Morphism of algebraic varieties

Last updated

In algebraic geometry, a morphism between algebraic varieties is a function between the varieties that is given locally by polynomials. It is also called a regular map. A morphism from an algebraic variety to the affine line is also called a regular function. A regular map whose inverse is also regular is called biregular, and the biregular maps are the isomorphisms of algebraic varieties. Because regular and biregular are very restrictive conditions – there are no non-constant regular functions on projective varieties – the concepts of rational and birational maps are widely used as well; they are partial functions that are defined locally by rational fractions instead of polynomials.

Contents

An algebraic variety has naturally the structure of a locally ringed space; a morphism between algebraic varieties is precisely a morphism of the underlying locally ringed spaces.

Definition

If X and Y are closed subvarieties of and (so they are affine varieties), then a regular map is the restriction of a polynomial map . Explicitly, it has the form: [1]

where the s are in the coordinate ring of X:

where I is the ideal defining X (note: two polynomials f and g define the same function on X if and only if f  g is in I). The image f(X) lies in Y, and hence satisfies the defining equations of Y. That is, a regular map is the same as the restriction of a polynomial map whose components satisfy the defining equations of .

More generally, a map f:XY between two varieties is regular at a pointx if there is a neighbourhood U of x and a neighbourhood V of f(x) such that f(U) ⊂ V and the restricted function f:UV is regular as a function on some affine charts of U and V. Then f is called regular, if it is regular at all points of X.

The composition of regular maps is again regular; thus, algebraic varieties form the category of algebraic varieties where the morphisms are the regular maps.

Regular maps between affine varieties correspond contravariantly in one-to-one to algebra homomorphisms between the coordinate rings: if f:XY is a morphism of affine varieties, then it defines the algebra homomorphism

where are the coordinate rings of X and Y; it is well-defined since is a polynomial in elements of . Conversely, if is an algebra homomorphism, then it induces the morphism

given by: writing

where are the images of 's. [lower-alpha 3] Note as well as [lower-alpha 4] In particular, f is an isomorphism of affine varieties if and only if f# is an isomorphism of the coordinate rings.

For example, if X is a closed subvariety of an affine variety Y and f is the inclusion, then f# is the restriction of regular functions on Y to X. See #Examples below for more examples.

Regular functions

In the particular case that Y equals A1 the regular maps f:XA1 are called regular functions, and are algebraic analogs of smooth functions studied in differential geometry. The ring of regular functions (that is the coordinate ring or more abstractly the ring of global sections of the structure sheaf) is a fundamental object in affine algebraic geometry. The only regular function on a projective variety is constant (this can be viewed as an algebraic analogue of Liouville's theorem in complex analysis).

A scalar function f:XA1 is regular at a point x if, in some open affine neighborhood of x, it is a rational function that is regular at x; i.e., there are regular functions g, h near x such that f = g/h and h does not vanish at x. [lower-alpha 5] Caution: the condition is for some pair (g, h) not for all pairs (g, h); see Examples.

If X is a quasi-projective variety; i.e., an open subvariety of a projective variety, then the function field k(X) is the same as that of the closure of X and thus a rational function on X is of the form g/h for some homogeneous elements g, h of the same degree in the homogeneous coordinate ring of (cf. Projective variety#Variety structure.) Then a rational function f on X is regular at a point x if and only if there are some homogeneous elements g, h of the same degree in such that f = g/h and h does not vanish at x. This characterization is sometimes taken as the definition of a regular function. [2]

Comparison with a morphism of schemes

If X = Spec A and Y = Spec B are affine schemes, then each ring homomorphism φ : BA determines a morphism

by taking the pre-images of prime ideals. All morphisms between affine schemes are of this type and gluing such morphisms gives a morphism of schemes in general.

Now, if X, Y are affine varieties; i.e., A, B are integral domains that are finitely generated algebras over an algebraically closed field k, then, working with only the closed points, the above coincides with the definition given at #Definition. (Proof: If f : XY is a morphism, then writing , we need to show

where are the maximal ideals corresponding to the points x and f(x); i.e., . This is immediate.)

This fact means that the category of affine varieties can be identified with a full subcategory of affine schemes over k. Since morphisms of varieties are obtained by gluing morphisms of affine varieties in the same way morphisms of schemes are obtained by gluing morphisms of affine schemes, it follows that the category of varieties is a full subcategory of the category of schemes over k.

For more details, see .

Examples

Properties

A morphism between varieties is continuous with respect to Zariski topologies on the source and the target.

The image of a morphism of varieties need not be open nor closed (for example, the image of is neither open nor closed). However, one can still say: if f is a morphism between varieties, then the image of f contains an open dense subset of its closure. (cf. constructible set.)

A morphism f:XY of algebraic varieties is said to be dominant if it has dense image. For such an f, if V is a nonempty open affine subset of Y, then there is a nonempty open affine subset U of X such that f(U) ⊂ V and then is injective. Thus, the dominant map f induces an injection on the level of function fields:

where the limit runs over all nonempty open affine subsets of Y. (More abstractly, this is the induced map from the residue field of the generic point of Y to that of X.) Conversely, every inclusion of fields is induced by a dominant rational map from X to Y. [3] Hence, the above construction determines a contravariant-equivalence between the category of algebraic varieties over a field k and dominant rational maps between them and the category of finitely generated field extension of k. [4]

If X is a smooth complete curve (for example, P1) and if f is a rational map from X to a projective space Pm, then f is a regular map XPm. [5] In particular, when X is a smooth complete curve, any rational function on X may be viewed as a morphism XP1 and, conversely, such a morphism as a rational function on X.

On a normal variety (in particular, a smooth variety), a rational function is regular if and only if it has no poles of codimension one. [lower-alpha 6] This is an algebraic analog of Hartogs' extension theorem. There is also a relative version of this fact; see .

A morphism between algebraic varieties that is a homeomorphism between the underlying topological spaces need not be an isomorphism (a counterexample is given by a Frobenius morphism .) On the other hand, if f is bijective birational and the target space of f is a normal variety, then f is biregular. (cf. Zariski's main theorem.)

A regular map between complex algebraic varieties is a holomorphic map. (There is actually a slight technical difference: a regular map is a meromorphic map whose singular points are removable, but the distinction is usually ignored in practice.) In particular, a regular map into the complex numbers is just a usual holomorphic function (complex-analytic function).

Morphisms to a projective space

Let

be a morphism from a projective variety to a projective space. Let x be a point of X. Then some i-th homogeneous coordinate of f(x) is nonzero; say, i = 0 for simplicity. Then, by continuity, there is an open affine neighborhood U of x such that

is a morphism, where yi are the homogeneous coordinates. Note the target space is the affine space Am through the identification . Thus, by definition, the restriction f |U is given by

where gi's are regular functions on U. Since X is projective, each gi is a fraction of homogeneous elements of the same degree in the homogeneous coordinate ring k[X] of X. We can arrange the fractions so that they all have the same homogeneous denominator say f0. Then we can write gi = fi/f0 for some homogeneous elements fi's in k[X]. Hence, going back to the homogeneous coordinates,

for all x in U and by continuity for all x in X as long as the fi's do not vanish at x simultaneously. If they vanish simultaneously at a point x of X, then, by the above procedure, one can pick a different set of fi's that do not vanish at x simultaneously (see Note at the end of the section.)

In fact, the above description is valid for any quasi-projective variety X, an open subvariety of a projective variety ; the difference being that fi's are in the homogeneous coordinate ring of .

Note: The above does not say a morphism from a projective variety to a projective space is given by a single set of polynomials (unlike the affine case). For example, let X be the conic in P2. Then two maps and agree on the open subset of X (since ) and so defines a morphism .

Fibers of a morphism

The important fact is: [6]

Theorem  Let f: XY be a dominating (i.e., having dense image) morphism of algebraic varieties, and let r = dim X  dim Y. Then

  1. For every irreducible closed subset W of Y and every irreducible component Z of dominating W,
  2. There exists a nonempty open subset U in Y such that (a) and (b) for every irreducible closed subset W of Y intersecting U and every irreducible component Z of intersecting ,

Corollary  Let f: XY be a morphism of algebraic varieties. For each x in X, define

Then e is upper-semicontinuous; i.e., for each integer n, the set

is closed.

In Mumford's red book, the theorem is proved by means of Noether's normalization lemma. For an algebraic approach where the generic freeness plays a main role and the notion of "universally catenary ring" is a key in the proof, see Eisenbud, Ch. 14 of "Commutative algebra with a view toward algebraic geometry." In fact, the proof there shows that if f is flat, then the dimension equality in 2. of the theorem holds in general (not just generically).

Degree of a finite morphism

Let f: XY be a finite surjective morphism between algebraic varieties over a field k. Then, by definition, the degree of f is the degree of the finite field extension of the function field k(X) over f*k(Y). By generic freeness, there is some nonempty open subset U in Y such that the restriction of the structure sheaf OX to f−1(U) is free as OY|U-module. The degree of f is then also the rank of this free module.

If f is étale and if X, Y are complete, then for any coherent sheaf F on Y, writing χ for the Euler characteristic,

[7]

(The Riemann–Hurwitz formula for a ramified covering shows the "étale" here cannot be omitted.)

In general, if f is a finite surjective morphism, if X, Y are complete and F a coherent sheaf on Y, then from the Leray spectral sequence , one gets:

In particular, if F is a tensor power of a line bundle, then and since the support of has positive codimension if q is positive, comparing the leading terms, one has:

(since the generic rank of is the degree of f.)

If f is étale and k is algebraically closed, then each geometric fiber f−1(y) consists exactly of deg(f) points.

See also

Notes

  1. Here is the argument showing the definitions coincide. Clearly, we can assume Y = A1. Then the issue here is whether the "regular-ness" can be patched together; this answer is yes and that can be seen from the construction of the structure sheaf of an affine variety as described at affine variety#Structure sheaf.
  2. It is not clear how to prove this, though. If X, Y are quasi-projective, then the proof can be given. The non-quasi-projective case strongly depends on one's definition of an abstract variety
  3. The image of lies in Y since if g is a polynomial in J, then, a priori thinking is a map to the affine space, since g is in J.
  4. Proof: since φ is an algebra homomorphism. Also,
  5. Proof: Let A be the coordinate ring of such an affine neighborhood of x. If f = g/h with some g in A and some nonzero h in A, then f is in A[h−1] = k[D(h)]; that is, f is a regular function on D(h).
  6. Proof: it's enough to consider the case when the variety is affine and then use the fact that a Noetherian integrally closed domain is the intersection of all the localizations at height-one prime ideals.

Citations

  1. Shafarevich 2013, p. 25, Def..
  2. Hartshorne 1997, Ch. I, § 3..
  3. Vakil, Foundations of algebraic geometry, Proposition 6.5.7.
  4. Hartshorne 1997, Ch. I,Theorem 4.4..
  5. Hartshorne 1997, Ch. I, Proposition 6.8..
  6. Mumford 1999 , Ch. I, § 8. Theorems 2, 3.
  7. Fulton 1998, Example 18.3.9..

Related Research Articles

In mathematics, an associative algebraA over a commutative ring K is a ring A together with a ring homomorphism from K into the center of A. This is thus an algebraic structure with an addition, a multiplication, and a scalar multiplication. The addition and multiplication operations together give A the structure of a ring; the addition and scalar multiplication operations together give A the structure of a module or vector space over K. In this article we will also use the term K-algebra to mean an associative algebra over K. A standard first example of a K-algebra is a ring of square matrices over a commutative ring K, with the usual matrix multiplication.

<span class="mw-page-title-main">Isomorphism</span> In mathematics, invertible homomorphism

In mathematics, an isomorphism is a structure-preserving mapping between two structures of the same type that can be reversed by an inverse mapping. Two mathematical structures are isomorphic if an isomorphism exists between them. The word isomorphism is derived from the Ancient Greek: ἴσοςisos "equal", and μορφήmorphe "form" or "shape".

The Riesz representation theorem, sometimes called the Riesz–Fréchet representation theorem after Frigyes Riesz and Maurice René Fréchet, establishes an important connection between a Hilbert space and its continuous dual space. If the underlying field is the real numbers, the two are isometrically isomorphic; if the underlying field is the complex numbers, the two are isometrically anti-isomorphic. The (anti-) isomorphism is a particular natural isomorphism.

In mathematics, rings are algebraic structures that generalize fields: multiplication need not be commutative and multiplicative inverses need not exist. Informally, a ring is a set equipped with two binary operations satisfying properties analogous to those of addition and multiplication of integers. Ring elements may be numbers such as integers or complex numbers, but they may also be non-numerical objects such as polynomials, square matrices, functions, and power series.

<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">Affine variety</span> Algebraic variety defined within an affine space

In algebraic geometry, an affine algebraic set is the set of the common zeros over an algebraically closed field k of some family of polynomials in the polynomial ring An affine variety or affine algebraic variety, is an affine algebraic set such that the ideal generated by the defining polynomials is prime.

<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, a scheme is a mathematical structure that enlarges the notion of algebraic variety in several ways, such as taking account of multiplicities and allowing "varieties" defined over any commutative ring.

In mathematics, an algebraic torus, where a one dimensional torus is typically denoted by , , or , is a type of commutative affine algebraic group commonly found in projective algebraic geometry and toric geometry. Higher dimensional algebraic tori can be modelled as a product of algebraic groups . These groups were named by analogy with the theory of tori in Lie group theory. For example, over the complex numbers the algebraic torus is isomorphic to the group scheme , which is the scheme theoretic analogue of the Lie group . In fact, any -action on a complex vector space can be pulled back to a -action from the inclusion as real manifolds.

In commutative algebra and field theory, the Frobenius endomorphism is a special endomorphism of commutative rings with prime characteristic p, an important class that includes finite fields. The endomorphism maps every element to its p-th power. In certain contexts it is an automorphism, but this is not true in general.

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.

In algebraic geometry, an algebraic variety or scheme X is normal if it is normal at every point, meaning that the local ring at the point is an integrally closed domain. An affine variety X (understood to be irreducible) is normal if and only if the ring O(X) of regular functions on X is an integrally closed domain. A variety X over a field is normal if and only if every finite birational morphism from any variety Y to X is an isomorphism.

In mathematics, in particular the subfield of algebraic geometry, a rational map or rational mapping is a kind of partial function between algebraic varieties. This article uses the convention that varieties are irreducible.

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 commutative algebra, an element b of a commutative ring B is said to be integral over a subring A of B if b is a root of some monic polynomial over A.

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 homotopy-theoretic generalization of a scheme in which classical commutative rings are replaced with derived versions such as cdgas, commutative simplicial rings, or commutative ring spectra.

In algebraic geometry, Lang's theorem, introduced by Serge Lang, states: if G is a connected smooth algebraic group over a finite field , then, writing for the Frobenius, the morphism of varieties

In algebraic geometry, the main theorem of elimination theory states that every projective scheme is proper. A version of this theorem predates the existence of scheme theory. It can be stated, proved, and applied in the following more classical setting. Let k be a field, denote by the n-dimensional projective space over k. The main theorem of elimination theory is the statement that for any n and any algebraic variety V defined over k, the projection map sends Zariski-closed subsets to Zariski-closed subsets.

References