Rational point

Last updated

In number theory and algebraic geometry, a rational point of an algebraic variety is a point whose coordinates belong to a given field. If the field is not mentioned, the field of rational numbers is generally understood. If the field is the field of real numbers, a rational point is more commonly called a real point.

Contents

Understanding rational points is a central goal of number theory and Diophantine geometry. For example, Fermat's Last Theorem may be restated as: for n > 2, the Fermat curve of equation has no other rational points than (1, 0), (0, 1), and, if n is even, (–1, 0) and (0, –1).

Definition

Given a field k, and an algebraically closed extension K of k, an affine variety X over k is the set of common zeros in Kn of a collection of polynomials with coefficients in k:

These common zeros are called the points of X.

A k-rational point (or k-point) of X is a point of X that belongs to kn, that is, a sequence of n elements of k such that for all j. The set of k-rational points of X is often denoted X(k).

Sometimes, when the field k is understood, or when k is the field of rational numbers, one says "rational point" instead of "k-rational point".

For example, the rational points of the unit circle of equation

are the pairs of rational numbers

where (a, b, c) is a Pythagorean triple.

The concept also makes sense in more general settings. A projective variety X in projective space over a field k can be defined by a collection of homogeneous polynomial equations in variables A k-point of written is given by a sequence of n + 1 elements of k, not all zero, with the understanding that multiplying all of by the same nonzero element of k gives the same point in projective space. Then a k-point of X means a k-point of at which the given polynomials vanish.

More generally, let X be a scheme over a field k. This means that a morphism of schemes f: XSpec(k) is given. Then a k-point of X means a section of this morphism, that is, a morphism a: Spec(k) → X such that the composition fa is the identity on Spec(k). This agrees with the previous definitions when X is an affine or projective variety (viewed as a scheme over k).

When X is a variety over an algebraically closed field k, much of the structure of X is determined by its set X(k) of k-rational points. For a general field k, however, X(k) gives only partial information about X. In particular, for a variety X over a field k and any field extension E of k, X also determines the set X(E) of E-rational points of X, meaning the set of solutions of the equations defining X with values in E.

Example: Let X be the conic curve in the affine plane A2 over the real numbers Then the set of real points is empty, because the square of any real number is nonnegative. On the other hand, in the terminology of algebraic geometry, the algebraic variety X over is not empty, because the set of complex points is not empty.

More generally, for a scheme X over a commutative ring R and any commutative R-algebra S, the set X(S) of S-points of X means the set of morphisms Spec(S) → X over Spec(R). The scheme X is determined up to isomorphism by the functor SX(S); this is the philosophy of identifying a scheme with its functor of points. Another formulation is that the scheme X over R determines a scheme XS over S by base change, and the S-points of X (over R) can be identified with the S-points of XS (over S).

The theory of Diophantine equations traditionally meant the study of integral points, meaning solutions of polynomial equations in the integers rather than the rationals For homogeneous polynomial equations such as the two problems are essentially equivalent, since every rational point can be scaled to become an integral point.

Rational points on curves

Much of number theory can be viewed as the study of rational points of algebraic varieties, a convenient setting being smooth projective varieties. For smooth projective curves, the behavior of rational points depends strongly on the genus of the curve.

Genus 0

Every smooth projective curve X of genus zero over a field k is isomorphic to a conic (degree 2) curve in If X has a k-rational point, then it is isomorphic to over k, and so its k-rational points are completely understood. [1] If k is the field of rational numbers (or more generally a number field), there is an algorithm to determine whether a given conic has a rational point, based on the Hasse principle: a conic over has a rational point if and only if it has a point over all completions of that is, over and all p-adic fields

Genus 1

It is harder to determine whether a curve of genus 1 has a rational point. The Hasse principle fails in this case: for example, by Ernst Selmer, the cubic curve in has a point over all completions of but no rational point. [2] The failure of the Hasse principle for curves of genus 1 is measured by the Tate–Shafarevich group.

If X is a curve of genus 1 with a k-rational point p0, then X is called an elliptic curve over k. In this case, X has the structure of a commutative algebraic group (with p0 as the zero element), and so the set X(k) of k-rational points is an abelian group. The Mordell–Weil theorem says that for an elliptic curve (or, more generally, an abelian variety) X over a number field k, the abelian group X(k) is finitely generated. Computer algebra programs can determine the Mordell–Weil group X(k) in many examples, but it is not known whether there is an algorithm that always succeeds in computing this group. That would follow from the conjecture that the Tate–Shafarevich group is finite, or from the related Birch–Swinnerton-Dyer conjecture. [3]

Genus at least 2

Faltings's theorem (formerly the Mordell conjecture) says that for any curve X of genus at least 2 over a number field k, the set X(k) is finite. [4]

Some of the great achievements of number theory amount to determining the rational points on particular curves. For example, Fermat's Last Theorem (proved by Richard Taylor and Andrew Wiles) is equivalent to the statement that for an integer n at least 3, the only rational points of the curve in over are the obvious ones: [0,1,1] and [1,0,1]; [0,1,−1] and [1,0,−1] for n even; and [1,−1,0] for n odd. The curve X (like any smooth curve of degree n in ) has genus

It is not known whether there is an algorithm to find all the rational points on an arbitrary curve of genus at least 2 over a number field. There is an algorithm that works in some cases. Its termination in general would follow from the conjectures that the Tate–Shafarevich group of an abelian variety over a number field is finite and that the Brauer–Manin obstruction is the only obstruction to the Hasse principle, in the case of curves. [5]

Higher dimensions

Varieties with few rational points

In higher dimensions, one unifying goal is the BombieriLang conjecture that, for any variety X of general type over a number field k, the set of k-rational points of X is not Zariski dense in X. (That is, the k-rational points are contained in a finite union of lower-dimensional subvarieties of X.) In dimension 1, this is exactly Faltings's theorem, since a curve is of general type if and only if it has genus at least 2. Lang also made finer conjectures relating finiteness of rational points to Kobayashi hyperbolicity. [6]

For example, the Bombieri–Lang conjecture predicts that a smooth hypersurface of degree d in projective space over a number field does not have Zariski dense rational points if dn + 2. Not much is known about that case. The strongest known result on the Bombieri–Lang conjecture is Faltings's theorem on subvarieties of abelian varieties (generalizing the case of curves). Namely, if X is a subvariety of an abelian variety A over a number field k, then all k-rational points of X are contained in a finite union of translates of abelian subvarieties contained in X. [7] (So if X contains no translated abelian subvarieties of positive dimension, then X(k) is finite.)

Varieties with many rational points

In the opposite direction, a variety X over a number field k is said to have potentially dense rational points if there is a finite extension field E of k such that the E-rational points of X are Zariski dense in X. Frédéric Campana conjectured that a variety is potentially dense if and only if it has no rational fibration over a positive-dimensional orbifold of general type. [8] A known case is that every cubic surface in over a number field k has potentially dense rational points, because (more strongly) it becomes rational over some finite extension of k (unless it is the cone over a plane cubic curve). Campana's conjecture would also imply that a K3 surface X (such as a smooth quartic surface in ) over a number field has potentially dense rational points. That is known only in special cases, for example if X has an elliptic fibration. [9]

One may ask when a variety has a rational point without extending the base field. In the case of a hypersurface X of degree d in over a number field, there are good results when d is much smaller than n, often based on the Hardy–Littlewood circle method. For example, the Hasse–Minkowski theorem says that the Hasse principle holds for quadric hypersurfaces over a number field (the case d = 2). Christopher Hooley proved the Hasse principle for smooth cubic hypersurfaces in over when n ≥ 8. [10] In higher dimensions, even more is true: every smooth cubic in over has a rational point when n ≥ 9, by Roger Heath-Brown. [11] More generally, Birch's theorem says that for any odd positive integer d, there is an integer N such that for all nN, every hypersurface of degree d in over has a rational point.

For hypersurfaces of smaller dimension (in terms of their degree), things can be more complicated. For example, the Hasse principle fails for the smooth cubic surface in over by Ian Cassels and Richard Guy. [12] Jean-Louis Colliot-Thélène has conjectured that the Brauer–Manin obstruction is the only obstruction to the Hasse principle for cubic surfaces. More generally, that should hold for every rationally connected variety over a number field. [13]

In some cases, it is known that X has "many" rational points whenever it has one. For example, extending work of Beniamino Segre and Yuri Manin, János Kollár showed: for a cubic hypersurface X of dimension at least 2 over a perfect field k with X not a cone, X is unirational over k if it has a k-rational point. [14] (In particular, for k infinite, unirationality implies that the set of k-rational points is Zariski dense in X.) The Manin conjecture is a more precise statement that would describe the asymptotics of the number of rational points of bounded height on a Fano variety.

Counting points over finite fields

A variety X over a finite field k has only finitely many k-rational points. The Weil conjectures, proved by André Weil in dimension 1 and by Pierre Deligne in any dimension, give strong estimates for the number of k-points in terms of the Betti numbers of X. For example, if X is a smooth projective curve of genus g over a field k of order q (a prime power), then

For a smooth hypersurface X of degree d in over a field k of order q, Deligne's theorem gives the bound: [15]

There are also significant results about when a projective variety over a finite field k has at least one k-rational point. For example, the Chevalley–Warning theorem implies that any hypersurface X of degree d in over a finite field k has a k-rational point if dn. For smooth X, this also follows from Hélène Esnault's theorem that every smooth projective rationally chain connected variety, for example every Fano variety, over a finite field k has a k-rational point. [16]

See also

Notes

  1. Hindry & Silverman (2000), Theorem A.4.3.1.
  2. Silverman (2009), Remark X.4.11.
  3. Silverman (2009), Conjecture X.4.13.
  4. Hindry & Silverman (2000), Theorem E.0.1.
  5. Skorobogatov (2001), section 6,3.
  6. Hindry & Silverman (2000), section F.5.2.
  7. Hindry & Silverman (2000), Theorem F.1.1.1.
  8. Campana (2004), Conjecture 9.20.
  9. Hassett (2003), Theorem 6.4.
  10. Hooley (1988), Theorem.
  11. Heath-Brown (1983), Theorem.
  12. Colliot-Thélène, Kanevsky & Sansuc (1987), section 7.
  13. Colliot-Thélène (2015), section 6.1.
  14. Kollár (2002), Theorem 1.1.
  15. Katz (1980), section II.
  16. Esnault (2003), Corollary 1.3.

Related Research Articles

<span class="mw-page-title-main">Elliptic curve</span> Algebraic curve

In mathematics, an elliptic curve is a smooth, projective, algebraic curve of genus one, on which there is a specified point O. An elliptic curve is defined over a field K and describes points in K2, the Cartesian product of K with itself. If the field's characteristic is different from 2 and 3, then the curve can be described as a plane algebraic curve which consists of solutions (x, y) for:

<span class="mw-page-title-main">Faltings's theorem</span> Curves of genus > 1 over the rationals have only finitely many rational points

Faltings's theorem is a result in arithmetic geometry, according to which a curve of genus greater than 1 over the field of rational numbers has only finitely many rational points. This was conjectured in 1922 by Louis Mordell, and known as the Mordell conjecture until its 1983 proof by Gerd Faltings. The conjecture was later generalized by replacing by any number field.

In mathematics, the Weil conjectures were highly influential proposals by André Weil (1949). They led to a successful multi-decade program to prove them, in which many leading researchers developed the framework of modern algebraic geometry and number theory.

<span class="mw-page-title-main">Abelian variety</span>

In mathematics, particularly in algebraic geometry, complex analysis and algebraic number theory, an abelian variety is a projective algebraic variety that is also an algebraic group, i.e., has a group law that can be defined by regular functions. Abelian varieties are at the same time among the most studied objects in algebraic geometry and indispensable tools for much research on other topics in algebraic geometry and number theory.

<span class="mw-page-title-main">Birational geometry</span> Field of algebraic geometry

In mathematics, birational geometry is a field of algebraic geometry in which the goal is to determine when two algebraic varieties are isomorphic outside lower-dimensional subsets. This amounts to studying mappings that are given by rational functions rather than polynomials; the map may fail to be defined where the rational functions have poles.

In mathematics, the étale cohomology groups of an algebraic variety or scheme are algebraic analogues of the usual cohomology groups with finite coefficients of a topological space, introduced by Grothendieck in order to prove the Weil conjectures. Étale cohomology theory can be used to construct ℓ-adic cohomology, which is an example of a Weil cohomology theory in algebraic geometry. This has many applications, such as the proof of the Weil conjectures and the construction of representations of finite groups of Lie type.

In mathematics, the Brauer group of a field K is an abelian group whose elements are Morita equivalence classes of central simple algebras over K, with addition given by the tensor product of algebras. It was defined by the algebraist Richard Brauer.

<span class="mw-page-title-main">Diophantine geometry</span> Mathematics of varieties with integer coordinates

In mathematics, Diophantine geometry is the study of Diophantine equations by means of powerful methods in algebraic geometry. By the 20th century it became clear for some mathematicians that methods of algebraic geometry are ideal tools to study these equations. Diophantine geometry is part of the broader field of arithmetic geometry.

In geometry, a hypersurface is a generalization of the concepts of hyperplane, plane curve, and surface. A hypersurface is a manifold or an algebraic variety of dimension n − 1, which is embedded in an ambient space of dimension n, generally a Euclidean space, an affine space or a projective space. Hypersurfaces share, with surfaces in a three-dimensional space, the property of being defined by a single implicit equation, at least locally, and sometimes globally.

In mathematics, a rational variety is an algebraic variety, over a given field K, which is birationally equivalent to a projective space of some dimension over K. This means that its function field is isomorphic to

In algebraic geometry, the Kodaira dimensionκ(X) measures the size of the canonical model of a projective variety X.

In algebraic geometry, a Fano variety, introduced by Gino Fano in, is a complete variety X whose anticanonical bundle KX* is ample. In this definition, one could assume that X is smooth over a field, but the minimal model program has also led to the study of Fano varieties with various types of singularities, such as terminal or klt singularities. Recently techniques in differential geometry have been applied to the study of Fano varieties over the complex numbers, and success has been found in constructing moduli spaces of Fano varieties and proving the existence of Kähler–Einstein metrics on them through the study of K-stability of Fano varieties.

<span class="mw-page-title-main">Tate conjecture</span>

In number theory and algebraic geometry, the Tate conjecture is a 1963 conjecture of John Tate that would describe the algebraic cycles on a variety in terms of a more computable invariant, the Galois representation on étale cohomology. The conjecture is a central problem in the theory of algebraic cycles. It can be considered an arithmetic analog of the Hodge conjecture.

This is a glossary of arithmetic and diophantine geometry in mathematics, areas growing out of the traditional study of Diophantine equations to encompass large parts of number theory and algebraic geometry. Much of the theory is in the form of proposed conjectures, which can be related at various levels of generality.

In algebraic geometry, the Chow groups of an algebraic variety over any field are algebro-geometric analogs of the homology of a topological space. The elements of the Chow group are formed out of subvarieties in a similar way to how simplicial or cellular homology groups are formed out of subcomplexes. When the variety is smooth, the Chow groups can be interpreted as cohomology groups and have a multiplication called the intersection product. The Chow groups carry rich information about an algebraic variety, and they are correspondingly hard to compute in general.

Arithmetic dynamics is a field that amalgamates two areas of mathematics, dynamical systems and number theory. Classically, discrete dynamics refers to the study of the iteration of self-maps of the complex plane or real line. Arithmetic dynamics is the study of the number-theoretic properties of integer, rational, p-adic, and/or algebraic points under repeated application of a polynomial or rational function. A fundamental goal is to describe arithmetic properties in terms of underlying geometric structures.

In algebraic geometry, a variety over a field k is ruled if it is birational to the product of the projective line with some variety over k. A variety is uniruled if it is covered by a family of rational curves. The concept arose from the ruled surfaces of 19th-century geometry, meaning surfaces in affine space or projective space which are covered by lines. Uniruled varieties can be considered to be relatively simple among all varieties, although there are many of them.

In arithmetic geometry, the Tate–Shafarevich groupШ(A/K) of an abelian variety A (or more generally a group scheme) defined over a number field K consists of the elements of the Weil–Châtelet group WC(A/K) = H1(GK, A) that become trivial in all of the completions of K (i.e. the p-adic fields obtained from K, as well as its real and complex completions). Thus, in terms of Galois cohomology, it can be written as

In mathematics and especially complex geometry, the Kobayashi metric is a pseudometric intrinsically associated to any complex manifold. It was introduced by Shoshichi Kobayashi in 1967. Kobayashi hyperbolic manifolds are an important class of complex manifolds, defined by the property that the Kobayashi pseudometric is a metric. Kobayashi hyperbolicity of a complex manifold X implies that every holomorphic map from the complex line C to X is constant.

In algebraic geometry, a smooth scheme over a field is a scheme which is well approximated by affine space near any point. Smoothness is one way of making precise the notion of a scheme with no singular points. A special case is the notion of a smooth variety over a field. Smooth schemes play the role in algebraic geometry of manifolds in topology.

References