Lectures On Height Zeta Functions: at The Confluence of Algebraic Geometry, Algebraic Number Theory, and Analysis
Lectures On Height Zeta Functions: at The Confluence of Algebraic Geometry, Algebraic Number Theory, and Analysis
Chambert-Loir
Part II
Lectures on height zeta functions: At the
confluence of algebraic geometry, algebraic
number theory, and analysis
Antoine Chambert-Loir
1 Introduction
1.1 Diophantine equations and geometry
1.1.1 Diophantine equations
Broadly speaking, arithmetic is the study of diophantine equations, that is, systems of polynomial
equations with integral coefficients, with a special emphasis on their solutions in rational integers.
Of course, there are numerous variants, the most obvious ones allowing to consider coefficients
and solutions in the field of rational numbers, or in more general number fields, or even in more
general fields, e.g., finite fields.
The reader should be warned that, in this generality, we are constrained by the undecidability
theorem of [32]: there is no general method, that is no algorithm, to decide whether or not any given
polynomial system has solutions in rational integers. Any mathematician working on diophantine
equations is therefore obliged to consider specific types of diophantine equations, in the hope that
such an undecidability issues do not apply within the chosen families of equations.
17
A. Chambert-Loir
Since a ball is convex, the leading term is easily seen to equal the volume of this ball; in other
words,
Card{x ∈ Zn ; kxk 6 B} ∼ B n V1 ,
where V1 is the volume of the unit ball in Rn . The study of the error term, however, is of a much
more delicate nature. For n = 2, there is an easy O(B)-bound which only requires the Lipschitz
property of the boundary of the unit ball; when the norm is euclidean, one can prove a O(B 2/3 )
bound using the positivity of the curvature of the boundary; however, the conjectured O(B 1/2 )-
bound remains open.
1.2.2 Heights
Consider the case where F = Q, the field of rational numbers. Any point x ∈ Pn (Q) can be
represented by n + 1 rational numbers, not all zero; however, if we multiply these homogeneous
coordinates by a common denominator, we see that we may assume them to be integers; we then
may divide them by their greatest common divisor and obtain a system of homogeneous coordinates
[x0 : · · · : xn ] made of n + 1 coprime integers. At this point, only one choice is left to us, namely
multiplying this system by −1.
Consequently, we may define the exponential height of x as H(x) = max(|x0 | , . . . , |xn |) and its
logarithmic height as h(x) = log H(x). (Observe the notation, popularized by Serge Lang: small
“h” for logarithmic height, capital “H” for exponential height.)
2n
N (B) ∼ B n+1
ζ(n + 1)
18
A. Chambert-Loir
Observe that −
βX 6 +
βX and that, when X = −
Pn , βX = +
βX = n + 1.
19
A. Chambert-Loir
abstract variety X, the obtained formulas force us to take into account the embedding of X in a
projective space Pn .
There are also more subtle geometric and arithmetic caveats, related not only to X, but to its
closed subschemes; we will consider these later.
20
A. Chambert-Loir
2 Heights
2.1 Heights over number fields
2.1.1 Absolute values on the field of rational numbers
Let us recall that an absolute value |·| on a field F is a function from F to R+ satisfying the
following properties:
Of course, the usual modulus on the field of complex numbers satisfies these properties, hence
is an absolute value on C. It induces an absolute value on any of its subfields, in particular on Q.
It is called archimedean since for any a ∈ Q such that |a| 6= 0, and any T > 0, there exists an
integer n such that |na| > T . We write |·|∞ for this absolute value.
In fact, the field of rational numbers possesses many other absolute values, namely the p-adic
absolute value, where p is any prime number. It is defined as follows: Any nonzero rational
number a can be written as pm u/v, where u and v are integers not divisible by p and m is a
rational integer; the integer m depends only on a and we define |a|p = p−m ; we also set |0|p = 0.
Using uniqueness of factorization of integers into prime numbers, it is an exercise to prove that
|·|p is an absolute value on Q. In fact, not only does the triangular inequality hold, but a stronger
form is actually true: the ultrametric inequality:
• the standard archimedean absolute value |·|∞ and its powers |·|s∞ for 0 < s 6 1;
• the p-adic absolute value |·|p and its powers |·|sp for 0 < s < ∞ and some prime number p.
21
A. Chambert-Loir
since the right hand side does not depend on the choice of a specific system of homogeneous
coordinates. Indeed, if we replace xi by axi , for some nonzero element of F , the right hand side
gets multiplied by ∏
|a|v = 1.
v∈Val(F )
For F = Q, this definition coincides with the one previously given. Indeed, we may assume
that the coordinates xi are coprime integers. Then, for any prime number p, |xi |p 6 1 (since the
xi s are integers), and one of them is actually equal to 1 (since they are not all divisible by p).
Consequently, the p-adic factor is equal to 1 and
Using a bit more of algebraic number theory, one can show that for any finite extension F 0 of F ,
and any x ∈ Pn (F ),
HF 0 (x) = HF (x)d , where d = [F 0 : F ].
22
A. Chambert-Loir
(f ε) ⊗ η = ε ⊗ (f η) = f (ε ⊗ η),
23
A. Chambert-Loir
Any line bundle L has an “inverse” L−1 for the tensor product; if ε is a frame of L on U , then
a frame of L−1 on U is ε−1 , again with the obvious compatibilities
(f ε)−1 = f −1 ε−1
2.2.4 Functoriality
Finally, if u : X → Y is a morphism of algebraic varieties and L is a line bundle on Y , there is
a line bundle u∗ L on X defined in such a way that the fibre of u∗ L at a point x ∈ X is the line
L(u(x)); similarly, if ε is a frame of L on an open set U of Y , then u∗ ε is a frame of u∗ L on u−1 (U ).
At the level of isomorphism classes, this induces a map u∗ : Pic(Y ) → Pic(X) which is a mor-
phism of Abelian groups.
where the hat indicates that one omits the corresponding factor.
On the intersections Ui ∩ Uj , one can check that the expressions si (x)⊗n+1 ωi and sj (x)⊗n+1 ωj
identify the one to the other if one uses the relation xi sj (x) = xj si (x) that we observed. This
implies that these expressions can be glued together as a section of OPn (n+1)⊗ωPn which vanishes
nowhere, thereby establishing the announced isomorphism.
24
A. Chambert-Loir
25
A. Chambert-Loir
Recalling that si is the section of OPn (1) corresponding to the ith homogeneous coordinate, we
define
|xi |v
ksi (x)kv = .
max(|x0 |v , . . . , |xn |v )
We observe that the right hand side does not depend on the choice of homogeneous coordinates.
Moreover, the product formula implies that
∏ ∏
HF (x) = ksi (x)k−1
v |xi |v = ksi (x)k−1
v .
v∈Val(F ) v∈Val(F )
In other words, we have given a formula for HF (x) as a product over Val(F ) where each factor is
well defined, independently of any choice of homogeneous coordinates.
The reader should not rush to conclusions: so far we have only exchanged the indeterminacy of
homogeneous coordinates with the choice of a specific index i, more precisely of a specific section si .
26
A. Chambert-Loir
By definition of an adelic metric, almost all of the terms are equal to 1. If t is any nonzero element
of L(x), there exists a ∈ F ∗ such that t = as(x); then ktkv = |a|v ks(x)kv for any v ∈ Val(F ). In
particular, ktkv = 1 for almost all v ∈ Val(F ) and
∏ ∏ ∏
ktk−1
v = kas(x)k−1
v = |a|−1 −1
v ks(x)kv
v∈Val(F ) v∈Val(F ) v∈Val(F )
∏ ∏ ∏
= |a|−1
v
ks(x)k−1
v
= ks(x)k−1
v = HL,s (x)
v∈Val(F ) v∈Val(F ) v∈Val(F )
27
A. Chambert-Loir
An isometry of adelically metrized line bundles is an isomorphism which preserves the metrics.
The set of isometry classes of metrized line bundles form a group, called the Arakelov-Picard group,
c
and denoted Pic(X). c
The above formula implies that the function from X(F ) × Pic(X) to R∗+ is
linear in the second variable.
Let us moreover assume that X is projective. A v-adic metric on the trivial line bundle is
characterized by the norm of its section 1, which is a nonvanishing continuous function ρv on X(Fv ).
Consequently, an adelic metric ρ on OX is given by a family (ρv ) of such functions which are
almost all equal to 1. (The projectivity assumption on X allows to write X as a finite union of
open subsets U such that X(Fv ) is the union of the sets U (ov ).) One has
∏
Hρ (x) = ρv (x)−1
v∈Val(F )
2.4.7 Functoriality
If f : X → Y is a morphism of algebraic varieties and L is a adelically metrized line bundle on Y ,
the line bundle f ∗ L on X has a natural adelic metric: indeed, the fibre of f ∗ L at a point x is
the line L(f (x)), for which we use its given norm. This construction is compatible with isometry
c ) → Pic(X).
classes and defines a morphism of Abelian groups f ∗ : Pic(Y c At the level of heights,
it implies the equality
HL (f (x)) = Hf ∗ L (x)
for any x ∈ X(F ).
28
A. Chambert-Loir
where the supremum runs over all couples (i, j) such that xi and xj are not both equal to 0. This
reduces us to proving the finitess assertion on P1 .
We also may assume that F is a Galois extension of Q and let G = {σ1 , . . . , σd } be its Galois
group. Let x ∈ F and let us consider the polynomial
∏
Px = (T − σ(x)) = T d + a1 T d−1 + · · · + ad ;
σ∈G
where the notation w | p means that the product ranges over all normalized absolute values w
on F which are equivalent to the p-adic absolute value. It follows that h([1 : σ1 (x) : · · · : σd (x)) 6
dh([1 : x]).
d
∏d which associates to a point [u0 : · · · : ud ] ∈ P the coefficients [v0 : · · · : vd ]
Moreover, the map
of the polynomial i=1 (u0 T − ui ) has degree d. This implies that
3 Manin’s problem
3.1 Counting functions and zeta functions
3.1.1 The counting problem
Let X be a variety over a number field F and let L be an ample line bundle with an adelic metric.
We have seen that for any real number B, the set of points x ∈ X(F ) such that HL (x) 6 B is
a finite set. Let NX (L; B) be its cardinality. We are interested in the asymptotic behaviour of
NX (L; B), when B grows to infinity. We are also interested in understanding the dependence on L.
29
A. Chambert-Loir
for any complex number s such that the series converges absolutely. In principle, and we will
eventually do so, one can omit the ampleness condition in that definition, but then the convergence
in some half-plane is not assured (and might actually fail).
In some cases, one can even establish terms of lower order, or prove explicit error terms.
However, these zeta functions are not as well behaved as the ones traditionally studied in alge-
braic number theory. Although some of them have a meromorphic extension to the whole complex
plane, like those of flag varieties (see §4.3.2 below, see also [23]), it happens quite often, for ex-
ample, that they have a natural boundary ; this is already the case for some toric surfaces, see [2,
Example 3.5.4]. Subtler analytic properties of the height zeta functions beyond the largest pole is
the subject of some recent investigations, see, e.g., [9].
30
A. Chambert-Loir
3. genus g > 2, ωX ample. By Mordell’s conjecture, first proved by [24], X(F ) is a finite
set.
The counting function distinguishes very clearly the first two cases. When X is the projective
line, NX (B) grows like a polynomial in B (depending on how the height is defined), and for X an
elliptic curve, NX (B) ≈ (log B)r/2 , where r is the rank of the Abelian group X(F ).
31
A. Chambert-Loir
As a consequence, the behaviour of the counting function can only be expected to reflect the
global geometry if one doesn’t count points in accumulating subvarieties. This leads to the notation
NU (L; B) and ZU (L; s), where U is any Zariski open subset in X.
To get examples of such behaviour, it suffices to blow-up a variety at one (smooth) rational
point P . It replaces the point P by a projective space E of dimension n − 1 if X has dimension n
which is likely to possess ≈ B a points of height 6 B, while the rest could be smaller. For example,
if X is an Abelian variety (generalization of elliptic curves in higher dimension), NX (L; B) only
grows like a power of log B.
32
A. Chambert-Loir
NU (L; B) ≈ B αX (L) (log B)t−1 . A stronger form of the conjecture of Batyrev–Manin also pre-
−1
dicts the order of this pole, at least when L = ωX . In that case, Batyrev and Manin conjecture
that t is the dimension of the real vector space Pic(X)R . (Recall that the Picard group of a Fano
variety has finite rank.)
3.3.2 (In)compatibilities
Some basic checks concerning that conjecture had been made by [25], for instance its compatibility
with products of varieties. However, [4] observed that the conjecture is not compatible with families
and produced a counterexample.
They consider the subscheme V of P3 × P3 defined by the equation x0 y03 + · · · + x3 y33 = 0 (where
[x0 : · · · : x3 ] are the homogeneous coordinates on the first factor, and [y0 : · · · : y3 ] are those on
the second). For fixed x ∈ P3 , the fibre Vx consisting of y ∈ P3 such that (x, y) ∈ V is a diagonal
cubic surface; in other words, V is the total space of the family of diagonal cubic surfaces.
Let p1 and p2 be the two projections from P3 × P3 to P3 . For any couple of integers (a, b), let
O(a, b) be the line bundle p∗1 OP3 (a)⊗p∗2 OP3 (b) on P3 ×P3 . The so-defined map from Z2 to Pic(P3 ×
P3 ) is an isomorphism of Abelian groups. Moreover, since V is an hypersurface in P3 × P3 , the
Lefschetz theorem implies that the restriction map induces an isomorphism from Pic(P3 × P3 )
to Pic(V ). The anticanonical line bundle ωV−1 of V corresponds to O(3, 1) (anticanonical of P3 ×P3
minus the class of the equation, that is (4, 4) − (1, 3)) and the conjecture of Batyrev–Manin
predicts that there should be ≈ B(log B) points of O(3, 1)-height 6 B in V (F ).
Let us fix some x ∈ P3 such that Vx is non-singular. This is a cubic surface, and its anticanonical
line bundle identifies with the restriction of OP3 (1). Thus, if tx = dimR Pic(Vx )R , the conjecture
predicts that there should be ≈ B(log B)tx −1 points of O(1)-height 6 B in Vx (F ).
As a consequence, if the conjecture holds for V and if tx > 2, then the fibre Vx would have
more points than what the total space seem to have. This doesn’t quite contradict the conjecture
however, because it explicitly allows to remove a finite union of subvarieties.
However, the rank of the Picard group can exhibit disordered behaviour in families; for example,
it may not be semi-continuous, and jump on a infinite union of subvarieties. This happens here,
since tx = 7 when F contains the cubic roots of unity and all the homogeneous coordinates of x
are cubes. The truth of the conjecture for V therefore requires to omit all such fibres Vx . But they
form an infinite union of disjoint subvarieties, a kind of accumulating subset which is not predicted
by the conjecture of Batyrev–Manin.
In particular, this conjecture is false, either for V , or for most cubic surfaces. By geometric
considerations, and using their previous work on toric varieties, [4] could in fact conclude that the
conjecture does not hold for V .
33
A. Chambert-Loir
3. a rational number related to the location of the anticanonical line bundle in the effective
cone Λeff of Pic(X)R ;
Let us give a short account of Peyre’s construction of a measure. The idea is to use the given
−1
adelic metric on ωX to construct a measure τX,v on the local spaces X(Fv ) for all places v of the
number field F , and then to consider a suitable (renormalized) product of these measures.
|α|v |f (x)|v
= dx1 . . . dxn
kαkv kαkv
is well-defined, independently of the choice of α. One can therefore glue all these local measures
and obtain a measure τX,v on X(Fv ).
converges absolutely. This convergence never holds, and one therefore needs to introduce conver-
gence factors to define a measure on X(AF ).
By a formula of [43], the equality τX,v (X(Fv )) = qv− dim X Card(X(kv )) holds for almost all finite
places v, where kv is the residue field of F at v, qv is its cardinality, and X(kv ) is the set of solutions
in kv of a fixed system of equations with coefficients in the ring of integers of F which defines X.
By Weil’s conjecture, established by [16], plus various cohomological computations, this implies
that
1
τX,v (X(Fv )) = 1 + Tr(Frobv | Pic(XF )R ) + O(qv−3/2 ),
qv
where Frobv is a “geometric Frobenius element” of the Galois group of F at the place v.
Let us consider the Artin L-function of the Galois-module P = Pic(XF )R : it is defined as the
infinite product
∏
L(s, P ) = Lv (s, P ), Lv (s, P ) = det(1 − qv−s Frobv |P )−1 .
v finite
34
A. Chambert-Loir
The product converges absolutely for Re(s) > 1, defines a holomorphic function in that domain.
Moreover, L(s, P ) has a meromorphic continuation to C with a pole of order t = dim Pic(X)R at
s = 1. Let
L∗ (1, P ) = lim L(s, P )(s − 1)−t ;
s→1
it is a positive real number.
Comparing the asymptotic expansions for Lv (1, P ) and τX,v (X(Fv )), one can conclude that the
following infinite product ∏( )
τX = L∗ (1, P ) Lv (1, P )−1 τX,v
v
3.3.6 Equidistribution
One important aspect of the language of metrized line bundles is that it suggests explicitly to look
at what happens when one changes the adelic metric on the canonical line bundle.
Assume for example that there is a number field F and an open subset U such that Manin’s con-
jecture (with precised constant as above) holds for any adelic metric on ωX . Then, [34] shows that
b
rational points in U (F ) of heights 6 B equidistribute towards the probability measure on X(AF )BM
proportional to τX . Namely, for any smooth function Φ on X(AF ),
∑ ∫
1 1
−1 Φ(x) −→ b
Φ(x) dτX (x).
NU (ωX ; B) x∈U (F ) τX (X(AF )BM b
) X(AF )BM
H (x)6B
ω −1
X
As we shall see in the next section, this strengthening by Peyre of the conjecture of Batyrev–
Manin is true in many remarkable cases, with quite nontrivial proofs.
In this formula, ζF is the Dedekind zeta function of the field F , r1 and 2r2 are the number of real
embeddings and complex embeddings of F , and DF is its discriminant.
In fact, it has been shown later by many authors, see e.g. [25], that the height zeta function
ZPn (L; s) is holomorphic for Re(s) > n + 1, has a meromorphic continuation to the whole complex
plane C, with a pole of order 1 at the point s = n+1 and no other pole on the line {Re(s) = n+1}.
To prove this estimate, one can sort points [x0 : · · · : xn ] in Pn (F ) according to the class of
the fractional ideal generated by (x0 , . . . , xn ). Constructing fundamental domains for the action of
units, the enumeration of such sets can be reduced to the counting of lattice points in homothetic
sets in R(n+1)(r1 +2r2 ) whose boundary has smaller dimension. The Möbius inversion formula is
finally used to take care of the coprimality condition.
35
A. Chambert-Loir
• Del Pezzo surfaces of degree 5, given as blow-ups of the plane in 4 points in general position
([8]);
This method suffers from an obvious drawback: 1) it is hard to take advantage of subtle geometric
properties of the situation studied; 2) it requires to know already much about the rational points,
e.g., to be able to parametrize them. However, this is so far the only way of dealing with the
(eventually singular) Del Pezzo surfaces which are not equivariant compactifications of the torus
Gm 2 or of the additive group Ga 2 .
To deal with the parametrization of the rational points, an essential tool is the universal torsor,
invented by [14], which is a quasi-projective variety lying over the given Del Pezzo surface. The
computation of a universal torsor is essentially equivalent to that of a multi-graded ring defined
in [15], the so-called Cox ring of the variety. In the case of the projective space Pn , the torsor
is An+1 \ {0}, mapping to Pn via (x0 , . . . , xn ) 7→ [x0 : · · · : xn ], and the Cox ring is the graded
polynomial ring in n + 1 variables.
In the context of Manin’s conjecture, universal torsors were first introduced by [35] for toric
varieties, in which case they are isomorphic to an open subset of an affine space. The parametriza-
tion of rational points which they allow to write proved to be very useful for the study of points
of bounded height, at least in contexts where other techniques, like harmonic analysis, do not
apply. Moreover, their computation is also related to the precise understanding of the numerically
effective cone in the Picard group.
[18] computes these universal torsors for smooth Del Pezzo surfaces (that is, blow-ups of the
projective plane in at most 8 points in general position). They were also computed for many
singular Del Pezzo surfaces, see, e.g., the article of [20] for an example leading to a case of Manin’s
conjecture. In that paper, the interested reader shall also find a list of known cases, mostly due to
Browning, de la Bretéche and Derenthal, with references to the articles; see also the more
recent preprint of [19]. Anyway, we apologize not to develop this important part of the subject
here, which would lead us to far of your topic.
36
A. Chambert-Loir
The circle method is restricted to cases where the number of variables n is very large in compar-
ison to the degrees d1 , . . . , dm . For instance, [5] establishes the desired result for an hypersurface
of degree d in Pn provided n > 2d (d − 1). However, when it applies, it gives very strong results,
including a form of the Hasse principle and equidistribution, see [34].
• The set VP of matrices with fixed characteristic polynomial P . Indeed, when the polyno-
mial P has distinct roots, any two such matrices are conjugate, hence VP is an homogeneous
space of GLn .
Various techniques allow to tackle the question of rational points of bounded height on such
varieties. These methods allow to derive asymptotic expansions for the counting function by
comparing it to the volume of a corresponding set in a real or adelic space, whose asymptotic
growth has to be established independently. They are also amenable to proving equidistribution
results, as well as establishing analyting properties of the height zeta function.
37
A. Chambert-Loir
considering a basis of W0 and extending it to a basis of V , it is easy to observe that any other
subspace of dimension d is of the form gW0 , for some automorphism g ∈ GL(V ); moreover,
gW0 = g 0 W0 if and only if g −1 g 0 W0 = W0 , that is, if g −1 g 0 belongs to the stabilizer P of W0 for
the action of GL(V ). In the basis that we fixed, P is the set of invertible upper-triangular block
matrices having only zeroes in the lower rectangle [d + 1, . . . , n] × [1, . . . , d].
More generally, in the language of algebraic groups, generalized flag varieties are quotients of
reductive groups by parabolic subgroups. The understanding of their height zeta functions is a
consequence of the observation, due to J. Franke, that these functions were studied extensively
in the theory of automorphic forms, being nothing else than generalized Eisenstein series. Modulo
some (not so obvious) translation, the results of R. P. Langlands on these Eisenstein series readily
imply the conjectures of Batyrev–Manin and Peyre (including equidistribution) for the flag
varieties, see [25] and [34].
If we use the right hand side of this formula to extend HL to a function on G(AF ), we obtain
that the height zeta function ZG (L; s) is an average over the discrete group G(F ) of a function
38
A. Chambert-Loir
HL−s on G(AF ) and one can use harmonic analysis and the spectral decomposition to under-
stand ZG (L; s).
Important cases of groups in which mathematicians have been able to use harmonic analysis to
prove the conjectures of Batyrev–Manin–Peyre are the following.
be the orthogonal of G(F ), that is the group of characters of G(AF ) which are trivial on G(F ).
This is a locally compact group and it carries a dual measure µ⊥ .
Any integrable function f ∈ L1 (G(AF )) has a Fourier transform, which is the function on G(A \ F)
\
defined as follows: for χ ∈ G(A F ),
∫
F (f ; χ) = f (g)χ(g) dµ(g).
G(AF )
∏ ∏
If f is a simple function, i.e., of the form f (g) = fv (gv ), then F (f ; χ) is a product Fv (fv ; χv )
of local Fourier transforms defined in an analogous manner, namely
∫
Fv (fv ; χv ) = fv (g)χv (g) dµv (g).
G(Fv )
The generalization of Poisson formula to this context states that if f is an integrable function
on G(AF ) such that moreover F (f ; ·) is integrable on G(F )⊥ , then
∑ ∫
f (g) = F (f ; χ) dµ⊥ (χ).
g∈G(F ) G(F )⊥
which in many cases proved to be a key tool towards establishing the desired analytic properties
of the height zeta function.
39
A. Chambert-Loir
The decomposition of the function HL as a product∏of function on the spaces G(Fv ) implies a
similar decomposition of Φ as a the product Φ(s) = v Φv (s), where
∫
Φv (s) = Fv (HL−s , 1) = HL (g)−s dµv (g).
G(Fv )
In the above mentioned cases, these functions have been often computed explicitly, resorting to
properties of algebraic groups. However, this can also be done in a very geometric manner thanks
to the fundamental observation: the integral Φv (s) on G(Fv ) can be viewed as a geometric analogue
−1
on X(Fv ) of Igusa’s local zeta functions. Assume for simplicity that L = ωX . In Section 5 we
will explain how to prove the following facts (see § 5.2.6):
• for almost all places v, Φv (s) can be explicitly computed (Denef’s formula);
∏
• the product Φv (s) converges for Re(s) > 1;
• this product has a meromorphic continuation which is governed by some Artin L-function.
We shall in fact explain that these facts apply in much more generality than that of algebraic
groups.
40
A. Chambert-Loir
canonical global section fL whose divisor is L. For any place v ∈ Val(F ), the function x 7→ kfL (x)kv
is continuous on X(Fv ) and vanishes precisely on L. Moreover, its sup-norm is equal to 1 for almost
all places v. ∏
Consequently, the infinite product v∈Val(F ) kfL (xv )kv converges to an element of R+ for any
adelic point x = (xv ) ∈ X(AF ). Let U = X \ L. If x is an adelic point of U , then kfL (xv )kv 6= 0 for
all v, and is in fact equal to 1 for almost all v. We thus can define the height of a point x ∈ U (AF )
to be equal to ∏
HL ((xv )) = kfL (xv )k−1
v .
v∈Val(F )
The resulting function on U (AF ) is continuous and admits a positive lower bound. Moreover, for
any B > 0, the set of points x ∈ U (AF ) such that HL (x) 6 B is compact.
converges absolutely. The limit will thus be a positive real number (unless some factor U (ov ) is
empty, which does not happen if we remove an adequate finite set of places in the product). The
existence of such factors is a mere triviality, since one could take for λv the inverse of τ(X,L),v (U (ov )).
Of course, we claim for a meaningful definition of a family (λv ), based on geometric or arithmetic
invariants of U .
The condition under which we can produce such factors is the following: X is proper, smooth,
geometrically connected, and the two cohomology groups H1 (X, OX ) and H2 (X, OX ) vanish. Let
us assume that this holds and let us define
∗
M 0 = H0 (UF , O ∗ )/F and M 1 = H1 (UF , O ∗ )/torsion;
in words, M 0 is the Abelian group of invertible functions on UF modulo constants, and M 1 is the
Picard group of UF modulo torsion. By the very definition as a cohomology group of UF , they
possess a canonical action of the Galois group ΓF of F /F . Moreover, M 0 is a free Z-module of finite
rank. Indeed, to an invertible function on UF , one may attach its divisor, which is supported by
the complementary subset, that is LF . This gives a morphism from M 0 to the free Abelian group
41
A. Chambert-Loir
generated by the irreducible components of LF and this map is injective because X is normal. It
follows that M 0 is a free Z-module of finite rank. Moreover, under the vanishing assertion above,
then M 1 is a free Z-module of finite rank too. We thus can consider the Artin L-functions of these
two ΓF -modules, L(s, M 0 ) and L(s, M 1 ).
Using Weil’s conjectures, proved by [16], to estimate the volumes U (ov ), in a similar manner to
what Peyre had done to define the global measure τX on X(AF ), we prove that the family (λv )
given by
Lv (1, M 0 )
λv =
Lv (1, M 1 )
if v is a finite place of F and λv = 1 if v is archimedean is a family of convergence factors. In other
words, the infinite product of measures
L∗ (1, M 1 ) ∏ Lv (1, M 0 )
dτ (xv )
L∗ (1, M 0 ) Lv (1, M 1 ) (X,L),v
v∈Val(F )
defines a Radon measure on the locally compact space U (AF ). We denote this measure by τ(X,L)
and call it the Tamagawa measure on U (AF ).
42
A. Chambert-Loir
(This function was denoted Φ in §4.4.) By definition, the measure τ(X,L) is (up to the convergence
factors and the global factor coming from Artin L-functions) the product of the local measures
τ(X,L),v on the analytic manifolds U (Fv ). Similarly, the integrated function x 7→ HL (x)−s is the
product of the local functions x 7→ kfL (x)ksv . Consequently, provided it converges absolutely, this
adelic integral decomposes as a product of integrals over local fields,
∫
L∗ (1, M 1 ) ∏ Lv (1, M 0 )
Z(s) = ∗ Z v (s), Z v (s) = kfL (x)ksv dτ(X,L),v (x).
L (1, M 0 ) Lv (1, M 1 ) U (Fv )
v∈Val(F )
where f is a polynomial and Φ is a smooth function with compact support on the affine n-space
over Fv . (See [30] for a very good survey on this theory.)
Our integrals Zv (s) are straightforward generalizations of Igusa’s local zeta functions. Indeed,
since τ(X,L),v = kfL k−1
v τX,v , we have
∫ ∫
s−1
Zv (s) = kfL (x)kv dτX,v (x) = kfL (x)ks−1
v dτX,v (x)
U (Fv ) X(Fv )
because L(Fv ) has measure 0 in X(Fv ). In our case, we integrate on a compact analytic manifold,
rather than a test function on the affine space; as always in Differential geometry, partitions of
unity and local charts convert a global integral into a finite sum of local ones. The absolute value
of a polynomial has been replaced by the norm of a global section of a line bundle; again, in
local coordinates, the function kfL kv is of the form |f |v ϕ, where f is a local equation of L and
ϕ a continuous non-vanishing function. However, because of these slight differences, we call such
integrals geometric Igusa zeta functions.
From now on, we will assume in this text that over the algebraic closure F , L is a divisor with
simple normal crossings. This means that the irreducible components of the divisor LF are smooth
and meet transversally; in particular, any intersection of part of these irreducible components is
either empty, or a smooth subvariety of the expected codimension. This assumption is only here for
the convenience of the computation. Indeed, by the theorem on resolution of singularities of [29],
there exists a proper birational morphism π : Y → X, with Y smooth, which is an isomorphism
on U and such that the inverse image of L (as a Cartier divisor) satisfies this assumption. We
may replace (X, L) by (Y, π ∗ L) without altering the definitions of Z, Zv , etc.. Let A be the set
of irreducible components of L; for α∑∈ A , let Lα be the corresponding component, and dα its
multiplicity in L. We thus have L = α∈A dα Lα .
This induces a stratification of X indexed by subsets A ⊂ A , the stratum XA being given by
XA = {x ∈ X ; x ∈ Lα ⇔ α ∈ A}.
In the sequel, we shall moreover assume that for any α ∈ A , Lα is geometrically irreducible.
This does not cover all cases, but the general case can be treated using a similar analysis, the
Dedekind zeta function of F being replaced by Dedekind zeta functions of finite extensions Fα
defined by the components Lα ; I refer the interested reader to [13] for details.
43
A. Chambert-Loir
As in [17], this is a consequence of the fact that for each point x̃ ∈ X(kv ), in the local analytic
description of the previous paragraph, we may find local coordinates xα and y such that ϕ ≡ ω = 1
which parametrize the open unit polydisk with “center x̃.” That this is at all possible is more or
less what is meant by “having good reduction”.
44
A. Chambert-Loir
More precisely, since Lα is geometrically irreducible and Xα is an open subset of it, it follows from
[16] that
qv− dim X Card(Xα (kv )) = qv−1 + O(qv−3/2 ).
(In fact, the estimates of [31] suffice for that.) Similarly, X∅ = X \ L = U , and
qv− dim X Card(X∅ (kv )) = 1 + O(qv−1/2 ).
In fact, the vanishing assumption on the cohomology groups H1 (X, OX ) and H2 (X, OX ) implies
that the remainder is even O(qv−1 ). By an integration formula of [43], this expression is precisely
equal to τ(X,L),v (U (ov )).
We therefore obtain that for any ε > 0, there exists δ > 0 such that if Re(s − 1) > −δ, then
∏ ( )
Zv (s) = τ(X,L),v (U (ov )) (1 − qv−1−dα (s−1) )−1 1 + O(qv−1−ε ) .
α∈A
If we multiply this estimate by the convergence factors λv that we have introduced, and which
satisfy λv = 1 + O(qv−1 ), we obtain that the infinite product
∏ ∏
λv Zv (s) ζF,v (1 + dα (s − 1))−1
v finite α∈A
converges absolutely and uniformly for Re(s − 1) > −δ/2; it defines a holomorphic function Φ(s)
on that half-plane. We have denoted by ζF,v the local factor at v of the Dedekind zeta function of
the number field F . Then, the equality
L∗ (1, M 1 ) ∏
Z(s) = λv Zv (s)
L∗ (1, M 0 )
v∈Val(F )
( )
L∗ (1, M 1 ) ∏ ∏
= ∗ ζF (1 + dα (s − 1)) Φ(s) Zv (s)
L (1, M 0 )
α∈A v archimedean
shows that Z converges absolutely, and defines a holomorphic function on {Re(s) > 1}. Since the
Dedekind zeta function ζF has a pole of order 1 at s = 1, Z(s) admits a meromorphic continuation
on the half-plane {Re(s) > 1 − δ/2}, whose only pole is at s = 1, with multiplicity Card(A ).
Comparing the Galois modules M 0 and M 1 for X and U , one can conclude that
∏
lim (s − 1)Card(A ) Z(s) = τX (X(AF )) d−1
α . (5.1)
s→1
α∈A
45
A. Chambert-Loir
converges for Re(s) > 1, defines a holomorphic function on this half-plane. It has a meromorphic
continuation on some half-plane {Re(s) > 1 − δ} (for some δ > 0) whose only pole is at s = 1, has
order Card(A ) and its asymptotic behaviour at s = 1 is given by Equation (5.1).
By a slight generalization of Ikehara’s Tauberian theorem, we conclude that V (B) satisfies the
following asymptotic expansion:
( ∏ )−1
V (B) ∼ (a − 1)! dα τX (X(AF ))B(log B)a−1 ,
α∈A
on X(AF ).
Acknowledgments
I would like to thank Yuri Tschinkel for his comments on a first version of this survey. Our
collaboration on the topics presented here began more than ten years ago. Let it be the occasion
to thank him heartily for having shared his views and projects with me.
I also thank the referee, as well as R. de la Bretéche, H. Oh and W. Veys for their suggestions.
46
A. Chambert-Loir
References
[1] V. V. Batyrev & Yu. I. Manin (1990), Sur le nombre de points rationnels de hauteur bornée
des variétés algébriques. Math. Ann., 286, 27–43.
[2] V. V. Batyrev & Yu. Tschinkel (1995), Rational points on bounded height on compactifi-
cations of anisotropic tori. Internat. Math. Res. Notices, 12, 591–635.
[3] V. V. Batyrev & Yu. Tschinkel (1996), Height zeta functions of toric varieties. Journal
Math. Sciences, 82 (1), 3220–3239.
[4] V. V. Batyrev & Yu. Tschinkel (1996), Rational points on some Fano cubic bundles. C.
R. Acad. Sci. Paris Sér. I Math., 323, 41–46.
[5] B. J. Birch (1962), Forms in many variables. Proc. London Math. Soc., 265A, 245–263.
[6] E. Bombieri & J. PILA (1989), The number of integral points on arcs and ovals. Duke Math.
J., 59 (2), 337–357.
[7] M. Borovoi & Z. Rudnick (1995), Hardy-Littlewood varieties and semisimple groups. Invent.
Math., 119 (1), 37–66.
[8] R. de la Bretèche (2002), Nombre de points de hauteur bornée sur les surfaces de del Pezzo
de degré 5. Duke Math. J., 113 (3), 421–464.
[9] R. De la Bretèche & P. Swinnerton-Dyer (2007), Fonction zêta des hauteurs associée
à une certaine surface cubique. Bull. Soc. Math. France, 135 (1), 65–92.
[10] T. D. Browning & D. R. Heath-Brown (2006), The density of rational points on non-
singular hypersurfaces. II. Proc. London Math. Soc. (3), 93 (2), 273–303. With an appendix
by J. M. Starr.
[12] A. Chambert-Loir & Yu. Tschinkel (2002), On the distribution of points of bounded height
on equivariant compactifications of vector groups. Invent. Math., 148, 421–452.
[13] A. Chambert-Loir & Yu. Tschinkel (2008), Tamagawa measures and volume estimates in
analytic and adelic geometry. Work in progress.
[14] J.-L. Colliot-Thélène & J.-J. Sansuc (1987), La descente sur les variétés rationnelles. II.
Duke Math. J., 54 (2), 375–492.
[15] D. A. Cox (1995), The homogeneous coordinate ring of a toric variety. J. Algebraic Geometry,
4 (1), 17–50.
[16] P. Deligne (1974), La conjecture de Weil, I. Publ. Math. Inst. Hautes Études Sci., 43, 273–
307.
[17] J. Denef (1987), On the degree of Igusa’s local zeta function. Amer. J. Math., 109, 991–1008.
[18] U. Derenthal (2006), On the Cox ring of Del Pezzo surfaces. ArXiv:math/060311
[math.AG].
47
A. Chambert-Loir
[20] U. Derenthal & Y Tschinkel Universal torsors over del Pezzo surfaces and rational points.
In Equidistribution in number theory, an introduction, volume 237 de NATO Sci. Ser. IIMath.
Phys. Chem., p. 169-196, Springer, Dordrecht.
[21] A. Eskin & C. McMullen (1993), Mixing, counting and equidistribution on Lie groups. Duke
Math. J., 71 (1), 181–209.
[22] A. Eskin, S. Mozes & N. Shah (1996), Unipotent flows and counting lattice points on ho-
mogeneous varieties. Ann. of Math., 143 (2), 253–299.
[23] D. Essouabri (2005), Analytic continuation of a class of zeta functions of heights. (Pro-
longements analytiques d’une classe de fonctions zêta des hauteurs et applications.). Bull.
Soc. Math. France, 133 (2), 297–329.
[24] G. Faltings (1983), Endlichkeitsätze für abelsche Varietäten über Zahlkörpern. Invent. Math.,
73 (3), 349–366.
[25] J. Franke, Yu. I. Manin & Yu. Tschinkel (1989), Rational points of bounded height on
Fano varieties. Invent. Math., 95 (2), 421–435.
[26] A. Gorodnik, F. Maucourant & H. Oh (2009), Manin’s and Peyre’s conjectures on rational
points and adelic mixing. Ann. Sci. École Norm. Sup. To appear.
[27] G. H. Hardy & E. M. Wright (1979), An introduction to the theory of numbers, The
Clarendon Press Oxford University Press, New York, fifth edition.
[28] D. R. Heayh-Brown (2002),The density of rational points on curves and surfaces, Ann. of
Math. (2), 155 (2), 553-595.
[29] H. Hironaka (1964), Resolution of singularities of an algebraic variety over a field of char-
acteristic zero. I, II. Ann. of Math. (2), 79, 109–203, 205–326.
[30] J.-I. Igusa (2000), An introduction to the theory of local zeta functions, AMS/IP Studies in
Advanced Mathematics 14, American Mathematical Society, Providence, RI.
[31] S. Lang & A. Weil (1954), Number of points of varieties in finite fields. Amer. J. Math., 76,
819–827.
[33] D. G. Northcott (1950), Periodic points on an algebraic variety. Ann. of Math., 51, 167–
177.
[34] E. Peyre (1995), Hauteurs et mesures de Tamagawa sur les variétés de Fano. Duke Math.
J., 79, 101–218.
[35] P. Salberger (1998), Tamagawa measures on universal torsors and points of bounded
height on Fano varieties. Nombre et répartition des points de hauteur bornée, E. Peyre,
Astérisque 251, 91–258.
48
A. Chambert-Loir
[36] P. Salberger (2007), On the density of rational and integral points on algebraic varieties. J.
Reine Angew.Math., 606, 123–147.
[37] S. Schanuel (1979), Heights in number fields. Bull. Soc. Math. France, 107, 433–449.
[39] J.-P. Serre (1997), Lectures on the Mordell-Weil theorem, Aspects of Mathematics, Friedr.
Vieweg & Sohn, Braunschweig, third . Translated from the French and edited by Martin
Brown from notes by Michel Waldschmidt, With a foreword by Brown and Serre.
[41] J. A. Shalika & Y. Tschinkel (2004), Height zeta functions of equivariant compactifications
of the Heisenberg group. Contributions to automorphic forms, geometry, and number theory,
743–771, Johns Hopkins Univ. Press, Baltimore, MD.
[42] J. Thunder (1992), An asymptotic estimate for heights of algebraic subspaces. Trans. Amer.
Math. Soc., 331, 395–424.
[43] A. Weil (1982), Adeles and algebraic groups, Progr. Math. 23, Birkhäuser.
49