1,045
Views
8
CrossRef citations to date
0
Altmetric
Original Article

Rota–Baxter operators and post-Lie algebra structures on semisimple Lie algebras

&
Pages 2280-2296 | Received 18 May 2018, Accepted 01 Oct 2018, Published online: 11 Jan 2019

Abstract

Rota–Baxter operators R of weight 1 on n are in bijective correspondence to post-Lie algebra structures on pairs (g,n), where n is complete. We use such Rota–Baxter operators to study the existence and classification of post-Lie algebra structures on pairs of Lie algebras (g,n), where n is semisimple. We show that for semisimple g and n, with g or n simple, the existence of a post-Lie algebra structure on such a pair (g,n) implies that g and n are isomorphic, and hence both simple. If n is semisimple, but g is not, it becomes much harder to classify post-Lie algebra structures on (g,n), or even to determine the Lie algebras g which can arise. Here only the case n=sl2(C) was studied. In this paper, we determine all Lie algebras g such that there exists a post-Lie algebra structure on (g,n) with n=sl2(C)sl2(C).

2010 MATHEMATICS SUBJECT CLASSIFICATION:

1. Introduction

Rota–Baxter operators were introduced by Baxter [Citation3] in 1960 as a formal generalization of integration by parts for solving an analytic formula in probability theory. Such operators R:AA are defined on an algebra A by the identity R(x)R(y)=R(R(x)y+xR(y)+λxy) for all x,yA, where λ is a scalar, called the weight of R. These operators were then further investigated, by Rota [Citation30], Atkinson [Citation1], Cartier [Citation16] and others. In the 1980s, these operators were studied in integrable systems in the context of classical and modified Yang–Baxter equations [Citation33,Citation4]. Since the late 1990s, the study of Rota–Baxter operators has made great progress in many areas, both in theory and in applications [Citation25,2,22,20,21,5,19].

Post-Lie algebras and post-Lie algebra structures also arise in many areas, e.g., in differential geometry and the study of geometric structures on Lie groups. Here, post-Lie algebras arise as a natural common generalization of pre-Lie algebras [Citation23,26,32,6,7,8] and LR-algebras [Citation9,Citation10], in the context of nil-affine actions of Lie groups, see [Citation11]. A detailed account of the differential geometric context of post-Lie algebras is also given in [Citation18]. On the other hand, post-Lie algebras have been introduced by Vallette [Citation34] in connection with the homology of partition posets and the study of Koszul operads. They have been studied by several authors in various contexts, e.g., for algebraic operad triples [Citation28], in connection with modified Yang–Baxter equations, Rota–Baxter operators, universal enveloping algebras, double Lie algebras, R-matrices, isospectral flows, Lie-Butcher series and many other topics [Citation2,Citation18,Citation19]. There are several results on the existence and classification of post-Lie algebra structures, in particular on commutative post-Lie algebra structures [Citation13–15].

It is well-known [Citation2] that Rota–Baxter operators R of weight 1 on n are in bijective correspondence to post-Lie algebra structures on pairs (g,n), where n is complete. In fact, RB-operators always yield PA-structures. So it is possible (and desirable) to use results on RB-operators for the existence and classification of post-Lie algebra structures.

The paper is organized as follows. In Section 2 we give basic definitions of RB-operators and PA-structures on pairs of Lie algebras. We summarize several useful results. For a complete Lie algebra n there is a bijection between PA-structures on (g,n) and RB-operators of weight 1 on n. The PA-structure is given by x·y={R(x),y}. Here we study the kernels of R and R+id. If g and n are not isomorphic, then both R and R+id have a nontrivial kernel. Moreover, if one of g or n is not solvable, then at least one of ker(R) and ker(R+id) is nontrivial.

In Section 3, we complete the classification of PA-structures on pairs of semisimple Lie algebras (g,n), where either g or n is simple. We already have shown the following in [Citation11]. If g is simple, and there exists a PA-structure on (g,n), then also n is simple, and we have gn with x·y=0 or x·y=[x,y]. Here we deal now with the case that n is simple. Again it follows that g and n are isomorphic. The proof via RB-operators uses results of Koszul [Citation27] and Onishchik [Citation29]. We also show a result concerning semisimple decompositions of Lie algebras. Suppose that g=s1+s2 is the vector space sum of two semisimple subalgebras of g. Then g is semisimple. As a corollary we show that the existence of a PA-structure on (g,n) for g semisimple and n complete implies that n is semisimple.

In Section 4, we determine all Lie algebras g which can arise by PA-structures on (g,n) with n=sl2(C)sl2(C). This turns out to be much more complicated than the case n=sl2(C), which we have done in [Citation11]. By Theorem 3.3 of [Citation12], g cannot be solvable unimodular. On the other hand, the result we obtain shows that there are more restrictions than that.

2. Preliminaries

Let A be a nonassociative algebra over a field K in the sense of Schafer [Citation31], with K-bilinear product A×AA,(a,b)ab. We will assume that K is an arbitrary field of characteristic zero, if not said otherwise.

Definition 2.1.

Let λK. A linear operator R:AA satisfying the identity (1) R(x)R(y)=R(R(x)y+xR(y)+λxy)(1) for all x,yA is called a Rota–Baxter operator on A of weight λ, or just RB-operator.

Two obvious examples are given by R = 0 and R=λid, for an arbitrary nonassociative algebra. These are called the trivial RB-operators. The following elementary lemma was shown in [Citation22], Proposition 1.1.12.

Lemma 2.2.

Let R be an RB-operator on A of weight λ. Then Rλid is an RB-operator on A of weight λ, and λ1R is an RB-operator on A of weight 1 for all λ0.

It is also easy to verify the following results.

Proposition 2.3.

[Citation5] Let R be an RB-operator on A of weight λ and ψAut(A). Then R(ψ)=ψ1Rψ is an RB-operator on A of weight λ.

Proposition 2.4.

[Citation22] Let B be a countable direct sum of an algebra A. Then the operator R defined on B by R((a1,a2,,an,))=(0,a1,a1+a2,a1+a2+a3,) is an RB-operator on B of weight 1.

Proposition 2.5.

Let B=AA and ψAut(A). Then the operator R defined on B by (2) R((a1,a2))=(0,ψ(a1))(2) is an RB-operator on B of weight 1. Furthermore the operator R defined on B by (3) R((a1,a2))=(a1,ψ(a1))(3) is an RB-operator on B of weight 1.

Proof.

Let x=(a1,a2) and y=(b1,b2). Then we have R(R(x)y+xR(y)+λxy)=R((0,ψ(a1)b2+(0,a2ψ(b1))+(a1b1,a2b2))=(0,ψ(a1b1))=(0,ψ(a1)ψ(b1))=R(x)R(y).

The second claim follows similarly.□

Proposition 2.6.

[Citation25] Let A=A1A2, R1 be an RB-operator of weight λ on A1, R2 be an RB-operator of weight λ on A2. Then the operator R:AA defined by R((a1,a2))=(R1(a1),R2(a2)) is an RB-operator of weight λ on A.

Proposition 2.7.

[Citation22] Let A=A1+̇A2 be the direct vector space sum of two subalgebras. Then the operator R defined on A by (4) R(a1+a2)=λa2(4) for a1A1 and a2A2 is an RB-operator on A of weight λ.

We call such an operator split, with subalgebras A1 and A2. Note that the set of all split RB-operators on A is in bijective correspondence with all decompositions A=A1+̇A2 as a direct sum of subalgebras.

Lemma 2.8.

[Citation5] Let R be an RB-operator of nonzero weight λ on an algebra A. Then R is split if and only if R(R+λid)=0.

Lemma 2.9.

Let A=A+̇A0+̇A+ be a direct vector space sum of subalgebras of A. Suppose that R is an RB-operator of weight λ on A0, A is an (R+id)(A0)-module and A+ is an R(A0)-module. Define an operator P on A by (5) P|A=0,P|A0=R,P|A+=λid.(5)

Then P is an RB-operator on A of weight λ.

Definition 2.10.

Let P be an RB-operator on A defined as above such that not both A and A+ are zero. Then P is called triangular-split.

We also recall the definition of post-Lie algebra structures on a pair of Lie algebras (g,n) over K, see [Citation11].

Definition 2.11.

Let g=(V,[ ,]) and n=(V,{ ,}) be two Lie brackets on a vector space V over K. A post-Lie algebra structure, or PA-structure on the pair (g,n), is a K-bilinear product x·y satisfying the identities: (6) x·yy·x=[x,y]{x,y},(6) (7) [x,y]·z=x·(y·z)y·(x·z),(7) (8) x·{y,z}={x·y,z}+{y,x·z}(8) for all x,y,zV.

Define by L(x)(y)=x·y the left multiplication operator of the algebra A=(V,·). By (8), all L(x) are derivations of the Lie algebra (V,{,}). Moreover, by (7), the left multiplication L:gDer(n)End(V),xL(x) is a linear representation of g.

If n is abelian, then a post-Lie algebra structure on (g,n) corresponds to a pre-Lie algebra structure on g. In other words, if {x,y}=0 for all x,yV, then the conditions reduce to x·yy·x=[x,y],[x,y]·z=x·(y·z)y·(x·z), i.e., x·y is a pre-Lie algebra structure on the Lie algebra g, see [Citation11].

Definition 2.12.

Let x·y be a PA-structure on (g,n). If there exists a φEnd(V) such that x·y={φ(x),y} for all x,yV, then x·y is called an inner PA-structure on (g,n).

The following result is proved in [Citation2], Corollary 5.6.

Proposition 2.13.

Let (n,{,},R) be a Lie algebra together with a Rota–Baxter operator R of weight 1, i.e., a linear operator satisfying {R(x),R(y)}=R({R(x),y}+{x,R(y)}+{x,y}) for all x,yV. Then x·y={R(x),y} defines an inner PA-structure on (g,n), where the Lie bracket of g is given by (9) [x,y]={R(x),y}{R(y),x}+{x,y}.(9)

Note that ker(R) is a subalgebra of n. For x,yker(R) we have R({x,y})=0. Recall that a Lie algebra is called complete, if it has trivial center and only inner derivations.

Proposition 2.14.

Let n be a Lie algebra with trivial center. Then any inner PA-structure on (g,n) arises by a Rota–Baxter operator of weight 1. Furthermore, if n is complete, then every PA-structure on (g,n) is inner.

Proof.

The first claim follows from Proposition 2.10 in [Citation11]. By Lemma 2.9 in [Citation11] every PA-structure on (g,n) with complete Lie algebra n is inner. The result can also be derived from the proof of Theorem 5.10 in [Citation2].□

Corollary 2.15.

Let n be a complete Lie algebra. Then there is bijection between PA-structures on (g,n) and RB-operators of weight 1 on n.

As we have seen, any inner PA-structure on (g,n) with Z(n)=0 arises by a Rota–Baxter operator of weight 1. For Lie algebra n with nontrivial center this need not be true.

Example 2.16.

Let (e1,e2,e3) be a basis of V and n=r2(K)K with {e1,e2}=e2. Then φ=(100010αβγ) defines an inner PA-structure on (g,n) by x·y={φ(x),y} with g=n, i.e., with [e1,e2]=e2. But φ is not always a Rota–Baxter operator of weight 1 for n. It is easy to see that this is the case if and only if β=0.

Proposition 2.17.

Let x·y be an inner PA-structure arising from an RB-operator R on n of weight 1. Then R is also an RB-operator of weight 1 on g, i.e., it satisfies [R(x),R[y)]=R([R(x),y]+[x,R(y)]+[x,y]) for all x,yV.

Proof.

Because of R([x,y])={R(x),R(y)} and the definition of [x,y] we have R([R(x),y]+[x,R(y)]+[x,y])={R(R(x)),R(y)}+{R(x),R(R(y))}+{R(x),R(y)}=[R(x),R(y)] for all x,yV.□

Corollary 2.18.

Let x·y={R(x),y} be a PA-structure on (g,n) defined by an RB-operator R of weight 1 on n. Denote by gi be the Lie algebra structure on V defined by [x,y]0={x,y},[x,y]i+1=[R(x),y]i[R(y),x]i+[x,y]i, for all i0. Then R defines a PA-structure on each pair (gi+1,gi).

We have [x,y]1=[x,y], and both R and R+id are Lie algebra homomorphisms from gi+1 to gi, see Proposition 7 in [Citation33]. Hence, we obtain a composition of homomorphisms giR+idRgi1R+idRR+idRg0

So the kernels ker(Ri) and ker((R+id)i) are ideals in gj for all 1ij.

For a Lie algebra g, denote by g(i) the derived ideals defined by g(1)=g and g(i+1)=[g(i),g(i)] for i1. An immediate consequence of Proposition 2.13 is the following observation.

Proposition 2.19.

Let x·y={R(x),y} be a PA-structure on (g,n) defined by an RB-operator R of weight 1 on n. Then we have dimg(i)dimn(i) for all i1.

Corollary 2.20.

Let x·y be a PA-structure on (g,n), where n is complete. Then we have dimg(i)dimn(i) for all i1. In particular, if n is solvable, so is g, and if g is perfect, so is n.

Proof.

By Corollary 2.15 this follows from the proposition.□

Proposition 2.21.

Let x·y={R(x),y} be a PA-structure on (g,n) defined by an RB-operator R of weight 1 on n. Then the following holds:

  1. If g and n are not isomorphic, then both R and R+id have a nontrivial kernel.

  2. If either g or n is not solvable, then at least one of the operators R and R+id has a nontrivial kernel.

Proof.

For (1), assume that ker(R)=0. Then R:gn is invertible, hence an isomorphism. This is a contradiction. The same is true for R+id. For (2) assume that ker(R)=ker(R+id)=0. Then R and R+id are isomorphisms from g to n, and gn. Then we can apply a result of Jacobson [Citation24] to the automorphism ψ:=(R+id)°R1 of n, because n is not solvable. We obtain a nonzero fixed point xn, so that 0=ψ(x)x=(R+id)R1(x)x=R1(x).

Since R is bijective, x = 0, a contradiction.□

Corollary 2.22.

Let n be a simple Lie algebra and R be an invertible RB-operator of nonzero weight λ on n. Then we have R=λid.

Proof.

By rescaling we may assume that R has weight 1. We obtain a PA-structure on (g,n) by Proposition 2.13, with Lie bracket (9) on g. Since n is not solvable, either R or R+id have a nontrivial kernel. But ker(R)=0 by assumption, so that ker(R+id) is a nontrivial ideal of n. Hence we have R+id=0.□

3. PA-structures on pairs of semisimple Lie algebras

We will assume that all algebras in this section are finite-dimensional. Let x·y be a PA-structure on (g,n) over C, where g is simple and n is semisimple. Then n is also simple, and both g and n are isomorphic, see Proposition 4.9 in [Citation11]. We have a similar result for n simple and g semisimple. However, its proof is more difficult than the first one.

Theorem 3.1.

Let x·y be a PA-structure on (g,n) over C, where n is simple and g is semisimple. Then g is also simple, and both g and n are isomorphic.

Proof.

By Corollary 2.15 we have x·y={R(x),y} for an RB-operator R of weight 1 on n. Assume that g and n are not isomorphic. By Proposition 2.21 (2) both ker(R) and ker(R+id) are proper nonzero ideals of g, with ker(R)ker(R+id)=0. So we have g=ker(R)ker(R+id)s with a semisimple ideal s. We have n=im(R)+im(R+id) because of x=R(x)+(R+id)(x) for all xn, and im(R)g/ker(R)ker(R+id)s,im(R+id)g/ker(R+id)ker(R)s.

This yields a semisimple decomposition n=(ker(R+id)s)+(ker(R)s).

Suppose that s is nonzero. Then both summands are not simple. This is a contradiction to Theorem 4.2 in Onishchik’s paper [Citation29], which says that at least one summand in a semisimple decomposition of a simple Lie algebra must be simple. Hence we obtain s=0,im(R)=ker(R+id),im(R+id)=im(R) and n=im(R)+̇im(R+id).

Then the main result of Koszul’s note [Citation27] implies that n=im(R)im(R+id), which is a contradiction to the simplicity of n. Hence g and n are isomorphic.□

If g is semisimple with only two simple summands, we can prove the same result for any field K of characteristic zero.

Proposition 3.2.

Let x·y be a PA-structure on (g,n), where n is semisimple, and g=s1s2 is the direct sum of two simple ideals of g. Then g and n are isomorphic.

The proof is the same as before. The only argument where we needed the complex numbers was the result of [Citation29], which we do not need here.

Let n=s1s2 be a direct sum of two simple isomorphic ideals s1 and s2. We would like to find all RB-operators of weight 1 on n such that g with bracket (9) is isomorphic to n.

Proposition 3.3.

All PA-structures on (g,n) with gn=s1s2, where s1 and s2 simple isomorphic ideals of n, arise by the trivial RB-operators or by one of the following RB-operators R on n, and ψAut(n), R((s1,s2))=(s1,ψ(s1)),R((s1,s2))=(0,ψ(s1)),R((s1,s2))=(s1,0)), up to permuting the factors and application of φ(R)=Rid to these operators.

Proof.

By Proposition 2.5 and Proposition 2.7 the given operators are RB-operators of weight 1 on n, because R is. By Proposition 2.21 at least one of ker(R) and ker(R+id) is nonzero. Suppose first that both ker(R) and ker(R+id) are zero. Then we have g=ker(R)ker(R+id) and n=ker(R)+̇ker(R+id). It is easy to see that ker(R) coincides with s1 or s2 by using the Theorem of Koszul [Citation27]. Applying φ if necessary, we can assume that ker(R)=s2. Then again by Koszul’s result we have R((s1,s2))=(ψ1(s1),ψ2(s1)) or R((s1,s2))=(ψ1(s1),0)) for some ψ1,ψ2Aut(n). Since im(R)=ker(R+id) we either have R((s1,s2))=(s1,ψ(s1)) or R((s1,s2))=(s1,0).

In the second case, one of the kernels is zero. Applying φ if necessary, we may assume that ker(R+id)=0 and ker(R)=s1. Then g/ker(R) is a simple Lie algebra, and Rid is an invertible RB-operator of weight 1 on g/ker(R). By Corollary 2.22 we obtain Rid=id, hence R = 0 on g/ker(R). This implies R2=0 on g. The projections of im(R) to s1 and s2 are either zero or an isomorphism on one factor. So we have R((s,0))=(0,ψ(s)) or R((s,0))=(ψ1(s),ψ2(s)) for some automorphisms ψ,ψ1,ψ2. But the second operator does not satisfy R2=0, and hence is impossible. Therefore we are done.□

Proposition 3.4.

Let x·y={R(x),y} be a PA-structure on (g,n) defined by an RB-operator R of weight 1 on n. Let n1=ker(Rn),n2=ker(R+id)n,n3=im(Rn)im((R+id)n) for n=dim(V). Then n=n1+̇n2+̇n3 with {n1,n3}n1,{n2,n3}n2, and n3 is solvable.

Proof.

We first show by induction that ker(Ri) is a subalgebra of n, and that {ker(Ri),im((R+id)i)}ker(Ri) for all i1. The case i = 1 goes as follows. We already know that ker(R) is a subalgebra of n. So we have to show that {ker(R),im(R+id)}ker(R). Let xker(R) and yn. Then by (6) we have {x,(R+id)(y)}={x,R(y)}+{x,y}=[x,y]+{y,R(x)}=[x,y], which is in ker(R), since this is an ideal in g. For the induction step ii+1 consider the iteration of the Lie bracket (9) for all i0, given by [x,y]i=[x,y]i+1[R(x),y]i[x,R(y)]i for all i0. Then {x,y}=[x,y]1[R(x),y]0[x,R(y)]0=[x,y]2[R2(x),y]02[R(x),y]02[R(x),R(y)]02[x,R(y)]0[x,R2(y)]0 and so on. Define a degree of a term [Rl(x),Rk(y)]m by l + k + m, and let x,yker(Ri+1). We can iterate the brackets, until the degree of every summand on the right-hand side will be greater than 3i, so that all summands either have a term Rl(x) with l > i, or a term Rk(y) with k > i, or all summands lie in [ker(Ri+1),ker(Ri+1)]i+1. By induction hypothesis, such terms will vanish for l > i or k > i, and since ker(Ri+1) is an ideal in gi+1, we have {x,y}ker(Ri+1), so that ker(Ri+1) is a subalgebra of n. The induction step for the second claim follows similarly.

Since the image of a subalgebra under the action of an RB-operator is a subalgebra, n1,n2 and their intersection n3 are subalgebras of n. We want to show that n=n1+̇n2+̇n3. Because of ker(Rn)im(Rn)=0 we have n=ker(Rn)+̇im(Rn). In the same way we have n=ker((R+id)n)+̇im((R+id)n). We obtain im(Rn)ker((R+id)n)+̇im(Rn)im((R+id)n)im(Rn).

We claim that ker((R+id)n)im(Rn), so that we have equality above. Indeed, for xker((R+id)n) we have by the binomial formula x+(nn1)R(x)++(n1)Rn1(x)=Rn(x)im(Rn).

Applying Rn1 we obtain Rn1(x)im(Rn) and x+nR(x)++(n2)Rn2(x)im(Rn).

Iterating this we obtain xim(Rn). This yields n=ker(Rn)+̇im(Rn)=ker(Rn)+̇ker((R+id)n)+̇im(Rn)im((R+id)n)=n1+̇n2+̇n3.

On n3 both operators R and R+id are invertible. By Proposition 2.21 part (2) it follows that n3 is solvable.□

Corollary 3.5.

The decomposition n=n1+̇n2+̇n3 induces a decomposition gi=n1+̇n2+̇n3 for each i1 with the same properties as in the Proposition. The Lie algebras (nj,[,]i) and (nj,[,]0) are isomorphic for j = 1, 2, 3.

Proof.

Since R and R+id are RB-operators on all gi, we obtain the same decomposition with the same subalgebras. Note that R+id is invertible on n1, R is invertible on n2 and both are invertible on n3. In order to show that (n1,[,]i is isomorphic to (n1,[,]0, we consider a chain of isomorphisms (n1,[,]n)R+id(n1,[,]n1)R+idR+id(n1,[,]0).

In a similar way we can deal with n2 and n3.□

Proposition 3.6.

Let g=s1+s2 be the vector space sum of two complex semisimple subalgebras of g. Then g is semisimple.

Proof.

Suppose that the claim is not true and let g be a counterexample of minimal dimension. Then g contains a nonzero abelian ideal a. Then we obtain g/a=s1/(s1a)+s2/(s2a).

Since s1a is an abelian ideal s1, it must be zero, i.e., s1a=0. In the same way we have s2a=0. Hence we obtain a semisimple decomposition of g/a with dim(g/a)<dim(g). If g/a is semisimple, this is a contradiction to the minimality of the counterexample g. Otherwise we may assume that g has one-dimensional solvable radical. Then g is reductive, and by Theorem 3.2 of [Citation29], there are no semisimple decompositions of a complex reductive non-semisimple Lie algebra. Hence we are done.□

Proposition 3.7.

Let x·y={R(x),y} be a PA-structure on (g,n) over C, where n is simple, defined by an RB-operator R of weight 1 on n, with associated Lie algebras gi for i=1,,n=dim(V). Assume that g0=n and gn are semisimple. Then all gi are isomorphic to n.

Proof.

Since n1 and n2 are kernels of homomorphisms, they are ideals in gn. The quotient gn/(n1+n2)n3 is semisimple and solvable by Proposition 3.4. Hence n3=0, and we obtain gn=ker(Rn)ker((R+id)n). Because of Corollary 3.5 we have the decomposition gi=ker(Rn)+̇ker((R+id)n) for all i < n, where all Lie algebras (ker(Rn),[,]i) are isomorphic, and all Lie algebras (ker((R+id)n),[,]i) are isomorphic. By Proposition 3.6 all gi are semisimple. By Koszul’s result [Citation27], all gi are isomorphic.□

Proposition 3.8.

Suppose that there is a post-Lie algebra structure on (g,n) over C, where g is semisimple and n is complete. Then n must be semisimple.

Proof.

By Corollary 2.15 the PA-structure is given by x·y={R(x),y}, where R is an RB-operator of weight 1 on n. If at least one of ker(R) and ker(R+id) is trivial, we obtain gn by Proposition 2.21, part (1). Otherwise n=im(R)+im(R+id) is the sum of two nonzero semisimple subalgebras. By Proposition 3.6 n is semisimple.□

4. PA-structures on (g,n) with n=sl2(C)×sl2(C)

In [Citation11], Proposition 4.7 we have shown that PA-structures with n=sl2(C) exist on (g,n) if and only if g is isomorphic to sl2(C), or to one of the solvable non-unimodular Lie algebras r3,λ(C) for λC{1}. In this section we want to show an analogous result for n=sl2(C)×sl2(C). Here we will use RB-operators on n and an explicit classification by Douglas and Repka [Citation17] of all subalgebras of n. This classification is up to inner automorphisms, but we will only need the subalgebras up to isomorphisms. Let us fix a basis (X1,Y1,H1,X2,Y2,H2) of n consisting of the following 4 × 4 matrices: X1=E12, Y1=E21, H1=E11E22, X2=E34, Y2=E43, H2=E33E44.

We use in .

Table 1. Complex 3-dimensional Lie algebras.

Table 2. Solvable subalgebras.

Among the family r3,λ(C),λ0 there are still isomorphisms. In fact, r3,λ(C)r3,μ(C) if and only if μ=λ1 or μ=λ. The list of subalgebras h of n is given as follows. We first list the solvable subalgebras, then the semisimple ones and the subalgebras with a nontrivial Levi decomposition.

Theorem 4.1.

Suppose that there exists a post-Lie algebra structure on (g,n), where n=sl2(C)sl2(C). Then g is isomorphic to one of the following Lie algebras, and all these possibilities do occur:

  1. sl2(C)sl2(C).

  2. sl2(C)r3,λ(C), λ1.

  3. r3,λ(C)r3,μ(C), (λ,μ)(1,1).

  4. r2(C)r2(C)r2(C).

  5. r2(C)(C3C)=x1,,x6 and Lie brackets, for α0,β0,1 [x1,x2]=x1, [x3,x6]=x3, [x4,x6]=αx4, [x5,x6]=βx5.

  6. C((r3,λ(C)C)C)=x1,,x6 and Lie brackets, for λ0,α0,1, [x2,x4]=x2, [x3,x4]=λx3, [x3,x6]=x3, [x5,x6]=αx5.

  7. (r3,λ(C)C2)C=x1,,x6 and Lie brackets, for λ0,α1,α20, and (λ,α1,α2)(1,α1,α11), [x1,x3]=x1, [x2,x3]=λx2, [x2,x6]=α1x2, [x4,x6]=x4, [x5,x6]=α2x5.

  8. (C2C2)C2=x1,,x6 and Lie brackets [x1,x5]=x1,[x2,x5]=α2x2, [x3,x5]=α4x3, [x4,x5]=α6x4,[x1,x6]=α1x1, [x2,x6]=α3x2, [x3,x6]=α5x3, [x4,x6]=α7x4,

with one of the following conditions:

  1. (a)α3=1, α5=α1α7, α6=α2α4,α1α21, α4,α70,1,

  2. (b)α4=α11, α5=α1, α6=α2(α11), α7=α1α3α12α2α3,α3α1α20, α10,1.

Proof.

By Corollary 2.15 it is enough to consider the RB-operators R of weight 1 on n. Then ker(R) and ker(R+id) are ideals in g. If R is trivial, or one of the kernels is trivial, then we have gn, which is type (1). So we assume that R is nontrivial, both ker(R) and ker(R+id) are nonzero, and dim(ker(R))dim(ker(R+id)). Then, for ng, either g has a nontrivial Levi decomposition, or g is solvable.

Case 1: Assume that g has a nontrivial Levi decomposition, i.e., that gsl2(C)r. We claim that sl2(C) is a direct summand of g, i.e., gsl2(C)r, and that r is not isomorphic to r3(C). Then we can argue as follows. Because of Remark 2.12 of [Citation12], g cannot be unimodular, except for gn. Thus r cannot be unimodular, so that g is isomorphic to sl2(C)r3,λ(C) with λ1. On the other hand, all such algebras do arise by Proposition 2.6 and Proposition 4.7 of [Citation11].

Case 1a: Suppose that sl2(C) is not contained in ker(R),ker(R+id) as a subalgebra. Then dim(ker(R+id))=1 and dim(ker(R)){1,2}. Let us assume, both have dimension 1. The other case goes similarly. Then we have r=x1,x2,x3,ker(R)=x1 and ker(R+id)=x2. Furthermore im(R)sl2(C)x2,x3 and im(R+id)sl2(C)x1,x3 are five-dimensional subalgebras of n. By , sl2(C) is a direct summand of them. This implies that sl2(C) is also a direct summand in g. Since both ker(R) and ker(R+id) are ideals in r, we can exclude that r is isomorphic to r3(C), and we are done.

Table 3. Semisimple subalgebras and Levi decomposable subalgebras.

Case 1b: sl2(C) is contained in one of ker(R),ker(R+id). Without loss of generality we may assume that sl2(C)ker(R). If ker(R)=sl2(C), then sl2(C) is an ideal of g, and we have gsl2(C)r, where rim(R)n is not isomorphic to r3(C) by , and we are done. Thus we may assume that dim(ker(R))4. If R splits with subalgebras ker(R) and ker(R+id), then gker(R)ker(R+id), and dim(ker(R))+dim(ker(R+id))=6. By , sl2(C) is a direct summand of ker(R), and hence of g. So we have again gsl2(C)r, and r is not isomorphic to r3(C). If R is not split, it remains to consider the case dim(ker(R))=4 and dim(ker(R+id))=1. We have r=x,y,z with ker(R)=sl2(C)x,ker(R+id)=y and [y,sl2(C)]=0. Assume that [z,sl2(C)]0. Then sl2(C) is not a direct summand of the five-dimensional subalgebra im(R+id) of n, which is a contradiction to . Thus we have gsl2(C)r. Since r has two disjoint one-dimensional ideals x and y, it is not isomorphic to r3(C).

Case 2: Assume that g is solvable. Then im(R) and im(R+id) are solvable subalgebras of n of dimension at most 4 by . So we have dim(ker(R))dim(ker(R+id))2. Thus we have the following four cases: (2a)dim(ker(R))=4,dim(ker(R+id))=2,(2b)dim(ker(R))=3,dim(ker(R+id))=3,(2c)dim(ker(R))=3,dim(ker(R+id))=2,(2d)dim(ker(R))=2,dim(ker(R+id))=2.

For the cases (2a) and (2b), R is split since the dimensions add up to 6. Then g is a direct sum of two solvable subalgebras, which are both isomorphic to subalgebras of n. So we have n=ker(R)+̇ker(R+id) and g=ker(R)ker(R+id).

Case 2a: Since we have only r2(C)r2(C) as four-dimensional solvable subalgebra of n, we have gr2(C)r2(C)C2, which is of type (3) for (λ,μ)=(0,0), or gr2(C)r2(C)r2(C), which is of type (4). Both cases can arise. For the first one we will show this in case (2b). For the second, it follows from Proposition 2.7 with n=X1,H1,X2,H2+̇Y1,Y2+H1.

Case 2b: We have gr3,λ(C)r3,μ(C). The case (λ,μ)=(1,1) cannot arise by Theorem 3.3 of [Citation11]. The cases (λ,μ)=(1,μ) for μ1 arise by Proposition 2.7 with n=X1,X2,H1H2+̇Y1,Y2,H1+μH2.

The other cases with λ,μ1 arise by Proposition 2.6 and Proposition 4.7 of [Citation11].

Case 2c: Here g is isomorphic to (r3,λ(C)r2(C))C or (r3,λ(C)C2)C. In the first case, r2(C)Cim(R) is a solvable subalgebra of n, hence isomorphic to r3,ν(C) by . So C acts trivially on r2(C), and im(R+id)r3,λ(C)Cr2(C)r2(C). Then gr2(C)r2(C)r2(C), which we have already considered in Case (2a). For (r3,λ(C)C2)C we need to distinguish λ=0 and λ0.

Case 2c, λ=0: By Proposition 2.3 we may assume that im(R+id)=X1,H1,X2,H2. Since ker(R) is an ideal of im(R+id) isomorphic to r2(C)C, we have ker(R)=X1,H1,X2. Let us consider the characteristic polynomial χR of the linear operator R acting on n. By assumption on the kernels, χR(t)=t3(t+1)2(tρ).

Case 2c, λ=0,ρ0,1: Then R(x6)=ρx6 for x6=H2+αH1+βX1+γX2. Since ker(R+id) is an abelian two-dimensional subalgebra of n, we have ker(R+id)=Y1+ν1X1+ν2H1,Y2+ν3X2+ν4H2.

We want to compute [x,y] for x = x6 and yker(R+id). By Proposition 2.13 we have, using R(x6)=ρx6 [x,y]={R(x),y}{R(y),x}+{x,y}={R(x),y}=ρ{x,y}.

For x6=H2+αH1+βX1+γX2 and yker(R+id) this yields, using the Lie brackets of n in the standard basis {X1,Y1,H1,X2,Y2,H2}, (10) [x6,Y1+ν1X1+ν2H1]=ρ((2αν12βν2)X12αY1+βH1),(10) (11) [x6,Y2+ν3X2+ν4H2]=ρ((2ν32γν4)X22Y2+γH2).(11)

Since ker(R+id) is an ideal in g and ρ0, both vectors lie again in ker(R+id). Comparing coefficients for the basis vectors we obtain β=2αν2, α(ν1+ν22)=0, γ=2ν4, ν3=ν42.

Suppose that α=0. Then x6=H22ν4X2 and X1,H1r2(C) is a direct summand of g. Therefore gr2(C)Cr3,μ(C) with C=Y1+ν1X1+ν2H1, r3,μ(C)=X2,H22ν4X2,Y2+ν4H2ν42X2,μ=(ρ+1)/ρ, which we have already considered above. Hence we may assume that α0 and ν1=ν22. Consider a new basis for g (note that we redefine x6) given by (x1,,x6)=(X1,12H1+ν2X1,X2,Y1+ν2H1ν22X1,Y2+ν4H2ν42X2,12ρ(H2+αH12αν2X12ν4X2)), with Lie brackets [x1,x2]=x1, [x1,x6]=ρ+1ραx1, [x3,x6]=ρ+1ρx3, [x4,x6]=αx4, [x5,x6]=x6.

This algebra is of type (5), if we replace x6 by x6+α(ρ+1)ρx2. It arises for the triangular-split RB-operator R with A=ker(R)=x1,x2,x3,ker(R+id)=x4,x5 and A0=x6, where x6=H22ν4X2, with the action R(x6)=ρx6.

Case 2c, λ=0,ρ=1: We may assume that there exists x6=Y2+v such that (R+id)(x6)=μ(H2+αH1+βX1+γX2) for some nonzero μ and some α,β,γC. Since ker(R+id) is an abelian subalgebra we obtain α=β=0 and ker(R+id)=H2+γX2,Y1+ν1X1+ν2H1. Then we may choose x6=Y2+κX2+ν3H1+ν4X1. Then [x6,H2+γX2]={R(x6),H2+γX2}={(R+id)(x6)x6,H2+γX2}={Y2+κX2,H2+γX2}=2Y2+2κX2+γH2.

This is not contained in ker(R+id), which is a contradiction to the fact that ker(R+id) is an ideal.

Case 2c, λ=0, ρ = 0: Then we have R(H2)=αH1+βX1+γX20 and ker(R+id)=Y1+ν1X1+ν2H1,Y2+ν3X2+ν4H2. Since [H2,Y2+ν1X1+ν2H1]={γX2,Y2+ν1X2+ν2H2} is in ker(R+id), we obtain γ=0. Since [H2,Y1+ν1X1+ν2H2]={αH1+βX1,Y1+ν1X1+ν2H1} is in ker(R+id), we obtain α(ν1+ν22)=0 and β=2αν2. Since R(H2)0 we have α0,ν1=ν22 and R(H2)=αH12αν2X1. Consider a new basis for g given by (x1,,x6)=(X1,12H1+ν2X1,X2,Y1+ν2H1ν22X1,Y2+ν3X2+ν4H2,12H2), with Lie brackets [x1,x2]=x1, [x1,x6]=αx1, [x3,x6]=x3, [x4,x6]=αx4.

This algebra is of type (3), if we replace x6 by x6αx2.

Case 2c, λ0: Then we have ker(R)=X1,X2,12(H1+λH2). We again have χR(t)=t3(t+1)2(tρ), where we distinguish the cases ρ0,1,ρ=1 and ρ = 0.

Case 2c, λ0,ρ0,1: Then we may assume that R(x6)=ρx6 for x6=H2+αH1+βX1+γX2. As ker(R+id) is abelian, we have ker(R+id)=Y1+ν1X1+ν2H1,Y2+ν3X2+ν4H2. Since V=ker(R)ker(R+id)x6, the two elements H1+λH2 and H2+αH1 need to be linearly independent, i.e., 1αλ0. By (10) and (11) we obtain γ=2ν4,β=2αν2,ν3=ν42, and α(ν1+ν22). Suppose that α=0. Then x6=H22ν4X2. Consider a new basis for g given by (x1,,x6)=(Y1+ν1X1+ν2H1,X1,X2,12(H1+λH2),Y2ν42X2+ν4H2,12(ρ+1)H2), with Lie brackets [x2,x4]=x2, [x3,x4]=λx3, [x3,x6]=x3, [x4,x6]=λν4x3, [x5,x6]=ρ1+ρx5.

This is an algebra of type (6), if we replace x4 by x4+λν4x3.

Now we assume that α0. Consider a new basis for g given by (x1,,x6)=(X1,X2,12(H1+λH2),Y2ν42X2+ν4H2,Y1ν22X1+ν2H1,12(ρ+1)(H22ν4X2+α(H12ν2X1))), with Lie brackets [x1,x3]=x1, [x2,x3]=λx2, [x1,x6]=αx1, [x2,x6]=x2,[x3,x6]=αν2x1λν4x2, [x4,x6]=δx4, [x5,x6]=αδx5, where δ=ρρ+1. Replacing x6 by 1δ(x6αν2x1ν4x2αx3) we obtain the Lie brackets [x1,x3]=x1, [x2,x3]=λx2, [x2,x6]=αx2, [x4,x6]=x4, [x5,x6]=αx5, where α=1αλδ=(ρ+1)(αλ1)ρ.

Note that α0 and ααλ1 by assumption. In other words, αα+1λ. Consider a new basis for g given by (x1,,x6)=(x2,x1,1λx3,x4,x5,x6αλx3), with Lie brackets [x1,x3]=x1, [x2,x3]=λx2, [x2,x6]=αλx2, [x4,x6]=x4, [x5,x6]=αx5, where λ=1λ. This is of type (7). Since r3,λ(C)r3,λ(C), one may check that we do not only have αα+1λ, but also αλα. For α+1λλα we obtain no restriction for α. However, for α+1λ=λα we obtain λ=1 or λ=α+1, which excludes both (λ,α,α)=(1,α,α1) and (λ,α,α)=(λ,λ1,1). Rewriting this in the parameters of the Lie brackets from type (7), we obtain all cases except for (λ,α,α)=(λ,λ1,1) with λ1. These PA-structures arise by a triangular-split RB-operator with A=ker(R),A+=ker(R+id) and A0=x6 with the action R(x6)=ρx6,ρ0,1.

Case 2c, λ0,ρ=1: This leads to a contradiction in the same way as case 2c with λ=0,ρ=1.

Case 2c, λ0,ρ = 0: We have R(H2)=αX1+βX2+γ(H1+λH2) and ker(R+id)=x4,x5 with x4=Y1+ν1X1+ν2H1,x5=Y2+ν3X2+ν4H2. Similarly to (10),(11) we obtain R(H2)=γ(H12ν2X1)+γλ(H22ν4X2). This implies that γ0 and x4=Y1ν22X1+ν2H1,x5=Y2ν42X2+ν4H2. By setting x1 = X1, x2 = X2, x3=12(H1+λH2) and x6=12γH2 we obtain a new basis for g with Lie brackets [x1,x3]=x1, [x2,x3]=λx2, [x1,x6]=x1, [x2,x6]=δx2,[x3,x6]=ν2x1+λ2ν4x2, [x4,x6]=x4, [x5,x6]=λx5, where δ=1+λγγ with δλ. Replacing x6 by x6+ν2x1+λν4x2+x3 we obtain the brackets [x1,x3]=x1, [x2,x3]=λx2, [x2,x6]=α1x2, [x4,x6]=x4, [x5,x6]=λx5 with α1=δ+λ=1γ. This is of type (7) with α2=λ. It arises by the triangular-split RB-operator with A=x1,x2,A+=x4,x5 and A0=u,v, with u=1γ(H22ν4X2) and v=H12ν2X1+λ(H22ν4X2), and the action R(u) = v, R(v) = 0.

Case 2d: Suppose that one of the kernels ker(R) and ker(R+id) is nonabelian. Without loss of generality, let us assume that ker(R)r2(C). Write g(ker(R)ker(R+id))a,b. Then ker(R)a is a three-dimensional solvable subalgebra of im(R+id). By , we see that it is isomorphic to r2(C)C. In this case there exist nonzero aker(R)a and bker(R)b such that [a,ker(R)]=[b,ker(R)]=0. Then gker(R)(ker(R+id)a,b) with ker(R)r2(C), and ker(R+id)a,br2(C)r2(C) by . Hence we obtain gr2(C)r2(C)r2(C), which is of type (4).

So we may assume that ker(R)ker(R+id)C2. Then the characteristic polynomial of R has the form χR(t)=t2(t+1)2(tρ1)(tρ2).

Case 2d, ρ1,ρ20,1: Suppose first that either ρ1ρ2 or that ρ1=ρ2 and the eigenspace is two-dimensional. Then by Proposition 3.4,n=ker(R)+̇ker(R+id)+̇x5,x6 with linearly independent eigenvectors x5,x6 corresponding to the eigenvalues ρ1 and ρ2. Since ker(R) is an abelian ideal in im(R+id)=X1,H1,X2,H2, we may assume that ker(R)=X1,X2 and [x5,x6]=0. The decomposition n=ker(R+id)+̇im(R+id) shows that ker(R+id) has a basis x3=Y1+αH1+ν3X1, x4=Y2+βH2+ν4X2. Since [x5,x6]=0, we have x5=H1+ν1X1+ξ1(H2+ν2X2), x6=H2+ν2X2+ξ2(H1+ν1X1) with ξ1ξ21. So we have by (10) and (11) x3=Y1ν12H1ν124X1, x4=Y2ν22H2ν224X2. Consider a basis for g given by (x1,,x6)=(X1,X2,x3,x4,12(1+ρ1)x5,12(1+ρ2)x6), with Lie brackets [x1,x5]=x1, [x1,x6]=ξ2x1, [x2,x5]=ξ1x2, [x2,x6]=x2,[x3,x5]=γx3, [x3,x6]=δξ2x3, [x4,x5]=γξ1x4, [x4,x6]=δx4, where γ=ρ1ρ1+1, δ=ρ2ρ2+1 with γ,δ0,1 and ξ1ξ21. This is type (8a). It arises by the triangular-split RB-operator R with A=X1,X2,A+=x3,x4 and A0=x5,x6, where R acts on A0 by R(x5)=ρ1x5 and R(x6)=ρ2x6. Note that for ν2=ξ2=0 and ξ10 we get type (7) without the restriction (λ,α1,α2)(λ,λ1,1) for λ1, which we had in Case 2c, λ0,ρ0,1.

Suppose now that ρ2=ρ10,1, and the eigenspace for ρ1 is one-dimensional. Let R(x5)=ρ1x5 and R(x6)=x5+ρ1x6. In the same way as before we have x5=H1+ν1X1+ξ(H2+ν2X2), x6=κ(H2+ν2X2) with κ0 and x3=Y1ν12H1ν124X1, x4=Y2ν22H2ν224X2. Consider a basis for g given by (x1,,x6)=(X1,X2,x3,x4,12(1+ρ1)x5,12(1+ρ1)x6), with Lie brackets [x1,x5]=x1, [x1,x6]=(γ+1)x1, [x2,x5]=ξx2, [x2,x6]=(κ+ξ+γξ)x2,[x3,x5]=γx3, [x3,x6]=(γ+1)x3, [x4,x5]=γξx4, [x4,x6]=(κγξγξ)x4, where γ=ρ1ρ1+10,1 and κ0. This is type (8b). It arises by the triangular-split RB-operator R with A=X1,X2,A+=x3,x4 and A0=x5,x6, where R acts on A0 by R(x5)=ρ1x5 and R(x6)=x5+ρ1x6.

Case 2d, ρ1=ρ2=0: We have g=ker(R+id)+̇im(R+id) and we can assume that ker(R)=X1,X2 and ker(R+id)=Y1+ν1X1+ν2H1,Y2+ν3X2+ν4H2. Suppose first that R(v)=X1 and R(w)=X2 for some v, w. Then [Y1+ν1X1+ν2H1,v]={Y1+ν1X1+ν2H1,X1}=H1+2ν2X1ker(R+id), which is a contradiction. Otherwise we see from the possible Jordan forms of R that there exist v, w with R(v)=αX1+βX20 and R(w) = v. This leads to a contradiction in the same way.

Case 2d, ρ1=0,ρ20,1: This case is analogous to the second part of the case before.

Case 2d, ρ1=0,ρ2=1: As above we may assume that im(R+id)=X1,X2,H1,H2 and ker(R)=X1,X2, and αH1+βH2+γX1+δX2ker(R+id)im(R+id) for some α,β,γ,δC. Since ker(R+id) is abelian, we may assume that ker(R+id)=H1+ν1X1,Y2+ν2X2+ν3H2 for some ν1,ν2,ν3C. Let vker(R2) such that R(v)=ν4X1+ν5X20. Then [v,Y2+ν2X2+ν3H2]={ν4X1+ν5X2,Y2+ν2X2+ν3H2}=ν5(H22ν3X2)ker(R+id) implies that ν5=0. By [v,H1+ν1X1]={ν4X1,H1+ν1X1}=2ν4X1ker(R+id) we obtain ν4=0, which is a contradiction to R(v)0.□

Remark 4.2.

The algebras from different types are nonisomorphic, except for algebras of type (8), which have intersections with types (3) and (7) for certain parameter choices.

Additional information

Funding

Dietrich Burde is supported by the Austrian Science Foundation FWF, grant P28079 and grant I3248. Vsevolod Gubarev acknowledges support by the Austrian Science Foundation FWF, grant P28079.

References

  • Atkinson, F. V. (1963). Some aspects of Baxters functional equation. J. Math. Anal. Appl. 7(1): 1–30.
  • Bai, C., Guo, L., Ni, X. (2010). Nonabelian generalized lax pairs, the classical Yang-Baxter equation and PostLie algebras. Commun. Math. Phys. 297(2):553–596.
  • Baxter, G. (1960). An analytic problem whose solution follows from a simple algebraic identity. Pacific J. Math. 10(3):731–742.
  • Belavin, A. A., Drinfel'd, V. G. (1983). Solutions of the classical Yang-Baxter equation for simple Lie algebras. Funct. Anal. Appl. 16(3):159–180.
  • Benito, P., Gubarev, V., Pozhidaev, A. (2018). Rota–Baxter operators on quadratic algebras. 23. arXiv:1801.07037
  • Burde, D. (1996). Affine structures on nilmanifolds. Int. J. Math. 07(05):599–616.
  • Burde, D., Dekimpe, K., Deschamps, S. (2005). The Auslander conjecture for NIL-affine crystallographic groups. Math. Ann. 332(1):161–176.
  • Burde, D. (2006). Left-symmetric algebras, or pre-Lie algebras in geometry and physics. Centr. Eur. J. Math. 4(3):323–357.
  • Burde, D., Dekimpe, K., Deschamps, S. (2009). LR-algebras. Contemp. Math. 491:125–140.
  • Burde, D., Dekimpe, K., Vercammen, K. (2010). Complete LR-structures on solvable Lie algebras. J. Group Theory. 13(5):703–719.
  • Burde, D., Dekimpe, K., Vercammen, K. (2012). Affine actions on Lie groups and post-Lie algebra structures. Linear Algebra Appl. 437(5):1250–1263.
  • Burde, D., Dekimpe, K. (2013). Post-Lie algebra structures and generalized derivations of semisimple Lie algebras. Mosc. Math. J. 13(1):1–18.
  • Burde, D., Dekimpe, K. (2016). Post-Lie algebra structures on pairs of Lie algebras. J. Algebra. 464:226–245.
  • Burde, D., Moens, W. A. (2016). Commutative post-Lie algebra structures on Lie algebras. J. Algebra. 467:183–201.
  • Burde, D., Moens, W. A., Dekimpe, K. (2017). Commutative post-Lie algebra structures and linear equations for nilpotent Lie algebras. 1–14. arXiv:1711.01964
  • Cartier, P. (1972). On the structure of free Baxter algebras. Adv. Math. 9(2):253–265.
  • Douglas, A., Repka, J. (2018). Subalgebras of the rank two semisimple Lie algebras. Linear Multilinear Algebra. 66(10):2049–2075.
  • Ebrahimi-Fard, K., Lundervold, A., Mencattini, I., Munthe-Kaas, H. Z. (2015). Post-Lie algebras and isospectral flows. SIGMA Symmetry Integrability Geom. Methods Appl. 11(93):16.
  • Gubarev, V. (2017). Universal enveloping Lie Rota–Baxter algebra of pre-Lie and post-Lie algebras. 1–13. arXiv:1708.06747
  • Gubarev, V., Kolesnikov, P. (2013). Embedding of dendriform algebras into Rota–Baxter algebras. Cent. Eur. J. Math. 11(2):226–245.
  • Guo, L., Keigher, W. (2000). Baxter algebras and shuffle products. Adv. Math. 150(1):117–149.
  • Guo, L. (2012). An Introduction to Rota–Baxter Algebra. Surveys of Modern Mathematics, Vol. 4. Somerville: Intern. Press, 226 pp.
  • Helmstetter, J. (1979). Radical d’une algèbre symétrique a gauche. Ann. Inst. Fourier. 29(4):17–35.
  • Jacobson, N. (1962). A note on automorphisms of Lie algebras. Pacific J. Math. 12(1):303–315.
  • Li, X. X., Hou, D. P., Bai, C. M. (2007). Rota–Baxter operators on pre-Lie algebras. J. Nonlinear Math. Phys. 14(2):269–289.
  • Kim, H. (1986). Complete left-invariant affine structures on nilpotent Lie groups. J. Differential Geom. 24(3):373–394.
  • Koszul, J. L. (1978). Variante d’un théoréme de H. Ozeki. Osaka J. Math. 15:547–551.
  • Loday, J.-L. (2008). Generalized bialgebras and triples of operads. Astrisque No. 320:116.
  • Onishchik, A. L. (1969). Decompositions of reductive Lie groups. Mat. Sbornik. 80(122):515–554.
  • Rota, G.-C. (1969). Baxter algebras and combinatorial identities I. Bull. Am. Math. Soc. 75(2):325–329.
  • Schafer, R. D. (1995). An Introduction to Nonassociative Algebras. New York, NY: Dover Publications, 166 pp.
  • Segal, D. (1992). The structure of complete left-symmetric algebras. Math. Ann. 293(1):569–578.
  • Semenov-Tyan-Shanskii, M. A. (1983). What is a classical R-matrix? Funktsional. Anal. i Prilozhen. 17(4):17–33.
  • Vallette, B. (2007). Homology of generalized partition posets. J. Pure Appl. Algebra. 208(2):699–725.