1 Introduction

Let \({\mathcal {A}}\) be an algebra over \({\mathbb {C}}\) the field of complex numbers and \({\mathcal {M}}\) be an \({\mathcal {A}}\)-bimodule. A linear map \(\delta \) from \({\mathcal {A}}\) into \({\mathcal {M}}\) is called a Jordan derivation if \(\delta (a^{2})=\delta (a)a+a\delta (a)\) for each a in \({\mathcal {A}}\). A linear map \(\delta \) from \({\mathcal {A}}\) into \({\mathcal {M}}\) is called a derivation if \(\delta (ab)=\delta (a)b+a\delta (b)\) for each ab in \({\mathcal {A}}\). Let m be an element in \({\mathcal {M}},\) the map \(\delta _{m}:{\mathcal {A}}\rightarrow {\mathcal {M}}, ~a\rightarrow \delta _{m}(a):=ma-am,\) is a derivation. A derivation \(\delta :{\mathcal {A}}\rightarrow {\mathcal {M}}\) is said to be an inner derivation when it can be written in the form \(\delta =\delta _{m}\) for some m in \({\mathcal {M}}.\) A fundamental result, due to Sakai [18], states that every derivation on a von Neumann algebra is an inner derivation.

An algebra \({\mathcal {A}}\) is called regular (in the sense of von Neumann) if for each a in \({\mathcal {A}}\) there exists b in \({\mathcal {A}}\) such that \(a=aba.\) Let \({\mathcal {R}}\) be a von Neumann algebra. We denote \(S({\mathcal {R}})\) and \(LS({\mathcal {R}})\), respectively, the algebras of all measurable and locally measurable operators affiliated with \({\mathcal {R}}.\) For a faithful normal semi-finite trace \(\tau \) on \({\mathcal {R}},\) we denote the algebra of all \(\tau \)-measurable operators from \(S({\mathcal {R}})\) by \(S({\mathcal {R}},\tau )\) (cf. [1, 4, 14]). If \({\mathcal {R}}\) is an abelian von Neumann algebra, then it is \(*\)-isomorphic to the algebra \(L^{\infty }(\Omega )=L^{\infty }(\Omega ,\Sigma ,\mu )\) of all (classes of equivalence of) essentially bounded measurable complex functions on a measurable space \((\Omega ,\Sigma ,\mu )\), and therefore, \(LS({\mathcal {R}})=S({\mathcal {R}})\cong L^{0}(\Omega ),\) where \(L^{0}(\Omega )=L^{0}(\Omega ,\Sigma ,\mu )\) is a unital commutative regular algebra of all measurable complex functions on \((\Omega ,\Sigma ,\mu ).\) In this case inner derivations on the algebra \(S({\mathcal {R}})\) are identically zero, i.e., trivial.

Ber et al. [9] obtain necessary and sufficient conditions for existence of non-trivial derivations on commutative regular algebras. In particular, they prove that the algebra \(L^{0}(0,1)\) of all measurable complex functions on the interval (0, 1) admits non-trivial derivations. Let \({\mathcal {R}}\) be a properly infinite von Neumann algebra. Ayupov and Kudaybergenov [4] show that every derivation on the algebra \(LS({\mathcal {R}})\) is an inner derivation.

In 1997, \(\breve{S}\)emrl [17] introduced 2-local derivations and 2-local automorphisms. A map \(\Delta : {\mathcal {A}} \rightarrow {\mathcal {M}}\) (not necessarily linear) is called a 2-local derivation if, for every \(x, y \in {\mathcal {A}}\), there exists a derivation \(D_{x, y}: {\mathcal {A}} \rightarrow {\mathcal {M}}\) such that \(D_{x, y}(x)=\Delta (x)\) and \(D_{x,y}(y) = \Delta (y)\). In particular, if, for every \(x, y \in {\mathcal {A}},\)\(D_{x, y}\) is an inner derivation, then we call \(\Delta \) is a 2-local inner derivation. Niazi and Peralta [15] introduce the notion of weak-2-local derivation (respectively, \(^{*}\)-derivation) and prove that every weak-2-local \(^{*}\)-derivation on \(M_{n}\) is a derivation. 2-local derivations and weak-2-local derivations have been investigated by many authors on different algebras and many results have been obtained in [3,4,5,6,7,8, 11, 13, 15,16,17, 19].

Let \({\mathcal {H}}\) be a infinite-dimensional separable Hilbert space. In [17] \(\breve{S}\)emrl shows that every 2-local derivation on \(\mathcal {B}({\mathcal {H}})\) is a derivation. Kim and Kim [13] give a short proof of that every 2-local derivation on a finite-dimensional complex matrix algebra is a derivation. Ayupov and Kudaybergenov [3] extend this result to an arbitrary von Neumann algebra. Ayupov et al. [5] prove that if \({\mathcal {R}}\) is a finite von Neumann algebra of type \(\text {I}\) without abelian direct summands, then each 2-local derivation on the algebra \(LS({\mathcal {R}})=S({\mathcal {R}})\) is a derivation. In the same paper, the authors also show that if \({\mathcal {R}}\) is an abelian von Neumann algebra such that the lattice of all projections in \({\mathcal {R}}\) is not atomic, then there exists a 2-local derivation on the algebra \(S({\mathcal {R}})\) which is not a derivation. Zhang and Li [19] construct an example of a 2-local derivation on the algebra of all triangular complex \(2\times 2\) matrices which is not a derivation.

Ayupov et al. [5] show that if \({\mathcal {A}}\) is a unital commutative regular algebra, then every 2-local derivation on the algebra \(M_{n}({\mathcal {A}}),\)\(n\ge 2,\) is a derivation. Ayupov and Arzikulov [8] show that if \({\mathcal {A}}\) is a unital commutative ring, then every 2-local inner derivation on \(M_{n}({\mathcal {A}}), ~n\ge 2,\) is an inner derivation. Let \({\mathcal {A}}\) be a unital Banach algebra and \({\mathcal {M}}\) be a unital \({\mathcal {A}}\)-bimodule. He et al. [11] prove that if every Jordan derivation from \({\mathcal {A}}\) into \({\mathcal {M}}\) is an inner derivation then every 2-local derivation from \(M_{n}({\mathcal {A}})\)\((n\ge 3)\) into \(M_{n}({\mathcal {M}})\) is a derivation.

Throughout this paper, \({\mathcal {A}}\) is an algebra with unit 1 over \({\mathbb {C}}\) and \({\mathcal {M}}\) is a unital \({\mathcal {A}}\)-bimodule. We say that \({\mathcal {A}}\)commutes with\({\mathcal {M}}\) if \(am=ma\) for every \(a\in {\mathcal {A}}\) and \(m\in {\mathcal {M}}\). From now on, \(M_{n}({\mathcal {A}})\), for \(n \ge 2,\) will denote the algebra of all \(n\times n\) matrices over \({\mathcal {A}}\) with the usual operations. By the way, we denote any element in \(M_{n}({\mathcal {A}})\) by \((a_{rs})_{n\times n},\) where \(r,s\in \{ 1,2,\ldots ,n \};\)\(E_{ij},\)\(i,j\in \{ 1,2,\ldots ,n \},\) the matrix units in \(M_{n}({\mathbb {C}})\); and \(x\otimes E_{ij},\) the matrix whose (ij)-th entry is x and zero elsewhere. We use \(A_{ij}\) for the (ij)-th entry of \(A\in M_{n}({\mathcal {A}})\) and denote \(\mathrm{diag}(x_{1},\ldots , x_{n})\) or \(\mathrm{diag}(x_{i})\) the diagonal matrix with entries \(x_{i}\in {\mathcal {A}}\), \(i\in {\{1,2,\ldots ,n\}},\) in the diagonal positions. Particularly, we denote \(\mathrm{diag}(x_{i})\) by \(\mathrm{diag}(x)\), where \(x_{i}=x\) for every \(i\in {\{1,2,\ldots ,n\}}.\)

Let \(\delta :{\mathcal {A}}\rightarrow {\mathcal {M}}\) be a derivation. Setting

$$\begin{aligned} \overline{\delta }((a_{ij})_{n\times n})=(\delta (a_{ij}))_{n\times n}, \end{aligned}$$
(1.1)

we obtain a well-defined linear operator from \(M_{n}({\mathcal {A}})\) into \(M_{n}({\mathcal {M}}),\) where \(M_{n}({\mathcal {M}})\) has a natural structure of \(M_{n}({\mathcal {A}})\)-bimodule. Moreover, \(\overline{\delta }\) is a derivation from \(M_{n}({\mathcal {A}})\) into \(M_{n}({\mathcal {M}}).\) If \({\mathcal {A}}\) is a commutative algebra, then the restriction of \(\overline{\delta }\) onto the center of the algebra \(M_{n}({\mathcal {A}})\) coincides with the given \(\delta .\)

In this paper we give characterizations of derivations, 2-local inner derivations and 2-local derivations from \(M_{n}({\mathcal {A}})\) into \(M_{n}({\mathcal {M}})\). In Sect. 2, we show that a derivation \(D: M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}}),\)\(n\ge 2\), can be decomposed as a sum of an inner derivation and a derivation induced by a derivation from \({\mathcal {A}}\) to \({\mathcal {M}}\) as (1.1), as follows:

$$\begin{aligned} D=D_{B}+\overline{\delta }. \end{aligned}$$

In addition, the representation of the above form is unique if and only if \({\mathcal {A}}\) commutes with \({\mathcal {M}}\). Let \({\mathcal {R}}\) be a finite von Neumann algebra of type \(\text {I}\) with center \(\mathcal {Z}\) and \(LS({\mathcal {R}})\) be the algebra of locally measurable operators affiliated with \({\mathcal {R}}.\) we prove that if the lattice \(\mathcal {Z}_{\mathcal {P}}\) of all projections in \(\mathcal {Z}\) is atomic, then every derivation \(D:{\mathcal {R}}\rightarrow LS({\mathcal {R}})\) is an inner derivation.

In Sect. 3, we consider 2-local inner derivations and 2-local derivations from \(M_{n}({\mathcal {A}})\) into \(M_{n}({\mathcal {M}})\). For the case that \({\mathcal {A}}\) commutes with \({\mathcal {M}}\), we obtain that every inner 2-local derivation from \(M_{n}({\mathcal {A}})\) into \(M_{n}({\mathcal {M}})\) is an inner derivation. In addition, if \({\mathcal {A}}\) is commutative, we prove that every 2-local derivation \(\Delta : M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}})\), \(n\ge 2\), is a derivation. Let \({\mathcal {R}}\) be an arbitrary von Neumann algebra without abelian direct summands. We also show every 2-local derivation \(\Delta : {\mathcal {R}}\rightarrow LS({\mathcal {R}})\) is a derivation.

2 Derivations

Let \({\mathcal {A}}\) be an algebra with unit 1 over \({\mathbb {C}}\) and \({\mathcal {M}}\) be a unital \({\mathcal {A}}\)-bimodule. Let \(D: M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}}),\)\(n\ge 2,\) be a derivation. Firstly, we define a map \(D^{ij}_{rs}: {\mathcal {A}}\rightarrow {\mathcal {M}}\) by

$$\begin{aligned} D^{ij}_{rs}(a)=[D(a\otimes E_{rs})]_{ij},\quad a\in {\mathcal {A}},~i,j,r,s\in {\{1,2,\ldots ,n\}}. \end{aligned}$$

For any \(a,b\in {\mathcal {A}}\) and some fixed \(m\in {\{1,2,\ldots ,n\}}, \) we have

$$\begin{aligned} D^{ij}_{rs}(ab)&=[D(ab\otimes E_{rs})]_{ij}\nonumber \\&=[D((a\otimes E_{rm})(b\otimes E_{ms}))]_{ij}\nonumber \\&=[D(a\otimes E_{rm})(b\otimes E_{ms})]_{ij}+[(a\otimes E_{rm})D(b\otimes E_{ms})]_{ij}\nonumber \\&=\delta _{js}[D(a\otimes E_{rm})]_{im}b+\delta _{ir}a[D(b\otimes E_{ms})]_{mj}, \end{aligned}$$

where \(\delta \) is the Kronecker’s delta. It follows that

$$\begin{aligned} D^{ij}_{rs}(ab)=\delta _{js}[D(a\otimes E_{rm})]_{im}b+\delta _{ir}a[D(b\otimes E_{ms})]_{mj}. \end{aligned}$$
(2.1)

For any \(m\in {\{1,2,\ldots ,n\}},\) we deduce from the equality (2.1) that

$$\begin{aligned} D^{mm}_{mm}(ab)=D^{mm}_{mm}(a)b+aD^{mm}_{mm}(b), \end{aligned}$$

thus \(D^{mm}_{mm}:{\mathcal {A}}\rightarrow {\mathcal {M}}\) is a derivation. We abbreviate the derivation \(D^{mm}_{mm}\) by \(D^{m}.\) Particularly, we denote the derivation \(D^{11}_{11}\) by \(D^{1}.\)

Theorem 2.1

Every derivation \(D: M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}}),\)\(n\ge 2,\) can be represented as a sum

$$\begin{aligned} D=D_{B}+\overline{\delta }, \end{aligned}$$
(2.2)

where \(D_{B}\) is an inner derivation implemented by an element \(B\in M_{n}({\mathcal {M}})\) and \(\overline{\delta }\) is a derivation of the form (1.1) induced by a derivation \(\delta \) from \({\mathcal {A}}\) into \({\mathcal {M}}\). Furthermore, if this representation is unique for every derivation D, then \({\mathcal {A}}\) commutes with \({\mathcal {M}}\) (i.e., \(am=ma\) for every \(a\in {\mathcal {A}},~m\in {\mathcal {M}}\)); and if \({\mathcal {A}}\) commutes with \({\mathcal {M}}\) then this representation is always unique.

Before the proof of Theorem 2.1, we first present the following lemma.

Lemma 2.2

For every \(~i,j,r,s,m\in {\{1,2,\ldots ,n\}}\) and every \(a\in {\mathcal {A}}\) the following equalities hold:

  1. (i)

    \(D^{ij}_{rs}=0,\)\(i\ne r\) and \(j\ne s,\)

  2. (ii)

    \(D^{ij}_{rj}(a)=D^{im}_{rm}(a)=D^{im}_{rm}(1)a,\) if \(i\ne r,\)

  3. (iii)

    \(D^{ij}_{is}(a)=D^{mi}_{ms}(a)=aD^{mj}_{ms}(1),\) if \(j\ne s,\)

  4. (iv)

    \(D^{im}_{jm}(1)=-D^{mj}_{mi}(1),\)

  5. (v)

    \(D^{ij}_{ij}(a)=D^{im}_{im}(1)a-aD^{jm}_{jm}(1)+D^{m}(a).\)

Proof

It obviously follows from (2.1) that statements \(\mathrm{(i)}\), \(\mathrm{(ii)}\) and \(\mathrm{(iii)}\) hold. We only need to prove \(\mathrm{(iv)}\) and \(\mathrm{(v)}.\)

\(\mathrm{(iv)}\): In the case \(i=j,\) we have

$$\begin{aligned} 0&=[D(1\otimes E_{ii})]_{ii}=[D((1\otimes E_{im})(1\otimes E_{mi}))]_{ii}\nonumber \\&=[D((1\otimes E_{im}))(1\otimes E_{mi})]_{ii}+[(1\otimes E_{im})D((1\otimes E_{mi}))]_{ii}\nonumber \\&=D^{im}_{im}(1)+D^{mi}_{mi}(1), \end{aligned}$$

i.e.,

$$\begin{aligned} D^{im}_{im}(1)=-D^{mi}_{mi}(1). \end{aligned}$$
(2.3)

For the case \(i\ne j,\) we have

$$\begin{aligned} 0&= D(0)=[D((1\otimes E_{ii})(1\otimes E_{jj}))]_{ij}\nonumber \\&=[D((1\otimes E_{ii}))(1\otimes E_{jj})]_{ij}+[(1\otimes E_{ii})D((1\otimes E_{jj}))]_{ij}\nonumber \\&=[D(1\otimes E_{ii})]_{ij}+[D(1\otimes E_{jj})]_{ij}\nonumber \\&=D^{ij}_{ii}(1)+D^{ij}_{jj}(1), \end{aligned}$$

i.e.,

$$\begin{aligned} D^{ij}_{jj}(1)=-D^{ij}_{ii}(1). \end{aligned}$$

By \(\mathrm{(ii)}\), \(\mathrm{(iii)}\) and equality (2.3), it follows that

$$\begin{aligned} D^{im}_{jm}(1)=-D^{mj}_{mi}(1). \end{aligned}$$

\(\mathrm{(v)}\): By equality (2.1), we have

$$\begin{aligned} D^{ij}_{ij}(a)=D^{im}_{im}(1)a+D^{mj}_{mj}(a), \end{aligned}$$
(2.4)

and

$$\begin{aligned} D^{ij}_{ij}(a)=D^{im}_{im}(a)+aD^{mj}_{mj}(1). \end{aligned}$$
(2.5)

Taking \(j=m\) in equality (2.4), we obtain that

$$\begin{aligned} D^{im}_{im}(a)=D^{im}_{im}(1)a+D^{m}(a). \end{aligned}$$
(2.6)

By equalities (2.3), (2.5) and (2.6), it follows that

$$\begin{aligned} D^{ij}_{ij}(a)=D^{im}_{im}(1)a-aD^{jm}_{jm}(1)+D^{m}(a). \end{aligned}$$

The proof is complete. \(\square \)

Now we are in position to prove Theorem 2.1.

Proof of Theorem 2.1

Let \((a_{rs})_{n\times n}\) be an arbitrary element in \(M_{n}({\mathcal {A}})\) and D be a derivation from \(M_{n}({\mathcal {A}})\) into \(M_{n}({\mathcal {M}})\). For any \(i,j\in {\{1,2,\ldots ,n\}},\) it follows from Lemma 2.2 that

$$\begin{aligned}{}[D((a_{rs})_{n\times n})]_{ij}&=\sum ^{n}_{r,s=1}D^{ij}_{rs}(a_{rs})\nonumber \\&=\sum ^{n}_{r=1}D^{ij}_{rj}(a_{rj})+\sum ^{n}_{s=1}D^{ij}_{is}(a_{is})-D^{ij}_{ij}(a_{ij})\nonumber \\&=\sum _{r\ne i}D^{ij}_{rj}(a_{rj})+\sum _{s\ne j}D^{ij}_{is}(a_{is})+D^{ij}_{ij}(a_{ij})\nonumber \\&=\sum _{r\ne i}D^{i1}_{r1}(1)a_{rj}+\sum _{s\ne j}a_{is}D^{1j}_{1s}(1)+D^{i1}_{i1}(1)a_{ij} \nonumber \\&\quad -\,a_{ij}D^{j1}_{j1}(1)+D^{1}(a_{ij})\nonumber \\&=\sum ^{n}_{r=1}D^{i1}_{r1}(1)a_{rj} -\sum ^{n}_{s=1}a_{is}D^{s1}_{j1}(1)+D^{1}(a_{ij})\nonumber \\&=\sum ^{n}_{k=1}\left( D^{i1}_{k1}(1)a_{kj} -a_{ik}D^{k1}_{j1}(1)\right) +D^{1}(a_{ij})\nonumber \\&=\left[ (D^{r1}_{s1}(1))_{n\times n}(a_{rs})_{n\times n}-(a_{rs})_{n\times n}(D^{r1}_{s1}(1))_{n\times n}\right] _{ij}\nonumber \\&\quad +\,\left[ \overline{D^{1}}((a_{rs})_{n\times n})\right] _{ij}, \end{aligned}$$

i.e.,

$$\begin{aligned}{}[D((a_{rs})_{n\times n})]_{ij}&=\left[ \left( D^{r1}_{s1}(1))_{n\times n}(a_{rs})_{n\times n}-(a_{rs})_{n\times n}(D^{r1}_{s1}(1)\right) _{n\times n}\right] _{ij}\nonumber \\&\quad +\left[ \overline{D^{1}}((a_{rs})_{n\times n})\right] _{ij}, \end{aligned}$$
(2.7)

where \((D^{r1}_{s1}(1))_{n\times n}\in M_{n}({\mathcal {M}})\) and \([(D^{r1}_{s1}(1))_{n\times n}]_{ij}=D^{i1}_{j1}(1).\) By equality (2.7), we have

$$\begin{aligned} D((a_{rs})_{n\times n})=\left[ (D^{r1}_{s1}(1))_{n\times n}(a_{rs})_{n\times n}-(a_{rs})_{n\times n}(D^{r1}_{s1}(1))_{n\times n}\right] +\left[ \overline{D^{1}}((a_{rs})_{n\times n})\right] . \end{aligned}$$

We denote \(B=(D^{r1}_{s1}(1))_{n\times n}\) and \(\delta =D^{1}.\) Therefore, every derivation \(D: M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}}),\)\(n\ge 2,\) can be represented as a sum

$$\begin{aligned} D=D_{B}+\overline{\delta }. \end{aligned}$$

Suppose that \(D_{M}\) is an inner derivation from \(M_{n}({\mathcal {A}})\) into \(M_{n}({\mathcal {M}})\) implemented by an element \(M\in M_{n}({\mathcal {M}}),\) and \(\overline{\zeta }\) is a derivation of the form (1.1) induced by a derivation \(\zeta \) from \({\mathcal {A}}\) into \({\mathcal {M}},\) such that \(D_{M}=\overline{\zeta }.\) The first step is to establish the following. \(\square \)

\(\mathbf Claim~1 \) If \({\mathcal {A}}\) commutes with \({\mathcal {M}}\), then \(D_{M}=\overline{\zeta }=0.\)

Proof of Claim 1

If \(i\ne j,~i,j\in {\{1,2,\ldots ,n\}},\) we have

$$\begin{aligned} 0=\overline{\zeta }(E_{ij})=D_{M}(E_{ij})=ME_{ij}-E_{ij}M. \end{aligned}$$

It follows that \(M_{ji}=0.\) Thus, M has a diagonal form, i.e., \(M=\mathrm{diag}(M_{kk}).\) Suppose that \(\overline{\zeta }\ne 0,\) then there exists an element \(a\in {\mathcal {A}}\) such that \(\zeta (a)\ne 0.\) Take \(A=\mathrm{diag}(a),\) then \( \overline{\zeta }(A)\ne 0.\) On the other hand,

$$\begin{aligned} \overline{\zeta }(A)=D_{M}(A)=\mathrm{diag}(M_{kk})\mathrm{diag}(a)-\mathrm{diag}(a)diag(M_{kk})=0. \end{aligned}$$

This is a contradiction. Thus, \(~\overline{\zeta }=0.\)\(\square \)

\(\mathbf Claim~2 \) If \({\mathcal {A}}\) does not commute with \({\mathcal {M}}\), then there exist \(D_{M}\) and \(\overline{\zeta },\) such that  \(D_{M}=\overline{\zeta }\ne 0.\)

Proof of Claim 2

By assumption, we can take \(a\in {\mathcal {A}}\) and \( m\in {\mathcal {M}}\) such that \(ma\ne am.\) We define a derivation \(\zeta :{\mathcal {A}}\rightarrow {\mathcal {M}}\) by \(\zeta (x)=mx-xm\) for every x in \({\mathcal {A}}.\) We denote \(M=\mathrm{diag}(m)\in M_{n}({\mathcal {M}}),\) then \(D_{M}\) is an inner derivation from \( M_{n}({\mathcal {A}})\) into \( M_{n}({\mathcal {M}}).\) Obviously, \(D_{M}=\overline{\zeta }~\) and \(~\overline{\zeta }(\mathrm{diag}(a))\ne 0.\) Thus, \(D_{M}=\overline{\zeta }\ne 0.\)

In the following, we show that the representation of the above form is unique if and only if \({\mathcal {A}}\) commutes with \({\mathcal {M}}.\)

Case 1 If \({\mathcal {A}}\) commutes with \({\mathcal {M}},\) we suppose that there exists a derivation \(D:M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}}),\)\(n\ge 2,\) which can be represented as \(D=D_{B_{1}}+\overline{\delta _{1}}=D_{B_{2}}+\overline{\delta _{2}}.\) This means that \(D_{B_{1}}-D_{B_{2}}=\overline{\delta _{2}}-\overline{\delta _{1}}.\) Since \(D_{B_{1}}-D_{B_{2}}=D_{B_{1}-B_{2}}\) and \(\overline{\delta _{2}}-\overline{\delta _{1}} =\overline{\delta _{2}-\delta _{1}},\) we have \(D_{B_{1}-B_{2}}=\overline{\delta _{2}-\delta _{1}}.\) It follows from Claim 1 that \(D_{B_{1}-B_{2}}=\overline{\delta _{2}-\delta _{1}}=0.\) i.e., \(D_{B_{1}}=D_{B_{2}}\) and \(\overline{\delta _{1}}=\overline{\delta _{2}}.\)

Case 2 If \({\mathcal {A}}\) does not commute with \({\mathcal {M}},\) by Claim 2, there exist derivations \(D_{M}\) and \(\overline{\zeta }\) from \(M_{n}({\mathcal {A}})\) into \(M_{n}({\mathcal {M}}),\)\(n\ge 2,\) such that \(D_{M}=\overline{\zeta }\ne 0.\) Let \(D:M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}}),\)\(n\ge 2,\) be an arbitrary derivation. By hypothesis, D can be represented as \(D=D_{B}+\overline{\delta }.\) We have \(D=D_{B}+\overline{\delta }=D_{B}+D_{M}- \overline{\zeta }+\overline{\delta }=D_{B+M}+\overline{\delta -\zeta }.\) This means that the derivation D can be represented as \(D=D_{B}+\overline{\delta },\) and as \(D=D_{B+M}+\overline{\delta -\zeta }\) too. Therefore, the representation of (2.2) is not unique for every derivation D. It follows from Cases 1 and 2 that the representation of (2.2) is unique if and only if \({\mathcal {A}}\) commutes with \({\mathcal {M}}\). The proof is complete. \(\square \)

As applications of Theorem 2.1, we obtain the following corollaries.

Corollary 2.3

The following statements are equivalent.

  1. (i)

    Every derivation \(\delta :{\mathcal {A}}\rightarrow {\mathcal {M}}\) is an inner derivation.

  2. (ii)

    Every derivation \(D:M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}}),\)\(n\ge 2,\) is an inner derivation.

Proof

If \(\delta :{\mathcal {A}}\rightarrow {\mathcal {M}}\) is an inner derivation, by the equality (1.1), obviously, \(\overline{\delta }:M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}}),\)\(n\ge 2,\) is an inner derivation.

\(\mathrm{(i)}\) implies \(\mathrm{(ii)}\): Let \(D:M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}}),\)\(n\ge 2,\) be an arbitrary derivation. By Theorem 2.1, D can be represented as a sum \(D=D_{M}+\overline{\delta },\) where \(D_{M}\) is an inner derivation. By hypothesis, \(\delta \) is an inner derivation from \({\mathcal {A}}\) into \({\mathcal {M}},\) and therefore, \(\overline{\delta }\) is an inner derivation. We know that the sum of two inner derivations is an inner derivation, this means that \(D:M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}}),\)\(n\ge 2,\) is an inner derivation.

\(\mathrm{(ii)}\) implies \(\mathrm{(i)}\): Suppose that \(\delta \) is a derivation from \({\mathcal {A}}\) into \({\mathcal {M}},\) then \(\overline{\delta }:M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}}),\)\(n\ge 2,\) is a derivation. By hypothesis, \(\overline{\delta }\) is an inner derivation. then the restriction of \(\overline{\delta }\) onto \(E_{11}M_{n}({\mathcal {A}})E_{11},\) the subalgebra of \(M_{n}({\mathcal {A}}),\) is an inner derivation. This means that \(\delta :{\mathcal {A}}\rightarrow {\mathcal {M}}\) is an inner derivation. \(\square \)

Corollary 2.4

Let \({\mathcal {A}}\) be a commutative unital algebra over \({\mathbb {C}}.\) Then every derivation on the matrix algebra \(M_{n}({\mathcal {A}})~(n\ge 2)\) is inner if and only if every derivation on \({\mathcal {A}}\) is identically zero, i.e., trivial.

Let \({\mathcal {R}}\) be a von Neumann algebra. Denote by \(S({\mathcal {R}})\) and \(LS({\mathcal {R}})\), respectively, the sets of all measurable and locally measurable operators affiliated with \({\mathcal {R}}.\) Then the set \(LS({\mathcal {R}})\) of all locally measurable operators with respect to \({\mathcal {R}}\) is a unital \(*\)-algebra when equipped with the algebraic operations of strong addition and multiplication and taking the adjoint of an operator and \(S({\mathcal {R}})\) is a solid \(*\)-subalgebra in \(LS({\mathcal {R}})\). If \({\mathcal {R}}\) is a finite von Neumann algebra, then \(S({\mathcal {R}})=LS({\mathcal {R}})\) (see, for example, [1, 4, 14]). Let \({\mathcal {A}}\) be a commutative algebra with unit 1 over \({\mathbb {C}}.\) We denote by \(\nabla \) the set \(\{e\in {\mathcal {A}}:e^{2}=e\}\) of all idempotents in \({\mathcal {A}}.\) For \(e,f\in \nabla \) we set \(e\le f\) if \(ef=e.\) Equipped with this partial order, lattice operations \(e\vee f=e+f-ef,\)\(e\wedge f=ef\) and the complement \(e^{\perp }=1-e,\) the set \(\nabla \) forms a Boolean algebra. A nonzero element q from the Boolean algebra \(\nabla \) is called an atom if \(0\ne e\le q,\)\(e\in \nabla ,\) imply that \(e=q.\) If given any nonzero \(e\in \nabla \) there exists an atom q such that \(q\le e,\) then the Boolean algebra \(\nabla \) is said to be atomic.

Let \({\mathcal {R}}\) be an abelian von Neumann algebra. Theorem 3.4 of [9] implies that every derivation on the algebra \(S({\mathcal {R}})\) is inner if and only if the lattice \({\mathcal {R}}_{\mathcal {P}}\) of all projections in \({\mathcal {R}}\) is atomic. If \({\mathcal {R}}\) is a properly infinite von Neumann algebra, in [4] the authors show that every derivation on the algebra \(LS({\mathcal {R}})\) is inner (see [4], Theorem 4.6). In the case of \({\mathcal {R}}\) is a finite von Neumann algebra of type \(\text {I}\), Theorem 3.5 of [4] shows that a derivation on the algebra \(LS({\mathcal {R}})\) is an inner derivation if and only if it is identically zero on the center of \({\mathcal {R}}.\)

As a direct application of Corollary 2.3, we obtain the following result.

Corollary 2.5

Let \({\mathcal {R}}\) be a finite von Neumann algebra of type \(\text {I}\) with center \(\mathcal {Z}.\) Then every derivation D on the algebra \(LS({\mathcal {R}})\) is inner if and only if the lattice \(\mathcal {Z}_{\mathcal {P}}\) of all projections in \(\mathcal {Z}\) is atomic.

Proof

Let \({\mathcal {R}}\) be a finite von Neumann algebra of type \(\text {I}\) with center \(\mathcal {Z}.\) There exists a family \(\{e_{n}\}_{n\in \mathcal {F}},\)\(\mathcal {F}\subseteq \mathbb {N},\) of central projections from \({\mathcal {R}}\) with \(\bigvee \nolimits _{n\in \mathcal {F}}e_{n}=1\) such that the algebra \({\mathcal {R}}\) is \(*\)-isomorphic with the \(C^{*}\)-product of von Neumann algebras \(e_{n}{\mathcal {R}}\) of type \(\text {I}_{n}\), respectively, \(n\in \mathcal {F},\) i.e., \({\mathcal {R}} \cong \bigoplus \nolimits _{n\in \mathcal {F}}e_{n}{\mathcal {R}}.\) By Proposition 1.1 of [1], we have that \(LS({\mathcal {R}}) \cong \prod \nolimits _{n\in \mathcal {F}}LS(e_{n}{\mathcal {R}}).\)

Suppose that D is a derivation on \(LS({\mathcal {R}})\) and \(\delta \) its restriction onto the center \(S(\mathcal {Z}).\) Since \(\delta \) maps each \(e_{n}S(\mathcal {Z})\) into itself, \(\delta \) generates a derivation \(\delta _{n}\) on \(e_{n}S(\mathcal {Z})\) for each \(n\in \mathcal {F}.\) By Proposition 1.5 of [1], \(LS(e_{n}{\mathcal {R}})\cong M_{n}(e_{n}S(\mathcal {Z}))\). Let \(\overline{\delta }_{n}\) be the derivation on the matrix algebra \(M_{n}(e_{n}S(\mathcal {Z}))\) defined as in (1.1). Put

$$\begin{aligned} \overline{\delta }(\{x_{n}\}_{n\in \mathcal {F}}) =\{\overline{\delta }_{n}(x_{n})\},\quad \{x_{n}\}_{n\in \mathcal {F}}\in LS({\mathcal {R}}). \end{aligned}$$
(2.8)

Then the map \(\overline{\delta }\) is a derivation on \(LS({\mathcal {R}}).\) Lemma 2.3 of [1] implies that each derivation D on \(LS({\mathcal {R}})\) can be uniquely represented in the form \(D=D_{B}+\overline{\delta },\) where \(D_{B}\) is an inner derivation and \(\overline{\delta }\) is a derivation given as (2.8).

If D is an arbitrary derivation on \(LS({\mathcal {R}})\) and \(\delta \) its restriction onto center \(S(\mathcal {Z}),\) by Theorem 3.4 of [9], the lattice \(\mathcal {Z}_{\mathcal {P}}\) is atomic if and only if \(\delta =0.\) We have \(\delta =0\) if and only if \(\delta _{n}=0\) for each \(n\in \mathcal {F}.\) By Corollary 2.3, \(\delta _{n}=0\) if and only if \(\overline{\delta _{n}}=0\) for each \(n\in \mathcal {F}.\) By equality (2.8), \(\overline{\delta _{n}}=0\) for each \(n\in \mathcal {F}\) if and only if \(\overline{\delta }=0.\) Therefore, every derivation on the algebra \(LS({\mathcal {R}})\) is inner derivation if and only if the lattice \(\mathcal {Z}_{\mathcal {P}}\) of all projections in \(\mathcal {Z}\) is atomic. The proof is complete. \(\square \)

Let \({\mathcal {R}}\) be a properly infinite von Neumann algebra and \({\mathcal {M}}\) be a \({\mathcal {R}}\)-bimodule of locally measurable operators. In [10], the authors show that every derivation \(D:{\mathcal {R}}\rightarrow {\mathcal {M}}\) is an inner derivation. In the case of \({\mathcal {R}}\) is a finite von Neumann algebra of type \(\text {I}\), we obtain the following result.

Theorem 2.6

Let \({\mathcal {R}}\) be a finite von Neumann algebra of type \(\text {I}\) with center \(\mathcal {Z}.\) If the lattice \(\mathcal {Z}_{\mathcal {P}}\) of all projections in \(\mathcal {Z}\) is atomic, then every derivation \(D:{\mathcal {R}}\rightarrow LS({\mathcal {R}})\) is an inner derivation.

Proof

Choose a central decomposition \(\{e_{n}\}_{n\in \mathcal {F}},\)\(\mathcal {F}\subseteq \mathbb {N},\) of the unity 1 such that \(e_{n}{\mathcal {R}}\) is a type \(\text {I}_{n}\) von Neumann algebra for each \(n\in \mathcal {F}.\) By hypothesis, it is easy to check that \(D(e_{n}{\mathcal {R}})\subseteq e_{n}LS({\mathcal {R}})\) for each \(n\in \mathcal {F}.\) Thus, we only need to show that the derivation D restricted to \(e_{n}{\mathcal {R}}\) is an inner derivation for each \(n\in \mathcal {F}.\)

Let \(e_{n}{\mathcal {R}}\) be a type \(\text {I}_{n} ~(n\in \mathcal {F})\) von Neumann algebra with center \(e_{n}\mathcal {Z}.\) It is well known that \(e_{n}{\mathcal {R}}\cong M_{n}(e_{n}\mathcal {Z}).\) We denote the center of \(S(e_{n}{\mathcal {R}})\) by \(\mathcal {Z}(S(e_{n}{\mathcal {R}}))\). By Proposition 1.2 of [1], we have \(\mathcal {Z}(S(e_{n}{\mathcal {R}}))=S(e_{n}\mathcal {Z}).\) By Proposition 1.5 of [1], \(LS(e_{n}{\mathcal {R}})=S(e_{n}{\mathcal {R}})\cong M_{n}(S(e_{n}\mathcal {Z})).\)

By assumption, the lattice \(\mathcal {Z}_{\mathcal {P}}\) of all projections in \(\mathcal {Z}\) is atomic. This implies that the lattice \(e_{n} \mathcal {Z}_{\mathcal {P}}\) is also atomic for each \(n\in \mathcal {F}.\) Statements \(\mathrm{(ii)}\) of Proposition 2.3 and \(\mathrm{(vi)}\) of Proposition 2.6 of [9] imply that every derivation \(\delta :e_{n}\mathcal {Z}\rightarrow S(e_{n}\mathcal {Z})\) is trivial. By Corollary 2.3, we have that every derivation from \(M_{n}(e_{n}\mathcal {Z})\) into \(M_{n}(S(e_{n}\mathcal {Z}))\) is inner. The proof is complete. \(\square \)

3 2-Local Derivations

This section is devoted to 2-local inner derivations and 2-local derivations from \(M_{n}({\mathcal {A}})\) into \(M_{n}({\mathcal {M}})\). Throughout this section, we always assume that \(\Delta : M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}})\) is a 2-local derivation. Firstly, we give the following lemma.

Lemma 3.1

For every 2-local derivation \(\Delta : M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}})\), \(n\ge 2\), there exists a derivation \(D: M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}})\) such that \(\Delta (E_{ij})=D(E_{ij})\) for all \(i,j\in {\{1,2,\ldots ,n\}}.\) In particular, if \(\Delta \) is a 2-local inner derivation, then D is an inner derivation.

Proof

Let \(\Delta : M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}})\), \(n\ge 2\), be a 2-local derivation. By Theorem 2.1, with the proof similar to the proof of Theorem 3 in [13], it is easy to check that there exists a derivation D such that \(\Delta (E_{ij})=D(E_{ij})\) for all \(i,j\in {\{1,2,\ldots ,n\}}.\)

Let \(\Delta \) be an inner 2-local derivation. We define two matrices S, T in \(M_{n}({\mathcal {A}})\) by

$$\begin{aligned} S=\sum ^{n}_{i=1}i1\otimes E_{ii},\quad T=\sum ^{n-1}_{i=1}E_{ii+1}. \end{aligned}$$

By assumption, there exists an inner derivation \(D: M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}})\) such that

$$\begin{aligned} \Delta (S)=D(S),\quad \Delta (T)=D(T). \end{aligned}$$

Replacing \(\Delta \) by \(\Delta -D\) if necessary, we may assume that \(\Delta (S)=\Delta (T)=0.\) Fixed \(i,j\in {\{1,2,\ldots ,n\}},\) by assumption, we can take two elements X,  Y in \(M_{n}({\mathcal {M}})\) such that

$$\begin{aligned} \Delta (E_{ij})=XE_{ij}-E_{ij}X,\quad 0=\Delta (S)=XS-SX, \end{aligned}$$

and

$$\begin{aligned} \Delta (E_{ij})=YE_{ij}-E_{ij}Y,\quad 0=\Delta (T)=YT-TY. \end{aligned}$$

It follows from \(XS=SX\) that X is a diagonal matrix. We denote X by \(\mathrm{diag}(x_{k}).\) The equality \(YT=TY\) implies that Y is of the form

$$\begin{aligned} Y=\left[ \begin{array}{llllll} \ y_{1}&{}\quad y_{2}&{}\quad y_{3}&{}\quad \cdot &{}\quad \cdot &{}\quad y_{n}\\ \ 0&{}\quad y_{1}&{}\quad y_{2}&{}\quad \cdot &{}\quad \cdot &{}\quad y_{n-1}\\ \ 0&{}\quad 0&{}\quad y_{1}&{}\quad \cdot &{}\quad \cdot &{}\quad y_{n-2}\\ \ \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots \\ \ \cdots &{}\quad \cdots &{}\quad \cdots &{}\quad \cdot &{}\quad y_{1} &{}\quad y_{2}\\ \ 0 &{}\quad 0 &{}\quad \cdots &{}\quad \cdot &{}\quad 0 &{}\quad y_{1}\\ \end{array} \right] . \end{aligned}$$

On the one side, we have

$$\begin{aligned} \Delta (E_{ij})=XE_{ij}-E_{ij}X=\mathrm{diag}(x_{k})E_{ij} -E_{ij}\mathrm{diag}(x_{k})=(x_{i}-x_{j})\otimes E_{ij}. \end{aligned}$$

On the other side, we have

$$\begin{aligned}{}[\Delta (E_{ij})]_{ij}=[YE_{ij}-E_{ij}Y]_{ij}=0. \end{aligned}$$

Therefore, \(\Delta (E_{ij})=0.\) The proof is complete. \(\square \)

Theorem 3.2

Suppose that \({\mathcal {A}}\) commutes with \({\mathcal {M}}.\) Then every 2-local inner derivation \(\Delta : M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}})\), \(n\ge 2\), is an inner derivation.

Proof

By Lemma 3.1, we may assume that \(\Delta (E_{ij})=0\) for all \(i,j\in {\{1,2,\ldots ,n\}}.\) For any \(A\in M_{n}({\mathcal {A}}),\) we take a pair (ji), \(j,i\in {\{1,2,\ldots ,n\}},\) by assumption, there exists an inner derivation \(D_{B}\), such that \(\Delta (A)=D_{B}(A)\) and \(0=\Delta (E_{ij})=D_{B}(E_{ij})\). We have

$$\begin{aligned} E_{ij}\Delta (A)E_{ij}&=E_{ij}D_{B}(A)E_{ij}\nonumber \\&=D_{B}(E_{ij}AE_{ij})-D_{B}(E_{ij})AE_{ij} -E_{ij}AD_{B}(E_{ij})=D_{B}(E_{ij}AE_{ij})\nonumber \\&=D_{B}(A_{ji}\otimes E_{ij})=D_{B}(\mathrm{diag}(A_{ji}, \ldots ,A_{ji})E_{ij})\nonumber \\&=D_{B}(\mathrm{diag}(A_{ji},\ldots ,A_{ji}))E_{ij} +\mathrm{diag}(A_{ji},\ldots ,A_{ji})D_{B}(E_{ij})\nonumber \\&=(B\mathrm{diag}(A_{ji},\ldots ,A_{ji})-\mathrm{diag}(A_{ji}, \ldots ,A_{ji})B)E_{ij}\nonumber \\&=0, \end{aligned}$$

i.e.,

$$\begin{aligned} E_{ij}\Delta (A)E_{ij}=0. \end{aligned}$$

Therefore,

$$\begin{aligned} E_{ji}(E_{ij}\Delta (A)E_{ij})E_{ji}=E_{jj}\Delta (A)E_{ii}=0, \end{aligned}$$

i.e.,

$$\begin{aligned}{}[\Delta (A)]_{ji}=0, \end{aligned}$$

for every \(j,i\in {\{1,2,\ldots ,n\}}.\) Hence \(\Delta (A)=0.\) The proof is complete. \(\square \)

Corollary 3.3

Suppose that \({\mathcal {A}}\) is a unital commutative algebra over \({\mathbb {C}}.\) Then every 2-local inner derivation \(\Delta : M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {A}})\), \(n\ge 2\), is an inner derivation.

Remark 3.4

The above result is proved in [8]. By comparison, our proof is more simple.

Suppose that \({\mathcal {A}}\) is an algebra over \({\mathbb {C}}\) and \(\mathcal {B}\) is a unital subalgebra in \({\mathcal {A}}.\) We denote the commutant of \(\mathcal {B}\) by \(\mathcal {B}^{\prime }=\{a\in {\mathcal {A}}:ab=ba, {for~every~}b\in \mathcal {B}\}\). Let \(\mathcal {C}\) be a submodule in \(\mathcal {B}^{\prime }.\) It follows from Theorem 3.2 that

Corollary 3.5

Every 2-local inner derivation \(\Delta : M_{n}(\mathcal {B})\rightarrow M_{n}(\mathcal {C})\), \(n\ge 2\), is an inner derivation.

Theorem 3.6

Suppose that \({\mathcal {A}}\) is a commutative algebra which commutes with \({\mathcal {M}}\). Then every 2-local derivation \(\Delta : M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}})\), \(n\ge 2\), is a derivation.

Proof

The proof is similar to the proof of Theorem 4.3 in [5]. We leave it to the reader. \(\square \)

Corollary 3.7

Suppose that \({\mathcal {A}}\) is a unital commutative algebra over \({\mathbb {C}}.\) Then every 2-local derivation \(\Delta : M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {A}})\), \(n\ge 2\), is a derivation.

If \({\mathcal {A}}\) is a non-commutative algebra, by Theorem 2.1 every derivation from \(M_{n}({\mathcal {A}})\) into \(M_{n}({\mathcal {M}})(n\ge 2)\) can be represented as a sum \(D=D_{B}+\overline{\delta }.\) In [7], the authors apply this representation of derivation to prove the following result.

Theorem 3.8

([7], Theorem 2.1) Let \({\mathcal {A}}\) be a unital Banach algebra and \({\mathcal {M}}\) be a unital \({\mathcal {A}}\)-bimodule. If every Jordan derivation from \({\mathcal {A}}\) into \({\mathcal {M}}\) is a derivation, then every 2-local derivation \(\Delta : M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {A}})\), \(n\ge 3\), is a derivation.

Theorem 3.9

Let \({\mathcal {A}}\) be a unital Banach algebra and \({\mathcal {M}}\) be a unital \({\mathcal {A}}\)-bimodule. If \(n\ge 6\) is a positive integer but not a prime number, then every 2-local derivation \(\Delta : M_{n}({\mathcal {A}})\rightarrow M_{n}({\mathcal {M}})\) is a derivation.

Proof

Suppose that \(n=rk\), where \(r\ge 3\) and \(k\ge 2.\) Then \(M_{n}({\mathcal {A}})\cong M_{r}(M_{k}({\mathcal {A}}))\) and \(M_{n}({\mathcal {M}})\cong M_{r}(M_{k}({\mathcal {M}})).\) In [2], the author proves that every Jordan derivation from \(M_{k}({\mathcal {A}})\) into \(M_{k}({\mathcal {M}})(k\ge 2)\) is a derivation ([2], Theorem 3.1). By Theorem 3.8, the proof is complete. \(\square \)

Let \({\mathcal {R}}\) be a type \(\text {I}_{n}~ (n\ge 2)\) von Neumann algebra with center \(\mathcal {Z}\) and \(\tau \) be a faithful normal semi-finite trace on \({\mathcal {R}}.\) We denote the centers of \(S({\mathcal {R}})\) and \(S({\mathcal {R}},\tau )\) by \(\mathcal {Z}(S({\mathcal {R}}))\) and \(\mathcal {Z}(S({\mathcal {R}},\tau ))\), respectively. By Proposition 1.2 of [1], we have \(\mathcal {Z}(S({\mathcal {R}}))=S(\mathcal {Z})\) and \(\mathcal {Z}(S({\mathcal {R}},\tau ))=S(\mathcal {Z},\tau _{\mathcal {Z}}),\) where \(\tau _{\mathcal {Z}}\) is the restriction of the trace \(\tau \) on \(\mathcal {Z}.\) By Propositions 1.4 and 1.5 of [1], \(S({\mathcal {R}})=LS({\mathcal {R}})\cong M_{n}(S(\mathcal {Z}))\) and \(S({\mathcal {R}},\tau )\cong M_{n}(S(\mathcal {Z},\tau _{\mathcal {Z}})).\)

As a direct application of Theorem 3.6, we have the following corollary.

Corollary 3.10

Suppose that \({\mathcal {R}}\) is a type \(\text {I}_{n}, n\ge 2,\) von Neumann algebra and \(\tau \) is a faithful normal semi-finite trace on \({\mathcal {R}}.\) Then we have

  1. (1)

      every 2-local derivation \(\Delta : {\mathcal {R}}\rightarrow LS({\mathcal {R}})\) is a derivation;

  2. (2)

      every 2-local derivation \(\Delta : {\mathcal {R}}\rightarrow S({\mathcal {R}},\tau )\) is a derivation.

Lemma 3.11

Let \(\Delta :{\mathcal {A}}\rightarrow {\mathcal {M}}\) be a 2-local derivation. If there exists a central idempotent e in \({\mathcal {A}}\) which commutates with \({\mathcal {M}},\) then \(\Delta (ea)=e\Delta (a),\) for each a in \({\mathcal {A}}.\)

Proof

For any \(a \in {\mathcal {A}},\) by assumption, there exists a derivation \(\delta :{\mathcal {A}}\rightarrow {\mathcal {M}}\) such that: \(\Delta (ea)=\delta (ea),\) and \(\Delta (a)=\delta (a).\) By assumption, e is a central idempotent in \({\mathcal {A}}\) which commutes with \({\mathcal {M}},\) it follows that \(\delta (e)=0.\) Then

$$\begin{aligned} \Delta (ea)=\delta (ea)=\delta (e)a+e\delta (a)=e\delta (a)=e\Delta (a). \end{aligned}$$

The proof is complete. \(\square \)

Theorem 3.12

Suppose that \({\mathcal {R}}\) is a finite von Neumann algebra of type \(\text {I}\) without abelian direct summands. Then every 2-local derivation \(\Delta : {\mathcal {R}}\rightarrow S({\mathcal {R}})= LS({\mathcal {R}})\) is a derivation.

Proof

By assumption, \({\mathcal {R}}\) is a finite von Neumann algebra of type \(\text {I}\) without abelian direct summands. Then there exists a family \(\{P_{n}\}_{n\in F},~F\subseteq \mathbb {N}{\setminus }{1},\) of orthogonal central projections in \({\mathcal {R}}\) with \(\sum _{n\in F} P_{n}=1,\) such that the algebra \({\mathcal {R}}\) is \(*\)-isomorphic with the \(C^{*}\)-product of von Neumann algebras \(P_{n}{\mathcal {R}}\) of type \(\text {I}_{n}\), respectively, \(n\in F.\) Then

$$\begin{aligned} P_{n}LS({\mathcal {R}})=P_{n}S({\mathcal {R}})=S(P_{n}{\mathcal {R}})\cong M_{n}(P_{n}Z({\mathcal {R}})),\quad n\in F. \end{aligned}$$

By Lemma 3.11, we have \(\Delta (P_{n}A)=P_{n}\Delta (A),\) for all \(A\in {\mathcal {R}}\) and each \(n\in F.\) This implies that \(\Delta \) maps each \(P_{n}{\mathcal {R}}\) into \(P_{n}S({\mathcal {R}}).\) For each \(n\in F,\) we define \(\Delta _{n}:P_{n}{\mathcal {R}}\rightarrow P_{n}S({\mathcal {R}})\) by

$$\begin{aligned} \Delta _{n}(P_{n}A)=P_{n}\Delta (A),\quad A\in {\mathcal {R}}. \end{aligned}$$

By assumption, it follows that \(\Delta _{n}\) is a 2-local derivation from \(P_{n}{\mathcal {R}}\) into \(P_{n}S({\mathcal {R}})\) for each \(n\in F.\) By (1) of Corollary 3.10, we have that \(\Delta _{n}\) is a derivation for each \(n\in F.\) Since \(\sum _{n\in F}P_{n}=1,\) it follows that \(\Delta \) is a linear mapping. For any \(A,B\in {\mathcal {R}},\) it follows \(\Delta _{n}\) is a derivation for each \(n\in F\) that

$$\begin{aligned} P_{n}\Delta (AB)&=\Delta _{n}(P_{n}AB)= \Delta _{n}(P_{n}A)P_{n}B+P_{n}A\Delta _{n}(P_{n}B)\nonumber \\&=P_{n}\Delta (A)B+P_{n}A\Delta (B)\nonumber \\&=P_{n}(\Delta (A)B+A\Delta (B)). \end{aligned}$$

By assumption, \(\sum _{n\in F}P_{n}=1,\) we get

$$\begin{aligned} \Delta (AB)=\Delta (A)B+A\Delta (B). \end{aligned}$$

Therefore, \(\Delta : {\mathcal {R}}\rightarrow S({\mathcal {R}})\) is a derivation. The proof is complete. \(\square \)

Ayupov et al. [7] have proved the following result. Now we give a different proof.

Theorem 3.13

([7], Theorem 3.1) Let \({\mathcal {R}}\) be an arbitrary von Neumann algebra without abelian direct summands and \(LS({\mathcal {R}})\) be the algebra of all locally measurable operators affiliated with \({\mathcal {R}}.\) Then every 2-local derivation \(\Delta : {\mathcal {R}}\rightarrow LS({\mathcal {R}})\) is a derivation.

Proof

Let \({\mathcal {R}}\) be an arbitrary von Neumann algebra without abelian direct summands. We know that \({\mathcal {R}}\) can be decomposed along a central projection into the direct sum of von Neumann algebras of finite type \(\text {I}\), type \(\text {I}_{\infty },\) type \(\text {II}\) and type \(\text {III}.\) By Lemma 3.11, we may consider these cases separately.

If \({\mathcal {R}}\) is a von Neumann algebra of finite type \(\text {I}\), Theorem 3.12 shows that every 2-local derivation from \({\mathcal {R}}\) into LS\(({\mathcal {R}})\) is a derivation.

If \({\mathcal {R}}\) is a von Neumann algebra of types \(\text {I}_{\infty },\)\(\text {II}\) or \(\text {III},\) then the halving Lemma ([12], Lemma 6.3.3) for type \(\text {I}_{\infty }\) algebras and ([12], Lemma 6.5.6) for types \(\text {II}\) or \(\text {III}\) algebras implies that the unit of \({\mathcal {R}}\) can be represented as a sum of mutually equivalent orthogonal projections \(e_{1},e_{2},\ldots ,e_{6}\) in \({\mathcal {R}}.\) It is well known that \({\mathcal {R}}\) is isomorphic to \(M_{6}({\mathcal {A}}),\) where \({\mathcal {A}}=e_{1}{\mathcal {R}}e_{1}.\) Further, the algebra \(LS({\mathcal {R}})\) is isomorphic to the algebra \(M_{6}(LS({\mathcal {A}})).\) Theorem 3.9 implies that every 2-local derivation from \({\mathcal {R}}\) into LS\(({\mathcal {R}})\) is a derivation. The proof is complete. \(\square \)