826
Views
3
CrossRef citations to date
0
Altmetric
Articles

On the subalgebra lattice of a Leibniz algebra

& ORCID Icon
Pages 255-267 | Received 23 Oct 2020, Accepted 13 Jul 2021, Published online: 02 Aug 2021

Abstract

In this paper, we begin to study the subalgebra lattice of a Leibniz algebra. In particular, we deal with Leibniz algebras whose subalgebra lattice is modular, upper semi-modular, lower semi-modular, distributive, or dually atomistic. The fact that a non-Lie Leibniz algebra has fewer one-dimensional subalgebras in general results in a number of lattice conditions being weaker than in the Lie case.

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 algebra. Every Lie algebra is a Leibniz algebra and every Leibniz algebra satisfying [x,x]=0 for every element is a Lie algebra.

Leibniz algebras were first considered by Bloh in [Citation8] and Loday in [Citation18], and nowadays they play an important role in several areas of mathematics such as homological algebra, algebraic K-theory, differential geometry, algebraic topology, noncommutative geometry, etc. As a result, the theory of these algebraic structures has been developing intensively in the last three decades and many important theorems for Lie algebras have been considered in the more general context of Leibniz algebras.

Now, the set of subalgebras of a nonassociative algebra forms a lattice under the operations and , where the join of two subalgebras is the subalgebra generated by their set-theoretic union, and the meet is the usual intersection. The relationship between the structure of a Lie algebra L and that of the lattice L(L) of all subalgebras of L has been studied by many authors. Much is known about modular subalgebras (modular elements in L(L)) through a number of investigations including [Citation1, Citation12, Citation13, Citation25–27]. Other lattice conditions, together with their duals, have also been studied. These include semi-modular, upper semi-modular, lower semi-modular, upper modular, lower modular, and their respective duals. For a selection of results on these conditions see [Citation9, Citation11, Citation14, Citation15, Citation17, Citation19, Citation21, Citation22, Citation24, Citation28].

The subalgebra lattice of a Leibniz algebra, however, is rather different; in a Lie algebra every element generates a one-dimensional subalgebra, whereas in a Leibniz algebra elements can generate subalgebras of any dimension. Thus, one could expect different results to hold for Leibniz algebras and, as we shall see, this is indeed the case. In Section 2, we show that cyclic Leibniz algebras are determined by their subalgebra lattice. In Section 3, we classify Leibniz algebras over a field of characteristic zero in which every subalgebra is an intersection of maximal subalgebras; an extra family arises in the non-Lie case.

In Section 4, we study upper semi-modular Leibniz algebras. There turn out to be many more of these than in the Lie algebra case; in particular, all nilpotent cyclic Leibniz algebras and a subset of the extraspecial Leibniz algebras introduced in [Citation16] belong to this class. Section 5 is devoted to lower semi-modular Leibniz algebras. The situation here turns out to be very similar to the Lie algebra case. The final section deals with modular Leibniz algebras. We see that this is a much larger class than in the Lie algebra case; in particular, any Leibniz algebra which is of the form EC, where E is an upper semi-modular extraspecial or a nilpotent cyclic Leibniz algebra and C is a central ideal is modular. However, these do not exhaust all of the modular Leibniz algebras even of dimension four.

We fix some notation and terminology. Unless otherwise stated, throughout the paper all algebras are assumed to be finite-dimensional. Algebra direct sums will be denoted by ⊕, whereas vector space direct sums will be denoted by +̇.

Let L be a Leibniz algebra over a field F. For a subset S of L, we denote by S the subalgebra generated by S. The Leibniz kernel is defined as Leib(L)=span{x2| xL}. Note that Leib(L) is the smallest ideal of L such that L/Leib(L) is a Lie algebra. Also, [L,Leib(L)]=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 (respectively solvable of derived length n) if Ln+1=0 but Ln0 (respectively 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, no matter how associated, is zero, but some product of n elements is non-zero (see, e.g., Proposition 5.4 of [Citation10]). The nilradical N(L) (respectively radical, R(L)) is the largest nilpotent (respectively solvable) ideal of L.

The Frattini ideal of L, ϕ(L), is the largest ideal contained in all maximal subalgebras of L; if ϕ(L)=0 we say that L is ϕ-free. The right centralizer of a subalgebra U of L is the set CLr(U)={xL|[U,x]=0}. It is easy to check that if U is an ideal of L then so is CLr(U). We denote by soc(L) the socle of L, that is, the sum of all minimal ideals of L.

2. Cyclic Leibniz algebras

A Leibniz algebra L is said to be cyclic if it is generated by a single element. In this case, L has a basis a,a2,,an (n > 1) and product [an,a]=α2a2++αnan. Let T be the matrix for Ra with respect to the above basis. Then T is the companion matrix for p(x)=xnαnxn1α2x=p1(x)n1ps(x)ns, where the pj’s are the distinct irreducible factors of p(x). We shall speak of these factors as being the distinct irreducible factors associated with L. The purpose of this short section is to characterize cyclic Leibniz algebras in terms of their subalgebra lattice. We shall need the following result, which is proved in [Citation6, Corollary 4.3].

Theorem 2.1.

The maximal subalgebras of the cyclic Leibniz algebra L are precisely the null spaces of rj(Ra), where rj(x)=p(x)/pj(x) and the pj(x) are the distinct irreducible factors associated with L for j=1,,s.

Theorem 2.2.

Let L be a Leibniz algebra over a field F with |F|>s. Then L has precisely s maximal subalgebras if and only if it is cyclic with s distinct associated irreducible factors.

Proof.

Necessity follows from Theorem 2.1. To prove sufficiency, let M1,,Ms denote the maximal subalgebras of L. Then i=1sMi cannot be the whole L, since |F|>s. Choose xLi=1sMi. Then x generates L.□

Corollary 2.3.

Let L be a Leibniz algebra over an infinite field. Then L is cyclic if and only if it has only finitely many maximal subalgebras.

3. Dually atomistic Leibniz algebras

We say that a Leibniz algebra L is dually atomistic if every subalgebra of L is an intersection of maximal subalgebras of L. It is easy to see that if L is dually atomistic then so is every factor algebra of L, and if L is dually atomistic then it is ϕ-free. Dually atomistic Lie algebras over a field of characteristic zero were classified in [Citation19]. An extra family arises for Leibniz algebras.

Lemma 3.1.

Let L be dually atomistic and let N be the nilradical of L. Then

  1. MN is an ideal of L for every maximal subalgebra M of L; and

  2. every subspace of N is an ideal of L, and N2=0.

Proof.

(i) The result is clear if NM, so suppose that NM. Then L=N+M and [L,NM]=[N+M,NM]N2+M2N                  Nϕ(L)+MNNM, using Theorem 6.5 of [Citation20] (or Theorem 2.4 of [Citation5]).

(ii) We have N2ϕ(L)=0, so every subspace of N is a subalgebra of L. Let S be any subspace of N. Then S=SN=(MMM)N=MM(MN), where M is the set consisting of all maximal subalgebras of L containing S. Therefore, S is an intersection of ideals of L, by (i), and so is itself an ideal of L.□

Proposition 3.2.

Let L be a solvable Leibniz algebra over any field F. Then L is dually atomistic if and only if one of the following conditions holds:

  1. L is an abelian or almost abelian Lie algebra;

  2. L=Leib(L)Fv, where v2=0 and [e,v]=e for every eLeib(L).

Proof.

Obviously, we can assume that L is not abelian. We first prove necessity. Let N be the nilradical of L. Then N is abelian by Lemma 3.1. As L is ϕ-free, from [Citation5, Theorems 2.4 and 2.6] it follows that L=NV for some subalgebra V of L. As L is solvable, all minimal ideals of L are abelian. Moreover, by Lemma 3.1, every subspace of N is an ideal of L. Therefore, by [Citation5, Theorem 2.4] and [Citation2, Lemma 1.9], we deduce that soc(L)=N=CLr(N). Note that the map θ:LDer(Fn),xRx|Fn is an endomorphism with kernel CLr(Fn) and so CLr(Fn) has codimension at most one in L. Let n1,n2N such that [ni,L]0, i = 1, 2. Since every subspace of N is an ideal of L, all elements of L act by scalar multiplication on N. It follows that CLr(Fn1)=CLr(Fn2)=CLr(N)=N, which allows us to conclude that V is one-dimensional.

Let vV,v0. As [N,v]0, there exists nN such that [n,v]=λn for some λ0. Replacing v by λ1v we can assume that [n,v]=n, so that Rv acts as the identity map on N. Now, if [v,n]=αn, then we have α2n=[v,[v,n]]=[v2,n][[v,n],v]=αn, so α=0,1.

Put I={nN|[v,n]=0},U={nN|[n,v]=[v,n]=n}. Then I and U are ideals of L and N=IU. Clearly, Leib(L)I. But, also, if nI then we have n=(v+n)2Leib(L). Finally, as every subspace of N is an ideal of L, we must have either I = 0 or U = 0, and the assertion follows.

We now prove the converse. If L is an almost abelian Lie algebra, then, by [Citation15, Proposition 1.1(3)], every subspace of L is a subalgebra, so L is obviously dually atomistic. Suppose next that L satisfies condition (ii) of the statement. Note that B + Fv is a maximal subalgebra of L for any maximal subspace B of Leib(L). Let H be a subalgebra of L. Let xH and write x=a+λv, for some aLeib(L) and λF. One has (a+λv)2=λa, hence aH. This proves that either H is contained in Leib(L) or is of the form H=A+Fv for some subalgebra A of Leib(L). In the latter case, we clearly have H=(BM2(B+Fv)), where M2 denote the set of all maximal subspaces of Leib(L) containing A. On the other hand, if HLeib(L), then we have H=Leib(L)(BM1(B+Fv)), where M1 denotes the set of all maximal subspaces of Leib(L) containing H. This completes the proof.□

The combination of Theorem 3.2 with the next result yields a full characterization of dually atomistic Lie algebras over fields of characteristic zero.

Proposition 3.3.

Let L be a dually atomistic Leibniz algebra over a field of characteristic zero. Then L is a three-dimensional non-split simple Lie algebra or is solvable.

Proof.

We have that L/Leib(L) is a solvable or three-dimensional non-split simple Lie algebra by [Citation19, Lemmas 1 and 2]. In the former case, L is solvable, so suppose that the latter case holds. Then, by [Citation3, Theorem 1], we have L=SI, where S is a three-dimensional non-split simple Lie algebra and I= Leib(L). Moreover, every subspace of I is an ideal of L, by Lemma 3.1 and the fact that L is ϕ-free. Let xI. Then [S,x]=0 and, for all s1,s2S,[x,s1]=λx,[x,s2]=μx for some λ,μF, whence [x,[s1,s2]]=[[x,s1],s2][[x,s2],s1]=λμxμλx=0.

It follows that [x,S]=[x,[S,S]]=0 and L=SI. Therefore L is a Lie algebra, and it follows from [Citation19, Lemma 1] that I = 0, as desired.□

4. Upper semi-modular Leibniz algebras

Let L be a Leibniz algebra. A subalgebra U of L is called upper semi-modular if U is a maximal subalgebra of U,B for every subalgebra B of L such that UB is maximal in B. We say that L is upper semi-modular if every subalgebra of L is upper semi-modular in L. The Lie algebras in this class were classified in [Citation15]. There are many more Leibniz algebras in this class.

We first establish the following result, which characterizes Leibniz algebras having a distributive lattice of subalgebras.

Proposition 4.1.

The only non-Lie distributive Leibniz algebras are the two-dimensional cyclic Leibniz algebras.

Proof.

Let L be a distributive Leibniz algebra with Leib(L)0. Then L/Leib(L) must be a distributive Lie algebra and so is one-dimensional. Now, if L has dimension greater than two, then Leib(L) is an abelian Lie subalgebra of dimension greater than one, and so L is not distributive. Hence, L is two-dimensional cyclic, and it has a basis x,x2 with multiplication [x2,x]=0 or [x2,x]=x2. In the former case, the only one-dimensional subalgebra is Fx2, and in the latter case the only one-dimensional subalgebras are Fx2 and F(xx2). In either case, L is distributive.□

The following is a straightforward exercise.

Lemma 4.2.

Let U be a subalgebra of the Leibniz algebra L, and let I be an ideal of L contained in U. Then U is upper semi-modular in L if and only if U/I is upper semi-modular in L/I.

Lemma 4.3.

Let L be a nilpotent cyclic Leibniz algebra. Then L is upper semi-modular.

Proof.

Let L be generated by a where [an,a]=0. Suppose that U is a subalgebra of L which is not in L2. Let 0uU. Then u=a+i=2nλiaiU for some λiF. But then it is easy to see that u,u2,,un are linearly independent, so U = L. Hence, all proper subalgebras of L are inside L2, which is abelian. The result follows.□

Proposition 4.4.

Let L be a non-Lie ϕ-free Leibniz algebra over any field F. Then L is upper semi-modular if and only L=Leib(L)+̇Fv, where Leib(L)=Fe1++Fer and [ei,v]=ei for 1ir, all other products being zero.

Proof.

Suppose first that L is upper semi-modular. Then L=NV where N is the nilradical, is abelian, and V is a subalgebra of L, by Theorems 2.4 and 2.6 of [Citation5]. Let vV,nN. Then [v,v]V Leib(L)VN=0 and Fv covers FvFn=0, so Fn is covered by n,v. It follows that [n,v],[v,n]Fn. As this is true for all vV, every subspace of N is an ideal of L. Then, as in Proposition 3.2, L is abelian or an almost abelian Leibniz algebra. Now suppose that V=UFv. Then Fv covers F(a+u)Fv, where aLeib(L),uU. Hence Fa is covered by a+u,v=Fa+Fu+Fv. It follows that u = 0 and we have the multiplication claimed.

Conversely, suppose that L has the form given. Then the subalgebras of L are of the form B or Fv + B where B is a subalgebra of Leib(L). If B,CLeib(L), they clearly satisfy the upper semi-modular condition. Consider B and Fv + C. Then B(Fv+C)=BC and B,Fv+C=Fv+B+C. If B covers BC, we have B=BC+Fb for some bB and B,Fv+C=Fv+C+Fb which covers Fv + C. If Fv + C covers BC then BC=C, whence CB and B,Fv+C=Fv+B which covers B.

Finally consider Fv + B and Fv + C. Then (Fv+B)(Fv+C)=Fv+BC and Fv+B,Fv+C=Fv+B+C. If Fv + B covers Fv+BC, then we have that B=BC+Fb and Fv+B+C=Fa+C+Fb which covers Fv + C. By symmetry this is sufficient in this case. Hence L is upper semi-modular.□

Corollary 4.5.

Let L be an upper semi-modular Leibniz algebra over an algebraically closed field. Then L is supersolvable.

Proof.

If L is a Lie algebra, then the assertion follows from [Citation15, Corollary 2.2]. On the other hand, if L is non-Lie, then by Proposition 4.4 we have that L/ϕ(L) is supersolvable, and hence so is L, by [Citation4, Theorems 3.9 and 5.2].□

The consideration of the three-dimensional non-split simple Lie algebras shows the hypothesis that the ground field is algebraically closed cannot be dropped in Corollary 4.5.

Lemma 4.6.

Let L be an upper semi-modular nilpotent Leibniz algebra. Then

  1. [Leib(L),x]xLeib(L) for all xL;

  2. [xi,y]Fxi+1++Fxn+1, for all xLeib(L),yL,2in, where xn+1=0,xn0. In fact, Ry|xLeib(L) has matrix (0000α3000αn1αn200αnαn1α30);

  3. xLeib(L) is an ideal of L for all xLeib(L);

  4. Let J={xL|x2=0}. Then J is an abelian ideal of L;

  5. [J,x]xLeib(L) and [x,J]Fxn for all xL, where xn+1=0,xn0.

Proof.

(i) Let yL. If y2x then [y2,x]xLeib(L). Hence suppose that y2x. Then Fy2x=0 is covered by Fy2, so x is covered by y2,x=x+Fy2. Since L is nilpotent, this implies that [y2,x]xLeib(L) and the result follows.

(ii) As L/Leib(L) is a nilpotent upper semi-modular Lie algebra, it follows from Theorem 2.2 of [Citation15] that L2Leib(L). Consequently, we have [x2,y]=[x,[x,y]]+[[x,y],x]=[[x,y],x]Fx2++Fxn+1 by (i). Suppose the result holds for i = k where k1. Then [xk+1,y]=[[xk,x],y]=[xk,[x,y]]+[[xk,y],x]Fxk+1++Fxn+1, which gives the result.

(iii) This is apparent from (ii).

(iv) Let x,yJ. Then FxFy=0 is covered by Fx, so x,y=Fx+Fy and [x,y]=[y,x]=0, whence J is an abelian subalgebra of L. Moreover, by Theorem 2.2 of [Citation15] we have L2Leib(L)J, so J is an ideal of L.

(v) Let yJ. If yx, then [y,x],[x,y]xLeib(L), so suppose that yx. Then xFy=0 is covered by Fy, so x,y=x+Fy. It follows that [y,x],[x,y]xLeib(L) again, which gives the first inclusion. Now 0=[x2,y]=[x,[x,y]]+[[x,y],x]=[[x,y],x], so [x,y]Fxn, yielding the second inclusion.□

Following [Citation16], we say that a Leibniz algebra L is extraspecial if dimZ(L)=1 and L/Z(L) is abelian. Then we have the following result.

Proposition 4.7.

An extraspecial Leibniz algebra L is upper semi-modular if and only if J={xL|x2=0} is an abelian ideal of L.

Proof.

The necessity follows from Lemma 4.6(iv). Conversely, suppose that J is an abelian ideal of L and let U, B be subalgebras of L for which UB is a maximal subalgebra of B. We need to show that U is a maximal subalgebra of U,B. We consider several cases.

Suppose first that UJ. If BJ then the result is clear. Therefore we can suppose that BJ. Then there exists bB such that [b,b]0, which implies that Z(L)B. It follows that B=UB+Z(L) and so U,B=U,Z(L)=U+Z(L) which covers U.

Suppose UJ, so that Z(L)U. Then U(B+Z(L))=UB+Z(L) so UZ(L)B+Z(L)Z(L) is covered by B+Z(L)Z(L).

It follows that UZ(L),B+Z(L)Z(L)=U,BZ(L) covers UZ(L).

Hence U,B covers U, completing the proof.□

5. Lower semi-modular Leibniz algebras

A subalgebra U of L is called lower semi-modular in L if UB is maximal in B for every subalgebra B of L such that U is maximal in U,B. We say that L is lower semi-modular if every subalgebra of L is lower semi-modular in L.

If U, V are subalgebras of L with UV, a J-series (or Jordan–Dedekind series) for (U, V) is a series U=U0U1Ur=V of subalgebras such that Ui is a maximal subalgebra of Ui+1 for 0ir1. This series has length equal to r. We shall call L a J-algebra if, whenever U and V are subalgebras of L with UV, all J-series for (U, V) have the same finite length, d(U, V). Put d(L)=d(0,L). The following is proved in similar manner to [Citation11, Lemma 5].

Proposition 5.1.

For a solvable Leibniz algebra the following are equivalent:

  1. L is lower semi-modular;

  2. L is a J-algebra; and

  3. L is supersolvable.

Proof.

(i)(ii): This is a purely lattice theoretic result (see [Citation7, Theorem V3]).

(ii)(iii): Since L is solvable, all chains from 0 to L will have length dimL, so all maximal subalgebras have codimension one in L. But then L is supersolvable, by [Citation4, Corollary 3.10].

(iii)(i) Let L be a supersolvable Leibniz algebra and let U, B be subalgebras of L such that U is maximal in U,B. Then U has codimension 1 in U,B, so U,B=U+B. But now dim(B/(UB)=dim((U+B)/U)=1, whence UB is maximal in B.□

The following results follow from the corresponding results in [Citation11].

Theorem 5.2.

(cf. [Citation11, Theorem 2]). Let L be a Leibniz algebra over a field F of characteristic 0. Then L is a J-algebra if and only if L=RS where the radical R is supersolvable and S is a semisimple Lie algebra which is a J-algebra. Moreover, if S is lower semi-modular then so is L.

Proof.

Let L be a J-algebra. By Levi’s Theorem [4], L=RS, where R is the radical and S is a semisimple Lie subalgebra of L. Moreover, R is supersolvable and [S,R]+[R,S]Leib(L), by Proposition 5.1. We claim that [R,S]=[S,R]=0.

Let L be a minimal counter-example. Suppose that R is not a minimal ideal of L. Let R/K be a chief factor of L. Then R=K+Fr for some rR, and the minimality implies that [K,S]=[S,K]=0 and [r,S]+[S,r]K. But then [S,r]=[S2,r][S,[S,r]]+[[S,r],S][S,K]+[K,S]=0.

Also, [r,S]=[r,S2][[r,S],S][K,S]=0.

Hence R=Leib(L) is a minimal ideal of L and S is a maximal subalgebra of L. Let M be a maximal subalgebra of S. Then d(L)=d(M)+2. But R + M is a maximal subalgebra of L, so d(L)=d(R+M)+1=d(R)+d(M)+1 since R + M is a J-algebra. It follows that d(R) = 1 so R = Fr is one-dimensional. As in the first paragraph of Proposition 3.2 we have that CLr(R) has codimension at most 1 in L. Suppose by contradiction that CLr(R)L. As RCLr(R), we have that CLr(R)S is an ideal of codimension 1 in S. Then, by Weyl’s Theorem, we have L=CLr(R)I, where I is a one-dimensional ideal of S, which is not possible as S is semisimple. Thus CLr(R)=L, in particular [R,S]=0. But [S,R]=[S,Leib(L)]=0, a contradiction. This establishes the claim.

The converse is as in [Citation11, Theorem 2].□

Corollary 5.3.

(cf. [Citation11, Corollary 1]). Let L be a Leibniz algebra over an algebraically closed field F of characteristic zero. Then the following are equivalent:

  1. L is lower semi-modular;

  2. L is a J-algebra; and

  3. L=RS where the radical R is supersolvable and S=sl2(F) or S = 0.

Theorem 5.4.

(cf. [Citation11, Theorem 3]). Let L be a Leibniz algebra over a field F of characteristic zero. Then L is lower semi-modular if and only if L=RS1Sn where the radical R is supersolvable and the Si are mutually non-isomorphic three-dimensional simple algebras for 1in, and also n1 when (F)F and the Si are indecomposable if (F)F.

Proof.

The necessity follows easily from Proposition 5.1, Theorem 5.2 and [Citation11, Theorem 3].

The converse follows from Theorem 5.2 and the fact that S=S1Sn is lower semi-modular, by [Citation11, Theorem 3].□

6. Modular Leibniz algebras

A subalgebra U of a Leibniz algebra L is called modular in L if the following two conditions hold: (1) U,VW=V,UW for all subalgebras V,WL with VW,(1) (2) U,VW=VW,U for all subalgebras V,WL with UW(2)

In particular, quasi-ideals are modular. We call L modular if every subalgebra of L is modular in L.

In the following result we establish when a cyclic Leibniz algebra is modular:

Proposition 6.1.

A cyclic Leibniz algebra L of dimension n is modular if and only if it is one of the following two types:

  1. nilpotent, so L=a where [ai,a]=ai+1 for 1in1, and all other products are zero; or

  2. solvable with L=a where [ai,a]=ai+1 for 1in1,[an,a]=an, and all other products are zero.

Proof.

Let {a,a2,,an} be a basis for L with [an,a]=α1a++αnan. The Leibniz identity shows that α1=0. Let T be the matrix for Ra with respect to this basis, so that it is the companion matrix for p(x)=xnαnxn1α2x=p1(x)n1ps(x)ns, where the pj are the distinct irreducible factors of p and p1(x)=x. Let L=W1Ws be the associated primary decomposition of L with respect to Ra. Then, as in [Citation6, Theorem 4.1], we have [Wj,Wk]=0 for 2j,ks;[W1,Wj]=0 for 2j;[Wj,W1]Wj for 1js.

(Note that [Citation6] concerns left Leibniz algebras, and there is a slight error in that paper in the first equation given here). Let wW=W2Ws. Then W1,wW=w,W1W=w, and so w is an ideal of L. Thus, L has a basis {x1,,xn1,w1,,wk} where [xi,a]=xi+1 for 1in11,[xn1,a]=0,[wi,a]=λiwi for 1ik. Then λiwi+λjwj=[wi+wj,a]=λ(wi+wj), where 0λF, since wi+wj is an ideal of L. Hence λi=λj=λ.

Let a=i=1n1μixi+i=1kνiwi. Then a2=i=1n11μixi+1+λi=1kνiwi,    an1=μ1xn1+λn11i=1kνiwi,an1+1=λn1i=1kνiwi.

Since L=a we have k = 0 or 1. Replacing a by (1/λ)a we have the multiplication given in (i) or (ii).

In case (i) all of the subalgebras are inside Leib(L), as in Lemma 4.3, and so it is easy to check that they are modular. Then suppose that case (ii) holds. Let U be a subalgebra which is not in Leib(L), and let u=a+i=2nλiaiU. Then u2=a2+i=2n1λiai+1+λnan,    un1=an1+(i=2nλi)an,un=(1+i=2nλi)an. Hence unun1=anan1U. But then un1un2=an1an2+λ2(anan1)U, so an1an2U. Similarly we have that aiai1U for 2in. Thus U has codimension 1 in L and so is a quasi-ideal and hence modular. All other subalgebras of L are inside Leib(L) and modularity is straightforward to check.□

The next result shows that the extraspecial Leibniz algebras described in Proposition 4.7 are indeed modular.

Proposition 6.2.

An extraspecial Leibniz algebra L is modular if and only if J={xL|x2=0} is an abelian ideal of L.

Proof.

If L is modular then it is upper semi-modular, so the necessity follows from Proposition 4.7.

Conversely, suppose that L is extraspecial and J={xL|x2=0} is an abelian ideal of L. It follows from the fact that modular nilpotent Lie algebras are abelian that every subalgebra of L is either inside J or is an ideal of L. The modular identities are then straightforward to check.□

In [Citation23] Towers defined a subspace U of a Leibniz algebra L to be a quasi-ideal if [U,V]+[V,U]U+V for every subspace V of L. Similarly we shall define a subalgebra U of L to be a weak quasi-ideal if [U,V]+[V,U]U+V for every subalgebra V of L. Then we have the following result.

Proposition 6.3.

Let L be a Leibniz algebra over an algebraically closed field. The following conditions are equivalent:

  1. L is modular;

  2. every subalgebra of L is a weak quasi-ideal of L; and

  3. [x,y]x+y for all x,yL.

Proof.

(i) (ii): Let U, V be subalgebras of L. Since L is modular, the intervals [U:U,V] and [UV:V] are isomorphic as lattices. As L is supersolvable (by Corollary 4.5) this implies that dimU,VdimU=dimVdimUV, whence dimU,V=dimU+dimVdimUV=dim(U+V).

It follows that U,V=U+V and U is a weak quasi-ideal of L.

(ii) (i): If we assume (ii) then the two modular identities (1) and (2) are easily checked.

(ii) (iii): If (ii) holds, then [x,y]x,y=x,yx+y.

(iii) (ii): Suppose that (iii) holds and let U, V be subalgebras of L. Then [u,v],[v,u]U+V for all uU,vV, whence U is a weak quasi-ideal of L.□

Remark 6.4.

In the proof of Proposition 6.3, the assumption that the ground field is algebraically closed is only used in the implication “(i) (ii)”. Therefore, the remaining implications remain valid over arbitrary fields.

Corollary 6.5.

Let L be a Leibniz algebra over a field F. Suppose that L=EZ, where Z is a central ideal of L and E is an extraspecial Leibniz algebra such that J={xL|x2=0} is an abelian ideal of L. Then L is modular.

Proof.

Note that every subalgebra of L is either inside J + Z or is an ideal of L. As a consequence, if U and V are subalgebras of L, then one has that [U,V]+[V,U]U+K, and the conclusion follows from Proposition 6.3 and Remark 6.4.□

Notice that the algebras described in Corollary 6.5 include those in which every subalgebra is an ideal, as described in [Citation16]; they are the ones for which J=Leib(L).

For our next result we shall need the following from [Citation6], which we include for the reader’s convenience.

Theorem 6.6.

([Citation6, Theorem 2.5]). Let L be a four-dimensional non-split non-Lie nilpotent Leibniz algebra with dim(L2)=2=dim(Leib(L)) and dim(L3)=0. Then, L is isomorphic to a Leibniz algebra spanned by x1,x2,x3,x4 with the nonzero products given by the following:

  • A14:[x1,x1]=x3,[x1,x2]=x4;

  • A15:[x1,x1]=x3,[x2,x1]=x4;

  • A16:[x1,x2]=x4,[x2,x1]=x3,[x2,x2]=x3;

  • A17:[x1,x1]=x3,[x1,x2]=x4,[x2,x1]=αx4,αC{1,0};

  • A18:[x1,x1]=x3,[x2,x1]=x4,[x1,x2]=αx3,[x2,x2]=x4,αC{1}; or

  • A19:[x1,x1]=x3,[x1,x2]=x3,[x2,x1]=x3+x4,[x2,x2]=x4.

In [Citation6] the authors are assuming that the algebras are defined over the complex field. However, all the basis changes made in the proof of this theorem are valid over any algebraically closed field of characteristic different from 2; the only place where 2 is a problem is in the choice of s in the penultimate line. By non-split they mean that L is not a direct sum of two non-zero ideals.

Proposition 6.7.

Let L be a Leibniz algebra over an algebraically closed field of characteristic different from 2 and suppose that L3=0. Then L is modular if and only if it is of the form given in Corollary 6.5.

Proof.

Suppose first that L is modular. We use induction on the (minimal) number of generators of L as a Leibniz algebra. If L is cyclic, then it is at most two-dimensional and so is clearly of the claimed form. Hence, suppose that L is generated by two elements, x, y, say. If L has dimension 3 or less, then it is clearly of the claimed form; if not, it must have dimension 4.

A four-dimensional nilpotent modular Leibniz algebra must have a basis x,x2,y,y2 with [x,y]=α1x2+α2y2 and [y,x]=β1x2+β2y2, by Proposition 6.3. Then L must be one of the algebras given in Theorem 6.6. None of these are modular, as is shown below A14:[x1,x2]=x4x1+x2=Fx1+Fx2+Fx3;A15:[x2,x1]=x4x1+x2=Fx1+Fx2+Fx3;A16:[x1,x2]=x4x1+x2=Fx1+Fx2+Fx3;A17:[x1,x2]=x4x1+x2=Fx1+Fx2+Fx3;A18:[x1x2,x1]=x3x4x1x2+x1=Fx1+Fx2+Fx3;A19:[x1x2,x1]=x4x1x2>+x1=Fx1+Fx2+Fx3.

That just leaves the split case. If L is a direct sum of a three-dimensional ideal and a one-dimensional ideal, then the latter is in the center and so this is of the claimed form. If it is a direct sum of two two-dimensional ideals, then it is of the form (Fx+Fx2)(Fy+Fy2) and [x+y,xy]=x2y2, which cannot be written as a direct sum of (x+y)2=x2+y2 and (xy)2=x2+y2.

Assume now that the result is true for n2, and suppose that L is generated by n + 1 elements. Then L=L1+Fx where L1 is of the form EC with E being an extraspecial Leibniz algebra and C being a central ideal of L. By considering pairs of generators and using the above we see that L2=(L1)2=Fz, say. Suppose there exists yL1Z(L1) such that x+yZ(L). Then L=L1F(x+y) is of the required form. Similarly, L=L1Fx if xZ(L). If neither, then L is of the form E˜C˜ where, E˜ is an extraspecial Leibniz algebra, C˜ is a central ideal of L, and Z(E˜)=Fz=E˜2, which completes the proof.□

Extending the above result to the case where Ln=0 and n > 3 is far from straightforward. It is easy to see that any Leibniz algebra of the form EC, where E is extraspecial in which J is an abelian ideal, or a nilpotent cyclic algebra and C is a central ideal, is modular. However, not every nilpotent modular Leibniz algebra L has this form, even if L is four-dimensional, as the following example shows:

Example 6.8.

Let L have basis x1,x2,x3,x4 and nonzero products [x1,x1]=x3,[x2,x2]=x4,[x1,x3]=x4 (This is A25 in [Citation6, Theorem 2.6]). Then it is straightforward to check that L is modular but is not of the form given in the previous paragraph.

It would also be good to remove the requirement of an algebraically closed field in Proposition 6.7. However, this is not straightforward either, as modularity is not preserved by extending the base field, as the following example shows.

Example 6.9.

Let L be the extraspecial Leibniz algebra L=Fx+Fy+Fz with [x,y]=z,[y,z]=z,x2=y2=z and z2=0, the other products being zero. By Proposition 6.3 and Remark 6.4, it is easy to see that L is modular over the real number field and non-modular over the complex number field. (For the latter conclusion, note that the elements x + iy and x − iy are in J but their sum is not.)

References