Group ring

Last updated

In algebra, a group ring is a free module and at the same time a ring, constructed in a natural way from any given ring and any given group. As a free module, its ring of scalars is the given ring, and its basis is the set of elements of the given group. As a ring, its addition law is that of the free module and its multiplication extends "by linearity" the given group law on the basis. Less formally, a group ring is a generalization of a given group, by attaching to each element of the group a "weighting factor" from a given ring.

Contents

If the ring is commutative then the group ring is also referred to as a group algebra, for it is indeed an algebra over the given ring. A group algebra over a field has a further structure of a Hopf algebra; in this case, it is thus called a group Hopf algebra.

The apparatus of group rings is especially useful in the theory of group representations.

Definition

Let be a group, written multiplicatively, and let be a ring. The group ring of over , which we will denote by , or simply , is the set of mappings of finite support ( is nonzero for only finitely many elements ), where the module scalar product of a scalar in and a mapping is defined as the mapping , and the module group sum of two mappings and is defined as the mapping . To turn the additive group into a ring, we define the product of and to be the mapping

The summation is legitimate because and are of finite support, and the ring axioms are readily verified.

Some variations in the notation and terminology are in use. In particular, the mappings such as are sometimes [1] written as what are called "formal linear combinations of elements of with coefficients in ":

or simply

[2]

Note that if the ring is in fact a field , then the module structure of the group ring is in fact a vector space over .

Examples

1. Let G = C3, the cyclic group of order 3, with generator and identity element 1G. An element r of C[G] can be written as

where z0, z1 and z2 are in C, the complex numbers. This is the same thing as a polynomial ring in variable such that i.e. C[G] is isomorphic to the ring C[]/.

Writing a different element s as , their sum is

and their product is

Notice that the identity element 1G of G induces a canonical embedding of the coefficient ring (in this case C) into C[G]; however strictly speaking the multiplicative identity element of C[G] is 1⋅1G where the first 1 comes from C and the second from G. The additive identity element is zero.

When G is a non-commutative group, one must be careful to preserve the order of the group elements (and not accidentally commute them) when multiplying the terms.

2. The ring of Laurent polynomials over a ring R is the group ring of the infinite cyclic group Z over R.

3. Let Q be the quaternion group with elements . Consider the group ring RQ, where R is the set of real numbers. An arbitrary element of this group ring is of the form

where is a real number.

Multiplication, as in any other group ring, is defined based on the group operation. For example,

Note that RQ is not the same as the skew field of quaternions over R. This is because the skew field of quaternions satisfies additional relations in the ring, such as , whereas in the group ring RQ, is not equal to . To be more specific, the group ring RQ has dimension 8 as a real vector space, while the skew field of quaternions has dimension 4 as a real vector space.

4. Another example of a non-abelian group ring is where is the symmetric group on 3 letters. This is not an integral domain since we have where the element is the transposition that swaps 1 and 2. Therefore the group ring need not be an integral domain even when the underlying ring is an integral domain.

Some basic properties

Using 1 to denote the multiplicative identity of the ring R, and denoting the group unit by 1G, the ring R[G] contains a subring isomorphic to R, and its group of invertible elements contains a subgroup isomorphic to G. For considering the indicator function of {1G}, which is the vector f defined by

the set of all scalar multiples of f is a subring of R[G] isomorphic to R. And if we map each element s of G to the indicator function of {s}, which is the vector f defined by

the resulting mapping is an injective group homomorphism (with respect to multiplication, not addition, in R[G]).

If R and G are both commutative (i.e., R is commutative and G is an abelian group), R[G] is commutative.

If H is a subgroup of G, then R[H] is a subring of R[G]. Similarly, if S is a subring of R, S[G] is a subring of R[G].

If G is a finite group of order greater than 1, then R[G] always has zero divisors. For example, consider an element g of G of order |g| = m > 1. Then 1 - g is a zero divisor:

For example, consider the group ring Z[S3] and the element of order 3 g=(123). In this case,

A related result: If the group ring is prime, then G has no nonidentity finite normal subgroup (in particular, G must be infinite).

Proof: Considering the contrapositive, suppose is a nonidentity finite normal subgroup of . Take . Since for any , we know , therefore . Taking , we have . By normality of , commutes with a basis of , and therefore

.

And we see that are not zero, which shows is not prime. This shows the original statement.

Group algebra over a finite group

Group algebras occur naturally in the theory of group representations of finite groups. The group algebra K[G] over a field K is essentially the group ring, with the field K taking the place of the ring. As a set and vector space, it is the free vector space on G over the field K. That is, for x in K[G],

The algebra structure on the vector space is defined using the multiplication in the group:

where on the left, g and h indicate elements of the group algebra, while the multiplication on the right is the group operation (denoted by juxtaposition).

Because the above multiplication can be confusing, one can also write the basis vectors of K[G] as eg (instead of g), in which case the multiplication is written as:

Interpretation as functions

Thinking of the free vector space as K-valued functions on G, the algebra multiplication is convolution of functions.

While the group algebra of a finite group can be identified with the space of functions on the group, for an infinite group these are different. The group algebra, consisting of finite sums, corresponds to functions on the group that vanish for cofinitely many points; topologically (using the discrete topology), these correspond to functions with compact support.

However, the group algebra K[G] and the space of functions KG := Hom(G, K) are dual: given an element of the group algebra

and a function on the group f : GK these pair to give an element of K via

which is a well-defined sum because it is finite.

Representations of a group algebra

Taking K[G] to be an abstract algebra, one may ask for representations of the algebra acting on a K-vector space V of dimension d. Such a representation

is an algebra homomorphism from the group algebra to the algebra of endomorphisms of V, which is isomorphic to the ring of d × d matrices: . Equivalently, this is a left K[G]-module over the abelian group V.

Correspondingly, a group representation

is a group homomorphism from G to the group of linear automorphisms of V, which is isomorphic to the general linear group of invertible matrices: . Any such representation induces an algebra representation

simply by letting and extending linearly. Thus, representations of the group correspond exactly to representations of the algebra, and the two theories are essentially equivalent.

Regular representation

The group algebra is an algebra over itself; under the correspondence of representations over R and R[G] modules, it is the regular representation of the group.

Written as a representation, it is the representation gρg with the action given by , or

Semisimple decomposition

The dimension of the vector space K[G] is just equal to the number of elements in the group. The field K is commonly taken to be the complex numbers C or the reals R, so that one discusses the group algebras C[G] or R[G].

The group algebra C[G] of a finite group over the complex numbers is a semisimple ring. This result, Maschke's theorem, allows us to understand C[G] as a finite product of matrix rings with entries in C. Indeed, if we list the complex irreducible representations of G as Vk for k = 1, . . . , m, these correspond to group homomorphisms and hence to algebra homomorphisms . Assembling these mappings gives an algebra isomorphism

where dk is the dimension of Vk. The subalgebra of C[G] corresponding to End(Vk) is the two-sided ideal generated by the idempotent

where is the character of Vk. These form a complete system of orthogonal idempotents, so that , for j ≠ k, and . The isomorphism is closely related to Fourier transform on finite groups.

For a more general field K, whenever the characteristic of K does not divide the order of the group G, then K[G] is semisimple. When G is a finite abelian group, the group ring K[G] is commutative, and its structure is easy to express in terms of roots of unity.

When K is a field of characteristic p which divides the order of G, the group ring is not semisimple: it has a non-zero Jacobson radical, and this gives the corresponding subject of modular representation theory its own, deeper character.

Center of a group algebra

The center of the group algebra is the set of elements that commute with all elements of the group algebra:

The center is equal to the set of class functions, that is the set of elements that are constant on each conjugacy class

If K = C, the set of irreducible characters of G forms an orthonormal basis of Z(K[G]) with respect to the inner product

Group rings over an infinite group

Much less is known in the case where G is countably infinite, or uncountable, and this is an area of active research. [3] The case where R is the field of complex numbers is probably the one best studied. In this case, Irving Kaplansky proved that if a and b are elements of C[G] with ab = 1, then ba = 1. Whether this is true if R is a field of positive characteristic remains unknown.

A long-standing conjecture of Kaplansky (~1940) says that if G is a torsion-free group, and K is a field, then the group ring K[G] has no non-trivial zero divisors. This conjecture is equivalent to K[G] having no non-trivial nilpotents under the same hypotheses for K and G.

In fact, the condition that K is a field can be relaxed to any ring that can be embedded into an integral domain.

The conjecture remains open in full generality, however some special cases of torsion-free groups have been shown to satisfy the zero divisor conjecture. These include:

The case where G is a topological group is discussed in greater detail in the article Group algebra of a locally compact group.

Category theory

Adjoint

Categorically, the group ring construction is left adjoint to "group of units"; the following functors are an adjoint pair:

where takes a group to its group ring over R, and takes an R-algebra to its group of units.

When R = Z, this gives an adjunction between the category of groups and the category of rings, and the unit of the adjunction takes a group G to a group that contains trivial units: G × {±1} = {±g}. In general, group rings contain nontrivial units. If G contains elements a and b such that and b does not normalize then the square of

is zero, hence . The element 1 + x is a unit of infinite order.

Universal property

The above adjunction expresses a universal property of group rings. [2] [4] Let R be a (commutative) ring, let G be a group, and let S be an R-algebra. For any group homomorphism , there exists a unique R-algebra homomorphism such that where i is the inclusion

In other words, is the unique homomorphism making the following diagram commute:

Group ring UMP.svg

Any other ring satisfying this property is canonically isomorphic to the group ring.

Hopf algebra

The group algebra K[G] has a natural structure of a Hopf algebra. The comultiplication is defined by , extended linearly, and the antipode is , again extended linearly.

Generalizations

The group algebra generalizes to the monoid ring and thence to the category algebra, of which another example is the incidence algebra.

Filtration

If a group has a length function – for example, if there is a choice of generators and one takes the word metric, as in Coxeter groups – then the group ring becomes a filtered algebra.

See also

Representation theory

Category theory

Notes

  1. Milies & Sehgal (2002), pp. 129 and 131.
  2. 1 2 Milies & Sehgal (2002), p. 131.
  3. Passman, Donald S. (1976). "What is a group ring?". Amer. Math. Monthly. 83 (3): 173–185. doi:10.2307/2977018. JSTOR   2977018.
  4. "group algebra in nLab". ncatlab.org. Retrieved 2017-11-01.

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">Group representation</span> Group homomorphism into the general linear group over a vector space

In the mathematical field of representation theory, group representations describe abstract groups in terms of bijective linear transformations of a vector space to itself ; in particular, they can be used to represent group elements as invertible matrices so that the group operation can be represented by matrix multiplication.

In algebra, a homomorphism is a structure-preserving map between two algebraic structures of the same type. The word homomorphism comes from the Ancient Greek language: ὁμός meaning "same" and μορφή meaning "form" or "shape". However, the word was apparently introduced to mathematics due to a (mis)translation of German ähnlich meaning "similar" to ὁμός meaning "same". The term "homomorphism" appeared as early as 1892, when it was attributed to the German mathematician Felix Klein (1849–1925).

In mathematics, a product is the result of multiplication, or an expression that identifies objects to be multiplied, called factors. For example, 21 is the product of 3 and 7, and is the product of and . When one factor is an integer, the product is called a multiple.

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 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, in particular abstract algebra, a graded ring is a ring such that the underlying additive group is a direct sum of abelian groups such that . The index set is usually the set of nonnegative integers or the set of integers, but can be any monoid. The direct sum decomposition is usually referred to as gradation or grading.

<span class="mw-page-title-main">Quaternion group</span> Non-abelian group of order eight

In group theory, the quaternion group Q8 (sometimes just denoted by Q) is a non-abelian group of order eight, isomorphic to the eight-element subset of the quaternions under multiplication. It is given by the group presentation

In mathematics, a module is a generalization of the notion of vector space in which the field of scalars is replaced by a ring. The concept of module also generalizes the notion of abelian group, since the abelian groups are exactly the modules over the ring of integers.

<span class="mw-page-title-main">Lie algebra representation</span>

In the mathematical field of representation theory, a Lie algebra representation or representation of a Lie algebra is a way of writing a Lie algebra as a set of matrices in such a way that the Lie bracket is given by the commutator. In the language of physics, one looks for a vector space together with a collection of operators on satisfying some fixed set of commutation relations, such as the relations satisfied by the angular momentum operators.

In mathematics, a Hopf algebra, named after Heinz Hopf, is a structure that is simultaneously an algebra and a coalgebra, with these structures' compatibility making it a bialgebra, and that moreover is equipped with an antihomomorphism satisfying a certain property. The representation theory of a Hopf algebra is particularly nice, since the existence of compatible comultiplication, counit, and antipode allows for the construction of tensor products of representations, trivial representations, and dual representations.

<span class="mw-page-title-main">Irreducible representation</span> Type of group and algebra representation

In mathematics, specifically in the representation theory of groups and algebras, an irreducible representation or irrep of an algebraic structure is a nonzero representation that has no proper nontrivial subrepresentation , with closed under the action of .

In mathematics, Schur's lemma is an elementary but extremely useful statement in representation theory of groups and algebras. In the group case it says that if M and N are two finite-dimensional irreducible representations of a group G and φ is a linear map from M to N that commutes with the action of the group, then either φ is invertible, or φ = 0. An important special case occurs when M = N, i.e. φ is a self-map; in particular, any element of the center of a group must act as a scalar operator on M. The lemma is named after Issai Schur who used it to prove the Schur orthogonality relations and develop the basics of the representation theory of finite groups. Schur's lemma admits generalisations to Lie groups and Lie algebras, the most common of which are due to Jacques Dixmier and Daniel Quillen.

The representation theory of groups is a part of mathematics which examines how groups act on given structures.

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.

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. As an example, 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 differential geometry, a Lie-algebra-valued form is a differential form with values in a Lie algebra. Such forms have important applications in the theory of connections on a principal bundle as well as in the theory of Cartan connections.

Algebraic signal processing (ASP) is an emerging area of theoretical signal processing (SP). In the algebraic theory of signal processing, a set of filters is treated as an (abstract) algebra, a set of signals is treated as a module or vector space, and convolution is treated as an algebra representation. The advantage of algebraic signal processing is its generality and portability.

In mathematics, Maschke's theorem, named after Heinrich Maschke, is a theorem in group representation theory that concerns the decomposition of representations of a finite group into irreducible pieces. Maschke's theorem allows one to make general conclusions about representations of a finite group G without actually computing them. It reduces the task of classifying all representations to a more manageable task of classifying irreducible representations, since when the theorem applies, any representation is a direct sum of irreducible pieces (constituents). Moreover, it follows from the Jordan–Hölder theorem that, while the decomposition into a direct sum of irreducible subrepresentations may not be unique, the irreducible pieces have well-defined multiplicities. In particular, a representation of a finite group over a field of characteristic zero is determined up to isomorphism by its character.

In mathematics, Lie group–Lie algebra correspondence allows one to correspond a Lie group to a Lie algebra or vice versa, and study the conditions for such a relationship. Lie groups that are isomorphic to each other have Lie algebras that are isomorphic to each other, but the converse is not necessarily true. One obvious counterexample is and which are non-isomorphic to each other as Lie groups but their Lie algebras are isomorphic to each other. However, for simply connected Lie groups, the Lie group-Lie algebra correspondence is one-to-one.

References