Jacobson-Hersteins Commutativity Theorem

In this article, we present Jacobson-Hersteins Commutativity Theorem, which employs a classic reduction in its proof, that is, when trying to prove certain theorems about rings, it is possible to reduce one class of rings to another, simpler, class, in the following sense.

We shall now introduce the required definitions and intermediate results. They are mainly adopted from "A First Course in Noncommutative Rings", 2nd edition, T.Y. Lam.

The Jacobson radical (or the left radical) of a ring R, denoted by rad  R, is the intersection of all the maximal left ideals of R. If R 
eq 0, then maximal left ideals always exist by Zorns Lemma, and rad  R 
eq R. Otherwise if R = 0, then there exists no maximal left ideals. In this case, the Jacobson radical is defined to be zero.

The following result characterizes the elements of rad  R.

(1.1) Proposition. For any y in rad  R, the following are equivalent.

  1. y in rad  R.

  2. 1 - xy is left invertible for any x in R.

  3. yM = 0 for any simple left R-module M.

Proof. 1 Rightarrow 2. Suppose that y in rad  R and 1 - xy is not left invertible for some x in R. Then R cdot (1 - xy) is contained in a maximal left ideal mathfrak{m} of R. However, 1 - xy in mathfrak{m} and y in mathfrak{m} imply that 1 in mathfrak{m}, a contradiction.

2 Rightarrow 3. Suppose that 1 - xy is left invertible for any x in R and y does not kill m for some m in M. Then it follows from the simplicity of M that R cdot ym = M. In particular, x cdot ym = m for some x in R. Therefore (1 - xy)m = 0. By 2, m = 0, a contradiction.

3 Rightarrow 1. For any maximal left ideal mathfrak{m}, R/mathfrak{m} is a simple left R-module. By 3, y cdot R/mathfrak{m} = 0. It follows that y in mathfrak{m}. By definition, y in rad  R.

It is not hard to see that rad  R = igcap Ann(M), where M ranges over all the simple left R-modules. In particular, rad  R is an ideal of R.

Also, it is not hard to see that a fourth condition can be added to (1.1) to strengthen the second condition.

(1.2) Proposition. For any y in rad  R, the following are equivalent.

  1. 1 - xy is left invertible for any x in R.

  2. 1 - xyz is invertible for any x, z in R.

Proof. 1 Rightarrow 2. Since rad  R is an ideal, yz in rad  R. By (1.1), u(1 - xyz) = 1 for some u in R. Since rad  R is an ideal, xyz in rad  R. Again by (1.1), u = 1 + u(xyz) is left invertible. Since u is also right invertible, u is invertible and hence 1 - xyz is invertible.

2 Rightarrow 1. Clear.

Consequently, we have the following corollary.

(1.3) Corollary. The left radical of R agrees with its right radical.

The Jacobson radical leads to a new notion, which we now introduce.

Semiprimitivity. A ring R is called semiprimitive if rad  R = 0.

Observe that R/rad  R is always semiprimitive, as rad  (R/mathfrak{A}) cong (rad  R)/mathfrak{A} for any ideal mathfrak{A} of R.

Alternatively, semiprimivity can be characterized in the following way.

(1.4) Proposition. A ring R is semiprimitive if and only if R has a faithful semisimple left module M.

Proof. Suppose that M exists. Since rad  R acts as zero on all left simple R-modules, (rad R) cdot M = 0. Then the faithfulness of M implies that that rad  R = 0 . In other words, R is semiprimitive. Conversely, suppose that rad  R = 0. Let {M_i} be a complete set of mutually nonisomorphic simple left R-modules. Then M = igoplus M_i is semisimple, and

Ann(M) = igcap Ann(M_i) = rad  R. Since rad  R = 0, M is a faithful R-module.

The notion of a left primitive ring can be introduced in a similar manner.

Primitivity. A ring R is called left primitive if R has a faithful simple left module.

Although semiprimitivity is left-right symmetric, primitivity, on the other hand, is not. The first example was constructed by G. Bergman in 1964. See Errata: A Ring Primitive on the Right but Not on the Left for more information.

Before proceeding, we extend the notion of primitivity to ideals.

Primitivity. An ideal mathfrak{A} subseteq R is called left primitive if R/mathfrak{A} is left primitive.

And we have the following characterization of left primitive ideals.

(1.5) Proposition. An ideal mathfrak{A} in R is left primitive if and only if mathfrak{A} is the annihilator of a simple left R-module.

Proof. Suppose that mathfrak{A} = Ann(M), where M is a simple left R-module. Then the naturalR/mathfrak{A}-action on M is faithful. In other words, R/mathfrak{A} is a left primitive ring. Conversely, suppose that R/mathfrak{A} is left primitive and M is a faithful simple left R/mathfrak{A}-module. When considered as an R-module, M preserves simplicity. Moreover, Ann(M) = mathfrak{A}.

(1.5) implies that the Jacobson radical rad  R is the intersection of all the left primitive ideals in R.

By definition, a left primitive ring is always semiprimitive. Conversely, it is also possible to reduce a semiprimitive ring to a left primitive ring, in the following procedure.

Let {R_i} be rings, and epsilon: R 	o prod R_i be an injective ring homomorphism. epsilon represents R as a subdirect product of the R_is if each of the maps R 	o R_i obtained by composing epsilon with the coordinate projections is surjective. Moreover, a subdirect product representation is trivial if one of the maps R 	o R_i is an isomorphism.

(1.6) Theorem. A nonzero ring R is semiprimitive if and only if R is a subdirect product of left primitive rings.

Proof. Suppose that R is semiprimitive. Let {mathfrak{A}_i} be the family of left primitive ideals in R. Then igcap mathfrak{A}_i = 0 and R is a subdirect product of left primitive rings {R/mathfrak{A}_i}. Conversely, suppose that epsilon: R 	o prod R_i is a subdirect product representation, where R_is are left primitive rings. Then mathfrak{A}_i = ker(R 	o R_i) is a left primitive ideal, and igcap mathfrak{A}_i = 0. Since rad  R is contained in igcap mathfrak{A}_i, rad  R = 0.

We shall now define the notion of density, which will play an important role in the structure theorem of left primitive rings.

Acts Densely. Let R, k be rings, and V be an (R,k)-bimodule. R acts densely on V_k if for any f in End(V_k) and v_1, ..., v_n in V, there exists r in R such that rv_i = f(v_i) for all 1 leq i leq n.

(1.7) Lemma. In the above notation, if _RV is a semisimple left R-module and k = End(_RV). Then any R-submodule W of V is an End(V_k)-submodule.

Proof. Write V = W oplus W for some R-submodule W of V. Let e in End(_RV) be the projection of V on W with respect to this decomposition. Then for any f in End(_RV), f(W) = f(We) = (fW)e subseteq W, implying that W is an End(_RV)-submodule of V.

(1.7) prepares us for the following fundamental result.

(1.8) Density Theorem. Let R be a ring, and V be a semisimple left R-module. Then R acts densely on V_k, where k = End(_RV).

Proof. For any f in End(_RV) and any v_1, ..., v_n in V, define 	ilde{f}: 	ilde{V} 	o 	ilde{V} by 	ilde{f} = (f, f, ..., f), where 	ilde{V} = V^n. It is routine to verify that 	ilde{f} in End(	ilde{V}_{	ilde{k}}), where 	ilde{k} = End(_R	ilde{V}) cong M_n(End(_RV)) = M_n(k). Now consider the cyclic R-submodule 	ilde{W} of 	ilde{V} generated by (v_1,...,v_n) in 	ilde{V}. By (1.7), 	ilde{W} is stabilized by End(	ilde{V}_{	ilde{k}}), in particular, by 	ilde{f}. Therefore 	ilde{f}(v_1,...,v_n) = (f(v_1),...,f(v_n)), implying the existence of a desired r in R.

(1.9) Corollary. In the above notation, if V_k is finitely generated, then the natural map 
ho: R 	o End(V_k) is surjective.

Proof. Let v_1, ..., v_n in V be a finite set of generators of V_k. Then for any f in End(V_k), there exists r in R such that rv_i = f(v_i) for all 1 leq i leq n. Therefore rv = sum r(v_ia_i) = sum (rv_i)a_i = sum (f(v_i))a_i = f(sum{v_ia_i}) = f(v) for all v in V. By definition, f = 
ho(r).

Let k be a division ring, and V be a right vector space over k . A subset S subseteq End(V_k) is called m-transitive if for any set of n leq m linearly indepedent vectors v_1, ..., v_n and any other set of n vectors v_1, ..., v_n, there exists  s in S such that s(v_i) = v_i for all 1 leq i leq n. If S is m-transitive for all (finite) m, then S is called a dense set of linear transformations on V_k. The following result ensures that this definition is consistent with the one we introduced.

(1.10) Proposition. Let V be an (R,k)-bimodule, where R is a ring and k is a division ring. Then R acts densely on V_k if and only if 
ho(R) is a dense ring of linear transformations on V, where 
ho: R 	o End(V_k) is the natural ring homomorphism.

Proof. One direction is clear. For the other direction, assume that 
ho(R) is dense, and let f in End(V_k), v_1, ..., v_n in V. After a possible permutation, v_1, ..., v_m are linearly independent over k for some m leq n, and each v_i is a linear combination of v_1, ..., v_m for all m leq i leq n. Since 
ho(R) is m-transitive, there exists r in R such that rv_j = f(v_j) for all 1 leq j leq m. It follows that this is true when j is extended to 1, ..., n.

(1.8), (1.9), and (1.10) yield the following result, which characterizes the structure of left primitive rings completely. (1.11) is sometimes also referred as the Density Theorem.

(1.11) Theorem. Let R be a left primitive ring, V be a faithful simple left R-module, and k be the division ring End(_RV) (by Schurs Lemma). Then R is isomorphic to a dense ring of linear transformations on V_k. Moreover:

  1. If R is left Artinian, then n = dim_k{V} is finite and R cong M_n(k).
  2. If R is not left Artinian, then dim_k{V} is infinite, and for any n > 0 , there exists a subring R_n of R which admits a ring homomorphism onto M_n(k).

1 recovers the Artin-Wedderburns Theorem for left Artinian simple rings, as these are precisely the left Artinian left primitive rings, by a moments thought. A more generalized version of Artin-Wedderburns Theorem will be presented at the end of this article.

Proof. The first conclusion (before "moreover") follows easily from the faithfulness of _RV and the density of 
ho(R). For the second conclusion (after "moreover"), assume that dim_k{V} = n < infty. By (1.9), 
ho(R) 	woheadrightarrow End(V_k). And by faithfulness of _RV, 
ho(R) hookrightarrow End(V_k). In conclusion, R cong M_n(k). Since R is semisimple, R is clearly left Artinian. Next assume that dim_k{V} is infinite, and let v_1, ..., v_n be n linearly independent vectors, V_n = sum_{i = 1}^n{v_ik} for all 1 leq n < infty. Finally, let R_n = {r in R: r(V_n) subseteq V_n}, and mathfrak{A_n} = {r in R: r(V_n) = 0}. Then R_n/mathfrak{A}_n acts faithfully on the k-vector space V_n. By n-transitivity of R, any k-endomorphism of V_n can be realized as the action of some r in R_n. Therefore the natural map R_n/mathfrak{A}_n 	o End((V_n)_k) is an isomorphism, yielding R_n/mathfrak{A}_n cong M_n(k). Furthermore, by (n+1)-transitivity, there exists r in R such that rv_1 = ... = rv_n = 0 but rv_{n+1} 
eq 0. This induces a strictly descending sequence of left ideals mathfrak{A}_1 supset mathfrak{A}_2 supset mathfrak{A}_3 supset ..., so R is not left Artinian.

We shall now present the Jacobson-Hersteins Theorem.

Jacobson-Hersteins Theorem. A ring R is commutative if and only if for any a, b in R , there exists an integer n(a,b) > 1 such that (ab - ba)^{n(a,b)} = ab - ba.

Proof. In this proof, we assume the following theorem, which will be proved later.

(1.12) Theorem. Let R be a division ring such that for any a in R, a^{n(a)} = a for some integer n(a) > 1. Then R is commutative.

  1. (1.12) is true for any left primitive ring R. By (1.11), there exists a division ring k such that either R cong M_n(k) or for any m, there exists a subring R_m subseteq R which admits a ring homomorphism onto M_m(k). However, for m geq 2, M_m(k) cannot satisfy the required property "for any a, b in R , there exists an integer n(a,b) > 1 such that (ab - ba)^{n(a,b)} = ab - ba" by a routine check. Therefore R cong k. (1.12) applied.

  2. (1.12) is true for any semiprimitive ring R. By (1.6), there exists a subdirect produce representation R 	o prod{R_i}, where R_is are left primitive rings. Since each R_i is a homomorphic image of R, R_i is commutative by 1 and (1.12). Since R is isomorphic to a subring of prod{R_i}, R is commutative.
  3. (1.12) is true for any ring R. By 2, R/rad  R is commutative. Therefore for any a, b in R, ab - ba in rad  R. Since (ab - ba)^{n(a,b)} = ab - ba for some n(a,b) > 1, (ab - ba)(1 - (ab - ba)^{n(a,b) - 1}) = 0. However, (1 - (ab - ba)^{n(a,b) - 1}) in 1 + rad  R is a unit, by a moments reflection on (1.2), ab - ba must equal to zero.

Proof of (1.12).

Claim 1. A finite division ring R is a field.

Proof. Let F be the center of R. Then F is finite field or order q for some prime power q. Assume R is not field, or in other words, dim_F{D}  = n > 1. Write down the class equation for the finite group D^{	imes}: |D^{	imes}| = q^n - 1 = q - 1 + sum |D^{	imes}: C(a)^{	imes}|, and r = r(a) = dim_{F}C(a). Then r | n and it is possible to rewrite the class equation: |D^{	imes}| = q^n - 1 = q - 1 + frac{q^n -1}{q^r - 1}. Since r|n, q^n - 1 = Phi_n(q)(q^r - 1)h(q), where Phi_n(q) is the n-th cyclotomic polynomial, and h(q) in mathbb{Z}[q]. The class equation implies that frac{q^n-1}{q^r-1} is divisible by Phi_n(q) and hence q - 1 is divisible by Phi_n(q). In particular, q - 1 geq |Phi_n(q)| = prod|q - zeta|, where zeta ranges over all primitive n-th roots of unity. A contradiction.

Claim 2. A finite subgroup G of a division ring R of characteristic p > 0 is cyclic.

Proof. Let F be the prime field of R, and K = {sum a_ig_i: a_i in F, g_i in G}. Then K is a finite subring of R, which is a field by Claim 1. Since G is a subgroup of K^{	imes}, G is cyclic.

Claim 3. If all additive commutators are central in a division ring R, then R is a field.

Proof. It suffices to show that if y in R commutes with all additive commutators, then y in Z(R). If y 
otin Z(D), xy 
eq yx for some x in D. Consider x(xy) - (xy)x = x(xy - yx). Since y commutes with x(xy) - (xy)x and xy - yx 
eq 0, y commutes with x, a contradiction.

Claim 4. Let R be a division ring of characteristic p > 0. If a in R^{	imes} is noncentral and torsion, then there exists y in R^{	imes} such that yay^{-1} = a^i 
eq a for some i > 0. Moreover, y can be chosen to be an additive commutator in R.

Proof. This proof is removed from this article because of its length and my laziness.

By the assumption in (1.12), any nonzero additive commutator has finite order in the multiplicative group D^{	imes}. If R 
eq F = Z(R), then there exists some additive commutator a = bb - bb 
otin F by Claim 3. For any c in F^{	imes}, ca = (cb)b  - b(cb) is also a nonzero additive commutator. Since a and ca both have finite order, there exists an integer k > 0 such that 1 = a^k = (ca)^k = c^ka^k. This implies that char F = char R > 0 . Also, since a is non central and torsion, Claim 4 yields an additive commutator d in R^{	imes} such that dad^{-1} = a^i 
eq a, where i > 0. As y normalizes langle a 
angle, langle a 
angle cdot langle y 
angle is a finite subgroup of R^{	imes}. By Claim 2, langle a 
angle cdot langle y 
angle is commutative, a contradiction.

Proof of Artin-Wedderburns Theorem.

Statement. Let R be a semisimple ring. Then R cong M_{n_1}(D_1) 	imes ... 	imes M_{n_r}(D_r) for some division rings D_i and natural numbers n_i. Moreover, (n_1,D_1),...,(n_r,D_r) is uniquely determined, up to permutation.

Proof. Decompose _RR into a finite direct sum of minimal left ideals and group the minimal left ideals according to isomorphism types. Then _RR cong n_1V_1 oplus ... oplus n_rV_r, whereV_1, ..., V_r are mutually nonisormophic simple left R-modules. Since R-endomorphism of _RR is determined by right multiplication by elements of R, R cong End(_RR). Therefore R cong End(n_1V_1 oplus ... oplus n_rV_r) = End(n_1V_1) 	imes End(n_rV_r). By Schurs Lemma, each D_i = End(V_i) is a division ring. And by a routine verification, eachEnd(n_iV_i) cong M_{n_i}(D_i). In conclusion, we have the desired decomposition R cong M_{n_1}(D_1) 	imes ... 	imes M_{n_r}(D_r). The uniqueness of this decomposition follows easily from Jordan Holders Theorem.

Reference.

"A First Course in Noncommutative Rings", 2nd edition, T.Y. Lam.

推薦閱讀:

配極理論與調和點列
【解析幾何】雙聯立(齊次化處理)解決定點問題
【不等式】丟掉次數:伯努利不等式及其應用
二重積分
Pell方程

TAG:代數 | 抽象代數 | 數學 |