Autour de la méthode globale de déterminant
On the global determinant method
Résumé
In this paper, we build the global determinant method of Salberger by Arakelov geometry explicitly. As an application, we study the dependence on the degree of the number of rational points of bounded height in plane curves. We will also explain why some constants will be more explicit if we admit the Generalized Riemann Hypothesis.
Dans cet article, on construit la méthode globale de déterminant de Salberger par la géométrie d’Arakelov explicitement. Comme une application, on étudie la dépendance du degré du nombre de points rationnels de hauteur majorée dans courbes planes. On expliquera aussi pourquoi certaines constantes seront plus explicites si on admet l’hypothèse généralisée de Riemann.
1 Introduction
Let be a projective variety over a number field . For every rational point , we denote by the height (see (1) for the definition) of with respect to the above closed immersion, for example, the classic Weil height (cf. [25, §B.2, Definition]). Let
where and the embedding morphism is omitted. By the Northcott’s property, the cardinality is finite for a fixed .
In order to understand the density of the rational points of , it is an important approach to study the function with the variable . For different required properties of , numerous methods have been applied. In this article, we are interested in the uniform upper bound of for all with fixed degree and dimension, and for those satisfying certain common conditions.
1.1 Determinant mathod
In order to understand the function of the variable , we will introduce the so-called determinant method to study the number of rational points with bounded height in arithmetic varieties, which was proposed in [24].
1.1.1 Basic ideas and history
Traditionally, the determinant method is proposed over the rational number field to avoid some extra technical troubles. In [3] (see also [35]), Bombieri and Pila proposed a method of determinant argument to study plane affine curves. The monomials of a certain degree evaluated on a family of rational points in having the same reduction modulo some prime numbers form a matrix whose determinant is zero by a local estimate. By this method, they proved for all , where .
In [24], Heath-Brown generalized the method of [3] to the higher dimensional case. His idea is to focus on a subset of whose reductions modulo a prime number are a same regular point, and he proved that this subset can be covered by a bounded degree hypersurface which do not contain the generic point of . Then he counted the number of regular points over finite fields, and controled the regular reductions. In [5], Broberg generalized it to the case over arbitrary number fields.
In [41, 42], Serre asked whether is verified for all arithmetic varieties with a particular constant . In [24], Heath-Brown proposed a uniform version for all with and , which is called the dimension growth conjecture. He proved this conjecture for some special cases. Later, Browning, Heath-Brown and Salberger had some contributions on this subject, see [6, 7, 8] for the improvements of the determinant method and the proofs under certain conditions. In [39], Salberger considered the general reductions, and the multiplicities of rational points were taken into consideration, and he proved the dimension growth conjecture with certain conditions on the subvarieties of .
1.1.2 A global version
The so-called global determinant method was first introduced by Salberger in [40] in order to study the dimension growth conjecture mentioned above. In general, it allows one to use only one auxiliary hypersurface to cover the rational points of bounded height, and one needs to optimize the degree of this hypersurface. By the global version, he proved the dimension growth conjecture for and when is a curve.
1.1.3 The dependence on degree
Let be an geometrically integral variety of degree and dimension . We are also interested in the dependence of the uniform upper bound of on , in particular when is a plane curve ( and ).
In [47], Walkowiak studied this problem by counting integral points over . In [32, Théorème 2.10], Motte obtained an estimate, which has a better dependence on but a worse dependence on than that in [47, Théorème 1].
In fact, one is able to obtain a better dependence on by the global determinant method. In [10], Castryck, Cluckers, Dittmann and Nguyen improved [48] on giving an explicit dependence on . As applications, they obtained when is a plane curve, and a better partial result of the dimension growth conjecture than that in [40], and the estimates in [47] and [32] for the case of plane curves.
1.1.4 Formulation by Arakelov geometry
In [12, 13], H. Chen reformulated the works of Salberger [39] by Bost’s slope method from Arakelov geometry developed in [4]. In this formulation, H. Chen replaced the matrix of monomials by the evaluation map which sends a global section of a particular line bundle to its values on a family of rational points. By the slope inequalities, we can control the height of the evaluation map in the slope method, which replaces the role of Siegel’s lemma in controlling heights.
There are two advantages by the approach of Arakelov geometry. First, Arakelov geometry gives a natural conceptual framework for the determinant method over arbitrary number fields. Next, it is easier to obtain explicit estimates, since the constants obtained from the slope inequalities are given explicitly in general. I has a apple.
1.2 A global version with the formulation of Arakelov geometry
In this article, we will construct the global determinant method over an arbitrary number field by Arakelov geometry following the strategy of [12, 13]. As a direct application, we will study the problem of counting rational points in plane curves, and we consider how these upper bounds depend on the degree. Some ideas are inspired by [40, 48, 10].
1.2.1 Main results
First, we have the control of auxiliary hypersurface below in Corollary 5.3 and Corollary 5.3, which are deduced from Theorem 5.2.
1.2.2 Potential applications
Similar to the previous applications of the determinant, the above estimates are able to be applied to study the uniform upper bound of the number of rational points with bounded height, where we will initiate the induction on the dimension as usual and the study of the distribution of the loci of small degree in a variety (see [40, §4] for such an example, which considered the density of conics in a cubic surface). By the above operation, we are able to obtain estimates of general arithmetic varieties from those of hypersurfaces via a suitable linear projection.
Since our method works over arbitrary number field and gives an explicit estimate (or under some technical conditions), the further applications is possible to go well under the same conditions and also be explicit.
As a direct application, we have the following results on counting rational points of bounded height in plane curves in Theorem 6.1 and Theorem 6.2.
\theoname \the\smf@thm.
Theorem 6.1 generalizes [10, Theorem 2] over an arbitrary number field, and gives an explicit estimate under the assumption of the Generalized Riemann Hypothesis. Theorem 6.2 can be viewed as a projective analogue of [10, Theorem 3] over an arbitrary number field, and a better partial result of the conjecture of Heath-Brown referred at Remark 6.2. These two estimates are better than those given in [47, Théorème 1] and [32, Théorème 2.10].
1.2.3 The role of the Generalized Riemann Hypothesis
In this work, some explicit estimates of the distribution of primes ideals are applied. If we admit GRH (the Generalized Riemann Hypothesis) of the Dedekind zeta function of the base number field, we are able to obtain more explicit estimates, see [20], for example. Without the assumptions of GRH, it seems to be very difficult to obtain such explicit estimates over an arbitrary number field, since we do not know the zero-free region of the Dedekind zeta function. If we know the zero-free region clearly enough, for example, if we work on the rational number field or totally imaginary fields (see [44] and [22] respectively), or we just want an implicit estimate (see [37]), we do not need to suppose GRH.
1.3 Organization of article
This paper is organized as following. In §2, we provide some preliminaries to construct the determinant method. In §3, we formulate the global determinant method by the slope method. In §4, we give some useful estimates on the non-geometrically integral reductions, a count of multiplicities over finite fields, the distributions of some particular prime ideals, and the geometric Hilbert-Samuel function. In §5, we provide an explicit upper bound of the determinant and lower bounds of auxiliary hypersurfaces. In §6, we give two uniform upper bounds of rational points of bounded height in plane curves. In §7, under the assumption of GRH, we give some explicit estimates of the distribution of prime ideals of bounded norm in a ring of integers, and explain how to apply these explicit estimates in the global determinant to get more explicit estimates. In Appendix A, we will give an explicit lower bound of a useful function induced by the local Hilbert-Samuel function.
Acknowledgement
I would like to thank Prof. Per Salberger for introducing his brilliant work [40] to me, and for explaining to me some ingredients of his work. These discussions and suggestions play a significant role in this paper. I would also like to thank Prof. Stanley Yao Xiao for his suggestions on the study of the distribution of prime ideals. In particular, I would like to thank the anonymous referee for all the suggestions on revising the manuscript of the paper. Chunhui Liu was supported by Fundamental Research Funds for the Central Universities FRFCU5710010421.
2 Fundamental settings
In this section, we will introduce some preliminaries to understand the problem of counting rational points of bounded height. In particular, we will provide some basic notions in Arakelov geometry.
2.1 Counting rational points of bounded height
Let be a number field, and be its ring of integers. We denote by the set of finite places of , and by the set of infinite places of . In addition, we denote by the set of places of . For every and , we define the absolute value for each , extending the usual absolute values on or . Here denotes the -adic field , where is extended from under the extension .
Let . We define the height of in as
(1) |
We also define the logarithmic height of as
(2) |
which is invariant under the extensions over (cf. [25, Lemma B.2.1]).
Suppose that is a closed integral subscheme of , and is the closed immersion. For , we define , and usually we omit the closed immersion if there is no confusion. Next, we denote
By the Northcott’s property (cf. [25, Theorem B.2.3]), the cardinality is finite for a fixed .
2.2 A function induced by local Hilbert-Samuel functions
In this part, we will introduce a function induced by the local Hilbert-Samuel function of schemes at a closed point, and we will use this function in Proposition 3.1. For the motivation and background, see [39, §2] and [13, §3.2].
Let be a field, and be a closed subscheme of of pure dimension , which means all its irreducible components have the same dimension. Let be a closed point of . We denote by
(3) |
the local Hilbert-Samuel function of at the point with the variable , where is the maximal ideal of the local ring , and is the residue field of the local ring . For this function, we have the polynomial asymptotic
(4) |
when , and we define the positive integer as the multiplicity of point in .
We define the series as the increasing series of non-negative integers such that every integer appears exactly times in this series. For example, if , , then the series is
Let be the partial sum of the series , which is
(5) |
for all .
If is a hypersurface of , then by [26, Example 2.70 (2)], the local Hilbert-Samuel function of at the point defined in (3) is
In this case, we have the following explicit lower bound of , which is
(6) |
This lower bound has the optimal dominant term by the argument in [39, Main Lemma 2.5] and some other subsequent references. In Appendix A, we will provide a detailed proof of this lower bound.
2.3 Normed vector bundles
A normed vector bundle over is all the pairings , where:
-
—
is a projective -module of finite rank;
-
—
is a family of norms, where is a norm over which is invariant under the action of . We consider a complex place and its conjugation as two different places.
If for all , the norms are Hermitian, we say that is a Hermitian vector bundle over . If , we say that is a Hermitian line bundle.
Suppose that is a sub--module of . We say that is a saturated sub--module if is a torsion-free -module.
Let and be two Hermitian vector bundles. If is a saturated sub--module of and is the restriction of over for every , we say that is a sub-Hermitian vector bundle of over .
We say that is a quotient Hermitian vector bundle of over , if for every , the module is a projective quotient -module of and is the induced quotient space norm of .
For simplicity, we will denote by below.
2.4 Arakelov invariants
Let be a Hermitian vector bundle over , and be a -basis of . We will introduce some invariants in Arakelov geometry below.
2.4.1 Arakelov degree
The Arakelov degree of is defined as
where follows the definition in [11, 2.1.9] for all , and is the Gram matrix of the basis with respect to . We refer the readers to [17, 2.4.1] for a proof of the equivalence of the above two definitions. The Arakelov degree is independent of the choice of the basis by the product formula (cf. [33, Chap. III, Proposition 1.3]). In addition, we define
as the normalized Arakelov degree of , which is independent of the choice of .
2.4.2 Slope
Let be a non-zero Hermitian vector bundle over , and be the rank of . The slope of is defined as
In addition, we denote by the maximal value of slopes of all non-zero Hermitian subbundles, and by the minimal value of slopes of all non-zero Hermitian quotients bundles of .
2.4.3 Height of linear maps
Let and be two non-zero Hermitian vector bundles over , and be a non-zero homomorphism. The height of is defined as
where is the operator norm of induced by the above linear homomorphism with respect to .
We refer the readers to [4, Appendix A] for some equalities and inequalities on Arakelov degrees and corresponding heights of homomorphisms.
2.5 Arithmetic Hilbert-Samuel function
Let be a Hermitian vector bundle of rank over , and be the projective space which represents the functor from the category of commutative -algebras to the category of sets mapping all -algebra to the set of projective quotient -module of of rank . Let (or if there is no confusion) be the universal bundle, and we denote by (or ) the line bundle for simplicity. The Hermitian metrics on induce by quotient of Hermitian metrics (i.e. Fubini-Study metrics) on which define a Hermitian line bundle on .
For every , let
(7) |
and let be its rank over . In fact, we have
(8) |
For each , we denote by the norm over such that
(9) |
where is the corresponding Fubini-Study norm.
2.5.1 Metric of John
Next, we introduce the metric of John, see [45] for a systematic introduction of this notion. In general, for a given symmetric convex body , there exists the unique ellipsoid, called ellipsoid of John, contained in with the maximal volume.
For the -module and any place , we take the ellipsoid of John of its unit closed ball defined via the norm, and this ellipsoid induces a Hermitian norm, noted by . For every section , the inequality
(10) |
is verified by [45, Theorem 3.3.6].
2.5.2 Evaluation map
Let be an integral closed subscheme of , and be the Zariski closure of in . We denote by
(11) |
the evaluation map over induced by the closed immersion of in . We denote by the largest saturated sub--module of such that . When the integer is large enough, the homomorphism is surjective, which means .
The -module is equipped with the quotient metrics (from ) such that is a Hermitian vector bundle over , noted by this Hermitian vector bundle.
\definame \the\smf@thm (Arithmetic Hilbert-Samuel function).
Let be the Hermitian vector bundle over defined above from the map (11). We say that the function which maps the positive integer to is the arithmetic Hilbert-Samuel function of with respect to the Hermitian line bundle .
2.6 Height of rational points
In this part, we will define a height function of rational points by Arakelov geometry.
Let be a Hermitian vector bundle of rank over , , and be the Zariski closure of in . Let be the universal bundle equipped with the corresponding Fubini-Study metric at each , then is a Hermitian vector bundle over . We define the height of the rational point with respect to as
(12) |
We keep all the above notations. We choose
(13) |
where for every , is the -norm mapping to . In this case, we use the notations and for simplicity. We suppose that has a -rational projective coordinate , then we have (cf. [31, Proposition 9.10])
\remaname \the\smf@thm.
Let be a projective scheme, and . We define the height of as . We will just use the notation or if there is no confusion of the morphism and the Hermitian line bundle .
2.7 Height functions of arithmetic varieties
In this part, we will introduce several height functions of arithmetic varieties, which evaluate their arithmetic complexities.
2.7.1 Arakelov height
First, we will introduce a kind of height functions of arithmetic varieties defined by the arithmetic intersection theory developped by Gillet and Soulé in [18], which is first introduced by Faltings in [15, Definition 2.5], see also [43, III.6].
\definame \the\smf@thm (Arakelov height).
Let be a number field, be its ring of integers, be a Hermitian vector bundle of rank over , and be a Hermitian line bundle over . Let be a pure dimensional closed subscheme of of dimension , and be the Zariski closure of in . The Arakelov height of is defined as the arithmetic intersection number
where is the arithmetic first Chern class of (cf. [43, Chap. III.4, Proposition 1] for its definition). This height is noted by or .
2.7.2 Heights of hypersurfaces
Let be a hypersurface in of degree . By [23, Proposition 7.6 (d), Chap. I], is define by a homogeneous polynomial of degree . We define a height function of hypersurfaces by considering its polynomial of definition.
\definame \the\smf@thm (Naive height).
Let
We define the naive height of as
and
In addition, if is homogeneous and defines the hypersurface , we define the naive height of as
2.7.3 Comparison of height functions
In order to compare and for a hypersurface , we refer the following result in [28].
\propname \the\smf@thm.
Let be a hypersurface in of degree . With all the notations above, we have
where .
Démonstration.
Since is a hypersurface, the Chow variety of is just itself. Then we have the result from [28, Proposition 3.7] directly after some elementary calculations. ∎
3 Global determinant method for hypersurfaces
In the rest part of this article, unless specially mentioned, we suppose that is an integral hypersurface in , and is its Zariski closure in . In fact, is the generic fiber of . When we consider the height of a rational point embedded into , we use the definition in (12) by Arakelov geometry. Let be a maximal ideal of , and we denote by the fiber at .
Let be the rank of over , where is defined in §2.5. For the case where is a hypersurface of degree in , we have
Our main target of this section is to prove the following result.
\theoname \the\smf@thm.
We keep all the notations in §2.5 and this section. Let be a closed integral subscheme in , and be its Zariski closure in . Let be a finite family of maximal ideals of , and be a family of rational points of . For a fixed prime ideal of , let be the multiplicity of the point in , and we denote . If the inequality
(14) | |||
is verified, then there exists a section , which contains but does not contain the generic point of . In other words, can be covered by a hypersurfaces of degree which does not contain the generic point of .
3.1 Auxiliary results
We refer to some results in [12, 13], which are used in the reformulation of the determinant method by Arakelov geometry. We will also prove a new auxiliary lemma.
\propname \the\smf@thm ([12], Proposition 2.2).
Let be a Hermitian vector bundle of rank over , and be a family of Hermitian line bundles over . If
is an injective homomorphism, then there exists a subset of whose cardinality is such that the following equality
is verified, where is the canonical projection.
In order to benefit the readers, we will provide the details on the construction of certain local homomorphisms, which are introduced in [39, Lemma 2.4], see also [13, §3.2].
Let be an integral closed subscheme of and be the Zariski closure of in . Let be a maximal ideal of and . In this case, is a local algebra over . Let be a family of local homomorphisms of -algebras from to .
Let be a free sub--module of finite type of and let be the -linear homomorphism
Since is a local homomorphism of -algebras, it must be surjective. Let be the kernel of , then we have . Furthermore, since is a local ring and we suppose that is its maximal ideal, then we have . Moreover, since is a local homomorphism, we have . For each , is an -module of finite type.
In order to estimate its rank, we need the following result.
\lemmname \the\smf@thm.
With all the above notations and constructions, we have
Démonstration.
By definition, we have
Next, from the facts and , we claim that
is verified. In fact, for every , there exist and , such that . Then . Since , then . So we have . Conversely, since is the maximal ideal, then .
By the above fact, we have
which terminates the proof. ∎
By Nakayama’s lemma (cf. [30, Theorem 2.2]), we deduce that the rank of over is equal to the rank of over from the isomorphism in Lemma 3.1, which is the value of the local Hilbert-Samuel function defined in (3).
By this fact, we consider the filtration
of , which induces the filtration
(15) |
of whose -th sub-quotient is a free -module of rank smaller than or equal to .
We suppose that the reductions of all the above local homomorphisms modulo are same, which means all the composed homomorphisms
are the same for every , where the last arrow is the canonical reduction morphism modulo . We note . In this case, the restriction of on has its norm smaller than . In fact, for any , we have and hence we have .
By the above argument, we have the following result from [13, Lemma 3.2, Lemma 3.3].
\propname \the\smf@thm ([13], Proposition 3.4).
Let be a maximal ideal of and . Suppose that is a family of local -linear homomorphism from to whose reduction modulo are the same. Let be a free sub--module of finite type of and . Then for any integer , we have
(16) |
where , and is defined in (5).
The following lemma will be used in the global determinant estimate.
\lemmname \the\smf@thm.
Let be a normed field, be four normed vector spaces over , and be two -linear isomorphisms. Suppose and . We equipped
with the corresponding maximal value norms. Then we have
where the above is the norm of operators.
Démonstration.
By definition, the linear maps and are both scalar products by the corresponding determinants, and is the scalar product of the above two determinants. Then we have the result by definition directly. ∎
3.2 Proof of Theorem 3
Proof of Theorem 3.
Let be an integer larger than . We suppose that the global section predicted by Theorem 3 does not exist. Then the evaluation map
is injective. We can replace by one of its subsets such that is an isomorphism. From now on, we suppose is isomorphic, which means . Then by Proposition 3.1, we have
which implies
Now we estimate the height of . For every , we have
where the second inequality comes from the definition of metrics of John in §2.5.1.
Now we consider the case of . The homomorphism is induced by a homomorphism of -module
where is the Zariski closure of in for each . Then for every , we have .
We fix a maximal ideal of corresponding to , and decompose the set as the disjoint union
where all elements in modulo are the same point . If is empty for some , we define for simplicity. With the above notations, let
be an -basis of such that generates for all and . Since is a local ring, the -module is free, then there exists such a basis for a fixed maximal ideal . We denote by the sub--module of generated by .
4 Some quantitative estimates
In order to apply the global determinant method introduced in Theorem 3, we need to gather enough information on the term in it. For this target, we need to have a control of the reduction type of , an upper bound of when is geometrically integral, and a distribution of certain prime ideals of . We will also provide an explicit estimate of the geometric Hilbert-Samuel function of hypersurfaces.
4.1 Control of the non-geometrically integral reductions
Let be a geometrically integral hypersurface of degree , be its Zariski closure, and for every . By [21, Théorème 9.7.7], the set
(18) |
is finite.
Next, we introduce a numerical description of the set . In fact, there are fruitful results on this subject, but most of them are over rational number field . In [29], the estimate [38, Satz 4] was generalized over arbitrary number fields by using a height function in an adelic sense by the approach of [28, §3.4]. By [29, Proposition 4.1], we have
(19) |
where is the naive height of defined in Definition 2.7.2, , and the constant
In fact, we have .
4.2 Quantitative estimates over finite fields
In this subsection, we give an upper bound of the term for an arbitrary maximal ideal of , where is defined in the statement of Theorem 3. In this part, we consider this problem over arbitrary finite fields.
Let be the finite field with elements, be a geometrically integral hypersurface in of degree , and , where is the multiplicity of in defined via the local Hilbert-Samuel fuction in (4). Then we have
In order to estimate , we will consider the terms and separately.
4.2.1
For the estimate of , there are fruitful results on it. For our application, we have the following result deduced from [9, Corollary 5.6].
\propname \the\smf@thm.
Let be a geometrically integral hypersurface of degree over the finite field . When or , we have
Démonstration.
We consider this estimate case by case as following.
-
1.
If , we have . Then
-
2.
If , we have . Then
-
3.
If , by [9, Corollary 5.6], we have
∎
4.2.2
4.2.3
4.3 Distribution of certain prime ideals
In this part, we will consider some distributions of prime ideals of the ring of integers of number fields.
4.3.1 Distribution of prime ideals containing a fixed ideal
In this part, we will consider the distribution of certain maximal ideals of . First, we generalize [40, Lemma 1.10] over an arbitrary number field, where the former result works over only.
\lemmname \the\smf@thm.
Let be a proper ideal of , be an prime ideal of , and . Then we have
where the above sum takes all over the prime ideals contained in of .
Démonstration.
We will prove the inequality for the case of at first, and then we show the general case by it.
Case of . - In this case, we will repeat the proof of [40, Lemma 1.10] by Salberger, since this preprint is not easily available. Suppose that is generated by the positive square-free integer , and let be a positive integer such that . For the prime , let be the largest integer such that . By [44, Tome I, Corollaire 1.7] and [44, Tome I, Théorème 1.8], we have
and then we obtain
Let for , and then we accomplish the proof for .
Case of arbitrary number fields. - Let
where are distinct prime ideals of , and for all . Let the prime be the characteristic of the prime ideal , where as above. For a fixed prime , there are at most prime ideals of characteristic in . For all prime and , we have
Let be the product of all the different characteristics of , and we have by definition directly. Then by the above facts, we obtain
By the case of , we have
which proves the assertion. ∎
4.3.2 Distribution of prime ideals with bounded norm
Let , , and . In this part, we consider the of
(22) |
When , these are classic estimates of Chebyshev function (cf. [44, Tome I, Théorème 2.11]) and Mertens’ first theorem (cf. [44, Tome I, Théorème 1.8]). For the case of arbitrary number fields, a generalization of [44, Tome I, Théorème 2.11] is Landau’s prime ideals theorem (cf. [37, Theorem 2.2]), and a generalization of Mertens’ first theorem was deduced from this in [37, Lemma 2.3].
In this part, we will give a more explicit version of some results [37], which will be used in the application of Theorem 3.
By Landau’s prime ideal theorem (cf. [37, Theorem 2.2]), we have
where is a constant depending on . Then there exists a function of the number field and , such that
(23) |
where for all , and depends on only.
4.3.3 Distribution of non-geometrically integral reductions
In this part, we consider a distribution of the prime ideals modulo which the reductions are not geometrically integral. In the estimate below, the estimates of (22) will be involved.
\propname \the\smf@thm.
Démonstration.
We denote by the product of all maximal ideals in , and
Then by Lemma 4.3.1 and (19), we have
By (24), we have
where is defined in (24) depending on only.
Let
Since , then we have
where the last inequality is from (19). By combining the above two estimates, we terminate the proof. ∎
\remaname \the\smf@thm.
With all the notations and assumptions in Proposition 4.3.3. We have
4.4 An explicit estimate of the geometric Hilbert-Samuel function
In this part, we will provide an explicit lower bound of the geometric Hilbert-Samuel function of a projective hypersurface, which will be used in the application of the determinant method. The inequality
will be helpful in the calculation below.
\lemmname \the\smf@thm.
Let be a hypersurface of degree in . We denote by its geometric Hilbert-Samuel function with the variable . When , we have
and
Démonstration.
In fact, we have
when .
In order to obtain the lower bound, we have
Then we obtain
when .
On the other hand, we have
(28) |
which terminates the proof by an elementary calculation. ∎
5 An explicit estimate of determinant
In this section, we will give an upper bound of the degree of the auxiliary hypersurface determined by Theorem 3.
5.1 A uniform lower bound of arithmetic Hilbert-Samuel functions
Firstly, we refer to a result in [12], which is an application of the uniform lower bound of the arithmetic Hilbert-Samuel functions to the determinant method.
\propname \the\smf@thm ([12], Propoosition 2.12).
The uniform lower bound of for all will play a significant role in the construction of auxiliary hypersurfaces if we want to apply Proposition 5.1. In [14], David and Philippon give an explicit uniform lower bound of . This result is reformulated by H. Chen in [12, Theorem 4.8] by the language of the slope method. In fact, let be a closed integral subscheme of of dimension and degree , and be its Zariski closure in . The inequality
(29) |
is uniformly verified for all (see also [12, Remark 4.9] for some minor modifications), where follows the definition in Definition 2.7.1.
By Proposition 5.1, all the rational points with small heights in a projective variety can be covered by one hypersurface which does not contain the generic point of the original variety. The following result gives a numerical description of this observation.
\propname \the\smf@thm.
Let be an integral hypersurface of degree in , the constant and the constant
(30) |
If
then there exists a hypersurface of degree smaller than , which contains all rational points in but does not contain the generic point of , where we use the height function defined in (12).
Démonstration.
\remaname \the\smf@thm.
With all the notations in Proposition 5.1. By the arithmetic Hilbert-Samuel Theorem of arithmetic ample line bundles (cf. [19, Theorem 8], [49, Theorem 1.4] and [1, Théorème principal]), we have
for tends into infinity. So it is expected that we can obtain a better uniform lower bound of than that in (29). If we have a better explicit lower bound, we can improve the bound given in Proposition 5.1.
5.2 Estimate of the determinant
In the global determinant method, for each geometrically integral hypersurface, we allow only one auxiliary hypersurface to cover its rational points with bounded height not containing the generic point of the original hypersurface, and we optimize the degree of this auxiliary hypersurface.
In this part, we will give an upper bound of the degree of the auxiliary hypersurface determined in Theorem 3, where the size of non-geometrically integral reductions and the height of the original hypersurface will be involved.
Before the statement of the main theorem in this paragraph, we will introduce two constants depending on the number field and the positive integer , which will be used in the estimate of the determinant.
Let be a number field, we denote
(31) |
where is introduced in (23). By [37, Theorem 2.2], the above supremum exists and depends on only.
Let and be two integers, we denote
(32) |
where is defined in (24), and is defined in (26). By taking [37, Theorem 2.2] into the estimate of (26), the above supremum exists and depends on and only.
\theoname \the\smf@thm.
Let be a number field. Let be a geometrically integral hypersurface in of degree , and be the set of rational points in whose height is smaller than with respect to the above closed immersion, see (12) for the definition of the height function used above. Then there exists a hypersurface in of degree smaller than
which covers but does not contain the generic point of , where the constant
the constant is defined in (30), is defined in Proposition 4.3.3, is defined in (31), is defined in (32), and the height of is defined in Definition 2.7.2.
Démonstration.
By Proposition 5.1, we divide the proof into two parts.
I. Case of large height varieties. - If
where the constant is defined in (30) and is defined in Definition 2.7.2. Then by Proposition 5.1, can be covered by a hypersurface of degree no more than which does not contain the generic point of . By an elementary calculation, we obtain that is smaller than the bound provided in the statement of the theorem, for , , , and .
II. Case of small height varieties. - For the case of
which is equivalent to
we will treat it as following. We keep all the notations in Theorem 3, and we suppose from now on. We denote the set
and we apply Theorem 3 to the reductions at . If there does not exist such a hypersurface, then by Theorem 3 applied in the above sense, we have
From the explicit lower bound of provided at (29) and Proposition 5.1, we deduce
For the estimate of , we denote
Then by (24), we have
where the notation is introduced in Proposition 4.3.3, and is defined in (24).
For the term , it is equal to zero when . When , by (25), we have
By the above two estimates, we obtain
by combining the above two inequalities.
II-3. Deducing the contradiction. - We take (5.2) and (5.2) into (5.2), and we do some elementary calculations. Then the inequality
is uniformly verified for all .
When and , we have
by an elementary calculation. We take the above inequality into (5.2), and then we obtain
which deduces
with the constant in the statement of this theorem, and it leads to the contradiction. ∎
5.3 Control of auxiliary hypersurfaces
The following two upper bounds of the degree of the auxiliary hypersurface are deduced from Theorem 5.2 directly.
\coroname \the\smf@thm.
Démonstration.
Compared with Corollary 5.3, the result below has a better dependence on the degree of the original hypersurface but a little worse dependence on the bound of heights.
\coroname \the\smf@thm.
Démonstration.
If
then by Proposition 5.1, can be covered by a hypersurface of degree no more than which does not contain the generic point of . The upper bound of the degree satisfies the bound provided in the statement.
6 Counting rational points in plane curves
As applications of Corollary 5.3 and Corollary 5.3, we have the following uniform upper bounds of the number of rational points with bounded height in plane curves.
6.1 A generalization over an arbitrary number field
The following result generalizes [10, Theorem 2] over an arbitrary number field, which has the optimal dependence on the bound of heights.
\theoname \the\smf@thm.
Let be a geometrically integral curves in of degree . Then we have
where the constant is defined in Corollary 5.3. In addition, we have
6.2 A better dependence on the degree
In this part, we will provide another uniform upper bound of rational points with bounded height in plane curves. This result has a better dependence on the degree than that of Theorem 6.1, but a bit worse dependence on the bound of heights.
\theoname \the\smf@thm.
Let be a geometrically integral curves in of degree . Then we have
where the constant is defined in Corollary 5.3. In addition, we have
when .
Démonstration.
\remaname \the\smf@thm.
7 Explicit estimates under the assumption of GRH
Under the assumption of GRH (the Generalized Riemann Hypothesis) for the Dedekind zeta function of the number field , we have more explicit estimates of
where , , and . By these results, we are able to obtain more explicit estimates of in the global determinant method and more explicit estimates in the densities of rational points. In this section, we will give explicit estimates of the remainders of the above , and .
7.1 Explicit estimates of the distribution of prime ideals with bounded norms
In order to obtain explicit estimates, first we refer to a result in [20] under the assumption of GRH. Then by the same technique in [37], we get an explicit generalization of Mertens’ first theorem. If we do not need the explicit version, it is not necessary to assume GRH.
Let be the discriminant of the number field . By [20, Corollary 1.3], if , we have
under the assumption of GRH. Since , we obtain
(39) |
by Mathematica under the same assumption.
Same as the application of [37, Lemma 2.1] to the proof of [37, Lemma 2.3], by Abel’s summation formula, we have
From (39), we have
Then by an elementary calculation executed by Mathematica, we obtain
(40) |
for under the assumption of GRH.
The same goes for the above, we have
Then from (39), we obtain
Then after an elementary calculation executed by Mathematica, we have
(41) |
uniformly for all under the assumption of GRH.
7.2 Estimates of remainders
We compare (23) with (39), (24) with (40), and (25) with (41). Then we can suppose
in (23),
in (24), and
in (25) under the assumption of GRH. We take the above estimates of , and into Proposition 4.3.3, (31) and (32), then we obtain a more explicit estimate in Theorem 5.2. Hence the estimates in Corollary 5.3, Corollary 5.3, Theorem 6.1 and Theorem 6.2 are more explicit under the assumption of GRH.
\remaname \the\smf@thm.
Besides the case of , if we work over some other particular number fields, the assumption of the Generalized Riemann Hypothesis may not be obligatory. For example, in [22, Theorem 2], we are able to do it over totally imaginary fields. It depends on the understanding of the zero-free region of the Dedekind zeta function of the number field .
Annexe A An explicit lower bound of
In this appendix, we give an explicit lower bound of the function defined in (5) for the case of hypersurfaces.
In the following proof of Proposition A, the inequality
will be very useful, where and are two positive integers, and .
\propname \the\smf@thm.
Let be a hypersurface of , be a closed point in , and be the multiplicity of in induced by its local Hilbert-Samuel function. The function is defined in the equality (5). Then we have
Démonstration.
For the case of hypersurfaces, by [26, Example 2.70 (2)], we have
We define the function , then we have
Then we obtain
In order to get a lower bound of , we need to estimate the term
In fact, we have the estimate
By the equality (42), we obtain
where we use the estimate in the inequality (A). In addition, we obtain the inequality
In addition, we have
Then
Finally we have
for . Then we obtain the result. ∎
Références
- [1] A. Abbes & T. Bouche – « Théorème de Hilbert-Samuel “arithmétique” », Université de Grenoble. Annales de l’Institut Fourier 45 (1995), no. 2, p. 375–401.
- [2] M. Bhargava, A. Shankar, T. Taniguchi, F. Thorne, J. Tsimerman & Y. Zhao – « Bounds on 2-torsion in class groups of number fields and integral points on elliptic curves », Journal of the American Mathematical Society 33 (2020), no. 4, p. 1087–1099.
- [3] E. Bombieri & J. Pila – « The number of integral points on arcs and ovals », Duke Mathematical Journal 59 (1989), no. 2, p. 337–357.
- [4] J.-B. Bost – « Périodes et isogénies des variétés abéliennes sur les corps de nombres (d’après D. Masser et G. Wüstholz) », Astérisque (1996), no. 237, p. Exp. No. 795, 4, 115–161, Séminaire Bourbaki, Vol. 1994/1995.
- [5] N. Broberg – « A note on a paper by R. Heath-Brown: “The density of rational points on curves and surfaces” [Ann. of Math. (2) 155 (2002), no. 2, 553–595] », Journal für die Reine und Angewandte Mathematik 571 (2004), p. 159–178.
- [6] T. D. Browning & D. R. Heath-Brown – « The density of rational points on non-singular hypersurfaces. I », The Bulletin of the London Mathematical Society 38 (2006), no. 3, p. 401–410.
- [7] — , « The density of rational points on non-singular hypersurfaces. II », Proceedings of the London Mathematical Society. Third Series 93 (2006), no. 2, p. 273–303, With an appendix by J. M. Starr.
- [8] T. D. Browning, D. R. Heath-Brown & P. Salberger – « Counting rational points on algebraic varieties », Duke Mathematical Journal 132 (2006), no. 3, p. 545–578.
- [9] A. Cafure & G. Matera – « Improved explicit estimates on the number of solutions of equations over a finite field », Finite Fields and Their Applications 12 (2006), no. 2, p. 155–185.
- [10] W. Castryck, R. Cluckers, P. Dittmann & K. H. Nguyen – « The dimension growth conjecture, polynomial in the degree and without logarithmic factors », Algebra & Number Theory 14 (2020), no. 8, p. 2261–2294.
- [11] H. Chen – « Convergence des polygones de Harder-Narasimhan », Mémoires de la Société Mathématique de France 120 (2010), p. 1–120.
- [12] — , « Explicit uniform estimation of rational points I. Estimation of heights », Journal für die Reine und Angewandte Mathematik 668 (2012), p. 59–88.
- [13] — , « Explicit uniform estimation of rational points II. Hypersurface coverings », Journal für die Reine und Angewandte Mathematik 668 (2012), p. 89–108.
- [14] S. David & P. Philippon – « Minorations des hauteurs normalisées des sous-variétés des tores », Annali della Scuola Normale Superiore di Pisa. Classe di Scienze. Serie IV 28 (1999), no. 3, p. 489–543.
- [15] G. Faltings – « Diophantine approximation on abelian varieties », Annals of Mathematics. Second Series 133 (1991), no. 3, p. 549–576.
- [16] W. Fulton – Intersection theory, second éd., Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics], vol. 2, Springer-Verlag, Berlin, 1998.
- [17] H. Gillet & C. Soulé – « On the number of lattice points in convex symmetric bodies and their duals », Israel Journal of Mathematics 74 (1991), no. 2-3, p. 347–357.
- [18] H. Gillet & C. Soulé – « Arithmetic intersection theory », Institut des Hautes Études Scientifiques. Publications Mathématiques (1990), no. 72, p. 93–174 (1991).
- [19] — , « An arithmetic Riemann-Roch theorem », Inventiones Mathematicae 110 (1992), no. 3, p. 473–543.
- [20] L. Grenié & G. Molteni – « Explicit smoothed prime ideals theorems under GRH », Mathematics of Computation 85 (2016), no. 300, p. 1875–1899.
- [21] A. Grothendieck & J. Dieudonné – « Éléments de géométrie algébrique. IV. Étude locale des schémas et des morphismes de schémas. III », Institut des Hautes Études Scientifiques. Publications Mathématiques (1966), no. 28, p. 255.
- [22] M. Grześkowiak – « Explicit bound for the prime ideal theorem in residue classes », in International Conference on Number-Theoretic Methods in Cryptology, Springer, 2017, p. 48–68.
- [23] R. Hartshorne – Algebraic geometry, Springer-Verlag, New York, 1977, Graduate Texts in Mathematics, No. 52.
- [24] D. R. Heath-Brown – « The density of rational points on curves and surfaces », Annals of Mathematics. Second Series 155 (2002), no. 2, p. 553–595.
- [25] M. Hindry & J. H. Silverman – Diophantine geometry, Graduate Texts in Mathematics, vol. 201, Springer-Verlag, New York, 2000, An introduction.
- [26] J. Kollár – Lectures on resolution of singularities, Annals of Mathematics Studies, vol. 166, Princeton University Press, Princeton, NJ, 2007.
- [27] C. Liu – « Comptage des multiplicités dans une hypersurface sur un corps fini », arxiv:1606.09337 (2016), to appear in Kyoto Journal of Mathematics.
- [28] — , « Fibres non réduites d’un schéma arithmétique », Mathematische Zeitschrift 299 (2021), no. 3-4, p. 1275–1302.
- [29] — , « Control of the non-geometrically integral reductions », International Journal of Number Theory 18 (2022), no. 1, p. 47–60.
- [30] H. Matsumura – Commutative ring theory, second éd., Cambridge Studies in Advanced Mathematics, vol. 8, Cambridge University Press, Cambridge, 1989, Translated from the Japanese by M. Reid.
- [31] A. Moriwaki – Arakelov geometry, Translations of Mathematical Monographs, vol. 244, American Mathematical Society, Providence, RI, 2014, Translated from the 2008 Japanese original.
- [32] F. Motte – « De la géométrie à l’arithmétique en théorie inverse de Galois », Thèse, Université de Lille, Mai 2019.
- [33] J. Neukirch – Algebraic number theory, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 322, Springer-Verlag, Berlin, 1999, Translated from the 1992 German original and with a note by Norbert Schappacher, With a foreword by G. Harder.
- [34] M. Paredes & R. Sasyk – « Uniform bounds for the number of rational points on varieties over global fields », arXiv:2101.12174 (2021), to appear in Algebra and Number Theory.
- [35] J. Pila – « Density of integral and rational points on varieties », Astérisque (1995), no. 228, p. 4, 183–187, Columbia University Number Theory Seminar (New York, 1992).
- [36] H. Randriambololona – « Métriques de sous-quotient et théorème de Hilbert-Samuel arithmétique pour les faisceaux cohérents », Journal für die Reine und Angewandte Mathematik 590 (2006), p. 67–88.
- [37] M. Rosen – « A generalization of Mertens’ theorem », Journal of the Ramanujan Mathematical Society 14 (1999), p. 1–19.
- [38] W. Ruppert – « Reduzibilität ebener kurven. », Journal für die Reine und Angewandte Mathematik 369 (1986), p. 167–191.
- [39] P. Salberger – « On the density of rational and integral points on algebraic varieties », Journal für die Reine und Angewandte Mathematik 606 (2007), p. 123–147.
- [40] — , « Counting rational points on projective varieties », (2013), preprint.
- [41] J.-P. Serre – Lectures on the Mordell-Weil theorem, third éd., Aspects of Mathematics, Friedr. Vieweg & Sohn, Braunschweig, 1997, Translated from the French and edited by Martin Brown from notes by Michel Waldschmidt, With a foreword by Brown and Serre.
- [42] — , Topics in Galois theory, second éd., Research Notes in Mathematics, vol. 1, A K Peters, Ltd., Wellesley, MA, 2008, With notes by Henri Darmon.
- [43] C. Soulé – Lectures on Arakelov geometry, Cambridge Studies in Advanced Mathematics, vol. 33, Cambridge University Press, Cambridge, 1992, With the collaboration of D. Abramovich, J.-F. Burnol and J. Kramer.
- [44] G. Tenenbaum – Introduction à la théorie analytique et probabiliste des nombres, 4-ième éd., Échelles, Belin, 2015.
- [45] A. C. Thompson – Minkowski geometry, Encyclopedia of Mathematics and its Applications, vol. 63, Cambridge University Press, Cambridge, 1996.
- [46] F. Vermeulen – « Points of bounded height on curves and the dimension growth conjecture over », arxiv:2003.10988 (2020).
- [47] Y. Walkowiak – « Théorème d’irréductibilité de hilbert effectif », Acta Arithmetica 116 (2005), p. 343–362.
- [48] M. N. Walsh – « Bounded rational points on curves », International Mathematics Research Notices. IMRN (2015), no. 14, p. 5644–5658.
- [49] S. Zhang – « Positive line bundles on arithmetic varieties », Journal of the American Mathematical Society 8 (1995), no. 1, p. 187–221.