1 Introduction

Let R be a prime ring of characteristic different from 2 with center Z(R), extended centroid C, right Martindale quotient ring \(Q_r\), and symmetric Martindale quotient ring Q.

An additive map \(d:R\rightarrow R\) is a derivation on R if \(d(xy)=d(x)y+xd(y)\) for all \(x,y\in R\). Let \(a\in R\) be a fixed element. A map \(d:R\rightarrow R\) defined by \(d(x)=[a,x]=ax-xa\), \(x\in R\), is a derivation on R, which is called the inner derivation defined by a. Many results in the literature indicate how the global structure of a ring R is often tightly connected to the behavior of additive maps defined on R. A well-known result of Posner [18], states that if d is a derivation of R such that \([d(x),x]\in Z(R)\) for any \(x\in R\), then either \(d=0\) or R is commutative.

In this paper, we study the structure of prime and semiprime rings having skew derivations satisfying strong commutativity preserving conditions. Specifically, let \(\alpha \) be an automorphism of a ring R. An additive map \(D:R\rightarrow R\) is called an \(\alpha \)-derivation (or a skew derivation) on R if \(D (xy)=D(x)y+\alpha (x)D(y)\) for all \(x,y\in R\). In this case, \(\alpha \) is called an associated automorphism of D. Basic examples of \(\alpha \)-derivations are the usual derivations and the map \(\alpha -id\), where id denotes the identity map. Let \(b\in Q\) be a fixed element. Then a map \(D:R\rightarrow R\) defined by \(D(x)=bx-\alpha (x)b\), \(x\in R\), is an \(\alpha \)-derivation on R and it is called an inner \(\alpha \)-derivation (an inner skew derivation) defined by b. If a skew derivation D is not inner, then it is called outer.

If \(\mathcal {S}\subseteq \mathcal {R}\), the map \(F: \mathcal {R}\rightarrow \mathcal {R}\) is called commutativity preserving on \(\mathcal {S}\) if \([x,y]=0\) implies \([F(x),F(y)]=0\); it is called strong commutativity preserving (for brevity we will always say SCP) on \(\mathcal {S}\) if \([F(x),F(y)]=[x,y]\), for all \(x,y \in \mathcal {S}\).

In [1], Bell and Daif proved that if R is a semiprime ring admitting a derivation d which is SCP on the right ideal I of R, then \(I\subseteq Z\). The natural possibility when an additive map preserves commutativity appears in a paper by Bresar and Miers [2]. They showed that any additive map F which is SCP on a semiprime ring R is of the form \(F(x)=\lambda x+\mu (x)\), where \(\lambda \in C\), \(\lambda ^{2}=1\), and \(\mu :R \rightarrow C\) is an additive map of R into C.

Later in [15], Lin and Liu extended this result to Lie ideals, in case the ring R is prime. More precisely they proved that if L is a non-central Lie ideal of R and F is an additive map satisfying \([F(x),F(y)]-[x,y]\in C\) for all \(x,y \in L\), then \(F(x)=\lambda x+\mu (x)\), where \(\lambda \in C\), \(\lambda ^2=1\), and \(\mu :R\rightarrow C\), unless when \(char(R)=2\) and R satisfies the standard identity \(s_4\) of degree 4.

More recently, in [16] Liu showed that if R is a semiprime ring, I a non-zero right ideal of R, F, and G non-zero skew derivations of R, with associated automorphism \(\alpha \) such that \([F(x),G(y)]=[x,y]\) for all \(x,y \in I\), then \([\alpha (x),x]=0\) for any \(x\in I\). Moreover, if \(\alpha (I)\subseteq I\) then \(\sigma (I)\subseteq Z(R)\). Finally, it is proved that if R is prime then R is commutative.

Here we continue this line of investigation and we examine what happens in case F and G are skew derivations of R such that \([F(x),G(y)]_m=[x,y]^n\) for all \(x,y \in I\), where I is a non-zero ideal of R and \(m, n \ge 1\) are positive integers.

The results we obtained are the following:

Theorem 1

Let R be a prime ring of characteristic different from 2, C its extended centroid, Z(R) its center, I a non-zero ideal of R, F, and G non-zero skew derivations of R with associated automorphism \(\alpha \) and mn positive integers such that

$$\begin{aligned}{}[F(x),G(y)]_m=[x,y]^n \,\mathrm{{for \, all}}~ x,y \in I. \end{aligned}$$

Then R is commutative.

Theorem 2

Let R be a semiprime ring of characteristic different from 2, C its extended centroid, Z(R) its center, F and G non-zero skew derivations of R with associated automorphism \(\alpha \), mn positive integers such that

$$\begin{aligned}{}[F(x),G(y)]_m=[x,y]^n \,\mathrm{for \, all}~ x,y \in R. \end{aligned}$$

Then R is commutative.

2 Preliminaries

We denote the set of all skew derivations on Q by \(\mathrm{SDer}(Q)\). By a skew-derivation word, we mean an additive map \(\Delta \) of the form \(\Delta = d_{1},d_{2},\dots d_{m}\), with each \(d_{i}\in \mathrm{SDer}(Q)\). A skew-differential polynomial is a generalized polynomial with coefficients in Q, of the form \(\Phi (\Delta _{j}(x_{i}))\) involving non-commutative indeterminates \(x_{i}\) on which the derivations words \(\Delta _{j}\) act as unary operations. The skew-differential polynomial \(\Phi (\Delta _{j}(x_{i}))\) is called a skew-differential identity on a subset T of Q if it vanishes for any assignment of values from T to its indeterminates \(x_{i}\).

In order to prove our result, we need to recall the following known facts:

Fact 1

In [9], Chuang and Lee investigate polynomial identities with skew derivations. More precisely as a consequence of in [9, Theorem 1], we have that if D is an outer skew derivation of R which satisfies the generalized polynomial identity \(\Phi (x_i,D(x_j))\), then \(\Phi (x_i,y_j)\) is also a generalized polynomial identity for R, where \(x_i\) and \(y_j\) are distinct indeterminates.

Fact 2

Let R be a prime ring and I a two-sided ideal of R. Then I, R, and Q satisfy the same generalized polynomial identities with coefficients in Q (see [5]). Furthermore, I, R, and Q satisfy the same generalized polynomial identities with automorphisms [7, Theorem 1].

Fact 3

Recall that, in case \(char(R)=0\), an automorphism \(\alpha \) of Q is called Frobenius if \(\alpha (x)=x\) for all \(x\in C\). Moreover, in case \(char(R)=p\ge 2\), an automorphism \(\alpha \) is Frobenius if there exists a fixed integer t such that \(\alpha (x)=x^{p^t}\) for all \(x\in C\). In [7, Theorem 2], Chuang proves that if \(\Phi (x_i,\alpha (x_i))\) is a generalized polynomial identity for R, where R is a prime ring and \(\alpha \in Aut(R)\) an automorphism of R which is not Frobenius, then R also satisfies the non-trivial generalized polynomial identity \(\Phi (x_i,y_i)\), where \(x_i\) and \(y_i\) are distinct indeterminates.

Fact 4

Let R be a domain and \(\alpha \in Aut(R)\) an automorphism of R which is outer. In [13], Kharchenko proves that if \(\Phi (x_i,\alpha (x_i))\) is a generalized polynomial identity for R, then R also satisfies the non-trivial generalized polynomial identity \(\Phi (x_i,y_i)\), where \(x_i\) and \(y_i\) are distinct indeterminates.

Finally, let us mention that if R is a prime ring satisfying a non-trivial generalized polynomial identity and \(\alpha \) an automorphism of R such that \(\alpha (x)=x\) for all \(x\in C\), then \(\alpha \) is an inner automorphism of R [3, Theorem 4.7.4].

Lemma 1

Let R be a semiprime ring of characteristic different from 2, Q its symmetric Martindale quotient ring, F and G non-zero skew derivations of R, mn positive integers such that \(n\ge 2\) and

$$\begin{aligned} \left[ F(x),G(y)\right] _m=[x,y]^n~ \mathrm{{for \, all}}~ x,y \in R. \end{aligned}$$
(2.1)

Then R is commutative.

Proof

Assume first that R is prime. From the relation (2.1), we have both

$$\begin{aligned} \left[ F(x+z),G(y)\right] _m=[x+z,y]^n \end{aligned}$$
(2.2)

and

$$\begin{aligned} \left[ F(x+z),G(y)\right] _m=[F(x),G(y)]_m+[F(z),G(y)]_m=[x,y]^n+[z,y]^n. \end{aligned}$$
(2.3)

From (2.2) and (2.3), it follows that R satisfies the polynomial identity \([x+z,y]^n-[x,y]^n-[z,y]^n\). By Posner’s theorem [11, Theorem 2 p. 57, Lemma 1 p. 89], \(Q\subseteq M_k(E)\), the ring of \(k\times k\) matrices over a field E. Moreover, Q and \(M_k(E)\) satisfy the same polynomial identities. If \(k\ge 2\) and for \(x=e_{21}\), \(y=e_{11}\), \(z=e_{11}\) we have the contradiction

$$\begin{aligned} 0=\left[ e_{21}+e_{12},e_{11}\right] ^n- \left[ e_{21},e_{11}\right] ^n-[e_{12},e_{11}]^n=(e_{21}-e_{12})^n\ne 0. \end{aligned}$$

Thus, \(k=1\) and we have that Q is commutative, as well as R.

Let now R be semiprime. Since R is a semiprime ring for which \([x+z,y]^n-[x,y]^n-[z,y]^n\) is a polynomial identity, R is a subdirect product of prime rings \(R_\alpha \), each of which still satisfies the identity \([x+z,y]^n-[x,y]^n-[z,y]^n\). In this case, by the above argument, any \(R_\alpha \) is commutative. Thus, we conclude that R must be commutative. \(\square \)

Lemma 2

Let R be a non-commutative prime ring of characteristic different from 2, F and G non-zero skew derivations of R, m a positive integer such that

$$\begin{aligned}{}[F(x),G(y)]_m=[x,y]~\mathrm{{for \, all}}~x,y \in R. \end{aligned}$$

Then \(char(R)=p > 0\) and m is odd.

Proof

For any \(x,y \in R\) we have

$$\begin{aligned}{}[F(x),G(y+y)]_m=[x,y+y]=2[x,y] \end{aligned}$$
(2.4)

and also, by computing the m-th commutator

$$\begin{aligned}{}[F(x),G(2y)]_m=[F(x),2G(y)]_m=2^m[F(x),G(y)]_m=2^m[x,y]. \end{aligned}$$
(2.5)

Comparing (2.4) with (2.5) it follows \((2^m-2)[x,y]=0\) for all \(x,y \in Q\). Hence, since R is not commutative and \(char(R)\ne 2\), \(char(R)=p\ne 0\) (since \(2^m-2\equiv 0\), modulo p).

Moreover,

$$\begin{aligned} -[x,y]=[x,-y]=[F(x),G(-y)]_m=(-1)^m[F(x),G(y)]_m=(-1)^m[x,y] \end{aligned}$$

which implies that m must be an odd integer. \(\square \)

3 Results

Lemma 3

Let R be a non-commutative prime ring of characteristic different from 2, which is isomorphic to a dense subring of the ring of linear transformations of a vector space V over a division ring D, \(\alpha :R\rightarrow R\) an automorphism of R and \(q, u\in R\) such that

$$\begin{aligned}{}[qx,\alpha (y)u-uy]_m=[x,y] \end{aligned}$$

for all \(x,y\in R\). Then \(dim_DV=1\).

Proof

By Theorem 1 in [7], R and Q satisfy the same generalized polynomial identities with automorphisms and hence \([qx,\alpha (y)u-uy]_m-[x,y]\) is a generalized polynomial identity for Q. We assume that \(dim_DV\ge 2\) and prove that a contradiction follows.

By [12, p. 79], there exists a semilinear automorphism \(T\in End(V)\) such that \(\alpha (x)=TxT^{-1}\) for all \(x\in Q\). Hence, Q satisfies

$$\begin{aligned} \left[ qx,TyT^{-1}u-uy\right] _m=[x,y]. \end{aligned}$$

We notice that, if for any \(v\in V\) there exists \(\lambda _v \in D\) such that \(T^{-1}uv=v\lambda _v\), then, by a standard argument, it follows that there exists a unique \(\lambda \in D\) such that \(T^{-1}uv=v \lambda \), for all \(v\in V\) (see for example [8, Lemma 1]). In this case,

$$\begin{aligned} \alpha (x)uv=\left( TxT^{-1}\right) uv=Txv\lambda \end{aligned}$$

and

$$\begin{aligned} \biggl (\alpha (x)u-ux\biggr )v=T(xv\lambda )-uxv=T\left( T^{-1}uxv\right) -uxv=0 \end{aligned}$$

which implies that \(\biggl (\alpha (x)u-ux\biggr )V=(0)\). Thus \(\alpha (x)u-ux=0\) for any \(x\in R\), since V is faithful, and by our assumption \(0=[qx,\alpha (y)u-uy]_m=[x,y]\) for any \(x,y \in R\), which is a contradiction, since R is not commutative.

Therefore, there exists \(v\in V\) such that \(\{v,T^{-1}uv\}\) are linearly D-independent. By the density of Q, there exist \(r,s\in Q\) such that

$$\begin{aligned} rv=0,\quad rT^{-1}uv=v,\quad sv=T^{-1}uv,\quad sT^{-1}uv=\left( T^{-1}u\right) ^2v. \end{aligned}$$

Hence,

$$\begin{aligned} qrv=0,\quad \left( TsT^{-1}u-us\right) v=0,\quad [r,s]v=v \end{aligned}$$

and, by the main assumption, we get the contradiction

$$\begin{aligned} 0=\biggl ( \left[ qr,TsT^{-1}u-us\right] _m-[r,s]\biggr )v=-v\ne 0. \end{aligned}$$

\(\square \)

Lemma 4

Let R be a non-commutative prime ring of characteristic different from 2, C its extended centroid, Z(R) its center, I a non-zero ideal of R, F, and G non-zero skew derivations of R with associated automorphism \(\alpha \), m a positive integer such that

$$\begin{aligned}{}[F(x),G(y)]_m=[x,y]\quad \text{ for } \text{ all } \quad x,y \in I. \end{aligned}$$

Then there exist \(p,q,u \in Q\), with p invertible such that \(F(x)=pxp^{-1}q-qx\) and \(G(x)=pxp^{-1}u-ux\), for all \(x\in R\).

Proof

It is known that I, R, and Q satisfy the same generalized polynomial identities with skew derivations and automorphisms, so that \([F(x),G(y)]_m=[x,y]\), for all \(x,y \in Q\). Notice that if \(m=1\) then the result follows by [15]. Thus we consider \(m\ge 2\).

Fix \(y_0 \in R\) and denote \(z_0=G(y_0)\). Then for any \(x\in Q\) we get \([F(x),z_0]_m=[x,y_0]\). In case F is an outer skew derivation of R, it is known that Q satisfies \([t,z_0]_m-[x,y_0]\), and in particular for \(t=0\), we have \([x,y_0]=0\), for all \(x\in Q\). This implies \(y_0 \in C\). We may repeat this argument for any \(y_0\in Q\) and conclude that Q is commutative, a contradiction. Therefore, in the sequel, we always assume that F is an inner skew derivation of R. Thus, there exists \(0\ne q \in Q\) such that \(F(x)=\alpha (x)q-qx\), for all \(x\in R\).

Fix now \(x_0 \in R\) and denote \(z_0=F(x_0)\). Then for any \(y\in Q\) we get \([z_0,G(y)]_m=[x_0,y]\). In case G is an outer skew derivation of R, it is known that Q satisfies \([z_0,t]_m-[x_0,y]\), and in particular for \(t=0\), we have \([x_0,y]=0\) for all \(y\in Q\). This implies \(x_0 \in C\). We may repeat this argument for any \(x_0\in Q\) and conclude that Q is commutative, a contradiction. Therefore, in the sequel, we always assume that G is an inner skew derivation of R. Thus, there exists \(0\ne u \in Q\) such that \(G(x)=\alpha (x)u-ux\), for all \(x\in R\). Hence we have that

$$\begin{aligned}{}[\alpha (x)q-qx,\alpha (y)u-uy]_m=[x,y]~\mathrm{for\, all}~ x,y \in Q. \end{aligned}$$
(3.1)

We assume that \(\alpha \) is not inner. In this case

$$\begin{aligned}{}[tq-qx,\alpha (y)u-uy]_m-[x,y] \end{aligned}$$
(3.2)

is satisfied by Q.

If \(\alpha \) is not Frobenius, then by (3.2) it follows that Q satisfies the generalized identity

$$\begin{aligned}{}[tq-qx,zu-uy]_m-[x,y] \end{aligned}$$
(3.3)

and in particular for \(x=y=0\) in (3.3), \([tq,zu]_m=0\), for all \(t,z \in Q\). By applying the result in [14], and since \(q\ne 0\) and \(u\ne 0\), it follows that either Qq is a non-zero central left ideal of Q or Qu is a non-zero central left ideal of Q. In any case, Q is commutative, a contradiction.

Thus, we consider the case when \(\alpha \) is Frobenius. Again by (3.2), Q satisfies

$$\begin{aligned}{}[-qx,\alpha (y)u-uy]_m-[x,y]. \end{aligned}$$
(3.4)

In case both \(q\in C\) and \(u\in C\), by (3.4) we have that

$$\begin{aligned} -qu^m[x,\alpha (y)-y]_m-[x,y] \end{aligned}$$
(3.5)

is satisfied by Q. Then by the main Theorem in [6] and since \(qu^m\ne 0\), Q satisfies a non-trivial generalized polynomial identity. On the other hand, if either \(q\notin C\) or \(u\notin C\), then by (3.4) and again by the main Theorem in [6], Q satisfies a non-trivial generalized polynomial identity (hence Q is a GPI-ring in any case). Therefore, by [17, Theorem 3], Q is a primitive ring and it is a dense subring of the ring of linear transformations of a vector space V over a division ring D. Moreover, Q contains non-zero linear transformations of finite rank. By Lemma 3, it follows \(dim_DV=1\), that is Q is a division algebra which is finite dimensional over C. If C is finite, then Q is finite, so that Q is a commutative field, which is a contradiction. So, we assume in all that follows that C is infinite and \(char(Q)=p > 0\).

By using (3.4) in (3.2) it follows that

$$\begin{aligned}{}[tq,\alpha (y)u-uy]_m \end{aligned}$$
(3.6)

is satisfied by Q. Let \(s\ge 1\) be such that \(p^s\ge m\), and \(k=p^s\), then by (3.6) we have that Q satisfies

$$\begin{aligned} \biggl [tq,\alpha (y)u-uy\biggr ]_{k} \end{aligned}$$

that is

$$\begin{aligned} \biggl [tq,\biggl (\alpha (x)u-ux\biggr )^k\biggr ]=0~\mathrm{for \, all}~ t,x,\in Q. \end{aligned}$$
(3.7)

Since \(\alpha \) is Frobenius, \(\alpha (\gamma )=\gamma ^{p^h}\), for all \(\gamma \in C\) and some non-zero fixed integer h. Moreover, there exists \(\lambda \in C\) such that \(\lambda ^{p^h}\ne \lambda \), that is \(\lambda ^{p^h-1}\ne 1\).

In particular, we choose \(\gamma \in C\) such that \(\gamma =\lambda ^{p^h-1}\ne 0\). In the relation (3.7), we replace x by \(\lambda x\) and obtain that Q satisfies

$$\begin{aligned} \biggl [tq,\biggl (\lambda ^{p^h}\alpha (x)u-\lambda ux\biggr )^k\biggr ] \end{aligned}$$

that is

$$\begin{aligned} \biggl [tq,\biggl (\gamma \alpha (x)u-ux\biggr )^k\biggr ]. \end{aligned}$$

Let \(\Phi \) and \(\Omega \) be maps on Q, such that \(\Phi (x) = -ux\) and \(\Omega (x) = \alpha (x)u\), for any element x of Q. Thus it follows that \([tq,(\Phi (x)+\gamma \Omega (x))^k]=0\) for all \(x\in Q\). Expanding \((\Phi (x)+\gamma \Omega (x))^k\), we get

$$\begin{aligned} \sum _{i=0}^n \gamma ^i \left( \sum _{(i,k-i)}\varphi _1\cdot \varphi _2\cdots \varphi _k\right) =0 \end{aligned}$$

where the inside summations are taken over all permutations of \(k-i\) terms of the form \(\Phi (x)\) and i terms of the form \(\Omega (x)\). This means that each summation inside has exactly \(k-i\) terms of the form \(\Phi (x)\) and i terms of the form \(\Omega (x)\) but in some different order. For any \(j=0,\ldots ,k\), denote \(y_j=\sum _{(j,k-j)}\varphi _1\cdot \varphi _2\cdots \varphi _k\), then we can write

$$\begin{aligned} (\Phi (x)+\gamma \Omega (x))^k=y_0+\gamma y_1+\gamma ^2 y_2+\ldots +\gamma ^k y_k \end{aligned}$$

so that

$$\begin{aligned} \biggl [tq,y_0+\gamma y_1+\gamma ^2 y_2+\ldots +\gamma ^k y_k\biggr ]=0. \end{aligned}$$

that is

$$\begin{aligned}{}[tq,y_0]+\gamma [tq,y_1]+\gamma ^2[tq,y_2]+\ldots +\gamma ^k [tq,y_k]=0. \end{aligned}$$

Here, we denote by \(z_i=[tq,y_i]\), for any \(i=1,\ldots ,k\), then

$$\begin{aligned} z_0+\gamma z_1+\gamma ^2 z_2+\ldots +\gamma ^k z_k=0. \end{aligned}$$
(3.8)

Replacing in the previous argument \(\lambda \) successively by \(1,\lambda ,\lambda ^2,\ldots ,\lambda ^k\), the equation (3.8) gives the system of equations

$$\begin{aligned} \begin{aligned}&z_0+z_1+z_2+\cdots +z_k=0\\&z_0+\gamma z_1+\gamma ^2z_2+\cdots +\gamma ^kz_k=0\\&z_0+\gamma ^2z_1+\gamma ^4z_2+\cdots +\gamma ^{2k}z_k=0\\&z_0+\gamma ^3z_1+\gamma ^6z_2+\cdots +\gamma ^{3k}z_k=0\\&\cdots \cdots \cdots \cdots \\&z_0+\gamma ^kz_1+\gamma ^{2k}z_2+\cdots +\gamma ^{k^2}z_k=0. \end{aligned} \end{aligned}$$
(3.9)

Moreover, since C is infinite, there exist infinitely many \(\lambda \in C\) such that \(\lambda ^{i(p^h-1)}\ne 1\) for \(i=1,\ldots ,k\), that is there exist infinitely many \(\gamma =\lambda ^{p^h-1} \in C\) such that \(\gamma ^{i}\ne 1\) for \(i=1,\ldots ,k\). Hence, the Vandermonde determinant (associated with the system (3.9))

$$\begin{aligned} \left| \begin{array}{ccccc} 1 &{} 1 &{} \cdots &{} \cdots &{} 1\\ 1 &{} \gamma &{} \gamma ^2 &{} \cdots &{} \gamma ^k\\ 1 &{} \gamma ^2 &{} \gamma ^4 &{} \cdots &{} \gamma ^{2k}\\ \cdots &{} \cdots &{} \cdots &{} \cdots &{} \cdots \\ 1 &{} \gamma ^k &{} \gamma ^{2k} &{} \cdots &{} \gamma ^{k^2} \end{array} \right| =\prod _{0\le i < j \le k} (\gamma ^i - \gamma ^j) \end{aligned}$$

is not zero. Thus, we can solve the above system (3.9) and obtain \(z_i=0\) (\(i=0,\ldots ,k\)). In particular \(z_0=0\), that is Q satisfies

$$\begin{aligned} \biggl [tq,(-ux)^k\biggr ]. \end{aligned}$$
(3.10)

Since \(u\ne 0\) and \(q\ne 0\), then, by (3.10) and [14], either uQ is a non-zero central right ideal of Q or Qq is a non-zero central left ideal of Q. In any case, we get the contradiction that Q is commutative.

The previous argument shows that the automorphism \(\alpha \) must be inner, that is there exists an invertible element \(p \in Q\), such that \(\alpha (x)=pxp^{-1}\), for all \(x\in R\), as required. \(\square \)

The following result is an easy consequence of [4, Theorem 1]. It will be useful in the proof of our result:

Lemma 5

Let R be a prime ring of characteristic different from 2, Z(R) its center, C its extended centroid. Let p be an invertible element of R, d the inner derivation of R, which is induced by p, that is, \(d(x)=[p,x]\) for any \(x\in R\), and \(\beta (x)=pxp^{-1}\) for any \(x\in R\), the inner automorphism of R induced by p. Assume that F is a non-zero skew derivation of R with associated automorphism \(\beta \) and \(0\ne a \in R\) such that

$$\begin{aligned} aF(x)-F(x)a=d(x)~\mathrm{for\, all}~ x \in R. \end{aligned}$$

Then \(d=0\), \(\beta \) is the identity map on R, F is an ordinary derivation of R and \(a\in Z(R)\).

Proof

Firstly we notice that, since F is a skew derivation of R, \(F(xy)=F(x)y+\beta (x)F(y)\). Thus, F is both a right \((1,\beta )\)-generalized skew derivation and a left \((1,\beta )\)-generalized skew derivation of R, in the sense of [4]. Therefore, we may apply Theorem 1 in [4], and one of the following holds:

  1. (1)

    (case (i) of Theorem 1 in [4]) \(d(x)=apx-pxa\) for any \(x\in R\). Hence \(d(xy)=d(x)y+xd(y)=(apx-pxa)y+x(apy-pya)\), on the other hand \(d(xy)=apxy-pxya\). Comparing the previous identities, we get

    $$\begin{aligned} -px[y,a]-xapy+xpya=0. \end{aligned}$$
    (3.11)

    In particular, for \(x=y=p^{-1}\), it follows \(ap^{-1}-p^{-1}a=0\) which implies \(ap-pa=0\). Therefore \(d(x)=p[a,x]\), moreover by (3.11) we get \(-px[y,a]-xpay+xpya=0\), that is \([x,p][y,a]=0\). As an application of [19, Theorem 3], it follows that either \(p\in C\) or \(a\in C\), in any case \(d=0\).

  2. (2)

    (cases (ii) and (iii) of [4, Theorem 1]) There exists \(q\in Q\) such that \(d(x)=p[q,x]\), for all \(x\in R\). Hence \(d(xy)=d(x)y+xd(y)=p[q,x]y+xp[q,y]\), on the other hand \(d(xy)=p[q,xy]=p[q,x]y+px[q,y]\). Comparing the previous identities, we get \([p,x][q,y]=0\), for any \(x,y\in R\), that is either \(p\in C\) or \(q\in C\). In any case \(d=0\).

  3. (3)

    (case (iv) of [4, Theorem 1 ]) There exists \(q\in Q\) such that \(d(x)=px+p[q,x]\) for all \(x\in R\). Hence \(d(xy)=d(x)y+xd(y)=(px+p[q,x])y+x(py+p[q,y])\), on the other hand \(d(xy)=pxy+p[q,xy]=pxy+p[q,x]y+px[q,y]\). Comparing the previous identities, we get \(xpy+[p,x][q,y]=0\) for any \(x,y\in R\). In particular, for \(x=p^{-1}\), it follows the contradiction \(y=0\), for all \(y\in R\).

Since in any case \(p\in C\), \(\beta \) is the identity map on R and F is an ordinary derivation of R such that \([a,F(x)]=0\) for any \(x\in R\). Hence, by the first Posner’s theorem in [18], it follows \(a\in Z(R)\) or \(F(R)\subseteq Z(R)\). In this last case, by second Posner’s theorem in [18], R is commutative. In any case we obtain that \(a\in Z(R)\). \(\square \)

Lemma 6

Let \(R=M_t(C)\) be the ring of \(t\times t\) matrices over C, with \(char(R)=l\ne 0, 2\), \(m\ge 1\) be an odd integer, \(p,q,u \in R\) with p invertible such that \(F(x)=pxp^{-1}q-qx\), \(G(x)=pxp^{-1}u-ux\) and

$$\begin{aligned} \left[ pxp^{-1}q-qx,pyp^{-1}u-uy\right] _m-[x,y] \end{aligned}$$
(3.12)

for any \(x,y \in R\). Then t=1.

Proof

In (3.12) replace xy by \(p^{-1}x, p^{-1}y\), respectively, and denote \(a=p^{-1}q\), \(b=qp^{-1}\), \(c=p^{-1}u\), \(w=up^{-1}\). Therefore Q satisfies

$$\begin{aligned} \Phi (x,y)=\left[ xa-bx,yc-wy\right] _m-\left[ p^{-1}x,p^{-1}y\right] . \end{aligned}$$
(3.13)

In case \(a \in C\), then \(F=0\) and (3.12) implies \([x,y]=0\) for any \(x,y \in R\), that is R is commutative. On the other hand, if \(c \in C\), then \(G=0\) and again it follows that R is commutative. Therefore, here we may assume that \(t\ge 2\) and both \(a\notin C\) and \(c \notin C\). We prove that a number of contradiction follows.

Firstly, we notice that, for any inner automorphism \(\varphi \) of \(M_t(C)\), we have that

$$\begin{aligned} \left[ x\varphi (a)-\varphi (b)x,y\varphi (c)-\varphi (w)y\right] _m- \left[ \varphi (p^{-1})x,\varphi (p^{-1})y\right] \end{aligned}$$
(3.14)

is a generalized identity for R. We will make a frequent use of this fact.

As above, we denote by \(e_{ij}\) the usual matrix unit, with 1 in the (ij) entry and zero elsewhere, and say \(a=\sum _{kl} a_{kl}e_{kl}\), \(b=\sum _{kl}b_{kl}e_{kl}\), \(c=\sum _{kl}c_{kl}e_{kl}\) and \(w=\sum _{kl}w_{kl}e_{kl}\), for \(a_{kl}, b_{kl}, c_{kl}, w_{kl}\in C\).

Suppose \(t\ge 3\). In (3.13), we make the following choices: \(x=e_{ji}\), \(y=e_{kk}\), for any ijk different indices; moreover, we right multiply (3.13) by \(e_{jj}\) and left multiply by \(e_{kk}\). As a consequence we get \(a_{kj}c_{ij}^m=0\), that is \(a_{kj}c_{ij}=0\). Now, let \(\varphi (x)=(1+e_{ki})x(1-e_{ki})\) and denote \((a_{ij}')_{t\times t}\) the entries of the matrix \(\varphi (a)\), and \((c_{ij}')_{t\times t}\) the entries of the matrix \(\varphi (c)\). By the above computations, we get \(a_{kj}'c_{ij}'=0\), that is \((a_{kj}+a_{ij})c_{ij}=0\), which means \(a_{ij}c_{ij}=0\). Thus, by Proposition 1 in [10], it follows that either \(a \in C\) or \(c\in C\), in any case a contradiction.

Therefore, we finally consider the case \(t=2\) and \(R=M_2(C)\). For \(x=e_{ii}\) and \(y=e_{ji}\) in (3.13), with \(i\ne j\), right multiply (3.13) by \(e_{jj}\) and left multiply by \(e_{ii}\), it follows \(a_{ij}(c_{ij}+w_{ij})^m=0\), that is

$$\begin{aligned} a_{ij}(c_{ij}+w_{ij}). \end{aligned}$$
(3.15)

Let now \(x=y=e_{ij}\) in (3.13), with \(i\ne j\), and right multiply (3.13) by \(e_{ii}\) one has \((a_{ji}+c_{ji})^m=0\), that is

$$\begin{aligned} a_{ji}+c_{ji}~\mathrm{for \, all}~ i\ne j \end{aligned}$$
(3.16)

which means that \(a+c\) is a diagonal matrix. In this case, a standard argument shows that \(a+c\) is a central matrix, say \(a=\lambda -c\), for \(\lambda \in Z(R)\). Analogously, for \(x=y=e_{ij}\) in (3.13), and left multiply (3.13) by \(e_{jj}\) one has \((w_{ji}+b_{ji})^m=0\), that is

$$\begin{aligned} w_{ji}+b_{ji}~\mathrm{for \, all}~ i\ne j \end{aligned}$$

which means that \(w+b\) is a diagonal matrix and, as above, \(w+b\) is a central matrix, say \(b=\mu -w\), for \(\mu \in Z(R)\).

In other words, \(p^{-1}q=\lambda -p^{-1}u\) and \(qp^{-1}=\mu -up^{-1}\). Therefore, if either \(\lambda =0\) or \(\mu =0\) then \(q=-u\), \(\lambda =\mu =0\), and \(F=-G\). On the other hand, if both \(\lambda \ne 0\) and \(\mu \ne 0\), it follows both \(q=p\lambda -u\) and \(q=p\mu -u\), that is \(\lambda =\mu \) and easy computations show that \(F=-G\) in any case.

Now, we write \(v=c+w\) and let \(v=\sum _{kl}v_{kl}e_{kl}\), for \(v_{kl}\in C\). The next step is to prove that either v is diagonal or both a and c are diagonal matrices of R. To do this, we assume by contradiction that v is not diagonal, for example, let \(v_{12}\ne 0\), and prove that a contradiction follows. In this case, by (3.15) and (3.16) we get \(a_{12}=0\) and \(c_{12}=0\). Moreover, if \(v_{21}\ne 0\), then a and c are diagonal matrices and we are done. Thus we assume that \(v_{21}=0\).

Let \(\varphi (x)=(1+e_{12})x(1-e_{12})\) and \(\chi (x)=(1-e_{12})x(1+e_{12})\) and denote \(\varphi (a)=\sum _{kl} a_{kl}'e_{kl}\), \(\varphi (c)=\sum _{kl}c_{kl}'e_{kl}\), \(\varphi (v)=\sum _{kl}v'_{kl}e_{kl}\), \(\chi (a)=\sum _{kl} a_{kl}''e_{kl}\), \(\chi (c)=\sum _{kl}c_{kl}''e_{kl}\), \(\chi (v)=\sum _{kl}v''_{kl}e_{kl}\).

We notice that, if both \(v_{12}'\ne 0\) and \(v_{12}''\ne 0\), then \(a_{12}'=0\), \(a_{12}''=0\) and also \(c_{12}'=0\), \(c_{12}''=0\) that is \(a_{22}-a_{11}-a_{21}=0\), \(-a_{22}+a_{11}-a_{21}=0\), \(c_{22}-c_{11}-c_{21}=0\), \(-c_{22}+c_{11}-c_{21}=0\), implying \(a_{21}=0\) and \(c_{21}=0\), so that a and c are diagonal and we are done.

Thus we may assume (without loss of generality) \(v_{12}''=0\). We have proved that if \(v_{12}\ne 0\) then \(v_{12}=v_{22}-v_{11}\). Let \(\theta (x)=(1+e_{21})x(1-e_{21})\) and \(\theta (v)=\sum _{kl} v_{kl}'''e_{kl}\). Since \(v_{12}'''=v_{12}\ne 0\), then \(v_{12}'''=v_{22}'''-v_{11}'''\), that is \(v_{12}=-v_{22}+v_{11}\), implying again the contradiction \(v_{12}=0\).

The previous argument says that either v is diagonal or both a and c are diagonal matrices, and as above, we may conclude that either v is central or both a and c are central matrices of R.

Since if \(a\in Z(R)\) and \(c\in Z(R)\), then \(F=G=0\), which is a contradiction, then we assume in what follows that \(c+w=v=\nu \in Z(R)\), that is \(p^{-1}q+qp^{-1}=\nu \). For \(y=p\) in (3.12) it follows that

$$\begin{aligned} \biggl [pxp^{-1}q-qy,[p,u]\biggr ]_m=[x,p]. \end{aligned}$$
(3.17)

Assume that [pu] is an invertible matrix in \(M_2(C)\), thus \(0\ne [p,u]^2\in Z(R)\) and by computations one has

$$\begin{aligned} 2^{m-1}\biggl (\left( pxp^{-1}q-qx\right) [p,u]^m-[p,u] \left( pxp^{-1}q-qx\right) [p,u]^{m-1}\biggr )=[x,p]. \end{aligned}$$

Since [pu] is an invertible matrix and \(m-1\) is even, then \([p,u]^{m-1}\in Z(R)\) and

$$\begin{aligned} 2^{m-1}\biggl (\left( pxp^{-1}q-qx\right) [p,u]^m-[p,u]^m \left( pxp^{-1}q-qx\right) \biggr )=[x,p]. \end{aligned}$$

Since \(F\ne 0\) and \(2^{m-1}[p,u]^m\ne 0\), we may apply Lemma 5 and obtain the contradiction \(p\in Z(R)\).

Therefore [pu] is not an invertible matrix in \(M_2(C)\), i.e., \([p,u]^2=0\). Once again, for \(y=p\) in (3.12) and since \(m\ge 2\), we get

$$\begin{aligned} 0=\biggl [pxp^{-1}q-qx,[p,u]\biggr ]_m=[x,p] \end{aligned}$$
(3.18)

and as above we conclude with the contradiction \(p\in Z(R)\). \(\square \)

4 The proof of Theorem 1

In light of Lemma 1, we may assume \(n=1\). Since I, R, and Q satisfy the same generalized identities with skew derivations and automorphisms, we may assume

$$\begin{aligned}{}[F(x),G(y)]_m-[x,y]=0~\mathrm{for\,\, all}~ x,y \in Q. \end{aligned}$$
(4.1)

Moreover, by Lemma 4, there exists \(p,q,u \in Q\) with p invertible such that \(F(x)=pxp^{-1}q-qx\) and \(G(x)=pxp^{-1}u-ux\) for all \(x\in R\). Notice that, if either \(F=0\) or \(G=0\), then \([x,y]=0\) for all \(x,y \in Q\), which means that Q is commutative, as well as R. Thus we also assume both \(F\ne 0\) and \(G\ne 0\). Hence Q satisfies the generalized polynomial identity

$$\begin{aligned} \left[ pxp^{-1}q-qx,pyp^{-1}u-uy\right] _m-[x,y]. \end{aligned}$$
(4.2)

In (4.2) replace xy by \(p^{-1}x, p^{-1}y\), respectively, and denote \(a=p^{-1}q\), \(b=qp^{-1}\), \(c=p^{-1}u\), \(w=up^{-1}\). Therefore Q satisfies

$$\begin{aligned} \Phi (x,y)=[xa-bx,yc-wy]_m-\left[ p^{-1}x,p^{-1}y\right] . \end{aligned}$$
(4.3)

In case \(a \in C\), then \(F=0\). On the other hand, if \(c \in C\), then \(G=0\). Therefore, we may assume that both \(a\notin C\) and \(c \notin C\), in particular Q is not commutative. We will prove that a number of contradiction follows. Since both a and c are not central elements, then (4.3) is a non-trivial generalized identity for R as well as for Q. Hence, Q is a primitive ring dense of linear transformations over a vector space V over C.

Assume first that \(dim_CV=t\) is a finite integer. Thus, \(Q\cong M_k(C)\) and by Lemma 6 it follows the contradiction that Q is commutative.

Let now \(dim_CV=\infty \). Let \(y_1,y_2\in Q\). By Litoff’s theorem (see Theorem 4.3.11 in [3]) there exists an idempotent element \(e\in Q\) such that

$$\begin{aligned} y_1,y_2, a, b, c, w, p \in eQe \cong M_t(C) \end{aligned}$$

for some integer t. Of course \(\Phi (x,y)=0\) for all \(x,y \in eQe\). Thus by Lemma 6, either \(a \in Ce\), or \(c\in Ce\) or \([y_1,y_2]=0\). This means that either \(F(eQe)=0\) or \(G(eQe)=0\) or \([y_1,y_2]=0\). As above, if \(F(eQe)=0\) or \(G(eQe)=0\) then eQe is commutative. Therefore, in any case we get \([y_1,y_2]=0\). By the arbitrariness of \(y_1,y_2 \in Q\), it follows that \([y_1,y_2]=0\) for any \(y_1,y_2 \in Q\), that is Q is commutative, which is a contradiction.

5 The proof of Theorem 2

Also in this case, in light of Lemma 1, we may assume \(n=1\). Let P be a prime ideal of R. Set \(\overline{R}=R/P\) and write \(\overline{x}=x+P \in \overline{R}\), for all \(x\in R\). We start from

$$\begin{aligned} \biggl [\overline{F(x)},\overline{G(y)}\biggr ]_m=\overline{[x,y]}~\mathrm{for~ all}~ \overline{x}, \overline{y}\in \overline{R}. \end{aligned}$$
(5.1)

Case 1 \(F(P)\subseteq P\), \(\alpha (P)\not \subseteq P\).

In this case \(\overline{\alpha (P)}\) is an ideal of \(\overline{R}\). Moreover, for any \(x\in R\), \(p\in P\), \(F(px)=F(p)x+\alpha (p)F(x)\in P\), so that \(\alpha (p)F(x)\in P\) and \(\overline{\alpha (p)F(x)}=\overline{0}\). Since \(\overline{R}\) is prime and \(\overline{F(x)}\) annihilates a non-zero ideal of \(\overline{R}\), \(\overline{F(x)}=\overline{0}\), that is \(F(R)\subseteq P\). Thus \([F(x),G(y)]_m\in P\) for any \(x,y \in R\) and, by (5.1), it follows \([R,R]\subseteq P\).

Case 2 \(G(P)\subseteq P\), \(\alpha (P)\not \subseteq P\).

Identical computations as in CASE 1, imply \([R,R]\subseteq P\).

Case 3 \(F(P)\subseteq P\), \(\alpha (P)\subseteq P\).

In this case \(\overline{F}\) is a skew derivation of \(\overline{R}\). If \(G(P)\subseteq P\), then also \(\overline{G}\) is a skew derivation of \(\overline{R}\), and by the primeness of \(\overline{R}\) and Theorem 1, we have that \(\overline{R}\) is commutative, that is \([R,R]\subseteq P\).

Let now \(G(P)\not \subseteq P\). Then \(\overline{G(P)}\) is a non-zero ideal of \(\overline{R}\). For any \(x,y \in R\) and \(q\in P\), we get

$$\begin{aligned} \biggl [F(x),G(yq)\biggr ]_m=[x,yq]~\mathrm{for\,\, all} ~x,y\in R \end{aligned}$$

that is

$$\begin{aligned} \biggl [F(x),G(y)q+\alpha (y)G(q)\biggr ]_m=[x,yq]~\mathrm{for~all}~x,y\in R \end{aligned}$$

implying that

$$\begin{aligned} \biggl [F(x),\alpha (y)G(q)\biggr ]_m\in P~\mathrm{for~all}~x,y\in R \end{aligned}$$

and

$$\begin{aligned} \biggl [\overline{F(x)},\overline{\alpha (y)G(q)}\biggr ]_m=\overline{0}~\mathrm{for\,\,all}~ \overline{x}, \overline{y}\in \overline{R}. \end{aligned}$$
(5.2)

In particular, let \(h\ge 1\) be such that \(p^h\ge m\), then (5.2) is

$$\begin{aligned} \biggl [\overline{F(x)},\overline{\alpha (y)G(q)}^{p^h}\biggr ]=\overline{0}~\mathrm{for\,\,all}~ \overline{x}, \overline{y}\in \overline{R}. \end{aligned}$$
(5.3)

By the primeness of \(\overline{R}\) and since \(\overline{\alpha (R)G(P)}\) is a non-zero ideal of \(\overline{R}\), by applying the result in [14], it follows that either \(\overline{F(R)}\subseteq Z(\overline{R})\) or \(\overline{\alpha (R)G(P)}\subseteq Z(\overline{R})\). In the first case, by (5.1) we get \(\overline{[x,y]}=\overline{0}\). In the latter case it follows that \(\overline{R}\) is commutative. In any case \([R,R]\subseteq P\).

Case 4 \(G(P)\subseteq P\), \(\alpha (P)\subseteq P\).

In light of previous cases, we may also assume that \(F(P)\not \subseteq P\). In this case, \(\overline{G}\) is a skew derivation of \(\overline{R}\), moreover \(\overline{F(P)}\) is a non-zero ideal of \(\overline{R}\). For any \(x,y \in R\) and \(q\in P\), we get

$$\begin{aligned} \biggl [F(qx),G(y)\biggr ]_m=[qx,y],\quad \forall ~x,y\in R \end{aligned}$$

that is

$$\begin{aligned} \biggl [F(q)x+\alpha (q)F(x),G(y)\biggr ]_m\in P,\quad \forall ~x,y\in R \end{aligned}$$

implying that

$$\begin{aligned} \biggl [F(q)x,G(y)\biggr ]_m\in P,\quad \forall ~x,y\in R \end{aligned}$$

and

$$\begin{aligned} \biggl [\overline{F(P)R},\overline{G(R)}\biggr ]_m=\overline{0}. \end{aligned}$$
(5.4)

Since \(\overline{R}\) is prime and \(\overline{F(P)R}\) is a non-zero ideal of \(\overline{R}\), \(\overline{F(P)R}\), and \(\overline{R}\) satisfy the same generalized identities with skew derivation \(\overline{G}\) and automorphism \(\overline{\alpha }\). Therefore, by (5.4) we get

$$\begin{aligned} \biggl [\overline{R},\overline{G(R)}\biggr ]_m=\overline{0}. \end{aligned}$$
(5.5)

From (5.5) and (5.1) it follows again \(\overline{[x,y]}=\overline{0}\) for any \(x,y \in R\), that is \([R,R]\subseteq P\).

Case 5 \(F(P)\not \subseteq P\), \(G(P)\not \subseteq P\).

For any \(x,y \in R\) and \(q,u\in P\), by (5.1) we get

$$\begin{aligned} \biggl [F(xq),G(yu)\biggr ]_m=[xq,yu]\in P~\mathrm{for\, all}~x,y\in R \end{aligned}$$

that is

$$\begin{aligned} \biggl [F(x)q+\alpha (x)F(q),G(y)u+\alpha (y)G(u)\biggr ]_m\in P~\mathrm{for\, all}~x,y\in R \end{aligned}$$

implying that

$$\begin{aligned} \biggl [\alpha (x)F(q),\alpha (y)G(u)\biggr ]_m\in P~\text {for all}~x,y\in R \end{aligned}$$

and

$$\begin{aligned} \biggl [\overline{\alpha (R)F(P)},\overline{\alpha (R)G(P)}\biggr ]_m=\overline{0}. \end{aligned}$$

Assume first that \(\alpha (P)\subseteq P\), so that both \(\overline{F(P)}\) and \(\overline{G(P)}\) are non-zero ideals of \(\overline{R}\). On the other hand, if \(\alpha (P)\not \subseteq P\) then \(\overline{\alpha (R)F(P)}\) and \(\overline{\alpha (R)G(P)}\) are left ideals of \(\overline{R}\). In any case, for any \(\overline{x}\in \overline{\alpha (R)F(P)}\) and \(\overline{y}\in \overline{\alpha (R)G(P)}\), it follows \([\overline{x},\overline{y}]=\overline{0}\). Since \(\overline{R}\) is prime, we apply again a reduced version of main result in [14], and conclude that either \(\overline{\alpha (R)F(P)}\subseteq Z(\overline{R})\) or \(\overline{\alpha (R)G(P)}\subseteq Z(\overline{R})\). In any case \(\overline{R}\) contains a non-zero central ideal (either left or two-sided), so that \(\overline{R}\) is commutative, i.e., \(\overline{[x,y]}=\overline{0}\), for any \(x,y \in R\) and \([R,R]\subseteq P\).

Therefore in any case \([R,R]\subseteq P\), for any prime ideal P of R. Then \([R,R]\subseteq \bigcap _{i} P_{i}=(0)\) (where \(P_{i}\) are all prime ideals of R), that is R is commutative.