 .

p and no other poles.

By Riemann-Hurwitz, the

ramication degree

deg(Rx ) = 2(deg(x) + g 1) = 2(2 + 1 1) = 4,


p (x)

whereas the ramication indices


all

2.

for a degree two mapping are

Hence, the ramication divisor is of the form (cf.

14.1

for

notation)

Rx = [p1 ] + [p2 ] + [p3 ] + [p4 ]


ai = x(pi ) P1 . The {ai } are still
distinct points: by the form of Rx , [pi ] must occur with multiplicity
1
1
two in x ([ai ]); and since (deg(x) = 2 = ) deg(x ([ai ])) = 2, the
1
only possibility is x ([ai ]) = 2[pi ].
Now clearly one of the pi , say p4 , has to be p (as x has a double
pole there). So also a4 = , and we have the picture:
with

p1 , p2 , p3 , p4 M

distinct. Set

p1

p2

p3

a1

a2

a3

x
1

In the following,

{pi }

resp.

{ai }

means

i = 1, 2, 3.

26.1. CURVES OF GENUS 1

Next, notice that

x ai

283

is a local coordinate about

ai

P1 .

on

The

pi is simply that there is a local


pi on M such that1 zi2 = x ai . Dif-

meaning of a degree-2 ramication at


(holomorphic) coordinate about

loc

dx = 2zi dzi .

ferentiating gives

Again using Riemann-Roch, we have

a nonvanishing holomorphic form

1 (M ),

and by the Residue

Theorem

0=
Writing
(locally)

z :=

= dz ;

Resq (x ) = Resp (x ).

qM

p,

we have

+ h(z) (h

holomor-

for a local holomorphic coordinate at

since the residue vanishes,

loc 1

x=

loc

z2


2
1
2
0
term. Taking dierentials, dx =
+
h
(z)
dz . Put
phic) has no
3
z
z
together with the previous local computation, this tells us that dx has
divisor

(dx) =

3
X

!
[pi ]

3[p].

i=1

y0 := dx
K(M ) .

(y0 ) = (dx). If we put

Set
that

In light of the fact that

() = 0,

we have

3
Y
g(x) :=
(x ai ),
i=1

P3
2
then (g(x)) =
i=1 2[pi ]) 6[p] = 2(y0 ) = (y0 ). We
i=1 (x ai ) = (
g(x)
conclude that
has trivial divisor and so is some constant C , and
2
y0
dene y := y0 C so as to have
P3

y 2 g(x) = 0
on

M.

: M P2 for the morphism


dened by sending p 7 [0 : 0 : 1] and all other points q 7 [1 :
x(q) : y(q)]. The image (M ) is contained in the projective closure E
2
2
of {y g(x) = 0} (in P ), which is smooth due to distinctness of the
{ai }, and connected due to its irreducibility. By the usual arguments,
(M ) is open and closed in E , hence equals E . At this point we have
Now for the embedding.

1For

Write

2
zi it means
p that x ai = zi hi (zi ) where
h doesn't vanish at 0; and then we can put zi := zi hi (zi ).
2the coecient of 1 can be achieved by rescaling if needed
z2
an arbitrary choice of local coordinate

284
26. APPLICATIONS OF RIEMANN-ROCH, I: SPECIAL RIEMANN SURFACES

a diagram

C \ {p}

(26.1.1)

E \ {[0 : 0 : 1]} 

/ P2 \ {[0
PPP
lll
PPP
lll
l
PPP
|E
l
l
x
PPP 
lll
P'
lll
u
l
1
/ /

: 0 : 1]}

([Z : X : Y ]) := [Z : X].
is not injective, there exist

where
If

distinct points

q1 , q2 M \{p}

such

that

(q1 ) = (q2 ) =: Q;
applying

to this gives

x(q1 ) = x(q2 ) = (Q) =: ,


is not or one of the {ai }. Since deg(x) = 2, deg(x1 ([])) =
2 and we must have x1 () = {q1 , q2 }. From the equation for E it is evp
1
g())
ident that deg(|E ) = 2 also, with (|E ) () consisting of (,
p
g()). Clearly one of these points has to be Q. From
and (,
(26.1.1), it is also clear that q1 , q2 are the only points of M that can go
to these points. So whichever is not Q cannot get hit and fails to be
surjective, a contradiction.

in which

26.2. Hyperelliptic curves


Above we used the fact, for a genus one Riemann surface

M,

that

`(2[p]) = 2 > 1 = `([p]) for p M , to construct a degree-two mapping


x : M P1 . Now suppose M has genus 2: how to map it to P1 ? Well,

1
we have a basis {1 , 2 } (M ), and 2 produces a meromorphic
1
function, which does the job. By Poincar-Hopf, deg((1 )) = 2g 2 =
2, and so this map has two simple poles (or one double pole), hence
has degree two.
In terms of homogeneous coordinates, we might write

p 7 [1 (p) : 2 (p)],
loc

i = fi (z)dz in
terms of a local coordinate vanishing at p) simply [f1 (0) : f2 (0)]. If both
fi could simultaneously equal zero we would have a well-denedness

where the meaning of the right-hand side is (expressing

problem (which could be gotten around by taking a limit), but this


does not happen: we would have to have

i([p]) 2.

By Riemann-Roch

26.2. HYPERELLIPTIC CURVES

285

`([p]) = deg([p]) g + 1 + i([p]) 2, thereby producing an


1
isomorphism M P as in the proof of 25.2.4, contradicting g = 2.
this yields

This discussion hopefully motivates

M is hyperelliptic
1
morphism x : M P .

Definition 26.2.1. A Riemann surface

exists a (nonconstant) degree-two

Clearly, any genus

Now, let

i there

Riemann surface is hyperelliptic.

be hyperelliptic of any genus and consider what the

Riemann-Hurwitz formula has to say when applied to

x:

M = 2P1 rx
X
2 2g = 2 2
(p (x) 1),
pM
where

deg(x) = 2 = p (x) 2.

ramication points, and this is just

So the sum equals the number of

2g + 2:

Rx = [p1 ] + + [p2g+2 ].
P1 if necessary, we may as1
sume that none of the x(pi ) =: ai are 0 or . Put x ([]) =: [p] + [q]
1
0
0
and x ([0]) =: [p ] + [q ]. We have the picture
x

By composing

with an automorphism of

p1

a2

: M M denotes
M over P1 (cf. Exercise 1).
Lemma 26.2.2. Let

a2g+1 a2g

the involution exchanging the branches

(+1)-

and

(1)-

eigenspaces of

polynomial

m(t) = t 1.

diagonalizable.

= 1.

V =V

J.

Proof. With respect to any basis for

J: V
V into

be a nite-dimensional vector space,

an involution. Then we have a decomposition

the

a1

in which

p2g+1 p2g

P
of

p2

V,J

is a matrix with minimal

This has no repeated roots, and so

Moreover, since

J =

idV , any eigenvalue

is

satises

286
26. APPLICATIONS OF RIEMANN-ROCH, I: SPECIAL RIEMANN SURFACES

We apply this to the pullback map

: 1 (M ) 1 (M ).

Notice

1 (M )+ = {0} since such forms would be pullbacks of holomorphic


1
1
1

forms from P (cf. Exercise 2). Hence (M ) = (M ) and so


that

=
1 (M ).
Put D = (g + 1)[p] + (g + 1)[q] Div(M ). We have (Exercise 25.2)
i(D) = 0, so that Riemann-Roch gives

for all

`(D) = 2g + 2 g + 1 = g + 3.
Now apply the Lemma again, this time to
that

L(D)+

: L(D) L(D),

noting

contains the linearly independent set

{1, x, x2 , . . . , xg+1 }.
In fact,

LM (D)+ = x LP1 ((g + 1)[]),


|
{z
}
polynomials of
degree g + 1

and so the above set is a basis. Therefore

dim(L(D) ) = `(D) dim(L(D)+ )


= (g + 3) (g + 2) = 1,
y L(D) such
P2g+2
(y) = i=1 [pi ] D.

and there exists a nonzero


Claim 26.2.3.

Proof. Since the

and so

pi

that

y = y .

are ramication points,

(pi ) = pi

. But then

y(pi ) = ( yi )(pi ) = y((pi )) = y(pi )


P2g+2
y(pi ) = 0. That is, y 1 ([0]) i=1
[pi ], which implies
X 
deg((y)) = deg(y 1 ([0])) deg
[pi ] = 2g + 2.

On the other hand,

y L(D) = y 1 ([]) D =

deg((y)) = deg(y 1 ([])) deg(D) = 2g + 2.


So

deg((y)) is forced to equal 2g + 2,


[pi ] and y 1 ([]) = D.

3using
(e.g.

which means also that

y 1 ([0]) =


subscripts to denote which Riemann surface we are considering functions on

LM (D)

just means

L(D))

26.2. HYPERELLIPTIC CURVES

Set

287

2g+2

g(x) :=

(x ai ) K(M ) ,

i=1

Div(M ))
X
X
X
(g(x)) =
((xai )) = 2
[pi ](2g+2)x1 ([]) = 2
[pi ]2D = (y 2 ).

and compute (in

y 2 /g(x) has trivial


that (in K(M ))

But then
we have

divisor, and so is a constant. Rescaling

y,

y 2 g(x) = 0.
By considering the image of

: M P2
given by

m(6= p, q) 7 [1 : x(m) : y(m)]


and

p, q 7 [0 : 0 : 1],
we arrive at:
Theorem 26.2.4. Hyperelliptic Riemann surfaces are precisely the

normalizations of (plane) algebraic curves of the form


2g+2

(
Y 2 Z 2g =

(X ai Z)

P2 .

i=1

A basis of

1 (M )

is given by

j :=

Proof. We just need to show

xj1 dx
,
y

j = 1, . . . , g.

is holomorphic:

(j ) = (j 1)(x) + (dx) (y)


= {(j 1)([p0 ] + [q 0 ]) (j 1)([p] + [q])}
nX
o
+
[pi ] 2([p] + [q])
nX
o

[pi ] (g + 1)([p] + [q])


= (j 1) ([p0 ] + [q 0 ]) + (g j) ([p] + [q]) 0.

4note:

the singular point

[0 : 0 : 1] is not an ODP, so the construction of g

holomor-

phic dierentials that follows shouldn't be compared with the formulas you know
in that case. Also, it should be emphasized that the

ai

are distinct.

288
26. APPLICATIONS OF RIEMANN-ROCH, I: SPECIAL RIEMANN SURFACES

A hyperelliptic curve, by the way, is just an irredicuble projective


curve whose normalization is a hyperelliptic Riemann surface!
The rst two exercises below are ones you could have done long ago,
but ll in (very) small gaps in the proofs above. The same goes for the
third, if you had known the denition of hyperelliptic! The fourth does
make heavy use of Riemann-Roch.

Exercises

: M M 0 of compact
Riemann surfaces, with corresponding involution dened as follows: if p M is a ramication point of , (p) := p; otherwise,
1 ((p)) = {p, p} and (p) := p. Clearly = IdM (i.e. is an
involution) and = . Prove that : M M is a holomorphic map of Riemann surfaces. Since is injective and surjective
(why?), it follows that Aut(M ).
1

Continuing Exercise (1), let (M ) satisfy = . Prove

1
0
that = for some (M ).
2
Suppose that a dth-degree irreducible algebraic curve C P has
a point of multiplicity (d 2). Show that C is hyperelliptic.
Let M be a Riemann surface of genus two. In this problem you
will construct a realization of M as an algebraic curve, dierent to
that produced above. You will need to use that M is hyperelliptic,
1
with x : M P its degree-two mapping and the associated
involution. Take p and q (in contrast to the notation above) xed
non-ramication points on M with distinct images under x; let
and denote arbitrary points of M .
(a) Prove that `([] + []) = 1 unless () = . [Hint: otherwise

(1) Given a degree

(2)

(3)

(4)

holomorphic map

you get a dierent involution (why?). To see why this is a problem


you might consider the fact that

for all holomorphic

-symmetric.]
(b) For any points , on M , show i([] + [] [p] [q]) = 1 (as
opposed to 2) {, } =
6 {p, q}. [Hint: use (a), and consider
`([] + [] [p] [q]).]
2
(c) Use I([p] [q]) to construct a map : M P . [Hint:
compute i([p][q]).] You will need to check that is well-dened.
[Hint: compute i([p] [q] + []), using Exercise 25.2.]
forms implies their divisors are

EXERCISES

289

is injective o {p, q}, but that (p) = (q). [Hint:


using part (b), compute i([] + [] [p] [q]).]
(e) Show that there exists a meromorphic form I([p] [q])
with poles at both p and q .
(f ) Explain why the zero-divisor ()0 (the eective part of ())
1
2
is
of the intersection [divisor] of a line in P with C := (M ).
Prove that deg(()0 ) = 4 (easy). Assuming C is an algebraic curve
(this is GAGA), conclude that deg(C) = 4.
(g) Clearly (p) = (q) is a singularity of C . Prove it is the only
(d) Show that

one, and a double point. [Hint: assume otherwise, and produce a


genus zero normalization or similar.]

CHAPTER 27

Applications of Riemann-Roch, II: general


Riemann surfaces

Next we'll use Riemann-Roch to develop two methods for mapping


an arbitrary Riemann surface into a (usually higher-dimensional) projective space, with a nice application to curves of genus three.

The

second approach behaves dierently in the hyperelliptic and nonhyperelliptic cases, so we rst will want to convince ourselves that there are
nonhyperelliptic Riemann surfaces! To see this, we will start with an
heuristic argument for the number of complex parameters governing
Riemann surfaces, and show that the hyperelliptic ones have fewer parameters. But there's much more in this chapter, which should give a
glimpse of how rich the correspondence between algebraic curves and
Riemann surfaces really is.

27.1. Moduli
In algebraic geometry there is the notion of moduli spaces, which
parametrize structures of a prescribed sort modulo some equivalence
relation, such as smooth algebraic curves of degree
equivalence or Riemann surfaces of genus

up to projective

up to isomorphism. A

main point is that these spaces can be given algebraic structure themselves, i.e.

turned into algebraic varieties, in many cases.

Suitably

rened, the structure of these varieties is one of the hotter topics of


study around.

We shall only be concerned with the notion of moduli as a set of


local parameters (on the moduli space), and will say colloquially that
some structure has a certain number of moduli : e.g. genus

(resp.

0)

Riemann surfaces have one modulus (resp. zero moduli) since they can
all be expressed as

1so-called

C/Z h1, i (resp. P1 ) up to isomorphism.

Underlying

modular curves or modular varieties are a more specialized notion

with an arithmetic avor


291

292
27. APPLICATIONS OF RIEMANN-ROCH, II: GENERAL RIEMANN SURFACES

26.1 that not all Riemann surfaces of genus 6, 10, 15, etc.
2
embedded smoothly in P is a deep calculation of Riemann:

the claim in
can be

Theorem 27.1.1. Riemann surfaces of genus

phism) have

3g 3

[p]) = 0

(up to isomor-

moduli.

Proof. (Sketch) Consider a genus

eective divisor

g2

2g

of degree

for any point

p M.

on

M.

M , and any
i(D) = i(D

Riemann surface

By Exercise 25.2,

So Riemann-Roch says

`(D) = deg(D) g + 1 = g + 1,
`(D [p]) = g.
f L(D) and not in any of the nitely many L(D
[p]) for those p appearing in D; and so f 1 ([]) = D, which means
deg(f ) = deg(f 1 ([])) = 2g . Now Riemann-Hurwitz tells us about
Hence there exists

the ramication behavior:

M = deg(f ) P1 deg(Rf )
2 2g = 2g 2 rf
rf = 6g 2.
For almost all

the points in

Rf

will have multiplicity one (rami-

cations of order two) and lie over distinct points in


the branch locus

B P1

consists of

6g 2

P1 ,

meaning that

points. We want to use all

this data to compute the number of local deformation parameters of

M.
Look at this in a slightly more formal way, consider the set

S1

(M, f ) where M has genus g and f has degree 2g . This


maps to the set S2 of 2-tuples (M, D) where D 0 of degree 2g (take
D := f 1 ([])). From there you can map to the set S of Riemann
surfaces of genus g , by forgetting D . It's fairly clear that (xing M ) D
has 2g parameters, making dim(S2 ) dim(S) = 2g . Moreover, given
M and D, there are `(D) = g + 1 choices of parameter for f (to have
D as its poles), meaning dim(S1 ) dim(S2 ) = g + 1. Our argument
of 2-tuples

in the rst paragraph shows that the rst map is surjective (while the
second obviously is), and so
(27.1.1)

dim(S) = {dim(S1 ) g 1} 2g = dim(S1 ) 3g 1.


|
{z
}
dim(S2 )

27.1. MODULI

293

S1 to S3 , the set of (6g 2)-tuples


1
1
of (unordered) points on P , by taking f (Rf ) Div(P ). This map
1
is surjective since given a branch-point set in P you can construct an
On the other hand, you can map

2 and in fact

existence domain for an appropriate algebraic function,

the construction shows that there are only nitely many possibilities
for

M.

Moreover, it shows that a continuous family of degree-2g func-

tions on

with the same branch-point set gives rise to a continuous

M.

family of automorphisms of

But for

g 2 M

has only nitely

many automorphisms. So we see that this map is nite-to-1, and thus

dim(S1 ) = dim(S3 ) = 6g 2.

Plugging this in to (27.1.1), we get the

desired result.

It's much easier to count moduli for hyperelliptic and algebraic


plane curves.
Proposition 27.1.2. Hyperelliptic Riemann surfaces of genus

(up to isomorphism) have

2g 1

moduli.

Proof. They are essentially just the existence domains of the al-

qQ
2g+2

i=1 (z i ), and so are completely determined


2g+2
by the branch locus {i }i=1 . This has 2g + 2 parameters, but we have
1
to account for change of coordinate on P , which is by P GL2 (C), by
gebraic functions

subtracting

dim(P GL2 ) = 3.

Proposition 27.1.3. Smooth algebraic curves of degree

projective equivalence have

d+2
2

dimension

coordinates by

modulo

moduli.

Proof. A curve is determined by a polynomial in


d+2

S3d ,

which has

We have only to account for changing projective

GL3 (C),

which has dimension

9.

(Here

P GL3

is not

what is wanted, as we do want to quotient out the rescalings of the


equation.)

Now we can compare moduli, with two very interesting results.


First, consider the numbers you get for general Riemann surfaces of

1, 2, 3, 4, 5, 6 : the numbers of moduli are 1, 3, 6, 9, 12, 15. For


hyperelliptic ones, we have instead 1, 3, 5, 7, 9, 11. So while all genus

genus

2 Riemann surfaces are hyperelliptic, we have:

2like p(z a)(z b)(z c)(z d) (for g = 1), but more complicated (since g 2);
cf. [Griths and Harris], pp. 255-257

294
27. APPLICATIONS OF RIEMANN-ROCH, II: GENERAL RIEMANN SURFACES

Proposition 27.1.4. A general Riemann surface of genus

g 3

is non-hyperelliptic.
So we will need a more general method of realizing Riemann surfaces
as algebraic curves than that discussed in

26.2,

and that is what we

will do in the remainder of this Chapter.


Finally, look at those genera which correspond to nonsingular algebraic curves in

P2 of degree 3, 4, 5, 6, 7, . . .:

namely,

1, 3, 6, 10, 15, and

so on. The Riemann surfaces of these genera have (by Thm. 27.1.1)
moduli

1, 6, 15, 27, 42,etc.

But now look at the smooth algebraic

plane curves of the corresponding degrees (via Prop. 27.1.3): we get

1, 6, 12, 19, 27.

The case of genus

will be treated in

27.3.

Beyond

that, we have:

Proposition 27.1.5. Smooth algebraic plane curves of degree

d5

(d1)(d2)
do not yield all the Riemann surfaces of genus
 only a
2

special subset.

27.2. Projective embeddings


Let
degree

M be a Riemann surface of genus g . For any D Div(M ) of


> 2g 2, recall from Exercise 2 of Chap. 25 that i(D) = 0. By

Riemann-Roch, we then have

`(D) = deg(D) g + 1.
This will be used repeatedly in the argument below.
Now x a divisor

D=

pM

np [p] Div(M )

of degree

d 2g + 1.

We will dene an embedding (injective morphism of complex manifolds)

: M , Pdg .
d g g + 1, this can only give an embedding in P2 for g = 1.)
The support of D , which is the subset of M consisting of the points
aapearing in D (i.e. those p with nonzero np ), is written |D|.

(Since

3It's

very important to understand the argument in this section. Try slimming it

down (I've expressed it in a somewhat bloated manner) and writing it out for a
specic choice of

and

g (> 1,

say). (Also, if you are stuck on Chap. 26 Exercise

4, some of the steps are similar.)

27.2. PROJECTIVE EMBEDDINGS

295

d > 2g 2 and so `(D) = d g + 1.


basis of L(D), and dene for p
/ |D|

First o, certainly

{f0 , . . . , fdg }

(p) := [f0 (p) : : fdg (p)].

(27.2.1)
If

for a

Write

p |D|,

this is unsuitable since some functions may blow up (or

all functions may be required to vanish).


coordinate (vanishing at

Therefore if

is a local

to rst order), we put

(p) := [(z np f0 )(p) : : (z np fdg )(p)].

(27.2.2)

q in a neighborhood of p, [(z np f0 )(q) : : (z np fdg )(q)] gives


the same result as [f0 (q) : : fdg (q)], and so we have constructed
an analytic map . . . provided that (27.2.1-2) do not yield [0 : : 0]

For points

at any point. That is the central well-denedness issue, and we must


check it.

Now for

p, q M ,

each still have degree

notice that

> 2g 2.

D [p], D 2[p]

and

D [p] [q]

Therefore we have

`(D [p]) = d g
and

`(D 2[p]) = d g 1 = `(D [p] [q]),


with the immediate consequences
(27.2.3)

L(D [p]) ( L(D),

(27.2.4)

L(D [p] [q]) ( L(D [p]),

(27.2.5)

L(D 2[p]) ( L(D [p]).

p, q
/ |D|. Then
(27.2.3) says that there exists f L(D) not vanishing at p, meaning
that the {fi (p)} are not all zero; this makes well-dened on M \|D|.
Next, (27.2.4) gives us g L(D [p])\L(D [p] [q]), a function
vanishing at p but not q , forcing to take dierent values at p and q ;
hence is injective on M \|D|. Finally, (27.2.5) provides h L(D
[p])\L(D 2[p]), i.e. vanishing to exactly rst order at p, so that the
To interpret these, for simplicity rst assume

4Note

p |D|;
np = 0.

that (27.2.2) isn't just a special formula for

general than) (27.2.1) since for

p
/ |D|

we have

it contains (is more

296
27. APPLICATIONS OF RIEMANN-ROCH, II: GENERAL RIEMANN SURFACES

derivative of

hence that of

is nonzero there; together with the

M \|D| is smooth.
In order to extend these statements to all of M , we have to rene
the argument just a bit. For a general points p, q M , (27.2.3) tells
us that there exists a function f L(D) with p (f ) = np exactly;
(27.2.4) that some g L(D) has p (g) > np but q (g) = nq ; and
(27.2.5) that there exists an h L(D) with p (h) = np + 1 exactly.
injectivity result, this proves that the image of

These give precisely the well-denedness, injectivity, and smoothness


of image for the map described by (27.2.2). So the image is a compact
complex analytic curve

Pdg ,

which is algebraic by GAGA.

27.3. Canonical maps


Once again we consider a Riemann surface

g 2,

M,

this time of genus

and let

{1 , . . . , g } 1 (M )
be a basis. Instead of choosing a divisor and go through Riemann-Roch
to get a projective embedding from meromorphic functions, why not
just use these? Dene the canonical map

K : M Pg1
by

p 7 [1 (p) : : g (p)].
The meaning of this, as you would expect, is locally writing each

as

fi (z)dz , and taking [f1 (p) : : fg (p)]. This is well-dened, i.e. the
{i } do not all have a zero at p. Otherwise we would have I([p]) =
I(0) = g hence (by Riemann-Roch) L([p]) = 2, which we know to be
1
false for M not isomorphic to P .
Bottom line:

this looks quite promising, from the standpoint of

getting a convenient projective embedding. Or does it?

Example 27.3.1. For

hyperelliptic, consider the setting of The-

orem 26.2.4; we have


dx x dx
xg1 dx
K (p) =
:
: :
= [1 : x : : xg1 ].
y
y
y


27.3. CANONICAL MAPS

297

Notice that this looks a lot like the rational canonical map

from

Example 7.3.7. In fact, it factors


(27.3.1)

C @
@

@@
@@
@@


P
z<
z
zz
zz
zz f

g1

P1
with

deg(x) = 2,

and so does not give an embedding of

in

Pg1 .

All is not lost: the hyperelliptic case is very special (in a bad way),
and for

g3

we know that there are (lots of ) nonhyperelliptic curves.

Theorem 27.3.2. Let

Riemann surface of genus


(a)
(b)
(c)

K is nondegenerate;5
K (M ) is smooth; and
K is injective M

Proof. (a) Were

Pg1 ,

K be the canonical
g 2. Then

K (M )

map for an arbitrary

is nonhyperelliptic.

contained in a proper linear subspace of

this would produce a linear relation on the

{i }.

But they are

linearly independent by construction, being a basis!


(b) This is clear in the hyperelliptic case, by observing that the derivative of the (injective) rational canonical map is nowhere vanishing.
(We will return to this in the nonhyperelliptic case.)
(c) The implication 

 is already done (by contrapositive) in

Example 27.3.1.
Now let

be a local coordinate vanishing to rst order at a point

p M , and consider the linear functionals on 1 (M ) (= I(0)) given by



(p)
7
dz
 0
7
(p)
dz
.
.
.

 (k1)

(p).
dz
If the rst is zero on some given , then I([p]). If the rst and
second are zero, then I(2[p]). If all are zero, then I(k[p]).
Since k linear conditions cut out a subspace of codimension at most k ,
5cf.

the beginning of Chapter 7, and also 7.3.1.

298
27. APPLICATIONS OF RIEMANN-ROCH, II: GENERAL RIEMANN SURFACES

we have

i(k[p]) = dim I(k[p]) g k.


More precisely, we have

i(k[p]) = g k + a

and by Riemann-Roch

`(k[p]) = k g + 1 + i(k[p]) = 1 + a.
k = 1, there can be no redundancies in one linear
condition and so we have a = 0.


Suppose K ([p]) = K ([q]) for p 6= q . Then 7


(p) and 7
dz


1
(q) yield the same functional on (M ), up to a constant multiple;
dz
in particular they vanish on the same 's. So I([p]) = I([p]+[q]), which
yields i([p] + [q]) = i([p]) = g 1 and via R-R
For the special case

`([p] + [q]) = 2 g + 1 + i([p] + [q]) = 2.


Therefore there exists a nonconstant meromorphic function

F L([p]+

[q]). We know that L([p]) and L([q]) have only constant functions
(a = 0 when k = 1), and so F has to have the allowed simple pole at
both of p and q . Thus deg(F) = 2, and so M is hyperelliptic. This
completes the proof of (c).
(b, cont'd.)
have

M is non hyperelliptic.
a = 0 for k = 2. Consequently

Now assume

`(2[p]) = 1,

i.e.

Then we must

i(2[p]) = g 2
i([p]) = g 1),
coordinates at p,

(whilst
local

and we can arrange a basis of

1 (M )

so that in

loc

loc

1 = dz , 2 = zh2 (z)dz ,
loc

j = z 2 hj (z)dz (3 j g).
(Here the

hi (z)

are holomorphic, and

h2

doesn't vanish at

z = 0.)

The

canonical map takes the local form

K (z) = [1 : zh2 (z) : z 2 h3 (z) : : z 2 hg (z)]


with derivative

K (z) = [0 : h2 (z) + zh02 (z) : : : ]


which does not vanish at

p.

This gives the desired smoothness.

27.4. WEIERSTRASS POINTS

Consider a smooth, irreducible algebraic curve

299

and hyperplane

H = {W (Z) = 0}, both in Pn . (Here W is a homogeneous polynomial


in Z1 , . . . , Zn+1 of degree one, with ane form w .) One can dene an
intersection divisor C H on C in a way which extends what we have
2
done in P . If C is not necessarily smooth, then the divisor lives on

a normalization M of C and is denoted H ; it is given simply by


P

pM ordp ( w)[p]. We dene the degree of the curve to be the degree


of this divisor, called a hyperplane section :

deg(C) := deg( H).


Since any two hyperplane sections are rationally equivalent (why?),
any two hyperplane sections have the same degree, making

deg(C)

well-dened.
In the case at hand,

is

n = g 1. Hyperplane sections are


Pg
we write W (Z) =
i=1 i Zi , then

and

particularly interesting because if

H = (1 1 + + g g )
M!
K (M ) a

is a canonical divisor on
map, and its image

That's why

is called the canonical

canonical curve.

Proposition 27.3.3. Assume

Then the degree of the canonical curve

is nonhyperelliptic of genus

K (M ) P

g1

is

Proof. The assumption is necessary in order that

ize

K (M ).

g.

2g 2.
M

normal-

(In the hyperelliptic case, it is normalized by the ratio-

nal canonical map.)

deg(K) = 2g 2

deg(K (M )) = deg(K H) =
and that's it.


We then compute

by Poincar-Hopf,

And so, we nd that nearly all genus

3 curves have a nice embed-

ding into the projective plane.


Corollary 27.3.4. Every nonhyperelliptic genus

normalization of a smooth quartic curve in

curve is the

27.4. Weierstrass points


We began our discussion of Riemann-Roch with a naive analysis,
for a xed point

on a Riemann surface

M,

of what orders of pole

are possible if we are after a meromorphic function with its only pole
at

p.

To conclude, I will now briey explain the sense in which this

300
27. APPLICATIONS OF RIEMANN-ROCH, II: GENERAL RIEMANN SURFACES

can depend on the choice of

g>0

and not just the genus

of

M.

Assume

for what follows.

First note that

`(0) = 1

(constant functions), and

`([p]) = 1

by the

argument at the beginning of Chap. 25. By equation (25.2.2), we know


for each

that

0 `((k + 1)[p]) `(k[p]) 1.


(2g 1)[p] exceeds 2g 2,
i((2g 1)[p]) = 0, and so (by

On the other hand, since the degree of


we have (Chap.

25 Exercise 2) that

Riemann-Roch)

`((2g 1)[p]) = (2g 1) g + 1 = g.


More generally, for

k 2g 1,

the fact that

i(k[p]) = 0

yields

`(k[p]) = k g + 1.
`(k[p]) starts (at k = 0) at 1 and works its
way up to g in increments of 0 or 1, as k rises to 2g 1; thereafter it
increases by 1 whenever k does.
The situation with i(k[p]) is dual: it starts at g and works its way
down to 0 in decrements of 0 or 1, as k rises to 2g 1; and then it stays
at 0.
Now it turns out that at all but nitely many points, `(g[p]) = 1;
So the scenario is that

that is, all the increments are postponed as far as possible and the
sequence

`(k[p])

looks like

1, 1, 1, . . . , 1, 2, 3, . . . , g,

and so on.

Those

points where this is not the case are called the Weierstrass points of

M.

The simplest example I am aware of is, for a hyperelliptic curve,

2g + 2 xed points of the involution .


like 1, 1, 2, 2, 3, 3, etc.
the

For these the sequence looks

Exercises

K (p)
near p.

(1) Check that the denition of


choice of local coordinate

in

27.3

(2) Show that any smooth quartic curve in


genus

3),

P2

is independent of the

is a canonical curve (of

and hence also nonhyperelliptic.

(3) In this exercise you will prove a new Cayley-Bacharach type result
(in

P2 ):

if

C1

and

C2

(of degrees

and

assumed smooth and irreducible) meet in

n respectively, with C1
mn distinct points, and

EXERCISES

C3

301

m+n3) passes through p1 , . . . , pmn1 C1 C2 , then


through the remaining point p := pmn . (We shall write f1 ,

(of degree

it passes

f2 , f3

for the resp. dening homogeneous polynomials.) Start out

by assuming

C3

does not contain

p,

and follow these steps:

g denote the genus of C1 , and show that (m 3)m = 2g 2


m3
and g = dim(S3
).
f2 h
m3
(b) Let h S3
, set Fh :=
| K(C1 ) , and write (Fh ) =
f3 C 1
[p] + (h) D (this denes D Div(C1 )). Show that the map
S3m3 L(D)/C (here C =constant functions) given by h 7 Fh is
injective, and use this to put a lower bound on `(D).
(c) Find deg(D), i(D), and obtain a contradiction.
(a) Let

(4) A problem on automorphisms of canonical curves:


(a) Let

: C C

C of a linear automor
1
phism of P
. [Hint: consider the action of on (C).]
(b) Let M be a non hyperelliptic Riemann surface of genus 3 with
an involution . How many xed points does it have? What is the
genus

g.

be an automorphism of a canonical curve of

Prove that

is the restriction to

g1

genus of the quotient Riemann surface? [Hint: consider the canon-

ical embedding and apply (a);


linear automorphism on
(5)

is the restriction of what sort of

P ?]
Using the embedding of 27.2,

try to prove: (a) there exists, for an

arbitrary Riemann surface, an embedding onto a smooth curve in

P3
in

and (b) an immersion onto a curve with only ODP singularities

P2 .

[Hint: use suciently general projections of the complement

of a linear subspace in
do these curves have?

Pn

onto

P2

and

P3 .]

Also, (c) what degrees

CHAPTER 28

Abel's Theorem, part I

M a Riemann surface of genus


g 1, with closed paths (1-cycles)i giving a basis {[i ]}2g
i=1 for
H1 (M, Z). We have the Jacobian of M , which is the complex g -torus
Recall the setup from Chapter 20:

g
(1 (M ))
= C

.
J(M ) :=
H1 (M, Z)
M
The isomorphism is given by evaluating functionals against a basis

{1 , . . . , g } 1 (M ),

and

is called the period lattice. The Picard

group

Div 0 (M )
(K(M ) )

P ic0 (M ) :=

of degree-0 divisors modulo rational equivalence is the object we want


to understand. To this end, we had shown that the Abel-Jacobi map

AJ : P ic0 (M ) J(M )

D 7
1 D
is a well-dened homomorphism, where
writing some 1-chain
is that

AJ((f )) = 0

1 D

= D. The
f K(M ) .

with

for any

is just a compact way of

important content of this

By Abel's theorem we will henceforth mean the statement that

AJ

is injective, that is
(28.0.1)

AJ(D) = 0

D = (f ) for

some

f K(M ) ;

while the surjectivity will be known as Jacobi inversion: i.e.,


(28.0.2)

J(M ) [any functional on 1 (M ), up to periods],

divisor D inducing that functional via


().
1 D

given any point in


there exists a

These statements will be proved in Chap.

29.

Our aim here is just

to explain how Abel's theorem relates to Riemann-Roch and develop a


couple technical lemmas to be used in the sequel.

303

304

28. ABEL'S THEOREM, PART I

Before starting let's rene one aspect of the above picture just a
bit. Intersecting 1-cycles on

 or more precisely, intersecting trans-

verse representatives of homology classes  gives a perfect, unimodular


pairing


 : H1 (M, Z) H1 (M, Z) Z.

There is always a symplectic basis, that is one with the property that

i g+j = ij = g+j i
i j = 0 = i+g j+g
1 i, j g (where ij is the Kronecker delta). This is the situation
2g
pictured in 20.1. We will assume henceforth when writing {i }i=1 that

for

they form a symplectic basis for the rst homology.


We should also remark on what the Picard group is really doing

E , in P ic0 (E) we have [p+q][p][q]+[O]


0, where addition inside the brackets is the group law on E and outside
here. For an elliptic curve

the brackets means adding divisors. What this says is: while as divisors

E ) [p+q] 6= [p]+[q], working


modulo rational equivalence we do have [p + q] + [O] [p] + [q]. So
P ic eectively recovers the group law on E . Now, curves of higher

(i.e. in the free abelian group on points of

genus have no group law on points; but by linearizing points and


working modulo divisors of functions, we get a form of generalization
of the group law in genus

1.

Intriguingly, a more precise form of Jacobi

inversion in the next Chapter will tell us that this may almost be
seen as a group law on unordered

g -tuples

of points on

M.

28.1. From Riemann-Roch to Abel-Jacobi

D be a divisor on M ; we have been interested in the dimensions


vector spaces L(D) and I(D). In the interval

Let
of the

0 deg(D) 2g 2
is where anything of interest lies: outside this range, either

i(D)
(i)
(ii)

1each

`(D)

or

is zero. At the extremes, Abel's theorem will tell us:

`(D) when deg(D) = 0; and


i(D) when deg(D) = 2g 2.
intersection point contributes a

the orientation of

according to the right-hand rule and

given by the complex structure

28.1. FROM RIEMANN-ROCH TO ABEL-JACOBI

In case (i), if there is a meromorphic function

D 0,

305

f K(M )

with

(f ) +

then

deg((f ) + D) = deg((f )) + deg(D) = 0 + 0 = 0


=

(f ) + D = 0

In this event, there can be only one such

(f ) = D = (g)

rat

D 0.

(up to scale), as

(f /g) = 0

f /g

constant.

Together with similar reasoning in case (ii), and assuming Abel, this
argument proves
Proposition 28.1.1. (i) If

deg D = 0,

then

`(D) = 0

or

1;

and

rat

AJ(D) = 0 D 0 `(D) = 1.
(ii) If

deg D = 2g 2,

then

i(D) = 0

or

1;

and

rat

AJ(K D) = 0 D K `(D K) = 1 i(D) = 1.


Another point of contact with the last few chapters comes in the
context of canonical and hyperelliptic curves. First, x

qM

and look

at the mapping

uq : M J(M )
p
1
q.
.
p
7 AJ([p] [q]) =
.
p
g
q

Assuming Abel's theorem, we have (for genus


Proposition 28.1.2. (a)

uq

mod

M .

1)

is injective;

(b) its dierential yields the canonical map; and


(c) if

is hyperelliptic and

is a xed part of

symmetric with respect to the involution


Proof. (a) Assuming

p1 6= p2

and

u 7 u

of

, then uq (M )
J(M ).

uq (p1 ) = u2 (p2 ),

we have

AJ([p1 ] [p2 ]) = 0
Abel

f K(M )

f : M
P1

(=)

contradicting

g 1.

with

(f ) = [p1 ] [p2 ]

has degree one,

is

306

28. ABEL'S THEOREM, PART I

(b) Given

1 (M ),

we can consider

(p) Tp M.

By the funda-

mental theorem of calculus, the dierential

duq (p) : Tp M Tuq (p) J(M )


= Cg
loc

(1 (p), . . . , g (p)). (That is, if i = fi (z)dz , then duq (p)

sends
| 7 (f1 (0), . . . , fg (0)) Cg .) This associates a line in Cg to
z p
g1
each p M ; projectivizing clearly recovers K : M P
.
(c) Using ((x, y)) = (x, y), we have
p

(p)
dx
dx

y
y
q

q=(q)
.
.

.
.
=
uq ((p)) =
.
.

(p) xg1 dx
p xg1 dx
y
q
y
(q)

p
q dx
y

.
= uq (p).

.
=
.
p xg1 dx
q
y
is given by


In fact, in the hyperelliptic case it is clear from (c) that the xed
points of

map to

2-torsion

points of

J(M ).

28.2. Dierential forms of the third kind


There is a classical (and pass) terminology for meromorphic dierential forms on a Riemann surface: rst kind refers to holomorphic
forms; second kind to meromorphic forms with trivial residues (and
hence no simple poles); and third kind to everything else.

In this

section we'll pursue a method for constructing functions with a given


divisor (if possible). The title refers to the essential use we shall make
of meromorphic forms with prescribed (nonzero) residues.
Given

p, q M

i([p] [q]) = g (2) 1 + `([p] [q]) = g + 1 (> g),


|
{z
}
0

so there exists

I([p] [q])\1 (M ).
0 =

By the residue theorem,

Res () + Resq ()
| p
{z
}

both nonzero since poles simple

28.2. DIFFERENTIAL FORMS OF THE THIRD KIND

and so we can normalize

Resp =

so that

,
2 1

Resq =

1
.
2 1

, write () = ()0 () where ()0 , ()

For any meromorphic form

307

are the zero- and polar-divisors.

D Div 0 (M ), there

X
D I
[p]

Lemma 28.2.1. Given

exists

p|D|

such that

() =

[p]

and

Resp =

p|D|
Proof. Write

Resqk k =

2 1

D=

. Then

P
[pk ] [qk ], and
P
add
k =: .

ordp (D)

.
2 1

pick

so that

Respk k =


P
D = nj [Pj ] and D be as in Lemma 28.2.1, and consider
2g
a collection {i }i=1 of closed paths with support |i | M \|D|, such
2g
that their classes {[i ]}i=1 H1 (M, Z) yield a basis.
Next let

Lemma 28.2.2. If

D Z (i),

(28.2.1)

then (xing

Q M)

f (P ) := exp 2 1


D

yields a well-dened function

f K(M )

with

(f ) = D.

Cj denote

0
around the Pj . Given two paths Q.P and Q.P ,
X
X
0
Q.P Q.P = +
mj Cj +
`i i

Proof. We must check independence of path. Let

cular paths

where

is a (real-2-dimensional) closed region in

D =

mj

dD =

D =
Cj

0 = 0,

mj nj Z

M \|D|.

Now

cir-

308

28. ABEL'S THEOREM, PART I

ResPk D =

since

n
k , and
2 1

D Z

`i
i

by assumption (28.2.1). So for some



exp 2 1


Q.P D

 = exp 2 1 = 1,

exp 2 1
Q.P 0 D


and

For the divisor, let


neighborhood of

Pj

(with

be a holomorphic coordinate dened in a

z(Pj ) = 0),

loc

=
with

M \|D|.

is well-dened (and holomorphic) on

and write

n
dz
j
+ h(z)dz
2 1 z

holomorphic. Without loss of generality, we can assume that

lies in the neighborhood, with

(variable). Locally

z(Q) =: z0

f (z) = exp 2 1

(xed) and


D

Q
z

= exp 2 1

where

h(w)dw exp nj
z0

z0

= H(z)

z(P ) =: z

dw
w

exp (nj log z)


exp (nj log z0 )

is holomorphic and nonvanishing in our neighborhood (being

the exponential of something holomorphic). Finally, writing


H(z)
n , the above
z0 j
= H0 (z) z nj .
This makes it clear that

is meromorphic at

Pj

H0 (z) =

with

Pj (f ) = nj .
Doing this for each

j,

we conclude

(f ) =

nj [Pj ] = D.


EXERCISES

In the next chapter we will take


basis for

H1 ,

309

1 , . . . , 2g

to yield a symplectic

i.e.

i g+j = ij

delta)

20.1. It turns out


that the period vectors 1 , . . . , g associated to 1 , . . . , g are actually
2
linearly independent over C, and so the g g matrix they form is
invertible. Applying the inverse matrix to 1 , . . . , g , we may replace
them by {j } satisfying

for

1 i, j g.

(Kronecker

This is the situation pictured in

j = ij .
i
Given

and

as in Lemma 28.2.1, then, we can modify

f
D := D

g 
X

D i

so that

i = 1, . . . , g .

(28.2.2)

to

i=1

for

f
D = 0

We will prove that

AJ(D) = 0 =

there exists a further modication


Pg
c
f
D :=
D +
j=1 j j
with

so as to arm condition (28.2.1) (for

c
D Z (i = 1, . . . , 2g),

c
D ).

To attack (28.2.2), we need

the Riemann bilinear relations, our next topic.


The problems below are only loosely related to the material of his
chapter. The second one is rather open ended!

Exercises
(1) This problem, in which you will prove a version of Abel's theorem
for a singular cubic (not its normalization), is only loosely related

C as P1 with 0 identied to
dz
1
and coordinate z . We consider (C) to be spanned by
(even
z
1
though it isn't holomorphic) and H1 (C, Z) by S = unit curcle. Divisors must avoid the singularity, and meromorphic functions f
must have f (0) = f () 6= 0, .
to the chapter. Think of the cubic

2to

be proved in

29.1. Since the vectors are


R (which we already have).

independence over

non-real, this doesn't follow from

310

28. ABEL'S THEOREM, PART I

J(C)?
P
P
(b) Compute AJ(D) for D =
ni [zi ],
ni = 0.
(c) Show 0 = AJ((f )) J(C).
(a) What is

(d) Formulate and prove the injectivity statement (Abel's theorem). [Hint: the proof will use what you did in (c), even though
it's a converse, and so needn't be long.]
(2) Let

be a Riemann surface and

a (nonempty) nite set

of points.
(a) Dene divisors on the relative variety

ni [pi ] where no pi

(M, ) to be formal sums

; two of these are rationally equivalent

if their dierence is the divisor of f K(M ) which is 1 on all points


of . Construct an AJ map and Jacobian for (M, ). [Hint: the
1
case M = P , = {0, } should recover the results of exercise
lies in

(1).]
(b) Next consider the complement

Div(M )/Div(),

M \.

We dene divisors by

and rational equivalence by taking divisors of

meromorphic functions on

(and ignoring any poles/zeroes on

since that information is quaotiented out). Construct an


and Jacobian for

M \.

AJ

map

[Hint: note that there is no such thing as

degree of a divisor, since the points in

eectively have arbitrary

coecients. Or rather, using these points, you can make the degree
of any divisor zero! This should have some bearing on your choice
of path.]

CHAPTER 29

Abel's Theorem, part II

As suggested at the end of the previous chapter, on any Riemann

M,

surface

1 pairing on the level of homology

we can produce a perfect

h , i : H1 (M, Z) H1 (M, Z) Z

(29.0.3)

2 With respect to the basis


2g ) matrix3

by intersecting 1-cycles.
there, this has (2g

0 Ig
Ig 0

Q=

{j }2g
j=1

described

!
.

We can use this pairing to produce an isomorphism of dual spaces


(29.0.4)

H1 (M, C) = H1 (M, Z) C Hom(H1 (M, Z), C) = H 1 (M, C)


which is a special case of Poincar duality.
Recalling the isomorphisms

1
1 (M ) 1 (M ) HdR
(M, C) H 1 (M, C),
there is also a pairing (the cup-product)

H 1 (M, C) H 1 (M, C) C

1this means that the (bilinear) pairing is described (with respect to an integral basis
of

H1 (C, Z))

2More

, )

by an integrally invertible, i.e. unimodular, matrix.

precisely, any two given homology classes have representative 1-cycles (say,

which intersect transversely.

holomorphic coordinate

z = x + iy ,

p there is a local
v and v to the 1-

At each intersection point


and the tangent vectors

cycles (which have well-dened directions) can be wedged to produce an element

V2

Tp M (p 6= 0). The intersection is called positively


y
or negatively oriented depending upon the sign of p , and the intersection number
h[], []i is the number of positive intersection points minus the number of negative

v v = p x

ones.

3having

this particular intersection form is the denition of a symplectic basis of

H1 (M, Z)
311

312

29. ABEL'S THEOREM, PART II

induced on the level of 1-forms by

(, ) 7

.
M

Notice that since two holomorphic forms wedge to zero, this pairing
restricts to zero on

1 (M )1 (M ) (and 1 (M )1 (M )).

Yet another

pairing (the cap-product)

H1 (M, Z) H 1 (M, C) C

is induced by

(, ) 7

The restriction of this pairing to


the period matrix of chapter 20.

H1 (M, Z) 1 (M )

is caputred by

An important fact is that, under

(28.0.2), both of these integration-induced products are nothing but


complex-linear extensions of (28.0.1).

Assuming this compatibility, we can quickly derive the Riemann


bilinear relations as follows.

1 (M ),

If for any closed 1-form

we write

1 (M )

j () :=

,
j

then (28.0.2) identies

[] =

(29.0.5)

g
X

(j ()[j+g ] j+g ()[j ])

j=1

H 1 (M, C),
1 (M )
in

i.e.

as functionals on homology.

0=

(29.0.6)

M
by writing

g
X

One has for

(j ()j+g () j+g ()j ())

j=1

= h[], []i

and expanding both classes as in

(28.0.3). Similar reasoning together with the local computation

idz d
z = i(dx + idy) (dx idy) = i(2idx dy) = 2dx dy,
leads to
(29.0.7)

= i

0<i
M

g 
X
i=1

j ()j+g () j+g ()j () .

29.1. DERIVATION OF THE RIEMANN BILINEAR RELATIONS

313

This is all meant as motivation, though it can be made completely


rigorous.

We'll start the rst section with a more concrete, classical

proof of (29.0.4-5), without the compatibility assumptions on the three


bilinear pairings.

29.1. Derivation of the Riemann Bilinear Relations


We start by cutting

open to get the fundamental domain, a

simply-connected closed region

g+j

q
p

fundamental
domain

g+j
q
p

po

F. (Only a piece of it is shown in


1
of F be xed. Given (M ),
p
u(p) :=

with boundary
in the interior

j
q

the picture.) Let

p0

4 function on

F.

p0
then yields a well-dened (single-valued) holomorphic
If we take a second holomorphic form

1 (M ),

then

d(u) = = 0.
That is,

is a closed holomorphic form on

that

0 =
by Stokes's theorem.

with the consequence

d(u) =

u
F

Now, the picture above tells us that

is the

composition of paths

1 1
1 1
1 1
2g
g 2g g g+2
2 g+2 2 g+1
1 g+1 1 ,

4to be holomorphic on a closed set means that the function extends to a holomorphic
function on a slightly larger open set (which, in this case, would live on the universal
cover of

M ).

314

29. ABEL'S THEOREM, PART II

written right to left (with inverse meaning the reverse direction). So


the last integral becomes

g
X
=
{

(u(p) u(p )) +
{z }
j |

j=1

and, noting that

p0

q
p0

g
X

g+j

0<i
F

g+j

q
r0

r
q

this

g+j

i
, essentially the same computation yields
!

g
X
.

} =

u(i
) = i
| {z

Replacing

p
q

j=1

(u(r) u(r0 )) }
{z }
j+g |

by

d(u
)

g+j

j+g

j=1

So we have recovered (29.0.4-5).

2g
To reformulate this in matrix terms for any symplectic basis {j }j=1
g
1
th
of H1 (M, Z) and any basis {i }i=1 of (M ), notice that the (k, `)
entry of

2g

t
Q = 1

= g+1

is

g
X

0 Ig
Ig 0

2g 1

.
.
.

2g

.
.
g

2g

(j (k )g+j (` ) j (` )g+j (k ))

j=1
which is zero by (29.0.4); so

Q t = 0.

(29.1.1)
Similarly,

(29.1.2)
in the sense that

1 Q t Pi > 0

x R>0 for any x Cg .


x( 1Qt )

In particular,

the diagonal entries of (29.1.2) are positive real.

5recall

from Chapter 20 that

is the complex

g -vector

with

ith

entry

j (i )

29.1. DERIVATION OF THE RIEMANN BILINEAR RELATIONS

315

Remark 29.1.1. Consider any two symplectic integral bases

{j }

and

= {j0 }

(thought of as row-vectors), so that

0 = A
for some

A SL2g (Z).

Applying the basis

{i }

(viewed as a column-

vector of 1-forms) on the left yields

0 = A.
Furthermore, since both bases are symplectic we have

Q = t

and

Q = t 0 0 = t At A = t AQA;
that is,

A belongs to the symplectic group Sp2g (Z).

It is for this reason

that (29.1.1-2) are compatible with change of symplectic basis: e.g.,


assuming (29.1.1) we have

0 Qt 0 = AQt At = Qt = 0.

Now thinking of the

g 2g

period matrix as two

(29.1.3)

A B

gg

blocks, viz.

we have

Qt =

A B

Ig

Ig

A
t
B

!
= A tB B tA

and

= A t B B t A.

Qt
In these terms, (29.1.1) reads

A tB = B tA

(29.1.4)
while (29.1.2) becomes

(29.1.5)
If

0,

(v Cg ).

has nonzero kernel, then there exists

hence

that

v>0
1t v(At B B t A)

vA = 0

and

v = 0,
A

v Cg

satisfying

contradicting (29.1.5).

28.2.

Av =

It follows

is invertible, and so we have proved the statement on

independence asserted at the end of

C-linear

316

29. ABEL'S THEOREM, PART II

Applying

A1

to the left of

amounts to a change of the basis {i }

6
1
for (M ), viz.

A1 = A1

1
.
.
.

10

2g

.
.
.

2g

g0

g
If we apply it to (28.1.3), then we get

0 := A1 =

{i } is

Z ,

We can therefore always assume that

=
again as claimed in

28.2.
(

Ig

chosen so that

The bilinear relations (29.1.4-5) simplify to

(29.1.6)

Ig A1 B

Z = tZ
,
1(Z Z) > 0

which in particular tell us that the imaginary part

Im(Z) is a positive-

denite, real symmetric matrix.

29.2. Proof of Abel's Theorem


With the holomorphic basis as normalized above, we can now quickly
establish (28.2.2) and hence (28.0.1). Write

0)

and let

:= f
D

be as

28.2,

D=

ni [Pi ] (with

ni =

so that

(29.2.1)

ResPi () =

and

n
i
(i)
2 1

= 0 (j = 1, . . . , g).
j
For each

k = 1, . . . , g

set

uk (p) :=

k
p0

F, and let be a 1-chain (sum of paths) with = D. Then noting


P
D = ni ([Pi ] [p0 ]), we have

X
X

k =
ni uk (Pi ) = 2 1
Resp (uk )
on

6here

the product of

and

p|D|

is just the integral

29.3. PROOF OF JACOBI INVERSION

317

which by the Residue Theorem

29.1

X
(j (k )g+j () g+j (k )j ()) = g+k ().
| {z }
| {z }

uk =
F

AJ(D) = 0
every k
If

for

jk

2g
X

k =
Using

k = jk

= mk +

mj (j = 1, . . . , 2g )

then there are integers

g
X

k .

mj
j

j=1

Z = tZ

and

such that

(from (29.1.6)), this

mj+g j+g (k ) = mk +

g
X

mj+g k+g (j ).

j=1

j=1
Now

:=

g
X

mj+g j

j=1
is still an element of

I(

p|D| [p]) satisfying (29.2.1). Moreover, for

k {1, . . . , g}
k+g ()
= k+g ()

g
X

mj+g k+g (j )

j=1

g
X

mj+g k+g (j ) = mk Z

j=1

and

k ()
= k ()
| {z }
0

g
X
j=1

mj+g k (j ) = mk+g Z.
| {z }

kj


By Lemma 28.2.2, exp 2 1
now gives a meromorphic function
with (f ) = D .

29.3. Proof of Jacobi Inversion


To show that

AJ

is surjective, we will study the image of a certain

class of (degree zero) divisor on

M,

namely those of the form

[p1 ] + + [pd ] d[q]


given some xed point

qM

and natural number

d.

obviously in 1-to-1 correspondence with unordered

Such divisors are

d-tuples

of points

318

on

29. ABEL'S THEOREM, PART II

M,

in other words with elements of the

dth

symmetric power

d copies

z
}|
{
M

M
Symd M :=
.
(p1 , . . . , pd ) (p(1) , . . . , p(d) )
Sd
In order to be able to use complex analytic techniques we need to put
the structure of a

d-dimensional

complex manifold on this.

d = 2 and consider C instead


2
of a (compact) Riemann surface. The symmetric square Sym C is the
quotient of C C by the involution (z1 , z2 ) 7 (z2 , z1 ). What causes
To get the idea of how this works, take

diculty is the locus consisting of its xed points, i.e.


line.

Take two small open sets in

Sym M ,

the diagonal

one which intersects the

diagonal and one which does not:

z2

U
U

z1
Clearly

(z1 , z2 )

give local holomorphic coordinates on

U .

On

U ,

they are not well-dened, but their elementary symmetric polynomi-

2 (z1 , z2 ) = z1 z2 are. Moreover, these


functions generate all polynomials in z1 , z2 which are invariant un2
der the involution and hence well-dened on U Sym C. Taking
(w1 , w2 ) := (z1 + z2 , z1 z2 ) as the holomorphic coordinates there,9 the
transition function is then just (1 , 2 ). To see that this is invert2
ible on U , notice that in U the diagonal is dened by w1 = 4w2
2
(since z1 = z2 (z1 + z2 ) = 4z1 z2 ). Since U avoids this locus
p
w12 4w2 is well dened there and we can
(and is simply connected),

als

1 (z1 , z2 ) = z1 + z2

and

7elements are written either p + + p or {p , . . . , p }


1
d
1
d
8Had we started with M itself of dimension > 1, its symmetric

powers would be

singular complex analytic spaces, hence not manifolds. So what happens next is
special for

dim(M ) = 1.

9we could in fact take these as global coordinates, but this situation won't generalize
to

29.3. PROOF OF JACOBI INVERSION

dene

by

z =
1
z =
2

319

w12 4w2
2
w2 w12 4w2
2
w1 +

More generally, in a neighborhood of

{q1 , . . . , q1 ; . . . ; q` , . . . , q` } Symd M
| {z }
| {z }
k1 times

(where

P`

j=1

kj = d),

k` times

the local coordinate system is given in terms of

holomorphic coordinates

zj

on

qj ,

near each

by

{p1 , . . . , pk1 ; . . . ; pdk` +1 , . . . , pd } 7


| {z }
{z
}
|
all near q1

all near q`

(1 (z1 (p1 ), . . . , z1 (pk1 )) , . . . , k1 (z1 (p1 ), . . . , z1 (pk1 )) ; . . . ;


1 (z` (pdk` +1 ), . . . , z` (pd )) , . . . , k` (z` (pdk` +1 ), . . . , z` (pd ))) .
Inelegant, but it gets the job done.
Now let

be any divisor of degree

on

M,

and consider the map-

ping

D : P(L(D)) Symd M
which sends (for

f L(D))
[f ] 7 (f ) + D.

(f ) + D 0 by denition, and deg((f ) + D) = deg D = d; so


(f ) + D is of the form [p1 ] + + [pd ]. The map sends to the projective
equivalence class [f ], i.e.  f up to a constant multiple, to {p1 , . . . , pd }.)
(Here

Lemma 29.3.1.

is (a) injective and (b) holomorphic.

Definition 29.3.2. The linear system

10

|D| consists of all eective

M rationally equivalent to D. The Lemma evidently realizes


|D| = image(D ) as a subvariety of Symd M isomorphic to P`(D)1 .
divisors on

Proof. (of Lemma)

(f ) + D = (g) + D = (f ) = (g) = (f /g) = 0 = f /g


constant = [f ] = [g].
(b) To show D holomorphic in a neighborhood of [f0 ], augment f0
P`(D)
to a basis {f0 , f1 , . . . , f`(D) } L(D) and write f := f0 +
j=1 j fj
(a)

10the

notation

support of

D,

|D|

is unfortunately standard for both the linear system and the

two completely dierent concepts!

320

29. ABEL'S THEOREM, PART II

`(D)

{j }j=1 are the local holomorphic coordinates (on some small


U L(D)). Let p |D| |(f0 )|, with open neighborhood Np M and
local coordinate z (i.e. ordp (z) = 1). Set k := ordp (f0 ) + ordp (D), and
Wf0 ,p := Symk Np with coordinates 1 (z1 , . . . , zk ), . . . , k (z1 , . . . , zk ).
so that

We must show that the composition

U O(Np ) Wf0 ,p , Ck

7 f z ordp D 7

f z ordp D

1 z(p1 ()), . . . , z(pk ())

..
.

Np

z(p1 ()), . . . , z(pk ())

p1 () + + pk ()
is holomorphic, which in turn boils down to the statement that each

is holomorphic in each

j .

implicit function theorem; for

For

k = 1,

k > 1,

this is the holomorphic

it is this together with Rouch

and the Riemann extension theorem in a manner familiar from previous

chapters.

d divisor D (viewed as an
D = [p1 ] + + [pd ] with

Definition 29.3.3. An eective degree

element of
the

{pj }

Sym M )

is called general

distinct points of

M.

Now look at the Abel-Jacobi mapping

ud : Symd M J(C)
d
X

[p1 ] + + [pd ] 7 AJ

!
[pj ] d[q] ,

j=1
where

q M

is xed.

This is shown to be holomorphic by using

the fundamental theorem of calculus at general


Osgood and Riemann extension theorems.
taking a local lifting of the image of
The next result does not require

Lemma 29.3.4. The ber of

ud

u
D

to

D,

then applying the

(Boundedness is clear by

.)

to be general.

over

ud (D)

is

|D| (
= P`(D)1 ).

u for ud .)
Abel
u1 (u(D)) |D|: u(E) = u(D) = AJ(E D) = 0 = E D

is the divisor of some f K(M )


= (f ) + D = E 0 (since
d
E Sym M ) = f L(D) = E = D (f ) image(D ) = |D|.
Proof. (For simplicity write

29.3. PROOF OF JACOBI INVERSION

u1 (u(D)) |D|:

Given

E |D|,

there exists

321

f L(D)

such that

rat

E = (f ) + D = E D = (f ) 0 = 0 = AJ(E D) =
u(E) = u(D) = E u1 (u(D)).

D = [p1 ]+ +[pd ] is general, then writing zj for local coordinates
about each pj ,

(dud )D : TD Symd M Tu(D) (J(M ))
If

is computed by the matrix

z1

Pd zi

i=1
.
.
.
P
d

i=1
zd

zi
q

z1

1
..

zd

f1 (p1 )
.
.
.

..

f1 (pd )
where

g1

d=1

.
{p1 ,...,pd }

loc

is the canonical map and

. (For




i=1 q


.

.
.

Pd zi


i=1 q g
g

pj ) i = fi (zj )dzj , this


fg (p1 )
^
K (p1 )


.
.
,
.
.
=
.
.

^
fg (pd )
K (pd )

If we write locally (about each

Pd zi

g
^
K (pj ) C

is a lift of

K (pj )

this is just Proposition 28.1.2(b).) From this we see

that
rank


(dud )D = dim (span(K (p1 ), . . . , K (pd ))) + 1,

where span means the projective linear span in

Pg1 .

Taking

d = g,

we now have the following claim:

((dug )D ) = g for D = [p1 ]+ +[pg ] Symg M


d
Zariski open subset of Sym M .

Lemma 29.3.5. rank

generic,

11 i.e. in some

Proof. Choose

all of

g1

p1 , . . . , pg distinct with span (K (p1 ), . . . , K (pg )) =

. This is possible since the canonical map is always nonde-

g
generate by Theorem 27.3.2(a). Consequently rank ((du )D )
this holds more generally for

cause its failure is equivalent to

= g,

in an algebraic open set. This is be-

det(dug ) = 0,

which is an algebraic

condition which will hold on some codimension-one subvariety.

11general

may not be quite enough 

subvarieties of

Symg M

and

may have to avoid a larger number of

then just the ones where two or more

pj 's

coincide.

322

29. ABEL'S THEOREM, PART II

Theorem 29.3.6. [Jacobi inversion]

ug

is surjective and gener-

ically injective.

dug

Proof. By Lemma 29.3.5,

tangent spaces. So

Symd M

is generically an isomorphism of

takes an open ball about a general point

Symg M compact,
g
so image(u ) is both a closed analytic subvariety of J(M ) and contains
an open ball, and is therefore all of J(M ) (which is connected).
g
Since at a generic D , du is (in particular) injective, we see that
g 1
g
any such D is an isolated point of (u ) {u (D)}. But the latter is a
projective space by Lemma 29.3.4, and so the only way D is isolated
g 1
g
0
is if (u ) {u (D)} is isomorphic to P , i.e. is just D itself.

to an open ball. But

ug

is continuous and

Finally, to address (28.0.2) head-on, surjectivity of

AJ

follows from

the diagram

AJ

/ J(M )
OO
fMMM
MMM
ug
MM
D 7Dg[q] MM3 S

Div 0 (M )

(29.3.1)

Symg M

29.4. A nal remark on moduli


For any Riemann surface
tic basis of
for

(M )

H1 (M, Z),

(of genus

1)

with given symplec-

we know that there is a unique choice of basis

making the period matrix

over, we know by (29.1.6) that

imaginary part, i.e. belongs to the

of the form

Ig Z

. More-

is symmetric with positive denite

g th

Siegel upper half space

Hg := {Z Mg (C) | Z = t Z, Im(Z) > 0}.


Note that

H1

The Jacobian

H, the familiar upper half plane.


J(M ) is, of course, the quotient of Cg

is just

by the lattice

. More
generally, let Z be any g g complex matrix such that
Ig Z has
R-linearly independent column vectors. Writing Z for their Z-span,
g
we dene a complex torus by AZ := C /Z ; any complex g -torus is
given by integral linear combinations of the columns of

isomorphic to one of this form. A major result is the

29.4. A FINAL REMARK ON MODULI

323

Theorem 29.4.1. [Riemann Embedding Theorem]

AZ

is an

abelian variety (i.e., has a holomorphic embedding in projective space)


if and only if

belongs to

(Of course, any

Hg .

R-linearly

independent from

is in the upper

or lower half plane, so every complex 1-torus is algebraic; already for

g =2

this is false.) You can nd (eectively) two proofs in Griths

and Harris, one using generalized theta functions and the other using
Kodaira's embedding theorem.
For us, the implications of this theorem are:
(a)

Jacobians of Riemann surfaces of genus


eties of dimension

(b)

g;

Corollary 29.4.2. For

for

g 4,
For

are abelian vari-

and

abelian varieties of dimension

Since Riemann surfaces of genus

have

g(g+1)
moduli.
2

g 2 have 3g3 moduli, we conclude:

g < 4,

all abelian

g -folds

are Jacobians;

most of them are not.

g 4,

then, we have the problem of characterizing the Jaco-

bian locus in the moduli space

Hg /Sp2g (Z),

which is the (very di-

cult) Schottky problem. There are recent results describing this locus
in terms of the vanishing of theta functions.

12

12for example, see the seminar talk http://www.math.princeton.edu/~sam/papers/talk3s.pdf

Appendix: genera of singular curves

Start with an irreducible projective algebraic curve

C P2
of degree

d,

dened over

C.

We know how to piece the local normal-

izations about singular points together with the smooth part of


construct a Riemann surface

C ,

to

together with a map

: C P2
with image()

= C.

The genus formula of Chapter 14, derived from

a generic stereographic projection and the Riemann-Hurwitz formula,


said that

=
g(C)
if all singularities of
are

(d 1)(d 2)

(if any) are ordinary double points and there

such points.

Here we would like to be able to compute the genus of the normalization of an arbitrary irreducible curve, with singularities of any
order and type. There exist formulas when the singularities are ordinary,

13 but my preference is for methods over formulas, particularly

14
There are
when the methods allow you to treat more general cases.
two methods: the rst one computes the divisor of the pullback of a

P2 to C and applies Poincar-Hopf;


C to a line and applying Riemann-

meromorphic dierential 1-form on


the second is based on projecting

Hurwitz (like in the proof of the genus formula). Rather than trying to
state them formally, I'll use both methods to treat an example which
is suciently general that you'll be able to adapt the approaches to
any other curve.

13cf. 6.4

for the denition of an ordinary

k -tuple

the formula is Fulton's book on algebraic curves.

point. One book which proves

14and particularly when you can't remember formulas (welcome to my world)


325

326

APPENDIX: GENERA OF SINGULAR CURVES

So here is the ugly curve we will study: put

F (Z, X, Y ) := X 3 Z 3 + X 6 + Y 5 Z,
and

C := {F (Z, X, Y ) = 0} P2 ,
with ane form

x3 + x6 + y 5 = 0.

This is a degree

a smooth curve of this degree has genus

10.

6 (i.e.

sextic) curve;

That will not be the answer

here.
One immediately obvious singularity is at
projective coordinates

[Z : X : Y ]).

(0, 0)

(i.e.

[1 : 0 : 0]

in

The lowest-order homogeneous

term (of the ane equation, in coordinates vanishing at this point) is

x3 .

[1 : 0 : 0] is a triple point, but very denitely not


point of C . Ugly enough? Well, this turns out to

So

triple

an ordinary
be the only

singularity.

Method I: Poincar-Hopf
Set

 
dx
y


K1 (C)

(it will actually turn out to be in

although this is inessential for the method).

,
1 (C)

Poincar-Hopf tells us

. So we have to compute
deg(()) = 2g 2, where g = g(C)
P
Where might these {pi } lie in C ? Or rather,
() = mi [pi ] Div(C).
where might the {(pi )} lie on C ? There are four (not necessarily
that

disjoint) possibilities:
(1) on the intersections of
(2) at points where

(3) at singularities of

with the

x-axis,

i.e. in

has a vertical tangent, hence

C,

C {Y = 0};
in C {FY = 0};

i.e. in Sing(C); and

(4) on the line at innity, i.e. in

C {Z = 0}.

p () at p in one of these sets? For


dx
is zero on the line Y = 0; while for (2)
(1), the denominator of
y

Why might one expect nontrivial

the pullback of

dx

will be zero, since at such a point the curve has no

horizontal variation to rst order. You should always be suspicious


of (3) and (4). Conversely: on the smooth ane part of

C , dx

and

never blow up, and (1) and (2) are the only ways they can develop a
zero. So (1)-(4) are actually the only places where

can have a zero

or pole.
Now we go through these 4 sets of points for the particular curve
under consideration.

METHOD I: POINCAR-HOPF

327

(1): we look at the ane equaton and set

y = 0, which yields

1
(Here, 6 = exp(
).)
3

x3 + x6 = 0, hence x = 0, 6 , 6 , or 1.
While (0, 0) is a singular point and will be dealt with below, it is clear
dx
| will behave in the same way near the remaining three points:
that
y C
(1, 0), (6 , 0), and (6 , 0). We look in a neighborhood of (1, 0) on
C . Setting = x + 1 (or x = 1), the equation becomes in (, y):
0 = y 5 + ( 1)3 + ( 1)6 = y 5 3 + {higher-order

terms in

= y 5 3h(),
h(0) 6= 0.15 The local normalization
of C at (, y) = (0, 0) is

p
5
5
5
therefore t 7
t , t 3h(t ) , under which dx
= d
pulls back to
y
y

where

5)

d(t

3h(t5 )

5dt
= t3
5

3h(t5 )

which has a zero of order

at

t = 0.

So we

conclude that

1 [1:1:0] () = 3,
and similarly that

1 [1:6 :0] () = 1 [1:6 :0] () = 3.


(2): for vertical tangents or singularities we will have

0 = FY = 5Y 4 Z,
x-axis or along the line at . The
intersections with the x-axis other than [1 : 0 : 0] were just dealt with.
Any nonsingular intersections with {Z = 0} will be dealt with in step

so that these must occur along the

(4). So vertical tangents are subsumed under the other three categories.

(3): at a singular point we must have

0 = FY ,

0 = FX = 3X 2 Z 3 + 6X 5 = 3X 2 (Z 3 + 2X 3 ),
and

0 = FZ = 3X 3 Z 2 + Y 5 .
Z = 0 or Y = 0. If Z = 0 then the last two equations
imply X = Y = 0, a contradiction. If Y = 0 then the last equation
gives Z = 0 (no!) or X = 0; the latter works, and so [1 : 0 : 0] is
the only singular point. In local coordinates about (x, y) = (0, 0), our
5
3
6
5
3
curve is 0 = y + x + x = y + x h(x) (dierent h(x) from above,

We must have

15Sometimes

(though rarely) one may have to remember more about

types of problems, but not in this example.

in these

328

APPENDIX: GENERA OF SINGULAR CURVES

again

h(0) 6= 0),

which is locally irreducible and has a singularity of

order 3. Undeer the local normalization


back to

d(t5 )

t3 5

h(t5 )

5dt
=t
5
5

h(t )

t 7 (t5 , t3

p
5

h(t5 )),

dx
pulls
y

, and we conclude that

1 [1:0:0] () = 1.
(4):

C {Z = 0}

is the single point

[0 : 0 : 1].

We will need to

switch to ane coordinates vanishing at this point, namely

v=

x
y

X
(or conversely
Y

y=

1
,
u

x=

u=

1
y

Z
,
Y

v
):
u

y
x
We divide

Z 3 X 3 + X 6 + Y 5 Z = 0 by Y 6 , obtaining
 3  3  6
X
X
Z
Z
+
+ =0
Y
Y
Y
Y
v 6 + u3 v 3 + u = 0

which is a locally irreducible Weierstrass polynomial in

with (multi-

valued) roots of the form

s
v (u) :=

u3

(Use the quadratic formula to solve for


stituting in

t6

u6 4u

2
v3,

then take cube root.) Sub-

gives

36
6
18 + t3
+ t 4t
t30 4
3 t
3
6
=
v(t ) =
2
2
s

15 +
t30 4
3 t
=t
2
which is just t times some local analytic H(t) with H(0) 6= 0. So
v
d( u
)
dx
6
the normalization is t 7 (t , t H(t)) and
pulls back to
=
1
y
u


tH(t)
d( 6 )
t
= t6 d H(t)
= (tH 0 (t) 5H(t)) dt, which has neither zero nor
1
t5
s

t6

t18

METHOD II: RIEMANN-HURWITZ

pole at

t = 0.

329

Hence,

1 [0:0:1] () = 0.
Upshot: putting everything together,

() = [ 1 [1 : 0 : 0]]+3[ 1 [1 : 6 : 0]]+3[ 1 [1 : 1 : 0]]+3[ 1 [1 : 6 : 0]]


2g 2 = deg(()) = 1 + 3 + 3 + 3 = 10

g = 6.

Method II: Riemann-Hurwitz


Recall that this dealt with maps of Riemann surfaces

f : M N,
and told us that

M = deg(f ) N rf .
Here

deg(f )

is the mapping degree of

(the number of points in the

N ) and rf is the degree of the ramiRf := pM (p (f ) 1)[p].


2
, N = P1 , =stereographic projection
Now let q P \C , M = C
(P2 \{q}) P1 through q ; and take

preimage of a general point on

16
cation divisor

f : C P1
f := . Usually it is
[0 : 0 : 1] as q . In our case the

[1 : 0 : 0],
these not on C

to be given by

easiest to take

[0 : 1 : 0], or
is [0 : 1 : 0]. So

only one of

16locally

our projection looks like

about any

coodinates

z 7 z p (f )

z resp. w
= w.

p M and its image f (p) N , one has local holomorphic


(with z(p) = 0 resp. w(f (p)) = 0), in which f takes the form

330

APPENDIX: GENERA OF SINGULAR CURVES

P
C

[0:1:0]

(forgets xcoord.)
and the mapping degree is the number of intersection points of
and

P1

{x + x + y = 0} for general y0
= 2 2 0 = 2, so we have

 i.e.

deg(f ) = 6.

{y = y0 }

Obviously

2 2g = C = 6 2 rf = 12 rf
1
g = rf 5.
P2
So we will have to compute Rf =
mi [pi ] (or at least rf ) and the rst
issue to resolve is where the (pi ) can lie on C :
=

{F = 0} {FX = 0});
i.e. [1 : 0 : 0] for our

(1) points having horizontal tangents (subset of


(2) singular points

(FX = FY = FZ = 0)

example; and
(3)

L C
(1):

 i.e.

[0 : 0 : 1]

in our example.

0 = FX = 3X 2 (Z 3 + 2X 3 )

X = 0,
we
x= X
Z

has solutions other than

which corresponds to the singular point. Namely, writing

6
6
1

3 ,
3 ,
3 . Plugging this into the ane
2
2
2
52
53
54
5
1
1
5
equation of C yields y =
hence y =
5 ,
5 ,
5 ,
5 ,
5 . These
4
4
4
4
4
4
are independent of the choice amongst the 3 values for x, and so we
get

x3 +

1
2

=0

hence

x =

5 3 = 15 ramication points. As you may check, the intersections


between F = 0 and FX = 0 at these points are all of rst order, hence
get

correspond to ramications of order 2 and so make a contribution of

p (f ) 1 = 2 1 = 1 each to rf .
(2): Near (x, y) = (0, 0), the composition

t 7 (t5 , t3 H(t)) 7 t3 H(t)

EXERCISES

has

p (f ) = 3
(3): Near

hence contributes

to

331

rf .

(u, v) = (0, 0),

t 7 (t6 , tH(t)) 7 t6 .
| {z }
u,v

This is because

is supposed to take the

y -coordinate,

which is

1
u

here; but we have to compute the image in a holomorphic coordinate

p = [0 : 0 : 1]. So in fact u is the correct


map indeed has p (f ) = 6 and contributes 5 to rf .
rf = 15 1 + 2 + 5 = 22

vanishing at the image of


variable, and the

Conclusion:

g =

22
5 = 6,
2

conrming the previous computation.

Exercises
(1) Find the genus of the normalization of the curve
ane equation

2 3

x +y +y x

2 5
y
5

= 0.

C P2

with

Do this in 2 dierent

ways: (a) by using an appropriate projection, computing the degree


of the ramication divisor, and applying Riemann-Hurwitz; (b)
by computing the divisor of the pull-back of a meromorphic 1form on the normalization and applying Poincare-Hopf (try it with

dy
). [Hint: to do the local normalizations, rst make sure you are
x

dealing with an irreducible Weierstass polynomial

f (x, y) = 0

you may have to change variable, swap coordinates, divide out by


a unit (which can reduce the degree of the equation!), factor into
irreducibles, whatever. If you can't nd a multivalued solution

y(x)

by taking roots, using quadratic equation, and so on, you can always

f (x, y) is an irreducible Weierstrass polynomial


k
k
of degree k in y , then try substituting in t for x: write 0 = f (t , y)
and solve for y as a power-series in t, call this G(t). Then the local
k
normalization is t 7 (t , G(t)).]
use power series. If

You might also like