GTM 71 Riemann Surfaces (Hershel M. Farkas, Irwin Kra (Auth.) )
GTM 71 Riemann Surfaces (Hershel M. Farkas, Irwin Kra (Auth.) )
GTM 71 Riemann Surfaces (Hershel M. Farkas, Irwin Kra (Auth.) )
Kra
Riemann Surfaces
27 Figures
With 27
%
Springer-Verlag
New York Heidelberg Berlin
Hershel M. Farkas Irwin Kra
Department of Mathematics Department of Mathematics
The Hebrew University of Jerusalem S.U.M.Y. at Stony Brook
Jerusalem Stony Brook, NY 11794
Israel USA
Editorial Board
P. R. Halmos F. W. Gehring
Managing Editor University of Michigan
Indiana University Department of Mathematics
Department of Mathematics Ann Arbor, Michigan 48104
Bloomington, Indiana 47401 USA
USA
C. C. Moore
University of California
Department of Mathematics
Berkeley, California 94720
USA
The present volume is the culmination of ten years' work separately and joint-
ly. The idea of writinfthis book began with a set of notes for a course given
by one of the authors in 1970-1971 at the Hebrew University. The notes
were refined serveral times and used as the basic content of courses given sub-
sequently by each of the authors at the State University of New York at
Stony Brook and the Hebrew University.
In this book we present the theory of Riemann surfaces and its many dif-
ferent facets. We begin from the most elementary aspects and try to bring the
reader up to the frontier of present-day research. We treat both open and
closed surfaces in this book, but our main emphasis is on the compact case.
In fact, Chapters III, V, VI, and VII deal exclusively with compact surfaces.
Chapters land II are preparatory, and Chapter IV deals with uniformization.
All works on Riemann surfaces go back to the fundamental results of Rie-
mann, Jacobi, Abel, Weierstrass, etc. Our book is no exception. In addition
to our debt to these mathematicians of a previous era, the present work has
been influenced by many contemporary mathematicians.
At the outset we record our indebtedness to our teachers Lipman Bers and
Harry Ernest Rauch, who taught us a great deal of what we know about this
subject, and who along with Lars V. Ahlfors are responsible for the modem
rebirth of the theory of Riemann surfaces. Second, we record our gratitude
to our colleagues whose theorems we have freely written down without attri-
bution. In particular, some of the material in Chapter III is the work of
Henrik H. Martens, and some of the material in Chapters V and VI ultimately
goes back to Robert D. M. Accola and Joseph Lewittes.
We thank several colleagues who have read and criticized earlier versions
of the manuscript and made many helpful suggestions: Bernard Maskit,
Henry Laufer, Uri Srebro, Albert Marden, and Frederick P. Gardiner. The
errors in the final version are, however, due only to the authors. We also
thank the secretaries who typed the various versions: Carole Alberghine and
Estella Shivers.
CHAPTER I
Riemann Surfaces 9
1.1. Definitions and Examples 9
1.2. Topology of Riemann Surfaces 13
1.3. Differential Forms 19
1.4. Integration Formulae 26
CHAPTER II
Existence Theorems 30
ILL Hilbert Space Theory—A Quick Review 30
11.2. Weyl's Lemma 31
11.3. The Hilbert Space of Square Integrable Forms 37
11.4. Harmonic Differentials 43
11.5. Meromorphic Functions and Differentials 48
CHAPTER III
Compact Riemann Surfaces 52
111.1. Intersection Theory on Compact Surfaces 52
111.2. Harmonic and Analytic Differentials on Compact Surfaces 54
111.3. Bilinear Relations 62
111.4. Divisors and the Riemann–Roch Theorem 67
111.5. Applications of the Riemann–Roch Theorem 76
111.6. Abel's Theorem and the Jacobi Inversion Problem 86
111.7 , Hyperelliptic Riemann Surfaces 93
111.8. Special Divisors on Compact Surfaces 103
111.9. Multivalued Functions 119
Contents
CHAPTER IV
Uniformization 151
IV.!. More on Harmonic Functions (A Quick Review) 151
IV.2. Subharmonic Functions and Perron's Method 156
N.3. A Classification of Riemann Surfaces 163
I 1.4. The Uniformization Theorem for Simply Connected Surfaces 179
PLS. Uniformization of Arbitrary Riemann Surfaces 188
IV.6. The Exceptional Riemann Surfaces 192
IV.7. Two Problems on Moduli 196
PJ.8. Riemannian Metrics 198
I 1.9. Discontinuous Groups and Branched Coverings 205
IV. 10. Riemann–Roch—An Alternate Approach 222
N.11. Algebraic Function Fields in One Variable 226
CHAPTER V
Automorphisms of Compact Surfaces — Elementary Theory 241
VI. Hurwitz's Theorem 241
V.2. Representations of the Automorphism Group on Spaces of
Differentials 252
V.3. Representations of Aut Mon 111 (M) 269
V.4. The Exceptional Riemann Surfaces 276
CHAPTER VI
Theta Functions 280
VI. 1. The Riemann Theta Function 280
VI.2. The Theta Functions Associated with a Riemann Surface 286
VI.3. The Theta Divisor 291
CHAPTER VII
Examples 301
VII. 1. Hyperelliptic Surfaces (Once Again) 301
VII.2. Relations among Quadratic Differentials 311
VII.3. Examples of Non-hyperelliptic Surfaces 315
VII.4. Branch Points of Hyperelliptic Surfaces as Holomorphic Functions
of the Periods 326
VII.5. Examples of Prym Differentials 329
Bibliography 330
Index 333
Commonly Used Symbols
Z integers
0 rationals
,. real numb&
n-dimensional real Euclidean spaces
C" n-dimensional complex Euclidean spaces
Re real part
Im imaginary part
H absolute value
C infinitely differentiable (function or differential)
Jr(M) linear space of holmorphic q-differentials on M
.r(M) field of meromorphic functions on M
deg degree of divisor or map
L(D) linear space of the divisor D
r(D) dim L(D) = dimension of D
Q(D) space of meromorphic abelian differentials of the divisor D
i(D) dim Q(D) = index of specialty of D
[] greatest integer in
c(D) Clifford index of D
ordpf order off at P
ni(M) fundamental group of M
1-1 1(M) first (integral) homology group of M
J(M) Jacobian variety of M
II period matrix of M
M„ integral divisors of degree n on M
Mr. {D e Mn ;r(D-1 )._ r + 1.}
W. image of M. in J(M)
W'„ image of M:; in J(M)
Z canonical divisor
K vector of Riemann constants (usually)
tx transpose of the matrix X (vectors are usually written as columns;
thus for x e Or, rx is a row vector)
CHAPTER 0
An Overview
where T e C is fixed with Im t > 0, and in and n vary over the integers. (This
involves no loss of generality, because conjugating G in the automorphism
group of C does notechange the complex structure.) If we consider the
closarparallelogram it with vertices 0, 1, 1-+ -r, and r as shown in Figure 0.1,
then we see that
1. no two points of the interior of .41 are identified under G,
2. every point of C is identified to at least one point of II (.,/i is closed), and
3. each interior point on the line a (respectively, b) is identified with a unique
point on the line a' (respectively, b').
From these considerations, it follows rather easily that C/G is J1 with the
points on the boundary identified or just a torus. (IV.6.4)
These tori already exhibit a very important phenomenon. Every r E C,
with Im t > 0, determines a unique torus and every torus is constructed as
above. Given two such points T and r', when do they determine the same
torus? This is the simplest illustration of the general problem of moduli of
Riemann surfaces. (IV.7.3; VII.4)
The most interesting Riemann surfaces have the upper half plane as
universal covering space. The holomorphic self-mappings of U are z 1—*
(az + b)/(cz + d) with (a,b,c,d) E and det[ bd] > 0. We can normalize so
that ad — bc = 1. When we do this, the condition that the mapping be
fixed point free is that la + dl 2. It turns out that for subgroups of the
group of automorphisms of U, Aut U, the concepts of discontinuity and
discreteness agree. Hence the Riemann surfaces with universal covering
space U (and these are almost all the Riemann surfaces!) are precisely U/G
for discrete fixed point free subgroups G of Aut U. In this case, it turns out
that there exists a non-Euclidean (possibly with infinitely many sides and
Im z -O
Re Z=0
Figure 0.1
4 0 An Overview
and
i(A,B)P( A;.)
is positive definite. In this case one can demonstrate the existence of multi-
plicative holomorphic functions. These functions then embed the torus as
an algebraic variety in projective space. In our case the matrix P can always
be chosen as the intersection matrix of the cycles in the canonical homology
basis; that is, [_?
The finite number of exceptions are the Weierstrass points; they carry a
lot of information about the surface M. One of the fascinating aspects of
the study of Riemann surfaces is the ability to obtain such precise information
on our objects. (III.5)
We shall see how to use the existence of these Weierstrass points in order
to conclude that Aut M is always finite for g 2. (V.1)
Another object of study which is extremely important is the Jacobi=
variety J(M). It, together with the theory of Riemann's theta function, also
is a source of much information concerning M. (111.6; 111.8; 111.1 1; VI; VII)
CHAPTER1
Riemann Surfaces
In this chapter we define and give the simplest examples of Riemann surfaces.
We derive some basic properties of Riemann surfaces and of holomorphic
maps between compact surfaces. We assume the reader is familiar with the
elementary concepts in algebraic-topology and differential-geometry needed
for the study of Riemann surfaces. To establish notation, these concepts
are reviewed. The necessary surface topology is discussed. In later chapters
we will show how the complex structure can help obtain many of the needed
results about surface topology. The chapter ends with a development of
various integration formulae.
PROOF. Iff is not constant, then f(M) is open (because f is an open mapping)
and compact (because the continuous image of a compact set is compact).
Thus f(M) is a closed subset of N (since N is Hausdorff). Since M and N are
connected, f(M) = N. LI
Remark. Since holomorphicity is a local concept, all the usual local properties
of holomorphic functions can be used. Thus, in addition to the openess
property of holomorphic mappings used above, we know (for example) that
12 I Riemann Surfaces
E = IQ e N; E (b 1(P) + 1) n}.
Pef '(Q)
The "normal form" of the mapping f given by (1.6.1) shows that E, is open
in N. We show next that it is closed. Let Q = lirnk Qk with Qk e E„. Since
there are only finitely many points in N that are the images of ramification
points in M, we may assume that b f (P) = 0 for all P e f i (Qk), each k. Thus
f -1 (Qk) consists of n distinct points. Let P k1 , , P,„, be n points in f `(Q„).
Since M is compact, for each j, there is a subsequence of {Po} that con-
verges to a limit P. We may suppose that it is the entire sequence that
converges. The points P. need not, of course, be distinct. Clearly f (13 f) = Q,
and since f(Pki) = Qk, it follows (even if the points Pi are not distinct) that
Ep Er_. (Q) (bf (p)+1). n. Thus each E,, is either all of N or empty. Let
Q 0 e N be arbitrary and let m = Ep ef -I (Q) (bf (P)+ 1). Then 0 < m < co,
and since Q 0 e Em , Z„, = N. Since Qo # 1 , E„,, i must be empty.
1.2. Topology of Riemann Surfaces 13
Definition. The number m above, will be called the degree of f (= deg f),
and we will also say that f is an m-sheeted cover of N by M (or that f has m
sheets).
Remarks
1. If f is a non-constant meromorphic function on M, then (the theorem
asserts that) f has as many zeros as poles.
2. We have used the fact that (compact) Riemann surfaces are separable, in
order to conclude that it suffices to work with sequences rather than nets.
We will establish this in 1V.5.
3. The- above considetttions have established the fact that a single non-
constant meromorphic function completely determines the complex structure
of the Riemann surface. For iff e let(M)\ C. and P e M, and n - 1 = k(P),
then a local coordinate vanishing at P is given by
(f - f(P)) lin if f(P)
and
f— li n if f(P) = co.
1.1.7. Since an analytic function (on the plane) is smooth (Cœ), every
Riemann surface is a differentiable manifold. If {U,z} is a local coordinate
on the Riemann surface M, then x = Re z, y = Im z (z = x + iy) yield
smooth local coordinates on U. In 1.3 we shall make use of the underlying
Cm-structure of M.
1.2.1. We assume that the reader has been exposed to the general notions
of surface topology, in particular to the fundamental group and the simplicial
homology groups. We thus content ourselves with a brief review of these
14 I Riernann Surfaces
ideas. In this section the word curve on M will mean a continuous map c of
the closed interval I = [0,1] into M. The point c(0) will be called the initial
point of the curve, and c(1) will be called the terminal or end point of the
curve. Furthermore since we shall be primarily interested in compact
Riemann surfaces, we shall (in general) assume that the manifold is compact,
triangulable, and orientable. (All Riemann surfaces are triangulable and
orientable.)
1.2.2. n, (M)= Fundamental Group of M. If P, Q are two points of M and
c l and c, are two curves on M with initial point P and terminal point Q, we
say that c, is homotopic to c 2 (c 1 c2) provided there is a continuous map
h: I x I —• M with the properties h(t,0) = c h(t,l) = c h(0,u) = Pand
h(1,u) = Q (for all t, u e I).
If P is now any point of M, we consider all closed curves on M which pass
through P. This is the same as all curves on M with initial and terminal
point P. We say that two such curves c 1 , c2 are equivalent whenever they are
homotopic. The set of equivalence classes of closed curves through P forms a
group in the obvious manner. This group is called the fundamental group of
M based at P. It is easy to see that the fundamental group based at P and
the fundamental group based at Q are canonically isomorphic as groups.
The fundamental group of M, n,(M), is therefore defined to be the fundamental
group of M based at P, for any P e M.
in is, however, H(M) = H. = Z./B. and we call this group the nth simplicial
homology group (with integer coefficients). (By definition H„ = MI for n > 2.)
Let us now denote by [z] the equivalence class in H„ of z e Z. We shall
say that [:], j = 1, , /3„, are a basis for H„ provided each element of H.
can be written as an integral linear combination of the [z ] and provided the
integral equation > 1 ai[zi] = 0 implies a; = O. In this case we shall call
the number /3„ the nth-Betti number of the triangulated manifold.
It is very easy to describe the groups H o and H2, and thus the numbers
/30 and /32 . In fact it is apparent that /30 = 1, and that H o is isomorphic to the
integers. As far as H2 is concerned, a little thought shows that there are
exactly two possibiliti;§. If M is compact, then H2 is isomorphic to the
integeis and ,62 = 1. If M is not Compact, then H2 is trivial and /32 = O.
The only non-trivial case to consider is H 1 and ,6 1 . We have seen in the
previous paragraph that H o and H2 are independent of the triangulation.
The same is true for H, although this is not at all apparent. One way to see
this is to recall the fact that H, is isomorphic to the abelianized fundamental
group. We shall not prove this result here. Granting the result, however, and
using the normal forms for compact surfaces to be described in 1.2.5, it will
be easy to compute H 1 (M) and hence fi, for compact surfaces M.
This is also a good place to recall the monodromy theorem which states:
Let M* be a smooth unlimited covering manifold of M and cl , e 2 two curves on
M which are homotopic. Let cf, cl he lifts of c 1 , c 2 with the saine initial point.
Then el' is homotopic to el'. In particular, the curves cr and el' must have the
same end points.
If M* is a covering manifold of M with covering map f, then a homeo-
morphism h of M* onto itself with the property that foh=f is called a
covering transformation of M*. The set of covering transformations forms
a group, which is called transitive provided that whenever f(Pf) = f(PT) there
is a covering transformation h which maps Pt onto P. For the smooth unlim-
ited case, the group of covering transformations is transitive if and only if
n i(M*) is isomorphic to a normal subgroup of t 1(M) and in this case the
group of covering transformations is isomorphic to n i (M)ht i (M*).
Let us now assume that M is a Riemann surface. Then the definition of
covering manifold shows that M 5 has a unique Riemann surface structure
on it which makes f a holomorphic map. Furthermore, the group of covering
transformations consists now of conformal self maps of M. In the converse
direction things are not quite so simple. If M is a Riemann surface and G is a
group of conformal self maps of M, it is not necessarily the case that the orbit.
space M/G is even a manifold. However, if the group G operates discon-
tinuously on M, then M1G is a manifold and can be made into a Riemann
surface such that the natural map f :M M 1G is an analytic map of Riemann
surfaces. More details about these ideas will be found in IV.5 and IV.9.
If the sides a and a -1 follow one another in the polygon, and there is at
least one other side then you can remove both sides from the symbol and the
new symbol still is the symbol of a polygon which is a topological model
for M.
The polygon's sides have been labeled and so have the vertices of the
polygon. We now wish to transform the polygon into a polygon with all
vertices identified. This is done by cutting up the polygon and pasting in a
fairly straight forward fashion. To illustrate, suppose we have a vertex Q not
identified with a vertex P, as in Figure I.1. Make a cut joining R to P and
paste back along b to obtain Figure 1.2. We note that the number of Q
vertices has been decvased by one. Continuing, we end up after a finite
numbér of steps with a triangulation with-all the vertices identified.
Figure 1.2
b, .b, 1511 ( )
■ )
a,
-
Remark. The "pasting" process is, of course, not uniquely determined by the
symbol. For example, in the case of genus 1, after joining side a' a l , we
may twist the resulting cylinder by 2ir radians before identifying b -1 with 6 1 .
The "twisted" surface is, of course, homeomorphic to the "untwisted" one.
The homeomorphism is known as a Dehn twist.
1.2.6. Euler-Poincaré. The Euler-Poincaré characteristic z of compact
surfaces of genus g is given by x = ao - a 1 + a 2 , where ak is the number of k
simplices in the triangulation. A triangulation of the normal form gives
= 2 - 2g. The Euler-Poincaré characteristic is also given by fit, - fi1 + /3 2,
where fik is the kth Betti number. Thus the computations of the Betti numbers
in 1.2.4 and 1.2.6 yield an alternate verification of the value of x.
1.2.7. As an application of the topological invariance of the Euler- Poincaré
characteristic, we establish a beautiful formula relating various topological
indices connected with a holomorphic mapping between compact surfaces.
PROOF. Let S = {f(x); x e M and b f (x)> 01. Since S is a finite set, we can
triangulate N so that every point of S is a vertex of the triangulation. Assume
that this triangulation has F faces, E edges, and V vertices. Lift this triangula-
tion to M via the mapping f. The induced triangulation of M has nF faces,
flE edges, and n V — B vertices. We now compute the Euler—Poincaré charac-
teristic of each surface in two ways:
F— E + V = 2 — 2y
nF — nE + nV — B = 2 — 2g.
From the above we obtain
1 — g n(1 — 13/2.
1.2.8. We record now (using the same notation as above) several immediate
consequences.
Corollary 3
a. If g = 0, then y = O.
b. If 1 g = y, then either n = 1 and (thus) B = 0 or g = 1 and (thus) B = O.
ax ay
.R2)] [ [
(3.1.2)
[ g(z(1) )
1.3.3. Remark. To derive (3.2.3), we have made use of the "exterior" multi-
plication of forms. This multiplication satisfies the conditions: dx A dx =
0 = dy A dy, dx A dy = —dy A dx. The product of a k-form and an /-form
is a k + 1 form provided k + I < 2 and is the zero form (still k + 1) for
k + 1 > 2.
In view of the last remark, we let Ak denote the vector space of k-forms,
we see that Ak is a miklule over A° and that Ak = tol for k > 3. Further
A = Me/VeA2
is a graded anti-commutative algebra under the obvious multiplication of
forms.
1.3.4. A 0-form can be "integrated" over 0-chains; that is, over a finite
set of points. Thus, the "integral" of the function f over the 0-cycle
ENP„ P„e M, ni Z
is
En,f(P2).
A 1-form co can be integrated over 1-chains (finite unions of paths). Thus,
if the piece-wise differentiable path c is contained in a single coordinate disc
z = x + iy, c: I M (I = unit interval [0,1]), and do is given by (3.1.1), then
ri dx dy
w = Jo ff(x(t),y(t)) Tt + g(x(t),y(t))--d7} dt.
By the transition formula for o), (3.1.2), the above integral is independent
of choice of z, and by compactness the definition can be extended to arbitrary
piece-wise differentiable paths.
Similarly, a 2-form Q can be integrated over 2-chains, D. Again, restricting
to a single coordinate disc (and Q given by (3.1.3)),
1.3.5. For C'-forms (that is, forms whose coefficients are C' functions), we
introduce the differential operator d. Define
df = fx dx + fy dy
for C' functions f. For the C' 1-form co given by (3.1.1) we have (by definition)
= fD do).
fi = i(fx +
For 2-forms, the operators a and .0 are defined as the zero operator.
Recall. The equation fi = 0 is equivalent to the Cauchy-Riemann equations
for Re f, Im f ; that is, fi = 0 if and only if f is holomorphic.
It is easy to check that the operators on forms we have defined satisfy
d =a
1.3. Differential Forms 23
*:Ak
and ** = ( — 1)k. Further, if co is given in complex notation by (3.2.4 then
*a) = — iu(z) dz + iv(z) dT. (3.7.2)
The operation * defined on 1-forms co has the following geometric inter-
pertation. Iff is a C' function and z(s) = x(s) + iy(s)is the equation of a curve
parametrized by arc-length, then the differential df has the geometric
interpretation of being (0110-c) ds, where Offe-t is the directional derivative of
f in the direction of the tangent to the curve z. In this context, *df has the
geometric interpertation of being ( Of/an) ds, where Of left is the directional
derivative of f in the direction of the normal to the curve z. (Note that
ds = Idzi in this discussion.)
Remark. The reader should check that the above definitions (for example
(3.7.1)) are all well defined in the sense that *co is indeed a 1-form (that is,
it transforms properly under change of local coordinates).
1.3.8. Our principal interest will be in 1-forms. Henceforth all differential
forms are assumed to be 1-forms unless otherwise specified. A form co is called
exact if a) = df for some C2 function f on M; a) is called co-exact if *a) is
exact (if and only if co = *df for some C2 function f). We say that co is closed
provided it is C' and dco = 0; we say co is co-closed provided *ai is closed.
24 I Riemann Surfaces
At this point observe that we have a pairing between the homology group
H„ and the group of closed n-forms of class C'. This pairing is interesting
for n = 1, and we describe it only in this case. If c is a 1-cycle and co is a closed
1-form of class C2, define <c,co> = $, co.The homology group H, is defined by
Z,/B 1 , where Z 1 is the kernel of 6 and 13 1 is the image of the 2-chains in the
one chains. The operator d on functions of class C2 and differentials of class
C2 gives rise also to subgroups of the group of closed 1-forms. In particular
the exact differentials are precisely the image of C2 functions (0-forms) in
the group of 1-forms, and the closed forms themselves are the kernel of d
operating on C2 1-forms. Hence the quotient of closed 1-forms by exact
1-forms is a group and we have for compact surfaces a nonsingular pairing
between H I and this quotient group. We shall soon see that this quotient
group is isomorphic to the space of harmonic differentials. (See 11.3.6.)
PROOF. The forward implication has already been verified (every holomorphic
differential is harmonic). For the reverse, note that if co is given by (3.2.1)
and *o.) = ico, then (3.7.2) implies that co = u dz. Since do) = 0, u is
holomorphic.
IÔD CI)
Corollary 2. Let f be C' function and w a C' 1-form on the Riemann surface
M. If either f or co has compact support, then
ffm f do)— Sim o) df = 0. (4.1.2)
PROOF. If M is not compact, then take D to be compact and have nice bound-
ary so that either f or co vanishes on bD and use (4.1.1). If M is compact
cover M by a finite number of disjoint triangles Ll, j = 1, . . . , n. Over each
triangle (4.1.1) is valid. We obtain (4.1.2) by noting that
•E fro = 0,
J=1
on ie`
(coi,c0Do = SJ' wl D A *63 2- (4.2.2)
Using obvious notational conventions, we see that
= (02,04),
as is required for a Hilbert space inner product. Further,
= ffp *WI A — f2
SD =
= ld,p A * dip + d*3
Corollary. We have
(4.4.2)
ffp (9 — LIT) --= dD (q)*dtP *
PROOF. Rewrite (4.4.1) with replacing tP. Rewrite the resulting expression
by interchanging 9 and tp. Subtract one from the other, and use the fact
that (c19,d) = (dt1/4).
(d9,*dili) = 9 4.
PROOF. The notation is of Proposition 1.4.4. Apply Proposition 1.4.3 with
a = dtP (recall that d2 = 0).
Sao ff
= JD d(pAd-).
J0 f a = D df
Also,
df = fiD df.
Thus, also
4 _2 Sap Re(if)œ) fiD df
CHAPTER II
Existence Theorems
ilh — 911, cp E F.
11.1.3. Writing f = Ph, we see that we also have the following equivalent
form of the previous
,•- ...-
Theorem. Let F be a closed subs pace of a Hilbert space II. There exists a
unique linear mapping
P:H —+ F
satisfying:
a- VII = 1,
b. P2 = P (P is a projection), and
c. ker P = F1.
ffp 9 An = 0 (2.1.1)
L = A n'
and thus the necessity of (2.1.1) is established.
(2.1.4)
(r)
p(r)
11.2. Weyl's Lemma 33
Define on C x C
02
4 co(lz — 1) if z C,
7(z.;;) =j fez
0—
0 if z = C.
Since logiz — CI is a harmonic function on CVO, y(z,C) = 0 for lz — CI <
(also for lz — > E)•
Let ti be a C function with support in D i _ 2,. Consider
d
0(z) =.'ff, ( 0(1C -- zi)ii(C) 2i , ZE C. (2.1.5)
Thus
dc A dZ
ffic- zi e/2 CO(IC I)'40 —2i
= ce(z) +
34 II Existence Theorems
We have already shown that da exists (tp is C). Thus we may compute
formally. Now 01/0.7 can be computed formally without any difficulty (since
differentiating under the integral sign leads to a Lebesgue integrable function).
First
a a ,
, 1
IC — 2 z'
and thus
aci 1 rt . Ii(c) dc dr
(2.1.8)
1.11C — zI<e12 — Z — 2i
That
a2ct
4 = —/1 (2.1.9)
OzOz
will follow from
PROOF. Let e > 0 be chosen so that the closed ball of radius e about z is
contained in B. Let B, be the complement in B of this ball. Start with Stokes'
theorem
= 5.1.B. :c (cu(C)z )4 cg,
11.2. Weyl's Lemma 35
and obtain
2w
ff.., _ ,
z d-c•A = f B r, — Z —
4 0
L47. ze ie) dO.
— Z -2i
(2.3.1)
=— ff apiez dt- A dr
= -
1
- p(z)
4n c C - z -2i 4
by (2.2.1).
We now use condition (2.1.1) with ri = tfr (of the previous construction
via (2.1.5)). Thus we obtain
dz riT
0 = f fD 9 = -HD 9(z)p(z)
-2i
-I- If
D
9(z)
dz
aT,aZ —21
df
. ,
36 Il Existence Theorems
where (40 213/67.-- ôz) is given by (2.1.7). Hence for every C function p with
support in D, we have by Fubini's theorem
(In the above D(z) is just another symbol for D, and the (z) is supposed to
remind the reader that we are integrating with respect to z.) It thus follows
(because C`c functions of compact support are dense in the L. 2 functions) that
d:A d
= (o(), a.e. L'; E D. (2.3.3)
—2i
Clearly the left-hand side is C' (in ;), and thus the proof is complete.
has positive measure. Hence we may assume that one of these sets has
interior. If we now choose it to be non-negative and to have support in this
set, then we obtain a contradiction to (2.3.2) as in the arguments that estab-
lished sufficiency for a C2 function cp. Thus it could have been assumed to
have support in an arbitrarily small set to begin with, simplifying slightly
the reasoning at the beginning of this paragraph.
EXERCISE
Prove the following alternate form of Weyl's lemma: Let (f) be a measurable square
integrable function on the unit disk D. The function cp is holomorphic if and only if
This establishes necessity of (2.3.4). (The fact that tp is not defined on SD should
not cause any trouble.)
11.3. The Hilbert Space of Square Integrable Forms 37
(2) For sufficiency, first assume q, is C. Use that for every t; with compact support
(2.3.5) holds, and thus
— JD qv , az A = (m i d.: A fi O.
11.2.4. Exercise
Let f e L2([0,1]). Show that f equals almost everywhere a constant if and only if
j: f(x)glx)dx = 0
= d( Œ-T) —L ai = 0.
Thus sE E". Conversely, we have starting from the second equality
da A f = 0, all smooth f on M with compact support.
We let
H = n(E*)1,
and obtain the following orthogonal decomposition
L2(M) = E E* H. (3.2.1)
Note that from Proposition 11.3.2 we deduce that E and E* are orthog-
onal subspaces. It then follows that the direct sum E E* is closed and
thus also a Hilbert subspace of L2(M). By orthogonal decomposition
(Theorem MU), we therefore certainly have LAM) = (E e E*) (E e Ell,
and all we need to establish (3.2.1) is to verify the easy identity (E E*)i =
(E*)1. Comparing (3.2.1) with (3.1.1) we see that
= E* e H
= E H.
11.3.3. Let c be a simple closed curve on M. Cover c by a finite number
of coordinate disks and obtain a region S2 containing c. We call this region
0, a strip around c. By choosing Q sufficiently small, we may assume that
it is an annulus and that fl\e consists of two annuli ST, f2 +. We orient c
11.3. The Hilbert Space of Square Integrable Forms 39
0.
df on Q\c,
11c = 0 on (M \f2) u c.
Figure 11.5
(c0(c) = Cf A qc — CC df = df a
11.3.5. The most important result about OM) is contained in the following
(3.5.3)
PA n-
By Weyl's lemma, p is harmonic and hence C t . Applying this result to
*co (which also belongs to El n E*I) we see that q is C t . Hence co is of class
C t . Proposition 11.3.2 now yields that a) is closed and co-closed (that is,
harmonic).
Remark. The space L 2(M) can best be represented by the "three" dimen-
sional "orthogonal" diagram given in Figure 11.6.
E(exact)
E* (co-exact)
co-closed _
(harmonic)H
Figure 11.6
11.3. The Hilbert Space of Square Integrable Forms 41
Corollary
a. The L 2(M) closure of the square integrable closed (respectively, co-closed)
differentials is E H (E* H).
b. The square integrable smooth differentials are dense in L 2 (M).
b'. The smooth differentials with compact support are dense in L 2(M).
The verification of (a) and (b) is at this stage trivial. For example, if co is closed,
then co e .E*I =EeH. Conversely, if co e E H, then co = co, + (1) 2 with
E E and co2 E H. Thus, co2 is closed (it is harmonic) and co, = lim„ dl,,
with f„ smooth of compact support. We have shown that co is the limit of
closed differentials. "'" —
We cannot at this point establish (13'). We need to know that
M = U D.,
where
D„ is open,
D„ c D„ +1 ,
and
Cl D,, is compact
(a fact that will follow from 1V.5). Having such an "exhaustion" of M, we
construct a sequence of smooth functions {f„} on M with
0 f„ 1,
f„ = 1 on D„,
(support of f,,) c D„ 1 .
It clearly suffices to show that every E H can be approximated in L2(M)
by smooth forms with compact support. Now f„a E L 2(M) is smooth and
has compact support. By the Lebesgue dominated convergence theorem,
Figure 11.7
HA. Harmonic Differentials 43
L. nc = Q—P
lim f(Q) — lim f(Q) = 1
-
(here, of course, limQ_,,- means you approach P through C2'). This remark
together with Theorem 11.3.7 establishes the following
Theorem. If the Riemanrsurface M carries a-closed curve that does not sep-
arate, then there exists a closed non-exact differential in L 2(M) and therefore
a non-zero harmonic differential.
+ for Izi a,
h(z) = a
0 for izi > a.
In this convention the points in M\D are denoted by Uzi 11.)
We define another function 0 on M by setting
0(z) = h(z) for a/2,
and requiring 0 to be smooth in {Izi a}.
44 11 Existence Theorems
(d0,d9) = f fu p d p. (4.1.2)
= A *A
closed upper half plane such that U SD goes into a segment of the real axis.
By the reflection principle z is a local coordinate at P E M. Note that M is a
compact Riemann surface of genus 2g ± n — 1, and that there exists on M
an anti-conformal involution j such that j(D) = D' and j(P) = P for all
P E (5D. From now on we may forget about Mo and think of D as a domain
on its double M. (The above discussion is not strictly necessary for what
follows. It was introduced for its own sake.) We consider now the following
problem:
Fix a C2 function cp o defined on a neighborhood of Cl D. Among all functions
cp, C2 on a neighborhood of Cl D, find (if it exists) one u that has the same
boundary values (on SD) as cpo and minimizes 11411 0 over this class.
Assume that u is a harmonic function with the same boundary values as
(p o . We c. , ;npute for arbitrary (p with the same boundary values;
11d(o11 2 = (d(cp — u) + du, d(cp — u) + du)
= — 01 2 + 11(142 + 2 Re(d(q) — u),du).
Now use Proposition 1.4.4 to conclude that (d(cp — u),du) = 0 (since
— u = 0 on (5D and du = 0 on D). Thus
11411 2 = 11d((p — u)11 2 + I1d u 112 Ildu112.
Thus our problem has a unique solution—if we can find a harmonic
function with the same boundary values as (po . We shall in 1V.3 solve this
problem by Perron's method. Here we outline how the decomposition of
L2(D) given by (3.2.1) can be used to solve our problem. Finding a function
for which a given non-negative function on L 2(D) achieves a minimum is
known as the Dirichlet Principle.
Consider the function 9 0 . Since dcpo is exact (and thus closed), dcp o e
E (1) H. Now let co be the orthogonal projection of dcp onto H. We have
already seen that co is exact and harmonic. Thus co = du, for some harmonic
function u on D. Now d(u — cp 0) e E. Thus
(d(u — (p0),c() = 0, all a E H.
Let up,p2 be the function produced by Theorem 11.4.3. By letting a run
over the set of differentials
{dup,p2; P 1 E D', P2 E D'1,
one can show that u — q,,, is C 2 on a neighborhood of Cl D and that u —q, 0 =
0 on SD.
Theorem
a. Let P E M and let z be a local coordinate on M vanishing at P. For every
integer n > 1, there exists a meromorphic differential on M which is holo-
morphic on MVP} and with singularity 1/z"+ at P.
b. Give* two distinct pobtts P, and,P 2 on M and local coordinates z; vanishing
at pi, j = 1, 2, there exists a meromorphic differential a), holomorphic on
MVP,,P,}, with singularity 1/: at P 1 and singularity —1/z 2 at P2.
f(z) = E
n= N
50 11 Existence Theorems
Set f = co1 /co2 . Note that f has a pole at P 1 and a zero at P3.
1
E resp co = E f64i CO, (5.4.2)
Pet,' 2rti .1=
I
11.5. Meromorphic Functions and Differentials 51
EXERCISE
Using only formal manipulations of power series show that the residue of a meromorphic
abelian differential is well defined.
EXERCISE
Give an alternate proof of Proposition 1.1.6 in the special case that N=Cv too).
First reduce to showing that it suffices to establish that f has as many poles as zeros.
Then relate ordp f to resp(dflf).
This is one of the two most important chapters of this book. In it, we prove.
on the existence theorems of the previous chapter) the three most (based
important theorems concerning compact Riemann surfaces: the Riemann-
Roch theorem, Abel's theorem, and the Jacobi inversion theorem. Many
applications of these theorems are obtained; and the simplest compact
Riemann surfaces, the hyperelliptic ones, are discussed in great detail.
for all closed differentials 1. Since every cycle c on M is a finite sum of cycles
corresponding to simple closed curves, we conclude that to each such c,
we can associate a real closed differential tic with compact support such that
(1.1.1) holds.
Let a and b be two cycles on the Riemann surface M. We define the
intersection number of a and b by
a' b= S$M A tlb= ('1., —* %). (1.1.2)
Proposition. The intersection number is well defined and satisfies the following
properties (here a, b, c are cycles on M):
. -
a • b depends only on the homology classes of a and b, (1.1.3)
a • = —b a, (1.1.4)
(a+b).c=a•c+b-c, (1.1.5)
and
a •b (1.1.6)
Furthermore, a • b "counts" the number of times a intersects b.
PROOF. To show that (1.1.2) is well defined, choose and ti;, also satisfying
(1.1.1). We must verify
A
Note that while ti„ and are only closed, their difference pi', — = df,
where f is a C function with compact support. Thus it suffices to show
=(nb,*n.)= fa llb•
But Sa th contributes + 1 for each "intersection" of a with b. (This can be
verified as in 11.3.7 using Figure III.1.)
111.1.2. We now consider a compact Riemann surface of genus g 0, and
represent this surface by its symbol
aa' (genus 0),
fi
i=t
(t • bia; (genus g 1).
54 Ill Compact Riemann Surfaces
The sides of the polygon corresponding to the symbol give a basis for the
homology, H, = 11 1( M ), of M. Assume now that g 1. It is easy to check
that this basis has the following intersection properties
0 j k
• bk = CI ik = =
a; • ak = 0 = bi bk .
We can represent all this information in an intersection matrix J. This J is a
2g x 2g matrix of integers. If we label
= ai, j = 1, , g and = bJ_ g, j = g + 1, , 2g, (1.2.1)
then the ( j,k)-entry of J is the intersection number X; • Nk. Thus J is of the
form
0 11
[- ï OT
where 0 is the g x g zero matrix and I is the g x g identity matrix.
Any basis {N I, ... ,N 29) of H 1 with intersection matrix J will be called a
canonical homology basis for M. Given a canonical homology basis we can
use (1.2.1) to define the "a" and "b" curves. Note that we do not claim that
these curves come from a polygon in normal form.
U
N, œ ' • »f.„, '
If dim H > 2g, then cf) has a non-trivial kernel; that is, there exists an a E H,
all of whose periods are zero. Such an a must be the differential of a harmonic
function. Since there are no non-constant harmonic functions on a compact
surface, dim H 2g. The above argument also establishes the injectivity of
0. It remains to verify surjectivity. As we saw in the previous chapter,
(Theorem 11.3.7) it suffices to find a closed differential with period 1 over
a cycle j and period 0 over cycles Nk, k j.
Let
= ?hp j=1,...,g,
Œj = j = g + 1, . . . , 2g.
Then we see (by(1.1.1)) that for k = 1, . . . , g,
L a; = _j'Ç nj Aqak
tak • bi = bkj , j = 1, , g,
= ff„ =
—(ak • ni- g) = 0, = g + 1, , 2g,
and
bk • b; = 0, j = 1, . , g,
f j =g+1,...,2g.
We summarize the above information in
sak = 101, k = j, k = j — g,
, otherwise, fbk cc i = {01 ise.
, otherw
56 111 Compact Riemann Surfaces
In other words,
ftti, cci = bg„, j, k = 1, . .. , 2g. (2.1.1) 0
111.2.2. We have seen that the 2g x 2g matrix with (k, j)-entry a, is the
identity. We note that
ak A j = 1, , g,
Li = ffm Mk A nKi =
iL CCk A /j- g, = g + 1, . . . , 2g.
From this it is evident that the matrix whose (k, j)-entry is JIM; cc.; is of the
form [I fl = J. We conclude that
— *c()) = Xj•
We will investigate a companion matrix F whose (k,j)-entry is (a,,a;) =
ilk( ak A *at We note immediately that F is a real matrix. Further from 1.4,
ffm 0A 6- =,i[faf of
61
6—$
-
61
ofai 6]. (2.3.1)
=t f of ai bi
a—fbi ofa i •
11011 2= i=
[Jrdi Ofbi * 0 — bi
of * 0]
•
(2.3.5)
11002 = ff„ 0 A *0
1 OA df AO = ffa d(frh =
= [fa + fb j g + + fb;i fti
= ,
Let Zo E be arbitrary, then for z E
f(z)= f :00.
58 III Compact Riemann Surfaces
Letting z and z' denote two equivalent points on the sides ai and al of &It
we see that
k 1, . , g,
= fb,, *cci
{ —5 *ce — f *2
- i= AZ1,- g
k = g + 1, , 2g.
We show next that the real and symmetric matrix r is positive definite
(r > 0). Let
2g 2g
= E ift etik with k E C, E Rok i2 >0.
k=1 lo= 1
Recall that the differentials ak are real and that 11011 0 O. Thus, by (2.3.5),
g [ J
, 2g 2g 20 2g
0 E
J-1
E
k= 1
kCX/c S b,
'1= 1
Zi*cti E
k=1
scock fai 1=1E Zi*at]
Ski f.j *
2g g
= E
k.1=1
E
.1= 1
.11 fbj * at — Celt C41]
20
7"
= k,1=
E1 kl•
r = r —23
-1 i
k = 1, , g,
f Wk, = g + 1 , . . . , 2 9- (2.6.1)
( — il, —A 3).
and
ffh, A U= (3.0.1)
e f, — fb, e L rd.
g
fa= E (3.0.2)
0=
[Se,
e e (3.1.1)
obtain
(3.1.2)
Of course, (3.1.2) is just another way of saying that the matrix fl introduced
in 111.2.8 is symmetric.
1113.2. Next we let 0 = C; and a =L. We note that (df = Ci on Ji) by
Stokes' theorem (Formula (3.0.1))
55„, = IL
/zit Cj A Zk •
We conclude that
k bk
= —2i(Im nik)
(as usual it is the ( j,k)-entry of the matrix II). In particular,
1m 7r1; > 0.
Applying the same argument to
O= E Ckck, Ck E C,
k= 1
we conclude that
Im 11 > 0.
These facts have already been established in 111.2.
Remark. The above observation (plus the fact that dim Ye = g, where
Ye = dri(M)) can be used to give an alternate proof of the theorem that
there is a basis for Ye dual to a specific canonical homology basis. (The
reader is invited to do so!)
1113.4. We shall now adopt the following terminology: Recall that
meromorphic one-forms are called abelian differentials. The abelian differ-
entials which are holomorphic will be called of the first kind; while the
111 Compact Riemann Surfaces
Remark. Parts (a) of the proposition and its corollary have previously been
established (Proposition 111.2.8).
111.3.5.Let us consider abelian differentials of the third kind on M.
Choose two points P and Q on M. It involves no loss of generality to assume
that the canonical homology basis has representatives that do not contain
the points P and Q. Let us consider a differential r regular, (= holomorphic)
on M\{P,Q}, with
ordp t = — 1 = orda r,
(3.5.1)
r esp t = 1, res T = — 1.
Let c be a closed curve on M. It is, of course, no longer true that fc r depends
only on the homology class of the curve c. However (assuming P, Q are
not on the curves in question) if c and c' are homologous, then there is an
integer n such that
f t —f i= 27rin.
e
(3.5.2)
L ai fTPQ = fb , T PQ'
The line integral around the boundary of .ift can be evaluated by the residue
theorem, since
f(z) =
with the path of integration staying inside It. Thus we obtain
Similarly,
2ni sQ = fbj — ,E_ n.0 copQ,
where nil is the UM-entry of the period matrix H.
111.3.7. We treat now the case where
0= = Tits
(P, Q, R, S are all interior points of di). Here we cannot assert that 0 = df
on ilft. To get around this little obstruction, we cut it by joining a point 0 on
bdi to P by one curve and to Q by another curve. We obtain this way a simply
connected region 4 ' (see Figure 111.2).
In 4', 0 = df where
f(z) = 5:0 0. )
b,
Figure 111.2. Illustration for genus 2.
111.4. Divisors and the Riemann—Roch Theorem 67
a
Assume that is normalized so that it has zero periods over the cycles
a ,, a. Let
==E 40i)dz at P. (3.8.1)
1=0
Then, as before,
21 = PP - • IT, (4.1.1)
with P E M, x i e Z. We can write the divisor 21 as
21 = Pz(P) , ( 4.1 .1a)
PEA!
( ):X*(M)* Div(M)
from the multiplicative group of the field X(M) into the subgroup of divisors
of degree zero (see Proposition 1.1.6). A divisor in the image of ( ) is
called principal. The group of divisors modulo principal divisors is known
as the divisor class group. It is quite clear that the homomorphism, deg,
factors through to the divisor class group. The (normal) subgroup of principal
divisors introduces an equivalence relation on Div(M). Two divisors 91, 23
are called equivalent (21 $) provided that 21/23 is principal. If f E ir(110C,
then
fi(cc)=
PM
fl pmax{—ordp f,0)
111.4. Divisors and the Riemann-.Roch Theorem 69
defines the divisor of zeros of (or zero divisor of) f. Both divisors have the
same degree as the function f, and since
j:1(0)
(f)= ,
f 1 (x)
they are equivalent. More generally, for c E C,
f 1 (c),= ( f — c): 1 (0).
Thus the divisor class of f '(c) is independent of ce Cu ool. (It will be
clear from the context when f `(c) stands for a divisor or just for the under-
lying point set.)
One more remark about equivalent divisors: Let fi and f2 be mero-
morphic non-constant functions on M. Assume that fl = A o f2 for some
Miibius transformation A. Then f t (x)= f ' (A '(c)). Thus f '(oc)
PEM
Theorem. For 91E DiV(M), r(91) and 1(21)depend only on the divisor class of 91.
Furthermore, if 0 to is any abelian differential, then i(21) = r(21(to) 1 ) •
PROOF. Let 91 1 be equivalent to 912 (that is, 211 912-1 is a principal divisor
or 91,912-1 = (f) for some 0 f e AIM)). Then the mapping
L(912) a hf e L(91 1 )
establishes a C-linear isomorphism, proving that
r(91 1 ) = / ( 912). (4.6.1)
Next, the mapping
12(91) —
co e L(9.1(o.))-1)
111.4. Divisors and the Riemann-Roch Theorem 71
deg 91 = E n1 > O.
.1=
Note that C L(91 -1 ). Furthermore, if f E L(91 - 1 ), then f is regular on
MVP2 , ,P„,} and f has at worst a pole of order ni at P. Choosing a local
coordinate zi vanishing at Pi, we see that for such an f, the Laurent series
expansion at P is of the form
have zero "a" periods. We can easily compute the dimension of 00(91' -
Recall the abelian differentials 411 introduced in 111.3.8. For P = Pi and
2 < n < ni + 1, 1(; ) E 00(91' 1 ) • Thus
w =E dikz.,)dz i with d _ 1 =
by setting
(CO (d1,-2, • • • 41, — n, 1422— 2, • • • 42, — n2 — 11 • • • ,d„,, - 2> • • • ,d,„, — n„,— 1)•
If co E Kernel S, then w is an abelian differential of the first kind with zero
"a" periods, and hence co = 0 by Proposition 111.3.3. Hence ,§ is injective,
dim 120(91— 1 ) = deg 21, and every c0 E Q0(91") can be written uniquely as
—2
w=E E dikr(p. (4.8.2)
j= 1 k=nj1
(since each "b" curve imposes precisely one linear condition). Using the
"classical" linear algebra equation
r(91 - 1 ) = dim Image d + 1, (4.8.4)
we obtain from (4.8.3), the Riemann inequality,
r(9.1 -1 ) deg 21 — g + 1.
To obtain the Riemann-Roch equality we must evaluate dim Image d. We
use bilinear relation (3.8.2), to observe that the differential w of (4.8,2) has
zero bi period if and only if
m —2
276. E E 2(P. = 0,
) (4.8.5)
j= 1 k= —ni—i k + 1 dika
where {C I, ... ,C,} is a basis for the holomorphic differentials of the first
kind dual to our canonical homology basis and the power series expansion
of CI in terms of z; is given (in analogy to (3.8.1)) by
= E ce1)(p.01)dz, at Pr
s=0
111.4. Divisors and the Riemann—Roch Theorem 73
We recognize from (4.8.5) that the image of d is the kernel of a certain operator
from Cdeg to C. Consider the operator
T:(2(1) deg 9I
E dz.; at
k= 0
Represent the linear operator T with respect to the basis {Ç 1 , ,C9} of Q(1)
and the standard basis Cde" as the matrix
,
ceon(P i) 42)(P1) • %(A 13 1)
ag 1 (P2) ar(P2)
Œg )(P.) ceg)(P.)
g=0 q 0 1 - 2q
q>0 0
g=1 all q 1
g> 1 q<0 0
q=0 1
q =1
q>1 (2q - 1)(g - 1)
EXERCISE
Establish the above proposition for g = 0 and 1 without the use of the Riemann-Roch
theorem. (For g = O. write any co e i(M) as w = f de with f e Thus f is a
rational function. Describe the singularities of f.)
such that there does not exist a function f e f(M) holomorphic on MVP}
with a pole of order ni at P.
Remarks
1. The numbers appearing in the list (5.3.1) are called the "gaps" at P. Their
complement in the positive integers are called the "non-gaps". The
"non-gaps" clearly form an additive semi-group. There are precisely g
"non-gaps" in {2, ... ,2g} with 2g always a "non-gap." These are the
first g "non-gaps" in the semi-group of "non-gaps."
2. The Weierstrass "gap" theorem trivially holds for g = O. Since on the
sphere there is always a function with one (simple) pole, there are no
"gaps".
3. The Weierstrass "gap" theorem is a special case of a more general theorem
to be stated and proven in the next section.
111.5.4. We stay with the compact Riemann surface M of positive genus
g. Let
P,, P2, P3 , • • •
111.5. Applications of the Riemann-Roch Theorem 79
111.5.7. Proposition. If a i > 2, then for some j with 0 <j < g, we have
a; + Œg _; > 2g.
PROOF. We assume that ai + = 2g for all j with 0 < j < g. For q E
let [q] be the greatest integer 5. q. Then x i , 2;, . . . , [2g/;]; are "non-gaps"
<2g. If; > 2, then the above accounts for at most ig < g "non-gaps",
and there must be another one 5.2g. Let a be the first "non-gap" not
appearing in our previous enumeration. For some integer r, 1 r
[2g/a i ] < g, we must have
ra, < a < (r + 1)a i .
Thus we have "non-gaps"
11, 2 2 = 2a1, 1 = cc,
and by our assumption
= 2g — ra i , _ 0. 4 = 2g — a
(note that if r + 1 = g, then a = 2g and the last equality reads ao = 0—which
can be added consistently to our data). These are all the "non-gaps" which
are _ " and 52g. It follows that
a l + ag _ (, + = a, + 2g — a = 2g — (a — a i) > 2g — ra i =
It therefore follows that there is a "non-gap" < 2g, greater than ag _ , and not
in the list a9-19 • • • 9 a9-(9.+1)• This is an obvious contradiction.
Corollary. We have
g—1
E ce; g (g — 1 ),
I
with equality if and only if a, = 2.
PROOF. From Proposition 111.5.5, 2 ai _>.: 2g(g — 1). Furthermore if
a l = 2, then we have equality in the above by Proposition 111.5.6. If x i > 2,
we must have strict inequality by Proposition 111.5.7.
111.5. Applications of the Riemann-Roch Theorem 81
By induction, we can now construct the basis satisfying (5.8.1). The basis
adapted to z is (of course) not unique. We can make it unique as follows:
Let p; = ordz cp; . Consider the Taylor series expansion of (pi at z (in terms
of C)
= k=0
E akX z)k.
We may and hereafter do require that
1, k = j,
a- 10 k
where j, k = 1, . , n.
Then
ordz P = r(:).
PROOF. It is easy to see that a change of basis will lead to a non-zero constant
multiple of O. Hence we may assume, whenever necessary, that the basis
used is adapted to the point z. Let us abbreviate equation (5.8.3) by
0(z) = det[(p i(z),
[(Pi (p.
= f det
f(p't- + f'(p 1 f(p,,+ f 4.
= f det
[.fi7. 1 .10'n
remove from each column the appropriate multiple of (pi leaving us with the
previous expression equal to
•"
f(p'.
det ftp7 +2f'tp'i ••' fig + 21"0.
.fiPT -11 + ' • + — ilf (" -2),P'i • ' • f(P(- 1) -4- • ' • + — t " --2) (1):.
• •• (o.
- • • (p'.
We now repeat the same procedure to remove from each column the appro-
priate multiple of cp'j . This clearly terminates with the preceding equal to
det[cp„
We now turn our attention to the proof of the proposition which shall be
by induction on n. Clearly the result is true for n = 1. Let us now assume that
the proposition is true for n = k. Explicity we are thus assuming that
where pj = ord., pj. Consider now det[(Pi, ,(Pk+1]. It is clear from the
preceding remarks that
det[Ti, • • • ,49k4-1] = det[ 1 ,02/T1, • • •
Now the right-hand side is simply cp1+ 1 det[(T 2 /(p 1 )', ,frpk+ i ftp i r]. The
induction hypothesis now gives that
ordz {cpki + 1 det[(T2/TIY .•.,(4'k+ l(pi )'i}
,
k+1
= (k + 1)/./ 1 + E — 1) — 2)/
j=2
k+1
= 11 1 + E
1=2
j +
PROOF. By the hypothesis ord. cp) = pi. It follows from Corollary 1 that
T(z)= (jm; — j + 1) = o for an open dense set. Since (as we have pre-
viously remarked) pi j — 1 we have p = j — 1 for each j, on this open
dense set.
111.5.9. The considerations of the last paragraph apply (of course) to the
space YE(AI) of holomorphic q-differentials (q 1) on a compact Riemann
surface of genus g 1. A point P E M will be called a q-Weierstrass point
provided its weight with respect to yeq(m) is positive. A 1-Weierstrass point
is called simply a Weierstrass point or a classical Weierstrass point. It is clear
from the ideas in the previous paragraph that we have the following
a. i(P9)> 0, or
b. r(P 9) 2 (that is, at least one of the integers 2, ... , g is not a "gap").
111.5.11. We now finish the study of classical Weierstrass points (the case
q = 1).
Theorem. For g > 2, the weight of a point with respect to the holomorphic
abelian differentials is g(g — 1)/2. This bound is attained only for a point P
where the "non-gap" sequence begins with 2.
The above, of course, gives a trivial estimate on -OP). We need a better one.
Let 2 < cc i < Œ2 < • • • <; = 2g be the first g "non-gaps" at P. Then let 1 = n i
<n2 <<ng < 2g be the g-"gaps" at P. (That is, the sequence of nis is
the complement in {1, ... ,2g} of the sequence of xi's.) Then (recall 111.5.8)
zo g
r(P) = (ni — j) = j— ot; - E;
j=1 j=i j=1 j=1
2g - I g- I 3g
=E j —
j=1
(g — 1) — g(g — 1)
j=g+1
= g(g — 1)/2,
PROOF. The first inequality follows from (5.1 1.1) and the fact that the max-
imum weight of a Weierstrass point is g(g 1)/2 > O. "hie second from the
—
fact that the minimum weight of a Weierstrass point is 1 (so called simple
Weierstrass points). 0
Remark. The first equality is attained if and only if at every Weierstrass
point the "gap" sequence is 1, 3, ... , 2g — 1. These are the hyperelliptic sur-
faces to be studied in 111.7. The second equality is attained if and only if the
"gap" sequence at each Weierstrass point is 1, 2, . . , g 1, g + 1. Existence
—
S
Qk Ci jk , j, k = 1, 2, . , g.
We have seen in 111.2, that the matrix H with entries
= fbk Ci, j, k = 1, . , g,
111.6. Abel's Theorem and the Jacobi inversion Problem 87
90ln+ t) 49 (Mn) D • • D i)
We can also obtain a map that does not depend on the base point Po.
Let Div(°)(M) denote the divisors of degree zero of M; define
9:Div(M) J(M)
by setting
49 (D) = E
1=1
(p(P)— E q(Q)
i=1
for
D Pi • • Pr/Qi • • Q..
It is clear that if r = s, cp(D)is independent of the base point; that is, the map
cp:Div1°) (M) J(M)
J=1
œi = E 13P. 1 ).
E
J=1
Without loss of generality, we may assume that none of the points Pi, Q.
lie on the curves representing the canonical homology basis. Recall the
normalized abelian differentials tiv of the third kind introduced in 111.3.
Observe that df/f is an abelian differential of the third kind with simple
poles and
df
resp — = ordp f, for all P E m.
Thus
df
f CE E
.1=
Aitzpo)
(where P o is not any of the P. or Q. nor on the curves ai, bi) is an abelian
differential of the first kind. Hence, we can choose constants ci, j = 1, . . . , g,
111.6. Abel's Theorem and the Jacobi Inversion Problem 89
such that
df
J
L
j= 1 j=
E QiP0
1= 1
E q.
It follows that
df
Jaj
7 .=
and by the bilinear relations for the normalized differentials rpQ ,
df Pi
= 2nii( E E fli fQ")
J=1 p. E
j=1
where mi, ni are integers. It follows that the /th component of co(D) is
CP.; o c o, .• 1 df 1
j= 1
2i JPo
j= 1
JP0 ‘I = Jbl 7 j= 1 Cinil
= ni
J= 1
and thus q(D) = 0 mod(L(M)). We have used Po as the base point for cp.
Recall that since D is of degree 0, (MD) is independent of Po .
To prove the converse, we let D be given by (6.3.2) subject to (6.3.3).
Choose a point Q0 not equal to Po , nor any of the Pi nor any of the Q.; nor
lying on any of the curves aj, bp Set
k
P
r
P
g rP r \
f(P) = exp( E ai f
Qo '
tp p o — E fl•f
1 Qo
TQ
1 p°
± E c;k 0 6j)
.i= 1 j= 1
= exp fP T,
Qo
For f to be single-valued, we must have that J utf ,- ,b, - are of the form 2nin,
n E Z. Now (6.3.1) yields,
2
spp: 20
p( ,
= nt + E
t= 1=1
with n1, m; E Z, I = 1, . . . , g, j = 1, . . . , g. This means we can choose paths
yi and 7; joining Po to P; and Q; respectively, such that integrating over
these paths, leads to the above equation. It is clear that if we choose c; =
—2nim i, f will be single-valued. 0
EXERCISE
We outline an alternate proof of the necessity part of Abel's theorem. Let f be a non-
constant meromorphic function on M. For each aeCuf on), let f '(2) be viewed as
the integral divisor of degree deg f, consisting of the preimage of a. Then a cp(f ' (a))
is a holomorphic mapping of C In) into J(M). Since C uIx) is simply connected,
it lifts to a holomorphic mapping of C u { co} into C°. Since C u {co } is compact this
mapping is constant and thus (p( f '(0))=
The condition ((P i , ,Pg) = 0, however, gives us immediately that the rank
of the Jacobian is g. Therefore, the map
9: U, x • • • x Ug Cg
is a homeomorphism in a neighborhood of (0, ,0) by the inverse function
theorem; and thus covers a neighborhood U of K (in Cg or in J(M)). Let
,cy) a Cg. Then for a sufficiently large integer N, K + e/N e U,
and thus there exists a Q 1 , such that
9(Q 1 • • • Q.) = K + c/N
Or
N(9(Q1 ' • • Qd — K) = c.
Thus to show that c E Image cp, it suffices to show that there exists an integral
divisor D of degree g such that
(P(D)= N(4)(Q1, ,Q9) — K).
Consider the divisor (P i • • PON/W I • • Mil. The Riemann-Roch theo-
rem gives
( (P1 )= 1 + i(( Q' • •• • 1211)N1:1 )> 1.
• • Pg)
r
VQ1' • • Q1Pgol (P1 • • P
Hence, there is an f E L((P, • • P1/(Q,• • W P) ); that is, there is an
integral divisor D of degree g such that
D(P, • • • I'd'
(f) =
(Q1 . • 42,)."11 .
By Abel's theorem
9(D) + N91P1 ' • • Pd= N(49 (Qi • • Qd.
We have solved the Jacobi inversion problem:
111.6.7. Exercise
We have seen that every torus M is conformally equivalent to C/G where G is the
group generated by two elements zi— ■ z + 1 and z + T, Tm T > O. Further, the
origin of C/G may be made to correspond to any point in M.
(a) Every meromorphic function f on M can be viewed as a doubly periodic function
f (this identification should cause no confusion) on C; that is, a meromorphic
function J on C with le`
f(z + 1) = f(z) f(: + s), all e C.
The IVeierstrass p-function is defined by
1 1 1
p(z) = E — a — ms) 2 (n + ms)
2 ), - e C.
()*(O O)
(n.no)e Z.2
Note that p is an even function. Let P, Q e M. Let f e L(1/PaC. Show that there
exists x e Aut M (the group of conformal automorphisms 01M), 13 e Aut(C u {col),
such that
f=fiodoom
(For a discussion of Aut M, see V.4.)
(b) Show that every meromorphic differential on M is of the form f(z)dz where f is
a doubly periodic function. If we choose the loop corresponding to z z + 1 as
the "a-curve" and z 1--. z + r as the "b-curve", then we have a canonical homology
basis on M. The basis for .e(M) dual to this canonical homology basis is then
fdzI.
(c) From what we said above, p(z)dz is an abelian differential of the third kind with
zero residue and singularity 1/z 2 at the origin. Hence there exists a meromorphic
function on C such that C' = — p. This function cannot be doubly periodic
(why?). However, satisfies for all z E C
;(z + 1) = ;(z) + 1 ,
;(z + r) ;(z) + r12 ,
where and g 2 satisfy Legendre's equation
th r — 77 2 = 2ra. (6.7.1)
Derive (6.7.1) from (3.8.2).
surfaces are the ones for which the number of Weierstrass points is precisely
2g + 2. These surfaces thus show that the lower bound obtained in the
Corollary to Theorem 111.5.11 is sharp.
111.7.1. A compact Riemann surface M is called hyperelliptic provided
there exists an integral divisor D on M with
deg D = 2, r(D 1 ) 2.
Equivalently, M is hyperelliptic if and only if M admits a non-constant
meromorphic function with precisely 2 poles. If M has such a function, then
each ramification point has branch number 1, and hence the genus g of M
and the number B of branch points (= ramification points) off are related by
(using Riemann -Hurwitz)
B = 2g + 2.
Remarks
1. We can hence describe a hyperelliptic surface of genus g as a two-sheeted
covering of the sphere branched at 2g + 2 points.
2. Some authors restrict the term hyperelliptic to surfaces of genus 2 that
satisfy the above condition.
Remark. Surfaces of genus 1 are also called elliptic (tori). Surfaces of genus
zero admit, of course, functions of degree 1. Thus, hyperelliptic surfaces are
those which admit functions of lowest possible degree.
111.7.3. Let M be a hyperelliptic Riemann surface of genus > 2. Choose a
function : of degree 2 on M. Let f be another such function. We claim that f is
a Miibius transformation of z. Let the polar divisor of z be P ,Q, and the polar
divisor of f be P2 Q2. It suffices to show that P1 Q 1 P2Q 2 . For then there is
an h e YOM), such that multiplication by h establishes an isomorphism
between L(PT 1 Q t- ') and L(P 2-1 Q2-1 ). Since {1,4 and {1,f} are bases for
Hyperelliptic Riemann Surfaces 95
To establish the above equivalence, we observe first that the branch points
of f are precisely the Weierstrass points of M. To see this let P a M be a
branch point of f. Then f is locally two-to-one at P. Thus if f(P) = co, f has
a pole of order 2 at P (avl no other poles) .and
- then P is a Weierstrass point.
If f(P) co, then
f — f(P)
has a pole of order 2 at P, proving that P is a Weierstrass point on M. It now
follows by Proposition 111,5.6 that the "gap" sequence at any of the 2g + 2
branch points of f is
1, 3, ... ,
and thus the weight of any of these points is
(2k — 1) — k = g2
k= 2 2
Thus these 2g + 2 points contribute g(g2 — 11 to the sum of the weights of the
Weierstrass points. Since the sum of the weights of all the Weierstrass points
is precisely g(g2 — 1) there are no other Weierstrass points.
Let us choose any Weierstrass point P on M. We claim that the polar
divisor of f is equivalent to P2. If f(P) = co, there is nothing to prove.
Otherwise, look at the function 1/(f — f(P) = F. Its polar divisor is P2 which
is equivalent to the polar divisor of f since F is a Mdbius transformation of
f ( f l(co) f (N.
We have therefore established most of the following
= n (z - z(p)).
20-1-
(7.4.1)
b —1 0
element around a closed path changes sign if and only if the path has odd
winding number about ei (if it has even winding number, we return to the
original function element). Thus w changes sign if we continue it along a
simple closed path in Cu{co} that encloses an odd number of the ep If a
curve in C {ao}\{e 1 ....,e29 .,. 2 } begins at z and returns to this point,
enclosing an odd number of ei, then this curve must cross one of the "cuts"
and thus its lift to M is a curve joining the points P and Q, P Q, on M with
z(P) = z(Q). Thus w can be continued analytically along all paths in M.
Furthermore, continuation along any closed path in M (which must encircle
an even number of els when viewed in the plane) leads back to the original
value of w. Thus w is single-valued on M, and for any P, Q on M with P Q,
we have
z(P) w(P) = — w(Q).
The fact that (w) is given by (7.4.2) is an immediate consequence of (7.4.1).
(dz) = Pi • • P2g+2
Q?■21
we see that
dz\ rid ni
V V g2 V3V4.
w )
PROOF. Using the basis constructed in Corollary 1, the span of the products
has a basis consisting of
zi(dz) 2
w2 j = 0, , 2g — 2. (7.5.2)
Remarks
1. Note that 2g — 1 = 3g —3 if and only if g = 2.
2. Torobtain a basis fet the holomorphic quadratic differentials for a hyper-
elliptic surface of genus g> 2 we must add to the list in (7.5.2), the
differentials
:Ad: 12
j = 0, , g — 3.
w
EXERCISE
PROOF. Let Po be the base point (a Weierstrass point) and P another Weier-
strass point. Since 9(1'0= 0, we may assume P Po . We have seen that
there is a meromorphic function f on M whose polar divisor is P2 . Since
Po is a branch point of f, (f — f(P 0)) = POI'. Thus P1 — P2 and by
Abel's theorem,
2 40(P) = 49 (P2) = 49 (N) = a
EXERCISE
Let M be an arbitrary compact Riemann surface of genus g I. Let cp:M --+ J(M) be
the embedding of M into its Jacobian variety with base point Po . Assume that OP)
has order n, for some P E M, P Po . Show that there exists an f e or(M) with (1' = )
Pyrio . In particular if g> 1 and 2 n < g, then both P and Po are Weierstrass points
on M.
100 HI Compact Riemann Surfaces
Remark. The material of the next chapter (in particular IV.8 and IV.9) will
allow us to present at least two other proofs of the corollary to the above
proposition.
Corollary 1.1f g > 2, then the fixed points of the hyperelliptic involution are
the Weierstrass points.
We are interested in special divisors; that is, integral divisors D such that
there exists an integral divisor D* with
DD* = Z.
We call D* a complementary divisor of D.
Trivial (but Important) Remark, From the definition of c(D) it follows that
c(D) and deg D have the same parity (both are even or both are odd).
111.8.2. It is clear that the Clifford index depends only on the divisor class.
More is true.
D
r (i) 1) + r(D- 1)_ r(D r(151-152) .
PROOF. Observe that
L(D5 c L(DDT 1 DI 1 ).
This inclusion follows from the fact that D I D2 /D is integral and is
for j = 1, 2. Thus
L(D,- ) y L(D 2--1) L(DDT 1 D2-1 ), (8.3.2)
where y denotes linear span. Next we show
L(D 1 ) n L(D2-1 )= L(D-1 ). (8.3.3)
To verify (8.3.3) assume Di is given by (8.3.1). If f e L(Di 1) n L(Di 1 ) has a
pole at P of order a 1, then
a 5 ai(P), j= 1, 2.
Thus also
a 5 ininfai(P),OE2(P)},
and f e L(D'). We have established
L(13 -," 1 ) n L(DV) c L(D').
The reverse inclusion follows from the fact that Di > D, j = 1, 2. This ver-
ifies (8.3.3).
Finally, using (8.3.2) and (8.3.3) and a little linear algebra,
r(DT 1 )+r(Dji )—r(D - 1 )= dim L(DT 1 )+ dim L(D 2--1 )—dim(L(DT I )nL(DV))
= dim(L(D ') y L(132- 1 )),.. dim L(DD I-1 D2-1 )
= r(DDT 1 D 2-- 1 ).
= 2c((D,D*)).
This can be done since r(D - ) > 3. All we want is a function in L(131 -1 )
that vanishes at some point of D* and at one point not in D*. Let f be such a
function. Then (f)D is integral, equivalent to D, and satisfies (8.4.2). We
have produced a special divisor (13,D*) with 0 < deg(D,D*) < deg D and
c((D,D*)) = O. Thus 2 < deg(D,D*) 15. deg D — 2, and we can arrive at a
special divisor of degree 2, with Clifford index zero.
111.8.5 We now derive some consequences of Clifford's theorem. If
D e Div (.%I) is arbitrary with 0 < deg D :5_ 2g — 2, then c(D)> deg D unless
r(D - > 2. In the latter case (since there is a non-constant function f in
L(D')) we have D' =.(f)D is integral and equivalent to D. Since linearly
equivalent divisors have the same Clifford index, we have almost obtained
Corollary 1. Let D be a divisor on M with 0 < deg D < 2g — 2. Then c(D) > 0.
Equality occurs if and only if D is principal or canonical, unless M is a hyper-
elliptic Riemann surface.
PROOF. The remarks preceding the statement of the corollary show that unless
r(D - 1 ) > 2, c(D) deg D. If r(D -1 ) 2, we have D equivalent to an integral
divisor D' of the sanie degree. Proposition 111.8.2 allows us to assume with
no loss of generality that deg D g — 1. Now D' is a special divisor. The
fact that c(D') = c(D) > 0 follows from Clifford's theorem. If D is neither
principal nor canonical (and since as already stated we may restrict our atten-
tion to divisors of degree <g — 1, we are only interested in the case of D
not principal), we have c(D) = 0 gives r(D') = 1 + deg D. If deg D > 0
we have once again r(D - 1 ) > 2, and as before D is equivalent to a special
divisor and Clifford's theorem implies that M is hyperelliptic. If deg D = 0,
we have r(D -1 ) = 1 and D is principal.
Thus either Ç, or (say I: 3) must vanish at at least 3 of the zeros of(,. Thus
C 1 /C 3 is a non-constant function with at most three poles. This function gives
rise to a special divisor D of degree 3 with r(D -1 )> 2. Theorem 111.8.4
shows that r(D - < 2. Hence r(D - = 2.
111.8.7. Theorem. Let M be a compact Riemann surface of genus 4. Then one
and only one of the following holds:
a. M is hyperelliptic.
b. M has a function f of degree 3 such that (f)= with D, 91 integral and
132 — Z. Any other function of degree 3 on M is a fractional linear trans-
formation off.
c. M has exactly two functions of degree 3 that are not Meibius transformations
of each other.
PROOF. Proposition 111.7.10 implies that (a) cannot occur simultaneously
with (b) or (c). We have already shown that every compact surface of genus
4 admits a non-constant function of degree <3. Note that for any divisor
D of degree 3 on M (of genus 4), we have
r(D - i(D). (8.7.1)
Suppose we are in case (b). Let us assume there is an integral divisor
D 1 0 D with deg D t = 3 and r(Di- = 2. Let ft E L(DT %C. If D 1 is equiv-
alent to D, then ft is a Möbius transformation off . To see this, let h e
be such that (h)D i = D. Then L(D - 1 ) = 1,f } and L(Dt- = {h, fh} = {1,
110 III Compact Riemann Surfaces
Remark. Cases (b) and (c) correspond to the relations of rank 3 and 4 of
(8.6.4) respectively. In case (c) we have (f) = 91/D, (f,) = 11,1D, = D2/D 1 .
Consider the identity
(DD,)(2191 1 ) = (91D,)(D91 1 ).
There are holomorphic differentials j = 1, , 4, satisfying
(Ci) = Bpi, (C4)= (MO= D 21 i,
9.1
( ( 2) = (KO = MIL = 219f ,,
D I 91
((s) (f 1/7:4) = 5 D91 1 = D ,21.
1112. Special Divisors on Compact Surfaces Ill
D,
(4:2)= D2, (47.1)= DD,, ( .3)= (l.f1)= DD, — = D1
D '
(where (fi ) = D I/D). Thus, we obtain the relation of rank 3 (after adjusting
constants)
= C2C3.
Again, the three differentials in question are independent, since D, 0 D.
Thus if al1 + a21 2 + a3C3 = 0, choosing a point P e D, P D,, gives
a 3 = O. Similarly, choosing a point P c D,, P 0 D, gives a2 = O.
Case I: B is not the polar divisor of a function. Then there is at least one
P E B such that r(PB -1 )= r(B'). Let B = B'P and observe that
2 deg B' g — 2,
and
c(B') = deg B' — 2r((13') - + 2
= deg B — 1 — 2r(B - + 2 = c(B) — 1 = 0.
Hence the surface is hyperelliptic by Clifford's theorem.
Case II: B is the polar divisor of a function. In this case we apply Corollary 1
and Corollary 1 to Clifford's theorem (111.8.5), to obtain
fa(Pi)
rank [ fi(P1) 2
f1(P2) • fa(Pk)i
Having now chosen P,,.. ,P,,_ such that no PI appears in A'B' and such
that det(f„,* (P;)), m = 1, , k — 1,j = 1, . . . , k — 1 (k s) does not vanish,
we consider
.1.01) • • • fk(P1)
det
i) ' • • fk(Pi---11
.îi(P) • _
The linear independence of ffi, ,f,) assures us that the determinant
does not vanish identically so that we can choose P, not to appear in A'B'
and such that det(f„,(P i)), m = 1, . . . , k, j =1, . . . , k does not vanish. Hence
1) • • • fd( 1) = k.
rank [
f,(Pk) ••• fd(Pk)
This verifies the existence of a divisor D of degree s with the required
properties and in fact, shows that almost all divisors D of degree s would
work,
Finally, we need show the existence of divisors A — A' and B B' such
that (A,B)/D is integral. To this end consider L(D/A'). Clearly (for s
r(DIA') t + 1 — 1, and thus there is a non-constant function f E L(D/ A')
such that (f)= DI,/A' and we may take A = D11. Similarly there is an
integral divisor 12 such that B may be chosen as D/ 2 .
and hence,
AB
1 deg( --.) 2g< — 3.
(A ,B)
Thus, by Clifford's theorem,
AB
= 1 = c((A,B)).
c((A ,B)
We now consider cases:
Case I: r(11(A,B))= 1.
From-the definition (*Clifford index,
1 ---- c((A,B))= deg(.4,B).
But Proposition 111.8.9 implies that
c( 4112) = c(AB) deg(A,B) — c((A,B)) + 2(1 — s) = 2(1 — s).
Since s > 1, we see by Clifford's theorem that s = 1,20 Z, and r(91 - 1 ) = 3.
Furthermore,
1 = c(21) = deg 91 — 4,
shows that deg Z = 2 deg 91 = 10 or g = 6.
Case II: r(11(A,B))._ 2 (thus deg(A,B) 3).
In this case we may assume that (A,B) is the polar divisor of a function.
(Note first that if P E AI appears in the divisor (A,B), then
1\
r 1.
((A,B)) = r ((A ,B)
Otherwise,
=
deg(A,B) 1 2r(1 2 = c((A,B)) — 1 = 0,
P (A,B))+
contradicting that M is not hyperelliptic. A function belonging to
111.8.12. To change the pace slightly, we prove a result in linear algebra. For
the proposition of this section, we will need some elementary results from
algebraic geometry.
To begin with, the set of r x r symmetric matrices can be viewed in a
natural way as a vector space of dimension 112-r(r + 1) which we identify
with Ç(2)(+ 1)
tA 01 F A Bi ['AA P:
L`B `Di LO j LC D ['BA 'BB LC BS
From the above it fo lows that A is non-singular CAA = A implies (det A) 2 =
det A 0 0). We claim that A, B uniquely determine C, B. It is clear that
C =T. Further, since fi = 'AB and A is non-singular b = 'BB =
t i(rAA) -I fi=q1-Â - '11. Conversely, given a non-singular symmetric A as
above, it can be written as 'AA for some A. Also given an arbitrary IT as
above, we can define a matrix t of the above form. Thus we see that dim V,
is equal to the sum of the dimension of all possible A and the dimension
of all possible
111.8. Special Divisors on Compact. Surfaces 117
k p + 1)p(r — p) r(r + 1) 1. 0
aikfifk = O.
j.k=1
and f 11/13 has not more than n/2 zeros and is not constant. We can set
B = the divisor of zeros of f, and observe that 1/f E L(B -1 ).
E afimg , = o
j.k=
111.8.15. The next lemma is both technically very useful, and explains
what it means for an integral divisor to have positive dimension. Roughly
it says that an integral divisor D has dimension >s if and only if D has
s — 1 "free points", and gives a precise meaning to this statement.
Remark. The lemma could have been established at the end of 111.4. We
have delayed its appearance because its proof uses techniques of this sec-
tion. In fact, we have already used and proved part of the lemma in 111.8.10.
PROOF. To prove the sufficiency, we assume r(1/D) = d and that (f1 , ,fd )
is a basis for .1(1/D). As in 111.8.10, we construct a divisor D' = Q, • • • Q,_
such that rank (fk (Qi)), k = 1, . , d, j = 1, . , s — 1, is precisely
min {d, s — 1) and such that none of the points Oi appear in D. If d s — 1
we would have that the only function which vanished at Q1,..., Q 1 in
L(1/D) would be the zero function contradicting the fact that we can choose
an integral divisor D" such that D'D" D. Hence d > s.
To show necessity assume that r(D -1 )= s> 1. Note that for arbitrary
divisor 21 and arbitrary point P E M, r(P91 1 ) > r( 21 - 1) — 1. Now let D' be
integral of degree — 1. By the above remark, r(D'/D) s — (s — 1) = 1.
Now choose f e L(D'/D). Thus there exists a divisor D" such that D'D" D.
where
eval( f,P) = f(P)
proj(f,P) = P.
The verification of the above claims is routine, and hence left to the
reader. In this connection the reader should see also IVA.
111.9.2. We shall be interested exclusively in a restricted class of compo-
nents g7 as above. Namely, we require that
i. proj be surjective, and that
ii. for every path c:I —> M, and every f e ...42-• with proj(f) = c(0), there
exists a (necessarily) unique path a:I 3-- with E(0) = f and c = proj 0 E.
We shall call e the analytic continuation of E(0) along c.
Note that by the Monodromy theorem E(1) depends only on the homotopy
class of the path c and the point E(0). (In the language of 1.2.4, we are con-
sidering only those components Y which are smooth unlimited covering
manifolds of M.)
111.9.3. Let M be a compact Riemann surface and rc,(M), its fundamental
group. By a character x on ic,(M) we mean a homomorphism of ic,(M) into
the multiplicative subgroup of C, C* = C\10). Since the range is commuta-
tive, a character x is actually a homomorphism
x : Ii i(M) C*.
111.9. Multivalued Functions 121
= esk or ak = Sk ,
and
t cgr,
= v -,
Re c,x ,
= euk or E Re cirri, = uk . (9.9.4)
1=1
To see that this cho'ce is indeed possible and in fact unique, recall that we
can write the matrix 17 = (ink)
k ) as
11= X + iY
with Y positive definite, and thus non-singular. To solve (9.9.4) we write
c = r(c,, ,eg), = t(a l , ,;), etc.,. , and note that we want to solve
Re[(X + iY)(a + ill)] = u.
Since we have already chosen a = s the equation we wish to solve is
Re[(X + iY)(s + il3)] = u,
124 Ill Compact Riemann Surfaces
z(b) = exp(fli), j = 1, , g,
and fixing e> 0, we may assume in addition
where 2 = (2,, ... ,2g) e Ug is a variable point, and Po U. For fixed I' we
get a multiplicative function with character zf satisfying (9.11.1). This is,
of course, a consequence of the fact that fak rP = 0, for k = 1, . . . , g. Further-
more,
Mk) = exp E bk T. . = exp>2 2rci f 1') r,k (9.11.3)
by the bilinear relation (3.6.3), where {C,, ..; g } is the normalized basis
for the abelian differentials of the first kind dual to the given holomogy
basis.
Now (9.11.3) obviously defines a holomorphic mapping 40: Ug Cg
= (0 1, • • • , ( Pg)
9g(zg)_
126 III Compact Riemann Surfaces
Remark. The reader should notice the similarity between the proof of the
above theorem and the proof of the Jacobi inversion theorem.
Remark. The isomorphisms established in (9.12.2) and (9.12.3) are also useful
in their own right.
111.9.13 An important class of characters are the so-called nth-integer
characteristics. Let {a i, b 1 , . . ,b,} be a canonical homology basis on
111.9. Multivalued Functions 127
with ere; integers between 0 and n — 1. The symbol [], will be called an
nth-integer characteristic. It determines a normalized character x on M via
X(a) = exp(27tiej/n),
x(b3) = exp(27ris;/n), j = I, . . . , g.
P. le' _ • -.'
It is clear that for every multiplicative function f belonging to such a
character x, Ifl is a (single-valued) function on M, and f itself lifts to a function
on an n-sheeted covering surface of M. Furthermore, we can construct such
an f by taking nth roots of a meromorphic function h on M provided (h)
is an nth power of a divisor on M (that is, provided the order of zeros and
poles of h are integral multiples of n).
111.9.14. Proposition. For x E Char M,
if x is inessential,
ix (1 ) = { gg _ I if x is essential.
PROOF. The Riemann-Roch theorem says:
r - i(l) = —g + 1 + i,(1).
Thus, i1(1) ..- g — 1, and ir(1) __ g. Furthermore. i,(1) = g if and only if there
is an f E L 1 _ 1 (1), f # O. Since such an f must be a unit if it is not identically
zero, we are done.
111.9.15. We end this section with the statement of Abel's theorem for mul-
tiplicative functions belonging to a character x. The proof is omitted since
it is exactly the same as the proof already studied (in 111.6.3).
111.9.17. Exercise
(1) Let P E M and let z be a local coordinate vanishing at P. Let n e Z, n > 1. By a
principal part (of a meromorphic function) at P we mean a rational function of z of
the form
f( ,)= akz4.
-n
(5) Using the normalized abelian differentials of the second kind introduced in 111.3,
show that
(a) Every system of principal parts is the system of principal parts of an additive
function.
(b) For every x e Hom(11 1 (M),C) there is an additive function belonging to z.
(c) Let { al, b1, ,b9} be a canonical homology basis on M. A homomor-
phism x e Hom(H,(M).C) is called normalized if x(c(i) = 0, j = 1, , g. Show
that every system of principal parts belongs to a unique normalized homomor-
phism.
(6) Prove that a system F of principal parts is the system of principal parts of a (single-
val ued) mcromorphic (OM ion on M if and on ly if F induces the zero linear functional
on
and
R. P, Q, are three distinct points on M.
(8) Let F be a system of principal parts at {P I , ... ,P.}. Let Q be arbitrary but distinct
from Pi, j = 1, , m. Define for P # P, P # 0,
E(P)= — F(T pi!),
Show that E agrees on M ,P„„Q} with the unique (up to additive constant)
additive function with F as its system of principal parts and belonging to a
normalized homomorphism.
projective space
0:M pd
130 III Compact Riemann Surfaces
= 2 — g + 1 + r(PQ/Z)= 2.
1)
Thus M must be hyperelliptic. Next we assume q> 1, and compute
r(Z - qPQ)= q(2g — 2) — 2— g + 1 + r(Z9-1 /PQ)
= (2q — 1)(g — 1) — 2 + r(Zq - 2 /PQ).
Now for (10.2.2) to hold we must have that
r(Zi - 1 /PQ) =I;
which implies that
deg(Z"/PQ) 0.
This last statement is equivalent to
(q — 1)(g — 1) 1.
Since g 2 and q> 2, this is only possible for g = 2 = q. Note that the above
argument also establishes that for each P E M, 2 is a "q-gap" except if g = 2 =
q or q = 1 and M is hyperelliptic.
111.10.3. To study the excluded cases, we represent a hyperelliptic surface
M by
w2 = (z — el ) • • (z — e29 + 2)
with distinct ei. We have seen that a basis for dri(m) is then
j dz zdz zg dz}
tw w
Again
IdZ 2 (12 2 2 C/Z 2
W2 W 2 w2 1'
0(P) = (1,z(P),z(P)2)
is independent of w, and the 2-canonical map (for genus 2) is two-to-one, and
not of maximal rank at the Weierstrass points. 0
111.10.4. Since the image of a compact manifold under an analytic mapping
of maximal rank is a sub-manifold, we have obtained
132 111 Compact Riemann Surfaces
w = E fi(z)dzi,
with is, holomorphic. In particular, to is closed. Since
I71 111 af
we also have
= (14- , all k, j = 1, . . . , m.
ez, cz j
Thus, letting e E R2", it follows by (11.1.1) that there is a unique vector E C"
such that
Re E xi du, = ck , k = 1,.. . , 2n. (11.2.1)
ak
Now (11.2.1) shows that the differentials du,, du,, , du„ are linearly inde-
pendent over C. Furthermore, given any holomorphic 1-differential on T,
there exists a differential co = x dui such that
Define a function F on T by
w=E f3(z)dz.
1=1
III] I. More on the lacobian Variety 135
R2" C"
which is simply another way of saying All = 17M.
It follows immediately from the previous remarks that a holomorphic
injective map of one complex torus into another torus of the same dimension
is necessarily biholomorphic. If H i and n, are the period matrices of the
two tori, the map can be represented by a matrix A :C" C" such that
A/7 2 = 11 2 M for some M as above. It is necessarily the case that both A and
M are non-singular. Thus we have established the following
Further,
11 2 M E SL(2n,Z).
Thus
1 1 1M = Afl ,M = A17 2 = 112 M 2 =
and hence (since H " C" is an isomorphism)
MiM = MM2.
It thus follows, since M is non-singular, that trace M i = trace M2.
3,- Ejkk, J = 1, . . . n,
k=1
111.1 1.8. We have seen in 111.6, that the mapping 9 extends to divisors
on M. Let W. be the image in J(M) of the integral divisors of degree n. Set
by definition Wc, = {0}. It is thcn clear that
WN WN+1,
(because tp(D) = cp(DP0) for every divisor D) and 147, = J(M)(Jacobi inversion
theorem).
Let W„ be the set of points in J(M) which are images of integral divisors
D that satisfy the two conditions deg D = n, and r(D - ')> r + 1. Denote
by K the image under of the canonical (integral) divisors. By Abel's theorem,
K consists of one point.
Proposition. 1K) =
111.1 1.9. Let M be a Riemann surface and n > 0 an integer. The n-fold
Cartesian product, M", is naturally a complex manifold of dimension n. By
M. we denote the set of integral divisors of degree n on M. We topologize
and give a complex structure to M. as follows: Let Y„ be the symmetric
group on n letters. An element o- E .99„ acts on M" by:
= ( — 1) E
= ( _ 1)2 EZiZk
Pck
(„ = ( - 1rZi • :in .
Notice that C.; is (locally) a holomorphic function on M. and that the ordered
set {C I , ,C„} determines the unordered set {z i, ,z„} uniquely as the
roots of the polynomial z" + C i f + • • • + C„ = O. Hence =
is a local homeomorphism of a neighborhood of p(P 1 , 'P.) E M„ and
thus serves as a local coordinate at p(P 1 , ,P„) E M„.
111.1 1. More on the Jacobian Variety 139
t„ k = 1, , n.
j=1
M„
commutes, then f is holomorphic if and only if fi is.
Remarks
F(z,, ,z„) =
k=0 j=1
and conclude that in terms of the local coordinate (t 1 , ,t„) on C'115' „,
OF 1 dif
= a..
at dZi z=0
E m = n.
I
Consider the map
: M,,
in a neighborhood of the n-tuple A made up of elch Qs appearing ink times.
Let 6" be 3 pulled back to M,,. A neighborhood of the point A consists of
points
Q1, Q1, • • . Qi, • • • , Q,1 2 , •••,Qi,
s • • • Qfn,,
E
k= 1
mkhk(zk)
defines a function on M„ whose differential is 3". (To see this recall the
definition of the pullback of a differential. The differential can be written
as df with f a linear function on Cg. Now hk is simply f o cp. The differential
of f ocps is 3", where 9„ is (I) viewed as a map of M„ into J(M). Since f is
linear, f (9 JP • • P.)) = f(9( 13 1) + • • • + (p(P.)) = .1((P(P 1)) + " • +
f(T(P s)).) Assume now that .5" vanishes at the point A. Since the pullback
of 3" to M" vanishes at we must have dh k/dz k = 0 at zk = 0 for k = 1, . . . , s.
We need more accurate information about the differential 6' in the case that
at least one mk > 1. In this case, the function
hk(zu) + • • • + hk(z.,k) (11.10.1)
can be expressed as a power series in the elementary symmetric functions
of the second kind of the Ink variables z1 . If we write
hk(zk) = E
v=1
111.11. More on the Jacobian Variety 141
we see that the coefficient of the /-th symmetric function of the second kind
in (11.10.1) is precisely a,. We see therefore that
6" = d( E Mkhk),
k= 1
Let us denote by Z a local coordinate vanishing at Qk and let Ci, the j-th
normalized differential of first kind, have the power series expansion
E arzt dzk
(
I= 0
will have zeros only close to the zeros off ° or close to the distinct points of D.
To see that this is the case, we first observe that if P is a zero of f° of multi-
plicity m then f has m zeros in a neighborhood of P. It is, however, possible
that new zeros are introduced which are not in these neighborhoods. We
claim that the new zeros must be in neighborhoods of the distinct points of D.
(Delete from M the open neighborhoods of the distinct zeros of f ° and the
111.1 1. More on the Jacobian Variety 143
Remarks
1. We saw in the proof that the rank of the Jacobian of ço at D E M. can be
written as g — i(D).
2. A "generic" (see 111.6.5) divisor D E M„, n < g, has index of specialty
i(D)= g — n. Thus W. for n g has dimension precisely n (as expected),
by Remark (1).
111.11.12. By the inverse function theorem it follows that at the image of an
integral divisor D of degree n < g with r(D - = 1, there are local coordinates
• , .79 for J(M) so that the points of W„ are given by the equations
• = • • • = = 0. We say in this case that W„ is regularly embedded at
9(D).
Conversely, if W„ is regularly embedded at a point u e J(M), then there are
g — n linearly independent holomorphic differentials on J(M) whose pull-
backs to M. vanish at the pre-image of u on M.. We have seen (Lemma
111.1 1.10) that such a holomorphic differential corresponds to a differential
6 on M such that (3) is a multiple of D for all D E 9 -1 (u). Now if u E W,1„
then for any Q E M there is a divisor D of degree n containing Q such that
(p(D)= u. Hence the differential 6 must vanish identically on M. But then
it vanishes identically on J(M). Now if g — n 1, and W. is regularly im-
bedded at u, there is at least one differential on J(M) that vanishes on W.
and is not identically zero. We have established the
Remarks
1. The above considerations allow us to describe more intrinsically the
tangent space to J(M). Let u 1 . .... u9 be the canonical coordinates on Cg.
Then these also serve as local coordinates at an arbitrary point x e J(M)=
Co/G. In terms of these coordinates, a natural basis for the (complex)
tangent space Tx (J(M)) of the manifold J(M) at the point x is given by
the vectors a/aui, , a/aug. On the cotangent space 'T(.1(M)), the
144 III Compact Riemann Surfaces
g a g y
E ai .---)x
bi du)=
(E b E ai i,
( i=i u uj „,=„1 i= 1
Remark. The proposition shows that the sobvarieties W J(M) for t < g
are as far from being translation invariant as possible. Specifically an
inclusion 141, + u c Wu W, = {0 } .
111.11.16. Proposition
a. For any r, t 0 and any a, b e J(M), we have
(W, + a) + (W, + h) + a + b).
b. For 0 r t g — 1 and any a, b E J(M), (W, + a)e (W,. + b) =
Remarks
1. The condition that t < g — 1 in (b) is necessary because W, = J(M) for
t g. Hence for any S c J(M), es=
J(M) for t g.
2. A special case of the proposition is also worthy of mention here:
= Wg-1 e Wg-2 =
1,6 W9-2
K/9_ 1 - u.
Thus W1 M can be recovered from 14/9 _ 2 and Wg_ Hence M is
determined by W8 _ 1 and Wg _ 2.
111.11.17. Proposition. For any n > 0 and any r > 0,
Wr„ = W„_, e(-- W,), whenever r < n, and (11.17.1)
W rn = Ø whenever r > n. (11.17.2)
PROOF. Let x E W. Then x = (p(D), D E M., and r(D') r + 1. The
Riemann—Roch theorem gives r(D") = n — g + 1 + i(D). This is clearly
impossible when r > n.
We have previously seen (Lemma 111.8.15) that u E Wr„ if and only if
for every v E W, there is a if E W, such that u = u + e. Hence
= n
ye W,.
(w„_, y). (11.17.3)
Remarks
1. Formula (11.17.3) can be reformulated as u E —(W, — u) c W,,„.
Clifford's theorem gives a necessary condition that Wr„ not be empty.
111.11. More on the Jacobian Variety 147
The condition is that 2r < n or that r < n — r. Hence the above gives
a geometric interpretation of this theorem.
2. It is an immediate consequence of the proposition (in the form of
Equation (11.17.3)) that the subsets W: are complex analytic subvarieties
of J(M).
3. If we define W„ to be the empty set for n < O, then (11.17.2) becomes a
special case of (11.17.1), and (11.17.3) is always valid. We shall adopt
this convention.
If the two divisors are not identical, then r(P,77 • • P; > 2 and u E W , 1 .
The reverse inclusion is trivial.
Note that the first g products are linearly independent and so are the last
g products. Let A, and A2 denote the subspaces of d 02 (M ) spanned by
these products. Now dim(A, n A 2) = dim A, + dim A2 — dim(A, v A 2),
and dim(A, n A 2) = I. To verify the last assertion write (coi) = Dpi,j = 1, 2,
and note that D, and D2 are relatively prime integral divisors of degree g
150 In Compact Riemann Soi faces
= 2g — 1 + g — dim(A v A2 V A 3).
We show that dirnt (A, v A 2) n A 3) = 2. and in fact the space in question is
spanned by w 1 w 3 , co,w 3 . Note that if i) e(.4 1 v A,) n A3, then q = (:30) 3 =
co, ± ," 2a) 2 . Since the right-hand side vanishes at D, it follows that (C 3) =
DD. Thus 4:3 = X1011 ± X2(02 (with Xi E C) and the dimension is 2, as claimed.
Thus dim(A, v A2 V A3) = 3g — 3 = dim y" 2(M). This concludes the proof
for g = 2.
For the case g = 3, we start by choosing two elements co l , (0 2 E Ye i (M)
such that co, and co, have precisely one common zero at Po , and a) 3 E Ye l (M)
such that (03 does not vanish at the zeros of w 1 co 2 .
Let {f„ ,f39 _ 3 } be a basis of if 2 (M) with each fi a 2-fold product of
elements of Ye i (M). Let A ; he the (3g — 3)-dimensional space spanned by
3. As before
EXERCISE
Recall the exercises of 111.7.5 and complete the above discussion for hyperelliptic
surfaces (including the case g = 2).
CHAPTER IV
Uniformizati on
This chapter has two purposes. The first and by far the most important is to
prove the uniformization theorem for Riemann surfaces. This theorem
describes all simply connected Riemann surfaces and hence with the help
of topology, all Riemann surfaces.
The second purpose is to give different proofs for the existence of mero-
morphic functions on Riemann surfaces. These proofs will not need the
topological facts we assumed in Chapter II (triangulability of surfaces). As a
matter of fact, all the topology can be quickly recovered from the complex
structure.
This chapter also contains a discussion of the exceptional surfaces (those
surfaces with abelian fundamental groups), an alternate proof of the
Riemann—Roch theorem, and a treatment of analytic continuation (algebraic
functions on compact surfaces).
1 ine).
= a, + E ianeine+ Le'
n1L
Multiplying the above by e -i" and integrating, we get
f(z)=
17, f 2x g(ren[l + 2 E re()
tz 1
re' + z
= j ' g(rele) . dB (1.2.1)
2ir 0 2 re' - z '
and
g(z)= Re f(z)= — s g( re") Re w dB
2n 0 re - z
1 $27r r2 - 1:1 2
= — g(ren .0 2 dO. (1.2.2)
27r 0 Ire' - z1
The expression (r2 - 1z1 2)/Irew - z1 2 is known as the Poisson kernel (for the
disc of radius r about the origin). It has the following important properties:
2. r2 - 1z12
d19 = 1, (1.2.3)
2n
10J - 21 2 Izi <
1V. 1. More on Harmonic Functions (A Quick Review) 153
r2 - Izi 2
> 0 for fzi < r. (1.2.4)
r2 -
urn
z -re°o 2n
I 9-
fr-?-1
8°I>— 12 dB = 0 for 0 < I < (1.2.5)
Il <r
From the reproducing formula (1.2.1) we also get a formula for the har-
monic conjugate of g that vanishes at z = 0, namely,
1 r 2norea) , _ re' + z
Im f(z) = 1." o - z d0
Jo
1 nit re-"z - reit
=— g(re0) d0. (1.2.6)
-z' 2
IV.1.3. Theorem. The Dirichlet problem for the disc has a unique solution;
that is, given a continuous function f defined on {Izi = r}, there exists a con-
tinuous function F on {Izi rl such that Fis harmonic in IN < ;1 and F(ren =
f(ren, 0 0 2 1t.
PROOF. Without loss of generality f is real-valued. Since real harmonic func-
tions satisfy the maximum and minimum principle, uniqueness is obvious.
For existence, one sets for fzi < r
s 21 r2 — 1:12 .
F(z) = f(re")d0 (1.3.1)
2n 0 -
and uses the properties of the Poisson kernel (1.2.3)-(1.2.5). El
Corollary. If f is a continuous function on a domain D c C, and f satisfies
the mean-value property in D, then f is harmonic in D.
PROOF. Again without loss of generality f is real-valued. Solve the Dirichlet
problem for f — zol = rl with {iz - z ol c D. Call the solution F.
Then F - f satisfies both minimum and maximum principles, since the
mean value property is all that one needs to prove these principles. Hence
F =f on {iz - zol r), and f is harmonic.
u) such that
1 u(: 1 )
— c, all :1 , :2 E Dl.
C U(.72)
This can be seen from the fact that the limit function is continuous and
necessarily has the mean-value property.
are both open. Hence one of them is empty. The same inequalities show that
the convergence is locally uniform. Thus the result follows from the previous
proposition.
IV.1.7. Theorem. Let u be a harmonic function on IO <z < 11. Then there
exist constants a, 13 such that for 0 <r < 1,
2x
o u(reie)d0 = a log r + (1.7.1)
IV.!. More on Harmonic Functions (A Quick Review) 155
PROOF. Recall Formula I (4.4.2). For u 1 , u, harmonic and 0 < p < r (and
the usual counterclockwise orientation for the circles) v‘e have:
(– fi is, of course, the value of the right-hand side of the above equation for
r = p, which we take to be a fixed value). Thus
Œ = fi r *du.
Corollary 2. If u is harmonic and bounded in (0 < 121 < 11, then u can be
extended as a harmonic function to lizi < I).
PROOF. Since a = O. *du = O. Thus (we assume u is real, this involves
no loss of generality) there exists an analytic function f on { 0 < jzi < 1},
with u = Re f on {0 < Izi < 1 } . Set F=exp f. Since IFI
= exp u, and u is
bounded, so is IF1. By the Riemann removable singularity theorem F can be
extended to izi < 1. Since a pole or an essential singularity off is an essential
singularity for F, f has a removable singularity at z = O. El
i f 02E u(end0
u(0) -3Tr
1 r,n
u(0) r2ir 14" Li0 Iu(0) u(k) (ei° ) do,
-i---n Jo
27c
proving that u(') is subharmonic. 0
Then r E C 2(M) and „1r = 4.1u + 4E> 0. Thus, by the first part, y is sub-
harmonic. Hence, for every K, a conformal disc on M, y< or ,
Remarks
1. In most applications the functions y satisfying (2.6.2) and (2.6.3) will be
u( K) and max{u 1 ,142 }, respectively.
2. If F is a Perron family on M, if K is a conformal disc in M, and if u; E 37-",
j = 1, , n, then there is a V e Y such that viK is harmonic and y
= 1, . , n.
and prove the theorem for K. Let it :JILT= 1 be a dense set of points in M. (We
cannot at this point choose a countable dense set in our original M, since
we do not know )et that every Riemann surface is second countable.) For
each ], choose a sequence such that
u(z) = lim V k(2
k--•
Assume u + cc. Without loss of generality we assume that u(z i ) < + cc.
By IIarnack's principle
W = lim yk (2.6.5)
Remark. The above proof also showed how to obtain the function u. Let
K be an arbitrary compact subset of M. Cover K by finitely many discs
{K; = 1, . . ,n}. For each j, there is a sequence of increasing harmonic
functions {VA } such that
u = lirn Vi,, on K
and
2. lim sup u(Q) f(P).
Q
PROOF OF (1). If f(P) = mo , there is nothing to prove. So assume f(P) > trio .
Choose e> 0 such that f(P) — e > m o . There is then a neighborhood N(P)
such that
f(Q) f(P) — & for all Q E N(P) n SD.
Let # be a normalized barrier at P which is 1 outside N(P). Set
w(Q) = — (f(P) — mo — 6)P(Q) +f(P) — e, Q E CI D.
Clearly, 11? E C(CI D) and w is subharmonic. For Q E Cl D,
w(Q) f(P) — a < mi,
and
w(Q) = mofl(Q) + (f(P) — 0(1 — I3(Q))
moll(Q) + mo( 1 — fl(Q)) = mo.
Finally, for Q E SD,
w(Q) = mo f(Q), Q N(P),
and
w(Q) f(P) — & f(Q), Q E N(P).
We have shown that w E F. Hence, w(Q) u(Q), all Q E D, and thus
lim inf u(Q) w(P) = f(P) — e.
PROOF OF (2). There is nothing to verify if f(P) = in,. Assume f(P) < m,.
Choose e > 0 so that f(P) + e < in s . Choose N =- N(P), a relatively compact
neighborhood of P, such that
f(0). f(P) + e for all Q e N(P) n D.
Let y cF be arbitrary. We claim
r(Q) — (m, — f(P) — e)fl(Q) f(P) + E, Q E N(P) n D. (2.8.1)
(Here # is again a normalized barrier of P, that is F_-- 1 outside N(P).) Since
the function on the left of the inequality is subharmonic, it suffices to check
the inequality on ii(N D). If Q e iS(N D), then either (i) Q e 6N n
Cl D or (ii) Q E Cl N n 6D. In case (i), the left-hand side of (2.8.1) satisfies
=r(Q) nil +f(P) + e f(P) + e.
In case (ii) we have the estimates (for the same quantity)
v(Q) f(Q) f(P) -F e.
We have thus verified (2.8.1). Thus for Q E N(P) n D
v(Q) fiP) + (M1 — f(P) 6)13
and hence also
1 4 0) .f(P) + & + f(P)
From this last inequality we conclude
him sup u(Q) 5_ f(P) + e.
Hence
u < m, e,
and since e is arbitrary, we are donc.
i. to is harmonic on M\K,
= 1 on (5(M\K), and
iii. O <w < 1 on M\K.
Note that 4r., contains ro and is hence non-empty. Furthermore, co, has all
the properties of co. We claim that co l < (.75 for all C15 enjoying the properties
of the theorem. Consider
ve
Assume that v I M\K' = 0 for some compact set K'. Thus
on 5(K",K)
(since Co 0 on SK', u = 0 on 61(' ; & > v on SK). Hence Co' > v on (K'\K) k.)
(M\K'), and since v is arbitrary (4) 1 on M\K.
EXERCISE
Let D c C be a domain bounded by finitely many disjoint Jordan curves C 1 , , Cn.
The harmonic measure of CJ(j = 1, . ,n) with respect to D is classically defined as the
unique harmonic function co) in D that is continuous on Cl D and has boundary values
k = 1, • • . , n. Show that this definition makes sense and relate this concept to the
one introduced in IV.3.4. Show that co l + • • • + con = 1 and discuss the properties of the
(n — 1) by (n — 1) "period" matrix Li *dcok.
Inequality (3.7.2) implies that g(Q) < oo for Q 0 P. Thus, by Perron's prin-
ciple, g is harmonic in MVP}. The fact that g > 0 on MVP} follows from
the corresponding fact that u > 0 on M\{P} for all u e „F. Next we show
(3.6.3). Note that for 1z1 < r, u E
r A log r
u(z) + logz
11 + log r + log r = ; . .
4.• A, 7 1 4— 1
Thus also
„ log r
g(z) + log for 1z1 r.
—1
Hyperbolic
for the Green's function, and let Q0 E D\{P}. It suflices to show that
#(120) gu(P,120) — E
for all e > 0. For each Q e iSD, there exists a neighborhood UQ of Q in
Cl D\{ Q0} such that gr, < e on UQ . Set
U= U U.
QeSD
Then on D\U we have — gp> —e since we have this estimate on S(D\U)=
SU. In particular, j(Q0) — gD(P,Q0) —E. (Intuitively ge, is necessarily the
right choice since it has the smallest possible value on SD, namely 0.)
1
= —v — Lui (log r)*dv — v(re i9)- rd0
ui r
2
Similarly,
= f v(rel°)
o dB = 27tv(0) 27rg0(Q,P).
The last corollary in IV.2.2 shows that the functions in F are subharmonic
on MVP.
Let K be any conformal disc on M\ {P}. Let D E 9. Choose D* e 3 such
that D* c D t.) K. Thus ga.(P,•) ge,(P,•) and ga.(P,•)lInt K is harmonic.
Similarly, if D i , D2 e 9, we can choose D E with D D 1 U D2 and observe
that gb(P,•) goi(P,•), j = 1, 2. Thus the family F is a Perron family. By
Perron's Principle
y(Q)= sup fgp(P,Q)1, Q E M \ (,11
DE 9
shows that 7(Q) < + cc, all Q e M \ {P } . The last inequality also shows that
y g(P,•) on MVP}.
Since y is a competitor for the Green's function (it clearly satisfies (3.6.2) and
(3.6.3)), y g(P,•) on M \{P}. We have thus shown that
7(Q) = g(P,Q) = sup {g D(P,Q)}, Q e M.
Deg,
Hence for Q, P e D
go (P,Q) = gp(Q,P) g(Q,P).
g(Q,P).
Reversing the roles of P and Q gives the opposite inequality. Thus the
proof of the theorem is complete.
172 IV Uniforrnization
EXERCISE
is a conformal mapping and g is the Green's function on N with singularity at f(P) for
some P e M, then g 0 f is the Green's function on M with singularity at P.
function for the domain D. Recall that u = f on SD and that Ag(P0 ,•) =
0 on D\ p301 while g(P,•)= 0 on SD. Thus
f * dg(P 0 ,• ) — r (u(z) * dg (13 G ,z) — g(P 0,z) * du(:)) = — g(PG,• )0.
Remarks
1. The theorem should be contrasted with Theorem II. 4.1 (and its com-
panions). What is important for us is not the more general singularity at P
that we can produce, but the boundedness statement (3.11.2).
1V.3. A Classi fication of Riemann Surfaces 173
then
il (grad u), 115_ cm
IP —
._
with c a universal constant.
PROOF. First: by grad u we mean the vector (u,,u,.), and by its norm we mean
max{1 14,cl, luyi }. Choose r > 0 such that the closed disc of radius r about P
is contained in D. The function u, is harmonic in D. Thus (without loss of
generality P = 0), by the mean-value property for harmonic functions
lux(0)1 =
ff izi <, ux dx dy
1
r2 — y2 , y) — u(— r2 — y2, y))dy
nr2
/nt 4m
< 2r =
icr trr
From the above we see that c = 4/n.
EXERCISE
Reprove the above lemma using the methods and result of Lemma IV. 3.12.
fax
PROOF. Without loss of generality we may assume (by adding a constant)
that u > 0 on M\K. Let z be the local coordinate corresponding to K so
that bK = = 11. Since
it suffices to assume that u is harmonic in fizi rl, r < 1 and to show that
=1
J!
Let g be the collection of all domains D such that K cc D cc M and bD
is analytic. Let up be the solution to the Dirichlet problem on CI(D\K,.) with
uo = ul5K, and uo iSD = 0, where K,. = {Izi< r}. Extend ur, to be zero
outside D. Let
= fuo ;D e
and
y = sup Up.
0E9
IV.3. A Classification of Riemann Surfaces 175
The reasoning in Theorem IV.3.10 shows .9" is a Perron family on IVACI K,.
For every D e up u; hence y < u. Now u — y is a bounded harmonic
function on M \CI K,. By Theorem IV.3.3, the maximum and minimum of
this function occur on 6K,. Thus u = y. We have shown that
u = lim u p = sup up on tizi > rl.
DE2 DE 2
Let cop be the solution to the Dirichlet problem on CI(D\K,) with w, 6K. = 1
and w0 15D = O. Then (as a special case of the previous argument)
lim cop = sup coD = 1 on izi > r.
DEge.Deg
-
Now note that by the reflection principle, up and cop have harmonic exten-
sions (which we do not use) across 6D. It thus follows that *du *chop
are defined and smooth on CI(D\K,.).
Now use the remark following Perrons principle (IV.2.6), and choose
sequences of domains A") c 2, {DIP} c g, such that
u = lim upço, 1 = lim cop(p,
=— LK ( *dup,_u*dwo .
We know that Re FP(z) = uP(z), p < Izi <R. Now both FP and f have
Laurent series expansions on p < Izi <R, say
Thus, taking real parts, we obtain (after changing the names of the constants
in the Fourier series) for z
. P
CCo co
UP(teie) — R e f(te) = — +
2
E
„=
(04t. + cos nO
CO
o (uP(ten — Re
f(ten)d0 =1f) .
1 j .?..c •
— o (uP(te) — Re f(te)) sin k0 dO = + fi t kt —k , k = 1, 2, ...
(3.16.3)
We first use each of the above equations for t = p (uP(pe) = Re f(pe i8)) to
obtain (k = 1,2,3, ...)
= 0,
Œk= _ p 2koc c, (3.16.4)
= P 2kfifc•
We let
rn = sup {I f(z)i ; = 1 } ,
m(p) = sup{luP(z)1; z = 1 } .
Using Equation (3.16.3) for t = 1, we obtain
'at + 2(m(p) + m),
Ifl + 13P_k i 2(m(p) + m).
Combining the above result with (3.16.4) we have:
1411 1 — P 211 5- 2(m(P) + m),
141 1 — P 2I 2 (m(P) + m);
and for p <
Xi 4(m(p) + m),
Ififl 4(m(p) + m).
11/.3. A Classification of Riemann Surfaces 177
We relabel {P1 , 1) 2 , P3 ,• -} as flPi 1 , P i2,1) 13.. ..1. Now we can choose a sub-
sequence fo n
tr 21 , r- 221r 23 , • • •I of {P1149 124'13, • • • } so that uP2k converges uni-
formly to u on tr Izi r2 1. By induction the sequence f. 1.3, • • -}
satiNfies
Pi- 1.i > >•'•,
P- 1.k < rk, k = 1, 2, ... ,
and
lim uPi -- 1.k = u on fi lzi
defines a meromorphic function with a pole (of order > 1) at P, and a zero
(of order 1) at P2.
M is elliptic or parabolic. Choose u; harmonic on AP, [Pil such that
14 = Re 1 + y(z)
zi
with y; harmonic in a neighborhood of zero. Define 9 as before. This time
ordp, 9 5 —2 and ordp, cp 2.
PROOF. Let
f be any non-constant meromorphic function on M. Let
{P 1 ,P2 , .} be the set of poles and critical values (those P E M with df(.1))=
0) of f. Let z; be a local coordinate vanishing at P i with the sets {14 < 1}
all disjoint. Let 0 < r, < r2 < 1 and let coi be a C function with 0 _5 coj 5 1,
o.); = 1 on {Izil r 1 } and a); = 0 on { Izil r2 }.
Define arc length ds by
ds2 = Idf12 (i — to) + E w) ldzi12.
Remark. We shall (see IV.8) be able to do much better. We shall show that the
metric may actually be chosen to have constant curvature.
Corollary 3. Every Riemann surface has a countable basis for its topology.
While it is possible to prove the last two corollaries at this point, we shall
delay the proof until after IV.5 when we will be able to give shorter proofs.
PROOF OF THEORFNI (The proof is almost shorter than the statement of the
theorem and has implicitly been established in the section on multivalued
functions.) Let c: [0,1] M be a continuous path with c(0) = Po . Let
0 = tt e [0,1]; ço can be continued along c, = cl [0,t] }.
It is easy to see that 0 is non-empty (0 E 0), open (without any hypothesis),
and closed (by use of (a) or (b)). Note also that the analytic continuation
satisfies (a) or (b).
Let c be a closed path beginning at P o . Let (p i be the continuation of ço
along c. when, assuming itypothesis, (a), we haye
log 191 = u = loglp i near Po .
Thus there is a z(e)E C, = 1, such that
(p i = x(c) (p near Po . (4.2.1)
Since (p i depends only on the homotopy class of c, x determines a homo-
morphism
x:ir l (M,P0) --■
from the fundamental group of M (based at P o) into the unit circle (viewed
as a multiplicative subgroup of C). Since S' is a commutative group, the
kernel of x contains all commutators and thus x determines a homomorphism
from the first homology group
H i(M) —> S 1 . (4.2.2)
In case (b), (4.2.1) is replaced by
(Pi = p + i(c)
with x(c)e R, and (4.2.2) is replaced by a homomorphism
CI
Remarks
1. The two sphere 5 2 is of course elliptic. If ,S 2 \D is a point, then (via a
Möbius transformation) D C. The plane C is parabolic. (Prove this
directly. This is also a consequence of Theorem IV.4.5.)
182 IV Li itiform izat i on
Then u is harmonic and >0 except at the (isolated) zeros of (p. Also
1
-1 °g1001
n + log ICI
is harmonic for ICI small. Let y e .F, where .F is the family of subharmonic
functions defined in the proof of Theorem IV.3.7, with P replaced by S
(and z by C). Let D be the support of y, and D o be D with small discs about
the (finite number of) singularities of u deleted such that u - y > 0 on (D\Do).
By the maximum principle u y on D o . Hence, on M\ {S}. We conclude that
1
- - logl(p(T)I = u(T) sup v(T) = g(S,T)
ve.f
= -loglf(S,T)1,
Or
IP(T)I If(S,T)I. (4.4.2)
Setting T = R, we get (by (4.4.2))
we conclude that
cp(T) = yf(S,T), y E C, IXI = 1.
We rewrite the above equation as
f(R,S) — f(R,T)
y f(S,T) = ' ,
1 — f(R,S)f(R,T)
and deduce that
f(R,S) f(R,T)44. f(S,T) = øS = T.
We have shown that f(P,•) is an injective holomorphic function of M into it
We now define
= h(f(Q)), Q 6 M.
Then F is a holomorphic function of M into LI with a simple zero at P and no
others. Thus —logIFI is a competing function for the Green's function and
we have
— g(P,Q) = —logif(P.Q)1,
Or
IRO if(Q)J.
Hence we conclude that
h(z)
1 for 1,71 small,
and
h'( 0)1 1
contradicting (4.4.3) El
g(z) = + a0 + a,(.-: — : 0) + • •
(— - -.0)
7./(Q) + 6 '
Setting Q = P0 , we see that y = 0 (we may thus take 6 = 1) and thus (because
we know the singularity of f and f at P0)1 = I. D
0 1, Q2 , - •
be a maximal set of non-equivalent components of Q. It is clear that (as point
sets) the orbit spaces
S-2/G and la)
IV.5.3. Since SL(2,C) inherits a topology from its imbedding into C4,
PL(2,C) S L(2,C)/ + I is a topological group. A group G c PL(2,C) is
called discrete if it is a discrete subset of the topological space PL(2,C).
It is obvious that
az + b ;
G = fzi— ■ ad — bc = land a,b,c,d eZ[i]
cz + d
Remarks
1. We will see in IV.6 that for most Riemann surfaces M, 11-4 U.
2. At this point, it is quite easy to give alternate proofs of Corollaries 2-4
in IV.3.18. A stronger form of Corollary 1 will be established in IV.8.
IV.5.7. If i = U, then the group G is, of course, a Fuchsian group. If A = C
or C u {co } , then G can have at most one limit point (G acts discontinuously
on C). Thus we have established the following
Remark. The theorem thus shows that compact surfaces are uniformized
by groups of the first kind.
Theorem
a. The only simply connected Riemann surfaces are the ones conformally
equivalent to C u {on}, C, or 4 = {z E C, IZI < 11.
b. The only surfaces with tr i(M) Z are (conformally equivalent to) C* =
CVO}, 4* = AVID}, or A, = {Z E C; r <Izi < 1 } , 0 < r < 1.
c. The only surfaces with tr i(M);-.--.. Z Z (the commutative free group on
two generators) are the tori C/G, where G is generated by z + 1 and
z + r, Im r > O.
d. For all other surfaces M, it,(M) is not abelian.
The remainder of this section is devoted to the proof of this theorem.
IV.6.3. Theorem. The only Riemann surface M which has as universal covering
the sphere, is the sphere itself.
PROOF. The covering group of M would necessarily have fixed points if
i(A1) { 1 }.
111 .6.4. The fixed point free elements in Aut C are of the form z Z + a,
a E C. Since a covering group of a Riemann surface must be discrete, we
see that (consult Ahlfors' book Complex Analysis) we have the following
Remark. Using Fuchsian groups, one can determine generators and relations
for fundamental groups of surfaces.
Thus C(z) = Az with A e R, A > 0, A o 1. Thus Aut U does not contain any
non-hyperbolic loxodromic elements, and hence the covering group of a
Riemann surface (whose universal covering space is U) consists only of
parabolic and hyperbolic elements.
[Aa Ab ] = [Aa b ] .
in
c d Ac d
IV.6. The Exceptional Riemann Surfaces 195
IV.6.7. Theorem. Let M be a Riemann surface with ir,(M) Z G31 Z, then the
holomorphic universal covering space of M is C.
PROOF. Assume the covering space is U. The covering group of M is an
abelian group on two generators. Let A be one of the generators. If A is
parabolic, then we may assume A(z) = z + 1. Let B be another free generator.
Since B commutes with A, B must also be parabolic. Thus B(z)= z +
fi E RAI The group generated by A and B is not discrete.
So assume A is hyperbolic (it cannot be anything else if it is not parabolic).
We may assume A(z) = Iz it> L The other generator B must be of the
,
form B(z) = pz with p > 1 and 2" # pm for all (n,m) E Z Z \ {O}. Again,
such a group cannot be discrete (take logarithms to transform to a problem
in discrete modules). D
IV.6.8. To finish the proof of Theorem IV.6.1. we must establish the following
is given by
• expi2ni.:).
For z = re', 0 < 0 < 27, we define
log z = log r + i0
(the principal branch of the logarithm). The covering map (for the case
of hyperbolic generator)
U —3 1.1,
is then given by
log :
f:mi-+ M2
IV.7. Two Problems on Moduli 197
11/.7.2. How many conformally distinct annuli are there? W have seen
that every annulus can be written as U/G where G is the group generated
by :i—z, A e > 1jf A l and A2 are two such annuli with the cor-
responding ,1 1 and A2 , when do they determirre the same conformal annulus?
They determine the same annulus if and only if there is an element T a Aut U
such that
T(2 i z) = A2T(z), z e U.
From this it follows rather easily that A i =A. (by direct examination or
using the fact that the trace of a Möbius transformation is invariant under
conjugation).
We conclude
Of course, we could substitute for the word "annuli" the words "domains
in C u apl with two non-degenerate boundary components."
•C
°i .
gll
' T2
The mappingf is conformal (thus affine: that is, z' --4 az + b), and it induces
an isomorphism
G1 g g e G2.
We abbreviate the Mtibius transformation z + c by c. Thus
a = 9(1) = al + [3'1. 2 with a, /3 e
(7.3.1)
at, = 0(r 1 ) = 71 + 6r2 with y, Se Z.
198 iv uraformization
2(z)idzi, z e M. (8.1.1)
We set
2
2(z) = , for M = C L.) {
1+
).(z) = 1, for M = C,
2
2(z) = 1 _ 1z12 , for M = = {z e C; izi < 1}.
The definition of A for M = C L.) {co} is, of course, only valid for z o.
At infinity, invariance leads to the form of 2 in terms of the local coordinate
= 1/z.
We will now explain each of the above metrics.
(here 4 is the Laplacian, not the unit disc). A calculation shows that
K= +1,
and that the area of C u {00 } in the metric is
4
Area(C u loop = ffc ).(z)2 dx dy = fo2n Sox' + r2)2 rdrd9 = 4n.
Proposition. An element T e Aut(C u {,,0 } ) is an isometry in the metric
(2/(1 + 1z1 2))1dzi if and only if as a matrix
[a +?
T= lar ic1 2 = 1' (8.2.2)
c 5
200 IV Uniformization
PROOF. Write
ra bl
ad — bc = 1.
Lc di'
Of course, T is an isometry if and only if
il(Tz)IT'(z)1 = ZEC u {X},
EXERCISE
Determine the geodesics of the metric of constant positive curvature on the sphere.
IV.8.3. The metric idzi on C needs little explanation. It is, of course, the
Euclidean metric. It has constant curvature zero. Geodesics are the straight
lines; and automorphisms of C of the form
z + b, 6 e R, b e C,
are isometries in the metric.
IV.8. Riemannian Metrics 201
IV.8.4. We turn now to the most interesting case: the unit disk 4. Let us
try to define a metric invariant under the full automorphism group Aut 4.
The metric should be conformally equivalent to the Euclidean metric; hence
of the form ,1.(z)Id4 Set 40) = 2 and note that ;1(A0)124'(0)1 = ).(0) for all
A e Aut 4 that fix 0 (since A(z)= e lez). Let z o e 4 be arbitrary and choose
A e Aut J such that A(0) = z o. Define (to have an invariant metric)
= 2144'(0)1 - 1 .
Note that A(z) = (z + z0)/(1 + 70z) (it suffices to consider only motions of
this form). From which we conclude that
._
2
1 2
A simple calculation shows that this metric is indeed invariant under Aut 4.
(In particular, the elements of Aut A are isometries in this metric.)
Before proceeding let us compute the formula for the metric on the upper
half plane U. Since we want the metric X(z)id:( to be invariant under Aut U,
we must require
X(z) = 2IA'(0)1 -1 ,
where A is a conformal map of 4 onto U taking 0 onto zo . A calculation
similar to our previous one shows that
1(z) = iml
Computations now show that the metric we have defined has constant
curvature — 1.
Let c: I —* A be a smooth path. Then the length of this path 1(c) is defined
by
1(c). sol .1.(c(t))1e(t)idt. (8.4.1)
We can now introduce a distance function on J by defining
d(a,b)= inf{/(c); c is a piecewise smooth path
joining a to b in J } . (8.4.2)
Take a = 0 and b = x, real, 0 x < 1, and compute the length 1(c) of an
arbitrary piecewise smooth path c joining 0 to x,
2 ,
1(c) = fo 1 _ ic(t)12 ic (oldt
2 Ça
> $o1 1 — =
c(t)2 t =
2
_ u2
1+x 1+x
= log log .
1—x 1—x
202 IV Uniformization
We conclude tint
+
d(0,.x) = log
1— x
and that c(t) = tx, 0 t < 1, is the unique geodesic joining 0 to x. We now
let 0 # b = zed be arbitrary (and keep a = 0). Since
e y
1 + IzI
d(0„-..) = log (8.4.3)
1 — lzr
and 00= I:, 0 t 1, is the unique geodesic joiniru0 to
Proposition
a. The distance induced by the metric 2: (1 !:i 2 );ii.:1 is ef,mplete.
b. The topology induced by this metric is equivalent to the usual topology on J.
c. The geodesics in the metric are the arcs of circles orthogonal to the unit
circle {z C; lzi = 1).
d. Given any two points a, b E .1, a 0 b, there is a unique geodesic between them.
PROOF. Note first that (8.4.31 shows that every Euclidean circle with center
at the origin is also a non-Euclidean circle and vice-versa. Also note that
. d(0,z)
lim
Izi
showing that at the origin the topologies agree. Since there is an isometry
taking an arbitrary point z e d to the origin, the topologies agree everywhere.
This establishes (b).
We have already shown that the unique geodesic between 0 and b is the
segement of the straight line joining these two points. Now let z i and z 2 be
arbitrary. Choose A E Aut d such that A(: 1 ) = 0. Let b = A(z 2). The geodesic
between z, and z2 is A 1 (c), where c is the portion of the diameter joining
0 to b. Since A' preserves circles and angles, (c) and (d) are verified.
Now let {z„} be a Cauchy sequence in the d-metric. Then {d(0,z)} is
bounded. Hence {z„) is also Cauchy in the Euclidean metric and tiza ll is
bounded away from 1. This completes the proof of the proposition.
1
Area D = ).(z) 2 1dz A dui
Case I. The triangle has one cusp at (see Figure IV.1), and one right angle.
The angles of the triangle are hence: 0, it/2, a. (Note we are not excluding the
possibility that x = 0; that is, a cusp at B.) The area of the triangle is then
dy a dx
dx Jo -- arc sin
- = 7r —
.or — P
Case 11. The triangle has a cusp at co. (See Figure IV.2.) It has angles 0, a, )3.
We are not excluding the possibility that either a = 0 or /3 = 0 or both
= fi = 0. Choose any point D on the arc AB and draw a geodesic "through
-P a
Figure 1V.2
204 IV Uniform ization
D and co." The area of the triangle is the sum of the areas of two triangles:
2
The case in which A and B lie on the same side of D is treated similarly.
Case III. If the triangle has a cusp, then it is equivalent under Aut U to the
one treated in Case II.
Case IV. The triangle has no cusps. (See Figure IV.3.) Extend the geodesic
AB until it intersects at D. Join C to D by a non-Euclidean straight line.
The area in question is the difference between the areas of two triangles and
thus equals
Figure IV.3
Remarks
I. If a Riemann surface M carries a Riemannian metric, then the metric can
be used to define an element of area dA and a curvature K. The Gauss—
Bonnet formula gives
j=1 )
v
struction that is very useful in studying function groups. (These are groups
that have an invariant component.) Let
p:b *1)
be the holornorphic universal covering map of D and define
F {yeAutb;poy=gopforsomegeGI.
It is now easy to check that 5fr DIG (thus G is discontinuous if and only
if F is). Furthermore, define a homomorphism
p*:r --+ G
by
p 0 y = p *( 7) 0 p, y E F,
and let H = Kernel p*. Then
{1} —J1 F G -+ {1 }
1
and (n > 0) since
1 1
- 5 1 — — < 1,
2 v
we conclude that
n > 1, n < 4.
Thus n = 2 or n = 3. If n = 2, then
21 1
= y1 ;72:
1 12
1.
2 2 N
121
Thus
11
— <— or 6. > v3.
6 1, 3"
Thus v 3 = 3,4 or 5.
We summarize below the signatures and cardinalities of the groups G
that could possibly act on Cu { x }.
Signature of G IG1
(O;—) 1
(0; v,v) y (25v<co)
t0;2.2,v) 2' (2 < x)
(0;2,3,3) 12
(0;2,3,4) 24
(0:2,3,5) 60
IV.9.4. Before proceeding to the next special case, we must establish the
following
(K — 1)(K2 — 1)
C(z) z +
K
which is parabolic.
EXERCISE
Let M be a Riemann surface whose universal co% eriug space is the unit disk and let
P E M. Then the stabili7er of P in any subgroup of Aut A/ is finite. Is the assertion true
for other Riematin surfaces?
IV.9.5. Next we determine the groups G that can possibly act discon-
tinuously on C. Since G leaves C invariant, G c Aut C, and the elements g
of G are of the form
+ b. (9.5.1)
Observe that a, if 01, must be a root of unity, since a is the multiplier of
g and g must in this case be elliptic of finite order. Map the element g e G
of (9.5.1) onto a e 5'. This map is a homomorphism since
blr 131 r aa nI3 +
LO i][0 1 j = 1 j.
Let Go be the kernel of this homomorphism. Clearly G o is the unique maximal
fixed point free (normal) subgroup of G.
If Go is trivial, then by the preceding lemma, all elements of G must
fix a common point : 0 e C. We conclude that G is a finite cyclic group
acting discontinuously on C t.i {co}. We will not consider such groups to
be acting on C.
We consider now the case with Go non-trivial. It is then easily seen that
Go is a free group on 1 or 2 generators.
Assume that G o is cyclic. Then (without loss of generality G o is generated
by zi-+ z + 1) C, G0 is equivalent to the punctured plane C* (via the map
eriz), and G,'G o acts as a group of conformal automorphisms of C/G 0.
The automorphisms of C* are of the form zl-- ■ kz and k/z (see V.4.3).
The former lift to automorphisms of C of the form (zi—oz + (log k)/2ni) and
must therefore be in Go ; the latter, to automorphisms of the form zi-0 —z +
(log k)/2ni whose square must be in Go . Thus G/Go is trivial or isomorphic
to Z2. In the second case, G consists of mappings of the form zi— ■ ±z + n,
n e Z.
We "draw" fundamental domains for the groups encountered above.
G generated by zi— ■ z + 1 (Figure IV.4)
m z =0
Re z=0 Re z I
Figure IV.4. Signature (0;oo,co).
IV.9. Discontinuous Groups and Branched Coverings 211
/////
•
•/./ •
• z=0
z=
ReizO Re z=1
Figure 1V.5. Signature (0:2,2,x)) wheie the first two twos correspond to the points
z=0andz=1.
We now consider the case where G o has rank 2. Then we may assume
that G o is generated by the two translations z z + 1, z 1-• z + 2, 1m r > 0,
and C/G o is a torus. Without loss of generality, we may assume IT! 1.
Again G:Go acts as a group of automorphisms of C/Go .
Assume now that G/Go is non-trivial. Let z + 1 and h:zi- ■ az + b
be parabolic and elliptic elements of G. We conclude that
hog=11':z1-+:+a
is an element of G o , or that the multipliers of elements of G are periods of
Go. Since there are only finitely many periods on the unit circle, we can
find a "primitive" multiplier K = e24' " with p c Z, i> 1 and p maximal.
By conjugating G once again (by an element zi - + c) by a conjugation
that does not destroy the previous normalization, we may assume that G
contains the element z Kz. Hence, we see that G/G o is a finite cyclic group
4. The group G hence consists of motions of the form
zi-■ z + n + MT, V = 0, , - 1, n Z, e Z.
Now we let y be the genus of C/G(=(C/G0)/(G/G 0 )). Let x, = 0, x 2 , .. • ,
be the fixed points of G/Go on C/Go and let v l , , V,. be the orders of the
respective stability subgroups. We use Riemann-Hurwitz (see V(1.3.1))
1
0 = 2y - 2 + (1 - - ), (9.5.2)
vi
and conclude that if r > 0 (G/Go non-trivial) then y = O. Thus (9.5.2) becomes
2= - -1 ).
vi
Let N be the order of GIGO . Then
2 vi N.
Since 1 > (1 - 1/y;) r = 3 or 4. If r = 4, then y, = v 2 = v 3 = v 4 = 2,
and the ramification is completely accounted for by the element - z,
and G/Go
212 IV Uniformizatiou
Im z - 0
Re Z=0
Figure 1V.6. Signature (1;—).
Z=T
,
T /
Z = ,
/ç -;
1m z =0
1 ' -2 z
Re z =0
Figure IV.7. Signature (02,2,2,2) where the twos correspond to the points z = 0, z =
Z = r/2, z = (T + 1) 72.
V1 V2 V3 G/G0
2 3 6 16
2 4 4 14
3 3 3 13
K = n + MT.
Im z = 0
Re z=0
Figure IV.8. Signature (02,3.6) where the two Corresponds to the point z = the
three corresponds to the point z = (K + 1)/3 and the six corresponds to the point = 0.
Fundamental domain _An- {z 1—* K": + n + niK} K = e 2 ,a14 = i (Figure IV.9)
Im z =0
Re Z=0
Figure IV.9. Signature (0;2,4,4) where the two corresponds to z = I, one four corre-
sponds to z = (1 + 0/2 and the last four corresponds to z = 0.
Fundamental domain for {zi—+ + n + mK}, K = e2 ' 1" 3 (Figure IV.10)
Im z =0
Re Z =0
Figure IV.I0. Signature (0;3,3,3) where the threes correspond to the points z = 0,
z = (K + 2)/3, z = (2K + 1) 73. (Each of these three points is equivalent to a corre-
sponding point in the figure.)
Remark. We have established above the existence and uniqueness of all the
elementary groups with one limit point. We have also determined all pos-
sible signatures that yield zero characteristic.
214 IV Uniformization
Lemma. Let G he a Fitehsian group operating on the upper half plane U. Assume
that P is a puncture on U/G, and that the natural projection it is unratnified
over a deleted neighborhood of P. Then there exists a deleted neighborhood
D of P in U/G, a disc b in U, and a parabolic element T e G such that D
.5/G 0 , where G o is the cyclic group generated by T, and two points in b are
equivalent under G if and only if they are equivalent under G o .
G0 = G; = 151.
IV.9.8. To establish the converse of the above lemma, and to show that
D can always be chosen to be the half plane ft z e C; 1m:> I} we prove the
following
(in particular, the sequence A„) is distinct). Now we choose 7 so that 0 <
(1 — < y and lai < y. By choice, la c)! =la! < y. Suppose Ian; <y for
some n, then also
+ = 11 — anc„I 1 + Ian' Ic 1 = I + I a . . I Icl'
< 1 ± ylc1 2" < 1 + yid < y.
Hence, by induction
la,,I < y for all n.
From the relation
an+ , = 1 — a„c„
and (9.8.1), we conclude that
lim a„ = 1.
Thus also
1 1 m b„= Jim = 1,
11 -.00
and
lira d = Ern (1 + a„_ l c._ 1 ) = 1.
,
n -4 co co
216 Iv uniformization
W.9.9. Lemma. Let G he a Fuchsian group operating on the upper half plane
U. Assume that G contains a parabolic element T with fixed point x E t'
Then G„, the stabilizer fx in G is infinite cyclic.
PROOF. We claim that every element A e Gx is parabolic. Without loss of
generality x = co. Assume that G œ contains an element A with another
fixed point which may be taken to be the point 1 (since it must be on the real
axis because A cannot be elliptic). Thus we have T(z) = z + a, a e tR, and
A(z) = Œ(z — 1) + 1, a e R, a > O. Without loss of generality we may take
0 < a < 1 (replace ,4 by A - 1 , if necessary). Now A(z) = a"(z — 1) + 1, and
hence
A„ o T o A; '(z) = z + aan.
Thus we hive cousu wied a distinct sequence in G approaching the identity.
We have shown that every element of G, is of the form z z + a. Taking,
the minimum of all such positive a's, we obtain a generator for G. 0
i. u (z)idzi
iz — T.1
for U which is clearly invariant under G (since G c Aut U). Thus we may
project ).[; (z)idzi to U/G and obtain a metric with singularities at the images
of the elliptic fixed points. We want to show that these singularities are not
too bad in the sense that the Poincare metric on U/G is locally square inte-
grable at these points and that the same is true even in a deleted neighborhood
of a puncture. Let
n:U U/G
be the canonical projection. Denote the Poincaré metric on U G by ;.(Z)IdZI
in terms of the local coordinate Z. If Z can be expressed as a function of
z e U, then of course, the relation
I(Z)IdZi = 4(z)idzi
allows us to solve for A
- Z).
Let us assume that .70 e U is an elliptic fixed point. By conjugation we
may replace U by Li and assume that zo = O. The stabilizer G o is then gen-
erated by an element of the form z e21"1"z, for some v e Z, y > 2. Thus
Z = z' is a local coordinate at n(0). Since
2
A,(z) —
(1 —
we see that
21Z1(1/")-1
.T.(Z) =
v(1 —
Since (1/v) — 1 > —1, ;7(Z) 2 IdZ A d.ZI is integrable in a sufficiently small
neighborhood of Z = O.
It remains to investigate a neighborhood of a puncture. Here it is more
convenient to assume that G acts on U, and that the puncture corresponds
218 IV Uniformization
IZI log
Area(UfG) = — yk)
k= 1
co
= nc 2 — E bk,
k= 1
commutes. Let
M' = Ix e M; .(x) I = 1 }
= Ix e M; i ool = 11.
Let U' = ir-1(M) and LP, = 7r1-1 (M). Choose g e U', e such that
n(g) = niCg = b e M'. Let r = ir*(iri(U',E)) and ri= ri(ri(triSi)). It is
quite easy to see that F1 F (and hence F l = F). From this it follows that
it and ir i are equivalent coverings. El
222 IV Uniformization
Remark. The above theorems also establish the existence of finite Kleinian
groups.
EXERCISE
Prove that the theorem is not true for the two excluded cases.
E + i)
2( J=
by Theorem IV.3.11. Not all the functions u e H have single valued harmonic
conjugates. Let c; be a small circle around P. Notice that
f. *du = 0
224 Iv Uniformization
Ecxj + 1 — g].
2[ i =
Thus the dimension over C of the space of meromorphic functions on M
with poles of order <al at Pi is
ai + 1— g.
sj + 1— g
J=1
= deg D + 1 — g.
D = fl ro(P).
Pem
Remark. The above is certainly a shorter and more elegant proof of the
Riemann-Roch theorem than the one in 111.4. The earlier proof was, in our
opinion, more transparent than this one.
N 4-1 )n , aN O. (11.1.2)
n=
If there are infinitely many negative integers n with a„ 0, then .97 has an
essential analytic singularity at O. Let us assume that the resulting F is
meromorphic.
./(:)= E
n= N
k. 01.2.2)
E
n=N
anEn Znik
We exclude the constant series (a) = a0) from ,-/1*. In the above, k E Z, k > 1.
If k> 1 we assume that there exists an n> 0 with a„ 0 and n/k Z (this
can always be achieved), and that k has been chosen as small as possible
with this property. We call such a k the ramification index of co, ram co.
IV.11. Algebraic Function Fields in One Variable 1 29
The space di* comes equipped with two functions (into C {co}):
..//* P" ■ C u{x}
eval
Cuf }
if co is given by (11.3.1),
proj co = 1:0
ao- co if co is given by (11.3.2),
oo if N < 0,
eval 0) —{tio if N =-. 0,
0 if N > O.
Before proceeding, we must introduce an equivalence relation on de. We
shall say that o 0 . //* is equivalent to (33 e de (given by (11.3.1) or (11.3.2)
with coefficients rin, Y o, provided
proj = proj Co, ram co = ram Co = k,
and there exists an E E C with ek = 1 such that
= n = N, N + 1, . .
From now on ../t* will stand for the previously defined object with the same
symbol modulo this equivalence relation. Note that ram, proj, and even
eval are well defined on de (the new di* !).
We topologize de as follows. Let co c, be given by (11.3.1). Assume that
the series coverges for If — Zol < a. Let 0 < r < a. We define a neighborhood
U(coo ,r) of coo . Let zi e C with 0 < Iz i — zol < r. The "multivalued" function
(Z — z 0) 1 3' determines k single valued analytic functions converging in
{lz — z i I < a — r}—thus k distinct Taylor series in (z — z 1 ). Substituting
these Taylor series into (11.3.1) we obtain k convergent Laurent series. The
collection of all such Laurent series (for all z 1 with 0 <1z, — zo l < r) plus
the original w o forms the neighborhood U(coo ,r) of coo . Next assume that
coo is given by (11.3.2) where the series converges for Izi > 1/e. Choose, again,
O < r < E. For Iz i l > 1/r, the "multivalued" function z -13` determines k
convergent Taylor series in (z — z 1 ) with radius of convergence tir — Va.
Substitute these functions into (11.3.2) and proceed as before.
We see that with this "topology" (we will show next that the sets we have
introduced do form a basis for a topology), a deleted neighborhood of any
point coo E de can be chosen to consist only of (.0 E .,/i * with proj w 0 00,
eval ai co, and ram w = 1.
Let CO E U(ai 1 ,r 1 ) n U(w 2 ,r2). Assume that proj w = zo, proj w 1 = z 1 , and
proj co 2 = z2 all belong to C, and that co is given by (11.3.1). We must find
an r> 0 such that U(ca,r) c U(ah,r,) n U(w 2 ,r2). Let us first assume that
230 IV Uniformization
EXERCISE
conformal involution on JP'. lise this fact to prove that the analytic configuration of a
function and its inverse are conformally equivalent.
= E como.) ;(z)
‘,/:;
a(z) = (— E (0„,(440,,2(z) ••• co„„(z)
re, nv = 1
n i < n2 <• • •<nv
PROOF. Let (p be the map of the previous theorem. Since the image of tp
is a compact component of .*(C u { xp}), z, w satisfy an algebraic equation
by Theorem IV.11.4. 0
Remark. The mapping 4, of Theorem IV.11.5 is one-to-one if, for example,
the pair z, w separate points on M. The converse is not true. We begin with a
Definition. We say that z, w E X(M)\C form a primitive pair provided the
map cp of Theorem IV. 11.5 is one-to-one.
IV.11.7. Proposition. If M is a compact Riemann surface and z E JC(MK,
then there exists w E ,r(M)\C such that z, w form a primitive pair.
PROOF. Let n = deg z. Choose a e C such that z -1 (a) consists of n distinct
points x l , , x„. By the Riemann inequality, there exists a non-constant
meromorphic function wi on M such that
w; has a pole of order v; 1 at xi, and
w; is holomorphic on Mqx;}.
234 IV Uniformization
fiXk) = E
;=o
aiMw(xk)i, k = 1, . , n.
Note that
det W.
j = 0, , n — 1,
ajg) = det W'
IV.11. Algebraic Function Fields in One Variable 235
where
w(x ••
w(x•,) •••
w(x) w(x„r 1
is the Vandermonde determ*nant (and thus det W = 11„ i (w(.x;)— w(xk))
and 1,17; is the matrix obtained by substituting the column
for the (j + 1)-st column of W. The functions ai(C) are independent of the
ordering of the x1. They are defined except for finitely many e C. It is clear
that thy extend to nterornorphic functions of :- -thus rational functions.
Equation (11.10.1) holds now for all but a finite number of points. Hence,
everywhere by continuity.
Corollary. If 11 is a compact suffice and z , w is a primitire pair as above,
satisfying the irreducible equation P, then
11M) C(rgivl/P(Z.Iv).
PROOF. The last equality means that f(M) is isomorphic (as a field) to an
f=
o
Thus f defines an element of ,if(m). Since P(z,w) goes to the zero element,
of AIM), this mapping factors to a mapping from C(z)[w] into AIM). This
mapping is surjective by the proposition, and injective because the domain
is a field and the mapping is non-trivial. (:3
IV.11.11. Example. Consider a hyperelliptic surface M of genus g > 0. Let
z E X(M) with deg z = 2. We have seen (1 11.7) that z has 2g + 2 branch
points. Now Proposition IV.11.7 shows that there exists a w such that the
pair z and w are primitive. By Proposition IV.11.6, z and w satisfy a poly-
nomial equation
w2 — 2aw + c = 0,
where a, c E C(Z). Completing the square in the usual manner, we rewrite
the above as
(w — a)2 + c — a2 = 0.
We make now a birational transformation
(w — a) = w,, z = z1,
and obtain
WI2 = a2 — c.
236 IV Uniformization
a2 - C b2 n (: - 0,
1=1
with distinct ei e C and b E C(z). We now define one more birational
transformation
11,7, =
b
and obtain the equation (dropping the subscript 2)
w2 - 1-1 (:: -
j= 1
Since all our birational transformations kept = fixed, the complex numbers
ej can, of course, be identified. They are the finite branch values of Thus
r = 2g +1 f J.:, is a branch value, and r 2g + 2 if Y.: is not a branch value.
We have now reproven Proposition 111.7.4 without the use of "paste
and scissors".
1 — g > 1,
D
and
- = deg D — g > O.
r (P
and set
ho
fo = + h i.
ho(Po)
rf= fl of
f -- 1, ,Y ; E C,
f - 13;
atui thus.
v(r o f) = > v(f
;=o
E vcr —
j0
By Lemma IV.11.13, v(f — A) = 0 unless ). = 0, in which case v(f — =
v(f ) = 1. This establishes (11.15.2). Let 13 1 , , P. be the zeros of f listed
according to their multiplicities. Let F E It(M) be arbitrary. By Propositions
1V.11.6 and IV.11.9, F satisfies an equation
n-
F" + E ri( f)Fi = 0, with r; rational functions.
;=o
Thus, by (11.15.2)
nv(F) min tordo r; + jv(F), j = 0, , n — 1).
If v(F) <O, then ordo r; <O for some j = 0, , n — 1. Since r;(0) is a sym-
metric function of the values of F at the points P1 , , P„, we see that F
must have a pole at one of these points. Similarly, if v(F) > 0, then v(F -1 ) < 0
and F must vanish at one of these points. We conclude that v(F) = 0 when-
ever 0 0 F(P;) 0 co, j = 1, . . . , n. From now on we assume (without loss of
generality) that (P 1 , ,P} are the distinct zeros of f. (We are changing the
meaning of the symbol n.) Let h be a function holomorphic and non-zero
at P 1 ,. , P„. Further, choose an h that separates these points with dh not
zero at these points. We have v(h) = O. Consider the function
D. (h — h (p))°rdp, f
IV.11. Algebraic Function Fidds in One Variable 239
FIT= (h _
so-
)j (pordp; F
and
= 0.
This contradiction establishes the continuity of F*.
We show next that F* is holoinorphic. Let x e M be arbitrary. Choose
an f E .f(N) such that f is univalent in a disc U about F*(x). Clearly Ff C.
There is thus a disc V about x such that
F*(V) c U and (Ff)(V) c f(U).
Corolla, y 1
a. Let F:.-K(M)-- ■ 11M) be the identity, then F*:M 114 is also the identity.
b. If dr(M i ) 11: .Y((:t 2)L: A/(M 3 ) are C-algebra homomorphisms, then
(F2 o F 1 )* = F o F.
V.1.2. We assume now that g >_ 2. Let W(M) be the (finite) set of Weier-
strass points on M. Recall that W(M) consists of at least 2g + 2 points and
precisely 2g + 2 points if and only if M is hyperelliptic.
V.1.3. Having seen that Aut M is a finite group, we naturally want to get
a bound on its order.
N 84(g — 1).
PROOF. Consider (recall the discussion in 111.7.8) the holomorphic projection
(abbreviate Aut M by G)
-+ M/G.
V. t. Hurwitz's Theorem 243
Note that v; > 2 and thus < 1 — 1/v; < 1. The rest of the proof consists of
an analysis of (1.3.4 It is clear (since we may assume that N > 1) that g > y.
We consider possibilities:
Case I: y 2.
In this case we obtain from (1.3.1) that
2g — 2 > 2N or N g — 1.
Case II: y = 1.
In this case (1.3.1) becomes
1
2g — 2 = N E (i - - ). (1.3.2)
If r = 0, then also g = 1 (we assumed g > 1). This is the basic fact (the left
hand side of (1.3.1) is 2) that will be used repeatedly. Thus (1.3.2) implies
2g — 2 IN or N 4(g — 1).
Case III: y = 0.
We rewrite (1.3.1) as
1
2(g — 1) = N( E (i _-_)_
vi
2 (1.3.3)
Clearly v 3 > 3 (otherwise the right hand side of (1.3.3) is negative). Further-
more, v 2 3. 11v3 7, then (1.3.3) yields
2(g — 1) + + 2) or N 84(g — 1).
EXERCISE
We outline below an alternate proof of Hurwitz's theorem.
(1) Represent M as U/F where U is the upper half plane and F is a fixed point free
Fuchsian group.
(2) Show that N(F)=-- normalizer of F in A ut U is Fuchsian and that A ut M N(F)/F.
(3) Use the fact that it: u/r U/N(F) is holomorphic to conclude that N(F) is of
finite type over U.
(4) Observe that
deg it = [N(r):1],
where deg it = degree of 7C, and [N(F): = the index of F in N(T).
(5) Show that
Area(U/T)
— [N(T):F].
Area(U/N(F))
V.I. Hurwitz's Theorem 245
(6) Prove that for any Fuchsian group F of finite type over U, we have
Area(U/F) H
it .
Remark. The fact that f is invariant under T shows that T has finite order.
V.1.5. In this section, M is a compact surface of genus g 2.
Clearly these two are fixed under T. At infinity (7, = 1/z is a good local
coordinate, and the function elements over oo are
be the canonical projection, and let y be the genus of .11 The Riemann—
Hurwitz formula yields
2g — 2 = n(27 — 2) + (n — 1 )v.(T). (1.7.1)
Let PE M be a fixed point of T and is it(P) E /171. There is anon-constant
function f Ar(k) such that fis holomorphic on :1-4- \ {P } and Jhas a pole
at P of order Sy + 1. Let f = f o IC f is the lift of Ito Af). Then f e
(
f is holomorphic on M\ `A, and ordp f n(y + 1). We now use (1.7.1) and
the hypothesis v(T) > 4 to conclude
2g — 2 -= n(2y — 2) + (n — 1)v(T) > n(2y — 2) + 4(n — 1),
Or
g + 1 > n(y + 1).
It follows that P is a Weierstrass point on M.
V.1.4 that f; = fi o T for each j and thus T(Pi) = P. We have thus verified
the first part of (1.8.1). In particular, v(T) > v(T). We rewrite
g> n26. + (n — 1)2
in the form (add (n — 1)g to both sides and simplify)
g — ng g + n — 1
(1.8.5)
n— 1 > n
We continue with
n—1
20 n— 1 2g
= 2 +-- +-2-- > 2 +—.
n
(The first strict inequality being a consequence of (1.8.5).) By Proposition
V.1.5,
2g
v(T) 2+
ord T —
and hence ord T n, finishing the proof of (1.8.1).
Assume (ird T = ord T. 13kith T and 1' are in the stability subgroup (of
Aut M) of P 1 . Since this stability subgroup is a cyclic rotation group
(Corollary to Propoition 111.7.7), it is clear that <T> = <D>.
Next, let B e Aut M and set C = To B. Then ord C = ord T; and,
by the inequality on g, v(C) = v(D) = 2 + 2(g — n),/(n — 1) > 2n(j + 1).
Thus by (1.8.2) C e <t> or < t> is normal in Aut M. We have established
(1.8.3) and hence also (1.8.4).
(with Gp, the stability subgroup of G at P e M), with a similar formula for
Bi. Now if Gp is nontrivial, it is cyclic and generated by some T e G. Clearly
V.I. Hurwitz's Theorem 251
B = X B.
PROOF. Let y = y(T) and use Riemann-Hurwitz with respect to the projec-
tion 7r:M = 114/<T> onto a surface of genus g
2g - 2 = n(24 -2) + v(n - 1). (1.11.1)
Assume n > 2g.
If g 2, then n(2g - 2) 2n 4g; a contradiction.
If g= 1, then v 0 0 and v(n - 1) 2g -1: again a contradiction.
If g = 0, then v > 3. Assume first that v > 4, then -2n + v(n - 1) >
2n - 4 > 4g - 4; a contradiction.
If y = 3, then (1.11.1) reads 2g - 2 = n - 3 or n = 2g + 1.
Remark. Proposition V.1.11 has shown that the maximal prime order of
an automorphism is 2g + 1 and that in this case the automorphism T has
3 fixed points and Mg T> is the Riemann sphere. Examples of such Riemann
surfaces are those defined by the equations
w29+ 1 = (Z — e i )(z - e 2)22(z - e3)2',
where e l, e2 , e3 are distinct complex numbers and a,, a 2, 2 3 are positive
integers such that a, + a2 + a, 0 mod(2g + 1), and 2g + 1 is a prime
252 V Autornorphisms of Compact Surfaces--Elementary Thcory
(M) = ..rfq(M)
q=o
is a commutative graded algebra.
Let T e Aut M and 9 e )t‘g(M), then T acts on 9 to produce
Tp = 9 0 T -1 .
(Say that 9 in terms of the local coordinate z vanishing at P e M is given by
p(z)dz. Choose a local coordinate vanishing at TP. Assume that in terms
of these local coordinates T . ' is given by z = f( . Then Ty) in terms of the
)
T(TIY)2)= (T(PI)(T<P2).
d. For 0 9 e Yeq(M), Te Aut M, P e M,
ord Tp T9 = ordp <p.
V.2. Representations of the Automorphism Group on Spaces of Differentials 253
= (z e t ) • (z — e 2g , 2 ),
dz
j = 1, , g.
from which it follows clearly that J9 = 9 for all g, e .1( 2(M). For g 2 and
g > 2 even, we consider
(de )=
wq -t wq -t•
Proposition. For q 1
Yft(M) Yeq(A 71.,Z (q)). (2.2.2)
PROOF'. The proof is analogous to the one given in 111.5.2. We use the iso-
morphism of the proposition and the Riemann-Roch theorem. Let Z he a
canonical divisor on 11-4-. It is easy to see that
ZN )
L(
where is a q-differentfal and (9) = Zq. Now for g = 1, Z11) = 1 and thus
Jr'oa,z"))= ye9'(171)= the g-dimensional space of abelian differentials of
the first kind on M. For g > 2, we compute using Riemann-Roch
= deg( g + 1 + t --
Po) Z (q'
1 Zq
= g(2ff — 2) + E [g(1 — — — 1) +
PE sf v(P)
Thus to complete the proof of the corollary, it suffices to show that
i(Zq/ZN )) = 0 which follows from the inequality deg(11,,Z 41) > 2g — 2. We
have already used the fact that
1
deg(-K g—) = q(24 — 2) ± E [41 —
Z(q ) PE
z. -
If g = 1, then we saw in the proof of Hurwitz's theorem that for some P e
(2.2.5)
v(P) 2, thus the sum in (2.2.5) is greater than or equal to [q/2] I. For
g = 0, we must show that the sum in (2.2.5) exceeds 2(q — 1). Now
z Fq ( i i1 z fq q
PER L v(P)A p la 1 v(P) v(P)
1
= (q l) PL (1 — ; j) .
But we saw in the proof of Hurwitz's theorem (equation (1.3.3)) that
Ep Esr, — (1/v(P))) > 2.
Remark. Let g = 2 and let n = the number of points 1 e la with v(P) > 1,
then
dim yew = 3g — 3 +n 0).
256 Y Automorphisms of Compact Surfaces—Elementary Theory
Proposition. There exists a basis for ..)(4 (M ) such that the action of T on
.Yt'l(M) is represented by a diagonal matrix.
PROOF. Since Aut M is a finite group, there is an integer n such that T" = 1.
Thus (because Aut M is represented on Yeq(M)) T" = 1 as a matrix with
respect to any basis for g'q(M). In particular, A." — 1 is an annihilating
polynomial for T and the minimal polynomial (a factor of A" — 1) has distinct
roots. Thus, by a well known result of linear algebra, T can be diagonalized.
We can say something more. Each entry on the diagonal must be an n-th
root of unity. We will continue with this analysis in the next section.
Remark. We shall actu:.!ly be able (in most cases) to exhibit a basis for
,Yt'q( M) with respect to which T is diagonal.
Since EAa dim EA= dim AIM) and (g — 1) = n(g — 1), we see that all the
n-th roots of unity must be eieenvalues and that their multiplicities are
given by the above formulae.
V.2.5. We extend next the considerations of the previous section to the
general case (G = is assumed to have fixed points). The development
given below is due to I. Guerrero.
To fix notation (in addition to what was introduced in V.2.4), we let
ni = dim Eo , j = 0, , n — 1.
Formulae (2.2.3) and (2.2.4) compute n,. To compute n j for 1 j — 1,
we partition the branch vet X of 7r.; M M/G into a disjoint union
n-1
X = U X i,
1=1
where
X, = {P e M; = P and Tk P P for 0 < k 51— 1}.
We claim that if Xi 0 0, then /In (1 divides n). Suppose P e X, and 2 5 1
n — 1. Write n = pl + k, 0 < k < 1 — 1. Then VT = P, and thus k = O.
Hence we see 11n. Furthermore, the orbit of each point P e X, consists of
the 1 points P, TP. . , T''P and are also in Xi. It is thus clear that if for
non-empty X i , we set
n(X I ) = {Q.1; 1 x i}, 1 < 1 < n — 1,
then n(X,) consists of x, distinct points and X, consists of lx, points. We
define x, to be zero for empty X,. We can choose a local coordinate z for
each point in it -1 (Q„„) such that is given by
where 1, is a primitive nfi-root of unity (note that n,,,, depends only on Q„,,
and not on the choice of P E
Suppose now that 0 o (po e E„, I <j n — 1. As before, in this case we
also have
n = r(-
1 ),
Di
9= E safki)dzq.
k=0
258 y Automorphisrns of Compact Surfaces—Elementary Theory
.511,111.i = X .144-7k .
k= 0 k o
In particular, for each k = 0, 1, 2, ... ,
sik1nti q - = 0.
Choose the unique integer such that
1 < 41j < n11 and q1 , =
Note that /1„,1(n ,1) = n/I (this defines ).„,i„) and we set (for further use) ;.„,to = 0.
We conclude that
=0 unless k + q ).„„.*mod - .
I
n)
Let xm, be the order of rp o at ir -1 (Q„,,) (this integer only depends on Q„,,
and not on the point P we choose in its preimage). By our previous remarks
there is an integer /3„,,i such that
n
ml = /3?NU• - +-
.n q 0
(note that /3„,u 0).
Let Q 1 , .. , Q, be the projections to M/G of the zeros of cp o not in X. Let
= ord„, (00 ço o . We note that a function _re /G) is such that fcp c, e
drq(M), where f = o rr, if and only if
j. / is holomorphic except at the points Qk, Q„,,,
ordok -xk, and
iii. ord/ + (q - )..01014
Now we know how to define the divisor Di (which depends on To):
D = n 421,k Fl 11
k= I m= 1
QAmi . +Lams' — 9110/11]
ml •
Remark. With our convention for ] = 0, the formula for D valid for all j,
0 < j < n - 1.
We return to (2.5.3) and continue our calculation (using (2.5.2) and Riemanti-
Hurwitz)
xi „ q])
deeD; = E + E n/1 n11
k= 1 lin m= 1
v. " —1
(n 2k + ) L
n s =t lin m=1 lin m=1 nit
1 1 1
= ;i 2q(g ;77 1, -r-1 (20
lin
1
+ (2g — 2)(q — 1) = 2g— 2 + !(2 — 2)(q — 1)
>. 14. — 2.
D= fl 01/4 H
k=1
II
lin m= 1
The lift (p of gi will have a zero of order ak at each preimage of Qk, and a
zero of order
a n (n
P .a.; + - = 1 2,21
1 /
at each point of 7t -1 (Q.). Thus 9 will have the same divisor as </Jo. But this
means that (p is a constant multiple of (po . Since 9 is G-invariant, q, e E 1 .
Thiscontradicts the fact that T o e Ed, 1 < n — 1.
We have concluded that if Ed is non-trivial, then it has dimension
r(1/D;) = deg I) + 1 — g, except for the case q = 1,j = 0, and no fixed points.
Let us turn to the case q = 1. Note that for j > 0, [().„,u — 1)1(n11)] = O.
260 v A utomorphistns of Compact Surfaces--Elementary Theory
deg Di - E E 1 (Anii - 1 ).
n lin m=1
r(-)= deg Di + 1 - g= 4 - 1 + - E (n -
Di n lin rn ,-1
XI
=g- i
—
E
Iii, m=1
1 ;•ntlj• (2.5.4)
We compute next
Recall that n, is a primitive (n//)-root of unity and that ii,trj efi. For
j = 1, 2, ... , (n11), WI is the set of all (n//)-roots of unity. We thus conclude
that
1-'4 1 mtj
is a permutation of 11, ... ,(n/01 with o-(nI1) = (n/1). Further for j = k(n11) +
r, Amu = Therefore
j=0 ii
1 x, (1 1
= n(4 - 1) + 1 + (n - 1) E x, - - y E - ni + - n2
tin n lin m=1 2 2
1
= n(g - 1) + 1 +(n-1):Ci
L
= 9.
We finally conclude that once again for each j, ei is an eigenvalue (and we
have computed its multiplicity). Note that some of the multiplicities may be
zero (for example, if g= 0). It involves no loss of generality to consider these
as eigenvalues with zero multiplicity.
V.2. Representations of the Automorphism Group on Spaces of Differentials 261
= deg Di + 1 —
Di
Xi 2 +
=E2k+EE(mi
k=1 Iln m=1 n/I
mu+[na'
n/1
41)+1 —
x,„ ;
2q(g — 1)
+E E ( 1- - mu +rmun/I- 1+1 -
fin m=1 n/I
x,
n
) [ n/1
+1—g.
il,, Iln m = 1 nil
Next we compute
n— 1 (
nI 1 _
Recall now that {).„,u ; j = 1, . ,n/i} is a permutation of {1, ... ,n/1}. Thus
the above summed from j = 1 to j = n/I yields
E (i - ( 1 + ix, = E 1 1)/x,--2
- 1
In iln lin
E d= - 1,
1=1
trT=g+1 F
fin n lin m=1 j=1
1 xi n-1
E (2.7.1)
!in n lin m=1 j=1
Amii6i=j„,“&+;.,„,2& 2 + • • • +
l= 1
20/1)
▪ 1. mi 1E 1+ 1 + • • • ± A rni(nI0e
l-1
=EE& k=0
k(n[1)-i-
E ckm/i)
nil I— I
=E
.11 k=o
E 1
Amusi= —2 m1(nlI) =
i=
E
.i=1
E
i=1
Àrniinntti
We now return to (2.7.1) and continue with our computation after changing
notation: x1 = t = number of fixed points of T and nrni
x, - -1 E E n - -1 E'1 —n
tr T = 1 —E
lin n in m=1 I nm 1 1—
= 1 - E x, + E x, + E 1
tin lin m=1 1 —
1 #1
1
=I—t+ m=1 _ ev„,
ev-
= E
m = 1 1 - by-
i i j±
n/I
e.
I li .= 1 j= 0 n/L
The first sum, as before, is equal to zero. Thus we conclude (simplying further
as in V.2.6)
xr n- 1 zi n-1 v (I)
1
tr T= -- E/ + n/I
n lin m=1 j0 m=1 j= 0
1 x' — n v v
x' n v
-1 — vo) .
= — 1 E 1(--f) x, E
n tin
1#1
n 1— 0 m n/I f.
As we remarked before,
E
j= 1
ymud = mIlE Ym12 62 ymi(n/I) 0
Iwo+ . . • ± y rnione
= y(1 + ell + 8 2(n/l) + • • • + 0 - 1)(n/1)) = 0,
264 v A utomorphisms of Compact Surfaces—Elementary Theory
For 1. 1
n-1 n-
E Yrni.,e1 =E
j= 1 j= 1
= E1 [11 ""
I Vi
=E
1= 1
m-1
[
J
•
—
n
1 ,..,1
v- I
q;',. 1
--
i= 1 — 1— qm ,
We can finally complete the calculation:
xi 1
tr T -- E x, + E
II. m. 1 1 — riml /in m=1 1 — nm1
1* 1 1* 1
e V,V
=E
m=1 1 — em .
V.2.9. We summarize our results in the following
Theorem (Eichler Trace Formula). Let T be an automorphism of order n > 1
of a compact Riemann surface M of genus g> 1. Represent T by a matrix via
its action on .Yeq(M). Let t be the number offixed points of T. Let s = exp(2ni/n).
Let P„ . . . , Pt be the fixed points of T. For each m --= 1, . . . , t choose a local
coordinate z at P. and an integer v„, such that 1 < v. n — 1 and such that
T -1 is given near P. by
T -1 :z
(note that v m must be relatively prime to n). Then
t Ev.
E _ sv„, for q = 1,
tr T = 1 + m=11
and en v
tr T = E for q> 1,
m=1 1 — s v."
where 0 < y n is chosen as the unique integer such that q = pn + v with
p e Z. In each case the sum is taken to be zero whenever t = O.
Note that only the fixed points of T contribute to tr T (not the points that
are fixed by a power of T but not T itself).
V.2. Representations of the Automorphism Group on Spaces of Differentials 265
Theorem. Let T e Aut M be of prime order n, and assume that T fixes a non
point P. Then 2 < v(T) 4. Further, the genus g of MI<T>-Weirsta
is given by g = [0], and writing g = 4n + r(0 < r n — 1), there are only
three possibilities:
a. r = 0, g = gn, v(T) = 2,
b. r = i(n — 1), g = ('+ 5)n —4, v(T) = 3, or
c. r = n — 1, g = (4 + 1)n — 1, v(T) = 4.
PROOF. Since T fixes a non Weierstrass point, we have 7i = j, j = 1, . . . , g
and the action of T on .r1 (M) is given by the diagonal matrix
diag(e, e2, ,89)
266 V Automorphisms of Compact Surfaces—Elementary Theory
2g — 2 = (2q - 2)n + (n — 1) or
If (p• e .Ye i (ill) and T9 = 9, then ordp 9 > 0. Since it is not the case that for
all 9 E .f l (M) ord, q, > 0, we can find a 9 e Yfl(M) with T9 = e9, e 0 1,
e" = 1, ordp 9 = 0. Since 9 does not vanish at P (the unique fixed point of T)
and (y)) is invariant under T, 2g — 2 = kn for some positive integer k. Thus
we see that
1 = (2g — 1) — (2g — 2) = n(2:4 — 1) — nk = n(2:41 — 1 — k).
In particular, n cannot be > 1, and this contradiction establishes the theorem. _
Remark. The Riemann-Hurwitz formula that we applied used the fact that
n is prime. This is essential; for the theorem is not true if n is not a prime.
V.212. We consider now the representation of an automorphism T of
prime order n on oveq(M), q> 1. We assume that P is a fixed point of T and
that P is not a g-Weierstrass point. Then the representation of T on drq(M) is
diag(eq, ,8(2q-1)(9-1)+" ).
where OE is the genus of MX T>. Combining these results with the Riemann-
Hurwitz formula (2.10.1) we see that
(2kn + 1){1( - 1) + Iv(n - 1)1 = f(2kn + 1)(4 - 1) + vk(n + r,
with 0 < n - 1. Thus's.
2r
v =— -
n-1
and there is just one possibility (recall v(T) 0 1) r = n - 1 and y = 2, and we
have established the following
Theorem. Let T e Aut M be of prime order n with v(T) > 2. Then every fixed
point of T is a q- Weierstrass point for every q> 1, q 1 mod n.
v.2.13. We can now reprove and generalize slightly some of the results on
automorphisms of hyperelliptic surfaces (compare with 111.7.11).
This is, however, the representation off on ye' (AI). Since the representation
of Aut M on Je l (M) is faithful T = J.
Next we consider an automorphism T of prime order n, 2 <n < 2g, that
fixes a Weierstrass point. Its action on .X"(M) is given by (e = exp(21ti/n))
diag(s, ,e 4- 1 ).
Write 2g — 1 = In + r, 0 < r < n — 1. Then (since g is the multiplicity of the
eigenvalue 1) 1 even implies œ= 1/2 and 1 odd implies that œ = 1(1 + 1).
Using Riemann-Hurwitz (in a by now familiar way) we see that
r+1 2g+ 1
!even v(T) =2+ =3 (r = n 2).=
rt ,
n—1 2g- + 1
1 odd v(T) = 1 + n r 1 = 2 (r = n 1) n =
(2œ + 1 — r = r2 —1.
Since n> 2, the last equation implies r2 = 1 or r2 > 3. Since r2 < v(T), we
see that r2 = 1. Thus T has a fixed point P that is not a Weierstrass point.
Since, J is central in Aut M,
T(J(P)) = J(T(P))= J(P),
and J(P) (0 P) is also a fixed point.
If T fixes two points, both must be Weierstrass points by an argument
similar to the one given above. 1:1
Remark. The definition is meaningful and the set is indeed a group. All
that we need verify is that if X G Sp(g,Z) so is It is easy to check that
if X =[ ] and Xe Sp(g.Z) then =[ This last statement
].
S
r, T w= w. (3.1.4)
dual to ' {x} ; that is, .f„, ock = bkj , k, j = 1, . , 2g. It is clear that this basis
is unique and we have, as before, jr.• = 1„, = bu . Let the 2g x 2g
matrix X = (xik) represent the action of T on `{K}. We then have
2g
I = (61) = (fT., Tat) = (fat xjk.k Tai)= (El xik f Kk Tai).
Let us denote by T the matrix whose (I,k) entry is f,,„ Tat . Thus the above
equation reads
I = 7-1X,
Or
T=
Since Tzi is determined by its periods we have established the following
V.3.3. In this section we wish to strengthen Theorem V.3.1 and its corollary.
We prove the following
Theorem. Let M be of genus (I > 2 and assume Te Aut M satisfies T(a i)= a„,
T(a2) a2 , T(b,)= b, and '7'(b 2) = 6 2 in H ,(M). Then T = 1.
Thus we see that b must be an integral multiple of some eu). Since the
components of b are relatively prime, b = ± e°, for some j. 0
We can assume without loss of generality that u v. (Let X' = JoX. Then
X' e Sp(n,Z) and X1';,] = [4'].) We shall now show that y = O.
V.3. Representation of Aut Mon 111 (M) 273
Let us for the moment assume that T has a fixed point say P. Let Q E M
be arbitrary, then we conclude from (3.1.4) that
j = 1, 2, g + 1, g + 2.
We consider the surface i = M/< T> and the natural projection it: M
Let j = 1, 2, g + 1, or g + 2. Because of T(Œ) = cti, the differentials a; project
to harmonic differentials itai on M. (This assertion is not quite obvious.
Its verification is left to the reader.) If Zr is a closed curve on gr, then we
lift it to a curve c on M beginning at some Q E M and ending at TIV. By
(3.3.2) we see that
fe 7rŒi e Z, j 1, 2, g + 1, g + 2.
(Note that since we are only interested in the homology classes of the curves,
we may assume that c does not pass through any fixed points of T.) Let us
construct a canonical homology basis {di , on M and a
274 V Automorphisms of Compact Surfaces—Elementary Theory
X = [e C]=[ b • '"]
of integers mod n.
Since 9 is surjective the vector (s,e') has relatively prime components.
We review the above situation in the language of multiplicative functions.
Construct on Sa a multiplicative function f belonging to the n-characteristic
x. This function lifts to a single valued function on the cover corresponding
to the homomorphism
By a change of homology basis (since Z„ is commutative we may work
with 1-1 1 (1N) instead of It i (k)) we may assume that x is of the form [I; 8:::
This follows immediately from our second lemma. More precisely, if x is
then we can consider the change of homology basis effected by a matrix
V.3. Representation of Aut M on Ii i (M) 275
X(c) = (EM [ n ].
[e 1)]
= jth entry of
0
It is now a simple matter to complete the proof. We examine the cover
corresponding to X = .:: g]. It is clear that if di, ... ,
is the homology basis for :V/ such that X = B g ::: 8], then di , 2, ,
(n — WTI lift to open curves on M and nil, lifts to a closed curve on M.
Further b 1 has n disjoint lifts all of which are homologous on M. Each of
the other cycles also have n disjoint lifts on M; however, these are all
homologously independent. The action of T on M fixes the lift of na, and
permutes the lifts of ; however, since the lifts are all homologous we can
say that (as far as the induced action on homology) T fixes the lift of b 1 . No
other cycles are fixed. Hence we finally conclude that T must be the identity
and this concludes the proof of the theorem.
V.3.4. In this paragraph we strengthen (in another direction) Theorem
V.3.1. The group Aut M acts not only on H 1 (M) but also on H i(M,4),
the first homology group on M with coefficients in the integers mod n.
Choose T E Aut M. We have seen that T acts on H ,(M) by a square integral
matrix A E Sp(g,Z) and A has finite order m. We first prove the following
discrete. We will establish a result which will hold with any reasonable
concept of discreteness for Aut M. We state it as a
1. C fœl,
2. C,
3. 4 = {z E C, iZi < 1 } ,
4. C* = (01,
5. A* ..4\{O},
6. = {z E C, r < Izi < 1), 0 < r < 1,
7. a torus, C/G, where G is the free group generated by the two translations
7.1.—oz+
V.4.2. In any standard book on complex analysis, the reader will find a
description of the automorphism groups of the three simply connected
Riemann surfaces. We have used this information already in IV.5.2. We
repeat it here:
Aut(C {x}) PL(2,C)
Aut C P4(2,C)
Aut 4 Aut U PL(2,1/).
The above are clearly not discrete subgroups of PL(2,C). As a matter of fact
they are Lie groups. The first two are complex Lie groups (of complex
dimensions 3 and 2, respectively). The last is a real Lie group (of real di-
mension 3).
V.4.3. It is clear that every automorphism A of C* extends to one of
C {CO}. Furthermore, A must either fix 0 and oo or interchange these
points. Thus the automorphism group of C* consists of Möbius transforma-
tions of the form
z kz or
V.4.7. We examine next the torus. Let the covering group G of the torus
T be generated by z 1-0 z + 1 and zi—• z + t 1m t > O. It is quite easy to see
,
a = 1 X ± NI X 2 — 4
2 2
with x the trace of the elliptic element X e SL(2,7L)/+I.
Since SL(2,1) has only finitely many conjugacy classes of elliptic elements,
the number of a's is finite. (Since 0 < x2 G 4, we must have al = 1. As an
exercise prove this another way.)
V.4.8. It remains to show that we have exhausted all M with Aut M not
discrete. We may restrict our attention to those surfaces whose universal
covering space is the unit disc (or the upper half plane U).
Let G be the covering group of M. Suppose there exists a distinct sequence
fn e N(G) such that lim„ f„-= 1. Choose 1 0 A E G and observe that
lim„f„.Aof„-1 .21 -1 = 1. Since f, o A o f, o A'eG, there exists an N
such that f„o A o f1 o A 1 = 1 for n > N. Thus f,, commutes with A for
large n. If A is parabolic, then by Lemma IV.6.6 so is f„ for large n; and A
and f„ have a common fixed point. Since 1 0 A E G was arbitrary, G is a
commutative parabolic group and by Theorem IV.6.1, M is the punctured
disc.
Similarly, if A is hyperbolic, every element of G is hyperbolic and com-
mutes with A. Appealing to the same earlier theorem, we see that M is an
annulus.
CHAPTER vi
Theta Functions
(because N is an integer vector). Thus for all but finitely many terms, the
general term of (1.1.1) is bounded in absolute value by
exp( — lir). 011N11 2),
and our series is majorized by
g 1
E exp(
2
10. 04
J=1 nj=
which converges since each factor converges (by the Cauchy root test, for
example).
We have shown that a is a holomorphic function on C9 x We shall
be mostly interested in this function with fixed "C E In this case, we will
abbreviate 0(:,r) by 0(4
(because 'N E Z)
= exp 2ni[—` /Az — 4tyrid0(z,r)
(because (N + p) is just as good a summation index as N).
VI. I. The Riemann Theta Function 283
Corollary 2. Let t the k-th column of s, and Tkk the (k,k)-entry of T. Then
rkk
0(z + r(k),r)= exp 2ni[ - z - —]0(:,r), all : e tg, all r E S9, (1.2.3)
k 2
where z k is the k-th component of:.
..- -
PROOF. Take y' = 0, y = e(". 0
Remark. It is also easy to verify that 0 is an even function of z. Thus in addi-
tion to our basic formula (1.2.1), we also have
and that
Tkk
0(z + e + T( ) = exp 2n[
i - z k - ek - — 10(z + e). (1.4.2)
2
The connection between theta functions and the theory of multiply periodic
functions on Cg is seen rather clearly when one considers two distinct points,
d and e E Cg, and sets
0(z + e)0(7 - e)
= f(z). (1.4.3)
d)
It is clear from (1.4.1) and (1.4.2) that f(z) in (1.4.3) is a periodic meromorphic
function on Cg with periods eti), rtn, j = 1, . . . , g.
Recall that the columns of the g x 2g matrix (I,r) are linearly independent
over R. Hence, every e e Cg can be uniquely expressed as e = le' + TE with
e, e' points in R9. Hence 0(z + e) = 0(z + + TO. This suggests that for any
e, e' E R9 we define a function on Cg xC.A. by
2s 1
1(z,r) = yexp 274- t(N + c)r(N + e) + `(N + e)(: +
2
(1.4.4)
0[2s' NeZe
We observe at once that
=
and in general
26 I 1
0[ 1(z r) = exp 276[- :A/TN + ' N(z + — 'et& + tez +
2er N is 2 2
, 1,
= exp .ari[- 'STE t EZ t E816(Z + + TE,T). (1.4.5)
2
From the last equality it follows that for integral p, p' e Zg, we have (because
of (1.2.1))
1
0[28 1(Z + lie T/4, = exp 27ri[- ETE t EZ t EE' t EGI '
2s' 2
x 0(z + le' + te + + TA, i)
1
exp 21rik'sre + + tee'] exp 21i[e(ii + Ty)
1
- '1(z + le' + Te) - - '114(1 0(z + le' + se, T)
2
1 2e
= exp 24--
'tap- f izz + tp's - tpte1[ 1
2 2e'
(1.4.6)
V1.1. The Ricmann Theta Function 285
In particular, we have
26
0 [ 1(z + e(*),T) = exp 27ri[ek] Or] (1.4.7)
2e
and
E
0[26 ](z + z (k),T) = exp 4 0 2-- e, (Z,T). (1.4.8)
2e' zk
VI.1.5. For us the most important case is when 24 and 24 E Z. Let e, e' e Z9.
Then OM (z,t) is called the first order theta- function with integer characteristic
D.1
Proposition. The first order theta function with integer characteristic [e] has
the following properties:
or+ 2v 1 (zr) = ex p 2 7t i —
tcv' 0 (z (1.5.4)
E' 2 [e
(Here y and y' are integer vectors in Z9.)
PROOF. Equations (1.5.1) and (1.5.2) are simply the restatements of (1.4.7)
and (1.4.8). Equations (1.5.3) and (1.5.4) are derived in the same way as (1.4.5),
(1.2.4), and (1.2.1). LI
The importance of the statements (1.5.3) and (1.5.4) is that we immediately
have the following
Corollary. Up to sign there are exactly 229 different first order theta functions
with integer characteristics. Of these 29- '(29 + 1) are even functions, while
29- '(29 — 1) are odd. These 229 functions can be thought to correspond to the
229 points of order 2 in Cg/<1,0, where at> means the group of translations
of C9 generated by the columns of (./,T).
PROOF. The only part needing some comment is the number of even and
odd functions. This is established by an easy induction argument. LI
286 VI Theta Functions
EXAMPLE. When g -- 1 the three even functions arc: O[g](z,r), 0[?](:.-r), and
OW(z,r). The odd function is 0[1](z,r). When g 2 the six odd functions are:
[i o i
](:t)
0,,
o[ o[i ].7A (
Remarks
1. For the first order theta functions with integer characteristic, 0[:.](::,r),
the even ones are precisely those for which t e.E' 0 (mod 2) and the odd
ones those for which iCe' -Z4 1 (mod 2).
[ ]( ,T) + — t+
= exp 2ni [—
re
— —Z
re
--
1,(
2. Formula (1.4.5) applied to the case of integer characteristic yields
1 re
Z+ / — +T—
e
S . ; = S4a, Bb A + Bll,
and hence
C' = (A + B17) -1 C,
and
17' = C' =Sc. Db (A + BI7) -1 C = (A + B17) -1 (C + D17).
Thus II and 17' are related by an element of the symplectic modular group
of degree g. The relation among the corresponding theta functions involves
then the study of modular forms for the symplectic modular group (which
will not be pursued here).
VI.2.3. We shall for the remainder of this chapter assume that the
Riemann surface M has on it a fixed canonical homology basis ' {a,b} =
{a 1 , . ,b9}, and we will study the theta functions associated with
M and qa,b1 as functions on M. We do this by recalling that we have intro-
duced in 111.6 a map 9:M J(M) = Cg/L(M), where L(M) is the lattice
(over Z) generated by the 2g columns of the matrix (1,17). Recall that q was
defined by choosing a point Po e M and setting 9(P) = C, where C =
t (C i , ,C9) is the basis of OW) dual to ' {a,b}. We now consider 0 0 9 and
in this way view 0 or 0[ ] , where [ e ] is an integer characteristic, as a func-
tion on M. The reader is, of course, aware that 0 0 9 is not single valued on
M, since 9 is not single valued (as a function into C9) on M, but depends on
the path of integration. We have seen (Proposition 111.6.1) that the map 9
is well defined into J(M), and therefore the function 0 o q has a very simple
multiplicative behavior. The behavior of this function is not quite as simple
as the behavior of the functions considered in 111.9, since here the multiplier
288 VI Theta Functions
where f - denotes the value of f on the cycle ak-1 or bk-1 . It follows imme-
diately from (2.3.1) and (2.3.2) that if P is a point on ak
Ck _.... -n kk
f - (P) = exp 2n[
i --- —.
2 2
while if P is a point on 6,,
= 1 [ r (4) df_ _ df -) r df _ df
27ri kL
:
=1 Jak f 9 ibk 9 f -
where (p - plays the same role for 9 as f - served for f. We once again use
the relation between f - and f and the fact that for P a point on a,,, 9 - =
9 n(k), while for P a point on bk, 9 = (p - e ). Thus
1 Fr df
= Liak 9 y— fd - 27d9' (z)
+ n (k))(i k dz)
fbk Op - e- df
- 9 _ df -]-
n(k) dff + 2nin(k) 9(z)dz + 27ci99;,(z)dz
,‘L= ' a k L
1 g f 00 df -
+—
27ri =
e Ef-•
We now need to evaluate integrals of the form f,,k (df/f) and fbk (df/f).
Since df/f = d log f, the integral in question is simply log f(P 1 ) - log
where P 1 = P2 are the initial and terminal points of the cycle a„ (or kJ.
(The difference is not zero, since we must use different "branches" of f.) We
now once again use (2.3.1) and (2.3.2) to obtain for P i and P2 initial and
290 VI Theta Functions
= 1 g
- it 2ni + 2nink + n(k)2ni + f ahcp(p(z)dz
t
k, bri
1 9
4ni
E k= 1
Ek
e0127,i(--
2
nkk - pk) + 2itim k],
-—
2
where mk and nk are integers. It therefore follows that
1 df
=—
27r1 Li 9 f
g E ek) s'
n(k) (1 — nk )
2 2
2kk (pet)
+ I (pcp'k dz - e( 5--
or finally
E
= - TI - - / - + lin + lm - K,
2 2
where
9
K=-
k
E [f ak
7rkk
cpcpkdz - em(-- + (p k(P,))].
2
(2.4.1)
Remark and Definition. The vector K depends, of course, on the choice of the
canonical homology basis and the choice of the base point P o of the map (p.
(There is an illusory dependence on the base point P, of n 1(M) which can
be dispensed with by choosing P, = Po in the above.) The vector K is
known as the vector of Riemann constants. We have proved the following
There are now two possibilities to consider. Either i/c is identically zero
or not. In the former case, we have for each k = 1, . , g,
0((p(P,) — (OP, • • • P9) + K)) = O(—(P 1 • • • P„ • • • Pg)- K)
= 0((P(P • • • Pk • • P g) + K)= 0
VI.3.2. We now observe that we can really prove a bit more than we have
claimed in VI.3.1. The points = P, • • Ps and co Q, • • • Q, utilized in
the above proof, were fairly arbitrary in the sense that any other points
t;', co' e M, which are sufficiently close to and w would have worked as
well. We also found that e = cp(T • • T 9 ,P2 • • • Ps) e M 9 _ 1 has at least
s — 1 "arbitrary" points P2 , , Ps in it. Let D' = ' • • Ps be arbitrary but
V1.3. The Theta Divisor 293
= — 4) (Q1 • • Os--1 6) — K;
PROOF. All but the last remark is contained in the discussion preceding the
statement of the theorem. In the discussion we only showed i(C) s; however,
the statement i(;) = s is an immediate consequence of the one preceding it.
If i(C)> s, we would have O( W3 — W5 — e) 0 contrary to the hypothesis.
Theorem
a. If e e J(M) and -L=. 0, then e = cp(D) + K for some D e M. with i(D)=
s > 1, and s is the least integer such that
0 1 4 s+i — — O.
b. If e = cp(D)+ K, D e M9, then tk # 0 if and only if i(D)= 0 and D is the
=. 0 if and only if i(D)> 0.
divisor of zeros of O. In particular, III 7..
294 VI Theta Functions
PROOF. Since the lemma clearly holds for r = 0, we take r> O. We have
0(p(P 1 ' • • P.) — (P(Q, • " Q,)— e)= 0 for all D = Pi • • • Pr M„ all D' =
Q i • • Q, e M,. Thus
0(40(P) + 402 • • • Pd — (.0 (Qt • ' e)
vanishes identically on M as a function of P. If we now expand this function
in a Taylor series about the point Q, and set P = Q1, we find the coefficient
of a local coordinate at Q i to be
vanish at ‘;'. We would rather view this for the moment as a function on M,
and in order to do so we consider
g(P) = exp(S
P
E 'Cm)
k „ ex (f" TR 4
P° 1
where co = R 1 • • • R„ T = T, • • • T 5. It is clear that g(P) and f(P) have the
same zeros and poles and the same multiplicative behavior. Hence it follows
immediately that f(P) = cg(P), with c a constant which may depend on
VI.3. The Theta Divisor 297
= c[E(R 1) (3.5.5)
azi 0)(R P 2 ' • • P — Q(r) — eg;( 1? i)
32e
J.
E
g
k=
(q)(P3 • • •
:k 1uZjC
Ps) — (1)(R 3 • • Rs) — e)C(R 1 )Ck (R 2)
It follows from (3.5.3) that the right side of (3.5.7) does not vanish. Hence
the same is true for the left side and therefore it is surely not the case that all
sth order partial derivatives of 0 vanish at — e (and therefore also at e).
298 VI Theta Functions
by the Riemann- Roch theorem implies Kw) >: 1. and therefore igco) = 1.
Hence ,:co is the divisor of a holomorphic differential.
Conversely, suppose d is a divisor of degree 2g — 2 on M such that
q(J) = —2K. It follows by Abel's theorem that there is a function f on M
with divisor z1/(Œ), where is a holomorphic differential on M and (a) denotes
the divisor of Thus fa is a differential on M with divisor (fa) = J.
111.8.13 and Theorem 111.8.7 (also for hyperelliptic surfaces of genus 3). We
conclude from Theorem 111.1 1.19 that
dim O zer. = g — 1,
dim esup„ zero = g — 2,
and
g— 4 dim esing g — 3,
and hence the inclusions of (3.7.1) are all proper. The points corresponding
to vanishing of the 0-function, but not identically vanishing (on the surface),
are in
(417,\W) + K = (J(M)\W;) + K.
300 VI Theta Functions
The sets ezer,„ and E) only on the period matrix 11. The fact that
such a set is non-empty can, of course, be translated by the Riemann vanishing
theorem and the Riemann—Roch theorem into a statement about existence
of meromorphic functions (of a certain type) on M.
It should also be remarked that Ozer° and esing are independent of the
base point of the map (p:M J(M). Thus °sing describes those points
e E J(M) corresponding to the identically vanishing of ti/ for all choices of the
base point (see also VII.2.1).
CHAPTER VII
Examples
= n
2g+ 2
k= 1
(z — e,,), ek 0 ep for k j. (1.0.1)
bk Pk J0k Pk f Pk
P2k
= 2 J. = 2 f., ç.
sk r2k - 1
Thus up to homology
1
(P ,51 ) = - (70) + • • • + it(k) + eu ) + e(k+1)), k = 1, 2, . . . , g - 1,
2
)
(1)
(P(P2g+ i) = 1-2 (7 +• + 7/(g) eu )),
1
41P2, +2)
304 VII Examples
A change of basis for abelian differentials of the first kind, multiplies the
above matrix by a non-singular matrix. Hence to compute the rank of (1.1.2),
we may choose a convenient basis for .e(M). We shall use the basis given
by 111(7.5.1):
zi dz
j = 0, . . . , g — 1.
1 e3 es-
1
g dz
T—r 1 e$
= 11 — (P2)+ I) •
i= 1 w •
„..„ 9+ I
1 e294- 1 c 294 1
The last matrix is, of course, the Vandermonde matrix which has already
been encountered (in IV.11.10), and is non-singular. We have hence shown
(III.11.11) that i(D) = 0, and that (p. establishes an isomorphism between a
neighborhood of D in Mg and its image in J(M).
VII.! . Hyperelliptic Surfaces (Once Again) 305
It2kk
K= — (pcp'k dz — e (k) (-
k-=1
where we assumed that the base point for 7t 1 (M) is chosen to agree with the
base point of the map ça:M J(M)—which in our case was selected to be
a Weierstrass, point P1 . It is rather difficult to evaluate K directly. We know
(Theorem V1.3.6) that — 2K is the image in J(M) of the canonical divisor
class. Since P?g' is the divisor of an abelian differential of the first kind,
the canonical class gets nipped into 0 by (p. Thus K is a half-périod. We
want to know which half-period. We shall shoir- that
K= 2 i+ 1). (1.2.1)
i=
Assume for the moment that we knew that 0(0) 0 0, for the 0-function in-
troduced in Chapter VI. In this case 0 o cp does not vanish identically on M
(this is the fact we really need), and hence 0 0 cp has g zeros Q1,. . . , Qg on M.
These zeros satisfy
f + K = 0.
J=1
(recall Theorem VI.2.4).
We have seen that (P) is a half-period. Such a half-period can be written
as PM + Hs), where rej is an integer characteristic. Thus the half-periods
can be classified as odd or even, depending on the parity of the characteristic.
Up to a constant non-zero factor,
19 06' +
—
Note that conversely, the identity (1.2.1) shows that 0 does not vanish
identically on M. For if 0 vanished identically on M, then (p(P3P3 • • • P20.1)-1-
K = 0 and i(P 3 • • P2 g+ j) > 0 by Theorem VI.3.3. This contradicts the
previously established fact that i(P3 • • - P 2g+ ,) = 0. These remarks also
show that 0 vanishes at 0 if and only if 0 vanishes identically on M. The fact
that an even theta function vanishes at 0 if and only if it vanishes identically
on M is also a consequence of the Riemann vanishing theorem.
306 VII Examples
0, 2, ... , 2g — 2.
Thus
i(pv 1 11(g + 1), if g is odd,
)= if g is even.
19,
(Using the greatest integer notation, we can write that the order of vanishing
is always [1(g + 1)].)
For D e Mg _ 1 , r(D - 1 ) = i(D), and Clifford's theorem (111.8.4) shows that
— 2 i•
We have established that for a hyperelliptic surface there are points at which
the 0-function vanishes to the maximum order possible. We have shown
that such vanishing occurs at a point of order 2 in J(M).
As a matter of fact, given any non-negative integer n [ -(g + 1)], there
is a point e e J(M) such that e has order 2 and 0 vanishes at e to precisely
order n. To see this, we have to construct for n> 0, divisors C E M9 _1 with
C containing only Weierstrass points and i(C) = n. Then the point e =
(p(() + K will have the desired property (for n = 0, we, of course, need a
divisor C e Mg with i(C) = 0). The calculation in VII.1.1 established the
following fact: If D e M, and D consists of distinct Weierstrass points, then
i(D) = 0. From this observation it follows that if (for 0 r g)D E M9 _,. and
Hyperelliptic Surfaces (Once Again) 307
cp(Pik )+ K
k=1
is a point of order two in J(M) at which the 0-function does not vanish
(choosing jk = 2k + 1, we obtain the previously used fact that 0(0) 0 0).
Choosing g — 1 distinct integers as above (J is now permitted to be 1), we
see that 7) vanishes to orelér 1 at •• •
g- 1
E (p(pik ) + K.
k= 1
Theorem. Let M be a compact surface of genus 4. Then one and only one of
the following holds:
°sing is 1-dimensional,
b. O consists of precisely one point (a point of order 2 in J(M)), or
c. 951 n, consists of precisely two points (a 0 0, a e J(M), and — a), neither one
of which is of order 2.
Each case corresponds to the case indexed by the same letter in Theorem
111.8.7.
PROOF. If M is hyperelliptic then () sing is one-dimensional (and conversely).
In fact, in this case, ()sin, is analytically equivalent to If, q(M) and hence
to M (EXERCISE). Thus case (a) is disjoint from the other two cases.
Proposition 111.8.6 showed that WI is non-empty. Assume a e 47 + K.
Assume also b e W + K. The proof of Theorem 111.8.7 showed that
+ b = 0 unless M is hyperelliptic (recall that the image of the canonical
class in J(M) is —2K (Theorem VI.3.6)). Hence the only possibilities are
a = b (and hence 2a = 0) or WI contains two points and b = — a 0 0. In
the latter case a cannot be of order 2. LI
and we must add a set of g — r numbers in {2,3, ,2g + 2} that are not in
ti 1 ,i2 , Assume this set is {k 1 , Then
(p(Pii ph • • Pi.) + K = cp(Pk ,Pk2 • • •
V11.2. Relations among Quadratic Differentials 311
Recall that the set fo(P 2), ,9(P2 „,)) contains exactly g odd half-periods
and g + 1 even half periods. Thus in the set
2.3 2g + 2',
there are at least g — r indices corresponding to odd half-periods, and at
least g — r + 1 corresponding to even half-periods. We choose s odd points
and t=g—r—s even points. What is the parity of this divisor? Using the
remark at the end of VII.1.2, we see that this divisor has the same parity as
+ —
s+ = s + -1(9 — r)(g — r — 1).
2 s ._
Thus if1(g — r)(g— r —1) is even, we choose s to be odd, and if 1(g — r)(g — r— 1)
is odd we choose s to be even. We have enough room in our set to accomplish
this.
+ (e)Ci"(P0) + 3 E (ePPogigo)
04, j,k GUJUk j
e3o
+ ;;(PoK k(Pogi(P0) -3-i+ • • • • (2.2.2)
CUJOUkClii
E= A a: ° OW/EC! = 0. (2.2.4)
Uk 041
PROOF. Equation (2.2.3) follows from (2.2.2) and the previous observation
that the first partials of 0 vanish. The vanishing of the coefficient of z3 in
(2.2.2) yields
a20 1,3o
E E Gui C/Uk GUI (e), =
, ( e)cick + j,k,i (2.2.5)
a20 030
3 E euiatik +..
j,k,1 aU aU
eU k
0. (2.2.6)
If we subtract (2.2.6) from (2.2.5) and use the fact that 0 is an even func-
tion, then we obtain (2.2.4).
Remark. There are instances when the above proposition is of little value.
For example, if e e &sing , and all the second partials of 0 at e vanish (e e
W.2 _ 1 + K), then (2.2.3) yields no information. It could also happen that all
VII.2. Relations among Quadratic Differentials 313
the second order partial derivatives do not vanish at e, but all the third
order partial derivatives do vanish. This takes place when e e °sing is an
even point of order 2. Why? (EXERCISE.)
EXERCISE
How much of the above carries over to cases (a) and (b) of Theorem VII. 1.6?
au,uuk
(e(z)) = atij aU k
(eo) + (E S UJ SIk SL (eo)(Po)
(v
a3o 7, v
au,au, au„`"(P 0' 4- IL:. Otim auS41 -O
0
ufônk(e0)i(Pog„,(P0))f-
2
(2.4.3)
Inserting (2.4.3) into (2.4.2), we obtain
4 5 20
= 0,
j,k= I aUjaUk(e°"k
-3n
E E au, Su"SU,, (e ° g 1(Po) C.4= 0,
and
030 ,940
E (E
j,k 1 autaujauk
(eAl(P0) E G um au,,,ui cuk (eogi(Po)m(P0))
- CiCk = O.
varies the base point Po). The problem of specifying three such relations is
apparently still open.
EXERCISE
Is the map M esing, that we have constructed, surjective?
VII.2.5. We shall discuss briefly in this paragraph a question intimately
related to the question of relations among the products of the normalized
abelian differentials of the first kind. Let 110 e 4. be a period matrix of a
compact Riemann surface of genus 4. (Recall that is the Siegel upper
half space of genus 4 introduced in V1.1.1.) Theorem VII.1.6 guarantees
that (51 ; is non-empty. Assume we are, again, in case (c). By the remark
following 111.8.7, the rank of the associated relation (2.2.3) among the abelian
differentials must be 4. Choose a (small) neighborhood N of Flo e Z4. What
is a necessary condition for // E N to be a period matrix of a compact
surface of genus 4? Clearly, the 0-function for II must have the property
that esing 0 Ø. We proceed to write down an equation in 54 that expresses
this condition. Consider the system of equations on C4 x
eo , u= 1. 2, 3, 4. ( 2.5.1)
ez.
The hypothesis that /70 is a period matrix tells us that (z°,I70) solves (2.5.1),
whenever z° e °sin , for /70 . The condition that (2.2.3) have rank four, tells
us that the Jacobian matrix of the system (2.5.1) with respect to the z-variables
has rank 4, and thus the implicit function theorem asserts that we can solve
for z = t(z 1,z 2 ,z 3 ,z4) as holomorphic functions of t in a neighborhood of
(z °,17 0) and that
= 0 for n = 1, 2, 3, 4, t e N.
A necessary analytic condition for t to be a period matrix of a compact
surface is now easily written down:
F(r)= Ci(z(T),T) = 0, t e N.
, j = 0, , k — 2,
and
zi dz
-v-v---2, / = 0, .. . , 2k — 2,
eh
t
W2
= Q3lQ3(2k-I-2)
We shall consider the special case k = 2 (hence, g = 4). Thus the basis for
Je l (M) is:
dz dz dz d-
-, (3.1.2)
ww w
The above basis is adapted (recall 111.5.8) to the point Q 1 . As a matter
of fact, the Weierstrass "gap" sequence at Q i is:
1, 2, 4, 7. (3.1.3)
VII.3. Examples of Non-hyperelliptic Surfaces 317
Note that the above is also the "gap" sequence at Q2, Q3 and Pi , j = 1, 2, 3.
Thus each of these 6 points is a Weierstrass point of weight 4. We have
thus accounted for 24 of the 60 Weierstrass Points (here we arc counting
each Weierstrass point according to its multiplicity). The remaining 36
Weierstrass points will occur in groups of three. Each such group of three
will project to the same point in C u { 3o} by the map z. This follows from
the fact that M has an automorphism of period three that interchanges the
sheets of the cover.
We continue with the case k = 2 in (3.1.1). Since M is of genus 4,
by Theorem VII.1.6, A7t is either hyperelliptic (impossible, because M
carries a function of degree 3, by Proposition 111.7.10) or °sing consists
of one point of order 2 or es,,,, consists of precisely two points (neither
one of order 2). Since Qf is the divisor of the abelian differential z 2 (dz,42 ),
the vector K of Riemann constants with respect to the base point Q 1 , is
a point of order 2 of J(M) by Theorem VI.3.6. Further, since
K =9{Q?) ± K,
K is a zero of the theta function. Also; looking at the "gap" sequence (3.1.3),
we see that
i(Q1) = 2,
which shows that K e 0,ing . We have shown that °sins
consists of the single
point K. It must be the case that there is a relation of rank 3 among the
products of the abelian differentials of the first kind. The relation is easily
exhibited:
( 7. dZy = ( z2 dZ2 )(dZ2 ) .
w2 (3.2.1)
W
ÇO4
_ , 3 dz dz dz
Ç07 Z
dz
95 - (P6Th -,
w-
then we can write down the following maximal set of linearly independent
relations
(,03(P7 = 9496, 94 ( 7 =
VII.3.4. The Riemann surface M of (3.1.1) is an example of a surface
with an automorphism T (of period 3, in our case) such that M,i<T>
C u tool. We have seen that for these surfaces, contains a point of
() sin ,
order 2. It may seem at first glance that the existence of points of order 2
in 0,ing is caused by the presence of the automorphism. This conclusion is,
in general, false. Consider, for example, the surface M defined by
1V3 = Z Z -
( 1)(z — ;. 1 )2(z — ;1 2)2(z — ). 3)2,
(z)= 42?
Q3
(w) —
Q1Q2P PiP23■
421
and
(dz)= QiQiPiPiPi
421
A basis for if l (M) is thus
dz dz dz dz
z (z — ,1 1 )(z — ,1 2 )(z — z(z — 2 1 )(z — ). 2)(z —
w w
VII.3. Examples of Non-hyperelliptic Surfaces 319
From the above we see that the possible orders of zeros of abelian differentials
of the first kind at Q 1 are:
0, 1, 3, 4,
and thus the "gap" sequence at Q 1 is:
1, 2, 4, 5.
In particular, there is no differential in 3r1 (M) all of whose zeros are at
Q 1 . Now, the point K belongs to e sing (because f(Q1) = 2), and if °sing
a point of order 2, K would have to be that point. Hence 2K = O. contaied
By Theorem VI.3.6, weewould also have y9(0) = 0 = —2K or that 421
is canonical. We have seen that this does not occur in our example.
VII.3.5. There is, however, one case where the presence of an auto-
morphism T produces points of order 2 in 0„,,g : the case when T has period 2.
Suppose now that M is a surface of genus g 2, and T e Aut M has period 2
and v(T) fixed points. The genus 4-. of MgT> is computed (recall V.1.9) by
= v(T) = 2g + 2 —
(and M is called #-hyperelliptic).
To show that 0,, n, contains a point of order two, it suffices to find an
w e Ye 1 (M) with even order zeros
(0.)) = Pi • • •
such that
i(Pi • • P9 _ 1 ) 2.
For then e cp(P, • • • P9 _ 1 ) + K e esing , and 2e = cp((co)) + 2K = O.
If # = 0, then M is hyperelliptic. Then for g 3, there is a point of order 2
in °sin, and the order of vanishing of the 0-function at this point is 1(g + 1).
Hence we now assume g> O.
VII.3.6. Let us generalize by considering an automorphism T of prime
order N with MI<T> of positive genus. Consider the action of T on Y i(M).
Since MgT> has positive genus #, 1 is an eigenvalue. Since M has bigger
genus than M MgT>, there is another eigenvalue e = 1, a 0 1). In other
words, there are holomorphic differentials co, co, e .e.(m) such that
Tco = co, Tco, = scoi , co O.
The differential w projects to A7i; while co, projects to a multivalued differ-
ential on M. The function f = co i/co on M projects to an N-valued function
on M.
To describe the structure of the divisors (w) and (co 1 ), we let P1 , , P,
be the fixed points of T (v = v(T) could be zero). Calculations similar to
the ones performed in V.2 show that
ordp, (.0 = Ns; + N — 1, r•eZ
1
r•>- 0•
320 VII Examples
-.. X. (3.6.3)
(/) =
Every multivalued function on A-4. whose divisor has the same fractional
exponent at 11; as in (3.6.3) and that has the same image in J(M) as (1),
lifts to a meromorphic function on M.
Consider the divisor D on AI
D= •• 131:1-i'4N .
We now pose the following problem: Does there exist an integral divisor
X on k such that
v t•
deg X = - -
and
4 (X 2) — - • • .1)„1- '-"' LI) (3.6.6)
equals the point of order N in J(.1) determined by the divisor of the
function _T?
The theory of the Jacobi inlersion problem asserts that we can always
solve the above problem, and in fact do so with [-I (1 - ti/N)] - 1
free points. The condition that we shall need is
R= (1 — -t--1) 2,
•■••
which occurs by (3.6.4) whenever
v(T) 2N.
Furthermore, there are (2N)24 different solutions to (3.6.5) corresponding
to the (2N) 24 points of order 2N in J(k). Only 22i of these solutions will
satisfy (3.6.6).
There are now two cases to consider:
R is an even integer ( 2), (3.6.7)
and
R is an odd integer ( 3). (3.6.8)
In case (3.6.7),
x2N
prtl . AN
We conclude that
p+11x2N
prt, prvr-tv A N
R = E (i - =
N 2
and thus X has precisely
r4_(7_111
v ( T)
i(Z1)
Remark. The above analysis for N = 2 and v = 0 does not give points of
order two in esi„,. A slightly different and in some sense simpler analysis,
however, does.
(z,w)i—* (1 — z w). ,
(= = Q1Q2023 ,24 = i t%
More generally for any ceCk..) 13c1, '(e) is a canonical divisor. It follows
from these observations that the "gap" sequence at any 1=1» = 1,2,3,4) is
1, 2, 5.
(Note that the Pi are defined by z(Pi)4 = 1.) Thus each 13 is a Weierstrass
point of weight 2. We have accounted for 8 of the 24 Weierstrass points.
Similarly, the points Qi (j = 1,2,3,4) are Weierstrass points of weight 2, since
dw dw dw
z2 '
is also a basis of Ye l (M). Finally, the points Q. (j = 5,6,7,8) are also
Weierstrass points of weight 2, since
z 1)
(z,w) 1—*(—, —
ww
is an automorphism (of period 2) of the surface M that interchanges the
zeros and poles of w. We have thus accounted for all the Weierstrass points
on M. It is a surface of genus 3 with 12 distinct Weierstrass points each
of weight 2.
VII.3.9. The next to last example of this section is the most complicated
one. It will be used to show that the upper bound on the number of (distinct)
Weierstrass points g3 — g on a surface of genus g (Theorem 111.5.11) is
324 VII Examples
On the surface, the function z is of degree 7 (and w is of degree 3). The function
z is ramified over 0, 1, co (with ramification number 7). A calculation using
Riemann-Hurwitz now shows that M has genus 3. In our usual short hand:
-
Pi
(0=
P7131
(d-) =
Qt
and
P i Pj
I'
(w)= —
42 31
From the above formulas we see that the differentials
dz dz dz
(z — 1)-6 (3.9.2)
JT is given by
(z,w)i— ■ fez, — e 3 w),
so that the only possible fixed points of JT are P, = z -1 (0) or Qt, Q z the
two points lying over co. Since T is of prime order 5, the Riemann-Hurwitz
formula gives 2 = 5(2 4- — 2) + 4v(T), with v(T) as usual the number of fixed
points of T. The only way this can be satisfied is with g = 0 and v(T) -- -, 3.
Hence T fixes P 1 , Q 1 , and Q2. It is therefore obvious that JT fixes P 1 , but
cannot fix either Q, or Q2, because JTQ, = JQ, = Q2 and JTQ 2 =
= Q1.
The reader should now recall Theorem V.2.11.
Pj+ 2 = Z j = 1, . , 2g — I, P2 9 +2 = Z
We have seen in VII.1.2, that using P1 as a base point for 9: M J(M), the
vector K of Riemann constants is given by
- g
K= —
1 g—1 1
+1
2
1 1
where 17 is the period matrix for the canonica homology basis constructed
in VIL 1.1. We shall show that the Ai are holomorphic functions of the entries
rcki of H. We will accomplish this by expressing the function z in terms of
0-functions.
The function z e or(M) is characterized (up to a constant multiple) by
the property that it has a double pole at P29+ 2, a double zero at 1' 1 , and is
VII.4. Branch Points of Hyperelliptic Surfaces 327
by virtue of (1.2.1) and the fact that ç(P2) = 0 for every Weierstrass point
P on M. We have computed (p(P3) in
= Vie t) + et." 71- e(2)).
Similarly,
(P(P20 +2P5P7 • • P29+1) K = 9(P29 . 2 P3) = lop + e(2)),
We also observe that
i(P1P5P7 • • • P29+ I) = = i(P2 9 +2P5P7 • ' • P2g+t)
According to Theorem VI.2.4, the numerator vanishes precisely (it does not
vanish identically by Theorem VI.3.3) at P1, P5, P7,. . . , 1329+1 and the
denominator at P29+2, P5, P7,. . . , P2g _, 1 . In particular, the function (4.2.1)
vanishes to first order at P, and has a simple pole at P20. 2 . Examining the
multiplicative behavior of the above function, we see that
02 [1 oo • • •
1 1 O • -
f(P)=
02 F1 o o •
Lo 1 0 • •
f = cz, c e C\(01.
The constant c is evaluated by f(P 2) = cz(P2) = c. Thus we see that
1 1 0 • •• 01 (0p2),H)
o2[
1 1 0 • -
= f(P2) = 02[1 0 0 • • 0 2,
0 1 0 • •• OI<P(P ) H)
328 VII Examples
and
1 0 0 — 0 1 1 (i ) 1 0 0 ol
0' (0(11,././)
0 1 0 • • ' 0] ( .,.. 7 . 11) o2 [1 1 0 0j
,111 — -
1 0 0 . - • 01 1 100 01
O2 — re ,1 11)02
[1 • [0 1 0 0 (9(P),n)
0 0 •• - 0-1 02 1- 1 0 0 •• • 01
02r_
O I 0 • - 0] Li 1 o — 0
(4.2.2)
0 0 0 ••• 01 192 r1 0 0 •••
02 (P
1 1 0 •• • 0 _I LO 1 0 • • • 00] (P),17 )
02 [0 0 0 ..• 0] [0 0 0
0-,
LO 1 0 0 0 0 0 0.]
—
00 0 0 ••• 01 02 [0 0 0-,_
02[
0••• 0 0 0,
By replacing the point of order 2 that started this whole procedure (for
example, use
49(PiP3P7* • • P2g + K
instead of
0(P IP5P7 P29+ 1 ) K),
we can get similar formulae for for the other values of j. We have hence
established the following
VII.4.3. The fact that there are many ways to express the function z in
terms of 0-functions leads to useful and interesting relations among 0-
constants. We will, however, not pursue this fascinating subject.
V11.5. Examples of Prym Differentials 329
the differentials
dz dz , dz
- z . . . , z- - - —
w w
form a basis for abelian differentials of the first kind.
On M we can construct (locally) the function y given by
= z(z — 1) n (z — 4),
2g - 3
k. I
15. Fuchs. J. W. H. J.: Topics in the theory of functions of one complex variable. Van
Nostrand: Princeton, New Jersey 1967
16. Griffiths J. P., Harris, J.: Principles of algebraic geometry. Wiley: New York 1978
17. Gunning, R. C.: Lectures on Rieinann surfaces. Mathematical N&-:es. Princeton
Univ. Press: Princeton, New Jersey 1967
18. Gunning, R. C.: Lectures on vector bundles over Riemann surfaces. Mathematical
Notes. Princeton Univ. Press: Princeton, New Jersey 1967
19. Gunning, R. C.: Lectures on Riemann surfaces: Jacobi varieties. Mathematical
Notes.Princeton Univ. Press: Princeton, New Jersey 1972
20. Gunning, R. C.: Riemann surfaces and generalized theta functions. Springer:
New York, Heidelberg, Berlin 1976
21. Hensel, K., Landsberi, G.: Theorie der algebraishen Funktionen einer Variabeln.
Teubner: Leipzig 1902
1 2. Hurwitz, A., Courant, R.: Vorlesungen über allgemeine Funktionen theorie und
elliptische Funktionen. Springer: Berlin 1929
23. Igusa,1-1.: Theta functions. Springer: New York. Heidelberg, Berlin 1972
24. Klein, F.: Über Riemann's Theorie der algebraischen Funktionen und ihrer
Integrale. Teubner: Leipzig 1882
25. Kra, I.: Automorphic forms and Kleinian groups. Benjamin: Reading, Massa-
chusetts 1972
26. Krazer, A.: Lehrbuch der Thetafunktionen. Teubner: Leipzig 1903
27. Krushkal, S. L.: Quasiconformal mappings and Riemann surfaces. Winston &
Sons: Washington, D.C. 1979
28. Kunzi, H. P.: Quasikonforrne Abbilduntlen: Springer: Berlin, GOttingen, Heidel-
berg 1960
29. Lang, S.: Introduction to algebraic and abelian functions. Addison—Wesley:
Reading, Massachusetts 1972
30. Lang, S.: Elliptic functions. Addison—Wesley: Reading, Massachusetts 1973
31. Lehner, J.: Discontinuous groups and automorphic functions. Mathematical
Surveys, Number VIII. American Mathematical Society : Providence, Rhode
Island 1964
32. Lehner, J.: A short course in automorphic funetions. Holt: New York 1966
33. Lehto, O., Virtanen, K. I.: Quasiconformal mappings in the plane. Springer:
New York, Heidelberg, Berlin 1973
34. Magnus, J. W.: Noneuclidean tesselations and their groups. Academic Press:
New York 1974
35. Mumford, D.: Curves and their Jacobians. The Univ. of Michigan Press: Ann
Arbor 1975
36. Nevanlinna, R.: Uniformisierung. Springer: Berlin, Gottingen, Heidelberg 1953
37. Pfluger, A.: Theorie der Riemannschen Fliichen. Springer: Berlin, Gottingen,
Heidelberg 1957
38. Rauch, H. E., Farkas, H. M.: Theta functions with applications to Riemann
surfaces. Williams & Wilkins: Baltimore, Maryland 1974
39. Rauch, H. E., Lebowitz, A.: Elliptic functions, theta functions, and Riemann
surfaces. Williams & Wilkins: Baltimore, Maryland 1973
40. Riemann, B.: Gesammelte Mathematische Werke. Dover: New York 1953
41. Schiffer, M., Spencer, D. C.: Functionals of finite Riemann surfaces. Princeton
Univ. Press: Princeton, New Jersey 1954
42. Springer, G.: Introduction to Riemann surfaces. Addison—Wesley: Reading,
Massachusetts 1957
43. Siegel, C. L.: Topics in complex function theory. Wiley—Interscience, New York.
Vol. I, 1969, Elliptic functions and uniformization theory. Vol. II, 1971, Auto-
morphic functions and abelian integrals. Vol. III, 1973, Abelian functions and
modular functions of several variables
332 Bibliography