361
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Weak c-ideals of Leibniz algebras

&
Pages 4676-4685 | Received 02 Mar 2023, Accepted 12 May 2023, Published online: 25 May 2023

Abstract

A subalgebra B of a Leibniz algebra L is called a weak c-ideal of L if there is a subideal C of L such that L=B+C and BCBL where BL is the largest ideal of L contained in B. This is analogous to the concept of a weakly c-normal subgroup, which has been studied by a number of authors. We obtain some properties of weak c-ideals and use them to give some characterizations of solvable and supersolvable Leibniz algebras generalizing previous results for Lie algebras. We note that one-dimensional weak c-ideals are c-ideals, and show that a result of Turner classifying Leibniz algebras in which every one-dimensional subalgebra is a c-ideal is false for general Leibniz algebras, but holds for symmetric ones.

2020 Mathematics Subject Classification:

1 Introduction

An algebra L over a field F is called a Leibniz algebra if, for every x,y,zL, we have [x,[y,z]]=[[x,y],z][[x,z],y].

In other words the right multiplication operator Rx:LL:y[y,x] is a derivation of L. As a result such algebras are sometimes called right Leibniz algebras, and there is a corresponding notion of left Leibniz algebras, which satisfy [x,[y,z]]=[[x,y],z]+[y,[x,z]].

Clearly the opposite of a right (left) Leibniz algebra is a left (right) Leibniz algebra, so, in most situations, it does not matter which definition we use. A symmetric Leibniz algebra L is one which is both a right and left Leibniz algebra and in which [[x,y],[x,y]]=0 for all x,yL. This last identity is only needed in characteristic two, as it follows from the right and left Leibniz identities otherwise (see [Citation6, Lemma 1]). Symmetric Leibniz algebras L are flexible, power associative and have x3=0 for all xL (see [Citation5, Proposition 2.37]), and so, in a sense, are not far removed from Lie algebras.

Every Lie algebra is a Leibniz algebra and every Leibniz algebra satisfying [x,x]=0 for every element is a Lie algebra. They were introduced in 1965 by Bloh [Citation3] who called them D-algebras, though they attracted more widespread interest, and acquired their current name, through work by Loday and Pirashvili [Citation7, Citation8]. They have natural connections to a variety of areas, including algebraic K-theory, classical algebraic topology, differential geometry, homological algebra, loop spaces, noncommutative geometry and physics. A number of structural results have been obtained as analogues of corresponding results in Lie algebras.

The Leibniz kernel is the set I= span {x2:xL}. Then I is the smallest ideal of L such that is a Lie algebra. Also [L,I]=0.

We define the following series: L1=L,Lk+1=[Lk,L](k1) and L(0)=L,L(k+1)=[L(k),L(k)](k0).

Then L is nilpotent of class n (resp. solvable of derived length n) if Ln+1=0 but Ln0 (resp. L(n)=0 but L(n1)0) for some nN. It is straightforward to check that L is nilpotent of class n precisely when every product of n+1 elements of L is zero, but some product of n elements is non-zero.The nilradical, N(L), (resp. radical, R(L)) is the largest nilpotent (resp. solvable) ideal of L.

The subalgebra lattice of a Lie algebra has been extensively studied over many decades. More recently, the properties of the subalgebra lattice of a Leibniz algebra have been investigated in [Citation11], and similar results for restricted Lie algebras have been carried out in [Citation10]. In [Citation14] we introduced the concept of a weak c-ideal, which is analogous to the concept of a weakly c-normal subgroup as introduced by Zhu, Guo and Shum in [Citation16] and which has since been further studied by a number of authors. It is a generalization of the concept of a c-ideal, as introduced in [Citation13]. Here we investigate the extent to which the results in [Citation14] can be extended to Leibniz algebras.

In Section 2 we give some basic properties of weak c-ideals; in particular, it is shown that weak c-ideals inside the Frattini subalgebra of a Leibniz algebra L are necessarily ideals of L. In Section 3 we first show that all maximal subalgebras of L are weak c-ideals of L if and only if L is solvable, and that L has a solvable maximal subalgebra that is a weak c-ideal if and only if L is solvable. Unlike the corresponding results for c-ideals, it is necessary to restrict the underlying field to characteristic zero, as was shown by an example in [Citation14]. Finally we have that if all maximal nilpotent subalgebras of L are weak c-ideals, or if all Cartan subalgebras of L are weak c-ideals and F has characteristic zero, then L is solvable.

In Section 4 we show that if L is a solvable symmetric Lie algebra over a general field and every maximal subalgebra of each maximal nilpotent subalgebra of L is a weak c-ideal of L then L is supersolvable. If each of the maximal nilpotent subalgebras of L has dimension at least two then the assumption of solvability can be removed. Similarly if the field has characteristic zero and L is not three-dimensional simple then this restriction can be removed.

In the final section we see that every one-dimensional subalgebra is a weak c-ideal if and only if it is a c-ideal, and go on to study the class of Leibniz algebras in which every one-dimensional subalgebra is a c-ideal. It is shown that the cyclic subalgebras in this class are at most two dimensional. There is a characterization of all of the algebras in this class given by Turner in [Citation15, Theorem 3.2.9], but an example is given to show that this result is false. A number of properties of such algebras are given, but a full classification appears complicated. However, it is shown finally that Turner’s result does hold for symmetric Leibniz algebras.

Throughout, the term “Leibniz algebra” will refer to a finite-dimensional right Leibniz algebra over a field F. If A, B are subalgebras of L with AB, the centraliser of A in B, CB(A)={bB|[b,A]+[A,b]=0}. The normaliser of A in B, NB(A)={bB|[b,A]+[A,b]A}. Algebra direct sums will be denoted by , whereas direct sums of the vector space structure alone will be written as +̇. Subsets will be denoted by “” and proper subsets by “”.

2 Preliminary results

Definition 1.

[Citation9] Let L be a Leibniz algebra over a field F and B be a subalgebra of L. We call B a subideal of L if there is a chain of subalgebras B=BtBt1B0=Lsuch that Bi is an ideal of Bi1 for each 1it.

Definition 2.

[Citation9] Let L be a Leibniz algebra and H a subalgebra of L. Then H is called a c-ideal of L if there is an ideal K of L such that L=H+K and HK is contained in the core of H (with respect to L), denoted by HL, where this is the largest ideal of L contained in H.

Definition 3.

A subalgebra B of a Leibniz algebra L is a weak c-ideal of L if there exists a subideal C in L such that L=B+C and BCBL.

Definition 4.

A Leibniz algebra L is called weakly c-simple if L does not contain any weak c-ideals other than L, the trivial subalgebra 0 and the Leibniz kernel I. It is simple if these same three subalgebras are the only ideals of L.

Lemma 2.1.

Let L be a Leibniz algebra. Then the following statements are valid:

  1. Let B be a subalgebra of L . If B is a c-ideal of L then B is a weak c-ideal of L.

  2. L is weakly c-simple if and only if L is simple.

  3. If B is a weak c-ideal of L and K is a subalgebra with BKL then B is a weak c-ideal of K .

  4. If A is an ideal of L and AB then B is a weak c-ideal of L if and only if B/A is a weak c-ideal of L/A.

Proof.

  1. This is apparent from the definition.

  2. Suppose first that L is simple and let B be a weak c-ideal with BL. Then by definition of weak c-ideal L=B+C and BCBL

where C is a subideal of L. But since L is simple BL must be 0 or I. If BL=0 then C=L, since C0 and is a subideal of L, whence B=0. If BL=I then, similarly, C=L and hence B=I. Hence L is weakly c-simple.

Conversely, suppose L is weakly c-simple. Then, since every ideal of L is a weak c-ideal, L must be simple.

  1. If B is a weak c-ideal of L, then there exists a subideal C of L such that L=B+C and BCBL.

Now K=KL=K(B+C)=B+(KC). Since C is a subideal of L there exists a chain of subalgebras C=CnCn1C0=Lwhere Cj is an ideal of Cj1 for each 1jn. If we intersect each term in this chain with K we get CK=CnKCn1KC0K=LK=K and obviously CjK is an ideal of Cj1K for each 0jn. Thus CK is a subideal of K. Also, B(CK)BK and so that B is a weakc-ideal of K.

  1. Suppose first that B/A is a weakc-ideal of L/A. Then there exists a subideal C/A of L/A such that LA=BA+CA and BACA(BA)L/A=BLA.

It follows that L=B+C and BCBL where C is a subideal of L.

Suppose conversely that A is an ideal of L with AB and B is a weakc-ideal of L. Then there exists a C subideal of L such that L=B+C and BCBL.

Since A is an ideal and AB the factor algebra LA=B+CA=BA+C+AAwhere (C+A)/A is a subideal of L/A and BAC+AA=B(C+A)A=A+BCABLA=(BA)L/A

so B/A is a weakc-ideal of L/A.

The Frattini subalgebra F(L) of a Leibniz algebra L is the intersection of all of the maximal subalgebras of L. The Frattini ideal, F(L)L, of L is denoted by ϕ(L).

The next result is a generalization of [Citation13, Proposition 2.2] and the same proof works, but we will include it for completeness.

Proposition 2.2.

Let B, C be subalgebras of L with BF(C). If B is a weak c-ideal of L, then B is an ideal of L and Bφ(L).

Proof.

Suppose that L=B+K where K is a subideal of L and BKBL. Then C=CL=C(B+K)=B+CK=CKsince BF(C). Hence, BCK, giving B=BKBL and B is an ideal of L. It then follows from [Citation12, Lemma 4.1] that Bφ(L).

An ideal A is complemented in L if there is a subalgebra U of L such that L=A+U and AU=0. We adapt this to define a subideal complement as follows:

Definition 5.

Let L be a Leibniz algebra and let B be a subalgebra of L. Then B has a subideal complement in L if there is a subideal C of L such that L=B+C and BC=0.

Then we have the following lemma:

Lemma 2.3.

If B is a weak c-ideal of a Leibniz algebra L, then B/BL has a subideal complement in L/BL. Conversely, if B is a subalgebra of L such that B/BL has a subideal complement in L/BL, then B is a weak c-ideal of L.

Proof.

Let B be a weak c-ideal of L. Then there exists a subideal C of L such that B+C=L and BCBL. If BL=0 then BC=0 and C is a subideal complement of B in L. So, assume that BL0, then we can construct the factor algebras B/BL and (C+BL)/BL. If we intersect these two factor algebras; BBLC+BLBL=B(C+BL)BL=BL+(BC)BLBLBL=0

Hence, (C+BL)/BL is a subideal complement of B/BL in L/BL. Conversely, if K is a subideal of L such that K/BL is a subideal complement of B/BL in L/BL then we have that LBL=BBL+KBL and BBLKBL=0

Then L=B+K and BK=0BL, whence B is a weak c-ideal of L. □

3 Some characterizations of solvable algebras

Theorem 3.1.

Let L be a Leibniz algebra over a field F of characteristic zero and let B be an ideal of L. Then B is solvable if and only if every maximal subalgebra of L not containing B is a weak c-ideal of L.

Proof.

Suppose first that B is solvable and let M be a maximal subalgebra of L not containing B. Then there exists kN such that B(k+1)M, but B(k)M. Clearly L=M+B(k) and B(k)M is an ideal of L, so B(k)MML. It follows that M is a c-ideal and hence a weak c-ideal of L.

Conversely, suppose that every maximal subalgebra of L not containing B is a weak c-ideal of L. Let M/I be a maximal subalgebra of L/I not containing (B+I)/I. Then M is a maximal subalgebra of L not containing B, and so is a weak c-ideal of L. Hence M/I is a weak c-ideal of L/I, by Lemma 2.1. Since L/I is a Lie algebra, it follows from [Citation14, Theorem 3.2] that (B+I)/I is solvable. It follows that B+I, and hence B, is solvable. □

Corollary 3.2.

Let L be a Leibniz algebra over a field F of characteristic zero. Then L is solvable if and only if every maximal subalgebra of L is a weak c-ideal of L.

Lemma 3.3.

Let L=U+C be a Leibniz algebra, where U is a solvable subalgebra of L and C is a subideal of L. Then there exists n0N such that L(n0)C.

Proof.

This is the same as for [Citation14, Lemma 3.5]. □

Theorem 3.4.

Let L be a Leibniz algebra over a field F of characteristic zero. Then L has a solvable maximal subalgebra that is a weak c-ideal of L if and only if L is solvable.

Proof.

Suppose first that L has a solvable maximal subalgebra M that is a weak c-ideal of L. We show that L is solvable. Let L be a minimal counter-example. Then there is a subideal K of L such that L=M+K and MKML. If ML0 then L/ML is solvable, by the minimality assumption, and ML is solvable, whence L is solvable, a contradiction. It follows that ML=0 and L=M+̇K. If R is the solvable radical of L then RML=0, so L is a semisimple Lie algebra, by [Citation2, Theorem 1]. But now, for all n1, L=L(n)KL, by Lemma 3.3, a contradiction. The result follows.

The converse follows from Corollary 3.2. □

Theorem 3.5.

Let L be a Leibniz algebra over a field of characteristic zero such that all maximal nilpotent subalgebras are weak c-ideals of L. Then L is solvable.

Proof.

Suppose that L is not solvable but that all maximal nilpotent subalgebras of L are weak c-ideals of L. Let L=R+̇S be the Levi decomposition of L, where S0 is a semisimple Lie algebra ([Citation2, Theorem 1]). Let B be a maximal nilpotent subalgebra of S and U be a maximal nilpotent subalgebra of L containing it. Then there is a subideal C of L such that L=U+C and UCUL. It follows from Lemma 3.3 that S=S(n0)L(n0)C, and so BUCUL, whence SUL0. But SUL is an ideal of S and so is semisimple. Since U is nilpotent, this is a contradiction. □

Definition 6.

A Cartan subalgebra of a Leibniz algebra L is a nilpotent subalgebra C such that C=NL(C). Over a field of characteristic zero such subalgebras certainly exist (see [Citation1, Section 6]).

The following result is a generalization of a result of Dixmier in [Citation4].

Lemma 3.6.

Let L be a Leibniz algebra over a field of characteristic zero with non-zero Levi factor S. If H is a Cartan subalgebra of S and B is a Cartan subalgebra of its centralizer in the solvable radical of L, then H+B is a Cartan subalgebra of L.

Proof.

Let R be the solvable radical of L. Then (H+B)r=Hr+Br for all r1, so H+B is a nilpotent subalgebra of L. Let xNL(H+B) and put x=s+r where sS and rR. Then [x,H]+[H,x]=[s,H]+[H,s]+[r,H]+[H,r]H+B,so [H,s]+[s,H]H, whence sNS(H)=H. Moreover, [x,B]+[B,x]=[s,B]+[B,s]+[r,B]+[B,r]H+B,

so [B,r]+[r,B]B, since [s,B]+[B,s]=0, whence rNR(B)=B. Thus, NL(H+B)=H+B and H+B is a Cartan subalgebra of L. □

Theorem 3.7.

Let L be a Leibniz algebra over a field of characteristic zero in which every Cartan subalgebra of L is a weak c-ideal of L. Then L is solvable.

Proof.

Suppose that every Cartan subalgebra of L is a weak c-ideal of L and that L has a non-zero Levi factor S. Let H be a Cartan subalgebra of S and let B be a Cartan subalgebra of its centralizer in the solvable radical of L. Then C=H+B is a Cartan subalgera of L, by Lemma 3.6, and there is a subideal K of L such that L=C+K and CKCL. Now there is an n02 such that L(n0)K by Lemma 3.3. But SL(n0)K and so CSCKCL giving CSCLS=0,a contradiction. It follows that S=0 and, hence, that L is solvable. □

4 Some characterizations of supersolvable algebras

In this section we will restrict attention to symmetric Leibniz algebras. We know of no examples of a Leibniz algebra which is not symmetric and for which the results are false, but have been unable to establish them in that more general case. First we need the following lemma which holds in the general case.

Lemma 4.1.

[Citation15, Lemma 5.1.2] Let L be a Leibniz algebra over any field, let A be an ideal of L and U/A be a maximal nlpotent subalgebra of L/A. Then, U=C+A where C is a maximal nilpotent subalgebra of L.

Theorem 4.2.

Let L be a solvable symmetric Leibniz algebra over any field F in which every maximal subalgebra of each maximal nilpotent subalgebra of L is a weak c-ideal of L. Then L is supersolvable.

Proof.

Let L be a minimal counter-example. If I=0 the result follows from [Citation14, Theorem 4.5], so suppose that I0 and let A be a minimal ideal of L contained in I. Since [L,I]=[I,L]=0, dimA=1. Let U/A be a maximal nilpotent subalgebra of L/A and let B/A be a maximal subalgebra of U/A. Then U=C+A where C is a maximal nilpotent subalgebra of L, by Lemma 4.1. But C+A is nilpotent, so AC=U. Hence, B is a maximal subalgebra of C and there is a subideal K of L such that L=B+K and BKBL. Now LA=BA+K+AA and K+AA is a subideal of LA.

Moreover, BAK+AA=A+BKAA+BLA(BA)L/A.

It follows that L/A satisfies the same hypothesis as L, and so L/A is supersolvable, by the minimality of L. Hence, L is supersolvable. □

If L has no one-dimensional maximal nilpotent subalgebras, we can remove the solvability assumption from the above result provided that F has characteristic zero.

Corollary 4.3.

Let L be a symmetric Leibniz algebra over a field F of characteristic zero in which every maximal nilpotent subalgebra has dimension at least two. If every maximal subalgebra of each maximal nilpotent subalgebra of L is a weak c-ideal of L, then L is supersolvable.

Proof.

Let N be the nilradical of L, and let xN. Then xC for some maximal nilpotent subalgebra C of L. Since dimC>1, there is a maximal subalgebra B of C with xB. Then there is a subideal K of L such that L=B+K and BKBLCLN. Clearly, xK, since otherwise xBKN.

Now L/I=(B+I)/I+(K+I)/I where (B+I)/I is nilpotent and (K+I)/I is a subideal of L/I. It follows from [Citation14, Lemma 4.2] that LrK+I for some rN. Hence Lr+1K. We have shown that if xN there is a subideal K of L with xK and Lr+1K. Suppose that L is not solvable. Then there is a semisimple Levi factor S of L. Choose xS. Then xS=Sr+1K, a contradiction. Thus L is solvable and the result follows from Theorem 4.2. □

If L has a one-dimensional maximal nilpotent subalgebra, then we can also remove the solvability assumption from Corollary 4.3, provided that the underlying field F has again characteristic zero and L is not three-dimensional simple.

Corollary 4.4.

Let L be a symmetric Leibniz algebra over a field F of characteristic zero. If every maximal subalgebra of each maximal nilpotent subalgebra of L is a weak c-ideal of L, then L is supersolvable or three-dimensional simple.

Proof.

If every maximal nilpotent subalgebra of L has dimension at least two, then L is supersolvable by Corollary 4.3. So we need only consider the case where L has a one-dimensional maximal nilpotent subalgebra say Fx. Suppose first that I=0. Then L is a Lie algebra and the result follows from [Citation14, Corollary 4.7].

So now let I0 and let L be a minimal-counter-example. Then L has a minimal ideal AI. As in the proof of Theorem 4.2, L/A satisfies the same hypothesis as L and so is supersolvable or three-dimensional simple. In the former case, L is solvable and so is supersolvable, by Theorem 4.2. In the latter case, L=AS where S is three-dimensional simple, by Levi’s Theorem. But now L is a Lie algebra and the result follows again from [Citation14, Corollary 4.7]. □

5 Leibniz algebras in which every one-dimensional subalgebra is a weak c-ideal

Proposition 5.1.

For a one-dimensional subalgebra Fx of a Leibniz algebra L the following are equivalent:

  1. Fx is a weak c-ideal of L ;

  2. Fx is a c-ideal of L ; and

  3. either Fx is an ideal of L , or there is an ideal B of L such that L=B+̇Fx and xL2 .

Proof.

(i) and (ii) are equivalent since a subideal of codimension one in L is an ideal.

If (ii) holds, then there is an ideal B in L such that L=Fx+B, and FxB(Fx)L=0 or Fx. The former implies that L=B+̇Fx and xL2B; the latter implies that Fx is an ideal of L. Hence (iii) holds. The converse is clear. □

Definition 7.

Put J=xL|x2=0. Note that IJ.

Corollary 5.2.

Let L be a Leibniz algebra over any field. Then every one-dimensional subalgebra is a c-ideal if and only if L2JAsoc(L).

Proof.

Clearly, if Fx is a one-dimensional subalgebra of L, then xJ. It follows from Proposition 5.1 that if every one-dimensional subalgebra of L is a c-ideal, then L2JAsoc(L).

So suppose that L2JAsoc(L), and let Fx be a one-dimensional subalgebra of L. If xL2, then Fx is an ideal of L. If xL2, let B be a subspace containing L2 which is complementary to Fx. Then B is an ideal of L and L=Fx+̇B. Hence, Fx is a c-ideal, by Proposition 5.1 (iii). □

Proposition 5.3.

Let L=x be a cyclic Leibniz algebra. Then every one-dimensional subalgebra of L is a c-ideal if and only if dimL2.

Proof.

Suppose that every one-dimensional subalgebra of L is a c-ideal and that dimL>1. If Fx2 is not an ideal of L, there is an ideal B of L such that L=Fx2+B. But then x+λx2B for some λF, whence x2=[x,x+λx2]B, a contradiction. Thus Fx2 is an ideal of L, x3=λx2 for some λF and L=Fx+Fx2.

Suppose conversely that dimL=2. Then L=Fx+Fx2, where x3=0 or x3=x2. In the former case, the only one-dimensional subalgebra is Fx2 and that is an ideal of L. In the latter case, the one-dimensional subalgebras are Fx2, which is an ideal, and F(xx2), which is complemented by Fx2. □

In [Citation15] the following result appears.

Theorem 5.4.

Let L be a Leibniz algebra over any field F . Then all one-dimensional subalgebras of L are c-ideals of L if and only if:

  1. L3=0 ; or

  2. L=AB , where A is an abelian ideal of L and B is an almost abelian ideal of L .

Proof.

See [Citation15, Theorem 3.2.9, p. 26]. □

Turner defines a subalgebra B of a Leibniz algebra L to be almost abelian if B=Fx+̇D where, D is abelian and [d,x]D for all dD. (She actually has the products the other way around as she is dealing with left Leibniz algebras, whereas, here we are concerned with right Leibniz algebras.) However, this definition is problematic as it appears to be assumed in the proof that [d,x]=d for all dD, and that does not follow from the definition. Also, nothing is said about [x, d]. Elsewhere in the literature there have been defined two types of almost abelian Leibniz algebras: B=Fx+̇D is called an almost abelian Lie algebra if [d,x]=[x,d]=d for all dD, and is an almost abelian non-Lie Leibniz algebra if [d,x]=d for all dD, all other products being zero in each case.

Moreover, the result is false, as the following example shows.

Example 5.1.

Let L be the three-dimensional Leibniz algebra over a field of characteristic different from 2 with basis a, b, x and non-zero products a2=b, [a,x]=[x,a]=12a, [b,x]=b. Then (αa+βb+γx)2=0 if and only if α2=βγ. If α=0 then, either β0, in which case Fx is complemented by the ideal Fa+Fb, or γ=0, in which case Fb is an ideal of L. If α0, then F(αa+βb+α2βx) is complemented by the ideal Fa+Fb. It follows that every one-dimensional subalgebra is a c-ideal. However, L is not of the form given in Theorem 5.4.

In fact, the structure of Leibniz algebras in which all one-dimensional subalgebras are c-ideals can be more complicated than is claimed by Theorem 5.4. The best that we can achieve currently is the following.

Lemma 5.5.

Let L be a Leibniz algebra in which every one-dimensional subalgebra is a c-ideal. Then

  1. all minimal abelian ideals are one dimensional;

  2. if Asoc(L)=Fa1Far and xL , then aiZ(L) or [ai,x]=λxai for some 0λxF , 1ir .

  3. Let Asoc(L)=Z(L)D where [a,x]=λxa for all aD , xL . Then, either D=0 or L=(Z(L)D)+̇C+̇Fx where [D,C]=[C,D]=0 , (Z(L)D)+̇C is an ideal of L and [a,x]=a for all aD .

Proof.

  1. Let A be a minimal abelian ideal of L and let aA. If AFa then there is an ideal K of L such that L=Fa+̇K. But AK=0 so A=Fa, a contradiction.

  2. Suppose that [ai,x]=λai and [aj,x]=μaj where λμ. Then F(ai+aj) is not an ideal of L and so there is an ideal M of L such that L=F(ai+aj)+̇M. Clearly, one of ai and aj does not belong to M. Suppose that aiM, so L=FaiM and aiZ(L).

  3. Let Λ:LF be given by Λ(x)=λx. This is a linear transformation. Hence, either Im Λ=0, in which case D=0, or else L= Ker Λ+̇Fx and λx=1. Put L=(ZD)+̇C+̇Fx, where C Ker Λ. Let dD, cC with [c,d]=λd. Then λ2d=λ[c,d]=[c[c,d]]=[c2,d][[c,d],c]=0,

so λ=0. Hence [D,C]=[C,D]=0. It is straightforward to check that Ker Λ is an ideal of L.

However, we can retrieve Theorem 5.4 for symmetric Leibniz algebras.

Theorem 5.6.

Let L be a symmetric Leibniz algebra over any field F . Then all one-dimensional subalgebras of L are c-ideals of L if and only if:

  1. L3=0 ; or

  2. L=AB , where A is an abelian ideal of L and B is an almost abelian Lie ideal of L .

Proof.

Suppose that all one-dimensional subalgebras of L are c-ideals of L. First note that, if x,yL, then [x,y]2L2J, so L2Asoc(L), by Corollary 5.2. Also, L must have the structure given in Lemma 5.5 (iii). If D=0 then L2Z(L) and (i) holds. So suppose that D0.

Let 0aD. If zZ, then F(z+a) is not an ideal of L and so z+aL2. But a=[a,x]L2, so zL2 and ZL2=0. It follows that L2=D. Now, if cC, we have that c2L2=D, so c3=c2. But c3=0, by [Citation5, Proposition 2.17]. If [c,x]=0, then [x,c]L2=D and so [x,c]=[[x,c],x]=[x,[c,x]]=0, since L is also a flexible algebra, by [Citation5, Proposition 2.17] again. Hence Fc is an ideal of L. But then cL2C=DC=0.

So suppose that [c,x]0. Then [c[c,x],x]=0 and [x,c[c,x]]=[x,c][x,[c,x]]=[x,c][[x,c],x]=0, using the flexible law again. Also, [c[c,x],c[c,x]]=[c,[c,x]][[c,x],c]=[c,[c,x]][c,[x,c]]=0, since [c,x]+[x,c]I. It follows that F(c[c,x]) is an ideal of L and hence is inside L2=D. Thus, cCD=0, so C=0.

Finally, for all aD, [a,x]=a. Also, [a,x]+[x,a]I, so 0=[[a,x],x]+[[x,a],x]=[a,x]+[x,a] and [x,a]=a. Thus, L is as described in (ii). □

Acknowledgments

The authors are grateful to the referee for her/his careful reading of the paper and for helpful comments.

References

  • Barnes, D. W. (2011). Some theorems on Leibniz algebras. Commun. Algebra 39:2463–2472. DOI: 10.1080/00927872.2010.489529.
  • Barnes, D. W. (2012). On Levi’s Theorem for Leibniz algebras. Bull. Aust. Math. Soc. 86:184–185. DOI: 10.1017/S0004972711002954.
  • Bloh, A. (1965). On a generalization of the concept of Lie algebra. Dokl. Akad. Nauk SSSR. 165:471–473.
  • Dixmier, J. (1956). Sous-algebres de Cartan et decompositions de Levi dans les algebres de Lie (French). Proc. Trans. Royal Soc. Can. Section III (3) 50:17–21.
  • Feldvoss, J. (2017). Leibniz algebras as nonassociative algebras. In: Vojtĕchovský, P., Bremner, M. R., Scott Carter, J., Evans, A. B., Huerta, J., Kinyon, M. K., Eric Moorhouse, G., Smith, J. D. H., eds. Nonassociative Mathematics and its Applications, Contemporary Mathematics, Vol. 721. Providence, RI: American Mathematical Society, pp. 115–149.
  • Jibladze, M., Pirashvili, T. (2020). Lie theory for symmetric Leibniz algebras. J. Homotopy Related Struct. 15:167–183. DOI: 10.1007/s40062-019-00248-x.
  • Loday, J.-L. (1993). Une version non commutative des algèbres de Lie: les algèbres de Leibniz. Enseign. Math. (2) 39(3–4):269–293.
  • Loday, J.-L., Pirashvili, T. (1993). Universal enveloping algebras of Leibniz algebras and (co)homology. Math. Ann. 296(1):139–158. DOI: 10.1007/BF01445099.
  • Misra, K. C., Stitzinger, E., Yu, X. (2021). Subinvariance in Leibniz algebras. J. Algebra 567(1):128–138. DOI: 10.1016/j.jalgebra.2020.09.031.
  • Páez-Guillán, P., Siciliano, S., Towers, D. A. (2023). On the subalgebra lattice of a restricted Lie algebra. Linear Algebra Appl. 660:47–65. DOI: 10.1016/j.laa.2022.12.004.
  • Siciliano, S., Towers, D. A. (2022). On the subalgebra lattice of a Leibniz algebra. Commun. Algebra 50:255–267. DOI: 10.1080/00927872.2021.1956510.
  • Towers, D. A. (1973). A Frattini theory for algebras. Proc. London Math. Soc. 27:440–462. DOI: 10.1112/plms/s3-27.3.440.
  • Towers, D. A. (2009). C-ideals of Lie algebras. Commun. Algebra 37:4366–4373. DOI: 10.1080/00927870902829023.
  • Towers, D. A., Ciloglu, Z. (2021). Weak c-ideals of a Lie algebra. Turk. J. Math. 45:1940–1948.
  • Turner, B. N. (2016). Some criteria for solvable and supersolvable Leibniz algebras. Ph.D. thesis, North Carolina State University. http://www.lib.ncsu.edu/resolver/1840.16/11085.
  • Zhu, L., Guo, W., Shum, K. P. (2002). Weakly c-normal subgroups of finite groups and their properties. Commun. Algebra 30:5505–5512. DOI: 10.1081/AGB-120015666.