Linear system of divisors

Last updated
A linear system of divisors algebraicizes the classic geometric notion of a family of curves, as in the Apollonian circles. Apollonian circles.svg
A linear system of divisors algebraicizes the classic geometric notion of a family of curves, as in the Apollonian circles.

In algebraic geometry, a linear system of divisors is an algebraic generalization of the geometric notion of a family of curves; the dimension of the linear system corresponds to the number of parameters of the family.

Contents

These arose first in the form of a linear system of algebraic curves in the projective plane. It assumed a more general form, through gradual generalisation, so that one could speak of linear equivalence of divisors D on a general scheme or even a ringed space (X, OX). [1]

Linear system of dimension 1, 2, or 3 are called a pencil , a net, or a web, respectively.

A map determined by a linear system is sometimes called the Kodaira map.

Definitions

Given a general variety , two divisors are linearly equivalent if

for some non-zero rational function on , or in other words a non-zero element of the function field . Here denotes the divisor of zeroes and poles of the function .

Note that if has singular points, the notion of 'divisor' is inherently ambiguous (Cartier divisors, Weil divisors: see divisor (algebraic geometry)). The definition in that case is usually said with greater care (using invertible sheaves or holomorphic line bundles); see below.

A complete linear system on is defined as the set of all effective divisors linearly equivalent to some given divisor . It is denoted . Let be the line bundle associated to . In the case that is a nonsingular projective variety, the set is in natural bijection with [2] by associating the element of to the set of non-zero multiples of (this is well defined since two non-zero rational functions have the same divisor if and only if they are non-zero multiples of each other). A complete linear system is therefore a projective space.

A linear system is then a projective subspace of a complete linear system, so it corresponds to a vector subspace W of The dimension of the linear system is its dimension as a projective space. Hence .

Linear systems can also be introduced by means of the line bundle or invertible sheaf language. In those terms, divisors (Cartier divisors, to be precise) correspond to line bundles, and linear equivalence of two divisors means that the corresponding line bundles are isomorphic.

Examples

Linear equivalence

Consider the line bundle on whose sections define quadric surfaces. For the associated divisor , it is linearly equivalent to any other divisor defined by the vanishing locus of some using the rational function [2] (Proposition 7.2). For example, the divisor associated to the vanishing locus of is linearly equivalent to the divisor associated to the vanishing locus of . Then, there is the equivalence of divisors

Linear systems on curves

One of the important complete linear systems on an algebraic curve of genus is given by the complete linear system associated with the canonical divisor , denoted . This definition follows from proposition II.7.7 of Hartshorne [2] since every effective divisor in the linear system comes from the zeros of some section of .

Hyperelliptic curves

One application of linear systems is used in the classification of algebraic curves. A hyperelliptic curve is a curve with a degree morphism . [2] For the case all curves are hyperelliptic: the Riemann–Roch theorem then gives the degree of is and , hence there is a degree map to .

grd

A is a linear system on a curve which is of degree and dimension . For example, hyperelliptic curves have a since defines one. In fact, hyperelliptic curves have a unique [2] from proposition 5.3. Another close set of examples are curves with a which are called trigonal curves. In fact, any curve has a for . [3]

Linear systems of hypersurfaces in a projective space

Consider the line bundle over . If we take global sections , then we can take its projectivization . This is isomorphic to where

Then, using any embedding we can construct a linear system of dimension .

Linear system of conics

Characteristic linear system of a family of curves

The characteristic linear system of a family of curves on an algebraic surface Y for a curve C in the family is a linear system formed by the curves in the family that are infinitely near C. [4]

In modern terms, it is a subsystem of the linear system associated to the normal bundle to . Note a characteristic system need not to be complete; in fact, the question of completeness is something studied extensively by the Italian school without a satisfactory conclusion; nowadays, the Kodaira–Spencer theory can be used to answer the question of the completeness.

Other examples

The Cayley–Bacharach theorem is a property of a pencil of cubics, which states that the base locus satisfies an "8 implies 9" property: any cubic containing 8 of the points necessarily contains the 9th.

Linear systems in birational geometry

In general linear systems became a basic tool of birational geometry as practised by the Italian school of algebraic geometry. The technical demands became quite stringent; later developments clarified a number of issues. The computation of the relevant dimensions the Riemann–Roch problem as it can be called can be better phrased in terms of homological algebra. The effect of working on varieties with singular points is to show up a difference between Weil divisors (in the free abelian group generated by codimension-one subvarieties), and Cartier divisors coming from sections of invertible sheaves.

The Italian school liked to reduce the geometry on an algebraic surface to that of linear systems cut out by surfaces in three-space; Zariski wrote his celebrated book Algebraic Surfaces to try to pull together the methods, involving linear systems with fixed base points. There was a controversy, one of the final issues in the conflict between 'old' and 'new' points of view in algebraic geometry, over Henri Poincaré's characteristic linear system of an algebraic family of curves on an algebraic surface.

Base locus

The base locus of a linear system of divisors on a variety refers to the subvariety of points 'common' to all divisors in the linear system. Geometrically, this corresponds to the common intersection of the varieties. Linear systems may or may not have a base locus – for example, the pencil of affine lines has no common intersection, but given two (nondegenerate) conics in the complex projective plane, they intersect in four points (counting with multiplicity) and thus the pencil they define has these points as base locus.

More precisely, suppose that is a complete linear system of divisors on some variety . Consider the intersection

where denotes the support of a divisor, and the intersection is taken over all effective divisors in the linear system. This is the base locus of (as a set, at least: there may be more subtle scheme-theoretic considerations as to what the structure sheaf of should be).

One application of the notion of base locus is to nefness of a Cartier divisor class (i.e. complete linear system). Suppose is such a class on a variety , and an irreducible curve on . If is not contained in the base locus of , then there exists some divisor in the class which does not contain , and so intersects it properly. Basic facts from intersection theory then tell us that we must have . The conclusion is that to check nefness of a divisor class, it suffices to compute the intersection number with curves contained in the base locus of the class. So, roughly speaking, the 'smaller' the base locus, the 'more likely' it is that the class is nef.

In the modern formulation of algebraic geometry, a complete linear system of (Cartier) divisors on a variety is viewed as a line bundle on . From this viewpoint, the base locus is the set of common zeroes of all sections of . A simple consequence is that the bundle is globally generated if and only if the base locus is empty.

The notion of the base locus still makes sense for a non-complete linear system as well: the base locus of it is still the intersection of the supports of all the effective divisors in the system.

Example

Consider the Lefschetz pencil given by two generic sections , so given by the scheme

This has an associated linear system of divisors since each polynomial, for a fixed is a divisor in . Then, the base locus of this system of divisors is the scheme given by the vanishing locus of , so

A map determined by a linear system

Each linear system on an algebraic variety determines a morphism from the complement of the base locus to a projective space of dimension of the system, as follows. (In a sense, the converse is also true; see the section below)

Let L be a line bundle on an algebraic variety X and a finite-dimensional vector subspace. For the sake of clarity, we first consider the case when V is base-point-free; in other words, the natural map is surjective (here, k = the base field). Or equivalently, is surjective. Hence, writing for the trivial vector bundle and passing the surjection to the relative Proj, there is a closed immersion:

where on the right is the invariance of the projective bundle under a twist by a line bundle. Following i by a projection, there results in the map: [5]

When the base locus of V is not empty, the above discussion still goes through with in the direct sum replaced by an ideal sheaf defining the base locus and X replaced by the blow-up of it along the (scheme-theoretic) base locus B. Precisely, as above, there is a surjection where is the ideal sheaf of B and that gives rise to

Since an open subset of , there results in the map:

Finally, when a basis of V is chosen, the above discussion becomes more down-to-earth (and that is the style used in Hartshorne, Algebraic Geometry).

Linear system determined by a map to a projective space

Each morphism from an algebraic variety to a projective space determines a base-point-free linear system on the variety; because of this, a base-point-free linear system and a map to a projective space are often used interchangeably.

For a closed immersion of algebraic varieties there is a pullback of a linear system on to , defined as [2] (page 158).

O(1) on a projective variety

A projective variety embedded in has a natural linear system determining a map to projective space from . This sends a point to its corresponding point .

See also

Related Research Articles

The Riemann–Roch theorem is an important theorem in mathematics, specifically in complex analysis and algebraic geometry, for the computation of the dimension of the space of meromorphic functions with prescribed zeros and allowed poles. It relates the complex analysis of a connected compact Riemann surface with the surface's purely topological genus g, in a way that can be carried over into purely algebraic settings.

<span class="mw-page-title-main">Algebraic variety</span> Mathematical object studied in the field of algebraic geometry

Algebraic varieties are the central objects of study in algebraic geometry, a sub-field of mathematics. Classically, an algebraic variety is defined as the set of solutions of a system of polynomial equations over the real or complex numbers. Modern definitions generalize this concept in several different ways, while attempting to preserve the geometric intuition behind the original definition.

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

In algebraic geometry, a projective variety over an algebraically closed field k is a subset of some projective n-space over k that is the zero-locus of some finite family of homogeneous polynomials of n + 1 variables with coefficients in k, that generate a prime ideal, the defining ideal of the variety. Equivalently, an algebraic variety is projective if it can be embedded as a Zariski closed subvariety of .

In mathematics, especially in algebraic geometry and the theory of complex manifolds, coherent sheaves are a class of sheaves closely linked to the geometric properties of the underlying space. The definition of coherent sheaves is made with reference to a sheaf of rings that codifies this geometric information.

In mathematics, Kähler differentials provide an adaptation of differential forms to arbitrary commutative rings or schemes. The notion was introduced by Erich Kähler in the 1930s. It was adopted as standard in commutative algebra and algebraic geometry somewhat later, once the need was felt to adapt methods from calculus and geometry over the complex numbers to contexts where such methods are not available.

In mathematics, deformation theory is the study of infinitesimal conditions associated with varying a solution P of a problem to slightly different solutions Pε, where ε is a small number, or a vector of small quantities. The infinitesimal conditions are the result of applying the approach of differential calculus to solving a problem with constraints. The name is an analogy to non-rigid structures that deform slightly to accommodate external forces.

In mathematics, a distinctive feature of algebraic geometry is that some line bundles on a projective variety can be considered "positive", while others are "negative". The most important notion of positivity is that of an ample line bundle, although there are several related classes of line bundles. Roughly speaking, positivity properties of a line bundle are related to having many global sections. Understanding the ample line bundles on a given variety X amounts to understanding the different ways of mapping X into projective space. In view of the correspondence between line bundles and divisors, there is an equivalent notion of an ample divisor.

In algebraic geometry, divisors are a generalization of codimension-1 subvarieties of algebraic varieties. Two different generalizations are in common use, Cartier divisors and Weil divisors. Both are derived from the notion of divisibility in the integers and algebraic number fields.

<span class="mw-page-title-main">Blowing up</span>

In mathematics, blowing up or blowup is a type of geometric transformation which replaces a subspace of a given space with the space of all directions pointing out of that subspace. For example, the blowup of a point in a plane replaces the point with the projectivized tangent space at that point. The metaphor is that of zooming in on a photograph to enlarge part of the picture, rather than referring to an explosion.

In mathematics, the Hirzebruch–Riemann–Roch theorem, named after Friedrich Hirzebruch, Bernhard Riemann, and Gustav Roch, is Hirzebruch's 1954 result generalizing the classical Riemann–Roch theorem on Riemann surfaces to all complex algebraic varieties of higher dimensions. The result paved the way for the Grothendieck–Hirzebruch–Riemann–Roch theorem proved about three years later.

In mathematics, the tautological bundle is a vector bundle occurring over a Grassmannian in a natural tautological way: for a Grassmannian of -dimensional subspaces of , given a point in the Grassmannian corresponding to a -dimensional vector subspace , the fiber over is the subspace itself. In the case of projective space the tautological bundle is known as the tautological line bundle.

In mathematics, especially in algebraic geometry and the theory of complex manifolds, the adjunction formula relates the canonical bundle of a variety and a hypersurface inside that variety. It is often used to deduce facts about varieties embedded in well-behaved spaces such as projective space or to prove theorems by induction.

In algebraic geometry, a line bundle on a projective variety is nef if it has nonnegative degree on every curve in the variety. The classes of nef line bundles are described by a convex cone, and the possible contractions of the variety correspond to certain faces of the nef cone. In view of the correspondence between line bundles and divisors, there is an equivalent notion of a nef divisor.

In algebraic geometry, a morphism of schemes generalizes a morphism of algebraic varieties just as a scheme generalizes an algebraic variety. It is, by definition, a morphism in the category of schemes.

In mathematics, Arakelov theory is an approach to Diophantine geometry, named for Suren Arakelov. It is used to study Diophantine equations in higher dimensions.

In mathematics, the Segre class is a characteristic class used in the study of cones, a generalization of vector bundles. For vector bundles the total Segre class is inverse to the total Chern class, and thus provides equivalent information; the advantage of the Segre class is that it generalizes to more general cones, while the Chern class does not. The Segre class was introduced in the non-singular case by Segre (1953).. In the modern treatment of intersection theory in algebraic geometry, as developed e.g. in the definitive book of Fulton (1998), Segre classes play a fundamental role.

In mathematics, especially in algebraic geometry and the theory of complex manifolds, coherent sheaf cohomology is a technique for producing functions with specified properties. Many geometric questions can be formulated as questions about the existence of sections of line bundles or of more general coherent sheaves; such sections can be viewed as generalized functions. Cohomology provides computable tools for producing sections, or explaining why they do not exist. It also provides invariants to distinguish one algebraic variety from another.

The concept of a Projective space plays a central role in algebraic geometry. This article aims to define the notion in terms of abstract algebraic geometry and to describe some basic uses of projective spaces.

This is a glossary of algebraic geometry.

In algebraic geometry, a derived scheme is a pair consisting of a topological space X and a sheaf either of simplicial commutative rings or of commutative ring spectra on X such that (1) the pair is a scheme and (2) is a quasi-coherent -module. The notion gives a homotopy-theoretic generalization of a scheme.

References

  1. Grothendieck, Alexandre; Dieudonné, Jean. EGA IV, 21.3.
  2. 1 2 3 4 5 6 Hartshorne, R. 'Algebraic Geometry', proposition II.7.2, page 151, proposition II.7.7, page 157, page 158, exercise IV.1.7, page 298, proposition IV.5.3, page 342
  3. Kleiman, Steven L.; Laksov, Dan (1974). "Another proof of the existence of special divisors". Acta Mathematica. 132: 163–176. doi: 10.1007/BF02392112 . ISSN   0001-5962.
  4. Arbarello, Enrico; Cornalba, Maurizio; Griffiths, Phillip (2011). Geometry of algebraic curves. Grundlehren der Mathematischen Wissenschaften. Vol. II, with a contribution by Joseph Daniel Harris. Heidelberg: Springer. p. 3. doi:10.1007/978-1-4757-5323-3. ISBN   978-1-4419-2825-2. MR   2807457.
  5. Fulton, William (1998). "§ 4.4. Linear Systems". Intersection Theory. Springer. doi:10.1007/978-1-4612-1700-8_5.