1 Introduction

Let \(\mathcal {A}\) be an associative algebra over the field \(\mathbb F\) (\(\mathbb R\) or \(\mathbb C)\) and \(\mathcal M\) be an \(\mathcal A\)-bimodule. A linear mapping \(\delta \) from \(\mathcal A\) into \(\mathcal M\) is called a derivation if \(\delta (AB)=\delta (A)B+A\delta (B)\) for each AB in \(\mathcal A\), and \(\delta \) is called an inner derivation if there exists an element M in \(\mathcal M\) such that \(\delta (A)=AM-MA\) for every A in \(\mathcal A\). Clearly, every inner derivation is a derivation. In [13, 24], Kadison and Sakai independently proved that every derivation on a von Neumann algebra is inner. In [6], Chernoff proved that every derivation from a standard operator algebra \(\mathcal A\) into B(X) is inner for a Banach space X. In [8], Christensen showed that every derivation on nest algebras is inner.

In 1990, Kadison [14], Larson and Sourour [15] independently introduced the concept of local derivations. A linear mapping \(\delta \) from \(\mathcal {A}\) into \(\mathcal {M}\) is called a local derivation if for every A in \(\mathcal {A}\) there exists a derivation \(\delta _A\) (depending on A) from \(\mathcal {A}\) into \(\mathcal {M}\) such that \(\delta (A)=\delta _A(A)\). In [14], Kadison proved that every continuous local derivation from a von Neumann algebra into its dual Banach module is a derivation. In [15], Larson and Sourour proved that every local derivation on B(X) is a derivation for a Banach space X. In [12], Johnson proved that every local derivation from a \(C^*\)-algebra into its Banach bimodule is a derivation. In [29], Zhu and Xiong proved that every local derivation from a unital standard operator algebra \(\mathcal A\) into B(X) is a derivation.

In 1997, \(\check{\text {S}}\)emrl [25] introduced the concept of 2-local derivations. A mapping (not necessarily linear) \(\delta \) from \(\mathcal {A}\) into \(\mathcal {M}\) is called a 2-local derivation if for each AB in \(\mathcal {A}\), there exists a derivation \(\delta _{A,B}\) (depending on AB) from \(\mathcal {A}\) into \(\mathcal {M}\) such that \(\delta (A)=\delta _{A,B}(A)\) and \(\delta (B)=\delta _{A,B}(B)\). In [25], \(\check{\text {S}}\)emrl proved that every 2-local derivation on \(B(\mathcal H)\) is a derivation for a separable Hilbert space H. In [2], Ayupov and Kudaybergenov proved that every 2-local derivation on a von Neumann algebra is a derivation. In [10], we showed that every 2-local derivation from a standard operator algebra \(\mathcal A\) into B(X) is a derivation.

A linear mapping \(\delta \) from \(\mathcal A\) into \(\mathcal M\) is called a Lie derivation if \(\delta ([A,B])=[\delta (A),B]+[A,\delta (B)]\) for each AB in \(\mathcal A\), where \([A,B]=AB-BA\) is called a commutator on \(\mathcal A\). A Lie derivation \(\delta \) is said to be standard if it can be decomposed as \(\delta =d+\tau \), where d is a derivation from \(\mathcal {A}\) into \(\mathcal M\) and \(\tau \) is a linear mapping from \(\mathcal {A}\) into \(\mathcal {Z}(\mathcal {M},\mathcal A)\) with \(\tau ([A,B])=0\) for each AB in \(\mathcal {A}\), where \(\mathcal {Z}(\mathcal {M},\mathcal A)=\{M\in \mathcal {M}:MA=AM~\text{ for } \text{ every }~A~\text{ in }~\mathcal {A}\}\).

An interesting problem is to identify those algebras on which every Lie derivation is standard. In [22], Mathieu and Villena proved that every Lie derivation on a \(C^*\)-algebra is standard. In [7], Cheung characterized Lie derivations on triangular algebras. In [20, 21], Lu studied Lie derivations on CDCSL algebras and reflexive algebras, respectively. In [3], Benkovi\({\check{\text {c}}}\) proved that every Lie derivation on a matrix algebra \(M_n(\mathcal A)\) is standard, where \(n\ge 2\) and \(\mathcal A\) is a unital algebra.

Similarly to local derivations and 2-local derivations, in [4], Chen et al. introduced the concepts of local Lie derivations and 2-local Lie derivations. A linear mapping \(\delta \) from \(\mathcal {A}\) into \(\mathcal {M}\) is called a local Lie derivation if for every A in \(\mathcal {A}\) there exists a Lie derivation \(\delta _{A}\) (depending on A) from \(\mathcal {A}\) into \(\mathcal {M}\) such that \(\delta (A)=\delta _{A}(A)\). A mapping (not necessarily linear) \(\delta \) from \(\mathcal {A}\) into \(\mathcal {M}\) is called a 2-local Lie derivation if for every AB in \(\mathcal {A}\) there exists a Lie derivation \(\delta _{A,B}\) (depending on AB) from \(\mathcal {A}\) into \(\mathcal {M}\) such that \(\delta (A)=\delta _{A,B}(A)\) and \(\delta (B)=\delta _{A,B}(B)\).

In [4], Chen et al. study local Lie derivations and 2-local Lie derivations on B(X). In [5], Chen and Lu proved that every local Lie derivation on nest algebras is a Lie derivation. In [18, 19], Liu and Zhang proved that under certain conditions every local Lie derivation on triangular algebras is a Lie derivation, and every local Lie derivation on factor von Neumann algebras with dimension exceeding 1 is a Lie derivation. In [9], He et al. proved that every local Lie derivation on some algebras such as finite von Neumann algebras, nest algebras, Jiang–Su algebras and UHF algebras is a Lie derivation, and every 2-local Lie derivation on on some algebras such as factor von Neumann algebras, Jiang–Su algebra and UHF algebras is also a Lie derivation. In [16, 17], Liu proved that under certain conditions every local Lie derivation on generalized matrix algebras is a Lie derivation, and he showed that every 2-local Lie derivation of nest subalgebras of factors is a Lie derivation.

A linear mapping \(\delta \) from \(\mathcal A\) into \(\mathcal M\) is a Lie triple derivation if \(\delta ([[A,B],C])=[[\delta (A),B],C]+[[A,\delta (B)],C]+[[A,B],\delta (C)]\) for each A, B and C in \(\mathcal A\). We call [[AB], C] a double commutator on \(\mathcal A\). It is clear that every Lie derivation is a Lie triple derivation. A Lie triple derivation \(\delta \) from \(\mathcal A\) into \(\mathcal M\) is said to be standard if it can be decomposed as \(\delta =d+\tau \), where d is a derivation from \(\mathcal {A}\) into \(\mathcal M\) and \(\tau \) is a linear mapping from \(\mathcal {A}\) into \(\mathcal {Z}(\mathcal {M},\mathcal {A})\) with \(\tau ([[A,B],C])=0\) for each A, B and C in \(\mathcal A\).

Similarly to Lie derivations, the authors always consider the problem of identifying those algebras on which every Lie triple derivation is standard. In [23], Miers proved that if \(\mathcal A\) is a von Neumann algebra with no central abelian summands, then every Lie triple derivation on \(\mathcal A\) is standard. In [11], Ji and Wang proved that every continuous Lie triple derivation on TUHF algebras is standard. In [28], Zhang et al. proved that if \(\mathcal N\) is a nest on a complex separable Hilbert space \(\mathcal H\), then every Lie triple derivation on the nest algebra \(\mathrm {Alg}\,\mathcal {N}\) is standard. In [27], Yu and Zhang studied the Lie triple derivations on commutative subspace lattice algebras. In [3], Benkovi\({\check{\text {c}}}\) showed that if \(\mathcal A\) is a unital algebra with a nontrivial idempotent, then under suitable assumptions every Lie triple derivation d on \(\mathcal A\) is of the form \(d=\Delta +\delta +\tau \), where \(\Delta \) is a derivation on \(\mathcal A\), \(\delta \) is a Jordan derivation on \(\mathcal A\) and \(\tau \) is a linear mapping from \(\mathcal A\) into its center \(\mathcal {Z}(\mathcal {A})\) that vanishes on \([[\mathcal A,\mathcal A],\mathcal A]\). In [1], Ashraf and Akhtar proved that every Lie triple derivation on a generalized matrix algebra is standard. In [26], Wani proved that every Lie triple derivation from standard operator algebra into itself is standard.

Now we give the concepts of local Lie triple derivations and 2-local Lie triple derivations. A linear mapping \(\delta \) from \(\mathcal {A}\) into \(\mathcal {M}\) is called a local Lie triple derivation if for every A in \(\mathcal {A}\) there exists a Lie triple derivation \(\delta _{A}\) (depending on A) from \(\mathcal {A}\) into \(\mathcal {M}\) such that \(\delta (A)=\delta _{A}(A)\). A mapping (not necessarily linear) \(\delta \) from \(\mathcal {A}\) into \(\mathcal {M}\) is called a 2-local Lie triple derivation if for every AB in \(\mathcal {A}\) there exists a Lie triple derivation \(\delta _{A,B}\) (depending on AB) from \(\mathcal {A}\) into \(\mathcal {M}\) such that \(\delta (A)=\delta _{A,B}(A)\) and \(\delta (B)=\delta _{A,B}(B)\).

In this paper, we always suppose that X is a Banach space over the field \(\mathbb F\) (\(\mathbb R\) or \(\mathbb C)\). Denote by B(X) the set of all linear mappings on X and by F(X) the set of all finite rank operators on X. A subalgebra \(\mathcal A\subseteq B(X)\) is called a standard operator algebra if \(F(X)\subseteq \mathcal A\). Suppose that \(\delta \) is a mapping from \(\mathcal A\) into B(X). In Sect. 2, we prove that if \(\delta \) is a Lie triple derivation, then \(\delta \) is standard. In Sect. 3, we prove that if \(\delta \) is a local Lie triple derivation and \(\mathrm {dim}(X)\ge 3\), then \(\delta \) is a Lie triple derivation. In Sect. 4, we prove that if \(\delta \) is a 2-local Lie triple derivation, then \(\delta =d+\tau \), where d is a derivation and \(\tau \) is a homogeneous mapping from \(\mathcal A\) into \(\mathbb {F}I\) such that \(\tau (A+B)=\tau (A)\) for each AB in \(\mathcal A\) where B is a sum of double commutators.

We shall review some simple properties of rank one operators and finite rank operators. Denote by \(X^*\) the set of all bounded linear functionals on X. For each x in X and f in \(X^{*}\), one can define an operator \(x\otimes f\) by \((x\otimes f)y=f(y)x\) for every y in X. Obviously, \(x\otimes f\in B(X)\). If both x and f are nonzero, then \(x\otimes f\) is an operator of rank one. The following properties are evident and will be used frequently in this paper.

Proposition 1.1

Suppose that X is a Banach space and \(\mathcal A\subseteq B(X)\) is a standard operator algebra. For each xy in X, fg in \(X^{*}\) and AB in B(X), the following statements hold:

  1. (1)

    \((x\otimes f)A=x\otimes (fA)\) and \(A(x\otimes f)=(Ax)\otimes f\);

  2. (2)

    \((x\otimes f)(y\otimes g)=f(y)(x\otimes g)\);

  3. (3)

    \(\mathcal {Z}(B(X),\mathcal A)=\mathbb {F}I\).

2 Lie triple derivations

In this section, we choose \(x_0\in X\) and \(f_0\in X^*\) such that \(f_0(x_0)=1\), and denote by I the unit operator in B(X). For the convenience of expression, we give some symbols firstly. Let \(P_1=x_0\otimes f_0\) and \(P_2=I-P_1\). It is easy to see that \(P_1\) and \(P_2\) are two idempotents in B(X). Denote \(P_i\mathcal A P_j\) and \(P_iB(X)P_j\) by \(\mathcal A_{ij}\) and \(B(X)_{ij}\), respectively, denote \(P_iAP_j\) by \(A_{ij}\) for every A in \(\mathcal A\), where \(1\le i,j\le 2\).

Lemma 2.1

\(P_1AP_1=f_0(Ax_0)P_1=f_0(P_1AP_1x_0)P_1\) for every A in B(X). Moreover, \(B(X)_{11}\) is commutative.

Proof

For every A in B(X), by Proposition 1.1 (1) and (2), we have

$$\begin{aligned} P_1AP_1=x_0\otimes f_0Ax_0\otimes f_0=f_0(Ax_0)x_0\otimes f_0=f_0(Ax_0)P_1. \end{aligned}$$
(2.1)

Replacing A by \(P_1AP_1\) in (2.1), we get

$$\begin{aligned} P_1AP_1=P_1P_1AP_1P_1=f_0(P_1AP_1x_0)P_1. \end{aligned}$$

It follows that \(B(X)_{11}\) is commutative. \(\square \)

Lemma 2.2

  1. (1)

    If \(BA_{21}=0\) for every \(A_{21}\) in \(\mathcal A_{21}\), then \(BP_2=0\).

  2. (2)

    If \(A_{12}B=0\) for every \(A_{12}\) in \(\mathcal A_{12}\), then \(P_2B=0\).

Proof

(1) Let \(A_{21}=P_2x\otimes f_0P_1\), where x is an arbitrary element in X. We obtain

$$\begin{aligned} 0=BP_2x\otimes f_0P_1x_0=f_0(P_1x_0)BP_2x=BP_2x. \end{aligned}$$

It follows that \(BP_2=0\).

(2) Let \(A_{12}=P_1x_0\otimes fP_2\), where f is an arbitrary element in \(X^*\). We obtain

$$\begin{aligned} 0=P_1x_0\otimes fP_2Bx=f(P_2Bx)P_1x_0=f(P_2Bx)x_0 \end{aligned}$$

for every x in X. It follows that \(f(P_2Bx)=0\) for each \(f\in X^*\) and x in X. Thus, \(P_2B=0\).

\(\square \)

Next we consider Lie triple derivations from a unital standard operator algebra \(\mathcal A\) into B(X). The following theorem is the main result in this section.

Theorem 2.3

Let X be a Banach space and \(\mathcal A\subseteq B(X)\) be a unital standard operator algebra. If \(\delta \) is a Lie triple derivation \(\delta \) from \(\mathcal A\) into B(X), then \(\delta \) is standard.

Before we prove Theorem 2.3, we present some lemmas.

Lemma 2.4

\(\delta (I)\in \mathbb {F}I\).

Proof

Let P be an idempotent in \(\mathcal A\). We have

$$\begin{aligned} 0=\delta ([[I,P],P])=[[\delta (I),P],P]=[\delta (I)P-P\delta (I),P]=\delta (I)P+P\delta (I)-2P\delta (I)P. \end{aligned}$$

Multiplying the above equation by P from the right, we obtain \(P\delta (I)P=\delta (I)P\). It means that \((I-P)\delta (I)P=0\). Thus, \(P_1\delta (I)P_2=P_2\delta (I)P_1=0\); it follows that \(\delta (I)\in B(X)_{11}+B(X)_{22}\). By Lemma 2.1, we know that \(\mathcal A_{11}\) is commutative, so \([\delta (I),A_{11}]=0\) for every \(A_{11}\) in \(\mathcal A_{11}\). In the following, we show

$$\begin{aligned}{}[\delta (I),A_{22}]=[\delta (I),A_{12}]=[\delta (I),A_{21}]=0 \end{aligned}$$

for every \(A_{22}\) in \(\mathcal A_{22}\), \(A_{12}\) in \(\mathcal A_{12}\) and \(A_{21}\) in \(\mathcal A_{21}\).

For each AB in \(\mathcal A\), we have

$$\begin{aligned}{}[[A,B],\delta (I)]=\delta ([[A,B],I])-[[A,\delta (B)],I]-[[\delta (A),B],I]=0. \end{aligned}$$

By \(A_{12}=[P_1,A_{12}]\) and \(A_{21}=[A_{21},P_1]\), we have

$$\begin{aligned}{}[\delta (I),A_{12}]=[\delta (I),A_{21}]=0. \end{aligned}$$
(2.2)

By (2.2), it follows that

$$\begin{aligned} 0=[\delta (I),A_{22}B_{21}]=[\delta (I),A_{22}]B_{21}+A_{22}[\delta (I),B_{21}]=[\delta (I),A_{22}]B_{21} \end{aligned}$$

for every \(A_{22}\) in \(\mathcal A_{22}\) and \(B_{21}\) in \(\mathcal A_{21}\). By Lemma 2.2, we have \([\delta (I),A_{22}]P_{2}=0\). By \(\delta (I)\in B(X)_{11}+B(X)_{22}\), we obtain \([\delta (I),A_{22}]\in B(X)_{22}\), it follows that \([\delta (I),A_{22}]=0\). Hence by Proposition 1.1 (3), we have \(\delta (I)\in \mathcal Z(B(X),\mathcal A)=\mathbb {F}I\). \(\square \)

Lemma 2.5

\(P_1\delta (P_1)P_1+P_2\delta (P_1)P_2\in \mathbb {F}I\).

Proof

By Lemma 2.1, we know that \(P_1\delta (P_1)P_1=\lambda P_1\), where \(\lambda =f_0(P_1\delta (P_1)P_1x_0)\in \mathbb {F}\). Let x be in X and let \(P_2x\otimes f_0P_1=A_{21}\). It follows that

$$\begin{aligned} -\delta (A_{21})&=\delta ([[P_2,A_{21}],P_2])\nonumber \\&=[[\delta (P_2),A_{21}],P_2]+[[P_2,\delta (A_{21})],P_2]+[[P_2,A_{21}],\delta (P_2)]\nonumber \\&=-A_{21}\delta (P_2)P_2-P_2\delta (P_2)A_{21}+A_{21}\delta (P_2)+2P_2\delta (A_{21})P_2\nonumber \\&\quad -\delta (A_{21})P_2-P_2\delta (A_{21})+A_{21}\delta (P_2)-\delta (P_2)A_{21}. \end{aligned}$$
(2.3)

Multiplying (2.3) by \(P_2\) from the left and by \(P_1\) from the right, we obtain

$$\begin{aligned} P_2\delta (P_2)A_{21}=A_{21}\delta (P_2)P_1. \end{aligned}$$

That is,

$$\begin{aligned} P_2\delta (P_2)P_2x\otimes f_0P_1=P_2x\otimes f_0P_1\delta (P_2)P_1. \end{aligned}$$
(2.4)

By letting both sides of (2.3) act on \(x_0\) in X, we have

$$\begin{aligned} f_0(P_1x_0)P_2\delta (P_2)P_2x=f_0(P_1\delta (P_2)P_1x_0)P_2x. \end{aligned}$$

Since \(f_0(P_1x_0)=f_0(x_0)=1\), it follows that

$$\begin{aligned} P_2\delta (P_2)P_2=f_0(P_1\delta (P_2)P_1x_0)P_2. \end{aligned}$$
(2.5)

By Lemma 2.4, we know that \(\delta (I)\in \mathbb {F}I\). It follows that

$$\begin{aligned} P_2\delta (I)P_2=\delta (I)P_2=\delta (I)f_0(x_0)P_2=f_0(\delta (I)x_0)P_2=f_0(P_1\delta (I)P_1x_0)P_2. \end{aligned}$$

Now replacing \(\delta (P_2)\) by \(\delta (I)-\delta (P_1)\) in (2.5), we obtain

$$\begin{aligned} P_2\delta (P_1)P_2=f_0(P_1\delta (P_1)P_1x_0)P_2=\lambda P_2. \end{aligned}$$

This implies \(P_1\delta (P_1)P_1+P_2\delta (P_1)P_2=\lambda (P_1+P_2)=\lambda I\). \(\square \)

Let \(G=P_1\delta (P_1)P_2-P_2\delta (P_1)P_1\) and define a mapping \(\Delta \) from \(\mathcal A\) into B(X) by

$$\begin{aligned} \Delta (A)=\delta (A)-[A,G] \end{aligned}$$

for every A in \(\mathcal A\). Obviously, \(\Delta \) is also a Lie triple derivation from \(\mathcal A\) into B(X). Moreover,

$$\begin{aligned} \Delta (P_1)=\delta (P_1)-[P_1,G]=P_1\delta (P_1)P_1+P_2\delta (P_1)P_2 \end{aligned}$$

and, by Lemma 2.5, we know that \(\Delta (P_1)\in \mathbb {F}I\). In Lemmas 2.6, 2.7 and 2.8, we show some properties of \(\Delta \).

Lemma 2.6

\(\Delta (\mathcal A_{ij})\subseteq B(X)_{ij}\), where \(1\le i,j\le 2\) and \(i\ne j\).

Proof

Since \(\Delta (P_1)\in \mathbb {F}I\), for each \(A_{12}\) in \(\mathcal A_{12}\), we have

$$\begin{aligned} \Delta (A_{12})&=\Delta ([[A_{12},P_1],P_1])\nonumber \\&=[[\Delta (A_{12}),P_{1}],P_1]+[[A_{12},\Delta (P_1)],P_1]+[[A_{12},P_1],\Delta (P_1)]\nonumber \\&=[[\Delta (A_{12}),P_{1}],P_1]\nonumber \\&=P_1\Delta (A_{12})P_2+P_2\Delta (A_{12})P_1. \end{aligned}$$
(2.6)

In the following, we show that \(P_{2}\Delta (A_{12})P_1=0\).

Let \(B_{12}\) be in \(\mathcal A_{12}\), then \([A_{12},B_{12}]=0\). Thus,

$$\begin{aligned} 0=\Delta (0)=\Delta ([[A_{12},B_{12}],C])&=[[\Delta (A_{12}),B_{12}],C]+[[A_{12},\Delta (B_{12})],C]\nonumber \\&=[[\Delta (A_{12}),B_{12}]+[A_{12},\Delta (B_{12})],C] \end{aligned}$$

for every C in \(\mathcal A\). It means that \(J=[\Delta (A_{12}),B_{12}]+[A_{12},\Delta (B_{12})]\in \mathbb {F}I\). Since \(A_{12}=[P_1,A_{12}]\), we have

$$\begin{aligned}{}[\Delta (A_{12}),B_{12}]&=J-[A_{12},\Delta (B_{12})]=J-[[P_1,A_{12}],\Delta (B_{12})]\nonumber \\&=J-(\Delta ([[P_1,A_{12}],B_{12}])-[[\Delta (P_1),A_{12}],B_{12}]-[[P_1,\Delta (A_{12})],B_{12}])\nonumber \\&=J+[[P_1,\Delta (A_{12})],B_{12}]. \end{aligned}$$

By (2.6), we have

$$\begin{aligned}{}[P_1\Delta (A_{12})P_2+P_2\Delta (A_{12})P_1,B_{12}]&=J+[[P_1,P_1\Delta (A_{12})P_2+P_2\Delta (A_{12})P_1],B_{12}]\nonumber \\&=J+[P_1\Delta (A_{12})P_2-P_2\Delta (A_{12})P_1,B_{12}]. \end{aligned}$$

Hence

$$\begin{aligned}{}[P_2\Delta (A_{12})P_1,B_{12}]=\frac{1}{2}J\in \mathbb {F}I. \end{aligned}$$

It is well known that \([P_2\Delta (A_{12})P_1,B_{12}]=0\). Thus, \(P_2\Delta (A_{12})B_{12}=B_{12}\Delta (A_{12})P_1=0\) for every \(B_{12}\) in \(\mathcal A_{12}\). By Lemma 2.2, we know that \(P_{2}\Delta (A_{12})P_1=0\). Similarly, we have \(\Delta (\mathcal A_{21})\subseteq B(X)_{21}\). \(\square \)

Lemma 2.7

\(\Delta (\mathcal A_{11})\subseteq \mathbb {F}I.\)

Proof

For every \(A_{11}\) in \(\mathcal A_{11}\), by Lemma 2.1, we have

$$\begin{aligned} \Delta (A_{11})=\Delta (P_1A_{11}P_1)=\Delta (f_0(A_{11}x_0)P_1)=f_0(A_{11}x_0)\Delta (P_1). \end{aligned}$$

Since \(\Delta (P_1)\in \mathbb {F}I\), it follows that \(\Delta (A_{11})\in \mathbb {F}I\). \(\square \)

Lemma 2.8

\(\Delta (A_{22})-f_0(\Delta (A_{22})x_0)I\in B(X)_{22}\) for every \(A_{22}\) in \(\mathcal A_{22}\). In particular, \(\Delta (P_2)=f_0(\Delta (P_2)x_0)I.\)

Proof

Through simple calculation, we get

$$\begin{aligned} 0=\Delta ([[P_1,A_{22}],P_1])&=[[P_1,\Delta (A_{22})],P_1]=-P_1\Delta (A_{22})P_2-P_2\Delta (A_{22})P_1. \end{aligned}$$

It follows that \(\Delta (A_{22})\in B(X)_{11}+ B(X)_{22}\). By Lemma 2.1, we obtain

$$\begin{aligned} \Delta (A_{22})=P_1\Delta (A_{22})P_1+P_2\Delta (A_{22})P_2=f_0(\Delta (A_{22})x_0)P_1+P_2\Delta (A_{22})P_2, \end{aligned}$$

that is,

$$\begin{aligned} \Delta (A_{22})-f_0(\Delta (A_{22})x_0)I=-f_0(\Delta (A_{22})x_0)P_2+P_2\Delta (A_{22})P_2\in B(X)_{22}. \end{aligned}$$

Since \(\Delta (P_2)=\Delta (I)-\Delta (P_1)\in \mathbb {F}I\), we have

$$\begin{aligned} \Delta (P_2)-f_0(\Delta (P_2)x_0)I\in \mathbb {F}I\cap B(X)_{22}=\{0\}. \end{aligned}$$

Thus, \(\Delta (P_2)=f_0(\Delta (P_2)x_0)I.\) \(\square \)

In the following, we prove Theorem 2.3.

Proof

Define two mappings \(\tau \) and D on from \(\mathcal A\) into B(X) by

$$\begin{aligned} \tau (A)=f_0(P_1AP_1x_0)\Delta (P_1)+f_0(\Delta (P_2AP_2)x_0)I \end{aligned}$$

and

$$\begin{aligned} D(A)=\Delta (A)-\tau (A) \end{aligned}$$

for every A in \(\mathcal A\). It is clear that \(\tau \) is a linear mapping from \(\mathcal A\) into \(\mathcal Z(B(X),\mathcal A)\) and D is a linear mapping from \(\mathcal A\) into B(X). Moreover, according to the previous lemmas and the definitions of \(\tau \) and D, we have

  1. (1)

    \(D(A_{ij})=\Delta (A_{ij})\in B(X)_{ij}\) for every \(A_{ij}\) in \(\mathcal A_{ij}\), where \(1\le i,j\le 2\) and \(i\ne j\);

  2. (2)

    \(D(P_1)=D(P_2)=D(I)=0\);

  3. (3)

    \(D(A_{11})=0\) for every \(A_{11}\) in \(\mathcal A_{11}\);

  4. (4)

    \(D(A_{22})\in B(X)_{22}\) for every \(A_{22}\) in \(\mathcal A_{22}\).

To prove that \(\Delta \) is standard, it is sufficient to show that D is a derivation and \(\tau ([[A,B],C])=0\) for each A, B and C in \(\mathcal A\).

In the following we show

$$\begin{aligned} D(A_{ij}B_{sk})=D(A_{ij})B_{sk}+A_{ij}D(B_{sk}) \end{aligned}$$

for every \(A_{ij}\) in \(\mathcal A_{ij}\) and \(B_{sk}\) in \(\mathcal A_{sk}\), where \(1\le i,j,s,k\le 2\).

Since \(D(\mathcal A_{ij})\in B(X)_{ij}\), we have

$$\begin{aligned} D(A_{ij}B_{sk})=D(A_{ij})B_{sk}+A_{ij}D(B_{sk}) \end{aligned}$$

for \(j\ne s\). Thus, we only need to prove the following 8 cases:

  1. (1)

    \(D(A_{11}B_{11})=D(A_{11})B_{11}+A_{11}D(B_{11})\);

  2. (2)

    \(D(A_{11}B_{12})=D(A_{11})B_{12}+A_{11}D(B_{12})\);

  3. (3)

    \(D(A_{12}B_{22})=D(A_{12})B_{22}+A_{12}D(B_{22})\);

  4. (4)

    \(D(A_{21}B_{11})=D(A_{21})B_{11}+A_{21}D(B_{11})\);

  5. (5)

    \(D(A_{22}B_{21})=D(A_{22})B_{21}+A_{22}D(B_{21})\);

  6. (6)

    \(D(A_{22}B_{22})=D(A_{22})B_{22}+A_{22}D(B_{22})\);

  7. (7)

    \(D(A_{12}B_{21})=D(A_{12})B_{21}+A_{12}D(B_{21})\);

  8. (8)

    \(D(A_{21}B_{12})=D(A_{21})B_{12}+A_{21}D(B_{12})\).

Since \(D(A_{11})=0\) for every \(A_{11}\) in \(\mathcal A_{11}\), case (1) is trivial.

For each AB in \(\mathcal A\), by \(\Delta (A)-D(A)=\tau (A)\in \mathcal Z(B(X),\mathcal A)\), we have \([\Delta (A),B]=[D(A),B]\). Therefore

$$\begin{aligned} D(A_{11}B_{12})&=\Delta (A_{11}B_{12})=-\Delta ([[P_1,B_{12}],A_{11}])\\&=-[[P_1,\Delta (B_{12})],A_{11}]-[[P_1,B_{12}],\Delta (A_{11})]\\&=-[\Delta (B_{12}),A_{11}]-[B_{12},\Delta (A_{11})]\\&=[A_{11},D(B_{12})]+[D(A_{11}),B_{12}]\\&=A_{11}D(B_{12})+D(A_{11})B_{12} \end{aligned}$$

for each \(A_{11}\) in \(\mathcal A_{11}\) and \(B_{12}\) in \(\mathcal A_{12}\). Thus, case (2) holds. The cases (3), (4) and (5) are similar to case (2), so we omit the proofs.

For every \(C_{21}\) in \(\mathcal A_{21}\), according to case (5), we have the following two equations:

$$\begin{aligned} D(A_{22}B_{22}C_{21})=D(A_{22}B_{22})C_{21}+A_{22}B_{22}D(C_{21}) \end{aligned}$$
(2.7)

and

$$\begin{aligned} D(A_{22}B_{22}C_{21})&=D(A_{22})B_{22}C_{21}+A_{22}D(B_{22}C_{21})\nonumber \\&=D(A_{22})B_{22}C_{21}+A_{22}D(B_{22})C_{21}+A_{22}B_{22}D(C_{21}) \end{aligned}$$
(2.8)

for each \(A_{22},B_{22}\) in \(\mathcal A_{22}\). Comparing (2.7) and (2.8), we have

$$\begin{aligned} D(A_{22}B_{22})C_{21}=D(A_{22})B_{22}C_{21}+A_{22}D(B_{22})C_{21}. \end{aligned}$$

It follows that \((D(A_{22}B_{22})-D(A_{22})B_{22}-A_{22}D(B_{22}))C_{21}=0\) for every \(C_{21}\) in \(\mathcal A_{21}\). By Lemma 2.2 and \(D(A_{22})\in \mathcal A_{22}\), we know that

$$\begin{aligned} D(A_{22}B_{22})-D(A_{22})B_{22}-A_{22}D(B_{22})=0. \end{aligned}$$

Finally, we show cases (7) and (8). Let \(A_{12}\) be in \(\mathcal A_{12}\) and \(B_{21}\) be in \(\mathcal A_{21}\). Through simple calculation, we obtain

$$\begin{aligned}&\Delta ([[A_{12},P_2],B_{21}])-D([[A_{12},P_2,],B_{21}])\\&\quad =[[\Delta (A_{12}),P_2],B_{21}]+[[A_{12},P_2],\Delta (B_{21})]-D([[A_{12},P_2,],B_{21}])\\&\quad =[\Delta (A_{12}),B_{21}]+[A_{12},\Delta (B_{21})]-D[A_{12},B_{21}]\\&\quad =[D(A_{12}),B_{21}]+[A_{12},D(B_{21})]-D(A_{12}B_{21}-B_{21}A_{12})\\&\quad =D(A_{12})B_{21}-B_{21}D(A_{12})+A_{12}D(B_{21})-D(B_{21})A_{12}-D(A_{12}B_{21})+D(B_{21}A_{12})\\&\quad =(D(A_{12})B_{21}+A_{12}D(B_{21})-D(A_{12}B_{21}))+(D(B_{21}A_{12})-B_{21}D(A_{12})-D(B_{21})A_{12}). \end{aligned}$$

Since \(\Delta ([A_{12},B_{21}])-D([A_{12},B_{21}])\) belongs to \(\mathbb {F}I\), we may assume that

$$\begin{aligned} \Delta ([A_{12},B_{21}])-D([A_{12},B_{21}])=\lambda I \end{aligned}$$

holds for some \(\lambda \) in \(\mathbb {F}\). That is,

$$\begin{aligned} \lambda I&=(D(A_{12})B_{21}+A_{12}D(B_{21})-D(A_{12}B_{21}))\nonumber \\&\quad +(D(B_{21}A_{12})-B_{21}D(A_{12})-D(B_{21})A_{12}). \end{aligned}$$
(2.9)

Since \(D(\mathcal A_{ij})\in B(X)_{ij}\), we get

$$\begin{aligned} D(A_{12})B_{21}+A_{12}D(B_{21})-D(A_{12}B_{21})\in B(X)_{11} \end{aligned}$$

and

$$\begin{aligned} D(B_{21}A_{12})-B_{21}D(A_{12})-D(B_{21})A_{12}\in B(X)_{22}. \end{aligned}$$

Multiplying (2.9) by \(P_1\) and \(P_2\) respectively from the right, we obtain the following two equations:

$$\begin{aligned} D(A_{12}B_{21})=D(A_{12})B_{21}+A_{12}D(B_{21})-\lambda P_1 \end{aligned}$$
(2.10)

and

$$\begin{aligned} D(B_{21}A_{12})=B_{21}D(A_{12})+D(B_{21})A_{12}+\lambda P_2. \end{aligned}$$
(2.11)

By case (2) and equation (2.10), we obtain

$$\begin{aligned}&D(A_{12}B_{21}A_{12})\nonumber \\&\quad =D(A_{12}B_{21})A_{12}+A_{12}B_{21}D(A_{12})\nonumber \\&\quad =D(A_{12})B_{21}A_{12}+A_{12}D(B_{21})A_{12}-\lambda A_{12}+A_{12}B_{21}D(A_{12}). \end{aligned}$$
(2.12)

By case (3) and equation (2.9), we obtain

$$\begin{aligned}&D(A_{12}B_{21}A_{12})\nonumber \\&\quad =D(A_{12})B_{21}A_{12}+A_{12}D(B_{21}A_{12})\nonumber \\&\quad =D(A_{12})B_{21}A_{12}+A_{12}D(B_{21})A_{12}+A_{12}B_{21}D(A_{12})+\lambda A_{12}. \end{aligned}$$
(2.13)

Comparing (2.12) and (2.13), we have \(\lambda A_{12}=0\). Thus, \(\lambda =0\). By (2.10) and (2.9), cases (7) and (8) hold.

By cases (1)–(8), this implies immediately that D is a derivation. Now we show that \(\tau ([[A,B],C])=0\) for each A, B and C in \(\mathcal A\). Indeed,

$$\begin{aligned} \tau ([[A,B],C])&=\, \Delta ([[A,B],C])-D([[A,B],C])\nonumber \\&=\, [[\Delta (A),B],C]+[[A,\Delta (B)],C]+[[A,B],\Delta (C)]-D([[A,B],C])\nonumber \\&=\, [[D(A),B],C]+[[A,D(B)],C]+[[A,B],D(C)]-D([[A,B],C])\nonumber \\&=\, 0.\nonumber \end{aligned}$$

It follows that \(\Delta (A)=D(A)+\tau (A)\) is a standard Lie triple derivation from \(\mathcal A\) into B(X). Define a linear mapping from \(\mathcal A\) into B(X) by

$$\begin{aligned} d(A)=D(A)+[A,G] \end{aligned}$$

for every A in \(\mathcal A\). Thus, we have

$$\begin{aligned} \delta (A)=\Delta (A)+[A,G]=D(A)+\tau (A)+[A,G]=d(A)+\tau (A), \end{aligned}$$

where d is a derivation from \(\mathcal A\) into B(X) and \(\tau \) is a linear mapping from \(\mathcal A\) into \(\mathcal Z(B(X),\mathcal A)\) such that \(\tau ([[A,B],C])=0\) for each A, B and C in \(\mathcal A\). \(\square \)

For a non-unital standard operator algebra, the following result holds.

Corollary 2.9

Let X be a Banach space and \(\mathcal A\subseteq B(X)\) be a non-unital standard operator algebra. If \(\delta \) is a Lie triple derivation \(\delta \) from \(\mathcal A\) into B(X), then \(\delta \) is standard.

Proof

Denote the unital algebra \(\mathcal A\oplus \mathbb {F}I\) by \(\widetilde{\mathcal A}\). Thus, \(\widetilde{\mathcal A}\) is a unital standard operator algebra. Define a linear mapping \(\widetilde{\delta }\) from \(\widetilde{\mathcal A}\) into B(X) by

$$\begin{aligned} \widetilde{\delta }(A+\lambda I)=\delta (A) \end{aligned}$$

for every A in \(\mathcal A \) and \(\lambda \) in \(\mathbb F\). Through a simple calculation, it is easy to show that \(\widetilde{\delta }\) is also a Lie triple derivation. By Theorem 2.3, we know that \(\widetilde{\delta }\) is standard, and so is \(\delta \). \(\square \)

3 Local Lie triple derivations

In this section, we study local Lie triple derivations and the following theorem is the main result.

Theorem 3.1

Let X be a Banach space of dimension at least 3 and \(\mathcal A\subseteq B(X)\) be a unital standard operator algebra. If \(\delta \) is a local Lie triple derivation \(\delta \) from \(\mathcal A\) into B(X), then \(\delta \) is a Lie triple derivation.

Proof

For every A in B(X), there is a Lie triple derivation \(\delta _A\) from \(\mathcal A\) into B(X) such that \(\delta (A)=\delta _A(A).\) By Theorem 2.3, we know \(\delta _A(A)\) is standard, then there exist a derivation \(d_A\) from \(\mathcal A\) into B(X) and a scalar operator \(\tau _A(A)\) in \(\mathbb {F}I\) such that \(\delta (A)=d_A(A)+\tau _A(A).\) By [6, Corollary 3.4], we know that \(d_A\) is an inner derivation, then there exists an element \(T_A\) in B(X) such that \(d_A(A)=[A,T_A]\). Thus, we have

$$\begin{aligned} \delta (A)=d_A(A)=[A,T_A]+\tau _A(A). \end{aligned}$$

We claim that \(\tau _A(A)\) is unique. In fact, if

$$\begin{aligned} \delta (A)=[A,S_A]+\tau '_A(A) \end{aligned}$$

for some \(S_A\) in B(X) and \(\tau '_A(A)\) in \(\mathbb {F}I\), then

$$\begin{aligned}{}[A,S_A-T_A]=\tau _A(A)-\tau '_A(A)=\lambda I \end{aligned}$$

for some \(\lambda \) in \(\mathbb {F}\). It is well known that \(\tau _A(A)=\tau '_A(A)\). Hence we can define a mapping from \(\mathcal A\) into \(\mathbb {F}I\) by

$$\begin{aligned} \tau (A)=\tau _A(A) \end{aligned}$$

for every A in \(\mathcal {A}\). Moreover, by the definition of \(\tau \) and Theorem 2.3, we know that \(\tau (A)=\tau _A(A)=0\) if A is a sum of double commutators.

For each x in X and f in \(X^*\), define \(\psi (x,f)=\tau (x\otimes f)\). Then we have

$$\begin{aligned} \delta (x\otimes f)=[x\otimes f,T_{x\otimes f}]+\psi (x,f) \end{aligned}$$
(3.1)

for some \(T_{x\otimes f}\) in B(X). In the following we show that \(\psi (x,f)\) is a bilinear mapping.

Firstly, we show the homogeneity of \(\psi \). For each x in X, f in \(X^*\) and \(\lambda \) in \(\mathbb {F}\), by (3.1), we have

$$\begin{aligned} \delta (x\otimes f)=[x\otimes f,T_{x\otimes f}]+\psi (x,f)~~ \mathrm {and} ~~\delta (\lambda x\otimes f)=[\lambda x\otimes f,T_{\lambda x\otimes f}]+\psi (\lambda x,f). \end{aligned}$$

By \(\delta (\lambda x\otimes f)=\lambda \delta (x\otimes f)\), we infer

$$\begin{aligned}{}[\lambda x\otimes f,T_{x\otimes f}-T_{\lambda x\otimes f}]=\psi (\lambda x,f)-\lambda \psi (x,f)\in \mathbb {F}I. \end{aligned}$$

Thus, \(\lambda \psi (x,f)=\psi (\lambda x,f)\). This proved that \(\psi \) is homogenous in the first variable. In the same way, we can show that \(\psi \) is homogenous in the second variable.

Secondly, we show that \(\psi (x,f)\) is biadditive. We note that \(\psi (x,f)=0\) for x in X and f in \(X^*\) with \(f(x)=0\). Indeed, we may choose an element z in X such that \(f(z)=1\), then \(x\otimes f=[[x\otimes f,z\otimes f],z\otimes f]\) is a double commutator and hence \(\psi (x,f)=\tau (x\otimes f)=0\).

Let \(x_1, x_2\) be in X and f be in \(X^*\). If both \(x_1\) and \(x_2\) belong to \(\mathrm {ker}f\), then

$$\begin{aligned} \psi (x_1,f)=\psi (x_2,f)=\psi (x_1+x_2,f)=0 \end{aligned}$$

and so

$$\begin{aligned} \psi (x_1+x_2,f)=\psi (x_1,f)+\psi (x_2,f). \end{aligned}$$

If one of \(x_1\) and \(x_2\) is not in \(\mathrm {ker}f\), then \(\mathrm {dim}(\mathrm {span}\{x_1,x_2\}\cap \mathrm {ker}f)\le 1\). Since \(\mathrm {dim}(X)\ge 3\), we know that \(\mathrm {dim}(\mathrm {ker}f)\ge 2\). Thus, we can take \(y\in \mathrm {ker}f\) such that \(y\notin \mathrm {span}\{x_1,x_2\}\). By (3.1), we have the following equations:

$$\begin{aligned} \delta (x_1\otimes f)y=\psi (x_1,f)y+\mu _1x_1~~~~\delta (x_2\otimes f)y=\psi (x_2,f)y+\mu _2x_2 \end{aligned}$$

and

$$\begin{aligned} \delta ((x_1+x_2)\otimes f)y=\psi (x_1+x_2,f)y+\mu (x_1+x_2) \end{aligned}$$

for some \(\mu ,\mu _1,\mu _2\in \mathbb {F}\). Since \(\delta \) is an additive mapping, we know that

$$\begin{aligned} (\psi (x_1+x_2,f)-\psi (x_1,f)-\psi (x_2,f))y=\mu _1x_1+\mu _2x_2-\mu (x_1+x_2). \end{aligned}$$

Since \(y\notin \) span \(\{x_1,x_2\}\), it follows that

$$\begin{aligned} \psi (x_1+x_2,f)=\psi (x_1,f)+\psi (x_2,f). \end{aligned}$$

It means that \(\psi \) is additive in the first variable.

Let \(f_1,f_2\) be in \(X^*\) and x be in X. If \(x\in \mathrm {ker}f_1\cap \mathrm {ker}f_2\), then

$$\begin{aligned} \psi (x,f_1+f_2)=\psi (x,f_1)=\psi (x,f_2)=0 \end{aligned}$$

and so

$$\begin{aligned} \psi (x,f_1+f_2)=\psi (x,f_1)+\psi (x,f_2). \end{aligned}$$

If \(x\notin \mathrm {ker}f_1\cap \mathrm {ker}f_2\), then we can take \(z\in \mathrm {ker}f_1\cap \mathrm {ker}f_2\) which is linearly independent of x, By (3.1), we have

$$\begin{aligned} \delta (x\otimes f_1)z=\psi (x,f_1)z+\lambda _1x~~~~\delta (x\otimes f_2)z=\psi (x,f_2)z+\lambda _2x \end{aligned}$$

and

$$\begin{aligned} \delta (x\otimes (f_1+f_2))z=\psi (x,f_1+f_2)z+\lambda x \end{aligned}$$

for some \(\lambda ,\lambda _1,\lambda _2\in \mathbb {F}\). Since \(\delta \) is an additive mapping, we know that

$$\begin{aligned} (\psi (x,f_1+f_2)-\psi (x,f_1)-\psi (x,f_2))z=(\lambda _1+\lambda _2-\lambda )x. \end{aligned}$$

Since z and x are linearly independent, it follows that

$$\begin{aligned} \psi (x,f_1+f_2)=\psi (x,f_1)+\psi (x,f_2). \end{aligned}$$

The next goal is to show that there is an element J in B(X) such that

$$\begin{aligned} \delta (x\otimes f)=[x\otimes f,J]+\psi (x,f) \end{aligned}$$

for every rank one operator \(x\otimes f\) in B(X).

For each x in X and f in \(X^*\), define

$$\begin{aligned} \phi (x,f)=[x\otimes f,T_{x\otimes f}]=\delta (x\otimes f)-\psi (x,f). \end{aligned}$$
(3.2)

It is easy to see that \(\phi (x,f)\) is a bilinear mapping and \(\phi (x,f)\mathrm {ker}f\subseteq \mathbb {F}x\). Hence by [21, Proposition 1.1], there are two linear mappings \(T:X\rightarrow X\) and \(S^*:X^*\rightarrow X^*\) such that

$$\begin{aligned} \phi (x,f)=[x\otimes f,T_{x\otimes f}]=Tx\otimes f+x\otimes S^*f \end{aligned}$$
(3.3)

for each x in X and f in \(X^*\). It follows that

$$\begin{aligned} (T+T_{x\otimes f})x\otimes f=x\otimes (T^*_{x\otimes f}-S^*)f \end{aligned}$$
(3.4)

for each x in X and f in \(X^*\).

We claim that \(S^*=-T^*\). We only have to show that \(S^*f(x)=-f(Tx)\) for each x in X and f in \(X^*\). It is trivial if one of x and f is zero. Suppose that neither of x and f is zero. If both sides of (3.4) are zeros, then

$$\begin{aligned} (T+T_{x\otimes f})x=(T^*_{x\otimes f}-S^*)f=0. \end{aligned}$$

It follows that

$$\begin{aligned} S^*f(x)=T^*_{x\otimes f}f(x)=f(T_{x\otimes f}x)=-f(Tx). \end{aligned}$$

If both sides of (3.4) are not zeros, then we have

$$\begin{aligned}{}[(T+T_{x\otimes f})x\otimes f]^2=[x\otimes (T^*_{x\otimes f}-S^*)f]^2, \end{aligned}$$

that is,

$$\begin{aligned} f((T+T_{x\otimes f})x)((T+T_{x\otimes f})x\otimes f)=((T^*_{x\otimes f}-S^*)f)(x)(x\otimes (T^*_{x\otimes f}-S^*)f). \end{aligned}$$

It follows that

$$\begin{aligned} f((T+T_{x\otimes f})x)=((T^*_{x\otimes f}-S^*)f)(x) \end{aligned}$$

and then \(S^*f(x)=-f(Tx)\). Consequently, we always have \(S^*=-T^*\). By (3.2) and (3.3), we have

$$\begin{aligned} \delta (x\otimes f)=Tx\otimes f+x\otimes S^*f +\psi (x,f)=[x\otimes f,-T]+\psi (x,f) \end{aligned}$$

for every \(x\otimes f\) in \(\mathcal A\). Let \(J=-T\) and by \(\psi (x,f)=\tau (x\otimes f)\in \mathbb {F}I\), we obtain

$$\begin{aligned} \delta (A)=[A,J]+\lambda _{A}I \end{aligned}$$
(3.5)

for every \(A=x\otimes f\) in \(\mathcal A\) and some \(\lambda _{A}\in \mathbb {F}\). Finally, we show that

$$\begin{aligned} \delta (A)=[A,J]+\lambda _AI \end{aligned}$$

holds for every A in \(\mathcal {A}\). Suppose that PQ are two idempotents of rank one and let \(P^\perp =I-P\), \(Q^\perp =I-Q\). By Proposition 1.1(1) and (3.5), it follows that

$$\begin{aligned} \delta (A)&=\delta (PA+P^\perp AQ+P^\perp AQ^\perp )\nonumber \\&=\delta (PA)+\delta (P^\perp AQ)+\delta (P^\perp AQ^\perp )\nonumber \\&=[PA,J]+\lambda _{PA}I+[P^\perp AQ,J]+\lambda _{P^\perp AQ}I+[P^\perp AQ^\perp ,T_{P^\perp AQ^\perp }]+\lambda _{P^\perp AQ^\perp }I\nonumber \\&=PAJ-JPA+P^\perp AQJ-JP^\perp AQ+[P^\perp AQ^\perp ,T_{P^\perp AQ^\perp }]+\lambda _A I\nonumber \\&=PAJ-JAQ+P^\perp AQJ-JPAQ^\perp +[P^\perp AQ^\perp ,T_{P^\perp AQ^\perp }]+\lambda _A I, \end{aligned}$$
(3.6)

where \(\lambda _A=\lambda _{PA}+\lambda _{P^\perp AQ}+\lambda _{P^\perp AQ^\perp }\). Multiplying (3.6) by P on the left and by Q on the right, we have

$$\begin{aligned} P\delta (A)Q=P[A,J]Q+\lambda PQ, \end{aligned}$$

that is,

$$\begin{aligned} P(\delta (A)-[A,J]-\lambda _A I)Q=0. \end{aligned}$$

By the arbitrariness of P and Q, it follows that \(\delta (A)=[A,J]+\lambda _A I\), where J is a fixed element and \(\lambda _A\) is depends on A. By the uniqueness of \(\tau \), we know that \(\tau (A)=\lambda _A I\) and \(\tau \) is a linear mapping from \(\mathcal A\) into \(\mathbb {F}I\) vanishing on every double commutator, which means that \(\delta \) is a Lie triple derivation. \(\square \)

Corollary 3.2

Let X be a Banach space of dimension at least 3 and \(\mathcal A\subseteq B(X)\) be a non-unital standard operator algebra. If \(\delta \) is a local Lie triple derivation \(\delta \) from \(\mathcal A\) into B(X), then \(\delta \) is a Lie triple derivation.

Proof

Denote the unital algebra \(\mathcal A\oplus \mathbb {F}I\) by \(\widetilde{\mathcal A}\). Thus, \(\widetilde{\mathcal A}\) is a unital standard operator algebra. Define a linear mapping \(\widetilde{\delta }\) from \(\widetilde{\mathcal A}\) into B(X) by

$$\begin{aligned} \widetilde{\delta }(A+\lambda I)=\delta (A) \end{aligned}$$

for every A in \(\mathcal A \) and \(\lambda \) in \(\mathbb F\).

Since \(\delta \) is a local Lie triple derivation from \(\mathcal A\) into B(X), for each \(A\in \mathcal A \) and \(\lambda \in \mathbb F\), there exists a Lie triple derivation \(\delta _A\) such that \(\delta (A)=\delta _A(A)\). Define a linear mapping \(\widetilde{\delta _A}\) from \(\widetilde{\mathcal A}\) into B(X) by

$$\begin{aligned} \widetilde{\delta _A}(B+\lambda I)=\delta _A(B) \end{aligned}$$

for every B in \(\mathcal A \) and \(\lambda \) in \(\mathbb F\). It is easy to show that \(\widetilde{\delta _A}\) is also a Lie triple derivation. Moreover, we have

$$\begin{aligned} \widetilde{\delta }(A+\lambda I)=\delta (A)=\delta _A(A)=\widetilde{\delta _A}(A+\lambda I). \end{aligned}$$

It means that \(\widetilde{\delta }\) is a local Lie triple derivation from \(\widetilde{\mathcal A}\) into B(X). By the result of the case that \(\mathcal A\) contains the unit, \(\widetilde{\delta }\) is a Lie triple derivation. Hence \(\delta \) is also a Lie triple derivation. \(\square \)

4 2-Local Lie triple derivations

In this section, we study the 2-local Lie triple derivations and the following theorem is the main result.

Theorem 4.1

Let X be a Banach space and \(\mathcal A\subseteq B(X)\) be a unital standard operator algebra. If \(\delta \) is a 2-local Lie triple derivation from \(\mathcal A\) into B(X), then \(\delta =d+\tau \), where d is a derivation and \(\tau \) is a homogeneous mapping from \(\mathcal A\) into \(\mathbb {F}I\) such that \(\tau (A+B)=\tau (A)\) for each AB in \(\mathcal A\) where B is a sum of double commutators.

Proof

Similarly to the proof of Theorem 3.1, we can show that \(\delta \) has a unique decomposition at each point A in \(\mathcal {A}\), i.e.

$$\begin{aligned} \delta (A)=\delta _A(A)=d_A(A)+\tau _A(A), \end{aligned}$$

where \(\delta _{A}\) is a Lie triple derivation, \(d_A\) is a derivation and \(\tau _A\) is a linear mapping from \(\mathcal {A}\) into \(\mathbb {F}I\) such that \(\tau _A[[X,Y],Z]=0\) each X, Y and Z in \(\mathcal A\).

Thus, we can define

$$\begin{aligned} d(A)=d_A(A)~~\mathrm {and}~~\tau (A)=\tau _A(A) \end{aligned}$$

for every A in \(\mathcal {A}\).

In the following we show that d is a derivation and \(\tau \) is a homogeneous mapping. Given A and B in \(\mathcal {A}\), there exists a Lie triple derivation \(\delta _{A,B}\) from \(\mathcal A\) into B(X) such that

$$\begin{aligned} \delta (A)=\delta _{A,B}(A)=d_{A,B}(A)+\tau _{A,B}(A), \end{aligned}$$

and

$$\begin{aligned} \delta (B)=\delta _{A,B}(B)=d_{A,B}(B)+\tau _{A,B}(B), \end{aligned}$$

where \(d_{A,B}\) + \(\tau _{A,B}\) is the standard decomposition of \(\delta _{A,B}\). By the uniqueness of the decomposition, \(d(A)=d_{A,B}(A)\) and \(d(B)=d_{A,B}(B)\). Hence d is a 2-local derivation and by [10, Theorem 3.1], we know d is a derivation from \(\mathcal A\) into B(X).

For every A in \(\mathcal A\) and \(\lambda \) in \(\mathbb {F}\), there exists a Lie triple derivation \(\delta _{A,\lambda A}\) from \(\mathcal A\) into B(X) such that

$$\begin{aligned} \delta (A)=\delta _{A,\lambda A}(A)~~\mathrm {and}~~\delta (\lambda A)=\delta _{A,\lambda A}(\lambda A). \end{aligned}$$

It follows that \(\delta \) is homogeneous, and so is \(\tau \).

Moreover, for each AB in \(\mathcal A\) where B is a sum of double commutators, there is a linear mapping \(\tau _{A,A+B}\) from \(\mathcal {A}\) into \(\mathbb {F}I\) vanishing on every double commutator such that

$$\begin{aligned} \tau (A+B)=\tau _{A,A+B}(A+B)=\tau _{A,A+B}(A)=\tau (A). \end{aligned}$$

\(\square \)