From relative Rota-Baxter operators and relative averaging operators on Lie algebras to relative Rota-Baxter operators on Leibniz algebras: a uniform approach

Rong Tang Department of Mathematics, Jilin University, Changchun 130012, Jilin, China tangrong@jlu.edu.cn , Yunhe Sheng Department of Mathematics, Jilin University, Changchun 130012, Jilin, China shengyh@jlu.edu.cn and Friedrich Wagemann Laboratoire de Mathematiques Jean Leray, University de Nantes, France wagemann@math.univ-nantes.fr
Abstract.

In this paper, first we construct two subcategories (using symmetric representations and antisymmetric representations) of the category of relative Rota-Baxter operators on Leibniz algebras, and establish the relations with the categories of relative Rota-Baxter operators and relative averaging operators on Lie algebras. Then we show that there is a short exact sequence describing the relation between the controlling algebra of relative Rota-Baxter operators on a Leibniz algebra with respect to a symmetric (resp. antisymmetric) representation and the controlling algebra of the induced relative Rota-Baxter operators (resp. averaging operators) on the canonical Lie algebra associated to the Leibniz algebra. Finally, we show that there is a long exact sequence describing the relation between the cohomology groups of a relative Rota-Baxter operator on a Leibniz algebra with respect to a symmetric (resp. antisymmetric) representation and the cohomology groups of the induced relative Rota-Baxter operator (resp. averaging operator) on the canonical Lie algebra.

Key words and phrases:
Rota-Baxter operator, averaging operator, cohomology, Lie algebra, Leibniz algebra

1. Introduction

The purpose of this paper is to provide a unified approach to study relative Rota-Baxter operators and relative averaging operators on Lie algebras.

1.1. Relative Rota-Baxter operators on Lie algebras

G. Baxter introduced the concept of a Rota-Baxter commutative algebra [8] in his study of fluctuation theory in probability. A Rota-Baxter operator of weight zero on an associative algebra (A,) is a linear map :AA such that

(x)(y)=((x)y+x(y)),x,yA.

These operators with general weights have been found many applications in recent years, including the algebraic approach of Connes-Kreimer [13] to renormalization of perturbative quantum field theory, quantum analogue of Poisson geometry [48], twisting on associative algebras [49], dendriform algebras and associative Yang-Baxter equations [1]. The notion of Rota-Baxter operators on Lie algebras was introduced independently in the 1980s as the operator form of the classical Yang-Baxter equation. A linear operator :𝔤𝔤 on a Lie algebra 𝔤 is called a Rota-Baxter operator of weight zero if the following condition is satisfied:

[(x),(y)]𝔤=([(x),y]𝔤+[x,(y)]𝔤),x,y𝔤.

Moreover, Kupershmidt introduced the notion of a relative Rota-Baxter operator (also called an 𝒪-operator) on a Lie algebra 𝔤 with respect to arbitrary representation in [29]. Note that a skew-symmetric classical r-matrix is a relative Rota-Baxter operator on a Lie algebra with respect to the coadjoint representation. Relative Rota-Baxter operators play important roles in the study of integrable systems [6, 40], provide solutions of the classical Yang-Baxter equation in the semidirect product Lie algebra and give rise to pre-Lie algebras [5]. See [21] for more details. Recently, the deformation and homotopy theory of relative Rota-Baxter Lie algebras and relative Rota-Baxter associative algebras were established in [15, 30, 44, 45].

1.2. Averaging operators on Lie algebras

An averaging operator on an associative algebra (A,) is a linear operator :AA such that

(x)(y)=((x)y)=(x(y)),x,yA.

In the last century, many studies on averaging operators were done for various special algebras, such as function spaces, Banach algebras, and the topics and methods were largely analytic [7, 12, 24, 39]. Recently, it was found that commutative double algebra structures [20] on a vector space V can be described by symmetric averaging operators on the associative algebra End(V). Averaging operators can be defined on algebras over arbitrary binary operad [1]. In particular, a linear operator :𝔤𝔤 on a Lie algebra 𝔤 is called an averaging operator if the following condition is satisfied:

[(x),(y)]𝔤=([(x),y]𝔤),x,y𝔤.

There is a close relation between the classical Yang-Baxter equation, conformal algebras and averaging operators on Lie algebras [26].

Averaging operators on Lie algebras are also called embedding tensors in some mathematical physics literatures. Embedding tensors and their associated tensor hierarchies provide an algebraic and efficient way to construct supergravity theories and further to construct higher gauge theories (see e.g. [10, 23, 28, 42]). Recently the controlling algebra and the cohomology theory of embedding tensors were established in [41] using the higher derived bracket. Meanwhile the cohomology and homotopy theory of averaging associative algebras were studied in [52].

The concept of Rota-Baxter operators and averaging operators were further studied on the level of algebraic operads in [3, 35, 36, 37, 50, 51]. The action of the Rota-Baxter operator on a binary quadratic operad splits the operad, and the action of the averaging operator on a binary quadratic operad duplicates the operad. Since the splitting and duplication processes are in Koszul duality, one can regard the Rota-Baxter operator and averaging operator to be Koszul dual to each other.

1.3. Relative Rota-Baxter operators on Leibniz algebras

Leibniz algebras were first discovered by Bloh who called them D-algebras [9]. Then Loday rediscovered this algebraic structure and called them Leibniz algebras with the motivation in the study of the periodicity in algebraic K-theory [31, 33, 34]. Averaging operators on Lie algebras give rise to Leibniz algebras, which can be viewed as the duplication of Lie algebras. From the viewpoint of the operad theory, leibniz algebras are defined as the algebras over the Leibniz operad which is the duplicator of the Lie operad [36]. More intrinsically, let 𝒫 be a binary quadratic operad. The algebra duplicates by an averaging operator on the 𝒫-algebra is a Du(𝒫)-algebra. On the other hand, the algebra splits by a Rota-Baxter operator on the 𝒫-algebra is a BSu(𝒫)-algebra. Since Du(𝒫) is the Koszul dual of BSu(𝒫!) [36], the duplication construction of an averaging operator is a kind of Koszul dual of the splitting construction of a Rota-Baxter operator [3, 36]. The (co)homology and homotopy theories of Leibniz algebras were established in [18, 33, 38, 2]. Recently Leibniz algebras were studied from different aspects due to applications in both mathematics and physics. In particular, integration of Leibniz algebras were studied in [11, 14] and deformation quantization of Leibniz algebras was studied in [16]. As the underlying algebraic structures of embedding tensors and Courant algebroids, Leibniz algebras also have application in higher gauge theories [10, 28, 43] and homogeneous spaces [25]. The notion of relative Rota-Baxter operators on Leibniz algebras was introduced in [46] to study Leibniz bialgebras. Moreover, the cohomology theory of relative Rota-Baxter operators on Leibniz algebras was given in [47] and used to classify linear deformations and formal deformations.

1.4. Outline of the paper

In this paper, we propose a unified approach to study relative Rota-Baxter operators and relative averaging operators on Lie algebras using relative Rota-Baxter operators on Leibniz algebras. First we observe that there are forgetful functors from the categories of relative Rota-Baxter operators and relative averaging operators on Lie algebras to the category of relative Rota-Baxter operators on Leibniz algebras. Conversely, we construct two functors from certain subcategories of the category of relative Rota-Baxter operators on Leibniz algebras to the categories of relative Rota-Baxter operators and relative averaging operators on Lie algebras, and show that they are left adjoint for the aforementioned forgetful functors. Then we show that the controlling algebras of relative Rota-Baxter operators and relative averaging operators on Lie algebras can be derived from the controlling algebra of relative Rota-Baxter operators on Leibniz algebras. Finally, we give a long exact sequence to describe the relation between the cohomology groups of a relative Rota-Baxter operator on a Leibniz algebra with respect to a symmetric (resp. antisymmetric) representation and the cohomology groups of the induced relative Rota-Baxter operator (resp. averaging operator) on the canonical Lie algebra. The concrete relations can be summarized by the following diagrams:

𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖲Theorem 2.8descendent𝖱𝖡𝖮𝖫𝗂𝖾descendentLie algebrasIdLie algebras𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖠Theorem 2.12descendent𝖠𝖮𝖫𝗂𝖾descendentLeibniz algebrasIdLeibniz algebras.

The paper is organized as follows. In Section 2, using symmetric representations and antisymmetric representations, we construct two subcategories of the category of relative Rota-Baxter operators on Leibniz algebras, and establish the relations with the categories of relative Rota-Baxter operators and relative averaging operators on Lie algebras. In Section 3, we show that there is a short exact sequence describing the relation between the controlling algebra of relative Rota-Baxter operators on a Leibniz algebra with respect to a symmetric representation and the controlling algebra of the induced relative Rota-Baxter operators on the canonical Lie algebra. For relative averaging operators, we establish a similar result. In Section 4, first we introduce the Loday-Pirashvili cohomology of a relative Rota-Baxter operator on a Lie algebra and then use a long exact sequence to describe the relation between the cohomology groups of a relative Rota-Baxter operator on a Leibniz algebra with respect to a symmetric representation and the Loday-Pirashvili cohomology groups of the induced relative Rota-Baxter operator on the canonical Lie algebra. Similarly, there is a long exact sequence describing the relation between the cohomology groups of a relative Rota-Baxter operator on a Leibniz algebra with respect to an antisymmetric representation and the cohomology groups of the induced relative averaging operator on the canonical Lie algebra.

In this paper, we work over an algebraically closed field 𝕂 of characteristic 0 and all the vector spaces are over 𝕂 and finite-dimensional.

2. Relations between the categories 𝖠𝖮𝖫𝗂𝖾, 𝖱𝖡𝖮𝖫𝗂𝖾 and 𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓

In this section, we establish the relations between certain subcategories of the category of relative Rota-Baxter operators on Leibniz algebras and the category of relative Rota-Baxter operators on Lie algebras as well as the category of relative averaging operators on Lie algebras.

2.1. Relative Rota-Baxter operators and relative averaging operators

Definition 2.1.

([29]) A relative Rota-Baxter operator on a Lie algebra (𝔤,[,]𝔤) with respect to a representation (V;ρ) is a linear map T:V𝔤 satisfying the following quadratic constraint:

(1) [Tu,Tv]𝔤=T(ρ(Tu)(v)ρ(Tv)(u)),u,vV.

Let T:V𝔤 (resp. T:V𝔤) be a relative Rota-Baxter operator on the Lie algebra (𝔤,[,]𝔤) (resp. (𝔤,{,}𝔤)) with respect to the representation (V;ρ) (resp. (V,ρ)). A homomorphism from T to T is a pair (ϕ,φ), where ϕ:𝔤𝔤 is a Lie algebra homomorphism, φ:VV is a linear map such that

(2) Tφ = ϕT,
(3) φρ(x)(u) = ρ(ϕ(x))(φ(u)),x𝔤,uV.

In particular, if ϕ and φ are invertible, then (ϕ,φ) is called an isomorphism.

We denote by 𝖱𝖡𝖮𝖫𝗂𝖾 the category of relative Rota-Baxter operators on Lie algebras.

Definition 2.2.

([1, 39]) A relative averaging operator on a Lie algebra (𝔤,[,]𝔤) with respect to a representation (V;ρ) is a linear map T:V𝔤 satisfying the following quadratic constraint:

(4) [Tu,Tv]𝔤=T(ρ(Tu)(v)),u,vV.

Let T:V𝔤 (resp. T:V𝔤) be a relative averaging operator on the Lie algebra (𝔤,[,]𝔤) (resp. (𝔤,{,}𝔤)) with respect to the representation (V;ρ) (resp. (V,ρ)). A homomorphism from T to T is a pair (ϕ,φ), where ϕ:𝔤𝔤 is a Lie algebra homomorphism, φ:VV is a linear map such that

(5) Tφ = ϕT,
(6) φρ(x)(u) = ρ(ϕ(x))(φ(u)),x𝔤,uV.

In particular, if ϕ and φ are invertible, then (ϕ,φ) is called an isomorphism.

We denote by 𝖠𝖮𝖫𝗂𝖾 the category of relative averaging operators on Lie algebras.

A (left) Leibniz algebra is a vector space 𝔊 together with a bilinear operation [,]𝔊:𝔊𝔊𝔊 such that

[x,[y,z]𝔊]𝔊=[[x,y]𝔊,z]𝔊+[y,[x,z]𝔊]𝔊,x,y,z𝔊.

A representation of a Leibniz algebra (𝔊,[,]𝔊) is a triple (V;ρL,ρR), where V is a vector space, ρL,ρR:𝔊𝔤𝔩(V) are linear maps such that for all x,y𝔊,

ρL([x,y]𝔊) = [ρL(x),ρL(y)],
ρR([x,y]𝔊) = [ρL(x),ρR(y)],
ρR(y)ρL(x) = ρR(y)ρR(x).

Here [,] is the commutator Lie bracket on 𝔤𝔩(V). There are two important kinds of representations of Leibniz algebras:

  • a representation (V;ρL,ρR) is called symmetric if ρR=ρL;

  • a representation (V;ρL,ρR) is called antisymmetric if ρR=0.

For any representation (V;ρL,ρR) of a Leibniz algebra 𝔊, we define a vector subspace Vanti which is spanned by all elements ρL(x)v+ρR(x)v with x𝔊 and vV. Directly from ρR(y)ρL(x)=ρR(y)ρR(x) for all x,y𝔊, we deduce that (Vanti;ρL,ρR) is an antisymmetric representation. The quotient representation V/Vanti is then symmetric, denoted by Vsym, and we have an exact sequence

(7) 0VantiVVsym0.

Now we recall the notion of a relative Rota-Baxter operator on a Leibniz algebra.

Definition 2.3.

([46]) Let (V;ρL,ρR) be a representation of a Leibniz algebra (𝔊,[,]𝔊). A linear operator T:V𝔊 is called a relative Rota-Baxter operator on (𝔊,[,]𝔊) with respect to (V;ρL,ρR) if T satisfies:

(8) [Tu,Tv]𝔊=T(ρL(Tu)v+ρR(Tv)u),u,vV.

Let T:V𝔊 (resp. T:V𝔊) be a relative Rota-Baxter operator on the Leibniz algebra (𝔊,[,]𝔊) (resp. (𝔊,{,}𝔊)) with respect to the representation (V;ρL,ρR) (resp. (V;ρL,ρR)). A homomorphism from T to T is a pair (ϕ,φ), where ϕ:𝔊𝔊 is a Leibniz algebra homomorphism, φ:VV is a linear map such that for all x𝔊,uV,

(9) Tφ = ϕT,
(10) φρL(x)(u) = ρL(ϕ(x))(φ(u)),
(11) φρR(x)(u) = ρR(ϕ(x))(φ(u)).

In particular, if ϕ and φ are invertible, then (ϕ,φ) is called an isomorphism from T to T.

We denote by 𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓 the category of relative Rota-Baxter operators on Leibniz algebras. Moreover, we denote by 𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖲 and 𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖠 the categories of relative Rota-Baxter operators on Leibniz algebras with respect to symmetric representations and antisymmetric representations respectively.

2.2. Relation to crossed modules

Note that relative Rota-Baxter operators on Lie algebras and Leibniz algebras T:V𝔤 resp. T:V𝔊 as well as relative averaging operators on Lie algebras T:V𝔤 give rise to a bracket on V.

For a relative Rota-Baxter operator T:V𝔤 on a Lie algebra 𝔤, the bracket

[v,w]T:=ρ(Tv)wρ(Tw)v

for all v,wV renders V a Lie algebra such that T:V𝔤 becomes a morphism of Lie algebras.

In the same vein, given a relative Rota-Baxter operator T:V𝔊 on a Leibniz algebra 𝔊, the bracket

[v,w]T:=ρL(Tv)w+ρR(Tw)v

for all v,wV renders V a Leibniz algebra such that T:V𝔊 becomes a morphism of Leibniz algebras.

Still in the same vein, given a relative averaging operator T:V𝔤 on a Lie algebra 𝔤, the bracket

[v,w]T:=ρ(Tv)w

for all v,wV renders V a Leibniz algebra such that T:V𝔤 is a morphism of Leibniz algebras.

The next proposition shows that with an equivariance condition, the action of 𝔤 (resp. 𝔊) on V becomes an action by derivations with respect to these brackets. This kind of equivariance condition is present, for example, in Loday-Pirashvili’s tensor category of linear maps, see [32].

Proposition 2.4.
  1. (1)

    Let T:V𝔤 be a relative Rota-Baxter operator on a Lie algebra and suppose that T is equivariant in the sense that T(ρ(x)v)=[x,Tv]𝔤 for all x𝔤 and all vV. We deduce that ρ(x) for all x𝔤 is a derivation of the Lie algebra (V,[,]T). Thus, we obtain that ρ:𝔤Der(V) is homomorphism of Lie algebras.

  2. (2)

    Let T:V𝔊 be a relative Rota-Baxter operator on a Leibniz algebra and suppose that T is equivariant in the sense that T(ρL(x)v)=[x,Tv]𝔊 and T(ρR(x)v)=[Tv,x]𝔊 for all x𝔤 and all vV. Then 𝔊 acts on (V,[,]T) by derivations (in the Leibniz sense), i.e. for all x𝔊 and all v,wV

    ρL(x)[v,w]T=[ρL(x)v,w]T+[v,ρL(x)w]T,
    ρR(x)[v,w]T=[v,ρR(x)w]T[w,ρR(x)v]T,
    [ρL(x)v,w]T+[ρR(x)v,w]T=0.

We deduce that the pair (ρL(x),ρR(x)) for all x𝔊 is a biderivation [31] of the Leibniz algebra (V,[,]T). For relative averaging operators, equivariance will imply that the left action is by derivations, but defining the right operation to be the opposite of the left action or zero will not render this an action by derivations in the Leibniz sense in general.

Note that this proposition does not give rise to crossed modules, because the requirements ρ(Tv)w=[v,w] in the Lie case and ρL(Tv)w=[v,w]=ρR(w)v in the Leibniz case for v,wV are not fullfilled. In the case of relative averaging operators, this requirement is fullfilled, but the action is not by derivations in general. On the other hand, it is well known that crossed modules of Lie algebras provide examples of relative averaging operators, see Example 2.14 in [41].

2.3. Relations between the categories 𝖱𝖡𝖮𝖫𝗂𝖾 and 𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖲

First we construct a functor F from the category 𝖱𝖡𝖮𝖫𝗂𝖾 of relative Rota-Baxter operators on Lie algebras to the category 𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖲 of relative Rota-Baxter operators on Leibniz algebras with respect to symmetric representations.

On objects, the functor F is defined as follows. Let T:V𝔤 be a relative Rota-Baxter operator on the Lie algebra (𝔤,[,]𝔤) with respect to the representation (V;ρ). We view the Lie algebra (𝔤,[,]𝔤) as a Leibniz algebra. Moreover the Lie algebra representation ρ gives rise to a symmetric representation (V;ρ,ρ) of the Leibniz algebra (𝔤,[,]𝔤). Then the equality (1) means that T:V𝔤 is a relative Rota-Baxter operator on the Leibniz algebra (𝔤,[,]𝔤) with respect to the symmetric representation (V;ρ,ρ).

On morphisms, the functor F is defined as follows. Let T:V𝔤 (resp. T:V𝔤) be a relative Rota-Baxter operator on the Lie algebra (𝔤,[,]𝔤) (resp. (𝔤,{,}𝔤)) with respect to the representation (V;ρ) (resp. (V,ρ)). Let (ϕ,φ) be a homomorphism from the relative Rota-Baxter operator T to T. Then it is obvious that (ϕ,φ) is also a homomorphism between the above induced relative Rota-Baxter operators on Leibniz algebras.

It is straightforward to see that F defined above is indeed a functor.

In the sequel, we construct a functor G from the category 𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖲 to the category 𝖱𝖡𝖮𝖫𝗂𝖾.

Let (𝔊,[,]𝔊) be a Leibniz algebra. Denote by 𝖫𝖾𝗂𝖻(𝔊) the ideal of squares spanned by all elements [x,x]𝔊 for all x𝔊. We call 𝖫𝖾𝗂𝖻(𝔊) the Leibniz kernel of (𝔊,[,]𝔊). Observe that 𝖫𝖾𝗂𝖻(𝔊)=Vanti for the adjoint representation (V=𝔊,ρL,ρR) with ρL(x)=[x,] and ρR(x)=[,x]. The following result is obvious.

Lemma 2.5.

Let (𝔊,[,]𝔊) be a Leibniz algebra. Then

𝔊Lie:=𝔊/𝖫𝖾𝗂𝖻(𝔊)

is naturally a Lie algebra in which we denote the Lie bracket by [,]𝔊Lie.

We call 𝔊Lie the canonical Lie algebra associated to the Leibniz algebra (𝔊,[,]𝔊). Observe that 𝔊Lie=Vsym for the adjoint representation (V=𝔊,ρL,ρR). The equivalence class of x𝔊 in 𝔊Lie will be denoted by x¯. Moreover, 𝖫𝖾𝗂𝖻(𝔊) is contained in the left center of the Leibniz algebra 𝔊. We denote by

(12) pr:𝔊𝔊Lie

the natural projection from the Leibniz algebra 𝔊 to the canonical Lie algebra 𝔊Lie. Obviously, the map pr preserves the bracket operation, i.e. the following equality holds:

(13) pr[x,y]𝔊=[pr(x),pr(y)]𝔊Lie,x,y𝔊.

Let (V;ρL,ρR) be a representation of a Leibniz algebra (𝔊,[,]𝔊). Then ρL:𝔊𝔤𝔩(V) is a Leibniz algebra homomorphism. We deduce that

(14) ρL(𝖫𝖾𝗂𝖻(𝔊))=0.

Thus, there is exactly one Lie algebra homomorphism θ:𝔊Lie𝔤𝔩(V) which is defined by

(15) θ(x¯)=ρL(x),x𝔊,

such that following diagram of Leibniz algebra homomorphisms commute:

𝔊prρL𝔊Lieθ𝔤𝔩(V).

It implies that (V;θ) is a representation of the canonical Lie algebra 𝔊Lie.

Proposition 2.6.

Let T:V𝔊 be a relative Rota-Baxter operator on a Leibniz algebra (𝔊,[,]𝔊) with respect to a symmetric representation (V;ρL,ρR=ρL). Then T¯:=prT:V𝔊Lie is a relative Rota-Baxter operator on the canonical Lie algebra 𝔊Lie with respect to the representation (V;θ).

Proof.

Applying the projection pr:𝔊𝔊Lie to the both sides of (8), we obtain

[prT(u),prT(v)]𝔊Lie = prT(ρL(Tu)vρL(Tv)u)
= prT(θ(prT(u))vθ(prT(v))u),

which implies that prT:V𝔊Lie is a relative Rota-Baxter operator on the canonical Lie algebra 𝔊Lie with respect to the representation (V;θ).

Proposition 2.7.

Let T:V𝔊 (resp. T:V𝔊) be a relative Rota-Baxter operator on the Leibniz algebra (𝔊,[,]𝔊) (resp. (𝔊,{,}𝔊)) with respect to the symmetric representation (V;ρL,ρR=ρL) (resp. (V;ρL,ρR=ρL)). Let (ϕ,φ) be a homomorphism of relative Rota-Baxter operators on Leibniz algebras from T to T. Then (ϕ¯,φ) is a homomorphism of relative Rota-Baxter operators on Lie algebras from T¯ to T¯, where the Lie algebra homomorphism ϕ¯:𝔊Lie𝔊Lie is defined by

ϕ¯(x¯):=ϕ(x)¯,x𝔊.
Proof.

By ϕ(𝖫𝖾𝗂𝖻(𝔊))𝖫𝖾𝗂𝖻(𝔊), we obtain that ϕ induces the Lie algebra homomorphism ϕ¯:𝔊Lie𝔊Lie. Then it is straightforward to deduce that (ϕ¯,φ) is a homomorphism of relative averaging operators from T¯ to T¯.

Now we are ready to give the main result in this subsection.

Theorem 2.8.

There is a functor G from the category 𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖲 of relative Rota-Baxter operators on Leibniz algebras with respect to symmetric representations to the category 𝖱𝖡𝖮𝖫𝗂𝖾 of relative Rota-Baxter operators on Lie algebras, such that G is a left adjoint for F.

Proof.

Let T:V𝔊 be a relative Rota-Baxter operator on a Leibniz algebra (𝔊,[,]𝔊) with respect to a symmetric representation (V;ρL,ρR=ρL). By Proposition 2.6, T¯:=prT:V𝔊Lie is a relative Rota-Baxter operator on the canonical Lie algebra 𝔊Lie with respect to the representation (V;θ). Thus, on objects, we define

G(T):=T¯.

Let (ϕ,φ) be a homomorphism of relative Rota-Baxter operators on Leibniz algebras from T to T. By Proposition 2.7, (ϕ¯,φ) is a homomorphism of relative Rota-Baxter operators from T¯ to T¯. Thus, on morphisms, we define

G(ϕ,φ):=(ϕ¯,φ).

Then it is straightforward to see that G is indeed a functor. Moreover, let 𝔤 be a Lie algebra, we have 𝖫𝖾𝗂𝖻(𝔤)=0. Thus, we obtain that GF=Id𝖱𝖡𝖮𝖫𝗂𝖾. Thus, we have the identity natural transformation ε (the counit of the adjunction)

ε=IdId𝖱𝖡𝖮𝖫𝗂𝖾:Id𝖱𝖡𝖮𝖫𝗂𝖾=GFId𝖱𝖡𝖮𝖫𝗂𝖾.

Moreover, for any relative Rota-Baxter operator K:W𝔥 on a Lie algebra 𝔥 with respect to a representation (W;ρ), relative Rota-Baxter operator T:V𝔊 on a Leibniz algebra 𝔊 with respect to a symmetric representation (V;ρL,ρR=ρL) and relative Rota-Baxter operator homomorphism (χ,ξ) from G(T) to K, we deduce that χpr is a Leibniz algebra homomorphism from 𝔊 to 𝔥 and

Kξ=(χpr)T
ξ(ρL(x)u)=(15)ξ(θ(x¯)u)=ρ(χ(x¯))ξ(u)=ρ((χpr)(x))ξ(u),x𝔊,uV.

Thus, there is exactly one relative Rota-Baxter operator homomorphism (χpr,ξ) from T to F(K) such that the following diagram commutes:

K=GF(K)(Id𝔥,IdW)KG(T),G(χpr,ξ)(χ,ξ)

which implies that G is a left adjoint for F. The proof is finished.

Remark 2.9.

The functors G:𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖲𝖱𝖡𝖮𝖫𝗂𝖾 and F:𝖱𝖡𝖮𝖫𝗂𝖾𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖲 are induced by those establishing the equivalence of categories between modules over a Lie algebra 𝔤 and symmetric Leibniz representations over 𝔤, viewed as a Leibniz algebra. From this, one may deduce that the pair (F,G) constitutes almost an adjoint equivalence. In fact, F is fully faithful, because the counit of the adjunction ϵ:GFId is a natural isomorphism for each object of 𝖱𝖡𝖮𝖫𝗂𝖾. On the other hand, the unit of the adjunction η:IdFG is not an isomorphism. For an object T:V𝔊 in 𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖲, (FG)(T) corresponds to viewing prT:V𝔊Lie again as an object in 𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖲. We have a natural relative Rota-Baxter operator homomorphism (pr,IdV) from T to prT

VIdVT𝔊prVprT𝔊Lie,

but it is not an isomorphism in general.

Note that one may add the data which arises from the passage from 𝔊Lie to 𝔊 to the category 𝖱𝖡𝖮𝖫𝗂𝖾 in order to make it an equivalence of categories. The data consists of the isomorphism class of an antisymmetric representation of 𝔊Lie (corresponding to 𝖫𝖾𝗂𝖻(𝔊)) and of the Loday-Pirashvili cohomology class of a 2-cocycle characterizing the abelian extension

(16) 0𝖫𝖾𝗂𝖻(𝔊)𝔊𝔊Lie0.

Given a relative Rota-Baxter operator T:V𝔤 together with a 𝔤-module A (viewed as an antisymmetric representation of the Leibniz algebra 𝔤) and a 2-cocycle αZ2(𝔤,A), we obtain a relative Rota-Baxter operator T:V𝔊 for the Leibniz algebra 𝔊 constructed from (𝔤,A,α) with A=𝖫𝖾𝗂𝖻(𝔊) and V viewed as a symmetric representation of 𝔊, because 𝖫𝖾𝗂𝖻(𝔊) acts trivially from the left and thus also trivially from the right as the representation V is symmetric. This construction depends on the choice of the cocycle and the module A, but becomes natural when passing to cohomology and isomorphism classes.

2.4. Relations between the categories 𝖠𝖮𝖫𝗂𝖾 and 𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖠

First we construct a functor from the category 𝖠𝖮𝖫𝗂𝖾 of relative averaging operators on Lie algebras to the category 𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖠 of relative Rota-Baxter operators on Leibniz algebras with respect to antisymmetric representations.

On objects, the functor is defined as follows. Let T:V𝔤 be a relative averaging operator on a Lie algebra (𝔤,[,]𝔤) with respect to a representation (V;ρ). We view the Lie algebra (𝔤,[,]𝔤) as a Leibniz algebra. Moreover the Lie algebra representation ρ gives rise to an antisymmetric representation (V;ρ,0) of the Leibniz algebra (𝔤,[,]𝔤). Then the equality (4) means that T:V𝔤 is a relative Rota-Baxter operator on the Leibniz algebra (𝔤,[,]𝔤) with respect to the antisymmetric representation (V;ρ,0).

On morphisms, the functor is defined as follows. Let T:V𝔤 (resp. T:V𝔤) be a relative averaging operator on a Lie algebra (𝔤,[,]𝔤) (resp. (𝔤,{,}𝔤)) with respect to a representation (V;ρ) (resp. (V,ρ)). Let (ϕ,φ) be a homomorphism from the relative averaging operator T to T. Then it is obvious that (ϕ,φ) is also a homomorphism between the above induced relative Rota-Baxter operators.

It is straightforward to see that defined above is indeed a functor.

In the sequel, we construct a functor 𝒢 from the category 𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖠 to the category 𝖠𝖮𝖫𝗂𝖾.

Proposition 2.10.

Let T:V𝔊 be a relative Rota-Baxter operator on a Leibniz algebra (𝔊,[,]𝔊) with respect to an antisymmetric representation (V;ρL,ρR=0). Then T¯:=prT:V𝔊Lie is a relative averaging operator on the Lie algebra 𝔊Lie with respect to the representation (V;θ).

Proof.

Applying the projection pr:𝔊𝔊Lie to the both sides of (8), we obtain

[prT(u),prT(v)]𝔊Lie=prT(ρL(Tu)v)=(15)prT(θ(prT(u))v),

which implies that prT:V𝔊Lie is an averaging operator on the Lie algebra 𝔊Lie with respect to the representation (V;θ).

Similar to Proposition 2.7, we have the following result.

Proposition 2.11.

Let T:V𝔊 (resp. T:V𝔊) be a relative Rota-Baxter operator on the Leibniz algebra (𝔊,[,]𝔊) (resp. (𝔊,{,}𝔊)) with respect to the antisymmetric representation (V;ρL,ρR=0) (resp. (V;ρL,ρR=0)). Let (ϕ,φ) be a homomorphism of relative Rota-Baxter operators on Leibniz algebras from T to T. Then (ϕ¯,φ) is a homomorphism of relative averaging operators from T¯ to T¯, where the Lie algebra homomorphism ϕ¯:𝔊Lie𝔊Lie is defined by

ϕ¯(x¯):=ϕ(x)¯,x𝔊.

Now we are ready to give the main result in this subsection.

Theorem 2.12.

There is a functor 𝒢 from the category 𝖱𝖡𝖮𝖫𝖾𝗂𝖻𝗇𝗂𝗓𝖠 of relative Rota-Baxter operators on Leibniz algebras with respect to antisymmetric representations to the category 𝖠𝖮𝖫𝗂𝖾 of relative averaging operators on Lie algebras, such that 𝒢 is a left adjoint for .

Proof.

Let T:V𝔊 be a relative Rota-Baxter operator on a Leibniz algebra (𝔊,[,]𝔊) with respect to an antisymmetric representation (V;ρL,ρR=0). By Proposition 2.10, T¯:=prT:V𝔊Lie is a relative averaging operator on the Lie algebra 𝔊Lie with respect to the representation (V;θ). Thus, on objects, we define

𝒢(T):=T¯.

Let (ϕ,φ) be a homomorphism of relative Rota-Baxter operators on Leibniz algebras from T to T. By Proposition 2.11, (ϕ¯,φ) is a homomorphism of relative averaging operators from T¯ to T¯. Thus, on morphisms, we define

𝒢(ϕ,φ):=(ϕ¯,φ).

Then it is straightforward to see that 𝒢 is indeed a functor.

By the similar argument as Theorem 2.8, we can show that 𝒢 is a left adjoint for . We omit details.

Remark 2.13.

Like in Remark 2.9, the pair of functors and 𝒢 form almost an adjoint equivalence.

3. Relations between the controlling algebras

In this section, we establish the relation between the controlling algebra of relative Rota-Baxter operators on a Leibniz algebra with respect to a symmetric (resp. antisymmetric) representation and the controlling algebra of the relative Rota-Baxter (resp. averaging) operators on the canonical Lie algebra.

A permutation σ𝕊n is called an (i,ni)-shuffle if σ(1)<<σ(i) and σ(i+1)<<σ(n). If i=0 or n, we assume σ=Id. The set of all (i,ni)-shuffles will be denoted by 𝕊(i,ni). The notion of an (i1,,ik)-shuffle and the set 𝕊(i1,,ik) are defined analogously.

A degree 1 element θ𝔤1 is called a Maurer-Cartan element of a differential graded Lie algebra (k𝔤k,[,],d) if it satisfies the Maurer-Cartan equation: dθ+12[θ,θ]=0. The set of Maurer-Cartan elements in a dgla 𝔤 will be denoted by MC(𝔤).

Let 𝔤 be a vector space. We consider the graded vector space C(𝔤,𝔤)=n1Cn(𝔤,𝔤)=n1Hom(n𝔤,𝔤). An element PCp+1(𝔤,𝔤) is defined to have degree p. The Balavoine bracket on the graded vector space C(𝔤,𝔤) is given by:

(17) [P,Q]𝖡=P¯Q(1)pqQ¯P,PCp+1(𝔤,𝔤),QCq+1(𝔤,𝔤),

where P¯QCp+q+1(𝔤,𝔤) is defined by

(18) P¯Q=k=1p+1PkQ,

and k is defined by

(PkQ)(x1,,xp+q+1)
= σ𝕊(k1,q)(1)(k1)q(1)σP(xσ(1),,xσ(k1),Q(xσ(k),,xσ(k+q1),xk+q),xk+q+1,,xp+q+1).

It is well known that

Theorem 3.1.

([4, 19]) With the above notations, (C(𝔤,𝔤),[,]𝖡) is a graded Lie algebra. Its Maurer-Cartan elements (as a differential graded Lie algebra with zero differential) are precisely the Leibniz algebra structures on 𝔤.

3.1. The controlling algebra of relative Rota-Baxter operators on a Leibniz algebra

Let (V;ρVL,ρVR) be a representation of a Leibniz algebra (𝔊,[,]𝔊). Then there is a Leibniz algebra structure on 𝔊V given by

(19) [x+u,y+v]=[x,y]𝔊+ρVL(x)v+ρVR(y)u,x,y𝔤,u,vV.

This Leibniz algebra is called the semidirect product of 𝔊 and (V;ρVL,ρVR), and denoted by 𝔊ρVL,ρVRV. We denote the above semidirect product Leibniz multiplication by μ. Consider the graded vector space

C(V,𝔊):=n1Cn(V,𝔊)=n1Hom(nV,𝔊),

where an element gCn(V,𝔊) is defined to be of degree n.

Theorem 3.2.

([46]) With the above notations, (C(V,𝔊),{,}V) is a graded Lie algebra, where the graded Lie bracket {,}V:Cm(V,𝔊)×Cn(V,𝔊)Cm+n(V,𝔊) is given by the derived bracket as following:

{g1,g2}V = (1)m1[[μ,g1]𝖡,g2]𝖡,g1Cm(V,𝔊),g2Cn(V,𝔊).

More precisely, we have

{g1,g2}V(v1,v2,,vm+n)
= k=1mσ𝕊(k1,n)(1)(k1)n+1(1)σg1(vσ(1),,vσ(k1),ρVL(g2(vσ(k),,vσ(k+n1)))vk+n,vk+n+1,,vm+n)
+k=2m+1σ𝕊(k2,n,1)σ(k+n2)=k+n1(1)kn(1)σg1(vσ(1),,vσ(k2),ρVR(g2(vσ(k1),,vσ(k+n2)))vσ(k+n1),vk+n,,vm+n)
+k=1mσ𝕊(k1,n1)(1)(k1)n(1)σ[g2(vσ(k),,vσ(k+n2),vk+n1),g1(vσ(1),,vσ(k1),vk+n,,vm+n)]𝔊
+σ𝕊(m,n1)(1)mn+1(1)σ[g1(vσ(1),,vσ(m)),g2(vσ(m+1),,vσ(m+n1),vm+n)]𝔊
+k=1nσ𝕊(k1,m)(1)m(k+n1)(1)σg2(vσ(1),,vσ(k1),ρVL(g1(vσ(k),,vσ(k+m1)))vk+m,vk+m+1,,vm+n)
+k=1nσ𝕊(k1,m,1)σ(k+m1)=k+m(1)m(k+n1)+1(1)σg2(vσ(1),,vσ(k1),ρVR(g1(vσ(k),,vσ(k1+m)))vσ(k+m),vk+m+1,,vm+n).

Moreover, its Maurer-Cartan elements (where the differential is taken to be zero again) are relative Rota-Baxter operators on the Leibniz algebra (𝔊,[,]𝔊) with respect to the representation (V;ρVL,ρVR).

In fact, this graded Lie algebra construction is functorial. More precisely, we consider the category 𝔊-𝖱𝖾𝗉 of representations of the Leibniz algebra (𝔊,[,]𝔊). Then, we have a contravariant functor 𝖢𝔊 from the category 𝔊-𝖱𝖾𝗉 to the category 𝖦𝗅𝖺 of graded Lie algebras.

Let (W;ρWL,ρWR) and (V;ρVL,ρVR) be two representations of the Leibniz algebra (𝔊,[,]𝔊). A homomorphism from (W;ρWL,ρWR) to (V;ρVL,ρVR) is a linear map ϕ:WV such that

(20) ϕ(ρWL(x)w) = ρVL(x)ϕ(w),
(21) ϕ(ρWR(x)w) = ρVR(x)ϕ(w),x𝔊,wW.

Let ϕ:WV be a homomorphism from the representation (W;ρWL,ρWR) to (V;ρVL,ρVR). Define a linear map Φ:Hom(nV,𝔊)Hom(nW,𝔊),n1, by

(22) Φ(f):=fϕn,fHom(nV,𝔊).
Proposition 3.3.

With the above notations, Φ is a homomorphism from the graded Lie algebra (C(V,𝔊),{,}V) to (C(W,𝔊),{,}W).

Proof.

It follows directly from (20) and (21). We omit the details.

Moreover, we have the following theorem.

Theorem 3.4.

Theorem 3.2 and Proposition 3.3 give us a contravariant functor 𝖢𝔊 from the category 𝔊-𝖱𝖾𝗉 to the category 𝖦𝗅𝖺 of graded Lie algebras,

  • on objects, the functor 𝖢𝔊: 𝔊-𝖱𝖾𝗉 𝖦𝗅𝖺 is defined by

    (23) 𝖢𝔊((V;ρVL,ρVR)) = (C(V,𝔊),{,}V),
  • on morphisms, the functor 𝖢𝔊: 𝔊-𝖱𝖾𝗉 𝖦𝗅𝖺 is defined by

    (24) 𝖢𝔊(WϕV) = (C(V,𝔊),{,}V)Φ(C(W,𝔊),{,}W),

where (W;ρWL,ρWR) and (V;ρVL,ρVR) are representations of the Leibniz algebra (𝔊,[,]𝔊) and ϕHom𝔊-𝖱𝖾𝗉(W,V).

Proof.

Let (V;ρVL,ρVR) be a representation of the Leibniz algebra (𝔊,[,]𝔊). For the identity homomorphism IdV:VV, we have 𝖢𝔊(IdV)=IdC(V,𝔊). Moreover, let ϕ:WV and ψ:VU be two homomorphisms of representations of the Leibniz algebra (𝔊,[,]𝔊). For n1 and θHom(Un,𝔊), we have

𝖢𝔊(ψϕ)θ =(22) θ(ψϕ)n
= (θψn)ϕn
= (𝖢𝔊(ϕ)𝖢𝔊(ψ))θ.

Thus, we deduce that 𝖢𝔊 is a contravariant functor.

As a consequence, the short exact sequence (7)

0VantiϕVψVsym0

associated to the 𝔊-representation (V;ρVL,ρVR) induces homomorphisms of graded Lie algebras

(C(Vsym,𝔊),{,}Vsym)Ψ(C(V,𝔊),{,}V)Φ(C(Vanti,𝔊),{,}Vanti),

which do not necessarily form an exact sequence of graded Lie algebras. By passage to the corresponding Maurer-Cartan sets, we obtain a sequence of induced maps

(25) MC(C(Vsym,𝔊))ΨMC(C(V,𝔊))ΦMC(C(Vanti,𝔊)),

which describe (loosely) how a relative Rota-Baxter operator on the Leibniz algebra (𝔊,[,]𝔊) with respect to (V;ρVL,ρVR) is assembled from a relative Rota-Baxter operator on the Lie algebra 𝔊Lie and a relative averaging operator on 𝔊Lie, according to Theorems 2.8 and 2.12, see also Corollaries 3.6 and 3.9.

The (not necessarily exact) sequence (25) is the effect in the module-variable of the standard short exact sequence. In order to investigate also the effect in the Leibniz-algebra-variable of the standard short exact sequence (26) (which will be the subject of Theorem 3.5), we need some more preparation.

3.2. The controlling algebra of relative Rota-Baxter operators on a Lie algebra

Associated to any Leibniz algebra (𝔊,[,]𝔊), we have an exact sequence of Leibniz algebras

(26) 0𝖫𝖾𝗂𝖻(𝔊)𝔦𝔊pr𝔊Lie0.

It is an abelian extension of 𝔊Lie by 𝖫𝖾𝗂𝖻(𝔊). Moreover the induced representation of 𝔊Lie on 𝖫𝖾𝗂𝖻(𝔊) is an antisymmetric representation. Observe that the sequence (26) is the special case of the sequence (7) for the adjoint representation of 𝔊.

Let (V;ρVL,ρVR) be a representation of a Leibniz algebra (𝔊,[,]𝔊). Thus, there is exactly one Lie algebra homomorphism θV:𝔊Lie𝔤𝔩(V), such that following diagram of Leibniz algebra homomorphisms commutes:

(31)

Consider the category 𝔊-𝖲𝖱𝖾𝗉 of symmetric representations of the Leibniz algebra (𝔊,[,]𝔊), which is a subcategory of 𝔊-𝖱𝖾𝗉. Then we have the following functors

s:𝔊-𝖲𝖱𝖾𝗉𝖫𝖾𝗂𝖻(𝔊)-𝖲𝖱𝖾𝗉
𝒫s:𝔊-𝖲𝖱𝖾𝗉𝔊Lie-𝖲𝖱𝖾𝗉

which are given by

  • the functor s: 𝔊-𝖲𝖱𝖾𝗉 𝖫𝖾𝗂𝖻(𝔊)-𝖲𝖱𝖾𝗉, which is defined on objects and on morphisms respectively by

    (32) s((V;ρVL,ρVL)) = (V;ρVL𝔦,ρVL𝔦),
    (33) s(WϕV) = WϕV,
  • the functor 𝒫s: 𝔊-𝖲𝖱𝖾𝗉 𝔊Lie-𝖲𝖱𝖾𝗉, which is defined on objects and on morphisms respectively by

    (34) 𝒫s((V;ρVL,ρVL)) = (V;θV,θV),
    (35) 𝒫s(WϕV) = WϕV,

for symmetric representations (W;ρWL,ρWL) and (V;ρVL,ρVL) of the Leibniz algebra (𝔊,[,]𝔊) and ϕHom𝔊-𝖲𝖱𝖾𝗉(W,V).

Thus, we have three functors 𝖢𝖫𝖾𝗂𝖻(𝔊)s,𝖢𝔊 and 𝖢𝔊Lie𝒫s from 𝔊-𝖲𝖱𝖾𝗉 to 𝖦𝗅𝖺, where the functor 𝖢𝔊 is given in Theorem 3.4. Moreover, for any symmetric representation (V;ρVL,ρVL) of the Leibniz algebra (𝔊,[,]𝔊), we define

αV:(𝖢𝖫𝖾𝗂𝖻(𝔊)s)(V;ρVL,ρVL)𝖢𝔊(V;ρVL,ρVL)

and

βV:𝖢𝔊(V;ρVL,ρVL)(𝖢𝔊Lie𝒫s)(V;ρVL,ρVL)

as following:

(36) αV(g) = g,gHom(nV,𝖫𝖾𝗂𝖻(𝔊)),
(37) βV(f) = prf,fHom(nV,𝔊).
Theorem 3.5.

With the above notations, α is a natural transformation from the functor 𝖢𝖫𝖾𝗂𝖻(𝔊)s to 𝖢𝔊, and β is a natural transformation from the functor 𝖢𝔊 to 𝖢𝔊Lie𝒫s. Moreover, for any symmetric representation (V;ρVL,ρVL) of the Leibniz algebra (𝔊,[,]𝔊), we have the following short exact sequence of graded Lie algebras:

(38) 0(𝖢𝖫𝖾𝗂𝖻(𝔊)s)(V;ρVL,ρVL)αV𝖢𝔊(V;ρVL,ρVL)βV(𝖢𝔊Lie𝒫s)(V;ρVL,ρVL) 0.
Proof.

Since 𝖫𝖾𝗂𝖻(𝔊) is an ideal of the Leibniz algebra (𝔊,[,]𝔊), it follows that C(V,𝖫𝖾𝗂𝖻(𝔊)) is a subalgebra of 𝖢𝔊(V;ρVL,ρVL). Thus the linear embedding map αV is a graded Lie algebra homomorphism. Let ϕ:WV be a homomorphism of symmetric representations of (𝔊,[,]𝔊). It is straightforward to obtain the following commutative diagram:

which implies that α is a natural transformation.

Since pr is a Leibniz algebra homomorphism and the definition (15) of θV, we deduce that βV is a homomorphism of graded Lie algebras. Moreover, let ϕ:WV be a homomorphism of symmetric representations of (𝔊,[,]𝔊) and fHom(nV,𝔊). Then we have

(prf)ϕn=pr(fϕn),

which implies that we have the following commutative diagram:

Thus, β is a natural transformation.

Since V is vector space over the field 𝕂, for any positive integer n, the functor Hom(nV,) is an exact functor from the category of vector spaces over 𝕂 to itself. Moreover, we have αV=𝔦 and βV=pr. By the exact sequence of Leibniz algebras (26), we obtain the short exact sequence of graded Lie algebras (38). The proof is finished.

Corollary 3.6.

Let (V;ρVL,ρVL) be a symmetric representation of a Leibniz algebra (𝔊,[,]𝔊). The Maurer-Cartan elements of the graded Lie algebra (𝖢𝔊Lie𝒫s)(V;ρVL,ρVL) are exactly relative Rota-Baxter operators on the Lie algebra 𝔊Lie with respect to the representation (V,θV) given in (15).

Let (V;ρ) be a representation of a Lie algebra (𝔤,[,]𝔤). Then (V;ρ,ρ) is a symmetric representation of the Leibniz algebra (𝔤,[,]𝔤). By 𝖫𝖾𝗂𝖻(𝔤)=0, we obtain a graded Lie algebra 𝖢𝔤(V;ρ,ρ) whose Maurer-Cartan elements are the relative Rota-Baxter operators on the Lie algebra (𝔤,[,]𝔤) with respect to the representation (V;ρ). On the other hand, in [44] the authors construct another graded Lie algebra on the graded vector subspace 𝒞(V,𝔤):=k1Hom(kV,𝔤), where the graded Lie bracket is given by

P,Q(v1,v2,,vm+n)
= σ𝕊(n,1,m1)(1)σP(ρ(Q(vσ(1),,vσ(n)))vσ(n+1),vσ(n+2),,vσ(m+n))
+(1)mnσ𝕊(m,1,n1)(1)σQ(ρ(P(vσ(1),,vσ(m)))vσ(m+1),vσ(m+2),,vσ(m+n))
(1)mnσ𝕊(m,n)(1)σ[P(vσ(1),,vσ(m)),Q(vσ(m+1),,vσ(m+n))]𝔤

for all PHom(mV,𝔤) and QHom(nV,𝔤). Its Maurer-Cartan elements are also relative Rota-Baxter operators on 𝔤 with respect to the representation (V;ρ).

Proposition 3.7.

The graded Lie algebra (𝒞(V,𝔤),,) is a subalgebra of 𝖢𝔤(V;ρ,ρ).

Proof.

Let (V;ρ) be a representation of a Lie algebra (𝔤,[,]𝔤). The semidirect product Lie algebra 𝔤ρV is the same as the semidirect product Leibniz algebra 𝔤ρ,ρV. We denote this semidirect product multiplication by μ. Recall that the Nijenhuis-Richardson bracket [,]𝖭𝖱 associated to the direct sum vector space 𝔤V gives rise to a graded Lie algebra (k1Hom(k(𝔤V),𝔤V),[,]𝖭𝖱). The graded Lie algebra (𝒞(V,𝔤),,) is also obtained via the derived bracket [27]:

P,Q=(1)m1[[μ,P]𝖭𝖱,Q]𝖭𝖱,PHom(mV,𝔤),QHom(nV,𝔤).

Since (k1Hom(k(𝔤V),𝔤V),[,]𝖭𝖱) is a graded Lie subalgebra of (n1Hom(n(𝔤V),𝔤V),[,]𝖡), we deduce that

P,Q = (1)m1[[μ,P]𝖭𝖱,Q]𝖭𝖱=(1)m1[[μ,P]𝖡,Q]𝖡={P,Q}V,

which implies that (𝒞(V,𝔤),,) is a subalgebra of (C(V,𝔤),{,}V)=𝖢𝔤(V;ρ,ρ).

3.3. The controlling algebras of relative averaging operators on a Lie algebra

Denote by 𝔊-𝖠𝖲𝖱𝖾𝗉 the category of antisymmetric representations of the Leibniz algebra (𝔊,[,]𝔊), which is a subcategory of 𝔊-𝖱𝖾𝗉. Then we have the following functors

a:𝔊-𝖠𝖲𝖱𝖾𝗉𝖫𝖾𝗂𝖻(𝔊)-𝖠𝖲𝖱𝖾𝗉
𝒫a:𝔊-𝖠𝖲𝖱𝖾𝗉𝔊Lie-𝖠𝖲𝖱𝖾𝗉

which are given by

  • the functor a: 𝔊-𝖠𝖲𝖱𝖾𝗉 𝖫𝖾𝗂𝖻(𝔊)-𝖠𝖲𝖱𝖾𝗉, which is defined on objects and on morphisms respectively by

    (42) a((V;ρVL,0)) = (V;ρVL𝔦,0),
    (43) a(WϕV) = WϕV,
  • the functor 𝒫a: 𝔊-𝖠𝖲𝖱𝖾𝗉 𝔊Lie-𝖠𝖲𝖱𝖾𝗉, which is defined on objects and on morphisms respectively by

    (44) 𝒫a((V;ρVL,0)) = (V;θV,0),
    (45) 𝒫a(WϕV) = WϕV,

for antisymmetric representations (W;ρWL,0) and (V;ρVL,0) of the Leibniz algebra (𝔊,[,]𝔊) and ϕHom𝔊-𝖠𝖲𝖱𝖾𝗉(W,V).

Thus, we have three functors 𝖢𝖫𝖾𝗂𝖻(𝔊)a,𝖢𝔊 and 𝖢𝔊Lie𝒫a from the category 𝔊-𝖠𝖲𝖱𝖾𝗉 to the category 𝖦𝗅𝖺. Moreover, for any antisymmetric representation (V;ρVL,0) of the Leibniz algebra (𝔊,[,]𝔊), we define

αV:(𝖢𝖫𝖾𝗂𝖻(𝔊)a)(V;ρVL,0)𝖢𝔊(V;ρVL,0)

and

βV:𝖢𝔊(V;ρVL,0)(𝖢𝔊Lie𝒫a)(V;ρVL,0)

as following:

αV(g) = g,gHom(nV,𝖫𝖾𝗂𝖻(𝔊)),
βV(f) = prf,fHom(nV,𝔊).

Similar to Theorem 3.5, we have the following result.

Theorem 3.8.

With above notations, α is a natural transformation from the functor 𝖢𝖫𝖾𝗂𝖻(𝔊)a to 𝖢𝔊, and β is a natural transformation from the functor 𝖢𝔊 to 𝖢𝔊Lie𝒫a. Moreover, for any antisymmetric representation (V;ρVL,0) of the Leibniz algebra (𝔊,[,]𝔊), we have the following short exact sequence of graded Lie algebras:

0(𝖢𝖫𝖾𝗂𝖻(𝔊)a)(V;ρVL,0)αV𝖢𝔊(V;ρVL,0)βV(𝖢𝔊Lie𝒫a)(V;ρVL,0) 0.
Proof.

The proof is parallel to that of Theorem 3.5, so we omit the details.

Corollary 3.9.

Let (V;ρVL,0) be an antisymmetric representation of a Leibniz algebra (𝔊,[,]𝔊). The Maurer-Cartan elements of the graded Lie algebra (𝖢𝔊Lie𝒫a)(V;ρVL,0) are exactly relative averaging operators on the Lie algebra 𝔊Lie with respect to the representation (V,θV) given in (15).

Let (V;ρ) be a representation of a Lie algebra (𝔤,[,]𝔤). Then (V;ρ,0) is an antisymmetric representation of the Leibniz algebra (𝔤,[,]𝔤). By 𝖫𝖾𝗂𝖻(𝔤)=0, we obtain a graded Lie algebra 𝖢𝔤(V;ρ,0) whose Maurer-Cartan elements are the relative averaging operators on the Lie algebra (𝔤,[,]𝔤) with respect to the representation (V;ρ). This graded Lie algebra is exactly the same as the one given in [41].

4. Relations between the cohomologies

In this section, we establish the relations between the cohomology groups of a relative Rota-Baxter operator on a Leibniz algebra with respect to a symmetric (resp. antisymmetric) representation and the cohomology groups of the induced relative Rota-Baxter (resp. averaging) operator on the canonical Lie algebra.

4.1. Cohomology of relative Rota-Baxter operators on Leibniz algebras

Definition 4.1.

([33]) Let (V;ρL,ρR) be a representation of a Leibniz algebra (𝔊,[,]𝔊). The Loday-Pirashvili cohomology of 𝔊 with coefficients in V is the cohomology of the cochain complex (C(𝔊,V)=k=0+Ck(𝔊,V),), where Ck(𝔊,V)=Hom(k𝔊,V) and the coboundary operator :Ck(𝔊,V)Ck+1(𝔊,V) is defined by

(f)(x1,,xk+1) = i=1k(1)i+1ρL(xi)f(x1,,xi^,,xk+1)+(1)k+1ρR(xk+1)f(x1,,xk)
+1i<jk+1(1)if(x1,,xi^,,xj1,[xi,xj]𝔊,xj+1,,xk+1),

for all x1,,xk+1𝔊. The resulting cohomology is denoted by (𝔊,V).

Theorem 4.2.

([47]) Let T be a relative Rota-Baxter operator on a Leibniz algebra (𝔊,[,]𝔊) with respect to (V;ρL,ρR). Define [,]T:2VV by

(46) [u,v]T=ρL(Tu)v+ρR(Tv)u,u,vV.

Then (V,[,]T) is a Leibniz algebra. Moreover, define ϱL,ϱR:V𝔤𝔩(𝔊) by

(47) ϱL(u)x=[Tu,x]𝔊TρR(x)u,ϱR(u)x=[x,Tu]𝔊TρL(x)u,uV,x𝔊.

Then (𝔊;ϱL,ϱR) is a representation of the Leibniz algebra (V,[,]T).

Let T:Cn(V,𝔊)Cn+1(V,𝔊) be the corresponding Loday-Pirashvili coboundary operator of the Leibniz algebra (V,[,]T) with coefficients in the representation (𝔊;ϱL,ϱR). More precisely, T:Cn(V,𝔊)Cn+1(V,𝔊) is given by

(Tf)(v1,,vn+1)
= i=1n(1)i+1[Tvi,f(v1,,v^i,,vn+1)]𝔊i=1n(1)i+1TρR(f(v1,,v^i,,vn+1))vi
+(1)n+1[f(v1,,vn),Tvn+1]𝔊+(1)nTρL(f(v1,,vn))vn+1
+1i<jn+1(1)if(v1,,v^i,,vj1,ρL(Tvi)vj+ρR(Tvj)vi,vj+1,,vn+1).
Definition 4.3.

([47]) Let T be a relative Rota-Baxter operator on a Leibniz algebra (𝔊,[,]𝔊) with respect to a representation (V;ρL,ρR). The cohomology of the cochain complex

(C(V,𝔊)=k=0+Ck(V,𝔊),T)

is taken to be the cohomology for the relative Rota-Baxter operator T.

We denote the set of k-cocycles by 𝒵RBk(T), the set of k-coboundaries by RBk(T) and the k-th cohomology group by

(48) RBk(T)=𝒵RBk(T)/RBk(T).

Up to a sign, the coboundary operators T coincides with the differential operator {T,}V defined by using the Maurer-Cartan element T.

Theorem 4.4.

([47]) Let T be a relative Rota-Baxter operator on the Leibniz algebra (𝔊,[,]𝔊) with respect to the representation (V;ρL,ρR). Then we have

Tf=(1)n1{T,f}V,fHom(nV,𝔊),n=1,2,.

4.2. The Loday-Pirashvili cohomology of relative Rota-Baxter operators on Lie algebras

In [44] the authors established a cohomology theory of a relative Rota-Baxter operator on a Lie algebra as the Chevalley-Eilenberg cohomology of a Lie algebra. Thus we will refer to it as the Chevalley-Eilenberg cohomology of a relative Rota-Baxter operator on a Lie algebra. More precisely, let T be a relative Rota-Baxter operator on a Lie algebra 𝔤 with respect to a representation (V;ρ). Define ϱ:V𝔤𝔩(𝔤) by

ϱ(u)(x):=[Tu,x]+Tρ(x)(u),x𝔤,uV.

Then (𝔤;ϱ) is a representation of the Lie algebra (V,[,]T) on the vector space 𝔤. We denote by ((V,𝔤),d𝖢𝖤) the Chevalley-Eilenberg cochain complex of the Lie algebra (V,[,]T) with coefficients in the representation (𝔤;ϱ), where (V,𝔤)=k0k(V,𝔤) and k(V,𝔤)=Hom(kV,𝔤). The cohomology of the cochain complex ((V,𝔤),d𝖢𝖤) is called the Chevalley-Eilenberg cohomology of the relative Rota-Baxter operator T. Denote the set of k-cocycles by 𝒵𝖢𝖤k(T) and the set of k-coboundaries by 𝖢𝖤k(T). Denote by

(49) 𝖢𝖤k(T)=𝒵𝖢𝖤k(T)/𝖢𝖤k(T),k0,

the corresponding k-th cohomology group.

As explained in Section 2, a relative Rota-Baxter operator T:V𝔤 on a Lie algebra (𝔤,[,]𝔤) with respect to a representation (V;ρ) can be viewed as a special relative Rota-Baxter operator on the Leibniz algebra (𝔤,[,]𝔤) with respect to the symmetric representation (V;ρ,ρ). Thus, the above approach to define the cohomology of a relative Rota-Baxter operator on a Leibniz algebra can be applied to define a cohomology of a relative Rota-Baxter operator on a Lie algebra. More precisely, let T:V𝔤 be a relative Rota-Baxter operator on a Lie algebra (𝔤,[,]𝔤) with respect to a representation (V;ρ). Then (V,[,]T) is a Leibniz algebra, where the Leibniz bracket [,]T is defined by

[u,v]T=ρ(Tu)vρ(Tv)u,u,vV.

Moreover, (𝔤;ϱL,ϱR) is a symmetric representation of the Leibniz algebra (V,[,]T), where ϱL,ϱR:V𝔤𝔩(𝔤) is defined by

ϱL(u)x=[Tu,x]𝔤+Tρ(x)u,ϱR(u)x=[x,Tu]𝔤Tρ(x)u.

Let T:Cn(V,𝔤)Cn+1(V,𝔤) be the corresponding Loday-Pirashvili coboundary operator of the Leibniz algebra (V,[,]T) with coefficients in the symmetric representation (𝔤;ϱL,ϱR).

Definition 4.5.

Let T:V𝔤 be a relative Rota-Baxter operator on a Lie algebra (𝔤,[,]𝔤) with respect to a representation (V;ρ). The cohomology of the cochain complex (n=0+Cn(V,𝔤),T) is called the Loday-Pirashvili cohomology of the relative Rota-Baxter operator T.

We denote the set of k-cocycles by 𝒵𝖫𝖯k(T), the set of k-coboundaries by 𝖫𝖯k(T) and the k-th cohomology group by

(50) 𝖫𝖯k(T)=𝒵𝖫𝖯k(T)/𝖫𝖯k(T).
Remark 4.6.

Note that the monomorphism k(V,𝔤)=Hom(kV,𝔤)Ck(V,𝔤)=Hom(kV,𝔤) induced from the natural epimorphism kVkV induces isomorphisms

𝖢𝖤0(T)𝖫𝖯0(T)and𝖢𝖤1(T)𝖫𝖯1(T)

and a long exact sequence

0 𝖢𝖤2(T)𝖫𝖯2(T)rel0(T)
𝖢𝖤3(T)𝖫𝖯3(T)rel1(T

where rel(T) is the cohomology of the relative complex Crel(V,𝔤) which is by definition (up to a degree shift) the cokernel of the monomorphism Hom(kV,𝔤)Hom(kV,𝔤). These matters as well as a spectral sequence linking rel(T) and 𝖫𝖯(T) in still another way can be found in [18]. In particular, Theorem 2.6 in [18] shows how the vanishing of Chevalley-Eilenberg cohomology implies the vanishing of Loday-Pirashvili cohomology.

Remark 4.7.

Note furthermore that the comparison between Chevalley-Eilenberg cohomologies and Loday-Pirashvili cohomologies makes sense more generally for any relative Rota-Baxter operator T:V𝔊 over a Leibniz algebra 𝔊 as soon as the representation of 𝔊 on V is symmetric, because then the bracket [,]T on V is a Lie bracket.

In case the induced representation of (V,[,]T) on 𝔊 is not symmetric, one may use the short exact sequence

0𝔊anti𝔊𝔊sym0

and the relation of Loday-Pirashvili cohomology with values in an antisymmetric representation to the cohomology with values in a symmetric representation (see [18]) in order to reduce the computations to the case of symmetric coefficients.

Let T:V𝔊 be a relative Rota-Baxter operator on a Leibniz algebra 𝔊 with respect to a symmetric representation (V;ρL,ρR=ρL). By Proposition 2.6, T¯:=prT:V𝔊Lie is a relative Rota-Baxter operator on the Lie algebra 𝔊Lie with respect to the representation (V;θ) given by (15). Now we establish the relationship between the cohomology groups of the relative Rota-Baxter operator T and the Loday-Pirashvili cohomology groups of the relative Rota-Baxter operator T¯.

Lemma 4.8.

Let T:V𝔊 be a relative Rota-Baxter operator on a Leibniz algebra 𝔊 with respect to a symmetric representation (V;ρL,ρR=ρL). Then (𝖫𝖾𝗂𝖻(𝔊);ϱL,ϱR) is a subrepresentation of (𝔊;ϱL,ϱR).

Proof.

Since ρR=ρL, by (14), we have

ϱL(u)[y,y]𝔊=[Tu,[y,y]𝔊]𝔊+ρL([y,y]𝔊)u=[[Tu,y]𝔊,y]𝔊+[y,[Tu,y]𝔊]𝔊,

for all y𝔊, which implies that ϱL(u)(𝖫𝖾𝗂𝖻(𝔊))𝖫𝖾𝗂𝖻(𝔊).

By the fact that 𝖫𝖾𝗂𝖻(𝔊) is contained in the left center of the Leibniz algebra 𝔊, we have

ϱR(u)[y,y]𝔊=[[y,y]𝔊,Tu]𝔊TρL([y,y]𝔊)u=0,

which implies that ϱR(u)(𝖫𝖾𝗂𝖻(𝔊))=0𝖫𝖾𝗂𝖻(𝔊). Therefore, (𝖫𝖾𝗂𝖻(𝔊);ϱL,ϱR) is a subrepresentation of (𝔊;ϱL,ϱR).

Note that even though the Leibniz algebra (V,[,]T) becomes a Lie algebra if T is a relative Rota-Baxter operator on a Leibniz algebra 𝔊 with respect to a symmetric representation, the representation given in Theorem 4.2 is still a representation of (V,[,]T) as a Leibniz algebra, and ϱRϱL. But the quotient representation (𝔊Lie;ϱ¯L,ϱ¯R) is symmetric, where ϱ¯L:V𝔤𝔩(𝔊Lie) and ϱ¯R:V𝔤𝔩(𝔊Lie) are given by

(51) ϱ¯L(u)x¯=ϱL(u)x¯,ϱ¯R(u)x¯=ϱR(u)x¯,uV,x𝔊.
Proposition 4.9.

Let T:V𝔊 be a relative Rota-Baxter operator on a Leibniz algebra 𝔊 with respect to a symmetric representation (V;ρL,ρR=ρL). Then the quotient representation (𝔊Lie;ϱ¯L,ϱ¯R) is a symmetric representation of (V,[,]T) as a Leibniz algebra, i.e. ϱ¯R=ϱ¯L.

Proof.

For all uV,x𝔊, we have

ϱL(u)x+ϱR(u)x = [Tu,x]𝔊TρR(x)u+[x,Tu]𝔊TρL(x)u
= [Tu,x]𝔊+[x,Tu]𝔊𝖫𝖾𝗂𝖻(𝔊).

Thus, we deduce that ϱ¯R=ϱ¯L.

Let ¯T:Cn(V,𝔊Lie)Cn+1(V,𝔊Lie) be the corresponding Loday-Pirashvili coboundary operator of the Leibniz algebra (V,[,]T) with coefficients in the quotient representation (𝔊Lie;ϱ¯L,ϱ¯R).

Proposition 4.10.

Let T be a relative Rota-Baxter operator on a Leibniz algebra (𝔊,[,]𝔊) with respect to a symmetric representation (V;ρL,ρR=ρL). Then

¯T=T¯.

That is, the Loday-Pirashvili cochain complex (n=0+Cn(V,𝔊Lie),T¯) associated to the relative Rota-Baxter operator T¯ and the cochain complex (n=0+Cn(V,𝔊Lie),¯T) obtained by using the quotient representation (𝔊Lie;ϱ¯L,ϱ¯R) are the same.

Proof.

Obviously, the spaces of n-cochains are the same. Thus we only need to show that the induced Leibniz algebra structure [,]T¯ on V by the relative Rota-Baxter operator T¯ is the same as [,]T, and the representations (𝔊Lie;ϱL,ϱR) and (𝔊Lie;ϱ¯L,ϱ¯R) are the same.

For all u,vV, by (15), we have

[u,v]T¯=θ(T¯u)vθ(T¯v)u=ρL(Tu)vρL(Tv)u=[u,v]T.

For any uV,x𝔊, we have

ϱL(u)(x¯)=[T¯u,x¯]𝔊Lie+T¯θ(x¯)u=[Tu,x]𝔊¯+TρL(x)u¯=ϱ¯L(u)x¯,
ϱR(u)(x¯)=[x¯,T¯u]𝔊LieT¯θ(x¯)u=[x,Tu]𝔊¯TρL(x)u¯=ϱ¯R(u)x¯.

Thus ¯T=T¯.

Let T:V𝔊 be a relative Rota-Baxter operator on a Leibniz algebra 𝔊 with respect to a symmetric representation (V;ρL,ρR=ρL). By Lemma 4.8, we have the following short exact sequence of representations of the Leibniz algebra (V,[,]T):

(52) 0(𝖫𝖾𝗂𝖻(𝔊);ϱL,ϱR)𝔦(𝔊;ϱL,ϱR)pr(𝔊Lie;ϱ¯L,ϱ¯R) 0.

Moreover, we have the following result:

Theorem 4.11.

Let T be a relative Rota-Baxter operator on a Leibniz algebra (𝔊,[,]𝔊) with respect to a symmetric representation (V;ρL,ρR=ρL). Then there is a short exact sequence of the cochain complexes:

00Hom(nV,𝖫𝖾𝗂𝖻(𝔊))αVHom(n+1V,𝖫𝖾𝗂𝖻(𝔊))αVHom(nV,𝔊)βVTHom(n+1V,𝔊)βVHom(nV,𝔊Lie)T¯Hom(n+1V,𝔊Lie),00

where αV and βV are given by (36) and (37) respectively.

Consequently, there is a long exact sequence of the cohomology groups:

(53) n(V,𝖫𝖾𝗂𝖻(𝔊))n(αV)RBn(T)n(βV)𝖫𝖯n(T¯)cnn+1(V,𝖫𝖾𝗂𝖻(𝔊)),

where the connecting map cn is defined by cn([h])=[αV1(T(βV1(h)))], for all [h]𝖫𝖯n(T¯).

Proof.

For any gHom(nV,𝖫𝖾𝗂𝖻(𝔊)) and v1,,vn+1V, we have

(αV(g))(v1,,vn+1)
=(36) (g)(v1,,vn+1)
= i=1n(1)i+1[Tvi,g(v1,,v^i,,vn+1)]𝔊+(1)nTρL(g(v1,,vn))vn+1
+1i<jn+1(1)ig(v1,,v^i,,vj1,ρL(Tvi)vj+ρR(Tvj)vi,vj+1,,vn+1)
= (T(αV(g)))(v1,,vn+1),

which implies that αV is a homomorphism of cochain complexes. For n=0 and x𝔊,vV, we have

(βV(Tx))(v)=βV([x,Tv]𝔊+TρL(x)v)=(37)ϱ¯R(v)x¯=(T¯βV(x))(v).

For n1, by Theorem 4.4 and Theorem 3.5, for all fHom(nV,𝔊), we have

βV(Tf)=(1)n1βV{T,f}V=(1)n1{βV(T),βV(f)}V=T¯βV(f).

Thus, we deduce that βV is a homomorphism of cochain complexes. Since V is vector space over the field 𝕂, for any positive integer n, the functor Hom(nV,) is an exact functor from the category of vector spaces over 𝕂 to itself. Moreover, we have αV=𝔦 and βV=pr. By the short exact sequence (52) of representations of the Leibniz algebra (V,[,]T), we obtain the short exact sequence of the above cochain complexes. The proof is finished.

Remark 4.12.

In the situation of Theorem 4.11, the short exact sequence of cochain complexes can be completed into a diagram

0000(V,𝖫𝖾𝗂𝖻(𝔊))C(V,𝖫𝖾𝗂𝖻(𝔊))Crel(V,𝖫𝖾𝗂𝖻(𝔊))00(V,𝔊)C(V,𝔊)Crel(V,𝔊)00(V,𝔊Lie)C(V,𝔊Lie)Crel(V,𝔊Lie)0000

which leads then to a corresponding exact diagram in cohomology (starting from degree 2, while there are isomorphisms in degree 0 and 1 as before).

4.3. Cohomology of relative averaging operators on Lie algebras

As explained in Section 2, a relative averaging operator T:V𝔤 on a Lie algebra (𝔤,[,]𝔤) with respect to a representation (V;ρ) can be viewed as a special relative Rota-Baxter operator on the Leibniz algebra (𝔤,[,]𝔤) with respect to the antisymmetric representation (V;ρ,0). Thus, the above approach to define the cohomology of relative Rota-Baxter operators on Leibniz algebras can be applied to define the cohomology of relative averaging operators. More precisely, let T:V𝔤 be a relative averaging operator on a Lie algebra (𝔤,[,]𝔤) with respect to a representation (V;ρ). Then (V,[,]T) is a Leibniz algebra, where the Leibniz bracket [,]T is defined by

[u,v]T=ρ(Tu)v,u,vV.

Moreover, (𝔤;ϱL,ϱR) is a representation of the Leibniz algebra (V,[,]T), where ϱL,ϱR:V𝔤𝔩(𝔤) is defined by

ϱL(u)x=[Tu,x]𝔤,ϱR(u)x=[x,Tu]𝔤Tρ(x)u.

Let T:Cn(V,𝔤)Cn+1(V,𝔤) be the corresponding Loday-Pirashvili coboundary operator of the Leibniz algebra (V,[,]T) with coefficients in the representation (𝔤;ϱL,ϱR). More precisely, T:Cn(V,𝔤)Cn+1(V,𝔤) is given by

(Tf)(v1,,vn+1) = i=1n(1)i+1[Tvi,f(v1,,v^i,,vn+1)]𝔤
+(1)n+1[f(v1,,vn),Tvn+1]𝔤+(1)nTρ(f(v1,,vn))vn+1
+1i<jn+1(1)if(v1,,v^i,,vj1,ρ(Tvi)vj,vj+1,,vn+1).
Definition 4.13.

Let T:V𝔤 be a relative averaging operator on a Lie algebra (𝔤,[,]𝔤) with respect to a representation (V;ρ). The cohomology of the cochain complex (n=0+Cn(V,𝔤),T) is defined to be the cohomology of the relative averaging operator T.

We denote the set of k-cocycles by 𝒵AOk(T), the set of k-coboundaries by AOk(T) and the k-th cohomology group by

(54) AOk(T)=𝒵AOk(T)/AOk(T).

See [41] for more details about the cohomology theory of a relative averaging operator on a Lie algebra.

Let T:V𝔊 be a relative Rota-Baxter operator on the Leibniz algebra 𝔊 with respect to an antisymmetric representation (V;ρL,ρR=0). By Proposition 2.10, T¯:=prT:V𝔊Lie is a relative averaging operator on the Lie algebra 𝔊Lie with respect to the representation (V;θ) given by (15). Now we establish the relationship between the cohomology groups of the relative Rota-Baxter operator T and the cohomology groups of the relative averaging operator T¯.

Similarly as Lemma 4.8, we have the following result.

Lemma 4.14.

Let T:V𝔊 be a relative Rota-Baxter operator on a Leibniz algebra 𝔊 with respect to an antisymmetric representation (V;ρL,ρR=0). Then (𝖫𝖾𝗂𝖻(𝔊);ϱL,ϱR) is a subrepresentation of (𝔊;ϱL,ϱR).

By Lemma 4.14, we have the quotient representation (𝔊Lie;ϱ¯L,ϱ¯R), where ϱ¯L:V𝔤𝔩(𝔊Lie) and ϱ¯R:V𝔤𝔩(𝔊Lie) are given by (51). Let ¯T:Cn(V,𝔊Lie)Cn+1(V,𝔊Lie) be the corresponding Loday-Pirashvili coboundary operator of the Leibniz algebra (V,[,]T) with coefficients in the quotient representation (𝔊Lie;ϱ¯L,ϱ¯R). Similar to Proposition 4.10, we have the following result.

Proposition 4.15.

Let T be a relative Rota-Baxter operator on a Leibniz algebra (𝔊,[,]𝔊) with respect to an antisymmetric representation (V;ρL,ρR=0). Then

¯T=T¯.

That is, the cochain complex (n=0+Cn(V,𝔊Lie),T¯) associated to the relative averaging operator T¯ and the cochain complex (n=0+Cn(V,𝔊Lie),¯T) obtained by using the quotient representation (𝔊Lie;ϱ¯L,ϱ¯R) are the same.

Proof.

Obviously, the spaces of n-cochains are the same. Thus we only need to show that the induced Leibniz algebra structure [,]T¯ on V by the relative averaging operator T¯ is the same as [,]T, and the representations (𝔊Lie;ϱL,ϱR) and (𝔊Lie;ϱ¯L,ϱ¯R) are the same.

For all u,vV, by (15), we have

[u,v]T¯=θ(T¯u)v=ρL(Tu)v=[u,v]T.

For any uV,x𝔊, we have

ϱL(u)x¯=[T¯u,x¯]𝔊Lie=[Tu,x]𝔊¯=ϱ¯L(u)x¯,
ϱR(u)x¯=[x¯,T¯u]𝔊LieT¯θ(x¯)u=[x,Tu]𝔊¯TρL(x)u¯=ϱ¯R(u)x¯.

Thus, ¯T=T¯.

Let T:V𝔊 be a relative Rota-Baxter operator on a Leibniz algebra 𝔊 with respect to an antisymmetric representation (V;ρL,ρR=0). By Lemma 4.14, we have the following short exact sequence of representations of the Leibniz algebra (V,[,]T):

(55) 0(𝖫𝖾𝗂𝖻(𝔊);ϱL,ϱR)𝔦(𝔊;ϱL,ϱR)pr(𝔊Lie;ϱ¯L,ϱ¯R) 0.

Similar to Theorem 4.11, we have the following result describing the relation among various cohomology groups.

Theorem 4.16.

Let T be a relative Rota-Baxter operator on a Leibniz algebra (𝔊,[,]𝔊) with respect to an antisymmetric representation (V;ρL,ρR=0). Then we have a long exact sequence of cohomology groups:

(56) n(V,𝖫𝖾𝗂𝖻(𝔊))n(αV)RBn(T)n(βV)AOn(T¯)cnn+1(V,𝖫𝖾𝗂𝖻(𝔊)).

4.4. Some computations of cohomologies

Example 4.17.

Let 𝔤𝔩(n,) be the Lie algebra of all n×n matrices over . Then T=Id:𝔤𝔩(n,)𝔤𝔩(n,) is a relative Rota-Baxter operator on the Lie algebra 𝔤𝔩(n,) with respect to the representation (𝔤𝔩(n,);ρ), where ρ:𝔤𝔩(n,)𝔤𝔩(𝔤𝔩(n,)) is defined by

ρ(A)B=AB,A,B𝔤𝔩(n,).

Then (𝔤𝔩(n,),[,]T) is the original Lie algebra 𝔤𝔩(n,). Moreover, (𝔤𝔩(n,);ϱ) is a representation of the Lie algebra 𝔤𝔩(n,), where ϱ:𝔤𝔩(n,)𝔤𝔩(𝔤𝔩(n,)) is defined by

(57) ϱ(A)B:=[TA,B]+Tρ(B)(A)=AB=ρ(A)B,A,B𝔤𝔩(n,).

Thus, the cohomology of the relative Rota-Baxter operator T is the Chevalley-Eilenberg cohomology of 𝔤𝔩(n,) with coefficients in the representation (𝔤𝔩(n,);ρ). In fact, the Lie algebra 𝔤𝔩(n,) is the direct sum 𝔤𝔩(n,)=𝔰𝔩(n,), where the factor is the center and consists of diagonal matrices. The Hochschild-Serre formula [22] reads then

𝖢𝖤n(𝔤𝔩(n,),𝔤𝔩(n,))=p+q=n𝖢𝖤p(𝔰𝔩(n,))𝖢𝖤q(,𝔤𝔩(n,))𝔰𝔩(n,)
=𝖢𝖤n(𝔰𝔩(n,))𝖢𝖤0(,𝔤𝔩(n,))𝔰𝔩(n,)𝖢𝖤n1(𝔰𝔩(n,))𝖢𝖤1(,𝔤𝔩(n,))𝔰𝔩(n,).

The invariant space 𝖢𝖤0(,𝔤𝔩(n,))𝔰𝔩(n,) is zero, because 𝖢𝖤0(,𝔤𝔩(n,))=0. Furthermore, we have 𝖢𝖤1(,𝔤𝔩(n,))=0, because all elements are coboundaries. Therefore all the cohomology groups are zero.

Example 4.18.

Let 𝔱(n,) be the Lie algebras of all n×n upper triangular matrices over . Then T=Id:𝔱(n,)𝔱(n,) is a relative Rota-Baxter operator on the Lie algebra 𝔱(n,) with respect to the representation (𝔱(n,);ρ), where ρ:𝔱(n,)𝔤𝔩(𝔱(n,)) is defined by

ρ(A)B=AB,A,B𝔱(n,).

Thus, the cohomology of the relative Rota-Baxter operator T is the Chevalley-Eilenberg cohomology of 𝔱(n,) with coefficients in the representation (𝔱(n,);ρ). Recall that the Lie algebra 𝔱(n,) is an extension

0𝔫(n,)𝔱(n,)𝔡(n,)0,

where 𝔫(n,) is the Lie algebras of all n×n upper triangular nilpotent matrices and 𝔡(n,) is the Lie algebras of all n×n diagonal matrices.

The Hochschild-Serre spectral sequence [22] associated to this extension converges towards 𝖢𝖤(𝔱(n,),𝔱(n,)) and has as its second page

E2p,q=𝖢𝖤p(𝔡(n,),𝖢𝖤q(𝔫(n,),𝔱(n,))).

Let us compute this E2-term for n=2. The Lie algebra 𝔫(2,) is 1-dimensional and we have only non zero contributions from q=0,1. We have 𝖢𝖤0(𝔫(2,),𝔱(2,))=𝔱(2,)𝔫(2,) of dimension 2 generated by

(1000)and(0100).

From there, we obtain using Dixmier’s Theorem 1 of [17] for the nilpotent Lie algebra 𝔡(2,) with values in a module which does not contains the trivial module

𝖢𝖤p(𝔡(2,),𝖢𝖤0(𝔫(2,),𝔱(2,)))=0,

for all p. On the other hand, 𝖢𝖤1(𝔫(2,),𝔱(2,))𝔡(2,) as all maps in Hom(𝔫(2,),𝔱(2,))𝔱(2,) are cocycles and the coboundaries correspond to 𝔫(2,). From there, we obtain again with Dixmier’s Theorem 1

𝖢𝖤p(𝔡(2,),𝖢𝖤1(𝔫(2,),𝔱(2,)))=0,

for all p. Thus again all the cohomology groups are zero. For general n, the whole cohomology will still vanish in case the cohomology spaces 𝖢𝖤q(𝔫(n,),𝔱(n,)) do not contain the trivial 𝔡(n,)-module.

Example 4.19.

Let 𝔫(n,) be the Lie algebras of all n×n upper triangular nilpotent matrices over . Then T=Id:𝔫(n,)𝔫(n,) is a relative Rota-Baxter operator on the Lie algebra 𝔫(n,) with respect to the representation (𝔫(n,);ρ), where ρ:𝔫(n,)𝔤𝔩(𝔫(n,)) is defined by

ρ(A)B=AB,A,B𝔫(n,).

Thus, the cohomology of the relative Rota-Baxter operator T is the Chevalley-Eilenberg cohomology of 𝔫(n,) with coefficients in the representation (𝔫(n,);ρ). Since 𝔫(n,) is a nilpotent Lie algebra and the 𝔫(n,)-module 𝔫(n,) does contain the trivial 𝔫(n,)-module (namely the submodule generated by the matrix with non-zero entry in the upper right corner), we have by Dixmier’s Theorem 2 of [17] that

dim𝖢𝖤0(𝔫(n,),𝔫(n,))1,dim𝖢𝖤n(n1)/2(𝔫(n,),𝔫(n,))1,

and

dim𝖢𝖤i(𝔫(n,),𝔫(n,))2,0<i<n(n1)2.
Example 4.20.

([41]) Let (𝔊,[,]𝔊) be a Leibniz algebra. Then the natural projection

T=pr:𝔊𝔊Lie

gives a relative averaging operator on the Lie algebra 𝔊Lie with respect to the representation (𝔊;ρ), where ρ:𝔊Lie𝔤𝔩(𝔊) is defined by

ρ(x¯)y=[x,y]𝔊,x,y𝔊.

Then (𝔊,[,]T) is the original Leibniz algebra (𝔊,[,]𝔊). Moreover, (𝔊Lie;ϱL,ϱR) is a representation of the Leibniz algebra (𝔊,[,]T), where ϱL,ϱR:𝔊𝔤𝔩(𝔊Lie) is defined by

ϱL(x)y¯=[x¯,y¯]𝔊Lie,ϱR(x)y¯=[y¯,x¯]𝔊Lie[y,x]𝔊¯=0.

Thus, the cohomology of the relative averaging operator T is the Loday-Pirashvili cohomology of (𝔊,[,]𝔊) with coefficients in the antisymmetric representation (𝔊Lie;ϱL,ϱR).

Let now (𝔊,[,]𝔊) be a finite-dimensional semisimple Leibniz algebra over a field of characteristic zero. By Theorem 4.3 of [18], we obtain that p(𝔊,𝔊Lie)={0} for p2 and an exact sequence

0(𝔊Lie)anti0(𝔊,𝔊Lie)𝔊Lie𝔊LieHom𝔊(𝔊,𝔊Lie)1(𝔊,𝔊Lie)0.

As 𝔊Lie is already antisymmetric and 0(𝔊,𝔊Lie)=𝔊Lie, the first (non-trivial) map in this sequence is an isomorphism. Furthermore 𝔊Lie𝔊Lie={0}, because 𝔊Lie is semisimple. Therefore the last (non-trivial) map in this sequence is also an isomorphism (as in Lemma 1.4 of [18]).

Acknowledgements. This research is supported by NSFC (11922110,12001228).

References

  • [1] M. Aguiar, Pre-Poisson algebras. Lett. Math. Phys. 54 (2000), 263-277.
  • [2] M. Ammar and N. Poncin, Coalgebraic approach to the Loday infinity category, stem differential for 2n-ary graded and homotopy algebras. Ann. Inst. Fourier (Grenoble). 60 (2010), 355-387.
  • [3] C. Bai, O. Bellier, L. Guo and X. Ni, Spliting of operations, Manin products and Rota-Baxter operators. Int. Math. Res. Not. (2013), 485-524.
  • [4] D. Balavoine, Deformations of algebras over a quadratic operad. Contemp. Math. AMS, 202 (1997), 207-234.
  • [5] C. Bai, A unified algebraic approach to the classical Yang-Baxter equation. J. Phys. A: Math. Theor. 40 (2007), 11073-11082.
  • [6] C. Bai, L. Guo and X. Ni, Nonabelian generalized Lax pairs, the classical Yang-Baxter equation and PostLie algebras. Comm. Math. Phys. 297 (2010), 553-596.
  • [7] C. Barnett, Averaging operators in noncommutative Lp spaces. I. Glasgow Math. J. 24 (1983), 71-74.
  • [8] G. Baxter, An analytic problem whose solution follows from a simple algebraic identity. Pacific J. Math. 10 (1960), 731-742.
  • [9] A. Bloh, On a generalization of the concept of Lie algebra. (Russian) Dokl. Akad. Nauk SSSR 165 (1965), 471-473.
  • [10] R. Bonezzi and O. Hohm, Leibniz gauge theories and infinity structures. Comm. Math. Phys. 377 (2020), 2027-2077.
  • [11] M. Bordemann and F. Wagemann, Global integration of Leibniz algebras. J. Lie Theory 27 (2017), 555-567.
  • [12] B. Brainerd, On the structure of averaging operators. J. Math. Anal. Appl. 5 (1962), 347-377.
  • [13] A. Connes and D. Kreimer, Renormalization in quantum field theory and the Riemann-Hilbert problem. I. The Hopf algebra structure of graphs and the main theorem. Comm. Math. Phys. 210 (2000), 249-273.
  • [14] S. Covez, The local integration of Leibniz algebras. Ann. Inst. Fourier (Grenoble) 63 (2013), 1-35.
  • [15] A. Das, Deformations of associative Rota-Baxter operators. J. Algebra 560 (2020), 144-180.
  • [16] B. Dherin and F. Wagemann, Deformation quantization of Leibniz algebras. Adv. Math. 270 (2015), 21-48.
  • [17] J. Dixmier, Cohomologie des algèbres de Lie nilpotentes. Acta Sci. Math. (Szeged) 16 (1955), 246-25
  • [18] J. Feldvoss and F. Wagemann, On Leibniz cohomology. J. Algebra 569 (2021), 276-317.
  • [19] A. Fialowski and A. Mandal, Leibniz algebra deformations of a Lie algebra. J. Math. Phys. 49 (2008), 093511, 11 pp.
  • [20] M. E. Goncharov and P. S. Kolesnikov, Simple finite-dimensional double algebras. J. Algebra 500 (2018), 425-438.
  • [21] L. Guo, An introduction to Rota-Baxter algebra. Surveys of Modern Mathematics, 4. International Press, Somerville, MA; Higher Education Press, Beijing, 2012. xii+226 pp.
  • [22] G. Hochschild and J. P. Serre, Cohomology of Lie algebras. Ann. of Math. (2) 57 (1953), 591-603.
  • [23] O. Hohm and H. Samtleben, Leibniz-Chern-Simons theory and phases of exceptional field theory. Comm. Math. Phys. 369 (2019), 1055-1089.
  • [24] C. B. Huijsmans and B. de Pagter, Averaging operators and positive contractive projections. J. Math. Anal. Appl. 113 (1986), 163-184.
  • [25] M. Kinyon and A. Weinstein, Leibniz algebras, Courant algebroids, and multiplications on reductive homogeneous spaces. Amer. J. Math. 123 (2001), 525-550.
  • [26] P. S. Kolesnikov, Homogeneous averaging operators on semisimple Lie algebras. Algebra Logic 53 (2014), 510-511.
  • [27] Y. Kosmann-Schwarzbach, From Poisson algebras to Gerstenhaber algebras. Ann. Inst. Fourier (Grenoble) 46 (1996), 1243-1274.
  • [28] A. Kotov and T. Strobl, The embedding tensor, Leibniz-Loday algebras, and their higher gauge theories. Comm. Math. Phys. 376 (2020), 235-258.
  • [29] B. A. Kupershmidt, What a classical r-matrix really is. J. Nonlinear Math. Phys. 6 (1999), 448-488.
  • [30] A. Lazarev, Y. Sheng and R. Tang, Deformations and homotopy theory of relative Rota-Baxter Lie algebras. Comm. Math. Phys. 383 (2021), 595-631.
  • [31] J. L. Loday, Une version non commutative des algèbres de Lie: les algèbres de Leibniz. Enseign. Math. (2) 39 (1993), 269-293.
  • [32] J. L. Loday and T. Pirashvili, The tensor category of linear maps and Leibniz algebras. (English summary) Georgian Math. J. 5 (1998), 263-276.
  • [33] J. L. Loday and T. Pirashvili, Universal enveloping algebras of Leibniz algebras and (co)homology. Math. Ann. 296 (1993), 139-158.
  • [34] J. L. Loday, Algebraic K-theory and the conjectural Leibniz K-theory. K-Theory 30 (2003), 105-127.
  • [35] J. Pei, C. Bai and L. Guo, Splitting of Operads and Rota-Baxter Operators on Operads. Appl. Categor. Struct. 25 (2017), 505-538.
  • [36] J. Pei, C. Bai, L. Guo and X. Ni, Replicating of binary operads, Koszul duality, Manin products and average operators. New Trends in Algebras and Combinatorics (2020), 317-353.
  • [37] J. Pei and L. Guo, Averaging algebras, Schroder numbers, rooted trees and operads. J. Algebraic Combin. 42 (2015), 73-109.
  • [38] T. Pirashvili, On Leibniz homology. Ann. Inst. Fourier (Grenoble) 44 (1994), 401-411.
  • [39] G. C. Rota, Ten mathematics problems I will never solve. Mitt. Dtsch. Math. Ver. 2 (1998), 45-52.
  • [40] M. Semonov-Tian-Shansky, What is a classical r-matrix? Funct. Anal. Appl. 17 (1983), 259-272.
  • [41] Y. Sheng, R. Tang and C. Zhu, The controlling L-algebras, cohomologies and homotopy of embedding tensors and Lie-Leibniz triples. Comm. Math. Phys. 386 (2021), 269-304.
  • [42] T. Strobl, Leibniz-Yang-Mills gauge theories and the 2-Higgs mechanism. Phys. Rev. D 99 (2019), 115026.
  • [43] T. Strobl and F. Wagemann, Enhanced Leibniz algebras: structure theorem and induced Lie 2-algebra. Comm. Math. Phys. 376 (2020), 51-79.
  • [44] R. Tang, C. Bai, L. Guo and Y. Sheng, Deformations and their controlling cohomologies of 𝒪-operators. Comm. Math. Phys. 368 (2019), 665-700.
  • [45] R. Tang, C. Bai, L. Guo and Y. Sheng, Homotopy Rota-Baxter operators and post-Lie algebras. J. Noncommut. Geom. 17 (2023), 1-35.
  • [46] R. Tang and Y. Sheng, Leibniz bialgebras, relative Rota-Baxter operators and the classical Leibniz Yang-Baxter equation. J. Noncommut. Geom. 16 (2022), 1179-1211.
  • [47] R. Tang, Y. Sheng and Y. Zhou, Deformations of relative Rota-Baxter operators on Leibniz algebras. Int. J. Geom. Methods Mod. Phys. 17 (2020), 2050174, 21 pp.
  • [48] K. Uchino, Quantum analogy of Poisson geometry, related dendriform algebras and Rota-Baxter operators. Lett. Math. Phys. 85 (2008), 91-109.
  • [49] K. Uchino, Twisting on associative algebras and Rota-Baxter type operators. J. Noncommut. Geom. 4 (2010), 349-379.
  • [50] K. Uchino, Derived bracket construction and Manin products. Lett. Math. Phys. 93 (2010), 37-53.
  • [51] B. Vallette, Manin products, Koszul duality, Loday algebras and Deligne conjecture. J. Reine Angew. Math. 620 (2008), 105-164.
  • [52] K. Wang and G. Zhou, Cohomology theory of averaging algebras, L-structures and homotopy averaging algebras. arXiv:2009.11618.