1 Introduction

Throughout this paper, all rings are associative with identity. Let \(R\) be a ring, we use \(N(R)\), \(J(R)\), \(E(R)\), \(Z(R)\), and \(U(R)\) to denote the set of all nilpotent elements, the Jacobson radical, the set of all idempotent elements, the center, and the set of all invertible elements of \(R\), respectively. For any nonempty subset \(X\) of a ring \(R\), \(r(X)=r_R(X)\) and \(l(X)=l_R(X)\) denote the right annihilator of \(X\) and the left annihilator of \(X\), respectively.

Following [6], a ring \(R\) is called CN if \(N(R)\subseteq Z(R)\). Clearly, commutative rings and reduced rings (thai is, a ring \(R\) with \(N(R)=0\)) are \(CN\).

A theorem of Herstein [10] stated that a ring \(R\) which satisfies the identity \((xy)^n=x^ny^n\), where \(n\) is a fixed positive integer greater than 1, must have nil commutator ideal. In [3], Bell proved that if \(R\) is an \(n\)-torsion-free ring with identity 1 and satisfies the two identities \((xy)^n=x^ny^n\) and \((xy)^{n+1}=x^{n+1}y^{n+1}\), then \(R\) is commutative. In [1], Khuzam proved that if \(R\) is \(n(n-1)\)-torsion-free ring with 1 and satisfies the identity \((xy)^n=x^ny^n\), then \(R\) is commutative. In [2], Khuzam proved that if \(R\) is a semiprime ring in which for each \(x\) in \(R\) there exists a positive integer \(n=n(x)> 1\) such that \((xy)^n=x^ny^n\) for all \(y\in R\), then \(R\) is commutative. In [11], Ligh and Richou proved that if \(R\) is a ring with 1 which satisfies the identities: \((xy)^k=x^ky^k\), \(k=n, n+1, n+2\), where \(n\) is a positive integer, then \(R\) is commutative. The purpose of this note is to generalize these results.

2 Main Results

We begin with the following theorem which generalizes [6, Theorem 5].

Theorem 2.1

The following conditions are equivalent for a ring \(R\):

  1. (1)

    \(R\) is a \(CN\) ring;

  2. (2)

    For any \(a\in N(R)\), there exists \(n=n(a)\ge 2\) such that \(a-a^n\in Z(R)\);

  3. (3)

    For any \(a\in N(R)\) and \(b\in R\), there exists \(c=c(a, b)\in R\) such that \([ a-a^2c, b]=0\).

Proof

\((1)\Longrightarrow (i), i=2,3\) are trivial.

\((2)\Longrightarrow (1)\) Assume that \(a\in N(R)\) with \(a^m=0\) for some \(m\ge 2\). By (2), there exists \(n_1=n_1(a)\ge 2\) such that \(a-a^{n_1}\in Z(R)\). Since \(a^{n_1}\in N(R)\), by (2), there exists \(n_2=n_2(a^{n_1})\ge 2\) such that \(a^{n_1}-a^{n_1n_2}\in Z(R)\). Continuing this process, there exists \(n_s=n_s(a^{n_1n_2\cdots n_{s-1}})\ge 2\) such that \(a^{n_1n_2\cdots n_{s-1}}-a^{n_1n_2\cdots n_{s-1}n_s}\in Z(R)\) and \(n_1n_2\cdots n_{s-1}n_s\ge m\). Hence \(a^{n_1n_2\cdots n_{s-1}n_s}=0\) and \(a=a-a^{n_1n_2\cdots n_{s-1}n_s}=(a-a^{n_1})+(a^{n_1}-a^{n_1n_2})+\cdots +(a^{n_1n_2\cdots n_{s-1}}-a^{n_1n_2\cdots n_{s-1}n_s})\in Z(R)\).

\((3)\Longrightarrow (1)\) Assume that \(a\in N(R)\) with \(a^n=0\) for some \(n\ge 2\). By induction on \(n\), we claim that \(a\in Z(R)\). For each \(x\in R\), by (3), there exists \(c=c(a, x)\in R\) such that \([a-a^2c, x]=0\). If \(n=2\), then \(a\in Z(R)\), we are done. Now we assume that \(n> 2\) and assume that for each \(y\in N(R)\) with the index of nilpotence at most \(n-1\), we have \(y\in Z(R)\). Since \((a^2)^{n-1}=0\), by the induction hypothesis, \(a^2\in Z(R)\). For any \(z\in (a^2)=a^2R\), we have \(z^{n-1}\in a^{2(n-1)}R=0\), so \(z\in Z(R)\) by the induction hypothesis. This implies \(a^2R\subseteq Z(R)\). Hence \(0=[a-a^2c, x]=[a, x]\) for any \(x\in R\), so \(a\in Z(R)\), this shows that \(R\) is \(CN\). \(\square \)

A ring \(R\) is called NLI if \(N(R)\) is a \(Lie-\)ideal of \(R\) (that is, for any \(a\in N(R)\) and \(b\in R\), \(ab-ba\in N(R)\) and \(N(R)\) is an additive subgroup of \(R\)). Clearly, \(NI\) rings (that is, \( N(R)\) forms an ideal of \(R\) ) are \(NLI\). A ring \(R\) is called QCN if for any \(a\in N(R)\) and \(b\in R\), there exist \(n=n(a, b)> 1\) and \(c\in R\) such that \(ab-ba=(ab-ba)^nc\). Clearly, \(CN\) rings are \(QCN\).

Theorem 2.2

The following conditions are equivalent for a ring \(R\):

  1. (1)

    \(R\) is a \(CN\) ring;

  2. (2)

    \(R\) is a \(QCN\) \(NI\) ring;

  3. (3)

    \(R\) is a \(QCN\) \(NLI\) ring.

Proof

\((1)\Longrightarrow (2)\Longrightarrow (3)\) is trivial. \((3)\Longrightarrow (1)\) Assume that \(a\in N(R)\) and \(b\in R\). Since \(R\) is a \(QCN\) ring, \(ab-ba=(ab-ba)^nc\) for some \(n=n(a, b)\) and \(c\in R\). Since \(R\) is an \(NLI\) ring, \(ab-ba\in N(R)\). Let \(m\ge 1\) such that \((ab-ba)^m=0\). Clearly, \((n-1)m+1\ge m\). Since \(ab-ba=(ab-ba)^{(n-1)m+1}c^m\), \(ab-ba=0\). Hence \(a\in Z(R)\) and \(R\) is a \(CN\) ring. \(\square \)

In preparation for the proof of our next theorem, we first state the following known lemma ([12, Lemma 2]).

Lemma 2.1

Let \(x, y\in R\). Suppose that for some positive integer \(n\), \(xy^n=0=x(1+y)^n\). Then \(x=0\).

Theorem 2.3

The following conditions are equivalent for a ring \(R\):

  1. (1)

    \(R\) is a \(CN\) ring;

  2. (2)

    For any \(x\in N(R)\) and \(y\in R\), \(((1+x)y)^{n+k}=(1+x)^{n+k}y^{n+k}\), where \(n\) is a fixed positive integer and \(k=0, 1,2\);

  3. (3)

    For any \(x\in N(R)\) and \(y\in R\), \(((1+x)y)^{n+k}=y^{n+k}(1+x)^{n+k}\), where \(n\) is a fixed positive integer and \(k=0, 1, 2\).

Proof

\((1)\Longrightarrow (i), i=2, 3\) are trivial. \((2)\Longrightarrow (1)\) Assume that \(x\in N(R)\) and \(y\in R\). Then by the hypothesis,

$$\begin{aligned} (1+x)^{n+1}y^{n+1}&= (1+x)^ny^n(1+x)y. \end{aligned}$$
(2.1)
$$\begin{aligned} (1+x)^{n+2}y^{n+2}&= (1+x)^{n+1}y^{n+1}(1+x)y. \end{aligned}$$
(2.2)

Since \(1+x\) is invertible in \(R\), (2.1) gives

$$\begin{aligned} (xy^n-y^nx)y=0, \end{aligned}$$
(2.3)

(2.2) gives

$$\begin{aligned} (xy^{n+1}-y^{n+1}x)y=0. \end{aligned}$$
(2.4)

Multiply (2.3) on the left by \(y\), one gets

$$\begin{aligned} (yxy^n-y^{n+1}x)y=0. \end{aligned}$$
(2.5)

From (2.4) and (2.5) we have

$$\begin{aligned} (xy-yx)y^{n+1}=0. \end{aligned}$$
(2.6)

Since (2.6) holds for all \(y\in R\), substitute \(y+1\) for \(y\), to get

$$\begin{aligned} (xy-yx)(1+y)^{n+1}=0. \end{aligned}$$
(2.7)

From (2.6), (2.7) and Lemma 2.1, we have

$$\begin{aligned} xy=yx. \end{aligned}$$
(2.8)

Hence \(R\) is \(CN\).

\((3)\Longrightarrow (1)\) Suppose that \(x\in N(R)\) and \(y\in R\). Since \(((1+x)y)^{n+1}=(1+x)y((1+x)y)^n\), by the hypothesis, we have

$$\begin{aligned} y^{n+1}(1+x)^{n+1}&= (1+x)y^{n+1}(1+x)^n, \end{aligned}$$
(2.9)
$$\begin{aligned} y^{n+2}(1+x)^{n+2}&= (1+x)y^{n+2}(1+x)^{n+1}. \end{aligned}$$
(2.10)

Since \(1+x\) is invertible in \(R\), (2.9) gives

$$\begin{aligned} xy^{n+1}=y^{n+1}x, \end{aligned}$$
(2.11)

(2.10) gives

$$\begin{aligned} xy^{n+2}=y^{n+2}x. \end{aligned}$$
(2.12)

Multiply (2.11) on the left by \(y\), from (2.12), one gets

$$\begin{aligned} (xy-yx)y^{n+1}=0. \end{aligned}$$
(2.13)

Similar to the proof of \((2)\Longrightarrow (1)\), we have

$$\begin{aligned} xy=yx. \end{aligned}$$
(2.14)

Hence \(R\) is \(CN\). \(\square \)

Let \(R\) be a \(CN\) ring. Then for any \(n\ge 2\) and any \(a\in N(R)\) and \(b\in R\), we have \((ab)^n=a^nb^n=b^na^n\). But the converse is not true in general.

Example 2.1

Let \(D\) be a division ring and \(R=\left( \begin{array}{cc}D&{}D\\ 0&{}D\end{array}\right) \). Then \(N(R)=\left( \begin{array}{cc}0&{}D\\ 0&{}0\end{array}\right) \) with \(N(R)^2=0\). Since \(N(R)\) is an ideal of \(R\), for any \(n\ge 2\), any \(A\in N(R)\) and \(B\in R\), we have \((AB)^n=A^nB^n=B^nA^n=0\). But \(R\) is not \(CN\).

Theorem 2.4

Let \(R\) be a \(CN\) ring and \(n\ge 1\). If for any \(x, y\in R\backslash N(R)\), \((xy)^{n+k}=x^{n+k}y^{n+k}\), where \(k=0, 1,2\), then \(R\) is commutative.

Proof

It follows immediately from the result in [11].\(\square \)

Lemma 2.2

Let \(R\) be a semiprime ring and \(n\ge 2\). If for any \(a\in N(R)\) and \(b\in R\), \((ab)^n=a^nb^n\), then \(R\) is reduced.

Proof

Let \(a\in R\) with \(a^n=0\). Then \((ax)^n=0\) for each \(x\in R\). If \(a\ne 0\), then \(aR\) is a nonzero nil right ideal of \(R\) satisfying the identity \(z^n=0\) for all \(z\in aR\). Now by [8, Lemma 1.1], \(R\) has a nonzero nilpotent ideal which is a contradiction since \(R\) is semiprime. Thus \(a=0\), this implies \(R\) is reduced. \(\square \)

Theorem 2.5

Let \(R\) be a semiprime ring and \(n\ge 1\). If for any \(x\in R\backslash J(R)\) and \( y\in R\), \((xy)^{n+k}=x^{n+k}y^{n+k}\), where \(k=0, 1\), then \(R\) is commutative.

Proof

If \(N(R)\cap J(R)=0\), then by Lemma 2.2, \(R\) is reduced. If \(N(R)\cap J(R)\ne 0\), then there exists \(0\ne a\in N(R)\cap J(R)\) with \(a^2=0\). By the hypothesis, for any \(y\in R\), we have

$$\begin{aligned} (1+a)^{n+k}(ya)^{n+k}=((1+a)ya)^{n+k}. \end{aligned}$$
(2.15)

Clearly, for any \(i\ge 1\), one gets

$$\begin{aligned} ((1+a)ya)^i=(1+a)(ya)^i. \end{aligned}$$
(2.16)

Hence

$$\begin{aligned} (1+a)^{n+k}(ya)^{n+k}=(1+a)(ya)^{n+k}. \end{aligned}$$
(2.17)

Since \(1+a\) is invertible in \(R\) and \((1+a)^i=1+ia\) for each \(i\ge 1\), we have

$$\begin{aligned} (n+k-1)a(ya)^{n+k}=0, k=0, 1. \end{aligned}$$
(2.18)

This implies

$$\begin{aligned} a(ya)^{n+1}=0. \end{aligned}$$
(2.19)

Hence

$$\begin{aligned} (ay)^{n+2}=0. \end{aligned}$$
(2.20)

This leads to \(aR\) is a nonzero nil right ideal of \(R\) satisfying the identity \(z^{n+2}=0\) for all \(z\in aR\). Now by [8, Lemma 1.1], \(R\) has a nonzero nilpotent ideal which is a contradiction since \(R\) is semiprime. Thus \(N(R)\cap J(R)=0\) and so \(R\) is reduced.

Now suppose \(x, y\in R\). If \(x, 1+x\notin J(R)\), then by the hypothesis, we have

$$\begin{aligned}&\displaystyle x^{n+1}y^{n+1}=x^ny^nxy, \end{aligned}$$
(2.21)
$$\begin{aligned}&\displaystyle (x+1)^{n+1}y^{n+1}=(x+1)^ny^n(x+1)y. \end{aligned}$$
(2.22)

They give

$$\begin{aligned} x^n(xy^n-y^nx)y&= 0. \end{aligned}$$
(2.23)
$$\begin{aligned} (x+1)^n(xy^n-y^nx)y&= 0. \end{aligned}$$
(2.24)

From Lemma 2.1, (2.23) and (2.24), one gets

$$\begin{aligned} (xy^n-y^nx)y=0. \end{aligned}$$
(2.25)

If \(x\in J(R)\), then \(x+1\) is invertible in \(R\), so

$$\begin{aligned} (x+1)^{n+1}y^{n+1}=(x+1)^ny^n(x+1)y. \end{aligned}$$
(2.26)

This gives

$$\begin{aligned} (xy^n-y^nx)y=0. \end{aligned}$$
(2.27)

If \(x\notin J(R)\) and \(1+x\in J(R)\), then \(x\) is invertible in \(R\), so (2.21) implies

$$\begin{aligned} (xy^n-y^nx)y=0. \end{aligned}$$
(2.28)

Hence, in any case, we have

$$\begin{aligned} (xy^n-y^nx)y&= 0, \end{aligned}$$
(2.29)
$$\begin{aligned} (y(xy^n-y^nx))^2&= 0. \end{aligned}$$
(2.30)

Since \(R\) is reduced, one gets

$$\begin{aligned} y(xy^n-y^nx)=0. \end{aligned}$$
(2.31)

Clearly, for any \(r\in R\), we have

$$\begin{aligned} ((xy^n-y^nx)ry)^2=0. \end{aligned}$$
(2.32)

Hence, for any \(r\in R\), we have

$$\begin{aligned} (xy^n-y^nx)ry=0. \end{aligned}$$
(2.33)

that is,

$$\begin{aligned} (xy^n-y^nx)Ry=0. \end{aligned}$$
(2.34)

Thus

$$\begin{aligned} (xy^n-y^nx)R(xy^n-y^nx)=0. \end{aligned}$$
(2.35)

Since \(R\) is semiprime, one gets

$$\begin{aligned} xy^n=y^nx. \end{aligned}$$
(2.36)

Since \(R\) has no nonzero nil ideals, by [9, Theorem 2], \(R\) is commutative.\(\square \)

Corollary 2.1

Let \(R\) be a primitive ring and \(n\ge 1\). If for any \(x, y\in R\), \((xy)^k=x^ky^k\), where \(k=n, n+1\), then \(R\) is a field.

Proof

By Theorem 2.5, \(R\) is commutative. We claim that \(R\) is a division ring. If not, there exists a subring \(S\) of \(R\) such that \(S\) is isomorphic to \(2\times 2\) full matrix ring \(M_2(D)\) over a division ring \(D\). Clearly, for any \(x, y\in S\), \((xy)^k=x^ky^k\), where \(k=n, n+1\), hence \(M_2(D)\) satisfies the same conditions. Now let \(A=\left( \begin{array}{cc}0&{}1\\ 0&{}0\end{array}\right) , B=\left( \begin{array}{cc}0&{}0\\ 1&{}0\end{array}\right) \), \((AB)^{n+1}\ne A^{n+1}B^{n+1}\), which is a contradiction. Hence \(R\) is a division ring, and so \(R\) is a field.\(\square \)

Theorem 2.6

\(R\) is a \(CN\) ring if and only if for some positive integer \(n\ge 1, m> 1\) and any \(x\in R\) and \(y\in N(R)\), \([xy-x^ny^m, x]=0\).

Proof

One direction is clear.

Now assume that \(n\ge 1, m> 1\) such that for any \(x\in R\) and \(y\in N(R)\), we have \([xy-x^ny^m, x]=0\). Since \(y\in N(R)\), there exists \(p\ge 1\) such that \(y^{m^p}=0\). The equation \([xy-x^ny^m, x]=0\) gives

$$\begin{aligned} x[x, y]=x^n[x, y^m]. \end{aligned}$$
(2.37)

Since \(y^m\in N(R)\), substitute \(y^m\) for \(y\) in (2.37), one gets

$$\begin{aligned} x^2[x, y]=x^{n+1}[x, y^m]=x^n(x[x, y^m])=x^{2n}[x, y^{m^2}]. \end{aligned}$$
(2.38)

Hence

$$\begin{aligned} x^p[x, y]=x^{np}[x, y^{m^p}]. \end{aligned}$$
(2.39)

This implies

$$\begin{aligned} x^p[x,y]=0. \end{aligned}$$
(2.40)

Since (2.40) holds for all \(x\in R\), substitute \(x+1\) for \(x\) and use Lemma 2.1, we have \([x, y]=0\). Hence \(R\) is \(CN\). \(\square \)

Theorem 2.7

Let \(R\) be a ring and \(n\ge 1\). If for any \(x\in R\backslash N(R)\) and \(y\in R\), \((xy)^k=x^ky^k\), \(k=n, n+1, n+2\), then \(R\) is commutative.

Proof

Suppose that \(x, y\in R\). If \(x\in N(R)\), then \(1+x\) is invertible. By the hypothesis,

$$\begin{aligned} (1+x)^ky^k=((1+x)y)^k, k=n, n+1, n+2. \end{aligned}$$
(2.41)

Hence, one gets

$$\begin{aligned} (1+x)^{n-1}y^n&= y((1+x)y)^{n-1}, \end{aligned}$$
(2.42)
$$\begin{aligned} (1+x)^ny^{n+1}&= y((1+x)y)^n, \end{aligned}$$
(2.43)
$$\begin{aligned} (1+x)^{n+1}y^{n+2}&= y((1+x)y)^{n+1}. \end{aligned}$$
(2.44)

Multiply (2.42) on the right by \((1+x)y\), from (2.43), one gets

$$\begin{aligned} y^nxy=xy^{n+1}. \end{aligned}$$
(2.45)

Multiply (2.43) on the right by \((1+x)y\), from (2.44), one gets

$$\begin{aligned} y^{n+1}xy=xy^{n+2}.\end{aligned}$$
(2.46)

Multiply (2.45) on the left by \(y\), from (2.46), one gets

$$\begin{aligned} (xy-yx)y^{n+1}=0.\end{aligned}$$
(2.47)

If \(x\notin N(R)\), then by the hypothesis, one gets

$$\begin{aligned} (xy)^k=x^ky^k, k=n, n+1, n+2.\end{aligned}$$
(2.48)

If \(1+x\in N(R)\), then \(x\) is invertible in \(R\). Similar to the proof of above, (2.48) implies

$$\begin{aligned} (xy-yx)y^{n+1}=0.\end{aligned}$$
(2.49)

If \(1+x\notin N(R)\), then one has

$$\begin{aligned} ((1+x)y)^k=(1+x)^ky^k, k=n, n+1, n+2.\end{aligned}$$
(2.50)

Similar to the proof of Theorem 2.6, (2.48) and (2.50) imply

$$\begin{aligned} (xy-yx)y^{n+1}=0. \end{aligned}$$
(2.51)

Hence, (2.47), (2.49) and (2.51) imply that in any case, one has

$$\begin{aligned} (xy-yx)y^{n+1}=0. \end{aligned}$$
(2.52)

Substitute \(y+1\) for \(y\) in (2.52), one gets

$$\begin{aligned} (xy-yx)(y+1)^{n+1}=0\end{aligned}$$
(2.53)

By Lemma 2.3, (2.52), and (2.53), one obtains \(xy=yx\). Thus \(R\) is commutative. \(\square \)

Following [5], an element \(x\) of \(R\) is called weakly clean if \(x=u+e\) or \(x=u-e\) for some \(u\in U(R)\) and \(e\in E(R)\). The ring \(R\) is said to be weakly clean if all of its elements are weakly clean. Clean rings are weakly clean. But the converse is not true because of the example \(Z_{(3)}\cap Z_{(5)}\) where \(Z_{(p)}=\{ \frac{r}{s}| p \) does not divide \(s\} \). An element \(x\) of \(R\) is called weakly exchange if there exists \(e\in E(R)\) such that \(e\in xR\) and \(1-e\in (1-x)R\) or \(1-e\in (1+x)R\). The ring \(R\) is said to be weakly exchange if all of its elements are weakly exchange. Clearly, exchange elements are weakly exchange. Checking carefully the proof of [5, Theorem 2.1], we find that weakly clean elements are weakly exchange, so weakly clean rings and exchange rings are all weakly exchange. In fact, [5, Theorem 2.1] showed that Abel weakly exchange rings are weakly clean. In this paper, we obtain that \(NLI\) weakly exchange rings are weakly clean.

Theorem 2.8

Let \(R\) be an \(NLI\) ring and \(x\in R\).

  1. (1)

    If \(x\) is weakly exchange, then \(x\) is weakly clean.

  2. (2)

    If \(x\) is exchange, then \(x\) is clean.

  3. (3)

    If \(R\) is a weakly exchange ring, then \(R\) is a weakly clean ring.

  4. (4)

    If \(R\) is an exchange ring, then \(R\) is a clean ring.

Proof

(1) Let \(e\in E(R)\) such that \(e\in xR\) and \(1-e\in (1-x)R\) or \(1-e\in (1+x)R\). Write \(e=xy\) for some \(y=ye\in R\). If \(1-e\in (1-x)R\), then let \(1-e=(1-x)z\) for some \(z=z(1-e)\in R\). By computing, we have \((x-(1-e))(y-z)=1-(1-e)y-ez\). Since \(R\) is a \(NLI\) ring and \((1-e)y=(1-e)ye\in N(R)\), \((1-e)yez-ez(1-e)y\in N(R)\), that is, \((1-e)yz-ezy\in N(R)\). Hence there exists \(n\ge 1\) such that \(((1-e)yz-ezy)^n=0\). By computing, we have \(((1-e)yz)^n+(-1)^n(ezy)^n=0\), this leads to \(((1-e)yz)^n=(ezy)^n=0\). Since \((ez+(1-e)y)^2=ezy+(1-e)yz\), \((ez+(1-e)y)^{2n}=(ezy)^n+((1-e)yz)^n=0\). Hence \(1-ez-(1-e)y\in U(R)\), that is, \((x-(1-e))(y-z)\in U(R)\). Let \(u\in R\) such that \(((x-(1-e))(y-z))u=1\). Let \(g=((y-z)u)(x-(1-e))\). Then \((x-(1-e))g=x-(1-e)\) and \(g^2=g\). Let \(h=(x-(1-e))-g(x-(1-e))\). Then \(hg=h\), \(gh=0\) and \(h^2=0\). Since \(R\) is an \(NLI\) ring, \((y-z))uh-h(y-z))u\in N(R)\), that is, \((y-z))uh-(1-g)\in N(R)\). Hence there exists \(n\ge 1\) such that \(((y-z))uh-(1-g))^n=0\), this gives \(1-g=d(y-z)uh\) for some \(d\in R\). Thus \(1-g=d(y-z))uh=d(y-z))uhg=(1-g)g=0\), so \(((y-z)u)(x-(1-e))=g=1\). Hence \(x-(1-e)\in U(R)\). If \(1-e\in (1+x)R\), then let \(1-e=(1+x)w\) for some \(w=w(1-e)\in R\). By computing, we have \((x+(1-e))(y+w)=1+(1-e)y-ew\). Similarly, we obtain \(1+(1-e)y-ew\in U(R)\), this gives \(x+(1-e)\in U(R)\). We are done. (2) It has been have shown in (1).

(3) and (4) are immediate corollaries of (1) and (2), respectively.\(\square \)

A ring \(R\) is called a generalized CN ring if for any \(a\in N(R)\) and \(b\in R\), \(ab=0\) implies \(aRb=0\) or there exists \(c\in R\) such that \(0\ne acb\in Z(R)\). Clearly, \(CN\) rings are generalized \(CN\). But the converse is not true. For example, the ring \(R\) in Example 2.5 is a generalized \(CN\) ring, but \(R\) is not \(CN\).

A ring \(R\) is called left SF if every simple left \(R-\)module is flat. In [13, Remark 3.13], it is shown that if \(R\) is a reduced left \(SF\) ring, then \(R\) is strongly regular. We can generalize this result as follows.

Proposition 2.1

Let \(R\) be a generalized \(CN\) ring. If \(R\) is a left \(SF\) ring, then \(R\) is a strongly regular ring.

Proof

Let \(a\in R\) with \(aRa=0\). If \(a\ne 0\), then there exists a maximal left ideal \(M\) of \(R\) such that \(r(aR)\subseteq M\). Since \(R\) is a left \(SF\) ring, \(R/M\) is flat as left \(R-\)module. Since \(a\in r(aR)\subseteq M\), \(a=am\) for some \(m\in M\). Since \(R\) is a generalized \(CN\) ring and \(a(1-m)=0\), \(aR(1-m)=0\), or there exists \(c\in R\) such that \(0\ne ac(1-m)\in Z(R)\). If \(aR(1-m)=0\), then \(1-m\in r(aR)\subseteq M\), this gives \(1=(1-m)+m\in M\), a contradiction. Thus there exists \(c\in R\) such that \(0\ne ac(1-m)\in Z(R)\). Since \((ac(1-m))^2=0\), there exists a maximal left ideal \(N\) of \(R\) such that \(l(ac(1-m))\subseteq N\). Since \(R\) is a left \(SF\) ring, \(R/N\) is flat as left \(R-\)module. Then since \(ac(1-m)\in N\), \(ac(1-m)=ac(1-m)n\) for some \(n\in N\). Since \(ac(1-m)\in Z(R)\), \(ac(1-m)=nac(1-m)\), this leads to \(1-n\in l(ac(1-m))\subseteq N\), which implies \(1=(1-n)+n\in N\), a contradiction. Hence \(a=0\), which implies \(R\) is a semiprime ring. Now let \(b\in R\) with \(b^2=0\). Since \(R\) is a generalized \(CN\) ring, either \(bRb=0\) or there exists \(c\in R\) such that \(0\ne bab\in Z(R)\). If there exists \(c\in R\) such that \(0\ne bcb\in Z(R)\), then \(bcbRbcb=bcbbcbR=0\), so \(bcb=0\) because \(R\) is semiprime, which is a contradiction. Hence \(bRb=0\), also, the semiprimeness of \(R\) implies \(b=0\). Hence \(R\) is a reduced ring. By [13, Remark 3.13], \(R\) is a strongly regular ring. \(\square \)

Following [4], a ring \(R\) is said to be semiperiodic if for each \(x\in R \backslash (J(R)\cup Z(R))\), there exist \(m, n \in \mathbb {Z}\), of opposite parity, such that \(x^n-x^m\in N(R)\). Clearly, the class of semiperiodic rings contains all commutative rings, all Jacobson radical rings, and certain non-nil periodic rings.

Lemma 2.3

Let \(R\) be a generalized \(CN\) ring. If \(R\) is a semiperiodic ring, then \(N(R)\subseteq J(R)\)

Proof

Let \(a\in N(R)\) with \(a^k = 0\), and let \(x\in R\). If \(ax\in J(R)\), then \(ax\) is right quasiregular; and if \(ax\in Z(R)\), then \(ax\) is nilpotent and again \(ax\) is right quasiregular. Suppose, then, that \(ax\notin J(R)\cup Z(R)\), in which case [4, Lemma 2.3(iii)] gives \(q\in \mathbb {Z}^+\) and an idempotent \(e\) of form \(axy\) such that \((ax)^q = (ax)^qe\). Since \(e=axy=eaxy=ea(1-e)xy+eaexy=ea(1-e)xy+ea^2(xy)^2=ea(1-e)xy+ea^2(1-e)(xy)^2+ea^2e(xy)^2 =ea(1-e)xy+ea^2(1-e)(xy)^2+ea^3(xy)^3=\cdots =\Sigma _{i=1}^{k-1}ea^i(1-e)(xy)^i+ea^k(xy)^k=\Sigma _{i=1}^{k-1}ea^i(1-e)(xy)^i \). For any \(z\in R\), \(ez(1-e)\in N(R)\) and \((ez(1-e))^2=0\). Since \(R\) is a generalized \(CN\) ring, either \(ez(1-e)Rez(1-e)=0\) or there exists \(c\in R\) such that \(0\ne ez(1-e)cez(1-e)\in Z(R)\). If there exists \(c\in R\) such that \(0\ne ez(1-e)cez(1-e)\in Z(R)\), then \(ez(1-e)cez(1-e)=(ez(1-e)cez(1-e))(1-e)=(1-e)ez(1-e)cez(1-e)=0\), which is a contradiction. Hence \(ez(1-e)Rez(1-e)=0\), which implies \(ez(1-e)\in J(R)\) for any \(z\in R\). Therefore \(e=\Sigma _{i=1}^{k-1}ea^i(1-e)(xy)^i\in J(R)\), this leads to \(e=0\) and \((ax)^q=0\), which shows that \( ax\) is right quasi-regular. Thus \(a\in J(R)\).\(\square \)

Theorem 2.9

If \(R\) is a generalized \(CN\) semiperiodic ring, then \(R/J(R)\) is commutative.

Proof

By [4, Theorem 4.3], \(R\) is either commutative or periodic, so we may assume \(R\) is periodic. Since \(J(R)\) contains no nonzero idempotents, \(J(R)\) is contained in \(N(R)\) and hence \(J(R)=N(R)\) by Lemma 2.3. Thus \(R/J(R)=R/N(R)\) is reduced; and since \(R/N(R)\) is also semiperiodic, it is commutative by [4, Theorem 4.4]. \(\square \)

Theorem 2.10

Let \(R\) be a generalized \(CN\) semiperiodic ring. Then

  1. (1)

    \(N(R)\) is an ideal of \(R\).

  2. (2)

    If \(J(R)\ne N(R)\), then \(R\) is commutative.

Proof

In the proof of Theorem 2.9, we obtain that if \(R\) is not commutative, then \(J(R)=N(R)\). Hence \((2)\) holds and \((1)\) also holds for noncommutative ring \(R\). But also if \(R\) is commutaive, \(N(R)\) is an ideal; hence \((1)\) holds in any case. \(\square \)