Decomposition of a module

Last updated

In abstract algebra, a decomposition of a module is a way to write a module as a direct sum of modules. A type of a decomposition is often used to define or characterize modules: for example, a semisimple module is a module that has a decomposition into simple modules. Given a ring, the types of decomposition of modules over the ring can also be used to define or characterize the ring: a ring is semisimple if and only if every module over it is a semisimple module.

Contents

An indecomposable module is a module that is not a direct sum of two nonzero submodules. Azumaya's theorem states that if a module has an decomposition into modules with local endomorphism rings, then all decompositions into indecomposable modules are equivalent to each other; a special case of this, especially in group theory, is known as the Krull–Schmidt theorem.

A special case of a decomposition of a module is a decomposition of a ring: for example, a ring is semisimple if and only if it is a direct sum (in fact a product) of matrix rings over division rings (this observation is known as the Artin–Wedderburn theorem).

Idempotents and decompositions

To give a direct sum decomposition of a module into submodules is the same as to give orthogonal idempotents in the endomorphism ring of the module that sum up to the identity map. [1] Indeed, if , then, for each , the linear endomorphism given by the natural projection followed by the natural inclusion is an idempotent. They are clearly orthogonal to each other ( for ) and they sum up to the identity map:

as endomorphisms (here the summation is well-defined since it is a finite sum at each element of the module). Conversely, each set of orthogonal idempotents such that only finitely many are nonzero for each and determine a direct sum decomposition by taking to be the images of .

This fact already puts some constraints on a possible decomposition of a ring: given a ring , suppose there is a decomposition

of as a left module over itself, where are left submodules; i.e., left ideals. Each endomorphism can be identified with a right multiplication by an element of R; thus, where are idempotents of . [2] The summation of idempotent endomorphisms corresponds to the decomposition of the unity of R: , which is necessarily a finite sum; in particular, must be a finite set.

For example, take , the ring of n-by-n matrices over a division ring D. Then is the direct sum of n copies of , the columns; each column is a simple left R-submodule or, in other words, a minimal left ideal. [3]

Let R be a ring. Suppose there is a (necessarily finite) decomposition of it as a left module over itself

into two-sided ideals of R. As above, for some orthogonal idempotents such that . Since is an ideal, and so for . Then, for each i,

That is, the are in the center; i.e., they are central idempotents. [4] Clearly, the argument can be reversed and so there is a one-to-one correspondence between the direct sum decomposition into ideals and the orthogonal central idempotents summing up to the unity 1. Also, each itself is a ring on its own right, the unity given by , and, as a ring, R is the product ring

For example, again take . This ring is a simple ring; in particular, it has no nontrivial decomposition into two-sided ideals.

Types of decomposition

There are several types of direct sum decompositions that have been studied:

Since a simple module is indecomposable, a semisimple decomposition is an indecomposable decomposition (but not conversely). If the endomorphism ring of a module is local, then, in particular, it cannot have a nontrivial idempotent: the module is indecomposable. Thus, a decomposition with local endomorphism rings is an indecomposable decomposition.

A direct summand is said to be maximal if it admits an indecomposable complement. A decomposition is said to complement maximal direct summands if for each maximal direct summand L of M, there exists a subset such that

[7]

Two decompositions are said to be equivalent if there is a bijection such that for each , . [7] If a module admits an indecomposable decomposition complementing maximal direct summands, then any two indecomposable decompositions of the module are equivalent. [8]

Azumaya's theorem

In the simplest form, Azumaya's theorem states: [9] given a decomposition such that the endomorphism ring of each is local (so the decomposition is indecomposable), each indecomposable decomposition of M is equivalent to this given decomposition. The more precise version of the theorem states: [10] still given such a decomposition, if , then

  1. if nonzero, N contains an indecomposable direct summand,
  2. if is indecomposable, the endomorphism ring of it is local [11] and is complemented by the given decomposition:
    and so for some ,
  3. for each , there exist direct summands of and of such that .

The endomorphism ring of an indecomposable module of finite length is local (e.g., by Fitting's lemma) and thus Azumaya's theorem applies to the setup of the Krull–Schmidt theorem. Indeed, if M is a module of finite length, then, by induction on length, it has a finite indecomposable decomposition , which is a decomposition with local endomorphism rings. Now, suppose we are given an indecomposable decomposition . Then it must be equivalent to the first one: so and for some permutation of . More precisely, since is indecomposable, for some . Then, since is indecomposable, and so on; i.e., complements to each sum can be taken to be direct sums of some 's.

Another application is the following statement (which is a key step in the proof of Kaplansky's theorem on projective modules):

To see this, choose a finite set such that . Then, writing , by Azumaya's theorem, with some direct summands of and then, by modular law, with . Then, since is a direct summand of , we can write and then , which implies, since F is finite, that for some J by a repeated application of Azumaya's theorem.

In the setup of Azumaya's theorem, if, in addition, each is countably generated, then there is the following refinement (due originally to Crawley–Jónsson and later to Warfield): is isomorphic to for some subset . [12] (In a sense, this is an extension of Kaplansky's theorem and is proved by the two lemmas used in the proof of the theorem.) According to ( Facchini 1998 ), it is not known whether the assumption " countably generated" can be dropped; i.e., this refined version is true in general.

Decomposition of a ring

On the decomposition of a ring, the most basic but still important observation, known as the Wedderburn-Artin theorem is this: given a ring R, the following are equivalent:

  1. R is a semisimple ring; i.e., is a semisimple left module.
  2. for division rings , where denotes the ring of n-by-n matrices with entries in , and the positive integers , the division rings , and the positive integers are determined (the latter two up to permutation) by R
  3. Every left module over R is semisimple.

To show 1. 2., first note that if is semisimple then we have an isomorphism of left -modules where are mutually non-isomorphic minimal left ideals. Then, with the view that endomorphisms act from the right,

where each can be viewed as the matrix ring over , which is a division ring by Schur's Lemma. The converse holds because the decomposition of 2. is equivalent to a decomposition into minimal left ideals = simple left submodules. The equivalence 1. 3. holds because every module is a quotient of a free module, and a quotient of a semisimple module is semisimple.

See also

Notes

  1. Anderson & Fuller 1992 , Corollary 6.19. and Corollary 6.20.
  2. Here, the endomorphism ring is thought of as acting from the right; if it acts from the left, this identification is for the opposite ring of R.
  3. Procesi 2007 , Ch.6., § 1.3.
  4. Anderson & Fuller 1992 , Proposion 7.6.
  5. ( Jacobson 2009 , A paragraph before Theorem 3.6.) calls a module strongly indecomposable if nonzero and has local endomorphism ring.
  6. Anderson & Fuller 1992 , § 32.
  7. 1 2 Anderson & Fuller 1992 , § 12.
  8. Anderson & Fuller 1992 , Theorrm 12.4.
  9. Facchini 1998 , Theorem 2.12.
  10. Anderson & Fuller 1992 , Theorem 12.6. and Lemma 26.4.
  11. Facchini 1998 , Lemma 2.11.
  12. Facchini 1998 , Corollary 2.55.

Related Research Articles

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.

In abstract algebra, the direct sum is a construction which combines several modules into a new, larger module. The direct sum of modules is the smallest module which contains the given modules as submodules with no "unnecessary" constraints, making it an example of a coproduct. Contrast with the direct product, which is the dual notion.

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 ring theory, a branch of mathematics, an idempotent element or simply idempotent of a ring is an element a such that a2 = a. That is, the element is idempotent under the ring's multiplication. Inductively then, one can also conclude that a = a2 = a3 = a4 = ... = an for any positive integer n. For example, an idempotent element of a matrix ring is precisely an idempotent matrix.

In algebra, the Wedderburn–Artin theorem is a classification theorem for semisimple rings and semisimple algebras. The theorem states that an (Artinian) semisimple ring R is isomorphic to a product of finitely many ni-by-ni matrix rings over division rings Di, for some integers ni, both of which are uniquely determined up to permutation of the index i. In particular, any simple left or right Artinian ring is isomorphic to an n-by-n matrix ring over a division ring D, where both n and D are uniquely determined.

In abstract algebra, a module is indecomposable if it is non-zero and cannot be written as a direct sum of two non-zero submodules.

In mathematics, especially in the area of abstract algebra known as module theory, a semisimple module or completely reducible module is a type of module that can be understood easily from its parts. A ring that is a semisimple module over itself is known as an Artinian semisimple ring. Some important rings, such as group rings of finite groups over fields of characteristic zero, are semisimple rings. An Artinian ring is initially understood via its largest semisimple quotient. The structure of Artinian semisimple rings is well understood by the Artin–Wedderburn theorem, which exhibits these rings as finite direct products of matrix rings.

<span class="mw-page-title-main">Semisimple Lie algebra</span> Direct sum of simple Lie algebras

In mathematics, a Lie algebra is semisimple if it is a direct sum of simple Lie algebras..

The direct sum is an operation between structures in abstract algebra, a branch of mathematics. It is defined differently, but analogously, for different kinds of structures. To see how the direct sum is used in abstract algebra, consider a more elementary kind of structure, the abelian group. The direct sum of two abelian groups and is another abelian group consisting of the ordered pairs where and . To add ordered pairs, we define the sum to be ; in other words addition is defined coordinate-wise. For example, the direct sum , where is real coordinate space, is the Cartesian plane, . A similar process can be used to form the direct sum of two vector spaces or two modules.

In mathematics, in the field of abstract algebra, the structure theorem for finitely generated modules over a principal ideal domain is a generalization of the fundamental theorem of finitely generated abelian groups and roughly states that finitely generated modules over a principal ideal domain (PID) can be uniquely decomposed in much the same way that integers have a prime factorization. The result provides a simple framework to understand various canonical form results for square matrices over fields.

In mathematics, a separable algebra is a kind of semisimple algebra. It is a generalization to associative algebras of the notion of a separable field extension.

Schur–Weyl duality is a mathematical theorem in representation theory that relates irreducible finite-dimensional representations of the general linear and symmetric groups. It is named after two pioneers of representation theory of Lie groups, Issai Schur, who discovered the phenomenon, and Hermann Weyl, who popularized it in his books on quantum mechanics and classical groups as a way of classifying representations of unitary and general linear groups.

In category theory, a branch of mathematics, a Krull–Schmidt category is a generalization of categories in which the Krull–Schmidt theorem holds. They arise, for example, in the study of finite-dimensional modules over an algebra.

In mathematics, the Krull–Schmidt theorem states that a group subjected to certain finiteness conditions on chains of subgroups, can be uniquely written as a finite direct product of indecomposable subgroups.

Module theory is the branch of mathematics in which modules are studied. This is a glossary of some terms of the subject.

In abstract algebra, a uniserial moduleM is a module over a ring R, whose submodules are totally ordered by inclusion. This means simply that for any two submodules N1 and N2 of M, either or . A module is called a serial module if it is a direct sum of uniserial modules. A ring R is called a right uniserial ring if it is uniserial as a right module over itself, and likewise called a right serial ring if it is a right serial module over itself. Left uniserial and left serial rings are defined in an analogous way, and are in general distinct from their right counterparts.

In algebra, Weyl's theorem on complete reducibility is a fundamental result in the theory of Lie algebra representations. Let be a semisimple Lie algebra over a field of characteristic zero. The theorem states that every finite-dimensional module over is semisimple as a module

In mathematics, semi-simplicity is a widespread concept in disciplines such as linear algebra, abstract algebra, representation theory, category theory, and algebraic geometry. A semi-simple object is one that can be decomposed into a sum of simple objects, and simple objects are those that do not contain non-trivial proper sub-objects. The precise definitions of these words depends on the context.

In mathematics, specifically in representation theory, a semisimple representation is a linear representation of a group or an algebra that is a direct sum of simple representations. It is an example of the general mathematical notion of semisimplicity.

In abstract algebra, Kaplansky's theorem on projective modules, first proven by Irving Kaplansky, states that a projective module over a local ring is free; where a not-necessarily-commutative ring is called local if for each element x, either x or 1 − x is a unit element. The theorem can also be formulated so to characterize a local ring.

References