In this post, I will rely on this last post for results and notations, for example will also mean a compact Riemann surface. We will note
the canonical bundle of
, i.e. the bundle with sheaf of sections
the holomorphic 1-forms on
. We will also note
the dimension of
by which we really mean
the j’th cohomology group of
in the sheaf of holomorphic sections of
. I’ll also write
to mean
where
is the line bundle associated with the divisor
, and sometimes the line bundle
might get written
accordingly. We will sometimes also suppress the
and simply write
for
. Let us first state the Riemann-Roch theorem in the form that we aim to get to:
Theorem (Riemann-Roch): For any holomorphic line bundle
, the following relation holds:
.
We will freely use the fact that any line bundle on a compact Riemann surface actually comes from a divisor, which can be seen by Jacobi’s inversion theorem, and define the degree of simply as the degree of its associated divisor. To be more precise, we will do the proof for line bundles coming from generic divisors
with the
‘s distinct. We will first prove that
which is relatively easy. The hard part will be
Theorem (Serre duality): There is a natural coupling
which induces an isomorphism
In particular, this tells us that and Riemann-Roch follows.
Proof of the first part: We show the first part by induction on the degree of If
, i.e. if
is trivial, then by the maximum principle
and
. So
and we are done with the base case. Now, suppose that we can represent as
for a divisor
with the plus or minus sign in accordance with the sign of
, and with all
‘s distinct.
Suppose by induction that the statement is true for . Then it suffices to show it is true for the line bundles associated to
and
for some divisor
of degree
not containing
. In the case of positive degree, we use the short exact sequence
.
Note that this short exact sequence was also considered in my last post with trivial . Here we do the usual identification between sections
with meromorphic functions
such that
and the first arrow corresponds to the inclusion of families of these spaces of meromorphic functions. We get a long exact sequence
hence adding alternating dimensions we find
.
Using our induction assumption, i.e. that , we deduce that
which is what we wanted.
For the second case, when , we use another but related short exact sequence:
.
A similar argument shows that
and we are done.
QED
Before attacking Serre duality, recall that for any line bundle there is an operator
where is the sheaf of smooth
-forms with values in
. It is the unique such operator satisfying Leibniz rule
and such that
is exact. It is called the del-bar operator associated to . Analogously to last post, we get a Dolbeault isomorphism
.
The identifications we discovered at the end of last post still hold in this context, with the necessary adjustments. Given two line bundles , this isomorphism permits us to define a pairing
by setting
.
In particular, with and
, we get Serre’s pairing
where the last isomorphism is (which follows from the same argument as for
) composed with integration
.
Theorem (Serre duality): This pairing is non-degenerate.
Proof: We will proceed by induction as in the proof of Riemann-Roch. We say that satisfies Serre duality if Serre’s pairing induces an isomorphism
and
.
Note that satisfies Serre duality if and only if
does.
The base case is the statement that the pairing given by
is non-degenerate, which is just the isomorphism
given by integration coupled with the fact that
is injective.
Suppose now by induction that all line bundles with
satisfy Serre’s duality. Like in Riemann-Roch, it suffices to show that
and
satisfy Serre duality (where
with
not containing
). We will now identify
with meromorphic sections of
with at worst a simple pole at
. We thus get the short exact sequence
where the last arrow is (recall that the residue of a meromorphic function at
is naturally an element of the holomorphic tangent space
, as discussed in my last post and under our identifications,
). We obtain in this way our first exact sequence
.
Recall from last post that if for
, then
for
an extension of
in the neighborhood
of
.
Similarly, we see elements of as holomorphic sections of
vanishing at
, and we have the following short exact sequence of sheaves:
where now the last arrow is simply . This gives our second exact sequence
.
Dualizing the first exact sequence, we get the following diagram:
where the vertical arrows are given by Serre’s pairing. By the Five Lemma and by using our induction hypothesis, it suffices to show that this diagram is commutative. The only difficult parts are obviously the two squares in the middle.
Actually, to be exactly precise, we need to consider instead of just
and similarly
.
For the third square, consider (so
is a holomorphic 1-form and
a section of
). We need to see that for all
, we have
.
But we have seen in last post that if near
, then after identifying
with
, we have
. Thus if
, we find
which, since is supported near
, by using Stokes theorem and Cauchy’s formula, is equal to
for a small circle around
. But
, so this shows the third square commutes.
The fourth square is similar. Take . Then unwinding the definition of
, we see that
if in some coordinates around
. Then under the Dolbeault isomorphism, this class corresponds to
and it acts on
by
where , for
and
. Like for the third square, this integral can be rewritten
for a small loop around
and
the residue of
in the
coordinate. But this is exactly what we needed since
.
In this post, will mean a compact Riemann surface (without boundary).
Let us start by considering a form of the Mittag-Leffler problem: Given a point , can we find a global meromorphic function
such that
and
? i.e. such that the only poles of
is a simple pole at
with residue
in some coordinates. Since this would give a biholomorphic function
, this is only possible if the genus of
is zero, but let us see how this obstruction manifests itself.
We will consider a covering of
where
is a coordinate patch centered at
with (holomorphic) coordinate
, while
is just
. We can reformulate the question as follows: does there exist
such that
and
? NB: Here
means the restriction
.
Before going further, note that the residue of a meromorphic function is only invariantly defined as a tangent vector: if has a simple pole at
, then we define, for
,
where is any extension of
, and where
is a well-defined residue for any meromorphic 1-form defined around
, where
is a small loop around
. Another way to see this is to take
a holomorphic coordinate around
and
a meromorphic function with a simple pole at
, say
for
holomorphic. Then if
is another choice of coordinates, then
so
for some holomorphic . In other words,
i.e.
is well-defined as an element of .
First approach: Our first approach to solving the Mittag-Leffler problem is with Cech cohomology: What we want is to find a global meromorphic function such that on
we have
. Thus we see
, as a Cech 1-cochain in the sheaf of holomorphic functions and ask that
on
for
, i.e. that
defines a coboundary in
. Indeed,
would then be holomorphic on
with
near
.
Another way to realize this is via the following short exact sequence of sheaves:
where is the sheaf of meromorphic functions having at worst a simple pole at
and where we consider
as the associated skyscraper sheaf. From the induced long exact sequence we get
and in this formulation, finding a meromorphic function with at worst a simple pole at with residue
is possible if and only if
. But unwinding the definition of
, we see that
iff
such that
and
in
. Thus we can take
and
, and
.
We once again come to the conclusion that the problem is solvable exactly when is a coboundary.
Second approach: Consider a smooth cut-off function supported in
and identically equal to 1 near
. Then
and the probem reduces to finding
such that
. Note that
near
so we can consider
as an element of , i.e. a
-form on
. Reformulating this, we can solve the problem exactly when
.
To relate this to our first approach, we note the sheaf of
-forms and use the short exact sequence of sheaves
which gives the exact sequence
.
In other words, this gives an isomorphism
Unwinding the definition of , we see that
iff
in
, for
.
This isomorphism is the simplest instance of the so-called Dolbeault isomorphism which is a similar isomorphism
valid for arbitrary compact complex manifolds and complex vector bundle
, where
is the sheaf of holomorphic p-forms with values in
.We can recap our discussion with the following:
Proposition: The obstruction to solving the Mittag-Leffler problem coming from our second approach,
which is represented by the 1-cocycle
, corresponds under Dolbeault’s isomorphism to
, the obstruction from our first approach.
In this post, I will discuss Jacobi’s inversion theorem. It is a follow up in a series of posts about Riemann surfaces, the first of which being this one. This theorem tells us in what sense the Abel-Jacobi map is surjective.
Theorem (Jacobi’s inversion): Let
be a compact Riemann surface of genus
and take any
. Then for any
, we can find
points
such that
.
In other words, for
a basis of
, for all
, there is
and paths
from
to
such that
for all
.
With this theorem coupled with Abel’s theorem, we will have completely proved the exactness of
.
Lemma 1: The set of effective divisors of degree
on
is a compact complex manifold.
Proof of lemma 1: Consider the action of the permutation group on
(
times).
The quotient, denoted , inherits the quotient topology making
a continuous map. Clearly,
is in bijection with
. Suppose for a moment that
. We will write an element of
as a sum
to indicate the unimportance of the ordering (note that the
‘s are not necessarily distinct). We consider the map
that takes to the d-tuple
consisting of the coefficents of the monic polynomial having
as its roots, i.e
if
.
In other words, where
is the ith elementary symmetric polynomial. By the fundamental theorem of algbera,
is a bijection, and in fact a homeomorphism.
The difficulty is in seeing that is also continuous. To see this, take an open set
and consider
. In other words,
with
.
We need to show that there is an open set . Write
which, when defined, is the number of zeros of the polynomial inside
. We can choose an open set
containing
small enough so that
is defined for all
. Then since
is continuous and takes only integer values, we have
for all
.
But this means that for all , all the roots of
are in
so that
as we wanted.
This gives the structure of a complex manifold of dimension
, in fact biholomorphic to
.
Now we put a similar complex manifold structure on for arbitrary
. Consider
and take holomorphic charts
around
on
such that
if
and
if
. We obviously have an injective mapping
where , by doing as above;
.
The verifications that this gives a homeomorphism onto an open set of is just as in the earlier case. These maps thus provide
with a holomorphic atlas.
QED
Fixing a base point we get a (holomorphic) injection
and thus holomorphic mappings
(mod
)
by composing and
. Jacobi’s inversion theorem says that
is surjective.
Lemma 2: Let be a holomorphic map between two compact connected complex manifolds of the same dimension. If
is not everywhere singular, i.e. the Jacobian matrix
is not identically zero, then
is necessarily surjective.
Proof of lemma 2: This is immediate from the proper mapping theorem which says that if is an analytic subvariety, then
is an analytic subvariety of
. In this case,
would be a compact subvariety containing an open set which would mean
. Griffiths-Harris presents a more elementary proof which does not use the rather deep proper mapping theorem:
Consider a volume form on
. Since
is not identically
and since
preserves the orientation (being holomorphic), we have
.
Since for any we have
the volume form is exact in
and
for some -form
on
. But then if
, we have
,
which contradicts earlier considerations.
QED
To prove the theorem, we thus have to show that is not everywhere singular.
Proof of the theorem: At points such that the
‘s are distinct, the quotient map
is locally a biholomorphism. So choosing disjoint charts
in
centered at
, we get a chart
.
In such coordinates, for near
, we have
mod
.
So
.
But with near
, this is
hence
The jacobian matrix of near
is thus given by
.
It suffices to see that this matrix is of full rank for some choice of and basis
. We simply do Gauss reduction: Choose
such that
. Then subtracting a multiple of
to
, we make it so that
.
The ‘s are still a basis of
and we continue like this, choosing a
such that
etc… We eventually arrive at the form
which is of maximal rank since for all
. This shows that
is not everywhere singular, so it is surjective by lemma 2.
QED
Putting everything together, we showed that the set of (isomorphism classes) of topologically trivial holomorphic line bundles (i.e. with zero first Chern class) has the structure of a complex torus
of dimension
. Indeed, the group
consisting of those (isomorphism classes of) holomorphic line bundles is isomorphic to
, which by Abel’s and Jacobi’s theorem is isomorphic to
.
Note that in fact the fibres of consist of projective spaces. Indeed, if
, then by Abel’s theorem the fiber is
the set of effective divisors linearly equivalent to
, which corresponds to the projectivisation of
. It can be shown that generically the fiber of
is a point and that
is a birational map.
In last post we proved Abel’s theorem and as a corollary we saw that any compact Riemann surface of genus
(a smooth elliptic curve) is biholomorphic to a complex torus, the biholomorphism being given by the Abel-Jacobi mapping:
which is given by
(mod
)
where some ,
is a holomorphic 1-form and
is the period lattice
for
the two period vectors
for
giving a basis for
.
In particular, every such Riemann surface has a group structure and we will see that this description of the group structure on an elliptic curve is the same as the usual one given for elliptic curves in
given by a cubic polynomial. By the usual group structure I mean the following: recall that by Bézout’s theorem, if you intersect a line in
with the zero locus of a cubic polynomial, you get (counting multiplicities) exactly 3 points. Thus to add 2 points
, you consider the line between them and declare that
if
is the third point of intersection of this line and the elliptic curve.
After a preliminary discussion of projective embeddings of Riemann surfaces, we will see that in fact any genus 1 compact Riemann surface can be embedded in as the zero locus of some cubic polynomial. In particular when the initial surface is a complex torus
, this embedding is given by the so-called Weierstrass
-functions. We will then see that this embedding is actually an inverse to the Abel-Jacobi mapping.
First the remarks on projective embeddings of compact Riemann surfaces. Suppose is a holomorphic line bundle over a compact complex manifold
and
are a basis for
the global holomorphic sections of
. Suppose that there is no point on
such that every sections in
vanish simultaneously. Then for every
, we have a non-zero vector
.
By choosing a trivialisation, we can consider this vector as sitting in . Of course this vector will depend on the choice of trivialisation of
around
but different choices of trivialisations will only change the vector by a complex multiple. We thus get a map
defined by
.
Kodaira’s embedding theorem states that when is a positive line bundle (i.e. admits a connection of positive curvature), there exists a
such that for
, the map
is well-defined (the global sections of are never all vanishing) and is an embedding of
.
We can in fact give a sharper version of Kodaira’s embedding theorem for compact Riemann surfaces (you can find a detailed discussion of this on pages 214, 215 of Griffiths-Harris):
Theorem: Let
be a compact Riemann surface and
a holomorphic line bundle. If
, then
is well-defined and an embedding.
In this proposition, the degree of a line bundle is just under the identification
or equivalently, since if
then the Poincaré dual
, the degree is
. Also,
is the canonical bundle of
and
.
The case that will interest us is when the genus of is
. What we show is that in fact we can take
. Indeed, since
because
is trivial (
is topologically a torus), for any
the above theorem tells us that for the bundle
, the map
gives an embedding of
into
for
If
, then
are 3 linearly independent global holomorphic sections so
. On the other hand, a global section of
corresponds to a meromorphic function on
having poles only at
and at worst of degree
. Such a function is completely determined by the first four coefficients in its Laurent expension around
:
(the difference of two such functions having all of these 4 numbers equal would be a global holomorphic function vanishing at hence everywhere). Moreover, two such functions having the same
and
must have the same
because otherwise their difference would be a meromorphic function having only one pole at
and of order 1, which is impossible if the genus is not zero. So we need at most 3 parameters to determine such a function, hence
. This shows
.
We can write this embedding explicitly as follows: by Kodaira’s vanishing theorem, we have so considering the long exact sequence associated to the short exact sequence
where is the skyscraper sheaf at
, we find
so there exists a meromorphic function
on
having a double pole at
and no other poles. Since
, there is also on
a nonzero holomorphic 1-form
. But as remarked earlier, since the canonical bundle
is trivial, the divisor
has zero Poincaré dual so is itself zero, i.e.
is nowhere vanishing. We consider
. It is a meromorphic form with only one pole at
, of order 2, and since the sum of the residues of a meromorphic 1-form must be zero,
.
So if is a local holomorphic coordinate near
, we can write
near
after possibly multiplying by a constant and adding a constant. Considering also the meromorphic function , which is holomorphic except at
where it has a triple pole, we get a global meromorphic function
which is
near
for the right choice of ‘s. Since
and
are three linearly independant global sections of
, they form a basis and thus the embedding
is given by
.
We can use this to explicitly describe as the zero set of a polynomial in
: around
, we have
and
.
Thus the meromorphic function has at most a simple pole at
and no other poles, so is constant. This tells us that
is included in the zero locus of the polynomial
which can be rewritten by making suitable linear changes in the coordinates by
for . Note that here
is a set of affine coordinates in
on the open set
. Since there are both topological torus (the second by the degree-genus formula), they must be equal, i.e. the zero locus of this polynomial in
is exactly
.
We will show that in fact the above construction gives an inverse for the Abel-Jacobi map which realises such a non-singular planar curve as a complex torus.
Consider and
as above, with their pole at
. Take
as a basis for
and
the coordinate on
such that
. In this context, the function
is the so-called Weierstrass
-function. Its derivative
is denoted
. If the Laurent expansions of
around
contained a term of odd degree, then killing the order 2 pole by
we would obtain a noncontant holomorphic function on
. So it has no term of odd degree and we can write
.
We find the relation
,
where and
. The corresponding embedding is then
where
,
and is the zero locus of the polynomial
,
written in suitable affine coordinates. On the other hand, the Abel-Jacobi map from to
is given by
(mod
)
with let’s say . Indeed,
so
is a non-zero holomorphic 1-form on
. Moreover,
is actually inverse to the embedding
because
.
We have thus found an inverse for the Abel-Jacobi mapping of a smooth elliptic curve in .
We are now in a position to discuss the group structure from different point of views. Any Riemann surface of genus
inherits a group structure simply by letting
for
There is also a group structure induced by an embedding as above :
if
are colinear in
. These are in fact the same. To see this, consider
as embedded in
, take 3 points
and denote
the corresponding points in the complex torus. Then by Abel’s theorem,
iff
with
.
But suppose for
, i.e. with
the line joining
(in affine coordinates), we also have
. Then
is a meromorphic function on the torus having divisor so the function
has divisor
. Thus if
are collinear, we have
and conversely, if there is a meromorphic function
on
such that
, then
is a meromorphic function with
and this forces
because otherwise we would have a meromorphic function having a simple pole at
and no other poles.
In the first post of this series, I defined the Abel-Jacobi map from a compact Riemann surface
of genus
to its jacobian
. Recall that the jacobian of
is defined to be the complex torus
where
is the period lattice
for the period vectors
associated to some basis of the space
of holomorphic 1-forms, and where
are cycles giving a canonical basis of
. For
a divisor of degree 0, the Abel-Jacobi map is then
(mod
).
The goal of this post of to prove the following result:
Theorem (Abel): Let
and suppose
is a basis of
. Denote by
the meromorphic functions on
. Then
for some
if and only if
.
Proof: The only if part is quite easy. Suppose is the divisor associated to a meromorphic function
, i.e.
for
respectively the zeros and poles of
. The idea is to view
as a holomorphic function
.
Then has a well-defined degree
which is the number of points (counted with multiplicity) in a fiber
. Let’s note
the divisor which is the fiber of
at
. Then the degree of
is zero for all
, and
(mod
)
for some chain in
with
. Since the set of points in
vary analytically with
(
is a locally
with
), we obtain a holomorphic function
by defining
.
But since is simply connected, it lifts to a holomorphic function
which must be constant by the maximum principle. Hence
itself is constant, which means in particular that
, i.e. that
.
Since is additive and
, this tells us that
when
, which is what we wanted to show.
The converse is harder. Given a divisor of degree
, where
and the
are all distinct, such that
, we search for a meromorphic function
such that
. We first reduce the problem to the existence of a certain differential of the third kind:
Lemma 1: There is an with
if and only if there is a meromorphic 1-form
such that
1. ,
i.e. has a simple pole exactly at the zeros and poles of
;
2. and
,
i.e. the residues at those poles are given by the order of ;
3. for any loop
in
.
The correspondence being given by
and
for any .
Proof of lemma 1: Let be such a meromorphic function and define
as above. Then since
is locally given by
for some
, we have
locally, with
. So
has no zero and a simple pole at every zero and pole of
, with residue at a point equal to the order of
at that point. Conditions 1. and 2. are thus satisfied for this
. To see that the periods of
are integers, we write
for a loop in
where
are small loops around
and
. Then
.
Conversely, given a meromorphic 1-form satisfying conditions 1., 2. and 3., we let
be the function
for some not in
. Note that condition 3. insures that this function is well defined. Near one of
‘s simple pole
, we can write
for some never-vanishing holomorphic function around
. Then for
sufficiently near
, we have in these coordinates
which gives
for never vanishing holomorphic functions around
. Similarly around a point
we find
and we conclude that
,
concluding the proof of the lemma
QED
To construct such a meromorphic 1-form, we will use another lemma:
Lemma 2: Given a finite number of points on
and complex numbers
such that
, there exists a meromorphic 1-form
with only simple poles having its poles exactly at the points
with residue
at
.
Proof of lemma 2: We consider the exact sequence of sheaves
where is the sheaf of meromorphic 1-forms having only simple poles exactly at the points
, i.e. if
, then
, and where
is the skyscraper sheaf around
. By Kodaira-Serre duality (see for example p. 153 in Griffiths & Harris, replacing
with the trivial line bundle),
where the last isomorphism comes from the maximum principle. Then the long exact sequence gives the exact sequence
i.e.
so the residue map has 1-dimensional cokernel, i.e. the image of is of codimension at most 1. But if
, then by the residue theorem
hence the image of
by the residue map, which is a linear subspace of codimension at most 1, is contained in the hyperplane
which is of codimension 1. Hence these two sets are equal, i.e.
which is exactly what the lemma says.
QED
Back to the proof of the theorem: Given our divisor of degree 0, this last lemma tells us we can find a meromorphic 1-form
satisfying conditions 1. and 2. of the first lemma. All that remains to be done is to show that we can perturb this
such that its periods are integers, i.e. such that it satisfies condition 3. of the lemma. Since the biholomorphism class of
is independant of the choice of basis
for
and
for
, we may take
a canonical basis and
normalised with respect to this basis, i.e. such that
for
.
Recall from last post that such a choise of basis for is possible as a consequence of the reciprocity law. Let
satisfy conditions 1. and 2. and denote its periods by
(
).
By correcting with a linear combinations of the
‘s, we can suppose its A-periods vanish. Indeed, take
.
Then still satisfies conditions 1. and 2. because the
‘s are holomorphic, but clearly for
, we have
The game is now to add an integral linear combination of the ‘s to
to make its B-periods integers. By the reciprocity law,
.
So since for
and the basis
is normalised, by condition 2. we find
for all
for a proper choice of path in this last integral (take path from to
circling the right amount of times along the
‘s to incorporate the
‘s in the integral). Let us denote by
those paths on which we integrate in this last expression. Since
by hypothesis, there exists a cycle
with
such that
for all
(this is the definition of being 0 in the jacobian). Then we have
for all
.
The periods of are thus
and
for . We can now correct
for it to satisfy conditions 1. 2. and having integral B-periods by taking
.
Indeed, for , the periods of
are
and
.
But by Riemann’s first bilinear relation, the expression in parentheses above vanishes for all , so
. We have thus found a meromorphic 1-form
satisfying conditions 1, 2 and 3 of lemma 1, concluding the proof of Abel’s theorem
QED
Corollary: For a compact Riemann surface of genus , the Abel-Jacobi map gives a holomorphic embedding of
into
(i.e. it is injective and has maximal rank 1 everywhere)
Proof: To see that is injective, suppose that
for two distinct points
, i.e. that
. Then by Abel’s theorem, there is a meromorphic function
having divisor
. We see this meromorphic function a holomorphic function
.
Recall that any holomorphic function between two compact Riemann surfaces is in fact a branched cover. In particular, has a degree
(the degree is the number of points in the fibers of
, which, counting with multiplicity, does not depend on the point in the image). Since
, we have
. But this would mean that
for every
. So
would be a biholomorphism, which constradicts the fact that the genus is not
.
To see that has maximal rank everywhere, we just compute its differential. Let
be a holomorphic coordinate centered at
and write
. Writing the basis of holomorphic one-forms in this chart as
, we have
so
.
Since is a basis for
, they never all vanish simultaneously, so
has maximal rank 1 everywhere.
QED
Corollary: Every smooth Riemann surface of genus one is biholomorphic to a complex torus
.
Proof: This is immediate from last corollary: if , the Abel-Jacobi map gives a biholomorphism
.
QED
This is a follow up to this post where I defined the Abel-Jacobi map of a compact Riemann surface
of genus
to its jacobian
. My first goal will be to demonstrate Abel’s theorem, which goes like this:
Theorem (Abel): For
and
a basis for
the space of holomorphic 1-forms, the divisor
is principal, i.e.
for some meromorphic function
if and only if
, i.e. iff
.
As a corollary, we will have that is in fact a smooth analytic embedding of
into its Jacobian. Before proving the theorem, we will establish in this post a reciprocity law relating the periods of a holomorphic 1-form and a meromorphic 1-form having only simple poles. These two types of 1-forms are classically called differentials of the first and third kind, respectively. Recall the notion of canonical basis for
from last post.
Proposition (First Reciprocity Law):Let
be cycles inducing a canonical basis of
and suppose
are respectively differential of the first and third kind. Let
be the poles of
. Then the following relation holds:
.
On the right hand side, the integral is taken on any path inside
which is simply connected and
is any base point.
Proof: The region is the standard polygonal representation of the Riemann surface
, which is a
-sided plane polygonal with sides labelled
etc… with the corresponding orientation. This is hard to convey without drawing pictures so you should refer to some book if you’ve never seen this, maybe a book which treats the classification of closed surfaces like Munkres. Think of the standard way to represent a torus as a quotient of the square (which is a
-polygonal when
) but with more sides.
Anyways, since is simply connected, we can choose
and define without ambiguity the function
for . Then
is a holomorphic mapping on the closure of
with
. By construction, if
and
are two points of
that are identified in
on the cycle
, then
and similarly for and
that are identified in
, we have
.
Using this, we find that
and similarly
.
Moreover, by the residue theorem, we have
.
Equating these two last equations, we obtain the first reciprocity law.
QED
A first consequence of this result is that it permits us to choose a very nice basis for the holomorphic 1-forms . This will be a consequence of the
Corollary 1: For
a non-zero holomorphic 1-form, we have
.
Proof: Take . Then
for like above. So
.
So by integrating like above, we find
so letting gives the result since
is positive.
QED
A normalised basis for
with respect to the basis
for
will be a basis such that
is
if
and
if not. Corollary 1 permits us to always choose a normalised basis. Indeed, consider the linear mapping
defined by
.
Then iff
for all
which by corollary 1 would mean
.
Recall from last post the period matrix defined by
where the columns
are the vectors
translated. By choosing a normalised basis, the period matrix becomes
for some matrix
consisting of the B-periods.
Corollary 2 (Riemann’s bilinear relations):
1) First bilinear relation: If
are two differentials of the first kind (i.e. holomorphic), then the reciprocity law tells us
.
In particular, with
a normalised basis,
,
i.e.
is a symmetric matrix.
2) Second bilinear relation: Im
, i.e. the matrix consisting of the imaginary parts of the coefficients of
is positive definite.
Proof: The first bilinear relation is immediate from the first reciprocity law. For the second, we proceed as in the proof of the first corollary and write
.
But by the first bilinear relation this last expression is equal to
.
QED
In the next couple of posts, I will write about Riemann surfaces and their Jacobian, following Griffiths and Harris. To each compact Riemann surface of genus
, we will associate a
-dimensional complex torus, which is a complex manifold of the form
for
a discrete full-rank integral lattice in
.
Recall that by Dolbeault’s theorem, on a compact complex manifold there is an isomorphism
.
So on a Riemann surface of genus
, taking
-cohomology class we find that
. Hence the dimension is
because
so since
.
Let us first consider a genus
compact Riemann surface, i.e. a torus. Then
is 1-dimensional, let
be a generator. If
and
are two points on the torus, of course the integral
is not well-defined as a complex number, it depends on the path chosen from to
on which we integrate. Instead of fixing a path to make sense of this integral, we can change the space on which it lives to account for this indeterminacy: if
are two paths from
to
, their difference will be a closed loop representing a cycle in
. The idea is to view the integral as a point in the quotient
.
More precisely, for a compact Riemann surface of genus
, we consider cycles
which give a basis for
. We will suppose that
is a canonical basis, i.e. that
and
for
, where
means the intersection product. This just means that
intersects
once positively and intersects no other cycles, think of the canonical cycles on the g-holed torus. The first half of the basis
is the
-cycles and the
-cycles are the others,
. Let’s also fix a basis
for
. Then the periods are the vectors
.
The period matrix is the matrix
having the periods as columns:
.
Note that the periods are -linearly independent so they generate a full-dimensional lattice
.
To see this, recall that since , the forms
generate
. So if we had a relation
with
, this would give
and
for all
so since the pairing is non-degenerate, this would give a relation
, which is impossible since
give a basis for the homology group.
The Jacobian of the Riemann surface is defined to be the complex torus
.
Intrinsically, is just the quotient
where we identify
with its image under the natural inclusion which takes a class of closed loops
and maps it to the functional
. This space is the right one to consider integrals on
in the sense that the vector
is not well-defined as a point of but it is if we consider is as living on
, i.e. modulo the period lattice
.
After choosing a base point , we can define a mapping
by
(mod
).
In fact, given a divisor , we can extend this map by letting
. In particular, for divisors of degree 0
,
the map is independent of the base point and
(mod
).
When considered on divisors of degree 0, this is the Abel-Jacobi map
.
The first goal will be to show Abel’s theorem, which says that the kernel of is exactly the divisors of meromorphic functions on
. Recall the short exact sequence
where is the space of all meromorphic functions on
. If we restrict this to
and the corresponding flat line bundles
, Abel’s theorem gives an injection
.
Our second goal will be to show Jacobi’s inversion theorem which essentially says that this injection is in fact an isomorphism. From this point of view, we see that the -dimensional torus
naturally parametrizes flat line bundles on
. In fact, Torelli’s theorem says that
, considered as a principally polarized abelian variety, entirely determines the Riemann surface
we started with. Jacobian varieties are thus key in the study of compact Riemann surfaces.
Let be a compact oriented Riemannian manifold of dimension
with boundary
. The aim of this note is to define the divergence and Laplacian operators on
and to clarify the validy and meaning of various formulas such as integration by parts
or Green’s formula
which are well-known to hold for domains in .
The Laplace operator acting on functions defined over is usually defined by the simple formula
.
Obviously, this definition is no good to define anything on a manifold, so we formulate it in a more geometric way as
.
In a beginning calculus course, this expression is usually understood as some kind of matrix multiplication, and we formally see that it holds. But this equation tells us what the Laplacian really is. Indeed, the divergence of a vector field is a real-valued function that at a point measures the amount of infinitesimal dilation an infinitesimal object placed at
would experience if it was to flow infinitesimally along the vector field. This is the geometric interpretation of the Laplacian. On a Riemannian manifold, we can define the gradient of a function by duality via its exterior derivative. Thus we only need a notion of divergence. From our informal discussion, it makes sense to define the divergence of a vector field
on an oriented Riemannian manifold as the infinitesimal change of measurement of volumes when the volume form is flowing along
, that is we want
where denotes the volume form associated to the metric and
the one-parameter group of diffeomorphisms associated to
. The right side of this equation is called the Lie derivative of
along
and is noted
. Since the bundle of top forms of an oriented Riemannian manifold is trivialised by
, i.e. every top form is of the form
for a unique
function
, we actually get a well-defined function
by the relation
,
given of course that there actually is a volume form, i.e. that our manifold is orientable. We thus define the Laplacian acting on smooth functions as where
is the gradient of
, caracterized by the relation
.
The only thing that remains undefined is the normal vector field . It is not hard to check that if all the transition functions of some atlas on
have positive Jacobian determinants, then their restriction to the boundary also have positive Jacobian determinants, enough so that the boundary is also orientable. Supposing without loss of generality that all charts containing a point of the boundary have their image lying in only one half-space of
, say
, we can define a notion of “inward-pointing vector” tangent to the boundary. Indeed, at a point
of the boundary, a tangent vector in
will be called “inward-pointing” if its image in any chart is contained in the half-space
. At each point of the boundary, we decompose
orthogonally with the metric, and we define the normal unit inward-pointing vector field
as the unit vector field lying in the second component of this decomposition that is everywhere inward-pointing. Note that this gives a canonical volume form on the boundary, given by
where
is the interior product.
Before we start to state and prove theorems, let’s recall Cartan’s formula which says that for every vector field and differential form
,
.
This identity tells us that because the volume form
is of course closed. This will be useful when paired with Stoke’s theorem, which says that for all forms
of degree
on
,
where is the inclusion. We can now prove the generalization of Gauss’s divergence theorem:
Theorem 1: With the notation introduced above,
for all vector fields
.
Proof: By Cartan’s formula and Stoke’s theorem, we get
.
But if is an orthonormal basis of
for
, then we see that
is equal to
because and
for all
. Thus at all points of the boundary we have the equality
.
We can now conclude from the first equation in the proof.
QED
Recall that the integration by parts formula for functions follows from Leibniz rule
and from the fundamental theorem of calculus:
.
To get the generalized version for oriented Riemannian manifolds, we need to establish the following Leibniz rule for the divergence operator:
Lemma: For all functions
and vector fields
,
.
Proof: This again follows from Cartan’ identity. Indeed, by the very definition of the interior product, we get that for any differential form
so we get
.
But by the usual Leibniz rule for the exterior derivative, this is equal to
hence
so it suffices to see that . We proceed as in theorem 1. Let
be an orthonormal basis of
at some point
. Then
which is equal to
.
This shows that
but is precisely
.
QED
Coupled with the divergence theorem, this immediately yields the
Theorem: For all vector fields
and smooth functions
, there is an integration by parts formula
.
Note that for a closed Riemannian manifold, with , this shows that
i.e. that minus the divergence operator is kind of a formal adjoint to the gradient operator. The so-called Green formulas are a simple application of integration by parts. Recall that the Laplacian of a smooth function is defined as
and that
is the inward-pointing vector field on the boundary. We will denote
by
.
Theorem: (Green formulas) For any two functions
,
and hence
.
Proof: Integrating by parts, we get
hence the first formula. The second evidently follows from the first.
QED
As an immediate application, we can show the
Proposition: Consider the equation
on a compact Riemannian manifold with boundary
subject to either Dirichlet (i.e.
) or Neumann (i.e.
) boundary conditions. Then if a smooth solution exists, it is unique in the case of Dirichlet conditions and unique up to a constant in the case of Newmann conditions.
Proof: Consider two solutions and
to the problem and consider their difference
. Then since
, it follows that
. Integrating by parts, we thus find
so if either the Dirichlet or Neumann boundary conditions are satisfied, the integral on the right vanishes and we deduce that is a constant, i.e. that
. In the case of Dirichlet boundary conditions, the function
has to vanish on the boundary hence everywhere, so
.
QED
In the same vein, if is a harmonic function on a closed Riemannian manifold (i.e.
and
), then
so the only harmonic functions on a closed Riemannian manifold are the constants.
From now on, let us denote by
and
by
. Another immediate application of integration by parts yield results about the spectrum of
:
Proposition: Let
be a closed Riemannian manifold. Then (with our definition), the eigenvalues of the Laplacian are non-positive. Moreover, eigenfunctions corresponding to different eigenvalues are orthogonal for the
inner product.
Proof: For the first assertion, suppose . Then integrating by parts gives
.
The second assertion follows from Green’s formula since if the eigenfunctions and
correpond to eigenvalues
and
, then
.
QED
This last proposition is not surprising in view of the fact that integrating by parts actually shows that is formally self-adjoint with respect to the
inner product on a closed manifold. Indeed, in this context Green’s formula reads
The first assertion of this proposition is the reason why geometers often define the Laplacian as
, in order to get a positive spectrum. Finally, from the first equation of the proof, we see that eigenvalues of (minus) the Laplacian satisfy
with is their associated eigenfunctions, which one could recognize as a renormalized Dirichlet energy functional. With this point of view, the variational min/max principle which says that the
‘st eigenvalue is given by the infimum of that functional over the functions orthogonal to the eigenfunctions associated to the first
eigenvalues makes plenty of sense. This min/max principle in turn explains the link between the first eigenvalue of the Laplacian and the important Poincaré’s inequality, which says that for some constant
, all smooth functions integrating to
(i.e. orthogonal to the constants) satisfy
,
the optimal constant being exactly , attained by the first eigenfunctions of the Laplacian.
In this post I introduce the notion of a Fredholm operator between two Hilbert spaces, following part 1 of Booss & Bleecker’s book Topology and Analysis. These operators end up providing some very nice bridges between topological ideas and functional analytic ideas. For example the Atiyah-Jänich theorem says that the space of Fredholm operators on the complex Hilbert space represents the K functor in topological K-theory. It is also at the heart of index theory (see the Atiyah-Singer index theorem), where some deep analogies between ideas coming from the study of PDEs and ideas from geometry are established. In the hope of understanding this all one day, let’s start with those Fredholm operators.
We will place ourselves in the context of a complex Hilbert space having a countable orthonormal basis. Note that all such Hilbert spaces are isometrically isomorphic to
. We will denote
the Banach algebra of bounded linear operators on
. A Fredholm operator is an operator
with closed range and finite dimensional kernel and cokernel. We will denote the set of Fredholm operators by
. The index of a Fredholm operator
is defined as
.
For example, the situation for an operator between two finite dimensional spaces, the situation is quite simple: We have
since
and
.
These operators originally appeared in the theory of integral equations. The condition of being Fredholm means that the equation has a finite dimensional space of solutions and that there are a finite number of linear relations we can impose on
to make sure that
is solvable (because
has to be in the orthogonal complement of the cokernel).
We can also phrase the index of in terms of its adjoint. Recall that by Riesz’s representation theorem, for each
we have the adjoint operator
that satisfies, for all
,
such that is an isometry. It is an easy exercice to check that
and thus
. In fact, this holds for any bounded linear operator with closed range. We can then give an alternative definition: An operator
is Fredholm if its image is closed and both
and
are finite dimensional. The index is then
. Here is another example:
Proposition: The operator
for
a finite-rank operator (ie. an operator with finite dimensional image) is Fredholm with index 0.
Proof: We first show that is Fredholm. Let
. We proceed in two steps: (1) show that
and then (2) that
. For (1), if
, then for
, we have
so . For (2), first note that
is orthogonal to
. To see this, take
and
. We get
as wanted.
Now define a linear map by
. This is well defined since
. Since
, this map is injective and (2) follows.
Consider now the following diagram:
Clearly the rows are exact. To see that it is commutative and that the columns are well-defined, it is enough to show that . to see this, let
and
. Then
so . Now since
is finite dimensional, we have
and since
, it’s index is also 0. Computing the alternating sum of the dimensions we get from the snake lemma, we finally conclude that
so that
.
QED
Fredholm operators are often encoutered in the study of PDE’s and more classically in the theory of integral equations. Since for
, the equation
is solvable for
iff
. In particular, if the index of
is 0, i.e.
, then we get the so-called Fredholm alternative : Either (1) the inhomogeneous equation
has a unique solution for every
or (2) the homogeneous equation
has
linearly independant solutions and there are
such that if
for
then
has a solution.
For example, if for
some closed interval, then it can be shown that for
,
is a Fredholm operator of index 0. It follows that the equation
has a solution if and only if is
-orthogonal to every solution
of the homogeneous adjoint equation
.
This principle can also be seen to be behind the fact that the laplace equation can be solved for
on a compact manifold iff
.
The fact that the integral equation above gives a Fredholm operator of index 0 is a consequence of the compactness of and of the
Theorem (Riesz, 1918): For any compact operator , the operator
is Fredholm with index 0.
In the next post, I will prove this after introducing compact operators and then discuss a close relation between compact and Fredholm operators given by Atkinson’s theorem.
Hilbert’s theorem 90 is the 90’th theorem in Hilbert’s Zahlbericht (meaning number report according to google translate), which is a famous report on the state of algebraic number theory at the end of the nineteenth century. It is a basic theorem shedding some light on cyclic field extensions (a Galois extension with a cyclic Galois group) and it seems to lead to some really big ideas in algebraic number theory. As a corollary of this, we directly get a parametrization of the rational points on the circle, hence of integral pythagorean triples.
Given a finite extension , we can naturally consider
as a
-vector space. We can use this point of view to define a notion of norm in field extensions. For any
, we define the norm of
with respect to this extension as
where is multiplication by
viewed as a
-linear transformation on
. It also makes sense to speak of the trace of an element with respect to a field extension and we actually get an analogous “additive” Hilbert 90 without further difficulty.
Take for example the complex numbers as an extension of the reals. If , then in the basis
the application
is represented by the matrix
so we get
. A similar argument shows that in a quadratic number field of the form
for
an integer which is not a square, we find
.
Lemma: For a simple algebraic extension , we have
where
is the constant term of the minimal polynomial of
in
.
Proof: Let be the minimal polynomial of
, say of degree
and recall that
can be seen as the field of
-polynomials of degree less than
where multiplication is done modulo
. Then
is identified with the class of
. A nice basis for this space is
. Here multiplication by
acts on these vectors as
(everything modulo
). But
so multiplication by
is represented by
.
This matrix clearly has determinant hence the assertion.
QED
Note that it also follows from the proof that . To get a better description of that norm, we need another lemma:
Lemma: For a finite extension and
, let
. Then
.
Proof: If is a basis of
as a vector space over
and
is the canonical basis of
over
, then
for
and
is a basis for
over
. Ordering this basis by
, the matrix of multiplication by
is given by the block diagonal matrix
where is the matrix representing multiplication by
in
as in the first lemma.
QED
The analogous result for the trace is that . These two lemmas let us interpret the norm and trace in a clarifying way:
Proposition: For a Galois extension and
, we have
and
.
Proof: If are the roots of the minimal polynomial
of
whose constant term we denote by
and
is of degree
over
, then since
, the preceding lemmas tell us that
.
Now, recall that acts transitively on the set of roots of
since it is irreducible. Denoting by
the stabilizer of a root
for this action, we have
and since there is only one orbit, all stabilizers are conjugate of each others so
for all root
. This shows that to each root of
there corresponds at least
elements of
. Since
, this correspondence is bijective and we have
as we wanted.
QED
We are now in a position to state and prove our main theorem, which characterizes unit norm elements in a cyclic extension:
Theorem (Hilbert 90): Let be a cyclic (Galois) extension of degree
, take
a generator for the Galois group and
. Then
if and only if
for some
.
One direction is immediate. Indeed, for any and
, we have
(where the products are all taken over ). For the other direction, we need a lemma:
Lemma: If are pairwise distinct
-automorphisms of
, then they are linearly independant over
.
Proof of the lemma: Let be a minimal linear relation among the automorphisms. Since they are pairwise distinct, there exists
such that
. Then for all
, since
, we have also that
. We obtain in this way that for all
,
and
so subtracting give us which is a non-trivial relation (since we took
so that
). But this relation is shorter than our initial relation, which we supposed to be minimal. This is a contradiction and the lemma is proven.
Proof of the theorem: Suppose and define
as
. Then it suffices to find a fixed point of
. It is easy to see that for
, we have
.
In particular, so if
, we have
for all
. In particular, the order of
is
. Since moreover
is
-linear over
, we get a representation
by
. From this point of view, it is easy to determine if
has a fixed point. Indeed, for any group
and any representation
, recall that
is the projection of onto the trivial summand of the representation. So to show that a group action has a fixed point, it suffices to show that this map is not identically
. But in our case we have
and since is a generator for the Galois group, the automorphisms
are linearly independent by the lemma so
is not identically zero because that would give us a non-trivial linear relation between them. This shows that
for some
and that
.
QED
Note that the analogous result for the trace is that in a cyclic extension, is and only if
. As a cool application of this theorem, we can give a parametrization of pythagorean triples. Indeed, consider the equation
for
not a square. Recall that the norm of
is
(this is easy to see with the finer interpretation of the norm as a product over the Galois group). So rational solutions of our equation correspond exactly to the elements of unit norm in . By Hilbert’s theorem 90 and since conjugation generates the Galois group, these elements are those of the form
.
Since multiplying and
by some rational number does not change the expression, we can take
and
as integers. Therefore, the rational solutions to the equation are
where . Letting
, we find all the rational points on the unit circle, which give us the integer pythagorean triples by the correspondence
.