21.10.2014 Views

brauer groups, tamagawa measures, and rational points

brauer groups, tamagawa measures, and rational points

brauer groups, tamagawa measures, and rational points

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

BRAUER GROUPS,<br />

TAMAGAWA MEASURES,<br />

AND RATIONAL POINTS<br />

ON ALGEBRAIC VARIETIES<br />

Jörg Jahnel


Jörg Jahnel<br />

Mathematisches Institut, Bunsenstraße 3–5, D–37073 Göttingen, Germany.<br />

E-mail : jahnel@uni-math.gwdg.de<br />

Url : http://www.uni-math.gwdg.de/jahnel<br />

2000 Mathematics Subject Classification. —<br />

11G35, 14F22, 16K50, 11-04, 14G25, 11G50.<br />

Acknowledgements. I wish to acknowledge with gratitude my debt to Y. Tschinkel. Most of the work described in this<br />

book was initiated by his numerous mathematical questions. During the years he spent at Göttingen, he always shared<br />

his ideas in an extraordinarily generous manner.<br />

I further wish to express my deep gratitude to my friend <strong>and</strong> collegue Stephan Elsenhans. He influced this book in<br />

many ways, directly <strong>and</strong> indirectly. It is no exaggeration to say that most of what I know about computer programming,<br />

I learned from him. The experiments which are described in Part C were carried out as a joint work. I am also indebted<br />

to Stephan for proofreading.<br />

Special thanks go to my parents, Regina <strong>and</strong> Alfred Jahnel, for their unflaggingbelief that, despite their incomprehension<br />

about what I do, I must be doing that very well. Finally, I would like to thank all the people who constantly supported<br />

<strong>and</strong> encouraged me over the years. I risk doing someone a disservice by not naming all of them here, but plead<br />

paucity of space. Let me mention particularly Ulrich Stuhler, H.-S. Holdgrün, Thomas Fröhlich, Stefanie Schmidt,<br />

<strong>and</strong> Carsten Thiel.<br />

The computerpart of the workdescribed in this bookwas executed on theSun Fire V20z Serversof the GaußLaboratory<br />

for Scientific Computing at the Göttingen Mathematical Institute. The author is grateful to Y. Tschinkel for the<br />

permission to use these machines as well as to the system administrators for their support.<br />

J. J.<br />

Göttingen, Lower Saxony,<br />

September 24, 2008


BRAUER GROUPS, TAMAGAWA MEASURES,<br />

AND RATIONAL POINTS<br />

ON ALGEBRAIC VARIETIES<br />

Abstract. — We study existence <strong>and</strong> asymptotics of <strong>rational</strong> <strong>points</strong> on algebraic<br />

varieties of Fano <strong>and</strong> intermediate type. In the first part, we study the various<br />

versions of the Brauer group. We explain why a Brauer class may serve as an<br />

obstruction to weak approximation or even to the Hasse principle. In the second<br />

part, we discuss to some extent the concept of a height <strong>and</strong> formulate Manin’s<br />

conjecture on the asymptotics of <strong>rational</strong> <strong>points</strong> on Fano varieties. The final<br />

part describes numerical experiments.<br />

Habilitationsschrift<br />

vorgelegt von<br />

Jörg Jahnel<br />

aus Eisenberg<br />

Göttingen, Sommer 2008


CONTENTS<br />

Table of contents . . ....................................................<br />

iii<br />

Notation <strong>and</strong> fundamental conventions . . .............................. vii<br />

Notation <strong>and</strong> conventions . . ............................................ vii<br />

References <strong>and</strong> Citation . . .............................................. ix<br />

Introduction . . ..........................................................<br />

xi<br />

Part A. The Brauer group . . ............................................ 1<br />

I. The Brauer-Severi variety associated with a central simple algebra . . 3<br />

1. Introductory remarks . . .............................................. 3<br />

2. Non-abelian group cohomology . . ................. . . ............... 5<br />

3. Galois descent . . . ......................... .......................... 8<br />

4. Central simple algebras <strong>and</strong> non-abelian H 1 . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

5. Brauer-Severi varieties <strong>and</strong> non-abelian H 1 . . ........................ 28<br />

6. Central simple algebras <strong>and</strong> Brauer-Severi varieties . . . ............... 35<br />

7. Functoriality . . ........................... ........................... 39<br />

8. The functor of <strong>points</strong> . . .............................................. 53<br />

II. On the Brauer group of a scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

1. Azumaya algebras . . ........................ ........................ 61<br />

2. The Brauer group . . .................................................. 65<br />

3. The cohomological Brauer group . . .................................. 66<br />

4. The relation to the Brauer group of the function field . . . . . . . . . . . . . . 71<br />

5. The Brauer group <strong>and</strong> the cohomological Brauer group . . . . . . . . . . . . 74<br />

6. Orders in central simple algebras . . ................. ................. 76<br />

7. The Theorem of Ausl<strong>and</strong>er <strong>and</strong> Goldman . . . . . . . . . . . . . . . . . ......... 84


iv<br />

CONTENTS<br />

8. Examples . . .......................................................... 86<br />

III. An application: The Brauer-Manin obstruction . . .................. 99<br />

1. Adelic <strong>points</strong> . . ...................................................... 99<br />

2. The Brauer-Manin obstruction . . .................................... 103<br />

3. Technical lemmata . . ........................ ........................ 106<br />

4. Computing the Brauer-Manin obstruction – The general strategy . . 110<br />

5. The examples of Mordell . . .......................................... 113<br />

6. The “first case” of diagonal cubic surfaces . . .......................... 128<br />

Part B. Heights . . ...................................................... 147<br />

IV. The concept of a height . . ............................................ 149<br />

1. The naive height on the projective space overÉ. . . . . . . . . . . . . . . . . . . . 149<br />

2. Generalization to number fields . . . ................ ................. 151<br />

3. Geometric interpretation . . .......................................... 155<br />

4. Basic arithmetic intersection theory . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . 160<br />

5. An intersection product on singular arithmetic varieties . . . . . . . . . . . . 165<br />

6. The arithmetic Hilbert-Samuel formula . . ............................ 169<br />

7. The adelic Picard group . . ............................................ 173<br />

V. On the distribution of small <strong>points</strong> on abelian <strong>and</strong> toric varieties . . 189<br />

1. Introduction . . ...................................................... 189<br />

2. The dynamics of f . . ................................................ 192<br />

3. Perturbing almost semiample metrics . . .............................. 199<br />

4. Equidistribution . . ......................... ......................... 204<br />

VI. Conjectures on the asymptotics of <strong>points</strong> of bounded height . . .... 211<br />

1. A heuristic . . ........................................................ 211<br />

2. The conjecture of Lang . . ............................................ 214<br />

3. The conjecture of Batyrev <strong>and</strong> Manin . . .............................. 216<br />

4. The conjecture of Manin . . .......................................... 220<br />

5. Peyre’s constant . . .................................................... 224<br />

Part C. Numerical experiments . . ...................................... 241<br />

VII. Points of bounded height on cubic <strong>and</strong> quartic threefolds . . ...... 243<br />

1. Introduction — Manin’s Conjecture . . ................................ 243<br />

2. Computing the Tamagawa number . . ................................ 247


CONTENTS<br />

v<br />

3. On the geometry of diagonal cubic threefolds . . . . . . . . . . . . . . . . . . . . . . 252<br />

4. Accumulating subvarieties . . ........................................ 254<br />

5. Results . . ............................................................ 259<br />

VIII. On the smallest point on a diagonal quartic threefold . . . . . . . . . . 265<br />

1. A computer experiment . . .......................................... 265<br />

2. A negative result . . .................................................. 268<br />

3. The fundamental finiteness property . . .............................. 270<br />

IX. General cubic surfaces . . ............................................ 281<br />

1. Rational <strong>points</strong> on cubic surfaces . . .................................. 281<br />

2. Background . . ........................................................ 284<br />

3. The Galois operation on the 27 lines . . .............................. 287<br />

4. Computation of Peyre’s constant . . .................................. 290<br />

5. Numerical Data . . .................................................... 292<br />

6. A concrete example . . ................................................ 297<br />

X. On the smallest point on a diagonal cubic surface . . ................ 301<br />

1. Introduction . . ...................................................... 301<br />

2. The factors α <strong>and</strong> β . . ....................... . ...................... 305<br />

3. Splitting the Picard group . . .......................................... 306<br />

4. A technical lemma . . ................................................ 310<br />

5. A negative result . . .................................................. 311<br />

6. The fundamental finiteness property . . .............................. 316<br />

XI. The Diophantine Equation x 4 + 2y 4 = z 4 + 4w 4 . . . . . . . . . . . . . . . . . . 329<br />

1. Introduction . . ...................................................... 329<br />

2. Congruences . . ...................................................... 330<br />

3. Naive methods . . .................................................... 332<br />

4. An algorithm to efficiently search for solutions . . .................... 332<br />

5. General formulation of the method . . ................................ 335<br />

6. Improvements I – More congruences . . .............................. 336<br />

7. Improvements II – Adaption to our hardware . . . . . . . . . . . . . . . . . . . . . . 341<br />

8. The solution found . . ................................................ 348<br />

XII. New sums of three cubes . . ........................................ 349<br />

1. Introduction . . ...................................................... 349<br />

2. Elkies’ method . . .......................... .......................... 350<br />

3. Implementation . . .................................................... 350


vi<br />

CONTENTS<br />

4. Results . . ............................................................ 351<br />

Appendix . . .............................................................. 353<br />

1. A script in GAP . . .................................................... 353<br />

2. The list . . ............................................................ 356<br />

Bibliography . . .......................................................... 363<br />

Index . . ................................. ................................. 383


NOTATION AND FUNDAMENTAL<br />

CONVENTIONS<br />

Notation <strong>and</strong> conventions<br />

We follow st<strong>and</strong>ard notation <strong>and</strong> conventions from Algebra, Algebraic Number<br />

Theory, <strong>and</strong> Algebraic Geometry. More precisely:<br />

i) We writeÆ,,É,Ê, <strong>and</strong>for the sets of natural numbers, integers, <strong>rational</strong><br />

numbers, real numbers, <strong>and</strong> complex numbers, respectively.<br />

ii) For a group G <strong>and</strong> elements σ 1 , . . . , σ n ∈ G, we denote the subgroup<br />

generated by σ 1 , . . . , σ n by 〈σ 1 , . . . , σ n 〉 ⊆ G.<br />

If G is abelian then G n ⊆ G is the subgroup consisting of all elements of torsion<br />

dividing n.<br />

iii) If a group G operates on a set M then M G denotes the invariants. We write<br />

M σ instead of M 〈σ〉 .<br />

iv) All rings are assumed to be associative.<br />

v) If R is a ring then R op denotes the opposite ring. I.e., the ring that coincides<br />

with R as an abelian group but in which one has xy = z when one had yx = z<br />

in R.<br />

vi) For R a ring with unit, R ∗ denotes the multiplicative group of invertible<br />

elements in R.<br />

vii) All homomorphisms between rings with unit are supposed to respect the<br />

unit elements.<br />

viii) By a field, we always mean a commutative field. I.e., a commutative ring<br />

with unit every non-zero element of which is invertible. If K is a field then K sep<br />

<strong>and</strong> K denote a fixed separable closure <strong>and</strong> a fixed algebraic closure, respectively.<br />

ix) A ring with unit every non-zero element of which is invertible is called a<br />

skew field.<br />

x) If R is a commutative ring with unit then an R-algebra is always understood<br />

to be a ring homomorphism j : R → A the image of which is contained in the<br />

center of A. An R-algebra j : R → A is denoted simply by A when there seems


viii<br />

NOTATION AND FUNDAMENTAL CONVENTIONS<br />

to be no danger of confusion. An R-algebra being a skew field is also called a<br />

division algebra.<br />

xi) If σ : R → R is an automorphism of R then A σ denotes the R-algebra<br />

R −→ σ<br />

R −→ j<br />

A. If M is an R-module then we put M σ := M ⊗ R R σ . M σ is an<br />

R σ -module as well as an R-module.<br />

xii) All central simple algebras are assumed to be finite dimensional over a<br />

base field.<br />

xiii) For K a number field, we write O K to denote the ring of integers in K.<br />

If ν ∈ Val(K ) is a non-archimedean valuation ν ∈ Val(K ) then the ν-adic<br />

completion of K is denoted by K ν <strong>and</strong> its ring of integers by O Kν .<br />

In the particular case that K =É, we denote by ν p the normalized p-adic<br />

valuation corresponding to a prime number p.<br />

xiv) For R a commutative ring, we denote by SpecR the affine scheme constituted<br />

by its spectrum.<br />

xv) The projective space of relative dimension n over a scheme X will be denoted<br />

by P n X . We omit the subscript when there is no danger of confusion.<br />

xvi) If X is a scheme over a scheme T <strong>and</strong> Y is a T -scheme then we also write<br />

X Y for the fiber product X × T Y. If Y = SpecR is affine then write X R instead<br />

of X SpecR .<br />

xvii) For X a scheme over a scheme T , we denote by X t the fiber of X<br />

over t ∈ T .<br />

If C is a scheme over the integer ring O Kν of the completion of the number<br />

field K with respect to the valuation ν then we write C ν for the special fiber.<br />

In the particular case ν = ν p , we write C p instead of C νp .<br />

If C is a scheme over the integer ring O of a number field K then we use same<br />

the notation, not mentioning the base change to O Kν .<br />

xviii) For R any commutative ring, A a commutative R-algebra, <strong>and</strong> X an<br />

R-scheme, a morphism x : SpecA → X of R-schemes is also called an A-valued<br />

point on X. If A is a field then we also adopt more conventional language <strong>and</strong><br />

speak of a point defined or <strong>rational</strong> over A. The set of all A-valued <strong>points</strong> on X<br />

will be denoted by X (A).<br />

xix) If C is a scheme over a valuation ring O <strong>and</strong> x ∈ C (O) then the reduction<br />

of x is denoted by x.


NOTATION AND FUNDAMENTAL CONVENTIONS<br />

ix<br />

References <strong>and</strong> Citation<br />

When we refer to a definition, proposition, theorem, etc., in the same chapter<br />

then we simply rely on the corresponding numbering within the chapter.<br />

Otherwise, we add the number of the chapter.<br />

For the purpose of citation, the articles <strong>and</strong> books being used are encoded in<br />

the manner specified by the bibliography. In addition, we mostly give the<br />

number of the relevant section <strong>and</strong> subsection or the number of the definition,<br />

proposition, theorem, etc. Normally, we do not mention page numbers.


INTRODUCTION<br />

Here, in the midst of this sad <strong>and</strong> barren l<strong>and</strong>scape of the Greek accomplishments<br />

in arithmetic, suddenly springs up a man with youthful energy:<br />

Diophantus. Where does he come from, where does he go to? Who were his<br />

predecessors, who his successors? We do not know. It is all one big riddle.<br />

He lived in Alex<strong>and</strong>ria. If a conjecture were permitted, I would say he was<br />

not Greek; . . . if his writings were not in Greek, no-one would ever think<br />

that they were an outgrowth of Greek culture . . . .<br />

HERMANN HANKEL (1874, translated by N. Schappacher)<br />

Diophantine equations have a long history. More than two thous<strong>and</strong> years ago,<br />

Diophantus of Alex<strong>and</strong>ria considered, among many others, the equations<br />

x 2 + y 2 = z 2 ,<br />

(∗)<br />

y(6 − y) = x 3 − x ,<br />

<strong>and</strong><br />

y 2 = x 2 + x 4 + x 8 .<br />

In Diophantus’ book “Arithmetica”, we find the formula<br />

(<br />

(p 2 − q 2 )λ, 2pqλ, (p 2 + q 2 )λ ) (†)<br />

which generates infinitely many solutions of (∗). For the second <strong>and</strong><br />

third of the equations mentioned, Diophantus gives particular solutions,<br />

namely (1/36, 1/216) <strong>and</strong> (1/2, 9/16), respectively.<br />

In general, a polynomial equation in several indeterminates where solutions<br />

are sought in integers or <strong>rational</strong> numbers, is called a Diophantine equation,<br />

in honour of Diophantus. Diophantus himself was interested in solutions in<br />

positive integers or positive <strong>rational</strong> numbers. Contrary to the point of view<br />

usually adopted today, he did not accept negative numbers.<br />

It is remarkable that algebro-geometric methods have often been fruitful in order<br />

to underst<strong>and</strong> a Diophantine equation. For example, there is a simple geometric<br />

idea behind formula (†).


xii<br />

INTRODUCTION<br />

Indeed, since the equation is homogeneous, it suffices to look for solutions<br />

of X 2 + Y 2 = 1 in <strong>rational</strong>s. This equation defines the unit circle. For every<br />

t ∈Ê, there is the line “x = −ty + 1” going through the point (1, 0).<br />

An easy calculation shows that the second point where this line meets the<br />

unit circle is given by ( 1 − t<br />

2<br />

2t<br />

)<br />

1 + t 2, . (‡)<br />

1 + t 2<br />

As every point on the unit circle may be connected with (1, 0) by a line,<br />

one sees that the parametrization (‡) yields every <strong>rational</strong> point on the circle<br />

(except for (−1, 0) for which the form of line equation given is not adequate).<br />

Consequently, formula (†) delivers essentially every solution of equation (∗),<br />

a fact which was seemingly not known to the ancient mathematicians. The morphism<br />

P 1 −→ P 2<br />

(p : q) ↦→ ( (p 2 − q 2 ) : 2pq : (p 2 + q 2 ) )<br />

provides a <strong>rational</strong> parametrization of the plane conic C given by the equation<br />

x 2 + y 2 = z 2 in P 2 . Essentially the same method works for every conic in<br />

the plane. Further, it may be extended to several classes of singular curves of<br />

higher degree.<br />

Every Diophantine equation defines an algebraic variety X in an affine or projective<br />

space. There is a one-to-one correspondence between solutions of the<br />

Diophantine equation <strong>and</strong>É-<strong>rational</strong> <strong>points</strong> on X. We will prefer geometric<br />

language to number theoretic throughout this book.<br />

The cases when there is an obvious <strong>rational</strong> parametrization are in some sense<br />

best possible. But even when there is nothing like that, algebraic geometry often<br />

yields a guideline of which behaviour to expect, whether there will be no, a few<br />

or many solutions.<br />

The Kodaira classification distinguishes between Fano varieties, varieties of intermediate<br />

type, <strong>and</strong> varieties of general type (at least under the additional<br />

assumption that X is non-singular). It does not use any specifically arithmetic<br />

information, but only information about X as a complex variety. Nevertheless,<br />

there is overwhelming evidence for a strong connection between the<br />

classification of X according to Kodaira <strong>and</strong> its set of <strong>rational</strong> <strong>points</strong>.<br />

To make a vague statement, on a Fano variety, there are infinitely many <strong>rational</strong><br />

<strong>points</strong> expected while, on a variety of general type, there are only finitely many<br />

<strong>rational</strong> <strong>points</strong> or even none at all. More precisely, there is the conjecture<br />

that, on a Fano variety, there are always infinitely many <strong>rational</strong> <strong>points</strong> after a<br />

suitable finite extension of the ground field. On the other h<strong>and</strong>, for varieties of<br />

general type, there is the conjecture of Lang. It states that there are only finitely


INTRODUCTION<br />

xiii<br />

many <strong>rational</strong> <strong>points</strong> outside the union of all closed subvarieties which are not<br />

of general type.<br />

Another method to analyze a Diophantine equation is given by congruences.<br />

Kurt Hensel provided a more formal framework for this method by his invention<br />

of the p-adic numbers. As one is working over local fields, this might be<br />

called the local method.<br />

Consider, for example, the Diophantine equation<br />

x 3 + 7y 3 + 49z 3 + 2u 3 + 14v 3 + 98w 3 = 0 . (§)<br />

It has no solution inÉ6 except for (0, 0, 0, 0, 0, 0). This may be seen by a repeated<br />

application of an argument modulo 7. A more formal reason is the fact that the<br />

projective algebraic variety defined by (§) has no point defined overÉ7.<br />

One might ask whether or to what extent solvability overÉp for every prime<br />

number p together with solvability in real numbers implies the existence of<br />

a <strong>rational</strong> solution. This question has been very inspiring for research over<br />

many decades. As early as 1785, A.-M. Legendre gave an affirmative answer for<br />

equations of the type<br />

q(x, y, z) = 0<br />

where q is a ternary quadratic form. Legendre’s result was generalized<br />

to quadratic forms in arbitrarily many variables by Hasse <strong>and</strong> Minkowski.<br />

The term “Hasse principle” was coined to describe the phenomenon.<br />

A totally different sort of examples where the Hasse principle is valid is provided<br />

by the circle method originally developed by G. H. Hardy <strong>and</strong> J. E. Littlewood.<br />

The circle method uses tools from complex analysis to study the asymptotics<br />

of the number of <strong>points</strong> of bounded height on complete intersections in a very<br />

high-dimensional projective space. It provides an asymptotic formula <strong>and</strong> an<br />

error term. The main term is of the form<br />

τB n+1−d 1− ... −d r<br />

for a complete intersection of multidegree (d 1 , . . . , d r ) in P n . The reader might<br />

want to consult [Va] for a description of the method <strong>and</strong> references to the<br />

original literature.<br />

The exponent of the main term allows a beautiful algebro-geometric interpretation.<br />

The anticanonical sheaf on a complete intersection of multidegree<br />

(d 1 , . . . , d r ) in P n is precisely O(n + 1 − d 1 − . . . − d r )| X . This means,<br />

when working with an anticanonical height instead of the naive height, the<br />

circle method proves linear growth for theÉ-<strong>rational</strong> <strong>points</strong>.<br />

The coefficient τ of the main term is a product of p-adic densities together with<br />

a factor corresponding to the archimedean valuation.


xiv<br />

INTRODUCTION<br />

Unfortunately, it is necessary to make very restrictive assumptions on the number<br />

of variables in comparison with the degrees of the equations. These assumptions<br />

on the dimension of the ambient projective space are needed in order to<br />

ensure that the provable error term is smaller than the main term. On might,<br />

nevertheless, hope that there is a similar asymptotic under much less restrictive<br />

conditions. This is the origin of Manin’s conjecture.<br />

However, aswasobservedbyJ. Franke, Yu. I. Manin, <strong>and</strong>Y. Tschinkel[F/M/T],<br />

the main term as described above is not compatible with the formation of<br />

direct products. Already on a variety as simple as P 1 × P 1 , the growth of the<br />

number of theÉ-<strong>rational</strong> <strong>points</strong> is actually asymptotically equal to τB log B.<br />

This may be seen by a calculation which is completely elementary.<br />

Thus, in general, the asymptotic formula has to be modified by a log-factor.<br />

Franke, Manin, <strong>and</strong> Tschinkel suggest the factor log rkPic(X )−1 B <strong>and</strong> prove that<br />

this factor makes the asymptotic formula compatible with direct products.<br />

Furthermore, it turns out that the coefficient τ has to be modified<br />

when rkPic(X ) > 1. There appears an additional factor which is today<br />

called α(X ). This factor is defined by a beautiful yet somewhat mysterious<br />

elementary geometric construction.<br />

Another problem is that the Hasse principle does not hold universally. Consider<br />

the following elementary example which was given by C.-E. Lind in 1940.<br />

Lind [Lin] dealt with the Diophantine equation<br />

2u 2 = v 4 − 17w 4<br />

defining an algebraic curve of genus 2. It is obvious that this equation is nontrivially<br />

solvable in reals <strong>and</strong> it is easy to check that it is non-trivially solvable<br />

inÉp for every prime number p.<br />

On the other h<strong>and</strong>, there is no solution in <strong>rational</strong>s except for (0, 0, 0).<br />

Indeed, assume the contrary. Then, there is a solution in integers such<br />

that gcd(u, v, w) = 1. For such a solution, one clearly has ∤u. Since 2 is a<br />

square but not a fourth power modulo 17, we conclude that ( u<br />

17)<br />

= −1. On the<br />

other h<strong>and</strong>, for every odd prime divisor p of u, one has v 4 −17w 4 ≡ 0 (mod p).<br />

This shows ( ) (<br />

17<br />

p = 1. By the low of quadratic reciprocity, p<br />

17)<br />

= 1. Altogether,<br />

( u<br />

17)<br />

= 1 which is a contradiction.<br />

One might argue that this example is not too interesting since, on a curve of<br />

genus g ≥ 2, there can be only finitely manyÉ-<strong>rational</strong> <strong>points</strong>. Thus, it might<br />

happen that there are none of them without any particular reason.<br />

However, several other counterexamples to the Hasse principle had been invented.<br />

Some of them were Fano varieties. For example, Sir P. Swinnerton-<br />

Dyer[S-D62] <strong>and</strong> L.-J. Mordell [Mord] (cf. Chapter III, Section 5) constructed


INTRODUCTION<br />

xv<br />

examples of cubic surfaces violating the Hasse principle. A few years later,<br />

J. W. S. Cassels <strong>and</strong> M. J. T. Guy [Ca/G] as well as A. Bremner [Bre] even found<br />

isolated examples of diagonal cubic surfaces showing that behaviour. Typically,<br />

the proofs were a bit less elementary than Lind’s in that sense that they<br />

required not the quadratic but the cubic or biquadratic reciprocity low.<br />

In the late sixties of the 20th century, Yu. I. Manin [Man] made the remarkable<br />

discovery that all the known counterexamples to the Hasse principle could<br />

be explained in a uniform manner. There was actually a class α ∈ Br(X ) in<br />

the Brauer group of the underlying algebraic variety responsible for the lack<br />

ofÉ-<strong>rational</strong> <strong>points</strong>.<br />

This may be explained as follows. The Brauer group ofÉis relatively complicated.<br />

One has, by virtue of global class field theory,<br />

Br(SpecÉ) = ker ( M<br />

s :<br />

2/−→É/) .<br />

pprimeÉ/⊕ M<br />

σ : K→Ê1<br />

Here, s is just the summation. The summ<strong>and</strong>É/corresponding to the prime<br />

number p is nothing but Br(SpecÉp) while the last summ<strong>and</strong> is Br(SpecÊ).<br />

Let α ∈ Br(X ) be any Brauer class of a variety X overÉ. An adelic point<br />

x = (x ν ) ν∈Val(É) ∈ X (É)<br />

defines restrictions of α to Br(SpecÊ) <strong>and</strong> Br(SpecÉp) for each p. If the<br />

sum of all invariants is different from zero then, according to the computation<br />

of Br(SpecÉ), x may not be approximated byÉ-<strong>rational</strong> <strong>points</strong>.<br />

As α ∈ Br(X ) then “obstructs” x from being approximated by <strong>rational</strong> <strong>points</strong>,<br />

the expression Brauer-Manin obstruction became the general st<strong>and</strong>ard for this<br />

famous observation of Manin.<br />

In the counterexamples to the Hasse principle which were known to Manin in<br />

those days, one typically had a Brauer class the restrictions of which had a totally<br />

degenerate behaviour. For example, on Lind’s curve, there is a Brauer class α<br />

such that its restriction is independent of the choice of the adelic point. α restricts<br />

to zero in Br(SpecÊ) <strong>and</strong> Br(SpecÉp) for p ≠ 17 but non-trivially<br />

to Br(SpecÉ17). This suffices to show that there is noÉ-<strong>rational</strong> point on<br />

that curve.<br />

In general, the Brauer-Manin obstruction defines a subset X (É) Br ⊆ X (É)<br />

consisting of the adelic <strong>points</strong> which are not affected by the obstruction. At least<br />

for cubic surfaces, there is a conjecture of J.-L. Colliot-Thélène stating that<br />

X (É) Br is equal to the set of all adelic <strong>points</strong> which may actually be approximated<br />

byÉ-<strong>rational</strong> <strong>points</strong>.<br />

Thus,X(É) Br = ∅whileX(É) ≠ ∅meansthatX isaprovencounterexample<br />

to the Hasse principle. If X (É) Br X (É) then we have a counterexample to


xvi<br />

INTRODUCTION<br />

weak approximation. If Colliot-Thélène’s conjecture were true then one could<br />

say that all cubic surfaces which are counterexamples to the Hasse principle or<br />

to weak approximation are of this form.<br />

The Brauer group of an algebraic variety X over an algebraically nonclosed<br />

field k admits, according to the Hochschild-Serre spectral sequence,<br />

a canonical filtration in three steps. The first step is given by the image<br />

of Br(Speck) in Br(X ). Second, Br(X )/ Br(Speck), has a subgroup canonically<br />

isomorphic to H 1( Gal(k/k), Pic(X k<br />

) ) . The remaining subquotient is a subgroup<br />

of Br(X k<br />

) Gal(k/k) . It turns out that only the group H 1( Gal(k/k), Pic(X k<br />

) )<br />

is relevant for the Brauer-Manin obstruction.<br />

In the cases where the circle method is applicable, the Noether-Lefschetz Theorem<br />

shows that Pic(X ) =with trivial Galois operation. Consequently,<br />

H 1( Gal(k/k), Pic(X ) ) = 0 which is clearly sufficient for the absence<br />

of the Brauer-Manin obstruction. This coincides perfectly well with the observation<br />

that the circle method always proves equidistribution.<br />

By consequence, in a conjectural generalization of the results proven by the<br />

circle method, one can work with X (É) Br instead of X (É) without making<br />

any change in the proven cases. However, in the cases where weak<br />

approximation fails, this does not give the correct answer as was observed<br />

by D. R. Heath-Brown [H-B92a], in 1992. On a cubic surface such that<br />

H 1( Gal(k/k), Pic(X ) ) =/3<strong>and</strong> a non-trivial Brauer class excludes two<br />

thirds of the adelic <strong>points</strong>, there are nevertheless as many <strong>rational</strong> <strong>points</strong> as<br />

naively expected. Even more, E. Peyre <strong>and</strong> Y. Tschinkel [Pe/T] showed experimentally<br />

that if H 1( Gal(k/k), Pic(X ) ) =/3<strong>and</strong> the Brauer class does not<br />

exclude any adelic point then there are three times more <strong>rational</strong> <strong>points</strong> than expected.<br />

Correspondingly, in E. Peyre’s [Pe95a] definition of the conjectural<br />

constant τ, there appears an additional factor β(X ) := #H 1( Gal(k/k), Pic(X k<br />

) ) .<br />

This book is concerned with Diophantine equations from the theoretical <strong>and</strong><br />

experimental <strong>points</strong> of view. It is divided into three parts. Part A deals with the<br />

concepts of a Brauer group <strong>and</strong> their applications. In the first chapter, we begin<br />

with the simplest particular case <strong>and</strong> recall the Brauer group of a field. As an<br />

abstract group, we have Br(k) = H 2( Gal(k/k), k ∗) . Taking this as a definition,<br />

it is not hard to show that Br(X ) classifies two a priori rather different sorts<br />

of objects. Namely, on one h<strong>and</strong>, central simple algebras over k <strong>and</strong>, on the<br />

other, Brauer-Severi varieties over k.<br />

WerecallindetailthemachineryofGaloisdescent. Asanapplication, weexplain<br />

why both central simple algebras of dimension n 2 splitting over an extension<br />

field L as well as Brauer-Severi varieties of dimension n splitting over L are<br />

classified by H 1( Gal(L/K ), PGL n (L) ) . In Section I.7, there is given a functor


INTRODUCTION<br />

xvii<br />

from central simple algebras to Brauer-Severi varieties which yields the classical<br />

correspondence on objects.<br />

The second chapter considers A. Grothendieck’s generalization of the<br />

Brauer group to the case of an arbitrary scheme. We recall the concept of a<br />

sheaf of Azumaya algebras on a scheme <strong>and</strong> explain how such a sheaf of algebras<br />

gives rise to a class in the étale cohomology Br ′ (X ) := H 2 ét (X,m). This is what<br />

is called the cohomological Brauer group. On the other h<strong>and</strong>, a rather naive<br />

generalization of the definition for fields yields the concept of the Brauer group.<br />

One has Br(X ) ⊆ Br ′ (X ). In general, the two are not equal to each other.<br />

In Section II.7, we give a proof for the Theorem of Ausl<strong>and</strong>er <strong>and</strong> Goldman<br />

stating that Br(X ) = Br ′ (X ) in the case of a smooth surface. This result was<br />

originally shown in [A/G] before the actual invention of schemes. The proof<br />

of Ausl<strong>and</strong>er <strong>and</strong> Goldman was formulated in the language of Brauer <strong>groups</strong> for<br />

commutative rings. However, all the arguments given carry over immediately<br />

to the case of a scheme.<br />

The chapter is closed by computations of Brauer <strong>groups</strong> in particular examples.<br />

In the case of a variety over an algebraically non-closed field, we study the<br />

relationship of Br(X ) with H 1( Gal(k/k), Pic(X k<br />

) ) . We prove Manin’s formula<br />

expressing the latter cohomology group in terms of the Galois operation on<br />

a specific set of divisors. For smooth cubic surfaces, one may work with the<br />

classes given by the 27 lines.<br />

This leads to the result of Sir P. Swinnerton-Dyer [S-D93] that, for a smooth<br />

cubic surface, H 1( Gal(k/k), Pic(X k<br />

) ) is one of the <strong>groups</strong> 0,/2,/3,<br />

(/2) 2 , <strong>and</strong> (/3) 2 . Swinnerton-Dyer’s proof filled the entire article [S-D93]<br />

<strong>and</strong> was later modified by P. K. Corn in his thesis [Cor].<br />

We discovered that Swinnerton-Dyer’s result may be obtained in a manner which<br />

is rather brute force but very simple. The Galois group acting on the 27 lines on<br />

a smooth cubic surface is a subgroup of W (E 6 ). There are only 350 conjugacy<br />

classes of sub<strong>groups</strong> of W (E 6 ). We computed H 1( Gal(k/k), Pic(X k<br />

) ) in each of<br />

these cases using GAP. This took 28 seconds of CPU time.<br />

As an application of Brauer <strong>groups</strong>, the third chapter is concerned with the<br />

Brauer-Manin obstruction. We recall the notion of an adelic point <strong>and</strong> define the<br />

local <strong>and</strong> global evaluation maps. An adelic point x = (x ν ) ν∈Val(É) is “obstructed”<br />

from being approximated by <strong>rational</strong> <strong>points</strong> if the global evaluation map ev gives<br />

a non-zero value ev(α, x) for a certain Brauer class α ∈ Br(X ).<br />

We then describe a strategy on how the Brauer-Manin obstruction may be<br />

explicitly computed in concrete examples. We carry out this strategy for two<br />

special types of cubic surfaces.<br />

The first type is given as follows. Let p 0 ≡ 1 (mod 3) be a prime number<br />

<strong>and</strong> K/Ébe the unique cubic field extension contained in the cyclotomic


xviii<br />

INTRODUCTION<br />

extensionÉ(ζ p0 )/É. Fix the explicit generator θ ∈ K given by<br />

θ := trÉ(ζ p0 )/K(ζ p0 − 1) = −2n + ∑<br />

i∈(∗ p0 ) 3 ζ i p 0<br />

for n := p 0−1<br />

. Then, consider the cubic surface X ⊂ P 3É, given by<br />

6<br />

3 ( )<br />

T 3 (a 1 T 0 + d 1 T 3 )(a 2 T 0 + d 2 T 3 ) = T0 + θ (i) T 1 + (θ (i) ) 2 T 2 .<br />

Here, a 1 , a 2 , d 1 , d 2 ∈. The θ (i) denote the three images of θ under Gal(K/É).<br />

Proposition III.5.3 provides criteria to verify that such a surface is smooth<br />

<strong>and</strong> has p-adic <strong>points</strong> for every prime p. More importantly, the Brauer-Manin<br />

obstruction can be understood completely explicitly. At least for a generic<br />

choice of a 1 , a 2 , d 1 , <strong>and</strong> d 2 , one has that H 1( Gal(k/k), Pic(X k<br />

) ) =/3.<br />

Further, there is a class α ∈ Br(X ) with the following property. For an adelic<br />

point x = (x ν ) ν , the value of ev(α, x) depends only on the component x νp0 .<br />

Write x νp0 =: (t 0 : t 1 : t 2 : t 3 ). Then, one has ev(α, x) = 0 if <strong>and</strong> only if<br />

a 1 t 0 + d 1 t 3<br />

t 3<br />

is a cube in∗ p0<br />

. Note that p 0 ≡ 1 (mod 3) implies that only every third element<br />

of∗ p0<br />

is a cube.<br />

∏<br />

i=1<br />

Observe that the reduction of X modulo p 0 is given by<br />

T 3 (a 1 T 0 + d 1 T 3 )(a 2 T 0 + d 2 T 3 ) = T 3<br />

0 .<br />

This means, there are three planes intersecting in a triple line. NoÉp 0<br />

-<strong>rational</strong><br />

point may reduce to the triple line. Thus, there are three different planes, a<br />

Ép 0<br />

-<strong>rational</strong> point may reduce to. The value of ev(α, x) depends only the plane<br />

to which its component v νp0 is mapped under reduction.<br />

For instance (cf. Example III.5.24), for p 0 = 19, consider the cubic surface X<br />

given by<br />

3 ( )<br />

T 3 (T 0 + T 3 )(12T 0 + T 3 ) = T0 + θ (i) T 1 + (θ (i) ) 2 T 2 .<br />

Then, in19, the cubic equation<br />

∏<br />

i=1<br />

T (1 + T )(12 + T ) − 1 = 0<br />

has the three solutions 12, 15, <strong>and</strong> 17. However, in19, 13/12 = 9, 16/15 = 15,<br />

<strong>and</strong> 18/17 = 10 which are three non-cubes. This shows that X (É) = ∅. It is<br />

easy to check that X (É) ≠ ∅. Therefore, X is an example of a cubic surface<br />

violating the Hasse principle.<br />

We construct a number of similar examples. For instance, Example III.5.24<br />

describes a cubic surface X such that H 1( Gal(k/k), Pic(X k<br />

) ) =/3but the


INTRODUCTION<br />

xix<br />

generating Brauer class does not exclude a single adelic point. One would expect<br />

that X satisfies weak approximation. Recall that, in similar examples, E. Peyre<br />

<strong>and</strong> Y. Tschinkel [Pe/T] showed experimentally that there are three times more<br />

É-<strong>rational</strong> <strong>points</strong> than expected.<br />

The historically first cubic surface which could be proven to be a counterexample<br />

to the Hasse principle was provided by Sir P. Swinnerton-Dyer [S-D62].<br />

We recover Swinnerton-Dyer’s example (cf. Example III.5.27) for p 0 = 7,<br />

d 1 = d 2 = 1, a 1 = 1, <strong>and</strong> a 2 = 2. L. J. Mordell [Mord] generalized Swinnerton-Dyer’s<br />

work by giving a series of examples for p 0 = 7 <strong>and</strong> a series of<br />

examples for p 0 = 13. Yu. I. Manin mentions Mordell’s examples explicitly in<br />

his book [Man]. He explains these counterexamples to the Hasse principle by a<br />

Brauer class. We generalize Mordell’s examples further to the case that p 0 is an<br />

arbitrary prime such that p 0 ≡ 1 (mod 3).<br />

We conclude Chapter III by a section on diagonal cubic surfaces. For these, the<br />

Brauer-Manin obstruction was investigated in the monumental work [CT/K/S]<br />

of J.-L. Colliot-Thélène, D. Kanevsky, <strong>and</strong> J.-J. Sansuc. We present an explicit<br />

computation of the Brauer-Manin obstruction under a congruence condition<br />

which corresponds more or less to the “first case” of [CT/K/S]. Our argumentation<br />

is, however, shorter <strong>and</strong> simpler than the original one. The point is<br />

that we make use of the fact that H 1( Gal(k/k), Pic(X k<br />

) ) may be only 0,/2,<br />

/3, (/2) 2 , or (/3) 2 . Further, the group (/3) 2 appears only once in<br />

a very particular case. Thus, in order to prove H 1( Gal(k/k), Pic(X k<br />

) ) =/3,<br />

it is almost sufficient to construct an element of order three.<br />

The second part of this book is devoted to the various concepts of a height.<br />

In the fourth chapter, we start with the naive height forÉ-<strong>rational</strong> <strong>points</strong> on<br />

projective space. Then, our goal is to give an overview of the theories which<br />

provide natural generalizations of this simple concept.<br />

The very first generalization is the naive height for <strong>points</strong> in projective space<br />

defined over a finite extension ofÉ. Then, following André Weil, we introduce<br />

the concept of a height defined by an ample invertible sheaf. This is a height<br />

function which is defined only up to a bounded summ<strong>and</strong>.<br />

To overcome this difficulty, one has to work with arithmetic varieties <strong>and</strong><br />

metrized invertible sheaves. Arithmetic varieties are schemes projective<br />

over Spec. Actually, this leads to a beautiful geometric interpretation of<br />

the naive height.<br />

Indeed, let X be a projective variety overÉ<strong>and</strong> X a projective model of X<br />

over Spec. Fix a hermitian line bundle L on X . Then, according to the valuative<br />

criterion of properness, everyÉ-<strong>rational</strong> point x on X extends uniquely<br />

to a-valued point x : Spec→X. The height function with respect to L is


xx<br />

INTRODUCTION<br />

then given by<br />

h L<br />

(x) := ̂deg ( x ∗ L ) .<br />

Here, ̂deg denotes the Arakelov degree of a hermitian line bundle over Spec.<br />

It turns out that this coincides exactly with the naive height when one works<br />

with X = P n, L = O(1), <strong>and</strong> the minimum metric which is defined by<br />

∣ ‖l‖ min := min<br />

l ∣∣∣<br />

i=0, ... ,n<br />

∣ .<br />

X i<br />

In general, h L<br />

admits a fundamental finiteness property as soon as L is ample.<br />

This is the starting point of arithmetic intersection theory, a fascinating theory<br />

which we may only touch upon. We recall the main definitions <strong>and</strong> results<br />

of the arithmetic intersection theory due to H. Gillet <strong>and</strong> C. Soulé [G/S 90,<br />

G/S 92, S/A/B/K]. As an application, we consider the concept of a height<br />

introduced by J.-B. Bost, H. Gillet, <strong>and</strong> C. Soulé [B/G/S]. This is a height<br />

function for cycles of arbitrary dimension. For a 0-dimensional cycle, the<br />

Bost-Gillet-Soule height coincides with the usual height function defined by the<br />

hermitian line bundle.<br />

Wecloseour overviewofarithmeticintersectiontheoryby a sectiononthe arithmetic<br />

Hilbert-Samuel formula. We formulate the formula as Theorem IV.6.5.<br />

Asaconsequence, weproveS. Zhangs’s[Zh95b, (5.2)]theorem ofsuccessiveminima.<br />

Arithmetic intersection theory seems to be a very general framework. It is,<br />

however, still not general enough. An obvious problem is that the naive height<br />

on P n is not covered. This problem is of technical nature. To define the intersection<br />

product of two hermitian line bundles (or, more correctly, of the<br />

corresponding arithmetic cycles), Gillet <strong>and</strong> Soulé had to assume that all hermitian<br />

metrics are smooth. Unfortunately, the minimum metric is continuous,<br />

but not smooth.<br />

A theory which is general enough to cover the Bost-Gillet-Soule height as well<br />

as the naive height is provided by S. Zhang’s adelic intersection theory [Zh95a].<br />

We give a definition for a version of the adelic Picard group ˜Pic(X ). Elements of<br />

˜Pic(X ) are called integrably metrized invertible sheaves. Then, we construct<br />

the corresponding adelic intersection product. As an application, we explain<br />

the height function defined by an integrably metrized invertible sheaf.<br />

In Chapter 5, we use adelic intersection theory to give a unified proof for two<br />

equidistribution theorems for <strong>points</strong> of small height. On one h<strong>and</strong>, there is<br />

the equidistribution theorem of L. Szpiro, E. Ullmo <strong>and</strong> S. Zhang [S/U/Z]<br />

for small <strong>points</strong> on abelian varieties. On the other h<strong>and</strong>, the completely parallel<br />

theorem for the naive height on projective space which was established


INTRODUCTION<br />

xxi<br />

by Yuri Bilu [Bilu97]. The main result of Chapter 5 is formulated as Theorem<br />

V.1.14. A. Chambert-Loir’s equidistribution result [C-L] for quasi-split<br />

semi-abelian varieties can easily be deduced from this theorem.<br />

Chapter VI is devoted to some of the most popular conjectures concerning<br />

<strong>rational</strong> <strong>points</strong> on projective algebraic varieties. We discuss Lang’s conjecture,<br />

the conjecture of Batyrev <strong>and</strong> Manin, <strong>and</strong>, most notably, Manin’s conjecture<br />

about the asymptotics of <strong>points</strong> of bounded height on Fano varieties (Conjecture<br />

VI.5.37). A large part of the chapter is concerned with E. Peyre’s Tamagawa<br />

type number τ (X ), the coefficient expected in the asymptotic formula.<br />

We discuss in detail all factors appearing in the definition of τ (X ).<br />

In particular, we give a number of examples for which we explicitly compute<br />

the factor α(X ). We mainly consider smooth cubic surfaces of arithmetic Picard<br />

rank two.<br />

Part C collects several reports on practical experiments. Each chapter is concerned<br />

with a particular problem for a specific variety or family of varieties.<br />

We tried to keep the chapters as self-contained as possible. Each one starts with<br />

an introduction recalling those parts of the general theory described in parts A<br />

<strong>and</strong> B which are necessary to follow that chapter.<br />

In Chapter VII, we describe our investigations regarding the families<br />

“ax 3 = by 3 +z 3 +v 3 +w 3 ”, a, b = 1, . . . , 100, <strong>and</strong> “ax 4 = by 4 +z 4 +v 4 +w 4 ”,<br />

a, b = 1, . . . , 100, of projective algebraic threefolds. We report numerical evidence<br />

for the conjecture of Manin in the refined form due to E. Peyre.<br />

Our experiments included searching for <strong>points</strong>, computing the Tamagawa number,<br />

<strong>and</strong> detecting the accumulating subvarieties. Concerning the programmer’s<br />

efforts, detection of accumulating subvarieties was the most difficult part of<br />

this project. For example, one the cubic threefolds, the following non-obvious<br />

lines have been found. These are the only non-obvious lines we know <strong>and</strong> the<br />

only ones containing a point of height less than 5000.<br />

TABLE 1. Sporadic lines on cubic threefolds<br />

a b Smallest point Point s.t. x = 0<br />

19 18 (1 : 2 : 3 : -3 : -5) (0 : 7 : 1 : -7 : -18)<br />

21 6 (1 : 2 : 3 : -3 : -3) (0 : 9 : 1 : -10 : -15)<br />

22 5 (1 : -1 : 3 : 3 : -3) (0 : 27 : -4 : -60 : 49)<br />

45 18 (1 : 1 : 3 : 3 : -3) (0 : 3 : -1 : 3 : -8)<br />

73 17 (1 : 5 : -2 : 11 : -15) (0 : 27 : -40 : 85 : -96)<br />

We describe all the computations which were done as well some background on<br />

the geometry of cubic <strong>and</strong> quartic threefolds.


xxii<br />

INTRODUCTION<br />

In the eighth chapter, we continue our study of the family of quartic threefolds<br />

by comparing the height m(X ) of the smallest <strong>rational</strong> point with Peyre’s constant.<br />

We prove that there is no constant C such that m(X ) < C<br />

τ<br />

for every<br />

(X )<br />

diagonal quartic threefold, not even for those of type ax 4 = y 4 + z 4 + v 4 + w 4 .<br />

1<br />

We also prove that, at least for diagonal quartic threefolds, the reciprocal<br />

τ (X )<br />

1<br />

behaves like a height function. I.e.,<br />

τ<br />

admits a fundamental finiteness property.<br />

(X )<br />

Chapter IX is devoted to the study of general cubic surfaces. A general cubic<br />

surface is described by twenty coefficients. With current technology, it is impossible<br />

to study all cubic surfaces with coefficients below a reasonable bound.<br />

For that reason, we decided to work with coefficient vectors provided by a<br />

r<strong>and</strong>om number generator. Our first sample consists of 20 000 surfaces with coefficients<br />

r<strong>and</strong>omly chosen in the interval [0, 50]. The second sample consists of<br />

20 000 surfaces with r<strong>and</strong>omly chosen coefficients from the interval [−100, 100].<br />

We verified explicitly that each of the surfaces studied is smooth. Then, we<br />

proved that, for each surface, the full Galois group W (E 6 ) acts on the 27 lines.<br />

This implies that the Picard rank is always equal to 1. Also, the Brauer-Manin obstruction<br />

is not present on any of the surfaces considered. The algorithm to<br />

check that the group acting on the 27 lines is indeed W (E 6 ) is described in detail.<br />

Further, we searched for <strong>rational</strong> <strong>points</strong> <strong>and</strong> computed the Tamagawa number,<br />

thereby giving numerical evidence for Manin’s conjecture.<br />

In the next chapter, we return to the more st<strong>and</strong>ard case of diagonal cubic surfaces.<br />

The experiments areanalogous to those described in Chapters VII<strong>and</strong> VIII<br />

for diagonal cubic <strong>and</strong> quartic threefolds. The theory is, however, more complicated.<br />

The geometric Picard rank is equal to 7 <strong>and</strong>, in the generic case,<br />

there is a Brauer-Manin obstruction to weak approximation excluding precisely<br />

two thirds of the adelic <strong>points</strong>. The factors α(X ) <strong>and</strong> β(X ) appearing in the<br />

definition of Peyre’s constant are not always the same <strong>and</strong> need to be considered.<br />

We demonstrate experimentally the connection of Peyre’s constant with the<br />

height m(X ) of the smallest <strong>rational</strong> point. Under the Generalized Riemann<br />

Hypothesis, we prove that there is no constant C such that m(X ) < C<br />

τ for (X )<br />

every diagonal cubic surface. We also prove that, for diagonal cubic surfaces, the<br />

1<br />

reciprocal<br />

τ behaves like a height function. I.e., 1<br />

(X ) τ<br />

admits a fundamental<br />

(X )<br />

finiteness property.<br />

Chapter XI is concerned with the Diophantine equation<br />

x 4 + 2y 4 = z 4 + 4w 4 . ()<br />

This equation gives an example of a K3 surface X defined overÉ. It is an open<br />

question whether there exists a K3 surface overÉwhich has a finite non-zero<br />

number ofÉ-<strong>rational</strong> <strong>points</strong>.


INTRODUCTION<br />

xxiii<br />

X might be a c<strong>and</strong>idate for a K3 surface with this property. (1 : 0 : 1 : 0)<br />

<strong>and</strong> (1 : 0 : (−1) : 0) are two obvious <strong>rational</strong> <strong>points</strong>. Sir P. Swinnerton-<br />

Dyer [Poo/T, Problem/Question 6.c)] had publicly posed the problem to find a<br />

third <strong>rational</strong> point on X. But no <strong>rational</strong> <strong>points</strong> different from the two obvious<br />

ones had been found in experiments carried out by several people.<br />

We explain our approach to efficiently search forÉ-<strong>rational</strong> <strong>points</strong> on algebraic<br />

varieties defined by a decoupled equation. It is based on hashing, a method from<br />

Computer Science. In the particular case of a surface in P 3É, our algorithm is of<br />

complexity essentially O(B 2 ) for a search bound of B.<br />

In the final implementation, we could work with the search bound B = 10 8 .<br />

We discovered the following solution of the Diophantine equation ().<br />

1 484 801 4 + 2 · 1 203 120 4 = 9 050 910 498 475 648 046899201<br />

1 169 407 4 + 4 · 1 157 520 4 = 9 050 910 498 475 648 046899201<br />

Up to changes of sign, this is the only non-obvious solution of () we know <strong>and</strong><br />

the only non-obvious solution of height less than 10 8 [EJ2, EJ3].<br />

The last chapter is a report on our approach to the three cubes problem. To be<br />

more precise, the problem is whether every <strong>rational</strong> integer n ≢ 4, 5 (mod 9)<br />

can be written as a sum of three integral cubes.<br />

Using an implementation of Elkies’ method, we found, among many others,<br />

the following equations.<br />

156 = 26 577 110 807 569 3 − 18 161 093 358 005 3 − 23 381 515 025 762 3<br />

318 = 1 970 320 861 387 3 + 1 750 553 226 136 3 − 2 352 152 467 181 3<br />

= 30 828 727 881 037 3 + 27 378 037 791 169 3 − 36 796 384 363 814 3<br />

366 = 241 832 223 257 3 + 167 734 571 306 3 − 266 193 616 507 3<br />

420 = 8 859 060 149 051 3 − 2 680 209 928 162 3 − 8 776 520 527 687 3<br />

564 = 53 872 419 107 3 − 1 300 749 634 3 − 53 872 166 335 3<br />

758 = 662 325 744 409 3 + 109 962 567 936 3 − 663 334 553 003 3<br />

= 83 471 297 139 078 3 + 77 308 024 343 011 3 − 101 433 242 878 565 3<br />

789 = 18 918 117 957 926 3 + 4 836 228 687 485 3 − 19 022 888 796 058 3<br />

894 = 19 868 127 639 556 3 + 2 322 626 411 251 3 − 19 878 702 430 997 3<br />

948 = 103 458 528 103 519 3 + 6 604 706 697 037 3 − 103 467 499 687 004 3<br />

For none of the numbers on the left, a decomposition into a sum of three cubes<br />

had been known before.<br />

Göttingen,<br />

Summer 2008


PART A<br />

THE BRAUER GROUP


CHAPTER I<br />

THE BRAUER-SEVERI VARIETY<br />

ASSOCIATED WITH<br />

A CENTRAL SIMPLE ALGEBRA ∗<br />

No attention should be paid to the fact that algebra <strong>and</strong> geometry are<br />

different in appearance.<br />

OMAR KHAYYÁM (1070, translated by A. R. Amir-Moez)<br />

1. Introductory remarks<br />

This chapter is devoted to the correspondence between central simple algebras<br />

<strong>and</strong> Brauer-Severi varieties.<br />

This is classical material. Only Section 7 is exceptional to a certain extent.<br />

There, we will give a functor from central simple algebras to Brauer-Severi<br />

varieties which yields the classical correspondence on objects.<br />

Some history. —<br />

Central simple algebras were studied intensively by many mathematicians at the<br />

end of the 19th <strong>and</strong> in the first half of the 20th century. We refer the reader to<br />

N. Bourbaki [Bou-A, Note historique] for a detailed account on the history of<br />

the subject <strong>and</strong> mention only a few important milestones here. The structure<br />

of central simple algebras (being finite dimensional over a field K) is fairly easy.<br />

They are full matrix rings over division algebras the center of which is equal<br />

to K. This was finally discovered by J. H. Maclagan-Wedderburn in 1907<br />

[MWe08] after several special cases had been treated before. T. Molien [Mol]<br />

had considered the case of-algebras already in 1893 <strong>and</strong> the case ofÊ-algebras<br />

had been investigated by E. Cartan [Car]. J. H. Maclagan-Wedderburn himself<br />

had proven the structure theorem for central simple algebras over finite fields<br />

in 1905 [MWe05, Di05].<br />

(∗) This chapter is a revised version of the article: The Brauer-Severi variety associated with a<br />

central simple algebra, Linear Algebraic Groups <strong>and</strong> Related Structures 52(2000), 1-60.


4 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

In 1929, R. Brauer ([Bra], see also [Deu] <strong>and</strong> [A/N/T]), using E. Noether’s<br />

ideas about crossed products of algebras, found the group structure on the<br />

set of similarity classes of central simple algebras over a field K. He constructed,<br />

in today’s language, an isomorphism to the Galois cohomology<br />

group H 2( Gal(K sep /K ), (K sep ) ∗) . Further, he discussed the structure of this<br />

group in the case of a number field.<br />

Relative versions of central simple algebras over base rings instead of fields were<br />

introduced by G. Azumaya [Az] <strong>and</strong> M. Ausl<strong>and</strong>er <strong>and</strong> O. Goldman [A/G].<br />

The case of an arbitrary base scheme was considered by A. Grothendieck in his<br />

famous Bourbaki talks “Le groupe de Brauer” [GrBrI, GrBrII, GrBrIII].<br />

Brauer-Severi varieties are twisted forms of the projective space. The term<br />

“varieté de Brauer” appeared for the first time in 1944 in the article [Ch44]<br />

of F. Châtelet. Nevertheless, F. Severi had proven already in 1932 that a Brauer-<br />

Severi variety over a field K admitting a K-valued point is necessarily isomorphic<br />

to the projective space.<br />

It should be mentioned that Brauer-Severi varieties are important in a number<br />

of applications. The first <strong>and</strong> definitely most striking one is the proof of the<br />

theorem of Merkurjev-Suslin [Me/S] on the cotorsion of K 2 of a field.<br />

The Merkurjev-Suslin theorem has an interesting further application. It provides<br />

a certain underst<strong>and</strong>ing of the (torsion part of the) Chow <strong>groups</strong> of certain<br />

algebraic varieties. The reader should consult the work of J.-L. Colliot-Thélène<br />

[CT91] for information about that.<br />

As a second kind of application, Brauer-Severi varieties often appear in constructions<br />

of varieties with particular properties. For instance, the famous example<br />

due to M. Artin <strong>and</strong> D. Mumford [A/M] of a threefold which is uni<strong>rational</strong><br />

but not <strong>rational</strong> is a variety fibered over a <strong>rational</strong> surface such that the generic<br />

fiber is a conic without <strong>rational</strong> <strong>points</strong>. For more historical details, especially<br />

on the work of F. Châtelet, we refer the reader to the article of J.-L. Colliot-<br />

Thélène [CT88].<br />

The correspondence between central simple algebras <strong>and</strong> Brauer-Severi varieties<br />

was first observed by E. Witt [Wi35] <strong>and</strong> H. Hasse in the special case of<br />

quaternion algebras <strong>and</strong> plane conics.<br />

Here, we are going to present in detail the most elementary approach to this correspondence.<br />

It is based on non-abelian group cohomology. The main observation<br />

is that central simple algebras of dimension n 2 over a field K as well<br />

as (n − 1)-dimensional Brauer-Severi varieties over K can both be described<br />

by classes in one <strong>and</strong> the same cohomology set H 1( Gal(K sep /K ), PGL n (K sep ) ) .<br />

Note that this approach was promoted by J.-P. Serre in his books [Se62, chap. X,


Sec. 2] NON-ABELIAN GROUP COHOMOLOGY 5<br />

§§5,6] <strong>and</strong> [Se73, Remarque III.1.3.1]. It was, however, known to F. Châtelet before.<br />

A second approach is closer to A. Grothendieck’s style. One can give a direct<br />

description of a functor of <strong>points</strong> P XA : {K-schemes} → {sets} in terms of data<br />

of the central simple algebra A. It is possible to prove representability by a<br />

projective scheme using brute force. This approach is presented explicitly in<br />

the book of I. Kersten [Ke]. For a very detailed account, the reader may consult<br />

the Ph.D. thesis of F. Henningsen [Hen]. We are going to prove that these two<br />

approaches are equivalent. In fact, we will compute the functor of <strong>points</strong> of the<br />

variety given by the first approach <strong>and</strong> show that it is naturally isomorphic to<br />

the functor usually taken as the starting point for the second approach.<br />

There is a third approach which we only mention here. It works via algebraic<br />

<strong>groups</strong> <strong>and</strong> can be used to produce twisted forms not only of the projective<br />

space but of any homogeneous space G/P where G is a semisimple algebraic<br />

group <strong>and</strong> P ⊂ G a parabolic subgroup (see [K/R]).<br />

2. Non-abelian group cohomology<br />

In this section, we recall elementary facts about non-abelian group cohomology.<br />

I.e., cohomology of discrete <strong>groups</strong> with non-abelian coefficients. Non-abelian<br />

Galois cohomology will be the central tool for this chapter.<br />

2.1. Definition. –––– Let G be a finite group.<br />

i) A G-set E is a set equipped with a G-operation from the left.<br />

A morphism of G-sets, a G-morphism for short, is a map i : E → F of G-sets<br />

such that the diagram<br />

G × E ·<br />

E<br />

commutes.<br />

id×i<br />

<br />

G × F<br />

· F<br />

ii) A G-set E carrying a group structure such that g (xy) = g x g y for every g ∈ G<br />

<strong>and</strong> x, y ∈ E is called a G-group. Here, we write g x := g ·x for x ∈ E <strong>and</strong> g ∈ G.<br />

If the group underlying E is abelian then E is called a G-module.<br />

2.2. Definition. –––– Let G be a finite group.<br />

i) For a G-set E, one puts<br />

H 0 (G, E) := E G .<br />

I.e., the zeroth cohomology set of G with coefficients in E is equal to the subset of<br />

G-invariants.<br />

i


6 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

ii) Let A be a G-group.<br />

Then, a cocycle from G to A is a map G → A, g ↦→ a g such that a gh = a g · ga<br />

h<br />

for each g, h ∈ G.<br />

Two cocycles a, a ′ are said to be cohomologous if there exists some b ∈ A such<br />

that a ′ g = b −1 · a g · gb for every g ∈ G.<br />

This is an equivalence relation <strong>and</strong> the quotient set, the first cohomology set of G<br />

with coefficients in A, is denoted by H 1 (G, A). This is a pointed set as the map<br />

g ↦→ e defines a cocycle, the so-called trivial cocycle.<br />

2.3. Remarks. –––– i) If E is a G-group then H 0 (G, E) is a group.<br />

ii) For a cocycle a, a ′ such that a ′ g := b−1·a g · gb for each g ∈ G is a cocycle, too.<br />

iii) H 0 (G, A) <strong>and</strong> H 1 (G, A) are covariant functors in A. If i : A → A ′ is a<br />

morphism of G-sets (a morphism of G-<strong>groups</strong>) then the induced map(s) will be<br />

denoted by i ∗ : H 0 (G, A) → H 0 (G, A ′ ) (<strong>and</strong> i ∗ : H 1 (G, A) → H 1 (G, A ′ )).<br />

iv) If A is abelian then the notions defined above coincide with the usual<br />

group cohomology. One of the possible descriptions for H ∗ (G, A) is the cohomology<br />

of the complex<br />

with differentials<br />

0 −→ A d<br />

−→ Map(G, A)<br />

d<br />

−→ Map(G 2 , A)<br />

dϕ(g 1 , . . . , g n+1 ) := g 1<br />

ϕ(g 2 , . . . , g n+1 )<br />

+<br />

n<br />

∑<br />

j=1<br />

d<br />

−→ . . .<br />

(−1) j ϕ(g 1 , . . . , g j g j+1 , . . . , g n+1 )<br />

+ (−1) n+1 ϕ(g 1 , . . . , g n ) .<br />

2.4. Proposition. –––– Let G be a finite group.<br />

a) Let A ⊆ B be a G-subgroup <strong>and</strong> B/A the set of left cosets. Then, there is a natural<br />

exact sequence of pointed sets<br />

1 → H 0 (G, A) → H 0 (G, B) → H 0 (G, B/A) δ → H 1 (G, A) → H 1 (G, B) .<br />

b) If A ⊆ B is a normal G-subgroup then there is a natural exact sequence of<br />

pointed sets<br />

1 → H 0 (G, A) → H 0 (G, B) → H 0 (G, B/A) δ → H 1 (G, A) → . . .<br />

. . . → H 1 (G, B) → H 1 (G,B/A) .


Sec. 2] NON-ABELIAN GROUP COHOMOLOGY 7<br />

c) If A ⊆ B is a G-module lying in the center of B then there is a natural exact<br />

sequence of pointed sets<br />

1 → H 0 (G, A) → H 0 (G, B) → H 0 (G, B/A) δ → H 1 (G, A) → . . .<br />

. . . → H 1 (G, B) → H 1 (G, B/A) → δ′<br />

H 2 (G, A) .<br />

Here,theabeliangroupH 2 (G, A)isconsidered asapointed setwiththeunitelement.<br />

Proof. The connection homomorphism δ is defined in the following way.<br />

Let x ∈ H 0 (G, B/A). Take a representative x ∈ B for x <strong>and</strong> put a s := x −1 · sx.<br />

This is a cocycle <strong>and</strong> its equivalence class is denoted by δ(x). The definition<br />

of δ(x) is independent of the x chosen.<br />

In the situation of c), the map δ ′ is given as follows. Let x ∈ H 1 (G, B/A).<br />

Choose a cocycle (x g ) g∈G that represents x <strong>and</strong> lift each x g to some x g ∈ B.<br />

Put a(g 1 , g 2 ) := g 1 xg2 · x −1<br />

g 1 g 2 · x g1 . This is a 2-cocycle with values in A <strong>and</strong> its<br />

equivalence class is denoted by δ ′ (x). The definition of δ ′ (x) is independent of<br />

the choices.<br />

Exactness has to be checked at each entry, separately.<br />

2.5. Remark. –––– A sequence (A, a)<br />

to be exact in (B, b) if i(A) = j −1 (c).<br />

i<br />

→ (B, b)<br />

□<br />

j<br />

→ (C, c) of pointed sets is said<br />

2.6. Definition. –––– Let h : G ′ → G be a homomorphism of finite <strong>groups</strong>.<br />

Then, for an arbitrary G-set E, one has a natural pull-back map<br />

h ∗ : H 0 (G, E) → H 0 (G ′ , E).<br />

If E is a G-group then the pull-back map is a group homomorphism.<br />

For an arbitrary G-group A, there is the natural pull-back map<br />

which is a morphism of pointed sets.<br />

h ∗ : H 1 (G, A) → H 1 (G ′ , A)<br />

2.7. Remark. –––– If h is the inclusion of a subgroup then the pull-back<br />

res G′<br />

G := h∗ is usually called the restriction map.<br />

If h : G ′ ֒→ G is the canonical projection to a quotient group then inf G′<br />

G := h ∗<br />

is said to be the inflation map.<br />

The composition of res G′<br />

G<br />

or infG′ G with some extension of the G ′ -set E (the<br />

G ′ -group A) is usually called the restriction, respectively inflation, too.


8 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

2.8. Remark. –––– Non-abelian group cohomology may easily be extended to<br />

the case where G is a profinite group <strong>and</strong> A a discrete G-set (respectively<br />

G-group) on which G operates continuously. Indeed, put<br />

H i (G, A) := lim H i (G/G ′ , A G′ )<br />

−→G<br />

′<br />

fori = 0<strong>and</strong>1. Here, thedirectlimitistakenover theinflationmaps<strong>and</strong>G ′ runs<br />

through the open normal sub<strong>groups</strong> G ′ ⊆ G with finite quotient.<br />

3. Galois descent<br />

3.1. Definition. –––– Let L/K be a finite Galois extension <strong>and</strong> σ ∈ Gal(L/K ).<br />

a) Then, a map i : V 1 → V 2 of L-vector spaces is said to be σ-linear if, for every<br />

v, w ∈ V 1 <strong>and</strong> λ ∈ L, one has<br />

i(v + w) = i(v) + i(w)<br />

<strong>and</strong><br />

i(λv) = σ(λ)i(v) .<br />

b) Let π 1 : X 1 → SpecL <strong>and</strong> π 2 : X 2 → SpecL be L-schemes. We say that<br />

f : X 1 → X 2 is a morphism of L-schemes twisted by σ if the diagram<br />

X 1<br />

f<br />

X 2<br />

π 2<br />

π 1<br />

<br />

SpecL<br />

S(σ)<br />

SpecL<br />

commutes. Here, S(σ): SpecL → SpecL denotes the morphism of affine<br />

schemes induced by σ −1 : L → L.<br />

3.2. Theorem. –––– Let L/K be a finite Galois extension <strong>and</strong> G := Gal(L/K ).<br />

Then,<br />

i) there are the following equivalences of categories,<br />

⎧<br />

⎫<br />

⎨L-vector spaces<br />

⎬<br />

{K-vector spaces} −→ with a G-operation from the left<br />

⎩<br />

⎭ ,<br />

such that every σ ∈ G operates σ-linearly<br />

⎧<br />

⎫<br />

⎨L-algebras<br />

⎬<br />

{K-algebras} −→ with a G-operation from the left<br />

⎩<br />

⎭ ,<br />

such that every σ ∈ G operates σ-linearly


Sec. 3] GALOIS DESCENT 9<br />

⎧ ⎫ ⎧ ⎫<br />

⎨ central simple ⎬ ⎨ central simple algebras over L ⎬<br />

algebras<br />

⎩ ⎭ −→ with a G-operation from the left<br />

⎩<br />

⎭ ,<br />

over K<br />

such that every σ ∈ G operates σ-linearly<br />

⎧<br />

⎫<br />

{ } ⎨ commutative L-algebras<br />

⎬<br />

commutative<br />

−→ with a G-operation from the left<br />

K-algebras ⎩<br />

⎭ ,<br />

such that every σ ∈ G operates σ-linearly<br />

⎧ ⎫ ⎧ ⎫<br />

⎨ commutative ⎬ ⎨ commutative L-algebras with unit ⎬<br />

K-algebras<br />

⎩ ⎭ −→ with a G-operation from the left<br />

⎩<br />

⎭ ,<br />

with unit such that every σ ∈ G operates σ-linearly<br />

A ↦→ A ⊗ K L,<br />

ii) there is the following equivalence of categories,<br />

⎧<br />

⎫<br />

quasi-projective L-schemes<br />

{ } ⎪⎨ with a G-operation from the left ⎪⎬<br />

quasi-projective<br />

−→ by morphisms of K-schemes ,<br />

K-schemes<br />

such that every σ ∈ G operates<br />

⎪⎩<br />

⎪⎭<br />

by a morphism twisted by σ<br />

X ↦→ X × SpecK SpecL .<br />

iii) Let X be a K-scheme <strong>and</strong> r be a natural number. Then there are the following<br />

equivalences of categories,<br />

⎧<br />

⎫<br />

quasi-coherent sheaves M<br />

{ }<br />

quasi-coherent<br />

−→<br />

sheaves on X<br />

{ }<br />

locally free sheaves<br />

−→<br />

of rank r on X<br />

on X × SpecK SpecL<br />

⎪⎨<br />

⎪⎬<br />

together with a system (ι σ ) σ∈G<br />

of isomorphisms ι σ : x ∗ σ M → M ,<br />

satisfying ι ⎪⎩<br />

τ ◦ x ∗ τ (ι σ) = ι στ ⎪⎭<br />

for every σ, τ ∈ G<br />

⎧<br />

⎫<br />

locally free sheaves M of rank r<br />

on X × SpecK SpecL<br />

⎪⎨<br />

⎪⎬<br />

together with a system (ι σ ) σ∈G<br />

of isomorphisms ι σ : x ∗ σ M → M ,<br />

satisfying ι ⎪⎩<br />

τ ◦ x ∗ τ (ι σ) = ι στ ⎪⎭<br />

for every σ, τ ∈ G<br />

F ↦→ M := π ∗ F.<br />

Here the morphisms in the categories are those respecting all the extra structures.<br />

π : X × SpecK SpecL −→ X is the canonical morphism <strong>and</strong><br />

x σ : X × SpecK SpecL −→ X × SpecK SpecL


10 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

denotes the morphism induced by S(σ): SpecL → SpecL.<br />

Proof. In each case we have to prove that the functor given is fully faithful <strong>and</strong><br />

essentially surjective. Full faithfulness is proven in Propositions 3.7, 3.8, <strong>and</strong><br />

3.9, respectively. Propositions 3.3, 3.5, <strong>and</strong> 3.6 show essential surjectivity. □<br />

3.3. Proposition (Galois descent-algebraic version). —–<br />

Let L/K be a finite Galois extension of fields <strong>and</strong> G := Gal(L/K ).<br />

Let W be a vector space (an algebra, a central simple algebra, a commutative<br />

algebra, a commutative algebra with unit, ... ) over L together with an operation<br />

T : G × W → W of G from the left. Assume that T respects all extra structures<br />

<strong>and</strong> that for each σ ∈ G the action of σ is a σ-linear map T σ : W → W .<br />

Then, there is a vector space V (an algebra, a central simple algebra, a commutative<br />

algebra, a commutative algebra with unit, ... ) over K such that there is an<br />

isomorphism<br />

V ⊗ K L −→ b<br />

∼=<br />

W .<br />

Here, V ⊗ K L is equipped with the G-operation induced by the canonical operation<br />

on L. b respects all extra structures <strong>and</strong> the operation of G.<br />

Proof. Define V := W G . This is clearly a K-vector space (a K-algebra, a<br />

commutative K-algebra, a commutative K-algebra with unit). If W is a central<br />

simple algebra over L then V is a central simple algebra over K. The latter can<br />

not be seen directly but follows immediately from the formula W G ⊗ K L = W<br />

which we prove below. For that, let {l 1 , . . . , l n } be a K-basis of L.<br />

The assertion follows from the claim below.<br />

□<br />

3.4. Claim. –––– There exist an index set J <strong>and</strong> a subset {x j | j ∈ J } ⊂ W G such<br />

that {l i x j | i ∈ {1, . . . , n}, j ∈ J } is a K-basis of W .<br />

Proof. By Zorn’s Lemma, there exists a maximal subset {x j | j ∈ J } ⊂ W G<br />

such that {l i x j | i ∈ {1, . . . , n}, j ∈ J } ⊂ W is a system of K-linearly<br />

independent vectors. Assume, that system is not a basis of W . Then,<br />

〈l i x j | i ∈ {1, . . . , n}, j ∈ J 〉 K<br />

is a proper G-invariant L-sub-vector space of W <strong>and</strong> one can choose an element<br />

x ∈ W \ 〈l i x j | i ∈ {1, . . . , n}, j ∈ J 〉 K . For every l ∈ L, the sum<br />

∑ T σ (lx) = ∑ σ(l) · T σ (x)<br />

σ∈G σ∈G


Sec. 3] GALOIS DESCENT 11<br />

is G-invariant. Further, by linear independence of characters, the matrix<br />

⎛<br />

⎞<br />

σ 1 (l 1 ) . . . σ 1 (l n )<br />

⎜<br />

⎝<br />

.<br />

· · · ..<br />

⎟ · · · ⎠<br />

σ n (l 1 ) . . . σ n (l n )<br />

is of maximal rank. In particular, there is some l ∈ L such that the image of<br />

x β := ∑ σ(l) · σ(x)<br />

σ∈G<br />

in W /〈l i x j | i ∈ {1, . . . , n}, j ∈ J 〉 K is not equal to zero. Therefore,<br />

{l i x j | i ∈ {1, . . . , n}, j ∈ J } ∪ {l i x β | i ∈ {1, . . . , n} }<br />

is a K-linearly independent system of vectors contradicting the maximality<br />

of {x j | j ∈ J }.<br />

□<br />

3.5. Proposition (Galois descent-geometric version). —–<br />

Let L/K be a finite Galois extension of fields <strong>and</strong> G := Gal(L/K ).<br />

Further, let Y be a quasi-projective L-scheme together with an operation of G from<br />

the left by twisted morphisms. I.e., such that the diagrams<br />

Y<br />

T σ<br />

Y<br />

SpecL<br />

S(σ)<br />

SpecL<br />

commute where S(σ): SpecL → SpecL is the morphism induced by σ −1 : L → L.<br />

Then, there exists a quasi-projective K-scheme X such that there is an isomorphism<br />

of L-schemes<br />

X × SpecK SpecL −→ f<br />

∼=<br />

Y .<br />

Here, X × SpecK SpecL is equipped with the G-operation induced by the operation<br />

on SpecL <strong>and</strong> f is compatible with the operation of G.<br />

Proof. Affine Case. Let Y = SpecB be an affine scheme. The G-operation<br />

on SpecB corresponds to a G-operation from the right on B such that each<br />

σ ∈ G operates σ −1 -linearly. Define a G-operation from the left on B<br />

by σ · b := bσ −1 . The assertion follows immediately from Proposition 3.3.<br />

GeneralCase. ByLemma3.10, thereexistsanaffineopencovering {Y 1 , . . . , Y n }<br />

of Y by G-invariant schemes. Galois descent yields affine K-schemes<br />

X 1 , . . . , X n such that there are isomorphisms<br />

X i × SpecK SpecL ∼ =<br />

−→ Y i


12 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

<strong>and</strong> affine K-schemes X ij , 1 ≤ i < j ≤ n, such that there are isomorphisms<br />

X ij × SpecK SpecL ∼ =<br />

−→ Y i ∩ Y j .<br />

It is to be shown that every X ij admits canonical, open embeddings into X i<br />

<strong>and</strong> X j .<br />

In each case, on the level of rings we have a homomorphism A⊗ K L −→ B ⊗ K L<br />

<strong>and</strong> an isomorphism B ⊗ K L ∼ =<br />

−→ (A ⊗ K L) f such that their composition is the<br />

localization map. Clearly, f may be assumed to be G-invariant. I.e., we may<br />

suppose f ∈ A. Consequently, B ⊗ K L ∼ = A f ⊗ K L <strong>and</strong>, by consideration of<br />

the G-invariants on both sides, B ∼ = A f .<br />

The cocycle relations are obvious. Hence, we can glue the affine schemes<br />

X 1 , . . . , X n along the affine schemes X ij for 1 ≤ i < j ≤ n to obtain the<br />

scheme X desired. Lemma 3.12.ix) below completes the proof. □<br />

3.6. Proposition (Galois descent for quasi-coherent sheaves). —–<br />

Let L/K be a finite Galois extension of fields <strong>and</strong> G := Gal(L/K ).<br />

Further, let X be a K-scheme, π : X × SpecK SpecL → X the canonical morphism<br />

<strong>and</strong> x σ : X × SpecK SpecL → X × SpecK SpecL be the morphism induced by<br />

S(σ): SpecL → SpecL.<br />

Let M bea quasi-coherentsheaf over X × SpecK SpecL together with a system (ι σ ) σ∈G<br />

of isomorphisms ι σ : x ∗ σM → M which are compatible in the sense that for each<br />

σ, τ ∈ G there is the relation ι τ ◦ x ∗ τ(ι σ ) = ι στ .<br />

Then, there exists a quasi-coherent sheaf F over X such that there is an isomorphism<br />

under which the canonical isomorphism<br />

π ∗ F ∼ =<br />

−→ M<br />

i σ : x ∗ σπ ∗ F = (πx σ ) ∗ F = π ∗ F id<br />

−→ π ∗ F<br />

is identified with ι σ for each σ. I.e., the diagrams<br />

b<br />

x ∗ σπ ∗ F<br />

x∗ σ(b)<br />

x ∗ σM<br />

are commutative.<br />

i σ<br />

<br />

π ∗ F<br />

b<br />

M<br />

Proof. First step. Assume X ∼ = SpecA to be an affine scheme.<br />

Then, M = ˜M for some (A ⊗ K L)-module M. We have<br />

x ∗ σ M =<br />

˜<br />

M ⊗ (A⊗K L) (A ⊗ K L σ−1 ) = M ˜ ⊗ L L σ−1 .<br />

ι σ


Sec. 3] GALOIS DESCENT 13<br />

Hence, x ∗ σM = ˜M σ−1 where M σ−1 coincides with M as an A-module. But its<br />

structure of an L-vector space is given by<br />

l ·M σ −1 m := σ−1 (l) ·M m.<br />

Consequently, the isomorphism ι σ : x ∗ σM → M is induced by an A-module<br />

isomorphism j σ : M → M which is σ −1 -linear. The compatibility relations,<br />

required above, are easily translated into the condition that the maps j σ form a<br />

G-operation on M from the right.<br />

Define a G-operation from the left on M by σ ·m := j σ −1(m). By Galois descent<br />

for vector spaces, theK-vector spaceM G ofG-invariants satisfies M G ⊗ K L ∼ = M.<br />

Thisisalsoanisomorphism ofA-modulesastheG-operationonM iscompatible<br />

with the A-operation. Putting F := ˜M G , we obtain a quasi-coherent sheaf<br />

over X such that π ∗ F ∼ = M .<br />

The commutativity of the diagram is a consequence of Proposition 3.3.<br />

Second step. In general, consider an affine open covering<br />

for X α<br />

∼ = SpecRα .<br />

X = [ α∈I<br />

X α<br />

For every intersection X α1 ∩ X α2 , we consider an affine open covering<br />

{X α1 ,α 2 ,β ∼ = SpecR α1 ,α 2 ,β | β ∈ J α1 ,α 2<br />

}.<br />

By the affine case, for each α ∈ I, we are given an R α -module M α such that<br />

π ∗ ˜M α<br />

∼ = M |Xα × SpecK Spec L. Further, for each triple (α 1 , α 2 , β) with α 1 , α 2 ∈ I<br />

<strong>and</strong> β ∈ J α1 ,α 2<br />

, we have R α1 ,α 2 ,β-modules M α1 ,α 2 ,β satisfying<br />

π ∗ ˜ Mα1 ,α 2 ,β ∼ = M | Xα1 ,α 2 ,β× SpecK SpecL .<br />

The construction of these modules is compatible with restriction to affine subschemes.<br />

Therefore, by Proposition 3.9, we get isomorphisms<br />

i α1 ,α 2 ,β : M α1 ⊗ Rα1 R α1 ,α 2 ,β<br />

∼=<br />

−→ M α2 ⊗ Rα2 R α1 ,α 2 ,β.<br />

It is clear that, for every α 1 , α 2 , α 3 ∈ I <strong>and</strong> every β 1 ∈ J α2 ,α 3<br />

, β 2 ∈ J α3 ,α 1<br />

,<br />

<strong>and</strong> β 3 ∈ J α1 ,α 2<br />

, these isomorphisms are compatible on the triple intersection<br />

X α1 ,α 2 ,β 3<br />

∩ X α3 ,α 1 ,β 2<br />

∩ X α2 ,α 3 ,β 1<br />

. I.e., we can glue the quasi-coherent sheaves ˜M α<br />

along the ˜M α1 ,α 2 ,β to obtain the quasi-coherent sheaf M desired. □<br />

3.7. Proposition (Galois descent for homomorphisms). —–<br />

Let L/K be a finite Galois extension of fields <strong>and</strong> G := Gal(L/K ). Then, it is<br />

equivalent


14 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

i) to give a homomorphism r : V → V ′ of K-vector spaces (of algebras over K,<br />

of central simple algebras over K, of commutative K-algebras, of commutative<br />

K-algebras with unit, ... ),<br />

ii) to give a homomorphism r L : V ⊗ K L → V ′ ⊗ K L of L-vector spaces (of algebras<br />

over L, ofcentralsimplealgebrasover L, ofcommutative L-algebras, ofcommutative<br />

L-algebras with unit, ... ) which is compatible with the G-operations. I.e., such that<br />

for each σ ∈ G the diagram<br />

V ⊗ K L<br />

r L<br />

V ′ ⊗ K L<br />

commutes.<br />

σ<br />

V ⊗ K L<br />

r L<br />

σ<br />

V ′ ⊗ K L<br />

Proof. If r is given then one defines r L := r ⊗ K L. Clearly, if r is a ring<br />

homomorphism then r L is, too.<br />

Conversely, in order to construct r from r L , note that the commutativity of<br />

the diagrams implies that r L is compatible with G-invariants. Further, we<br />

know (V ⊗ K L) G = V <strong>and</strong> (V ′ ⊗ K L) G = V ′ . Thus, we obtain a K-linear<br />

map r : V −→ V ′ . If r L is a ring homomorphism then its restriction r is, too.<br />

It is clear that the two procedures described are inverse to each other.<br />

□<br />

3.8. Proposition (Galois descent for morphisms of schemes). —–<br />

Let L/K be a finite Galois extension of fields <strong>and</strong> G := Gal(L/K ).<br />

Then, it is equivalent<br />

i) to give a morphism of K-schemes f : X → X ′ ,<br />

ii) to give a morphism of L-schemes f L : X × SpecK SpecL → X ′ × SpecK SpecL<br />

which is compatible with the G-operations. I.e., such that for each σ ∈ G the<br />

diagram<br />

X × SpecK SpecL<br />

f L<br />

X ′ × SpecK SpecL<br />

commutes.<br />

σ<br />

X × SpecK SpecL<br />

f L<br />

σ<br />

X ′ × SpecK SpecL<br />

Proof. Let f be given. Then, one defines f L := f × SpecK SpecL.<br />

Conversely, in order to construct f from f L , observe that the question is<br />

local in X ′ <strong>and</strong> X. Thus, we may assume we are given a homomorphism


Sec. 3] GALOIS DESCENT 15<br />

r L : A ′ ⊗ K L −→ A ⊗ K L of L-algebras with unit making the diagrams<br />

A ′ ⊗ K L<br />

r L<br />

A ⊗ K L<br />

σ<br />

A ′ ⊗ K L<br />

r L<br />

σ<br />

A ⊗K L<br />

commute. This is exactly the situation covered by Proposition 3.7.<br />

It is clear that the two processes described are inverse to each other.<br />

□<br />

3.9. Proposition (Galois descent for morphisms of quasi-coherent sheaves). —–<br />

Let L/K be a finite Galois extension of fields <strong>and</strong> G := Gal(L/K ). Further, let X<br />

be a K-scheme <strong>and</strong> π : X × SpecK SpecL → X be the canonical morphism.<br />

Then, it is equivalent<br />

i) to give a morphism r : F → G of coherent sheaves over X,<br />

ii) to give a morphism r L : π ∗ F → π ∗ G of quasi-coherent sheaves over<br />

X × SpecK SpecL which is compatible with the G-operations. I.e., such that for<br />

each σ ∈ G the diagram<br />

commutes.<br />

π ∗ F<br />

x ∗ σ<br />

<br />

π ∗ F<br />

Proof. If r is given then one defines r L := π ∗ r.<br />

r L<br />

r L<br />

π ∗ G<br />

x ∗ σ<br />

<br />

π ∗ G<br />

Conversely, if r L is given then the question to construct r is local in X. Thus, assume<br />

A is a commutative ring with unit <strong>and</strong> M <strong>and</strong> N are A-modules. We are<br />

given a homomorphism r L : M ⊗ K L → N ⊗ K L of A ⊗ K L-modules such that<br />

the diagrams<br />

r L<br />

M ⊗ K L N ⊗ K L<br />

σ<br />

M ⊗ K L<br />

r L<br />

σ<br />

N ⊗K L<br />

commute for each σ ∈ G. We get a morphism r : M → N as M <strong>and</strong> N are the<br />

A-modules of G-invariants on the left <strong>and</strong> right h<strong>and</strong> side, respectively.<br />

The two procedures described above are inverse to each other.<br />

3.10. Lemma. –––– Let L be a field <strong>and</strong> Y be a quasi-projective L-scheme equipped<br />

with an operation of some finite group G acting by morphisms of schemes.<br />

Then, there exists a covering of Y by G-invariant affine open subsets.<br />

Proof. Let y ∈ Y be an arbitrary closed point. Everything which is needed is an<br />

affine open G-invariant subset containing y. For that, we choose an embedding<br />


16 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

i : Y ֒→ P N L . By Sublemma 3.11, there exists a hypersurface H y such that<br />

H y ⊃ i(Y ) \ i(Y ) <strong>and</strong> i(σ(y)) ∉ H y for every σ ∈ G. Here, i(Y ) denotes the<br />

closure of i(Y ) in P N L . By construction, the morphism<br />

i| Y \i −1 (H y ) : Y \ i −1 (H y ) −→ P N L \ H y<br />

is a closed embedding. As P N L \ H y is an affine scheme, Y \ i −1 (H y ) must be<br />

affine, too. Hence,<br />

O y := \<br />

σ −1 (Y \ i −1 (H y )) ⊂ Y<br />

σ∈G<br />

is the intersection of finitely many affine open subsets in a quasi-projective <strong>and</strong>,<br />

therefore, separated scheme. Thus, O y is an affine open subset. By construction,<br />

O y is G-invariant <strong>and</strong> contains y.<br />

□<br />

3.11. Sublemma. –––– Let L be a field <strong>and</strong> Z P N L be a Zariski-closed subset.<br />

Further, let p 1 , . . . , p n ∈ P N L be finitely many closed <strong>points</strong> not contained in Z.<br />

Then, there exists a hypersurface H ⊂ P N L which contains Z but does not contain<br />

any of the <strong>points</strong> p 1 , . . . , p n .<br />

Proof. We will give two proofs, an elementary one <strong>and</strong> a more canonical one<br />

which uses cohomology of coherent sheaves.<br />

First proof. Let S := L[X 0 , . . . , X N ] be the homogeneous coordinate ring for<br />

the projective space P N L . S is a graded L-algebra. For d ∈Æ, we will denote by<br />

S d the L-vector space of homogeneous elements of degree d.<br />

We proceed by induction. The case n = 0 is trivial. Thus, assume that the<br />

assertion is true for n − 1 <strong>and</strong> consider n <strong>points</strong> p 1 , . . . , p n . By induction<br />

hypothesis, there exists a homogeneous element s ∈ S, i.e., a hypersurface<br />

H := V (s) of a certain degree d, such that H ⊇ Z <strong>and</strong> p 1 , . . . , p n−1 ∉ H .<br />

We may assume p n ∈ H as, otherwise, the proof would be complete.<br />

Z ∪ {p 1 } ∪ . . . ∪ {p n−1 } is a Zariski closed subset of P N L not containing p n.<br />

Therefore, there exist some d ′ ∈Æ<strong>and</strong> some homogeneous s ′ ∈ S d ′ such that<br />

V (s ′ ) ⊇ Z ∪ {p 1 } ∪ . . . ∪ {p n−1 } but p n ∉ V (s ′ ). Every hypersurface<br />

V (a · s ′d + b · s d ′ ) for a, b ∈ L non-zero contains Z but neither p 1 , . . . , p n−1 ,<br />

nor p n .<br />

Second proof. Tensoring the canonical exact sequence<br />

0 −→ I {p1 , ... ,p n } −→ O X −→ O {p1 , ... ,p n } −→ 0<br />

with the ideal sheaf I Z yields an exact sequence<br />

0 −→ I {p1 , ... ,p n }∪Z −→ I Z −→ O {p1 , ... ,p n } −→ 0


Sec. 3] GALOIS DESCENT 17<br />

of coherent sheaves on P N L .<br />

For each d ∈, we tensor with the invertible sheaf O(d) <strong>and</strong> find a long<br />

cohomology exact sequence<br />

Γ(P N L , I Z(d)) −→ Γ(P N L , O {p 1 , ... ,p n }(d)) −→ H 1 (P N L , I {p 1 , ... ,p n }∪Z(d)) .<br />

By Serre’s vanishing theorem [Ha77, Theorem III.5.2],<br />

for d ≫ 0. Hence, there is a surjection<br />

H 1 (P N L , I {p 1 , ... ,p n }∪Z(d)) = 0<br />

Γ(P N L , I Z (d)) −→ Γ(P N L , O {p1 , ... ,p n }(d)) ∼ = κ(p 1 ) ⊕ · · · ⊕ κ(p n ) .<br />

That means that there exists a global section s of I Z (d) which does not vanish<br />

in any of the <strong>points</strong> p 1 , . . . , p n . The section s defines a hypersurface of degree<br />

d in P N L containing Z but not containing any of the <strong>points</strong> p 1 , . . . , p n . □<br />

3.12. Lemma (A. Grothendieck <strong>and</strong> J. Dieudonné). —–<br />

LetL/K beafinitefield extension<strong>and</strong> X beaK-schemesuchthatX × SpecK SpecL is<br />

i) reduced,<br />

ii) irreducible,<br />

iii) quasi-compact,<br />

iv) locally of finite type,<br />

v) of finite type,<br />

vi) locally Noetherian,<br />

vii) Noetherian,<br />

viii) proper,<br />

ix) quasi-projective,<br />

x) projective,<br />

xi) affine,<br />

or<br />

xii) regular.<br />

Then, X admits the same property.<br />

Proof. Let π : X × SpecK SpecL → X denote the canonical morphism. For iii)<br />

through xi), we may assume L/K to be Galois. Put G := Gal(L/K ).<br />

i) If 0 ≠ s ∈ Γ(U,O X ) were a nilpotent section of the structure sheaf over some<br />

open subset U ⊆ X then π ♯ (s) ∈ Γ(U × SpecK SpecL, O X ×SpecK SpecL) would be a<br />

nilpotent non-zero local section of the structure sheaf of X × SpecK SpecL.


18 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

ii) If X = X 1 ∪ X 2 were a decomposition into two closed proper subschemes<br />

then X × SpecK SpecL = X 1 × SpecK SpecL ∪ X 2 × SpecK SpecL would be, too.<br />

iii) Let {U α | α ∈ A } be an arbitrary affine open covering of X. Then,<br />

{U α × SpecK SpecL | α ∈ A }<br />

is an affine open covering of X × SpecK SpecL. Quasi-compactness guarantees<br />

the existence of a finite sub-covering {U α × SpecK SpecL | α ∈ A 0 }.<br />

Then, {U α | α ∈ A 0 } is a finite affine open covering for X.<br />

iv) We may assume X = SpecA to be affine. Then, by assumption, A ⊗ K L<br />

is a finitely generated L-algebra. Let {b 1 , . . . , b n } be a system of generators<br />

for A ⊗ K L. Decomposing into elementary tensors, if necessary, we may<br />

assume without restriction that all b i = a i ⊗ l i are elementary. Consider the<br />

homomorphism<br />

π : K [X 1 , . . . , X n ] → A<br />

of K-algebras given by X i ↦→ a i . π induces a surjection after tensoring with L.<br />

It is, therefore, surjective as ⊗ K L is a faithful functor.<br />

v) This is just the combination of iii) <strong>and</strong> iv).<br />

vi) Let X α<br />

∼ = SpecAα be an affine open subscheme of X. We have to<br />

show that A α is Noetherian under the hypothesis that A α ⊗ K L is. Assume<br />

I 1 ⊂ I 2 ⊂ I 3 ⊂ . . . is an ascending chain of ideals in A α which does<br />

not stabilize. It induces an ascending chain I 1 ⊗ K L ⊂ I 2 ⊗ K L ⊂ I 3 ⊗ K L ⊂ . . .<br />

of ideals in A α ⊗ K L that does not stabilize, either. This is a contradiction.<br />

vii) This is the combination of iii) <strong>and</strong> vi).<br />

viii) By virtue of v), X is of finite type. We apply the valuative criterion<br />

for properness. Consider a commutative diagram<br />

U<br />

i<br />

T<br />

X<br />

SpecK<br />

where T = SpecR is the spectrum of a valuation ring, U = SpecF is the<br />

spectrum of its quotient field <strong>and</strong> i is the canonical morphism. Taking the base<br />

change to SpecL, we find<br />

U × SpecK SpecL<br />

X × SpecK SpecL<br />

ι<br />

i<br />

<br />

T × SpecK SpecL<br />

SpecL .<br />

Here, the morphism ι in the diagonal is the unique one making the diagram commute.<br />

Note that U × SpecK SpecL = Spec(F ⊗ K L) is no longer the spectrum


Sec. 3] GALOIS DESCENT 19<br />

of a field but rather the union of finitely many spectra of fields. Similarly,<br />

T × SpecK SpecL = Spec(R ⊗ K L) is a finite union of spectra of valuation rings.<br />

Further, there is a natural G-operation on the whole diagram without ι. As ι is<br />

uniquely determined by the condition that it makes the diagram commute, the<br />

diagrams<br />

ι<br />

T × SpecK SpecL X × SpecK SpecL<br />

σ<br />

T × SpecK SpecL<br />

ι<br />

σ<br />

X × SpecK SpecL<br />

are commutative for each σ ∈ G. By Proposition 3.8, ι is the base change of a<br />

morphism T → X.<br />

ix) <strong>and</strong> x) Taking v) <strong>and</strong> viii) into account, everything left to be shown<br />

is the existence an ample invertible sheaf on X. By assumption, there<br />

is an ample invertible sheaf M on X × SpecK SpecL. For σ ∈ G let<br />

x σ : X × SpecK SpecL → X × SpecK SpecL be the morphism of schemes induced by<br />

S(σ): SpecL → SpecL, i.e., by σ −1 on coordinate rings. The invertible sheaves<br />

x ∗ σ M are ample, too. Therefore, M := N υ∈G x ∗ υM is an ample invertible sheaf.<br />

For each σ ∈ G, there are canonical identifications<br />

ι σ : x ∗ σ M = O υ∈G<br />

x ∗ σ x∗ υ M = O υ∈G<br />

x ∗ υσ M<br />

id<br />

−→ O υ∈G<br />

x ∗ υ M = M .<br />

Obviously, these are compatible in the sense that, for each σ, τ ∈ G, there is<br />

the relation ι τ ◦x ∗ τ(ι σ ) = ι στ . By Galois descent for locally free sheaves, there is<br />

an invertible sheaf L ∈ Pic(X ) such that π ∗ L ∼ = M . Lemma 3.13 shows that<br />

L is ample.<br />

xi) We suppose that X × SpecK SpecL ∼ =<br />

−→ SpecB is affine. The isomorphism<br />

defines a G-operation from the right on the L-algebra B such that each σ ∈ G<br />

operates σ −1 -linearly. Define a G-operation from the left on B by σ·b := b·σ −1 .<br />

Proposition 3.3 provides us a K-algebra A such that there is an isomorphism<br />

X × SpecK SpecL ∼ =<br />

∼=<br />

−→ SpecB = Spec(A ⊗ K L) −→ SpecA × SpecK SpecL ,<br />

compatible with the G-operations. Proposition 3.5 implies X ∼ = SpecA.<br />

xii) Let (A, m) be the local ring at a point p ∈ X. By vi), we may assume (A, m)<br />

is Noetherian. We have to show A is regular. Let us give two proofs for that,<br />

the st<strong>and</strong>ard one using Serre’s homological characterization of regularity <strong>and</strong> an<br />

elementary one.<br />

First proof. Assume, by contradiction, (A, m) were not regular. Then, [Mat,<br />

Theorem 19.2] shows gl. dim A = ∞. Further, by [Mat, §19, Lemma 1],


20 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

proj. dim A<br />

A/m = ∞ <strong>and</strong><br />

On the other h<strong>and</strong>, we have<br />

sup {i | Tor A i (A/m, A/m) ≠ 0 } = ∞.<br />

Tor A i (A/m, A/m) ⊗ K L = Tor A i (A/m, A/m) ⊗ A (A ⊗ K L)<br />

= Tor A i (A/m, (A ⊗ K L/m ⊗ K L))<br />

= Tor A⊗ KL<br />

i ((A ⊗ K L/m ⊗ K L), (A ⊗ K L/m ⊗ K L))<br />

= 0<br />

for i > dim A ⊗ K L as A ⊗ K L is regular. This is a contradiction.<br />

Note that we only used that π is faithfully flat.<br />

Second proof. Put d := dim A. The ring A ⊗ K L is not local, in general.<br />

But the quotient (A ⊗ K L)/(m⊗ K L) ∼ = A/m⊗ K L is a direct product of finitely<br />

many fields since L/K is a finite, separable field extension. The quotients<br />

(A ⊗ K L)/(m n ⊗ K L) are Artin rings as they are flat of relative dimension zero<br />

over A/m n . Consequently, they are direct products of Artin local rings,<br />

(A ⊗ K L)/(m n ⊗ K L)<br />

∼=<br />

−→ A (n)<br />

1<br />

× · · · × A (n)<br />

l .<br />

Under this isomorphism, (m ⊗ K L)/(m n ⊗ K L) is mapped to the product of the<br />

maximal ideals m (n)<br />

1<br />

× · · · × m (n)<br />

l as it is a nilpotent ideal <strong>and</strong> the quotient has to<br />

be a direct product of fields.<br />

For each i ∈ {1, . . . , l}, let f i ∈ A ⊗ K L be the element which maps to the<br />

st<strong>and</strong>ard vector e i = (0, . . . , 0, 1, 0, . . . , 0). Then,<br />

A (n)<br />

i<br />

∼= ((A ⊗ K L)/(m n ⊗ K L)) fi<br />

∼ = (A ⊗K L) fi /m n (A ⊗ K L) fi .<br />

By assumption, (A ⊗ K L) fi is a regular local ring <strong>and</strong> m(A ⊗ K L) fi is its maximal<br />

ideal. (A ⊗ K L) fi is of dimension d as it is flat of relative dimension zero<br />

over A. The Hilbert-Samuel function of a regular local ring is st<strong>and</strong>ard. We have<br />

length ( A (n)<br />

i<br />

) = length ((A ⊗K L) fi /m n (A ⊗ K L) fi )<br />

= length ((A ⊗ K L) fi /(m(A ⊗ K L) fi ) n )<br />

( ) n + d − 1<br />

= .<br />

d<br />

Consequently, for the function H A such that H A (n) := dim K (A/m n ), one gets<br />

(<br />

H A (n) = dim L (A ⊗K L)/(m n ⊗ K L) )<br />

( )<br />

= dim L A<br />

(n)<br />

1<br />

× · · · × A (n)<br />

l


Sec. 3] GALOIS DESCENT 21<br />

=<br />

=<br />

l<br />

∑<br />

i=1<br />

l<br />

∑<br />

i=1<br />

( )<br />

dim L A<br />

(n)<br />

i<br />

[(<br />

A<br />

(n)<br />

i<br />

/m (n)<br />

i<br />

) ] ( )<br />

: L · length A<br />

(n)<br />

i .<br />

)<br />

. How-<br />

This shows that H A is a constant multiple of n ↦→ ( n+d−1<br />

d<br />

ever, since H A (1) = 1, there is no ambiguity about constant factors. A is regular.<br />

□<br />

3.13. Lemma. –––– Let L/K be an arbitrary field extension, X a K-scheme of<br />

finite type, π : X × SpecK SpecL → X the canonical morphism <strong>and</strong> L ∈ Pic(X ) an<br />

invertible sheaf. Suppose that the pull-back π ∗ L ∈ Pic(X × SpecK SpecL) is ample.<br />

Then, L is ample.<br />

Proof. We have to prove that, for every coherent sheaf F on X <strong>and</strong> every<br />

closed point x ∈ X, the canonical map px n : Γ(X, F ⊗ L ⊗n ) → F x ⊗ Lx<br />

⊗n is<br />

surjective for n ≫ 0. For this, it is obviously sufficient to prove surjectivity of<br />

But, as is easy to see,<br />

p n x ⊗ K L : Γ(X, F ⊗ L ⊗n ) ⊗ K L → (F x ⊗ L ⊗n<br />

x ) ⊗ K L .<br />

Γ(X, F ⊗ L ⊗n ) ⊗ K L = Γ(X × SpecK SpecL, π ∗ F ⊗ π ∗ L ⊗n )<br />

while (F x ⊗ L n<br />

x ) ⊗ K L = Γ(π −1 (x), π ∗ F ⊗ π ∗ L ⊗n ). Here, π −1 (x) denotes the<br />

fiber of π above x. For n ≫ 0, the map p n x ⊗ K L is surjective as π ∗ L is ample.<br />

□<br />

3.14. Lemma. –––– Let R be a ring, F ≠ 0 a free R-module of finite rank <strong>and</strong> M<br />

an arbitraryR-module. Suppose that M⊗ R F is a locally free R-module of finite rank.<br />

Then, M is locally free of finite rank.<br />

Proof. The R-module M⊗ R F is locally free of finite rank. Therefore, there exists<br />

some affine open covering {SpecR f1 , . . . , SpecR fn } of SpecR such that each<br />

(M ⊗ R F )⊗ R R fi = (M ⊗ R R fi ) ⊗ R F<br />

is a free R fi -module. M ⊗ R R fi = M fi is a direct summ<strong>and</strong>. Therefore, M fi is a<br />

projective R fi -module.<br />

We have a surjection (M ⊗ R F )⊗ R R fi ։ M fi , the kernel K of which is a direct<br />

summ<strong>and</strong> of (M ⊗ R F ) ⊗ R R fi , too. In particular, K is a finitely generated<br />

R fi -module. Hence, M fi is finitely presented <strong>and</strong>, therefore, locally free by<br />

[Mat, Theorem 7.12 <strong>and</strong> Theorem 4.10].<br />


22 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

3.15. Remark. –––– Galois descent is a central technique in André Weil’s foundation<br />

of Algebraic Geometry. The Grothendieck school gave a far-reaching<br />

generalization of it, the so-called faithful flat descent. It turns out that Galois<br />

descent is sufficient for our presentation of the relationship between central simple<br />

algebras <strong>and</strong> Brauer-Severi varieties. For that reason, we decided to present<br />

it in detail. Faithful flat descent will play a main role in the next chapter.<br />

4. Central simple algebras <strong>and</strong> non-abelian H 1<br />

We are going to make use of the following well-known facts about central<br />

simple algebras.<br />

4.1. Lemma (J. H. Maclagan-Wedderburn, R. Brauer). —– Let K be a field.<br />

a) Let A be a central simple algebra over K. Then, there exist a skew field D with<br />

center K <strong>and</strong> a natural number n such that A ∼ = M n (D) is isomorphic to the full<br />

algebra of n × n-matrices with entries in D.<br />

b) Let L be a field extension of K <strong>and</strong> A be a central simple algebra over K. Then,<br />

A ⊗ K L is a central simple algebra over L.<br />

c) Assume K to be separably closed. Let D be a skew field being finite dimensional<br />

over K whose center is equal to K. Then, D = K.<br />

Proof. See the st<strong>and</strong>ard literature, for example [Lan93], [Bou-A], or [Ke].<br />

4.2. Remarks. –––– a) Let A be a central simple algebra over a field K.<br />

i) The proof of Lemma 4.1.a) shows that in the presentation A ∼ = M n (D) the<br />

skew field D is unique up to isomorphism of K-algebras <strong>and</strong> the natural number<br />

n is unique.<br />

ii) A ⊗ K K sep is isomorphic to a full matrix algebra over K sep . In particular,<br />

dim K A is a perfect square. The natural number ind(A) := √ dim K (D) is called<br />

the index of A.<br />

b) Let A 1 , A 2 be central simple algebras over a field K. Then A 1 ⊗ K A 2 can be<br />

shown to be a central simple algebra over K. Further, if A is a central simple<br />

algebra over a field K then A ⊗ K A op ∼ = Aut K-Vect (A). I.e., it is isomorphic to a<br />

matrix algebra.<br />

c) Two central simple algebras A 1<br />

∼ = Mn1 (D 1 ), A 2<br />

∼ = Mn2 (D 2 ) over a field K are<br />

said to be similar if the corresponding skew fields D 1 <strong>and</strong> D 2 are isomorphic<br />

as K-algebras. This is an equivalence relation on the set of all isomorphism<br />

classes of central simple algebras over K. The tensor product induces a group<br />

structure on the set of similarity classes of central simple algebras over K, this<br />

is the so-called Brauer group Br(K ) of the field K.<br />


Sec. 4] CENTRAL SIMPLE ALGEBRAS AND NON-ABELIAN h 1 23<br />

4.3. Definition. –––– Let K be a field <strong>and</strong> A be a central simple algebra over K.<br />

A field extension L of K admitting the property that A ⊗ K L is isomorphic to<br />

a full matrix algebra is said to be a splitting field for A. In this case, one says that<br />

A splits over L.<br />

4.4. Lemma (Theorem of Skolem-Noether). —–<br />

Let R be a commutative ring with unit. Then, GL n (R) operates on M n (R) by<br />

conjugation,<br />

(g, m) ↦→ gmg −1 .<br />

If R = L is a field then this defines an isomorphism<br />

PGL n (L) := GL n (L)/L ∗ ∼ =<br />

−→ Aut L (M n (L)) .<br />

Proof. One has L = Zent(M n (L)). Therefore, the mapping is well-defined <strong>and</strong><br />

injective.<br />

Surjectivity. Let j : M n (L) → M n (L) be an automorphism. We consider the<br />

algebra<br />

M := M n (L) ⊗ L M n (L) op ( ∼ = M n 2(L)).<br />

M n (L) gets equipped with the structure of a left M-module in two ways.<br />

(A ⊗ B) • 1 C := A · C · B<br />

(A ⊗ B) • 2 C := j(A) · C · B<br />

Two M n 2(L)-modules of the same L-dimension are isomorphic, as the<br />

n 2 -dimensional st<strong>and</strong>ard L-vector space equipped with the canonical operation<br />

of M n 2(L) is the only simple left M n 2(L)-module <strong>and</strong> there are no non-trivial<br />

extensions. Thus, there is an isomorphism h : (M n (L), • 1 ) → (M n (L), • 2 ).<br />

Let us put I := h(E) to be the image of the identity matrix.<br />

M ∈ M n (L) we have<br />

h(M ) = h((E ⊗ M ) • 1 E) = (E ⊗ M ) • 2 h(E) = h(E) · M = I · M.<br />

In particular, I ∈ GL n (L). Therefore,<br />

I · M = h(M ) = h((M ⊗ E) • 1 E) = (M ⊗ E) • 2 h(E) = j(M ) · I<br />

For every<br />

for each M ∈ M n (L) <strong>and</strong> j(M ) = IMI −1 .<br />

□<br />

4.5. Definition. –––– Let n be a natural number.<br />

i) If K is a field then we will denote by Az K n the set of all isomorphism classes<br />

of central simple algebras A of dimension n 2 over K.


24 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

ii) Let L/K be a field extension. Then Azn<br />

L/K will denote the set of all isomorphism<br />

classes of central simple algebras A which are of dimension n 2 over K<br />

<strong>and</strong> split over L. Obviously, Az K n := S Az L/K<br />

L/K<br />

n .<br />

4.6. Theorem (cf. J.-P. Serre: Corps locaux [Se62, chap. X, §5]). —–<br />

Let L/K be a finite Galois extension of fields, G := Gal(L/K ) its Galois group,<br />

<strong>and</strong> n ∈Æ.<br />

Then, there is a natural bijection of pointed sets<br />

a = a L/K<br />

n<br />

: Az L/K<br />

n<br />

∼=<br />

−→ H 1 (G, PGL n (L)) ,<br />

A ↦→ a A .<br />

Proof. Let A be a central simple algebra over K which splits over L,<br />

The diagrams<br />

A ⊗ K L ∼ =<br />

−→ M n (L) .<br />

f<br />

A ⊗ K L f M n (L)<br />

σ<br />

σ<br />

do not commute, in general.<br />

A ⊗ K L f M n (L)<br />

For each σ ∈ G, define a σ ∈ PGL n (L) by putting ( f ◦ σ) = a σ ◦ (σ ◦ f ).<br />

It turns out that<br />

f ◦ στ = ( f ◦ σ) ◦ τ<br />

= a σ ◦ (σ ◦ f ) ◦ τ<br />

= a σ ◦ σ ◦ ( f ◦ τ )<br />

= a σ ◦ σ ◦ (a τ ◦ (τ ◦ f ))<br />

= a σ ◦ σ a τ ◦ (στ ◦ f ) .<br />

I.e., a στ = a σ · σa τ <strong>and</strong> (a σ ) σ∈G is a cocycle.<br />

If one starts with another isomorphism f ′ : A⊗ K L −→ M n (L) then there exists<br />

some b ∈ PGL n (L) such that f = b ◦ f ′ . The equality ( f ◦ σ) = a σ ◦ (σ ◦ f )<br />

implies<br />

f ′ ◦ σ = b −1 ◦ f ◦ σ = b −1 · a σ ◦ (σ ◦ (b ◦ f ′ )) = b −1 · a σ · σb ◦ (σ ◦ f ′ ) .<br />

Thus, the isomorphism f ′ yields a cocycle cohomologous to (a σ ) σ∈G . The<br />

mapping a is well-defined.


Sec. 4] CENTRAL SIMPLE ALGEBRAS AND NON-ABELIAN h 1 25<br />

Injectivity. Assume that A <strong>and</strong> A ′ are chosen in such a way that the construction<br />

above yields one <strong>and</strong> the same cohomology class a A = a A ′ ∈ H 1 (G, PGL n (L)).<br />

After choosing suitable isomorphisms f <strong>and</strong> f ′ , one has the equalities<br />

( f ◦ σ) = a σ ◦ (σ ◦ f ) <strong>and</strong> ( f ′ ◦ σ) = a σ ◦ (σ ◦ f ′ ) in the diagram<br />

A ⊗ K L f f<br />

M n (L) ′<br />

A ′ ⊗ K L<br />

σ<br />

A ⊗ K L f<br />

σ<br />

M n (L)<br />

f ′<br />

σ<br />

A ′ ⊗ K L .<br />

Consequently, f ◦ σ ◦ f −1 ◦ σ −1 = f ′ ◦ σ ◦ f ′−1 ◦ σ −1 <strong>and</strong>, therefore,<br />

f ◦ σ ◦ f −1 ◦ f ′ ◦ σ −1 ◦ f ′−1 = id .<br />

The outer part of the diagram commutes. Taking the G-invariants on both sides<br />

yields A ∼ = A ′ .<br />

Surjectivity. Let a cocycle (a σ ) σ∈G for H 1 (G, PGL n (L)) be given. We define a<br />

new G-operation on M n (L) as follows.<br />

Let σ ∈ G act as<br />

a σ ◦ σ : M n (L)<br />

σ a<br />

−→ M n (L) −→<br />

σ<br />

Mn (L) .<br />

Note that this is a σ-linear mapping. Further, one has<br />

(a σ ◦ σ) ◦ (a τ ◦ τ ) = a σ ◦ σ a τ ◦ στ = a στ ◦ στ .<br />

I.e., we constructed a group operation from the left. Galois descent yields the<br />

desired algebra.<br />

□<br />

4.7. Corollary. –––– Let L/K be a finite Galois extension of fields <strong>and</strong> let n be a<br />

natural number.<br />

a) Let L ′ be a field extension of L such that L ′ /K is Galois, too. Then, the following<br />

diagram of morphisms of pointed sets commutes,<br />

Az L/K<br />

n<br />

a L/K<br />

n<br />

H 1 (Gal (L/K ), PGL n (L))<br />

nat. incl.<br />

inf Gal (L′ /K)<br />

Gal(L/K)<br />

Az L′ /K<br />

n<br />

a L′ /K<br />

n<br />

H 1( Gal(L ′ /K ), PGL n (L ′ ) ) .


26 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

b)LetK ′ beanintermediatefield oftheextensionL/K. Then,thefollowingdiagram<br />

of morphisms of pointed sets commutes,<br />

Az L/K<br />

n<br />

a L/K<br />

n<br />

H 1( Gal(L/K ), PGL n (L) )<br />

⊗ K K ′ <br />

res Gal(L/K′ )<br />

Gal(L/K)<br />

Az L/K ′<br />

n<br />

a L/K ′<br />

n<br />

H 1( Gal(L/K ′ ), PGL n (L) ) .<br />

Proof. These are direct consequences of the construction of the bijections a ∗ n .<br />

□<br />

4.8. Corollary. –––– Let K be a field <strong>and</strong> n be a natural number. Then, there is a<br />

unique natural bijection<br />

a = a K n : Az K n −→ H 1( Gal(K sep /K ), PGL n (K sep ) )<br />

such that a K n | Az<br />

L/K<br />

n<br />

= a L/K<br />

n for each finite Galois extension L/K in K sep .<br />

Proof. In order to get connected to the definition of the cohomology of a<br />

profinite group, only one technical point is to be proven. That is the formula<br />

PGL n (K sep ) Gal(K sep /K ′) = PGL n (K ′ )<br />

for K ⊆ K ′ ⊆ K sep any intermediate field. To see this, observe that the<br />

exact sequence<br />

1 −→ (K sep ) ∗ −→ GL n (K sep ) −→ PGL n (K sep ) −→ 1<br />

induces a long exact sequence<br />

1→(K ′ ) ∗ → GL n (K ′ ) → PGL n (K sep ) Gal(K sep /K ′) → H 1( Gal(K sep /K ′ ), (K sep ) ∗)<br />

in cohomology. Finally, the right entry vanishes by Hilbert’s Theorem 90.<br />

Cf. Lemma 5.11.<br />


Sec. 4] CENTRAL SIMPLE ALGEBRAS AND NON-ABELIAN h 1 27<br />

4.9. Proposition. –––– Let K be a field <strong>and</strong> m <strong>and</strong> n be natural numbers.<br />

Then, there is a commutative diagram<br />

Az K n<br />

a K n<br />

H 1( Gal(K sep /K ), PGL n (K sep ) )<br />

A↦→M m (A)<br />

Az K mn<br />

(i n mn) ∗<br />

a K mn<br />

H 1( Gal(K sep /K ), PGL mn (K sep ) ) .<br />

Here, (imn) n ∗ is the map induced by the block-diagonal embedding<br />

imn n : PGL n (K sep ) −→ PGL mn (K sep )<br />

⎛ ⎞<br />

E 0 · · · 0<br />

0 E · · · 0<br />

E ↦→ ⎜<br />

⎝ . .<br />

..<br />

⎟ . . ⎠ .<br />

0 0 · · · E<br />

Proof. Let A ∈ Az K n . By the construction above, a cycle representing<br />

the cohomology class a K n (A) is given as follows. Choose an isomorphism<br />

f : A ⊗ K K sep → M n (K sep ) <strong>and</strong> put a σ := ( f ◦ σ) ◦ (σ ◦ f ) −1 ∈ Aut(M n (K sep ))<br />

for each σ ∈ Gal(K sep /K ).<br />

On the other h<strong>and</strong>, for M m (A) ∈ Az K mn one may choose the isomorphism<br />

M m ( f ): M m (A) ⊗ K K sep = M m (A ⊗ K K sep ) −→ M m (M n (K sep )) ∼ = M mn (K sep ).<br />

For each σ ∈ Gal(K sep /K ), this yields the automorphism ã σ of M m (M n (K sep ))<br />

which operates as a σ on each block. If a σ is given by conjugation with a matrix<br />

A σ then ã σ is given by conjugation with<br />

⎛ ⎞<br />

A σ 0 · · · 0<br />

0 A σ · · · 0<br />

⎜<br />

⎝ . .<br />

..<br />

⎟ . . ⎠ .<br />

0 0 · · · A σ<br />

This is exactly what was to be proven.<br />

□<br />

4.10. Remark. –––– The proposition above shows<br />

Br(K ) ∼ = lim −→n<br />

H 1( Gal(K sep /K ), PGL n (K sep ) ) .


28 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

Further, for each m <strong>and</strong> n, there is a commutative diagram of exact sequences<br />

as follows,<br />

1 (K sep ) ∗ GL n (K sep )<br />

PGL n (K sep )<br />

1<br />

1 (K sep ) ∗ GL mn (K sep ) PGL mn (K sep ) 1 .<br />

We note that (K sep ) ∗ is mapped into the centers of GL n (K sep ) <strong>and</strong> GL mn (K sep ), respectively.<br />

Therefore, there are boundary maps to the second group cohomology<br />

group <strong>and</strong> they are compatible with each other to give a map<br />

lim H<br />

−→ 1( Gal(K sep /K ), PGL n (K sep ) ) −→ H 2( Gal(K sep /K ), (K sep ) ∗) .<br />

n<br />

It is not complicated to show that this map is injective <strong>and</strong> surjective. Cf. Corollary<br />

II.5.3.<br />

j n mn<br />

5. Brauer-Severi varieties <strong>and</strong> non-abelian H 1<br />

5.1. Definition. –––– Let K be a field.<br />

i) A scheme X over K is called a Brauer-Severi variety if there exists a finite,<br />

separable field extension L/K such that X × SpecK SpecL is isomorphic to a<br />

projective space P N L .<br />

ii) A field extension L of K admitting the property that X × SpecK SpecL ∼ = P N L<br />

for some n ∈Æis said to be a splitting field for X. In this case, one says that<br />

X splits over L.<br />

5.2. Proposition. –––– Let X be a Brauer-Severi variety over a field K. Then,<br />

i) X is a variety. I.e., X is reduced <strong>and</strong> irreducible as a scheme.<br />

ii) X is projective <strong>and</strong> regular.<br />

iii) X is geometrically integral.<br />

iv) One has Γ(X, O X ) = K.<br />

v) K is algebraically closed in the function field X (K ).<br />

Proof. We will denote the dimension of X by N .<br />

i) <strong>and</strong> ii) are direct consequences of Lemma 3.12.i),ii), x) <strong>and</strong> xii).<br />

iii) is clear from the definition.<br />

iv) We have<br />

Γ(X, O X ) ⊗ K K sep = Γ(X × SpecK SpecK sep , O X ×SpecK SpecK sep)<br />

= Γ(P N K sep, O P N K sep)<br />

= K sep .


Sec. 5] BRAUER-SEVERI VARIETIES AND NON-ABELIAN h 1 29<br />

Consequently, Γ(X, O X ) = K.<br />

v) Assume, g ∈ K (X ) is a <strong>rational</strong> function which is algebraic over K. We fix<br />

an algebraic closure K of K.<br />

The pull-back g K<br />

is a <strong>rational</strong> function on X × SpecK SpecK ∼ = P N which is<br />

K<br />

algebraic over K. Thus, g K<br />

is a constant function. Consequently, g itself has<br />

definitely no poles. I.e., g is a regular function on X. By iv), we see g ∈ K. □<br />

5.3. Lemma. –––– Let R be a commutative ring with unit.<br />

a) Then, there is an operation of the group GL n (R) on P n−1<br />

R<br />

R-schemes as follows.<br />

by morphisms of<br />

A ∈ GL n (R) gives rise to the morphism given by the graded automorphism<br />

R[X 0 , . . . , X n−1 ] −→ R[X 0 , . . . , X n−1 ]<br />

f (X 0 , . . . , X n−1 ) ↦→ f ((X 0 , X 1 , . . . , X n−1 ) · A t )<br />

of the homogeneous coordinate ring.<br />

b) If R = L is a field then this induces an isomorphism<br />

PGL n (L)<br />

∼=<br />

−→ Aut L-schemes (P n−1<br />

L ) .<br />

Proof. a) is clear. For b), see [Ha77, Example II.7.1.1]. The proof given there<br />

works equally well without the assumption on L to be algebraically closed. □<br />

5.4. Remark. –––– Let S be any commutative R-algebra with unit. Then, on<br />

the set P n−1<br />

R (S) of S-valued <strong>points</strong>, the definition above yields the naive operation<br />

⎛ ⎞<br />

a 11 . . . a 1n<br />

⎜<br />

⎝<br />

.<br />

. ..<br />

⎟ . ⎠<br />

a n1 . . . a nn<br />

⊤<br />

↓<br />

(x 0 : . . . : x n−1 ) ↦→ ( (a 11 x 0 + . . . + a 1n x n−1 ) : . . . : (a n1 x 0 + . . . + a nn x n−1 ) )<br />

of GL n (S).<br />

5.5. Definition. –––– Let r be natural number.<br />

i) If K is a field then we will denote by BS K r the set of all isomorphism classes of<br />

Brauer-Severi varieties X of dimension r over K.<br />

ii) Let L/K be a field extension. Then, BS L/K<br />

r will denote the set of all isomorphism<br />

classes of Brauer-Severi varieties X over K which are of dimension r <strong>and</strong><br />

split over L. Obviously, BS K r := S BS L/K<br />

r .<br />

L/K


30 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

5.6. Theorem (cf. J.-P. Serre: Corps locaux [Se62, chap. X, §6). —–<br />

Let L/K be a finite Galois extension, G := Gal(L/K ) its Galois group, <strong>and</strong> n be<br />

a natural number. Then, there exists a natural bijection of pointed sets<br />

α = α L/K<br />

n−1 : BSL/K n−1<br />

∼=<br />

−→ H 1 (G, PGL n (L)),<br />

X ↦→ α X .<br />

Proof. Let X be a Brauer-Severi variety over K that splits over L,<br />

X × SpecK SpecL ∼ =<br />

−→ P n−1<br />

L .<br />

On X × SpecK SpecL, as well as on P n−1<br />

L , there are operations of G from<br />

the left by morphisms of K-schemes. The action of σ ∈ G is induced by<br />

S(σ): SpecL → SpecL in both cases. Unfortunately, the diagrams<br />

do not commute, in general.<br />

X × SpecK SpecL<br />

σ<br />

X × SpecK SpecL<br />

f<br />

f<br />

f<br />

P n−1<br />

L<br />

σ<br />

P n−1<br />

L<br />

For each σ ∈ G, define α σ ∈ PGL n (L) by putting ( f ◦ σ) = α σ ◦ (σ ◦ f ).<br />

It turns out that<br />

I.e., (α σ ) σ∈G is a cocycle.<br />

f ◦ στ = ( f ◦ σ) ◦ τ<br />

= α σ ◦ (σ ◦ f ) ◦ τ<br />

= α σ ◦ σ ◦ ( f ◦ τ )<br />

= α σ ◦ σ ◦ (α τ ◦ (τ ◦ f ))<br />

= α σ ◦ σ α τ ◦ (στ ◦ f ) .<br />

If one starts with another isomorphism f ′ : X × SpecK SpecL −→ P n−1<br />

L then<br />

there exists some b ∈ PGL n (L) such that f = b ◦ f ′ . The equality<br />

( f ◦ σ) = α σ ◦ (σ ◦ f ) implies<br />

f ′ ◦ σ = b −1 ◦ f ◦ σ = b −1 · α σ ◦ (σ ◦ (b ◦ f ′ )) = b −1 · α σ · σb ◦ (σ ◦ f ′ ) .<br />

Thus, the isomorphism f ′ yields a cocycle cohomologous to (α σ ) σ∈G . By consequence,<br />

the mapping α is well-defined.<br />

Injectivity. Assume X <strong>and</strong> X ′ are chosen such that the same cohomology class<br />

α X = α X ′ ∈ H 1 (G, PGL n (L)) arises. After choosing suitable isomorphisms f<br />

<strong>and</strong> f ′ , one has the formulas ( f ◦σ) = α σ ◦ (σ ◦ f ) <strong>and</strong> ( f ′ ◦σ) = α σ ◦ (σ ◦ f ′ )


Sec. 5] BRAUER-SEVERI VARIETIES AND NON-ABELIAN h 1 31<br />

in the diagram<br />

X × SpecK SpecL<br />

σ<br />

X × SpecK SpecL<br />

f<br />

f<br />

P n−1<br />

L<br />

σ<br />

P n−1<br />

L<br />

f ′<br />

f ′<br />

X ′ × SpecK SpecL<br />

σ<br />

X ′ × SpecK SpecL .<br />

Therefore, f ◦ σ ◦ f −1 ◦ σ −1 = f ′ ◦ σ ◦ f ′−1 ◦ σ −1 <strong>and</strong>, consequently,<br />

f ◦ σ ◦ f −1 ◦ f ′ ◦ σ −1 ◦ f ′−1 = id .<br />

The outer part of the diagram commutes. Galois descent yields X ∼ = X ′ .<br />

Surjectivity. Let a cocycle (α σ ) σ∈G for H 1 (G, PGL n (L)) be given. We define a<br />

G-operation on P n−1<br />

L by letting σ ∈ G operate as<br />

α σ ◦ σ : P n−1<br />

L<br />

σ<br />

−→ P n−1<br />

L<br />

α σ<br />

−→ P<br />

n−1<br />

L .<br />

This is a group operation as (α σ ) σ∈G is a cocycle. The geometric version of<br />

Galois descent yields the desired variety.<br />

□<br />

5.7. Corollary. –––– Let L/K be a finite Galois extension of fields <strong>and</strong> n be a<br />

natural number.<br />

a) Let L ′ be a field extension of L such that L ′ /K is Galois, too. Then, the following<br />

diagram of morphisms of pointed sets commutes.<br />

BS L/K<br />

n−1<br />

α L/K<br />

n−1<br />

H 1( Gal(L/K ), PGL n (L) )<br />

nat. incl.<br />

inf Gal(L′ /K)<br />

Gal(L/K)<br />

BS L′ /K<br />

n−1<br />

α L′ /K<br />

n−1<br />

H 1( Gal(L ′ /K ), PGL n (L ′ ) )<br />

b) LetK ′ beanintermediatefieldoftheextensionL/K. Then,thefollowingdiagram<br />

of morphisms of pointed sets commutes.<br />

BS L/K<br />

n−1<br />

α L/K<br />

n−1<br />

H 1( Gal(L/K ), PGL n (L) )<br />

× SpecK SpecK ′ <br />

BS L/K ′<br />

n−1<br />

α L/K ′<br />

res Gal(L/K′ )<br />

Gal(L/K)<br />

n−1<br />

H 1( Gal(L/K ′ ), PGL n (L) )<br />

Proof. These are direct consequences of the construction of the mappings α ∗ n−1.<br />


32 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

5.8. Corollary. –––– Let K be a field <strong>and</strong> n be a natural number. Then, there is a<br />

natural bijection<br />

α = α K n−1 : BSK n−1 −→ H 1( Gal(K sep /K ), PGL n (K sep ) )<br />

such that α K n−1 | BS<br />

L/K<br />

n−1<br />

= α L/K<br />

n−1 for each finite Galois extension L/K in K sep .<br />

Proof. The proof follows the same lines as the proof of Corollary 4.8.<br />

□<br />

5.9. Proposition (F. Severi, cf. J.-P. Serre: [Se62, chap. X, §6, Exc. 1]). —–<br />

Let r ∈Æ. Assume that X is a Brauer-Severi variety of dimension r over a field K.<br />

If X (K ) ≠ ∅ then X ∼ = P r K .<br />

Proof. Let L/K be a finite Galois extension which is a splitting field for X.<br />

Denote its Galois group by G. Choose an isomorphism<br />

X × SpecK SpecL ∼ =<br />

−→ P r L .<br />

Here, one may assume without restriction that the L-valued point<br />

x L ∈ X × SpecK SpecL (L)<br />

induced by the K-valued point x ∈ X (K ) is mapped to (1 : 0 : . . . : 0) ∈ P r L(L).<br />

Therefore, the cohomology class α X ∈ H 1( G, PGL r+1 (L) ) is given by a cocycle<br />

(α σ ) σ∈G such that every α σ admits (1 : 0 : . . . : 0) as a fixed point.<br />

Consequently, α X belongs to the image of H 1( G, F/L ∗) under the natural<br />

homomorphism where<br />

⎧⎛<br />

⎞<br />

⎫<br />

⎪⎨ a 11 . . . a 1,r+1<br />

⎪⎬<br />

⎜<br />

F := ⎝<br />

.<br />

. ..<br />

⎟<br />

. ⎠ ∈ GL r+1 (L)<br />

a 21 = a 31 = . . . = a r+1,1 = 0<br />

⎪⎩<br />

a r+1,1 . . . a r+1,r+1<br />

∣<br />

⎪⎭ .<br />

But it is obvious that<br />

⎧⎛<br />

⎞⎫<br />

1 a 12 . . . a 1,r+1<br />

⎪⎨<br />

F/L ∗ ∼ = F ′ 0 a 22 . . . a 2,r+1<br />

⎪⎬<br />

:= ⎜<br />

⎝<br />

.<br />

. . ..<br />

⎟ ⊆ GL r+1 (L) .<br />

. ⎠<br />

⎪⎩<br />

⎪⎭<br />

0 a r+1,2 . . . a r+1,r+1<br />

Further, the natural homomorphism H 1 (G, F ′ ) → H 1 (G, PGL r+1 (L)) factors<br />

via H 1 (G, GL r+1 (L)). The assertion follows from Lemma 5.11. □<br />

5.10. Remark. –––– One can easily show that even H 1 (G, F ′ ) = 0. Indeed,<br />

F 1 := { (a ij ) 1≤i,j≤r+1 | a ij = 0 for i ≥ 2, i ≠ j;<br />

a 11 = a 22 = · · · = a r+1,r+1 = 1 }


Sec. 5] BRAUER-SEVERI VARIETIES AND NON-ABELIAN h 1 33<br />

is a normal G-subgroup of F ′ . Thus, F ′ admits a filtration by G-sub<strong>groups</strong>,<br />

each normal in the succeeding one, such that all the subquotients occurring are<br />

isomorphic either to GL r (L) or to (L, +).<br />

5.11. Lemma (Theorem Hilbert 90). —– Let L/K be a finite Galois extension,<br />

G := Gal(L/K ) its Galois group, <strong>and</strong> n ∈Æ. Then, H 1 (G, GL n (L)) = 0.<br />

Proof. Let (a σ ) σ∈G be a cocycle with values in GL n (L). We define a G-operation<br />

from the left on L n as follows.<br />

We let σ ∈ G act as<br />

a σ ◦ σ : L n<br />

σ<br />

−→ L n<br />

a σ<br />

−→ L n .<br />

Then, it is clear that a σ ◦ σ is a σ-linear map. Galois descent yields a K-vector<br />

space V such that there is an isomorphism<br />

making the diagrams<br />

V ⊗ K L ∼ =<br />

−→<br />

b<br />

L n<br />

V ⊗ K L<br />

σ<br />

V ⊗ K L<br />

b<br />

b<br />

L n<br />

L n a σ ◦σ<br />

commute. In particular, one has that dim K V = n. The choice of an isomorphism<br />

V ∼ =<br />

−→ K n<br />

provides us with an element b ∈ GL n (L) such that b ◦ σ = a σ ◦ σ ◦ b. I.e.,<br />

a σ = b ◦ σ ◦ b −1 ◦ σ −1 = b ◦ σ (b −1 )<br />

for all σ ∈ G. (a σ ) σ∈G is cohomologous to the trivial cocycle.<br />

□<br />

5.12. Definition. –––– Let K be a field, r ∈Æ, <strong>and</strong> X be a Brauer-Severi variety<br />

of dimension r. Then, a linear subspace of X is a closed subvariety Y ⊂ X such<br />

that<br />

Y × SpecK SpecK sep ⊂ X × SpecK SpecK sep ∼ = P<br />

r<br />

K sep<br />

is a linear subspace of the projective space.<br />

This property is independent of the isomorphism chosen.<br />

5.13. Remark. –––– A K-valued point would be a zero-dimensional linear subspace.<br />

But, except for the trivial case X ∼ = P r K<br />

, there are no K-valued <strong>points</strong>.


34 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

Nevertheless, it may happen that there exist linear subspaces Y X of higher dimension.<br />

They can be investigated by cohomological methods generalizing the<br />

argument given above.<br />

5.14. Proposition (F. Châtelet <strong>and</strong> M. Artin). —– Let K be a field, r, d ∈Æ,<br />

X a Brauer-Severi variety of dimension r, <strong>and</strong> Y a linear subspace of dimension d.<br />

Then, the natural boundary maps send the cohomology classes<br />

α K r (X ) ∈ H 1( Gal(K sep /K ), PGL r+1 (K sep ) )<br />

<strong>and</strong><br />

α K d (Y ) ∈ H 1( Gal(K sep /K ), PGL d+1 (K sep ) )<br />

to the same class in the cohomological Brauer group H 2( Gal(K sep /K ), (K sep ) ∗) .<br />

Proof. Let L be a common splitting field for X <strong>and</strong> Y. We may assume that<br />

L/K is a finite Galois extension. Denote its Galois group by G.<br />

We choose an isomorphism X × SpecK SpecL ∼ =<br />

−→ P r L . We may assume without<br />

restriction that the linear subspace<br />

Y × SpecK SpecL ⊂ X × SpecK SpecL ∼ = P r L<br />

is given by the homogeneous equations X d+1 = · · · = X r = 0. Then, the<br />

cohomology class α L/K<br />

r (X ) ∈ H 1( G, PGL r+1 (L) ) is given by a cocycle (α σ ) σ∈G<br />

such that “X d+1 = · · · = X r = 0” is an invariant subspace for every α σ .<br />

Consequently, α L/K<br />

r (X ) is in the image of H 1( G, F/L ∗) under the natural<br />

homomorphism for<br />

{( )<br />

E1 H<br />

F :=<br />

∈ GL r+1 (L)<br />

0 E 2<br />

∣ E }<br />

1 ∈ GL d+1 (L), E 2 ∈ GL r−d (L),<br />

.<br />

H ∈ M (d+1)×(r−d) (L)<br />

F comes equipped with the homomorphism of G-<strong>groups</strong><br />

p : F −→ GL d+1 (L),<br />

( )<br />

E1 H<br />

↦→ E 1 .<br />

0 E 2<br />

Thus, we obtain a commutative diagram that connects the three central inclusions,<br />

GL d+1 (L)<br />

L ∗<br />

p<br />

L ∗<br />

F<br />

L ∗ GL r+1 (L) .<br />

i


Sec. 6] CENTRAL SIMPLE ALGEBRAS AND BRAUER-SEVERI VARIETIES 35<br />

Therefore, there is the following commutative diagram uniting the boundary<br />

maps,<br />

H 1 (G, PGL d+1 (L)) H 2 (G, L ∗ )<br />

p ∗<br />

<br />

H 1 (G, F/L ∗ )<br />

H 2 (G, L ∗ )<br />

i ∗<br />

<br />

H 1 (G, PGL r+1 (L)) H 2 (G, L ∗ ).<br />

Finally, recall that we have a cohomology class α ∈ H 1 (G, F/L ∗ ) such that<br />

i ∗ (α) = αr<br />

L/K (X ) <strong>and</strong> p ∗ (α) = αd L/K (Y ). □<br />

6. Central simple algebras <strong>and</strong> Brauer-Severi varieties<br />

6.1. Theorem. –––– Let n ∈Æ, K a field <strong>and</strong> A a central simple algebra over K<br />

of dimension n 2 .<br />

a) Then,thereexistsaBrauer-SeverivarietyX A ofdimension (n−1)overK satisfying<br />

condition (∗). (∗) determines X A uniquely up to isomorphism of K-schemes.<br />

(∗)If L is a splitting field for A which is finite <strong>and</strong> Galois over K then L is a<br />

splitting field for X A , too. There is one <strong>and</strong> the same cohomology class<br />

associated with A <strong>and</strong> X A .<br />

a A = α XA ∈ H 1( Gal(L/K ), PGL n (L) )<br />

b) The assignment X : A ↦→ X A admits the following properties.<br />

i) X is compatible with extensions K ′ /K of the base field. I.e.,<br />

X A⊗K K ′ ∼ = XA × SpecK SpecK ′ .<br />

ii) L is a splitting field for A if <strong>and</strong> only if L is a splitting field for X A .<br />

Proof. a) Uniqueness is clear from the results of the preceding sections.<br />

Existence. Choose a finite Galois extension L of K which is a splitting field<br />

for A. Take condition (∗) as a definition for X A . This is independent of the<br />

choice of L by Corollaries 4.7.a) <strong>and</strong> 5.7.a).


36 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

b.i) Let K ′ /K be an arbitrary field extension. We have the obvious diagram of<br />

field extensions<br />

LK ′ <br />

<br />

K ′ <br />

L<br />

K<br />

<br />

<strong>and</strong> the canonical inclusion Gal(LK ′ /K ′ ) ⊆ Gal(L/K ). Under the constructions<br />

given above, one assigns to the central simple algebra A ⊗ K K ′ <strong>and</strong> the<br />

Brauer-Severivariety X × SpecK SpecK ′ the restrictions of the cohomology classes<br />

assigned to A, respectively X. I.e.,<br />

a (A⊗K K ′ ) = res Gal(LK ′ /K ′ )<br />

Gal(L/K )<br />

(a A ),<br />

α (X ×SpecK SpecK ′) = resGal(LK ′ /K ′ )<br />

Gal(L/K )<br />

(α X ).<br />

ii) The two statements are equivalent to<br />

<strong>and</strong><br />

res Gal(LK ′ /K ′ )<br />

Gal(L/K )<br />

(a A ) ∈ H 1( Gal(LK ′ /K ′ ), PGL n (LK ′ ) ) = 0<br />

res Gal(LK ′ /K ′ )<br />

Gal(L/K )<br />

(α XA ) ∈ H 1( Gal(LK ′ /K ′ ), PGL n (LK ′ ) ) = 0,<br />

respectively. As we have a A = α XA , the assertion follows.<br />

□<br />

6.2. Corollary. –––– Let K be a field <strong>and</strong> A be a central simple algebra over K.<br />

Then, K ′ is a splitting field for A if <strong>and</strong> only if X A (K ′ ) ≠ ∅.<br />

Proof. “=⇒” is trivial <strong>and</strong> “⇐=” is an easy consequence of Proposition 5.9. □<br />

6.3. Corollary. –––– i) Let K be a field <strong>and</strong> n ∈Æ. Then, X induces a bijection<br />

X K n : AzK n → BS K n−1 .<br />

ii) Let L/K be a field extension. Then, X induces a bijection<br />

X L/K<br />

n<br />

: Az L/K<br />

n<br />

→ BS L/K<br />

n−1 .<br />

iii) These mappings are compatible with extensions of the base field. I.e., the diagram<br />

Az K n<br />

X K n<br />

BS K n−1<br />

⊗ K K ′ <br />

Az K ′<br />

n<br />

X K ′<br />

n<br />

BS K ′<br />

n−1<br />

× SpecK SpecK ′


Sec. 6] CENTRAL SIMPLE ALGEBRAS AND BRAUER-SEVERI VARIETIES 37<br />

commutes for every field extension K ′ /K.<br />

Proof. i) follows immediately from Theorem 6.1.a). ii) is a consequence of<br />

Theorem 6.1.a) together with Theorem 6.1.b.ii). iii) is simply a reformulation<br />

of Theorem 6.1.b.i).<br />

□<br />

6.4. Remark. –––– It may happen that two Brauer-Severi varieties X 1 , X 2 are<br />

bi<strong>rational</strong>ly equivalent but not isomorphic. S. A. Amitsur [Am55] proved in<br />

this case that the corresponding central simple algebras A 1 <strong>and</strong> A 2 generate the<br />

same cyclic subgroup of Br(K ). It is an open question whether the converse<br />

is true. Interesting partial results have been obtained by P. Roquette [Roq64]<br />

<strong>and</strong> S. L. Tregub [Tr].<br />

6.5. Proposition. –––– Let K be a field, n ∈Æ, <strong>and</strong> A a central simple algebra of<br />

dimension n 2 over K. Then, there is an isomorphism<br />

x A : Aut K (A)<br />

∼=<br />

−→ Aut K-schemes (X A ).<br />

Proof. Choose a splitting field L for A which is finite <strong>and</strong> Galois over K.<br />

Put G := Gal(L/K ). Choose an isomorphism A ⊗ K L −→ f<br />

M n (L). Then, for<br />

each σ ∈ G, there is a commutative diagram<br />

A ⊗ K L<br />

f<br />

M n (L)<br />

σ<br />

A ⊗ K L<br />

f<br />

a σ ◦σ<br />

M n (L) .<br />

The L-linear maps a σ : M n (L) → M n (L) form a cocycle for the cohomology<br />

class a A ∈ H 1 (G, PGL n (L)).<br />

Further, by Galois descent, to give an element of Aut K (A) is equivalent to giving<br />

an element of PGL n (L) which is invariant under a σ ◦ σ for every σ ∈ G.<br />

As α XA = a A , there exists an isomorphism X A × SpecK SpecL −→ f ′<br />

P n−1<br />

L such<br />

that the diagrams<br />

X A × SpecK SpecL<br />

f ′<br />

P n−1<br />

L<br />

σ<br />

X A × SpecK SpecL<br />

f ′<br />

P n−1<br />

L<br />

commute. Therefore, by Galois descent for morphisms of schemes, it is equivalent<br />

to give an element of Aut K-schemes (X A ) or to give an element of PGL n (L)<br />

which is invariant under a σ ◦ σ for every σ ∈ G.<br />

□<br />

a σ ◦σ


38 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

6.6. Proposition (F. Châtelet <strong>and</strong> M. Artin). —– Let K be a field, n, d ∈Æ,<br />

<strong>and</strong> A be a central simple algebra of dimension n 2 over K.<br />

Then, the Brauer-Severi variety X A associated with A admits a linear subspace of<br />

dimension d if <strong>and</strong> only if d ≤ n − 1 <strong>and</strong><br />

d ≡ −1<br />

(mod ind(A)).<br />

Proof. We write A = M m (D) with a skew field D <strong>and</strong> put e := ind(A).<br />

Clearly, n = me.<br />

“=⇒” Let H ⊂ X A be a linear subspace of dimension d. By Proposition 5.14,<br />

we know H ∼ = X A ′ for some central simple algebra A ′ which is similar to A.<br />

I.e.,A ′ ∼ = M k (D)for acertaink ∈Æ. Itfollowsthatdim A ′ = k 2·dim D = k 2·e 2<br />

<strong>and</strong> dim H = dim X A ′ = k · e − 1. This implies the congruence desired.<br />

Further, we have dim X A = n − 1. Consequently,<br />

d = dim H ≤ dim X A = n − 1.<br />

“⇐=”LetL beasplittingfieldforA whichisfinite<strong>and</strong> GaloisoverK. Denoteits<br />

Galois group by G. Further, let k be the natural number such that d = k ·e −1.<br />

By assumption, k · e − 1 = d ≤ n − 1 = m · e − 1, hence k ≤ m.<br />

We consider the cohomology class a D ∈ H 1 (G, PGL e (L)), associated with D.<br />

By Proposition 4.9, one has a A = (i e me ) ∗(a D ) where i e me : PGL e(L) → PGL me (L)<br />

is the block-diagonal embedding. Let (a σ ) σ∈G be a cocycle representing the<br />

cohomology class a D . Then, (i e me (a σ)) σ∈G is a cocycle that represents a A .<br />

We define a G-operation on P me−1<br />

L<br />

by letting σ ∈ G act as<br />

i e me (a σ) ◦ σ : P me−1<br />

L<br />

σ<br />

P me−1<br />

L<br />

i e me(a σ )<br />

This is a group operation since (i e me (a σ)) σ∈G is a cocycle.<br />

P me−1<br />

L .<br />

The geometric version of Galois descent yields the Brauer-Severi variety X A .<br />

Further, the G-operation keeps invariant the linear subspace defined by the<br />

homogeneous equations X ke = X ke+1 = · · · = X me−1 = 0. Thus, Galois<br />

descent may be applied to this subspace, too. It yields a variety Y of dimension<br />

ke − 1 = d.<br />

Galois descent for morphisms of schemes, applied to the canonical embedding,<br />

yields a morphism Y −→ X A . This is a closed immersion as that property<br />

descends under faithful flat base change. Consequently, Y is a linear subvariety<br />

of X A of the dimension desired.<br />


Sec. 7] FUNCTORIALITY 39<br />

7. Functoriality<br />

7.1. Remark. –––– The preceding results suggest that X : A ↦→ X A should<br />

somehow be a functor. This is indeed the case. But for that, the construction<br />

of the Brauer-Severi variety associated with a central simple algebra, which was<br />

given above, is not sufficient. The problem is that X A is determined by its<br />

cohomology class only up to isomorphism, not up to unique isomorphism.<br />

Thus, in order to make X a functor, it would still be necessary to make choices.<br />

For that reason, we aim at a more natural description of X A .<br />

7.2. Lemma (A. Grothendieck). —– Let n ∈Æ<strong>and</strong> R a commutative ring<br />

with unit. Then, there is a bijection<br />

{ }<br />

submodules M in R<br />

κ R : P n−1<br />

R (R) −→ G(R) :=<br />

n such that R n /M<br />

is a locally free R-module of rank (n − 1)<br />

subject to the conditions below.<br />

i) κ R is natural in R. I.e., for every ring homomorphism i : R → R ′ , the diagram<br />

commutes.<br />

P n−1<br />

R (R)<br />

P n−1 (i)<br />

<br />

P n−1<br />

R ′ (R ′ )<br />

κ R<br />

κ R ′<br />

G(R)<br />

G(R ′ )<br />

G(i): M ↦→M ⊗ R R ′<br />

ii) For every a ∈ PGL n (R), the canonical operations of a on P n−1<br />

R (R) <strong>and</strong> G(R)<br />

are compatible with κ R . That means that the diagrams<br />

commute.<br />

P n−1<br />

R (R)<br />

a<br />

P n−1<br />

R<br />

(R)<br />

κ R<br />

κ R<br />

G(R)<br />

a<br />

G(R)<br />

Here, a ∈ PGL n (R) acts on a submodule M ⊂ R n by matrix multiplication from<br />

the left. I.e., by M ↦→ a · M for a ∈ GL n (R) a representative of a.<br />

Proof. A submodule M ⊆ R n such that the quotient R n /M is locally free of<br />

rank (n − 1) defines an R-valued point in P n−1<br />

R .<br />

To see this, let first m ⊆ R be an arbitrary maximal ideal. Then, Rm/M n m is a<br />

free R m -module of rank (n − 1). In particular, it is projective. Hence, M m is<br />

a direct summ<strong>and</strong> of Rm n <strong>and</strong>, therefore, projective <strong>and</strong> of finite presentation.<br />

Consequently, by [Mat, Theorem 7.12], M m is a free R m -module of rank one.<br />

As R n /M is R-flat, there is an exact sequence<br />

0 −→ M ⊗ R (R/m) −→ R n /mR n −→ R n /(M + mR n ) −→ 0.


40 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

Thus, the canonical map M/mM → R n /mR n is injective. I.e., mM = mR n ∩M.<br />

In particular, M can not be contained in mR n . Hence, the R m -module M m<br />

which is free of rank one is not contained in mRm.<br />

n Consequently, if<br />

M m = 〈(r 0 , . . . , r n−1 )〉 then there is some α ∈ {0, . . . , n − 1} such that<br />

r α ∈ R m is a unit.<br />

This is equivalent to<br />

R n m/(M m + e 0 · R m + · · · + e α−1 · R m + e α+1 · R m + · · · + e n−1 · R m ) = 0<br />

where e 0 , . . . , e n−1 ∈ R n m denote the st<strong>and</strong>ard elements with an index shift<br />

by (−1). The latter is an open condition on m by [Mat, Theorem 4.10].<br />

Therefore, there exists some f ∈ R\m such that<br />

R n f = M f + e 1 · R f + · · · + e α−1 · R f + e α+1 · R f + · · · + e n · R f .<br />

I.e., such that the (α + 1)-th projection M f → R f is surjective.<br />

As, on both sides, there are locally free modules of rank one, this must<br />

be an isomorphism. Consequently, M f is free <strong>and</strong> there is a generator of<br />

type (r 0 , . . . , r α−1 , 1, r α+1 , . . . , r n−1 ). Taking the entries as homogeneous<br />

coordinates, we get a morphism SpecR f → P n−1<br />

R .<br />

It is easy to see that all these can be glued together to give a morphism<br />

SpecR → P n−1<br />

R of R-schemes.<br />

Conversely, an R-valued point SpecR → P n−1<br />

R defines a quotient module of R n<br />

of rank (n − 1) as follows. Cover SpecR by affine, open subsets<br />

SpecR =<br />

n−1<br />

[<br />

α=0<br />

SpecR α<br />

such that for each α the image i(SpecR α ) is contained in the st<strong>and</strong>ard affine<br />

set “X α ≠ 0”. Then, i| SpecRα can be given in the form<br />

(r 0 : r 1 : . . . : r α−1 : 1 : r α+1 : . . . : r n−1 ) .<br />

But R n α/(r 0 , r 1 , . . . , r α−1 , 1, r α+1 , . . . , r n−1 ) is a free R α -module of rank (n−1)<br />

for trivial reasons.<br />

The compatibility stated as i) follows directly from the construction given.<br />

ii) is clear.<br />

7.3. Definition. –––– Let R be a commutative ring with unit.<br />

i) Let X be an R-scheme. Then, by<br />

P X : {R-schemes} op → {sets}<br />


Sec. 7] FUNCTORIALITY 41<br />

we will denote the functor defined by<br />

T ↦→ Mor R-schemes (T , X )<br />

on objects <strong>and</strong> by composition on morphisms.<br />

ii) Let F : {R-schemes} op → {sets} be any functor. If, for some R-scheme X,<br />

∼=<br />

there is an isomorphism ι: F −→ P X then we will say F is represented by X.<br />

Having fixed the isomorphism ι then, by Yoneda’s Lemma, the K-scheme representing<br />

the functor is determined up to unique isomorphism.<br />

7.4. Remarks. –––– a) The functors P X depend on X in a natural manner.<br />

I.e., if p : X → X ′ is a morphism of R-schemes then there is a morphism of<br />

functors p ∗ : P X → P X ′ given by composition with p. Therefore, there is a<br />

covariant functor<br />

P : {R-schemes} −→ Fun ({R-schemes} op , {sets})<br />

described by X ↦→ P X on objects.<br />

b) The functor P X induced by an R-scheme X is exactly what in category theory<br />

is usually called the Hom-functor. Nevertheless, we will follow the st<strong>and</strong>ards in<br />

Algebraic Geometry <strong>and</strong> refer to it as the functor of <strong>points</strong>. Note that P X (T ) is<br />

the set of all T -valued <strong>points</strong> on X.<br />

7.5. Corollary (A. Grothendieck). —– Let n ∈Æ<strong>and</strong> R be a commutative ring<br />

with unit. Then, the functor<br />

P n−1<br />

R : {R-schemes} −→ {sets}<br />

T ↦→ P n−1<br />

R (T ) :=<br />

is represented by the R-scheme P n−1<br />

R .<br />

⎧<br />

⎫<br />

⎨ subsheaves M in OT n such that ⎬<br />

OT n /M is a locally free<br />

⎩<br />

⎭<br />

O T -module of rank (n − 1)<br />

Proof. It is clear that both functors satisfy the sheaf axiom for Zariski coverings.<br />

Therefore, it is sufficient to construct the isomorphism on the full subcategory<br />

of affine R-schemes. This is exactly what is done in Lemma 7.2. □<br />

7.6. Corollary-Definition. –––– Let n ∈Æ, L be a field, <strong>and</strong> F be an L-vector<br />

space of dimension n. Put F := ˜F to be the coherent sheaf associated to F<br />

on SpecL.


42 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

Then, the functor<br />

P(F ): {L-schemes} −→ {sets}<br />

⎧<br />

⎫<br />

⎨ subsheaves M ⊂ π ∗ F such that ⎬<br />

(π : T → SpecL) ↦→ P(F )(π) := π ∗ F/M is a locally free<br />

⎩<br />

⎭<br />

O T -module of rank (n − 1)<br />

is representable by an L-scheme which is isomorphic to P n−1<br />

L . We will denote<br />

that scheme by P(F ) <strong>and</strong> call it the projective space of lines in F.<br />

7.7. Remark. –––– If L ′ is a field containing L then P(F ⊗ L L ′ ) is naturally<br />

isomorphic to the functor P(F )| L ′ -schemes. Therefore, there is a canonical isomorphism<br />

P(F ⊗ L L ′ ∼=<br />

) −→ P(F ) × SpecL SpecL ′<br />

between the representing objects. For each L ′ -scheme U, the diagram of natural<br />

mappings<br />

Mor L ′ -schemes(U, P(F )) {M ⊂ πU ∗ F }<br />

∼=<br />

<br />

Mor L ′ -schemes(U, P(F ⊗ L L ′ )) {M ⊂ πU ∗ F }<br />

commutes.<br />

Consequently, for each L-scheme T , there is a commutative diagram<br />

Mor L-schemes (T , P(F ))<br />

{M ′ ⊂ π ∗ T F }<br />

comp. with projection<br />

<br />

Mor L ′ -schemes(T × SpecL SpecL ′ , P(F ))<br />

pull-back<br />

∼=<br />

<br />

Mor L ′ -schemes(T × SpecL SpecL ′ , P(F ⊗ L L ′ )) {M ⊂ π ∗ T × SpecL Spec L ′F } .<br />

Note that the column on the left is, up to the canonical isomorphism above,<br />

exactly the base extension from SpecL to SpecL ′ .<br />

7.8. Definition. –––– Let L/K be a finite Galois extension of fields <strong>and</strong><br />

G := Gal(L/K ) be its Galois group.<br />

i) By L-Vect K , we will denote the category of all finite dimensional<br />

L-vector spaces. As morphisms, we allow all injections which are σ-linear<br />

for a certain σ ∈ G.<br />

ii) L-Vect K 0 is a subcategory ofL-Vect K which consists of the same class of objects.<br />

In L-Vect K 0 , only the bijections are taken as morphisms.


Sec. 7] FUNCTORIALITY 43<br />

7.9. Definition. –––– Let H be a group. Then, by H we will denote the category<br />

consisting of exactly one object ∗ such that Mor(∗, ∗) = H .<br />

7.10. Definition. –––– Let L/K be a finite Galois extension of fields <strong>and</strong> G be<br />

its Galois group. By Sch L/K , we will denote the category of all L-schemes with<br />

morphisms twisted by any element of G.<br />

We note that Sch L/K has a canonical structure of a fibered category over G.<br />

The pull-back under σ ∈ G is given by X ↦→ X × SpecL SpecL σ−1 .<br />

7.11. Lemma. –––– Let L/K be a finite Galois extension of fields <strong>and</strong> G be its<br />

Galois group.<br />

i) Then, P is a covariant functor from L-Vect K to Sch L/K . Here, a σ-linear<br />

monomorphism i : F → F ′ for σ ∈ G induces a morphism<br />

which is twisted by σ.<br />

i ∗ = P(i): P(F ) −→ P(F ′ )<br />

ii) LetF beanL-vectorspace<strong>and</strong> i : F → F bethemultiplicationwithx ∈ L, x ≠ 0.<br />

Then, i ∗ : P(F ) → P(F ) is equal to the identity morphism.<br />

Proof. i) Let i : F → F ′ be a σ-linear monomorphism. We have to consider<br />

the diagram<br />

P(F ) P(F ′ )<br />

SpecL<br />

S(σ)<br />

SpecL .<br />

By Yoneda’s Lemma, there has to be constructed a natural transformation<br />

i + : P(F ) → P(F ′ )(S(σ) ◦ · ) .<br />

I.e., for each L-scheme π : T → SpecL there is a mapping<br />

i + (π): P(F )(π) → P(F ′ )(S(σ) ◦ π)<br />

to be given such that for each morphism p : π 1 → π 2 of L-schemes the diagram<br />

(†)<br />

P(F )(π 2 )<br />

i + (π 2 )<br />

P(F ′ )(S(σ) ◦ π 2 )<br />

commutes.<br />

P(F )(p)<br />

P(F )(π 1 )<br />

i + (π 1 )<br />

P(F ′ )(S(σ)◦p)<br />

P(F ′ )(S(σ) ◦ π 1 )


44 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

We construct i + (π) as follows. There is the description<br />

P(F )(π) = M ⊂ π ∗ F | π ∗ F/M is locally free of rank (dim F − 1) }<br />

{ }<br />

M ⊂ (S(σ) ◦ π)<br />

=<br />

∗ S(σ −1 ) ∗ F |<br />

(S(σ) ◦ π) ∗ S(σ −1 ) ∗ F/M is locally free of rank (dim F − 1) .<br />

Here, S(σ −1 ) ∗ F = ˜F σ for F σ equal to F as an abelian group <strong>and</strong> equipped<br />

with the scalar multiplication given by l · f := σ(l) f . Hence, i gives rise<br />

to an L-linear monomorphism i : F σ → F ′ <strong>and</strong>, therefore, to a morphism<br />

ĩ : S(σ −1 ) ∗ F → F ′ of sheaves on SpecL.<br />

If M is a subsheaf of (S(σ) ◦ π) ∗ S(σ −1 ) ∗ F such that (S(σ) ◦ π) ∗ S(σ −1 ) ∗ F/M<br />

is locally free of rank (dim F − 1) then ( (S(σ) ◦ π) ∗ (ĩ) ) (M ) is a subsheaf of<br />

(S(σ)◦π) ∗ F ′ such that the quotient is locally free of rank (dim F ′ −1). This gives<br />

rise to a morphism of functors i + : P(F ) −→ P(F ′ )(S(σ) ◦ · ) as desired.<br />

By consequence, we have a morphism i ∗ : P(F ) −→ P(F ′ ) of schemes making<br />

the diagram (†) commute.<br />

ii) is clear from the construction.<br />

7.12. Corollary. –––– Let L/K be a finite Galois extension of fields <strong>and</strong><br />

G := Gal(L/K ) its Galois group.<br />

i) Then, P can be made into a contravariant functor from L-Vect K 0 to Sch L/K .<br />

Here, a σ-linear isomorphism i : F → F ′ for σ ∈ G induces a morphism<br />

which is twisted by σ −1 .<br />

i ∗ : P(F ′ ) → P(F )<br />

ii) LetF beanL-vectorspace<strong>and</strong> i : F → F bethemultiplicationwithx ∈ L, x ≠ 0.<br />

Then, i ∗ : P(F ) → P(F ) is equal the identity morphism.<br />

Proof. For an isomorphism i : F → F ′ , put i ∗ := i −1<br />

∗ . □<br />

7.13. Remark. –––– The morphisms i ∗ can also be constructed directly. Indeed,<br />

the task is to give a natural transformation i + : P(F ′ ) → P(F )(S(σ −1 ) ◦ · ).<br />

There are the descriptions<br />

<strong>and</strong><br />

P(F ′ )(π) = { M ⊂ π ∗ F ′ | π ∗ F ′ /M is locally free of rank (dim F ′ − 1) }<br />

P(F )(S(σ −1 ) ◦ π) =<br />

{ }<br />

M ⊂ π ∗ S(σ −1 ) ∗ F |<br />

π ∗ S(σ −1 ) ∗ .<br />

F/M is locally free of rank (dim F − 1)<br />

If M is a subsheaf of π ∗ F ′ such that the quotient π ∗ F ′ /M is locally free of<br />

rank (dim F ′ − 1) = (dim F − 1) then π ∗ (ĩ) −1 (M ) is a subsheaf of π ∗ S(σ −1 ) ∗ F<br />


Sec. 7] FUNCTORIALITY 45<br />

such that the quotient is locally free of rank (dim F − 1). This gives rise to a<br />

morphism of functors i + : P(F ′ ) −→ P(F )(S(σ −1 ) ◦ · ) as desired.<br />

7.14. Remark. –––– If L is a field <strong>and</strong> A is a central simple algebra of dimension<br />

n 2 over L then all the non-zero, simple, left A-modules are isomorphic<br />

to each other. Further, if l is a non-zero, simple, left A-module then each<br />

automorphism of l is given by multiplication with an element of the center<br />

of A.<br />

Hence, every automorphism of l induces the identity map on P(l). In particular,<br />

two arbitrary isomorphisms i 1 , i 2 : l → l ′ between non-zero, simple,<br />

left A-modules induce one <strong>and</strong> the same isomorphism i ∗ 1 = i ∗ 2 : P(l′ ) → P(l).<br />

Thismeans, A determinesP(l)notonlyuptoisomorphism, butuptouniqueisomorphism.<br />

7.15. Definition. –––– Let r ∈Æ<strong>and</strong> L/K be a finite Galois extension.<br />

Denote the Galois group Gal(L/K ) by G.<br />

i) By Mat L/K<br />

r , we will denote the category of all split central simple algebras of<br />

dimension r 2 over L. I.e., of all algebras isomorphic to M r (L). As morphisms,<br />

wetakeallhomomorphismsofK-algebraswhichareσ-linear for acertain σ ∈ G<br />

<strong>and</strong> preserve the unit element.<br />

ii) P L/K<br />

r will denote the subcategory of Sch L/K consisting of all L-schemes isomorphic<br />

to the projective space P r L. As morphisms, P L/K<br />

r allows only the isomorphisms.<br />

7.16. Remark. –––– In Mat K r<br />

two objects are isomorphic.<br />

, every morphism is an isomorphism <strong>and</strong> every<br />

7.17. Proposition. –––– Let n ∈Æ<strong>and</strong> L/K be a finite Galois extension.<br />

Denote the Galois group Gal(L/K ) by G.<br />

: Mat L/K<br />

n → P L/K<br />

n−1 .<br />

is given by A ↦→ P(l) where l ⊂ A is a non-zero, simple,<br />

i) There is an equivalence of categories Ξ L/K<br />

n<br />

On objects, Ξ L/K<br />

n<br />

left A-module. If i : A → A ′ is a morphism in Mat L/K<br />

n<br />

morphism of schemes induced by the canonical homomorphism l → A ′ ⊗ A l.<br />

then Ξ L/K<br />

n<br />

(i) is the<br />

ii) If i : A → A ′ is σ-linear for σ ∈ G then Ξ L/K<br />

n (i) is a morphism twisted by σ.<br />

Proof. If l is a non-zero, simple, left A-module then A ′ ⊗ A l is a non-zero, simple,<br />

left A ′ -module. Both are n-dimensional L-vector spaces. Up to isomorphism,<br />

these modules are unique. Therefore, the morphism of schemes<br />

i l∗ := Ξ L/K<br />

n<br />

(i): Ξn<br />

L/K (A) = P(l) −→ P(A ′ ⊗ A l) = Ξn L/K (A ′ )


46 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

induced by the homomorphism<br />

i l : l → A ′ ⊗ A l,<br />

x ↦→ 1 ⊗ x<br />

is well-defined.<br />

If i : A → A ′ is a σ-linear homomorphism then l → A ′ ⊗ A l is a σ-linear<br />

homomorphism of vector spaces. By Lemma 7.11, the morphism Ξn<br />

L/K (i) is<br />

twisted by σ. This proves ii).<br />

As i is invertible, Ξ L/K<br />

n<br />

(i) is an isomorphism of schemes. Consequently, Ξn<br />

L/K is<br />

is an equivalence of<br />

a functor between the categories described. To prove Ξn<br />

L/K<br />

categories, we have to show it is fully faithful <strong>and</strong> essentially surjective. As in<br />

P L/K<br />

n−1<br />

every two objects are isomorphic, essential surjectivity is clear.<br />

For full faithfulness, it will suffice to prove that<br />

Ξ L/K<br />

n | Aut(A) : Aut(A) −→ Aut(X A )<br />

is an isomorphism of <strong>groups</strong> in the case A = M n (L). For that, note<br />

<strong>and</strong><br />

Thus, the functor Ξ L/K<br />

n<br />

Aut(A) = PGL n (L) ⋊ G<br />

Aut(Ξn<br />

L/K (A)) = Aut K-schemes (P n−1<br />

L ) = PGL n (L) ⋊ G .<br />

induces on Aut(A) a group homomorphism<br />

Ξ: PGL n (L) ⋊ G → PGL n (L) ⋊ G .<br />

By Lemma 7.2.ii), the restriction of Ξ to PGL n (L) is the identity. Statement<br />

ii) shows that the quotient map G → G is the identity, too. Consequently,<br />

Ξn L/K | Aut(A) is an isomorphism. □<br />

7.18. Corollary. –––– The categories Mat L/K<br />

n<br />

<strong>and</strong> P L/K<br />

n−1<br />

are anti-equivalent, too.<br />

There is an equivalence of categories ˇΞ n<br />

L/K : Mat L/K<br />

n → (P L/K<br />

n−1 )op given on objects by<br />

A ↦→ P(l) where l ⊂ A is a non-zero, simple, left A-module.<br />

If i : A → A ′ is σ-linear for σ ∈ G then ˇΞ L/K<br />

n (i) is twisted by σ −1 .<br />

Proof. Put simply ˇΞ n<br />

L/K (i) := (Ξn L/K (i)) −1 . □<br />

7.19. Proposition. –––– Let n ∈Æ, K a field <strong>and</strong> A be a central simple algebra of<br />

dimension n 2 over K. Then, the Brauer-Severi variety X A associated with A may<br />

be described as follows.<br />

Let L be a splitting field for A which is finite <strong>and</strong> Galois over K. By covariant<br />

functoriality, the canonical Gal(L/K )-operation on A ⊗ K L induces an operation


Sec. 7] FUNCTORIALITY 47<br />

on the projective space Ξn<br />

L/K (A ⊗ K L). Here, σ ∈ Gal(L/K ) acts by a morphism<br />

twisted by σ. The geometric version of Galois descent yields the K-scheme X A .<br />

Proof. Choose an isomorphism f : A ⊗ K L → M n (L). Then, there is a cocycle<br />

(a σ ) σ∈G from G to PGL n (L) such that for each σ ∈ G the diagram<br />

A ⊗ K L<br />

f<br />

M n (L)<br />

to the whole situation, we obtain com-<br />

commutes. Applying the functor Ξ L/K<br />

n<br />

mutative diagrams<br />

Ξ L/K<br />

n (A ⊗ K L)<br />

Ξ L/K<br />

n ( f )<br />

<br />

P n−1<br />

L<br />

σ<br />

a σ ◦σ<br />

A ⊗ K L<br />

f<br />

M n (L)<br />

σ <br />

Ξ L/K<br />

n (A ⊗ K L)<br />

a σ ◦σ<br />

P n−1<br />

L<br />

Ξ L/K<br />

n ( f )<br />

such that the vertical arrows are isomorphisms. Galois descent on the lower half<br />

of the diagram yields the description of X A given in Section 6. Galois descent<br />

on the upper half of the diagram yields the description claimed. □<br />

7.20. Corollary. –––– Let L/K be a field extension <strong>and</strong> A be a central simple<br />

algebra over K.<br />

i) Then, there is a canonical isomorphism of L-schemes<br />

ξ L/K<br />

A : X A⊗K L −→ X A × SpecK SpecL .<br />

are com-<br />

ii) For field extensions L ′ /L/K, the isomorphisms ξ L′ /K<br />

A , ξ L′ /L<br />

A , <strong>and</strong> ξ L/K<br />

A<br />

patible. I.e., the diagram<br />

X A⊗K L ′ ξ L′ /K<br />

A<br />

<br />

commutes.<br />

ξ L′ /L<br />

X<br />

A<br />

A × SpecK SpecL ′<br />

<br />

X A⊗K L × SpecL SpecL ′<br />

ξ L/K<br />

A × SpecLSpecL ′<br />

Proof. Let n 2 = dim A. We choose a splitting field L ′′ for A containing L ′ .<br />

Then, X A , X A⊗K L, <strong>and</strong> X A⊗K L ′ are constructed from<br />

Ξ L′′ /K<br />

n (A ⊗ K L ′′ ) = Ξ L′′ /L<br />

n (A ⊗ K L ′′ ) = Ξ L′′ /L ′<br />

n (A ⊗ K L ′′ )<br />

by Galois descent using compatible descent data.<br />


48 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

7.21. Definition. –––– Let r ∈Æ<strong>and</strong> K be a field.<br />

i) By Az K r , we will denote the category of all central simple algebras of dimension<br />

r 2 over K. As morphisms, we take the homomorphisms of K-algebras<br />

preserving the unit element. Note, in Az K r every morphism is an isomorphism.<br />

Az /K<br />

r will denote the category of all central simple algebras of dimension r 2<br />

over any field extension of K. As morphisms, we take the K-algebra homomorphisms<br />

which preserve the unit element.<br />

ii) By BS K r , we will denote the category of all Brauer-Severi varieties of dimension<br />

r over K. The isomorphisms of K-schemes are taken as morphisms.<br />

BS /K<br />

r<br />

will denote the category of all Brauer-Severi varieties of dimension r over<br />

any field extension of K. As morphisms, we take the compositions of an<br />

isomorphism of K-schemes with a morphism of type<br />

id × Speca : X × SpecL SpecL ′ −→ X.<br />

Here, a : L → L ′ is a homomorphism of fields.<br />

7.22. Theorem. –––– Let K be a field <strong>and</strong> n ∈Æ.<br />

i) There is an equivalence of categories<br />

X /K<br />

n<br />

satisfying the following condition.<br />

: Az/K n −→ (BS /K<br />

n−1 )op<br />

For each field extension L/K, on isomorphism classes, the functor X /K<br />

n<br />

bijection Xn L : Az L n → BS L n−1 from Corollary 6.3.i).<br />

induces an equiva-<br />

ii) In particular, for each field extension L/K, the functor X /K<br />

n<br />

lence of categories X L n : Az L n → (BS L n−1) op .<br />

Proof. First step. Construction of the functor.<br />

induces the<br />

For A ∈ Ob(Az /K<br />

n ), we put X /K<br />

n (A) := X A . We are going to make use of the<br />

intrinsic description of X A given in Proposition 7.19.<br />

If i : A → A ′ is a morphism in Az n<br />

/K then, by restriction to the centers, i induces<br />

a homomorphism i| Z(A) : Z(A) → Z(A ′ ) of fields extending K. Therefore, there<br />

is a unique factorization<br />

A cZ(A′ )<br />

A<br />

A ⊗ Z(A) Z(A ′ )<br />

j<br />

A ′<br />

of i via the canonical inclusion c Z(A′ )<br />

A .<br />

By Corollary 7.20, one has the canonical isomorphism<br />

ξ Z(A′ )/Z(A)<br />

A : X A⊗Z(A) Z(A ′ ) −→ X A × SpecZ(A) SpecZ(A ′ ).


Sec. 7] FUNCTORIALITY 49<br />

We put X /K<br />

n (c Z(A′ )<br />

A ) to be the morphism of schemes induced from ξ Z(A′ )/Z(A)<br />

A by<br />

projection to the first factor.<br />

In order to construct X /K<br />

n (i) as a functor on the category Az/K n , it remains<br />

to describe it on the full subcategories Az K 1<br />

n for all field extensions K 1 /K.<br />

That means, we are left with the case that Z(A) = Z(A ′ ) <strong>and</strong> i| Z(A) is the identity.<br />

In order to ensure functoriality, the construction has to be made compatibly<br />

with field extensions. I.e., such that the diagram<br />

X A ′ ⊗ K1 K 2<br />

X /K<br />

n (i⊗ K1 K 2 ) <br />

X /K<br />

n (c K 2<br />

A ′ ) <br />

X A ′<br />

X /K<br />

n (i)<br />

X A⊗K1 K 2<br />

X A<br />

X /K<br />

n (c K 2<br />

A )<br />

commutes for every field K 1 containing K, every morphism i : A → A ′ in Az K 1<br />

n ,<br />

<strong>and</strong> every extension K 2 /K 1 of fields containing K.<br />

For this, we choose a splitting field L for A. We may assumeL is finite <strong>and</strong> Galois<br />

over K 1 . As i is automatically an isomorphism, L is a splitting field for A ′ , too.<br />

Thus, weobtainaGaloisinvarianthomomorphismi⊗ K1 L : A⊗ K1 L → A ′ ⊗ K1 L.<br />

Applying the functor Ξ L/K 1<br />

n , one gets a Galois invariant morphism of schemes<br />

X A ′ × SpecK1 SpecL = Ξ L/K 1<br />

n (A ′ ⊗ K1 L)<br />

Ξ L/K 1<br />

Ξ L/K 1<br />

n (i⊗ L M )<br />

n (A ⊗ K1 L) = X A × SpecK1 SpecL .<br />

Galois descent for morphisms of schemes yields the morphism<br />

X /K<br />

n (i): X A ′ → X A<br />

desired. X /K<br />

n (i) is an isomorphism of schemes as i is. Consequently, X /K<br />

n is a<br />

functor between the categories stated.<br />

By construction, X /K<br />

n is essentially surjective.<br />

We also note that, for each field extension K ′ /K, X /K<br />

n induces a functor<br />

X K ′<br />

n : AzK ′<br />

n −→ (BS K ′<br />

n−1 )op .<br />

Second step. Full faithfulness on automorphisms.<br />

Let us deal with the following statement which is a special case of full faithfulness<br />

of X /K<br />

n .<br />

(‡) Let K ′ be a field containing K <strong>and</strong> A be a central simple algebra of dimension<br />

n 2 over K ′ . Then, the functor X K ′<br />

n induces an isomorphism of <strong>groups</strong><br />

x A := X K ′<br />

n | AutK ′ (A) : Aut K ′(A) −→ Aut K ′ -schemes(X A ).


50 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

In the case A = M n (K ′ ), this was proven in Proposition 7.17.<br />

Let now A be a general central simple algebra of dimension n 2 over K ′ . Choose a<br />

splitting field L for A which is finite <strong>and</strong> Galois over K ′ .<br />

Injectivity. Assume a, a ′ ∈ Aut K ′(A) induce one <strong>and</strong> the same morphism<br />

a ∗ = a ′∗ ∈ Aut K ′ -schemes(X A ) .<br />

There are unique homomorphisms a L , a ′ L : A⊗ K L → A⊗ K L of central simple<br />

algebras over L making the diagrams<br />

A ⊗ K L<br />

a L A ⊗K L A ⊗ K L<br />

a ′ L<br />

A ⊗ K L<br />

c L A<br />

A<br />

a<br />

A ,<br />

c L A<br />

commute. Applying the functor X K ′<br />

n , we obtain the commutative diagrams below,<br />

X A⊗K L<br />

(c L A )∗ <br />

a ∗ L<br />

a ∗ =a ′∗<br />

X A⊗K L<br />

(c L A )∗<br />

c L A<br />

A<br />

X A⊗K L<br />

(c L A )∗ <br />

a ′ L ∗<br />

a ′<br />

a ∗ =a ′∗<br />

A<br />

X A⊗K L<br />

X A X A , X A X A .<br />

(cA L)∗ is, up to isomorphism, the canonical morphism X A × SpecK SpecL → X A<br />

from the fiber product to its first factor. Therefore, the morphism<br />

a ∗ ◦ (cA L)∗<br />

= a ′∗ ◦ (cA L)∗ admits a unique factorization into a morphism of<br />

L-schemes <strong>and</strong> (cA) L ∗ . This implies a ∗ L = a L ′ ∗ .<br />

SinceA⊗ K L ∼ = M n (L), wemayconcludea L = a L ′ from this. Theequalitya = a′<br />

follows immediately.<br />

By consequence, the homomorphism x A is injective.<br />

Surjectivity. Let p : X A → X A be an automorphism of the K ′ -scheme X A .<br />

(cA L)∗ is, up to isomorphism, the canonical projection X A × SpecK SpecL → X A<br />

from the fiber product to its first factor. Therefore, the composition<br />

p ◦ (cA L)∗ factors uniquely via (cA L)∗ . I.e., there exists a unique morphism<br />

p L : X A⊗K L → X A⊗K L of L-schemes such that there is a commutative diagram<br />

c L A<br />

(c L A )∗<br />

X A⊗K L<br />

p L<br />

X A⊗K L<br />

(c L A )∗ <br />

p<br />

X A X A .<br />

As A⊗ K L ∼ = M n (L), Proposition 7.17 shows there exists some b ∈ Aut(A⊗ K L)<br />

such that p L = b ∗ .<br />

(c L A )∗


Sec. 7] FUNCTORIALITY 51<br />

For σ ∈ Gal(L/K ′ ), let id × σ : A ⊗ K L → A ⊗ K L be the corresponding<br />

automorphism of A ⊗ K L. Clearly, (id × σ) ◦ cA L = cA L . Therefore, the diagram<br />

X A⊗K L<br />

b ∗ =p L<br />

X A⊗K L<br />

(id×σ) ∗ <br />

X A⊗K L<br />

b ∗ =p L<br />

X A⊗K L<br />

(id×σ) ∗<br />

commutes, too. Hence,<br />

(c L A )∗ <br />

p<br />

X A X A .<br />

(c L A )∗<br />

b ∗ = (id × σ) ∗ ◦ b ∗ ◦ (id × σ −1 ) ∗ = ((id × σ −1 ) ◦ b ◦ (id × σ)) ∗ .<br />

The injectivity of (‡) shows b = (id × σ −1 ) ◦ b ◦ (id × σ). I.e., b is invariant<br />

with respect to the operation of Gal(L/K ′ ).<br />

By Galois descent for homomorphisms, there exists a homomorphism<br />

a : A → A of central simple algebras such that the diagram<br />

A ⊗ K L<br />

b<br />

A ⊗ K L<br />

commutes.<br />

c L A<br />

A<br />

a<br />

By consequence, there is a commutative diagram on the level of Brauer-Severi<br />

varieties as follows,<br />

X A⊗K L<br />

(c L A )∗ <br />

b ∗ =p L<br />

a ∗<br />

A<br />

c L A<br />

X A⊗K L<br />

X A X A .<br />

A direct comparison with the definition of p L shows p ◦ (c L A) ∗ = a ∗ ◦ (c L A) ∗ .<br />

Indeed, both compositions are equal to (c L A )∗ ◦ p L . Since (c L A )∗ is dominant, this<br />

implies p = a ∗ .<br />

The homomorphism x A is surjective, too.<br />

Third step. Isomorphisms.<br />

Let K ′ ⊇ K be a field <strong>and</strong> A <strong>and</strong> A ′ be central simple algebras of dimension n 2<br />

over K ′ such that A ∼ = A ′ . Then, X K ′<br />

n induces a bijection<br />

(c L A )∗<br />

Iso K ′ (A, A ′ ) −→ Iso K ′ -schemes(X A ′, X A ) .<br />

This is an immediate consequence of the results from the previous step.


52 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

Fourth step. Faithfulness.<br />

Assume the homomorphisms i 1 , i 2 : A → A ′ induce the same morphism of<br />

schemes i ∗ 1 = i ∗ 2 : X A ′ → X A.<br />

We first observe that i 1 <strong>and</strong> i 2 give rise to the same homomorphism<br />

i 1 | Z(A) = i 2 | Z(A) : Z(A) → Z(A ′ )<br />

between centers. Indeed, X A <strong>and</strong> X A ′ are Brauer-Severi varieties over Z(A)<br />

<strong>and</strong> Z(A ′ ), respectively. Further, by Proposition 5.2.iv), Γ(X A , O XA ) = Z(A)<br />

<strong>and</strong> Γ(X A ′, O XA ′) = Z(A ′ ). I.e., one can recover the centers of A <strong>and</strong> A ′ from<br />

the Brauer-Severi varieties associated with them. By the construction of X /K<br />

n<br />

on morphisms, the pull-back on global sections<br />

(i ∗ 1 )♯ = (i ∗ 2 )♯ : Z(A) = Γ(X A , O XA ) −→ Γ(X A ′, O XA ′<br />

) = Z(A ′ )<br />

is equal to the homomorphism Z(A) → Z(A ′ ) given by i 1 , respectively i 2 , when<br />

restricted to the centers.<br />

Consequently, both i 1 <strong>and</strong> i 2 can be factorized via the canonical homomorphism<br />

c Z(A′ )<br />

A : A → A ⊗ Z(A) Z(A ′ ) .<br />

Let j 1 , j 2 : A⊗ Z(A) Z(A ′ ) → A ′ bethehomomorphismsofcentralsimplealgebras<br />

over Z(A ′ ) such that j 1 ◦ c Z(A′ )<br />

A = i 1 <strong>and</strong> j 2 ◦ c Z(A′ )<br />

A = i 2 . Both j 1 <strong>and</strong> j 2 are<br />

isomorphisms as they are homomorphisms of central simple algebras over the<br />

same base field.<br />

On the other h<strong>and</strong>, the morphisms<br />

j ∗ 1 , j ∗ 2 : X A ′ −→ X A⊗Z(A) Z(A ′ ) = X A × SpecZ(A) SpecZ(A ′ )<br />

coincide. Indeed, their projections to SpecZ(A ′ ) do, both being the structural<br />

morphism. Further, the projections to X A are equal to i ∗ 1 = i ∗ 2 .<br />

This implies j 1 = j 2 .<br />

Fifth step. Fullness.<br />

LetA, A ′ ∈ Ob (Az /K<br />

n )<strong>and</strong> f : X A ′ → X A beamorphism inthecategoryBS /K<br />

n−1 .<br />

The corresponding map of global sections is a homomorphism Z(A) → Z(A ′ )<br />

of fields. From the definition of the category BS /K<br />

n−1<br />

, we see that f gives rise to<br />

an isomorphism of Z(A ′ )-schemes<br />

such that f = (c Z(A′ )<br />

A ) ∗ ◦ f .<br />

f : X A ′ −→ X A × SpecZ(A) SpecZ(A ′ ) = X A⊗Z(A) Z(A ′ )


Sec. 8] THE FUNCTOR OF POINTS 53<br />

By virtue of Corollary 6.3, the central simple algebras A ′ <strong>and</strong> A ⊗ Z(A) Z(A ′ )<br />

over Z(A ′ ) are isomorphic. In the third step, we showed there is an isomorphism<br />

a : A ⊗ Z(A) Z(A ′ ) −→ A ′<br />

such that f = a ∗ . Consequently, f = (c Z(A′ )<br />

A ) ∗ ◦ a ∗ = (a ◦ c Z(A′ )<br />

A ) ∗ is in the<br />

image of X /K<br />

n on morphisms. X /K<br />

n is full. □<br />

7.23. Corollary. –––– Let K be a field <strong>and</strong> n ∈Æ. Then, there is an equivalence<br />

of categories<br />

ˇX K n : AzK n −→ BS K n−1 .<br />

Proof. Compose X K n with the equivalence of categories ι: BSK n−1 → (BS K n−1 )op<br />

given by the identity on objects <strong>and</strong> by ι(g) := g −1 on morphisms. □<br />

8. The functor of <strong>points</strong><br />

8.1. Remark. –––– This section deals with the contravariant functor of <strong>points</strong><br />

on the category of all K-schemes defined by the Brauer-Severi variety X A .<br />

It turns out that this functor may be described completely explicitly in terms of<br />

the central simple algebra A.<br />

Thus, there is a different method to introduce X A . One could start with the<br />

functor <strong>and</strong> has to prove its representability by a scheme. It seems that this<br />

method is closer to A. Grothendieck’s style in Algebraic Geometry than the<br />

approach presented here. Unfortunately, the proof of representability is not<br />

trivial at all. It is presented in detail in [Hen] or [Ke].<br />

8.2. Definition. –––– Let K be a field, n ∈Æ, <strong>and</strong> A ∈ Ob (Az K n ).<br />

Then, by<br />

I A : {K-schemes} op → {sets} ,<br />

we will denote the functor given by<br />

{ }<br />

sheaves of right ideals J in π<br />

∗<br />

T ↦→<br />

T A such that<br />

πT ∗ A /J is a locally free O T -module of rank n 2 − n<br />

on objects <strong>and</strong> by pull-back on morphisms.<br />

Here, A := Ã is the sheaf of O K-algebras associated with A on SpecK.<br />

π = π T : T → SpecK denotes the structural morphism.


54 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

8.3. Remark. –––– The various functors I A depend on A in a natural manner.<br />

I.e., if i : A → A ′ is a morphism in Az K n then there is a morphism of functors<br />

i ∗ : I A ′ → I A given by the inverse image<br />

J ↦→ π ∗ T (ĩ) −1 (J ) .<br />

In other words, there exists a contravariant functor<br />

I : Az K n −→ Fun ({K-schemes} op , {sets})<br />

given on objects by A ↦→ I A .<br />

8.4. Theorem. –––– Let K be a field <strong>and</strong> n ∈Æ. Then, there is an isomorphism<br />

between the contravariant functors<br />

ι: I −→ P ◦ X K n<br />

I, P ◦ X K n : AzK n −→ Fun ({K-schemes} op , {sets}) .<br />

8.5. Remark. –––– For a fixed central simple algebra A of dimension n 2 over K,<br />

Theorem 8.4 asserts that there is an isomorphism ι A : I A −→ P XA between<br />

the functors<br />

I A , P XA : {K-schemes} op → {sets} ,<br />

described in Definitions 8.2 <strong>and</strong> 7.3.<br />

This means, the T -valued <strong>points</strong> on X A are in natural bijection to the<br />

sheaves J ⊂ π ∗ T A of right ideals such that π∗ T A /J is a locally free O T -<br />

module of rank n 2 − n. Further, if i : A → A ′ is a morphism of n 2 -dimensional<br />

central simple algebras over K then the diagram<br />

I A ′<br />

ι A ′<br />

P XA ′<br />

is commutative.<br />

i ∗ [X K n (i)] ∗<br />

ι A<br />

I A PXA<br />

8.6. Remark. –––– I A is canonically a subfunctor of the Graßmann functor<br />

Grass n2<br />

n 2 −n parametrizing (n 2 − n)-dimensional quotients of the n 2 -dimensional<br />

st<strong>and</strong>ard vector space. Thus, one has an embedding X A ֒→ Grass n2<br />

n 2 −n into<br />

the Graßmann scheme. This is the key observation for a direct proof of the<br />

representability of I A [Hen, Ke].


Sec. 8] THE FUNCTOR OF POINTS 55<br />

8.7. Definition. –––– Let G be a group.<br />

By Sets G , we will denote the category of all mappings M → G where M is an<br />

arbitrary set. As morphisms from f 1 : M 1 → G to f 2 : M 2 → G, we allow all<br />

mappings making the diagram<br />

M 1<br />

M 2<br />

f 2<br />

f 1<br />

<br />

G<br />

· g<br />

commutative for a certain g ∈ G. Here, · g : G → G denotes multiplication<br />

by g from the right.<br />

8.8. Notation. –––– Let L/K be a finite Galois extension. Then, Sch L/K<br />

+ will<br />

denote the full subcategory of Sch L/K (cf. Definition 7.10) consisting of all<br />

non-empty L-schemes.<br />

8.9. Fact-Definition. –––– Let L/K be a finite Galois extension <strong>and</strong> G be its<br />

Galois group.<br />

Then, for each X ∈ Ob (Sch L/K ), the contravariant functor<br />

G<br />

Hom Sch<br />

L/K ( · , X ): Sch L/K<br />

+ −→ {sets}<br />

factors canonically via Sets G . The resulting functor<br />

P L/K<br />

X<br />

will be called the functor of <strong>points</strong> of X.<br />

Proof. Each morphism T → X in Sch L/K<br />

+<br />

: Sch L/K<br />

+ −→ Sets G<br />

is twisted by a unique element σ ∈ G.<br />

□<br />

8.10. Remark. –––– For that, we need the assumption T ≠ ∅. On the other<br />

h<strong>and</strong>, X = ∅ could have been allowed since there are no morphisms at all in<br />

this case.<br />

8.11. Remarks. –––– a) P L/K<br />

X<br />

is closely related to the ordinary functor of<br />

<strong>points</strong> P X introduced in Definition 7.3. For each non-empty L-scheme T ,<br />

one has P X (T ) = [P L/K<br />

X<br />

(T )] −1 (id).<br />

b) There is a covariant functor<br />

given on objects by X ↦→ P L/K<br />

X<br />

.<br />

P L/K : Sch L/K → Fun ((Sch L/K<br />

+ )op , Sets G )


56 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

c) The functor Hom L/K Sch<br />

( · , X ) remembers all information on X as an object<br />

+<br />

of Sch L/K<br />

+ .<br />

The functor of <strong>points</strong><br />

P L/K<br />

X<br />

: Sch L/K<br />

+ −→ Sets G<br />

carries additional information related to the structure of Sch L/K<br />

+ as a fibered category.<br />

8.12. Lemma. –––– Let n ∈Æ, L/K be a finite Galois extension of fields <strong>and</strong><br />

G := Gal(L/K ) its Galois group.<br />

of functors between the com-<br />

i) Then, there is an isomorphism jn<br />

L/K<br />

position<br />

P L/K<br />

n<br />

: Mat L/K<br />

n<br />

Ξ L/K<br />

n<br />

P L/K<br />

n−1<br />

embedding<br />

: P L/K<br />

n<br />

−→ I L/K<br />

n<br />

Sch L/K<br />

+<br />

P L/K Fun ((Sch L/K<br />

+ )op , Sets G )<br />

<strong>and</strong> the functor In<br />

L/K given by<br />

⎛ ⎧<br />

⎫⎞<br />

A ↦→ ⎝T ↦→ G ⎨ sheaves of right ideals J in (S(σ) ◦ π T ) ∗ A ⎬<br />

such that (S(σ) ◦ π<br />

⎩<br />

T ) ∗ A /J<br />

⎠<br />

σ∈G is a locally free O T -module of rank (n 2 ⎭<br />

− n)<br />

on objects <strong>and</strong> by pull-back on morphisms.<br />

Here, A := Ã denotes the sheaf of O L-algebras associated with A on SpecL.<br />

π T : T → SpecL is the structural morphism. The set on the right h<strong>and</strong> side is<br />

equipped with the canonical map to G.<br />

ii) Let L ′ /L be a finite field extension such that L ′ /K is Galois. Then, the isomorphism<br />

j L/K<br />

n is compatible with base extension from L to L ′ .<br />

I.e., for every A ∈ Ob (Mat L/K<br />

n<br />

P L/K<br />

n (A)(T )<br />

) <strong>and</strong> every T ∈ Ob (Sch L/K<br />

+ ), the diagram<br />

Mor Sch<br />

L/K (T , Ξn<br />

L/K (A))<br />

j L/K<br />

n (A)(T )<br />

I L/K<br />

n (A)(T )<br />

. × SpecL SpecL ′ <br />

Mor Sch<br />

L ′ /K (T× SpecL SpecL ′ , Ξ L′ /K<br />

n (A⊗ L L ′ ))<br />

pull-back<br />

j I L′ /K<br />

n (A⊗ L L ′ )(T× SpecL SpecL ′ )<br />

P L′ /L<br />

n (A ⊗ L L ′ )(T × SpecL SpecL ′ )<br />

commutes. j is an abbreviation for j L′ /K<br />

n (A ⊗ L L ′ )(T × SpecL SpecL ′ ).


Sec. 8] THE FUNCTOR OF POINTS 57<br />

Proof. i) For each A ∈ Ob (Mat L/K<br />

n ), both functors satisfy the sheaf axiom for<br />

Zariski coverings. Thus, it will suffice to verify the assertion on affine schemes.<br />

Let R be a commutative L-algebra with unit. By construction, we have<br />

Ξ L/K<br />

n (A) = P(l) for l a non-zero, simple, left A-module. By Corollary-<br />

Definition 7.6, the set of all R-valued <strong>points</strong> on P(l) is in natural bijection<br />

to the set of all submodules M ⊂ l⊗ L R such that the quotient l⊗ L R/M is a<br />

locally free R-module of rank (n − 1).<br />

As an R-module, A⊗ L R is isomorphic to the direct sum of n copies of l⊗ L R.<br />

Under this isomorphism, an R-submodule M ⊆ l ⊗ R determines a right<br />

ideal M ⊆ A ⊗ L R according to the definition M := L n<br />

i=1 M. This gives a<br />

natural bijection between the set of all right ideals in A ⊗ L R <strong>and</strong> the set of<br />

all R-submodules M ⊆ l. Obviously,<br />

(A ⊗ L R)/M ∼ = ((l ⊗ L R)/M ) n ∼ = ((l ⊗L R)/M ) ⊗ R R n<br />

as R-modules. Thus, if (l⊗ L R)/M is locally free of rank (n−1) then (A⊗ L R)/M<br />

is locally free of rank (n 2 − n). Conversely, if (A ⊗ L R)/M is locally free of<br />

rank (n 2 − n) then, by Lemma 3.14, (l⊗ L R)/M is locally free. Clearly, it is of<br />

rank (n − 1).<br />

To give a morphism of L-schemes f : SpecR → Ξn<br />

L/K (A) which is twisted by<br />

some σ ∈ G is equivalent to giving a commutative diagram<br />

SpecR<br />

π SpecR<br />

<br />

SpecL<br />

id<br />

S(σ)<br />

SpecR<br />

S(σ)◦π SpecR<br />

<br />

SpecL<br />

f<br />

<br />

Ξ L/K<br />

n (A)<br />

id SpecL .<br />

π L/K Ξ n (A)<br />

Therefore, f becomes a morphism of L-schemes in the ordinary sense if one<br />

just changes the structural morphism of SpecR into S(σ) ◦ π.<br />

Consequently, the functor of <strong>points</strong> P L/K is isomorphic to the functor given<br />

Ξ L/K<br />

R (A)<br />

in the assertion.<br />

ii) Again, we need the concrete description of Ξ n<br />

./K . We have Ξ L/K<br />

n (A) = P(l)<br />

for l a non-zero, simple, left A-module. Further, Ξ L′ /K<br />

n (A⊗ L L ′ ) = P(l⊗ L L ′ ).<br />

The claim now easily follows from Remark 7.7.<br />

□<br />

8.13. Remark. –––– Lemma 8.12 states, in particular, that for A ∈ Ob (Mat L/K<br />

the ordinary functor of <strong>points</strong> of the L-scheme Ξn<br />

L/K (A) is isomorphic to<br />

{ }<br />

sheaves of right ideals J in π<br />

∗<br />

T ↦→<br />

T A such that<br />

πT ∗ A /J is a locally free O T -module of rank (n 2 .<br />

− n)<br />

n )<br />

I.e., for every L-algebra R, the set of R-valued <strong>points</strong> on Ξ L/K<br />

n (A) is naturally<br />

isomorphic to the set of all right ideals I in A⊗ L R such that A⊗ L R/I is a locally<br />

free R-module of rank (n 2 − n).


58 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

8.14. Proof of Theorem 8.4. –––– For every A ∈ Ob (Az K n ), we have to construct<br />

an isomorphism ι A : I A → P XA of functors. ι A has to be natural in A.<br />

We will proceed in three steps.<br />

First step. The case A ∼ = M n (K ).<br />

We have A⊗ K R ∼ = M n (R). For X A , there is the intrinsic description given in<br />

Proposition 7.19. X A = P(l) where l denotes a non-zero, simple, left A-module.<br />

The assertion is true by Lemma 8.12. We note explicitly that P(l) = Ξ K/K ′<br />

n (A)<br />

for every subfield K ′ ⊂ K such that K/K ′ is finite Galois. In particular, the<br />

isomorphism ι A : I A → P XA is compatible with automorphisms of A which are<br />

only K ′ -linear.<br />

Second step. Reduction to T -valued <strong>points</strong> for affine T .<br />

We return to the general case that A is an arbitrary central simple algebra<br />

over K of dimension n 2 . The functors I A <strong>and</strong> P XA satisfy the sheaf axioms for<br />

Zariski coverings. Hence, it suffices to consider the affine case.<br />

There is to be constructed a natural isomorphism ι A : I A → P XA of functors on<br />

the full subcategory of affine schemes. I.e., for each commutative K-algebra R<br />

with unit, an isomorphism<br />

{ }<br />

right ideals J in A ⊗K R such that A ⊗<br />

I A (R) :=<br />

K R/J<br />

is a locally free R-module of rank n 2 − n<br />

↓ι A (R)<br />

P XA (R) := Mor K-schemes (SpecR, X A ) .<br />

For a homomorphism r : R → R ′ of K-algebras, the corresponding diagram<br />

I A (R)<br />

ι A (R)<br />

<br />

P XA (R)<br />

is supposed to commute.<br />

Third step. Galois descent.<br />

Let L be a splitting field for A.<br />

Put G := Gal(L/K ).<br />

I A (r)<br />

P XA (r)<br />

I A (R ′ )<br />

ι A (R ′ )<br />

P XA (R ′ )<br />

Assume that L/K is finite <strong>and</strong> Galois.<br />

By Theorem 3.2.ii), there is a bijection<br />

{ }<br />

p : SpecR → XA |<br />

P XA (R) =<br />

p morphism of K-schemes<br />

⎧<br />

⎫<br />

⎨ p : SpecR ⊗ K L → X A⊗K L |<br />

⎬<br />

∼= p morphism of L-schemes,<br />

⎩<br />

⎭ , (§)<br />

p compatible with the G-operations on both sides<br />

natural in the K-algebra R.


Sec. 8] THE FUNCTOR OF POINTS 59<br />

As A⊗ K L is isomorphic to the full matrix algebra, we have<br />

X A⊗K L = Ξn<br />

L/K (A ⊗ K L) = P(l)<br />

where l is a non-zero, simple, right A⊗ K L-module. Hence, there is a second<br />

natural bijection<br />

⎧<br />

⎫<br />

⎨ p : SpecR ⊗ K L → X A⊗K L |<br />

⎬<br />

p morphism of L-schemes,<br />

⎩<br />

⎭<br />

p compatible with the G-operations on both sides<br />

⎧<br />

⎫<br />

J ⊂ (A⊗ K L) ⊗ L (R⊗ K L) |<br />

⎪⎨ J right ideal,<br />

⎪⎬<br />

∼= ((A⊗ K L) ⊗ L (R⊗ K L))/J locally free (R⊗ K L)-module, . ()<br />

rk R⊗K L ((A⊗ K L) ⊗ L (R⊗ K L))/J = (n<br />

⎪⎩<br />

2 − n) ,<br />

J invariant with respect to the G-operation<br />

⎪⎭<br />

Indeed, this is exactly the result of the first step when we note that the isomorphism<br />

of functors I A⊗K L → P XA⊗K is compatible with K-linear automorphisms<br />

L<br />

of A⊗ K L.<br />

Finally, there is a natural bijection<br />

⎧<br />

⎫<br />

J ⊂ (A⊗ K L) ⊗ L (R⊗ K L) |<br />

⎪⎨ J right ideal,<br />

⎪⎬<br />

((A⊗ K L) ⊗ L (R⊗ K L))/J locally free (R⊗ K L)-module,<br />

rk R⊗K L ((A⊗ K L) ⊗ L (R⊗ K L))/J = n<br />

⎪⎩<br />

2 − n ,<br />

J invariant with respect to the G-operation<br />

⎪⎭<br />

⎧<br />

⎫<br />

J ⊂ A⊗ K R |<br />

⎪⎨<br />

⎪⎬<br />

J right ideal,<br />

∼=<br />

= I<br />

(A⊗ ⎪⎩ K R)/J locally free R-module, A (R) (‖)<br />

⎪⎭<br />

rk R (A⊗ K R)/J = (n 2 − n)<br />

by Lemma 8.16. I.e., by Galois descent for right ideals.<br />

Indeed, everything is clear except for the statements on the ranks of the quotients.<br />

For that, if (A⊗ K R)/J is a locally free R-module of rank (n 2 − n) then<br />

((A⊗ K L) ⊗ L (R⊗ K L))/J ∼ = ((A⊗ K R)/J ) ⊗ R (R⊗ K L)<br />

is a locally free (R⊗ K L)-module of the same rank. On the other h<strong>and</strong>, if<br />

((A⊗ K R)/J ) ⊗ R (R⊗ K L)<br />

is a locally free (R ⊗ K L)-module of rank (n 2 − n) then it is a locally free<br />

R-module, too. By Lemma 3.14, (A ⊗ K R)/J is locally free as an R-module.<br />

Clearly, it is of rank (n 2 − n).


60 THE BRAUER-SEVERI VARIETY OF A CENTRAL SIMPLE ALGEBRA [Chap. I<br />

We finally note that the construction of ι A given is functorial in A. For that,<br />

one first has to choose a splitting field L A ⊃ K for each A ∈ Ob (Az K n ) in such a<br />

way that L A depends only on the isomorphism class of A in Az K n . If i : A → A ′<br />

is a morphism in Az K n then A <strong>and</strong> A′ are automatically isomorphic. We execute<br />

the constructions of ι A <strong>and</strong> ι A ′ via the splitting field chosen.<br />

The natural bijections (§) <strong>and</strong> (‖) are applications of descent <strong>and</strong>, therefore,<br />

compatible with the morphisms induced by i. For (), apply Lemma 8.12.i).<br />

The proof is complete.<br />

8.15. Remark. –––– ι A is independent of the choice of the splitting field made<br />

in the proof. It is not difficult to deduce this from Lemma 8.12.ii).<br />

8.16. Lemma (Galois descent for right ideals). —–<br />

Let L/K be a finite Galois extension. Denote by G := Gal(L/K ) its Galois group.<br />

Further, let A be a K-algebra <strong>and</strong> I ⊂ A ⊗ K L a right ideal invariant under the<br />

canonical operation of G on A⊗ K L.<br />

Then, there is a unique right ideal I ⊂ A such that I = I ⊗ K L.<br />

Proof. I inherits from A⊗ K L a structure of an L-algebra. Further, I is naturally<br />

equipped with an operation of G by homomorphisms of K-algebras. σ ∈ G acts<br />

by a σ-linear map.<br />

By the algebraic version of Galois descent, there exists a K-algebra I such<br />

that I = I ⊗ K L.<br />

Furthermore, the canonical homomorphism<br />

I = I ⊗ K L ֒→ A⊗ K L<br />

of L-algebras is compatible with the G-operations on both sides. Therefore,<br />

by Galois descent for homomorphisms, we get a homomorphism I → A<br />

of K-algebras. As the functor ⊗ K L is exact <strong>and</strong> faithful, that homomorphism<br />

is necessarily injective.<br />

Consider I ⊂ A as a subring. For every a ∈ A, the multiplication ·a : I → I<br />

from the right is compatible with the operation of G. Hence, it descends to a homomorphism<br />

I → I. That homomorphism is compatible with multiplication<br />

by a from the right on the whole of A. Consequently, I ⊂ A is a right ideal.<br />

Uniqueness is clear.<br />

□<br />


CHAPTER II<br />

ON THE BRAUER GROUP OF A SCHEME<br />

1. Azumaya algebras<br />

1.1. Definition. –––– Let X be any scheme.<br />

Then, a sheaf of Azumaya algebras or simply an Azumaya algebra over X is a<br />

locally free sheaf A of locally finite rank over X equipped with the structure of<br />

a sheaf of O X -algebras such that the following condition is fulfilled.<br />

For every closed point x ∈ X, one has that A (x) := A ⊗ X k(x) is a central<br />

simple algebra over the residue field k(x).<br />

1.2. Example. –––– If X = Speck is the spectrum of a field then an Azumaya<br />

algebra over X is nothing but a central simple algebra over k.<br />

1.3. Proposition. –––– Let X be any scheme <strong>and</strong> A be any locally free sheaf of<br />

locally finite rank on X having the structure of a sheaf of algebras.<br />

Then, A is an Azumaya algebra over X if <strong>and</strong> only if the canonical homomorphism<br />

ι A : A ⊗ X A op −→ End(A ) ,<br />

a ⊗ b<br />

is an isomorphism of locally free sheaves.<br />

↦→ (x ↦→ axb)<br />

Here, a, b, x ∈ A (U ) denote sections of A over an arbitrary open subset U of X.<br />

Proof. “=⇒” Let A be an Azumaya algebra. Then, A (x) is a central simple<br />

algebra over k(x) for every closed point x ∈ X. By virtue of Remark I.4.2.b),<br />

this implies that ι A is an isomorphism at every closed point. It is therefore an<br />

isomorphism of locally free sheaves.<br />

“⇐=” Let x ∈ X be a closed point. We need to show that A (x) is a central<br />

simple algebra over k(x).<br />

By assumption, we know that<br />

A (x) ⊗ k(x) [A (x)] op ∼ = End(k(x) n ) ∼ = M n (k(x))


62 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

for n := rk x (A ). Further, M n (k(x)) is a central simple algebra over k(x).<br />

If A (x) were not central, Z(A (x)) = L k(x), then<br />

L = L ⊗ k(x) k(x) ⊆ A (x) ⊗ k(x) [A (x)] op ∼ = Mn (k(x))<br />

would be contained in the center of M n (k(x)). This is a contradiction.<br />

Assume, finally, A (x) were not simple. Then, it would contain a proper twosided<br />

ideal (0) I A (x). But, under this assumption, I ⊗ k(x) [A (x)] op would<br />

be a non-trivial, proper two-sided ideal in A (x) ⊗ k(x) [A (x)] op ∼ = M n (k(x)).<br />

We obtained a contradiction. A is, therefore, a central simple algebra.<br />

1.4. Definition. –––– Let X be any scheme <strong>and</strong> A be a locally free sheaf on X<br />

having the structure of a sheaf of algebras.<br />

a) Then, for a point x ∈ X, we will say that A is Azumaya in x if A (x) is a<br />

central simple algebra over k(x).<br />

b) The set of all <strong>points</strong> x ∈ X such that A is not Azumaya in x will be called<br />

the non-Azumaya locus of A .<br />

1.5. Corollary. –––– Let X be any scheme <strong>and</strong> A be any locally free sheaf on X<br />

having the structure of a sheaf of algebras.<br />

a) Then, the non-Azumaya locus T ⊂ X of A is the support of a Cartier divisor.<br />

In particular, it is a closed set in X.<br />

b) Assume, X is regular <strong>and</strong> connected. Then, the non-Azumaya locus of A is either<br />

the whole of X or a subset pure of codimension one.<br />

Proof. a) The assertion is local in X. We may, therefore, assume that A is a free<br />

sheaf, say, of rank n.<br />

ι A : A ⊗ X A op ∼ = O<br />

n 2<br />

X −→ End(A ) ∼ = OX<br />

n2<br />

is then a morphism of free sheaves which are both of rank n 2 .<br />

Whether ι A ⊗ X k(x) is an isomorphism is equivalent to the non-vanishing of its<br />

determinant in x. In the trivialization, the latter is a section s of the structure<br />

sheaf O X . div s is, by definition, the support of a Cartier divisor.<br />

b) This is an immediate consequence of a). □<br />

1.6. Fact. –––– Let A <strong>and</strong> B be two Azumaya algebras on a scheme X.<br />

Then, their tensor product A ⊗ OX B as locally free sheaves over X, equipped with<br />

the obvious structure of an O X -algebra, is again an Azumaya algebra.<br />

Proof. On sections over an open subset U ⊆ X, the algebra structure is given by<br />

(a ⊗ b)(a ′ ⊗ b ′ ) := aa ′ ⊗ bb ′ .<br />


Sec. 1] AZUMAYA ALGEBRAS 63<br />

It is clear that A ⊗ OX B is a locally free O X -module.<br />

The property of being Azumaya may be tested in closed <strong>points</strong>. One has<br />

(A ⊗ OX B)(x) = A ⊗ OX B⊗ OX k(x) = (A ⊗ OX k(x)) ⊗ k(x) (B⊗ OX k(x)) .<br />

On the right h<strong>and</strong> side, the tensor product of two central simple algebras is<br />

again a central simple algebra.<br />

□<br />

1.7. Fact. –––– Let f : X → Y be a morphism of schemes <strong>and</strong> A be an Azumaya<br />

algebra on Y.<br />

Then, the pull-back f ∗ A of A as a locally free sheaf, equipped with the obvious<br />

structure of an O X -algebra, is an Azumaya algebra on X.<br />

Proof. As the assertion is local in both, Y <strong>and</strong> X, we may assume that<br />

f : X = SpecS → Y = SpecR is a morphism of affine schemes <strong>and</strong> that<br />

A is given by an R-algebra A which is free of finite rank as an R-module.<br />

Then, f ∗ A is given by the S-algebra A ⊗ R S. It algebra structure is given,<br />

analogously to the one on the tensor product above, by<br />

(a ⊗ s)(a ′ ⊗ s ′ ) := aa ′ ⊗ ss ′ .<br />

Again, we test the property of being Azumaya in closed <strong>points</strong>.<br />

For that, let x ∈ X <strong>and</strong> put y := f (x). Thus, k(x) is a field extension of k(y).<br />

What we have to show is that A⊗ R S ⊗ S k(x) is a central simple algebra over k(x).<br />

But,<br />

A⊗ R S ⊗ S k(x) = A⊗ R k(x) = A⊗ R k(y)⊗ k(y) k(x)<br />

<strong>and</strong> A⊗ R k(y) is, by assumption, a central simple algebra over k(y).<br />

The assertion follows from Lemma I.4.1.b).<br />

□<br />

1.8. Proposition. –––– Let f : X → Y be a morphism of schemes which is faithfully<br />

flat. Further, let A be a quasi-coherent sheaf on Y, equipped with the structure<br />

of a sheaf of O Y -algebras.<br />

Then, A is an Azumaya algebra on Y if <strong>and</strong> only if f ∗ A is an Azumaya algebra<br />

on X.<br />

Proof. “=⇒” This is Fact 1.7.<br />

“⇐=” The assumption that f ∗ A is an Azumaya algebra on X includes that<br />

f ∗ A is a locally free O X -module, locally of finite rank. We aim first at showing<br />

that this implies A is a locally free O Y -module, locally of finite rank.<br />

For that, let y ∈ Y <strong>and</strong> choose x ∈ X such that f (x) = y. We have a flat local<br />

homomorphism of local rings i : O Y,y → O X,x . By [Mat, Theorem 7.2], O X,x is<br />

faithfully flat over O Y,y .


64 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

Further, ( f ∗ A ) x = A y ⊗ OY,y O X,x is a locally free O X,x -module of finite rank.<br />

Lemma I.3.14 implies that A y is indeed a locally free O Y,y -module <strong>and</strong> of finite<br />

rank. This means, A is a locally free O Y -module, locally of finite rank.<br />

In the rest of the proof, flatness will no more be used.<br />

To verify A is an Azumaya algebra over Y, we will use the criterion provided<br />

by Proposition 1.3. We have f ∗ A = A ⊗ OY O X <strong>and</strong>, therefore,<br />

f ∗ A ⊗ OX [ f ∗ A ] op = (A ⊗ OY O X ) ⊗ OX (O X ⊗ OY [A ] op )<br />

∼= A ⊗ OY [A ] op ⊗ OY O X<br />

= f ∗ (A ⊗ OY [A ] op ) .<br />

Further,<br />

End( f ∗ A ) = f ∗ (End(A ))<br />

<strong>and</strong> the isomorphisms are compatible in such a way that<br />

f ∗ ι A = ι f ∗ A .<br />

Thus, we are given that f ∗ ι A = ι f ∗ A is an isomorphism of locally free sheaves.<br />

We have to prove this implies that the original ι A was an isomorphism.<br />

For that, note that det( f ∗ ι A ) = f # det(ι A ). Thus, we know that f # det(ι A ) is<br />

a nowhere vanishing section of the invertible sheaf<br />

f ∗ [Λ max (A ⊗ OY [A ] op ) ∨ ⊗ Λ max (End(A ))] .<br />

We have to show that det(ι A ) is nowhere vanishing.<br />

Again, for y ∈ Y, we choose x ∈ X such that f (x) = y. Then, we have a local<br />

homomorphism of local rings i : O Y,y → O X,x . As we work in stalks, we may<br />

choose a trivialization <strong>and</strong> write det(ι A ) ∈ O Y,y . Then,<br />

f # det(ι A ) = i(det(ι A )) ∈ O X,x .<br />

If det(ι A ) were vanishing at y then this would mean simply det(ι A ) ∈ m Y,y .<br />

Then, i(det(ι A )) ∈ m X,x . I.e., f # det(ι A ) vanishes at x.<br />

This is a contradiction.<br />

1.9. Proposition. –––– Let f : X → Y be a morphism of schemes which is faithfully<br />

flat <strong>and</strong> quasi-compact.<br />

Then, it is equivalent to give<br />

a) an Azumaya algebra A over Y,<br />

b) an Azumaya algebra B over X together with an isomorphism<br />

Φ: pr ∗ 1 B → pr∗ 2 B<br />


Sec. 2] THE BRAUER GROUP 65<br />

of Azumaya algebras on X × Y X which satisfies<br />

on X × Y X × Y X.<br />

pr ∗ 31 (Φ) = pr∗ 32 (Φ) ◦ pr∗ 21 (Φ)<br />

Sketch of Proof. “a) =⇒ b)” Put B := f ∗ A .<br />

“b) =⇒ a)” This is what is called faithful flat descent.<br />

The existence of A as a quasi-coherent O Y -module satisfying B := f ∗ A follows<br />

directly from [K/O74, Theorem II.3.2]. The additional algebra structure descends,<br />

too. This may easily be seen from the construction of the descent module<br />

given in the proof. It is shown in [K/O74, Theorem II.3.4].<br />

A is an Azumaya algebra over Y by virtue of Proposition 1.8.<br />

1.10. Remark. –––– This means that an Azumaya algebra over a scheme X<br />

may be described by local gluing data with respect to the fpqc-topology or any<br />

weaker topology.<br />

Intuitively, B describes the desired Azumaya algebra on each set of a cover.<br />

Φ collects the gluing isomorphisms on intersections of two sets of the cover.<br />

The condition on X × Y X × Y X is a cocycle condition encoding that the gluing<br />

maps be compatible.<br />

We will use this approach for constructing Azumaya algebras from data local in<br />

the étale topology.<br />

□<br />

2. The Brauer group<br />

2.1. Definition. –––– Let X be any scheme.<br />

Two Azumaya algebras A <strong>and</strong> B on X are said to be similar if there exist two<br />

locally free O X -modules E <strong>and</strong> F, both everywhere of positive rank, such that<br />

A ⊗ OX End(E ) ∼ = B ⊗ OX End(E ′ ) .<br />

2.2. Remarks. –––– i) Similarity is an equivalence relation.<br />

For that, the only point which is not entirely obvious is transitivity. To achieve<br />

this, assume<br />

A ⊗ OX End(E ) ∼ = B ⊗ OX End(E ′ ) <strong>and</strong> B ⊗ OX End(E ′′ ) ∼ = C ⊗ OX End(E ′′′ ) .


66 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

Then,<br />

A ⊗ OX End(E ⊗ OX E ′′ ) ∼ = A ⊗ OX End(E ) ⊗ OX End(E ′′ )<br />

∼= B ⊗ OX End(E ′ ) ⊗ OX End(E ′′ )<br />

∼= B ⊗ OX End(E ′′ ) ⊗ OX End(E ′ )<br />

∼= C ⊗ OX End(E ′′′ ) ⊗ OX End(E ′ )<br />

∼= C ⊗ OX End(E ′′′ ⊗ OX E ′′′ ) .<br />

ii) Similarity is compatible with the tensor product of Azumaya algebras <strong>and</strong><br />

with pull-back.<br />

2.3. Definition. –––– Let X be any scheme.<br />

The set of all similarity classes of Azumaya algebras on X is called the Brauer<br />

group of X <strong>and</strong> denoted by Br(X ).<br />

2.4. Remarks. –––– i) The similarity class of an Azumaya algebra A will be<br />

denoted by [A ].<br />

ii) A binary operation on Br(X ) is given by [A ] ∗ [A ′ ] := [A ⊗ OX A ′ ] .<br />

iii) For this binary operation, [End(E )] is the neutral element. Here, E is an<br />

arbitrary locally free sheaf, nowhere of rank zero.<br />

iv) The inverse element of [A ] is given by [A op ]. This shows that Br(X ) is<br />

indeed a group.<br />

3. The cohomological Brauer group<br />

3.1. Lemma. –––– Let (R, m) be a Henselian local ring <strong>and</strong> A be an Azumaya<br />

algebra over SpecR. Assume<br />

A ⊗ OSpecR O SpecR/m<br />

∼ = End(O<br />

n<br />

SpecR/m )<br />

for a certain n ∈Æ. Then, A ∼ = End(O n SpecR ).<br />

Proof. We denote by A the R-algebra corresponding to A . In this notation,<br />

we, therefore, have A ⊗ R R/m ∼ = M n (R/m). On the right h<strong>and</strong> side, choose an<br />

idempotent matrix ǫ of rank one.<br />

Let a ∈ A be such that a := (a mod mA) = ǫ. Then, R[a] is a finite<br />

commutative R-algebra. As R is Henselian, [SGA4, Exp. VIII, 4.1] shows that<br />

R[a] is a direct product of local rings. From R[a]⊗ R k = k[ǫ] ∼ = k ⊕ k, we see<br />

that R[a] ∼ = R/I ⊕ R/J is actually composed of two local R-algebras. In k ⊕ k,<br />

let ǫ correspond to the element (1, 0). Thus, the element e ∈ R[a] mapped<br />

to (1, 0) under the isomorphism R[a] → R/I ⊕ R/J is an idempotent lifting ǫ.


Sec. 3] THE COHOMOLOGICAL BRAUER GROUP 67<br />

Consider the homomorphism of R-algebras<br />

φ: A −→ End R (Ae) ,<br />

a ↦→ (be ↦→ abe) .<br />

Both R-modules are free of the same rank. It is classically known that φ⊗ R k<br />

is an isomorphism. Nakayama’s lemma ensures that φ is an isomorphism itself.<br />

□<br />

3.2. Corollary. –––– Let (R, m) be a strictly Henselian local ring <strong>and</strong> A be an<br />

Azumaya algebra over SpecR.<br />

Then, A ∼ = End(O n SpecR ).<br />

Proof. k = R/m is a separably closed field. By virtue of Remark I.4.2.b), we<br />

know that<br />

A ⊗ OSpecR O SpecR/m<br />

∼ = End(O<br />

n<br />

SpecR/m )<br />

is a full matrix algebra over R/m.<br />

□<br />

3.3. Lemma (Theorem of Skolem-Noether, version over local rings). —–<br />

Let R be a commutative ring with unit. Then, GL n (R) operates on M n (R) by<br />

conjugation,<br />

(g, m) ↦→ gmg −1 .<br />

If R = (R, m) is a local ring then this defines an isomorphism<br />

PGL n (R) := GL n (R)/R ∗ ∼ =<br />

−→ Aut R (M n (R)) .<br />

Proof. (Compare Lemma I.4.4 where the special case R is a field was treated.)<br />

One has R = Zent(M n (R)). Therefore, the mapping is well-defined <strong>and</strong> injective.<br />

Surjectivity. Let j : M n (R) → M n (R) be an automorphism. We consider the<br />

algebra<br />

M := M n (R) ⊗ R M n (R) op ( ∼ = End R-mod (M n (R)) ∼ = M n 2(R)) .<br />

M n (R) gets equipped with the structure of a left M-module in two ways.<br />

(A ⊗ B) • 1 C := A · C · B<br />

(A ⊗ B) • 2 C := j(A) · C · B<br />

We will denote the resulting M-modules by N 1 <strong>and</strong> N 2 , respectively.


68 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

Our first claim is that N 1 is a projective M-module. This means, we have to<br />

show that M n (R) is projective when equipped with its canonical structure as<br />

an End R-mod (M n (R))-module.<br />

For that, consider a morphism i : M n (R) → R of R-modules which is a section<br />

of the canonical inclusion mapping r to r·E. Note such a section exists because<br />

M n (R) is a projective R-module.<br />

Then, the surjection of End R-mod (M n (R))-modules End R-mod (M n (R)) → M n (R),<br />

given by f ↦→ f (1), admits the section<br />

m ↦→ (m ′ ↦→ i(m ′ )m) .<br />

This shows M n (R) is a projective End R-mod (M n (R))-module.<br />

Over the field R/m, it is known that two M n 2(R/m)-modules of the same<br />

R/m-dimension are isomorphic. Thus, there is an isomorphism<br />

As M 1 is projective, the map<br />

M 1<br />

g : M 1 ⊗ R R/m → M 2 ⊗ R R/m .<br />

π<br />

−→ M 1 ⊗ R R/m −→ g<br />

M 2 ⊗ R R/m<br />

may be lifted to a homomorphism h : M 1 → M 2 of End R-mod (M n (R))-modules.<br />

h⊗ R R/m is an isomorphism. Nakayama’s lemma implies that h is an isomorphism<br />

itself.<br />

Let us put I := h(E) to be the image of the identity matrix.<br />

M ∈ M n (R) we have<br />

h(M ) = h((E ⊗ M ) • 1 E) = (E ⊗ M ) • 2 h(E) = h(E) · M = I · M.<br />

In particular, I ∈ GL n (R). Therefore,<br />

I · M = h(M ) = h((M ⊗ E) • 1 E) = (M ⊗ E) • 2 h(E) = j(M ) · I<br />

For every<br />

for each M ∈ M n (R) <strong>and</strong> j(M ) = IMI −1 .<br />

□<br />

3.4. Proposition –––– Let X be any quasi-compact scheme. Then, the set of all isomorphismclassesofrankn<br />

2 AzumayaalgebrasonX isgivenbytheČechcohomology<br />

set Ȟ 1 ét (X, PGL n).<br />

Proof. We know by Proposition 1.9 that Azumaya algebras may be described<br />

by gluing data local in the étale topology.<br />

Let A be any Azumaya algebra over X. For every point x ∈ X, Corollary<br />

3.2 yields that A is isomorphic to matrix algebra after strict Henselization,<br />

A ⊗ OX OX,x sh ∼ = M n (OX,x sh ). To describe the isomorphism, only finitely many


Sec. 3] THE COHOMOLOGICAL BRAUER GROUP 69<br />

data on O sh<br />

X,x are necessary. For this reason, there exists an affine scheme U x,<br />

which is an étale neighbourhood of x, such that A | Ux<br />

∼ = End(O<br />

n<br />

Ux<br />

). By virtue of<br />

quasi-compactness, one may select finitely many of the U i covering the whole<br />

of X.<br />

All these étale neighbourhoods together form gluing data for A since the morphism<br />

U x1<br />

F . . .<br />

F<br />

Uxl → X is faithfully flat <strong>and</strong> quasi-compact.<br />

The conditions for gluing data formulated in Proposition 1.9 are equivalent to<br />

giving a Čech cocycle for Ȟ 1 ét (X, Aut(M n)). Different gluing data describing the<br />

same algebra differ by a coboundary.<br />

The Theorem of Skolem-Noether implies that the sheaf Aut(M n ) is the same<br />

as PGL n .<br />

□<br />

3.5. Remark. –––– On a quasi-compact scheme X, the set of all isomorphism<br />

classes of rank n locally free sheaves is given by the Čech cohomology set<br />

Ȟ 1 ét (X, GL n).<br />

This is proven in the same way as Proposition 3.4.<br />

3.6. Notation. –––– Let X be any scheme.<br />

i) We will denote the cohomology class corresponding to an Azumaya algebra<br />

A over X by c(A ) ∈ Ȟ 1 ét (X, PGL n).<br />

ii) We will denote the cohomology class corresponding to a locally free sheaf E<br />

over X by cl(E ) ∈ Ȟ 1 ét (X, GL n).<br />

iii) Recall that there is a short exact sequence<br />

0 −→m −→ GL n −→ PGL n −→ 0<br />

of sheaves. In cohomology, it induces a long exact sequence of pointed sets<br />

Ȟ 1 ét (X, GL n)<br />

ι<br />

−→ Ȟ 1 ét (X, PGL d<br />

n) −→ H 2 ét (X,m).<br />

3.7. Facts. –––– Let X be a quasi-compact scheme.<br />

i) For a locally free sheaf E on X, one has that<br />

ι(cl(E )) = c(End(E )) .<br />

ii) For two Azumaya algebras A <strong>and</strong> B on X, of ranks n <strong>and</strong> m, respectively,<br />

one has<br />

c(A ⊗ OX B) = c(A )∗c(B) .<br />

Here, the operation ∗ is induced by the canonical mapping<br />

PGL n × PGL m −→ PGL nm


70 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

given by<br />

GL n ×GL m = Aut(n a )×Aut(m a ) −→ Aut(n a ⊗m a ) = Aut(nm<br />

a ) = GL nm .<br />

iii) For α ∈ Ȟ 1 ét (X, PGL n) <strong>and</strong> β ∈ Ȟ 1 ét (X, PGL n), we have<br />

d(α∗β) = d(α) + d(β) .<br />

Proof. The verification of these compatibilities is purely technical <strong>and</strong> hence<br />

omitted.<br />

□<br />

3.8. Proposition. –––– Let X be any quasi-compact scheme.<br />

Then, the boundary maps d induce an injective group homomorphism<br />

Proof. First step. i X is well-defined.<br />

i X : Br(X ) −→ H 2 ét (X,m) .<br />

Suppose we are given two Azumaya algebras A <strong>and</strong> B over X which are<br />

similar. This means, we have two locally free sheaves, E <strong>and</strong> F, such that<br />

A ⊗ OX End(E ) ∼ = B ⊗ OX End(E ′ ). In particular,<br />

c(A ⊗ OX End(E )) = c(B ⊗ OX End(E ′ )) .<br />

(∗)<br />

The compatibilities above yield<br />

hence<br />

c(A ⊗ OX End(E )) = c(A )∗c(End(E )) = c(A )∗ι(cl(E )) ,<br />

d(c(A )) = d(c(A ))+d(ι(cl(E ))) = d(c(A )∗ι(cl(E ))) = d(c(A ⊗ OX End(E ))) .<br />

Completely analogously, one shows that d(c(B)) = d(c(B ⊗ OX End(E ′ ))) .<br />

Formula (∗), therefore, yields the claim.<br />

Second step. i X is a group homomorphism.<br />

We have d(c(A ⊗ OX B)) = d(c(A )∗c(B)) = d(c(A )) + d(c(B)) .<br />

Third step. i X is an injection.<br />

Suppose, we have an Azumaya algebra A such that d(c(A )) = 0. Then, according<br />

to the exactness of the long cohomology sequence, c(A ) = ι(cl(E )) for<br />

a locally free sheaf E on X. Thus, A ∼ = End(E ) <strong>and</strong> [A ] = 0 ∈ Br(X ). □<br />

3.9. Definition. –––– Let X be any scheme. Then, H 2 ét (X,m) is called the<br />

cohomological Brauer group of X. It will be denoted by Br ′ (X ).


Sec. 4] THE RELATION TO THE BRAUER GROUP OF THE FUNCTION FIELD 71<br />

3.10. Remark. –––– i X is, in general, not an isomorphism of <strong>groups</strong>. The first<br />

counterexample to surjectivity has been constructed by A. Grothendieck<br />

in [GrBrII, Remarque 1.11.b].<br />

i X is, however, an isomorphism in a number of particular cases. One such case<br />

are smooth algebraic surfaces. This was essentially known to M. Ausl<strong>and</strong>er<br />

<strong>and</strong> O. Goldman [A/G] before Grothendieck’s invention of the general Brauer<br />

group for schemes, may be even before the actual invention of the concept of<br />

a scheme.<br />

In Section 7, we will present the result of Ausl<strong>and</strong>er <strong>and</strong> Goldman in a more<br />

up-to-date formulation. Although this is not the most general result known<br />

today, we hope it may serve as a good illustration for what Br(X ) <strong>and</strong> Br ′ (X )<br />

are <strong>and</strong> which methods might be used in order to compare them.<br />

4. The relation to the Brauer group of the function field<br />

4.1. –––– In this section, we always assume that X is an integral scheme.<br />

Denote by g : SpecK = η → X the inclusion of the generic point.<br />

4.2. Fact. –––– Let X be an integral scheme which is regular, separated, <strong>and</strong> quasicompact.<br />

Then,<br />

Br(X ) ⊆ Br ′ (X ) ⊆ Br ′ (K )<br />

where K := Q(X ) denotes the function field of X.<br />

Proof. Denote by g : η → X the inclusion of the generic point. We have the<br />

short exact sequence<br />

0 −→m,X −→ g ∗m,K −→ Div X −→ 0<br />

of sheaves in the étale topology. Here, Div X = L<br />

H 1 ét (X, Div X ) = M<br />

It follows, at first, that<br />

codim x=1<br />

codimx=1<br />

H 1 ét ({x},) = M<br />

codimx=1<br />

H 2 ét (X,m) ⊆ H 2 ét (X, g ∗m,K ).<br />

Second, we consider the Leray spectral sequence<br />

i x∗x which implies<br />

Hom(G x ,) = 0 .<br />

E p,q<br />

2<br />

:= H p<br />

ét (X, Rq g ∗m,K ) =⇒ H n ét (SpecK,m,K).<br />

By virtue of Hilbert’s Theorem 90, for the first higher direct image, we<br />

have R 1 g ∗m,K = 0. This implies there is a short exact sequence


72 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

0 −→ H 2 ét (X, g ∗m,K ) −→ H 2 ét (SpecK,m,K) −→ H 0 ét (X, R2 g ∗m,K ) −→ 0 .<br />

(†)<br />

of terms of lower order.<br />

□<br />

4.3. Remark. –––– One may therefore think of Br(X ) as follows. It consists of<br />

those classes of the usual Br(Q(X )) which allow an extension over the whole<br />

of X.<br />

4.4. Lemma. –––– Let X be an integral scheme which is separated <strong>and</strong> quasicompact.<br />

Then, for each q > 0, the sheaf R q g ∗m,K is uniquely l-divisible for every<br />

l ∈prime to the residue characteristics of X.<br />

Proof. The inclusion g : η → X is the inverse limit, in the category of all<br />

X-schemes, of the open embeddings g U : U → X for U ⊆ X affine open.<br />

The assumption that X is separated makes sure that all the transition maps<br />

g U ′ ,U : U ′ ⊆<br />

−→ U are affine. In this situation, from [SGA4, Exp. VII, Corollaire<br />

5.11], we see that<br />

R q g ∗ µ l = lim R q g U∗ −→U<br />

µ l .<br />

As the g U are open embeddings, <strong>and</strong> hence smooth, they are acyclic [SGA4,<br />

Exp. XV, Théorème 2.1]. In particular, for q > 0 <strong>and</strong> l prime to the residue<br />

characteristics of X, we have R q g U∗ µ l = R q g U∗ g U ∗ µ l,X = 0.<br />

Altogether,<br />

R q g ∗ µ l = 0 .<br />

The Kummer sequence implies that the multiplication map<br />

l : R q g ∗m,K → R q g ∗m,K<br />

is an isomorphism for every q > 0.<br />

□<br />

4.5. Proposition. –––– Let X be an integral scheme which is regular, separated, <strong>and</strong><br />

quasi-compact.<br />

a) If char(Q(X )) = p > 0 then H 0 ét (X, R2 g ∗m,K ) is a p-power torsion group.<br />

b) If all residue characteristics of X are equal to zero then H 0 ét (X, R2 g ∗m,K ) = 0.<br />

Proof. H 2 ét (SpecK,m,K) = H 2( Gal(K/K ), K ∗) is a torsion group directly by<br />

its definition. As we have a surjection H 2 ét (SpecK,m,K) → H 0 ét (X, R2 g ∗m,K ),<br />

the latter is a torsion group, too.


Sec. 4] THE RELATION TO THE BRAUER GROUP OF THE FUNCTION FIELD 73<br />

b) For every l > 0, the sheaf R 2 g ∗m,K is uniquely l-divisible. Therefore, the<br />

multiplication map<br />

l : H 0 ét (X, R2 g ∗m,K ) → H 0 ét (X, R2 g ∗m,K )<br />

is an isomorphism. As H 0 ét (X, R2 g ∗m,K ) is a torsion group, this implies<br />

H 0 ét (X, R2 g ∗m,K ) = 0.<br />

a) Here, the same argument still works for l prime to p. This shows<br />

H 0 ét (X, R2 g ∗m,K ) has no prime-to-p torsion.<br />

□<br />

4.6. Corollary. –––– Let X be an integral scheme which is regular, separated, <strong>and</strong><br />

quasi-compact. Assume that the residue characteristics of X are equal to zero.<br />

Then, for g : η → X the inclusion of the generic point, one has<br />

H 2 ét (X, g ∗m,K ) ∼ = H 2 ét (SpecK,m,K ) .<br />

Proof. This follows from the exact sequence (†) together with Proposition 4.5.b).<br />

□<br />

4.7. Remark. –––– The same result is true for schemes of dimension at most<br />

one in any characteristic as is often shown in the literature. See, e.g., [GrBrIII,<br />

formule (2.2)].<br />

4.8. Proposition. –––– Let X be an integral scheme which is regular, separated, <strong>and</strong><br />

quasi-compact. Assume that<br />

i) dim X ≤ 1 or<br />

ii) the residue characteristics of X are equal to zero.<br />

Then, there is a short exact sequence<br />

0 −→ Br ′ (X ) −→ Br ′ (SpecQ(X )) −→ M<br />

Hom(G x ,É/) .<br />

codim x=1<br />

Proof. Consider once more the short exact sequence<br />

0 −→m,X −→ g ∗m,K −→ Div X −→ 0<br />

of sheaves in the étale topology. Since H 1 ét (X, Div X ) = 0, as shown above, we<br />

have the following fragment of the long exact sequence<br />

0 −→ H 2 ét (X,m,X ) −→ H 2 ét (X, g ∗m,K ) −→ H 2 ét (X, Div X )<br />

in cohomology. Here, the entry on the left h<strong>and</strong> side is Br ′ (X ), according to<br />

its very definition. The term in the middle is isomorphic to Br ′ (Q(X )) as was


74 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

shown in Corollary 4.6. For the term on the right h<strong>and</strong> side, we have<br />

H 2 ét (X, Div X ) = M<br />

H 2 ét ({x},)<br />

codim x=1<br />

= M<br />

H 2 (G x ,)<br />

codim x=1<br />

= M<br />

Hom(G x ,É/) .<br />

codim x=1<br />

□<br />

4.9. Corollary. –––– Let X be an integral scheme which is regular, separated, <strong>and</strong><br />

quasi-compact. Assume that<br />

i) dim X ≤ 1 or<br />

ii) the residue characteristics of X are equal to zero.<br />

Then, one has<br />

Br ′ (X ) = \<br />

codimx=1<br />

where the intersection takes place in Br ′ (Q(X )).<br />

Proof. “⊆” is clear.<br />

Br ′ (Spec O X,x )<br />

“⊇” From Proposition 4.8, we see that all obstructions against an extension of<br />

a central simple algebra over Q(X ) to the whole of X are given by the <strong>points</strong><br />

of codimension one. For SpecO X,x instead of X, the same is true. □<br />

4.10. Remark. –––– Under certain hypotheses on X, there is the same formula<br />

for the Brauer group instead of the cohomological Brauer group. Compare Theorem<br />

7.1.i), below.<br />

5. The Brauer group <strong>and</strong> the cohomological Brauer group<br />

5.1. –––– In order to compare the two Brauer <strong>groups</strong> in relatively simple cases,<br />

the following lemma is helpful.<br />

5.2. Lemma. –––– Let X = SpecR for a local ring R <strong>and</strong> γ ∈ Br ′ (X ) an element<br />

in the cohomological Brauer group. Assume there exists a morphism π : Y → X of<br />

schemes which is finite <strong>and</strong> faithfully flat such that γ ∈ ker(Br ′ (X ) → Br ′ (Y )).<br />

Then, one has even γ ∈ Br(X ).<br />

Proof. We have Y = SpecS. Nakayama’s lemma implies that S ∼ = R n is free<br />

as an R-module. As π is finite, it has no higher direct images <strong>and</strong> the Leray<br />

spectral sequence collapses to<br />

H 2 ét (Y,m) = H 2 ét (X, π ∗m) .


Sec. 5] THE BRAUER GROUP AND THE COHOMOLOGICAL BRAUER GROUP 75<br />

We also have the long exact sequence<br />

H 1 ét (X, π ∗m/m) −→ H 2 ét (X,m) −→ H 2 ét (X, π ∗m)<br />

at our disposal. γ vanishes in H 2 ét (X, π ∗m). It belongs, therefore, to the image<br />

of H 1 ét (X, π ∗m/m).<br />

The canonical map S ∗<br />

of sheaves<br />

֒→ Aut(R n ) = GL n (R) induces a homomorphism<br />

π ∗m ֒→ GL n .<br />

Finally, we note that the diagram<br />

H 1 ét (X, π ∗m/m)<br />

H 2 ét (X,m)<br />

Ȟ 1 ét (X, GL n/m)<br />

Ȟ 2 ét (X,m)<br />

commutes. Hence, γ is in the image of Ȟ 1 ét (X, GL n/m) = Ȟ 1 ét (X, PGL n).<br />

□<br />

5.3. Corollary. –––– One has Br(SpecK ) = Br ′ (SpecK ) for every field K.<br />

Proof. Let γ ∈ Br ′ (SpecK ) = H 2( Gal(K/K ), K ∗) . Then, there is a finite Galois<br />

extension L/K such that γ ∈ H 2( Gal(L/K ), L ∗) , already. When restricted<br />

to SpecL, the cohomology class γ vanishes.<br />

□<br />

5.4. Corollary. –––– One has Br(SpecR) = Br ′ (SpecR) = Br(SpecR/m) for<br />

every Henselian local ring (R, m). In particular, Br(SpecR) = 0 for R strictly local.<br />

Proof. The equality Br ′ (SpecR) = Br ′ (Spec R/m) follows directly from a<br />

general result on étale cohomology [SGA4, Exp. VIII, Corollaire 8.6].<br />

Thus, let γ ∈ Br ′ (SpecR). For its image in Br ′ (Spec R/m), choose a splitting<br />

field l which is finite over k := R/m. Let a be a primitive element <strong>and</strong> f ∈ k[X ]<br />

be its minimal polynomial. Then, l ∼ = k[X ]/( f ). Choose a lift F ∈ R[X ] of f .<br />

The local ring (S, n) := R[X ]/(F ) is finite <strong>and</strong> flat over R. S is Henselian<br />

by [SGA4, Exp. VIII, 4.1.ii)]. Under pull-back to SpecS, the cohomology<br />

class γ vanishes since Br ′ (SpecS) = Br ′ (Spec S/n).<br />

□<br />

5.5. Corollary. –––– One has Br(SpecR) = Br ′ (SpecR) for every discrete valuation<br />

ring R.<br />

Proof. Let γ ∈ Br ′ (SpecR) ⊆ Br ′ (SpecQ(R)). We know that γ vanishes after<br />

a finite, separable field extension L/Q(R). Take for S the integral closure of R<br />

in L. Then, γ vanishes in Br ′ (SpecS) ⊆ Br ′ (L) <strong>and</strong> S is finite <strong>and</strong> flat over R.<br />


76 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

5.6. Remark. –––– For X an integral scheme which is regular, separated, <strong>and</strong><br />

quasi-compact in characteristic zero, one therefore has<br />

Br ′ (X ) = \<br />

Br(SpecO X,x )<br />

in Br(Q(X )).<br />

codimx=1<br />

6. Orders in central simple algebras<br />

i. Orders over a general scheme. —<br />

6.1. Definition. –––– Let X be an integral scheme <strong>and</strong> A be a central simple<br />

algebra over its quotient field Q(X ).<br />

By an order in A over X, one means an O X -algebra A which is coherent as<br />

an O X -module <strong>and</strong> has A as its stalk at the generic point.<br />

6.2. Remark. –––– Assume the integral scheme X is locally Noetherian.<br />

Then, for every central simple algebra A over Q(X ) there exists an order<br />

over X.<br />

Indeed, an order may be constructed as follows. Choose a set {a 1 , . . . , a l } of<br />

generators of A as a Q(X )-vector space. In the sheaf g ∗ A on X, these generate<br />

a coherent subsheaf G . Then, the sheaf (G : G ) ⊆ g ∗ A which is defined by<br />

(G : G )(U ) := {s ∈ g ∗ A(U ) | G (U )·s ⊆ G (U ) }<br />

for every open U ⊆ X is, by definition, an O X -algebra. Clearly, it is coherent<br />

<strong>and</strong> has A as its stalk at the generic point. It is, therefore, an order.<br />

6.3. Remarks. –––– i) If X = SpecR is the spectrum of a Noetherian, integrally<br />

closed domain then every global section s ∈ Γ(X, A ) ⊆ A of every order A is<br />

integral over R.<br />

ii) On a central simple algebra A over a field K, there is the reduced trace<br />

tr: A → K. It admits the property that<br />

(x, y) ↦→ tr(xy)<br />

is a non-degenerate bilinear form [Re, Theorem (9.26)].<br />

Further, if R ⊆ K is integrally closed <strong>and</strong> s a section of an order over SpecR<br />

then tr(s) ∈ R.


Sec. 6] ORDERS IN CENTRAL SIMPLE ALGEBRAS 77<br />

6.4. Lemma. –––– Let X be an integral scheme which is Noetherian <strong>and</strong> normal.<br />

Then, every order A over X in a central simple algebra A over Q(X ) is contained<br />

in a maximal order.<br />

Proof. Let SpecR ⊆ X be affine open. Then, A | SpecR is generated by finitely<br />

many global sections s 1 , . . . , s l . Clearly, s 1 , . . . , s l generate A over K.<br />

If A ′ ⊃ A | SpecR is any order <strong>and</strong> t any global section then ts 1 , . . . , ts l are<br />

global sections <strong>and</strong> tr(ts 1 ), . . . , tr(ts l ) ∈ R. This list of conditions defines a<br />

finite R-module, i.e. a coherent sheaf à over SpecR containing every order<br />

which contains A.<br />

Since X is covered by finitely many open affine sets, every ascending chain of<br />

orders in A over X stops after finitely many steps.<br />

□<br />

6.5. Lemma. –––– Let X be an integral scheme which is Noetherian <strong>and</strong> normal.<br />

Let A be a maximal order over X in a central simple algebra A over Q(X ).<br />

Then, for every (not necessarily closed) point x ∈ X, A | SpecOX,x is a maximal order<br />

in A over Spec O X,x .<br />

Proof. For X affine, this follows from [Re, Theorem (1.11)].<br />

For the general case, it remains to verify that A | V is maximal for every affine<br />

open subscheme SpecR = V ⊆ X. We shall do this by contradiction.<br />

Assume, there would be a larger order B over V . Then, there exists some<br />

non-zero r ∈ R such that r·B ⊆ A . Extend r·B to a coherent subsheaf E ⊆ A<br />

on X in the obvious way. I.e., a local section s ∈ E (U ) is, by definition, a<br />

section of A such that s| U ∩V ∈ (r·B)(U ∩ V ).<br />

Now, consider the order (E : E ) over X. Its restriction to V is easily understood.<br />

One has (E : E )| V = ((r ·B) : (r ·B)) = (B : B) = B as r<br />

commutes with B <strong>and</strong> B is an order. In particular, A | V ⊂ B has the property<br />

that E | V ·A | V ⊆ E | V . Every local section s ∈ (E ·A )(U ) is, therefore, a local<br />

section of A with the additional property that s| U ∩V ∈ E (U ∩ V ). By construction<br />

of E , this means s ∈ E (U ). Hence, E ·A ⊆ E <strong>and</strong> A ⊆ (E : E ).<br />

Since A is a maximal order, this implies A = (E : E ).<br />

Together with (E : E )| V = B, this shows A | V = B in contradiction to our<br />

assumption, B would be a larger order.<br />

(This is nothing but a geometrization of the proof for [Re, Theorem (1.11)].)<br />

□<br />

6.6. –––– Our interest in maximal orders comes from the following proposition.


78 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

6.7. Proposition. –––– Let X be a Noetherian, normal, integral scheme with<br />

generic point η <strong>and</strong> A be an Azumaya algebra over X.<br />

Then, A is a maximal order in A η over X.<br />

Proof. It is clear from the definition that A is an order.<br />

If it were not maximal then there would be a larger order B. B admits a local<br />

section s ∉ A (U ) over a certain subscheme U. We may therefore assume that<br />

X = SpecR is affine, A = Ã, <strong>and</strong> B = ˜B for B A.<br />

We have A⊗ R A op = M n (R) for a certain n ∈Æ. Then, O := B⊗ R A op M n (R)<br />

would be a larger order. I.e., the order M n (R) would not be maximal.<br />

Choose M ∈ O \M n (R). There is one entry x not in R. Elementary matrix<br />

operations produce a diagonal matrix<br />

⎛ ⎞<br />

x 0 · · · 0<br />

0 x · · · 0<br />

⎜<br />

⎝<br />

.<br />

. . ..<br />

⎟ . ⎠ ∈ O .<br />

0 0 · · · x<br />

This element alone generates the commutative R-algebra R[x] which can not be<br />

finite since R is integrally closed. This is a contradiction.<br />

□<br />

6.8. Remark. –––– The same approach yields that M n (A ) is a maximal order<br />

as soon as A is.<br />

ii. Orders over a discrete valuation ring. —<br />

6.9. Notation. –––– In this subsection, we will consider the following situation.<br />

– (R, m) is a discrete valuation ring. By π, we denote a generator of m.<br />

– K := Q(R) is its quotient field,<br />

– A is a central simple algebra over K.<br />

As usual, ̂R <strong>and</strong> ̂K are the completions of R <strong>and</strong> K.<br />

For x ∈ A, we will denote by N A/K (x) := det(·x : A → A) the (non-reduced)<br />

norm of x.<br />

Finally, we will simply formulate that O is an Azumaya algebra (an order)<br />

over R when Õ is an Azumaya algebra (an order) over SpecR.<br />

6.10. Lemma. –––– Assume R to be complete <strong>and</strong> let D = A be a (finite) division<br />

algebra over K. Then,<br />

x ∈ D is integral over R ⇐⇒ N D/K (x) ∈ R .


Sec. 6] ORDERS IN CENTRAL SIMPLE ALGEBRAS 79<br />

Proof. K (x) is a finite field extension of K. In D, two K (x)-modules K (x)·y 1<br />

<strong>and</strong> K (x)·y 2 are either equal or disjoint. Thus, there is a decomposition<br />

rM<br />

D = K (x) · y i<br />

i=1<br />

of D into a direct sum. The multiplication map ·x : D → D is a block matrix.<br />

Consequently,<br />

N D/K (x) = [N K (x)/K (x)] r .<br />

It is, therefore, sufficient to show the assertion for N K (x)/K instead of N D/K .<br />

As K is complete, there is a unique extension of the discrete valuation from K<br />

to K (x). It is given by<br />

ν K (x) (y) :=<br />

1<br />

[K (x) : K ] ν K (N K (x)/K (y)) .<br />

The elements y of valuation ν K (x) (y) ≥ 0 form a finite R-algebra <strong>and</strong> are,<br />

therefore, integral. If ν K (x) (y) < 0 then a relation y n +a n−1 y n−1 + . . . +a 0 = 0<br />

would contradict the ultrametric triangle inequality.<br />

□<br />

6.11. Proposition. –––– Assume R to be complete. If D = A is a (finite) division<br />

algebra over K then, in D, there is exactly one maximal order over R. This is the set<br />

of all elements which are integral over R.<br />

O = {x ∈ D | N D/K (x) ∈ R}<br />

Proof. The only thing not yet proven is that O is actually an order. For this, in<br />

turn, we still have to verify that O is a ring.<br />

Recall that the valuation on D is given by<br />

ν D (x) :=<br />

1<br />

[D : K ] ν K (N D/K (x)) .<br />

The assertion would follow if we could verify the usual formulas<br />

ν D (xy) = ν D (x) + ν D (y) <strong>and</strong> ν D (x + y) ≥ min{ν D (x), ν D (y)}<br />

in the non-commutative situation, too.<br />

For this, note that det(·xy) = det((·x) ◦ (·y)) = det(·x) det(·y) implies<br />

N D/K (xy) = N D/K (x)N D/K (y) from which the first formula follows.<br />

For the ultrametric triangle inequality, we remark first that it is true in the<br />

case x <strong>and</strong> y commute. The point is, K (x, y) is then a finite field extension<br />

of K. One has ν D (·) = ν K (x,y) (·) <strong>and</strong> the inequality carries over from the<br />

commutative case.


80 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

In general, we have x −1 y ∈ O or y −1 x ∈ O. Assume without restriction<br />

that x −1 y ∈ O. Then, since 1 commutes with x −1 y,<br />

<strong>and</strong><br />

ν D (1 + x −1 y) ≥ 0 = min{ν D (1), ν D (x −1 y)}<br />

ν D (x + y) = ν D (x·(1 + x −1 y)) = ν D (x) + ν D (1 + x −1 y) ≥ ν D (x) .<br />

x −1 y ∈ O is equivalent to ν D (x) = min{ν D (x), ν D (y)}.<br />

□<br />

6.12. Remark. –––– For I ⊂ O the maximal ideal, O/I is a division algebra.<br />

O is Azumaya if <strong>and</strong> only if I = mO.<br />

6.13. Proposition. –––– Assume R to be complete. LetD be a (finite) division<br />

algebra over K <strong>and</strong> consider the Azumaya algebra A = M n (D) for a certainn ∈Æ.<br />

Then,<br />

i) ∆ := M n (O) is a maximal order in A.<br />

ii) In ∆, every left ideal is projective as a ∆-module.<br />

Proof. i) is already proven. Cf. Remark 6.8.<br />

ii) Let I ⊆ ∆ be any left ideal.<br />

We observe first that ∆/m∆ ∼ = M n (O/m) is a matrix algebra over a skew field.<br />

In particular, every ∆/m∆-module is projective as a ∆/m∆-module. Consequently,<br />

every ∆/m∆-module is of cohomological dimension ≤ 1 as a ∆-module.<br />

This is true, in particular, for I/mI.<br />

For every i ≥ 1 <strong>and</strong> every ∆-module M, we have the fragments<br />

Ext i ∆ (I, M ) ·π<br />

−→ Ext i ∆ (I, M )<br />

d<br />

−→ Ext i+1<br />

∆ (I/πI, M )<br />

of the Ext-long exact sequence. Here, as i +1 ≥ 2, the rightmost term vanishes.<br />

This means, E := Ext i ∆<br />

(I, M ) has the property that E/mE = 0.<br />

If M is a finite ∆-module then E is a finite ∆-module <strong>and</strong> hence also a finite R-<br />

module. Nakayama’s lemma implies that Ext i ∆ (I, M ) = E = 0. For the general<br />

case, we note that Ext commutes with filtered direct limits in the second argument.<br />

This yields Ext i ∆ (I, M ) = 0 for i ≥ 1, unconditionally. I is projective.<br />

□<br />

6.14. Corollary. –––– Let R be a complete discrete valuation ring <strong>and</strong> A be an<br />

arbitrary Azumaya algebra over K = Q(R). Then, there exists a maximal order<br />

∆ ⊂ A over R such that every left ideal in ∆ is projective.<br />

Proof. There exists a division algebra D such that A = M n (D).<br />


Sec. 6] ORDERS IN CENTRAL SIMPLE ALGEBRAS 81<br />

6.15. Remark. –––– We want to abolish the completeness assumption. It will<br />

turn out that Corollary 6.14 is true for arbitrary discrete valuation rings.<br />

6.16. Example. –––– It is, however, not true that every division algebra over a<br />

non-complete discrete valuation ring has a unique maximal order.<br />

Consider, for example, the quaternion algebraÀ:=É·1 ⊕É·i ⊕É·j ⊕É·k<br />

with i 2 = j 2 = k 2 = −1 <strong>and</strong> ij = −ji = k, etc. This is a division algebra<br />

over K =É. InÉ, consider the discrete valuation ring R =(5) corresponding<br />

to the 5-adic valuation. This leads to ̂É=É5.<br />

We have thatÀ⊗ÉÉ5 ∼ = M 2 (É5) splits since (−1) is a square inÉ5. Lemma 6.18<br />

will show that there is a one-to-one correspondence between maximal orders<br />

inÀover(5) <strong>and</strong> maximal orders in M 2 (É5) over5. In particular, there are<br />

several maximal orders.<br />

6.17. Sublemma. –––– Let V be a finite-dimensional K-vector space <strong>and</strong> put<br />

̂V := V ⊗ K ̂K.<br />

Then, there is a bijection<br />

{R-lattices of full rank in V } −→ { ̂R-lattices of full rank in ̂V } ,<br />

Γ ↦→ i<br />

Γ ⊗ R ̂R ,<br />

Proof. We put n := dim V .<br />

Γ ∩ V ι ← Γ .<br />

“ι◦i = id” This follows from the fact that ̂R is faithfully flat over R.<br />

“i◦ι = id” Choose a basis for V .<br />

Then, an ̂R-lattice of full rank in ̂V is given by n linearly independent vectors<br />

in ̂K n , i.e. by a matrix M ∈ GL n (̂K ).<br />

The lattice is not altered by column operations. I.e., by the interchange of two<br />

columns or by adding λ times a column to another column for λ ∈ ̂R. It is not<br />

changed either by the multiplication of a column with a unit of ̂R.<br />

The goal is to verify that these operations allow to transform M into a matrix<br />

in GL n (K ).<br />

Column operations alone may bring M into upper triangular form. Multiplying<br />

the columns by units, we obtain a matrix as follows,<br />

⎛<br />

⎞<br />

π a 1<br />

0 0 . . . 0<br />

c 21 π a 2<br />

0 . . . 0<br />

c 31 c 32 π a 3<br />

. . . 0<br />

.<br />

⎜<br />

⎝<br />

.<br />

. . . ..<br />

⎟<br />

. ⎠<br />

c n1 c n2 c n3 . . . π a n


82 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

The entries under the diagonal are not yet in K. We transform them inductively,<br />

row-by-row. For the first row, there is nothing left to do. So assume, the<br />

entries in rows 1, . . . , k − 1 are already in K. Write c kj = u kj π e j<br />

with a<br />

unit u kj . The operations allowed may alter c kj by any element from ̂Rπ a k .<br />

Since ̂Rπ e j /̂Rπ a k ∼ = Rπ<br />

e j/Rπ a k, the assertion follows.<br />

□<br />

6.18. Lemma. –––– There is a bijection<br />

{ (Maximal) orders in A } −→ { (Maximal) orders in A ⊗ K ̂K } ,<br />

∆ ↦→ ∆ ⊗ R ̂R .<br />

Proof. In our situation, an order is the same as a lattice of full rank which<br />

inherits a ring structure from the central simple algebra. It is, therefore, to<br />

be shown that the operations described in the sublemma above preserve such<br />

ring structures.<br />

This is obvious for ∆ ↦→ ∆ ⊗ R ̂R.<br />

For the other direction, the map is given by Λ ↦→ Λ ∩ A. We note that<br />

to have a ring structure means to contain the unit element <strong>and</strong> closedness<br />

under multiplication. The intersection clearly preserves the unit element.<br />

Closedness under multiplication carries over to ∆ := Λ ∩A, too. Indeed, by the<br />

sublemma, Λ = ∆⊗ R ̂R. If we had ∆ ⊗ R ∆ ։ I <strong>and</strong> I/∆ ≠ 0 then this would<br />

remain true after the faithful flat base change from R to ̂R.<br />

□<br />

6.19. Proposition. –––– LetR be any discretevaluation ring <strong>and</strong> A bean arbitrary<br />

Azumaya algebra over K = Q(R). Then, there exists a maximal order ∆ ⊂ A<br />

over R such that every left ideal in ∆ is projective.<br />

Proof. If R is complete then this is Corollary 6.14.<br />

In general, we have a maximal order Λ ⊂ A⊗ R ̂R over ̂R such that every left<br />

ideal in Λ is projective. Lemma 6.18 provides us with a maximal order ∆ ⊂ A<br />

such that ∆⊗ R ̂R = Λ. We claim, in ∆, every left ideal I is projective.<br />

Indeed, I⊗ R̂R is an ideal in ∆⊗ R̂R <strong>and</strong> therefore projective. Projectivity descends<br />

under this faithful flat map as always<br />

Ext i R (M,N)⊗ R ̂R = Ext îR (M ⊗ R ̂R, N ⊗ R ̂R) .<br />

□<br />

6.20. Lemma. –––– In ∆, every left ideal is principal.<br />

Proof. Let I ⊆ ∆ be any left ideal. Then, KI is a left ideal in A.<br />

Hence, KI = Ae = K ∆e for some idempotent e. In particular, I <strong>and</strong> ∆e<br />

are free R-modules of the same rank.


Sec. 6] ORDERS IN CENTRAL SIMPLE ALGEBRAS 83<br />

From this, it follows that the ideals I/mI <strong>and</strong> (∆e)/m(∆e) in the central simple<br />

algebra ∆/m∆ are isomorphic.<br />

As I <strong>and</strong> ∆e are both projective, we may lift a pair of mutually inverse isomorphisms<br />

to homomorphisms f : I → ∆e <strong>and</strong> g : ∆e → I. f ◦g <strong>and</strong> g ◦ f are<br />

automorphisms by virtue of Nakayama’s lemma. The assertion follows. □<br />

6.21. Proposition. –––– Let R be a discrete valuation ring, K = Q(R), <strong>and</strong> A be<br />

an Azumaya algebra over K.<br />

Then, all maximal orders in A over R are conjugate to each other.<br />

Proof. Let ∆ <strong>and</strong> ∆ ′ be two such orders. We may assume that, in ∆, every left<br />

ideal is principal. Put<br />

F := {x ∈ A | x∆ ′ ⊆ ∆} .<br />

This is a left ∆-module in A, hence a broken ideal. It follows that F = ∆t for a<br />

suitable t ∈ A.<br />

∆ <strong>and</strong> ∆ ′ are orders. Hence, they are full lattices. This implies, there exists<br />

r ∈ R\{0} such that r∆ ′ ⊆ ∆. We have r ∈ F. Consequently, t ∈ A is an<br />

invertible element.<br />

F is, by definition, a right ∆ ′ -module. This shows<br />

t∆ ′ ⊆ ∆t ,<br />

hence ∆ ′ t −1 ⊆ ∆t. Maximality yields the assertion.<br />

□<br />

6.22. Corollary. –––– Let R be a discrete valuation ring, K = Q(R), A an<br />

Azumaya algebra over K, <strong>and</strong> Λ ⊂ A be a maximal order over R.<br />

Then, in Λ, every left ideal is principal.<br />

Proof. Combine Lemma 6.20 with Proposition 6.21.<br />

□<br />

6.23. Remark. –––– There are analogous results for right ideals.<br />

6.24. Remarks. –––– i) If A allows an extension A which is an Azumaya algebra<br />

over SpecR then A is necessarily a maximal order (Proposition 6.7).<br />

Proposition 6.21 shows that every maximal order is Azumaya in this case.<br />

ii) If A does not allow an extension as an Azumaya algebra then the same is true<br />

for M n (A). Indeed, for Λ a maximal order in A, we have that Λ⊗Λ op → End(Λ)<br />

is not surjective. This remains true when going over from Λ to the full matrix<br />

algebra M n (Λ).


84 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

7. The Theorem of Ausl<strong>and</strong>er <strong>and</strong> Goldman<br />

7.1. Theorem (M. Ausl<strong>and</strong>er <strong>and</strong> O. Goldman). —–<br />

Let X be an integral scheme which is regular, separated, <strong>and</strong> quasi-compact. Assume<br />

that the dimension of X is at most two.<br />

i) Then, one has<br />

Br(X ) = \<br />

codimx=1<br />

where the intersection takes place in Br(Q(X )).<br />

ii) In particular, Br(X ) = Br ′ (X ).<br />

Br(Spec O X,x )<br />

Proof. Recall that, by Corollary 5.5, Br(Spec O X,x ) = Br ′ (Spec O X,x ). Further,<br />

Br(Q(X )) = Br ′ (Q(X )).<br />

By virtue of the general Fact 4.2, we have that Br(X ) ⊆ Br(Q(X ))<br />

<strong>and</strong> Br(Spec O X,x ) ⊆ Br(Q(X )). Thus, the natural maps Br(X ) → Br(Spec O X,x )<br />

are all injective. This shows the “⊆”-part of i).<br />

The same argument is true for the cohomological Brauer group. We obtain that<br />

Br ′ (X ) ⊆ \<br />

Br(Spec O X,x ) . (‡)<br />

codimx=1<br />

This, in turn, makes sure that the second assertion immediately follows<br />

from the first. Indeed, the inclusion (‡) together with i) immediately yields<br />

Br ′ (X ) ⊆ Br(X ). The inclusion the other way round was established in Fact 4.2.<br />

The main part is to prove “⊇”. For that, let α ∈ Br ′ (Q(X )) such that<br />

α ∈ \<br />

Br(Spec O X,x ) .<br />

codim x=1<br />

We choose an Azumaya algebra A over Q(X ) corresponding to the class α.<br />

By Lemma 6.4, there exists a maximal O X -order A in A. We claim, A is a<br />

locally free sheaf.<br />

This may be tested locally. Let x ∈ X be any point on X. We write R := O X,x<br />

<strong>and</strong> denote the maximal ideal in R by m. The maximality property of A implies<br />

that A is equal to its bidual. In particular, A x = A ∨∨<br />

x .<br />

We consider the short exact sequence<br />

0 −→ M −→ F 0 −→ A ∨<br />

x −→ 0<br />

of O X,x -modules with F 0 free. Dualizing yields<br />

0 A ∨∨<br />

x<br />

F ∨ 0<br />

M ∨ . . . .<br />

A


Sec. 7] THE THEOREM OF AUSLANDER AND GOLDMAN 85<br />

M ∨ := Hom R (M,R) is a torsion-free R-module. Thus, M ∨ → M ∨ ⊗ R K<br />

is injective. We see, M ∨ may be embedded into a K-vector space. It, therefore,<br />

allows an embedding into a free R-module F 1 , too,<br />

0 −→ A x −→ F ∨ 0 −→ F 1 −→ Q −→ 0 .<br />

R is of cohomological dimension ≤ 2 [Mat, Theorem 19.2]. To establish<br />

that A x is free, it is sufficient to verify that Tor R i (A x , R/m) = 0 for i = 1<br />

<strong>and</strong> i = 2 [Mat, §19, Lemma 1].<br />

For this, writing Q 1 := ker(F 1 → Q), we see<br />

Tor R 2 (A x, R/m) = Tor R 3 (Q 1, R/m) = 0<br />

<strong>and</strong><br />

Tor R 1 (A x, R/m) = Tor R 2 (Q 1, R/m) = Tor R 3 (Q, R/m) = 0 .<br />

A is a locally free sheaf.<br />

Corollary 1.5 shows that the non-Azumaya locus of A is a closed subset T ⊆ X,<br />

pure of codimension one. We have to show that T = ∅.<br />

Assume the contrary. Then, there is an irreducible divisor D = {x} ⊆ T for x<br />

a codimension one point. By assumption, A allows an extension to Spec O X,x<br />

as an Azumaya algebra B. Furthermore, we may consider A | SpecOX,x . This is a<br />

maximal order in A over Spec O X,x . As maximal orders, B <strong>and</strong> A | SpecOX,x are<br />

conjugate by Proposition 6.21. In particular, A | SpecOX,x is an Azumaya algebra.<br />

Thus, the assumption D = {x} ⊆ T leads to a contradiction.<br />

7.2. Remark. –––– Using more refined methods, it was shown that the two<br />

Brauer <strong>groups</strong> coincide in several more cases.<br />

The most general result in this direction is due to O. Gabber ([Ga, Hoo] or<br />

[K/O80]). It shows that Br(X ) = Br ′ (X ) tors if X = U ∪ V is the union of two<br />

affine schemes such that U ∩ V is again affine.<br />

Quite recently, S. Schröer [Schr] showed Br(X ) = Br ′ (X ) tors for any quasicompact,<br />

separated, geometrically normal algebraic surface.<br />

If one assumes X to be regular <strong>and</strong> separated then the subscript may be<br />

=/2<br />

omitted<br />

in view of Fact 4.2. A. Grothendieck knew that Br ′ (X ) may be non-torsion,<br />

even for an affine normal surface over, in 1965/66, already [GrBrII, Remarque<br />

1.11.b].<br />

D. Edidin, B. Hassett, A. Kresch, <strong>and</strong> A. Vistoli [E/H/K/V] (see also [Bt])<br />

constructed an example of a non-separated surface such that Br ′ (X )<br />

but Br(X ) = 0. X is the union of two affine schemes, regular in codimension<br />

one, but not normal.<br />


86 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

8. Examples<br />

i. Some rings <strong>and</strong> fields from number theory. —<br />

8.1. Examples. –––– Let p be a prime <strong>and</strong> K be a finite field extension ofÉp.<br />

i) Then, Br(SpecK ) =É/.<br />

This is shown by local class field theory [Se67, Sec. 1, Theorems 1 <strong>and</strong> 2].<br />

Classically, the isomorphism is given as follows. It turns out that<br />

Br(SpecK ) = H 2( Gal(K nr /K ), (K nr ) ∗)<br />

for K nr the maximal unramified extension. The valuation ν : K nr →induces<br />

H 2( Gal(K nr /K ), (K nr ) ∗) ∼ =<br />

−→ H 2( Gal(K nr /K ),) = H 2 (̂,)<br />

= Hom(̂,É/) =É/.<br />

ii) For O K , the ring of integers, one has Br(Spec O K ) = 0.<br />

Indeed, by the description of Br(SpecK ), the canonical homomorphism<br />

Br ′ (SpecK ) −→ Hom ( Gal(p/p),É/)<br />

from Proposition 4.8 is bijective.<br />

8.2. Examples. –––– i) One has Br(SpecÊ) = 1 2/.<br />

ii) One has Br(Spec) = 0.<br />

8.3. Examples. –––– Let K be a number field.<br />

i) Then,<br />

Br(SpecK ) = ker ( M<br />

s :<br />

pprimeÉ/⊕ M<br />

2/−→É/) .<br />

σ : K→Ê1<br />

Here, s is just the summation.<br />

→É/<br />

This is shown by global class field theory [Ta67, 11.2]. The isomorphism is<br />

induced by the canonical maps<br />

i ν : Br(SpecK ) → Br(SpecK ν )<br />

for ν ∈ Val(K ) composed with the invariant map inv ν : Br(SpecK ν )<br />

(or to 1 2/or 0, respectively) from Example 8.1.i).


Sec. 8] EXAMPLES 87<br />

ii) Denote by r the number of real embeddings of K. Then, for the integer<br />

ring O K , one has<br />

{ (<br />

1<br />

Br(Spec O K ) = 2/) r−1 if r ≥ 1 ,<br />

0 otherwise .<br />

In particular, Br(Spec) = 0.<br />

Here, the canonical homomorphism<br />

Br ′ (SpecK ) −→<br />

M<br />

codimx=1<br />

Hom(G x ,É/)<br />

from Proposition 4.8 is nothing but the projection<br />

ker ( M<br />

M<br />

s :<br />

2/−→É/) −→<br />

ii. Geometric examples. —<br />

pprimeÉ/⊕ M<br />

σ : K→Ê1<br />

p primeÉ/.<br />

8.4. Lemma. –––– Let n ∈Æ<strong>and</strong> X be any scheme such that n is invertible on X.<br />

Then, there is the short exact sequence<br />

0 −→ Pic(X )/n Pic(X )<br />

c 1<br />

−→ H<br />

2<br />

ét (X, µ n) −→ Br ′ (X ) n −→ 0 .<br />

Proof. This follows from the long exact sequence in étale cohomology associated<br />

to the Kummer sequence<br />

0 −→ µ n −→m<br />

n<br />

−→m −→ 0 .<br />

□<br />

8.5. Proposition. –––– Let k ⊆be an algebraically closed field <strong>and</strong> X be a<br />

scheme which is regular <strong>and</strong> proper over k.<br />

Then, Br ′ (X ) = (É/) rkH 2 (X (),)−rkNS(X )<br />

⊕ H 3 (X (),) tors .<br />

Proof. Br ′ (X ) is torsion in view of Fact 4.2.<br />

For the middle term in the exact sequence above, by [SGA4, Exp. XVI, Corollaire<br />

1.6], we have H 2 ét (X, µ n) ∼ = H 2 ét (X, µ n ). Further, the comparison theorem<br />

[SGA4, Exp. XI, Théorème 4.4] shows H 2 ét (X, µ n ) ∼ = H 2 (X (),/n).<br />

On the other h<strong>and</strong>, by the Theorem of Murre <strong>and</strong> Oort [SGA6, Exp. XII,<br />

Corollaires 1.2 et 1.5.a)], the Picard functor is representable by a scheme Pic X/k .<br />

This is a group scheme, hence smooth over k. It is the disjoint union of quasiprojective<br />

k-schemes. Among other properties, one has Pic X/k (k) = Pic(X )<br />

<strong>and</strong> Pic X/k () = Pic(X).<br />

The connected components of Pic X/k are in bijection with NS(X ).<br />

Base change from k todoes not change these components. In particular,<br />

NS(X ) = NS(X).


88 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

Corresponding to this, Pic(X ) is an extension of the Néron-Severi group NS(X )<br />

which is finitely generated by a divisible group,<br />

0 −→ Pic 0 (X ) −→ Pic(X ) −→ NS(X ) −→ 0 .<br />

This shows Pic(X )/n Pic(X ) ∼ = NS(X )/nNS(X ) ∼ = NS(X)/nNS(X).<br />

By consequence, we may assume from now on that k =.<br />

Doing this, we first observe that the homomorphism<br />

Pic(X )/n Pic(X ) −→ H 2 ét (X, µ n)<br />

induced by the Kummer sequence is compatible with the usual first Chern class<br />

homomorphism coming from the exponential sequence. Indeed, there is the<br />

commutative diagram<br />

0<br />

2πi O X<br />

exp O ∗ X<br />

0<br />

exp( 2πi<br />

n ·)<br />

0 µ n O ∗ X<br />

exp( 1 n ·)<br />

(·) n O ∗ X<br />

of exact sequences of sheaves which induces the commutative diagram<br />

Pic(X )<br />

Pic(X )<br />

=<br />

0<br />

= H 1 (X, OX ∗ ) H 2 (X,)<br />

=<br />

= H 1 (X, O ∗ X ) H 2 (X, µ n ) .<br />

By the Lefschetz theorem on (1, 1)-classes [G/H, p. 163], the cokernel of the<br />

first Chern class homomorphism<br />

c 1 : Pic(X ) −→ H 2 (X,)<br />

is always torsion-free. Writing r := rkH 2 (X (),) <strong>and</strong> ρ := rkNS(X ), we<br />

therefore have coker c 1<br />

∼ =r−ρ . Thus, the cokernel of the induced homomorphism<br />

c 1 ⊗/n: Pic(X )/n Pic(X ) → H (X,)⊗/n<br />

2<br />

is isomorphic to (/n) r−ρ .<br />

Further, the universal coefficient theorem [Sp, Chap. 5, Sec. 5, Theorem 10]<br />

yields the short exact sequence<br />

0 → H 2 (X (),) ⊗/n−→ H 2 (X (),/n) −→ H 3 (X (),) n → 0 .<br />

Thus, from Lemma 8.4 we see that there is a surjective canonical homomorphism<br />

Br ′ (X ) n → H 3 (X (),) n the kernel of which is isomorphic to<br />

coker(c 1 ⊗/n) ∼ = (/n) r−ρ .


Sec. 8] EXAMPLES 89<br />

Altogether, there is a short exact sequence<br />

0 −→ H −→ Br ′ (X ) −→ H 3 (X (),) tors −→ 0<br />

where #H n = n r−ρ for each n ∈Æ. A simple application of the structure<br />

theorem for finite abelian <strong>groups</strong> shows that (É/) r−ρ is the only torsion<br />

group having this property.<br />

Finally, there are no non-trivial extensions of a finite torsion group by (É/) r−ρ .<br />

□<br />

8.6. Remark. –––– A generalization to an arbitrary algebraically closed field k<br />

of characteristic zero is not difficult. Indeed, to describe X only finitely many<br />

data are needed. Thus, X ∼ = X 0 × Speck0 Speck for a certain scheme X 0 over a<br />

field k 0 which is finitely generated overÉ. k 0 allows an embedding into.<br />

8.7. Examples. –––– Let k ⊆be an algebraically closed field.<br />

i) Let X be a smooth proper curve over k. Then, Br(X ) = 0.<br />

(Actually, in this case, one has even Br(Q(X )) = 0 by the Theorem of Tsen<br />

[SGA4 1 , Arcata, III, Théorème (2.3)].)<br />

2<br />

ii) Let X be a <strong>rational</strong> surface over k (proper, not necessarily minimal).<br />

Then, Br(X ) = 0.<br />

iii) Let X be a K3 surface over k. Then, Br(X ) = (É/) 22−rkPic(X ) .<br />

iv) Let X be an Enriques surface over k. Then, Br(X ) =/2.<br />

8.8. Fact. –––– Let k ⊆be an algebraically closed field <strong>and</strong> X be a smooth<br />

complete intersection of dimension ≥ 3 in P n k .<br />

Then, Br ′ (X ) = 0.<br />

Proof. The Lefschetz hyperplane theorem [Bot, Corollary of Theorem 1]<br />

implies that H 1 (X (),) = 0 <strong>and</strong> H 2 (X (),) =. From the universalcoefficient<br />

theorem for cohomology [Sp, Chap. 5, Sec. 5, Theorem 3], we<br />

deduce H 2 (X (),) =<strong>and</strong><br />

H 3 (X (),) ∼ = Hom(H 3 (X (),),) .<br />

In particular, H 3 (X (),) is torsion-free.<br />

□<br />

iii. Varieties over a number field or local field. —<br />

8.9. –––– In this subsection, we deal with the case that X is a scheme over an<br />

algebraically non-closed field K. The relevant cases for us are that K is either a<br />

number field or a local field.


90 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

It is not true that the homomorphism Br(SpecK ) → Br(X ), induced by the<br />

structural map, is always injective. For this, an additional hypothesis is needed.<br />

8.10. Proposition. –––– Let K be a field <strong>and</strong> π : X → SpecK be any scheme<br />

over K. Assume that<br />

a) either K is a local field <strong>and</strong> X (K ) ≠ ∅<br />

b) or K is a number field <strong>and</strong> ∏ X (K ν ) ≠ ∅.<br />

ν∈Val(K )<br />

Then, the canonical map π ∗ : Br(SpecK ) → Br(X ) is injective.<br />

Proof. a) A K-valued point x : SpecK → X induces a section of π ∗ .<br />

b) Put S := F ν SpecK ν . The assumption says that X has an S-valued point.<br />

Therefore, the natural homomorphism<br />

i : Br(K ) → Br(S) = ∏ Br(K ν )<br />

ν<br />

factors via π ∗ . i is injective by Example 8.3.i).<br />

8.11. Proposition. –––– Let K be a number field or a local field <strong>and</strong><br />

π : X → SpecK be a geometrically integral scheme which is proper over K.<br />

i) Then, there is an exact sequence<br />

0 −→ Pic(X )<br />

i<br />

−→ Pic(X K<br />

) Gal(K/K ) π<br />

−→ Br(SpecK ) −→<br />

∗<br />

π ∗<br />

−→ ker(Br ′ (X ) i′<br />

→ Br ′ (X K<br />

)) −→ H 1( Gal(K/K ), Pic(X K<br />

) ) −→ 0 .<br />

Here, i <strong>and</strong> i ′ are induced by the natural morphism X K<br />

→ X.<br />

ii) If π ∗ : Br(SpecK ) → Br ′ (X ) is injective then there is the short exact sequence<br />

0 → Br(SpecK) → π∗<br />

ker(Br ′ (X ) → i′<br />

Br ′ (X K<br />

)) → H 1( Gal(K/K ), Pic(X K<br />

) ) → 0 .<br />

Further, Pic(X ) ∼ = Pic(X K<br />

) Gal(K/K ) .<br />

Proof. Consider the Hochschild-Serre spectral sequence<br />

in étale cohomology.<br />

E p,q<br />

2<br />

:= H p( Gal(K/K ), H q ét (X K ,m) ) =⇒ H n ét (X,m)<br />

As X is integral <strong>and</strong> proper over K, we have that H 0 ét (X K ,m) = K ∗ . Further,<br />

H 1 ét (X K ,m) = Pic(X K<br />

) <strong>and</strong> H 2 ét (X K ,m) = Br ′ (X K<br />

).<br />

This yields E 0,1<br />

2<br />

= Pic(X K<br />

) Gal(K/K ) <strong>and</strong> E 2,0<br />

2<br />

= Br(SpecK ). Furthermore,<br />

E 1,1<br />

2<br />

= H 1( Gal(K/K ), Pic(X K<br />

) ) <strong>and</strong> E 0,2<br />

2<br />

= Br ′ (X K<br />

) Gal(K/K ) .<br />

The classical Theorem Hilbert 90 shows that E 1,0<br />

2<br />

= H 1( Gal(K/K ), K ∗) = 0.<br />


Sec. 8] EXAMPLES 91<br />

In addition, if K is a number field then one has E 3,0<br />

2<br />

= H 3( Gal(K/K ), K ∗) = 0<br />

by [Ta67, Section 11.4]. Otherwise, if K is a local field then, by class field theory<br />

H 3( Gal(L/K ), L ∗) ∼ = H<br />

1 ( Gal(L/K ),) = 0 for every finite extension of K.<br />

Thus, H 3( Gal(K/K ), K ∗) = 0.<br />

Finally, E 1 = H 1 ét (X,m) = Pic(X ) <strong>and</strong> E 2 = H 2 ét (X,m) = Br ′ (X ).<br />

The sequence in i) is nothing but a sequence of lower order terms in the spectral<br />

sequence E.<br />

ii) directly follows from i).<br />

8.12. Corollary. –––– Let K be a number field or local field <strong>and</strong> π : X → SpecK<br />

be a geometrically integral scheme which is proper over K.<br />

i) Then, Br(X ) is a countable group.<br />

ii) Suppose that rk NS(X ) = rkH 2 (X (),) <strong>and</strong> rkH 1 (X (),) = 0. Then,<br />

Br(X )/π ∗ Br(SpecK ) is finite.<br />

Proof. i) Br(X K<br />

) is a countable group by Proposition 8.5. It is therefore sufficient<br />

to verify that ker(Br(X ) → Br(X K<br />

)) is a countable group. We will even show<br />

that ker(Br ′ (X ) → Br ′ (X K<br />

)) is countable.<br />

It is known from Example 8.3.i) that Br(SpecK ) is countable. We are left with<br />

proving H 1( Gal(K/K ), Pic(X K<br />

) ) is a countable group.<br />

There is a short exact sequence of Gal(K/K )-modules<br />

0 −→ Pic 0 (X K<br />

) −→ Pic(X K<br />

) −→ NS(X K<br />

) −→ 0 .<br />

Here, H 1( Gal(K/K ), NS(X K<br />

) ) is finite since NS(X K<br />

) is a finitely generated<br />

abelian group acted upon by a finite quotient of Gal(K/K ).<br />

On the other h<strong>and</strong>, Pic 0 (X K<br />

) is divisible <strong>and</strong> all the <strong>groups</strong> Pic 0 (X K<br />

) n are finite.<br />

The Kummer sequence<br />

0 −→ Pic 0 (X K<br />

) n −→ Pic(X K<br />

)<br />

n<br />

−→ Pic(X K<br />

) −→ 0<br />

induces a surjection H 1( Gal(K/K ), Pic 0 (X K<br />

) n<br />

) → H<br />

1 ( Gal(K/K ), Pic 0 (X K<br />

) ) n .<br />

In particular, H 1( Gal(K/K ), Pic 0 (X K<br />

) ) n is finite <strong>and</strong> H 1( Gal(K/K ), Pic 0 (X K<br />

) )<br />

is countable.<br />

ii) Here, the assumption rk NS(X ) = rkH 2 (X (),) makes sure that Br(X K<br />

)<br />

is finite. It is therefore sufficient to verify that<br />

is finite. Again, we will even show that<br />

ker(Br(X ) → Br(X K<br />

))/π ∗ Br(SpecK )<br />

ker(Br ′ (X ) → Br ′ (X K<br />

))/π ∗ Br(SpecK ) ∼ = H 1( Gal(K/K ), Pic(X K<br />

) )<br />


92 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

is finite.<br />

For this, the assumption rkH 1 (X (),) = 0 implies that Pic(X K<br />

) = NS(X K<br />

).<br />

The assertion follows.<br />

□<br />

8.13. Remark. –––– The case that Br ′ (X K<br />

) = 0 is of particular interest. For<br />

example, this happens if X K<br />

is a <strong>rational</strong> surface, a K3 surface, or a smooth<br />

complete intersection of dimension ≥ 3 in P n (cf. Subsection ii).<br />

K<br />

In this case, the short exact sequence in assertion ii) simply means<br />

Br ′ (X )/π ∗ Br(SpecK) ∼ = H 1( Gal(K/K ), Pic(X K<br />

) ) .<br />

8.14. Fact. –––– LetK beanumber field or alocalfield <strong>and</strong>X beasmoothcomplete<br />

intersection of dimension ≥ 3 in P n K .<br />

i) Then, Br(X ) = Br ′ (X ).<br />

ii) The homomorphism π ∗ : Br(SpecK ) → Br(X ) induced by the structural map<br />

is surjective.<br />

iii) If X (K ) ≠ ∅ for K a local field or ∏ X (K ν ) ≠ ∅ for K a number field<br />

ν∈Val(K )<br />

then Br(X ) = Br(SpecK ).<br />

Proof. The Lefschetz hyperplane theorem implies that Pic(X K<br />

) =. Therefore,<br />

H 1( Gal(K/K ), Pic(X K<br />

) ) = H 1( Gal(K/K ),) = 0 .<br />

Since Br ′ (X K<br />

) = 0, Proposition 8.11 shows that π ∗ : Br(SpecK ) → Br ′ (X )<br />

is surjective.<br />

In particular, π ∗ does not factorize via any proper subgroup of Br ′ (X ).<br />

Hence, Br(X ) = Br ′ (X ). This proves i) <strong>and</strong> ii).<br />

iii) follows now from Proposition 8.10.<br />

8.15. –––– In the situation of a curve over a local field, Br(X )/π ∗ Br(SpecK )<br />

has been computed by S. Lichtenbaum. The following result is often referred<br />

to as Lichtenbaum duality.<br />

8.16. Theorem (Lichtenbaum). —– Let p beaprimenumber,K afiniteextension<br />

ofÉp, <strong>and</strong> X be a geometrically integral curve which is proper <strong>and</strong> smooth over K.<br />

Then, the canonical pairing<br />

Br(X ) × X (K ) −→É/,<br />

(α, x) ↦→ inv νp (α| x )<br />

□<br />

induces an isomorphism Br(X )/π ∗ ∼=<br />

Br(SpecK ) −→ Hom(Pic 0 (X ),É/).<br />

Proof. This is [Lic, Corollary 1].<br />


Sec. 8] EXAMPLES 93<br />

iv. Manin’s formula. —<br />

8.17. –––– If X K<br />

is a <strong>rational</strong> surface then H 1( Gal(K/K ), Pic(X K<br />

) ) is often<br />

effectively computable using the following result due to Yu. I. Manin.<br />

8.18. Proposition (Manin, [Man,кIV,ÈÖÐÓÒº¿]). —–<br />

Let X be a regular, integral scheme which is of finite type over a field k. Suppose<br />

further, we are given a finite, G := Gal(k sep /k)-invariant set {D i } of divisors<br />

on X := X × Speck Speck sep generating Pic(X ).<br />

Denote by S ⊆ Div(X ) the group generated by {D i } <strong>and</strong> by S 0 ⊆ S the subgroup<br />

of all principal divisors. Finally, let H ⊆ G be a normal subgroup acting trivially<br />

on {D i }.<br />

Assume,<br />

i) Pic(X ) is a free abelian group <strong>and</strong><br />

ii) there is a perfect pairing<br />

Then, there is a canonical isomorphism<br />

H 1 (G, Pic(X ))<br />

Pic(X ) × Pic(X ) −→.<br />

i<br />

−→ Hom((NS ∩ S 0 )/NS 0 ,É/) .<br />

Here, N : S → S denotes the norm map on S as a G/H -module.<br />

Proof. By Lemma 8.19, we have<br />

Ĥ −1 (G/H , Pic(X )) ∼ = (NS ∩ S 0 )/NS 0 .<br />

To this relation, we apply the duality theorem [C/E, Chap. XII, Corollary 6.5].<br />

It shows that<br />

Ĥ 0 (G/H , Hom(Pic(X ),É/)) ∼ = Hom(Ĥ −1 (G/H , Pic(X )),É/)<br />

It remains to construct a canonical isomorphism<br />

= Hom((NS ∩ S 0 )/NS 0 ,É/) .<br />

Ĥ 0 (G/H , Hom(Pic(X ),É/)) ∼ = H 1 (G, Pic(X )) .<br />

For this, since Pic(X ) is free, we have a short exact sequence of G/H -modules<br />

0 −→ Hom(Pic(X ),) −→ Hom(Pic(X ),É) −→ Hom(Pic(X ),É/) −→ 0 .<br />

As Hom(Pic(X ),É) is uniquely divisible, this shows<br />

Ĥ 0 (G/H , Hom(Pic(X ),É/)) ∼ = H 1 (G/H , Hom(Pic(X ),)) .


94 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

We finally apply the perfect pairing on Pic(X ).<br />

□<br />

8.19. Lemma. –––– Let G be a finite group <strong>and</strong> D be a finite G-set. Consider the<br />

free abelian group S over D as a G-module. Then, for every G-submodule S 0 ⊆ S,<br />

there is a canonical isomorphism<br />

Ĥ −1 ∼=<br />

(G, P) −→ (NS ∩<br />

=<br />

S 0 )/NS 0 .<br />

Proof. We first calculate Ĥ −1 (G, S) directly, according to the definition. Observe<br />

that S is a direct sum of G-modules of the form[G/H ] for normal<br />

sub<strong>groups</strong> H ⊆ G. For these, one has<br />

H 0 (G,[G/H ]) ∼<br />

<strong>and</strong> the isomorphism is given by the map[G/H ] →sending a formal<br />

sum to the sum of all coefficients. Further, the homomorphism→[G/H ]<br />

mapping 1 to ∑ g∈G/H g has no kernel. Thus, Ĥ −1 (G, S) = 0.<br />

The short exact sequence<br />

0 −→ S 0 −→ S −→ P −→ 0<br />

of G-modules therefore induces a long exact sequence<br />

0 −→ Ĥ −1 (G, P) −→ Ĥ 0 (G, S 0 ) −→ Ĥ 0 (G, S) .<br />

on Tate cohomology. We may consequently write<br />

Ĥ −1 (G, P) = ker(S G 0 /NS 0 → S G /NS) ∼ = (NS ∩ S 0 )/NS 0 .<br />

□<br />

8.20. Remark. –––– The assumptions are fulfilled if X is a proper surface such<br />

that H 1 (X (),) = 0 <strong>and</strong> H 2 (X, O X ) = 0. For example, X may be a (not<br />

necessarily minimal) <strong>rational</strong> surface or a K3 surface.<br />

The first condition implies H 1 (X (),) = 0 <strong>and</strong> H 2 (X (),) is torsion-free.<br />

In particular, the first Chern class<br />

c 1 : Pic(X) → H 2 (X (),)<br />

is injective. The second condition makes sure it is a surjection. As H 2 (X (),)<br />

is torsion-free, Poincaré duality yields that the intersection pairing is perfect.<br />

8.21. –––– Let us consider the case that X is a smooth cubic surface in a bit<br />

more detail.


Sec. 8] EXAMPLES 95<br />

If is well known that, on a cubic surface X over an algebraically closed field,<br />

there are exactly 27 lines D 1 , . . . , D 27 . Further, Pic(X ) ∼ =7 is generated by<br />

the classes of these lines.<br />

The set L := {D 1 , . . . , D 27 } is equipped with the intersection product<br />

〈 , 〉: L ×L → {−1, 0, 1}. The pair (L , 〈 , 〉) is the same for all smooth<br />

cubic surfaces. It is well known [Man,кIV,ÌÓÖѽººµ] that the group<br />

of permutations of L respecting 〈 , 〉 is isomorphic to W (E 6 ). We fix such<br />

an isomorphism.<br />

Further, we put S ∼ =27 to be the free group over L <strong>and</strong><br />

{ ∣ }<br />

∣∣<br />

S 0 := ∑ a i D i ∑a i 〈D i , D j 〉 = 0 for j = 1, . . . , 27 .<br />

i i<br />

Let X be a smooth cubic surface over a number field K. Then, Gal(K/K )<br />

operates canonically on the set L X of the 27 lines on X K<br />

. Fix a bijection<br />

∼ =<br />

i X : L X −→ L respecting the intersection pairing. This induces a group homomorphism<br />

ι X : Gal(K/K ) → W (E 6 ). We denote its image by G ⊂ W (E 6 ).<br />

As L X is clearly Gal(K/K )-invariant, we may apply Proposition 8.18. It shows<br />

H 1( Gal(K/K ), Pic(X K<br />

) ) = H 1 (G, S/S 0 ) = Hom ( (NS ∩ S 0 )/NS 0 ,É/) .<br />

The right h<strong>and</strong> side depends only on the conjugacy class of the subgroup<br />

G ⊆ W (E 6 ). Actually, it depends only on the decomposition of L<br />

into G-orbits.<br />

8.22. Remark. –––– Almost as a byproduct, we may write down a similar<br />

formula for another arithmetic invariant of the cubic surface X, namely for the<br />

Picard rank. Indeed, Proposition 8.11.i) shows<br />

rkPic(X ) = rkH 0( Gal(K/K ), Pic(X K<br />

) )<br />

<strong>and</strong>, therefore, rk Pic(X ) = rk(S/S 0 ) G = rkN(S/S 0 ) = rk[(NS + S 0 )/S 0 ]. I.e.,<br />

rkPic(X ) = rkNS/(NS ∩ S 0 )<br />

= rkNS − rk(NS ∩ S 0 ) .<br />

Observe that rkNS is the number of Gal(K/K ) orbits into which the<br />

27 lines are decomposed. The group NS ∩ S 0 has to be computed anyway<br />

for H 1( Gal(K/K ), Pic(X K<br />

) ) .<br />

8.23. Explicit computation. –––– There are exactly 350 conjugacy classes of<br />

sub<strong>groups</strong> in W (E 6 ). UsingGAP, we computed the right h<strong>and</strong> side in each case.<br />

The result is the following list.


96 ON THE BRAUER GROUP OF A SCHEME [Chap. II<br />

TABLE 1. H 1 (G,Pic) <strong>and</strong> rk Pic(X ) for smooth cubic surfaces<br />

1 #U = 1 [ ], #H^1 = 1, Rk(Pic) = 7, Orbits [ 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1,<br />

1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1 ]<br />

2 #U = 2 [ 2 ], #H^1 = 1, Rk(Pic) = 6, Orbits [ 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1,<br />

1, 2, 2, 2, 2, 2, 2 ]<br />

3 #U = 2 [ 2 ], #H^1 = 4 [ 2, 2 ], Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 2, 2, 2, 2, 2, 2, 2,<br />

2, 2, 2, 2 ]<br />

...........................................<br />

347 #U = 1440 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 12, 15 ]<br />

348 #U = 1920 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 10, 16 ]<br />

349 #U = 25920 [ ], transitive<br />

/2<br />

350 #U = 51840 [ 2 ], transitive<br />

/3<br />

0 for 257 classes,<br />

for 65 classes,<br />

for 16 classes,<br />

(/2) 2 for 11 classes,<br />

The code written in GAP as well as the full list are reproduced in the appendix.<br />

8.24. –––– It turns out that H 1( Gal(K/K ), Pic(X K<br />

) ) is isomorphic to<br />

(/3) 2<br />

for one class.<br />

The computations took approximately 28 seconds of CPU time.<br />

In 25 of the 257 classes, we have H 1( Gal(K/K ), Pic(X K<br />

) ) = 0 for the trivial<br />

reason that the operation of G on the 27 lines is transitive. Among these classes,<br />

there are the general cubic surfaces. I.e., the case that G = W (E 6 ). We will<br />

consider general cubic surfaces in more detail in Chapter IX.<br />

A different case where H 1( Gal(K/K ), Pic(X K<br />

) ) = 0 is that when G = 0.<br />

This means that all 27 lines are defined over K.<br />

The case where H 1( Gal(K/K ), Pic(X K<br />

) ) = (/3) 2 was described already<br />

by Yu. I. Manin [Man,кVI,ËÐ×ØÚº]. It occurs in a situation<br />

when G ∼ =/3splits the 27 lines into nine orbits of three lines each. This is<br />

realized, for example, by the diagonal cubic surface “x 3 + y 3 + z 3 + dw 3 = 0”<br />

in P 3 K when K ( √ 3 d)/K is a cubic Galois extension.<br />

8.25. Remark. –––– That only these five <strong>groups</strong> may appear was known to<br />

Sir P. Swinnerton-Dyer in 1993, already. Swinnerton-Dyer’s proof fills almost<br />

the entire article [S-D93]. A different proof is given in the Ph.D. thesis of<br />

P. K. Corn [Cor]. That proof consists of a mixture of mathematical arguments<br />

with Computer work.<br />

A purely computational approach was, seemingly, considered rather hopeless at<br />

that time (cf. [Cor, Proposition 1.3.11] <strong>and</strong> the remarks before).


Sec. 8] EXAMPLES 97<br />

8.26. Remark. –––– We find a Picard rank of<br />

1 for 137 classes,<br />

2 for 133 classes,<br />

3 for 62 classes,<br />

4 for twelve classes,<br />

5 for four classes,<br />

6 for one class,<br />

7 for one class.<br />

8.27. Remark. –––– In the next chapter, we will use the list given in Table 1<br />

in order to describe the Brauer-Manin obstruction for a large class of diagonal<br />

cubic surfaces.


CHAPTER III<br />

AN APPLICATION:<br />

THE BRAUER-MANIN OBSTRUCTION<br />

I was at first almost frightened when I saw such mathematical force made<br />

to bear upon the subject, <strong>and</strong> then wondered to see that the subject stood it<br />

so well.<br />

MICHAEL FARADAY (1857, in a letter to J. C. Maxwell)<br />

1. Adelic <strong>points</strong><br />

i. The concept of an adelic point. —<br />

1.1. Definition. –––– Let K be a number field <strong>and</strong> X be a scheme over K.<br />

Then, an adelic point on X is a morphism SpecK → X of K-schemes from<br />

the spectrum of the adele ring [Cas67, Sec. 14]. The set of all adelic <strong>points</strong> on X<br />

is denoted by X (K).<br />

1.2. Remark. –––– The scheme SpecK is not Noetherian.<br />

1.3. Remark. –––– K is a subring ofK via the diagonal homomorphism.<br />

Thus, every K-valued point on X induces an adelic point.<br />

1.4. Remarks. –––– i)K is canonically a sub-K-algebra of ∏ ν K ν . Therefore,<br />

every adelic point on X gives rise to a ∏ ν K ν -valued point.<br />

ii) A ∏ ν K ν -valued point induces a K ν -valued point for every ν ∈ Val(K ).<br />

If the scheme X is separated <strong>and</strong> quasi-compact then the ∏ ν K ν -valued <strong>points</strong><br />

are in a canonical bijection with ∏ ν X (K ν ). This is a technical point which we<br />

defer to Lemma 3.2.


100 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

1.5. Lemma. –––– Let K be a number field <strong>and</strong> X be a separated K-scheme.<br />

Then, the sequence x = (x ν ) ν ∈ ∏ ν X (K ν ) determines the corresponding adelic<br />

point, if it exists, uniquely.<br />

Proof. We first observe that the set<br />

[<br />

SpecK ν ⊂ SpecK<br />

ν∈Val(K )<br />

is Zariski dense. Indeed, a base of the open subsets is provided by the sets D( f )<br />

for 0 ≠ f = ( . . . , f ν , . . . ) ∈K. If f ν ≠ 0 then SpecK ν is contained<br />

in D( f ).<br />

Now suppose that two morphisms x 1 , x 2 : SpecK → X induce the<br />

same element in ∏ ν X (K ν ). This means simply that x 1 <strong>and</strong> x 2 coincide<br />

on S ν∈Val(K ) SpecK ν .<br />

In the category of schemes, there exists a universal morphism i : V → SpecK<br />

such that x 1 ◦i = x 2 ◦i. According to [EGA, Chapitre I, Proposition (5.2.5)],<br />

i is a closed embedding.<br />

Therefore, x 1 <strong>and</strong> x 2 coincide on the Zariski closure of S ν∈Val(K ) SpecK ν . This is<br />

the whole of SpecK.<br />

□<br />

1.6. Lemma. –––– Let K be a number field, O K its ring of integers, <strong>and</strong> X be a<br />

separated O K -scheme of finite type. Denote by X its generic fiber.<br />

Then, a ∏ ν K ν -valued point x = (x ν ) ν ∈ ∏ ν X (K ν ) is induced by an adelic point if<br />

<strong>and</strong> only if all but finitely many of the x ν extend to O Kν -valued <strong>points</strong> on X .<br />

Proof. “⇐=” We are given a finite set S of valuations, including all archimedean<br />

ones, <strong>and</strong> morphisms SpecO Kν → X for ν ∉ S <strong>and</strong> SpecK ν → X for ν ∈ S.<br />

To ease notation, we write R ν := K ν for ν ∈ S <strong>and</strong> R ν := O Kν for ν ∉ S.<br />

Thus, we have morphisms SpecR ν → X . We claim that these may be put<br />

together to give a morphism SpecR → X for R := ∏ ν∈Val(K ) R ν . The assertion<br />

then follows as there is a canonical morphism SpecK → SpecR.<br />

For that, we first note that the assertion is true when X = SpecA is an<br />

affine scheme. Indeed, to give a ring homomorphism A → ∏ ν∈Val(K ) R ν is the<br />

same as giving a system of ring homomorphisms {A → R ν } ν∈Val(K ) .<br />

We cover X by finitely many affine schemes X j<br />

∼ = SpecAj , j ∈ {1, . . . , n}.<br />

Given a system of morphisms {SpecR ν → X } ν∈Val(K ) , we may find a decomposition<br />

Val(K ) = S 1 ∪ . . . ∪ S n<br />

into mutually disjoint subsets such that SpecR ν maps to X j if ν ∈ S j . For this,<br />

observe that all the rings R ν are local.


Sec. 1] ADELIC POINTS 101<br />

Now we use the assumption that the X j are affine. For j fixed, the morphisms<br />

SpecR ν → X j for ν ∈ S j give rise to a morphism<br />

Finally, observe that<br />

nG<br />

j=1<br />

since the index set is finite.<br />

Spec ( ∏<br />

ν∈S j<br />

R ν<br />

) → Xj ⊆ X .<br />

Spec ( )<br />

∏ R ν ∼=<br />

( )<br />

Spec ∏ R ν<br />

i∈S j ν∈Val(K )<br />

“=⇒” Cover X by affine schemes X 1 , . . . , X n <strong>and</strong> write<br />

X j = Spec O K [T j1 , . . . , T jlj ]/I j .<br />

The adelic point given induces homomorphisms of O K -algebras<br />

ϕ j : O K [T j1 , . . . , T jlj ]/I j −→ (K ) fj<br />

where ( f 1 , . . . , f n ) = (1). Fix adeles g 1 , . . . , g n such that g 1 f 1 + . . . +g n f n = 1<br />

<strong>and</strong> write<br />

h jk<br />

:= ϕ j (T jk )<br />

where h jk ∈K <strong>and</strong> e jk ≥ 0.<br />

f e jk<br />

j<br />

We let S be the set of all valuations ν ∈ Val(K ) which are either archimedean<br />

or such that not all of the adeles g j <strong>and</strong> h jk are integral at ν. This is a finite set<br />

of valuations.<br />

Assume ν ∉ S. Then, for one adele f i , we certainly have that ‖( f i ) ν ‖ ν ≥ 1.<br />

Since (h jk ) ν is integral, we see that [ϕ j (T jk )] ν ∈ O Kν for every k. ϕ j induces a<br />

morphism SpecO K → X j as claimed.<br />

□<br />

1.7. Definition (Topology on X (K)). —– Let K be a number field <strong>and</strong> X be<br />

a separated K-scheme of finite type.<br />

Then, X allows a model X which is separated <strong>and</strong> of finite type over the integer<br />

ring O K . The sets X (K ν ) carry natural topologies as ν-adic analytic spaces.<br />

For ν non-archimedean we have the subsets X (O Kν ) ⊆ X (K ν ) carrying the<br />

subspace topology.<br />

We equip the set X (K) of all adelic <strong>points</strong> on X with the restricted topological<br />

product topology [Cas67, Sec. 13]. A basis is provided by all direct products<br />

∏<br />

ν<br />

Γ ν<br />

where Γ ν ⊆ X (K ν ) is open for all ν <strong>and</strong> Γ ν = X (O Kν ) for almost all ν.


)É<br />

102 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

1.8. Remarks. –––– a) For two different models X 1 <strong>and</strong> X 2 , the isomorphism<br />

=<br />

(X 1 )É∼ −→ X ∼ =<br />

−→ (X 2<br />

may be extended to an open neighbourhood of the generic fiber. This implies<br />

X 1 (O Kν ) ∼ = X 2 (O Kν ) for all but finitely many valuations ν. Hence, the definition<br />

is independent on the choice of the model. The existence of a model is<br />

provided by Lemma 3.4.a).<br />

b) X (K) is a Hausdorff topological space satisfying the second axiom of countability.<br />

i) If X = A 1 then there is a bijection between A 1 (K ) <strong>and</strong> the adele ringK.<br />

The topology we defined on A 1 (K) coincides with the adele topology onK<br />

defined in [Cas67, Sec. 14].<br />

ii) If X = A 1 \{0} then X (K) corresponds to the idele group. It is, however,<br />

equipped with the restriction of the adele topology, not with the idele topology.<br />

1.9. Lemma. –––– Let K be a number field <strong>and</strong> X be a scheme proper over K.<br />

Then,<br />

X (K) = ∏ X (K ν )<br />

ν∈Val(K )<br />

equipped with the Tychonov topology.<br />

Proof. Lemma 3.4.b) guarantees the existence of a model X which is proper<br />

over O K except for a finite set S of special fibers. The valuative criterion for<br />

properness implies that X (O Kν ) = X (K ν ) for ν ∉ S.<br />

□<br />

ii. Weak approximation <strong>and</strong> the Hasse principle. —<br />

1.10. Definition. –––– Let X be a separated scheme which is of finite type over<br />

a number field K.<br />

a) One says that X satisfies the Hasse principle if the condition X (K) ≠ ∅<br />

implies X (K ) ≠ ∅.<br />

b) X is said to satisfy weak approximation if X (K ) is dense in X (K).<br />

1.11. Theorem. –––– a) The following classes of projective schemes satisfy the<br />

Hasse principle.<br />

i) Smooth projective quadrics,<br />

ii) Brauer-Severi varieties,<br />

iii) smooth projective cubics in P nÉfor n ≥ 9,<br />

iv) smooth complete intersections of two quadrics in P n K<br />

for n ≥ 8.


Sec. 2] THE BRAUER-MANIN OBSTRUCTION 103<br />

b) The following projective schemes satisfy weak approximation.<br />

i) Projective varieties <strong>rational</strong> over K,<br />

for n ≥ 6 satisfy-<br />

ii) smooth complete intersections X of two quadrics in P n K<br />

ing X (K ) ≠ ∅.<br />

1.12. Remark. –––– These lists are not meant to be exhaustive. Further notes<br />

<strong>and</strong> references to the literature may be found in [Sk, Sec. 5.1].<br />

2. The Brauer-Manin obstruction<br />

2.1. –––– Weak approximation <strong>and</strong> even the Hasse principle are not always satisfied.<br />

Isolated counter examples have been known for a long time.<br />

Genus one curves violating the Hasse principle have been constructed by<br />

C.-E. Lind [Lin] as early as 1940 <strong>and</strong> by E. S. Selmer [Sel] in 1951.<br />

The first example of a cubic surface not fulfilling the Hasse principle is due to<br />

Sir P. Swinnerton-Dyer [S-D62]. A series of examples generalizing Swinnerton-Dyer’s<br />

work has been given by L. J. Mordell [Mord]. We will generalize<br />

Mordell’s examples even further in Section 5.<br />

Diagonal cubic surfaces which are counterexamples to the Hasse principle have<br />

been constructed by J. W. S. Cassels <strong>and</strong> M. J. T. Guy [Ca/G] as well as A. Bremner<br />

[Bre]. These investigations were systematized by J.-L. Colliot-Thélène,<br />

D. Kanevsky, <strong>and</strong> J.J. Sansuc [CT/K/S]. We will discuss this topic in more<br />

detail in Section 6.<br />

A singular quartic surface which is a counterexample to the Hasse principle has<br />

been given by V. A. Iskovskikh [Is].<br />

A method to explain all these examples in a unified manner was provided by<br />

Yu. I. Manin in his book on cubic forms [Man]. This is what is nowadays called<br />

the Brauer-Manin obstruction.<br />

2.2. Definition. –––– Let K be a number field <strong>and</strong> X be a scheme separated <strong>and</strong><br />

of finite type over K.<br />

i) Then, for each ν ∈ Val(K ), there is the local evaluation map, given by<br />

ev ν : Br(X ) × X (K ν ) −→É/,<br />

(α, x) ↦→ inv ν (α| x ) .<br />

ii) Further, there is the global evaluation map or Manin map<br />

ev: Br(X ) × X (K) −→É/,<br />

( )<br />

α, (xν ) ν ↦→ ∑ev ν (α, x ν ) .<br />

ν


104 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

2.3. Proposition. –––– Let K be a number field <strong>and</strong> π : X → SpecK be a scheme<br />

which is separated <strong>and</strong> of finite type over K.<br />

Then,<br />

a) The local evaluation map is<br />

i) additive in the first variable <strong>and</strong><br />

ii) continuous in the second variable.<br />

iii) If α = π ∗ a for a ∈ Br(SpecK ) then ev ν (α, ·) = inv ν (a) is constant.<br />

iv) For each α ∈ Br(X ), there exists a finite set S of valuations including all archimedean<br />

ones such that<br />

ev ν (α, x) = 0<br />

whenever ν ∉ S <strong>and</strong> x ∈ X (O Kν ).<br />

b) The Manin map is well-defined. Further, it is<br />

i) additive in the first variable <strong>and</strong><br />

ii) continuous in the second variable.<br />

iii) In the first variable, ev factors via Br(X )/π ∗ Br(SpecK ).<br />

iv) Further,<br />

if x ∈ X (K ) ⊆ X (K).<br />

Proof. a) i) <strong>and</strong> iii) are obvious.<br />

ev(α, x) = 0<br />

ii) Let α ∈ Br(X ) <strong>and</strong> x ∈ X (K ν ). We have to show that ev ν (α, ·) is constant in<br />

a neighbourhood of x in the ν-adic topology.<br />

By iii), we may assume that ev ν (α, x) = 0. Then, there is an Azumaya algebra A<br />

over X such that [A ] = α <strong>and</strong> A | x<br />

∼ = Mn (K ν ). By Corollary II.5.4, this implies<br />

A | SpecO h<br />

XKν ,x<br />

∼ = Mn (O h X Kν ,x)<br />

for the Henselization of the local ring at x ∈ X Kν . To define this isomorphism,<br />

only finitely many data are required. Thus, there exist an étale morphism<br />

f : X ′ → X <strong>and</strong> a K ν -valued point x ′ ∈ X ′ such that f (x ′ ) = x<br />

<strong>and</strong> A | X ′<br />

∼ = Mn (O X ′).<br />

f induces a local isomorphism of ν-adic analytic spaces.<br />

A | y<br />

∼ = Mn (K ν ) for y in a ν-adic neighbourhood of x.<br />

In particular,<br />

iv) By Lemma 3.4.a), there exists a model X of X which is separated <strong>and</strong> of<br />

finite type over O K . We may assume that X is reduced.<br />

Let A be an Azumaya algebra over X such that [A ] = α. By Lemma 3.5,<br />

there is an open subset X ◦ ⊆ X such that A may be extended to an Azumaya<br />

algebra à over X ◦ .


Sec. 2] THE BRAUER-MANIN OBSTRUCTION 105<br />

The closed subset X \X ◦ ⊂ X does not meet the generic fiber. Since X is of<br />

finite type, X \X ◦ is contained in finitely many special fibers. We put S to be<br />

the set of the corresponding valuations together with all archimedean valuations.<br />

We have to prove that ev ν (α, x) = 0 for ν ∉ S <strong>and</strong> x ∈ X (K ν ). The point x<br />

induces a morphism x : Spec O Kν → X of O K -schemes. The assumption ν ∉ S<br />

implies that x actually maps to X ◦ .<br />

Over X ◦ , there is the Azumaya algebra à extending A . The restriction<br />

of [A ] ∈ Br(X ) along x : SpecK ν → X coincides with the restriction<br />

of [Ã ] ∈ Br(X ◦ x<br />

) along SpecK ν −→ X −→ ⊂<br />

X . The latter factorizes<br />

via Br(Spec O Kν ). We note, finally, that Br(Spec O Kν ) = 0 by virtue of Example<br />

8.1.ii).<br />

b) In order to show that ev is well-defined, one has to verify that the sum is<br />

always finite. This, in turn, immediately follows from a.iv).<br />

i) <strong>and</strong> iii) are direct consequences from a.i) <strong>and</strong> a.iii).<br />

ii) This follows from a.ii) together with a.iv).<br />

iv) is a consequence of the description of Br(K ) given in Example II.8.3.i).<br />

2.4. Remark. –––– Let x ∈ X (K) be an adelic point. If there exists a Brauer<br />

class α ∈ Br(X ) such that ev(α, x) ≠ 0 then x can not be approximated by a<br />

sequence of K-valued <strong>points</strong>. The Brauer class α therefore “obstructs” the adelic<br />

point x from being approximated by <strong>rational</strong> <strong>points</strong>. This justifies the name<br />

Brauer-Manin obstruction.<br />

2.5. Notation. –––– Let K be a number field <strong>and</strong> π : X → SpecK be a scheme<br />

separated <strong>and</strong> of finite type over K.<br />

For α ∈ Br(X ), we write<br />

We define<br />

X (K) α := {x ∈ X (K) | ev(α, x) = 0} .<br />

X (K ) Br := \<br />

X (K ) α .<br />

α∈Br(X )<br />

2.6. Remarks. –––– a) According to Proposition 2.3.b.ii), X (K) Br is always a<br />

closed subset of X (K).<br />

b) Proposition 2.3.b.iv) shows X (K ) ⊆ X (K) Br .<br />

2.7. Remark. –––– As shown in Corollary II.8.12.ii), there are many particular<br />

cases in which Br(X )/π ∗ Br(K ) is actually a finite group.<br />

In this case, one might choose a finite system α 1 , . . . , α n ∈ Br(X ) generating<br />

Br(X )/π ∗ Br(SpecK ). This leads to X (K) Br = X (K) α 1 ∩ . . . ∩ X (K) α n<br />

.<br />


106 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

By Proposition 2.3.a.iv), only finitely many places of K are involved in the<br />

evaluation of α 1 , . . . , α n .<br />

2.8. Definition. –––– Let K be a number field <strong>and</strong> π : X → SpecK be a scheme<br />

separated <strong>and</strong> of finite type over K.<br />

a) If X (K) Br ≠ X (K) then one says that, on X, there is a Brauer-Manin<br />

obstruction to weak approximation.<br />

b) If X (K) Br = ∅ <strong>and</strong> X (K) ≠ ∅ then one says that, on X, there is a<br />

Brauer-Manin obstruction to <strong>rational</strong> <strong>points</strong> on X or that there is a Brauer-Manin<br />

obstruction to the Hasse principle on X.<br />

2.9. Remark. –––– In the case that X (K ) Br ≠ ∅, many authors use the somewhat<br />

confusing formulation that the Brauer-Manin obstruction would be empty.<br />

2.10. Remark. –––– There is clearly no Brauer-Manin obstruction when<br />

Br(X )/π ∗ Br(SpecK ) = 0, neither to the Hasse principle nor to weak approximation.<br />

We know several such cases from the computations described above. For example,<br />

as indicated in II.8.23, we have Br(X )/π ∗ Br(SpecK ) = 0 for X a general<br />

cubic surface or a smooth cubic surface such that all 27 lines are defined over K.<br />

By Fact II.8.14, we have Br(X )/π ∗ Br(SpecK ) = 0 when X is a smooth complete<br />

intersection of dimension ≥ 3 in P n K .<br />

2.11. Definition. –––– If the statement<br />

X (K ) Br ≠ ∅ =⇒ X (K ) ≠ ∅<br />

is true for a certain class of separated K-schemes of finite type then one says that<br />

the Brauer-Manin obstruction is the only obstruction to the Hasse principle for<br />

that class.<br />

2.12. Conjecture (Colliot-Thélène, cf. [CT/S, Conjecture C]). —–<br />

The Brauer-Manin obstruction is the only obstruction to the Hasse principle for<br />

smooth cubic surfaces.<br />

3. Technical lemmata<br />

3.1. –––– Let (K i ) i∈Æbe a sequence of fields. Then, there is the canonical<br />

morphism of schemes<br />

G<br />

ι:<br />

i −→ Spec<br />

i∈ÆSpecK ( )<br />

∏ i .<br />

i∈ÆK


Sec. 3] TECHNICAL LEMMATA 107<br />

3.2. Lemma. –––– LetX beaschemewhichisseparated<strong>and</strong>quasi-compact. Then,ι<br />

induces a bijection<br />

X ( ∏<br />

i∈ÆK i<br />

) −→ ∏<br />

i∈ÆX(K i ) .<br />

Proof. We first note that the assertion is true when X = SpecA is an<br />

affine scheme. Indeed, to give a ring homomorphism A → ∏ i∈ÆK i is the<br />

same as giving a system of ring homomorphisms (A → K i ) i∈Æ.<br />

Surjectivity. We cover X by finitely many affine schemes X j<br />

∼ = SpecAj ,<br />

j ∈ {1, . . . , n}. Given a system of morphisms (SpecK i → X ) i∈Æ, we may find<br />

a decomposition<br />

Æ=S 1 ∪ . . . ∪ S n<br />

into mutually disjoint subsets such that SpecK i maps to X j if i ∈ S j .<br />

Now we use the assumption that the X j are affine. For j fixed, the morphisms<br />

SpecK i → X j for i ∈ S j give rise to a morphism<br />

Finally, observe that<br />

nG<br />

j=1<br />

since the index set is finite.<br />

Spec ( ∏<br />

i∈S j<br />

K i<br />

) → Xj ⊆ X.<br />

Spec ( ∏<br />

i∈S j<br />

K i<br />

) ∼= Spec<br />

(<br />

∏<br />

i∈ÆK i<br />

)<br />

Injectivity. Suppose two morphisms g 1 , g 2 : Spec(∏ i∈ÆK i ) → X induce the<br />

same morphism on F i∈ÆSpecK i . Proposition (5.2.5) of [EGA, Chapitre I] implies<br />

that g 1 <strong>and</strong> g 2 coincide on the Zariski closure of F i∈ÆSpecK i .<br />

We claim that F i∈ÆSpecK i is actually dense in Spec(∏ i∈ÆK i ). Indeed, a base of<br />

the open subsets is provided by the sets D( f ) for<br />

0 ≠ f = ( . . . , f ν , . . . ) ∈ ∏<br />

i∈ÆK i .<br />

If f ν ≠ 0 then SpecK ν is contained in D( f ). The claim follows.<br />

□<br />

3.3. Remark. –––– The assertion of the lemma is certainly not true in general<br />

for X an arbitrary scheme. Indeed, Yoneda’s lemma would then imply<br />

that ι: F i∈ÆSpecK i → Spec(∏ i∈ÆK i ) is an isomorphism.<br />

ι induces, however, not even a bijection on closed <strong>points</strong>. In fact, there is<br />

the ideal ⊕ i∈ÆK i ⊂ ∏ i∈ÆK i . According to the lemma of Zorn, this ideal is<br />

contained in a maximal ideal which is different from those in F i∈ÆSpecK i .


108 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

3.4. Lemma. –––– Let K be a number field <strong>and</strong> X be a scheme which is separated<br />

<strong>and</strong> of finite type over K.<br />

a) Then, there exists a model X of X. This means that X is a scheme which is<br />

separated <strong>and</strong> of finite type over O K <strong>and</strong> fulfills X × SpecOK SpecK ∼ = X.<br />

b) If X is proper over K then X is proper over Spec(O K ) f for some 0 ≠ f ∈ O K .<br />

Proof. First step. Existence of a model.<br />

X is given by gluing together finitely many affine K-schemes U i = SpecR i .<br />

The intersections U i ∩ U j are affine. Thus, gluing takes place along isomorphisms<br />

ϕ ij : Spec(R i ) gij −→ Spec(R j ) gji .<br />

The K-algebras may be described by generators <strong>and</strong> relations in the form<br />

R i = K [T i1 , . . . , T iji ]/(h i1 , . . . , h iki ). In the polynomials h ik , g ij , only finitely<br />

many denominators are occurring. We, therefore, have R i = R ′ i⊗ OK K for some<br />

finitely generated O K -algebras R ′ i such that g ji ∈ R ′ i .<br />

The ϕ ij induce isomorphisms ϕ # ij : (R′ j ) g ji<br />

⊗ OK K → (R i ′) g ij<br />

⊗ OK K. For some suitable<br />

0 ≠ f ij ∈ O K , the isomorphism ϕ # ij<br />

is the base extension of an isomorphism<br />

(R ′ j ) g ji<br />

⊗ OK (O K ) fij → (R i ′) g ij<br />

⊗ OK (O K ) fij .<br />

Putting R i := (R i) ′ f for f := ∏ f ij , the ϕ ij therefore extend to isomorphisms<br />

i,j<br />

˜ϕ ij : Spec(R i ) gij −→ Spec(R j ) gji .<br />

The cocycle relations “˜ϕ ik = ˜ϕ jk ◦ ˜ϕ ij ” are satisfied after restriction to the<br />

generic fiber. Since the schemes Spec(R i ) gij are of finite type, this implies<br />

that these relations are still fulfilled when the R i are localized only by some<br />

suitable 0 ≠ f ′ ∈ O K .<br />

We constructed gluing data for an O K -scheme X . X is clearly of finite type.<br />

Second step. Separatedness.<br />

Denote by ∆ ⊂ X × SpecOK X the Zariski closure of the diagonal. Then, the<br />

diagonal morphism δ : X → ∆ is an isomorphism on the generic fiber. Since X<br />

<strong>and</strong> ∆ are O K -schemes of finite type, δ is an isomorphism outside finitely many<br />

special fibers. We delete the corresponding special fibers from X .<br />

Third step. Properness.<br />

Assume first that X is reduced. Then, by Chow’s lemma, we have a surjective<br />

map X ′ → X from a projective K-scheme X ′ ⊆ P n K . Let X ′ be the Zariski<br />

closure of X ′ in P n O K<br />

equipped with the induced reduced structure. This is a<br />

model of X ′ which is proper over Spec O K .


Sec. 3] TECHNICAL LEMMATA 109<br />

Let X be any model of X in the sense of the previous steps. Then, there<br />

exists some 0 ≠ f ∈ O K such that the morphism X ′ → X extends to a morphism<br />

of (O K ) f -schemes g : X ′ × SpecOK Spec(O K ) f → X . [EGA, Chapitre II,<br />

Corollaire (5.4.3.ii)] shows that im g is proper over Spec(O K ) f . This implies that<br />

im g ⊆ X × SpecOK Spec(O K ) f is a closed subscheme containing the generic fiber.<br />

In particular, the closure ofX in X × SpecOK Spec(O K ) f is proper over Spec(O K ) f .<br />

Return to the general case. Let X be any model of X in the sense of the<br />

two steps above. We may assume that X is equal to the closure of X.<br />

Then, there exists some 0 ≠ f ∈ O K such that X red × SpecOK Spec(O K ) f is<br />

proper over Spec(O K ) f . By virtue of [EGA, Chapitre II, Corollaire (5.4.6)], this<br />

suffices for X × SpecOK Spec(O K ) f being proper.<br />

□<br />

3.5. Lemma. –––– Let K be a number field, X a reduced scheme which is separated<br />

<strong>and</strong> of finite type over the integer ring O K , <strong>and</strong> A be an Azumaya algebra over the<br />

generic fiber X.<br />

Then, there exist an open subset X ◦ ⊆ X containing X <strong>and</strong> an Azumaya algebra<br />

à over X ◦ which extends A .<br />

Proof. First step. Extending A to a coherent sheaf.<br />

Let j : X → X be the embedding of the generic fiber. Then, j ∗ A is a quasicoherent<br />

sheaf on X . We claim, j ∗ A contains a coherent subsheaf F such that<br />

F | X = A .<br />

Assume that would not be the case. We construct an increasing sequence of<br />

subsheaves of j ∗ A recursively as follows.<br />

Put F 1 := 〈1〉 for the section 1 ∈ Γ(X , j ∗ A ). Having F n already constructed,<br />

we have, by our assumption, F n | X A . We observe that X is a Noetherian<br />

scheme. Therefore, j ∗ A is the union of its coherent subsheaves [Ha77,<br />

Chap. II, Exercise 5.15.a)]. By consequence, there exists a coherent sheaf<br />

G ⊆ j ∗ A such that G | X ⊈ F n | X . We put F n+1 := F n + G .<br />

This means, (F n | X ) n∈Æis a strictly increasing sequence of coherent subsheaves<br />

of A . This is a contradiction.<br />

Second step. Extending A to a locally free sheaf.<br />

We may assume X to be connected. Then,<br />

ψ(x) := dim k(x) F X ⊗ OX k(x)<br />

is constant on the generic fiber X. As ψ is upper semicontinuous, there is<br />

an open subset X ′ ⊆ X containing the generic fiber such that ψ is constant<br />

on X ′ . By reducedness, this suffices for F | X ′ being locally free.<br />

Third step. Extending A to an Azumaya algebra.<br />

j ∗ A carries a natural structure of a sheaf of O X -algebras.


110 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

F ⊂ j ∗ A contains the constant section 1 by the construction given in the<br />

first step. To obtain a sheaf of algebras, we must establish closedness under multiplication.<br />

The image F ′ of the multiplication map F ×F → j ∗ A is a coherent subsheaf.<br />

On the generic fiber, we have F ′ | X ⊆ F | X . This implies Q := (F ′ + F )/F<br />

is a coherent sheaf on X such that Q| X = 0. Upper semicontinuity shows that<br />

Q vanishes on an open neighbourhood X ′′ of the generic fiber. F | X ′ ∩X ′′ is a<br />

locally free sheaf of O X ′ ∩X ′′-algebras.<br />

By Corollary 1.5, the non-Azumaya locus is again a closed subset.<br />

□<br />

4. Computing the Brauer-Manin obstruction – The general strategy<br />

4.1. –––– In this section, we want to illustrate that the Manin pairing ev(α, x)<br />

is effectively computable in certain cases.<br />

We will assume that X is a geometrically integral scheme which is proper over<br />

a number field K. Further, we suppose that Br ′ (X K<br />

) = 0 <strong>and</strong> Br(X ) = Br ′ (X ).<br />

Under these assumptions, Br(X )/π ∗ Br(SpecK ) ∼ = H 1( Gal(K/K ), Pic(X K<br />

) ) .<br />

These assumptions are fulfilled, e.g., for X a smooth proper surface such that<br />

X K<br />

is a (not necessarily minimal) <strong>rational</strong> surface.<br />

We may also suppose X (K) ≠ ∅ as, otherwise, the Brauer-Manin obstruction<br />

would not be of much interest.<br />

4.2. –––– Usually, α ∈ Br(X ) is not given itself, but only its image<br />

α ∈ Br(X )/π ∗ Br(SpecK ) ∼ = H 1( Gal(K/K ), Pic(X K<br />

) ) .<br />

In order to compute ev(α, x) = ∑ ν ev ν (α, x ν ), it is therefore necessary to explicitly<br />

lift α to Br(X ). Note that the ev ν , as opposed to ev itself, do not factor<br />

via Br(X )/π ∗ Br(SpecK).<br />

4.3. Lemma. –––– LetK beanumberfield<strong>and</strong>π: X → SpecK beageometrically<br />

integral scheme which is proper over K.<br />

Then, there is a natural isomorphism<br />

ker(Br ′ (X ) → i′<br />

Br ′ (X K<br />

))/π ∗ ∼=<br />

Br(SpecK ) −→ H 1( Gal(K/K ), Pic(X K<br />

) )


Sec. 4] COMPUTING THE BRAUER-MANIN OBSTRUCTION 111<br />

induced by the commutative diagram<br />

Br(SpecK )<br />

Br(SpecK )<br />

π ∗<br />

0 ker(Br ′ (X ) → i′<br />

Br ′ (X K )) H 2( Gal(K/K ),Q(X K ) ∗) H 2( Gal(K/K ),Div(X K ) )<br />

0 H 1( Gal(K/K ),Pic(X K ) ) H 2( Gal(K/K ),Q(X K ) ∗ /K ∗) H 2( Gal(K/K ),Div(X K ) ) .<br />

Here, the bottom row is part of the long exact sequence of cohomology associated to<br />

the exact sequence<br />

0 −→ Q(X K<br />

) ∗ /K ∗ −→ Div(X K<br />

) −→ Pic(X K<br />

) −→ 0 .<br />

The middle column is part of the long exact cohomology sequence associated to the<br />

exact sequence<br />

0 −→ K ∗ −→ Q(X K<br />

) ∗ −→ Q(X K<br />

) ∗ /K ∗ −→ 0 .<br />

Finally, the middle row is obtained when mapping the short exact sequence from<br />

Proposition II.4.8 to<br />

0 −→ H 2( Gal(Q(X K<br />

)/K ), Q(X K<br />

) ∗) −→ H 2( Gal(Q(X K<br />

)/K ), Q(X K<br />

) ∗) −→ 0<br />

0<br />

<strong>and</strong> taking kernels.<br />

□<br />

4.4. Remark. –––– The isomorphism given in Lemma 4.3 is the same as the<br />

isomorphism provided by the Hochschild-Serre spectral sequence (Proposition<br />

II.8.11). This is asserted in [Lic, Sec. 2].<br />

4.5. –––– The diagram above leads to the following general strategy.<br />

General strategy for computing ev(α, x).<br />

a) Compute the image of α in H 2( Gal(K/K ), Q(X K<br />

) ∗ /K ∗) .<br />

b) Lift that to a cohomology class in H 2( Gal(K/K ), Q(X K<br />

) ∗) .<br />

c) For each ν,<br />

i) restrict to H 2( Gal(K ν /K ν ), Q(X Kν<br />

) ∗) .<br />

ii) In a neighbourhood U of x ν such that Pic(U Kν<br />

) = 0, use the exact sequence<br />

H 2( Gal(K ν /K ν ), Γ(U Kν<br />

, O ∗ U Kν<br />

) ) −→ H 2( Gal(K ν /K ν ), Q(X Kν<br />

) ∗) −→<br />

−→ H 2( Gal(K ν /K ), Div(U Kν<br />

) )


112 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

in order to lift to H 2( Gal(K ν /K ν ), Γ(U Kν<br />

, O ∗ U Kν<br />

) ) .<br />

iii) Apply the evaluation map Γ(U Kν<br />

, O ∗ U Kν<br />

) → K ∗ ν at x ν to get a cohomology<br />

class in H 2( Gal(K ν /K ν ), K ∗ ν)<br />

.<br />

iv) Take its invariant inÉ/.<br />

d) Take the sum of all these invariants.<br />

4.6. Remark. –––– In practice, there are several problems with that strategy.<br />

First of all, our skills in dealing with general Galois cohomology<br />

classes are limited. Anyway, for some suitable finite field extension L/K,<br />

α ∈ H 1( Gal(K/K ), Pic(X K<br />

) ) is the image under inflation of a class in<br />

H 1( Gal(L/K ), Pic(X L ) ) .<br />

4.7. Observations (Inflation from a finite quotient Gal(L/K )). —–<br />

a) There is an exact sequence<br />

0 −→ Q(X L ) ∗ /L ∗ −→ Div(X L ) −→ Pic(X L ) −→ 0<br />

inducing a map H 1( Gal(L/K ), Pic(X L ) ) → H 2( Gal(L/K ), Q(X L ) ∗ /L ∗) .<br />

b) ThehomomorphismH 2( Gal(L/K ), Q(X L ) ∗) → H 2( Gal(L/K ), Q(X L ) ∗ /L ∗)<br />

is not surjective, in general. It is, however, when Gal(L/K ) is cyclic. Indeed, in<br />

this case there is a non-canonical isomorphism<br />

H 3( Gal(L/K ), L ∗) ∼ = H<br />

1 ( Gal(L/K ), L ∗)<br />

<strong>and</strong> the latter group vanishes by Hilbert’s Theorem 90.<br />

c) The restriction in i) goes to H 2( Gal(L w /K ν ), Q(X Lw ) ∗) for w a prime above ν.<br />

In step ii), one can work with U = Spec O XLw ,x ν<br />

. There is the exact sequence<br />

H 2( Gal(L w /K ν ), Γ(U Lw , O ∗ U Lw<br />

) ) −→ H 2( Gal(L w /K ν ), Q(X Lw ) ∗) −→<br />

−→ H 2( Gal(L w /K ), Div(U Lw ) )<br />

such that one may lift to H 2( Gal(L w /K ν ), Γ(U Lw , O ∗ U Lw<br />

) ) .<br />

4.8. Remark. –––– To summarize, assume Gal(L/K ) is cyclic <strong>and</strong> we start with<br />

a class in the image under inflation of H 1( Gal(L/K ), Pic(X L ) ) . Then, all the<br />

cohomology classes obtained according to the general strategy are in the image<br />

of H 2( Gal(L/K ), ·) or H 2( Gal(L w /K ν ), ·),<br />

respectively.<br />

4.9. –––– Thus, we assume that G := Gal(L/K ) is a cyclic group of order n.<br />

In this case, H 2 (G,) ∼ =/n<strong>and</strong> the cup product with a generator<br />

induces an isomorphism Ĥ q (G, A) → Ĥ q+2 (G, A) for all integers q <strong>and</strong>


Sec. 5] THE EXAMPLES OF MORDELL 113<br />

all G-modules. We fix a generator of H 2 (G,) in order to determine these<br />

isomorphisms uniquely.<br />

Note that there is no distinguished generator of H 2 (G,) unless a generator<br />

of G is fixed. Thus, the isomorphisms discussed above are not canonical.<br />

Nevertheless, there is a commutative diagram, analogous to the one in<br />

Lemma 4.3, with H 2 replaced by Ĥ 0 <strong>and</strong> H 1 replaced by Ĥ −1 .<br />

4.10. –––– We therefore have the following plan.<br />

Plan for computing ev(α, x).<br />

One has<br />

Ĥ −1 (G, Pic(X L )) ∼ = [Div 0 (X L ) G ∩ N Div(X L )]/N Div 0 (X L ) .<br />

The map from step a) becomes the canonical map to Div 0 (X L ) G /N Div 0 (X L ).<br />

Step b) is the lift to Q(X L ) G /NQ(X L ) = Q(X K )/NQ(X L ) under f ↦→ div f .<br />

For each ν, steps c.i), ii) <strong>and</strong> iii) amount to the evaluation<br />

Q(X K )/NQ(X L ) −→ K ∗ ν/NL ∗ w ,<br />

f ↦→ [ f (x ν )]<br />

at x ν . Note that this map is well-defined although a representative in Q(X K )<br />

might have a pole at x ν .<br />

Finally, in step c.iv), one has to apply the chosen isomorphism backwards,<br />

K ∗ ν/NL ∗ w = Ĥ 0 (G, L ∗ w ) ∼ =<br />

−→ H 2 (G, L ∗ w ) inv ν<br />

−→É/.<br />

ev(α, x) is the sum of all these local invariants.<br />

5. The examples of Mordell<br />

i. Formulation of the results. —<br />

5.1. –––– In this section, we present a series of examples of cubic surfaces<br />

overÉfor which the Hasse principle fails. Our series generalizes the examples<br />

of Mordell [Mord]. It was observed by Yu. I. Manin [Man,кVI,º½¹º]<br />

himself that the failure of the Hasse principle in Mordell’s examples may be<br />

explained by the Brauer-Manin obstruction.


114 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

5.2. –––– Let p 0 ≡ 1 (mod 3) be a prime number <strong>and</strong> K/Ébe the unique<br />

cubic field extension contained in the cyclotomic extensionÉ(ζ p0 )/É. We fix<br />

the explicit generator θ ∈ K given by<br />

More concretely, we write<br />

where n is given by p 0 = 6n + 1.<br />

θ := trÉ(ζ p0 )/K (ζ p0 − 1) .<br />

θ = −2n + ∑<br />

i∈(∗ p0 ) 3 ζ i p 0<br />

5.3. Proposition. –––– Consider the cubic surface X ⊂ P 3É, given by<br />

T 3 (a 1 T 0 + d 1 T 3 )(a 2 T 0 + d 2 T 3 ) =<br />

3<br />

∏<br />

i=1<br />

(<br />

T0 + θ (i) T 1 + (θ (i) ) 2 T 2<br />

)<br />

.<br />

Here, a 1 , a 2 , d 1 , d 2 ∈. The θ (i) are the images of θ under Gal(K/É).<br />

i) Let A ≡ 1 (mod 3) be the unique integer such that 4p 0 = A 2 + 27B 2 .<br />

Suppose that d 1 d 2 ≠ 0, a 1 ≠ 0, a 2 ≠ 0, a 1 /d 1 ≠ a 2 /d 2 , <strong>and</strong> that<br />

T (a 1 + d 1 T )(a 2 + d 2 T ) + 4p 0−A 2<br />

(p 2 0 −3p 0−A) 2 = 0<br />

has no multiple zeroes. Then, X is smooth.<br />

ii) Assume that p 0 ∤ d 1 d 2 , that gcd(a 1 , d 1 ) <strong>and</strong> gcd(a 2 , d 2 ) contain only prime factors<br />

which decompose in K, <strong>and</strong> that T (a 1 + d 1 T )(a 2 + d 2 T ) − 1 = 0 has at least one<br />

simple zero inp 0<br />

. Then, X (É) ≠ ∅.<br />

iii) Assume that X is smooth. Suppose further that p 0 ∤ d 1 d 2 <strong>and</strong> gcd(d 1 , d 2 ) = 1.<br />

Then, there is a class α ∈ Br(X ) with the following property.<br />

For an adelic point x = (x ν ) ν , the value of ev(α, x) depends only on the component<br />

x νp0 . Write x νp0 =: (t 0 : t 1 : t 2 : t 3 ). Then, one has ev(α, x) = 0 if <strong>and</strong><br />

only if<br />

a 1 t 0 + d 1 t 3<br />

t 3<br />

is a cube in∗ p0<br />

.<br />

5.4. Remarks. –––– i) K/Éis an abelian cubic field extension. It is totally<br />

ramified at p 0 <strong>and</strong> unramified at all other primes. A prime p is completely<br />

decomposed in K if <strong>and</strong> only if p is a cube modulo p 0 .<br />

ii) We have ∏<br />

3 ( )<br />

T0 + θ (i) T 1 + (θ (i) ) 2 T 2 = NK/É(T 0 + θT 1 + θ 2 T 2 ) .<br />

i=1<br />

iii) The reduction of T (a 1 + d 1 T )(a 2 + d 2 T ) + 4p 0−A 2<br />

modulo p<br />

(p0 2−3p 0−A) 2 0 is exactly<br />

T (a 1 + d 1 T )(a 2 + d 2 T ) − 1.


Sec. 5] THE EXAMPLES OF MORDELL 115<br />

5.5. Remark. –––– Let (t 0 : t 1 : t 2 : t 3 ) ∈ X (Ép 0<br />

). Then, for the reduction,<br />

one has (t 0 : t 3 ) = (1 : s) for s a solution of T (a 1 + d 1 T )(a 2 + d 2 T ) − 1 = 0.<br />

On the other h<strong>and</strong>, if s is a simple solution then, by Hensel’s lemma, there<br />

exists (t 0 : t 1 : t 2 : t 3 ) ∈ X (Ép 0<br />

) such that (t 0 : t 3 ) = (1 : s).<br />

The possible values of s are in bijection with the up to three planes a p 0 -adic<br />

point on X may reduce to.<br />

We will show in Fact 5.19 that noÉp 0<br />

-valued point on X reduces to the triple line<br />

“T 0 = T 3 = 0”.<br />

5.6. Observation. –––– As<br />

a 1 t 0 + d 1 t 3<br />

t 3<br />

= a 1 + d 1 s<br />

s<br />

the value of ev(α, x) depends only on the plane to which its component x νp0 is<br />

mapped under reduction.<br />

5.7. Remark. –––– Since s is a solution of T (a 1 + d 1 T )(a 2 + d 2 T ) − 1 = 0, we<br />

have s ≠ 0 <strong>and</strong> a 1 + d 1 s ≠ 0. Hence, a 1+d 1 s<br />

s<br />

is well defined <strong>and</strong> non-zero inp 0<br />

.<br />

ii. Smoothness. —<br />

5.8. Lemma. –––– The minimal polynomial of θ is<br />

T 3 + p 0 T 2 p 0 − 1<br />

+ p 0 T + p3 0 − 3p0 2 − Ap 0<br />

.<br />

3 27<br />

5.9. Remark. –––– This is an Eisenstein polynomial. We see once more that<br />

K/Éis totally ramified at p 0 . Further, ν(θ) = 1 3 for ν the extension of ν p 0<br />

to K.<br />

5.10. Proof of the lemma. ––––∗ p0<br />

is decomposed into three cosets C 1 , C 2 ,<br />

<strong>and</strong> C 3 when factored by (∗ p0<br />

) 3 . We have<br />

θ (i) = −2n + ∑<br />

j∈C i<br />

ζ j p 0<br />

.<br />

As the sum over all p 0 -th roots of unity vanishes, this immediately implies<br />

θ (1) + θ (2) + θ (3) = −p 0 .<br />

Further, multiplying terms <strong>and</strong> adding-up shows that<br />

θ (1) θ (2) + θ (2) θ (3) + θ (3) θ (1) = 12n 2 + 2(−2n) ∑ ζ j p 0<br />

+ p 0 − 1<br />

j∈∗ p0<br />

3<br />

∑<br />

[ ] 2 p0 − 1<br />

= 12 + 2 p 0 − 1<br />

− p 0 − 1<br />

6 3 3<br />

= p 0 · p0 − 1<br />

.<br />

3<br />

,<br />

j∈∗ p0<br />

ζ j p 0


116 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

In order to establish θ (1) θ (2) θ (3) = −p3 0 +3p2 0 +Ap 0<br />

27<br />

, multiplying expressions leads to<br />

θ (1) θ (2) θ (3) = −8n 3 + 4n 2 ∑ ζ j p 0<br />

− 2n p 0 − 1<br />

j∈∗ p0<br />

3<br />

∑ ζ j p 0<br />

+<br />

j∈∗ p0<br />

( )(<br />

+ ∑ ζ j p 0<br />

j∈C 1<br />

( )( )( )<br />

= −8n 3 + ∑ ζ j p 0 ∑ ζ j p 0 ∑ ζ j p 0<br />

.<br />

j∈C 1 j∈C 2 j∈C 3<br />

)(<br />

∑ ζ j p 0<br />

j∈C 2<br />

)<br />

∑ ζ j p 0<br />

j∈C 3<br />

To calculate the remaining triple product is more troublesome as one needs to<br />

know how often a product ζ j 1<br />

p 0<br />

ζ j 2<br />

p 0<br />

ζ j 3<br />

p 0<br />

for j i ∈ C i is equal to one.<br />

Sublemma 5.11 shows that this happens exactly p−1 · p0+1+A times. As the<br />

3 9<br />

expression is invariant under Gal(É(ζ p0 )/É) <strong>and</strong> there are all in all (p−1)3 summ<strong>and</strong>s,<br />

we find<br />

27<br />

(<br />

)(<br />

∑ ζ j p 0<br />

j∈C 1<br />

)(<br />

∑ ζ j p 0<br />

j∈C 2<br />

)<br />

∑ ζ j p 0<br />

= p 0 − 1<br />

(p 0 + 1 + A)<br />

j∈C 3<br />

27<br />

+ 1<br />

27 [(p 0 − 1) 2 − (p 0 + 1 + A)] ∑<br />

j∈∗ p0<br />

ζ j p 0<br />

= 3p 0 − 1 + Ap 0<br />

.<br />

27<br />

Further, −8n 3 = −8( p 0−1<br />

) 3 = −p3 0 +3p2 0 −3p 0+1<br />

. The assertion follows. □<br />

6 27<br />

5.11. Sublemma. –––– Let D be a non-cube in∗ p0<br />

. Then,<br />

#{(x, y) ∈2 p0<br />

| Dx 3 + D 2 y 3 = 1} = p 0 + 1 + A .<br />

Proof. The number of solutions inp 0<br />

of such an equation may be counted<br />

using Jacobi sums. As in [I/R, Chapter 8, §3], we see<br />

#{(x, y) | Dx 3 + D 2 y 3 = 1} =<br />

where χ is a fixed cubic character.<br />

=<br />

2 2<br />

∑ ∑ ∑<br />

i=0 j=0 a+b=1<br />

2<br />

∑<br />

i=0<br />

χ i (a/D)χ j (b/D 2 )<br />

2<br />

∑ χ i (1/D)χ j (1/D 2 )J (χ i , χ j ) .<br />

j=0<br />

The summ<strong>and</strong>s for i = j are the same as in the case D = 1. If i = 0 <strong>and</strong> j ≠ 0<br />

or vice versa then the corresponding summ<strong>and</strong> is 0. Finally, by [I/R, Chapter 8,<br />

§3, Theorem 1.c)], we have J (χ, χ 2 ) = J (χ 2 , χ) = −1. Altogether,<br />

#{(x, y) | Dx 3 + D 2 y 3 = 1} = #{(x, y) | x 3 + y 3 = 1} + 3 .


Sec. 5] THE EXAMPLES OF MORDELL 117<br />

The claim now follows from a theorem of C. F. Gauß [I/R, Chapter 8, §3,<br />

Theorem 2)].<br />

□<br />

5.12. Notation. –––– We will write<br />

F (T 0 , T 1 , T 2 , T 3 ) := T 3 (a 1 T 0 +d 1 T 3 )(a 2 T 0 +d 2 T 3 )−<br />

3<br />

∏<br />

i=1<br />

(<br />

T0 +θ (i) T 1 +(θ (i) ) 2 T 2<br />

)<br />

.<br />

5.13. Proof of Proposition 5.3.i) –––– Since d 1 d 2 ≠ 0, the projection to the first<br />

three coordinates induces a morphism of K-schemes q : X → P 2Éwhich is finite<br />

of degree three.<br />

The conditions ∂F<br />

∂T 1<br />

= 0 <strong>and</strong> ∂F<br />

∂T 2<br />

= 0 for a singular point x depend only on q(x).<br />

Let us first analyze them.<br />

Writing l i := T 0 + θ (i) T 1 + (θ (i) ) 2 T 2 , we get the system of equations<br />

the solution of which is<br />

θ (1) l 2 l 3 + θ (2) l 3 l 1 + θ (3) l 1 l 2 = 0 ,<br />

θ (1)2 l 2 l 3 + θ (2)2 l 3 l 1 + θ (3)2 l 1 l 2 = 0 ,<br />

(l 2 l 3 : l 3 l 1 : l 1 l 2 ) = (θ (2) θ (3) (θ (2) − θ (3) ) : θ (3) θ (1) (θ (3) − θ (1) ) : θ (1) θ (2) (θ (1) − θ (2) )) .<br />

As ̟ : (T 0 : T 1 : T 2 ) → (l 2 l 3 : l 3 l 1 : l 1 l 2 ) is a quadratic transformation, there are<br />

exactly four <strong>points</strong> in P 2 which fulfill that condition. These are the three <strong>points</strong><br />

of indeterminacy of ̟ which we denote by P 1 , P 2 , <strong>and</strong> P 3 <strong>and</strong> a non-trivial<br />

solution P.<br />

P 1 , P 2 , <strong>and</strong> P 3 are given by l i = l j = 0 for a pair of indices i ≠ j. This implies<br />

that these <strong>points</strong> are not contained in the line “T 0 = 0”. The fiber of q over<br />

any of these three <strong>points</strong> is therefore given by T 3 (a 1 + d 1 T 3 )(a 2 + d 2 T 3 ) = 0<br />

which shows that q is unramified. Consequently, the <strong>points</strong> over P 1 , P 2 , <strong>and</strong> P 3<br />

are smooth <strong>points</strong> of X.<br />

It remains to consider the fiber of P. For that point, we find<br />

( )<br />

θ<br />

(1)<br />

(l 1 : l 2 : l 3 ) =<br />

θ (2) − θ : θ (2)<br />

(3) θ (3) − θ : θ (3)<br />

.<br />

(1) θ (1) − θ (2)<br />

The linear system of equations<br />

T 0 + θ (1) T 1 + (θ (1) ) 2 T 2 =<br />

T 0 + θ (2) T 1 + (θ (2) ) 2 T 2 =<br />

T 0 + θ (3) T 1 + (θ (3) ) 2 T 2 =<br />

θ (1)<br />

θ (2) − θ (3) ,<br />

θ (2)<br />

θ (3) − θ (1) ,<br />

θ (3)<br />

θ (1) − θ (2)


118 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

has the obvious solution<br />

(T 0 : T 1 : T 2 )<br />

= (−3θ (1) θ (2) θ (3) : 2[θ (1) θ (2) + θ (2) θ (3) + θ (3) θ (1) ] : [−(θ (1) + θ (2) + θ (3) )])<br />

( p<br />

3<br />

= 0 − 3p0 2 − Ap 0<br />

: 2 )<br />

9 3 (p 0 − 1)p 0 : p 0<br />

(<br />

)<br />

6(p 0 − 1)<br />

= 1 :<br />

p0 2 − 3p 0 − A : 9<br />

p0 2 − 3p .<br />

0 − A<br />

which is unique since the coefficient matrix is V<strong>and</strong>ermonde.<br />

Using Lemma 5.8again, a direct calculation which is conveniently done inmaple<br />

shows that<br />

(<br />

1 + 6(p )<br />

0 − 1) 9<br />

p0 2 − 3p 0 − A θ(i) +<br />

p0 2 − 3p 0 − A (θ(i) ) 2 = − 4p 0 − A 2<br />

(p0 2 − 3p 0 − A) . 2<br />

3<br />

∏<br />

i=1<br />

Therefore, the fiber of q over P is given by<br />

4p 0 − A<br />

T 3 (a 1 + d 1 T 3 )(a 2 + d 2 T 3 ) +<br />

2<br />

(p0 2 − 3p 0 − A) = 0 . 2<br />

By assumption, this equation has no multiple solutions. Therefore, q is unramified<br />

over P. The <strong>points</strong> above P are hence smooth <strong>points</strong> of X. □<br />

5.14. Remark. –––– The fact that X is smooth is not mentioned in the original<br />

literature, neither in [Mord], nor in [Man].<br />

iii. Existence of an adelic point. —<br />

5.15. Proof of Proposition 5.3.ii) –––– We have to show that X (Éν) ≠ ∅ for<br />

every valuation ofÉ.<br />

X (Ê) ≠ ∅ is obvious. For a prime number p, in order to show X (Ép) ≠ ∅, we<br />

use Hensel’s lemma. It is sufficient to verify that the reduction X p has a smooth<br />

p-valued point.<br />

Case 1: p = p 0 .<br />

As ν p0 (θ) > 0, the reduction X p0 is given by T 3 (a 1 T 0 +d 1 T 3 )(a 2 T 0 +d 2 T 3 ) = T0 3.<br />

This is the union of three planes meeting in the line given by T 0 = T 3 = 0. By assumption,<br />

one of them has multiplicity one <strong>and</strong> is defined overp 0<br />

. It contains<br />

p0 2 smooth <strong>points</strong>.<br />

Case 2: p ≠ p 0 , p ∤ d 1 d 2 .<br />

It suffices to show that there is a smoothp-valued point on the intersection X ′<br />

p<br />

of X p with the hyperplane “T 0 = 0”. This curve is given by the equation<br />

∏<br />

d 1 d 2 T 3<br />

3 = θ (1) θ (2) θ (3) 3<br />

i=1<br />

(T 1 + θ (i) T 2 ) .


Sec. 5] THE EXAMPLES OF MORDELL 119<br />

If p ≠ 3 then this equation defines a smooth genus one curve. It has anp-valued<br />

point by Hasse’s bound.<br />

If p = 3 then the projection X ′<br />

p → P 1 given by (T 1 : T 2 : T 3 ) ↦→ (T 1 : T 2 ) is oneto-one<br />

onp-valued <strong>points</strong>. At least one of them is smooth since ∏ 3 i=1 (T +θ(i) )<br />

is a separable polynomial.<br />

Case 3: p ≠ p 0 , p|d 1 d 2 .<br />

X ′ p := X p ∩ “T 0 = 0” is given by 0 = θ (1) θ (2) θ (3) ∏ 3 i=1 (T 1+θ (i) T 2 ). In particular,<br />

x = (0 : 0 : 0 : 1) ∈ X p (p). We may assume that x is singular.<br />

Then, X p is given as Q(T 0 , T 1 , T 2 )T 3 +K (T 0 , T 1 , T 2 ) = 0 for Q a quadratic form<br />

<strong>and</strong> K a cubic form. If Q ≢ 0 then there is anp-<strong>rational</strong> line l through x such<br />

that Q| l ≠ 0. Hence, l meets X p twice in x <strong>and</strong> once in anotherp-valued point<br />

which is smooth.<br />

Otherwise, (F mod p) does not depend on T 3 , i.e., the left h<strong>and</strong> side of the<br />

equation of X vanishes modulo p. This means, one of the factors on the left<br />

h<strong>and</strong> side vanishes modulo p. Say, we have a 1 ≡ d 1 ≡ 0 (mod p).<br />

Then, by assumption, p decomposes completely in K. At such a prime, X p ′ is<br />

the union of three lines which are all defined overp, different from each other,<br />

<strong>and</strong> meeting in one point. We have plenty of smooth <strong>points</strong>.<br />

□<br />

iv. Construction of a Brauer class. —<br />

5.16. –––– We write G := Gal(K/É).<br />

According to the general strategy, described in the section above, an element<br />

of H 1 (G, Pic(X K )) ⊆ H 1( Gal(É/É), Pic(XÉ) ) may be given by a <strong>rational</strong><br />

function f ∈ Q(X ) such that div( f ) ∈ N Div(X K ). Such a function is<br />

f := a 1T 0 + d 1 T 3<br />

T 3<br />

.<br />

Indeed, “a 1 T 0 + d 1 T 3 = 0” defines a triangle, the three lines of which are given<br />

by that equation <strong>and</strong> T 0 + θ (i) T 1 + (θ (i) ) 2 T 2 = 0 for i = 1, 2, or 3, respectively.<br />

Considered as a Weil divisor, this triangle is the norm of the divisor given by<br />

one of the lines.<br />

We write [ f ] for the Brauer class defined by f .<br />

5.17. Remarks. –––– i) One might work as well with [ f ′ ] for f ′ := a 2T 0 +d 2 T 3<br />

T 3<br />

or<br />

even with both, [ f ] <strong>and</strong> [ f ′ ], as Manin does in [Man,кVI,ÈÖÐÓÒº].<br />

However,<br />

[ ]<br />

[ f ]·[ f ′ (a1 T 0 + d 1 T 3 )(a 2 T 0 + d 2 T 3 )<br />

] =<br />

=<br />

Consequently, X (É)<br />

T3<br />

2<br />

[ f ] = X (É)<br />

[ f ′] .<br />

[ ( )]<br />

T0 + θT 1 + θ 2 T 2<br />

N<br />

= 0 .<br />

T 3


=/3<br />

120 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

ii) Under very mild hypotheses on X, one has Br(X )/π ∗ Br(SpecÉ) ∼<br />

<strong>and</strong> [ f ] is a generator. We will discuss this point in Subsection vi.<br />

5.18. Lemma. –––– Let ν be any valuation ofÉdifferent from ν p0 . Then,<br />

for every adelic point x ∈ X (É).<br />

Proof. First step: Elementary cases.<br />

ev ν ([ f ], x) = 0<br />

If ν = ν ∞ then ev ν ([ f ], x) = 0 since #G = 3 while only the values 1 2 <strong>and</strong> 0<br />

are possible.<br />

If ν = ν p <strong>and</strong> p is decomposed in K then every element ofÉ∗ p is a norm.<br />

Therefore, ev ν ([ f ], x) = 0.<br />

Second step: Preparations.<br />

It remains to consider the case that p remains prime in K. We have to show that<br />

a 1 t 0 +d 1 t 3<br />

t 3<br />

is a norm for each point (t 0 : t 1 : t 2 : t 3 ) ∈ X (Ép). An element w ∈É∗ p<br />

is a norm if <strong>and</strong> only if 3|ν p (w).<br />

It might happen that θ is not a unit in K ν . As K ν /Ép is unramified, there exists<br />

t ∈É∗ p such that θ := tθ ∈ K ν is a unit. The surface X ′ given by<br />

T 3 (a 1 T 0 + d 1 T 3 )(a 2 T 0 + d 2 T 3 ) =<br />

3<br />

∏<br />

i=1<br />

(<br />

T0 + θ (i) T 1 + (θ (i) ) 2 T 2<br />

)<br />

is isomorphic to X. The map ι: (t 0 : t 1 : t 2 : t 3 ) ↦→ (t 0 :<br />

isomorphism X → X ′ which leaves the <strong>rational</strong> function a 1T 0 +d 1 T 3<br />

T 3<br />

Hence, we may assume without restriction that θ ∈ K ν is a unit.<br />

Third step: The case θ is a unit.<br />

We may assume t 0 , t 1 , t 2 , t 3 ∈p are coprime.<br />

t 1t<br />

:<br />

t 2<br />

t 2<br />

: t 3 ) is an<br />

unchanged.<br />

We could have ∏ 3 i=1(<br />

t0 +θ (i) t 1 + (θ (i) ) 2 t 2<br />

) = 0 only when t0 = t 1 = t 2 = 0 since<br />

θ generates a cubic field extension ofÉp. Then, t 3 (a 1 t 0 + d 1 t 3 )(a 2 t 0 + d 2 t 3 ) = 0<br />

implies t 3 = 0 which is a contradiction.<br />

Therefore, both sides of the equation are different from zero. In particular, we<br />

automatically have t 3 ≠ 0.<br />

If ν ( t 3 (a 1 t 0 + d 1 t 3 )(a 2 t 0 + d 2 t 3 ) ) = 0 then a 1t 0 +d 1 t 3<br />

t 3<br />

is clearly a norm. Otherwise,<br />

we have<br />

ν ( ∏<br />

3 (t 0 + θ (i) t 1 + (θ (i) ) 2 t 2 ) ) > 0 .<br />

i=1<br />

This implies that ν(t 0 ), ν(t 1 ), ν(t 2 ) > 0. Then, t 3 must be a unit.<br />

From the equation of X, we deduce ν(d 1 d 2 ) > 0. If ν(d 2 ) > 0 then, according<br />

to the assumption, d 1 is a unit. This shows ν(a 1 t 0 + d 1 t 3 ) = 0 from which the<br />

assertion follows.


Sec. 5] THE EXAMPLES OF MORDELL 121<br />

Thus, assume ν(d 1 ) > 0. Then, d 2 is a unit <strong>and</strong>, therefore, ν(a 2 t 0 + d 2 t 3 ) = 0.<br />

Further, we note that 3|ν ( ∏ 3 i=1 (t 0 + θ (i) t 1 + (θ (i) ) 2 t 2 ) ) since the right h<strong>and</strong> side<br />

is a norm. By consequence,<br />

3|ν ( t 3 (a 1 t 0 + d 1 t 3 )(a 2 t 0 + d 2 t 3 ) ) .<br />

Altogether, we see that 3|ν(a 1 t 0 + d 1 t 3 ) <strong>and</strong> 3|ν ( )<br />

a 1 t 0 +d 1 t 3<br />

t 3<br />

. The claim follows.<br />

□<br />

5.19. Fact. –––– i) The reduction X p0 of X at p 0 is given by<br />

T 3 (a 1 T 0 + d 1 T 3 )(a 2 T 0 + d 2 T 3 ) = T 3<br />

0 .<br />

ii) Over the algebraic closure, X p0 is the union of three planes meeting in a triple line<br />

“T 0 = T 3 = 0”.<br />

iii) NoÉp 0<br />

-valued point on X reduces to the triple line.<br />

Proof. i) <strong>and</strong> ii) are clear.<br />

iii) Let (t 0 : t 1 : t 2 : t 3 ) ∈ X (Ép 0<br />

). We may assume t 0 , t 1 , t 2 , t 3 ∈p 0<br />

are coprime.<br />

We write ν for the extension of ν p0 to K. Then ν(t 0 ) ≥ 1 <strong>and</strong> ν(t 3 ) ≥ 1 together<br />

imply ν(t 3 (a 1 t 0 + d 1 t 3 )(a 2 t 0 + d 2 t 3 )) ≥ 3. On the other h<strong>and</strong>,<br />

ν ( ∏<br />

3 (t 0 + θ (i) t 1 + (θ (i) ) 2 t 2 ) )<br />

i=1<br />

is equal to 1 or 2, since t 1 or t 2 is a unit <strong>and</strong> ν(θ (i) ) = 1 3 .<br />

□<br />

5.20. Lemma. –––– Let (t 0 : t 1 : t 2 : t 3 ) ∈ X (Ép 0<br />

).<br />

Then, ev νp0 ([ f ], (t 0 : t 1 : t 2 : t 3 )) = 0 if <strong>and</strong> only if a 1t 0 +d 1 t 3<br />

t 3<br />

is a cube in∗ p0<br />

.<br />

Proof. (p 0 ) = p 3 is totally ramified. InÉp 0<br />

, there is a uniformizer which is<br />

a norm. Further, a p 0 -adic unit u is a norm if <strong>and</strong> only if u := (u mod p 0 ) is a<br />

cube in∗ p0<br />

.<br />

We have that a 1t 0 +d 1 t 3<br />

t 3<br />

is automatically a p 0 -adic unit. Indeed, suppose that<br />

t 0 , t 1 , t 2 , t 3 ∈p 0<br />

are coprime. Modulo p 0 , the equation of X is<br />

T 3 (a 1 T 0 + d 1 T 3 )(a 2 T 0 + d 2 T 3 ) = T 3<br />

0 .<br />

As no point may reduce to the singular line, we see that t 0 ≠ 0. This implies<br />

t 3 ≠ 0 <strong>and</strong> a 1 t 0 + d 1 t 3 ≠ 0 which is the assertion.<br />


122 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

v. Examples. —<br />

5.21. Corollary. –––– Let p 0 ≡ 1 (mod 3) be a prime number <strong>and</strong> consider the<br />

cubic surface X ⊂ P 3É, given by<br />

T 3 (a 1 T 0 + d 1 T 3 )(a 2 T 0 + d 2 T 3 ) =<br />

3<br />

∏<br />

i=1<br />

(<br />

T0 + θ (i) T 1 + (θ (i) ) 2 T 2<br />

)<br />

.<br />

Here, a 1 , a 2 , d 1 , d 2 ∈<strong>and</strong> assume that a 1 ≠ 0, a 2 ≠ 0, a 1 /d 1 ≠ a 2 /d 2 , p 0 ∤d 1 d 2 ,<br />

<strong>and</strong> that gcd(a 1 , d 1 ) <strong>and</strong> gcd(a 2 , d 2 ) contain only prime factors which decompose<br />

in K.<br />

i) Assume that<br />

T (a 1 + d 1 T )(a 2 + d 2 T ) − 1 = 0<br />

has three different zeroes t (1) , t (2) , t (3) ∈p 0<br />

.<br />

If a 1+d 1 t (1)<br />

, a 1+d 1 t (2)<br />

, <strong>and</strong> a 1+d 1 t (3)<br />

are non-cubes in∗ t (1) t (2) t (3)<br />

p0<br />

then X (É) = ∅. On X, there<br />

is a Brauer-Manin obstruction to the Hasse principle.<br />

If exactly one of the three expressions is a cube in∗ p0<br />

then, on X, there is a Brauer-<br />

Manin obstruction to weak approximation.<br />

ii) Assume that<br />

T (a 1 + d 1 T )(a 2 + d 2 T ) − 1 = 0<br />

has exactly one zero t ∈p 0<br />

which is simple.<br />

a 1 +d 1 t<br />

t<br />

If is not a cube in∗ p0<br />

then X (É) = ∅. On X, there is a Brauer-Manin<br />

obstruction to the Hasse principle.<br />

5.22. Remark. –––– In the case of three different solutions, it is impossible that<br />

exactly two of the expressions a 1+d 1 t (1)<br />

, a 1+d 1 t (2)<br />

, <strong>and</strong> a 1+d 1 t (3)<br />

are cubes. Indeed, a<br />

t (1) t (2) t (3)<br />

direct calculation shows<br />

a 1 + d 1 t (1) + d 1 t<br />

·a1 (2) + d 1 t<br />

·a1 (3)<br />

= d 3<br />

t (1) t (2) t (3) 1 .<br />

This means, the Brauer-Manin obstruction might exclude no, one, or all three<br />

of the planes, the reduction X p0 consists of, but not exactly two of them.<br />

5.23. Example. –––– Let p 0 ≡ 1 (mod 3) be a prime number <strong>and</strong> assume that<br />

d 1 <strong>and</strong> d 2 are non-cubes in∗ p0<br />

such that d 1 d 2 is a cube, p 0 |a 1 <strong>and</strong> p 0 |a 2 . Further,<br />

suppose a 1 ≠ 0, a 2 ≠ 0, <strong>and</strong> a 1 /d 1 ≠ a 2 /d 2 , as well as that gcd(a 1 , d 1 )<br />

<strong>and</strong> gcd(a 2 , d 2 ) contain only prime factors which decompose in K.<br />

Then, the cubic surface X given by<br />

T 3 (a 1 T 0 + d 1 T 3 )(a 2 T 0 + d 2 T 3 ) =<br />

is a counterexample to the Hasse principle.<br />

3<br />

∏<br />

i=1<br />

(<br />

T0 + θ (i) T 1 + (θ (i) ) 2 T 2<br />

)


Sec. 5] THE EXAMPLES OF MORDELL 123<br />

Indeed, one has<br />

a 1 t 0 + d 1 t 3<br />

t 3<br />

≡ d 1 (mod p 0 )<br />

which is a non-cube. The assumption that d 1 d 2 is a cube is important to<br />

guarantee X (Ép 0<br />

) ≠ ∅.<br />

More concretely, for p 0 = 19, a counterexample to the Hasse principle is<br />

given by<br />

T 3 (19T 0 + 5T 3 )(19T 0 + 4T 3 ) =<br />

3<br />

∏<br />

i=1<br />

(<br />

T0 + θ (i) T 1 + (θ (i) ) 2 T 2<br />

)<br />

= T0 3 − 19T0 2 T 1 + 133T0 2 T 2 + 114T 0 T1<br />

2<br />

− 1 539T 0 T 1 T 2 + 5 054T 0 T2 2 − 209T1<br />

3<br />

+ 3 971T1 2 T 2 − 23 826T 1 T2 2 + 43 681T2 3 .<br />

5.24. Example. –––– For p 0 = 19, consider the cubic surface X given by<br />

3 ( )<br />

T 3 (T 0 + T 3 )(12T 0 + T 3 ) = ∏ T0 + θ (i) T 1 + (θ (i) ) 2 T 2 .<br />

i=1<br />

Then, X (É) ≠ ∅ but X (É) = ∅. On X, there is a Brauer-Manin obstruction<br />

to the Hasse principle.<br />

Indeed, in19, the cubic equation<br />

T (1 + T )(12 + T ) − 1 = 0<br />

has the three solutions 12, 15, <strong>and</strong> 17. However, in19, 13/12 = 9, 16/15 = 15,<br />

<strong>and</strong> 18/17 = 10 which are three non-cubes.<br />

5.25. Example. –––– For p 0 = 19, consider the cubic surface X given by<br />

T 3 (T 0 + T 3 )(6T 0 + T 3 ) =<br />

3<br />

∏<br />

i=1<br />

(<br />

T0 + θ (i) T 1 + (θ (i) ) 2 T 2<br />

)<br />

.<br />

Then, on X, there is a Brauer-Manin obstruction to weak approximation.<br />

Indeed, in19, the cubic equation<br />

T (1 + T )(6 + T ) − 1 = 0<br />

has the three solutions 8, 9, <strong>and</strong> 14. However, in19, 10/9 = 18 is a cube while<br />

9/8 = 13 <strong>and</strong> 15/14 = 16 are non-cubes.<br />

The smallestÉ-<strong>rational</strong> point on X is (14 : 15 : 2 : (−7)). Note that indeed<br />

T 3 /T 0 = −7/14 ≡ 9 (mod 19).


124 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

5.26. Example. –––– For p 0 = 19, consider the cubic surface X given by<br />

3 ( )<br />

T 3 (T 0 + T 3 )(2T 0 + T 3 ) = T0 + θ (i) T 1 + (θ (i) ) 2 T 2 .<br />

Then, on X, there is a Brauer-Manin obstruction to the Hasse principle.<br />

Indeed, in19, the cubic equation<br />

∏<br />

i=1<br />

T (1 + T )(2 + T ) − 1 = 0<br />

has T = 5 as its only solution. The other two solution are conjugate to each<br />

other in19 2. However, in19, 6/5 = 5 is a non-cube.<br />

5.27. Example (Swinnerton-Dyer [S-D62]). —– For p 0 = 7, consider the cubic<br />

surface X given by<br />

T 3 (T 0 + T 3 )(T 0 + 2T 3 ) =<br />

3<br />

∏<br />

i=1<br />

(<br />

T0 + θ (i) T 1 + (θ (i) ) 2 T 2<br />

)<br />

= T0 3 − 7T0 2 T 1 + 21T0 2 T 2 + 14T 0 T1 2 − 77T 0 T 1 T 2<br />

+ 98T 0 T2 2 − 7T1 3 + 49T1 2 T 2 − 98T 1 T2 2 + 49T2 3 .<br />

Then, on X, there is a Brauer-Manin obstruction to the Hasse principle.<br />

Indeed, in7, the cubic equation<br />

T (1 + T )(1 + 2T ) − 1 = 0<br />

has T = 5 as its only solution. However, in7, 6/5 = 4 is not a cube.<br />

5.28. Remark. –––– From each of the examples given, by adding multiplies<br />

of p 0 to the coefficients a 1 , d 1 , a 2 , <strong>and</strong> d 2 , a family of surfaces arises which are<br />

of similar nature. In Example 5.23, some care has to be taken to keep a 1 <strong>and</strong> a 2<br />

different from zero <strong>and</strong> a 1 /d 1 different from a 2 /d 2 . In all examples, gcd(a 1 , d 1 )<br />

<strong>and</strong> gcd(a 2 , d 2 ) need some consideration.<br />

5.29. Remark (Lattice basis reduction). —– The norm form in the p 0 = 19 examples<br />

produces coefficients which are rather large. An equivalent form with<br />

smaller coefficients may be obtained using lattice basis reduction. In its simplest<br />

form, this means the following.<br />

For the rank-2 lattice inÊ3 , generated by the vectors v 1 := (θ (1) , θ (2) , θ (3) ) <strong>and</strong><br />

v 2 := ( (θ (1) ) 2 , (θ (2) ) 2 , (θ (3) ) 2) , in fact {v 1 , v 2 + 7v 1 } is a reduced basis. Therefore,<br />

the substitution T 1 ′ := T 1−7T 2 simplifies the norm form. Actually, we find<br />

3<br />

∏<br />

i=1<br />

(<br />

T0 + θ (i) T 1 + (θ (i) ) 2 T 2<br />

)<br />

=T<br />

3<br />

0 −19T0 2 T 1+114T ′<br />

0 T ′<br />

1 2 +57T 0 T ′<br />

1 T 2−133T 0 T2<br />

2<br />

− 209T ′<br />

1 3 − 418T ′<br />

1 2 T 2 + 1045T ′<br />

1 T 2 2 − 209T2 3 .


Sec. 5] THE EXAMPLES OF MORDELL 125<br />

vi. Calculation of the complete Br(X )/π ∗ Br(SpecÉ). —<br />

5.30. Notation. –––– Let X be a cubic surface in P 3 given by<br />

for six linear forms l 1 , l 2 , l 3 , λ 1 , λ 2 , λ 3 .<br />

l 1 l 2 l 3 = λ 1 λ 2 λ 3<br />

Then, we write L ij for the line on X given by l i = λ j = 0.<br />

The subgroup of Pic(X ) generated by these nine lines will be denoted by P.<br />

Further, we let S be the free abelian group over all lines L ij .<br />

5.31. Facts. –––– i) P ⊂ Pic(X ) is a lattice of rank five <strong>and</strong> discriminant 3.<br />

ii) The kernel S 0 of the canonical homomorphism S → P is generated by<br />

D 1 := L 11 +L 12 +L 13 −L 21 −L 22 −L 23 , D 2 := L 11 +L 12 +L 13 −L 31 −L 32 −L 33 ,<br />

D 3 := L 11 + L 13 − L 22 − L 32 , <strong>and</strong> D 4 := L 11 + L 12 − L 23 − L 33 .<br />

Proof. We have D 1 = div(l 1 /l 2 ), D 2 = div(l 1 /l 3 ), D 3 = div(l 1 /λ 2 ),<br />

<strong>and</strong> D 4 = div(l 1 /λ 3 ). Thus, D 1 , D 2 , D 3 , D 4 ∈ ker(S → P). It is easy to<br />

check that D 1 , D 2 , D 3 , D 4 are linearly independent. This shows that rkP ≤ 5.<br />

On the other h<strong>and</strong>, the intersection matrix of L 11 , L 22 , L 23 , L 32 , <strong>and</strong> L 33 is<br />

⎛<br />

⎞<br />

−1 0 0 0 0<br />

0 −1 1 1 0<br />

⎜ 0 1 −1 0 1<br />

⎟<br />

⎝ 0 1 0 −1 1⎠ .<br />

0 0 1 1 −1<br />

The determinant of this matrix is equal to 3 which is a square-free integer.<br />

Assertion i) is proven.<br />

It remains to verify that D 1 , D 2 , D 3 , <strong>and</strong> D 4 generate ker(S → P). For this, we<br />

have to show that the canonical injection<br />

is actually bijective.<br />

〈L 11 , L 22 , L 23 , L 32 , L 33 〉 −→ S/〈D 1 , D 2 , D 3 , D 4 〉<br />

For surjectivity, observe that, modulo 〈D 1 , D 2 , D 3 , D 4 〉, the remaining generators<br />

are given by L 12 ≡ −L 11 + L 23 + L 33 , L 13 ≡ −L 11 + L 22 + L 32 ,<br />

L 21 ≡ L 11 + L 12 + L 13 − L 22 − L 23 ≡ −L 11 + L 32 + L 33 , <strong>and</strong>, finally,<br />

L 31 ≡ L 11 + L 12 + L 13 − L 32 − L 33 ≡ −L 11 + L 22 + L 23 .<br />

□<br />

5.32. Proposition. –––– Let K/Ébe a Galois extension of degree three. Assume<br />

that λ 1 , λ 2 , λ 3 are defined over K <strong>and</strong> form a Galois orbit <strong>and</strong> that l 1 , l 2 , l 3<br />

are linear forms defined overÉ.<br />

Then, Ĥ −1( Gal(K/É), P ) =/3<strong>and</strong> [l 1 /l 2 ] is a generator.


126 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

Proof. By Lemma II.8.19, we have a canonical isomorphism<br />

Ĥ −1( Gal(K/É), P ) ∼ =<br />

−→ (NS ∩ S 0 )/NS 0 .<br />

Let us calculate NS ∩ S 0 .<br />

D 1 = N (L 11 − L 21 ) <strong>and</strong> D 2 = N (L 11 − L 31 ) are clearly norms. If<br />

aD 3 + bD 4 = (a + b)L 11 + bL 12 + aL 13 − aL 22 − bL 23 − aL 32 − bL 33<br />

is a norm then the coefficients of L 31 , L 32 , <strong>and</strong> L 33<br />

Hence, a = b = 0. Consequently,<br />

must coincide.<br />

NS ∩ S 0 = 〈D 1 , D 2 〉 .<br />

Further, N (D 1 ) = 3D 1 , N (D 2 ) = 3D 2 , <strong>and</strong> N (D 3 ) = N (D 4 ) = D 1 + D 2 .<br />

Hence,<br />

NS 0 = 〈3D 1 , 3D 2 , D 1 + D 2 〉 .<br />

The assertion follows.<br />

□<br />

5.33. –––– Return to the case of a cubic surface given an equation of type<br />

3 ( )<br />

T 3 (a 1 T 0 + d 1 T 3 )(a 2 T 0 + d 2 T 3 ) = T0 + θ (i) T 1 + (θ (i) ) 2 T 2 .<br />

To deduce information on H 1( Gal(É/É), Pic(XÉ) ) from H 1 (G, P) ∼ =/3,<br />

we need some underst<strong>and</strong>ing of the remaining 18 lines.<br />

OverÉ(θ), there are defined at least nine lines. The list of the 350 conjugacy<br />

classes of sub<strong>groups</strong> of W (E 6 ), established using GAP (cf. II.8.23 <strong>and</strong> the appendix),<br />

contains only four classes which fix nine or more lines. The corresponding<br />

extract looks like this.<br />

1 #U = 1 [ ], #H^1 = 1, Rk(Pic) = 7, Orbits [ 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1,<br />

1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1 ]<br />

2 #U = 2 [ 2 ], #H^1 = 1, Rk(Pic) = 6, Orbits [ 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1,<br />

1, 2, 2, 2, 2, 2, 2 ]<br />

7 #U = 3 [ 3 ], #H^1 = 1, Rk(Pic) = 5, Orbits [ 1, 1, 1, 1, 1, 1, 1, 1, 1, 3, 3, 3, 3,<br />

3, 3 ]<br />

24 #U = 6 [ 2 ], #H^1 = 1, Rk(Pic) = 5, Orbits [ 1, 1, 1, 1, 1, 1, 1, 1, 1, 3, 3, 3, 3,<br />

3, 3 ]<br />

Thus, the remaining 18 lines are defined over an extension ofÉ(θ) which may<br />

be of degree 6, 3, 2, or 1.<br />

We see that rkPic(XÉ(θ)) ≥ 5. Equality holds if <strong>and</strong> only if the field of definition<br />

of the 27 lines is of degree 3 or 6 overÉ(θ).<br />

We also observe that, in any case,<br />

∏<br />

i=1<br />

H 1( Gal(É/É(θ)), Pic(XÉ) ) = 0 .


Sec. 5] THE EXAMPLES OF MORDELL 127<br />

5.34. Remark. –––– For every concrete choice of the coefficients, one may<br />

determine the field of definition of the 27 lines by a Gröbner base calculation.<br />

It turned out that it was of degree 6 overÉ(θ) in every example, we tested.<br />

Thus, degree 6 seems to be the generic case.<br />

If the field of definition of the 27 lines is a degree 6 or degree 3 extension ofÉ(θ)<br />

then we may describe H 1( Gal(É/É), Pic(XÉ) ) , completely.<br />

5.35. Proposition. –––– Let p 0 ≡ 1 (mod 3) be a prime number, θ (i) as above,<br />

<strong>and</strong> X ⊂ P 3Ébe the cubic surface given by<br />

T 3 (a 1 T 0 + d 1 T 3 )(a 2 T 0 + d 2 T 3 ) =<br />

3<br />

∏<br />

i=1<br />

(<br />

T0 + θ (i) T 1 + (θ (i) ) 2 T 2<br />

)<br />

for a 1 , a 2 , d 1 , d 2 ∈. Assume that the field of definition of the 27 lines on X is of<br />

degree 3 or 6 overÉ(θ).<br />

Then, Br(X )/π ∗ Br(SpecÉ) ∼ =/3<strong>and</strong> [ a 1T 0 +d 1 T 3<br />

T 3<br />

] is a generator.<br />

Proof. Since H 1( Gal(É/É(θ)), Pic(XÉ) ) = 0, the inflation map<br />

H 1( Gal(É(θ)/É), Pic(XÉ) Gal(É/É(θ) ) ) −→ H 1( Gal(É/É), Pic(XÉ) )<br />

is an isomorphism. Hence, in view of Proposition 5.32, it will suffice to<br />

show that P = Pic(XÉ) Gal(É/É(θ)) .<br />

“⊆” is obvious.<br />

“⊇” Our assumption implies that rkPic(XÉ(θ)) = 5. Further, from Proposition<br />

II.8.11, we see that Pic(XÉ(θ)) is always a subgroup of finite index<br />

in Pic(XÉ) Gal(É/É(θ)) . Hence, rkPic(XÉ) Gal(É/É(θ)) = 5, too.<br />

By Fact 5.31.i), we have rkP = 5. Thus, P ⊆ Pic(XÉ) Gal(É/É(θ)) is a sublattice<br />

of finite index. Since DiscP = 3 is square-free, the claim follows. □<br />

5.36. Example. –––– For p 0 = 19, consider the cubic surface X given by<br />

T 3 (T 0 + T 3 )(7T 0 + T 3 ) =<br />

3<br />

∏<br />

i=1<br />

(<br />

T0 + θ (i) T 1 + (θ (i) ) 2 T 2<br />

)<br />

.<br />

Then, on X, there is no Brauer-Manin obstruction to weak approximation.<br />

Indeed, a Gröbner base calculation shows that the 27 lines on X are defined over<br />

a degree 6 extension ofÉ(θ). Thus, Br(X )/π ∗ Br(SpecÉ) ∼ =/3. A generator<br />

is given by the class α := [ T 0+T 3<br />

T 3<br />

].<br />

However, in19, the cubic equation<br />

T (1 + T )(7 + T ) − 1 = 0


128 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

has the only solution T = 11. Hence, everyÉ19-valued point on X has a<br />

reduction of the form (1 : y : z : 11). For the local evaluation, we find<br />

= 12/11 = 8 ∈19 which is a cube.<br />

T 0 +T 3<br />

T 3<br />

Br<br />

Thus, ev(x, α) = 0 for every adelic point x ∈ X (É). By consequence,<br />

X (É) = X (É). One might expect that X satisfies weak approximation.<br />

5.37. Remarks (Degenerate cases). —– i) If the field of definition of the 27 lines<br />

is quadratic overÉ(θ) then Br(X )/π ∗ Br(SpecÉ) = 0.<br />

In fact, in this case, we have a cyclic group G of order 6 acting on the 27 lines.<br />

There are seven conjugacy classes of cyclic <strong>groups</strong> of order 6 in W (E 6 ). We have<br />

the following list.<br />

26 #U = 6 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 6, 6, 6, 6 ]<br />

27 #U = 6 [ 2, 3 ], #H^1 = 4 [ 2, 2 ], Rk(Pic) = 1, Orbits [ 3, 6, 6, 6, 6 ]<br />

28 #U = 6 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 4, Orbits [ 1, 1, 1, 2, 2, 2, 3, 3, 3, 3, 6 ]<br />

30 #U = 6 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 2, 2, 6, 6, 6 ]<br />

32 #U = 6 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 3, 3, 3, 3, 6, 6 ]<br />

36 #U = 6 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 2, 2, 3, 3, 6, 6 ]<br />

39 #U = 6 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 6, 6, 6, 6 ]<br />

As the element of order two fixes 15 lines, we may not have more than two<br />

orbits of size 6. This implies the claim.<br />

Actually, we are in class No. 32.<br />

ii) If the 27 lines are defined overÉ(θ) then, seemingly, there are two cases.<br />

One might either have rk Pic(X ) = 3. Then, Br(X )/π ∗ Br(SpecÉ) = 0.<br />

Or, otherwise, rk Pic(X ) = 1. In this case, Br(X )/π ∗ Br(SpecÉ) = (/3) 2 .<br />

α := [ a 1T 0 +d 1 T 3<br />

T 3<br />

] is one of the generators.<br />

We do not know of an example overÉin which any of these degenerate<br />

cases occurs.<br />

6. The “first case” of diagonal cubic surfaces<br />

i. The statements. —<br />

6.1. –––– In this section, we show how to compute the Brauer-Manin obstruction<br />

for a large class of diagonal cubic surfaces. More precisely, we are concerned<br />

with smooth diagonal cubic surfaces in P 3É, given by an equation of the form<br />

a 0 T 3<br />

0 + a 1 T 3<br />

1 + a 2 T 3<br />

2 + a 3 T 3<br />

3 = 0<br />

for a 0 , . . . , a 3 ∈\{0}, such that the following additional condition is satisfied.<br />

(∗) There exists a prime number p 0 such that p 0 |a 3 but neither p 3 0|a 3 nor p 0 |a i<br />

for i = 0, 1, 2.


Sec. 6] THE “FIRST CASE” OF DIAGONAL CUBIC SURFACES 129<br />

6.2. –––– X has a model over Specgiven by the same equation. We will<br />

denote that model by X .<br />

On X, there is the smooth genus one curve E given by the equation T 3 = 0.<br />

If p 0 ≠ 3 then assumption (∗) implies that the reduction X p0 is a cone over E p0 .<br />

There is the mapping<br />

red: X (Ép 0<br />

) −→ E p0 (p 0<br />

)<br />

(x 0 : x 1 : x 2 : x 3 ) ↦→ ((x 0 mod p 0 ) : (x 1 mod p 0 ) : (x 2 mod p 0 )) .<br />

6.3. Remark. –––– The Brauer-Manin obstruction on diagonal cubic surfaces<br />

was studied by J.-L.Colliot-Thélène, D. Kanevsky, <strong>and</strong> J.-J.Sansuc in [CT/K/S].<br />

For the case when (∗) is fulfilled, the main result asserts that X (É) Br is exactly<br />

one third of the whole of X (É) in a sense made precise.<br />

More concretely, there is the following theorem.<br />

∅. Sup-<br />

6.4. Theorem (Colliot-Thélène, Kanevsky, <strong>and</strong> Sansuc). —–<br />

Let X ⊂ P 3Ébe →É/<br />

a smooth diagonal cubic surface such that X (É) ≠<br />

pose that (∗) is fulfilled for a certain prime number p 0 .<br />

a) Then, Br(X )/π ∗ Br(SpecÉ) ∼ =/3.<br />

b) The image of the evaluation map<br />

ev: Br(X ) × X (É)<br />

is 1 3/.<br />

c) Choose an adelic point (x ν ) ν ∈ X ().<br />

Let α ∈ Br(X ) be such that its image in Br(X )/π ∗ Br(SpecÉ) is non-trivial.<br />

Assume that α is a normalized modulo π ∗ Br(SpecÉ) in such a way that<br />

ev νp0 (α, x νp0 ) = 0.<br />

i) Assume p 0<br />

→É/<br />

≠ 3.<br />

Then, the local evaluation map<br />

ev νp0 : Br(X ) × X (Ép 0<br />

)<br />

has the following property.<br />

ev νp0 (α, ·) is the composition of red with a surjective group homomorphism<br />

(E p0 (p 0<br />

), x νp0 ) → 1 3/.<br />

In particular, ev νp0 (α, x) depends only on the reduction x ∈ X (p 0<br />

).<br />

Further, we have<br />

#{y ∈ X (p 0<br />

) | ev νp0 (α, y)=0} = #{y ∈ X (p 0<br />

) | ev νp0 (α, y)=1/3}<br />

= #{y ∈ X (p 0<br />

) | ev νp0 (α, y)=2/3} .


130 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

ii) Let p 0<br />

→É/<br />

= 3.<br />

Then, the local evaluation map<br />

ev ν3 : Br(X ) × X (Ép 0<br />

)<br />

has the following properties.<br />

The image of ev ν3 (α, ·): X (É3) →É/is 1 3/.<br />

There exists a positive integer n such that ev ν3 (α, x) depends only on the reduction<br />

x (3n )<br />

∈ X (/3 n).<br />

For every y 0 ∈ X (3), we have<br />

In particular,<br />

#{y ∈ X (/3 n) | y = y 0 , ev ν3 (α, y)=0}<br />

= #{y ∈ X (/3 n) | y = y 0 , ev ν3 (α, y)=1/3}<br />

= #{y ∈ X (/3 n) | y = y 0 , ev ν3 (α, y)=2/3} .<br />

#{y ∈ X (/3 n) | ev ν3 (α, y)=0} = #{y ∈ X (/3 n) | ev ν3 (α, y)=1/3}<br />

= #{y ∈ X (/3 n) | ev ν3 (α, y)=2/3} .<br />

6.5. Corollary. –––– Let X ⊂ P 3Ébe a smooth diagonal cubic surface such<br />

that X (É) ≠ ∅. Suppose that (∗) is fulfilled for a certain prime number p 0 .<br />

a) Then, on X, there is a Brauer-Manin obstruction to weak approximation.<br />

b) There is, however, no Brauer-Manin obstruction to the Hasse principle.<br />

6.6. Remark. –––– Colliot-Thélène, Kanevsky, <strong>and</strong> Sansuc verify the same behaviour<br />

in their “first case” which is a bit more general than assumption (∗).<br />

Another particular case in which the same result is true is when p 0 ∤ a 0 , a 1 , but<br />

p 0 |a 2 , a 3 , p 2 0 ∤a 2, a 3 , <strong>and</strong> a 2 /a 3 is a p 0 -adic cube.<br />

The three authors also show that, in the “Second case”, there exist diagonal cubic<br />

surfaces such that X (É) ≠ ∅ <strong>and</strong> X (É) Br = X (É) or ∅. From Theorem 6.4,<br />

it is clear that this may happen only under rather restrictive conditions. For example,<br />

every prime number dividing one of the coefficients must necessarily<br />

divide a second one.<br />

6.7. Remark. –––– The statement that the three values are equally distributed<br />

was not formulated in [CT/K/S].


Sec. 6] THE “FIRST CASE” OF DIAGONAL CUBIC SURFACES 131<br />

6.8. Remark. –––– In order to prove Br(X )/π ∗ Br(SpecÉ) ∼ =/3, we make<br />

heavy use of the list of all possible quotients Br(X )/π ∗ Br(SpecÉ) established<br />

by help of the computer. Cf. II.8.23 <strong>and</strong>, for the list in full, the appendix.<br />

ii. Calculating Br(X )/π ∗ Br(SpecÉ). —<br />

6.9. Notation. –––– Following Manin [Man], we will denote the lines on X<br />

as follows. We fix third roots of a 0 , . . . , a 3 , once <strong>and</strong> for all.<br />

Then, L k (m, n) is the line given by<br />

a 1 3 i T i + ζ m 3 a 1 3<br />

j<br />

T j = 0 ,<br />

a 1 3<br />

k T k + ζ n 3 a 1 3<br />

3<br />

T 3 = 0 .<br />

Here, k ∈ {0, 1, 2}, i, j ∈ {0, 1, 2}\{k}, i < j, <strong>and</strong> m, n ∈/3.<br />

6.10. –––– In this notation, the 45 triangles on X may easily be described.<br />

i) The planes “a 1 3<br />

k T k + ζ n 3 a 1 3<br />

3<br />

T 3 = 0”, for k ∈ {0, 1, 2}, n ∈/3, define 9 triangles<br />

∆ 1 kn . They consist of L k(0, n), L k (1, n), <strong>and</strong> L k (2, n).<br />

ii) Analogously, the planes “a 1 3 i T i + ζ m 3 a 1 3<br />

j<br />

T j = 0”, for i, j ∈ {0, 1, 2}, i < j,<br />

<strong>and</strong> m ∈/3, define 9 triangles ∆ 2 km . They consist of L k(m, 0), L k (m, 1),<br />

<strong>and</strong> L k (m, 2) for k ∈ {0, 1, 2} \ {i, j}.<br />

iii) Finally, the planes given by a 1 3<br />

0<br />

T 0 + ζ a 3 a 1 3<br />

1<br />

T 1 + ζ b 3 a 1 3<br />

2<br />

T 2 + ζ c 3 a 1 3<br />

3<br />

T 3 = 0 for<br />

a, b, c ∈ {0, 1, 2} define 27 triangles ∆ 3 abc . Those consist of L 0(b − a, c),<br />

L 1 (b, c − a), <strong>and</strong> L 2 (a, c − b).<br />

6.11. Notation. –––– We write L :=É( 3 √<br />

a3 /a 0 , 3 √<br />

a2 /a 0 , 3 √<br />

a1 /a 0 , ζ 3 )/É) <strong>and</strong><br />

K :=É( 3 √<br />

a2 /a 0 , 3 √<br />

a1 /a 0 , ζ 3 ). Then, G := Gal(L/É) is the Galois group<br />

acting on the 27 lines.<br />

We have the subgroup G 1 := Gal(L/K ) which is isomorphic to/3by consequence<br />

of assumption (∗). We choose a generator σ of G 1 .<br />

Finally, we write τ for the involution ζ 3 ↦→ ζ3 2 fixing the three third roots.<br />

Then, 〈σ, τ〉 ∼ = S 3 .<br />

6.12. Proposition. –––– The restriction<br />

is not the zero map.<br />

H 1 (G, Pic(XÉ)) −→ H 1 (〈σ〉, Pic(XÉ))


132 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

Proof. The assumption (∗) guarantees that [É( 3 √<br />

a3 /a 0 , ζ 3 ) :É] = 6. The adjunctions<br />

of 3 √<br />

a1 /a 0 <strong>and</strong> 3 √<br />

a2 /a 0 might or might not lead to further cubic extensions.<br />

Thus, there will be three cases to distinguish, #G may be either 6,<br />

or 18, or 54.<br />

We take for S the free group generated by the 27 lines <strong>and</strong> use the canonical<br />

isomorphisms H 1 (G, Pic(XÉ)) ∼ = Hom((NG S ∩ S 0 )/N G S 0 ,É/) <strong>and</strong><br />

H 1 (〈σ〉, Pic(XÉ)) ∼ = Hom((N 〈σ〉 S ∩ S 0 )/N 〈σ〉 S 0 ,É/). The restriction is then<br />

induced by the norm map N G/〈σ〉 : N 〈σ〉 S → N G S.<br />

We will start with the divisor D := div ( (T 0 + (a 1 /a 0 ) 1 3 T1 )/(T 0 + (a 2 /a 0 ) 1 3 T2 ) )<br />

which is the norm of L 2 (0, 0) − L 1 (0, 0). We have to show that N G/〈σ〉 D is not<br />

the norm N G of a principal divisor.<br />

First case: #G = 54.<br />

This is the generic case. We find that<br />

N G/〈σ〉 D = 6 div ( (a 0 T 3<br />

0 + a 1 T 3<br />

1 )/(a 0 T 3<br />

0 + a 2 T 3<br />

2 ) )<br />

= 6 ∑<br />

(m,n)∈(/3) 2<br />

L 2 (m, n) − 6<br />

∑<br />

(m,n)∈(/3) 2<br />

L 1 (m, n) .<br />

On the other h<strong>and</strong>, the group S 0 of all principal divisors in S is generated by the<br />

pairwise differences of all triangles [Man,кVI,ÄÑѺ].<br />

We have<br />

N G ∆ 1 kn = 18 ∑L k (i, j) ,<br />

(i,j)∈(/3) 2<br />

N G ∆km 2 = 18 ∑L k (i, j) ,<br />

(i,j)∈(/3) 2<br />

N G ∆ 3 abc = 6<br />

∑<br />

(i,j)∈(/3) 2<br />

L 0 (i, j) + 6<br />

∑<br />

(i,j)∈(/3) 2<br />

L 1 (i, j) + 6<br />

<strong>and</strong> see that N G/〈σ〉 D is not generated by these elements.<br />

Second case: #G = 18.<br />

∑<br />

(i,j)∈(/3) 2<br />

L 2 (i, j)<br />

Assume without restriction that a 0 = a 1 <strong>and</strong> a 2 /a 0 is not a cube inÉ∗ .<br />

Then, we find<br />

N G/〈σ〉 D = div ( (T0 3 + T1 3 )6 /(T0 3 + (a 2 /a 0 ) 3 T2 3 )2)<br />

= 6[L 2 (0, 0) + L 2 (0, 1) + L 2 (0, 2)] − 2 ∑L 1 (i, j) .<br />

(i,j)∈(/3) 2<br />

On the other h<strong>and</strong>,<br />

⎧<br />

⎪⎨<br />

N G L k (m, n) =<br />

⎪⎩<br />

2 ∑ L k (i, j) if k ≠ 2 ,<br />

(i,j)∈(/3) 2<br />

6 ∑<br />

j∈/3L k (m, j) if k = 2 .


Sec. 6] THE “FIRST CASE” OF DIAGONAL CUBIC SURFACES 133<br />

Hence,<br />

N G ∆ 1 kn = 6 ∑L k (i, j),<br />

⎧<br />

⎪⎨<br />

N G ∆ 2 km =<br />

⎪⎩<br />

N G ∆ 3 abc = 2<br />

(i,j)∈(/3) 2<br />

6 ∑ L k (i, j) if k ≠ 2 ,<br />

(i,j)∈(/3) 2<br />

18 ∑<br />

j∈/3L k (m, j) if k = 2 ,<br />

∑<br />

(i,j)∈(/3) 2<br />

L 0 (i, j) + 2<br />

∑<br />

(i,j)∈(/3) 2<br />

L 1 (i, j) + 6 ∑ 2 (a, j) .<br />

j∈/3L<br />

Factoring modulo all summ<strong>and</strong>s of type L 2 <strong>and</strong> ignoring the round brackets,<br />

these expressions are 54L 0 , 54L 1 , <strong>and</strong> 18L 0 + 18L 1 . They do not generate<br />

(−18)L 1 .<br />

Third case: #G = 6.<br />

Here, we find that<br />

On the other h<strong>and</strong>,<br />

N G/〈σ〉 D = 2 div ( (T 0 + (a 1 /a 0 ) 1 3 T1 )/(T 0 + (a 2 /a 0 ) 1 3 T2 ) )<br />

= 2 ∑<br />

i∈/3L 2 (0, i) − 2 ∑<br />

i∈/3L 1 (0, i).<br />

N G ∆ 1 kn = 2 ∑<br />

i∈/3L k (0, i) + 2 ∑<br />

i∈/3L k (1, i) + 2 ∑<br />

i∈/3L k (2, i)<br />

N G ∆ 2 km = 6 ∑<br />

i∈/3L k (m, i)<br />

N G ∆ 3 abc = 2 ∑<br />

i∈/3L 0 (b − a, i) + 2 ∑<br />

i∈/3L 1 (b, i) + 2 ∑<br />

i∈/3L 2 (a, i) .<br />

Ignoring the round brackets, those are 18L k for k ∈ {0, 1, 2}<strong>and</strong> 6L 0 +6L 1 +6L 2<br />

which do not generate 6L 2 − 6L 1 .<br />

□<br />

6.13. Corollary. –––– The restriction map<br />

H 1 (G, Pic(XÉ)) −→ H 1 (〈σ, τ〉, Pic(XÉ))<br />

is an isomorphism. The <strong>groups</strong> are isomorphic to/3.<br />

Proof. H 1 (〈σ〉, Pic(XÉ)) is purely 3-torsion. Sir P. Swinnerton Dyer’s list<br />

(cf. II.8.24) shows it is either/3or (/3) 2 . Further, we know that the<br />

restriction H 1 (G, Pic(XÉ)) → H 1 (〈σ〉, Pic(XÉ)) is not the zero map. Thus, the<br />

list implies that H 1 (G, Pic(XÉ)), too, is nothing but/3or (/3) 2 .


134 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

The restriction H 1 (G, Pic(XÉ)) → H 1 (〈σ〉, Pic(XÉ)) factors via<br />

H 1 (G, Pic(XÉ)) → H 1 (〈σ, τ〉, Pic(XÉ)).<br />

Thus, the latter is not the zero map, either. Again, a view on the list makes sure<br />

that H 1 (〈σ, τ〉, Pic(XÉ)) may be only/3or (/3) 2 .<br />

At this point, recall that (/3) 2 occurs only when the group acting nontrivially<br />

on Pic(XÉ) is of order three. Since τ ∈ 〈σ, τ〉 ⊂ G is of order two <strong>and</strong><br />

acting non-trivially, we are not in this case.<br />

The assertion follows.<br />

6.14. Proposition. –––– The restriction map<br />

H 1 (G, Pic(XÉ)) −→ H 1 (〈σ〉, Pic(XÉ)) τ<br />

is an isomorphism. The <strong>groups</strong> are isomorphic to/3.<br />

Proof. It suffices to show that the restriction map<br />

H 1 (〈σ, τ〉, Pic(XÉ)) −→ H 1 (〈σ〉, Pic(XÉ)) τ<br />

is an isomorphism. We know, the group on the left h<strong>and</strong> side is isomorphic<br />

to/3. The group on the right h<strong>and</strong> side is clearly a 3-torsion group.<br />

We consider the inflation-restriction spectral sequence<br />

E p,q<br />

2<br />

:= H p (〈τ〉, H q (〈σ〉, Pic(XÉ))) =⇒ H n (〈σ, τ〉, Pic(XÉ))<br />

<strong>and</strong> obtain the following exact sequence of terms of lower order,<br />

0 −→ H 1 (〈τ〉, Pic(XÉ) σ ) −→ H 1 (〈σ, τ〉, Pic(XÉ)) −→<br />

−→ H 1 (〈σ〉, Pic(XÉ)) τ −→ H 2 (〈τ〉, Pic(XÉ) σ ) .<br />

Since the homomorphism considered is encircled by 2-torsion <strong>groups</strong>, the proof<br />

is complete.<br />

→É/<br />

□<br />

iii. Some observations. —<br />

6.15. Lemma. –––– The image of the Manin map<br />

ev: Br(X ) × X (É) is contained in 1 3/.<br />

□<br />

Proof. By Proposition 2.3.b.i) <strong>and</strong> iii), the Manin map is additive in the first<br />

variable <strong>and</strong> factors via Br(X )/π ∗ Br(K ) ∼ =/3.<br />


Sec. 6] THE “FIRST CASE” OF DIAGONAL CUBIC SURFACES 135<br />

6.16. Lemma. –––– Let α ∈ Br(X ) be any Brauer class <strong>and</strong> (C, x 0 ) ⊂ XÉp 0<br />

be a<br />

smooth elliptic curve.<br />

Then, one of the following two statements is true.<br />

i) α| C ∈ Br(C )/π ∗ Br(SpecÉ) is zero. Then, ev νp0 (α, ·) is constant on C (Ép 0<br />

).<br />

ii) α| C ∈ Br(C )/π ∗ Br(SpecÉ) is non-zero.<br />

Then, there is a surjective group homomorphism g : C (Ép 0<br />

) → 1 3/such that<br />

for every x ∈ C (Ép 0<br />

).<br />

ev νp0 (α, x) = g(x) + ev νp0 (α, x 0 )<br />

Proof. i) is clear.<br />

ii) follows directly from Theorem II.8.16.<br />

□<br />

6.17. Lemma. –––– Let p 0 ≠ 3 be a prime number, (C, O) an elliptic curve<br />

overÉp 0<br />

3/<br />

, <strong>and</strong> C a minimal model of C overp 0<br />

. Assume that there are no<br />

Ép 0<br />

-valued <strong>points</strong> on C with singular reduction.<br />

Then, every continuous group homomorphism<br />

g : C (Ép 0<br />

) −→ 1<br />

factors via the reduction map C (Ép 0<br />

) → C (p 0<br />

) modulo p 0 .<br />

Proof. Continuity means that g factors via the reduction C (Ép 0<br />

) → C (/p0)<br />

n<br />

for a certain n ∈Æ. The/p0-valued n <strong>points</strong> on C reducing to the neutral<br />

element form a p 0 -group.<br />

□<br />

6.18. Proposition. –––– Suppose that p 0 ≠ 3. Then, for α ∈ Br(X ), the local<br />

evaluation map ev νp0 (α, ·) factors via<br />

red: X (Ép 0<br />

) −→ E p (p 0<br />

)<br />

(x 0 : x 1 : x 2 : x 3 ) ↦→ ((x 0 mod p 0 ) : (x 1 mod p 0 ) : (x 2 mod p 0 )) .<br />

Proof. Let (x 0 : x 1 : x 2 : x 3 ) ∈ X (Ép 0<br />

). We may assume without restriction that<br />

x 0 , x 1 , x 2 , x 3 ∈p 0<br />

are coprime. Then, assumption (∗) implies that x 0 , x 1 , or x 2<br />

is a unit. Assume, again without restriction, that x 2 is a unit. Then, x 0 <strong>and</strong> x 1<br />

can not both be multiples of p 0 . Assume p 0 ∤x 1 .<br />

Hensel’s lemma ensures that there exists a unique t ∈p 0<br />

such that<br />

t ≡ x 2 (mod p 0 ) <strong>and</strong> (x 0 : x 1 : t : 0) ∈ X (Ép 0<br />

). We claim that<br />

ev νp0 (α, (x 0 : x 1 : x 2 : x 3 )) = ev νp0 (α, (x 0 : x 1 : t : 0)) .<br />

Proof of the claim: Both <strong>points</strong> are contained in the intersection of X with the<br />

hyperplane “x 1 T 0 − x 0 T 1 = 0”. This is a genus one curve C. It is given by the


136 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

equation<br />

[a 0 (x 0 /x 1 ) 3 + a 1 ]T 3<br />

1 + a 2 T 3<br />

2 + a 3 T 3<br />

3 = 0.<br />

The assumptions imply that the first coefficient is not divisible by p 0 .<br />

As p 0 |a 3 , the curve C has bad reduction at p 0 . The equation above defines a<br />

minimal model C ofC overp. Thereare noÉp 0<br />

-valued <strong>points</strong> with singular reduction.<br />

C p0<br />

3/<br />

consists of three lines meeting in the singular point (0 : 0 : 0 : 1).<br />

It may happen that all three lines are defined overp 0<br />

or that one line is defined<br />

overp 0<br />

<strong>and</strong> the other two overp 2 <strong>and</strong> conjugate to each other.<br />

We choose a basepoint x ∈ C (Ép 0<br />

). Then, by Lemma 6.15,<br />

ev νp0 (α, ·)| C (Ép 0<br />

) : C (Ép 0<br />

) −→ 1<br />

differs from a continuous homomorphism of <strong>groups</strong> just by a constant summ<strong>and</strong>.<br />

Lemma 6.17 shows that ev νp0 (α, ·)| C (Ép 0<br />

) factors via the reduction<br />

map C (Ép 0<br />

) → C reg<br />

p 0<br />

(p 0<br />

).<br />

If only one line of C p0 is defined overp 0<br />

then we have #C reg<br />

p 0<br />

(p 0<br />

) = p 0 .<br />

Every group homomorphism to 3/is 1 constant which implies the claim in<br />

this case.<br />

Otherwise, we have #C reg<br />

p 0<br />

(p 0<br />

) = 3p 0 , each line having exactly p 0 smooth <strong>points</strong>.<br />

Let x ′ be the third point of intersection of the line tangent to C in x with C.<br />

Then, the group structure on C (Ép 0<br />

) has the property that P+Q+R = 2·x+x ′<br />

if <strong>and</strong> only if P, Q, <strong>and</strong> R are collinear. In particular, P + Q is the third point<br />

on the line through R <strong>and</strong> the neutral element.<br />

The reductions of three collinear <strong>points</strong> are either belonging to the same line<br />

of C p0 or to three different lines. By consequence, the subgroup of C reg<br />

p 0<br />

(p 0<br />

) of<br />

order p 0 is provided by the line containing the neutral element. The three lines<br />

form its cosets.<br />

As the reductions of (x 0 : x 1 : x 2 : x 3 ) <strong>and</strong> (x 0 : x 1 : t : 0) are on the same line,<br />

this implies the claim.<br />

Now, we apply Theorem II.8.16 <strong>and</strong> Lemma 6.17 to the elliptic curve E given<br />

by T 3 = 0. This shows<br />

ev νp0<br />

(<br />

α, (x0 : x 1 : t : 0) ) = g ( (x 0 mod p 0 ) : (x 1 mod p 0 ) : (t mod p 0 ) )<br />

for a map g : E p0 (p 0<br />

) → 3/. 1 Since (t mod p 0 ) = (x 2 mod p 0 ), the proof<br />

is complete.<br />

□<br />

6.19. Remark. –––– In the case p 0 ≠ 3, the only assertion still to be proven is<br />

that ev νp0 (α, ·)| E(Ép 0<br />

) is non-constant. For that, according to Theorem II.8.16,<br />

it suffices to show that α| E ∈ Br(E)/π ∗ Br(SpecÉ) is non-zero. We will verify<br />

this in the next subsection in Proposition 6.24.


Sec. 6] THE “FIRST CASE” OF DIAGONAL CUBIC SURFACES 137<br />

The next lemma, although true in general, will be relevant for us only in the<br />

case p 0 = 3.<br />

6.20. Lemma. –––– Let p be an arbitrary prime number <strong>and</strong> α ∈ Br(X ) be any<br />

Brauer class. Then, there exists a positive integer n such that ev νp (α, x) depends only<br />

on the reduction x (pn )<br />

∈ X (/p n).<br />

Proof. This follows directly from the continuity of ev νp with respect to the<br />

right argument.<br />

□<br />

iv. Evaluating on the elliptic curve. —<br />

6.21. Notation. –––– For the fields K <strong>and</strong> L as in 6.11, we fix a valuation ν of K<br />

lying above ν p0 <strong>and</strong> an extension w of ν to L.<br />

6.22. Lemma. –––– Fix an isomorphism ι: H 2 (〈σ〉,) ∼ =/3. Under the<br />

periodicity isomorphism<br />

H 1 (〈σ〉, Pic(X Kν<br />

))<br />

∼=<br />

−→ Ĥ −1 (〈σ〉, Pic(X Kν<br />

))<br />

= (N Div(X Lw ) ∩ Div 0 (X Kν ))/N Div 0 (X Lw )<br />

induced by ι, a representative of a generator is given by<br />

Proof. In fact,<br />

f := (T 0 + (a 1 /a 0 ) 1 3 T1 )/(T 0 + (a 2 /a 0 ) 1 3 T2 ) .<br />

div( f ) = L 2 (0, 0) + L 2 (0, 1) + L 2 (0, 2) − L 1 (0, 0) − L 1 (0, 1) − L 2 (0, 2)<br />

= N (L 2 (0, 0) − L 1 (0, 0)) .<br />

Further, the norms of the principal divisors are generated by the pairwise differences<br />

of<br />

N ∆ 1 kn = ∑<br />

i∈/3L k (0, i) + ∑<br />

i∈/3L k (1, i) + ∑<br />

i∈/3L k (2, i) ,<br />

N ∆ 2 km = 3 ∑<br />

i∈/3L k (m, i) ,<br />

N ∆ 3 abc = ∑ 0 (b − a, i) + ∑ 1 (b, i) + ∑ 2 (a, i) .<br />

i∈/3L i∈/3L i∈/3L Ignoring the round brackets, these elements are 9L k for k ∈ {0, 1, 2}<br />

<strong>and</strong> 3L 0 + 3L 1 + 3L 2 . They do not generate 3L 2 − 3L 1 .<br />

□<br />

6.23. Remark. –––– Theperiodicityisomorphism iscompatiblewiththe action<br />

of the involution τ. As f is τ-invariant, it represents a non-zero cohomology<br />

class in H 1 (〈σ〉, Pic(X Kν<br />

)) τ .


138 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

6.24. Proposition. –––– Assume p 0 ≠ 3. Then, the pull-back<br />

is not the zero map.<br />

Br(X )/π ∗ Br(SpecÉ) −→ Br(EÉp 0<br />

)/π ∗ Br(SpecÉp 0<br />

)<br />

Proof. It suffices to show that the restriction<br />

Br(X )/π ∗ Br(SpecÉ) −→ Br(E Kν )/π ∗ Br(SpecK ν )<br />

is not the zero map.<br />

It factors via [Br(X Kν )/π ∗ Br(SpecK ν )] τ ∼ = H 1 (〈σ〉, Pic(X Kν<br />

)) τ for which,<br />

by Lemma 6.22, we have a generator f in explicit form.<br />

It is, therefore, sufficient to show that there exists a point x ∈ E(K ν ) such<br />

that f (x) ∈ K ∗ ν is not in the image of the norm map N : L ∗ w → K ∗ ν.<br />

As L w /K ν is totally ramified, this follows directly from the lemma below.<br />

□<br />

6.25. Lemma. –––– Let p ≠ 3 be a prime number, K be a finite field extension<br />

ofÉp, <strong>and</strong> L/K a totally ramified Galois extension of degree three. Further, let C<br />

be the genus one curve over K given by x 3 + y 3 + z 3 = 0.<br />

Then, the range of the <strong>rational</strong> function f := (x + y)/(x + z) on C (K ) contains<br />

an element which is not a norm under N : L ∗ → K ∗ .<br />

Proof. The residue field k := O K /m K is finite <strong>and</strong> #k ≡ 1 (mod 3). We have<br />

to show that the range of the <strong>rational</strong> function f = (x + y)/(x + z) on C (k)<br />

contains a non-cube in k ∗ .<br />

For this, we write P := ((−1) : 1 : 0) <strong>and</strong> Q := ((−1) : 0 : 1).<br />

Then, div f = 3(P) − 3(Q). Choosing an arbitrary point O ∈ C (k) as the<br />

neutral element fixes a group law on C (k). We have the 3-division point P − Q<br />

on C.<br />

Consider the isogeny<br />

i : C ′ ·3<br />

:= C/〈P − Q〉 −→ C.<br />

i corresponds to the field extension k(C ′ ) = k(C ) ( √ )<br />

3 f /k(C ). We claim that<br />

→<br />

i ( C ′ (k) ) ∪ {P, Q} = {x ∈ C (k) | f (x) ∈ (k ∗ ) 3 } ∪ {P, Q} .<br />

Indeed, togiveapointx ∈ C (k)isequivalenttogivingavaluationν x : k(C )<br />

such that, for the corresponding residue field, one has O x /m x<br />

∼ = k. x belongs to<br />

the image of i(k) if <strong>and</strong> only if this valuation splits completely in k(C ′ ).<br />

Thus, our task is to show that i(k): C ′ (k) → C (k) is not surjective. Note that<br />

C (k) contains at least the nine <strong>points</strong> given by (1 : (−ζ3 i ) : 0) <strong>and</strong> permutations.


Sec. 6] THE “FIRST CASE” OF DIAGONAL CUBIC SURFACES 139<br />

The exact sequence of Gal(k/k)-modules<br />

induces a long exact sequence<br />

0 −→ µ 3 −→ C ′ (k) −→ C (k) −→ 0<br />

C ′ (k) −→ C (k) −→ k ∗ /(k ∗ ) 3 −→ 0 .<br />

Observe, by Hasse’s bound, every genus one curve has a point over k. Thus,<br />

there are no non-trivial torsors over C ′ <strong>and</strong> we have H 1( Gal(k/k), C ′ (k) ) = 0.<br />

The assertion follows.<br />

□<br />

v. Completing the proof for p 0 = 3. —<br />

6.26. Observation. –––– We have K ν =É3(ζ 3 ) or K ν =É3(ζ 3 , 3 √<br />

2).<br />

√ 3√<br />

Proof. By construction, K ν =É3(ζ 3 , 3 a, b) for a, b units inÉ∗ 3 . Modulo third<br />

powers, units inÉ∗ √ 3 fall into √ three classes, represented by 1, 2, <strong>and</strong> 4, respectively.<br />

AsÉ3(ζ 3 , 3 2) =É3(ζ 3 , 3 4), the assertion follows.<br />

□<br />

6.27. Notation. –––– For F ∈3, we denote the genus one curve, given by the<br />

equation T 3 = FT 0 on X, by E F . In particular, E 0 = E in our previous notation.<br />

The projection of E F to the first three coordinates is “Ax 3 0 + a 1 x 3 1 + a 2 x 3 2 = 0”<br />

for A := a 0 + F 3 a 3 . The assumptions 3 ∤a 0 <strong>and</strong> 3|a 3 make sure that A is a<br />

3-adic unit.<br />

We write E F for the model of E F over Spec3 given by T 3 = FT 0 on X .<br />

6.28. Proposition. –––– For every F ∈3, the pull-back<br />

is not the zero map.<br />

Br(X )/π ∗ Br(SpecÉ) −→ Br(E FÉ3 )/π∗ Br(SpecÉ3)<br />

Proof. It suffices to show that the restriction<br />

is not the zero map.<br />

Br(X )/π ∗ Br(SpecÉ) −→ Br(E F K ν<br />

)/π ∗ Br(SpecK ν )<br />

It factors via [Br(X Kν )/π ∗ Br(SpecK ν )] τ ∼ = H 1 (〈σ〉, Pic(X Kν<br />

)) τ for which,<br />

by Lemma 6.22, we have a generator [ f ] in explicit form.<br />

It is, therefore, sufficient to show that there exists a point x ∈ E F (K ν ) such that<br />

f (x) ∈ Kν ∗ is not in the image of the norm map N : L ∗ w → Kν. ∗ Since a 2 /a 1 is a<br />

cube in Kν ∗ , this follows directly from the lemma below.<br />


140 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

6.29. Lemma. –––– Let A ∈ O Kν be different from zero <strong>and</strong> C be the genus one<br />

curve over K ν given by Ax 3 + y 3 + z 3 = 0.<br />

Then, the range of the <strong>rational</strong> function f := (x + y)/(x + z) on C (K ν ) contains<br />

an element which is not a norm under N : L ∗ w → K ∗ ν.<br />

Proof. First step: Description of the norms.<br />

Let π be a uniformizing element of K ν . Then, according to [Se62, Chap. V,<br />

Proposition 5], there exists a positive integer t such that the homomorphism<br />

N n : (1 + m n L w<br />

)/(1 + m n+1<br />

L w<br />

) −→ ( 1 + (π) n) / ( 1 + (π) n+1)<br />

is a bijection for n < t <strong>and</strong> three-to-one for n = t. In particular, there is a<br />

unit α ∈ O ∗ K ν<br />

such that every element<br />

is not a norm.<br />

For t, there is the formula<br />

w ≡ 1 + απ t (mod π t+1 )<br />

t = ν L(D Lw /K ν<br />

)<br />

2<br />

where D Lw /K ν<br />

denotes the different of L w /K ν .<br />

Second step: Calculation of the different.<br />

Case 1: K ν =É3(ζ 3 , 3 √<br />

2).<br />

Then, L w =É3(ζ 3 , 3 √<br />

2,<br />

3√<br />

3). For the discriminants, one calculates<br />

d Lw /É3 = (3)37 <strong>and</strong> d Kν /É3 = (3)7 . This results in ν L (D Lw /K ν<br />

) = 16 <strong>and</strong> t = 7.<br />

For comparison, ν K (3) = 6.<br />

Case 2: K ν =É3(ζ 3 ).<br />

− 1<br />

Then, L w =É3(ζ 3 , 3 √<br />

3),É3(ζ 3 , 3 √<br />

6), orÉ3(ζ 3 , 3 √<br />

12).<br />

For each possibility the discriminant is the same, d Lw /É3 = (3)11 . Together with<br />

d Kν /É3 = (3), this shows ν L(D Lw /K ν<br />

) = 8 <strong>and</strong> t = 3.<br />

For comparison, ν K (3) = 2.<br />

Third step: Construction of the point.<br />

We claim that, on C, there is a K-valued point (x : y : z) such that x = π t ,<br />

y ≡ −1/α (mod π), <strong>and</strong> z ≡ 1/α (mod π). Then, the assertion follows since<br />

is not a norm.<br />

x + y<br />

x + z ≡ πt − 1/α<br />

π t + 1/α ≡ απt − 1<br />

απ t + 1 ≡ −(1 + απt ) (mod π t+1 )


Sec. 6] THE “FIRST CASE” OF DIAGONAL CUBIC SURFACES 141<br />

To show the existence of the point, we choose y ≡ −1/α (mod π), arbitrarily.<br />

Then, the equation<br />

g(Z) := Z 3 + (y 3 + Aπ 3t ) = 0<br />

has a solution Z ≡ −y (mod π) by Hensel’s lemma since<br />

ν K (g(−y)) = ν K (Aπ 3t ) ≥ 3t ,<br />

ν K (g ′ (−y)) = ν K (3y 2 ) = ν K (3) ,<br />

<strong>and</strong> 3t > 2ν K (3).<br />

□<br />

6.30. Proposition. –––– Assume that a 2 /a 1 ∈É∗ 3 is a non-cube <strong>and</strong> let<br />

t = (t 0 : t 1 : t 2 : t 3 ) ∈ X (3) be a point with t 0 ≠ 0. Choose n ∈Æsuch<br />

that ev νp (α, x) depends only on the reduction x (pn )<br />

∈ X (/p n).<br />

Then, for any F ∈3 such that (F mod 3) = t 3 /t 0 , one has<br />

#{y ∈ E F (/3 n) | y = t, ev ν3 (α, y)=0}<br />

= #{y ∈ E F (/3 n) | y = t, ev ν3 (α, y)=1/3}<br />

= #{y ∈ E F (/3 n) | y = t, ev ν3 (α, y)=2/3} .<br />

Proof. Proposition 6.28 shows that α| E ∈ Br(E FÉ3<br />

FÉ3 )/π∗ Br(SpecÉ3) is different<br />

from zero. Thus, α| E yields a surjective homomorphism of <strong>groups</strong><br />

FÉ3<br />

E F (É3) → 3/. 1 From this, we immediately see<br />

#{y ∈ E F (/3 n) | ev ν3 (α, y) = 0} = #{y ∈ E F (/3 n) | ev ν3 (α, y) = 1/3}<br />

We have to show the same for <strong>points</strong> reducing to t.<br />

= #{y ∈ E F (/3 n) | ev ν3 (α, y) = 2/3} .<br />

A unit u ∈3 is a cube if <strong>and</strong> only if u ≡ ±1 (mod 9). Permuting coordinates<br />

<strong>and</strong> changing a sign, if necessary, we may therefore assume that a 1 ≡ 1 (mod 9)<br />

<strong>and</strong> a 2 ≡ 7 (mod 9).<br />

E F is given by Ax 3 0 + a 1 x 3 1 + a 2 x 3 2 = 0 for A := a 0 + F 3 a 3 . In principle, there<br />

are three cases.<br />

First case: A ≡ ±7 (mod 9).<br />

We may assume A ≡ 7 (mod 9). Then, modulo 3, allÉ3-valued <strong>points</strong> on E F<br />

reduce to (1 : 0 : (−1)). Hence, the assertion is true in this case.<br />

Second case: A ≡ ±4 (mod 9).<br />

This is impossible as 4x 3 0 + 1x 3 1 + 7x 3 2 = 0 allows no solutions inÉ3 3 except<br />

for (0, 0, 0).<br />

Third case: A ≡ ±1 (mod 9).


142 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

Assume without restriction that A ≡ 1 (mod 9). AÉ3-valued point on E F<br />

may reduce either to (1 : (−1) : 0) or to (1 : 1 : 1).<br />

Take (1 : (−a) : 0) ∈ E F (É3), fora thethirdrootofA/a 1 , astheneutralelement.<br />

Then, Lemma 6.31 shows that the reduction map E F (É3) → E F (3) gives rise<br />

to a surjective group homomorphism E F (É3) →/2.<br />

As 2 is prime relative to 3, the asserted equidistribution of ev ν3 (α, ·) holds in<br />

each class, separately.<br />

□<br />

6.31. Lemma. –––– Consider the elliptic curve C overÉ3, given by<br />

Take (1 : (−1) : 0) as its basepoint.<br />

x 3 + y 3 + 7z 3 = 0 .<br />

C has a minimal Weierstraß model such that C 0 (É3), the group of <strong>points</strong> with<br />

non-singular reduction, coincides with the set of the <strong>points</strong> reducing (naively)<br />

to (1 : (−1) : 0). Further, C (É3)/C 0 (É3) ∼ =/2.<br />

Proof. The substitutions<br />

lead to the Weierstraß equation<br />

x ′ := −21z , y ′ := 63<br />

2 (x − y) , z′ := x + y<br />

z ′ y ′2 = x ′3 − 72 3 3<br />

4 z′3 .<br />

If the reduction of (x : y : z) is (1 : 1 : 1) then (x ′ : y ′ : z ′ ) reduces to (0 : 0 : 1).<br />

This is the cusp.<br />

On the other h<strong>and</strong>, consider the case that x = 1, z = 3k, <strong>and</strong>, therefore,<br />

y ≡ −1−7·3 2 k 3 (mod 27). Dividing the substitution formulas by 63, we obtain<br />

x ′ = −k, y ′ ≡ 1+ 7 2 32 k 3 (mod 27), i.e., y ′ ≡ 1 (mod 9), <strong>and</strong>z ′ = −k 3 (mod 3).<br />

In particular, y ′ is always a unit. The reduction of (x ′ : y ′ : z ′ ) is never equal to<br />

the cusp.<br />

It remains to show that C (É3)/C 0 (É3) ∼ =/2<strong>and</strong> that the Weierstraß model<br />

found is minimal. For this, we follow Tate’s algorithm [Ta75].<br />

We have the affine equation<br />

Y 2 = X 3 − 72 3 3<br />

4 .<br />

According to [Ta75, Summary], we are in case 6). The polynomial P is given<br />

by P(T ) = T 3 − 72 . It has a triple zero modulo 3.<br />

4<br />

We observe that 72<br />

≡ 1 (mod 9) <strong>and</strong> writeu for the 3-adic unit such 4 thatu3 = 72<br />

. 4<br />

The substitution X ′ := X − 3u leads to Y 2 = X ′3 + 9uX ′2 + 27u 2 X ′ .


Sec. 6] THE “FIRST CASE” OF DIAGONAL CUBIC SURFACES 143<br />

J. Tate’s equation (8.1) becomes<br />

Y 2 2 = 9X 3 2 + 9uX 2 2 + 3u 2 X 2 .<br />

Here, Y =: 9Y 2 <strong>and</strong> X ′ =: 9X 2 . Equation (9.1) is<br />

3Y 2 3 = 3X 3 2 + 3uX 2 2 + u 2 X 2<br />

for Y 2 =: 3Y 3 . As u 2 is a unit, we see that the Weierstraß model is minimal <strong>and</strong><br />

of Kodaira type III ∗ . [Ta75, table on p. 46] indicates C (É3)/C 0 (É3) ∼ =/2.<br />

□<br />

6.32. Proposition. –––– Let X be any diagonal cubic surface fulfilling (∗).<br />

Choose n ∈Æsuch that ev νp (α, x) depends only on the reduction<br />

x (pn )<br />

∈ X (/p n).<br />

Then, for every t = (t 0 : t 1 : t 2 : t 3 ) ∈ X (3), we have<br />

#{y ∈ X (/3 n) | y = t, ev ν3 (α, y)=0}<br />

= #{y ∈ X (/3 n) | y = t, ev ν3 (α, y)=1/3}<br />

= #{y ∈ X (/3 n) | y = t, ev ν3 (α, y)=2/3} .<br />

Proof. The case a 0 ≡ a 1 ≡ a 2 (mod 9) is treated in Example 6.33.<br />

We may therefore assume that two of the quotients a 0 /a 1 , a 1 /a 2 , <strong>and</strong> a 2 /a 0 ,<br />

say a 1 /a 2 <strong>and</strong> a 2 /a 0 , are non-cubes inÉ∗ 3 . It is impossible that t 0 = t 1 = 0.<br />

Again without restriction, assume t 0 ≠ 0.<br />

We consider the projection<br />

π : X −→ P 1 ,<br />

(x 0 : x 1 : x 2 : x 3 ) ↦→ (x 3 : x 0 ) .<br />

On aÉ3-valued point x reducing to t, the bi<strong>rational</strong> map π is defined.<br />

We have π(x) = (F : 1) for some F ∈3 such that (F mod 3) = t 3 /t 0 .<br />

By Proposition 6.30, the asserted equality is true in every fiber of π. It is<br />

therefore true in general.<br />

□<br />

6.33. Example. –––– Let d ∈<strong>and</strong> consider the cubic surface overÉgiven by<br />

T0 3 + T1 3 + T2 3 + 3dT3 3 = 0 .<br />

=/3<br />

By the results shown, we have<br />

Br(X )/π ∗ Br(SpecÉ) = H 1( Gal(L/K ), Pic(X L ) ) ∼


144 AN APPLICATION: THE BRAUER-MANIN OBSTRUCTION [Chap. III<br />

for L =É(ζ 3 , 3 √<br />

3d) which is equal to the field of definition of the 27 lines<br />

on X. As K =É(ζ 3 ) does not contain any cubic extensions, the Brauer-<br />

Manin obstruction may be described completely explicitly, without relying on<br />

Lichtenbaum’s duality.<br />

In fact,<br />

H 1( Gal(L/É), Pic(X L ) ) ∼ = H<br />

1 ( 〈σ〉, Pic(X L ) ) τ<br />

= [Br(XÉ(ζ 3 ))/π ∗ Br(SpecÉ(ζ 3 ))] τ .<br />

For the latter, we have the explicit generator [ f ] for f := (T 0 +T 1 )/(T 0 +T 2 ).<br />

This means, for x = (x 0 : x 1 : x 2 : x 3 ) ∈ X (É3(ζ 3 )) <strong>and</strong> ν ∈ Val(É(ζ 3 )) the<br />

extension of ν 3 , we have<br />

(<br />

ev ν ([ f ], x) = i ν (t0 + t 1 )/(t 0 + t 2 ) ) .<br />

Here, i ν is the homomorphism<br />

É3(ζ 3 ) ∗ −→É3(ζ 3 ) ∗ /NL ∗ ∼=<br />

w −→ Ĥ 0 (〈σ〉, L ∗ ∼=<br />

w) −→ H 2 (〈σ〉, L ∗ w) inv ν<br />

−→É/<br />

for w the extension of ν to L.<br />

However, we want to considerÉ3-valued <strong>points</strong>, notÉ3(ζ 3 )-valued ones.<br />

Thus, let α ∈ Br(X ) be a Brauer class mapping to [ f ] under restriction.<br />

Then, for x ∈ X (É3),<br />

2 · ev ν3 (α, x) = ev ν ([ f ], x) .<br />

Fortunately, multiplication by 2 is an automorphism of 1 3/.<br />

On X, two kinds ofÉ3-valued <strong>points</strong> may be distinguished.<br />

First kind: x 0 , x 1 , <strong>and</strong> x 2 are units.<br />

Then, x 0 ≡ x 1 ≡ x 2 (mod 3) <strong>and</strong> d ∤ 3. We have<br />

ev ν3 (α, x) = 0 ⇐⇒ x 1 ≡ x 2 (mod 9) .<br />

Second kind: Among x 0 , x 1 , <strong>and</strong> x 2 , there is an element which is a multiple of 3.<br />

Suppose 3|x 0 . Then, x 1 ≡ −x 2 (mod 3). We have<br />

ev ν3 (α, x) = 0 ⇐⇒ 2x 0 + x 1 + x 2 (mod 9) .<br />

6.34. Corollary. –––– If x 3 0 +x 3 1 +x 3 2+3x 3 3 = 0 for x 0 , . . . , x 3 ∈then we have<br />

the following non-trivial congruences.<br />

i) If x 0 , x 1 , <strong>and</strong> x 2 are units then<br />

x 0 ≡ x 1 ≡ x 2 (mod 9) .


Sec. 6] THE “FIRST CASE” OF DIAGONAL CUBIC SURFACES 145<br />

ii) If 3|x 0 then<br />

x 0 ≡ 0 (mod 9) .<br />

Proof. This is the particular case d = 1. Then, L =É(ζ 3 , 3 √<br />

3) is unramified at<br />

every prime different from 3. As ν p<br />

(<br />

(t0 + t 1 )/(t 0 + t 2 ) ) is always divisible by 3,<br />

we have ev νp (α, x) = 0 for every p ≠ 3.<br />

The congruences obtained from ev ν3 (α, x) = 0 imply the congruences asserted.<br />

□<br />

6.35. Remark. –––– Thus, we have shown that, for the equation<br />

T 3<br />

0 + T 3<br />

1 + T 3<br />

2 + 3T 3<br />

3 = 0<br />

overÉ, the restrictions coming from the Brauer-Manin obstruction are exactly<br />

the congruences obtained by R. Heath-Brown [H-B92a] in 1993.<br />

Note that Heath-Brown used rather different methods. His main tool is the law<br />

of cubic reciprocity.


PART B<br />

HEIGHTS


CHAPTER IV<br />

THE CONCEPT OF A HEIGHT<br />

Equations are just the boring part of mathematics. I attempt to see things<br />

in terms of geometry.<br />

STEPHEN HAWKING (A Biography (2005) by Kristine Larsen, p. 43)<br />

1. The naive height on the projective space overÉ<br />

1.1. –––– Heights have been studied by number theorists for a very long time.<br />

A height is a function measuring the size or, more precisely, the arithmetic complexity<br />

of certain objects. These objects are classically solutions of Diophantine<br />

equations or <strong>rational</strong> <strong>points</strong> on an algebraic variety. A height then might answer<br />

the question how many bits one would need in order to store the solution or<br />

the point on a computer.<br />

More recently, starting with G. Faltings’ ideas of heights on moduli spaces, it<br />

became more common to consider heights for more complicated objects such<br />

as cycles.<br />

1.2. Definition. –––– For (x 0 : . . . : x n ) ∈ P n (É), put<br />

Here,<br />

H naive (x 0 : . . . : x n ) := max<br />

i=0,...,n |x′ i| .<br />

(x ′ 0 : . . . : x′ n ) = (x 0 : . . . : x n )<br />

such that all x ′ i are integers <strong>and</strong> gcd(x′ 0 , . . . , x′ n ) = 1.<br />

The function H naive : P nÉ(É) →Êis called the naive height.<br />

1.3. Fact (Fundamental finiteness). —– For every C ∈Ê, there are only finitely<br />

many <strong>points</strong> x ∈ P nÉ(É) such that<br />

H naive (x) < C .<br />

Proof. For eachcomponentof (x 0, ′ . . . , x n), ′ wehave −C < x i ′ < C. Thus, there<br />

are only finitely many choices.<br />


150 THE CONCEPT OF A HEIGHT [Chap. IV<br />

1.4. –––– The naive height is probably the simplest function one might think<br />

of which fulfills the fundamental finiteness property. For more general height<br />

functions, the fundamental finiteness property will always be required.<br />

1.5. Remark. –––– Let X ⊂ P nÉbe a subvariety. Then, everyÉ-<strong>rational</strong> point<br />

on X is also aÉ-<strong>rational</strong> point on P n .<br />

We call the restriction of H naive to X (É) the naive height on X.<br />

Is is obvious that the naive height on X fulfills the fundamental finiteness property.<br />

1.6. Notation. –––– i) For a prime number p, denote by ‖ . ‖ p the normalized<br />

p-adic valuation.<br />

I.e., for x ∈É\{0}, let x = ±2 v 2 ·3<br />

v<br />

v pk<br />

3 · . . . · p<br />

k<br />

prime factors. Then, put<br />

‖x‖ p := p −v p<br />

.<br />

Further, ‖0‖ p := 0.<br />

be its decomposition into<br />

ii) We use ‖ . ‖ ∞ as an alternative notation for the usual absolute value,<br />

{ x if x ≥ 0 ,<br />

‖x‖ ∞ :=<br />

−x if x < 0 .<br />

1.7. Fact. –––– For p a prime number or infinity, ‖ . ‖ p is indeed a valuation.<br />

This means, for x, y ∈É,<br />

i) ‖x‖ p ≥ 0,<br />

ii) ‖x‖ p = 0 if <strong>and</strong> only if x = 0,<br />

iii) ‖xy‖ p = ‖x‖ p · ‖y‖ p ,<br />

iv) ‖x + y‖ p ≤ ‖x‖ p + ‖y‖ p .<br />

For p ≠ ∞, one even has that ‖x + y‖ p ≤ max{‖x‖ p , ‖y‖ p }.<br />

1.8. Fact (Product formula). —– For x ∈É\{0}, one has<br />

∏ ‖x‖ p = 1 .<br />

p prime or ∞<br />

1.9. Lemma. –––– Let (x 0 : . . . : x n ) ∈ P nÉ(É). Then,<br />

[<br />

H naive (x 0 : . . . : x n ) =<br />

∏<br />

p prime or ∞<br />

]<br />

max ‖x i‖ p<br />

i=0,...,n<br />

Proof. The product formula implies that the right h<strong>and</strong> side remains unchanged<br />

when (x 0 : . . . : x n ) is replaced by (λx 0 : . . . : λx n ) for λ ≠ 0. Thus, we may<br />

suppose that all x i are integers <strong>and</strong> gcd(x 0 , . . . , x n ) = 1.<br />

.


Sec. 2] GENERALIZATION TO NUMBER FIELDS 151<br />

The assumptions that x 0 , . . . , x n ∈<strong>and</strong> gcd(x 0 , . . . , x n ) = 1 imply<br />

max<br />

i=0,...,n ‖x i‖ p = 1<br />

for every prime number p. Hence, the formula on the right h<strong>and</strong> side may be<br />

simplified to max i=0,...,n |x i |.<br />

This is precisely the assertion.<br />

1.10. Remark. –––– Despite being so primitive, the naive height is actually<br />

sufficient for most applications. For example, in the numerical experiments<br />

described in Part C, we will always work with the naive height.<br />

□<br />

2. Generalization to number fields<br />

i. The definition. —<br />

2.1. –––– Let K be a number field. I.e., K is a finite extension ofÉ. It is<br />

well known from algebraic number theory [Cas67] that there is a set Val(K ) of<br />

normalized valuations ‖ . ‖ ν on K satisfying the following conditions.<br />

a) The functions ‖ . ‖ ν : K →Êare indeed valuations. I.e., for x, y ∈É,<br />

i) ‖x‖ ν ≥ 0,<br />

ii) ‖x‖ ν = 0 if <strong>and</strong> only if x = 0,<br />

iii) ‖xy‖ ν = ‖x‖ ν · ‖y‖ ν ,<br />

iv) ‖x + y‖ ν ≤ ‖x‖ ν + ‖y‖ ν .<br />

b) There is the product formula<br />

∏ ‖x‖ ν = 1<br />

∈Æ<br />

ν∈Val(K )<br />

for every x ∈ K \{0}.<br />

2.2. –––– Further, for L/K a degree d extension of number fields, the sets<br />

Val(K ) <strong>and</strong> Val(L) are compatible in the following sense.<br />

i) For every ‖ . ‖ µ ∈ Val(L), there are a valuation ‖ . ‖ ν ∈ Val(K ) <strong>and</strong> d µ<br />

such that<br />

‖ . ‖ µ | K = ‖ . ‖ d µ<br />

ν .<br />

In this case, it is said that ‖ . ‖ µ is lying above ‖ . ‖ ν .<br />

ii) For every ‖ . ‖ ν ∈ Val(K ), there are only a finite number of valuations<br />

‖ . ‖ µ1 , . . . , ‖ . ‖ µl ∈ Val(L) lying above ‖ . ‖ ν . One has<br />

l<br />

∑<br />

i=1<br />

d µi = d .


152 THE CONCEPT OF A HEIGHT [Chap. IV<br />

This implies<br />

for every x ∈ K.<br />

‖x‖ d ν =<br />

l<br />

∏<br />

i=1<br />

‖x‖ µi<br />

2.3. –––– A valuation is called archimedean if it lies above the valuation ‖ . ‖ ∞<br />

ofÉ. Otherwise, it is called non-archimedean.<br />

If ν is non-archimedean then one has the ultrametric triangle inequality<br />

‖x + y‖ ν ≤ max{‖x‖ ν , ‖y‖ ν } .<br />

2.4. Definition. –––– Let K be a number field of degree d. Then, for<br />

(x 0 : . . . : x n ) ∈ P n (K ), one puts<br />

[ ] 1<br />

H naive (x 0 : . . . : x n ) := ∏ max ‖x d<br />

i‖ ν .<br />

i=0,...,n<br />

ν∈Val(K )<br />

This height function is the number field version of the naive height on P n . It is<br />

usually called the absolute height.<br />

2.5. Lemma. –––– Let K be a number field <strong>and</strong> (x 0 : . . . : x n ) ∈ P n (K ). Further,<br />

let L ⊃ K be a finite extension.<br />

Then, the absolute height H naive (x 0 : . . . : x n ) remains unchanged when<br />

(x 0 : . . . : x n ) is considered as an L-<strong>rational</strong> point.<br />

Proof. Put d ′ := [L : K ]. Then, by the properties of the valuations, we have<br />

[<br />

H naive (x 0 : . . . : x n ) = ∏<br />

ν∈Val(K )<br />

= ∏<br />

ν∈Val(K )<br />

= ∏<br />

µ∈Val(L)<br />

] 1<br />

max ‖x d<br />

i‖ ν<br />

i=0,...,n<br />

[<br />

∏<br />

µ∈Val(L)<br />

µ above ν<br />

[<br />

] 1<br />

max ‖x dd<br />

i‖ ′<br />

µ<br />

i=0,...,n<br />

] 1<br />

max ‖x dd<br />

i‖ ′<br />

µ<br />

i=0,...,n<br />

which is exactly the formula for H naive (x 0 : . . . : x n ) considered as an<br />

L-<strong>rational</strong> point.<br />

□<br />

2.6. Proposition (Northcott). —– Let B, D ∈Ê. Then,<br />

is a finite set.<br />

{x ∈ P n (É) | x ∈ P n (K ) for [K :É] < D <strong>and</strong> H naive (x) < B}<br />

Proof. We may work with the number fields K of a fixed degree d.


Sec. 2] GENERALIZATION TO NUMBER FIELDS 153<br />

For x, we choose homogeneous coordinates such that some coordinate equals 1.<br />

Then, it is clear that, for every valuation ‖ . ‖ ν <strong>and</strong> every index i, we have<br />

max{‖x 0 ‖ ν , . . . , ‖x n ‖ ν } ≥ max{1, ‖x i ‖ ν } .<br />

Multiplying over all ν <strong>and</strong> taking the d-th root therefore shows<br />

Hence, it suffices to verify that the set<br />

H naive (x 0 : . . . : x n ) ≥ H naive (1 : x i ) .<br />

{(1 : x) ∈ P 1 (É) | (1 : x) ∈ P 1 (K ) for [K :É] = d <strong>and</strong> H naive (1 : x) < B}<br />

is finite.<br />

For this, wewriteσ 1 (x), . . . , σ d (x) ∈Éfor theelementarysymmetricfunctions<br />

in the conjugates of x. According to Vieta, x is a zero of the polynomial<br />

F x (T ) := T d − σ 1 (x)T d−1 + σ 2 (x)T d−2 + . . . + (−1) d σ d (x) ∈É[T ] .<br />

Lemma 2.7 shows that<br />

(<br />

H naive 1 : σr (x) ) ( d<br />

)<br />

( d<br />

)<br />

≤ · H naive (1 : x) rd ≤ · B rd .<br />

r r<br />

Thus, by Fact 1.3, we know that for each σ r (x) there are only finitely many possibilities.<br />

Hence, there are only finitely many possibilities for the polynomial F x<br />

<strong>and</strong>, therefore, only finitely many possibilities for x.<br />

This completes the proof.<br />

2.7. Lemma. –––– Let K be number field of degree d which is Galois overÉ.<br />

For x ∈ K, denote by σ 1 (x), . . . , σ d (x) ∈Éthe elementary symmetric functions<br />

in the conjugates of x. Then,<br />

(<br />

H naive 1 : σr (x) ) ( d<br />

)<br />

≤ · H naive (1 : x) rd .<br />

r<br />

Proof. We denote the conjugates of x by x 1 , . . . , x d . Let ν be any valuation<br />

of K. Then,<br />

∥ ∥∥∥∥ν<br />

‖σ r (x)‖ ν =<br />

∥<br />

∑ x i1 · . . . · x ir<br />

1≤i 1


154 THE CONCEPT OF A HEIGHT [Chap. IV<br />

By consequence,<br />

max{1, ‖σ r (x)‖ ν } ≤ C (d, r, ν) · max max{1, ‖x i ‖ ν } r<br />

i<br />

≤ C (d, r, ν) · ∏ max{1, ‖x i ‖ ν } r .<br />

i<br />

Multiplying over all valuations of K yields<br />

(<br />

H naive 1 : σr (x) )d ( d<br />

) d<br />

≤ ·<br />

r<br />

∏ H naive (1 : x i ) rd<br />

i<br />

( d<br />

) d<br />

= · Hnaive (1 : x) rd2<br />

r<br />

since conjugate <strong>points</strong> have the same height. This inequality is equivalent to<br />

the assertion.<br />

□<br />

2.8. –––– For most theoretical investigations <strong>and</strong>, may be, some practical ones,<br />

it is actually better to work with the logarithm of the absolute height.<br />

2.9. Definition. –––– For (x 0 : . . . : x n ) ∈ P n (É), put<br />

h naive (x 0 : . . . : x n ) := log H naive (x 0 : . . . : x n ) .<br />

The height function h naive is called the logarithmic height.<br />

ii. Application: The height defined by an ample invertible sheaf. —<br />

2.10. Definition. –––– Let X be a projective variety over a number field K<br />

<strong>and</strong> L be an ample invertible sheaf on X.<br />

Then, a height function h L induced by L is given as follows. Let m ∈Æbe such<br />

that L ⊗m is very ample. Put<br />

h L (x) := 1 m h (<br />

naive iL ⊗m(x) )<br />

for x ∈ P a closed point. Here, i L ⊗m : P → P N denotes a closed embedding<br />

defined by L ⊗m .<br />

2.11. Lemma. –––– Let X be a projective variety over a number field K <strong>and</strong> L<br />

be an ample invertible sheaf on X. If h (1)<br />

L <strong>and</strong> h(2) L are two height functions induced<br />

by L then there is a constant C such that<br />

for every x ∈ X (É).<br />

| h (1)<br />

L (x) − h(2) L (x)| ≤ C


Sec. 3] GEOMETRIC INTERPRETATION 155<br />

Proof. First, it is clear that the k-uple embedding fulfills<br />

h naive<br />

(<br />

ρk (x) ) = k·h naive (x)<br />

for every k ∈Æ<strong>and</strong> x ∈ P N (É). Hence, if h L is defined using L ⊗m then the<br />

same height function may be defined using L ⊗km .<br />

We may therefore assume that h (1)<br />

L<br />

<strong>and</strong> h(2) L<br />

are defined using the same tensor<br />

L<br />

differ by an automor-<br />

⊗m<br />

power of L . Then, the two embeddings i (1)<br />

L<br />

<strong>and</strong> i (2)<br />

⊗m<br />

phism ι: P N → P N .<br />

We have to verify that<br />

| h naive<br />

(<br />

ι(x)<br />

) − hnaive (x)| ≤ C<br />

for every x ∈ P N (É). ι is explicitly given by<br />

(x 0 : . . . : x N ) ↦→ (x ′ 0 : . . . : x ′ N )<br />

for some a = (a ij ) ij ∈ M N +1 (K ).<br />

:= (a 00 x 0 + . . .+ a 0N x N ) : . . . : (a N 0 x 0 + . . .+ a NN x N )<br />

For all but finitely many ν ∈ Val(K ), we have that ν is non-archimedean<br />

<strong>and</strong> ‖a ij ‖ ν = 1 for all i <strong>and</strong> j. This yields<br />

max<br />

i=0,...,N ‖x′ i‖ w ≤ max ‖x i‖ w<br />

i=0,...,N<br />

for every number field L <strong>and</strong> every w ∈ Val(L) lying above such a valuation.<br />

For every other valuation of K, we find a constant D ν such that<br />

The desired inequality<br />

follows immediately from this.<br />

max<br />

i=0,...,N ‖x′ i‖ w ≤ D d w/ν<br />

ν<br />

h naive<br />

(<br />

ι(x)<br />

) − hnaive (x) ≤<br />

· max<br />

i=0,...,N ‖x i‖ w .<br />

∑<br />

ν∈Val(K )<br />

log D ν<br />

For the inequality the other way round, one may work with ι −1 instead of ι.<br />

□<br />

3. Geometric interpretation<br />

3.1. –––– Heights are more closely related to modern algebraic geometry than<br />

this might seem from the definitions given in the sections above. The geometric<br />

interpretation of the concept of a height is the starting point of arithmetic<br />

intersection theory, a fascinating theory which we may only touch upon here.


156 THE CONCEPT OF A HEIGHT [Chap. IV<br />

The reader is advised to consult the articles [G/S 90, G/S 92] of H. Gillet<br />

<strong>and</strong> C. Soulé, the textbook [S/A/B/K], <strong>and</strong> the references therein. More recent<br />

developments are described by J. I. Burgos Gil, J. Kramer, <strong>and</strong> U. Kühn<br />

in [B/K/K].<br />

The particular case of the arithmetic intersection theory on a curve over a<br />

number field had been developed earlier. The articles of S. Yu. Arakelov [Ara]<br />

<strong>and</strong> G. Faltings [Fa84] present the point of view taken before around 1990 which<br />

is a bit different from today’s.<br />

3.2. –––– We return to the assumption that the ground field is K =É. This is<br />

done mainly in order to ease notation. The theory would work equally well<br />

over an arbitrary number field.<br />

3.3. Definition. –––– An arithmetic variety is an integral scheme which is projective<br />

<strong>and</strong> flat over Spec.<br />

3.4. Definition. –––– A hermitian line bundle on an arithmetic variety X is a<br />

pair (L , ‖ . ‖) consisting of an invertible sheaf L ∈ Pic(X ) <strong>and</strong> a continuous<br />

hermitian metric on the line bundle Lassociated with L on the complex<br />

space X ().<br />

All hermitian line bundles on X form an abelian group which is denoted<br />

by ̂Pic C 0 (X ).<br />

For the group of all smooth (C ∞ ) hermitian line bundles, one writes ̂Pic(X ).<br />

̂Pic(X ) is usually called the arithmetic Picard group.<br />

3.5. Example. –––– The projective space P nover Specis an arithmetic variety.<br />

On P n, there is the tautological invertible sheaf O(1). It is generated by<br />

the global sections X 0 , . . . , X n ∈ Γ(P n, O(1)).<br />

Let us show two explicit hermitian metrics on O(1)making O(1) to a hermitian<br />

line bundle.<br />

i) The Fubini-Study metric is given by<br />

for l ∈ Γ(P n, O(1)).<br />

‖l‖ FS :=<br />

√<br />

(X0<br />

ii) The minimum metric is given by<br />

‖l‖ min := min<br />

l<br />

1<br />

) 2<br />

+ . . . + ( X n<br />

) 2<br />

l<br />

i=0,...,n<br />

∣ l ∣∣∣<br />

∣ .<br />

X i


Sec. 3] GEOMETRIC INTERPRETATION 157<br />

In the case of the Fubini-Study metric, the formula should be interpreted in<br />

such a way that ‖l‖ FS (x) = 0 at all <strong>points</strong> x ∈ P n () where l vanishes. Similarly,<br />

in the definition of ‖l‖ min (x), the minimum is actually to be taken over<br />

all i such that x i ≠ 0.<br />

The Fubini-Study metric is smooth (C ∞ ). The minimum metric is continuous<br />

but not even C 1 .<br />

3.6. Definition (Arakelov degree). —– For a hermitian line bundle (L , ‖ . ‖)<br />

on Spec, define its Arakelov degree by<br />

̂deg (L , ‖ . ‖) := log #(L /sL ) − log ‖s‖<br />

for s ∈ Γ(Spec, L ) a non-zero section.<br />

3.7. Remark. –––– L is associated with a free-module L of rank one. Lis<br />

the-vector space L⊗. Thus, we work with a non-zero element s ∈ L <strong>and</strong><br />

with the norm ‖s ⊗ 1‖.<br />

The definition is independent of the choice of s since<br />

#(L /nsL ) = n·#(L /sL ) ,<br />

therefore log #(L /nsL ) = log #(L /sL ) + log n.<br />

log ‖ns‖ = log ‖s‖ + logn.<br />

On the other h<strong>and</strong>,<br />

3.8. Fact. –––– Let L be a hermitian line bundle on Spec. If there is a section<br />

s ∈ Γ(Spec, L ) of norm less than one then ̂deg L > 0.<br />

Proof. This follows immediately from the definition.<br />

3.9. Fact. –––– For two hermitian line bundles L 1 , L 2 on Spec, one has<br />

̂deg (L 1 ⊗ L 2 ) = ̂deg L 1 + ̂deg L 2 .<br />

Proof. Apply the definition to arbitrary non-zero sections s 1 ∈ Γ(Spec, L 1 ),<br />

s 2 ∈ Γ(Spec, L 2 ), <strong>and</strong> to their tensor product s 1 ⊗s 2 ∈ Γ(Spec, L 1 ⊗L 2 ). □<br />

3.10. –––– AÉ-<strong>rational</strong> point on the projective space is actually a morphism<br />

x : SpecÉ→P nof schemes. The valuative criterion for properness implies<br />

that there is a unique extension to a morphism from Spec,<br />

x<br />

SpecÉ n P<br />

<br />

<br />

∩ <br />

<br />

<br />

x<br />

Spec.<br />


158 THE CONCEPT OF A HEIGHT [Chap. IV<br />

3.11. Observation (Arakelov). —– Let x ∈ P n (É) <strong>and</strong> x : Spec→P nits<br />

extension to Spec. Then,<br />

h naive (x) = ̂deg ( x ∗ (O(1), ‖ . ‖ min ) ) .<br />

Proof. Write x in coordinates as (x 0 : . . . : x n ) for x 0 , . . . , x n ∈. Then, the<br />

pull-back of X i ∈ Γ(P n, O(1)) is x i ∈ Γ(Spec, x ∗ (O(1))). Since O(1) is generated<br />

by the global sections X i , we have<br />

x ∗ (O(1)) = 〈x 0 , . . . , x n 〉 = gcd(x 0 , . . . , x n ) ·.<br />

We choose an index i 0 such that x i0 ≠ 0. Using this section, we obtain<br />

̂deg (x ∗ (O(1)), ‖ . ‖) = log # ( x ∗ (O(1))/x i0 x ∗ (O(1)) ) − log ‖x i0 ‖<br />

which is exactly the logarithmic height of x.<br />

= log # ( gcd(x 0 , . . . , x n )/x i0) − log min<br />

∣ = log<br />

x i0<br />

∣gcd(x 0 , . . . , x n ) ∣ + log max<br />

x i ∣∣∣<br />

i=0,...,n<br />

∣x i0<br />

max |x i|<br />

i=0,...,n<br />

= log<br />

gcd(x 0 , . . . , x n )<br />

i=0,...,n<br />

∣ x i0 ∣∣∣<br />

∣ x i<br />

□<br />

3.12. –––– This motivates the following general definition.<br />

Definition. Let X be an arithmetic variety <strong>and</strong> L be a hermitian line bundle<br />

on X .<br />

Then, the height with respect to L of theÉ-<strong>rational</strong> point x ∈ X (É) is given by<br />

h L<br />

(x) = ̂deg ( x ∗ L ) .<br />

Here, x : Spec→X is the extension of x to Spec.<br />

3.13. Examples. –––– i) Let X = P n<strong>and</strong> let L be the invertible sheaf O(1)<br />

equipped with the minimum metric. Then, as shown in 3.11, the height with<br />

respect to L is the naive height.<br />

ii) Let X = P n<strong>and</strong> let L be the invertible sheaf O(1) equipped with the<br />

Fubini-Study metric. Then, the height with respect to L is the l 2 -height which<br />

is given by<br />

√<br />

h l 2(x 0 : . . . : x n ) := log x<br />

0 ′2 + . . . + x′2 n<br />

for<br />

(x ′ 0 : . . . : x′ n ) = (x 0 : . . . : x n )


Sec. 3] GEOMETRIC INTERPRETATION 159<br />

projective coordinates such that all x ′ i are integers <strong>and</strong> gcd(x ′ 0, . . . , x ′ n) = 1.<br />

3.14. Lemma. –––– Let f : X → Y be a morphism of arithmetic varieties <strong>and</strong><br />

L be a hermitian line bundle on Y .<br />

Then, for every x ∈ X (É),<br />

h f ∗ L (x) = h L ( f (x)) .<br />

Proof. Let x be the extension of x to Spec. Then, f ◦x is the extension of f (x)<br />

to Spec. Thus,<br />

h f ∗ L (x) = ̂deg ( x ∗ ( f ∗ L ) ) = ̂deg ( ( f ◦x) ∗ L ) = h L<br />

( f (x)) .<br />

□<br />

3.15. Proposition. –––– Let X be an arithmetic variety.<br />

a) Let ‖ . ‖ 1 <strong>and</strong> ‖ . ‖ 2 be hermitian metrics on one <strong>and</strong> the same invertible sheaf L .<br />

Then, there is a constant C such that<br />

for every x ∈ X (É).<br />

|h (L ,‖ .‖1 )(x) − h (L ,‖ .‖2 )(x)| < C<br />

b) Let L 1 <strong>and</strong> L 2 be two hermitian line bundles. Then, for every x ∈ X (É),<br />

h (L1 ⊗L 2 ) (x) = h L 1<br />

(x) + h L2<br />

(x) .<br />

c) LetL beahermitianlinebundlesuchthattheunderlyinginvertiblesheafisample.<br />

Then, for every C ∈Ê, there are only finitely many <strong>points</strong> x ∈ X (É) such that<br />

h L<br />

(x) < C .<br />

Proof. a) For x the extension of x to Spec, we have<br />

h (L ,‖ .‖1 )(x) − h (L ,‖ .‖2 )(x) = ̂deg ( x ∗ L , x ∗ ‖ . ‖ 1<br />

) − ̂deg<br />

(<br />

x ∗ L , x ∗ ‖ . ‖ 2<br />

)<br />

.<br />

Working with a non-zero section s ∈ Γ(Spec, x ∗ L ), we see that the latter difference<br />

is equal to<br />

log #(x ∗ L /sx ∗ L ) − log ‖s‖ 1 (x) − ( log #(x ∗ L /sx ∗ L ) − log ‖s‖ 2 (x) )<br />

= log ‖ . ‖ 2(x)<br />

‖ . ‖ 1 (x) .<br />

Since X () is compact, there is a positive constant D such that<br />

1<br />

D < ‖ . ‖ 2(x)<br />

‖ . ‖ 1 (x) < D<br />

for every x ∈ X (). This implies the claim.


160 THE CONCEPT OF A HEIGHT [Chap. IV<br />

b) Clearly, for every x ∈ X (É), one has<br />

h (L1 ⊗L 2 ) (x) = ̂deg ( x ∗ (L 1 ⊗L 2 ) ) = ̂deg ( (x ∗ L 1 )⊗(x ∗ L 2 ) )<br />

= ̂deg (x ∗ L 1 ) + ̂deg (x ∗ L 2 ) = h L1<br />

(x) + h L2<br />

(x) .<br />

c) There is some n ∈Æsuch that L ⊗n is very ample. Part b) shows that<br />

it suffices to verify the assertion for L ⊗n . Thus, we may assume that L is<br />

very ample.<br />

Let i : X ֒→ P n be the closed embedding induced by L . Then, L = i ∗ O(1).<br />

It follows from Tietze’s Theorem that there is a hermitian metric on O(1) such<br />

that L = i ∗ O(1). Then, h L<br />

(x) = h O(1)<br />

(i(x)). It, therefore, suffices to show<br />

fundamental finiteness for the height function h O(1)<br />

on P n (É).<br />

Part a) together with Observation 3.11 shows that h O(1)<br />

differs from h naive by a<br />

bounded summ<strong>and</strong>. Fact 1.3 yields the assertion.<br />

□<br />

4. Basic arithmetic intersection theory<br />

4.1. Remark. –––– It should be noticed that there is a strong formal analogy of<br />

the concept of a height on an arithmetic variety to the concept of a degree in<br />

algebraic geometry over a ground field. The only obvious difference is that the<br />

role of the sections of an invertible sheaf is now played by small sections, say, of<br />

norm less than one. Nevertheless, it seems that the height of a point is actually<br />

some sort of arithmetic intersection number.<br />

This is an idea which has been formalized first by S. Yu. Arakelov [Ara] for twodimensional<br />

arithmetic varieties <strong>and</strong> later by H. Gillet <strong>and</strong> C. Soulé [G/S 90]<br />

for arithmetic varieties of arbitrary dimension.<br />

Let us briefly recall the most important definitions <strong>and</strong> some fundamental results<br />

from arithmetic intersection theory.<br />

4.2. Definition (Gillet-Soulé). —– Let X be a regular arithmetic variety<br />

<strong>and</strong> i ∈Æ.<br />

i) Then, an arithmetic cycle of codimension i on X is a pair (Z, g Z ) consisting a<br />

codimensioni cycleZ ontheschemeX together withaGreen’scurrent g Z forZ.<br />

The latter means that g Z ∈ D p−1,p−1 (X ()) is a real current on the complex<br />

manifold X () invariant under the complex conjugation F ∞ : X () → X ()<br />

such that<br />

ω Z := 1<br />

2πi ∂∂g Z + δ Z<br />

is a smooth differential form on X (). Here, δ Z is the current given by<br />

integration along Z() ⊆ X ().


Sec. 4] BASIC ARITHMETIC INTERSECTION THEORY 161<br />

The abelian group of all codimension i arithmetic cycles on X is denoted<br />

by Ẑi (X ).<br />

ii) A codimension i arithmetic cycle on X is said to be linearly equivalent to<br />

zero if it belongs to the subgroup of Ẑi (X ) generated by all arithmetic cycles<br />

of types<br />

(0, ∂u)<br />

for u ∈ D p−2,p−1 (X ()),<br />

(0, ∂v)<br />

for v ∈ D p−1,p−2 (X ()), <strong>and</strong><br />

(div f , − log | f | 2 )<br />

for f a non-zero <strong>rational</strong> function on an integral subscheme of X of codimension<br />

i − 1.<br />

The abelian group of all codimension i arithmetic cycles on X which are<br />

linearly equivalent to zero is denoted by ̂R i (X ).<br />

iii) The i-th arithmetic Chow group of X is defined by<br />

ĈH i (X ) := Ẑi (X )/̂R i (X ) .<br />

=<br />

4.3. Remark. –––– − log | f | 2 is indeed a Green’s current for div f . This fact is<br />

classically known as the Poincaré-Lelong equation [G/H, p. 388].<br />

4.4. Example. –––– For every regular arithmetic variety X , we have<br />

ĈH 0 (X ) = CH 0 (X )<br />

since, on X (), there are no currents of type (−1, −1).<br />

4.5. Example-Definition. –––– The arithmetic degree<br />

is an isomorphism.<br />

̂deg : ĈH1 (Spec) −→Ê<br />

[ (∑<br />

i<br />

a i (p i ), r )] ↦→ 1 2 r + ∑a i log p i<br />

i<br />

Indeed, the homomorphism ̂deg is well-defined since (div n, − log |n| 2 ) is<br />

mapped to zero. It is injective <strong>and</strong> surjective as one has CH 1 (Spec) = 0<br />

for the usual Chow group <strong>and</strong> [D 0,0 (pt)] − =Ê.


162 THE CONCEPT OF A HEIGHT [Chap. IV<br />

4.6. Example. –––– Obviously,<br />

for i ≥ 2.<br />

ĈH i (Spec) = CH i (Spec) = 0<br />

4.7. Definition. –––– Let X be a regular arithmetic variety <strong>and</strong> L ∈ ̂Pic(X )<br />

be a hermitian line bundle. Then, the first arithmetic Chernclass of L is given by<br />

for a non-zero section s of L .<br />

ĉ 1 (L ) := [(div s, − log ‖s‖ 2 )] ∈ ĈH1 (X )<br />

4.8. Lemma. –––– For every regular arithmetic variety X , the first arithmetic<br />

Chern class provides an isomorphism<br />

ĉ 1 : ̂Pic(X ) −→ ĈH1 (X ) .<br />

Proof. It follows form the Poincaré-Lelong equation that the homomorphism<br />

ĉ 1 is well-defined.<br />

To show injectivity, assume [(div s, − log ‖s‖ 2 )] = 0 ∈ ĈH1 (X ) for a non-zero<br />

<strong>rational</strong> section s of a hermitian line bundle L . Since there are no (0, 0)-<br />

currents of types ∂u or ∂v, this means (div s, − log ‖s‖ 2 ) = (div f , − log ‖ f ‖ 2 )<br />

for a <strong>rational</strong> function f .<br />

Hence, s/f is a global section of L without zeroes or poles <strong>and</strong> of constant<br />

norm one. This implies that L represents the neutral element of ̂Pic(X ).<br />

For surjectivity, let any arithmetic 1-cycle (Z, g Z ) be given. The invertible<br />

sheaf O(Z) may be equipped with a smooth hermitian metric. This reduces the<br />

question to the case Z = 0.<br />

This means, we consider an arithmetic 1-cycle (0, g). I.e., g is a (0, 0)-current<br />

such that ω := ∂∂g is a smooth (1, 1)-form. We claim this implies already that<br />

g is smooth.<br />

Indeed, the ∂∂-lemma [G/H, p. 149] yields a smooth function g ′ such<br />

that ω = ∂∂g ′ . Hence, ∂∂(g ′ − g) = 0. This means that the current g ′ − g<br />

vanishes on all test forms of type ∂∂η. By the ∂∂-lemma, these are all d-closed<br />

forms of top degree. By consequence of Poincaré duality for de Rham cohomology,<br />

g ′ − g is the current associated with a constant function.<br />

For the arithmetic 1-cycle (0, g), we therefore have (0, g) = ĉ 1 (O X ) where O X<br />

is the structure sheaf equipped with the smooth hermitian metric given<br />

by ‖ f ‖ := e − g 2 | f |.<br />

□<br />

4.9. Remark. –––– For a hermitian line bundle L ∈ ̂Pic(Spec), its Arakelov<br />

degree as defined in the previous section is the same as ̂deg ( ĉ 1 (L ) ) .


)⊗É<br />

Sec. 4] BASIC ARITHMETIC INTERSECTION THEORY 163<br />

4.10. –––– H. Gillet <strong>and</strong> C. Soulé [G/S 90, Theorem 4.2.3] define an intersection<br />

product<br />

·: ĈHi (X ) × ĈHj (X ) −→ ĈHi+j (X<br />

with the following properties.<br />

i) “·” is commutative <strong>and</strong> associative.<br />

ii) If Y <strong>and</strong> Z are irreducible subvarieties meeting properly, i.e.,<br />

codim(Y ∩ Z) ≥ codim Y + codim Z ,<br />

then<br />

[(Y, g Y )] · [(Z, g Z )] = [Y ·Z, g Y ∗ g Z ] .<br />

Here, Y ·Z is the usual intersection cycle defined by the Tor-formula [Fu84,<br />

Section 20.4]. The ∗-product of two Green’s currents is morally given by<br />

the formula<br />

g Y ∗ g Z := g Y ·δ Z + ω Y ·g Z .<br />

This definition is obviously problematic as g Y ·δ Z is formally a product of<br />

two currents. It may be justified in the expected manner when g Y is a Green’s<br />

form of log type along Y [G/S 90, 2.1.3]. The latter is always fulfilled if Y is a<br />

cycle of codimension one.<br />

4.11. –––– For a morphism f : X → Y of regular arithmetic varieties,<br />

H. Gillet <strong>and</strong> C. Soulé [G/S 90, 4.4.3] define a pull-back homomorphism<br />

It has the properties below.<br />

f ∗ : ĈHi (Y ) −→ ĈHi (X ) .<br />

i) For a smooth hermitian line bundle L ∈ ̂Pic(Y ), one has<br />

ii) For y ∈ ĈHi (Y ) <strong>and</strong> z ∈ ĈHj (Y ),<br />

f ∗ ĉ 1 (L ) = ĉ 1 ( f ∗ L ) .<br />

f ∗ (y·z) = f ∗ y · f ∗ z .<br />

iii) For two morphisms f : X → Y <strong>and</strong> g : Y → Z of regular arithmetic<br />

varieties, the equality<br />

is true.<br />

(g ◦ f ) ∗ = f ∗ ◦ g ∗ : ĈHi (Z ) −→ ĈHi (X )


164 THE CONCEPT OF A HEIGHT [Chap. IV<br />

4.12. –––– For a morphism f : X → Y of regular arithmetic varieties such<br />

that fÉ: XÉ→YÉis smooth, there is a push-forward homomorphism<br />

with the following properties.<br />

i) For [(Z, g Z )] ∈ ĈHi (X ), one has<br />

f ∗ : ĈHi (X ) −→ ĈHi−dimX+dim Y (Y )<br />

f ∗ [(Z, g Z )] = [( f ∗ Z, f ∗ g Z )] .<br />

Here, f ∗ Z is the usual push-forward cycle [Fu84, Section 1.4] <strong>and</strong> f ∗ g Z is the<br />

push-forward of g Z as a current.<br />

ii) There is the projection formula<br />

for x ∈ ĈHi (X ) <strong>and</strong> y ∈ ĈHj (Y ).<br />

f ∗ (x · f ∗ y) = f ∗ x · y<br />

iii) For two morphisms f : X → Y <strong>and</strong> g : Y → Z of regular arithmetic<br />

varieties such that fÉ<strong>and</strong> gÉare smooth, one has that<br />

(g ◦ f ) ∗ = g ∗ ◦ f ∗ : ĈHi (X ) −→ ĈHi−dimX+dimZ (Z ) .<br />

4.13. Remarks. –––– It requires rather hard work to establish all the results<br />

described in 4.10 through 4.12. There are two different sorts of problems.<br />

i) On the scheme-theoretic side, it is already a problem to define the intersection<br />

of two cycles a ∈ CH i (X ) <strong>and</strong> b ∈ CH j (X ) in the ordinary Chow <strong>groups</strong>.<br />

The point is that on a scheme of finite type over, there is no moving lemma<br />

available except for divisors. Thus, in order to define the intersection product<br />

of two cycles, a highly indirect approach must be chosen.<br />

H. Gillet <strong>and</strong> C. Soulé use the fact that the Brown-Gersten-Quillen spectral<br />

sequence degenerates after tensoring withÉ. The Chow <strong>groups</strong> of X appear<br />

as part of the E 2 -terms while the spectral sequence converges versus the algebraic<br />

K-<strong>groups</strong> K ∗ (X ). The product structure on K ∗ (X ) provided by the<br />

tensor product of vector bundles then induces the desired intersection product<br />

·: CH i (X ) × CH j (X ) −→ CH i+j (X )⊗É.<br />

Many more details are given in [S/A/B/K, Chapter I].<br />

Today, there is a second approach to this intersection product which is based<br />

on the resolution of singularities by alterations found by A. J. de Jong [dJo].<br />

This method is rather indirect, too. It yields the same intersection product as<br />

the K-theory approach.


Sec. 5] AN INTERSECTION PRODUCT ON SINGULAR ARITHMETIC VARIETIES 165<br />

One would prefer to have an intersection product<br />

·: CH i (X ) × CH j (X ) −→ CH i+j (X ) .<br />

In other words, one would like to get rid of the tensor product withÉ.<br />

For schemes over, there is no such intersection product with reasonable<br />

properties known up to now.<br />

ii) On the other h<strong>and</strong>, it requires enormous technical efforts to justify the<br />

definitions made on the complex-analytic side. H. Gillet <strong>and</strong> C. Soulé show that<br />

for every arithmetic cycle there is an equivalent one which is of the form (Y, g Y )<br />

for g Y a Green’s form of log type along Y. This is necessary to give the ∗-product<br />

of Green’s currents a precise meaning.<br />

Still, a lot of work has to be done in order to verify all the properties of<br />

the intersection product which are asserted. Here, the hardest point is to<br />

prove associativity.<br />

4.14. –––– Over the years, a lot of progress has been made on this part of<br />

the theory. Several people tried to define the arithmetic intersection product in<br />

a more canonical manner. It turned out that the Gillet-Soulé intersection theory<br />

may be understood as a particular case of a more general framework. From such<br />

a more general point of view, the ∗-product of two Green’s currents appears as<br />

being induced by the cup product on Deligne-Beilinson cohomology. Such an<br />

approach therefore reduces commutativity <strong>and</strong> associativity of the intersection<br />

product into formalities.<br />

The clearest <strong>and</strong> most thorough framework for arithmetic intersection theory<br />

on regular arithmetic varieties seen up to now has been provided by J. I. Burgos<br />

Gil, J. Kramer, <strong>and</strong> U. Kühn [B/K/K].<br />

4.15. Remark. –––– One of the main drawbacks of the arithmetic intersection<br />

theory developed so far is that the push-forward is defined only for morphisms<br />

which are smooth on the generic fiber. Replacing in Definition 4.2 the requirement<br />

that ω Z has to be smooth by a slightly weaker condition, one obtains<br />

a theory which is free of this drawback. Such theories are due to J. I. Burgos<br />

Gil [Bu] <strong>and</strong> A. Moriwaki [Mori]. The reader may find a description<br />

in [B/K/K, Theorem 6.35].<br />

5. An intersection product on singular arithmetic varieties<br />

5.1. –––– For arbitrary, i.e., possibly singular, arithmetic varieties, arithmetic<br />

intersection theory is considerably less far developed. The following theorem is<br />

a consequence from the theory presented so far which is valid in more generality.


166 THE CONCEPT OF A HEIGHT [Chap. IV<br />

5.2. Theorem. –––– a) For every arithmetic variety π : X → Specsuch that its<br />

generic fiber XÉis smooth of dimension d, there is a unique-multilinear map<br />

pGS X : ̂Pic(X ) × . . . × ̂Pic(X ) −→Ê<br />

} {{ }<br />

(d+1) times<br />

satisfying the following conditions.<br />

i) If X is regular then<br />

p X GS (L 1, . . . , L d+1 ) = ̂deg ( π ∗<br />

(ĉ1 (L 1 ) · . . . · ĉ 1 (L d+1 ) ))<br />

for every L 1 , . . . , L d+1 ∈ ̂Pic(X ).<br />

ii) If f : X → Y is a generically finite morphism of arithmetic varieties such that<br />

XÉ<strong>and</strong> YÉare smooth then<br />

p X GS ( f ∗ L 1 , . . . , f ∗ L d+1 ) = deg f · p Y GS (L 1, . . . , L d+1 )<br />

for every L 1 , . . . , L d+1 ∈ ̂Pic(Y ).<br />

b) pGS X is symmetric.<br />

Proof. a) Let π : X → Specbe any arithmetic variety such that its generic<br />

fiber XÉis smooth of dimension d.<br />

Uniqueness: Let r : X ′ → X be a generically finite morphism of arithmetic<br />

varieties such that X ′ is regular. The existence of such a morphism is provided<br />

by the main result of [dJo]. Then, conditions i) <strong>and</strong> ii) imply that, necessarily,<br />

p X GS (L 1, . . . , L d+1 ) := 1<br />

degr ̂deg ( (π◦r) ∗<br />

(ĉ1 (r ∗ L 1 ) · . . . · ĉ 1 (r ∗ L d+1 ) )) .<br />

Existence: We take formula (∗) as the definition for p GS .<br />

One has to show that the definition does not depend on the choice of r.<br />

For this, suppose there are given two generically finite morphisms r ′ : X ′ → X<br />

<strong>and</strong> r ′′ : X ′′ → X such that X ′ <strong>and</strong> X ′′ are regular.<br />

The fiber product X ′ × X X ′′ is generically finite over X . Again by virtue<br />

of [dJo], there is a generically finite morphism X ′′′ → X ′ × X X ′′ such<br />

that X ′′′ is regular. Thus, we are reduced to the case that r ′′ = r ′ ◦ g<br />

for g : X ′′ → X ′ a morphism of regular schemes.<br />

Then, r ′′∗ L i = g ∗ r ′∗ L i for every i. We may now apply the results from<br />

the previous section as g is a morphism of regular arithmetic varieties.<br />

Thus, ĉ 1 (r ′′∗ L i ) = g ∗( ĉ 1 (r ′∗ L i ) ) <strong>and</strong><br />

ĉ 1 (r ′′∗ L 1 ) · . . . · ĉ 1 (r ′′∗ L d+1 ) = g ∗( ĉ 1 (r ′∗ L 1 ) · . . . · ĉ 1 (r ′∗ L d+1 ) ) .<br />

(∗)


Sec. 5] AN INTERSECTION PRODUCT ON SINGULAR ARITHMETIC VARIETIES 167<br />

The projection formula implies<br />

(π◦r ′′ ) ∗<br />

(<br />

g<br />

∗ ( ĉ 1 (r ′∗ L 1 ) · . . . · ĉ 1 (r ′∗ L d+1 ) ))<br />

= (π◦r ′ ◦g) ∗<br />

(<br />

g<br />

∗ ( ĉ 1 (r ′∗ L 1 ) · . . . · ĉ 1 (r ′∗ L d+1 ) ))<br />

= (π◦r ′ ) ∗ g ∗<br />

(<br />

g<br />

∗ ( ĉ 1 (r ′∗ L 1 ) · . . . · ĉ 1 (r ′∗ L d+1 ) ))<br />

= (π◦r ′ ) ∗<br />

[<br />

(g∗ 1) · (ĉ<br />

1 (r ′∗ L 1 ) · . . . · ĉ 1 (r ′∗ L d+1 ) )] .<br />

This yields the claim as g ∗ 1 = deg g.<br />

p X GS is-multilinear by definition.<br />

It fulfills condition i) as one may work with r = id in the case that X is regular.<br />

ii) Choose a generically finite morphism r : X ′ → X such that X ′ is regular.<br />

The definitions of p X GS ( f ∗ L 1 , . . . , f ∗ L d+1 ) <strong>and</strong> p Y GS (L 1, . . . , L d+1 ) using r<br />

<strong>and</strong> f ◦r, respectively, differ only by the degree divided out.<br />

b) This follows from formula (∗) together with the commutativity <strong>and</strong> associativity<br />

of the intersection product of H. Gillet <strong>and</strong> C. Soulé [cf. 4.10.i)]. □<br />

5.3. Definition. –––– Let X be an arithmetic variety <strong>and</strong> L ∈ ̂Pic(X ) be a<br />

hermitian line bundle.<br />

Then, for every cycle Z ∈ Z(XÉ) of dimension i, one defines its height by<br />

h BGS<br />

L<br />

(Z) := pZ GS(L | Z<br />

, . . . , L | } {{ Z<br />

) . }<br />

(i+1) times<br />

Here, Z is the Zariski closure of Z in X equipped with its induced reduced structure.<br />

5.4. Remark. –––– This definition for the height of a cycle of arbitrary<br />

dimension is, in slightly different form, due to J.-B. Bost, H. Gillet,<br />

<strong>and</strong> C. Soulé [B/G/S, Section 3.1.1]. It fulfills a positivity condition <strong>and</strong> a<br />

very strong fundamental finiteness statement.<br />

5.5. Theorem. –––– Let i ∈Æ, X be an arithmetic variety, <strong>and</strong> L ∈ ̂Pic(X ) be<br />

a hermitian line bundle.<br />

a) If the curvature form c 1 (L ) is positive <strong>and</strong> some positive power L ⊗n is generated<br />

by global sections of norm less than one then<br />

for every effective cycle Z on XÉ.<br />

h BGS<br />

L (Z) > 0


168 THE CONCEPT OF A HEIGHT [Chap. IV<br />

b) If L is ample on X then for every B ∈Êthere are only finitely many effective<br />

cycles Z ∈ Z i (XÉ) such that deg L<br />

(Z) < B <strong>and</strong><br />

h BGS<br />

L (Z) < B .<br />

Proof. These are [B/G/S, Proposition 3.2.4 <strong>and</strong> Theorem 3.2.5].<br />

□<br />

5.6. Remark. –––– There are rather few computational results for the Bost-<br />

Gillet-Soulé height h BGS<br />

L for cycles of positive dimension.<br />

One of them is the height of projective space itself. In [B/G/S, Section 3.3], it<br />

is shown that<br />

Z<br />

h (O(1),‖ .‖FS )(P n ) = h (O(1),‖ .‖FS )(P n−1 ) − log ‖x 0 ‖ FS ωFS n .<br />

As the latter integral may easily be computed to 1 2<br />

h (O(1),‖ .‖FS )(P n ) = 1 2<br />

n<br />

∑<br />

k=1<br />

k<br />

∑<br />

m=1<br />

P n ()<br />

n<br />

1<br />

∑<br />

m<br />

m=1<br />

1<br />

m = n + 1<br />

2<br />

n<br />

∑<br />

m=1<br />

, this leads to<br />

1<br />

m − n 2 .<br />

5.7. Definition. –––– Let X be an arithmetic variety, X its generic fiber,<br />

<strong>and</strong> L ∈ ̂Pic(X ) be a hermitian line bundle.<br />

Then, for a closed point x ∈ X, one defines its absolute height by<br />

1 ( )<br />

h L<br />

(x) :=<br />

[L :É] hBGS L (x) .<br />

Here, L is the field of definition of x. (x) denotes the 0-dimensional cycle on X<br />

associated with x.<br />

5.8. Remarks. –––– i) Theorem 5.2.a.ii) implies that this definition is independent<br />

of the choice of L.<br />

ii) Since x ∈ Z 1 (X ) is a one-dimensional cycle, this definition actually does not<br />

involve the arithmetic intersection product. Only the arithmetic degree is used.<br />

For this reason, there is an explicit formula for the dependence of the height on<br />

the choice of the metric. Indeed,<br />

h (L ,‖ .‖ ′ )(x) = h (L ,‖ .‖) (x) − 1<br />

[L :É] ∑<br />

σ : L→log ‖ . ‖′( σ(x) )<br />

‖ . ‖ ( σ(x) ) .<br />

In particular, there is a constant C such that<br />

| h (L ,‖.‖ ′ )(x) − h (L ,‖.‖) (x)| < C<br />

for every x ∈ X. This naturally generalizes Lemma 2.11 <strong>and</strong> Proposition 3.15.


Sec. 6] THE ARITHMETIC HILBERT-SAMUEL FORMULA 169<br />

6. The arithmetic Hilbert-Samuel formula<br />

6.1. Definition. –––– Let X be a Noetherian scheme <strong>and</strong> h: X →Êbe<br />

any function. For i ∈Æ, we define the i-th successive minimum of h by<br />

e i (h) :=<br />

sup<br />

inf<br />

codim X Y=i x∈X \Y<br />

h(x) .<br />

Here, Y runs through all (possibly reducible) closed subsets of X of codimension<br />

i. x runs through the closed <strong>points</strong>.<br />

6.2. Remark. –––– A priori, one has that e i (h) ∈ [−∞, ∞]. Further, directly<br />

from the definition, we see<br />

e 1 (h) ≥ e 2 (h) ≥ . . . ≥ e dim X+1 (h) = inf<br />

x∈X h(x) .<br />

6.3. Lemma. –––– Let X be an arithmetic variety the generic fiber of which is X.<br />

Let L ∈ ̂Pic(X ) <strong>and</strong> suppose that L is ample. Then,<br />

for every i ∈Æ.<br />

e i (h L<br />

) > −∞<br />

Proof. First, we note that it is sufficient to show e dim X+1 (h L<br />

) > −∞.<br />

For this, we may suppose that L is very ample. Then, L is generated by<br />

global sections s 1 , . . . , s k ∈ Γ(X , L ). Further, on L , there exists a hermitian<br />

metric ‖ . ‖ ′ such that c 1 (L , ‖ . ‖ ′ ) is positive. Scale ‖ . ‖ ′ by a positive constant<br />

factor such that ‖s 1 ‖ ′ , . . . , ‖s k ‖ ′ < 1.<br />

Then, by Theorem 5.5.a), h (L ,‖ .‖ ′ )(x) > 0 for every closed point x ∈ X. Remark<br />

5.8.ii) shows that there is a constant C such that<br />

h L<br />

(x) > C<br />

for every x ∈ X. Consequently, e dim X+1 (h L<br />

) > −∞.<br />

□<br />

6.4. Notation. –––– Let X be an arithmetic variety, X its generic fiber,<br />

(L , ‖ . ‖) ∈ ̂Pic(X ), <strong>and</strong> n ∈Æ.<br />

Further, let s ∈ Γ(X, LÉ) ⊗n be any global section. We define its norm by<br />

‖s‖ := sup ‖s‖(x) .<br />

x∈X ()<br />

This induces a norm on theÊ-vector space Γ(X, LÉ)⊗ÉÊ.<br />

⊗n<br />

On Γ(X, LÉ)⊗ÉÊ, ⊗n we consider the normalized Lebesgue measure µ n such<br />

that the unit ball<br />

B n := {s ∈ Γ(X, LÉ)⊗ÉÊ|‖s‖ ⊗n<br />

≤ 1}


É)⊗ÉÊ<br />

170 THE CONCEPT OF A HEIGHT [Chap. IV<br />

is of measure one.<br />

Further, we observe that<br />

Γ(X , L ⊗n ) ⊂ Γ(X, L ⊗n<br />

is a lattice of maximal rank.<br />

6.5. Theorem (Arithmetic Hilbert-Samuel formula). —–<br />

Let X be an arithmetic variety, X its generic fiber, <strong>and</strong> L ∈ ̂Pic(X ).<br />

Putd := dim X. Assume thatX is smooth <strong>and</strong> the curvature formc 1 (L ) is positive.<br />

Then,<br />

− log covol µn Γ(X , L ⊗n ) ∼ pX GS (L , . . . , L ) n d+1 .<br />

(d + 1)!<br />

Proof. When X is regular, this result may be deduced from the arithmetic<br />

Riemann-Roch Theorem proven by H. Gillet <strong>and</strong> C. Soulé [G/S 92].<br />

See also [Fa92]. This, in turn, is a very deep theorem. Its proof combines the<br />

work of several people including, besides H. Gillet <strong>and</strong> C. Soulé, D. Quillen,<br />

J.-M. Bismut, <strong>and</strong> G. Lebeau.<br />

A direct proof for the arithmetic Hilbert-Samuel formula which does not rely<br />

on the arithmetic Riemann-Roch Theorem has been provided by A. Abbes<br />

<strong>and</strong> T. Bouche [A/B]. This proof works for singular arithmetic varieties, too.<br />

□<br />

6.6. Corollary (Theorem of successive minima, cf. [Zh95b, (5.2)]). —–<br />

Let X be an arithmetic variety, X its generic fiber, <strong>and</strong> L ∈ ̂Pic(X ).<br />

Putd := dim X. Assume that X is smooth <strong>and</strong> the curvature formc 1 (L ) is positive.<br />

Then,<br />

e 1 (h L<br />

) ≥ pX GS (L , . . . , L )<br />

(d + 1) deg L<br />

X .<br />

Proof. Write (L , ‖ . ‖) for L .<br />

When one multiplies ‖ . ‖ by a constant factor e −C , the summ<strong>and</strong> C is added<br />

to the height function h L<br />

. Therefore, the term e 1 (h L<br />

) on the left h<strong>and</strong> side is<br />

changed by C, too. On the other h<strong>and</strong>, from Lemma 6.9, we see that the right<br />

h<strong>and</strong> side is changed by the same summ<strong>and</strong>.<br />

Consequently, it will suffice to show that e 1 (h L<br />

) ≥ 0 under the assumption<br />

p X GS (L , . . . , L ) > 0 .<br />

For this, we first observe that the usual Hilbert-Samuel formula yields a constant<br />

D such that<br />

dim Γ(X, L ⊗n É)⊗ÉÊ


Sec. 6] THE ARITHMETIC HILBERT-SAMUEL FORMULA 171<br />

Hence, by the arithmetic Hilbert-Samuel formula,<br />

⊗n dim Γ(X,L<br />

log [2<br />

É)⊗ÉÊ· covol µn Γ(X , L ⊗n )]<br />

= (log 2) dim Γ(X, LÉ)⊗ÉÊ+log ⊗n covol µn Γ(X , L ⊗n )<br />

< D(log 2)n d − 1 2<br />

< 0<br />

p X GS (L , . . . , L )<br />

(d + 1)!<br />

n d+1<br />

for n ≫ 0. Exponentiating, we see<br />

⊗n dim Γ(X,L<br />

2<br />

É)⊗ÉÊ· covol µn Γ(X , L ⊗n ) < 1 = vol B n .<br />

This shows that we may apply Minkowski’s lattice point theorem [B/S,<br />

кII, §4]. The unit ball B n contains a lattice point which is different from<br />

the origin.<br />

In other words, there is a non-zero section s ∈ Γ(X , L ⊗n ) such that ‖s‖ < 1.<br />

We now claim that h L<br />

(x) ≥ 0 for every x ∈ X \div s| X . This clearly suffices for<br />

the asserted inequality.<br />

To verify the claim, let L be the field of definition of the point x <strong>and</strong> x its<br />

Zariski closure in X . Then,<br />

h L<br />

(x) = 1 ( ) 1<br />

[L :É] hBGS L (x) =<br />

[L :É] px GS (L | x) .<br />

The subscheme x might be singular but there is a morphism i : Spec O L → x of<br />

schemes which is generically one-to-one. Theorem 5.2 shows<br />

p x GS(L | x ) = p SpecO L<br />

GS<br />

(i ∗ L ) = ̂deg π ∗ ĉ 1 (i ∗ L )<br />

for π : Spec O L → Specthe structural morphism.<br />

By our assumption, i # s ∈ Γ(Spec O L , i ∗ L ) is a non-zero section. Thus,<br />

ĉ 1 (i ∗ L ) = [(div i # s, − log ‖i # s‖ 2 )] .<br />

Here, div i # s ∈ Z 0 (Spec O L ) is an effective cycle. − log ‖i # s‖ 2 is a function taking<br />

only positive values on the 0-dimensional manifold Hom fields (L,). Hence,<br />

π ∗ ĉ 1 (i ∗ L ) = [(π ∗ div i # s, π ∗ (− log ‖i # s‖ 2 ))] =: [(z, r)]<br />

for an effective cycle z ∈ Z 0 (Spec) <strong>and</strong> r > 0.<br />

It follows directly from Example-Definition 4.5 that ̂deg [(z, r)] > 0.<br />


172 THE CONCEPT OF A HEIGHT [Chap. IV<br />

6.7. Remark. –––– Using the same techniques, S. Zhang [Zh95b, (5.2)] also<br />

shows the inequality<br />

p X GS (L , . . . , L )<br />

deg L<br />

X<br />

6.8. Remark. –––– By definition,<br />

≥ e 1 (h L<br />

) + . . . + e d+1 (h L<br />

) .<br />

p X GS (L , . . . , L ) = hBGS (X ) .<br />

The theorem of successive minima therefore provides a comparison of the height<br />

of the variety X itself with the heights of the <strong>points</strong> on it.<br />

In order to simplify formulas, S. Zhang [Zh95b] normalizes the height of a<br />

closed subvariety Z ⊆ X to be<br />

h L<br />

(Z) :=<br />

h BGS<br />

L (Z)<br />

(d + 1) deg L<br />

X .<br />

This normalization has to be used with some care. The problem is that the<br />

absolute height of a closed point is not the same as the normalized height of the<br />

corresponding 0-cycle.<br />

6.9. Lemma. –––– Let X be an arithmetic variety, X its generic fiber, <strong>and</strong><br />

(L , ‖ . ‖) ∈ ̂Pic(X ). Write d := dim X. Then, for every C > 0,<br />

( ) ( )<br />

(L , C ·‖ . ‖), . . . , (L , C ·‖ . ‖) = p<br />

X<br />

GS (L , ‖ . ‖), . . . , (L , ‖ . ‖)<br />

p X GS<br />

− (d + 1)(log C ) deg L<br />

X .<br />

Proof. By [dJo], we may choose a regular arithmetic variety X ′ <strong>and</strong> a generically<br />

finite morphism f : X ′ → X which surjects to X . Theorem 5.2.ii) implies<br />

that the assertion for X <strong>and</strong> L is equivalent to the assertion for X ′ <strong>and</strong> f ∗ L .<br />

We are therefore reduced to the case that X is regular.<br />

Then, by Theorem 5.2.i),<br />

( )<br />

(L , C ·‖ . ‖), . . . , (L , C ·‖ . ‖)<br />

p X GS<br />

= ̂deg π ∗<br />

(ĉ1 (L , C ·‖ . ‖) · . . . · ĉ 1 (L , C ·‖ . ‖) )<br />

= ̂deg π ∗<br />

(<br />

(ĉ1 (L , ‖ . ‖)+[(0, −2 logC)]) · . . . · (ĉ 1 (L , ‖ . ‖)+[(0, −2 logC)]) ) .<br />

Here, π : X → Specdenotes the structural morphism.<br />

Next, we apply the binomial theorem. The description of the intersection<br />

product given in 4.10 immediately shows<br />

[(0, −2 logC)]·[(Z, g Z )] = [(0, −2(logC )δ Z )]


Sec. 7] THE ADELIC PICARD GROUP 173<br />

for (Z, g Z ) an arbitrary arithmetic cycle. In particular,<br />

[(0, −2 logC )]·[(0, −2 logC)] = 0 .<br />

By consequence, we find<br />

( )<br />

(L , C ·‖ . ‖), . . . , (L , C ·‖ . ‖)<br />

p X GS<br />

( )<br />

= pGS<br />

X (L , ‖ . ‖), . . . , (L , ‖ . ‖)<br />

+ (d + 1) ̂deg π ∗<br />

(<br />

[(0, −2 logC )] · ĉ1 (L , ‖ . ‖) · . . . · ĉ 1 (L , ‖ . ‖) ) .<br />

Here, ĉ 1 (L , ‖ . ‖) · . . . · ĉ 1 (L , ‖ . ‖) = [(Z, g Z )] for Z a cycle of dimension one.<br />

The restriction of Z to the generic fiber is a 0-cycle of degree deg L<br />

X. Therefore,<br />

π ∗<br />

(<br />

[(0, −2 logC )] · [(Z, gZ )] ) = π ∗ [(0, −2(logC )δ Z )]<br />

= [(0, −2(logC ) deg L<br />

X )] .<br />

The claim now follows from Example-Definition 4.5.<br />

□<br />

7. The adelic Picard group<br />

7.1. –––– The arithmetic intersection theory discussed so far seems to be<br />

very general. It is, however, still not general enough. An obvious problem<br />

is that the naive height on P n is not covered. The requirement that<br />

(L , ‖ . ‖) = L ∈ ̂Pic(X ) includes that ‖ . ‖ has to be smooth. For the naive<br />

height, we want to include the minimum metric. This means, we need a generalization<br />

to continuous hermitian metrics.<br />

This may be done in a rather ad hoc manner by taking into account certain<br />

limits of smooth hermitian metrics. A way to formalize this as well as the<br />

dependence of the height on the choice of a model is provided by S. Zhang’s<br />

adelic Picard group [Zh95a].<br />

i. The local case. Metrics induced by a model. —<br />

7.2. –––– Let K be an algebraically closed valuation field. The cases we have<br />

in mind are K =Ép for a prime number p <strong>and</strong> K =É∞ =.<br />

We will denote the valuation of x ∈ K by |x|. In the case K =Ép, we assume<br />

| . | is normalized by |p| = 1 . We also write ν(x) := − log |x|.<br />

p<br />

7.3. Definition. –––– Let X be a K-scheme. Then, by a metric on an invertible<br />

sheaf L ∈ Pic(X ), we mean a system of K-norms on the K-vector spaces L (x)<br />

for x ∈ X (K ).


174 THE CONCEPT OF A HEIGHT [Chap. IV<br />

7.4. Remark. –––– If K =then a metric on L is the same as a (possibly<br />

discontinuous) hermitian metric.<br />

7.5. Definition. –––– Assume K to be non-archimedean <strong>and</strong> let O K be the ring<br />

of integers in K. Further, let X be a K-scheme <strong>and</strong> L ∈ Pic(X ).<br />

Then, by a model of (X, L ) we mean a triple (X , ˜L , n) consisting of a natural<br />

number n, a flat projective scheme π : X → O K such that X K<br />

∼ = X, <strong>and</strong> an<br />

invertible sheaf ˜L ∈ Pic(X ) fulfilling ˜L | X<br />

∼ = L ⊗n .<br />

7.6. Example. –––– Assume K to be non-archimedean, let O K be the ring of<br />

integers in K, <strong>and</strong> let X be a K-scheme equipped with an invertible sheaf L .<br />

Then, a model (X , ˜L , n) of (X, L ) induces a metric ‖ . ‖ on L as follows.<br />

x ∈ X (K ) has a unique extension x : Spec O K → X . Then, x ∗˜L is a projective<br />

O K -module of rank 1. Each l ∈ L (x) induces l ⊗n ∈ L ⊗n (x) <strong>and</strong>, therefore, a<br />

<strong>rational</strong> section of x ∗˜L . Put<br />

‖l‖(x) :=<br />

[<br />

inf<br />

{<br />

|a| | a ∈ K, l ∈ a · x ∗˜L }] 1<br />

n<br />

. (†)<br />

7.7. Definition. –––– The metric ‖ . ‖ given by (†) is called the metric on L<br />

induced by the model (X , ˜L , n).<br />

7.8. Remark. –––– Note here that O K is, in general, a non-discrete valuation<br />

ring. In particular, O K will usually be non-Noetherian.<br />

Nevertheless, projectivity includes being of finite type [EGA, Chapitre II, Définition<br />

(5.5.2)]. This means, for the description of X , only a finite number of<br />

elements from O K are needed.<br />

In the particular case K =Ép, the group ν(K ) is isomorphic to (É, +).<br />

Thus, for any finite set {a 1 , . . . , a s } ⊂ , there exists a discrete valuation ring<br />

OÉp<br />

O ⊆ containing a OÉp<br />

1, . . . , a s .<br />

By consequence, X is the base change of some scheme which is projective over<br />

a discrete valuation ring.<br />

7.9. Definition. –––– Let K be an algebraically closed valuation field. Assume<br />

K to be non-archimedean.<br />

Then, a metric ‖ . ‖ on L ∈ Pic(X ) is called continuous, respectively bounded,<br />

if ‖ . ‖ = f · ‖ . ‖ ′ for ‖ . ‖ ′ a metric induced by some model <strong>and</strong> f a function<br />

on X (K ) which is continuous or bounded, respectively.<br />

7.10. Remark. –––– If K =then we adopt the concepts of bounded, continuous,<br />

<strong>and</strong> smooth metrics in their the usual meaning from complex geometry.<br />

Note that smooth metrics are continuous <strong>and</strong> that continuous metrics are automatically<br />

bounded in the case K =.


Sec. 7] THE ADELIC PICARD GROUP 175<br />

ii. The global case. Adelicly metrized invertible sheaves. —<br />

7.11. Definition. –––– Let X be a projective variety overÉ<strong>and</strong> m ∈Æ.<br />

Then, by a model of X over Spec[ 1 ], we mean a scheme X which is projective<br />

m<br />

<strong>and</strong> flat over Spec[ 1 ] such that the generic fiber of X is isomorphic to X.<br />

m<br />

7.12. Definition. –––– Let X be a projective variety overÉ<strong>and</strong> L ∈ Pic(X )<br />

be an invertible sheaf.<br />

a) Then, an adelic metric on L is a system<br />

‖ . ‖ = {‖ . ‖ ν } ν∈Val(É)<br />

of continuous <strong>and</strong> bounded metrics on ∈ ) such that<br />

LÉν<br />

Pic(XÉν<br />

i) for each ν ∈ Val(É), the metric ‖ . ‖ ν is Gal(Éν/Éν)-invariant,<br />

ii) for some m ∈Æ, there exist a model X of X over Spec[ 1 ], an invertible<br />

m<br />

sheaf ˜L ∈ Pic(X ), <strong>and</strong> a natural number n such that<br />

˜L | X<br />

∼ = L<br />

⊗n<br />

<strong>and</strong>, for all prime numbers p ∤m, the metric ‖ . ‖ νp is induced by (X p , ˜L | Xp , n).<br />

b) Aninvertiblesheafequipped withan adelic metriciscalled an adeliclymetrized<br />

invertible sheaf.<br />

All adelicly metrized invertible sheaves on X form an abelian group which will<br />

be denoted by Pic ad (X ).<br />

7.13. Notation. –––– Let X be a model of X over Spec. Then, taking the<br />

induced metric yields two natural homomorphisms<br />

i X : ̂Pic(X ) → Pic ad (X ) ,<br />

a X : ker(Pic(X ) → Pic(X )) ⊗É→Pic ad (X ) .<br />

Further, one has the forgetful homomorphism<br />

v : Pic ad (X ) → Pic(X ) .<br />

7.14. Notation. –––– The models of X together with all bi<strong>rational</strong> morphisms<br />

between them form an inverse system of schemes. This is a filtered inverse<br />

system since, for two models X <strong>and</strong> X ′ , the closure of the diagonal<br />

∆ ⊂ X × SpecÉX ⊂ X × SpecX ′ projects to both of them.<br />

Thus, the arithmetic Picard <strong>groups</strong> ̂Pic(X ) for all models X of X form a<br />

filtered direct system. The injections ̂Pic(X ) ֒→ Pic ad (X ) fit together to yield<br />

an injection<br />

ι X : lim −→ ̂Pic(X ) ֒→ Pic ad (X ) .


176 THE CONCEPT OF A HEIGHT [Chap. IV<br />

Similarly, the usual Picard <strong>groups</strong> Pic(X ) form a filtered direct system, too.<br />

We get a homomorphism<br />

α X : lim −→<br />

ker(Pic(X ) → Pic(X )) ⊗É−→ Pic ad (X ) .<br />

7.15. Definition (Metric on v −1 (L ) ⊆ Pic ad (X )). —–<br />

Let X be a projective variety overÉ. On X, let (L , ‖ . ‖) <strong>and</strong> (L , ‖ . ‖ ′ ) be two<br />

adelicly metrized invertible sheaves with the same underlying sheaf.<br />

Then, the distance between (L , ‖ . ‖) <strong>and</strong> (L , ‖ . ‖ ′ ) is given by<br />

δ ( (L , ‖ . ‖), (L , ‖ . ‖ ′ ) ) ( )<br />

:= ∑ δ ν ‖ . ‖ν , ‖ . ‖ ′ ν<br />

ν∈Val(É)<br />

for<br />

( δ ν ‖ . ‖ν , ‖ . ‖ ν) ′ := sup<br />

∣ log ‖ . ‖′ ν(x)<br />

‖ . ‖ ν (x) ∣ .<br />

x∈X (Éν )<br />

7.16. Lemma. –––– δ is a metric on the set v −1 (L ) of all metrizations of L .<br />

Proof. We have to show that the sum is always finite.<br />

For this, we note first that the metrics ‖ . ‖ ν <strong>and</strong> ‖ . ‖ ′ ν are bounded by definition.<br />

Therefore, each summ<strong>and</strong> is finite.<br />

We may thus ignore a finite set S of primes <strong>and</strong> assume that ‖ . ‖ <strong>and</strong> ‖ . ‖ ′<br />

are given by triples (X , ˜L , n) <strong>and</strong> (X ′ , ˜L ′ , n ′ ), respectively, in the sense of<br />

Definition 7.12.a.ii).<br />

=<br />

The isomorphism XÉ∼<br />

−→ X ∼ =<br />

−→ X ′Émay be extended to an open neighbourhood<br />

of the generic fiber. Therefore, enlarging S if necessary, we have<br />

an isomorphism X ′ ∼ =<br />

−→ X of schemes over Spec\S. Further, the<br />

triple (X , ˜L , n) may be replaced by (X , ˜L ⊗n′ , nn ′ ) without any change of<br />

the induced metric. Thus, without restriction, n = n ′ .<br />

To summarize, we are reduced to the case that<br />

XÉ‖ . ‖ <strong>and</strong> ‖ . ‖ ′ are given<br />

by (X , ˜L , n) <strong>and</strong> (X , ˜L ′ , n). We have an isomorphism<br />

= ˜L | XÉ∼<br />

−→ L ⊗n ∼ =<br />

−→ ˜L ′ |<br />

which may be extended to an open neighbourhood of the generic fiber. Therefore,<br />

in the definition of δ ( (L , ‖ . ‖), (L,‖.‖ ′ ) ) , all the summ<strong>and</strong>s vanish,<br />

except finitely many.<br />

Positivity, symmetry, <strong>and</strong> the triangle inequality are clear.<br />

7.17. Remark. –––– It is convenient to consider two adelicly metrized invertible<br />

sheaves with different underlying sheaves as of distance infinity. Then, the<br />

distance δ is no longer a metric but only a separated écart in the sense of N. Bourbaki<br />

[Bou-T, §1].<br />


Sec. 7] THE ADELIC PICARD GROUP 177<br />

7.18. Lemma. –––– Let f : X → Y be a morphism of projective varieties overÉ.<br />

i) Then, the homomorphism f ∗ : Pic ad (Y ) → Pic ad (X ) is continuous with respect<br />

to the metric topology.<br />

ii) Even more,<br />

δ( f ∗ L 1 , f ∗ L 2 ) ≤ δ(L 1 , L 2 )<br />

for arbitrary adelicly metrized invertible sheaves L 1 , L 2 ∈ Pic ad (Y ).<br />

Proof. ii) is obvious. i) follows immediately from ii).<br />

□<br />

iii. The adelic Picard group. —<br />

7.19. Definition. –––– Let X be a regular, projective variety overÉ.<br />

a) We call an adelicly metrized invertible sheaf (L , ‖ . ‖) on X semipositive if<br />

there exist a model X of X over Spec<strong>and</strong> a smooth hermitian line bundle<br />

(L , ‖ . ‖ ∞ ) ∈ ̂Pic(X ) fulfilling<br />

i) (L , ‖ . ‖) = i X (L , ‖ . ‖ ∞ ),<br />

ii) L | Xp<br />

is nef for every prime number p, <strong>and</strong><br />

iii) the curvature form c 1 (L, ‖ . ‖ ∞ ) is non-negative.<br />

b) By the semipositive cone C ≥0 ⊂ X<br />

Picad (X ), we mean the set of all adelicly<br />

metrized invertible sheaves which may be written as positiveÉ-linear combinations<br />

of semipositive elements.<br />

7.20. Definition. –––– Let X be a regular, projective variety overÉ.<br />

We define the adelic Picard group ˜Pic(X ) ⊆ Pic ad (X ) to be the subgroup generated<br />

by C ≥0<br />

X . Here, overlining indicates the topological closure in Picad (X ) with<br />

respect to the metric topology.<br />

If (L , ‖ . ‖) ∈ ˜Pic(X ) then ‖ . ‖ is called an integrable metric on L . (L , ‖ . ‖) is<br />

then said to be an integrably metrized invertible sheaf.<br />

7.21. Remark. –––– Since C ≥0<br />

X<br />

is a surjection.<br />

diff : C ≥0<br />

X<br />

is closed under addition, the difference map<br />

× C ≥0<br />

X −→ ˜Pic(X )<br />

7.22. Definition. –––– Let X be a regular, projective variety overÉ.<br />

We define the integrable topology on ˜Pic(X ) to be the finest topology such that<br />

the difference map diff is continuous.


178 THE CONCEPT OF A HEIGHT [Chap. IV<br />

7.23. Remarks. –––– i) The integrable topology on ˜Pic(X ) is finer than the<br />

metric topology.<br />

ii) We will consider ˜Pic(X ) as being equipped with the integrable topology.<br />

Not with the metric topology which is the same as the topology as a subspace<br />

of Pic ad (X ).<br />

iii) A map i : ˜Pic(X ) → M to any topological space is continuous if <strong>and</strong> only if<br />

i ◦ diff : C ≥0<br />

X<br />

× C ≥0<br />

X<br />

−→ M<br />

is continuous with respect to the metric topology.<br />

7.24. Lemma. –––– Let X be a regular, projective variety overÉ. Then, the<br />

difference map<br />

diff : C ≥0 × X<br />

C ≥0 −→ ˜Pic(X X<br />

)<br />

is open.<br />

Proof. Let U ⊆ C ≥0 × X<br />

C ≥0<br />

X<br />

be any open set. We have to show that<br />

diff −1 (diff(U )) is open. But this is clear since<br />

diff −1 (diff(U )) = [ [ ] [ ]<br />

U + (x, x) ∩ C<br />

≥0 ≥0<br />

X<br />

×CX<br />

x∈Pic ad (X )<br />

<strong>and</strong> addition is continuous with respect to the metric topology.<br />

7.25. Lemma. –––– Let X be a regular, projective variety overÉ. Then, the<br />

topological closures of C ≥0<br />

X<br />

with respect to the metric topology <strong>and</strong> the integrable<br />

topology coincide.<br />

Proof. Let C ≥0<br />

≥0<br />

X<br />

be the closure of CX<br />

with respect to the metric topology <strong>and</strong><br />

C be the closure with respect to the integrable topology. Since the integrable<br />

topology is finer, we have C ⊆ C ≥0<br />

X .<br />

On the other h<strong>and</strong>, let l ∈ C ≥0<br />

X . Then, there is a sequence (l i) i in C ≥0<br />

X<br />

which<br />

converges to l. As the canonical map diff(·, 0): C ≥0 → ˜Pic(X X<br />

) is continuous,<br />

we see that l i → l with respect to the integrable topology. Hence, l ∈ C. □<br />

7.26. Definition. –––– Let X be a smooth, projective variety overÉ<br />

<strong>and</strong> L ∈ Pic(X ).<br />

a) Let p be a prime number.<br />

Then, a metric ‖ . ‖ on will be called semipositive if there is a sequence<br />

LÉp<br />

(‖ . ‖ i ) i≥0 of metrics on L satisfying the following conditions.<br />

i) (‖ . ‖ i ) i≥0 converges uniformly to ‖ . ‖.<br />

ii) Each ‖ . ‖ i is induced by a triple (X × Specp, Spec(p)<br />

pr ∗ 1˜L , n) where X is a<br />

scheme over(p) :=p ∩É, ˜L ∈ Pic(X ), <strong>and</strong> ˜L | Xp<br />

is nef.<br />


Sec. 7] THE ADELIC PICARD GROUP 179<br />

b) We will call a metric ‖ . ‖ on Lpositive if ‖ . ‖ is smooth <strong>and</strong> the curvature<br />

form c 1 (L , ‖ . ‖) is positive.<br />

‖ . ‖ is said to be almost semiample if there is a sequence of positive metrics<br />

(‖ . ‖ i ) i≥0 on Lthat converges uniformly to ‖ . ‖.<br />

7.27. Lemma. –––– Let X be a regular, projective variety overÉ. Then, one has<br />

C ≥0<br />

X = {(L , {‖ . ‖ ν} ν∈Val(É)) ∈ Pic ad (X ) |<br />

‖ . ‖ ν semipositive for ν non-archimedean,<br />

Proof. We consider C ≥0<br />

X<br />

“⊆” is clear from Definitions 7.19 <strong>and</strong> 7.26.<br />

‖ . ‖ ν almost semiample for ν archimedean}.<br />

as the closure with respect to the metric topology.<br />

“⊇” Let (L , {‖ . ‖ ν } ν ) ∈ Pic ad (X ) be such that ‖ . ‖ νp is semipositive for every<br />

prime number p <strong>and</strong> ‖ . ‖ ν∞ is almost semiample. The assertion says that, for<br />

every ε > 0, there exists (L , {‖ . ‖ ′ ν} ν ) ∈ C ≥0<br />

X<br />

such that<br />

δ ( (L , {‖ . ‖ ν } ν ), (L , {‖ . ‖ ′ ν} ν<br />

L<br />

) ) < ε .<br />

The archimedean valuation. The hermitian metric ‖ . ‖ ν∞ is almost semiample.<br />

This means that, for every ε > 0, there is a hermitian metric ‖ . ‖ ′ ν ∞<br />

on<br />

which is smooth <strong>and</strong> positively curved such that<br />

δ ν∞<br />

( ‖ . ‖ν∞ , ‖ . ‖ ′ ν ∞<br />

) < ε .<br />

Good primes. By Definition 7.12, we have a finite set S of bad prime numbers,<br />

a model X of X over Spec\S, an invertible sheaf ˜L ∈ Pic(X ), <strong>and</strong> a natural<br />

number n such that<br />

i) ˜L | X<br />

∼ = L ⊗n <strong>and</strong><br />

ii) ‖ . ‖ νp is induced by (X ν , ˜L | Xν , n) for all prime numbers p ∉ S.<br />

Bad primes. Let p ∈ S. According to Definition 7.26.a), for every ε > 0, there<br />

exists an induced metric ‖ . ‖ ′ ν p<br />

such that<br />

δ νp<br />

( ‖ . ‖νp , ‖ . ‖ ′ ν p<br />

) < ε .<br />

More precisely, ‖ . ‖ ′ ν p<br />

is induced by a triple<br />

(X p × Spec(p) Specp, pr ∗ 1˜L (p) , n p )<br />

such that X p is a model of over(p), ˜L (p) ∈ Pic(M XÉ(p) p ), <strong>and</strong> ˜L (p) | Xp<br />

is nef.


180 THE CONCEPT OF A HEIGHT [Chap. IV<br />

Gluing. We claim that X <strong>and</strong> all the X p for p ∈ S may be glued together along<br />

the generic fiber X to a scheme X ′ over Spec. Further, for<br />

N := n ∏ n p ,<br />

p∈S<br />

the tensor powers ˜L ⊗N/n <strong>and</strong> [˜L (p) ] ⊗N/n p<br />

fit together to give an invertible<br />

sheaf ˜L ′ on X .<br />

Then, {‖ . ‖ ′ ν} ν is defined by<br />

(L , {‖ . ‖ ′ ν} ν ) := 1 N i X (˜L ′ , ‖ . ‖ ′ ν ∞<br />

) .<br />

This guarantees that ‖ . ‖ ′ ν p<br />

<strong>and</strong> ‖ . ‖ νp coincide for all p ∉ S. In particular,<br />

˜L ′ | Xp<br />

is automatically nef for these primes. For the bad primes p ∈ S, the<br />

same property follows immediately from the construction.<br />

Gluing may, however, not be done directly as X ⊂ X is not an open set.<br />

Instead, we may extend each X p to a scheme X (p) over Spec[ 1 ] for some<br />

m<br />

m ∈Æwhich is not a multiple of p. We may assume that m is divisible by all<br />

bad primes different from p.<br />

The isomorphism between the generic fibers of X <strong>and</strong> X (p) extends to an open<br />

neighbourhood of X. More precisely, there exists some f p ∈not divisible by<br />

any of the bad primes such that the induced bi<strong>rational</strong> map<br />

is an isomorphism.<br />

X (p) × SpecSpec[ 1<br />

p f p<br />

] −→ X × SpecSpec[ 1 f p<br />

]<br />

Putting f := ∏ p∈S f p <strong>and</strong> U (p) := X (p) × SpecSpec[ 1 ], we have an isomorphism<br />

f<br />

ϕ p : U (p) × SpecSpec[ 1 ] −→ X × p SpecSpec[ 1 ] . f<br />

The definition<br />

ϕ p1 ,p 2<br />

:= ϕ −1<br />

p 2<br />

◦ϕ p1 : U (p1 ) × SpecSpec[ 1 p 1<br />

] −→ U (p2 ) × SpecSpec[ 1 p 2<br />

]<br />

makes the gluing data complete.<br />

The cocycle relations are fulfilled since the corresponding relations are satisfied<br />

on the generic fiber. Observe, U (p) for p ∈ S <strong>and</strong> X are integral schemes.<br />

By consequence, the U (p) <strong>and</strong> X may be glued together to give a-scheme X ′ .<br />

To glue the invertible sheaves together, we need isomorphisms<br />

ι p : ϕ ∗ p˜L ⊗N/n | U −→ [˜L (p) ] ⊗N/n p<br />

| U(p) × SpecSpec[ 1 p ]<br />

extending the isomorphism over X.


Sec. 7] THE ADELIC PICARD GROUP 181<br />

Generally, there is an open neighbourhood O ⊆ U (p) × SpecSpec[ 1 p ] of X<br />

such that ι p | O exists <strong>and</strong> is an isomorphism. Adding to f a suitable set of prime<br />

factors guarantees that O = U (p) × SpecSpec[ 1 p ].<br />

This completes the proof.<br />

7.28. Remark (concerning the notion of a semipositive metric). —–<br />

Condition ii) in Definition 7.26 seems to be somewhat unnatural. One might<br />

want to allow arbitrary models overp or, even more generally, over the integer<br />

ring O L of a finite extension L ofÉp.<br />

It turns out from a st<strong>and</strong>ard descent argument that finite extensions actually<br />

make no difference. Indeed, a metric is supposed to be Gal(Ép/Ép)-invariant.<br />

Thus, it may be shown by descent that for every model over O L there is an<br />

equivalent model overp. Similarly, for every model over the integer ring of a<br />

finite extension ofÉ(p), there is an equivalent model over(p).<br />

The descent needed here may actually be seen as a particular case of faithful<br />

flat descent [K/O74] or as a slight generalization of Proposition I.3.5<br />

to O K -schemes instead of K-schemes.<br />

It seems, however, that allowing the completion leads to a more general notion<br />

of semipositivity. We do not know whether Lemma 7.27 is still true for<br />

that notion.<br />

7.29. Lemma. –––– Let X be a regular, projective variety overÉ.<br />

a) Then, the natural homomorphisms<br />

<strong>and</strong><br />

factorize via ˜Pic(X ).<br />

ι X : lim −→ ̂Pic(X ) → Pic ad (X )<br />

α X : lim −→<br />

ker(Pic(X ) → Pic(X )) ⊗É→Pic ad (X )<br />

b) The subgroup 〈 im ι X , im α X 〉 is dense in ˜Pic(X ).<br />

Proof. a) Let X be any model of X over Spec. We have to show that the<br />

homomorphisms i X <strong>and</strong> a X factorize via ˜Pic(X ).<br />

Since X is a projective scheme, every (L , ‖ . ‖) ∈ ̂Pic(X ) may be written<br />

as (L 1 , ‖ . ‖ 1 )⊗ OX (L 2 , ‖ . ‖ 2 ) ∨ where both oper<strong>and</strong>s are ample invertible sheaves<br />

with positive metrics. Further, we note that ampleness of L i implies that L i | Xp<br />

is ample on Xp<br />

for every p (cf. [EGA, Chapitre II, Corollaire (4.5.10.ii)]).<br />

Hence,<br />

i X (L , ‖ . ‖) = i X (L 1 , ‖ . ‖ 1 ) − i X (L 2 , ‖ . ‖ 2 ) ∈ diff(C ≥0<br />

X , C ≥0<br />

X ) .<br />


182 THE CONCEPT OF A HEIGHT [Chap. IV<br />

On the other h<strong>and</strong>, let k ∈Æ<strong>and</strong> L ∈ ker(Pic(X ) → Pic(X )) be given.<br />

We may write L = L 1 ⊗ OX L2 ∨ for two ample invertible sheaves such<br />

that L 1 | X<br />

∼ = L2 | X is k-divisible in Pic(X ). Then,<br />

a X (L ⊗ 1/k) = 1 k ·i X (L 1 , ‖ . ‖) − 1 k ·i X (L 2 , ‖ . ‖) ∈ diff(C ≥0<br />

X , C ≥0<br />

X )<br />

independently of the metric ‖ . ‖ chosen. Note that 1 k ·i X (L i , ‖ . ‖) ∈ Pic ad (X )<br />

since L 1 | X is k-divisible.<br />

b) Since the difference map diff : C ≥0<br />

X<br />

× C ≥0<br />

X<br />

surjective, it suffices to prove that 〈 im ι X , im α X 〉∩C ≥0<br />

X<br />

show that even 〈 im ι X , im α X 〉 ⊇ C ≥0<br />

X .<br />

For this, let (L , ‖ . ‖) ∈ C ≥0<br />

X<br />

. By definition, we have<br />

→ ˜Pic(X ) is continuous <strong>and</strong><br />

≥0<br />

is dense in C . We will<br />

(L , ‖ . ‖) = a 1 i X (L 1 , ‖ . ‖ 1 ) + . . . + a n i X (L n , ‖ . ‖ n )<br />

for a certain model X of X, (L i , ‖ . ‖ i ) ∈ ̂Pic(X ), <strong>and</strong> <strong>rational</strong> numbers a i ≥ 0.<br />

We choose a common denominator k of all the a i . Then,<br />

(L , ‖ . ‖) ⊗k = i X (L , ‖ . ‖ ′ )<br />

for some (L , ‖ . ‖ ′ ) ∈ ̂Pic(X ). In particular, L ⊗k ∼ = L | X .<br />

We may write L = R ⊗k ⊗ T for R ∈ Pic(X ) an arbitrary extension of L<br />

<strong>and</strong> T ∈ ker(Pic(X ) → Pic(X )). Then,<br />

(L , ‖ . ‖) = i X (R, ‖ . ‖ ′ 1 k ) + aX (T ⊗ 1/k) .<br />

X<br />

The assertion is proven.<br />

□<br />

iv. Pull-back <strong>and</strong> intersection product. —<br />

7.30. Lemma. –––– Let f : X → Y be a morphism of projective varieties overÉ.<br />

a) Then, there is a natural homomorphism<br />

f ∗ : ˜Pic(Y ) → ˜Pic(X )<br />

(L , ‖ . ‖) ↦→ ( f ∗ L , f ∗ ‖ . ‖) .<br />

f ∗ admits the property that f ∗( )<br />

CY<br />

≥0 ⊆ C<br />

≥0<br />

X .<br />

b) f ∗ is continuous.<br />

Proof. a) It is clear that if ‖ . ‖ is an adelic metric then f ∗ ‖ . ‖ is an adelic metric.<br />

According to Definition 7.20, for the existence of f ∗ : ˜Pic(Y ) → ˜Pic(X ), it will<br />

be sufficient to show<br />

f ∗( )<br />

CY<br />

≥0 ⊆ C<br />

≥0<br />

X .


Sec. 7] THE ADELIC PICARD GROUP 183<br />

Let us first verify that f ∗ (CY ≥0 ) ⊆ C ≥0<br />

X .<br />

≥0<br />

For this, let (L , ‖ . ‖) ∈ CY .<br />

I.e., (L , ‖ . ‖) = i Y (L , ‖ . ‖ ∞ ) for a smooth hermitian line bundle (L , ‖ . ‖ ∞ )<br />

on a model Y of Y such that L | Yp<br />

is nef for every prime number p <strong>and</strong> the<br />

curvature form c 1 (L, ‖ . ‖ ∞ ) is non-negative.<br />

Fix an arbitrary model X ′ of X. We define X to be the closure of the<br />

graph of f in X ′ ×Y . Then, the projection to the second factor yields a<br />

morphism f : X → Y which extends f . This yields<br />

f ∗ (L , ‖ . ‖) = i X ( f ∗ L , f ∗ ‖ . ‖ ∞ ) .<br />

The pull-back of a nef invertible sheaf is always nef <strong>and</strong> the pull-back of<br />

a smooth non-negatively curved hermitian line bundle is smooth <strong>and</strong> nonnegatively<br />

curved. Hence, we indeed have f ∗ (CY ≥0 ) ⊆ C ≥0<br />

X .<br />

Further, by Lemma 7.18, the homomorphism f ∗ : Pic ad (Y ) → Pic ad (X ) is continuous<br />

with respect to the metric topology. Lemma 7.25 yields the assertion.<br />

b) f ∗ : CY<br />

≥0 → C ≥0<br />

X<br />

is continuous with respect to the metric topology. Thus, we<br />

have a continuous map<br />

C ≥0<br />

Y<br />

× C ≥0<br />

Y<br />

f ∗ ×f<br />

−→ ∗<br />

C ≥0 × X<br />

C ≥0 diff<br />

X<br />

−→ Pic ad (X ) .<br />

The definition of the integrable topology yields the claim.<br />

□<br />

7.31. Theorem. –––– Let X be a regular, projective variety overÉ.<br />

i) Then, there is exactly one continuous-multilinear map, the adelic intersection<br />

product<br />

pZ X : ˜Pic(X ) × . . . × ˜Pic(X )<br />

Ê<br />

−→Ê,<br />

} {{ }<br />

dim X+1 times<br />

such that for every model X of X over Specthe diagram<br />

˜Pic(X ) × . . . × ˜Pic(X<br />

pZ<br />

X<br />

)<br />

commutes.<br />

ii) p X Z<br />

is symmetric.<br />

i dimX+1<br />

X<br />

̂Pic(X ) × . . . × ̂Pic(X )<br />

<br />

Proof. i) By Theorem 5.2.ii), the intersection products pGS X are compatible with<br />

pull-backs under bi<strong>rational</strong> maps. We obtain a-multilinear map p such that<br />

p X GS


Ê<br />

184 THE CONCEPT OF A HEIGHT [Chap. IV<br />

the diagrams<br />

p<br />

im ι X × . . . × im ι X<br />

i dimX+1<br />

X<br />

<br />

p X<br />

̂Pic(X ) × . . . × ̂Pic(X<br />

GS<br />

)<br />

commute.<br />

Every element of im α X has an integral multiple in im ι X . Therefore, p extends<br />

uniquely to a-multilinear map<br />

p : 〈 im ι X , im α X 〉 × . . . × 〈 im ι X , im α X 〉 −→Ê.<br />

By Lemma 7.29.b), 〈 im ι X , im α X 〉 is a dense subset of ˜Pic(X ). This implies that<br />

pZ X is unique.<br />

To show the existence, we need that p is continuous. This may be tested locally.<br />

Let<br />

We write<br />

l = (l 0 , . . . , l dimX ) ∈ 〈 im ι X , im α X 〉 × . . . × 〈 im ι X , im α X 〉<br />

⊆ ˜Pic(X ) × . . . × ˜Pic(X ) .<br />

(l 0 , . . . , l dimX ) = (l (1)<br />

0 , . . . , l(1) dim X ) − (l(2) 0 , . . . , l(2) dimX )<br />

for l (i)<br />

0 , . . . , l(i) ∈ dimX C ≥0<br />

(i)<br />

X<br />

<strong>and</strong> define U<br />

j<br />

in C ≥0<br />

X<br />

for ε := 1. By Lemma 7.24,<br />

to be the ε-neighbourhood of l (i)<br />

j<br />

U l := diff(U (1)<br />

0 , U (2)<br />

(1)<br />

0<br />

) × . . . × diff(U<br />

dim X , U (2)<br />

is an open neighbourhood of (l 0 , . . . , l dimX ).<br />

dimX )<br />

It is therefore sufficient to verify that p is continuous on U l . This is a direct<br />

consequence of Lemma 7.32. Indeed, for every (L , ‖ . ‖) ∈ U (i)<br />

j<br />

the restriction<br />

:= L | X is always the same. Put<br />

L (i)<br />

j<br />

M 0 :=<br />

2 2<br />

∑ ∑<br />

i 1 =1 i 2 =1<br />

. . .<br />

2<br />

∑<br />

i dimX =1<br />

deg ( c 1 (L (i 1)<br />

1<br />

) · . . . · c 1 (L (i dimX )<br />

dim X )) .<br />

Then, p is Lipschitz continuous on U l with respect to the first variable. M 0 is a<br />

Lipschitz constant.<br />

Lipschitz continuity in the other variables may be shown analogously.<br />

ii) This follows immediately from the uniqueness together with the symmetry<br />

of the intersection products pGS X proven in Theorem 5.2.b).<br />


Sec. 7] THE ADELIC PICARD GROUP 185<br />

7.32. Lemma. –––– Let π : X → Specbe an arithmetic variety. Suppose its<br />

generic fiber X is regular <strong>and</strong> put d := dim X.<br />

Let L 1 , . . . , L d ∈ ̂Pic(X ). Suppose that L 1 | , . . . , L Xp<br />

d| Xp<br />

are nef for every<br />

prime number p. Further, consider (L , ‖ . ‖) ∈ ̂Pic(X ) such that L | X<br />

∼ = OX .<br />

Then,<br />

|p X GS(<br />

(L , ‖ . ‖), L1 , . . . , L d<br />

) | ≤<br />

≤ δ ( i X (L , ‖ . ‖), 0 ) · deg ( c 1 (L 1 | X ) · . . . · c 1 (L d | X ) )<br />

for 0 ∈ Pic ad (X ) the neutral element.<br />

Proof. The assertion is compatible with pull-back under a generically finite, surjective<br />

morphism. Thus, we may assume without restriction that X is regular.<br />

The asserted inequality then reads<br />

∣<br />

∣̂deg π ∗<br />

(ĉ1 (L , ‖ . ‖) · ĉ 1 (L 1 ) · . . . · ĉ 1 (L d ) )∣ ∣ ≤<br />

To show this, we write first<br />

≤ δ ( i X (L , ‖ . ‖), 0 ) · deg ( c 1 (L 1 | X ) · . . . · c 1 (L d | X ) ) .<br />

ĉ 1 (L , ‖ . ‖) =: [(D, g)] .<br />

Here, D = div s for s the <strong>rational</strong> section of L such that s| X = 1 ∈ Γ(X, O X ).<br />

In particular, D is a divisor supported in the special fibers. Further,<br />

g := − log ‖s‖ 2 .<br />

It is sufficient to treat the cases (0, g) <strong>and</strong> (D, 0).<br />

First Case: D = 0.<br />

We have ĉ 1 (L 1 ) · . . . · ĉ 1 (L d ) = [(Z, g Z )] for Z a cycle of dimension<br />

one. The restriction of Z to the generic fiber is a 0-cycle of degree<br />

deg ( c 1 (L 1 | X ) · . . . · c 1 (L d | X ) ) . Therefore, by the description of the arithmetic<br />

intersection product given in 4.10,<br />

(<br />

π ∗ [(0, g)] · ĉ1 (L 1 ) · . . . · ĉ 1 (L d ) ) (<br />

= π ∗ [(0, g)] · [(Z, gZ )] )<br />

= π ∗ [(0, gδ Z )] .<br />

The absolute value of the arithmetic degree is bounded by<br />

1<br />

2 sup |g(x)|·degZÉ≤ sup |log ‖1‖(x)| · deg ( c 1 (L 1 | X ) · . . . · c 1 (L d | X ) ) .<br />

x∈X ()<br />

x∈X ()<br />

Second Case: g = 0.<br />

We may assume that D is contained in one special fiber, say, in Xp .


186 THE CONCEPT OF A HEIGHT [Chap. IV<br />

Assume first that D = (V ) for an irreducible component V of of maximal<br />

dimension. Then, by the definition of the push-forward <strong>and</strong> the arithmetic<br />

Xp<br />

degree, we have<br />

̂deg π ∗<br />

(<br />

[(D, 0)] · ĉ1 (L 1 ) · . . . · ĉ 1 (L d ) ) = (log p)·deg ( c 1 (L 1 | V )·. . . ·c 1 (L d | V ) ) .<br />

We observe, if D is effective then the expression on the right h<strong>and</strong> side is nonnegative.<br />

Indeed, the intersection number of nef divisors is always non-negative<br />

by a theorem of S. Kleiman [Laz, Example 1.4.16].<br />

By consequence, if<br />

for r ∈Êthen<br />

∣ ̂deg<br />

(<br />

π ∗ [(D, 0)] · ĉ1 (L 1 ) · . . . · ĉ 1 (L d ) )∣ ∣ ≤<br />

) ≤ D ≤ ) (‡)<br />

−r(Xp r(Xp<br />

≤ r(log p) · deg ( c 1 (L 1 | Xp ) · . . . · c 1(L d | Xp ))<br />

= r(log p) · deg ( c 1 (L 1 | X ) · . . . · c 1 (L d | X ) ) (§)<br />

since the fibers are linearly equivalent to each other.<br />

We choose the minimal number r such that property (‡) is fulfilled.<br />

This means, there is some ξ ∈ Xp<br />

such that O X ,ξ (D) = O X ,ξ (±rXp ).<br />

In particular, r = s/t is a <strong>rational</strong> number. We assume the plus sign without<br />

restriction. Then, O X ,ξ (D) ⊗t = O X ,ξ (Xp )⊗s .<br />

The definition of D shows that L = O X (D). By consequence,<br />

L ⊗t<br />

ξ<br />

= O X,ξ (D) ⊗t = O X ,ξ (Xp )⊗s .<br />

Definition 7.7implies that ‖1‖ p (x) = p −r for x ∈ X specializing to ξ. Hence, by<br />

Lemma-Definition 7.15,<br />

δ νp<br />

(<br />

iX (L , ‖ . ‖), 0 ) ≥ |log ‖1‖ p (x)| = r log p .<br />

Formula (§) yields the assertion.<br />

□<br />

7.33. Lemma. –––– Let f : X → Y be a generically finite morphism of projective<br />

varieties overÉ. Then, for every L 1 , . . . , L dim Y+1 ∈ ˜Pic(Y ),<br />

p X Z ( f ∗ L 1 , . . . , f ∗ L dim Y+1 ) = p Y Z (L 1, . . . , L dim Y+1 ) .<br />

Proof. This follows directly from Theorem 5.2.a.ii) together with<br />

Lemma 7.29.b).<br />


Sec. 7] THE ADELIC PICARD GROUP 187<br />

7.34. Remark. –––– One might want to enlarge ˜Pic(X ). This is easily possible<br />

by omitting condition ii) in Definition 7.12. This would not destroy any of the<br />

properties proven except for the explicit description of the closure C ≥0<br />

X<br />

given<br />

in Lemma 7.27.<br />

7.35. Remark. –––– A priori, S. Zhang’s adelic Picard group [Zh95a] is somewhat<br />

smaller.<br />

S. Zhang works with a more restrictive notion of convergence. Instead of<br />

the closure with respect to the metric topology, he works with all limits of<br />

sequences ( {‖ . ‖ ν,i } ν<br />

)i<br />

of adelic metrics which are convergent in the following<br />

sense.<br />

(‖ . ‖ ν,i ) i is a constant sequence for all but finitely many valuations ν. It converges<br />

uniformly for the other ν.<br />

The proof of Lemma 7.27 indicates, however, that this should make no difference.<br />

v. Examples. —<br />

7.36. Example. –––– Let K be a number field. Then,<br />

I.e., there is an exact sequence<br />

˜Pic(SpecK ) = ̂Pic(Spec O K ) .<br />

0 −→ Cl K −→ ˜Pic(SpecK ) ˜deg<br />

−→Ê−→ 0 .<br />

Here, the class group Cl K is a discrete subgroup of ˜Pic(SpecK ).<br />

homeomorphism on every cofactor.<br />

˜deg is a<br />

Proof. The models of SpecK are the spectra of the orders of K. All induced<br />

metrics come from the maximal order.<br />

□<br />

7.37. Example. –––– Let V be a semistable projective curve overÉ.<br />

Then, there is an exact sequence<br />

where= L<br />

ν∈Val(É)<br />

0 →→ ˜Pic(V ) → Pic(V ) → 0<br />

C 0 (V (Éν)).<br />

Here, C 0 (V (É∞)) = C(V ()) F ∞<br />

is the space of all continuous F ∞ -invariant<br />

functions on V ().<br />

For p a prime number, C 0 (V (Ép)) is given as follows.<br />

We choose a minimal semistable model V of V over SpecÉp. Let D 1 , . . . , D r<br />

be the irreducible divisors in the special fiber. Further, for each k ∈Æ, we


→Ê<br />

188 THE CONCEPT OF A HEIGHT [Chap. IV<br />

define a countable disjoint covering {U i<br />

(k) } i∈Æof V (Ép) by the requirement<br />

that x, y ∈ V (Ép) are in the same set if they coincide modulo p k .<br />

Then, C 0 (V (Ép)) consists of all Gal(Ép/Ép)-invariant functions V (Ép)<br />

of the type<br />

r<br />

∞<br />

∑ α i ‖1‖ (V ,O(Di )) + ∑ ∑ β ikl χ U<br />

(k) .<br />

i<br />

i=1<br />

l=1 ik<br />

Here, ‖ . ‖ (V ,O(Di )) denotes the induced metric, induced by the model (V , O(D i )).<br />

χ U<br />

(k)<br />

i<br />

is the characteristic function of U (k)<br />

i<br />

. α i <strong>and</strong> β ikl are real numbers.<br />

We require that the inner sum is finite for each l ∈Æ<strong>and</strong> that the outer series<br />

converges uniformly.<br />

vi. Adelic heights. —<br />

7.38. Definition. –––– Let X be a regular, projective variety overÉ<strong>and</strong><br />

L ∈ ˜Pic(X ) be an integrably metrized invertible sheaf.<br />

Then, the absolute height with respect to L of an L-valued point x ∈ X (L)<br />

for L a number field is given by<br />

1<br />

h L<br />

(x) :=<br />

[L :É] ˜deg L | x .<br />

7.39. Remark. –––– Lemma 7.33 shows that this definition is independent of<br />

the choice of L.<br />

7.40. Corollary (Theorem of successive minima). —–<br />

Let X be a regular, projective variety overÉ<strong>and</strong> L ∈ ˜Pic(X ). Assume<br />

a) L ∈ C ≥0<br />

X<br />

or<br />

b) L ∈ ˜Pic(X ) is such that ‖ . ‖ p is semipositive for all p ≠ ∞ <strong>and</strong> ‖ . ‖ ∞ is<br />

almost semiample.<br />

Then,<br />

e 1 (h L<br />

) ≥ pX Z (L , . . . , L )<br />

(dim X + 1) deg L<br />

X ≥ e 1(h L<br />

) + . . . + e dim X+1 (h L<br />

)<br />

.<br />

dim X + 1<br />

Proof. Cf. [Zh95a, Theorem 1.10].<br />

a) From Theorem 7.31, we see that Corollary 6.6, together with Remark 6.7,<br />

immediately implies the assertion for L ∈ C ≥0<br />

X . Further, pX Z <strong>and</strong> ˜deg are continuous.<br />

This shows that the inequalities are true for L ∈ C ≥0<br />

X , too.<br />

b) Lemma 7.27 says that L = (L , ‖ . ‖) ∈ C ≥0<br />

X<br />

is equivalent to the statement<br />

that (L , ‖ . ‖) ∈ ˜Pic(X ) such that ‖ . ‖ p is semipositive for all p ≠ ∞ <strong>and</strong> ‖ . ‖ ∞<br />

is almost semiample.<br />


CHAPTER V<br />

ON THE DISTRIBUTION OF SMALL POINTS<br />

ON ABELIAN AND TORIC VARIETIES ∗<br />

1. Introduction<br />

i. Classical results. —<br />

1.1. –––– In 1972, J.-P. Serre proved the following remarkable result.<br />

Theorem (Serre). Let K be an algebraic number field <strong>and</strong> E be an elliptic curve<br />

over K without complex multiplication over K. Then, for almost every prime<br />

number l, the Galois group Gal(K/K ) acts transitively on the l-torsion <strong>points</strong> of E.<br />

1.2. Notation. –––– For x ∈ E(K ), we denote by δ x the Dirac measure associated<br />

to its Galois orbit. I.e., if x ∈ E(F ) for some number field F ⊇ K then<br />

for each ϕ ∈ C 0 (E()).<br />

δ x (ϕ) :=<br />

1<br />

♯ Gal(F/K )<br />

∑<br />

σ : F֒→ϕ(σ(x))<br />

1.3. –––– In this language, Serre’s theorem immediately implies the following<br />

corollary.<br />

Corollary. Let K be an algebraic number field <strong>and</strong> E be an elliptic curve over K<br />

without complex multiplication over K. Further, let (x i ) i∈Æbe a sequence of torsion<br />

<strong>points</strong> in E which does not contain any constant subsequence.<br />

Then, the associated sequence (δ xi ) i∈Æof <strong>measures</strong> on E() converges weakly to the<br />

Haar measure of volume one.<br />

(∗) This chapter is a revised version of the article: On the distribution of small <strong>points</strong> on abelian<br />

<strong>and</strong> toric varieties, Preprint.


190 DISTRIBUTION OF SMALL POINTS [Chap. V<br />

1.4. –––– In 1997, L. Szpiro, E. Ullmo <strong>and</strong> S. Zhang [S/U/Z] showed a generalization<br />

of this corollary from torsion <strong>points</strong> (which are of Néron-Tate height<br />

zero) to <strong>points</strong> of a small positive height. The three authors heavily make use<br />

of the powerful methods of Arakelov geometry as presented in Chapter IV.<br />

Their result is as follows.<br />

i∈Æ<br />

1.5. Theorem (Szpiro, Ullmo, Zhang). —– Let A be an abelian variety over<br />

some number field K <strong>and</strong> (x i ) i∈Æbe a sequence of closed <strong>points</strong> in A. Assume (x i )<br />

converges to the generic point of A in the sense of the Zariski topology. Suppose further<br />

that lim h NT (x i ) = 0.<br />

i→∞<br />

Then, the associated sequence (δ xi ) i∈Æof <strong>measures</strong> on A() converges weakly to the<br />

Haar measure of volume one.<br />

1.6. –––– Surprisingly, there is a completely parallel theorem for the naive<br />

height on projective space which was proven by Yuri Bilu [Bilu97]. In fact,<br />

for P 1 , he proved this result already in 1988 [Bilu88]. It would have been easy<br />

to deduce the general result from this.<br />

1.7. Theorem (Bilu). —– Let (x i ) i∈Æbe a sequence of closed <strong>points</strong> in P nÉ.<br />

Assume that (x i ) i∈Æconverges to the generic point of P n in the sense of the<br />

Zariski topology. Assume further that lim<br />

i→∞<br />

h naive (x i ) = 0.<br />

Then, the associated sequence (δ xi ) i∈Æof <strong>measures</strong> on P n () converges weakly to the<br />

Haar measure of volume one on (S 1 ) n ⊂n ⊂ P n ().<br />

ii. Canonical heights. —<br />

1.8. General situation. –––– Let P be a projective variety overÉequipped<br />

with an ample invertible sheaf L ∈ Pic(P). Further, we assume there is given<br />

a self map f : P → P such that there is some isomorphism Φ: L ⊗d ∼ =<br />

−→ f ∗ L .<br />

1.9. Definition. –––– The canonical height function h f ,L on P is given by<br />

1<br />

h f ,L (x) := lim<br />

n→∞ d h ( n L f (n) (x) )<br />

for x ∈ P a closed point. Here, f (n) means the n-fold iteration of f . h L denotes<br />

the height defined by the invertible sheaf L introduced in Definition IV.2.10.<br />

1.10. Remark. –––– The limit process defining h f ,L is a generalization of the<br />

classical one for the Néron-Tate height [C/S, Chapter VI, §4].<br />

We will see in Example 1.16.ii) that the naive height is a canonical height, too.


Sec. 1] INTRODUCTION 191<br />

1.11. Remark. –––– h L is defined only up to a bounded summ<strong>and</strong>. Nevertheless,<br />

h f ,L is independent of the choice of a particular representative. Indeed, if<br />

|h (1)<br />

L (x) − h(2) L (x)| ≤ C for every x ∈ P then<br />

1 (<br />

∣d n h(1) L f (n) (x) ) − 1 (<br />

d n h(2) L f (n) (x) )∣ ∣∣ ≤ C d . n<br />

iii. An equidistribution result. —<br />

1.12. –––– The goal of the present chapter is to prove a common generalization<br />

of Theorems 1.5 <strong>and</strong> 1.7. We will give an Arakelovian approach to a situation<br />

which covers the case of the Néron-Tate height as well as the naive height.<br />

1.13. Situation. –––– Let P be a regular projective variety overÉcontaining<br />

a group scheme G ⊆ P as an open dense subset. Further, let a morphism<br />

f : P → P be given <strong>and</strong> assume that<br />

i) the unit e is a repelling fixed point. I.e., all eigenvalues of the tangent<br />

map T e f : T e G → T e G are of absolute value strictly bigger than 1.<br />

ii) f | G is a group homomorphism f | G : G → G.<br />

iii) There is some compact, Zariski dense subgroup K ⊆ G() which is both<br />

backward <strong>and</strong> forward invariant under f.<br />

Finally, let L ∈ Pic(P) be ample <strong>and</strong> Φ: L ⊗d ∼ =<br />

−→ f ∗ L be an isomorphism.<br />

This guarantees that we have the canonical height h f ,L .<br />

1.14. Theorem (Equidistribution on P()). —– Let P, f , L , <strong>and</strong> Φ are as described<br />

in 1.13.<br />

Let (x i ) i∈Æbe a sequence of closed <strong>points</strong> in P. Assume that (x i ) i∈Æconverges<br />

to the generic point of P in the sense of the Zariski topology. Further, suppose<br />

that h f ,L (x i ) → 0.<br />

Then, the associated sequence (δ xi ) i of <strong>measures</strong> on P() converges weakly to the<br />

measure τ on P() being the zero measure on P() \ K <strong>and</strong> the Haar measure of<br />

volume one on K.<br />

1.15. Remark. –––– For a general self map f , iterating will lead to fractals.<br />

Unfortunately, not much is known about the corresponding fractal heights.<br />

1.16. Examples. –––– The condition that e is a repelling fixed point is fulfilled,<br />

in particular, if G is commutative <strong>and</strong> f : g ↦→ g l is the homomorphism raising<br />

to the l-th power for l ≥ 2. This includes the two st<strong>and</strong>ard cases. Namely,<br />

i) let P = G = A be an abelian variety <strong>and</strong> L ∈ Pic(A) a symmetric, ample<br />

invertible sheaf. One has<br />

f = [l]: A → A


192 DISTRIBUTION OF SMALL POINTS [Chap. V<br />

<strong>and</strong> puts K = A(). Then, independently of the choice of l, h f ,L is the<br />

Néron-Tate height corresponding to L .<br />

ii) Let P = P n , G =n+1 ∼<br />

m /m =n m <strong>and</strong> L = O(1). We have<br />

f = [l]: (x 0 , . . . , x n ) ↦→ (x l 0 , . . . , xl n )<br />

<strong>and</strong> put K := U (1) n . Here, h f ,L is the naive height.<br />

iii) One may combine abelian varieties <strong>and</strong> projective spaces <strong>and</strong> consider a<br />

split semi-abelian variety A × P n . Here, G = A ×n m , K = A × U (1),<br />

<strong>and</strong> L = S ⊠O(p) for S a symmetric, ample invertible sheaf on A. One may<br />

put f := [l] × [l 2 ].<br />

A. Chambert-Loir’s equidistribution result [C-L] for quasi-split semi-abelian<br />

varieties can easily be deduced from this.<br />

iv) Consider a free abelian group N of finite rank <strong>and</strong> a finite <strong>rational</strong> polyhedral<br />

decomposition {σ α } α of N ⊗Ê. These data define a proper toric variety X.<br />

The multiplication map N → N , n ↦→ ln induces a self-morphism<br />

f = [l]: X → X.<br />

Further, consider a function g : N ⊗Ê→Êwhich is strictly convex in the<br />

sense of [K/K/M/S, Chapter I, Theorem 13]. Then, L := F g is a T -invariant<br />

ample invertible sheaf on X. One has [l] ∗ F g<br />

∼ = Flg = F ⊗l<br />

g .<br />

In this situation, P = X, G = T ∼ =n m is the Zariski dense torus in X, <strong>and</strong><br />

K ⊂ T is its maximal compact subgroup.<br />

1.17. Remark. –––– The latter example clearly contains the case of the naive<br />

height on projective space. Notice that it also contains the blow-up Bl (0:0:0) (P 2 )<br />

of the projective plane in the origin <strong>and</strong> the canonical height defined by the<br />

ample divisor nH − mE for n, m ∈Æ, n > m.<br />

1.18. Remark. –––– For abelian varieties, our proof coincides with that of<br />

Szpiro, Ullmo, <strong>and</strong> Zhang. In the case of projective space, we provide a more<br />

geometric proof for Bilu’s theorem.<br />

2. The dynamics of f<br />

i. Arakelovian interpretation of the canonical height. —<br />

2.1. –––– Let P, f , L , <strong>and</strong> Φ are as described in 1.8.<br />

To underst<strong>and</strong> the canonical height better, one has to use the language of<br />

Arakelov geometry introduced in the previous chapter.


Sec. 2] THE DYNAMICS OF f 193<br />

First, recall that, in Definition 1.9, we may replace h L by any other height<br />

function which differs only by a bounded summ<strong>and</strong>. For example, we may<br />

work with h ′ L (x) := 1 m h l 2 (<br />

iL ⊗m(x) ) .<br />

Further, there is a projective model P of P over Specto which L ⊗m extends.<br />

Indeed, L ⊗m defines a closed embedding i L ⊗m : P → P N <strong>and</strong> one may<br />

put P := i L ⊗m(P).<br />

Then, L := O(1)| P is an extension of L ⊗m to P. We equip L with the<br />

restriction of the Fubini-Study metric. This guarantees that, for every closed<br />

point x ∈ P,<br />

h L ,0 (x) := h L (x) = 1 m h (L ,‖ .‖)(x) .<br />

Here, h (L ,‖ .‖) is the absolute height defined by the smooth hermitian line bundle<br />

(L , ‖ . ‖), cf. Definition IV.5.7.<br />

The recursion is given by<br />

h L ,i+1 (x) := 1 d h L ,i(<br />

f (x)<br />

)<br />

.<br />

Unfortunately, the isomorphism Φ: L ⊗d ∼ =<br />

−→ f ∗ L does not extend to P. If it<br />

would then we could put ‖ . ‖ 0 := ‖ . ‖ <strong>and</strong>, recursively,<br />

This would then lead to<br />

‖ . ‖ i := Φ −1 f ∗ ‖ . ‖ 1/d<br />

i−1 .<br />

h L ,i (x) = 1 m h (L ,‖ .‖ i )(x) .<br />

2.2. –––– To make the plan above precise, one has to work in the more flexible<br />

context of adelic Picard <strong>groups</strong> introduced in Section IV.7.<br />

There is the adelic metric ‖ . ‖ (0)<br />

∼ on L ⊗m given by<br />

(L ⊗m , ‖ . ‖ (0)<br />

∼ ) := i P(L , ‖ . ‖ 0 ) .<br />

Since L is very ample <strong>and</strong> ‖ . ‖ 0 is a restriction of the Fubini-Study metric, we<br />

see that (L ⊗m , ‖ . ‖ ∼ (0))<br />

∈ C ≥0<br />

P .<br />

Further, we define, recursively,<br />

‖ . ‖ ∼ (i) := (Φ⊗m ) −1 f ∗[ ]<br />

‖ . ‖ ∼<br />

(i−1) 1<br />

d<br />

.<br />

Lemma IV.7.30.a) shows (L ⊗m , ‖ . ‖ ∼ (i) ) ∈ C ≥0<br />

P<br />

for every i ∈Æ.<br />

We also observe that the isomorphism Φ does extend to an open neighbourhood<br />

of the generic fiber of P. Therefore, the sequence (‖ . ‖ ∼ (i) ) i is actually constant<br />

at all but finitely many valuations.


194 DISTRIBUTION OF SMALL POINTS [Chap. V<br />

Furthermore, Lemma IV.7.18 implies that<br />

δ ( (L ⊗m , ‖ . ‖ (i)<br />

∼ ), (L ⊗m , ‖ . ‖ (i+1)<br />

∼ )) ≤ 1 d i ·δ( (L ⊗m , ‖ . ‖ (0)<br />

∼ ), (L ⊗m , ‖ . ‖ (1)<br />

∼ )) .<br />

This means, for every ν ∈ Val(É), the sequence (‖ . ‖ (i)<br />

ν ) i of metrics is uniformly<br />

convergent.<br />

The limit ‖ . ‖ ∼ clearly fulfills all the requirements of Definition IV.7.12.<br />

I.e., ‖ . ‖ ∼ is an adelic metric on L ⊗m . Even more,<br />

L := (L ⊗m , ‖ . ‖ ∼ ) ∈ C ≥0<br />

P<br />

⊆ ˜Pic(P) .<br />

By Lemma IV.7.27, ‖ . ‖ νp is semipositive for every prime number p ≠ ∞ <strong>and</strong><br />

‖ . ‖ ν∞ is almost semiample.<br />

Finally, we have<br />

h f ,L = 1 m h L = 1 m h (L ⊗m ,‖ .‖ ∼ ) .<br />

2.3. Lemma. –––– Let P, f , L , <strong>and</strong> Φ are as described in 1.8. Then, f is<br />

automatically finite of degree d dimP .<br />

Proof. P is a projective variety overÉ<strong>and</strong> f : P → P is a morphism such<br />

that L ⊗d ∼ = f ∗ L for some ample L ∈ Pic(P).<br />

In particular, f ∗ L is ample which implies f is quasi-finite. Since f is a projective<br />

morphism, this is sufficient for finiteness. Further, we have deg L<br />

P ≠ 0<br />

<strong>and</strong> deg f ∗ L P = deg L<br />

P = d dimP deg ⊗d L<br />

P.<br />

□<br />

2.4. Lemma. –––– Let P, f , L , <strong>and</strong> Φ are as described in 1.8. Then,<br />

i) p P Z(<br />

(L ⊗m , ‖ . ‖ ∼ ), . . . , (L ⊗m , ‖ . ‖ ∼ ) ) = 0,<br />

ii) e 1 (h L<br />

) = . . . = e dim P+1 (h L<br />

) = 0.<br />

Proof. i) For every i ∈Æ, write<br />

(<br />

P i := pZ<br />

P (L ⊗m , ‖ . ‖ ∼ (i) ), . . . , (L ⊗m , ‖ . ‖ ∼ (i) )) .<br />

Then, since f is of degree d dimP , Lemma 7.33 shows<br />

(<br />

d dimP P i−1 := pZ<br />

P ( f ∗ L ⊗m , f ∗ ‖ . ‖ ∼<br />

(i−1) ), . . . , ( f ∗ L ⊗m , f ∗ ‖ . ‖ ∼ (i−1) ) ) = 0 .<br />

Applying the isomorphism Φ ⊗m <strong>and</strong> dividing each of the (dim P +1) operators<br />

by d yields<br />

P i−1<br />

d<br />

= ( pP Z (L ⊗m , ‖ . ‖ ∼ (i) ), . . . , (L ⊗m , ‖ . ‖ ∼ (i) )) = P i .<br />

(P i ) i is, therefore, a zero sequence.<br />

ii) Lemma IV.6.3 implies that there is a constant C such that<br />

h L ,0 (x) > C


Sec. 2] THE DYNAMICS OF f 195<br />

for everyx ∈ P. Consequently, h L ,i (x) > C/d i <strong>and</strong> h L<br />

(x) ≥ 0 for everyx ∈ P.<br />

This means<br />

e dim P+1 (h L<br />

) ≥ 0 .<br />

Corollary IV.7.40 together with the inequality<br />

e 1 (h L<br />

) ≥ . . . ≥ e dim P+1 (h L<br />

)<br />

implies the claim.<br />

□<br />

2.5. Remark. –––– As the construction of ‖ . ‖ ∼ depends on the iterated pullbacks<br />

under f , one has to underst<strong>and</strong> the dynamics of the system ( f i ) i≥1 . It will<br />

turn out to be sufficient for our purposes to pay attention to the infinite place.<br />

I.e., to the dynamics of that system on P().<br />

ii. Analyzing the dynamics of f . —<br />

2.6. –––– Let us analyze the special situation described in 1.13.<br />

i) G ⊆ P is an open dense subset being a group scheme such that f has a<br />

restriction m := f | G : G → G which is a homomorphism of group schemes.<br />

Therefore, kerm is a group scheme of finite order d dimP .<br />

As f is surjective, im m is a priori Zariski dense. Since m is a homomorphism,<br />

it is surjective, too.<br />

There is a compact subgroup K ⊆ G() which is both forward <strong>and</strong> backward<br />

invariant under m. In particular, all the finite <strong>groups</strong> ker(m◦ . . . ◦m) are<br />

contained in K.<br />

ii) All eigenvalues of the tangent map T e m have absolute value strictly bigger<br />

than 1. Therefore, on G(), there is a left invariant Riemannian metric µ<br />

such that<br />

q := ‖(T e m) −1 ‖ max < 1 .<br />

Indeed, based on the Jordan normal form, one finds easily a hermitian scalar<br />

product on T e G() satisfying the analogous inequality. We take its real part.<br />

By transport of structure, we find a left invariant Riemannian metric µ on G().<br />

We will denote by δ the metric on G() given by the lengths of the<br />

µ-shortest paths.<br />

2.7. Convention. –––– In order to simplify notation, we will often write f<br />

<strong>and</strong> m instead of f<strong>and</strong> mwhen there is no danger of confusion.


196 DISTRIBUTION OF SMALL POINTS [Chap. V<br />

2.8. Lemma. –––– a) m induces a mapping<br />

which is a diffeomorphism.<br />

m : G()/K → G()/K<br />

b) m is exp<strong>and</strong>ing for the Riemannian metric µ on G()/K induced by µ.<br />

More precisely,<br />

for any z 1 , z 2 ⊆ G()/K.<br />

δ ( m(z 1 ), m(z 2 ) ) ≥ 1 q δ(z 1, z 2 )<br />

Proof. a) m is well-defined since K is forward invariant under m.<br />

Backward invariance of K implies that m is even an injection. Further,<br />

as m : G() → G() is a surjection, the same is then true for m.<br />

Finally, the tangent map<br />

T e m : T e G()/K → T e G()/K<br />

is exp<strong>and</strong>ing, too. In particular, T e m is invertible. This yields that m is a<br />

local diffeomorphism.<br />

b) The preimage m −1 (w) of a path w in G()/K is a path, too. All tangent<br />

maps of m are exp<strong>and</strong>ing by a factor ≥ 1 q . Therefore,<br />

l ( m −1 (w) ) ≤ q · l(w) .<br />

This implies the claim.<br />

□<br />

2.9. Corollary. –––– Let G() = G()∪{∞} be the one-point compactification<br />

of G().<br />

Then, for each x ∈ G()\K, the sequence (x i ) i such that x 0 = x <strong>and</strong> x i+1 = m(x i )<br />

converges to ∞. In other words, K is the Julia set of the dynamical system (m i ) i≥1<br />

on G().<br />

Proof. We have to show that for each compact set L ⊆ G() there are only<br />

finitely many i ∈Æsuch that x i ∈ L. Replacing L by KL, if necessary, we may<br />

assume that L is left K-invariant.<br />

This means, we are given a compact set L ⊆ G()/K <strong>and</strong> have to verify that<br />

x i ∈ L for only finitely many values of i.<br />

As m : G()/K → G()/K is exp<strong>and</strong>ing <strong>and</strong> x 0 ≠ [K ], we have<br />

δ(x i , [K ]) → ∞ .<br />

However, L is a compact set. In particular, L is bounded.<br />


Sec. 2] THE DYNAMICS OF f 197<br />

2.10. Corollary. –––– The union over all the <strong>groups</strong> ker(m ◦ . . . ◦ m) is dense<br />

in K.<br />

Proof. Otherwise, the topological closure of<br />

K ′ := [ ker(m ◦ . . . ◦ m)<br />

} {{ }<br />

i<br />

i times<br />

would be a compact Lie group properly contained in K. It is obvious that K ′ is<br />

forward <strong>and</strong> backward invariant under f. Therefore, the map induced by f<br />

on the compact homogeneous space K/K ′ would be injective <strong>and</strong> exp<strong>and</strong>ing.<br />

□<br />

2.11. Notation. –––– For N ∈Ê, we will write U N (K ) for the set<br />

{x ∈ G() | δ(x, K ) ≤ N } .<br />

2.12. Proposition –––– Let µ i be the measure induced by the smooth differential<br />

form<br />

c 1 (L , ‖ . ‖ ∞) (i) ∧ . . . ∧ c 1 (L , ‖ . ‖ ∞)<br />

(i)<br />

of type (dim P, dimP) on P(). Then, the sequence (µ i ) i converges weakly to<br />

the measure µ which is the zero measure on P()\K <strong>and</strong> the Haar measure of<br />

volume deg L<br />

P on K.<br />

Proof. First step. Some observations.<br />

Each µ i is of volume deg L<br />

P. Further, the sequence (µ i ) i obeys the recursive low<br />

Second step. µ i | P()\K → 0.<br />

µ i+1 = 1<br />

d dim P m∗ µ i .<br />

For this, let g ∈ C 0 (P() \K) be a continuous function with compact support.<br />

We consider the sequence (g i ) i such that g 0 := g <strong>and</strong> g i+1 := 1 f<br />

d dimP ∗ g i . Here, f ∗ ,<br />

the push-forward of a function, is given by summation over the fibers of f .<br />

By definition, µ i (g) = µ 0 (g i ). Thus, let us show µ 0 (g i ) → 0 for i → ∞.<br />

For this, we note that |g| is bounded by some constant C. This implies |g i | ≤ C<br />

for every i. Further, there is some ε > 0 such that supp(g) ⊆ P()\U ε (K ).<br />

Consequently,<br />

Z<br />

|µ i (g)| = |µ 0 (g i )| ≤ C χ (P()\U ε/q i ) dµ 0 .<br />

Here, χ A denotes the characteristic function of a measurable set A. The integr<strong>and</strong><br />

converges monotonically to χ P()\G() which is of integral zero. Therefore,<br />

the Theorem of Beppo Levi yields µ i (g) → 0.<br />

P()


198 DISTRIBUTION OF SMALL POINTS [Chap. V<br />

Third step. The assertion in general.<br />

Let g ∈ C 0 (G()) be a continuous function with compact support. Again, we<br />

put g 0 := g <strong>and</strong> g i+1 := 1 m<br />

d dimP ∗ g i such that we have µ i (g) = µ 0 (g i ). As the sequence<br />

(g i ) i is uniformly bounded, it suffices to show that it converges pointwise<br />

to the constant<br />

I := 1 Z<br />

g dρ .<br />

volK<br />

We claim that (g i ) i converges even uniformly in every compact set A ⊆ G().<br />

For this, we may assume without loss of generality that A = U N (K ) for<br />

some N ∈Ê. Further, note that<br />

g i (x) =<br />

K<br />

1<br />

♯ kerm i<br />

∑<br />

k∈kerm i g(ky)<br />

for any y such that f (y) = x. Here, we clearly have (m i ) −1 (A) ⊆ U Nq i(K )<br />

<strong>and</strong> Nq i → 0 for i → ∞.<br />

In addition, g is uniformly continuous on A <strong>and</strong> m is exp<strong>and</strong>ing. Thus, for each<br />

ε > 0, there is some i ∈Æsuch that, for two arbitrary <strong>points</strong> x, x ′ ∈ A, one<br />

can find y, y ′ such that f (y) = x, f (y ′ ) = x ′ , <strong>and</strong> δ(y, y ′ ) < ε. Consequently,<br />

max g i | A − min g i | A<br />

tends to zero for i → ∞.<br />

Finally, for every i ∈Æ, I is the mean value of g i on K.<br />

□<br />

iii. Some further observations. —<br />

2.13. Caution. –––– It is not true in general that the curvature current, given by<br />

c 1 (L, ‖ . ‖ ∞ ) := 1 ∂∂ log ‖s‖2<br />

2πi<br />

for a non-zero <strong>rational</strong> section of L, vanishes on P() \K. Nevertheless, one<br />

has at least the following.<br />

2.14. Lemma. –––– The curvature current c 1 (L, ‖ . ‖ ∞ ) has the properties below.<br />

a) The restriction c 1 (L, ‖ . ‖ ∞ )| G() is left <strong>and</strong> right K-invariant.<br />

b) It satisfies the equation f ∗ c 1 (L, ‖ . ‖ ∞ ) = d · c 1 (L, ‖ . ‖ ∞ ).<br />

Proof. a) By construction, c 1 (L, ‖ . ‖ ∞,i ) is invariant under ker(m i).<br />

Thus, c 1 (L, ‖ . ‖ ∞ ), which is the weak limit of that sequence of currents,<br />

is invariant under S i ker(m i). Invariance under any k ∈ K follows since<br />

k i → k implies that the test forms k i ω converge to kω in the Schwartz space.<br />

b) is clear. □


Sec. 3] PERTURBING ALMOST SEMIAMPLE METRICS 199<br />

3. Perturbing almost semiample metrics<br />

3.1. –––– One of the most natural ways to produce a new adelic metric from<br />

an old one is to replace ‖ . ‖ ∞ by ‖ . ‖ ∞ · exp(−g) for some continuous function<br />

g ∈ C (P()). As we want to use Corollary IV.7.40 as a fundamental tool,<br />

it is necessary to underst<strong>and</strong> for which g this metric is almost semiample.<br />

3.2. Notation. –––– In this section we will continue to use the notations of 1.13.<br />

Further, we denote by S the set of all continuous functions g ∈ P() such that<br />

‖ . ‖ ∞ · exp(−g) is almost semiample.<br />

3.3. Lemma. –––– a) One has C ∈ S for every constant C.<br />

b) Let g 1 , g 2 ∈ S <strong>and</strong> 0 < a < 1. Then, ag 1 + (1 − a)g 2 ∈ S.<br />

c) If g ∈ S then 1 d f ∗ g ∈ S.<br />

d) Let g ∈ S be a function such that supp g ⊆ G(). Then, for each k ∈ K, one<br />

has g · k ∈ S for<br />

{ g(kx) if x ∈ G() ,<br />

(g · k)(x) :=<br />

0 otherwise .<br />

e) S is closed under uniform convergence.<br />

Proof. e) is trivial.<br />

a) ‖ . ‖ ∞ is almost semiample. Thus, there is a sequence (‖ . ‖ i ∞ ) i of smooth<br />

hermitian metrics with strictly positive curvature on Lwhich is uniformly<br />

convergent to ‖ . ‖ ∞ . The metrics ‖ . ‖ ∞,i · exp(−C ) have the same curvatures<br />

<strong>and</strong> converge uniformly to ‖ . ‖ ∞ · exp(−C ).<br />

b) Let (‖ . ‖ i ′) i <strong>and</strong> (‖ . ‖ i ′′)<br />

i be sequences uniformly convergent versus<br />

‖ . ‖ ∞ · exp(−g 1 ) <strong>and</strong> ‖ . ‖ ∞ · exp(−g 2 ), respectively. Put<br />

‖ . ‖ i := ‖ . ‖ ′a<br />

i<br />

′′ (1−a)<br />

· ‖ . ‖ i .<br />

These metrics have positive curvature <strong>and</strong>, uniformly,<br />

‖ . ‖ i → ‖ . ‖ ∞ · exp ( − ag 1 − (1 − a)g 2<br />

)<br />

.<br />

c) Put ‖ . ‖ ∞ ′ := ‖ . ‖ ∞ · exp(−g). Then,<br />

(<br />

Φ −1 f ∗ ‖ . ‖ ′ d<br />

1<br />

∞ = Φ −1 f ∗ d<br />

‖ . ‖ 1 ∞ · exp − 1 )<br />

d f ∗ g<br />

(<br />

= ‖ . ‖ ∞· exp − 1 )<br />

d f ∗ g .<br />

Therefore, if (‖ . ‖ ∞,i ′ ) i is a sequence of metrics with positive curvature<br />

which converges uniformly to ‖ . ‖ ∞ ′ then (Φ −1 f ∗ ‖ . ‖ ′ 1 d<br />

∞,i) i converges uniformly<br />

to ‖ . ‖ ∞ · exp(− 1 f ∗ g).<br />

d<br />

d) We denote by<br />

e k : G() → G()


200 DISTRIBUTION OF SMALL POINTS [Chap. V<br />

multiplication by k from the left.<br />

By e), we may assume that k ∈ ker(m j ) for some j ∈Æ. To show<br />

‖ . ‖ ∞ · exp(−g · k)<br />

is almost semiample, it suffices to consider the hermitian metric<br />

‖ . ‖ d j<br />

∞ · exp(−d j g · k)<br />

on L.<br />

⊗d j<br />

Let us consider the restriction to G(), first. Iterated application of Φ induces<br />

an isomorphism<br />

e ∗ k L | ⊗d j ∼ G() = e ∗ k ( f j ) ∗ L| ∼ G() = ( f j ) ∗ L| ∼ ⊗d j<br />

G() = L| G() . (∗)<br />

Here, the isomorphism in the middle is canonical as f j ◦e k = f j .<br />

By construction 2.2, ‖ . ‖ ∞ is the uniform limit of a sequence (‖ . ‖ ∞ (i))<br />

i such that,<br />

for every i ≥ j,<br />

e ∗ k<br />

[ ] ‖ . ‖<br />

(i) d j<br />

∞<br />

under the identification (∗). By consequence,<br />

= [ ]<br />

‖ . ‖ ∞<br />

(i) d j<br />

e ∗ k‖ . ‖ d j<br />

∞ = ‖ . ‖ d j<br />

∞ ,<br />

too, which shows<br />

[ ‖ . ‖<br />

d j<br />

∞ · exp(−d j g) ] = ‖ . ‖ d j<br />

∞ · exp(−d j g · k) .<br />

e ∗ k<br />

As almost semiampleness is preserved under holomorphic pull-back, we see that<br />

‖ . ‖ d j<br />

∞ · exp(−d j g · k) is almost semiample on G().<br />

On the other h<strong>and</strong>, there is some neighbourhood U of P() \G() such<br />

that g| U = 0. This implies −d j g · k = 0 <strong>and</strong>, therefore,<br />

‖ . ‖ d j<br />

∞ · exp(−d j g · k) = ‖ . ‖ d j<br />

∞<br />

on U. In particular, ‖ . ‖ d j<br />

∞ · exp(−d j g · k) is almost semiample on U.<br />

The assertion follows from Proposition 3.4.<br />

□<br />

3.4. Proposition –––– Let M be a projective complex manifold, G ∈ Pic(M ) be<br />

ample, <strong>and</strong> {U 1 , . . . , U n } an open covering. Let ‖ . ‖ 1 , . . . , ‖ . ‖ n be continuous<br />

hermitian metrics on G | U1 , . . . , G | Un such that, for some δ > 0, the<br />

sets D 1,δ , . . . , D n,δ , given by<br />

D i,δ := {x ∈ U i | ‖ . ‖ i,(x) ≤ (1 + δ) · min{‖ . ‖ 1,(x) , . . . , ‖ . ‖ n,(x) }} ⊆ U i


Sec. 3] PERTURBING ALMOST SEMIAMPLE METRICS 201<br />

are compact. Assume, each ‖ . ‖ i is almost semiample.<br />

Then, the continuous hermitian metric ‖ . ‖ := min i ‖ . ‖ i is almost semiample.<br />

Proof. We may assume that G is very ample.<br />

For every i, let (‖ . ‖ ij ) j be a sequence of smooth <strong>and</strong> positively curved metrics<br />

on G | Ui which converges uniformly to ‖ . ‖ i . Put<br />

‖ . ‖ − j<br />

:= min{‖ . ‖ 1j , . . . , ‖ . ‖ nj } .<br />

For j ≫ 0, ‖ . ‖ − j<br />

is a continuous hermitian metric on the whole of G . The sequence<br />

(‖ . ‖ − j ) j converges uniformly to ‖ . ‖.<br />

We choose an embedding i : M ֒→ P N such that i ∗ O(1) ∼ = G . Further, we<br />

extend the smooth metrics ‖ . ‖ ij to smooth metrics ‖ . ‖ ′ ij<br />

on O(1) which are<br />

defined on open subsets W i ⊇ U i of P N .<br />

Under the tautological action PGL n () × P N → P N , each γ ∈ PGL n ()<br />

defines an automorphism e γ : P N → P N such that there is a natural identification<br />

eγ ∗ O(1) ∼ = O(1). We find an open neighbourhood O of e ∈ PGL n () such<br />

that, for every γ ∈ O <strong>and</strong> every i ∈ {1, . . . , n}, the pull-back eγ ∗ ‖ . ‖<br />

ij ′ is<br />

well-defined on D i,δ <strong>and</strong><br />

eγ‖ ∗ . ‖ ′ ij | Ui ∩eγ −1 (W i )<br />

has strictly positive curvature in every point of D i,δ .<br />

We claim that, for each γ ∈ O, the perturbation<br />

‖ . ‖ −,γ<br />

j<br />

:= e ∗ γ(<br />

min{‖ . ‖<br />

′<br />

1j , . . . , ‖ . ‖ ′ nj} ) | M<br />

of ‖ . ‖ − j<br />

has a positive curvature current on each holomorphic curve inside M.<br />

Indeed, this is a local statement. Let x ∈ M. Fix a holomorphic section<br />

0 ≠ s ∈ Γ(U,G) defined in some neighbourhood U of x. Then,<br />

c 1 (G , ‖ . ‖ −,γ<br />

j<br />

)| U = −dd c log ( ‖s‖ −,γ ) 2<br />

j<br />

= dd c( max{− logeγ‖s‖ ∗ ′ 1j 2 , . . . , − loge∗ γ‖s‖ nj} ) ′ 2 .<br />

Observe that the maximum of a finite system of plurisubharmonic functions is<br />

plurisubharmonic, again.<br />

Now use Sobolev’s averaging procedure. For every non-negative smooth function<br />

ϕ ≠ 0 on PGL n () such that supp ϕ ⊆ O, one gets an approximation<br />

/<br />

Z<br />

Z<br />

‖ . ‖ −,ϕ<br />

j<br />

:= ϕ(γ) · ‖ . ‖ −,γ<br />

j<br />

dρ γ ϕ(γ)dρ γ<br />

PGL n ()<br />

PGL n ()<br />

of ‖ . ‖ − j . Here, ρ is the left Haar measure on PGL n().


202 DISTRIBUTION OF SMALL POINTS [Chap. V<br />

Consider a sequence (ϕ k ) k of functions as above which converges weakly to the<br />

delta distribution δ e . Every ‖ . ‖ −,ϕ k<br />

j<br />

is smooth <strong>and</strong> positively curved. The sequence<br />

(‖ . ‖ −,ϕ k) k k<br />

converges uniformly to ‖ . ‖.<br />

3.5. Corollary. –––– If g 1 , g 2 ∈ S then g := max{g 1 , g 2 } ∈ S.<br />

3.6. Corollary. –––– Let l ∈Æ, 0 < a < 1, <strong>and</strong> 0 ≠ s ∈ Γ(P(), L). ⊗l Then,<br />

Proof. Consider the function g on<br />

g := max{a log ‖s‖ l , 0} ∈ S .<br />

U := {x ∈ P() | log ‖s‖(x) >−1}<br />

given by g(x) := a log ‖s‖ l (x). We have to show that the hermitian metric<br />

‖ . ‖ ′ ∞ := ‖ . ‖l ∞ · exp(−g)<br />

on L| ⊗l U is the limit of a uniformly convergent sequence of smooth metrics<br />

with positive curvature.<br />

We put ‖ . ‖ ′ ∞,i := ‖ . ‖l ∞,i/ ( ‖s‖ l ∞,i) a. It is clear that, uniformly,<br />

Further,<br />

‖ . ‖ ′ ∞,i → ‖ . ‖ ′ ∞ .<br />

c 1 (L| U , ‖ . ‖ ′ ∞,i ) = ddc (− log ‖s‖ ′2 ∞,i )<br />

= dd c( − log(‖s‖ l ∞,i )2(1−a))<br />

= (1 − a)l · c 1 (L| U , ‖ . ‖ ∞,i ) > 0 . □<br />

□<br />

3.7. Lemma. –––– Let g ∈ S suchthatsupp g ⊆ G(). Consider the K-invariant<br />

function g given by<br />

{<br />

max g(kx) if x ∈ G() ,<br />

g(x) := k∈K<br />

g(x) otherwise .<br />

Then, g ∈ S, too.<br />

Proof. g is the limit of the uniformly convergent sequence (g i ) i , defined by<br />

g i (x) := max g(kx) .<br />

k∈ker(m i )<br />


Sec. 3] PERTURBING ALMOST SEMIAMPLE METRICS 203<br />

3.8. Proposition –––– Let ε > 0 <strong>and</strong> U ⊆ G() an open set containing K.<br />

Then, there is some non-negative K-invariant function 0 ≠ g ∈Ê+·S such that<br />

i) g(e) = 1,<br />

ii) max x∈P() g(x) ≤ 1 + ε, <strong>and</strong><br />

iii) supp g ⊆ U.<br />

Proof. Put D := P \G. For some j ≫ 0, the coherent sheaf L ⊗j ⊗I D has a<br />

section which does not vanish in e ∈ G. I.e., there is a section s ∈ Γ(P, L ⊗j )<br />

vanishing in D but not in e.<br />

Using Corollary 3.6, we see<br />

1<br />

}<br />

˜g C := max{<br />

2j log ‖Cs‖j , 0<br />

{ 1<br />

= max<br />

2j log ‖s‖j + logC }<br />

, 0 ∈ S<br />

2j<br />

for every C > 0. It is clear that, for each C, supp ˜g C ⊆ G() is a compact set.<br />

We put<br />

<strong>and</strong><br />

A := max<br />

x∈K<br />

1<br />

2j log ‖s‖j (x)<br />

1<br />

B := max<br />

x∈P() 2j log ‖s‖j (x) .<br />

Then, we choose C such that log C + A > 0 <strong>and</strong><br />

logC + B<br />

log C + A ≤ 1 + ε .<br />

Further, let g 0 be the K-invariant function associated to ˜g C . I.e.,<br />

{<br />

max ˜g C (kx) if x ∈ G() ,<br />

g 0 := k∈K<br />

g(x) otherwise .<br />

By construction, g 0 (e) > 0 <strong>and</strong><br />

max g 0 (x)/g 0 (e) ≤ 1 + ε .<br />

x∈P()<br />

Further, supp g 0 ⊆ K ·supp ˜g C is compact. Lemma 3.7 shows that g 0 ∈ S.<br />

Finally, we define a sequence (g i ) i of functions on P() by putting, recursively,<br />

g i+1 := 1 d f ∗ g i . Then, one clearly has g i (e) = g 0 (e) > 0 <strong>and</strong><br />

max g i(x)/g i (e) = max g 0(x)/g 0 (e) ≤ 1 + ε<br />

x∈P() x∈P()<br />

for every i ∈Æ. Lemma 3.3.c) implies that g i ∈ S.<br />

As supp g 0 is compact, there is some N ∈Êsuch that supp g 0 ⊆ U N (K ).<br />

This yields, by Lemma 2.8.b), that supp g i ⊆ U Nq i(K ).<br />

Therefore, the function g, given by g(x) := g i (x)/g i (e) for some i ≫ 0, has all<br />

the properties desired.<br />


204 DISTRIBUTION OF SMALL POINTS [Chap. V<br />

4. Equidistribution<br />

4.1. Definition. –––– Let X be a projective variety overÉ. Then, a sequence<br />

(x i ) i of closed <strong>points</strong> on X is called generic if no infinite subsequence is<br />

contained in a proper closed subvariety of X.<br />

4.2. Remark. –––– In other words, (x i ) i is generic if it converges to the<br />

generic point with respect to the Zariski topology.<br />

4.3. Definition. –––– Let X be a projective variety overÉ<strong>and</strong> L ∈ ˜Pic(X ).<br />

Suppose that e dim X+1 (h L<br />

) = 0.<br />

Then, a sequence (x i ) i of closed <strong>points</strong> on X is called small if h L<br />

(x i ) → 0.<br />

4.4. Lemma. –––– Let X be a projective variety overÉ<strong>and</strong><br />

L = (L , ‖ . ‖) ∈ ˜Pic(X )<br />

be such that ‖ . ‖ p is semipositive for every prime number p <strong>and</strong> ‖ . ‖ ∞ is almost semiample.<br />

Assume e dim X+1 (L ) = 0.<br />

Then, pZ X (L , . . . , L ) = 0 is equivalent to the existence of a sequence of closed<br />

<strong>points</strong> on X which is generic <strong>and</strong> small.<br />

Proof. “=⇒” There are not more than countably many closed subvarieties<br />

X 1 , X 2 , . . . ⊂ X. Further, Corollary IV.7.40 implies e 1 (h L<br />

) = 0. Therefore,<br />

we may choose a sequence (x i ) i of closed <strong>points</strong> on X such that<br />

x i ∈ X \X 1 \X 2 \ . . . \X i<br />

<strong>and</strong> h L<br />

(x i ) < 1 i . (x i) i is generic <strong>and</strong> small.<br />

“⇐=” Let (x i ) i be a sequence of closed <strong>points</strong> on X which is generic <strong>and</strong> small.<br />

Let Y ⊂ X be a closed subset of codimension one. Then, only finitely many<br />

of the x i are contained in Y. Hence, inf x∈X \Y h L<br />

(x) = 0. As this is true for<br />

every Y, we see e 1 (h L<br />

) = 0. Corollary IV.7.40 yields pZ X (L , . . . , L ) = 0.<br />

□<br />

4.5. Lemma. –––– Let P, f , L = (L ⊗m , ‖ . ‖ ∼ ), <strong>and</strong> Φ be as described<br />

in 1.13. Denote by S the set of all continuous functions g ∈ C (P()) such that<br />

‖ . ‖ ∞· exp(−g) is almost semiample.<br />

Let ϕ ∈ C (P()). Then, for every ε > 0, there exist a function ϕ 1 ∈ C (P())<br />

supported in P()\K <strong>and</strong> a function ϕ 2 ∈Ê+·S such that<br />

‖ϕ − ϕ 1 − ϕ 2 ‖ max < ε .


Sec. 4] EQUIDISTRIBUTION 205<br />

Proof. The set<br />

T := {h ∈ C (K ) | h = ϕ| K for some ϕ ∈Ê+·S}<br />

fulfills all the assumptions of Sublemma 4.6, except for closedness under uniform<br />

convergence.<br />

Indeed, Lemma 3.3.a) <strong>and</strong> b) imply that assumptions i) <strong>and</strong> ii) are satisfied.<br />

Corollary 3.6 is general enough to guarantee assumption iii). Lemma 3.3.d)<br />

yields assumption iv). Finally, Corollary 3.5 implies that assumption v) is fulfilled.<br />

Therefore, there is some ϕ 2 ∈Ê+·S such that |ϕ(x)−ϕ 2 (x)| < ε for every x ∈ K.<br />

2<br />

A decomposition of the unit adapted to a suitable open covering {P()\K, U }<br />

of P() yields some ϕ 1 ∈ C (P()) such that supp ϕ 1 ⊆ P()\K <strong>and</strong><br />

∣ ( ϕ(x) − ϕ 2 (x) ) − ϕ 1 (x) ∣ < ε<br />

for every x ∈ P().<br />

□<br />

4.6. Sublemma. –––– Let L be a compact group <strong>and</strong> T ⊆ C (L) a set of continuous<br />

real-valued functions such that the following conditions are fulfilled.<br />

i) T contains all the constant functions.<br />

ii) T is closed under addition <strong>and</strong> multiplication by positive constants.<br />

iii) For each x ∈ L such that x ≠ e, there is some g ∈ T fulfilling<br />

g(e) > g(x) .<br />

iv) For each x ∈ L <strong>and</strong> g ∈ T , the shift g · x is also in T .<br />

v) If g ∈ T then g + ∈ T for<br />

g + (x) := max {g(x), 0} .<br />

vi) T is closed under uniform convergence.<br />

Then, T = C (L) contains all continuous functions.<br />

Proof. We fix a Haar measure ρ on L. Then, the conditions ii,) iv) <strong>and</strong> vi)<br />

together imply the following.<br />

(†) If h ∈ T <strong>and</strong> g is a non-negative, measurable, <strong>and</strong> bounded function then,<br />

for the convolution, we have g∗h ∈ T .


206 DISTRIBUTION OF SMALL POINTS [Chap. V<br />

Hence, it suffices to show that, for each open set U ⊆ L containing e, there<br />

is some non-negative function g U ∈ T such that g U ≠ 0 <strong>and</strong> supp g U ⊆ U.<br />

For this, by i), ii), <strong>and</strong> v), we only need a function g ∈ T having its maximum<br />

entirely in U.<br />

Assume, for some open U 0 , there would be no such function. Then, for<br />

every g ∈ T having its maximum in e, there is a non-empty set<br />

A g := {y ∈ L\U 0 | g(y) = g(e)} ⊆ L\U 0<br />

where this maximum is taken, too. If<br />

\<br />

g∈T<br />

A g = ∅<br />

then, by compactness, already a finite intersection A g1 ∩ . . . ∩ A gn would<br />

be empty. Then, g 1 + . . . + g n ∈ S had its maximum in U 0 , only.<br />

Consequently,<br />

is a non-empty set. Put<br />

A U0 := \<br />

g∈T<br />

A := {e} ∪ [<br />

A g<br />

U ∋e,U open<br />

By construction, this set has the property below.<br />

A U .<br />

(‡) Every function g ∈ T assuming its maximum in e is necessarily constant<br />

on A. Further, A ⊆ L is the largest set containing e with this property.<br />

The property (‡) implies that A is a closed set.<br />

For y ∈ A <strong>and</strong> g ∈ T , the shift g·y −1 has its maxima in yA. Therefore, A = yA<br />

for every y ∈ A. This means that A ⊆ L is a subgroup.<br />

Fix a Haar measure ρ A on A <strong>and</strong> denote by i : A → L the natural inclusion.<br />

Property (‡) together with i), ii) <strong>and</strong> v) implies that there is a sequence of<br />

functions of T such that the associated distributions converge weakly to i ∗ ρ A .<br />

Then, property (†) guarantees together with vi) that every A-left invariant<br />

continuous function is an element of T .<br />

Now, we use condition iii). Choose some x 0 ∈ A different from e. There exists<br />

some g ∈ T such that g(e) > g(x 0 ). Shifting by an element of A, if necessary,<br />

we may assume that g(e) = max x∈A g(x).<br />

Define another continuous function M on L by<br />

M (x) := − max<br />

a∈A g(ax) .


Sec. 4] EQUIDISTRIBUTION 207<br />

Obviously, M is A-invariant. Thus, M ∈ T . Consequently, ˜g := g + M ∈ T ,<br />

too.<br />

However, ˜g(x) ≤ 0 for every x ∈ L, ˜g(e) = 0, <strong>and</strong> ˜g(x 0 ) < 0 together form a<br />

contradiction to property (‡).<br />

□<br />

4.7. Proposition. –––– Let P, f , L , <strong>and</strong> Φ be as described in 1.13.<br />

Let (x i ) i be a sequence of closed <strong>points</strong> on P which is generic <strong>and</strong> small. Assume, some<br />

non-negative ϕ ∈ C (P()) fulfills either supp ϕ ⊆ P()\K or ϕ ∈Ê+·S.<br />

Then,<br />

lim inf<br />

i→∞<br />

Z<br />

P()<br />

ϕ dδ xi ≥<br />

Z<br />

P()<br />

ϕ dτ .<br />

Here, τ is the zero measure on P() \K <strong>and</strong> the Haar measure of volume one on K.<br />

Proof. For ϕ such that supp ϕ ⊆ P() \K, the right h<strong>and</strong> side is zero. Thus, in<br />

this case, the assertion is clear.<br />

Let ϕ ∈Ê+·S. Then, for any positive λ ∈Ê, let ‖ . ‖ ∼ λ be the adelic metric<br />

on L ⊗m given by ‖ . ‖ λ p := ‖ . ‖ p for the non-archimedean valuations <strong>and</strong><br />

by ‖ . ‖ ∞ λ := ‖ . ‖ ∞ · exp(−λϕ) for the archimedean valuation.<br />

For λ → 0, the adelic metric ‖ . ‖ λ fulfills the assumptions of Corollary IV.7.40.<br />

Further, one clearly has<br />

Z<br />

h (L ⊗m ,‖ .‖ λ )(x i ) = h L<br />

(x i ) + λ ϕ dδ xi .<br />

As (x i ) i is generic <strong>and</strong> h L<br />

(x i ) → 0, we obtain, according to the very definition<br />

of e 1 ,<br />

Z<br />

λ·liminf ϕ dδ xi ≥ e 1 (h (L ⊗m ,‖.‖ λ )) . (§)<br />

On the other h<strong>and</strong>,<br />

=<br />

1<br />

(dim X + 1)c 1 (L ⊗m<br />

i→∞<br />

X ()<br />

É)<br />

dim X pX Z<br />

1<br />

(dim X + 1)m dimX deg L<br />

X pX Z (L , . . . , L )<br />

X ()<br />

(<br />

(L ⊗m , ‖ . ‖ λ ∼ ), . . . , (L ⊗m , ‖ . ‖ λ ∼ ))<br />

1<br />

+<br />

m dim X deg L<br />

X ·<br />

(<br />

(OX , ‖ . ‖ ∞· exp(−λϕ)), (L ⊗m , ‖ . ‖ ∼ ), . . . , (L ⊗m , ‖ . ‖ ∼ ) )<br />

+ O(λ 2 ) .<br />

p X Z


208 DISTRIBUTION OF SMALL POINTS [Chap. V<br />

By virtue of Lemma 4.4, we have p X Z (L , . . . , L ) = 0. Further,<br />

1 (<br />

m dimX deg L<br />

X ·pX Z (OX , ‖ . ‖ ∞· exp(−λϕ)), (L ⊗m , ‖ . ‖ ∼ ), . . . , (L ⊗m , ‖ . ‖ ∼ ) )<br />

= λ lim<br />

i→∞<br />

I i (ϕ)<br />

for<br />

I i (ϕ) := 1 ( )<br />

m dim X pX Z (OX , ‖ . ‖ ∞· exp(−ϕ)), (L ⊗m , ‖ . ‖ ∼ (i) ), . . . , (L ⊗m , ‖ . ‖ ∼ (i) ) .<br />

} {{ }<br />

dim X times<br />

At this level, we can make the arithmetic intersection product explicit <strong>and</strong> find<br />

I i (ϕ) =<br />

Z<br />

P()<br />

ϕc 1 (L , ‖ . ‖ (i)<br />

∞) ∧ . . . ∧ c 1 (L , ‖ . ‖ (i)<br />

∞) .<br />

Corollary IV.7.40 therefore yields in view of Proposition 2.12<br />

Z<br />

e 1 (h (L ⊗m ,‖.‖ λ )) ≥ λ· ϕ dτ . ()<br />

P()<br />

The equations (§) <strong>and</strong> () together imply the assertion.<br />

4.8. Remark. –––– The argument for the case ϕ ∈Ê+·S is due to E. Ullmo.<br />

Cf. [S/U/Z] or [Zh98].<br />

4.9. Theorem (Equidistribution on P()). —– Let P, f , L , <strong>and</strong> Φ be as described<br />

in 1.13.<br />

Then, for each sequence (x i ) i of closed <strong>points</strong> on P which is generic <strong>and</strong> small, the<br />

associated sequence of <strong>measures</strong> (δ xi ) i converges weakly to the measure τ which is the<br />

zero measure on P()\K <strong>and</strong> the Haar measure of volume one on K.<br />

Proof. First step. δ xi | P()\K → 0.<br />

By Proposition 3.8, for every ε 1 , ε 2 > 0, there is some non-negative g ε1 ,ε 2<br />

∈ S<br />

such that supp g ε1 ,ε 2<br />

⊆ U ε1 (K ),<br />

□<br />

<strong>and</strong><br />

Then, by Proposition 4.7,<br />

max<br />

x∈P() g ε 1 ,ε 2<br />

≤ 1 + ε 2 ,<br />

Z<br />

P()<br />

Z<br />

lim inf<br />

i→∞<br />

P()<br />

g ε1 ,ε 2<br />

dτ = 1 .<br />

g ε1 ,ε 2<br />

dδ xi ≥ 1 .


Sec. 4] EQUIDISTRIBUTION 209<br />

This yields<br />

But<br />

Consequently,<br />

Z<br />

lim sup<br />

i→∞<br />

1 − g ε1 ,ε 2<br />

(x) ≥<br />

P()<br />

(1 − g ε1 ,ε 2<br />

) dδ xi ≤ 0 .<br />

{ 1 if x ∈ P()\U ε1 (K ) ,<br />

−ε 2 if x ∈ U ε1 (K ) .<br />

This shows<br />

lim sup<br />

i→∞<br />

[<br />

δxi<br />

(<br />

P()\U ε1 (K ) ) − ε 2 δ xi<br />

(<br />

Uε1 (K ) )] ≤ 0 .<br />

lim sup<br />

i→∞ δ x i<br />

(<br />

P()\U ε1 (K ) ) ≤ ε 2 .<br />

As the latter is true for every ε 2 > 0, we have<br />

lim δ (<br />

x i P()\U ε1 (K ) ) = 0 .<br />

i→∞<br />

Further, this formula is true for each ε 1 > 0. Hence, (δ xi | P()\K ) i converges<br />

weakly to the zero measure.<br />

Second step. The assertion in general.<br />

Let ϕ ∈ C (P()) be an arbitrary continuous function. As we can interchange<br />

the roles of ϕ <strong>and</strong> (−ϕ), it suffices to prove<br />

Z<br />

Z<br />

lim inf ϕ dδ xi ≥ ϕ dτ.<br />

i→∞<br />

P()<br />

By Lemma 4.5, we have continuous functions ϕ 1 <strong>and</strong> ϕ 2 on P() such that<br />

supp ϕ 1 ⊆ P()\K <strong>and</strong> ϕ 2 ∈Ê+·S.<br />

The result of the first step shows<br />

lim<br />

Z<br />

i→∞<br />

P()<br />

i→∞<br />

P()<br />

P()<br />

ϕ 1 dδ xi = 0 .<br />

Further, by Proposition 4.7,<br />

Z<br />

lim inf ϕ 2 dδ xi ≥<br />

Consequently,<br />

Z<br />

lim inf ϕdδ xi ≥<br />

i→∞<br />

P()<br />

Z<br />

P()<br />

The assertion follows since ε > 0 is arbitrary.<br />

Z<br />

P()<br />

ϕ 2 dτ .<br />

ϕdτ − ε .<br />

4.10. Corollary. –––– Let X ⊂ P be a closed subvariety <strong>and</strong> (x i ) i a sequence of<br />

closed <strong>points</strong> on X which is generic <strong>and</strong> small.<br />


210 DISTRIBUTION OF SMALL POINTS [Chap. V<br />

Then, the sequence (δ xi | X ()\K) i converges weakly to the zero measure.<br />

Proof. As there are only finitely many i such that x i ∉ G, let us assume that<br />

the whole sequence (x i ) i is contained in G. Then, we choose some K-invariant<br />

metric on G() <strong>and</strong> assume that, for some ε > 0, we would have<br />

lim inf<br />

i→∞ δ x i<br />

(<br />

X ()\U ε (K ) ) = δ > 0 .<br />

We will construct another sequence (y i ) i which is generic <strong>and</strong> small on the<br />

whole of P.<br />

For this, note first that we have<br />

h L<br />

(x) = 1 ( )<br />

h<br />

deg f L f (x) .<br />

Consequently, h L<br />

is invariant under shift by any torsion point<br />

t ∈ K tor =<br />

j∈Æker(m [ j ) .<br />

Further, all the sequences (t·x i ) i fulfill<br />

lim inf<br />

i→∞ δ t·x i<br />

(<br />

P()\U ε (K ) ) = δ .<br />

Finally, K ⊆ G() is assumed to be Zariski dense. Therefore, for each i, the<br />

union S {t·x i } is a Zariski dense subset of P.<br />

t∈K tor<br />

Finally, we make use of the fact there are only countably many proper closed<br />

subvarieties P 0 , P 1 , P 2 , . . . ⊂ P. We can choose a sequence (y i ) i such that, for<br />

every i ∈Æ, one has<br />

y i ∈ [<br />

{t·x i }<br />

<strong>and</strong><br />

t∈K tor ,i≥0<br />

y i ∈ P \P 0 \ . . . \P i .<br />

Then,<br />

δ yi<br />

(<br />

P()\U ε (K ) ) > δ/2<br />

for i ≫ 0 <strong>and</strong> h L<br />

(y i ) → 0. This is a contradiction to Theorem 4.9.<br />

□<br />

4.11. Remark. –––– This corollary shows, in particular, the following.<br />

If X ∩ K = ∅ then<br />

p X Z (L , . . . , L ) > 0 .<br />

In other words, there are no small <strong>and</strong> generic sequences on X.


CHAPTER VI<br />

CONJECTURES ON THE ASYMPTOTICS OF<br />

POINTS OF BOUNDED HEIGHT<br />

I have no satisfaction in formulas unless I feel their numerical magnitude.<br />

WILLIAM THOMSON 1ST BARON KELVIN,<br />

(Life (1943) by Sylvanus Thompson, p. 827)<br />

1. A heuristic<br />

1.1. –––– Let X be a projective variety overÉ. Then, one of the most natural<br />

questions to ask is<br />

(∗) Does X have finitely many or infinitely manyÉ-<strong>rational</strong> <strong>points</strong>?<br />

1.2. –––– If the answer is that #X (É) < ∞ then one might ask for the precise<br />

number. Otherwise, ifX(É) is infinite then one may ask for the asymptotics<br />

of the <strong>rational</strong> <strong>points</strong> with respect to a certain height function H on X (É).<br />

(†) What is the asymptotics of the function N X,H given by<br />

for B → ∞?<br />

N X,H (B) := #{x ∈ X (É) | H(x) < B}<br />

1.3. –––– If X is a hypersurface of degree d in P n then there is a statistical<br />

heuristic for the asymptotics of N X,Hnaive .<br />

1.4. Statistical heuristic. –––– Let X be a hypersurface of degree d in P nÉ.<br />

Then,<br />

for some positive constant C.<br />

N X,Hnaive (B) ∼ C ·B n+1−d<br />

“Proof.” On P n , the total number ofÉ-<strong>rational</strong> <strong>points</strong> of height < B is ∼B n+1 .<br />

Indeed, these <strong>points</strong> may be given in the form<br />

(x 0 : . . . : x n )


212 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

for x i ∈such that x i ∈ [−B, B]. There are ∼ B n+1 such (n + 1)-tuples <strong>and</strong><br />

the probability that gcd(x 0 , . . . , x n ) = 1 is positive.<br />

Further, X is given by a homogeneous form F of degree d. Its range of values<br />

on [−B, B] n+1 is, a priori, [−DB d , DB d ] for a certain positive constant D.<br />

We assume that the values of F are equally distributed in that range. The value<br />

zero is then hit ∼B n+1−d times.<br />

□<br />

1.5. Remark. –––– This heuristic is definitely an interesting guideline about<br />

which behaviour to expect. It is, however, too naive in oder to work well in<br />

all cases.<br />

1.6. Example (Too few <strong>points</strong> – p-adic unsolvability). —–<br />

Consider the cubic fourfold X in P 5Égiven by the equation<br />

x 3 + 7y 3 + 49z 3 + 2u 3 + 14v 3 + 98w 3 = 0 .<br />

Then, X (É7) = ∅ which implies X (É) = ∅.<br />

Note that the statistical heuristic would predict cubic growth for the number<br />

ofÉ-<strong>rational</strong> <strong>points</strong> x on X such that H naive (x) < B.<br />

1.7. Example (Too few <strong>points</strong> – The Brauer-Manin obstruction). —–<br />

Consider the cubic surface X in P 3Égiven by the equation<br />

Here,<br />

w(x + w)(12x + w) =<br />

3<br />

∏<br />

i=1<br />

(<br />

x + θ (i) y + (θ (i) ) 2 z ) .<br />

θ := −38 + ∑<br />

i∈(∗ 19 ) 3 ζ i 19 ∈É(ζ 19 )<br />

<strong>and</strong> θ (i) are the three conjugates of θ inÉ.<br />

As shown in Example III.5.24, X (É) ≠ ∅ but X (É) = ∅. On X, there is a<br />

Brauer-Manin obstruction to the Hasse principle.<br />

In this case, the statistical heuristic would predict linear growth for the number<br />

ofÉ-<strong>rational</strong> <strong>points</strong> x on X such that H naive (x) < B.<br />

1.8. Examples (Too many <strong>points</strong> – Accumulating subvarieties). —–<br />

i) Consider the smooth threefold X ⊂ P 4Égiven by<br />

x 4 + y 4 = z 4 + v 4 + w 4 .<br />

Then, the statistical heuristic predicts linear growth. I.e., there should be ∼CB<br />

<strong>rational</strong> <strong>points</strong> of height


Sec. 1] A HEURISTIC 213<br />

ii) Other examples of interest are provided by the diagonal cubic threefolds<br />

V a,1 ∈ P 4Éof the form<br />

V a,1 : ax 3 = y 3 + z 3 + v 3 + w 3 .<br />

Here, quadratic growth is predicted by the statistical heuristic.<br />

However, V a,1 clearly contains the lines given by v = −w <strong>and</strong><br />

(x : y : z) = (x 0 : y 0 : z 0 )<br />

for (x 0 : y 0 : z 0 ) aÉ-<strong>rational</strong> point on the curve<br />

E a : ax 3 = y 3 + z 3 .<br />

This is a twisted Fermat cubic. One has the <strong>rational</strong> point (0 : 1 : (−1)) ∈ E a (É).<br />

Therefore, E a is an elliptic curve overÉ. The lines described form a cone over<br />

an elliptic curve. There are ∼B 2É-<strong>rational</strong> <strong>points</strong> of height < B alone on<br />

this cone.<br />

Examples of twisted Fermat cubics of rank up to 3 have been known to<br />

Ernst S. Selmer as early as 1951 [Sel]. For example, E 6 (É) is of rank one,<br />

generated by (21 : 17 : 37).<br />

1.9. Remark. –––– These examples grew out of our study of diagonal cubic <strong>and</strong><br />

quartic threefolds which is described in Chapter VII.<br />

1.10. Geometric interpretation. –––– The exponent n + 1 − d appearing in<br />

the statistical heuristic may be positive, zero, or negative. According to this<br />

distinction, there are three cases.<br />

This distinction into three cases coincides perfectly well with the Kodaira classification<br />

of projective varieties into Fano varieties, varieties of intermediate type,<br />

<strong>and</strong> varieties of general type. Indeed, the anticanonical divisor (−K ) on a<br />

degree d hypersurface in P n is exactly O(n + 1 − d).<br />

1.11. Remark. –––– The Kodaira classification is a purely geometric one.<br />

It does not make use of any arithmetic information on the projective variety<br />

considered. Only the complex algebraic variety produced by base change<br />

tois exploited.<br />

It is a very remarkable observation that whether a projective variety X is Fano,<br />

of intermediate type, or of general type seemingly has a lot of influence on the<br />

set ofÉ-<strong>rational</strong> <strong>points</strong> on X.


214 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

1.12. –––– More concretely, expectations are as follows.<br />

First case. X is a variety of general type.<br />

By definition, this means that the canonical invertible sheaf K is ample.<br />

In the case of a hypersurface, this corresponds to the case that n + 1 − d < 0.<br />

Here, the statistical heuristic is rather illogical. It states that for a large height<br />

bound we expect less <strong>points</strong> than for a small one. In the limit for B → ∞, we<br />

have B n+1−d → 0.<br />

One would therefore expect only very fewÉ-<strong>rational</strong> <strong>points</strong> on X. This is<br />

exactly the content of the conjecture of Lang.<br />

Second case. X is a variety of intermediate type.<br />

In the case of a hypersurface, this corresponds to the case that n + 1 − d = 0.<br />

Here, the statistical heuristic states that there should be a constant number of<br />

<strong>points</strong>, independently of the height bound B.<br />

One would therefore expect that there are a fewÉ-<strong>rational</strong> <strong>points</strong> on X. A more<br />

precise statement is given by the conjecture of Batyrev <strong>and</strong> Manin.<br />

Third case. X is a Fano variety.<br />

By definition, this means that the anticanonical invertible sheaf K ∨ is ample.<br />

In the case of a hypersurface, this corresponds to the case that n + 1 − d > 0.<br />

Here, one would expect that there are a lot ofÉ-<strong>rational</strong> <strong>points</strong> on X. A by far<br />

more precise formulation is given by the conjecture of Manin.<br />

2. The conjecture of Lang<br />

2.1. –––– The conjecture of Langdeals with the case of a variety of general type.<br />

There are actually several versions of it [Lan86].<br />

2.2. Conjecture (Lang). —– Let X be a smooth, projective variety of general type<br />

over a number field K. Then,<br />

i) (Weak Lang conjecture)<br />

the set of <strong>rational</strong> <strong>points</strong> X (K ) is not Zariski dense in X.<br />

ii) (Strong Lang conjecture)<br />

Thereis aZariski closed subset Z ⊂ X such that, for any finite field extension L ⊇ K,<br />

one has that X (L)\Z(L) is finite.<br />

2.3. Conjecture (Geometric Lang conjecture). —–<br />

LetX beasmooth, projective variety of generaltype over a field K of characteristic 0.<br />

Then, there is a proper Zariski closed subset Z(X ) ⊂ X, called the Langian exceptional<br />

set, which is the union of all positive dimensional subvarieties which are not<br />

of general type.


Sec. 2] THE CONJECTURE OF LANG 215<br />

2.4. Remark. –––– These conjectures are strongly interrelated. If the geometric<br />

Lang conjecture is true then the weak Lang conjecture implies the strong<br />

Lang conjecture.<br />

2.5. –––– Lang’s conjectures are known to be true when X is a subvariety of<br />

an abelian variety. This was proven by G. Faltings in [Fa91].<br />

Faltings’ 1991 result includes the case that X is a curve of general type. I.e., a<br />

curve of genus g ≥ 2. This particular case of Lang’s conjecture has been popular<br />

for decades as the Mordell conjecture. The Mordell conjecture was proven<br />

by G. Faltings in 1983 [Fa83].<br />

2.6. Examples. –––– If X is a curve then there is no Langian exceptional set Z.<br />

Weak <strong>and</strong> strong Lang conjecture are therefore equivalent.<br />

This is no longer the case for surfaces of general type.<br />

i) For example, the quintic surface given by<br />

x 5 + y 5 + z 5 + w 5 = 0<br />

in P 3Écontains the line “x = −y, z = −w” on which there are infinitely many<br />

É-<strong>rational</strong> <strong>points</strong>.<br />

ii) Another example of a surface of general type containing a projective line is<br />

provided by the Godeaux surface [Bv, Example X.3.4].<br />

iii) Let V 3<br />

a,b be the cubic threefold in P4Égiven by<br />

ax 3 = by 3 + z 3 + v 3 + w 3 .<br />

Then, the moduli space L a,b of the lines onV a,b is a surface of general type [Cl/G,<br />

Lemma 10.13].<br />

If the cubic curve<br />

E a,b : ax 3 + by 3 + z 3 = 0<br />

contains a <strong>rational</strong> point then, on V a,b , there are the lines given by v = −w<br />

<strong>and</strong> (x : y : z) = (x 0 : y 0 : z 0 ) for (x 0 : y 0 : z 0 ) ∈ E a,b (É). We call these lines the<br />

obvious lines on V a,b .<br />

Inother words, ifE a,b (É) ≠ ∅thenthesurfaceL a,b , whichisofgeneraltype, containsacopyoftheelliptic<br />

curveE a,b . ThereareinfinitelymanyÉ-<strong>rational</strong> <strong>points</strong><br />

on E a,b , for example for a = 6 <strong>and</strong> b = 1.<br />

2.7. An experiment. –––– We searched systematically forÉ-<strong>rational</strong> lines on<br />

the cubic threefolds V a,b for a, b = 1, . . . , 100, a ≥ b. Our method is<br />

described in detail in Chapter VII, Section 4. It guarantees that every line which<br />

contains a point of height


216 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

The results may be interpreted as providing numerical evidence for Lang’s conjecture.<br />

Indeed, aÉ-<strong>rational</strong> point on L a,b corresponds to aÉ-<strong>rational</strong> line<br />

on V a,b . Points on the elliptic curves lying on L a,b correspond to the obvious<br />

lines.<br />

In agreement with Lang’s conjecture, only very few non-obvious lines<br />

were found. On all the varieties V a,b for a, b = 1, . . . , 100, a ≥ b together,<br />

there are only 42 non-obvious lines containing a point of height 0, there exists a Zariski open subset X ◦ ⊆ X such that<br />

for B → ∞.<br />

N X ◦ ,H L<br />

(B) = #{x ∈ X ◦ (k) | H L (x) < B} ≪ B r+ε<br />

3.2. –––– Here, H L is the exponential of the height h L defined by L as<br />

introduced in Definition IV.2.10. It is determined only up to a factor which is<br />

bounded above <strong>and</strong> below by positive constants.<br />

The conjecture is consistent with these changes of the height function since B r+ε<br />

<strong>and</strong> (CB) r+ε differ by a constant factor.


Sec. 3] THE CONJECTURE OF BATYREV AND MANIN 217<br />

3.3. Remark. –––– This conjecture was first formulated by V. V. Batyrev <strong>and</strong><br />

Yu. I. Manin in [B/M]. An excellent presentation may be found in the survey<br />

article [Pe02] by E. Peyre.<br />

3.4. Fact (Varieties of general type). —– The conjecture of Batyrev <strong>and</strong> Manin<br />

implies the weak Lang conjecture.<br />

Proof. Indeed, if X is a variety of general type then the canonical invertible<br />

sheaf K itself is ample <strong>and</strong> we may work with L := K .<br />

Then, K + (−1)L = 0 is an effectiveÊ-divisor. Hence, for every ε > 0, the<br />

conjecture of Batyrev <strong>and</strong> Manin yields that<br />

#{x ∈ X ◦ (k) | H L (x) < B} ≪ B −1+ε<br />

for a suitable Zariski open subset X ◦ ⊂ X. In the limit for B → ∞, this shows<br />

that X ◦ (k) is empty.<br />

The set X (k) is therefore not Zariski dense in X.<br />

3.5. –––– Let X be a smooth Fano variety. Then, the anticanonical invertible<br />

sheaf K ∨ is ample <strong>and</strong> we may consider an anticanonical height which is defined<br />

by L := K ∨ .<br />

3.6. Fact (Fano varieties). —– For X a smooth Fano variety, the conjecture of<br />

Batyrev <strong>and</strong> Manin yields that<br />

N X ◦ ,H K ∨(B) = #{x ∈ X ◦ (k) | H K ∨(x) < B} ≪ B 1+ε<br />

□<br />

for a suitable Zariski open subset X ◦ ⊂ X.<br />

Proof. In this situation, K + (−K ) = 0 is an effectiveÊ-divisor.<br />

□<br />

3.7. Remarks. –––– i) In the case of a Fano hypersurface, this assertion fits<br />

perfectly well with the statistical heuristic.<br />

ii) However, for Fano varieties, the conjecture of Manin describes the growth<br />

of N X ◦ ,H K ∨<br />

much more precisely. The most interesting case of the conjecture<br />

of Batyrev <strong>and</strong> Manin is, therefore, that of a variety of intermediate type.<br />

ii. Varieties of intermediate type. —<br />

3.8. –––– Let X be a smooth, projective, minimal surface of Kodaira dimension<br />

0. Fix an ample invertible sheaf L ∈ Pic(X ).<br />

Then, the conjecture of Batyrev <strong>and</strong> Manin states that, for every ε > 0, there<br />

exists a Zariski open subset X ◦ ⊂ X such that<br />

N X ◦ ,H L<br />

(B) = #{x ∈ X ◦ (k) | H L (x) < B} ≪ B ε .


)⊗Ê<br />

218 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

Indeed, 12K is linearly equivalent to zero in any of the four cases of the Kodaira<br />

classification. Hence,<br />

[K ] ∈ NS(X<br />

is the class of an effectiveÊ-divisor on X.<br />

3.9. Example. –––– If X is an abelian variety then the conjecture of Batyrev<br />

<strong>and</strong> Manin is true.<br />

Indeed, if the rank of X (k) is equal to r then<br />

N X ◦ ,H L<br />

(B) ∼ C ·log r/2 B .<br />

3.10. Remark. –––– On the other h<strong>and</strong>, for K3 surfaces, the conjecture of<br />

Batyrev <strong>and</strong> Manin is open.<br />

For special K3 surfaces, most notably for Kummer surfaces such that the associated<br />

abelian surface is a product of two elliptic curves, a particular result has<br />

been obtained by D. McKinnon [McK]. If d is the minimal degree of a <strong>rational</strong><br />

curve on X then McKinnon shows N X ◦ ,H L<br />

(B) ≪ B 2/d for X ◦ the complement<br />

of the union of all <strong>rational</strong> curves on X of degree d.<br />

3.11. Observation. –––– There are examples of K3 surfaces over a number<br />

field k which contain infinitely many <strong>rational</strong> curves defined over k [Bill, Sec. 3].<br />

Let X be such a K3 surface <strong>and</strong> let C 1 , C 2 , . . . be <strong>rational</strong> curves on X.<br />

Denote their degrees by d 1 , d 2 , . . . . Assume that the curves are listed in such a<br />

way that the degrees are in ascending order.<br />

Then, on<br />

X ◦<br />

l := X \C 1 \ . . . \C l−1 ,<br />

there are still ≥ c l B 2/d l<br />

k-<strong>rational</strong> <strong>points</strong> to be expected. Indeed, X ◦<br />

l contains a<br />

non-empty, open subset of C l .<br />

In other words, there is no way to choose a uniform Zariski open subset X ◦ ⊂ X<br />

such that<br />

N X ◦ ,H L<br />

(B) = #{x ∈ X ◦ (k) | H L (x) < B} ≪ B ε<br />

for every ε > 0. In order to fulfill the conjecture of Batyrev <strong>and</strong> Manin, one<br />

therefore has to choose X ◦ in dependence of ε.<br />

3.12. Example. –––– Consider the diagonal quartic surface X given in P 3Éby<br />

the equation<br />

x 4 + 2y 4 = z 4 + 4w 4 .


Sec. 3] THE CONJECTURE OF BATYREV AND MANIN 219<br />

On this K3 surface, there are theÉ-<strong>rational</strong> <strong>points</strong> (1 : 0 : ±1 : 0) <strong>and</strong><br />

(±1 484 801 : ±1 203 120 : ±1 169 407 : ±1 157 520) .<br />

These are the onlyÉ-<strong>rational</strong> <strong>points</strong> onX known <strong>and</strong> the onlyÉ-<strong>rational</strong> <strong>points</strong><br />

of (naive) height less than 10 8 [EJ2, EJ3].<br />

A systematic search forÉ-<strong>rational</strong> <strong>points</strong> on the K3 surface X is described in<br />

Chapter XI.<br />

3.13. Remark. –––– In general, not much is known about the arithmetic<br />

of K3 surfaces. Nevertheless, in 1981, F. Bogomolov formulated a very optimistic<br />

conjecture.<br />

3.14. Conjecture (F. Bogomolov, cf. [Bo/T]). —– Let X be a K3 surface over a<br />

number field k. Then, every k-<strong>rational</strong> point on X lies on <strong>rational</strong> curve C ⊂ X<br />

(defined over the algebraic closure k).<br />

iii. A somewhat different example. —<br />

3.15. Example. –––– Let X be the blow-up of P 2 in nine <strong>points</strong> P 1 , . . . , P 9 in<br />

general position.<br />

Denote by E := E 1 + . . . + E 9 the sum of the corresponding nine exceptional<br />

lines. According to the Nakai-Moizhezon criterion, a divisor aL − bE is<br />

ample if <strong>and</strong> only if a > 3b > 0. Further, we have K = −3L + E.<br />

The conjecture of Batyrev <strong>and</strong> Manin therefore asserts that, for every ε > 0,<br />

there exists a Zariski open subset X ◦ ⊂ X such that<br />

N X ◦ ,H O(aL−bE)<br />

(B) ≪ B 3/a+ε .<br />

On the other h<strong>and</strong>, assuming aL − bE to be very ample for simplicity, the<br />

embedding ι: (X<br />

(<br />

։)P<br />

) 2 ֒→ P N is given by homogeneous forms of degree a.<br />

Hence, H O(aL−bE) ι(x) ≤ c · H<br />

a<br />

naive(x) for some constant c which shows the<br />

lower bound<br />

N X ◦ ,H O(aL−bE)<br />

(B) = Ω(B 3/a )<br />

for any Zariski open subset X ◦ ⊂ X.<br />

3.16. –––– On X, there are infinitely many exceptional curves D such that<br />

D 2 = −1 <strong>and</strong> DK = −1. For every d ∈Æ, there are finitely many of the<br />

type dL − a 1 E 1 − . . . − a 9 E 9 .


220 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

Relative to aL − bE, their degrees are<br />

(dL − a 1 E 1 − . . . − a 9 E 9 )(aL − bE) = da − b(a 1 + . . . + a 9 )<br />

= da − b(3d − 1)<br />

= d(a − 3b) + b .<br />

Since a − 3b > 0, this expression tends to infinity for d → ∞. Only finitely<br />

many of the exceptional curves are of degree < 2 a which is equivalent to ≫B3/a<br />

3<br />

<strong>rational</strong> <strong>points</strong> of height


Sec. 4] THE CONJECTURE OF MANIN 221<br />

4.2. Lemma. –––– Let k ⊆be an algebraically closed field of characteristic 0<br />

<strong>and</strong> X be a smooth, projective variety over k. Assume that X is Fano.<br />

Then, for every prime number l,<br />

i) H 1 ét (X,l) = 0.<br />

ii) The first Chern class induces an isomorphism<br />

∼=<br />

Pic(X −→ H )⊗l<br />

2 ét (X,l(1)) .<br />

Proof. Let first i ∈Æbe arbitrary. By [SGA4, Exp. XVI, Corollaire<br />

1.6], we have, for every k ∈Æ, H i ét (X, µ l k) ∼ = H i ét (X, µ l k). Further,<br />

the comparison theorem [SGA4, Exp. XI, Théorème 4.4] shows<br />

that H i ét (X, µ l k) ∼ = H i (X (),/l k).<br />

On the other h<strong>and</strong>, by the universal coefficient theorem for cohomology [Sp,<br />

Chap. 5, Sec. 5, Theorem 10],<br />

H i (X (),/l k) ∼ = H i (X (),)⊗/l k⊕Tor 1 (H i+1 k<br />

(X (),),/l k)<br />

∼= H i (X (),)⊗/l k⊕H i+1 (X (),) l k .<br />

Here, the transition maps H i+1 (X (),) l k+1 → H i+1 (X (),) l k are given by<br />

multiplication by l. A finite composition of them is the zero map. This implies<br />

H i ét (X,l(1)) ∼ = lim ←−<br />

H i (X (),)⊗/l<br />

= H i (X . (‡)<br />

(),)⊗l<br />

i) Since H 1 (X (), O X ()) = 0, the long exact cohomology sequence associated<br />

to the exponential sequence yields that H 1 (X (),) = 0. Formula (‡) implies<br />

the claim.<br />

ii) Here, we use both, H 1 (X (), O X ()) = 0 <strong>and</strong> H 2 (X (), O X ()) = 0.<br />

The long exact cohomology sequence associated to the exponential sequence<br />

then shows that<br />

c 1 : Pic(X ) −→ H 2 (X (),)<br />

is an isomorphism. Tensoring yields isomorphisms<br />

Pic(X )⊗/l k.<br />

<strong>and</strong> going over to the inverse limit implies the assertion.<br />

k−→ H 2 (X (),)⊗/l<br />


222 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

4.3. Remarks. –––– i) The first Chern class in étale cohomology is defined using<br />

the Kummer sequence. Recall that there is the commutative diagram<br />

0<br />

2πi O X<br />

exp O ∗ X<br />

0<br />

exp( 2πi<br />

n ·)<br />

0 µ n O ∗ X<br />

exp( 1 n ·)<br />

(·) n O ∗ X<br />

showing that this agrees with the definition based on the exponential sequence.<br />

ii) The tensor product does in general not commute with inverse limits. However,<br />

for a finitely generated-module A,<br />

k∼ = A⊗l<br />

lim<br />

←− A⊗/l<br />

is the l-adic completion of A [Mat, Theorem 8.7].<br />

=<br />

0<br />

ii. Fixing a particular anticanonical height. —<br />

4.4. Remark. –––– The anticanonical height H K ∨ on X is determined only up<br />

to a certain factor which is bounded above <strong>and</strong> below by positive constants.<br />

To be able to make any statement on the value of the constant τ, we have to fix<br />

a particular height function.<br />

4.5. –––– For this, according to Definition IV.7.38, it is necessary to choose an<br />

adelic metric ‖ . ‖ = {‖ . ‖ ν } ν∈Val(É) on K ∨ such that (K ∨ , ‖ . ‖) ∈ ˜Pic(X ).<br />

For the remainder of this section, we fix such an adelic metric once <strong>and</strong> for all.<br />

4.6. Definition. –––– Put<br />

H K ∨(x) := exph (K ∨ ,‖.‖)(x)<br />

for h (K ∨ ,‖ .‖) theabsoluteheightwithrespecttotheintegrablymetrizedinvertible<br />

sheaf (K ∨ , ‖ . ‖) in the sense of Definition IV.7.38.<br />

We will call this height function the anticanonical height defined by the adelic<br />

metric given in 4.5.<br />

4.7. –––– Most height functions occurring in practice are a lot simpler than<br />

the general theory. For this reason, it is probably wise to recall an elementary<br />

particular case.<br />

Choose<br />

i) a projective embedding ι: X → P NÉsuch that K ∨ ∼ = ι ∗ O(d) for some d ∈Æ,<br />

<strong>and</strong><br />

ii) a continuous hermitian metric ‖ . ‖ ∞ on K.<br />


Sec. 4] THE CONJECTURE OF MANIN 223<br />

Then, the topological closure of ι(X ) in P Nis an arithmetic variety X which<br />

is a model of X over Spec.<br />

Put<br />

H K ∨(x) := exp h (O(d)|X ,‖ .‖ ∞ )(x)<br />

for h (O(d)|X ,‖.‖ ∞ ) the height function with respect to the hermitian line bundle<br />

(O(d)| X , ‖ . ‖ ∞ ) ∈ ̂Pic C 0 (X ) in the sense of Definition IV.3.12.<br />

4.8. –––– This height function corresponds to the adelic metric on K ∨ which<br />

is given by the following construction. (Cf. Example IV.7.6.)<br />

i) ‖ . ‖ ∞ is part of the data given.<br />

ii) For a prime number p, the metric ‖ . ‖ p is given as follows.<br />

Let x ∈Ép. There is a unique extension x : SpecOÉp → X of x. Then, x∗ O(d)<br />

is a projective O K -module of rank 1. Each l ∈ O(d)(x) induces a <strong>rational</strong> section<br />

of x ∗ O(d). Put<br />

‖l‖(x) := inf {| a| | a ∈ K, l ∈ a · x ∗ O(d)}.<br />

4.9. Remarks. –––– i) To define the metric ‖ . ‖ p for a particular prime p, a<br />

model L of X over Spec[ 1 ] for p ∤m is sufficient.<br />

m<br />

ii) For every adelic metric ‖ . ‖ = {‖ . ‖ ν } ν∈Val(É) on K ∨ , there exist a model X<br />

of X over Spec[ 1 ] for a certain m ∈Æ<strong>and</strong> an extension of K ∨ to X such<br />

m<br />

that ‖ . ‖ νp is induced by that model for every p ∤m.<br />

iii. The conjecture. —<br />

4.10. –––– The conjecture of Manin deals with the anticanonical height H K ∨<br />

on a Fano variety.<br />

4.11. Conjecture (Manin). —– Let X be a smooth, projective variety overÉ.<br />

Assume that X is Fano.<br />

Then, there exist a positive integer r, a real number τ, <strong>and</strong> a Zariski open subset<br />

X ◦ ⊆ X such that<br />

for B → ∞.<br />

N X ◦ ,H K ∨(B) = #{x ∈ X ◦ (k) | H K ∨(x) < B} ∼ τB log r B<br />

4.12. Remark. –––– The factor log r B is new in comparison with the statistical<br />

heuristic given in 1.4. It is known for a long time that such a factor is, in general,<br />

necessary. In fact, J. Franke, Yu. I. Manin, <strong>and</strong> Y. Tschinkel [F/M/T] showed<br />

in 1989 that Manin’s conjecture becomes compatible with direct products of<br />

Fano varieties only when a suitable log r B-factor is added.


224 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

4.13. Remark. –––– At least in the case that X is a surface, it is expected<br />

that r = rkPic(X ) − 1. There are, however, counterexamples to this formula<br />

in dimension three [Ba/T].<br />

On the other h<strong>and</strong>, in Chapter VII, we will present numerical evidence<br />

for Manin’s conjecture for diagonal cubic <strong>and</strong> quartic threefolds. For those,<br />

rkPic(X ) = 1 <strong>and</strong> our experiments indicate that r = 0.<br />

4.14. –––– In [Pe95a], E. Peyre refined Manin’s conjecture by providing an<br />

explicit description for the value of the coefficient τ. Peyre’s constant is a<br />

product of several factors which we will subsequently explain.<br />

5. Peyre’s constant<br />

i. The factor α. —<br />

5.1. Definition (Cf. [Pe95a, Définition 2.4]). —–<br />

Let X be a projective algebraic variety overÉ. Choose an isomorphism<br />

ι: Pic(X )/ Pic(X ) tors<br />

∼ =<br />

−→t .<br />

Identify Pic(X )⊗ÊwithÊt according to ι.<br />

Further, let Λ eff (X ) ⊂ Pic(X )⊗Ê=Êt be the cone generated by the effective<br />

divisors. Consider the dual cone Λ ∨ eff (X ) ⊂ (Êt ) ∨ . Then,<br />

α(X ) := t · vol {x ∈ Λ ∨ eff<br />

| 〈x, −K 〉 ≤ 1 } .<br />

Here, 〈 . , . 〉 denotes the tautological pairing 〈 . , . 〉: (Êt ) ∨ ×Êt →Ê. vol is the<br />

ordinary Lebesgue measure on (Êt ) ∨ .<br />

5.2. Remark. –––– In [Pe/T, Definition 2.5], α(X ) is defined by an integral.<br />

An elementary calculation shows that the two definitions are equivalent.<br />

5.3. Example. –––– Suppose Pic(X ) =. Denote by [L] ∈ Pic(X ) the ample<br />

generator. Let δ ∈Æbe such that [−K ] = δ[L]. Then, α(X ) = 1/δ.<br />

In particular, one has α(X ) = 1 for every smooth cubic surface such that<br />

rkPic(X ) = 1. Indeed, (−K ) 2 = 3 is square-free. Therefore, [−K ] ∈ Pic(X ) is<br />

not divisible.


Sec. 5] PEYRE’S CONSTANT 225<br />

5.4. Example. –––– α(P 1 × P 1 ) = 1/4.<br />

Indeed, Pic(P 1 ×P 1 ) =L 1 ⊕L 2 . The effective cone is generated by L 1 <strong>and</strong> L 2 .<br />

In the dual space,<br />

Λ ∨ eff (P1 × P 1 ) = {al 1 + bl 2 | a, b ≥ 0} .<br />

Further, −K = 2L 1 + 2L 2 . The condition 〈x, −K 〉 ≤ 1 is therefore equivalent<br />

to 2a + 2b ≤ 1.<br />

The area of the triangle with vertices (0, 0), (1/2, 0), <strong>and</strong> (0, 1/2) is equal to 1/8.<br />

5.5. Example. –––– Let X be the blow-up of P 2 in aÉ-<strong>rational</strong> point.<br />

Then, α(X ) = 1/6.<br />

Here, Pic(X ) =L ⊕E. The effective cone is generated by E <strong>and</strong> L − E.<br />

In the dual space,<br />

Λ ∨ eff (X ) = {al + be | b ≥ 0, a − b ≥ 0} .<br />

Further, −K = 3L−E. The condition 〈x, −K 〉 ≤ 1 is equivalent to 3a−b ≤ 1.<br />

The area of the triangle with vertices (0, 0), (1/3, 0), <strong>and</strong> (1/2, 1/2) is equal<br />

to 1/12.<br />

5.6. Example. –––– Let X be the blow-up of P 2 in sixÉ-<strong>rational</strong> <strong>points</strong> which<br />

form an orbit under Gal(É/É). Then, α(X ) = 4/3.<br />

Again, Pic(X ) =L⊕E. Here, the effective cone is generated by E, L−1/3E,<br />

<strong>and</strong> 2L − 5/6E. I.e., by E <strong>and</strong> L − 5/12E.<br />

In the dual space,<br />

Λ ∨ eff (X ) = {al + be | b ≥ 0, a − 5/12b ≥ 0} .<br />

Further, −K = 3L − E. The condition 〈x, −K 〉 ≤ 1 is therefore equivalent<br />

to 3a − b ≤ 1.<br />

The area of the triangle with vertices (0, 0), (1/3, 0), <strong>and</strong> (5/3, 4) is equal to 2/3.<br />

5.7. Example. –––– Let X be a smooth cubic surface overÉ. Assume the<br />

orbit lengths of the 27 lines under the Gal(É/É)-operation are [1, 10, 16].<br />

Then, α(X ) = 1.<br />

Note that this is the generic case of a cubic surface containing aÉ-<strong>rational</strong> line.<br />

The computation using GAP discussed in II.8.23 shows that rk Pic(X ) = 2.<br />

Compare the list given in the appendix. We claim that Pic(X ) =K ⊕E<br />

for E theÉ-<strong>rational</strong> line.


∈<br />

226 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

Indeed, K <strong>and</strong> E are linearly independent since K 2 = 3 <strong>and</strong> E 2 = −1. Further,<br />

if aK + bE ∈ Pic(X ) then intersecting with K shows that 3a − b<br />

while intersecting with a line skew to E shows −a ∈. Altogether, a, b ∈.<br />

Write D 1 for the sum of the ten lines meeting E <strong>and</strong> D 2 for the sum of the sixteen<br />

lines skew to E. As D 1 K = −10 <strong>and</strong> D 1 E = 10, we have D 1 = −5K − 5E.<br />

Similarly, D 2 K = −16 <strong>and</strong> D 2 E = 0 imply D 2 = −4K + 4E.<br />

The effective cone is generated by E, D 1 , <strong>and</strong> D 2 . The calculations show that E<br />

<strong>and</strong> −K − E form a simpler system of generators.<br />

In the dual space,<br />

Λ ∨ eff (X ) = {ak + be | b ≥ 0, −a − b ≥ 0} .<br />

Further, the condition 〈x, −K 〉 ≤ 1 is equivalent to −a ≤ 1.<br />

The area of the triangle with vertices (0, 0), (−1, 0), <strong>and</strong> (−1, 1) is equal to 1/2.<br />

5.8. Example. –––– Let X be a smooth cubic surface overÉ. Assume the<br />

orbit lengths of the 27 lines under the Gal(É/É)-operation are [2, 5, 10, 10].<br />

Then, α(X ) = 3/2.<br />

Note that the two lines conjugate to each other are skew. Indeed, if they<br />

were contained in a plane then the third line contained in that plane would be<br />

É-<strong>rational</strong>. This example describes the generic case of a cubic surface containing<br />

two skew lines which are conjugate to each other over a quadratic number field.<br />

Again, the computation usingGAP discussed in II.8.23 shows that rkPic(X ) = 2.<br />

We claim that Pic(X ) =K ⊕E for E := E 1 + E 2 the sum of the two lines<br />

conjugate to each other.<br />

Indeed, the discriminant of the lattice spanned by K <strong>and</strong> E is<br />

−2 −2<br />

∣ −2 3∣ = −10 .<br />

As this number is non-zero, we see that K <strong>and</strong> E are linearly independent.<br />

Since (−10) is a square-free integer, the lattice may not be refined.<br />

Write D 1 for the sum of the five lines meeting E 1 <strong>and</strong> E 2 , D 2 for the sum of<br />

the ten other lines meeting E 1 or E 2 , <strong>and</strong> D 3 for the sum of the ten lines not<br />

meeting E at all. Then, D 1 K = −5 <strong>and</strong> D 1 E = 10 imply that D 1 = −3K −2E.<br />

Similarly, D 2 K = −10 <strong>and</strong> D 2 E = 10 show that D 2 = −4K − E. Finally,<br />

D 3 K = −10, D 3 E = 0, <strong>and</strong> D 3 = −2K + 2E.<br />

The effective cone is generated by E, D 1 , D 2 , <strong>and</strong> D 3 . The calculations yield E<br />

<strong>and</strong> −3K − 2E as a simpler system of generators.<br />

In the dual space,<br />

Λ ∨ eff (X ) = {ak + be | b ≥ 0, −3a − 2b ≥ 0} .


Sec. 5] PEYRE’S CONSTANT 227<br />

Further, the condition 〈x, −K 〉 ≤ 1 is equivalent to −a ≤ 1.<br />

The area of the triangle with vertices (0, 0), (−1, 0), <strong>and</strong> (−1, 3/2) is 3/4.<br />

5.9. Remarks. –––– i) Let X be a smooth cubic surface overÉcontaining a<br />

É-<strong>rational</strong> line. Assume that rk Pic(X ) = 2. Then, α(X ) = 1.<br />

Indeed, the arguments given in Example 5.7 still show that Pic(X ) =K ⊕E.<br />

On the other h<strong>and</strong>, the Gal(É/É)-orbits of the 27 lines might break into<br />

smaller pieces. For example, assume there is an orbit consisting of k of the<br />

ten lines meeting E. This causes another generator D 1 ′ of the effective cone.<br />

However, we have that D 1 ′ K = −k <strong>and</strong> D 1 ′ E = k. Therefore,<br />

is simply a scalar multiple of D 1 .<br />

D ′ 1 = − k 2 K − k 2 E = k 10 D 1<br />

Analogously, an orbit consisting of some of the sixteen lines skew to E leads to<br />

a scalar multiple of D 2 . This ensures that Λ ∨ eff<br />

(X ) is the same as in the generic<br />

case described in Example 5.7.<br />

ii) Similarly, let X be any smooth cubic surface overÉcontaining two skew<br />

lines conjugate over a quadratic number field. Assume that rk Pic(X ) = 2.<br />

Then, α(X ) = 3/2.<br />

5.10. Remark. –––– For smooth cubic surfaces in general, the value of α depends<br />

only on the orbit structure of the 27 lines under the Gal(É/É)-operation.<br />

In particular, there are only 350 cases corresponding to the conjugacy classes of<br />

sub<strong>groups</strong> of W (E 6 ). The examples given above actually cover the lion’s share<br />

of these cases.<br />

In fact, 137 of the 350 conjugacy classes of sub<strong>groups</strong> of W (E 6 ) lead to Picard<br />

rank one. Then, α(X ) = 1 as shown in Example 5.3.<br />

Further, 133 conjugacy classes yield Picard rank two. The argument of Remark<br />

5.9.i) alone covers 98 of them. Eight further conjugacy classes are treated<br />

by Remark 5.9.ii).<br />

This might indicate that α(X ) may be effectively computed for each of the<br />

350 conjugacy classes of sub<strong>groups</strong> of W (E 6 ) which could appear as the Galois<br />

<strong>groups</strong> acting on the 27 lines. This has, however, not yet been done.<br />

In the cases of high Picard rank, the computation of the volume of the resulting<br />

polytope should be the main challenge. For example, László Lovász [Lo]<br />

explains that good algorithms for the exact computation of such volumes are impossible.<br />

His suggestion is to use a Monte-Carlo method instead.<br />

The case of maximal rank has, however, been settled.


228 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

5.11. Example. –––– Let X be a smooth cubic surface over the <strong>rational</strong>s such<br />

that rk Pic(X ) = 7. I.e., such that each of the 27 lines on X is defined overÉ.<br />

Then, α(X ) = 1<br />

120 .<br />

This has been proven in the Ph.D. thesis of U. Derenthal [Der, Theorem 8.3].<br />

ii. The factor β. —<br />

5.12. Definition. –––– Let X be a projective variety overÉ. Then, β(X ) is<br />

defined as<br />

β(X ) := #H 1( Gal(É/É), Pic(XÉ) ) .<br />

5.13. Remarks. –––– We studied H 1( Gal(É/É), Pic(XÉ) ) in Chapter II. Recall<br />

the following facts.<br />

i) If Pic(XÉ) =then β(X ) = 1.<br />

ii) In the particular case that X is a smooth cubic surface, β(X ) depends only<br />

on the conjugacy class of the subgroup G ⊆ W (E 6 ) defined by the operation<br />

of Gal(É/É) on the 27 lines. Actually, β(X ) depends only on the decomposition<br />

of the 27 lines into Galois orbits.<br />

We computed β(X ) for each of the 350 conjugacy classes of sub<strong>groups</strong> of W (E 6 ).<br />

The possible values are β(X ) = 1, 2, 3, 4, <strong>and</strong> 9. More details are given in II.8.23<br />

<strong>and</strong> the list presented in the appendix.<br />

iii. The value of L at 1. —<br />

5.14. Lemma. –––– Let X be a projective algebraic variety overÉ. Denote by<br />

L( . , χ Pic(XÉ)) the Artin L-function of the Gal(É/É)-representation Pic(XÉ)⊗.<br />

Then, for t the Picard rank of X,<br />

is a real number different from zero.<br />

lim<br />

s→1<br />

(s − 1) t L(s, χ Pic(XÉ))<br />

Proof. The Gal(É/É)-representation Pic(XÉ)⊗contains the trivial representation<br />

t times as a direct summ<strong>and</strong>. Therefore, L(s, χ Pic(XÉ)) = ζ(s) t ·L(s, χ P )<br />

<strong>and</strong><br />

lim<br />

s→1<br />

(s − 1) t L(s, χ Pic(XÉ)) = L(1, χ P ) .<br />

Here, ζ denotes the Riemann zeta function <strong>and</strong> P is a Gal(É/É)-representation<br />

which does not contain trivial components. [Mu-y, Corollary 11.5 <strong>and</strong> Corollary<br />

11.4] show that L( . , χ P ) has neither a pole nor a zero at 1. □


Sec. 5] PEYRE’S CONSTANT 229<br />

iv. The p-adic <strong>measures</strong>. —<br />

5.15. Definition. –––– Let p be a prime number <strong>and</strong> x ∈ X (Ép) be arbitrary.<br />

Choose local coordinates t 1 , . . . , t n on X × SpecÉSpecÉp in a neighbourhood<br />

of x. These define a morphism ofÉp-schemes<br />

ι x : U x −→ A nÉp<br />

from a Zariski open neighbourhood U x of x such that ι(x) = (0, . . . , 0) <strong>and</strong> ι is<br />

étale at x.<br />

The tensor field (ι −1 ) [ ∂<br />

]<br />

x,Ép ∗ ∂t 1<br />

∧ . . . ∧ ∂<br />

∂t n<br />

is the restriction to the p-adic <strong>points</strong><br />

of a section of the anticanonical sheaf K ∨Ép .<br />

In a neighbourhood of x, we define the measure ω p by<br />

[<br />

∥ ∂<br />

∥(ι −1 ) x,Ép ∗ ∧ . . . ∧ ∂ ]∥ ∥∥ · ι<br />

∗<br />

∂t 1 ∂t (dt x,Ép 1 · . . . · dt n ) .<br />

n<br />

Here, each dt i denotes a copy of the Haar measure onÉp normalized in the<br />

usual manner.<br />

5.16. Remark. –––– This definition is independent of the choice of the coordinates<br />

t 1 , . . . , t n . Indeed, for t ′ 1, . . . , t ′ n forming another system of local<br />

coordinates near x, we have<br />

dt ′ 1 · . . . · dt ′ n = det ∂(t′ 1 , . . . , t′ n )<br />

∂(t 1 , . . . , t n ) · dt 1 . . . dt n<br />

<strong>and</strong><br />

∂<br />

∂t ′ 1<br />

∧ . . . ∧ ∂<br />

∂t ′ n<br />

= det ∂(t 1, . . . , t n )<br />

∂(t<br />

1 ′, . . . , t′ n ) · ∂<br />

∧ . . . ∧ ∂ .<br />

∂t 1 ∂t n<br />

5.17. Lemma. –––– For a primenumber p, suppose that the metric ‖ . ‖ p is induced<br />

by a model X of X. Then, the measure ω p on X (Ép) is given as follows.<br />

Let a ∈ X (/p k) <strong>and</strong> put<br />

Then,<br />

U (k)<br />

a := {x ∈ X (Ép) | x ≡ a (mod p k ) } .<br />

ω p (U a (k) ) = lim #{ y ∈ X (/p n) | y ≡ a (mod p k ) }<br />

.<br />

n→∞ p n dim X<br />

Proof. This is simply a more concrete reformulation of the abstract definition.<br />

□<br />

5.18. Corollary. –––– Let p be a prime such that X is smooth over p. Then, for<br />

every a ∈ X (/p),<br />

ω p (U a (1) ) = 1<br />

p . dim X


230 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

Proof. This is a consequence of Hensel’s lemma.<br />

□<br />

5.19. Remark. –––– ω p may as well be described in terms of the affine cone<br />

over X. For example, if X is a hypersurface of degree d in P n then one has<br />

(<br />

ω ) p X (Ép) = 1 − p−k<br />

· lim<br />

1 − p #C (dim X+1)n<br />

−1 X (/p<br />

n)/p<br />

n→∞<br />

for k := n + 1 − d. A proof is given in [Pe/T, Corollary 3.5].<br />

5.20. Definition (The measure τ p on X (Ép)). —– Let p be a prime number.<br />

We define the measure τ p on X (Ép) by<br />

τ p := det ( 1 − p −1 Frob p | Pic(XÉ) I p) · ωp .<br />

Here, Pic(XÉ) I p<br />

denotes the fixed module under the inertia group.<br />

5.21. Remark. –––– In the easiest case that Pic(XÉ) =, one has<br />

det ( )<br />

1 − p −1 Frob p | Pic(XÉ) I 1<br />

p = 1 −<br />

p .<br />

v. The real measure. —<br />

5.22. Definition (The measure τ ∞ nÊ<br />

on X (Ê)). —– Let x ∈ X (Ê) be arbitrary.<br />

Choose local coordinates t 1 , . . . , t n on X × SpecÉSpecÊin a neighbourhood<br />

of x. These define a morphism ofÊ-schemes<br />

ι x : U x −→ A<br />

from a Zariski open neighbourhood U x of x such that ι(x) = (0, . . . , 0) <strong>and</strong> ι is<br />

étale at x. Consequently, ι x,Ê: U x (Ê) →Ên is a diffeomorphism near x.<br />

[<br />

The tensor field (ιx,Ê) −1 ∂<br />

]<br />

∗ ∂t 1<br />

∧ . . . ∧ ∂<br />

∂t n<br />

is the restriction to the real <strong>points</strong> of a<br />

section of the anticanonical sheaf K.<br />

∨<br />

In a neighbourhood of x, we define the measure τ ∞ by the differential form<br />

[<br />

∥ ∂<br />

∥(ι −1<br />

x,Ê) ∗ ∧ . . . ∧ ∂ ]∥ ∥∥ · ι<br />

∗<br />

∂t 1 ∂t x,Ê(dt 1 ∧ . . . ∧ dt n ) .<br />

n<br />

5.23. Remark. –––– Again, this definition is independent of the choice of the<br />

coordinates t 1 , . . . , t n . If t ′ 1, . . . , t ′ n form another system of local coordinates<br />

near x then<br />

dt ′ 1 ∧ . . . ∧ dt ′ n = det ∂(t′ 1 , . . . , t′ n )<br />

∂(t 1 , . . . , t n ) · dt 1 ∧ . . . ∧ dt n<br />

<strong>and</strong><br />

∂<br />

∂t ′ 1<br />

∧ . . . ∧ ∂<br />

∂t ′ n<br />

= det ∂(t 1, . . . , t n )<br />

∂(t<br />

1 ′, . . . , t′ n) · ∂<br />

∧ . . . ∧ ∂ .<br />

∂t 1 ∂t n


Sec. 5] PEYRE’S CONSTANT 231<br />

5.24. –––– In the situation that X is a hypersurface, the measure τ ∞ may be<br />

described by far more concretely in terms of the Leray measure.<br />

5.25. Definition. –––– Let f ∈Ê[X 0 , . . . , X n ] be a homogeneous polynomial<br />

such that (grad f )(x 0 , . . . , x n ) ≠ 0 for every (x 0 , . . . , x n ) different from<br />

the origin.<br />

Then, the Leray measure on “ f = 0” is given by the formula<br />

1<br />

ω Leray =<br />

‖ grad f ‖ ω hyp .<br />

Here, ω hyp denotes the usual hypersurface measure.<br />

5.26. Lemma. –––– Let f ∈Ê[X 0 , . . . , X n ] be as in Definition 5.25 <strong>and</strong><br />

U ⊂Ên+1 be an open subset such that ∂ f<br />

∂x 0<br />

does not vanish on U.<br />

Then, on U ∩“ f = 0”, the Leray measure is given up to sign by the differential form<br />

1<br />

| ∂f/∂x 0 | dx 1 ∧ . . . ∧ dx n .<br />

Proof. This is an immediate consequence of Sublemma 5.27 below.<br />

5.27. Sublemma. –––– Let U ⊂Ên+1 be an open subset <strong>and</strong> f : U →Êbe a<br />

smooth function such that ∂ f<br />

∂x 0<br />

does not vanish.<br />

Then,thehypersurfacemeasureon“f = 0”isgivenuptosignbythedifferentialform<br />

‖ grad f ‖<br />

| ∂f/∂x 0 | dx 1 ∧ . . . ∧ dx n .<br />

Proof. It is well known that, having resolved the equation<br />

f (x 0 , . . . , x n ) = 0<br />

by x 0 , the hypersurface measure is given by the form<br />

√<br />

1 + (∂x 0 /∂x 1 ) 2 + . . . + (∂x 0 /∂x n ) 2 dx 1 ∧ . . . ∧ dx n .<br />

For every i = 1, . . . , n, the equation f ( )<br />

x 0 (x 1 , . . . , x n ), x 1 , . . . , x n = 0<br />

immediately yields ∂f ∂x 0<br />

∂x 0 ∂x i<br />

+ ∂f<br />

∂x i<br />

= 0 . In other words,<br />

∂x 0<br />

= − ∂f / ∂f<br />

.<br />

∂x i ∂x i ∂x 0<br />

Altogether, we find the differential form<br />

√<br />

1 +<br />

n<br />

∑<br />

i=1<br />

( ∂f<br />

∂x i<br />

/ ∂f<br />

∂x 0<br />

) 2<br />

dx 1 ∧ . . . ∧ dx n<br />

for the hypersurface measure. This is exactly the assertion.<br />

□<br />


232 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

5.28. Notation. –––– Let X ⊆ P nÉbe a projective variety. We denote by CX<br />

the affine cone over X.<br />

A metric ‖ . ‖ on O(1)| X () defines a compact subset<br />

N ‖.‖ := {x = (x 0 , . . . , x n ) ∈ CX (Ê) | | x 0 | ≤ ‖X 0 ‖(x), . . . , | x n | ≤ ‖X n ‖(x)}<br />

of the affine cone which is symmetric to the origin. Note that the conditions<br />

| x i | ≤ ‖X i ‖(x) <strong>and</strong> | x j | ≤ ‖X j ‖(x) are equivalent to each other as long<br />

as x i , x j ≠ 0.<br />

5.29. Examples. –––– i) Consider on O(1)| X () the minimum metric ‖ . ‖ min<br />

from Example IV.3.5. Then,<br />

N ‖ .‖min<br />

= {x = (x 0 , . . . , x n ) ∈ CX (Ê) | | x 0 | ≤ 1, . . . , | x n | ≤ 1}<br />

is a hypercube.<br />

ii) Similarly, for the l 2 -metric,<br />

is the unit ball.<br />

N ‖.‖l 2 = {x = (x 0, . . . , x n ) ∈ CX (Ê) | x 2 0 + . . . + x 2 n ≤ 1}<br />

5.30. Proposition. –––– Let F ∈É[X 0 , . . . , X n ] be a homogeneous polynomial<br />

of degree d <strong>and</strong> X ⊂ P nÉbe the hypersurface defined by the equation F = 0.<br />

According to the adjunction formula, there is a canonical isomorphism<br />

ι: O(n − d + 1)<br />

∼=<br />

−→ K ∨<br />

X .<br />

The metric ‖ . ‖ on K ∨<br />

X induces a metric ι−1 ‖ . ‖ 1<br />

n+d−1 on O(1)|X (). Let<br />

N := N ι −1 ‖ .‖<br />

be the subset defined by ι −1 ‖ . ‖ 1<br />

n+d−1 .<br />

1<br />

n−d+1<br />

⊂ CX (Ê)<br />

Then, in terms of N <strong>and</strong> the Leray measure, one may write<br />

τ ∞ (U ) = n − d + 1 Z<br />

ω Leray<br />

2<br />

for every measurable set U ⊆ X (Ê).<br />

Proof. First step: Explicit description of ι.<br />

CU ∩N<br />

The assertion is local in X. We may therefore assume without restriction that<br />

X 0 ≠ 0 <strong>and</strong> ∂F<br />

∂X 1<br />

≠ 0. Then, t 2 , . . . , t n for t i := X i /X 0 form a local system of<br />

coordinates on X.


Sec. 5] PEYRE’S CONSTANT 233<br />

When putting F = X d 0 f (t 1, . . . , t n ), we find<br />

∂F<br />

= X d−1 ∂f<br />

0 .<br />

∂X 1 ∂t 1<br />

Under the Poincaré residue map [G/H, p. 147], the differential form<br />

∂f<br />

∂t 1<br />

df<br />

f<br />

∧ dt 2 ∧ . . . ∧ dt n = ∂f<br />

∂t 1<br />

dt 1 ∧ . . . ∧ dt n<br />

on P nis mapped to dt 2 ∧ . . . ∧ dt n . Dually, under this correspondence, the<br />

tensor field ∂<br />

∂t 2<br />

∧ . . . ∧ ∂<br />

∂t n<br />

on X () is identified with<br />

f ∂<br />

∧ . . . ∧ ∂ = 1 F ∂<br />

∧ . . . ∧ ∂ .<br />

∂t 1 ∂t n X<br />

∂F 0 ∂t<br />

∂X 1 ∂t n 1<br />

Furthermore, the Euler sequence identifies ∂<br />

∂t 1<br />

∧ . . . ∧ ∂<br />

∂t n<br />

with the global section<br />

X n+1<br />

0 ∈ Γ ( P n , O(n + 1) ) .<br />

Altogether,<br />

ι −1 ( ∂<br />

∂t 2<br />

∧ . . . ∧ ∂<br />

∂t n<br />

)<br />

= 1<br />

∂F<br />

∂X 1<br />

X n 0 .<br />

Second step: The measure τ ∞ of Peyre.<br />

Peyre’s measure τ ∞ on X (Ê) is therefore given by the differential form<br />

1<br />

∥ ∂F ∥ dt 2 ∧ . . . ∧ dt n = | X0 d−1 / ∂F<br />

∂X 1<br />

|·‖X n−d+1<br />

0 ‖dt 2 ∧ . . . ∧ dt n . (§)<br />

∂X 1<br />

X n 0<br />

On the other h<strong>and</strong>, the subset N ⊂ CX (Ê) is given by<br />

| x 0 | ≤ A := ‖X 0 ‖ .<br />

Therefore, according to Fubini’s theorem,<br />

Z<br />

CU ∩N<br />

ω Leray =<br />

=<br />

=<br />

Z<br />

CU<br />

|x 0 |≤A<br />

Z A<br />

−A<br />

1<br />

| ∂F<br />

x n−1<br />

0<br />

x d−1<br />

0<br />

∂X 1<br />

| dx 0 ∧ dx 2 ∧ . . . ∧ dx n<br />

dx 0 ·<br />

Z<br />

CU<br />

x 0 =1<br />

2<br />

n − d + 1 · An−d+1 ·<br />

1<br />

| ∂F<br />

∂X 1<br />

| dx 2 ∧ . . . ∧ dx n<br />

Z<br />

CU<br />

x 0 =1<br />

1<br />

| ∂F<br />

∂X 1<br />

| dx 2 ∧ . . . ∧ dx n .<br />

Note here that CX is an (n −1)-dimensional cone while the integr<strong>and</strong> is homogeneous<br />

of degree −(d − 1).


234 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

Consequently,<br />

n − d + 1<br />

2<br />

Z<br />

CU ∩N<br />

ω Leray =<br />

Z<br />

CU<br />

x 0 =1<br />

‖X 0 ‖ n−d+1<br />

dx<br />

| ∂F 2 ∧ . . . ∧ dx n<br />

∂X 1<br />

|<br />

which, according to formula (§), is exactly equal to τ ∞ (U ).<br />

□<br />

5.31. Remarks. –––– i) This result allows a generalization to complete intersections.<br />

ii) If X is a cubic surface then n − d + 1 = 1.<br />

On the other h<strong>and</strong>, consider the case that X is a hypersurface in P n for n ≥ 4.<br />

Then, Pic(X ) = 1 <strong>and</strong> α(X ) = 1 . We therefore have that<br />

n−d+1<br />

α(X )·τ ∞ (U ) = 1 Z<br />

ω Leray<br />

2<br />

for every measurable set U ⊆ X (Ê).<br />

vi. The Tamagawa measure. —<br />

CU ∩N<br />

5.32. Definition. –––– The Tamagawa measure τ H on the set X (É) of<br />

<strong>points</strong> on X is defined to be the product measure<br />

τ H := ∏ τ ν .<br />

ν∈Val(É)<br />

adelic<br />

5.33. Remark. –––– X is projective. Therefore, one has that<br />

X (É) = ∏ X (Éν) .<br />

ν∈Val(É)<br />

5.34. Lemma. –––– The infinite product<br />

(<br />

∏ τ )<br />

ν X (Éν)<br />

ν∈Val(É)<br />

is absolutely convergent. In particular, the infinite product measure ∏ τ ν is well defined.<br />

ν∈Val(É)<br />

Proof. Absolute convergence may not be destroyed by a finite set of factors.<br />

Thus, assume that almost all of the metrics ‖ . ‖ νp of the adelic metric<br />

‖ . ‖ = {‖ . ‖ ν } ν∈Val(É) are induced by a model X of X. Further, we may<br />

restrict the infinite product to all prime numbers p such that X is smooth<br />

over p.<br />

Corollary 5.18 assures that<br />

(<br />

τ ) p X (Ép) = det ( )<br />

1 − p −1 Frob p | Pic(XÉ) I #X (p)<br />

p · . ()<br />

pdim X


Sec. 5] PEYRE’S CONSTANT 235<br />

Further, smoothness over p implies that<br />

Pic(XÉ) I p<br />

= Pic(XÉ) .<br />

We denote the eigenvalues of Frob p on Pic(XÉ) by λ 1 , . . . , λ r . As every<br />

divisor on is actually defined over a finite field extension, all these are roots<br />

Xp<br />

of unity.<br />

Since Pic(XÉ) is a-module of finite rank, the characteristic polynomial<br />

of Frob p is monic with integral coefficients. This shows, if λ is an eigenvalue<br />

then λ = 1 λ is an eigenvalue of Frob p, too.<br />

Clearly, one has<br />

det ( 1 − p −1 Frob p | Pic(XÉ) I p) = (1 − λ1 p −1 ) · . . . · (1 − λ r p −1 )<br />

= 1 − (λ 1 + . . . + λ r )p −1 + E<br />

where<br />

| E| <<br />

( ( k k<br />

p<br />

2)<br />

−2 + . . . + p<br />

k)<br />

−k<br />

for p > 2k.<br />

< k2<br />

p 2 + k3<br />

p 3 + . . .<br />

< 2k2<br />

p 2<br />

In order to determine #X (p), we use the Lefschetz trace formula in étale cohomology<br />

[SGA4 1 , Rapport, Théorème 3.2]. This yields<br />

2<br />

for every prime l ≠ p.<br />

dimX p<br />

#X (p) =<br />

∑<br />

i=0<br />

(−1) i tr ( )<br />

Frob p | H i ét ,Él) (‖)<br />

(Xp<br />

Here, according to the Weil conjectures proven by P. Deligne [Del, Théorème<br />

(1.6)], every eigenvalue of the Frobenius on H i ét (Xp ,Él) is of absolute<br />

value p i/2 .<br />

The theorem on smooth base change implies a comparison theorem even for<br />

the unequal characteristic case [SGA4, Exp. XVI, Corollaire 2.5 <strong>and</strong> Exp. XV,<br />

Théorème 2.1]. We therefore know that<br />

H i ét (Xp ,Él) ∼ = H i ét (XÉ,Él) .<br />

According to [SGA4, Exp. XVI, Corollaire 1.6 <strong>and</strong> Exp. XI, Théorème 4.4], the<br />

latter is isomorphic to the usual cohomology H i (X (),Él).


236 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

Let C be a constant such that dim H i (X (),Él) ≤ C for every i ∈Æ. Then,<br />

dim H i ét (Xp ,Él) ≤ C<br />

for every i ∈Æ, too. Further, Lemma 4.2 describes the first <strong>and</strong> second<br />

cohomologies of a Fano variety more concretely. We have H 1 ét (Xp ,Él) = 0<br />

while the first Chern class induces an isomorphism<br />

Pic(XÉ)⊗Él<br />

∼= H 2 ét (XÉ,Él(1))<br />

∼= H 2 ét (Xp ,Él(1)) .<br />

This isomorphism is compatible with the operation of Frob p . Consequently,<br />

the Frobenius eigenvalues on H 2 ét (Xp ,Él) are λ 1 p, . . . , λ r p.<br />

2 dim X −1<br />

Poincaré duality shows that Hét (Xp ,Él) = 0. Further, by [Del, (2.4)],<br />

Poincaré duality is compatible with the Frobenius eigenvalues. Therefore, the<br />

eigenvalues of the Frobenius on H 2 dimX−2 ,Él) are<br />

ét<br />

(Xp<br />

λ 1 p dimX−1 , . . . , λ r p dimX−1 .<br />

For the number ofp-<strong>rational</strong> <strong>points</strong>, this yields, according to (‖),<br />

where<br />

X (p) = p dim X + (λ 1 + . . . + λ r )p dim X −1 + Dp dim X<br />

| D| < C [p −3/2 + p −2 + p −5/2 + . . . ] < Cp −3/2 1<br />

1 − p −1/2 < 4Cp−3/2 .<br />

Altogether,<br />

τ p<br />

(<br />

X (Ép) )<br />

= (1 − (λ 1 + . . . + λ r )p −1 + E) · (1 + (λ 1 + . . . + λ r )p −1 + D)<br />

= 1 + E + D + (λ 1 + . . . + λ r ) 2 p −2 + (E − D)(λ 1 + . . . + λ r )p −1 + ED .<br />

For p > 4C 2 , we certainly have<br />

| E| < 2k2<br />

p < 2C 2<br />

2 p 2<br />

Therefore, τ p<br />

(<br />

X (Ép) ) = 1 + X where<br />

= 4C · C<br />

2<br />

4C<br />

<<br />

p3/2·p1/2 p . 3/2<br />

| X | < 8C<br />

p + C 2<br />

3/2 p + 8C 2 2<br />

16C<br />

+ 2 p5/2 p 3<br />

< 8C + C 2<br />

2C + 8C 2<br />

4C 2 + 16C 2<br />

8C 3<br />

p 3/2


Sec. 5] PEYRE’S CONSTANT 237<br />

= 2 + 17<br />

2 C + 2 C<br />

p 3/2 .<br />

Since 3/2 > 1, the infinite product is absolutely convergent.<br />

□<br />

vii. The constant of E. Peyre. —<br />

5.35. Definition (E. Peyre, [Pe/T, Definition 2.4]). —–<br />

Peyre’s constant or Peyre’s Tamagawa type number is defined as<br />

τ (X ) := α(X )·β(X ) · lim<br />

s→1<br />

(s − 1) t L(s, χ Pic(XÉ)) · τ H<br />

(<br />

X (É)<br />

for t = rk Pic(X ).<br />

Br)<br />

5.36. Remark. –––– Here, X (É) denotes the part which is not<br />

affected by the Brauer-Manin obstruction. The Brauer-Manin obstruction was<br />

discussed in detail in Chapter III. The precise definition of X (É) Br is given<br />

in III.2.5. According to Proposition III.2.3.b.ii), X (É) Br is a closed subset<br />

of X (É) <strong>and</strong> therefore measurable.<br />

Br ⊆ X (É)<br />

5.37. –––– Thus, there is the following conjecture which refines Conjecture<br />

4.11.<br />

Conjecture (Manin-Peyre). Let X be a smooth, projective variety overÉ.<br />

AssumethatX isFano. DenotebyH K ∨ theanticanonicalheightfunctionintroduced<br />

in Definition 4.6.<br />

a) If dim X ≤ 2 or X ∈ P n is a complete intersection then there exist a real<br />

number τ <strong>and</strong> a Zariski open subset X ◦ ⊆ X such that<br />

for B → ∞.<br />

N X ◦ ,H K ∨(B) = #{x ∈ X ◦ (k) | H K ∨(x) < B} ∼ τB log rk Pic(X )−1 B<br />

b) Suppose that, for X,thereexist a realnumber τ <strong>and</strong> aZariski opensubset X ◦ ⊆ X<br />

such that<br />

for B → ∞.<br />

N X ◦ ,H K ∨(B) = #{x ∈ X ◦ (k) | H K ∨(x) < B} ∼ τB log rk Pic(X )−1 B<br />

Then, τ = τ (X ) is Peyre’s constant.


238 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

5.38. Remarks (Some motivation). —– i) Morally, the conjecture of Manin in<br />

the refined form due to Peyre states the following. The (conjectural) constant<br />

describing the growth of the number ofÉ-<strong>rational</strong> <strong>points</strong> on X is equal to a<br />

regularized product over the densities of the p-adic <strong>points</strong> on X together with<br />

the density of the real <strong>points</strong>.<br />

That there is such a connection is actually not very surprising. For instance,<br />

a low density of p-adic <strong>points</strong> on X for a particular prime p implies strong<br />

congruence conditions forÉ-<strong>rational</strong> <strong>points</strong>.<br />

ii) More precise results have been obtained by the classical circle method [Bir].<br />

For complete intersections in a very high-dimensional projective space, the<br />

circle method provides an asymptotic formula for the number ofÉ-<strong>rational</strong><br />

<strong>points</strong> on X <strong>and</strong> an error term. The assumptions on the dimension of the<br />

ambient projective space are rather restrictive. They are needed in order to<br />

ensure that the provable error term is smaller than the main term.<br />

The coefficient of the main term is a product of p-adic densities together with a<br />

factor corresponding to the archimedean valuation. Unlike the p-adic densities,<br />

the latter factor does not coincide in general with E. Peyre’s factor τ ∞<br />

(<br />

X (Ê) ) .<br />

Rather, it is an integral over the Leray measure (at least in the case that X is<br />

a hypersurface).<br />

By Proposition 5.30, the factor corresponding to the archimedean valuation<br />

may be rewritten in the form α(X ) · τ ∞<br />

(<br />

X (Ê) ) . Therefore, one sees that the<br />

coefficient of the main term provided by the circle method is equal to<br />

Here,<br />

(<br />

∏τ ) (<br />

p X (Ép) · α(X ) · τ ) (<br />

∞ X (Ê) = α(X ) · τ H X (É) p<br />

τ p<br />

(<br />

X (Ép) ) =<br />

) .<br />

(<br />

1 − 1 )<br />

#{ y ∈ X (/p n) | y ≡ a (mod p k ) }<br />

· lim<br />

.<br />

p n→∞ p n dim X<br />

The (1 − 1 ) are convergence-generating factors.<br />

p<br />

In the cases where the circle method is applicable, one always has Pic(X ) ∼ =.<br />

It seems very natural to use (1 − 1 p )t for t = rk Pic(XÉ) as the convergencegenerating<br />

factors for the general case.<br />

In the definition of Peyre’s constant, this is done <strong>and</strong> more than that. Indeed, the<br />

Gal(É/É)-representation Pic(XÉ) I p ⊗contains the trivial representation<br />

t times. Denote the complementary summ<strong>and</strong> by P. Then, the factors used in<br />

Definition 5.20 may be decomposed as<br />

det ( (<br />

)<br />

1 − p −1 Frob p | Pic(XÉ) I p = 1 − 1 t<br />

· det<br />

p) ( 1 − p −1 Frob p | P ) .


Sec. 5] PEYRE’S CONSTANT 239<br />

The factors det ( 1 − p −1 Frob p | P ) form the Euler product for<br />

1/L(s, χ P )<br />

at s = 1. The value of the function L at 1 is introduced into Definition 5.35<br />

only in order to cancel these factors out.<br />

iii) The definition of the factor α(X ) is somehow more complicated <strong>and</strong>, may<br />

be, more mysterious than those of the other parts. Some motivation for the<br />

appearance of such a factor is given by the circle method.<br />

There is, however, another point which is at least as important. The definition<br />

of α(X ) implies that the conjecture of Manin-Peyre is compatible with direct<br />

products of Fano varieties.<br />

iv) For a smooth cubic surface of the “first case” of Colliot-Thélène, Kanevsky,<br />

<strong>and</strong> Sansuc, the assertion of Theorem III.6.4.c) implies that<br />

(<br />

τ ) H X (É) .<br />

On the other h<strong>and</strong>, β(X ) = 3 such that one might want to simplify Peyre’s<br />

formula to<br />

) . (∗∗)<br />

Br) = 1 3 ·τ H(<br />

X (É)<br />

τ (X ) := α(X ) · lim<br />

s→1<br />

(s − 1) t L(s, χ Pic(XÉ)) · τ H<br />

(<br />

X (É)<br />

Clearly, this is also true in all cases when β(X ) = 1.<br />

Formula (∗∗) is, however, wrong in general. For example, when X is a smooth<br />

cubic surface on which there is a Brauer-Manin obstruction to the Hasse principle<br />

then we need τ (X ) = 0 <strong>and</strong> this is not provided by the simplified formula.<br />

Furthermore, in [Pe/T], E. Peyre <strong>and</strong> Y. Tschinkel reported strong numerical<br />

evidencefor Peyre’sformulaincasessimilar toExampleIII.5.36. I.e., ifβ(X ) = 3<br />

but the Brauer-Manin obstruction does not exclude any adelic point on X then<br />

there are three times moreÉ-<strong>rational</strong> <strong>points</strong> on X than naively expected.<br />

5.39. Remark (Known cases). —– As indicated above, Manin’s conjecture in<br />

the refined form due to Peyre is established for smooth complete intersections<br />

of multidegree d 1 , . . . , d n in the case that the dimension of V is very large<br />

compared to d 1 , . . . , d n [Bir]. This is the classical circle method.<br />

Further, Conjecture 5.37 is proven for projective spaces <strong>and</strong> quadrics. Finally,<br />

there are a number of further particular cases in which Manin’s conjecture<br />

is known to be true. See, e.g., the survey given in [Pe02, Sec. 4].<br />

5.40. Remark (Numerical evidence). —– R. Heath-Brown [H-B92a] as well as<br />

E. Peyre <strong>and</strong> Y. Tschinkel [Pe/T] reported numerical evidence for Conjecture<br />

5.37 for isolated examples of smooth cubic surfaces.


240 CONJECTURES ON POINTS OF BOUNDED HEIGHT [Chap. VI<br />

In Chapter VII, we present numerical evidence for Manin’s conjecture in the case<br />

of the threefolds V e<br />

a,b given by axe = by e + z e + v e + w e in P 4Éfor e = 3 <strong>and</strong> 4.<br />

In Chapter IX, we test Manin’s conjecture numerically for general cubic surfaces.<br />

5.41. Remark. –––– It is not all trivial that the conditionally convergent Euler<br />

product<br />

1<br />

det ( 1 − p −1 Frob p | P )<br />

∏<br />

p<br />

really converges versus L(1, χ P ). The point is that the value L(1, χ P ) is actually<br />

defined by analytic continuation.<br />

5.42. Remark. –––– The idea behind Peyre’s constant is in fact an old one.<br />

The main term of the asymptotic formula provided by the circle method was<br />

used before, for example, by D. R. Heath-Brown in his experimental investigations<br />

on cubic surfaces where weak approximation fails [H-B92a].<br />

Actually, Heath-Brown’s point of view was a little different. He considered<br />

the factors (1 − 1 p )t as convergence-generating factors growing out of the circle<br />

method. The resulting conditionally convergent infinite product was treated<br />

like a definition. The value of the function L at 1 appeared, too, but only as<br />

part of a numerical method to speed up convergence.<br />

E. Peyre’s approach is most likely inspired by considerations like those<br />

in [H-B92a]. It has, however, the potential to work in much more generality.


PART C<br />

NUMERICAL EXPERIMENTS


CHAPTER VII<br />

POINTS OF BOUNDED HEIGHT ON CUBIC<br />

AND QUARTIC THREEFOLDS ∗<br />

. . . , one by one, or all at once.<br />

W. S. GILBERT AND A. SULLIVAN: The Yeomen of the Guard (1888)<br />

1. Introduction — Manin’s Conjecture<br />

i. Summary. —<br />

1.1. –––– For the families ax 3 = by 3 + z 3 + v 3 + w 3 , a, b = 1, . . . , 100, <strong>and</strong><br />

ax 4 = by 4 +z 4 +v 4 +w 4 , a, b = 1, . . . , 100, of projective algebraic threefolds,<br />

we test numerically the conjecture of Manin (in the refined form due to Peyre,<br />

cf. Conjecture VI.5.37) about the asymptotics of <strong>points</strong> of bounded height on<br />

Fano varieties.<br />

This includes searching for <strong>points</strong>, computing the Tamagawa number, <strong>and</strong> detecting<br />

the accumulating subvarieties. The goal of this chapter is describe these<br />

computations as well some background on the geometry of cubic <strong>and</strong> quartic<br />

threefolds.<br />

ii. Manin’s conjecture. —<br />

1.2. –––– Let X be a projective algebraic variety overÉ. We fix an embedding<br />

ι: X → P nÉ. In this situation, recall there is the well-known naive<br />

height H naive : X (É) →Êgiven by<br />

H naive (P) := max<br />

i=0,...,n | x i| .<br />

Here, (x 0 : . . . : x n ) := ι(P) ∈ P n (É) where the projective coordinates are<br />

integers satisfying gcd(x 0 , . . . , x n ) = 1.<br />

(∗) This chapter is a revised <strong>and</strong> slightly extended version of the article: The Asymptotics of<br />

Points of Bounded Height on Diagonal Cubic <strong>and</strong> Quartic Threefolds, in: Algorithmic number<br />

theory, Lecture Notes in Computer Science 4076, Springer, Berlin 2006, 317–332, joint with<br />

A.-S. Elsenhans.


244 POINTS OF BOUNDED HEIGHT ON THREEFOLDS [Chap. VII<br />

1.3. –––– It is of interest to ask for the asymptotics of the number ofÉ-<strong>rational</strong><br />

<strong>points</strong> on X of bounded naive height. This applies particularly to Fano varieties<br />

as those are expected to have many <strong>rational</strong> <strong>points</strong> (at least after a finite extension<br />

of the ground-field).<br />

Simplest examples of Fano varieties are complete intersections in P nÉof a multidegree<br />

(d 1 , . . . , d r ) such that d 1 + . . . + d r ≤ n. In this case, the conjecture<br />

of Manin discussed in general in Chapter VI reads as follows.<br />

1.4. Conjecture. –––– Let X ⊆ P nÉbe a non-singular complete intersection of<br />

multidegree (d 1 , . . . , d r ). Assumedim X ≥ 3<strong>and</strong>k := n+1−d 1 − . . . −d r > 0.<br />

Then, there exists a Zariski open subset X ◦ ⊆ X such that<br />

#{x ∈ X ◦ (É) | H naive (x) k < B} ∼ τB<br />

(∗)<br />

for a constant τ.<br />

1.5. Example. –––– Let X ⊂ P 4Ébe a smooth hypersurface of degree 4.<br />

Conjecture 1.4 predicts ∼ τB <strong>rational</strong> <strong>points</strong> of height < B. However, the<br />

hypersurface x 4 + y 4 = z 4 + v 4 + w 4 contains the line given by x = z, y = v,<br />

<strong>and</strong> w = 0 on which there is quadratic growth, already. This explains the<br />

necessity of the restriction to a Zariski open subset X ◦ ⊆ X.<br />

1.6. Remark. –––– Conjecture 1.4 is proven for P n , linear subspaces, <strong>and</strong><br />

quadrics. Further, it is established [Bir] in the case that the dimension of X<br />

is very large compared to d 1 , . . . , d r . The general version (Conjecture VI.5.37)<br />

is known to be true in a number of further particular cases. A rather complete<br />

list may be found in the survey article [Pe02, Sec. 4].<br />

In this chapter, we present numerical evidence for Conjecture 1.4 in the case of<br />

the varieties X e a,b given by axe = by e + z e + v e + w e in P 4Éfor e = 3 <strong>and</strong> 4.<br />

1.7. Remark. –––– By the Noether-Lefschetz Theorem, the assumptions made<br />

on X imply that Pic(XÉ) ∼ =[Ha70, Corollary IV.3.2]. This is no longer true<br />

in dimension two. See Remark 1.14.ii) for more details.<br />

iii. The constant. —<br />

1.8. –––– Conjecture 1.4 is compatible with results obtained by the classical<br />

circle method (e.g. [Bir]). Motivated by this, E. Peyre provided a description<br />

of the constant τ expected in (∗). We formulated the definition for the general<br />

case in VI.5.35.<br />

In the situation considered here, Pic(XÉ) ∼ =implies that β(X ) = 1 <strong>and</strong> there<br />

is no Brauer-Manin obstruction on X. Peyre’s constant is therefore equal to the


Sec. 1] INTRODUCTION — MANIN’S CONJECTURE 245<br />

Tamagawa-type number<br />

τ (X ) := α(X ) · ∏ (1 − 1 ) ω p p(X (Ép)) .<br />

p∈È∪{∞}<br />

In this formula, α(X ) := 1/k for k ∈Æsuch that O(−K ) ∼ = O(k).<br />

The measure ω p is given in local p-adic analytic coordinates x 1 , . . . , x d by<br />

∂<br />

∥ ∧ . . . ∧<br />

∂ ∥ ∥∥∥p<br />

dx 1 . . . dx d .<br />

∂x 1 ∂x d<br />

Here, each dx i denotes a Haar measure onÉp which is normalized in the<br />

∂<br />

usual manner.<br />

∂x 1<br />

∧ . . . ∧ ∂<br />

∂x d<br />

is a section of O(−K ).<br />

1.9. –––– For p finite, one has a natural model X ⊆ P of X given by the<br />

np<br />

defining equation. This induces the metric ‖ . ‖ p on O(k). It is almost immediate<br />

from the definition that<br />

(<br />

ω ) #X (/p m)<br />

p X (Ép) = lim .<br />

m→∞ p m dim X<br />

1.10. Remark. –––– There is another description of ω p of interest for finite p.<br />

One has<br />

(<br />

ω ) p X (Ép) = 1 − p−k<br />

lim<br />

1 − p #CX (/p m)/p (d+1)m .<br />

−1 m→∞<br />

Here, CX denotes the affine cone over X . For a proof, see [Pe/T, Corollary<br />

3.5].<br />

1.11. –––– The hermitian metric ‖ . ‖ ∞ corresponding the naive height H naive<br />

is given by ‖ . ‖ ∞ := ‖ . ‖ k min on O(−K ) ∼ = O(k). Here, ‖ . ‖ min is the hermitian<br />

metric on O(1) defined by<br />

‖x i ‖ min := inf<br />

j=0,...,n | x i/x j | .<br />

The factor ω ∞<br />

(<br />

X (É∞) ) may then be described as follows.<br />

1.12. Lemma. –––– If X ⊂ P n is a hypersurface defined by the equation f = 0<br />

then<br />

(<br />

α(X ) · ω ) ∞ X (Ê) = 1 Z<br />

ω Leray .<br />

2<br />

The Leray measure ω Leray on<br />

f (x 0 ,...,x n )=0<br />

|x 0 |,...,|x n |≤1<br />

{(x 0 , . . . , x n ) ∈Ên+1 | f (x 0 , . . . , x n ) = 0}


246 POINTS OF BOUNDED HEIGHT ON THREEFOLDS [Chap. VII<br />

1<br />

is related to the usual hypersurface measure by the formula ω Leray =<br />

‖ grad f ‖ ω hyp.<br />

On the other h<strong>and</strong>, one may also write<br />

1<br />

ω Leray =<br />

| ∂ f<br />

∂x i<br />

(x 0 , . . . , x n )| dx 0 ∧ . . . ∧ ̂dx i ∧ . . . ∧ dx n .<br />

Proof. The equivalence of the two descriptions of the Leray measure is shown<br />

in Lemma VI.5.26. The main assertion is proven in Proposition VI.5.30. □<br />

1.13. Remark. –––– As the description of the constant τ given here grew out<br />

of a definition given in canonical terms, it is no surprise that τ is invariant<br />

under scaling.<br />

This can also be seen directly. When we multiply f by a prime number p then<br />

τ p gets multiplied by a factor of p. On the other h<strong>and</strong>, τ ∞ gets multiplied by a<br />

factor of 1/p <strong>and</strong> all the other factors remain unchanged.<br />

1.14. Remark. –––– There are several ways to generalize Conjecture 1.4.<br />

These are presented in detail in Chapter VI. For completeness, let us recall<br />

the following facts.<br />

i) One may consider more general heights corresponding to the tautological<br />

invertible sheaf O(1). This includes to<br />

a) replace the minimum metric by an arbitrary continuous hermitian metric<br />

on O(1). This would affect the domain of integration for the factor at infinity.<br />

b) multiply H naive (x) with a function that depends on the reduction of x modulo<br />

some N ∈Æ. This augments Conjecture 1.4 by an equidistribution statement.<br />

ii) Instead of complete intersections, one may consider arbitrary projective Fano<br />

varieties X. Then, H k naive needs to be replaced by a height defined by the<br />

anticanonical sheaf O(−K X ).<br />

If Pic(XÉ) ̸∼ =then the description of the constant C gets more complicated in<br />

several ways. First, there is an additional factor β := #H 1( Gal(É/É), Pic(XÉ) ) .<br />

Further, instead of (1 − 1 ), one has to use more complicated convergencegenerating<br />

factors. (Cf. Remark VI.5.38.) Finally, the Tamagawa measure has<br />

p<br />

to be taken not of the full variety X (É) but of the subset which is not affected<br />

by the Brauer-Manin obstruction.<br />

If Pic(X ) ̸∼ =already overÉthen the right h<strong>and</strong> side of (∗) has to be replaced<br />

by CB log t B. For the exponent of the log-term, there is the expectation that<br />

t = rk Pic(X ) − 1. There are, however, examples [Ba/T] in dimension three in<br />

which the exponent is larger.<br />

The definition of α is rather complicated, in general. The factor α depends on<br />

the structure of the effective cone in Pic(X ) <strong>and</strong> on the position of (−K X ) in it.<br />

(Cf. Definition VI.5.1).


Sec. 2] COMPUTING THE TAMAGAWA NUMBER 247<br />

2. Computing the Tamagawa number<br />

i. Counting <strong>points</strong> over finite fields. —<br />

2.1. –––– We consider the projective varieties X e a,b<br />

given by<br />

ax e = by e + z e + v e + w e<br />

in P . We assume a, b ≠ 0 (<strong>and</strong> p ∤e) in order to avoid singularities. Observe<br />

that, even for large p, these are at most e 2 varieties up to obviousp-<br />

4p<br />

isomorphism as∗ p consists of no more than e cosets modulo (∗ p ) e .<br />

It follows from the Weil conjectures, proven by P. Deligne [Del, Théorème (8.1)],<br />

that<br />

#X e a,b (p) = p 3 + p 2 + p + 1 + E e a,b<br />

where the error-term E e a,b may be estimated by | Ee a,b | ≤ C e p 3/2 .<br />

Here, C 3 = 10 <strong>and</strong> C 4 = 60 as dim H 3 (X 3 ,Ê) = 10 for every smooth cubic<br />

threefold<strong>and</strong>dim H 3 (X 4 ,Ê) = 60for everysmoothquarticthreefoldinP 4 ().<br />

These dimensions result from the Weak Lefschetz Theorem together with<br />

F. Hirzebruch’s formula [Hi, Satz 2.4] for the Euler characteristic which actually<br />

works in much more generality.<br />

2.2. Remark. –––– Suppose e = 3 <strong>and</strong> p ≡ 2 (mod 3). Then,<br />

#X 3 a,b (p) = #X 1 a,b (p)<br />

as gcd(p − 1, 3) = 1. Similarly, for e = 4 <strong>and</strong> p ≡ 3 (mod 4), one has<br />

gcd(p − 1, 4) = 2 <strong>and</strong><br />

#X 4 a,b (p) = #X 2 a,b (p).<br />

In these cases, the error term vanishes <strong>and</strong><br />

#X e a,b (p) = p 3 + p 2 + p + 1.<br />

In the remaining cases p ≡ 1 (mod 3) for e = 3 <strong>and</strong> p ≡ 1 (mod 4) for e = 4,<br />

our goal is to compute the number ofp-<strong>rational</strong> <strong>points</strong> on X e a,b . As X e a,b ⊆ P4 ,<br />

there would be an obvious O(p 4 )-algorithm. We can do significantly better<br />

than that.<br />

2.3. Definition. –––– Let K be a field <strong>and</strong> let x ∈ K n <strong>and</strong> y ∈ K m be two vectors.<br />

Then, their convolution z := x∗y ∈ K n+m−1 is defined to be z k := ∑ x i y j .<br />

i+j=k


248 POINTS OF BOUNDED HEIGHT ON THREEFOLDS [Chap. VII<br />

2.4. Theorem (FFT convolution). —– Let n = 2 l <strong>and</strong> K be a field which contains<br />

the 2n-th roots of unity. Then, the convolution x ∗ y of two vectors x, y of<br />

length ≤ n can be computed in O(n logn) steps.<br />

Proof. The idea is to apply the Fast Fourier Transform (FFT) [Fo, Satz 20.3].<br />

The connection to the convolution is shown in [Fo, Satz 20.2] or [C/L/R,<br />

Theorem 32.8].<br />

□<br />

Theorem 2.4 is the basis for the following algorithm.<br />

2.5. Algorithm (FFT point counting on X e a,b<br />

). —–<br />

i) Initialize a vector x[0 . . . p] with zeroes.<br />

ii) Let r run from 0 to p − 1 <strong>and</strong> increase x[r e mod p] by 1.<br />

iii) Calculate ỹ := x ∗ x ∗ x by FFT convolution.<br />

iv) Normalize by putting y[i] := ỹ[i] + ỹ[i + p] + ỹ[i + 2p] for each<br />

i ∈ { 0, . . . , p − 1 }.<br />

v) Initialize N as zero.<br />

vi) (Counting <strong>points</strong> with first coordinate ≠ 0)<br />

Let j run from 0 to p − 1 <strong>and</strong> increase N by y[(a − bj 4 ) mod p].<br />

vii) (Counting <strong>points</strong> with first coordinate 0 <strong>and</strong> second coordinate ≠ 0)<br />

Increase N by y[(−b) mod p].<br />

viii) (Counting <strong>points</strong> with first <strong>and</strong> second coordinate 0)<br />

Increase N by (y[0] − 1)/(p − 1).<br />

ix) Return N as the number of allp-valued <strong>points</strong> on X e a,b .<br />

2.6. Remark. –––– For the running-time, step iii) is dominant. Therefore, the<br />

running-time of Algorithm 2.5 is O(p log p).<br />

To count, for fixed e <strong>and</strong> p,p-<strong>rational</strong> <strong>points</strong> on X e a,b<br />

with varying a <strong>and</strong> b,<br />

one needs to execute the first four steps only once. Afterwards, one may<br />

perform steps v) through ix) for all pairs (a, b) of elements from a system<br />

of representatives for∗ p /(∗ p ) e . Note that steps v) through ix) alone are of<br />

complexity O(p).<br />

2.7. –––– We ran this algorithm for all primes p ≤ 10 6 <strong>and</strong> stored the cardinalities<br />

in a file.<br />

2.8. Remark. –––– There is a formula for #X e a,b (p) in terms of Jacobi sums.<br />

A skilful manipulation of these sums should lead to another efficient algorithm<br />

which serves the same purpose as Algorithm 2.5.


Sec. 2] COMPUTING THE TAMAGAWA NUMBER 249<br />

ii. The local factors at finite places. —<br />

2.9. –––– We are interested in the Euler product<br />

τ e a,b,fin := ∏ −<br />

p∈È(1 1 )<br />

p<br />

lim<br />

m→∞<br />

#X e<br />

a,b<br />

(/pm)<br />

.<br />

p 3m<br />

2.10. Lemma. –––– a) (Good reduction)<br />

If p ∤ abe then the sequence (#X e<br />

a,b (/p m)/p 3m ) m∈Æis constant.<br />

b) (Bad reduction)<br />

i) If p divides ab but not e then the sequence (#X e<br />

a,b (/p m)/p 3m ) m∈Æbecomes<br />

stationary as soon as p m divides neither a nor b.<br />

ii) If p = 2 <strong>and</strong> e = 4 then the sequence (#X e<br />

a,b (/p m)/p 3m ) m∈Æbecomes<br />

stationary as soon as 2 m does not divide 8a or 8b.<br />

iii) If p = 3 <strong>and</strong> e = 3 then the sequence (#X e<br />

a,b (/p m)/p 3m ) m∈Æbecomes<br />

stationary as soon as 3 m divides neither 3a nor 3b.<br />

iii. An estimate. —<br />

2.11. Theorem. –––– For every pair (a, b) of integers such that a, b ≠ 0, the<br />

Euler product τ e a,b,fin<br />

is convergent.<br />

Proof. Cf. Lemma VI.5.34 where this is proven in more generality.<br />

Let p be a prime bigger than | a|, | b|, <strong>and</strong> e. Then, the factor at p is<br />

τ p := (1 − 1 p )(1 + p + p2 + p 3 + D p p 3/2 )/p 3<br />

where |D p | ≤ C e for C 3 = 10 <strong>and</strong> C 4 = 60, respectively. As sums are easier to<br />

estimate than products, we take a look at the logarithm,<br />

log τ p = D p<br />

p 3/2 + O(p−5/2 ) .<br />

Taking the logarithm, we consider ∑ p log τ p . In the case e = 3, the sum is<br />

effectively over the primes p = 1 (mod 3). If e = 4 then summation extends<br />

over all primes p = 1 (mod 4). In either case, we take a sum over one-half of


250 POINTS OF BOUNDED HEIGHT ON THREEFOLDS [Chap. VII<br />

all primes. This leads to the following estimate,<br />

[ ]<br />

Ce<br />

∑ | log τ p | ≤ ∑<br />

p≥N<br />

p≥N<br />

p + 3/2 O(p−5/2 ) ∼ C Z ∞<br />

e<br />

2<br />

≤<br />

N<br />

C e<br />

2 logN<br />

1<br />

t 3/2 logt dt<br />

Z ∞<br />

N<br />

t −3/2 dt =<br />

C e<br />

√ N logN<br />

.<br />

2.12. Remark. –––– We are interested in an∣explicit upper bound for<br />

∣∣∣∣ ∣ ∑ log τ p .<br />

p≥10 6<br />

Using Taylor’s formula, one gets<br />

∣ ∣ ∑<br />

D ∣∣∣∣<br />

p<br />

log τ p − ∑ ≤ 10 −8 .<br />

p<br />

p≥10 6 p≥10 3/2 6<br />

Since D p<br />

is zero for p ≡ 3 (mod 4) (or p ≡ 2 (mod 3)), the sum should be compared<br />

with log(ζ K (3/2)). Here, ζ K is the Dedekind zeta function of K =É(i)<br />

p 3/2<br />

orÉ(ζ 3 ), respectively. This yields<br />

∑<br />

p≥10 6<br />

p≡1(mod 4)<br />

√<br />

1<br />

≤ log<br />

p3/2 (1 − 2 −3/2 ) −1/2 · ∏<br />

p≡3(mod 4)<br />

ζÉ(i)(3/2)<br />

(1 − p −3 ) −1/2 · ∏(1 − p −3/2 ) −1<br />

p


Sec. 2] COMPUTING THE TAMAGAWA NUMBER 251<br />

numbers had been precomputed using FFT point counting (Algorithm 2.5).<br />

The algorithm below is based on the fact that the vast majority of the factors<br />

actually do not need to be computed. They are available from a list.<br />

2.15. Algorithm (Compute an approximate value for τ 3 (τ 4 a,b,fin a,b,fin<br />

)). —–<br />

i) Let p run over all prime numbers such that p ≡ 2 (mod 3) (p ≡ 3 (mod 4))<br />

<strong>and</strong> p ≤ N <strong>and</strong> calculate the product of all values of (1 − 1/p 4 ).<br />

ii) Compute the factor corresponding to p = 3 (p = 2) by Lemma 2.10.b).<br />

iii) Let p run over all prime numbers such that p ≡ 1 (mod 3) (p ≡ 1 (mod 4))<br />

<strong>and</strong> p ≤ N . Calculate the product of the factors described below.<br />

If p|ab then the corresponding factor is given by Lemma 2.10.b). Otherwise,<br />

compute the e-th power residue-symbols of a <strong>and</strong> b <strong>and</strong> look up the<br />

precomputed factor for thisp-isomorphism class of varieties in the list.<br />

iv) Multiply the two products from steps i) <strong>and</strong> iii) <strong>and</strong> the factor from step ii)<br />

with each other. Correct the product by taking the bad primes p ≡ 2 (mod 3)<br />

(p ≡ 3 (mod 4)) into consideration.<br />

2.16. Remark. –––– When we meet a bad prime p, we have to count<br />

/p m-valued <strong>points</strong> on X e<br />

a,b<br />

. This is done by an algorithm which is very<br />

similar to Algorithm 2.5.<br />

2.17. –––– We used Algorithm 2.15 to compute the Euler products τ 3 a,b,fin<br />

<strong>and</strong> τ 4 a,b,fin for a, b = 1, . . . , 100. We did all calculations for N = 106 .<br />

Note that step i) had to be done only once for e = 3 <strong>and</strong> once for e = 4.<br />

v. The factor at the infinite place. —<br />

2.18. –––– In order to get an approximation for the integrals, we tried to use<br />

some st<strong>and</strong>ard methods from numerical analysis [Kr, Chapter 9], particularly<br />

the Gauß-Legendre formula. A direct application of this method is known to<br />

work well as long as the integr<strong>and</strong> is smooth enough <strong>and</strong> the dimension of the<br />

domain of integration is not too big.<br />

2.19. –––– For the quartic X 4 a,b<br />

, we have the integral<br />

over<br />

ω ∞<br />

(<br />

X<br />

4<br />

a,b (Ê) ) = 1<br />

4 4 √ a<br />

ZZZZ<br />

R<br />

1<br />

dy dz dv dw<br />

(by 4 + z 4 + v 4 + w 4 )<br />

3/4<br />

R := {(y, z, v, w) ∈Ê4 | |y|, |z|, |v|, |w| ≤ 1 <strong>and</strong> |by 4 + z 4 + v 4 + w 4 | ≤ a} .<br />

The integr<strong>and</strong> is singular in one point. We used a simple substitution to make<br />

it sufficiently smooth for numerical integration.


252 POINTS OF BOUNDED HEIGHT ON THREEFOLDS [Chap. VII<br />

On the other h<strong>and</strong>, for the cubic X 3 a,b<br />

, we have to consider<br />

ω ∞<br />

(<br />

X<br />

3<br />

a,b (Ê) ) = 1<br />

6 3 √ a<br />

ZZZZ<br />

R<br />

1<br />

dy dz dv dw<br />

(by 3 + z 3 + v 3 + w 3 )<br />

2/3<br />

for R := {(y, z, v, w) ∈Ê4 | |y|, |z|, |v|, |w| ≤ 1 <strong>and</strong> |by 3 +z 3 +v 3 +w 3 | ≤ a} .<br />

The difficulty here is the h<strong>and</strong>ling of the singularity of the integr<strong>and</strong>. It is located<br />

in the zero set of by 3 + z 3 + v 3 + w 3 in R which is a cone over a cubic surface.<br />

Since (by 3 +z 3 +v 3 +w 3 ) −2/3 is a homogeneous function, it is enough to integrate<br />

over the boundary of R. This reduces the problem to several three-dimensional<br />

integrals of functions having a two-dimensional singular locus. If a ≥ b + 3<br />

then R is a cube <strong>and</strong> the boundary of R is easy to describe. We restricted<br />

our attention to this case. We smoothed the singularities by separation of<br />

Puiseux expansions <strong>and</strong> substitutions. The resulting integrals were treated by<br />

the Gauß-Legendre formula [Kr].<br />

3. On the geometry of diagonal cubic threefolds<br />

i. Surfaces on a hypersurface in P 4 . —<br />

3.1. Lemma. –––– Let X ⊂ P 4 be any smooth hypersurface. Then, every (reduced<br />

but possibly singular) surface S ⊂ X is a complete intersection X ∩ H d with a<br />

hypersurface H d ⊂ P 4 .<br />

Proof. By the Noether-Lefschetz Theorem, we have Pic(X ) ∼ =. The surface<br />

S is a Weil divisor on X. Hence, O(S) = O(d) ∈ Pic(X ) for a<br />

certain d > 0. The restriction Γ(P 4 , O(d)) → Γ(X, O(d)) is surjective as<br />

H 1 (P 4 , O X (d − degX)) = 0 [Ha77, Theorem III.5.1.b)].<br />

□<br />

ii. Elliptic Cones. —<br />

3.2. –––– Let X ⊂ P 4 () be the diagonal cubic threefold given by the equation<br />

x 3 + y 3 + z 3 + v 3 + w 3 = 0. Fix ζ ∈such that ζ 3 = 1. Then, for every<br />

point (x 0 : y 0 : z 0 ) on the elliptic curve F : x 3 + y 3 + z 3 = 0, the line given by<br />

(x : y : z) = (x 0 : y 0 : z 0 ) <strong>and</strong> v = −ζw is contained in X. All these lines<br />

together form a cone CF over F the cusp of which is (0 : 0 : 0 : −ζ : 1). C F is<br />

a singular model of a ruled surface over an elliptic curve. This shows, there are<br />

no other <strong>rational</strong> curves contained in C F .


Sec. 3] ON THE GEOMETRY OF DIAGONAL CUBIC THREEFOLDS 253<br />

By permuting coordinates, one finds a total of thirty elliptic cones of that type<br />

within X. The cusps of these cones are usually named Eckardt <strong>points</strong> [Mu-e,<br />

Cl/G]. We call the lines contained in one of these cones the obvious lines lying<br />

on X. It is clear that there are an infinite number of lines on X running through<br />

each of the thirty Eckardt <strong>points</strong> (1 : −1 : 0 : 0 : 0), (1 : 0 : −1 : 0 : 0),<br />

. . . , (0 : 0 : 0 : 1 : −1), (1 : −e 2πi/3 : 0 : 0 : 0), . . . , (0 : 0 : 0 : 1 : −e −2πi/3 ).<br />

3.3. Proposition. (cf. [Mu-e, Lemma 1.18]). —– Let X ⊂ P 4 be the diagonal<br />

cubic threefold given by theequation x 3 +y 3 +z 3 +v 3 +w 3 = 0. Then, through each<br />

pointP ∈ X differentfromthethirtyEckardt<strong>points</strong>therearepreciselysixlinesonX.<br />

Proof. Let P = (x 0 : y 0 : z 0 : v 0 : w 0 ). A line l through<br />

P <strong>and</strong> another point Q = (x : y : z : v : w) is parametrized by<br />

(s : t) ↦→ ((sx 0 + tx) : . . . : (sw 0 + tw)). Comparing coefficients at s 2 t, st 2 ,<br />

<strong>and</strong> t 3 , we see that the condition that l lies on X may be expressed by the three<br />

equations below.<br />

x 2 0x + y 2 0y +z 2 0z + v 2 0v + w 2 0w = 0 (†)<br />

x 0 x 2 + y 0 y 2 + z 0 z 2 + v 0 v 2 + w 0 w 2 = 0 (‡)<br />

x 3 + y 3 + z 3 + v 3 + w 3 = 0 (§)<br />

The first equation means that Q lies on the tangent hyperplane H P at P while<br />

equation (§) just encodes that Q ∈ X. By Lemma 3.6, H P ∩ X is an irreducible<br />

cubic surface.<br />

On the other h<strong>and</strong>, the quadratic form q on the left h<strong>and</strong> side of equation (‡)<br />

is of rank at least 3 as P is not an Eckardt point. Therefore, q is not just the<br />

product of two linear forms. In particular, q| HP ≢ 0.<br />

As H P ∩X is irreducible, Z(q| HP ) <strong>and</strong> H P ∩X do not have a component in common.<br />

By Bezout’s theorem, their intersection in H P is a curve of degree 6. □<br />

3.4. Remark. –––– It may happen that some of the six lines coincide. Actually,<br />

it turns out that a line appears with multiplicity > 1 if <strong>and</strong> only if it is<br />

obvious [Mu-e, Lemma 1.19]. In particular, for a general point P the six lines<br />

through it are different from each other.<br />

Under certain exceptional circumstances it is possible to write down all six<br />

lines explicitly. For example, if P = ( 3 √ −4 : 1 : 1 : 1 : 1) then the line<br />

( 3 √ −4t : (t + s) : (t + is) : (t − s) : (t − is)) through P lies on X. Permuting the<br />

three rightmost coordinates yields all six lines.<br />

3.5. Remark. –––– The following lemma is a special case of Zak’s theorem [Za,<br />

Corollary 1.8] which is more elementary <strong>and</strong> completely sufficient for our purposes.<br />

We present it here for convenience of the reader.


254 POINTS OF BOUNDED HEIGHT ON THREEFOLDS [Chap. VII<br />

3.6. Lemma. –––– Let X ⊂ P 4 be the diagonal cubic threefold given by the equation<br />

x 3 +y 3 +z 3 +v 3 +w 3 = 0. Then, for any hyperplane H ⊂ P 4 , the intersection<br />

H ∩ X is irreducible <strong>and</strong> has at most finitely many singular <strong>points</strong>.<br />

Proof. The intersection of X with a hyperplane H is singular precisely in those<br />

<strong>points</strong> where H is tangent to X.<br />

The tangent hyperplane H at (x 0 : y 0 : z 0 : v 0 : w 0 ) ∈ X is given by<br />

x 2 0 x + y2 0 y + z2 0 z + v2 0 v + w2 0w = 0. From this formula, we see that H is<br />

tangent to X at all <strong>points</strong> of the form (±x 0 : ±y 0 : ±z 0 : ±v 0 : ±w 0 ) which<br />

happen to lie on X <strong>and</strong> at no others. By consequence, H ∩ X admits only a<br />

finite number of singular <strong>points</strong>.<br />

Every irreducible component of H ∩ X is a hypersurface in the projective<br />

3-space H . Two such components would intersect in a curve of singular <strong>points</strong>.<br />

Therefore, H ∩ X is necessarily irreducible.<br />

□<br />

3.7. Remark. –––– For a general point p ∈ X, the intersection of its tangent<br />

hyperplane with X admits exactly one singular point. Indeed, let<br />

p = (x 0 : y 0 : z 0 : v 0 : w 0 ). If (−x 0 : y 0 : z 0 : v 0 : w 0 ) ∈ X then x 0 = 0.<br />

If (−x 0 : −y 0 : z 0 : v 0 : w 0 ) ∈ X then x 0 + ζy 0 = 0 for ζ a third root of<br />

unity or ζ = 1. At this point, up to permutation of coordinates, the list of all<br />

possibilities is already complete. Multi-tangent hyperplanes are caused only by<br />

<strong>points</strong> lying on a finite arrangement of hyperplanes.<br />

4. Accumulating subvarieties<br />

i. The detection ofÉ-<strong>rational</strong> lines on the cubics. —<br />

4.1. –––– On a cubic threefold X 3 a,b<br />

, quadratic growth is predicted for the<br />

number ofÉ-<strong>rational</strong> <strong>points</strong> of bounded height. Lines are the only curves with<br />

such a growth rate.<br />

The moduli space of the lines on a cubic threefold is well-understood. It is a<br />

surface of general type [Cl/G, Lemma 10.13]. Nevertheless, we do not know<br />

of a method to find allÉ-<strong>rational</strong> lines on a given cubic threefold, explicitly.<br />

For that reason, we use the algorithm below which is an ir<strong>rational</strong>ity test for<br />

the six lines through a given point (x 0 : y 0 : z 0 : v 0 : w 0 ) ∈ X 3 a,b (É).<br />

4.2. Algorithm (Test the six lines through a given point for ir<strong>rational</strong>ity). —–<br />

i) Let p run through the primes from 3 to N .<br />

For each p, solve the system of equations (†), (‡), (§) (adapted to X 3 a,b ) in5 p .<br />

If the multiples of (x 0 , y 0 , z 0 , v 0 , w 0 ) are the only solutions then output that there<br />

is noÉ-<strong>rational</strong> line through (x 0 : y 0 : z 0 : v 0 : w 0 ) <strong>and</strong> terminate prematurely.


Sec. 4] ACCUMULATING SUBVARIETIES 255<br />

ii) If the loop comes to its regular end then output that the point is suspicious.<br />

It could possibly lie on aÉ-<strong>rational</strong> line.<br />

4.3. Remark. –––– We use a very naive O(p)-algorithm to solve the system<br />

of equations overp. If, say, x 0 ≠ 0 then it is sufficient to consider quintuples<br />

such that x = 0. We parametrize the projective plane given by (†). Then, we<br />

compute all <strong>points</strong> on the conic given by (†) <strong>and</strong> (‡). For each such point, we<br />

compute the cubic form on the left h<strong>and</strong> side of (§). When a non-trivial solution<br />

is found, we stop immediately.<br />

We carried out the ir<strong>rational</strong>ity test on everyÉ-<strong>rational</strong> point found on any<br />

of the cubics except for the <strong>points</strong> lying on an obvious line. We worked<br />

with N = 600. It turned out that suspicious <strong>points</strong> are rare <strong>and</strong> that, at least in<br />

our sample, each of them actually lies on aÉ-<strong>rational</strong> line.<br />

The lines found. We found only 42 non-obviousÉ-<strong>rational</strong> lines on all of the<br />

cubics X 3 a,b<br />

for 100 ≥ a ≥ b ≥ 1 together. Among them, there are only five<br />

essentially different ones. We present them in the table below. The list might<br />

be enlarged by two, as X21,6 3 <strong>and</strong> X 22,5 3 may be transformed into X 48,21 3 <strong>and</strong> X 40,22 3 ,<br />

respectively, by an automorphism of P 4 . Further, each line has six pairwise<br />

different images under the obvious operation of the group S 3 .<br />

TABLE 1. Sporadic lines on the cubic threefolds<br />

a b Smallest point Point s.t. x = 0<br />

19 18 (1 : 2 : 3 : -3 : -5) (0 : 7 : 1 : -7 : -18)<br />

21 6 (1 : 2 : 3 : -3 : -3) (0 : 9 : 1 : -10 : -15)<br />

22 5 (1 : -1 : 3 : 3 : -3) (0 : 27 : -4 : -60 : 49)<br />

45 18 (1 : 1 : 3 : 3 : -3) (0 : 3 : -1 : 3 : -8)<br />

73 17 (1 : 5 : -2 : 11 : -15) (0 : 27 : -40 : 85 : -96)<br />

4.4. Remark. –––– It is a priori unnecessary to search for accumulating surfaces,<br />

at least if we assume the conjectures of Batyrev/Manin <strong>and</strong> Lang formulated in<br />

Chapter VI.<br />

First of all, only <strong>rational</strong> surfaces are supposed to accumulate that many <strong>rational</strong><br />

<strong>points</strong> that it could be seen through our asymptotics of O(B 2 ). Indeed, a surface<br />

which is abelian or bielliptic may not have more than O(log t B) <strong>points</strong> of<br />

height < B. Non-<strong>rational</strong> ruled surfaces accumulate <strong>points</strong> in curves, anyway.<br />

Further, it is expected (Conjecture VI.3.1) that K3 surfaces, Enriques surfaces,<br />

<strong>and</strong> surfaces of Kodaira dimension one may have no more than O(B ε ) <strong>points</strong><br />

of height < B outside a finite union of <strong>rational</strong> curves. For surfaces of general<br />

type, finally, expectations are even stronger (Conjecture VI.2.2).


256 POINTS OF BOUNDED HEIGHT ON THREEFOLDS [Chap. VII<br />

A <strong>rational</strong> surface S is, up to exceptional curves, the image of a <strong>rational</strong><br />

map ϕ: P 2 X ⊂ P 4 . There is a bi<strong>rational</strong> morphism ε: P → P 2 such<br />

that ϕ := ϕ ◦ ε is a morphism of schemes. ε is given by a sequence of blowingups<br />

[Bv, Theorem II.11]. ϕ is defined by the linear system |dH − E| where<br />

d := deg ϕ, H is a hyperplane section, <strong>and</strong> E is the exceptional divisor. On the<br />

other h<strong>and</strong>, K := K P = −3H + E. Therefore, if d ≥ 3 then<br />

H naive (ϕ(p)) = H dH −E (p) = H 3H −<br />

3<br />

d E(p)d/3 ≥ c · H −K (p) d/3<br />

for p ∉ supp(E). Manin’s conjecture implies there are O((B log t B) 3 d ) = o(B 2 )<br />

<strong>points</strong> of height < B on the Zariski-dense subset ϕ(P \ supp(E)) ⊆ S.<br />

It remains to show that there are no <strong>rational</strong> maps ϕ: P 2 <br />

X of degree<br />

d ≤ 2. Indeed, under this assumption, deg ϕ(P 2 ) ≤ 4. This implies,<br />

by virtue of Lemma 3.1, that ϕ(P 2 ) is necessarily a hyperplane section X ∩ H .<br />

Lemma 3.6 shows that X ∩ H contains only finitely many singular <strong>points</strong>.<br />

It is, however, well known that cubic surfaces in 3-space which are the image<br />

of P 2 under a quadratic map have a singular line [Bv, Corollary IV.8].<br />

ii. The detection ofÉ-<strong>rational</strong> conics on the quartics. —<br />

4.5. –––– On a quartic threefold, linear growth is predicted for the number of<br />

É-<strong>rational</strong> <strong>points</strong> of bounded height. The assumption b > 0 ensures that there<br />

are noÊ-<strong>rational</strong> lines contained in X 4 a,b<br />

. The only other curves with at least<br />

linear growth one could think about are conics.<br />

We do not know of a method to find allÉ-<strong>rational</strong> conics on a given quartic<br />

threefold, explicitly. Worse, we were unable to create an efficient routine to test<br />

whether there is aÉ-<strong>rational</strong> conic through a given point. The resulting system<br />

of equations seems to be too complicated to h<strong>and</strong>le.<br />

Conics through two <strong>points</strong>. A conic Q through (x 0 : y 0 : z 0 : v 0 : w 0 )<br />

<strong>and</strong> (x 1 : y 1 : z 1 : v 1 : w 1 ) may be parametrized in the form<br />

(s : t) ↦→ ((λx 0 s 2 + µx 1 t 2 + xst) : . . . : (λw 0 s 2 + µw 1 t 2 + wst))<br />

for some x, y, z, v, w, λ, µ ∈. The condition that Q is contained in X 4 a,b leads<br />

to a system G of seven equations in x, y, z, v, w, <strong>and</strong> λµ. The phenomenon<br />

that λ <strong>and</strong> µ do not occur individually is explained by the fact that they are not<br />

invariant under the automorphisms of P 1 which fix 0 <strong>and</strong> ∞.<br />

4.6. Algorithm (Test for conic through two <strong>points</strong>). —– i) Let p run through<br />

the primes from 3 to N .


Sec. 4] ACCUMULATING SUBVARIETIES 257<br />

In the exceptional case that G could allow a solution such that p|x, y, z, v, w but<br />

p 2 ∤ λµ, do nothing. Otherwise, solve G in6 p . If (0, 0, 0, 0, 0, 0)is the only solution<br />

then output that there is noÉ-<strong>rational</strong> conic through (x 0 : y 0 : z 0 : v 0 : w 0 )<br />

<strong>and</strong> (x 1 : y 1 : z 1 : v 1 : w 1 ) <strong>and</strong> terminate prematurely.<br />

ii) If the loop comes to its regular end then output that the pair is suspicious.<br />

It could possibly lie on aÉ-<strong>rational</strong> conic.<br />

4.7. –––– To solve the system G in6 p , we use an O(p)-algorithm. Actually,<br />

comparison of coefficients at s 7 t <strong>and</strong> st 7 yields two linear equations in x, y,<br />

z, v, <strong>and</strong> w. We parametrize the projective plane I given by them. Comparison<br />

of coefficients at s 6 t 2 <strong>and</strong> s 2 t 6 leads to a quadric O <strong>and</strong> an equation<br />

λµ = q(x, y, z, v, w)/M with a quadratic form q over<strong>and</strong> an integer<br />

M ≠ 0. The case p|M sends us to the next prime immediately. Otherwise,<br />

we compute all <strong>points</strong> on the conic I ∩ O. For each of them, we<br />

test the three remaining equations. When a non-trivial solution is found, we<br />

stop immediately.<br />

Conics through three <strong>points</strong>. Three <strong>points</strong> P 1 , P 2 , <strong>and</strong> P 3 on X 4 a,b<br />

define a projective<br />

plane P. The <strong>points</strong> together with the two tangent lines P∩ T P1 <strong>and</strong> P∩ T P2<br />

determine a conic Q, uniquely. It is easy to transform this geometric insight<br />

into a formula for a parametrization of Q. We then need a test whether a conic<br />

given in parametrized form is contained in X 4 a,b<br />

. This part is algorithmically<br />

simple but requires the use of multiprecision integers.<br />

Detecting conics. For each quartic X 4 a,b<br />

, we tested every pair ofÉ-<strong>rational</strong><br />

<strong>points</strong> of height < 100 000 for a conic through them. The existence of a<br />

conic through (P, Q) is equivalent to the existence of a conic through (gP, gQ)<br />

for g ∈ (/2) 4 ⋊S 3 ⊆ Aut(X 4 a,b<br />

). This reduces the running time by a factor of<br />

about 96. Further, pairs already known to lie on the same conic were excluded<br />

from the test.<br />

For each pair (P, Q) found suspicious, we tested the triples (P, Q, R) for R running<br />

through theÉ-<strong>rational</strong> <strong>points</strong> of height < 100 000, until a conic was found.<br />

Due to the symmetries, one finds several conics at once. For each conic detected,<br />

all <strong>points</strong> on it were marked as lying on this conic.<br />

Actually, there were a few pairs found suspicious through which no conic could<br />

be found. In any of these cases, it was easy to prove by h<strong>and</strong> that there is actually<br />

noÊ-<strong>rational</strong> conic passing through the two <strong>points</strong>. This means, we detected<br />

every conic which meets at least two of the <strong>rational</strong> <strong>points</strong> of height < 100 000.<br />

Concerning our programming efforts, this was the most complex part of the<br />

entire project.


258 POINTS OF BOUNDED HEIGHT ON THREEFOLDS [Chap. VII<br />

The conics found. Up to symmetry, we found a total of 1 664É-<strong>rational</strong> conics<br />

on all of the quartics X 4 a,b<br />

for 1 ≤ a, b ≤ 100 together.<br />

Among them, 1 538 are contained in a plane of type z = v+w <strong>and</strong> Yx−Xy = 0<br />

for (X, Y, t) a <strong>rational</strong> point on the genus one curve aX 4 − bY 4 = 2t 2 . Further,<br />

there are 93 conics which are slight modifications of the above with y<br />

interchanged with z, v, or w. This is possible if b is a fourth power.<br />

There is a geometric explanation for the occurrence of these conics. The hyperplane<br />

given by z = v + w intersects X 4 a,b<br />

in a surface S with the two singular<br />

<strong>points</strong> (0 : 0 : −1 : e ±2πi/3 : e ∓2πi/3 ). The linear projection π : S P 1 to the<br />

first two coordinates is undefined only in these two <strong>points</strong>. Its fibers are plane<br />

quartics which split into two conics as (v+w) 4 +v 4 +w 4 = 2(v 2 +vw+w 2 ) 2 . After<br />

resolution of singularities, the two conics become disjoint. ˜S is a ruled surface<br />

over a twofold cover of P 1 ramified in the four <strong>points</strong> such that ax 4 − by 4 = 0.<br />

I.e., over a curve of genus one.<br />

In the case a is twice a square, a different sort of conics comes from the equations<br />

v = z + Dy <strong>and</strong> w = Ly when (L, D) is a point on the affine genus three curve<br />

C b : L 4 + b = D 4 . Here, by 4 + z 4 + v 4 + w 4 becomes twice a full square<br />

after the substitutions. This explains why this particular intersection of X 4 a,b<br />

with a plane splits into two conics. We found 28 conics of this type. C b has<br />

aÉ-<strong>rational</strong> point for b = 5, 15, 34, 39, 65, 80, <strong>and</strong> 84. The conics actually<br />

admit aÉ-<strong>rational</strong> point for a = 2, 18, 32, <strong>and</strong> 98.<br />

The remaining five conics are given as follows. For a = 3, 12, 27, or 48<br />

<strong>and</strong> b = 10, intersect with the plane given by v = y + z <strong>and</strong> w = 2y + z.<br />

For a = 17 <strong>and</strong> b = 30, put v = 2x + y <strong>and</strong> w = x + 3y + z.<br />

4.8. Remark. –––– Again, it is not necessary to search for accumulating surfaces.<br />

Here, <strong>rational</strong> maps ϕ: P 2 X ⊂ P 4 such that deg ϕ ≤ 3 need to be<br />

taken into consideration. We claim, such a map is impossible.<br />

If deg ϕ = 3 then we had ϕ: (λ : µ : ν) ↦→ (K 0 (λ, µ, ν) : . . . : K 4 (λ, µ, ν))<br />

where K 0 , . . . , K 4 are cubic forms defined overÉ. K 0 = 0 defines a plane cubic<br />

which has infinitely many real <strong>points</strong>, automatically. As the image of ϕ<br />

is assumed to be contained in X 4 a,b , we have that K 0(λ, µ, ν) = 0 implies<br />

K 1 (λ, µ, ν) = . . . = K 4 (λ, µ, ν) = 0 for λ, µ, ν ∈Ê. By consequence,<br />

K 1 , . . . , K 4 are divisible by K 0 (or by a linear factor of K 0 in the case it is<br />

reducible) <strong>and</strong> ϕ is not of degree three.<br />

For deg ϕ ≤ 2, we had deg ϕ(P 2 ) ≤ 4 such that ϕ(P 2 ) = X ∩ H is a hyperplane<br />

section. Lemma 3.6 shows it has at most finitely many singular <strong>points</strong>.<br />

On the other h<strong>and</strong>, a quartic in P 3 which is the image of a quadratic map from P 2<br />

is a Steiner surface. It is known [Ap, p. 40] to have one, two, or (in generic case)<br />

three singular lines.


Sec. 5] RESULTS 259<br />

5. Results<br />

i. A technology to find solutions of Diophantine equations. —<br />

5.1. –––– In Chapter XI (cf. [EJ2] <strong>and</strong> [EJ3]), we describe a modification<br />

of D. Bernstein’s [Be] method to search efficiently for all solutions of naive<br />

height < B of a Diophantine equation of the particular form<br />

f (x 1 , . . . , x n ) = g(y 1 , . . . , y m ) .<br />

The expected running-time of our algorithm is O(B max{n,m} ). Its basic idea is<br />

as follows.<br />

5.2. Algorithm (Search for solutions of a Diophantine equation). —–<br />

i) (Writing)<br />

Evaluate f on all <strong>points</strong> of the cube {(x 1 , . . . , x n ) ∈n | |x i | < B} of<br />

dimension n. Store the values within a hash table H .<br />

ii) (Reading)<br />

Evaluate g on all <strong>points</strong> of the cube {(y 1 , . . . , y m ) ∈m | |y i | < B}. For each<br />

value, start a search in order to find out whether it occurs in H . When a<br />

coincidence is detected, reconstruct the corresponding values of x 1 , . . . , x n <strong>and</strong><br />

output the solution.<br />

5.3. Remark. –––– In the case of a variety X e a,b<br />

, the running-time is obviously<br />

O(B 3 ). We decided to store the values of z e +v e +w e into the hash table.<br />

Afterwards, we have to look up the values of ax e − by e .<br />

In this form, the algorithm would lead to a program in which almost the entire<br />

running-time is consumed by the writing part. Observe, however, the following<br />

particularity of our method. When we search on up to O(B) threefolds,<br />

differing only by the values of a <strong>and</strong> b, simultaneously, then the running-time<br />

is still O(B 3 ).<br />

5.4. Remark (Running times). —– We worked with B = 5 000 for the cubics<br />

<strong>and</strong> B = 100 000 for the quartics. In either case, we dealt with all threefolds<br />

arising for a, b = 1, . . . , 100, simultaneously.<br />

The by far largest portion of the running time was spent on point search on<br />

the quartics. All in all, this took around 155 days of CPU time. This is<br />

approximately only three times longer than searching on a single threefold<br />

had lasted. Searching for conics was done within 32 days. In comparison with<br />

this, the corresponding computations for the cubics could be done in a negligible<br />

amount of time. The reason for this is simply that for the cubics the search


260 POINTS OF BOUNDED HEIGHT ON THREEFOLDS [Chap. VII<br />

bound was by far lower. A program with integrated line detection took us<br />

approximately ten days.<br />

To compute Peyre’s constants, the precomputation was actually the main part.<br />

Point counting using FFT took 100 hours for the cubics <strong>and</strong> 100 hours for<br />

the quartics. For the final computation, the running-time was a quarter of an<br />

hour for either exponent.<br />

ii. The results for the cubics. —<br />

5.5. –––– We counted allÉ-<strong>rational</strong> <strong>points</strong> of height less than 5 000 on the<br />

threefolds X 3 a,b where a, b = 1, . . . , 100 <strong>and</strong> b ≤ a. Note that X 3 a,b ∼ = X 3 b,a .<br />

Points lying on one of the elliptic cones or on a sporadicÉ-<strong>rational</strong> line<br />

in X a,b were excluded from the count. The smallest number of <strong>points</strong> found is<br />

3 930 278 for (a, b) = (98, 95). The largest numbers of <strong>points</strong> are 332 137 752<br />

for (a, b) = (7, 1) <strong>and</strong> 355 689 300 in the case that a = 1 <strong>and</strong> b = 1.<br />

On the other h<strong>and</strong>, for each threefold X 3 a,b<br />

whereas a, b = 1, . . . , 100<br />

<strong>and</strong> b + 3 ≤ a, we calculated the expected number of <strong>points</strong> <strong>and</strong> the quotients<br />

# { <strong>points</strong> of height < B found } / # { <strong>points</strong> of height < B expected }.<br />

Let us visualize the quotients by two histograms.<br />

250<br />

250<br />

200<br />

200<br />

150<br />

150<br />

100<br />

100<br />

50<br />

50<br />

0 0.88 0.9 0.92 0.94 0.96 0.98 1 1.02 0 0.88 0.9 0.92 0.94 0.96 0.98 1 1.02<br />

FIGURE 1. Distribution of the quotients for B = 1 000 <strong>and</strong> B = 5 000.


Sec. 5] RESULTS 261<br />

The statistical parameters are listed in the table below.<br />

TABLE 2. Parameters of the distribution in the cubic case<br />

B = 1 000 B = 2 000 B = 5 000<br />

mean value 0.98179 0.98854 0.99383<br />

st<strong>and</strong>ard deviation 0.01274 0.00823 0.00455<br />

iii. The results for the quartics. —<br />

5.6. –––– We counted allÉ-<strong>rational</strong> <strong>points</strong> of height less than 100 000 on<br />

the threefolds X 4 a,b<br />

where a, b = 1, . . . , 100. It turns out that on 5 015 of<br />

these varieties, there are noÉ-<strong>rational</strong> <strong>points</strong> occurring at all as the equation<br />

is unsolvable inÉp for some small p. In this situation, Manin’s conjecture is<br />

true, trivially.<br />

For the remaining varieties, the <strong>points</strong> lying on a knownÉ-<strong>rational</strong> conic in X a,b<br />

were excluded from the count. Table 3 shows the quartics sorted by the numbers<br />

of <strong>points</strong> remaining.<br />

TABLE 3. Numbers of <strong>points</strong> of height < 100 000 on the quartics.<br />

a b # <strong>points</strong> # not on conic # expected<br />

(by Manin-Peyre)<br />

29 29 2 2 135<br />

58 87 288 288 272<br />

58 58 290 290 388<br />

87 87<br />

. .<br />

386<br />

.<br />

386<br />

.<br />

357<br />

.<br />

34 1 9938976 5691456 5 673000<br />

17 64 5708664 5708664 5 643000<br />

1 14 7205502 6361638 6 483000<br />

3 1 12657056 7439616 7 526000<br />

We see that the variation of the quotients is higher than in the cubic case.


262 POINTS OF BOUNDED HEIGHT ON THREEFOLDS [Chap. VII<br />

20<br />

15<br />

10<br />

5<br />

0 0.7 0.8 0.9 1 1.1 1.2 1.3<br />

FIGURE 2. Distribution of the quotients for B = 100 000.<br />

7<br />

7<br />

6<br />

6<br />

5<br />

5<br />

4<br />

4<br />

3<br />

3<br />

2<br />

2<br />

1<br />

1<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

FIGURE 3. Distribution of the quotients for B = 1 000 <strong>and</strong> B = 10 000.<br />

The statistical parameters are listed in the table below.


Sec. 5] RESULTS 263<br />

TABLE 4. Parameters of the distribution in the quartic case<br />

B = 1 000 B = 10 000 B = 100 000<br />

mean value 0.9853 0.9957 0.9982<br />

st<strong>and</strong>ard deviation 0.3159 0.1130 0.0372<br />

iv. Interpretation of the result. —<br />

5.7. –––– The results suggest that Manin’s conjecture should be true for the<br />

two families of threefolds considered. In the cubic case, the st<strong>and</strong>ard deviation<br />

is by far smaller than in the case of the quartics. This, however, is not very<br />

surprising as on a cubic there tend to be much more <strong>rational</strong> <strong>points</strong> than on a<br />

quartic. This makes the sample more reliable.<br />

5.8. Remark. –––– The data we collected might be used to test the sharpening<br />

of the asymptotic formula (∗) suggested by Sir P. Swinnerton-Dyer [S-D05].<br />

5.9. Question. –––– Our calculations seem to indicate that the number of <strong>rational</strong><br />

<strong>points</strong> often approaches its expected value from below. Is that more than<br />

an accidental effect?


CHAPTER VIII<br />

ON THE SMALLEST POINT ON A<br />

DIAGONAL QUARTIC THREEFOLD ∗<br />

In the course of your work, you will from time to time encounter the situation<br />

where the facts <strong>and</strong> the theory do not coincide. In such circumstances,<br />

young gentlemen, it is my earnest advice to respect the facts.<br />

IGOR IVANOVICH SIKORSKY<br />

1. A computer experiment<br />

1.1. –––– Let X ⊆ P nÉbe a Fano variety defined overÉ. If X (Éν) ≠ ∅ for<br />

every ν ∈ Val(É) then it is natural to ask whether X (É) ≠ ∅ (Hasse’s principle).<br />

Further, it would be desirable to have an a-priori upper bound for the height<br />

of the smallestÉ-<strong>rational</strong> point on X as this would allow to effectively decide<br />

whether X (É) ≠ ∅ or not.<br />

When X is a conic, Legendre’s theorem on zeroes of ternary quadratic forms<br />

proves the Hasse principle <strong>and</strong>, moreover, yields an effective bound for the<br />

smallest point. For quadrics of arbitrary dimension, the same is true by an<br />

observation due to J. W. S. Cassels [Cas55]. Further, there is a theorem of<br />

C. L. Siegel [Si73, Satz 1] which provides a generalization to hypersurfaces<br />

defined by norm equations. For more general Fano varieties, no theoretical<br />

upper bound is known for the height of the smallestÉ-<strong>rational</strong> point. Some of<br />

these varieties fail the Hasse principle.<br />

In this chapter, we present some theoretical <strong>and</strong> experimental results concerning<br />

the height of the smallestÉ-<strong>rational</strong> point on quartic hypersurfaces in P 4É.<br />

1.2. –––– There is the conjecture, due to Yu. I. Manin, that the number ofÉ-<strong>rational</strong><br />

<strong>points</strong> of anticanonical height < B on a Fano variety X is asymptotically<br />

equal to τB log rkPic(X )−1 B, for B → ∞. A detailed description of Manin’s<br />

conjecture is given in Chapter VI.<br />

(∗) This chapter is a revised version of the article: On the smallest point on a diagonal quartic<br />

threefold, J. Ramanujan Math. Soc. 22(2007), 189–204, joint with A.-S. Elsenhans.


266 ON THE SMALLEST POINT ON A DIAGONAL QUARTIC THREEFOLD [Chap. VIII<br />

In the particular case of a quartic threefold, the anticanonical height is the<br />

same as the naive height. Further, rkPic(X ) = 1 <strong>and</strong>, finally, the coefficient<br />

τ ∈Êequals the Tamagawa-type number τ (X ) introduced by E. Peyre (Definition<br />

VI.5.35).<br />

Thus, one expects ∼τ (X )B <strong>points</strong> of height < B. Assuming equidistribution,<br />

the height of the smallest point should be ∼ 1<br />

τ<br />

. Being a bit optimistic, this<br />

(X )<br />

might lead to the expectation that m(X ), the height of the smallestÉ-<strong>rational</strong><br />

point on X, is always less than C<br />

τ<br />

for a certain absolute constant C.<br />

(X )<br />

1.3. –––– To test this expectation, we computed the Tamagawa number <strong>and</strong><br />

ascertained the smallestÉ-<strong>rational</strong> point for each of the quartic threefolds<br />

X (−a,b)<br />

4<br />

⊂ P 4Égiven by ax 4 = by 4 + z 4 + v 4 + w 4 for a, b = 1, . . . , 1000.<br />

On 516 820 of these varieties, there are noÉ-<strong>rational</strong> <strong>points</strong> as the equation is<br />

unsolvable inÉp for p = 2, 5, or 29. (Note that, for each prime p different<br />

from 2, 5, or 29, there are p-adic <strong>points</strong> already on the Fermat quartic given<br />

by z 4 + v 4 + w 4 = 0.) On each of the remaining varieties,É-<strong>rational</strong> <strong>points</strong><br />

were found. In other words, there are no counterexamples to the Hasse principle<br />

in this family.<br />

The computation of Tamagawa numbers is explained in Chapter VII, Section 2.<br />

The methods to systematically search for solutions of Diophantine equations<br />

we applied are described in Chapter VII, Section 5.<br />

The results are summarized by the diagrams below.<br />

a, b = 1 . . . 100 a, b = 1 . . . 1000<br />

FIGURE 1. Height of smallest point versus Tamagawa number<br />

It is apparent from the diagrams that the experiment agrees with the expectation<br />

above. The slope of a line tangent to the top right of each of the scatter<br />

plots is indeed near (−1). However, we show in Section 2 that, in general, the<br />

inequality m(X ) < C<br />

τ<br />

does not hold. The following remains a logical possibility.<br />

(X )


Sec. 1] A COMPUTER EXPERIMENT 267<br />

1.4. Question. –––– For every ε > 0, does there exist a constant C (ε) such<br />

that, for each quartic threefold,<br />

m(X ) <<br />

C (ε)<br />

τ (X ) 1+ε ?<br />

1.5. Peyre’s constant. –––– Recall from the previous chapter that, in our case,<br />

E. Peyre’s Tamagawa-type number may be defined as an infinite product<br />

τ (X ) := ∏ τ ν (X ) .<br />

ν∈Val(É)<br />

Indeed, for a smooth quartic threefold X, one has O(−K ) = O(1). In particular,<br />

α(X ) = 1.<br />

For a prime number p, the definition of the local factor may be simplified to<br />

(<br />

τ p (X ) := 1 − 1 )<br />

#X (/p<br />

· m)<br />

lim<br />

p m→∞ p 3m<br />

for X the model in P ngiven by the equation defining X.<br />

τ ∞ (X ) is described in full generality in Definition VI.5.22. In the case of<br />

the diagonal quartic threefold X (a 0, ... ,a 4 )<br />

given by a 0 x 4 0 + . . . + a 4 x 4 4 = 0<br />

(a 0 < 0, a 1 , . . . , a 4 > 0) in P 4É, this yields<br />

τ ∞ (X (a 0, ... ,a 4 ) ) = 1<br />

4 4 √<br />

|a0 |<br />

ZZZZ<br />

where the domain of integration is<br />

R<br />

1<br />

(a 1 y 4 + a 2 z 4 + a 3 v 4 + a 4 w 4 ) 3/4 dy dz dv dw (∗)<br />

R := { (y, z, v, w) ∈ [−1, 1] 4 | | a 1 y 4 + a 2 z 4 + a 3 v 4 + a 4 w 4 | ≤ |a 0 | } .<br />

1.6. –––– In the case of diagonal quartic threefolds, there is an estimate for<br />

1<br />

m(X ) in terms of τ (X ). Namely,<br />

τ<br />

admits a fundamental finiteness property.<br />

(X )<br />

More precisely, in Section 3, we will show the following results.<br />

Theorem. Leta = (a 0 , . . . , a 4 )beavectorsuchthata 0 , . . . , a 4 ∈,a 0 < 0, <strong>and</strong><br />

a 1 , . . . , a 4 > 0. Denote by X a the quartic in P 4Égiven by a 0 x 4 0+ . . . +a 4 x 4 4 = 0.<br />

Then, for each ε > 0 there exists a constant C (ε) > 0 such that<br />

1<br />

τ (X a ) ≥ C (ε) · H( 1<br />

a 0<br />

: . . . :<br />

1<br />

a 4<br />

) 1<br />

4<br />

−ε<br />

.<br />

Corollary (Fundamental Finiteness). For each B > 0, there are only finitely many<br />

quartics X a : a 0 x 4 0 + . . . + a 4 x 4 4 = 0 in P 4Ésuch that a 0 < 0, a 1 , . . . , a 4 > 0,<br />

<strong>and</strong> τ (X a ) > B.


268 ON THE SMALLEST POINT ON A DIAGONAL QUARTIC THREEFOLD [Chap. VIII<br />

Corollary (An inefficient search bound). There exists a monotonically decreasing<br />

function S : (0, ∞) → [0, ∞), the search bound, satisfying the following condition.<br />

Let X a be the quartic threefold given by the equation a 0 x 4 0 + . . . + a 4 x 4 4 = 0.<br />

Assume a 0 < 0, a 1 , . . . , a 4 > 0, <strong>and</strong> X a (É) ≠ ∅. Then, X a admits aÉ-<strong>rational</strong><br />

point of height ≤ S(τ (X a )).<br />

Proof. One may simply put S(t) := max min H(P).<br />

τ (X a )≥t P∈X a (É) X a (É)≠∅<br />

□<br />

In other words, we have m(X a ) ≤ S(τ (X a )) as soon as X a (É) ≠ ∅.<br />

2. A negative result<br />

For a ∈Æ, let X (−a) ⊂ P 4Ébe given by ax 4 = y 4 + z 4 + v 4 + w 4 <strong>and</strong> let<br />

m(X (−a) ) := min { H(x : y : z : v : w) | (x : y : z : v : w) ∈ X (−a) (É)}<br />

be the smallest height of aÉ-<strong>rational</strong> point on X (−a) . We compare m(X (−a) )<br />

with the Tamagawa type number τ (−a) := τ (X (−a) ).<br />

2.1. Notation. –––– For a prime number p <strong>and</strong> integers y, z, . . . , not all of<br />

which are equal to zero, we write gcd p<br />

(y, z, . . . ) for the largest power of p<br />

dividing all of the y, z, . . . .<br />

2.2. Theorem. –––– There is no constant C such that<br />

for all a ∈Æ.<br />

m(X (−a) ) < C<br />

τ (−a)<br />

Proof. The proof consists of several steps.<br />

First step. For a ≥ 4, one has τ (−a)<br />

∞ = 1 4√ a<br />

I where I is an integral independent<br />

of a.<br />

This follows immediately from formula (∗) above.<br />

Second step. For the height of the smallest point, we have m(X (−a) ) ≥ 4 √ a<br />

4 .<br />

|x| ≥ 1 yields y 4 + z 4 + v 4 + w 4 ≥ a <strong>and</strong> max{|y|, |z|, |v|, |w|} ≥ 4 √ a<br />

4 .<br />

Third step. There are two positive constants C 1 <strong>and</strong> C 2 such that, for all a ∈Æ,<br />

C 1 < ∏<br />

τ (−a)<br />

p prime<br />

p>13,p∤a<br />

p < C 2 .


Sec. 2] A NEGATIVE RESULT 269<br />

For a prime p of good reduction, Hensel’s lemma shows<br />

(<br />

τ p (X (−a) ) = 1 − 1 )<br />

· #X (−a) (p)<br />

.<br />

p p 3<br />

Further, for the number of <strong>points</strong> on a non-singular variety over a finite<br />

field, there are excellent estimates provided by the Weil conjectures, proven<br />

by P. Deligne. In our situation, [Del, Théorème (8.1)] may be directly applied.<br />

It shows #X (−a) (p) = p 3 + p 2 + p + 1 + E (−a) with an errorterm<br />

|E (−a) | ≤ 60p 3/2 . Note that dim H 3 (X,Ê) = 60 for every smooth quartic<br />

threefold X in P 4[Hi, Satz 2.4].<br />

Consequently,<br />

1 −<br />

60(1 − 1/p)<br />

p 3/2<br />

− 1 p 4 ≤ τ (−a)<br />

p<br />

≤ 1 +<br />

60(1 − 1/p)<br />

p 3/2 − 1 p 4 .<br />

Here, the left h<strong>and</strong> side is positive for p > 13. The infinite product over all<br />

1 − 60(1−1/p) − 1 (respectively 1 + 60(1−1/p) − 1 ) is convergent.<br />

p 3/2 p 4 p 3/2 p 4<br />

Fourth step. There is a sequence (a i ) i∈Æof<br />

i∈Æ<br />

natural numbers such that<br />

( )<br />

∏ τ (−a i)<br />

p<br />

p prime<br />

p≤13 or p|ai<br />

is unbounded.<br />

Let C ∈Êbe given. We will show that<br />

∏<br />

τ p<br />

(−a)<br />

p prime<br />

p≤13 or p|a<br />

> C<br />

when a := p 1 · . . . · p r is a product of sufficiently many different primes<br />

p i ≡ 3 (mod 4) fulfilling a ≡ 1 (mod M ) for M := 16 · 3 · 5 · 7 ·11·13.<br />

Let p be a prime such that p|a. We made sure that p 4 ∤a. Then, for any<br />

point (x : y : z : v : w) ∈ X (−a) (/p n), the assumption p | y, z, v, w<br />

would imply p 4 |ax 4 <strong>and</strong> p|x. Therefore, gcd p<br />

(y, z, v, w) = 1. Further,<br />

y 4 + z 4 + v 4 + w 4 ≡ 0 (mod p).<br />

As p ≡ 3 (mod 4), the number of solutions of that congruence is the same as<br />

that of y 2 + z 2 + v 2 + w 2 ≡ 0 (mod p). Since p remains prime in[i], this<br />

quadratic form is a direct sum of two norm forms. Its number of zeroes in4 p<br />

is therefore equal to 1 + (p − 1)(p + 1) 2 .<br />

gcd p<br />

(y, z, v, w) = 1 implies that Hensel’s lemma is applicable. It shows<br />

#X (−a) (/p n) = pn · (p − 1)(p + 1) 2 p 3(n−1)<br />

(p − 1)p n−1 = p 3n−2 (p + 1) 2 .


270 ON THE SMALLEST POINT ON A DIAGONAL QUARTIC THREEFOLD [Chap. VIII<br />

Hence, τ (−a)<br />

p<br />

= (1 − 1 p ) (p+1)2<br />

p 2<br />

∏<br />

p prime<br />

p≤13 or p|a<br />

τ (−a)<br />

p =<br />

= 1 + 1 − 1 − 1 . We may consequently write<br />

p p 2 p 3<br />

r (<br />

1 + 1 − 1 − 1 )<br />

·<br />

p i pi<br />

2 p<br />

∏τ (−a)<br />

i<br />

3 p .<br />

∏<br />

i=1<br />

The second product is over p = 2 <strong>and</strong> a finite number of primes of good reduction.<br />

The value of the product depends only on the residue of a modulo<br />

M = 16·3·5·7·11·13by virtue of Lemma VII.2.10.a) <strong>and</strong> b.ii). In particular,<br />

our assumption a ≡ 1 (mod M ) implies<br />

T := ∏<br />

τ p<br />

(−a)<br />

p prime<br />

p≤13<br />

is a constant. It is clear that T > 0 as the equation x 4 = y 4 + z 4 + v 4 + w 4<br />

admits a non-zero solution in/mfor any m > 1.<br />

Itremains to show that there exists a set {p 1 , . . . , p r } of primes p i ≡ 3 (mod 4)<br />

such that p i > 13, r (<br />

1 + 1 − 1 − 1 )<br />

≥ C p i pi<br />

2 pi<br />

3 T , (†)<br />

∏<br />

i=1<br />

<strong>and</strong> p 1 · . . . · p r ≡ 1 (mod M ).<br />

1<br />

Condition (†) is easily satisfied as the series ∑ p≡3 (mod 4) diverges. We find a<br />

p<br />

set {p 1 , . . . , p s } of prime numbers p i > 13 of the form p i ≡ 3 (mod 4) such<br />

− 1 p 3 i<br />

p prime<br />

p≤13<br />

) ≥<br />

C<br />

T . Enlarging {p 1, . . . , p s } makes that product<br />

(<br />

that ∏ s i=1 1 +<br />

1<br />

p i<br />

− 1 pi<br />

2<br />

even bigger. We may therefore assume p 1 · . . . · p s ≡ 3 (mod 4).<br />

The numbers M <strong>and</strong> p 1 · . . . · p s are relatively prime. By Dirichlet’s prime number<br />

theorem, there exists a prime p s+1 , larger than each of the p 1 , . . . , p s , such<br />

that p 1 · . . . · p s+1 ≡ 1 (mod M ). This shows, in particular, p s+1 ≡ 3 (mod 4).<br />

The assertion follows.<br />

Conclusion. The four steps together show that m(X (−a) )·τ (−a) is unbounded.<br />

□<br />

3. The fundamental finiteness property<br />

i. An estimate for the factors at the finite places. —<br />

3.1. Notation. –––– i) For a prime number p <strong>and</strong> an integer x ≠ 0, we<br />

put x (p) := p ν p(x) . Note x (p) = 1/‖x‖ p for the normalized p-adic valuation.<br />

ii) By putting ν(x) := min<br />

ξ∈p<br />

to/p r.<br />

x=(ξ mod p r )<br />

ν(ξ), we carry the p-adic valuation fromp over<br />

Note that any 0 ≠ x ∈/p rhas the form x = ε·p ν(x) where ε ∈ (/p r) ∗ is<br />

a unit. Clearly, ε is unique only in the case ν(x) = 0.


Sec. 3] THE FUNDAMENTAL FINITENESS PROPERTY 271<br />

3.2. Definition. –––– For (a 0 , . . . , a 4 ) ∈5 , r ∈Æ, <strong>and</strong> ν 0 , . . . , ν 4 ≤ r, put<br />

S (r)<br />

ν 0 , ... ,ν 4 ;a 0 , ... ,a 4<br />

:= {(x 0 , . . . ,x 4 ) ∈ (/p r) 5 |<br />

ν(x 0 ) = ν 0 , . . . ,ν(x 4 ) = ν 4 ; a 0 x 4 0 + . . . + a 4 x 4 4 = 0 ∈/p r} .<br />

For the particular case ν 0 = . . . = ν 4 = 0, we will write<br />

I.e.,<br />

Z (r)<br />

a 0 , ... ,a 4<br />

:= S (r)<br />

0, ... ,0;a 0 , ... ,a 4<br />

.<br />

Z (r)<br />

a 0 , ... ,a 4<br />

= {(x 0 , . . . , x 4 ) ∈ [(/p r) ∗ ] 5 | a 0 x 4 0 + . . . + a 4 x 4 4 = 0 ∈/p r} .<br />

We will use the notation z (r)<br />

a 0 , ... ,a 4<br />

:= #Z (r)<br />

a 0 , ... ,a 4<br />

.<br />

3.3. Sublemma. –––– If p k |a 0 , . . . , a 4 <strong>and</strong> r > k then we have<br />

z (r)<br />

a 0 , ... ,a 4<br />

= p 5k · z (r−k)<br />

a 0 /p k , ... ,a 4 /p k .<br />

Proof. Since a 0 x 4 0 + . . . + a 4x 4 4 = pk (a 0 /p k · x 4 0 + . . . + a 4/p k · x 4 4 ), there is a<br />

surjection<br />

ι: Z a (r)<br />

0 , ... ,a 4<br />

−→ Z (r−k) ,<br />

a 0 /p k , ... ,a 4 /p k<br />

given by (x 0 , . . . , x 4 ) ↦→ ((x 0 mod p r−k ), . . . , (x 4 mod p r−k )). The kernel of<br />

the homomorphism of modules underlying ι is (p r−k/pr) 5 . □<br />

3.4. Lemma. –––– Assume gcd p<br />

(a 0 , . . . , a 4 ) = p k . Then, there is an estimate<br />

z (r)<br />

a 0 , ... ,a 4<br />

≤ 8p 4r+k .<br />

Proof. Suppose first that k = 0. This means, one of the coefficients is prime<br />

to p. Without restriction, assume p ∤a 0 .<br />

For any (x 1 , . . . , x 4 ) ∈ (/p r) 4 , there appears an equation of the form<br />

a 0 x 4 0 = c. For p odd, it cannot have more than four solutions in (/p r) ∗<br />

as this group is cyclic. On the other h<strong>and</strong>, in the case p = 2, we have<br />

(/2 r) ∗ ∼ =/2 r−2×/2<strong>and</strong> up to eight solutions are possible.<br />

The general case follows directly from Sublemma 3.3. Indeed, if k < r then<br />

z (r)<br />

a 0 , ... ,a 4<br />

= p 5k · z (r−k)<br />

a 0 /p k , ... ,a 4 /p k ≤ p 5k · 8p 4(r−k) = 8p 4r+k .<br />

On the other h<strong>and</strong>, if k ≥ r then the assertion is completely trivial since<br />

z (r)<br />

a 0 , ... ,a 4<br />

= #Z (r)<br />

a 0 , ... ,a 4<br />

< p 5r ≤ p 4r+k < 8p 4r+k .<br />


272 ON THE SMALLEST POINT ON A DIAGONAL QUARTIC THREEFOLD [Chap. VIII<br />

3.5. Remark. –––– The proof shows that in the case p ≠ 2 the same inequality<br />

is true with coefficient 4 instead of 8. If p ≡ 3 (mod 4) then one could even<br />

reduce the coefficient to 2. Unfortunately, these observations do not lead to a<br />

substantial improvement of our final result.<br />

3.6. Lemma. –––– Let r ∈Æ<strong>and</strong> ν 0 , . . . , ν 4 ≤ r. Then,<br />

#S ν (r)<br />

0 , ... ,ν 4 ;a 0 , ... ,a 4<br />

= z(r) · ϕ(p r−ν p 4ν 0a 0 , ... ,p 4ν 0<br />

4a ) · . . . · ϕ(p<br />

r−ν 4)<br />

4<br />

.<br />

ϕ(p r ) 5<br />

Proof. As p 4ν 0 a0 x 4 0 + . . . + p4ν 4 a4 x 4 4 = a 0(p ν 0 x0 ) 4 + . . . + a 4 (p ν 4 x4 ) 4 , we have<br />

a surjection<br />

π : Z (r) −→ S (r)<br />

p 4ν 0a 0 , ... ,p 4ν 4a 4<br />

ν 0 , ... ,ν 4 ;a 0 , ... ,a 4<br />

,<br />

given by (x 0 , . . . , x 4 ) ↦→ (p ν 0 x0 , . . . , p ν 4 x4 ).<br />

For i = 0, . . . , 4, consider the mapping ι:/p r→/pr, x 0 ↦→ p ν i<br />

x 0 .<br />

If ν i = r then ι is the zero map. All ϕ(p r ) = (p − 1)p r−1 units are mapped<br />

to zero. Otherwise, observe that ι is p ν i<br />

: 1 on its image. Further, ν(ι(x)) = ν i<br />

if <strong>and</strong> only if x is a unit. By consequence, π is (K (ν 0)<br />

· . . . · K (ν4) ) : 1 when<br />

we put K (ν) := p ν for ν ≠ r <strong>and</strong> K (r) := (p − 1)p r−1 . Summarizing, we could<br />

have written K (ν) := ϕ(p r )/ϕ(p r−ν ). The assertion follows.<br />

□<br />

3.7. Corollary. –––– Let (a 0 , . . . , a 4 ) ∈ (\{0}) 5 . Then, for the local factor<br />

:= τ p (X (a 0, ... ,a 4 ) ), one has<br />

τ (a 0, ... ,a 4 )<br />

p<br />

τ (a 0, ... ,a 4 )<br />

p<br />

r<br />

r→∞<br />

∑<br />

ν 0 , ... ,ν 4 =0<br />

= lim<br />

Proof. Remark VI.5.19 shows that<br />

τ (a 0, ... ,a 4 )<br />

p<br />

Lemma 3.6 yields the assertion.<br />

z (r) · ϕ(p r−ν 0<br />

) · . . . · ϕ(p r−ν 4<br />

)<br />

p 4ν 0a 0 , ... ,p 4ν 4a 4<br />

.<br />

p 4r · ϕ(p r ) 5<br />

r<br />

#S ν<br />

r→∞<br />

∑<br />

(r)<br />

0 , ... ,ν 4 ;a 0 , ... ,a 4<br />

ν 0 , ... ,ν 4 =0<br />

= lim<br />

p 4r .<br />

3.8. Proposition –––– Let (a 0 , . . . , a 4 ) ∈ (\{0}) 5 . Then, for each ε such<br />

that 0 < ε < 1 , one has<br />

4<br />

(<br />

τ (a 0, ... ,a 4 ) 1<br />

)( 1<br />

) 4 (<br />

p ≤ 8<br />

· a<br />

(p)<br />

1 − 1 1 − 1 0<br />

· . . . · a (p) ) 1−ε (<br />

4<br />

3<br />

a (p) ) ε.<br />

4<br />

p 1−4ε p ε<br />

Proof. By Lemma 3.4,<br />

z (r)<br />

p 4ν 0a 0 , ... ,p 4ν 4a 4<br />

/p 4r ≤ 8 gcd p<br />

(p 4ν 0<br />

a 0 , . . . , p 4ν 4<br />

a 4 )<br />

= 8 gcd ( p 4ν 0<br />

a (p)<br />

0 , . . . , p4ν 4<br />

a (p)<br />

4<br />

)<br />

.<br />


Sec. 3] THE FUNDAMENTAL FINITENESS PROPERTY 273<br />

(<br />

(p)) Writing k i := ν p (a i ) = ν p a i , we see<br />

z (r)<br />

p 4ν 0a 0 , ... ,p 4ν 4a 4<br />

/p 4r ≤ 8 gcd(p 4ν 0+k 0<br />

, . . . , p 4ν 4+k 4<br />

)<br />

= 8p min{4ν 0+k 0 , ... ,4ν 4 +k 4 } .<br />

We estimate the minimum by a weighted arithmetic mean with weights 1−ε<br />

, <strong>and</strong> ε,<br />

1−ε<br />

4 , 1−ε<br />

, 1−ε<br />

4 4<br />

min{3ν 0 + k 0 , . . . , 3ν 3 + k 3 }<br />

This shows<br />

≤ 1 − ε<br />

3<br />

· (3ν 0 + k 0 ) + 1 − ε · (3ν 1 + k 1 )<br />

3<br />

+ 1 − ε · (3ν 2 + k 2 ) + ε(3ν 3 + k 3 )<br />

3<br />

= (1 − ε)(ν 0 + ν 1 + ν 2 ) + 3εν 3<br />

+ 1 − ε (k 0 + k 1 + k 2 ) + εk 3 .<br />

3<br />

z (r) /p 4r ≤ 8p (1−ε)(ν 0+ ... +ν 3 )+4εν 4 + 1−ε<br />

p 4ν 0a 0 , ... ,p 4ν 4 (k 0+ ... +k 3 )+εk 4<br />

4a 4<br />

(a<br />

= 8p (1−ε)(ν 0+ ... +ν 3 )+4εν4 (p)<br />

·<br />

0<br />

· . . . · a (p) ) 1−ε (<br />

4<br />

3<br />

a (p) ) ε.<br />

4<br />

We may therefore write<br />

τ (a 0, ... ,a 4 )<br />

p<br />

≤ 8 ( a (p)<br />

0<br />

· . . . · a (p) ) 1−ε (<br />

4<br />

3<br />

a (p)<br />

4<br />

r<br />

r→∞<br />

∑<br />

ν 0 , ... ,ν 4 =0<br />

· lim<br />

) ε<br />

p (1−ε)(ν 0+ ... +ν 3 )+4εν 4 · ϕ(p<br />

r−ν 0) · . . . · ϕ(p<br />

r−ν 4)<br />

.<br />

ϕ(p r ) 5<br />

Here, the term under the limit is precisely the product of four copies of the<br />

finite sum r<br />

p<br />

∑<br />

(1−ε)ν · ϕ(p r−ν r−1<br />

) 1<br />

=<br />

ν=0<br />

ϕ(p r )<br />

∑<br />

ν=0<br />

(p ε ) + p 1<br />

ν p − 1 (p ε ) r<br />

<strong>and</strong> one copy of the finite sum<br />

r<br />

p<br />

∑<br />

4εν · ϕ(p r−ν r−1<br />

)<br />

=<br />

ν=0<br />

ϕ(p r )<br />

∑<br />

ν=0<br />

1<br />

(p 1−4ε ) + p<br />

ν<br />

1<br />

p − 1 (p 1−4ε ) . r<br />

For r → ∞, geometric series do appear while the additional summ<strong>and</strong>s tend<br />

to zero.<br />

□<br />

4 ,<br />

3.9. Remark. –––– Unfortunately, the constants<br />

(<br />

C p (ε) 1<br />

)( 1<br />

) 4<br />

:= 8<br />

1 − 1 1 − 1 p 1−4ε p ε


274 ON THE SMALLEST POINT ON A DIAGONAL QUARTIC THREEFOLD [Chap. VIII<br />

have the property that the product ∏ p C p<br />

(ε) diverges. On the other h<strong>and</strong>, we<br />

have at least that C p<br />

(ε) is bounded for p → ∞, say C p (ε) ≤ C (ε) .<br />

3.10. Lemma. –––– Let C > 1 be any constant. Then, for each ε > 0, one has<br />

∏<br />

p prime<br />

p|x<br />

for a suitable constant c (depending on ε).<br />

C ≤ c · x ε<br />

Proof. This follows directly from [Nat, Theorem 7.2] together with [Nat,<br />

Section 7.1, Exercise 7].<br />

□<br />

3.11. Proposition –––– For each ε > 0, there exists a constant c such that<br />

∏ τ (a 0, ... ,a 4 )<br />

p ≤ c · | a 0 · . . . · a 4 | 4 1 · ∏<br />

p prime<br />

for all (a 0 , . . . , a 4 ) ∈ (\{0}) 5 .<br />

p prime<br />

min ‖a i‖ 1 4 −ε<br />

p<br />

i=0, ... ,4<br />

Proof. As already noticed in the third step of the proof of Theorem 2.2, the<br />

product over all primes of good reduction is bounded by consequence of the<br />

Weil conjectures. It, therefore, remains to show that<br />

∏ τ (a 0, ... ,a 4 )<br />

p ≤ c · |a 0 · . . . · a 4 | 4 1 · ∏<br />

p prime<br />

p|2a 0 ... a 4<br />

p prime<br />

For this, we may assume that ε is small, say ε < 1 4 .<br />

Then, by Proposition 3.8, we have at first<br />

min ‖a i‖ 1 4 −ε<br />

i=0, ... ,4<br />

p .<br />

τ (a 0, ... ,a 4 )<br />

p<br />

≤ C (ε)<br />

p<br />

= C (ε)<br />

p<br />

·<br />

·<br />

(a(p)<br />

0<br />

· . . . · a (p) ) 1<br />

4 − ε5<br />

3<br />

· (a (p)<br />

4 ) 4 5 ε<br />

(a(p)<br />

0<br />

· . . . · a (p) ) 1<br />

3 a(p) 4<br />

− ε5<br />

4<br />

· (a (p)<br />

4 )− 1 4 +ε .<br />

Here, the indices 0, . . . , 4 are interchangeable. Hence, it is even allowed to<br />

write<br />

τ (a 0, ... ,a 4 )<br />

p<br />

≤ C (ε)<br />

p<br />

= C (ε)<br />

p<br />

·<br />

·<br />

(a(p)<br />

0<br />

· . . . · a (p)<br />

4<br />

(a(p)<br />

0<br />

· . . . · a (p)<br />

4<br />

) 1<br />

4<br />

− ε5 · (max<br />

) 1<br />

4<br />

− ε5 · min<br />

i<br />

i<br />

)<br />

a (p) −<br />

1<br />

4<br />

+ε<br />

i<br />

‖a i ‖ 1 4 −ε<br />

p .<br />

Now, we multiply over all prime divisors of 2a 0 · . . . · a 4 . Thereby, on the<br />

right h<strong>and</strong> side, we may twice write the product over all primes since the two


Sec. 3] THE FUNDAMENTAL FINITENESS PROPERTY 275<br />

rightmost factors are equal to one for p ∤a 0 · . . . · a 4 , anyway.<br />

∏ τ (a 0, ... ,a 4 )<br />

p<br />

p prime<br />

p|2a 0 ... a 4<br />

≤ ∏ C p<br />

(ε)<br />

p prime<br />

p|2a 0 ... a 4<br />

(<br />

(p)<br />

· ∏ a<br />

0<br />

· . . . · a (p) ) 1<br />

4<br />

− ε 5<br />

4<br />

· ∏<br />

p prime<br />

= ∏ C p (ε) · | a 0 · . . . · a 4 | 4 1 − ε 5 · ∏<br />

p prime<br />

p|2a 0 ... a 4<br />

p prime<br />

p prime<br />

min ‖a 4<br />

i‖ 1 −ε<br />

p<br />

i=0, ... ,4<br />

min ‖a i‖ 1 4 −ε<br />

p<br />

i=0, ... ,4<br />

when we observe that ∏ p a (p) = | a|. Further, we have C p<br />

(ε)<br />

Lemma 3.10,<br />

∏ C (ε) ≤ c · |2a 0 · . . . · a 4 | ε 5 .<br />

p prime<br />

p|2a 0 ... a 4<br />

We finally estimate 2 ε 5 by a constant. The assertion follows.<br />

≤ C (ε) <strong>and</strong>, by<br />

ii. A bound for the factor at the infinite place. — We want to estimate the integral<br />

τ (a 0, ... ,a 4 )<br />

∞ := τ ∞ (X (a 0, ... ,a 4 ) ) described in 1.5.<br />

3.12. Lemma. –––– There exist two constants C 1 <strong>and</strong> C 2 such that<br />

τ (a 0, ... ,a 4 )<br />

∞<br />

⎧<br />

1<br />

(<br />

⎨ 4√ C1 min(|a 0 |, 2a 4 ) 1/4) , if |a 0 | ≤ a 4 ,<br />

|a0 |a<br />

≤<br />

1·...·a 4<br />

(<br />

1<br />

⎩ 4√ C1 min(|a 0 |, 2a 4 ) 1/4 + C 2 a 1/4<br />

|a0 |a<br />

4<br />

log min(|a 0|,3a 1 )) a 1·...·a 4 4<br />

, otherwise,<br />

for all (a 0 , . . . , a 4 ) ∈Ê5 satisfying a 1 ≥ a 2 ≥ a 3 ≥ a 4 ≥ 1 <strong>and</strong> a 0 ≤ −1.<br />

Proof. A linear substitution shows<br />

τ (a 0, ... ,a 4 )<br />

∞ = 1 4 ·<br />

where<br />

ZZZZ<br />

1<br />

√<br />

| a0 |a 1 · . . . · a 4<br />

4<br />

1<br />

dy dz dv dw<br />

(y 4 + z 4 + v 4 + w 4 )<br />

3/4<br />

R (0)<br />

R (0) := {(y, z, v, w) ∈ [−a 1/4<br />

1<br />

, a 1/4<br />

1<br />

] × · · · × [−a 1/4<br />

4<br />

, a 1/4<br />

4<br />

]<br />

| |y 4 + z 4 + v 4 + w 4 | ≤ |a 0 |} .<br />

The integr<strong>and</strong> is non-negative. We cover R (0) ⊆ R 1 ∪ R 2 by two sets as follows,<br />

R 1 := {(y, z, v, w) ∈Ê4 | y 4 + z 4 + v 4 + w 4 ≤ min(| a 0 |, 2a 4 )} ,<br />

R 2 := {(y, z, v, w) ∈Ê4 | a 4 ≤ y 4 + z 4 + v 4 ≤ min(| a 0 |, 3a 1 ) <strong>and</strong><br />

□<br />

w ∈ [−a 1/4<br />

4<br />

, a 1/4<br />

4<br />

]} ,<br />

<strong>and</strong> estimate. In the case | a 0 | ≤ a 4 , the domain of integration is covered by R 1<br />

alone <strong>and</strong> we may omit R 2 , completely.


276 ON THE SMALLEST POINT ON A DIAGONAL QUARTIC THREEFOLD [Chap. VIII<br />

By homogeneity, we have<br />

ZZZZ<br />

1<br />

(y 4 + z 4 + v 4 + w 4 ) dy dz dv dw = ω 3/4 1 ·<br />

R 1<br />

min(| a 0 |,2a 4 ) 1/4<br />

Z<br />

0<br />

1<br />

r 3 · r3 dr<br />

= ω 1 · min(| a 0 |, 2a 4 ) 1/4<br />

where ω 1 is the three-dimensional hypersurface measure of the l 4 -unit hypersphere<br />

S 1 := {(y, z, v, w) ∈Ê4 | y 4 + z 4 + v 4 + w 4 = 1} .<br />

Further,<br />

ZZZZ<br />

1<br />

dy dz dv dw<br />

(y 4 + z 4 + v 4 + w 4 )<br />

3/4<br />

R 2<br />

ZZZ<br />

≤ 2a 1/4 1<br />

4<br />

dy dz dv<br />

(y 4 + z 4 + v 4 )<br />

3/4<br />

R 3<br />

where<br />

R 3 := {(y, z, v) ∈Ê3 | a 4 ≤ y 4 + z 4 + v 4 ≤ min(| a 0 |, 3a 1 )} .<br />

The latter integral may be treated in much the same way as the one above.<br />

We see<br />

ZZZ<br />

1<br />

(y 4 + z 4 + v 4 ) dy dz dv = ω 3/4 2 ·<br />

R 3<br />

min(|a 0 |,3a 1 ) 1/4<br />

Z<br />

a 1/4<br />

4<br />

1<br />

r 3 · r2 dr<br />

where ω 2 is the usual two-dimensional hypersurface measure of the l 4 -unit sphere<br />

Finally,<br />

min(| a 0 |,3a 1 ) 1/4<br />

Z<br />

a 1/4<br />

4<br />

S 2 := {(y, z, v) ∈Ê3 | y 4 + z 4 + v 4 = 1} .<br />

1<br />

r 3 · r2 dr = log min(| a 0|, 3a 1 ) 1/4<br />

a 1/4<br />

4<br />

= 1 4 log min(| a 0|, 3a 1 )<br />

a 4<br />

. □<br />

3.13. Proposition –––– For every ε > 0, there exists a constant C such that<br />

τ (a 0, ... ,a 4 )<br />

∞<br />

≤ C · | a 0 · . . . · a 4 | − 1 4 +ε · min<br />

i=0, ... ,4 ‖a i‖ 1 4<br />

∞<br />

for each (a 0 , . . . , a 4 ) ∈5 satisfying a 0 < 0 <strong>and</strong> a 1 , . . . , a 4 > 0.<br />

Proof. We assume without restriction that a 1 ≥ . . . ≥ a 4 . There are two cases<br />

to be distinguished.


Sec. 3] THE FUNDAMENTAL FINITENESS PROPERTY 277<br />

First case. | a 0 | ≤ a 4 .<br />

Then, by Lemma 3.12, we have<br />

τ (a 0, ... ,a 4 )<br />

∞ ≤ | a 0 · . . . · a 4 | − 1 4 · C1 min{| a 0 |, 2a 4 } 1 4<br />

= C 1 · | a 0 · . . . · a 4 | − 1 4 · | a0 | 1 4<br />

= C 1 · | a 0 · . . . · a 4 | − 1 4 · min<br />

i=0, ... ,4 ‖a i‖ 1 4<br />

∞ .<br />

Second case. | a 0 | > a 4 .<br />

Here, Lemma 3.12 shows<br />

τ (a 0, ... ,a 4 )<br />

∞ ≤ | a 0 · . . . · a 4 | − 1 4<br />

≤ | a 0 · . . . · a 4 | − 1 4<br />

= | a 0 · . . . · a 4 | − 1 4 · | a4 | 1 4<br />

(<br />

C 1 min{|a 0 |, 2a 4 } 1 4 + C2 a 1/4<br />

(<br />

C 1 (2a 4 ) 1 4 + C2 a 1/4<br />

4<br />

log | a 0|<br />

(<br />

C 1 2 1 4 + C2 log | a 0|<br />

a 4<br />

)<br />

4<br />

log min{|a 0|, 3a 1 }<br />

)<br />

a 4<br />

)<br />

a 4<br />

(<br />

= | a 0 · . . . · a 4 | − 4 1 · min ‖a i‖ 1 4<br />

∞ · C 1 2 4 1 + C2 log | a )<br />

0|<br />

i=0, ... ,4 a 4<br />

≤ | a 0 · . . . · a 4 | − 1 4 · min<br />

i=0, ... ,4 ‖a i‖ 1 4<br />

∞ · (C<br />

1 2 1 4 + C2 log | a 0 · . . . · a 4 | )<br />

≤ | a 0 · . . . · a 4 | − 1 4 · min<br />

i=0, ... ,4 ‖a i‖ 1 4<br />

∞ · (C<br />

3 · | a 0 · . . . · a 4 | ) ε<br />

.<br />

□<br />

iii. The Tamagawa number. —<br />

3.14. Proposition –––– For every ε > 0, there exists a constant C > 0 such that<br />

) 1<br />

1 4<br />

a 4<br />

1<br />

τ ≥ C · H( 1<br />

a 0<br />

: . . . :<br />

(a 0, ... ,a 4 )<br />

| a 0 · . . . · a 4 | ε<br />

for each (a 0 , . . . , a 4 ) ∈5 satisfying a 0 < 0 <strong>and</strong> a 1 , . . . , a 4 > 0.<br />

Proof. By Proposition 3.13, we have<br />

τ (a 0, ... ,a 4 )<br />

∞<br />

On the other h<strong>and</strong>, by Theorem 3.11,<br />

≤ C 1 · | a 0 · . . . · a 4 | − 1 4 + ε 2 · min<br />

i=0, ... ,4 ‖a i‖ 1 4<br />

∞ .<br />

∏ τ (a 0, ... ,a 4 )<br />

p ≤ C 2 · | a 0 · . . . · a 4 | 4 1 · ∏<br />

p prime<br />

p prime<br />

min ‖a i‖ 1 4 − ε 2<br />

i=0, ... ,4<br />

p .


278 ON THE SMALLEST POINT ON A DIAGONAL QUARTIC THREEFOLD [Chap. VIII<br />

It follows that<br />

τ (a 0, ... ,a 4 )<br />

≤ C 3·|a 0 · . . . · a 4 | ε 2 · ∏<br />

<strong>and</strong><br />

p prime<br />

1<br />

τ (a 0, ... ,a 4 )<br />

≥ 1 C 3<br />

·<br />

= 1 C 3<br />

·<br />

= 1 C 3<br />

·<br />

[ ]<br />

min ‖a 4<br />

i‖ 1 p · min ‖a i‖ 1 −<br />

ε<br />

4<br />

∞ · ∏ min ‖a 2<br />

i‖ p<br />

i=0, ... ,4 i=0, ... ,4 i=0, ... ,4<br />

[<br />

∏<br />

p prime<br />

] −<br />

1<br />

min ‖a 4<br />

i‖ p ·<br />

i=0, ... ,4<br />

| a 0 · . . . · a 4 | ε 2 · ∏<br />

p prime<br />

∥ 1 ∏ max ∥ 1 ∥∥ 4<br />

i=0, ... ,4<br />

a i<br />

· max<br />

p<br />

p prime<br />

| a 0 · . . . · a 4 | ε 2 · ∏<br />

p prime<br />

H ( 1<br />

a 0<br />

: . . . :<br />

[<br />

| a 0 · . . . · a 4 | ε 2 · ∏<br />

p prime<br />

p prime<br />

[ ] −<br />

1<br />

min ‖a 4<br />

i‖ ∞<br />

i=0, ... ,4<br />

[<br />

min<br />

‖a ] −<br />

ε<br />

i‖ 2 p<br />

i=0, ... ,4 ∥ 1 ∥ 1 ∥∥ 4<br />

a i<br />

∞<br />

i=0, ... ,4<br />

[<br />

max<br />

i=0, ... ,4 a(p) i<br />

) 1<br />

1 4<br />

a 4<br />

It is obvious that max<br />

i=0, ... ,4 a(p) i ≤ | a (p)<br />

0<br />

· . . . · a (p)<br />

4 | <strong>and</strong><br />

This shows<br />

max<br />

i=0, ... ,4 a(p) i<br />

∏ | a (p)<br />

0<br />

· . . . · a (p)<br />

4 | = | a 0 · . . . · a 4 |.<br />

p prime<br />

] ε<br />

2<br />

] ε<br />

2<br />

.<br />

1<br />

τ ≥ 1 H ( ) 1<br />

1 1<br />

a<br />

·<br />

0<br />

: . . . : 4<br />

a 4 (a 0, ... ,a 4 )<br />

C 3 | a 0 · . . . · a 4 | ε 2 · | a 0 · . . . · a 4 | ε 2<br />

= 1 · H( ) 1<br />

1 1<br />

a 0<br />

: . . . : 4<br />

a 4<br />

. □<br />

C 3 | a 0 · . . . · a 4 | ε<br />

3.15. Lemma. –––– Let (a 0 : . . . : a 4 ) ∈ P 4 (É) be any point such that<br />

a 0 ≠ 0, . . . , a 4<br />

∈<br />

≠ 0. Then,<br />

H(a 0 : . . . : a 4 ) ≤ H( 1 1<br />

a 0<br />

: . . . :<br />

a 4<br />

) 4 .<br />

Proof. First, observe that (a 0 : . . . : a 4 ) ↦→ ( )<br />

1<br />

1<br />

a 0<br />

: . . . :<br />

a 4<br />

is a welldefined<br />

map. Hence, we may assume without restriction that a 0 , . . . , a 4<br />

<strong>and</strong> gcd(a 0 , . . . , a 4 ) = 1. This yields H(a 0 : . . . : a 4 ) = max | a i|.<br />

i=0, ... ,4<br />

On the other h<strong>and</strong>, ( 1 1<br />

a 0<br />

: . . . :<br />

a 4<br />

) = (a 1 a 2 a 3 a 4 : . . . : a 0 a 1 a 2 a 3 ). Consequently,<br />

H ( 1<br />

a 0<br />

: . . . :<br />

1<br />

a 4<br />

) ≤ [ max<br />

i=0,...,4 | a i|] 4 = H(a 0 : . . . : a 4 ) 4 .


Sec. 3] THE FUNDAMENTAL FINITENESS PROPERTY 279<br />

From this, the asserted inequality emerges when the roles of a i <strong>and</strong> 1 a i<br />

are interchanged.<br />

□<br />

3.16. Corollary. –––– Let a 0 , . . . , a 4 ∈such that gcd(a 0 , . . . , a 4 ) = 1. Then,<br />

| a 0 · . . . · a 4 | ≤ H ( 1<br />

a 0<br />

: . . . :<br />

1<br />

a 4<br />

) 20.<br />

Proof. Observe that | a 0 · . . . · a 4 | ≤ max | a i| 5 = H(a 0 : . . . : a 4 ) 5 <strong>and</strong> apply<br />

i=0, ... ,4<br />

Lemma 3.15.<br />

□<br />

3.17. Theorem. –––– For each ε > 0, there exists a constant C (ε) > 0 such that,<br />

for all (a 0 , . . . , a 4 ) ∈5 satisfying a 0 < 0 <strong>and</strong> a 1 , . . . , a 4 > 0,<br />

1<br />

τ ≥ C (ε) · H( ) 1<br />

1 1<br />

(a 0, ... ,a 4 ) a 0<br />

: . . . : 4<br />

−ε<br />

a 4<br />

.<br />

Proof. We may assume that gcd(a 0 , . . . , a 4 ) = 1. Then, by Proposition 3.14,<br />

1<br />

τ ≥ C · H( ) 1<br />

1 1<br />

a 0<br />

: . . . : 4<br />

a 4<br />

.<br />

(a 0, ... ,a 4 )<br />

|a 0 · . . . · a 4 | ε 20<br />

(<br />

Corollary 3.16 yields | a 0 · . . . · a 4 | ε 20 ≤ H 1<br />

)<br />

1 ε.<br />

a 0<br />

: . . . :<br />

a 4<br />

3.18. Corollary (Fundamental finiteness). —– For each B > 0, there are only<br />

finitely many quartics X (a 0, ... ,a 4 ) : a 0 x 4 0 + . . . + a 4 x 4 4 = 0 in P 4Ésuch that a 0 < 0,<br />

a 1 , . . . , a 4 > 0, <strong>and</strong> τ (a 0, ... ,a 4) > B.<br />

Proof. This is an immediate consequence of the comparison to the naive height<br />

established in Theorem 3.17.<br />

□<br />


CHAPTER IX<br />

GENERAL CUBIC SURFACES ∗<br />

Anyone who considers arithmetical methods of producing r<strong>and</strong>om digits is,<br />

of course, in a state of sin.<br />

JOHN VON NEUMANN (1951)<br />

1. Rational <strong>points</strong> on cubic surfaces<br />

1.1. –––– The arithmetic of cubic surfaces is a fascinating subject. To a large<br />

extent, it was initiated by the work of Yu. I. Manin, particularly by his fundamental<br />

<strong>and</strong> influential book on Cubic Forms [Man].<br />

In this chapter, we study the distribution of <strong>rational</strong> <strong>points</strong> on general cubic<br />

surfaces overÉ. The main problems are<br />

i) Existence ofÉ-<strong>rational</strong> <strong>points</strong>,<br />

ii) Asymptotics ofÉ-<strong>rational</strong> <strong>points</strong>,<br />

iii) The height of the smallest point.<br />

i. Existence of <strong>rational</strong> <strong>points</strong>. —<br />

1.2. –––– Let X be an algebraic variety defined overÉ. Recall that the<br />

Hasse principle is said to hold for X if<br />

X (É) = ∅ ⇐⇒ ∃ ν ∈ Val(É): X (Éν) = ∅ .<br />

For quadrics in P nÉ, the Hasse principle holds by the famous Theorem of Hasse-<br />

Minkowski.<br />

It is, however, well-known that for smooth cubic surfaces overÉthe Hasse principle<br />

does not hold, in general. In all known examples, this is explained by the<br />

Brauer-Manin obstruction which was discussed in detail in Chapter III.<br />

(∗) This chapter is a revised version of the article: Experiments with general cubic surfaces, to<br />

appear in: The Manin Festschrift, joint with A.-S. Elsenhans.


282 GENERAL CUBIC SURFACES [Chap. IX<br />

ii. Asymptotics of <strong>rational</strong> <strong>points</strong>. —<br />

1.3. –––– On the asymptotics of <strong>rational</strong> <strong>points</strong> of bounded height, there is<br />

Manin’s conjecture which was described in Chapter VI.<br />

1.4. Conjecture (Manin). —– Let X be an arbitrary Fano variety overÉ<strong>and</strong> H<br />

be an anticanonical height on X. Then, there exist a dense, Zariski open subset<br />

X ◦ ⊆ X <strong>and</strong> a constants r <strong>and</strong> τ such that<br />

#{x ∈ X ◦ (É) | H(x) < B} ∼ τB log r B<br />

(∗)<br />

for B → ∞.<br />

In the case that X is a surface, it is expected that r = rkPic(X ) − 1.<br />

1.5. Remark (Known cases). —– Conjecture 1.4 is established for smooth<br />

complete intersections of multidegree d 1 , . . . , d n in the case that the dimension<br />

of X is very large compared to d 1 , . . . , d n [Bir]. Further, it is proven for projective<br />

spaces <strong>and</strong> quadrics. Finally, there are a number of further special cases<br />

in which Manin’s conjecture is known to be true. See, e.g., [Pe02, Sec. 4].<br />

In Chapter VII, numerical evidence for Conjecture 1.4 is presented in the case<br />

of the threefolds X e a,b given by axe = by e + z e + v e + w e in P 4Éfor e = 3 <strong>and</strong> 4.<br />

1.6. Remark (Peyre’s constant). —– Motivated by results obtained by the classical<br />

circle method, E. Peyre refined Manin’s conjecture by a conjectural value for<br />

the leading coefficient τ. We formulated a general description of Peyre’s constant<br />

in Definition VI.5.35.<br />

1.7. –––– We are interested in smooth cubic surfaces. Let us explain<br />

Peyre’s constant when X is the smooth surface in P 3Édefined by a homogeneous<br />

cubic polynomial f ∈[X 0 , X 1 , X 2 , X 3 ].<br />

Consider the anticanonical height which is the same as the naive height. Assume<br />

that rkPic(X ) = 1 <strong>and</strong> H 1( Gal(É/É), Pic(XÉ) ) = 0.<br />

Then, Peyre’s constant is equal to the Tamagawa type number τ given by<br />

where<br />

τ p =<br />

τ := ∏ τ p<br />

p∈È∪{∞}<br />

(<br />

1 − 1 )<br />

#X (/p m)<br />

· lim<br />

p m→∞ p 2m


Sec. 1] RATIONAL POINTS ON CUBIC SURFACES 283<br />

for p finite <strong>and</strong><br />

τ ∞ = 1 2<br />

Z<br />

x∈[−1,1] 4<br />

f (x)=0<br />

1<br />

dS .<br />

‖(grad f )(x)‖ 2<br />

Here, X ⊂ P 3is the integral model of X defined by the polynomial f .<br />

dS denotes the usual hypersurface measure on the cone CX (Ê), considered as a<br />

hypersurface inÊ4 .<br />

1.8. Remarks. –––– i) The assumption that rk Pic(X ) = 1 guarantees that there<br />

are no log-factors in the conjectural asymptotics.<br />

ii) According to Example VI.5.3, the assumption rkPic(X ) = 1 is sufficient<br />

for α(X ) = 1. On the other h<strong>and</strong>, H 1( Gal(É/É), Pic(XÉ) ) = 0 guarantees<br />

that there is no Brauer-Manin obstruction to weak approximation. Further,<br />

β(X ) = 1 <strong>and</strong> both additional factors do not need to be considered.<br />

In general, the definition of Peyre’s constant is more complicated, even for<br />

smooth cubic surfaces.<br />

1.9. Remark. –––– According to Definition VI.5.35, the local factors τ p<br />

are equipped with an additional factor det ( 1 − p −1 Frob p | p) I .<br />

Here, Pic(XÉ) I p<br />

denotes the fixed module under the inertia group. For compensation,<br />

another factor is added to Peyre’s constant, the value L(1, χ P ) of the<br />

Artin-L-function associated to the Gal(É/É)-representation P ⊂ Pic(XÉ)⊗<br />

complementary to the trivial summ<strong>and</strong>.<br />

Inour case, therepresentationP turnsouttobeirreducibleofranksix. Thefinite<br />

quotient of Gal(É/É) acting on P is isomorphic to W (E 6 ) of order 51 840.<br />

We do not even make an attempt to compute the exact value of L. Instead, we<br />

work with the conditionally convergent series described.<br />

iii. The smallest point. —<br />

1.10. –––– It would be desirable to have an a-priori upper bound for the height<br />

of the smallestÉ-<strong>rational</strong> point on X as this would allow to effectively decide<br />

whether X (É) ≠ ∅ or not.<br />

When X is a conic, Legendre’s theorem on zeroes of ternary quadratic forms<br />

yields an effective bound for the smallest point. For quadrics of arbitrary dimension,<br />

the same is true by an observation due to J. W. S. Cassels [Cas55].<br />

Further, there is a theorem of C. L. Siegel [Si73, Satz 1] which provides a generalization<br />

to hypersurfaces defined by norm equations. This certainly includes<br />

some special cubic surfaces but, in general, no theoretical upper bound is known<br />

for the height of the smallestÉ-<strong>rational</strong> point on a cubic surface.


284 GENERAL CUBIC SURFACES [Chap. IX<br />

1.11. Remark. –––– If one had an error term [S-D05] for (∗) uniform over all<br />

cubic surfaces X of Picard rank 1 then this would imply that the height m(X )<br />

of the smallestÉ-<strong>rational</strong> point is always less than<br />

C<br />

τ<br />

for certain constants<br />

(X ) α<br />

α > 1 <strong>and</strong> C > 0.<br />

The investigations on quartic threefolds made in Chapter VIII (cf. [EJ5]) indicate<br />

that one might have even m(X ) < C (ε) for any ε > 0. The same inequality<br />

τ (X ) 1+ε<br />

is suggested by the experiments with diagonal cubic surfaces which we will<br />

describe in the next chapter.<br />

Assuming equidistribution, one would expect that the height of the smallestÉ-<strong>rational</strong><br />

point on X should be even ∼ 1<br />

τ . An inequality of the<br />

(X )<br />

form m(X ) < C<br />

τ<br />

is, however, known to be wrong for diagonal quartic threefolds.<br />

(Cf. Theorem VIII.2.2.) Under the Generalized Riemann Hypothesis,<br />

(X )<br />

Theorem X.5.3 will show that m(X ) < C<br />

τ<br />

is wrong for diagonal cubic surfaces,<br />

(X )<br />

too.<br />

iv. The results. —<br />

1.12. –––– We consider two families of cubic surfaces which are produced by<br />

a r<strong>and</strong>om number generator. For each of these surfaces, we do the following.<br />

i) We verify that the Galois group acting on the 27 lines is equal to W (E 6 ).<br />

This implies that rk Pic(X ) = 1 <strong>and</strong> H 1( Gal(É/É), Pic(XÉ) ) = 0.<br />

ii) We compute E. Peyre’s constant τ (X ).<br />

iii) Up to a certain bound for the anticanonical height, we count allÉ-<strong>rational</strong><br />

<strong>points</strong> on the surface X.<br />

Thereby, we establish the Hasse principle for each of the surfaces considered.<br />

Further, we test numerically the conjecture of Manin, in the refined form due to<br />

E. Peyre, on the asymptotics of <strong>points</strong> of bounded height. Finally, we study the<br />

behaviour of the height of the smallestÉ-<strong>rational</strong> point versus E. Peyre’s constant.<br />

This means, we test the estimates formulated in Remark 1.11.<br />

2. Background<br />

i. The 27 lines. —<br />

2.1. –––– A non-singular cubic surface defined overÉcontains exactly 27 lines.<br />

The symmetries of the configuration of the 27 lines respecting the intersection<br />

pairing are given by the Weyl group W (E 6 ). The order of the group #W (E 6 )<br />

is 51 840.


Sec. 2] BACKGROUND 285<br />

These are st<strong>and</strong>ard facts which may be found in many places in the literature.<br />

We discussed them before in Chapter II, 8.21 ff. Observe the list of the 350 conjugacy<br />

classes of W (E 6 ) generated using GAP which is given in the appendix.<br />

2.2. Remark. –––– The group W (E 6 ) admits the following properties.<br />

i) W (E 6 ) has 3 4·5 = 405 2-Sylow sub<strong>groups</strong>.<br />

ii) W (E 6 ) has 2 5·5 = 160 3-Sylow sub<strong>groups</strong>.<br />

iii) W (E 6 ) has 2 4·3 4 = 1296 5-Sylow sub<strong>groups</strong>.<br />

2.3. Fact. –––– LetX beasmoothcubic surface definedoverÉ<strong>and</strong> let K bethe field<br />

of definition of the 27 lines on X. Then K is a Galois extension ofÉ. The Galois<br />

group Gal(K/É) is a subgroup of W (E 6 ).<br />

Proof. This was shown in Chapter II, 8.21.<br />

□<br />

2.4. Remark. –––– W (E 6 ) contains a subgroup U of index two which is<br />

isomorphic to the simple group of order 25 920. It is of Lie type B 2 (3),<br />

i.e. U ∼ = Ω 5 (3) ⊂ SO 5 (3). In U, the stabilizer of two non-intersecting lines<br />

is the alternating group A 5 .<br />

2.5. Remark. –––– On the other h<strong>and</strong>, the operation of W (E 6 ) on the 27 lines<br />

gives rise to a transitive permutation representation ι: W (E 6 ) → S 27 . It turns<br />

out that the image of ι is contained in the alternating group A 27 . We will call<br />

an element σ ∈ W (E 6 ) even if σ ∈ U <strong>and</strong> odd, otherwise. This should not be<br />

confused with the sign of ι(σ) ∈ S 27 which is always even.<br />

2.6. Remark. –––– The operation of the group W (E 6 ) on the 27 lines satisfies<br />

the properties below.<br />

i) It is transitive.<br />

ii) Fix a line l 1 . Then, the remaining 26 lines split into two classes. There are<br />

10 lines which intersect l 1 <strong>and</strong> 16 lines which do not. The stabilizer of l 1 acts<br />

transitively on each class.<br />

iii) Choose a second line l 2 which does not intersect l 1 . Then, there are precisely<br />

five lines g 1 , . . . , g 5 intersecting both l 1 <strong>and</strong> l 2 . The stabilizer of l 1<br />

<strong>and</strong> l 2 is isomorphic to S 5 . An isomorphism is given by the operation on the<br />

lines g 1 , . . . , g 5 .<br />

In particular, the group order is indeed 27·16·5! = 51 840.<br />

ii. The Brauer-Manin obstruction. —<br />

2.7. –––– For Fanovarieties, allknownobstructionsagainsttheHasseprinciple<br />

are explained by the following observation.


286 GENERAL CUBIC SURFACES [Chap. IX<br />

2.8. Observation (Manin). —– Let π : X → SpecÉbe a non-singular variety<br />

overÉ. Choose an element α ∈ Br(X ) from its Brauer group.<br />

Then, anyÉ-<strong>rational</strong> point x ∈ X (É) gives rise to an adelic point (x ν ) ν ∈ X (AÉ)<br />

satisfying the condition<br />

ev(α, x) :=<br />

∑ inv(α| xν ) = 0 .<br />

ν∈Val(É)<br />

Here, inv: Br(Éν) →É/(respectively inv: Br(Ê) → 1 2/) denotes the canonical<br />

isomorphism.<br />

Proof. This is the Brauer-Manin obstruction which was studied in detail in Chapter<br />

III. The assertion was proven in Proposition III.2.3.b.iv).<br />

□<br />

2.9. –––– Recall the following facts.<br />

i) Proposition III.2.3.a.ii) shows that inv(α| xν ) depends continuously<br />

on x ν ∈ X (Éν). Further, by Proposition III.2.3.b.ii), for every proper variety<br />

X overÉ<strong>and</strong> each α ∈ Br(X ), there exists a finite set S ⊂ Val(É) such that<br />

inv(α| xν ) = 0 for every ν ∉ S <strong>and</strong> x ν ∈ X (Éν).<br />

This implies that the Brauer-Manin obstruction, if present, is an obstruction to<br />

the principle of weak approximation.<br />

ii) For certain varieties X, it happens, however, that the Brauer-Manin obstruction<br />

is not an obstruction to the Hasse principle. The point is that it possibly<br />

excludes only part of the adelic <strong>points</strong>.<br />

This is true, e.g., for the cubic surface given in Example III.5.25. Likewise for<br />

all diagonal cubic surfaces of the “first case” according to the classification<br />

of J.-L. Colliot-Thélène, D. Kanevsky, <strong>and</strong> J.-J. Sansuc [CT/S]. See Theorem<br />

III.6.4.<br />

iii) It is obvious that altering α ∈ Br(X ) by some Brauer class π ∗ ρ for ρ ∈ Br(É)<br />

does not change the obstruction defined by α. By consequence, it is only the<br />

factor group Br(X )/π ∗ Br(É) which is relevant for the Brauer-Manin obstruction.<br />

As was shown in Proposition II.8.11 <strong>and</strong> Remark II.8.13, the latter<br />

is canonically isomorphic to H 1( Gal(É/É), Pic(XÉ) ) . In particular,<br />

if H 1( Gal(É/É), Pic(XÉ) ) = 0 then there is no Brauer-Manin obstruction<br />

on X.<br />

2.10. –––– For a smooth cubic surface X, the geometric Picard group Pic(XÉ)<br />

is generated by the classes of the 27 lines on XÉ. Its first cohomology group<br />

can be described in terms of the Galois action on these lines. Indeed, by<br />

Proposition II.8.18, there is a canonical isomorphism<br />

H 1( Gal(É/É), Pic(XÉ) ) ∼ = Hom<br />

(<br />

(NF ∩ F0 )/NF 0 ,É/) . (†)


Sec. 3] THE GALOIS OPERATION ON THE 27 LINES 287<br />

Here, F ⊂ Div(XÉ) is the group generated by the 27 lines, F 0 ⊂ F denotes<br />

the subgroup of principal divisors, <strong>and</strong> N is the norm map under the operation<br />

of Gal(É/É)/H , H being the stabilizer of F.<br />

2.11. Remark. –––– Consider the particular case when the Galois group acts<br />

transitively on the 27 lines. Then, (†) shows that H 1( Gal(É/É), Pic(XÉ) ) = 0.<br />

In particular, there is no Brauer-Manin obstruction in this case.<br />

It is expected that the Hasse principle holds for all cubic surfaces such that<br />

H 1( Gal(É/É), Pic(XÉ) ) = 0. A statement which is very close to the<br />

Hasse principle was conjectured by J.-L. Colliot-Thélène <strong>and</strong> J.-J. Sansuc [CT/S,<br />

Conjecture C].<br />

3. The Galois operation on the 27 lines<br />

Let X be a smooth cubic surface defined overÉ<strong>and</strong> let K be the field of<br />

definition of the 27 lines on X. By Fact 2.3, K/Éis a Galois extension <strong>and</strong><br />

the Galois group G := Gal(K/É) is a subgroup of W (E 6 ). For general cubic<br />

surfaces, G is actually equal to W (E 6 ). To verify this for particular examples,<br />

the following lemma is useful.<br />

3.1. Lemma. –––– Let H ⊆ W (E 6 ) be a subgroup which acts transitively on the<br />

27 lines <strong>and</strong> contains an element of order five. Then, either H is the subgroup<br />

U ⊂ W (E 6 ) of index two or H = W (E 6 ).<br />

Proof. H ∩ U still acts transitively on the 27 lines <strong>and</strong> still contains an element<br />

of order five. Thus, we may suppose H ⊆ U.<br />

Assume that H U. Denote by k the index of H in U. The natural action<br />

of U on the set of cosets U/H yields a permutation representation i : U → S k .<br />

As U is simple, i is necessarily injective. In particular, since #U ∤ 8!, we see<br />

that k > 8. Let us consider the stabilizer H ′ ⊂ H of one of the lines. As H acts<br />

transitively, it follows that #H ′ = #H = #U . We distinguish two cases.<br />

27<br />

= 960<br />

27·k k<br />

First case: k > 16. Then, k ≥ 20 <strong>and</strong> #H ′ ≤ 48. This implies that the<br />

5-Sylow subgroup is normal in H ′ . Its conjugate by some σ ∈ H therefore depends<br />

only on σ ∈ H /H ′ . By consequence, the number n of 5-Sylow sub<strong>groups</strong><br />

in H is a divisor of #H /#H ′ = 27. Sylow’s congruence n ≡ 1 (mod 5) yields<br />

that n = 1.<br />

Let H 5 ⊂ H be the 5-Sylow subgroup. Then, ι(H 5 ) ⊂ S 27 is generated by a<br />

product of disjoint 5-cycles leaving at least two lines fixed. It is, therefore, not<br />

normal in the transitive group ι(H ). This is a contradiction.<br />

Second case: 9 ≤ k ≤ 16. We have k | 960. On the other h<strong>and</strong>, the assumption<br />

5 | #H implies 5 ∤k. This shows, there are only two possibilities, k = 12


288 GENERAL CUBIC SURFACES [Chap. IX<br />

<strong>and</strong> k = 16. As, in U, there is no subgroup of index eight or less, H ⊂ U<br />

must be a maximal subgroup. In particular, the permutation representation<br />

i : U → S k is primitive.<br />

Primitive permutation representations of degree up to 20 have been classified<br />

already in the late 19th century. It is well known that no group of order 25 920<br />

allows a faithful primitive permutation representation of degree 12 or 16 [Sim,<br />

Table 1].<br />

□<br />

3.2. Remark. –––– The sub<strong>groups</strong> of the simple group U have been completely<br />

classified by L. E. Dickson [Di04] in 1904. There are 114 conjugacy classes of<br />

proper sub<strong>groups</strong> H ≠ {e} in U. It would not be complicated to deduce the<br />

lemma from Dickson’s list.<br />

3.3. –––– Let the smooth cubic surface X be given by a homogeneous equation<br />

f = 0 with integral coefficients. We want to compute the Galois group G.<br />

An affine part of a general line l can be described by four coefficients a, b, c, d<br />

via the parametrization<br />

l: t ↦→ (1 : t : (a + bt) : (c + dt)) .<br />

l is contained in S if <strong>and</strong> only if it intersects S in at least four <strong>points</strong>. This implies<br />

that<br />

f (l(0)) = f (l(∞)) = f (l(1)) = f (l(−1)) = 0<br />

is a system of equations for a, b, c, d which encodes that l is contained in S.<br />

By a Gröbner base calculation in SINGULAR, we compute a univariate polynomial<br />

g of minimal degree belonging to the ideal generated by the equations. If g<br />

is of degree 27 then the splitting field of g is equal to the field K of definition<br />

of the 27 lines on X. We then use van der Waerden’s criterion [Poh/Z, Proposition<br />

2.9.35]. This means, we factor g modulo various primes. If all irreducible<br />

factors are distinct then the degrees of the factors describe the cycle structure of<br />

an element in the Galois group of g.<br />

If the various factorizations are incompatible [Coh, Section 3.5.2] then we get<br />

the information that g is irreducible. The Galois group G then acts transitively<br />

on the 27 lines. Furthermore, if we find an element of order five <strong>and</strong> an element<br />

which is not contained in the index two subgroup then G is actually equal<br />

to W (E 6 ). More precisely, our algorithm works as follows.<br />

3.4. Algorithm (Verifying G = W (E 6 )). —– Given the equation f = 0 of a<br />

smooth cubic surface, this algorithm verifies G = W (E 6 ).


Sec. 3] THE GALOIS OPERATION ON THE 27 LINES 289<br />

i) Compute a univariate polynomial 0 ≠ g ∈[d] of minimal degree such that<br />

g ∈ ( f (l(0)), f (l(∞)), f (l(1)), f (l(−1))) ⊂É[a, b, c, d]<br />

where l: t ↦→ (1 : t : (a + bt) : (c + dt)).<br />

If g is not of degree 27 then terminate with an error message. In this case, the<br />

coordinate system for the lines is not sufficiently general. If we are erroneously<br />

given a singular cubic surface then the algorithm will fail at this point.<br />

ii) Factor g modulo all primes below a given limit. Ignore the primes dividing<br />

the leading coefficient of g.<br />

iii) If one of the factors is multiple then go to the next prime immediately.<br />

Otherwise, check whether the decomposition type corresponds to one of the<br />

cases listed below,<br />

A := {(9, 9, 9)} , B := {(1, 1, 5, 5, 5, 5, 5), (2, 5, 5, 5, 10)} ,<br />

C := {(1, 4, 4, 6, 12), (2, 5, 5, 5, 10), (1, 2, 8, 8, 8)} .<br />

iv) If each of the cases occurred for at least one of the primes then output the<br />

message “The Galois group is equal to W (E 6 ).” <strong>and</strong> terminate.<br />

Otherwise, output “Can not prove that the Galois group is equal to W (E 6 ).”<br />

3.5. Remark. –––– The cases above are functioning as follows.<br />

a) Case B shows that the order of the Galois group is divisible by five.<br />

b) Cases A <strong>and</strong> B together guarantee that g is irreducible. Therefore,<br />

by Lemma 3.1, A <strong>and</strong> B prove that G contains the index two subgroup<br />

U ⊂ W (E 6 ).<br />

c) Case C is a selection of the most frequent odd conjugacy classes in W (E 6 ).<br />

3.6. Remark. –––– One could replace cases B <strong>and</strong> C by their common element<br />

(2, 5, 5, 5, 10). This would lead to a simpler but less efficient algorithm.<br />

3.7. Remark. –––– Actually, a decomposition type as considered in step iii)<br />

does not always represent a single conjugacy class in W (E 6 ). Two elements<br />

ι(σ), ι(σ ′ ) ∈ S 27 might be conjugate in S 27 via a permutation τ ∉ ι(W (E 6 )).<br />

For example, as is easily seen using GAP, the decomposition type (3, 6, 6, 6, 6)<br />

falls into three conjugacy classes two of which are even <strong>and</strong> one is odd (cf. Remark<br />

2.4). However, all the decomposition types searched for in Algorithm 3.4<br />

do represent single conjugacy classes.


290 GENERAL CUBIC SURFACES [Chap. IX<br />

3.8. Remark. –––– Since we expect G = W (E 6 ), we can estimate the probability<br />

of each case by the Chebotarev density theorem. Case A has a probability<br />

of 1 . This is the lowest value among the three cases.<br />

9<br />

3.9. Remark. –––– As we do not use the factors of g explicitly, it is enough<br />

to compute their degrees <strong>and</strong> to check that each of them occurs with multiplicity<br />

one. This means, we only have to compute gcd(g(X ), g ′ (X ))<br />

<strong>and</strong> gcd(g(X ), X pd −X ) inp[X ] for d = 1, 2, . . . , 13 [Coh, Algorithms 3.4.2<br />

<strong>and</strong> 3.4.3].<br />

4. Computation of Peyre’s constant<br />

i. The Euler product. —<br />

4.1. –––– We want to compute the product over all τ p . For a finite place p,<br />

we have<br />

(<br />

τ p = 1 − 1 )<br />

X (/p m)<br />

· lim .<br />

p m→∞ p 2m<br />

If the reduction is smooth then the sequence under the limit is constant<br />

Xp<br />

by virtue of Hensel’s Lemma. Otherwise, it becomes stationary after finitely<br />

many steps.<br />

We approximate the infinite product over all τ p by the finite product taken over<br />

the primes less than 300. Numerical experiments show that the contributions of<br />

larger primes do not lead to a significant change. (Compare the values calculated<br />

for the concrete example in Section 6.)<br />

ii. The factor at the infinite place. —<br />

4.2. –––– We want to compute<br />

τ ∞ = 1 2<br />

where the domain of integration is given by<br />

Z<br />

R<br />

1<br />

‖ grad f ‖ 2<br />

dS<br />

R = {(x, y, z, w) ∈ [−1, 1] 4 | f (x, y, z, w) = 0} .<br />

Here, dS denotes the usual hypersurface measure on R, considered as a hypersurface<br />

inÊ4 . Thus, τ ∞ is given by a three-dimensional integral.<br />

Since f is a homogeneous polynomial, we may reduce to an integral over the<br />

boundary of R which is a two-dimensional domain. In our particular case, we


Sec. 4] COMPUTATION OF PEYRE’S CONSTANT 291<br />

have deg f = 3. Then, a direct computation leads to<br />

Z<br />

Z<br />

τ ∞ =<br />

dA +<br />

‖( ∂ f<br />

‖( ∂ f<br />

Z<br />

+<br />

1<br />

, ∂ f , ∂ f )‖ R 0 ∂y ∂z ∂w 2<br />

1<br />

‖( ∂ f , ∂ f , ∂ f )‖ R 2 ∂x ∂y ∂w 2<br />

where the domains of integration are<br />

Z<br />

dA +<br />

1<br />

, ∂ f , ∂ f )‖ R 1 ∂x ∂z ∂w 2<br />

1<br />

‖( ∂ f , ∂ f , ∂ f )‖ R 3 ∂x ∂y ∂z 2<br />

R i = {(x 0 , x 1 , x 2 , x 3 ) ∈ [−1, 1] 4 | x i = 1 <strong>and</strong> f (x 0 , x 1 , x 2 , x 3 ) = 0} .<br />

dA denotes the two-dimensional hypersurface measure on R i , considered as a<br />

hypersurface inÊ3 .<br />

We therefore have to integrate a smooth function over a compact part of a<br />

smooth two-dimensional submanifold inÊ3 . To do this, we approximate the<br />

domain of integration by a triangular mesh.<br />

4.3. Algorithm (Generating a triangular mesh). —–<br />

Given the equation f = 0 of a smooth surface inÊ3 , this algorithm constructs<br />

a triangular mesh approximating the part of the surface which is contained in a<br />

given cube.<br />

i) We split the cube into eight smaller cubes <strong>and</strong> iterate this procedure a predefined<br />

number of times, recursively. During recursion, we exclude those<br />

cubes which obviously do not intersect the manifold. To do this, we estimate<br />

‖grad f ‖ 2 on each cube.<br />

ii) Then, each resulting cube is split into six simplices.<br />

iii) For each edge of each simplex which intersects the manifold, we compute<br />

an approximation of the point of intersection. We use them as the vertices of<br />

the triangles to be constructed. This leads to a mesh consisting of one or two<br />

triangles per simplex.<br />

4.4. Remark. –––– Similar algorithms are popular in the computer graphics<br />

community.<br />

4.5. –––– The next step is to compute the contribution of each triangle ∆ i to<br />

the integral. For this, we use some adaption of the midpoint rule.<br />

We approximate the integr<strong>and</strong> g by its value g(C i ) at the barycenter C i of<br />

the triangle. Note that this point usually lies outside the surface given by f = 0.<br />

Algorithm 4.3 guarantees only that the three vertices of each facet are contained<br />

in that surface. We map it to the manifold along the flow of the gradient<br />

field grad f .<br />

dA<br />

dA


292 GENERAL CUBIC SURFACES [Chap. IX<br />

The product g(C i )A(∆ i ) seems to be a reasonable approximation of the contribution<br />

of ∆ i to the integral. We correct by an additional factor, the cosine of the<br />

angle between the normal vector of the triangle <strong>and</strong> the gradient vector grad f<br />

at the barycenter C. We calculated the contribution of each triangle by some<br />

adaption of the midpoint rule. The problem is our algorithm makes sure for<br />

each facet that its three vertices are contained in the manifold R i . This means<br />

that its barycenter is usually outside R i . We map it to the manifold along the<br />

flow of the gradient field grad f .<br />

4.6. Remark. –––– In our application, these correctional factors are close to 1<br />

<strong>and</strong> seem to converge versus 1 when the number of recursions is growing.<br />

This is, however, not a priori clear. H. A. Schwarz’s cylindrical surface [Schw]<br />

constitutes a famous example of a sequence of triangulations where the triangles<br />

become arbitrarily small <strong>and</strong> the factors are nevertheless necessary for<br />

correct integration.<br />

We use the method described above to approximate the value of τ ∞ . In Algorithm<br />

4.3, we work with six recursions.<br />

4.7. Remark. –––– Our method of numerical integration is a combination of<br />

st<strong>and</strong>ard algorithms for 2.5-dimensional mesh generation <strong>and</strong> two dimensional<br />

integration which are described in the literature [Hb].<br />

On a triangle, we integrate linear functions correctly. This indicates that the<br />

method should converge of second order. The facts that we work with the<br />

area of a linearized triangle <strong>and</strong> that the barycenters C i are located in a certain<br />

distance from the manifold generate errors of the same order.<br />

5. Numerical Data<br />

i. The computations carried out. —<br />

5.1. –––– A general cubic surface is described by twenty coefficients. With current<br />

technology, it is impossible to study all cubic surfaces with coefficients<br />

below a reasonable bound. For that reason, we decided to work with coefficient<br />

vectors provided by a r<strong>and</strong>om number generator. Our first sample consists<br />

of 20 000 surfaces with coefficients r<strong>and</strong>omly chosen in the interval [0, 50].<br />

The second sample consists of 20 000 surfaces with r<strong>and</strong>omly chosen coefficients<br />

from the interval [−100, 100].<br />

These limits were, of course, chosen somewhat arbitrarily. There is, at least,<br />

some reason not to work with too large limits as this would lead to low values


Sec. 5] NUMERICAL DATA 293<br />

of τ. (The reader might want to compare Theorem X.6.21 where this is rigorously<br />

proven in a different situation.) Low values of τ are undesirable as they<br />

require high search bounds in order to satisfactorily test Manin’s conjecture.<br />

We verified explicitly that each of the surfaces studied is smooth. For this,<br />

we inspected a Gröbner base of the ideal corresponding to the singular locus.<br />

The computations were done in SINGULAR.<br />

Then, using Algorithm 3.4, <strong>and</strong> a prime limit of 1000, we proved that, for each<br />

surface, the full Galois group W (E 6 ) acts on the 27 lines. The largest prime used<br />

was 457. This means that all our examples are general from the Galois point<br />

of view. By consequence, their Picard ranks are equal to 1. Further, according<br />

to Remark 2.11, the Brauer-Manin obstruction is not present on any of the<br />

surfaces considered.<br />

Almost as a byproduct, we verified that no two of the 40 000 surfaces are isomorphic.<br />

Actually, when running part ii) of Algorithm 3.4, we wrote the<br />

decomposition types found into a file. Primes at which the algorithm failed<br />

were labeled by a special marker. A program, written in C, ran in an iterated<br />

loop over all pairs of surfaces <strong>and</strong> looked for a prime at which the decomposition<br />

types differ. The largest prime needed to distinguish two surfaces was 73.<br />

We counted allÉ-<strong>rational</strong> <strong>points</strong> of height less than 250 on the surfaces of the<br />

first sample. It turns out that, on two of these surfaces, there are noÉ-<strong>rational</strong><br />

<strong>points</strong> occurring as the equation is unsolvable inÉp for some small p. In this<br />

situation, Manin’s conjecture is true, trivially.<br />

On each of the remaining surfaces, we found at least oneÉ-<strong>rational</strong> point.<br />

228 examples contained less than ten <strong>points</strong>. On the other h<strong>and</strong>, 1213 examples<br />

contained at least one hundredÉ-<strong>rational</strong> <strong>points</strong>. The largest number of <strong>points</strong><br />

found was 335.<br />

For the second sample, the search bound was 500. Again, on two of these<br />

surfaces, there are noÉ-<strong>rational</strong> <strong>points</strong> occurring as the equation is unsolvable in<br />

a certainÉp. There were 202 examples containing between one <strong>and</strong> nine <strong>points</strong>.<br />

1857 examples contained at least one hundredÉ-<strong>rational</strong> <strong>points</strong>. The largest<br />

number of <strong>points</strong> found was 349.<br />

To find theÉ-<strong>rational</strong> <strong>points</strong>, we used a 2-adic search method which works<br />

as follows. Let a cubic surface X be given. Then, in a first step, we determined<br />

on X all <strong>points</strong> defined over/512(respectively/1024). Then, for each<br />

of the <strong>points</strong> found we checked which of its lifts to P 3 () actually lie on X.<br />

This leads to an O(B 3 )-algorithm which may be efficiently implemented in C.<br />

Furthermore, using the method described in Section 4, we computed an approximation<br />

of Peyre’s constant for each surface.


294 GENERAL CUBIC SURFACES [Chap. IX<br />

5.2. Remark. –––– To search for theÉ-<strong>rational</strong> <strong>points</strong>, there are algorithms<br />

which are asymptotically faster. For example, Elkies’ method which is O(B 2 )<br />

<strong>and</strong> implemented in Magma. A practical comparison shows, however, that<br />

Elkies’ method is not yet faster for our relatively low search bounds.<br />

ii. The density results. —<br />

5.3. –––– For each of the surfaces considered we calculated the quotient<br />

#{ <strong>points</strong> of height < B found } / #{ <strong>points</strong> of height < B expected }.<br />

Let us visualize the distribution of the quotients by some histograms.<br />

2.5<br />

2.5<br />

2<br />

2<br />

1.5<br />

1.5<br />

1<br />

1<br />

0.5<br />

0.5<br />

0<br />

0.5 1 1.5 2 2.5 0<br />

0.5 1 1.5 2 2.5<br />

First sample, B =125 First sample, B =250<br />

FIGURE 1. Distribution of the quotients for the first sample


Sec. 5] NUMERICAL DATA 295<br />

2.5<br />

2.5<br />

2<br />

2<br />

1.5<br />

1.5<br />

1<br />

1<br />

0.5<br />

0.5<br />

0<br />

0.5 1 1.5 2 2.5 0<br />

0.5 1 1.5 2 2.5<br />

Second sample, B =250 Second sample, B =500<br />

FIGURE 2. Distribution of the quotients for the second sample<br />

Some statistical parameters are as follows.<br />

TABLE 1. Parameters of the distribution for the first sample<br />

search bound 125 250<br />

mean 0.99993 0.99887<br />

st<strong>and</strong>ard deviation 0.23558 0.16925<br />

TABLE 2. Parameters of the distribution for the second sample<br />

search bound 250 500<br />

mean 1.00093 0.99943<br />

st<strong>and</strong>ard deviation 0.22527 0.16158<br />

iii. The results for the smallest point. —<br />

5.4. –––– For each of the surfaces in our samples, we determined the height<br />

m(X ) of its smallest point. We visualize the behaviour of m(X ) in the diagrams<br />

below.<br />

At the first glance, it looks very natural to consider the distribution of the values<br />

of m(X ) versus the Tamagawa type number τ (X ). In view of the inequalities<br />

asked for in the introduction, it seems, however, to be better to make a slight<br />

modification <strong>and</strong> plot the product m(X )τ (X ) instead of m(X ) itself.


296 GENERAL CUBIC SURFACES [Chap. IX<br />

m<br />

1000<br />

m<br />

1000<br />

100<br />

100<br />

10<br />

10<br />

1<br />

0.001 0.01 0.1 1 10<br />

First sample<br />

τ<br />

1<br />

0.001 0.01 0.1 1 10<br />

Second sample<br />

τ<br />

FIGURE 3. The smallest height of a <strong>rational</strong> point versus the Tamagawa number<br />

5.5. Conclusion –––– Our experiments suggest that, for general cubic surfaces<br />

X overÉ, the following assertions hold.<br />

i) There are no obstructions to the Hasse principle.<br />

ii) Manin’s conjecture is true in the form refined by E. Peyre.<br />

Further, it is apparent from the diagrams in Figure 3 that the experiment agrees<br />

with the expectation for the heights of the smallest <strong>points</strong> formulated in Remark<br />

1.11. Indeed, for both samples, the slope of a line tangent to the top right<br />

of the scatter plot is near (−1). This indicates that even the strong form of the<br />

estimate should be true. I.e., m(X ) < C (ε)<br />

τ (X ) 1+ε for any ε > 0.<br />

5.6. Running Times –––– The largest portion of the running time was spent on<br />

the calculation of the Euler products. It took 20 days of CPU time to calculate<br />

all 40 000 Euler products for p < 300.<br />

For comparison, we estimated all the integrals, using six recursions, within<br />

36 hours. Further, it took eight days to systematically search for all <strong>points</strong> of<br />

height less than 500 on the surfaces of the second sample. Search for <strong>points</strong> of<br />

height less than 250 on the surfaces of the first sample took only one day.<br />

When running Algorithm 3.4, the lion’s share of the time was used for the<br />

computation of the univariate degree 27 polynomials. This took approximately<br />

seven days of CPU time.<br />

In comparison with that, all other parts were negligible. It took only twelve<br />

minutes to ensure that all 40 000 surfaces are smooth. The C program verifying<br />

that no two of the surfaces are isomorphic to each other ran approximately<br />

80 seconds.


Sec. 6] A CONCRETE EXAMPLE 297<br />

6. A concrete example<br />

i. The Example. —<br />

6.1. Example –––– Let us conclude the article by some results on the particular<br />

cubic surface X given by<br />

x 3 + 2xy 2 + 11y 3 + 3xz 2 + 5y 2 w + 7zw 2 = 0 . (‡)<br />

Example (‡) was not among the surfaces produced by the r<strong>and</strong>om number generator.<br />

Our intention is just to present the output of our algorithms in a specific<br />

(<strong>and</strong> not too artificial) example <strong>and</strong>, most notably, to show the intermediate<br />

results of Algorithm 3.4.<br />

A Gröbner base calculation inMagma shows that X has bad reduction at p = 2,<br />

3, 7, 23, <strong>and</strong> 22 359 013 270 232 677. The idea behind that calculation is the same<br />

as described above for the verification of smoothness. The only difference is<br />

that we consider Gröbner bases overinstead ofÉ.<br />

6.2. The Galois group –––– The first step of Algorithm 3.4 works well on X.<br />

I.e., the polynomial g is indeed of degree 27. Its coefficients become rather large.<br />

The absolutely largest one is that of d 13 . It is equal to 38 300 982 629 255 010.<br />

The leading coefficient of g is 5 3 · 7 12 . We find case A at p = 373. The common<br />

decomposition type (2, 5, 5, 5, 10) of the cases B <strong>and</strong> C occurs at p = 19, 31, 59,<br />

61, 191, 199, <strong>and</strong> 223.<br />

Consequently, X is an explicit example of a smooth cubic surface overÉ<br />

admitting the property that the Galois group which acts on the 27 lines is equal<br />

to W (E 6 ).<br />

6.3. Remark. –––– ThefirstsuchexampleshavebeenconstructedbyT. Ekedahl<br />

[Ek, Theorem 2.1].<br />

6.4. Remark. –––– Our example (‡) is different from Ekedahl’s. Indeed, in<br />

Ekedahl’s examples, the Frobenius Frob 11 acts on the 27 lines as an element<br />

of the conjugacy class C 15 ⊂ W (E 6 ) (in Sir P. Swinnerton-Dyer’s numbering).<br />

In our case, however, the first two steps of Algorithm 3.4 show that Frob 11<br />

yields the decomposition type (1, 1, 1, 1, 1, 2, 4, 4, 4, 4, 4). This corresponds to<br />

the class C 18 [Man,кIV, §9,Ìн]. Note that Ekedahl’s examples, as<br />

well as ours, have good reduction at p = 11.


298 GENERAL CUBIC SURFACES [Chap. IX<br />

ii. The computation of Peyre’s constant. —<br />

6.5. –––– As an approximation of the Euler product, we get<br />

∏ τ p ≈ 0. 729 750 .<br />

p


Sec. 6] A CONCRETE EXAMPLE 299<br />

TABLE 4. Points on X of height ≤ 20<br />

Point Height Point Height<br />

( 0 : 0 : 0 : 1) 1 ( 4 : -6 : -13 : 2) 13<br />

( 0 : 0 : 1 : 0) 1 ( 5 : 5 : 0 : -14) 14<br />

( 1 : 2 : -3 : -2) 3 (10 : -14 : 12 : 11) 14<br />

( 3 : -2 : -3 : 2) 3 ( 0 : 7 : -16 : 7) 16<br />

( 0 : 7 : -6 : -7) 7 (16 : -8 : 3 : -4) 16<br />

( 0 : 4 : -3 : 8) 8 ( 6 : -9 : 16 : 3) 16<br />

( 5 : -5 : 0 : 8) 8 (12 : 7 : -6 : 17) 17<br />

( 2 : -8 : 8 : 7) 8 (14 : -9 : -2 : 17) 17<br />

(10 : -5 : 0 : -1) 10 ( 6 : -3 : -18 : 7) 18<br />

( 0 : 5 : 0 : -11) 11 ( 9 : -6 : -1 : 18) 18<br />

( 8 : 6 : -11 : -8) 12 ( 3 : 6 : -19 : -6) 19<br />

(12 : -6 : -4 : -3) 12 ( 8 : 7 : -4 : 19) 19<br />

( 9 : -12 : 9 : 10) 12<br />

iv. The field of definition of the 27 lines. —<br />

6.7. –––– Having done the Gröbner base calculation in Algorithm 3.4.i),<br />

the 27 lines may be computed at high precision. This allows to find the 45 triangles<br />

on X,explicitly. We calculated a degree 45 resolvent G of the degree 27 polynomial<br />

g the zeroes of which are all the sumsa i1 +a i2 +a i3 for l i1 , l i2 , l i3 representing<br />

three lines which form a triangle. Here, l i : t ↦→ (1 : t : (a i +b i t) : (c i +d i t))<br />

denote parametrizations of the 27 lines. As G ∈[X ], our floating point<br />

calculation is in fact exact.<br />

6.8. Proposition. –––– The unique quadratic subfield in the field K of definition of<br />

the 27 lines on X isÉ(√<br />

−23 · 22 359 013 270 232 677<br />

)<br />

.<br />

Proof. K is unramified at all places of good reduction of X. This leaves us with<br />

only 2 6 − 1 = 63 possibilities for the quadratic subfieldÉ( √ d). To exclude 62<br />

of them is algorithmically easy.<br />

Indeed, for a good prime p, there is a way to compute ( d<br />

p)<br />

without knowledge<br />

of d. We factor the degree 45 resolvent G modulo p. If p divides the<br />

leading coefficient or there are multiple factors then we get no answer. Otherwise,<br />

( d<br />

p) = ±1 depending on whether the decomposition type found is even<br />

or odd in S 45 .<br />

It turns out that it is sufficient to do this for p = 13, 17, 19, 29, 31, <strong>and</strong> 53.<br />


300 GENERAL CUBIC SURFACES [Chap. IX<br />

6.9. Proposition. –––– The field extension K/Éis ramified exactly at p = 2, 3, 7,<br />

23, <strong>and</strong> 22 359 013 270 232 677.<br />

Proof. It remains to verify ramification at p = 2, 3, <strong>and</strong> 7. For that, we<br />

computed in Magma the p-adic factorization of g. The decomposition types are<br />

(3, 24) for p = 2 <strong>and</strong> 3 <strong>and</strong> (1, 1, 1, 4, 4, 4, 4, 8) for p = 7.<br />

Let Z p be the decomposition field of p. If p were unramified then Gal(K/Z p )<br />

would be a cyclic group, i.e. Gal(K/Z p ) = 〈σ〉 for some σ ∈ W (E 6 ). On the<br />

other h<strong>and</strong>, on the 27 lines, the orbit structure under the operation of Gal(K/Z p )<br />

isthesameasunder theoperationofGal(Ép/Ép). Thereis, however, noelement<br />

in W (E 6 ) which yields the decomposition type (3, 24) or (1, 1, 1, 4, 4, 4, 4, 8)<br />

[Man,кIV, §9,Ìн].<br />

□<br />

6.10. Remark. –––– This shows that there is no integral model of X which is<br />

smooth over p = 2, 3, 7, 23, or 22 359 013 270 232 677.


CHAPTER X<br />

ON THE SMALLEST POINT ON A<br />

DIAGONAL CUBIC SURFACE ∗<br />

All life is an experiment. The more experiments you make the better.<br />

RALPH WALDO EMERSON (1842)<br />

1. Introduction<br />

1.1. –––– In Chapter VIII, we presented some theoretical <strong>and</strong> experimental<br />

results concerning the height of the smallestÉ-<strong>rational</strong> point on diagonal quartic<br />

threefolds in P 4É. In this chapter, we will consider diagonal cubic surfaces in P 3É.<br />

It will turn out that the analogies to the case treated before are enormous.<br />

However, there are a few <strong>points</strong> indicating that the case of a diagonal cubic<br />

surface is technically more complicated.<br />

An obvious difference is that the Picard rank of a quartic threefold is always<br />

equal to one. By consequence, H 1( Gal(É/É), Pic(XÉ) ) is automatically the<br />

trivial group. Both these observations are wrong for diagonal cubic surfaces.<br />

Hence, the factors α <strong>and</strong> β in the definition of Peyre’s constant (Definition<br />

VI.5.35)will not always be the same <strong>and</strong> need to be considered. Further, one<br />

has to expect that a Brauer-Manin obstruction is present.<br />

1.2. –––– The conjecture of Manin states that the number ofÉ-<strong>rational</strong> <strong>points</strong><br />

of anticanonical height < B on a Fano variety X is asymptotically equal to<br />

τB log rkPic(X )−1 B for B → ∞.<br />

In the particular case of a cubic surface, the anticanonical height is the same<br />

as the naive height. Further, the coefficient τ ∈Êequals the Tamagawa-type<br />

number τ (X ) introduced by E. Peyre.<br />

Thus, one expects at least ∼τ (X )B <strong>points</strong> of height


302 ON THE SMALLEST POINT ON A DIAGONAL CUBIC SURFACE [Chap. X<br />

this might lead to the expectation that m(X ), the smallest height of aÉ-<strong>rational</strong><br />

point on X, is always less than C<br />

τ<br />

for a certain absolute constant C.<br />

(X )<br />

1.3. –––– To test this expectation, we computed the Tamagawa number <strong>and</strong><br />

ascertained the smallestÉ-<strong>rational</strong> point for each of the cubic surfaces given by<br />

ax 3 + by 3 + 2z 3 + w 3 = 0<br />

for a = 1, . . . , 3000 <strong>and</strong> b = 1, . . . , 300.<br />

Thereby, we restricted our considerations to the cases that<br />

i) a <strong>and</strong> b are odd,<br />

ii) there exists an odd prime p dividing a but not b such that 3 ∤ ν p (a), or<br />

iii) there exists an odd prime p dividing b but not a such that 3 ∤ν p (b).<br />

This guarantees that we are in the “first case” according to the classification of<br />

J.-L. Colliot-Thélène [CT/K/S] <strong>and</strong> his coworkers. In this case, the effect of<br />

the Brauer-Manin obstruction is clear. Precisely two thirds of the adelic <strong>points</strong><br />

are excluded.<br />

The results are summarized by the diagram below.<br />

FIGURE 1. Height of smallest point versus Tamagawa number<br />

It is apparent from the diagram that the experiment agrees with the expectation<br />

above. The slope of a line tangent to the top right of the scatter plot is<br />

indeed near (−1).<br />

However, we will show in Section 5 that, as in the threefold case, the inequality<br />

m(X ) < C<br />

τ<br />

does not hold in general. The following remains a logical possibility.<br />

(X )


Sec. 1] INTRODUCTION 303<br />

1.4. Question. –––– For every ε > 0, does there exist a constant C (ε) such<br />

that, for each cubic surface,<br />

m(X ) <<br />

C (ε)<br />

τ (X ) 1+ε ?<br />

1.5. Peyre’s constant. –––– Recall that, in Definition VI.5.35, E. Peyre’s Tamagawa-type<br />

number was defined as<br />

τ (X ) := α(X )·β(X ) · lim<br />

s→1<br />

(s − 1) t L(s, χ Pic(XÉ)) · τ H<br />

(<br />

X (É)<br />

for t = rk Pic(X ).<br />

The factor β(X ) is simply defined as<br />

β(X ) := #H 1( Gal(É/É), Pic(XÉ) ) .<br />

Br)<br />

α(X ) is given as follows. Let Λ eff (X ) ⊂ Pic(X )⊗Êbe the cone generated by<br />

the effective divisors. Identify Pic(X )⊗ÊwithÊt via a mapping induced by an<br />

∼=<br />

isomorphism Pic(X ) −→t . Consider the dual cone Λ ∨ eff (X ) ⊂ (Êt ) ∨ . Then,<br />

α(X ) := t · vol {x ∈ Λ ∨ eff | 〈x, −K 〉 ≤ 1 } .<br />

L( · , χ Pic(XÉ)) denotes the Artin L-function of the Gal(É/É)-representation<br />

Pic(XÉ) ⊗which contains the trivial representation t times as a direct summ<strong>and</strong>.<br />

Therefore, L(s, χ Pic(XÉ)) = ζ(s) t · L(s, χ P ) <strong>and</strong><br />

lim<br />

s→1<br />

(s − 1) t L(s, χ Pic(XÉ)) = L(1, χ P )<br />

where ζ denotes the Riemann zeta function <strong>and</strong> P is a representation which<br />

does not contain trivial components. [Mu-y, Corollary 11.5 <strong>and</strong> Corollary 11.4]<br />

show that L(s, χ P ) has neither a pole nor a zero at s = 1.<br />

Finally, τ H is the Tamagawa measure on the set X (É) of adelic <strong>points</strong> on X<br />

<strong>and</strong> X (É) Br ⊆ X (É) denotes the part which is not affected by the Brauer-<br />

Manin obstruction.<br />

1.6. The Tamagawa measure. –––– As X is projective, we have<br />

X (É) = ∏ X (Éν).<br />

ν∈Val(É)<br />

τ H is defined to be a product measure τ H := ∏ τ ν .<br />

ν∈Val(É)<br />

For a prime number p, the local measure τ p on X (Ép) is given as follows.<br />

Let X ⊆ P 3be the model of X given by the defining cubic equation.<br />

For a ∈ X (/p k), put<br />

U (k)<br />

a := {x ∈ X (Ép) | x ≡ a (mod p k ) } .


304 ON THE SMALLEST POINT ON A DIAGONAL CUBIC SURFACE [Chap. X<br />

Then,<br />

τ p (U (k)<br />

a ) := det(1 − p−1 Frob p |Pic(XÉ) I p<br />

)<br />

#{ y ∈ X (/p<br />

· m) | y ≡ a (mod p k ) }<br />

lim<br />

.<br />

m→∞ p m dim X<br />

Here, Pic(XÉ) I p<br />

denotes the fixed module under the inertia group.<br />

τ ∞ is described in Definition VI.5.22. In the case of a cubic surface defined by<br />

the equation f = 0, this yields<br />

τ ∞ (U ) = 1 Z<br />

ω Leray<br />

2<br />

CU<br />

|x 0 |, ... ,|x 3 |≤1<br />

for U ⊂ X (Ê). Here, ω Leray is the Leray measure on the cone CX (Ê) associated<br />

to the equation f = 0.<br />

The Leray measure is related to the usual hypersurface measure by the formula<br />

ω Leray = 1<br />

‖ grad f ‖ ω hyp.<br />

1.7. The results. –––– In the case of diagonal cubic surfaces, there is an estimate<br />

1<br />

for m(X ) in terms of τ (X ). Namely,<br />

τ<br />

admits a fundamental finiteness property.<br />

More precisely, in Section 6, we will show the following<br />

(X )<br />

theorem.<br />

Theorem. Let a = (a 0 , . . . , a 3 ) ∈ (\{0}) 4 be a vector. Denote by X a the cubic<br />

surface in P 3Égiven by a 0 x 3 0 + . . . + a 3 x 3 3 = 0. Then, for each ε > 0 there exists a<br />

constant C (ε) > 0 such that<br />

1<br />

τ (X a ) ≥ C (ε) · H( 1<br />

a 0<br />

: . . . :<br />

1<br />

a 3<br />

) 1<br />

3<br />

−ε<br />

.<br />

Corollary (Fundamental finiteness). For each T > 0, there are only finitely many<br />

diagonal cubic surfaces X a : a 0 x 3 0 + . . . + a 3 x 3 3 = 0 in P 3Ésuch that τ (V a ) > T .<br />

Corollary (An inefficient search bound). There exists a monotonically decreasing<br />

function F : (0, ∞) → [0, ∞), the search bound, satisfying the following condition.<br />

Let X a be the cubic surface given by the equation a 0 x 3 0 + . . . + a 3 x 3 3 = 0.<br />

Assume X a (É) ≠ ∅. Then, X a admits aÉ-<strong>rational</strong> point of height ≤F (τ (X a )).<br />

Proof. One may simply put F (t) := max min H(P).<br />

τ (X a )≥t P∈X a (É) X a (É)≠∅<br />

□<br />

In other words, we have m(X a ) ≤ F (τ (X a )) as soon as X a (É) ≠ ∅.


Sec. 2] THE FACTORS α AND β 305<br />

2. The factors α <strong>and</strong> β<br />

2.1. –––– Recall that on a smooth cubic surface X over an algebraically closed<br />

field, there are exactly 27 lines. For the Picard group, which is isomorphic to7 ,<br />

the classes of these lines form a system of generators.<br />

2.2. Notation. –––– i) ThesetL ofthe27linesisequippedwith theintersection<br />

product 〈 , 〉: L ×L → {−1, 0, 1}. The pair (L , 〈 , 〉) is the same for all<br />

smooth cubic surfaces. It is well known from Chapter II, 8.21 that the group<br />

of permutations of L respecting 〈 , 〉 is isomorphic to W (E 6 ). We fix such<br />

an isomorphism.<br />

Denote by F ⊂ Div(X ) the group generated by the 27 lines <strong>and</strong> by F 0 ⊂ F the<br />

subgroup of principal divisors. Then, F is equipped with an operation of W (E 6 )<br />

such that F 0 is a W (E 6 )-submodule. We have Pic(X ) ∼ = F/F 0 .<br />

ii) If X is a smooth cubic surface overÉthen Gal(É/É) operates canonically<br />

on the set L X of the 27 lines on XÉ. Fix a bijection i X : L X −→ L<br />

∼ =<br />

respecting the intersection pairing. This induces a group homomorphism<br />

ι X : Gal(É/É) → W (E 6 ). We denote its image by G ⊂ W (E 6 ).<br />

2.3. Lemma. –––– There is a constant C such that, for all smooth cubic surfaces X<br />

overÉ,<br />

1 ≤ β(X ) ≤ C.<br />

Proof. By definition, β(X ) = #H 1( Gal(É/É), Pic(XÉ) ) . Using the notation<br />

just introduced, we may write H 1( Gal(É/É), Pic(XÉ) ) = H 1 (G, F/F 0 ).<br />

Note that this cohomology group is always finite. Indeed, since G is a finite<br />

group <strong>and</strong> F/F 0 is a finite[G]-module, the description via the st<strong>and</strong>ard<br />

complex shows it is finitely generated. Further, it is annihilated by #G.<br />

H 1 (G, F/F 0 ) depends only on the subgroup G ⊂ W (E 6 ) occurring. For that,<br />

there are finitely many possibilities. This implies the claim.<br />

□<br />

2.4. Remark. –––– A more precise consideration (cf. Proposition II.8.18) yields<br />

a canonical isomorphism<br />

H 1( Gal(É/É), Pic(XÉ) ) ∼ = Hom<br />

(<br />

(NF ∩ F0 )/NF 0 ,É/) .<br />

Here, N is the norm map under the operation of G.<br />

As an application of this, one may inspect the 350 conjugacy classes of sub<strong>groups</strong><br />

of W (E 6 ) using GAP. See Chapter II, 8.23, or the complete list of the values<br />

of H 1( Gal(É/É), Pic(XÉ) ) given in the appendix. The calculations show that<br />

the lemma is actually true for C = 9.


306 ON THE SMALLEST POINT ON A DIAGONAL CUBIC SURFACE [Chap. X<br />

2.5. Lemma. –––– There are positive constants C 1 <strong>and</strong> C 2 such that, for all smooth<br />

cubic surfaces X overÉsatisfying X (É) ≠ ∅,<br />

C 1 ≤ α(X ) ≤ C 2 .<br />

Proof. Again, we claim that α(X ) is completely determined by the group<br />

G ⊂ W (E 6 ). Thus, suppose that we do not have the full information available<br />

about what surface X is but are given the group G only.<br />

The assumption X (É) ≠ ∅ makes sure that Pic(X ) ∼ = Pic(XÉ) G [K/T, Remark<br />

3.2.ii)]. We may therefore write Pic(X ) ∼ = (F/F 0 ) G . The effective cone<br />

Λ eff (X ) ⊂ Pic(X )⊗∼ = (F/F0 ) G ⊗is generated by the symmetrizations<br />

of the classes l 1 , . . . , l 27 of the 27 lines in F. In particular, it is determined<br />

by G, completely. Further, we have K = − 1(l 9 1 + . . . + l 27 ). These data are<br />

sufficient to compute α(X ) according to its very definition.<br />

□<br />

2.6. Remark. –––– Here, we do not know the optimal values of C 1 <strong>and</strong> C 2 in<br />

explicit form. α(X ) has not yet been computed in all cases.<br />

2.7. Remark. –––– In the experiment, we work entirely with cubic surfaces of<br />

Picard rank one overÉ. This may easily be seem from Theorem 3.6, below.<br />

Therefore, we always have α(X ) = 1. Further, Theorem III.6.4.a) implies<br />

that β(X ) = 3.<br />

3. Splitting the Picard group<br />

3.1. –––– In the case of the diagonal cubic surface X (a 0,...,a 3 )<br />

⊂ P 3É, given<br />

by a 0 x 3 0 + . . . + a 3x 3 3 = 0 for a 0, . . . , a 3 ∈\{0}, the 27 lines on X (a 0,...,a 3 )<br />

may easily be written down explicitly. Indeed, for each pair (i, j) ∈ (/3) 2 ,<br />

the system<br />

√ √<br />

3 a0 x 0 + ζ3<br />

i 3 a1 x 1 = 0<br />

√ √<br />

3 a2 x 2 + ζ j 3<br />

3 a3 x 3 = 0<br />

of equations defines a line on X (a 0,...,a 3 ) . Decomposing the index set<br />

{0, . . . , 3} differently into two subsets of two elements each yields<br />

all the lines. √In particular, √ we √ see that<br />

)<br />

the 27 lines may be defined<br />

over K =É( ζ3 , 3 a1 /a 0 , 3 a2 /a 0 , 3 a3 /a 0 .


Sec. 3] SPLITTING THE PICARD GROUP 307<br />

3.2. Fact. –––– Let p be a prime number <strong>and</strong> a 0 , . . . , a 3 be integers not divisible<br />

by p. Then,<br />

#X (a 0, ... ,a 3 ) (p) =<br />

⎧<br />

p<br />

⎪⎨<br />

2 + ( 1 + χ 3 (a 0 a 1 a 2 2 a2 3 ) + χ 3(a 2 0 a2 1 a 2a 3 )<br />

+ χ 3 (a 0 a 2 1 a 2a 2 3 ) + χ 3(a 2 0 a 1a 2 2 a 3)<br />

+ χ 3 (a 0 a 2 1<br />

⎪⎩<br />

a2 2 a 3) + χ 3 (a 2 0 a 1a 2 a 2 3 )) p + 1 if p ≡ 1 (mod 3) ,<br />

p 2 + p + 1 if p ≡ 2 (mod 3) .<br />

Here, in the case p ≡ 1 (mod 3), χ 3 :∗ p →denotes a cubic residue character.<br />

Proof. If p ≡ 2 (mod 3) then every residue class modulo p has a<br />

unique cubic root. Therefore, the map X (a 0, ... ,a 3 ) (p) → P 2 (p) given by<br />

(x : y : z : w) ↦→ (x : y : z) is bijective. This shows #X (a 0, ... ,a 3 ) (p) = p 2 +p+1.<br />

Turn to the case p ≡ 1 (mod 3). It is classically known that, on a degree m<br />

diagonal variety, the number ofp-<strong>rational</strong> <strong>points</strong> for p ≡ 1 (mod m) may<br />

be determined using Jacobi sums. The formula given follows immediately<br />

from [I/R, Chapter 10, Theorem 2] together with the well-known relation<br />

g(χ 3<br />

É<br />

)g(χ 2 3 ) = p for cubic Gauß sums.<br />

□<br />

3.3. Lemma. –––– Let a 0 , . . . , a 3<br />

É<br />

∈\{0}. Then, for each prime p such that<br />

p ∤ 3a 0 · . . . · a 3 ,<br />

χ<br />

(a Pic(X 0 ,...,a 3 )<br />

)(Frob p ) = tr ( )⊗)<br />

Frob p | Pic(X (a 0,...,a 3 )<br />

⎧<br />

χ 3 (a 0 a 1 a<br />

⎪⎨<br />

2 2 a2 3 ) + χ 3(a 2 0 a2 1 a 2a 3 )<br />

+ χ 3 (a 0 a 2 1<br />

=<br />

a 2a 2 3 ) + χ 3(a 2 0 a 1a 2 2 a 3)<br />

+ χ 3 (a 0 a 2 1<br />

⎪⎩<br />

a2 2 a 3) + χ 3 (a 2 0 a 1a 2<br />

É ⊗<br />

a 2 3 ) + 1 if p ≡ 1 (mod 3) ,<br />

1 if p ≡ 2 (mod 3) .<br />

Proof. As we have good reduction, the trace of Frob p on Pic(X (a 0,...,a 3 )<br />

)<br />

is the same as that of Frob on Pic(X (a 0,...,a 3 )<br />

) ⊗. Further, for the number<br />

p<br />

of <strong>points</strong> on a non-singular cubic surface over a finite field, the Lefschetz trace<br />

formula can be made completely explicit [Man,кIV,ÌÓÖѺ½]. It shows<br />

#X (a 0, ... ,a 3 ) (p) = p 2 + p · tr ( Frob | Pic(X (a ⊗) 0, ... ,a 3 )<br />

) + 1 .<br />

p<br />

The explicit formulas for the numbers of <strong>points</strong> given in Fact 3.2 therefore yield<br />

the assertion.<br />

□<br />

√<br />

3.4. Notation. –––– For A an integer, denote the fieldÉ(ζ 3 , 3 A) by K.<br />

Further, let G := Gal(K/É), H := Gal(K/É(ζ 3 )), <strong>and</strong> χ: H → 〈ζ 3 〉 ⊂∗<br />

be a primitive character. Then, we write ν K := ind G H<br />

(χ) for the induced character<br />

<strong>and</strong> V K for the corresponding G-representation.


308 ON THE SMALLEST POINT ON A DIAGONAL CUBIC SURFACE [Chap. X<br />

If K is of degree three overÉ(ζ 3 ) then V K is an irreducible rank two representation<br />

of G ∼ = S 3 . Otherwise, K =É(ζ 3 ). Then, V K ∼ =⊕M splits<br />

into the direct sum of a trivial <strong>and</strong> a non-trivial one-dimensional representation<br />

of H ∼ =/2.<br />

We will freely consider V K as a Gal(É/É)-representation.<br />

3.5. Lemma. –––– Let A be any integer. Then, for a prime p not dividing A,<br />

we have<br />

{<br />

νÉ(ζ 3 , 3√ A) χ3 (A) + χ<br />

(Frob p ) =<br />

3 (A) if p ≡ 1 (mod 3) ,<br />

0 if p ≡ 2 (mod 3) .<br />

Proof. The primitive character is unique up to conjugation by an element of G.<br />

Therefore, the induced character λ is well-defined.<br />

The Kummer pairing allows to make a definite choice for χ as follows. Fix an<br />

embedding σ :É(ζ 3 ) →. Then, put χ(g) := σ ( g( 3 √ A)/<br />

3√ A<br />

)<br />

.<br />

If p ≡ 2 (mod 3) then p remains prime inÉ(ζ 3 ). This means, Frob p acts<br />

non-trivially onÉ(ζ 3 ), i.e. Frob p ∈ G\H . Since H is a normal subgroup in G,<br />

the induced character vanishes on such an element.<br />

For p ≡ 1 (mod 3), we have that (p) splits inÉ(ζ 3 ). Let us write (p) = pp.<br />

The choice of p is equivalent to the choice of a homomorphism ι: 〈ζ 3 〉 →∗ p .<br />

The Frobenius Frob p is determined only up to conjugation, we may choose<br />

Frob p = Frob p ∈ H . Then, directly by the definition of an induced<br />

character, νÉ(ζ 3 , 3√ A) (Frob p ) = χ(Frob p ) + χ(Frob p ). We need to show<br />

that χ(Frob p ) = χ 3 (A) or χ(Frob p ) = χ 3 (A).<br />

For this, by the choice made above, we have<br />

χ(Frob p ) := σ ( Frob p ( 3 √<br />

A)/<br />

3√<br />

A<br />

)<br />

.<br />

After reduction modulo p, we may write<br />

Frob( 3 √<br />

A)/<br />

3√<br />

A = (<br />

3√<br />

A) p / 3 √<br />

A = A<br />

p−1<br />

3 .<br />

√<br />

Therefore, Frob p ( 3 A)/<br />

3√ A = ι −1 (A p−1<br />

3 ) which shows<br />

É<br />

χ(Frob p ) = σ ( ι −1 (A p−1<br />

3 ) ) .<br />

É<br />

That final formula is a definition for a cubic residue character at A. □<br />

3.6. Theorem. –––– Let a 0 , . . ., a 3 ∈\{0}. Then, the Gal(É/É)-representation<br />

Pic(X (a 0,...,a 3 )<br />

) ⊗splits into the direct sum<br />

Pic ( X (a 0,...,a 3<br />

)<br />

) ⊗∼ =⊕V K 1<br />

⊕ V K 2<br />

⊕ V K 3


Sec. 3] SPLITTING THE PICARD GROUP 309<br />

√<br />

where K 1 , K 2 , <strong>and</strong> K 3 denote the fieldsÉ(ζ 3 , 3 a0 a 1 a 2 2 a2 3 ),É(ζ √<br />

√ 3 , 3 a0 a 2 1 a 2a 2 3 ),<br />

<strong>and</strong>É(ζ 3 , 3 a0 a 2 1 a2 2 a É<br />

3), respectively.<br />

Proof. We will show that the representations on both sides have the same character.<br />

For that, by virtue of the Chebotarev density theorem, it suffices to<br />

consider the values at the Frobenii Frob p for p ∤ 3a 0 · . . . · a 3 .<br />

For the representation on the left h<strong>and</strong><br />

É<br />

side, χ<br />

(a Pic(X 0 ,...,a 3 )<br />

)(Frob p ) has been computed<br />

in Lemma 3.3. For the representation on the right h<strong>and</strong> side, Lemma 3.5<br />

shows that exactly the same formula is true.<br />

□<br />

3.7. Corollary. –––– Let a 0 , . . . , a 3 ∈\{0} be integers,<br />

P := Pic ( X (a 0, ... ,a 3<br />

)<br />

) ⊗,<br />

considered as a Gal(É/É)-representation,<br />

√<br />

<strong>and</strong> denote by χ P be the associated<br />

character. Put K 1 :=É(ζ 3 , 3 a0 a 1 a 2 2 a2 3 ), K √<br />

√ 2 :=É(ζ 3 , 3 a0 a 2 1 a 2a 2 3<br />

), <strong>and</strong><br />

K 3 :=É(ζ 3 , 3 a0 a 2 1 a2 2 a 3). Then, for the Artin conductor N χP of χ P , we have<br />

where<br />

N 2 χ P<br />

= D(K 1 ) D(K 2 ) D(K 3 )/(−27) ,<br />

D(K ) :=<br />

{ Disc(K/É) if [K :É(ζ 3 )] = 3 ,<br />

−27 if K =É(ζ 3 ) .<br />

Proof. We have to show N 2 = D(K )/(−3). Assume first that [K :É(ζ<br />

ν K 3 )] = 3.<br />

Then, the conductor-discriminant formula [Ne, Chapter VII, Section (11.9)]<br />

shows Disc(K/É) = NN M N 2 <strong>and</strong> −3 = Disc(É(ζ<br />

ν K 3 )/É) = NN M which<br />

together yield the assertion. In the opposite case, we have V K =⊕M<br />

<strong>and</strong> N ν K = NN M = −3.<br />

□<br />

3.8. Lemma. –––– Let a <strong>and</strong> b be integers different from zero. Then,<br />

Proof. We have, at first,<br />

|Disc (É(ζ 3 , 3 √<br />

ab2 )/É) | ≤ 3 9 a 4 b 4 .<br />

| Disc (É(ζ 3 , 3 √<br />

ab2 )/É) | ≤ | Disc<br />

(É(ζ 3 )/É) |3 · Disc (É( 3 √<br />

ab2 )/É) 2<br />

= 27 · Disc (É( 3 √<br />

ab2 )/É) 2.<br />

Further, by [Mc, Chapter 2, Exercise 41], we know<br />

|Disc (É( 3 √<br />

ab2 )/É) | ≤ 3 3 a 2 b 2 .<br />

This shows<br />

|Disc (É(ζ 3 , 3 √<br />

ab2 )/É) | ≤ 3 9 a 4 b 4 .<br />


É<br />

310 ON THE SMALLEST POINT ON A DIAGONAL CUBIC SURFACE [Chap. X<br />

3.9. Corollary. –––– Let a 0 , . . . , a 3 ∈\{0} be integers,<br />

P := Pic ( X (a 0, ... ,a 3<br />

)<br />

) ⊗,<br />

considered as a Gal(É/É)-representation, <strong>and</strong> χ P be the associated character.<br />

Then, for the Artin conductor N χP of χ P , we have the estimate<br />

| N χP | ≤ 3 12 (a 0 · . . . · a 3 ) 6 .<br />

Proof. Lemma 3.8 shows | D(K i )| ≤ 3 9 (a 0 · . . . · a 3 ) 4 for i = 1, 2, <strong>and</strong> 3.<br />

The assertion follows immediately from this.<br />

□<br />

4. A technical lemma<br />

4.1. Sublemma. –––– a) (Good reduction)<br />

If p ∤ 3a 0 · . . . · a 3 then the sequence ( #X (a 0,...,a 3<br />

n∈Æ<br />

) (/p n)/p2n) n∈Æis constant.<br />

b) (Bad reduction)<br />

i) If p divides a 0· . . . ·a 3 but not 3 then the sequence ( #X (a 0,...,a 3 ) (/p n)/p2n) becomes stationary as soon as p n does not divide any of the coefficients a 0 , . . . , a 3 .<br />

ii) If p = 3 then the sequence ( #X (a 0,...,a 3 ) (/p n)/p2n) n∈Æbecomes stationary as<br />

soon as 3 n does not divide any of the numbers 3a 0 , . . . , 3a 3 .<br />

4.2. Lemma. –––– There are two positive constants C 1 <strong>and</strong> C 2 such that, for<br />

all a 0 , . . ., a 3 ∈\{0},<br />

Proof. Cf. Lemma VI.5.34.<br />

C 1 < ∏<br />

p prime<br />

p∤3a 0·····a 3<br />

τ p<br />

(<br />

X<br />

(a 0 , ... ,a 3 ) (Ép) ) < C 2 .<br />

For a prime p of good reduction, Sublemma 4.1 shows<br />

τ p<br />

(<br />

X<br />

(a 0 , ... ,a 3 ) (Ép) ) = det ( 1 − p −1 Frob p | Pic(XÉ) ) · #X (a 0, ... ,a 3 ) (p)<br />

p 2 .<br />

Further, for the number of <strong>points</strong> on a non-singular cubic surface over a finite<br />

field, the Lefschetz trace formula can be made completely explicit [Man,<br />

кIV,ÌÓÖѺ½]. It shows<br />

#X (a 0, ... ,a 3 ) (p) = p 2 + p · tr ( Frob p | Pic(XÉ) ) + 1 .


Sec. 5] A NEGATIVE RESULT 311<br />

Denoting the eigenvalues of the Frobenius on Pic(XÉ) by λ 1 , . . . , λ 7 , we find<br />

τ p<br />

(<br />

X<br />

(a 0 , ... ,a 3 ) (Ép) )<br />

= (1 − λ 1 p −1 )(1 − λ 2 p −1 ) · . . . · (1 − λ 7 p −1 )<br />

· [1 + (λ 1 + · · · + λ 7 )p −1 + p −2 ]<br />

= (1 − σ 1 p −1 + σ 2 p −2 ∓ . . . − σ 7 p −7 )(1 + σ 1 p −1 + p −2 )<br />

= 1 + (1 − σ 2 1 + σ 2 )p −2 − (σ 1 − σ 1 σ 2 + σ 3 )p −3 ±<br />

± . . . − (σ 5 − σ 1 σ 6 + σ 7 )p −7 + (σ 6 − σ 1 σ 7 )p −8 − σ 7 p −9<br />

where σ i denote the elementary symmetric functions in λ 1 , . . . , λ 7 .<br />

We know |λ i | = 1 for all i. Estimating very roughly, we have |σ j | ≤ ( 7 j ) ≤ 7j<br />

<strong>and</strong> see<br />

1 − 99p −2 − 7·99p −3 − . . . − 7 7·99p −9 ≤ τ p<br />

(<br />

X<br />

(a 0 , ... ,a 3 ) (Ép) ) ≤<br />

≤ 1 + 99p −2 + 7·99p −3 + . . . + 7 7 · 99p −9 .<br />

I.e., 1 − 99p −2 1 < τ (<br />

1−7/p p X<br />

(a 0 , ... ,a 3 ) (Ép) ) < 1 + 99p −2 1<br />

1−7/p<br />

product over all 1 − 99p −2 1<br />

1−7/p<br />

The left h<strong>and</strong> side is positive for p > 13.<br />

(respectively 1 + 99p−2<br />

1<br />

1−7/p<br />

. The infinite<br />

) is convergent.<br />

For the small primes remaining,<br />

we need a better lower bound. For this, note that a cubic<br />

surface over<br />

(<br />

a finite fieldp always has at least onep-<strong>rational</strong> point.<br />

This yields τ p X<br />

(a 0 , ... ,a 3 ) (Ép) ) ≥ (1 − 1/p) 7 /p 2 > 0.<br />

□<br />

4.3. Remark. –––– The correctional factors det ( )<br />

1 − p −1 Frob p | Pic(XÉ) I p<br />

are all positive. Indeed, for a pair of complex conjugate eigenvalues, we have<br />

(1 − λp −1 )(1 − λp −1 ) = |1 − λp −1 | 2 > 0 <strong>and</strong> an eigenvalue of 1 or (−1)<br />

contributes a factor 1 ± p −1 > 0. Consequently, we always have<br />

(<br />

1 − 1 p) 7<br />

< det<br />

(<br />

1 − p −1 Frob p | Pic(XÉ) I p<br />

) <<br />

(<br />

1 + 1 p) 7.<br />

5. A negative result<br />

5.1. –––– For an integer q ≠ 0, let X (q) ⊂ P 3Ébe the cubic surface given by<br />

qx 3 + 4y 3 + 2z 3 + w 3 = 0 <strong>and</strong> let<br />

m(X (q) ) := min { H(x : y : z : w) | (x : y : z : w) ∈ X (q) (É)}<br />

be the smallest height of aÉ-<strong>rational</strong> point on X (q) . We compare m(X (q) ) with<br />

the Tamagawa type number τ (q) := τ (X (q) ).


312 ON THE SMALLEST POINT ON A DIAGONAL CUBIC SURFACE [Chap. X<br />

5.2. Lemma. –––– There is a constant C with the following property. For each<br />

pair (a, b)of natural numbers satisfying gcd(a, b) = 1, thereare two prime numbers<br />

p 1 , p 2 ≡ a (mod b), p 1 ≠ p 2 , such that p 1 , p 2 < C ·b 5.5 .<br />

Proof. Linnik’s Theorem in the version of R. Heath-Brown [H-B92b] shows<br />

there is one prime with the property stated. To get a second one, the following<br />

simple trick helps.<br />

One of the numbers a <strong>and</strong> a + b is relatively prime to 4b. Assume without<br />

restriction that gcd(a, 4b) = 1. Then, we find p 1 ≡ a (mod 4b) <strong>and</strong><br />

p 2 ≡ a + 2b (mod 4b) such that p 1 , p 2 < ˜C ·(4b) 5.5 .<br />

□<br />

5.3. Theorem. –––– Assume the Generalized Riemann Hypothesis. Then, there is<br />

no constant C such that<br />

m(X (q) ) < C<br />

τ (q)<br />

for all q ∈\{0}.<br />

Proof. We will construct a sequence (q i ) i∈Æof prime numbers such that<br />

q i ≡ 1 (mod 72) <strong>and</strong> m(X (qi) ) τ (q i)<br />

→ ∞ for i → ∞. The proof will consist<br />

of several steps.<br />

Firststep. The prime q i divides exactly one of the four coefficients in the equation<br />

defining X (qi) . In this case, it is known by the work of J.-L. Colliot-Thélène <strong>and</strong><br />

his coworkers [CT/K/S, Proof of Proposition 2] that precisely<br />

τ H<br />

(<br />

X (É)<br />

It is therefore sufficient to verify that<br />

Br) = 1 3 τ H(<br />

X (É)<br />

) .<br />

m(X (qi) ) · α(X (qi) )·β(X (qi) )·lim (s −1) t L(s, χ )· (<br />

(qi)<br />

s→1 Pic(X ∏ τ<br />

É) ν X<br />

(q i ) (Éν) ) → ∞.<br />

ν∈Val(É)<br />

Second step. If q is a prime different from 2 then we have rkPic(X (q) ) = 1.<br />

[Man,кIII,ÍÔÖÒÒº½¾] shows that X (q) is a minimal cubic surface<br />

overÉ. By [Man,кIV,ÌÓÖѺ½], we have Pic(X (q) ) =.<br />

Third step. We have α(X (qi) ) = 1 <strong>and</strong> β(X (qi) ) = 3.<br />

α(X (qi) ) = 1 follows immediately from rk Pic(X (qi) ) = 1. β(X (qi) ) can<br />

be computed by the method indicated in Remark 2.4. We work with a<br />

group G ⊂ W (E 6 ) of order 18 which decomposes the 27 lines into three<br />

orbits of nine lines each. An easy calculation shows<br />

Hom ( (NF ∩ F 0 )/NF 0 ,É/) ∼=/3.<br />

Fourth step. For the height of the smallest point, we have m(X (q) ) ≥ 3 √<br />

q<br />

7 .


Sec. 5] A NEGATIVE RESULT 313<br />

There are no <strong>rational</strong> solutions of the equation 4y 3 + 2z 3 + w 3 = 0 as<br />

this is impossible, 2-adically. √ |x| ≥ 1 yields |4y 3 + 2z 3 + w 3 | ≥ q <strong>and</strong><br />

max{|y|, |z|, |w|} ≥ q 3 . 7<br />

Fifth step. For |q| ≥ 7, one has τ ∞<br />

(<br />

X (q) (Ê) ) = 1<br />

3√<br />

|q|<br />

I where I is independent<br />

of q.<br />

By Lemma 6.15, we have<br />

τ ∞ (X (q) ) = 1<br />

4 3 √<br />

|q|<br />

Z<br />

CX (1, ... ,1) (Ê)<br />

|x 0 |≤ 3√ |q|,|x 1 |≤ 3√ 4,|x 2 |≤ 3√ 2,|x 3 |≤1<br />

ω CX (1, ... ,1) (Ê)<br />

Leray .<br />

Since |x 1 | ≤ 3 √<br />

4, |x2 | ≤ 3 √<br />

2, |x3 | ≤ 1, <strong>and</strong><br />

|x 0 | = 3 √<br />

| x 3 1 + x3 2 + x3 3 | ≤ 3 √| x 1 | 3 + | x 2 | 3 + | x 3 3 | ,<br />

the condition |x 0 | ≤ 3 √<br />

|q| is actually empty.<br />

(<br />

Sixth step. There is a positive constant C such that ∏ τ p X (q) (Ép) ) > C for<br />

p prime<br />

every prime q ≡ 1 (mod 72).<br />

By Lemma 4.2, we have C 1 > 0 such that<br />

∏<br />

p prime<br />

p≠2,3,q<br />

τ p<br />

(<br />

X (q) (Ép) ) > C 1 .<br />

It, therefore, remains to give lower bounds for the factors τ 2<br />

(<br />

X (q) (É2) ) ,<br />

τ 3<br />

(<br />

X (q) (É3) ) , <strong>and</strong> τ q<br />

(<br />

X (q) (Éq) ) .<br />

As 2 ∤q, by virtue of Sublemma 4.1 we have, τ 2<br />

(<br />

X (q) (É2) ) = 1 2 7 · #X (q) (/8)<br />

64<br />

.<br />

Further,<br />

#X (q) (/8) ≥ 1<br />

since q ≡ 1 (mod 8) implies (1 : 0 : 0 : (−1)) ∈ X (q) (/8).<br />

Similarly, τ 3<br />

(<br />

X (q) (É3) ) = ( 2 3 )7 · #X (q) (/9)<br />

81<br />

. Again, q ≡ 1 (mod 9) makes sure<br />

that (1 : 0 : 0 : (−1)) ∈ X (q) (/9) <strong>and</strong> #X (q) (/9) ≥ 1.<br />

For the prime q, we argue a bit differently. First,<br />

det ( 1 − q −1 Frob p | Pic(X É) (q) ) I q ≥ (1 − 1/q) 7 ≥ (72/73) 7 .<br />

Furthermore, the reduction of X (q) modulo q is the cone over the elliptic<br />

curve given by 4y 3 + 2z 3 + w 3 = 0. Therefore, on X (q) there are at<br />

least (q − 2 √ q + 1)(q − 1) smooth <strong>points</strong> defined overq. As Hensel’s lemma


314 ON THE SMALLEST POINT ON A DIAGONAL CUBIC SURFACE [Chap. X<br />

may be applied to them, we get<br />

#X (q) (/q n)<br />

lim<br />

≥ (q − 2√ q + 1)(q − 1)<br />

(<br />

> 1 − 2 )(<br />

√ 1 − 1 )<br />

n→∞ q 2n q 2 q q<br />

≥ 72 (<br />

1 − √ 2 ).<br />

73 73<br />

Seventh step. There is a sequence (q i ) i∈Æof primes such that q i ≡ 1 (mod 72)<br />

<strong>and</strong> [lim (s − 1)L(s, χ (qi)<br />

s→1 Pic(X<br />

)] → ∞ for i → ∞.<br />

É)<br />

Since rk Pic(X (qi) ) = 1, the representation Pic ( É)<br />

X (q i) ⊗contains exactly one<br />

trivial summ<strong>and</strong>. Hence,<br />

L ( )<br />

s, χ (qi) Pic(X = ζ(s) · L(s, χP (qi))<br />

É)<br />

for a representation P (q i) not containing trivial components. Our goal is, therefore,<br />

to show L(1, χ P (qi)) → ∞ for i → ∞.<br />

DenotebyP i thei-thprimenumber p suchthat p ≡ 1 (mod 3). For eachi ∈Æ,<br />

choose a cubic residue r i modulo P i . Then, define q i to be the second smallest<br />

prime number such that<br />

q i ≡ r j (mod P j )<br />

for all j ∈ {1, . . . , i} <strong>and</strong> q i ≡ 1 (mod 72).<br />

This system of simultaneous congruences is solvable by the Chinese Remainder<br />

Theorem. Its solutions form an arithmetic progression with common difference<br />

72P 1 · . . . · P i . By Chebyshev’s inequalities, we know<br />

72P 1 · . . . · P i ≤ 72e θ(P i) < 72e (2 log2)P i<br />

.<br />

According to our definition, q i is the second smallest prime in this arithmetic<br />

progression. From this, we clearly have that q i > 72P 1 · . . . · P i → ∞<br />

for i → ∞. On the other h<strong>and</strong>, Lemma 5.2 shows<br />

for certain constants C 1 <strong>and</strong> C 2 .<br />

q i ≤ C 1 · (72e (2 log2)P i<br />

) 5.5 = C 2 e (11 log2)P i<br />

Corollary 3.9 gives us an estimate for the Artin conductor of the character χ P (qi).<br />

We see N χP ≤ 3 12 (a<br />

(qi) 0 · . . . · a 3 ) 6 = 3 12 8 6 qi 6 ≤ C 3 e (66 log2)P i<br />

for another constant<br />

C 3 . Consequently,<br />

logN χP (qi )<br />

≤ (66 log2)P i + logC 3 .<br />

We observe that (log N χP (qi) )1/2 ≤ P i for i sufficiently large. We assume from<br />

now on that this inequality is fulfilled.


Sec. 5] A NEGATIVE RESULT 315<br />

Recall that P (q i) is actually the direct sum of representations which are induced<br />

from one-dimensional characters (Theorem 3.6). By consequence, it is known<br />

that the Artin L-function L( · , χ P (qi)) is entire. Since we also assume the<br />

Generalized Riemann Hypothesis, we may apply the estimate of W. Duke [Duk,<br />

Proposition 5]. It shows<br />

Here,<br />

logL(1, χ P (qi)) =<br />

∑<br />

p


316 ON THE SMALLEST POINT ON A DIAGONAL CUBIC SURFACE [Chap. X<br />

| D(K 3 )| = |Disc (É(ζ 3 , 3 √<br />

qi )/É) |<br />

= Disc (É( 3 √<br />

qi )/É)2<br />

· |N ( Disc (É(ζ 3 , 3 √<br />

qi )/É( 3 √<br />

qi ) )) |<br />

≥ Disc (É( 3 √<br />

qi )/É) 2.<br />

According to [Mc, Chapter 2, Exercise 41], we know<br />

| Disc (É( 3 √<br />

qi )/É) | ≥ 3q<br />

2<br />

i . □<br />

5.4. Remark. –––– Note that the estimate for L(1, χ P (qi)) is the only point where<br />

we use the Generalized Riemann Hypothesis.<br />

6. The fundamental finiteness property<br />

6.1. –––– In this section, we return to the case of a general diagonal cubic<br />

surface X (a 0, ... ,a 3 )<br />

⊂ P 3Égiven by a 0 x 3 0 + . . . + a 3x 3 3 = 0. Our goal is to<br />

establish the estimate for τ (a 0, ... ,a 3 ) := τ (X (a 0, ... ,a 3 ) ) formulated as Theorem 1.7.<br />

For this, in the subsections below, we will give an individual estimate for each<br />

of the factors occurring in the definition of τ (X (a 0, ... ,a 3 ) ).<br />

i. An estimate for the L-factor. —<br />

É<br />

)<br />

6.2. Proposition. –––– For each ε > 0, there exist positive constants c 1 <strong>and</strong> c 2<br />

such that<br />

c 1 · |a 0 · . . . · a 3 | −ε < lim (s − 1) t L ( s, χ<br />

s→1<br />

É<br />

)<br />

(a Pic(X 0 , ... ,a 3 ) < c2 · |a 0 · . . . · a 3 | ε<br />

for all (a 0 , . . . , a 3 ) ∈ (\{0}) 4 . Here, t = rk Pic(X ).<br />

Proof. The Galois representation Pic(X (a 0, ... ,a 3 )<br />

) ⊗contains the trivial representation<br />

t times as a direct summ<strong>and</strong>. Therefore,<br />

L ( )<br />

s, χ (a Pic(X 0 , ... ,a 3 ) = ζ(s)t · L(s, χ P )<br />

É<br />

)<br />

where ζ denotes the Riemann zeta function <strong>and</strong> P is a representation which<br />

does not contain trivial components. All we need to show is<br />

c 1 · |a 0 · . . . · a 3 | −ε < L(1, χ P ) < c 2 · |a 0 · . . . · a 3 | ε .<br />

L( · , χ P ) is the product of at most three factors of the form L( · , λ) where λ is<br />

the non-trivial Dirichlet character ofÉ(ζ 3 )/É<strong>and</strong> at most three factors which<br />

are Artin-L-functions L( · , ν K ) for K a purely cubic field extension ofÉ(ζ 3 )


Sec. 6] THE FUNDAMENTAL FINITENESS PROPERTY 317<br />

as above, sayK =É( ζ3 , 3 √<br />

a0 a 1 a 2 2 a2 3)<br />

. AsL(1, λ)does not depend on a0 , . . . , a 3 ,<br />

at all, it will suffice to show<br />

for each ε > 0.<br />

c 1 (ε) · |a 0 · . . . · a 3 | −ε < L(1, ν K ) < c 2 (ε) · |a 0 · . . . · a 3 | ε<br />

V K is the only irreducible two-dimensional representation of Gal(K/É) ∼ = S 3 .<br />

For that reason, by virtue of [Ne, Chapter VII, Corollary (10.5)], we have<br />

ζ K (s) = ζÉ(s) · L(s, λ) · L(s, ν K ) 2<br />

= ζÉ(ζ 3 )(s) · L(s, ν K ) 2<br />

for a complex variable s. It, therefore, suffices in our particular situation to<br />

estimate the residue res s=1 ζ K (s) of the Dedekind zeta function of K.<br />

An estimate from above has been given by C. L. Siegel. In view of the analytic<br />

class number formula, his [Si69, Satz 1] gives<br />

res ζ K (s) < C (log Disc(K/É)) 5<br />

s=1<br />

≤ C [log(3 9 a 4 0a 4 1a 4 2a 4 3)] 5<br />

= C [4 log |a 0 · . . . · a 3 | + 9 log 3] 5<br />

for a certain constant C. The final term is less than c 2 (ε) · |a 0 · . . . · a 3 | ε for<br />

every ε > 0.<br />

On the other h<strong>and</strong>, H. M. Stark [St, formula (1)] shows<br />

res ζ K (s) > C (ε)·Disc(K/É) −ε/4<br />

s=1<br />

for every ε > 0 which implies res<br />

s=1 ζ K (s) > c 1 (ε)·|a 0 · . . . · a 3 | −ε . □<br />

ii. An estimate for the factors at the finite places. —<br />

6.3. Notation. –––– We will adopt the notation from Chapter VIII, 3.1. More<br />

precisely,<br />

i) For a prime number p <strong>and</strong> an integer x ≠ 0, we put x (p) := p ν p(x) .<br />

Note x (p) = 1/‖x‖ p for the normalized p-adic valuation.<br />

ii) By putting ν(x) := min<br />

ξ∈p<br />

to/p r.<br />

x=(ξ mod p r )<br />

ν(ξ), we carry the p-adic valuation fromp over<br />

Note that any 0 ≠ x ∈/p rhas the form x = ε·p ν(x) where ε ∈ (/p r) ∗ is<br />

a unit. Clearly, ε is unique only in the case ν(x) = 0.


318 ON THE SMALLEST POINT ON A DIAGONAL CUBIC SURFACE [Chap. X<br />

6.4. Definition. –––– For (a 0 , . . . , a 3 ) ∈4 , r ∈Æ, <strong>and</strong> ν 0 , . . . , ν 3 ≤ r, put<br />

N (r)<br />

ν 0 , ... ,ν 3 ;a 0 , ... ,a 3<br />

:= { (x 0 , . . . ,x 3 ) ∈ (/p r) 4 |<br />

ν(x 0 ) = ν 0 , . . . ,ν(x 3 ) = ν 3 ; a 0 x 3 0 + . . . + a 3 x 3 3 = 0 ∈/p r}.<br />

For the particular case ν 0 = . . . = ν 3 = 0, we will write<br />

I.e.,<br />

Z (r)<br />

a 0 , ... ,a 3<br />

:= N (r)<br />

0, ... ,0;a 0 , ... ,a 3<br />

.<br />

Z (r)<br />

a 0 , ... ,a 3<br />

= { (x 0 , . . . , x 3 ) ∈ [(/p r) ∗ ] 4 | a 0 x 3 0 + . . . + a 3 x 3 3 = 0 ∈/p r}.<br />

We will use the notation z (r)<br />

a 0 , ... ,a 3<br />

:= #Z (r)<br />

a 0 , ... ,a 3<br />

.<br />

6.5. Sublemma. –––– If p k |a 0 , . . . , a 3 <strong>and</strong> r > k then we have<br />

z (r)<br />

a 0 , ... ,a 3<br />

= p 4k · z (r−k)<br />

a 0 /p k , ... ,a 3 /p k .<br />

Proof. Since a 0 x 3 0 + . . . + a 3 x 3 3 = p k (a 0 /p k · x 3 0 + . . . + a 3 /p k · x 3 3 ), there is a<br />

surjection<br />

ι: Z a (r)<br />

0 , ... ,a 3<br />

−→ Z (r−k) ,<br />

a 0 /p k , ... ,a 3 /p k<br />

given by (x 0 , . . . , x 3 ) ↦→ ( (x 0 mod p r−k ), . . . , (x 3 mod p r−k ) ) . The kernel of<br />

the homomorphism of modules underlying ι is (p r−k/pr) 4 . □<br />

6.6. Lemma. –––– Assume gcd p<br />

(a 0 , . . . , a 4 ) = p k . Then, there is an estimate<br />

z (r)<br />

a 0 , ... ,a 4<br />

≤ 3p 3r+k .<br />

Proof. Suppose first that k = 0. This means, one of the coefficients is prime<br />

to p. Without restriction, assume p ∤a 0 .<br />

For any (x 1 , x 2 , x 3 ) ∈ (/p r) 3 , there appears an equation of the form a 0 x 3 0 = c.<br />

It cannot have more than three solutions in (/p r) ∗ . Indeed, for p odd, this<br />

follows directly from the fact that (/p r) ∗ is a cyclic group. On the other<br />

h<strong>and</strong>, in the case p = 2, we have (/2 r) ∗ ∼ =/2 r−2×/2. Again, there<br />

are only up to three solutions possible.<br />

The general case may now easily be deduced from Sublemma 6.5. Indeed, if<br />

k < r then<br />

z (r)<br />

a 0 , ... ,a 3<br />

= p 4k · z (r−k)<br />

a 0 /p k , ... ,a 3 /p k ≤ p 4k · 3p 3(r−k) = 3p 3r+k .<br />

On the other h<strong>and</strong>, if k ≥ r then the assertion is completely trivial since<br />

z (r)<br />

a 0 , ... ,a 3<br />

= #Z (r)<br />

a 0 , ... ,a 3<br />

< p 4r ≤ p 3r+k < 3p 3r+k .<br />


Sec. 6] THE FUNDAMENTAL FINITENESS PROPERTY 319<br />

6.7. Remark. –––– The proof shows that in the case p ≡ 2 (mod 3) one could<br />

reduce the coefficient to 1. Unfortunately, this observation does not lead to a<br />

substantial improvement of our final result.<br />

6.8. Lemma. –––– Let r ∈Æ<strong>and</strong> ν 0 , . . . , ν 3 ≤ r. Then,<br />

#N ν (r)<br />

0 , ... ,ν 3 ;a 0 , ... ,a 3<br />

= z(r) · ϕ(p r−ν 0<br />

) · . . . · ϕ(p r−ν 3<br />

)<br />

p 3ν 0a 0 , ... ,p 3ν 3a 3<br />

.<br />

ϕ(p r ) 4<br />

Proof. As p 3ν 0 a0 x 3 0 + . . . + p 3ν 3 a3 x 3 3 = a 0 (p ν 0 x0 ) 3 + . . . + a 3 (p ν 3 x3 ) 3 , we have<br />

a surjection<br />

π : Z (r) −→ N (r)<br />

p 3ν 0a 0 , ... ,p 3ν 3a 3<br />

ν 0 , ... ,ν 3 ;a 0 , ... ,a 3<br />

,<br />

given by (x 0 , . . . , x 3 ) ↦→ (p ν 0 x0 , . . . , p ν 3 x3 ).<br />

For i = 0, . . . , 3, consider the mapping ι:/p r→/pr, x ↦→ p ν i<br />

x.<br />

If ν i = r then ι is the zero map. All ϕ(p r ) = (p − 1)p r−1 units are mapped<br />

to zero. Otherwise, observe that ι is p ν i<br />

: 1 on its image. Further, ν(ι(x)) = ν i<br />

if <strong>and</strong> only if x is a unit. By consequence, π is (K (ν 0)<br />

· . . . · K (ν3) ) : 1 when<br />

we put K (ν) := p ν for ν < r <strong>and</strong> K (r) := (p − 1)p r−1 . Summarizing, we could<br />

have written K (ν) := ϕ(p r )/ϕ(p r−ν ). The assertion follows.<br />

□<br />

6.9. Corollary. –––– Let (a 0 , . . . , a 3 ) ∈ (\{0}) 4 . Then, for the local factor<br />

τ p<br />

(<br />

X<br />

(a 0 , ... ,a 3 ) (Ép) ) , one has<br />

(<br />

τ p X<br />

(a 0 , ... ,a 3 ) (Ép) ) = det ( )<br />

1 − p −1 Frob p | Pic(XÉ) I p<br />

r<br />

r→∞<br />

∑<br />

ν 0 ,...,ν 3 =0<br />

· lim<br />

z (r) · ϕ(p r−ν p 3ν 0a 0 , ... ,p 3ν 0<br />

3a ) · . . . · ϕ(p<br />

r−ν 3)<br />

3<br />

.<br />

p 3r · ϕ(p r ) 4<br />

Proof. By Remark VI.5.19, we have<br />

τ p<br />

(<br />

X<br />

(a 0 , ... ,a 3 ) (Ép) ) = det ( 1 − p −1 Frob p | Pic(XÉ) I p )<br />

Lemma 6.8 yields the assertion.<br />

r<br />

#N ν<br />

r→∞<br />

∑<br />

(r)<br />

0 ,...,ν 3 ;a 0 ,...,a 3<br />

ν 0 ,...,ν 3 =0<br />

· lim<br />

p 3r .<br />

□<br />

6.10. Proposition. –––– Let (a 0 , . . . , a 3 ) ∈ (\{0}) 4 . Then, for each ε such<br />

that 0 < ε < 1 , one has<br />

3<br />

(<br />

τ p X<br />

(a 0 , ... ,a 3 ) (Ép) ) (<br />

≤ 1+ 1 ) (<br />

1<br />

)( 1<br />

) 3·(<br />

(p)<br />

) 1−ε<br />

7·3 ( a<br />

p 1 − 1 1 − 1 0 a(p) 1 a(p) 3<br />

2<br />

a (p) ) ε<br />

3 .<br />

p 1−3ε p ε


320 ON THE SMALLEST POINT ON A DIAGONAL CUBIC SURFACE [Chap. X<br />

Proof. We use the formula from Corollary 6.9. By Remark 4.3, the first factor<br />

is at most (1 + 1/p) 7 . Further, by Lemma 6.6,<br />

z (r)<br />

p 3ν 0a 0 , ... ,p 3ν 3a 3<br />

/p 3r ≤ 3 gcd p<br />

(p 3ν 0<br />

a 0 , . . . , p 3ν 3<br />

a 3 )<br />

(<br />

(p)) Writing k i := ν p (a i ) = ν p a i , we see<br />

= 3 gcd ( p 3ν 0<br />

a (p)<br />

0 , . . . , p3ν 3<br />

a (p)<br />

3<br />

z (r)<br />

p 3ν 0a 0 , ... ,p 3ν 3a 3<br />

/p 3r ≤ 3 gcd(p 3ν 0+k 0<br />

, . . . , p 3ν 3+k 3<br />

)<br />

= 3p min{3ν 0+k 0 , ... ,3ν 3 +k 3 } .<br />

We estimate the minimum by a weighted arithmetic mean with weights 1−ε<br />

1−ε<br />

, <strong>and</strong> ε,<br />

, 1−ε<br />

3 3<br />

min{3ν 0 + k 0 , . . . , 3ν 3 + k 3 }<br />

This shows<br />

≤ 1 − ε<br />

3<br />

· (3ν 0 + k 0 ) + 1 − ε · (3ν 1 + k 1 )<br />

3<br />

+ 1 − ε · (3ν 2 + k 2 ) + ε(3ν 3 + k 3 )<br />

3<br />

= (1 − ε)(ν 0 + ν 1 + ν 2 ) + 3εν 3<br />

)<br />

.<br />

+ 1 − ε (k 0 + k 1 + k 2 ) + εk 3 .<br />

3<br />

z (r) /p 3r ≤ 3p (1−ε)(ν 0+ν 1 +ν 2 )+3εν 3 + 1−ε<br />

p 3ν 0a 0 , ... ,p 3ν 3 (k 0+k 1 +k 2 )+εk 3<br />

3a 3<br />

(a<br />

= 3p (1−ε)(ν 0+ν 1 +ν 2 )+3εν3 (p)<br />

) 1−ε<br />

· ( 0 a(p) 1 a(p) 3<br />

2<br />

a (p) ) ε.<br />

3<br />

We may therefore write<br />

τ p<br />

(<br />

X<br />

(a 0 , ... ,a 3 ) (Ép) ) ≤<br />

r<br />

r→∞<br />

∑<br />

ν 0 , ... ,ν 3 =0<br />

· lim<br />

(<br />

1 + 1 ) 7 ( · 3 a<br />

(p)<br />

) 1−ε (<br />

0<br />

p<br />

a(p) 1 a(p) 3<br />

2<br />

a (p) ) ε<br />

3<br />

p (1−ε)(ν 0+ν 1 +ν 2 )+3εν 3 · ϕ(p<br />

r−ν 0) · . . . · ϕ(p<br />

r−ν 3)<br />

ϕ(p r ) 4 .<br />

Here, the term under the limit is precisely the product of three copies of the<br />

finite sum r<br />

p<br />

∑<br />

(1−ε)ν · ϕ(p r−ν r−1<br />

) 1<br />

=<br />

ν=0<br />

ϕ(p r )<br />

∑<br />

ν=0<br />

(p ε ) + p 1<br />

ν p − 1 (p ε ) r<br />

<strong>and</strong> one copy of the finite sum<br />

r<br />

p<br />

∑<br />

3εν · ϕ(p r−ν )<br />

ν=0<br />

ϕ(p r )<br />

=<br />

r−1<br />

∑<br />

ν=0<br />

1<br />

(p 1−3ε ) + p<br />

ν<br />

1<br />

p − 1 (p 1−3ε ) . r<br />

3 ,


Sec. 6] THE FUNDAMENTAL FINITENESS PROPERTY 321<br />

For r → ∞, geometric series do appear while the additional summ<strong>and</strong>s tend<br />

to zero.<br />

□<br />

6.11. Remark. –––– Unfortunately, the constants<br />

) 7 (<br />

C p (ε) := · 3<br />

(<br />

1 + 1 p<br />

1<br />

)( 1<br />

) 3<br />

1 − 1 1 − 1 p 1−3ε p ε<br />

diverges. On the other h<strong>and</strong>, we<br />

is bounded for p → ∞, say C p (ε) ≤ C (ε) .<br />

have the property that the product ∏ p C (ε)<br />

p<br />

have at least that C (ε)<br />

p<br />

6.12. Lemma. –––– Let C > 1 be any constant. Then, for each ε > 0, one has<br />

∏<br />

p prime<br />

p|x<br />

for a suitable constant c (depending on ε).<br />

C ≤ c · x ε<br />

Proof. This follows directly from [Nat, Theorem 7.2] together with [Nat,<br />

Section 7.1, Exercise 7].<br />

□<br />

6.13. Proposition. –––– For each ε such that 0 < ε < 1 , there exists a constant c<br />

3<br />

such that<br />

(<br />

∏τ p X<br />

(a 0 , ... ,a 3 ) (Ép) ) ≤ c · |a 0 · . . . · a 3 | 3 1 − ε 8 · ∏<br />

p prime<br />

for all (a 0 , . . . , a 3 ) ∈ (\{0}) 4 .<br />

p prime<br />

min ‖a i‖ 1 3 −ε<br />

p<br />

i=0, ... ,3<br />

Proof. The product over all primes of good reduction is bounded by virtue of<br />

Lemma 4.2.a). It, therefore, remains to show that<br />

∏<br />

p prime<br />

p|3a 0 ... a 3<br />

(<br />

τ p X<br />

(a 0 , ... ,a 3 ) (Ép) ) ≤ c · |a 0 · . . . · a 3 | 3 1 − ε 8 · ∏ min ‖a i‖ 1 3 −ε<br />

i=0, ... ,3<br />

For this, by Proposition 6.10, we have at first<br />

τ p<br />

(<br />

X<br />

(a 0 , ... ,a 3 ) (Ép) ) ≤ C (ε)<br />

p<br />

= C (ε)<br />

p<br />

·<br />

·<br />

p prime<br />

(a(p)<br />

) 1<br />

0 a(p) 1 a(p) 3 − ε4<br />

2<br />

· (a (p)<br />

(a(p)<br />

0 a(p) 1 a(p) 2 a(p) 3<br />

3 ) 3 4 ε<br />

) 1<br />

3<br />

− ε4 · (a (p)<br />

3 )− 1 3 +ε .<br />

Here, the indices 0, . . . , 3 are interchangeable. Hence, it is even allowed to<br />

write<br />

(<br />

τ p X<br />

(a 0 , ... ,a 3 ) (Ép) ) (a<br />

≤ C p<br />

(ε) (p)<br />

) 1<br />

·<br />

0 a(p) 1 a(p) 2 a(p) 3<br />

− ε4<br />

3<br />

· (max )<br />

a (p) −<br />

1<br />

3<br />

+ε<br />

i i<br />

)<br />

= C p<br />

(ε)<br />

1<br />

· 3<br />

− ε4 · min ‖a i ‖ 1 3 −ε<br />

p .<br />

(a(p)<br />

0 a(p) 1 a(p) 2 a(p) 3<br />

Now, we multiply over all prime divisors of a 0 · . . . · a 3 . Thereby, on the<br />

right h<strong>and</strong> side, we may twice write the product over all primes since the two<br />

i<br />

p .


322 ON THE SMALLEST POINT ON A DIAGONAL CUBIC SURFACE [Chap. X<br />

rightmost factors are equal to one for p ∤ 3a 0 · . . . · a 3 , anyway.<br />

(<br />

τ p X<br />

(a 0 , ... ,a 3 ) (Ép) )<br />

∏<br />

p prime<br />

p|3a 0 ... a 3<br />

≤ ∏ C p<br />

(ε)<br />

p prime<br />

p|3a 0 ... a 3<br />

(<br />

(p)<br />

) 1<br />

· ∏ a<br />

0 a(p) 1 a(p) 2 a(p) 3 − ε 4<br />

3<br />

· ∏<br />

p prime<br />

= ∏ C p (ε) · | a 0 · . . . · a 3 | 1 3 − ε 4 · ∏<br />

p prime<br />

p|3a 0 ... a 3<br />

p prime<br />

p prime<br />

min ‖a i‖ 1 3 −ε<br />

p<br />

i=0, ... ,3<br />

min ‖a 3<br />

i‖ 1 −ε<br />

p<br />

i=0, ... ,3<br />

when we observe that ∏ p a (p) = | a|. Further, we have C p<br />

(ε)<br />

Lemma 6.12,<br />

∏ C (ε) ≤ c · |3a 0 · . . . · a 3 | ε 8 .<br />

p prime<br />

p|3a 0 ... a 3<br />

We finally estimate 3 ε 8 by a constant. The assertion follows.<br />

iii. An estimate for the factor at the infinite place. —<br />

6.14. Corollary. –––– Let a 0 , . . . , a 3 ∈Ê\{0}. Then,<br />

[<br />

ω CX (a 0 , ... ,a 3 ) (Ê) 1<br />

Leray = dx<br />

3| a 0 |x 2 1 ∧ dx 2 ∧ dx 3<br />

].<br />

0<br />

Proof. We apply Lemma VI.5.26 to U =Ê4 <strong>and</strong><br />

f (x 0 , . . . , x 3 ) := a 0 x 3 0 + . . . + a 3 x 3 3 .<br />

≤ C (ε) <strong>and</strong>, by<br />

Note that {(x 0 , . . . , x 3 ) ∈ CX (a 0,...,a 3 ) (Ê) | x 0 = 0} is a zero set according to<br />

the Leray measure as it is for the hypersurface measure.<br />

□<br />

6.15. Lemma. –––– Let a 0 , . . . , a 3 ∈Ê\{0}. Then,<br />

(<br />

τ ∞ X<br />

(a 0 , ... ,a 3 ) (Ê) ) 1<br />

= √<br />

2 3 | a0 · . . . · a 3 |<br />

Z<br />

CX (1, ... ,1) (Ê)<br />

|x 0 |≤ 3√ |a 0 |, ... ,|x 3 |≤ 3√ |a 3 |<br />

ω CX (1, ... ,1) (Ê)<br />

Leray .<br />

(<br />

Proof. According to the definition of τ ∞ X<br />

(a 0 , ... ,a 3 ) (Ê) ) <strong>and</strong> the corollary above,<br />

we need to show<br />

Z<br />

1 1<br />

dx 1 ∧dx 2 ∧ dx 3<br />

6 |a 0 |<br />

x 2<br />

CX (a 0 , ... ,a 3 ) 0<br />

(Ê)<br />

|x 0 |≤1, ... ,|x 3 |≤1<br />

Z<br />

1<br />

1<br />

= √ dX<br />

6 3 | a0 · . . . · a 3 | X 2 1 ∧ dX 2 ∧ dX 3 .<br />

CX (1, ... ,1) 0<br />

(Ê)<br />

|X 0 |≤ 3√ |a 0 |, ... ,|X 3 |≤ 3√ |a 3 |<br />


Sec. 6] THE FUNDAMENTAL FINITENESS PROPERTY 323<br />

For that, consider the linear mapping l : CX (a 0, ... ,a 3<br />

√ √ ) (Ê) → CX (1, ... ,1) (Ê) given<br />

by (x 0 , . . . , x 3 ) ↦→ ( 3 a0 x 0 , . . . , 3 a3 x 3 ). Then,<br />

( 1<br />

) 3√<br />

l ∗ a1 a 2 a 3 1<br />

dX<br />

X0<br />

2 1 ∧ dX 2 ∧ dX 3 = dx<br />

x 2 1 ∧ dx 2 ∧ dx 3 .<br />

0<br />

This immediately yields the assertion when we take into consideration that<br />

orientations are chosen in such a way that both integrals are positive. □<br />

a 2/3<br />

0<br />

6.16. Proposition. –––– For real numbers 0 < b 0 ≤ b 1 ≤ b 2 ≤ b 3 , we have<br />

Z<br />

(<br />

ω CX (1, ... ,1) (Ê)<br />

Leray ≤ 64 + 64<br />

3 log 3 + 1 3√<br />

3ω2<br />

)b 0 + 64b 0 log b 1<br />

3<br />

b 0<br />

CX (1, ... ,1)(Ê)<br />

| x 0 |≤b 0 , ... ,|x 3 |≤b 3<br />

where ω 2 is the two-dimensional hypersurface measure on the l 3 -unit sphere<br />

S 2 := { (x 1 , x 2 , x 3 ) ∈Ê3 | | x 1 | 3 + | x 2 | 3 + | x 3 | 3 = 1 } .<br />

Proof. First step. We cover the domain of integration by 25 sets as follows.<br />

We put R 0 := [−b 0 , b 0 ] 4 ∩ CX (1, ... ,1) (Ê). Further, for each σ ∈ S 4 , we set<br />

R σ := { (x 0 , . . . , x 3 ) ∈Ê4 | | x σ(0) | ≤ · · · ≤ |x σ(3) |, | x i | ≤ b i , b 0 ≤ | x σ(3) | }<br />

Second step. One has R R σ<br />

ω CX (1, ... ,1) (Ê)<br />

Leray<br />

∩ CX (1, ... ,1) (Ê) .<br />

≤ R R id<br />

ω CX (1, ... ,1) (Ê)<br />

Leray for every σ ∈ S 4 .<br />

Consider the map i σ :Ê4 →Ê4 given by (x 0 , . . . , x 3 ) ↦→ (x σ(0) , . . . , x σ(3) ).<br />

Since CX (1, ... ,1) (Ê) is defined by a symmetric cubic form, it is invariant under i σ .<br />

We claim that i σ (R σ ) ⊆ R id .<br />

Indeed, let (x 0 , . . . , x 3 ) ∈ R σ . Then, i σ (x 0 , . . . , x 3 ) = (x σ(0) , . . . , x σ(3) )<br />

has the properties | x σ(0) | ≤ . . . ≤ | x σ(3) | <strong>and</strong> b 0 ≤ | x σ(3) |. In order to show<br />

i σ (x 0 , . . . , x 3 ) ∈ R id , all we need to verify is | x σ(i) | ≤ b i for i = 0, . . . , 3.<br />

For this, we use that the b i are sorted. We have | x σ(3) | ≤ b σ(3) ≤ b 3 .<br />

Further, | x σ(2) | ≤ b σ(2) <strong>and</strong> | x σ(2) | ≤ |x σ(3) | ≤ b σ(3) , one of which is at<br />

most equal to b 2 . Similarly, | x σ(1) | ≤ b σ(1) , | x σ(1) | ≤ |x σ(2) | ≤ b σ(2) ,<br />

<strong>and</strong> | x σ(1) | ≤ |x σ(3) | ≤ b σ(3) , the smallest of which is not larger than b 1 .<br />

Finally, | x σ(0) | ≤ b σ(0) , | x σ(0) | ≤ |x σ(1) | ≤ b σ(1) , | x σ(0) | ≤ |x σ(2) | ≤ b σ(2) ,<br />

<strong>and</strong> | x σ(0) | ≤ |x σ(3) | ≤ b σ(3) which shows | x σ(0) | ≤ b 0 .<br />

As x 3 0+ . . . +x 3 3 is a symmetric form, the Leray measure on CX (1, ... ,1) (Ê) is invariant<br />

under the canonical operation of S 4 on CX (1, ... ,1) (Ê) ⊂Ê4 . This means,<br />

we have (i σ ) ∗ ω CX (1, ... ,1) (Ê)<br />

Leray = ω CX (1, ... ,1) (Ê)<br />

Leray for each σ ∈ S 4 .


324 ON THE SMALLEST POINT ON A DIAGONAL CUBIC SURFACE [Chap. X<br />

Altogether,<br />

Z<br />

ω CX (1, ... ,1) (Ê)<br />

Leray<br />

R σ<br />

≤<br />

Z<br />

i −1 σ (R id )<br />

Z<br />

ω CX (1, ... ,1) (Ê)<br />

Leray =<br />

(i σ ) ∗ ω CX (1, ... ,1) (Ê)<br />

Leray =<br />

R id<br />

Z<br />

R id<br />

ω CX (1, ... ,1) (Ê)<br />

Leray .<br />

Third step. We have R R 0<br />

ω CX (1, ... ,1) (Ê)<br />

Leray ≤<br />

3√ 1 3ω2 b<br />

3 0 .<br />

By virtue of Corollary 6.14, we have<br />

Z<br />

Leray = 1 3<br />

R 0<br />

ω CX (1, ... ,1) (Ê)<br />

= 1 3<br />

Z<br />

1<br />

x 2 3<br />

R 0<br />

ZZZ<br />

π(R 0 )<br />

dx 0 ∧ dx 1 ∧ dx 2<br />

1<br />

(x 3 0 + x3 1 + x3 2 )2/3 dx 0 dx 1 dx 2<br />

where π : CX (1, ... ,1) (Ê) →Ê3 , (x 0 , x 1 , x 2 , x 3 ) ↦→ (x 0 , x 1 , x 2 ), denotes the projection<br />

to the first three coordinates.<br />

We enlarge the domain of integration to<br />

R ′ := { (x 1 , x 2 , x 3 ) ∈Ê3 | | x 0 | 3 + | x 1 | 3 + | x 2 | 3 ≤ 3b 3 0 } .<br />

Then, by homogeneity, we see<br />

ZZZ<br />

R ′ 1<br />

(x 3 0 + x3 1 + x3 2 )2/3 dx 0 dx 1 dx 2 = ω 2 ·<br />

Fourth step. We have R<br />

Observe<br />

R id<br />

ω CX (1, ... ,1) (Ê)<br />

3√<br />

3b0<br />

Z<br />

0<br />

1<br />

r 2 · r2 dr = ω 2 · 3√<br />

3b0 .<br />

Leray ≤ ( 8 3 + 8 9 log 3)b 0 + 8 3 b 0 log b 1<br />

| x 3 | = ∣<br />

√x 3 3 0 + x3 1 + x3 2 ∣ ≤<br />

√| 3 x 0 | 3 + | x 1 | 3 + | x 2 | 3 .<br />

For (x 0 , . . . , x 3 ) ∈ R id , this implies | x 3 | ≤ 3 √<br />

3 | x2 | <strong>and</strong> | x 2 | ≥ b 0 / 3 √<br />

3. We find<br />

Z<br />

Leray = 1 3<br />

R id<br />

ω CX (1, ... ,1) (Ê)<br />

Z<br />

1<br />

x 2 3<br />

R id<br />

≤ 1 Z<br />

1<br />

3 x 2 2<br />

R id<br />

< 1 3<br />

≤ 1 3<br />

Z b 0<br />

dx 0 ∧ dx 1 ∧ dx 2<br />

dx 0 ∧ dx 1 ∧ dx 2<br />

Z<br />

−b 0 |x 1 |∈[| x 0 |,b 1 ]<br />

Z b 0 Z<br />

−b 0 |x 1 |∈[|x 0 |,b 1 ]<br />

Z<br />

|x 2 |≥b 0 / 3√ 3<br />

|x 2 |≥|x 1 |<br />

1<br />

x 2 2<br />

dx 2 dx 1 dx 0<br />

2<br />

√<br />

max{b 0 / 3 3, | x1 |} dx 1 dx 0<br />

b 0<br />

.


Sec. 6] THE FUNDAMENTAL FINITENESS PROPERTY 325<br />

≤ 2 3<br />

⎡<br />

⎢<br />

⎣<br />

Z b 0<br />

Z<br />

−b 0 |x 1 |∈[|x 0 |,b 0 / 3√ 3]<br />

3√<br />

3<br />

dx 1 dx 0 +<br />

b 0<br />

Z b 0<br />

Z<br />

−b 0 |x 1 |∈[b 0 / 3√ 3,b 1 ]<br />

⎤<br />

1<br />

| x 1 | dx ⎥<br />

1 dx 0 ⎦<br />

≤ 2 3 · 4b2 0<br />

3√<br />

3 ·<br />

3√<br />

3<br />

+ 2 Z b 0<br />

2 log<br />

b 0 3<br />

−b 0<br />

3√<br />

3b1<br />

b 0<br />

dx 0<br />

= 8 3 b 0 + 8 3 b 0 log<br />

=<br />

3√<br />

3b1<br />

b 0<br />

( 8<br />

3 + 8 9 log 3 )<br />

b 0 + 8 3 b 0 log b 1<br />

b 0<br />

.<br />

□<br />

6.17. Corollary. –––– For every ε > 0, there exists a constant C such that<br />

τ ∞<br />

(<br />

X<br />

(a 0 , ... ,a 3 ) (Ê) ) ≤ C · | a 0 · . . . · a 3 | − 1 3 +ε · min<br />

i=0, ... ,3 ‖a i‖ 1 3<br />

∞<br />

for each (a 0 , . . . , a 3 ) ∈ (\{0}) 4 .<br />

Proof. We assume without restriction that | a 0 | ≤ . . . ≤ | a 3 |. Then,<br />

Lemma 6.15 <strong>and</strong> Proposition 6.16 together show that, for certain explicit positive<br />

constants C 1 <strong>and</strong> C 2 ,<br />

(<br />

τ ∞ X<br />

(a 0 , ... ,a 3 ) (Ê) ) (<br />

√ )<br />

≤ | a 0 · . . . · a 3 | − 3 1 · C 1 | a 0 | 3 1 + C2 | a 0 | 3 1 log<br />

3<br />

| a 1 |<br />

| a 0 |<br />

= | a 0 · . . . · a 3 | − 1 3 · | a0 | 1 3<br />

(<br />

C 1 + 1 3 C 2 log | a 1|<br />

| a 0 |<br />

≤ | a 0 · . . . · a 3 | − 3 1 · min ‖a i‖ 1 3<br />

∞<br />

i=0, ... ,3<br />

(<br />

· C 1 + 1 )<br />

3 C 2 log | a 0 · . . . · a 3 | .<br />

There is a constant C such that C 1 + 1 C 3 2 log | a 0 · . . . · a 3 | ≤ C | a 0 · . . . · a 3 | ε<br />

for every (a 0 , . . . , a 3 ) ∈ (\{0}) 4 .<br />

□<br />

)<br />

iv. The Tamagawa number. —<br />

6.18. Proposition. –––– For every ε > 0, there exists a constant C > 0 such that<br />

1<br />

τ ≥ C · H( ) 1<br />

1 1<br />

a 0<br />

: . . . : 3<br />

a 3 (a 0, ... ,a 3 )<br />

|a 0 · . . . · a 3 | ε<br />

for each (a 0 , . . . , a 3 ) ∈ (\{0}) 4 .


326 ON THE SMALLEST POINT ON A DIAGONAL CUBIC SURFACE [Chap. X<br />

Proof. We may assume that ε is small, say ε < 2 . Then, immediately from the<br />

3<br />

definition of τ (a 0, ... ,a 3 ) , we have<br />

τ (a 0, ... ,a 3 )<br />

= α(X (a 0, ... ,a 3 ) )·β(X (a 0, ... ,a 3 ) ) · lim<br />

s→1<br />

(s − 1) t L ( s, χ Pic(X<br />

(a 0 , ... ,a 3 )<br />

≤ α(X (a 0, ... ,a 3 ) )·β(X (a 0, ... ,a 3 ) ) · lim<br />

s→1<br />

(s − 1) t L ( s, χ Pic(X<br />

(a 0 , ... ,a 3 )<br />

É<br />

)<br />

= α(X (a 0, ... ,a 3 ) )·β(X (a 0, ... ,a 3 ) ) · lim<br />

s→1<br />

(s − 1) t L ( s, χ Pic(X<br />

(a 0 , ... ,a 3 )<br />

)<br />

)<br />

É<br />

(<br />

·τ H X<br />

(a 0 , ... ,a 3 ) (É) Br)<br />

)<br />

)<br />

(<br />

·τ H X<br />

(a 0 , ... ,a 3 ) (É) )<br />

)<br />

)<br />

·∏<br />

ν∈Val(É)<br />

τ ν<br />

(<br />

X<br />

(a 0 , ... ,a 3 ) (Éν) ) .<br />

Let us collect estimates for the factors. First, by Proposition 6.2, we have<br />

lim<br />

s→1<br />

(s − 1) t L ( s, χ Pic(X<br />

(a 0 , ... ,a 3 )<br />

) < c1 · | a 0 · . . . · a 3 | ε 16<br />

for a certain constant c 1 . Further, Proposition 6.13 yields<br />

(<br />

∏ τ p X<br />

(a 0 , ... ,a 3 ) (Ép) ) ≤ c 2 · | a 0 · . . . · a 3 | 3 1 − ε<br />

16 · ∏<br />

p prime<br />

Finally, Corollary 6.17 shows<br />

p prime<br />

min ‖a 3<br />

i‖ 1 − ε 2<br />

i=0, ... ,3<br />

τ ∞<br />

(<br />

X<br />

(a 0 , ... ,a 3 ) (Ê) ) ≤ C · | a 0 · . . . · a 3 | − 1 3 + ε 2 · min<br />

i=0, ... ,3 ‖a i‖ 1 3<br />

∞ .<br />

p .<br />

We assert that the three inequalities together imply the following estimate for<br />

Peyre’s constant τ (a 0, ... ,a 3) = τ (X (a 0, ... ,a 3 ) ),<br />

τ (a 0, ... ,a 3 ) ≤ C 3 · | a 0 · . . . · a 3 | ε 2 · ∏<br />

p prime<br />

min ‖a i‖ 1 3 p<br />

i=0, ... ,3<br />

[ ]<br />

· min ‖a i‖ 1 −<br />

ε<br />

3<br />

∞ · ∏ min ‖a 2<br />

i‖ p .<br />

i=0, ... ,3 i=0, ... ,3<br />

p prime<br />

Indeed, this is trivial in the case τ (a 0, ... ,a 3 )<br />

= 0. Otherwise, X (a 0, ... ,a 3 ) has an<br />

adelic point <strong>and</strong> we may estimate the factors α <strong>and</strong> β by constants as shown in<br />

Section 2. By consequence,<br />

1<br />

τ (a 0,...,a 3 )<br />

≥ 1 C 3<br />

·<br />

[<br />

∏<br />

p prime<br />

] −<br />

1<br />

min ‖a 3<br />

i‖ p ·<br />

i=0,...,3<br />

[<br />

] −<br />

1<br />

min ‖a 3<br />

i‖ ∞<br />

i=0,...,3<br />

[<br />

| a 0 · . . . · a 3 | ε 2 · ∏ min ‖a ] −<br />

ε<br />

2<br />

i‖ p<br />

p prime i=0,...,3


Sec. 6] THE FUNDAMENTAL FINITENESS PROPERTY 327<br />

= 1 C 3<br />

·<br />

= 1 C 3<br />

·<br />

∏<br />

max<br />

p prime i=0,...,3<br />

∥ 1 ∥ 1 ∥∥ 3<br />

a i p<br />

[<br />

| a 0 · . . . · a 3 | ε 2 · ∏<br />

p prime<br />

· max<br />

i=0,...,3<br />

∥ 1 ∥ 1 ∥∥ 3<br />

a i<br />

∞<br />

max<br />

i=0,...,3 a(p) i<br />

] ε<br />

2<br />

H ( ) 1<br />

1 1<br />

a 0<br />

: . . . : 3<br />

a 3<br />

[ ] ε .<br />

| a 0 · . . . · a 3 | ε 2<br />

2 · ∏ max<br />

p prime i=0,...,3 a(p) i<br />

It is obvious that max<br />

i=0,...,3 a(p) i ≤ | a (p)<br />

0<br />

· . . . · a (p)<br />

3 | <strong>and</strong><br />

This shows<br />

∏ | a (p)<br />

0<br />

· . . . · a (p)<br />

3 | = | a 0 · . . . · a 3 | .<br />

p prime<br />

1<br />

τ (a 0,...,a 3 ) ≥ 1 C 3<br />

·<br />

H ( ) 1<br />

1 1<br />

a 0<br />

: . . . : 3<br />

a 3<br />

| a 0 · . . . · a 3 | ε 2 · | a 0 · . . . · a 3 | ε 2<br />

= 1 C 3<br />

· H( 1<br />

a 0<br />

: . . . :<br />

1<br />

a 3<br />

) 1<br />

3<br />

| a 0 · . . . · a 3 | ε . □<br />

6.19. Lemma. –––– Let (a 0 : . . . : a 3 ) ∈ P 3 (É) be any point such that<br />

a 0 ≠ 0, . . . , a 3<br />

∈<br />

≠ 0. Then,<br />

H(a 0 : . . . : a 3 ) ≤ H( 1 1<br />

a 0<br />

: . . . :<br />

a 3<br />

) 3 .<br />

Proof. First, observe that (a 0 : . . . : a 3 ) ↦→ ( )<br />

1<br />

1<br />

a 0<br />

: . . . :<br />

a 3<br />

is a welldefined<br />

map. Hence, we may assume without restriction that a 0 , . . . , a 3<br />

<strong>and</strong> gcd(a 0 , . . . , a 3 ) = 1. This yields H(a 0 : . . . : a 3 ) = max |a i|.<br />

i=0,...,3<br />

On the other h<strong>and</strong>, ( 1 1<br />

a 0<br />

: . . . :<br />

a 3<br />

) = (a 1 a 2 a 3 : . . . : a 0 a 1 a 2 ). Consequently,<br />

H ( 1<br />

a 0<br />

: . . . :<br />

1<br />

a 3<br />

) ≤ [ max<br />

i=0,...,3 |a i|] 3 = H(a 0 : . . . : a 3 ) 3 .<br />

From this, the asserted inequality emerges when the roles of a i <strong>and</strong> 1 a i<br />

are interchanged.<br />

□<br />

6.20. Corollary. –––– Let a 0 , . . . , a 3 ∈such that gcd(a 0 , . . . , a 3 ) = 1. Then,<br />

|a 0 · . . . · a 3 | ≤ H ( 1<br />

a 0<br />

: . . . :<br />

1<br />

a 3<br />

) 12.<br />

Proof. Observe that | a 0 · . . . · a 3 | ≤ max | a i| 4 = H(a 0 : . . . : a 3 ) 4 <strong>and</strong> apply<br />

i=0,...,3<br />

Lemma 6.19.<br />


328 ON THE SMALLEST POINT ON A DIAGONAL CUBIC SURFACE [Chap. X<br />

6.21. Theorem. –––– For each ε > 0, there exists a constant C (ε) > 0 such that,<br />

for all (a 0 , . . . , a 3 ) ∈ (\{0}) 4 ,<br />

1<br />

τ ≥ C (ε) · H( ) 1<br />

1 1<br />

(a 0,...,a 3 ) a 0<br />

: . . . : 3<br />

−ε<br />

a 3<br />

.<br />

Proof. We may assume that gcd(a 0 , . . . , a 3 ) = 1. Then, by Proposition 6.18,<br />

1<br />

τ ≥ C (ε) · H( ) 1<br />

1 1<br />

a 0<br />

: . . . : 3<br />

a 3<br />

.<br />

(a 0,...,a 3 )<br />

| a 0 · . . . · a 3 | ε 12<br />

(<br />

Corollary 6.20 yields | a 0 · . . . · a 3 | ε 12 ≤ H 1<br />

)<br />

1 ε.<br />

a 0<br />

: . . . :<br />

a 3<br />

6.22. Corollary (Fundamental finiteness). —– For each T > 0, there are only<br />

finitely many diagonal cubic surfaces X (a 0,...,a 3 ) : a 0 x 3 0 + . . . +a 3 x 3 3 = 0 in P 3Ésuch<br />

that τ (a 0,...,a 3) > T .<br />

Proof. This is an immediate consequence of the comparison to the naive height<br />

established in Theorem 6.21.<br />

□<br />


CHAPTER XI<br />

THE DIOPHANTINE EQUATION<br />

x 4 + 2y 4 = z 4 + 4w 4∗<br />

Hash, x. There is no definition for this word—nobody knows what hash is.<br />

AMBROSE BIERCE: The Devil’s Dictionary (1906)<br />

1. Introduction<br />

1.1. –––– An algebraic curve C of genus g > 1 admits at most a finite number<br />

ofÉ-<strong>rational</strong> <strong>points</strong>. On the other h<strong>and</strong>, for genus one curves, #C (É) may be<br />

zero, finite non-zero, or infinite. For genus zero curves, one automatically has<br />

#C (É) = ∞ as soon as C (É) ≠ ∅.<br />

1.2. –––– In higher dimensions, there is a conjecture, due to S. Lang, stating<br />

that if X is a variety of general type over a number field then all but finitely many<br />

of its <strong>rational</strong> <strong>points</strong> are contained in the union of closed subvarieties which are<br />

not of general type (cf. Conjecture VI.2.2). On the other h<strong>and</strong>, abelian varieties<br />

(as well as, e.g., elliptic <strong>and</strong> bielliptic surfaces) behave like genus one curves.<br />

I.e., #X (É) may be zero, finite non-zero, or infinite. Finally, <strong>rational</strong> <strong>and</strong> ruled<br />

varieties comport in the same way as curves of genus zero in this respect.<br />

This list does not yet exhaust the classification of algebraic surfaces, to say<br />

nothing of dimension three or higher. In particular, the following problem is<br />

still open.<br />

1.3. Problem. –––– Does there exist a K3 surface X overÉwhich has a finite<br />

non-zero number ofÉ-<strong>rational</strong> <strong>points</strong>? I.e., such that 0 < #X (É) < ∞?<br />

1.4. Remark. –––– This question was posed by Sir P. Swinnerton-Dyer as<br />

Problem/Question 6.a) in the problem session to the workshop [Poo/T]. We are<br />

not able to give an answer to it.<br />

(∗) This chapter collects material from the articles:<br />

The Diophantine equation x 4 + 2y 4 = z 4 + 4w 4 , Math. Comp. 75(2006), 935–940 <strong>and</strong><br />

The Diophantine equation x 4 + 2y 4 = z 4 + 4w 4 — a number of improvements, Preprint,<br />

both joint with A.-S. Elsenhans.


330 THE DIOPHANTINE EQUATION x 4 + 2y 4 = z 4 + 4w 4 [Chap. XI<br />

1.5. –––– One possible c<strong>and</strong>idate for a K3 surface with the property<br />

0 < #X (É) < ∞ is given by the following<br />

Problem. Find a third point on the projective surface X ⊂ P 3 defined by<br />

x 4 + 2y 4 = z 4 + 4w 4 .<br />

1.6. Remarks. –––– i) Problem 1.5 is also due to Sir P. Swinnerton-Dyer<br />

[Poo/T, Problem/Question 6.c)]. It was raised, in particular, during his<br />

talk [S-D04, very end of the article] at the Göttingen Mathematisches Institut<br />

on June 2 nd , 2004.<br />

ii) x 4 +2y 4 = z 4 +4w 4 is a homogeneous quartic equation. It, therefore, defines<br />

a K3 surface X in P 3 . As trivial solutions of the equation, we consider those<br />

corresponding to theÉ-<strong>rational</strong> <strong>points</strong> (1:0:1:0) <strong>and</strong> (1:0:(−1):0).<br />

iii) Our main result is the following theorem which contains an answer to<br />

Problem 1.5.<br />

1.7. Theorem. –––– The diagonal quartic surface X in P 3 given by<br />

x 4 + 2y 4 = z 4 + 4w 4<br />

(∗)<br />

admits precisely tenÉ-<strong>rational</strong> <strong>points</strong> having integral coordinates within the hypercube<br />

|x|, |y|, |z|, |w| < 10 8 .<br />

These are (±1:0:±1:0) <strong>and</strong> (±1 484 801:±1 203 120:±1169407:±1157520).<br />

1.8. Remark. –––– This result does clearly not exclude the possibility that<br />

#X (É) is actually finite. It might indicate, however, that a proof for this<br />

property is deeper than one originally hoped for.<br />

2. Congruences<br />

2.1. –––– It seems natural to first try to underst<strong>and</strong> the congruences<br />

x 4 + 2y 4 ≡ z 4 + 4w 4 (mod p) (†)<br />

modulo some prime number p. For p = 2 <strong>and</strong> 5, one finds that all primitive<br />

solutions insatisfy<br />

a) x <strong>and</strong> z are odd,<br />

b) y <strong>and</strong> w are even,<br />

c) y is divisible by 5.


Sec. 2] CONGRUENCES 331<br />

For other primes, it follows from the Weil conjectures, proven by P. Deligne<br />

[Del], that the number of solutions of the congruence (†) is<br />

#C X (p) = 1 + (p − 1)(p 2 + p + 1 + E) = p 3 + E(p − 1) .<br />

Here, E is an error-term which may be estimated by |E| ≤ 21p.<br />

Indeed, consider the projective variety X overÉdefined by (∗). It has good reduction<br />

at every prime p ≠ 2. Therefore, [Del, Théorème (8.1)] may be applied<br />

to the reduction X p . This yields #X p (p) = p 2 + p + 1 + E <strong>and</strong> |E| ≤ 21p.<br />

We note that dim H 2 (X ,Ê) = 22 for every complex surface X of type K3 [Bv,<br />

p. 98].<br />

2.2. –––– Another question of interest is to count the numbers of solutions<br />

to the congruences x 4 + 2y 4 ≡ c (mod p) <strong>and</strong> z 4 + 4w 4 ≡ c (mod p) for a<br />

certain c ∈.<br />

This means to count thep-<strong>rational</strong> <strong>points</strong> on the affine plane curves C l c <strong>and</strong> C r c<br />

defined overp by x 4 + 2y 4 = c <strong>and</strong> z 4 + 4w 4 = c, respectively. If p ∤ c<br />

<strong>and</strong> p ≠ 2 then these are smooth curves of genus three.<br />

By the work of André Weil [We48, Corollaire 3 du Théorème 13], the numbers<br />

ofp-<strong>rational</strong> <strong>points</strong> on their projectivizations are given by<br />

#C l c (p) = p + 1 + E l <strong>and</strong> #C r c (p) = p + 1 + E r ,<br />

where the error-terms can be bounded by |E l |, |E r | ≤ 6 √ p. There may be up<br />

to fourp-<strong>rational</strong> <strong>points</strong> on the infinite line. For our purposes, it suffices to<br />

notice that both congruences admit a number of solutions which is close to p.<br />

The case p|c, p ≠ 2, is slightly different since it corresponds to the case of a<br />

reducible curve. The congruence x 4 + ky 4 ≡ 0 (mod p) admits only the trivial<br />

solution if (−k) is not a biquadratic residue modulo p. Otherwise, it has exactly<br />

1 + (p − 1) gcd(p − 1, 4) solutions.<br />

Finally, if p = 2 then #C l 0 (2) = #C l 1 (2) = #C r 0 (2) = #C r 1 (2) = 2.<br />

2.3. Remark. –––– The number of solutions of the congruence (†) is<br />

#C X (p) = ∑ #Cc l (p) · #Cc r (p).<br />

c∈p<br />

Hence, the formulas just mentioned yield an elementary estimate for that count.<br />

They show once more that the dominating term is p 3 . The estimate for the<br />

error is, however, less sharp than the one obtained via the more sophisticated<br />

methods in 2.1.


332 THE DIOPHANTINE EQUATION x 4 + 2y 4 = z 4 + 4w 4 [Chap. XI<br />

3. Naive methods<br />

3.1. –––– The most naive method to search for solutions of (∗) is probably<br />

the following. Start with the set<br />

{(x, y, z, w) ∈|0≤x, y, z, w ≤ N }<br />

<strong>and</strong> test the equation for every quadruple.<br />

Obviously this method requires about N 4 steps. It can be accelerated using the<br />

congruence conditions for primitive solutions noticed above.<br />

3.2. –––– A somewhat better method is to start with the set<br />

{x 4 + 2y 4 − 4w 4 | x, y, w ∈, 0 ≤ x, y, w ≤ N }<br />

<strong>and</strong> to search for fourth powers. This set has about N 3 Elements, <strong>and</strong> the<br />

algorithm takes about N 3 steps. Again, it can be sped up by the above congruence<br />

conditions for primitive solutions. We used this approach for a trial-run<br />

with N = 10 4 .<br />

An interesting aspect of this algorithm is the optimization by further congruences.<br />

Suppose x <strong>and</strong> y are fixed, then about one half or three-quarter of the<br />

values for w are no solutions to the congruence modulo a new prime. Following<br />

this way, one can find more congruences for w <strong>and</strong> the size of the set may<br />

be reduced by a constant factor.<br />

4. An algorithm to efficiently search for solutions<br />

i. The basic idea. —<br />

4.1. –––– We need to compute the intersection of two sets<br />

{x 4 + 2y 4 | x, y ∈, 0 ≤ x, y ≤ N } ∩ {z 4 + 4w 4 | z, w ∈, 0 ≤ z, w ≤ N } .<br />

Both have about N 2 elements.<br />

It is a st<strong>and</strong>ard problem in Computer Science to find the intersection of two<br />

sets which both fit into memory. Using the congruence conditions modulo 2<br />

<strong>and</strong> 5, one can reduce the size of the first set by a factor of 20 <strong>and</strong> the size of the<br />

second set by a factor of 4.


Sec. 4] AN ALGORITHM TO EFFICIENTLY SEARCH FOR SOLUTIONS 333<br />

ii. Some details. —<br />

4.2. –––– The two sets described above are too big, at least for our computers<br />

<strong>and</strong> interesting values of N . Therefore, we introduce a prime number p p which<br />

we call the page prime.<br />

Define the sets<br />

<strong>and</strong><br />

L c := {x 4 + 2y 4 | x, y ∈, 0 ≤ x, y ≤ N , x 4 + 2y 4 ≡ c (mod p p )}<br />

R c := {z 4 + 4w 4 | z, w ∈, 0 ≤ z, w ≤ N , z 4 + 4w 4 ≡ c (mod p p )} .<br />

This means, the intersection problem is divided into p p pieces <strong>and</strong> the sets L c<br />

<strong>and</strong> R c fit into the computer’s memory if p p is big enough. We worked with<br />

N = 2.5 · 10 6 <strong>and</strong> chose p p = 30 011.<br />

For every value of c, our program computes L c <strong>and</strong> stores this set in a hash table.<br />

Then, it determines the elements of R c <strong>and</strong> looks them up in the table. Assuming<br />

uniform hashing, the expected running-time of this algorithm is O(N 2 ).<br />

4.3. Remark –––– An important further aspect of this approach is that the<br />

problem may be attacked in parallel on several machines. The calculations<br />

for one particular value of c are independent of the analogous calculations for<br />

another one. Thus, it is possible, say, to let c run from 0 to (p p − 1)/2 on one<br />

machine <strong>and</strong>, at the same time, from (p p + 1)/2 to (p p − 1) on another.<br />

iii. Some more details. —<br />

iii.1. The page prime. —<br />

4.4. –––– For each value of c, it is necessary to find the solutions of the congruences<br />

x 4 +2y 4 ≡ c (mod p p ) <strong>and</strong> z 4 +4w 4 ≡ c (mod p p ) in an efficient manner.<br />

We do this in a rather naive way by letting y (w) run from 0 to p p − 1.<br />

For each value of y (w), we compute x 4 (z 4 ). Then, we extract the fourth root<br />

modulo p p .<br />

Note that the page prime fulfills p p ≡ 3 (mod 4). Hence, the fourth roots of<br />

unity modulo p are just ±1 <strong>and</strong>, therefore, a fourth root modulo p p , if it exists,<br />

is unique up to sign. This makes the algorithm easier to implement.


334 THE DIOPHANTINE EQUATION x 4 + 2y 4 = z 4 + 4w 4 [Chap. XI<br />

4.5. –––– Actually, we do not execute any modular powering operation or even<br />

computation of fourth roots in the lion’s share of the running time. For more<br />

efficiency, all fourth powers <strong>and</strong> all fourth roots modulo p p are computed <strong>and</strong><br />

stored in an array during an initialization step. Thus, the main speed limitation<br />

to find all solutions to a congruence modulo p p is, in fact, the time it takes to<br />

look up values stored in the machine’s main memory.<br />

iii.2. Hashing. —<br />

4.6. –––– We do not compute L c <strong>and</strong> R c directly, because this would require<br />

the use of multi precision integers within the inner loop. Instead, we choose<br />

two other primes, the hash prime p h <strong>and</strong> the control prime p c , which fit into<br />

the 32-bit registers of our computers. All computations are done modulo p h<br />

<strong>and</strong> p c .<br />

More precisely, for each pair (x, y) considered, the expression<br />

(<br />

(x 4 + 2y 4 ) mod p h<br />

)<br />

defines its position in the hash table. In other words, we hash pairs (x, y)<br />

whereas (x, y) ↦→ ( (x 4 + 2y 4 ) mod p h<br />

)<br />

plays the role of the hash function.<br />

For each pair (x, y), we write two entries into the hash table, namely the value<br />

of ( (x 4 + 2y 4 ) mod p c<br />

)<br />

<strong>and</strong> the value of y.<br />

In the main computation, we worked with the numbers p h = 25 000 009 for the<br />

hash prime <strong>and</strong> p c = 400 000 009 for the control prime.<br />

4.7. –––– Note that, when working with a particular value of c, there are<br />

around p p pairs ((x mod p p ), (y mod p p )) which fulfill the required congruence<br />

Therefore, approximately<br />

x 4 + 2y 4 ≡ c (mod p p ) .<br />

p p · (N /2<br />

p p<br />

· N /10<br />

p p<br />

) =<br />

N 2<br />

20p p<br />

values will be written into the table. For our choices,<br />

N 2<br />

20p p<br />

≈ 10 412 849 which<br />

means that the hash table will get approximately 41.7% filled.<br />

Like for many other rules, there is an exception to this one. If c = 0 then approximately<br />

1 + (p p − 1) gcd(p p − 1, 4) pairs ((x mod p p ), (y mod p p )) may<br />

satisfy the congruence<br />

x 4 + 2y 4 ≡ 0 (mod p p ) .<br />

As p p ≡ 3 (mod 4) this is not more than 2p p − 1 <strong>and</strong> the hash table will be<br />

filled not more than about 83.3%.


Sec. 5] GENERAL FORMULATION OF THE METHOD 335<br />

4.8. –––– To resolve collisions within the hash table, we use an open addressing<br />

method. We are not particularly afraid of clustering <strong>and</strong> choose linear probing.<br />

We feel free to use open addressing as, thanks to the Weil conjectures, we<br />

have a priori estimates available for the load factor.<br />

4.9. –––– The program makes frequent use of fourth powers modulo p h <strong>and</strong> p c .<br />

Again, we compute these data in the initialization part of our program <strong>and</strong> store<br />

them in arrays, once <strong>and</strong> for all.<br />

5. General formulation of the method<br />

5.1. –––– The method described in the previous section is actually a systematic<br />

method to search for solutions of a Diophantine equation. It works efficiently<br />

when the equation is of the form<br />

f (x 1 , . . . , x n ) = g(y 1 , . . . , y m ) .<br />

We find all solutions which are contained within the (n + m)-dimensional cube<br />

{(x 1 , . . . , x n , y 1 , . . . , y m ) ∈n+m | | x i |, |y i | ≤ B} .<br />

The expected running-time of the algorithm is O(B max{n,m} ).<br />

5.2. –––– The basic idea may be formulated as follows.<br />

Algorithm H.<br />

i) Evaluate f on all <strong>points</strong> of the n-dimensional cube<br />

Store the values within a set L.<br />

ii) Evaluate g on all <strong>points</strong> of the cube<br />

{(x 1 , . . . , x n ) ∈n | |x i | ≤ B} .<br />

{(y 1 , . . . , y m ) ∈m | |y i | ≤ B}<br />

of dimension m. For each value start a search in order to find out whether<br />

it occurs in L. When a coincidence is detected, reconstruct the corresponding<br />

values of x 1 , . . . , x n <strong>and</strong> output the solution.<br />

5.3. Remarks. –––– a) In fact, we are interested in the very particular Diophantine<br />

equation<br />

x 4 + 2y 4 = z 4 + 4w 4<br />

which was suggested by Sir Peter Swinnerton-Dyer.


336 THE DIOPHANTINE EQUATION x 4 + 2y 4 = z 4 + 4w 4 [Chap. XI<br />

b.i) In the form stated above, the main disadvantage of Algorithm H is that it<br />

requires an enormous amount of memory. Actually, the set L is too big to be<br />

stored in the main memory even of our biggest computers, already when the<br />

value of B is only moderately large.<br />

For that reason, we introduced the idea of paging. We choose a page prime p p <strong>and</strong><br />

work with the sets L r := {s ∈ L | s ≡ r (mod p p )} for r = 0, . . . , p p − 1, separately.<br />

At the cost of some more time spent on initializations, this yields<br />

a reduction of the memory space required by a factor of 1 p p<br />

.<br />

ii) The sets L r were implemented in the form of a hash table with open addressing.<br />

iii) It is possible to achieve a further reduction of the running-time <strong>and</strong> the memory<br />

required by making use of some obvious congruence conditions modulo 2<br />

<strong>and</strong> 5.<br />

5.4. –––– The goal of the remainder of this chapter is to describe an improved<br />

implementation of Algorithm H which we used in order to find<br />

all solutions of x 4 + 2y 4 = z 4 + 4w 4 contained within the hypercube<br />

{(x, y, z, w) ∈4 | |x|, |y|, |z|, |w| ≤ 10 8 }.<br />

6. Improvements I – More congruences<br />

6.1. –––– The most obvious way to further reduce the size of the sets L r <strong>and</strong><br />

to increase the speed of Algorithm H is to find further congruence conditions<br />

for solutions <strong>and</strong> evaluate f <strong>and</strong> g only on <strong>points</strong> satisfying these conditions.<br />

As the equation, we are interested in, is homogeneous, it is sufficient to restrict<br />

consideration to primitive solutions.<br />

6.2. –––– It should be noticed, however, that this idea is subject to strict limitations.<br />

If we were using the most naive O(B n+m )-algorithm then, for more or less<br />

every l ∈Æ, the congruence f (x 1 , . . . , x n ) ≡ g(y 1 , . . . , y m ) (mod l) caused a<br />

reduction of the number of (n + m)-tuples to be checked. For Algorithm H,<br />

however, the situation is by far less fortunate.<br />

One may gain something only if there are residue classes (r mod l) which are<br />

represented by f , but not by g, or vice versa. Values, the residue class of which<br />

is not represented by g, do not need to be stored into L r . Values, the residue<br />

class of which is not represented by f , do not need to be searched for.<br />

Unfortunately, if l is prime <strong>and</strong> not very small then the Weil conjectures ensure<br />

that all residue classes modulo l are represented by both f <strong>and</strong> g. In this case, the<br />

idea fails completely. The same is, however, not true for prime powers l = p k .<br />

Hensel’s Lemma does not work when all partial derivatives ∂f<br />

∂x i<br />

(x 1 , . . . , x n ),


Sec. 6] IMPROVEMENTS I – MORE CONGRUENCES 337<br />

respectively ∂g<br />

∂y i<br />

(y 1 , . . . , y m ), are divisible by p. This makes it possible that<br />

certain residue classes (r mod p k ) are not representable although (r mod p) is.<br />

i. The prime 5. Congruences modulo 625. —<br />

6.3. –––– In the algorithm described in the previous chapter, we made use<br />

of the fact that y is always divisible by 5. However, at this point, one can<br />

do a lot better. When one takes into consideration that a 4 ≡ 1 (mod 5) for<br />

every a ∈not divisible by 5, a systematic inspection shows that there are<br />

actually two cases.<br />

Either, 5|w. Then, 5∤x <strong>and</strong> 5∤z. Or, otherwise, 5|x. Then, 5∤z <strong>and</strong> 5∤w.<br />

Note that, in the latter case, one indeed has z 4 + 4w 4 ≡ 1 + 4 ≡ 0 (mod 5).<br />

6.4. –––– The Case 5|w. We call this case “N” <strong>and</strong> use the letter N at a<br />

prominent position in the naming of the relevant files of source code. N st<strong>and</strong>s<br />

for “normal”. To consider this case as the ordinary one is justified by the fact<br />

that all primitive solutions known actually belong to it. Note, however, that<br />

we have no theoretical reason to believe that this case should in whatever sense<br />

be better than the other one.<br />

In case N, we rearrange the equation to f N (x, z) = g N (y, w) where<br />

f N (x, z) := x 4 − z 4 <strong>and</strong> g N (y, w) := 4w 4 − 2y 4 .<br />

As y <strong>and</strong>w are both divisible by 5, we get g N (y, w) = 4w 4 −2y 4 ≡ 0 (mod 625).<br />

Consequently, f N (x, z) ≡ 0 (mod 625).<br />

This yields an enormous reduction of the set L r . To see this, recall 5∤x <strong>and</strong> 5∤z.<br />

That means, for x, there are precisely ϕ(625) possibilities in/625. Further,<br />

for each such value, the congruence z 4 ≡ x 4 (mod 625) may not have<br />

more than four solutions. All in all, there are 4 · ϕ(625) = 2 000 possible pairs<br />

(x, z) ∈ (/625) 2.<br />

Further, these pairs are very easy to find, computationally. The fourth<br />

roots of unity modulo 625 are ±1 <strong>and</strong> ±182. For each x ∈/625∗ , put<br />

z := (±x mod 625) <strong>and</strong> z := (±182x mod 625).<br />

We store the values of f N into the set L r . Only 2 000 out of 625 2 values (0.512%)<br />

need to be computed <strong>and</strong> stored. Then, each value of g N is looked up in L r .<br />

Here, as y <strong>and</strong> w are both divisible by 5, only one value out of 25 (4%) needs to<br />

be computed <strong>and</strong> searched for.<br />

6.5. –––– The Case 5|x. We call this case “S” <strong>and</strong> use the letter S at a prominent<br />

position in the naming of the relevant files of source code. S st<strong>and</strong>s for<br />

“Sonderfall” which means “exceptional case”. It is not known whether there<br />

exists a solution belonging to case S.


338 THE DIOPHANTINE EQUATION x 4 + 2y 4 = z 4 + 4w 4 [Chap. XI<br />

Here, we simply interchange both sides of the equation. Define<br />

f S (z, w) := z 4 + 4w 4 <strong>and</strong> g S (x, y) := x 4 + 2y 4 .<br />

As x <strong>and</strong> y are divisible by 5, we get x 4 + 2y 4 ≡ 0 (mod 625) <strong>and</strong>, therefore,<br />

z 4 + 4w 4 ≡ 0 (mod 625).<br />

Again, this congruence allows only 4 · ϕ(625) = 2 000 solutions<br />

(z, w) ∈ (/625) 2<br />

<strong>and</strong> these pairs are easily computable, too. The fourth roots of (−4)<br />

in/625are ±181 <strong>and</strong> ±183. For each x ∈/625∗ , one has to consider<br />

z := (±181x mod 625) <strong>and</strong> z := (±183x mod 625).<br />

We store the values of f S into the set L r . Then, we search through L r for the<br />

values of g S . As above, only 2 000 out of 625 2 values need to be computed<br />

<strong>and</strong> stored <strong>and</strong> one value out of 25 needs to be computed <strong>and</strong> searched for.<br />

ii. The prime 2. —<br />

6.6. –––– Any primitive solution is of the form that x <strong>and</strong> z are odd while y<br />

<strong>and</strong> w are even.<br />

6.7. –––– In case S, there is no way to do better than that as both f S <strong>and</strong> g S<br />

represent (r mod 2 k ) for k ≥ 4 if <strong>and</strong> only if r ≡ 1 (mod 16).<br />

IncaseN, thesituationissomewhatbetter. g N (y, w) = 4w 4 −2y 4 isalwaysdivisible<br />

by 32 while f N (x, z) = x 4 −z 4 ≡ 0 (mod 32), as may be seen by inspecting<br />

the fourth roots of unity modulo 32, implies the condition x ≡ ±z (mod 8).<br />

This may be used to halve the size of L r .<br />

iii. The prime 3. —<br />

6.8. –––– Looking for further congruence conditions, a primitive solution<br />

must necessarily satisfy, we did not find any reason to distinguish more cases.<br />

But there are a few more congruences which we used in order to reduce the size<br />

of the sets L r .<br />

To explain them, let us first note two theorems on binary quadratic forms.<br />

They may both be easily deduced from [H/W, Theorems 246 <strong>and</strong> 247].<br />

6.9. Theorem. –––– The quadratic formsq 1 (a, b) := a 2 +b 2 , q 2 (a, b) := a 2 −2b 2 ,<br />

<strong>and</strong> q 3 (a, b) := a 2 + 2b 2 admit the property below.<br />

Suppose n 0 := q i (a 0 , b 0 ) is divisible by a prime p which is not represented by q i .<br />

Then, p|a 0 <strong>and</strong> p|b 0 .


Sec. 6] IMPROVEMENTS I – MORE CONGRUENCES 339<br />

6.10. Theorem. –––– A prime number p is representedbyq 1 , q 2 , or q 3 , respectively,<br />

if <strong>and</strong> only if (0 mod p) is represented in a non-trivial way. In particular,<br />

i) p is represented by q 1 if <strong>and</strong> only if p = 2 or p ≡ 1 (mod 4).<br />

ii) p is represented by q 2 if <strong>and</strong> only if p = 2 or ( 2<br />

p) = 1. The latter means<br />

p ≡ 1, 7 (mod 8).<br />

iii) p is represented by q 3 if <strong>and</strong> only if p = 2 or ( −2<br />

p<br />

to p ≡ 1, 3 (mod 8).<br />

6.11. Remark –––– There is the obvious asymptotic estimate<br />

Further,<br />

♯{q i (a, b) | a, b ∈, q i (a, b) ∈È, q i (a, b) ≤ n} ∼<br />

) = 1. The latter is equivalent<br />

n<br />

2 logn .<br />

n<br />

♯{q i (a, b) | a, b ∈, |q i (a, b)| ≤ n} ∼ C i √<br />

logn<br />

where C 1 , C 2 , <strong>and</strong> C 3 are constants which can be expressed explicitly by Euler<br />

products. (For q 1 , this is worked out in [Brü, Satz (1.8.2)]. For the other<br />

forms, J. Brüdern’s argument works in the same way without essential changes.)<br />

6.12. Congruences modulo 81. ––––<br />

In case N,<br />

g N (y, w) = (2w 2 ) 2 − 2(y 2 ) 2 = q 2 (2w 2 , y 2 )<br />

where q 2 does not represent the prime 3. Therefore, if 3|g N (y, w) then 3|2w 2<br />

<strong>and</strong> 3|y 2 which implies y <strong>and</strong> w are both divisible by 3. By consequence, if<br />

3|g N (y, w) then, automatically, 81|g N (y, w).<br />

If 3| f N (x, z) but 81∤ f N (x, z) then f N (x, z) does not need to be stored into L r .<br />

Further, if 3|x <strong>and</strong> 3|z then f N (x, z) does not need to be stored, either, as it<br />

cannot lead to a primitive solution. This reduces the size of the set L r by a<br />

factor of 1 + 4 · 1<br />

9 3 (1 − 1 ) = 131 ≈ 53.9%.<br />

3 81 243<br />

In case S, the situation is the other way round.<br />

f S (z, w) = (z 2 ) 2 + (2w 2 ) 2 = q 1 (z 2 , 2w 2 )<br />

<strong>and</strong> q 1 does not represent the prime 3. Therefore, if 3| f S (z, w) then 3|z 2 <strong>and</strong><br />

3|2w 2 which implies that z <strong>and</strong> w are both divisible by 3 <strong>and</strong> 81| f S (z, w).<br />

We use this in order to reduce the time spent on reading. If 3|g S (x, y) but<br />

81∤g S (x, y) or if 3|x <strong>and</strong> 3|y then g S (x, y) does not need to be searched for.<br />

Although modular operations are not at all fast, the reduction of the number of<br />

attempts to read by 53.9% is highly noticeable.


340 THE DIOPHANTINE EQUATION x 4 + 2y 4 = z 4 + 4w 4 [Chap. XI<br />

iv. Some more hypothetical improvements. —<br />

6.13. –––– i) In the argument for case N given above, p = 3 might be replaced<br />

by any other prime p ≡ 3, 5 (mod 8).<br />

In case S, the same argument as above works for every prime p ≡ 3 (mod 8).<br />

For primes p ≡ 5 (mod 8), the strategy could be reversed. q 3 is a binary<br />

quadratic form which represents (0 mod p) only in the trivial manner. Therefore,<br />

if p|g S (x, y) then p|x <strong>and</strong> p|y. It is unnecessary to store f S (z, w) if p|z<br />

<strong>and</strong> p|w or if p| f S (z, w) but p 4 ∤ f S (z, w).<br />

i ′ ) Each argument mentioned may be extended to some primes p ≡ 1 (mod 8).<br />

For example, in case N, what is actually needed is that 2 is not a fourth power<br />

modulo p. This is true, e.g., for p = 17, 41, <strong>and</strong> 97, but not for p = 73 <strong>and</strong> 89.<br />

ii) f N <strong>and</strong> f S do not represent the residue classes of 6, 7, 10, <strong>and</strong> 11 modulo 17.<br />

g N <strong>and</strong> (−g S ) do not represent 1, 3, <strong>and</strong> 9 modulo 13. This could be used to<br />

reduce the load for writing as well as reading.<br />

6.14. Remarks. –––– a) We did not implement these improvements as it seems<br />

the gains would be marginal or the cost of additional computations would<br />

even dominate the effect. It is, however, foreseeable that these congruences will<br />

eventually become valuable when the speed of the CPU’s available will continue<br />

to grow faster than the speed of memory. Observe that alone the congruences<br />

noticed in a) could reduce the amount of data to be stored into L to a size<br />

asymptotically less than εB 2 for any ε > 0.<br />

b) For every prime p different from 2, 5, 13, <strong>and</strong> 17, the quartic forms f N , g N ,<br />

f S , <strong>and</strong> g S represent all residue classes modulo p. This means, ii) may not be<br />

carried over to any further primes.<br />

This can be seen as follows. Let b be equal to f N , f S , g N , or g S . (0 mod p) is<br />

represented by b, trivially. Otherwise, b(x, y) = r defines an affine curve<br />

C r of genus three with at most four <strong>points</strong> on the infinite line. The Weil<br />

conjectures [We48, Corollaire 3 du Théorème 13] imply that [(p+1−6 √ p)−4]<br />

is a lower bound for the number ofp-<strong>rational</strong> <strong>points</strong> on C r . This is a positive<br />

number as soon as p ≥ 43. In this case, every residue class (r mod p) is<br />

represented, at least, once.<br />

For the remaining primes up to p = 41, an experiment shows that all residue<br />

classes modulo p are represented by f N , f S , g N , as well as g S .


Sec. 7] IMPROVEMENTS II – ADAPTION TO OUR HARDWARE 341<br />

7. Improvements II – Adaption to our hardware<br />

i. A 64 bit based implementation of the algorithm. —<br />

7.1. –––– We migrated the implementation of Algorithm H from a 32 bit<br />

processor to a 64 bit processor. This means, the new hardware supports addition<br />

<strong>and</strong> multiplication of 64 bit integers. Even more, every operation on (unsigned)<br />

integers is automatically modulo 2 64 .<br />

From this, various optimizations of the implementation described in Section 4<br />

are almost compelling. The basic idea is that 64 bits should be enough to<br />

define hash value <strong>and</strong> control value by selection of bits instead of using (notoriously<br />

slow) modular operations. Hash value <strong>and</strong> control value are two integers<br />

significantly less than 2 32 which should be independent on each other.<br />

Note, however, that the congruence conditions modulo 2 imposed imply that<br />

x 4 ≡ z 4 ≡ 1 (mod 16) <strong>and</strong> 2y 4 ≡ 4w 4 ≡ 0 (mod 16). This means, the four<br />

least significant bits of f <strong>and</strong> g may not be used as they are always the same.<br />

7.2. –––– The description of the algorithm below is based on case S, case N<br />

being completely analogous.<br />

Algorithm H64.<br />

I. Initialization. Fix B := 10 8 . Initialize a hash table of 2 27 = 134 217 728<br />

integers, each being 32 bit long. Fix the page prime p p := 200 003.<br />

Further, define two functions, the hash function h <strong>and</strong> the control function c,<br />

which map 64 bit integers to 27 bit integers <strong>and</strong> 31 bit integers, respectively, by<br />

selecting certain bits. Do not use any of the bits twice to ensure h <strong>and</strong> c are<br />

independent on each other <strong>and</strong> do not use the four least significant bits.<br />

II) Loop. Let r run from 0 to p p − 1 <strong>and</strong> execute steps A. <strong>and</strong> B. for each r.<br />

A. Writing. Build up the hash table, which is meant to encode the set L r ,<br />

as follows.<br />

a) Find all pairs (z, w) of non-negative integers less than or equal to B which<br />

satisfy z 4 + 4w 4 ≡ r (mod p p ) <strong>and</strong> all the congruence-conditions for primitive<br />

solutions, listed above. (Make systematic use of the Chinese remainder theorem.)<br />

b) Execute steps i) <strong>and</strong> ii) below for each such pair.<br />

i) Evaluate f S (z, w) := (z 4 + 4w 4 mod 2 64 ).<br />

ii) Use the hash value h( f S (z, w)) <strong>and</strong> linear probing to find a free place in the<br />

hash table <strong>and</strong> store the control value c( f S (z, w)) there.


342 THE DIOPHANTINE EQUATION x 4 + 2y 4 = z 4 + 4w 4 [Chap. XI<br />

B. Reading. Search within the hash table, as follows.<br />

a) Find all pairs (x, y) of non-negative integers less than or equal to B which satisfy<br />

x 4 + 2y 4 ≡ r (mod p p ) <strong>and</strong> all the congruence conditions for primitive solutions,<br />

listed above. (Make systematic use of the Chinese remainder theorem.)<br />

b) Execute steps i) <strong>and</strong> ii) below for each such pair.<br />

i) Evaluate g S (x, y) := (x 4 + 2y 4 mod 2 64 ) on all <strong>points</strong> found in step a).<br />

ii) Search for the control value c(g S (x, y)) in the hash table, starting at the<br />

hash value h(g S (x, y)) <strong>and</strong> using linear probing, until a free position is found.<br />

Report all hits <strong>and</strong> the corresponding values of x <strong>and</strong> y.<br />

7.3. Remarks (Some details of the implementation). —– i) The fourth powers<br />

<strong>and</strong> fourth roots modulo p p are computed during the initialization part of the<br />

program <strong>and</strong> stored into arrays because arithmetic modulo p p is slower than<br />

memory access.<br />

ii) The control value is limited to 31 bits as it is implemented as a signed integer.<br />

We use the value (−1) as a marker for an unoccupied place in the hash table.<br />

iii) In contrast to our previous programs, we do not precompute large tables of<br />

fourth powers modulo 2 64 because an access to these tables is slower than the<br />

execution of two multiplications in a row (at least on our computer).<br />

iv) It is the impact of the congruences modulo 625, 8, <strong>and</strong> 81, described above,<br />

that the set of pairs (y, w) [(x, y)] to be read is significantly bigger than the<br />

set of pairs (x, z) [(z, w)] to be written. They differ actually by a factor of<br />

625 2<br />

· 243<br />

2000·25 112<br />

6252<br />

· 2 ≈ 33.901 in case N <strong>and</strong> · 112<br />

2 000·25 243<br />

≈ 3.601 in case S.<br />

As a consequence of this, only a small part of the running-time is spent on writing.<br />

The lion’s share is spent on unsuccessful searches within L.<br />

7.4. Remarks (Post-processing). —– i) Most of the hits found in the hash table<br />

actually do not correspond to solutions of the Diophantine equation. Hits indicate<br />

only a similarity of bit-patterns. Thus, for each pair of x <strong>and</strong> y reported, one<br />

needs to check whether a suitable pair of z <strong>and</strong> w does exist. We do this by recomputing<br />

z 4 +4w 4 for all z <strong>and</strong> w which fulfill the given congruence conditions<br />

modulo p p <strong>and</strong> powers of the small primes.<br />

Although this method is entirely primitive, only about 3% of the total runningtime<br />

is actually spent on post-processing. One reason for this is that postprocessing<br />

is not called very often, on average only once on about five pages.<br />

For those pages, the writing part of the algorithm needs to be recapitulated.<br />

This is, however, not time-critical as only a small part of the running-time is<br />

spent on writing, anyway.<br />

ii) An interesting alternative for post-processing would be to apply the theory of<br />

binary quadratic forms. The obvious strategy is to factorize x 4 +2y 4 completely


Sec. 7] IMPROVEMENTS II – ADAPTION TO OUR HARDWARE 343<br />

into prime powers <strong>and</strong> to deduce from the decomposition all pairs (a, b) such<br />

that a 2 + b 2 = x 4 + 2y 4 . Then, one may check whether for one of them both<br />

a <strong>and</strong> b are perfect squares.<br />

2<br />

7.5. Remark. –––– The migration to a more bit-based implementation led to<br />

an increase of the speed of our programs by a factor of approximately 1.35.<br />

ii. Adaption to the memory architecture of our computer. —<br />

7.6. –––– The factor of 1.35 is less than what we actually hoped for. For that<br />

reason, we made various tests in order to find out what the limiting bottleneck<br />

of our program is. It turned out that the major slowdown is the access of the<br />

processor to main memory.<br />

Our programs are, in fact, doing only two things, integer arithmetic <strong>and</strong> memory<br />

access. The integer execution units of modern processors are highly optimized<br />

circuits <strong>and</strong> several of them work in parallel inside one processor. They<br />

work a lot faster than main memory does. In order to reach a further improvement,<br />

it will therefore be necessary to take the architecture of memory into<br />

closer consideration.<br />

7.7. The Situation. –––– Computer designers try to bridge the gap between the<br />

fast processor <strong>and</strong> the slow memory by building a memory hierarchy which<br />

consists of several cache levels.<br />

The cache is a very small <strong>and</strong> fast memory inside the processor. The first<br />

cache level, called L1 cache, of our processor consists of a data cache <strong>and</strong> an<br />

instruction cache. Both are 64 kbyte in size. The cache manager stores the most<br />

recently used data into the cache in order to make sure a second access to them<br />

will be fast.<br />

If the cache manager does not find necessary data within the L1 cache then the<br />

processor is forced to wait. In order to deliver data, the cache management first<br />

checks the L2 cache which is 1024 kbyte large. It consists of 16384 lines of<br />

64 byte, each.<br />

7.8. Our Program. –––– Our program fits into the instruction cache, completely.<br />

Therefore, no problem should arise from this.<br />

When we consider the data cache, however, the situation is entirely different.<br />

The cache manager stores the 1024 most recently used memory lines, each being<br />

64 byte long, within the L1 data cache.<br />

This strategy is definitely good for many applications. It guarantees main<br />

memory may be scanned at a high speed. On the other h<strong>and</strong>, for our application,<br />

it fails completely. The reason is that access to our 500 Mbyte hash table is


344 THE DIOPHANTINE EQUATION x 4 + 2y 4 = z 4 + 4w 4 [Chap. XI<br />

completely r<strong>and</strong>om. An access directly to the L1 cache happens in by far less<br />

than 0.1% of the cases. In all other cases, the processor has to wait.<br />

Even worse, it is clear that in most cases we do not even access the L2 cache.<br />

This means, the cache manager needs to access main memory in order to transfer<br />

the corresponding memory line of 64 byte into the L1 cache. After this, the<br />

processor may use the data. In the case that there is no free line available<br />

within the L1 cache, the cache manager must restore old data back to main<br />

memory, first. This process takes us 60 nanoseconds, at least, which seems to<br />

be short, but the processor could execute more than 100 integer instructions<br />

during the same time.<br />

The philosophy for further optimization must, therefore, be to adapt the programs<br />

as much as possible to our hardware, first of all to the sizes of the L1 <strong>and</strong><br />

L2 caches.<br />

7.9. Programmer’s position. –––– Unfortunately, the whole memory hierarchy<br />

is invisible from the point of view of a higher programming language, such asC,<br />

since such languages are designed for being machine-independent. Further, the<br />

hardware executes the cache management in an automatic manner. This means,<br />

even by programming in assembly, one cannot control the cache completely<br />

although some new assembly instructions such as prefetch allow certain direct<br />

manipulations.<br />

7.10. A way out. –––– A practical way, nonetheless to gain some influence on<br />

thememoryhierarchy, istorearrangethealgorithm inanapparentlynonsensical<br />

manner, thereby making memory access less chaotic. One may then hope that<br />

the automatic management of the cache, when confronted with the modified<br />

algorithm, is able to react more properly. This should allow the program to<br />

run faster.<br />

iii. Our first trial. —<br />

7.11. –––– Our first idea for this was to work with two arrays instead of one.<br />

Algorithm M.<br />

i) Store the values of f into an array <strong>and</strong> the values of g into a another one.<br />

Write successively calculated values into successive positions. It is clear that this<br />

part of the algorithm is not troublesome as it involves a linear memory access<br />

which is perfectly supported by the memory management.<br />

ii) Then, use Quicksort in order to sort both arrays. In addition to being<br />

fast, Quicksort is known to have a good memory locality when large arrays<br />

are sorted.


Sec. 7] IMPROVEMENTS II – ADAPTION TO OUR HARDWARE 345<br />

iii) In a final step, search for matches by going linearly through both arrays as<br />

in Mergesort.<br />

7.12. Remark –––– Unfortunately, the idea behind Algorithm M is too simple<br />

to give it any chance of being superior to the previous algorithms. However, it is<br />

a worthwhile experiment. Indeed, our implementation of Algorithm M causes<br />

at least 30 times more memory transfer compared with the previous programs<br />

but, actually, it is only three times slower. This indicates that our approach<br />

is reasonable.<br />

iv. Hashing with partial presorting. —<br />

7.13. –––– Our final algorithm is a combination of sorting <strong>and</strong> hashing. An important<br />

aspect of it is that the sorting step has to be considerably faster than the<br />

Quicksort algorithm. For that reason, we adopted some ideas from linear-time<br />

sorting algorithms such as Radix sort or Bucket sort.<br />

7.14. –––– The algorithm works as follows. Again, the description is based on<br />

case S, case N being analogous.<br />

Algorithm H64B.<br />

I. Initialization. Fix B := 10 8 . Initialize a hash table H of 2 27 = 134 217 728<br />

integers, each being 32 bit long. Fix the page prime p p := 200 003.<br />

In addition, initialize 1024 auxiliary arrays A i each of which may contain<br />

2 17 = 131 072 long (64 bit) integers.<br />

Further, define two functions, the hash function h <strong>and</strong> the control function c,<br />

which map 64 bit integers to 27 bit integers <strong>and</strong> 31 bit integers, respectively, by<br />

selecting certain bits. Do not use any of the bits twice to ensure h <strong>and</strong> c are<br />

independent on each other <strong>and</strong> do not use the four least significant bits.<br />

Finally, let h (10) denote the function mapping 64 bit integers to integers<br />

within [0, 1023] which is given by the ten most significant bits of h. In other<br />

words, for every x, h (10) (x) is the same as h(x) shifted to the right by 17 bits.<br />

II) Outer Loop. Let r run from 0 to p p − 1 <strong>and</strong> execute A. <strong>and</strong> B. for each r.<br />

A. Writing. Build up the hash table, which is meant to encode the set L r ,<br />

as follows.<br />

a) Preparation. Find all pairs (z, w) of non-negative integers less than or equal<br />

to B which satisfy z 4 + 4w 4 ≡ r (mod p p ) <strong>and</strong> all the congruence-conditions<br />

for primitive solutions, listed above. (Make systematic use of the Chinese<br />

remainder theorem.)<br />

b) Inner Loop. Execute steps i) – iii) below for each such pair.


346 THE DIOPHANTINE EQUATION x 4 + 2y 4 = z 4 + 4w 4 [Chap. XI<br />

i) Evaluate f S (z, w) := (z 4 + 4w 4 mod 2 64 ).<br />

ii) Do not store f S (z, w) into the hash table, immediately. Put<br />

first.<br />

i := h (10) ( f S (z, w)),<br />

iii) Add f S (z, w) to the auxiliary array A i . Maintain A i as an unordered list.<br />

I.e., always write to the lowest unoccupied address.<br />

If there is no space left in A i then output an error message <strong>and</strong> abort the algorithm.<br />

c) Storing. Let i run from 0 to 1023. For each i let j run through the addresses<br />

occupied in A i .<br />

For fixed i <strong>and</strong> j, extract from the 64 bit integer A i [j] the 27 bit hash value<br />

h(A i [j]) <strong>and</strong> the 31 bit control value c(A i [j]).<br />

Usethehashvalueh(A i [j])<strong>and</strong>linear probingtofindafreeplaceinthehashtable<br />

<strong>and</strong> store the control value c(A i [j]) there.<br />

d) Clearing up. Clear the auxiliary arrays A i for all i ∈ [0, 1023] to make them<br />

available for reuse.<br />

B. Reading. Search within the hash table, as follows.<br />

a) Preparation. Find all pairs (x, y) of non-negative integers less than or equal<br />

to B which satisfy x 4 + 2y 4 ≡ r (mod p p ) <strong>and</strong> all the congruence conditions<br />

for primitive solutions, listed above. (Make systematic use of the Chinese<br />

remainder theorem.)<br />

b) Inner Loop. Execute steps i) – iii) below for each such pair.<br />

i) Evaluate g S (x, y) := (x 4 + 2y 4 mod 2 64 ).<br />

ii) Do not look up g S (x, y) in the hash table, immediately. Put<br />

first.<br />

i := h (10) (g S (x, y)),<br />

iii) Add g S (x, y) to the auxiliary array A i . Maintain A i as an unordered list.<br />

I.e., always write to the lowest unoccupied address.<br />

If there is no space left in A i then call d[i] <strong>and</strong> add g S (x, y) to A i , afterwards.<br />

c) Searching. Clearing all buffers. Let i run from 0 to 1023. For each i, call d[i].<br />

When this is finished, go to the next iteration of the outer loop.


Sec. 7] IMPROVEMENTS II – ADAPTION TO OUR HARDWARE 347<br />

Subroutine d[i]) Clearing a buffer. Let j run through the addresses occupied<br />

in A i . For fixed j, search for the control value c(A i [j]) within the hash table H ,<br />

starting at the hash value h(A i [j]) <strong>and</strong> using linear probing, until a free place<br />

is found. Report all hits <strong>and</strong> the corresponding values of x <strong>and</strong> y.<br />

Having done this, declare A i to be empty.<br />

7.15. Remark –––– The auxiliary arrays A i play the role of a buffer. Thus, one<br />

could say that we introduced some buffering into the management of the hash table<br />

H . However, this description misses the point.<br />

What is more important is that the values of f S to be stored into L r are partially<br />

sorted according to the 10 most significant bits of h( f S (z, w)) by putting them<br />

into the auxiliary arrays A i . When the hash table is then built up, the records<br />

arrive almost in order. The same is true for reading.<br />

What we actually did is, therefore, to introduce some partial presorting into the<br />

management of the hash table.<br />

7.16. Remark –––– It is our experience that each auxiliary array carries more<br />

or less the same load. In particular, in step II.A.b.iii), when the buffers are filled<br />

up for writing, a buffer overflow should never occur. For this reason, we feel<br />

free to treat this possibility as a fatal error.<br />

v. Running time. —<br />

7.17. –––– Algorithm H64B uses about three times more memory than our<br />

previous algorithms but our implementation runs almost three times as fast.<br />

It was this factor which made it possible to attack the bound B = 10 8 in a<br />

reasonable amount of time.<br />

The final version of our programs took almost exactly 100 days of CPU time<br />

on an AMD Opteron 248 processor. This time is composed almost equally<br />

of 50 days for case N <strong>and</strong> 50 days for case S. The main computation was<br />

executed in parallel on two machines in February <strong>and</strong> March, 2005.<br />

7.18. Why is this algorithm faster? –––– To answer this question, one has to<br />

look at the impact of the cache. For the old program, the cache memory was<br />

mostly useless. For the new program, the situation is completely different.<br />

When the auxiliary arrays are filled in step II.A.b.iii) <strong>and</strong> II.B.b.iii), access to<br />

these arrays is linear. There are only 1024 of them which is exactly the number<br />

of lines in the L1 cache. When an access does not hit into that innermost cache<br />

then the corresponding memory line is moved to it <strong>and</strong> the next seven accesses<br />

to the same auxiliary array are accesses to that line. Altogether, seven of eight<br />

memory accesses hit into the L1 cache.


348 THE DIOPHANTINE EQUATION x 4 + 2y 4 = z 4 + 4w 4 [Chap. XI<br />

When an auxiliary array is emptied in step II.A.c) or II.B.d[i]), the situation<br />

is similar. There are a high number of accesses to a very short segment of the<br />

hash table. This segment fits completely into the L2 cache. It has to be moved<br />

into that cache, once. Then, it can be used many times. Again, access to the<br />

auxiliary array is linear <strong>and</strong> a hit into the L1 cache occurs in seven of eight cases.<br />

All in all, for Algorithm H64B, most memory accesses are hits into the cache.<br />

This means, at the cost of some more data transfer altogether, we achieved that<br />

main memory may be mostly used at the speed of the cache.<br />

8. The solution found<br />

8.1. –––– Ironically, the premature version of our algorithm as described in<br />

Section 4 already solved Problem 1.5.<br />

The improvements made it possible to shift the search bound to 10 8 which<br />

was far beyond our original expectations. However, they did not produce any<br />

new solution.<br />

Thus, let us conclude this chapter by some comments on the simpler approach.<br />

8.2. –––– Test versions of the program were written in Delphi.<br />

The definitive version was written in C. It took about 130 hours of CPU time<br />

on a 3.00 GHz Pentium 4 processor with 512 kbyte cache memory. The main<br />

computation was executed in parallel on two machines during the very first<br />

days of December, 2004.<br />

8.3. –––– Instead of looking for solutions of x 4 + 2y 4 = z 4 + 4w 4 , the algorithm<br />

searches, in fact, for solutions to the corresponding simultaneous congruences<br />

modulo p p <strong>and</strong> p c which, in addition, fulfill that ( (x 4 + 2y 4 ) mod p h<br />

)<br />

<strong>and</strong> ( (z 4 + 4w 4 ) mod p h<br />

)<br />

are “almost equal”.<br />

To this modified problem, we found approximately 3800 solutions such that<br />

(y, w) ≠ (0, 0). These congruence solutions were checked by an exact computation<br />

using O. Forster’s [Fo] Pascal-style multi-precision interpreter language<br />

ARIBAS.<br />

8.4. –––– Among the congruence solutions, exact equality occurred only once.<br />

This solution is as follows.<br />

==> 1484801**4 + 2 * 1203120**4.<br />

-: 90509_10498_47564_80468_99201<br />

==> 1169407**4 + 4 * 1157520**4.<br />

-: 90509_10498_47564_80468_99201


CHAPTER XII<br />

NEW SUMS OF THREE CUBES ∗<br />

There are three cubes (levels) in total, each of which are connected by a<br />

single door. Players have the goal of moving the character from room to<br />

room, cube to cube in an attempt to find the final exit door of all three cubes.<br />

If this door is found the character will appear to leave the cube, walk across<br />

the table surface <strong>and</strong> vanish. The game then begins again.<br />

JULIAN OLIVER: levelHead, A 3D Spatial Memory Game (2007)<br />

1. Introduction<br />

It is a long st<strong>and</strong>ing problem as to whether every <strong>rational</strong> integer<br />

n ≢ 4, 5 (mod 9) can be written as a sum of three integral cubes. According<br />

to the web page [Be-list] of Daniel Bernstein, the first attacks by computer<br />

were carried out as early as 1955.<br />

Nevertheless, for example for n = 3, there is still no solution known apart from<br />

the obvious ones: (1, 1, 1), (4, 4, −5), (4, −5, 4), <strong>and</strong> (−5, 4, 4). For n = 30, the<br />

first solution was found by N. Elkies <strong>and</strong> his coworkers in 2000 [El]. It is interesting<br />

to note that, in 1992, D. R. Heath-Brown [H-B92a] had made a prediction<br />

on the density of the solutions for n = 30 without knowing any solution explicitly.<br />

Over the years, a number of algorithms have been developed in order to attack<br />

the general problem. An excellent overview concerning the various approaches<br />

invented up to around 2000 has been given in [B/P/T/Y], published recently.<br />

The historically first algorithm which has a complexity of O(B 1+ε ) for a search<br />

bound of B is the method of R. Heath-Brown [H/L/R]. The best algorithm<br />

presently known is Elkies’ method described in [El].<br />

(∗) This chapter is a revised version of the article: New sums of three cubes, to appear in:<br />

Math. Comp., joint with A.-S. Elsenhans.


350 NEW SUMS OF THREE CUBES [Chap. XII<br />

2. Elkies’ method<br />

This<br />

√<br />

algorithm is geometric<br />

√<br />

in nature. The idea is to cover the curve<br />

Y = 3 1 − X 3<br />

, X ∈ [0, 1/ 3 2], by very small parallelograms which we call flagstones.<br />

The algorithm finds all <strong>rational</strong> <strong>points</strong> of the particular form (x/z, y/z)<br />

which are contained in one of the flagstones for z ∈Æup to a given bound B.<br />

(This means we are searching for triples such that x 3 + y 3 − z 3 will be small.)<br />

For each flagstone, this is equivalent to the detection of all <strong>points</strong> of the st<strong>and</strong>ard<br />

lattice3 that are contained in a certain pyramid. The problem is that, viewed<br />

in the st<strong>and</strong>ard coordinates, this pyramid has an enormous height in comparison<br />

with the two other dimensions. Thus, it has an extremely sharp apex. Searching<br />

naively for lattice <strong>points</strong> in such a pyramid would be highly inefficient.<br />

The idea to overcome this difficulty is to work in coordinates more<br />

adapted to this pyramid. The drawback in this case is that the basis<br />

{ (1, 0, 0), (0, 1, 0), (0, 0, 1) } of the st<strong>and</strong>ard lattice is then far from being<br />

reduced in whatever sense. One needs to apply lattice basis reduction. Having<br />

done that, searching for lattice <strong>points</strong> within the pyramid is essentially<br />

equivalent to a search for small <strong>points</strong> of the lattice. For this, one may use the<br />

well-known algorithm of Fincke-Pohst [F/P].<br />

The size of the flagstones is somewhat arbitrary. Smaller flagstones mean that<br />

more time is required for lattice basis reductions. Larger ones lead to more time<br />

being spent on the algorithm of Fincke-Pohst. The optimum size depends on<br />

details of the implementation.<br />

3. Implementation<br />

Our implementation of Elkies’ method is written in C <strong>and</strong> C++. We took care<br />

that only initialization parts of the code were written inC++ or made use of the<br />

multi precision floats of GMP.<br />

The time-critical parts were written in plain C making some use of the instruction<br />

asm. It turned out that, for most of the computations, 128-bit fixed-point<br />

arithmetic was sufficiently precise. We realized the 128-bit fixed-point numbers<br />

as arrays consisting of two long ints. The arithmetic of the fixed-point<br />

numbers was implemented in such a way that all loops (of length two) were<br />

manually unrolled.<br />

For lattice basis reduction, we implemented a version of LLL for three-dimensional<br />

lattices. It turned out that adjacent flagstones led to similar reduced bases.<br />

That is why enormous savings could be achieved by doing LLL incrementally.


Sec. 4] RESULTS 351<br />

We start theLLL computation for the next flagstone with a reduced basis of the<br />

previous one <strong>and</strong> not with a naive basis.<br />

Another substantial improvement was realized in the Fincke-Pohst part.<br />

Here, one has to compute many adjacent values of the same cubic polynomial<br />

in three variables. An implementation of a difference scheme reduced most<br />

of these computations to a few additions of values obtained before.<br />

Some details. — We searched systematically for solutions of x 3 + y 3 + z 3 = n<br />

where the positive integer n < 1000 is neither a cube nor twice a cube <strong>and</strong><br />

| x|, |y|, |z| < 10 14 . The length of the flagstones was chosen dynamically. It was<br />

around 8.4 · 10 −12 near x = 0 <strong>and</strong> around 6.6 · 10 −14 near x = 1/ 3 √<br />

2. The area<br />

of the flagstones was essentially constant at a value near 1.7 · 10 −40 . This led to<br />

a total number of a bit more than 10 13 flagstones to deal with.<br />

We chose the widths of the flagstones such that all <strong>points</strong> in a horizontal<br />

distance of < 10 −30 from the curve are contained in one of the flagstones.<br />

This should make sure that all solutions of heights between 10 11 <strong>and</strong> 10 14 are<br />

certainly found. Indeed, if we arrange variables such that | x| ≤ |y| ≤ |z|<br />

then the point (| x/z|, |y/z|) is at a horizontal distance from the curve of,<br />

√<br />

in<br />

ds<br />

first order approximation, 1/3<br />

|<br />

ds s=(1−X 3 ) · n/|z| 3 . Since X := | x/z| < 1/ 3 2,<br />

the derivative is always less than 0.53.<br />

The whole search took around ten months of CPU time. Only 14% of that<br />

time was spent on lattice basis reductions. The lion’s share was spent searching<br />

for small lattice <strong>points</strong>. I.e., on our implementation of the algorithm of Fincke-<br />

Pohst.<br />

4. Results<br />

In comparison with the computations of [B/P/T/Y] <strong>and</strong> the lists, dating back<br />

to 2001 <strong>and</strong> published in [Be-list], 3519 new solutions have been found.<br />

Among them, there are the following. For each of the nine numbers on the<br />

left, no solution had been given before, neither in D. Bernstein’s lists, nor<br />

in [B/P/T/Y].<br />

156 = 26 577 110 807 569 3 − 18 161 093 358 005 3 − 23 381 515 025 762 3<br />

318 = 1 970 320 861 387 3 + 1 750 553 226 136 3 − 2 352 152 467 181 3<br />

= 30 828 727 881 037 3 + 27 378 037 791 169 3 − 36 796 384 363 814 3<br />

366 = 241 832 223 257 3 + 167 734 571 306 3 − 266 193 616 507 3<br />

420 = 8 859 060 149 051 3 − 2 680 209 928 162 3 − 8 776 520 527 687 3


352 NEW SUMS OF THREE CUBES [Chap. XII<br />

564 = 53 872 419 107 3 − 1 300 749 634 3 − 53 872 166 335 3<br />

758 = 662 325 744 409 3 + 109 962 567 936 3 − 663 334 553 003 3<br />

= 83 471 297 139 078 3 + 77 308 024 343 011 3 − 101 433 242 878 565 3<br />

789 = 18 918 117 957 926 3 + 4 836 228 687 485 3 − 19 022 888 796 058 3<br />

894 = 19 868 127 639 556 3 + 2 322 626 411 251 3 − 19 878 702 430 997 3<br />

948 = 103 458 528 103 519 3 + 6 604 706 697 037 3 − 103 467 499 687 004 3<br />

For 13 values of n, for which exactly one solution was known, we found a<br />

second one. Among those, there is n = 30. The second solution for n = 30 is<br />

30 = 3 982 933 876 681 3 − 636 600 549 515 3 − 3 977 505 554 546 3 .<br />

A second <strong>and</strong> a third solution for n = 75 are<br />

75 = 2 576 191 140 760 3 + 1 217 343 443 218 3 − 2 663 786 047 493 3<br />

= 59 897 299 698 355 3 − 47 258 398 396 091 3 − 47 819 328 945 509 3 .<br />

On the other h<strong>and</strong>, the highest numbers of solutions found are 93 for n = 792<br />

<strong>and</strong> 85 for n = 720.<br />

This fits well with some speculations made in [H-B92a]. D. R. Heath-Brown<br />

predicts ∼A(n) logB essentially different solutions of height


APPENDIX<br />

Diophantus, with this name which is frequent in Greece, was a true Greek,<br />

disciple of Greek science, if one who towers high above his contemporaries.<br />

He was Greek in what he accomplished, as well as in what he was not able<br />

to accomplish. But we must not forget that Greek science, as it conquered<br />

the East from Alex<strong>and</strong>ria . . . , brought new ideas back home from these<br />

campaigns, that Greek mathematics as such has ceased to pick up whatever<br />

it found worth picking up here <strong>and</strong> there.<br />

MORITZ CANTOR (1907, translated by N. Schappacher)<br />

1. A script in GAP<br />

This is the script in GAP which was used to compute H 1( G, Pic(XÉ) )<br />

<strong>and</strong> rkPic(X ) for all <strong>groups</strong> which may operate on the 27 lines upon a smooth<br />

cubic surface. The mathematical background is explained in Section 8, Subsection<br />

iv of Chapter II.<br />

TABLE 1. A GAP script<br />

# A script in GAP to compute H^1(G, Pic) = (NS \cap S_0) / NS_0 <strong>and</strong> the Picard rank for all<br />

# Galois <strong>groups</strong> of a cubic surface.<br />

# Here, S is the free abelian group on the 27 lines, S_0 \subset S the subgroup of all<br />

# principal divisors, <strong>and</strong> N: S --> S the norm map.<br />

# We compute the intersection matrix of the 27 lines.<br />

# g must be a faithful representation of W(E_6) in S_27.<br />

SchnittMatrixBerechnen := function (g)<br />

local i, j, mat, len, tmp;<br />

mat := NullMat(27, 27);<br />

for i in [1..27] do<br />

for j in [1..27] do<br />

if (i = j) then<br />

mat[i][j] := -1;<br />

else<br />

tmp := [i, j]; Sort(tmp);<br />

len := Size(Orbit(g, tmp, OnSets));<br />

# Operation of W(E_6) on unordered pairs of lines.<br />

# A pair of skew lines has orbit length 27*8.<br />

# A pair of intersecting lines has orbit length 27*5.<br />

if (len = 27*5) then<br />

mat[i][j] := 1;<br />

fi;<br />

fi;<br />

od;<br />

od;<br />

return mat;


354 APPENDIX<br />

end;<br />

# Compute the 27x27-matrix representing the norm map N: S --> S.<br />

NBerechnen := function (orb, u)<br />

local N, akt, j, k, l;<br />

# The matrix N.<br />

N := NullMat(27, 27);<br />

for j in [1..Size(orb)] do<br />

akt := orb[j];<br />

for k in [1..Size(akt)] do<br />

for l in [1..Size(akt)] do<br />

N[akt[k]][akt[l]] := Size(u) / Size(akt);<br />

od;<br />

od;<br />

od;<br />

return N;<br />

end;<br />

# Build up a matrix containing a minimal system of generators of NS in its columns.<br />

IBerechnen := function (orb, u)<br />

local my_I, j, k, akt;<br />

# The matrix I.<br />

my_I := NullMat(27, Size(orb));<br />

for j in [1..Size(orb)] do<br />

akt := orb[j];<br />

for k in [1..Size(akt)] do<br />

# The line akt[k] lies in orbit akt numbered j.<br />

my_I[akt[k]][j] := Size(u) / Size(akt);<br />

od;<br />

od;<br />

return my_I;<br />

end;<br />

# A separate routine in order to make a lattice base from the column vectors of a matrix.<br />

NormMat := function (m)<br />

return TransposedMat(HermiteNormalFormIntegerMat(BaseIntMat(TransposedMat(m))));<br />

end;<br />

# The discriminant of a lattice. The lattice needs not be maximal.<br />

# The lattice base is supposed to be given in the column vectors of m.<br />

# It is important that m is a lattice base. The function will return 0 for a linearly<br />

# dependent system.<br />

Disc := function (m)<br />

return Determinant(TransposedMat(m) * m);<br />

end;<br />

# We compute H^1 of the Picard group <strong>and</strong> the arithmetic Picard rank using Manin’s formulas.<br />

h1_pic := function (schnitt_matrix, u)<br />

local orb, I_mat, AI, N, K, linke_seite, rechte_seite, disc_links,<br />

disc_rechts, index, rang, mul, scal_links, schnitt;<br />

# Everything depends only on the combinatorial structure of the orbits.<br />

orb := Orbits(u, [1..27]);<br />

I_mat := IBerechnen(orb, u); # I_mat represents NS.<br />

AI := schnitt_matrix * I_mat;<br />

N := NBerechnen(orb, u); # N represents the norm map.<br />

# Normalization: We choose all integral vectors in the kernel.


APPENDIX 355<br />

# Computation of NS \cap S_0.<br />

K := NullspaceIntMat(TransposedMat(AI));<br />

# K collects the relations among the generators of NS modulo S_0.<br />

linke_seite := NormMat(I_mat * TransposedMat(K));<br />

# linke_seite represents the lattice NS \cap S_0.<br />

K := NullspaceIntMat(TransposedMat(schnitt_matrix));<br />

# The rows of K are generators of S_0.<br />

rechte_seite := NormMat(N * TransposedMat(K));<br />

# The columns of N * TransposedMat(K) are generators of NS_0.<br />

# The discriminats.<br />

disc_links := Disc(linke_seite);<br />

disc_rechts := Disc(rechte_seite);<br />

index := RootInt(disc_rechts / disc_links);<br />

rang := Size(orb) - Size(TransposedMat(rechte_seite));<br />

# The order of H^1(G, Pic).<br />

# The arithmetic Picard rank.<br />

AppendTo("h1pic.txt"," #H^1 = ", index);<br />

if index > 3 then<br />

# We have to compute the primary decomposition.<br />

# This code suffices since index may be only 1, 2, 3, 4, or 9.<br />

if index = 4 then mul := 2; else mul := 3; fi;<br />

scal_links := mul * linke_seite;<br />

schnitt := TransposedMat(BaseIntersectionIntMats(TransposedMat(rechte_seite),<br />

TransposedMat(scal_links)));<br />

if Disc(schnitt) = Disc(scal_links) then<br />

AppendTo("h1pic.txt"," [ ", mul, ", ", mul, " ]");<br />

else<br />

AppendTo("h1pic.txt"," More complicated than the direct sum of Z/", mul, "Z");<br />

fi;<br />

fi;<br />

AppendTo("h1pic.txt", ", Rk(Pic) = ", rang, ",");<br />

end;<br />

# Adds the orbit structure of the lines to the list.<br />

bahnstruktur := function (u)<br />

local ol;<br />

ol := ShallowCopy(OrbitLengths(u, [1..27]));<br />

Sort(ol);<br />

AppendTo("h1pic.txt", " Orbits ", ol);<br />

end;<br />

# Our group taken from the library.<br />

we6 := TransitiveGroup(27, 1161);<br />

schnitt_matrix := SchnittMatrixBerechnen(we6);<br />

# The sub<strong>groups</strong>.<br />

ugv := ConjugacyClassesSub<strong>groups</strong>(we6);;<br />

# We compute H^1(G, Pic) for all Galois <strong>groups</strong>.<br />

for i in [1..Size(ugv)] do<br />

u := Representative(ugv[i]);<br />

AppendTo("h1pic.txt", i, " #U = ", Size(u), " ");<br />

AppendTo("h1pic.txt", AbelianInvariants(u), ",");<br />

if IsTransitive(u) <strong>and</strong> (Size(u) > 1) then<br />

AppendTo("h1pic.txt", " transitive");<br />

else<br />

h1_pic(schnitt_matrix, u);<br />

bahnstruktur(u);<br />

fi;<br />

AppendTo("h1pic.txt", "\n");<br />

od;


356 APPENDIX<br />

2. The list<br />

This is the list produced by theGAP script. For each subgroup of W (E 6 ), we give<br />

its abelian quotient, the corresponding values of H 1( G, Pic(XÉ) ) <strong>and</strong> rkPic(X ),<br />

<strong>and</strong> the orbit structure of the 27 lines.<br />

We make use of this list in several ways. For our presentation of the Brauer-<br />

Manin obstruction in the case of diagonal cubic surfaces given in Section III.6,<br />

the list is of fundamental importance. On the other h<strong>and</strong>, we also use it in the<br />

arguments given in III.5.33 <strong>and</strong> Remarks III.5.37.<br />

It is further applied in Examples VI.5.7 <strong>and</strong> VI.5.8. Here, the goal is simply to<br />

show that one has Picard rank 2 in the situations considered.<br />

Finally, we give reference to the list in Remarks VI.5.13.ii) <strong>and</strong> X.2.4 in order<br />

to prove that the factor β is bounded from above by 9.<br />

TABLE 2. H 1 (G, Pic) <strong>and</strong> rk Pic(X ) for smooth cubic surfaces<br />

1 #U = 1 [ ], #H^1 = 1, Rk(Pic) = 7, Orbits [ 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1,<br />

1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1 ]<br />

2 #U = 2 [ 2 ], #H^1 = 1, Rk(Pic) = 6, Orbits [ 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1,<br />

1, 2, 2, 2, 2, 2, 2 ]<br />

3 #U = 2 [ 2 ], #H^1 = 4 [ 2, 2 ], Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 2, 2, 2, 2, 2, 2, 2,<br />

2, 2, 2, 2 ]<br />

4 #U = 2 [ 2 ], #H^1 = 1, Rk(Pic) = 5, Orbits [ 1, 1, 1, 1, 1, 1, 1, 2, 2, 2, 2, 2, 2, 2,<br />

2, 2, 2 ]<br />

5 #U = 2 [ 2 ], #H^1 = 1, Rk(Pic) = 4, Orbits [ 1, 1, 1, 2, 2, 2, 2, 2, 2, 2, 2, 2, 2,<br />

2, 2 ]<br />

6 #U = 3 [ 3 ], #H^1 = 9 [ 3, 3 ], Rk(Pic) = 1, Orbits [ 3, 3, 3, 3, 3, 3, 3, 3, 3 ]<br />

7 #U = 3 [ 3 ], #H^1 = 1, Rk(Pic) = 5, Orbits [ 1, 1, 1, 1, 1, 1, 1, 1, 1, 3, 3, 3, 3,<br />

3, 3 ]<br />

8 #U = 3 [ 3 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 3, 3, 3, 3, 3, 3, 3, 3, 3 ]<br />

9 #U = 4 [ 4 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 4, 4, 4, 4, 4, 4 ]<br />

10 #U = 4 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 5, Orbits [ 1, 1, 1, 1, 1, 1, 1, 2, 2, 2, 2, 2, 2,<br />

2, 2, 4 ]<br />

11 #U = 4 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 4, Orbits [ 1, 1, 1, 1, 1, 2, 2, 2, 4, 4, 4, 4 ]<br />

12 #U = 4 [ 2, 2 ], #H^1 = 4 [ 2, 2 ], Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 2, 2, 4, 4, 4, 4 ]<br />

13 #U = 4 [ 4 ], #H^1 = 4 [ 2, 2 ], Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 4, 4, 4, 4, 4 ]<br />

14 #U = 4 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 2, 2, 2, 4, 4, 4, 4 ]<br />

15 #U = 4 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 4, Orbits [ 1, 1, 1, 2, 2, 2, 2, 2, 2, 4, 4, 4 ]<br />

16 #U = 4 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 2, 2, 2, 2, 2, 4, 4, 4 ]<br />

17 #U = 4 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 2, 4, 4, 4, 4, 4 ]<br />

18 #U = 4 [ 4 ], #H^1 = 1, Rk(Pic) = 4, Orbits [ 1, 1, 1, 1, 1, 2, 4, 4, 4, 4, 4 ]<br />

19 #U = 4 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 2, 2, 2, 4, 4, 4, 4 ]<br />

20 #U = 4 [ 4 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 2, 4, 4, 4, 4, 4 ]<br />

21 #U = 4 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 4, 4, 4, 4, 4 ]<br />

22 #U = 4 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 4, Orbits [ 1, 1, 1, 2, 2, 2, 2, 2, 2, 2, 2, 4, 4 ]<br />

23 #U = 5 [ 5 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 5, 5, 5, 5, 5 ]<br />

24 #U = 6 [ 2 ], #H^1 = 1, Rk(Pic) = 5, Orbits [ 1, 1, 1, 1, 1, 1, 1, 1, 1, 3, 3, 3, 3,<br />

3, 3 ]<br />

25 #U = 6 [ 2 ], #H^1 = 4 [ 2, 2 ], Rk(Pic) = 1, Orbits [ 3, 3, 3, 6, 6, 6 ]<br />

26 #U = 6 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 6, 6, 6, 6 ]<br />

27 #U = 6 [ 2, 3 ], #H^1 = 4 [ 2, 2 ], Rk(Pic) = 1, Orbits [ 3, 6, 6, 6, 6 ]<br />

28 #U = 6 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 4, Orbits [ 1, 1, 1, 2, 2, 2, 3, 3, 3, 3, 6 ]<br />

29 #U = 6 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 3, 3, 6, 6, 6 ]<br />

30 #U = 6 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 2, 2, 6, 6, 6 ]<br />

31 #U = 6 [ 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 2, 2, 6, 6, 6 ]<br />

32 #U = 6 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 3, 3, 3, 3, 6, 6 ]


APPENDIX 357<br />

33 #U = 6 [ 2 ], #H^1 = 1, Rk(Pic) = 4, Orbits [ 1, 1, 1, 2, 2, 2, 3, 3, 3, 3, 6 ]<br />

34 #U = 6 [ 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 3, 3, 3, 3, 3, 3, 3, 6 ]<br />

35 #U = 6 [ 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 2, 2, 3, 3, 6, 6 ]<br />

36 #U = 6 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 2, 2, 3, 3, 6, 6 ]<br />

37 #U = 6 [ 2 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 3, 3, 3, 6, 6, 6 ]<br />

38 #U = 6 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 3, 3, 6, 6, 6 ]<br />

39 #U = 6 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 6, 6, 6, 6 ]<br />

40 #U = 8 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 8, 8, 8 ]<br />

41 #U = 8 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 2, 2, 2, 8, 8 ]<br />

42 #U = 8 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 3, Orbits [ 1, 1, 1, 4, 4, 4, 4, 4, 4 ]<br />

43 #U = 8 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 2, 2, 8, 8 ]<br />

44 #U = 8 [ 2, 2 ], #H^1 = 4 [ 2, 2 ], Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 4, 4, 4, 4, 4 ]<br />

45 #U = 8 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 4, 4, 8, 8 ]<br />

46 #U = 8 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 4, Orbits [ 1, 1, 1, 2, 2, 2, 2, 2, 2, 4, 4, 4 ]<br />

47 #U = 8 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 2, 2, 2, 4, 4, 8 ]<br />

48 #U = 8 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 3, Orbits [ 1, 1, 1, 4, 4, 4, 4, 4, 4 ]<br />

49 #U = 8 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 4, Orbits [ 1, 1, 1, 1, 1, 2, 4, 4, 4, 4, 4 ]<br />

50 #U = 8 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 4, 8, 8 ]<br />

51 #U = 8 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 4, 8, 8 ]<br />

52 #U = 8 [ 2, 4 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 8, 8 ]<br />

53 #U = 8 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 2, 4, 4, 4, 4, 4 ]<br />

54 #U = 8 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 4, 4, 4, 4, 8 ]<br />

55 #U = 8 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 4, 8, 8 ]<br />

56 #U = 8 [ 2, 4 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 4, 8, 8 ]<br />

57 #U = 8 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 2, 4, 8, 8 ]<br />

58 #U = 8 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 2, 4, 4, 4, 8 ]<br />

59 #U = 8 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 2, 4, 4, 4, 8 ]<br />

60 #U = 8 [ 2, 4 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 4, 4, 8 ]<br />

61 #U = 8 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 2, 4, 4, 4, 4, 4 ]<br />

62 #U = 8 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 8, 8 ]<br />

63 #U = 8 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 2, 4, 8, 8 ]<br />

64 #U = 8 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 4, 4, 8 ]<br />

65 #U = 8 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 4, 4, 8 ]<br />

66 #U = 8 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 8, 8 ]<br />

67 #U = 8 [ 8 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 8, 8 ]<br />

68 #U = 8 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 2, 4, 4, 4, 8 ]<br />

69 #U = 8 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 4, 4, 4, 8 ]<br />

70 #U = 9 [ 3, 3 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 3, 3, 3, 3, 3, 3, 9 ]<br />

71 #U = 9 [ 3, 3 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 9, 9, 9 ]<br />

72 #U = 9 [ 3, 3 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 3, 3, 3, 9, 9 ]<br />

73 #U = 9 [ 9 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 9, 9 ]<br />

74 #U = 10 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 2, 5, 5, 5, 10 ]<br />

75 #U = 10 [ 2, 5 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 2, 5, 5, 5, 10 ]<br />

76 #U = 10 [ 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 5, 5, 5, 5, 5 ]<br />

77 #U = 12 [ 3 ], #H^1 = 1, Rk(Pic) = 4, Orbits [ 1, 1, 1, 1, 1, 4, 4, 4, 4, 6 ]<br />

78 #U = 12 [ 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 6, 6, 12 ]<br />

79 #U = 12 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 4, Orbits [ 1, 1, 1, 2, 2, 2, 3, 3, 3, 3, 6 ]<br />

80 #U = 12 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 2, 2, 6, 6, 6 ]<br />

81 #U = 12 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 3, 3, 6, 12 ]<br />

82 #U = 12 [ 2, 2 ], #H^1 = 4 [ 2, 2 ], Rk(Pic) = 1, Orbits [ 3, 6, 6, 6, 6 ]<br />

83 #U = 12 [ 2, 2, 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 2, 4, 6, 12 ]<br />

84 #U = 12 [ 3, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 4, 6, 12 ]<br />

85 #U = 12 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 2, 4, 6, 12 ]<br />

86 #U = 12 [ 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 4, 6, 12 ]<br />

87 #U = 12 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 3, 3, 4, 6, 6 ]<br />

88 #U = 12 [ 2, 2, 3 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 3, 3, 4, 6, 6 ]<br />

89 #U = 12 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 2, 2, 3, 3, 6, 6 ]<br />

90 #U = 12 [ 3, 4 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 12, 12 ]<br />

91 #U = 12 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 6, 6, 12 ]<br />

92 #U = 12 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 6, 6, 12 ]<br />

93 #U = 12 [ 2, 2, 3 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 6, 6, 12 ]<br />

94 #U = 12 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 3, 3, 6, 6, 6 ]<br />

95 #U = 12 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 6, 6, 12 ]<br />

96 #U = 12 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 2, 4, 6, 12 ]<br />

97 #U = 12 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 3, 3, 4, 6, 6 ]


358 APPENDIX<br />

98 #U = 12 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 6, 6, 6, 6 ]<br />

99 #U = 16 [ 2, 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 2, 2, 16 ]<br />

100 #U = 16 [ 2, 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 3, Orbits [ 1, 1, 1, 4, 4, 4, 4, 4, 4 ]<br />

101 #U = 16 [ 2, 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 4, 16 ]<br />

102 #U = 16 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 8, 8, 8 ]<br />

103 #U = 16 [ 2, 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 4, 8, 8 ]<br />

104 #U = 16 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 4, 4, 8, 8 ]<br />

105 #U = 16 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 16 ]<br />

106 #U = 16 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 4, 4, 8, 8 ]<br />

107 #U = 16 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 16 ]<br />

108 #U = 16 [ 4, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 16 ]<br />

109 #U = 16 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 8, 8 ]<br />

110 #U = 16 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 2, 4, 8, 8 ]<br />

111 #U = 16 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 16 ]<br />

112 #U = 16 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 8, 8 ]<br />

113 #U = 16 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 16 ]<br />

114 #U = 16 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 8, 8 ]<br />

115 #U = 16 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 16 ]<br />

116 #U = 16 [ 2, 2, 4 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 8, 8 ]<br />

117 #U = 16 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 4, 16 ]<br />

118 #U = 16 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

119 #U = 16 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 4, 8, 8 ]<br />

120 #U = 16 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 4, 16 ]<br />

121 #U = 16 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 4, 4, 8, 8 ]<br />

122 #U = 16 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

123 #U = 16 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 4, 4, 8 ]<br />

124 #U = 16 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 8, 8 ]<br />

125 #U = 16 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 2, 4, 4, 4, 8 ]<br />

126 #U = 16 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 8, 8 ]<br />

127 #U = 16 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 8, 8 ]<br />

128 #U = 18 [ 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 3, 3, 3, 3, 3, 3, 9 ]<br />

129 #U = 18 [ 2, 3 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 3, 3, 3, 9, 9 ]<br />

130 #U = 18 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 3, 3, 3, 3, 3, 3, 9 ]<br />

131 #U = 18 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 3, 3, 18 ]<br />

132 #U = 18 [ 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 3, 3, 18 ]<br />

133 #U = 18 [ 2, 3 ], #H^1 = 4 [ 2, 2 ], Rk(Pic) = 1, Orbits [ 6, 6, 6, 9 ]<br />

134 #U = 18 [ 2, 3, 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 3, 6, 6, 9 ]<br />

135 #U = 18 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 3, 6, 6, 9 ]<br />

136 #U = 18 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

137 #U = 18 [ 2 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 9, 9, 9 ]<br />

138 #U = 18 [ 2 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 3, 6, 9, 9 ]<br />

139 #U = 18 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 6, 18 ]<br />

140 #U = 18 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 3, 6, 6, 9 ]<br />

141 #U = 18 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 6, 6, 6, 9 ]<br />

142 #U = 18 [ 2, 3, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 6, 18 ]<br />

143 #U = 18 [ 2, 3 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 3, 6, 9, 9 ]<br />

144 #U = 18 [ 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 9, 9 ]<br />

145 #U = 18 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

146 #U = 20 [ 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 2, 5, 10, 10 ]<br />

147 #U = 20 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 2, 5, 5, 5, 10 ]<br />

148 #U = 20 [ 4 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 5, 5, 5, 10 ]<br />

149 #U = 24 [ 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

150 #U = 24 [ 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 2, 6, 8, 8 ]<br />

151 #U = 24 [ 2, 3 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 2, 6, 8, 8 ]<br />

152 #U = 24 [ 2 ], #H^1 = 1, Rk(Pic) = 4, Orbits [ 1, 1, 1, 1, 1, 4, 4, 4, 4, 6 ]<br />

153 #U = 24 [ 2, 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

154 #U = 24 [ 3 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 8, 8, 8 ]<br />

155 #U = 24 [ 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

156 #U = 24 [ 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 4, 4, 6, 8 ]<br />

157 #U = 24 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 6, 6, 12 ]<br />

158 #U = 24 [ 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 12, 12 ]<br />

159 #U = 24 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 6, 6, 12 ]<br />

160 #U = 24 [ 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 6, 8, 8 ]<br />

161 #U = 24 [ 2, 3 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 12, 12 ]<br />

162 #U = 24 [ 2, 3 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 12, 12 ]


APPENDIX 359<br />

163 #U = 24 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 6, 6, 12 ]<br />

164 #U = 24 [ 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 12, 12 ]<br />

165 #U = 24 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 4, 4, 6, 8 ]<br />

166 #U = 24 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 6, 8, 8 ]<br />

167 #U = 24 [ 2, 2, 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 4, 6, 12 ]<br />

168 #U = 24 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 3, 3, 4, 6, 6 ]<br />

169 #U = 24 [ 8 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

170 #U = 24 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 2, 4, 6, 12 ]<br />

171 #U = 24 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 4, 6, 12 ]<br />

172 #U = 24 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 12, 12 ]<br />

173 #U = 24 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 4, 6, 12 ]<br />

174 #U = 24 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 4, 6, 12 ]<br />

175 #U = 24 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 4, 6, 12 ]<br />

176 #U = 24 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 6, 6, 12 ]<br />

177 #U = 27 [ 3, 3 ], transitive<br />

178 #U = 27 [ 3, 3, 3 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 9, 9, 9 ]<br />

179 #U = 27 [ 3, 3 ], transitive<br />

180 #U = 32 [ 2, 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 8, 8, 8 ]<br />

181 #U = 32 [ 2, 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 2, 2, 4, 16 ]<br />

182 #U = 32 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 8, 8, 8 ]<br />

183 #U = 32 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 16 ]<br />

184 #U = 32 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

185 #U = 32 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 16 ]<br />

186 #U = 32 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

187 #U = 32 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 4, 4, 8, 8 ]<br />

188 #U = 32 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 16 ]<br />

189 #U = 32 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 16 ]<br />

190 #U = 32 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

191 #U = 32 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

192 #U = 32 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 16 ]<br />

193 #U = 32 [ 2, 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 8, 8 ]<br />

194 #U = 32 [ 2, 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 16 ]<br />

195 #U = 36 [ 4 ], #H^1 = 4 [ 2, 2 ], Rk(Pic) = 1, Orbits [ 6, 6, 6, 9 ]<br />

196 #U = 36 [ 2, 2 ], #H^1 = 4 [ 2, 2 ], Rk(Pic) = 1, Orbits [ 6, 6, 6, 9 ]<br />

197 #U = 36 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 3, 3, 3, 3, 3, 3, 9 ]<br />

198 #U = 36 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 3, 3, 18 ]<br />

199 #U = 36 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 3, 6, 6, 9 ]<br />

200 #U = 36 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 3, 6, 6, 9 ]<br />

201 #U = 36 [ 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 6, 6, 6, 9 ]<br />

202 #U = 36 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 6, 6, 6, 9 ]<br />

203 #U = 36 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 6, 9, 12 ]<br />

204 #U = 36 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 6, 18 ]<br />

205 #U = 36 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 6, 18 ]<br />

206 #U = 36 [ 2, 2, 3 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 6, 9, 12 ]<br />

207 #U = 36 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 6, 18 ]<br />

208 #U = 36 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 3, 6, 6, 9 ]<br />

209 #U = 36 [ 2, 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 6, 18 ]<br />

210 #U = 36 [ 2, 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 6, 18 ]<br />

211 #U = 36 [ 2, 2 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 3, 6, 9, 9 ]<br />

212 #U = 36 [ 2, 2, 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 3, 6, 6, 9 ]<br />

213 #U = 36 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

214 #U = 40 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 2, 5, 10, 10 ]<br />

215 #U = 48 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 2, 6, 16 ]<br />

216 #U = 48 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

217 #U = 48 [ 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 6, 16 ]<br />

218 #U = 48 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 2, 2, 6, 8, 8 ]<br />

219 #U = 48 [ 2, 2, 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 2, 6, 16 ]<br />

220 #U = 48 [ 3, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 6, 16 ]<br />

221 #U = 48 [ 2, 2, 3 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 4, 6, 8, 8 ]<br />

222 #U = 48 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 4, 6, 8, 8 ]<br />

223 #U = 48 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 12, 12 ]<br />

224 #U = 48 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 2, 2, 4, 4, 6, 8 ]<br />

225 #U = 48 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 2, 6, 8, 8 ]<br />

226 #U = 48 [ 2, 2, 3 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 12, 12 ]<br />

227 #U = 48 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 12, 12 ]


360 APPENDIX<br />

228 #U = 48 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 12, 12 ]<br />

229 #U = 48 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 6, 6, 12 ]<br />

230 #U = 48 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 2, 6, 16 ]<br />

231 #U = 48 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 4, 6, 8, 8 ]<br />

232 #U = 48 [ 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

233 #U = 48 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 12, 12 ]<br />

234 #U = 48 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 12, 12 ]<br />

235 #U = 48 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

236 #U = 48 [ 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

237 #U = 48 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

238 #U = 48 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 4, 6, 12 ]<br />

239 #U = 54 [ 2 ], transitive<br />

240 #U = 54 [ 2 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 9, 9, 9 ]<br />

241 #U = 54 [ 2, 3 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 9, 9, 9 ]<br />

242 #U = 54 [ 2, 3, 3 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 9, 9, 9 ]<br />

243 #U = 54 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

244 #U = 54 [ 2, 3, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

245 #U = 54 [ 2, 3 ], transitive<br />

246 #U = 54 [ 2, 3 ], transitive<br />

247 #U = 60 [ ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 6, 6, 15 ]<br />

248 #U = 60 [ ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 5, 5, 5, 10 ]<br />

249 #U = 64 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 8, 8, 8 ]<br />

250 #U = 64 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

251 #U = 64 [ 2, 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 4, 4, 16 ]<br />

252 #U = 64 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

253 #U = 64 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

254 #U = 64 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

255 #U = 64 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

256 #U = 72 [ 2, 2 ], #H^1 = 4 [ 2, 2 ], Rk(Pic) = 1, Orbits [ 6, 6, 6, 9 ]<br />

257 #U = 72 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 6, 9, 12 ]<br />

258 #U = 72 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 6, 9, 12 ]<br />

259 #U = 72 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 6, 6, 6, 9 ]<br />

260 #U = 72 [ 3, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

261 #U = 72 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 6, 9, 12 ]<br />

262 #U = 72 [ 2, 4 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 6, 9, 12 ]<br />

263 #U = 72 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 3, 3, 6, 6, 9 ]<br />

264 #U = 72 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 6, 18 ]<br />

265 #U = 80 [ 5 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 10, 16 ]<br />

266 #U = 81 [ 3, 3 ], transitive<br />

267 #U = 96 [ 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

268 #U = 96 [ 3 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 8, 8, 8 ]<br />

269 #U = 96 [ 2, 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

270 #U = 96 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 2, Orbits [ 1, 4, 6, 8, 8 ]<br />

271 #U = 96 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 2, 6, 16 ]<br />

272 #U = 96 [ 4 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

273 #U = 96 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

274 #U = 96 [ 2, 2, 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 6, 16 ]<br />

275 #U = 96 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 6, 16 ]<br />

276 #U = 96 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 6, 16 ]<br />

277 #U = 96 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 6, 16 ]<br />

278 #U = 96 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 6, 16 ]<br />

279 #U = 96 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 3, 12, 12 ]<br />

280 #U = 108 [ 2, 2 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 9, 9, 9 ]<br />

281 #U = 108 [ 2, 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

282 #U = 108 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

283 #U = 108 [ 2, 2, 3 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 9, 9, 9 ]<br />

284 #U = 108 [ 2, 2 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 9, 9, 9 ]<br />

285 #U = 108 [ 4 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

286 #U = 108 [ 4 ], transitive<br />

287 #U = 108 [ 3, 4 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

288 #U = 108 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

289 #U = 108 [ 2, 2 ], transitive<br />

290 #U = 108 [ 2, 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

291 #U = 120 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 6, 6, 15 ]<br />

292 #U = 120 [ 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 12, 15 ]


APPENDIX 361<br />

293 #U = 120 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 2, 5, 10, 10 ]<br />

294 #U = 120 [ 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 5, 5, 5, 10 ]<br />

295 #U = 120 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 2, 5, 10, 10 ]<br />

296 #U = 120 [ 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 12, 15 ]<br />

297 #U = 128 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

298 #U = 144 [ 2, 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 6, 9, 12 ]<br />

299 #U = 144 [ 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

300 #U = 160 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 10, 16 ]<br />

301 #U = 162 [ 2, 3 ], transitive<br />

302 #U = 162 [ 2, 3 ], transitive<br />

303 #U = 162 [ 2 ], transitive<br />

304 #U = 192 [ 2 ], #H^1 = 1, Rk(Pic) = 3, Orbits [ 1, 1, 1, 8, 8, 8 ]<br />

305 #U = 192 [ 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

306 #U = 192 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

307 #U = 192 [ 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

308 #U = 192 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

309 #U = 192 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

310 #U = 192 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

311 #U = 192 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 4, 6, 16 ]<br />

312 #U = 216 [ 2, 2, 2 ], #H^1 = 3, Rk(Pic) = 1, Orbits [ 9, 9, 9 ]<br />

313 #U = 216 [ 2, 2 ], transitive<br />

314 #U = 216 [ 2, 4 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

315 #U = 216 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

316 #U = 216 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

317 #U = 216 [ 2, 2 ], transitive<br />

318 #U = 216 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

319 #U = 216 [ 8 ], transitive<br />

320 #U = 216 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

321 #U = 216 [ 2, 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

322 #U = 240 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 2, 5, 10, 10 ]<br />

323 #U = 240 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 12, 15 ]<br />

324 #U = 288 [ 3, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

325 #U = 320 [ 4 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 10, 16 ]<br />

326 #U = 324 [ 3 ], transitive<br />

327 #U = 324 [ 2, 2 ], transitive<br />

328 #U = 360 [ ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 6, 6, 15 ]<br />

329 #U = 384 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

330 #U = 384 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 2, 8, 16 ]<br />

331 #U = 432 [ 2, 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 9, 18 ]<br />

332 #U = 432 [ 2, 2 ], transitive<br />

333 #U = 576 [ 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

334 #U = 576 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

335 #U = 576 [ 2, 3 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

336 #U = 648 [ 2 ], transitive<br />

337 #U = 648 [ 2 ], transitive<br />

338 #U = 648 [ 2, 3 ], transitive<br />

339 #U = 648 [ 3 ], transitive<br />

340 #U = 720 [ 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 12, 15 ]<br />

341 #U = 720 [ 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 12, 15 ]<br />

342 #U = 720 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 6, 6, 15 ]<br />

343 #U = 960 [ ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 10, 16 ]<br />

344 #U = 1152 [ 2, 2 ], #H^1 = 1, Rk(Pic) = 1, Orbits [ 3, 24 ]<br />

345 #U = 1296 [ 2 ], transitive<br />

346 #U = 1296 [ 2, 2 ], transitive<br />

347 #U = 1440 [ 2, 2 ], #H^1 = 2, Rk(Pic) = 1, Orbits [ 12, 15 ]<br />

348 #U = 1920 [ 2 ], #H^1 = 1, Rk(Pic) = 2, Orbits [ 1, 10, 16 ]<br />

349 #U = 25920 [ ], transitive<br />

350 #U = 51840 [ 2 ], transitive


BIBLIOGRAPHY<br />

[A/B]<br />

[A/V]<br />

[Al]<br />

[AMD]<br />

[Am55]<br />

Abbes, A. et Bouche, T.: Théorème de Hilbert-Samuel “arithmétique”,<br />

Ann. Inst. Fourier 45(1995), 375–401<br />

Abramovich, D. <strong>and</strong> Voloch, J. F.: Lang’s conjectures, fibered powers,<br />

<strong>and</strong> uniformity, New York J. Math. 2(1996), 20–34<br />

Albert, A.: Structure of Algebras, Revised Printing, Amer. Math.<br />

Soc. Coll. Pub., Vol. XXIV, Amer.Math. Soc., Providence, R.I. 1961<br />

Software Optimization Guide for AMD Athlon TM 64 <strong>and</strong> AMD<br />

Opteron TM Processors, Rev. 3.04, AMD, Sunnyvale (CA) 2004<br />

Amitsur, S. A.: Generic splitting fields of central simple algebras,<br />

Annals of Math. 62(1955), 8–43<br />

[Am72] Amitsur, S. A.: On central division algebras, Israel J. Math. 12<br />

(1972), 408–420<br />

[Am81]<br />

[Ap]<br />

[Ara]<br />

Amitsur, S. A.: Generic splitting fields, in: van Oystaeyen, F. <strong>and</strong><br />

Verschoren, A. (Eds.): Brauer <strong>groups</strong> in ring theory <strong>and</strong> algebraic<br />

geometry, Proceedings of a conference held at Antwerp 1981, Lecture<br />

Notes Math. 917, Springer, Berlin, Heidelberg, New York<br />

ÖÐÓڸ˺ºÌÓÖÔÖ×ÕÒÚÞÓÖÓÚÒÖѹ<br />

1982, 1–24<br />

´½µ¸½½ß½½¾¸ ØÕ×ÓÔÓÚÖÒÓ×ظÁÞÚººÆÙËËËʸËÖºÅغ¿<br />

Apéry, F.: Models of the real projective plane, Vieweg, Braunschweig<br />

1987<br />

English translation: Arakelov, S. Yu.: Intersection theory of divisors<br />

on an arithmetic surface, Math. USSR Izv. 8(1974), 1167–1180


364 BIBLIOGRAPHY<br />

[ArE27a]<br />

[ArE27b]<br />

[A/N/T]<br />

Artin, E.: Über einen Satz von J. H. Maclagan-Wedderburn, Abh.<br />

Math. Sem. Univ. Hamburg 5(1927), 245–250<br />

Artin, E.: Zur Theorie der hyperkomplexen Zahlen, Abh. Math.<br />

Sem. Univ. Hamburg 5(1927), 251–260<br />

Artin, E., Nesbitt, C. J., <strong>and</strong> Thrall, R.: Rings with minimum<br />

condition, Univ. of Michigan Press, Ann Arbor 1948<br />

[ArM81a] Artin, M.: Left Ideals in Maximal Orders, in: van Oystaeyen, F.<br />

<strong>and</strong> Verschoren, A. (Eds.): Brauer <strong>groups</strong> in ring theory <strong>and</strong> algebraic<br />

geometry, Proceedings of a conference held at Antwerp<br />

1981, Lecture Notes Math. 917, Springer, Berlin, Heidelberg, New<br />

York 1982, 182–193<br />

[ArM81b]<br />

[A/M]<br />

[A/G]<br />

Artin, M. (Notes by Verschoren, A.): Brauer-Severi varieties, in:<br />

van Oystaeyen, F. <strong>and</strong> Verschoren, A. (Eds.): Brauer <strong>groups</strong> in<br />

ring theory <strong>and</strong> algebraic geometry, Proceedings of a conference<br />

held at Antwerp 1981, Lecture Notes Math. 917, Springer, Berlin,<br />

Heidelberg, New York 1982, 194–210<br />

Artin, M. <strong>and</strong> Mumford, D.: Some elementary examples of uni<strong>rational</strong><br />

varieties which are not <strong>rational</strong>, Proc. London Math. Soc. 25<br />

(1972), 75–95<br />

Ausl<strong>and</strong>er, M. <strong>and</strong> Goldman, O.: The Brauer group of a commutative<br />

ring, Trans. Amer. Math. Soc. 97(1960), 367–409<br />

[Az] Azumaya, G.: On maximally central algebras, Nagoya Math. J. 2<br />

(1951), 119–150<br />

[B/H/P/V] Barth, W., Hulek, K., Peters, C., <strong>and</strong> Van de Ven, A.: Compact<br />

complex surfaces, Second edition, Springer, Berlin 2004<br />

[B/M]<br />

[Ba/T]<br />

[Bv]<br />

Batyrev, V. V. <strong>and</strong> Manin, Yu. I.: Sur le nombre des <strong>points</strong> rationnels<br />

de hauteur borné des variétés algébriques, Math. Ann. 286<br />

(1990), 27–43<br />

Batyrev, V. V. <strong>and</strong> Tschinkel, Y.: Rational <strong>points</strong> on some Fano<br />

cubic bundles, C. R. Acad. Sci. Paris 323(1996), 41–46<br />

Beauville, A.: Complex algebraic surfaces, LMS Lecture Note Series<br />

68, Cambridge University Press, Cambridge 1983


BIBLIOGRAPHY 365<br />

[B/P/T/Y] Beck, M., Pine, E., Tarrant, W., <strong>and</strong> Yarbrough Jensen, K.: New<br />

integer representations as the sum of three cubes, Math. Comp. 76<br />

(2007), 1683–1690<br />

[B-gh]<br />

van den Bergh, M.: The Brauer-Severi scheme of the trace ring of<br />

generic matrices, in: Perspectives in ring theory (Antwerp 1987),<br />

NATO Adv. Sci. Inst. Ser. C: Math. Phys. Sci. 233, Kluwer Acad.<br />

Publ., Dordrecht 1988, 333–338<br />

[Be] Bernstein, D. J.: Enumerating solutions to p(a) + q(b) =<br />

r(c) + s(d), Math. Comp. 70(2001), 389–394<br />

[Be-list]<br />

Bernstein, D. J.: threecubes, Web page available at the URL:<br />

http://cr.yp.to/threecubes.html<br />

[Bt] Bertuccioni, I.: Brauer <strong>groups</strong> <strong>and</strong> cohomology, Arch. Math. 84<br />

ÐÓÖÓÚ׸ººÐÙ¸º℄ÊÚÒÓÑÖÒÓÖ×ÔÖ¹<br />

(2005), 406–411<br />

ÐÒÐÖÕ×Õ×ÐÐÞÒÕÒÓÑÙÖÙÙ¸<br />

[BB] Białynicki-Birula, A.: A note on deformations of Severi-Brauer<br />

varieties<br />

¹¾¸½¾ Î׬ºÆÚÙËËʸËÖº¬Þº¹ÅغÆÚÙ½´½µ¸<br />

<strong>and</strong> relatively minimal models of fields of genus 0, Bull.<br />

Acad. Polon. Sci. Sér. Sci. Math. Astronom. Phys. 18(1970), 175–176<br />

[Bilu88]<br />

[Bilu97]<br />

Bilu, Yu.: Limit distribution of small <strong>points</strong> on algebraic tori, Duke<br />

Math. J. 89(1997), 465-476<br />

[Bill] Billard, H.: Propriétés arithmétiques d’une famille de surfaces K3,<br />

Compositio Math. 108(1997), 247–275<br />

[Bir] Birch, B. J.: Forms in many variables, Proc. Roy. Soc. Ser. A 265<br />

(1961/1962), 245–263<br />

[Bl-ard]<br />

[Bl-et]<br />

[Bo/T]<br />

Blanchard, A.: Les corps non commutatifs, Presses Univ. de France,<br />

Vendôme 1972<br />

Blanchet, A.: Function fields of generalized Brauer-Severi varieties,<br />

Comm. in Algebra 19(1991), 97–118<br />

Bogomolov, F. <strong>and</strong> Tschinkel, Y.: Rational curves <strong>and</strong> <strong>points</strong> on<br />

K3 surfaces, Amer. J. Math. 127(2005), 825–835


ÓÖÚÕ¸ºÁºÖÚÕ¸ÁºÊºÌÓÖÕ×иÌÖØ ÞÒ¸ÆÙ¸ÅÓ×Ú½<br />

366 BIBLIOGRAPHY<br />

[B/S]<br />

English translation: Borevich, Z. I. <strong>and</strong> Shafarevich, I. R.: Number<br />

theory, Academic Press, New York-London 1966<br />

[B/G/S]<br />

Bost, J.-B., Gillet, H., Soulé, C.: Heights of projective varieties <strong>and</strong><br />

positive Green forms, J. Amer. Math. Soc. 7(1994), 903-1027<br />

[Bot] Bott, R.: On a theorem of Lefschetz, Michigan Math. J. 6(1959),<br />

211–216<br />

[Bou-T]<br />

[Bou-A]<br />

Bourbaki, N.: Éléments de Mathématique, Livre III: Topologie<br />

générale, Chapitre IX, Hermann, Paris 1948<br />

Bourbaki, N.: Éléments de Mathématique, Livre II: Algèbre,<br />

Chapitre VIII, Hermann, Paris 1958<br />

[Bra] Brauer, R.: Über Systeme hyperkomplexer Zahlen, Math. Z. 30<br />

(1929), 79–107<br />

[Bre]<br />

[Brü]<br />

[Bu]<br />

[B/K/K]<br />

[Car]<br />

[C/E]<br />

[Cas55]<br />

[Cas67]<br />

Bremner, A.: Some cubic surfaces with no <strong>rational</strong> <strong>points</strong>, Math.<br />

Proc. Cambridge Philos. Soc. 84(1978), 219–223<br />

Brüdern, J.: Einführung in die analytische Zahlentheorie, Springer,<br />

Berlin 1995<br />

Burgos Gil, J. I.: Arithmetic Chow rings, Ph.D. thesis, Barcelona<br />

1994<br />

Burgos Gil, J. I., Kramer, J., <strong>and</strong> Kühn, U.: Cohomological arithmetic<br />

Chow rings, J. Inst. Math. Jussieu 6(2007), 1–172<br />

Cartan, E.: Œvres complètes, Partie II, Volume I: Nombres complexes,<br />

Gauthier-Villars, 107–246<br />

Cartan, H. <strong>and</strong> Eilenberg, S.: Homological Algebra, Princeton<br />

Univ. Press, Princeton 1956<br />

Cassels, J. W. S.: Bounds for the least solutions of homogeneous<br />

quadratic equations, Proc. CambridgePhilos. Soc. 51(1955), 262–264<br />

Cassels, J. W. S.: Global fields, in: Algebraic number theory, Edited<br />

by J. W. S. Cassels <strong>and</strong> A. Fröhlich, Academic Press <strong>and</strong> Thompson<br />

Book Co., London <strong>and</strong> Washington 1967


BIBLIOGRAPHY 367<br />

[Ca/G]<br />

[C-L]<br />

[Cha]<br />

Cassels, J. W. S. <strong>and</strong> Guy, M. J. T.: On the Hasse principle for<br />

cubic surfaces, Mathematika 13(1966), 111–120.<br />

Chambert-Loir, A.: Points de petite hauteur sur les variétés semiabéliennes,<br />

Ann. Sci. École Norm. Sup. 33(2000), 789–821<br />

Chase, S. U.: Two remarks on central simple algebras, Comm. in<br />

Algebra 12(1984), 2279–2289<br />

[Ch44] Châtelet, F.: Variations sur un thème de H. Poincaré, Annales E.<br />

N. S. 61(1944), 249–300<br />

[Ch54/55]<br />

[Cl/G]<br />

[Coh]<br />

[CT83]<br />

[CT88]<br />

[CT91]<br />

[CT/K/S]<br />

[CT/S]<br />

[CT/SD]<br />

Châtelet, F.: Géométrie diophantienne et théorie des algèbres,<br />

Séminaire Dubreil, 1954–55, exposé 17<br />

Clemens, C. H. <strong>and</strong> Griffiths, P. A.: The intermediate Jacobian of<br />

the cubic threefold, Annals of Math. 95(1972), 281–356<br />

Cohen, H.: A course in computational algebraic number theory,<br />

Springer, Berlin, Heidelberg 1993<br />

Colliot-Thélène, J.-L.: Hilbert’s theorem 90 for K 2 , with application<br />

to the Chow group of <strong>rational</strong> surfaces, Invent. Math. 35<br />

(1983), 1–20<br />

Colliot-Thélène, J.-L.: Les gr<strong>and</strong>s thèmes de François Châtelet,<br />

L’Enseignement Mathématique 34(1988), 387–405<br />

Colliot-Thélène, J.-L.: Cycles algébriques de torsion et K-théorie<br />

algébrique, in: Arithmetic Algebraic Geometry (Trento 1991), Lecture<br />

Notes Math. 1553, Springer, Berlin, Heidelberg 1993, 1–49<br />

Colliot-Thélène, J.-L., Kanevsky, D., <strong>and</strong> Sansuc, J.-J.: Arithmétique<br />

des surfaces cubiques diagonales, in: Diophantine approximation<br />

<strong>and</strong> transcendence theory (Bonn 1985), Lecture Notes in<br />

Math. 1290, Springer, Berlin 1987, 1–108<br />

Colliot-Thélène, J.-L. <strong>and</strong> Sansuc, J.-J.: On the Chow <strong>groups</strong> of<br />

certain <strong>rational</strong> surfaces: A sequel to a paper of S. Bloch, Duke<br />

Math. J. 48(1981), 421–447<br />

Colliot-Thélène, J.-L. <strong>and</strong> Swinnerton-Dyer, Sir P.: Hasse principle<br />

<strong>and</strong> weak approximation for pencils of Severi-Brauer <strong>and</strong> similar<br />

varieties, J. für die Reine und Angew. Math. 453(1994), 49–112


368 BIBLIOGRAPHY<br />

[C/L/R]<br />

[Cor]<br />

[C/S]<br />

Cormen, T., Leiserson, C., <strong>and</strong> Rivest, R.: Introduction to algorithms,<br />

MIT Press <strong>and</strong> McGraw-Hill, Cambridge <strong>and</strong> New York<br />

1990<br />

Corn, P. K.: Del Pezzo surfaces <strong>and</strong> the Brauer-Manin obstruction,<br />

Ph.D. thesis, Harvard 2005<br />

Cornell, G. <strong>and</strong> Silverman, J. H.: Arithmetic geometry, Papers<br />

from the conference held at the University of Connecticut,<br />

Springer, New York 1986<br />

[Del] Deligne, P.: La conjecture de Weil I, Publ. Math. IHES 43(1974),<br />

273–307<br />

[Der]<br />

[Deu]<br />

[Di04]<br />

[Di05]<br />

[Dr]<br />

[Duk]<br />

Derenthal, U.: Geometry of universal torsors, Ph.D. thesis, Göttingen<br />

2006<br />

Deuring, M.: Algebren, Springer, Ergebnisse der Math. und ihrer<br />

Grenzgebiete, Berlin 1935<br />

Dickson, L. E.: Determination of all the sub<strong>groups</strong> of the known<br />

simple group of order 25 920, Trans. Amer. Math. Soc. 5(1904), 126–<br />

166<br />

Dickson, L. E.: On finite algebras, Nachrichten von der Königlichen<br />

Gesellschaft der Wissenschaften zu Göttingen, Math.-Phys. Klasse,<br />

1905, 358–393<br />

Draxl, P. K.: Skew fields, London Math. Soc. Lecture Notes Series<br />

81, Cambridge University Press, Cambridge 1983<br />

Duke, W.: Extreme values of Artin L-functions <strong>and</strong> class numbers,<br />

Compositio Math. 136(2003), 103–115<br />

[E/H/K/V] Edidin, D., Hassett, B., Kresch, A., <strong>and</strong> Vistoli, A.: Brauer <strong>groups</strong><br />

<strong>and</strong> quotient stacks, Amer. J. Math. 123(2001), 761–777<br />

[Ek]<br />

[EGA]<br />

Ekedahl, T.: An effective version of Hilbert’s irreducibility theorem,<br />

in: Séminaire de Théorie des Nombres, Paris 1988–1989,<br />

Progr. Math. 91, Birkhäuser, Boston 1990, 241–249<br />

Grothendieck, A. et Dieudonné, J.: Éléments de Géométrie Algébrique,<br />

Publ. Math. IHES 4, 8, 11, 17, 20, 24, 28, 32(1960–68)


BIBLIOGRAPHY 369<br />

[El]<br />

[EJ1]<br />

[EJ2]<br />

[EJ3]<br />

[EJ4]<br />

[EJ5]<br />

[EJ6]<br />

[EJ7]<br />

[EJ8]<br />

[EJ9]<br />

[Fa83]<br />

Elkies, N. D.: Rational <strong>points</strong> near curves <strong>and</strong> small nonzero<br />

|x 3 − y 2 | via lattice reduction, in: Algorithmic number theory<br />

(Leiden 2000), Lecture Notes in Computer Science 1838, Springer,<br />

Berlin 2000, 33–63<br />

Elsenhans, A.-S. <strong>and</strong> Jahnel, J.: The Fibonacci sequence modulo<br />

p 2 —An investigation by computer for p < 10 14 , Preprint<br />

Elsenhans, A.-S. <strong>and</strong> Jahnel, J.: The Diophantine equation<br />

x 4 + 2y 4 = z 4 + 4w 4 , Math. Comp. 75(2006), 935–940<br />

Elsenhans, A.-S. <strong>and</strong> Jahnel, J.: The Diophantine equation<br />

x 4 + 2y 4 = z 4 + 4w 4 — a number of improvements, Preprint<br />

Elsenhans, A.-S. <strong>and</strong> Jahnel, J.: The Asymptotics of Points of<br />

Bounded Height on Diagonal Cubic <strong>and</strong> Quartic Threefolds, in:<br />

Algorithmic number theory, Lecture Notes in Computer Science<br />

4076, Springer, Berlin 2006, 317–332<br />

Elsenhans, A.-S. <strong>and</strong> Jahnel, J.: On the smallest point on a diagonal<br />

quartic threefold, J. Ramanujan Math. Soc. 22(2007), 189–204<br />

Elsenhans, A.-S. <strong>and</strong> Jahnel, J.: Experiments with general cubic<br />

surfaces, to appear in: The Manin Festschrift<br />

Elsenhans, A.-S. <strong>and</strong> Jahnel, J.: On the smallest point on a diagonal<br />

cubic surface, Preprint<br />

Elsenhans, A.-S. <strong>and</strong> Jahnel, J.: New sums of three cubes, to appear<br />

in: Math. Comp.<br />

Elsenhans, A.-S. <strong>and</strong> Jahnel, J.: List of solutions of x 3 +y 3 +z 3 = n<br />

for n < 1 000 neither a cube nor twice a cube, Web page available at<br />

the URL: http://www.uni-math.gwdg.de/jahnel/Arbeiten/<br />

Liste/threecubes_20070419.txt<br />

Faltings, G.: Endlichkeitssätze für abelsche Varietäten über Zahlkörpern,<br />

Invent. Math. 73(1983), 349–366<br />

[Fa84] Faltings, G.: Calculus on arithmetic surfaces, Annals of Math. 119<br />

(1984), 387–424<br />

[Fa91]<br />

Faltings, G.: The general case of S. Lang’s conjecture, in: Barsotti<br />

Symposium inAlgebraicGeometry(AbanoTerme1991), Perspect.<br />

Math. 15, Academic Press, San Diego 1994, 175–182


370 BIBLIOGRAPHY<br />

[Fa92]<br />

[F/P]<br />

[Fo]<br />

[F/M/T]<br />

[Fu84]<br />

[Ga]<br />

[G/T]<br />

[G/S 90]<br />

[G/S 92]<br />

[Gi]<br />

[G/H]<br />

[GrBrI]<br />

[GrBrII]<br />

Faltings, G.: Lectures on the arithmetic Riemann-Roch theorem,<br />

Annals of Mathematics Studies 127, Princeton University Press,<br />

Princeton 1992<br />

Fincke, U. <strong>and</strong> Pohst, M.: Improved methods for calculating vectors<br />

of short length in a lattice, including a complexity analysis,<br />

Math. Comp. 44(1985), 463–471<br />

Forster, O.: Algorithmische Zahlentheorie, Vieweg, Braunschweig<br />

1996<br />

Franke, J., Manin, Yu. I., <strong>and</strong> Tschinkel, Y.: Rational <strong>points</strong> of<br />

bounded height on Fano varieties, Invent. Math. 95(1989), 421–435<br />

Fulton, W.: Intersection theory, Springer, Ergebnisse der Math.<br />

und ihrer Grenzgebiete, 3. Folge, B<strong>and</strong> 2, Berlin 1984<br />

Gabber, O.: Some theorems on Azumaya algebras, in: The Brauer<br />

group (Les Plans-sur-Bex 1980), Lecture Notes Math. 844, Springer,<br />

Berlin, Heidelberg, New York 1981, 129–209<br />

Gatsinzi, J.-B. <strong>and</strong> Tignol, J.-P.: Involutions <strong>and</strong> Brauer-Severi varieties,<br />

Indag. Math. 7(1996), 149–160<br />

Gillet, H., Soulé, C.: Arithmetic intersection theory, Publ. Math.<br />

IHES 72(1990), 93-174<br />

Gillet, H., Soulé, C.: An arithmetic Riemann-Roch theorem, Invent.<br />

Math. 110(1992), 473–543<br />

Giraud, J.: Cohomologie non abélienne, Springer,Grundlehren der<br />

math. Wissenschaften 179, Berlin, Heidelberg, New York 1971<br />

Griffiths, P. <strong>and</strong> Harris, J.: Principles of algebraic geometry, John<br />

Wiley & Sons, New York 1978<br />

Grothendieck, A.: Le groupe de Brauer, I: Algèbres d’Azumaya et<br />

interprétations diverses, Séminaire Bourbaki, 17 e année 1964/65,<br />

n o 290<br />

Grothendieck, A.: Le groupe de Brauer, II: Théorie cohomologique,<br />

Séminaire Bourbaki, 18 e année 1965/66, n o 297


BIBLIOGRAPHY 371<br />

[GrBrIII]<br />

[Hab]<br />

[Hai]<br />

[Hb]<br />

[H/W]<br />

[Ha70]<br />

[Ha77]<br />

[H-B92a]<br />

[H-B92b]<br />

[H/L/R]<br />

[H/P]<br />

[Hen]<br />

Grothendieck, A.: Le groupe de Brauer, III: Exemples et compléments,<br />

in: Grothendieck, A.: Dix exposés sur la Cohomologie<br />

des schémas, North-Holl<strong>and</strong>, Amsterdam <strong>and</strong> Masson, Paris 1968,<br />

88–188<br />

Haberl<strong>and</strong>, K.: Galois cohomology of algebraic number fields,<br />

With two appendices by H. Koch <strong>and</strong> Th. Zink, VEB Deutscher<br />

Verlag der Wissenschaften, Berlin 1978<br />

Haile, D. E.: On central simple algebras of given exponent, J. of<br />

Algebra 57(1979), 449–465<br />

H<strong>and</strong>book of numerical analysis, edited by P. G. Ciarlet <strong>and</strong><br />

J. L. Lions, Vols. II–IV, North-Holl<strong>and</strong> Publishing Co., Amsterdam<br />

1991–1996<br />

Hardy, G. H. <strong>and</strong> Wright, E. M.: An introduction to the theory of<br />

numbers, Fifth edition, Oxford University Press, New York 1979<br />

Hartshorne, R.: Ample subvarieties of algebraic varieties, Lecture<br />

Notes Math. 156, Springer, Berlin-New York 1970<br />

Hartshorne, R.: Algebraic Geometry, Graduate Texts in Mathematics<br />

52, Springer, New York 1977<br />

Heath-Brown, D. R.: The density of zeros of forms for which<br />

weak approximation fails, Math. Comp. 59(1992), 613–623<br />

Heath-Brown, D. R.: Zero-free regions for Dirichlet L-functions,<br />

<strong>and</strong> the least prime in an arithmetic progression, Proc. London<br />

Math. Soc. 64(1992), 265–338<br />

Heath-Brown, D. R., Lioen, W. M., <strong>and</strong> te Riele, H. J. J.: On<br />

solving the diophantine equation x 3 + y 3 + z 3 = k on a vector<br />

computer, Math. Comp. 61(1993), 235–244<br />

Hennessy, J. L. <strong>and</strong> Patterson, D. A.: Computer Architecture:<br />

A Quantitative Approach, 2 nd ed., Morgan Kaufmann, San Mateo<br />

(CA) 1996<br />

Henningsen, F.: Brauer-Severi-Varietäten und Normrelationen<br />

von Symbolalgebren, Doktorarbeit (Ph. D. thesis), available in electronic<br />

form at http://www.biblio.tu-bs.de/ediss/data/<br />

20000713a/20000713a.pdf


372 BIBLIOGRAPHY<br />

[Herm]<br />

[Hers]<br />

[Heu]<br />

[Hi]<br />

[H/W]<br />

[H/J/S]<br />

[Hoo]<br />

[I/R]<br />

[Is]<br />

[Jac]<br />

[J1]<br />

Hermann, G.: Die Frage der endlich vielen Schritte in der Theorie<br />

der Polynomideale, Math. Ann. 95(1926), 736–788<br />

Herstein, I. N.: Noncommutative rings, The Math. Association of<br />

America (Inc.), distributed by John Wiley <strong>and</strong> Sons, Menasha, WI<br />

1968<br />

Heuser, A.: Über den Funktionenkörper der Normfläche einer<br />

zentral einfachen Algebra, J. für die Reine und Angew. Math. 301<br />

(1978), 105–113<br />

Hirzebruch, F.: Der Satz von Riemann-Roch in Faisceau-theoretischer<br />

Formulierung: Einige Anwendungen und offene Fragen,<br />

Proc. Int. Cong. of Mathematicians, Amsterdam 1954, vol. III,<br />

457–473, Erven P. Noordhoff N.V. <strong>and</strong> North-Holl<strong>and</strong> Publishing<br />

Co., Groningen <strong>and</strong> Amsterdam 1956<br />

Hoffman, J. W. <strong>and</strong> Weintraub, S. H.: Four dimensional symplectic<br />

geometry over the field with three elements <strong>and</strong> a moduli space of<br />

abelian surfaces, Note Mat. 20(2000/01), 111–157<br />

Hoffmann, N., Jahnel, J., <strong>and</strong> Stuhler, U.: Generalized vector bundles<br />

on curves, J. für die Reine und Angew. Math. 495(1998), 35-60<br />

Hoobler, R. T.: A cohomological interpretation of Brauer <strong>groups</strong><br />

Á×ÓÚ׸κºÃÓÒØÖÔÖÑÖÔÖÒÔÙÀ××Ð××¹<br />

of rings, Pacific J. Math. 86(1980), 89–92<br />

ØÑÝÚÙÚÖØÕÒÝÓÖÑÓØÔØÔÖÑÒÒݸÅع ÑØÕ×Ñؽ¼´½½µ¸¾¿ß¾<br />

Irel<strong>and</strong>, K. F. <strong>and</strong> Rosen, M. I.: A classical introduction to modern<br />

number theory, Graduate Texts in Mathematics 84, Springer, New<br />

York <strong>and</strong> Berlin 1982<br />

English translation: Iskovskih, V. A.: A counterexample to the<br />

Hasseprinciplefor systemsoftwoquadraticformsinfivevariables,<br />

Mathematical Notes 10(1971), 575–577<br />

Jacobson, N.: Structure of rings, Revised edition, Amer. Math.<br />

Soc. Coll. Publ. 37, Amer. Math. Soc., Providence, R.I. 1964<br />

Jahnel, J.: Line bundles on arithmetic surfaces <strong>and</strong> intersection<br />

theory, manuscripta math. 91(1996), 103-119


BIBLIOGRAPHY 373<br />

[J2]<br />

[J3]<br />

[J4]<br />

[J5]<br />

[dJo]<br />

Jahnel, J.: Heights for line bundles on arithmetic varieties, manuscripta<br />

math. 96(1998), 421-442<br />

Jahnel, J.: A height function on the Picard group of singular<br />

Arakelov varieties, in: Algebraic K-Theory <strong>and</strong> Its Applications,<br />

Proceedings of the Workshop <strong>and</strong> Symposium held at ICTP<br />

Trieste, September 1997, edited by H. Bass, A. O. Kuku <strong>and</strong><br />

C. Pedrini, World Scientific, Singapore 1999, 410-436<br />

Jahnel, J.: The Brauer-Severi variety associated with a central<br />

simple algebra, Linear Algebraic Groups <strong>and</strong> Related Structures 52<br />

(2000), 1-60<br />

Jahnel, J.: On the distribution of small <strong>points</strong> on abelian <strong>and</strong> toric<br />

varieties, Preprint<br />

de Jong, A. J.: Smoothness, semi-stability <strong>and</strong> alterations, Publ.<br />

Math. IHES 83(1996), 51–93<br />

[K/K/M/S] Kempf, G., Knudsen, F. F., Mumford, D., <strong>and</strong> Saint-Donat, B.:<br />

Toroidal embeddings I, Lecture Notes in Mathematics 339,<br />

Springer, Berlin-New York 1973<br />

[Ke]<br />

[K/R]<br />

[K/O74]<br />

[K/O80]<br />

[K/T]<br />

Kersten, I.: Brauergruppen von Körpern, Vieweg, Braunschweig<br />

1990<br />

Kersten, I. <strong>and</strong>Rehmann, U.: Genericsplittingofreductive<strong>groups</strong>,<br />

Tôhoku Math. J. 46(1994), 35–70<br />

Knus, M. A. <strong>and</strong> Ojanguren, M.: Théorie de la descente et algèbres<br />

d’Azumaya, Springer, Lecture Notes Math. 389, Berlin, Heidelberg,<br />

New York 1974<br />

Knus, M. A. <strong>and</strong> Ojanguren, M.: Cohomologie étale et groupe<br />

de Brauer, in: The Brauer group (Les Plans-sur-Bex 1980), Lecture<br />

Notes Math. 844, Springer, Berlin, Heidelberg, New York 1981,<br />

210–228<br />

Kresch, A. <strong>and</strong> Tschinkel, Y.: Effectivity of Brauer-Manin obstructions,<br />

Preprint<br />

[Kr] Kress, R.: Numerical analysis, Graduate Texts in Mathematics 181,<br />

Springer, New York 1998


374 BIBLIOGRAPHY<br />

[Lan86]<br />

[Lan93]<br />

[Laz]<br />

[Le]<br />

[Lic]<br />

[Lin]<br />

[L/O]<br />

[Lo]<br />

[MWe05]<br />

[MWe08]<br />

[Man]<br />

[Mc]<br />

[Mat]<br />

Lang, S.: Hyperbolic <strong>and</strong> Diophantine analysis, Bull. Amer. Math.<br />

Soc. 14(1986), 159–205<br />

Lang, S.: Algebra, Third Edition, Addison-Wesley, Reading, MA<br />

1993<br />

Lazarsfeld, R.: Positivity in algebraic geometry I, Ergebnisse<br />

der Mathematik und ihrer Grenzgebiete (3. Folge) 48, Springer,<br />

Berlin 2004<br />

Lelong, P.: Fonctions plurisousharmoniques et formes différentielles<br />

positives, Gordon & Breach, Paris 1968<br />

Lichtenbaum, S.: Duality theorems for curves over p-adic fields,<br />

Invent. Math. 7(1969), 120–136<br />

Lind, C.-E.: Untersuchungen über die <strong>rational</strong>en Punkte der<br />

ebenen kubischen Kurven vom Geschlecht Eins, Doktorarbeit<br />

(Ph.D. thesis), Uppsala 1940<br />

Lorenz, F. <strong>and</strong> Opolka, H.: Einfache Algebren und projektive<br />

Darstellungen über Zahlkörpern, Math. Z. 162(1978), 175–182<br />

Lovász, L.: How to compute the volume?, Jahresbericht der DMV,<br />

Jubiläumstagung 100 Jahre DMV (Bremen 1990), 138–151<br />

ÅÒÒ¸ºÁºÃÙÕ×ÓÖÑÝÐÖ¸ÓÑØÖ¸Ö¹<br />

Maclagan-Wedderburn, J. H.: A theorem on finite algebras, Trans.<br />

ÑظÆÙ¸ÅÓ×Ú½¾<br />

Amer. Math. Soc. 6(1905), 349–352<br />

Maclagan-Wedderburn, J. H.: On hypercomplex numbers, Proc.<br />

London Math. Soc., 2nd Series, 6(1908), 77–118<br />

English translation: Manin, Yu. I.: Cubic forms, algebra, geometry,<br />

arithmetic, North-Holl<strong>and</strong> Publishing Co. <strong>and</strong> American Elsevier<br />

Publishing Co., Amsterdam-London <strong>and</strong> New York 1974<br />

Marcus, D. A.: Number fields, Springer, New York-Heidelberg<br />

1977<br />

Matsumura, H.: Commutative ring theory, Cambridge Studies in<br />

Advanced Mathematics 8, Cambridge University Press, Cambridge<br />

1986


BIBLIOGRAPHY 375<br />

[McK]<br />

McKinnon, D.: Counting <strong>rational</strong> <strong>points</strong> on K3 surfaces, J. Number<br />

Theory 84(2000), 49–62<br />

[Me-a]<br />

ÅÖÙÖÚ¸ºËºÇ×ØÖÓÒÖÙÔÔÝÖÙ«ÖÔÓиÁÞÚº<br />

O’Meara, O. T.: Introduction to quadratic forms, Springer, Berlin,<br />

ºÆÙËËËÊËÖºÅغ´½µ¸¾ß¸<br />

New York 1963<br />

[Me83] Merkurjev, A. S.: Brauer <strong>groups</strong> of fields, Comm. in Algebra 11<br />

(1983), 2611–2624<br />

[Me85]<br />

Englishtranslation: Merkurjev, A. S.: StructureoftheBrauer group<br />

[Me93]<br />

[Me/S]<br />

ÅÖÙÖÚ¸ºËºÑÒÙØÝØÓÕÑÒÓÓÓÖÞËÚÖ¹ ÖÙ«Ö¸Î×ØÒ˺¹ÈØÖÙÖºÍÒÚºÅغź×ØÖÓ¹ ÒÓѺ½¿¹¾´½¿µ¸½ß¿<br />

of fields, Math. USSR-Izv. 27(1986), 141–157<br />

ÅÖÙÖÚ¸ºËºËÙ×ÐÒ¸ººK¹ÓÓÑÓÐÓÑÒÓÓ¹ ÓÖÞËÚÖ¹ÖÙ«ÖÓÑÓÑÓÖÞÑÒÓÖÑÒÒÓÓÚݹ ÕظÁÞÚººÆÙËËËÊËÖºÅغ´½¾µ¸½¼½½ß½¼¸<br />

English translation: Merkurjev, A. S.: Closed <strong>points</strong> of Brauer-<br />

Severi varieties, Vestnik St.PetersburgUniv.Math. 26-2(1993), 44–46<br />

English translation: Merkurjev, A. S. <strong>and</strong> Suslin, A. A.: K-<br />

cohomology of Severi-Brauer varieties <strong>and</strong> the norm residue homomorphism,<br />

Math. USSR-Izv. 21(1983), 307–340<br />

[deM/F]<br />

[deM/I]<br />

[Mi]<br />

[Mol]<br />

[Mord]<br />

de Meyer, F. <strong>and</strong> Ford, T. J.: On the Brauer group of surfaces <strong>and</strong><br />

subrings ofk[x, y], in: van Oystaeyen, F. <strong>and</strong> Verschoren, A. (Eds.):<br />

Brauer <strong>groups</strong> in ring theory <strong>and</strong> algebraic geometry, Proceedings<br />

of a conference held at Antwerp 1981, Lecture Notes Math. 917,<br />

Springer, Berlin, Heidelberg, New York 1982, 211–221<br />

de Meyer, F. <strong>and</strong> Ingraham, E.: Separable algebras over commutative<br />

rings, Lecture Notes Math. 181, Springer, Berlin, Heidelberg,<br />

New York 1971<br />

Milne, J. S.: Étale Cohomology, Princeton University Press, Princeton<br />

1980<br />

Molien, T.: Über Systeme höherer complexer Zahlen, Math. Ann.<br />

41(1893), 83–156<br />

Mordell, L. J.: On the conjecture for the <strong>rational</strong> <strong>points</strong> on a cubic<br />

surface, J. London Math. Soc. 40(1965), 149–158


376 BIBLIOGRAPHY<br />

[Mori]<br />

[Mu-e]<br />

[Mu-y]<br />

[Nak]<br />

[Nat]<br />

[Ne]<br />

[Noe]<br />

[Or]<br />

[Or/S]<br />

[Or/V]<br />

[Pe95a]<br />

[Pe95b]<br />

Moriwaki, A.: Intersection pairing for arithmetic cycles with degenerate<br />

Green currents, E-print AG/9803054<br />

Murre, J. P.: Algebraic equivalence modulo <strong>rational</strong> equivalence<br />

on a cubic threefold, Compositio Math. 25(1972), 161–206<br />

Murty, M. R.: Applications of symmetric power L-functions, in:<br />

Lectures on automorphic L-functions, Fields Inst. Monogr. 20,<br />

Amer. Math. Soc., Providence 2004, 203–283<br />

Nakayama, T.: Divisionsalgebren über diskret bewerteten perfekten<br />

Körpern, J. für die Reine und Angew. Math. 178(1937), 11–13<br />

Nathanson, M. B.: Elementary methods in number theory, Graduate<br />

Texts in Mathematics 195, Springer, New York 2000<br />

Neukirch, J.: Algebraic number theory, Grundlehren der Mathematischen<br />

Wissenschaften 322, Springer, Berlin 1999<br />

Noether, M.: Rationale Ausführung der Operationen in der Theorie<br />

der algebraischen Funktionen, Math. Ann. 23(1884), 311–358<br />

Orzech, M.: Brauer Groups <strong>and</strong> Class Groups for a Krull Domain,<br />

in: van Oystaeyen, F. <strong>and</strong> Verschoren, A. (Eds.): Brauer <strong>groups</strong> in<br />

ring theory <strong>and</strong> algebraic geometry, Proceedings of a conference<br />

held at Antwerp 1981, Lecture Notes Math. 917, Springer, Berlin,<br />

Heidelberg, New York 1982, 66–90<br />

Orzech, M. <strong>and</strong> Small, C.: The Brauer group of commutative rings,<br />

Lecture Notes in Pure <strong>and</strong> Applied Mathematics 11, Marcel Dekker,<br />

New York 1975<br />

Orzech, M. <strong>and</strong> Verschoren, A.: Some Remarks on Brauer Groups<br />

of Krull Domains, in: van Oystaeyen, F. <strong>and</strong> Verschoren, A. (Eds.):<br />

Brauer <strong>groups</strong> in ring theory <strong>and</strong> algebraic geometry, Proceedings<br />

of a conference held at Antwerp 1981, Lecture Notes Math. 917,<br />

Springer, Berlin, Heidelberg, New York 1982, 91–95<br />

Peyre, E.: Hauteurs et mesures de Tamagawa sur les variétés de<br />

Fano, Duke Math. J. 79(1995), 101–218<br />

Peyre, E.: Products of Severi-Brauer varieties <strong>and</strong> Galois cohomology,<br />

in: Proc. Symp. Pure Math. 58, Part 2: K-theory <strong>and</strong> algebraic<br />

geometry: connections with quadratic forms <strong>and</strong> division algebras


BIBLIOGRAPHY 377<br />

(Santa Barbara, CA 1992), Amer. Math. Soc., Providence, R.I. 1995,<br />

369–401<br />

[Pe02]<br />

[Pe/T]<br />

[Pi]<br />

[Poh/Z]<br />

[Poi]<br />

[Poo/T]<br />

[Qu]<br />

[Re]<br />

[Roq63]<br />

[Roq64]<br />

[Rou]<br />

Peyre, E.: Points de hauteur bornée et géométrie des variétés<br />

(d’après Y. Manin et al.), Séminaire Bourbaki 2000/2001,Astérisque<br />

282(2002), 323–344<br />

Peyre, E. <strong>and</strong> Tschinkel, Y.: Tamagawa numbers of diagonal cubic<br />

surfaces, numerical evidence, Math. Comp. 70(2001), 367–387<br />

Pierce, R. S.: Associative Algebras, Graduate Texts in Mathematics<br />

88, Springer, New York 1982<br />

Pohst, M. <strong>and</strong> Zassenhaus, H.: Algorithmic algebraic number theory,<br />

Encyclopedia of Mathematics <strong>and</strong> its Applications 30, Cambridge<br />

University Press, Cambridge 1989<br />

Poincaré, H.: Sur les propriétés arithmétiques des courbes algébriques,<br />

J. Math. Pures et Appl. 5 e série, 7(1901), 161–234<br />

Poonen, B. <strong>and</strong> Tschinkel, Y. (eds.): Arithmetic of higherdimensional<br />

algebraic varieties, Proceedings of the Workshop on<br />

Rational <strong>and</strong> Integral Points of Higher-Dimensional Varieties held<br />

in Palo Alto, CA, December 11–20, 2002, Birkhäuser, Progress in<br />

Mathematics 226, Boston 2004<br />

Quillen, D.: Higher algebraic K-theory I, in: Algebraic K-theory<br />

I, Higher K-theories, Lecture Notes Math. 341, Springer, Berlin,<br />

Heidelberg, New York 1973, 85–147<br />

Reiner, I.: Maximal orders, Academic Press, London-New York<br />

1975<br />

Roquette, P.: On the Galois cohomology of the projective linear<br />

group <strong>and</strong> its applications to the construction of generic splitting<br />

fields of algebras, Math. Ann. 150(1963), 411–439<br />

Roquette, P.: Isomorphisms of generic splitting fields of simple<br />

algebras, J. für die Reine und Angew. Math. 214/215(1964), 207–226<br />

Roux, B.: Sur le groupe de Brauer d’un corps local à corps résiduel<br />

imparfait, in: Groupe d’Etude d’Analyse Ultramétriuqe, Publ.<br />

Math. de l’Univ. de Paris VII 29(1985/86), 85–97


378 BIBLIOGRAPHY<br />

[Sal80]<br />

[Sal81]<br />

[Sat]<br />

[Scha]<br />

[Sch/St]<br />

[Schr]<br />

[Schw]<br />

Saltman, P. J.: Division algebras over discrete valued fields, Comm.<br />

in Algebra 8(1980), 1749–1774<br />

Saltman, P. J.: The Brauer group is torsion, Proc. Amer. Math. Soc.<br />

81(1981), 385–387<br />

Satake, I.: On the structure of Brauer group of a discretely-valued<br />

complete field, Sci. Papers Coll. Gen. Ed. Univ. Tokyo 1(1951), 1–10<br />

Scharlau, W.: Quadratic <strong>and</strong> Hermitian forms, Springer, Grundlehren<br />

der math. Wissenschaften 270, Berlin, Heidelberg, New<br />

York 1985<br />

Schönhage, A. <strong>and</strong> Strassen, V.: Schnelle Multiplikation großer<br />

Zahlen, Computing 7(1971), 281–292<br />

Schröer, S.: There are enough Azumaya algebras on surfaces, Math.<br />

Ann. 321(2001), 439–454<br />

Schwarz, H. A.: Sur une définition erronée de l’aire d’une surface<br />

courbe, Communication faite à M. Charles Hermite, 1881–<br />

82, in: Gesammelte Mathematische Abh<strong>and</strong>lungen, Zweiter B<strong>and</strong>,<br />

Springer, Berlin 1890, 309–311<br />

[Sed] Sedgewick, R.: Algorithms, Addison-Wesley, Reading 1983<br />

[See97]<br />

[See99]<br />

[Sel]<br />

[Se67]<br />

[SGA3]<br />

Seelinger, G.: A description of the Brauer-Severi scheme of trace<br />

rings of generic matrices, J. of Algebra 184(1996), 852–880<br />

Seelinger, G.: Brauer-Severi schemes of finitely generated algebras,<br />

Israel J. Math. 111(1999), 321–337<br />

Selmer, E. S.: The Diophantine equation ax 3 +by 3 +cz 3 = 0, Acta<br />

Math. 85(1951), 203–362<br />

Serre, J.-P.: Local class field theory, in: Algebraic number theory,<br />

Edited by J. W. S. Cassels <strong>and</strong> A. Fröhlich, Academic Press <strong>and</strong><br />

Thompson Book Co., London <strong>and</strong> Washington 1967<br />

Demazure, A. et Grothendieck, A.: Schemas en groupes, Séminaire<br />

de Géométrie Algébrique du Bois Marie 1962–64 (SGA3), Lecture<br />

Notes Math. 151, 152, 153, Springer, Berlin, Heidelberg, New York<br />

1970


BIBLIOGRAPHY 379<br />

[SGA4]<br />

[SGA4 1 2 ]<br />

[SGA6]<br />

Artin, M., Grothendieck, A. et Verdier, J.-L. (avec la collaboration<br />

de Deligne, P. et Saint-Donat, B.): Théorie des topos et cohomologie<br />

étale des schémas, Séminaire de Géométrie Algébrique du Bois<br />

Marie 1963–1964 (SGA4), Lecture Notes in Math. 305, Springer,<br />

Berlin, Heidelberg, New York 1973<br />

Deligne, P. (avec la collaboration de Boutot, J. F., Grothendieck,<br />

A., Illusie, L. et Verdier, J. L.): Cohomologie Étale, Séminaire<br />

de Géométrie Algébrique du Bois Marie (SGA4 1 ), Lecture Notes<br />

2<br />

Math. 569, Springer, Berlin, Heidelberg, New York 1977<br />

Berthelot, P., Grothendieck, A. et Illusie, L. (Avec la collaboration<br />

de Ferr<strong>and</strong>, D., Jouanolou, J. P., Jussila, O., Kleiman, S., Raynaud,<br />

M. et Serre, J. P.): Théorie des intersections et théorème de<br />

Riemann-Roch, Séminaire de Géométrie Algébrique du Bois Marie<br />

1966-67 (SGA6), Lecture Notes in Math. 225, Springer, Berlin-New<br />

York 1971<br />

[Se62] Serre, J.-P.: Corps locaux, Hermann, Paris 1962<br />

[Se72]<br />

[Se73]<br />

[Sev]<br />

[Si69]<br />

[Si73]<br />

[Sim]<br />

[Sk]<br />

Serre, J.-P.: Propriétés galoisiennes des <strong>points</strong> d’ordre fini des<br />

courbes elliptiques, Invent. Math. 15(1972), 259-331<br />

Serre, J.-P.: Cohomologie Galoisienne, Lecture Notes in Mathematics<br />

5, Quatrième Edition, Springer, Berlin 1973<br />

Severi, F.: Un nuovo campo di richerche nella geometria sopra una<br />

superficie e sopra una varietà algebrica, Mem. della Acc. R. d’Italia<br />

3(1932), 1–52 = Opere matematiche, vol. terzo, Acc. Naz. Lincei,<br />

Roma, 541–586<br />

Siegel, C. L.: Abschätzung von Einheiten, Nachr. Akad. Wiss. Göttingen,<br />

Math.-Phys. Kl. II 1969(1969), 71–86<br />

Siegel, C. L.: Normen algebraischer Zahlen, Nachr. Akad. Wiss.<br />

Göttingen, Math.-Phys. Kl. II 1973(1973), 197–215<br />

Sims, C. C.: Computational methods in the study of permutation<br />

<strong>groups</strong>, in: Computational Problems in Abstract Algebra (Proc.<br />

Conf., Oxford 1967), Pergamon, Oxford 1970, 169–183<br />

Skorobogatov, A.: Torsors <strong>and</strong> <strong>rational</strong> <strong>points</strong>, Cambridge Tracts<br />

in Mathematics 144, Cambridge University Press, Cambridge 2001


380 BIBLIOGRAPHY<br />

[Sm]<br />

Smart, N. P.: The algorithmic resolution of Diophantine equations,<br />

London Mathematical Society Student Texts 41, Cambridge<br />

University Press, Cambridge 1998<br />

[S/A/B/K] Soulé, C., Abramovich, D., Burnol, J.-F., <strong>and</strong> Kramer, J.: Lectures<br />

on Arakelov geometry, Cambridge Studies in Advanced Mathematics<br />

8, Cambridge University Press, Cambridge 1992<br />

[Sp]<br />

[St]<br />

[Su]<br />

[Sw]<br />

[S-D62]<br />

[S-D93]<br />

[S-D04]<br />

[S-D05]<br />

[Sze]<br />

[Szp]<br />

ËÙ×ÐÒ¸ººÃÚØÖÒÓÒÒÝÓÑÓÑÓÖÞÑÐÔÓÐÙÒ¹<br />

Spanier, E. H.: Algebraic topology, McGraw-Hill, New York-Toronto-London<br />

1966<br />

ÒÓÒ¸ÓкºÆÙËËËʾ´½¾µ¸¾¾ß¾¸<br />

Stark, H. M.: Some effective cases of the Brauer-Siegel theorem,<br />

Invent. Math. 23(1974), 135–152<br />

English translation: Suslin, A. A.: The quaternion homomorphism<br />

for the function field on a conic, Soviet Math. Dokl. 26(1982), 72–77<br />

Swiȩcicka, J.: Small deformations of Brauer-Severi schemes, Bull.<br />

Acad. Polon. Sci. Sér. Sci. Math. Astronom. Phys. 26(1978), 591–593<br />

Swinnerton-Dyer, Sir P.: Two special cubic surfaces, Mathematika<br />

9(1962), 54–56<br />

Swinnerton-Dyer, Sir P.: The Brauer group of cubic surfaces, Math.<br />

Proc. Cambridge Philos. Soc. 113(1993), 449–460<br />

Swinnerton-Dyer, Sir P.: Rational <strong>points</strong> on fibered surfaces, in:<br />

Tschinkel, Y. (ed.): Mathematisches Institut, Seminars 2004, Universitätsverlag,<br />

Göttingen 2004, 103–109<br />

Swinnerton-Dyer, Sir P.: Counting <strong>points</strong> on cubic surfaces II,<br />

in: Geometric methods in algebra <strong>and</strong> number theory, Progr.<br />

Math. 235, Birkhäuser, Boston 2005, 303–309<br />

Szeto, G.: Splitting Rings for Azumaya Quaternion Algebras, in:<br />

van Oystaeyen, F. <strong>and</strong> Verschoren, A. (Eds.): Brauer <strong>groups</strong> in<br />

ring theory <strong>and</strong> algebraic geometry, Proceedings of a conference<br />

held at Antwerp 1981, Lecture Notes Math. 917, Springer, Berlin,<br />

Heidelberg, New York 1982, 118–125<br />

Szpiro, L.: Sur les propriétés numériques du dualisant relatif<br />

d’une surface arithmétique, The Grothendieck Festschrift, vol. III,<br />

Birkhäuser, Boston 1990


BIBLIOGRAPHY 381<br />

[S/U/Z]<br />

[Ta67]<br />

[Ta75]<br />

[Ti81]<br />

[Ti87]<br />

[Tr]<br />

[Tu]<br />

[Ul]<br />

[Va]<br />

[Ve]<br />

Szpiro, L., Ullmo, E., Zhang, S.: Equirépartition des petits <strong>points</strong>,<br />

Invent. Math. 127(1997), 337-347<br />

Tate, J.: Global class field theory, in: Algebraic number theory,<br />

Edited by J. W. S. Cassels <strong>and</strong> A. Fröhlich, Academic Press <strong>and</strong><br />

Thompson Book Co., London <strong>and</strong> Washington 1967<br />

Tate, J.: Algorithm for determining the type of a singular fiber in an<br />

elliptic pencil, Modular functions of one variable IV (Proc. Internat.<br />

Summer School, Antwerp 1972), Lecture Notes in Math. 476,<br />

Springer, Berlin 1975, 33–52<br />

Tignol, J.-P.: Sur les decompositions des algèbres à divisions en<br />

produit tensoriel d’algèbres cycliques, in: van Oystaeyen, F. <strong>and</strong><br />

Verschoren, A. (Eds.): Brauer <strong>groups</strong> in ring theory <strong>and</strong> algebraic<br />

geometry, Proceedings of a conference held at Antwerp 1981, Lecture<br />

Notes Math. 917, Springer, Berlin, Heidelberg, New York<br />

ÌÖٸ˺ĺÇÖÓÒÐÒÓ«ÚÚÐÒØÒÓ×ØÑÒÓÓÓ¹ ÖÞÖÙ«Ö¹ËÚÖ¸Í×ÔÅغÆÙ¹´¾¾µ´½½µ¸<br />

1982, 126–145<br />

¾½ß¾½<br />

Tignol, J.-P.: On the corestriction of central simple algebras, Math.<br />

Z. 194(1987), 267–274<br />

English translation: Tregub, S. L.: On the bi<strong>rational</strong> equivalence<br />

of Brauer-Severi varieties, Russian Math. Surveys 46-6(1991), 229<br />

Turner, S.: The zeta function of a bi<strong>rational</strong> Severi-Brauer scheme,<br />

Bol. Soc. Brasil. Mat. 10(1979), 25–50<br />

Ullmo, E.: Positivité et discrétion des <strong>points</strong> algébriques des<br />

courbes, Annals of Math. 147(1998), 167-179<br />

Vaughan, R. C.:The Hardy-Littlewood method, Cambridge Tracts<br />

in Mathematics 80, Cambridge University Press, Cambridge 1981<br />

Verschoren, A.: A check list on Brauer <strong>groups</strong>, in: van Oystaeyen,<br />

F. <strong>and</strong> Verschoren, A. (Eds.): Brauer <strong>groups</strong> in ring theory<br />

<strong>and</strong> algebraic geometry, Proceedings of a conference held at<br />

Antwerp 1981, Lecture Notes Math. 917, Springer, Berlin, Heidelberg,<br />

New York 1982, 279–300<br />

[Wan] Wang, S.: On the commutator group of a simple algebra, Amer. J.<br />

Math. 72(1950), 323–334


382 BIBLIOGRAPHY<br />

[War]<br />

[We48]<br />

Warner, F. W.: Foundations of differentiable manifolds <strong>and</strong> Lie<br />

<strong>groups</strong>, Scott, Foresman <strong>and</strong> Co., Glenview-London 1971<br />

Weil, A.: Sur les courbes algébriques et les variétés qui s’en déduisent,<br />

Actualités Sci. Ind. 1041, Hermann et Cie., Paris 1948<br />

[We56] Weil, A.: The field of definition of a variety, Amer. J. of Math. 78<br />

(1956), 509–524<br />

[Wi35]<br />

ÒÕÚ׸κÁºÈÖÒÔÀ××ÐÖÙÔÔÖÙ«ÖÔÓÐ<br />

Witt, E.: Über ein Gegenbeispiel zum Normensatz, Math. Z. 39<br />

(1935), 462–467<br />

[Wi37]<br />

¾¼´½µ¸¿¼ß¿ ÙÒÒÔÖÓÞÚÒÑÒÓÓÓÖÞËÚÖ¹ÖÙ«Ö ÔÖÓØÚÒÝÚÖ¸ÌÖÙÝÅغÁÒ×غÑÒËØÐÓÚ<br />

Witt, E.: Schiefkörper über diskret bewerteten Körpern, J. für die<br />

Reine und Angew. Math. 176(1937), 153–156<br />

[Ya]<br />

[Yu] Yuan, S.: On the Brauer <strong>groups</strong> of local fields, Annals of Math. 82<br />

(1965), 434–444<br />

[Za]<br />

Zak, F. L.: Tangents <strong>and</strong> secants of algebraic varieties, AMS Translations<br />

of Mathematical Monographs 127, Amer. Math. Soc., Providence<br />

1993<br />

[Zh95a] Zhang, S.: Small <strong>points</strong> <strong>and</strong> adelic metrics, J. Alg. Geom. 4(1995),<br />

281-300<br />

[Zh95b]<br />

[Zh98]<br />

Zhang, S.: Positive line bundles on arithmetic varieties, J. Amer.<br />

Math. Soc. 8(1995), 187-221<br />

Zhang, S.: Equidistribution of small <strong>points</strong> on abelian varieties,<br />

Annals of Math. 147(1998), 159-165


INDEX<br />

A<br />

Abbes, A. . . . . . . . . . . . . . . . . . . . . . . . . . . . 170<br />

abelian surface . . . . . . . . . . . . . . . . . . 218, 255<br />

abelian variety . . . . . . . . . . . . . . . . . . 218, 329<br />

accumulating subvariety xxi, 212, 254f, 258<br />

accumulating surface. . . . . . . . . . . . .255, 258<br />

adelic intersection product . . . . . . . . xx, 183<br />

adelic intersection theory. . . .xx, 183 – 188<br />

adelic metric.175, 182, 187, 193f, 199, 207,<br />

222f<br />

adelic Picard group . xx, 173, 177, 187, 193<br />

adelic point . . xv – xix, xxii, 99ff, 105, 118,<br />

128, 234, 239, 286, 302 f, 326<br />

adelic topology . . . . . . . . . . . . . . . . . . . . . . 101<br />

adelicly metrized invertible sheaf . . . . 175ff<br />

adjunction formula . . . . . . . . . . . . . . . . . . 232<br />

affine cone . . . . . . . . . . . . . . . . . . . . . . 232, 283<br />

algebraic surface . . . . . . . . . . . . . . . . . . . 71, 85<br />

algebraic surfaces<br />

classification of . . . . . . . . . . . . . . . . . . . . 329<br />

Algorithm<br />

Elkies’ method . . . . . . . . . . xxiii, 294, 349<br />

FFT point counting. . . . . . .248, 251, 260<br />

for numerical integration.251, 291f, 298<br />

obvious . . . . . . . . . . . . . . . . . . . . . . . . . . . 247<br />

of Fincke-Pohst. . . . . . . . . . . . . . . . . . .350 f<br />

of Tate. . . . . . . . . . . . . . . . . . . . . . . . . . . .142<br />

to compute approximate value for Peyre’s<br />

constant . . . . . . . . . . . . . . . 251, 290, 293<br />

to detect conics on quartic threefold 257<br />

to detect lines on cubic threefold . . 255,<br />

260<br />

to generate triangular mesh. . . . . . . . .291<br />

to search for solutions of Diophantine<br />

equation.xxi, xxiii, 259, 284, 294, 296,<br />

332 – 348<br />

naive . . . . . . . . . . . . . . . . . . . . . . . . . . . 332<br />

to solve system of equations over finite<br />

field. . . . . . . . . . . . . . . . . . . . . . . .255, 257<br />

to test for conic through two <strong>points</strong>.256<br />

to test lines for ir<strong>rational</strong>ity . . . . . . . 254f<br />

to test whetherconic is contained in quartic<br />

threefold. . . . . . . . . . . . . . . . . . . . .257<br />

to verify Galois group acting on 27 lines<br />

xxii, 288f, 293, 296f, 299<br />

2-adic Hensel lift . . . . . . . . . . . . . . . . . . 293<br />

algorithms, for computation of volume 227<br />

alteration. . . . . . . . . . . . . . . . . . . . . . . . . . . .164<br />

AMD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347<br />

Amir-Moez, A. R.. . . . . . . . . . . . . . . . . . . . . .3<br />

Amitsur, S. A. . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

apex . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350<br />

Arakelov, S. Yu. . . . . . . . . . . . . . . . . . 156, 160<br />

Arakelov degree . . . . . . . . . . . . . xx, 157, 162<br />

Arakelov geometry . . . . . . . . . . . . . . 190, 192<br />

archimedean valuation . . . . . . . . . . . . . . . 152<br />

ARIBAS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348<br />

arithmetic Chern class. . . . . . . . . . . . . . . .162<br />

arithmetic Chow group. . . . . . . . .161 – 164<br />

arithmetic cycle . . . . . . . . xx, 160f, 165, 173


384 INDEX<br />

linearly equivalent to zero. . . . . . . . . .161<br />

arithmetic degree . . . . . . . . . . 161, 168, 185f<br />

arithmetic Hilbert-Samuel formula . . . . . xx<br />

arithmetic intersection product . . 163, 165,<br />

168, 185, 208<br />

arithmetic intersection theory . . . . . xx, 155<br />

arithmetic Picard group . . . . . . . . . . . . . . 156<br />

arithmetic variety . . . . . . . . . xix, 156, 158 ff,<br />

162 – 167, 169f, 172, 185<br />

singular . . . . . . . . . . . . . . . . . . . . . 165 – 168<br />

Artin, M. . . . . . . . . . . . . . . . . . . . . . . . 4, 34, 38<br />

Artin conductor. . . . . . . . . . . . . . . .309f, 314<br />

Artin L-function. . . . . .228, 303, 314ff, 326<br />

assembly. . . . . . . . . . . . . . . . . . . . . . . .344, 350<br />

Ausl<strong>and</strong>er-Goldman, Theorem of xvii, 71,<br />

84<br />

automorphisms of P n−1 . . . . . . . . . . . 29, 155<br />

Azumaya algebra . . 61 – 70, 78, 80, 82f, 85,<br />

104, 109<br />

B<br />

bad<br />

prime . . . . . . . . . . . . . . . . . . . . . . . 179f, 251<br />

reduction . . . . . . . . . . . . . . . . . . . . .136, 297<br />

barycenter . . . . . . . . . . . . . . . . . . . . . . . . . . 292<br />

Batyrev, V. V. . . . . . . . . . . . . . . . . . . . . . . . .217<br />

Batyrev <strong>and</strong> Manin, Conjecture ofxxi, 214,<br />

216 – 219, 255<br />

Bernstein, D. . . . . . . . . . . . . . . 259, 349, 351f<br />

Bezout’s theorem . . . . . . . . . . . . . . . . . . . . 253<br />

bielliptic surface. . . . . . . . . . . . . . . . .255, 329<br />

Bierce, A. . . . . . . . . . . . . . . . . . . . . . . . . . . . 329<br />

Bilu, Yu.. . . . . . . . . . . . . . . . . . . .xxi, 190, 192<br />

equidistribution theorem ofxxi, 190, 192<br />

biquadratic reciprocity low . . . . . . . . . . . . xv<br />

Bismut, J.-M. . . . . . . . . . . . . . . . . . . . . . . . . 170<br />

Bogomolov, F. . . . . . . . . . . . . . . . . . . . . . . . 219<br />

Conjecture of . . . . . . . . . . . . . . . . . . . . . 219<br />

Bost, J.-B. . . . . . . . . . . . . . . . . . . . . . . . . xx, 167<br />

Bouche, T.. . . . . . . . . . . . . . . . . . . . . . . . . . .170<br />

Bourbaki, N.. . . . . . . . . . . . . . . . . . . . .3 f, 176<br />

Brauer, R. . . . . . . . . . . . . . . . . . . . . . . . . . . 4, 22<br />

Brauer group . . . . . . .xv ff, 22, 66, 71 f, 74 ff,<br />

84 – 97, 286<br />

cohomological. . . . .xvii, 70 – 76, 84 – 97<br />

of cubic surface. . . . . . . . . . . . . . . . . . . . .96<br />

Brauer-Manin obstruction . . . . . . . . . . . xv ff,<br />

xix, xxii, 97, 103, 105 f, 110, 113, 122,<br />

128f, 144f, 212, 237, 239, 244, 246,<br />

281, 286f, 293, 301ff<br />

is the only obstruction . . . . . . . . . . . . . 106<br />

to <strong>rational</strong> <strong>points</strong> . . . . . . . . . . . . . . . . . . 106<br />

to the Hasse principle . . . xviii, 106, 113,<br />

122ff, 130, 212, 239, 286<br />

to weak approximation. .xxii, 106, 122f,<br />

127, 130, 283, 286<br />

Brauer-Severi variety. . . . .xvi f, 3 f, 22, 28 f,<br />

32 – 39, 46, 48, 51 ff, 102<br />

that splits . . . . . . . . . . . . . . . . . . . . xvi, 28ff<br />

Bremner, A.. . . . . . . . . . . . . . . . . . . . . .xv, 103<br />

Brown-Gersten-Quillen spectral sequence<br />

164<br />

Bucket sort. . . . . . . . . . . . . . . . . . . . . . . . . .345<br />

buffer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346f<br />

Burgos Gil, J. I. . . . . . . . . . . . . . . . . . 156, 165<br />

C<br />

C . . . . . . . . . . . . . . . . . . . . . . 296, 344, 348, 350<br />

C++ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350<br />

Cantor, M. . . . . . . . . . . . . . . . . . . . . . . . . . . 353<br />

Cartan, E. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

Cassels, J. W. S. . . . . . . . . . .xv, 103, 265, 283<br />

Čech cocycle . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

Čech cohomology. . . . . . . . . . . . . . . . . . .68ff<br />

center. . . . . . . . . . .3, 7, 22, 28, 45, 48, 52, 62<br />

central simple algebra . . . . . . . . . . . . . . . xvi f,<br />

3 ff, 10, 22 f, 35 – 39, 45 – 54, 58, 61ff,<br />

74, 76, 78, 82 f<br />

index of. . . . . . . . . . . . . . . . . . . . . . . . . . . .22<br />

that splits . . . . . . . . . . . . . . xvi, 23f, 45, 81<br />

Chambert-Loir, A. . . . . . . . . . . . . . . . xxi, 192<br />

equidistribution theorem of . . . . xxi, 192<br />

Châtelet, F. . . . . . . . . . . . . . . . . . . . . 4f, 34, 38<br />

Chebotarev density theorem .290, 309, 315<br />

Chebyshev’s inequalities. . . . . . . . . . . . . .314<br />

Chinese remainder theorem. . . .341f, 345f<br />

circle method . . . . . xiii, xvi, 238ff, 244, 282<br />

class number formula . . . . . . . . . . . . . . . . 317<br />

classification of algebraic surfaces . . . . . 329<br />

clustering . . . . . . . . . . . . . . . . . . . . . . . . . . . 335


INDEX 385<br />

cocycle . . . . . . . . . . . 6f, 24f, 30 – 34, 37 f, 47<br />

cohomologous to another . . . . . 6, 24, 30<br />

trivial . . . . . . . . . . . . . . . . . . . . . . . . . . . 6, 33<br />

cocycle relations . . . . . . . . . . . . . 12, 108, 180<br />

local in étale topology. . . . . . . . . . . . . . .65<br />

cohomological Brauer group . xvii, 70 – 76,<br />

84 – 97<br />

Colliot-Thélène, J.-L..xv f, xix, 4, 103, 106,<br />

129f, 239, 286f, 302, 312<br />

Conjecture of . . . . . . . . . . . . . . . . . . . . . 106<br />

Colliot-Thélène <strong>and</strong> Sansuc, Conjecture of<br />

287<br />

Colliot-Thélène, Kanevsky, <strong>and</strong> Sansuc,<br />

Theorem of. . . . . . . . . . . . . . . . . . . .129 f<br />

collision. . . . . . . . . . . . . . . . . . . . . . . . . . . . .335<br />

computer graphics . . . . . . . . . . . . . . . . . . . 291<br />

conductor-discriminant formula . . . . . . 309<br />

congruencexiii, xix, 38, 144 f, 238, 269, 287,<br />

314, 330 – 334, 336 – 342, 345f, 348<br />

conic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256ff<br />

Conjecture<br />

of Batyrev <strong>and</strong> Manin . . . . . . . . . xxi, 214,<br />

216 – 219, 255<br />

of Bogomolov . . . . . . . . . . . . . . . . . . . . . 219<br />

of Colliot-Thélène . . . . . . . . . . . . . . . . . 106<br />

of Colliot-Thélène <strong>and</strong> Sansuc . . . . . . 287<br />

of Lang . . . . . . . . . xii, xxi, 214ff, 255, 329<br />

geometric. . . . . . . . . . . . . . . . . . . . . . .214<br />

strong . . . . . . . . . . . . . . . . . . . . . . . . . . 214<br />

weak. . . . . . . . . . . . . . . . . . . . . . .214, 217<br />

of Manin . . . . . . xiv, xxi, 214, 223f, 239 f,<br />

243f, 246, 256, 261, 263, 265, 282, 284,<br />

293, 296, 301<br />

of Manin-Peyre . . . . . . . . . . . 237, 239, 261<br />

of Mordell . . . . . . . . . . . . . . . . . . . . . . . . 215<br />

control function. . . . . . . . . . . . . . . . .341, 345<br />

convolution. . . . . . . . . . . . . . . . . . . .205, 247 f<br />

Corn, P. K. . . . . . . . . . . . . . . . . . . . . . . xvii, 96<br />

cubic reciprocity low . . . . . . . . . . . . . xv, 145<br />

cubic surface. . . . . . . . . . . . . . .xv – xix, xxi f,<br />

94 – 97, 103, 106, 113f, 122, 125 – 130,<br />

143, 224 – 228, 234, 239 f, 252f, 256,<br />

281 – 328, 353<br />

diagonal . xxii, 103, 128ff, 143, 301 – 328<br />

general . . . . . . . . xxii, 106, 240, 281 – 300<br />

minimal . . . . . . . . . . . . . . . . . . . . . . . . . . 312<br />

cubic threefold . . . . xxi, 215, 240, 243 – 263<br />

diagonal . . . . . . . . . . . . . . . . . xxi f, 213, 224<br />

curvature . . . . . . . . . . . . . . . . . . . . . . 199, 201f<br />

current . . . . . . . . . . . . . . . . . . . . . . . 198, 201<br />

form . . . . . . . . . . . .167, 170, 177, 179, 183<br />

cusp . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252<br />

D<br />

de Jong, A. J. . . . . . . . . . . . . . . . . . . . . . . . . 164<br />

decomposition type . . . . . . . . . . . . . . . . . . 289<br />

decoupled equation. . . . . . . . .xxiii, 259, 335<br />

Dedekind zeta function . . . . . . . . . . 250, 317<br />

Deligne, P. . . . . . . . . . . . . . 165, 235, 247, 269<br />

Deligne-Beilinson cohomology. . . . . . . .165<br />

∂∂-lemma. . . . . . . . . . . . . . . . . . . . . . . . . . .162<br />

Derenthal, U.. . . . . . . . . . . . . . . . . . . . . . . .228<br />

descent. . .xvi, 8, 10 – 15, 19, 22, 25, 31, 33,<br />

37 f, 47, 49, 51, 58 ff, 64f, 181<br />

Dickson, L. E. . . . . . . . . . . . . . . . . . . . . . . . 288<br />

’s list. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .288<br />

Dieudonné, J.. . . . . . . . . . . . . . . . . . . . . . . . .17<br />

difference scheme . . . . . . . . . . . . . . . . . . . . 351<br />

Diophantine equation . . . . . . . . . . . xi – xiv,<br />

xvi, xxi ff, 149, 259, 266, 284, 294, 296,<br />

329, 332 – 348<br />

decoupled . . . . . . . . . . . . . . . xxiii, 259, 335<br />

Diophantus. . . . . . . . . . . . . . . . . . . . . . .xi, 353<br />

Dirichlet character . . . . . . . . . . . . . . . . . . . 316<br />

Dirichlet’s prime number theorem . . . . 270<br />

distance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176<br />

division algebra . . . . . . . . . . . . . . . . . . 78 – 81<br />

Duke, W.. . . . . . . . . . . . . . . . . . . . . . . . . . . .315<br />

dynamical system . . . . . . . . . . . . . . . . . . . . 195<br />

E<br />

Eckardt point. . . . . . . . . . . . . . . . . . . . . . . .253<br />

Edidin, D. . . . . . . . . . . . . . . . . . . . . . . . . . . . .85<br />

effective cone . . . . . . . . . . . . . . . . . . 224 – 227<br />

Eisenstein polynomial. . . . . . . . . . . . . . . .115<br />

Ekedahl, T. . . . . . . . . . . . . . . . . . . . . . . . . . . 297<br />

elementary symmetric functions. .153, 311<br />

Elkies, N. . . . . . . . . . . . . . . . . . . . . . . . . . . . 349<br />

’ method . . . . . . . . . . . . . . . xxiii, 294, 349f


386 INDEX<br />

elliptic cone . . . . . . . . . . . . . . . . . . . . . . . . . 252<br />

elliptic curve . . . . . . . . . . . . . . . . . . . . . . . . 137<br />

elliptic surface . . . . . . . . . . . . . . . . . . . . . . . 329<br />

Elsenhans, A.-S. . . .243, 265, 281, 301, 329,<br />

349<br />

Emerson, R. W.. . . . . . . . . . . . . . . . . . . . . .301<br />

Enriques surface . . . . . . . . . . . . . . . . . . . . . . 89<br />

equidistribution . . . xvi, xx, 142, 190f, 208,<br />

246, 266, 284, 301<br />

theorem of Bilu . . . . . . . . . . . xxi, 190, 192<br />

theorem of Chambert-Loir . . . . . xxi, 192<br />

theorem of Szpiro, Ullmo, <strong>and</strong> Zhangxx,<br />

190<br />

étale cohomology . . . . . . . . . . . . 75, 222, 235<br />

étale morphism . . . . . . . . . . . . . . . . . . . . .229f<br />

étale neighbourhood. . . . . . . . . . . .69, 87, 90<br />

étale topology . . . . . . . . . . . . . . 65, 68, 71, 73<br />

Euler product. . . . . . . . .249 ff, 290, 296, 298<br />

Euler sequence. . . . . . . . . . . . . . . . . . . . . . .233<br />

even . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285<br />

exp<strong>and</strong>ing map . . . . . . . . . . . . . . . . . . . . . . 196<br />

exponential sequence . . . . . . . . . . . . . 88, 222<br />

F<br />

factor α . . . . xiv, xxi, 224 – 228, 234, 238f,<br />

245f, 283, 301, 306, 326<br />

factor β . . .xvi, 228, 244, 283, 301, 305, 326<br />

Faltings, G. . . . . . . . . . . . . . . . . .149, 156, 215<br />

Fano variety . . . . . . . . . . . . . . . . . xii, xiv, xxi,<br />

213f, 217, 220f, 223, 236 f, 239, 243f,<br />

246, 265, 282, 285, 301<br />

Faraday, M.. . . . . . . . . . . . . . . . . . . . . . . . . . .99<br />

Fermat cubic . . . . . . . . . . . . . . . . . . . . . . . . 213<br />

FFT convolution . . . . . . . . . . . . . . . . . . . . 248<br />

FFT point counting . . . . . . . . . . . . . 251, 260<br />

Fincke-Pohst, algorithm of . . . . . . . . . . 350f<br />

first arithmetic Chern class . . . . . . . . . . . 162<br />

first cohomology set . . . . . . . . . . . . . . . . . . . 6<br />

flagstone . . . . . . . . . . . . . . . . . . . . . . . . . . . 350f<br />

Forster, O. . . . . . . . . . . . . . . . . . . . . . . . . . . 348<br />

fpqc-topology . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

Franke, J. . . . . . . . . . . . . . . . . . . . . . . . xiv, 223<br />

Frobenius eigenvalues. . . . . . . . . . . .236, 311<br />

Fubini-Study metric. . . . . . . . . . . .156ff, 193<br />

functor of <strong>points</strong> . . . . . . . . . . . 5, 41, 53, 55 ff<br />

functor represented by a scheme . . . . . . . 41<br />

fundamental finiteness . xx, xxii, 149 f, 160,<br />

167, 194, 267, 270, 279, 304, 316, 328<br />

G<br />

G-group . . . . . . . . . . . . . . . . . . . . . . . . 5 – 8, 34<br />

G-morphism . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

G-set . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 – 8, 94<br />

Gabber, O. . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

GAP xvii, 95f, 126, 225f, 285, 289, 305, 353,<br />

356<br />

Gauß-Legendre formula . . . . . . . . . . . . . 251f<br />

Gauß sum . . . . . . . . . . . . . . . . . . . . . . . . . . . 307<br />

Generalized Riemann Hypothesis284, 312,<br />

315f<br />

genus one curve . 119, 129, 135, 138ff, 258,<br />

329<br />

Gilbert, W. S.. . . . . . . . . . . . . . . . . . . . . . . .243<br />

Gillet, H.. . . . .xx, 156, 160, 163ff, 167, 170<br />

global class field theory . . . . . . . . . xv, 86, 91<br />

global evaluation map. .see also Manin map<br />

gluing data . . . . . . . . . . . . . . 65, 68f, 108, 180<br />

GMP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350<br />

Godeaux surface . . . . . . . . . . . . . . . . . . . . . 215<br />

good<br />

prime . . . . . . . . . . . . . . . . . . . . . . . . 179, 299<br />

reduction . 269f, 274, 297, 299, 307, 310,<br />

321, 331<br />

Gröbner base . . . . . . 127, 288, 293, 297, 299<br />

Graßmann<br />

functor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

scheme. . . . . . . . . . . . . . . . . . . . . . . . . . . . .54<br />

Green’s<br />

current . . . . . . . . . . . . . . . . . 160f, 163, 165<br />

form of log type . . . . . . . . . . . . . . 163, 165<br />

Grothendieck, A. . . xvii, 4f, 17, 22, 39, 41,<br />

53, 71, 85<br />

Guy, M. J. T. . . . . . . . . . . . . . . . . . . . . . xv, 103<br />

H<br />

Hankel, H. . . . . . . . . . . . . . . . . . . . . . . . . . . . xi<br />

hardware . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341<br />

Hardy, G. H. . . . . . . . . . . . . . . . . . . . . . . . . xiii<br />

hash function . . . . . . . . . . . . . . .334, 341, 345<br />

hash table . . . . . . . . . . . . . . 259, 335, 345, 347<br />

hashing. . . . . . . . . . . . . . . . . . . .xxiii, 334, 345


INDEX 387<br />

uniform . . . . . . . . . . . . . . . . . . . . . . . . . . 333<br />

Hasse, H. . . . . . . . . . . . . . . . . . . . . . xiii, 4, 265<br />

’s bound . . . . . . . . . . . . . . . . . . . . . . 119, 139<br />

Hasse principle. . .xiii – xvi, xix, 102f, 106,<br />

113, 265f, 281, 284f, 287<br />

Brauer-Manin obstruction to . xviii, 106,<br />

113, 122ff, 130, 212, 239, 286<br />

counterexample to . . . . . xviii, 103, 122ff<br />

obstruction to . . . . . . . . . . . . . . . . . . . . .296<br />

Hassett, B.. . . . . . . . . . . . . . . . . . . . . . . . . . . .85<br />

Hawking, S. . . . . . . . . . . . . . . . . . . . . . . . . . 149<br />

Heath-Brown, D. R. . . . xvi, 145, 239f, 312,<br />

349, 352<br />

’s congruences . . . . . . . . . . . . . . . . . . . . . 145<br />

height<br />

absolute 152, 154, 168, 172, 188, 193, 222<br />

adelic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188<br />

anticanonical . . xiii, 217, 222, 237, 265 f,<br />

282, 284, 301<br />

Bost-Gillet-Soulé . . . . . . . . . . . . . . xx, 167 f<br />

canonical. .190, 192, 217, 222, 237, 265 f,<br />

282, 284, 301<br />

canonical . . . . . . . . . . . . . . . . . . . . . . . . . xiii<br />

defined by adelic metric . . . . . . . . . . . . 222<br />

defined by an invertible sheaf. . .xix, 154<br />

l 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .158<br />

logarithmic . . . . . . . . . . . . . . . . . . . 154, 158<br />

Néron-Tate . . . . . . . . . . . . . . . . . . . . . . 190 f<br />

naive . . . . xiii, xix f, 149 – 152, 158, 173,<br />

190ff, 244ff, 266, 279, 282, 301, 328<br />

normalized. . . . . . . . . . . . . . . . . . . . . . . .172<br />

of smallest point . . . . . . . . xxii, 281, 295 f<br />

with respect to hermitian line bundlexix,<br />

158<br />

Hensel, K. . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii<br />

’s lemma . . . .xiii, 66ff, 75, 104, 115, 118,<br />

135, 141, 230, 269, 290, 313, 336<br />

hermitian line bundle .xx, 156 – 159, 162 f,<br />

167f, 183, 223<br />

continuous. . . . . . . . . . . . . . . . . . . . . . . .157<br />

smooth . . . . . . . . . . . . . . . . . . . . . . . . . . . 157<br />

hermitian metric . . xx, 156, 159f, 174, 200,<br />

202, 245<br />

almost semiample . . 179, 188, 194, 199ff,<br />

204<br />

bounded . . . . . . . . . . . . . . . . . . . . . . . . . . 174<br />

continuous . . . . 156, 173f, 200f, 222, 246<br />

positive . . . . . . . . . . . . . . . . . . . . . . . . . . . 179<br />

smooth . . . . . . . . . . . . . . . . . 162, 173f, 199<br />

heuristic . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211<br />

Hirzebruch, F. . . . . . . . . . . . . . . . . . . . . . . . 247<br />

Hochschild-Serre spectral sequence xvi, 90,<br />

111<br />

hypersurface measure. . . . . . . . . . .231, 322f<br />

I<br />

index<br />

of central simple algebra . . . . . . . . . . . . 22<br />

induced<br />

character. . . . . . . . . . . . . . . . . . . . . . . . .307f<br />

representation. . . . . . . . . . . . . . . . . . . .307f<br />

inertia group . . . . . . . . . . . . . . . . . . . . . . . . 283<br />

inflation . . . . . . . . . . . . . . . . 7 f, 112, 127, 134<br />

integrable metric. . . . . . . . . . . . . . . . . . . . .177<br />

integrable topology. . . . . . . . . . . . .177f, 183<br />

integrably metrized invertible sheaf. . . .xx,<br />

177, 188<br />

intersection of two sets . . . . . . . . . . . . . . .332<br />

invariant map. . . . . . . . . . . . . . . . . . . . . . . . .86<br />

ir<strong>rational</strong>ity test for lines . . . . . . . . . . . . 254f<br />

isomorphism test<br />

J<br />

for cubic surfaces . . . . . . . . . . . . . . . . . . 296<br />

Jacobi sum . . . . . . . . . . . . . . . . . 116, 248, 307<br />

Julia set . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196<br />

K<br />

K3 surface . . . . xxii f, 89, 92, 94, 218ff, 255,<br />

329f<br />

Kanevsky, D.. . . . . .xix, 103, 129f, 239, 286<br />

Kelvin, W. Thomson 1st Baron . . . . . . . 211<br />

Kersten, I. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

Khayyám, O. . . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

Kleiman, S.. . . . . . . . . . . . . . . . . . . . . . . . . .186<br />

Kodaira classification . . . . . . . . . . . . 213, 255<br />

Kodaira dimension. . . . . . . . . . . . . . . . . . .255<br />

Kramer, J. . . . . . . . . . . . . . . . . . . . . . . 156, 165<br />

Kresch, A.. . . . . . . . . . . . . . . . . . . . . . . . . . . .85


388 INDEX<br />

Kühn, U. . . . . . . . . . . . . . . . . . . . . . . . 156, 165<br />

Kummer pairing . . . . . . . . . . . . . . . . . . . . . 308<br />

Kummer sequence . . . . . . . . 72, 87f, 91, 222<br />

Kummer surface . . . . . . . . . . . . . . . . . . . . . 218<br />

L<br />

L1 cache . . . . . . . . . . . . . . . . . . . . . .343f, 347f<br />

L2 cache . . . . . . . . . . . . . . . . . . . . . . . 343f, 348<br />

l 3 -unit sphere . . . . . . . . . . . . . . . . . . . . . . . . 323<br />

Lang, S. . . . . . . . xii, xxi, 214 – 217, 255, 329<br />

’s conjecture. . . . .xii, xxi, 214ff, 255, 329<br />

geometric. . . . . . . . . . . . . . . . . . . . . . .214<br />

strong . . . . . . . . . . . . . . . . . . . . . . . . . . 214<br />

weak. . . . . . . . . . . . . . . . . . . . . . .214, 217<br />

Langian exceptional set. . . . . . . . . . . . . .214f<br />

Larsen, K. . . . . . . . . . . . . . . . . . . . . . . . . . . . 149<br />

lattice basis reduction. . . . . . . . . . .124, 350f<br />

Lebeau, G. . . . . . . . . . . . . . . . . . . . . . . . . . . 170<br />

Lebesgue measure . . . . . . . . . . . . . . . 169, 224<br />

Lefschetz hyperplane theorem . . . . . . . . . 89<br />

Lefschetz theorem on (1, 1)-classes. . . . . .88<br />

Lefschetz trace formula. . . . . .235, 307, 310<br />

Legendre, A.-M.. . . . . . . . . . . . .xiii, 265, 283<br />

’s Theorem . . . . . . . . . . . . . . . xiii, 265, 283<br />

Lemma<br />

of Grothendieck . . . . . . . . . . . . . . . . . . . .39<br />

of Grothendieck <strong>and</strong> Dieudonné. . . . .17<br />

of Hensel . . . xiii, 66 ff, 75, 104, 115, 118,<br />

135, 141, 230, 269, 290, 313, 336<br />

of Wedderburn <strong>and</strong> Brauer . . . . . . . . . . 22<br />

of Yoneda. . . . . . . . . . . . . . . . . . . . . . . . .107<br />

of Zorn. . . . . . . . . . . . . . . . . . . . . . . .10, 107<br />

Leray measure . . 231f, 238, 245f, 304, 322f<br />

Lichtenbaum, S. . . . . . . . . . . . . . . . . . . . . . . 92<br />

Theorem of . . . . . . . . . . . . . . . . . . . . . . . . 92<br />

Lichtenbaum duality. . . . . . . . . . . . . .92, 144<br />

Lind, C.-E. . . . . . . . . . . . . . . . . . . . . . xiv f, 103<br />

line . . . xii, xviii, 115, 117 ff, 121, 125f, 128,<br />

131, 136, 212f, 215f, 219, 225ff, 244,<br />

252 – 258, 260, 285, 287f, 299<br />

non-obvious. . . . . . . . . .xxi, 216, 255, 260<br />

obvious . . . . . . . . . . . . . xxi, 215 f, 253, 255<br />

sporadic. . . . . . . . . . . . . .xxi, 216, 255, 260<br />

–s, 27 on cubic surface . . see also 27 lines<br />

on cubic surface<br />

linear probing . . . . . . . . . . . . . . . . . . . . . . . 335<br />

linear subspace. . . . . . . . . . . . . . .33 f, 38, 244<br />

Linnik’s Theorem. . . . . . . . . . . . . . . . . . . .312<br />

Littlewood, J. E. . . . . . . . . . . . . . . . . . . . . . xiii<br />

LLL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350f<br />

local evaluation map . . . . . . 103f, 129f, 135<br />

local measure . . . . . . . . . . . . . . . . . . . . . . . . 230<br />

log-factor. . . . . . . . . . . . . . . . . . . . . . . .xiv, 223<br />

Lovasz, L. . . . . . . . . . . . . . . . . . . . . . . . . . . . 227<br />

M<br />

Maclagan-Wedderburn, J. H.. . . . . . . . .3, 22<br />

Magma .. . . . . . . . . . . . . . . . . . . . . . . . . 294, 297<br />

Manin, Yu. I. . . . xiv f, xix, 93, 96, 103, 113,<br />

217, 223, 281<br />

’s conjecture . . .xiv, xxi, 214, 223f, 239f,<br />

243f, 246, 256, 261, 263, 265, 282, 284,<br />

293, 296, 301<br />

’s formula . . . . . . . . . . . . . . . . . . . . . . . . . . 93<br />

Manin map . . . . . . . . . . . xvii, 103f, 110, 134<br />

Manin-Peyre, Conjecture of. .237, 239, 261<br />

maple.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .118<br />

Maxwell, J. C. . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

McKinnon, D.. . . . . . . . . . . . . . . . . . . . . . .218<br />

memory architecture. . . . . . . . . . . . . . . . .343<br />

Mergesort . . . . . . . . . . . . . . . . . . . . . . . . . . . 345<br />

Merkurjev, A. S. . . . . . . . . . . . . . . . . . . . . . . . 4<br />

Merkurjev-Suslin<br />

Theorem of . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

mesh generation<br />

2.5-dimensional. . . . . . . . . . . . . . . . . . . .292<br />

metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173<br />

bounded . . . . . . . . . . . . . . . . . . . . . . . . . . 174<br />

continuous. . . . . . . . . . . . . . . . . . . . . . . .174<br />

induced by a model . . . . . . . . . . . . . . . . 174<br />

minimum metric. . . . . . .xx, 156ff, 173, 246<br />

Minkowski, H. . . . . . . . . . . . . . . . . . . . . . . xiii<br />

’s lattice point theorem. . . . . . . . . . . . .171<br />

model. . . . . . . . . . .xix, 101f, 104, 108 f, 129,<br />

135f, 139, 142 f, 173ff, 177, 179, 181ff,<br />

187f, 193, 220, 223, 229, 234, 245, 252,<br />

267, 283, 300, 303<br />

modular operation . . . . . . . . . . . . . . . . . . . 341


INDEX 389<br />

Molien, T. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

Monte-Carlo method. . . . . . . . . . . . . . . . .227<br />

Mordell, L. J. . . . . . . . xiv, xix, 103, 113, 215<br />

’s conjecture. . . . . . . . . . . . . . . . . . . . . . .215<br />

’s examples of cubic surfaces . . . xiv, 103,<br />

113 – 128<br />

Moriwaki, A.. . . . . . . . . . . . . . . . . . . . . . . .165<br />

morphism twisted by σ. . . . . . . . . . . . . . . . .8<br />

moving lemma. . . . . . . . . . . . . . . . . . . . . . .164<br />

Mumford, D. . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

N<br />

Néron-Severi group. . . . . . . . . . . . . . . .88, 91<br />

nef . . . . . . . . . . . . . . . . . . 177 – 180, 183, 185 f<br />

Noether, E.. . . . . . . . . . . . . . . . . . . . . . . . . . . .4<br />

Noether-Lefschetz Theorem . xvi, 244, 252<br />

non-archimedean valuation . . . . . . . . . . . 152<br />

non-Azumaya locus . . . . . . . . . . . 62, 85, 110<br />

non-obvious line . . . . . . . . xxi, 216, 255, 260<br />

norm<br />

of global section . . . . . . . . . . . . . . . . . . . 169<br />

norm variety . . . . . . . . . . . . . . . . . . . . . . . . 265<br />

normalized valuation. . . . . . . . . . . . . . . . .150<br />

Northcott’s theorem . . . . . . . . . . . . . . . . . 152<br />

numerical integration . . . . . . 251, 291f, 298<br />

O<br />

obvious line . . . . . . . . . . . xxi, 215f, 253, 255<br />

odd . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285<br />

Oliver, J. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 349<br />

open addressing. . . . . . . . . . . . . . . . . . . . . .335<br />

Opteron processor . . . . . . . . . . . . . . . . . . . 347<br />

order. . . . . . . . . . . . . . . . . . .76 – 79, 82ff, 187<br />

maximal . . . . . . . . . . . . . . . 77 – 83, 85, 187<br />

P<br />

p-adic measure . . . . . . . . . . . . . . . . . . . . . 229 f<br />

p-adic unsolvability . . . . . . . . . . . . . . . . . . 212<br />

p-adic valuation . . . . . . . . . . . . . . . . . 270, 317<br />

p-adic numbers . . . . . . . . . . . . . . . . . . . . . . xiii<br />

page prime . . . . . . . . . . . . . 333, 336, 341, 345<br />

paging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 336<br />

parametrization . . . . . . . . . . . . . . . . . . . . . . xii<br />

Pari................................250<br />

partial presorting . . . . . . . . . . . . . . . . . . . . 347<br />

Pentium 4 processor . . . . . . . . . . . . . . . . . 348<br />

periodicity isomorphism, for cohomology<br />

of cyclic group . . . . . . . . . . . . . . . . . . 137<br />

permutation representation. . . . . . . . . . .285<br />

primitive. . . . . . . . . . . . . . . . . . . . . . . . . .288<br />

Peyre, E. . xvi, xix, xxi, 217, 224, 237 – 240,<br />

244, 266f, 282, 284, 296, 298, 301, 303<br />

Peyre’s constant . . . . . . . . . . . . . . . . xxi f, 224,<br />

237 – 240, 243ff, 247, 260, 266ff, 277,<br />

282ff, 295f, 298, 301 f, 311, 325f<br />

Peyre’s Tamagawa type number . . . . see also<br />

Peyre’s constant<br />

Picard functor . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

Picard group. . . . . . . . . . . . . . . . . . . . . . . . . .91<br />

Picard rank<br />

of cubic surface. . . . . . . . . . . . . . . . . . . . .97<br />

Picard scheme . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

Poincaré residue map. . . . . . . . . . . . . . . . .233<br />

Poincaré-Lelong equation. . . . . . . . . . . . .162<br />

post-processing . . . . . . . . . . . . . . . . . . . . . . 342<br />

presorting, partial. . . . . . . . . . . . . . . . . . . .345<br />

prime number theorem<br />

Dirichlet’s. . . . . . . . . . . . . . . . . . . . . . . . .270<br />

principal divisor<br />

norm of. . . . . . . . . . . . . . . . . . . . . . . . . . .137<br />

product formula . . . . . . . . . . . . . . . . . . . . . 150<br />

projective plane. . . . . . . . . . . . . . . . . . . . . .257<br />

projective space of lines. . . . . . . . . . . . . . . .42<br />

Proposition<br />

of Châtelet-Artin. . . . . . . . . . . . . . . .34, 38<br />

of Severi . . . . . . . . . . . . . . . . . . . . . . . . . . . 32<br />

Puiseux expansion . . . . . . . . . . . . . . . . . . . 252<br />

pull-back<br />

of Azumaya algebra. . . . . . . . . . . . . . . . .63<br />

on adelic Picard group . . . . . . . . . . . . . 182<br />

on arithmetic Chow group . . . . . . . . 163f<br />

on group cohomology . . . . . . . . . . . . . . . 7<br />

pure cubic field<br />

discriminant of. . . . . . . . . . . . . . . .309, 316<br />

push-forward<br />

on arithmetic Chow group . . . . . . . . 164f<br />

pyramid. . . . . . . . . . . . . . . . . . . . . . . . . . . . .350<br />

Q<br />

quadratic form . . . . . . . . . . . . . . . . 338ff, 342


390 INDEX<br />

quadratic reciprocity low . . . . . . . . . . . . . xiv<br />

quadric. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .257<br />

quartic surface<br />

diagonal . . . . . . . . . . . . . . . . . . . . . . . . . . 218<br />

quartic threefold.xxi f, 240, 243 – 279, 284,<br />

301<br />

diagonal . . . . . . . . . . . . xxi f, 213, 224, 284<br />

Quicksort . . . . . . . . . . . . . . . . . . . . . . . . . .344f<br />

Quillen, D.. . . . . . . . . . . . . . . . . . . . . . . . . .170<br />

R<br />

Radix sort . . . . . . . . . . . . . . . . . . . . . . . . . . .345<br />

r<strong>and</strong>om number generator. . .xxii, 284, 292<br />

<strong>rational</strong> <strong>points</strong><br />

Brauer-Manin obstruction to . . . . . . . 106<br />

<strong>rational</strong> surface . . . . . . . . . . 4, 89, 92ff, 255f<br />

<strong>rational</strong> variety . . . . . . . . . . . . . . . . . . . . . . 329<br />

Reading . . . . . . . . . . . . . 259, 339f, 342, 346 ff<br />

reciprocity low<br />

biquadratic. . . . . . . . . . . . . . . . . . . . . . . . .xv<br />

cubic . . . . . . . . . . . . . . . . . . . . . . . . . . xv, 145<br />

quadratic . . . . . . . . . . . . . . . . . . . . . . . . . . xiv<br />

reduced trace . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

representability, of a functor . . . . . . . . . . . 53<br />

restriction . . . . . . . . . . . . . . . . . . . . . 7, 36, 134<br />

Riemann zeta function . . . . . . . . . . . . . . . 316<br />

ruled surface. . . . . . . . . . . . . . . .252, 255, 258<br />

ruled variety. . . . . . . . . . . . . . . . . . . . . . . . .329<br />

S<br />

Sansuc, J.-J. . . . . . . xix, 103, 129f, 239, 286f<br />

Schappacher, N. . . . . . . . . . . . . . . . . . . xi, 353<br />

Schröer, S.. . . . . . . . . . . . . . . . . . . . . . . . . . . .85<br />

Schwarz, H. A. . . . . . . . . . . . . . . . . . . . . . . 292<br />

’s cylindrical surface . . . . . . . . . . . . . . . 292<br />

search bound . . . xxiii, 260, 268, 284, 293 ff,<br />

304, 348f<br />

searching for <strong>rational</strong> <strong>points</strong> . . . . . . . see also<br />

searching for solutions of Diophantine<br />

equation<br />

searching for solutions of Diophantine<br />

equation.xxi, xxiii, 259, 284, 294, 296,<br />

332 – 348<br />

naive . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332<br />

selection of bits . . . . . . . . . . . . . . . . . . . . . . 341<br />

semi-abelian variety<br />

split . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192<br />

semipositive. . . . . .177ff, 181, 188, 194, 204<br />

semipositive cone . . .177ff, 181 – 184, 187f<br />

sequence<br />

generic. . . . . . . . . . . . . . . . . . . . . . . . . . . .204<br />

small . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204<br />

Serre, J.-P. . . . . . . . . . . . . . . . . . . . . . . . . . 4, 189<br />

’s homological characterization of regularity<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

’s vanishing theorem . . . . . . . . . . . . . . . . 17<br />

Theorem of . . . . . . . . . . . . . . . . 24, 30, 189<br />

Severi, F. . . . . . . . . . . . . . . . . . . . . . . . . . . . 4, 32<br />

sheaf of Azumaya algebras . . . . . . . . xvii, 61<br />

Siegel, C. L. . . . . . . . . . . . . . . . . 265, 283, 317<br />

’s estimate. . . . . . . . . . . . . . . . . . . . . . . . .317<br />

σ-linear . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

Sikorsky, I. I. . . . . . . . . . . . . . . . . . . . . . . . . 265<br />

similarity<br />

of Azumaya algebras . . . . . . . . . . . 65 f, 70<br />

of central simple algebras . . . . . . 4, 22, 38<br />

simple group . . . . . . . . . . . . . . . . . . . . . . . . 285<br />

SINGULAR.. . . . . . . . . . . . . . . . . . . . . .288, 293<br />

64 bit processor. . . . . . . . . . . . . . . . . . . . . .341<br />

Skolem-Noether, Theorem of . . . 23, 67, 69<br />

smallest point . . . . . . . . . . . . . . . . . . . 283, 295<br />

solutions of Diophantine equation<br />

algorithm to search for. . .xxi, xxiii, 259,<br />

284, 294, 296, 332 – 348<br />

naive . . . . . . . . . . . . . . . . . . . . . . . . . . . 332<br />

sorting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345<br />

Soulé, C. . . . . . xx, 156, 160, 163ff, 167, 170<br />

splitting field . . 23, 28, 32, 34ff, 38, 46f, 49,<br />

58, 60, 75, 288<br />

sporadic line. . . . . . . . . . . .xxi, 216, 255, 260<br />

∗-product of Green’s currents. . . . .163, 165<br />

Stark, H. M.. . . . . . . . . . . . . . . . . . . . . . . . .317<br />

statistical parameters . . . . . . . . . . . 261ff, 295<br />

Steiner surface . . . . . . . . . . . . . . . . . . . . . . . 258<br />

successive differences . . . . . . . . . . . . . . . . . 298<br />

successive minima.xx, 169f, 172, 188, 194f<br />

Sullivan, A.. . . . . . . . . . . . . . . . . . . . . . . . . .243<br />

surface . . . . . . . . . . . . . . . . . . . 85, 94, 224, 252<br />

abelian . . . . . . . . . . . . . . . . . . . . . . . 218, 255


INDEX 391<br />

algebraic. . . . . . . . . . . . . . . . . . . . . . . .71, 85<br />

bielliptic. . . . . . . . . . . . . . . . . . . . . .255, 329<br />

cubic. . .xv – xix, xxi f, 94 – 97, 103, 106,<br />

113f, 122, 125 – 130, 143, 224 – 228,<br />

234, 239f, 252f, 256, 281 – 328, 353<br />

diagonal . . . . . . . . xxii, 103, 128 ff, 143,<br />

301 – 328<br />

general . . . . . . xxii, 106, 240, 281 – 300<br />

elliptic . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329<br />

Enriques . . . . . . . . . . . . . . . . . . . . . . . . . . . 89<br />

Godeaux . . . . . . . . . . . . . . . . . . . . . . . . . .215<br />

in P 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . .xxiii<br />

K3 . . . xxii f, 89, 92, 94, 218ff, 255, 329ff<br />

Kummer . . . . . . . . . . . . . . . . . . . . . . . . . . 218<br />

non-minimal . . . . . . . . . . . . . . . . . . . . . . 220<br />

non-separated . . . . . . . . . . . . . . . . . . . . . . 85<br />

of general type. . . . . . . . . . . . . . .215, 254 f<br />

of Kodaira dimension one . . . . . . . . . . 255<br />

of Kodaira dimension zero . . . . . . . . . 217<br />

<strong>rational</strong> . . . . . . . . . . . . . . . 4, 89, 92ff, 255 f<br />

ruled. . . . . . . . . . . . . . . . . . . . .252, 255, 258<br />

Steiner. . . . . . . . . . . . . . . . . . . . . . . . . . . .258<br />

surfaces<br />

classification of . . . . . . . . . . . . . . . . . . . . 329<br />

Suslin, A. A. . . . . . . . . . . . . . . . . . . . . . . . . . . .4<br />

suspicious<br />

pair. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .257<br />

point . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255<br />

Swinnerton-Dyer, Sir P. xiv, xvii, xix, xxiii,<br />

96, 103, 124, 133, 263, 297, 329f, 335<br />

’s list. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .133<br />

Sylow subgroup . . . . . . . . . . . . . . . . . 285, 287<br />

Szpiro, L. . . . . . . . . . . . . . . . . . . . xx, 190, 192<br />

Szpiro, Ullmo, <strong>and</strong> Zhang, equidistribution<br />

theorem of . . . . . . . . . . . . . . . . . . xx, 190<br />

T<br />

Tamagawa measure. . . . . . . . . .234, 246, 303<br />

Tamagawa number see also Peyre’s constant<br />

Tate, J. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143<br />

’s algorithm . . . . . . . . . . . . . . . . . . . . . . . 142<br />

Tate cohomology . . . . . . . . . . . . . . . . . . . . . 94<br />

tensor field . . . . . . . . . . . . . . . . . . . . . . . . . . 230<br />

test for conic through two <strong>points</strong> . . . . . 256<br />

Theorem<br />

Hilbert 90 . . . . . . . . . . . 26, 33, 71, 90, 112<br />

of Ausl<strong>and</strong>er-Goldman. . . . . .xvii, 71, 84<br />

of Bezout . . . . . . . . . . . . . . . . . . . . . . . . . 253<br />

of Bilu . . . . . . . . . . . . . . . . . . . xxi, 190, 192<br />

of Chambert-Loir . . . . . . . . . . . . . xxi, 192<br />

of Colliot-Thélène, Kanevsky, <strong>and</strong> Sansuc<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . 129f<br />

of Hasse-Minkowski . . . . . . . . . . . . . . . 281<br />

of Legendre. . . . . . . . . . . . . . .xiii, 265, 283<br />

of Lichtenbaum . . . . . . . . . . . . . . . . . . . . 92<br />

of Linnik . . . . . . . . . . . . . . . . . . . . . . . . . 312<br />

of Merkurjev-Suslin . . . . . . . . . . . . . . . . . . 4<br />

of Noether-Lefschetz . . . . . . . . . . 244, 252<br />

of Northcott . . . . . . . . . . . . . . . . . . . . . . 152<br />

of Serre . . . . . . . . . . . . . . . . . . . . 24, 30, 189<br />

of Skolem-Noether. . . . . . . . . . .23, 67, 69<br />

of successive minima . . . . . . . . . . 170, 188<br />

of Szpiro, Ullmo, <strong>and</strong> Zhang. . . .xx, 190<br />

of Tietze . . . . . . . . . . . . . . . . . . . . . . . . . . 160<br />

of Tsen. . . . . . . . . . . . . . . . . . . . . . . . . . . . .89<br />

of Zak . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253<br />

Weak Lefschetz. . . . . . . . . . . . . . . . . . . .247<br />

Thompson, S. . . . . . . . . . . . . . . . . . . . . . . . 211<br />

threefold . . . . . . . . . . . . . . . . . . . 212, 282, 302<br />

cubic . . . . . . . . . . .xxi, 215, 240, 243 – 263<br />

diagonal. . . . . . . . . . . . . . .xxi f, 213, 224<br />

quartic. . . .xxi f, 240, 243 – 279, 284, 301<br />

diagonal . . . . . . . . . . xxi f, 213, 224, 284<br />

Tietze’s Theorem . . . . . . . . . . . . . . . . . . . . 160<br />

Tor-formula . . . . . . . . . . . . . . . . . . . . . . . . . 163<br />

toric variety . . . . . . . . . . . . . . . . . . . . . . . . . 192<br />

Tregub, S. L.. . . . . . . . . . . . . . . . . . . . . . . . . .37<br />

triangular mesh . . . . . . . . . . . . . . . . . . . . . . 291<br />

Tschinkel, Y. . . . . . . . .xiv, xvi, xix, 223, 239<br />

Tsen’s theorem. . . . . . . . . . . . . . . . . . . . . . . .89<br />

27 lines on cubic surfacexvii, xxii, 95f, 106,<br />

126ff, 131f, 144, 225 – 228, 284 – 288,<br />

293, 297, 299 f, 305f, 312, 353<br />

2.5-dimensional mesh generation. . . . . .292<br />

Tychonov topology . . . . . . . . . . . . . . . . . . 102<br />

U<br />

Ullmo, E. . . . . . . . . . . . . . . . xx, 190, 192, 208


392 INDEX<br />

V<br />

valuation<br />

archimedean . . . . . . . . . . . . . . . . . . . . . . 152<br />

lying above another. . . . . . . . . . . . . . . .151<br />

non-archimedean . . . . . . . . . . . . . . . . . . 152<br />

van der Waerden’s criterion . . . . . . . . . . . 288<br />

variety . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii<br />

Fano . . . . . . . . . . . . . . . . . . . . . . xii, xiv, xxi,<br />

213f, 217, 220f, 223, 236 f, 239, 243f,<br />

246, 265, 282, 285, 301<br />

of general type xii f, 213 ff, 217, 254 f, 329<br />

of intermediate type . . . xii, 26, 31, 213f,<br />

217, 297<br />

Vistoli, A. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

von Neumann, J. . . . . . . . . . . . . . . . . . . . . 281<br />

W<br />

weak approximation. . .xvi, xix, xxii, 102f,<br />

106, 128, 240<br />

Brauer-Manin obstruction to . . xxii, 106,<br />

122f, 127, 130, 283, 286<br />

counterexample to . . . . . . . . . . . . . . . . . xvi<br />

Weak Lefschetz Theorem. . . . . . . . . . . . .247<br />

Weil, A. . . . . . . . . . . . . . . . . . . . . . xix, 22, 331<br />

Weil conjectures . . . 235, 247, 269, 274, 331<br />

for curves . . . . . . . . . . . . . . . . . . . . 331, 340<br />

Witt, E. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

Writing . . . . . . . . . . . . . 259, 340ff, 345, 347f<br />

Y<br />

Yoneda’s lemma. . . . . . . . . . . . . . . . . . . . . .107<br />

Z<br />

Zak’s theorem . . . . . . . . . . . . . . . . . . . . . . . 253<br />

zeroth cohomology set . . . . . . . . . . . . . . . . . 5<br />

Zhang, S. . . . . . . . . . . xx, 172f, 187, 190, 192<br />

Zorn’s lemma . . . . . . . . . . . . . . . . . . . . 10, 107

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!