2000 Mathematics Subject Classification. Primary 11G10, 11G20, 11G25; Secondary 11N36.
Weil numbers generated by other Weil numbers and torsion fields of abelian varieties
E. Kowalski
Universite Bordeaux I A2X, 351, cours de la Liberation, 33405 Talence Cedex, France E-mail address : emmanuel.kowalski@math.u-bordeaux1.fr
-
Abstract.
Using properties of the Frobenius eigenvalues, we show that, in a precise sense, “most” isomorphism classes of (principally polarized) simple abelian varieties over a finite field are characterized, up to isogeny, by the sequence of their division fields, and a similar result for “most” isogeny classes. Some global cases are also treated.
1 Introduction
Let
be a power of a prime number and let
be a
-Weil number, i.e., an algebraic integer such that for every automorphism
of
we have
. Let
be the multiplicative group generated inside
by the Galois-conjugates of
. We are interested in the multiplicative group structure of
and particularly in the set (say
) of
-Weil numbers inside
. Clearly,
contains all the conjugates of
, and we wish to know when there is equality.
Our motivation relates to abelian varieties over finite fields. Let
be such an abelian variety over a field with
elements. Weil proved that all eigenvalues of the Frobenius endomorphism
of
are
-Weil numbers. We denote by
the multiplicative group that they generate. If
is simple, then
“is” an algebraic integer and
.
We also denote by
the set of
-Weil numbers in
.
In [K1,Th.3.4] , it is shown that given abelian varieties
and
, the condition
is equivalent with the fact that, for all
coprime with some integer
(which may depend on
and
), we have
, in which case we say that
and
are isokummerian. In the case of a simple variety, therefore, if
is reduced to the conjugates of
, any variety
satisfying the above condition must be isogenous to a power of
.
Some precise results (e.g., for product of elliptic curves) are given in [K1] , and also an example due to Serre of two abelian varieties
and
over a finite field, simple and non-isogenous over
, such that
. (In particular, in such a case
contains strictly the set of conjugates of
).
Our main result shows that for “most” abelian varieties over finite fields,
is the set of eigenvalues of
. What “most” means has to be specified, of course, and there are actually at least two natural ways of doing this, taking
up to isomorphism or isogeny.
We will consider both possibilities, using a common lemma and specialized arguments.
In the isomorphism case, we use rather deep results of Mumford and Chavdarov on what “most” isomorphism classes of abelian varieties over
look like; it is quite appealing that we use here both
-adic methods having to do with ordinariness and
-adic methods related to monodromy of
-adic sheaves. Those also allow us to derive some results for abelian varieties over number fields, although they are conditional on ordinarity assumptions. In the isogeny case, the method is more elementary, based on lattice-point counting, using results of Howe and DiPippo, and the multidimensional large sieve inequality.
We now state precisely our main result. Let
be an integer,
with
prime and
. We introduce the following:
-
is the set of isomorphism classes of principally polarized abelian varieties of dimension
defined over
;
-
is the subset of those varieties
such that any
isokummerian with
is isogenous to a power of
;
-
is the set of isogeny classes of abelian varieties of dimension
defined over
;
-
is the subset of isogeny classes such that any
isokummerian to
is (isogenous to) a power of
.
Theorem 1.1.
Let
and
with
prime and
. We have
|
(1.1)
|
and
|
(1.2)
|
Acknowledgments. The results of Chavdarov [C] which are crucial for this paper were mentioned by N. Katz during a lecture; shortly afterward a question by U. Zannier made me realize that those results could be quite useful to study
and
and improve on [K1] . I thank them both for these lucky coincidences...
2 Determination of
in a special case
In this section we consider only
-Weil numbers, and give a criterion for
to be reduced to the conjugates of
. For simplicity we assume that
does not have real conjugates, hence
is of even degree
. Let
be the Galois closure of
. For every conjugate
of
,
is also a conjugate of
; if we fix an embedding
, we have
, the complex conjugate of
.
Proposition 2.1.
Let
be prime,
with
. Let
be a
-Weil number such that
. Let
denote the Galois group of the Galois closure
of
. Let
,
, be the Galois conjugates of
in
, in complex conjugate pairs. Assume that:
(1) For all
,
,
and
are coprime in
.
(2) For all
,
, there exists
such that
and
for
.
Then
is the set of conjugates of
.
-
Proof.
Let
be the restriction of complex conjugation, so that
and
for every
. Thus setting
we have for
the relations
|
(2.1)
|
Fix a prime ideal
in
dividing
. Since
and
,
are coprime by assumption, we see that
divides one and only one of
and
. We renumber/pair the conjugates so that
, and
for
. Notice this doesn't affect the existence of
,
with properties as stated for the new numbering.
Let
where
is the valuation on
associated to
. Notice that by coprimality again we have
|
(2.2)
|
Let now
be a
-Weil number. We can write
with
. We deduce from
that
|
(2.3)
|
and from this we notice in particular that the sum
can not be zero, in particular not all the
can be zero.
We have
,
, which translate to
Dividing by
, summing and comparing with ( 2.3 ), we see that one of
and
is equal to
and the other is equal to
.
Now we consider
. This is an algebraic integer and therefore
,
. We have
so using ( 2.2 ) these two conditions translate to
| |
| |
which means that
for all
.
Now consider
with
. We have
by ( 2.1 ), hence the integrality conditions
,
mean
| |
| |
The first of these shows that at most one
can be equal to
; the second that at most one
can be equal to
. Both of these can not occur because that would give
, which is impossible. So either there exists exactly one
with
, and the other
are
, which gives
(because one must have
,
); or there exists exactly one
with
(and the other
are
), which gives
(because then
,
). □
Remark 2.2.
Since we actually solved the equations in terms of the parameters
uniquely (for a given
), we have also proved that
,
, form a free generating set of
under the assumptions of the proposition. In particular, the rank of
is then equal to
.
Remark 2.3.
Proposition 2.1 also applies to prove that if
is a product of pairwise geometrically non-isogenous elliptic curves over a finite field
with
elements, the only
-Weil numbers in
are the conjugates of the Frobenius elements for the
(see [
K1,Th.3.4,(5)]
). It also gives back in this case the lemma of Spiess used to prove this statement in loc. cit.
To apply Proposition 2.1 to a simple abelian variety
with Frobenius
, we need criteria for the two conditions involved. Here we start by Condition (1), which has to do with the “behavior at
” (since all primes dividing
are above
in
) of the Frobenius of
.
Recall that an abelian variety
of dimension
over a field
of characteristic
is called ordinary if
, which is the maximal number of
-torsion points therecan be in characteristic
. If
is a finite field with
elements, then
is ordinary if and only if the middle coefficient of the characteristic polynomial of Frobenius is coprime with
(see e.g. [DH] ). The following lemma is certainly well-known.
Lemma 2.4.
Let
with
prime,
, let
be a simple ordinary abelian variety. Then for any eigenvalue
of the Frobenius of
, we have
in the Galois closure of
.
-
Proof.
Let
,
, be the conjugates of
with
. Assume there exists
and a prime ideal
in
dividing both
and
. The middle coefficient
of the characteristic polynomial of Frobenius is given by
In this sum, if
or
, we have
by assumption. Otherwise,
and
are both chosen inside the set
with
elements, and
. Thus
. If
, then
Thus we find that
divides all the terms in the sum giving
. Since
, this means
. Hence
is not ordinary, and the result follows by contraposition. □
This implies that any ordinary abelian variety satisfies Condition (1) of Proposition 2.1 .
Note that the examples of Serre in [K1] are not ordinary (since their endomorphism rings are not commutative, which is another consequence of ordinarity, see e.g. [W,§7] ).
In analogy with
,
, we now denote
-
the set of isomorphism classes of principally polarized ordinary abelian varieties of dimension
defined over
;
-
the set of isogeny classes of ordinary abelian varieties of dimension
defined over
.
Now we come to Condition (2), where there is also a simple sufficiency criterion.
Lemma 2.5.
Let
be a
-Weil number such that
and such that the Galois group of
over
is isomorphic to
, the Weyl group of
. Then
satisfies Condition (2) of Proposition 2.1 .
-
Proof.
Recall that
(the Galois group of a “generic” polynomial
of degree
such that
) can be identified with the group of permutations of
pairs
,
, such that the couples
are stable. In the case of the Galois group of
, the pairs can be identified with the pairs of conjugates
, which shows that
can be identified with a subgroup of
. If it is equal to
, the existence of the required elements
is obvious. □
Corollary 2.6.
Let
with
prime and
. For any simple ordinary abelian variety
of dimension
such that the Galois group
of
is isomorphic to
, the set of
-Weil numbers in
is equal to the set of conjugates of
.
This is immediate from Proposition 2.1 , Lemma 2.4 and Lemma 2.5 . Again we denote:
-
the set of isomorphism classes of absolutely simple principally polarized abelian varieties of dimension
defined over
such that the Galois group of
is isomorphic to
;
-
the set of isogeny classes of absolutely simple abelian varieties of dimension
defined over
such that the Galois group of
is isomorphic to
.
3 General abelian varieties up to isomorphism
We now apply Proposition 2.1 to “generic” isomorphism classes of abelian varieties of dimension
. More precisely, one has to consider (for instance) the moduli space
of abelian varieties of dimension
with a principal polarization, which is known to be irreducible of dimension
over
.
For Condition (1) of Proposition 2.1 , we use Lemma 2.4 . It is known that generic abelian varieties are ordinary (see [ON] 1
.) Thus in
, there exists a dense Zariski open subset
such that the polarized abelian variety parameterized by any
is ordinary. (See also [CL,§5] for a sketch; roughly speaking, ordinarity is an open condition, and we know that ordinary abelian varieties of any dimension exist, for instance products of ordinary elliptic curves).
Proposition 3.1.
Let
with
prime,
. We have
-
Proof.
Mumford's result gives this for the corresponding counting of isomorphism classes of principally polarized abelian varieties with some rigidifying structure; then one deduces the statement above by dividing by the number of choices for the rigidifying data, and dealing with possible extra automorphisms, as done for instance in [KS,10.7,11.3] . □
Remark 3.2.
This is much weaker than what the result of Mumford implies: since the space of abelian varieties is of dimension
and the space of non-ordinary abelian varieties must be of dimension
, we have for
| |
| |
Condition (2) is not so easy to treat. We use Lemma 2.5 , and the crucial fact is that Chavdarov [C] has shown that “most” abelian varieties
with
are simple and satisfy the assumptions of that lemma.
Proposition 3.3.
Let
with
prime and
. Then we have
-
Proof.
This follows from [C,Th.2.1] , applied to a suitably “rigidified” universal family of principally polarized abelian varieties of dimension
over
, after eliminating as before the extra factor counting the rigidifying parameters (compare again [KS,11.3] ). The monodromy groups modulo
involved in applying Chavdarov's Theorem are as large as possible for
because (for instance), it is already the case for the families of jacobians of hyperelliptic curves considered in [KS,Th.11.0.4] , which are the same as those in [C,Ex.2.4] (this is where characteristic
enters). This is due to J.K. Yu (unpublished).
See also below for more discussion of these examples. □
We now deduce from Proposition 3.1 and Proposition 3.3 the first main result of this paper.
Theorem 3.4.
Let
with
prime and
. For
, let
be set of isomorphism classes of principally polarized absolutely simple abelian varieties
of dimension
such that
and
is equal to the set of conjugates of
. Then we have
Informally: “most” abelian varieties
of dimension
over
with
large are simple, ordinary, the group
is isomorphic to
and the only
-Weil numbers in
are
and its conjugates.
By the criterion stated in the introduction for two varieties to be isokummerian over a finite field, we see that this theorem is equivalent with the first part ( 1.1 ) of Theorem 1.1 .
Remark 3.5.
As in [
C]
or [
KS]
, it would be very interesting to have a corresponding result with
and
; and (as in those cases) this seems very hard.
On the other hand, introducing some analytic ideas (a bilinear form estimate for representations of
-adic sheaves and “old-fashioned” large sieve as in [
G]
and Section 4 ), it is possible to improve Proposition 3.3 in some cases (in particular, if
satisfies
) to obtain a sharper estimate
for
; see [
K2,Cor.6.4]
.
If one does not wish to deal with the moduli space, one can apply Chavdarov's theorem to any algebraic family of principally polarized abelian varieties over a finite field
for which the monodromy group mod
is equal to
for almost all
, provided one can check that ordinarity is generic in that family. The simplest example are provided by taking an algebraic family of curves and then the associated jacobian family, which has a canonical principal polarization. If one takes the universal family of curves, then the generic ordinarity is a result of Miller, who gives explicit examples of ordinary curves of every genus and characteristic, so that the result follows from the openness of ordinarity and the irreducibility of the moduli space of curves. The fact that the corresponding monodromy group is
follows again in characteristic
from the examples of families of hyperelliptic curves of [C,Ex.2.4] (see also [KS,10.2] ).
It is natural to want to give similar explicit equations of families of curves which are both generically ordinary and have monodromy
. However note that Miller's families
fail the monodromy test (because they fail the diophantine irreducibility test, see [KS,Lemma10.1.15] , as a simple computation shows). On the other hand, the author couldn't find references to the ordinarity for the families with large monodromy of loc. cit. For the moment, we merely state the following fairly easy result:
Proposition 3.6.
Let
be a prime number,
an integer. Put
if
,
otherwise.
(1) The
-parameter family
of smooth projective curves of genus
over
given by compactification of the affine family
over the open subset
is generically ordinary and has geometric monodromy group modulo
equal to
for
,
.
(2) In particular, there exists
and
such that the
-parameter family
of curves of genus
over
given by
with
is generically ordinary and has geometric monodromy group modulo
equal to
for
,
.
(3) If
and
is even, one can in fact take
, so the family
with
, where
is the group of
-roots of unity, is generically ordinary and has geometric monodromy group modulo
equal to
for
,
.
We now come to some global consequences that follow also from other results of Chavdarov's paper. Those have the virtue of concerning individual abelian varieties, as the exceptions become a set of primes of density
which does not affect (for instance) the Isogeny Theorem.
Proposition 3.7.
Let
be a number field. Let
be
,
, or an odd integer. Let
be an abelian variety of dimension
such that
and such that the set of primes of good reduction
of
where the reduction of
modulo
is ordinary is of density
. Then for any abelian variety
,
is isokummerian to
if and only if
is isogenous to a power of
.
-
Proof.
By Chavdarov's “horizontal” version of his result ([C,Cor.6.9] ), the assumption of
ensures that for all prime ideals
in a set of primes of density
, the reduced variety
is ordinary, absolutely simple and its Frobenius has Galois group
. By Corollary 2.6 , it follows that
must be isogenous to a power of
for any such
. The dimension of
fixes a
such that
for all primes in a set of density
. Then by Faltings's Isogeny Theorem, it follows that
over
. □
The assumption of ordinarity at almost all places for varieties with
is widely expected to hold, but few results are known. For elliptic curves, it is quite easy, but this case of the proposition is already treated in [K1] without this assumption. Here is another situation that can be treated unconditionally (compare with Ogus's theorem quoted in [CL,Th.6.3] ):
Proposition 3.8.
Let
be an abelian surface over
with
.
Then the set of primes of good ordinary reduction for
is of density
.Hence any
is isokummerian to
if and only if
is isogenous to a power of
.
-
Proof.
We use Serre's
-adic methods [S] . Let
be a prime and
the
-adic representation associated to
. Serre has shown (this is already used in the proof of Chavdarov's horizontal theorem) that the image of
is dense. Consider the exterior square
. It is an
-adic representation of rank
and “weight”
, and it is faithful, so the closure
of the image of
is again isomorphic to
.
Moreover, for any prime
of good reduction, the properties of
and standard algebra show that the trace of the image by
of a Frobenius element
at
is the middle coefficient
of the characteristic polynomial of the Frobenius of
modulo
.
Hence, by the characterization of ordinarity already stated,
has ordinary reduction at
if and only if this trace
is not divisible by
.
However, by the Riemann Hypothesis for
modulo
, we have
so if
is not ordinary at
, the trace must belong to set
. Let
be any of these thirteen values. We claim that the set of primes
for which
is of density
. Clearly this implies the proposition.
The proof of the claim is easy: since
, if
satisfies the stated condition then we have
Using
and simple computations, it is easy to see that
is a closed subset of
of Minkowski dimension
(see [S,§3] for the definition of Minkowski, or
-dimension). Hence by Theorem 10 of loc. cit., the set of primes with
is of density
. □
In a general higher dimensional situations (over
, say), the non-ordinary primes are such that the trace
of the
-th exterior power of the representation on the Tate module are divisible by
, which for
allows an unbounded number of values of
(for
,
with
). Even using explicit forms of the Chebotarev density theorem (on GRH) to detect each value, the uniformity is not sufficient to obtain any non-trivial result.
Remark 3.9.
In [
K1]
, the question of the “splitting behavior” of a simple abelian variety
at all primes is also raised: is it true that the reduction modulo
of
remains simple for almost all
? In fact, the “horizontal” statements of Chavdarov already deal with this. For instance, this holds if
has the property that the Galois group of the field
generated by the points of
-torsion of
is equal to
for
large.
4 General abelian varieties up to isogeny
Since Weil numbers, ordinarity, and having Galois group
are all isogeny-invariant properties of abelian varieties, it is natural to ask for analogs of the results of the previous section for isogeny classes of abelian varieties, instead of isomorphism classes. Going directly from one to the other is not easy, since finding the number of isomorphism classes in an isogeny class is a quite delicate question, typically related with class numbers (as can be seen most easily in the case of elliptic curves), see [W,§4.3] .
However, we can use results of DiPippo and Howe to deal directly with isogeny classes.
Note then that it is not necessary to introduce a polarization. This is rather satisfactory since not all isogeny classes contain a principally polarized one; see for instance [H,Th.1.3] ; however it is proved there (Th. 1.2) that any isogeny class of odd-dimensional abelian varieties over a finite field contains a principally polarized one.
Proposition 4.1.
Let
,
with
prime and
. We have
and
|
(4.1)
|
for some constant
.
This is proved by DiPippo and Howe in [DH] (see Theorem 1.1), in fact in a much more precise form. Note in particular that this says intuitively that the “dimension” of the space of isogeny classes of abelian varieties of dimension
is
, half that of the moduli space.
Proposition 4.2.
We have
Using Lemma 2.5 , this shows that the analogue of Theorem 3.4 holds for isogeny classes, and therefore that the second part ( 1.2 ) of Theorem 1.1 holds.
To prove Proposition 4.1 , DiPippo and Howe identify the set of isogeny classes considered with a set of lattice points in a region
. We argue similarly for Proposition 4.2 , except that we do not need to be so precise because we only look for an upper bound on the number of isogeny classes with “smaller” Galois group, which is a question of probabilistic Galois theory. It is straightforward to adapt here the method of Gallagher [G] based on the large sieve inequality. It has already been shown, using those methods, that self-reciprocal polynomials of degree
and bounded height have generically
as Galois group (see [DDS] ), but our parameter set is different.
Let
be an abelian variety of dimension
over a number field. The characteristic polynomial of Frobenius
of
is of degree
with real roots of even multiplicity and complex roots arising in pairs
. Therefore one can write
with
. To
we associate the vector
.
Lemma 4.3.
Let
, let
with
prime and
. For any abelian variety
, the vector
above satisfies
where
-
Proof.
This is obvious by the Riemann Hypothesis and the definition of
. □
The analytic ingredient we need is the following consequence of the large sieve inequality.
Lemma 4.4.
Let
. For
, let
and let
Let
and for all primes
, let
be a finite set of cardinality
. Let
and for any
let
denote the number of
such that
. Then we have
the implied constant depending only on
.
-
Proof.
We derive this from the following multidimensional (trigonometric) large sieve inequality: for any finite set of vectors
such that
for two elements
in
(where
is the distance in
), and for any complex numbers
defined for
, we have
|
(4.2)
|
where the implied constant depends only on
. This is a special case of [Hu,Th.1] .
To obtain the lemma from this, proceed as in Lemma A of [G] , which we repeat for convenience: let
be the characteristic function of
, and expand it in Fourier series
Thus we have
|
(4.3)
|
We have for
|
(4.4)
|
Denote by
the inner sum. We now write by Cauchy's inequality and ( 4.3 )
| |
| |
| |
and applying the trigonometric large sieve inequality ( 4.2 ) with the trivial spacing estimate for distinct vectors
,
,
,
, this gives
so
As, by ( 4.4 ), we have
we are done. □
-
Proof of Proposition 4.2 .
First, for any
-tuple
in a ring
, we denote
and
Let
be an abelian variety and
the characteristic polynomial of Frobenius for
and
the Galois group of its splitting field, which can be seen (in possibly many ways) as a subgroup of
. By Lemma 2 of [DDS] , we have
if
contains a
-cycle, a
-cycle, a
-cycle and a
-cycle.
For
, let
denote the number of lattice points
in the region
defined in Lemma 4.3 such that the polynomial
is either reducible or such that the Galois group
of the splitting field of
, seen as a subgroup of
again, does not contain an
-cycle. By the observation above and Lemma 4.3 , it follows that the number
of isogeny classes of abelian varieties
with
not having Galois group
satisfies
For each
, we know from classical algebraic number theory (see e.g. [vdW,§61] ) that if the polynomial
reduces modulo some prime
to a polynomial
which factorizes as a product of
distinct linear factors and a single irreducible polynomial of degree
, then
contains an
-cycle. Therefore, choosing
arbitrary and putting
for
, we see that for
such that
does not contain an
-cycle we have
for all
. With notation as in Lemma 4.4 with
(so
), we have therefore
, and the large sieve inequality implies by positivity that
where the implied constant depends only on
. However by Lemma 3 of [DDS] (see p.
269, or compare [G,p.96,l.10] ) we have for
the lower bound
where
is the number of
-cycles in
, where the implied constant depend only on
. Thus we get by the Prime Number Theorem (Chebychev's elementary lower-bound estimate suffices) that
We choose
, so that
and
hence
with an implied constant depending only on
. By comparison with ( 4.1 ), we see that Proposition 4.2 is proved. □
Remark 4.5.
The bound obtained from the large sieve estimate may seem quite poor because of the choice of a rather small
, constrained by the smallest
. One may certainly expect that having a small Galois group would be of “codimension” at least
, which would mean essentially
. There is a similar discrepancy between what is proved and what is expected in other problems of probabilistic Galois theory.
Remark 4.6.
In contrast with the isomorphism case, the results above do not yield examples of “thinner” families of isogeny classes which would be ordinary and have the
as associated Galois group. Most notably, it is by no means clear how to prove the analogue of ( 1.2 ) where the isogeny classes are jacobians of curves of genus
(equivalently, where arbitrary Weil numbers are replaced by those associated with curves).
Distinguishing jacobians among abelian varieties over a finite field is a deep unsolved problem.
References
-
A. Chambert-Loir: Cohomologie cristalline: un survol, Exposition. Math. 16 (1998), 333–382.
-
N. Chavdarov: The generic irreducibility of the numerator of the zeta function in a family of curves with large monodromy, Duke Math. J. 87 (1997), 151–180.
-
S. Davis, W. Duke and X. Sun: Probabilistic Galois theory of reciprocal polynomials, Exposition. Math. 16 (1998), no. 4, 263–270.
-
S. DiPippo and E. Howe: Real polynomials with all roots on the unit circle and abelian varieties over finite fields, J. Number Theory 73 (1998), 426–450; Corrig., J. Number Theory 83 (2000), 182.
-
P.X. Gallagher: The large sieve and probabilistic Galois theory, in Proc. Sympos. Pure Math., Vol. XXIV, Amer. Math. Soc. (1973), 91–101.
-
E. Howe: Principally polarized ordinary abelian varieties over finite fields, Trans. Amer. Math. Soc. 347 (1995), 2361–2401.
-
M.N. Huxley: The large sieve inequality for algebraic number fields, Mathematika 15 (1968), 178–187.
-
N. Katz and P. Sarnak: Random matrices, Frobenius eigenvalues and monodromy, A.M.S Colloquium Publ. 45, 1999.
-
E. Kowalski: Some local-global applications of Kummer theory, manuscripta math. 111 (2003), 105–139.
-
E. Kowalski: The large sieve, monodromy and zeta functions of curves, preprint (2005), arXiv:math/NT0503714.
-
L. Miller: Curves with invertible Hasse-Witt matrix, Math. Annalen 197 (1972), 123–127.
-
D. Mumford: Bi-extensions of formal groups, in Algebraic Geometry (Internat. Colloq., Tata Inst. Fund. Res., Bombay, 1968), Oxford Univ. Press, 307–322.
-
P. Norman and F. Oort: Moduli of abelian varieties, Ann. of Math. (2) 112 (1980), no. 3, 413–439.
-
J-P. Serre: Quelques applications du théorème de densité de Chebotarev, Publ. Math. I.H.E.S 54 (1981), 123–201.
-
B.L. van der Waerden: Moderne algebra, vol I, Springer 1935.
-
W. Waterhouse: Abelian Varieties over Finite Fields, Ann. scient. Éc. Norm. Sup. 4ème série, 2 (1969), 521–560.
Universite Bordeaux I A2X, 351, cours de la Liberation, 33405 Talence Cedex, France E-mail address : emmanuel.kowalski@math.u-bordeaux1.fr