123
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Post-Lie algebra structures for perfect Lie algebras

ORCID Icon, &
Received 18 Nov 2023, Accepted 26 Feb 2024, Published online: 15 May 2024

Abstract

We study the existence of post-Lie algebra structures on pairs of Lie algebras (g,n), where one of the algebras is perfect non-semisimple, and the other one is abelian, nilpotent non-abelian, solvable non-nilpotent, simple, semisimple non-simple, reductive non-semisimple or complete non-perfect. We prove several nonexistence results, but also provide examples in some cases for the existence of a post-Lie algebra structure. Among other results we show that there is no post-Lie algebra structure on (g,n), where g is perfect non-semisimple, and n is sl3(C). We also show that there is no post-Lie algebra structure on (g,n), where g is perfect and n is reductive with a 1-dimensional center.

2020 MATHEMATICS SUBJECT CLASSIFICATION:

1 Introduction

Post-Lie algebras and Post-Lie algebra structures (or PA-structures) on pairs of Lie algebras have been studied in many areas of mathematics during the last years. PA-structures are a natural generalization of pre-Lie algebra structures on Lie algebras, which arise among other things from affine manifolds and affine actions on Lie groups, crystallographic groups, étale affine representations of Lie algebras, quantum field theory, operad theory, Rota-Baxter operators, and deformation theory of rings and algebras. There is a large literature on pre-Lie and post-Lie algebras, see for example [Citation5–9, Citation12, Citation14, Citation18] and the references therein. For a survey on pre-Lie algebra respectively post-Lie algebra structures see [Citation2, Citation11].

In the present article we study the existence question of post-Lie algebra structures on pairs of Lie algebras (g,n), where one Lie algebra is perfect and the other one is abelian, nilpotent, solvable, simple, semisimple, reductive, complete or perfect. In all but three cases we can solve the existence question and generalize our previous results for semisimple Lie algebras to perfect Lie algebras. However, for these three cases of (g,n), namely where g is perfect non-semisimple, and n is either nilpotent, simple, or semisimple, we are not able solve the existence question in general. We conjecture that there do not exist post-Lie algebra structures in these cases. For some special families of examples we can prove this conjecture.

The outline of this paper is as follows. In the second section we provide basic results on perfect Lie algebras, including a classification of complex perfect Lie algebras of dimension n9. We also recall the basic notions for post-Lie algebra structures. In the third section we study the existence question for pairs (g,n) where g is perfect. If n is nilpotent, then we show for perfect Lie algebras g of dimension 6, that there exist no post-Lie algebra structures on (g,n). We also prove, using the classification of complex perfect Lie algebras in low dimension, that there is no post-Lie algebra structure on (g,n), where n=sl3(C). We find post-Lie algebra structures on (g,n) for examples of reductive Lie algebras n, and show that such structures do not exist for reductive Lie algebras n with a 1-dimensional center. In the fourth section we study the existence question for pairs (g,n) where n is perfect. Here we often find post-Lie algebra structures and are able to solve the existence problem in all cases.

2 Preliminaries

Let g be a finite-dimensional Lie algebra over a field K. Denote by Z(g) the center of g, by rad(g) the solvable radical of g, and by nil(g) the nilradical of g. A Lie algebra g is called perfect, if g=[g,g]. Every semisimple Lie algebra over a field of characteristic zero is perfect. The converse does not hold. It is well known that the solvable radical of a perfect Lie algebra is nilpotent. One can also give a necessary and sufficient condition for a Levi decomposition of g, so that g is perfect.

Proposition 2.1.

Let g=srad(g) be a Levi decomposition of a Lie algebra g and consider V=r/[r,r] with r=rad(g) as an s-module. Then g is perfect if and only if V does not contain the trivial 1-dimensional s-module.

The lowest-dimensional example of a complex perfect non-semisimple Lie algebra is sl2(C)V(2) of dimension 5. Here sl2(C)=span{e1,e2,e3} has Lie brackets given by [e1,e2]=e3, [e1,e3]=2e1, [e2,e3]=2e2, and V(2)=span{e4,e5} is the natural representation of sl2(C). The Lie brackets of sl2(C)V(2) are given by [e1,e2]=e3,   [e2,e3]=2e2,   [e3,e5]=e5.[e1,e3]=2e1,   [e2,e4]=e5,[e1,e5]=e4,   [e3,e4]=e4,his Lie algebra is perfect and has a non-trivial solvable radical. Its center is trivial. We also give an example of a non-semisimple perfect Lie algebra with non-trivial center. For this, let sl2(C)=span{e1,e2,e3} and n3(C)=span{e4,e5,e6} be the Heisenberg Lie algebra, with [e4,e5]=e6. Consider the following semidirect sum of sl2(C) and n3(C), given by the following Lie brackets in the basis (e1,,e6): [e1,e2]=e3,   [e2,e3]=2e2,   [e3,e5]=e5,[e1,e3]=2e1,   [e2,e4]=e5,   [e4,e5]=e6.[e1,e5]=e4,   [e3,e4]=e4,

We denote this Lie algebra by sl2(C)n3(C).

Example 2.2.

The Lie algebra g=sl2(C)n3(C) is perfect, but not semisimple. It has a 1-dimensional center.

Indeed, we have Z(g)=span{e6}, and g is not semisimple. The nilradical n3(C) is isomorphic to V(2)V(1) as sl2(C)-module, where V(n) denotes the irreducible sl2(C)-module of dimension n1. It follows from the Lie brackets that g is perfect. We also can derive this from Proposition 2.1. The quotient V=n3/[n3,n3] is isomorphic to V(2) and does not contain the trivial 1-dimensional module V(1). Hence g is perfect.

In the study of PA-structures for perfect Lie algebras we are also interested in a classification of perfect Lie algebras in low dimensions. Turkowski has classified Lie algebras with non-trivial Levi decomposition up to dimension 8 over the real numbers in [Citation16], where he lists explicit Lie brackets for all algebras. From this work it is not difficult to derive a classification of complex perfect Lie algebras of dimension n8. We need to add one Lie algebra though, which Turkowski has not in his list. It is the complexification of the algebra L8,13ε for ε=0, isomorphic to sl2(C)(V(2)n3(C)). Turkowski only allows ε=±1.

There is another classification given by Alev, Ooms and Van den Bergh in [Citation1], namely the classification of non-solvable algebraic Lie algebras of dimension n8 over an algebraically closed field of characteristic zero. It also contains explicit Lie brackets for all algebras. Since perfect Lie algebras are algebraic, we obtain again a list of complex perfect non-semisimple Lie algebras of dimension n8. This list coincides with the (corrected) one by Turkowski. Let n5(C) denote the 5-dimensional Heisenberg Lie algebra, with basis (e1,,e5) and Lie brackets [e1,e2]=[e3,e4]=e5.

Denote by f2,3(C) the free-nilpotent Lie algebra with 2 generators and nilpotency class 3, with basis (e1,,e5) and Lie brackets [e1,e2]=e3, [e1,e3]=e4, [e2,e3]=e5.

The classification result is as follows.

Proposition 2.3.

Every complex perfect non-semisimple Lie algebra of dimension n8 is isomorphic to one of the following Lie algebras:

Turkowski has also classified in [Citation17] real and complex Lie algebras with non-trivial Levi decomposition in dimension 9. This yields a list of complex perfect Lie algebras of dimension 9. Denote by f3,2(C) the free-nilpotent Lie algebra with 3 generators and nilpotency class 2, and by A6,4(C) the 2-step nilpotent Lie algebra with basis (e1,,e6) and Lie brackets [e1,e4]=e5, [e2,e3]=e5, [e3,e4]=e6.

Then we have the following result.

Proposition 2.4.

Every complex perfect non-semisimple Lie algebra of dimension 9 is isomorphic to one of the following Lie algebras:

In particular the nilradical of such an algebra always has nilpotency class c2.

We recall the definition of a post-Lie algebra structure on a pair of Lie algebras (g,n) over a field K, see [Citation5]:

Definition 2.5.

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: (1) x·yy·x=[x,y]{x,y}(1) (2) [x,y]·z=x·(y·z)y·(x·z)(2) (3) x·{y,z}={x·y,z}+{y,x·z}(3) for all x,y,zV.

Define by L(x)(y)=x·y the left multiplication operators of the algebra A=(V,·). By (3), all L(x) are derivations of the Lie algebra (V,{,}). Moreover, by (2), 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, or left-symmetric structure on g. In other words, if {x,y}=0 for all x,yV, then the conditions reduce to (4) x·yy·x=[x,y](4) (5) [x,y]·z=x·(y·z)y·(x·z),(5) i.e., x·y is a pre-Lie algebra structure on the Lie algebra g.

For semisimple Lie algebras n we have the following result on PA-structures on pairs (g,n), see Proposition 2.14 in [Citation5].

Proposition 2.6.

Let n be a semisimple Lie algebra. Then a pair (g,n) admits a PA-structure if and only if there is an injective Lie algebra homomorphism φ:gnn such that the map (p1p2)|g:gn is bijective. Here pi:nnn denotes the projection onto the i-th factor for i = 1, 2.

Let us denote the composition of pi and φ by ji=φ°pi:gn for i = 1, 2.

We also recall the following results, see Propositions 2.14 and 2.21 in [Citation9].

Proposition 2.7.

Let (g,n) be a pair of Lie algebras, where n is complete. Then there is a bijection between PA-structures on (g,n) and Rota-Baxter operators R of weight 1 on n. Every such PA-structure is then of the form x·y={R(x),y}. If g and n are not isomorphic, then both ker(R) and ker(R+id) are nonzero ideals in g.

Here a Rota-Baxter algebra operator, for a nonassociative algebra over a field K, of weight λK is a linear operator R:AA satisfying R(x)R(y)=R(R(x)y+xR(y)+λxy)for all x,yA.

3 PA-structures with g perfect

In this section we study the existence question of PA-structures on pairs of complex Lie algebras (g,n), where g is perfect non-semisimple. We consider 7 different cases for n, namely (a) n is abelian, (b) n is nilpotent non-abelian, (c) n is solvable non-nilpotent, (d) n is simple, (e) n is semisimple non-simple, (f) n is reductive non-semisimple, and (g) n is complete non-perfect.

We start with case (a).

Proposition 3.1.

There is no PA-structure on a pair (g,n), where g is perfect and n is abelian.

Proof.

A PA-structure on a pair (g,n), where n is abelian, corresponds to a left-symmetric (or pre-Lie algebra) structure on g. By Corollary 21 of [Citation15] there is no such structure on a perfect Lie algebra. □

For case (b) we only have partial results so far. In Proposition 3.6 of [Citation7] we have proved that there is no post-Lie algebra structure on (g,n), where g is perfect of dimension 5, namely g=sl2(C)V(2), and n is nilpotent. We can generalize this result to perfect Lie algebras g of dimension 6. According to Proposition 2.3, the non-semisimple perfect Lie algebras of dimension 6 are given by sl2(C)V(3) and sl2(C)n3(C).

Proposition 3.2.

Let (g,n) be a pair of Lie algebras, where g is either sl2(C)V(3) or sl2(C)n3(C), and n is nilpotent. Then there is no PA-structure on (g,n).

Proof.

Let g=sl2(C)a and assume that there exists a PA-structure on (g,n), with the homomorphism L:gDer(n) given as in Definition 2.5. Let φ:g⏧n⋊Der(n) be the embedding defined by x(x,L(x)). We claim that ker(L)sl2(C)=0. Suppose that this intersection is nonzero. Since it is an ideal in the simple Lie algebra sl2(C), this implies that ker(L)sl2(C)=sl2(C), so that sl2(C)ker(L). In particular, we have L(s) = 0 for all ssln(C). By axiom (1) in Definition 2.5 it follows that [s,t]{s,t}=s·tt·s=0for all s,tsl2(C). Hence n has a subalgebra isomorphic to sl2(C). This is impossible, because n is nilpotent, so that the claim follows.

We have shown in the proof of Theorem 3.3 in [Citation12] that the Lie algebra nh, with h=L(g)Der(n),

has a direct vector space sum decomposition nh=φ(g)h.

Since φ(g) and h are homomorphic images of a perfect Lie algebra, they are perfect. Hence nh is perfect. Let s=L(sl2(C)) and r=L(a). Then we have h=sr. Since ker(L)sl2(C)=0, s is nonzero and semisimple. Hence dim(s)=3 and ssl2(C). So hn=(sr)nis perfect and has nilradical m=rn. By Proposition 2.1, m/[m,m] does not contain the trivial 1-dimensional s-module V(1). Since nm, also n/[n,n]m/[m,m] does not contain the trivial 1-dimensional s-module V(1). Hence sn is a perfect Lie algebra of dimension 9. By Proposition 2.4, the nilpotency class c(n) of n is at most 2. Proposition 4.2 of [Citation5] says, that if (g,n) admits a post-Lie algebra structure, and c(n)2, then g admits a pre-Lie algebra structure. Since g is perfect, this is impossible by Corollary 21 of [Citation15]. □

Let us again state the last result used in the proof, see also Proposition 3.3 in [Citation7].

Proposition 3.3.

Let (g,n) be a pair of Lie algebras, where g is perfect and n is 2-step nilpotent. Then there exists no PA-structure on (g,n).

In case (c), we have proved the following result in Proposition 4.4 of [Citation5].

Proposition 3.4.

There is no PA-structure on a pair (g,n), where g is perfect and n is solvable non-nilpotent.

For case (d) we start with low-dimensional simple Lie algebras n. There is no pair (g,n) with nsl2(C) and g perfect non-semisimple, since the only perfect Lie algebra in dimension 3 is simple. The next case is to consider pairs (g,n), where nsl3(C) and g is a perfect non-semisimple Lie algebra of dimension 8. We start with the following result.

Lemma 3.5.

Let i:sl2(C)V(2)sl3(C) be an injective Lie algebra homomorphism. By conjugating with a matrix in GL3(C) we may assume that the image of i is of the form im(i)={(a1a2a3a4a1a5000)sl3(C)}, or im(i)={(b1b20b3b10b4b50)sl3(C)}.

Proof.

The vector space C3 becomes an sl2(C)-module via the embedding i restricted to sl2(C). Hence either C3V(2)V(1), or C3V(3) as sl2(C)-module. In the first case we can choose a basis of C3 such that i(sl2(C))={(a1a20a3a10000)sl3(C)},by using the natural representation i(e1)=E12, i(e2)=E21 and i(e3)=E11E22 for the basis (e1,e2,e3) of sl2(C). A short computation shows that when such a representation extends to sl2(C)V(2), we obtain one of the forms for im(i) as described above.

In the second case we may assume that i(e1)=E12+2E23, i(e2)=E21+2E32 and i(e3)=2E112E33. It is easy to see that this representation does not extend to one of sl2(C)V(2). □

Lemma 3.6.

Let j:sl2(C)sl3(C) be an injective Lie algebra homomorphism. Denote by ci:gl3(C)C3 the projection of a matrix in gl3(C) to its i-th column, and by ri:gl3(C)C3 the projection to its i-th row. Then none of the linear maps ci°j,ri°j:sl2(C)C3 for i = 1, 2, 3 is bijective.

Proof.

It is enough to show the claim for columns. We obtain the result for rows by applying the isomorphism of Lie algebras sl3(C)sl3(C), given by XXT, to the result for columns. We will give the proof for c3°j. The other two cases are similar. Note that C3 is an sl2(C)-module via j. Hence the map ci°j is actually the map sl2(C)C3, xx·(001).

Because of dimsl2(C)=dimC3 the annihilator of any vector is non-trivial by Lemma 4.1 in [Citation6], so that the map c3·j is not injective. □

Lemma 3.7.

Let s1 and s2 be Lie algebras isomorphic to sl2(C). Then the ideals of the Lie algebra g=s1(s2V(2)) are given by 0,s1,s2V(2),s1V(2),V(2),g.

Proof.

It is clear that all of these subspaces are ideals in g. Conversely, assume that a is an ideal in g. If as10, then s1a and a/s1 is an ideal of g/s1s2V(2). But the only ideals of s2V(2) are 0, V(2) and s2V(2). So, if as10 then all ideals are given by s1, s1V(2) and g.

Now suppose that as1=0. We claim that then as2V(2). Indeed, suppose that there exists an element x in a(s2V(2)). We can write x=x1+x2 with x1s1 and x2s2V(2), where x10. There exists a ys1 such that [y,s1]0, so that 0[y,x]=[y,x1+x2]=[y,x1]s1a, which is a contradiction. Hence we have as2V(2), which leads to the ideals 0, V(2) and s2V(2). □

Lemma 3.8.

There is no direct vector space decomposition sl3(C)=ab with subalgebras a and b of sl3(C) satisfying asl2(C)V(2) and bsl2(C).

Proof.

Assume that there is such a decomposition sl3(C)=ab. Then after applying a base change we may assume by Lemma 3.5 that a={(a1a2a3a4a1a5000)sl3(C)}, or a={(b1b20b3b10b4b50)sl3(C)}

As sl3(C) is a direct vector space sum of a and b we must have that the row projection map r3:bC3 is bijective in the first case, and the column projection map c3:bC3 is bijective in the second case. However, by Lemma 3.6, this is impossible. □

We can now apply these lemmas to PA-structures on pairs (g,n) with n=sl3(C), where g has a Levi subalgebra isomorphic to sl2(C)sl2(C).

Proposition 3.9.

Let g=sl2(C)(sl2(C)V(2)) and n=sl3(C). Then there is no PA-structure on the pair (g,n).

Proof.

Assume that there exists a PA-structure on (g,n) with n=sl3(C). Let us write g=s1(s2V(2)). By Proposition 2.6 there exists an injective Lie algebra homomorphism j:gnn, x(j1(x),j2(x)) such that j1j2 is a bijective linear map. We will examine the possible kernels of j1. Since ker(j1) is an ideal in g, it must be one of the six possibilities given in Lemma 3.7.

Case 1: ker(j1)=0. Then j1:gn is an isomorphism. Since g is not semisimple, this is a contradiction.

Case 2: ker(j1)=g. Then j1 is the zero map. Since j1j2 has to be bijective, j2:gn is an isomorphism. This is impossible.

Case 3: ker(j1)=V(2). Then the representation j1|s1s2:s1s2sl3(C)is faithful. However, the smallest dimension of a faithful representation of s1s2 is equal to 4, see [Citation3], Proposition 2.5. This is a contradiction.

Case 4: ker(j1)=s1V(2). Then j2|s1V(2)=(j2j1)|s1V(2):s1V(2)sl3(C)has to be injective. However, s1V(2) contains a 3-dimensional abelian subalgebra, whereas sl3(C) does not. This is a contradiction.

Case 5: ker(j1)=s2V(2). Then (s2V(2))ker(j2)=0. Since ker(j2) is an ideal, which is nonzero as in case 1, we have ker(j2)=s1. So for every xg we can write x=x1+x2 with x1s1 and x2s2V(2). Then we have (j1j2)(x)=j1(x1)j2(x2), and j1j2 is injective if and only if im(j1)im(j2)=0. This is equivalent to sl3(C)=im(j1)im(j2)with im(j1)s1 and im(j2)s2V(2), which is a contradiction to Lemma 3.8.

Case 6: ker(j1)=s1. Then s1 is not contained in ker(j2). So from the six possibilities for the ideal ker(j2), there are left 0, V(2) and s2V(2). We already know that ker(j2) must be nonzero. Also, ker(j2)=V(2) leads to a contradiction as in case 3. Finally ker(j2)=sV(2) and ker(j1)=s1 is exactly the symmetric situation to case 5, and so also leads to a contradiction. □

Now we can prove the following result.

Theorem 3.10.

Let (g,n) be a pair of Lie algebras, where g is perfect non-semisimple and n=sl3(C). Then there is no PA-structure on (g,n).

Proof.

Denote by sa Levi subalgebra of g. Then either ssl2(C)sl2(C) or ssl2(C). In the first case we have gsl2(C)(sl2(C)V(2)) by Proposition 2.3. Then the claim follows by Proposition 3.9. In the second case, by Proposition 2.3, gsl2(C)r, where r is isomorphic to one of the following five Lie algebras V(5),V(2)V(3),V(2)n3(C),n5(C),f2,3(C).

Again we are using the maps ji:gn, and assume that j1j2 is bijective. In the first two cases, either j1 or j2 must be injective on the factor V(3), respectively V(5). This contradicts the fact that sl3(C) does not contain an abelian subalgebra of dimension n3.

Assume that rV(2)n3(C). We will look again at the possibilities for ker(j1). If it has a non-trivial Levi factor, then ker(j1)=g, so that j2:gn is bijective. This is a contradiction. Hence we may assume that ker(j1) is solvable, so that it is contained in V(2)n3(C). Here we can view ker(j1) both as a subalgebra and as a submodule of V(2)n3(C). As an sl2(C)-submodule, V(2)n3(C) is isomorphic to V(2)V(2)V(1), because n3(C)V(2)V(1), see Example 2.2. The submodule V(2)V(2) cannot occur as an ideal, since there is no subalgebra corresponding to it. So we have the following possibilities for ker(j1) as a submodule:

Case 1: ker(j1)=0. Then j1 is an isomorphism. This is a contradiction.

Case 2: ker(j1)V(2). Then j1|sl2(C)n3(C) is injective. This is impossible, because sl2(C)n3(C) has a 3-dimensional abelian subalgebra, see Example 2.2, but sl3(C) does not have one.

Case 3: ker(j1)n3(C). Since ker(j1)ker(j2)=0, it follows that j2 induces an injective homomorphism sl2(C)ker(j1)sl3(C), which is impossible as in case 2.

Case 4: ker(j1)V(2)V(1), with V(1)Z(n3(C)). Then j2 is injective on ker(j1), which is impossible, because V(2)V(1) is an abelian subalgebra of dimension 3.

Case 5: ker(j1)V(2)n3(C). Then j2|V(2)n3(C) is injective. This is impossible, because V(2)n3(C) has a 3-dimensional abelian subalgebra, but sl3(C) does not have one.

Assume that rn5(C). We claim that at least one of the maps j1, j2 must be injective on n5(C). Otherwise j1(Z(n5(C))=j2(Z(n5(C))=0, so that (j1j2)|Z(n5(C))=0, which is a contradiction to the fact that j1j2 is bijective. So we may assume that j1 or j2 is injective. This is impossible since n5(C) contains a 3-dimensional abelian subalgebra, but sl3(C) does not have one.

Finally assume that rf2,3(C). Then again at least one of the maps j1, j2 must be injective on r. If j1 is not injective on r, then Z(r)ker(j1)0. We claim that Z(r)ker(j1). In fact, every ideal a of g satisfying aZ(r)0 also satisfies Z(r)a. To see this, note that that Z(r)=[r,[r,r]], and that the action of sl2(C) on [r,r]/[r,[r,r]] is trivial, since the quotient is 1-dimensional. It follows that the action of sl2(C) on Z(r) coincides with the action on r/[r,r]. By Proposition 2.1 it has no trivial 1-dimensional submodule, since g is perfect. This also implies that Z(r) has no trivial 1-dimensional submodule, so that dim(aZ(r))2. Since dimZ(r)=2 it follows that Z(r)a. Hence if both j1 and j2 are not injective on r, the center Z(r) is contained in ker(j1) and ker(j2), so that (j1j2)|Z(r)=0. This is a contradiction. Consequently, ji:f2,3(C)sl3(C) is an injection for some i, contradicting the fact that f2,3(C) has a 3-dimensional abelian subalgebra, but sl3(C) does not have one. □

It is not clear how to generalize this proof for other simple Lie algebras n.

For case (e) we can prove the following general result.

Proposition 3.11.

Let (g,n) be a pair of Lie algebras, where g is perfect non-semisimple and n is semisimple. Assume that we have a Levi decomposition g=sV, where s is a simple subalgebra and V is an irreducible s-module, considered as abelian Lie algebra. Then there is no PA-structure on (g,n).

Proof.

Suppose that there exists a PA-structure on (g,n). Then by Proposition 2.7 it is of the form x·y={R(x),y} for a Rota-Baxter operator R:gn of weight 1 on n. Moreover, since g and n are not isomorphic, both ker(R) and ker(R+id) are nonzero ideals of g with ker(R)ker(R+id)=0.

We will show that the only ideals of g=sV are 0,V,g. Then it is clear that V is contained in the above intersection, so that V = 0. This is a contradiction. So let a be an ideal of g. Then we obtain a Levi decomposition for a by a=(as)(aV).

Since as is an ideal in s, and s is simple, we have either as=0 or as=s. Also, since aV is an s-submodule of V and V is irreducible, we have either aV=0 or aV=V.

Case 1: aV=0. Then a=as is either zero or s. It follows that either a=0 or a=s. But s is not an ideal in g, so that we obtain a=0.

Case 2: aV=V. Then a=(as)V, which means either a=V or a=g.

So we obtain V = 0 and hence a contradiction. □

For n=sl2(C)sl2(C) we obtain a further result for case (e). In Theorem 4.1 of [Citation9] we have classified all Lie algebras g, such that the pair (g,n) with n=sl2(C)sl2(C) admits a PA-structure. Here we have used the theory of Rota-Baxter operators. It is easy to see that none of the eight cases for g in this classification yields a perfect, non-semisimple Lie algebra. Hence we obtain the following result.

Proposition 3.12.

Let (g,n) be a pair of Lie algebras, where g is perfect non-semisimple and n=sl2(C)sl2(C). Then there is no PA-structure on (g,n).

For case (f) consider the perfect non-semisimple Lie algebra g=sl2(C)V(2). Let (e1,,e5) be a basis of g with sl2(C)=span{e1,e2,e3}, V(2)=span{e4,e5}, and Lie brackets given as follows: [e1,e2]=e3,   [e2,e3]=2e2,   [e3,e4]=e4,[e1,e3]=2e1,   [e2,e4]=e5,   [e3,e5]=e5,[e1,e5]=e4.

Example 3.13.

The pair of Lie algebras (g,n)=(sl2(C)V(2),sl2(C)C2) admits a PA-structure given by e1·e5 =e4,e2·e4=e5,e3·e4=e4,e3·e5=e5.

Here n is reductive with a 2-dimensional center. Such examples are impossible when n is reductive with a 1-dimensional center, as the following result shows.

Proposition 3.14.

Let (g,n) be a pair of Lie algebras, where g is perfect and n is reductive with a 1-dimensional center. Then there is no PA-structure on (g,n).

Proof.

Assume that there exists a PA-structure x·y=L(x)(y) on (g,n). Then by Proposition 2.11 in [Citation5] we have an injective Lie algebra homomorphism φ:g⏧n⋊Der(n), x(x,L(x)).

Writing h=L(g) we obtain a direct vector space decomposition nh=φ(g)h. Note that h is nonzero. Since g is perfect and φ and L are homomorphisms, φ(g) and h are perfect subalgebras of nh. Hence also φ(g)h is perfect, see the proof of Lemma 2.3 in [Citation13], so that nh is perfect. By assumption we have n=[n,n]Z(n), where s=[n,n] is semisimple and Z(n) is 1-dimensional. Since the commutator and the center of n are characteristic ideals in n, and n is an ideal in nh, both [n,n] and Z(n) are ideals in nh. We claim that [Z(n),h]=0 for the Lie bracket in nh. Since Z(n) is an ideal in nh, we have [h,Z(n)]Z(n), so that Z(n) is a 1-dimensional h-module. However, for a perfect Lie algebra, every 1-dimensional module is trivial. The proof is the same as for a semisimple Lie algebra. Hence we obtain [Z(n),h]=0. It follows that [n,h]=[s+Z(n),h]=[s,h]s,since s is an ideal in nh. Because nh is perfect, we have n+h=[n+h,n+h]=[n,n]+[n,h]+[h,h]=s+h.

However, because of n=sZ(n) we have dim(n)=dim(s)+1. Since nh=sh=0, this implies dim(n+h)=dim(s+h)+1. This is a contradiction to n+h=s+h. □

For case (g) we have the following result.

Proposition 3.15.

Let (g,n) be a pair of Lie algebras, where g is perfect and n is complete non-perfect. Then there is no PA-structure on (g,n).

Proof.

Assume that there exists a PA-structure on (g,n). Since n is complete, this PA-structure is given by x·y={R(x),y} for a Rota-Baxter operator R of weight 1. Because g is perfect, it follows by Corollary 2.20 in [Citation9] that n is also perfect. This is a contradiction. □

4 PA-structures with n perfect

In this section we study the existence question of PA-structures on pairs of complex Lie algebras (g,n), where n is perfect non-semisimple. We consider 7 different cases for g, namely (a) g is abelian, (b) g is nilpotent non-abelian, (c) g is solvable non-nilpotent, (d) g is simple, (e) g is semisimple non-simple, (f) g is reductive non-semisimple, and (g) g is complete non-perfect.

For case (a) we have the following result.

Proposition 4.1.

There is no PA-structure on a pair (g,n), where g is abelian and n is perfect.

Proof.

Any PA-structure on (g,n) with g abelian corresponds to an LR-structure on n. However, every Lie algebra admitting an LR-structure is 2-step solvable by Proposition 2.1 in [Citation4]. Hence there exists no PA-structure on (g,n). □

For case (b) we have the following result.

Proposition 4.2.

There is no PA-structure on a pair (g,n), where g is nilpotent non-abelian and n is perfect.

Proof.

Assume that there is a PA-structure on (g,n). Since g is nilpotent, n must be solvable by Proposition 4.3 in [Citation5]. This is a contradiction. □

For case (c), consider the perfect non-semisimple Lie algebra n=sl2(C)V(2) with the Lie brackets given before Example 3.13. We have a decomposition n=n1n2 into subalgebras n1=span{e1,e4,e5} and n2=span{e2,e3}. Then by Propositions 2.7 and 2.13 in [Citation9], the Rota-Baxter operator R given by R(x+y)=y for all xn1,yn2 defines a PA-structure on the pair (g,n), where gn1n2n3(C)r2(C) is solvable non-nilpotent. The matrix of R is given by R=(0000001000001000000000000),and the Lie brackets of g are given by [e1,e5]=e4 and [e2,e3]=2e2.

Example 4.3.

The pair of Lie algebras (g,n) with gn3(C)r2(C) and n=sl2(C)V(2) admits a PA-structure given by e2·e1=e3,   e3·e1=2e1,   e3·e4=e4,e2·e3=2e2,   e3·e2=2e2,   e3·e5=e5,e2·e4=e5.

For the cases (d) and (e) we have the following result.

Proposition 4.4.

There is no PA-structure on a pair (g,n), where g is semisimple and n is perfect non-semisimple.

Proof.

Assume that there exists a PA-structure on (g,n), where g is semisimple. Then by Theorem 3.3 in [Citation12], g is isomorphic to n. This is a contradiction. □

For case (f) we have the following example.

Example 4.5.

The pair of Lie algebras (g,n)=(sl2(C)C2,sl2(C)V(2)) admits a PA-structure given by e4·e2=e5,e5·e1=e4,e4·e3=e4,e5·e3=e5.

Here we use the Lie brackets for sl2(C)V(2) as in Example 3.13, and the standard Lie brackets of sl2(C)=span{e1,e2,e3} for g. This PA-structure can also be realized by the Rota-Baxter operator R=(0000000000000000001000001)for the decomposition n=n1n2, where n1=span{e1,e2,e3}, n2=span{e4,e5}, n=sl2(C)V(2) and x·y={R(x),y}.

Finally, for case (g) we have the following example.

Example 4.6.

The pair of Lie algebras (g,n)=(sl2(C)r2(C),sl2(C)V(2)) admits a PA-structure given by e4·e2=e5,e5·e1=e4,e5·e4=e4,e4·e3=e4,e5·e3=e5,e5·e5=e5.

Here g is complete non-perfect, and the Lie brackets of g are given by the standard brackets for sl2(C)=span{e1,e2,e3}, and by [e4,e5]=e4 for r2(C)=span{e4,e5}.

5 The existence question

We summarize the existence results for post-Lie algebra structures from the previous sections and from the papers [Citation5–10, Citation12] as follows.

Theorem 5.1.

The existence table for post-Lie algebra structures on pairs (g,n) is given as follows:

A checkmark only means that there is some non-trivial pair (g,n) of Lie algebras with the given algebraic properties admitting a PA-structure. A dash means that there does not exist any PA-structure on such a pair. Recall that the classes are (to avoid unnecessary overlap) abelian, nilpotent non-abelian, solvable non-nilpotent, simple, semisimple non-simple, reductive non-semisimple and complete non-perfect.

Additional information

Funding

Dietrich Burde and Mina Monadjem are supported by the Austrian Science Foundation FWF, grant P 33811. Karel Dekimpe is supported by Methusalem grant Meth/21/03 - long term structural funding of the Flemish Government.

References

  • Alev, J., Ooms, A. I., Van den Bergh, M. (2000). The Gelfand-Kirillov conjecture for Lie algebras of dimension at most eight. J. Algebra 227(2):549–581. DOI: 10.1006/jabr.1999.8192.
  • Burde, D. (2006). Left-symmetric algebras, or pre-Lie algebras in geometry and physics. Central Eur. J. Math. 4(3):323–357.
  • Burde, D., Moens, W. A. (2007). Minimal faithful representations of reductive Lie algebras. Archiv der Math. 89(6):513–523.
  • Burde, D., Dekimpe, K., Deschamps, S. (2009). LR-algebras. Contemp. Math. 491:125–140.
  • Burde, D., Dekimpe, K., Vercammen, K. (2012). Affine actions on Lie groups and post-Lie algebra structures. Linear Algebra Appl. 437(5):1250–1263. DOI: 10.1016/j.laa.2012.04.007.
  • 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. DOI: 10.1016/j.jalgebra.2016.05.026.
  • Burde, D., Ender, C., Moens, W. A. (2018). Post-Lie algebra structures for nilpotent Lie algebras. Int. J. Algebra Comput. 28(5):915–933. DOI: 10.1142/S0218196718500406.
  • Burde, D., Gubarev, V. (2019). Rota–Baxter operators and post-Lie algebra structures on semisimple Lie algebras. Commun. Algebra 47(5):2280–2296.
  • Burde, D., Gubarev, V. (2020). Decompositions of algebras and post-associative algebra structures. Int. J. Algebra Comput. 30(3):451–466. DOI: 10.1142/S0218196720500071.
  • Burde, D. (2021). Crystallographic actions on Lie groups and post-Lie algebra structures. Commun. Math. 29(1): 67–89. DOI: 10.2478/cm-2021-0003.
  • Burde, D., Dekimpe, K., Monadjem, M. (2022). Rigidity results for Lie algebras admitting a post-Lie algebra structure. Int. J. Algebra Comput. 32(08):1495–1511. DOI: 10.1142/S0218196722500679.
  • Burde, D., Moens, W. A. (2022). Semisimple decompositions of Lie algebras and prehomogeneous modules. J. Algebra 604:664–681. DOI: 10.1016/j.jalgebra.2022.04.015.
  • Ebrahimi-Fard, K., Lundervold, A., Mencattini, I., Munthe-Kaas, H. (2015). Post-Lie algebras and isospectral flows. SIGMA Symmetry Integrability Geom. Methods Appl. 11:Paper 093, 16 pp. DOI: 10.3842/SIGMA.2015.093.
  • Helmstetter, J. (1979). Radical d’une algèbre symétrique a gauche. Ann. Inst. Fourier 29:17–35. DOI: 10.5802/aif.764.
  • Turwokski, P. (1988). Low-dimensional real Lie algebras. J. Math. Phys. 29(10):2139–2144.
  • Turkowski, P. (1992). Structure of real Lie algebras. Linear Algebra Appl. 171:197–212. DOI: 10.1016/0024-3795(92)90259-D.
  • Vallette, B. (2007). Homology of generalized partition posets. J. Pure Appl. Algebra 208, no. 2 (2007), 699–725.