1 Introduction and statement of the main results

Let \(\mathcal {A}\) be an algebra. For \(A, B \in \mathcal {A}\), \(A\circ B=AB+BA\) is the Jordan product and \([A,B]=AB-BA\) is the Lie product. The map \(\Phi {:}\,\mathcal {A}\longrightarrow \mathcal {A}\) is Lie product preserving if \( \Phi [A,B]=[\Phi (A),\Phi (B)]\) for all \(A, B \in \mathcal {A}\). This kind of maps are closely related to Lie homomorphisms (linear maps preserving Lie products) which have been studied by many authors. For example see [2, 5, 7, 10, 11] and their references. Šemrl [17] characterized bijective maps preserving the Lie product on B(X). Zhang and Zhang [19] characterized the structure of bijective maps preserving the Lie product on factor von Neumann algebras, with a different method. Also the problem of characterizing maps preserving the commutativity (maps preserving zero Lie products), studied intensively (see [1, 2, 5, 7, 13, 17] and references therein).

In [1], Bell and Daif gave the conception of strong commutativity preserving maps. We say that a map \(\Phi {:}\,\mathcal {A}\longrightarrow \mathcal {A}\) is strong commutativity preserving if \([\Phi (A),\Phi (B)]=[A,B]\), for all \(A, B\in \mathcal {A}\). Note that a strong commutativity preserving map must be commutativity preserving, but the reverse is not true generally. These maps are also called strong Lie product preserving maps. Brešar and Miers [5] proved that every strong commutativity preserving additive map \(\Phi \) on a semiprime ring \(\mathcal {R}\) is of the form \(\Phi (A) =\lambda A + \mu (A)\), where \(\lambda \in \mathcal {C}\), the extended centroid of \(\mathcal {R}\), \(\lambda ^2 = 1\), and \(\mu {:}\,\mathcal {R} \longrightarrow \mathcal {C}\) is an additive map. In [9], Lin and Liu obtained a similar result on a noncentral Lie ideal of a prime ring. In [14, 15], Qi and Hou gave a complete characterization of strong commutativity preserving surjective maps (without the assumption of additivity) on prime rings and triangular algebras, respectively.

Let \(\mathcal {A}\) be a *-ring. For any \(A, B \in \mathcal {A}\), \([A, B]_* = AB-BA^*\) denotes the skew Lie product of A and B. This kind of product is found playing a more and more important role in some research topics such as representing quadratic functionals with sesquilinear functionals, and its study has attracted many authors attention (see [3, 6, 18] and the references therein). Molnar [12] initiated the systematic study of this skew Lie product, and studied the relation between subspaces and ideals of \(B(\mathcal {H})\), the algebra of all bounded linear operators acting on a Hilbert space \(\mathcal {H}\).

Recall that a map \(\Phi {:}\, \mathcal {A}\longrightarrow \mathcal {A}\) is called zero skew Lie product preserving, if \(\Phi (A)\Phi (B)-\Phi (B)\Phi (A)^*=0\) whenever \(AB-BA^* = 0\) for any \(A, B \in \mathcal {A}\). Additive or linear maps preserving zero skew Lie products on various rings and algebras had been studied by many authors (see [1] and the references therein). More specially, \(\Phi \) is strong skew commutativity preserving, if \([\Phi (A), \Phi (B)]_* = [A, B]_*\) for all \(A, B\in \mathcal {A}\). It is obvious that strong skew commutativity preserving maps must be zero skew Lie product preserving. However, the reverse is not true in general. In [8], Cui and Park proved that, if \(\mathcal {A}\) is a factor von Neumann algebra, then every strong skew commutativity preserving map \(\Phi \) on \(\mathcal {A}\) has the form \(\Phi (A) = \Psi (A) + h(A)I\) for all \(A\in \mathcal {A}\), where \(\Psi {:}\, \mathcal {A}\longrightarrow \mathcal {A}\) is a linear bijective map satisfying \(\Psi (A)\Psi (B)-\Psi (B)\Psi (A)^* = AB-BA^*\) for all \(A, B\in \mathcal {A}\) and h is a real linear functional of \(\mathcal {A}\) with \(h(0) = 0\); particularly, if \(\mathcal {A}\) is of type I, then \(\Phi (A) = cA + h(A)I\) for each \(A\in \mathcal {A}\), where \(c \in \{-1, 1\}\). Recently, Qi and Hou [18], generalized the above result to von Neumann algebras without central summand of type \(I_1\).

The purpose of the present paper is to continue these studies and to consider a special type of nonlinear strong skew Jordan product preserving maps on von Neumann algebras.

Note that a subalgebra \(\mathcal {A}\) of \(B(\mathcal {H})\) is called a von Neumann algebra when it is closed under the weak operator topology. A von Neumann algebra \(\mathcal {A}\) is called factor when its center is trivial. We denote the real and the imaginary part of an operator T by Re(T) and Im(T), respectively, i.e. \(Re(T) = (T+T^*)/2\) and \(Im(T) = (T-T^*)/2i\). Recall that a von Neumann algebra is the closed linear span of the set of its projections. In this paper we will use this fact several times. For convenience, from here let \(\mathcal {Z}(\mathcal {A})\) denote the center of \(\mathcal {A}\), \(\mathcal {P}(\mathcal {A})\) the set of all projections in \(\mathcal {A}\), \(\mathcal {A}_{ sa}\) the space of all Hermitian elements in \(\mathcal {A}\) and define \(Q=I-P\) for an arbitrary projection P.

We are now ready to state the main results of the paper.

Theorem 1.1

Let \(\mathcal {H}\) be a complex Hilbert space and \(\mathcal {A}\) be a von Neumann algebra acting on \(\mathcal {H}\). Assume that a surjective mapping \(\Phi \) on \(\mathcal {A}\) satisfies the condition

$$\begin{aligned} \Phi (T)\Phi (P)+\Phi (P)\Phi (T)^*=TP+PT^*, \end{aligned}$$
(1.1)

for all \(T\in \mathcal {A}\) and every \(P\in \mathcal {P}(\mathcal {A})\). Then \(\Phi (I)\in \mathcal {Z}(\mathcal {A}) \) and there exists a map \(f{:}\,{\mathcal {A}_{ sa}} \longrightarrow \mathcal {Z}(\mathcal {A})\) such that f(T) is a selfadjoint element in \(\mathcal {Z}(\mathcal {A})\) and \(\Phi (T)=T\Phi (I)+if(T)\) for all \(T\in \mathcal {A}_{ sa}\). More precisely, \(f(T)=Im(\Phi (T))\). In particular, for all \(T\in \mathcal {A}_{ sa}\) for which \(\Phi (T)\) is in \(\mathcal {A}_{ sa}\), we have \(\Phi (T)=T\Phi (I)\).

Theorem 1.2

Let \(\mathcal {H}\) be a complex Hilbert space and \(\mathcal {A}\) be a factor von Neumann algebra acting on \(\mathcal {H}\). Assume that a surjective mapping \(\Phi \) on \(\mathcal {A}\) satisfies the condition 1.1 for all \(T\in \mathcal {A}\) and every \(P\in \mathcal {P}(\mathcal {A})\). Then there is a real map \(\tau \) on \(\mathcal {A}\) such that \(\Phi (T)=T+i\tau (T) I\) for all \(T\in \mathcal {A}\) or \(\Phi (T)=-T+i\tau (T)I\).

2 The proof of the main results

Proof of Theorem 1.1

We will complete the proof by checking several steps.

Step 1 (i) \(\Phi (0)=0\),

(ii) \(\Phi \) is injective on \(\mathcal {A}_{ sa}\).

  1. (i)

    For every \(T\in \mathcal {A}\) we have \(T0+0T^*=0\). By Eq. 1.1 we obtain that \(\Phi (T)\Phi (0)+\Phi (0)\Phi (T)^*=0\). Because of the surjectivity of \(\Phi \) we have \(I \Phi (0)+\Phi (0)I^*=0\) and therefore \(\Phi (0)=0\).

  2. (ii)

    Assume that TS are selfadjoint elements in \(\mathcal {A}\) such that \(\Phi (T)=\Phi (S)\). By Eq. 1.1 we have

    $$\begin{aligned} \Phi (T)\Phi (I)+\Phi (I)\Phi (T)^*=TI+IT, \end{aligned}$$

    and

    $$\begin{aligned} \Phi (S)\Phi (I)+\Phi (I)\Phi (S)^*=SI+IS. \end{aligned}$$

Subtracting these two latter equations one gets \(T-S=0\) or \(T=S\), means that \(\Phi \) is injective on selfadjoint elements in \(\mathcal {A}\).

Step 2 For all \(P\in \mathcal {P}(\mathcal {A})\), we have

  1. (i)

    \(\Phi (P)^*=\Phi (P),\)

  2. (ii)

    \(\Phi (I)=\Phi (P)+\Phi (Q)\),

  3. (iii)

    \(\Phi (P)^2=P.\)

  1. (i)

    Since \(\Phi \) is surjective so there exists \(T\in \mathcal {A}\) such that \(\Phi (T)=I\). Therefore by Eq. 1.1

    $$\begin{aligned} I\Phi (P)+\Phi (P)I^*=TP+PT^* \end{aligned}$$

    or \(2\Phi (P)=TP+PT^*\). Obviously, we have \(2\Phi (P)^*=TP+PT^*\) and this implies that \(\Phi (P)^*=\Phi (P)\).

  2. (ii)

    Similarly to the part (i), there exists \(T\in \mathcal {A}\) such that \(\Phi (T)=I\) and

    $$\begin{aligned} 2\Phi (P)=TP+PT^*, \end{aligned}$$
    (2.1)

    for all \(P\in \mathcal {P}(\mathcal {A})\). Let \(P=Q\) in the above equation, thus

    $$\begin{aligned} 2\Phi (Q)=TQ+QT^*. \end{aligned}$$
    (2.2)

    Since \(P+Q=I\), so adding these two equations implies that

    $$\begin{aligned} 2(\Phi (P)+\Phi (Q))=T+T^*. \end{aligned}$$
    (2.3)

    On the other hand by substituting P by I in Eq. 1.1 one gets

    $$\begin{aligned} 2\Phi (I)=T+T^*. \end{aligned}$$
    (2.4)

    Comparing Eqs. 2.3 and 2.4 we obtain that \(\Phi (I)=\Phi (P)+\Phi (Q)\).

  3. (iii)

    Let \(T=P\) in Eq. 1.1. By part (i) it is clear that \(\Phi (P)^2=P\).

Step 3 \(\Phi (P)\Phi (Q)=\Phi (Q)\Phi (P)=0\) and \(\Phi (P)Q=Q\Phi (P)=0\), for all \(P\in \mathcal {P}(\mathcal {A})\).

Let \(T=\Phi (P)\Phi (Q)\), so by part (iii) of step 2 and the fact that \(\Phi (P)Q=Q\Phi (P)\) for all \(P\in \mathcal {P}(\mathcal {A})\), we have

$$\begin{aligned} TT^*=\Phi (P)\Phi (Q)^2\Phi (P)=Q\Phi (P)^2. \end{aligned}$$

Again by using step 2, \(TT^*=0\). Similarly, we can show that \(T^*T=0\). Consequently, \(T=\Phi (P)\Phi (Q)=0\).

The second part can be obtained easily from the first part and step 2.

Step 4 \(\Phi (I)\in \mathcal {Z}(\mathcal {A})\).

Let \(T=I\) in Eq. 1.1, so by part (i) of step 2 we have

$$\begin{aligned} \Phi (I)\Phi (P)+\Phi (P)\Phi (I)=2P, \end{aligned}$$

for all \(P\in \mathcal {P}(\mathcal {A})\). Multiplying by \(\Phi (I)\) from the left and right, respectively, one gets

$$\begin{aligned} \Phi (P)+\Phi (I)\Phi (P)\Phi (I)=2\Phi (I)P \end{aligned}$$

and

$$\begin{aligned} \Phi (I)\Phi (P)\Phi (I)+\Phi (P)=2P\Phi (I). \end{aligned}$$

These last two equations imply that \(\Phi (I)P=P\Phi (I)\) for all \(P\in \mathcal {P}(\mathcal {A})\). As we mentioned in Sect. 1, each operator in a von Neumann algebra is a limit of finite linear combinations of projections in \(\mathcal {A}\). Thus we have \(\Phi (I)T=T\Phi (I)\) for all \(T\in \mathcal {A}\), means that \(\Phi (I)\) is in the center of \(\mathcal {A}\).

Step 5 \(\Phi (P)=P\Phi (I)\), for all \(P\in \mathcal {P}(\mathcal {A})\).

Let \(\Phi (P)\Phi (I)=S\). By step 2, part (ii), we have \(\Phi (P)(\Phi (P)+\Phi (Q))=S\). Since \(\Phi (P)^2=P\) and \(\Phi (P)\Phi (Q)=0\) one gets that \(P=S\). Consequently, \(\Phi (P)\Phi (I)=P\). Multiplying by \(\Phi (I)\) from the right we obtain \(\Phi (P)=P\Phi (I)\).

Step 6 There is a map \(f{:}\,{\mathcal {A}_{ sa}} \longrightarrow \mathcal {Z}(\mathcal {A})\) such that \(f(T)=Im(\Phi (T))\) and \(\Phi (T)=T\Phi (I)+if(T)\) for all \(T\in \mathcal {A}_{ sa}\). In particular, for all \(T\in \mathcal {A}_{ sa}\) for which \(\Phi (T)\) is in \(\mathcal {A}_{ sa}\) we have \(f(T)=0\) i.e. \(\Phi (T)=T\Phi (I)\).

For every Hermitian element \(T\in \mathcal {A}\), Eq. 1.1 implies that

$$\begin{aligned} \Phi (T)\Phi (P)+\Phi (P)\Phi (T)^*=TP+PT, \end{aligned}$$
(2.5)

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

Let \(P=I\) in the above equation and multiply by \(\Phi (I)\) from the right. This implies that \(Re \Phi (T)=T\Phi (I)\). So we have \(\Phi (T)=T\Phi (I)+iU\), where U is the imaginary part of \(\Phi (T)\). Now let \(\Phi (T)=T\Phi (I)+iU\) in Eq. 2.5. Since \(\Phi (P)=P\Phi (I)\) it is easily seen that \(UP=PU\), for all \(P\in \mathcal {P}(\mathcal {A})\). It means that U is in the center of \(\mathcal {A}\). So there exists a map \(f{:}\,{\mathcal {A}_{ sa}} \longrightarrow \mathcal {Z}(\mathcal {A})\) such that \(f(T)=U=Im(\Phi (T))\) and \(\Phi (T)=T\Phi (I)+if(T)\) for all \(T\in \mathcal {A}_{ sa}\).

Now suppose that T and \(\Phi (T)\) are in \(\mathcal {A}_{ sa}\). Clearly, \(f(T)=0\) and \(\Phi (T)=T\Phi (I)\). This completes the proof of Theorem 1.1. \(\square \)

Proof of Theorem 1.2

By checking the proof of Theorem 1.1, steps 1–4 still hold. We complete the proof using several further steps.

Step 1 \(\Phi (I)=I\) or \(\Phi (I)=-I\).

By step 4 of the proof of Theorem 1.1, we know that \(\Phi (I)\) is in the center of \(\mathcal {A}\). Since \(\mathcal {A}\) is a factor, one can conclude that \(\Phi (I)\in \mathbb {C}I\). By the facts that \(\Phi (I)^*=\Phi (I)\) and \(\Phi (I)^2=I\), it is easily seen that \(\Phi (I)=I\) or \(\Phi (I)=-I\).

From here we suppose that \(\Phi (I)=I\) without loss of generality.

Step 2 \(\Phi (P)=P\) for all \(P\in \mathcal {P}(\mathcal {A})\).

It is obvious because of step 5 of the proof of Theorem 1.1 and the fact that \(\Phi (I)=I\).

Step 3 \(\Phi \) preserves skew selfadjoint elements of \(\mathcal {A}\).

Suppose that T is a skew selfadjoint element in \(\mathcal {A}\), that is \(T+T^*=0\). By Eq. 1.1 we have

$$\begin{aligned} \Phi (T)\Phi (I)+\Phi (I)\Phi (T)^*=0. \end{aligned}$$

Since \(\Phi (I)=I\) it is obvious that \(\Phi (T)=-\Phi (T)^*\).

Step 4 There exists a real functional \(\tau _1\) on selfadjoint elements in \(\mathcal {A}\) such that \(\Phi (T)=T+i\tau _1(T) I\) for all \(T\in \mathcal {A}_{ sa} \).

Similarly to step 6 of the proof of Theorem 1.1 and by the fact that \(\Phi (I)=I\), we can show that \(\Phi (T)=T+iU\), where U is in the center of \(\mathcal {A}\). Since \(\mathcal {A}\) is a factor and U is selfadjoint so there exists \(\alpha \in \mathbb {R}\), corresponding to T, such that \(U=\alpha I\), Hence we conclude that there is a real functional \(\tau _1{:}\, {\mathcal {A}_{ sa}} \longrightarrow \mathbb {R}\) such that \(\Phi (T)=T+i\tau _1(T) I\).

Step 5 There exists a real functional \(\tau _2\) on selfadjoint elements in \(\mathcal {A}\) such that \(\Phi (iT)=iT+i\tau _2(T) I\) for all \(T\in \mathcal {A}_{ sa} \).

Assume that T is a selfadjoint element in \(\mathcal {A}\), so iT is skew selfadjoint. By step 3 we know that \(\Phi \) preserves skew selfadjoint elements, thus \(\Phi (iT)^*=-\Phi (iT)\), so Eq. 1.1 implies

$$\begin{aligned} \Phi (iT)\Phi (P)-\Phi (P)\Phi (iT)=iTP-iPT, \end{aligned}$$
(2.6)

for all \(P\in \mathcal {P}(\mathcal {A})\). Since \(\phi (P) = P\) in the above equation, therefore we have

$$\begin{aligned} (\Phi (iT)-iT)P=P(\Phi (iT)-iT), \end{aligned}$$

for every \(P\in \mathcal {P}(\mathcal {A})\). Since \(\mathcal {A}\) is a von Neumann algebra thus \(\Phi (iT)-iT \in \mathcal {Z}(\mathcal {A})\). \(\mathcal {A}\) is a factor, consequently, there exists \(\lambda \in \mathbb {C}\) corresponding to iT such that \(\Phi (iT)-iT=\lambda I \). Note that \(\Phi (iT)-iT\) is skew selfadjoint, therefore \(\lambda =i\beta \) for \(\beta \in \mathbb {R}\). These statements imply that there is a real functional \(\tau _2\) on selfadjoint elements of \(\mathcal {A}\) such that \(\Phi (iT)=iT+i\tau _2 (T)I\).

Step 6 \(\Phi (T)=T+i\tau (T) I\) for all \(T\in \mathcal {A}\), where \(\tau \) is a real functional on \(\mathcal {A}\) such that \(\tau (P)=0\) for every \(P\in \mathcal {P}(\mathcal {A})\).

It is obvious by steps 4 and 5. Clearly, \(\tau \) is a real functional and \(\tau (P)=0\) for every \(P\in \mathcal {P}(\mathcal {A})\).

Similarly, it can be shown that if \(\Phi (I)=-I\) then \(\Phi \) is of the form \(\Phi (T)=-T+i\tau (T)I\) for all \(T\in \mathcal {A}\). \(\square \)