1 Introduction

The main purpose of the present paper is to discuss the problem of characterizing the strong skew commutativity preserving maps and the strong 2-skew commutativity preserving maps on operator algebras. However, we discuss first in a general pure algebraic frame and then apply to operator algebras.

Let \({\mathcal {R}}\) be a ring. For any elements, \(A,B\in {\mathcal {R}}\), define \([A,B]_0=A, [A,B]_1=AB-BA\), and inductively \([A,B]_k=[[A,B]_{k-1},B]\), where \(k\ge 1\) is a positive integer. A map \(\Phi :{\mathcal {R}} \rightarrow {\mathcal {R}}\) is said to be strong k-commutativity preserving if \([\Phi (A),\Phi (B)]_k = [A,B]_k\) for all \(A,B\in {\mathcal {R}}\). The problem of characterizing strong k-commutativity preserving maps has been studied intensively on various rings and operator algebras (for example, ref. [1,2,3,4,5,6,7,8]). It was shown in [5] (resp. [6]) that a surjective map on a unital prime ring \({\mathcal {R}}\) containing a nontrivial idempotent is a strong commutativity preserving map (resp. a strong 2-commutativity preserving map) if and only if it has the form: \(A\mapsto \lambda A+f(A)\), where \(\lambda \in {\mathcal {C}}\) with \(\lambda ^2=1\) (resp. \(\lambda ^3=1\)), \(f:{\mathcal {R}}\rightarrow {{\mathcal {Z}}}({\mathcal {R}})\) is a map, \({\mathcal {C}}\) is the extended centroid of \({\mathcal {R}}\), and \({{\mathcal {Z}}}({\mathcal {R}})\) is the center of \({\mathcal {R}}\).

Recall that a ring \({\mathcal {R}}\) is called a \(*\)-ring or a ring with involution if there is an additive map \(*: \mathcal {R}\rightarrow {\mathcal {R}}\) satisfying \((AB)^{*}=B^{*}A^{*}\) and \((A^{*})^{*}=A\) for all \(A, B\in {\mathcal {R}}\). For \(A,B\in {\mathcal {R}}\), the skew Lie product or the skew commutator of A and B is denoted by \({}_*[A,B]=AB-BA^{*}\) (Ref. [9], where \({}_*[A,B]\) is denoted by \([A,B]_*\)). The concept of skew Lie product is found playing an important role in some research topics, and their study has attracted many authors’ attention (Ref. [10,11,12,13,14,15,16,17,18]). Similar to the commutator case stated in previous paragraph, a map \(\Phi :{{\mathcal {R}}} \rightarrow {{\mathcal {R}}}\) is said to be strong skew commutativity preserving if \({}_*[\Phi (A), \Phi (B)]\) = \({}_*[A, B]\) for any \(A, B\in {{\mathcal {R}}}\). For the problem of how to characterize the strong skew commutativity preserving maps, it was shown in [12] that a nonlinear surjective map \(\Phi \) on a factor von Neumann algebra \({\mathcal {A}}\) is strong skew commutativity preserving if and only if it has the form \(A\mapsto \Psi (A)+h(A)I\), where \(\Psi \) is a strong skew commutativity preserving bijective linear map and \(h:\mathcal A\rightarrow \mathbb {R}\) is a functional with \(h(0) = 0\). As usual, \(\mathbb R\) and \({\mathbb {C}}\) denote the real number field and the complex number field, respectively. In [16], a more precise form of the above surjective map was achieved: \(A\mapsto ZA\), where \(Z\in {\mathcal {Z}}_S\), the set of all self-adjoint elements in the center of \({\mathcal {A}}\). In a pure algebraic frame, it was also shown in [16] that if \({\mathcal {R}}\) is a prime \(*\)-ring which contains a nontrivial symmetric idempotent, then the strong skew commutativity preserving surjective maps on \({\mathcal {R}}\) have the form \(\Phi (A)=ZA+f(A)\) for all \(A\in {{\mathcal {R}}}\), where \(Z\in {\mathcal {Z}}_S({\mathcal {R}})\) and \(f:{\mathcal {R}}\rightarrow \mathcal Z_S({\mathcal {R}})\). Recently, [17] and [18] generalized this result to unital \(*\)-rings \({\mathcal {R}}\) which contain a nontrivial symmetric idempotent P satisfying \( A\mathcal {R}P=\{0\}\Rightarrow A=0\) and \( A{\mathcal {R}}(I-P ) =\{ 0\}\Rightarrow A = 0\).

Let \({\mathcal {R}}\) be a \(*\)-ring. For any positive integer \(k\ge 1\), like the concept of k-commutator, the k-skew commutator of a pair AB in \({\mathcal {R}}\) is naturally defined inductively by \({}_*[A,B]_{k}={{}_*[A,{{}_*[A,B]_{k-1}}]}\) with \({}_*[A,B]_{0}=B,\,{{}_*[A,B]_{1}}={{}_*[A,B]}=AB-BA^*\). The k-skew commutator has some interesting properties. For instance, if \({\mathcal {R}}\) is unital, prime and of characteristic not 2, then \({}_*[A,B]_2=0\) holds for all B if and only if \(A\in {\mathcal Z}={{\mathcal {Z}}}({\mathcal {R}})\) and \(A=A^*\), that is, \(A\in {\mathcal Z_S}\) (Lemma 4.3); if, in addition, \({\mathcal {R}}\) contains a nontrivial symmetric idempotent, then \({}_*[B,A]_2=0\) holds for all B if and only if \(A=0\) (Lemma 4.2).

Let \({\mathcal {R}}\) be a \(*\)-ring. Recall that a map \(\Phi :{{\mathcal {R}}} \rightarrow {{\mathcal {R}}}\) is said to be strong k-skew commutativity preserving if \({}_*[\Phi (A), \Phi (B)]_k={{}_*[A, B]_k}\) holds for all \(A, B\in {{\mathcal {R}}}\). That is, the strong k-skew commutativity preserving maps are those maps that preserve the k-skew commutators. It is interesting to ask how to characterize the strong k-skew commutativity preserving maps. In the present paper, we focus our attention on the case \(1\le k\le 2\).

Our aim is twofold. The first one is to sharp the results in [16,17,18] by proving the symmetric center valued map f is 0, and then get a characterization of the strong skew commutativity preserving maps. Let \({\mathcal {R}}\) be a unital \(*\)-ring which contains a nontrivial symmetric idempotent P satisfying \( A{\mathcal {R}}P=0\Rightarrow A=0\) and \( A{\mathcal {R}}(I-P ) = 0\Rightarrow A = 0\), where I denotes the unit. We show that a surjective map \(\Phi :{{\mathcal {R}}}\rightarrow {{\mathcal {R}}}\) is strong skew commutativity preserving if and only if there exists \(Z \in {{\mathcal {Z}}}_S\) with \(Z^2=I\) such that \(\Phi (A)=ZA\) for all \(A\in {\mathcal {R}}\) (Theorem 2.1). Particularly, if \({\mathcal {R}}\) is prime, we have further that \(Z\in \{-I,I\}\). The second one, as the main part of the paper, is to characterize the strong 2-skew commutativity preserving maps. Let \({\mathcal {R}}\) be a unital prime \(*\)-ring with the unit I and containing a nontrivial symmetric idempotent. We show that a surjective map \(\Phi :\mathcal {R}\rightarrow {\mathcal {R}}\) is strong 2-skew commutativity preserving if and only if there exists an element \(\lambda \in {\mathcal {C}}_{S}\) with \(\lambda ^{3}=I\) such that \(\Phi (A)=\lambda A\) for all \(A\in {{\mathcal {R}}}\). Applying to some operator algebras such as prime \(\hbox {C}^*\)-algebras, factor von Neumann algebras and indefinite self-adjoint standard operator algebras, we get a characterization of the identity map. It is somewhat surprising by our results that the 2-skew commutator is a complete invariant for the identity map for the operator algebras mentioned above. A characterization of strong 2-skew commutativity preserving maps on symmetric standard operator algebras is also established.

The paper is organized as follows. Section 2 is devoted to sharping the results in [16,17,18] and characterizing the strong skew commutativity preserving maps. In Sect. 3, we give a characterization of strong 2-skew commutativity preserving maps on unital prime \(*\)-rings containing a nontrivial symmetric idempotent (Theorem 3.1) and its corollaries applying to some operator algebras. Section 4 is devoted to proving Theorem 3.1 for the case when the ring is 2-torsion-free; while Sect. 5 is devoted to proving Theorem 3.1 for the case when the characteristic of the ring is 2.

2 Strong Skew Commutativity Preserving Maps

In this section, we sharp the main results in [16,17,18] by proving the symmetric center valued map f vanishes and consequently get a characterization of strong skew commutativity preserving maps.

Recall that a nontrivial symmetric idempotent P is an element so that \(P^2=P=P^*\) and \(P\not \in \{0,I\}\), where I is the unit.

Theorem 2.1

Let \({\mathcal {R}}\) be a unital ring with involution \(*\). Assume that \({\mathcal {R}}\) contains a nontrivial symmetric idempotent P which satisfies:

  • \((\mathbb {Q}_1)\ A{\mathcal {R}}P = \{0\}\) implies \(A = 0, (\mathbb {Q}_2)\ A{\mathcal {R}}(I-P) =\{ 0\}\) implies \(A = 0\).

Let \(\Phi :{\mathcal {R}}\rightarrow {\mathcal {R}}\) be a surjective map. Then \(\Phi \) is strong skew commutativity preserving if and only if there exists an element \(Z \in {\mathcal {Z}}_{S }(\mathcal {R})\) with \(Z^2 = I\) such that \(\Phi (A)=ZA\) for all \(A\in {\mathcal R}\).

Proof

The “if” part is obvious. To check the “only if” part, assume that \(\Phi \) is strong skew commutativity preserving. By [17, 18], there exists a map \(f:{\mathcal {R}}\rightarrow \mathcal {Z}_S({\mathcal {R}})\) and an element \(Z\in {\mathcal {Z}}_S(\mathcal {R})\) with \(Z^2=I\) such that \(\Phi (A)=ZA+f(A)\) for all \(A\in \mathcal {R}\). Now we need to show \(f(A)=0\) for all \(A\in {\mathcal {R}}\). Assume, on the contrary, there exists \(B\in {\mathcal {R}}\) such that \(f(B)\not =0\). For any \(A\in {\mathcal {R}}\), as \(\Phi \) is strong skew commutativity preserving, we have

$$\begin{aligned} {}_*[A,B]= & {} {{}_*[\Phi (A),\Phi (B)]}={{}_*[ZA+f(A),ZB+f(B)]}\\= & {} {{}_*[ZA,ZB]}+{{}_*[ZA,f(B)]}, \end{aligned}$$

which implies

$$\begin{aligned} Zf(B)(A-A^*)=(A-A^*)Zf(B)=0 \end{aligned}$$

for all \(A\in {\mathcal {R}}\). Since \(Z^2=I\),

$$\begin{aligned} f(B)(A-A^*)=(A-A^*)f(B)=0 \end{aligned}$$
(2.1)

for all \(A\in {\mathcal {R}}\).

Obviously, Eq. (2.1) implies that

$$\begin{aligned} f(B)PA(I-P)+f(B)(I-P)A^*P=0 \end{aligned}$$

holds for every \(A\in {{\mathcal {R}}}\). By the Peirce decomposition relative to P, this implies \(f(B)PA(I-P)=0\) for all \(A\in {\mathcal R}\). Hence, by (\({{\mathbb {Q}}}_2\)), we have \(f(B)P=0\). Since \(f(B)\in {{\mathcal {Z}}}({{\mathcal {R}}})\), this entails \(f(B)AP=Af(B)P=0\) for all \(A\in {{\mathcal {R}}}\), so that, by (\({{\mathbb {Q}}}_1\)), \(f(B)=0\), a contradiction. Hence \(f=0\) and \(\Phi (A)=ZA\) for all A. \(\square \)

In the case when the ring is prime, a more concrete form is achieved.

Corollary 2.2

Let \({\mathcal {R}}\) be a unital prime ring with involution \(*\) and containing a nontrivial symmetric idempotent. Let \(\Phi :{\mathcal {R}}\rightarrow {\mathcal {R}}\) be a surjective map. Then \(\Phi \) is strong skew commutativity preserving if and only if either \(\Phi (A)= A\) for all \(A\in {{\mathcal {R}}}\), or \(\Phi (A)= - A\) for all \(A\in {{\mathcal {R}}}\).

Proof

By Theorem 2.1, there exists a central symmetric element Z with \(Z^2=I\) such that \(\Phi (A)=ZA\) for each \(A\in {\mathcal {R}}\). If \(Z\not \in \{-I,I\}\), then \(I+Z\not =0, I-Z\not =0\) and \((I+Z)(I-Z)=0\). It follows that \((I+Z)A(I-Z)=(I+Z)(I-Z)A=0\) holds for all \(A\in {\mathcal {R}}\), which contradicts the primeness of \({\mathcal {R}}\). \(\square \)

Corollary 2.2 can be applied to prime \(\hbox {C}^*\)-algebras, factor von Neumann algebras as well as indefinite self-adjoint standard operator algebras like what we have done in the next section for characterizing strong 2-skew commutativity preserving maps (see Corollaries 3.23.7). We leave these to the readers.

3 Strong 2-skew Commutativity Preserving Maps

Begining from this section, we discuss the problem of how to characterize strong 2-skew commutativity preserving maps on prime \(*\)-rings and operator algebras.

Let \({\mathcal {R}}\) be a ring. Denote by \({\mathcal {Q}}=\mathcal {Q}_\mathrm{ml}({\mathcal {R}})\) the maximal left ring of quotients. The center \({\mathcal {C}}={\mathcal {C}}({\mathcal {R}})={\mathcal Z}({\mathcal {Q}})\) of \({\mathcal {Q}}\) is called the extended centroid of \({\mathcal {R}}\). Note that if \({\mathcal {R}}\) is prime, then \({\mathcal {Q}}\) is also prime and \({\mathcal {C}}\) is a field. If \({\mathcal {R}}\) is a prime \(*\)-ring, the symmetric extended centroid \({\mathcal {C}}_S=\{\lambda \in {\mathcal {C}}:\lambda =\lambda ^*\}\) is also a field. Moreover, \({\mathcal {Z}}({\mathcal {R}})\subseteq \mathcal {C}\) (Ref. [22] for details). Bear in mind that

$$\begin{aligned} {}_*[A,B]_2=A^2B-2ABA^*+B(A^*)^2. \end{aligned}$$

The following is our main result.

Theorem 3.1

Let \({\mathcal {R}}\) be a unital prime ring with involution \(*\) and containing a nontrivial symmetric idempotent. Let \(\Phi :{\mathcal {R}}\rightarrow {\mathcal {R}}\) be a surjective map. Then \(\Phi \) is strong 2-skew commutativity preserving if and only if there exists \(\lambda \in {{\mathcal {C}}}_S\) with \(\lambda ^3=I\) such that \(\Phi (A)=\lambda A\) for all \(A\in {{\mathcal {R}}}\).

Denote by char\(({\mathcal {R}})\) the characteristic of the ring \({\mathcal {R}}\). The proof of Theorem 3.1 will be given in Sect. 4 for the case char\(({\mathcal {R}})\not =2\) and in Sect. 5 for the case char\(({{\mathcal {R}}})=2\).

Now we give some applications of Theorem 3.1 to operator algebras.

By [19, Corollary 2.4], if \({\mathcal {A}}\) is a prime \(\hbox {C}^*\)-algebra, then its extended centroid \({{\mathcal {C}}}({\mathcal A})={\mathbb {C}}\). Thus, \({{\mathcal {C}}}_S({\mathcal {A}})={{\mathbb {R}}}\) and the following corollary is immediate, which also gives a characterization of the identity map on such prime \(\hbox {C}^*\)-algebras.

Corollary 3.2

Let \({\mathcal {A}}\) be a unital prime \(\hbox {C}^*\)-algebra containing a nontrivial self-adjoint idempotent and \(\Phi :{{\mathcal {A}}}\rightarrow {\mathcal {A}}\) a surjective map. Then \(\Phi \) is strong 2-skew commutativity preserving if and only if \(\Phi \) is the identity map.

Recall that a von Neumann algebra \({{\mathcal {M}}}\) is a \(\hbox {C}^*\)-subalgebra of some \({{\mathcal {B}}}(H)\), the algebra of all bounded linear operators acting on a complex Hilbert space H, which satisfies the double commutant property: \({\mathcal M}^{\prime \prime }={\mathcal {M}}\), where \({{\mathcal {M}}}^\prime =\{T: T\in {{\mathcal {B}}}(H)\ \text{ and }\ TA=AT\ \forall A\in {{\mathcal {M}}}\}\) and \({{\mathcal {M}}}^{\prime \prime }=\{{{\mathcal {M}}}^\prime \}^\prime \). \({{\mathcal {M}}}\) is called a factor if its center \({\mathcal Z}({{\mathcal {M}}})={{\mathcal {M}}}\cap {{\mathcal {M}}}^\prime ={{\mathbb {C}}}I\). It is well known that von Neumann algebras are unital \(\hbox {C}^*\)-algebras containing many nontrivial self-adjoint idempotents, and every factor von Neumann algebra is prime. So the following corollary is obvious from Corollary 3.2.

Corollary 3.3

Let \({{\mathcal {M}}}\) be a factor von Neumann algebra and \(\Phi : {{\mathcal {M}}}\rightarrow {{\mathcal {M}}}\) be a surjective map. Then \(\Phi \) is strong 2-skew commutativity preserving if and only if \(\Phi \) is the identity map.

Before the statement of our next corollary, let us recall some more notions. Denote by \({{\mathbb {F}}}\) the real field \({{\mathbb {R}}}\) or complex field \({{\mathbb {C}}}\). An indefinite inner product space means a linear space H over \({{\mathbb {F}}}\) equipped with a non-degenerate sesqui-linear Hermite functional \(\langle \cdot , \cdot \rangle \). Let \((H, \langle \cdot , \cdot \rangle )\) be an indefinite inner product space. If there exist a positive subspace \(H_+\) and a negative subspace \(H_-\) such that \( H = H_+ \oplus H_- \) and \((H_+, \langle \cdot , \cdot \rangle )\) is a Hilbert space when \(\langle \cdot , \cdot \rangle \) is restricted to \(H_+\), while \((H_-, -\langle \cdot , \cdot \rangle )\) is a Hilbert space when \(-\langle \cdot , \cdot \rangle \) is restricted to \(H_-\), we call H a complete indefinite inner product space or Krein space (Ref. [20]). (\(H, \langle \cdot ,\cdot \rangle \)) is a Krein space if and only if \((H, \langle \cdot ,\cdot \rangle _0)\) is a Hilbert space, where the inner product \(\langle \cdot ,\cdot \rangle _0\) is defined by

$$\begin{aligned} \langle x,y\rangle _0=\left\{ \begin{array}{ll} \langle x,y\rangle &{}\quad \mathrm{if}\ x,y\in H_+; \\ -\langle x,y\rangle &{}\quad \mathrm{if}\ x,y\in H_-; \\ 0 &{}\quad \mathrm{if}\ x\in H_{\pm }, y\in H_{\mp }. \end{array}\right. \end{aligned}$$

The set of all bounded linear operators on H with respect to the indefinite inner product \( \langle \cdot ,\cdot \rangle \) is the same as \({{\mathcal {B}}}(H, \langle \cdot ,\cdot \rangle _0)\), which is still denoted by \({{\mathcal {B}}}(H)\). For \(A\in {{\mathcal {B}}}(H)\), denote by \(A^\dagger \) the indefinite adjoint of A, which is determined by \(\langle x, A^\dagger y\rangle =\langle Ax,y\rangle \) for all \(x,y\in H\). Clearly, when \(\langle \cdot ,\cdot \rangle \) is an inner product itself, \(A^\dag =A^*\), the usual adjoint operator of A. Then \(A\mapsto A^\dag \) is clearly an involution on \({\mathcal B}(H)\). An indefinite self-adjoint standard operator algebra \({\mathcal {A}}\) on a real or complex Krein space H is a subalgebra of \({{\mathcal {B}}}(H)\) which contains the identity operator I and all finite rank operators, and satisfies \(A\in {{\mathcal {A}}}\Rightarrow A^\dag \in {{\mathcal {A}}}\). Obviously, every indefinite self-adjoint standard operator algebra \({\mathcal {A}}\) is a unital involution prime algebra with involution \(\dag \) and containing many nontrivial self-adjoint idempotents, of which \({{\mathcal {Z}}}_S({{\mathcal {A}}})={{\mathbb {R}}}I\). In addition, by [21], for standard operator algebras, Lemma 4.1 is valid if we replace “linearly independent over \({{\mathcal {C}}}\)” and “linear combination over \({\mathcal {C}}\)” simply by “linearly independent” and “linear combination.” Then, by a check the proof of Theorem 3.1 in Sect. 4, the following corollary is true.

Corollary 3.4

Let H be a real or complex Krein space with \(\dim H\ge 2\) and \({\mathcal A}\subseteq {{\mathcal {B}}}(H)\) be an indefinite self-adjoint standard operator algebra. Assume that \(\Phi :{{\mathcal {A}}} \rightarrow {{\mathcal {A}}}\) is a surjective map. Then \(\Phi \) is strong 2-skew commutativity preserving, that is, \(\Phi \) satisfies

$$\begin{aligned} \Phi (A)^2\Phi (B)-2\Phi (A)\Phi (B)\Phi (A)^\dag +\Phi (B)(\Phi (A)^\dag )^2=A^2B-2ABA^\dag +B(A^\dag )^2 \end{aligned}$$

for any \(A,B\in {\mathcal {A}}\), if and only if \(\Phi \) is the identity map.

Let H be a real or complex Hilbert space with inner product \(\langle \cdot ,\cdot \rangle \). For any invertible self-adjoint operator \(J\in {{\mathcal {B}}}(H), \langle \cdot ,\cdot \rangle _J\) defined by \(\langle x,y\rangle _J=\langle Jx,y\rangle \) is an indefinite inner product and \((H,\langle \cdot ,\cdot \rangle _J)\) is a Krein space. For \(A\in {{\mathcal {B}}}(H), A^\dag =J^{-1}A^*J\). If \(A^\dag =A\), we say that A is J-self-adjoint. Furthermore, a standard operator algebra \({{\mathcal {A}}}\subseteq {{\mathcal {B}}}(H)\) is called a J-self-adjoint standard operator algebra if \(A\in {{\mathcal {A}}} \) implies \(A^\dag =J^{-1}A^*J\in {{\mathcal {A}}}\).

Corollary 3.4′ 1

Let H be a Hilbert space and \(J\in {{\mathcal {B}}}(H) \) be an invertible self-adjoint operator. Let \({\mathcal {A}}\) be a J-self-adjoint standard operator algebra and \(\Phi :{{\mathcal {A}}}\rightarrow {{\mathcal {A}}}\) a surjective map. Then, \(\Phi \) satisfies

$$\begin{aligned}&\Phi (A)^2\Phi (B)-2\Phi (A)\Phi (B)J^{-1}\Phi (A)^*J+\Phi (B)J^{-1}(\Phi (A)^*)^2J\\&\quad =A^2B-2ABJ^{-1}A^*J+BJ^{-1}(A^*)^2J \end{aligned}$$

for any \(A,B\in {\mathcal {A}}\), if and only if \(\Phi \) is the identity map.

Particularly, we have

Corollary 3.5

Let \({{\mathcal {M}}}_n({\mathbb {F}})\) be the algebra of \(n\times n\) matrices over the real or complex field \({\mathbb {F}}\) and \(J\in {{\mathcal {M}}}_n({\mathbb {F}})\) an invertible self-adjoint matrix. Let \(\Phi : {{\mathcal {M}}}_n({\mathbb {F}})\rightarrow {\mathcal M}_n({\mathbb {F}})\) be a surjective map. Then \(\Phi \) satisfies

$$\begin{aligned}&\Phi (A)^2\Phi (B)-2\Phi (A)\Phi (B)J^{-1}\Phi (A)^*J+\Phi (B)J^{-1}(\Phi (A)^*)^2J\\&\quad =A^2B-2ABJ^{-1}A^*J+BJ^{-1}(A^*)^2J \end{aligned}$$

for any \(A,B\in {\mathcal {A}}\), if and only if \(\Phi \) is the identity map.

Given an orthonormal basis of a complex Hilbert space H, the transpose map \(A\mapsto A^t\) with respect to the basis is an involution operation on \({{\mathcal {B}}}(H)\). A standard operator algebra \({{\mathcal {A}}}\subseteq {{\mathcal {B}}}(H)\) is said to be symmetric (with respect to this basis) if \(A\in {\mathcal A}\Rightarrow A^t\in {{\mathcal {A}}}\). Thus, every symmetric standard operator algebra is a unital involution prime algebra with the involution being taken the transpose and, in this situation, \({{\mathcal {Z}}}_S({{\mathcal {A}}})={{\mathcal {Z}}}({\mathcal {A}})={\mathbb {C}}I\). By Theorem 3.1 and the reason mentioned before Corollary 3.4, we have

Corollary 3.6

Let H be a complex Hilbert space with \(\dim H\ge 2\) and \({\mathcal A}\subseteq {{\mathcal {B}}}(H)\) be a symmetric standard operator algebra with respect to some given orthonormal basis of H. Assume that \(\Phi :{{\mathcal {A}}} \rightarrow {{\mathcal {A}}}\) is a surjective map. Then \(\Phi \) satisfies

$$\begin{aligned} \Phi (A)^2\Phi (B)-2\Phi (A)\Phi (B)\Phi (A)^t+\Phi (B)(\Phi (A)^t)^2=A^2B-2ABA^t+B(A^t)^2 \end{aligned}$$

for any \(A,B\in {\mathcal {A}}\) if and only if there exists \(\lambda \in \{1, e^{i\frac{2\pi }{3}}, e^{i\frac{4\pi }{3}}\}\) such that \(\Phi (A) = \lambda A\) for all \(A \in {{\mathcal {A}}}\).

Corollary 3.7

Let \({{\mathcal {M}}}_n({\mathbb {C}})\) be the algebra of \(n\times n\) complex matrices, and let \(\Phi : {\mathcal M}_n({\mathbb {C}})\rightarrow {{\mathcal {M}}}_n({\mathbb {C}})\) be a surjective map. Then \(\Phi \) satisfies

$$\begin{aligned} \Phi (A)^2\Phi (B)-2\Phi (A)\Phi (B)\Phi (A)^t+\Phi (B)(\Phi (A)^t)^2=A^2B-2ABA^t+B(A^t)^2 \end{aligned}$$

for any \(A,B\in {{\mathcal {M}}}_n({\mathbb {C}})\), if and only if there exists \(\lambda \in \{1, e^{i\frac{2\pi }{3}}, e^{i\frac{4\pi }{3}}\}\) such that \(\Phi (A) = \lambda A\) for all \(A \in {{\mathcal {M}}}_n(\mathbb C)\).

Corollaries 3.6 and 3.7 also reveal that, in general, for a unital prime \(*\)-ring \({\mathcal {R}}, \lambda \in {{\mathcal {C}}}_S\) and \(\lambda ^3=1\) may not imply \(\lambda =1\). Hence the maps in Theorem 3.1 may not be the identity map.

4 Proof of Theorem 3.1: Char\(({{\mathcal {R}}})\not =2\)

In this section, we prove Theorem 3.1 for the case when the characteristic of \({\mathcal {R}}\) is not 2. To do this, we need three lemmas.

The first lemma comes from ( [22, Theorem A.7]).

Lemma 4.1

Let \({{\mathcal {R}}}\) be a prime ring, and let \(A_i, B_i, C_j, D_j\in {{\mathcal {R}}}, i=1,2,\ldots , n\) and \(j=1,2,\ldots , m\), such that \(\sum {{}_{i=1}^n} A_{i}T B_{i} = \sum {{}_{j=1}^m} C_{j}TD_{j}\) for all \( T\in {{\mathcal {R}}}\). If \(A_{1},\ldots ,A_{n}\) are linearly independent over \({\mathcal {C}}\) (the extended centroid of \({\mathcal {R}}\)), then each \(B_{i}\) is a linear combination of \(D_{1},\ldots ,D_{m}\) over \({\mathcal {C}}\). Similarly, if \(B_{1},\ldots ,B_{n}\) are linearly independent over \({\mathcal {C}}\), then each \(A_{i}\) is a linear combination of \(C_{1},\ldots ,C_{m}\) over \({\mathcal {C}}\). In particular, if \(ATB = BTA\) for all \(T\in {{\mathcal {A}}}\), then A and B are linearly dependent over \({\mathcal {C}}\).

The next two lemmas give some interesting properties possessed by 2-skew commutators.

Lemma 4.2

Let \({\mathcal {R}}\) be a unital prime \(*\)-ring with involution \(*\) and containing a nontrivial symmetric idempotent. Then, for \(S\in {{\mathcal {R}}}, {{}_*[X,S]_2}=0\) for all \(X\in {\mathcal {R}}\) if and only if \(S=0\).

Proof

The “if” part is obvious. Let us check the “only if ” part.

Let P be a nontrivial symmetric idempotent in \({\mathcal {R}}\). From \({}_*[PX,S]_2(I-P)=0\), we get \(PXPXS(I-P)=0\) holds for all \(X\in {{\mathcal {R}}}\). In particular, with \(X=P\), we have \(PS(I-P)=0\). Also, substituting \(X+P\) we find

$$\begin{aligned} PXPS(I-P)+PXS(I-P)=PX(P+I)S(I-P)=0, \end{aligned}$$

so that \((P+I)S(I-P)=0\) since \({\mathcal {R}}\) is prime and \(P\not =0\). Thus one gets \(S(I-P)=(P+I)S(I-P)-PS(I-P)=0\). By an analogous process, starting from \({}_*[(I-P)X,S]_2P=0\), one finds \(SP=0\). Hence \(S=SP+S(I-P)=0\), as desired. \(\square \)

Lemma 4.3

Let \({\mathcal {R}}\) be a prime \(*\)-ring of characteristic not 2 and \(S\in {{\mathcal {R}}}\). Then \({}_*[S,X]_2=0\) for any \(X\in {\mathcal {R}}\) if and only if \(S\in {\mathcal {Z}}_{S}({\mathcal {R}})\).

Proof

To check the “if” part, assume that \(S\in \mathcal {Z}_{S}({\mathcal {R}})\). It is easily seen that

$$\begin{aligned} {}_*[S,X]_2 = S^{2}X-2SXS^{*}+X{S^{*}}^{2}=(S-S)^{2}X=0 \end{aligned}$$

holds for all \( X\in {\mathcal {R}}.\)

Next we check the “only if” part.

Assume that \(S\not =0\) satisfies that

$$\begin{aligned} {}_*[S,X]_2 = S^{2}X-2SXS^{*}+X{S^{*}}^{2} = 0 \end{aligned}$$

for all \(X \in {\mathcal {R}}\); that is,

$$\begin{aligned} S^{2}X+X{S^{*}}^{2}=2SXS^{*} \end{aligned}$$
(4.1)

for all \(X \in {\mathcal {R}}\).

By Eq. (4.1) and Lemma 4.1, either \(\{I,S^2\}\) is linearly dependent over \({\mathcal {C}}\) (i.e., \(S^2\in {{\mathcal {Z}}}({{\mathcal {R}}})\)) or in particular there exists \(\alpha \in {{\mathcal {C}}}\) such that \(I=\alpha S^*\) (\(\alpha \not =0\)), so that \(S={\alpha ^{-1}}^* I\in {\mathcal Z}({{\mathcal {R}}})\). In any case, we have \(S^2\in {\mathcal Z}({{\mathcal {R}}})\). Thus Eq. (4.1) can be rewritten as \(X(S^2+{S^*}^2)=SX2S^*\). Again by Lemma 4.1, either \(2S^*=0\) (and \(S=0\) because char\(({{\mathcal {R}}})\not =2\)) or \(\{2S^*\}\) is linearly independent over \({\mathcal {C}}\) and there exists \(\beta \in {\mathcal {C}}\) such that \(S=\beta I\). In any case, \(S\in {{\mathcal {Z}}}({{\mathcal {R}}})\). Lastly, by Eq. (4.1) again, \((S^2-2SS^*+{S^*}^2)X=(S-S^*)^2X=(S-S^*)X(S-S^*)=0\) for every \(X\in {\mathcal {R}}\), which implies, since \({\mathcal {R}}\) is prime, that \(S-S^*=0\), i.e., \(S\in {{\mathcal {Z}}}_S({\mathcal {R}})\). \(\square \)

Now we are at a position to give our proof of the main theorem for the case that the ring is 2-torsion-free.

Proof of Theorem 3.1:

Char\(({\mathcal {R}})\not =2\). The “if” part of the theorem is obvious. In the sequel, we always assume that Char\(({\mathcal {R}})\not =2\) and \(\Phi :{\mathcal {R}} \rightarrow \mathcal {R}\) is a surjective strong 2-skew commutativity preserving map. We will check the “only if” part by several steps.

Step 1.\(\Phi \) is additive.

Note that, for any \(A,B,C\in {{\mathcal {R}}}\), we have \({}_*[C,A+B]_2={{}_*[C,A]_2}+{{}_*[C,B]_2}\). Then,

$$\begin{aligned}&{{}_*[\Phi (C), \Phi (A + B) - \Phi (A) - \Phi (B)]_2}\\&\quad = {{}_*[\Phi (C), \Phi (A + B)]_2} - {{}_*[\Phi (C), \Phi (A)]_2} -{{}_*[\Phi (C), \Phi (B)]_2}\\&\quad = {{}_*[C, A+B]_2 -{{}_*[C, A]_2} - {{}_*[C, B]_2}}\\&\quad = 0. \end{aligned}$$

By the surjectivity of \(\Phi \) and Lemma 4.2, the above equation implies \(\Phi (A+B)=\Phi (A)+\Phi (B)\). So \(\Phi \) is additive.

Step 2. For any nontrivial symmetric idempotent P and any \(X\in {\mathcal {R}}\), we have

$$\begin{aligned} {}_*[\Phi (P),X]_2\in P{\mathcal {R}}(I-P)+(I-P)\mathcal {R}P. \end{aligned}$$
(4.2)

Let \(P\in {\mathcal {R}}\) be a nontrivial symmetric idempotent. By the surjectivity of \(\Phi \), for every \(X\in {\mathcal {R}}\) there exists \(Y\in {\mathcal {R}}\) such that \(X=\Phi (Y)\). Hence \({}_*[\Phi (P),X]_2={{}_*[\Phi (P),\Phi (Y)]_2}={{}_*[P,Y]_2}\in {{\mathcal {R}}}_{12}+{{\mathcal {R}}}_{21}\), as it is easily checked.

Step 3. For any \(A\in {{\mathcal {R}}}\), we have \(\Phi (A^*)=\Phi (A)^*\).

Note that, for any \(A,B\in {{\mathcal {R}}}\), we have

$$\begin{aligned} {{}_*[B,A]_2}^*= {{}_*[B,A^*]_2}. \end{aligned}$$

Given \(A\in {{\mathcal {R}}}\), it follows that

$$\begin{aligned} {}_*[\Phi (B),\Phi (A)^*]_2{=}{{}_*[\Phi (B),\Phi (A)]_2}^*={{}_*[B,A]_2}^* ={{}_*[B,A^*]_2}{=}{{}_*[\Phi (B),\Phi (A^*)]_2} \end{aligned}$$

holds for all \(B\in {\mathcal {R}}\). Since \(\Phi \) is surjective, we see from the above equation that \({}_*[X, \Phi (A^*)-\Phi (A)^*]_2=0\) for all \(X\in {\mathcal {R}}\). Then, applying Lemma 4.2 ensures that \(\Phi (A^*)=\Phi (A)^*\).

Step 4. For any nontrivial symmetric idempotent \(P\in {\mathcal {R}}\), there exist \(\lambda , \mu \in {\mathcal {C}}_S\) such that \(\Phi (P) = \lambda P + \mu \) with \(\lambda \not =0\).

Let \(P\in {\mathcal {R}}\) be any nontrivial symmetric idempotent. Write \(P_1=P, P_2=I-P\) and \({{\mathcal {R}}}_{ij}=P_i{{\mathcal {R}}}P_j\). Then, there are \(S_{ij}\in {{\mathcal {R}}}_{ij}\) so that \(\Phi (P_1)=S_{11}+S_{12}+S_{21}+S_{22}\).

Taking \(X = X_{11}\in {\mathcal {R}}_{11}\) and applying Step 2, we have

$$\begin{aligned} 0=P_2{{}_*[\Phi (P_1),X_{11}]}_2P_2 =-2P_2\Phi (P_1)X_{11}{\Phi (P_1)^*}P_2 =-2S_{21}X_{11}S_{21}^*, \end{aligned}$$

which implies that \(S_{21}=0\) because \({\mathcal {R}}\) is prime and char\({\mathcal {R}}\ne 2\). Similarly, \(S_{12}=0\).

Hence one has \(\Phi (P) = S_{11} + S_{22}\). Let us determine \(S_{11}\) and \(S_{22}\) below.

Taking \( X= X_{11}\in {{\mathcal {R}}}_{11}\) and applying Step 2, we have

$$\begin{aligned} {}_*[ \Phi (P_1),X_{11}]_2 = S_{11}^2X_{11}-2S_{11}X_{11}S_{11}^*+X_{11}{S_{11}^*}^2=0 \end{aligned}$$

holds for all \(X_{11}\in {\mathcal {R}}_{11}\). That is \({}_*[S_{11},X_{11}]_2=0\) for all \(X_{11}\in {\mathcal {R}}_{11}\). By Lemma 4.3, \(S_{11}\in {\mathcal {Z}}_S({\mathcal {R}}_{11})\) since \({{\mathcal {R}}}_{11}\) is prime of characteristic not 2. It follows that \(S_{11}XP_1=P_1XS_{11}\) for all \(X\in {\mathcal {R}}\). Then, by Lemma 4.1, there exists some \(\lambda _1\in {\mathcal {C}}\) such that \(S_{11}=\lambda _1P_1\). Since \(S_{11}=S_{11}^*, (\lambda _1-\lambda _1^{*})P_1=S_{11}-S_{11}^*=0\), which entails that \(\lambda _1=\lambda _1^{*}\). So \(\lambda _1\in {\mathcal {C}}_S\). Analogously, there exists some \(\lambda _2\in {\mathcal {C}}_S\) such that \(S_{22}=\lambda _2P_2\). It follows that \(\Phi (P_1)=\lambda _1P_1+\lambda _2P_2=\lambda P + \mu I\), where \(\lambda =\lambda _1-\lambda _2,\ \mu =\lambda _2\).

Finally, we still need to prove that \(\lambda \ne 0\). On the contrary, we have \(\Phi (P_1)= \mu I\in {\mathcal {Z}}_S\), which implies \(X_{12}+X_{21}={{}_*[P_1, X]_2}={{}_*[\Phi (P_1),\Phi (X)]_2}=0\) for any \(X=X_{11}+X_{12}+X_{21}+X_{22}\in {\mathcal {R}}\), contradicting the primeness of \({\mathcal {R}}\). The proof of the step is completed.

Note that \({\mathcal {C}}\) is a field since \({\mathcal {R}}\) is prime. So \(\lambda \in {\mathcal {C}}\) is invertible. In the sequel, denote \(\alpha =\lambda ^{-1}\) and fix a nontrivial symmetric idempotent \(P_1\) and let \(P_2=I-P_1\).

Step 5. For any \(X_{ij}\in {\mathcal {R}}_{ij}\) with \(1\le i\ne j\le 2\), we have \(\Phi (X_{ij})=\alpha ^2 X_{ij}\).

For any \(X_{12}\in {\mathcal {R}}_{12}\), write \(\Phi (X_{12})= S_{11}+S_{12}+S_{21}+S_{22}\) with \(S_{ij}\in {{\mathcal {R}}}_{ij}\). Then

$$\begin{aligned} X_{12}=\,&{{}_*[P_1,X_{12}]_2}\!=\!{{}_*[\Phi (P_1),\Phi (X_{12})]_2}\!=\! {{}_*[\lambda P_1\!+\!\mu , \Phi (X_{12})]_2}={{}_*[\lambda P_1,\Phi (X_{12})]_2}\\ =\,&\lambda ^2 P_1 \Phi (X_{12})-2\lambda P_1\Phi (X_{12})(\lambda P_1)+\Phi (X_{12})(\lambda ^2 P_1)\\=\,&\lambda ^2(S_{12}+S_{21})\in {\mathcal {R}}_{12}. \end{aligned}$$

Hence we get \(\lambda ^2S_{12}=X_{12}\) and \( \lambda ^2S_{21}=0\), which implies that \(S_{21}=0\) and \(S_{12}=(\lambda ^2)^{-1}X_{12}=\alpha ^2 X_{12}\). It follows that \(\Phi (X_{12})=S_{11}+\alpha ^2 X_{12}+S_{22}\).

For any \(Y=Y_{11}+Y_{21}+Y_{22}\in {\mathcal {R}}\), by the surjectivity of \(\Phi \), there exists an element \(T=T_{11}+T_{12}+T_{21}+T_{22}\in {\mathcal {R}}\) such that \(\Phi (T)=Y\). So we have

$$\begin{aligned} {{}_*[\Phi (X_{12}),Y]_2}={{}_*[X_{12},T]_2}=-2X_{12}T_{22}X_{12}^*. \end{aligned}$$
(4.3)

Multiplying by \(P_2\) from both sides of Eq. (4.3) gives

$$\begin{aligned} P_2{{}_*[\Phi (X_{12}),Y]_2}P_2=P_2{{}_*[X_{12},T]_2}P_2=0. \end{aligned}$$

Thus, a simple computation reveals

$$\begin{aligned} S_{22}^2Y_{22}-2S_{22}Y_{22}S_{22}^*+Y_{22}{S_{22}^*}^2=P_2{{}_*[\Phi (X_{12}),Y]_2}P_2=0, \end{aligned}$$

which implies that

$$\begin{aligned} _*[S_{22}, Y_{22}]_2=0 \end{aligned}$$

holds for any \(Y_{22}\in {\mathcal {R}}_{22}\). So, by Lemma 4.3, \(S_{22}\in \mathcal {Z}_S({\mathcal {R}}_{22})\) and hence, \(S_{22}=\mu _2 P_2\) for some \(\mu _2\in {\mathcal {C}}_S \).

Take \(Y=Y_{21}\). Multiplying by \(P_2\) from the left side and \(P_1\) from the right of Eq. (4.3) gives

$$\begin{aligned} S_{22}^2Y_{21}+Y_{21}( S^*_{11})^2-2S_{22}Y_{21}S_{11}^*=0, \end{aligned}$$

which implies that \(Y_{21}(S_{11}^*-\mu _2 P_1)^2=0\) holds for all \(Y_{21}\in {\mathcal R}_{21}\). By the primeness of \({\mathcal {R}}\), we get \((S_{11}^*-\mu _2 P_1)^2=0\).

In summary, we have shown that, for any \(X_{12}\in {\mathcal R}_{12}\), there exists \(\mu (X_{12})\in {{\mathcal {C}}}_S\) such that

$$\begin{aligned} \Phi (X_{12})=\alpha ^2 X_{12}+S_{11}(X_{12})+\mu (X_{12})P_2, \end{aligned}$$
(4.4)

where \(S_{11}(X_{12})\in {{\mathcal {R}}}_{11}\) satisfies that

$$\begin{aligned} (S_{11}(X_{12})-\mu (X_{12})P_1)^2=0 \end{aligned}$$
(4.5)

In a similar way, one can show that, for any \(X_{21}\in {\mathcal R}_{21}\), there exists \(\mu (X_{21})\in {{\mathcal {C}}}_S\) such that

$$\begin{aligned} \Phi (X_{21})=\alpha ^2 X_{21}+\mu (X_{21})P_1+S_{22}(X_{21}), \end{aligned}$$
(4.6)

where \(S_{22}(X_{21})\in {{\mathcal {R}}}_{22}\) satisfies that

$$\begin{aligned} (S_{22}(X_{21})-\mu (X_{21})P_2)^2=0 \end{aligned}$$
(4.7)

Now, by Step 3, for any \(X_{12}\in {{\mathcal {R}}}_{12}\), Eqs. (4.4)–(4.7) entail that

$$\begin{aligned}&\alpha ^2X_{12}^*+(S_{11}(X_{12}))^*+\mu (X_{12})P_2=\Phi (X_{12})^*\\&\quad =\Phi (X_{12}^*)= \alpha ^2 X_{12}^*+\mu (X_{12}^*)P_1+S_{22}(X_{12}^*). \end{aligned}$$

It follows that \((S_{11}(X_{12}))^*=\mu (X_{12}^*)P_1\) and consequently, by Eq. (4.5), \(S_{11}(X_{12})=\mu (X_{12}^*)P_1=\mu (X_{12})P_1\). So

$$\begin{aligned} \Phi (X_{12})=\alpha ^2 X_{12} + \mu (X_{12}) P_1+\mu (X_{12}) P_2=\alpha ^2 X_{12} + \mu (X_{12}). \end{aligned}$$

By now we have shown that, for any \(X_{12}\in {{\mathcal {R}}}_{12}\), there exists \(\mu (X_{12})\in {{\mathcal {C}}}_S\) such that \(\Phi (X_{12})=\alpha ^2 X_{12}+\mu (X_{12})\). Next we prove that \(\mu (X_{12})=0\) holds for all \(X_{12}\).

For any \(Y_{12}\in {{\mathcal {R}}}_{12}\), we have

$$\begin{aligned} {}_*[P_1+Y_{12},X_{12}]_2=X_{12}-X_{12}Y_{12}^* \end{aligned}$$

and, by Step 1 and Step 4,

$$\begin{aligned}&X_{12}-X_{12}Y_{12}^* \\&\quad = {}_*[\Phi (P_1\!+\!Y_{12}),\Phi (X_{12})]_2={{}_*[\lambda P_1+\mu \!+\!\alpha ^2 Y_{12}+\mu (Y_{12}), \alpha ^2 X_{12}\!+\!\mu (X_{12})]_2}\\&\quad ={}_*[\lambda P_1+\alpha ^2 Y_{12}, \alpha ^2 X_{12}]_2+{{}_*[\lambda P_1+\alpha ^2 Y_{12}, \mu (X_{12})]_2}\\&\quad = \alpha ^2[(\lambda ^2P_1+\lambda \alpha ^2 Y_{12})X_{12}-2(\lambda P_1+\alpha ^2 Y_{12})X_{12}(\lambda P_1+\alpha ^2 Y_{12}^*)\\&\qquad +\,X_{12}(\lambda ^2P_1+\lambda \alpha ^2 Y_{12}^*)]\\&\qquad +\,\mu (X_{12})[(\lambda ^2P_1+\lambda \alpha ^2 Y_{12})-2(\lambda P_1+\alpha ^2 Y_{12})(\lambda P_1+\alpha ^2 Y_{12}^*)\\&\qquad +\,(\lambda ^2P_1+\lambda \alpha ^2 Y_{12}^*)]\\&\quad = \alpha ^2\lambda (\lambda X_{12}-\alpha ^2 X_{12}Y_{12}^*)+\mu (X_{12})(\alpha ^2\lambda Y_{12}+\alpha ^2\lambda Y_{12}^*-2\alpha ^4Y_{12}Y_{12}^*). \end{aligned}$$

This implies that \(\mu (X_{12})\alpha ^2\lambda Y_{12}^*=0\) for any \(Y_{12}\in {{\mathcal {R}}}_{12}\), which forces \(\mu (X_{12})=0\). Therefore, \(\Phi (X_{12})=\alpha ^2 X_{12}\) holds for each \(X_{12}\in {{\mathcal {R}}}_{12}\).

Similarly, for any \(X_{21}\in {{\mathcal {R}}}_{21}\) we have \(\Phi (X_{21})=\alpha ^2 X_{21}\).

Step 6. For any \(X_{ii}\in {\mathcal {R}}_{ii}, \Phi (X_{ii})=\lambda X_{ii}, i=1,2\).

Again, we only prove the assertion of Step 6 for the case \(i=1\). The case \(i=2\) is dealt with similarly.

For arbitrarily given \(X_{11}\in {\mathcal {R}}\), write \(\Phi (X_{11})=S_{11}+S_{12}+S_{21}+S_{22}\).

Since \( 0={{}_*[P_1, X_{11}]_2}={{}_*[\Phi (P_1), \Phi (X_{11})]_2}={{}_*[\lambda P_1+\mu , \Phi (X_{11})]_2}={{}_*[\lambda P_1, \Phi (X_{11})]_2} \), one can get \(\lambda ^2(S_{12}+S_{21})=0\), which entails that \(S_{12}=S_{21}=0\) and consequently, \(\Phi (X_{11})=S_{11}+S_{22}\).

Then, for any \(Y_{12}\in {\mathcal {R}}_{12}\), by Step 5, we have

$$\begin{aligned} \begin{array}{rl} 0=&{}{}_*{{[P_1+Y_{12}, X_{11}]}_2}={{}_*{[\Phi (P_1+Y_{12}), \Phi (X_{11})]}_2}\\ = &{} {{}_*[\lambda P_1+\mu +\alpha ^2Y_{12}, \Phi (X_{11})]_2}={{}_*[\lambda P_1+\alpha ^2 Y_{12}, S_{11}+S_{22}]_2}\\ =&{} (\lambda ^2P_1+\lambda \alpha ^2 Y_{12})(S_{11}+S_{22})-2(\lambda P_1+\alpha ^2 Y_{12})(S_{11}+S_{22})(\lambda P_1+\alpha ^2 Y_{12}^*)\\ &{} +\,(S_{11}+S_{22})(\lambda ^2P_1+\lambda \alpha ^2 Y_{12}^*)\\ =&{} \lambda \alpha ^2 Y_{12}S_{22}-2\alpha ^4Y_{12}S_{22}Y_{12}^*+\lambda \alpha ^2 S_{22}Y_{12}^*. \end{array} \end{aligned}$$

This implies that \(\lambda \alpha ^2 Y_{12}S_{22}=0\) holds for any \(Y_{12}\in {\mathcal {R}}_{12}\). By the primeness of \({\mathcal {R}}\), one sees that \(S_{22}=0\).

Now, for any \(Y_{21}\in {\mathcal {R}}_{21}\), we have

$$\begin{aligned}&Y_{21}X_{11}+X_{11}Y_{21}^*-2Y_{21}X_{11}Y_{21}^*={{}_*{{[P_1+Y_{21}, X_{11}]}_2}}\\&\quad ={{}_*{[\Phi (P_1+Y_{21}), \Phi (X_{11})]}_2}={{}_*[\lambda P_1+\alpha ^2Y_{21}, S_{11}]_2}\\&\quad =\lambda \alpha ^2 Y_{21}S_{11}+\lambda \alpha ^2S_{11}Y_{21}^*-2\alpha ^4Y_{21}S_{11}Y_{21}^*. \end{aligned}$$

This entails that \(Y_{21}X_{11}=\lambda \alpha ^2 Y_{21}S_{11}\) holds for any \(Y_{21}\in \mathcal {R}_{21}\). Recall that \(\alpha =\lambda ^{-1}\). So one gets \(P_2Y (S_{11}-\lambda X_{11})=0\) for all \(Y\in {\mathcal {R}}\), and therefore, \(S_{11}-\lambda X_{11}=0\) as \({\mathcal {R}}\) is prime. Consequently, \(\Phi (X_{11})=\lambda X_{11}\) as desired.

Step 7.\(\lambda ^3=I\) and \(\Phi (A)= \lambda A\) for all \(A\in {\mathcal {R}}\).

For any \(X_{11},Y_{11}\in {{\mathcal {R}}}_{11}\), by Step 6 we have

$$\begin{aligned} {}_*[X_{11},Y_{11}]_2={{}_*[\Phi (X_{11}),\Phi (Y_{11})]_2}={{}_*[\lambda X_{11},\lambda Y_{11}]_2}=\lambda ^3{{}_*[X_{11},Y_{11}]_2} \end{aligned}$$

which ensures that \(\lambda ^3=I\). Consequently, for any \(X_{ij}\in {{\mathcal {R}}}_{ij}\) with \(i\not =j\), by Step 5, \(\Phi (X_{ij})=\alpha ^2 X_{ij}=\lambda X_{ij}\).

Then, for any \(A\in {\mathcal {R}}\), writing \(A=A_{11}+A_{12}+A_{21}+A_{22}\) with \(A_{ij}\in {{\mathcal {R}}}_{ij}\), by Step 1, one gets

$$\begin{aligned} \begin{array}{rl} \Phi (A)=&{}\Phi (A_{11})+\Phi (A_{12})+\Phi (A_{21})+\Phi (A_{22})\\ = &{} \lambda (A_{11}+A_{12}+A_{21}+A_{22})=\lambda A. \end{array} \end{aligned}$$

The proof of Theorem 3.1 for the case char\(({{\mathcal {R}}})\not =2\) is finished. \(\square \)

5 Proof of Theorem 3.1: Char\(({{\mathcal {R}}})=2\)

In this section, we always assume that the characteristic of the ring is 2. Note that Lemma 4.1 and Lemma 4.2 are still true for prime \(*\)-rings of characteristic 2, but Lemma 4.3 is not valid in this situation. What we can have is the following lemma.

Lemma 5.1

Let \({\mathcal {R}}\) be a prime \(*\)-ring of characteristic 2 and \(S\in {{\mathcal {R}}}\). Then \({}_*[S,X]_2=0\) holds for any \(X\in {\mathcal {R}}\) if and only if \(S^2\in \mathcal {Z}_{S}({\mathcal {R}})\).

Proof

Since \({\mathcal {R}}\) has characteristic 2, \({}_*[ A,B]_2=A^2B+B(A^*)^2\). Thus if \(S^2\in {\mathcal {Z}}_{S}(\mathcal {R})\), we always have

$$\begin{aligned} {}_*[S,X]_2=S^2X+X(S^*)^2=2XS^2=0. \end{aligned}$$

Conversely, assume that \({}_*[S,X]_2=S^2X+X(S^*)^2=0\) holds for all \(X\in {\mathcal {R}}\). By Lemma 4.1, there exists \(\alpha \in {\mathcal C}\), the extended centroid of \({\mathcal {R}}\), such that \(S^2=\alpha I\). It follows that \(X(\alpha I+(S^*)^2)=0\) for all X. By the primeness of \({\mathcal {R}}, (S^*)^2=-\alpha I=-S^2=S^2\). Thus we have \(S^2\in {{\mathcal {Z}}}_S\). \(\square \)

Proof of Theorem 3.1:

Char\(({{\mathcal {R}}})=2\). We only need to check the “only if” part.

Let P be a nontrivial symmetric idempotent in \({\mathcal {R}}\), and let \(\Phi : {{\mathcal {R}}}\rightarrow {{\mathcal {R}}}\) be a map that is surjective and strong 2-skew commutativity preserving. It is easily checked that the assertions of Step 1–Step 3 in Sect. 4 are still valid for the case Char\(({{\mathcal {R}}})=2\). So \(\Phi \) is additive, \(\Phi (A^*)=\Phi (A)^*\) and

$$\begin{aligned} {}_*[\Phi (P),X]_2\in P{\mathcal {R}}(I-P)+(I-P){\mathcal {R}}P \end{aligned}$$

holds for every \(X\in {\mathcal {R}}\). We shall complete the remainder of the proof by several steps.

Step 1. There exist \(\lambda , \mu \in {{\mathcal {C}}}_S\) such that \(\Phi (P)^2=\lambda P+\mu \) with \(\lambda \not =0\).

Again, write \(P_1=P, P_2=I-P\) and \({{\mathcal {R}}}_{ij}=P_i{\mathcal R}P_j, i,j\in \{1,2\}\). Then \(\Phi (P_1)^2=S_{11}+S_{12}+S_{21}+S_{22}\) for some \(S_{ij}\in {{\mathcal {R}}}_{ij}\). Note that \(\Phi (P_1)=\Phi (P_1)^*\) as \(P_1^*=P_1\).

For any \(X_{12}\in {{\mathcal {R}}}_{12}\), since

$$\begin{aligned} 0=P_2 {{}_*[\Phi (P_1), X_{12}]_2}P_2=P_2(\Phi (P_1)^2X_{12}+X_{12}(\Phi (P_1)^*)^2)P_2=P_2\Phi (P_1)^2X_{12}, \end{aligned}$$

we see that \(S_{21} X_{12}=0\) holds for each \(X_{12}\). Then it follows from the primeness of \({\mathcal {R}}\) we have \(S_{21}=0\). Similarly, one can check that \(S_{12}=0\). Hence \(\Phi (P_1)^2=S_{11}+S_{22}\).

For any \(X_{11}\in {{\mathcal {R}}}_{11}\), there exists \(Y=Y_{11}+Y_{12}+Y_{21}+Y_{22}\) such that \(\Phi (Y)=X_{11}\) as \(\Phi \) is surjective. Then

$$\begin{aligned} Y_{12}+Y_{21}={{}_*[P_1,Y]_2}={{}_*[\Phi (P_1),X_{11}]_2}=S_{11}X_{11}+X_{11}S_{11}. \end{aligned}$$

This means that \(S_{11}X_{11}+X_{11}S_{11}=0\). Therefore, for any \(X_{11}\) we have

$$\begin{aligned} S_{11}=\lambda _1 P_1 \end{aligned}$$

for some \(\lambda _1\in {{\mathcal {C}}}_S\) by Lemma 4.1. Similarly,

$$\begin{aligned} S_{22}=\mu P_2 \end{aligned}$$

for some \(\mu \in {{\mathcal {C}}}_S\). Consequently,

$$\begin{aligned} \Phi (P_1)^2=\lambda P_1+\mu \end{aligned}$$

with \(\lambda =\lambda _1-\mu \in {{\mathcal {C}}}_S\). Note that \(\lambda \not =0\) because otherwise, by Lemma 5.1, \(P_1X+XP_1={{}_*[P_1,X]_2}={{}_*[\Phi (P_1),\Phi (X)]_2}=0\) for every \(X\in {{\mathcal {R}}}\), which is impossible by primeness.

Step 2.\( \Phi ({{\mathcal {R}}}_{ii})\subset {\mathcal R}_{11}+{{\mathcal {R}}}_{22}\) and \(\Phi ({{\mathcal {R}}}_{ij})\subset {{\mathcal {R}}}_{11}+{{\mathcal {R}}}_{ij}+{{\mathcal {R}}}_{22}\) for \(i\not =j. \)

For any \(X\in {\mathcal {R}}\), since, by Step 1,

$$\begin{aligned} \begin{array}{rl} P_1X+XP_1=&{}{{}_*[P_1,X]_2}={{}_*[\Phi (P_1),\Phi (X)]_2}\\ =&{}(\lambda P_1+\mu )\Phi (X)+\Phi (X)(\lambda P_1+\mu )=\lambda (P_1\Phi (X)+\Phi (X)P_1),\end{array} \end{aligned}$$

we have

$$\begin{aligned} P_1XP_2=\lambda P_1\Phi (X)P_2\quad \mathrm{and}\quad P_2XP_1=\lambda P_2\Phi (X)P_1. \end{aligned}$$
(5.1)

Clearly, Eq. (5.1) implies that

$$\begin{aligned} \Phi ({{\mathcal {R}}}_{ii})\subset {{\mathcal {R}}}_{11}+{{\mathcal {R}}}_{22}\quad \mathrm{and}\quad \Phi ({{\mathcal {R}}}_{ij})\subset {{\mathcal {R}}}_{11}+{\mathcal R}_{ij}+{{\mathcal {R}}}_{22}\ \ \mathrm{for}\ i\not =j. \end{aligned}$$

Step 3.\(\Phi (X_{ij})=\lambda ^{-1}X_{ij}\) for any \(X_{ij}\in {{\mathcal {R}}}_{ij}\) with \(i\not =j\).

If \(X_{12}\in {{\mathcal {R}}}_{12}\), then, by Step 2, we have \(\Phi (X_{12})=S_{11}+\lambda ^{-1}X_{12}+S_{22}\) for some \(S_{ii}\in {{\mathcal {R}}}_{ii}, i=1,2\). We have to show that \(S_{11}=S_{22}=0\). To do this, take any \(W_{21}\in {{\mathcal {R}}}_{21}\) and take Y so that \(\Phi (Y)=P_1+W_{21}\). By Step 2, it is easily checked that \(Y\in {{\mathcal {R}}}_{11}+{{\mathcal {R}}}_{21}+{\mathcal R}_{22}\) and \(P_2{{}_*[Y,X_{12}]_2}P_1=0\). Therefore, we have

$$\begin{aligned} \begin{array}{rl} 0=&{}P_2{{}_*[\Phi (Y),\Phi (X_{12})]_2}P_1=P_2{{}_*[P_1+W_{21},S_{11}+\lambda ^{-1}X_{12}+S_{22}]_2}P_1\\ =&{} P_2((P_1+W_{21})(S_{11}+\lambda ^{-1}X_{12}+S_{22})+(S_{11}+\lambda ^{-1}X_{12}+S_{22})(P_1+W_{21}^*))P_1\\ =&{}W_{21}S_{11}. \end{array} \end{aligned}$$

As \(W_{21}\) is arbitrary, the primeness of \({\mathcal {R}}\) implies that \(S_{11}=0\). Similarly, considering \(P_1+W_{12}\) gives \(S_{22}=0\). Hence \(\Phi (X_{12})=\lambda ^{-1}X_{12}\) holds for every \(X_{12}\in {{\mathcal {R}}}_{12}\).

The case for \(X_{21}\in {{\mathcal {R}}}_{21}\) can be dealt with similarly.

Step 4. There exists \(\delta \in {{\mathcal {C}}}_S\) with \(\delta ^3=I\) such that \(\Phi (X_{ij})=\delta X_{ij}\) for each \(X_{ij}\in {{\mathcal {R}}}_{ij}, i,j\in \{1,2\}\).

Write \(\Phi (P_1)=S_{11}+S_{22}\) and note that \(\Phi (P_1)^*=\Phi (P_1)\). For any \(X_{11}\in {{\mathcal {R}}}_{11}\) and \(X_{12}\in {{\mathcal {R}}}_{12}\), it follows from Step 2 and Step 3 that \(\Phi (X_{11})=A_{11}+A_{22}\) for some \(A_{ii}\in {\mathcal R}_{ii}, i=1,2\) and

$$\begin{aligned}\begin{array}{rl} 0=&{}P_2{{}_*[P_1+X_{12},X_{11}]_2}P_1=P_2{{}_*[\Phi (P_1)+\Phi (X_{12}),\Phi (X_{11})]_2}P_1\\ =&{}P_2(\Phi (P_1) +\lambda ^{-1}X_{12})^2\Phi (X_{11})+\Phi (X_{11})(\Phi (P_1)+\lambda ^{-1}X_{12}^*)^2)P_1\\ =&{}P_2(\lambda P_1+\mu +\lambda ^{-1}(S_{11}+S_{22})X_{12}+\lambda ^{-1}X_{12}(S_{11}+S_{22}))(A_{11}+A_{22})P_1\\ &{} +\,P_2(A_{11}+A_{22})(\lambda P_1+\mu + \lambda ^{-1}X_{12}^*(S_{11}+S_{22})+\lambda ^{-1}(S_{11}+S_{22})X_{12}^*)P_1\\ =&{} \lambda ^{-1}(A_{22}X_{12}^*S_{11}+A_{22}S_{22}X_{12}^*). \end{array} \end{aligned}$$

Consequently,

$$\begin{aligned} S_{11}X_{12}A_{22}^*=X_{12}S_{22}A_{22}^* \end{aligned}$$
(5.2)

holds for all \(X_{12}\in {{\mathcal {R}}}_{12}\). Taking \(X_{11}=P_1\) in Eq. (5.2) and applying Lemma 4.1 again, since \(P_1\not =0\), one gets

$$\begin{aligned} S_{22}^2=\alpha S_{22} \end{aligned}$$

for some \(\alpha \in {{\mathcal {C}}}_S\). Suppose \(S_{22}\not =0\) to reach a contradiction; then, we get \(S_{11}=\alpha P_1\) and \(\alpha A_{22}^*=S_{22}A_{22}^* \) from Eq. (5.2) by primeness. Clearly, \(\alpha \not =0\) as \(\alpha ^2 P_1=S_{11}^2=\lambda P_1\not =0\). Since \(S_{22}\not =0, \mu \not =0\). Then it follows from the above equations that \(S_{22}=\alpha ^{-1}\mu P_2=\alpha P_2\). Thus we have \(\Phi (P_1)=S_{11}+S_{22}=\alpha \) which implies that \(\mu =\Phi (P_1)^2=\lambda P_1+\mu \), a contradiction since \(\lambda \not =0\). So \(S_{22}=0\). Due to \(S_{11}^2+S_{22}^2=\Phi (P_1)^2=(\lambda +\mu ) P_1+\mu P_2\), we get \(\mu =0\) and \(S_{11}\not =0\). Then, it follows from Eq. (5.2) we get \(S_{11}X_{12}A_{22}^*=0\) for every \(X_{12}\), which forces \(A_{22}=0\). Therefore, \(\Phi (X_{11})=A_{11}\) for every \(X_{11}\in {{\mathcal {R}}}_{11}\).

Similarly, for every \(X_{22}\in {{\mathcal {R}}}_{22}\) we can show that \(\Phi (X_{22})=B_{22}\in {{\mathcal {R}}}_{22}\).

Now, for any \(X_{12}\in {{\mathcal {R}}}_{12}\) and \(X_{22}\in {\mathcal R}_{22}\) we have

$$\begin{aligned}&X_{12}X_{22}+X_{22}X_{12}^*={{}_*[P_1+X_{12}, X_{22}]_2}\\&\quad ={}_*[\Phi (P_1)+\Phi (X_{12}), \Phi (X_{22})]_2={{}_*[S_{11}+\lambda ^{-1}X_{12}, B_{22}]_2}\\&\quad = (S_{11}+\lambda ^{-1}X_{12})^2B_{22}+B_{22}(S_{11}+\lambda ^{-1}X_{12}^*)^2\\&\quad = \lambda ^{-1}(S_{11}X_{12}B_{22}+B_{22}X_{12}^*S_{11}), \end{aligned}$$

which implies that

$$\begin{aligned} \lambda ^{-1}S_{11}X_{12}B_{22}=X_{12}X_{22} \end{aligned}$$
(5.3)

holds for all \(X_{12}\in {{\mathcal {R}}}_{12}\). Put \(\Phi (P_2)=T_{22}\). Since \(\lambda ^{-1}T_{22}\not =0\), taking \(X_{22}=P_2\) in Eq. (5.3) we find by Lemma 4.1 that

$$\begin{aligned} S_{11}=\delta P_1 \end{aligned}$$

for some \(\delta \in {{\mathcal {C}}}_S\). In addition, by primeness we get \(\delta ^2=\lambda \) from \(\delta ^2P_1=S_{11}^2=\lambda P_1\).

Moreover, for any \(X_{21}\in {{\mathcal {R}}}_{21}\),

$$\begin{aligned} X_{21}X_{11}+X_{11}X_{21}^*= & {} {{}_*[P_1+X_{21},X_{11}]_2}\\= & {} {{}_*[\Phi (P_1+X_{21}),\Phi (X_{11})]_2} = {{}_*[\delta P_1+\lambda ^{-1}X_{21}, A_{11}]_2}\\= & {} \delta \lambda ^{-1}(X_{21}A_{11}+A_{11}X_{21}^*)=\delta ^{-1}(X_{21}A_{11}+A_{11}X_{21}^*). \end{aligned}$$

It follows that \(X_{21}X_{11}=\delta ^{-1}X_{21}A_{11}\) for all \(X_{21}\) and consequently, \(A_{11}=\delta X_{11}\).

In addition, because

$$\begin{aligned} {}_*[X_{11},Y_{11}]_2={{}_*[\Phi (X_{11}),\Phi (Y_{11})]_2}={{}_*[\delta X_{11},\delta Y_{11}]_2}=\delta ^3{{}_*[X_{11},Y_{11}]_2} \end{aligned}$$

holds for any \(X_{11},Y_{11}\in {{\mathcal {R}}}_{11}\), we see that \(\delta ^3=I\). Thus we have \(\lambda ^{-1}=\delta ^{-2}=\delta \) and \(\Phi (X_{ij})=\lambda ^{-1}X_{ij}=\delta X_{ij}\) holds for any \(X_{ij}\) with \(i\not =j\).

Analogously, the consideration of \({}_*[P_1+X_{12},X_{22}]_2\) gives \(\Phi (X_{22})=\delta X_{22}\) for any \(X_{22}\in {{\mathcal {R}}}_{22}\).

Step 5.\(\Phi (A)=\delta A\) for any \(A\in {{\mathcal {R}}}\).

This follows directly from Step 4 by additivity.

This completes the proof of Theorem 3.1 for the case char\(({{\mathcal {R}}})=2\). \(\square \)