Chapter 11

Finiteness Conditions for Rings and Modules

A scientist worthy of the name, above all a mathematician, experiences in his work the same impression as an artist; his pleasure is as great and of the same nature.

—Henri Poincaré

The field img of complex numbers is a two-dimensional vector space over img A ring that is also a vector space over a field F is called an algebra over F. In the nineteenth century many attempts were made to describe the division algebras that are finite dimensional over img and in particular those (like img) that are fields. Certainly the three-dimensional examples would have had applications to physics, but after looking in vain for such an algebra, the first success came in 1843 when W.R. Hamilton discovered the ring img of quaternions, a four-dimensional algebra that, surprisingly, was not commutative. It was not until 1878 that G. Frobenius showed that there is no three-dimensional example, and that the only possible associative examples are img, and img Meanwhile, G. Grassmann described many examples that were rings but not necessarily division rings, of which the matrix algebras, constructed in 1858 by A. Cayley, were an important special case. The next event in the development of the theory came in 1907 when J.H.M. Wedderburn gave the first characterization of the simple finite dimensional algebras. Finally, in 1927, E. Artin extended Wedderburn's theorem to a result about rings by eliminating the dependence on the fact that the ring is finite dimensional as an algebra, replacing it with finiteness conditions on the set of left ideals. These seminal results mark the beginning of the theory of noncommutative ring theory, and we prove them in this chapter.

11.1 Wedderburn's Theorem

If F is a field, a ring R is called an F-algebra if it is a vector space over F and t(ab) = a(tb) for all t imgF and all a, b img R. Hence, img is a two-dimensional img-algebra and the ring Mn(F) of all n × n matrices over F is an n2-dimensional F-algebra. The ring img of real quaternions is a four-dimensional algebra over img that has the distinction of being a division ring (every nonzero element is a unit). Wedderburn's theorem asserts that if R is a simple, finite dimensional algebra then RMn(D) for some n ≥ 1 and some division ring D. We are going to prove an extension of this theorem that removes the restriction that R is an algebra. The first task is to find an appropriate “ finiteness condition” on R to replace the finite dimensional requirement.

Finiteness Conditions

If R is a ring, recall (Section 7.1) that a left R-module is an additive abelian group with a left R-action rx img M, r img R, x img M, such that the axioms for a vector space are satisfied. In the nineteenth and early twentieth centuries, most of the modules studied were finite dimensional vector spaces. Emmy Noether, and later Emil Artin, realized that the right way to extend this finite dimensional condition was to use finiteness conditions on the set of submodules. Lemma 1 below identifies the most important of these.

If img is a nonempty set of submodules of a module M, then img is said to be maximal in img if img implies that K = N. Similarly N is called minimal in img if img implies that K = N. For example, the maximal ideals in a ring are the maximal members of img is an ideal of R and AR}.

Lemma 1. If M is a module, consider the following conditions where the Ki denote submodules of M. Then (ACC) img (MAX) and (DCC) img (MIN).

(ACC) If K1K2img then Kn = Kn+1 =img for some n.

(MAX) Every nonempty set of submodules of M has a maximal member.

(DCC) If K1K2img then Kn = Kn+1 =img for some n.

(MIN) Every nonempty set of submodules of M has a minimal member.

Proof. If (MAX) holds then (ACC) holds where Kn is any maximal member of {Ki img i ≥ 1}. Conversely, suppose that img is a nonempty set of submodules of M that has no maximal member. Choose img Since K1 is not maximal in img there exists img such that K1K2. But K2 is also not maximal in img so we obtain K1K2K3 for some img This process continues indefinitely 129 to violate (ACC). This proves that (ACC) img (MAX); the proof that (DCC) img (MIN) is analogous. img

The notations ACC and DCC refer to the ascending and descending chain conditions, on submodules, respectively. A module M is called noetherian if it has the ACC, and M is called artinian if it has the DCC. If R is an F-algebra with unity 1 then any left module RM becomes an F-space via the action t x = (t1)x for all t img F and x img M. Hence every submodule of M is a subspace, so finite dimensional modules are both noetherian and artinian.130

Example 1. Regarded as a module over itself, img is noetherian but not artinian.

Proof. img is not artinian as img shows. For the converse, suppose K1K2img are subgroups of img Then K = ∪ Ki is a subgroup and so img for some img by Theorem 7 §2.4. But then k img Kn for some n and it follows easily that Kn = Kn+1 = img = K. Hence img is noetherian. img

If img is a prime, let img. This is an additive subgroup of img and one verifies that the groups img are the only subgroups of X containing img (Exercise 9). The factor group img is called the Prüfer p-group, and this shows that the only subgroups of img are

img,

where img for each k. Furthermore, o(xk) = pk and pxk+1 = xk hold for each k. Hence img is an infinite group, but every proper subgroup is finite. Clearly

Example 2. The Prüfer group img is artinian but not noetherian as a img-module.

The following basic properties of the ACC and the DCC will be needed.

Lemma 2. If NM are modules then M is noetherian (artinian) if and only if the same is true of both N and M/N.

Proof. We prove the noetherian case; the other is analogous. If M has the ACC then N has the ACC because submodules of N are submodules of M. On the other hand, every submodule X of M/N has the form X = K/N for some submodule K of M containing N by Theorem 5 §8.1. Hence every ascending chain in M/N takes the form K1/NK2/Nimg where K1K2img in M. It follows that M/N has the ACC.

Conversely, let K1K2img be a chain from M. If both N and M/N have the ACC then the chains NK1NK2img and img both terminate, so there exists n ≥ 1 such that NKn = NKn+1 = img and img (whence N + Kn = N + Kn+1 = img). Since KiKi+1 for each i, we have Ki+1 ∩ (Ki + N) = Ki + (Ki+1N) by the modular law (Theorem 2 §7.1). So if in we have

img

This proves that M is noetherian. img

Corollary. A sum M = M1 + img + Mn of modules is artinian (noetherian) if and only if the same is true of each Mi.

Proof. We prove only the artinian case, the other being analogous. If M is artinian, so are its submodules Mi by Lemma 2. Conversely, if n ≥ 2, assume inductively that K = M2 + img + Mn is artinian. By Corollary 2 of Theorem 1 §7.1, it follows that M/K = (M1 + K)/KM1/(M1K) is also artinian by Lemma 2, being an image of M1. Hence, M is artinian, again by Lemma 2, as required. img

Let M1, . . ., Mn be modules. The external direct sum M = M1imgMn is artinian (noetherian) if and only if the same is true of each Mi by the Corollary because M is a sum of submodules img isomorphic to the Mi.

Because the img-action on an abelian group is naturally written on the left, we discussed only left R-modules in Chapter 7. However, an additive abelian group M is called a right R-module (written MR) if R acts on the right: That is, for any r img R and x img M, an element xr img M is defined such that

(x + y)r = xr + yr,  x(r + s) = xr + xs,  (xr)s = x(rs),  and x1 = x

hold for all x, y img M and all r, s img R.131 With this, the definition of submodules and homomorphisms of right modules are analogous to those for left modules. Moreover, the analogues of theorems about left modules in Section 7.1 go through verbatim for right modules. Note that the distinction does not matter for a commutative ring R because a left R -module M becomes a right module if we define the action by x · r = rx for all r img R and x img M.

A ring R is called left artinian (left noetherian) if RR is artinian (noetherian), with similar definitions on the right. The submodules of RR (RR) are the left (right) ideals.

Example 3. In each case consider the subring R of img

(1) img is left noetherian but not right noetherian.

(2) img is left artinian but not right artinian.

Solution.

(1) If img then A is an ideal of R and img is noetherian as a ring by the Corollary to Lemma 2 (verify). Hence, R/A is noetherian as a left R-module (the left R-submodules of R/A are just the left ideals of the ring R/A). Since RA is also noetherian (verify), it follows by Lemma 2 that R is left noetherian. But the sequence img of right ideals shows that R is not right noetherian. This proves (1).

(2) Observe that img is not finite dimensional as a img -space (otherwise img would be countable contradicting Cantor's theorem from set theory). Hence, there exist img-subspaces X1X2img in img But then img are right ideals of R, proving that R is not right artinian. The proof that R is left artinian is analogous to the argument in (1).

We note in passing that, despite Example 3, every left artinian ring is automatically left noetherian; this result is called the Hopkins–Levitzky theorem, proved independently in 1939 by Charles Hopkins and Jacob Levitzki.

A left R-module RM is said to be simple if it is nonzero and satisfies the following equivalent conditions:

1. The only submodules of M are 0 and M.

2. Rx = M for all 0 ≠ x img M.

Thus, the simple img-modules are the cyclic groups img where img is a prime. Every simple module RM is principal, that is, M = Rx for some x img M. If R = D is a division ring the converse holds: If DM = Dx is simple, then MDD via dxd. In particular, the simple modules over a field are the one-dimensional vector spaces.

As for groups, a series M = M0M1imgMn = 0 of submodules of a module M is called a composition series of length n for M if all the factors Mi/Mi+1 are simple modules (see Section 9.1).

Theorem 1. Let M ≠ 0 be a module.

1 M has a composition series if and only if it is both noetherian and artinian.

2. Jordan–Hölder Theorem. Any two composition series for M have the same length, and the factors can be paired so that corresponding factors are isomorphic.

Proof.

(1) If M = M0M1imgMn = 0 is a composition series then Mn−1 and Mn−2/Mn−1 are both simple, so Mn−2 is noetherian and artinian by Lemma 2. Then the same is true of Mn−3 because Mn−3/Mn−2 is simple. Continuing we see that every Mk (including M0 = M) is noetherian and artinian.

Conversely, let M be noetherian and artinian. Since M is artinian, let K1M be a simple submodule. If K1M, choose K2 minimal in the set {K img KK1}. Then K2K1 ⊃ 0 and K2/K1 is simple. If K2M, let K3K2K1 ⊃ 0 where K3/K2 simple. Since M is noetherian, this process cannot continue indefinitely, so some Kn = M and we have created a composition series.

(2) The analogue of the proof of the Jordan-Hölder theorem for arbitrary groups (Theorem 1 §9.1) goes through. img

The length n of any composition series for M is called the composition length of M and denoted length (M). Note that Lemma 2 and Theorem 1 combine to show that, if KM are modules, then M has a composition series if and only if both K and M/K have composition series. Moreover, in this case the proof of Theorem 2 §9.1 goes through to show that

length (M) = length (K) + length (M/K), (*)

and that the composition factors of M are exactly those of K and M/K.

The finitely generated vector spaces over a field are called finite dimensional, and an Theorem 6 §6.1 shows that they all have a finite basis. The same is true for any division ring, but we give a different proof using the Jordan–Hölder theorem.

Corollary. Let D be a division ring and let DM be a module. Then

1. M is finitely generated if and only if it has a finite basis.

2. Any two bases of M contain the same number of elements, say n.

3. Every nonzero submodule of M has a basis of at most n elements.

Proof. (1) Since M is finitely generated, let n ≥ 1 be minimal such that M has a set {x1, img, xn} such that M = ΣiDxi. If Σidixi = 0 and dk ≠ 0 for some k, then (since img it follows that M = ΣikDxi contradicting the minimality of n. So {x1, . . ., xn} is independent and hence is a basis of M. This proves (1).

(2) If {x1, . . ., xn} is a basis, then M = Dx1imgDxn where Dxi is simple, and we obtain a composition series

MDx2imgDxnimgDxn−1DxnDxn ⊃ 0

for M. Hence n is the composition length of M, proving (2).

(3) If KDM is a submodule, then K has a composition series by Lemma 2 and Theorem 1, and the composition length is at most n by (*). img

If D is a division ring, the number of elements in any finite basis of a module DM is called the dimension of M and denoted dim M.132 With this, most of the theorems about finite dimensional vector spaces (Section 6.1) go through for finitely generated modules over D.

Endomorphism Rings

If R is a ring and M and N are two R-modules, recall that a map α: MN is called R-linear (or an R-morphism) if α(x + y) = α(x) + α(y) and α(rx) = (x) for all x, y img M and all r img R. Many results about rings (in particular, Wedderburn's theorem) arise from representing them as rings of R-linear maps.

If M and N are two modules write

hom (M, N) = {α img α: MN, α is R-linear}.

If α, β img hom (M, N), define α + β and −α by

(α + β)(x) = α(x) + β(x)  and (− α)(x) = − α(x), for all x img M.

These are R-linear and make hom (M, N) into an abelian group. Furthermore, composition of maps distributes over this addition:

  γ(α + β) = γα + γβ whenever Mα,β N → γ K are R-linear,

  (α + β)δ = αδ + βδ whenever img are R-linear.

All these routine verifications are left to the reader.

Our interest here is in a special case: If M is any module, an R-linear map α: MM is called an endomorphism of M, and we write

end M = hom (M, M).

The additive abelian group end M becomes a ring, called the endomorphism ring of M, if we define addition as above and use composition of maps as the multiplication. Again we leave to the reader the routine verifications of the ring axioms, and that end M ≅ end N whenever MN. Note that the unity of end N is the identity map 1M.

If S is a ring, denote the ring of all n × n matrices over S by Mn(S). Wedderburn's theorem asserts that certain rings are isomorphic to Mn(D) where D is a division ring. The next result shows how matrix rings arise as endomorphism rings. If M is a module, recall that Mn denotes the external direct sum of n copies of M.

Lemma 3. Let RM be a module and write S = end M for the endomorphism ring. Then end(Mn) ≅ Mn(S) as rings for each integer n ≥ 1.

Proof. Identify Mn with the set of n-tuples from M, and let img be the standard maps: σi(x) = (0, . . ., x, . . ., 0), where x is in position i, and πi(x1, x2, . . ., xn) = xi. Then one verifies that

img  and  πjσi = δij1M, for all j and i,

where δij = 0 or 1 according as ij or i = j. 133 Given α img end (Mn), we have πiασj img S for all i and j, so we define

θ: end (Mn) → Mn(S) by θ(α) = [πiασj].

It routine to see θ(α + β) = θ(α) + θ(β) for all α and β, and that img the identity matrix. The (i, j) entry of the matrix θ(α)θ(β) is

Σk(πiασk)(πkβσj) = πiαkσkπk)βσj = πiαβσj,

so θ(α)θ(β) = θ(αβ). Thus, θ is a ring homomorphism; we claim it is an isomorphism.

If θ(α) = 0 then πiασj = 0 for all i and j, so

img

This shows that θ is one-to-one, and it remains to show that θ is onto. To this end, let [γij] img Mn(S), and define α: MnMn by

img.

Then, for every x img M we have

πiασj(x) = πi(α(0, . . ., x, . . ., 0)) = πi(γ1j(x), γ2j(x), . . ., γnj(x)) = γij(x).

It follows that πiασj = γij for all i and j, that is θ(α) = [γij]. This shows that θ is onto, and so proves the lemma. img

If V is an n-dimensional vector space over a field F, then VFn so Lemma 3 shows that end (V) ≅ Mn(F), a familiar fact from linear algebra.

We say that a mapping img acts as a left operator x img β(x) because we write β on the left of its argument. If img we have the composite map αβ: MK where αβ(x) = α[β(x)] for all x img M. Hence, because α and β are left operators, the notation αβ means “ first β then α”. This is somewhat unfortunate because the order gets reversed, but the use of left operators is very common. On the other hand, we could write a map img on the right of its argument, so that x img . In this case the composite img is given by x(βα) = ()α, so we write it as βα: MK. In particular, βα and means “ first β then α ” in the same order as the arrows. Hence composition means a different thing depending on whether the maps act on the left or the right, and we must always be clear which we are using. Lemma 4 below gives a good illustration of the use of right operators.

An element e img R is called an idempotent if e2 = e. Note that 0 and 1 are idempotents in any ring, and they are the only ones in a division ring (in fact in a domain). If e img R is an idempotent, we define eReR as follows:

eRe = {ere img r img R} = {s img R img es = s = se} = eRRe.

Then eRe is a ring with unity e, called the corner ring corresponding to e. The name comes from the following example: If img where F is a field, and if img then e2 = e and img

These corner rings arise as endomorphism rings in a natural way.

Lemma 4. If R is a ring and e2 = e img R, then eReend(Re) where endomorphisms of Re act as right operators.

Proof. Given a img eRe, define a right operator

ρa: ReRe by a = xa, for all x img Re.

Note that xa img Re because a img eReRe. Then ρa preserves addition and, in fact ρa img end(Re) because (rx)ρa = (rx)a = r(xa) = r(a) for all r img R and x img Re.134 Hence, we may define

θ: eReend(Re) by θ(a) = ρa for each a img eRe.

We show that θ is a ring isomorphism. We have θ(e) = ρe = 1Re, so θ preserves the unity. If a, b img eRe, we have ρa+b = ρa + ρb (verify) so θ preserves addition. Also ρab = ρaρb because ab = x(ab) = (xa)b = (a)ρb = x(ρaρb) for all x img Re, so θ preserves multiplication. Hence θ is a ring homomorphism. Moreover, θ is one-to-one because θ(a) = 0 implies that ρa = 0, that is, xa = 0 for all x img Re. Taking x = e gives that ea = 0, so a = 0 because a img eRe.

To show that θ is onto, let λ: ReRe be R -linear, and define a = . Then ae = a because a img Re, and ea = e() = (e2)λ = = a. Hence a img eRe. Finally, if x img Re then x = xe, so = (xe)λ = x() = xa = a. Since this holds for all x img Re, we have λ = ρa = θ(a), so θ is onto. This completes the proof. img

Taking e = 1 in Lemma 4 gives a very important special case.

Corollary. If R is a ring then R ≅ end (RR) as rings where the endomorphisms of RR are right multiplications by elements of R.

Wedderburn's Theorem

Wedderburn's theorem asserts that every simple, left artinian ring is isomorphic to a matrix ring over a division ring. The division ring comes from the next result, due in 1905 to Issai Schur.

Lemma 5. Schur's Lemma. Let M and N be simple modules.

1. If α: MN is R-linear, then either α = 0 or α is an isomorphism.

2. end M is a division ring.

Proof. (1) Observe that ker α and im α = α(M) are submodules of M and N, respectively. If α ≠ 0 then ker αM and α(M) ≠ 0, so ker α = 0 and α(M) = N by simplicity. Hence α is an isomorphism, proving (1). Now (2) follows if N = M. img

In 1907, J.H.M. Wedderburn proved the following important theorem for finite dimensional algebras. A ring R is called simple if the only ideals are 0 and R.

Theorem 3. Wedderburn's Theorem. The following conditions are equivalent for a ring R:

1. R is a simple ring that is left artinian.

2. R is a simple ring that has a simple left ideal.

3. RMn(D) for some n ≥ 1 and some division ring D.

4. The right–left analogues of (1) and (2).

Furthermore, the integer n is uniquely determined by R, as is the division ring D up to isomorphism.

Proof. (1)⇒(2) is clear, and (4) follows by the left–right symmetry in (3).

(2)⇒(3). Let L be a simple left ideal of R, and recall that LR is the set of all finite sums of products ba where b img L and a img R. Then LR is an ideal of R containing L, so LR = R because R is a simple ring. In particular, let 1 = b1a1 + b2a2 + img + bnan where n ≥ 1, and where bi img L and ai img R for each i. Assume that n is the smallest positive integer with this property. The reader can verify that

R = La1 + La2 + img + Lan. (*)

Note that Lai ≠ 0 for each i because biai ≠ 0 by the minimality of n. Hence LLai because the map x img xai is a nonzero, onto, R-morphism LLai, and L is simple. In particular each Lai is simple.

Now we claim that the sum in (*) is direct. By Theorem 3 §7.1, we must show that Lak ∩ (ΣikLai) = 0 for each k. Suppose not. Then Lak ∩ (ΣikLai) is a nonzero left deal contained in Lak. Hence, Lak ∩ (ΣikLai) = Lak because Lak is simple. But then Lak ⊆ ΣikLai, and it follows that R = ΣikLai, contrary to the minimality of n. Hence (*) is a direct sum.

Since LaiL for each i, (*) gives img But then Lemma 3 and the Corollary to Lemma 4 and give

img

Since end L is a division ring by Schur's lemma, (3) follows with D = end L.

(3)⇒(1). The ring Mn(D) is simple by the Corollary to Theorem 7 §3.3 so, given (3), it remains to show that Mn(D) is left artinian. But Mn(D) is a finite dimensional vector space over D (in fact the dimension is n2), and so is artinian as a D-space. Hence, the ring Mn(D) is left artinian because every left ideal is a D -subspace. This proves (1).

Uniqueness. The fact that RRLn shows that n is the composition length of RR and so is uniquely determined by R. To show that D is uniquely determined, we prove that D ≅ end K for any simple left ideal K of R. But the proof of (2)⇒(3) shows that RKn, and hence that LnKn. We have maps img, where τ is an isomorphism, σ1 is the inclusion, and the πi are the projections. Then πiτσ1 ≠ 0 for some i because τσ1(L) ≠ 0. Hence LK by Schur's lemma, and so end K ≅ end L = D, as required. img

Remark. The “left artinian” condition in (1) of Theorem 3 cannot be replaced with “ left noetherian”. In fact, there exists a simple, left and right noetherian domain that contains no simple left ideal. It is called the first Weyl algebra, after Hermann Weyl, and can be described roughly as the ring of polynomials over img in noncommuting indeterminates x and y which satisfy the condition that xyyx = 1. This ring first arose in quantum mechanics as an algebra generated by position and momentum operators.

Exercises 11.1

1. Show that a module RM is simple if and only if MR/L for some maximal left ideal L.

2. Show that the following are equivalent for a ring R: (1) R is a division ring; (2) every principal module Rx ≠ 0 is simple; (3) RR is simple.

3. If aR = bR where a, b img R, show that there is an R-isomorphism σ: RaRb such that σ(a) = b.

4. Given RM and m img M define λm: RM by λm(a) = am for all a img R.

a. Show that λm is R-linear for each m img M.

b. Show that m img λm is an abelian group isomorphism M → hom (RR, M).

5. If R is left noetherian (left artinian), show that the same is true of the corner ring eRe for any idempotent e = e2 in R. [Hint: If LeRe is a left idel, consider RL.]

6. Show that the following are equivalent for a ring R: (1) R is left noetherian (left artinian); (2) Mn(R) is left noetherian (left artinian). [Hint: Mn(R) is a free left R -module.]

7. If R is left noetherian, show that every finitely generated left R-module is left noetherian. [Hint: Lemma 2 and Theorem 5 §7.1.]

8. Let R be left artinian. If X is a subset of a left module RM, define the annihilator ann(X) = {a img R img ax = 0 for all x img X}.

a. Show that ann(M) = ann (X) for some finite subset XM.

b. If ann (M) = 0, show that RR is isomorphic to a submodule of M.

9. Complete the solution of Example 2 as follows: If img is a prime and we write img k ≥ 0}, show that the only subgroups of X that contain img are img for k ≥ 0. [Hint: If img Y a subgroup of X, choose img in Y where m and pn are relatively prime and n is maximal.]

10. Let KM be modules. If KNM where N is a submodule, show that N/K is a submodule of M/K, and that every submodule img of M/K has the form img for some submodule NK. [Hint: Theorem 5 § 8.1.]

11. Let RM be a module and let π2 = π img end M.

a. Show that M = π(M) ⊕ ker π.

b. If M = NK, N and K submodules, show that π2 = π img end M exists such that N = π(M) and K = ker π.

c. Show that π(M) = ker(1 − π) and (1 − π)(M) = ker π.

12. Show that a module M is noetherian if and only if every submodule is finitely generated.

13. Call a module M finite dimensional if it contains no infinite direct sum of nonzero submodules, and call M indecomposable if it is not a direct sum M = AB where A and B are both nonzero submodules.

a. Show that M is finite dimensional if it is either noetherian or artinian.

b. If M is finite dimensional show that M = N1N2imgNk where each Nj is indecomposable.

14. Suppose R is a ring for which RR is finite dimensional (preceding exercise). If ab = 1 in R show that ba = 1. [Hint: Write ba = e and show that e2 = e and RRRa = Re.]

15. Show that the converse of (2) of Schur's lemma is not true. That is find a module K such that end K is a division ring but K is not simple. [Hint: If F is a field, consider img let img and consider Re.]

16. Extend Lemma 4 as follows: If e2 = e and f2 = f are idempotents in a ring R, then eRf≅ hom(Re, Rf) as additive abelian groups. [Hint: If a img eRf and x img Re then xa img Rf.]

17. Let M be a module and let αimg end(M).

a. If M is noetherian and α is onto, show that α is one-to-one.

b. If M is artinian and α is one-to-one, show that α is onto.

[Hint: We have chains ker (α)⊆ ker (α2)⊆ img and α(M) ⊇ α2(M) ⊇ img.]

18. Fitting's Lemma. Suppose that the module M is both artinian and noetherian. If αimg end (M) show that there exists n ≥ 1 such that M = αn(M)⊕ ker (αn). [Hint: ker (α)⊆ ker (α2)⊆ img and α(M) ⊇ α2(M) ⊇ img.]

19. Let R be an n -dimensional algebra over a field F, and fix a basis {u1, u2, . . ., un} of R. Define θ: RMn(F) by θ(r) = [rij], where img

a. Show that θ is a one-to-one ring homomorphism.

The image θ(R) is called the regular representation of R; in each case find it.

b. R = img, F = img, basis {1, i}.

c. R = img, F = img, basis {1, i, j, k}.

d. Basis {1, u} where 12 = 1, 1u = u = u1 and u2 = 0.

3. Basis {1, u} where 12 = 1, 1u = u = u1 and u2 = 1.

11.2 The Wedderburn–Artin Theorem

The proof of Wedderburn's theorem (Theorem 3 §11.1) shows that if R is a simple ring containing a simple left ideal L, then R = L1L2imgLn where the Li are isomorphic, simple left ideals (in fact LiL for each i). Wedderburn actually treated the case where the Li are not necessarily all isomorphic, albeit in the special case when R is a finite dimensional algebra. We deal with this general case in this section, in the context of rings.

The work necessitates a look at modules that are the sum of a (possibly infinite) family of simple submodules. This investigation provides an extension of the well-known theory of vector spaces over any division ring. However, not surprisingly, transfinite methods are required.

The technique we need is called Zorn's lemma. Let img be a nonempty family of subsets of some set. A chain from img is a (possibly infinite) set img where either XiXj or XjXi for any i, j img I. We say that img is inductive if the union of any chain from img is again in img Zorn's lemma asserts that every inductive family img has a maximal member, that is a set img such that img implies Y = Z. A more general version of Zorn's lemma is discussed in Appendix C.

Semisimple Modules

Let R be a ring, and let {Mi img i img I} be a (possibly infinite) family of submodules of a module M. The sum ΣiimgIMi of these submodules is defined to be the set of all finite sums of elements of the Mi; more formally

ΣiimgIMi = {x1 + img + xm img m ≥ 1, xj img Mj for some j img I}.

This is a submodule of M that contains every Mi. The following lemma is the extension of Theorem 3 §7.1 to the case of infinite sets of submodules.

Lemma 1. The following are equivalent for submodules {Mi img i img I} of a module:

1. Every element of ΣiimgIMi is uniquely represented as a sum of elements of Mi for distinct i.

2. The only way a sum of elements from distinct Mi can equal 0 is if each of the elements is zero.

3. ΣiimgJMi is direct for every finite subset JI.

Proof. (1) ⇒ (2). If x1 + img + xm = 0 = 0 + img + 0 then each xi = 0 by (1).

(2) ⇒ (3). If img where the ij are distinct, then N is direct by (2) and Theorem 3 §7.1.

(3) ⇒ (1). If x img M let x = x1 + img + xm = y1 + img + yk where img and img Inserting zeros where necessary, we may assume that xi and yi are in img for each i, and that these img are distinct. Hence, x1 + img + xm = y1 + img + yk in img for distinct ijby (3). Thus, xi = yi for each i by Theorem 3 §7.1, proving (1). img

In this case we say that ΣiimgIMi is a direct sum, and we write it as ⊕iimgIMi. One reason for introducing these infinite direct sums here is that they provides the language needed to generalize some of the results about vector spaces.

If D is a division ring and DM is an module, let B = {bi img i img I} be a (possibly infinite) set of nonzero elements of M. Then B is said to span M if M = ΣiimgIDbi. The set B is called independent if Σribi = 0, ri img D, implies that ri = 0 for each i, equivalently if M = ⊕ iimgIDbi (since D is a division ring). A set B that is both independent and spans M is called a basis of M. Furthermore, each principal module Dbi is simple (again because D is a division ring). In this form, these facts can be stated much more generally.

It is shown in Appendix C that every module M over a division ring D has a basis. By the above discussion this means that M is a direct sum of simple submodules. In general, a module RM over a ring R is called semisimple if it is the direct sum of a (possibly infinite) family of simple submodules. The following example shows that RR is semisimple if R is the 2 × 2 matrix ring over a division ring.

Example 1. Let D be a division ring, and let img If img and img, show that L1 and L2 are each simple left ideals of R, that R = L1L2, and that L1L2 as R-modules.

Solution. L1 and L2 are left ideals by the definition of matrix multiplication, and it is clear that R = L1 + L2 and L1L2 = 0. Hence R = L1L2. We show that L1 is simple; the proof for L2 is similar. So let img say a ≠ 0. Given img we have img so L1 = Rx. A similar argument shows that L1 = Rx when b ≠ 0.

Finally, if img and img then it is a routine matter to show that the maps L1L2 given by x img xa and L2L1 given by x img xb are mutually inverse R-isomorphisms. Hence L1L2.

We are going to give several characterizations of semisimple modules, and the following notion is essential. A module M is said to be complemented if every submodule K is a direct summand, that is if M = KN for some submodule N. Hence every simple module is complemented, as is every finite dimensional vector space (Theorem 8(3) §6.1). A submodule N of a module M is called proper if NM, and a maximal proper submodule is called simply a maximal submodule.

Lemma 2. Let M be a complemented module. Then

1. Every submodule N of M is complemented.

2. Every proper submodule of M is contained in a maximal submodule.

Proof. (1) If KN let KK1 = M. Then N = N ∩ (KK1) = K ⊕ (NK1) by the modular law (Theorem 2 §7.1).

(2) If xN, write img is a submodule, NP and xP}. Then img is nonempty and inductive (verify) so, by Zorn's lemma, choose a submodule P maximal in img Since M is complemented, let M = PQ; we show that P is maximal by showing that Q is simple. If not, let AQ where A ≠ 0, Q. By (1) write Q = AB. Then PA and PB both strictly contain P, so neither is in img Hence x lies in both by the maximality of P, say x = p1 + a = p2 + b where p1, p2 img P, a img A and b img B. Since a img Q and b img Q, and since PQ is direct, it follows that a = b. But then a = b img AB = 0, so x = p1 img P, a contradiction. img

Lemma 3. Let M = ΣiimgIMi where each Mi is simple. If N is any submodule of M there exists JI such that M = N ⊕ (⊕ jimgJMj). In particular, M is complemented.

Proof. If N = M take J = ∅. If NM then some Mj img N so NMj = 0 (because Mj is simple). Hence, img is nonempty where

img is a direct sum}.

Let {Ji} be a chain from img and write J = ∪ Ji. To see that J is in img let

img,

where x img N and each img Since {Ji} is a chain, there exists some Jk containing all these it so (since img it follows that img for each t. Hence N + ΣiimgJMi is direct, which shows that img

Hence, by Zorn's lemma, img has a maximal member J0. If img it remains to show that L = M. Since M = ΣMi, we show that MiL for each i img I. This is clear for all i img J0. If i0 img IJ0, then img so img is not direct, that is, img Hence, img as img is simple, so img as required. img

With this we can characterize the semisimple modules.

Theorem 1. The following conditions are equivalent for a module M ≠ 0:

1. M = ΣiimgIMi, where each Mi is a simple submodule.

2. M = ⊕ iimgIMi, where each Mi is a simple submodule.

3. M is complemented.

4. Every maximal submodule of M is a direct summand, and every proper submodule is contained in a maximal submodule.

Proof. (1)⇒(2) and (2)⇒(3). These are by Lemma 3, the first with N = 0.

(3)⇒(4). This is clear by Lemma 2.

(4)⇒(1). Let S denote the sum of all simple submodules of M (take S = 0 if there are none). Suppose SM. By (4) let SN where N is maximal in M, and write M = NK for some submodule K, again by (4). Then K is simple (because KM/N) so KSN, contrary to KN = 0. So S = M and (1) follows. img

A module M ≠ 0 is called semisimple if it satisfies the conditions in Theorem 1, We also regard 0 as a semisimple module. Hence every module over a division ring is semisimple, and the ring in Example 1 is semisimple as a left module over itself. Lemma 3 gives

Corollary 1. If M is semisimple so also is every submodule and image of M. More precisely, if NM = ⊕ iimgIMi where each Mi is simple, then

N ≅ ⊕ iimgJMi  and  M/N ≅ ⊕ iimgIJMi for some JI.

Note that we need not have N = ⊕ iimgJMi in Corollary 1. As an example, let img where F is a field. Then M = M1M2 where img and img But if N = {(a, a) img a img F} then Mi6 ⊆ N for i = 1, 2.

Corollary 2. If M is finitely generated then M is semisimple if and only if every maximal submodule is a direct summand.

Proof. By Theorem 1, it remains to show that every proper submodule KM is contained in a maximal submodule of M. To this end, let img we show that img contains maximal members (they are then maximal in M). So let {Xi img k img I} be a chain from img and write X = ∪ iimgIXi. By Zorn's lemma it suffices to show that XM. But if X = M, let M = Rm1 + img + Rmk (as M is finitely generated). Then each img for some it img I. But the Xi are a chain so there exists k img I such that img for each it. Hence each mt img Xk, so MXk, a contradiction. Hence XM after all. img

Lemma 4. The following are equivalent for a semisimple module M:

1. M is finitely generated.

2. M = M1M2imgMn where each Mi is simple.

3. M is artinian.

4. M is noetherian.

Finally the decomposition in (2) is unique: If M = N1N2imgNm, where each Ni is simple then m = n and (after relabeling) NiMi for each i.

Proof. (1)⇒(2). If M = ⊕ iMi where the Mi are simple, the generators of M all lie in a finite sum img by (1).

(2)⇒(3) and (2)⇒(4). By (2), MM2imgMnimgMn ⊃ 0 is a composition series for M so (3) and (4) follow from Theorem 1 §11.1.

(3)⇒(1) and (4)⇒(1). If M = ⊕ iimgIMi where I is infinite, we may assume that {1, 2, 3, img } ⊆ I. Then (M1M2M3img) ⊃ (M2M3img) ⊃ img contradicts (3) and M1M1M2img contradicts (4).

Finally, as we saw above, (2) gives rise to a composition series M with factors M1, M2, . . ., Mn. Hence, the last statement in the corollary follows from the Jordan–Hölder theorem (Theorem 1 §11.1). img

Note that Lemma 4 proves the uniqueness of the number of elements in a finite basis of a module over any division ring (so we can speak of dimension). In fact, a version of the uniqueness actually goes through for arbitrary infinite direct sum decompositions of a semisimple module, a result beyond the scope of this book.

Homogeneous Components

Before proceeding we need a technical lemma.

Lemma 5. Let M = ⊕ iimgIMi with each Mi simple and let KM be a simple submodule, Then there exist i1, i2, . . ., im in I such that img and img for each t = 1, 2, . . ., m.

Proof. Choose 0 ≠ y img K, and write img, where img for each t. Then img so, because K is a simple module, img Because this is a direct sum, we can define img as follows: If x img K take img, where img and img for each j. Then αt is R-linear for each t, and αt ≠ 0 because img So αt is an isomorphism by Schur's lemma. img

If KM are modules with K simple, define the homogeneous component H(K) of M generated by K as follows:

H(K) = Σ{XM img Xis a submodule and XK}.

We say that H(K) = 0 if M contains no copy of K. Hence, if KM is simple then H(K) is a submodule of M containing every submodule isomorphic to K (and hence K itself). In fact every simple submodule of H(K) is isomorphic to K.

Lemma 6. Let KM be modules with K simple. The following are equivalent for a simple submodule L M:

1. LK.

2. H(L) = H(K).

3. LH(K).

Proof. (1)⇒(2). This is because XL if and only XK by (1).

(2)⇒(3). Here LH(L) = H(K) by (2).

(3)⇒(1). If LH(K) = Σ{XM img XK}, then LX for some XK by Lemma 5. This proves (1). img

A submodule N of M is said to be fully invariant in M if α(N) ⊆ N for every endomorphism α: MM. Clearly 0 and M are fully invariant for every module M. Recall that the endomorphisms of RR are just the right multiplications by elements of R (Corollary to Lemma 4 §11.1). It follows that a left ideal L of R is fully invariant in RR if and only if L is an ideal of R.

We can now prove the main structure theorem for semisimple modules. For convenience, we say that submodules K and N meet if KN ≠ 0.

Theorem 2. Let M be a semisimple module. Then the following hold:

1. M is the direct sum of its homogeneous components.

2. Each homogeneous component of M is fully invariant in M.

3. Each fully invariant submodule of M is the direct sum of the homogeneous components it meets.

Proof. Let {H(Ki) img i img I} be the distinct homogeneous components of M.

(1) M = ΣiimgIH(Ki) because M is semisimple; to see that this sum is direct we show that H(Kt) ∩ [ΣitH(Ki)] = 0 for each t img I, and invoke Lemma 1. But if H(Kt) ∩ [ΣitH(Ki)] ≠ 0 then it contains a simple submodule L (being semisimple). Then LKt by Lemma 5 because LH(Kt); and LKi for some it by Lemma 5 because L ⊆ ΣitH(Ki). Hence, KtLKi and so H(Kt) = H(Ki) by Lemma 6, contrary to the choice of the H(Ki).

(2) Consider H(K), where KM is simple. If α img end M we must show that α[H(K)] ⊆ H(K), that is α(L) ⊆ H(K) whenever LM and LK. But since L is simple, either α(L) = 0 or α(L) ≅ LK by Schur's lemma. Either way, α(L) ⊆ H(K).

(3) Let the Ki be as above, and let NM be fully invariant. We show that N = Σ{H(Ki) img H(Ki) ∩ N ≠ 0}. Every simple submodule of M is in some H(Ki) by Lemma 5, so N ⊆ Σ{H(Ki) img H(Ki) ∩ N ≠ 0} because N is semisimple. Conversely, if H(Ki) ∩ N ≠ 0 then there exists UH(Ki) ∩ N such that UKi. If LKi is arbitrary then LU by Lemma 5, so let σ: UL be an isomorphism. Since M is complemented (Theorem 1), write M = UU1, and define α: MM by α(u + u1) = σ(u). Then L = σ(U) = α(U) ⊆ α(N) ⊆ N, where α(N) ⊆ N because N is fully invariant. Since LKi was arbitrary, it follows that H(Ki) ⊆ N. img

A semisimple module M is called homogeneous if it has only one homogeneous component, that is (by Lemma 6) if all simple submodules of M are isomorphic. In particular if D is a division ring and R = M2(D), then Example 1 shows that RR is a homogeneous, semisimple module. In fact much more is true as we shall see below.

Free and Projective Modules

The Wedderburn–Artin theorem shows that rings R such that RR is semisimple as a left R-module are isomorphic to finite direct products of matrix rings over division rings. Other equivalent conditions on R are also considered.

Finitely generated free and projective modules were discussed in Section 7.1. However, the Wedderburn–Artin theorem involves arbitrary projective modules, so we pause to review these notions. Let RW be a module, and let B be a set of nonzero elements of W. We make three definitions:

B is said to generate W if  img

B is called independent  if  img implies each ri = 0.

B called a basis of W if it is independent and generates W.

A module RW that has a basis is called a free module. Note that the second condition above implies that RwiRR for each i img B (see the corollary to Theorem 1 §7.1). Hence, every free module is isomorphic to a direct sum of copies of R. The finitely generated free modules in Section 7.1 are examples, but free modules with bases of arbitrary size can easily be constructed.

If I is any nonempty set and R is any ring, there exists a free R -module with a basis indexed by I. If i img ri is any function IR, write it as img Hence img if and only if ri = si for each i img I, and we call ri the ith component of img We call img an I-sequence from R. Define

img for all but finitely many i img I }.

If I has n elements it is clear that R(I) = Rn. In general, R(I) becomes a left R-module with componentwise operations:

img

If ei denotes the I-sequence with ith component 1 and all other components 0, then it is a routine matter to check that {ei img i img I} is a basis of R(I), called the standard basis. Hence, R(I) = ⊕ iimgIRei, where ReiRR for each i.

Now let {xi img i img I} be a generating set for a module RM, that is, M = ΣiimgIRxi. Let W be a free module with a basis img indexed by I (for example W = R(I)). Given r1, r2, . . ., rn in R, define

β: WM  by img

Because B is a basis, this map is well defined, and it is evidently onto. Since every module M has a generating set (the set of all nonzero elements, for example), this proves the first part of

Lemma 7. Let M denote any left R -module.

1. M is an image of a free module.

2. If α: MW is onto and W is free, then ker α is a direct summand of M.

Proof. We proved (1) above. As to (2), let img be a basis of W. As α is onto, let img, xi img M for each i. By the above discussion, there exists β: WM such that img for each i. Then img for each i, so αβ = 1W because the img generate W. But then M = ker αβ(W) by Lemma 8 below, proving (2). img

Lemma 8. If α: MP is onto, the following conditions are equivalent:

1. There exists β: PM such that αβ = 1P.

2. ker α is a direct summand of M, in fact M = ker αβ(P).

In this case the map α is said to split.

Proof. (1) ⇒ (2). If αβ = 1P as in (1), let m img M. As α(m) img P, we have α(m) = 1P[α(M)] = αβα(m). Thus mβα(m) img ker α, so M = ker α + β(P). But if m img ker αβ(P), let m = β(p), p img P. Then 0 = α(m) = αβ(p) = p, so m = β(p) = β(0) = 0. This proves that ker αβ(P) = 0 and so proves (2).

(2) ⇒ (1). Given (2), let M = ker αQ. Observe that P = α(M) = α(Q), so we define β: PM as follows: if p img P and p = α(q), q img Q, define β(p) = q. This is well defined because if p = α(q1), q1 img Q, then qq1 img Q ∩ ker α = 0. Now given p = α(q) in P then αβ(p) = α(q) = p, proving (1).

A module RP is called projective if it satisfies the condition:

img is R -linear and onto then ker α is a direct summand of M.

Hence all free modules are projective by Lemma 7 (but see Example 3 below). Also, P is projective if and only if every onto R-morphism MP splits. We need

Lemma 9. If P is projective and QP, then Q is projective.

Proof. Let α: MQ be onto. If σ: QP is an isomorphism then σα: MP is onto so ker α = ker σα is a direct summand of M. Hence, Q is projective.

Theorem 3. The following conditions on a module RP are equivalent:

1. P is projective.

2. P is isomorphic to a direct summand of a free module.

3. If α is onto in the diagram, then γ exists such that αγ = β.

4. If img is onto there exists img such that αβ = 1P.

img

Proof. (1)⇒(2). If img is onto, then W = ker αQ for some Q by (1). Hence, P = α(P) ≅ W/ker αQ, proving (2).

(2)⇒(3). Since being projective is preserved under isomorphism, we may assume that W = PQ is free. Let π: WP be the projection defined by π(p + q) = p for p img P and q img Q, and let img be a basis of W. Given α as in the diagram, we must construct γ: PM such that αγ = β.

Note that img for each i img I. Since α: MN is onto, there exists mi img M such that img But the img are a basis of W, so there exists λ: WM such that img for each i. Hence,

img,  for each i.

It follows that αλ = βπ because the img generate W. Finally, let γ: PM be the restriction of λ to P, that is, γ(p) = λ(p) for all p img P. Compute

αγ(p) = αλ(p) = βπ(p) = β(p), for all p img P,

because π(p) = p. Hence, αγ = β, as required.

(3)⇒(4). Take N = P and β = 1P in the diagram.

(4)⇒(1). This is Lemma 8.

Example 2. A module RP is a principal projective module if and only if img for some e2 = e img R.

Solution. If P = Rx is projective, define α: RRx by α(r) = rx for all r img R. Then α is onto so RR = ker αL for some left ideal L. Hence PR/ker αL. But direct summands of RR have the form Re for some idempotent e by Example 6 §7.1, so we have PRe for some e2 = e img R.

Conversely, if e2 = e then ReR(1 − e) = R is free, so Re is projective by Theorem 3.

Example 3. There exist projective modules that are not free.

Solution. Let F be a field, and let R = F × F. Then e = (1, 0) is an idempotent in R, so Re = {(a, 0) img a img R} is projective. But dim F(Re) = 1 so Re is not free since any free R -module has F-dimension at least 2 (because dim FR = 2).

The Wedderburn–Artin Theorem

Let A be a left ideal of a ring R. If X is a nonempty subset of a module RM, define AX to be the set of all finite sums of elements ax where a img A and x img X. This is a submodule of M, and it is easy to verify that

(A + B)X = AX + BX and A(BX) = (AB)X

hold for all left ideals A and B. Note that RX = X if X is a submodule of M. Taking M = R in the above discussion shows that multiplication of left ideals is associative: A(BC) = (AB)C for any left ideals A, B, and C.

An ideal A is called nilpotent if An = 0 for some n≥ 1; equivalently if any product of n elements of A is zero. The next result, proved in 1942 by Richard Brauer, characterizes the non-nilpotent, simple left ideals.

Lemma 9. Brauer's Lemma. Let K be a simple left ideal of a ring R. Then either K2 = 0 or K = Re for some nonzero idempotent e2 = e img K.

Proof. Assume that K2 ≠ 0, so that Ka ≠ 0 for some 0 ≠ a img K. Since KaK is a left ideal, it follows that Ka = K because RK is simple. In particular ea = a for some 0 ≠ e img K. Hence e2a = ea, so e2e img B = {b img K img ba = 0}. Moreover, B is a left ideal and BK because Ka ≠ 0, so B = 0, again by simplicity. This means that e2 = e, and so ReK because e img K. But e ≠ 0 so Re = K by a third appeal to the simplicity of RK. This is what we wanted. img

We now turn to another property of rings that plays a prominent role in the Wedderburn-Artin theorem. A ring R is called semiprime if it satisfies the following equivalent conditions (the routine verifications are left to the reader):

1. If A2 = 0, A a left (or right, or two-sided) ideal, then A = 0.

2. If An = 0, n ≥ 1, A a left (or right, or two-sided) ideal, then A = 0.

3. If aRa = 0 where a img R, then a = 0.

Hence, every ring with no nonzero nilpotent elements is semiprime, and the converse is true for commutative rings. A product R = R1 × R2 × img × Rn of rings is semiprime if and only if each Ri is semiprime (using condition (3)). A matrix ring Mn(R) is semiprime if and only if R is semiprime because the ideals of Mn(R) all have the form Mn(A) for some ideal A of R by Lemma 3 §3.3.

The next theorem gives several useful characterizations of when a ring R is semisimple as a left module over itself.

Theorem 4. If R is a ring the following conditions are equivalent:

1. RR is semisimple

2. Every left R-module is semisimple.

3. Every left R-module is projective.

4. R is left artinian and semiprime.

Proof. (1) ⇒ (2). Each module RM is an image of a free module, and free modules are semisimple by (1). So (2) follows from Corollary 1 of Theorem 1.

(2) ⇒ (3). Given a module M, let α: NM be an onto R-linear map. Then ker α is a summand of N because N is semisimple by (2). This proves (3).

(3) ⇒ (1). Let L be any left ideal of R, and consider the coset map ϕ: RR/L. Then L = ker ϕ so L is a direct summand of RR because R/L is projective by (3). Hence RR is complemented, and so is semisimple by Theorem 1.

(1) ⇒ (4). First, R is left artinian by Lemma 4 because RR is finitely generated. Suppose A is an ideal of R with A2 = 0. By Theorem 1 R = AL, L a left ideal, so ALAL = 0. Hence, A = AR = A2 + AL = 0 + 0 = 0, so R is semiprime.

(4) ⇒ (1). If RR is not semisimple, let L be minimal among nonzero left ideals of R that are not semisimple (by (4)). Since L ≠ 0, let KL be a simple left ideal, again by (4). Then K2 ≠ 0 because R is semiprime, so K = Re for some 0 ≠ e2 = e by Brauer's lemma. Since R = KR(1 − e) and KL, we obtain

L = K ⊕ (LR(1 − e)) by the modular law (Theorem 2 §7.1).

Hence, there are two cases: either LR(1 − e) = 0 (in which case L = K is simple) or LR(1 − e) is semisimple (by the minimality of L). Either way, L is semisimple, a contradiction. So RR is semisimple. img

Note that we cannot replace “left artinian” by “left noetherian” in (4) of Theorem 4. In fact the ring img of integers is noetherian and semiprime, but it is not semisimple.

The first, and possibly the most important application of the theory of semisimple modules is to prove the following fundamental theorem: If R is a ring then RR is semisimple if and only of R is a finite direct product of matrix rings over division rings. This result was first proved in 1908 by Wedderburn for finite dimensional algebras. Then in 1927, Artin replaced the finite dimensional hypothesis by the descending chain condition on left ideals.

Theorem 5. Wedderburn–Artin Theorem. The following conditions are equivalent for a ring R:

1. RR is semisimple.

2. RR is semisimple.

3. img for division rings Di.

Moreover, the integers k, n1, . . ., nk in (3) are uniquely determined by R as are the division rings Di up to isomorphism.

Proof. We need prove only (1) img (3) by the right–left symmetry of (3).

(1)⇒(3). Let H1, H2, . . ., Hm be the homogeneous components of RR. Hence

img

by Theorem 2. Moreover, each Hi is an ideal of R (being fully invariant), so RH1 × H2 × img × Hm as rings by Theorem 7 §3.4 (and induction). So, by Wedderburn's theorem, it remains to show that Hi is left artinian and simple for each i. Write H = Hi for convenience. The ring R is left artinian (by Lemma 4 since it is semisimple), so the same is true of RH. But the left ideals of the ring H are exactly the left ideals of R that happen to be contained in H (verify). Hence HH is artinian. Finally, if A ≠ 0 is an ideal of the ring H, then A is an ideal of R because H = Re where e2 = e is central in R (verify). Hence, A is fully invariant in RR. But then Theorem 2 shows that A = Σ{Hi img AHi ≠ 0}, and it follows that A = H. Thus, the ring H is left artinian and simple, and (3) follows by Wedderburn's theorem.

(3)⇒(1). If D is a division ring, the ring Mn(D) = K1K2imgKn where Kj is the left ideal of all matrices with only column j nonzero. It is a routine verification (see Example 1) that each Kj is a simple left ideal of Mn(D), so Mn(D) is semisimple. Hence, (1) follows from (3).

Uniqueness. Suppose img is another such decomposition where each Bi is a division ring. Lemma 10 below shows that k = l and, after relabeling, that img for each i = 1, 2, img, k. We can now apply the uniqueness in Wedderburn's theorem (Theorem 3 §11.1). img

The rings in Theorem 5 are called semisimple rings. The next result shows that this property is inherited by several related rings—the proofs are Exercises 10, 11, and 12.

Corollary. If R is a semisimple ring, so also is every matrix ring Mn(R), every factor ring R/A, and every corner ring eRe where e2 = e.

Note that subrings of semisimple rings need not be semisimple (consider img

Lemma 10. Let R1 × R2 × img × RkS1 × S2 × img × Sl as rings, where each Ri and each Sj is a simple ring with a left composition series. Then k = l and, after relabeling, RiSi for each i.

Proof. Write img and img We may assume that kl. Let σ: RS be a ring isomorphism. The ideals of S are all of the form ΠAj, where Aj is an ideal of Sj for each j (Exercise 4). Hence, since σ(R1) is a simple ideal of ΠSj, it must be one of the Sj; by relabeling assume that σ(R1) = S1. Similarly σ(R2) = Sj for some j, and j ≠ 1 because σ(R1) ≠ σ(R2), σ is one-to-one, and R1R2 = 0. So, after relabeling, let σ(R2) = S2. Continue to obtain S = σ(R1) × img × σ(Rk) × Sk+1 × img × Sl. Hence, SR × (Sk+1 × img × Sl) so, since S and R have the same left composition length, the Jordan–Hölder theorem (Theorem 1 §11.1) shows that Sk+1 × img × Sl has length zero. Hence, k = l and Si = σ(Ri) ≅ Ri for each i. This completes the proof. img

Wedderburn's 1908 version of Theorem 5 was a breakthrough. 135 To quote Emil Artin, “ This extraordinary result has excited the fantasy of every algebraist and still does so to this day. Very great efforts have been directed toward a deeper understanding of its meaning.” 136 When the Wedderburn-Artin theorem appeared in 1927 it was a landmark in algebra. It influenced a generation of ring theorists and has inspired many generalizations.

Wedderburn's theorem asserts that a left artinian simple ring is a matrix ring over a division ring. In 1945, Nathan Jacobson, and independently Claude Chevalley, extended Wedderburn's theorem by dropping the artinian hypothesis and showing that a simple ring with a simple left ideal must be isomorphic to a “ dense” subring of the ring of endomorphisms of a vector space over a division ring. This result is called the density theorem.

Also in 1945, Jacobson showed that the intersection J of all the maximal left ideals of R always equals the intersection of the maximal right ideals of R. He then proved that J is the largest ideal with the property that 1 + a is a unit for all a img J, extending work of Sam Perlis done in 1942 for finite dimensional algebras over a field. The ideal J is called the Jacobson radical of the ring R. It is known that, if R is left (or right) artinian, the factor ring R/J is semisimple, and idempotents can be lifted modulo J in the sense that if a2a img J then there exists e2 = e such that ae img J. In 1960 this led Hyman Bass to carry the theory further. He called a ring R semiperfect if R/J is semisimple and idempotents can be lifted modulo J, and he showed that many properties of left artinian rings carry over to these semiperfect rings.137

Finally, the Wedderburn–Artin theorem shows that a left artinian semiprime ring is semisimple, and another natural question is what happens if we replace the left artinian condition by the requirement that the ring be left noetherian. In 1960, Alfred Goldie proved a fundamental structure theorem for the semiprime left noetherian rings. There is a way of embedding certain rings into a ring of left quotients, a noncommutative version of the construction of the field of quotients of an integral domain. Goldie showed that every left noetherian, semiprime ring has a semisimple ring of left quotients.

Exercises 11.2

1. Describe the semisimple img -modules, and the homogeneous ones.

2. Let R be a ring and let img.

1. If R is a semisimple ring and L is a left ideal, show that L = Re for some e2 = e(so L is principal). [Hint: R is complemented.]

2. In general, if Ra = Re with e2 = e, show that aR = fR for some f2 = f. [Hint: Show that ata = a where e = ta, and use f = at.]

3. Let MR be a right R-module, and write E = end(MR).

1. Show that M is a left E -module via α · x = α(x) for all α img E and x img M.

2. Show that EM is simple if and only if the only fully invariant submodules of MR are 0 and M.

4. If A is an ideal of a product R = R1 × R2 × img × Rn of rings, show that A has the form A = A1 × A2 × img × An, where Ai is an ideal of Ri for each i.

5. If N1, . . ., Nm are all maximal submodules of a module M, show that M/(∩ Ni) is semisimple.

6. Let RM = H1H2imgHn where each Hi is fully invariant in M. Show that end M≅ end(H1)× end(H2img × end(Hn) as rings.

7. If M is a finitely generated, semisimple module, show that end M is a semisimple ring. [Hint: Preceding exercise and Lemma 3 §7.1.]

8. Let K be a simple module. If M is any module define HM(K) to be the sum of all submodules of M isomorphic to K, where HM(K) = 0 if M has no submodule isomorphic to K. If α: MN is an R -linear map, show that α[HM(K)] ⊆ HN(K).

9. Show that every domain with a simple left ideal is a division ring. [Hint: Brauer's lemma.]

10. If R is semisimple show that Mn(R) is semisimple for all n ≥ 1. [Hint: Theorem 4.]

11. If R is semisimple show that R/A is a semisimple ring for all ideals A of R.

12. If R is semisimple show that eRe is semisimple if e2 = e img R. [Hint: Theorem 4.]

13. If R is semiprime and e2 = e img R, show that the following are equivalent: (1) Re is a simple left ideal; (2) eRe is a division ring; (3) eR is a simple right ideal. [Hint: Lemma 4 §11.1.]

14. If img a field, let img Show that eRe is a field, but Re is not simple (so the converse of Brauer's lemma is false). Show that eR is simple.

15. If RR is semisimple, show that R is right and left noetherian. [Hint: Lemma 4.]

16. Let R be a semiprime ring.

a. If L and M are left ideals, show that LM = 0 if and only if ML = 0.

b. If A and B are ideals show that AB = 0 if and only if AB = 0.

c. If A is an ideal, and r img R, show that rA = 0 if and only if Ar = 0.

17. If R = R1 × R2 × img × Rn, where the Ri are rings. Show that R is semiprime if and only if each Ri is semiprime.

18. If R is semiprime show that eRe is also semiprime for any e2 = e img R.

19. Show that a ring R is semiprime if and only if Mn(R) is semiprime for some (any) n ≥ 1.

20. A ring R is called prime if AB = 0, A, B ideals, implies that A = 0 or B = 0.

a. Show that the commutative prime rings are the integral domains.

b. Show that a ring R is a prime and left artinian if and only if RMn(D) for some n ≥ 1 and some division ring D.

21. If P1, P2, . . ., Pn, are projective modules, show that img is projective.

22. If M is a left module, define the socle of M, denoted soc(M), to be the sum of all the simple submodules of M. (Take soc(M) = 0 if M contains no simple submodule). Show that

a. soc(M) is fully invariant in M.

b. If NM is a submodule then soc(N) = N∩ soc(M).

c. If M = N1N1 then soc(M) = soc(N1)⊕ soc(N2).

Notes

129. Actually this requires a set-theoretical theorem called transfinite recursion. This is discussed in Appendix D.

130. The ACCP designation used extensively in Chapter 5 is just the ACC applied to the set of all principal ideals in an integral domain.

131. These are the analogues of the axioms M1–M4 in Section 7.1.

132. In fact, every nonzero module over a division ring has a basis (and so is free). The proof requires Zorn's lemma (Example 2, Appendix C).

133. img is often called the Kronecker delta.

134. If we wrote ρa as a left operator then img need not be equal.

135. Maclagan Wedderburn, J.H. On hypercomplex numbers, Proceeding of the London Mathematical Society, Series 2, 6 (1908), 77–118.

136. Artiin, E. The influence of J.H.M Wedderburn on the development of modern algebra, Bulletin of the American Mathematical Society 56 (1950), 65–72.

137. Yes, there are perfect rings, indeed left and right perfect rings. They were also introduced by Bass, but a discussion of these rings is beyond the scope of this book.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
3.20.238.187