This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\alttitle

Autour de la méthode globale de déterminant

On the global determinant method

Chunhui Liu Institute for Advanced Study in Mathematics
Harbin Institute of Technology
150001 Harbin
P. R. China
chunhui.liu@hit.edu.cn
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.

{altabstract}

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 XKnX\hookrightarrow\mathbb{P}^{n}_{K} be a projective variety over a number field KK. For every rational point ξX(K)\xi\in X(K), we denote by HK(ξ)H_{K}(\xi) the height (see (1) for the definition) of ξ\xi with respect to the above closed immersion, for example, the classic Weil height (cf. [25, §B.2, Definition]). Let

S(X;B)={ξX(K)HK(ξ)B},S(X;B)=\{\xi\in X(K)\mid H_{K}(\xi)\leqslant B\},

where B1B\geqslant 1 and the embedding morphism is omitted. By the Northcott’s property, the cardinality #S(X;B)\#S(X;B) is finite for a fixed BB\in\mathbb{R}.

In order to understand the density of the rational points of XX, it is an important approach to study the function #S(X;B)\#S(X;B) with the variable B+B\in\mathbb{R}^{+}. For different required properties of #S(X;B)\#S(X;B), numerous methods have been applied. In this article, we are interested in the uniform upper bound of #S(X;B)\#S(X;B) for all XKnX\hookrightarrow\mathbb{P}^{n}_{K} with fixed degree and dimension, and for those satisfying certain common conditions.

1.1 Determinant mathod

In order to understand the function #S(X;B)\#S(X;B) of the variable B+B\in\mathbb{R}^{+}, 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 \mathbb{Q} 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 S(X;B)S(X;B) having the same reduction modulo some prime numbers form a matrix whose determinant is zero by a local estimate. By this method, they proved #S(X;B)δ,ϵB2/δ+ϵ\#S(X;B)\ll_{\delta,\epsilon}B^{2/\delta+\epsilon} for all ϵ>0\epsilon>0, where δ=deg(X)\delta=\deg(X).

In [24], Heath-Brown generalized the method of [3] to the higher dimensional case. His idea is to focus on a subset of S(X;B)S(X;B) 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 XX. 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 #S(X;B)XBdim(X)(logB)c\#S(X;B)\ll_{X}B^{\dim(X)}(\log B)^{c} is verified for all arithmetic varieties XX with a particular constant cc. In [24], Heath-Brown proposed a uniform version #S(X;B)d,δ,ϵBd+ϵ\#S(X;B)\ll_{d,\delta,\epsilon}B^{d+\epsilon} for all ϵ>0\epsilon>0 with δ=deg(X)\delta=\deg(X) and d=dim(X)d=\dim(X), 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 XX.

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 deg(X)=δ4\deg(X)=\delta\geqslant 4 and #S(X;B)δB2δlogB\#S(X;B)\ll_{\delta}B^{\frac{2}{\delta}}\log B when XX is a curve.

In [48], Walsh refined the global determinant method in [40], and he removed the logB\log B term in [40] when XX is a curve.

1.1.3 The dependence on degree

Let XnX\hookrightarrow\mathbb{P}^{n}_{\mathbb{Q}} be an geometrically integral variety of degree δ\delta and dimension dd. We are also interested in the dependence of the uniform upper bound of #S(X;B)\#S(X;B) on δ\delta, in particular when XX is a plane curve (n=2n=2 and d=1d=1).

In [47], Walkowiak studied this problem by counting integral points over \mathbb{Z}. In [32, Théorème 2.10], Motte obtained an estimate, which has a better dependence on BB but a worse dependence on δ\delta than that in [47, Théorème 1].

In fact, one is able to obtain a better dependence on δ\delta by the global determinant method. In [10], Castryck, Cluckers, Dittmann and Nguyen improved [48] on giving an explicit dependence on δ\delta. As applications, they obtained #S(X;B)δ4B2/δ\#S(X;B)\ll\delta^{4}B^{2/\delta} when XX 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.

In [34], the work [10] was generalized over an arbitrary global field. Before [34] was announced, Vermeulen studied the case over 𝔽q(t)\mathbb{F}_{q}(t) in [46].

Besides the uniform bounds of rational points, [10] also studied the bound of 22-torsion points of the class group of number fields, which improved the work [2, Theorem 1.1] of Bhargava, Shankar, Taniguchi, Thorne, Tsimerman and Y. Zhao.

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.

\theoname \the\smf@thm.

Let XX be a geometrically integral hypersurface in Kn\mathbb{P}^{n}_{K} of degree δ\delta, and

an=n(n1)δ1/(n1)a_{n}=\dfrac{n}{{(n-1)\delta^{1/(n-1)}}}

be a constant depending on nn. Then there is a hypersurface of degree ϖ\varpi which covers S(X;B)S(X;B) but does not contain the generic point of XX. In addition, we have

ϖK,nδ3Ban\varpi\ll_{K,n}\delta^{3}B^{a_{n}}

in Corollary 5.3, and

ϖK,nδ31/(n1)Banmax{logB[K:],1}\varpi\ll_{K,n}\delta^{3-{1}/{(n-1)}}B^{a_{n}}\max\left\{\frac{\log B}{[K:\mathbb{Q}]},1\right\}

in Corollary 5.3.

If we assume the Generalized Riemann Hypothesis, the above constants depending on KK and nn will be given explicitly in Corollary 5.3 and Corollary 5.3.

Since we apply the approach of Arakelov geometry in this article, we do not use the technique of "change of coordinate" in [48, §3] and [10, §3.4] any longer. Instead, we are able to obtain a uniform estimate directly.

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.

Let XX be a geometrically integral plane curve in K2\mathbb{P}^{2}_{K} of degree δ\delta. Then we have

#S(X;B)Kδ4B2/δ\#S(X;B)\ll_{K}\delta^{4}B^{{2}/{\delta}}

in Theorem 6.1, and

#S(X;B)Kδ3B2/δlogB\#S(X;B)\ll_{K}\delta^{3}B^{{2}/{\delta}}\log B

in Theorem 6.2.

If we assume the Generalized Riemann Hypothesis, the above constants depending on KK will be given explicitly in Theorem 6.1 and Theorem 6.2.

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 \mathbb{Q} 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 KK be a number field, and 𝒪K\mathcal{O}_{K} be its ring of integers. We denote by MK,fM_{K,f} the set of finite places of KK, and by MK,M_{K,\infty} the set of infinite places of KK. In addition, we denote by MK=MK,fMK,M_{K}=M_{K,f}\sqcup M_{K,\infty} the set of places of KK. For every vMKv\in M_{K} and xKx\in K, we define the absolute value |x|v=|NKv/v(x)|v1[Kv:v]|x|_{v}=\left|N_{K_{v}/\mathbb{Q}_{v}}(x)\right|_{v}^{\frac{1}{[K_{v}:\mathbb{Q}_{v}]}} for each vMKv\in M_{K}, extending the usual absolute values on p\mathbb{Q}_{p} or \mathbb{R}. Here v\mathbb{Q}_{v} denotes the pp-adic field p\mathbb{Q}_{p}, where vv is extended from pp under the extension K/K/\mathbb{Q}.

Let ξ=[ξ0::ξn]Kn(K)\xi=[\xi_{0}:\cdots:\xi_{n}]\in\mathbb{P}^{n}_{K}(K). We define the height of ξ\xi in Kn\mathbb{P}^{n}_{K} as

HK(ξ)=vMKmax0in{|ξi|v[Kv:v]}.H_{K}(\xi)=\prod_{v\in M_{K}}\max_{0\leqslant i\leqslant n}\left\{|\xi_{i}|_{v}^{[K_{v}:\mathbb{Q}_{v}]}\right\}. (1)

We also define the logarithmic height of ξ\xi as

h(ξ)=1[K:]logHK(ξ),h(\xi)=\frac{1}{[K:\mathbb{Q}]}\log H_{K}(\xi), (2)

which is invariant under the extensions over KK (cf. [25, Lemma B.2.1]).

Suppose that XX is a closed integral subscheme of Kn\mathbb{P}^{n}_{K}, and ϕ:XKn\phi:X\hookrightarrow\mathbb{P}^{n}_{K} is the closed immersion. For ξX(K)\xi\in X(K), we define HK(ξ)=HK(ϕ(ξ))H_{K}(\xi)=H_{K}(\phi(\xi)), and usually we omit the closed immersion ϕ\phi if there is no confusion. Next, we denote

S(X;B)={ξX(K)|HK(ξ)B}, and N(X;B)=#S(X;B).S(X;B)=\{\xi\in X(K)|H_{K}(\xi)\leqslant B\},\hbox{ and }N(X;B)=\#S(X;B).

By the Northcott’s property (cf. [25, Theorem B.2.3]), the cardinality N(X;B)N(X;B) is finite for a fixed B1B\geqslant 1.

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 kk be a field, and XX be a closed subscheme of kn\mathbb{P}^{n}_{k} of pure dimension dd, which means all its irreducible components have the same dimension. Let ξ\xi be a closed point of XX. We denote by

Hξ(s)=dimκ(ξ)(𝔪X,ξs/𝔪X,ξs+1)H_{\xi}(s)=\dim_{\kappa(\xi)}\left(\mathfrak{m}_{X,\xi}^{s}/\mathfrak{m}_{X,\xi}^{s+1}\right) (3)

the local Hilbert-Samuel function of XX at the point ξ\xi with the variable ss\in\mathbb{N}, where 𝔪X,ξ\mathfrak{m}_{X,\xi} is the maximal ideal of the local ring 𝒪X,ξ\mathcal{O}_{X,\xi}, and κ(ξ)\kappa(\xi) is the residue field of the local ring 𝒪X,ξ\mathcal{O}_{X,\xi}. For this function, we have the polynomial asymptotic

Hξ(s)=μξ(X)(d1)!sd1+o(sd1)H_{\xi}(s)=\frac{\mu_{\xi}(X)}{(d-1)!}s^{d-1}+o(s^{d-1}) (4)

when s+s\rightarrow+\infty, and we define the positive integer μξ(X)\mu_{\xi}(X) as the multiplicity of point ξ\xi in XX.

We define the series {qξ(m)}m0\{q_{\xi}(m)\}_{m\geqslant 0} as the increasing series of non-negative integers such that every integer ss\in\mathbb{N} appears exactly Hξ(s)H_{\xi}(s) times in this series. For example, if Hξ(0)=1H_{\xi}(0)=1, Hξ(1)=2,Hξ(2)=4,Hξ(3)=5,H_{\xi}(1)=2,H_{\xi}(2)=4,H_{\xi}(3)=5,\ldots, then the series {qξ(m)}m0\{q_{\xi}(m)\}_{m\geqslant 0} is

{0,1,1,2,2,2,2,3,3,3,3,3,4,}.\{0,1,1,2,2,2,2,3,3,3,3,3,4,\ldots\}.

Let {Qξ(m)}m0\{Q_{\xi}(m)\}_{m\geqslant 0} be the partial sum of the series {qξ(m)}m0\{q_{\xi}(m)\}_{m\geqslant 0}, which is

Qξ(m)=qξ(0)+qξ(1)++qξ(m)Q_{\xi}(m)=q_{\xi}(0)+q_{\xi}(1)+\cdots+q_{\xi}(m) (5)

for all mm\in\mathbb{N}.

If XX is a hypersurface of kn\mathbb{P}^{n}_{k}, then by [26, Example 2.70 (2)], the local Hilbert-Samuel function of XX at the point ξ\xi defined in (3) is

Hξ(s)=(n+s1s)(n+sμξ(X)1sμξ(X)).H_{\xi}(s)={n+s-1\choose s}-{n+s-\mu_{\xi}(X)-1\choose s-\mu_{\xi}(X)}.

In this case, we have the following explicit lower bound of Qξ(m)Q_{\xi}(m), which is

Qξ(m)>((n1)!μξ(X))1n1(n1n)mnn1n3+2n2+n42n(n+1)m.Q_{\xi}(m)>\left(\frac{(n-1)!}{\mu_{\xi}(X)}\right)^{\frac{1}{n-1}}\left(\frac{n-1}{n}\right)m^{\frac{n}{n-1}}-\frac{n^{3}+2n^{2}+n-4}{2n(n+1)}m. (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 Spec𝒪K\operatorname{Spec}\mathcal{O}_{K} is all the pairings E¯=(E,(.v)vMK,)\overline{E}=\left(E,\left(\|\raisebox{1.72218pt}{.}\|_{v}\right)_{v\in M_{K,\infty}}\right), where:

  • EE is a projective 𝒪K\mathcal{O}_{K}-module of finite rank;

  • (.v)vMK,\left(\|\raisebox{1.72218pt}{.}\|_{v}\right)_{v\in M_{K,\infty}} is a family of norms, where .v\|\raisebox{1.72218pt}{.}\|_{v} is a norm over E𝒪K,vE\otimes_{\mathcal{O}_{K},v}\mathbb{C} which is invariant under the action of Gal(/Kv)\operatorname{Gal}(\mathbb{C}/K_{v}). We consider a complex place and its conjugation as two different places.

If for all vMK,v\in M_{K,\infty}, the norms (.v)vMK,\left(\|\raisebox{1.72218pt}{.}\|_{v}\right)_{v\in M_{K,\infty}} are Hermitian, we say that E¯\overline{E} is a Hermitian vector bundle over Spec𝒪K\operatorname{Spec}\mathcal{O}_{K}. If rk𝒪K(E)=1\operatorname{rk}_{\mathcal{O}_{K}}(E)=1, we say that E¯\overline{E} is a Hermitian line bundle.

Suppose that FF is a sub-𝒪K\mathcal{O}_{K}-module of EE. We say that FF is a saturated sub-𝒪K\mathcal{O}_{K}-module if E/FE/F is a torsion-free 𝒪K\mathcal{O}_{K}-module.

Let E¯=(E,(.E,v)vMK,)\overline{E}=\left(E,\left(\|\raisebox{1.72218pt}{.}\|_{E,v}\right)_{v\in M_{K,\infty}}\right) and F¯=(F,(.F,v)vMK,)\overline{F}=\left(F,\left(\|\raisebox{1.72218pt}{.}\|_{F,v}\right)_{v\in M_{K,\infty}}\right) be two Hermitian vector bundles. If FF is a saturated sub-𝒪K\mathcal{O}_{K}-module of EE and .F,v\|\raisebox{1.72218pt}{.}\|_{F,v} is the restriction of .E,v\|\raisebox{1.72218pt}{.}\|_{E,v} over F𝒪K,vF\otimes_{\mathcal{O}_{K},v}\mathbb{C} for every vMK,v\in M_{K,\infty}, we say that F¯\overline{F} is a sub-Hermitian vector bundle of E¯\overline{E} over Spec𝒪K\operatorname{Spec}\mathcal{O}_{K}.

We say that G¯=(G,(.G,v)vMK,)\overline{G}=\left(G,\left(\|\raisebox{1.72218pt}{.}\|_{G,v}\right)_{v\in M_{K,\infty}}\right) is a quotient Hermitian vector bundle of E¯\overline{E} over Spec𝒪K\operatorname{Spec}\mathcal{O}_{K}, if for every vMK,v\in M_{K,\infty}, the module GG is a projective quotient 𝒪K\mathcal{O}_{K}-module of EE and .G,v\|\raisebox{1.72218pt}{.}\|_{G,v} is the induced quotient space norm of .E,v\|\raisebox{1.72218pt}{.}\|_{E,v}.

For simplicity, we will denote by EK=E𝒪KKE_{K}=E\otimes_{\mathcal{O}_{K}}K below.

2.4 Arakelov invariants

Let E¯\overline{E} be a Hermitian vector bundle over Spec𝒪K\operatorname{Spec}\mathcal{O}_{K}, and {s1,,sr}\{s_{1},\ldots,s_{r}\} be a KK-basis of EKE_{K}. We will introduce some invariants in Arakelov geometry below.

2.4.1 Arakelov degree

The Arakelov degree of E¯\overline{E} is defined as

deg^(E¯)\displaystyle\widehat{\deg}(\overline{E}) =\displaystyle= vMK[Kv:v]logs1srv\displaystyle-\sum_{v\in M_{K}}[K_{v}:\mathbb{Q}_{v}]\log\left\|s_{1}\wedge\cdots\wedge s_{r}\right\|_{v}
=\displaystyle= log(#(E/𝒪Ks1++𝒪Ksr))12vMK,[Kv:v]logdet(si,sjv,1i,jr),\displaystyle\log\left(\#\left(E/\mathcal{O}_{K}s_{1}+\cdots+\mathcal{O}_{K}s_{r}\right)\right)-\frac{1}{2}\sum_{v\in M_{K,\infty}}[K_{v}:\mathbb{Q}_{v}]\log\det\left(\langle s_{i},s_{j}\rangle_{v,1\leqslant i,j\leqslant r}\right),

where s1srv\left\|s_{1}\wedge\cdots\wedge s_{r}\right\|_{v} follows the definition in [11, 2.1.9] for all vMK,v\in M_{K,\infty}, and si,sjv,1i,jr\langle s_{i},s_{j}\rangle_{v,1\leqslant i,j\leqslant r} is the Gram matrix of the basis {s1,,sr}\{s_{1},\ldots,s_{r}\} with respect to vMK,v\in M_{K,\infty}. 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 {s1,,sr}\{s_{1},\ldots,s_{r}\} by the product formula (cf. [33, Chap. III, Proposition 1.3]). In addition, we define

deg^n(E¯)=1[K:]deg^(E¯)\widehat{\deg}_{n}(\overline{E})=\frac{1}{[K:\mathbb{Q}]}\widehat{\deg}(\overline{E})

as the normalized Arakelov degree of E¯\overline{E}, which is independent of the choice of KK.

2.4.2 Slope

Let E¯\overline{E} be a non-zero Hermitian vector bundle over Spec𝒪K\operatorname{Spec}\mathcal{O}_{K}, and rk(E)\operatorname{rk}(E) be the rank of EE. The slope of E¯\overline{E} is defined as

μ^(E¯):=1rk(E)deg^n(E¯).\widehat{\mu}(\overline{E}):=\frac{1}{\operatorname{rk}(E)}\widehat{\deg}_{n}(\overline{E}).

In addition, we denote by μ^max(E¯)\widehat{\mu}_{\max}(\overline{E}) the maximal value of slopes of all non-zero Hermitian subbundles, and by μ^min(E¯)\widehat{\mu}_{\min}(\overline{E}) the minimal value of slopes of all non-zero Hermitian quotients bundles of E¯\overline{E}.

2.4.3 Height of linear maps

Let E¯\overline{E} and F¯\overline{F} be two non-zero Hermitian vector bundles over Spec𝒪K\operatorname{Spec}\mathcal{O}_{K}, and ϕ:EKFK\phi:\;E_{K}\rightarrow F_{K} be a non-zero homomorphism. The height of ϕ\phi is defined as

h(ϕ)=1[K:]vMKlogϕv,h(\phi)=\frac{1}{[K:\mathbb{Q}]}\sum_{v\in M_{K}}\log\|\phi\|_{v},

where ϕv\|\phi\|_{v} is the operator norm of ϕv:EKKvFKKv\phi_{v}:E\otimes_{K}K_{v}\rightarrow F\otimes_{K}K_{v} induced by the above linear homomorphism with respect to vMKv\in M_{K}.

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 ¯\overline{\mathcal{E}} be a Hermitian vector bundle of rank n+1n+1 over Spec𝒪K\operatorname{Spec}\mathcal{O}_{K}, and ()\mathbb{P}(\mathcal{E}) be the projective space which represents the functor from the category of commutative 𝒪K\mathcal{O}_{K}-algebras to the category of sets mapping all 𝒪K\mathcal{O}_{K}-algebra AA to the set of projective quotient AA-module of 𝒪KA\mathcal{E}\otimes_{\mathcal{O}_{K}}A of rank 11. Let 𝒪()(1)\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1) (or 𝒪(1)\mathcal{O}(1) if there is no confusion) be the universal bundle, and we denote by 𝒪()(D)\mathcal{O}_{\mathbb{P}(\mathcal{E})}(D) (or 𝒪(D)\mathcal{O}(D)) the line bundle 𝒪()(1)D\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)^{\otimes D} for simplicity. The Hermitian metrics on \mathcal{E} induce by quotient of Hermitian metrics (i.e. Fubini-Study metrics) on 𝒪()(1)\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1) which define a Hermitian line bundle 𝒪()(1)¯\overline{\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)} on ()\mathbb{P}(\mathcal{E}).

For every D+D\in\mathbb{N}^{+}, let

ED=H0((),𝒪()(D)),E_{D}=H^{0}\left(\mathbb{P}(\mathcal{E}),\mathcal{O}_{\mathbb{P}(\mathcal{E})}(D)\right), (7)

and let r(n,D)r(n,D) be its rank over 𝒪K\mathcal{O}_{K}. In fact, we have

r(n,D)=(n+DD).r(n,D)={n+D\choose D}. (8)

For each vMK,v\in M_{K,\infty}, we denote by .v,sup\|\raisebox{1.72218pt}{.}\|_{v,\sup} the norm over ED,v=ED𝒪K,vE_{D,v}=E_{D}\otimes_{\mathcal{O}_{K},v}\mathbb{C} such that

sED,v,sv,sup=supx(K)v()s(x)v,FS,\forall\;s\in E_{D,v},\;\|s\|_{v,\sup}=\sup_{x\in\mathbb{P}(\mathcal{E}_{K})_{v}(\mathbb{C})}\|s(x)\|_{v,\mathrm{FS}}, (9)

where .v,FS\|\raisebox{1.72218pt}{.}\|_{v,\mathrm{FS}} 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 CC, there exists the unique ellipsoid, called ellipsoid of John, contained in CC with the maximal volume.

For the 𝒪K\mathcal{O}_{K}-module EDE_{D} and any place vMK,v\in M_{K,\infty}, we take the ellipsoid of John of its unit closed ball defined via the norm.v,sup\|\raisebox{1.72218pt}{.}\|_{v,\sup}, and this ellipsoid induces a Hermitian norm, noted by .v,J\|\raisebox{1.72218pt}{.}\|_{v,J}. For every section sEDs\in E_{D}, the inequality

sv,supsv,Jr(n,D)sv,sup\|s\|_{v,\sup}\leqslant\|s\|_{v,J}\leqslant\sqrt{r(n,D)}\|s\|_{v,\sup} (10)

is verified by [45, Theorem 3.3.6].

2.5.2 Evaluation map

Let XX be an integral closed subscheme of (K)\mathbb{P}(\mathcal{E}_{K}), and 𝒳\mathscr{X} be the Zariski closure of XX in ()\mathbb{P}(\mathcal{E}). We denote by

ηX,D:ED,K=H0((K),𝒪(D))H0(X,𝒪(K)(1)|XD)\eta_{X,D}:\;E_{D,K}=H^{0}\left(\mathbb{P}(\mathcal{E}_{K}),\mathcal{O}(D)\right)\rightarrow H^{0}\left(X,\mathcal{O}_{\mathbb{P}(\mathcal{E}_{K})}(1)|_{X}^{\otimes D}\right) (11)

the evaluation map over XX induced by the closed immersion of XX in (K)\mathbb{P}(\mathcal{E}_{K}). We denote by FDF_{D} the largest saturated sub-𝒪K\mathcal{O}_{K}-module of H0(𝒳,𝒪()(1)|𝒳D)H^{0}\left(\mathscr{X},\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)|_{\mathscr{X}}^{\otimes D}\right) such that FD,K=Im(ηX,D)F_{D,K}=\operatorname{Im}(\eta_{X,D}). When the integer DD is large enough, the homomorphism ηX,D\eta_{X,D} is surjective, which means FD=H0(𝒳,𝒪()(1)|𝒳D)F_{D}=H^{0}(\mathscr{X},\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)|_{\mathscr{X}}^{\otimes D}).

The 𝒪K\mathcal{O}_{K}-module FDF_{D} is equipped with the quotient metrics (from E¯D\overline{E}_{D}) such that FDF_{D} is a Hermitian vector bundle over Spec𝒪K\operatorname{Spec}\mathcal{O}_{K}, noted by F¯D\overline{F}_{D} this Hermitian vector bundle.

\definame \the\smf@thm (Arithmetic Hilbert-Samuel function).

Let F¯D\overline{F}_{D} be the Hermitian vector bundle over Spec𝒪K\operatorname{Spec}\mathcal{O}_{K} defined above from the map (11). We say that the function which maps the positive integer DD to μ^(F¯D)\widehat{\mu}(\overline{F}_{D}) is the arithmetic Hilbert-Samuel function of XX with respect to the Hermitian line bundle 𝒪(1)¯\overline{\mathcal{O}(1)}.

2.6 Height of rational points

In this part, we will define a height function of rational points by Arakelov geometry.

Let ¯\overline{\mathcal{E}} be a Hermitian vector bundle of rank n+1n+1 over Spec𝒪K\operatorname{Spec}\mathcal{O}_{K}, P(K)(K)P\in\mathbb{P}(\mathcal{E}_{K})(K), and 𝒫()(𝒪K)\mathcal{P}\in\mathbb{P}(\mathcal{E})(\mathcal{O}_{K}) be the Zariski closure of PP in ()\mathbb{P}(\mathcal{E}). Let 𝒪()(1)¯\overline{\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)} be the universal bundle equipped with the corresponding Fubini-Study metric at each vMK,v\in M_{K,\infty}, then 𝒫𝒪()(1)¯\mathcal{P}^{*}\overline{\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)} is a Hermitian vector bundle over Spec𝒪K\operatorname{Spec}\mathcal{O}_{K}. We define the height of the rational point PP with respect to 𝒪()(1)¯\overline{\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)} as

h𝒪()(1)¯(P)=deg^n(𝒫𝒪()(1)¯).h_{\overline{\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)}}(P)=\widehat{\deg}_{n}\left(\mathcal{P}^{*}\overline{\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)}\right). (12)

We keep all the above notations. We choose

¯=(𝒪K(n+1),(.v)vMK,),\overline{\mathcal{E}}=\left(\mathcal{O}_{K}^{\oplus(n+1)},\left(\|\raisebox{1.72218pt}{.}\|_{v}\right)_{v\in M_{K,\infty}}\right), (13)

where for every vMK,v\in M_{K,\infty}, .v\|\raisebox{1.72218pt}{.}\|_{v} is the 2\ell^{2}-norm mapping (t0,,tn)(t_{0},\ldots,t_{n}) to |v(t0)|2++|v(tn)|2\sqrt{|v(t_{0})|^{2}+\cdots+|v(t_{n})|^{2}}. In this case, we use the notations Kn=(K)\mathbb{P}^{n}_{K}=\mathbb{P}(\mathcal{E}_{K}) and 𝒪Kn=()\mathbb{P}^{n}_{\mathcal{O}_{K}}=\mathbb{P}(\mathcal{E}) for simplicity. We suppose that PP has a KK-rational projective coordinate [x0::xn][x_{0}:\cdots:x_{n}], then we have (cf. [31, Proposition 9.10])

h𝒪()(1)¯(P)\displaystyle h_{\overline{\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)}}(P) =\displaystyle= vMK,f[Kv:v][K:]log(max1in|xi|v)\displaystyle\sum\limits_{v\in M_{K,f}}\frac{[K_{v}:\mathbb{Q}_{v}]}{[K:\mathbb{Q}]}\log\left(\max\limits_{1\leqslant i\leqslant n}|x_{i}|_{v}\right)
+12vMK,[Kv:v][K:]log(j=0n|v(xj)|2).\displaystyle\;\;+\frac{1}{2}\sum\limits_{v\in M_{K,\infty}}\frac{[K_{v}:\mathbb{Q}_{v}]}{[K:\mathbb{Q}]}\log\left(\sum\limits_{j=0}^{n}|v(x_{j})|^{2}\right).
\remaname \the\smf@thm.

We compare the logarithmic height h(.)h(\raisebox{1.72218pt}{.}) defined in (2) and the height h𝒪()(1)¯(.)h_{\overline{\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)}}(\raisebox{1.72218pt}{.}) defined in (12) by Arakelov geometry, where ¯\overline{\mathcal{E}} is defined in (13). In fact, by an elementary calculation, the inequality

|h(P)h𝒪()(1)¯(P)|12log(n+1)\left|h(P)-h_{\overline{\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)}}(P)\right|\leqslant\frac{1}{2}\log(n+1)

is uniformly verified for all PKn(K)P\in\mathbb{P}^{n}_{K}(K).

Let ψ:XKn\psi:X\hookrightarrow\mathbb{P}^{n}_{K} be a projective scheme, and PX(K)P\in X(K). We define the height of PP as h𝒪()(1)¯(ψ(P))h_{\overline{\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)}}(\psi(P)). We will just use the notation h𝒪()(1)¯(P)h_{\overline{\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)}}(P) or h(P)h(P) if there is no confusion of the morphism ψ\psi and the Hermitian line bundle 𝒪()(1)¯\overline{\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)}.

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 KK be a number field, 𝒪K\mathcal{O}_{K} be its ring of integers, ¯\overline{\mathcal{E}} be a Hermitian vector bundle of rank n+1n+1 over Spec𝒪K\operatorname{Spec}\mathcal{O}_{K}, and ¯\overline{\mathcal{L}} be a Hermitian line bundle over ()\mathbb{P}(\mathcal{E}). Let XX be a pure dimensional closed subscheme of (K)\mathbb{P}(\mathcal{E}_{K}) of dimension dd, and 𝒳\mathscr{X} be the Zariski closure of XX in ()\mathbb{P}(\mathcal{E}). The Arakelov height of XX is defined as the arithmetic intersection number

1[K:]deg^(c^1(¯)d+1[𝒳]),\frac{1}{[K:\mathbb{Q}]}\widehat{\deg}\left(\widehat{c}_{1}(\overline{\mathcal{L}})^{d+1}\cdot[\mathscr{X}]\right),

where c^1(¯)\widehat{c}_{1}(\overline{\mathcal{L}}) is the arithmetic first Chern class of ¯\overline{\mathcal{L}} (cf. [43, Chap. III.4, Proposition 1] for its definition). This height is noted by h¯(X)h_{\overline{\mathcal{L}}}(X) or h¯(𝒳)h_{\overline{\mathcal{L}}}(\mathscr{X}).

\remaname \the\smf@thm.

With all the notations in Definition 2.5.2 and Definition 2.7.1. By [36, Théorème A], we have

h𝒪(1)¯(X)=limD+deg^n(F¯D)Dd+1/(d+1)!.h_{\overline{\mathcal{O}(1)}}(X)=\lim_{D\rightarrow+\infty}\frac{\widehat{\deg}_{n}(\overline{F}_{D})}{D^{d+1}/(d+1)!}.

2.7.2 Heights of hypersurfaces

Let XX be a hypersurface in Kn\mathbb{P}^{n}_{K} of degree δ\delta. By [23, Proposition 7.6 (d), Chap. I], XX is define by a homogeneous polynomial of degree δ\delta. We define a height function of hypersurfaces by considering its polynomial of definition.

\definame \the\smf@thm (Naive height).

Let

f(T0,,Tn)=(i0,,in)n+1ai0,,inT0i0TninK[T0,,Tn].f(T_{0},\ldots,T_{n})=\sum\limits_{(i_{0},\ldots,i_{n})\in\mathbb{N}^{n+1}}a_{i_{0},\ldots,i_{n}}T_{0}^{i_{0}}\cdots T_{n}^{i_{n}}\in K[T_{0},\ldots,T_{n}].

We define the naive height of f(T0,,Tn)f(T_{0},\ldots,T_{n}) as

HK(f)=vMKmax(i0,,in)n+1{|ai0,,in|v}[Kv:v],H_{K}(f)=\prod_{v\in M_{K}}\max\limits_{(i_{0},\ldots,i_{n})\in\mathbb{N}^{n+1}}\left\{|a_{i_{0},\ldots,i_{n}}|_{v}\right\}^{[K_{v}:\mathbb{Q}_{v}]},

and

h(f)=1[K:]logHK(f).h(f)=\frac{1}{[K:\mathbb{Q}]}\log H_{K}(f).

In addition, if f(T0,,Tn)f(T_{0},\ldots,T_{n}) is homogeneous and defines the hypersurface XKnX\hookrightarrow\mathbb{P}^{n}_{K}, we define the naive height of XX as

HK(X)=HK(f) and h(X)=h(f).H_{K}(X)=H_{K}(f)\hbox{ and }h(X)=h(f).

2.7.3 Comparison of height functions

In order to compare h𝒪(1)¯(𝒳)h_{\overline{\mathcal{O}(1)}}(\mathscr{X}) and h(X)h(X) for a hypersurface XX, we refer the following result in [28].

\propname \the\smf@thm.

Let XX be a hypersurface in Kn\mathbb{P}^{n}_{K} of degree δ\delta. With all the notations above, we have

δ(12δlog((n+1)(δ+1))+12n)\displaystyle-\delta\left(\frac{1}{2\delta}\log\left((n+1)(\delta+1)\right)+\frac{1}{2}\mathcal{H}_{n}\right) \displaystyle\leqslant h(X)h𝒪(1)¯(𝒳)\displaystyle h(X)-h_{\overline{\mathcal{O}(1)}}(\mathscr{X})
\displaystyle\leqslant δ(log2+5log(n+1)12n),\displaystyle\delta\left(\log 2+5\log(n+1)-\frac{1}{2}\mathcal{H}_{n}\right),

where n=1++1n\mathcal{H}_{n}=1+\cdots+\frac{1}{n}.

Démonstration.

Since XX is a hypersurface, the Chow variety of XX is just XX 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 XX is an integral hypersurface in Kn\mathbb{P}^{n}_{K} , and 𝒳\mathscr{X} is its Zariski closure in 𝒪Kn\mathbb{P}^{n}_{\mathcal{O}_{K}}. In fact, XSpecKX\rightarrow\operatorname{Spec}K is the generic fiber of 𝒳Spec𝒪K\mathscr{X}\rightarrow\operatorname{Spec}\mathcal{O}_{K}. When we consider the height h(P)h(P) of a rational point PX(K)P\in X(K) embedded into Kn\mathbb{P}^{n}_{K}, we use the definition in (12) by Arakelov geometry. Let 𝔭\mathfrak{p} be a maximal ideal of 𝒪K\mathcal{O}_{K}, and we denote by 𝒳𝔭=𝒳×Spec𝒪KSpec𝔽𝔭𝔽𝔭\mathscr{X}_{\mathfrak{p}}=\mathscr{X}\times_{\operatorname{Spec}\mathcal{O}_{K}}\operatorname{Spec}\mathbb{F}_{\mathfrak{p}}\rightarrow\mathbb{F}_{\mathfrak{p}} the fiber at 𝔭\mathfrak{p}.

Let r1(n,D)r_{1}(n,D) be the rank of FDF_{D} over 𝒪K\mathcal{O}_{K}, where FDF_{D} is defined in §2.5. For the case where XX is a hypersurface of degree δ\delta in Kn\mathbb{P}^{n}_{K}, we have

r1(n,D)=(n+Dn)(n+Dδn).r_{1}(n,D)={n+D\choose n}-{n+D-\delta\choose n}.

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 XX be a closed integral subscheme in kn\mathbb{P}^{n}_{k}, and 𝒳\mathscr{X} be its Zariski closure in 𝒪Kn\mathbb{P}^{n}_{\mathcal{O}_{K}}. Let {𝔭j}jJ\{\mathfrak{p}_{j}\}_{j\in J} be a finite family of maximal ideals of 𝒪K\mathcal{O}_{K}, and {Pi}iI\{P_{i}\}_{i\in I} be a family of rational points of XX. For a fixed prime ideal 𝔭\mathfrak{p} of 𝒪K\mathcal{O}_{K}, let μξ(𝒳𝔭)\mu_{\xi}(\mathscr{X}_{\mathfrak{p}}) be the multiplicity of the point ξ\xi in 𝒳𝔭\mathscr{X}_{\mathfrak{p}}, and we denote n(𝒳𝔭)=ξ𝒳(𝔽𝔭)μξ(𝒳𝔭)n(\mathscr{X}_{\mathfrak{p}})=\sum\limits_{\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}})}\mu_{\xi}(\mathscr{X}_{\mathfrak{p}}). If the inequality

supiIh(Pi)<μ^(F¯D)Dlogr1(n,D)2D\displaystyle\;\;\sup_{i\in I}h(P_{i})<\frac{\widehat{\mu}(\overline{F}_{D})}{D}-\frac{\log r_{1}(n,D)}{2D} (14)
+1[K:]jJ((n1)!1n1(n1)r1(n,D)1n1nDn(𝒳𝔭j)1n1n3+2n2+n42Dn(n+1))logN(𝔭j)\displaystyle+\frac{1}{[K:\mathbb{Q}]}\sum_{j\in J}\left(\frac{(n-1)!^{\frac{1}{n-1}}(n-1)r_{1}(n,D)^{\frac{1}{n-1}}}{nDn(\mathscr{X}_{\mathfrak{p}_{j}})^{\frac{1}{n-1}}}-\frac{n^{3}+2n^{2}+n-4}{2Dn(n+1)}\right)\log N(\mathfrak{p}_{j})

is verified, then there exists a section sED,Ks\in E_{D,K}, which contains {Pi}iI\{P_{i}\}_{i\in I} but does not contain the generic point of XX. In other words, {Pi}iI\{P_{i}\}_{i\in I} can be covered by a hypersurfaces of degree DD which does not contain the generic point of XX.

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 E¯\overline{E} be a Hermitian vector bundle of rank r>0r>0 over Spec𝒪K\operatorname{Spec}\mathcal{O}_{K}, and {L¯i}iI\{\overline{L}_{i}\}_{i\in I} be a family of Hermitian line bundles over Spec𝒪K\operatorname{Spec}\mathcal{O}_{K}. If

ϕ:EKiILi,K\phi:\;E_{K}\rightarrow\bigoplus\limits_{i\in I}L_{i,K}

is an injective homomorphism, then there exists a subset I0I_{0} of II whose cardinality is rr such that the following equality

μ^(E¯)=1r(iI0μ^(L¯i)+h(r(prI0ϕ)))\widehat{\mu}(\overline{E})=\frac{1}{r}\left(\sum_{i\in I_{0}}\widehat{\mu}(\overline{L}_{i})+h\left(\wedge^{r}\left(\operatorname{pr}_{I_{0}}\circ\phi\right)\right)\right)

is verified, where prI0:iILi,KiI0Li,K\operatorname{pr}_{I_{0}}:\;\bigoplus\limits_{i\in I}L_{i,K}\rightarrow\bigoplus\limits_{i\in I_{0}}L_{i,K} 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 XX be an integral closed subscheme of Kn\mathbb{P}^{n}_{K} and 𝒳\mathscr{X} be the Zariski closure of XX in 𝒪Kn\mathbb{P}^{n}_{\mathcal{O}_{K}}. Let 𝔭\mathfrak{p} be a maximal ideal of 𝒪K\mathcal{O}_{K} and ξ𝒳(𝔽𝔭)\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}}). In this case, 𝒪𝒳,ξ\mathcal{O}_{\mathscr{X},\xi} is a local algebra over 𝒪K,𝔭\mathcal{O}_{K,\mathfrak{p}}. Let (fi)1im(f_{i})_{1\leqslant i\leqslant m} be a family of local homomorphisms of 𝒪K,𝔭\mathcal{O}_{K,\mathfrak{p}}-algebras from 𝒪𝒳,ξ\mathcal{O}_{\mathscr{X},\xi} to 𝒪K,𝔭\mathcal{O}_{K,\mathfrak{p}}.

Let EE be a free sub-𝒪K,𝔭\mathcal{O}_{K,\mathfrak{p}}-module of finite type of 𝒪𝒳,ξ\mathcal{O}_{\mathscr{X},\xi} and let ff be the 𝒪K,𝔭\mathcal{O}_{K,\mathfrak{p}}-linear homomorphism

(fi|E)1im:E𝒪K,𝔭m.(f_{i}|_{E})_{1\leqslant i\leqslant m}:E\rightarrow\mathcal{O}_{K,\mathfrak{p}}^{\oplus m}.

Since f1f_{1} is a local homomorphism of 𝒪K,𝔭\mathcal{O}_{K,\mathfrak{p}}-algebras, it must be surjective. Let 𝔞\mathfrak{a} be the kernel of f1f_{1}, then we have 𝒪𝒳,ξ/𝔞𝒪K,𝔭\mathcal{O}_{\mathscr{X},\xi}/\mathfrak{a}\cong\mathcal{O}_{K,\mathfrak{p}}. Furthermore, since 𝒪𝒳,ξ\mathcal{O}_{\mathscr{X},\xi} is a local ring and we suppose that 𝔪ξ\mathfrak{m}_{\xi} is its maximal ideal, then we have 𝔪ξ𝔞\mathfrak{m}_{\xi}\supseteq\mathfrak{a}. Moreover, since f1f_{1} is a local homomorphism, we have 𝔞+𝔭𝒪𝒳,ξ=𝔪ξ\mathfrak{a}+\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}=\mathfrak{m}_{\xi}. For each jj\in\mathbb{N}, 𝔞j/𝔞j+1\mathfrak{a}^{j}/\mathfrak{a}^{j+1} is an 𝒪𝒳,ξ/𝔞𝒪K,𝔭\mathcal{O}_{\mathscr{X},\xi}/\mathfrak{a}\cong\mathcal{O}_{K,\mathfrak{p}}-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

𝔽𝔭𝒪K,𝔭(𝔞j/𝔞j+1)\displaystyle\mathbb{F}_{\mathfrak{p}}\otimes_{\mathcal{O}_{K,\mathfrak{p}}}(\mathfrak{a}^{j}/\mathfrak{a}^{j+1}) \displaystyle\cong (𝔞/𝔞𝔭𝒪𝒳,ξ)j/(𝔞/𝔞𝔭𝒪𝒳,ξ)j+1\displaystyle\left(\mathfrak{a}/\mathfrak{a}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}\right)^{j}/\left(\mathfrak{a}/\mathfrak{a}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}\right)^{j+1}
\displaystyle\cong (𝔪ξ/𝔪ξ𝔭𝒪𝒳,ξ)j/(𝔪ξ/𝔪ξ𝔭𝒪𝒳,ξ)j+1.\displaystyle\left(\mathfrak{m}_{\xi}/\mathfrak{m}_{\xi}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}\right)^{j}/\left(\mathfrak{m}_{\xi}/\mathfrak{m}_{\xi}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}\right)^{j+1}.
Démonstration.

By definition, we have

𝔽𝔭𝒪K,𝔭(𝔞j/𝔞j+1)(𝔞/𝔞𝔭𝒪𝒳,ξ)j/(𝔞/𝔞𝔭𝒪𝒳,ξ)j+1.\mathbb{F}_{\mathfrak{p}}\otimes_{\mathcal{O}_{K,\mathfrak{p}}}(\mathfrak{a}^{j}/\mathfrak{a}^{j+1})\cong\left(\mathfrak{a}/\mathfrak{a}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}\right)^{j}/\left(\mathfrak{a}/\mathfrak{a}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}\right)^{j+1}.

Next, from the facts 𝔞+𝔭𝒪𝒳,ξ=𝔪ξ\mathfrak{a}+\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}=\mathfrak{m}_{\xi} and 𝔞𝔪ξ\mathfrak{a}\subseteq\mathfrak{m}_{\xi}, we claim that

𝔞+𝔪ξ𝔭𝒪𝒳,ξ=𝔪ξ(𝔞+𝔭𝒪𝒳,ξ)=𝔪ξ\mathfrak{a}+\mathfrak{m}_{\xi}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}=\mathfrak{m}_{\xi}\cap(\mathfrak{a}+\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi})=\mathfrak{m}_{\xi}

is verified. In fact, for every x𝔪ξ=𝔪ξ(𝔞+𝔭𝒪𝒳,ξ)x\in\mathfrak{m}_{\xi}=\mathfrak{m}_{\xi}\cap(\mathfrak{a}+\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}), there exist v𝔞v\in\mathfrak{a} and w𝔭𝒪𝒳,ξw\in\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}, such that x=v+wx=v+w. Then w=xv𝔪ξw=x-v\in\mathfrak{m}_{\xi}. Since v,x𝔪ξv,x\in\mathfrak{m}_{\xi}, then w𝔪ξ𝔭𝒪𝒳,ξw\in\mathfrak{m}_{\xi}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}. So we have x𝔞+𝔪ξ𝔭𝒪𝒳,ξx\in\mathfrak{a}+\mathfrak{m}_{\xi}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}. Conversely, since 𝔪ξ\mathfrak{m}_{\xi} is the maximal ideal, then 𝔞+𝔪ξ𝔭𝒪𝒳,ξ𝔪ξ\mathfrak{a}+\mathfrak{m}_{\xi}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}\subseteq\mathfrak{m}_{\xi}.

By the above fact, we have

𝔞/𝔞𝔭𝒪𝒳,ξ𝔞/𝔞(𝔪ξ𝔭𝒪𝒳,ξ)(𝔞+𝔪ξ𝔭𝒪𝒳,ξ)/𝔪ξ𝔭𝒪𝒳,ξ𝔪ξ/𝔪ξ𝔭𝒪𝒳,ξ,\mathfrak{a}/\mathfrak{a}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}\cong\mathfrak{a}/\mathfrak{a}\cap(\mathfrak{m}_{\xi}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi})\cong(\mathfrak{a}+\mathfrak{m}_{\xi}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi})/\mathfrak{m}_{\xi}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}\cong\mathfrak{m}_{\xi}/\mathfrak{m}_{\xi}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi},

which terminates the proof. ∎

By Nakayama’s lemma (cf. [30, Theorem 2.2]), we deduce that the rank of 𝔞j/𝔞j+1\mathfrak{a}^{j}/\mathfrak{a}^{j+1} over OK,𝔭O_{K,\mathfrak{p}} is equal to the rank of (𝔪ξ/𝔪ξ𝔭𝒪𝒳,ξ)j/(𝔪ξ/𝔪ξ𝔭𝒪𝒳,ξ)j+1\left(\mathfrak{m}_{\xi}/\mathfrak{m}_{\xi}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}\right)^{j}/\left(\mathfrak{m}_{\xi}/\mathfrak{m}_{\xi}\cap\mathfrak{p}\mathcal{O}_{\mathscr{X},\xi}\right)^{j+1} over 𝔽𝔭\mathbb{F}_{\mathfrak{p}} from the isomorphism in Lemma 3.1, which is the value of the local Hilbert-Samuel function Hξ(j)H_{\xi}(j) defined in (3).

By this fact, we consider the filtration

𝒪𝒳,ξ=𝔞0𝔞𝔞2𝔞j𝔞j+1\mathcal{O}_{\mathscr{X},\xi}=\mathfrak{a}^{0}\supseteq\mathfrak{a}\supseteq\mathfrak{a}^{2}\supseteq\cdots\supseteq\mathfrak{a}^{j}\supseteq\mathfrak{a}^{j+1}\supseteq\cdots

of 𝒪𝒳,ξ\mathcal{O}_{\mathscr{X},\xi}, which induces the filtration

:E=E𝔞0E𝔞E𝔞2E𝔞jE𝔞j+1\mathcal{F}:\;E=E\cap\mathfrak{a}^{0}\supseteq E\cap\mathfrak{a}\supseteq E\cap\mathfrak{a}^{2}\supseteq\cdots\supseteq E\cap\mathfrak{a}^{j}\supseteq E\cap\mathfrak{a}^{j+1}\supseteq\cdots (15)

of EE whose jj-th sub-quotient E𝔞j/E𝔞j+1E\cap\mathfrak{a}^{j}/E\cap\mathfrak{a}^{j+1} is a free 𝒪K,𝔭\mathcal{O}_{K,\mathfrak{p}}-module of rank smaller than or equal to Hξ(j)H_{\xi}(j).

We suppose that the reductions of all the above local homomorphisms f1,,fmf_{1},\ldots,f_{m} modulo 𝔭\mathfrak{p} are same, which means all the composed homomorphisms

𝒪𝒳,ξfi𝒪K,𝔭𝔽𝔭\mathcal{O}_{\mathscr{X},\xi}\stackrel{{\scriptstyle f_{i}}}{{\longrightarrow}}\mathcal{O}_{K,\mathfrak{p}}\rightarrow\mathbb{F}_{\mathfrak{p}}

are the same for every i=1,,mi=1,\ldots,m, where the last arrow is the canonical reduction morphism modulo 𝔭\mathfrak{p}. We note N(𝔭)=#𝔽𝔭N(\mathfrak{p})=\#\mathbb{F}_{\mathfrak{p}}. In this case, the restriction of ff on E𝔞jE\cap\mathfrak{a}^{j} has its norm smaller than N(𝔭)jN(\mathfrak{p})^{-j}. In fact, for any 1im1\leqslant i\leqslant m, we have fi(𝔞)𝔭𝒪K,𝔭f_{i}(\mathfrak{a})\subseteq\mathfrak{p}\mathcal{O}_{K,\mathfrak{p}} and hence we have fi(𝔞j)𝔭j𝒪K,𝔭f_{i}(\mathfrak{a}^{j})\subseteq\mathfrak{p}^{j}\mathcal{O}_{K,\mathfrak{p}}.

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 𝔭\mathfrak{p} be a maximal ideal of 𝒪K\mathcal{O}_{K} and ξ𝒳(𝔽𝔭)\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}}). Suppose that {fi}1im\{f_{i}\}_{1\leqslant i\leqslant m} is a family of local 𝒪K,𝔭\mathcal{O}_{K,\mathfrak{p}}-linear homomorphism from 𝒪𝒳,ξ\mathcal{O}_{\mathscr{X},\xi} to 𝒪K,𝔭\mathcal{O}_{K,\mathfrak{p}} whose reduction modulo 𝔭\mathfrak{p} are the same. Let EE be a free sub-𝒪K,𝔭\mathcal{O}_{K,\mathfrak{p}}-module of finite type of 𝒪𝒳,ξ\mathcal{O}_{\mathscr{X},\xi} and f=(fi|E)1imf=(f_{i}|_{E})_{1\leqslant i\leqslant m}. Then for any integer r1r\geqslant 1, we have

rfKN(𝔭)Qξ(r),\|\wedge^{r}f_{K}\|\leqslant N(\mathfrak{p})^{-Q_{\xi}(r)}, (16)

where N(𝔭)=#(𝒪K/𝔭)N(\mathfrak{p})=\#(\mathcal{O}_{K}/\mathfrak{p}), and Qξ(r)Q_{\xi}(r) is defined in (5).

The following lemma will be used in the global determinant estimate.

\lemmname \the\smf@thm.

Let (K,|.|)(K,|\raisebox{1.72218pt}{.}|) be a normed field, E1,E2,F1,F2E_{1},E_{2},F_{1},F_{2} be four normed vector spaces over KK, f1:E1F1f_{1}:E_{1}\rightarrow F_{1} and f2:E2F2f_{2}:E_{2}\rightarrow F_{2} be two KK-linear isomorphisms. Suppose dimK(E1)=dimK(F1)=r1\dim_{K}(E_{1})=\dim_{K}(F_{1})=r_{1} and dimK(E2)=dimK(F2)=r2\dim_{K}(E_{2})=\dim_{K}(F_{2})=r_{2}. We equipped

f1f2:E1E2F2F2f_{1}\oplus f_{2}:E_{1}\oplus E_{2}\rightarrow F_{2}\oplus F_{2}

with the corresponding maximal value norms. Then we have

r1+r2(f1f2)=r1f1r2f2,\left\|\wedge^{r_{1}+r_{2}}\left(f_{1}\oplus f_{2}\right)\right\|=\|\wedge^{r_{1}}f_{1}\|\cdot\|\wedge^{r_{2}}f_{2}\|,

where the above .\|\raisebox{1.72218pt}{.}\| is the norm of operators.

Démonstration.

By definition, the linear maps r1f1\wedge^{r_{1}}f_{1} and r2f2\wedge^{r_{2}}f_{2} are both scalar products by the corresponding determinants, and r1+r2(f1f2)\wedge^{r_{1}+r_{2}}\left(f_{1}\oplus f_{2}\right) is the scalar product of the above two determinants. Then we have the result by definition directly. ∎

3.2 Proof of Theorem 3

We are going to prove Theorem 3, and some ideas of the proof below are inspired from [13, §3].

Proof of Theorem 3.

Let DD be an integer larger than 11. We suppose that the global section predicted by Theorem 3 does not exist. Then the evaluation map

f:FD,KiIPi𝒪Kn(D)f:\;F_{D,K}\rightarrow\bigoplus_{i\in I}P_{i}^{*}\mathcal{O}_{\mathbb{P}^{n}_{K}}(D)

is injective. We can replace II by one of its subsets such that ff is an isomorphism. From now on, we suppose ff is isomorphic, which means #I=r1(n,D)=rkFD,K\#I=r_{1}(n,D)=\operatorname{rk}F_{D,K}. Then by Proposition 3.1, we have

μ^(F¯D)=1r1(n,D)(DiIh(Pi)+h(r1(n,D)f)),\widehat{\mu}(\overline{F}_{D})=\frac{1}{r_{1}(n,D)}\left(D\sum_{i\in I}h(P_{i})+h\left(\wedge^{r_{1}(n,D)}f\right)\right),

which implies

μ^(F¯D)DsupiIh(Pi)+1Dr1(n,D)h(r1(n,D)f).\frac{\widehat{\mu}(\overline{F}_{D})}{D}\leqslant\sup_{i\in I}h(P_{i})+\frac{1}{Dr_{1}(n,D)}h\left(\wedge^{r_{1}(n,D)}f\right).

Now we estimate the height of r1(n,D)f\wedge^{r_{1}(n,D)}f. For every vMK,v\in M_{K,\infty}, we have

1r1(n,D)logr1(n,D)fvlogfvlogr1(n,D),\frac{1}{r_{1}(n,D)}\log\|\wedge^{r_{1}(n,D)}f\|_{v}\leqslant\log\|f\|_{v}\leqslant\log\sqrt{r_{1}(n,D)},

where the second inequality comes from the definition of metrics of John in §2.5.1.

Now we consider the case of vMK,fv\in M_{K,f}. The homomorphism ff is induced by a homomorphism of 𝒪K\mathcal{O}_{K}-module

FDiI𝒫i𝒪𝒪Kn(D),F_{D}\rightarrow\bigoplus\limits_{i\in I}\mathcal{P}_{i}^{*}\mathcal{O}_{\mathbb{P}^{n}_{\mathcal{O}_{K}}}(D),

where 𝒫i\mathcal{P}_{i} is the Zariski closure of PiP_{i} in 𝒳\mathscr{X} for each iIi\in I. Then for every vMK,fv\in M_{K,f}, we have logr1(n,D)fv0\log\|\wedge^{r_{1}(n,D)}f\|_{v}\leqslant 0.

We fix a maximal ideal 𝔭\mathfrak{p} of 𝒪K\mathcal{O}_{K} corresponding to vMK,fv\in M_{K,f}, and decompose the set {𝒫i}iI\{\mathcal{P}_{i}\}_{i\in I} as the disjoint union

{𝒫i}iI=ξ𝒳(𝔽𝔭){𝒫l,ξ}l=1mξ,\{\mathcal{P}_{i}\}_{i\in I}=\bigcup_{\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}})}\{\mathcal{P}_{l,\xi}\}_{l=1}^{m_{\xi}},

where all elements in {𝒫l,ξ}l=1mξ\{\mathcal{P}_{l,\xi}\}_{l=1}^{m_{\xi}} modulo 𝔭\mathfrak{p} are the same point ξ𝒳(𝔽𝔭)\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}}). If {𝒫l,ξ}l=1mξ\{\mathcal{P}_{l,\xi}\}_{l=1}^{m_{\xi}} is empty for some ξ𝒳(𝔽𝔭)\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}}), we define mξ=0m_{\xi}=0 for simplicity. With the above notations, let

ξ𝒳(𝔽𝔭){sl,ξ}l=1mξ\bigcup_{\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}})}\{s_{l,\xi}\}_{l=1}^{m_{\xi}}

be an 𝒪K,𝔭\mathcal{O}_{K,\mathfrak{p}}-basis of FD,𝔭F_{D,\mathfrak{p}} such that f(sl,ξ)f(s_{l,\xi}) generates 𝒫l,ξ𝒪𝒪Kn(D)\mathcal{P}_{l,\xi}^{*}\mathcal{O}_{\mathbb{P}^{n}_{\mathcal{O}_{K}}}(D) for all l=1,,mξl=1,\ldots,m_{\xi} and ξ𝒳(𝔽𝔭)\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}}). Since 𝒪K,𝔭\mathcal{O}_{K,\mathfrak{p}} is a local ring, the 𝒪K,𝔭\mathcal{O}_{K,\mathfrak{p}}-module FD,𝔭F_{D,\mathfrak{p}} is free, then there exists such a basis for a fixed maximal ideal 𝔭\mathfrak{p}. We denote by FD,ξF_{D,\xi} the sub-𝒪K,𝔭\mathcal{O}_{K,\mathfrak{p}}-module of FD,𝔭F_{D,\mathfrak{p}} generated by {sl,ξ}l=1mξ\{s_{l,\xi}\}_{l=1}^{m_{\xi}}.

By Proposition 3.1, we have

logrk(FD,ξ)f|FD,ξ𝔭Qξ(rk(FD,ξ))logN(𝔭).\log\left\|\wedge^{\operatorname{rk}(F_{D,\xi})}f|_{F_{D,\xi}}\right\|_{\mathfrak{p}}\leqslant-Q_{\xi}(\operatorname{rk}(F_{D,\xi}))\log N(\mathfrak{p}).

By definition, we have

FD,𝔭=ξ𝒳(𝔽𝔭)FD,ξ,F_{D,\mathfrak{p}}=\bigoplus_{\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}})}F_{D,\xi},

and

r1(n,D)=ξ𝒳(𝔽𝔭)rk(FD,ξ).r_{1}(n,D)=\sum_{\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}})}\operatorname{rk}(F_{D,\xi}). (17)

Then from the above construction, by applying Lemma 3.1 and Proposition 3.1 respectively, we obtain

logr1(n,D)f𝔭=ξ𝒳(𝔽𝔭)logrk(FD,ξ)f|FD,ξ𝔭ξ𝒳(𝔽𝔭)Qξ(rk(FD,ξ))logN(𝔭).\log\left\|\wedge^{r_{1}(n,D)}f\right\|_{\mathfrak{p}}=\sum_{\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}})}\log\left\|\wedge^{\operatorname{rk}(F_{D,\xi})}f|_{F_{D,\xi}}\right\|_{\mathfrak{p}}\leqslant-\sum\limits_{\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}})}Q_{\xi}\left(\operatorname{rk}(F_{D,\xi})\right)\log N(\mathfrak{p}).

In order to estimate the term

1r1(n,D)ξ𝒳(𝔽𝔭)Qξ(rk(FD,ξ)),\frac{1}{r_{1}(n,D)}\sum\limits_{\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}})}Q_{\xi}\left(\operatorname{rk}(F_{D,\xi})\right),

by (6), we have

ξ𝒳(𝔽𝔭)Qξ(rk(FD,ξ))\displaystyle\sum\limits_{\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}})}Q_{\xi}\left(\operatorname{rk}(F_{D,\xi})\right)
>\displaystyle> ξ𝒳(𝔽𝔭)(((n1)!μξ(𝒳𝔭))1n1(n1n)rk(FD,ξ)nn1n3+2n2+n42n(n+1)rk(FD,ξ))\displaystyle\sum_{\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}})}\left(\left(\frac{(n-1)!}{\mu_{\xi}(\mathscr{X}_{\mathfrak{p}})}\right)^{\frac{1}{n-1}}\left(\frac{n-1}{n}\right)\operatorname{rk}(F_{D,\xi})^{\frac{n}{n-1}}-\frac{n^{3}+2n^{2}+n-4}{2n(n+1)}\operatorname{rk}(F_{D,\xi})\right)
=\displaystyle= (n1)!1n1(n1)nξ𝒳(𝔽𝔭)rk(FD,ξ)nn1μξ(𝒳𝔭)1n1n3+2n2+n42n(n+1)r1(n,D).\displaystyle\frac{(n-1)!^{\frac{1}{n-1}}(n-1)}{n}\sum_{\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}})}\frac{\operatorname{rk}(F_{D,\xi})^{\frac{n}{n-1}}}{\mu_{\xi}(\mathscr{X}_{\mathfrak{p}})^{\frac{1}{n-1}}}-\frac{n^{3}+2n^{2}+n-4}{2n(n+1)}r_{1}(n,D).

By (17) and Hölder’s inequality, we have

ξ𝒳(𝔽𝔭)rk(FD,ξ)nn1μξ(𝒳𝔭)1n1(ξ𝒳(𝔽𝔭)rk(FD,ξ))nn1(ξ𝒳(𝔽𝔭)μξ(𝒳𝔭))1n1=r1(n,D)nn1n(𝒳𝔭)1n1,\sum_{\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}})}\frac{\operatorname{rk}(F_{D,\xi})^{\frac{n}{n-1}}}{\mu_{\xi}(\mathscr{X}_{\mathfrak{p}})^{\frac{1}{n-1}}}\geqslant\frac{\left(\sum\limits_{\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}})}\operatorname{rk}(F_{D,\xi})\right)^{\frac{n}{n-1}}}{\left(\sum\limits_{\xi\in\mathscr{X}(\mathbb{F}_{\mathfrak{p}})}\mu_{\xi}(\mathscr{X}_{\mathfrak{p}})\right)^{\frac{1}{n-1}}}=\frac{r_{1}(n,D)^{\frac{n}{n-1}}}{n(\mathscr{X}_{\mathfrak{p}})^{\frac{1}{n-1}}},

where n(𝒳𝔭)n(\mathscr{X}_{\mathfrak{p}}) is defined in the statement of Theorem 3. Then we obtain the inequality

μ^(F¯D)DsupiIh(Pi)+logr1(n,D)2D\displaystyle\;\;\frac{\widehat{\mu}(\overline{F}_{D})}{D}\leqslant\sup_{i\in I}h(P_{i})+\frac{\log r_{1}(n,D)}{2D}
1[K:]jJ((n1)!1n1(n1)r1(n,D)1n1nDn(𝒳𝔭j)1n1n3+2n2+n42Dn(n+1))logN(𝔭j),\displaystyle-\frac{1}{[K:\mathbb{Q}]}\sum_{j\in J}\left(\frac{(n-1)!^{\frac{1}{n-1}}(n-1)r_{1}(n,D)^{\frac{1}{n-1}}}{nDn(\mathscr{X}_{\mathfrak{p}_{j}})^{\frac{1}{n-1}}}-\frac{n^{3}+2n^{2}+n-4}{2Dn(n+1)}\right)\log N(\mathfrak{p}_{j}),

which leads to a contradiction. ∎

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 n(𝒳𝔭)n(\mathscr{X}_{\mathfrak{p}}) in it. For this target, we need to have a control of the reduction type of 𝒳𝒪KnSpec𝒪K\mathscr{X}\hookrightarrow\mathbb{P}^{n}_{\mathcal{O}_{K}}\rightarrow\operatorname{Spec}\mathcal{O}_{K}, an upper bound of n(𝒳𝔭)n(\mathscr{X}_{\mathfrak{p}}) when 𝒳𝔭Spec𝔽𝔭\mathscr{X}_{\mathfrak{p}}\rightarrow\operatorname{Spec}\mathbb{F}_{\mathfrak{p}} is geometrically integral, and a distribution of certain prime ideals of 𝒪K\mathcal{O}_{K}. 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 XKnX\hookrightarrow\mathbb{P}^{n}_{K} be a geometrically integral hypersurface of degree δ\delta, 𝒳𝒪KnSpec𝒪K\mathscr{X}\hookrightarrow\mathbb{P}^{n}_{\mathcal{O}_{K}}\rightarrow\operatorname{Spec}\mathcal{O}_{K} be its Zariski closure, and 𝒳𝔽𝔭=𝒳×Spec𝒪KSpec𝔽𝔭Spec𝔽𝔭\mathscr{X}_{\mathbb{F}_{\mathfrak{p}}}=\mathscr{X}\times_{\operatorname{Spec}\mathcal{O}_{K}}\operatorname{Spec}\mathbb{F}_{\mathfrak{p}}\rightarrow\operatorname{Spec}\mathbb{F}_{\mathfrak{p}} for every 𝔭Spm𝒪K\mathfrak{p}\in\operatorname{Spm}\mathcal{O}_{K}. By [21, Théorème 9.7.7], the set

𝒬(𝒳)={𝔭Spm𝒪K|𝒳𝔽𝔭Spec𝔽𝔭 is not geometrically integral}\mathcal{Q}(\mathscr{X})=\left\{\mathfrak{p}\in\operatorname{Spm}\mathcal{O}_{K}|\;\mathscr{X}_{\mathbb{F}_{\mathfrak{p}}}\rightarrow\operatorname{Spec}\mathbb{F}_{\mathfrak{p}}\hbox{ is not geometrically integral}\right\} (18)

is finite.

Next, we introduce a numerical description of the set 𝒬(𝒳)\mathcal{Q}(\mathscr{X}). In fact, there are fruitful results on this subject, but most of them are over rational number field \mathbb{Q}. 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

1[K:]𝔭𝒬(𝒳)logN(𝔭)(δ21)h(X)+C(n,δ),\frac{1}{[K:\mathbb{Q}]}\sum_{\mathfrak{p}\in\mathcal{Q}(\mathscr{X})}\log N(\mathfrak{p})\leqslant(\delta^{2}-1)h(X)+C(n,\delta), (19)

where h(X)h(X) is the naive height of XX defined in Definition 2.7.2, N(𝔭)=#(𝒪K/𝔭)N(\mathfrak{p})=\#(\mathcal{O}_{K}/\mathfrak{p}), and the constant

C(n,δ)=(δ21)(3logδ+δlog3+log(n+δδ)).C(n,\delta)=(\delta^{2}-1)\left(3\log\delta+\delta\log 3+\log{n+\delta\choose\delta}\right).

In fact, we have C(n,δ)nδ3C(n,\delta)\ll_{n}\delta^{3}.

4.2 Quantitative estimates over finite fields

In this subsection, we give an upper bound of the term n(𝒳𝔭)n(\mathscr{X}_{\mathfrak{p}}) for an arbitrary maximal ideal 𝔭\mathfrak{p} of 𝒪K\mathcal{O}_{K}, where n(𝒳𝔭)n(\mathscr{X}_{\mathfrak{p}}) is defined in the statement of Theorem 3. In this part, we consider this problem over arbitrary finite fields.

Let 𝔽q\mathbb{F}_{q} be the finite field with qq elements, XX be a geometrically integral hypersurface in 𝔽qn\mathbb{P}^{n}_{\mathbb{F}_{q}} of degree δ\delta, and n(X𝔽q)=ξX(𝔽q)μξ(X)n(X_{\mathbb{F}_{q}})=\sum\limits_{\xi\in X(\mathbb{F}_{q})}\mu_{\xi}(X), where μξ(X)\mu_{\xi}(X) is the multiplicity of ξ\xi in XX defined via the local Hilbert-Samuel fuction in (4). Then we have

n(X𝔽q)=#X(𝔽q)+ξX(𝔽q)(μξ(X)1).n(X_{\mathbb{F}_{q}})=\#X(\mathbb{F}_{q})+\sum\limits_{\xi\in X(\mathbb{F}_{q})}\left(\mu_{\xi}(X)-1\right).

In order to estimate n(X𝔽q)n(X_{\mathbb{F}_{q}}), we will consider the terms #X(𝔽q)\#X(\mathbb{F}_{q}) and ξX(𝔽q)(μξ(X)1)\sum\limits_{\xi\in X(\mathbb{F}_{q})}\left(\mu_{\xi}(X)-1\right) separately.

4.2.1

For the estimate of #X(𝔽q)\#X(\mathbb{F}_{q}), 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 X𝔽qnX\hookrightarrow\mathbb{P}^{n}_{\mathbb{F}_{q}} be a geometrically integral hypersurface of degree δ\delta over the finite field 𝔽q\mathbb{F}_{q}. When qδ2q\leqslant\delta^{2} or q27δ4q\geqslant 27\delta^{4}, we have

#X(𝔽q)qn1nδ2qn32.\#X(\mathbb{F}_{q})-q^{n-1}\leqslant n\delta^{2}q^{n-\frac{3}{2}}.
Démonstration.

We consider this estimate case by case as following.

  1. 1.

    If qδq\leqslant\delta, we have #X(𝔽q)#n(𝔽q)=qn++1\#X(\mathbb{F}_{q})\leqslant\#\mathbb{P}^{n}(\mathbb{F}_{q})=q^{n}+\cdots+1. Then

    #X(𝔽q)qn1nqnnδ2qn32.\#X(\mathbb{F}_{q})-q^{n-1}\leqslant nq^{n}\leqslant n\delta^{2}q^{n-\frac{3}{2}}.
  2. 2.

    If δ+1qδ2\delta+1\leqslant q\leqslant\delta^{2}, we have #X(𝔽q)δ#n1(𝔽q)=δ(qn1++1)\#X(\mathbb{F}_{q})\leqslant\delta\#\mathbb{P}^{n-1}(\mathbb{F}_{q})=\delta(q^{n-1}+\cdots+1). Then

    #X(𝔽q)qn1(δ1)qn1+δ(qn2++1)nδ2qn32.\#X(\mathbb{F}_{q})-q^{n-1}\leqslant(\delta-1)q^{n-1}+\delta(q^{n-2}+\cdots+1)\leqslant n\delta^{2}q^{n-\frac{3}{2}}.
  3. 3.

    If q27δ4q\geqslant 27\delta^{4}, by [9, Corollary 5.6], we have

    #X(𝔽q)qn1(δ1)(δ2)qn32+(5δ2+δ+1)qn2nδ2qn32.\#X(\mathbb{F}_{q})-q^{n-1}\leqslant(\delta-1)(\delta-2)q^{n-\frac{3}{2}}+(5\delta^{2}+\delta+1)q^{n-2}\leqslant n\delta^{2}q^{n-\frac{3}{2}}.

\remaname \the\smf@thm.

With all the notations in Proposition 4.2.1. When δ2nqnδ4\delta^{2}\ll_{n}q\ll_{n}\delta^{4}, by [9, Corollary 5.6], we have

#X(𝔽q)qn1(δ1)(δ2)qn32+B(n,δ)qn2,\#X(\mathbb{F}_{q})-q^{n-1}\leqslant(\delta-1)(\delta-2)q^{n-\frac{3}{2}}+B(n,\delta)q^{n-2},

where the constant satisfies B(n,δ)nδ4B(n,\delta)\ll_{n}\delta^{4}. It seems that the constant B(n,δ)B(n,\delta) could have a better dependence on δ\delta, but up to the author’s knowledge, we do not know the answer.

4.2.2

For the term ξX(𝔽q)(μξ(X)1)\sum\limits_{\xi\in X(\mathbb{F}_{q})}\left(\mu_{\xi}(X)-1\right), by [27, Theorem 5.1], we have

ξX(𝔽q)(μξ(X)1)\displaystyle\sum\limits_{\xi\in X(\mathbb{F}_{q})}\left(\mu_{\xi}(X)-1\right) \displaystyle\leqslant 12ξX(𝔽q)μξ(X)(μξ(X)1)\displaystyle\frac{1}{2}\sum\limits_{\xi\in X(\mathbb{F}_{q})}\mu_{\xi}(X)\left(\mu_{\xi}(X)-1\right)
\displaystyle\leqslant (n1)22δ(δ1)max{δ1,q}n2,\displaystyle\frac{(n-1)^{2}}{2}\delta(\delta-1)\max\{\delta-1,q\}^{n-2},

which has the optimal dependances on δ\delta and max{δ1,q}\max\{\delta-1,q\} when qδ1q\geqslant\delta-1.

4.2.3

We combine Propostion 4.2.1 and the estimate (4.2.2). When qδ2q\leqslant\delta^{2} or q27δ4q\geqslant 27\delta^{4}, we have

n(X𝔽q)qn1+n2δ2max{q,δ1}n32n(X_{\mathbb{F}_{q}})\leqslant q^{n-1}+n^{2}\delta^{2}\max\left\{q,\delta-1\right\}^{n-\frac{3}{2}}

by an elementary calculation. In addition, we have

1n(X𝔽q)1n11qn2δ2max{q,δ1}32\frac{1}{n(X_{\mathbb{F}_{q}})^{\frac{1}{n-1}}}\geqslant\frac{1}{q}-\frac{n^{2}\delta^{2}}{\max\left\{q,\delta-1\right\}^{\frac{3}{2}}} (21)

under the same assumption of qq and δ\delta as above.

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 𝒪K\mathcal{O}_{K}. First, we generalize [40, Lemma 1.10] over an arbitrary number field, where the former result works over \mathbb{Z} only.

\lemmname \the\smf@thm.

Let 𝔞\mathfrak{a} be a proper ideal of 𝒪K\mathcal{O}_{K}, 𝔭\mathfrak{p} be an prime ideal of 𝒪K\mathcal{O}_{K}, and N(𝔞)=#(𝒪K/𝔞)N(\mathfrak{a})=\#(\mathcal{O}_{K}/\mathfrak{a}). Then we have

1[K:]𝔭𝔞logN(𝔭)N(𝔭)loglog(N(𝔞))+2,\frac{1}{[K:\mathbb{Q}]}\sum_{\mathfrak{p}\supseteq\mathfrak{a}}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})}\leqslant\log\log\left(N(\mathfrak{a})\right)+2,

where the above sum takes all over the prime ideals contained in 𝔞\mathfrak{a} of 𝒪K\mathcal{O}_{K}.

Démonstration.

We will prove the inequality for the case of K=K=\mathbb{Q} at first, and then we show the general case by it.

Case of K=K=\mathbb{Q}. - In this case, we will repeat the proof of [40, Lemma 1.10] by Salberger, since this preprint is not easily available. Suppose that 𝔞\mathfrak{a} is generated by the positive square-free integer π\pi, and let mm be a positive integer such that mπm\leqslant\pi. For the prime pp, let vp(m)v_{p}(m) be the largest integer such that pvp(m)mp^{v_{p}(m)}\mid m. By [44, Tome I, Corollaire 1.7] and [44, Tome I, Théorème 1.8], we have

mp|πlogppp|πlogpp|πvp(m!)logppπvp(m!)logp=logm!mlogm,m\sum_{p|\pi}\frac{\log p}{p}-\sum_{p|\pi}\log p\leqslant\sum_{p|\pi}v_{p}(m!)\log p\leqslant\sum_{p\leqslant\pi}v_{p}(m!)\log p=\log m!\leqslant m\log m,

and then we obtain

p|πlogpplogm+1mp|πlogplogm+1mlogπ.\sum_{p|\pi}\frac{\log p}{p}\leqslant\log m+\frac{1}{m}\sum_{p|\pi}\log p\leqslant\log m+\frac{1}{m}\log\pi.

Let m=[logπ]m=[\log\pi] for π2\pi\geqslant 2, and then we accomplish the proof for K=K=\mathbb{Q}.

Case of arbitrary number fields. - Let

𝔞=𝔭1v𝔭1(𝔞)𝔭kv𝔭k(𝔞),\mathfrak{a}=\mathfrak{p}_{1}^{v_{\mathfrak{p}_{1}}(\mathfrak{a})}\cdots\mathfrak{p}_{k}^{v_{\mathfrak{p}_{k}}(\mathfrak{a})},

where 𝔭1,,𝔭k\mathfrak{p}_{1},\ldots,\mathfrak{p}_{k} are distinct prime ideals of 𝒪K\mathcal{O}_{K}, and v𝔭i(𝔞)+v_{\mathfrak{p}_{i}}(\mathfrak{a})\in\mathbb{N}^{+} for all i=1,,ki=1,\ldots,k. Let the prime pip_{i} be the characteristic of the prime ideal 𝔭i\mathfrak{p}_{i}, where i=1,,ki=1,\ldots,k as above. For a fixed prime pp, there are at most [K:][K:\mathbb{Q}] prime ideals of characteristic pp in 𝒪K\mathcal{O}_{K}. For all prime pp and f+f\in\mathbb{N}^{+}, we have

logpfpflogpp.\frac{\log p^{f}}{p^{f}}\leqslant\frac{\log p}{p}.

Let P(𝔞)P(\mathfrak{a}) be the product of all the different characteristics of 𝔭1,,𝔭k\mathfrak{p}_{1},\ldots,\mathfrak{p}_{k}, and we have P(𝔞)N(𝔞)P(\mathfrak{a})\leqslant N(\mathfrak{a}) by definition directly. Then by the above facts, we obtain

𝔭𝔞logN(𝔭)N(𝔭)[K:]p|P(𝔞)logpp.\sum_{\mathfrak{p}\supseteq\mathfrak{a}}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})}\leqslant[K:\mathbb{Q}]\sum_{p|P(\mathfrak{a})}\frac{\log p}{p}.

By the case of K=K=\mathbb{Q}, we have

p|P(𝔞)logpploglogP(𝔞)+2loglogN(𝔞)+2,\sum_{p|P(\mathfrak{a})}\frac{\log p}{p}\leqslant\log\log P(\mathfrak{a})+2\leqslant\log\log N(\mathfrak{a})+2,

which proves the assertion. ∎

4.3.2 Distribution of prime ideals with bounded norm

Let x+x\in\mathbb{R}^{+}, 𝔭Spm𝒪K\mathfrak{p}\in\operatorname{Spm}\mathcal{O}_{K}, and N(𝔭)=#(𝒪K/𝔭)N(\mathfrak{p})=\#(\mathcal{O}_{K}/\mathfrak{p}). In this part, we consider the of

θK(x)=N(𝔭)xlogN(𝔭),ψK(x)=N(𝔭)xlogN(𝔭)N(𝔭),ϕK(x)=N(𝔭)xlogN(𝔭)N(𝔭)32.\theta_{K}(x)=\sum\limits_{N(\mathfrak{p})\leqslant x}\log N(\mathfrak{p}),\;\psi_{K}(x)=\sum\limits_{N(\mathfrak{p})\leqslant x}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})},\;\phi_{K}(x)=\sum\limits_{N(\mathfrak{p})\leqslant x}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})^{\frac{3}{2}}}. (22)

When K=K=\mathbb{Q}, 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

θk(x)=x+OK(xeclogx),\theta_{k}(x)=x+O_{K}\left(xe^{-c\sqrt{\log x}}\right),

where cc is a constant depending on KK. Then there exists a function ϵ1(K,x)\epsilon_{1}(K,x) of the number field KK and x+x\in\mathbb{R}^{+}, such that

|θK(x)x|ϵ1(K,x),\left|\theta_{K}(x)-x\right|\leqslant\epsilon_{1}(K,x), (23)

where ϵ1(K,x)=OK(xeclogx)\epsilon_{1}(K,x)=O_{K}\left(xe^{-c\sqrt{\log x}}\right) for all x+x\in\mathbb{R}^{+}, and cc depends on KK only.

By [37, Lemma 2.3], we have

ψK(x)=logx+OK(1),\psi_{K}(x)=\log x+O_{K}(1),

which is obtained by Abel’s summation formula applied in [37, Lemma 2.1]. Then there exists a function ϵ2(K)\epsilon_{2}(K) of the number field KK, such that

|ψK(x)logx|ϵ2(K).\left|\psi_{K}(x)-\log x\right|\leqslant\epsilon_{2}(K). (24)

Same as the application of [37, Lemma 2.1] to the proof of [37, Lemma 2.3], we have

ϕK(x)=θK(x)x32+322xθK(t)t52𝑑t\phi_{K}(x)=\frac{\theta_{K}(x)}{x^{\frac{3}{2}}}+\frac{3}{2}\int_{2}^{x}\frac{\theta_{K}(t)}{t^{\frac{5}{2}}}dt

by [37, Lemma 2.1]. Then by (23), we have

|ϕK(x)1x322x1t32𝑑t|ϵ1(K,x)x32+322xϵ1(K,t)t52𝑑t.\left|\phi_{K}(x)-\frac{1}{\sqrt{x}}-\frac{3}{2}\int_{2}^{x}\frac{1}{t^{\frac{3}{2}}}dt\right|\leqslant\frac{\epsilon_{1}(K,x)}{x^{\frac{3}{2}}}+\frac{3}{2}\int_{2}^{x}\frac{\epsilon_{1}(K,t)}{t^{\frac{5}{2}}}dt.

Then by an elementary calculation, there exists a function ϵ3(K,x)\epsilon_{3}(K,x) of the number field KK and x+x\in\mathbb{R}^{+}, such that

|ϕK(x)322+2x|ϵ3(K,x),\left|\phi_{K}(x)-\frac{3}{2}\sqrt{2}+\frac{2}{\sqrt{x}}\right|\leqslant\epsilon_{3}(K,x), (25)

where

ϵ3(K,x)=ϵ1(K,x)x32+322xϵ1(K,t)t52𝑑t.\epsilon_{3}(K,x)=\frac{\epsilon_{1}(K,x)}{x^{\frac{3}{2}}}+\frac{3}{2}\int_{2}^{x}\frac{\epsilon_{1}(K,t)}{t^{\frac{5}{2}}}dt. (26)

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.

Let XX be a geometrically integral hypersurface of degree δ\delta in Kn\mathbb{P}^{n}_{K}, and 𝒳\mathscr{X} be its Zariski closure in 𝒪Kn\mathbb{P}^{n}_{\mathcal{O}_{K}}. Let 𝔭Spm𝒪K\mathfrak{p}\in\operatorname{Spm}\mathcal{O}_{K}, N(𝔭)=#(𝒪K/𝔭)N(\mathfrak{p})=\#(\mathcal{O}_{K}/\mathfrak{p}),

𝒬(𝒳)={𝔭Spm𝒪K|N(𝔭)>27δ4 and 𝒳𝔭Spec𝔽𝔭 not geometrically integral},\mathcal{Q}^{\prime}(\mathscr{X})=\{\mathfrak{p}\in\operatorname{Spm}\mathcal{O}_{K}|\;N(\mathfrak{p})>27\delta^{4}\hbox{ and }\mathscr{X}_{\mathfrak{p}}\rightarrow\operatorname{Spec}\mathbb{F}_{\mathfrak{p}}\hbox{ not geometrically integral}\},

and

b(𝒳)=𝔭𝒬(𝒳)exp(logN(𝔭)N(𝔭)).b^{\prime}(\mathscr{X})=\prod_{\mathfrak{p}\in\mathcal{Q}^{\prime}(\mathscr{X})}\exp\left(\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})}\right).

Then we have

b(𝒳)\displaystyle b^{\prime}(\mathscr{X}) \displaystyle\leqslant exp(2ϵ2(K)3log3+[K:])\displaystyle\exp\left(2\epsilon_{2}(K)-3\log 3+[K:\mathbb{Q}]\right)
(δ2δ4)(h(X)+(3logδ+δlog3+log(n+δδ))),\displaystyle\;\cdot(\delta^{-2}-\delta^{-4})\left(h(X)+\left(3\log\delta+\delta\log 3+\log{n+\delta\choose\delta}\right)\right),

where h(X)h(X) is defined in Definition 2.7.2, and ϵ2(K)\epsilon_{2}(K) is defined in (24).

Démonstration.

We denote by 𝒫(𝒳)\mathcal{P}^{\prime}(\mathscr{X}) the product of all maximal ideals in 𝒬(𝒳)\mathcal{Q}^{\prime}(\mathscr{X}), and

c(𝒳)=(δ21)(h(X)+(3logδ+δlog3+log(n+δδ))).c^{\prime}(\mathscr{X})=(\delta^{2}-1)\left(h(X)+\left(3\log\delta+\delta\log 3+\log{n+\delta\choose\delta}\right)\right).

Then by Lemma 4.3.1 and (19), we have

1[K:]logb(𝒳)=1[K:]𝔭𝒬(𝒳)logN(𝔭)N(𝔭)\displaystyle\frac{1}{[K:\mathbb{Q}]}\log b^{\prime}(\mathscr{X})=\frac{1}{[K:\mathbb{Q}]}\sum_{\mathfrak{p}\in\mathcal{Q}^{\prime}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})}
\displaystyle\leqslant 1[K:]27δ4<N(𝔭)c(𝒳)logN(𝔭)N(𝔭)+1[K:]𝔭𝒬(𝒳)N(𝔭)>c(𝒳)logN(𝔭)c(𝒳).\displaystyle\frac{1}{[K:\mathbb{Q}]}\sum_{27\delta^{4}<N(\mathfrak{p})\leqslant c^{\prime}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})}+\frac{1}{[K:\mathbb{Q}]}\sum_{\begin{subarray}{c}\mathfrak{p}\in\mathcal{Q}^{\prime}(\mathscr{X})\\ N(\mathfrak{p})>c^{\prime}(\mathscr{X})\end{subarray}}\frac{\log N(\mathfrak{p})}{c^{\prime}(\mathscr{X})}.

By (24), we have

1[K:]27δ4<N(𝔭)c(𝒳)logN(𝔭)N(𝔭)\displaystyle\frac{1}{[K:\mathbb{Q}]}\sum_{27\delta^{4}<N(\mathfrak{p})\leqslant c^{\prime}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})}
=\displaystyle= 1[K:]N(𝔭)c(𝒳)logN(𝔭)N(𝔭)1[K:]N(𝔭)27δ4logN(𝔭)N(𝔭)\displaystyle\frac{1}{[K:\mathbb{Q}]}\sum_{N(\mathfrak{p})\leqslant c^{\prime}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})}-\frac{1}{[K:\mathbb{Q}]}\sum_{N(\mathfrak{p})\leqslant 27\delta^{4}}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})}
\displaystyle\leqslant 1[K:](logc(𝒳)4logδ+2ϵ2(K)3log3),\displaystyle\frac{1}{[K:\mathbb{Q}]}\left(\log c^{\prime}(\mathscr{X})-4\log\delta+2\epsilon_{2}(K)-3\log 3\right),

where ϵ2(K)\epsilon_{2}(K) is defined in (24) depending on KK only.

Let

𝒬(𝒳)={𝔭Spm𝒪K|𝒳𝔽𝔭Spec𝔽𝔭 is not geometrically integral}.\mathcal{Q}(\mathscr{X})=\left\{\mathfrak{p}\in\operatorname{Spm}\mathcal{O}_{K}|\;\mathscr{X}_{\mathbb{F}_{\mathfrak{p}}}\rightarrow\operatorname{Spec}\mathbb{F}_{\mathfrak{p}}\hbox{ is not geometrically integral}\right\}.

Since 𝒬(𝒳)𝒬(𝒳)\mathcal{Q}^{\prime}(\mathscr{X})\subseteq\mathcal{Q}(\mathscr{X}), then we have

1[K:]𝔭𝒬(𝒳)N(𝔭)>c(𝒳)logN(𝔭)c(𝒳)1[K:]c(𝒳)𝔭𝒬(𝒳)logN(𝔭)1,\frac{1}{[K:\mathbb{Q}]}\sum_{\begin{subarray}{c}\mathfrak{p}\in\mathcal{Q}^{\prime}(\mathscr{X})\\ N(\mathfrak{p})>c^{\prime}(\mathscr{X})\end{subarray}}\frac{\log N(\mathfrak{p})}{c^{\prime}(\mathscr{X})}\leqslant\frac{1}{[K:\mathbb{Q}]c^{\prime}(\mathscr{X})}\sum_{\mathfrak{p}\in\mathcal{Q}(\mathscr{X})}\log N(\mathfrak{p})\leqslant 1,

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

b(𝒳)n,Kmax{δ2h(X),δ1}.b^{\prime}(\mathscr{X})\ll_{n,K}\max\left\{\delta^{-2}h(X),\delta^{-1}\right\}.

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

(Nm+1)mm!(Nm)(N(m1)/2)mm!\frac{(N-m+1)^{m}}{m!}\leqslant{N\choose m}\leqslant\frac{\left(N-(m-1)/2\right)^{m}}{m!}

will be helpful in the calculation below.

\lemmname \the\smf@thm.

Let XX be a hypersurface of degree δ\delta in Kn\mathbb{P}^{n}_{K}. We denote by r1(n,D)r_{1}(n,D) its geometric Hilbert-Samuel function with the variable DD. When Dδ+1D\geqslant\delta+1, we have

r1(n,D)1n1δ(n1)!n1D(δ2)δ(n1)!n1,r_{1}(n,D)^{\frac{1}{n-1}}\geqslant\sqrt[n-1]{\frac{\delta}{(n-1)!}}D-(\delta-2)\sqrt[n-1]{\frac{\delta}{(n-1)!}},

and

r1(n,D)1n1δ(n1)!n1D+n2δ(n1)!n1.r_{1}(n,D)^{\frac{1}{n-1}}\leqslant\sqrt[n-1]{\frac{\delta}{(n-1)!}}D+\frac{n}{2}\sqrt[n-1]{\frac{\delta}{(n-1)!}}.
Démonstration.

In fact, we have

r1(n,D)=(n+Dn)(n+Dδn)r_{1}(n,D)={n+D\choose n}-{n+D-\delta\choose n}

when Dδ+1D\geqslant\delta+1.

In order to obtain the lower bound, we have

r1(n,D)\displaystyle r_{1}(n,D) =\displaystyle= j=1δ(Dδ+n1+jn1)δ(Dδ+2)n1(n1)!\displaystyle\sum_{j=1}^{\delta}{D-\delta+n-1+j\choose n-1}\geqslant\frac{\delta(D-\delta+2)^{n-1}}{(n-1)!}
\displaystyle\geqslant δ(n1)!Dn1δ(δ2)(n1)!Dn2.\displaystyle\frac{\delta}{(n-1)!}D^{n-1}-\frac{\delta(\delta-2)}{(n-1)!}D^{n-2}.

Then we obtain

r1(n,D)1n1\displaystyle r_{1}(n,D)^{\frac{1}{n-1}} \displaystyle\geqslant δ(n1)!n1D(1δ2D)1n1\displaystyle\sqrt[n-1]{\frac{\delta}{(n-1)!}}D\left(1-\frac{\delta-2}{D}\right)^{\frac{1}{n-1}}
\displaystyle\geqslant δ(n1)!n1D(δ2)δ(n1)!n1\displaystyle\sqrt[n-1]{\frac{\delta}{(n-1)!}}D-(\delta-2)\sqrt[n-1]{\frac{\delta}{(n-1)!}}

when Dδ+1D\geqslant\delta+1.

On the other hand, we have

r1(n,D)=j=1δ(Dδ+n1+jn1)δ(D+n2)n1(n1)!=δDn1(n1)!(1+n2D)n1,r_{1}(n,D)=\sum_{j=1}^{\delta}{D-\delta+n-1+j\choose n-1}\leqslant\frac{\delta(D+\frac{n}{2})^{n-1}}{(n-1)!}=\frac{\delta D^{n-1}}{(n-1)!}\left(1+\frac{n}{2D}\right)^{n-1}, (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).

We keep all the notations in §2.5. Let XX be a closed integral subscheme of Kn\mathbb{P}^{n}_{K}, Z=(Pi)iIZ=(P_{i})_{i\in I} be a family of rational points, and

ϕZ,D:FD,KiIPi𝒪Kn(D)\phi_{Z,D}:F_{D,K}\rightarrow\bigoplus_{i\in I}P^{*}_{i}\mathcal{O}_{\mathbb{P}^{n}_{K}}(D)

be the evaluation map. If we have the inequality

supiIh𝒪(1)¯(Pi)<μ^max(F¯D)D12Dlogr1(n,D),\sup_{i\in I}h_{\overline{\mathcal{O}(1)}}(P_{i})<\frac{\widehat{\mu}_{\max}(\overline{F}_{D})}{D}-\frac{1}{2D}\log r_{1}(n,D),

where r1(n,D)=rk(FD)r_{1}(n,D)=\operatorname{rk}(F_{D}) and the height function h𝒪(1)¯(.)h_{\overline{\mathcal{O}(1)}}(\raisebox{1.72218pt}{.}) is defined in (12), then the homomorphism ϕZ,D\phi_{Z,D} is not able to be injective.

The uniform lower bound of μ^(F¯D)\widehat{\mu}(\overline{F}_{D}) for all D1D\geqslant 1 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 μ^(F¯D)\widehat{\mu}(\overline{F}_{D}). This result is reformulated by H. Chen in [12, Theorem 4.8] by the language of the slope method. In fact, let XX be a closed integral subscheme of Kn\mathbb{P}^{n}_{K} of dimension dd and degree δ\delta, and 𝒳\mathscr{X} be its Zariski closure in 𝒪Kn\mathbb{P}^{n}_{\mathcal{O}_{K}}. The inequality

μ^(F¯D)Dd!δ(2d+2)d+1h𝒪(1)¯(𝒳)log(n+1)2d\frac{\widehat{\mu}(\overline{F}_{D})}{D}\geqslant\frac{d!}{\delta(2d+2)^{d+1}}h_{\overline{\mathcal{O}(1)}}(\mathscr{X})-\log(n+1)-2^{d} (29)

is uniformly verified for all D2(nd)(δ1)+d+2D\geqslant 2(n-d)(\delta-1)+d+2 (see also [12, Remark 4.9] for some minor modifications), where h𝒪(1)¯(𝒳)h_{\overline{\mathcal{O}(1)}}(\mathscr{X}) 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 XX be an integral hypersurface of degree δ\delta in Kn\mathbb{P}^{n}_{K}, the constant n=1++1n\mathcal{H}_{n}=1+\cdots+\frac{1}{n} and the constant

C1(n)=(2n)n(n1)!(log2+5log(n+1)12n)32log(n+1)2n1.C_{1}(n)=-\frac{(2n)^{n}}{(n-1)!}\left(\log 2+5\log(n+1)-\frac{1}{2}\mathcal{H}_{n}\right)-\frac{3}{2}\log(n+1)-2^{n-1}. (30)

If

logB[K:]<(n1)!δ(2n)nh(X)+C1(n),\frac{\log B}{[K:\mathbb{Q}]}<\frac{(n-1)!}{\delta(2n)^{n}}h(X)+C_{1}(n),

then there exists a hypersurface of degree smaller than 2δ+n12\delta+n-1, which contains all rational points in S(X;B)S(X;B) but does not contain the generic point of XX, where we use the height function defined in (12).

Démonstration.

If there does not exist such a hypersurface, the evaluation map ϕZ,D\phi_{Z,D} in Proposition 5.1 is injective. On the other hand, by (29), Proposition 2.7.3 and the fact that

r1(n,D)(n+Dn)(n+1)Dr_{1}(n,D)\leqslant{n+D\choose n}\leqslant(n+1)^{D}

is uniformly verified for all n,D1n,D\geqslant 1, we obtain

logB[K:]\displaystyle\frac{\log B}{[K:\mathbb{Q}]} <\displaystyle< (n1)!δ(2n)nh(X)+C1(n)\displaystyle\frac{(n-1)!}{\delta(2n)^{n}}h(X)+C_{1}(n)
\displaystyle\leqslant (n1)!δ(2n)nh𝒪(1)¯(𝒳)32log(n+1)2n1\displaystyle\frac{(n-1)!}{\delta(2n)^{n}}h_{\overline{\mathcal{O}(1)}}(\mathscr{X})-\frac{3}{2}\log(n+1)-2^{n-1}
\displaystyle\leqslant μ^max(F¯D)D12Dlogr1(n,D),\displaystyle\frac{\widehat{\mu}_{\max}(\overline{F}_{D})}{D}-\frac{1}{2D}\log r_{1}(n,D),

which contradicts to Proposition 5.1. ∎

\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

μ^(F¯D)=h𝒪(1)¯(𝒳)nδD+o(D)\widehat{\mu}(\overline{F}_{D})=\frac{h_{\overline{\mathcal{O}(1)}}(\mathscr{X})}{n\delta}D+o(D)

for DD tends into infinity. So it is expected that we can obtain a better uniform lower bound of μ^(F¯D)\widehat{\mu}(\overline{F}_{D}) 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 KK and the positive integer n2n\geqslant 2, which will be used in the estimate of the determinant.

Let KK be a number field, we denote

κ1(K)=supx+ϵ1(K,x)x,\kappa_{1}(K)=\sup_{x\in\mathbb{R}^{+}}\frac{\epsilon_{1}(K,x)}{x}, (31)

where ϵ1(K,x)\epsilon_{1}(K,x) is introduced in (23). By [37, Theorem 2.2], the above supremum exists and κ1(K)\kappa_{1}(K) depends on KK only.

Let δ1\delta\geqslant 1 and n2n\geqslant 2 be two integers, we denote

κ2(K,n)=supδ1{3log3+2n233+2n2δ2ϵ3(K,27δ4)+2ϵ2(K)},\kappa_{2}(K,n)=\sup_{\delta\geqslant 1}\left\{-3\log 3+\frac{2n^{2}}{3\sqrt{3}}+2n^{2}\delta^{2}\epsilon_{3}(K,27\delta^{4})+2\epsilon_{2}(K)\right\}, (32)

where ϵ2(K)\epsilon_{2}(K) is defined in (24), and ϵ3(K,x)\epsilon_{3}(K,x) is defined in (26). By taking [37, Theorem 2.2] into the estimate of (26), the above supremum exists and κ2(K,n)\kappa_{2}(K,n) depends on KK and nn only.

\theoname \the\smf@thm.

Let KK be a number field. Let XX be a geometrically integral hypersurface in Kn\mathbb{P}^{n}_{K} of degree δ2\delta\geqslant 2, and S(X;B)S(X;B) be the set of rational points in XX whose height is smaller than BB with respect to the above closed immersion, see (12) for the definition of the height function used above. Then there exists a hypersurface in Kn\mathbb{P}^{n}_{K} of degree smaller than

eC2(n,K)Bn/((n1)δ1/(n1))δ41/(n1)b(𝒳)HK(X)n!(n1)(2n)nδ11/(n1)e^{C_{2}(n,K)}B^{{n}/{\left((n-1)\delta^{1/(n-1)}\right)}}\delta^{4-{1}/{(n-1)}}\frac{b^{\prime}(\mathscr{X})}{H_{K}(X)^{\frac{n!}{(n-1)(2n)^{n}}\delta^{-1-{1}/{(n-1)}}}}

which covers S(X;B)S(X;B) but does not contain the generic point of XX, where the constant

C2(n,K)\displaystyle C_{2}(n,K) =\displaystyle= nC1(n)[K:](n1)2n1+κ2(K,n)+log(n1)!n1\displaystyle\frac{nC_{1}(n)[K:\mathbb{Q}]}{(n-1)\sqrt[n-1]{2}}+\kappa_{2}(K,n)+\frac{\log(n-1)!}{n-1}
+3+n3+2n2+n42(n21)(n1)!n1(1+n4)(1+κ1(K)),\displaystyle+3+\frac{n^{3}+2n^{2}+n-4}{2(n^{2}-1)\sqrt[n-1]{(n-1)!}}\left(1+\frac{n}{4}\right)\left(1+\kappa_{1}(K)\right),

the constant C1(n)C_{1}(n) is defined in (30), b(𝒳)b^{\prime}(\mathscr{X}) is defined in Proposition 4.3.3, κ1(K)\kappa_{1}(K) is defined in (31), κ2(K,n)\kappa_{2}(K,n) is defined in (32), and the height HK(X)H_{K}(X) of XX 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

logB[K:]<(n1)!δ(2n)nh(X)+C1(n),\frac{\log B}{[K:\mathbb{Q}]}<\frac{(n-1)!}{\delta(2n)^{n}}h(X)+C_{1}(n),

where the constant C1(n)C_{1}(n) is defined in (30) and h(X)h(X) is defined in Definition 2.7.2. Then by Proposition 5.1, S(X;B)S(X;B) can be covered by a hypersurface of degree no more than 2δ+n12\delta+n-1 which does not contain the generic point of XX. By an elementary calculation, we obtain that 2δ+n12\delta+n-1 is smaller than the bound provided in the statement of the theorem, for n2n\geqslant 2, δ1\delta\geqslant 1, b(𝒳)1b^{\prime}(\mathscr{X})\geqslant 1, and HK(X)1H_{K}(X)\geqslant 1.

II. Case of small height varieties. - For the case of

logB[K:](n1)!δ(2n)nh(X)+C1(n),\frac{\log B}{[K:\mathbb{Q}]}\geqslant\frac{(n-1)!}{\delta(2n)^{n}}h(X)+C_{1}(n),

which is equivalent to

h(X)δ(2n)n(n1)!(logB[K:]C1(n)),h(X)\leqslant\frac{\delta(2n)^{n}}{(n-1)!}\cdot\left(\frac{\log B}{[K:\mathbb{Q}]}-C_{1}(n)\right),

we will treat it as following. We keep all the notations in Theorem 3, and we suppose D3δlogδ+n12δ+n1D\geqslant 3\delta\log\delta+n-1\geqslant 2\delta+n-1 from now on. We denote the set

(𝒳)\displaystyle\mathcal{R}(\mathscr{X}) =\displaystyle= {𝔭Spm𝒪K| 27δ4N(𝔭)r1(n,D)1n1,\displaystyle\{\mathfrak{p}\in\operatorname{Spm}\mathcal{O}_{K}|\;27\delta^{4}\leqslant N(\mathfrak{p})\leqslant r_{1}(n,D)^{\frac{1}{n-1}},
𝒳𝔭Spec𝔽𝔭 is geometrically integral},\displaystyle\;\mathscr{X}_{\mathfrak{p}}\rightarrow\operatorname{Spec}\mathbb{F}_{\mathfrak{p}}\hbox{ is geometrically integral}\},

and we apply Theorem 3 to the reductions at (𝒳)\mathcal{R}(\mathscr{X}). If there does not exist such a hypersurface, then by Theorem 3 applied in the above sense, we have

logB[K:]μ^(F¯D)Dlogr1(n,D)2D\displaystyle\frac{\log B}{[K:\mathbb{Q}]}\geqslant\frac{\widehat{\mu}(\overline{F}_{D})}{D}-\frac{\log r_{1}(n,D)}{2D}
+1[K:]𝔭(𝒳)((n1)!1n1(n1)r1(n,D)1n1nDn(𝒳𝔭)1n1n3+2n2+n42Dn(n+1))logN(𝔭).\displaystyle\;+\frac{1}{[K:\mathbb{Q}]}\sum_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\left(\frac{(n-1)!^{\frac{1}{n-1}}(n-1)r_{1}(n,D)^{\frac{1}{n-1}}}{nDn(\mathscr{X}_{\mathfrak{p}})^{\frac{1}{n-1}}}-\frac{n^{3}+2n^{2}+n-4}{2Dn(n+1)}\right)\log N(\mathfrak{p}).

From the explicit lower bound of μ^(FD)\widehat{\mu}(F_{D}) provided at (29) and Proposition 5.1, we deduce

logB[K:](n1)!δ(2n)nh(X)+C1(n)\displaystyle\frac{\log B}{[K:\mathbb{Q}]}-\frac{(n-1)!}{\delta(2n)^{n}}h(X)+C_{1}(n)
\displaystyle\geqslant (n1)!1n1(n1)r1(n,D)1n1nD[K:]𝔭(𝒳)logN(𝔭)n(𝒳𝔭)1n1\displaystyle\frac{(n-1)!^{\frac{1}{n-1}}(n-1)r_{1}(n,D)^{\frac{1}{n-1}}}{nD[K:\mathbb{Q}]}\sum_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{n(\mathscr{X}_{\mathfrak{p}})^{\frac{1}{n-1}}}
n3+2n2+n42nD(n+1)[K:]𝔭(𝒳)logN(𝔭).\displaystyle\;-\frac{n^{3}+2n^{2}+n-4}{2nD(n+1)[K:\mathbb{Q}]}\sum_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\log N(\mathfrak{p}).

II-1. Estimate of 𝔭(𝒳)logN(𝔭)n(𝒳𝔭)1n1\sum\limits_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{n(\mathscr{X}_{\mathfrak{p}})^{\frac{1}{n-1}}}. - In order to estimate 𝔭(𝒳)logN(𝔭)n(𝒳𝔭)1n1\sum\limits_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{n(\mathscr{X}_{\mathfrak{p}})^{\frac{1}{n-1}}} in (5.2), by (21), we have

𝔭(𝒳)logN(𝔭)n(𝒳𝔭)1n1𝔭(𝒳)logN(𝔭)N(𝔭)n2δ2𝔭(𝒳)logN(𝔭)N(𝔭)32.\sum\limits_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{n(\mathscr{X}_{\mathfrak{p}})^{\frac{1}{n-1}}}\geqslant\sum_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})}-n^{2}\delta^{2}\sum_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})^{\frac{3}{2}}}.

For the estimate of 𝔭(𝒳)logN(𝔭)N(𝔭)\sum\limits_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})}, we denote

𝒬(𝒳)\displaystyle\mathcal{Q}^{\prime}(\mathscr{X}) =\displaystyle= {𝔭Spm𝒪K| 27δ4N(𝔭)r1(n,D)1n1,\displaystyle\{\mathfrak{p}\in\operatorname{Spm}\mathcal{O}_{K}|\;27\delta^{4}\leqslant N(\mathfrak{p})\leqslant r_{1}(n,D)^{\frac{1}{n-1}},
𝒳𝔽𝔭Spec𝔽𝔭 is not geometrically integral}.\displaystyle\;\mathscr{X}_{\mathbb{F}_{\mathfrak{p}}}\rightarrow\operatorname{Spec}\mathbb{F}_{\mathfrak{p}}\hbox{ is not geometrically integral}\}.

Then by (24), we have

𝔭(𝒳)logN(𝔭)N(𝔭)=27δ4N(𝔭)r1(n,D)1n1logN(𝔭)N(𝔭)𝔭𝒬(𝒳)logN(𝔭)N(𝔭)\displaystyle\sum_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})}=\sum_{27\delta^{4}\leqslant N(\mathfrak{p})\leqslant r_{1}(n,D)^{\frac{1}{n-1}}}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})}-\sum_{\mathfrak{p}\in\mathcal{Q}^{\prime}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})}
\displaystyle\geqslant 1n1logr1(n,D)3log34logδ2ϵ2(K)log(b(𝒳)),\displaystyle\frac{1}{n-1}\log r_{1}(n,D)-3\log 3-4\log\delta-2\epsilon_{2}(K)-\log\left(b^{\prime}(\mathscr{X})\right),

where the notation b(𝒳)b^{\prime}(\mathscr{X}) is introduced in Proposition 4.3.3, and ϵ2(K)\epsilon_{2}(K) is defined in (24).

For the term 𝔭(𝒳)logN(𝔭)N(𝔭)32\sum\limits_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})^{\frac{3}{2}}}, it is equal to zero when r1(n,D)1n127δ4r_{1}(n,D)^{\frac{1}{n-1}}\leqslant 27\delta^{4}. When r1(n,D)1n1>27δ4r_{1}(n,D)^{\frac{1}{n-1}}>27\delta^{4}, by (25), we have

𝔭(𝒳)logN(𝔭)N(𝔭)32\displaystyle\sum\limits_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})^{\frac{3}{2}}} \displaystyle\leqslant 27δ4N(𝔭)r1(n,D)1n1logN(𝔭)N(𝔭)32\displaystyle\sum\limits_{27\delta^{4}\leqslant N(\mathfrak{p})\leqslant r_{1}(n,D)^{\frac{1}{n-1}}}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})^{\frac{3}{2}}}
\displaystyle\leqslant 233δ22r1(n,D)12(n1)+2ϵ3(K,27δ4).\displaystyle\frac{2}{3\sqrt{3}\delta^{2}}-2r_{1}(n,D)^{-\frac{1}{2(n-1)}}+2\epsilon_{3}(K,27\delta^{4}).

By the above two estimates, we obtain

𝔭(𝒳)logN(𝔭)n(𝒳𝔭)1n1\displaystyle\sum\limits_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\frac{\log N(\mathfrak{p})}{n(\mathscr{X}_{\mathfrak{p}})^{\frac{1}{n-1}}}
\displaystyle\geqslant 1n1logr1(n,D)3log34logδlog(b(𝒳))2ϵ2(K)\displaystyle\frac{1}{n-1}\log r_{1}(n,D)-3\log 3-4\log\delta-\log\left(b^{\prime}(\mathscr{X})\right)-2\epsilon_{2}(K)
n2δ2(233δ22r1(n,D)12(n1)+2ϵ3(K,27δ4))\displaystyle\;-n^{2}\delta^{2}\left(\frac{2}{3\sqrt{3}\delta^{2}}-2r_{1}(n,D)^{-\frac{1}{2(n-1)}}+2\epsilon_{3}(K,27\delta^{4})\right)

by combining the above two inequalities.

II-2. Estimate of 𝔭(𝒳)logN(𝔭)\sum\limits_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\log N(\mathfrak{p}). - For the estimate of 𝔭(𝒳)logN(𝔭)\sum\limits_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\log N(\mathfrak{p}), by (23), we have

1D𝔭(𝒳)logN(𝔭)1DN(𝔭)r1(n,D)1n1logN(𝔭)\displaystyle\frac{1}{D}\sum\limits_{\mathfrak{p}\in\mathcal{R}(\mathscr{X})}\log N(\mathfrak{p})\leqslant\frac{1}{D}\sum\limits_{N(\mathfrak{p})\leqslant r_{1}(n,D)^{\frac{1}{n-1}}}\log N(\mathfrak{p})
\displaystyle\leqslant 1D(r1(n,D)1n1+ϵ1(K,r1(n,D)1n1)),\displaystyle\frac{1}{D}\left(r_{1}(n,D)^{\frac{1}{n-1}}+\epsilon_{1}\left(K,r_{1}(n,D)^{\frac{1}{n-1}}\right)\right),

where ϵ1(K,x)\epsilon_{1}(K,x) is defined in (23).

II-3. Deducing the contradiction. - We take (5.2) and (5.2) into (5.2), and we do some elementary calculations. Then the inequality

logB[K:](n1)!δ(2n)nh(X)+C1(n)\displaystyle\frac{\log B}{[K:\mathbb{Q}]}-\frac{(n-1)!}{\delta(2n)^{n}}h(X)+C_{1}(n)
\displaystyle\geqslant (n1)!1n1(n1)n[K:]r1(n,D)1n1D(1n1logr1(n,D)log(b(𝒳))4logδ\displaystyle\frac{(n-1)!^{\frac{1}{n-1}}(n-1)}{n[K:\mathbb{Q}]}\cdot\frac{r_{1}(n,D)^{\frac{1}{n-1}}}{D}\Bigg{(}\frac{1}{n-1}\log r_{1}(n,D)-\log\left(b^{\prime}(\mathscr{X})\right)-4\log\delta
 3log32n233+2n2δ2r1(n,D)12(n1)2n2δ2ϵ3(K,27δ4)2ϵ2(K))\displaystyle\;3\log 3-\frac{2n^{2}}{3\sqrt{3}}+\frac{2n^{2}\delta^{2}}{r_{1}(n,D)^{\frac{1}{2(n-1)}}}-2n^{2}\delta^{2}\epsilon_{3}(K,27\delta^{4})-2\epsilon_{2}(K)\Bigg{)}
n3+2n2+n42n(n+1)[K:]r1(n,D)1n1D(1+ϵ1(K,r1(n,D)1n1)r1(n,D)1n1)\displaystyle\;-\frac{n^{3}+2n^{2}+n-4}{2n(n+1)[K:\mathbb{Q}]}\cdot\frac{r_{1}(n,D)^{\frac{1}{n-1}}}{D}\left(1+\frac{\epsilon_{1}\left(K,r_{1}(n,D)^{\frac{1}{n-1}}\right)}{r_{1}(n,D)^{\frac{1}{n-1}}}\right)

is uniformly verified for all D3δlogδ+n12δ+n1D\geqslant 3\delta\log\delta+n-1\geqslant 2\delta+n-1.

From (4.4) in Lemma 4.4, we have

1n1logr1(n,D)logD+1n1logδ1n1log(n1)!\frac{1}{n-1}\log r_{1}(n,D)\geqslant\log D+\frac{1}{n-1}\log\delta-\frac{1}{n-1}\log\left(n-1\right)! (37)

when D2δ+n1D\geqslant 2\delta+n-1.

We take Lemma 4.4 and (37) into (5.2), and by the fact

ϵ1(K,r1(n,D)1n1)r1(n,D)1n1κ1(K)\frac{\epsilon_{1}\left(K,r_{1}(n,D)^{\frac{1}{n-1}}\right)}{r_{1}(n,D)^{\frac{1}{n-1}}}\leqslant\kappa_{1}(K)

and

3log32n233+2n2δ2r1(n,D)12(n1)2n2δ2ϵ3(K,27δ4)2ϵ2(K)κ2(n,K),3\log 3-\frac{2n^{2}}{3\sqrt{3}}+\frac{2n^{2}\delta^{2}}{r_{1}(n,D)^{\frac{1}{2(n-1)}}}-2n^{2}\delta^{2}\epsilon_{3}(K,27\delta^{4})-2\epsilon_{2}(K)\geqslant-\kappa_{2}(n,K),

we obtain

n(n1)δn1(logB[K:](n1)!δ(2n)nh(X)+C1(n)[K:])\displaystyle\frac{n}{(n-1)\sqrt[n-1]{\delta}}\left(\log B-\frac{[K:\mathbb{Q}](n-1)!}{\delta(2n)^{n}}h(X)+C_{1}(n)[K:\mathbb{Q}]\right)
\displaystyle\geqslant (1δ2D)(logD(41n1)logδlog(b(𝒳))\displaystyle\left(1-\frac{\delta-2}{D}\right)\Bigg{(}\log D-\left(4-\frac{1}{n-1}\right)\log\delta-\log\left(b^{\prime}(\mathscr{X})\right)
κ2(n,K)log(n1)!n1)\displaystyle\;-\kappa_{2}(n,K)-\frac{\log(n-1)!}{n-1}\Bigg{)}
n3+2n2+n42(n21)(n1)!n1(1+n4)(1+κ1(K)).\displaystyle\;-\frac{n^{3}+2n^{2}+n-4}{2(n^{2}-1)\sqrt[n-1]{(n-1)!}}\cdot\left(1+\frac{n}{4}\right)\left(1+\kappa_{1}(K)\right).

When D3δlogδ+n12δ+n1D\geqslant 3\delta\log\delta+n-1\geqslant 2\delta+n-1 and δ2\delta\geqslant 2, we have

δ2D(logD(41n1)logδlog(n1)!n1log(b(𝒳))κ2(K,n))\displaystyle\frac{\delta-2}{D}\Bigg{(}\log D-\left(4-\frac{1}{n-1}\right)\log\delta-\frac{\log(n-1)!}{n-1}-\log\left(b^{\prime}(\mathscr{X})\right)-\kappa_{2}(K,n)\Bigg{)}
\displaystyle\leqslant δ2DlogD3\displaystyle\frac{\delta-2}{D}\log D\leqslant 3

by an elementary calculation. We take the above inequality into (5.2), and then we obtain

n(n1)δn1(logB[K:](n1)!δ(2n)nh(X)+C1(n)[K:])\displaystyle\frac{n}{(n-1)\sqrt[n-1]{\delta}}\left(\log B-\frac{[K:\mathbb{Q}](n-1)!}{\delta(2n)^{n}}h(X)+C_{1}(n)[K:\mathbb{Q}]\right)
\displaystyle\geqslant logD(41n1)logδlog(b(𝒳))κ2(K,n)3log(n1)!n1\displaystyle\log D-\left(4-\frac{1}{n-1}\right)\log\delta-\log\left(b^{\prime}(\mathscr{X})\right)-\kappa_{2}(K,n)-3-\frac{\log(n-1)!}{n-1}
n3+2n2+n42(n21)(n1)!n1(1+n4)(1+κ1(K)),\displaystyle\;-\frac{n^{3}+2n^{2}+n-4}{2(n^{2}-1)\sqrt[n-1]{(n-1)!}}\cdot\left(1+\frac{n}{4}\right)\left(1+\kappa_{1}(K)\right),

which deduces

logD\displaystyle\log D \displaystyle\leqslant nlogB(n1)δn1[K:]n!δ1+1n1(n1)(2n)nh(X)+log(b(𝒳))\displaystyle\frac{n\log B}{(n-1)\sqrt[n-1]{\delta}}-\frac{[K:\mathbb{Q}]n!}{\delta^{1+\frac{1}{n-1}}(n-1)(2n)^{n}}h(X)+\log\left(b^{\prime}(\mathscr{X})\right)
+(41n1)logδ+C2(n,K)\displaystyle+\left(4-\frac{1}{n-1}\right)\log\delta+C_{2}(n,K)

with the constant C2(n,K)C_{2}(n,K) 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.

We keep all the notations and conditions in Theorem 5.2. Then there exists a hypersurface of degree smaller than

C3(n,K)δ3Bn/((n1)δ1/(n1)),C_{3}(n,K)\delta^{3}B^{{n}/{\left((n-1)\delta^{1/(n-1)}\right)}},

which covers S(X;B)S(X;B) but does not contain the generic point of XX, where the constant

C3(n,K)=eC2(n,K)(n+6)(n1)(2n)nexp(2ϵ2(K)3log3+[K:])n!,C_{3}(n,K)=e^{C_{2}(n,K)}\frac{(n+6)(n-1)(2n)^{n}\exp\left(2\epsilon_{2}(K)-3\log 3+[K:\mathbb{Q}]\right)}{n!},

and C2(n,K)C_{2}(n,K) is defined in Theorem 5.2.

Démonstration.

By the upper bound of b(𝒳)b^{\prime}(\mathscr{X}) given in Proposition 4.3.3, we have

b(𝒳)\displaystyle b^{\prime}(\mathscr{X}) \displaystyle\leqslant exp(2ϵ2(K)3log3+[K:])\displaystyle\exp\left(2\epsilon_{2}(K)-3\log 3+[K:\mathbb{Q}]\right)
(δ2δ4)(h(X)+(3logδ+δlog3+log(n+δδ)))\displaystyle\cdot(\delta^{-2}-\delta^{-4})\left(h(X)+\left(3\log\delta+\delta\log 3+\log{n+\delta\choose\delta}\right)\right)
\displaystyle\leqslant exp(2ϵ2(K)3log3+[K:])δ2(h(X)+(3δ+2δ+δlog(n+1)))\displaystyle\exp\left(2\epsilon_{2}(K)-3\log 3+[K:\mathbb{Q}]\right)\delta^{-2}\left(h(X)+(3\delta+2\delta+\delta\log(n+1))\right)
\displaystyle\leqslant (n+6)exp(2ϵ2(K)3log3+[K:])max{δ2h(X),δ1}.\displaystyle(n+6)\exp\left(2\epsilon_{2}(K)-3\log 3+[K:\mathbb{Q}]\right)\max\{\delta^{-2}h(X),\delta^{-1}\}.

We denote by GK(X)=HK(X)n!(n1)(2n)nδ11n11G_{K}(X)=H_{K}(X)^{\frac{n!}{(n-1)(2n)^{n}}\delta^{-1-\frac{1}{n-1}}}\geqslant 1 for simplicity, where the last inequality is obtained by definition directly. Then by an elementary calculation, we have

b(𝒳)HK(X)n!(n1)(2n)nδ11n1\displaystyle\frac{b^{\prime}(\mathscr{X})}{H_{K}(X)^{\frac{n!}{(n-1)(2n)^{n}}\delta^{-1-\frac{1}{n-1}}}}
\displaystyle\leqslant (n+6)exp(2ϵ2(K)3log3+[K:])max{δ1+1n1(n1)(2n)nn![K:]logGK(X),δ1}GK(X)\displaystyle(n+6)\exp\left(2\epsilon_{2}(K)-3\log 3+[K:\mathbb{Q}]\right)\frac{\max\left\{\frac{\delta^{-1+\frac{1}{n-1}}(n-1)(2n)^{n}}{n![K:\mathbb{Q}]}\log G_{K}(X),\delta^{-1}\right\}}{G_{K}(X)}
\displaystyle\leqslant (n+6)(n1)(2n)nexp(2ϵ2(K)3log3+[K:])n!δ1+1n1.\displaystyle\frac{(n+6)(n-1)(2n)^{n}\exp\left(2\epsilon_{2}(K)-3\log 3+[K:\mathbb{Q}]\right)}{n!}\delta^{-1+\frac{1}{n-1}}.

We have the assertion by taking the above estimate into Theorem 5.2. ∎

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.

We keep all the notations and conditions in Theorem 5.2. Then there exists a hypersurface of degree smaller than

C3(n,K)δ31n1HK(X)n!(n1)(2n)nδ11n1Bn/((n1)δ1/(n1))max{logB[K:],1},\frac{C^{\prime}_{3}(n,K)\delta^{3-\frac{1}{n-1}}}{H_{K}(X)^{\frac{n!}{(n-1)(2n)^{n}}\delta^{-1-\frac{1}{n-1}}}}B^{{n}/{\left((n-1)\delta^{1/(n-1)}\right)}}\max\left\{\frac{\log B}{[K:\mathbb{Q}]},1\right\},

which covers S(X;B)S(X;B) but does not contain the generic point of XX, where the constant

C3(n,K)=eC2(n,K)C1(n)(n+6)(2n)nexp(2ϵ2(K)3log3+[K:])(n1)!,C^{\prime}_{3}(n,K)=-e^{C_{2}(n,K)}\frac{C_{1}(n)(n+6)(2n)^{n}\exp\left(2\epsilon_{2}(K)-3\log 3+[K:\mathbb{Q}]\right)}{(n-1)!},

C1(n)C_{1}(n) is defined in (30), C2(n,K)C_{2}(n,K) is defined in Theorem 5.2, and HK(X)H_{K}(X) is defined in Definition 2.7.2.

Démonstration.

If

logB[K:]<(n1)!δ(2n)nh(X)+C1(n),\frac{\log B}{[K:\mathbb{Q}]}<\frac{(n-1)!}{\delta(2n)^{n}}h(X)+C_{1}(n),

then by Proposition 5.1, S(X;B)S(X;B) can be covered by a hypersurface of degree no more than 2δ+n12\delta+n-1 which does not contain the generic point of XX. The upper bound of the degree satisfies the bound provided in the statement.

If

logB[K:](n1)!δ(2n)nh(X)+C1(n),\frac{\log B}{[K:\mathbb{Q}]}\geqslant\frac{(n-1)!}{\delta(2n)^{n}}h(X)+C_{1}(n),

which is equivalent to

h(X)δ(2n)n(n1)!logB[K:]δ(2n)n(n1)!C1(n),h(X)\leqslant\frac{\delta(2n)^{n}}{(n-1)!}\cdot\frac{\log B}{[K:\mathbb{Q}]}-\frac{\delta(2n)^{n}}{(n-1)!}C_{1}(n),

then we deal with it as following. Same as the proof of Corollary 5.3, we have

b(𝒳)(n+6)exp(2ϵ2(K)3log3+[K:])max{δ2h(X),δ1},b^{\prime}(\mathscr{X})\leqslant(n+6)\exp\left(2\epsilon_{2}(K)-3\log 3+[K:\mathbb{Q}]\right)\max\{\delta^{-2}h(X),\delta^{-1}\},

where b(𝒳)b^{\prime}(\mathscr{X}) is the same as that in Proposition 4.3.3 and Theorem 5.2. Then we have

b(𝒳)HK(X)n!(n1)(2n)nδ11n1\displaystyle\frac{b^{\prime}(\mathscr{X})}{H_{K}(X)^{\frac{n!}{(n-1)(2n)^{n}}\delta^{-1-\frac{1}{n-1}}}}
\displaystyle\leqslant (n+6)exp(2ϵ2(K)3log3+[K:])δ1HK(X)n!(n1)(2n)nδ11n1max{(2n)n(n1)!(logB[K:]C1(n)),1}\displaystyle\frac{(n+6)\exp\left(2\epsilon_{2}(K)-3\log 3+[K:\mathbb{Q}]\right)\delta^{-1}}{H_{K}(X)^{\frac{n!}{(n-1)(2n)^{n}}\delta^{-1-\frac{1}{n-1}}}}\max\left\{\frac{(2n)^{n}}{(n-1)!}\left(\frac{\log B}{[K:\mathbb{Q}]}-C_{1}(n)\right),1\right\}
\displaystyle\leqslant C1(n)(n+6)(2n)nexp(2ϵ2(K)3log3+[K:])(n1)!HK(X)n!(n1)(2n)nδ11n1δ1max{logB[K:],1},\displaystyle\frac{-C_{1}(n)(n+6)(2n)^{n}\exp\left(2\epsilon_{2}(K)-3\log 3+[K:\mathbb{Q}]\right)}{(n-1)!H_{K}(X)^{\frac{n!}{(n-1)(2n)^{n}}\delta^{-1-\frac{1}{n-1}}}}\delta^{-1}\max\left\{\frac{\log B}{[K:\mathbb{Q}]},1\right\},

and we obtain the assertion by taking the above inequality into Theorem 5.2. ∎

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 XX be a geometrically integral curves in K2\mathbb{P}^{2}_{K} of degree δ\delta. Then we have

#S(X;B)C3(2,K)δ4B2/δ,\#S(X;B)\leqslant C_{3}(2,K)\delta^{4}B^{{2}/{\delta}},

where the constant C3(2,K)C_{3}(2,K) is defined in Corollary 5.3. In addition, we have

#S(X;B)Kδ4B2/δ.\#S(X;B)\ll_{K}\delta^{4}B^{{2}/{\delta}}.
Démonstration.

We apply the Bézout Theorem in the intersection theory (cf. [16, Proposition 8.4]) to XX and the auxiliary hypersurface determined in Corollary 5.3 for the case of n=2n=2, and then we obtain the result. ∎

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 XX be a geometrically integral curves in K2\mathbb{P}^{2}_{K} of degree δ\delta. Then we have

#S(X;B)C3(2,K)δ3B2/δmax{logB[K:],1},\#S(X;B)\leqslant C^{\prime}_{3}(2,K)\delta^{3}B^{{2}/{\delta}}\max\left\{\frac{\log B}{[K:\mathbb{Q}]},1\right\},

where the constant C3(2,K)C^{\prime}_{3}(2,K) is defined in Corollary 5.3. In addition, we have

#S(X;B)Kδ3B2/δlogB\#S(X;B)\ll_{K}\delta^{3}B^{{2}/{\delta}}\log B

when Bexp([K:])B\geqslant\exp([K:\mathbb{Q}]).

Démonstration.

This is the same application of Bézout Theorem in the intersection theory (cf. [16, Proposition 8.4]) to Corollary 5.3 as that of Corollary 6.1 when n=2n=2, where we take HK(X)1H_{K}(X)\geqslant 1 defined in Definition 2.7.2 into consideration. ∎

\remaname \the\smf@thm.

It seems that the upper bound given in Theorem 6.1 and Theorem 6.2 are not optimal. Actually, for a geometrically integral plane curve X2X\hookrightarrow\mathbb{P}^{2}_{\mathbb{Q}} of degree δ\delta, Heath-Brown conjectured the uniform upper bound

#S(X;B)δ2B2/δ.\#S(X;B)\ll\delta^{2}B^{{2}/{\delta}}.

By the examples given in [10, §6], the exponent 22 of δ\delta in the above conjecture would be optimal.

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 KK, we have more explicit estimates of

θK(x)=N(𝔭)xlogN(𝔭),ψK(x)=N(𝔭)xlogN(𝔭)N(𝔭), and ϕK(x)=N(𝔭)xlogN(𝔭)N(𝔭)32,\theta_{K}(x)=\sum\limits_{N(\mathfrak{p})\leqslant x}\log N(\mathfrak{p}),\;\psi_{K}(x)=\sum\limits_{N(\mathfrak{p})\leqslant x}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})}\hbox{, and }\phi_{K}(x)=\sum\limits_{N(\mathfrak{p})\leqslant x}\frac{\log N(\mathfrak{p})}{N(\mathfrak{p})^{\frac{3}{2}}},

where x+x\in\mathbb{R}^{+}, 𝔭Spm𝒪K\mathfrak{p}\in\operatorname{Spm}\mathcal{O}_{K}, and N(𝔭)=#(𝒪K/𝔭)N(\mathfrak{p})=\#(\mathcal{O}_{K}/\mathfrak{p}). 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 θK(x)\theta_{K}(x), ψK(x)\psi_{K}(x) and ϕK(x)\phi_{K}(x).

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 ΔK\Delta_{K} be the discriminant of the number field KK. By [20, Corollary 1.3], if x3x\geqslant 3, we have

|θK(x)x|\displaystyle\left|\theta_{K}(x)-x\right| \displaystyle\leqslant x((12πlog(18.8xlog2x)+2.3)logΔK\displaystyle\sqrt{x}\Bigg{(}\left(\frac{1}{2\pi}\log\left(\frac{18.8x}{\log^{2}x}\right)+2.3\right)\log\Delta_{K}
+(18πlog2(18.8xlog2x)+1.3)[K:]+0.3logx+14.6)\displaystyle\;+\left(\frac{1}{8\pi}\log^{2}\left(\frac{18.8x}{\log^{2}x}\right)+1.3\right)[K:\mathbb{Q}]+0.3\log x+14.6\Bigg{)}
\displaystyle\leqslant xlog2x(logΔK+[K:])(12π(log2x)log(18.8xlog2x)+3.6\displaystyle\sqrt{x}\log^{2}x\left(\log\Delta_{K}+[K:\mathbb{Q}]\right)\Bigg{(}\frac{1}{2\pi(\log^{2}x)}\log\left(\frac{18.8x}{\log^{2}x}\right)+3.6
+18π(log2x)log2(18.8xlog2x)+0.31logx+14.61log2x)\displaystyle\;+\frac{1}{8\pi(\log^{2}x)}\log^{2}\left(\frac{18.8x}{\log^{2}x}\right)+0.3\frac{1}{\log x}+14.6\frac{1}{\log^{2}x}\Bigg{)}

under the assumption of GRH. Since x3x\geqslant 3, we obtain

|θK(x)x|528xlog2x(logΔK+[K:])\left|\theta_{K}(x)-x\right|\leqslant 528\sqrt{x}\log^{2}x\left(\log\Delta_{K}+[K:\mathbb{Q}]\right) (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

ψK(x)=θK(x)x+2xθK(t)t2𝑑t.\psi_{K}(x)=\frac{\theta_{K}(x)}{x}+\int_{2}^{x}\frac{\theta_{K}(t)}{t^{2}}dt.

From (39), we have

|ψK(x)12x1tdt|528(logΔK+[K:])(log2xx+2xlog2tt32dt).\left|\psi_{K}(x)-1-\int_{2}^{x}\frac{1}{t}dt\right|\leqslant 528\left(\log\Delta_{K}+[K:\mathbb{Q}]\right)\left(\frac{\log^{2}x}{\sqrt{x}}+\int_{2}^{x}\frac{\log^{2}t}{t^{\frac{3}{2}}}dt\right).

Then by an elementary calculation executed by Mathematica, we obtain

|ψK(x)logx|9550(logΔK+[K:])\left|\psi_{K}(x)-\log x\right|\leqslant 9550\left(\log\Delta_{K}+[K:\mathbb{Q}]\right) (40)

for x3x\geqslant 3 under the assumption of GRH.

The same goes for the above, we have

ϕK(x)=θK(x)x32+322xθK(t)t52𝑑t.\phi_{K}(x)=\frac{\theta_{K}(x)}{x^{\frac{3}{2}}}+\frac{3}{2}\int_{2}^{x}\frac{\theta_{K}(t)}{t^{\frac{5}{2}}}dt.

Then from (39), we obtain

|ϕK(x)1x322x1t32dt|792(logΔK+[K:])(log2xx+2xlog2tt2dt).\left|\phi_{K}(x)-\frac{1}{\sqrt{x}}-\frac{3}{2}\int_{2}^{x}\frac{1}{t^{\frac{3}{2}}}dt\right|\leqslant 792\left(\log\Delta_{K}+[K:\mathbb{Q}]\right)\left(\frac{\log^{2}x}{x}+\int_{2}^{x}\frac{\log^{2}t}{t^{2}}dt\right).

Then after an elementary calculation executed by Mathematica, we have

|ϕK(x)322+2x12|2516log2xx(logΔK+[K:])\left|\phi_{K}(x)-\frac{3}{2}\sqrt{2}+\frac{2}{x^{\frac{1}{2}}}\right|\leqslant 2516\frac{\log^{2}x}{x}\left(\log\Delta_{K}+[K:\mathbb{Q}]\right) (41)

uniformly for all x+x\in\mathbb{R}^{+} 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

ϵ1(K,x)=528xlog2x(logΔK+[K:])\epsilon_{1}(K,x)=528\sqrt{x}\log^{2}x\left(\log\Delta_{K}+[K:\mathbb{Q}]\right)

in (23),

ϵ2(K)=9550(logΔK+[K:])\epsilon_{2}(K)=9550\left(\log\Delta_{K}+[K:\mathbb{Q}]\right)

in (24), and

ϵ3(K,x)=2516log2xx(logΔK+[K:])\epsilon_{3}(K,x)=2516\frac{\log^{2}x}{x}\left(\log\Delta_{K}+[K:\mathbb{Q}]\right)

in (25) under the assumption of GRH. We take the above estimates of ϵ1(K,x)\epsilon_{1}(K,x), ϵ2(K)\epsilon_{2}(K) and ϵ3(K,x)\epsilon_{3}(K,x) 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 K=K=\mathbb{Q}, 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 KK.

Annexe A An explicit lower bound of Qξ(r)Q_{\xi}(r)

In this appendix, we give an explicit lower bound of the function Qξ(r)Q_{\xi}(r) defined in (5) for the case of hypersurfaces.

In the following proof of Proposition A, the inequality

(Nm+1)mm!(Nm)(N(m1)/2)mm!\frac{(N-m+1)^{m}}{m!}\leqslant{N\choose m}\leqslant\frac{\left(N-(m-1)/2\right)^{m}}{m!}

will be very useful, where NN and mm are two positive integers, and Nm1N\geqslant m\geqslant 1.

\propname \the\smf@thm.

Let XX be a hypersurface of kn\mathbb{P}^{n}_{k}, ξ\xi be a closed point in XX, and μξ\mu_{\xi} be the multiplicity of ξ\xi in XX induced by its local Hilbert-Samuel function. The function Qξ(r)Q_{\xi}(r) is defined in the equality (5). Then we have

Qξ(r)\displaystyle Q_{\xi}(r) >\displaystyle> ((n1)!μξ)1n1(n1n)rnn1n3+2n2+n42n(n+1)r.\displaystyle\left(\frac{(n-1)!}{\mu_{\xi}}\right)^{\frac{1}{n-1}}\left(\frac{n-1}{n}\right)r^{\frac{n}{n-1}}-\frac{n^{3}+2n^{2}+n-4}{2n(n+1)}r.
Démonstration.

For the case of hypersurfaces, by [26, Example 2.70 (2)], we have

Hξ(s)=(n+s1s)(n+sμξ1sμξ).H_{\xi}(s)={n+s-1\choose s}-{n+s-\mu_{\xi}-1\choose s-\mu_{\xi}}.

We define the function Uξ(k)=Hξ(0)++Hξ(k)U_{\xi}(k)=H_{\xi}(0)+\cdots+H_{\xi}(k), then we have

Uξ(k)\displaystyle U_{\xi}(k) =\displaystyle= j=0k(n+j1j)j=0k(n+jμξ1jμξ)\displaystyle\sum\limits_{j=0}^{k}{n+j-1\choose j}-\sum\limits_{j=0}^{k}{n+j-\mu_{\xi}-1\choose j-\mu_{\xi}}
=\displaystyle= (n+kn)(n+kμξn).\displaystyle{n+k\choose n}-{n+k-\mu_{\xi}\choose n}.

Then we obtain

Qξ(Uξ(k))\displaystyle Q_{\xi}(U_{\xi}(k)) =\displaystyle= j=0njHξ(j)\displaystyle\sum\limits_{j=0}^{n}jH_{\xi}(j)
=\displaystyle= j=0kj(j+n1n1)j=0kj(n+jμξ1n1)\displaystyle\sum\limits_{j=0}^{k}j{j+n-1\choose n-1}-\sum\limits_{j=0}^{k}j{n+j-\mu_{\xi}-1\choose n-1}
=\displaystyle= n(k+nn+1)n(kμξ+nn+1)μξ(n+kμξn).\displaystyle n{k+n\choose n+1}-n{k-\mu_{\xi}+n\choose n+1}-\mu_{\xi}{n+k-\mu_{\xi}\choose n}.

Let r]Uξ(k1),Uξ(k)]r\in]U_{\xi}(k-1),U_{\xi}(k)]. By the definition of Qξ(r)Q_{\xi}(r) in the equality (5), we have the inequality

Qξ(Uξ(k1))Qξ(r)Qξ(Uξ(k)).Q_{\xi}(U_{\xi}(k-1))\leqslant Q_{\xi}(r)\leqslant Q_{\xi}(U_{\xi}(k)).

So we have

Qξ(r)\displaystyle Q_{\xi}(r) =\displaystyle= Qξ(Uξ(k1))+k(rUξ(k1))\displaystyle Q_{\xi}(U_{\xi}(k-1))+k(r-U_{\xi}(k-1)) (42)
=\displaystyle= n(n+k1n+1)n(n+kμξ1n+1)μξ(n+kμξ1n)\displaystyle n{n+k-1\choose n+1}-n{n+k-\mu_{\xi}-1\choose n+1}-\mu_{\xi}{n+k-\mu_{\xi}-1\choose n}
+krk(n+k1n)+k(n+kμξ1n)\displaystyle+kr-k{n+k-1\choose n}+k{n+k-\mu_{\xi}-1\choose n}
=\displaystyle= kr+(n+kμξn+1)(n+kn+1).\displaystyle kr+{n+k-\mu_{\xi}\choose n+1}-{n+k\choose n+1}.

In order to get a lower bound of Qξ(r)Q_{\xi}(r), we need to estimate the term

(n+kμξn+1)(n+kn+1).{n+k-\mu_{\xi}\choose n+1}-{n+k\choose n+1}.

In fact, we have the estimate

[(n+kn+1)(n+kμξn+1)]/Uξ(k1)\displaystyle\left[{n+k\choose n+1}-{n+k-\mu_{\xi}\choose n+1}\right]\Big{/}U_{\xi}(k-1)
=\displaystyle= [(n+kn+1)(n+kμξn+1)]/[(n+k1n)(n+kμξ1n)]\displaystyle\left[{n+k\choose n+1}-{n+k-\mu_{\xi}\choose n+1}\right]\Big{/}\left[{n+k-1\choose n}-{n+k-\mu_{\xi}-1\choose n}\right]
=\displaystyle= (n+k)(n+k1)k(n+kμξ)(kμξ)(n+1)[(n+k1)k(n+kμξ1)(kμξ)]\displaystyle\frac{(n+k)(n+k-1)\cdots k-(n+k-\mu_{\xi})\cdots(k-\mu_{\xi})}{(n+1)[(n+k-1)\cdots k-(n+k-\mu_{\xi}-1)\cdots(k-\mu_{\xi})]}
=\displaystyle= ((n+k)(n+k1)k1(n+k2)(k1)(n+kμξ1)(kμξ)(n+kμξ))/\displaystyle\left(\frac{(n+k)(n+k-1)}{k-1}\frac{(n+k-2)\cdots(k-1)}{(n+k-\mu_{\xi}-1)\cdots(k-\mu_{\xi})}-(n+k-\mu_{\xi})\right)\Big{/}
([n+k1k1[(n+k2)(k1)(n+kμξ1)(kμξ)]1](n+1))\displaystyle\left(\left[\frac{n+k-1}{k-1}\left[\frac{(n+k-2)\cdots(k-1)}{(n+k-\mu_{\xi}-1)\cdots(k-\mu_{\xi})}\right]-1\right](n+1)\right)
=\displaystyle= 1n+1[n+k+μξn+k1k1(n+k2)(k1)(n+kμξ1)(kμξ)1]\displaystyle\frac{1}{n+1}\left[n+k+\frac{\mu_{\xi}}{\frac{n+k-1}{k-1}\cdot\frac{(n+k-2)\cdots(k-1)}{(n+k-\mu_{\xi}-1)\cdots(k-\mu_{\xi})}-1}\right]
\displaystyle\leqslant 1n+1[n+k+μξ(n+k1n+kμξ1)n1]\displaystyle\frac{1}{n+1}\left[n+k+\frac{\mu_{\xi}}{\left(\frac{n+k-1}{n+k-\mu_{\xi}-1}\right)^{n}-1}\right]
=\displaystyle= 1n+1[n+k+(n+kμξ1)n(n+k1)n1++(n+kμξ1)n1]\displaystyle\frac{1}{n+1}\left[n+k+\frac{(n+k-\mu_{\xi}-1)^{n}}{(n+k-1)^{n-1}+\cdots+(n+k-\mu_{\xi}-1)^{n-1}}\right]
\displaystyle\leqslant 1n+1(n+k+n+kμξ1n)\displaystyle\frac{1}{n+1}\left(n+k+\frac{n+k-\mu_{\xi}-1}{n}\right)
=\displaystyle= (n+1)k+n2+nμξ1n(n+1).\displaystyle\frac{(n+1)k+n^{2}+n-\mu_{\xi}-1}{n(n+1)}.

By the equality (42), we obtain

Qξ(r)\displaystyle Q_{\xi}(r) =\displaystyle= kr+(n+kμξn+1)(n+kn+1)\displaystyle kr+{n+k-\mu_{\xi}\choose n+1}-{n+k\choose n+1}
\displaystyle\geqslant kr(n+1)k+n2+nμξ1n(n+1)Uξ(k1)\displaystyle kr-\frac{(n+1)k+n^{2}+n-\mu_{\xi}-1}{n(n+1)}U_{\xi}(k-1)
>\displaystyle> kr(n+1)k+n2+nμξ1n(n+1)r\displaystyle kr-\frac{(n+1)k+n^{2}+n-\mu_{\xi}-1}{n(n+1)}r
=\displaystyle= (n21)kn2n+μξ+1n(n+1)r\displaystyle\frac{(n^{2}-1)k-n^{2}-n+\mu_{\xi}+1}{n(n+1)}r
=\displaystyle= (n1nkn2+nμξ1n(n+1))r,\displaystyle\left(\frac{n-1}{n}k-\frac{n^{2}+n-\mu_{\xi}-1}{n(n+1)}\right)r,

where we use the estimate Uξ(k1)<rU_{\xi}(k-1)<r in the inequality (A). In addition, we obtain the inequality

rUξ(k)\displaystyle r\leqslant U_{\xi}(k) =\displaystyle= (n+kn)(n+kμξn)=j=1μξ(n+kjn1)\displaystyle{n+k\choose n}-{n+k-\mu_{\xi}\choose n}=\sum\limits_{j=1}^{\mu_{\xi}}{n+k-j\choose n-1}
\displaystyle\leqslant 1(n1)!j=1μξ(k+n2j+1)n1.\displaystyle\frac{1}{(n-1)!}\sum\limits_{j=1}^{\mu_{\xi}}(k+\frac{n}{2}-j+1)^{n-1}.

In addition, we have

j=1μξ(k+n2j+1)n1μξ(k+n2)n1.\sum\limits_{j=1}^{\mu_{\xi}}(k+\frac{n}{2}-j+1)^{n-1}\leqslant\mu_{\xi}\left(k+\frac{n}{2}\right)^{n-1}.

Then

k1μξn1(n1)!rn1n2.k\geqslant\frac{1}{\sqrt[n-1]{\mu_{\xi}}}\sqrt[n-1]{(n-1)!r}-\frac{n}{2}.

Finally we have

Qξ(r)\displaystyle Q_{\xi}(r) >\displaystyle> (n1n(1μξn1(n1)!rn1n2)n2+nμξ1n(n+1))r\displaystyle\left(\frac{n-1}{n}\left(\frac{1}{\sqrt[n-1]{\mu_{\xi}}}\sqrt[n-1]{(n-1)!r}-\frac{n}{2}\right)-\frac{n^{2}+n-\mu_{\xi}-1}{n(n+1)}\right)r
\displaystyle\geqslant ((n1)!μξ)1n1(n1n)rnn1n3+2n2+n42n(n+1)r,\displaystyle\left(\frac{(n-1)!}{\mu_{\xi}}\right)^{\frac{1}{n-1}}\left(\frac{n-1}{n}\right)r^{\frac{n}{n-1}}-\frac{n^{3}+2n^{2}+n-4}{2n(n+1)}r,

for μξ1\mu_{\xi}\geqslant 1. 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. FultonIntersection 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. HartshorneAlgebraic 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. SilvermanDiophantine geometry, Graduate Texts in Mathematics, vol. 201, Springer-Verlag, New York, 2000, An introduction.
  • [26] J. KollárLectures 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. MatsumuraCommutative 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. MoriwakiArakelov 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. NeukirchAlgebraic 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. SerreLectures 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. SouléLectures 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. TenenbaumIntroduction à la théorie analytique et probabiliste des nombres, 4-ième éd., Échelles, Belin, 2015.
  • [45] A. C. ThompsonMinkowski 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 𝔽q[t]\mathbb{F}_{q}[t]  », 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.