Length of a module

Last updated

In algebra, the length of a module over a ring is a generalization of the dimension of a vector space which measures its size. [1] page 153 It is defined to be the length of the longest chain of submodules. For vector spaces (modules over a field), the length equals the dimension. If is an algebra over a field , the length of a module is at most its dimension as a -vector space.

Contents

In commutative algebra and algebraic geometry, a module over a Noetherian commutative ring can have finite length only when the module has Krull dimension zero. Modules of finite length are finitely generated modules, but most finitely generated modules have infinite length. Modules of finite length are called Artinian modules and are fundamental to the theory of Artinian rings.

The degree of an algebraic variety inside an affine or projective space is the length of the coordinate ring of the zero-dimensional intersection of the variety with a generic linear subspace of complementary dimension. More generally, the intersection multiplicity of several varieties is defined as the length of the coordinate ring of the zero-dimensional intersection.

Definition

Length of a module

Let be a (left or right) module over some ring . Given a chain of submodules of of the form

one says that is the length of the chain. [1] The length of is the largest length of any of its chains. If no such largest length exists, we say that has infinite length. Clearly, if the length of a chain equals the length of the module, one has and

Length of a ring

The length of a ring is the length of the longest chain of ideals; that is, the length of considered as a module over itself by left multiplication. By contrast, the Krull dimension of is the length of the longest chain of prime ideals.

Properties

Finite length and finite modules

If an -module has finite length, then it is finitely generated. [2] If R is a field, then the converse is also true.

Relation to Artinian and Noetherian modules

An -module has finite length if and only if it is both a Noetherian module and an Artinian module [1] (cf. Hopkins' theorem). Since all Artinian rings are Noetherian, this implies that a ring has finite length if and only if it is Artinian.

Behavior with respect to short exact sequences

Suppose

is a short exact sequence of -modules. Then M has finite length if and only if L and N have finite length, and we have

In particular, it implies the following two properties

Jordan–Hölder theorem

A composition series of the module M is a chain of the form

such that

A module M has finite length if and only if it has a (finite) composition series, and the length of every such composition series is equal to the length of M.

Examples

Finite dimensional vector spaces

Any finite dimensional vector space over a field has a finite length. Given a basis there is the chain

which is of length . It is maximal because given any chain,

the dimension of each inclusion will increase by at least . Therefore, its length and dimension coincide.

Artinian modules

Over a base ring , Artinian modules form a class of examples of finite modules. In fact, these examples serve as the basic tools for defining the order of vanishing in intersection theory. [3]

Zero module

The zero module is the only one with length 0.

Simple modules

Modules with length 1 are precisely the simple modules.

Artinian modules over Z

The length of the cyclic group (viewed as a module over the integers Z) is equal to the number of prime factors of , with multiple prime factors counted multiple times. This follows from the fact that the submodules of are in one to one correspondence with the positive divisors of , this correspondence resulting itself from the fact that is a principal ideal ring.

Use in multiplicity theory

For the needs of intersection theory, Jean-Pierre Serre introduced a general notion of the multiplicity of a point, as the length of an Artinian local ring related to this point.

The first application was a complete definition of the intersection multiplicity, and, in particular, a statement of Bézout's theorem that asserts that the sum of the multiplicities of the intersection points of n algebraic hypersurfaces in a n-dimensional projective space is either infinite or is exactly the product of the degrees of the hypersurfaces.

This definition of multiplicity is quite general, and contains as special cases most of previous notions of algebraic multiplicity.

Order of vanishing of zeros and poles

A special case of this general definition of a multiplicity is the order of vanishing of a non-zero algebraic function on an algebraic variety. Given an algebraic variety and a subvariety of codimension 1 [3] the order of vanishing for a polynomial is defined as [4]

where is the local ring defined by the stalk of along the subvariety [3] pages 426-227, or, equivalently, the stalk of at the generic point of [5] page 22. If is an affine variety, and is defined the by vanishing locus , then there is the isomorphism

This idea can then be extended to rational functions on the variety where the order is defined as [3]

which is similar to defining the order of zeros and poles in complex analysis.

Example on a projective variety

For example, consider a projective surface defined by a polynomial , then the order of vanishing of a rational function

is given by

where

For example, if and and then

since is a unit in the local ring . In the other case, is a unit, so the quotient module is isomorphic to

so it has length . This can be found using the maximal proper sequence

Zero and poles of an analytic function

The order of vanishing is a generalization of the order of zeros and poles for meromorphic functions in complex analysis. For example, the function

has zeros of order 2 and 1 at and a pole of order at . This kind of information can be encoded using the length of modules. For example, setting and , there is the associated local ring is and the quotient module

Note that is a unit, so this is isomorphic to the quotient module

Its length is since there is the maximal chain

of submodules. [6] More generally, using the Weierstrass factorization theorem a meromorphic function factors as

which is a (possibly infinite) product of linear polynomials in both the numerator and denominator.

See also

Related Research Articles

In commutative algebra, the prime spectrum of a commutative ring R is the set of all prime ideals of R, and is usually denoted by ; in algebraic geometry it is simultaneously a topological space equipped with the sheaf of rings .

In mathematics, the endomorphisms of an abelian group X form a ring. This ring is called the endomorphism ring of X, denoted by End(X); the set of all homomorphisms of X into itself. Addition of endomorphisms arises naturally in a pointwise manner and multiplication via endomorphism composition. Using these operations, the set of endomorphisms of an abelian group forms a (unital) ring, with the zero map as additive identity and the identity map as multiplicative identity.

In mathematics, K-theory is, roughly speaking, the study of a ring generated by vector bundles over a topological space or scheme. In algebraic topology, it is a cohomology theory known as topological K-theory. In algebra and algebraic geometry, it is referred to as algebraic K-theory. It is also a fundamental tool in the field of operator algebras. It can be seen as the study of certain kinds of invariants of large matrices.

<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, the annihilator of a subset S of a module over a ring is the ideal formed by the elements of the ring that give always zero when multiplied by each element of S.

In mathematics, specifically abstract algebra, an Artinian ring is a ring that satisfies the descending chain condition on (one-sided) ideals; that is, there is no infinite descending sequence of ideals. Artinian rings are named after Emil Artin, who first discovered that the descending chain condition for ideals simultaneously generalizes finite rings and rings that are finite-dimensional vector spaces over fields. The definition of Artinian rings may be restated by interchanging the descending chain condition with an equivalent notion: the minimum condition.

In mathematics, a Cohen–Macaulay ring is a commutative ring with some of the algebro-geometric properties of a smooth variety, such as local equidimensionality. Under mild assumptions, a local ring is Cohen–Macaulay exactly when it is a finitely generated free module over a regular local subring. Cohen–Macaulay rings play a central role in commutative algebra: they form a very broad class, and yet they are well understood in many ways.

In mathematics, in particular in the theory of schemes in algebraic geometry, a flat morphismf from a scheme X to a scheme Y is a morphism such that the induced map on every stalk is a flat map of rings, i.e.,

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 algebraic geometry, a closed immersion of schemes is a morphism of schemes that identifies Z as a closed subset of X such that locally, regular functions on Z can be extended to X. The latter condition can be formalized by saying that is surjective.

In mathematics, the Grothendieck group, or group of differences, of a commutative monoid M is a certain abelian group. This abelian group is constructed from M in the most universal way, in the sense that any abelian group containing a homomorphic image of M will also contain a homomorphic image of the Grothendieck group of M. The Grothendieck group construction takes its name from a specific case in category theory, introduced by Alexander Grothendieck in his proof of the Grothendieck–Riemann–Roch theorem, which resulted in the development of K-theory. This specific case is the monoid of isomorphism classes of objects of an abelian category, with the direct sum as its operation.

In algebraic geometry, Proj is a construction analogous to the spectrum-of-a-ring construction of affine schemes, which produces objects with the typical properties of projective spaces and projective varieties. The construction, while not functorial, is a fundamental tool in scheme theory.

In algebraic geometry, an étale morphism is a morphism of schemes that is formally étale and locally of finite presentation. This is an algebraic analogue of the notion of a local isomorphism in the complex analytic topology. They satisfy the hypotheses of the implicit function theorem, but because open sets in the Zariski topology are so large, they are not necessarily local isomorphisms. Despite this, étale maps retain many of the properties of local analytic isomorphisms, and are useful in defining the algebraic fundamental group and the étale topology.

In mathematics, a π-system on a set is a collection of certain subsets of such that

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.

In mathematics, Hochschild homology (and cohomology) is a homology theory for associative algebras over rings. There is also a theory for Hochschild homology of certain functors. Hochschild cohomology was introduced by Gerhard Hochschild (1945) for algebras over a field, and extended to algebras over more general rings by Henri Cartan and Samuel Eilenberg (1956).

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 mathematics, a sheaf of O-modules or simply an O-module over a ringed space (X, O) is a sheaf F such that, for any open subset U of X, F(U) is an O(U)-module and the restriction maps F(U) → F(V) are compatible with the restriction maps O(U) → O(V): the restriction of fs is the restriction of f times the restriction of s for any f in O(U) and s in F(U).

References

  1. 1 2 3 "A Term of Commutative Algebra". www.centerofmathematics.com. pp. 153–158. Archived from the original on 2013-03-02. Retrieved 2020-05-22. Alt URL
  2. "Lemma 10.51.2 (02LZ)—The Stacks project". stacks.math.columbia.edu. Retrieved 2020-05-22.
  3. 1 2 3 4 Fulton, William, 1939- (1998). Intersection theory (2nd ed.). Berlin: Springer. pp. 8–10. ISBN   3-540-62046-X. OCLC   38048404.{{cite book}}: CS1 maint: multiple names: authors list (link) CS1 maint: numeric names: authors list (link)
  4. "Section 31.26 (0BE0): Weil divisors—The Stacks project". stacks.math.columbia.edu. Retrieved 2020-05-22.
  5. Hartshorne, Robin (1977). Algebraic Geometry. Graduate Texts in Mathematics. Vol. 52. New York, NY: Springer New York. doi:10.1007/978-1-4757-3849-0. ISBN   978-1-4419-2807-8. S2CID   197660097.
  6. "Section 10.120 (02MB): Orders of vanishing—The Stacks project". stacks.math.columbia.edu. Retrieved 2020-05-22.