4

Basic Structure Theory of Rings

The structure theory of rings, which we shall consider in this chapter, is an outgrowth of the structure theory of finite-dimensional algebras over a field that was developed during the latter part of the nineteenth century and the early part of this century. The discovery of quaternions by Hamilton in 1843 led to the construction of other “hypercomplex number systems,” that is, finite dimensional algebras over images and images, including some non-associative ones (e.g., the octonions of Cayley and Graves). The problem of classifying such systems was studied intensively during the second half of the nineteenth century by a number of mathematicians: Molien, Frobenius, Cartan, and others. Their results constituted a very satisfactory theory for algebras over images and images, which included a complete classification of an important subclass, the semi-simple algebras. This structure theory was extended to finite dimensional algebras over an arbitrary field by Wedderburn in 1908. In 1927, Artin, stimulated by Emmy Noether’s earlier strikingly successful simplification of ideal theory of commutative rings by the introduction of chain conditions, and motivated by the needs of arithmetic in finite dimensional algebras, extended the Wedderburn theory to rings satisfying chain conditions for one-sided ideals. In 1945 this author developed a structure theory of rings without finiteness assumptions. The principal ingredients of this theory were a new definition of a radical, the concepts of primitivity and semi-primitivity (formerly called semisimplicity), and a density theorem (due independently to Chevalley) that constituted an extensive generalization of a classical theorem of Burnside.

In this chapter we shall reverse the historical order of development by first giving an account of the general theory and then specializing successively to Artin’s theory and to Wedderburn’s theory of finite-dimensional algebras over fields. The Wedderburn structure theorem for simple algebras reduces the classification of these algebras to division algebras. The proper vehicle for studying these is the Brauer group of algebra classes with composition defined by tensor products. We shall give an introduction to the study of these groups and related results on central simple algebras. At the end of the chapter we shall apply the results to the study of Clifford algebras. These are needed to round out the structure theory of orthogonal groups that was presented in Chapter 6 of BAI. Further applications of the theory to representation theory of finite groups will be given in the next chapter.

4.1   PRIMITIVITY AND SEMI-PRIMITIVITY

In the development of ring theory that we shall give, rings of endomorphisms of abelian groups and concretizations of rings as rings of endomorphisms play a predominant role. We recall that if M is an abelian group (written additively), the set of endomorphisms of M has a natural structure of a ring. Thus we obtain End M (or Endimages M), the ring of endomorphisms of M. By a ring of endomorphisms we mean a subring of End M for some abelian group M. We define a representation ρ of a ring R to be a homomorphism of R into a ring End M of endomorphisms of an additive abelian group M. A representation ρ of R acting on M (that is, with codomain End M) defines a left R-module structure on M by specifying that the action of R on M is given through ρ:

images

for aR, xM. Equation (1) defines a left module action since ρ(a) is an endomorphism and a images ρ(a) is a ring homomorphism. Conversely, if M is a left module for the ring R, then M is an abelian group and M affords a representation ρ = ρM of R by defining aM for a ∈ R to be the endomorphism

images

Then ρ : a images aM is a homomorphism of R into End M, hence a representation of R.

By an irreducible representation of a ring we shall mean a representation ρ for which the associated module M (as just defined) is irreducible: M ≠ 0 and M and 0 are the only submodules of M. The structure theory that we shall develop is based on two articles of faith: irreducible representations are the best kinds of representations and the best behaved rings are those that have enough irreducible representations to distinguish between elements of the ring, that is, given a pair of distinct elements a, b of R, there exists an irreducible representation ρ such that ρ(a) ≠ ρ(b). Evidently this is equivalent to the simpler condition that given any a ≠ 0 in R there exists an irreducible representation ρ of R such that ρ(a) ≠ 0.

Before we proceed to the main definitions, it will be useful to collect some elementary facts on representations and modules. Because of the connection between representations and left modules that we have noted, we give preference in this chapter to left modules. Accordingly, in this chapter the unadorned term “module” will always mean left module.

If ρ is a representation of the ring R acting in M, the kernel of ρ is evidently the ideal

images

(bx = ρ(b)x). We shall call the representation ρ (and occasionally the module M) faithful if ker ρ = 0. For any x ∈ M we have the order ideal or annihilator of x, annRx = {b ∈ R|bx = 0}. This is a left ideal and the cyclic submodule Rx generated by x is isomorphic to R/annRx where R is regarded as left R-module in the natural way. It is evident that

images

If f is a homomorphism of a ring S into R and ρ is a representation of R acting on M, then ρf is a representation of S acting on M. The corresponding module action of S on M is given by

images

for bS. As in the last chapter, the notations SM and RM serve to distinguish M as S-module from M as R-module (as well as to indicate that the action is on the left). In most cases, it will be clear from the context which of these is intended and “M” will be an adequate notation. It is clear that

images

where, of course, f−1( ) denotes the inverse image under f. If the representation ρ of R is faithful, then annSM = f−1(0) = kerf

Let B be an ideal of R contained in annRM. Then M becomes an images = R/B module by defining the action of the coset a + B on xM by

images

Since bx = 0 for every bB, xM, it is clear that (6) is a well-defined action. It is immediate also that this is a module action of images on M and it is clear that a subset N of M is an R-submodule if and only if it is an images-submodule. In other words, RM and images have the same lattice of submodules. The relation between the kernels of the representations of R and of images is given by the formula

images

We now proceed to introduce two of our main definitions.

DEFINITION 4.1.   A ring R is called primitive if it has a faithful irreducible representation. R is called semi-primitive if for any a ≠ 0 in R there exists an irreducible representation ρ such that ρ(a) ≠ 0

We shall be interested also in representations that are completely reducible in the sense that the corresponding module M is completely reducible, that is, M = ∑Mα where the Mα are irreducible submodules. We establish first two characterizations of semi-primitivity in the following

PROPOSITION 4.1.   The following conditions on a ring R are equivalent: (1) R is semi-primitive. (2) R has a faithful completely reducible representation. (3) R is a subdirect product of primitive rings (see p. 69).

Proof. (1) images (2). For each a ≠ 0 in R let Ma be an irreducible module such that for the representation ρMa we have ρ Ma(a) ≠ 0. Form M = imagesa≠0Ma. This is a completely reducible module and it is clear that

images

Hence ρM is a faithful completely reducible representation for R.

(2) images (3). Suppose ρ is a faithful completely reducible representation for R so M = ∑Mα, Mα irreducible, for the corresponding module M. We have 0 = annRM = imagesαannRMα, and annRMα is an ideal in R. Hence R is a subdirect product of the rings Rα = R/annRMα. On the other hand, since (annRMα)Mα = 0, Mα can be regarded as an irreducible module for R and the representation of Rα defined by Mα is faithful. Hence Rα is primitive and R is a subdirect product of primitive rings.

(3) images (1). Let R be a subdirect product of the primitive rings Rα. For each Rα let ρα be a faithful irreducible representation of Rα. We have the canonical homomorphism πα of R onto Rα and imagesα ker πα = 0. We have the irreducible representation ραπα of R whose kernel is ker πα. Accordingly, we have a family of irreducible representations {ραπα} of R such that imagesα ker ραπα = 0. Then R is semi-primitive. images

The definitions we have given thus far involve objects (representations and modules) that are external to the ring R. It is easy and useful to replace the definitions by internal characterizations. Let I be a left ideal of the ring R. We define

images

It is clear that if we put M = R/I and regard this as a left R-module then

images

It follows from this or directly from (9) that (I : R) is an ideal. Moreover, by(9), (I : R) ⊂ I and (I : R) contains every ideal of R contained in I. In other words, (I : R) is the (unique) largest ideal of R contained in I. We can now give the following characterization of primitivity and semi-primitivity.

PROPOSITION 4.2. A ring R is primitive if and only if R contains a maximal left ideal I that contains no non-zero ideal of R. A ring R is semi-primitive if and only if R ≠ 0 and imagesI(I : R) = 0 where the intersection is taken over all maximal left ideals I of R.

Proof.   If R is primitive, we have an irreducible module M such that annRM = 0. Now M images R/I for a maximal left ideal I (p. 117). Then (I : R) = annRR/I = annRM = 0. Thus I is a maximal left ideal containing no non-zero ideal of R. Conversely, if the condition holds, we have a maximal left ideal I such that (I : R) = 0. Then M = R/I provides a faithful irreducible representation of R, so R is primitive.

Now let R be semi-primitive and let {ρα} be a set of irreducible representations of R that is adequate for distinguishing the elements of R. Then images ker ρα = 0. If Ma is the module for ρα then Mα images R/Iα, Iα a maximal left ideal in R. Hence (Iα : R) = annRMα = ker ρα and imagesα(Iα : R) = 0. A fortiori, images(I : R) taken over all of the maximal left ideals of R is 0. The converse follows by retracing the steps. images

The internal characterizations of primitivity and semi-primitivity have some important consequences that we shall now record. The first of these is

COROLLARY 1.   Any simple ring (≠0) is primitive.

Proof.   Any ring R ≠ 0 contains a maximal left ideal I (p. 68). If R is simple, (I : R) = 0. Hence R is primitive. images

Next we obtain characterizations of primitive and semi-primitive commutative rings.

COROLLARY 2.   Let R be a commutative ring. Then R is primitive if and only if R is a field and R is semi-primitive if and only if it is a subdirect product of fields.

Proof.   If R is a commutative primitive ring, R contains a maximal left ideal I containing no ideal ≠0 of R. Since R is commutative, I is an ideal. Hence I = 0. Then 0 is a maximal ideal in R, which means that R ≠ 0 and 0 and R are the only ideals in the commutative ring R. Then R is a field. Conversely, it is clear that any field satisfies the first condition of Proposition 4.2. Hence any field is primitive. It now follows from Proposition 4.1.3 that R is semi-primitive if and only if it is a subdirect product of fields. images

We shall now give some examples of primitive and semi-primitive rings. A convenient way of constructing examples of rings is as rings of endomorphisms of abelian groups. If R is a ring of endomorphisms of an abelian group M, then M is an R-module in the natural way in which the action ax, aR, xM is the image of x under a. Obviously, the corresponding representation is the injection of R into End M and so this is faithful. Hence if R acts irreducibly on M ≠ 0 in the sense that there is no subgroup of M other than 0 and M that is stabilized by R, then R is a primitive ring. Several of the following examples are of this type.

EXAMPLES

1. The ring L of linear transformations of a finite dimensional vector space over a division ring is simple (p. 171). Hence L is primitive.

2. Let V be a vector space over a division ring Δ that need not be finite dimensional. We claim that the ring L of linear transformations in V over Δ acts irreducibly on V and hence is a primitive ring. To prove irreducibility, we have to show that if x is any non-zero vector and y is any vector, there exists an lL such that Ix = y. Now this is clear, since we can take x to be an element in a base ( for V over Δ, and given any base (eα) and corresponding elements fα for every eα, then there exists a linear transformation l such that leα = fα. If V is infinite dimensional, the subset F of transformations l with finite dimensional range (l(V) finite dimensional) is a non-zero ideal in L. Since 1 images F, FL. Hence L is not simple, so this is an example of a primitive ring that is not simple. If the dimensionality of V is countably infinite, it is easy to see that F is the only proper non-zero ideal of L. Hence L/F is simple and thus primitive.

3. If (x1,x2,x3,…) is a base for a vector space V over a division ring Δ, then a linear transformation l is determined by its action on the base. We can write

images

where the λij∈ Δ and for a given i only a finite number of the λij are ≠0. Thus l can be described by a matrix images = (λij), which is row finite in the sense that there are only a finite number of non-zero entries in each row. The set of row-finite matrices with entries in Δ is a ring under the usual matrix addition and multiplication, and the map I images images where images is the matrix of l relative to the base (xi) is an anti-isomorphism. Now let R be the set of linear transformations whose matrices relative to the given base have the form

images

where Φ is a finite square block and λ ∈ Δ. The set of matrices of the form (12) is a subring of the ring of row-finite matrices. Hence R is a subring of the ring of linear transformations L. R acts irreducibly on V. For, if x is a non-zero vector and y is any vector, then we can write x = ξ1x1 +… + ξnxn, y = η1x1 + … + ηnxn for some n where the ξl, ηi ∈ Δ. Thus x and y are contained in a finite dimensional subspace V′ of V. There exists a linear transformation l′ of V′ into itself such that l′x = y and l′ can be extended to a linear transformation l of V contained in R (for example, by specifying that lxi = 0 if i > n). Then lx = y. Hence V is irreducible as R-module and so R is primitive.

4. Let V be as in example 3 with Δ = F, a field, and let p and q be the linear transformations in V over F such that

images

Then pqxi = xi for all i and qpx1 = 0, qpxi = xi for i > 1. Hence

images

Let R be the set of linear transformations

images

where the αijx. The first relation in (14) implies that R is closed under multiplication, so R is a ring of endomorphisms of V. Now put

images

The matrix of eij has a 1 in the (j,i)-position, 0’s elsewhere. It follows that the matrix ring corresponding to R includes all the matrices (12). Thus R contains the ring given in example 3 as a subring. Hence the present R acts irreducibly on V, so R is primitive.

5. For any ring R we have the regular representation ρR whose module is the additive group of R on which the ring R acts by left multiplication (see BAI, pp. 422–426). This representation is faithful and the submodules of R are the left ideals. Now let A be an algebra over the field F. Then the regular representation ρA of A provides a monomorphism of A into the algebra L of linear transformations of the vector space A/F. Thus any algebra over a field can be imbedded in a primitive ring. We shall now show that A is also a homomorphic image of a primitive ring. Since A is an arbitrary algebra over a field, this will show that there is not much that can be said about homomorphic images of primitive rings. For our construction we take V = imagesA, a direct sum of a countably infinite number of copies of the vector space A. Let A act on V in the obvious way and identify A with the corresponding algebra of linear transformations in V. For n = 1, 2, 3,… let Vn be the subspace of V of elements of the form (x1,x2,…,xn,0, 0,…), xi ∈ A, and let Ln be the set of linear transformations that map Vn into itself and annihilate all of the elements (0,…,0,xn+ 1,xn+ 2,…). Then L1L2. Put L = imagesLi and let R be the ring of linear transformations generated by A and L. Then R acts irreducibly on V, so R is primitive. Moreover, R = A+L, AL = 0, and L is an ideal in R. Then R/L images A, so the given algebra A is a homomorphic image of the primitive algebra R.

6. The ring images is semi-primitive since imagespprime(p) = 0 and images/(p) is a field. Hence images is a subdirect product of fields, so images is semi-primitive. Similarly any p.i.d. with an infinite number of primes is semi-primitive.

EXERCISES

        1. Let F{x,y} be the free algebra generated by x and y, and let K be the ideal in F{x,y} generated by xy – 1. Show that F{x,y}/K is isomorphic to the algebra R defined in example 4 above.

        2. (Samuel.) Let V,F be as in example 4 and let S be the algebra of linear transformations generated by p as in example 4 and the linear transformation r such that rxi = xi2 + 1. Show that S is primitive and S is isomorphic to the free algebra F{x,y}.

        3. Let F{x,y} be as in exercise 1 and let K be the ideal generated by xyyxx. Assume that F is of characteristic 0. Show that R = F{x,y}/K is primitive.

        4. Let R be a ring containing an ideal N ≠ 0 that is nilpotent in the sense that there exists an integer m such that Nm = 0 (that is, the product of any m elements of N is 0). Show that R is not semi-primitive.

        5. Show that if R is primitive then the matrix ring Mn(R) is primitive.

        6. Show that if R is primitive and e is an idempotent element ≠0 in R, then the ring eRe (with unit e) is primitive.

        7. Show that if R and R′ are Morita similar (p. 179), then R is primitive if and only if R′ is primitive.

        8. Let F be a field, P = F[t1,…, tn], the ring of polynomials in n indeterminates with coefficients in F. Show that for any i = 1, 2,…,n, there exists a unique derivation Di of P such that

images

(see exercise 11, p. 147). For fP let images denote the multiplication g images fg in P. Show that if D is a derivation in P and fP, then imagesD is a derivation in P. Show that the derivations of P into P are the maps images where Di is as above and flP. Let R be the ring of endomorphisms of the additive group of P generated by the derivations and the multiplications of P. Show that if F is of characteristic 0, then R is an irreducible ring of endomorphisms and hence R is primitive.

        9. Let F be a field of characteristic 0 and F{xl,…, xn, y1,…, yn} the free algebra over F determined by the 2n elements xl,…, xn, yl,…, yn. Let I be the ideal in F{xl,…, xn,…, yl,…, yn} generated by the elements

images

1 ≤ i,jn. The algebra Wn = F{xl,…, xn, yl,…, yn}/I is called a Weyl algebra. Show that if P and R are as in exercise 8, then the homomorphism η of F{x1,…,xn, y1,…,yn} into R such that xiimages imagesi, yt images Dl, 1 ≤ in, is surjective and ker η = l Hence conclude that Wn is primitive.

     10. Show that Wn is simple. (Suggestion: First treat the case in which n = 1.)

4.2   THE RADICAL OF A RING

From the point of view that we took in the previous section, the purpose of defining the radical of a ring is to isolate that part of the ring whose elements are mapped into 0 by every irreducible representation of the ring. Accordingly, we introduce the following

DEFINITION 4.2.   The (Jacobson) radical of a ring is the intersection of the kernels of the irreducible representations of the ring.

Evidently, the radical, rad R, of the ring R is an ideal in R.

We shall call an ideal P of R primitive (in R) if R/P is a primitive ring. Let l be a maximal left ideal of R and let P = (I : R). Then M = R/I is an irreducible module for R whose annihilator is P. Hence M can be regarded as an images = R/P-module by (6) and M is irreducible as images-module. By (7), annimagesM = annRM/P = P/P = 0. Hence M is a faithful irreducible module for images and so images is primitive. Then, by definition, P is a primitive ideal in R. Conversely, let P be a primitive ideal in R. Then we have an irreducible module M for images = R/P such that the associated representation is faithful. Regarding M as R-module via (4) we see that M is an irreducible R-module such that annRM = P (see(5). Now M images R/I where I is a maximal left ideal of R and hence annRR/I = P. Since annRR/I = (I : R), we have P = (I: R). We therefore have the

LEMMA 1.   An ideal P of R is primitive in R if and only if P = (I : R) for some maximal left ideal I of R.

We can now give our first internal characterization of the radical.

PROPOSITION 4.3. (1) rad R is the intersection of the primitive ideals of R. (2) rad R is the intersection of the maximal left ideals of R.

Proof. (1) By definition,

images

where {M} is the class of irreducible modules of R. Since M images R/I and ann R/I = (I : R), we have

images

where I runs over the maximal left ideals of R. By the lemma, this can be written also as

images

where P runs over the primitive ideals in R. This proves (1). To prove (2), we recall that for any module M we have annRM = imagesx∈M annRx = imagesx ≠ 0 annRx. If M is irreducible, this expresses annRM as an intersection of maximal left ideals. Hence, by (17), rad R is an intersection of maximal left ideals, so rad RimagesI where I ranges over the maximal left ideals of R. On the other hand, since I ⊃ (I: R), imagesIimages(I : R) = rad R so

images

where the intersection is taken over the set of maximal left ideals of R. images

We prove next

PROPOSITION 4.4. (1) R is semi-primitive if and only if rad R = 0. (2) If R ≠ 0, then R/rad R is semi-primitive and rad R is contained in every ideal B of R such that R/B is semi-primitive.

Proof. (1) If rad R = 0, then imagesP = 0 for the primitive ideals of R. Then R is a subdirect product of the primitive rings R/P and R is semi-primitive by Proposition 4.1. Conversely, if R is semi-primitive, R is a subdirect product of primitive rings and so there exists a set of primitive ideals whose intersection is 0. Then rad R = imagesP = 0 if P runs over the set of primitive ideals of R.

(2) If B is an ideal of R, then any ideal of images = R/B has the form P/B where P is an ideal of R containing B. Since R/P images images/images where images = P/B, it is clear that images is primitive in images if and only if P is primitive in R. Since images is semiprimitive if and only if imagesimages = 0, images primitive in images, it follows that images is semiprimitive if and only if B is the intersection of all of the primitive ideals of R containing B. Then B ⊃ rad R, by (18′). If B = radR, (18′) implies that B is an intersection of primitive ideals. Then R/B = R/rad R is semi-primitive. This proves (2). images

We shall give next an important element characterization of the radical. For this purpose, we introduce a number of definitions. First, we call an element z of a ring R left (right) quasi-regular in R if 1 — z has a left (right) inverse in R. Evidently, left (right) quasi-regularity is equivalent to the following: the principal left (right) ideal R(1 — z) ((1 — z)R) = R. If z is both left and right quasi-regular, then z is called quasi-regular. A left (right) ideal I is called quasiregular if all of its elements are left (right) quasi-regular.

If z is left quasi-regular, 1 — z has a left inverse that we can write as 1 — z′. Then (1 — z′) (í — z) = 1 gives the relation

images

The binary product z′ O z = z + z′ — zz defines the circle composition O in the ring R. Let σ denote the map x images 1 — x in R. Then σ2 = 1, so σ is bijective. We have

images

This implies that the circle composition is associative. Also since σ(0) = 1, 0oz = z = zo0. Thus (R,o,0) is a monoid.

LEMMA 2. (1) If I is a quasi-regular left (right) ideal, then I is a subgroup of (R, o, 0). (2) Nilpotent elements are quasi-regular.

Proof. (1) Let z ∈ l. Then we have a z′ ∈ R such that z′ Ο z = 0. Then z + z′ — zz = 0 and z′ = zzzI. Hence there exists a z″ such that z″ O z′ = 0. Since (R,O,0) is a monoid, the equations z′ O z = 0 = z″ O z′ imply that z″ = z, so z′ is the inverse of z in (R, O, 0) and z′ ∈ I. Since I is closed under O and contains 0, statement (1) is clear.

(2) If zn = 0, then (1 — z) (1 + z + z2 + … + zn−1) = 1 = (1 + z + z2 + … + zn−1) (1 — z). Hence z is quasi-regular. images

An ideal (left, right ideal) is called nil if all of its elements are nilpotent. The second part of Lemma 2 shows that any nil left (right) ideal is quasi-regular.

We can now give the following element characterizations of rad R.

PROPOSITION 4.5. (1) radR is a quasi-regular left ideal that contains every quasi-regular left ideal of R. (2) rad R is the set of elements z such that az is left quasi-regular for every aR.

Proof. (1) Suppose z ∈ rad R is not left quasi-regular. Then R(l —z) ≠ R, so this left ideal can be imbedded in a maximal left ideal I. Since rad R is the intersection of the maximal left ideals of R, zI. But 1 – zR(1 – z) ⊂ I, so 1 = 1 – z + zI contrary to I images R. This contradiction shows that every z ∈ rad R is left quasi-regular, so it proves that rad R is a quasi-regular left ideal. Now let Z be any quasi-regular left ideal of R. If Z images z rad R, then there exists a maximal left ideal I such that Z images I. Then I + Z images I and so I + Z = R. Then 1 = b + z for some bI, zZ, and 1 = (1 — z)−1bI contrary to I images R. Thus ZR and part (1) is proved.

(2) If z ∈ rad R, then az ∈ rad R for any a in R, so az is left quasi-regular. Conversely, suppose z satisfies this condition. Then Rz is a quasi-regular left ideal. Hence Rz ⊂ rad R by (1) and z ∈ rad R. images

The concept of primitivity of a ring and of an ideal in a ring is not left-right symmetric. For, there exist primitive rings that are not right primitive in the sense that they have no irreducible right modules M′ such that the associated anti-representation ρ′ :aimages ρ′(a), where ρ′(a)x = xa, is monic. The first examples of such rings were given by George Bergman. In spite of this lack of symmetry in the concept of primitivity, the concepts of the radical and semi-primitivity are symmetric. Evidently, everything we have done can be carried over to antirepresentations, right modules, and right ideals. Suppose we denote the corresponding right radical as rad′ R. Then we have the analogue of Proposition 4.5 for rad′ R. On the other hand, by Lemma 2, the elements of rad R and rad′ R are quasi-invertible. Hence, by Proposition 4.5 and its analogue for right ideals, we have rad′ R ⊂ rad R and rad Rrad′ R. Thus rad R = rad′ R. This permits us to give a number of additional characterizations of rad R. We give these and summarize the ones we obtained before in

THEOREM 4.1.   We have the following characterizations of the radical of a ring: radR is

            1. the intersection of the primitive ideals of R,

            2. the intersection of the maximal left ideals of R,

            3. the intersection of the right primitive ideals of R,

            4. the intersection of the maximal right ideals of R,

            5. a quasi-regular left ideal containing every quasi-regular left ideal,

            6. the set of elements z such that az is left quasi-regular for every aR,

            7. a quasi-regular right ideal containing every quasi-regular right ideal,

            8. the set of elements z such that za is right quasi-regular for every aR.

It is clear also that rad R contains every nil right ideal as well as every nil left ideal of R. Moreover, since right semi-primitivity (defined in the obvious way) is equivalent to rad′R = 0, it is clear that a ring is semi-primitive if and only if it has a faithful completely reducible anti-representation and if and only if it is a subdirect product of right primitive rings. Another way of putting the left-right symmetry is that if we regard R and Rop as the same sets, then rad R = rad Rop and if R is semi-primitive, then so is Rop.

EXERCISES

        1. Show that a ring R is a local ring (p. 111) if and only if R/rad R is a division ring and that if this is the case, then rad R is the ideal of non-units of R.

        2. Determine the radical of images/(n), n ≥ 0.

        3. (McCrimmon.) Show that z ∈ rad R if and only if for every aR there exists a w such that z + w = waz = zaw.

        4. Show that ab is quasi-regular if and only if ba is quasi-regular.

        5. Call an element a of a ring R von Neumann regular if there exists a b such that aba = a. Show that the only element in rad R having this property is 0. (A ring is called von Neumann regular if all of its elements are von Neumann regular. Hence this exercise implies that such rings are semi-primitive.)

        6. Let R be a ring such that for any a ∈ R there is an integer n(a) > 1 such that an(a) = a. Show that R is semi-primitive.

        7. Let e be a non-zero idempotent in a ring R. Show that rad eRe = e(rad R)e = eRe ∩ rad R. (Sketch of proof: Show that every element of e(rad R)e is quasiregular in eRe, so e(rad R)e ⊂ rad eRe. Next let z ∈ rad eRe and let aR. Write a = eae + ea(1 – e) + (1 – e)ae + (1 – e)a(1 – e) (the two-sided Peirce decomposition). Then za = zeae + zea( 1 –e) and zeae has a quasi-inverse z′ in eRe. Then za O z′ = zea(1 – e), which is nilpotent, thus quasi-regular. Hence za is right quasiregular. Hence z ∈ rad ReRe and rad eRee(rad R)e.)

        8. Show that for any ring R, rad Mn(R) = Mn(rad R). (Sketch of proof: We have a 1–1 correspondence B images Mn(B) of the set of ideals of R onto the set of ideals of Mn(R) (see exercise 1, p. 171). Let {ei|1 ≤ i, jn} be a set of matrix units for Mn(R). We have the isomorphism r images (r1)e11 of R onto e11 Mn(R)e11. Under this the image of rad R is rade11Mn(R)e11 = e11 radMn(R)e11 (by exercise 7) = e11Mn(B)e11. It follows that B = rad R.)

        9. Show that if R and R′ are (Morita) similar, then in any correspondence between the ideals of R and R′ given by Morita I, the radicals of R and R′ are paired.

     10. Prove that any ultraproduct of semi-primitive (primitive) rings is semi-primitive (primitive).

4.3   DENSITY THEOREMS

If V is a vector space over a division ring Δ, then a set S of linear transformations in V is called dense if for any given finite set of linearly independent vectors x1,x2, …, xn and corresponding vectors y1, y2, …, yn there exists an lS such that lxi = yi 1 ≤ in. If V is finite dimensional, then we can take the xi to be a base and yi = axi for any given linear transformation a. Then we have an lS such that lxi = axi 1 ≤ in, from which it follows that l = a. Hence the only dense set of linear transformations in a finite dimensional vector space is the complete set of linear transformations.

In this section we shall prove a basic theorem that permits the identification of any primitive ring with a dense ring of linear transformations. We consider first a more general situation in which we have an arbitrary ring R and a completely reducible (left) module M for R. Let R′ = EndRM, R″ = EndR.M where M is regarded in the natural way as a left R′-module. We shall prove first the

DENSITY THEOREM FOR COMPLETELY REDUCIBLE MODULES.   Let M be a completely reducible module for a ring R. Let R′ = EndR M, R″ = EndR M. Let {xu,…, xn} be a finite subset of M, a″ an element of R″. Then there exists an aR such that axi = a″xi, 1 ≤ in.

The proof we shall give is due to Bourbaki and is based on the following pair of lemmas.

LEMMA 1.   Any R-submodule N of M is an R″-submodule.

Proof.   Since M is completely reducible as R-module, we can write M = N images P where P is a second submodule (p. 121). Let e be the projection on N determined by this decomposition. Then eR′ and N = e(M). Hence if a″R″, then a″(N) = ae(M) = ea″(M) ⊂ N. Thus N is a submodule of R″ M. images

LEMMA 2.   Let M be a module, M(n) the direct sum of n copies of M, n = 1, 2,… Then EndRM(n) is the set of maps (ul, u2, …, un) images (v1 v2, …, vn) where vl = ∑, a′ijuj, a′ijR′ = EndRM. Moreover, for any a″ ∈ R″ = EndRM, the map (ul, u2, …, un)images(a″u1, a″u2, …, a″un) is contained in the ring of endomorphisms of M(n) regarded as a left EndRMn-module.

Proof.     Let l ∈ EndRM(n), and consider the action of I on the element (0, …, 0, ui, 0, …, 0) where ui is in the ith place. We have 1(0,…, 0, ui0, …, 0) = (u1i, u2i, …, uni), which defines the map ui images uji, 1 ≤ jn. This is contained in EndRM. Denoting this map as a′ji we obtain l(u1, …, un) = images. It is clear also that any map of this form is contained in EndRM(n). This proves the first assertion. The second follows, since for any a″ ∈ R″ the map (u1, u2, …, un) images (au1, au2, …, aun) commutes with every map (u1, u2, …, un) images images, ajiR′. images

We can now give the

Proof of the theorem. Suppose first that n = 1. Then N = Rx1 is an R -submodule, hence an R″ -submodule. Since x1N, ax1N = Rx1. Hence we have an aR such that ax1 = ax1. Now suppose n is arbitrary. Consider M(n) the direct sum of n copies of the R-module M. The complete reducibility of M implies that M(n) is completely reducible. By Lemma 2, if a″ ∈ R″, the map (ul, u2, …, un) (au1, au2, …, aun) is in the ring of endomorphisms of M(n), regarded as left EndRM(n) module. We apply the result established for n = 1 to this endomorphism and the element x = (x1, …, xn) of M(n). This gives an element aR such that ax = (axl, …, axn). Then axi = axi, 1 ≤ in, as required. images

There is a good deal more that can be done at the level of completely reducible modules (see the exercises below). However, for the sake of simplicity, we shall now specialize to the case that is of primary interest for the structure theory: a faithful irreducible representation of a ring. In this case we have the

DENSITY THEOREM FOR PRIMITIVE RINGS.   A ring R is primitive if and only if it is isomorphic to a dense ring of linear transformations in a vector space over a division ring.

Proof.   Suppose R is a primitive ring, so R has a faithful irreducible representation ρ acting in M. By Schur’s lemma, Δ = EndRM is a division ring. Since ρ is faithful, R is isomorphic to ρ(R), which is a ring of linear transformations in M regarded as a vector space over Δ. Now let x1, …, xn be linearly independent vectors in M, y1, …, yn arbitrary vectors. Since the x’s can be taken to be part of a base, there exists a linear transformation l such that lxi = yi, 1 ≤ in. Since l ∈ EndΔM and Δ = EndRM, it follows from the previous theorem that there exists an aR such that axi = lxi = yi, 1 ≤ in. Then ρ(R) is a dense ring of linear transformations in M over Δ. Conversely, suppose R is isomorphic to a dense ring of linear transformations in a vector space M over a division ring Δ. If ρ is the given isomorphism, then putting ax = ρ(a)x for aR, xM, makes M an R-module. Moreover, this is irreducible since if x ≠ 0 and y is arbitrary, we have an aR such that ax = y. Hence ρ is a faithful irreducible representation of R and so R is primitive. images

For some applications it is important to have an internal description of the division ring Δ in the foregoing theorem. This can be obtained by identifying an irreducible R-module with an isomorphic one of the form R/I where I is a maximal left ideal of R. We need to determine EndR(R/I). We shall do this for any left ideal I or, equivalently, for any cyclic module. We define the idealizer of I in R by

images

It is clear that this is a subring of R containing I as an ideal and, in fact, B contains every subring of R having this property. We have

PROPOSITION 4.6.   If I is a left ideal in R, then EndR(R/I) is anti-isomorphic to B/I where B is the idealizer of I in R.

Proof.   Since 1 + I is a generator of R/I, f ∈ EndR(R/I) is determined by f(1 + I) = b + I = b(1 + l) as the map

images

If al, a + I = I = 0 in R/I and hence its image ab + I = I. Then abl. Thus b ∈ the idealizer B of I. Conversely, let bB and consider the homomorphism of the free R-module R into R/I sending 1 images b + I. Any dl is mapped into db + l = l by this homomorphism. Hence we have an induced homomorphism of R/I into R/I sending 1 + I images b + I. Thus any bB determines an endomorphism of R/I given by (22). Denote the correspondence between elements of B and endomorphisms of R/I determined in this way by g. Evidently g is additive and g(1) = 1. Also if b1, b2B, we have g(b1b2)(1 + l) = b1b2 + I and g(b2)g(b1) (1 + l) = g(b2)(b1 + l) = b1b2 + I. Hence g is an antihomomorphism of B onto EndR(R/I). The ideal in B of elements mapped into 0 by this map is l. Hence B/I and EndR(R/I) are anti-isomorphic. images

The reader should be warned that in spite of this nice determination of the division ring Δ in the density theorem, given by Proposition 4.6, this division ring is not an invariant of the primitive ring. A given primitive ring may have non-isomorphic irreducible modules giving faithful representations and even some irreducible modules whose endomorphism division rings are not isomorphic. There are important cases in which the division ring is an invariant of the primitive ring. For example, as we shall show in the next section, this is the case for artinian primitive rings. Our next theorem gives the structure of these rings.

THEOREM 4.2.   The following conditions on a ring R are equivalent:

            (1) R is simple, left artinian, and non-zero.

            (2) R is primitive and left artinian.

            (3) R is isomorphic to a ring EndΔM where M is a finite dimensional vector space over a division ring Δ

Proof.   (1) images (2) is clear.

(2) images (3). Let M be a faithful irreducible module for R, Δ = EndRM. We claim that M is finite dimensional over Δ. Otherwise, we have an infinite linearly independent set of elements {xi|i = 1, 2, 3,…} in M. Let Ij = annRxj so Ij is a left ideal. Evidently, images is the subset of R of elements annihilating x1,…, xn. By the density theorem there exists an aR such that ax1 = = axn = 0, axn + l ≠ 0. Hence images Then

images

is an infinite properly descending sequence of left ideals in R contrary to the artinian property of R. Thus M is finite dimensional over Δ and if we let (x1, x2, …, xn) be a base for M over Δ, the density property implies that for any y1, …, yn there exists an aR such that axi = yi, 1 ≤ in. Hence if ρ is the representation determined by M, ρ(R) = EndΔM so R is isomorphic to the ring of linear transformations in the finite dimensional vector space M over Δ.

(3) images (1). This has been proved in proving the Wedderburn-Artin theorem (p. 171). It can also be seen directly by using the anti-isomorphism of EndΔM with Mn(Δ) and recalling that if R is any ring, the map B images Mn(B) is an isomorphism of the lattice of ideals of R with the lattice of ideals of Mn(R) (exercise 8, p. 103 of BAI; or exercise 1, p. 171). The result then follows from the simplicity of Δ. images

EXERCISES

        1. Let R be a primitive ring, M a faithful irreducible module for R, Δ = EndRM. Show that either R images Mk(Δ) for some n or for any k = 1, 2, 3,…, R contains a subring having Mk(Δ) as a homomorphic image.

        2. Use exercise 1 to show that if R is a primitive ring such that for any aR there exists an integer n(a) > 1 such that an(a) = a, then R is a division ring.

        3. Call a set S of linear transformations k-fold transitive if for any l ≤ k linearly independent vectors x1,…, xl and arbitrary vectors y1,…, yl there exists an aS such that axi = yi 1 ≤ il. Show that a ring of linear transformations that is two-fold transitive is dense.

In the next three exercises, M = RM is completely reducible, R′ = EndRM, R″ = EndR M.

        4. Show that any homomorphism of a submodule RN of RM into RM can be extended to an endomorphism of RM (an element of R′) and that if RN is irreducible, any non-zero homomorphism of RN into RM can be extended to an automorphism of RM.

        5. Show that if x ≠ 0 is an element of an irreducible submodule RN of RM, then R′x is an irreducible submodule of R′ M. (Hint: Let y = a′x ≠ 0, a′ ∈ R′, and consider the map u images a′u of RN = Rx into RM.)

        6. Show that RM is completely reducible and that its dimensionality (defined on p. 125) is finite if R is left artinian. Show also that in this case R″ = ρ(R) where ρ is the representation defined by M.

        7. Let M be a completely reducible R-module, B an ideal in R. Show that the following conditions are equivalent: (i) BM = M. (ii) If xM satisfies Bx = 0, then x = 0. (iii) For any x, xBx.

        8. Prove the following extension of the density theorem for completely reducible modules: If R, M, and B are as in exercise 7 and B satisfies (i), (ii), (iii), then for any a″ ∈ R″ and x1,…, xnM there exists a bB such that bxi = a″ xi, 1 ≤ in.

        9. Use exercise 8 to prove the simplicity of the ring of linear transformations in a finite dimensional vector space over a division ring.

4.4   ARTINIAN RINGS

In this section we shall derive the classical results on structure and representations of (left) artinian rings. We shall show first that the radical of such a ring is nilpotent. We recall that if B and C are ideals of a ring R, then BC = images. This is an ideal in R. Also, if D is another ideal, then (BC)D = B(CD). We define Bk inductively by B1 = B, Bi = Bi − 1 B and B is called nilpotent if Bk = 0 for some k. This is equivalent to the following: the product of any k elements of B is 0. In particular, bk = 0 for every B. Hence B nilpotent implies that B is a nil ideal. On the other hand, it is easy to give examples of nil ideals that are not nilpotent (exercise 1, at the end of this section).

THEOREM 4.3.   The radical of a (left) artinian ring is nilpotent.

Proof.   Let N = rad R, R artinian. By induction, NN2N3 ⊃ …. Since these are left ideals, there is a k such that PNk = Nk + 1 = …. Then P = P2 = N2k. N is nilpotent if and only if P = 0. Suppose this is not the case and consider the set S of left ideals I of R such that (1) IP and (2) PI ≠ 0. Since P has these properties, S is not vacuous, and since R is artinian, S contains a minimal element I. Evidently, I contains an element b such that Pb ≠ 0. Then Pb is a left ideal contained in I and in P and P(Pb) = P2b = Pb ≠ 0. Hence Pb = I by the minimality of I in S. Since bI, we have a zP such that zb = b or (1 − z)b = 0. Since zP ⊂ rad R, z is quasi-regular. Hence (1 − z)−1 exists and b = (1 − z)−1(l − z)b = 0. This contradicts Pb ≠ 0 and proves (rad R)k = P = 0. images

We recall that the radical of any ring contains all quasi-regular one-sided ideals, hence all nil one-sided ideals of the ring. The foregoing result implies that all such ideals in an artinian ring are nilpotent. We shall now derive the main result on the structure of semi-primitive artinian rings. For the proof we shall need the

LEMMA.   Let M be an artinian module that is a subdirect product of irreducible modules. Then M is a direct sum of a finite number of irreducible submodules.

Proof.   The second hypothesis is equivalent to the following: M contains a set of submodules {Nα} such that ∩Nα = 0 and every Mα = M/Nα is irreducible. Consider the set of finite intersections images of images. The artinian condition on M implies that there exists a minimal element in this set, say, N1 ∩ … ∩ Nm. Then for any Nα, Nα ∩ (N1 ∩ … ∩ Nm) = N1∩ … ∩ Nm so NαN1 ∩ … ∩ Nm. Since imagesNα = 0, we have Nl ∩ … ∩ Nm = 0. Then we have a monomorphism of M into images. Thus M is isomorphic to a submodule of the completely reducible module imagesMi. Then M is completely reducible and is a direct sum of irreducible modules isomorphic to some of the Mi. images

We shall now call a ring semi-simple if it is a subdirect product of simple rings. We have the following easy consequence of the preceding lemma.

PROPOSITION 4.7.   If R is semi-simple and R satisfies the minimum condition for two-sided ideals, then R is a direct sum imagesn1Ri of simple rings Ri.

Proof.   Let M(R) be the ring of endomorphisms of the additive group of R generated by the left and the right multiplications x images ax and x images xa. This ring is called the multiplication ring of R. The additive group R can be regarded in the natural way as M(R)-module and when this is done, the submodules of M(R)R are the ideals of the ring R. The hypothesis that R is semi-simple is equivalent to the following: R as M(R)-module is a subdirect product of irreducible M(R)-modules. The other hypothesis of the lemma is also fulfilled. The conclusion then follows from the lemma. images

We can now prove the main

STRUCTURE THEOREM FOR SEMI-PRIMITIVE ARTINIAN RINGS.   The following conditions on a ring R are equivalent:

            (1) R is artinian and contains no nilpotent ideals ≠ 0.

            (2) R is semi-primitive and artinian.

            (3) R is semi-simple and artinian.

            (4) RR is a completely reducible R-module.

            (5) R is a direct sum of a finite number of rings Ri each of which is isomorphic to the ring of linear transformations of a finite dimensional vector space over a division ring.

The implication (1) images (5) is called the Wedderburn-Artin theorem.

Proof. (1) images (2). This is clear, since the radical of an artinian ring is a nilpotent ideal containing all nilpotent ideals and rad R = 0 is equivalent to semi-primitivity.

(2) images (3) is clear from Theorem 4.2, which establishes the equivalence of simplicity and primitivity for artinian rings.

(2) ⇒ (4). Since rad R is the intersection of the maximal left ideals, semi-primitivity of R implies that RR is a subdirect product of irreducible modules. Then (4) follows from the lemma.

(4) ⇒ (2). The semi-primitivity is clear, since the representation determined by RR is faithful and completely reducible. Also we have R = ∑, a minimal left ideal. Then images. Then for aR, we have a = a1 = images and hence R = I1 imagesI2 imagesimages Il for a subset of the Ij. Then it is clear that

images

is a composition series for RR. The existence of such a series implies that R is artinian and noetherian. Then (2) holds.

(3) images (5). Evidently the artinian property implies that R satisfies the minimum condition for two-sided ideals. Hence, by Proposition 4.7, R is a direct sum of simple rings. Thus images where the Ri are ideals and are simple rings. Then for ij, RiRjRiRj = 0. It follows that any left ideal of Ri is a left ideal of R; hence, each Ri is artinian. Then Ri is isomorphic to the ring of linear transformations of a finite dimensional vector space over a division ring. Hence (5) holds.

(5) images(2). If images where the Ri are ideals and Ri is isomorphic to EndΔl Mi, Mi a finite dimensional vector space over the division ring Δl, then Rl is primitive artinian by Theorem 4.2, and R is a subdirect product of primitive rings, so R is semi-primitive. Since Ri is a submodule of RR and the R-submodules of Ri are left ideals, RRi is artinian. Hence R is artinian. Thus (2) holds. This completes the proof. images

Let R be any ring and suppose R = R1 imagesimagesRS where the Ri are ideals that are indecomposable in the sense that if images ideals, then either R′. = 0 or R″ = 0. Now let B be any ideal in R. Then B = B1 = BR = BR1 +…+ BRS and since BRiRi, we have B = BR1 imagesimages BRS. Also BRi is an ideal of R, so if B is indecomposable then we have BRl = B for some i. Similarly, we have B = RjB for some j and since RiRj = 0 if ij, we have B = RiB. It follows that if images is a second decomposition of R as a direct sum of indecomposible ideals, then s = t and the R′j are, apart from order, the same as the Ri. This applies in particular to the decomposition of a semi-simple artinian ring as direct sum of ideals that are simple rings. We see that there is only one such decomposition. Accordingly, we call the terms Ri of the decomposition the simple components of R.

The structure theorem for semi-primitive artinian rings (or Theorem 4.2) shows that if R is simple artinian, then R images EndΔM, M a finite dimensional vector space over the division ring Δ. We now consider the question of uniqueness of M and Δ. The problem can be posed in the following way: Suppose EndΔ1M1 and EndΔ2M2 are isomorphic for two finite dimensional vector spaces Mi over Δi; what does this imply about the relations between the two spaces? We shall see that the vector spaces are semi-linearly isomorphic. We recall that if σ is an isomorphism of Δ1 onto Δ2, then a map s : M1M2 is called a a-semi-linear map if

images

for x,yM1 δ ∈ Δ1 (BAI, p. 469). It is immediately apparent that if s is bijective, then s–1 is a σ‒ 1 -semi-linear map of M2 onto M1. We shall say that M1 and M2 are semi-linearly isomorphic if there exists a semi-linear isomorphism ( = bijective semi-linear map) of M1 onto M2. This assumes tacitly that the underlying division rings Δ1 and Δ2 are isomorphic. Moreover, if {xα} is a base for M1 over Δ1 and s is a semi-linear isomorphism of Ml onto M2, then {Sxα} is a base for M2 over Δ2. Hence the spaces have the same dimensionality. Conversely, let M1 over Δ1 and M2 over Δ2 have the same dimensionality and assume that Δ1 and Δ2 are isomorphic. Let {xα}, {yα} be bases for M1 and M2 respectively, indexed by the same set, and let σ be an isomorphism of Δ1 onto Δ2. Define the map s of M1 into M2 by

images

One checks directly that s is σ-semi-linear and bijective. Hence we see that M1 over Δ1 and M2 over Δ2 are semi-linearly isomorphic if and only if they have the same dimensionality and Δ1 and Δ2 are isomorphic.

The key result for the isomorphism theorem (and for representation theory as well) is the following

LEMMA 1.   Any two irreducible modules for a simple artinian ring are isomorphic.

Proof.   We choose a minimal left ideal I of R, so I is an irreducible R-module. The result will follow by showing that any irreducible module M for R is isomorphic to I. Since R is simple, the representation of R given by M is faithful. Hence, since I ≠ 0, there exists an xM such that Ix ≠ 0. We have the homomorphism b images bx of I into M. Since Ix ≠ 0, the homomorphism is not 0. By Schur’s lemma, it is an isomorphism images

We shall require also the following

LEMMA 2.   Let M be a vector space over a division ring Δ, R = EndΔM the ring of linear transformations of M over Δ. Then the centralizer of R in the ring of endomorphisms End M of M as abelian group is the set of maps δ′ :u images δu, δ ∈ Δ. Moreover, δ images δ′ is an isomorphism of Δ onto this centralizer.

Proof.   Since M is a completely reducible Δ-module, the first statement is a special case of the density theorem. We can also give a simple direct proof of this result: Let d be an endomorphism of (M, + ,0) that centralizes EndΔM. Let x ≠ 0 in M. Then dxΔx. Otherwise, x and dx are linearly independent, and hence there is an l ∈ EndΔM such that Ix = 0 and l(dx) ≠ 0. Since l(dx) = d(lx) = 0, this is impossible. Hence for any x ≠ 0 we have a δxΔ such that dx = δxx. Let y be a second non-zero element of M and choose l ∈ EndΔM such that lx = y. Then δyy = dy = dlx = ldx = xx = δxlx = δxy. Hence δx = δy for all non-zero x and y, which implies that d has the form δ′ : u images δu. It is clear that δ images δ′ is an isomorphism. images

We can now prove the

ISOMORPHISM THEOREM FOR SIMPLE ARTINIAN RINGS.   Let Mibe a finite dimensional vector space over a division ring Δi, i = 1, 2, and let g be an isomorphism of R1 = EndΔ1M1 onto R2 = EndΔ2M2. Then there exists a semi-linear isomorphism s of Ml onto M2 such that

images

for all aR1.

Proof.   We consider M1 as R1 = EndΔiM1 module in the natural way and regard M2 as R1-module by defining ay = g(a)y for aR1, yM2. These two modules for the simple artinian ring R1 are irreducible. Hence they are isomorphic. Accordingly, we have a group isomorphism s of M1 onto M2 such that for any xM1, aR1 we have s(ax) = a(sx). Since ay = g(a)y, yM2, this gives the operator relation g(a) = sas−1 which is (25).

Since S is an isomorphism of M1 onto M2, the map b images sbs− 1, b ∈ End Ml is an isomorphism of the endomorphism ring End M1 of the abelian group M1 onto the endomorphism ring End M2 of the abelian group M2. Since this maps EndΔ1 M1 onto EndΔ2M2, its restriction to the centralizer of End Δ1 M1 in End M1 is an isomorphism of this centralizer onto the centralizer of EndΔ2M2 in End M2. By Lemma 2, the two centralizers are isomorphic respectively to Δ1 and Δ2 under the maps sending δi ∈ Δi into the endomorphism δ′ : ui images δiui. Hence we have an isomorphism σ of Δ1 onto Δ2 that is the composite of images with the inverse of δ2 images δ2. Thus we have 1s− 1 = (σδ1)′ and hence for any xM1 we have images. This shows that s is a σ-semi-linear isomorphism of M1 onto M2. images

An immediate consequence of this theorem is that if EndΔ1M1 and EndΔ2M2 are isomorphic, then M1 and M2 are semi-linearly isomorphic. Conversely, if s is a semi-linear isomorphism of M1 onto M2 then sas− 1 ∈ End,Δ2 M2 for any a ∈ EndΔ1M1 and it is immediate that a images sas− 1 is an isomorphism of EndΔ1M1 onto EndΔ2M2. We have seen that M1 and M2 are semi-linearly isomorphic if and only if M1 and M2 have the same dimensionality and Δ1 and Δ2 are isomorphic. This shows that a simple artinian ring is determined up to isomorphism by an integer n, the dimensionality of M, and the isomorphism class of a division ring Δ.

The isomorphism theorem for simple artinian rings also gives a determination of the automorphism group of such a ring. If we take R = EndΔM where M is a finite dimensional vector space over the division ring Δ, then the isomorphism theorem applied to EndΔM and EndΔM states that the automorphisms of R are the maps

images

where sG, the group of bijective semi-linear transformations of the vector space M. Denoting the map in (26) as Is, we have the epimorphism s images Is of G onto Aut R, the group of automorphisms of R. The kernel of this epimorphism is the set of bijective semi-linear transformations s such that sas−1 = a for every aR = EndΔM. By Lemma 2, any such map has the form δ′ : x images δx, δΔ. We note also that if δ ≠ 0, then δ′ is a semi-linear automorphism of M whose associated automorphism of Δ is the inner automorphism iδ: α images δαδ−1.This follows from

images

Thus the kernel of s images Is is the group Δ*′ of multiplications x images δx determined by the non-zero elements δ of Δ. Evidently, our results on automorphisms can be stated in the following way.

COROLLARY.   We have an exact sequence

images

where Δ* is the multiplicative group of non-zero elements of Δ, G is the multiplicative group of bijective semi-linear transformations of the vector space M, Aut R is the group of automorphisms of R = End ΔM, i is the map δ images δ′, and I is the map s images Is.

Of course, this implies that Aut R images G/Δ*′.

A reader who is familiar with the fundamental theorem of projective geometry (see BAI, pp. 468–473) should compare the results obtained here on isomorphisms and automorphisms of simple artinian rings with the fundamental theorem and its consequences. We remark that the present results could be derived from the geometric theorem. However, it would be somewhat roundabout to do so.

We shall derive next the main result on representations of artinian rings. This is the following

THEOREM 4.4.   Any module for a semi-simple artinian ring R is completely reducible and there is a 1−1 correspondence between the isomorphism classes of irreducible modules for R and the simple components of the ring. More precisely, if images where the Rj are the simple components and Ij is a minimal left ideal in Rj, then {I1,…,IS} is a set of representatives of the isomorphism classes of irreducible R-modules.

Proof.   Write each simple component as a sum of minimal left ideals of R. This gives R = ∑Ik where Ik is a minimal left ideal of R contained in a simple component. Now let M be an R-module and let x be any non-zero element of M. Then x = 1xRx = ∑Ikx. By Schur’s lemma, either Ikx = 0 or Ikx is an irreducible module isomorphic to Ik. Now images and this shows that M is a sum of irreducible submodules that are isomorphic to minimal left ideals of the simple components of R. Hence M is completely reducible and if M is irreducible, it is isomorphic to a minimal left ideal I of a simple component of R. To complete the proof, we need to show that if I is a minimal left ideal in Rj and I′ is a minimal left ideal in Rj′, then I images I′ as R-modules if and only if j = j′. Suppose first that j = j′. Then I and I′ are Rj-isomorphic by Lemma 1 on p. 205. Since RkI = 0 = RkI′ for kj′, it follows that I and I′ are R-isomorphic. Next assume jj′. Then RjI = I and RjI = 0; hence I and I′ are not isomorphic as R-modules. images

We also have a definitive result on irreducible representations for arbitrary (left) artinian rings. This is

THEOREM 4.5.   Let R be (left) artinian, N = rad R, and images where the imagesi are the simple components. For each j, images, let imagesj be a minimal left ideal of imagesj. Then imagesj is an irreducible module for R, images (as R-modules) if jj′, and any irreducible R-module is isomorphic to one of the imagesj.

Proof.   Any irreducible R-module M is annihilated by N, so it can be regarded as an irreducible images-module. Conversely, any irreducible images-module is an irreducible R-module. It is clear also that irreducible R-modules are isomorphic if and only if they are isomorphic as images-modules. Hence the theorem is an immediate consequence of Theorem 4.4. images

EXERCISES

        1. If z is a nilpotent element of a ring, then the index of nilpotency is the smallest t such that zt = 0, zt−1 ≠ 0. A similar definition applies to nilpotent ideals. Note that if N is a nilpotent ideal, then the indices of nilpotency of the elements of N are bounded. Use this to construct an example of a nil ideal that is not nilpotent.

        2. Prove that the center of any simple ring is a field.

        3. Show that if images where the Ri are simple, then the center C of R is images where Ci = CRi is the center of Ri. Hence C is a commutative semi-simple artinian ring. Note that the Ri are determined by the simple components Ci of C since Ri = ClR.

        4. Let R be left artinian, N = rad R, and assume that Nt = 0, Nt−1 ≠ 0. Show that R/N and Nl/Ni + 1, 1 ≤ it − 1, are completely reducible modules for R of finite dimensionality. Use this to prove that RR has a composition series and hence that R is left noetherian.

The next four exercises are designed to give conditions for anti-isomorphism of simple artinian rings and to determine the anti-isomorphisms and involutions of such rings.

        5. Let V be a vector space over a division ring Δ that has an anti-automorphism j:δ images images. A sesquilinear form on V (relative to j) is a map f of V × V into Δ such that

images

for x,yV, δΔ. Call f non-degenerate if f (z,y) = 0 for all y implies z = 0 and f (x, z) = 0 for all x implies z = 0. Show that if V is finite dimensional and f is non-degenerate, then for a linear transformation l in V over Δ there exists a (left) adjoint relative to f defined to be a linear transformation l′ such that f(lx, y) = f(x, I′y) for all x,y. Show that l images l′ is an anti-isomorphism and that this is an involution if f is hermitian in the sense that f(y,x) = fimages for all x,y or anti-hermitian: f (y,x) = − fimages.

        6. Let V and Δ be as in exercise 5 and let V* be the dual space homΔ(V,Δ). This is a right vector space over Δ, hence a left vector space over Δop. If l is a linear transformation of V over Δ, let l* be the transposed linear transformation in V*. Show that 1 images l* is an anti-isomorphism of EndΔV onto End V*Δ. Hence show that if Vi,i = 1, 2, is a finite dimensional vector space over a division ring Δi then EndΔ1V1 and EndΔ2V2 are anti-isomorphic if and only if Δ1 and Δ2 are antiisomorphic and V1 and V2 have the same dimensionality.

        7. Let V be a finite dimensional vector space over a division ring Δ and assume EndΔV has an anti-automorphism J. Note that J(l) images l* is an isomorphism of EndΔV onto EndΔopV* where V* is regarded in the usual way as left vector space over Δop (αx* = x*a). Hence there exists a semi-linear isomorphism s of V over Δ onto V* over Δop such that

images

for l ∈ EndΔ V. If x, yV, define

images

Show that f is a non-degenerate sesquilinear form on V corresponding to the antiautomorphism σ that is the isomorphism of Δ onto Δop associated with s and that J is the adjoint map relative to f.

        8. (Continuation of exercise 7.) Show that if J is an involution (J2 = 1), then either f is skew hermitian or f can be replaced by ρf, ρ ≠ 0, in Δ so that pf is hermitian and J is the adjoint map relative to ρf (Sketch: Note that for any vV, the map x images σ−lf (v, x) is a linear function, so there exists a v′V such that f(v,x) = images, xV. Show that for any uV the map l:x images f(x,u)v is in EndΔV and its adjoint relative to f is y images σ−1f(v,y)u = f(y,v′)u. Conclude from J2(l) = l that there exists a δ ≠ 0 in Δ such that images for all x,yV. Show that images = δ−1 and if δ ≠ − 1, then ρf for ρ = δ + 1 is hermitian and J is the adjoint map relative to ρf.)

        9. Note that the Wedderburn-Artin theorem for simple rings has the following matrix form: Any simple artinian ring is isomorphic to a matrix ring Mn(Δ′), Δ′ a division ring (Δ′ ≅ Δop if Δ is as in the statement of the theorem). Note that the isomorphism theorem implies that if Δ′1 and Δ′2 are division rings, then Mn1(Δ′1) ≅ Mn2(Δ′2) implies that n1 = n2 and Δ′1 ≅ Δ′2. Show that the isomorphism theorem implies also that the automorphisms of Mn(Δ′) have the form A images S(σA)S−1 where SMn(Δ′), σ is an automorphism of Δ, and σA = (σ(αij)) for A = (αij).

     10. Formulate the results of exercises 6–8 in matrix form.

4.5   STRUCTURE THEORY OF ALGEBRAS

In this section we shall extend the structure theory of rings that we developed in sections 1–4 to algebras over commutative rings. We shall then specialize to the case of finite dimensional algebras over a field.

We recall that an algebra A over a commutative ring K is a (left) K-module and a ring such that the module A and the ring A have the same additive group (A, + ,0) and the following relations connecting the K-action and ring multiplication hold:

images

for a, bA, kK (p. 44). We have the ring homomorphism

images

of K into A (p. 138). The image under ε is K1 = {Kl |kK}. This is contained in the center of A, and ka for kK, aA, coincides with (k1)a = a(k1). Conversely, given a ring R and a subring K of the center of R, R becomes a K- algebra by defining ka for kK, aR, to be the ring product ka. We remark also that rings are images-algebras in which na for nimages, aR, is the nth multiple of a. Hence the theory of rings is a special case of that of algebras.

The prime examples of K-algebras are the algebras EndKM where M is a K- module for the commutative ring K. Here kf for kK,f ∈ EndkM, is defined by (kf) (x) = kf(x) =f(kx), xM (p. 139). If A is an arbitrary K-algebra, then we define a representation of A to be a (K-algebra) homomorphism of A into an algebra EndKM. If ρ is a representation of A acting in the K-module M, then we define ax = ρ(a)x, aA, xM, to make M an A-module as well as a K- module. Since ρ(ka) = (a), kK, and ρ(a) is a K-endomorphism, we have

images

for kK, aA, xM. We now give the following

DEFINITION 4.3.   If A is an algebra over the commutative ring K, a (left) A- module is an abelian group written additively that is both a (left) K-module and a (left) A-module such that (30) holds for all kK, aA, xM.

If M is a module for A, we obtain a representation ρM of A by defining ρM(a) for aA to be the map x images ax. This is contained in EndkM, and a images ρM(a) is an algebra homomorphism. It follows from (30) that kx = (kl)x. Hence, if M and N are modules for the algebra A, then any homomorphism f of M into N regarded as modules for the ring A is also a K-homomorphism. It is clear also that any A-submodule is also a K-submodule and this gives rise to a factor module M/N for the algebra A in the sense of Definition 4.3. We remark also that if M is a ring A-module, then M is a K-module in which kx = (k1)x, and (30) holds; hence, this is an algebra A-module. The algebra A itself is an A-module in which the action of A is left multiplication and the action of K is the one given in the definition of A. The submodules are the left ideals. These are the same as the left ideals of A as ring. If kK and f ∈ homA(M,N) where M and N are modules for A, then kf is defined by (kf)(x) = kf(x), xM. This is contained in hom A(M, N), and this action of K on homA(M,N) is a module action. It is immediate that EndAM is an algebra over K.

We can now carry over the definitions and results on the radical, primitivity, and semi-primitivity to algebras. An algebra is called primitive if it has a faithful irreducible representation, semi-primitive if there are enough irreducible representations to distinguish elements. The radical is defined to be the intersection of the kernels of the irreducible representations of the algebra. The results we proved for rings carry over word for word to algebras. We have observed that if A is an algebra over K, then the left ideals of A as ring are also left ideals of A as algebra and, of course, the converse holds. Since the radical of A as algebra or as ring is the intersection of the maximal left ideals, it is clear that rad A is the same set with the same addition and multiplication whether A is regarded as ring or as algebra. Similarly, A is primitive (semiprimitive) as algebra if and only if it is primitive (semi-primitive) as ring.

The structure and representation theory of artinian rings carries over to algebras. In particular, the theory is applicable to algebras that are finite dimensional over fields, since any left ideal in such an algebra is a subspace and the descending chain condition for subspaces is satisfied. For the remainder of this section, except in some exercises, we consider this classical case: A a finite dimensional algebra over a field F. Then rad A is a nilpotent ideal containing every nil one-sided ideal of A. If rad A = 0, then A images where the Ai are ideals that are simple algebras. If A is simple, then A is isomorphic to the algebra of linear transformations in a finite dimensional vector space M over a division algebra Δ. Here M is any irreducible module for A and Δ is the division algebra EndAM given by Schur’s lemma. We recall also that M can be taken to be A/I where I is a maximal left ideal of A and Δ is anti-isomorphic to B/I where B is the idealizer of I (Proposition 4.7, p. 203). Hence Δ is a finite dimensional division algebra over F. Since EndΔM for M n-dimensional over Δ is anti-isomorphic to Mn(Δ) and Mn(Δ) is anti-isomorphic to Mnop), it is clear that A is a finite dimensional simple algebra over F if and only if it is isomorphic to an algebra Mn(Δ′), Δ′ a finite dimensional division algebra. These results were proved by Wedderburn in 1908.

In BAI, pp. 451–454, we determined the finite dimensional division algebras over algebraically closed, real closed, and finite fields. We showed that if F is algebraically closed, then F is the only finite dimensional division algebra over F and if F is real closed, then there are three possibilities for finite dimensional algebras over F: (1) F, (2) F(images), and (3) the quaternion algebra over F with base (1,i,j,k) such that i2 = − 1 = j2 and ij = k = −ji. If F is finite, then one has Wedderburn’s theorem that the finite dimensional algebras over F are fields. There is one of these for every finite dimensionality and any two of the same (finite) dimensionality are isomorphic (see BAI, pp. 287–290). Combining these results with the structure theorems, one obtains a complete determination of the finite dimensional semi-simple algebras over algebraically closed, real closed, or finite fields.

The results on representations of artinian rings are also applicable to finite dimensional algebras over fields. Thus if A over F is semi-simple, then any representation of A is completely reducible and we have a 1−1 correspondence between the isomorphism classes of irreducible modules and the simple components of the algebra.

One can apply the theory of algebras to the study of arbitrary sets of linear transformations in a finite dimensional vector space V over a field F. If S is such a set, we let Env S denote the subalgebra of EndF V generated by S. We call this the enveloping algebra of S. Evidently, Env S consists of the linear combinations of 1 and the products s1s2sr, siS, and the dimensionality dim Env Sn2 where n = dim V. The injection map into EndFV is a representation of Env S. The theory of algebras is applicable to the study of S via this representation. We shall now apply this method and the density theorem to obtain a classical theorem of the representation theory of groups. This is

BURNSIDE’S THEOREM.   Let G be a monoid of linear transformations in a finite dimensional vector space V over an algebraically closed field F that acts irreducibly on V in the sense that the only subspaces stabilized by G are V and 0. Then G contains a base for EndF V over F.

Proof.   The hypothesis is that G contains 1 and G is closed under multiplication. Then Env G = FG, the set of F-linear combinations of the elements of G. Then V is an irreducible module for A = Env G and A′ = End A V is a division algebra by Schur’s lemma. Evidently, A′ is the centralizer of A in EndF V, so A′ is a finite dimensional algebra over F. Since F is algebraically closed, we have A′ = F1. By the density theorem, EndA′V = A. Hence A = FG = EndFV. Evidently this implies that G contains a base for EndFV. images

If S is an arbitrary set of linear transformations in V over F that acts irreducibly, then Burnside’s theorem can be applied to the submonoid G generated by S. The result one obtains is that there exists a base for EndFV consisting of 1 and certain products of the elements of S.

EXERCISES

In these exercises A is an algebra over a field F. In exercises 1–5, A may be infinite dimensional; thereafter all vector spaces are finite dimensional.

        1. Let aA and let α1, …, αm be elements of F such that a − αl,l is invertible in A. Show that either a is algebraic over F or the elements (a − αi,1)−1, 1 ≤ im, are linearly independent over F.

        2. Let A be a division algebra with a countable base over an uncountable algebraically closed field F (e.g., F = images). Show that A = F.

        3. Let A be finitely generated over an uncountable algebraically closed field F and let M be an irreducible A-module. Show that EndAM = F.

        4. Show that the elements of rad A are either nilpotent or transcendental over F.

        5. Let A be finitely generated over an uncountable field F. Show that rad A is a nil ideal.

        6. Assume F contains n distinct nth roots of 1 and let ε be a primitive one of these. Let s and t be the linear transformations of the vector space V/F with base (x1, x2, …, xn) whose matrices relative to this base are respectively

images

Verify that

images

and that Env {s, t} = EndF V. Let u and v be linear transformations of a vector space W/F such that un = 1 = vn, uv = εvu. Apply the representation theory of simple algebras to show that dim W = r dim V and there exists a base for W over F such that the matrices of u and v relative to this base are respectively diag {σ, σ,…, σ} and diag {τ, τ, …, τ} .

        7. If a is a linear transformation in a vector space V over F with base (x1, …, xn) and axi = ∑αijxj, then the trace of a, tr a = ∑αii, is independent of the choice of the base (BAI, p. 196). Moreover, tv a is the sum of the characteristic roots of a. Let

images

Show that t(a,b) is a non-degenerate symmetric bilinear form on EndF V (BAI, p. 346).

        8. Let G, V, F be as in Burnside’s theorem. Suppose tr G = {tr a|aG} is a finite set. Show that G is finite and |G| ≤ |tr G|n2, n = dim V. (Hint: G contains a base (a1,…, an2) for EndFV by Burnside’s theorem. Consider the map a images (t(a, a1), t(a, a2,…, t(a, an2)) where t(a, b) = tr ab as in exercise 7.)

        9. A linear transformation is called unipotent if it has the form 1 + z where z is nilpotent. Let G be a monoid of unipotent linear transformations in a finite dimensional vector space V over an algebraically closed field F. Prove Kolchins theorem that V has a base relative to which the matrix of every aG has the form

images

(Hint: Let V = V1V2 ⊃ … ⊃ Vs + 1 = 0 be a composition series for V as A = FG-module and apply exercise 7 to the induced transformations in the irreducible modules Vi/Vi + 1.) Extend the theorem to arbitrary base fields F.

     10. (Burnside.) Let G be a group of linear transformations in a finite dimensional vector space V over an algebraically closed field F such that there exists an integer m not divisible by the characteristic of F such that am = 1 for all aG. Prove that G is finite. (Hint: If G acts irreducibly, the theorem follows from exercise 8. Otherwise, we have a subspace U ≠ 0, ≠ V stabilized by G. Use induction on the dimensionality to conclude that the induced groups of transformations in U and in V/U are finite. Let K1 and K2 be the kernels of the homomorphisms of G onto the induced groups in U and V/U. Then K1 and K2 and hence K1K2 are of finite index in G. Finally, show that K1K2 = 1.)

4.6   FINITE DIMENSIONAL CENTRAL SIMPLE ALGEBRAS

An algebra A over a field F is called central if its center C = F1 = {α1|αF}. If A is a finite dimensional simple algebra over A, then AMn(Δ) where Δ is a finite dimensional division algebra over Δ. The center C of A is isomorphic to the center of Δ and hence C is a field (cf. exercise 2, p. 209). Now A can be regarded as an algebra over C. When this is done, A becomes finite dimensional central simple over C. As we shall see in this section and the next, the class of finite dimensional central simple algebras has some remarkable properties. To facilitate their derivation it will be useful to collect first some simple results on tensor products of modules and algebras over fields that will be needed in the main part of our discussion.

We note first that since any F-module ( = vector space over F) is free, it is projective and thus flat (p. 154). Hence if we have an injective linear map i: V′ → V of F-modules, then i images 1 and 1 images i are injective maps of V′ images UV images U and of U images V′ into U images V respectively for any F-module U. If {xz|α ∈ I} is a base for V/F, then any element of U images V can be written in one and only one way as a finite sum images. The fact that it is such a sum is clear since any element of U images V has the form images. The uniqueness follows from the isomorphism of images (Proposition 3.3, p. 131), which implies that if images, then every images. Then the isomorphisms images imply that aα images xα = 0 if and only if az = 0. Similarly, if {yβ|β ∈ J} is a base for U, then every element of U images V can be written in one and only one way as images. It is clear also that if {xz|α ∈ I} is a base for V and {yβ|β} ∈ J} is a base for U, then {yβ images xz|βj,α ∈ I} is a base for U images V. Hence the dimensionality [U images V : F] = [U : F][V : F].

If A and B are algebras over the field F, then A images B contains the subalgebras 1 images B = {1 images b|bB} and A images 1 = {a images 1|aA}. These are isomorphic to B and A respectively under the maps b images 1 images b and a images a images 1. Hence we can identify B with the subalgebra 1 images B and A with the subalgebra A images 1 of A images B. If we write a for a images 1 and b for 1 images b, then we have ab = ba and A images B = AB = BA.

We shall now prove the following result, which in effect gives internal characterizations of tensor products of algebras over a field.

PROPOSITION 4.8.   If A and B are subalgebras of an algebra D over a field F, then DA imagesFB if the following conditions are satisfied:

            (1) ab = ba for all aA,bB.

            (2) There exists a base (xα) for A such that every element of D can be written in one and only one way as images.

If D is finite dimensional, then condition (2) can be replaced by

            (2′) D = AB and [D: F] = [A: F] [B: F].

Proof.   The first condition implies that we have an algebra homomorphism of A images B into D such that a images b images ab, aA, bB (p. 144). The second condition insures that this is bijective, hence, an isomorphism. It is immediate that (2′) images (2) if D is finite dimensional. images

As a first application of this criterion, we prove

PROPOSITION 4.9.   If B is an algebra over F, then Mn(B) ≅ Mn(F)imagesFB.

Proof.   This follows by applying the criterion to the subalgebras Mn(F) and B1 = {bl|bB} where

images

If {eij|1 ≤ i, jn} is the usual base of matrix units for Mn(F), then (b1)eij = eij(b1). Hence condition (1) holds. Moreover, (b1)eij is the matrix that has b in the (i,j)-position and 0’s elsewhere. This implies condition (2). images

Next we prove

PROPOSITION 4.10.   Mm(F)imagesMn(F) ≅ Mmn(F).

Proof.   This follows by applying the criterion to the subalgebras consisting of the matrices

images

where bMm(F), and of the matrices

images

respectively. We leave the details to the reader. images

If A is any algebra, we define Ae = A imagesFAop and we call this the enveloping algebra of A. If A is a subalgebra of another algebra B, then we have a natural module action of Ae on B defined by

images

where the aiA, biAop, yB. It is clear from the universal property of the tensor product that we have a homomorphism of Ae into B such that images Hence (31) is well-defined. Direct verification shows that it is a module action. In particular, we have a module action of Ae on A. The submodules of A as Ae-module are the ideals of A. Hence if A is simple, then A is Ae-irreducible. We recall that the endomorphisms of A regarded as left (right) A-module in the natural way are the right (left) multiplications x images xa (x images ax). It follows that EndAeA is the set of maps that are both left and right multiplications x images cx = xd. Putting x = 1, we obtain d = c; hence EndAeA is the set of maps x images cx, c in the center. If A is central, this is the set of maps x images (α1)x = αx, αF. We shall now apply this to prove our first main theorem on finite dimensional central simple algebras.

THEOREM 4.6.   If A is a finite dimensional central simple algebra over a field F, then images where n = dim A.

Proof.   We regard A as Ae-module as above. Then A is irreducible and EndAeA = F. Also A is finite dimensional over F. Hence by the density theorem Ae maps onto EndFA. Since both Ae and EndFA have dimensionality n2, we have an isomorphism of Ae onto EndFA. Since EndFAMn(F), the result follows. images

We consider next how a finite dimensional central simple algebra A sits in any algebra B (not necessarily finite dimensional) containing A as subalgebra. The result is

THEOREM 4.7.   Let A be a finite dimensional central simple subalgebra of an algebra B. Then BA imagesFC where C is the centralizer of A in B. The map I images AI is a bijection of the set of ideals of C with the set of ideals of B. Moreover, the center of B coincides with the center of C.

Proof.   We regard B as Ae-module as above. By the preceding theorem, Ae is simple. Hence B is a direct sum of Ae-irreducible modules all of which are isomorphic to A (since A is Ae-irreducible and any two irreducible Ae-modules are isomorphic). Now observe that the generator 1 of A as Ae-module satisfies (a images 1)1 = a1 = 1a = (l images a)1 and (a images 1)1 = 0 implies a = 0. Hence in any irreducible Ae-module, we can choose a c such that (a images 1)c = (1 images a)c and (a images 1)c = 0 implies a = 0. Applying this to B as Ae-module, we see that we can write B = imagesAcα where acα = cαa for all aA and acα = 0 implies a = 0. Then cαC and any element of B can be written in one and only one way as a finite sum ∑aαcα, aα ∈ A.If cCwe have c = ∑aαcα and ac = ca implies aαa = aaα. It follows that every aαF1 and c ∈ ∑Fcα. This implies that C = ∑Fcα and {cα} is a base for C. It is now clear from Proposition 4.8 that A imagesFCB.

Now let I be an ideal in C. Then Al is an ideal in B = AC. Moreover, AIC = I. For, if (x1 = l, x2,…, xn) is a base for A, then the canonical isomorphism of A imagesFC onto B (p. 144) implies that any element of B can be written in one and only one way as images. Then the elements of AI have the form ∑dixi, diI, and if such an element is in C, it also has the form c1x1, c1C. It follows that CAl is the set of elements d1x1 = d1I. Thus CAI = I, which implies that the map I images AI of the ideals of C into ideals of B is injective. To see that it is surjective, let I′ be any ideal of B. Then I′ is a submodule of B as Ae-module. Hence I′ = ∑Adβ where dβI = I′ ∩ C, which implies that I′ = AI Thus the map I images Al is also surjective.

It remains to prove that the center of B is the center of C. Since the center of B is contained in C, it is contained in the center of C. On the other hand, if c is contained in the center of C, it commutes with every element of B = AC and so it is in the center of B. images

It follows from the preceding theorem that B is simple if and only if C is simple and B is central if and only if C is central. The theorem can be applied to tensor products in which one of the factors is finite dimensional central simple. Thus let B = A imagesFC where A is finite dimensional central simple. We identify A and C with the corresponding subalgebras A images 1 and 1 images C of A images C. Now we claim that C is the centralizer of A in B. For, let {yβ} be a base for C over F. Then any element of B can be written in one and only one way as b = ∑aβyβ, aβA, and ab = ba for aA is equivalent to aaβ = aβa for every aβ. This implies that b commutes with every aA if and only if b = ∑αβyβ αβF, that is, if and only if bC. We can now apply Theorem 4.7 to obtain the

COROLLARY 1.   Let A be a finite dimensional central simple algebra over F, C an arbitrary algebra over F. Then the map I images A imagesFI is a bijection of the set of ideals of C with the set of ideals of B = A imagesFC. Moreover, the center of B coincides with the center of C (identified with the corresponding subalgebra of B).

An important special case of this result is

COROLLARY 2.   Let A be a finite dimensional central simple algebra over F, C any algebra over F. Then A imagesFC is simple if C is simple and A imagesFC is central if C is central.

Iteration of Corollary 2 has the following consequence.

COROLLARY 3.   The tensor product A1 imagesFA2 imagesFimagesFArof a finite number of finite dimensional central simple algebras is finite dimensional central simple.

Of particular interest are tensor products in which one of the factors is an extension field E of the base field F. If A is any algebra over F, then A imagesFE can be regarded as an algebra over E. This algebra over E is denoted as AE and is called the algebra obtained from A by extending the base field to E (see exercise 13, p. 148). If (xα) is a base for A over F, then by identifying A with A images 1 in AE we can say that () is a base for AE over E. Hence the dimensionality of A over F is the same as that of AE over E. It follows also that if K is an extension field of E, then (AE)KAK. If B is a second algebra over F with base (yβ), then A imagesFB has the base (xαyβ) ( = (xα images yβ)) over F and this is also a base for (A imagesFB)E over E and AE imagesEBE over E. It follows that (A imagesFB)EAE imagesEBE. (More general results are indicated in exercise 13, p. 148.)

If A is finite dimensional central simple over F, then Corollary 2 to Theorem 4.7 shows that AE is finite dimensional central simple over E. We recall also that Mn(F) is finite dimensional central simple over F. These are the simplest examples of central simple algebras. We now give

DEFINITION 4.4.   The matrix algebra Mn(F) is called a split central simple algebra over F. If A is finite dimensional central simple over F, then an extension field E of F is called a splitting field for A if AE is split over E (AEMn(E)for some n).

Evidently any extension field of F is a splitting field of Mn(F) and if E is a splitting field for A and B, then E is a splitting field for the central simple algebra A imagesFB. This follows since AEMn(E) and BEMm(E) imply (A imagesFB)EAE imagesEBEMn(E) imagesEMm(E) ≅ Mnm(E). We note also that if E is a splitting field for A, then it is a splitting field for Aop. For, (Aop)E = (AE)opMn(E)op, and since Mn(E) has an anti-automorphism, Mn(E)opMn(E). If AMr(B) where B is central simple over F then E is a splitting field for A if and only if it is a splitting field for B. We have AEMr(BE) since Mr(B) ≅ Mr(F) imagesFB and Mr(B)EMr(F)E imagesEBEMr(BE). Now BEMs(Δ) where Δ is a central division algebra over E. Then AEMrs(Δ). If E splits A (that is, is a splitting field for A) then AEMn(E) and hence Mn(E) ≅ Mrs(Δ) which implies that rs = n and Δ = E. Then BEMS(E) and E splits B. The converse is clear from Proposition 4.10. In particular, if AMr(Δ) where Δ is a central division algebra over F then E splits A if and only if E splits Δ. Thus it suffices to consider splitting fields for central division algebras. We remark finally that any extension field K of a splitting field is a splitting field. This is clear since (AE)KAK and Mn(E)KMn(K).

If E is an algebraically closed extension field of F, then the only finite dimensional simple algebras over E are the algebras Mn(E) (p. 213). Accordingly, any algebraically closed extension field E of F is a splitting field for every finite dimensional central simple algebra over F. We shall see later that every field can be imbedded in an algebraically closed field (section 8.1). When this result is available, it proves the existence of splitting fields for finite dimensional central simple algebras.

We shall now prove the existence of finite dimensional splitting fields for any finite dimensional central simple algebra A. Writing A = Mn(Δ) where Δ is a finite dimensional central division algebra, we shall show that any maximal subfield of Δ is a splitting field. This will follow from the following result (which we shall improve in Theorem 4.12).

THEOREM 4.8.   If Δ is a finite dimensional central division algebra over F, then a finite dimensional extension field E of F is a splitting field for Δ if and only if E is a subfield of an algebra A = Mr(Δ) such that the centralizer CA(E) = E.

Proof.   Suppose the condition holds: E is a subfield of A = Mr(Δ) such that CA(E) = E. We can identify A with EndΔ′V where V is an r-dimensional vector space over Δ′ = Δop. Then V is a Δ′ imagesFE module such that (d images e)x = dex = edx for d ∈ Δ′, eEA. Since Δ′ imagesFE is simple, the representation of Δ′ imagesFE in V is faithful and we can identify Δ′ imagesFE with the corresponding ring of endomorphisms in V. Since Δ′ imagesFE is finite dimensional simple over F, V is completely reducible as Δ′ imagesFE module and so we can apply the density theorem. Now EndΔV = A so EndΔ′ images EV is the centralizer of E in A. By hypothesis, this is E. Now V is finite dimensional over F since it is finite dimensional over Δ′ and Δ′ is finite dimensional over F. Thus V is finite dimensional over E. Hence, by the density theorem, Δ′ imagesFE is isomorphic to the complete algebra of linear transformations in V over E. If the dimensionality [V : E] = n, then Δ′ imagesFEMn(E). Thus E is a splitting field for Δ′ over F and hence of Δ over F.

Conversely, suppose Δ imagesFE images Mn(E). Then also Δ′ imagesFE = Mn(E). Let V be an irreducible module for Δ′ imagesFE. Since Δ′ imagesFEMn(E), V is an n-dimensional vector space over E and Δ′ imagesFE can be identified with EndEV. Also V is a vector space over Δ′ and if [V : Δ′] = r, then since E centralizes Δ′, E ⊂ EndΔ′V. The centralizer of E in EndΔ′ V is contained in the centralizer of Δ′ imagesE in EndEV. Since Δ′ imagesE = EndEV, this is E. Hence CEndΔ′V(E) = E and CMn(Δ)(E) = E. images

The existence of a maximal subfield of Δ is clear since [Δ : F] < ∞. We now have the following corollary, which proves the existence of finite dimensional splitting fields for any finite dimensional central simple algebra.

COROLLARY.   Let A be finite dimensional central simple over F and let A images Mr(Δ) where Δ is a division algebra. Then any maximal subfield E of Δ is a splitting field for A.

Proof. It suffices to show that E is a splitting field for Δ. Let E′ = CΔ(E). Then E′E and if E′E, we can choose cEimages E. The division algebra generated by E and c is commutative, hence this is a subfield of Δ properly containing E, contrary to the maximality of E. Thus CΔ(E) = E. Hence Theorem 4.8 with A = Δ shows that E is a splitting field. images

If E is a splitting field for A so that AEMn(E), then [A : F] = [AE : E] = [Mn(E) : E] = n2. Thus we see that the dimensionality of any central simple algebra is a square. The square root n of this dimensionality is called the degree of A.

We shall prove next an important theorem on extension of homomorphisms:

THEOREM 4.9.   Let A be a simple subalgebra of a finite dimensional central simple algebra B. Then any homomorphism of A into B can be extended to an inner automorphism of B.

Proof.   We form the algebra E = A imagesFBop, which is finite dimensional and simple. Any module for E is completely reducible and any two irreducible E-modules are isomorphic. It follows that any two E-modules of the same finite dimensionality over F are isomorphic. We now make B into an E-module in two ways. In the first, the action is (∑ai images bi)x = ∑aixbi and in the second it is images where f is the given homomorphism of A into B. Clearly these are module actions and the two modules are isomorphic. Hence there exists a bijective linear transformation l of B over F such that

images

for all aiA, x,biB. In particular, l(x)b = l(xb) for all x,bB. It follows that l has the form x images dx where d is an invertible element of B. We have also al(x) = l(f(a)x) for all aA, xB. Then adx = df(a)x. Putting x = 1 we obtain f(a) = d− lad. Hence f can be extended to the inner automorphism b images d−1bd in B. images

If we take A = F1 in Theorem 4.9 we obtain the important

COROLLARY (Skolem-Noether). Any automorphism of a finite dimensional central simple algebra is inner.

We prove next an important double centralizer theorem for central simple algebras.

THEOREM 4.10.   Let A be a semi-simple subalgebra of a finite dimensional central simple algebra B. Then the double centralizer CB(CB(A)) = A.

Proof. Consider the algebra B imagesBop and its subalgebra A imagesBop. If we identify B with B images 1, then it is easily seen as in the proof of Theorem 4.7 that B ∩ (A imagesBop) = A. Now regard B as B images Bop-module in the usual way. By the proof of Theorem 4.6, the algebra B images Bop is isomorphic to the algebra of linear transformations of B over F of the form x imagesbixb′i where bi,b′iB. Then the result that B ∩ (A imagesBop) = A implies that if a linear transformation of B is simultaneously of the form x imagesaixbi where aiA, biB, and of the form x images bx, bB, then b = aA. We now consider the algebra A images Bop. We claim that this is semi-simple. To see this we write A = A1 imagesimages AS where the Ai are simple ideals. Then A images Bop = (A1 images Bop) imagesimages(As images Bop) and Ai images Bop is an ideal in A images Bop. Since Ai is simple and Bop is finite dimensional central simple, Ai images Bop is simple by Corollary 2, p. 219. Thus A images Bop is a direct sum of ideals that are simple algebras. Hence A images Bop is semi-simple. Now regard B as A images Bop-module by restricting the action of Bimages Bop to the subalgebra A images Bop. Since A images Bop is semi-simple, B is completely reducible as E = A images Bop -module. Now consider E′ = EndEB. If we use the fact that Ihe endomorphisms of B as right B-module are the left multiplications, we see that E′ is the set of maps x images cx, cCB(A). Since B is finite dimensional over F and B is completely reducible as A images Bop-module, it follows from the density theorem that EndE′B is the set of maps images Now let c′CB(CB(A)). Then x images c′x commutes with x images cx for cCB(A) and with x images xb for bB. It follows that x images c′x is in EndE′B. Hence x images c′x also has the form x imagesaixbi, aiA, biB. We have seen that this implies that c′A. Thus we have proved that CB(CB(A)) ⊂ A. Since ACB(CB(A)) is clear, we have CB(CB(A)) = A. images

The foregoing result does not give us any information on the structures of the various algebras involved in the proof or on their relations. We shall now look at this question and for the sake of simplicity, we confine our attention to the most interesting case in which A is a simple subalgebra of the central simple algebra B. The extension to A semi-simple is readily made and will be indicated in the exercises.

Assume A simple. Then E = A images Bop is simple, so E images Mr(Δ) where Δ is a division algebra. Let d = [Δ : F]. Then we have

images

If {eij |1 ≤ i,jr} is the usual set of matrix units for Mr(Δ), then it is clear that Δ ≅ e11Mr(Δ)e11 which is just the set of matrices with (1, l)-entry in Δ and other entries 0. Also, Mr(Δ) is a direct sum of the r minimal left ideals Mr(Δ)eii and these are isomorphic modules for the algebra Mr(Δ). Each of these irreducible modules has dimensionality r2d/r = rd over F. Now consider B as E = A images Bop-module. Since E is simple, B as E-module is a direct sum of, say, s irreducible submodules all of which are isomorphic to minimal left ideals of E. Since EMr(Δ), the dimensionality over F of any irreducible E-module is rd. Hence we have

images

By (33) and (34) we see that s|r and

images

Now consider CB(A) ≅ E′ = EndEB. Since B is a direct sum of s irreducible E-modules and these are isomorphic to minimal left ideals Ee, e2 = e, of E, we can determine the structure of EndEB, hence of CB(A), by using the following

LEMMA.   Let e be a non-zero idempotent in an algebra E, I = Ee, I(s) the direct sum of s copies of I. Then EndEI(s) is isomorphic to Ms((eEe)op).

Proof.   By Lemma 2, p. 198, EndEI(s) consists of the maps (u1,…, us) images (v1,…, vs) where vi = ∑ja′ijuj and a′ij ∈ EndEI. This gives a map (a′ij) images η(a′ij) of Ms(EndEI) into EndEI(s). Direct verification shows that this is an isomorphism. Next we consider EndEI = EndEEe. As in the proof of Lemma (1), p. 180, EndEEe is the set of maps ae images aebe, a, bE. In this way we obtain a map of eEe into EndEEe such that ebe images η(ebe):ae images aebe. Direct verification shows that this is an anti-isomorphism. Hence EndEI ≅ (eEe)op and EndEI(s)Ms((eEe)op). images

Applying this lemma to B as E-module, we see that EndEBMs((e11Mr(Δ)e11)op) ≅ Msop). Hence CB(A) ≅ Msop) and

images

Then

images

We summarize these results in

THEOREM 4.11.   Let A be a simple subalgebra of a finite dimensional central simple algebra B. Then [B : f] = [A : F] [CB(A) : F] and A images BopMr(Δ), Δ a division algebra, and CB(A) ≅ MS(Aop). Moreover, s|r, [B : F] = rsd where d = [Δ : F] and [A : F] = r/s.

The relation [B : F] = [A : F][CB(A) : F] shows that if A is a field, then CB(A) = A images [B : F] = [A : F]2. This gives the following precise determination of the finite dimensional splitting fields of a central division algebra.

THEOREM 4.12.   A finite dimensional extension field E of F is a splitting field for a central division algebra Δ if and only if [E : F] = nr where n is the degree of Δ, and E is isomorphic to a subfield of Mr(Δ).

EXERCISES

        1. Let Δ be a finite dimensional central division algebra over F, E a maximal subfield of Δ. Show that [Δ : F] = [E : F]2.

        2. (Proof of Wedderburn’s theorem on finite division rings.) Let Δ be a finite division ring, F the center of Δ so Δ is a central division algebra over F. Show that if E1 and E2 are maximal subfields of Δ, then these are conjugates in the sense that there exists an s ∈ Δ such that E2 = sE1s−1. Conclude from this that the multiplicative group Δ* of non-zero elements of Δ is a union of conjugates of the subgroup E1*. Hence conclude that Δ = E1 = F is commutative (see exercise 7, p. 77 of BAI).

        3. (Proof of Frobenius’ theorem on real division algebras.) Let Δ be a finite dimensional division algebra over images. If Δ is commutative, then either Δ = images or Δ = images. If Δ is not commutative, its center is either images or images; images is ruled out since it is algebraically closed and so has no finite dimensional division algebras over it. Now suppose Δ is central over images and Δ images images. Then Δ contains an element i such that i2 = − 1 and since images = images(i) has an automorphism such that i imagesi, there exists a j in Δ such that ji = − ij. Show that the subalgebra generated by i and j is Hamilton’s quaternion algebra images. Use exercise 1 to prove that [Δ : images] =4 and that Δ = images.

        4. Let E be a cyclic field over F, [E : F] = n, Gal E/F, the Galois group of E/F, = images. Let (E,s,γ) be the subalgebra of Mn(E) generated by all of the diagonal matrices

images

bE, and the matrix

images

where γ is a non-zero element of F. Note that b images images is a monomorphism and identify b with images. Then we have the relations

images

and every element of (E,s,γ) can be written in one and only one way in the form

images

where the blE. Prove that (E, s, γ) is central simple and [E, s, γ) : F] = n2.

        5. Let E,F,G,s be as in exercise 4. Note that s ∈ EndFE and for any bE, x images bx is in EndFE. Identifying this with b, we have sb = s(b)s. Show that the algebra L of linear transformations in E over F generated by the multiplications bE and s is isomorphic to the algebra (E, s, 1) of exercise 4. Hence conclude that L = EndFE and (E, s, 1) ≅ Mn(F).

        6. Prove that (E, s, γ) ≅ Mn(F) if and only if γ is the norm of an element of E. (Hint: Use Theorem 4.9, p. 222). Hence show that if n = p is a prime and γ is not a norm in E, then (E, s, γ) is a division algebra.

        7. Let E = images(r) where r is a root of x3 + x2 − 2x − 1. Srhow that E/images is cyclic (see BAI, p. 236, exercise 1). Show that if γ is an integer not divisible by 8, then γ is not a norm of any element of E. Hence conclude from exercise 6 that (E, s, γ) is a division algebra.

        8. Let A be an algebra over F with base (u1,…, un) and let a images ρ(a) be the regular matrix representation determined by this base. Put N(a) = det ρ(a) (see BAI, p. 403). Show that A is a division algebra if and only if N(a) ≠ 0 for every a ≠ 0 in A. Use this criterion to show that if t is an indeterminate, then A is a division algebra if and only if AF(t) is a division algebra. Generalize to show that if t1,…,tr are indeterminates, then AF(t1,…, tl) is a division algebra if and only if A is a division algebra.

        9. Show that if Δ1 and Δ2 are finite dimensional division algebras over F and Δ1 is central, then Δ1 imagesF Δ2 = Mr(E) where E is a division algebra and r|[Δi : F], i = 1, 2. Hence show that Δ1 images Δ2 is a division algebra if ([Δ1 : F], [Δ2 : F]) = 1.

     10. If A is a subalgebra of an algebra B, a linear map D of A into B is called a derivation of A into B if

images

for aiA. Form B(2) and define

images

for aA, (x,y) ∈ B(2). Verify that B(2) is an A-module with this action. Show that B(2) is an A images Bop-module in which

images

     11. A derivation of B into B is called a derivation in B. If dB, the map x images [d, x] = dxxd is a derivation in B called the inner derivation determined by d. Prove that if A is a semi-simple subalgebra of a finite dimensional central simple algebra B, then any derivation of A into B can be extended to an inner derivation of B.

     12. Let A be a semi-simple subalgebra of a finite dimensional central simple algebra B. Show that CB(A) is semi-simple and investigate the relation between the structure of A, A images Bop and CB(A).

4.7   THE BRAUER GROUP

The results on tensor products of finite dimensional central simple algebras given in the previous section lead to the introduction of an important group that was first defined by R. Brauer in 1929. The elements of this group are similarity classes of finite dimensional central simple algebras. If A and B are such algebras, we say that A is similar to B (AB) if there exist positive integers m and n such that Mm(A) ≅ Mn(B) or, equivalently, Mm(F) images AMn(F) images B. The relation of similarity is evidently reflexive and symmetric. It is also transitive since if

images

and

images

then

images

Note that we have used associativity and commutativity of tensor multiplication as well as the formula Mn(F) images Mm(F) = Mmn(F), which was established in Proposition 4.10, p. 217. We now see that ∼ is an equivalence relation. Let [A] denote the similarity class of A, that is, the equivalence class of finite dimensional central simple algebras similar to A. Suppose AA′ and BB′ so we have positive integers m, m′, n, n′ such that A images Mn(F) ≅ A′ images Mn′(F) and B images Mm(F) ≅ B′ images Mm′(F). Then A images B images Mnm(F) ≅ A′ images B′ images Mn′m′,(F). Hence A imagesBA′ images B′. Thus we have a well-defined binary composition of similarity classes given by

images

Evidently this is associative and commutative. Moreover, the set of matrix algebras Mn(F) constitutes a similarity class 1 and this acts as unit since A ∼ A images Mn(F). We have seen in Theorem 4.6, p. 218, that A images Aop ∼ 1. Hence [A][Aop] = 1 = [Aop][A].

If A is finite dimensional central simple over F, we can write AMn(F) images Δ for Δ, a finite dimensional central division algebra. Conversely, if Δ is as indicated, then Mn(F) images Δ is finite dimensional central simple over F. We recall that the isomorphism theorem for simple artinian rings (p. 206) implies that if Mn(Δ) ≅ Mm(Δ′) for division algebras Δ and Δ′, then m = n and Δ ≅ Δ′. It follows that the division algebra Δ in the formula AMn(F) images Δ is determined up to isomorphism by A. Hence a similarity class [A] contains a single isomorphism class of finite dimensional central division algebras and distinct similarity classes are associated with non-isomorphic division algebras. Now the class of subalgebras of the matrix algebras Mn(F), n = 1, 2, 3,… is a set and every algebra over F is isomorphic to a member of this set. Hence the isomorphism classes of algebras over F constitute a set. This set has as subset the isomorphism classes of central division algebras over F. Hence the similarity classes of central simple algebras over F is a set. Our results show that this set is a group under the composition defined by (44). This group is denoted as Br(F) and is called the Brauer group of the field F. It is clear that the determination of Br(F) for a field F is equivalent to a complete classification of the finite dimensional central division algebras over F.

If F is algebraically closed or is a finite field, then Br(F) = 1. If F = images, then Br(F) is a cyclic group of order two by Frobenius’ theorem. We shall determine Br(F) for F a p-adic field in Chapter 9 (section 9.14). One of the most important achievements of algebra and number theory in the 1930’s was the determination of Br(images) and, more generally, Br(F) for F a number field (that is, a finite dimensional extension field of images). This involves some deep arithmetic results.

Let E be an extension field of F. If A is finite dimensional central simple over F, then AE is finite dimensional central simple over E. We have (A imagesFB)EAE images EBE and Mn(F)EMn(E). This implies that we have a homomorphism of Br(F) into Br(E) sending [A] into [AE]. The kernel, which we shall denote as Br(E/F) is the set of classes [A] such that AE ∼ 1 (1 for F is customary here), that is, the [A] such that A is split by E. We shall consider this group in Chapter 8 for the case in which E is a finite dimensional Galois extension field of F.

EXERCISE

        1. Use Theorem 4.11 (p. 224) to show that if A is finite dimensional central simple over F and E is a subfield of A/F, then CA(E) ∼ AE (in Br(E)).

4.8   CLIFFORD ALGEBRAS

In this section we apply some of the results on central simple algebras to the study of certain algebras—the Clifford algebras—defined by quadratic forms. These algebras play an important role in the study of quadratic forms and orthogonal groups. The results on these matters that we require can be found in BAI, Chapter 6.

Let V be a finite dimensional vector space over a field F equipped with a quadratic form Q. We recall the definition: Q is a map of V into the base field F such that

            1. Q(αx) = α2Q(x), αF, xV.

            2. B(x, y) ≡ Q(x + y)−Q(x)−Q(y) is bilinear.

Evidently, B(x, y) is symmetric, that is, B(y,x) = B(x, y) and B(x, x) = 2Q(x). We use Q to define an algebra in the following manner.

DEFINITION 4.5.   Let V be a vector space over a field F, Q a quadratic form on V. Let images be the tensor algebra defined by V (p. 140) and let KQ be the ideal in T(V) generated by all of the elements of the form

images

Then we define the Clifford algebra of the quadratic form Q to be the algebra

images

If aT(V), we write images and we have the map i : x images images of V into C(V,Q). Since V generates T(V), i(V) generates the Clifford algebra. We have images. Moreover, we claim that we have the following universality of the map i: If f is a linear map of V into an algebra A such that

images

then there exists a unique algebra homomorphism g of C(V,Q) into A such that

images

is commutative. To see this, we recall the basic property of the tensor algebra that any linear map f of V into an algebra has a unique extension to an algebra homomorphism f′ of T(V) into A (see p. 140). Now for the given f, the kernel, ker f′, contains every element x images xQ(x)1, xV, since images. Hence the ideal KQker f′ and so we have the induced homomorphism images. In particular, g(images) = f′ (x) = f(x), which is the commutativity of (46). The uniqueness of g is clear since the images generate C(V,Q).

The ideal KQ defining C(V,Q) contains every

images

Since images we see that

images

Equivalently, we have the relations

images

in C(V,Q) as well as images2 = Q(x)1. We can use these to prove

LEMMA 1.   If the elements ul, u2,…, un span the vector space V over F, then the elements

images

span the vector space C(V,Q) over F.

Proof.   Since T(V) is generated by V and the ui span V, it is clear that the ui generate T(V), so every element of T(V) is a linear combination of 1 and monomials in the ui of positive degree. We now call 1 and the monomials images standard. Let S be the set of these. We proceed to prove by induction on i and on the degree of uS that uiu is congruent modulo the ideal KQ to a linear combination of standard monomials of degree ≤deg u + 1. This is clear if deg u = 0. Now suppose u = ui1ui2uir, r ≥ 1. If i = 1, the result is clear if i1 > 1 and if il = 1, then images (mod KQ). Now let i > 1. If ii1 the result follows as in the case i = 1. Hence assume i > i1. Then, by (47)

images

By the degree induction, uiui2uir is congruent modulo KQ to a linear combination of standard monomials of degree ≤r. Then induction on the subscript i implies that ui1uiui2uir and hence uiui1ui2uir is congruent modulo KQ to a linear combination of standard monomials of degree ≤ r + 1. The result we have proved implies that if C′ is the subspace of C(V, Q) spanned by the elements images then images. This implies that C′C′C′ so C′ is a subalgebra of C(V,Q). Since C′ contains images = {images| xV} and images generates C(V, Q), we have C′ = C(V, Q). images

Evidently, the lemma implies that if dim V = n, then dim C(V, Q) ≤ 2n, which is the number of standard monomials in the ui.

For the sake of simplicity, we assume in the remainder of the text of this section that char F ≠ 2. The extension of the results to the characteristic two case will be indicated in the exercises. We assume first that B is non-degenerate and we separate off the lowest dimensional cases n = 1, 2.

n = 1. We have dim C(V, Q) ≤ 2 and if V = Fu, then u−2 = Q(u)1 and Q(u) ≠ 0, since B is non-degenerate. Let A = F[t]/(t2Q(u)1) where t is an indeterminate. Then A has the base (1, images) where images = t + (t2Q(u) 1). Then images2 = Q(u) 1, which implies that the linear map f of V into A such that f(u) = images satisfies f(x)2 = Q(x)images. Hence we have a homomorphism of C(V, Q) into A such that images. This is surjective. Since dim A = 2 and dim C(V,Q) ≤ 2, we have dim C(V,Q) = 2 and our homomorphism is an isomorphism. If Q(u) is not a square in F, then t2Q(u) 1 is irreducible in F[t] and C(V, Q) is a field. If Q(u) = β2, βF, then images satisfies

images

Hence e = −(2β)−1e′ is an idempotent ≠ 0, 1 in C(V,Q). Then C(V, Q) = Fe images F(1 − e), a direct sum of two copies of F. Hence we have the following lemma.

LEMMA 2.   If B(x, y) ,is non-degenerate and n = 1, then dim C(V, Q) = 2 and if V = Fu, then C(V, Q) is afield or a direct sum of two copies of F according as Q(u) is not or is a square in F.

n = 2. Choose an orthogonal base (u,v) for V. Then Q(u)Q(v) ≠ 0. Let C′ be the Clifford algebra of Fu relative to the restriction Q′ of Q to Fu. Then C′ has a base (1, images) where images2 = Q(u) 1. Now consider the matrices

images

in M2(C′). We have

images

so u′v′ + v′u′ = 0. Also u2 = Q(u) 1, v′2 = Q(v) 1. It follows that A = F1 + Fu′ + Fv′ + Fu′ v′ is a subalgebra of M2(C′). It is clear from the form of the matrices that 1, u′ v′, u′v′ are linearly independent. Hence dim A = 4. The relations on u′ and v′ imply that the linear map f of V into A such that u images u′, v images v′ satisfies f(x)2 = Q(x) 1, xV. Hence we have a homomorphism g of C(V, Q) into A such that g(images) = u′, g(images) = v′. Then g is surjective and since dim C(V, Q)≤4 and dim A = 4, it follows that g is an isomorphism and dim C(V, Q) = 4.

An algebra A over a field F of characteristic ≠2 is called a (generalized) quaternion algebra if A is generated by two elements i and j such that

images

where α, β ∈ F. We denote this algebra as (α, β) or (α, β)/F if we need to call attention to the base field F. We now prove

LEMMA 3.   Any quaternion algebra is four-dimensional central simple over F.

Proof.   The relations (49) imply that any element of A is a linear combination of the elements 1, i, j, ij. Now suppose λl + μi + vj + ρij = 0 for λ, μ, v, ρF. If we multiply this relation on the left by i and on the right by i−1 = α−1i, we obtain λ1 + μivjρij = 0. Then λ1 + μi = 0 and vj + ρij = 0. Multiplication of these on the left by j and on the right by j−1 = β−1 j then gives λ1 − μi = 0 and vjρij = 0. Then λ1 = μi = vj = ρij = 0 and λ = μ = v = ρ = 0. Hence the elements 1, i, j, ij are linearly independent and so these constitute a base for A. Hence dim.A = 4. Now let I be an ideal ≠ A in A and let images = 1 + I, images = i + I, images = j + I. Then images. Hence A/I is a quaternion algebra and so dim A/I = 4. Then I = 0 and hence A is simple. It remains to show that the center of A is F1. Let imagescenter of A. Then ci = ic implies that v = ρ = 0. The fact that λ1 + μi commutes with j implies that μ = 0. Hence c = λ1 and the center is F1. images

The result we obtained before on C(V, Q) in the case n = 2 can now be stated in the following way.

LEMMA 4.   If B(x, y) is non-degenerate and n = 2, then C(V, Q) is a quaternion algebra.

Proof.   We had an isomorphism of C(V, Q) with the algebra A generated by u′ and v′ such that u′2 = Q(u) 1 ≠ 0, v′2 = Q(v) 1 ≠ 0, and u′v′ = − v′u′. Then A and, hence, C(V, Q) are quaternion algebras. images

We recall that if B(x, y) is a symmetric bilinear form on a vector space V and (u1, u2,…, un) is a base for V, then δ = det (B(ui, uj)) is called a discriminant of B. A change of base replaces δ by δβ2 where β ≠ 0 is the determinant of the matrix giving the change of base. B is non-degenerate if and only if δ ≠ 0. We recall also that if U is a subspace of V on which the restriction of B to U is non-degenerate, then V = U images Uimages. Moreover, the restriction of B to Uimages is nondegenerate.

We shall now prove the following factorization property.

LEMMA 5.   Let B(x, y) be non-degenerate and dim V ≥ 3. Let U be a twodimensional subspace of V on which the restriction of B to U is non-degenerate. Write V = U images Uimages and let Q′ and Q″ denote the restrictions of Q to U and Uimages respectively. Then

images

where δ′ is a discriminant of the restriction of B to U.

Proof.   We shall prove (50) by using two universal map properties to produce inverse isomorphisms between C(V, Q) and the right-hand side of (50). We denote the canonical maps of U into C(U, Q′) and of Uimages into C(Uimages, − δ′Q″) as i′ : y images y′ and i′ : z images z″ respectively. Let (u, v) be an orthogonal base for U and put images (in C(V, Q)). We have images, so images where δ′ is the discriminant of the restriction of B to U defined by the base (u, v)). Also if yU and zUimages, then images hence

images

Since y−2 = Q(y) 1 and images, the universal map property of Clifford algebras implies that we have homomorphisms of C(U, Q′) and C(Uimages, − δ′Q″) into C(V, Q) sending images respectively. The elements images and images generate the images under the homomorphisms and, by (51), these elements commute. Hence the images under our homomorphisms centralize each other, so by the universal map property of tensor products, we have a homomorphism h of images into C(V, Q) such that

images

Now consider the element d′ = 2u′v′ in C(U, Q′). The calculations made before show that d′y′ = −y′d′ and d′2 = −δ′ 1, so d′ is invertible in C(U, Q′). Next consider the element images. We have

images

Hence by the universal map property of C(V, Q) we have a homomorphism g of C(V, Q) into images such that

images

Checking on generators, by (52) and (53), we see that gh = 1 on images and hg = 1 on C(V, Q). Hence C(V, Q) ≅ images. images

We are now ready to prove the main theorem on the structure of Clifford algebras.

THEOREM 4.13.   Let Q be a quadratic form of an n-dimensional vector space over afield F of characteristic ≠2. Then

        (1) dim C(V, Q) = 2n and if (u1, u2,…, un) is a base for V, the elements

images

constitute a base for C(V, Q) over F.

        (2) The canonical map i : x images images of V into C(V, Q) is injective.

        (3) If the bilinear form B associated with Q is non-degenerate, then C(V, Q) is central simple if n is even and if n = 2v + 1, vimages, then C(V, Q) is either simple with a two-dimensional field as center or is a direct sum of two isomorphic central simple algebras according as (−1)′ 2δ, δ a discriminant of B, is not or is a square in F.

Proof.   We note first that the second statement in (1) and statement (2) are consequences of the dimensionality relation dim C(V, Q) = 2n. For, by Lemma 1, the elements in (54) span C(V, Q). Since the number of these elements is ≤ 2n, if dim C(V, Q) = 2n, their number is 2n (that is, they are distinct) and they form a base. This implies also that the imagesi, 1 ≤ in, are linearly independent and hence the linear map i : x images images is injective.

Next we prove the dimensionality relation and (3) in the non-degenerate case. If n = 1, the results follow from Lemma 2 and if n = 2, they follow from Lemmas 3 and 4. Now assume n > 2. Then we can pick a two-dimensional subspace U on which the restriction of B is non-degenerate. Then V = U images Uimages and the restriction of B to Uimages is non-degenerate. By Lemma 5, C(V, Q) ≅ C(U, Q′) images C(Uimages, − δ′Q″) where δ′ is a discriminant of the restriction of B to U. Moreover, C(U, Q′) is a quaternion algebra, hence, is fourdimensional central simple. Using induction on the dimensionality, we may assume the results for the quadratic form −δ′Q″ on Uimages. Then dim C(Uimages, − δ′Q″) = 2n − 2 and this algebra is central simple if n − 2 is even. Hence dim C(V, Q) = 222n − 2 = 2n, and by Corollary 2 to Theorem 4.7, p. 219, this algebra is central simple if n is even. If n = 2v + 1, n − 2 = 2(v − 1) + 1 and the induction hypothesis implies that if δ″ is a discriminant of the restriction of B to Uimages, then C(Uimages, − δ′Q″) is simple with two-dimensional center or is a direct sum of two isomorphic central simple algebras according as (− l)v − 1 2( − δ′)n − 2 δ″ is not or is a square in F. Accordingly, C(V, Q) is simple with two-dimensional center or is a direct sum of two isomorphic central simple algebras according as (− l)v − 1 2( − δ)n − 2 δ″ is not or is a square. Now

images

and since n − 3 is even this is a square if and only if (− 1)v is a square. This proves (3).

It remains to prove the dimensionality formula in the degenerate case. For this purpose, we shall imbed V in a finite dimensional space W with a quadratic form that has a non-degenerate associated bilinear form and is an extension of Q. To do this we write V = Vimages images U for some subspace U. Then the restriction of B to U is non-degenerate. Now put W = Vimages images U images (Vimages)* where (Vimages)* is the space of linear functions on Vimages. Let x = z + y + f where zVimages ,yU, and f ∈ (Vimages)* and define

images

It is readily seen that images is a quadratic form on W whose associated symmetric bilinear form images is non-degenerate. Let x images images be the canonical map of W into C(W, images). It follows from the universal property of C(V, Q) that we have a homomorphism of C(V, Q) into C(W, images) such that images images images for xV. Let (u1,…, un, un + 1,…, uq) be a base for W such that (u1,…, un) is a base for V. Since images is non-degenerate, the elements images, are distinct and linearly independent. Then this holds also for the elements images. Since the homomorphism of C(V, Q) into C(W, images) maps imagesi1imagesir into imagesi1imagesir, the elements imagesi1imagesir, i1 < … < ir, 1 ≤ rn, are linearly independent. Hence dim C(V, Q) = 2n. images

Since the map i : x images images of V into C(V, Q) is injective, we can identify V with the corresponding subspace of C(V, Q). Hence from now on we assume VC(V, Q). If U is a subspace of V, then the subalgebra of C(V, Q) generated by U can be identified with C(U, Q′) where Q′ is the restriction of Q to U. This is clear from the last part of the proof of Theorem 4.13. For, if (u1uir) is a base for U over F, then the argument shows that the elements 1, ui1, … ,uir, i1 < i2 < … < ir, 1 ≤ rm, are linearly independent and this implies that the canonical homomorphism of C(U, Q′) into C(V, Q) is a monomorphism.

It is clear from the definitions that if Q = 0, then C(V, Q) is the exterior algebra E(V) defined by V (p. 141). The results (1) and (2) of Theorem 4.13 give another proof of properties of E(V) that were derived in BAI, pp. 411–414.

We remark finally that the proof of statement (3) in Theorem 4.13 yields a stronger result than we stated in this theorem. We state this as the following