These are my live-TeXed notes (reorganized according to my tastes) for the second part of the course Introduction to Riemann surfaces and Teichmuller theory by Professor Feng Luo, June 28 — July 14, at Math Science Center of Tsinghua University.
Classification of Riemann Surfaces
Surfaces (assumed to be connected and orientable) of finite type (i.e., with finitely generated fundamental group) have a complete topological classification due to Mobius ([1, Theorem 1.1]):
(Topological classification of surfaces)
- Any closed surface is homeomorphic to the connected sum of a sphere with tori, where is called the genus.
- Any compact surface is obtained from a closed surface by removing disjoint disks, where is the number of boundary components.
- Any surface is obtained from a compact surface by removing points from the interior, where is the number of punctures.
Among these, , , , are the only cases with abelian fundamental group . However, each single topological class may contain various analytic structures.
Let
be a topological surface. Denote by
the set of biholomorphism classes of Riemann surfaces that are homeomorphic to
. This is called
Riemann's moduli space.
We can restate the uniformization theorem as follows.
(Uniformization Theorem)
Let be a Riemann surface with nontrivial fundamental group. From general algebraic topology property, we know that there exists a simply connected surface and a covering map such that acts freely and discontinuously on : for any , is a homeomorphism and if and only if for some ; for any there exists a neighborhood of such that for any nontrivial . In particular, for any nontrivial , we have for any . is a Riemann surface such that is analytic. The Uniformization Theorem tells us that is either , or .
The main goal of this section is to prove the following classification of Riemann surfaces with simple topology.
To prove Theorem 3, we need the following two lemmas.
- For any , we have .
- Any continuous map can be lifted to such that , where and are the universal covering maps. Furthermore, if is a homeomorphism then is so. Obviously, if is analytic then is so, hence if then .
(a) is direct. For (b), since
is simply connected and
is a covering map, we choose
to be the lifting of
. Furthermore, suppose
is a homeomorphism with
. Taking
,
and a lift
,
, we get a lifting
similarly. Then
so
and
are both the liftings of
sending
to
. The uniqueness of the lifting ensures that
are inverses.
¡õ
Let
, then
for any
if and only if all eigenvalues of
are real.
Since
, the eigenvalues are of the form
. The condition that the eigenvalues are real implies that
. The fixed points condition implies that
, namely
, too.
¡õ
Now we are in a position to prove Theorem 3.
(Proof of Theorem 3)
(a) Since every Mobius transformation
has a fixed point in
,
cannot act on
freely.
(b) Suppose . The automorphism has no fixed points, hence and , namely . If , then by . If with , then topologically is . We claim that . If is a rational multiple of , then for , therefore , which contradicts the fact that and are generators of . If is an irrational multiple of , then is dense in the line , so that has limit points which contradicts the discontinuity. So where is a lattice. A similar argument shows that cannot have more than two generators.
(c) Suppose .
- If , with , then every fixed point of is inside by freeness. After conjugating by some , we may assume that has the form , where (by Lemma 2) or .
If , then . So by If , then , where . Then is biholomorphic to where , by composing the logarithm map, the strip-scaling map and the exponential map
- If , then since . Up to conjugation, we may assume that or (). If , then and , so cannot act discontinuously; if , then and with . If , then ; if , then cannot act discontinuously.
We claim that no two 's are not biholomorphically equivalent. If then by Riemann's theorem on removable singularities. If , suppose is a biholomorphism, then . By Lemma 1, lifts to and . Comparing the eigenvalues, we know , therefore . The claim follows.
¡õ
.
where
consists of two points
and
.
The second part of the corollary follows directly from Theorem
3 and the first part follows Theorem
3 along with the following lemma.
¡õ
If
then
if and only if
are equivalent under the action of
.
: by change of lattice basis and Remark
1.
: composing
by an automorphism of
, we may assume
. Let
be the lifting of
such that
using Lemma
1. Hence
, which implies that
and
are equivalent under the action of
. Moreover,
ensures that they are actually equivalent under
.
¡õ
By now, we have completely described Riemann's moduli spaces for all topological surfaces with abelian fundamental groups.
Quasi-conformal maps
We fix the following conventions in this section: are open in . is a - smooth function. Denote , , , and so on. The following proposition can be thought of the chain rule with respect the variables and .
Let
and
be
-smooth,
. Then
- .
- .
The derivative
of
is given by
.
By definition and Proposition
1, we have
The lemma follows.
¡õ
It is easy to see that every -linear map is given by , where satisfies and . Moreover, . If , then . It follows that the Jacobian of is equal to . So preserves orientation if and only if .
Geometrically, sends circles to ellipses. The following easy lemma describes which of the radii of a circle are sent to the axes of the corresponding ellipse.
For
(assume that
), we have
The second equality holds if and only if
, if and only if
. The first equality holds if and only if
, if and only if
.
The
dilatation of
, denoted by
, is the ratio of axes of the ellipse
, which is equal to
.
Note that will switch the major axis and the minor axis. Also, it is easy to see that and .
The
Beltrami coefficient is defined to be
. We have
and
.
Assume
preserves orientation. Define
to be the dilatation of
and
to be the Beltrami coefficient of
. Then
.
Suppose
is
-smooth, then
where
.
By definition,
The lemma follows.
¡õ
If
, then
is conformal and
. If
, then
is conformal and
.
Say
is a
quasi-conformal map if
is an orientation-preserving homeomorphism,
-smooth outside a discrete set
and there exists
such that
for any
. Denote
, then
.
The concept of quasi-conformal maps can be generalized to Riemann surfaces using local coordinates. A natural question is to find the minimum dilatation quasi-conformal maps (in given a homotopy class) between Riemann surfaces with different analytic structures. This is the content of Teichmuller theory and the main theorem is the following.
(Teichmuller Theorem)
Suppose
are compact Riemann surfaces of genus
and
is a given homeomorphism. Then there exists a unique quasi-conformal map
such that
is homotopy equivalent to
and
has the smallest dilatation among all such
. Moreover,
is a constant independent of
.
This theory is motivated by the classical Grotzsch problem.
(Grotzsch problem)
Let
be rectangles with vertices
and size
,
. Consider all
-smooth diffeomorphism
such that
. We have:
where
is the affine map.
For any
, let
be the horizontal line segment
, then
So using the triangle inequality, we obtain
Now the Jacobian
, so
by Cauchy's inequality. The equality holds if and only if
is affine:
is a constant and
is a constant, therefore
and
are constants. The theorem now follows.
¡õ
Let
be an annulus of module
(
). Then
by
, where
. So if
is a
-smooth diffeomorphism, using the same proof as in Theorem
6, we know that
and the equality holds if and only if the lift
is affine given by
.
How can we generalize the solution to the Grotzsch problem to arbitrary Riemann surfaces?
Teichmuller used quadratic differentials to solve this problem.
Quadratic differentials
A holomorphic 1-form
on
is an analytic function
such that
for any
. Similarly, a quadratic differential
on
is an analytic function
such that
for any
. More generally, for a Riemann surface
with analytic charts
, a
holomorphic 1-form on
is a collection of analytic functions
such that
, i.e.,
. Similarly, a
quadratic differential on
is a collection of analytic functions such that
.
There are no holomorphic 1-forms on
(
). All holomorphic 1-forms and quadratic differentials on
are of the forms
and
where
(
).
A holomorphic 1-form at gives an -linear map from the tangent space to . Similarly, a quadratic differential at gives a quadratic form from to . A quadratic differential provides the following geometric data.
- A horizontal direction in , namely such that . So .
- A vertical direction in , namely such that . So .
- A flat Riemannian metric near with , given by . This metric is conformal to the standard Euclidean metric, i.e., angles in are the same as angles on the Riemann surface .
- is locally isometric to the Euclidean space . Indeed, given , define , then and and .
(Teichmuller)
Suppose
is a nonzero quadratic differential on a Riemann surface
, then for any
, there exists an analytic coordinate
at
with
such that
in
is given by
where
. We call this analytic coordinate the
natural coordinate for
.
We may work locally and assume that
,
and
where
is analytic. If
, we have just proved that
where
. Now suppose that
. Let
where
is analytic and
. Our goal is to write
as
where
is analytic,
and
. Namely, we would like to solve the equation
Formally, it is equivalent to
or
We need to justify that this is well-defined. Suppose
where
. Then
Hence
is single-valued and satisfies Equation
1.
¡õ
For a nonzero quadratic differential
, we denote
.
The metric
near
is a cone with cone angle
. The cone minus the vertex has a flat metric. Suppose the cone angle is
, then the pullback metric of
via
is
. For example, the metric
is obtained by gluing three half planes along the origin.
Teichmuller Uniqueness Theorem
Let be a Riemann surface, a quadratic differential on and , where . Then there exists a new Riemann surface such that the underlying set of and are the same, and there exists a quadratic differential on such that is -quasi-conformal. Let us construct the analytic charts for as follows. Pick any point .
Type I . By Theorem 8, there exists a natural coordinate of at with and . Define a new analytic chart for by , in other words, , .
Note that if another is analytic with and such that , then , which implies that . So if and are two natural coordinates of Type I for , the the transition function between them is analytic: we know that where is a constant, therefore , so . Hence , is a well-defined quadratic differential on .
Type II . Choose a natural coordinate at such that by Theorem 8. Define a chart for by This is well-defined since the right hand side is and , so we can just choose the principal branch. Moreover, is a homeomorphism . Indeed, , so
We claim that the transition function for a Type I chart and a Type II chart is analytic. Suppose at . We may assume that . Then we find that in , so . Hence is analytic and .
The following uniqueness theorem can be regarded as the general answer to the Grotzsch problem.
(Teichmuller Uniqueness Theorem)
Let
be a compact Riemann surface with a nonzero quadratic differential
. Let
be the new Riemann surface. If
is a quasi-conformal map homotopic to the identity map, then
and the equality holds if and only if
.
Denote for simplicity.
(Key Lemma)
Suppose
is an embedded horizontal arc in
and
such that
,
are homotopic relative to
. Then the length of
is no greater than the length of
in
.
The problem occurs due to the presence of singularities (negative curvatures). To prove this Key Lemma, we need the following version of Gauss-Bonnet Theorem.
(Gauss-Bonnet)
Let
be a polyhedral surface obtained by isometric gluing of Euclidean triangles. For a vertex
of
, define its curvature
Then
.
(Proof of Key Lemma 7)
Let us minimize the length of
. We may assume that
is a piece-wise geodesic path homotopic to
such that the vertices of
are in
. Moreover, any vertex other than
and
has cone angle at least
(otherwise we can decrease the length), so
for these vertices. Now by the Gauss-Bonnet Theorem, we get
, impossible unless
.
¡õ
Let
be a
-quasi-conformal map homotopic to the identity map. Then there exists a constant
such that for any horizontal arc
in
, we have
Let
be a
-smooth homotopy between
and
. Let
. Using Key Lemma
7, since
is also horizontal in
and
, we know that
This completes the proof.
¡õ
Let
be the partial derivative with respect to the natural coordinate
in
,
. Let
be measured in
. Then
Let
be
minus all horizontal lines through
(several branches at one singular point), then
has full measure in
. For any
and
, define
be the horizontal line centered at
of length
. Then
is simple since the horizontal direction is unique outside
. We choose the arc length parametrization, namely
, so
, where
under the natural coordinate. Using Key Lemma
7, we know that
where
is a constant. But by the chain rule
we know that
So
The lemma then follows by letting
.
¡õ
(Proof of Theorem 9)
By Lemma
9 and Cauchy's inequality, we have
since
when
is compact. The theorem follows.
¡õ
Teichmuller spaces
Let us return to the study of Riemann's moduli space consisting of all Riemann surfaces homeomorphic to a fixed topological surface . We observe that it is the equivalent to study all complex structures on .
Let
Let
be a homeomorphism, then we can pull
back to another complex structure
on
. Define
Then the Riemann moduli space
.
We say two complex structures
and
are Teichmuller equivalent if there is a biholomorphism
such that
is homotopic to the identity map. The set of all Teichmuller classes is called the
Teichmuller space, denoted by
. Define
Then
is the same as
.
The group
is called the
mapping class group of
. Then
.
There is a subtle problem concerning "markings" on the Teichmuller space since we fix the underlying set .
What is the space of all congruence classes of triangles in the plane? Let
be a fixed triangle. Then
which is isomorphic to
. But the Riemann moduli space
is isomorphic to
.
Instead of fixing the underlying set , we may consider the equivalence classes of marked Riemann surfaces.
Fix a topological surface
. A
marked Riemann surface by
is a pair
where
is a homeomorphism.
and
is Teichmuller equivalent if there exists a biholomorphism
such that
are homotopic. So the Teichmuller space
is the set of all Teichmuller equivalent classes of
.
Our goal is to show that is isomorphic to the Fricke space Fricke showed that can be viewed as a subset in a -dimensional manifold , and then one can prove that is actually homeomorphic to (cf. Uniformization Theorem 14).
Hyperbolic geometry
This subsection is a crash course on the basics of hyperbolic geometry. Define a Riemannian metric on .
The distance
Since the cross ratio is invariant under a Mobius transformation, we have
where
are the intersection points of the semicircle through
with the
-axis.
Let be a compact Riemann surface of genus . The uniformization theorem tells us that . The metric is invariant under , so it induces a Riemannian metric on (thus such an is called a hyperbolic surface) and now it makes sense to talk about lengths on .
Suppose
is a closed hyperbolic surface, then for
, we have
, or equivalently,
has two distinct real eigenvalues (which are called
hyperbolic).
By the proof of Lemma
2, we know that
. If
, after conjugation, we may assume that
. Since
is compact, we can find finitely many small balls
covering
(so
's are contractible). By Lebesgue's Lemma, there exists
such that any loop of length less than
in
is in some
(hence homotopic to a point). But for
, consider the image of
, whose length is
. It is homotopically nontrivial, a contradiction.
¡õ
Uniformization theorem
We are now going to talk about the remaining two main theorems, namely the uniformization theorem and the measurable Riemann mapping theorem. Teichmuller asserted that the Teichmuller space is isomorphic to , but this result was not rigorously proved until the work of Ahlfors and Bers in 1950s. We will utilize some corollaries of the measurable Riemann mapping theorem (MRM) to prove the uniformization theorem in this section and discuss the MRM in the last section.
Let be a closed surface of genus . Let be the universal covering of , then there exists a biholomorphism such that acts freely and discontinuously on . Moreover, induces a biholomorphism on the quotient . In summary, we get a map
The image is called the Frick space of . By Remark 10 , the map is injective. So we can identify with .
(Fricke)
is a subset of the
dimensional smooth manifold
.
Let
. Let
be a discrete faithful representation and
,
. We get
After conjugation, we may assume that
, where
by Lemma
10. Suppose
. By the proof of Theorem
3, we know
in
, since
is nonabelian. But
, so
(consider
) and
(consider 0).
Now commutes with , so after conjugation we get
We may assume for . If we can solve uniquely from Equation 2, then . Indeed, we compute that
and one can show that , , , can be solved successively using the fact that , .
¡õ
By the Riemann-Roch Theorem 7, is a dimensional -vector space and has a natural -norm. Let . Denote , where . Define the Teichmuller map
Now the Teichmuller Uniqueness Theorem 9 can be restated as follows.
(Teichmuller Uniqueness Theorem)
is injective.
If
and
, then there exists a biholomorphism
such that
and
are homotopic. By the Teichmuller Uniqueness Theorem
9, we know that
therefore
. It follows from the construction of
that
.
¡õ
Combining Theorem 11 and Theorem 12, we obtain the following picture:
The next two lemmas will be proved using MRM in the next section.
is continuous and proper.
is connected.
Now the uniformization theorem will follow easily from the above two lemmas and Brouwer's famous invariance of domain theorem.
(Brouwer's Invariance of Domain)
Let
be an open set in
and
be a one-to-one continuous map into a manifold of the same dimension, then
is open in
and
is a homeomorphism.
(Uniformization Theorem)
is a homeomorphism. In particular, the Teichmuller space
and the Fricke space
are homeomorphic to
.
By the continuity in Lemma
11 and Invariance of Domain Theorem
13, we know that the image of
is open in
. By the properness in Lemma
11and the fact that the target
is a locally compact Hausdorff space, we find that
is a closed map and thus the image of
is also closed in
. But
is connected by Lemma
12, hence the image of
is the whole of
, which completes the proof.
¡õ
Measurable Riemann Mapping Theorem
As promised in the last section, we shall discuss the Measurable Riemann Mapping Theorem and utilize it to finish the proof of the uniformization theorem.
Let be a quasi-conformal map. By definition, the supremum norm of the Beltrami coefficients . The following inverse question asks about the existence of such quasi-conformal maps.
(Beltrami Equation)
For
with
, is there any quasi-conformal map
such that
?
The Measurable Riemann Mapping Theorem gives affirmative answers to this question.
(Measurable Riemann Mapping Theorem)
Let
,
, then there exists a unique quasi-conformal map
satisfying the normalization
,
and
such that
.
This deep theorem was proved by Lavrentev for -functions and by Morrey for measurable functions .
MRM can be also generalized to Riemann surfaces. From Sullivan's geometric point of view, the MRM simply says the following: place an ellipse on each tangent space on a Riemann surface, then there exists a unique quasi-conformal map sending this ellipse field to a circle field. In other words, for a pair , we need to construct a new Riemann surface satisfying the above geometric condition.
Suppose
are Riemann surfaces and
is a quasi-conformal map. Then the
Beltrami differential (of
)
is well-defined on
.
Suppose
, where
are analytic. Let
. By the chain rule, we have
Then
So
is well-defined.
¡õ
Now let be the vector space of all Beltrami differentials on . It has a natural -norm given by . We have a pairing So can be viewed as the dual of in some sense.
Looking at the universal covering and using the normal MRM Theorem 15, one can prove the following MRM on Riemann Surfaces.
(MRM on Riemann Surfaces)
Let
be a Riemann surface and
be a Beltrami differential on
with
. Then there exists a Riemann surface
and a quasi-conformal map
such that
and
is unique up to biholomorphism.
Finally, here comes the proof of Lemma 11 and Lemma 12, which in turn completes the proof of Theorem 14.
(Proof of Lemma 11)
By Remark
15,
is the composition of the continuous MRM given by Theorem
16 and the continuous map
, hence continuous. If it is not proper, suppose
satisfies
(
) and
converges to
. Then we get quasi-conformal maps
such that
. But
for some constant
as
converges to
, which contradicts the assumption that
.
¡õ
(Proof of Lemma 12)
We use MRM to prove that
is connected. Namely, we need to show that there is a continuous process from one
to another
such that every matrix representation of
is discrete and faithful. This can be done by using
.
¡õ
References
[1]Benson Farb and Dan Margalit, A Primer on Mapping Class Groups, http://www.math.utah.edu/~margalit/primer/.
[2]Lars V. Ahlfors, Lectures on Quasiconformal Mappings (University Lecture Series), American Mathematical Society, 2006.