Ims All
Ims All
Ims All
– An Introduction –
written by
BERND KREUSSLER
Again, many more things could have been included here. For
example, higher dimensional Abelian Varieties and the Abel-Jacobi
map which naturally emerge in the study of curves of higher genus are
not mentioned. The idea of a scheme over an arbitrary commutative
ring with unity are definitely beyond the scope of this article. To
introduce the ideas of a moduli space and of a universal object would
be a natural next step after the introduction of the Poincaré bundle.
A higher dimensional analogue of an elliptic curve would be a so-
called K3-surface. Their study has much in common with the theory
of elliptic curves but they couldn’t be touched either. There are
many excellent textbooks available, among which are [3, 5, 16].
References
[1] N.I. Akhiezer, Elements of the theory of elliptic functions. Translations of
Mathematical Monographs 79, AMS (1990)
[2] T. Bridgeland, Derived categories of coherent sheaves. Proceedings of the
international congress of mathematicians (ICM), Madrid, Spain, August
22–30, 2006, Volume II, 563–582 (2006)
[3] P. Griffiths, J. Harris, Principles of algebraic geometry. John Wiley & Sons
(1978)
[4] D. Huybrechts, M. Lehn, The geometry of moduli spaces of sheaves. Aspects
of Mathematics E 31, Vieweg (1997)
[5] F. Kirwan, Complex algebraic curves. London Mathematical Society Stu-
dent Texts 23, Cambridge University Press (1992)
[6] A.W. Knapp, Elliptic curves. Mathematical Notes (Princeton) 40, Princeton
University Press (1992)
[7] N. Koblitz, Algebraic Aspects of Cryptography. Springer (2004)
ELLIPTIC CURVES – AN INTRODUCTION 5
BERND KREUSSLER
1. Pythagoras
The aim of this introductory section is to recall the well-known
geometric construction of all Pythagorean triples of integers. Three
integers a, b, c ∈ Z form a Pythagorean triple, if
a 2 + b 2 = c2 .
Almost everybody knows the Pythagorean triple (3, 4, 5) and many
know (5, 12, 13). However, not everybody has come across (8, 15, 17)
or (20, 21, 29).
Clearly, if n ∈ Z and (a, b, c) is such a triple, (na, nb, nc) will also
be one. In this way, starting with the well known triple (3, 4, 5) we
obtain (6, 8, 10), (−3, −4, −5), (12, 16, 20) etc.
Note that a prime number which divides two of the three inte-
gers in a Pythagorean triple automatically divides the third in the
triple. Therefore, it is enough to find all Pythagorean triples in
which any two of the three integers are co-prime. We shall call such
a Pythagorean triple reduced. Because the only Pythagorean triple
with c = 0 is (a, b, c) = (0, 0, 0), we shall assume in the sequel c 6= 0.
This allows us to introduce the new variables
a b
x= and y = .
c c
Using these coordinates, the search for reduced Pythagorean triples
translates into the problem to find all rational solutions of the equa-
tion
x2 + y 2 = 1.
7
8 BERND KREUSSLER
In other words, we would like to find all points on the unit circle
whose coordinates are rational.
The key observation is that a line which connects two points with
rational coordinates always has a rational slope. Therefore, we shall
look at all lines in the plane which pass through the point (0, −1)
and which have rational slope r ∈ Q.
y = 3x − 1
x2 + y 2 = 1
bc 3 4
5, 5
bc
(0, −1)
2r r2 − 1
x= 2 and y = 2 .
r +1 r +1
2r r 2 −1
The map which sends r ∈ Q to the point r2 +1 , r2 +1 on the unit
circle gives a parametrisation of the set of all rational points on this
curve. This completely solves our problem.
If we wish to derive a complete description of all Pythagorean
triples of integers, we start by writing the slope r as r = uv with co-
prime integers u, v. Using symmetry, we may assume r > 1. More
precisely, switching from r to −r corresponds to a sign change of
x, whereas a sign change of y is achieved by switching from r to
1
r . Thus, we assume u > v > 0 and u, v co-prime. Under these
SOLVING CUBIC EQUATIONS IN TWO VARIABLES 9
u
assumptions, r = v produces the point with coordinates
2r 2uv r2 − 1 u2 − v 2
x= = and y = = .
r2 + 1 u2 + v 2 r2 + 1 u2 + v 2
Now it is not hard to see that each reduced Pythagorean triples in
which a is odd can be written as
u2 − v 2 u2 + v 2
(a, b, c) = uv, ,
2 2
with u > v > 0, both odd and co-prime. Up to interchanging a and
b this gives us all reduced Pythagorean triples, because a and b are
co-prime, hence at least one of these to integers is odd. For small
values of u, v we obtain the following table
u v a b c u v a b c
3 1 3 4 5 7 5 35 12 37
5 1 5 12 13 9 1 9 40 41
5 3 15 8 17 9 3 27 36 45
7 1 7 24 25 9 5 45 28 53
7 3 21 20 29 9 7 63 16 65
2. A cubic example
The aim of this section is to find integer solutions of cubic equa-
tions by using the geometric idea used in the previous section. We
shall explain this method through the following example
b2 c = 4a3 − 4ac2 + c3 .
a
As before, we assume c 6= 0 and introduce new coordinates x = c
and y = cb in which the above equation becomes
y 2 = 4x3 − 4x + 1. (1)
This can be rewritten as
(y − 1)(y + 1) = 4(x + 1)x(x − 1).
In this form it is obvious that we have the following six solutions
(−1, ±1), (0, ±1), (1, ±1).
Question: Are these all the solutions of equation (1)?
It is not hard to produce a sketch of this curve in the real plane.
This can be done through the following step-by-step approach. First,
we draw the graph of the cubic polynomial 4x3 − 4x + 1. The inter-
section points with the x-axis can be found with Cardano’s formula.
10 BERND KREUSSLER
This polynomial has three real roots because its discriminant is posi-
tive. To get the second picture, we remove all points from the graph
which have negative y-coordinate. The next picture is produced
by applying the square root function. Finally, the cubic curve is ob-
tained by adding in the mirror image along the x-axis, because (x, y)
is on this curve if and only if (x, −y) is so.
bc bc bc
bc bc bc
The six marked points in the picture are the points we found before.
If we seek to find more rational points on this curve, we may try to
use lines with rational slope which pass through one of the known
points. This leads to a quadratic equation the solutions of which
correspond to two further intersection points of this line with the
curve given by equation (1).
bc
For example, the line with slope 1 which passes through the point
(0, 1) has the equation y = x+1. The x-coordinates of its intersection
with our curve are the solutions of the equation (x+1)2 = 4x3 −4x+1
or equivalently 4x3 − x2 − 6x = 0. The known solution corresponds
SOLVING CUBIC EQUATIONS IN TWO VARIABLES 11
bc
bc
Pbc bcR
Q bc bc bc
y = 2x + 1
bc
S
bc
bc P
Q bc
bc
P +Q
This example shows that we actually need to know only one ra-
tional point on our cubic in order to get started. As before, we can
then produce many other points. Using the notation suggested by
Definition 5, we obtain here:
b−4P
7P
bc
b bc
P bc2P
−3P
b
5P bc
3P bc b b−2P
b
−7P
bc4P
We shall see in the next section that this is in fact the structure
of an Abelian group in which for each point T , −T is the mirror
image of T with respect to the x-axis. The line which connects an
arbitrary point T on our cubic with its mirror image −T is a vertical
line. Because T + (−T ) = 0, we expect all these lines to go through
the neutral element of this group. Therefore, we shall look for the
neutral element “at infinity”. This can be made more precise with
the aid of the projective plane P2 , introduced in the following section.
formulate the next definition for any field K. The reader who is not
familiar with the concept of a field may substitute Q or C for K.
Definition 7. The projective plane P2 (K) over the field K is the set
of all equivalence classes (z0 : z1 : z2 ) of non-zero vectors (z0 , z1 , z2 ) ∈
K3 . Two such vectors (z0 , z1 , z2 ) and (w0 , w1 , w2 ) are equivalent if
and only if there exits a non-zero λ ∈ K such that (z0 , z1 , z2 ) =
λ(w0 , w1 , w2 ). This implies
(z0 : z1 : z2 ) = (λz0 : λz1 : λz2 ) for all λ 6= 0.
The notation (z0 : z1 : z2 ) for the equivalence class of the vector
(z0 , z1 , z2 ) is chosen in order to suggest that we are dealing with the
ratios between the three numbers z0 , z1 and z2 only. A similar con-
struction, of course, can be carried out in any dimension to produce
Pn (K) for all n ≥ 1. The one-dimensional case is particularly easy.
If K = C it leads to the Riemannian sphere. Notations used for the
Riemannian sphere are S 2 = C ∪ ∞ = C and P1 (C), the notation
we are going to use here. Its points are equivalence classes (z0 : z1 )
of non-zero vectors (z0 , z1 ) ∈ C2 . All points in P1 (C) with z0 = 0
are equivalent to ∞ = (0 : 1). Any point with z0 6= 0 is equiva-
lent to (1 : z) where z = zz10 . This gives a bijection between C and
P1 (C) \ {∞}. A neighbourhood of ∞ would be the set of all those
points of P1 (C) which have z1 6= 0. This is again in bijection with
C by using w = zz10 . The relationship between these two patches of
P1 (C) is given by w = z1 . This actually makes P1 (C) into a complex
manifold of dimension one, the simplest compact Riemann surface.
The local structure of P2 (K) can be studied in a similar way.
To this end, we define the three basic open sets which cover P2 (K)
completely
U0 := {(z0 : z1 : z2 ) | z0 6= 0} ⊂ P2 (K)
U1 := {(z0 : z1 : z2 ) | z1 6= 0} ⊂ P2 (K)
U2 := {(z0 : z1 : z2 ) | z2 6= 0} ⊂ P2 (K).
2
Each of these sets is in bijection with
K . For example, the map
U0 → K2 given by (z0 : z1 : z2 ) 7→ zz10 , zz02 has as its inverse the
map K2 → U0 which sends (ξ1 , ξ2 ) to (1 : ξ1 : ξ2 ).
z
Similarly, on U1 we can work with affine coordinates ηj = zj1 ,
j = 0, 2 and on U2 we have ζk = zzk2 , k = 0, 1. The gluing maps
16 BERND KREUSSLER
l0 z 0 + l1 z 1 + l2 z 2 = 0
with l0 , l1 , l2 ∈ K but not all three equal to zero. Because λl0 , λl1 , λl2
define the same line in P2 (K) if λ 6= 0, the set of all lines in P2 (K)
is another P2 (K), called the dual projective plane and sometimes
denoted P2 (K)∨ .
Each line in P2 (K) is isomorphic to P1 (K). The three lines at
infinity introduced before are also lines in this sense, because Lj was
given by the equation zj = 0. In particular, the line L2 corresponds
to the point (0 : 0 : 1) ∈ P2 (K)∨ . Any other line, with coefficients
(0 : 0 : 1) 6= (l0 : l1 : l2 ) ∈ P2 (K)∨ intersects U2 in an ordinary line.
The equation of this intersection is
l0 x + l1 y = −l2
Remark 9. This result is true for any field K and any cubic equation
of the form
z12 z2 = z03 + pz0 z22 + qz23 (2)
3 2
with p, q ∈ K satisfying ∆ = −16(4p + 27q ) 6= 0. If the characteris-
tic of K is not equal to two or three (i.e. if 1 + 1 6= 0 and 1 + 1 + 1 6= 0
in K), every regular cubic with a point over K can be given by such
an equation. When working in characteristic zero (e.g. K = Q or
K = C), we can change coordinates so that (2) becomes
z12 z2 = 4z03 − g2 z0 z22 − g3 z23 . (3)
The discriminant of such an equation is ∆ = g23 − 27g32 . A cubic
equation of the form (3) is called a Weierstraß equation, named after
Karl Weierstraß (1815–1897). The coefficient 4 at z03 is used because
it appears in the differential equation of the Weierstraß ℘-function.
(See the article by M. Franz [5].)
The most basic structure result about the group E(Q) was shown
in 1922 by Mordell [17].
Theorem 10 (Mordell). If E is given by (2) with p, q ∈ Q and
4p3 + 27q 2 6= 0 then the Abelian group E(Q) is finitely generated.
Remark 11. Theorem 10 has been generalised by A. Weil to Abelian
varieties of arbitrary dimension over any number field [23]. Therefore
Mordell’s Theorem is also known as the Mordell-Weil Theorem and
the group E(Q) is sometimes called the Mordell-Weil group.
The curve studied in section 2 has E(Q) ∼ = Z with generator
P = (0, 1). The discriminant of this curve is ∆ = 37.
Remark 12. The assumption 4p3 + 27q 2 6= 0 in Mordell’s Theorem
is crucial. If 4p3 + 27q 2 = 0 the cubic polynomial x3 + px + q has
a multiple root and this gives rise to a singular point on the cubic
curve given by (2). This changes the situation completely, because
singular cubics are rational. More explicitly, suppose −4p3 = 27q 2
and p, q ∈ Q \ {0}. A straightforward calculation shows that, under
these assumptions,
2
3 3q 3q
x + px + q = x − x+ .
p 2p
3q
This implies that − 2p , 0 is a singular point of the cubic which
means that each line in P2 (Q) that passes through this point will
SOLVING CUBIC EQUATIONS IN TWO VARIABLES 19
have at most one other intersection point with the cubic curve (2).
Just as in the case of the circle in section 1 this produces a bijection
between Q (the set of slopes) and all rational points on a singular
cubic apart from the singular point. This shows that the non-singular
rational points on a singular cubic form a group which is not finitely
generated.
bc
Example 13. Look at the singular cubic z12 z2 = 4z03 − 3z0 z22 +
z23 , which has discriminant ∆ = 33 − 27(−1)2 = 0. On U2 , using
coordinates x, y as before, its equation is
y 2 = 4x3 − 3x + 1.
From 4x3 − 3x + 1 = (x + 1)(2x − 1)2 we see that the singular point
has coordinates 12 , 0 . Any line with rational slope r through this
point will have equation y = r x − 12 , hence the new intersection
2
point will be found by solving r2 (2x − 1)2 = (x + 1)(2x − 1)2 .
Therefore, the coordinates of this point are given by
r2 − 4 r3 − 6r
x= and y= .
4 4
The main difference between this and the non-singular case is that we
cannot find such a closed formula for all rational solutions in the non-
singular case. This is explained by the involvement of transcendental
functions such as theta functions and the Weierstraß ℘-function (see
the article [5]).
20 BERND KREUSSLER
4. Further results
In this section we collect some general results known about the
Mordell-Weil group E(Q). We also discuss several normal forms of
plane cubic curves.
Let us look at a cubic curve given by an equation of the form
y 2 + a1 xy + a3 y = x3 + a2 x2 + a4 x + a6 .
24 BERND KREUSSLER
xy − wz = 0
y 2 − a2 w2 + x2 − a2 z 2 = 0.
SOLVING CUBIC EQUATIONS IN TWO VARIABLES 27
References
[1] M. Artebani, I. Dolgachev, The Hesse pencil of plane cubic curves,
arXiv:math.AG/0611590
[2] J.W.S. Cassels, Diophantine equations with special reference to elliptic
curves. J. Lond. Math. Soc. 41, 193–291 (1966); Corrigenda. Ibid. 42, 183
(1967)
[3] H.M. Edwards, A normal form for elliptic curves. Bull. Amer. Math. Soc.
44, 393–422 (2007)
[4] T. Ekedahl, One semester of elliptic curves. EMS Series of Lectures in Math-
ematics, European Mathematical Society Publishing House. (2006)
[5] M. Franz, Theta Functions, this issue.
[6] I. Garcı́a-Selfa, M. A. Olalla, J. M. Tornero, Computing the rational torsion
of an elliptic curve using Tate normal form. J. Number Theory 96, No. 1,
76–88 (2002)
[7] R.V. Gurjar et al., Elliptic curves. Praveshika Series. New Delhi: Narosa
Publishing House/dist. by the AMS (2006)
[8] R. Hartshorne, Algebraic geometry. Graduate Texts in Mathematics 52,
Springer (1977)
[9] O. Hesse, Über die Wendepuncte der Curven dritter Ordnung. J. Reine
Angew. Math. 28, 97–107 (1844)
28 BERND KREUSSLER
MADEEHA KHALID
1. Manifolds
Throughout this article we work over C, the field of complex num-
bers. We denote Pn (C) by Pn .
An n dimensional complex manifold M is a topological space
which locally looks like Cn . This means that there exists an open
cover Uα and co-ordinate maps φα : Uα → Cn such that φα φβ−1 :
φβ (Uα ∩ Uβ ) → Cn is holomorphic for all α, β. Similarly a function
f on an open set U ⊂ M is holomorphic if for all α, f ◦ φ−1 α is
n
holomorphic on φα (Uα ∩ U ) ⊂ C . A map f : M → N between two
complex manifolds is holomorphic if it is given in terms of local holo-
morphic co-ordinates on N by holomorphic functions. Open subsets,
products of complex manifolds and suitable quotients of complex
manifolds are also complex manifolds.
The simplest example of a one dimensional complex manifold is
just C itself. Then there is P1 (isomorphic to the Riemann sphere)
C/Λ
P
aijk ξ1 j ξ2 k . So C ∩ U0 is the zero locus of the holomorphic func-
tion F0 (ξ1 , ξ2 ). The calculations for the other charts U1 and U2 are
similar.
We say that p ∈ C ∩ U0 is a smooth point, if at least one of the
(ξ1 ,ξ2 ) ∂F0 (ξ1 ,ξ2 )
partial derivatives ∂F0∂ξ 1
, ∂ξ2 is not equal to zero at p. We
say C is smooth if every point in C is a smooth point. If C is smooth
then in fact it is a submanifold of P2 . Another example of a smooth
curve is the curve given locally by ξ1 n + ξ2 n = 1. In homogeneous
coordinates it is the zero locus of z1 n + z2 n = z0 n and is known as
the Fermat curve.
There is a nice formula that computes the genus of a smooth plane
curve C of degree d namely g = (d−1)(d−2) 2 . So if C has degree d
where d ≥ 3 then g ≥ 1 and hence C is not isomorphic to P1 . The
genus of the Fermat curve is (n−1)(n−2) 2 so for n = 1 and 2 it is
1
isomorphic to P while for n = 3 it has genus one and is a complex
torus.
The curve C ′ in P2 given by the equation z2 z1 2 − z0 3 + z0 2 z2 = 0
is not smooth as all the partial derivatives vanish at (0 : 0 : 1). We
say (0 : 0 : 1) is a singular point of C ′ .
These notions of smooth points and singular points can also be
extended to higher dimensional manifolds. In fact just as the im-
plicit and inverse function hold in the differentiable case, so do their
analytic versions. For example if V is a hypersurface given locally
as the zero set of a single holomorphic function f and the jacobian
matrix of f has rank 1 everywhere then V is a manifold of dimension
n − 1.
bc
R
bc P
Q bc
P + Q := R̄
bc
3. Complex torus
In this section we relate E to the one dimensional complex torus
given as the quotient C/Λ of C by a rank 2 lattice Λ in C. Since
34 MADEEHA KHALID
Zx1
R(x, y)dx
x0
of the point (x, y), sums of such integrals (known as Abelian sums)
XZ
xi
R(x, y)dx
x0
ZP
dx
.
y
O
4. Divisors
In the previous section we saw that a plane cubic curve E is
isomorphic to a complex torus C/Λ. Now C/Λ inherits a group
structure from C and hence induces a group structure on E via the
isomorphism. In Section 2 we defined a group operation on E using
geometry. How do these two compare?
The answer is: they coincide! In the subsequent sections we de-
scribe a proof which weaves together some pretty ideas from algebraic
geometry. To do so we have to first introduce an important notion
in algebraic geometry which is that of a divisor. In the case of a
curve it has a simple description.
Definition 3. Let C be a smooth curve in P2 . A divisor on C is a
formal finite linear combination D = a1 · P1 + · · · + am · Pm of points
Pi ∈ C with integer coefficients ai .
38 MADEEHA KHALID
Pm of a divisor D = a1 · P1 + · · · + am · Pm is
Definition 4. The degree
defined to be deg D = i=1 ai and this gives a group homomorphism
deg : Div(C) → Z.
Remark 5. The notion of a divisor extends also to higher dimen-
sional manifolds. In that case a divisor is a linear combination of
subsets given locally by zero sets of irreducible holomorphic func-
tions.
The group Div(C) is very large, even in the one-dimensional case.
Therefore we introduce the sub-group of principal divisors. The ben-
efit is that the factor group of all divisors modulo principal divisors
is finitely generated. This factor group will prove to be useful in Sec-
tion 6 as well. In order to explain the definition of principal divisors,
we need the notion of the order of a function at a point P
Let f be a holomorphic function on an open set U ⊂ C. Let
P ∈ U and let x be the local co-ordinate on U such that P is given
by x − λ for some λ ∈ C. The order of f at P , denoted ordP (f ), is
the largest integer a ∈ Z such that locally
f (x) = (x − λ)a · h(x)
where h is a holomorphic function with h(λ) 6= 0. Since f is holomor-
phic a is non negative. Note that for g, h any holomorphic functions
ordP (gh) = ordP (g) + ordP (h).
We would like to include the cases when ordP (f ) is negative. To
do so we have to include what are known as meromorphic functions.
A function f on C is called a meromorphic function if it can be writ-
ten locally as a ratio hg , where g 6= 0 and h are holomorphic functions
which do not have a common zero. Then, by using a Laurent series
expansion for f at P , we see that ordP (f ) = ordP (g) − ordP (h). So
ordP (f ) is negative if ordP (g) < ordP (h).
Collecting zeros and poles of a global meromorphic function f
gives us a natural way to associated a divisor to it.
Definition 6. Let f be a meromorphic function on C. Then the
divisor of f , called a principal divisor and denoted (f ), is given by
X
(f ) = ordP f · P.
P ∈C
GROUP LAW ON THE CUBIC CURVE 39
5. Line Bundles
Divisors are closely tied together to another geometric notion
which is that of a line bundle. A line bundle is a rank 1 holomorphic
vector bundle (Definition 10). In this section we discuss the relations
between line bundles and divisors.
Let us for the moment refer back to Example 8. The homoge-
neous coordinates z0 , z1 of P2 are also natural homogeneous coordi-
nates on L2 , since L2 = {(z0 : z1 : 0) ∈ P2 }. Our aim is to associate
40 MADEEHA KHALID
1
and is defined as follows. For w = z ∈ V0 ∩ V1 and (w, λ) ∈ V1 × C,
(z, µ) ∈ V0 × C we define
(w, λ) ∼ (z, µ) ⇐⇒ µ = z 2 λ.
Note that we define the “patching” condition using z 2 , the nowhere
vanishing function on V0 ∩ V1 relating the two local descriptions of
D above.
This new manifold L is an example of a line bundle, (see Def. 10
below) and is often denoted O(D). The collection {(f0 , V0 ), (f1 , V1 )}
of local defining functions for D = 2 · O defines a section (see subsec-
tion 5.2) of O(D) and the function z 2 relating these local functions
on the overlap V0 ∩ V1 is a transition function of O(D).
We now give the general definition of a line bundle.
π
Definition 10. Let M be a complex manifold. A line bundle L → M
is a holomorphic vector bundle of rank 1. That is
(1) L is a complex manifold such that for any x ∈ M, π −1 (x) =
Lx is equipped with the structure of a one dimensional com-
plex vector space.
(2) The projection mapping π : L → M is holomorphic.
(3) There is an open cover {Uα } of M and biholomorphic maps,
φα : π −1 (Uα ) → Uα × C, compatible with the projections
onto Uα , such that the restriction to the fibre φα : Lx →
{x} × C is linear for all x ∈ Uα . The pair (φα , Uα ) is called
a trivialisation of L over Uα .
Lx LUα ⊂ L
π
b Uα ⊂ M
x
42 MADEEHA KHALID
This gives f ∗ L its manifold structure over the open set f −1 (U ). The
transition functions for f ∗ L are the pull backs f ∗ (gαβ ) := gαβ ◦ f of
the transition functions gαβ of L.
Remark 15. If D is a divisor on N with local defining functions
{(hα , Uα )}, we can pull it back to a divisor f ∗ D on M with local
defining functions {(hα ◦ f, f −1 (Uα ))}. If L = O(D), then f ∗ (L) =
O(f ∗ D).
5.1. Group structure on the set of all line bundles. The ten-
sor product of C with itself, C ⊗ C is C again. Similarly given two
line bundles L1 and L2 with transition functions gαβ and hαβ respec-
tively, we can define the tensor product L1 ⊗ L2 and get a new line
bundle L. The fibres of L are just the tensor product of fibres of L1
and L2 . The transition functions tαβ of L are therefore the product
of the transition functions of L1 and L2 , i.e. for all x ∈ Uα ∩ Uβ
tαβ (x) = gαβ (x)hαβ (x).
This defines a binary operation on the set of line bundles. Tensoring
with the trivial bundle O gives the same bundle back, so it is the
neutral element of the group structure. Associated to each line bun-
dle L with transition functions gαβ , there is another line bundle L∗
−1
whose transition functions are gαβ . It is called the dual bundle of L.
∗
Since L ⊗ L = O, the dual bundle is like the inverse of L. Hence
we get a group structure on the isomorphism classes of line bundles
on M . This group is called the Picard group of M denoted Pic(M ).
In the next section we describe Pic(E) for E a smooth cubic curve
in P2 .
5.2. Sections of a line bundle. A section s of a line bundle L is a
holomorphic map s : M → L such that π◦s = Id. Locally this means
we have an open cover Uα and a collection of holomorphic functions
sα : Uα → C such that
sα (x) = gαβ (x) · sβ (x) ∀ x ∈ Uα ∩ Uβ .
An example of a section is given in Example 9. It may be the case
that a line bundle does not have any holomorphic sections. Lo-
cal holomorphic sections always exist but they may not satisfy the
patching condition on overlaps.
For instance consider the line bundle OP1 (−1) as in Example 14.
Suppose it has a local holomorphic section s1 (w) on U1 where s1 (w)
is a holomorphic function. Then on U0 ∩ U1 it transforms to s0 (z) =
GROUP LAW ON THE CUBIC CURVE 45
6. Poincaré Bundle
Now we restrict attention to the case of a plane cubic curve E and
ask ourselves the following question. What does the set Pic0 (E) of
all degree zero line bundles on E look like?
Here by degree of a line bundle we mean the degree of its associ-
ated divisor (Definition 4). In the case of P1 up to isomorphism there
is only one line bundle of degree zero and that is the trivial bundle.
However on E there are many non-trivial line bundles having degree
zero as we shall soon see. In fact they form a family parametrised
by E.
First let’s take a point P ∈ E. This is a divisor of degree 1 on
E, and it defines a line bundle O(P ). Now choose another point Q
distinct from P and take the divisor P − Q. This has degree zero
and correspondingly defines a line bundle O(P − Q). One could ask
is O(P − Q) isomorphic to the trivial bundle?
If so then by Lemma 17 there would exist some global meromor-
phic function f on E such that P − Q = (f ). This means that f
has exactly a pole of order 1 at Q and a zero of order 1 at P and no
other poles or zeroes. But then we can define a bijective map
E → P1
x 7→ (f (x) : 1)
Under this mapping Q maps to the point at ∞ = (1 : 0) on P1 . Since
f is meromorphic with exactly one pole and holomorphic elsewhere
it is an isomorphism. But E has genus 1 while P1 has genus 0 so
they cannot be isomorphic. Therefore P − Q ≁ 0, i.e. P , Q are
48 MADEEHA KHALID
References
[1] N. H. Abel Recherches sur les fonctions elliptiques J. reine angew. Math.
2, pages 101-181 (1827).
[2] J. B. Bost, An introduction to compact Riemann surfaces, jacobians and
Abelian varieties, From number theory to Physics, M.Waldschmidt et.al
(eds), Springer Verlag (1992).
GROUP LAW ON THE CUBIC CURVE 51
MARINA FRANZ
1. Introduction
First of all, we want to analyse periodic complex functions f :
C → C with respect to a lattice Λ. So let us fix once and for all
a complex number τ ∈ C with Im τ > 0 and consider the lattice
Λ := Z ⊕ τ Z ⊂ C.
τ V 1 + τ
0 1
def X
θ(z + 1) = exp(πin2 τ ) exp(2πin(z + 1))
n∈Z
X
= exp(πin2 τ + 2πinz + 2πin)
n∈Z
X
= exp(πin2 τ ) exp(2πinz) exp(2πin)
| {z }
n∈Z
=1 for all n∈Z
X
= exp(πin2 τ ) exp(2πinz)
n∈Z
def
= θ(z)
def X
θ(z + τ ) = exp(πin2 τ ) exp(2πin(z + τ ))
n∈Z
X
= exp(πin2 τ + 2πinz + 2πinτ )
n∈Z
if we complete the square and rearrange the summands then
X
= exp πin2 τ + 2πinτ + πiτ − πiτ
n∈Z
+2πinz + 2πiz − 2πiz)
X
= exp(−πiτ − 2πiz) exp(πi(n + 1)2 τ ) exp(2πi(n + 1)z)
n∈Z
if we make a simple index shift m = n + 1 then
X
= exp(−πiτ − 2πiz) exp(πim2 τ ) exp(2πimz)
m∈Z
def
= exp(−πiτ − 2πiz)θ(z)
Hence the basic theta function is not periodic with respect to the
τ -direction as in general exp(−πiτ − 2πiz) 6= 1.
56 MARINA FRANZ
= exp(−πip2 τ − 2πipz)
X
· exp(πi(n + p)2 τ ) exp(2πi(n + p)z)
n∈Z
if we make a simple index shift m = n + p then
X
= exp(−πip2 τ − 2πipz) exp(πim2 τ ) exp(2πimz)
m∈Z
def
= exp(−πip2 τ − 2πipz)θ(z)
Hence the basic theta function θ is quasi-periodic with
θ(z + λ) = θ(z + pτ + q)
= exp(−πip2 τ − 2πipz)θ(z)
for all λ = pτ + q ∈ Λ and z ∈ C.
Definition. We define
e(λ, z) := exp(−πip2 τ − 2πipz)
and call this the automorphy f actor.
Remark. We have e(λ1 + λ2 , z) = e(λ1 , z + λ2 )e(λ2 , z) for all λ1 ,
λ2 ∈ Λ.
THETA FUNCTIONS 57
· e(λ2 , z)
= exp(−πip21 τ − 2πip1 (z + λ2 ))e(λ2 , z)
def
= e(λ1 , z + λ2 )e(λ2 , z)
θ(z + ξ) = θ(z + aτ + b)
def X
= exp(πin2 τ ) exp(2πin(z + aτ + b))
n∈Z
X
= exp(πin2 τ + 2πinz + 2πinaτ + 2πinb)
n∈Z
X
θ(z + ξ) = exp πin2 τ + 2πinaτ + πia2 τ − πia2 τ
n∈Z
+2πin(z + b) + 2πia(z + b) − 2πia(z + b))
= exp(−πia2 τ − 2πia(z + b))
X
· exp(πi(n + a)2 τ ) exp(2πi(n + a)(z + b))
n∈Z
P
Note that the sum n∈Z exp(πi(n + a)2 τ ) exp(2πi(n + a)(z + b))
looks very similar to the sum in the definition of our basic theta
function above.
X
θξ (z) := θξ (τ )(z) := exp(πi(n + a)2 τ ) exp(2πi(n + a)(z + b))
n∈Z
X
θξ (z) = exp(πi(n + a)2 τ ) exp(2πi(n + a)(z + b)) (2)
n∈Z
for all z ∈ C.
(3)
θξ (z + λ) = exp(πia2 τ + 2πia(z + λ + b))θ(z + λ + ξ)
(1)
= exp(πia2 τ + 2πia(z + λ + b))e(λ, z + ξ)θ(z + ξ)
(3)
= exp(πia2 τ + 2πia(z + λ + b))e(λ, z + ξ)
· exp(−πia2 τ − 2πia(z + b))θξ (z)
= exp(2πiaλ) exp(−πip2 τ − 2πip(z + ξ))θξ (z)
= exp(2πiaλ − πip2 τ − 2πip(z + ξ))θξ (z)
θξ (z + λ) = θaτ +b (z + pτ + q)
= exp(2πiaλ − πip2 τ − 2πip(z + ξ))θξ (z)
Proof.
θσ (−z) = θ 12 τ + 21 (−z)
" 2 !
def X 1
= exp πi n + τ
2
n∈Z
1 1
exp 2πi n + −z +
2 2
if we make a simple index shift m = −n − 1 then
" 2 !
X 1
= exp πi −m − τ
2
m∈Z
1 1
exp 2πi −m − −z +
2 2
" !
X 2
1
= exp πi m + τ
2
m∈Z
1 1 1
exp 2πi m + z+ − 2πi m +
2 2 2
" !
X 2
1
= exp πi m + τ
2
m∈Z
1 1
exp 2πi m + z+
2 2
exp(−2πim) exp(−πi) ]
| {z }
=−1 for all m∈Z
def
= −θσ (z)
α : [0, 1] → C; t 7→ w + t
β : [0, 1] → C; t 7→ w + 1 + tτ
γ : [0, 1] → C; t 7→ w + (1 − t) + τ
δ : [0, 1] → C; t 7→ w + (1 − t)τ
w+τ γ w+1+τ
×0
δ
β
-
w α w+1
Note
γ(t) = w + (1 − t) + τ = α(1 − t) + τ
and
δ(t) = w + (1 − t)τ = β(1 − t) − 1
Z
1 θσ′
We want to show that (z) dz = 1.
2πi ∂Vw θσ
THETA FUNCTIONS 63
Z Z
1 θσ′ 1 θσ′
(z) dz = 1 − (z) dz
2πi γ θσ 2πi α θσ
and
Z Z
1 θσ′ 1 θσ′
(z) dz = − (z) dz
2πi δ θσ 2πi β θσ
Z Z 1
1 θσ′ 1 θσ′
(z) dz = (γ(t))γ ′ (t) dt
2πi γ θσ 2πi 0 θσ
1 ′Z
1 θσ
= (α(1 − t) + τ )(−1) dt
2πi 0 θσ
Z ′
1 θσ
=− (z + τ ) dz
2πi α θσ
Z ′
1 eσ (τ, z)θσ (z) + eσ (τ, z)θσ′ (z)
=− dz
2πi α eσ (τ, z)θσ (z)
Z ′ Z ′
1 eσ (τ, z) 1 θσ
=− dz − (z) dz
2πi α eσ (τ, z) 2πi α θσ
when we use
1
eσ (τ, z) = exp(2πi τ − πiτ − 2πi(z + σ))
2
then
Z
1 exp′ (−2πi(z + σ))
=− dz
2πi α exp(−2πi(z + σ))
Z ′
1 θσ
− (z) dz
2πi α θσ
Z Z ′
1 1 θσ
=− −2πi dz − (z) dz
2πi α 2πi α θσ
Z ′
1 θσ
=1− (z) dz
2πi α θσ
64 MARINA FRANZ
Z Z 1
1 θσ′ 1 θσ′
(z) dz = (δ(t))δ ′ (t) dt
2πi δ θσ 2πi 0 θσ
Z
1 ′
1 θσ
= (β(1 − t) − 1)(−τ ) dt
2πi 0 θσ
Z ′
1 θσ
=− (z − 1) dz
2πi β θσ
Z ′
1 eσ (−1, z)θσ (z) + eσ (−1, z)θσ′ (z)
=− dz
2πi β eσ (−1, z)θσ (z)
Z ′ Z ′
1 eσ (−1, z) 1 θσ (z)
=− dz − dz
2πi β eσ (−1, z) 2πi β θσ (z)
when we use
1
eσ (−1, z) = exp(−2πi )
2
then
Z Z ′
1 exp′ (−πi) 1 θσ (z)
=− dz − dz
2πi β exp(−πi) 2πi β θσ (z)
Z ′
1 θσ (z)
=− dz
2πi β θσ (z)
Then we have
Z Z ′ Z ′
1 θσ′ 1 θσ 1 θσ
(z) dz = (z) dz + (z) dz
2πi ∂Vw θσ 2πi α θσ 2πi β θσ
Z ′ Z ′
1 θσ 1 θσ
+ (z) dz + (z) dz
2πi γ θσ 2πi δ θσ
=1
Proof. We know θσ (z) = 0 if and only if z ∈ Λ and all the zeros are
simple. Hence
(3)
θξ (z) = 0 ⇔ exp(πia2 τ + 2πia(z + b))θ(z + ξ) = 0
(3)
⇔ exp πia2 τ + 2πia(z + b)
2
1
· exp −πi τ
2
1 1 1 1
−2πi z+ξ− τ − +
2 2 2 2
· θσ (z + ξ − σ) = 0
⇔z+ξ−σ ∈Λ
⇔z ∈σ−ξ+Λ
In particular we have θ(z) = 0 if and only if z ∈ σ + Λ.
w+τ γ w+1+τ
×
D a α
0 i0 6 i0
a × i
i1
δ bj× β
1
Dj′ 0bj× βj0
0 6
×
a
- i2
w α w+1
hence
Z Z Z
1 f′ 1 f′ 1 f′
z (z) dz = z (z) dz + z (z) dz
2πi ∂Vw f 2πi α f 2πi β f
Z Z
1 f′ 1 f′
+ z (z) dz + z (z) dz
2πi γ f 2πi δ f
Z ′ Z ′
1 f 1 f
= −τ (z) dz + (z) dz ∈ Λ
2πi α f 2πi β f
1
R f′ 1
R f′
since 2πi β f (z) dz, 2πi α f (z) dz ∈ Z.
Secondly we show that
Xn Xm Z
1 f′
ni − mj = (z) dz
i=1 j=1
2πi ∂Vw
f
=0
Again the first equality is clear, since Vw contains a representative
for every zero and pole of f in C/Λ.
The second equality follows from:
Z ′ Z 1 ′
1 f 1 f
(z) dz = (γ(t))γ ′ (t) dt
2πi γ f 2πi 0 f
Z 1 ′
1 f
= (α(1 − t) + τ )(−1) dt
2πi 0 f
Z 1 ′
1 f
=− (α(1 − t)) dt
2πi 0 f
Z ′
1 f
=− (z) dz
2πi α f
and
Z Z 1
1 f′ 1 f′
(z) dz = (δ(t))δ ′ (t) dt
2πi δ f 2πi 0 f
Z 1
1 f′
= (β(1 − t) − 1)(−τ ) dt
2πi 0 f
Z 1 ′
1 f
=− (β(1 − t))τ dt
2πi 0 f
Z ′
1 f
=− (z) dz
2πi β f
THETA FUNCTIONS 69
hence
Z Z Z ′
1 f′ 1 f′ 1 f
(z) dz = (z) dz + (z) dz
2πi ∂Vw f 2πi α f 2πi β f
Z ′ Z ′
1 f 1 f
+ (z) dz + (z) dz
2πi γ f 2πi δ f
=0
Qn ni
i=1 θσ (z − ai )
g : C → C; z →
7 Qm mj
j=1 θσ (z − bj )
and
Qn ni
i=1 θσ (z + τ − ai )
Qm
g(z + τ ) = mj
j=1 θσ (z + τ − bj )
Qn ni
i=1 (eσ (τ, z − ai )θσ (z − ai ))
Qm
= mj
j=1 (eσ (τ, z − bj )θσ (z − bj ))
Qn Qn
i=1 eσ (τ, z − ai )ni i=1 θσ (z − ai )ni
= Qm mj
Qm mj
j=1 eσ (τ, z − bj ) j=1 θσ (z − bj )
Qn ni
i=1 eσ (τ, z − ai )
= Qm mj
· g(z)
j=1 e σ (τ, z − b j )
70 MARINA FRANZ
but
Qn ni
Qn
i=1 e σ (τ, z − a i ) i=1 exp(−2πi(z − ai + σ))ni
Qm mj
= Qm mj
j=1 e σ (τ, z − b j ) j=1 exp(−2πi(z − bj + σ))
Qn ni
Q i=1 exp(−2πi(z + σ))
= m mj
j=1 exp(−2πi(z + σ))
Qn ni
i=1 exp(2πiai )
· Qm mj
j=1 exp(2πibj )
Pn
ni
exp(−2πi(z + σ)) i=1
= Pm
exp(−2πi(z + σ)) j=1 mj
Pn
exp(2πi i=1 ni ai )
· Pm
exp(2πi j=1 mj bj )
=1
So g(z + τ ) = g(z) as well. Hence g is doubly periodic w.r.t. Λ and
the function f : C/Λ → C with f ([z]) = g(z) is well-defined and a
solution.
Now suppose we are given two meromorphic functions f , g :
C/Λ → C with zeros [ai ] of order ni and poles [bj ] of order mj .
Then fg has no zeros or poles. Hence it is constant.
3. Weierstraß ℘-function
Now we want to capitalize on our work above. Therefore we con-
sider a very special periodic function, the Weierstraß ℘-function.
Definition. The W eierstraß ℘ − f unction is defined to be the
function ℘ : C → C given by
1 X 1 1
℘(z) = 2 + − 2
z (z − λ)2 λ
06=λ∈Λ
θ′
Note. The quotient θσσ isn’t doubly-periodic, but the derivative
′ ′
θσ
θσ is doubly-periodic.
′
θσ
To see this consider θσ (z + λ) for some λ = pτ + q ∈ Λ.
′
θσ′ (5) (eσ (λ, z)θσ (z))
(z + λ) =
θσ eσ (λ, z)θσ (z)
eσ (λ, z)θσ (z) + eσ (λ, z)θσ′ (z)
′
=
eσ (λ, z)θσ (z)
′ 2 ′
def exp (πiλ − πip τ − 2πip(z + σ))θσ (z) + eσ (λ, z)θσ (z)
=
eσ (λ, z)θσ (z)
−2πipeσ (λ, z)θσ (z) + eσ (λ, z)θσ′ (z)
=
eσ (λ, z)θσ (z)
θ′
= −2πip + σ (z)
θσ
′
θ
6= σ (z)
θσ
θ′ ′
θσ
as in general p 6= 0. From the equation θσσ (z + λ) = −2πip + θσ (z)
′ ′
θ
above it follows directly that θσσ is doubly-periodic.
′
′
θσ
If we add ℘ and θσ then we obtain
′ ′ X
θσ 1 1
℘(z)+ (z) = 2
− 2 + a1 + 2a2 z + 3a3 z 2 + . . .
θσ (z − λ) λ
06=λ∈Λ
′ ′
θ
From this sum we see directly that ℘ + θσσ doesn’t have any poles
′ ′
θ
in U . Hence ℘ + θσσ is holomorphic in a neighborhood of 0 and
thus holomorphic everywhere.
As it is in addition doubly-periodic
′
θ′
(since ℘ is as well as θσσ doubly-periodic) we know from our very
first lemma that it must be constant.
The Weierstraß ℘-function satisfies a number of equations and
differential equation. This feature makes the Weierstraß ℘-function
to be of interest. The most important differential equation that is
satisfied by the Weierstraß ℘-function is the following:
Theorem 10. The Weierstraß ℘ -function satisfies the differential
equation
℘′ (z)2 = c3 ℘(z)3 + c2 ℘(z)2 + c1 ℘(z) + c0
where the constants
X 1 X 1
c3 = 4 , c2 = 0 , c1 = −60 and c 0 = −140
λ4 λ6
06=λ∈Λ 06=λ∈Λ
This sum is absolutly convergent for all z ∈ C with |z| < |λ|; in
particular in a neigborhood of 0. P 1
To simplify the big sum from above we define sn := 06=λ∈Λ λn+2
for n ∈ N. Note that sn = 0 for all odd n ∈ N. We obtain
1
℘(z) = 2 + 2s1 z + 3s2 z 2 + 4s3 z 3 + 5s4 z 4 + . . .
z
1
= 2 + 3s2 z 2 + 5s4 z 4 + 7s6 z 6 . . .
z
which is true in a neigborhood of 0. With the constants
X 1 X 1
c3 = 4 , c2 = 0 , c1 = −60 and c 0 = −140
λ4 λ6
06=λ∈Λ 06=λ∈Λ
we obtain
1 c1 2 c0 4
℘(z) = − z − z + terms of higher order
z2 20 28
hence
2 c1 c0 3
℘′ (z) = − − z − z + terms of higher order
z 3 10 7
4 2c1 1 4c0
℘′ (z)2 = 6 + + + terms of higher order
z 5 z2 7
and
1 3c1 1 3c0
℘(z)3 = − − + terms of higher order
z6 20 z 2 28
Now consider
f (z) := ℘′ (z)2 − c3 ℘(z)3 − c1 ℘(z) − c0
The series of f has only positive powers of z. Hence f is holomorphic
around 0. Hence it is holomorphic everywhere. And as it is doubly-
periodic, it is constant. But the constant part of the series is 47 c0 +
3
4 · 28 c0 − c0 = 0. Hence f = 0.
and
e1 + e2 + e3 = 0
1
e 1 e 2 + e 1 e 3 + e 2 e 3 = c1
4
1
e 1 e 2 e 3 = − c0
4
where c0 and c1 are the constants from above.
Finally we will see how to use the Weierstraß ℘-function to give
a group structure to an elliptic curve.
Remark. If we consider the elliptic curve
C := {(x, y) ∈ C2 such that y 2 = c3 x3 + c2 x2 + c1 x + c0 }
for the constants
X 1 X 1
c3 = 4 , c2 = 0 , c1 = −60 and c0 = −140
λ4 λ6
06=λ∈Λ 06=λ∈Λ
References
[1] A. Gathmann Algebraic Geometry, Notes for a class taught at the Univer-
sity of Kaiserslautern 2002/2003, available at http://www.mathematik.uni-
kl.de/ gathmann/de/pub.html
[2] M. Khalid Group Law on the Cubic Curve, this issue
[3] D. Mumford Tata Lectures on Theta, Progress in Mathematics Vol 28,
Birkhauser Verlag, 1983
[4] G. Trautmann Complex Analysis II, Notes for a class taught at the Univer-
sity of Kaiserslautern 1996/1997
CIARA DALY
1. Introduction
Nowadays, vector bundles play an inportant role in many areas
of mathematics such as algebraic geometry, algebraic topology and
differential geometry, in the theory of partial differential equations.
The theory of vector bundles and the mathematical formalism
developed over the years, for the study of vector bundle related con-
cepts leads to the clarification or solution of many mathematical
problems. Some of the vector bundle related concepts are general-
isations of well-known classical notions. For instance, the notion of
a section of a vector bundle over a space X is a generalization of a
vector valued function on X.
One of the important problems is the problem of classification
of bundles. The problem of classification of vector bundles over an
elliptic curve (i.e. a nonsingular projective curve of arithmetic genus
one) has been completely solved by Atiyah in [1].
2. Preliminaries
For the purpose of this paper, X will denote a complex manifold,
unless otherwise specified. We will be working over C, the field of
complex numbers, throughout this paper.
75
76 CIARA DALY
ϕi
p−1 (Ui ) / U i × Cn
GG w
GG
GG www
p GG
w
G# www pr1
{w
Ui
commutes.
Moreover, ϕi takes the vector space Ex isomorphically onto {x} ×
n
C for each x ∈ Ui ; ϕi is called a trivialisation of E over U . Note
that for any pair of trivialisations ϕi and ϕj , the map
gij : Ui ∩ Vj → GL(n, C)
given by
gij (x) = ϕi ◦ (ϕj |{x}×Cn )−1
is holomorphic; the maps gij are called transition functions for E
relative to the trivialisations ϕi , ϕj . The transition functions of E
necessarily satisfy the identities
gij (x) · gji (x) = I for all x ∈ Ui ∩ Uj
f
E@ / E′
@@ }
@@ }}}
p @@@ }} ′
~}} p
X
commutes and for each point x ∈ X the map f |Ex : Ex → Ex′ is a
homomorphism of vector spaces.
Let E and E ′ be vector bundles over X with rank r and r′ , re-
spectively and let {Ui } be an open cover of X such that E and E ′ are
trivial over Ui for each i. A morphism f : E → E ′ can be described
locally by holomorphic functions, fi , as follows. For each i, using
trivialisations of E and E ′ , f induces maps
′
U i × Cr → U i × Cr , (x, v) 7→ (x, fi (x)v)
where fi : Ui → Matr′ ×r (C). These holomorphic functions necessar-
ily satisfy
′
fi (x) · gij (x) = gij (x) · fj (x) for all x ∈ Ui ∩ Uj
′
where gij and gij are transition functions of E and E ′ , respectively.
Note that a set of functions {fi } defines an isomorphism of vector
bundles if an only if fi (x) are invertible matrices for all i and x.
78 CIARA DALY
we get another exact sequence in which the last map is not surjective
in general
g
0 −−−−→ Hom(E, F ′ ) −−−−→ Hom(E, F ) −−−−→ Hom(E, F ′′ )
(2)
And so in a similar fashion to the way we defined cohomology
groups H i (E), we can define what are called the Ext groups, which
allow us to extend our short exact sequence (2) to a long exact
sequence as follows:
0 −−−−→ Hom(E, F ′ ) −−−−→ Hom(E, F ) −−−−→ Hom(E, F ′′ )
−−−−→ Ext1 (E, F ′ ) −−−−→ Ext1 (E, F ) −−−−→ ···
We say that Exti (E, ·) are the right derived functors of Hom(E, ·). So
in particular we have Ext0 (E, ·) = Hom(E, ·). We have the following
proposition to see the relationship between the cohomology groups
H i and the Ext groups (for a full proof of this proposition see [5]
Proposition 6.3).
Proposition 2.11. For any vector bundle E on a complex manifold
X we have:
Exti (OX , E) ∼
= H i (E) for all i ≥ 0.
Similarly we have:
Exti (E, OX ) ∼
= H i (E ∗ ) for all i ≥ 0.
Proof. Here we will just give a proof of the first statement, where
i = 0. Let f ∈ Hom(O, E), a fibre-wise holomorphic morphism such
that the following diagram commutes
f
X ×C /E.
GG ~
GG ~
GG ~~~
pr1 GG ~ p
G# ~~
~
X
Let s : X → E be a holomorphic section of E, i.e. p ◦ s = idx . Define
two linear maps
α : Hom(O, E) → H 0 (E)
and
β : H 0 (E) → Hom(O, E)
as follows: Define α(f )(x) := f (x, 1) and β(s)(x, λ) := λ · s(x).
We see that β(α(f )) = f as follows:
RANK TWO VECTOR BUNDLES ON ELLIPTIC CURVES 81
Remark 2.15. Recall from [7] that the degree of a line bundle is the
degree of it’s associated divisor and every line bundle can be written
as O(D), with D a divisor on X.
If E lies in an exact sequence of vector bundles on X as follows:
0 → E ′ → E → E ′′ → 0
then there is an isomorphism
det E ′ ⊗ det E ′′ ∼
= det E.
Since det E ′ and det E ′′ are line bundles we get deg E = deg E ′ +
deg E ′′ (as in general deg(L ⊗ L′ ) = deg L + deg L′ , where L and L′
line bundles). In other words, degree is additive on exact sequences.
As a special case, if a vector bundle E = L1 ⊕ L2 is the direct sum
of two line bundles L1 and L2 , then we have
0 → L1 → E → L2 → 0
with deg E = deg L1 + deg L2 .
Remark 2.16. Let E be a vector bundle of rank r over a complex
manifold X. If we tensor E with a line bundle L, then deg(E ⊗ L) =
deg E + r deg L. In particular for E of rank 2 we have deg(E ⊗ L) =
deg E + 2 deg L. To see this let’s look at an example:
Let E = L1 ⊕ L2 for line bundles L1 and L2 . Tensor this with
another line bundle L to get
E ⊗ L = (L1 ⊕ L2 ) ⊗ L = (L1 ⊗ L) ⊕ (L2 ⊗ L)
Now
deg(E ⊗ L) = deg(L1 ⊗ L) + deg(L2 ⊗ L)
= deg L1 + deg L + deg L2 + deg L
= deg L1 + deg L2 + 2 deg L
= deg E + 2 deg L.
In this way, when considering vector bundles of rank 2, it is enough
to consider vector bundles of degree −1 or 0 (or indeed, any even
and odd degree), then by tensoring with a line bundle of appropriate
degree we get all other degrees. We call this the “tensor product
trick”.
Definition 2.17. An exact sequence of vector bundles
0 → E ′ → E → E ′′ → 0
RANK TWO VECTOR BUNDLES ON ELLIPTIC CURVES 83
3. Vector bundles on P1
Before we move on to vector bundles on an elliptic curve (i.e. a
curve of genus one), it makes sense to look at vector bundles on a
curve of genus zero (P1 ). Let us now restate Lemma 2.22 in the case
of P1 , where genus g = 0, to see how the cohomology of line bundles
on P1 is particularly simple.
Lemma 3.1. Let L be a line bundle on P1 . Then we have the
following:
(a) H 0 (L) = 0 if deg L ≤ −1.
(b) H 1 (L) = 0 if deg L ≥ −1.
By Riemann-Roch we also have,
h0 (L) − h1 (L) = deg L + 1.
Remark 3.2. We have seen from [7] that for L a line bundle, L∗ is
the inverse of the line bundle L in the Picard group. We have also
seen that deg : Pic X → Z is a homomorphism (where Pic X denotes
the set of line bundles over X) and so we get deg L∗ = − deg L.
86 CIARA DALY
? ?
0 /M / L̃ /L /0
RANK TWO VECTOR BUNDLES ON ELLIPTIC CURVES 87
r−1
M
0→M →E→ L1 → 0
i=1
splits.
for all i = 2, . . . n.
0 = E0 ⊂ E1 ⊂ E2 ⊂ · · · ⊂ Ek = E
0 → E1 → E → E ′ → 0
0 ⊂ F2 ⊂ F3 ⊂ · · · ⊂ Fk = E ′
with
µ(F2 ) > µ(F3 /F2 ) > · · · > µ(Fk /Fk−1 )
? ?
0 / E1 / Ej / Fj / 0.
Proof. ([8] Proposition 5.4.2) Assume (Ei )i=1,...,n and (Fj )j=1,...,m
are two filtrations of E satisfying the conditions of Proposition 5.7
above. Now using the notation of Lemma 5.8 if we let E ′ := F1
we get µ(F1 ) ≤ µ(E1 ). Similarly if we allow E ′ := E1 , we get
µ(E1 ) ≤ µ(F1 ). Clearly then, µ(F1 ) = µ(E1 ).
Lemma 5.8 again implies E1 ⊂ F1 and F1 ⊂ E1 , hence E1 = F1 .
Using E/E1 and F/F1 we can proceed by induction as in the proof
of Proposition 5.7 to conclude that the filtration is unique.
The filtration of Proposition 5.7 is called the Harder-Narasimhan
filtration of E.
6. Conclusion
In conclusion, it is fair to say that the theory of vector bundles is
vast and indeed very interesting. We have seen how vector bundles
on P1 are not very complex, in the sense that they can be written as
a direct sum of line bundles. We have also seen a classification for
indecomposable rank 2 vector bundles on elliptic curves.
One could also go on to study higher rank vector bundles on
elliptic curves or on curves of a higher genus, or even on higher
dimensional complex manifolds. These are all very interesting in
their own right.
In Section 5, we studied slope stability for vector bundles on
curves. As was mentioned, stable bundles are required when con-
structing moduli spaces of vector bundles. For the reader interested
in stability, from here you could go on to study Bridgeland stability
conditions ([2] and [3]) and the space of all stability conditions on a
particular complex manifold (e.g. an elliptic curve).
RANK TWO VECTOR BUNDLES ON ELLIPTIC CURVES 97
References
[1] M. F. Atiyah: Vector bundles over an elliptic curve, Proc. London Math.
Soc. (3) 7 (1957) 414-452.
[2] T. Bridgeland: Stability Conditions on Triangulated Categories, preprint
arXiv:math.AG/0212237.
[3] T. Bridgeland: Spaces of Stability Conditions, arXiv:math.AG/0611510.
[4] A. Gathmann: Algebraic Geometry, Notes for a class taught at the Univer-
sity of Kaiserslauten (2002/2003) available at http://www.mathematik.uni-
kl.de/∼gathmann/class/alggeom-2002/main.pdf.
[5] R. Hartshorne: Algebraic Geometry, Graduate Texts in Mathematics,
Springer, (1977).
[6] D. Huybrechts, M. Lehn: The geometry of moduli spaces of sheaves Aspects
of Mathematics, E31. Friedr. Vieweg & Sohn, Braunschweig, (1997).
[7] M. Khalid: Group law on the cubic curve, this issue.
[8] J. Le Potier: Lectures on Vector Bundles Cambridge studies in advanced
mathematics, Cambridge University Press (1997).
[9] S. Mukai: An Introduction to Invariants and Moduli, Cambridge studies in
advanced mathematics, Cambridge University Press, (2003).
[10] C. A. Weibel: An introduction to homological algebra, Cambridge studies
in advanced mathematics, Cambridge University Press, (1994).