Cohomologies and deformations of weighted Rota-Baxter Lie algebras and associative algebras with derivations

Basdouri Imed \(^{1}\) 1, Sadraoui Mohamed Amin \(^{2}\) 2, Shuangjian Guo \(^{3}\) 3

1. University of Gafsa, Faculty of Sciences Gafsa, 2112 Gafsa, Tunisia
2. University of Sfax, Faculty of Sciences Sfax, BP 1171, 3038 Sfax, Tunisia
3. School of Mathematics and statistics,Guizhou University of Finance and Economics, Guiyang, 550025


Abstract

The purpose of the present paper is to investigate cohomologies and deformations of weighted Rota-Baxter Lie algebras as well as weighted Rota-Baxter associative algebras with derivations. First we introduce a notion of weighted Rota-Baxter LieDer and weighted Rota-Baxter AssDer pairs. Then we construct cohomologies of weighted Rota-Baxter LieDer pairs, weighted Rota-Baxter AssDer pairs and we discuss the relation between their cohmologies. Finally, as an application, we study deformations of both of them.

Key words:Lie algebras, Lie algebras with derivation, Rota-Baxter operators, cohomology, deformation .

1 Introduction↩︎

Rota-Baxter operators were first studied in the work of Baxter of the fluctuation theory in probability [1] and further was developed in [2]. These operators can be regarded as an algebraic abstraction representing the integral operator. Many papers have been devoted to various aspects of Rota-Baxter operators in many mathematical fields, like combinatorics [3], renormalization in quantum field theory [4], multiple zeta values in number theory [5], Yang Baxter equations [6], algebraic operad [7] and other papers. Rota-Baxter operators with arbitrary weight (also called weighted Rota-Baxter operators) was considered in [8][10].
Deformation theory of some algebraic structure goes back to Gerstenhaber [11] for associative algebras and Nijenhuis-Richardson [12] for Lie algebras. There are some advancements in deformation theory and cohomology theory of weighted Rota-Baxter algebras [13], [14]. More precisely, they considered weighted Rota-Baxter Lie algebras and associative algebras and define cohomology of them with coefficients in arbitrary Rota-Baxter representation.
Derivations have an important role to study many algebraic structures. Homotopy Lie algebras [15] and differential Galois theory [16] can be gained from derivations. Derivations also are important in control theory and gauge theories in quantum field theory [17], [18]. Recently, the cohomologies, extensions and deformations of Lie algebras with derivations (called LieDer pairs) were investigated in [19]. Then had been extended to associative algebras, Leibniz algebras, Lie triple systems, n-Lie algebras and compatible Lie algebras with derivations in [20][25]. Also derivations play new role in twisting associative and nonassociative algebras to obtain InvDer algebraic structure, for more details see [26]
Motivated by these works, we are generalizing the structre of Lie algebras with derivation [19] (respectively associative algebras with derivation [21]) and the structure of weighted Rota-Baxter Lie algebras [13] (respectively Rota-Baxter associative algebras [14]) to the cohomologies of weighted Rota-Baxter LieDer and AssDer pairs.
The paper is organized as follows. In section 2, we introduce a notion of a weighted Rota-Baxter LieDer pair and its representation. In section 3, we study cohomologies of weighted Rota-Baxter LieDer pairs and weighted Rota-baxter AssDer pairs. In section 4, we study deformation of two structures.

2 Weighted Rota-Baxter LieDer pairs↩︎

Let \(\mathfrak{L}=(L,[\cdot,\cdot])\) be a Lie algebra. A linear map \(R:L\rightarrow L\) is said to be a \(\lambda\)-weighted Rota-Baxter operator if \(R\) satisfies, for \(\lambda \in \mathbb{K}\) \[\label{RBO32eq1} [Rx,Ry]=R([Rx,y]+[x,Ry]+\lambda [x,y]),\quad \forall x,y \in L.\tag{1}\] According to the previous operator, in this section we introduce the notion of \(\lambda\)-weighted Rota-Baxter LieDer pairs (or simply weighted Rota-Baxter LieDer pairs if there is no confusion) which is a LieDer pair \((\mathfrak{L},\delta)\) equipped with a \(\lambda\)-weighted Rota-Baxter operator.
This notion is a generalization of weighted Rota-Baxter Lie algebra \((\mathfrak{L},R)\), see [13], [27][29] for more details, consisting of a Lie algebra \(\mathfrak{L}\) together with a \(\lambda\)-weighted Rota-Baxter operator on it.

2.1 Lie algebras equipped with a couple of derivations↩︎

In this subsection we study the structure of a Lie algebras \(\mathfrak{L}\) equipped with a couple of derivations \(\delta_1,\delta_2\) and we investigate its relation with the LieDer pairs. Recall first the definition of LieDer pair.

Definition 1. [19] A LieDer pair \((\mathfrak{L},\delta)\) is a Lie algebra \(\mathfrak{L}\) equipped with a derivation \(\delta\).

So a Lie algebra equipped with a derivation leads to construct a LieDer pair. Next we combine a Lie algebra \(\mathfrak{L}\) with two derivations \(\delta_1\) and \(\delta_2\) to construct a LieDer pair structure with an additionally condition.

Remark 1. Let \(\mathfrak{L}\) be a Lie algebra and \(\delta_1,\delta_2:L\rightarrow L\) be two derivations on \(\mathfrak{L}\). Then \(\delta_1\circ\delta_2\) is a derivation on \(\mathfrak{L}\) if and only if for all \(x,y\in L\) the following condition holds \[\label{condition32biderivation} [\delta_1x,\delta_2y]=[\delta_1y,\delta_2x]\tag{2}\]

Definition 2. Let \(\mathfrak{L}\) be a Lie algebra, \(\delta_1\) and \(\delta_2\) be two derivations on it. Then \((\mathfrak{L},\delta_1,\delta_2)\) is called a Lie-BiDer pair if \((\mathfrak{L},\delta_1\circ\delta_2)\) is a LieDer pair.

The previous definition means that \((\mathfrak{L},\delta_1,\delta_2)\) is a Lie-BiDer pair if and only if the condition 2 is satisfied.

Example 1. Let \(L\) be two dimentional Lie algebra such that \([e_1,e_2]=e_2\). Then \((\mathfrak{L},\delta_1,\delta_2)\) is a Lie-BiDer pair where, for all \(a,b,c,d\in \mathbb{K}\) \[\delta_1=\begin{pmatrix} 0 & 0 \\ a & b \end{pmatrix}\quad \text{ and }\quad \delta_2=\begin{pmatrix} 0 & 0 \\ c & d \end{pmatrix} \quad \text{ with }\quad ad=bc\]

Definition 3. Let \((\mathfrak{L},\delta_1,\delta_2)\) and \((\mathfrak{K},\partial_1,\partial_2)\) be two Lie-BiDer pairs. A Lie-BiDer homomorphism \(f:\mathfrak{L} \rightarrow \mathfrak{K}\) is a Lie algebra homomorphism from \(\mathfrak{L}\) to \(\mathfrak{K}\) such that \[\label{morphism32Lie-BiDer32pair} f\circ\delta_1=\partial_1\circ f,\quad f\circ\delta_2=\partial_2\circ f\tag{3}\]

Definition 4. Let \((\mathfrak{L},\delta_1,\delta_2)\) be a Lie-BiDer pair. A representation of it on a vector space \(V\) with respect to \((\varphi_1,\varphi_2)\in \mathrm{gl}(V)\) is a Lie algebra morphism \(\rho:L\rightarrow \mathrm{gl}(V)\) such that, for \(x\in L\), we have \[\label{representation32Lie-BiDer32pair} \rho(\delta_1x)\circ\varphi_2=-\rho(\delta_2x)\circ\varphi_1.\tag{4}\] Such a representation is denoted by \((\mathcal{V}=(V,\rho),\varphi_1,\varphi_2)\).

Example 2. Let \(\mathfrak{L}\) be a Lie algebra and a map \(\mathrm{ad}_x:L \rightarrow L\) defined by \[\mathrm{ad}_x(y)=[x,y],\quad \forall y\in L.\] Then \((L,\mathrm{ad},\delta_1,\delta_2)\) is a representation of the Lie-BiDer pair \((L,\delta_1,\delta_2)\) on \(L\) with respect to \(\delta_1\) and \(\delta_2\), it is called the adjoint representation.

Proposition 3. Let \((\mathfrak{L},\delta_1,\delta_2)\) be a Lie-BiDer pair and \((\mathcal{V},\varphi_1,\varphi_2)\) be a representation of it. Then \((L\oplus V,\delta_1\oplus\varphi_1,\delta_2\oplus\varphi_2)\) is a Lie-BiDer pair with the following Lie structure \[_{\ltimes}=[x,y]+\rho(x)b-\rho(y)a;\quad x,y \in L \text{ and }a,b \in V.\] and the maps \(\delta_\mathrm{i}\oplus\varphi_\mathrm{i}\) are given by \[\delta_\mathrm{i}\oplus\varphi_\mathrm{i}(x+a):=\delta_\mathrm{i}x+\varphi_\mathrm{i}(a),\quad \forall i\in \{1,2\}.\]

Such a Lie-BiDer pair is called the semi-direct product of \((\mathfrak{L},\delta_1,\delta_2)\) by a representation representation \((\mathcal{V},\varphi_1,\varphi_2)\) and it is denoted by \(L\ltimes_{\boldsymbol{BiDer}} V\).

Proof. We need just to prove that the condition 2 is satisfied for semi-direct product. \[\begin{align} _\ltimes&=[\delta_1x+\varphi_1(a),\delta_2y+\varphi_2(b)]_\ltimes\\ &=[\delta_1x,\delta_2y]+\rho(\delta_1x)\varphi_2(b)-\rho(\delta_2y)\varphi_1(a)\\ &\overset{\ref{condition32biderivation}}{=}[\delta_1y,\delta_2x]+\rho(\delta_1x)\varphi_2(b)-\rho(\delta_2y)\varphi_1(a)\\ &\overset{\ref{representation32Lie-BiDer32pair}}{=}[\delta_1y,\delta_2x]-\rho(\delta_2x)\varphi_1(b)+\rho(\delta_1y)\varphi_2(a)\\ &=[(\delta_1\oplus\varphi_1)(y+b),(\delta_2\oplus\varphi_2)(x+a)]_\ltimes \end{align}\] This complete the proof. ◻

2.2 Weighted Rota-Baxter LieDer pairs↩︎

In this subsection, we introduce a notion of \(\lambda\)-weighted Rota-Baxter LieDer pair (or simply weighted Rota-Baxter LieDer pair) and some basic definitions. Denote the Lie algebra \((L,[\cdot,\cdot])\) by \(\mathfrak{L}\) and its representation on a vector space \(V\) by \(\mathcal{V}=(V;\rho)\).
A weighted Rota-Baxter Lie algebra consists of a Lie algebra \(\mathfrak{L}\) equipped with a Rota-Baxter operator of weight \(\lambda\) denoted by \(R\).
Inspired by the notion of LieDer pair [19] and the definition of weighted Rota-Baxter Lie algebra [28] we introduce the following.

Definition 5. A weighted Rota-Baxter LieDer pair consists of a LieDer pair \((\mathfrak{L},\delta)\) together with a \(\lambda\)-weighted Rota-Baxter operator \(R\) such that \[\label{condition132RBLieDer32pair} R\circ \delta=\delta\circ R.\tag{5}\]

Example 4. Let \(\{e_1,e_2\}\) be a basis of a \(2\)-dimensional vector space \(L\) over \(\mathbb{K}\). Given a Lie structure \([e_1,e_2]=e_2\), then the triple \((\mathfrak{L},\delta,R)\) is a \(\lambda\)-weighted Rota-Baxter LieDer pair with \[\begin{align} \delta=\begin{pmatrix} 0 & 0 \\ 0 & a \end{pmatrix}\quad \text{ and } \quad R=\begin{pmatrix} 0 & 0 \\ 0 & b \end{pmatrix},\quad \forall a,b \in \mathbb{K}. \end{align}\]

Definition 6. A \(\lambda\)-weighted Rota-Baxter LieDer pair morphism from \((L_1,\delta_1,R_1)\) to \((L_2,\delta_2,R_2)\) is a Lie algebra morphism \(\varphi:L_1\rightarrow L_2\) such that the following identities holds, for all \(x,y\in L_1\) \[\begin{align} \varphi\circ \delta_1&=&\delta_2\circ\varphi,\tag{6}\\ \varphi\circ R_1&=&R_2\circ\varphi.\tag{7} \end{align}\]

Let \((\mathfrak{L},\delta)\) be a LieDer pair. Recall that a representation of it is a vector space \(V\) with two linear maps \(\rho:L\rightarrow gl(V)\) and \(\delta_\mathrm{V}:L\rightarrow L\) such that, forall \(x,y\in L\) \[\begin{align} \rho([x,y])&=&\rho(x)\circ \rho(y)-\rho(y)\circ \rho(x),\\ \delta_\mathrm{V}\circ \rho(x)&=&\rho(\delta x)+\rho(x)\circ \delta_\mathrm{V}. \end{align}\]

Definition 7. Let \((\mathfrak{L},\delta,R)\) be a weighted Rota-Baxter LieDer pair. A representation of it is a triple \((\mathcal{V},\delta_\mathrm{V},T)\) where \(T:V\rightarrow V\) is a linear map such that for all \(x\in L\) and \(u\in V\) \[\begin{align} \rho([x,y])&=&\rho(x)\circ \rho(y)-\rho(y)\circ \rho(x),\tag{8} \\ \delta_V\circ \rho(x)&=&\rho(\delta x)+\rho(x)\circ \delta_V,\tag{9}\\ \rho(Rx)(Tu)&=&T(\rho(Rx)(u)+\rho(x)(Tu)+\lambda \rho(x)u),\tag{10}\\ T\circ\delta_V&=&\delta_V\circ T. \tag{11} \end{align}\]

Example 5. Let \((L,\delta,R)\) be a \(\lambda\)-weighted Rota-Baxter LieDer pair and \((\mathcal{V},\delta_\mathrm{V},T)\) be a representation of it. Then for any scalar \(\mu\in \mathbb{K}\), the triple \((\mathcal{V},\delta_\mathrm{V},\mu T)\) is a representation of the \((\mu\lambda)\)-weighted Rota-Baxter LieDer pair \((L,\delta_\mathrm{V},\mu R)\).

Example 6. Let \((L,\delta,R)\) be a \(\lambda\)-weighted Rota-Baxter LieDer pair and \((\mathcal{V},\delta_\mathrm{V},T)\) be a representation of it. Then the quadruple \((\mathcal{V},\delta_\mathrm{V},-\lambda \mathrm{Id}_\mathrm{V}-T)\) is a representation of the \(\lambda\)-weighted Rota-Baxter LieDer pair \((L,\delta,-\lambda\mathrm{Id}_\mathrm{L}-R)\).

Example 7. Any \(\lambda\)-weighted Rota-Baxter LieDer pair \((L,\delta,R)\) is a representation of itself. Such a representation is called the adjoint representation.

Next, we construct the semi-direct product in the context of \(\lambda\)-weighted Rota-Baxter LieDer pair.

Proposition 8. Let \((\mathfrak{L},\delta,R)\) be a weighted Rota-Baxter LieDer pair and \((\mathcal{V},\delta_\mathrm{V},T)\) be a representation of it. Then \((L\oplus V,\delta\oplus \delta_{\mathrm{V}},R\oplus T)\) is a weighted Rota-Baxter LieDer pair where the Lie bracket on \(L\oplus V\) is given by \[_\ltimes:=[x,y]+\rho(x)b-\rho(y)a,\] and the derivation is given by \[\begin{align} \delta\oplus\delta_\mathrm{V}(x+a)=\delta x+\delta_\mathrm{V}a \end{align}\] and The \(\lambda\)-weighted Rota-Baxter LieDer pair is given by \[\begin{align} R\oplus T(x+a)=R x+Ta,\quad \forall x,y\in L \quad \forall a,b\in V. \end{align}\] We call such structure by the semi-direct product of the \(\lambda\)-weighted Rota-Baxter LieDer pair \((L,\delta,R)\) by a representation of it \((\mathcal{V},\delta_\mathrm{V},T)\) and denoted by \(L\ltimes_{\mathrm{R.B.LieDer}}V\).

Proof. According to the [28] and [19] the proof is straightforward. The idea is to show that \(\delta\oplus \delta_{\mathrm{V}}\) is a derivation on \(L\oplus V\) and to show that \(R\oplus T\) is a weighted Rota-Baxter operator. ◻

Proposition 9. Let \((\mathfrak{L},\delta,R)\) be a weighted Rota-Baxter LieDer pair. Then we have the triple \((L,[\cdot,\cdot]_\mathrm{R},\delta)\) is a LieDer pair with \[\label{induced32RBLieDer} [x,y]_R:=[Rx,y]+[x,Ry]+\lambda [x,y]\qquad{(1)}\] and it is denoted simply by \((\mathfrak{L}_\mathrm{R},\delta)\).

Proof. Let \(x,y\in L\) \[\begin{align} \delta([x,y]_R)&=\delta([Rx,y]+[x,Ry]+\lambda [x,y])\\ &=\delta([Rx,y])+\delta ([x,Ry]) +\lambda \delta([x,y])\\ &=[\delta\circ Rx,y]+[Rx,\delta y]+[\delta x,Ry]+[x,\delta\circ R]+\lambda [\delta x,y]+\lambda [x,\delta y]\\ &\overset{\eqref{condition132RBLieDer32pair}}{=}[R\circ \delta x,y]+[Rx,\delta y]+[\delta x,Ry]+[x,R\circ \delta]+\lambda [\delta x,y]+\lambda [x,\delta y]\\ &=\Big([R\circ \delta x,y]+ [\delta x,Ry]+\lambda [\delta x,y]\Big)+\Big([Rx,\delta y]+[x,R\circ \delta]+\lambda [x,\delta y] \Big)\\ &=[\delta x,y]_R+[x,\delta y]_R. \end{align}\] This complete the proof. ◻

Proposition 10. The triple \((\mathfrak{L}_R,\delta,R)\) is a weighted Rota-Baxter LieDer pair and the map \(R:L_R\rightarrow L\) is a morphism of weighted Rota-Baxter LieDer pair.

Proof. We have that \(R\) is a \(\lambda\)-weighted Rota-Baxter operator on L, it follows then from 1 that \[R([x,y]_R)=[Rx,Ry],\quad \forall x,y\in L.\] This complete the proof. ◻

Theorem 11. Let \((\mathfrak{L},\delta,R)\) be a weighted Rota-Baxter LieDer pair and \((\mathcal{V},\delta_{\mathrm{V}},T)\) be a representation of it. Define a map \[\label{rep32of32new32Rota-Baxter32LieDer32pairs} \widetilde{\rho}(x)(a)=\rho(Rx)(a)-T(\rho(x)(a)),\quad \text{for }x\in L , a\in V\qquad{(2)}\] Then \(\widetilde{\rho}\) defines a representation of the LieDer pair \((L_{\mathrm{R}},\delta)\) on \((\widehat{\mathcal{V}},\delta_{\mathrm{V}})=((V;\widetilde{\rho}),\delta_{\mathrm{V}})\) if and only if the too conditions 5 and 11 are satisfied. Moreover, \((\widehat{\mathcal{V}},\delta_{\mathrm{V}},T)\) is a representation of the weighted Rota-Baxter LieDer pair \((L_{\mathrm{R}},\delta,R)\).

Proof. We have already, from [28], that \(\widetilde{\rho}\) is a representation of \(L_\mathrm{R}\) on \(V\) in the context of Lie algebra structure. So we need just to prove that equation 9 holds for the representation \(\widetilde{\rho}\). Let \(x\in L\) and \(a\in V\) \[\begin{align} \delta_{\mathrm{V}}\circ \widetilde{\rho}(x)a&=\delta_{\mathrm{V}}\circ (\rho(Rx)a-T(\rho(x)a))\\ &=\delta_{\mathrm{V}}\circ \rho(Rx)a-\delta_{\mathrm{V}}(T(\rho(x)a))\\ &\overset{\ref{Rep32of32RBDer32pair1}}{=}\rho(\delta\circ Rx)a+\rho(Rx)\circ \delta_{\mathrm{V}}a-\delta_{\mathrm{V}}(T(\rho(x)a))\\ &\overset{\ref{Rep32of32RBDer32pair3}}{=}\rho(\delta\circ Rx)a+\rho(Rx)\circ \delta_{\mathrm{V}}a-T(\delta_{\mathrm{V}}(\rho(x)a))\\ &\overset{\ref{Rep32of32RBDer32pair1}}{=}\rho(\delta\circ Rx)a+\rho(Rx)\circ \delta_{\mathrm{V}}a-T(\rho(\delta x)a+\rho(x)\circ \delta_{\mathrm{V}}a)\\ &\overset{\ref{condition132RBLieDer32pair}}{=}\rho(R\circ\delta x)a-T(\rho(\delta x)a)+\rho(Rx)\circ \delta_{\mathrm{V}}a-T(\rho(x)\circ \delta_{\mathrm{V}}a)\\ &=\widetilde{\rho}(\delta x)a+\widetilde{\rho}(x)\circ \delta_{\mathrm{V}}a. \end{align}\] This means that \((\widetilde{\mathcal{V}},\delta_{\mathrm{V}})\) is a representation of the LieDer pair \((L_\mathrm{R},\delta)\). For the next result see (Theorem 2.17 in [28] ). ◻

3 Cohomology of weighted Rota-Baxter LieDer and AssDer pairs↩︎

In this section we introduce the cohomology of weighted Rota-Baxter LieDer and AssDer pairs.

3.1 Cohomology of weighted Rota-Baxter LieDer pairs↩︎

In this subsection, we first recall the Chevally-Eilenberg cohomology of Lie algebras, the cohomology of weighted Rota-Baxter Lie algebras [28] and the cohomology of LieDer pairs [19] with coefficients in an arbitrary representation. Then we define the cohomology of \(\lambda\)-weighted Rota-Baxter LieDer pairs.
Let \(\mathfrak{L}=(L,[\cdot,\cdot])\) be a Lie algebra, the Chevally-Eilenberg cohomology of \(\mathfrak{L}\) with cofficents in the representation \(\mathcal{V}\) is given by the cohomology of the cochain complex \(\{\mathrm{C}^{\star}(L;\mathcal{V}),\mathrm{d}\}\) where \(\mathrm{C}^{n}(L;\mathcal{V})=\mathrm{Hom}(\wedge^nL,V)\) for \(n\geq0\) and the coboundary map \[\mathrm {d}:C^n(L;\mathcal{V})\rightarrow C^{n+1}(L;\mathcal{V})\] is given by \[\begin{align} (\mathrm {d}(f_n))(x_1,\ldots,x_{n+1})&=\displaystyle\sum_{i=1}^{n+1}(-1)^{i+n}\rho(x_i)f_n(x_1,\ldots,\hat{x_i},\ldots,x_{n+1})\\ &+\displaystyle\sum_{1\leq i<j\leq n+1}(-1)^{i+j+n+1}f_n([x_i,x_j],x_1,\ldots,\hat{x_i},\ldots,\hat{x_j},\ldots,x_{n+1}) \end{align}\] for \(f_n\in C^n({L;\mathcal{V}})\) and \(x_1,\ldots,x_{n+1}\in L\).
Let \((\mathfrak{L},\delta)\) be a LieDer pair. Recall that the cohomology of LieDer pair \((\mathfrak{L},\delta)\) with coefficents in a representation \((\mathcal{V},\delta_{\mathrm{V}})\) is given as follow:
The set of LieDer pair \(0\)-cochains is \(0\) and the set of LieDer pair \(1\)-cochains is \(\mathfrak{C}^1_{\mathrm{LieDer}}(L;\mathcal{V})=\mathrm{Hom}(L,V)\). For \(n\geq2\), the set of LieDer pair \(n\)-cochains is given by \[\mathfrak{C}^n_{\mathrm{LieDer}}(L;\mathcal{V}):=C^n(L;\mathcal{V})\times C^{n-1}(L;\mathcal{V})\] For \(n\geq1\), define the following operator \(\partial:C^n(L;\mathcal{V})\rightarrow C^n(L;\mathcal{V})\) by \[\partial f_n=\displaystyle\sum_{i=1}^nf_n\circ (\mathbf{1}\otimes \cdots\otimes \underbrace{\delta}_{\text{i-th place}}\otimes \cdots\otimes \mathbf{1})-\delta_{\mathrm{V}}\circ f_n.\] Define \(\mathrm{D}:\mathfrak{C}^1_{\mathrm{LieDer}}(L;\mathcal{V})\rightarrow \mathfrak{C}^2_{\mathrm{LieDer}}(L;\mathcal{V})\) by \[\mathrm{D}f_1=(\mathrm {d}(f_1),(-1)^1\partial f_1),\quad \forall f_1\in \mathrm{Hom}(L,V).\] And for \(n\geq2\), define \(\mathrm{D}:\mathfrak{C}^n_{\mathrm{LieDer}}(L;\mathcal{V})\rightarrow \mathfrak{C}^{n+1}_{\mathrm{LieDer}}(L;\mathcal{V})\) by \[\mathrm{D}(f_n,g_{n-1})=(\mathrm{d}(f_n),\mathrm{d}(g_{n-1})+(-1)^n\partial f_n)\] for all \(f_n\in C^n(L;\mathcal{V})\) and \(g_{n-1}\in C^{n-1}(L;\mathcal{V})\).
Recall also the following equation \[\label{coboundary1} \mathrm{d}\circ \partial =\partial \circ \mathrm{d}\tag{12}\]

Let \((\mathfrak{L},R)\) be a Rota-Baxter Lie algebra, the cohomology of weighted Rota-Baxter Lie algebra with coefficients in a representation \((\mathcal{V},T)\) is given as follow as follow :
For each \(n\geq 0\), define an abelian group \(\mathfrak{C}^n_{\mathrm{R}}(L;\mathcal{V})\) by \[\mathfrak{C}^n_{\mathrm{R}}(L;V)= \left\{ \begin{array}{ll} C^0(L;\mathcal{V})=V,& \text{ if } n=0,\\ C^n(L;\mathcal{V})\oplus C^{n-1}(L_{\mathrm{R}};\widetilde{\mathcal{V}})=\mathrm{Hom}(\wedge^nL,V)\otimes\mathrm{Hom}(\wedge^{n-1}L,V),& \text{ if } n\geq1. \end{array} \right.\] Where \(\mathfrak{L}_{\mathrm{R}}=(L,[\cdot,\cdot]_{\mathrm{R}},R)\) is a Rota-Baxter Lie algebra with the bracket is defined in ?? and \(\widetilde{\mathcal{V}}=(V,\widetilde{\rho})\) is a representation of it with \(\widetilde{\rho}\) is defined in ?? .
The coboundary map is defined as \(\mathfrak{d}_{\mathrm{R}}:\mathfrak{C}^n_{\mathrm{R}}(L;\mathcal{V})\rightarrow \mathfrak{C}^{n+1}_{\mathrm{R}}(L;\mathcal{V})\) by \[\left\{ \begin{array}{ll} \mathfrak{d}_{\mathrm{R}}(v)=(\mathrm{d}(v),-v),& \text{ for } v\in \mathfrak{C}^0_{\mathrm{R}}(L;\mathcal{V})=V,\\ \mathfrak{d}_{\mathrm{R}}(f_n,g_{n-1})=(\mathrm{d}(f_n),-\mathrm{d}_{\mathrm{R}}(g_{n-1})-\Phi^n(f_n)),& \text{ for } (f_n,g_{n-1})\in\mathfrak{C}^n_{\mathrm{R}}(L;\mathcal{V}) . \end{array} \right.\] With \(\mathrm{d}_{\mathrm{R}}:C^n(L_{\mathrm{R}};\widetilde{\mathcal{V}})\rightarrow C^{n+1}(L_{\mathrm{R}};\widetilde{\mathcal{V}})\) is given by \[\begin{align} (\mathrm{d}_{\mathrm{R}}f_n)(x_1,\ldots,x_{n+1})&=\displaystyle\sum_{i=1}^{n+1}(-1)^{i+n} \;\widetilde{\rho}(x_i)f_n(x_1,\ldots,\hat{x_i},\ldots,x_{n+1})\\ &+\displaystyle\sum_{1\leq i<j\leq n+1}(-1)^{i+j+n+1}f_n([x_i,x_j]_{\mathrm{R}},x_1,\ldots,\hat{x_i},\ldots,\hat{x_j},\ldots,x_{n+1}) \end{align}\] and \(\Phi:C^n(L;\mathcal{V})\rightarrow C^n(L_{\mathrm{R}};\widetilde{\mathcal{V}})\), is inspired from [14], defined by

\[\label{cohomology32of32RBO} \left\{ \begin{array}{ll} \Phi=\mathbf{1}_{\mathrm{V}},& \\ \Phi(f_n)(x_1,\cdots,x_n)=f_n(Rx_1,\cdots,Rx_n)-\displaystyle\sum_{k=0}^{n-1}\lambda^{n-k-1}\displaystyle\sum_{i_1<\ldots<i_k}T\circ f_n(x_1,\cdots,R(x_{i_1}),\cdots,R(x_{i_k}),\cdots,x_n).& \end{array} \right.\tag{13}\]

is a morphism of cochain complex from \(\{C^{\star}(L;\mathcal{V}),\mathrm{d}\}\) to \(\{C^{\star}(L_{\mathrm{R}};\widetilde{\mathcal{V}}),\mathrm{d}\}\) i.e. \[\label{coboundary2} \mathrm{d}_{\mathrm{R}}\circ\Phi=\Phi\circ \mathrm{d},\quad \forall n\geq0.\tag{14}\] Also since \(L_{\mathrm{R}}\) is a Lie algebra and \(\mathrm{d}_{\mathrm{R}}\) its coboundary with respect to the representation \(\widetilde{\mathcal{V}}=(V;\widetilde{\rho})\) and \((L_{\mathrm{R}},\delta)\) is a LieDer pair then, from the cohomology of LieDer pair ([19]. Lemma(3.1)), we get \[\label{coboundary3} \partial\circ\mathrm{d}_{\mathrm{R}}=\mathrm{d}_{\mathrm{R}}\circ\partial\tag{15}\] Using all those tools we are in position to define the cohomology of \(\lambda\)-weighted Rota-Baxter LieDer pair \((\mathfrak{L},\delta,R)\) with coefficients in a representation \((\mathcal{V},\delta_\mathrm{V},T)\). The set of \(\lambda\)-weighted Rota-Baxter LieDer pair \(0\)-cochains is \(0\) and the set of \(\lambda\)-weighted Rota-Baxter LieDer pair \(1\)-cochains to be \(\mathfrak{C}^1_{\mathrm{RBLieDer}}(L,\mathcal{V})=\mathrm{Hom}(L,V)\).
For \(n\geq2\) define \(\lambda\)-weighted Rota-Baxter LieDer pair \(n\)-cochains by \[\mathfrak{C}^n_{\mathrm{RBLieDer}}(L;\mathcal{V}):=\mathfrak{C}^n_{\mathrm{LieDer}}(L;\mathcal{V})\otimes C^{n-1}(L_R;\widetilde{\mathcal{V}})\] Define \[\mathfrak{D}_{\mathrm{RBLieDer}}:\mathfrak{C}^n_{\mathrm{RBLieDer}}(L,\mathcal{V})\rightarrow \mathfrak{C}^{n+1}_{\mathrm{RBLieDer}}(L,\mathcal{V})\] as follow

  1. For \(n=1\) \[\begin{align} \mathfrak{D}_{\mathrm{RBLieDer}}&:&\mathfrak{C}^1_{\mathrm{RBLieDer}}(L,\mathcal{V})\rightarrow \mathfrak{C}^2_{\mathrm{RBLieDer}}(L,\mathcal{V}) \text{ is given by}\\ &&\mathfrak{D}_{\mathrm{RBLieDer}}(f)=(d(f),-\partial (f),-\Phi (f)),\quad \forall f\in \mathrm{Hom}(L,V). \end{align}\]

  2. For \(n\geq2\) \[\begin{align} \mathfrak{D}_{\mathrm{RBLieDer}}&:&\mathfrak{C}^n_{\mathrm{RBLieDer}}(L,\mathcal{V})\rightarrow \mathfrak{C}^{n+1}_{\mathrm{RBLieDer}}(L,\mathcal{V}) \text{ is given by}\\ &&\mathfrak{D}_{\mathrm{RBLieDer}}((f,g),h)=(d(f),d(g)+(-1)^n\partial (f),-d_R(h)-\Phi f),\quad \forall ((f,g),h)\in \mathrm{Hom}(L,V). \end{align}\]

Next in the following theorem we are in position to prove that \(\mathfrak{D}_{\mathrm{RBLieDer}}\) is a coboundary map.

Theorem 12. The map \(\mathfrak{D}_{\mathrm{RBLieDer}}\) is a coboundary operator, means that \[\mathfrak{D}_{\mathrm{RBLieDer}}\circ \mathfrak{D}_{\mathrm{RBLieDer}}=0.\]

Proof. For \(n\geq1\), using equations 12 and 14 \[\begin{align} \mathfrak{D}_{\mathrm{RBLieDer}}\circ\mathfrak{D}_{\mathrm{RBLieDer}}((f,g),h)&=\mathfrak{D}_{\mathrm{RBLieDer}}(d(f),d(g)+(-1)^n\partial (f),-d_R(h)-\Phi f)\\ &=(d^2(f),d^2(g)+(-1)^nd\circ\partial(f)+(-1)^{n+1}(f),d^2_R(h)+d_R\circ\Phi(f)-\Phi\circ d(f))\\ &=(0,0+(-1)^nd\circ\partial(f)+(-1)^{n+1}(f),0+d_R\circ\Phi(f)-\Phi\circ d(f))\\ &=(0,0,0) \end{align}\] This complete the proof. ◻

With respect to the representation \((\mathcal{V},\delta_V,T)\) we obtain a complex \(\{\mathfrak{C}^{\star}_{\mathrm{RBLieDer}}(L,\mathcal{V}),\mathfrak{D}_{\mathrm{RBLieDer}}\}\). Let \(\mathcal{Z}^n_{\mathrm{RBLieDer}}(L,\mathcal{V})\) and \(\mathcal{B}^n_{\mathrm{RBLieDer}}(L,\mathcal{V})\) denote the space of \(n\)-cocycles and \(n\)-coboundaries, respectively. Then we define the corresponding cohomology groups by \[\mathcal{H}^n_{\mathrm{RBLieDer}}(L,\mathcal{V}):=\frac{\mathcal{Z}^n_{\mathrm{RBLieDer}}(L,\mathcal{V})}{\mathcal{B}^n_{\mathrm{RBLieDer}}(L,\mathcal{V})},\quad \text{for } n\geq0.\] They are called the cohomolgy of \(\lambda\)-weighted Rota-Baxter LieDer pair \((\mathfrak{L},\delta,R)\) with coefficients in the representation \((\mathcal{V},\delta_{\mathrm{V}},T)\).

Remark 2. Recall that from [28], if \((\mathfrak{L},R)\) is a Rota-Baxter Lie algebra and \((\mathcal{V},T)\) is a representation of it. An element \(v\in V\) is in \(\mathrm{Z}^0_{RB}(L,\mathcal{V})\) if and only if \((d_{\mathrm{R}}(v),-v)=0\), this holds when \(v=0\) which means that \(\mathrm{H}^0_{RB}(L,\mathcal{V})=0\) and it coincide in our paper with \(\mathrm{H}^0_{RBLieDer}(L,\mathcal{V})=0=\mathrm{H}^0_{RB}(L,\mathcal{V})\).
Also \(\mathrm{H}^1_{RBLieDer}(L,\mathcal{V})\) coincide with \(\mathrm{H}^1_{RB}(L,\mathcal{V})=\frac{\mathrm{Der}(L,\mathcal{V})}{\mathrm{InnDer}(L,\mathcal{V})}\).

Proposition 13. Let \((\mathcal{V},\delta_{\mathrm{V}},T)\) be a representation of a \(\lambda\)-weighted Rota-Baxter LieDer pair \((\mathfrak{L},\delta,R)\). Then we have \[\mathcal{H}^1_{\mathrm{RBLieDer}}(L,\mathcal{V})=\{f;\quad f\in\mathcal{Z}^1(L,\mathcal{V}),\quad f\circ\delta=\delta_V\circ f,\quad f\circ R=T\circ f\}.\]

3.2 Cohomology of weighted Rota-Baxter AssDer pairs↩︎

In [14] authors defined the cohomology of a weighted Rota-Baxter associative algebra with coefficients in a Rota-Baxter bimodule. Later Das.A studied the associative algebra with derivation and denote it by AssDer pairs [21].
In this subsection we study the cohomology of \(\lambda\)-weighted Rota-Baxter AssDer pair and we show that this cohomolgy is related to the cohomology of Rota-Baxter LieDer pairs by suitable skew-symmetrization.
Let \((\mathfrak{A}=(A,\mu))\) be an associative algebra and \(\delta:A\rightarrow A\) be a derivation on it. Recall that an AssDer pair \((\mathfrak{A},\delta)\) is an associative algebra equipped with the derivation \(\delta\).
A linear map \(R:A \rightarrow A\) is said to be a \(\lambda\)-weighted Rota-Baxter operator on \(\mathfrak{A}\) if it satisfies \[\label{RBO32on32Ass32Algebras} \mu(Rx,Ry)=R(\mu(Rx,y)+\mu(x,Ry)+\lambda \mu(x,y)),\quad \forall x,y\in L.\tag{16}\] A bimodule over \(\mathfrak{A}\) consists of a vector space \(M\) together with two linear maps \(l:A\otimes M\rightarrow M\) and \(r:M\otimes A\rightarrow M\) such that for all \(x,y\in A\) and \(m\in M\) \[\begin{align} l(\mu(x,y),m)&=&l(x,l(y,m)),\\ r(l(x,m),y)&=&l(x,r(m,y))\\ r(r(m,x),y)&=&r(m,\mu(x,y)). \end{align}\] We will write \(xm\) instead of \(l(x,m)\) and \(mx\) instead of \(r(m,x)\) when there are no confusions.

Definition 8. A weighted Rota-Baxter AssDer pair consisits of an AssDer pair \((\mathfrak{A},\delta)\) together with a weighted Rota-Baxter operator such that \[R\circ \delta=\delta\circ R.\]

Definition 9. Let \((\mathfrak{A}_1,\delta_1,R_1)\) and \((\mathfrak{A}_2,\delta_2,R_2)\) be two \(\lambda\)-weighted Rota-Baxter AssDer pairs. Then the linear map \(f:A_1\rightarrow A_2\) is said to be a \(\lambda\)-weighted Rota-Baxter homomorphism if is satisfies the following, for \(x,y\in A\) \[\begin{align} f\circ\mu_1(x,y)&=&\mu_2(f(x),f(y)),\\ f\circ\delta_1&=&\delta_1\circ f,\\ f\circ R_1&=&R_1\circ f. \end{align}\]

Definition 10. Let \((\mathfrak{A},\delta,R)\) be a \(\lambda\)-weighted Rota-Baxter AssDer pair. A bimodule (representation) over it is a triple \((M,\delta_\mathrm{M},T)\) which is both a left and right module on \((\mathfrak{A},\delta,R)\) and \(M\) is an \(A\)-bimodule. This means the followings, for all \(x,y\in A\) and \(m\in M\) \[\begin{align} \delta_\mathrm{M}(xm)&=&\delta(x)m+x\delta_\mathrm{M}(m),\tag{17}\\ \delta_\mathrm{M}(mx)&=&\delta_\mathrm{M}(m)x+m\delta(x)\tag{18}\\ R(x)T(m)&=&T(R(x)m+xT(m)+\lambda xm)\\ T(m)R(x)&=&T(T(m)x+mR(x)++\lambda mx)\\ \delta_\mathrm{M}\circ T&=&T\circ \delta_\mathrm{M} \tag{19}. \end{align}\]

Proposition 14. Let \((\mathfrak{A},\delta,R)\) be a \(\lambda\)-weighted Rota-Baxter AssDer pair and \((M,\delta_\mathrm{M},T)\) be a representation of it. Then \((A\oplus M,\delta\oplus\delta_\mathrm{M},R\oplus T)\) is a \(\lambda\)-weighted Rota-Baxter AssDer pair where its structure is given by for all \(x,y\in A\) and \(m,n\in M\) \[\begin{align} \mu_\ltimes(x+m,y+n)&=&\mu(x,y)+xn+my,\\ (\delta\oplus\delta_\mathrm{M})(x+m)&=&\delta(x)+\delta_\mathrm{M}(m)\\ (R\oplus T)(x+m)&=&R(x)+T(m). \end{align}\]

Proof. the proof is easy to check because it is known that \(A\oplus M\) with the product \(\mu_\ltimes\) and the linear map \(\delta\oplus\delta_\mathrm{M}\) is a LieDer pair, see [21]. And it is well know that \(R\oplus T\) is a \(\lambda\)-weighted Rota-Baxter operator. ◻

Proposition 15. Let \((\mathfrak{A},\delta,R)\) be a weighted Rota-Baxter AssDer pair. Define a new binary operation as follow : \[\mu_\mathrm{R}:=\mu(x,Ry)+\mu(Rx,y)+\lambda \mu(x,y),\quad \forall x,y\in A.\] Then

  1. The operation \(\mu_\mathrm{R}\) forms an AssDer pair \((A,\mu_\mathrm{R},\delta)\) together with the derivation \(\delta\).

  2. The quadruple \((A,\mu_\mathrm{R},\delta,R)\) forms a weighted Rota-Baxter AssDer pair and it is denoted \((\mathfrak{A}_\mathrm{R},\delta,R)\).

  3. The map \(R:(A,\mu_\mathrm{R},\delta,R)\rightarrow(A,\mu,\delta,R)\) is a morphism of weighted Rota-Baxter AssDer pairs.

Proof. All we need to prove is that \(\delta\) is a derivation on the operation \(\mu_\mathrm{R}\), the rest of the proof see ([27],Theorem 1.1.17).
Let \(x,y\in A\) \[\begin{align} \delta(\mu_\mathrm{R}(x,y))&=\delta(\mu(x,Ry)+\mu(Rx,y)+\lambda \mu(x,y))\\ &=\mu(\delta x,Ry)+\mu(x,\delta\circ Ry)+\mu(\delta\circ Rx,y)+\mu(Rx,\delta y)+\lambda \mu(\delta x,y)+\lambda \mu(x,\delta y)\\ &\overset{\ref{condition132RBLieDer32pair}}{=}\mu(\delta x,Ry)+\mu(R\circ \delta x,y)+\lambda \mu(\delta x,y)+\mu(x,R\circ\delta y)+\mu(Rx,\delta y)+\lambda \mu(x,\delta y)\\ &=\mu_\mathrm{R}(\delta x,y)+\mu_\mathrm{R}(x,\delta y). \end{align}\] This complete the proof. ◻

Proposition 16. Let \((\mathfrak{A},\delta,R)\) be a weighted Rota-Baxter AssDer pair and \((M,\delta_\mathrm{M},T)\) be weighted Rota-Baxter bimodule over it. Define a left action \(l_\mathrm{R}\) and a right action \(r_\mathrm{R}\) of \(A\) on \(M\) as follows, for all \(x\in A,m\in M\) \[\begin{align} l_\mathrm{R}(x,m)&=&l(Rx,m)-T(L(x,m))\\ r_\mathrm{R}(m,x)&=&r(m,Rx)-T(r(m,x)). \end{align}\] Then these actions makes \(M\) into weighted Rota-Baxter bimodule over \((\mathfrak{A}_\mathrm{R},\delta,R)\) and denote it this new bimodule by \((M_\mathrm{R},\delta_M,T)\).

Proof. It is known that \(l_\mathrm{R}\) (respectively \(r_\mathrm{R}\)) is a left action (respectively right action) of \(A\) on \(M_\mathrm{R}\). So we need just to prove equations 17 and 18 .
Let \(x\in A,m\in M\) \[\begin{align} \delta_\mathrm{M}\circ l_\mathrm{R}&=\delta_\mathrm{M}(l(R(x),m)-T(l(x,m)))\\ &\overset{\ref{condition132RBLieDer32pair}}{=}l(R\circ \delta x,m)+l(Rx,\delta_\mathrm{M}(m))-T\circ \delta_\mathrm{M}(l(x,m))\\ &\overset{\ref{AssDer32rep325}}{=}l(R\circ \delta x,m)-T(l(\delta x,m))+l(Rx,\delta_\mathrm{M}(m))-T(l(x,\delta_\mathrm{M}(m)))\\ &=l_\mathrm{R}(\delta x,m)+l_\mathrm{R}(x,\delta_\mathrm{M}(m)) \end{align}\]Similarly to \(r_\mathrm{R}\). ◻

Next we will define a cohomology for weighted Rota-Baxter AssDer pairs.
Let \((\mathfrak{A},\delta)\) be an AssDer pair and \((M,\delta_M)\) be a representation of it. Recall from [14], the Hochshild cochain complex of \(\mathfrak{A}_\mathrm{R}\) with coefficients in \(M_\mathrm{R}\) is given as follow \[\mathrm{d_{R.Hoch}}:C^n(\mathfrak{A}_\mathrm{R},M_\mathrm{R})\rightarrow C^{n+1}(\mathfrak{A}_\mathrm{R},M_\mathrm{R}) \text{ by }\] \[\begin{align} \mathrm{d_{R.Hoch}}(f)(x_1,\ldots,x_{n+1})&= l_\mathrm{R}(x_1,f(x_2,\ldots,x_{n+1}))+\displaystyle\sum_{i=1}^n(-1)^if(x_1,\ldots,x_{i-1},\mu_\mathrm{R}(x_i,x_{i+1}),\ldots,x_{n+1})\\ &+(-1)^{n+1}r_\mathrm{R}(f(x_1,\ldots,x_n),x_{n+1}). \end{align}\] Define \(\partial\) the coboundary operator from \(C^n(A,M)\) to \(C^n(A,M)\) and it is given by \[\partial (f)=\displaystyle\sum_{i=1}^n f\circ (\mathbf{1}\otimes \ldots \delta \otimes \ldots \otimes \mathbf{1})-\delta_\mathrm{M}\circ f.\] First step let’s define the cohomology, in context of AssDer pair structure, of the weighted Rota-Baxter operator \(R\) with coefficients in the representation \((M_\mathrm{R},\delta_\mathrm{M},T)\).
Consider the cochain complex of the AssDer pair \((\mathfrak{A}_\mathrm{R},\delta)\) with coefficients in the representation \((M_\mathrm{R},\delta_M)\) \[\mathfrak{C}_\mathrm{AssDer}^\star(\mathfrak{A}_\mathrm{R},M_\mathrm{R}):=\oplus_{n\geq0}\mathfrak{C}_\mathrm{AssDer}^n(\mathfrak{A}_\mathrm{R},M_\mathrm{R})\] The space \(\mathfrak{C}_\mathrm{AssDer}^n(\mathfrak{A}_\mathrm{R},M_\mathrm{R})\) of \(n\)-cochains is defined as follow \[\left\{ \begin{array}{ll} \mathfrak{C}_\mathrm{AssDer}^0(\mathfrak{A}_\mathrm{R},M_\mathrm{R})=0,&\\ \mathfrak{C}_\mathrm{AssDer}^1(\mathfrak{A}_\mathrm{R},M_\mathrm{R})=\mathrm{Hom}(\mathfrak{A}_\mathrm{R},M_\mathrm{R}),&\\ \mathfrak{C}_\mathrm{AssDer}^n(\mathfrak{A}_\mathrm{R},M_\mathrm{R})=C^n(\mathfrak{A}_\mathrm{R},M_\mathrm{R})\oplus C^{n-1}(\mathfrak{A}_\mathrm{R},M_\mathrm{R}),& \text{for } n\geq2. \end{array} \right.\] With the coboundary map is given by \(\mathfrak{d}_{R.A.D}:\mathfrak{C}_\mathrm{R.AssDer}^n(\mathfrak{A}_\mathrm{R},M_\mathrm{R})\rightarrow \mathfrak{C}_\mathrm{R.AssDer}^{n+1}(\mathfrak{A}_\mathrm{R},M_\mathrm{R})\) as follow

\[\label{coboundary32AssDer32of32d46R} \mathfrak{d}_{R.A.D}(f,g)=(d_\mathrm{R.Hoch}(f),d_\mathrm{R.Hoch}(g)+(-1)^n\partial (f)),\quad \forall (f,g)\in \mathfrak{C}_\mathrm{R.AssDer}^n(\mathfrak{A}_\mathrm{R},M_\mathrm{R}).\tag{20}\]

By [19] and since \(d_\mathrm{R.Hoch}\) is Hochschild coboundary [14], it is easy to check that \[\begin{align} \mathfrak{d}_{R.A.D}\circ \mathfrak{d}_{R.A.D}&=&0,\\ d_\mathrm{R.Hoch}\circ \partial&=&\partial \circ d_\mathrm{R.Hoch}. \end{align}\] with the previous result we are able to define the cohomology of Rota-Baxter operator \(R\) with coefficients in \((M_\mathrm{R},\delta_\mathrm{M},T)\) in the context of AssDer pair structures.

Definition 11. Let \((\mathfrak{A},\delta,R)\) be a weighted Rota-Baxter AssDer pair and \((M,\delta_\mathrm{M},T)\) be a representation of it. Then the cocchain complex \((\mathfrak{C}_\mathrm{R.B.AssDer}(\mathfrak{A}_\mathrm{R},M_\mathrm{R}),\mathfrak{d}_{R.A.D})\) is called the cochain complex of weighted Rota-Baxter operator \(R\) with coefficients in \((M,\delta_\mathrm{M},T)\), denoted by \(\mathfrak{C}_\mathrm{R.AssDer}^\star(A,M)\) and its comology is denoted by \(\mathcal{H}_\mathrm{R.B.AssDer}(A,M)\), are called the cohomology of weighted Rota-Baxter operator \(R\) with coefficients in \((M,\delta_\mathrm{M},T)\).

Remark 3. When \((M,\delta_\mathrm{M},T)=(A,\delta,R)\) is the adjoint representation, then we denote the cochain complex of weighted Rota-Baxter operator \(R\) by \(\mathfrak{C}^\star_\mathrm{R.B.AssDer}(A)\) and its cohomology groups are denoted simply by \(\mathcal{H}^\star_\mathrm{R.B.AssDer}(A)\).

At the final step, we will combine the cohomology of AssDer pairs with the cohomology of weighted Rota-Bxater operators to define a cohomology theory for weighted Rota-Baxter AssDer pair.
Define the map \(\Phi^\star:C^\star(A,M)\rightarrow C^\star(\mathfrak{A}_\mathrm{R},M_\mathrm{R})\) by
\(\Phi^0=Id_\mathrm{M}\) and for \(n\geq1\) and \(f\in C^n(A,M)\) define \(\Phi^n(f)\in C^n(\mathfrak{A}_\mathrm{R},M_\mathrm{R})\) as follow \[\Phi^n(f)(x_1,\cdots,x_n)=f(Rx_1,\cdots,Rx_n)-\displaystyle\sum_{k=0}^{n-1}\lambda^{n-k-1}\displaystyle\sum_{i_1<\ldots<i_k}T\circ f(x_1,\cdots,R(x_{i_1}),\cdots,R(x_{i_k}),\cdots,x_n)\] From ([14], proposition 5.2), the map \(\Phi^\star\) is a chain map.

Definition 12. Let \((M,\delta_\mathrm{M},T)\) be a representation of a weighted Rota-Baxter AssDer pair \((\mathfrak{A},\delta,R)\). Define the cochain complex \((\mathfrak{C}^\star_\mathrm{R.B.AssDer}(A,M),\mathfrak{D}_\mathrm{R.B.AssDer})\) of weighted Rota-Baxter AssDer pair \((\mathfrak{A},\delta,R)\) with coeffecients in \((M,\delta_\mathrm{M},T)\) such that \[\mathfrak{C}^0_\mathrm{R.B.AssDer}(A,M)=0 \text{ and } \mathfrak{C}^n_\mathrm{R.B.AssDer}(A,M):=\mathfrak{C}^n_\mathrm{AssDer}(A,M)\oplus C^{n-1}(\mathfrak{A}_\mathrm{R},M_\mathrm{R}),\quad \forall n\geq1.\] With \(\mathfrak{C}^n_\mathrm{R.B.AssDer}(A,M)=(C^n(A,M)\oplus C^{n-1}(A,M))\oplus C^{n-1}(\mathfrak{A}_\mathrm{R},M_\mathrm{R})\).
And the differential \[\mathfrak{D}_\mathrm{R.B.AssDer}:\mathfrak{C}^n_\mathrm{R.B.AssDer}(A,M)\rightarrow \mathfrak{C}^{n+1}_\mathrm{R.B.AssDer}(A,M) \text{ is given by }\] \[\mathfrak{D}_\mathrm{R.B.AssDer}((f,g),h):=(d_\mathrm{Hoch}(f),d_\mathrm{Hoch}(g)+(-1)^n\partial f,-d_\mathrm{R.Hoch}(h)-\Phi^n(f)).\] Where \(d_\mathrm{Hoch}:C^n(A,M)\rightarrow C^{n+1}(A,M)\) is the coboundary map associated to the Hochschild cohomology [30].

The cohomology of weighted Rota-Baxter AssDer pair \((\mathfrak{C}^\star_\mathrm{R.B.AssDer}(\mathfrak{A}_\mathrm{R},M_\mathrm{R}),\mathfrak{D}_\mathrm{R.B.AssDer})\) of \((\mathfrak{A},\delta,R)\) with coefficients in \((M,\delta_\mathrm{M},T)\) is denoted by \(\mathcal{H}^\star_\mathrm{R.B.AssDer}(A)\)

3.3 Relation between cohomology of weighted Rota-Baxter AssDer and LieDer pairs↩︎

In this section, we show that the cohomology of Rota-Baxter LieDer pair is related to the cohomology of Rota-Baxter AssDer pairs via a suitable skew-symmetrization.

Proposition 17. Let \((\mathfrak{A},\delta,R)\) be a weighted Rota-Baxter AssDer pair. Then \((\mathfrak{A}_\mathrm{c},\delta,R)\) is a weighted Rota-Baxter LieDer pair, where \(\mathfrak{A}_\mathrm{c}=(A,[\cdot,\cdot]_\mathrm{c})\) such that \[_\mathrm{c}=\mu(x,y)-\mu(y,x),\quad \forall x,y \in A.\] (is called the skew-symmetrization).

Remark 4. Moreover, if \((M,\delta_\mathrm{M},T)\) is a representation of \((\mathfrak{A},\delta,R)\) then \((M_\mathrm{c},\delta_\mathrm{M},T)\) is a representation of the weighted Rota-Baxter LieDer pair \((\mathfrak{A}_\mathrm{c},\delta,R)\) with the representation is given as follow \[\rho(x)(m)=l(x,m)-r(m,x),\quad \forall x\in A, m\in M.\]

It is known that the standard skew-symmetrization gives rise to a morphism from the Hochschild cochain complex of an associative algebra to the Chevalley-Eilenberg cochain complex of the corresponding skewsymmetrized Lie algebra. Means that there is a morphism \(S_n\) from the \(n\)-th cochain group (in context of Hochschild cohomology) \(C^n(\mathfrak{A},M)\) to the \(n\)-th cochain group \(C^n(\mathfrak{A}_\mathrm{c},M_\mathrm{c})\) (in the context of Chevalley-Eilenberg cohomology), more general \(S_\star\) are called the skew-symmetrization maps.

Theorem 18. Let \((\mathfrak{A},\delta,R)\) be a weighted Rota-Baxter AssDer pair and \((M,\delta_\mathrm{M},T)\) be a representation of it. Then the collection of maps \[\mathcal{S}_\mathrm{n}:\mathfrak{C}^n_\mathrm{R.B.AssDer}(\mathfrak{A},M)\rightarrow \mathfrak{C}^n_\mathrm{R.B.LieDer}(\mathfrak{A}_\mathrm{c},M_\mathrm{c}),\quad \mathcal{S}_\mathrm{n}=(S_n,S_{n-1},S_{n-1})\] induces a morphism from the cohomology of \((\mathfrak{A},\delta,R)\) with coefficients in \((M,\delta_\mathrm{M},T)\) to the cohomology of \((\mathfrak{A}_\mathrm{c},\delta,R)\) with coefficients in \((M_\mathrm{c},\delta_\mathrm{M},T)\).

Proof. For \(((f,g),h)\in \mathfrak{C}^n_\mathrm{R.B.AssDer}(\mathfrak{A},M)\) we have \[\begin{align} \mathfrak{D}_\mathrm{R.B.LieDer}\circ \mathcal{S}_\mathrm{n} ((f,g),h)&=\mathfrak{D}_\mathrm{R.B.LieDer}(S_n(f),S_{n-1}(g),S_{n-1}(h))\\ &=(d\circ S_n(f),d\circ S_{n-1}(g)+(-1)^n\partial \circ S_n(f),-d_\mathrm{R}\circ S_{n-1}(h)-\Phi\circ S_n(f))\\ &=(S_{n+1}\circ d_\mathrm{Hoch}(f),S_n\circ d_\mathrm{Hoch}(g)+(-1)^nS_{n+1}\circ \partial (f),-S_n\circ d_\mathrm{R.Hoch}(h)-S_{n+1}\circ \Phi^n(f))\\ &=\mathcal{S}_{n+1}\circ \mathfrak{D}_\mathrm{R.B.AssDer}((f,g),h) \end{align}\] ◻

4 Aplication of cohomology: Formal deformations↩︎

In this section, we will study formal deformation of weighted Rota-Baxter LieDer and AssDer pairs.

4.1 Formal deformation of weighted Rota-Baxter LieDer pairs↩︎

In this subsection we will deform the Lie bracket on \(L\), the Rota-Baxter operator \(R\) and the derivation \(\delta\). We investigate the relation between such deformation and cohomology of \((L,\delta,R)\) with coefficients in the adjoint representation.
Let \((\mathfrak{L},\delta,R)\) be a Rota-Baxter LieDer pair, let \(\gamma\in C^2(L,L)=\mathrm{Hom}(\wedge^2L,L)\) be the element that correspponds to the Lie bracket on \(L\), i.e., \(\gamma(x,y)=[x,y]\) for \(x,y\in L\). Consider the space \(L[[t]]\) of the formal power series in \(t\) with coefficients from \(L\). Then \(L[[t]]\) is a \(\mathbb{K}[[t]]\)-module.

Definition 13. A formal \(1\)-parameter deformation of \((\mathfrak{L},\delta,R)\) consists of a triple \((\gamma_\mathrm{t},\delta_\mathrm{t},R_\mathrm{t})\) of three formal power series \[\begin{align} \gamma_\mathrm{t}&=&\displaystyle\sum_{i\geq0}\gamma_\mathrm{i}t^\mathrm{i},\quad \text{where } \gamma_\mathrm{i}\in \mathrm{Hom}(\wedge^2L,L) \text{ with } \gamma_0=\gamma,\\ \delta_\mathrm{t}&=&\displaystyle\sum_{i\geq0}\delta_\mathrm{i}t^\mathrm{i},\quad \text{where } \delta_\mathrm{i}\in \mathrm{Hom}(L,L) \text{ with } \delta_0=\delta,\\ R_\mathrm{t}&=&\displaystyle\sum_{i\geq0}R_\mathrm{i}t^\mathrm{i},\quad \text{where } R_\mathrm{i}\in \mathrm{Hom}(L,L) \text{ with } R_0=R. \end{align}\] Such that the \(\mathbb{K}[[t]]\)-bilinear map \(\gamma_\mathrm{t}\) defines a Lie algebra structure on \(L[[t]]\) and the two \(\mathbb{K}[[t]]\)-linear maps \(\delta_\mathrm{t},R_\mathrm{t}:L[[t]] \rightarrow L[[t]]\) are respectively a derivation and a Rota-Baxter opperator. This means that \((L[[t]]=(L[[t]],\gamma_\mathrm{t}),\delta_\mathrm{t,R_\mathrm{t}})\) is a Rota-Baxter LieDer pair over \(\mathbb{K}[[t]]\).

Thus, \((\gamma_\mathrm{t},\delta_{\mathrm{t}},R_\mathrm{t})\) is a formal \(1\)-parameter deformation of \((\mathfrak{L},\delta,R)\) if and only if, for \(x,y,z\in L\) \[\begin{align} \circlearrowleft_{x,y,z}\gamma_\mathrm{t}(x,\gamma_\mathrm{t}(y,z))&=&0,\\ \gamma_\mathrm{t}(R_\mathrm{t}x,R_\mathrm{t}y)&=&R_\mathrm{t}(\gamma_\mathrm{t}(R_\mathrm{t}x,y)+\gamma_\mathrm{t}(x,R_{\mathrm{t}}y)+\lambda\gamma_\mathrm{t}(x,y) ),\\ \delta_{\mathrm{t}}(\gamma_\mathrm{t}(x,y))&=&\gamma_\mathrm{t}(\delta_\mathrm{t}x,y)+\gamma_\mathrm{t}(x,\delta_{\mathrm{t}}y). \end{align}\] They are equivalent to the following systems of identities, for \(n\geq0\) and for all \(x,y,z\in L\), \[\begin{align} \displaystyle\sum_{i+j=n}\gamma_\mathrm{i}(x,\gamma_\mathrm{j}(y,z))+\gamma_\mathrm{i}(y,\gamma_\mathrm{j}(z,x))+\gamma_\mathrm{i}(z,\gamma_\mathrm{j}(x,y))&=&0,\tag{21}\\ \displaystyle\sum_{i+j=n}\Big(\delta_{\mathrm{i}}(\gamma_\mathrm{j}(x,y))-\gamma_\mathrm{j}(\delta_{\mathrm{i}}x,y)-\gamma_\mathrm{j}(x,\delta_{\mathrm{i}}y) \Big)&=&0,\tag{22}\\ \displaystyle\sum_{i+j+k=n}\Big(\gamma_\mathrm{i}(R_\mathrm{j}x,R_\mathrm{k}y) -R_\mathrm{i}(\gamma_\mathrm{j}(R_\mathrm{k}x,y)+\gamma_\mathrm{j}(x,R_\mathrm{k}y)+\lambda\gamma_\mathrm{j}(x,y))\Big)&=&0.\tag{23} \end{align}\]

Definition 14. Two formal deformations \((\gamma_\mathrm{t},\delta_{\mathrm{t}},R_\mathrm{t})\) and \((\gamma^{\prime}_\mathrm{t},\delta^\prime_{\mathrm{t}},R^\prime_\mathrm{t})\) of a Rota-Baxter LieDer pair \((L,\delta,R)\) are said to be equivalent if there is a formal isomorphism \[\varphi_\mathrm{t}=\displaystyle\sum_{i\geq0}\varphi_\mathrm{i}t^\mathrm{i}:L[[t]]\rightarrow L[[t]],\quad \text{where } \varphi_\mathrm{i}\in \mathrm{Hom}(L,L) \text{ and } \varphi_0=\mathrm{id}_L.\] Such that the \(\mathbb{K}[[t]]\)-linear map \(\varphi_\mathrm{t}\) is a morphism of Rota-Baxter LieDer pairs from \((L[[t]]^\prime,\delta^\prime_{\mathrm{t}},R^\prime_\mathrm{t})\) to \((L[[t]],\delta_{\mathrm{t}},R_\mathrm{t})\)

It means that the following identities holds \[\varphi_\mathrm{t}(\gamma^{\prime}_\mathrm{t}(x,y))=\gamma_\mathrm{t}(\varphi_\mathrm{t}(x),\varphi_\mathrm{t}(y)) \text{ and } \varphi_\mathrm{t}\circ \delta^\prime_{\mathrm{t}}=\delta_{\mathrm{t}} \circ \varphi_\mathrm{t} \text{ and } \varphi_\mathrm{t}\circ R^\prime_\mathrm{t}=R_\mathrm{t}\circ \varphi_\mathrm{t}.\] They can be expressed as follow, for \(n\geq0\) \[\begin{align} \displaystyle\sum_{i+j=n}\varphi_\mathrm{i}(\gamma^\prime_\mathrm{j}(x,y))&=&\displaystyle\sum_{i+j+k=n}\gamma_\mathrm{i}(\varphi_\mathrm{j}(x),\varphi_\mathrm{k}(y)),\tag{24}\\ \displaystyle\sum_{i+j=n}\varphi_\mathrm{i}\circ \delta^\prime_{\mathrm{j}}&=&\displaystyle\sum_{i+j=n}\delta_\mathrm{i}\circ \varphi_\mathrm{j},\tag{25}\\ \displaystyle\sum_{i+j=n}\varphi_\mathrm{i}\circ R^\prime_{\mathrm{j}}&=&\displaystyle\sum_{i+j=n}R_\mathrm{i}\circ \varphi_\mathrm{j}.\tag{26} \end{align}\]

Theorem 19. Let \((\gamma_\mathrm{t},\delta_{\mathrm{t}},R_\mathrm{t})\) be a formal \(1\)-parameter deformation of the Rota-Baxter LieDer pair \((L,\delta,R)\). Then \((\gamma_1,\delta_1,R_1)\in \mathfrak{C}^2_\mathrm{RBLieDer}(L,L)\) is a \(2\)-cocycle in the cohmology of \((L,\delta,R)\) with coefficients in the adjoint representation.

Proof. Since \((\gamma_\mathrm{t},\delta_{\mathrm{t}},R_\mathrm{t})\) is a formal \(1\)-parameter deformation, we have from 21 , 22 and 23 , for \(n\geq0\), that \[\circlearrowleft_{x,y,z}[x,\gamma_1(y,z)]+\circlearrowleft_{x,y,z}\gamma_1(x,[y,z])=0,\] and \[\gamma_1(Rx,Ry)+[R_1x,Ry]+[Rx,R_1y]=R_1([Rx,y]+[x,Ry])+R([R_1x,y]+[x,R_1y]+\gamma_1(Rx,y)+\gamma_1(x,Ry)),\] and \[\delta_1([x,y])+\delta(\gamma_1(x,y))-\gamma_1(\delta x,y)-[\delta_1 x,y]-\gamma_1[x,\delta y]-[x,\delta_1y]=0.\] The first identity means that \(\mathrm{d}\gamma_1=0\) and the second identity means that \(\mathrm{d}_\mathrm{R}(R_1)+\Phi^2(\gamma_1)=0\) and the last one means that \(\mathrm{d}(\delta_1)+\partial^2(\gamma_1)=0\) This implies that \[\begin{align} \mathfrak{D}_\mathrm{RBLieDer}((\gamma_1,\delta_1),R_1)&=(\mathrm{d}(\gamma_1),\mathrm{d}(\delta_1)+\partial(\gamma_1),-\partial(R_1)-\Phi(R_1))\\ &=(0,0,-\partial(R_1)-\Phi(R_1))\\ &=(0,0,0). \end{align}\] This complete the proof. ◻

Theorem 20. Let \((L,\delta,R)\) be a weighted Rota-Baxter LieDer pair. If \(\mathcal{H}_\mathrm{R.B.LieDer}(L,L)=0\) then any formal \(1\)-parameter deformation of \((\mathfrak{L},\delta,R)\) is equivalent to the trivial one \((\gamma^\prime_{\mathrm{t}}=\gamma,R^\prime_{\mathrm{t}}=R,\delta^\prime_{\mathrm{t}}=\delta)\).

Proof. Let \((\gamma_\mathrm{t},\delta_\mathrm{t},R_\mathrm{t})\) be a formal \(1\)-parameter deformation of the weighted Rota-Baxter LieDer pair \((\mathfrak{L},\delta,R)\). From theorem 19 we have that \((\gamma1,\delta1,R_1)\) is a \(2\)-cocycle. Then from the hypothesis there exists \(f\in \mathrm{Hom}(L,L)\) such that \[\label{equation32deformation} \mathfrak{D}_\mathrm{R.B.LieDer}(f)=(d(f),-\partial(f),-\Phi(f)).\tag{27}\] Means that \((\gamma1,\delta1,R_1)=(d(f),-\partial(f),-\Phi(f))\).
Let \(\phi_\mathrm{t}:L\rightarrow L\) be the map \(\phi_\mathrm{t}=id+\phi_1 t\). Then \((\overline{\gamma_\mathrm{t}}=\phi_\mathrm{t}^{-1}\circ \gamma_\mathrm{t}\circ(\phi_\mathrm{t}\otimes\phi_\mathrm{t}),\overline{\delta_\mathrm{t}}=\phi_\mathrm{t}^{-1}\circ\delta_\mathrm{t}\circ\phi_\mathrm{t},\overline{R_\mathrm{t}}=\phi_\mathrm{t}^{-1}\circ R_\mathrm{t}\circ\phi_\mathrm{t} )\) is a formal \(1\)-deformation of \((\mathfrak{L},\delta,R)\) equivalent to \((\gamma_\mathrm{t},\delta_\mathrm{t},R_\mathrm{t})\). By 27 we can easily check that \((\overline{\gamma_1}=\overline{\delta_1}=\overline{R_1}=0)\), means that \[\begin{align} \overline{\gamma_\mathrm{t}}&=&\gamma+\gamma_2t^2+...\\ \overline{\delta_\mathrm{t}}&=&\delta+\delta_2 t^2+...\\ \overline{R_\mathrm{t}}&=&R+R_2t^2+... \end{align}\] and by repeating the same argument we conculde that \((\gamma_\mathrm{t},\delta_\mathrm{t},R_\mathrm{t})\) is equivalent to \((\gamma^\prime_{\mathrm{t}}=\gamma,R^\prime_{\mathrm{t}}=R,\delta^\prime_{\mathrm{t}}=\delta)\). ◻

4.2 Formal deformation of weighted Rota-Baxter AssDer pairs↩︎

In this section we study formal deformation of weighted Rota-Baxter AssDer pairs.
Let \((\mathfrak{A},\delta,R)\) be a Rota-Baxter AssDer pair. A formal \(1\)-parameter deformation of \((\mathfrak{A},\delta,R)\) consists of three formal power series \[\begin{align} \mu_\mathrm{t}&=&\displaystyle\sum_{i\geq0}\mu_\mathrm{i}t^\mathrm{i},\quad \text{where } \mu_\mathrm{i}\in \mathrm{Hom}(\wedge^2L,L) \text{ with } \mu_0=\mu,\\ \delta_\mathrm{t}&=&\displaystyle\sum_{i\geq0}\delta_\mathrm{i}t^\mathrm{i},\quad \text{where } \delta_\mathrm{i}\in \mathrm{Hom}(L,L) \text{ with } \delta_0=\delta,\\ R_\mathrm{t}&=&\displaystyle\sum_{i\geq0}R_\mathrm{i}t^\mathrm{i},\quad \text{where } R_\mathrm{i}\in \mathrm{Hom}(L,L) \text{ with } R_0=R. \end{align}\] Then we say that \((\mu_\mathrm{t},\delta_\mathrm{t},R_\mathrm{t})\) is a formal \(1\)-parameter deformation of \((\mathfrak{A},\delta,R)\) if and only if \[\begin{align} \mu_\mathrm{t}(\mu_\mathrm{t}(x,y),z)&=&\mu_\mathrm{t}(x,\mu_\mathrm{t}(y,z)),\\ \mu_\mathrm{t}(R_\mathrm{t}x,R_\mathrm{t}y)&=&R_\mathrm{t}(\mu_\mathrm{t}(R_\mathrm{t}x,y)+\mu_\mathrm{t}(x,R_{\mathrm{t}}y)+\lambda\mu_\mathrm{t}(x,y) ),\\ \delta_{\mathrm{t}}(\mu_\mathrm{t}(x,y))&=&\mu_\mathrm{t}(\delta_\mathrm{t}x,y)+\mu_\mathrm{t}(x,\delta_{\mathrm{t}}y). \end{align}\] And they are equivalent to the followings \[\begin{align} \displaystyle\sum_{i+j=n}\mu_\mathrm{i}(\mu_\mathrm{j}(x,y),z)-\mu_\mathrm{i}(x,\mu_\mathrm{j}(y,z))&=&0,\tag{28}\\ \displaystyle\sum_{i+j=n}\Big(\delta_{\mathrm{i}}(\mu_\mathrm{j}(x,y))-\mu_\mathrm{j}(\delta_{\mathrm{i}}x,y)-\mu_\mathrm{j}(x,\delta_{\mathrm{i}}y) \Big)&=&0,\tag{29}\\ \displaystyle\sum_{i+j+k=n}\Big(\mu_\mathrm{i}(R_\mathrm{j}x,R_\mathrm{k}y) -R_\mathrm{i}(\mu_\mathrm{j}(R_\mathrm{k}x,y)+\mu_\mathrm{j}(x,R_\mathrm{k}y)+\lambda\mu_\mathrm{j}(x,y))\Big)&=&0.\tag{30} \end{align}\] For \(n=1\) we obtain \[\mu_1(\mu(x,y),z)+\mu(\mu_1(x,y),z)=\mu(x,\mu_1(y,z))+\mu_1(x,\mu(y,z))\] which means that \(d_\mathrm{Hoch}(\mu_1)=0\). And the second equation \[\delta(\mu_1(x,y))+\delta_1(\mu(x,y))=\mu(\delta_1x,y)+\mu_1(\delta x,y)+\mu(x,\delta_1y)+\mu_1(x,\delta y)\] means that \(d_\mathrm{Hoch}(\delta_1)+\partial \mu_1=0\). And the third equation \[\begin{align} &&\mu_1(Rx,Ry)-R_1(\mu(Rx,y)+\mu(x,Ry)+\lambda \mu(x,y))\\ &&+\mu(R_1x,Ry)-R(\mu_1(Rx,y)+\mu_1(x,Ry)+\lambda \mu_1(x,y))\\ &&+\mu(Rx,Ry)-R(\mu(R_1x,y)+\mu(x,R_1y)+\lambda \mu(x,y))=0 \end{align}\] which means that \(-d_\mathrm{R.Hoch}-\Phi^2(\mu_1)=0\). which leads us to the following

Proposition 21. Let \((\mu_\mathrm{t},\delta_\mathrm{t},R_\mathrm{t})\) be a formal deformation of a weighted Rota-Baxter AssDer pair \((\mathfrak{A},\delta,R)\). Then the linear term \((\mu_1,\delta_1,R_1)\) is a \(2\)-cocycle in the cohomology of the weighted Rota-Baxter AssDer pair \((\mathfrak{A},\delta,R)\) with coefficients in itself.

Definition 15. Two formal deformations \((\mu_\mathrm{t},\delta_{\mathrm{t}},R_\mathrm{t})\) and \((\mu^{\prime}_\mathrm{t},\delta^\prime_{\mathrm{t}},R^\prime_\mathrm{t})\) of a weighted Rota-Baxter AssDer pair \((\mathfrak{A},\delta,R)\) are said to be equivalent if there is a formal isomorphism \[\varphi_\mathrm{t}=\displaystyle\sum_{i\geq0}\varphi_\mathrm{i}t^\mathrm{i}:A[[t]]\rightarrow A[[t]],\quad \text{where } \varphi_\mathrm{i}\in \mathrm{Hom}(L,L) \text{ and } \varphi_0=\mathrm{id}_L.\] Such that the \(\mathbb{K}[[t]]\)-linear map \(\varphi_\mathrm{t}\) is a morphism of Rota-Baxter LieDer pairs from \((A[[t]]^\prime,\delta^\prime_{\mathrm{t}},R^\prime_\mathrm{t})\) to \((A[[t]],\delta_{\mathrm{t}},R_\mathrm{t})\)

It means that the following identities holds \[\varphi_\mathrm{t}(\mu^{\prime}_\mathrm{t}(x,y))=\mu_\mathrm{t}(\varphi_\mathrm{t}(x),\varphi_\mathrm{t}(y)) \text{ and } \varphi_\mathrm{t}\circ \delta^\prime_{\mathrm{t}}=\delta_{\mathrm{t}} \circ \varphi_\mathrm{t} \text{ and } \varphi_\mathrm{t}\circ R^\prime_\mathrm{t}=R_\mathrm{t}\circ \varphi_\mathrm{t}.\] It is equivalent to the followings \[\begin{align} \displaystyle\sum_{i+j=n}\varphi_\mathrm{i}(\mu^\prime_\mathrm{j}(x,y))&=&\displaystyle\sum_{i+j+k=n}\mu_\mathrm{i}(\varphi_\mathrm{j}(x),\varphi_\mathrm{k}(y)),\\ \displaystyle\sum_{i+j=n}\varphi_\mathrm{i}\circ \delta^\prime_{\mathrm{j}}&=&\displaystyle\sum_{i+j=n}\delta_\mathrm{i}\circ \varphi_\mathrm{j},\\ \displaystyle\sum_{i+j=n}\varphi_\mathrm{i}\circ R^\prime_{\mathrm{j}}&=&\displaystyle\sum_{i+j=n}R_\mathrm{i}\circ \varphi_\mathrm{j}. \end{align}\] For \(n=0\) we have \(\varphi_0=\mathrm{Id_\mathrm{A}}\) and for \(n=1\) we obtain \[\begin{align} \varphi_1\circ \mu^\prime+\mu_1^\prime&=&\mu_1+\mu\circ (\varphi_1\otimes \mathrm{Id_A})+\mu\circ (\mathrm{Id_A}\otimes \varphi_1),\tag{31}\\ \varphi_1\circ\delta^\prime+\delta^\prime_1&=&\delta_1+\delta\circ\varphi_1,\tag{32}\\ \varphi_1\circ R^\prime+R^\prime_1&=&R_1+R\circ\varphi_1.\tag{33} \end{align}\] Then equations 31 ,32 and 33 we obtain that \[(\mu^\prime_1,\delta^\prime_1,R^\prime_1)-(\mu_1,\delta_1,R_1):=\mathfrak{D}_\mathrm{R.B.AssDer}(\varphi_1)\] This leads us to the following result

Theorem 22. Tow formal \(1\)-parameter deformations of a weighted Rota-Baxter assDer pair \((\mathfrak{A},\delta,R)\) are cohomologous. Therefore, they correspond to the same cohomology class.

Definition 16. A formal deformation \((\mu_\mathrm{t},\delta_\mathrm{t},R_\mathrm{t})\) of a weighted Rota-Baxter AssDer pair \((\mathfrak{A},\delta,R)\) is said trivial if it is equivalent to \((\mu^\prime_\mathrm{t}=\mu;\delta^\prime_\mathrm{t}=\delta)\).

Theorem 23. If \(\mathcal{H}^2_\mathrm{R.B.AssDer}(A,A)=0\) then every formal deformation of the weighted Rota-Baxter AssDer pair \((\mathfrak{A},\delta,R)\) is trivial.

The authors would like to thank the referee for valuable comments and suggestions on this article.

References↩︎

[1]
G.Baxter. "An analytic problem whose solution follows from a simple algebraic identity." Pacific J. Math 10.3 (1960): 731-742.
[2]
G.C, Rota. "Baxter algebras and combinatorial identities. I." (1969): 325-329.
[3]
N.Gu, , Li Guo. "Generating functions from the viewpoint of Rota–Baxter algebras." Discrete Mathematics 338.4 (2015): 536-554.
[4]
A.Connes, D.Kreimer. "Renormalization in quantum field theory and the Riemann–Hilbert problem I: The Hopf algebra structure of graphs and the main theorem." Communications in Mathematical Physics 210.1 (2000): 249-273.
[5]
L.Guo, B.Zhang. "Renormalization of multiple zeta values." Journal of Algebra 319.9 (2008): 3770-3809.
[6]
C.Bai. "A unified algebraic approach to the classical Yang–Baxter equation." Journal of Physics A: Mathematical and Theoretical 40.36 (2007): 11073.
[7]
M.Aguiar. "On the associative analog of Lie bialgebras." Journal of Algebra 244.2 (2001): 492-532.
[8]
C.Bai, L.Guo, X.Ni. "Nonabelian generalized Lax pairs, the classical Yang-Baxter equation and PostLie algebras." Communications in Mathematical Physics 297.2 (2010): 553-596.
[9]
C.Bai, L.Guo, X.Ni. "Relative Rota–Baxter operators and tridendriform algebras." Journal of Algebra and Its Applications 12.07 (2013): 1350027.
[10]
J.Jiang, Yunhe Sheng, and Chenchang Zhu. "Lie theory and cohomology of relative Rota–Baxter operators." Journal of the London Mathematical Society 109.2 (2024): e12863.
[11]
M.Gerstenhaber. "On the deformation of rings and algebras." Annals of Mathematics (1964): 59-103.
[12]
A.Nijenhuis, R.W. Richardson. "Deformations of Lie algebra structures." Journal of Mathematics and Mechanics 17.1 (1967): 89-105.
[13]
A.Das. "Cohomology and deformations of weighted Rota–Baxter operators." Journal of Mathematical Physics 63.9 (2022).
[14]
K.Wang, G.Zhou. "Deformations and homotopy theory of Rota-Baxter algebras of any weight." arXiv preprint arXiv:2108.06744 (2021).
[15]
T.Voronov. "Higher derived brackets and homotopy algebras." Journal of pure and applied algebra 202.1-3 (2005): 133-153.
[16]
A.R.Magid. Lectures on differential Galois theory. No. 7. American Mathematical Soc., 1994.
[17]
V.Ayala, E.Kizil, I.d.Azevedo Tribuzy. "On an algorithm for finding derivations of Lie algebras." Proyecciones (Antofagasta) 31.1 (2012): 81-90.
[18]
V.Ayala, J.Tirao. "Linear control systems on Lie groups and controllability." Proceedings of symposia in pure mathematics. Vol. 64. American Mathematical Society, 1999.
[19]
R.Tang, Y.Frégier, Y.Sheng. "Cohomologies of a Lie algebra with a derivation and applications." Journal of Algebra 534 (2019): 65-99.
[20]
A.Das. "Leibniz algebras with derivations." Journal of Homotopy and Related Structures 16 (2021): 245-274.
[21]
A.Das, A.Mandal. "Extensions, deformation and categorification of \(\text{AssDer}\) pairs." arXiv preprint arXiv:2002.11415 (2020).
[22]
Q.Sun, Z.Wu. "Cohomologies of n-Lie Algebras with Derivations." Mathematics 9.19 (2021): 2452.
[23]
Q.Sun,C.Shan "Cohomologies and deformations of Lie triple systems with derivations." Journal of Algebra and Its Applications 23.03 (2024): 2450053.
[24]
Q.Sun,L.Zhen . "Cohomologies of relative Rota-Baxter Lie algebras with derivations and applications." Journal of Geometry and Physics 195 (2024): 105054.
[25]
I.Basdouri, S.Ghribi, M.A.Sadraoui, K.M.Satyendra."Maurer-Cartan characterization and cohomology of compatible LieDer and AssDer pairs." arXiv preprint arXiv:2311.14015 (2023).
[26]
I.Basdouri, E.Peyghan, M.A.Sadraoui, Twisted Lie algebras by invertible derivations, Asian-European Journal of mathematics, doi: 10.1142/S1793557124500414.
[27]
L.Guo. "An introduction to Rota-Baxter algebra. Surveys of modern mathematics, 4." International Press: Somerville, MA, USA (2012).
[28]
Das, Apurba. "Cohomology of weighted Rota-Baxter Lie algebras and Rota-Baxter paired operators." arXiv preprint arXiv:2109.01972 (2021).
[29]
S.K.Mishra, A.Das, S.K.Hazra. "Non-abelian extensions of Rota-Baxter Lie algebras and inducibility of automorphisms." Linear Algebra and its Applications 669 (2023): 147-174.
[30]
G.Hochschild. "On the cohomology groups of an associative algebra." Annals of Mathematics (1945): 58-67.

  1. Corresponding author, E-mail: basdourimed@yahoo. fr↩︎

  2. Corresponding author, E-mail: aminsadrawi@gmail.com↩︎

  3. Corresponding author, E-mail: shuangjianguo@126.com↩︎