Posts Tagged ‘prime ring’

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.

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.

Definition. A ring R is called reduced if it has no non-zero nilpotent element, i.e.

x \in R, \ x^n=0 \Longrightarrow x=0.

Note that R is reduced if and only if x \in R, \ x^2=0 \Longrightarrow x=0. To see this, suppose that x \in R is nilpotent and n is the smallest positive integer such that x^n=0. If n is even, then we’ll have (x^{n/2})^2=0 giving x^{n/2}=0, which contradicts minimality of n. If n > 1 is odd, then x^n=0 gives x^{n+1}=0 and so x^{(n+1)/2}=0, which again contradicts minimality of n. So n=1 and hence x=0, i.e. R is reduced. The converse is trivial.

Examples. Any domain, any subring of a domain, and any direct product of domains are clearly reduced. Also, by the Remark in this post, if R is a commutative reduced ring, then the polynomial ring R[x] is reduced too. If R is a ring with identity, then the matrix M_n(R), \ n \ge 2, with the standard basis \{e_{ij}\} is never reduced because, for example, e_{12}^2=0.

Remark 1. In a reduced ring R if xy=0, for some x,y \in R, then yRx=\{0\}. As a result, the left and the right annihilators of an element of R are equal.

Proof. For any z \in R we have (yzx)^2=0 and thus yzx=0. \ \Box

Remark 2. A ring R is a domain if and only if it is both prime and reduced.  Moreover, every reduced ring is semiprime.

Proof. Let x,y \in R. Suppose that R is a domain. Then R is reduced and if xRy=(0), then xy=0 and thus, since R is a domain, either x=0 or y=0. So R is a prime ring. Now suppose that R is both prime and reduced and xy=0. Then, by Remark 1, yRx=(0) and so, since R is prime, either y=0 or x=0. To prove that reduced rings are semiprime, suppose that xRx=(0), \ x \in R. Then x^2 \in xRx =(0) and so x=0. \ \Box

Remark 3. If R is reduced, then every idempotent of R is central.

Proof. Suppose that x \in R is an idempotent, i.e. x^2=x, and let y \in R. See that

(xy-xyx)^2=(yx-xyx)^2=0

and thus xy=yx=xyx. \ \Box

Remark 4. The intersection of all prime ideals of a reduced ring is zero.

Proof.  Suppose that the intersection of all prime ideals of R is non-zero and choose x \neq 0 in that intersection. Since R is reduced, x^n \neq 0 for all n \in \mathbb{N}. Let

S = \{x^n: \ n \in \mathbb{N} \},  \ \ \ \ \mathcal{A}=\{I \lhd R : \ I \cap S = \emptyset \}. 

Note that (0) \in \mathcal{A} and so \mathcal{A} \neq \emptyset. Thus, by Zorn’s lemma, \mathcal{A} has a maximal element, say P. To get a contradiction, we only need to show that P is a prime ideal:  suppose that I,J are some ideals of R which properly contain P and IJ \subseteq P. Since P is a maximal element of \mathcal{A}, we get I \notin \mathcal{A}, \ J \notin \mathcal{A}. Thus x^i \in I, \ x^j \in J, for some i,j \in \mathbb{N}. But then x^{i+j} \in IJ \subseteq P, which is absurd. \Box

Remark 5. In a ring R every prime ideal contains a minimal prime ideal. So the intersection of all prime ideals of R is equal to the intersection of all minimal prime ideals of R. If R is reduced, then the intersection of all minimal prime ideals of R is zero.

Proof. Let P be a prime ideal of R and \mathcal{B} the set of prime ideals of R contained in P. Consider \mathcal{B} with \supseteq. Clearly \mathcal{B} \neq \emptyset because P \in \mathcal{B}. Let \{Q_i \}_{i \in I} be a totally ordered subset of \mathcal{B}. Since \bigcap_{i \in I} Q_i \in \mathcal{B}, Zorn’s lemma gives us a prime ideal Q \subseteq P which is maximal in \mathcal{B}, i.e. Q is a minimal prime. \Box

See the second part of this post here.

Exercise 1 (Two generalizations of Remark 3). Let R be a ring. Show that

i) if R is reversible, then every idempotent of R is central (see Definition 2 and Example 1 in this post),

ii) if every nilpotent element of R is central, then every idempotent of R is central.
Hint. i) If x,y \in R, \ x^2=x, then x(xy-y)=(yx-y)x=0.

Exercise 2. Show that the center of a semiprime ring R is reduced.
Hint. If x is in the center of R, then xRx=Rx^2.

For the first part see here.

6) If R is simple, then Z(R)=Z(Q).

Proof. Let x=s^{-1}a \in Z(Q). Then from xs=sx we get sa=as and thus s^{-1}a=as^{-1}. Hence for every b \in R we’ll have s^{-1}ab=bs^{-1}a=bas^{-1}, which gives us abs=sba. Also, since R is simple, RsR=R, which means \sum_{i=1}^n b_isc_i = 1, for some b_i, \ c_i \in R. Thus

\sum_{i=1}^n sb_iac_i = \sum_{i=1}^n ab_isc_i = a=sx

and so x=\sum_{i=1}^n b_iac_i \in R. Therefore x \in Z(R) which proves Z(Q) \subseteq Z(R). Conversely, let b \in Z(R) and x=s^{-1}a \in Q. Since bs=sb, we have s^{-1}b=bs^{-1} and thus

bx=bs^{-1}a=s^{-1}ba=s^{-1}ab=xb

and so b \in Z(Q). \Box

7) The left uniform dimension of R and Q are equal.

Proof. We saw in the previous section that the left ideals of Q are exactly in the form QI, where I is a left ideal of R. Clearly \sum QI_i is direct iff \sum I_i is direct.

8) Let N be a nilpotent ideal of R and let I be the right annihilator of N in R. Then I is an essential left ideal of R and hence QI  is an essential left ideal of Q.

Proof. Let I be the right annihilator of N in R. For an essential left ideal J of R the left ideal QJ of Q is essential in Q because for every non-zero left ideal K of R : (0) \neq \ Q(J \cap K) \subseteq QJ \cap QK. So we only need to prove the first part of the claim. Let J be any non-zero left ideal of R and put

n=\min \{k \geq 0 : \ N^k J \neq (0) \}.

Then (0) \neq N^n J \subseteq I \cap J.

9) If Q is semisimple, then R is semiprime.

Proof. So we need to prove that R has no non-zero nilpotent ideal. Suppose that N is a nilpotent ideal of R and let I be the right annihilator of N in R. Since Q is semisimple, QI \oplus A = Q, for some left ideal A of Q. But, from the previous fact, we know that QI is essential in Q and thus A=(0), i.e. QI=Q. Thus s^{-1}a=1, for some a \in I=\text{r.ann}_R N. So s=a and Ns=Na=(0). Thus N=Nss^{-1}=(0).

We proved in the previous section that if R is prime, then Q is prime too.

10) If Q is simple, then R is prime.

Proof. Let I,J be two non-zero ideals of R. We need to show that IJ \neq (0). We have QIQ=Q, because I \neq (0) and Q is simple. Therefore 1=\sum_{i=1}^n x_ia_iy_i, for some x_i,y_i \in Q and a_i \in I. We can write x_i = s^{-1}b_i, for some b_i \in R. Then s=\sum_{i=1}^n b_ia_iy_i \in IQ. So IQ is a right ideal of Q which contains a unit. Thus IQ=Q. Similarly JQ=Q and hence IJQ=Q. As a result, IJ \neq (0). \Box

Throughout R is a ring with unity, S \subset R is a regular submonoid which is left Ore. So Q, the left quotient ring of R with respect to S, exists and R is a subring of Q. The following facts also holds for the “right” version:

 1) If I \lhd_{\ell} Q, then I \cap R \lhd_{\ell} R and if J \lhd_{\ell} Q, then I \cap R = J \cap R \Longleftrightarrow I=J. Also I=Q(I \cap R).

Proof. The first part is trivial. Now suppose that s^{-1}a=x \in I. Then a = sx \in I \cap R = J \cap R and so x=s^{-1}a \in J. Hence I \subseteq J. Similarly J \subseteq I. For the last part, if x = s^{-1}a \in I, then a=sx \in I \cap R and so x=s^{-1}a \in Q(I \cap R).

2) If I \lhd_{\ell} R, then QI \lhd_{\ell} Q and QI=\{s^{-1}a : \ s \in S, \ a \in I \}. It is also true that QI \cap R = \{a \in R : \ sa \in I \ \text{for some} \ s \in S \}.

Proof.  Again, the first part is trivial. To prove the non-trivial side of the second part, let x = \sum_{i=1}^n x_i a_i \in QI, where x_i \in Q and a_i \in I. As we’ve showed before, there exist some s \in S and b_i \in R such that x_i = s^{-1}b_i  and thus x = s^{-1}a, where a=\sum_{i=1}^n b_ia_i \in I. For the last part, suppose that s^{-1}b \in QI \cap R, where s \in S, \ b \in I, by the second part. So we have s^{-1}b=a \in R and thus sa = b \in I.

3) For any I, J \lhd_{\ell} R we have I \cap J = (0) \Longrightarrow QI \cap QJ = (0).

Proof. Let x=s^{-1}a = t^{-1}b \in QI \cap QJ, where a \in I, \ b \in J. We have ts^{-1} \in Q and thus ts^{-1}=u^{-1}c, for some u \in S, \ c \in R. Then, since ts^{-1}a=b, we get ca=ub \in I \cap J = (0), which gives us b=0 because u is a unit in Q (or because u is regular in R). Thus x = 0.

4) If R is left Noetherian, then so is Q. If I \lhd R and Q is left Noetherian, then QI \lhd Q.

Proof. For the first part, first note that, by 1) and 2), left ideals of Q are exactly in the form QI, where I is a left ideal of R. Let I \lhd_{\ell} R and I=\sum_{i=1}^n Ra_i. Then QI=\sum_{i=1}^n Qa_i. For the second part, let I be a two-sided ideal of R and s \in S. Then I \subseteq Is^{-1} \subseteq Is^{-2} \subseteq \cdots and thus QI \subseteq QIs^{-1} \subseteq QIs^{-2} \subseteq \cdots. So, since Q is left Noetherian, there exists some n \geq 0 such that QIs^{-n} = QIs^{-n-1}, and thus QI=QIs^{-1}. Therefore QIs^{-1}a = QIa \subseteq QI, for all s \in S, \ a \in R. Hence QIQ \subseteq QI and so QI is a two-sided ideal of Q. (In fact QIQ=QI.)

5) If R is prime, then so is Q.

Proof. Suppose that x_1Qx_2=(0), for some x_1=s_1^{-1}a_1, x_2=s_2^{-1}a_2 \in Q. Then s_1^{-1}a_1(as_2)s_2^{-1}a_2 = 0, for all a \in R. Thus s_1^{-1}a_1aa_2 = 0, which gives us a_1aa_2=0, i.e. a_1Ra_2=(0). Therefore either a_1=0 or a_2=0, which means either x_1=0 or x_2=0. This also proves that if R is semiprime, then so is Q.

Fact 1. Let R be a left primitive ring and let M be a faithful simple left R module. By Schur’s lemma D=\text{End}_R(M) is a division ring and M can be considered as a right vector space over D in the usual way. Let S=\text{End}_D(M) and define \varphi : R \longrightarrow S by \varphi(r)(x)=rx, for all r \in R and x \in M. Then \varphi is a well-defined ring homomorphism. Also \varphi is one-to-one because M is faithful. So R can be viewed as a subring of S.

Fact 2. Every left primitive ring R is prime.

Proof. Suppose M is a faithful simple left R module and I,J be two non-zero ideals of R with IJ=(0).  Now JM is a submodule of M and M is simple. Therefore either JM=(0) or JM=M. If JM=(0), then we get (0) \neq J \subseteq \text{ann}_R M = (0), which is nonsense. Finally, if JM=M, then we will have (0)=(IJ)M=I(JM)=IM. Thus I \subseteq \text{ann}_R M = (0) and so I=(0), a contradiction! \ \Box

Fact 3. A trivial result of Fact 2 is that the center of a left primitive ring is a commutative domain. A non-trivial fact is that every commutative domain is the center of some left primitive ring. For a proof of this see: T. Y. Lam,  A First Course in Noncommutative Ring Theory, page 183.

Fact 4. Let R be a prime ring and M a faithful left R module of finite length. Then R is left primitive.

Proof. Let (0)=M_0 \subset M_1 \subset \cdots \subset M_n=M be a composition series of M. Therefore M_k/M_{k-1} is a simple left R module for every 1 \leq k \leq n. We also let I_k=\text{ann}_R (M_k/M_{k-1}). Then each I_k is an ideal of R and it’s easy to see that I_1I_2 \cdots I_nM = (0). Thus I_1I_2 \cdots I_n = (0), because M is faithful. Hence I_{\ell} = (0), for some \ell, because R is prime. Therefore M_{\ell}/M_{\ell - 1} is a faithful simple left R module. \ \Box

Fact 5. Every left primitive ring R is semiprimitive.

Proof. Let M be a faithful simple left R module and J=J(R), as usual, be the Jacobson radical of R. The claim is that J=(0). So suppose that J \neq \{0\}  and choose 0 \neq x \in M. Then Rx=M, because M is simple, and so JM=Jx. Also either JM=(0), which is impossible because then J \subseteq \text{ann}_R M=(0), or JM=M. If Jx=JM=M, then ax =x, for some a \in J. Thus (1-a)x=0, which gives us the contradiction x = 0, because 1-a is invertible in R. \ \Box