Archive for the ‘Noncommutative Ring Theory Notes’ Category

For a ring R with identity, we denote by J(R), U(R), M_n(R), GL(n,R), the Jacobson radical of R, the group of units of R, the ring of n \times n matrices with entries from R, and the group of units of M_n(R), respectively.

Here we proved this basic result that if F is a finite field, then

|\text{GL}(n,F)|=\prod_{i=1}^n(|F|^n-|F|^{i-1}).

In the same post, we also showed that if R is a finite ring with identity, then

|U(R)|=|J(R)||U(R/J(R))|. \ \ \ \ \ \ \ \ \ \ (1)

We are now going to use (1) to find |\text{GL}(n,R)| for any finite commutative ring with identity R.

Theorem. Let R be a finite commutative ring with identity, and let \{ \mathfrak{m}_1, \cdots , \mathfrak{m}_k\} be the set of all maximal ideals of R. Let F_j:=R/\mathfrak{m}_j, \ 1 \le j \le k. Then

\displaystyle |\text{GL}(n,R)|=|J(R)|^{n^2}\prod_{j=1}^k\prod_{i=1}^n(|F_j|^n-|F_j|^{i-1}).

Proof (Y. Sharifi). Let \mathfrak{a}:=J(R). By the Chinese Remainder Theorem,

\displaystyle R/\mathfrak{a} =R/\bigcap_{j=1}^k\mathfrak{m}_j \cong \bigoplus_{j=1}^kR/\mathfrak{m}_j=\bigoplus_{j=1}^kF_j

and hence

\displaystyle M_n(R)/M_n(\mathfrak{a}) \cong M_n(R/\mathfrak{a}) \cong M_n\left(\bigoplus_{j=1}^kF_j\right) \cong \bigoplus_{j=1}^kM_n(F_j).

The above ring isomorphism gives the group isomorphism

\displaystyle U(M_n(R)/M_n(\mathfrak{a})) \cong U\left(\bigoplus_{j=1}^kM_n(F_j)\right)=\bigoplus_{j=1}^kU(M_n(F_j))=\bigoplus_{j=1}^k\text{GL}(n,F_j)

and so

\displaystyle |U(M_n(R)/M_n(\mathfrak{a}))|=\prod_{j=1}^k|\text{GL}(n,F_j)|=\prod_{j=1}^k\prod_{i=1}^n(|F_j|^n-|F_j|^i). \ \ \ \ \ \ \ \ \ \ (2)

On the other hand, by Example 8 in this post, M_n(\mathfrak{a})=J(M_n(R)) and so, by (1) applied to M_n(R) and also by (2),

\displaystyle |U(M_n(R))|=|M_n(\mathfrak{a})||U(M_n(R)/M_n(\mathfrak{a}))|=|\mathfrak{a}|^{n^2}\prod_{j=1}^k\prod_{i=1}^n(|F_j|^n-|F_j|^i). \ \Box

Example. Let m=\prod_{j=1}^kp_j^{r_j} be the prime factorization of an integer m \ge 2. Then

\displaystyle |\text{GL}(n, \mathbb{Z}/m\mathbb{Z})|=\prod_{j=1}^kp_j^{(r_j-1)n^2}(p_j^n-1)(p_j^n-p_j) \cdots (p_j^n-p_j^{n-1}).

Proof. Let R:=\mathbb{Z}/m\mathbb{Z}. The maximal ideals of R are \mathfrak{m}_j=p_j\mathbb{Z}/m\mathbb{Z}, \ 1 \le j \le k, and

J(R)=p_1p_2 \cdots p_k\mathbb{Z}/m\mathbb{Z}.

Therefore |J(R)|=\prod_{j=1}^kp_j^{r_j-1}. Also F_j:=R/\mathfrak{m}_j \cong \mathbb{Z}/p_j\mathbb{Z} and so |F_j|=p_j. The result now follows from the Theorem. \ \Box

Note. The above Example was posted on the Art of Problem Solving website as a problem yesterday; you can see the problem and my solution here.

For a division ring D, we denote by Z(D) and D^{\times} the center and the multiplicative group of D, respectively.

Let D_1 be a division ring, and suppose that D_2 is a proper subdivision ring of D_1, i.e. D_2 is a subring of D_1, \ D_2 \ne D_1, and D_2 itself is a division ring. Then D_2^{\times} is clearly a proper subgroup of D_1^{\times}. Now, one may ask: when exactly is D_2^{\times} a normal subgroup of D_1^{\times} ? The Cartan-Brauer-Hua theorem gives the answer: D_2^{\times} is a normal subgroup of D_1^{\times} if and only if D_2 \subseteq Z(D_1). In particular, D_2 must be a field. One side of the theorem is trivial: if D_2 \subseteq Z(D_1), then D_2^{\times} is obviously normal in D_1^{\times}. The other side of the theorem is not trivial, but it’s not hard to prove either. The proof is a quick result of the following simple yet strangely significant identity!

Hua’s Identity (Loo Keng Hua, 1949). Let D be a division ring, and let a,b \in D such that ab \ne ba. Then

a=(b^{-1}-(a-1)^{-1}b^{-1}(a-1))(a^{-1}b^{-1}a-(a-1)^{-1}b^{-1}(a-1))^{-1}.

Proof. First see that since ab \ne ba, the four elements a,a-1,b, and a^{-1}b^{-1}a-(a-1)^{-1}b^{-1}(a-1) are all nonzero hence invertible. Now,

a(a^{-1}b^{-1}a-(a-1)^{-1}b^{-1}(a-1))=b^{-1}a-a(a-1)^{-1}b^{-1}(a-1)

=b^{-1}a-(a-1+1)(a-1)^{-1}b^{-1}(a-1)=b^{-1}a-b^{-1}(a-1)-(a-1)^{-1}b^{-1}(a-1)

=b^{-1}-(a-1)^{-1}b^{-1}(a-1). \ \Box

Cartan-Brauer-Hua Theorem. Let D_1 be a division ring, and let D_2 be a proper subdivision ring of D_1. If D_2^{\times} is normal in D_1^{\times}, then D_2 \subseteq Z(D_1).

Proof. Suppose, to the contrary, that D_2 \nsubseteq Z(D_1). Let a \in D_1, b \in D_2 such that ab \ne ba. Then, since D_2^{\times} is a normal subgroup of D_1^{\times}, all the elements

b^{-1}, \ (a-1)^{-1}b^{-1}(a-1), \ a^{-1}b^{-1}a, \ (a-1)^{-1}b^{-1}(a-1)

are in D_2^{\times} and hence, by Hua’s identity, a \in D_2. So every element of D_1 \setminus D_2 commutes with b. Now let c \in D_1 \setminus D_2. Then ac \in D_1 \setminus D_2, because a \in D_2, and therefore both c,ac commute with b. But then a=acc^{-1} will also commute with b and that’s a contradiction. \ \Box

Note. There are other proofs of the Theorem; for example, here is a simple but not very well-known one.

We remarked here that the composition of derivations of a ring need not be a derivation. In this post, we prove this simple yet interesting result that if R is a prime ring of characteristic \ne 2 and if \delta_1,\delta_2 are nonzero derivations of R, then \delta_1\delta_2 can never be a derivation of R. But before getting into the proof of that, let me remind the reader of a little fact that tends to bug many students.

nTorsion Free vs Characteristic \ne n. Let R be a ring, and n \ge 2 an integer. Recall that we say that R is ntorsion free if a \in R, na=0 implies a=0. We say that \text{char}(R)=n if n is the smallest positive integer such that na=0 for all a \in R. It is clear that if R is n-torsion free, then \text{char}(R) \ne n. The simple point I’d like to make here is that the converse is not always true. That will make a lot more sense if you look at its contrapositive, in fact the contrapositive of a stronger statement: na=0 for some 0 \ne a \in R does not always imply that nR=(0). However, the converse is true if R is prime. First, an example to show that the converse in not always true.

Example 1. Consider the ring R=\mathbb{Z}_n \oplus \mathbb{Z}, and a=(1,0) \in R. Then a \ne 0 and na=0. So R is not n-torsion free. but \text{char}(R)=0 \ne n.

Now let’s show that the converse is true if R is prime.

Example 2. Let n \ge 2 be an integer and R a prime ring. Suppose that na=0 for some 0 \ne a \in R. Then nR=(0), i.e. nr=0 for all r \in R.

Proof. We have (0)=(na)Rr=aR(nr) and so, since a \ne 0 and R is prime, nr=0. \ \Box

Let’s now get to the subject of this post.

Lemma 1. Let R be a ring, and let \delta_1,\delta_2 be derivations of R. Then \delta_1\delta_2 is a derivation of R if and only if

\delta_1(a)b\delta_2(c)+\delta_2(a)b\delta_1(c)=0,

for all a,b,c \in R.

Proof. Since \delta_1\delta_2 is clearly additive, it is a derivation if and only if it satisfies the product rule, i.e.

\delta_1\delta_2(bc)=\delta_1\delta_2(b)c+b\delta_1\delta_2(c). \ \ \ \ \ \ \ (1)

On the other hand, since \delta_1,\delta_2 are derivations of R, we also have

\delta_1\delta_2(bc)=\delta_1(\delta_2(b)c+b\delta_2(c))=\delta_1(\delta_2(b)c)+\delta_1(b\delta_2(c))

=\delta_1\delta_2(b)c+\delta_2(b)\delta_1(c)+\delta_1(b)\delta_2(c)+b\delta_1\delta_2(c). \ \ \ \ \ \ \ (2)

So we get from (1),(2) that

\delta_1(b)\delta_2(c)+\delta_2(b)\delta_1(c)=0. \ \ \ \ \ \ \ \ (3)

Replacing b by ab in (3) gives

0=\delta_1(ab)\delta_2(c)+\delta_2(ab)\delta_1(c)=(\delta_1(a)b+a\delta_1(b))\delta_2(c)+(\delta_2(a)b+a\delta_2(b))\delta_1(c)

=\delta_1(a)b\delta_2(c)+\delta_2(a)b\delta_1(c)+a(\delta_1(b)\delta_2(c)+\delta_2(b)\delta_1(c))

=\delta_1(a)b\delta_2(c)+\delta_2(a)b\delta_1(c), \ \ \ \ \ \ \ \text{by} \ (3). \ \Box

Corollary. Let R be a 2-torsion free semiprime ring, and let \delta be a derivation of R. Then \delta^2 is a derivation of R if and only if \delta=0.

Proof. Suppose that \delta^2 is a derivation of R and let a \in R. Then choosing \delta_1=\delta_2=\delta and c=a in Lemma 1 gives 2\delta(a)b\delta(a)=0, for all b \in R. So, since R is 2-torsion free, \delta(a)b\delta(a)=0, for all b \in R. Thus \delta(a)R\delta(a)=(0) and hence \delta(a)=0, because R is semiprime. \ \Box

Lemma 2. Let R be a prime ring, and let \delta be a derivation of R such that \delta(a)b=0 for all a \in R and some b \in R. Then either b=0 or \delta=0.

Proof. Since \delta(a)b=0 for all a \in R, we have \delta(ca)b=0 for all a,c \in R and so

0=\delta(ca)b=(\delta(c)a+c\delta(a))b=\delta(c)ab+c\delta(a)b=\delta(c)ab.

So \delta(c)Rb=(0) for all c \in R and hence, since R is prime, either b=0 or \delta(c)=0 for all c \in R. \ \Box

Remark. Lemma 2 remains true if we replace the condition \delta(a)b=0 by b\delta(a)=0. The proof is similar, just this time replace a by ac.

Theorem (Edward C. Posner, 1957). Let R be a prime ring of characteristic \ne 2, and let \delta_1, \delta_2 be derivations of R. Then \delta_1\delta_2 is a derivation of R if and only if \delta_1=0 or \delta_2=0.

Proof. First note that, by Example 2, the condition \text{char}(R) \ne 2 is the same as saying that R is 2-torsion free. Now, suppose that \delta_1\delta_2 is a derivation of R and let a,b,c \in R. Applying Lemma 1 to a, \delta_2(c)b, c gives

\delta_1(a)\delta_2(c)b\delta_2(c)+\delta_2(a)\delta_2(c)b\delta_1(c)=0.

But by the identity (3) in Lemma 1, \delta_1(a)\delta_2(c)=-\delta_2(a)\delta_1(c) and so the above becomes

\delta_2(a)(\delta_2(c)b\delta_1(c)-\delta_1(c)b\delta_2(c))=0.

Thus, by Lemma 2, either \delta_2=0 or \delta_2(c)b\delta_1(c)-\delta_1(c)b\delta_2(c)=0. If \delta_2=0, we are done. So suppose that

\delta_2(c)b\delta_1(c)-\delta_1(c)b\delta_2(c)=0.

Adding the above identity to the identity \delta_1(c)b\delta_2(c)+\delta_2(c)b\delta_1(c)=0, which holds by Lemma 1, gives 2\delta_2(c)b\delta_1(c)=0. Hence \delta_2(c)b\delta_1(c)=0 and so \delta_2(c)R\delta_1(c)=(0). Thus, since R is prime, either \delta_1=0 or \delta_2=0. \ \Box

Example 3. The condition \text{char}(R) \ne 2 cannot be removed from the Theorem. Consider the polynomial ring R:=\mathbb{Z}_2[x], and the derivation \delta:=\frac{d}{dx}. Then \delta \ne 0 but \delta^2=0 is a derivation.

Note. Examples and the Corollary in this post are mine. I have also slightly simplified Posner’s proof of the Theorem.

All rings in this post are commutative with identity. For the basics on derivations of rings see this post and this post.

Let \delta be a derivation of a ring R. Since \delta is additive, \delta(na)=n\delta(a) for all integers n and all a \in R. So every derivation of a ring is a \mathbb{Z}-derivation. If R is \mathbb{Q}-algebra and r=\frac{m}{n} \in \mathbb{Q}, where m,n are integers and n \ne 0, then n\delta(ra)=\delta(nra)=\delta(ma)=m\delta(a) and so \delta(ra)=r\delta(a). Thus every derivation of a \mathbb{Q}-algebra is a \mathbb{Q}-derivation.

Definition. We say that a derivation \delta of a ring R is locally nilpotent if for every a \in R, there exists a positive integer n such that \delta^n(a)=0.

Example. Let R be the polynomial ring \mathbb{C}[x]. Then the derivation \delta:=\frac{d}{dx} is locally nilpotent because if p(x) \in R has degree n, then \delta^{n+1}(p(x))=0.

The following Theorem characterizes all \mathbb{Q}-algebras R for which there exists a locally nilpotent derivation \delta and a \in R such that \delta(a)=1. The polynomial ring in the above Example gives one of those algebras since, in that example, \delta(x)=1. It turns out that any such algebra is a polynomial ring!

Theorem. Let \delta be a locally nilpotent derivation of a \mathbb{Q}-algebra R, and let K:=\ker \delta. If there exists x \in R such that \delta(x)=1, then x is transcendental over K and R=K[x].

Proof. Suppose, to the contrary, that x is algebraic over K, and let n be the smallest positive integer such that \alpha_nx^n+\alpha_{n-1}x^{n-1}+ \cdots + \alpha_1x+\alpha_0=0 for some \alpha_i \in K, \ \alpha_n \ne 0. Then, since by the product rule, \delta(x^i)=ix^{i-1}\delta(x)=ix^{i-1}, we have

0=\delta( \alpha_nx^n+\alpha_{n-1}x^{n-1}+ \cdots + \alpha_1x+\alpha_0)= \alpha_n\delta(x^n)+\alpha_{n-1}\delta(x^{n-1})+ \cdots + \alpha_1\delta(x)

=n\alpha_n x^{n-1}+(n-1)\alpha_{n-1}x^{n-1}+ \cdots + \alpha_1,

which contradicts the minimality of n. So x is transcendental over K. We now show that R=K[x]. For a \in R, let \nu(a) be the smallest positive integer m such that \delta^m(a)=0. The proof is by induction over \nu(a), \ a \in R. If \nu(a)=1, then \delta(a)=0 and so a \in K \subset K[x]. Suppose now that a \in R and m:=\nu(a) \ge 2. Let

\displaystyle y:=\sum_{n=0}^{m-1}\frac{(-1)^n\delta^n(a)x^n}{n!}=a-bx,

where

\displaystyle b=\sum_{n=1}^{m-1}\frac{(-1)^{n-1}\delta^n(a)x^{n-1}}{n!}.

So a=y+bx and so we are done if we prove that b,y \in K[x].

Claim 1: y \in K.

Proof. We have

\displaystyle \delta(y)=\sum_{n=0}^{m-1}\frac{(-1)^n}{n!}\delta(\delta^n(a)x^n)=\sum_{n=0}^{m-1}\frac{(-1)^n}{n!}(\delta^{n+1}(a)x^n+\delta^n(a)\delta(x^n))

\displaystyle =\sum_{n=0}^{m-1}\frac{(-1)^n}{n!}(\delta^{n+1}(a)x^n+n\delta^n(a)x^{n-1})=\sum_{n=0}^{m-1}\frac{(-1)^n\delta^{n+1}(a)x^n}{n!}+\sum_{n=1}^{m-1}\frac{(-1)^n\delta^n(a)x^{n-1}}{(n-1)!}

\displaystyle =\sum_{n=0}^{m-2}\frac{(-1)^n\delta^{n+1}(a)x^n}{n!}+\sum_{n=0}^{m-2}\frac{(-1)^{n+1}\delta^{n+1}(a)x^n}{n!}=0

and so y \in K.

Claim 2: b \in K[x].

Proof. By the Leibniz formula,

\displaystyle \delta^{m-1}(b)=\sum_{n=1}^{m-1}\frac{(-1)^{n-1}}{n!}\delta^{m-1}(\delta^n(a)x^{n-1})=\sum_{n=1}^{m-1}\frac{(-1)^{n-1}}{n!}\sum_{k=0}^{m-1}\binom{m-1}{k}\delta^{n+k}(a)\delta^{m-k-1}(x^{n-1}).

Now notice that \delta^{n+k}(a)\delta^{m-k-1}(x^{n-1})=0 for all k,n because if n+k \ge m, then \delta^{n+k}(a)=0, and if n+k < m, then \delta^{m-k-1}(x^{n-1})=0. Thus \delta^{m-1}(b)=0 and so \nu(b) < \nu(a)=m. Hence, by our induction hypothesis, b \in K[x]. \ \Box

Exercise. Let \delta be a locally nilpotent derivation of a \mathbb{Q}-algebra R, and let c \in R. Let K:=\ker \delta, and define the map f: R \to R by

\displaystyle f(a):=\sum_{n=0}^{\infty}\frac{\delta^n(a)c^n}{n!},

for all a \in R. Show that f is a K-algebra homomorphism.

Note. The above Theorem is Theorem 2.8 in here. The proof I’ve given is essentially the same as the proof given in there; I just made the proof easier to follow.

See the first part of this post here. All rings in this post are assumed to have identity.

In the first part, we gave two examples of derivations on rings. We now give a couple of ways to make new derivations using given ones.

Example 1. Let R be a ring, and let c_1,c_2 \in Z(R), the center of R. If \delta_1,\delta_2 \in \text{Der}(R), then

i) c_1\delta_1+c_2\delta_2 \in \text{Der}(R),

ii) \delta_1\delta_2-\delta_2\delta_1 \in \text{Der}(R).

Proof. The first part is quite straightforward and so I’m going to leave it as an easy exercise. For the second part, we need to show that \delta_1\delta_2-\delta_2\delta_1 is additive and satisfies the product rule. So let a,b \in R. Then

\begin{aligned} (\delta_1\delta_2-\delta_2\delta_1)(a+b)=\delta_1\delta_2(a+b)-\delta_2\delta_1(a+b)= \delta_1(\delta_2(a)+\delta_2(b))-\delta_2(\delta_1(a)+\delta_1(b))\end{aligned}

=\delta_1\delta_2(a)+\delta_1\delta_2(b)-\delta_2\delta_1(a)-\delta_2\delta_1(b)=(\delta_1\delta_2-\delta_2\delta_1)(a)+(\delta_1\delta_2-\delta_2\delta_1)(b),

which proves that \delta_1\delta_2-\delta_2\delta_1 is additive. Now, for the product rule, we have

\begin{aligned}(\delta_1\delta_2-\delta_2\delta_1)(ab)=\delta_1(\delta_2(ab))-\delta_2(\delta_1(ab))=\delta_1(\delta_2(a)b+a\delta_2(b))-\delta_2(\delta_1(a)b+a\delta_1(b))\end{aligned}

=\delta_1(\delta_2(a)b)+\delta_1(a\delta_2(b))-\delta_2(\delta_1(a)b)-\delta_2(a\delta_1(b))

\begin{aligned} =\delta_1\delta_2(a)b+\delta_2(a)\delta_1(b)+\delta_1(a)\delta_2(b)+a\delta_1\delta_2(b)-\delta_2\delta_1(a)b-\delta_1(a)\delta_2(b)-\delta_2(a)\delta_1(b)-a\delta_2\delta_1(b)\end{aligned}

=\delta_1\delta_2(a)b+a\delta_1\delta_2(b)-\delta_2\delta_1(a)b-a\delta_2\delta_1(b)=(\delta_1\delta_2-\delta_2\delta_1)(a)b+a(\delta_1\delta_2-\delta_2\delta_1)(b),

proving that \delta_1\delta_2-\delta_2\delta_1 satisfies the product rule. \ \Box

Remark. By Example 1, ii), the commutator of two derivations is a derivation. However, the composition of two derivations need not be a derivation. For example, let R be the polynomial ring \mathbb{C}[x] and consider the derivation \delta:=\frac{d}{dx}. Then the second derivative, i.e. \delta^2, is not a derivation because if it was, then we would have \delta^2(x^2)=2x\delta^2(x)=0 but we know that \delta^2(x^2)=2.

Example 2. Let M_n(R) be the ring of n \times n matrices with entries from a ring R. Let \delta \in \text{Der}(R). Define the map \Delta: M_n(R) \to M_n(R) as follows: for any A=[a_{ij}] \in M_n(R), where a_{ij} is the (i,j)-entry of A, define \Delta(A)=[\delta(a_{ij})]. Then \Delta \in \text{Der}(M_n(R)).

Proof. Since \delta is additive, \Delta is clearly additive too. So we only need to show that \Delta satisfies the product rule. Let A=[a_{ij}], \ B=[b_{ij}] \in M_n(R). Let c_{ij},d_{ij} be, respectively, the (i,j)-entry of AB and the (i,j)-entry of \Delta(AB). Then c_{ij}=\sum_{k=1}^na_{ik}b_{kj} and so

\displaystyle \begin{aligned}  d_{ij}=\delta(c_{ij})=\sum_{k=1}^n\delta(a_{ik}b_{kj})=\sum_{k=1}^n(\delta(a_{ik})b_{kj}+a_{ik}\delta(b_{kj}))=\sum_{k=1}^n\delta(a_{ik})b_{kj}+\sum_{k=1}^na_{ik}\delta(b_{kj})\end{aligned},

which is the (i,j)-entry of \Delta(A)B+A\Delta(B). So \Delta(AB)= \Delta(A)B+A\Delta(B) hence the result. \ \Box

Next is to prove a familiar formula from Calculus, i.e. the Leibniz formula for the nth derivative of the product of two elements.

The Leibniz Formula. Let R be a ring, a,b \in R, and \delta \in \text{Der}(R). Then

\displaystyle \delta^n(ab)=\sum_{k=0}^n\binom{n}{k}\delta^k(a)\delta^{n-k}(b),

for all integers n \ge 0. Here \delta^0 is defined to be the identity map.

Proof. The proof is by induction over n. There is nothing to prove for n=0. Now, assuming that the formula holds for n, we have

\displaystyle \delta^{n+1}(ab)=\delta(\delta^n(ab))=\delta \left( \sum_{k=0}^n\binom{n}{k}\delta^k(a)\delta^{n-k}(b)\right)=\sum_{k=0}^n\binom{n}{k}\delta(\delta^k(a)\delta^{n-k}(b))

\displaystyle =\sum_{k=0}^n\binom{n}{k}(\delta^{k+1}(a)\delta^{n-k}(b)+\delta^k(a)\delta^{n-k+1}(b)), \ \ \ \ \ \ \text{by the product rule}

\displaystyle =\sum_{k=0}^n\binom{n}{k}\delta^{k+1}(a)\delta^{n-k}(b)+\sum_{k=0}^n\binom{n}{k}\delta^k(a)\delta^{n-k+1}(b)

\displaystyle =\sum_{k=1}^{n+1}\binom{n}{k-1}\delta^k(a)\delta^{n-k+1}(b)+\sum_{k=0}^n\binom{n}{k}\delta^k(a)\delta^{n-k+1}(b)

\displaystyle =\sum_{k=1}^{n+1}\binom{n}{k-1}\delta^k(a)\delta^{n-k+1}(b)+\sum_{k=1}^{n+1}\binom{n}{k}\delta^k(a)\delta^{n-k+1}(b)+a\delta^{n+1}(b)

\displaystyle =\sum_{k=1}^{n+1}\left[\binom{n}{k-1}+\binom{n}{k}\right]\delta^k(a)\delta^{n-k+1}(b)+a\delta^{n+1}(b)

\displaystyle =\sum_{k=1}^{n+1}\binom{n+1}{k}\delta^k(a)\delta^{n-k+1}(b)+a\delta^{n+1}(b)=\sum_{k=0}^{n+1}\binom{n+1}{k}\delta^k(a)\delta^{n-k+1}(b),

which proves the formula for n+1. \ \Box

Let me end this post with an example which is Remark 1.6.30 in Louis Rowen’s book Ring Theory (volume 1).

Example 3. Let R be a ring, and \delta \in \text{Der}(R). In the first part of this post, we showed that \ker \delta is a subring of R. Now, for any integer n \ge 0, let

K_n:=\ker \delta^n=\{a \in R: \ \delta^{n+1}(a)=0\}.

So \ker \delta =K_0. Then K:=\bigcup_{k=0}^{\infty}K_n is a subring of R and K_mK_n \subseteq K_{m+n}, for all m,n.

Proof. It is clear that (K_n,+) is an abelian group for all n, and K_0 \subseteq K_1 \subseteq \cdots. So (K,+) is an abelian group, and so we only need to show that K_mK_n \subseteq K_{m+n} for all m,n. Let a \in K_m and b \in K_n. We want to show that ab \in K_{m+n}. So \delta^{m+1}(a)=\delta^{n+1}(b)=0. We also have, by Leibniz formula,

\displaystyle \delta^{m+n+1}(ab)=\sum_{k=0}^{m+n+1}\binom{m+n+1}{k}\delta^k(a)\delta^{m+n+1-k}(b).

Now, if k \ge m+1, then \delta^k(a)=0 and if k \le m, then m+n+1-k \ge n+1 and therefore \delta^{m+n+1-k}(b)=0. Thus \delta^{m+n+1}(ab)=0 and so ab \in K_{m+n}. \ \Box

All rings in this post are assumed to have identity and Z(R) denotes the center of a ring R.

Those who have the patience to follow my posts have seen derivation on rings several times but they have not seen a post exclusively about the basics on derivations of rings because, strangely, that post didn’t exist until now.

In the polynomial ring \mathbb{C}[x], we have the familiar concept of differentiation with respect to x, i.e. \delta:= \frac{d}{dx}. This is a map form \mathbb{C}[x] \to \mathbb{C}[x] which is additive and satisfies the product rule. Also, \delta(c)=0 for all c \in \mathbb{C}. We now extend this concept to rings in general.

Definition 1. Let R be a ring. A derivation of R is any additive map \delta : R \to R that satisfies the product rule: \delta(ab)=\delta(a)b+a\delta(b), for all a,b \in R. We denote by \text{Der}(R) the set of all derivations of R.

We can now extend the familiar concept of differentiating polynomials to differentiating polynomials over any ring R. Note that the variable x is assumed to be in the center of R[x].

Example 1. Let R be a ring, and let R[x] be the ring of polynomials over R. Define the map \delta : R[x] \to R[x] by

\displaystyle \delta\left(\sum_{k=0}^na_kx^k\right)=\sum_{k=0}^nka_kx^{k-1}, \ \ \ \ \ n \in \mathbb{Z}_{\ge 0}, \ \ a_k \in R.

Then \delta \in \text{Der}(R[x]). The derivation \delta is usually denoted by \frac{d}{dx}. Note that, for the definition to make sense, the term ka_kx^{k-1} for k=0 is defined to be 0 and x^0 is defined to be 1.

Proof. We need to show that \delta is additive and also it satisfies the product rule. Let p(x), q(x) \in R[x]. So we can write p(x)=\sum_{k=0}^na_kx^k, \ q(x)=\sum_{k=0}^nb_kx^k. Then

\displaystyle \delta(p(x)+q(x))=\delta \left(\sum_{k=0}^n(a_k+b_k)x^k\right)=\sum_{k=0}^nk(a_k+b_k)x^{k-1}

\displaystyle =\sum_{k=0}^nka_kx^{k-1}+\sum_{k=0}^nkb_kx^{k-1}=\delta(p(x))+\delta(q(x)).

To prove the product rule, we have

\displaystyle \delta(p(x)q(x))=\delta \left(\sum_{k=0}^{2n}\sum_{j=0}^ka_jb_{k-j}x^k\right)=\sum_{k=0}^{2n}k\sum_{j=0}^ka_jb_{k-j}x^{k-1}, \ \ \ \ \ \ \ \ \ (*)

and

\displaystyle x(\delta(p(x))q(x)+p(x)\delta(q(x))=x\left(\sum_{k=0}^nka_kx^{k-1}\sum_{k=0}^nb_kx^k+\sum_{k=0}^na_kx^k\sum_{k=0}^nkb_kx^{k-1}\right)

\displaystyle \begin{aligned}=\sum_{k=0}^nka_kx^k\sum_{k=0}^nb_kx^k+\sum_{k=0}^na_kx^k\sum_{k=0}^nkb_kx^k=\sum_{k=0}^{2n}\sum_{j=0}^kja_jb_{k-j}x^k+\sum_{k=0}^{2n}\sum_{j=0}^k(k-j)a_jb_{k-j}x^k\end{aligned}

\displaystyle =\sum_{k=0}^{2n}k\sum_{j=0}^ka_jb_{k-j}x^k=x \delta(p(x)q(x)), \ \ \ \ \ \text{by} \ (*).

Therefore \delta(p(x)q(x))=\delta(p(x))q(x)+p(x)\delta(q(x)). \ \Box

Example 2. Let R be a ring, and let r \in R. Define the map \delta : R \to R by \delta(a)=ra-ar, for all a \in R. Then \delta \in \text{Der}(R) and it’s called an inner derivation.

Proof. We need to show that \delta is additive and also it satisfies the product rule. Let a,b \in R. Then

\delta(a+b)=r(a+b)-(a+b)r=ra-ar+rb-br=\delta(a)+\delta(b),

and

\begin{aligned} \delta(a)b+a\delta(b)=(ra-ar)b+a(rb-br)=rab-arb+arb-abr=rab-abr=\delta(ab). \ \Box \end{aligned}

Remarks. Let R be a ring, and let \delta \in \text{Der}(R). Let K:=\ker \delta=\{a \in R: \ \delta(a)=0\}.

i) \{0,1\} \subseteq K,

ii) K is a subring of R called the ring of constants of \delta,

iii) if a \in K is invertible in R, then a^{-1} \in K; so if R is a division ring, K is a division ring too,

iv) if r \in R, then \delta(ra)=r\delta(a) for all a \in R if and only if r \in K,

v) \delta(Z(R)) \subseteq Z(R),

vi) if \delta(a) \in Z(R), then \delta(a^n)=na^{n-1}\delta(a) for all integers n \ge 1, where, for n=1, we define a^0=1,

vii) The identity given in vi) may not hold true if \delta(a) \notin Z(R).

Proof. i) Since \delta is additive, \delta(0)=\delta(0+0)=\delta(0)+\delta(0)=2\delta(0), and so \delta(0)=0, hence 0 \in K. Also, by product rule, \delta(1)=\delta(1 \cdot 1)=\delta(1) \cdot 1 + 1 \cdot \delta(1)=2\delta(1), and so \delta(1)=0 hence 1 \in K.

ii) If a,b \in \ker \delta, then since \delta is additive, \delta(a+b)=\delta(a)+\delta(b)=0+0=0 and so a+b \in \ker \delta. Also, by product rule, \delta(ab)=\delta(a)b+a\delta(b)=0+0=0 and so ab \in \ker \delta.

iii) By i) and product rule, 0=\delta(1)=\delta(aa^{-1})=\delta(a)a^{-1}+a\delta(a^{-1})=a\delta(a^{-1}) and so \delta(a^{-1}) =0.

iv) If \delta(ra)=r\delta(a), for some r \in R and all a \in R, then a=1 gives \delta(r)=r\delta(1)=0 and so r \in K. Conversely, if r \in K, \ a \in R, then, by product rule, \delta(ra)=\delta(r)a+r\delta(a)=0+r\delta(a)=r\delta(a).

v) Let a \in Z(R), \ b \in R. Then

\delta(a)b+a\delta(b)=\delta(ab)=\delta(ba)=\delta(b)a+b\delta(a)=a\delta(b)+b\delta(a),

and so \delta(a)b=b\delta(a), proving that \delta(a) \in Z(R).

vi) First notice that, by v), the condition \delta(a) \in Z(R) is stronger than a \in Z(R). Now, the proof is by induction over n. It is clear for n=1. Assuming the identity holds for n, we have

\begin{aligned} \delta(a^{n+1})=\delta(a^na)=\delta(a^n)a+a^n\delta(a)=na^{n-1}\delta(a)a+a^n\delta(a)=na^n\delta(a)+a^n\delta(a)=(n+1)a^n\delta(a).\end{aligned}

vii) Consider R:=M_2(\mathbb{C}), the ring of 2 \times 2 matrices with complex entries. Let \delta be the inner derivation (see Example 2) of R corresponding to r=\begin{pmatrix}1 & 0 \\ 0 & 0\end{pmatrix}. Now, choose a=\begin{pmatrix}0 & 1\\ 1 & 0\end{pmatrix}. Then a^2 is the identity matrix and so \delta(a^2)=0. Also,

\delta(a)=ra-ar=\begin{pmatrix}0 & 1 \\ -1 & 0\end{pmatrix} \notin Z(R), \ \ \  2a\delta(a) =\begin{pmatrix}-2 & 0 \\ 0 & 2\end{pmatrix} \ne 0=\delta(a^2). \ \Box

Definition 2. Let R be a ring, \delta \in \text{Der}(R), and C a subring of R contained in the center of R. If C \subseteq \ker \delta, then, by Remark iv), \delta(ca+b)=c\delta(a)+\delta(b) for all c \in C,a,b \in R, i.e. \delta is C-linear. The set of all C-linear derivations of R is denoted by \text{Der}_C(R).

Exercise. Consider the commutative polynomial ring C]x] and let R:=C[x]/I, where I:=\langle x^2\rangle. Define the map \delta: R \to R by \delta(\alpha x+ \beta+I)=\alpha x + I, for all \alpha, \beta  \in C. Show that \delta \in \text{Der}_C(R).

In part two of this post, we will learn more about derivations on rings. Happy \pi day!

Rings in this post may or may not have identity. As always, the ring of n \times n matrices with entries from a ring R is denoted by M_n(R).

PI rings are arguably the most important generalization of commutative rings. This post is the first part of a series of posts about this fascinating class of rings.

Let C be a commutative ring with identity, and let C\langle x_1, \ldots ,x_n\rangle be the ring of polynomials in noncommuting indeterminates x_1, \ldots ,x_n and with coefficients in C. We will assume that each x_i commutes with every element of C. If n=1, then C\langle x_1, \ldots ,x_n\rangle is just the ordinary commutative polynomial ring C[x].
A monomial is an element of C\langle x_1, \ldots ,x_n\rangle which is of the form y_1y_2 \cdots y_k, where y_i \in \{x_1, \cdots , x_n\} for all i. The degree of a monomial y_1y_2 \cdots y_k is defined to be k. For example, x_1x_2x_1^3x_5^2 is a monomial of degree 7. So an element of f \in C\langle x_1, \ldots ,x_n\rangle is a C-linear combination of monomials, and we say that f is monic if the coefficient of at least one of the monomials of the highest degree in f is 1. For example, x_1^2+x_2-3x_2x_3x_2+2x_4^3 is not monic because none of the monomials of the highest degree, i.e. x_2x_3x_2 and x_4^3, have coefficient 1, but x_1^2+x_2-3x_2x_3x_2+x_4^3 is monic.

Definition 1. A ring R is called a polynomial identity ring, or PI ring for short, if there exists a positive integer n and a monic polynomial f \in \mathbb{Z}\langle x_1, \ldots ,x_n\rangle such that f(r_1, \cdots , r_n)=0 for all r_i \in R. We then say that R satisfies f or f is an identity of R.

Definition 2. Let C be a commutative ring with identity, and let R be a C-algebra. If, in Definition 1, we replace \mathbb{Z} with C, we will get the definition of a PI algebra. So R is called a PI algebra if there exists a positive integer n and a monic polynomial f \in C\langle x_1, \ldots ,x_n\rangle such that f(r_1, \cdots , r_n)=0 for all r_i \in R. Note that since every ring is a \mathbb{Z}-algebra, every PI ring is a PI algebra.

Example 1. Every commutative ring R is a PI ring.

Proof. Since R is commutative, r_1r_2-r_2r_1=0, for all r_1,r_2 \in R, and therefore R satisfies the monic polynomial f=x_1x_2-x_2x_1. \ \Box

Remark 1. A PI ring could satisfy many polynomials. For example a finite field of order q satisfies the polynomial in Example 1, because it is commutative, and it also satisfies the polynomial x^q-x. Another example is Boolean rings; they satisfy the polynomial in Example 1, because they are commutative, and they also satisfy the polynomial x^2-x.

Example 2. If C is a commutative ring with identity, then R:=M_2(C) is a PI ring.

Proof. Let A_1,A_2 \in R. Then \text{tr}(A_1A_2 - A_2A_1)=0 and so, by Cayley-Hamilton,

(A_1A_2 - A_2A_1)^2 = cI_2,

for some c \in C. Thus (A_1A_2-A_2A_1)^2 commutes with every element of R, i.e.,

(A_1A_2-A_2A_1)^2A_3-A_3(A_1A_2-A_2A_1)^2=0

for all A_1,A_2,A_3 \in R. Thus R satisfies the monic polynomial

f=(x_1x_2-x_2x_1)^2x_3 - x_3(x_1x_2 - x_2x_1)^2. \ \Box

Example 3. The division ring of real quaternions \mathbb{H} is a PI ring.

Proof. Recall that \mathbb{H}=\mathbb{R}+\mathbb{R}\bold{i}+\mathbb{R}\bold{j}+\mathbb{R}\bold{k}, where \bold{i}^2=\bold{j}^2=-1, \ \bold{ij}=-\bold{ji}=\bold{k}. Let

x=a+b\bold{i}+c\bold{j}+d\bold{k} \in \mathbb{H}, \ \ \ \ \ a,b,c,d \in \mathbb{R}.

It is easy to see that x^2-2ax+a^2+b^2+c^2+d^2=0. Thus, since a,b,c,d are in the center of \mathbb{H}, we get that y(x^2-2ax)=(x^2-2ax)y for all y \in \mathbb{H}. So yx^2-x^2y=2a(yx-xy) and hence, since a is central, (yx-xy)(yx^2-x^2y)=(yx^2-x^2y)(yx-xy) for all x,y \in \mathbb{H}. Therefore \mathbb{H} satisfies the monic polynomial f=(x_1x_2-x_2x_1)(x_1x_2^2-x_2^2x_1)-(x_1x_2^2-x_2^2x_1)(x_1x_2-x_2x_1). \ \Box

Remark 2. If R is a PI ring with identity, then R could satisfy a polynomial with a nonzero constant. For example, the ring \mathbb{Z}/n\mathbb{Z} satisfies the polynomial in Example 1, and it also satisfies the polynomial f=(x-1)(x-2) \cdots (x-n)+n. Now suppose that R has no identity, and f \in \mathbb{Z}\langle x_1, \ldots ,x_n\rangle has a nonzero constant m. So f=g + m, where g \in \in \mathbb{Z}\langle x_1, \ldots ,x_n\rangle has a zero constant. Now, what is f(r_1, \cdots , r_n), \ r_i \in R? Well, It is not defined because it is supposed to be g(r_1, \cdots , r_n)+m1_R but 1_R does not exist. So if R has no identity, all identities of R must have zero constants.

Example 4. Any subring or homomorphic image of a PI ring is a PI ring.

Proof. If R satisfies f, then obviously any subring of R satisfies f too. If S is a homomorphic image of R, then S \cong R/I for some ideal I of R. Now, it is clear that for any f(x_1, \cdots , x_n) \in \mathbb{Z}\langle x_1, \ldots ,x_n\rangle and any r_1, \cdots , r_n \in R, we have f(r_1+I, \cdots , r_n+I)=f(r_1, \cdots , r_n) + I. Hence if R satisfies f, then R/I satisfies f too. \ \Box

Example 5. If I is a nilpotent ideal of a ring R, and R/I is a PI ring, then R is a PI ring.

Proof. So I^m=(0) for some positive integer m. Now suppose that R/I satisfies a monic polynomial f(x_1, \cdots , x_n) \in \mathbb{Z}\langle x_1, \ldots ,x_n\rangle. Then

I=0_{R/I}=f(r_1+I, \cdots , r_n+I)=f(r_1, \cdots , r_n) + I

for all r_i \in R, and so f(r_1, \cdots , r_n) \in I. Thus (f(r_1, \cdots , r_n))^m \in I^m=(0) and hence R satisfies the monic polynomial f^m. \ \Box

Example 6. A finite direct product of PI rings is a PI ring.

Proof. By induction, we only need to prove that for a direct product two PI rings. So let R_1, R_2 be PI rings that, respectively, satisfy f,g \in \mathbb{Z}\langle x_1, \ldots ,x_n\rangle, and let R:=R_1 \times R_2. It is clear that for any polynomial h \in \mathbb{Z}\langle x_1, \ldots ,x_n\rangle and (r_i,s_i) \in R, \ 1 \le i \le n, we have

h((r_1,s_1), \cdots , (r_n,s_n))=(h(r_1, \cdots , r_n),h(s_1, \cdots , s_n)).

Therefore

f((r_1,s_1), \cdots , (r_n,s_n))g((r_1,s_1), \cdots , (r_n,s_n))=(f(r_1, \cdots , r_n)g(r_1, \cdots , r_n), f(s_1, \cdots , s_n)g(s_1, \cdots , s_n))

=(0_{R_1},0_{R_2})=0_R,

because f(r_1, \cdots , r_n)=0_{R_1} and g(s_1, \cdots , s_n)=0_{R_2}. So R satisfies the monic polynomial fg. \ \Box

Exercise. Let R be a ring with the center C. Suppose that for every r \in R, there exist a,b,c \in C such that r^3+ar^2+br+c=0. Show that R is a PI ring.
Hint. See the proof of Example 3.

Note. The reference for this post is Section 1, Chapter 13 of the book Noncommutative Noetherian Rings by McConnell and Robson. Example 3 was added by me.

Throughout this post, R is a ring with identity and Z(R) is its center. We denote by \mathcal{I}(R) the set of idempotents of R, i.e. r \in R such that r^2=r. Also, R[x] is the ring of polynomials in x with coefficients in R. Given r \in R, the left-multiplication map \ell_r: R \to R is defined by \ell_r(a)=ra, for all a \in R. Finally, ring homomorphisms in this post are not required to preserve the identity element.

Left multiplication maps \ell_r are clearly right R-module homomorphisms, i.e. \ell_r(ab)=\ell_r(a)b, for all a,b \in R. But for which r \in R is the map \ell_r a ring homomorphism? That’s probably not easy to answer in general, but I’m going to share some of my findings with you anyway. Before I do that, let me define a class of rings with a funny name which are related to our question.

Definition. A ring R is called abelian if \mathcal{I}(R) \subseteq Z(R), i.e. every idempotent of R is central.

Commutative rings are clearly abelian but abelian rings are not necessarily commutative. For example, by Remark 3 in this post, every reduced ring is abelian (in fact, by Exercise 1 in that post, every reversible ring is abelian). Part vi) of the following problem gives an equivalent definition of abelian rings: a ring R is abelian if and only if \ell_r is a ring homomorphism for every idempotent r of R.

Problem (Y. Sharifi). Let R be a ring, and consider the set

\mathcal{L}(R):=\{r \in R: \ \ell_r \ \text{is a ring homomorphism}\}.

Show that

i) \mathcal{L}(R)=\{r \in R: \ ra=rar, \ \forall a \in R\},

ii) if R is commutative, then \mathcal{L}(R)=\mathcal{I}(R),

iii) \mathcal{L}(R) is multiplicatively closed,

iv) given r \in \mathcal{I}(R), the set S_r:=\{a \in R: \ ra=rar\} is a subring of R,

v) Z(R) \cap \mathcal{I}(R) \subseteq \mathcal{L}(R) \subseteq \mathcal{I}(R),

vi) \mathcal{L}(R) = \mathcal{I}(R) if and only if R is abelian,

vii) if R is semiprime, then \mathcal{L}(R)= Z(R) \cap \mathcal{I}(R),

viii) if R is prime, then \mathcal{L}(R)=\{0,1\},

ix) \mathcal{L}(R) =R \cap \mathcal{L}(R[x]) and if R is commutative or semiprime, then \mathcal{L}(R[x])=\mathcal{L}(R),

x) \mathcal{L}\left(\prod_{i \in I}R_i\right)=\prod_{i \in I}\mathcal{L}(R_i) for any family of rings \{R_i\}_{i \in I}.

Solution. i) Since \ell_r is clearly additive, it is a ring homomorphism if and only if \ell_r(ab)=\ell_r(a)\ell_r(b), i.e. rab=rarb for all a,b \in R. Choosing b=1 gives ra=rar. Conversely, if ra=rar, then rab=rarb and so \ell_r is a ring homomorphism.

ii) By i), r \in \mathcal{L}(R) if and only if ra=r^2a, for all a \in R, which holds if and only if r^2=r, i.e r \in \mathcal{I}(R).

iii) Let r,s \in \mathcal{L}(R) and a \in R. Then, by i), rsar=rsa, \ sas=sa, and so rsars=rsas=rsa. Thus rs \in \mathcal{L}(R).

iv) If a,b \in S_r, then rab=rarb=rarbr=rabr and so ab \in S_r.

v) Let r \in \mathcal{L}(R). Then choosing a=1 in i) gives r=r^2 and so r \in \mathcal{I}(R). Now, let r \in Z(R) \cap \mathcal{I}(R), and a \in R. Then rar=rra=r^2a=ra and so r \in \mathcal{L}(R).

vi) If \mathcal{I}(R) \subseteq Z(R), then \mathcal{L}(R) = \mathcal{I}(R), by v). Suppose now that \mathcal{L}(R) = \mathcal{I}(R) and r \in \mathcal{I}(R). Then 1-r \in \mathcal{I}(R), and so, by i), ra=rar, \ (1-r)a=(1-r)a(1-r), for all a \in R. Thus

a-ra=(1-r)a=(1-r)a(1-r)=a-ar-ra+rar=a-ar,

which gives ra=ar and so r \in Z(R).

vii) By v), we only need to show that \mathcal{L}(R) \subseteq Z(R). So let r \in \mathcal{L}(R) and a,b \in R. If we show that (ra-ar)b(ra-ar)=0, we are done because that means (ra-ar)R(ra-ar)=(0) and so, since R is semiprime, ra=ar hence r \in Z(R). Now, to show that (ra-ar)b(ra-ar)=0, we have

(ra-ar)b(ra-ar)=rabra-rabar-arbra+arbar

and so, since by i), rabr=rab, \ rbr=rb, \ rbar=rba, we get that

(ra-ar)b(ra-ar)=raba-raba-arba+arba=0.

viii) Clearly the inclusion \{0,1\} \subseteq \mathcal{L}(R) holds for any ring R. Now, suppose that R is a prime ring and r \in \mathcal{L}(R), a \in R. Then, by i), ra(1-r)=0 and so rR(1-r)=(0), which gives r=0 or r=1.

ix) Let r \in R. Then, by i), r \in \mathcal{L}(R[x]) if and only if rf(x)=rf(x)r for all f(x) \in R[x] if and only if ra=rar for all a \in R, because x is central in R[x], if and only if r \in \mathcal{L}(R), by i).
If R is commutative, then \mathcal{L}(R)=\mathcal{I}(R) and so, since the polynomial ring R[x] is commutative too,

\mathcal{L}(R[x])=\mathcal{I}(R[x])=\mathcal{I}(R)=\mathcal{L}(R).

Note that the identity \mathcal{I}(R[x])=\mathcal{I}(R) is true by this post.
Now suppose that R is semiprime, and f(x)=\sum_{i=0}^nr_ix^i \in \mathcal{L}(R[x]), \ n \ge 1. So f(x)a=f(x)af(x) for all a \in R and hence, equating the coefficients of x^{2n} on both sides of the equality gives 0=r_nar_n. Thus r_nRr_n=(0) and so, since R is semiprime, r_n=0, contradiction. This proves that n=0 and hence f(x) \in R giving \mathcal{L}(R[x]) = \mathcal{L}(R).

x) Clear, by i). \ \Box

Example 1. This example shows that \mathcal{L}(R) is not always central. Let C be a commutative domain, and let R be the ring of 2 \times 2 upper triangular matrices with entries in C. Then, using the first part of the Problem, it is easy to see that

\mathcal{L}(R)=\left \{\begin{pmatrix}0 & 0 \\ 0 & 0\end{pmatrix}, \begin{pmatrix}1 & 0 \\ 0 & 1\end{pmatrix}, \begin{pmatrix}0 & c \\ 0 & 1\end{pmatrix}, \ c \in C\right \}.

Example 2. Let R be the ring in Example 1. Then \mathcal{L}(R[x]) \neq \mathcal{L}(R) because f(x)=r+sx \in \mathcal{L}(R[x]), where

r=\begin{pmatrix}0 & 0 \\ 0 & 1\end{pmatrix}, \ \ \ \ \ s=\begin{pmatrix}0 & 1 \\ 0 & 0\end{pmatrix}.

Example 3. If R is a semisimple ring, then |\mathcal{L}(R)|=2^n for some positive integer n.

Proof. By the Artin-Wedderburn theorem, R is a finite direct product of matrix rings over division rings. The result now follows from parts viii) and x) of the Problem. \ \Box

Throughout this post, R is a ring with identity, and M_n(R) is the ring of n \times n matrices with entries from R. Also, by ideal we always mean two-sided ideal.

Here we defined a prime ideal of R as a proper ideal P which satisfies this condition: if IJ \subseteq P for some ideals I,J of R, then I \subseteq P or J \subseteq P. By weakening this condition, we get a set of ideals of R that contains the set of prime ideals of R; we call these ideals semiprime.

Definition 1. A proper ideal Q of a ring R is called semiprime if for any ideal I of R, \ I^2 \subseteq Q implies that I \subseteq Q.

Now, what exactly are semiprime ideals in commutative rings?

Remark 1. In a commutative ring R, semiprime ideals are just radical ideals.

Proof. Since R is commutative, the condition I^2 \subseteq Q \implies I \subseteq Q for all ideals I of R, which is the definition of a semiprime ideal Q, is equivalent to the condition a^2 \in Q \implies a \in Q for all a \in R, which itself is equivalent to a^n \in Q \implies a \in Q for all a \in R and all positive integers n (why?). Now, the set

\{a \in R: \ a^n \in Q, \ \text{for some positive integer} \ n\}

is, by definition, \sqrt{Q}, the radical of Q, which is easily seen to be an ideal containing Q. If \sqrt{Q}=Q or, equivalently, \sqrt{Q} \subseteq Q, then we say that Q is a radical ideal. So semiprime ideals of a commutative ring are just radical ideals of the ring. \ \Box

Remark 2. It is clear from Definition 1 that any intersection of prime ideals is semiprime. We also know from commutative ring theory that radical ideals are exactly those ideals that are intersection of some prime ideals. By Remark 1, radical ideals of a commutative ring are exactly semiprime ideals of the ring. So it is natural to ask if, in general ring theory, it is true that every semiprime ideal is an intersection of some prime ideals. It turns out that the answer is positive. In fact some authors define a semiprime ideal as an ideal which is an intersection of some prime ideals and then they show this is equivalent to Definition 1. But I did not choose this approach here because it will complicate the proof of the following Proposition hence getting into the goal of this post which is introducing semiprime rings.

By Remark 1, an ideal Q in a commutative ring R is semiprime if and only if a^2 \in Q \implies a \in Q for all a \in R. In general ring theory, we have the following similar result (compare with Proposition 1 in this post!).

Proposition. A proper ideal Q of a ring R is semiprime if and only if for any a \in R, \ aRa \subseteq Q implies that a \in Q.

Proof. Suppose first that Q is semiprime and aRa \subseteq Q for some a \in R. Let I:=RaR. Then I is an ideal of R and I^2=RaRaR=R(aRa)R \subseteq Q. Thus I \subseteq Q, which gives a \in Q.
Conversely, suppose now that I^2 \subseteq Q for some ideal I of R and a \in I. Then

aRa \subseteq R(aRa)R=(RaR)(RaR) \subseteq I^2 \subseteq Q

and so a \in Q. Hence I \subseteq Q which proves that Q is semiprime. \ \Box

By Remark 1, (0) is a semiprime ideal of a commutative ring R if and only if \sqrt{(0)}=(0), i.e. if a^n=0 for some a \in R and some positive integer n, then a=0. In other words, (0) is a semiprime ideal of R if and only if R is reduced. In general ring theory, it is true that if R is reduced, then (0) is a semiprime ideal of R (Example 1). However the converse is not true (Example 2). If (0) is a semiprime ideal of a ring, then we will call the ring semiprime. So, in general ring theory, every reduced ring is semiprime but not every semiprime ring is reduced.

Definition 2. We say that a ring R is semiprime if (0) is a semiprime ideal of R, i.e. for any ideal I of R, if I^2=(0), then I=(0).

The Proposition gives the following simple yet useful test to check if a ring is semiprime or not.

Corollary. A ring R is semiprime if and only if for any a \in R, \ aRa=(0) implies that a=0.

Example 1. Every reduced ring R is semiprime.

Proof. If aRa=(0), for some a \in R, then a^2=0 and so a=0, because R is reduced. \ \Box

Example 2. Every prime ring is obviously semiprime and so the matrix ring M_n(k) over a field k is a semiprime ring which is not reduced.

Example 3. Every semiprimitive ring R is semiprime.

Proof. Suppose that aRa=(0), for some a \in R. Then (ra)^2=0 for all r \in R and hence 1-ra is invertible. Thus a \in J(R)=(0). \ \Box

Example 4. Let R be a ring and S:=M_n(R). Then S is semiprime if and only if R is semiprime. In particular, if R is reduced, then S is semiprime.

Proof. It’s the same as the proof for prime rings (Example 4). \ \Box

Exercise 1. Show that a proper ideal Q of a ring R is semiprime if and only if the ring R/Q is semiprime.

Exercise 2. Show that a proper ideal Q of a ring R is semiprime if and only if I^2 \subseteq Q implies that I \subseteq Q for any left ideal I of R.

Exercise 3. Show that the center of a semiprime ring R is reduced.
Hint. If a^2=0 for some central element of R, then aRa=Ra^2=(0).

Exercise 4. Show that a direct product of rings is semiprime if and only if each ring is semiprime.

Exercise 5. Show that a ring R is semiprime if and only if R has no nonzero nilpotent ideal, i.e if I^n=(0) for some ideal I of R and some positive integer n, then I=(0).
Hint. If n \ge 2 is the smallest positive integer n such that I^n=0, then (I^{n-1})^2=I^{2n-2}=(0).

Note. The references for this post are Chapter 4 of T. Y. Lam’s book A First Course in Noncommutative Rings and Chapter 3 of Goodearl & Warfield’s book An Introduction to Noncommutative Noetherian Rings.

Throughout this post, R is a ring with identity, and M_n(R) is the ring of n \times n matrices with entries from R. Also, by ideal we always mean two-sided ideal.

In commutative ring theory, a proper ideal P of a commutative ring R is said to be prime if IJ \subseteq P, where I,J are ideals of R, implies that I \subseteq P or J \subseteq P. The definition of a prime ideal remains the same even if R is not commutative.

Definition 1. A proper ideal P of a ring R is called prime if for any ideals I,J of R, \ IJ \subseteq P implies that I \subseteq P or J \subseteq P.

In commutative ring theory, we see that a proper ideal P is prime if and only if a,b \in R, \ ab \in P implies that a \in P or b \in P. In general ring theory, we have the following similar result.

Proposition 1. A proper ideal P of a ring R is prime if and only if for any a,b \in R, \ aRb \subseteq P implies that a \in P or b \in P.

Proof. Suppose first that P is prime and aRb \subseteq P for some a,b \in R. Let I:=RaR, \ J:=RbR. Then I,J are ideals of R and IJ=RaRbR=R(aRb)R \subseteq P. Thus either I \subseteq P, which gives a \in P, or J \subseteq P, which gives b \in P.
Conversely, suppose now that IJ \subseteq P for some ideals I,J of R and I \nsubseteq P. We need to show that J \subseteq P. Let a \in I \setminus P, \ b \in J. Then aRb \subseteq R(aRb)R=(RaR)(RbR) \subseteq IJ \subseteq P and so b \in P because a \notin P. Hence J \subseteq P which proves that P is prime. \ \Box

In commutative ring theory, if (0) is a prime ideal of a ring, the ring is called a commutative domain. In general ring theory, we still call a ring R a domain if a,b \in R, \ ab=0 implies that a=0 or b=0. It is true that if R is a domain, then (0) is a prime ideal of R (Example 1). However, in general, the converse is not true (see Example 2 and Proposition 2). If (0) is a prime ideal of a ring, then we will call the ring prime. So, in general ring theory, every domain is a prime ring but not every prime ring is a domain.

Definition 2. We say that a ring R is prime if (0) is a prime ideal of R, i.e. for any ideals I,J of R, if IJ=(0), then I=(0) or J=(0).

Proposition 1 gives the following simple yet useful test to check if a ring is prime or not. This is useful because it’s in terms of elements not ideals, which are usually not easy to find.

Corollary. A ring R is prime if and only if for any a,b \in R, \ aRb=(0) implies that a=0 or b=0.

Example 1. Every domain R is prime.

Proof. Suppose that aRb=(0) for some a,b \in R. Then ab=0 and so either a=0 or b=0, because R is a domain. \ \Box

Example 2. Every simple ring R is prime. In particular, every matrix ring M_n(k) over a field k is prime.

Proof. The only ideals of R are (0), R. So IJ=(0), where I,J are ideals of R, implies that I=(0) or J=(0). \ \Box

So not every prime ring is a domain. The following characterizes all prime rings that are domain.

Proposition 2. A ring is a domain if and only if it’s both prime and reduced.

Proof. See Remark 2 in this post. \ \Box

The following example generalizes Example 2.

Example 3. Every left primitive ring is prime.

Proof. See Fact 2 in this post. \ \Box

Example 4. Let R be a ring and S:=M_n(R). Then S is prime if and only if R is prime. In particular, if R is a domain, then S is prime.

Proof. It is a well-known fact that ideals of S are in the form M_n(I), where I is any ideal of R. Now, suppose that R is prime and KL=(0) for some ideals K,L of S. We have K=M_n(I), \ L=M_n(J) for some ideals I,J of R. So M_n(I)M_n(J)=(0) and hence IJ=(0), which gives I=(0) or J=(0), because R is prime. Therefore K=(0) or L=(0) and so S is prime. The proof for the converse is similar. \ \Box

Exercise 1. Show that a proper ideal P of a ring R is prime if and only if the ring R/P is prime.

Exercise 2. Show that a proper ideal P of a ring R is prime if and only if IJ \subseteq P implies that I \subseteq P or J \subseteq P, for any left ideals I,J of R.

Exercise 3. Show that the center of a prime ring R is a commutative domain.
Hint. If ab=0 for some central elements of R, then aRb=Rab=(0).

Exercise 4. Show that a direct product of two or more rings can never be a prime ring.
Hint. If R=R_1 \times R_2, and a=(1,0), \ b=(0,1), then aRb=(0).

Exercise 5. Show that every maximal ideal of a ring is prime.

Note. The reference for this post is the first few pages of Chapter 4 of T. Y. Lam’s book A First Course in Noncommutative Rings.