You are on page 1of 197

Introduction to Arakelov Theory

BOOKS OF INTEREST BY SERGE LANG

Geometry: A High School Course (with Gene Morrow)


This high school text, inspired by a researcher's educational interests and Gene
Morrow's experience as a high school teacher, presents geometry to the student in
an exemplary, accessible, and attractive form. The book emphasizes both the intel-
lectually stimulating parts of geometry, including physical and classical applications,
and routine arguments or computations.
MATH! Encounters with High School Students
This book is a faithful record of dialogues between Lang and high school students,
covering some of the topics in the Geometry book and others of the same mathemat-
ical level (areas and volumes of classical figures, pythagorean triples, infinities).
These encounters have been transcribed from tapes, and thus are true, authentic, and
alive.
The Beauty of Doing Mathematics: Three Public Dialogues
Here we have three dialogues given in three successive years between Lang and au-
diences at the Palais de la Decouverte (science museum) in Paris. The audiences
consisted of many types of persons, including some high school students, but pri-
marily people who just drop in at the Palais on a Saturday afternoon. Lang did
mathematics with them: prime numbers, elliptic curves, and Thurston's conjecture
on the classification of 3-manifolds. Again the encounters were taped, recording
Lang's success at putting a lay audience in contact with basic research problems of
mathematics. This was done by using a selection of topics which could be explained
from scratch in a self-contained way. The enthusiasm of the audience is partly evi-
denced by the fact that, the third time, more than 100 persons stayed for three and
one-half hours, dealing with the Thurston material.
Fundamentals of Diophantine Geometry
A systematic account of fundamentals, including the basic theory of heights, Roth
and Siegel's theorems, the Neron-Tate quadratic form, the Mordell-Weil theorem,
Weil and Neron functions, and the canonical form on a curve as it relates to the Ja-
cobian via the theta functions.
Introduction to Complex Hyperbolic Spaces
Since its introduction by Kobayashi, the theory of complex hyperbolic spaces has
progressed considerably. This book gives an account of some of the most important
results, such as Brody's theorem, hyperbolic imbeddings, curvature properties, and
some Nevalinna theory. It also includes Cartan's proof for the Second Main Theo-
rem, which was elegant and short.

*
OTHER BOOKS BY LANG PUBLISHED BY SPRINGER-VERLAG
Introduction to Arakelov Theory· Riemann-Roch Algebra (with William Fulton) •
Complex Multiplication • Introduction to Modular Forms • Elliptic Curves: Dio-
phantine Analysis· Modular Units (with Daniel Kubert) • Introduction to Algebraic
and Abelian Functions· Cyclotomic Fields • Elliptic Functions • Algebraic Num-
ber Theory· SL 2 (R) • Abelian Varieties • Differential Manifolds· Complex
Analysis· Undergraduate Analysis • Undergraduate Algebra· Linear Algebra •
Introduction to Linear Algebra· Calculus of Several Variables· First Course in
Calculus • Basic Mathematics· THE FILE
Serge Lang

Introduction to
Arakelov Theory

Springer Science+Business Media, LLC


Serge Lang
Department of Mathematics
Yale University
New Haven, CT 06520
U.S.A.

Mathematics Subject Classifications (1980): 11G05, 11010

Library of Congress Cataloging-in-Publication Data


Lang, Serge
Introduction to Arakelov theory I Serge Lang.
p. em.
Bibliography: p.
Includes index.
ISBN 978-1-4612-6991-5 ISBN 978-1-4612-1031-3 (eBook)
DOI 10.1007/978-1-4612-1031-3
1. Arithmetical algebraie geometry. 1. Title. II. Title:
Arakelov theory.
QA242.5.L36 1988
512'.72--de 19 88-15952

© 1988 by Springer Science+Business MediaNew York


Originally published by Springer-Verlag New York Inc. in 1985
Softcover reprint ofthe hardcover Ist edition 1985
AII rights reserved. This work may not be translated or copied in whole or in part without
the written permission ofthe publisher (Springer Science+Business Media, LLC),
except for brief excerpts in connection with reviews or scholarly analysis.
Use in connection with any form of information storage and retrieval, electronic adaptation,
computer software, or by similar or dissimilar methodology now known or hereafter
developed is forbidden.
The use of general descriptive names, trade names, trademarks, etc. in this publication, even
ifthe former are not especially identified, is not to be taken as a sign that such names, as
understood by the Trade Marks and Merchandise Marks Act, may accordingly be used
freely by anyone.

Typeset by Composition House Ltd., Salisbury, England.

9 8 7 6 54 32 1

ISBN 978-1-4612-6991-5
Foreword

The purpose of this book is to give an introduction to Arakelov theory.


This theory consists in transposing results of algebraic geometry,
especially numerical results, to the number field case, by completing the
set of points (or divisors) to include those at infinity, that is, the archi-
medean places. This is done by considering hermitian structures in addi-
tion to sheaf structures, so analysis becomes number theory at infinity.
One can then get the analogous results in a number of cases. For a
curve over the ring of integers of a number field, one gets for instance
the adjunction formula; the Hodge Index Theorem, treated here as in
Hriljac's paper, which means identifying the Arakelov intersection
numbers with the Neron-Tate form and using the positive definiteness of
this form rather than reproving it; and the Faltings Riemann-Roch
theorem on arithmetic surfaces. The proof of Riemann·-Roch here de-
pends on the adjunction formula, but does not depend on semistability
assumptions which are made systematically in Faltings, and subsequent
expositions (sometimes only implicitly). Faltings' other results on the
Noether formula are much deeper; they depend on the moduli spaces,
and I did not feel I could include them in this book. I have, however,
included Faltings' theorem on the positivity of the dualizing sheaf in the
semistable case.
The Riemann-Roch theorem has a formulation not only in terms of a
numerical relationship but in terms of a metric isomorphism of certain
sheaves, due to Faltings for metrics with the canonical Chern form.
Deligne [De] has pointed to the possibility of dealing with arbitrary
metrics. I have not entered into this aspect of the theorem. On the
other hand, the residue theorem and the discussion of the metrics
involved may be viewed as carrying out this more general study for
vi FOREWORD

adjunction. Readers wanting to see different treatments of the sheaf pair-


ing leading to the more sheafy Riemann-Roch for admissible metrics can
also look up Moret-Bailly [MB], who assumes semistability (but a proof
can be given without, as in the present book). Cf. Remark 3 at the end
of Chapter V, §3.
In addition, the formulation of Riemann -Roch as a metric isomorph-
ism of line sheaves should be carried out over the moduli space, as
already noted by Faltings [Fa 1] when he states, p. 415: "Of course our
Riemann-Roch theorem applies only to a single curve and not to a
family. But the technique of the proof can be generalized." Faltings also
points out how such a generalized version has applications to a refined
study of the calculus on arithmetic surfaces, involving the Noether for-
mula and beyond. At the moment, 1 did not find these topics ripe
enough for me to expand the present book to include them. For one
thing, appropriate basic texts on the moduli scheme are not yet available
for reference, e.g. Chai-Faltings, [A-C-G-H] Volume II, etc. However,
1 thank Vojta for providing me with an appendix which explains how
Arakelov theory, suitably developed, implies certain diophantine inequali-
ties, and more generally how Vojta's basic diophantine conjectures are
related to bounds for the canonical sheaf (so-called dualizing sheaf). The
level of exposition of this appendix is considerably higher than for the
rest of the book, but an attempt has been made to make it as accessible
as possible by repeating some definitions and giving some discussion of
concepts involved. Readers should also be aware of the development of
the higher dimensional intersection theory by Gillet-Soule, and the ana-
lytic contributions of Quillen, Bost, and Bismut, in course of publication.
I have tried to make the exposition as self-contained as possible, refer-
ring to books and systematic expositions of more elementary material
whenever necessary. For the Neron-Tate form, the local Neron symbols,
and the construction of the Neron functions (metrics on line bundles) on
curves, as pull backs from the Jacobian via theta functions, I could refer
to my Fundamentals of Diophantine Geometry. For various topics in
analysis, I was able to refer to appropriate texts, for instance, Griffiths-
Harris to construct Green's functions (solving a aJ-equation), and to
other books for other facts about PDEs. As for algebraic geometry, I
give a proof of the residue theorem due to Kunz-Waldi, based on
Kunz's book Kiihler Differentials, which is part of basic algebraic geo-
metry, and allows one to see directly the algebraic part of the adjunction
formula in the arithmetic case. I find this extremely satisfactory. It
would also be nice if eventually expositions of residues and the dualizing
sheaf in the case of local complete intersections gave a proper account of
this theorem, which can also be viewed as imbedded in the Grothendieck
theory, but it turns out that duality theory was not needed here.
The situation with respect to semi stability is much more disagreeable.
I have listed the results required for Faltings' theorem in Chapter V, §5,
FOREWORD Vll

and I have included those proofs for which I could find no convenient
reference. I am indebted to Mike Artin for his guidance to the literature
(especially some of his early papers in this respect), as well as for other
useful suggestions. I decided against including here a subbook giving a
systematic account of this basic material, which would have thrown
everything off balance; but eventually the appropriate text on such ele-
mentary topics, giving a systematic account of the basic properties of
semi stability among other things, will have to be written. (This is not a
threat.)
There are also some results which are proved in elementary or basic
texts for the geometric case, e.g. surfaces or varieties over an algebraical-
ly closed field, and whose proofs hold in the more general case of a vari-
ety over a Dedekind ring or over a discrete valuation ring. For instance,
Fulton explicitly advises the reader that essentially his entire book on
intersection theory holds in this more general context. The practice up
to now has been to write such books in the geometric case, and to add
the comments about the more general validity a posteriori. My feeling is
that the time has come to build in the greater generality from the start.
For instance, I think Hartshorne in his book could have waited a bit
longer before he went geometric, and sacrificed the more general ground
schemes which occur in the first part of his book for the sake of ground
fields. If he had, I would not have had to point out in some instances
that such and such a proof is valid in the more general situation. The
point holds not only for the intersection theory on surfaces, but for ques-
tions of resolution of singularities, minimal models, and the like.
Fulton emphasized, rightly, the importance of doing intersection
theory on singular schemes. Again, Hartshorne is too quick in making
regularity assumptions when treating curves and surfaces. He could have
treated a lot of the more general cases without any additional space.
So I have had to make compromises, depending on ad hoc judgments.
I hope enough people agree with these judgments to make the present
book useful.

New Haven, 1988 SERGE LANG

Acknowledgement. I thank Bill McCallum for a careful reading of the


manuscript and his help in proofreading.
Contents

Foreword v

CHAPTER I
Metrics and Chern Forms. 1
§l. Neron Functions and Divisors . . . . . . . . . . . . . . . . 1
§2. Metrics on Line Sheaves . . . . . . . . . . . . . . . . . . . 4
§3. The Chern Form of a Metric . . . . . . . . . . . 10
§4. Chern Forms in the Case of Riemann Surfaces . . . . . . . . . 14

CHAPTER II
Green's Functions on Riemann Surfaces 20
§l. Green's Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 21
~2. The Canonical Green's Function . . . . . . . . . . . . . . . . . . . . . . .. 28
§3. Some Formulas About the Green's Function. . . . . . . . . . . . . 30
§4. Coleman's Proof for the Existence of Green's Function 34
§5. The Green's Function on Elliptic Curves . . . . . . . . . . . . . . . 42

CHAPTER III
Intersections on an Arithmetic Surface 48
§l. The Chow Groups . . . . . . . . 48
§2. Intersections. . . . . . . . . . . . . . 55
§3. Fibral Intersections. . . . . . . . . . 58
§4. Morphisms and Base Change . . . 62
§5. Neron Symbols. . . . . . . . . . . . 66
CHAPTER IV
Hodge Index Theorem and the Adjunction Formula. 70
§l. Arakelov Divisors and Intersections . . . . . . . 71
§2. The Hodge Index Theorem. . . . . . . . . . . . . . . . . . 77
x CONTENTS

§3. Metrized Line Sheaves and Intersections . . . . . . 80


§4. The Canonical Sheaf and the Residue Theorem .. 87
§5. Metrizations and Arake1ov's Adjunction Formula. 97

CHAPTER V
The FaIlings Riemann-Roch Theorem . . . . . . . . 102
§1. Riemann Roch on an Arithmetic Curve. 102
§2. Volume Exact Sequences . . . . . . 105
§3. Faltings Riemann-Roch . . . . . . . 108
§4. An Application of Riemann-Roch. 118
§5. Semistability . . . . . . . . . . . . . . 120
§6. Positivity of the Canonical Sheaf. 128

CHAPTER VI
Faltings Volumes on Cohomology. 131
§1. Determinants . . . . . . . . . . . . . 131
§2. Determinant of Cohomology . . . . . 136
§3. Existence of the Faltings Volumes . 140
§4. Estimates for the Faltings Volumes. 146
§5. A Lower Bound for Green's Functions 149

APPENDIX by Paul Vojla


Diophantine Inequalities and Arakelov Theory. 155
§1. General Introductory Notions ... 156
§2. Theorems over Function Fields . 160
§3. Conjectures over Number Fields 163
§4. Another Height Inequality. 173
§5. Applications. . . .. . . . . . . . 176

References. . . . . . 179
Frequently Used Symbols. 183
Index . . . . . . . . . . . . . . 185
CHAPTER

Metrics and Chern Forms

In this first chapter we establish the language of metrics on invertible


sheaves, which is equivalent with the language of Neron functions or
Neron divisors. Classically, over the complex numbers, given a divisor
on a complex non-singular variety, one looks for a real-valued function g
which has a logarithmic singularity on the divisor, and satisfies some
PDE condition on aag. Such conditions can to some extent be expressed
algebraically over any complete, algebraically closed valued field, thus
giving rise to Neron functions. We shall give a description of the analytic
conditions precisely on Riemann surfaces, specializing Griffiths-Harris to
the case of dimension 1, in the language of metrics.
To shorten terminology, we shall use the words line sheaf instead of
invertible sheaf. Similarly, one might use the words vector sheaf to de-
note a locally free sheaf of finite rank.
In light of the use of certain functions in diophantine approximations
or Nevanlinna theory, I have given equal weight to Weil-Neron func-
tions or metrics in the exposition of this first chapter.

I, §1. NERON FUNCTIONS AND DIVISORS


We let X be a variety defined over an algebraically closed field k, com-
plete under an absolute value v. "Points" are in X(k). By convention,
for Z E k we define
v(Z) = -logizi.
Let D be a Cartier divisor on X. By a Weil function AD associated
with D we mean a function
A. D : X(k) - supp(D) -+ R
2 METRICS AND CHERN FORMS [I, §1]

from the complement of the support of D into the reals, satisfying the
following condition: If D is represented by a pair (VJ), where V is an
open set (Zariski open, that is), and f is a rational function, then there
exists a continuous function oc on V such that

AD(P) =v 0 f(P) + oc(P) for all P E V - suppeD).

Thus a Weil function has a logarithmic singularity along the divisor. We


do not require a Weil function to be smooth. For a number of diophan-
tine estimates, Weil functions are useful because we want to take sups
and infs of continuous functions. We call D the divisor of AD.

Example. Let X = pl and let D = (00) be the point at infinity. Let z


be the variable on k. Then

A(Z) = max(O, 10glzl) = log+ Izl

is a Weil function associated with D. This type of function, and similar


ones on higher dimensional projective spaces or varieties come up all the
time in approximation theories (diophantine or Nevanlinna, for instance).
On the other hand, to study Weil functions approximately, i.e.
mod 0(1), we actually have to define a more precise notion, that of
Neron function, and what it corresponds to in the context of metrized
line bundles or line sheaves. We shall now carry out the basic definitions
in this direction.

Let D be a Cartier divisor on X. On a covering family of open (Zar-


iski) sets {Vi} we can represent D by rational functions {fa or by the
pairs {(Vi'};)}. Instead of such pairs, we consider triples (V,/, oc) where

IX: V-+R

is a continuous function. (Or we could also take oc to be real analytic,


and we qualify accordingly.) Two such triples (V,/, IX) and (V, g, fJ) are
called compatible if the pairs (U,n and (U, g) represent the Cartier di-
visor D, so fg-l is a unit in iD(U n V), and if in addition

on Un V.

Note that v (jg-l)(p) = -logl(jg-l)(P)1 for P E Un V. A maximal


0

family of compatible triples, or a compatibility class of triples whose


open sets cover X is called a Neron divisor, whose Cartier divisor is D.
If we use the letter D for a Neron divisor, then we write (D) for its asso-
ciated Cartier divisor.
[I, §1] NERON FUNCTIONS AND DIVISORS 3

Neron divisors form a group, the group law being obtained by multi-
plying the rational functions f and adding the functions IX.
On the complement of the support of D (by definition, the support of
its Cartier divisor), we can define the Wei! function

AD(P) = v f(P)
0 + IX(P),
where (U,f, IX) is a representing triple for D, and P lies in U. The value
is independent of the choice of triple, as one sees directly from the defini-
tion of compatibility of triples.
Let f be a rational function on X, f #- O. We can define a Weil func-
tion Af to be AiP) = v f(P). The associated Neron divisor is repre-
0

sented on every open set U by (U, f, 0). We call Af or the associated


Neron divisor principal.
By a Neron family we mean an association

from Cartier divisors to Weil functions, satisfying the following condi-


tions:

NF 1. The map D ....... AD is a homomorphism.

NF 2. If D = (f) is the divisor of a rational function, then

AD = V 0 f - constant.

NF 3. If (U, f) represents D, then there exists a continuous (resp. real


analytic) function IX on U such that

AD = va f + IX.
NF 4. If X is of dimension 1, then for any two points P #- Q,

We can then define the function A on the product X x X outside of the


diagonal by

Then A is called the Neron function associated with the family. Converse-
ly, as we shall see at the end of §2, a Neron family on a curve can be
constructed from a function on X x X with divisor equal to the dia-
gonal, over the complex numbers. See the remarks at the end of §2.
4 METRICS AND CHERN FORMS [I, §2]

For any Neron family P'D}, if D = (f) IS the divisor of a rational


function, then we use the notation

2(f) = v 0 f - roCf),
so we denote by rif) the constant of NF 2. This constant will play an
essential role in the sequel, in defining the Arakelov divisor of a function
with respect to the given choice of Neron family. In fact, we shall see in
Chapter IV, §2, that this constant is the multiplicity at infinity of this
divisor. We can give a formula for rv(f) over the complex numbers as
follows. Suppose k = C, and J1 is a measure on X(k).

If J1 and ). are normalized such that

fX(k)
dJ1 = 1 and fQEX(k)
2(P, Q) dJ1(Q) = 0,

f
then
roCf) = -loglflv dJ1.
X(k)

The relation follows trivially from the normalizations and the definitions.

I, §2. METRICS ON LINE SHEAVES

Let fE be a line sheaf on X. Let {V;} be a covering of X such that


fEl Vi is free over (i)xl Vi' Let

be an isomorphism, i.e. a trivialization. Then

is an isomorphism, which can be identified with multiplication by a unit


in (i)(Vi n V j )*, also denoted by CPij' Note that {cpij} is the usual Cech
l-cocycle associated with the line sheaf.
Let s be a section of fE over some open set V. Under a trivialization
CPi as above, the section corresponds to a section of (i)x over Vj n V, in
other words, an element of (i)x(V n Vi) which we may denote by Si'
Then
[I,§2] METRICS ON LINE SHEAVES 5

Recall that if D is a Cartier divisor represented by the function h on


Vi' then the sheaf (9x(D) is such that over the open set Vj' the rational
function Ij-1 is a basis for (9x(D)(V j) over (9x(VJ Suppose 2 = (9x(D).
Then ({)i may be chosen to be multiplication by Ii' In such a case we
identify ({)j with h.

Example. Let X = pn and let 2 = (9p(l). Let Vi be the complement


of the hyperplane T; = 0, where To, ... :r,. are the variables. Then we can
take Si = TolT; to be the section whose divisor is {To = OJ, and then

({)ij = ~/T;.

A metric p on 2 amounts to the giving of a norm, or rather the


square of a norm, on each fiber, varying as smoothly as the conditions
prescribe (continuously, Coo, real analytic, ad lib). We define "varying
smoothly" by using charts. Suppose the family {(Vj, ({)jj)} respresents 2.
Suppose given for each i a function

such that on Vi n Vj we have

Then we say that the family of triples {(Vj, ({)jj' pJ} represents the metric.
It is clear how to define compatible families, or compatible triples in this
context, and the metric itself is an equivalence class of such covering
triples, or is the maximal family of compatible triples. We could also
write a representative family as {(Vj' ({)j, Pj)} using the isomorphisms ({)j
instead of the transition functions ({)jj. A line sheaf with a metric p will
be denoted by 2 p or as a pair (2, p), and will be called a metrized line
sheaf.

If S is a section of 2 over some open set V j containing a point P,


then we define

The value on the right-hand side is independent of the choice of Vj, as


one sees at once from the transformation law.
At a given point P, we have the stalk 2 p, which is a one-dimensional
free module over the local ring (9p, and we let

(mp = maximal ideal)


6 METRICS AND CHERN FORMS [I, §2]

be the fiber at P. Then 2'(P) is a one-dimensional vector space over the


ground field k. A metric P on 2' induces a norm on 2'(P)' Indeed, if
s E m p 2' p so s(P) = 0, then
ISi(P)1 = O.

For arbitrary s in 2' p the value Is(P)l p depends only on the residue class
of s mod m p 2' p, thus defining the norm on the fiber.

Example: the standard metric. Let X = pn = P be projective n-space.


Let To, ... , T" be the variables, let Vi be the open set T; -# 0, and let

be complex coordinates on Vi together with the constant 1. Then the


standard metric on lD p (1) is defined by the functions

Pi(Z) = I" Iz~i)12 = I" 11j/T;1 2 •


j=O j=O
on Vi'

Inverse

Let 2' be a line sheaf on X. If {(Vi' <Pij, Pi)} represents a metric P on


2', then {(V i ,<Pijl,Pi- 1 )} represents a metric on 2'-1, which will be
called the inverse, or dual metric. We also denote the inverse 2' - 1 by
!f! v.

Tensor product

Let 2', JI be line sheaves. Suppose that {(Vi' <Pij, p;)} represents a met-
ric on 2' and {(Vi> !/Iij' a)} represents a metric on JI. Then there exists
a unique metric on 2' ® JI represented by

Of course, two different covering families of open sets might at first rep-
resent the two metrics, but one can always find a common refinement,
and the above definition applies to define the tensor product metric, or as
we shall also say the product metric.

Functoriality

Let f: X' ~ X be a morphism, and let 2'p be a metrized line sheaf on X.


Then we can define in a natural way a metrized line sheaf, the pull back
f* 2'p
[I, §2] METRICS ON LINE SHEAVES 7

on X'. If 2 p is represented by {(VI> <Pii' Pi)} then f* 2 p is represented


by

Proposition 2.1. Let X be a quasi-projective variety, and let 2 be a


line sheaf on X. Then 2 admits a metric.

Proof There exist very ample sheaves vIt and vIt' such that

by elementary algebraic geometry. To say that vIt is very ample means


that vIt = (i)x(1) where the twist is taken relative to a projective imbed-
ding. Then we can give vIt the pull back metric from the standard met-
ric on (!)p(1), and similarly for viti. Taking the inverse and tensor
product gives a metric on 2 as desired.

Remark. Since the standard metric is as smooth as is possible (real


analytic over the complex numbers) it follows that its pull back is just
as smooth. Thus the metric on 2 in Proposition 2.1 is as smooth as
possible.

Let P be a metric on 2. Let s be a rational section of 2 with divisor


D. We also write this (s) = D. We define the Weil function associated
with the metric and s to be

A•. p(P) = -logls(P)l p , P¢supp(D).

Then in particular we can apply the above construction to the sheaf


2 = (!)x(D) with the rational section 1.
Suppose that X is complete, i.e. proper over k. Then two sections
having the same divisor differ by multiplication with a non-zero constant,
so their associated Weil functions differ by an additive constant.
Conversely, let AD be a Weil function with Cartier divisor D. Then we
can define a metric Pl.D on (!)x(D) as follows. Suppose {(Vi'};)} repre-
sents D. We then have the continuous function (Xi on Vi such that

for P E Vi and P rt suppeD). We may then define Pi on Vi by

(Xj(P) = t log plP) or

Thus we obtain a bijection between Weil functions whose divisor is D,


and metrics on the line sheaf (!)x(D).
8 METRICS AND CHERN FORMS [J, §2]

In the case when!£' = (!)x(D), and (U i ,/;, p;) represents a metric on !£'
in the neighborhood of a point P, the metric on the residue class field
(!)x(D)(p) can. be computed from the representative of the metric on U i,
and k(P) can be identified with (!)p/m p under the trivialization

given by fl-+ fl;·

The residue class field (!)x(D)(p) is merely the restriction (!)x(D)IP. In the
function field k(X) we have the function 1 which is a rational section of
(!)x(D), and is defined outside the support of D with value 1 in the re-
sidue class field. We can write I(P) for the value of this section at P. If
P ¢ suppeD), then

II(P)I; = II;(PW/Pi(P),

However, for P outside the support of D there is a representative triple


= 1. In this case we get simply
(U i , 1, Pi) with I;

and therefore we find:

Proposition 2.2. Let It be a Neron family DI-+A D. If P is a point not


lying in the support of D then

II(P)lpA..D = exp( -AD(P»,

Let f be a rational function, and let D, E be two Cartier divisors. Let


(!)x(D) and (!)x(E) be the corresponding line sheaves. Let P;',D and P;'.E be
the me tries associated with the Weil functions AD and AE.

Proposition 2.3. Suppose E = D + (f). Multiplication by f- 1 is a met-


ric isomorphism of (!)x(D) with metric P;',D and (!)x(E) with metric P;'.E if
and only if

or in the terminology of §1, y(f) = 0, where Y(f) = v f - 0 )"(J)'

Proof This is immediate from the definitions.

Note that for any constant c -1= 0 the divisors (f) and (cf) are equal.
Thus in a given class of functions modulo multiplication by a non-zero
constant, there is a unique system of multiples cf, with c determined up
to a scalar of absolute value 1, such that multiplication by cf is a metric
isomorphism between P;',D and P;',E'
[I, §2] METRICS ON LINE SHEAVES 9

In Chapter IV, §3, we shall deal with Arakelov divisors, and then we
shall recover the expected isomorphism under multiplication by f - 1.

Example: the case of curves

Suppose that X is a non-singular complete curve, defined over k.

Divisors on X can be identified with formal linear combinations of


points. One way of constructing a Neron family is to consider the sur-
face X x X. We let A be the diagonal, which is a divisor on X x X.
Let us fix a metric on (!Jx x x(A) with its associated Weil function A.1.
Since X x X has an automorphism of order 2, the symmetry inverting
the two factors, we can symmetrize the metric so that it is invariant
under this automorphism. If the metric is symmetric, then it follows that
the associated Weil function A.1 is symmetric.
For each point P E X(k), we can then imbed X in X x X by
X -+P X X,

and the pull back of A.1 is a Weil function Ap on X associated with the
divisor (P) and with the induced metric. Furthermore, the association

extended to arbitrary divisors by additivity satisfies the properties of a


Neron family except possibly for NF 2, that is, if D = (f) is globally the
divisor of a rational function, then

AD = V 0 f + constant.
In the next section and next chapter we shall see how to construct a
function on X x X which is the Weil function associated with the dia-
gonal, and such that this last condition NF 2 is also satisfied. To do
this, one requires an additional condition to normalize Weil functions up
to a constant, not only up to bounded functions. Over the complex
numbers, this particular condition will essentially be harmonicity, and we
shall use the fact that a harmonic function on a compact complex mani-
fold is constant. See Proposition 3.1, which is the heart of the matter.

For the rest of this chapter, and the entire Chapter II, we carry out
certain basic complex differential geometry, whose sole purpose is to
prove the existence of a Neron family, and certain admissible metrics
for use in the Hodge Index Theorem, the adjunction formula, and the
Riemann-Roch theorem. If the reader is willing to grant this existence,
and is interested only in the algebraic-geometric part of the proof, then
the rest of this chapter and Chapter II may be omitted.
10 METRICS AND CHERN FORMS [I, §3]

I, §3. THE CHERN FORM OF A METRIC


From now on, we suppose that we work over the complex numbers and
that the variety X is non-singular, so is a complex manifold, identifying
X with X(C).

For the convenience of the reader, we make a short table of the usual
0, a.First in terms of one complex analytic coordinate.
Let z = x + J"=1y be a local analytic coordinate. Then we have the
differential operators

~=~(~-J=1~)
oz2 ox oy
and ~ = ~2
oi
(i + J"=1 ~-).
ox oy

For any function I, we define

- 01_
and 01 = oi dz ,

so 0 and asend functions to I-forms. One step further yields

1
=- (0-21+02~
- dx A dy
2 ox 2 oy2
= ! III dx A dy,

where Il is the Laplacian, given in real coordinates by the usual expres-


sion. Note that

where d is the ordinary exterior derivative of differential forms.


We define the real operator de by

de =
J"=1 -
- - ( 0 - 0) = ----~-
1 (0 - 8)
4n 2nJ"=1 2
so that
[I, §3] THE CHERN FORM OF A METRIC 11

We shall use de in polar coordinates, which is

1 af 1 1 af
d<j = - - r de - - - - dr
4n ar 4n r ae

When we integrate around a circle, then the term with dr is O.

Of course, we have similar formulas in several variables. If Zl"" ,Zn


are complex coordinates, then we have the partial derivatives alaz i and
alazi • If h is a function, then

ah - ah
ah = I,-dz. and ah = I, a-Zi dz i •
aZ i '

As in one variables, we have

and 1
d e = -----
2n~
(a -
------.
2
a:]
By a form of type (1, 1) we mean a differential form which locally at
every point can be expressed as

where fij are smooth functions. If there is such a representation with re-
spect to one system of complex coordinates, then the form has such a
representation with respect to every choice of complex coordinates. One
defines forms of type (p, q) similarly, namely they are forms which locally
can be expressed as sums

"fIJ(z)
L.. dz·~l A'" A dz·'p A dz·Jl A'" AdzJq..

Now let f£ be a metrized line sheaf with COO metric p, represented by


the family {CUi' qJij' Pi)}' We define the Chern form of the metric to be

Since qJij is holomorphic non-zero, it follows that log ({Jij + log iPij is the
sum of holomorphic and anti-holomorphic terms, so
12 METRICS AND CHERN FORMS [I,§3]

Consequently the value dd log Pi is equal to dd c log Pj on Vi n V j , and


C

thus we have defined a global form c 1 (p), called the Chern form. This
Chern form is of type (1, 1).
The Chern class c 1 (2') is defined to be the class of c 1 (p) in H~R (de
Rham cohomology), which is the space of closed forms modulo exact
forms.
We recall that if s is a holomorphic section on an open set V repre-
sented by the function Si on Vi then

Therefore outside the support of s, we can write

independently of any coordinate representation, directly on the line sheaf.


We say that a (1, I)-form

is positive and we write w > 0 if the matrix h = (hij) is hermitian positive


for all values of z. By convention, here, positive will mean positive defin-
ite. We say explicitly semipositive if we want this weaker condition. The
condition is independent of the choice of holomorphic coordinates Zi.
A metric P is called positive if c 1(p) > o.
If X has dimension 1 then locally
;--.-
v' -1
w = ~ h dz 1\ dz.

We say that w is real if h is real valued. All these properties do not de-
pend on the choice of hoI om orphic coordinates.

Change of metrics

Directly from the definition, we see that the map

or

is a homomorphism from metrized line sheaves (under tensor product) to


(1, I)-forms on X.
[I, §3] THE CHERN FORM OF A METRIC 13

Let p, p' be COO metrics on the line sheaf 2, represented by


{(Vj, CfJjj' pJ} and {(Vj, CfJjj' p;)} respectively. Then

is a COO-function, independent of the choice of V j because the ICfJjj l2


cancel in the transformation law. Consequently there exists a function
IX E Coo(X, R) such that

Proposition 3.1. Assume that X is a compact complex connected mani-


fold. Let p, p' be metrics on a line sheaf having the same Chern form.
Then there exists a positive number y such that p' = yp"

Proof. Let a = p'I p. Then dd C log a = 0, and this equation holds for
the restriction of log a locally to every imbedded disc in X. Let P be a
maximum for log a. Then the restriction of log a to every imbedded disc
centered at P is harmonic, and so is constant since P is a maximum.
Hence log a is constant, as desired.

Proposition 3.2. Let f: X' ~ X be a morphism, and let 2 be a line


sheaf on X with positive metric p. Then f*p is semidefinite on f* 2. If
f is an imbedding, then f*p is definite, i.e. positive (according to our
convention).

Proof. This is an easy computation which is left to the reader.

Remark. Suppose JI = 2®m is the m-th tensor power of a line sheaf,


and that JI has a positive metric p. Then 2 has a positive metric pl/m.
This is useful in practice, when 2 is ample and JI very ample, and we
have found a positive metric on JI as in the next two propositions.
Taking the m-th root then defines a positive metric on fl.'.

Metrics on a sheaf on a variety will often be obtained by a pull back


from projective space pn. Let To, ... ,Tn be the projective coordinates,
and on the given open set Vj, complement of T j = 0, let

be complex coordinates together with the function 1. The standard met-


ric was defined on V j by
n

p(z) = L z.z•.
• =0
14 METRICS AND CHERN FORMS [I, §4]

Then its Chern form is computed using rules from freshman calculus for
the derivative of a product and quotient, to give

c 1 (p) =
J-1
2
- J=I
aa log p(z) = -
1
--2 (L h ij dZ j /\
_
dz)
n 2n p(Z)

where (hi) is the matrix

H = p(z)I - (ZjZ),

Proposition 3.3. The standard metric is positive.

Proof. It suffices to show that H is positive. For any complex vector


C = (co,," .c n ) we expand CH'C. and the Schwarz inequality immedi-
ately shows that for C =I- 0 and C j = 0 we have

CH'C > O.

This proves the proposition.

Proposition 3.4. Let !f! be an ample invertible sheaf on a variety X.


Then 51! admits a positive metric.

Proof. A sufficiently large positive multiple of 51! is very ample, and is


the pull back of (I:'p(1). The pull back of the standard metric to 51! gives
the desired positive metric.
Conversely, we have:

Kodaira's theorem. If 51! is a line sheaf on a complete non-singular vari-


ety, and 51! admits a positive metric, then 51! is ample.

This is more difficult to prove, cf. Griffiths-Harris, Chapter I, §4.

I, §4. CHERN FORMS IN THE CASE OF


RIEMANN SURFACES
In this sectioll, except at the end, we assume that X is a complete non-
singular variety of dimension lover the complex numbers, so a compact
Riemann surface.

We interpret some of the results in the present geometric context, and


indicate a general proof for the existence of Green's functions associated
with a divisor.
[I, §4J CHERN FORMS IN THE CASE OF RIEMANN SURFACES 15

Proposition 4.1. Let (J' be a metric on a line sheaf 2 on X. Then

Proof. This relation is given as an example of a higher dimensional


result in Griffiths-Harris, p. 144. One can repeat the proof in the case of
Riemann surfaces, when it becomes quite simple. Let s be a rational sec-
tion, so outside the support of s,

At each point P where s has a zero or pole, we put a small circle C(P, 8)
of radius 8 with respect to some choice of holomorphic coordinate, and
we apply Stokes' theorem to the complements of these discs. Then

At a point P, we can represent lsi; = fJh where f is mcromorphic at P


and h is smooth positive. Then

(0 - a) log lsi; = (0 - a) logf + (0 - a) 10gJ + (0 - a) log h.

But (o-a)logh involves h-1oh or h-1ah which is bounded locally,


and so the integral of this term tends to 0 as 8 ..... O. Also log f = 0 a
and 0 log J = O. Hence the integral at P except for a constant factor
amounts to

J=i 1m fC(P,t)
0 logf = J=i 1m fC(P,t)
f'lf(z) dz = 2nJ=im(P),

where m(P) is the multiplicity of s at the point P. One verifies at once


that the constant factors come out precisely so as to make the formula
of the proposition come out right, as desired.

Theorem 4.2. Let 2 be a line sheaf on X and let ({J be a real (1, 1)-
form such that Ix
({J = deg 2. Then there exists a metric p on It' such
that

and any two such metrics differ by a constant factor.


16 METRICS AND CHERN FORMS [I, §4]

Proof. The uniqueness up to a constant factor comes from Proposition


3.1. As for existence, let (J be some metric on 2. It will suffice to find a
real e"'-function fJ on X such that

because we then let p = eP(J. To solve this equation, we need two lem-
mas. The first is a special case of p. 149 of Griffiths-Harris. (See the
ddc-lemma below.)

Lemma 4.3. Let '1 be a (1, I)-form which is exact, that is, '1 = dw for
some C oo 1orm w. Then there exists a function fJ such that '1 = ddcfJ· If
'1 is real, then fJ can be taken to be real.

We shall also need the following special case of de Rham's theorem, in


top dimension on the Riemann surface.

Lemma 4.4. Let '1 be a Coo 2-form on X. Then '1 is exact, that is '1 =
dw for some 110rm w, if and only if

L '1 = 0.

Proof. De Rham's theorem says that

(2-forms)/d(l-forms) ~ H2(X, C),

and H2(X, C) is a I-dimensional vector space over C. By Stokes'


theorem, if w is a I-form, then

L dw = 0,

and on the other hand, the 2-forms I./J such that

LI./J=O

form a vector space of codimension 1 in the space of 2-forms, so by de


Rham's theorem, this proves the lemma.

Lemma 4.4 and Proposition 4.1 guarantee that f{J - c 1«(J) is exact, and
Lemma 4.3 gives the existence of fJ, thereby proving Theorem 4.2.
For the convenience of the reader, we reproduce the full statement of
the lemma on p. 149 of Griffiths-Harris. Actually, the statement given
there is not quite correct, so we repeat the result.
[I,§4J CHERN FORMS IN THE CASE OF RIEMANN SURFACES 17

Lemma 4.5. The ddc-Iemma. Let X be a compact Kahler manifold. Let


p, q ~ 1. Let 1'/ be a (p, q)-form on X, and assume that 1'/ is d-exact.
Then there exists a (p - 1, q - 1)10rm p such that

1'/ = ddcp.

If P = q and 1'/ is real, then we may take p to be real.

The proof in [G-HJ is a straightforward application of Hodge's decom-


position theorem. Observe that if a (p, q)-form 1'/ is d-exact, then frol?
the Hodge decomposition it follows that the form is also o-exact and 0-
exact.
Even if we want a result only on the Riemann surface X, we shall be
forced to deal with the product X x X, which has dimension 2, and we
shall need Lemma 4.5 in this context. We shall also need the generaliza-
tion of Proposition 4.1 and Theorem 4.2 in the higher dimension, so we
recall them and their proofs. See [G-HJ, p. 148.

Theorem 4.6. Let X be a projective non-singular variety of dimension n.


Let ,!£ be a line sheaf on X, with a metric p. Let s be a meromorphic
section of ,!£ with divisor (s) = D. Then for every closed (2n - 2)-form
IjJ we have

Proof. First, by additivity of both sides and the fact that D = L nw W


is a formal linear combination of irreducible divisors, we may assume
without loss of generality that D = W is an irreducible subvariety of X.
We may write

Let yew, a) be the set of points XEX such that Is(x) I > a. By Sard's
theorem (see, for instance, [La 4J, Chapter VIII, §1), for Lebesgue almost
all a > 0 the boundary sew, a) is smooth and yew, a) together with
sew, a) is a manifold with regular boundary. Since IjJ is assumed closed,
we get by Stokes' theorem

We now claim that the limit of the right-hand side is precisely


18 METRICS AND CHERN FORMS [I, §4]

for any (2n - 2)-form, not necessarily closed. Dealing with these more
general forms allows us to use a partition of unity on 1jJ, and thus we
may assume that the support of IjJ is contained in a small neighborhood
of some given point. If that point is not in W, then both the limit and
the integral over Ware O. So we can assume the point is in W. We
assume here the absolute convergence of the integrals, which follows
from the resolution of singularities, or can be proved as in Griffiths-
Harris, by branched covering like [G-H], pp.32 or 33. Then we may
assume that the support of IjJ lies in a coordinate neighborhood of a
regular point of W. In such a neighborhood, s corresponds to one of the
coordinate functions, say Zn' and there is a smooth positive function h
such that

So

In the coordinate neighborhood, SeW, a) is the set of points Z such that


a. Then one of the limits is 0, namely
IZn I =

lim
a~O
fS(W.a)
de log h 1\ IjJ = 0,

becauso:: de log h 1\ IjJ is a (2n - I)-form containing dZ n or dZn, and putting

we have
de log h 1\ IjJ IS( W, a) = a times a bounded form,

so the limit is 0 as asserted. Hence

lim
a~O
f
S(W,a)
de log lsi; 1\ IjJ

because ljJ(z l' ... ,zn _ l' Zn) = IjJ(Z l' ... ,Z.-l' 0) + 0(1) as a -> O. By what
amounts to the standard computation in one variable, the right-hand
side then is equal to

This concludes the proof.


[I, §4] CHERN FORMS IN THE CASE OF RIEMANN SURFACES 19

Theorem 4.7. Let X be a projective non-singular variety. Let Y be a


line sheaf on X with meromorphic section s such that (s) ,= D. Let cp be
a real closed (1, I)-form such that for all closed (2n - 2)-forms '" we
have

Then there exists a metric p on Y such that c 1 (p) = cp.


Proof. Let u be a metric on Y. By Theorem 4.6, we know that for
all "',

By de Rham duality, cp - c 1 (u) is a real (1, I)-form whose cohomology


class is O. Hence cp - c 1(u) is d-exact. By Lemma 4.5 there exists a real
function f3 such that

We let p = ef!u to conclude the proof.


CHAPTER II

Green's Functions on
Riemann Surfaces

In this chapter, we give some explicit formulas for Neron functions on


Riemann surfaces, and an explicit construction due to Coleman. These
are with respect to a canonical volume form. We shall also explain how
the functions change when we change the metrics. We work complex
analytically, in which case we can characterize the Neron functions in a
complex analytic fashion, thus finding classical objects called Green's
functions. We shall give the definition of Green's function and prove its
basic properties ab ovo. Actually, we give several proofs for some of the
basic theorems, depending on different explicit constructions. One of
them depends on the Hodge decomposition and harmonic forms, for
which a complete treatment is given in Griffiths-Harris. Another de-
pends on the smoothness of the Green's function, which may be con-
structed by theta functions. Different people at different times will use
the different techniques for different purposes. Since the construction of
the Green's function by theta functions is given in detail in [La 1],
Chapter 13, I do not reproduce this construction here. For the most
part, in the application to the intersection theory and Riemann-Roch
theorem, we use only the basic formal properties, and the construction of
a Green's function is irrelevant. In the proof of the existence of Faltings
volumes, given in Chapter VI, we need to relate the Green's function on
the curve with the Green's function on the Jacobian, associated with the
theta divisor. At this point, we shall make a reference to [La 1],
Chapter 13 for one particular property that is needed.
As a matter of terminology, by smooth we mean COO throughout.
[II, §1] GREEN'S FUNCTIONS 21

II, §1. GREEN'S FUNCTIONS

We let X be a compact Riemann surface, of a complete non-singular


curve over the complex numbers.
A form of type (1, 1) is then a differential form which locally at every
point can be expressed in terms of a coordinate z as

f(z, z)J=1 dz 1\ dz, with f smooth.

If f is real, then we say that the form is real. If f is positive, then we say
that the form is positive. By a volume form we mean a (1, I)-form which
is everywhere positive.
Let ({J be a real (1,1)-form, which is also assumed smooth or real
analytic throughout. We say that ({J is normalized if

f/ = 1.
In the sequel we always assume ({J normalized, without saying so expli-
citly any more. The important cases will occur when ({J is a volume
form. Thus by convention, a volume form is assumed normalized unless
otherwise specified.
Let D be a divisor on X. By a Green's function for D with respect to
({J we mean a function
9D: X - supp(D) --+ R

satisfying the following conditions:

GR 1. If D is represented by a rational function f on an open set U,


then there exists a smooth function ex on V such that for
P¢supp(D),
9D(P) = -loglf(PW + ex(P).
GR 2. ddc9D = (deg D)({J outside the support of D.

A function satisfying these two conditions is uniquely determined up


to an additive constant. For let 91' 92 be two such functions. Let
h = 91 - 92' Then h is harmonic on the complement of D, and being
sufficiently smooth on D by the continuity of the partial derivatives, h is
also harmonic at the points of D. Hence h is constant.
Finally, we require

This condition eliminates the remaining ambiguous constant.


22 GREEN'S FUNCTIONS ON RIEMANN SURFACES [II, §1]

Remark. Note that the Green's function is normalized so that

9 = 2A where A is a Neron function.

There is no consensus on the normalization. Faltings' 9 is - A in [Fa 1].


I use the present normalization so that certain formulas like Theorem 1.5
come out without an extra constant.

Let P, Q be points on the Riemann surface. We view P as a divisor


of degree 1. If P i= Q then the value gp(Q) is defined, and will be de-
noted by g(P, Q). Thus we view 9 as defined on the product X x X,
outside of the diagonal. Given the family {gp} of Green's functions for
all points P, we can then define gD for any divisor D by linearity, and
then gD is a Green's function for D.
The family {gph"x or the function 9 on X x X (outside the diagonal)
will also be called the Green's function associated with qJ, by abuse of
language.
For smooth functions a, f3 we note the formula

I da 1\ de f3 = df3 1\ dea. I
This is trivially proved, writing d = 0 + a and de = (0 - a)/4ni. In di-
mension 1, we have oa 1\ of3 = 0 and similarly for a, so the relation falls
out.
We shall use the following basic formula.

Lemma 1.1. Let P i= Q be two points, and let gp, gQ be functions sat-
isfying GR 1 and GR 2 with respect to smooth forms. Then

Proof Let z be an analytic coordinate at P or Q. Let a > 0, and let


U P,a and U Q,a be open sets on the Riemann surface corresponding to
discs around P and Q defined by Iz I < a. Let

Yea) = X - U P•a - UQ,a

On the complement of P and Q, we can write

On the other hand, by Stokes' theorem, if CCP, a) and CCQ, a) are the
circles of radius a around P and Q respectively, oriented counterclock-
[II, §1] GREEN'S FUNCTIONS 23

wise, then

fY(a)
dO) = - f C(P, a)
0) - fC(Q, a)
0).

We let
gp = -log r2 + v,
where v is CX) and r is the polar coordinate with respect to a given com-
plex coordinate in a neighborhood of P. Both in the current proof and
in later applications, we use a short integral table as follows.

Integral Table
If p is C'" in a neighborhood of P and rx = k log r + COO-function for
some constant k (which may be 0) then

lim
a ... O
f C(P, a)
rxdcp = O.

If P= log r2 + COO-function and rx is continuous then

lim r
a ... O JC(p, a)
rxdcp = rx(P).

By the integral table, we get the formula

f x
dO) = lim
a'" 0
f
Y(a)
dO) = gp(Q) - gQ(P),

As to the proof of the integral table, it comes from the expression of


dC in polar coordinates. For h smooth,

dCh _ 1. r oh dO on the circle of radius r.


- 2 or 2n

Then the integral table falls out of the definitions, since the integrals
which don't involve dCg p converge to 0 like a log a as a~. 0; and for the
integral with dCg p we have

ogp 2 .
- = - - + COO-functIOn.
or r

The desired integral falls out.


24 GREEN'S FUNCTIONS ON RIEMANN SURFACES [II, §1]

Theorem 1.2. Let {gp} be a family of functions on X satisfying GR 1


and GR 2. There exists a constant c such that 9 + c satisfies GR 3 if
and only if 9 is symmetric, that is gQ{P) = gP(Q) for all P =F Q. In par-
ticular, a Green's function is symmetric.

Proof From the formula of Lemma 1.1, if 9 is a Green's function,


then we get the symmetry; and conversely, from the symmetry we con-
clude that

is independent of P, whence adding an appropriate constant to 9 yields


GR 3. This concludes the proof.

The next result compares the Green's functions associated with two
different forms.

Proposition 1.3. Let ({Jl' ({J2 be real smooth (resp. real analytic) (1,1)-
forms on X such that

Let gl. g2 be the Green's functions associated with ({Jl' ({J2 respectively.
There exists a smooth (resp. real analytic) function p such that

ddCp = ({Jl - ({J2'

uniquely determined up to an additive constant. If the constant is nor-


malized so that

then for P =F Q we have

Proof The function p exists by Lemma 4.3 of Chapter I, and is as


smooth (resp. analytic) as the corresponding forms because ddc is an el-
liptic operator.* We fix one variable, say P. Then giP, Q) + P(Q) sat-

* We are using here the regularity theorem for solutions of elliptic equations. For the Coo
case, see, for instance, a complete treatment in Appendix 4 of my book SL2(R). For the
real analytic case, the theorem is due to Morrey-Nirenberg, and one needs additional esti-
mates on the coefficients of the Taylor expansion. The proof is reproduced elegantly and
briefly in Bers et al., PDE, Proceedings of the Summer Seminar, Boulder, Colorado, Inter-
science, New York, 1964, pp. 207-210.
[U, §1] GREEN'S FUNCTIONS 25

isfies GR 1 and GR 2, with respect to CfJl so differs from gl,P by a


constant. Thus it suffices to prove GR 3, namely that

fQEX
[g2CP, Q) + PCQ)]CfJlCQ) = - PCP),

where Q is the variable of integration. By the properties of Green's func-


tions and the normalization of p, the integral is equal to

- f/CfJ2 + Lg2(CfJl - CfJ2) =Ix - PddCg2 + g2 ddCP


= - Ix d(PdCg2 - g2 dC P)·

Let C(P, a) be a small circle around P. Then this last integral over X is
the limit of the integral taken outside the small disc, and we apply
Stokes' theorem to get the last expression equal to

By the integral table in the proof of Lemma 1.1, we find that the limit of
the integral is - PCP), as was to be shown.

The preceding proposition shows that if we have found one Green's


function with respect to some CfJ l' smooth or real analytic on X x X out-
side the diagonal, then we can find a Green's function with respect to
CfJ2; furthermore, every Green's function has this same property for any
other similar CfJ2' because the function

(P, Q) 1-+ PCP) + P(Q)

is smooth or real analytic on the whole of X x X.


In the next section, we shall find a function g on X x X -.6. such
that g has a logarithmic singularity of order 2 on the diagonal, such that

ddCg = <1>,

with some hermitian real analytic (1, I)-form, satisfying

with a constant form 1-1,


26 GREEN'S FUNCTIONS ON RIEMANN SURFACES [II, §l]

so that the forms <J>p are independent of P, on each vertical factor


P x X. Furthermore, we shall have

fQEX
g(P, Q).u(Q) = constant independent of P.

Thus we shall have proved:

Theorem 1.4. A Green's function exists with respect to any smooth


(resp. real analytic) normalized (l,l)-real form <p, and as a function on
X x X outside the diagonal, the Green's function is smooth (resp. real
analytic).

There is another proof of Theorem 1.4 as follows. One can construct


a Green's function by pull back from the Jacobian as in [La 1], Chapter
13, Theorem 5.2. The construction shows that the Green's function on
X x X is real analytic because the dependence on parameters of gp re-
sults from the construction by translations on the Jacobian, and transla-
tion is real analytic. See Theorem 5.1 in the case of genus 1.

For curves of genus 0, we can define a function on pl x pl in terms


of the affine coordinates (z, w) by

Iz - wl2
g(z, w) = -log -(l-+---'----zz--)(l---t-w~'

up to an appropriate additive constant. Restricting to any vertical, the


associated (1, I)-form is

J=1 dz /\ dz
<p(z) = 2~- (1 + ZZ)2'

which is independent of w. This is the Chern form of the standard met-


ric on pl.
We gave the definition of Green's functions by taking the derivative
ddCg in the ordinary manner, away from the singularities. One may also
ask for the behavior of g as a distribution. We define the functional
dd<[gp] on smooth functions f E C"'(X) by the formula

Similarly, we define [<p] by

[<p](f) = tf<P.
[II, §1] GREEN'S FUNCTIONS 27

Also we let bp be the Dirac distribution, that is, the functional such that
bp(f) = f(P). Then we have:

Theorem 1.5. Let g be the Green's function with respect to qJ. Then

or in other words, for all f E COO(X) we have

Proof As before, we have

on the complement of P in X. We apply Stokes' theorem to the region


outside a small circle of radius a around P and take the limit as a --+ 0.
The result drops out by the integral table of Lemma 1.1.

Thus the distribution property of the Green's function follows from


the seemingly weaker property GR 2.
Let us define the Laplacian I:lqJ with respect to a normalized volume
form qJ to be the operator on COO(X) such that

(I:lrpf)qJ = -dd)'.

By this definition, we can reformulate Theorem 1.5 as follows.

Corollary 1.6. Let qJ be a normalized volume form. If Jx f qJ = 0, then

This means that on the space of f orthogonal to the constants, convo-


lution with the Green's function g inverts the Laplace operator.
The space COO(X) admits a direct sum decomposition

COO(X) = C EB CJ.,

where CJ. is the space of functions whose integral with respect to qJ is 0,


because qJ is now a volume form, normalized to have integral lover X.
28 GREEN'S FUNCTIONS ON RIEMANN SURFACES [II, §2]

We define the Green's operator

G", = G: C"'(X) -+ C"'(X)

to be the inverse of the Laplace operator on C.i, and 0 on constants.


Then we may reformulate the above result as follows.

Theorem 1.7. Let lfJ be a normalized volume form. Let G be the Green's
operator associated with lfJ. Then G is obtained by convolution with the
Green's function, that is, for all fE C"'(X) we have

Gf(P) = f QeX
g(P, Q)f(Q)lfJ(Q)·

II, §2. THE CANONICAL GREEN'S FUNCTION

Let X be the Riemann surface of a complete non-singular curve of genus


q ~ 1. One can define a Hilbert space structure on the space of differen-
tials of first kind, denoted dfk, by the hermitian product

(lfJ,t/I)H~ Ix lfJ A if, = (lfJ,t/I)

Let lfJ1, ... ,lfJq be an orthonormal basis for the differentials of first kind.
Then the differential form

f.l
F1
=-2q- (lfJl A CP1 + ... + lfJq A cP q )

is independent of the choice of orthonormal basis, and will be called the


canonical 2-form or volume form on X. This form is a volume form
because if z is a local analytic coordinate, then for any dfk lfJ we have
lfJ(z) = fez) dz with f analytic, so

has a coefficient If(zW ~ 0 with respect to the oriented real coordinates


(x, y). Of course, f may have a finite number of zeros, but the canonical
2-form is everywhere > 0 since the dfk lfJl' ... ,lfJq have no common zero.
By definition of an orthonormal basis, we have
[II, §2] THE CANONICAL GREEN'S FUNCTION 29

Next we pass to the product X x X and the diagonal A, which is a


divisor on X x X. We let Pl, P2 be the projections on the first and sec-
ond factor of X x X respectively. Let

It is immediately verified that the pull back of <I> to any vertical P x X is


the constant canonical volume form Ji.. We shall have completed the proof
of Theorem 1.4 when we have shown:

Theorem 2.1. The form <I> is the Chern form of some metric on
(i)x x x(A)·

Proof We here meet a situation analogous to that of Theorem 4.2 of


Chapter I except in one higher dimension. We shall use the Hodge de-
composition, and in particular the fact that cohomology classes are re-
presented by harmonic forms [G-H], p. 116. By Theorem 4.7 of Chapter
I, to verify that <I> belongs to the diagonal, it suffices to verify that <I> is
closed (which is immediate) and that

fXxx
<1>/\1/1=11/1
.1

for all harmonic forms 1/1 on X x X. By the Kunneth formula, a basis


for the harmonic (1, 1)-forms over C is given by the "product" forms

The desired equality is then obtained by a direct computation which we


leave to the reader.

Theorem 2.2. Let P be a metric on (i)xxrtA) such that cl(p) = <1>. Let s
be a holomorphic section of (i)x x x(A) such that (s) = A, and let

After adding a constant to g (or equivalently changing s by a constant),


then g is the Green's function associated with the form 11 on X.

Proof Two sections having the same divisor differ by a multiplicative


constant, so anyhow g is determined up to an additive constant and g is
symmetric (the transposition of the factors of X x X leaves A invariant
30 GREEN'S FUNCTIONS ON RIEMANN SURFACES [II, §3]

and so multiplies s by a constant whose square is 1). For each point


P E X it is immediate from the definition of <I> that we have

<I>p = II on the factor P x X.

Thus for each P,


ddCg p = Jl.

Furthermore, 9 has a logarithmic singularity of order 2 on the diagonal


of X x X, so for each P, gp has a logarithmic singularity of order 2 at P
on X.
Finally, the integral

is constant, independent of P, because 9 is symmetric and we can apply


Lemma 1.1. Thus after adding a constant to 9 if necessary, we may as-
sume that for each P we have

By the uniqueness of the Green's function, we conclude that gp is the


Green's function on X at P, associated with the form Jl. Since the func-
tion 9 on X x X - A was defined in such a way as to make its smooth-
ness obvious, we also have completed the proof of Theorem 1.4.

II, §3. SOME FORMULAS ABOUT


THE GREEN'S FUNCTION

Let 9 be the Green's function associated with the canonical form Jl.

Let z denote a holomorphic coordinate with respect to the first factor,


and w a holomorphic coordinate with respect to the second factor. By
Theorem 2.2 in the neighborhood of a point (zo, zo) one has the expan-
SIOn

g(z, w) = -loglz - Wl2+ real analytic function in (z, w)


= -log«z - w)(i - w» + real analytic function.

Therefore

(1) awg(z, w) = (=_1


z- w
__= + h(z, W») dw,
where II is real analytic.
[II, §3] SOME FORMULAS ABOUT THE GREEN'S FUNCTION 31

If we want to use a notation independent of coordinates, we shall


write the 8, 8 operators in the form

and

with respect to the first variable P and second variable Q respectively, or


if we want to eliminate any reference to the variable we write

and

Theorem 3.1 (Reproduction formula). Let cp be a dfk on X. Then

In the integral on the left, P is the variable of integration, and Q is con-


stant.

Proof We have
d p (8 Q g /\ cp(P») = (8 p 8Q g) /\ cp(P).

Let U Q,a as before be a small disc of radius a around Q, with respect to


some analytic coordinate as in the previous section. By definition, the
integral on the left is the limit

I = lim I(a) where


a~O

Therefore by Stokes' theorem, we find:

Let z be an analytic coordinate at Q, with z(Q) = O. The boundary of


UQ,a relative to its complement in X is the circle, negatively oriented.
Write
cp(z) = f(z) di with f holomorphic.
Then by (1),

l(a) = r (~ + O(1»)f(Z) di /\ dw.


Jlzl~a z
Taking the limit as a approaches 0 and dividing by - 2ni yields f(O) dw,
which is the expression cp(Q) in terms of the coordinate w.
This proves the theorem.
32 GREEN'S FUNCTIONS ON RIEMANN SURFACES [II, §3]

Theorem 3.2. Let cp be a dfk on X, Then

8Q ( 8 p g(P, Q) " cp(P) = O.


JPEX
Proof The proof is analogous and is left to the reader.

Theorem 3.3. Let g be the Green's function on X x X associated with


the canonical volume form. Then on X x X outside d, we have

Let d be the diagonal map X -+ X x X. Then

Proof By the uniqueness of Green's functions, and of metrics having


a given Chern form, and by Theorem 2.1 the theorem follows immedia-
tely. Since Theorem 2.1 assumed general results via de Rham's theorems,
Hodge theorem, and the ddc-Iemma, we shall give another proof here, us-
ing only the definition of the Green's function on the one-dimensional X,
and the fact that when viewed on X x X the Green's function is smooth.
For this proof it is convenient to write the formula to be proved in the
form

Then

so

Next we note that

To see this, we apply a


p, and get
[II, §3] SOME FORMULAS ABOUT THE GREEN'S FUNCTION 33

Note that the singularity is removable, because in terms of local analytic


coordinates z, w at P, Q respectively, we have

o 0 2
OZ OW log Iz - wi = o.
To prove the theorem, it will suffice to prove:

Lemma 3.4.
q

op8 Q g = -1t L cp,(P) 1\ cp;(Q).


i= 1

Indeed, this will yield the middle terms in brackets as stated. Since
op8Q g is holomorphic in P and anti-holomorphic in Q by symmetry, the
Kunneth theorem for H2(X x X) implies that we can write
q

op8Q g = L cp;(P) 1\ I/I;(Q),


,= 1

where the forms l/Ii(Q) are of the first kind in the second variable Q. We
now evaluate the following integral in two ways. By the reproduction
formula,

On the other hand, using the fact that CPI' ... ,CPq form an orthonormal
basis, we also get the value
1--
- -I/I/Q).
1t

Hence I/I/Q) = -1tcp/Q). This proves the lemma and the first part of the
theorem. Letting P = Q proves the second part.

Theorem 3.5. Let p be the metric on (1) x x x(A) associated with the
Green's function of the canonical volume form. Then

Proof This IS immediate from the definition of the Chern form and
Theorem 3.3.

The next two sections are not logically necessary for what follows.
They give more explicit descriptions of the Green's function in some
cases.
34 GREEN'S FUNCTIONS ON RIEMANN SURFACES [II, §4]

II, §4. COLEMAN'S PROOF FOR THE


EXISTENCE OF GREEN'S FUNCTION

The explicit proof of existence in this section is due to Coleman. It is


based on differentials of second kind (dsk) and harmonic forms.
We begin by recalling a basic fact about the duality of differentials of
first kind (written dfk), namely the pairing

establishes a duality between the real vector spaces of differentials of first


kind and H I (X, R). Cf. [La 3], Chapter IV.
Let H be the complex vector space of all differential forms of type
rp + l[J, where rp, tfJ are dfk. Then H has dimension 2q. We call H the
space of harmonic forms. There is a unique C-bilinear map

H x HI (X, C) ~ C

such that for OJ E Hand Y E HI (X, Z) we have

(OJ, y) 1-+ lOJ.

Theorem 4.1. This pairing is a duality of complex vector spaces.

Proof Immediate from the duality over R which we just recalled.

We denote the space of differentials of first kind by H I • O, and we let


HO,l be the space of forms l[J, where tfJ is of first kind. Each of these
spaces has complex dimension q. We call HO,I the space of anti-holo-
morphic forms.
Note that the anti-holomorphic forms HO,t form the complex dual
space of Ht,o under the pairing given by the integral

(rp, l[J) 1-+ f


x
rp A l[J = } ,
v-1
<rp, tfJ>·

In the next theorem, we consider the space of Coleman forms, i.e. the
differential forms which can be expressed as linear combinations of forms
of type
hrp + algebraic differential form,
[II, §4J COLEMAN'S PROOF OF EXISTENCE 35

where cP is of first kind, and dh = r[J + '1 with '" of first kind and '1 of
second kind. For such forms, it makes sense to speak of their having a
pole at a point since the principal meromorphic part is well defined
modulo real analytic forms. It also makes sense to speak of the order of
the pole, and of the residue at a point.
We want to construct such forms having appropriate logarithmic
singularities, and pure imaginary periods. It may happen however that
only "real" periods are defined. So for any harmonic form on an open
set, we define its real part

Re(co) = Hco + w).


The following lemma is then useful.

Lemma 4.2. Let co be a harmonic differential on an open set, and let ')I
be a path in the open set. Then

Proof The formula is local, and the open set can be taken to be a
small disc in the complex plane. We may also prove the formula sepa-
rately when co is holomorphic and anti-holomorphic. But then co has a
primitive and the formula is obvious.

The next theorem gives the basic idea for the construction of a
Green's function.

Theorem 4.3. Let D be a divisor on X. Let CPh ... ,CPq be an orthonor-


mal basis for the dfk. There exists a unique form COD on X satisfying
the following conditions.
q
(i) COD = L h,cpi + an algebraic differential form where hi are functions
;= 1
such that dh i = c(ijJe - '1,) with '1i of second kind for all i, and some
constant c.
(ii) Re(co D) is closed, i.e. dRe(co D ) = 0 and its periods are O.
(iii) COD has poles of orders at most 1, and for all P,

Observe that prescribing the residues of the form in (iii) uniquely


determines the constant c in (i) if any such form exists.
Let us prove the uniqueness. Let co be the difference of two forms as
in the theorem. Then co is meromorphic by (i), holomorphic by (iii), and
36 GREEN'S FUNCTIONS ON RIEMANN SURFACES [II, §4]

o by (ii) and the real duality of differentials of first kind with HI (X, R).
This takes care of uniqueness.
Before proving existence, we indicate how the theorem can be used to
construct a Green's function for a divisor D. Let

This is a well-defined function since the periods of Re(wp) are O. Since


Wp has a pole of order 1 at P, with residue 1, the first condition GR 1 is
satisfied.
Next we have a routine computation:

oag = (0 + a)ag = -adg


= -a(WD + roD)
= -aWD

= -dWD

q
= - I. dh i 1\ CfJi
i= 1

cq
=--Jl
F1
by (i). Therefore GR 2 is satisfied if and only if the constant c has the
value

2n deg(D)
c=
q

The existence proof, to which we now come, shows that this can be
realized.

Let '7 be a dsk. On any simply connected domain we can integrate '7
from a fixed point 0 to a variable point P, and get a well-defined func-
tion except at the poles of '7, because the residues of '7 are O. In particu-
lar, we can do this locally, and such a function satisfies df = '7.
Let '71' '72 be two dsk. We define their scalar product

en], '72] = I. res p(fp'72),


p
[II, §4] COLEMAN'S PROOF OF EXISTENCE 37

where dIp ="1 in a neighborhood of P. Since in any case globally an


integral 11 of"1 is determined mod periods, and "2has no residues, we
"1
could write 11 instead of Ip to denote any of the different integrals of
up to a constant in a neighborhood of P.
Let us represent the Riemann surface as a quotient of a polygon t!I in
the usual manner, in such a way that
polygon. Then Cauchy's formula yields
"1' "2 have no poles on the

Let " be a dsk. Then" gives rise to a functional

by the pairing

for yeH 1(X, Z).

By Theorem 4.1, there exists a unique form H(,,) e H such that the func-
tional [H(,,)] is the same as the functional ["J. We call H(,,) the har-
monic projection of". In another language, this means that H(,,) has the
same periods as fl.

Theorem 4.4. For dsk "1' "2 we have

Proof Represent as usual the Riemann surface as a quotient of a


polygon t!I with sides ai' bi' -ai' -bi' On the interior of the polygon,
let hI' 11 be functions such that

and

They extend to the boundary in the usual manner. We note that har-
monic forms are closed, i.e. if w is harmonic, then

dw =0.

Indeed, d = iJ + a. If locally <p = h(z) dz with h holomorphic, then


a<p =0 because ah = 0 (Cauchy-Riemann). Similarly, iJi[J = O. Also
38 GREEN'S FUNCTIONS ON RIEMANN SURFACES [II, §4]

o<p = 8ip = 0 because each of these forms has dz /\ dz or di /\ di. Hence


dw = O. Thus we find

d(h 1 H(Yfz») = dhl /\ H(Yf2) + hi /\ dH(Yfz)


= dhl /\ H(Yfz),

so by Stokes' theorem,

(1)

Since H(Yfl) and Yfl have the same periods, it follows that hi - 11 is a
function on X. Hence for all i,

and similarly for the integral taken over (b i) + (- bi)' so by the preceding
equality (1) we get

(2)

We repeat the same procedure to find

and similarly taking the integral on (b;) + ( - bJ Taking the sum yields

(3)

We have already seen by Cauchy's theorem that this gives the desired
value [Yf 1, Yfzl This concludes the proof.

Corollary 4.5. The scalar product is skew-symmetric.

Proof Immediate from the formula in terms of the integral.

We recall that the space of de Rham cohomology, namely the space

HJR(X) = dsk/exact (by definition)


[II, §4J COLEMAN'S PROOF OF EXISTENCE 39

of differentials of second kind modulo exact differentials is of dimension


2q. The exact differentials are those of the form dj, where j is a rational
function on X, and obviously are in the kernel of the pairing ['11,I12J
smce the sum of the residues of a differential form is equal to O. The
map
H: dskjexact ~ H

into the harmonic forms is an isomorphism, because a differential form


of second kind '1 without periods has an integral which is a rational
function j on the curve, and then I] = dj, so the exact forms constitute
precisely the kernel.

Corollary 4.6. The pairing between dskjexact and itself given by


['11' I] 2J is non-degenerate.

Proof We use the formula of the theorem, together with the fact that
we can pick 1]2 such that

in light of the isomorphism H: dskjexact ~ harmonic forms. This proves


the corollary.

We have the inclusions

and dime HjH 1 • O = q.

We also have
dsk => H 1, ° EEl exact => exact.

In the duality, H 1 • O (namely dfk) is self-orthogonal. Consequently, the


pairing gives a natural identification

dskj(H1.0 + exact) ~ dual space of H1.o.


We can rephrase these remarks in terms of a dual basis as follows.

Lemma 4.7. Let ({Jl' ••• ,({Jq be a basis oj dfk. Then there exist
dsk '11' ... ,I]q such that:

(i) H(I]J E HO. l(X), so H(I]J is anti-holomorphic.


(ii) [I]j, ({JJ = fV
If the ({Jj are orthonormal, these '1i satisfy
40 GREEN'S FUNCTIONS ON RIEMANN SURFACES [II, §4]

Proof The first two statements have been proved. As to the third, we
carry out the computation, which involves the hermitian product and the
bilinear product:

cpJ = .~ Ix CPj /\ H(I'],)


- yCl-
()ij = [1'];, by Theorem 4.4

1 -
= - <CPj' H(I']J)
n

by definition of the hermitian product. This proves the lemma.

We now let CPI' '" ,CPq be an orthonormal basis for the dfk. We can
then define a function

which is well defined on X except at the singularities of 1'];, where h; has


poles. Then

Let D be a divisor on X. We want to construct W D. We start with


the form

We want to add a merom orphic form to WI so as to get the W D of


Theorem 4.3. We do this step by step.
By Lemma 4.7(ii) and the definition of the scalar product with
[1'];, CPJ = 1, we get

deg D
L res I
q
p h;cp; = -deg D.
q PEX ;=1

Lemma 4.8. For each point P of X suppose given a principal part for
an algebraic differential form locally at P, such that for almost all P the
principal part is 0, and the sum of the residues is O. Then there exists a
global algebraic differential form on X having the given principal part at
every point.

Proof First there exists a differential of third kind (dtk) having the
given residues, so after subtracting it we may assume without loss of gen-
erality that the residues are O. It also suffices to prove the lemma when
[II, §4] COLEMAN'S PROOF OF EXISTENCE 41

D is a single point. But then the existence of a dsk having the given
principal part at a given point is an immediate consequence of the
Riemann-Roch theorem.

By the lemma, there exists an algebraic differential form Wi such that


for all points P,
deg(D)
res p w
I

= --- res p L h,cp, + ordp(D)


q

q i= 1

and such that the form

has only simple poles with residues ordp(D) at P. Then W2 is real ana-
lytic on the complement of D, and if D is represented by a rational func-
tion f on an open set U, then

is real analytic on U. Thus W 2 has the right singularities.


It will now suffice to correct W 2 by a differential of first kind so that
the new form has pure imaginary periods. We have

deg(D)
dW 2 = - -
q
L H(l1i) 1\ CPi·
Changing signs and using the definition of the canonical volume form, as
well as Lemma 4.7(iii), we get

dW 2 = 2nJ=1 deg(D)Jl.
Thus dW 2 is a pure imaginary multiple of the real volume form Jl. We
obtain

because ji = Jl. Hence Re(w 2 ) is closed, so a period

depends only on the homology class of a closed path y. We want to add


a dfk to w 2 to make the periods pure imaginary.
42 GREEN'S FUNCTIONS ON RIEMANN SURFACES [II, §5]

We apply Lemma 4.2 with}' equal to a closed path, to find

so the real parts of the integrals of Wz over cycles are well defined on
the homology classes. Now we use the fact that there exists a differential
of first kind ({J having given real period matrix. We let

Then the periods of W3 are all pure imaginary. We have thus con-
structed a differential form satisfying the conditions of Theorem 4.3.
In principle this gives us the Green's function

and all we have to check is that the right constants appear. The matter
is either trivial or false at this point, and it is trivial as follows:

aag = (a + a)ag = - adg


= -a(W3 + c0 3 )
= -aW2

= - dW 2 = - 2nJ--l deg (D)fJ..


This concludes the proof.

II, §5. THE GREEN'S FUNCTION ON ELLIPTIC CURVES

Neron defined his function at infinity by means of theta functions. On


elliptic curves one has a q-expansion giving rise to certain explicit deter-
minations, cf. [La 2], Chapters 18 and 19. For our purposes here we can
make the presentation self-contained by defining ad hoc the relevant
functions as follows.
Let z = x + iy be a variable in C and let

't' = u + iv, v> 0,

be a variable in the upper half plane il. We define


and
[II, §5] THE GREEN'S FUNCTION ON ELLIPTIC CURVES 43

The second Bernoulli polynomial is

We define the Neron function

The reader who knows about Klein forms and Siegel functions will rea-
lize that this is minus the log of the absolute value of the Klein form.
Cf. [K-L], Chapter II, §1.
Since 1m r > 0 it follows at once that the product is absolutely con-
vergent and defines a real analytic function in both variables outside the
lattice points m + nr with m, n integers.
The form of the expression inside the absolute value sign immediately
shows that as a function of z, with fixed r,

A(Z + 1, r) = A(Z, r).


We also have
A(Z + r, r) = A(Z, r).

This is checked by keeping track of the first term in the products under
the change Z ~ Z + r, and using the formula

As a result, the function

Z~A(Z, r) for fixed r

is a function on the complex torus Cj[l, r] defined outside the lattice


points of the lattice [1, r], and having a logarithmic singularity of order
1 at the lattice points. In other words, it is a Neron function whose div-
isor is the origin.
We shall find the Chern form associated with A. We recall the rela-
tions:

ddy = J=1
2-1
n
oaf = ~ !Af dx
2n
A dy,

where A is the Laplacian, A = (ojoX)2 + (Ojoy)2. We let


V(r) = Vol[l, r] = 1m r
44 GREEN'S FUNCTIONS ON RIEMANN SURFACES [II, §5]

be the area of a fundamental domain of the lattice [I, r]. A function g is


a Green's function for a (1, I)-form w if

ddCg = OJ and

and g has a logarithmic singularity of order 2 at suitable points.


We let f1 be the (1, I)-form on C/[l, r] which is the scalar multiple of
dx 1\ dy such that the integral over the torus is 1. Thus

1
f1 = V dx 1\ dy

where V is the volume (area in this case) of a fundamental domain with


respect to the euclidean (flat) metric.

Theorem 5.1. Let g(z, r) = 2A.(z, r). Then g is the Green's function with
respect to f1 on C/[l, r], with singularity at O.

Remark. Before giving the proof, we notice that the Green's function
on the product, as a function of two variables, is then immediately
obtained by translation, namely

g(z, w) = g(z - w).

Note that by glvmg an explicit form as above, we have automatically


shown the analyticity outside the diagonal.

Proof The expression for g shows that it has the right logarithmic
singularity. We have to show that ddCg = f1 outside the singularity, and
that

where the integral is taken over a fundamental domain.


Let us start with the second condition about the integral. The func-
tion g is locally L 1 because of the mild logarithmic singularity, so the in-
tegral is absolutely convergent on any compact set of R2. For a
fundamental domain, we select the set of all points

r + sr, with 0< r, s < 1.

We change variables, letting

x=r+su and y = sv.


[II, §5] THE GREEN'S FUNCTION ON ELLIPTIC CURVES 45

Then
L )'(x, y) dx dy = ff ),(r + su, sv)v dr ds.

The log of the product expressing ), is a sum of logs involving three


expressions. It suffices to show that the integral of each one of these
expressions is O.
The first integral is

because the inner integral with respect to s is O.


Next we have the integral

for n ~ O.

But the log of the absolute value IS the real part of the log. Further-
more, since s > 0, we have

for n ~ O.

Thus the log of the absolute value is the real part of the usual series ex-
pressing the log, and it suffices to show that the integral of this series is
equal to O. For Iwl < 1 we have
00 wm
-log(1 - w) = I
m=l m

and the convergence is uniform on compact sets. For fixed s > 0 we


then have

This proves that 12 = O.


Similarly, using s < 1, we find that

This concludes the proof that

L),(X,y) dx dy = O.
46 GREEN'S FUNCTIONS ON RIEMANN SURFACES [II, §5]

Next we compute the Chern form. Recall that

;-:'
ddf = v2n- I caf = 11M dx /\ dy
2n 2

where Ll is the Laplacian.

Away from the lattice points, only the first term in the product for 9
under the log is not equal to the product of holomorphic times anti-
holomorphic terms, and so dd c of every term except the first is 0 away
from the lattice points. On the other hand, the first term is real analytic
(no singularity). The next lemma therefore suffices to prove that 9 IS a
Green's function with respect to fl, with singularity at the origin.

Lemma 5.2.

Proof Note that

is a smooth function, and therefore we can apply dd c directly to this


function to compute its value as a distribution. Using dd c in terms of
the Laplacian yields the desired expression fl.
In some applications it is useful to know also the distribution deter-
mined by 9 and we shall now determine that also.
If 'Y. is a locally L I-function, we define dd"['Y.] as a distribution by
transposition, that is

dd"['Y.J/J = ('Y.dcd/J.
JR2
Note. Usually when integrating by parts with one partial, there is a
minus sign when shifting the partial from one function to the other, but
here the Laplacian involves squares of partials, so the minus sign disap-
pears.

Similarly, we let [flJ be the distribution whose effect on a function f is


given by

[flJf = ( ffl·
JR2
[II, §5] THE GREEN'S FUNCTION ON ELLIPTIC CURVES 47

Theorem 5.3. Let b a be the Dirac distribution, ba(f) = f(a). Then

ddC[g] = [Il] - L ba •
aE[1.t]

Proof We have already computed the first term. The other terms are
given by the next lemmas.

Lemma 5.4. dd c [loglzI2] = bo (the Dirac distribution at the origin).

Proof This comes from Stokes' theorem applied to the region outside
a small circle centered at the origin, and the expression for d C in polar
coordinates given at the beginning of Chapter I, §3. It is the analogue
on euclidean space of Lemma 1.1.

Lemma 5.5. For each integer n we have

dd C [1ogl1 - q~qzI2] = L bm- nr


mEZ

Proof The only singularities are at the points m + m. Outside these


points, the function inside the brackets is a product of holomorphic
times anti-holomorphic, and so its dd c is O. Hence to prove the lemma,
we only have to test the left-hand side on a function with support in a
small neighborhood of m - m for an integer m. In light of the preceding
lemma, it suffices to prove that the left-hand side of the expression in the
lemma differs from the appropriate translation of dd c [loglzI2] by the
absolute value squared of a holomorphic function. But indeed, in a
neighborhood of m - m, the quotient

1- q~qz
z - (m - m)

is holomorphic since the numerator has a zero of order L. Hence

in a neighborhood of m - m. This proves the lemma, and also con-


cludes the proof of Theorems 5.1 and 5.3.

For an ade1ic version of the above, cf. Bergelson's thesis, An Adelic


Euler-MacLaurin Formula for Imaginary Quadratic Fields., Yale, 1980.
Using the Klein forms as in [K-L], one can presumably construct in
a similar fashion the Green's function for the cusps of a modular curve.
See also the last section of my book Introduction to Modular Forms.
CHAPTER III

Intersections on an
Arithmetic Surface

The purpose of this chapter is to recall briefly the intersection theory


that we shall need. This first arose in Shafarevich [Sh], and also Lich-
tenbaum [Li], which is a useful reference for other types of theorems
than those in Shafarevich, for instance, the projective imbedding of cer-
tain schemes, which we shall use if and when necessary.
The best foundations for intersection theory that I know are given in
Fulton [Fu], who deals with singular varieties. To have full flexibility, it
is useful to have some foundation results on singular varieties. Among
other things, even if we deal with a regular model over the ring of inte-
gers of a number field, making a base change introduces singularities.
One has then a choice: resolve them, with the technical consequences
that this implies (for us, only in Chapter V, §5), or deal with intersec-
tions on non-regular schemes. For this book, I have chosen to deal with
regular models for all the main theorems, this being still the simplest and
shortest approach as far as we go, and so I have reproduced some basic
proofs from the intersection theory. However, in the long run, one must
be ready to deal with the full generality of [Fu], in which case one can
just quote chapter and verse in his book. The key step in the singular
case is dealt with by Fulton in his Theorem 2.4, proving the commutati-
vity. Assuming regularity from the start avoids the difficulty of this step.

III, §1. THE CHOW GROUPS

We begin with basic definitions. Let X be a scheme, flat and proper


over spec(R), where R is a Dedekind ring or a field. All schemes are as-
sumed separated. We let Y = spec(R).
[III, §1] THE CHOW GROUPS 49

Let F be the quotient field of R. If R is a Dedekind ring, X is proper


over Y and integral, and X F is non-singular of dimension 1, then we say
that X is an arithmetic surface over R. For some purposes, we shall as-
sume that X is regular, but we do not do so at first.
An integral scheme of finite type over a regular ring will be called a
variety. If V is a variety, we let k(V) denote its function field, i.e. the re-
sidue class field of its generic point. If V is a variety over R, we also
denote k(V) by F(V).
If cp: V' ---+ V is a morphism of varieties, we let:

[V' : V] = 0 if k(V') is not finite over k(V);


[V': V] = degree of k(V') over k(V) otherwise.

An example of such a morphism arises from the normalization of a vari-


ety, in its function field or in a finite extension.

We assume throughout that the normalization of a variety is finite over


the variety.

This condition is satisfied in all applications, and the finiteness of inte-


gral closures in a finite extension is a property of so-called excellent
rings, see Matsumura [Mat 1]. It is a property stable under taking the
standard constructions of algebraic geometry: finitely generated exten-
sion, localization, and completion.
By a subvariety of X (over spec(R), always) we mean a variety to-
gether with a closed imbedding

Vc. X

over spec(R). By a prime, or irreducible Weil divisor we mean a subvar-


iety of codimension 1. A Weil divisor is an element of the free abelian
group generated by the prime divisors. A Weil divisor can thus be writ-
ten as a linear combination

L nyV or

where V ranges over the prime divisors, ny E Z, and ny = 0 for all but a
finite number of V. The notation [V] is to emphasize the free element
corresponding to the subvariety. If ny ~ 0 for all V, then we say that the
divisor is elfective.
Suppose that X is itself a variety. Let D be an effective Cartier div-
isor on X. Let V be a prime divisor on X. If D is locally defined by a
function fE lIJ y. x in a neighborhood of a generic point of V, then we de-
fine the order of the divisor

ordy(f) = ordy(D) = length(lIJylfllJy).


50 INTERSECTIONS ON AN ARITHMETIC SURFACE [III, §1]

The function fl-+ ordy(f) is a homomorphism on non-zero elements of


l!I y, as is immediately verified since for another such function g, the map
multiplication by g is injective. Hence

length(gl!ly/gfl!ly) = length(l!Iyffl!ly).

As a result, we can extend the order function to the multiplicative group


of all non-zero elements of k(X). Then we can define the order function
on Cartier divisors, effective or not, namely: If D = (f) in a neighbor-
hood of the generic point of a prime divisor V, then we let

ordy(D) = ordy(f).
We define the Weil divisor associated with a Cartier divisor D to be

[D] = L ordy(D)[V].
y

Let fE k(X), f # O. Then f defines a Cartier divisor (f) (or (fh if we


need to specify X in the notation), and the associated Weil divisor

[(f)] = [(fh] = L ordy(f)[V].


y

If X is regular, so Cartier and Weil divisors need not be distinguished,


then we may omit the brackets and simply write (f) or (fh for the div-
isor of f. The Cartier or Weil divisor as above are called the principal
divisors determined by f.
By a curve we mean a variety of dimension 1. By a curve on X we
mean a subvariety of dimension 1. The word dimension will have the ob-
vious meaning in the applications. For our purposes, we follow Fulton
[Fu], Chapter 20, p. 304. Let W be the image of the subvariety V in
spec(R). If W is a closed point x, we define

dim V = transcendence degree of keY) over k(x).

If W is not a point, then we define

dim V = transcendence degree(k(V)/F) + 1.

Note that Weil divisors on curves are elements of the free abelian
group generated by the closed points.
[III. §1] THE CHOW GROUPS 51

Lemma 1.1. Let X be a variety, and let X' be its normalization in a


finite extension of k(X). For f E k(X), f -:f. 0, we have

[X' : X] ordy(f) = L ordy.(f)[k(V'): k(V)].


y'

where the sum is over the subvarieties of x' lying above V.

Proof Suppose first that X' is the normalization of X in its function


field. Let A = @y,X and let A' be its integral closure in k(X). Then A'iA
is an A-module of finite length, and we have a commutative diagram of
inclusions:
A'

A fA '

fA

Since multiplication by f is an isomorphism, we get [A': A] = UA' :fA],


where the bracket denotes A-length. It follows that

[A :fA] = [A' :fA'].


By the Chinese Remainder Theorem, we have a direct sum decomposi-
tion
AIIfA' = E8 A~·lfA~.
y'

so the desired index falls out.

The general case of a finite extension then follows from the standard
formula relating the ramification indices and residue class degree for the
extension of a prime in a Dedekind ring in a finite extension.
The lemma will be applied frequently when X is a curve and when
the prime divisors are closed points.

Example. Let X be an arithmetic surface over spec(R), where R is a


Dedekind ring. Then Wei] divisors are of two types.
52 INTERSECTIONS ON AN ARITHMETIC SURFACE [III, §1]

Horizontal divisors. These are linear combinations of prime horizontal


divisors, and a prime horizontal Weil divisor is merely the Zariski clo-
sure in X of a rational point of X in some finite extension of K. If E is
such a prime divisor, then E = Spec(A), where A is finite over R. The
justification for this assertion is a theorem of Chevalley (cf. [EGA] III,
4.4.2) which says that a proper morphism I: Z' -+ Z such that I -l(Z) is
finite for all Z E Z is actually affine. If X is assumed regular, it follows
that E is regularly imbedded in X, that is, E is a locally complete inter-
section by a regular sequence of length 1. This puts a very useful condi-
tion on the type of rings A which can occur for such E. They are
"better" than a random order in a number field.

Vertical or fibral divisors. For each point y E Y = Spec(R) we have the


fiber X y (sometimes denoted by X (y» over the residue class field key),
which we denote also by k(v) for the valuation v corresponding to the
point y in Spec(R). Such a point y is merely a prime ideal p, so v = vp'
The fiber is smooth over key) for almost all y. For a finite number of y,
it may be much more complicated, but in any case, it has irreducible
components. A vertical Weil divisor on X is a linear combination of
such components. Unless otherwise specified, we always assume that y is
a closed point.

We return to the general case. By a p-cycle on X we mean an ele-


ment of the free abelian groups generated by the subvarieties of dimen-
sion p. The support IIX I of a p-cycle is the union of the varieties
occurring in IX with non-zero coefficient. We let

Zp(X) = group of p-cycles on X.

We define the subgroup of cycles rationally equivalent to 0 to be

Ratp(X) = subgroup of cycles IX satisfying the following condition.


There exists a finite number of subvarieties Wi of X, of di-
mension p + 1, and functions J; E k(Wi) such that

IX = L [(J;)].
Note that (J;) is the Cartier divisor of Ii on Wi, and [(J;)] is its asso-
ciated Weil divisor, which is a linear combination of subvarieties of di-
mension p, so it is a p-cycle.
We define the Chow group
[III, §1] THE CHOW GROUPS 53

Let qJ: X' --+ X be a morphism (over spec(R), as usual). Then qJ in-
duces a homomorphism, the direct image,

by letting
qJ*(V') = [V' : V]qJ(V').
For a composite qJ 0 !/J of morphisms, we have

Suppose X = {x} is a point. Then

CHo(x) = Z[x].

If X = spec(k) where k is a field, then we identify CHo(spec(k)) with Z.


In general, a proper morphism qJ: X' --+ X induces the homomorphism

such that
qJ*(L nAx']) = L nx·[k(x') : k(x)][x]
if the sum is taken over all x' with <p(x') = x. We shall see such sums
appear frequently in the sequel. If x = spec(k) we often omit [x] in the
formula.

Theorem 1.2. Let <p: X' --+ X be a proper surjective morphism of vari-
eties. Let f E k(X') be a rational function on X', f ¥- 0, and let [(f)] be
its Weil divisor on X'. Let N = N X'IX be the norm. Then

qJ*[(f)x,] = [(N(f)h] if dim X' = dim X,


=0 if dim X' > dim X.

Proof We shall consider first the special cases which arise in the ap-
plications. Suppose first dim X' = dim X, and suppose X' is the normal-
ization of X in the function field k(X). Then the formula is true, this
being a special case of Lemma 1.1. Suppose that X is normal and X' is
the normalization in a finite extension. Then the formula is a standard
fact of elementary theory of Dedekind rings, as in algebraic number
theory. For a proof in the classical style of algebraic number theory,
see [La 5], Chapter I, Proposition 22. For a proof in the style of lengths
of modules, see [Fu], Chapter 1, Proposition 1.4.
54 INTERSECTIONS ON AN ARITHMETIC SURFACE [III, §1]

Next, suppose that X and X' are both normal, and qJ: X' --+ X IS bi-
rational. Then
[(f)x-J = D' + D"
where a component V' of D' is such that qJ(V') is a prime divisor on X,
while a component V" of D" is such that qJ(V") has lower dimension.
The components V" disappear under qJ*. Suppose V' is of the first type.
Then (!.Iv'x = (!.Iv.x since a discrete valuation ring is maximal in the quo-
tient field. Conversely, given a discrete valuation ring (!.Iv,x of some
prime divisor on X, since qJ is assumed proper this ring induces a point
on X', and since qJ is a morphism, it follows that the point is in fact the
generic point of a prime divisor V', so (!.Iv, x = (!.Iv'x. From this the for-
mula of the theorem is obvious in the birational case, with both X, X'
normal.

The third important special case is given by the following standard re-
sult on curves, which is a reformulation of Theorem 1.2 for a complete
curve over a field.

Theorem 1.3. Let C be a complete curve over a field k. Let f E k( C), f


not algebraic over k. Then

Proof We sketch the proof which follows a technique similar to that


used in Lemma 1.1. The subfield kef) may be viewed as the function
field of the projective line pl. We then view f: C --+ pI as a finite morph-
Ism. We have two points 0 and r:f) on pI, and one verifies that the sums

L ordif)[k(x): k] and - L ordx(f)[k(x) : k]


x10 x1°o

are both equal to the degree [X: pI]. By symmetry, it suffices to prove
this equality just for the first sum, and then it amounts to Lemma 1.1
again.

The most general case of Theorem 1.2 follows formally from the
above special cases. Since these special cases will be the only ones occur-
ring in applications, we leave the rest of the proof to the reader. Cf.
[Fu], Proposition 1.4.
From Theorem 1.2, it follows immediately that if qJ: X' --+ X is a
proper surjective morphism over spec(R), then

qJ*: ZiX ') --+ ZiX )


maps RatiX') into Ratp(X), and consequently that qJ* is defined on the
Chow group CHp(X).
[III, §2] INTERSECTIONS 55

III, §2. INTERSECTIONS

Throughout this section, we let X be an arithmetic surface over spec(R),


where R is a Dedekind ring, and we let Y = spec(R). H~ now also as-
sume that X is regular.

Let D, E be effective divisors on X, and without common component.


Let x be a closed point of X. Suppose that D, E are represented by
functions f, g in a neighborhood of x. We define the intersection number

iiD,E) or iiD. E) = length({!\/(f, g)),


where the length is that of @x-module.

Proposition 2.1. The symbol iiD, E) is bilinear.

We recall the proof. Suppose D' is represented by f'. We have the


following sequence of ideals:

@x ::J (f, g) ::J (ff', g) ::J (g).

Let @x = @xI(g). Denote length(M/N) by [M: N]. Then

[(f, g) : (ff', g)] = [@J: @xff'] = [@x: @J'] = [@x : (f', g)],

because 1 is not a divisor of zero in @x. The desired additivity in the


first variable is then immediate.

For two divisors D, E without common component, we define their in-


tersection
D.E = L iAD, E)[x].
x

This sum can be viewed as a O-cycle on X or on the support of


IDI nlEI·
Lemma 2.2. Let C be a curve on X, let x be a closed point lying on C,
and let D be a divisor not having C as component. Suppose D = (f) in
a neighborhood of x. Then

where fl C is the restriction of f to C. In other words,

D.C=DIC.
56 INTERSECTIONS ON AN ARITHMETIC SURFACE [III, §2]

Proof Suppose C = (g) in a neighborhood of x. Then the desired


equality comes immediately from the definitions, using

and the order function.

Lemma 2.3. With the above notation, if t/J: C' -+ C is the normalization
of C in a finite extension, then

[C' : C]iiD, C) =I ord x ' t/J*(D)[k(x'): k(x)],


x'

where the sum is taken over all x' in C' lying above x. In other words,

t/J *t/J*(D) = [C' : C]D. C.

Proof Immediate consequences of Lemmas 1.1 and 2.2.

Let y be a closed point of Y. We define

i/D, E) = I iiD, E)[k(x) : key)]


x

where the sum is taken over all x E suppeD) (\ supp(E) lying above y.

Lemma 2.4. Let C be a curve on X, and having no common component


with D. Let U be an open set of X containing the (necessarily finite)
set of points x in IDI (\ ICI lying above y. Suppose D = (f) on U.
Then
i/D, C) = L ordx(fl C)[k(x) : key)].
x

If C is horizontal, then

where N k(C)IF is the norm.

Proof The first formula comes from Lemma 2.2 and the definition of
iy • Note that we can also write it on the normalization C' by Lemma
2.3, namely
i/D, C) = L ordAt/J*D)[k(x'): key)]
x'
[III,§2] INTERSECTIONS 57

and 1jJ*(D) is represented by IjJ 0 f The second formula comes from


Theorem 1.2.

Suppose that over the algebraic closure Fa the prime divisor C splits
into a sum of points

C= L (P;)
i= 1

which are conjugate to each other. Then we define

f(C) = n f(P;),

and extend f to arbitrary horizontal divisors by linearity. Then

Nk(C)/FUI c) = f(C).

It is important here to distinguish fl C and f(C) as defined above.


One may view X -+ Y as a family of curves n -ley) for y closed in Y,
in other words an "algebraic family". The fiber n-l(y) is defined locally
by one equation. Indeed, suppose we localize first R ail y. Let t be a
generator of the maximal ideal in the discrete valuation ring R y • Then
t = 0 is a local equation for n-l(y), which is thus locally principal.
On the other hand, let C be a horizontal curve in X. Then C is finite
over Y and C = spec(A) where A is finite over R, and torsion free as R-
module. Then A has a certain rank r which is called the rank, or degree,
of Cover Y, denoted by [C: Y].

Proposition 2.5. Let y be a closed point of Y and let C be a horizontal


curve on X. Then

In particular, this intersection is independent of y.

Proof We may assume R local at y. With the notation before the


proposition, the affine ring A has only a finite number of prime (maxi-
mal) ideals lying above y, and by the Chinese Remainder Theorem we
have

where x ranges over the maximal ideals of A above y. This relation


holds as a relation of k(y)-modules, and the result follows directly from
the definitions and Lemma 2.4.
58 INTERSECTIONS ON AN ARITHMETIC SURFACE [III, §3]

III, §3. FIBRAL INTERSECTIONS

We continue to assume that n: X --+ Y = spec(R) is an arithmetic surface,


and X is regular.

Theorem 3.1. Let C be a curve on X, contained in a fiber n- 1 (y). Let


D = (f) be a principal divisor on X without common component with C.
Then D. C is rationally equivalent to 0 on C. In particular, if C is a
curve over a field k finite over k(y), and l/!: C --+ spec(k) is induced by n,
then
l/!*(D.C) = 0,
and in particular,
iy(D.C) = o.
Proof This follows from Lemma 2.2 and Theorem 1.3.

We now define the self-intersection C. C as an element of CHo(C). In-


deed, we let f be a rational function on X such that C + (f) = D does
not have C as a component. Then let C. C be the class of D. C in
CHo(C). This class is well defined by Theorem 3.1. We can extend the
intersection D. E by linearity to any two divisors D, E which have no
horizontal component in common.
Let:

Divy(X) = free abelian group generated by the irreducible curves sup-


ported in the fiber n - 1(y).

We then obtain a pairing

defined as follows. Let c E CH 1(X) and E E Divy(X). Since X is regular,


there exists D 1 in c such that D 1 and E have no component in common.
We define

Theorem 3.1 shows that this intersection is well defined, independent of


the choice of D 1 • Instead of iy we also write deg (for degree) and abbre-
viate
iic, E) = deg(c. E) = (c. E).

Since there is only a finite number of irreducible components in n- 1 (y),


the free abelian group DiviX) is finitely generated. The fiber itself is a
[III, §3] FIBRAL INTERSECTIONS 59

divisor on X, and thus may be viewed as an element of Div/X), which


we denote by X y •

Interpretation with line sheaves. Let D be a divisor on X, and let Y


be a line sheaf isomorphic to (!)x(D). Let C be an irreducible component
of the fiber X y. Then Y I C is a line sheaf on C, and there exists a Car-
tier divisor a on C such that

Any two such Cartier divisors on C are linearly equivalent. Given an ef-
fective Cartier divisor 0 on C, we can define its Weil divisor as follows.
Let U be an open set of C such that the Cartier divisor is represented by
a function f For any point x of C we define

We define the Weil divisor

[0] = L ordia)(x),
x

and we define the degree

deg/o) = L ordx(a)[k(x): key)]·


x

This degree function extends naturally to all Cartier divisors, and is a


homomorphism into the integers. Then directly from the definitions and
Lemma 2.4, if CD is the class of D in Pic(X), we get

We also obtain a bilinear map

as follows. Let C, D be two fibral curves on X. If C =1= D, then we define

(C.D) = iiC.D).
If C = D, then we define

where c is the class of C in Pic(X). We can extend these values by bilin-


earity to the desired pairing, which is symmetric.
60 INTERSECTIONS ON AN ARITHMETIC SURFACE [III, §3]

Proposition 3.2. Let D be any fibral divisor. Then

and in particular

Proof Without loss of generality we may localize Y at y, and replace


R by the corresponding discrete valuation ring 0, because intersection
numbers are defined locally. Let t be a prime element of o. Then the
fiber X y is defined by the equation t = 0, so X y is a principal divisor
from which the desired assertion follows by Proposition 3.1.

We now see that our intersection pairing gives rise to a symmetric bi-
linear map

where the Q on the left means that we take the vector space generated
by, say, Divy(X) over Q. We shall study the signature of the associated
quadratic form.
Keeping the same notation, let

be the expression of a fiber as a sum of irreducible components with


multiplicities ni > O. Let

We call (a i ) the intersection matrix of the fiber.

For the rest of this section, we assume that X F is geometrically con-


nected. Then 7r*(!)x = (!)y and by Zariski's connectedness theorem, the
fibers Xy are connected. (See [Ha], Chapter III, Corollary 11.3.)

Proposition 3.3. The intersection matrix of the fiber has the following
properties.

(1) aij = aji for all i, j; that is, the matrix is symmetric.
(2) aij ~ 0 if i 0/ j.
(3) The row sums are 0, that is, I au = 0 for all i.
j

(4) For i 0/ j define i to be linked to j if aij > o. Then any two distinct
indices are connected by a chain of such linkings; that is, if i 0/ j
there is a sequence i = i l , i 2 , ... ,im =j such that iv is linked to iv+ l .

Proof The symmetry is obvious. If i 0/ j then C i 0/ C j so (C i . C j ) is


°
defined and ~ 0, which proves (2) since ni > for all i. The row sums
[III, §3] FIBRAL INTERSECTIONS 61

are 0 because L njC j = Xy and (D. Xy) = 0 by Proposition 3.2 for any
fibral divisor D. This proves (3). Property (4) immediately follows from
the fact that a fiber X y is connected. This concludes the proof.

Lemma 3.4. Let (a i) be a real symmetric matrix satisfying the condi-


tions of the preceding proposition. Then its corresponding quadratic
form is negative, and the null space is one-dimensional, spanned by
(1, 1, ... ,1).

Proof For any vector (... , Xi' ... ) we have

L aijXiXj = - L aiix i - Xj)2 ~ 0


~j i<j

after using (1), (2), (3) of Proposition 3.3. If Xi #- Xj for some pair r, ]
then a linking chain as in (4) shows that the sum over i < j has at least
one non-zero term and so the sum is #- O. This proves the lemma.

Proposition 3.5. Let D be a fibral divisor. Then (D2) ~ 0, and (D2) =0


if andonly if D = qX y for some rational number q.

Proof Let
D = LXiniCi

with rational Xi' Applying the lemma to (... , Xi' ... ) prov,es our theorem.

Since distinct fibers are orthogonal to each other, the theorem imme-
diately yields that if D is a fibral divisor such that (D2) = 0, then there
exist fibers F b ... ,Fm and rational numbers ql' ... ,qm such that

In particular, there exists a positive integer q and a divisor Z on Y such


that

The negativeness of the intersection product on fibral divisors is in


some sense the fibral part of the Hodge Index Theorem. We now come
to results which will be used in its globalization.
Let F be the quotient field of R as before. Let:

DivO(XF ) = group of divisors on X F of degree O.

This is the group of divisors on X F rational over F in older terminology,


and of course of degree O. For any irreducible divisor DE Div(X F ) its
62 INTERSECTIONS ON AN ARITHMETIC SURFACE [III, §4]

Zariski closure D in X is an irreducible horizontal divisor on X and


conversely every irreducible horizontal divisor arises in this way. We ex-
tend the association D 1---+ D by linearity to all of Div(X F). In particular,
D is defined for DE Divo(x F).

Theorem 3.6. There exists a unique linear map

such that for all DE DivO(X F) the divisor D + <l>y(D) is orthogonal to


DiviX).

Proof. By Theorem 3.5 the intersection pamng is a non-degenerate


symmetric bilinear form on the factor space QDiv y(X)/QX y. By Proposi-
tion 2.1, for fixed D of degree 0, the map

is a functional on this space, which can be represented by a unique ele-


ment of the space with respect to the intersection form by the non-
degeneracy. We define <l>y(D) to be minus this element to conclude the
proof of the theorem.

Corollary 3.7. Let f E k(X) be a rational function on X and let


D = (f)F he its divisor on X F. Then

(fh = D +L <l>iD) +L qyXy


y

with qy E Q.

Proof. The difference D - Uh is a fibral divisor, so the assertion fol-


lows from Theorem 3.6.

III, §4. MORPHISMS AND BASE CHANGE

Throughout this section we let X, X' be arithmetic surfaces over


spec(R) as before. We do not necessarily assume that the generic fiber
is geometrically irreducible, but we do assume X, X' regular. Let

(j): X' ---+ X

he a proper surjective morphism over Y = spec(R).

A morphism like (j) arises naturally as follows. Assume X F geomet-


rically irreducible. Let R' be the integral closure of R in a finite exten-
[III, §4] MORPHISMS AND BASE CHANGE 63

sion F' of the quotient field F of R, and let X' be a desingularization of


X x y Y', where Y' = spec(R'). Unfortunately, X x y Y' may not be regu-
lar.
Going back to cp as at the beginning of the section, let X* be the nor-
malization of X in k(X'). Then X* is finite over X, and cp factors:

<p

~
X -+X* -+X.
({J2

Now for inverse images. Let E be a curve on X. Then

cp*(E) = E' + fibral divisor,

where E' is defined as follows. The morphism cp is finite outside a


proper closed subset, say on an open set U. We let E' be the closure of
cp*(E) I U in X'. We call E' the proper inverse image of E, or the proper
transform of E. Viewing E as a Cartier divisor on X, its pull back cp!(E)
on X* is a Cartier divisor on X*, and has an associated Weil divisor

where ET are the irreducible components. If E is horizontal, then each


ET corresponds to a point of the generic fiber X F' such that k(Ef) is a
finite extension of k(E). The Zariski closure of ET in X' is an irreducible
divisor E; on X', and by the elementary theory of the splitting of a prime
in a finite extension we have the standard degree formula

L e;[k(E;) : k(E)] = [X' : X] = [X* : X],

which is just Lemma 1.1 applied to spec«(DE,X) and its normalization in


k(X'). Furthermore, the proper transform E' is precisely given by

Note that in going from X to X' via X*, we passed through the non-
regular scheme X*. We had to pay attention to the difference between a
Cartier divisor and a Weil divisor on X*, but we don't have to worry
about that on X'.
Note that if E is a horizontal divisor, then its proper transform E' is
also a horizontal divisor.
If D, E have no component in common, recall that

D.E = I iiD, E)[x] in CHo(IDI n lEI).


x
64 INTERSECTIONS ON AN ARITHMETIC SURFACE [III, §4]

If £ is a prime divisor and the support of D contains £, then

D. £ = class in CHo(£) represented by a divisor a


on £ such that (Ox(D) I£ >::; (OE(a).

Theorem 4.1 (Projection formula). Let D be a divisor on X, and let £'


be a divisor on X'. Let

!/J: <P - 1( ID I) n I £' I --> IDin <pC I £' I)

be the morphism induced by <po Then

in CHo(IDI n <p(I£'I)).

The next results prove the formula in the cases we need. Without loss
of generality, we can assume that £' is prime, so is a curve.

Theorem 4.2. Let V' be a curve on X'. Let D be a divisor on X. Let


V = <p(V'). Then
[V' : V]ixCD, V) if V is a curve, and D, V
~ ix,(<p*D, V')[k(x' ): k(x)] = { 0 intersect properly at x

if V is a point X.

In the second case, the sum on the left is to be interpreted as a sum


over all closed points of v' over k(x), and <p* D is replaced by a linearly
equivalent divisor not containing V'.

Proof Assume first that D, V intersect properly. Without loss of gen-


erality, we may assume that X = spec«(Ox,x) and D effective. Then D is
represented by a function f E (Ox, x. We consider the commutative dia-
gram:
V'~X'

V ---;----> X
)

where j, j' are the inclusions, and <PV' is the restriction of <p to V'. By
definition,
ix·(<p*D, V') = ordA(f <p)IV') 0

= ordA(f1 V) 0 <Pv.).
[III, §4] MORPHISMS AND BASE CHANGE 65

We multiply each side by [k(x'): k(x)] and use Lemma 2.3 to conclude
the proof when D, V intersect properly.

Note. If Viis a horizontal curve on X', then cp(V') is a horizontal


curve on X, and in particular is not a point. This is the case in the
applications.

Suppose next that V is a point. In this case, the sum on the left-hand
side of the formula is the degree of the line sheaf induced on V' over
k(x) by the pull back of (!J(D). But this pull back is trivial, since this pull
back can be taken via {x}. Hence this degree is 0, thus concluding the
proof of the theorem.

Theorem 4.3. Let X be an arithmetic surface over Y = spec(R), and


suppose X F geometrically irreducible. Let R' be the integral closure of
R in a finite extension F' of K, and let X' be a desingularization of
X x Y Y', where Y' = SpeC(R'). Let E be a prime divisor on X and let
E' be the proper transform of E on x'. Let D be a divisor on X not
containing E. Let y E Y. Then

[F': F]iy(D, E) = L iy.(cp*D, E') = L iy.(cp*D, cp*E),


y' y'

where the sum is taken over all points y' of Y' lying above y.

Proof This follows immediately from Theorem 4.2 and the standard
degree formula recalled in the previous discussion. Observe that the sec-
ond equality is due to the projection formula applied in the case when
cp(V') is a point, so the intersection of cp*D with the fibral divisor in cp*E
is O.

The next proposition gives us another useful property of base change


in a more trivial case.

Proposition 4.4. Let D, E be without common component. Let:

Rv = the completion of R at v = Vy for some closed y E Y;


y" = spec(Rv);

Xv = base extension of X to Y".

Let Dv. Ev be the extensions of D and E to Xv' Then


66 INTERSECTIONS ON AN ARITHMETIC SURFACE [III, §5]

This is a routine foundational result. First we observe that X v is


regular, because, for instance, if A = (!)x,x is the local ring of a closed
point on X, then by assumption A is regular, its completion is regular,
and this completion is also the completion of the base change A v, so Av
is regular. If f, g define D, E respectively locally at x, then they also de-
fine Dv , Ev and the proposition is then obvious.

Theorem 4.5. Assume that the generic fiber X F is geometrically con-


nected. Let Y' = spec(R') be the normalization of Y in a finite exten-
sion, and let X' be a desingularization of X x y Y'. Let t.p: X' -+ X
be the corresponding proper morphism. Let D be a divisor on X. Let
y be a closed point of Y and assume D ~ Div/X), that is, i/D, V) = 0
for all V E Divy(X). Then t.p*D ~ Divy,(X') for every y' E Y' lying
above y.

Proof After completing at y we may assume that R is complete and


so is R', and there is only one point of Y' above y. Then the assertion is
an immediate consequence of the projection formula.

The above theorem will be applied to the situation of Theorem 3.6,


where we defined the map

The inverse image t.p* D can be decomposed

t.p*D = D' + fibral divisor D",

where D' is the proper inverse image, and so

III, §5. NERON SYMBOLS

Let C be a curve, which in this section we assume to be a complete non-


singular curve, geometrically irreducible over a field F. Thus C = C F .
We let v be an absolute value on F.

We let:

DivO( C) = zg( C) = group of O-cycles or divisors of degree 0 on C.

DivO( C(F)) = subgroup of those divisors whose components are in


C(F) (the set of F-rational points of C).
[III, §5J NERON SYMBOLS 67

As usual, if DE Divo(C) is the divisor of a rational function J, and


E =I n;(P i ) is a O-cycle whose support is disjoint from D, we let
f(E) = TI f(P)"'.

We can summarize some of Neron's results in the next theorem as fol-


lows. Cf. [La IJ, Chapter 11, Theorems 3.6 and 3.7.

Theorem 5.1. Assume that C(F) is Zariski dense in C. To each pair of


elements D, E E DivO( C(F» with disjoint supports, one can define in one
and only one way a real number <D, E)v satisfying the following proper-
ties:

NS 1. The pairing is bilinear.

NS 2. rr D = (f) is principal, then <D, E)v = v f(E), where v(z) =


0

-loglzlv·

NS 3. The pairing is symmetric, that is, <D, E)v = <E, D)v.

NS 4. Fix PoE C(F) - suppeD). Then the map C(F) - suppeD) --> R
defined by

is continuous and locally bounded.

Note. Suppose F is complete with respect to v and locally compact.


Then locally bounded can be taken to mean the usual thing with respect
to the v-topology on X(F). In our applications we only need this case,
since as we shall see in the next paragraph, we may imbed a number
field F in its completion and pass to a finite extension. For the general
case, see [La IJ, Chapter 11, §3.

Now suppose that F c L is an extension (which in practice is finite,


or is the completion at v). Then a symbol <D, E)w is uniquely defined
for D, E E DivO( C(L», and any absolute value w extending v on L. By
the uniqueness property, the restriction of this symbol to pairs in
DivO( C(F» is the symbol with respect to F and v.
On the other hand, one may consider also cycles D, E whose absolute
components are not rational over F, but which are themselves rational
over F, that is, D, E E Divo(C). We define
68 INTERSECTIONS ON AN ARITHMETIC SURFACE [III, §5J

for any extension L over which D, E lie in DivO(C(L») and w extends v


in L. This value is independent of the choice of Land w. Whereas in
Theorem 5.1 dealing with rational points we need to assume C(F)
Zariski dense in C, in the present situation we do not have to make any
assumption on the existence of rational points. The field L could be
taken to be a finite extension of the completion, in which case C(L) is
certainly Zariski dense if there is one rational point, because there is a
whole neighborhood of rational points.
The following theorem from [Hr 1J does on curves what Neron did
for abelian varieties.

Theorem 5.2. Let F be the quotient field of the discrete valuation ring
R, with valuation v and closed point y. Suppose v normalized so that
v(z) = ordJz). Let X be an arithmetic surface over spec(R) as in §2.
For D, E E Divo(x F) with disjoint supports, we have

Proof Let C = X F' First suppose that we deal with the case when
C(F) is Zariski dense in C. Let D, E E DivO(C). It will suffice to prove
that the intersection pairing satisfies the properties of Theorem 5.1 char-
acterizing the Neron symbol. So let

This symbol is obviously bilinear. For NS 2, suppose D = (f)F' Then by


Corollary 3.7,

[D, EJ = iiD + <l>/D), E) = iiUh + qX y, E) = iiUh, E) = v feE).


0

As to the symmetry NS 3, we have the symmetric expression

because by construction, D + <l>y(D) is perpendicular to <l>iE). Finally,


NS 4 is satisfied because something stronger is true, namely, the symbol
is locally constant for the v-topology. For this we may pass to the com-
pletion Fv and Rv' We may also assume without loss of generality that
E is irreducible, so E is the Zariski closure of a rational point after pass-
ing to a finite extension if necessary, and we look at

Let P' be a point which is v-close to P in X(F J. Suppose that D is de-


fined locally by f = 0 in a neighborhood of P. Then the definition of the
[III, §5] NERON SYMBOLS 69

v-topology is such that v 0 f is continuous, and therefore va f(P') is con-


stant for P' in a neighborhood of P because of the discreteness of the
valuation. This proves the theorem when C(F) is Zariski dense. The
general case of the theorem is reduced to this one by passing to the
completion and a finite extension using Theorem 4.5. This concludes the
proof.

On the other hand, for archimedean absolute values we have:

Theorem 5.3. Let v be archimedean, and let A = Av be a N eron function.


Define

(P, Q)v = ).(P, Q)

for distinct points P, Q and extend this symbol to divisors D, E with


disjoint support uniquely satisfying bilinearity. Then the restriction of
this symbol to divisors of degree 0 satisfies the four properties NS 1
through NS 4, and is therefore the N eron symbol.

Proof The verification of the four properties is immediate. Note that


the constant in NF 2 disappears when we deal with divisors of degree O.

In the applications, the Neron function will be one half a Green's


function, according to our normalizations. If v comes from the absolute
value of a number field F, then we normalize the local symbol by:

((D, E'};v = (loglk(v)l)(D, E)v if v is finite


((D, E,};v = [Fv: Qv](D, E)v if v is archimedean.
CHAPTER IV

Hodge Index Theorem and


the Adjunction Formula

In a fundamental paper [Ara 2], Arakelov showed how to complete a


family of curves over the ring of integers of a number field by introduc-
a
ing the components at infinity, and getting divisor class group which in
many ways plays the role of the Picard group on complete surfaces.
The Hodge Index Theorem determines the signs of the intersection
form on the Arakelov divisor class group. It was done by Faltings
[Fa 1] in the semistable case and Hriljac [Hr 1]. We follow the latter,
by identifying the Arakelov intersection pairing with the Neron-Tate
form.
The label "adjunction formula" refers to the description of what hap-
pens to the canonical sheaf when restricted to a subvariety. We are con-
cerned here with a curve on an arithmetic surface. For a horizontal
curve, the theorem is due to Arakelov, and we formulate it for any a
priori given Neron family, not necessarily corresponding to the canonical
volume form at the places at infinity. For a vertical curve, the result is
standard, although there is no convenient reference if the curve is
singular, and lies on an arithmetic surface. However, the result depends
only on basic scheme theory in relative dimension 1. Hence we have
postponed discussing it until we need it for the proof of the Riemann-
Roch theorem in Chapter V, §3.
[IV, §l] ARAKELOV DIVISORS AND INTERSECTIONS 71

IV, §1. ARAKELOV DIVISORS AND INTERSECTIONS

Absolute values on number fields

Let F be a number field (finite extension of Q) and let R be its ring of


integers. We let Y = Spec(R). We let Val(F) denote the set of normal-
ized absolute values on F. If v E Val(F), then lal v is the absolute value
obtained by imbedding F in the completion F v' If z; is p-adic, then
Iplv = lip· If v is archimedean, then it is the ordinary absolute value.
We let

Iiali v = lal~v where

Then for a f= 0,

TI Iiali v = 1.
We let

Finally, we let Sa) be the set of absolute values at infinity.

Arithmetic surfaces

Let n: X ~ Spec(R) be a proper, fiat morphism, of relative dimension 1,


with X irreducible, integral, and such that if F is the quotient field of
R, then the generic fiber X F is geometrically irreducible. We also as-
sume that X is regular, so that X F is a complete non-singular geometri-
cally irreducible curve over F. If v is an absolute value on F we let

or

be the extension of X by the base change to F v or C v '

We shall repeat below some of the intersection formulas of the previ-


ous chapter in our present global setting, with the appropriate normal-
ization giving rise to global relations. We shall then define for any two
Arakelov divisors D, E without common component their intersection
number (D. E)v for all v, and then we define the total intersection number

(D. E);, = (D.E) = I (D.E}v·


72 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §1]

The intersection numbers (D. E)v for v E S", will be determined by the
choice of a Neron function Av.
We now pass to the definitions of the local intersection numbers.

Finite divisors, v finite.

Suppose first that D, E are two finite effective divisors without common
component. Let v = v" be a finite absolute value. For each closed point
x E X the residue class field k(x) is finite. Suppose k(x) is a finite exten-
sion of k(v). Let f, g represent D, E respectively in the local ring &x,x =
&x. Then we defined the intersection number iiD, E) by

iiD, E) = length of &xI(f, g) as &x-module,

and we define

(D. E)x = ix(D, E) 10glk(x)1 = log(&x: (f, g»,

where (&x: (f, g» is the finite index. Then we let

(D.E)v = I,(D.E)x·
xlv

INT 1. The symbol (D. E)v is bilinear and symmetric.

Suppose E is horizontal and irreducible. Let F~ be the algebraic clo-


sure of Fv' Let Xv = X ® F~ and Ev = E ® F~ be the extensions to F~,
so that Ev can be written as a sum

e
Ev = L (Q)
j= 1

of e distinct points, which are all conjugate over F. If f is a rational


function in F~(X), which has no zero or pole on E, then we define

e
f(Ev) = Df(Q)·
j=l

Note that f(Ev) is the norm


[IV, §1] ARAKELOV DIVISORS AND INTERSECTIONS 73

where fiE is the restriction of f to the irreducible divisor E, and k(E) is


the function field of E, which is finite over F, of degree e.
We shall also use the number e later, and we define the F-degree

degF(E) = e = [F(E) : F].

We note in passing that if D is a fibral divisor, we define its F-degree to


be 0, i.e. degF(D) = 0.

INT 2. Suppose D is represented by a rational function f on an open


set U containing the finite number of points x on X lying
above v in suppeD) n supp(E). If E is a horizontal irreducible
divisor, then

(D.E)v = - L 10gllflEII", = vif(Ev)),


i= 1

where the sum is taken over the primes Pi above v in the


number field k(E).

At the moment we are dealing with finite v, so we define

(D. E)fin = I (D. E)v'


v fin

INT 3. Let E be an irreducible divisor on X, so E has dimension 1,


and let E' be its normalization. Let

t/!:E'-X

be the natural morphism which is the composite of the normal-


ization morphism E' - E followed by the inclusion. Then for
any divisor D on X having no common component with E, we
have
(D. E)fin = degE • t/!*(D).

Here for any divisor D' on E' we define the arithmetic degree

degE·(D') = I ordp'(D') log Ik(P') I,


p'

where P' ranges over the closed points of E'.


The above property allows us to work with regular schemes, rather
than the possibly singular E (a singular curve on an arithmetic surface).
74 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §1]

INT 4. Let y E Y be a closed point. Let E be a jibral curve on X, i.e.


E is contained in X yo Let f be a rational function -=1= 0 whose
divisor does not contain E. Then

«(f) . E)fin = O.
In particular, the intersection product of any divisor with a fibral divisor
is defined on divisor classes modulo linear equivalence.

Next we let v E Sex)'

Throughout our discussion, we assume given for each v E Soo a Neron


function ).v in the sense of Chapter I, §1 so we have also the Neron
family P'D"J

We define the remammg intersections as follows, still for two finite


divisors D, E without common component.
If all components of D are vertical, we define (D. E)v = O.
Suppose D, E are horizontal and irreducible. We let Dv be the restric-
tion of D to the fiber X v over C v ' Then

D" = I
i= 1
(PJ

where the Pi are distinct points, which are all conjugate over F. Simi-
larly,

E" = I (Q).
j=l

We let N" = [Fv: Q,,] be the local degree, and we define

(D. E)" = N vAvCDv, Ev) = INvA,,(Pi, Qj).


i,j

Note that this intersection number depends on the original Neron func-
tion Av , and so should also be indexed by A, for instance by writing
(D. E);.,v' We omit the A frequently.

Vertical divisors at infinity. Let v E S 00' Then F v = R or C. We view


Xv as a complete non-singular curve over the complex numbers. By an
Arkelov divisor at infinity, we mean a formal linear combination

with rv E R.
[IV, §1] ARAKELOV DIVISORS AND INTERSECTIONS 75

We define the group of Arakelov divisors

Div AreX) = Divfin(X) EEl Div 00 (X),

where Div 00 eX) is the group of Arakelov divisors at infinity. Thus an


Arakelov divisor can be expressed uniquely as a sum D = Dfin + D oo '
where Dfin is an ordinary scheme divisor as in standard algebraic geo-
metry, and Doo is a formal linear combinations as above.
We then define the intersections when one of the divisors is a fiber Xv
at infinity, as follows. We let D be a finite divisor. We define:

(D.Xv)v = 0 if D is vertical;
(Xv,. Xv)v = 0 for all II, v' E Soo.
(D. X v)v = N v degF(D) if D is horizontal.

This last definition also applies to vertical divisors if we define degiD) =


o for any vertical divisor D. We had defined degF(D) for irreducible hor-
izontal divisors previously, and deg F is extended to all finite divisors by
linearity.
This concludes the definitions of intersection numbers.
Note that for the intersection of Xv and Xv' when v, v' E Soo we make
the intersection 0 even if v = v'. The reason for this is that for places not
at infinity, one has the "moving lemma": Given any finite divisors D, D'
there exists a rational function f such that CD) + (f) has no common
component with D'. This may not be true for the components at infinity,
where we allowed arbitrary real coefficients.

The Arakelov divisor of a function

Let f be a rational function in k(X). Then f has the ordinary scheme


divisor (fh, its finite part. We define the infinite part, depending on the
choice of Neron function A, to be

(fh = L NvyvCf)X v,
veSoo

where Yif) is the constant defined in Chapter I, §1, namely

and A(n.v is the Neron function. In this way we have associated an


Arakelov divisor
76 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §1]

to each rational function f We sometimes abbreviate

if the context is clear.

Theorem 1.1. Let D = (f)(X,).) be the Arakelov divisor of f. Then for


any divisor E with no finite component in common with D, we have

(D. E»). = O.

Proof Let (f);,. = L Nvyv(f)X v with the sum taken for v E Soo. Sup-
pose E horizontal irreducible. We have

(D.Eh = «(fh·E) + «(fh·E)


= L «(fh·E)v + L «(fh·E)v + L NvYv(f)degiE)
veSfin veS oo veSoo

= L NvV of(Ev) + L Nvl(f),v(Ev) + L Nvyv(f) degiE)


veSfin veS ao VESco

= L Nvvof(Ev) + L Nvvof(Ev)
veSfin veSCIO

=0 by the product formula.

This proves the theorem when E is horizontal irreducible. The cases


when E is vertical or a fiber at infinity follow directly and trivially from
the definitions, and will be left to the reader.

The group Div A.(X) = Div Ar contains the subgroup Rat(X,l), consist-
ing of the Arakelov divisors of functions f E k(X), f oF O. The factor
group is defined to be the Arakelov Chow group of Arakelov divisor
classes
CH 1 (X,1) = CH(X, 1) = Div Ar(X)/Rat(X, 1).

Given any two Arakelov divisors D, E there exists a rational function f


such that D + (f) and E have no common finite component. Then the
intersection «D + (f». E) is defined, and independent of the choice of f
by Theorem 1.1. Therefore the intersection product defines a symmetric
real-valued product on CH(X,l). We shall study the signature of this
product and relate it to the Neron symbols in the next section.
For the rest of this chapter, we let CH denote the group of divisor
classes. Thus on a complete curve over a field, CH is the Chow group
in dimension O. We want to use Pic for the group of isomorphism
classes of line sheaves, so we need two notations. Also for a curve over
a field, we let CHo be the group of divisor classes of degree O.
[IV, §2J THE HODGE INDEX THEOREM 77

Theorem 1.2. Suppose that X F is geometrically connected, i.e. F is alge-


braically closed in F(X). Let C1' C2 E CH(X, A). Let F' be a finite ex-
tension of F, and let X' be a desingularization of X x I' Y', where Y =
I

spec(R') is the normalization of Y in F'. Let qJ: X' -+ X be the corre-


sponding morphism. Then

Proof First we can find Arakelov divisors D, E in c l , C2 respectively


having no finite component in common. Then the formula for the finite
part has been proved in Theorem 4.3 of Chapter III, and for the A-part
of the intersections, it follows directly from the definitions.

IV, §2. THE HODGE INDEX THEOREM

The Hodge Index Theorem for arithmetic surfaces was done by Faltings
[Fa IJ in the semistable case and Hriljac [Hr 1]. We follow the latter.
We are interested in the intersection form on CH(X, A). Let D,
E E DivO(X F ). Then we define the global Neron symbol, which is none
other than minus the canonical height pairing, to be

where the sum is taken over all normalized absolute values on F. At


first this global symbol is defined only when D, E do not have any com-
ponents in common. But we defined X v' Xv' = 0 even when v = v' and
VES oo ' Given D, E there exists Dl linearly equivalent to D such that Dl
and E have no component in common except for v E S 00' and then
«.Dl' E'j) is defined, independent of the choice of Dl by Theorem 1.1, so
the global symbol is defined as a symmetric bilinear form on

Theorem 2.1 (Neron). The Neron symbol is negative on CHO(X F ), and


is negative definite on QCHO(XF ).

For a proof of this fundamental theorem due to Neron, cf. [La IJ,
Chapter 5, Theorems 5.2, 5.3, 7.2; and Chapter 11, Theorem 1.6.
We shall determine an orthogonal splitting of CH(X, A), exhibiting the
signature of the intersection form:

Theorem 2.2 (Hodge Index Theorem. Faltings-Hriljac). After tensoring


with Q, and modulo the kernel of the intersection form, the signature is
(+, -, ... , -).
However, the actual splitting will of course give more information.
78 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §2]

We identify
under DHD.

We have a direct sum decomposition (with y ranging over closed points


of Y):
Div Ar = Div(XF ) EB EB Divy EB Div 00

:::::> Rat(X, Je),

where this last group is the group of Arakelov divisors rationally equiva-
lent to 0, i.e. divisors of functions. We let the factor group of the last
two be denoted by

This yields an exact sequence

0-+ CHO(X, Je) -+ CH(X, Je) -+ Z -+ O.

Let R* be the group of units in R. Then

Div 00 n Rat(X, Je) = (R*)oo.

Let CHoo = Div,,)(R*)oo be the factor group of Div 00 by the image in


Div 00 of the units R*. We have a second exact sequence

0-+ CHoo -+ CH(X, Je) -+ CH(X) -+ o.

We observe that CH,Xl is orthogonal to Divo(X F) EB Divvet' and to it-


self, by definitions. In other words, CHoo is contained in the kernel of the
intersection pairing on CHO(X, Je).
After tensoring with Q, we obtain a direct orthogonal sum

OS 1.

where CHg n is the image in CHO(X, Je) of the finite divisors


whose component in Div(XF ) have degree o.

There remains to get an orthogonal splitting of QCHg n. To each


DE Div(XF ) we associate

(I + <D)D = D + L <DiD),
y
[IV, §2] THE HODGE INDEX THEOREM 79

where cD, is the orthogonalizing map of Chapter III, §2. Then / + cD in-
duces an imbedding

and Proposition 2.7 of Chapter III shows that we have a direct orthogo-
nal sum

OS 2. QCHg n = (/ + cD)QCHO(XF ) 1.. EB QDivy/QXy-


y

Furthermore, the sum over y is finite, for if X F has good reduction at y,


then the term on the right is 0.
By Chapter III, Proposition 2.5, the intersection form is negative de-
finite on the fibral sum, and by Neron's Theorem 2.1 and Chapter III,
Theorem 5.2, it is negative definite on the first term. This proves the
essential part of the Hodge Index Theorem, which we summarize in a
theorem.

Theorem 2.3. The subgroup QCHro is contained in the kernel of the in-
tersection pairing, and on QCHO(X, A.) modulo QCH oo ' the intersection
pairing is negative definite. The number of minus signs is equal to the
rank of the Mordell-Weil group plus the sum of the ranks of the
singular fibral summands QDivy/QXy for y ranging over the points
where X has bad reduction.

There remains but to look at the intersection of two dements which


arise from Arakelov divisors of non-zero degrees. Whereas in the classi-
cal case of surfaces over a field there is only one such element, here we
can pick a finite class c from CH(X F ) of degree > 0, say; and we can
pick a class e to be the class of a fiber Xv, v E Soo. After subtracting
from c a suitable element of CHg n we may assume without loss of gener-
ality that c is orthogonal to CH?in. We already know that e is orthogo-
nal to CHg n. There remains only to orthogonalize c and e.
Note that (c. e) "# 0. After multiplying e with a suitable real number,
we may assume without loss of generality that (c. e) = 1. Now it is a tri-
vial matter to get an orthogonalization of c, e. We choose real numbers
x and y such that

(c + xe)2 = 1 and (c + ye)2 = -1.

We can solve for x and y uniquely and let


1 - c2 1 + c2
e1 =c+-2- e and e2 = c - -2- e.

Then e 1 e 2 = 0, and the Hodge Index Theorem is proved.


80 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §3]

IV, §3. METRIZED LINE SHEAVES AND INTERSECTIONS

On an arithmetic curve

Let A be an order in some number field F(A), that is, A is a ring of


algebraic integers, finite over Z, of rank [F(A): Q]. We let

Z = spec(A).
Then Z is a curve, i.e. Z has dimension 1, and Z may have singularities.
Let D be a Cartier divisor on Z. Then we know how to associate a Weil
divisor [D] to D, and
[D] = L ordZ<D)[z],
z

where the sum is taken over the closed points. We define

If D = (f) locally at z, and f E lDz,z we have directly from the definitions

where the index is the number of elements in the factor module. For ar-
bitrary D (i.e. not necessarily effective) the degree is defined by additivity.
Let !/J: Z' --+ Z be the normalization of Z. Then the general formula
for orders on curves gives us:

degz(D) = L degA!/J*D),
z'lz

where the sum is taken over all points z' of Z' lying above z.
A line sheaf 2 on Z being coherent, there exists a module L =
HO(Z, 2) over A such that 2 = L-. The module L is locally free of
rank 1.
Let WE S",,(F(A». The base extension of .P to spec(F(A)w) is then a
one-dimensional vector space over F(A)w, which will be denoted by 2w.
We identify

By a metric p(w) on 2 we mean a norm on L w , so this norm satisfies


the triangle inequality and satisfies

for a E K(A)w, s E Lw'


[IV, §3J METRIZED LINE SHEAVES AND INTERSECTIONS 81

By a metrized line sheaf on Z we mean a line sheaf with a metric pew)


for each w E S 00 • We denote such a sheaf by a pair (2", p). Note that a
metric gives rise to the normalized function

where N w = [F(A)w: Qw].

When referring to isometries, we mean with respect to this normalized


function. We can further extend the metric to L ® A C w, where C w is the
complex field. The absolute value w over v is determined by [F(A)w: FvJ
imbeddings of F(A) over Fv into Cv. If 0': F(A) -+ C v is such an imbed-
ding, we also write p(O') instead of pew), and we then get a metric on the
sheaf 2"a determined by this imbedding. It is sometimes useful to deal
with the specific imbeddings 0'.
We let LF(A) = L ®A F(A), so LF(A) is a one-dimensional vector space
over F(A). We shall now define the degree of a metrized line sheaf.
Let s E LF(A)' S =I- 0, so s is a global rational section of 2". For each
closed point Z E Z there is an element gz E F(A) such that the stalk of 2"
at Z is principal:

We define the local degree at z to be

degs.zCL) = degs j2") = degigz);

and for WE SooCF(A»)

degs,w(L, p) = degs ,w(2", p) = -N w log Islp(w) = -log Ilsllp(w)'

We define the global degree to be

deg.(L, p) = deg.(2", p) = L deg j2") + L


s degs ,w(2", p).
z

If we change the section s to hs with some hE F(A), h =I- 0, then degs (2")
changes by
L degih) + L
z weS ao
-logllhll w = °
by the product formula. Indeed, the first sum over the closed points of
the scheme is the same as the similar sum taken over the closed points
of the normalization, and the product formula applies. Hence the degree
function is defined for an arbitrary metrized line sheaf (2", p), and we
therefore write the global degree as
deg z(2", p) = deg(2", p) = deg.(2", p)
without the subscript s, and with or without the subscript Z on deg.
82 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §3]

We note that the function

is a homomorphism. The operation of multiplication on pairs (.!f, p) is


that of tensor product both for the sheaf and the metric.
Let
cp: Z' --+ Z

be a proper surjective morphism of curves as above. Let (.!f, p) be a


metrized line sheaf on Z. Then we obtain a metrized line sheaf

cp*(.!f, p) = (.!f', p')


on Z' in a natural way. The pull back is already known for the finite
part, and if w' is an absolute value of F(Z') extending w, then p'(w') is
the norm on L ® F(Z')w' such that

for s E L, N(w', w) = [F(Z')w' : P(Z)w].


Then we have the expected formula:

Proposition 3.1. Let cp: Z' --+ Z be a proper surjective morphism of


curves as above. Then

deg z ' cp*(.!f, p) = [Z' : Z] degz(..'l', p).

Now consider the sheaf (!Jz(D). Locally at z this sheaf is locally free,
with basis f -1 for some f E peA) representing D locally. If.!f is an arbi-
trary line sheaf on Z, then .!f ~ (!Jz{D) for some Cartier divisor D.
From an isomorphism (in the sense of algebraic geometry)

and a metric on .!f we obtain a metric on (!Jz(D) such that the isomorph-
ism preserves the norms at each WE Soo(F(A».
Let .!f = L -. An isomorphism .!f --+ (!J z( D) induces an imbedding of L
in the function field F(A) of Z. Note that the Cartier divisor D is well
defined up to a principal Cartier divisor.
If .!f = (!Jz{D), then .!f is a subsheaf of the constant sheaf F(A), and L
is an A-submodule of peA). In this case, we may take s to be the ele-
ment 1 of the function field F(A). We shall write deg1 . .(.!f) or deg1.z(L)
for closed points z, and deg 1,w(.!f, p) or deg1.w(L, p) at infinity, Then
[IV, §3] METRIZED LINE SHEAVES AND INTERSECTIONS 83

and for WE Soo(F(A)),

In particular, we can always put what we shall call the trivial or standard
metric l' on (!)zfP) such that 11111. = 1. Then for each wESoo(F(A)) we
have

For convenience, we shall use the index (A: L). If L is contained in


A, then it is the usual index. In general, let a E A, a =F 0, be such that
aL c A. Then

(A . ) = (A : aLl
.L (A: aA)"

The quotient is independent of the choice of a. Cf. the diagram of


Lemma 1.1 in Chapter III. Entirely in terms of modules, we can express
the degree as follows.

Proposition 3.2. Let Z = spec(A), where A is a ring of integers in a


number field F(A), with A finite over Z. Let L c F(A) be a locally free
A-module, and !l' = L -, For each WE Soo(F(A)) suppose given a metric
pew) on L. Let A' be the integral closure of A in F(A). Then:

degz(!l',p) = -logN(LA') + L -log II 1 IIp(w)


WES",,(F(A))

= -log(A : L) + L -loglllllp(w)
w

= -log(A': LA') + L -loglllll p(w)·


w

Proof The proposition is a direct consequence of the ddinitions. The


index (A: L) is a product of the local indices (A z : L z ) for the closed
points z of Z, and from the definition,

so the formulas are clear.

Finally, we make the connection with Arakelov divisors on Z.


By an Arakelov divisor on Z we mean an element of the group

DivA.(Z) = Cartier divisors Et'l L R[w],


weSoc
84 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §3]

where the expression on the right is the vector space over R generated
by the free elements [w] corresponding to the archimedean absolute
values of F(A). An Arake10v divisor can then be written as a sum

D = Dfin + L r(w)[w], where Dfin is a Cartier divisor.


w

To each such Arakelov divisor we can associate a metric pew) on F(A),


namely the metric such that

-log I 1 IIp(w) = r(w).

In other words, the numbers r(w) determine the metric at w, compared


to the trivial metric such that 11111 w = 1. This association induces a ho-
momorphism

Div Ar(Z) -> Isomorphism classes of metrized line sheaves on Z.

If we define the degree of an Arakelov divisor as above to be

then this homomorphism preserves degrees. These statements follow im-


mediately from the definitions, and their verification will be left to the
reader.

On the arithmetic surface X

For each v E SC() we fix a Neron function Av, which we assume is a Weil
function on X v x X v with respect to the diagonal.

In Chapter I, §2 we considered metrized invertible sheaves over the


complex numbers. We can now tie up the notion globally on our
scheme X. Let .2? be an invertible sheaf on X, so that .2? ~ @x(D) for
some finite divisor D on X. For each v E Soo we let .Pv be the line sheaf
on X v obtained by base change from Y to F v' which we view ali natur-
ally contained in its algebraic closure Cv.
We say that .2? is a metrized line sheaf on X if we are given a metric
p(v) on .2?v for each v E S 00. Having fixed A, let

D = D fin + L rvXv
veSoo
[IV, §3] METRIZED LINE SHEA YES AND INTERSECTIONS 85

be an Arakelov divisor. The part at infinity is determined by the


numbers r v' V E S 00 • We define the associated metric on (!) x(D fin ) to be
the metric P;",D,v such that for P not in the support of D f :in we have

-loglll(P)ll p .<,D,v = NvAD,v(P) + r(v).

This is a change by r(v) of the metric of Chapter I, Proposition 2.2, and


corresponds to the additional data of the Arakelov part of the divisor D.
The metrics associated as above to Arakelov divisors will be called ..1,-
admissible, or simply admissible when A is fixed.
A metrized line sheaf on X will be denoted by the pair (L, p), or
simply !t'p. We say that (!t', p) is A-admissible, or simply admissible, if
(!t', p) is metrically isomorphic to some «(!)x(Dfin ), P;..,D)' For simplicity
we could also write this pair as

The first expression is the shortest, and D being an Arakelov divisor, the
first expression contains all the data of the other two.

Since A is fixed throughout the discussion, we can also omit the subscript
A, and write simply (!)x(D).

For the intersection theory, the notation with the Neron function was
more appropriate, since AlP, Q) gives the local intersection number of P,
Q viewed as sections. On the other hand, working with sheaves as we
shall do from now on makes the notation with the metric more useful.
One has to be able to go back and forth between the two" depending on
the context. The next theorem gives the precise connection between the
two. We let the Arakelov-Picard group be

Pic(X, A) = isomorphism classes of metrized line sheaves (!t', p) with ..1,-


admissible metrics p.

Theorem 3.3. The map

for DE Div Ar(X)

induces an isomorphism between CH(X, A) and Pic(X, A). Multiplication


by a rational function f -1 induces an isomorphism
86 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §3]

Proof The theorem is an immediate consequence of the definitions.


First for the finite part, the statement is part of what we assume about
very elementary algebraic geometry. At infinity, the correcting term (fh
in the Arakelov divisor of f is precisely what is needed to make multipli-
cation by f an isomorphism, in distinction to Proposition 3.2 of
Chapter I. In that proposition D was an ordinary divisor on X, whereas
now we are dealing with Arakelov divisors.

Intersections

We continue with the same hypotheses concerning the morphism

n: X -+ Y = Spec(R)

as stated at the beginning of §l.


Let E be an irreducible horizontal divisor on X. Then the embedding

E---->X

\/ Y

into X over Y is a local complete intersection, which we also call a regu-


lar imbedding, since E is a Cartier divisor on the regular scheme X.
Let (2, p) be a metrized line sheaf on X. For each v E SOC) we are
therefore given a metric Po on 2v. Suppose E = spec(A), so E is an
arithmetic curve Z as at the beginning of this section. For each
WE SOC)(F(A») we get a class of [F(A)",: Fv] conjugate imbeddings of F(A)
in C v over F v. Each such imbedding gives rise to a point P E X(C v), and
(2v, Po) induces a norm on the fiber 2v1P. Therefore by restriction, we
obtain a metrized line sheaf

(2, p)IE.

As we have seen at the beginning of the section, such a line sheaf has a
degree. For convenience of notation, this degree will also be denoted as
an intersection number

I deg E (2,p)IE = (2,p).E). I


Theorem 3.4. Let E be a horizontal curve on X. For DE Div A,(X) we
have
[IV, §4] THE CANONICAL SHEAF AND THE RESIDUE THEOREM 87

Proof Both sides depend only on the class of D in CH(X, A.) by


Theorem 3.3. So without loss of generality we may assume that Dfin
does not contain E as a component. Then the local intersections
numbers can be computed separately for v finite and v E S,IO' and we have
an equality locally:

(D.E)v = L deg1,w(Dx(D)IE if wi v, wESCX)(F(E»);


wlv

(D.E)y = L deg1,z(Dx(Dfin)IE,
zlY

where this second sum is taken over ZED r1 E lying above Y E spec(R).
We leave the detailed checking to the reader, since these relations are
immediate consequences of the definitions.

Remark. Of course, the index A. should appear on both sides of the


expressions in Theorem 3.4, since A. determines both the intersection at
infinity (D. E);. and the metric of (Dx(D);.. But we sometimes omit the A.,
and write the formula of Theorem 3.4 as

(D. E) = deg E (Dx(D)IE.

The difference between this formula and the scheme-theoretic analogue is


of course that here, we deal with an Arakelov divisor D, rather than D fin
on X.
It will be useful to deal with the intersection product in terms of line
sheaves without passing back to Arakelov divisors. Thus if (9', p),
(uI{, p') are metrized line sheaves on X with admissible metrics p, p' we
define their scalar product

«9', p), (uI{, p'» = (D. E)

if the line sheaves correspond to Arakelov divisors in the isomorphism of


Theorem 3.3.

IV, §4. THE CANONICAL SHEAF AND


THE RESIDUE THEOREM

The purpose of this section is to define the canonical sheaf and to prove
one of its basic properties, the residue theorem. We first recall quickly
the adjunction formula in the geometric case.
88 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §4]

Let X be a complete non-singular surface and let C be a non-singular


curve on the surface, over k. We let 0i = 0i/spec(k) and similarly for C.
The curve C is defined by a sheaf of ideals J, which are locally princi-
pal. We have the fundamental exact sequence

from the basic theory of derivations, and J / J2 is called the conormal


sheaf:

Taking the top exterior power, we get a natural isomorphism

or tensoring on the other side, we obtain the adjunction formula for sur-
faces:

where ~ = A 2 0i is called the canonical sheaf.


In the arithmetic case, we meet with two differences. The first has to
do with the triviality of horizontal differentiation. For any derivative d,
we have d(l) = O. Secondly, there is no ground field k. Thus we have to
do everything relative to the base.
We now develop the basic theory of the canonical sheaf that we shall
need.

Let X .... Y be a morphism of finite type, where Y = spec(R), and all we


need for R at the moment is R is Noetherian.

The following considerations apply to any closed imbedding

j:X .... s
with p: S ~ Y smooth, and X is a local complete intersection in S. This
means that the ideal sheaf J defining X in S is generated at every point
by a regular sequence, which we assume to be of the same length. The
conormal sheaf is the locally free sheaf
[IV, §4] THE CANONICAL SHEAF AND THE RESIDUE THEOREM 89

We then define the canonical sheaf of the imbedding to be the line sheaf

Here det = A'0P is the top exterior product. We recall below a proof that
is independent of the factorization po j, up to a natural isomorph-
illpoj

ism. We then write ill x /y instead of ill poj ' that is

ill x /y = det j*n~/y ® det 'fi s , x/


and call ill x /y a canonical sheaf, determined up to isomorphisms. The
proof that illpoj is independent of the factorization po j up to isomorph-
ism follows by applying the determinant homomorphism to the corre-
sponding (and stronger) statement concerning the tangent element in the
Grothendieck group, see Fulton-Lang [F-L], Chapter V, Proposition
7.1(i). For the convenience of the reader, we recall the argument from
scratch. First, recall that if

o--+ 8' --+ 8 --+ 8" --+ 0


is an exact sequence of vector sheaves (locally free sheaves of finite rank),
then there is a unique isomorphism

det 8' ® det 8" ~ det 8

such that, if t'l' ... ,t~ is a local basis of sections of 8', t'~, ... ,t~ are local
sections of 8 mapping onto a local basis l~, ... ,l~ of 8", then

(t'l 1\ ••• 1\ t~) ® (l~ /\ ... /\ l~) f--+ t'l /\ ... /\ t~ /\ t~ /\ ... /\ t~.

Now suppose given two imbeddings j: X --+ Sand j': X --+ S' in smooth
schemes over Y, we let T = S Xy S', so we get a commutative diagram

From the elementary theory of local complete intersections (see, for in-
stance, [F -L], Chapter IV, Proposition 3.9) we have an exact sequence

0--+ 'fi XIS --+ C(j X/T .... i*nhs --+ 0


90 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §4]

uTIS = I.* q'*nl


an d I'*nl US'/Y' I n t erch angmg
US'/Y = ]"*nl . S an d S' b y symme t ry,
and applying determinants (that is, N°P as previously indicated) gives the
desired isomorphism between (Opoj and (Op'oj"
Note that [F -L], Chapter IV, §1, §2, §3 provides an introduction to
the theory of local complete intersections. It was written with this pur-
pose in mind.

Remark on notation. We have written (0 as a capital letter because


sheaves are written with capital letters, and to distinguish the sheaf from
differential forms themselves. We may call (0 script omega, to distin-
guish it from 0, which denotes the ordinary differential sheaf.

In certain circumstances, a canonical sheaf is also a dualizing sheaf for


cohomology. See, for instance, [Ha 1], Chapter III, Theorem 7.11. This
property is irrelevant here. For our purposes a canonical sheaf replaces
the sheaf of ordinary differential forms when there is not enough smooth-
ness to make the latter serviceable. The following property follows im-
mediately from the definition.

CS 1. Let V be the open subset of X such that V -+ Y is smooth.


Then in a given isomorphism class, we can take

One defines a morphism f: X -+ Y to be regular if it is the composite


of a local complete intersection imbedding and a smooth morphism.
Then by applying the determinant to the basic exact sequences for the
sheaf of differential forms, or simply to the relation of [F - L], Chapter V,
Theorem 7.1(ii), we get:

CS 2. Let
f
Z-+X-+Y

be a composite of regular morphisms. Then

This last property may be viewed as an adjunction formula in scheme-


theoretic context, or as a vast generalization of the multiplicativity of the
different in towers.
A more general theory of the canonical sheaf will be developed by
Kunz and Waldi in a forthcoming monograph (AMS, Contemporary
Mathematics). In this chapter, we shall need a special case. Suppose now
that:
[IV, §4] THE CANONICAL SHEAF AND THE RESIDUE THEOREM 91

R is integral with quotient field F;

X is an integral scheme with function field F(X) = L;


X/Y is dominant and generically smooth, i.e. the field extension L/F is
separable.

In this case there is a canonical choice for OJ x /y as a subsheaf of the


constant sheaf det nltF (the sheaf of top degree rational differential forms
of X/Y). In fact, let z be the generic point of X. Then

is a subsheaf of the constant sheaf with fiber

Since the canonical sequence

is exact and ni/y,z = nltF' we obtain a natural isomorphism

which gives an imbedding

The image of OJ x /y in det nifF is independent of the factorization


X ~ S -+ Y. This follows by a slight modification of the proof that OJ x /y
is unique up to isomorphism.
In the sequel we regard OJ x /y as this intrinsic subsheaf of det ni/F'
and we may refer to this sheaf as the canonical sheaf. We then have

at all points x E X where X -+ Y is smooth, since

is an exact sequence of free @x,x-modules, whence


92 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §4]

The canonical sheaf for a horizontal curve

From here on, we let X be an arithmetic surface over a Dedekind ring


of characteristic zero, and perfect residue class fields at all the primes.
We assume X regular, but of course X -+ Y is not necessarily smooth.
We assume X F geometrically irreducible.

Let E be a horizontal curve on X. Then E is finite over Y, and E =


spec(A), where A is finite over R. The restriction EF of E to the generic
fiber X F is then a point, which we denote by P, on X F' The field

k(E) = k(P) = F(P)

is a finite extension of F. We have the canonical sheaves

and

Since X is regular, E is defined by a sheaf of ideals § which are locally


principal, and we have an equality

just as in the geometric case. If x is a closed point of E, then E is de-


fined locally by a single element t in (!)x.x, say, in other words

This same element t generates the maximal ideal

when we restrict to the generic fiber. Tn other words, we can use t as a


local parameter at the point P on X F • We have
[IV, §4J THE CANONICAL SHEAF AND THE RESIDUE THEOREM 93

Since E = spec(A) is finite over Y = spec(R), and is affine, the sheaf


WEIY can be expressed as

and M is a module locally principal over A. The first question is: What
"is" M? Theorem 4.1 will given an answer to this question, and should
be seen as a basic fact of algebraic geometry.
We define the dualizing module (the Dedekind complementary module)

WEIY = WAIR = {b E F(E) such that Tr(bA) c R}.

Here Tr denotes the trace from F(E) to F. We note that WAIR is given
with an imbedding in the function field F(E). The dualizing module
bears its name because it represents HomR(A, R) under the pairing given
by the trace. We shall deal with the case when WEIY is locally principal.
Then we let

so l)EIY is the different.


In the applications, we shall be in the number field casl~. Then we can
define the relative logarithmic discriminant to be

dEIY,fin = dEIY = 10g(WAIR : A) = 10g(A : l)AIR)'

Since WAIR::J A, the index on the right is the usual one. Then the rela-
tive discriminant can be computed in terms of the integral closure A' of
A (or on the normalization of E) to be

where N is the absolute norm. Cf. the diagram of Lemma 1.1, Chapter
III.

We shall define a canonical isomorphism, called the residue isomorph-


ism:

We do this pointwise. Let x E X be a closed point in E. Abbreviate


94 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §4]

The stalk :Ex is contained in the stalk of the generic fiber

Using the function t as local parameter at the point P, we can then use
dtjt as a basis for the stalk on the right-hand side over the local ring of
P on the generic fiber, so we have the inclusion

. ( 1 ) _ n dt
:Ex 4 :Ex,F 4 (lX/F,F ® (lJxF(EF) P - (lJP,XF ( '

Since :Ex is free of rank lover (lJx,x it follows that

Hence we have a natural injection, the residue mapping

Theorem 4.1 (Residue theorem). Let :E = W X/f @ (lJx(E). The image of


the residue map

is the stalk WE/f,x, and so we get an isomorphism

Proof. This theorem is imbedded in the general Grothendieck duality,


with latter material as in [A-K], [Ho], [KI]. I asked Lipman if there
was a direct known proof, and he gave me a proof in case the fibers are
generically simple on X, using the ideas of Kunz and directing me to
Kunz's work. Then Kunz himself communicated to me the following
short and elegant proof in general, which he found with R. Waldi. We
refer freely to Kunz's book Kahler Differentials [Ku].
The following lemma describes the stalk W X / f at x.

Lemma 4.2. Under the hypothesis of Theorem 4.1 we have

where A is a generator of the Kahler different (the Jacobian ideal) of


over R[t] (as defined in [Ku], §1O).
(lJx,x
[IV, §4] THE CANONICAL SHEAF AND THE RESIDUE THEOREM 95

Proof. We use the notation of the beginning of the section, with an


imbedding j: X -+ S in a smooth scheme S over Y. Consider the commu-
tative diagram with exact rows and columns:

0 0

j j
/x/,P; .$"

j j
n~s,rlR ® (9x,x ~ ~s,rlR ® (9x,x

j ja
0 - - (9x,xdt-- n~x,rIR )~ X,rlR[tl ----+0

j j
0 0

where $" is the kernel of oc. Since t is transcendental over F, and F has
characteristic 0, the module (9x,x dt is free, By the Snake Lemma, there
is an exact sequence

hence a natural isomorphism

and

The exact sequence

can be used to compute the Kahler different l)K«(9x,;</R[t]). Since


n~x,rIR[rl is a torsion module, the free modules $" and n~s,rIR ® (9x,x
have the same rank. Therefore the different is a principal ideal
96 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §4]

Moreover, we have
:f{ ® L = n~s,x/R ® L,

Under the natural injection ())x/y,X -+ nLF the module

is identified with an (i)x,x-submodule of det(:f{ ® L)V ® det(:f{ ® L) ~ L,


hence with a fractional (i)x,x-ideal in L, and this ideal is

as is easily seen. This proves Lemma 4,2,

We now return to the proof of Theorem 4.1. Since n~x,p/R[tl = 0, we


have

Thus Ll is a unit of (i)x,p, Moreover, by [Ku], 10.3b, we have

where .1 is the image of Ll in (i)E.x = (i)x.x/t(!)x.x'


By the lemma, we have

and the image of !l'x under res x is

It follows from [Ku], 10.17, that 1J,J..(i)E,x/R) equals the Dedekind differ-
ent 'nD«(!)E,x/R) = (WE/y );l, In fact, we may pass to the completions of
(i)E,x and the local ring of R lying under (i)E.X' Then we have a finite
ring extension which is a complete intersection, since (!)x.x/R was one by
the assumption of Theorem 4.1 saying that .$x is generated by a regular
sequence, By [Ku], 10.17, the differents agree for the completions. (Cf.
[Be].) It is easy to see then that they agree also before completion,
whence

This concludes the proof of Theorem 4.1, the residue theorem.


[IV, §5] METRIZATIONS AND ARAKELOV'S ADJUNCTION FORMULA 97

IV, §5. METRIZATIONS AND ARAKELOV'S


ADJUNCTION FORMULA

So far, except for the use of indices to compute the discriminant, we


could work over an arbitrary Dedekind ring. Now suppose we are in
the number field case. Having imbedded W EIy into the constant sheaf of
the function field of E by the residue, we can then give WEIY the trivial
metric r induced by the absolute values WE S ",,(F(E». Under these
absolute values, we have

-log 11111 w = O.

As in §3, we write the metrized canonical sheaf as

or simply

if we omit the r for simplicity. Unless otherwise specified, the metric on


WEIY will be this trivial one under the residue mapping.
Let K ElY be the Cartier divisor on E determined locally by the locally
principal module WEIY' We call K Ely the relative differential divisor.
Then

and we have the corresponding residue isomorphism


WEIY ~ (9(K Ely )·

Hence we can express the logarithmic discriminant as

Since the metrics are trivial, the degree of the canonical sheaf is then
entirely given by its finite components, and following thc~ definitions of
§3, we find:

Theorem 5.1

Next we give metrics to the sheaves W XIY and (9x(E). Since metrics de-
pend only on the base change of these sheaves to C v for each absolute
value v, we may replace X by X v' viewed as a complete non-singular
curve over Cv. Then the base change of W XIY to C v is simply nt. To
avoid double indices, we fix v, and we let

so C is a complete non-singular curve over the algebraically closed field


Cv, which we denote by k, again to avoid double indices.
98 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §5]

There is a canonical representation of n~/k as the conormal sheaf of


the diagonal imbedding

Let J be the ideal sheaf which is the kernel of multiplication

Let Ii be the diagonal. Then

and from the basic theory of differential forms, the natural map

is an isomorphism. Hence for each point Q E C(k) we have an isomorph-


ism

Given the absolute value v and the N eron function Av , which we assumed
to be a Weil function on X v x X v with respect to the diagonal:
Let n~/klQ have the metric induced by PA., -Q,v under the isomorphism

In other words, the metric on n~/k is the restriction of the metric P;., -A,v
on (1) x x x( - Ii) to Ii. This makes it clear that the metric varies smoothly
with Q.
e
Let Ev = I. P j be the splitting of E over C., so Ph ... ,Pe are distinct.
j= 1

Then
e
(1)C<E) = (8) (1)C<P j ).
j= 1

Let (1)c(E) have the metric P;',E v ,. = (8)j=l P;',PJ'.'

Let Q = Pi for some i. Then

9'. = (WX/f,v ® (1)x(E»v has the metric P;',Ev-P"v at the point Pi'
[IV, §5] METRIZATIONS AND ARAKELOV'S ADJUNCTION FORMULA 99

Note that Pi is not in the support of Ev - Pi' Hence from the defini-
tions, and Proposition 2.2 of Chapter I, we get:

Proposition 5.2. Let Pi be the point induced by an imbedding


(Ji: k(E) -> C v over F v , and inducing the absolute value w on k(E). Then

the sheaf 2'v = (W x /y ® (lJx(E»)v at the point Pi has the metric pew)
such that, if WE 2' v,p, is an element of the stalk, then

-log IlwIPillp(w) = -log IIresp/w)ll w + L N),.(Pi,P;),


i*i

The sum on the right-hand side of Proposition 4.3 leads us to define a


quantity which is the analogue at infinity for the discriminant at finite
places, namely we define the logarithmic A-discriminant to be

d",v(E) = L AvCPi , Pi) and d".F(E) = diE) = ~~ Nvd"jE).


i*i veS ao

Whereas the ordinary discriminant measures the extent to which two fin-
ite points come together in the residue class field, the quantity d",v mea-
sures how close they come together with the Av-measure at infinity. Each
quantity tends to + 00 as points are closer together. We then find:

Theorem 5.3 (Arakelov adjunction formula). Let E be a horizontal curve


on the arithmetic surface X. Then with the above prescribed metrics, we
have

Remark. In Theorem 5.1 we took (IJ(K E / y ) with the trivial metric. We


could give (lJE(K E/y ) a metric p, depending on A, in such a way that the
whole right-hand side could then be written as the degree of an Arakelov
divisor on E, or as the degree of a metrized line sheaf on E, namely

Then the formula would be formally identical with the formula in the
geometric case over an algebraically closed field. In any case, we can
also adjust the notation, and specify dE/y.fin for the finite part of the dis-
criminant, just as we specify

for the part at infinity, depending on our choice of Neron function. In


that case, we could simply write
100 HODGE INDEX THEOREM AND ADJUNCTION FORMULA [IV, §5]

in the usual systematic way. Depending on the context, we recommend


the use of such notation.

The proof of the adjunction formula has essentially been carried out.
The residue isomorphism was used to give a "trivial" metric to W E / y via
WilY in Theorem 4.1, while the extra term in the formula of Proposition
5.2, summed over i, gives exactly the quantity d ..(E). Hence the adjunc-
tion formula is a direct consequence of the definitions. The non-formal
content has been pushed into the statement and proof of Theorem 4.1.
Note that it was the relative discriminant which entered the adjunc-
tion formula. If we let the absolute logarithmic discriminant be the log of
the absolute discriminant:

then we have the formula from algebraic number theory

or dA /Z = d A / R + [F(A) : F]d R / Z .

Remark. Arakelov himself stated the adjunction formula under some


restrictive conditions. Several people have tried to formulate adjustments
to give the result in appropriate generality. Hriljac [Hr 2] made an ap-
plication of the generalized formula. Chin burg showed in the proof of
[Ch 1], Proposition 4.1, that the adjunction formula follows from the
Riemann-Roch formula. However, the proof of Riemann-Roch in Falt-
ings depends on semistability (which Faltings assumes throughout), and
other considerations which make Riemann-Roch deeper than adjunction;
and Chinburg in describing this proof uses flat base change for cohomo-
logy, overlooking the need for desingularization when he does not make
this assumption. After I pointed this out, Chinburg has told me that he
can make a direct study of how base change affects the Riemann-Roch
formula even in the non-semistable case. However, here we shall use ad-
junction to prove the Riemann-Roch formula, and we do not use base
change.

There is still one point which we have not considered, namely whether
the metrics on the sheaves over X are admissible. The metric we have
given to (7)x(E) is admissible by definition, but there is a problem about
the metric on W x /y. This problem occurs on the generic fiber, so it's a
question of the metric on ab, where C is a complete non-singular curve
over the complex numbers. We gave ab the metric such that the map
[IV, §5] METRIZATIONS AND ARAKELOV'S ADJUNCTION FORMULA 101

obtained by pull back from the diagonal

is a metric isomorphism for each point Q. The next theorem is also due
to Arakelov.

Theorem 5.4 If A(P, Q) = !g(P, Q) where g is the canonical Green's


function associated with the canonical form IJ., then the metric on nb is
A-admissible.

Proof. In this case, the Chern form in Theorem 3.3 of Chapter II pulls
back under the diagonal to a scalar multiple of the canonical form IJ..
And we know from Theorem 3.1 of Chapter I that two metrics which
have the same Chern form are constant multiples of each other. This
concludes the proof.

When we take A =!g and g is the Green's function associated with


the canonical form IJ., then we say that a metric is IJ.-admissible if it is
admissible with respect to A in the previous sense. Such IJ.-admissibility
will become important in the next chapter.
If the metrics of the adjunction formula are admissible, then the for-
mula can be rephrased in terms of divisors. So suppose that A = !g
where g is the Green's function corresponding to the canonical form IJ..
If K x /y is an Arakelov divisor in the class corresponding to OJ x /y with
the prescribed metric, then we call K x /y a canonical Arakelov divisor. By
Theorem 5.3, we can then reformulate the adjunction formula as follows.

Corollary 5.5. Assume that A = !g where g is the Green's function corre-


sponding to the canonical form IJ.. Then for a horizontal curve E, we
have
(K x /y ' E) + (E. E) = dEfY + diE).
The two intersection symbols on the left should of course carry the sub-
script A, which we omit for simplicity.

Corollary 5.6. Under the same hypotheses, if E is a section then

(Kx/y.E) + (E.E) = O.
Proof. In this case, both dEfY and diE) are trivially equal to O.
CHAPTER V

The Faltings Riemann-Roch


Theorem

In this chapter we prove the Faltings Riemann-Roch theorem, assuming


the existence of certain volumes on the cohomology of a line sheaf on a
curve over the complex numbers. The next chapter will be devoted to
proving the existence of these volumes by analytic means. Also we shall
postpone to the next chapter another result of analysis needed as a
lemma to justify one of Faltings' applications of his theorem. Thus the
present chapter constitutes a natural sequel in the style of algebraic geo-
metry with metrized line sheaves, continuing the ideas of the adjunction
formula.
As before, we first deal with an arithmetic curve and a module over a
ring of integers in a number field. We then deal with line sheaves on the
arithmetic surface, induce them on the curve, and compare Euler charac-
teristics, which is what Riemann-Roch does. The Euler characteristics in
this case have a component at infinity, which introduces Haar measures
and volumes, and ultimately the geometry of numbers.

V, §1. RIEMANN-ROCH ON AN ARITHMETIC CURVE

Let A be a ring of algebraic integers, finite over Z, and let

Z = spec(A).

We call Z an arithmetic curve. On @z we have the trivial metric r(w)


for each WE S ",(F(A», as in Chapter IV, §3. Let (!£!, p) be a metrized
line sheaf on Z. Let !f = L -, so L is locally free. For each absolute
[V, §1] RIEMANN-ROCH ON AN ARITHMETIC CURVE 103

value w we then have the p(w)-norm on L w, over F(A)w. We can imbed

on the diagonal as usual in algebraic number theory, and n = r 1 + 2r 2 •


We also have an imbedding

and under this imbedding, L is a lattice in L.t. Now for the measures:

Let F(A)w have Lebesgue measure if w is real;


let F(A)w have twice Lebesgue measure if w is complex;
let Lw have the unique measure such that if e 1 is a basis for Lw over
F(A)w, and x is a coordinate with respect to this basis, then the
measure on Lw is

where dv(x) is the measure on F(A)w as above.

We call the measure on Lw the normalized Haar measul·e. We then get


the product measure on the product space. We define the normalized
volume to be the one computed with respect to this measure and this
volume will be denoted by Vol p . The volume of a fundamental domain
for L in LR with respect to this measure will be denoted by

Suppose that L is imbedded as a submodule of the function field


F(A). Then each pew) is a metric on F(A)w itself, and the metric induces
a Haar measure on each such factor, which is equal to

1IlIIp(w) times the normalized Haar measure on F(A)w'

so the Haar measure on F(A)R induced by p is equal to

TI I 1IIp(w) times the normalized Haar measure on TI F(A)w·


w

We define the Euler characteristic

xUf', p) = -log Volp(LRlL).


104 THE FALTINGS RIEMANN-ROCH THEOREM [V, §1]

Theorem 1.1. Let d A / Z = log D A/Z be the log of the absolute value of
the absolute discriminant. Then

Proof This is a classical computation of algebraic number theory.


See, for instance, [La 5], Lemma 2, Chapter V, §2. Taking twice
Lebesgue measure at the complex absolute values gets rid of the factor
2'2 in this context.

Theorem 1.2 (Riemann-Roch). Let (se, p) be a metrized line sheaf on


Z. Then
x(se, p) - x(@z, ,) = degz(se, p).

Proof First we may imbed se in the constant sheaf of the function


field F(A) so that L is a submodule of F(A), locally free of rank lover A.
Also we may multiply L by a sufficiently divisible integer so that LeA,
for convenience. Each pew) induces a metric on F(A), differing from the
trivial metric by a constant, and whose log satisfies

-loglllll p(w) = deg 1 , w(L, p),

Note that LR = AR : : : : R", Since LeA we have for the trivial metric

Vol/LRiL) = (A: L) Vol.(A.JA)


so that

10g(A: L) - log Vol/LRiL) = -log Vol/ARiA) = X(@z, c),

But 10g(A: L) = -degfin(L) from the definitions, We add and subtract


Lw loglllllp(w) to
the left-hand side, and use the relations

-log Vol.(L.JL) -l)oglllll p(w) = -log Volp(LRI'L),


L loglllllp(w) = -degoo(L, p)
w

to conclude the proof

The Euler characteristics as above are in some sense absolute, Sup-


pose now that Z is finite over Y = spec(R), where R is the ring of
integers of the number field F as usual Then we may define the relative
Euler characteristic
[V, §2] VOLUME EXACT SEQUENCES 105

Here we wrote ft'" instead of (ft', p) to simplify the notation. Then we


also have the relative versions of Theorems 1.1 and 1.2.

Theorem 1.1 (relative). For (l}z with the trivial metric, we have

Proof This comes from the standard discriminant formula of alge-


braic number theory

where rkR(A) = [F(A) : F] is the R-rank of A.

Theorem 1.2 (Riemann-Roch relative). Again for (l}z with the trivial
metric,

Note that the right-hand side is unchanged from the absolute version,
because the term rkR(A)d Rlz cancels on the left-hand side.
These relative versions will be applied in the Faltings Riemann-Roch
theorem.

V, §2. VOLUME EXACT SEQUENCES

Let V be a vector space of dimension n over a field k. We can extend


the determinant

to an exact sequence of vector spaces

E: 0 ..... VO ..... VI ............. vm ..... 0


by defining

The tensor product on the right is canonically isomorphic to k.


Suppose k = R. We define a volume on V to be a metric p on det V.
Such a volume determines a Haar measure on V. Indeed, let {el' ... , en}
be a basis of V. Then the Haar measure is defined to be
106 THE FALTlNGS RIEMANN-ROCH THEOREM [V, §2]

where Xl' '" ,X n are the coordinates with respect to e l , .•• ,en' If M is a
lattice in V, then such a Haar measure determines a measure for a
fundamental domain VIM, denoted by Vol(VIM), and also called the
volume of VIM. For example, if M = L Ze;, then

Note that a positive definite scalar product on V gives rise to a


metric on det V. If we change the absolute value on V by a positive real
number c, then the corresponding volume changes by CdimV • This remark
can be applied to C, say, viewed as a real vector space of dimension 2.
Let E be an exact sequence of vector spaces as above, and suppose
given a volume on each vector space. The metric on each det Vi gives
rise to a metric on R by taking the product with the alternating sign,
and the metric on det E is therefore determined by the norm of 1 in R.
We define the volume discrepancy

]ICE) = log 111-

We say that E is volume exact if )I(E) = O.


Let M be a finitely generated module over Z. For simplicity of
language, we shall omit the qualifier "finitely generated" since all models
under consideration will have this property unless otherwise specified.
Given a volume on M R , we define the Euler characteristic of M to be

X(M) = -log Vol(Ma/M) + 10glMtorl,

where IMtorl = car M tor is the cardinality of the torsion subgroup of M,


and MR is the tensor product of M with Rover Z.
Consider an exact sequence of modules:

and the corresponding exact sequence obtained after tensoring with R:

Theorem 2.1. Given volumes on each M~, we have

In particular, X is additive over volume exact sequences.


[V, §2J VOLUME EXACT SEQUENCES 107

Proof It suffices to prove the theorem for short exact sequences

O ..... M' ..... M ..... M" ..... O.


Let
M;or c M' cDc M,

where D is the division group of M' in M, that is, DIM' is torsion, and
MID is free. Let {e 1 , ••• ,er} be a basis of M'IM;or' Then:

M' = [e l , ... ,erJ EB torsion,


M = DEB F, where F is free,

The bracket notation means that, say, [e l , ... ,erJ is the free submodule
with basis {el' ... ,er }. Then

so

By definition, the left-hand side is

(Vol MR/M')(Vol M;j M").


Let

M rr •• = M'IM;or'
Then
el 1\ '" 1\ er = ±(Dfree : Mrre.)v l 1\ ... 1\ Vr'
Hence

lel 1\ ... 1\ erl'lW r + l 1\ ... 1\ wnl"

= eY(Dfree: Mrre.)lv l 1\ ••• 1\ Vr 1\ W"+l 1\ ••. 1\ wnl,

and therefore
Vol(MR/M') Vol(M;jM")
eY = -::-::--:-':-~~-=-c:-=---'-------:==--,-
Vol(MRiM)(Dfre • : Mrree) .

Thus to prove what we want, it suffices that

(D . M' ) = IM;orIIM;~rl
fr.e . free IM I '
tor
108 THE FALTINGS RIEMANN-ROCH THEOREM [V, §3]

But trivially,

and so
IM;orIIM~orl IM;orI (Dfree : M~r.e)(Dtor: M;or)
IMtorl IDtorl
= (Dfree : M~r.e),
as was to be shown.

V, §3. FALTINGS RIEMANN-ROCH


Throughout this section except for Lemma 3.1, we let R be the ring of
integers of a number field F. We let Y = Spec(R), and X - Y is a flat,
projective morphism, with X regular. Thus X F is a projective non-
singular curve over F. We assume that X F is absolutely irreducible, of
genus q ~ 1.

The first lemma is quite general, so we restate the hypotheses com-


pletely in its statement.

Lemma 3.1. Let Y = Spec(R) where R is a Dedekind ring, and let


X - Y be a flat projective morphism of relative dimension d. For any
coherent sheaf:iF on X, we have

HP(X, :iF) =0 for p ~ d + 1.


Proof After localizing R at its primes, we are reduced to the case
when R is a discrete valuation ring. There exist homogeneous polyno-
mials F 0, ... ,Fd with coefficients in R such that the intersection

X n (F 0 = 0) n ... n (F d = 0) is empty.

Then X is covered by the affine open subsets F j #- 0 for j = 0, ... ,d, and
so by the definition with Cech cohomology, we have HP(~) = 0 for
p ~ d + 1, as desired.

Next we observe the simple fact that if !£ is a line sheaf on X such


that deg!£F > 0 on X F, then Hl(X, !£®") is a torsion module over R for
n sufficiently large. This is trivial from the standard fact that
H 1 (X F , !£'F) = 0 for n large, and the cohomology commutes with base
change in the case of flat base change (F is flat over R). This remark
will be used later. For the moment, we consider the cohomology groups
[V, §3] FALTINGS RIEMANN-ROCH 109

HO and Hl, which are the only non-trivial ones on our arithmetic surface
X as at the beginning of the section.
By Lemma 3.1 we have HP(X,!F) = 0 for !F coherent on X and
p ~ 2. Also HO(X,!F) and Hl(X,!F) are finitely generated modules over
R, while HO(X, ft') is free if ft' is a line sheaf. Then HP(X, ft')
(mod torsion) is a lattice in HP(X, ft')R (tensor product over Z), and we
have

HP(ft') ®z R = HP(!l')R = EB HP(Xp, ft'v) = EB (HP(X, ft'» ® Fp.


veSco veSoo

We define

Thus we also have for the determinants

det(Hft')"" = @ det(Hft')p.
veSoo

We shall apply §2 to the modules HO(X,!l') and Hl(X, !l'), which are
finite over R. For this we need volumes on MR where M is one of these
cohomology modules. Indeed, if M is a module finite over R, then we
can form
MR=M®zR= n M®Fp.
veSoo

Then M/ M tor is a lattice in MR' Each M v = M ® F p is a vector space


over Fp. For certain line sheaves ft' we have a natural volume on these
vector spaces as follows. We work at a single v, and we shall carry out
the discussion with a complete non-singular curve C over the complex
numbers.
Given a complex vector space with a hermitian metric there is an in-
duced volume obtained by regarding V as a vector space over the reals
and taking the determinant.
At the real places, the cohomology groups are real cohomology
groups, and the situation is similar, except that certain squares are miss-
ing.
Consider first n~, and HO( C, n~) which is a vector space of dimension
q (= genus of C) over C. As in Chapter II, §2 we have a canonical
hermitian product on this vector space, given by
110 THE FALTINGS RIEMANN-ROCH THEOREM [V, §3]

This hermitian product gives rise to a metric on HO(C, O~), called the
canonical metric. The corresponding metric on det HO( C, O~) is also
called the canonical metric, with corresponding canonical volume on HO.
We give HO(C, (DC> = C the trivial metric, such that 11111 = 1.
Since H 1 (C, (DC> is dual to HO(C, O~) we give it the dual volume,
which is compatible with C receiving the trivial metric.
Taking C = X" for each v E S co' we have then given a metric on
det H«(DC>, called the canonical metric on det H«(DC>.
We then have the following fundamental theorem of Faltings [Fa 1].
We follow Faltings' notation and for any point P we let 9'[P] denote
the sheaf on X which is 0 outside P and 9'IP at P. We may still
identify the two, but for some purposes we find the different notation
makes matters clearer.

Theorem 3.2. Let C be a complete non-singular curve over the complex


numbers. For every line sheaf 9' p with a metric p which is fl-admissible,
there exists a unique choice of hermitian metric on det H(9'p), called the
Faltings metric on det H(9'p), satisfying the following conditions:

F AL 1. If 9'p = (Dc with the trivial metric, then the F altings metric is
the canonical metric on det H(l!JC>.

F AL 2. An isometry 9'p --+ 9'~, induces an isometry

FAL 3. For each metrized 9'p and every point P, the short exact
sequence

o --+ 9'( - P) --+ 9' --+ 9'[P] --+ 0

induces an isometry

FAL 4. If cp: 2'p --+ !L'p' is a sheaf automorphism which multiplies the
norm by a constant c > 0, then the norm on det H(2'p) is mul-
tiplied by

where hP = dim HP( C, 2').

The proof of this theorem will be given in the next chapter. Here we
only draw consequences of the theorem.
[V, §3] FALTINGS RIEMANN~ROCH 111

All further references to metrics on det H(.!l') or volumes on H(.!l') will


be to the Faltings metrics or volumes of the above theorl~m, unless other-
wise specified.

We note that even if, say, HO(.!l') = Hl(.!l') = 0, the Fa.ltings metric on
det H(.!l'p) = C is not usually the trivial metric such that //1" = 1. We
shall meet later an important case when it is essential to have more in-
formation on such a metric.
Before going to the Faltings Riemann-Roch theorem, we state a
lemma which gives a useful consequence of FAL 3. In this lemma, we
need again the discriminant at infinity, so let g be the Green's function
of the canonical form fJ.. Let P l' ... ,p. be distinct points of C in C(C),
and E = L (P). We let the A.-discriminant be

DiE) = exp diE) where diE) = L A.(Pi , Pi) as before,


i*i

and A. = tg.
As usual, if .!l' is a line sheaf, we let


.!l'p(E) = .!l'pQ9 (J)C<P i );..
i= 1

Lemma 3.3. Let C be a complete non-singular curve ODer the complex


numbers. Let.!l'p be a line sheaf on C with a fJ.-admissible metric. Let
L
E = (Pi) as above. Then the sheaf exact sequence

o-+- .!l'( - E) -+- .!l' -+- EB .!l'[P;] -+- o.


induces a metric isomorphism


det H(.!l'p)'=' det H(.!l'p( -E» ® Q9 .!l'p[P;]. D;.(E)1/2.
i= 1

More precisely, the module-exact cohomology sequence II:

o-+- HO(.!l'i - E» ~ HO(.!l'p) ~ HO<EB .!l'p[P;])

~ Hl(.!l'i - E» ~ Hl{.!l'p) ~ 0 ~ O.

has the real volume discrepancy y(H) = -diE), in the notation of §2.
The volumes are of course the F altings volumes.
112 THE FALTINGS RIEMANN-ROCH THEOREM [V, §3]

Proof From F AL 3 and the exact sequence

o-+!t'( -PI - ... - P r) -+!t'( -PI - ... - P r - 1 )-+


!t'(-P 1 - .. ·-Pr - I )[Pr ]-+O
we get a metric isomorphism

det H!t'p( -Pl - ... - Pr - 1)

~ det H!t'i -PI _ ... - Pr)®!t'i -PI _ ... - Pr-l)[Pr].

For the last term on the right we have the metric isomorphism

and the metric on (9 c( - PI - ... - P r _ 1>..1 P r as in Chapter I, Proposition


2.2, is such that
-logI1(Pr )1 = L. - A(P
i<r
i, Pr)'

The rest of the proof is then concluded by induction.

Remark. Note that the sum over the indices i < j is precisely half the
sum over all indices i # j. This explains the exponent 1/2 in the first for-
mula, expressed in terms of the metric. In the second formula expressed
in terms of the volume discrepancy, there is an extra factor 2 coming
from the fact that the volume of C (as vector space over R) changes by
a square as explained in the preceding section.

Next we define the Euler characteristic of a metrized sheaf.


Let M be a module finite over R. Then M is also finite over Z. Sup-
pose M has a volume. Recall the absolute Euler characteristic

X(M, Z) = -log Vol(MalM) + 10glMtorl.


We define the relative Euler characteristic

X(M, R) = X(M, Z) - rankR(M)x(R, Z),

where R has the trivial metric, so the normalized volume as in §l. By


Theorem 2.1, both Euler characteristics are additive on volume exact
sequences.
Let !l'p be a Jl-admissible metrized line sheaf on X. We define the
relative Euler characteristic of this sheaf to be
[V, §3] FALTINGS RIEMANN-ROCH 113

The volumes on HOCK, 2'p) and Hl(X, 2'p) are the Faltings volumes. It
is sometimes practical to omit the index and write

and we thus also omit the R for simplicity, but it is the relative Euler
characteristics which are relevant in what follows.

Theorem 3.4 (FaItings Riemann-Roeh). Let the Neron function be A. =


tg where g is the Green's function belonging to the canonical form Jl at
each v E S<X/' Let D be an Arakelov divisor on X. Let K xly be an Arak-
elov divisor in the class corresponding to the metrized canonical sheaf
OJ XIY ' Then taking the intersections with respect to A., we have

Proof The proof is by an inductive procedure. We show that the


theorem is true for an Arakelov divisor D if and only jf it is true for
D + C, where C is an irreducible divisor, that is:
C = E is a horizontal curve on X;
C = Xv for some v;
C = irreducible vertical divisor, i.e. C is a vertical curve on X.
Case I
The first case is the most interesting, so suppose C = E is a horizontal
curve. We shall index the Euler characteristics by the ambient sheaf if
necessary, so we have both XXIV as defined before the theorem, and XEIY
as defined in §l. We can work formally just as in the case of surfaces
over the complex numbers. All the work has been dont: to justify the
formal use of intersection theory. Namely, using the relative Euler char-
acteristics on E, we have:

Lemma 3.5.
XEly(lDx(D) IE) = (D. E) - t[(E .E) + (Kxly.E) - d;.(E)].

Proof We have

by Riemann-Roch for curves


relative Theorem 1.2
= (D. E) - td Ely by relative Theorem 1.1
= (D. E) - H(E. E) + (K xly ' E) - d;.(E)]
by adjunction formula.
114 THE FALTINGS RIEMANN-ROCH THEOREM [V, §3]

Now we prove that Riemann-Roch for D implies Riemann-Roch for


D + E. We use the fact that X is additive on volume exact sequence, and
the discrepancy is now given by Lemma 3.3. Subtracting X«(!)x) from
both sides, we get:

X(@x(D + E» - X«(!)x) = X«(!)x(D» - X«(!)x) + X«(!)x(D + E)IE) - tdl(E)


= H(D.D) - (D. K xly )] + «D + E).E)
(by 3.5)
- t[(E . E) + (E. K XIY) - d;,(E)] - td,,(E)
= t[«D + E).(D + E» - «D + E). K xly )]

as was to be shown.
In fact the steps are reversible, so Riemann-Roch for D is equivalent
to Riemann-Roch for D + E if E is a horizontal curve on X.

Case 2
Next suppose that C = X v is a fiber at infinity. Then the right-hand side
changes by N v times

where q is the genus of the generic fiber. As for the left-hand side, the
metric on det H@x(D + Xv) differs from the metric on det H(!)x(D)
multiplicatively by

according to F AL 4. Hence directly from the definitions, the difference


between the Euler characteristics on the left-hand side is simply N v times

The ranks hO and hI are the same as the dimension of the cohomology
groups on the generic fiber, and hence Riemann-Roch for D is equivalent
to Riemann-Roch for D + X v if and only if

which is true by the basic algebraic geometry of non-singular curves.

Case 3
Finally, for the third case, the argument is similar except that we need
the same formalism for a possibly singular curve on the regular arithme-
tic surface. This formalism is essentially classical, except for the need to
[V, §3] FALTlNGS RIEMANN-ROCH 115

deal with singular curves on a surface, so naive versions of Riemann-


Roch on non-singular curves do not suffice. One needs the more general
theory of the dualizing sheaf and Grothendieck duality in special cases as
in Altman-Kleiman [A-K] or Hartshorne [Ha 1], Chapter III, Theorems
7.6 and 7.11. We recall these results.

First, let C be a proper curve over a field k, so C is an integral scheme


of dimension 1, proper over k. We also suppose that C is a local com-
plete intersection in some smooth scheme over k.

These assumptions are satisfied if C is a fibral curve on a regular arith-


metic surface and k = key), y E Y (closed point), since the surface itself is
regular, and is a local complete intersection in a smooth scheme S over
Y. Then C is a local complete intersection in the smooth fiber Sy by
[EGA], Chapter IV, 9.1.5.1(v) (see also [Mat 2], Theorem 21.2), because
both Sand Sy are regular, and C is regularly imbedded in S (via X).
For Cover k as above, we have a canonical sheaf

W C/k = WC/spec(k)

as in the general definition we gave in Chapter IV, §4. From the general
property of the canonical sheaf CS 2 we get:

Theorem 3.6 (Adjunction formula). Let X be a regular arithmetic sur-


face over Y= spec(R). Let C be a fibral curve in X y. Then

Let !l' be a line sheaf on C. Then we have the Euler characteristic

where hf(!l') = dimk HP( !l').

Of course, all cohomology groups are 0 for p ~ 2. We have the classical


analogues of the relative Theorems 1.1 and 1.2, namely:

Theorem 3.7 (Riemann-Roch on C). Let !l' be a line sheaf on C.


Then:
(1) XiC9d = -! degkW c/k;
(2) Xk(!l') - XiC9d = degk(!l')·

Proof The curve C is Cohen-Macaulay because it is a locally com-


plete intersection. Formula (2) is explicitly in [A-K], Chapter VIII,
Theorem 1.4, and [Ha 1], Chapter IV, Exercise 1.9, with their definition
of the dualizing sheaf. With that definition, (1) comes from the duality
116 THE FALTINGS RIEMANN-ROCH THEOREM [V, §3J

of [Ha 1J, Chapter III, Theorem 7.6. Here is the precise point where co-
homological duality becomes relevant. To identify the dualizing sheaf
from cohomology with our canonical sheaf, we can use [Ha IJ, Chapter
III, Theorem 7.11 and Corollary 7.12; see also Chapter II, Proposition
8.20. Note that the proof of Theorem 7.11 does not depend on projec-
tive space, only on the fact that the scheme is regularly imbedded in a
smooth scheme over k. Of course, we could also use all of projective
space if we accept the projective imbedding of X in P~ for some N as in
[LiJ.
Note that deg k W e /k is a classical, geometric analogue of the discrimin-
ant dEfY in Case 1. If k is a finite field, then we can define the more ex-
act analogue

where Ik I is the cardinality of k. Similarly,

We can now return to the proof of Case 3 proper, when C is a fibral


curve on the arithmetic surface X, and R is the ring of integers of the
number field F. We let Ke be a Cartier divisor on C such that We ~
(ge(Kc). Then for any Arakelov divisor D on X, the restriction (9x(D)IC
depends only on the finite part of the divisor, and we get:

Lemma 3.8.

X«(9x(D) IC») = (D. C) - H(C. C) + (K x /y • C)].


Proof We have

X«(9x(D)IC) = (D.C) + X«(9c) by Riemann-Roch on C


Theorem 3.7(2)
= (D. C) - ! deg We by Theorem 3.7(1)
= (D. C) - H(C.C) + (Kx/y·C)J by adjunction formula.

Instead of the exact sequence of Lemma 3.3 we now deal with the ex-
act sequence
o . . . .P( - C) ...... .P ...... .P IC ...... o.

The inclusion .P( - C) ...... .P gives rise to a metric isomorphism on the


generic fiber since .P I C is fibral, so in the present case, there is no dis-
crepancy in the metric isomorphisms on the cohomology. Then the sec-
ond part of the proof of Case 1 can be repeated verbatim, except for
[V, §3] FALTINGS RIEMANN-ROCH 117

omitting completely the term d;.(E). This concludes the proof of Faltings
Riemann-Roch.

Remark 1. Observe that we have used the classical formalism of inter-


sections on surfaces. In the present case, for fibral divisors on an arith-
metic surface, this formalism works without worrying about metrics or
number theory. All that is needed is the regularity of the: surface. Then,
for instance, if D is a fibral divisor on X, we can define: the arithmetic
genus by
p(D) - 1 = t[(D2) + (K . D)], where K = K x /y.

If D is effective, then P(D) = 1 - X«(!)D) by Lemma 3.8. In general, the


arithmetic genus satisfies the classical formalism:

p( -D) = (D2) - p(D) + 2,


p(C + D) = p(C) + p(D) + (C.D) - 1.

These formulas follow formally by arithmetic from the definition. In


these formulas, the intersection numbers are to be taken as the "geomet-
ric" ones, i.e. they are integers. We leave out the log of the order of the
residue class field. We shall use these formulas in the computations of §5
below.

Remark 2. It was part of intersection theory that the intersection


numbers behave well under extension of the base and the total inverse
images. One can ask whether the Euler characteristic behaves the same
way.

Lemma 3.9. Let M be a module finite over R. Let R' be the integral
closure of R in a finite extension F' of F, and let

M'=M®RR'.
Then
X(M', R') = [F' : F]X(M, R).

Proof. Immediate from the definitions.

Now suppose X' is a desingularization of X x y Y' where Y' =


spec(R'). Let ft'p be a metrized line sheaf on X, with a J.l-admissible
metric p. Let t/I: X' -+ X be the projection, which is generically finite. Is
it true that
118 THE FALTINGS RIEMANN-ROCH THEOREM [V, §4]

There is no problem about the volumes, which are computed on the gen-
eric fiber, but the torsion in HI presents a problem, and the answer is
NO in general. Faltings asserts the behavior as true in [Fa 1], but he
explicitly assumes that the arithmetic surface is semistable, from the very
beginning (e.g. p. 389). On the other hand, Chinburg [Ch 1], Remark
4.1, and Szpiro [Sz 1] do not mention semistability assumptions, and
thus their formulations are incorrect.
In §5, we shall define semistability formally, and we shall state the re-
levant theorem. Granting this, the reduction of the proof of Riemann-
Roch to the case when E is a section works without further ado in the
semis table case. We avoided base change in the proof of Riemann-Roch,
so we did not encounter the difficulties.

Remark 3. As shown by Faltings, the Riemann-Roch theorem has a


formulation as a metric isomorphism between certain sheaves. In addi-
tion, Deligne defines a pairing

Uf', Jt)f-+ (fe, Jt)

from metrized lines sheaves on X to a metrized line sheaf on the base


Y = spec(R). The Oeligne Riemann- Roch then asserts a metric
isomorphism

where det is the determinant, which will be recalled in Chapter VI, §2.
Taking degrees yields the Faltings Riemann-Roch as we have stated it.
In Falting's versions, the metrics are admissible, but in Deligne's version,
me tries need not be admissible. A proof can be given following the same
pattern as in this book, but working at the level of sheaves all the time.
For a version in the admissible case, see [MB], who always makes the
semistability assumption. Formula 6.13.2 in [MB] should have been
stated as the theorem, with the Faltings Riemann-Roch numerical formu-
la 6.13 as a corollary.

V, §4. AN APPLICATION OF RIEMANN-ROCH

Let D be an Arakelov divisor on X. We say that D is effective and write


D ~ 0 if all coefficients of D are ~ 0, including the coefficients r(v) of Xv
for v E S 00. The next theorem of Faltings will depend on an analytic esti-
mate, which will be used here, but whose proof is postponed to the next
chapter, where we consider the analytic side as a whole.
[V, §4] AN APPLICATION OF RIEMANN-ROCH 119

Theorem 4.1. Let D be an Arakelov divisor on X such: that (D2) > 0


and degF(D) > O. Then there exists no such that for n ~ no there exists
an effective Arakelov divisor E for which

nD ""' E.

Proof The problem here is at infinity. The theorem says that some
positive multiple of D is linearly equivalent to an effective divisor, even
at the infinite places. Suppose we have a function f on X such that

(f)Ar = E - nD,

and Erin ~ 0, so f E HO«(!Jx(nD fin ». We want, III addition, the positivity


condition

-f Xv
log Ilfllv.uv + nrv(D) ~ 0

or equivalently

flOg IIfllv.uv ~ nrJD).


By the convexity property of the integral
total measure 1, it suffices that
f log ~ log f on a space of

We want a non-zero lattice point f E HO(X, (!J(nD rin » in Bn' By Min-


kowski's theorem, it suffices that

Vol(B.} ~ 2 hO Vol(fundamental domain for HO(X, (!J(nD fin »),

or equivalently, taking logs and letting M = HO(X, (!J(nD fin ),

log Vol(B.} ~ hO(nDfin)log 2 + log Vol(M.JM).


120 THE FALTlNGS RIEMANN-ROCH THEOREM [V, §5]

We have hO(nDfin ) = O(n) by Riemann-Roch on the generic fiber. We re-


call that Hl(X, (9(nD fin ») is a torsion group for n large. From the defini-
tions, it follows that

log Vol(M.JM) = - x«(9(nD), R) -log IH 1 «(9(nD fin »torl


- rkRHo(nDcin)x(R, Z).

Hence it suffices to prove the inequality

log Vol(Bn) ~ O(n) - X«(9(nD), R)

= - !n 2 (D2) + O(n),
where this last inequality is by Faltings Riemann-Roch. It will be
proved in Theorem 4.1 of Chapter VI that given s, for n sufficiently large
we have

which is certainly sufficient to complete the proof.

Remark. For the convenience of the reader, we recall a proof that on


a measured space X with positive measure tt, total measure 1, and with a
function q> ~ 0, we have

flOg q> dtt ~ log f q> dtt·

Let c = Jq> dtt· If c = 0 then both sides are - 00. Assume c > O. We use
the inequality
log(l + h) ~ h for h > -1.

We put h = cplc - 1, and integrate

log q> = 10g( 1+ ~ - 1) + log c


to get what we want.

V, §5. SEMISTABILITY

Throughout this section we let X be a regular arithmetic surface, and


we assume that the residue class fields of R are perfect.
[V, §5] SEMISTABILITY 121

The purpose of this section is to develop the basic notions of semista-


bility. We say that X is semistable over Y, or semistable, if the following
conditions are satisfied.

SS 1. The generic X F is geometrically irreducible.

From Zariski's theorem, it follows that the fibers are geometrically


connected ([Ha 1], Chapter III, Corollary 11.5). Before stating the
next condition, we make the definition that the geometric fiber above a
closed point y E Y, is the base extension of the fiber to the algebraic
closure of k(y). Then an irreducible component of the: fiber Xy splits
into a finite number of conjugate geometric components in the geometric
fiber.

SS 2. The geometric fiber above a closed point y E Y is reduced and


has only ordinary double points.
SS 3. If C is a non-singular irreducible component of a geometric fiber,
and C has genus 0, then C meets the other components of the
geometric fiber in at least two points.

It follows from these three conditions that:

SS 4. A semistable X --4 Y is minimal.

This last property follows from the intersection properties SS 2 and SS 3


and Castelnuovo's criterion, which is carefully done in the case of non-
algebraically closed residue class field by Shafarevich [Sh], p. 102. In
particular, a semis table model does not contain any geometric fibral
component C with (C 2 ) = -1.
Since the residue class fields of R are assumed perfect, if we first make
a base change

where y is a closed point of Y, and @y is the completion of my, then fin-


ite unramified extensions of spec(@y) are in bijection with finite exten-
sions of the residue class fields. Base changing the arilthmetic surface
from spec(@y) to finite unramified extensions preserves the regularity of
the arithmetic surface.

Remark. Grothendieck defined the notion of semistability for abelian


varieties, and he proved that given an abelian variety, say over a number
field, there always exists a finite extension over which the abelian variety
is semistable. Using Grothendieck's theorem, Deligne-Mumford [D-M]
proved the corresponding theorem for curves. In other words, given a
122 THE FALTINGS RIEMANN-ROCH THEOREM [V, §5]

curve over a number field, there exists a finite extension F and a regular
arithmetic surface X over Y = spec(R) which is semistable. See also
Artin-Winters [A-W]. In what follows, we avoid the use of such theor-
ems by simply assuming that we are given a semistable model. However,
for applications one must of course use the theorem. One obtains a
bound for the degree and discriminant of the finite extension needed to
make a curve semis table by using some algebraic point to imbed the
curve in its Jacobian, and then adjoining the torsion points of suitably
large order to make the models semistable.

We shall now recall some facts about semi stability and base change.
We begin with two results of Artin.

Theorem 5.1. Let X ~ Y be a semistable regular arithmetic surface. Let


Y' = spec(R') where R' is the integral closure of R in a finite extension
F' of F. Then:
(a) The base change X Y' is normal.
(b) There exists a minimal desingularization X' of X Y"
(c) In this minimal resolution, the double points are resolved by geomet-
ric components which are curves isomorphic to pi, with self-intersec-
tion -2.

Proof Essentially a complete proof is given by Artin [Art 3], pp.


116-119. Since these notes were never published and are not now avail-
able, I shall copy those pages for the convenience of the reader.
For simplicity, we assume that R is a discrete valuation ring with
algebraically closed residue class field. To deal with finite residue class
field, one would have to superimpose the Castelnuovo criterion as in
Shafarevich [Sh], p. 102, taking the finiteness into account.
It is easy to see that the only non-regular points of X Y' lie above the
nodes of X (by node we mean ordinary double point, a standard termin-
ology). Above the nodes, X Y' is a locally complete intersection (defined
by one equation !), and hence Cohen-Macauley, so X Y' is normal by
Serre's criterion (see [Mat 1], Theorem 39, p. 125, following [EGA], IV,
2, p. 108), because no component of a fiber is multiple. Since X Y' is
Cohen - Macauley and generically smooth, the closed fiber X y' of X Y' is
reduced. The map X y' ~ X y is an isomorphism, trivially because we
assumed the residue class field algebraically closed.
Let X' ~ X Y' be a minimal resolution. Let X~, be the closed fiber, and
let C 1"" ,Cs be the connected components of the exceptional locus for
the map X' ~ X Y" Let C ij U = 1, ... ,r i ) be the irreducible components of
C i • Let X~, be the proper transform of X y ' in X'. Then

X~, = X~, + L aijCij with integers aij > O.


i,i
[V, §5] SEMIST ABILITY 123

It will now suffice to prove:


All integers aij are equal to 1, and the cycles C i consist of a configura-
tion as shown on the figure:

All intersections are transveral, Ctj is non-singular, and the self-inter-


sections are
(ct) = -2.
(We let the intersection numbers be the geometric ones, i.e. integers,
for this statement.)

The proof is by calculation. We shall use the formalism of the arith-


metic genus recalled after Lemma 3.8. We prove first:

Suppose that Z is an effective cycle> 0 on the .fiber X~, and


X~, - Z > O. If Z contains no exceptional curve of first kind, then

p(X~, - Z) < p(X~,).

Proof. Since X~, is orthogonal to all fibral components, we have

p(X~, - Z) = p(X~,) + t(Z2) - (K. Z».

By the negative definiteness of fibral intersections, modulo the total


fiber, and the fact that X~, has components with multiplicity 1, we get
(Z2) < 0, and (K. Z) ~ 0 because Z contains no exceptional curve. This
proves the claim.

Since X' was assumed minimal, it follows that

p(X~,) ~ p(x~, + ~ Cij),


"]

with equality if and only if all a ij = 1. From Remark 1 following Lemma


3.8, we get:

p(x~, + ~ C ii ) = p(X~,) + ~ p(C ij ) + ~ (X~" Cij)


',J J,) l.]
124 THE FALTINGS RIEMANN-ROCH THEOREM [V, §5]

Here:
p(C ti ) ;;;; 0,

L (X~,. C t) ;;;; 2s since C i lies above a node of X Y',


i,i

L (C ii · Cik) ;;;; r i - 1 since Ci is connected.


i<k

Moreover, p(X~,) ;;;; p(X y') - s, since separation of a node reduces p by 1


(s is at most equal to the number of nodes). So combining,
s s
p(X~,);;;; p(Xy') - s + 2s + L (ri - 1) - L ri = p(Xy,).
i= I i= I

But by the invariance of the Euler characteristic under specialization


([Ha 2], Chapter III, Corollary 9.10), if XF' is the generic fiber of X y',
then

Since XF' is also the generic fiber of X', we have p(X~.) = p(X y.). There-
fore all the above inequalities are equalities. Since p(C ti) = 0, we must
have (C~) < -1 by Castelnuovo's criterion. Since the total fiber is or-
thogonal to all fibral components, it follows that

But by the intersection numbers appearing in (.), the average values of


the numbers «X~, - C ii). Ci ) for i, j varying is 2. Hence

for all i, j,

and therefore (C~) = - 2. It is now easily checked that the configuration


of the figure is the only possible one. This concludes the proof.

As the reader sees, the proof is computational intersection theory on


the fibers.

The next theorems have to do with HI. We recall that for any
morphism of Noetherian schemes

f: S' --+ S,

the derived functor R1.(§i) on a coherent sheaf §i has the value

for every affine open set U of S.


[V, §5] SEMIST ABILITY 125

Theorem 5.2. Let X Y' be a finite base change of a semistable regular


arithmetic surface as in Theorem 5.1, and let f: X' ---. Xl"' be the minimal
desingularization. Then R 1f*(!Jx' = O.

Proof This theorem belongs to the foundations of the theory, and in


this case, we have a convenient reference in [Art 2], Proposition 1 and
Theorem 3, which give a criterion for a singularity to satisfy the condi-
tion R1f*(!Jr = O. In our notation, let

where Cij ranges over the resolving curves of a minimal desingulariza-


tion. By those references, it suffices to prove that p(Z) =, O. This comes
from the same sort of computation as in the proof of Theorem 5.1, using
the standard formulas for the arithmetic genus recalled after Lemma 3.8.
We can leave the arguments to the reader.

Let (!J x be a 2-dimensional local ring, which is normal and has alge-
braically closed residue class field. In [Art 2], Artin defines V = spec( (!J x)
to have a rational singularity if for any birational proper morphism

with regular V' (i.e. any desingularization of V) we have R If* (!Jy' = 0,


that is, Hl(V', (!Jy') = O. This condition is independent of the resolution f
because of the theorem on factorization of a birational proper morphism
of regular 2-dimensional schemes into blow ups [Sh] and [Li]. This can
be seen by following the proof given in [Ha 1], Chapter V, Proposition
3.4, which, although it is given over a ground field, actually applies to
the case over a discrete valuation ring, i,e. the arithmetic case. Theorem
5.2 can be stated by saying that the base change has only rational
singularities.
We state the next theorem for the record, because it has been used by
others implicitly in the base change method of proof for Riemann-Roch.
We shall make no use of it.

Proposition 5.3. Let X' ---. X Y' be the minimal desingularization of the
base change as above. Let cp: X' ---. X be the corresponding generically
finite morphism. Let 2 be a line sheaf on X. Then:

If R is the ring of integers of a number field and p is an admissible


metric on 2, then:
126 THE FAL TINGS RIEMANN~ROCH THEOREM [V, §5]

Proof The second statement follows at once from the first because
the Euler characteristic is defined partly in terms of a volume on the
generic fiber, in which case there is no difficulty in showing that this vol-
ume changes by the degree of the base extension; and partly in terms of
the torsion in HI, for which we require precisely the first statement for
the torsion group. As to this, we first note that since cohomology com-
mutes with flat base change ([Ha 1], Chapter III, Proposition 9.3) we are
reduced to proving the assertion for a desingularization when the base is
normal, with rational singularities, by Theorems 5.1 and 5.2. Then Lip-
man has pointed out to me that the proof is quite easy, namely: Let fey,
be the pull back of iE to X y', and let f: X' -> X y' be any desingulariza-
tion. Then we have a natural map

which is an isomorphism because X y' is normal. (The question is local


on the base, so it suffices to treat the case when fey. is the structure
sheaf, in which case the assertion is clear because f is proper.) But then
by Theorem 5.2 and elementary cohomology as in [Ha 1], Chapter III,
Exercise 8.1, directly from the Cech complex we have an isomorphism

This concludes the proof.

Next we deal with the canonical sheaf. When all the fibers of a regu-
lar arithmetic surface are generically smooth, as in the semistable case,
we can describe this sheaf as follows, following Arakelov. Let U be the
open set of smooth points over Y. Then

is the ordinary sheaf of 1-forms, and W UIY is a line sheaf on U. There


exists an extension of W UIY to a line sheaf on X. Indeed, since X is inte-
gral, we have

for some Cartier divisor D on U, and we can let the extension be ex(D)
where D is the Zariski closure of D in X. Such an extension is unique
by the following lemma.

Lemma 5.4. Let iE, ./1{ be two line sheaves on a normal variety. Let U
be the complement of a closed subset which has codimension ~ 2. Let
[V, §5] SEMISTABILITY 127

be an isomorphism. Then cp has a unique extension to an isomorphism


of ff with JI.

Proof The assertion is local, so after shrinking V in the neighborhood


of a point to a suitable open subset, we may assume U = V - W where
W is closed of codimension ~ 2, and ffy = JI y = (!Iy. An isomorphism
ffu --+ JI u is given by multiplication with some rational function. But a
rational function on a normal variety which is not morphic at a point
must have a divisorial pole passing through that point. By hypothesis,
this cannot happen, so the rational function is also morphic over W, so
over V. Hence the isomorphism of ffu to JI u extends to an isomorphism
of ffy to JI y, by multiplication with that same rational function, as de-
sired.

Thus under the assumption that there are only a finite number of
non-smooth points in the fibers, we can define the sheaf OJ XIY to be the
unique extension of ahlY to X, where U is the complement of the
singular points on the fibers.

The next result will be used in the proof of Theorem 6.1.

Proposition 5.5. Let X/Y be a semistable regular arithmetic surface, and


let X' /Y' be the minimal desingularization over a finite base change as
in Proposition 5.1. Let cp: X' --+ X be the natural projection. Then

Proof Since cp*OJ XIY and OJX'IY' agree on the generic :fiber, they differ
by fibral components, and

where D is a divisor supported in the union of the exceptional curves of


cp. To show D = 0, since the intersection form on exceptional fibral com-
ponents is negative definite, it suffices to prove that (D. C) = 0 for all fi-
bral components C such that cp(C) is a point. Since cp*OJ XIY has trivial
intersection with such fibral components, it suffices to prove that
OJX'(Y' IC is trivial. By the adjunction formula for each fibral component
in the minimal desingularization of a singularity on X Y' we have

because C ~ pI and the canonical class on pI is the class of degree - 2.


Hence OJX'IY'IC is trivial. This concludes the proof.
128 THE FALTINGS RIEMANN-ROCH THEOREM [V, §6]

The above results give an indication of necessary material that a basic


text in algebraic geometry must contain concerning semistable models.
Given the present status of the canonical sheaf and semistability, I found
it impracticable to include a sub book containing a complete, systematic
treatment of all the propositions we need to make Faltings' theorems go
through. By listing these propositions explicitly, I hope that I shall have
helped the writing of the appropriate text in algebraic geometry over a
Dedekind ring (or over a regular ring) when the time is ripe for it. At
least, I have tried not to slur over what is needed.

V, §6. POSITIVITY OF THE CANONICAL SHEAF

As far as I know, the next two theorems, due to Faltings, require semi-
stability for their validity.

Theorem 6.1. Assume X/Y semistable and q = genus of X F ~ 2. Let D


be an effective Arakelov divisor on X. Then:
(a) (Ki;y)degF(D) ~ 2q(2q - 2)(K x1y .D);
(b) (Ki/y) ~ o.
Proof We abbreviate K x /y = K. We first prove (a). We distinguish
cases, and the case when D is horizontal irreducible is the most interest-
ing. By Propositions 5.1 and 5.5, we can reduce the assertion to the case
when D is a section. By adjunction, we have for such D:

(D2) + (D.K) = 0.
Furthermore, (2q - 2)D - K has F -degree 0, so by the Hodge Index
Theorem,
«2q - 2)D - K)2 ~ 0,
whence
(K2) ~ 2(2q - 2)(D. K) + (2q - 2)2(D . K)
~ 2q(2q - 2)(D. K),

as desired. (In the function field case, this inequality is due to Szpiro.)

Suppose next that D = C is fibral irreducible in the fiber above y.


Then degF(C) = 0, so it suffices to show that (K. C) ~ 0. But if Pa(C) de-
notes the arithmetic genus of C, then Riemann-Roch and adjunction on
C yield the special case of Lemma 3.8, namely

2(Pa(C) - 1) 10glk(y)1 = (C 2 ) + (C. K),


[V, §6] POSITIVITY OF THE CANONICAL SHEAF 129

where x«(9d = 1 - Pa(C), We have (C 2 ) ~ 0 by Proposition 3.5 of


Chapter III. Suppose (K. C) < O. Then PaCe) = O. If (C 2 ) = 0, then by
the same Proposition 3.5 it follows that C is a rational multiple of the
fiber. In this case, in the divisor class of K we can find a divisor whose
horizontal part is effective (of F-degree 2q - 2), and hence has intersec-
tion ~ 0 with C. The vertical part is orthogonal to C, contradicting
(K. C) < O. Hence (C 2 ) < O. From the formula

(C 2 ) + (C. K) = -2 10glk(y)l,
it follows that iiC. C) = iy(C. K) = -1 and then by Castelnuovo's criter-
ion [Sh], p. 102, C must be contractible, contradicting SS 4, that X is a
minimal model. This concludes the proof of (a).
Next we prove (b). Let v be any place at infinity. Th~:re exists a real
number r such that

Indeed, we need (K2) + 2r(K. Xv) > 0, and on the generic fiber we have

so we can find r > 0 satisfying the desired condition. Then by Theorem


4.1 we have for some n > 0

n(K + rXv) '" E with E ~ 0 (Arakelov),

whence by the inequality of (a) which is linear in D, we get

(K2) 1 2
(K. (K + rXv» ~ 2 2 2 degF(K + rXv) = -2 (K).
q( q - ) q

Let r -+ ro + , where ro is the number such that

Then

whence
1 2 1 2
z(K ) = -ro(K.Xv) ~ 2q (K ),

which implies (K2) ~ 0, and concludes the proof of (b), as well as the
theorem.
130 THE FALTINGS RIEMANN-ROCH THEOREM [V, §6]

On the other hand, when the generic fiber has genus 1, we have the
following theorem.

Theorem 6.2. Let X! Y be semistable, with generic jiber of genus l.


Then

and in fact (K x /y ' C) = °


for every jibral curve C on X.

Proof The restriction of OJ x /y to the generic fiber is trivial, and hence


KX/y,fin consists only of fibral components. Thus to compute the self-in-
tersection (Ki/y) we may disregard completely the components at infini-
ty. We consider the intersection in a fiber X y • We let k = key), and for
simplicity of notation, we write the intersection number (D. C) = iy(D, C)
with fibral divisors D, C to mean the geometric intersection numbers,
since the order of the residue class tield is here irrelevant, and the prob-
lem has to do only with tibral intersection in X y' By the theory of Tate
curves, the fiber has a finite number of geometrically irreducible compo-
nents C 1 , ... ,Cm' each of which is isomorphic to a projective line, and
each of which intersects transversally exactly two other components, so
the geometrically irreducible components form an m-gon. Since the total
geometric fiber C 1 + ... + C m is orthogonal to all the components, it fol-
lows that for each geometrically irreducible component C we have
(C 2 ) = -2. By the adjunction formula of Theorem 3.7 on a tibral com-
ponent C, we conclude that

whence (K x /y . C) = 0. It follows that (Ki/y) = 0, as was to be shown.

The question arises when (Ki1y) = °for genus ~2. In the function
field case, this occurs precisely when the family is isotrivial, i.e. splits
after a finite base change. In the number field case when there is no
question of splitting, Szpiro thought this should imply that for curves of
genus ~ 2 (semis table) one should have (Ki/y) > 0. At the Arcata con-
ference in 1985, I raised the question whether the situation is not more
subtle: If the Jacobian has complex multiplication, does this imply that
(Ki/y) = 0, and also conversely? Computations prove to be difficult to
make, e.g. for the Fermat curve or its rational images, and certain modu-
lar curves. The idea, however, is that complex multiplication is the na-
tural way of forcing a curve to have no deformations, so it is a long shot
that this is the condition which corresponds to being isotrivial in the
function field case. There is not enough evidence today to raise this sug-
gestion to the status of a conjecture.
CHAPTER VI

Faltings Volumes on
Cohomology

The main purpose of this chapter is to prove the existe:nce of the Falt-
ings volumes on the determinant of the cohomology of a Riemann sur-
face.
The first two sections of this chapter belong to general algebraic geo-
metry. Determinants are playing an increasingly important role, and
should rapidly make it in basic texts. We develop ad hoc what we need
in §l. Next, we state some complements on the theory of the Jacobian
in §2, relying for some essential basic facts on proofs in [A-C-G-H].
Thus we follow the standard pattern of doing first the sheaf-theoretic re-
sults, and then metrizing them. The main existence theorem is proved in
§3.
We also give the proofs of another result of analysis which has been
used in the application of Riemann-Roch of Chapter V, §4. This result
depends on a volume computation given after the construction of the
Faltings volumes, and also on a lower bound estimate for the average
values of Green's functions. We prove this estimate in the last section.
It is an improvement of a Faltings result due to Elkies. Its proof is pure
analysis, independent of sheaf theory, so we have isolated it for easy re-
ference.

VI, §1. DETERMINANTS

The idea for proving the existence of the Faltipgs volumes on cohomo-
logy is to imbed the sheaves (1)(D) (with some convenient degree) in an
algebraic family, and to construct the determinants of cohomology for
the family. So we have to discuss determinants. We don't need much,
132 FALTINGS VOLUMES ON COHOMOLOGY [VI, §1]

but we shall describe somewhat more than we need to give the reader
some ideal of the general context of determinants. See [K-M] for a sys-
tematic development.
We first describe a special construction, which will be all that we
need.
Let p: C ~ S be a proper smooth morphism over a regular base S,
such that the fibers Cs (s E S) are non-singular curves of the same genus
g. Tn the application, S = X' or S = J is a Jacobian, and C = X x S is
simply a product.
Let 2:' be a line sheaf on C. Then p*2:' is a coherent sheaf on S, and
we have the derived sheaves R i p*2:', with i = 0, 1. For i ~ 2 all these
sheaves are 0 by the appropriate generalization of Lemma 3.1 of Chapter
V. In the application, the base can be taken to be an affine regular ring
as in the examples above, so the proof works in this case. Since R i p*2:'
need not be a vector sheaf in general, we cannot define the determinant
det Rp*2:' directly. Perhaps the simplest way is the following, which has
the advantage of being also the approach of [A-C-G-H], Chapter IV,
§3, for other purposes which will be of use to us later.

Lemma 1.1. Let D be an effective Cartier divisor on C which does not


contain any fibral component, and of sufficiently high degree. Then the
exact sequence
o ~ 2:' ~ 2:'(D) ~ 2:'(D)ID ~ 0
gives rise to an exact sequence with Rl p.2:'(D) = 0:

and the two middle sheaves (with D in them) are vector sheaves over S.

Proof. First D is a local complete intersection and so Cohen-Macau-


lay by [A-K], Chapter III, Corollary (4.5). By a standard criterion for
flatness, [A-K], Chapter V, Proposition 3.5, we know that D is flat over
S. Since D is also proper over S of relative dimension 0, it follows from
Chevalley's theorem that D is finite over S, so D is locally free over S.
Hence p*(2:'(D) ® (9 D) is a vector sheaf over S (because free over free is
free).
Next, by the standard theory of curves over a field we get
Hl(2:'(D)/C.) = 0 for all s E S, and since the Euler characteristic is con-
stant on flat families, it follows that the dimension of HO(2:'(D) IC s ) is
constant for s E S. By [Ha 1], Chapter III, Corollary 12.9, it follows that
p*2:'(D) is locally free on S, and that same reference also tells us that
R 1 p*2:'(D) = 0, as was to be shown.
[VI, §1] DETERMINANTS 133

We now want to define the determinant

We have to see that the right-hand side depends on D only up to a


natural isomorphism. Suppose D 1 , D2 are two divisors as above. After
taking D1 + D2 if necessary, we may assume without generality that
D1 ~ D 2 • Then we obtain two exact sequences as in the theorem, and an
exact diagram

o o

0-----* p*2' -----*


I I
p*2'(D 1) ----+ p*2'(D 1)ID 1 -----* R 1p*2' ----+ 0

1= I I 1=
0-----* p*2' -----* p*2'(D 2) -----* p*2'(D 2 )ID 2 ----+ R 1p*2' ----+ 0

I I
I
o
I
o
where %1, %2 are the cokernels. But the extreme left and extreme right
of the two horizontal sequences being the same, it follows that %1 ~ %2
in a natural way. From this we leave it to the reader to verify that the
natural map induced on the determinants by the central commutative
square is an isomorphism. This fact can easily be checked using
modules, and appropriate bases for the free modules which enter into
consideration, i.e. the middle ones.

We shall need a base change property for the middle sheaves in


Lemma 1.1, as in the next lemma.

Lemma 1.2. Let 2' be a line sheaf on C such that 5~ ICs has degree
~ 2g - 1 for all SES. Then:

(a) p*2' is a vector sheaf on S;


(b) p* commutes with base change on the determinan.t.
134 FALTINGS VOLUMES ON COHOMOLOGY [VI, §1]

Proof. Statement (a) is just a more general version of what we proved


previously for ii'(D). In the present case, we have HI(ii'1 C.) = 0 for all
s E S, and so the dimension of HO(ii'1 C.) is constant, so p*ii' is a vector
sheaf on S. Then if we compute the cohomology by the tech complex,
and qJ: T -> S is a base change with the cartesian diagram
<ll
--+

then the natural map

is an isomorphism, as in the proof of [Ha 2], III, Proposition 9.3, except


that in the present case, although the base change itself is not necessarily
flat, the exact sequence with the tech complex used to form cohomology
remains exact under tensor products, and also gives the cohomology
under base change. This proves (b), even in the stronger form that p*
commutes with base change for the cohomology sheaf functors Ri for
each i.
We shaH also need:

Lemma 1.3. Let .~ he a line sheaf on S. Assume S connected. Let


X = X(ii'.) be the Euler characteristic, which is constant. (Here ii'. is
the restriction of ii' to the fiber C., s E S.) Then

Proof. This is an immediate consequence of Lemma 1.1 and the pro-


jection formula

for any vector sheaf At on S. Cf. [Ha 1], Chapter III, Exercise 8.3. Note
that hO and hI separately may not behave nicely, but X = hO - hI does.

The above is all we shall need about determinants. But it may be


useful to the reader to see the general context, somewhat less ad hoc, as
shown in [K-M]. We summarize this context as follows.
Let p: X -> S be a proper morphism of schemes, always assumed
Noetherian. To each coherent sheaf :#' on X, flat over S, one can asso-
ciate a line sheaf denoted by det Rp*:#', satisfying the following proper-
ties.
[VI, §1] DETERMINANTS 135

DET 1. det Rp*ff is functorial for isomorphisms of coherent sheaves on


X.
DET 2. det Rp*ff commutes with base change, and is transitive with
respect to base change.
DET 3. If

is an exact sequence of coherent sheaves on X fiat over S then


there is an isomorphism

functorial with respect to base change and isomorphisms of


each sequences.
DET 4. Let

be a complex of vector sheaves, such that we have a homology


isomorphism

as in Mumford, Theorem of [Mu], 5. Then there is a natural


isomorphism

n
det Rp*ff ~ (8) (det tffi)®(-i)i
i= 1

commuting with base change.

In particular, if the cohomology sheaves Rip*ff (i ~ 0) are vector


sheaves, it follows from the preceding property that there is a natural
isomorphism
n
det Rp*ff ~ (8) (det Rip*ff)®(-l)'.
i= 1

DET 5. Suppose S is connected. Fix ff and let X be the constant


x(ff.) for s E S ([Ha 1], III, Corollary 9.10). If u is a section
of (r)t, then multiplication by u on ff induces an automorphism
det(u) according to DET 1, and we have

det(u) = uX•
136 FALTINGS VOLUMES ON COHOMOLOGY [VI, §2]

DET 6. If .,I{ is a line sheaf on S (assumed connected again), then


there is a natural isomorphism

The reader may prefer just to apply these properties systematically


rather than go through the ad hoc arguments given at the beginning. In
any case, as for other topics mentioned previously, the theory of deter-
minants will have to make it in basic algebraic geometry texts in the im-
mediate future.

VI, §2. DETERMINANT OF COHOMOLOGY

Let X be a complete non-singular curve, say over an algebraically closed


field. In this section, we derive a number of basic sheaf-theoretic lemmas
concerning divisor classes. The most useful reference for us now is
[A-C-G-H]. In the next section, we shall combine the lemmas with
Arakelov metrics to get the Faltings volumes.
We let J be the Jacobian of X. Thus J is isomorphic to the group of
divisor classes of degree O. We fix a point Po on X. Then there is a
unique canonical imbedding

cp:X-+J such that

We let r be a suitably large integer, and we let E be a fixed divisor on X


of degree
deg E = r + g - 1.

For P = (P l' .•. ,P,) E X' we let Dp = E - (P 1 + ... + P,). We then get a
map

by

Here CI denotes the linear equivalence class of the divisor, identified with
a point in J after we make the translation by (g - 1)Po.
By the theorem of the Jacobian as Picard variety, there is a line sheaf
2 on X x J called the Poincare sheaf, satisfying:

21 (P 0 x J) is trivial;
for each a E J, the line sheaf 21 X x {a} = 2a x {a} is a sheaf of de-
gree g - 1 on X;
[VI, §2] DETERMINANT OF COHOMOLOGY 137

and the map a f-+ 2'a gives an isomorphism between J and isomorph-
ism classes of line sheaves on X of degree g - 1.

If D is a divisor of degree d on X we denote S",(D) the image of


D - dP o in J. Thus if S",(D) = a and D has degree g - 1 then

2'IX x {a} ~ (!Jx(D).

For a proof in a style related with our present exposition, see


[A-C-G-H], Chapter IV, §2. We now consider the map <I> = id x CPr>
and the commutative diagram

We let p, q be the projections on the factors X' and J respectively.


On X x X'=X o X Xl X ... X X, (with X;=X for i=O, ... ,r) we
define a divisor 15 as follows. Let 15; be the divisor on X x X' which
projects on the diagonal of X 0 XXi' i ~ 1, and on the full factor X j for
j"# i. Thus
15 i = pr~i (diagonal of Xo x XJ,

Let
,
15 = E x X' - L 15 i.
i= 1

15.(x x P) = Dp x P,

and therefore

Thus

also parametrizes divisor classes of degree g - 1 on X, as does <1>*2'. A


standard lemma of elementary algebraic geometry tells Wi that they are
related in a trivial way.
138 FALTINGS VOLUMES ON COHOMOLOGY [VI, §2J

Lemma 2.1. There exists a line sheaf .A on xr such that

Proof See, for instance, [Ra 1J, III, Exercise 12.4. The lemma used to
be called the seesaw principle (on products).

For each sufficiently small open neighborhood U of a point in X', the


lemma implies that we have an isomorphism

I/Iv: <l>* 2'/(X x U) --+ lD(D)/(X xU),

and such an isomorphism is unique up to multiplication by a unit on


X x U, so a unit on U. Thus we obtain an induced isomorphism on co-
homology

Lemma 2.2. The induced isomorphism det Rl/lv is independent of the


choice of 1/1. and thus defines an isomorphism

Proof. This is a special case of Lemma 1.3, since det Rp*<l>* 2' has
weight hO - hI = o.

Lemma 2.3. With the previous notation, letting q: X x J --+ J be the


projection, we have a natural isomorphism

Proof. This is a special case of Lemma 1.2(b), applied to the auxiliary


sheaves p*2'(D) and p*(2'(D)/D) used to compute the determinant in
Lemma 1.1.

We shall now determine the sheaf det Rq*2'. First we make a remark
on an injective homomorphism of vector sheaves. Let Z be an integral
scheme, and let $0, $1 be vector sheaves on Z, of the same rank, with
an exact sequence

Then we can form the determinant

det T: det $0 --+ det $1.


[VI, §2] DETERMINANT OF COHOMOLOGY 139

Let W be the Cartier divisor defined by det T = O. Then we have a


natural isomorphism

Indeed, if {e} is a local basis of det E 1, then the image of det T is gener-
ated by some multiple se, with some section s, which defines the divisor
W locally, so the isomorphism is clear.

We shall apply this situation to the case when Z = J, and 8°, 8 1 are
used to compute the determinant of cohomology. For our purposes, the
theta divisor is defined as the sum of the curve taken 9 - 1 times on J,
i.e.
e = <p(X} + ... + <p(X} (g - 1 times).

(The convention fits the one of algebraists, so e = Wg .- 1 , where Wg - 1


has the same meaning everywhere; but [A-C-G-H] use e to denote the
divisor of the Riemann theta function.)

Lemma 2.4. There is a natural isomorphism

In fact, if we have a resolution as in Lemma 1.1, then q*ff' = 0, so we


have the exact sequence

with the vector sheaves 8°, 8 1 of the same rank. Furthermore, e is the
Cartier divisor defined by det T = o.

Proof. If a E J does not lie in the theta divisor e, then

by a basic property of the theta divisor, and so the map Ta is an


isomorphism. Hence the natural map T: CO --+ 8 1 is inject.ive since tf°, C 1
are locally free. Furthermore, det Ta is an isomorphism, so the sheaf
det Rq*ff' has a natural isomorphism with (!) J over the complement of e.
Since R 1 q*ff' is 0 at a generic point of J, it follows that R 1 q*ff' is a tor-
sion sheaf. From this and the preceding remark, we conclude that Co,
8 1 have the same rank and T is injective, i.e. q*ff' = O. That e itself is
defined as a Cartier divisor by det T = 0 follows from a general result of
Kempf, see [Ke] and [A-C-G-H], Corollary (4.5) of Chapter IV, p. 190.
140 FALTINGS VOLUMES ON COHOMOLOGY [VI, §3]

Their scheme Wd is defined precisely by the determinant det T, and we


apply the lemma with d = g - 1 and r = 0, in which case

since e is an irreducible divisor on J. This concludes the proof.

VI, §3. EXISTENCE OF THE FALTINGS VOLUMES

Throughout this section we let X be a complete non-singular curve over


the complex numbers. In Chapter V, Theorem 3.2, we stated the ex-
istence of Faltings metrics on

for every line sheaf with It-admissible metric p. The conditions listed in
that theorem were a convenient summary of what was used in Riemann-
Roch. We repeat the statement here for convenience.

Theorem 3.1. For every line sheaf fi' p with It-admissible metric p, there
exists a unique choice of hermitian metric on det H(fi'p) satisfying the
following conditions.
F AL 1. If fi' = 01 then the metric on

corresponds to the volume on HO(OD determined by the hermi-


tian form

<W, 1'/ >= ~-1 Ix W /\ if·

F AL 2. If fi'p --+ fi'p' is a sheaf automorphism which multiplies the norm


°
by c > then the Faltings norm on det H is multiplied by

And a metric isomorphism of line sheaves induces an isometry


on det H,
FAL 3. For a divisor D and a point P on X the sheaves lP(D - P) and
lP(D) being given the canonical metrics, so lP(D)[P] has the
canonical metric, the isomorphism

det H(lP(D)) ~ det H(lP(D - P)) ® (D(D)[P]


is an isometry.
[VI, §3] EXISTENCE OF THE FALTINGS VOLUMES 141

Proof. Every!l'p is metrically isomorphic to some (!J(D) with an Arak-


elov divisor D, so all we have to do is deal with the sheaves (!J(D). The
condition FAL 3 proves the uniqueness. Conversely, we shall now show
that we can use the relation in F AL 3 to prove the existence of volumes
on the cohomology satisfying F AL 3 and F AL 1.
We may start with the sheaf (!Jx with the canonical metric, and the
prescribed volumes on HO«(!Jx) and H1«(!JX) = HO(n1r. Inductively, if we
have constructed a volume on det H«(!J(D», we use the relation of FAL 3
to determine the volume on det H«(!J(D + P», say. The only question is
that such a determination is consistent, independent of the way we ob-
tain a divisor

or

by adding of subtracting points, and independent of the order in which


we add or subtract such points. Adding P l' ... ,Pr it will suffice to prove
the consistency of the definition if we make a transposition (Pi' Pi +1) 1--+
(P i + 1 , P;) for some i. Thus we are reduced to the case when we add two
points P, Q, subtract two points, or add one point and subtract another.
Say we add two points. Then we have to compare the mctrics we get on
(!J(D) starting with a given metric on (!J(D - P - Q) and adding P, Q one
way, then another.
In one case we get

det H«(!J(D» ~ det H«(!J(D - P - Q» ® (!J(D - Q)[P] ® lD(D)[Q],

and in the other case

det H«(!J(D» ~ det H«(!J(D - P - Q» ® (!J(D - P)[Q] ® (!J(D)[P].

By Lemma 3.3 of Chapter V, the sheaf isomorphism

(!J(D - P)[Q] ~ (!J(D)[Q]

induces an isomorphism on cohomology with a metric factor G(P, Q)1/2,


while the sheaf isomorphism

(!J(D - Q)[P] ~ (!J(D)[P]

induces an isomorphism on cohomology with a metric factor


G(P, Q)1/2 = G(Q, p)1/2. Hence the metrics we get on the determinant of
the cohomology are the same, independent of the order iI1L which we add
P and Q. The proof in the other cases is similar.

Any line sheaf with an admissible metric is metrically isomorphic to


some (!J(D) with a scalar multiple of the canonical metric, so we have
142 FALTINGS VOLUMES ON COHOMOLOGY [VI, §3]

satisfied FAL 1 and FAL 3. The problem is to prove that FAL 2 is sat-
isfied, and this is elaborate.
We have to show that if we have a metric isomorphism

then the induced map on the determinant of cohomology is metric pre-


serving. Using FAL 3, and adding or subtracting points, it suffices to
prove this property for divisors of any given degree, and we choose
degree 9 - 1, where 9 now denotes the genus of X. In that case, hO = hI
and we are thus supposed to prove that any sheaf isomorphism

induces an isometry. Recall that a sheaf isomorphism is obtained by


multiplication with a rational function f such that (f) = DI - D 2 • Or in
other words, we are supposed to prove that the metric we have defined
by FAL 1 and FAL 3 on det HlD(D) depends only on the equivalence
class of the divisor D. For this we put divisors of degree 9 - 1 in an
algebraic family, actually two families, parametrized respectively by prod-
ucts of X with itself and by the Jacobian. We then obtain a family of
metrics, and to show this family is constant, we show that the metric
arises from a pull back from the Jacobian, which parametrizes the
classes.
The algebraic part of this program was carried out in §2. Here we
put the metrics.
Lemmas 2.2, 2.3, and 2.4 give us isomorphisms

of line sheaves on X'. We let h be the Faltings metric on det RP.lD(D)


which is such that at a point P E X', the metric on the fiber

comes from the first two isomorphisms and the construction using
F AL 3. The isomorphism of det Rp* lD{D) with cp~lD i - 0) is such that
the section 1 of lD i - 0) corresponds to the image of 1 under the
isomorphism
C -='+ det HlD{Dp)

for those points P = (P l , ••. ,P,) such that

HO(X, lDx{Dp) = HI(X, lDx(Dp) = O.


[VI, §3J EXISTENCE OF THE FALTINGS VOLUMES 143

It will therefore suffice to prove that the FaItings metric on the fibers of
det Rp*(!)xxx.(D) is induced as the pull-back of a hermitian metric on
(!) i - 0) by CPr' and the other natural isomorphisms of Lemmas 2.2 and
2.4. We shall prove this in Theorem 3.3 below.
For this purpose, we recall that (!) i - 0) has a na turaI metric arising
from the theta functions. Indeed, the map cp: X -+ J induces an
isomorphism cP*: H°(J, O}) -+ HO(X, 01) on the holomorphic differentials.
Therefore there exists a basis {wi. ... ,w;} of holomorphic differentials on
J such that {cp*wf, ... ,cp*w:} is an orthonormal basis on the curve. We
then define the canonical (1, I)-form on J to be

Then by definition, cP* IlJ = Ilx is the canonical form on X.

Theorem 3.2. Let 0 be the theta divisor on J. There exists a metric h'
on (!)i - 0) such that

and such a metric is unique up to a constant factor.

The uniqueness results from Chapter I, Theorem 3.1, that if two met-
rics have the same Chern form, then one is equal to a constant multiple
of the other. The existence is proved in self-contained details in [La IJ,
Chapter 13, Theorem 1.1, Theorem 2.4, Proposition 3.1, and Theorem
4.1, expressed in terms of the Neron function associated with a divisor.
By the formula of Chapter I, Proposition 2.2, we know that if D is a div-
isor on J, then the metric P;., D on (!)iD) is defined in terms of the Neron
function by

for P not in the support of D. Then h' = P;., -8' For our limited pur-
poses, we won't need any other information on the metric h' besides the
statement of Theorem 3.2. To orient the reader, for other purposes, we
recall briefly that if F D is the normalized theta function associated with a
divisor, and defined uniquely up to a constant factor, then the corre-
sponding Neron or Green's function is defined by

where H is the associated hermitian form (see [La IJ, Chapter 13,
Theorem 1.1). Then the Chern form is
144 FALTINGS VOLUMES ON COHOMOLOGY [VI, §3]

In the application, we take D = -0. To get an explicit determination of


F -8' which is just Fel , we let () be the Riemann theta function. Then

for some point w we have the divisor relation «()w) = 0. We let F 8 be a


normalized theta function equivalent to ()w, i.e. F9 = ()w times a trivial
theta function. Cf. [La 1], Chapter 13, Theorem 4.1 and [La 3], Chapter
VI.

Theorem 3.3. Let % = det Rp/I>*!l' ~ cp~det Rq*!l' on X'. Then cp~h'
is a constant multiple of h under the isomorphism of Lemmas 2.2, 2.3
and 2.4.
As already noted, the proof of Theorem 3.3 will conclude the proof
that the Faltings metrics satisfy FAL 2. Then Faltings normalizes the
metric P;',8 (and so hi) by the condition

f
J(C)
1112p .. ,e ~(gll)g

g' rJ
= 2- g/2 .

In light of the isomorphism det Rq*!l' ~ @JC-0), the determinant


det H!l' a is canonically isomorphic to the fiber @JC - 0)a for a E J, and
thus can be endowed both with the Faltings metric and the metric com-
ing from PA, -9' By Theorem 3.3 there is a number <5(X) such that under
the isomorphism, we have
1,1 Fal = e- 6(X)/81·1 P.. , _e'
For a discussion of Faltings' <5 see [Fa 1], and Bost [Bo 2] for more in-
formation, as well as the appendix for the role of <5 in some applications.
Faltings views <5 as a function on the moduli space .JIg, and we note
that <5 should be a normalized, generalized Weil function with respect to
the divisor at infinity in a suitable compactification. The difference from
a Wei! function should be something like a log log term.

We now give the proof of Theorem 3.3. This is done by a computa-


tion of the Chern form in each case, and we carry out this computation
in two lemmas. The first lemma gives us an expression for c 1 (cp:h ' ),
which by Theorem 3.2 we already know is equal to -gcp:J1.J'
Lemma 3.4. Let Wk = cp*w~ for k = 1, ... ,g. Let CPr: X' -+ J be the map
as in §2. Then

cP~ J1.J = ~2-1


9
f k=l
[(.± )=1
pr1wk) 1\ (.±
)=1
pr1wk) ] .

where prj is the projection on the j-th factor.


Proof The map CPr is the composite of the natural imbedding X' 4 J'
by the product map cp('), the sum map S: J' -+ J, and a translation. The
[VI, §3] EXISTENCE OF THE FALTINGS VOLUMES 145

holomorphic differentials on J are invariant under translation. Under


the pull back by the sum map S, we have trivially
r
S *J_~
Wk - t... prj*J
Wk'
j= 1

because S = L prj, and for any homomorphisms IX, p: A -+ B of abelian


varieties, and a holomorphic differential form W on B we have

(IX + P)*w = IX*W + P*w.


Over the complex numbers this is obvious since homomorphisms are re-
presented by linear maps on the universal covering spac~. Taking pull
backs to the curve then proves the lemma.

Lemma 3.5. Let h be the Faltings metric on det Rp* (l)x x x.(D). Then

Proof Let E be as in §2 our fixed divisor of degree r + g - 1, and


Dp = E - (P 1 + ... + P r ), where P = (P 1> ••• , P r).

By Lemma 3.3 of Chapter V, the Faltings metric on det Rp*(I)(D) is such


that the exact sequence
r
0-+ (I)(Dp) -+ (l)x(E) -+ EB (l)x(E)[P j] -+ 0
j=l

gives rise to a metric isomorphism of complex line sheaves

® (l)x(E)[PJT8I - 1 • n G(P"
r
det H(I)x(Dp) ~ det H(I)x(E) ® p)1/2,
j= 1 i<j

where G is the exponential Green's function, namely log G = 2A.. As a


function of P E X r , the factor det H(I)x(E) is constant, and hence has zero
Chern form. The metric hE on (l)x(E) being admissible has Chern form

where Jl = Jlx is the canonical form. Since

if hj denotes the metric on l!Jx(E)[Pj]' then


146 FALTINGS VOLUMES ON COHOMOLOGY [VI, §4]

Finally, Theorem 3.3 of Chapter 2 gives us the additional value of the


Chern form corresponding to the Green factor, namely

L ddClog G(P;. Pj)'


i<j

The factor ~ disappears because the Chern form is defined from the
square of the norm of a section. Hence we find the sum of the two ex-
pressions:

-(r +9 - 1) L prrJi.
i= 1

For each integer m = 1, ... , r the inverse projection pr: Ji. occurs exactly
r - 1 times in the sum over i < j. This gives rise to a cancellation with
the first term, and yields

by Lemma 3.4.

This concludes the proof of Lemma 3.5, and as already stated, it also
concludes the proof of the existence of the FaItings metrics on the deter-
minant of cohomology.

VI, §4. ESTIMATES FOR THE FALTINGS VOLUMES

As before, we let X be our compact Riemann surface of genus 9 ~ 2, and


we let !l' be a line sheaf with a It-admissible metric p, where It is the
canonical (1, I)-form.
For the rest of this section, we suppose that

deg !l' ~ 2g - 1 so
[VI, §4] ESTIMATES FOR THE FALTINGS VOLUMES 147

Then H(!l') = HO(!l'), and we have the Faltings metric, and Fairings vol-
ume on det HO(!l'), which we shall denote by VoI Fal •
On the other hand, we have the L 2 -metric on HO(!l'), given by

whence the associated L 2 -volume denoted by Vol 2• on HO(!l').

Theorem 4.1 (Faltings). Under the above assumptions, let r ~ g so that


r + g - 1 ~ 2g - 1. Let!l' be a line sheaf on X with deg.9" = r + g - 1.
Let h(r) = (1/41t)r log r + OCr). Then on det H(!l') we have

Remark. Actually the constant factor on the right is an improvement


on Falting's result following from an analytic improvement by Elkies for
a lower bound of the Green's functions, given in §5. Th.~ r! wins over
e- h(r)/2. and so the Faltings volume goes to infinity with respect to the
L 2-volume. I don't know yet the significance of this.

Proof By the Riemann-Roch theorem, there exists an open subset U


of xr such that if (PI' ... ,Pr ) E U, then PI' ... ,Pr are distinct and

Therefore
det H!l'( -PI _ ... - P r ) = C

with the FaItings volume on C being determined by a positive real


number, which gives the norm of 1. We denote this number by
w(P 1 , ••• ,Pr ) so that we let by definition

I1I~al = w(P I' ... ,Pr )·

Lemma 4.2. The positive numbers w(P l' ... ,Pr ) are bounded from below
independently of !l' and r.

Proof Under the isomorphism of Lemmas 2.2, 2.3, 2.4 we have a met-
ric isomorphism

C = det H!l'( -PI - ... - P r ) ~ (9A -E»tpr(Pl, ••.• Pr)·


Under this isomorphism
148 FALTINGS VOLUMES ON COHOMOLOGY [VI, §4J

where Ae is a Neron function, and so this function is bounded from


below and tends to infinity on e. This proves the lemma.

By Chapter V, Lemma 3.3, we have a metric isomorphism

det HO 2' ~ w(P l' ... ,Pr )I!2 . nG(P


i<j
j, P j )I!2 . ® 2'[PJ.
r

i= 1

Furthermore, for each (P 1> ••. ,Pr) E U we have a linear isomorphism


r
HO(2') ~ EB2'[PJ given by S r--+ (S(P 1), ... ,S(Pr ».
;= 1

Let {S l' ... ,sr} be an orthonormal basis for HO 2' with respect to the
L2-norm. Let prj (j = 1, ... ,r) be the projection on the j-th factor of x r.
VVe have a vector sheaf
r
EB prj(2') on X'
j= 1

and its top exterior power. VVe can define a section j~ of the vector sheaf
by

Then

is a section of this top exterior power, and by definition we let

The above vector sheaf has the orthogonal sum of the hermitian prod-
ucts on each factor, and so its top exterior power has the corresponding
hermitian product, whence a metric so we can take the norm of the
determinant. Then we conclude that

VolFadHO 2' = w(P 1 , ... ,Pr)n G(Pj, P) ·ldet(sj(Pj »)12 Vol 2 1Ho 2'.
i<j

VVe now take the integral of both sides over U, fixing the volumes on
some set, say the unit ball for the L 2 -volume, so that this value is con-
stant. Since U is a Zariski open set, its boundary in x r has measure 0,
and so the integral over U is the same as the integral over x r. VVe use

n
the lower bound for w(P l' ... ,Pr) of Lemma 4.2, and the lower bound of
G(P j , P) which will be proved in Theorem 5.1 of the next section, and
[VI, §5] A LOWER BOUND FOR GREEN'S FUNCTIONS 149

we obtain:
Vol Fa Je- h(,)/2 Vol 2 ~ Lldet(s;(p)W.u(P1) /\ ... /\ .u(P,)

= f x'
Idet(s;(Pj »1 2 .u(P 1) /\ ... " .u(P,)

= r!

by a computation which will be carried out in Lemma 4.3, and which


will conclude the proof of the theorem.

Lemma 4.3. We have

f x'
Idet(s;(P)W.u(Pd /\ ... /\ .u(P,) = r!

Proof. We can write J; = pr!s; + ... + pr:s;. Then

11 /\ ... /\ fr = L e(o) pr! S"I /\ .•. /\ pr: s"r>


"
where u ranges over all permutations of {I, ... ,r} and f:(U) is the sign.
Then we find

111 /\ ... /\ frl2 = (11 /\ ... /\ fr,J1 /\ ... /\ fr)

n
'" t

= Le(o)e(r) (p r1s"j, p r1stj)


0',1' j

and therefore

But

so only the terms with u = r give a contribution in the preceding sum,


and those contribution add up to r!, thus concluding the proof.

VI, §5. A LOWER BOUND FOR GREEN'S FUNCTIONS

In this section we give a strengthening of Faltings' theorem by Elkies


(conjectured by Hriljac).
150 FAL TINGS VOLUMES ON COHOMOLOGY [VI, §5]

Theorem 5.1 (Elkies). Let X be a compact Riemann surface of genus


~ 1. Let 9 be the Green's function associated with the canonical volume
form Il. Then there exists a constant Co having the following property.
For all families of distinct points x = {Xl"" ,Xr } on X and real numbers
a = {a l , •.• ,ar } with aj ~ 0 for all i, we put a = aj and L
E(a, x) = L ajajg(x j, x).
i¢j

Then

E(a, .± at[IOg(~)
x) ~ - 41n ,=1 a,
+ co].

In particular, if aj = 1 for all i, then

L g(x i , Xj) ~
i¢j
-
n
1
-4 r log r - OCr).

Proof As in Elkies' exposition we first prove the analogous theorem


in the case of the circle R/Z to show the structure of the proof. Thus we
suppose that Xl"'" Xr are in R/Z, and we let the Bernoulli function be

B(z) = {Z}2 - {z} + i,


where {z} is the fractional part of z. We define

g(X, y) = tB(x - y).

In this case we want to prove the stronger inequality E ~ -l2 L aj.


We expand B(z) in a Fourier series, so we get the Fourier expansion
for 9 to be

Therefore

as was to be shown.
[VI, §5] A LOWER BOUND FOR GREEN'S FUNCTIONS 151

The analogous approach in higher dimensions uses th(: general ortho-


normal expansion

g(x, Y) ""
--
L I1 ep.lx)ep.ly),
where {ep;.} is an orthonormal family indexed by the positive eigenvalues
of the Laplacian d associated with /1. Referring back to Chapter II, §1,
Theorem 1.5, we know that the Laplacian satisfies

where (jx is the Dirac functional. In this case, the sum of the coefficients
l/A diverges. To make it converge, we replace 9 by
-t),
e --
glx, y) =LT ep;,(x)ep;'(y) with t > O.
y

I shall recall briefly below the justification for the analytic facts that we
are using to ensure convergence, and also the proof of the following
lemma.

Lemma 5.2. For all t > 0 and x -# y we have g(x, y) ~ !1lX, y) - t.

We may then prove Theorem 5.1 as follows.


Let tj (i = 1, ... , r) be positive numbers to be determined later. Then
by the lemma,

E = L ajajg(xi' Xj)
i*j

~ L ajaigt,+t/Xj, Xj) - (t i + tj»


i*j

= ~ ~ [~j e-(ti+tj).!aiajep.lXi)ep;,(X)] - i~j (t j + tj)(aiaj)

= ~~[I~e-tilaiep.lXi)r - ~e-2tilarlep;,(Xi)12]

+ 2(~ tja? - a ~ tiai)


~ - ~ a;(~ e-~tillep.lXi)12) - 2a ~ tia i

= -( ~ arg2t,(Xi, Xj) + 2a ~ tiaJ


152 FALTINGS VOLUMES ON COHOMOLOGY [VI, §5]

Since by Lemma 5.3 below

1
g/x, x) = - - log t + 0x(1),
4n

the optimal choice for t; is

1 a;
t·=--
• 8n a'

and with that choice we obtain the desired bound on E.

As Elkies shows, the same argument proves a similar theorem for all
compact Riemannian manifolds, and he also shows that the result is best
possible by explicitly showing how to place points to achieve the desired
lower bound. We do not reproduce these results here.
We shall now justify the analytic estimates needed for the above
proof. Elkies' proof of Lemma 5.2 held in two lines, assuming some facts
which are standard for analysts; but this book may also be read by alge-
braic geometers or number theorists who may need additional explana-
tions, for which I myself am indebted to Beals.
In the first place, by the Sobolev inequality of elliptic differential
equations, the eigenfunctions are uniformly bounded, independently of A,
on X. In addition, by a standard result of elliptic theory (see, for in-
stance, [Ta], Chapter XII, Theorem 2.1), the number of eigenvalues
(counting multiplicities) bounded by some number M is polynomial in
M, actually O(M2). [This goes essentially back to Hermann Weyl. Also
one can give an expression of the form cMn + O(M n - 1 ) on an n-dimen-
sional manifold, with the main term being classical, and the good error
term being due to Hormander [Ho], but we don't need anything so pre-
cise here.] So the expression for gt converges pointwise perfectly well for
t > 0, and can be differentiated term by term with respect to t.
There remains to prove Lemma 5.2. We first make some remarks on
the maximum principle.

Let u E C"(M x R~o) satisfy

au = Au
at
Then either u is constant, or for every t > 0 we have

sup u(x, t) < sup u(x, 0).


XEX xeX
[VI, §5] A LOWER BOUND FOR GREEN'S FUNCTIONS 153

This follows from [P- W], Theorem 3, p. 173, and the following remark,
p. 174. In fact, suppose sup u(x, t) is attained at xo. Let U be a co-
ordinate neighborhood of Xo and let V be an analytic disc in U contain-
ing Xo such that j7 c U. Applying the above theorem and remark to u
in V x R we have that either u is constant in j7 x [0, t], or there is a
point in j7 x [0, t) where u has value greater than u(xo, I); iterating this
we see that if u is not constant, then (*) holds.

Now for Lemma 5.2, let

We take the eigenfunctions to be real, which saves us the trouble to put


the complex conjugate over them. Given a test function ~~, let

where (', .) is the L 2 - inner product. Then as t -+ 0,

u(x, t) -+ L (1/1, cp;.)cpix) = I/I(x) - (1/1, 1).

In fact, the convergence is uniform with respect to x for ~~ E Coo.


For t > 0, we can differentiate freely and conclude that

The same equation is satisfied by

vex, t) = u(x, t) + (1/1, 1).


If 1/1 ;;:; 0, then v(x,O) = I/I(x) ;;:; 0, and it follows from the maximum prin-
ciple recalled at the beginning of the proof that vex, t) ;;:; 10 for all x and
t > 0. Thus we have

u(x, t) = vex, t) - (1/1, 1);;:; -(1/1,1) if 1/1 ;;:; n.


Now consider

as a function of (x, t). Again, as t -+ 0, this converges nicely to


<g(x, .), 1/1). We can differentiate in t for t > and we get °
a
Dr <gt(x, .), 1/1) = -<ht(x, ,),1/1) = -u(x, t) ~ (1/1, 1) if 1/1 ;;:; 0.
154 FALTINGS VOLUMES ON COHOMOLOGY [VI, §5]

Therefore

<gt(x, .), 1/1) = <g(x, .), 1/1) - {U(X, s) ds


~ <g(x, .), 1/1) + { (1/1,1) ds

= <g(x, .) + t, 1/1) for 1/1 ~ o.


Since gt(x, y) and g(x, y) are locally integrable in the variable y, and con-
tinuous for y i= x, it follows that

gt(x, y) ~ g(x, y) +t for all x i= y.

This proves Lemma 5.2.

Finally, we give the estimate for glx, x) used at the end.

Lemma 5.3. Uniformly in x, we have

1
gt(x, x) = - -- log t + 0(1).
4n

Proof This follows from the asymptotic expansion originally due to


Minakshisundaram and Pleijel, cf. [Mi], [Mi- P], and for a more recent
exposition [B-G- M], Chapter III, E.III.9. The asymptotic expansion
gives ht(x, x), and then gt can be obtained by a simple integration. See
also [McK -S].
APPENDIX by Paul Vojta

Diophantine Inequalities
and Arakelov Theory

In this appendix we consider a question of Parshin based on the Van de


Ven-Bogomolov-Miyaoka-Yau inequality between the Chern classes for
a compact complex surface. An arithmetic analogue of this inequality
would imply a bound on the heights of rational points on a curve of
genus g > 1, as well as the Fermat conjecture for sufficiently large expon-
ent.
The idea dates back to Par shin's 1968 paper [Pa 1], which is the first
to note that the Shafarevich conjecture implies the Mordell conjecture; at
the same time Parshin notes that it would also suffice to bound Ki in
the function field case. Commenting on Parshin's paper, Faltings wrote
in [Fa 1], p. 411: "As in [P] we could derive the Mordell conjecture if
we could bound <W x , W x ) from above in terms of K, {} and the set of
places v where X has bad reduction." One would then apply this bound
to the complex or arithmetic surfaces produced by the Kodaira-Parshin
construction relative to rational points on a fixed curve C of genus
greater than one. This then gives a bound on the height of the rational
point. Also, in an appendix to the Russian edition of Fundamentals of
Diophantine Geometry (appearing presently in AMS translation) Parshin
and Zarhin point out that the inequality ci ~ 3c 2 would suffice by the
same argument. In October of 1986, Parshin gave a talk in Paris, in
which he reformulated the Chern class inequality into a conjecture for
arithmetic surfaces, and showed that the resulting conjecture implied a
weak form of a conjecture of Szpiro between the conductors and discri-
minants of stable elliptic curves defined over Q; this in tum implies the
asymptotic Fermat conjecture by an argument due to Frey. Here we
give a more direct approach, using the effective bound for the heights of
156 DIOPHANTINE INEQUALITIES AND ARAKELOV THEORY [App. §l]

rational points on the curve X 4 + y4 = Z4 as a function of the normal-


ized discriminant of the field of definition of the point.
The first section sets some notation. In §2, we discuss the function
field case, which motivates the conjectures in the number field case. In
§3, we show how a certain upper bound on

(which may be called a canonical class inequality) implies a height in-


equality. We give another approach for such an implication in §4, sacri-
ficing concreteness in order to obtain uniformity in [F(P): F].
Conversely, we also show how conjectured height inequalities yield an in-
equality related to a canonical class inequality. Finally in §5 we give one
application to the curve X 4 + y4 = Z4, showing that the height inequali-
ties for this curve imply the asymptotic Fermat conjecture and a weak
form of the Masser-Oesterle abc conjecture.

§1. GENERAL INTRODUCTORY NOTIONS

Heights

Let X be a regular arithmetic surface over Y = Spec(R) where R is the


ring of integers of a number field F. We assume throughout that the
generic fiber is geometrically irreducible. If D is a divisor on X F, then we
have a height function

defined on the set of algebraic points of X F , and well defined by


D mod 0(1), i.e. modulo bounded functions. We say that hD is associated
with D. Furthermore, hD depends only on the rational equivalence class
of D. An example that will come up later is the case where the generic
fiber C of XjY is embedded in p2 and D is the restriction of a hyper-
plane section to C. Then the height may be defined as

where the sum is over all places of F(P) and [x: y: z] are homogeneous
coordinates for the image of P in p2.
On the other hand, heights can be defined via the methods of Arake-
lov intersection theory, as follows. If P E XF(Qa), then let Ep be the
Zariski closure of P in X, so that Ep is an irreducible horizontal divisor
said to correspond to P. Let D be an Arakelov divisor on X.
[ApP. §1] GENERAL INTRODUCTORY NOTIONS 157

Proposition 1.1. The function

1
PI-+[F(P):Q] (D.E p)

is a height function in the class of heights mod 0(1) associated with the
restriction of D to the generic fiber X F •

Proof The height is a sum of local Weil-Neron functions, and for


each absolute value w on F(P), the function PI-+(Dhor.Ep)w is a (normal-
ized) Weil function associated with the horizontal part Dhor ' For the ver-
tical part, we take into account that for any v on F,

Nv[F(P): F] for v infinite,


{
(Xv, Ep)v = 10glk(v)I[F(P):F] for v finite.

Therefore the terms corresponding to fibral components (finite or infin-


ite) do not matter, because their contribution (Dver.Ep)j[F(P):Q] re-
mains bounded.

Discriminants

For every number field F we define

1
d(F) = [F: Q] log DF/Q'

where DF/Q is the absolute value of the discriminant of the number


field F. We call d(F) the normalized logarithmic discriminant of F, or
simply the logarithmic discriminant if the context is clear. If F c F' then
d(F) ~ d(F'). If P E X ~Qa) we define

d(P) = d(F(P».

Stability

In §5 of Chapter V the definition of a semistable arithmetic surface was


given, and regularity was assumed as a part of the definition. There is
also a notion of a stable arithmetic surface. This is an arithmetic surface
(not necessarily regular) satisfying conditions SS l-SS 4, except that in
SS 3, a non-singular rational component of a geometric fiber must meet
the other components of the geometric fiber in at least three points. The
concepts of stability and semistability are closely related, in the sense
that a singularity on a stable model can be resolved by a sequence of
158 DIOPHANTINE INEQUALITIES AND ARAKELOV THEORY [ApP. §1]

blowings-up, glVlng a chain of rational (- 2)-curves as in the proof of


Theorem 5.1 of Chapter V. If all singularities are resolved in this way,
then the resulting surface is semistable. Conversely, all such chains on a
semis table model can be blown down, resulting in a stable model.
If X is a semis table model over Y, we denote by X# the correspond-
ing stable model.
If X# /Y is a stable family of curves of genus g> 1, and y is a closed
point of Y, then let

0: = number of double points on the (stable) geometric fiber over y.

If X/Y is only semistable, we define 0: by first reducing to the stable


model.
It is known ([Ara 1], p. 1293) that

Arakelov proves it in the complex case but it holds in arbitrary charac-


teristic.
If X/Y is a semis table model, then for each closed point y of Y, we
write

Oy = number of double points on the (semistable) geometric fiber over y.

Also let

If v is the valuation, corresponding to the closed point y of Y, then


we also write,
and

The direct image of the dualizing sheaf

In this appendix we will be using a number of properties of the direct


image 7r*W x /y, and we summarize some of its properties here. First of
all, since W XIY is torsion free and coherent, so is its direct image (see
[Ha 1], II, 5.8.1 for the latter). Hence 7r*W x / y is a vector sheaf over Y;
its rank is g. Also, it can be related to the structure of the relative Jaco-
bian-see [Ara 1]. Recall that the degree of a locally free sheaf is de-
fined as the degree of its highest exterior power so that

The number deg 7r",.W x /y is an invariant of the family X/Y measuring its
complexity in much the same way as the height measures the complexity
of a point on a variety.
[ApP. §1] GENERAL INTRODUCTORY NOTIONS 159

In fact, there exists a theory of moduli spaces of smooth curves; i.e.,


spaces whose closed points correspond bijectively to smooth curves.
More precisely, there exists a moduli space .A9 and a (:ompactification
.Ag with the following properties.

MOD 1. Stable families X# /Y correspond in a "natural" way with


morphisms mx: Y ~ .Ag •
MOD 2. The correspondence in MOD 1 IS compatible with base
change.
In order to put the above conditions into a more rigorous framework,
one should either define .A 9 as an algebraic stack, or attach additional
"level structure" (division points on the Jacobian) to the curves in ques-
tion. This is a quite deep subject area, and it would tak,~ another book
to deal with the question in sufficient detail. For the purposes of discus-
sion we will behave as if .Ag was a scheme rather than a stack, and
leave it up to the reader to keep in mind that often we will be using
actually a morphism from a cover of Y to a cover of .Ag instead of
directly from Y to .Ag • See also this book's Foreword.
With the above caution in mind, we will often also write m(X y) to de-
note the point mx(Y) for a point y on Y. This is well defined on .Ag
if Xy is smooth. If Y is a smooth projective complex curve and X/Y is a
stable or semis table family of curves of genus g, we then have the Falt-
ings height function on .Ag:

where '1Y denotes the generic point of Y.


There is a parallel theory of a moduli space d 9 of (principally polar-
ized) abelian varieties, and a morphism from .A 9 to d 9 corresponding to
the construction of the Jacobian. If J /Y is a family of a.belian varieties
over a smooth projective complex curve Y, then we also have the Falt-
ings height of J /Y:

hFalmj'1Y» = hFaP/Y) = deg 7T..OJ J / y.


If J is the Jacobian of a stable curve, then this definition is compatible
with the preceding definition since

The same definitions of the Faltings heights carry directly over to the
number field case once we have defined a metric on det 7T..OJ x / y ; for this
we use the canonical metric as in Chapter V, §3.

Remark. Deligne has found an example when deg 7T..OJ x /y can be neg-
ative, because of the Green's functions at infinity. This i8 of course un-
like the function field case, but this is of no consequence for §4.
160 DIOPHANTINE INEQUALITIES AND ARAKELOV THEORY [ApP. §2]

§2. THEOREMS OVER FUNCTION FIELDS


Throughout this section let n: X -> Y be a semistable family of curves
of genus 9 > lover a base Y, which is a smooth projective complex
curve.

The purpose of this section is to prove the following result which devel-
oped from the work of Van de Ven, Bogomolov, Miyaoka, Yau, and
Parshin.

Theorem 2.1 (Canonical class inequality). If X is a semistable model of


genus 9 over a base Y of genus q, then

(WI/ y ) ~ 3 I b: + (2g - 2) max(2q - 2,0).


YEY

Proof Let X # denote the associated stable model. For each singular
point P E X # let Z p be the inverse image in X; then Z p will be a chain
of ( - 2)-curves of length np. If q = 0 then let D be a divisor consisting
of two distinct smooth fibers of n. Otherwise let D = O. Let Kx denote
the canonical divisor of X. Then Kx + D is numerically effective; i.e., for
all curves C on X, the intersection number (K x + D. C) ~ O. Theorem
1.1 of [Miy 2] (with N' = E = Ip Z p and N = 0) then gives

(1) (Kx + D)2 ~ 3(c z (X) - e(D» - 3I (n


p
p + 1 __
np
1 -).
+1

Here e(D) is the Euler number of the curve D; it equals - 2(2g - 2) if


q = 0 and 0 otherwise. Thus, when D = 0, the left-hand side is c 1 (X)Z, so
this inequality strengthens the inequality d ~ 3c z as the last term is non-
positive. When D i= 0, the inequality can be regarded as the analogue of
ct ~ 3('2 for the non-compact surface X\D.
In any case, letting K x /y be a divisor such that W x /y ~ (l)x(K x /y), we
have

c2 (X) = I by + (2g - 2)(2q - 2)


y

= I b: +I np + (2g - 2)(2q - 2);


y p

<1.
IIp + 1
[ApP. §2] THEOREMS OVER FUNCTION FIELDS 161

The value for C2 above follows from [Ara 1], p. 1288. The theorem then
follows from the above equations and (1).

Corollary 2.2. Let s be the number of places of bad reduction. Then

(Wi/y) ~ (2g - 2)(2q - 2 + s).

Proof Note that the maximum has disappeared: one cannot have a
curve over A I with good reduction everywhere, because adjoining divi-
sion points to the Jacobian would produce a non-trivial etale cover of
AI.
Now, let Y' be a cover of Y of degree de, where d and e are natural
numbers and Y' is ramified to order exactly e at all points lying over
points of Y of bad reduction. Such a curve can be constructed by the
Kodaira - Parshin construction if q ~ 1 and by more elementary means if
q = O. (In the latter case we may require e to be odd.) Then, applying
Theorem 2.1 to the minimal desingularization of X x y Y' gives

de(Wi/y) ~ 3d L 1J: + (2g - 2)(de(2q - 2) + dee - l)s).


y

Taking e large then gives the result.

We note that the same method gives

(wi/y) ~ L min(31J:, 2g - 2) + (2g - 2)(2q _. 2)


y

if q ~ 1.

Lemma 2.3.

Proof By Noether's formula, we have

x(@x) = l2(Ki + c 2}
= l2«Wi/y) + 1J) + (g - l)(q - 1).

By [Ara 1], pp. 1287-1288,

(Note that E of [Ara 1] equals the dual of 1t* W x / y ; see p. 1293.) This
gives the lemma.
162 DIOPHANTINE INEQUALITIES AND ARAKELOV THEORY [ApP. §2]

Note that Lemma 2.3 is close to the form in which Noether's formula
carries over into the case of Arakelov theory-for the left-hand side use
instead deg det Rn*(W x /y), with notations and metrics as in Chapters V
and VI. See [Fa 1].
Therefore Corollary 2.2 becomes
g-l b
deg n*W x/y :;:;; -6- (2q - 2 + s) + 12·

Also, by a result of Xiao Gang [Xi], Corollary 1 of Theorem 2, p. 461,


we have

b :;:;; (8 + D deg n*Wx /y ·

See also Cornalba and Harris [C-H]. Therefore:

Theorem 2.4.

This agrees well with a result due to Arakelov (implicit in [Ara 1]),
that

where go is the dimension of the trace of the Jacobian of X over Y.


We shall now show how Conjecture 5.7.5 of [V], the (1, I)-form con-
jecture, implies a weak form of Theorem 2.4. For the convenience of the
reader, we shall restate the conjecture later in the context of number
fields, in §4. Here we merely refer to it.

Theorem 2.5. Assume Conjecture 5.7.5 of [V] holds. Fix g and e > O.
Fix a smooth projective complex curve Yo of genus qo, and a finite sub-
set S of closed points of Yo. Let Y denote a smooth projective curve of
genus q mapping onto Yo and let [Y: Yo] denote the degree of that
morphism. Then, as X varies over semistable families of curves of genus
g over Y with good reduction at all points of Y not lying over S,

(2)

and

(3) (Wi/y) :;:;; (6g + e)(2q - 2) + O([Y: Yo]).


[ApP. §3] CONJECTURES OVER NUMBER FIELDS 163

Proof For (2), use [V], (5.7.5)-(5.7.7), together with the definition

For (3), use (2) and Lemma 2.3, noting that {) ~ O.


All three of these bounds on deg 7t.OJ x /y then give the Shafarevich
conjecture (for curves over function fields with fixed degeneracies), pro-
vided one also shows that there are no infinite families with bounded
heights. In the number field case, if any analogue of the:~ bounds could
be proved, then one would have the Shafarevich conjecture without need-
ing any additional results.
Because of the connections with the Shafarevich conjecture, it appears
that an arithmetic analogue of Theorem 2.1 would still be quite deep.
Finally, we note that in all of the above discussion, the quantity
(2q - 2)/[Y: Yo] is the function field analogue of d(F); also an analogue
of s in the number field case would be

s = L 10glk(v)l.
veS

With these identifications, much of the above discussion ,~arries over into
the number field case (conjecturally, at least). The next two sections will
cover this in more detail.

§3. CONJECTURES OVER NUMBER FIELDS


Notation. Let F, F', F", etc. denote number fields with rings of integers
R, R', R" and let

Y = Spec(R), Y' = Spec(R'), Y" = Spec(R"),

respectively. We also continue to let X denote a semistable family of


curves of genus g over a base Y.

One basic height conjecture is stated in [V], 5.5.0.1, as follows.

Conjecture 3.1. Let C be a smooth projective curve defined over a


number field F. Let K denote the canonical divisor c.f C. Let e > O.
Then for P varying over all algebraic points of C, we have

hK(P) ~ (1 + e)d(P) + 0(1),


where the constant implicit in O( 1) depends on C and B, but is indepen-
dent of P.
164 DIOPHANTINE INEQUALITIES AND ARAKELOV THEORY [ApP. §3]

As shown in [V], Chapter 5, Appendix ABC, this conjecture implies a


large number of other diophantine conjectures, e.g., the Masser-Oesterle
abc conjecture, the Hall and Szpiro conjectures, etc. Conjecturally, the
factor 1 + E: is best possible. One can weaken the conjecture by replacing
1 + E: with larger factors. For instance, in the function field case I can
prove it with a factor of 2 + E:. Also, Theorems 4.5 and 3.5 hold also in
the function field case, giving theorems in the function field case corre-
sponding to Conjectures 4.4 and 3.4. In the last section, we shall remind
the reader how weaker forms of Conjecture 3.1 imply the asymptotic
Fermat conjecture and weaker forms of the abc conjecture.
We wish to see how Conjecture 3.1 is related to Arakelov theory. We
shall see that an upper bound on (Wi/y) proposed by Parshin, and ana-
logous to the Canonical Class Inequality 2.1, implies a weaker form of
Conjecture 3.1, where instead of 1 + E: we have a factor linear in g. Thus
an inequality in Arakelov theory implies a height inequality. Conversely,
in the next section, we shall show how a suitable conjectured height in-
equality implies a bound on deg 1r*W x /y . The relationship between
deg det 1[*W x / y and deg det R1r*W x / y
is not clear, and corresponds to the more subtle formulation of the
Noether formula in the number field case, as distinguished from the func-
tion field case in Lemma 2.3. Presumably a bound for (Wi/y) would
also emerge from the clarification of the situation.
We start by stating Parshin's analogue of Theorem 2.1. He prefers it
to be stated as a question rather than as a conjecture.

Question 3.2 (Parshin). Do there exist effectively computable positive


numbers ao, aI' a 2 , with aI' a 2 absolute constants, and a o depending on
g, such that for all number fields F and all semistable families X/Y of
genus g the inequality
(Wi/y) ~ a 2 L <5: 10glk(y)1 + a (2g -
I 2)[F: Q]d(F) + ao[F: Q]

holds? (The sum is taken over all closed points of Y.)


There is no clear understanding today of what the constants ao, al' a z
would be like. I would suggest that a l could be taken as 1 + E:, in which
case a o would also depend on 6. One could also formulate a conjectural
analogue of Corollary 2.2; it would hold if the above inequality holds.
We leave the details to the reader.
We note that the canonical class inequality in Question 3.2 is uniform
for all semistable families of a given genus g. In many applications, and
in particular in the rest of this section, we use uniformity over a smaller
family: we fix X/Y and consider the family {Xp} of ramified coverings of
X F obtained by the Parshin-Kodaira construction, as P varies over alge-
braic points of X.
[App. §3J CONJECTURES OVER NUMBER FIELDS 165

Notice that in Question 3.2 we are summing over finite places. This
differs from Parshin's formulation, in which he uses Faltings' definition of
(\ for archimedean v. (See the definition after Theorem 3.3 of Chapter
VI; we will not discuss it further here. See also [Fa IJ, Section 6, as well
as [B-BJ and [B-G-S].) The problem here is that we are dealing with
two concepts, b v and b:. Before discussing this in detail, we shall digress
briefly to discuss the general setup. Recall from Section I the discussion
of the moduli space vi! 9 of stable curves of genus g. Let v be an archi-
medean place of F; it corresponds to an injection (J: F c.. C. Applying (J
to the generic fiber of X/Y then produces a complex curve which we will
write as X()". This corresponds to a point m(X()") on vl!iC). The desired
functions b()" = b()",x/y (resp. b: = b:' x /y) should then be obtained by eva-
luating some function on vl!iC) at the point m(X()").
Now, returning to the matter at hand, we point out that Faltings' b"
is an analogue of our bv because he uses it for Noether's formula on a
semistable model. This makes sense also because bv at non-archimedean
places depends on the fiber over the local ring, and it makes sense to
translate a p-adic analytic definition into a complex analytic definition.
On the other hand, b: depends only on the special fiber" so on the mo-
duli space, it is a question of which Zariski closed subsets of the bound-
ary contain the point m(X v), At archimedean places, the generic fiber of
X /Y is assumed to be smooth, therefore m(X ()") never lies on the bound-
ary of the moduli space. Hence the archimedean analogue of b: should
be zero.
If the reader is not comfortable with the above argument, however,
the results of this section still remain valid for ba coming from any con-
tinuous function on vi! iC). Indeed, we will be applying the inequality of
Question 3.2 to surfaces X;/Y; obtained from (varying) algebraic points
P on a fixed X/Y via the Kodaira-Parshin construction. For any corre-
sponding embedding (J: F; ~ C, it is also true that Xp,o' coincides with
the curve obtained by applying the Kodaira-Parshin construction to the
point (J(P) on X a' In particular, m(X p,a) varies continuously with (J(P)
on vi! y(C). Since continuous functions on compact sets are bounded, b()"
would then be bounded for all archimedean places (J. Th.e bound, how-
ever, would necessarily depend on the generic fiber C of X/Y, and it is
not clear how effective this bound is.
The Parshin construction of ramified coverings of a given curve there-
fore leads us to formalize some properties of the families of curves ob-
tained in this way. Let:!l' = {X/Y} be a family of semistable arithmetic
surfaces. We shall say that :!l' is a limited family if it satisfies the follow-
ing conditions.

LIM 1. There exists Yo and a finite set of closed points So of Yo such


that Y ranges over finite covers of Yo and X/Y is smooth
over all points not lying above So.
166 DIOPHANTINE INEQUALITIES AND ARAKELOV THEORY [ApP. §3]

LIM 2. There exists a fixed compact subset of the moduli space


.Ag(C), containing all points m(X a ), as X varies over the
members of ~ and a varies over all embeddings a: F ..... C.

We may now reformulate a weaker form of the inequality proposed


by Parshin.

Conjecture 3.3. Let ~ be a limited family of semistable arithmetic sur-


faces. Then for all elements XjY in ~ we have

(WilY)
[F: Q] ~ a 1 (2g - 2)d(F) + a;t',
where a l is the constant of Question 3.2 and a;t' depends on !!l".

Conjecture 3.3 is implied by 3.2. Indeed, the term involving the


numbers <5 # can be absorbed into the term a;t' because each <5: is
bounded by 3g - 3. Also, if one uses a different definition of <5: at
archimedean places, then this contribution can also be absorbed into a;t'
by LIM 2.
The purpose of this section is to show that following conjecture fol-
lows from Conjecture 3.3.
Conjecture 3.4. Let C be a smooth projective curve defined over a
number field F, and let K denote the canonical divisor. Then for P vary-
ing over all algebraic points of C of bounded degree over F, we have

where a 1 is the same as in Question 3.2 and a4 depends vnly on C, F,


and [F(P): F].

Theorem 3.5. Conjecture 3.3 implies Conjecture 3.4.

It is still an open problem to show that a 4 can be effectively com-


puted in terms of the constants in Question 3.2, but that is likely to be
the case. The difficulty lies with proving bounds on the analytical quan-
tities used in the metrics at infinity.
We shall now prove Theorem 3.5. The inequality of Conjecture 3.4 is
unaffected by base change, so we may assume C corresponds to a semi-
stable family X/Yo
For the rest of this section we fix n: X ..... Y, a semistable family of
curves of genus g> 1, and we let C be the generic fiber of n. Let P
denote an algebraic point on C, let F' = k(P), and let X' be a minimal
desingularization of X x Y Y'. Let Ep denote the corresponding horizon-
tal divisor on X and let E~ denote an irreducible horizontal divisor on
X', of degree lover Y' and lying over Ep.
[App. §3] CONJECTURES OVER NUMBER FIELDS 167

Lemma 3.6. For each point P as above there exists a semistable arith-
metic surface

and a morphism fp: X; -+ X, such that the family

forms a limited family, and satisfies the following conditions (where we


omit the subscript P for simplicity):
(a) The generic fiber C" of X" has genus 4g - 2, and C" is ramified to
order 2 above P and unramified elsewhere;
(b) The degree [F": F] is bounded (depending only on X and
[F(P) : F]), and

deg f = 4[F": F];

(c) d(F");::;; id(P) + as, where as depends only on X and [F(P): F];
(d) We have

where E~ and E; are sections of n" lying over E;, and D" is
supported only on fibers of nil; and
(e) We have

(E;'. D") ;::;; a7 deg f, i = 1,2;


(E'{ .E;) ~ a8 deg f;

where a 6 , a 7 , as depend only on X and [F(P): F].

Proof The proof draws many of its ideas from Parshin and Szpiro
[Sz 3]. We break it down into six steps.

Also, let Sl denote the set of places of F where C has bad reduction,
together with all places lying over 2.
Step 1. Construct an (almost) etale cover of X. Let C' be an etale
cover of C of degree two. Let F 1 => F be a field over whil~h C' is defined
and attains semistable reduction. Let X 1 be a minimal desingularization
of X x y Y1. Then the function field of C' is obtained from the function
field of C by adjoining the square root of a function, say, k(C') =
Jh),
k( C)( where h is a rational function on Xl.
168 DIOPHANTINE INEQUALITIES AND ARAKELOV THEORY [ApP. §3]

Let F 2 be a field containing F 1 such that, for every place v of F 1 ly-


ing over S l' all places w of F 2 have the property that the corresponding
local fields F z . w contain all extensions of Fl." of degree at most [F':F].
Let X 2 be the semistable model of C obtained by desingularizing X x y Y2'
By construction, F~ = F 2(P) is unramified over F 2 at all places ly-
ing over S l' Therefore, X 2 x Y2 Y~ is also regular, and therefore the irre-
ducible horizontal divisor on X 2 lying over E p does not pass through
any of the double points on the fibers. Also let F 3 be a quadratic exten-
sion of F2 ramified at all places v of F 2 for which v 12 or the divisor (h)
is supported on a component of the fiber of X 2 over v. Note that none
of the choices made so far depends on P, only on [F(P): F].
Now let F~ = F iP) and let X 3 = X 2 X Y2 Y~. Let E3 be the irreduc-
ible horizontal divisor on X 3 lying over E~. We note that X 3 might not
be regular, but will be regular at all points of E 3 • Let X 4 be the nor-
malization of X 3 in the function field k(X 3)(Jh). Then we claim that, if
v is a place of F'J not lying over 2 or 00, then the two inverse images of
E3 in X 4 do not meet above v. Indeed, let x denote the point where E3
meets the fiber over v; the local ring will be isomorphic to

O.,[[TJ],

where (!" is the local ring of F'3 at v. This holds because the local ring is
regular and the fiber is irreducible at x. This ring is a UFD, so we can
factor has,
h= uteh~l ... h~r,

where u is a unit in (0 v[[T]], t is a uniformizer in (!) v' and hi"" A are


distinct primes. By assumption, e l , ... ,er must be even, and we may
therefore assume.

by absorbing squares into h. But also e must be even, as F 3 is ramified


over F 2 and h comes from a function on X 2' Thus, a local ring on the
inverse image of x in X 4 is etale over X 3 at x, so the inverse images of
x are distinct.
This analysis fails for places v 12, because the discriminant of X 2 - u is
no longer a unit in (!).,[[T]]. However, we can blow up the point x,
which corresponds to replacing T with tT'. After at most v(2) blowings-
up, we will find that there exists an element a E (!)v[[T]] such that a 2 ==
u mod 2 and therefore

a+~
2

is integral over (!j,,[[T]] with discriminant equal to a unit.


[App. §3] CONJECTURES OVER NUMBER FIELDS 169

Finally, let F 5 be a quadratic extension of F 3(P) such that P lifts to


two rational points on C(Fs). By the Chevalley-Weil theorem ([La 1],
Chapter 2, Theorem 8.1), d(F 5) - d(F 3) is bounded as po varies. Let X 5
be a minimal desingularization of X 4 X Y4 Ys , and let E s and E~ be the
irreducible horizontal divisors on X 5 lying over E 3 •
Thus, we have a regular family ?t: X 5 -+ Y5 , and the pull back of E'
(as a Cartier divisor) is E5 + E'5 + D 5 , where D5 is supported only on
fibers of ?t, and E 5 and E'5 do not meet. Also, note that when we desing-
ularized X 4 X Y Y5 to produce X 5' the singular points didl not lie over E'.
4

Therefore, D5 is supported only at fibers over 2 and 00. Also, the bad
reduction of each fiber over v is independent of P if v,r 2 and varies over
a finite set of possibilities if v 12. This family also has good reduction
outside of places over S l'

Step 2. Divide (!)(E5 - E~) in two. This is an invertible sheaf of degree


zero on the generic fiber; therefore there exists a finite extension F 6 of
F 5' of bounded degree, and ramified only at places above Sl' such that
the divisor E5 - E~ can be divided by two in PicO(C)(F 6)' Thus there
exists a rational function h' on X 6 = X 5 X Y, Y6 such that,

(h') = Es - E~ + 2D + D',
where D' is supported only on fibers of ?t6: X 6 -+ Y6 • Restricting D' to
smooth fibers gives a fractional ideal of R 6 ; by Minkowski-type compu-
tations, it is linearly equivalent to an ideal b of norm bounded by,

where N = [F 6 : Q] and DF6 is the absolute value of the discriminant of


F 6' See [La, 5], Chapter V, Theorem 4. Therefore we may assume that
D' is supported only on places over S 1> plus places occurring in the ideal
b. Let S2 be the set of places on which D' is supported and not lying
over Sl'
Step 3. Construct a ramified cover. Let F 7 be an extension of F 3 rami-
fied to even order over places in S l' above and beyond any ramification
indicated by the Chevalley-Weil argument in Step 1. Then F~ = F7F6 is
ramified to even order over F 6 at all places lying over S l.' Moreover, F 7
does not depend on P. Then, by the same arguments as in Step 1,
we can let X 7 be the normalization in F 7 k(X 6 )(fi) of a blow-up of
X6 x Y. Y~. Then inverse images of E5 and E~ will remain disjoint. Ex-
cept for primes above S2, components of fibers meeting E5 and E~ will
be reduced; fibers at primes above S2 will be smooth but may occur
with multiplicity one or two.
170 DIOPHANTINE INEQUALITIES AND ARAKELOV THEORY [ApP. §3]

Step 4. Pass to a semistable model. Let F s be an extension of F 6F 7


over which the 15-level structure of the Jacobian of X 7 is rational; this
has bounded degree and is only ramified over F 6 F 7 above places in S l'
Let F" be an extension field of F s over which X 7 attains stable reduc-
tion. Let X" be the resulting semistable model. By a moduli space argu-
ment, we see that F"/F s is unramified at places not lying over S1 or S2'
and is ramified to order at most two at places over S2'
Note that, when passing to the stable model, the neighborhood above
E5 and E~ will be unaffected except that some of the blowings-up of Step
3 will be undone (but the blowings-up of Step 1 will remain). Also note
that, by construction, conditions (a), (b), (c), and (d) of the lemma hold.
Step 5. Check condition (e). For the first equation, this amounts to
controlling D" in part (d). The divisor D" contains contributions related
to fibers over S l' as well as the archimedean fibers. As noted earlier, the
types of fibers over S 1 range over a finite set, therefore their contribution
to D" is bounded.
There is also a contribution to D" at the infinite places which
amounts to a Coo function on CIf, and therefore is bounded. Moreover,
this function varies continuously as P varies, therefore the contribution
to D" is again bounded. This is admittedly a weak point in the argu-
ment, but hopefully an expert in differential equations would be able to
produce an effective bound. For a few more explanations see the end of
the section.
The other two equations can be proved by similar arguments.
Step 6. Show that the surfaces X; form a limited family. Condition
LIM 1 is clear, since all surfaces X; have good reduction at places not
lying over S1' and S1 does not depend on P. Now consider condition
LIM 2. If we were to replace the base Y with a curve C and replace P
with the point ~ on C x C, defined over the function field k(C), then
the above construction would still be valid, except for the Minkowski
argument in Step 2. For that part (noting that S 1 = 0), we see that the
vertical fibers of (hi) define a divisor on C (the base) of even degree;
therefore it can be divided in two since C (the base) is defined over C
and Jac(C)(C) is a divisible group. Thus, we obtain a diagram as in (*)
in the proof of Theorem 4.5, and we take the compact subset of uHiC)
to be the image of the compact set C(C). Finally, it is necessary to
check that the process of taking a point P on C to X; is compatible
with the correspondence (*). We leave these details to the reader.
This concludes the proof of Lemma 3.6.

Lemma 3.7. With the above notations,


[ApP. §3] CONJECTURES OVER NUMBER FIELDS 171

Proof By part (e) of Lemma 3.6,

By the adjunction formula on X",

CE'?) = -(f*W x /y ® (!J(E'{ + E'2 + D").ED,

and likewise for (E'22). Therefore,

2(CE'{ + E'2)2) = - (f*W x /y ® (!J(D"). E'{ + E'2) + 2CE'{. E'2)


~ -(f*Wx/y.E'{ + E'2) - 2a 7 degf + 2as degf.

Combining this with the earlier equation gives

(Wi"w) > (W2) 3(f*Wx/y ·E'{ + E'2) _ a - a +a .


degf = XjY + 2 d eg f 6 7 8

But the difference between the pull back of E~ and 2(E'{ + E'2) is sup-
ported on exceptional divisors for the morphism from X" to X'; there-
fore,
2(f*Wx/y·E~ + E'2) _ (WX'/y,.E~)
degf - [F(P): F] .

Moreover,

Therefore,

(Wi"w) > (W2) 3(Wx/ y ·E p ) _ _ .


degf = X/y + 4[F(P): F] a6 a 7 + as·

This concludes the proof of Lemma 3.7.

Proof of Theorem 3.5. We combine Lemma 3.7 with the inequality of


Conjecture 3.3 applied to the family {X;/Y;}Pex to give

(W X / Y • Ep) :$; a (8g - 6) [F: Q] d(F")


[F(P):F] - 1 3

+ '34( -(W 2x /y) + a6 + a7 - as


) [F: Q]
+ ~-3- a!Z.
172 DIOPHANTINE INEQUALITIES AND ARAKELOV THEORY [ApP. §3]

Recall that degf = 4[F" : F] and that the genus in Conjecture 3.3 equals
4g - 2 for the present family. Consolidating a few constants and taking
note of condition (c) of Lemma 3.6 gives

The left-hand side is hK(P). and the proof is complete.

We shall now comment briefly on the argument at infinity at the end


of Step 5. We deal more generally with a finite morphism

f: C -+ C

of complete smooth curves over the complex numbers. We let g' and g
be the Green's functions on C and C respectively. viewed as functions
on the product minus the diagonal. Let

11' = Closure in C x C ofthe set of points (Q. Q')


such that f(Q) = f(Q') and Q -F Q'.
Then there exists a smooth function

a: C x C - 11' -+ R
such that
g'=go(fxf)+a.

Let R be the ramification divisor of f. so

R = L (e a, - l)(Q').
Q'

where the sum is taken over all complex points Q' of c. If Q is not a
ramification point of f. then g' - g (f x f) extends by continuity to
0

(Q. Q) since g' and go (f x f) have the same singularity at (Q. Q). Ex-
tending g' bilinearly to divisors. so

g'(Q. R) = L (e a, - l)g'(Q. Q').


Q'

we see that the function


fJ(Q) = g'(Q. R) + a(Q. Q)

is defined and smooth on the set of points Q ¢ R. But again by compar-


ing singularities. we see that
fJ: C-+R
[ApP. §4] ANOTHER HEIGHT INEQUALITY 173

is defined as a smooth function on all of C. Then the following ques-


tions arise.
(a) What are max f3 and min f3 on C?
(b) Suppose C varies in an algebraic family, so we write {C;}tET with
some parameter curve T. Then {C;} is the fibration of a surface S,
and the family {f3t} may be regarded as a function f3 on S. The
problem is to show that f3 is continuous, and to determine its
maximum and minimum on S.
The function f3 is precisely what is needed to compare the metrics
coming from the Green's functions on the sheaves

and f*W c ® l!J(R).

Such a comparison was needed in Step 5, where we dealt with the


Parshin family {C;}, or rather {C;'} where T is an etale covering of C.

§4. ANOTHER HEIGHT INEQUALITY

Just as Conjecture 3.1 arose from Nevanlinna theory, there is another


conjecture stemming from a different type of Nevanlinna theory in [V],
5.7.5, which we restate here briefly, assuming that the reader is ac-
quainted with some terminology of differential geometry.

Conjecture 4.1 (The (1, I)-form conjecture). Let V be a quasi-projective


variety defined over a number field contained in C and let D be a nor-
mal crossings divisor on V. Let w be a hermitian (1, I)-form on V\D,
whose holomorphic sectional curvatures are bounded from above by
-c < O. Let 2:' be a line sheaf on V such that c i (2:') ~ w. Let E be
an almost ample divisor on V. Then for all D-integralizable sets of
closed points P of V,

1
h!AP) ~ ~ d(P) + shEep) + 0(1).
c

All terms used in the statement of the conjecture are: defined in [V].
We only want to help the reader with a quick statement.

We want again to see how this conjecture relates to conjectures aris-


ing from Arakelov theory. Instead of using the Parshin Inequality 3.2 to
derive a height inequality as in Conjecture 3.4, we shall use the number
field analogue of (2) from Section 2. We shall see that the (1, I)-form
conjecture implies the analogue of (2), which in turn implies a streng-
thening of Conjecture 3.4. The analogue of (2) is the following.
174 DIOPHANTINE INEQUALITIES AND ARAKELOV THEORY [ApP. §4]

Conjecture 4.2. Fix a base Yo, g> 1, a finite set of finite places S, and
s > O. For all extensions Y of the base Yo and all semistable families
n: X --> Y of genus g, with good reduction at points of Y not lying over
S,

___1___ deg n (jJ


[F: Q] * :-s;
X/Y -
(~2 + s) d(F) + 0(1).

Theorem 4.3. The (1, I)-form conjecture implies Conjecture 4.2.

Proof We refer to the notation of [V], Example 5.7.6. Basically, Con-


jecture 4.2 is (5.7.7) of op. cit. First, though, it is necessary to indicate
how det n*(jJx/y is metrized. This is done via the Jacobian:

in turn the determinants are metrized by Faltings' canonical metric


(Chapter V, §3). As noted in [Fa 2], however, this metric extends to a
singular metric on any compactification of the moduli space d 9 of abe-
lian varieties of dimension g. It is possible to compactify d 9' obtaining
.rdg , such that the boundary is a normal crossings divisor D. The
singularities on the canonical metric are then known to be bounded by
O(log( Y. D)J for any infinite place v; therefore replacing the singular ca-
nonical metric with a smooth metric affects the left-hand side of inequali-
ty 4.2 by at most o(hD)' But also, by [V], 1.2.9h, hD ~ O(h K+ D).
Therefore, as noted in [V], the difference between the (smoothly
mctrized) height occurring in the (1, I)-form conjecture and the Faltings
height used in Conjecture 4.2 can be absorbed into the s.

Conjecture 4.4. Let C be a smooth projective curve defined over a


number .field F, and let K denote the canonical divisor. Fix 6 > O. Then
for P varying over all algebraic points of C, we have

hK(P) ~ (8g - 4 + 6) d(P) + OCt),

where the constant in 0(1) depends only on C, F, and f..

We note that the bounds do not depend on [F(P): FJ See also


[V], 5.5.1.

Theorem 4.5. Conjecture 4.2 implies Conjecture 4.4.


[ApP. §4] ANOTHER HEIGHT INEQUALITY 175

Proof By [Sz 4], X.1, there exists an etale cover C of C mapping into
(the moduli space of curves of genus 4g - 2):
J{ 4g _ 2

c
Let K (resp. K') denote the canonical divisor of C (resp. C). If P' E C
lies over PEe, then

hK(P) = hK,(P') + 0(1) by functoriality


~ (4 + B')hFa.{t(P'») + 0(1) see justification below
~ (8g - 4 + e) d(t(P'» + 0(1) by Conjecture 4.2
~ (8g - 4 + B) d(P') + 0(1) because F(t(P'») c F(P')
~ (8g - 4 + B) d(P) + 0(1) by Chevalley- 1iVeil.

To justify the first inequality, we note that the diagram (*) is obtained
by applying the Kodaira-Parshin construction as in L(:mma 3.6 to the
diagonal ,1 on C x C, regarded as a point on C rational over k(C). This
produces a surface n: W --+ C which also maps to C x C; the morphism
f: W --+ C x C has degree 4 deg p and is ramified only over,1. Thus

W~le' = !*(wrg;c/c ® (f)cxcC,1»;


(W~/C,) = (degf)«W2xc;d + (W cxc /c ',1) + i(,1.,1»)
= (degf)(O + (2g - 2) - ¥2g - 2»)

= i( deg f)(2g - 2)

= 3(deg p )(2g - 2).

Then, by Lemma 2.3,

deg n* WW/e' = !Cdeg p)(2g - 2).

Therefore, by functoriality of heights (the projection fommla, Chapter III,


Theorem 4.1; see also [La 1], Chapter 4, Theorem 5.1),

hFa1(t(P'» = hDCP') + 0(1)


for some divisor D of degree tcdeg p)(2g - 2) on C. Since deg K' =
(deg p)(2g - 2), we get the desired inequality by a property of heights
([La I], Chapter 4, Proposition 3.3.).
176 DIOPHANTINE INEQUALITIES AND ARAKELOV THEORY [ApP. §5]

One should also note that on page 264 of [Sz 4], the equation

initially holds up to a divisor supported on fibers, but then it can be


shown that this divisor has even degree and therefore it also is linearly
equivalent to a 2-divisible divisor.
This approach has the advantage of being elegant, and the constants
do not depend on [F(P): F], but it would likely be harder to prove
effective bounds in this case.

§5. APPLICATIONS

All the applications we can think of at this time in diophantine analysis


come from a height inequality. In [V], 5.7 we saw how the (1, I)-form
conjecture implies the Shafarevich conjecture that given a finite set of
places S of Y = Spec(R), there is only a finite number of isomorphism
classes of curves of given genus g> lover F with good reduction out-
side S.
A large number of other applications depend only on a weaker form
of a height inequality, essentially the form given in 3.4, namely for a
fixed curve Cover F,

hK(P) ~ a c d(P) + 0(1)

with P varying over the algebraic points of C, and a c is some constant,


depending on C. In [V], Chapter 5, §5 and Appendix, a number of con-
sequences were already listed, e.g. the abc conjecture, Hall and Szpiro
conjectures, etc. Here we repeat two of these applications just to show
how the constants enter in these consequences.
We apply the height inequality to the Fermat curve of degree 4. Let
a, b, and e be pairwise relatively prime integers with an + b" = en and n
odd. Let X be the curve X 4 + y4 = Z4 C p2, defined over Q. It has
genus g = 3. Then we have a point P = [a"/4: b"/4: C"/4] on X defined
over Q(yIa, ib, ifc)·

Lemma 5.1.

3
d(P) ~ 210g 2 +- L log p
4 plabc

~ 2 log 2 + ! log max( 1ai, 1b I, 1e I).


[ApP. §5] APPLICATIONS 177

Lemma S.2.

Proof. The canonical sheaf of C is the restriction to C of CD( 1) on


p2; this gives the lemma by Proposition 1.1 and the comments preceding
that proposition.

Lemma 5.3.
max(lal, Ibl, lei) ;?; 2n + 2.
Proof. We may assume that 0 < a < b < e. Letting x := e - 1, we have
b :S; x; a :S; x-I; therefore,

(x - I)" + x" ;?; (x + 1)";


f(x) == x" - 2.L
(,,-1)/2 (n)
2' Xli;?; O.
• =0 l

By "Descartes' rule of signs," f has a unique positive n~al root (); since
f(2n) < 0, () > 2n. Therefore e is an integer strictly greater than 2n + 1.
No doubt better elementary bounds exist.
Combining Lemmas 5.1 and 5.2 with Conjecture 3.4 gives

n 15al
4log max(lal, Ibl, lei) ~ a4 + a 9 + -2- (2 log 2 + £log max(lal, Ibl, leI».

Therefore, by Lemma 5.3,

135a 1 60a 1 log 2 + 4a 4 + 4a 9


n:S;--+ .
- 2 10g(2n + 2)

Likewise, we obtain the abe conjecture with a different exponent. Let


a, b, e be pairwise relatively prime integers with a + b + e = 0 and

N = Radical(abe) == n p.
plabc

We obtain
45a 1
log max(lal, Ibl, leI) ~ -2-log N + 6Oa 1 10g 2 + 4a4 + 4a9'

The derivation is similar.


178 DIOPHANTINE INEQUALITIES AND ARAKL ('Y THEORY [ApP. §5]

We note that here we obtained a weak form of the abc conjecture


from the weakening 3.4 of Conjecture 3.1. If we l.icd Conjecture 3.1 in-
stead, we would get the strong form of the abc ce'll.1ecture; see [V], Ap-
pendix 5.ABC.

Remark. As Lang points out, there might be applications of Arakelov


theory other than to diophantine analysis. For instance, one may trans-
late Theorems 1 and l' from [Xi] to give a conjectured condition under
which, for a semistable family X/Y, the image of t11,l: fundamental group

is trivial. Thus if a family X/Y satisfied a suita:,b bound on (WilY)'


then no coverings of X would come from a coverillg of the generic fiber.
References

[Ab] S. ABHYANKAR, Ramification Theoretic Methods in Algebraic Geo-


metry, Annals of Mathematics Studies, Vol. 43, Princeton, 1959.
[A-K] S. ALTMAN and S. KLEIMAN, Introduction to Grothendieck Duality
Theory, Lecture Notes in Mathematics, Vol. 146, Springer-Ver-
lag, 1970.
[Ara 1] S. ARAKELOV, "Families of algebraic curves with fixed degenera-
cies", Izv. Akad. Nauk SSSR Ser. Mat., 35, No.6 (1971); AMS
Translation 5 (1971), pp. 1277-1302.
[Ara 2] S. ARAKELOV, "Intersection theory of divisors on an arithmetic
surface", Izv. Akad. Nauk SSSR Ser. Mat., :J8, No.6 (1974);
AMS Translation, 8 No.6, (1974), pp. 1167-1180.
[A-C-G-H] E. ARBARELLO, M. CORNALBA, P. GRIFFITHS, and J. HARRIS, Geo-
metry of Algebraic Curves, Vol. I, Springer-Verlag, 1985.
[Art 1] M. ARTIN, "Some numerical criteria for contractibility of curves
on algebraic surfaces", Amer. J. Math., 84 (1962), pp. 485-496.
[Art 2] M. ARTIN, "On isolated rational singularities of surfaces", Amer.
J. Math., 88 (1966), pp. 129-136.
[Art 3] M. ARTIN, Grothendieck Topologies, Lecture Notl~s, Harvard, 1962.
[A-W] M. ARTIN and G. WINTERS, "Degenerate fibers and reduction of
curves", Topology, 10 (1971), pp. 373-383.
[Be 1] R. BERGER, Uber verschiedene Differentenbegriffe, Sitzungsber.
Heidelberger Akad. Wiss. Math.-Natur. KI., 1 Abh. 1960.
[Be 2] R. BERGER, "Ausdehnung von Derivationen und Schachtelung der
Differente", Math. Z., 78 (1962), pp. 97-115.
[Be 3] R. BERGER, "Differenten regularer Ringe", J. Rdne Angew. Math.
(1964), pp. 441-442.
[B-G-M] M. BERGER, P. GAUDUCHON, and E. MAZET, 1£ Spectre d'une Var-
iete Riemannienne, Lecture Notes in Mathematics, Vol. 194,
Springer-Verlag, 1971.
180 REFERENCES

[B-B] J. M. BISMUT and 1. B. BasT, "Quillen metrics and degenerating


Riemann surfaces", to appear
[B-G-S] J. M. BISMUT, H. GILLET, and C. SOULE, "Analytic torsion and
holomorphic determinant bundles", Comm. Math. Physics, 115
(1988).
1. Bott-Chern forms and analytic torsion, pp. 49-78
II. Direct images and Bott-Chern forms, pp. 79-126
III. Quillen metrics on holomorphic determinants, pp. 301-350
[Bo 1] 1. B. BOST, "Conformal and holomorphic anomalies on Riemann
surfaces and determinant line bundles", to appear
[Bo 2] J. B. BOST, "Fonctions de Green-Arakelov, fonctions theta et
courbes de genre 2", to appear.
[Ch IJ T. CHINBURG, "An Introduction to Arakelov Intersection Theory",
in Arithmetic Geometry (edited by G. Cornell and J. Silverman),
Springer-Verlag, 1986, pp. 289-307.
[Ch2] T. CHINBURG, "Minimal Models for Curves over Dedekind rings",
in Arithmetic Geometry (edited by G. Cornell and 1. Silverman),
Springer-Verlag, 1986, pp. 309-311.
[C-H] M. CORNALBA and J. HARRIS, "Divisor classes associated to fami-
lies of stable varieties, with applications to the moduli space of
curves", to appear
[De] P. DELlGNE, Le Determinant de la Cohomologie, Contemporary
Mathematics Vol. 67, American Mathematical Society, 1987.
[D-M] P. DELIGNE and D. MUMFORD, "The irreducibility of the space of
curves of given genus", Inst. Hautes Etudes Sci. Publ. Math., 36
(1969), pp. 75-109.
[EGA] Elements de Geometrie Algebrique, by A. Grothendieck, Institut
des Hautes Etudes Scientifiques.
[Fa 1] G. FALTINGS, "Calculus on arithmetic surfaces", Ann. oj Math.
119 (1984), pp. 387-424.
[Fa 2] G. FALTINGS, "Enlichkeitssatze fur abelschen Varietaten uber
Zahlkorpern", Invent. Math., 73 (1983), pp. 349-366. See also
the translation in Arithmetic Geometry (edited by G. Cornell
and J. Silverman), Springer-Verlag, 1986.
[Fu] W. FULTON, Intersection Theory, Springer-Verlag, 1984.
[F-LJ W. FULTON and S. LANG, Riemann-Roch Algebra, Springer-Verlag,
1985.
[G-H] P. GRIFFITHS and 1. HARRIS, Algebraic Geometry, Wiley-Inter-
science, 1978.
[G-S] H. GILLET and C. SOULE, "Arithmetic intersection theory", to
appear.
[Ha 1] R. HARTSHORNE, Algebraic Geometry, Springer-Verlag, 1977.
[Ha2J R. HARTSHORNE, Residues and Duality, Lecture Notes in Mathe-
matics, Vol. 20, Springer-Verlag, 1966.
[Ho] L. HORMANDER, "The spectral function of an elliptic operator",
Acta Math., 121 (1968), pp. 193-218.
[Hr 1] P. HRILJAC, "Heights and Arakelov's intersection theory", Amer. J.
Math., 107, No.1 (1985), pp. 23-38.
REFERENCES 181

[Hr 2] P. HRILJAC, "Splitting Fields of Principal Homogeneous Spaces",


1984-1985 New York Number Theory Seminar (edited by Chud-
novsky et al.), Lecture Notes in Mathematics, Vol. 1240,
Springer-Verlag, 1987, pp. 214-229.
[Ke] G. KEMPF, "On the geometry of a theorem of Riemann", Ann. of
Math., 98 (1973), pp. 178-185.
[KI] S. KLEIMAN, "Relative duality for quasi-coherent sheaves", Com-
positio Math., 41 (1980), pp. 39-60.
[K-M] F. KNUDSEN and D. MUMFORD, "The projectivity of the moduli
space of stable curves !", Math. Scand., 39 (1976), pp. 19-55.
[K-L] D. KUBERT and S. LANG, Modular Units, Springer-Verlag, 1981.
[Ku] E. KUNZ, Kiihler Differentials, Vieweg Verlag, 1986.
[La 1] S. LANG, Fundamentals of Diophantine Geometry, Springer-Verlag,
1983.
[La 2] S. LANG, Elliptic Functions, Addison-Wesley, 1973; Second Edi-
tion, Springer-Verlag, 1987.
[La 3] S. LANG, Introduction to Algebraic and Abelian Functions, Second
Edition, Springer-Verlag, 1982.
[La 4] S. LANG, Differential Manifolds, Addison-Wesley, 1972; reprinted
by Springer-Verlag, 1985.
[La 5] S. LANG, Algebraic Number Theory, Addison-Wesley 1970; re-
printed by Springer-Verlag, 1986.
[LiJ S. LICHTENBAUM, "Curves over discrete valuation rings", Amer. J.
Math. XC, No.2 (1968), pp. 380-405.
[Lip 1] 1. LIPMAN, "Rational singularities, with applications to algebraic
surfaces and unique factorizations", Inst. Hautes Etudes Sci.
Publ. Math., 36 (1969), pp. 195-279.
[Lip 2J J. LIPMAN, "Dualizing sheaves, differentials and residues on alge-
braic varieties", Asterisque, 117 (1984).
[Lip 3J 1. LIPMAN, Residues and Traces of Differential Forms via Hochs-
child Homology, Contemporary Mathematics, Vol. 61, American
Mathematical Society, 1987.
[Mat IJ H. MATSUMURA, Commutative Algebra, Second Edition, Benjamin,
1980.
[Mat 2J H. MATSUMURA, Commutative Ring Theory, Cambridge University
Press, 1986.
[MBJ L. MORET-BAILLY, "Metriques permises", Seminaire sur les pin-
ceaux Arithmetiques: La Conjecture de Mordell, Asterisque, 127
(1985).
[McK-S] H. McKEAN and I. SINGER, "Curvature and the eigenvalues of the
Laplacian", J. Differential Geom., 1 (1967), pp. 43-69.
[Mi] S. MINAKSHISUNDARAM, "Eigenfunctions on Riemannian mani-
folds", J. Indian Math. Soc., 17 (1953), pp. 159-165.
[Mi-P] S. MINAKSHISUNDARAM and A. PLEIJEL, "Some properties of the
eigenfunctions of the Laplace operator on Riemannian mani-
folds," Canadian J. Math., 1 (1949), pp. 242-256.
[Miy IJ Y. MIYAOKA, "On the Chern numbers of surfaC{:s of general type",
Invent. Math. 42 (1977), pp. 225-237.
182 REFERENCES

[Miy 2] Y. MTYAOKA, "The maximal number of quotient singularities on


surfaces with given numerical invariants", Math. Ann. 268
(1984), pp. 154-17l.
[Mu] D. MUMFORD, Abelian Varieties, Tata Institute, Oxford University
Press, 1970.
CPa 1] A. N. PARSHIN, "Algebraic curves over function fields !", Izv.
Akad. Nauk. 32 (1968), pp. 1191-1219.
CPa 2] Y. N PARSHIN, "The Bogomolov-Miyaoka-Yau inequality for the
arithmetical surfaces and its applications," Seminaire de Theorie
des Nombres, Paris, 1986-87.
M. PROTTER and H. F. WEINBERGER, Maximum Principle in Differ-
ential Equations, Prentice-Hall, 1967.
[Sh] I. SHAFAREVICH, On Minimal Models and Birational Transforma-
tions, Tata Institute, Bombay, 1966.
[Sz 1] L. SZPTRO, "Presentation de la theorie d'Arakelov", Current
Trends in Arithmetical Algebraic Geometry, Proceedings of a
Summer Conference held August 18-24, 1985, Contemporary
Mathematics, Vol. 67, American Mathematical Society, 1987, pp.
279-293.
[Sz 2] L. SZPIRO, Small Points and Torsion Points, Contemporary Mathe-
matics, Vol. 58, American Mathematical Society, 1986, pp.
251-260.
[Sz 3] L. SZPTRO, "Proprietes numeriques du faisceau dualisant relatif",
Semina ire sur les pinceaux de courbes de genre au moins deux,
Asterisque, 86 (1981).
[Sz 4] L. SZPTRO, Seminaire sur les pinceaux arithmetiques: La conjecture
de Mordell, Asterisque, 127 (1985).
[Ta] M. TAYLOR, Pseudo Differential Operators, Princeton University
Press, 1981.
[V] P. VOJTA, Diophantine Approximations and Value Distribution
Theory, Lecture Notes in Mathematics, Vol. 1239, Springer-
Verlag, 1987.
[Xi] XTAO GANG. "Fibered algebraic surfaces with low slope", Math.
Ann. 276 (1987), pp. 449-466.
Frequently Used Symbols

CH: Chow group 52


CH(X, A): Arakelov Chow group 76
c 1 (p): Chern form of metric 11
y(E): volume discrepancy 106
yv(f): component in Arakelov divisor of a function 8, 75
b(X): Faltings delta 144
by,b: : number of double points on geometric fiber 158
d, de: differential operators 10
D: discriminant 93, 111
d: log D 93, 99, 100, 111
dEfY 93
d;.: log A-discriminant 99, 111
d oo 99
1
d(P): [F(P): Q] log DF(p)fQ 157
deg F : F-degree 73
degz(D): local degree of a divisor 80
deg.(2', p): degree of a metrized line sheaf 81
deg z(2', p): global degree 81
Divo: divisors of degree zero 61, 77
Divy: fibral divisors 58
(D. E): intersection symbol 72
<D, E): Neron symbol 67, 77
g: genus or Green's function 21
X(2', p): Faltings Euler characteristic 103, 112
ix: intersection number 55
A, AD: Neron or Weil function 1, 3
184 FREQUENTL Y USED SYMBOLS

N v: local degree [Fv: Qv] 71


j1: canonical volume form 28
C(J*: direct image 53
<1>: canonical 2-form on X x X 29
Pic(X, 2): Arake10v Picard group 85
Rat(X, 2) 76
P: metric 5
P;",D: metric on (!J(D) defined by Weil function 7, 143
e: theta divisor 139
v(z): -log Izl" 1
vF(z) = -log Ilzllv 71
WEIY: Dedekind complementary module 93
WEIY: canonical sheaf or dualizing sheaf 89
Index

A Chern form 11, 33


Chow group 52, 76
Adjunction formula 88, 99, 115 Coleman forms 34
Admissible 85, 101, 140 Complementary module 93
Anti-holomorphic forms 34 Cornalba-Harris 162
Arakelov adjunction formula 99 Curve 50
Arakelov Chow group 76
Arakelov divisor 75, 83
Arakelov- Picard group 85 D
Arithmetic curve 80, 102
Arithmetic genus 117 dd c lemma 17
Arithmetic surface 49 Dedekind complementary module 93
Artin's theorem 122 Degree, of a curve over another 57
Associated metric 85 of an Arakelov divisor 84
of a Cartier divisor 80
of a divisor over a field 73
B of a line sheaf 81
Base change 62, 65, 122-127 of a Weil divisor on a curve 59, 73
Bernoulli polynomials 43, 150 Determinant 104, 133
Different 93
Differential sheaf 90
c Dimension 50
Canonical Arakelov divisor 101 Direct image 53
Canonical class inequality 156, .160 Discriminant 93, 97, 99, 111, 113, 157
Canonical form 28, 33, 143 Distributions 26, 47
Canonical Green function 29 Divisor 50, 59, 75, 83
Canonical height pairing 77 Divisors of degree zero 61, 67, 77
Canonical sheaf 89, 91 Dualizing module 93
base change 126
positivity 128 E
Cartier divisor 2, 50, 59
CH 52,76 Effective divisor 49, 118
Chern class 12 Eigenvalue expansion 150
186 INDEX

Elkies' theorem 150 Local complete intersection 86, 88


Elliptic curves 44, 130 Local degree 81
Euler characteristic 103, 104, 106, Logarithmic discriminant 93, 99, 100,
112, 115 157
base change 125
M
F
Metric on determinant (volume) 105
F AL conditions 110, 140 associated with Arakelov divisor 85
Faltings (j 144, 165 on line sheaf 5, 80
Faltings height 159 and Weil function 7, 85, 143
Faltings metric 110, 140, 142, 147 Metrized line sheaf 5
Faltings Riemann-Roch 113 on an arithmetic curve 81
F-degree 73 on an arithmetic surface 84
Fibral divisor 52 Minimal desingularization 122
Fibral intersections 60, 75, 79, 122,
130
N
G Negative intersection form 61, 77
Negative Neron symbol 77
Geometric fiber 121
Neron divisor 2
Global degree 81
Neron family 3
Global Neron symbol 77
Neron function 3, 69
GR conditions 21
on elliptic curve 43
Green's function 21. 29, 33 Neron symbol 67, 77
on elliptic curve 44
Normalized Haar measure 103
Green's operator 28 Normalized volume form 21, 103
NF conditions 3
H NS conditions 67
Haar measure 103, 105
Harmonic forms 34, 39 o
Height 77, 156
Height inequalities 158 Order 49
Height pairing 77 Orthogonalization 62, 79
Hodge Index 77
Horizontal divisor 52 P
Parshin 155, 164-167
I Parshin inequality 164
Intersection numbers 55, 64, 71, 72, Pic(X, A.) 85
74, 77, 86 Poincare sheaf 136
base change 65 Positive form or metric 12, 21
Inverse image 63, 64, 65, 77, 82, 125 Positivity of canonical sheaf 128
Irreducible divisor 49 Prime divisor 49
Principal divisor 50
Principal Neron divisor 3
K
Projection formula 64
Kunz-Waldi 94 Proper transform 63

L R
Laplacian 27 Rank 57
Limited family 165 Rat 52, 76
Line sheaf 1 Rational equivalence 52
INDEX 187

Rational singularity 125 T


Regular imbedding 86
Regular morphism 90 Tensor product of metrics 6
Relative differential divisor 97 Theta divisor 139
Relative Euler characteristic 104 Theta functions 143
Relative logarithmic discriminant 93, Trivial metric 83
157
Reproduction formula 31 V
Residue isomorphism 93
Residue theorem 94 Variety 49
Riemann-Roch on curves 104 115 Vector sheaf 1
Riemann-Roch Faltings 113 ' Vertical divisor 52
Volume, on a vector space 105
of fundamental domain 106
Volume discrepancy 106
s Volume exact 106, 111
Volume form 21
Self-intersection 58
Semipositive 12
Semistable 121
w
SS conditions 121 Weil divisor 49, 59
Stable 157 Weil function 1
Standard metric 6, 13, 83 and metric 7
Subvariety 49
Support 52 X
Symmetry 24
Szpiro 118, 128, 167 Xiao Gang 162, 178

You might also like