668
Views
1
CrossRef citations to date
0
Altmetric
Research Article

Left 3-Engel elements in locally finite 2-groups

&
Pages 4869-4882 | Received 10 Dec 2020, Accepted 10 May 2021, Published online: 05 Jun 2021

Abstract

We give an infinite family of examples that generalize the construction given by Noce, Tracey and Traustason of a locally finite 2-group G containing a left 3-Engel element x where xG, the normal closure of x in G, is not nilpotent. The construction is based on a family of Lie algebras that are of interest in their own right and make use of a classical theorem of Lucas, regarding when (mn) is even.

2020 Mathematics Subject Classification:

1. Introduction

Let G be a group. An element aG is a left Engel element in G, if for each xG there exists a non-negative integer n(x) such that [[[x,a],a],,a]n(x)=1.

If n(x) is bounded above by n then we say that a is a left n-Engel element in G. Throughout this paper we will assume that, when dealing with commutators or Lie products, these are left normed. Recall that the Hirsch-Plotkin radical of a group G is the subgroup generated by all the normal locally nilpotent subgroups of G and that this is also locally nilpotent. It is straightforward to see that any element of the Hirsch-Plotkin radical HP(G) of G is a left Engel element and the converse is known to be true for some classes of groups, including solvable groups and finite groups (more generally groups satisfying the maximal condition on subgroups) [Citation1, Citation3]. The converse is however not true in general and this is the case even for bounded left Engel elements. In fact whereas one sees readily that a left 2-Engel element is always in the Hirsch-Plotkin radical this is still an open question for left 3-Engel elements. Recently there has been a breakthrough and in [Citation7] it is shown that any left 3-Engel element of odd order is contained in HP(G). From [Citation12] one also knows that in order to generalize this to left 3-Engel elements of any finite order it suffices to deal with elements of order 2.

It was observed by William Burnside [Citation2] that every element in a group of exponent 3 is a left 2-Engel element and so the fact that every left 2-Engel element lies in the Hirsch-Plotkin radical can be seen as the underlying reason why groups of exponent 3 are locally finite. For groups of 2-power exponent there is a close link with left Engel elements. If G is a group of exponent 2n then it is not difficult to see that any element a in G of order 2 is a left (n+1)-Engel element of G (see the introduction of [Citation14] for details). For sufficiently large n we know that the variety of groups of exponent 2n is not locally finite [Citation6, Citation8]. As a result one can see (for example in [Citation14]) that it follows that for sufficiently large n we do not have in general that a left n-Engel element is contained in the Hirsch-Plotkin radical. Using the fact that groups of exponent 4 are locally finite [Citation11], one can also see that if all left 4-Engel elements of a group G of exponent 8 are in HP(G) then G is locally finite.

Swapping the role of a and x in the definition of a left Engel element we get the notion of a right Engel element. Thus an element aG is a right Engel element, if for each xG there exists a non-negative integer n(x) such that [[[a,x],x],,x]n(x)=1.

If n(x) is bounded above by n, we say that a is a right n-Engel element. By a classical result of Heineken [Citation5] one knows that if a is a right n-Engel element in G then a1 is a left (n+1)-Engel element.

In [Citation9] M. Newell proved that if a is a right 3-Engel element in G then aHP(G) and in fact he proved the stronger result that aG is nilpotent of class at most 3. The natural question arises whether the analogous result holds for left 3-Engel elements. In [Citation10] it is shown that this is not the case by giving an example of a locally finite 2-group with a left 3-Engel element a such that aG is not nilpotent. Moreover in [Citation4] an example is given, for each odd prime p, of a locally finite p-group containing a left 3-Engel element x where xG is not nilpotent.

In this paper we extend the example above, of a 2-group, to an infinite family of examples. The construction will be based on a family of Lie algebras that generalize the Lie algebra given in [Citation13]. These algebras are of interest in their own right and will make use of a classical theorem of Lucas. Before stating Lucas’s Theorem we need some notation.

Let p be a prime and consider non-negative integers m and n written in base p m=m0+m1p++mk1pk1+mkpkn=n0+n1p++nk1pp1+nkpk, where 0m0,,mk,n0,,nkp1. We introduce a partial order p, where npm if and only if nimi, for 0ik.

Theorem 1.1

(Lucas’ Theorem). The binomial coefficient (mn) is divisible by p if and only if npm.

Remark.

Notice that when p = 2 we get that the binomial coefficient (mn) is odd if and only if n2m.

2. The Lie algebra L

In this section we construct a family of Lie algebras that extend the example given in [Citation13]. The construction makes an interesting use of Lucas’ Theorem.

Let F be the field of order 2 and let n=2m2, for m2. Consider the (n+2)-dimensional vector space L=Fv(0)+Fv(1)++Fv(n1)+Fw+Fx. We equip L with a binary product where v(i)·v(j)=(j+1ni)v(ij)v(i)·w=w·v(i)=v(i+1),i=0,1,,n2v(n1)·w=w·v(n1)=wv(i)·x=x·v(i)=0w·x=x·w=v(0), such that iji+j(modn1) and ij{0,,n2}, and where ww=xx=0. We then extend the product linearly on L. The next theorem is our first main result.

Theorem 2.1.

L is a Lie algebra over F.

Proof.

Let 0i,j,kn1 and suppose that i+1=a0+2a1++2m1am1j+1=b0+2b1++2m1bm1k+1=c0+2c1++2m1cm1n+1(i+1)=(1a0)+2(1a1)++2m1(1am1)n+1(j+1)=(1b0)+2(1b1)++2m1(1bm1)n+1(k+1)=(1c0)+2(1c1)++2m1(1cm1), where 0ai,bi,ci1 for 0im1. In order to show that the product is alternating, it only remains to see that v(i)·v(i)=0 and v(i)·v(j)=v(j)·v(i) (recall that the characteristic is 2). Firstly v(i)·v(i)=(i+1n+1(i+1))v(ii).

In order for the product to be zero we know by Lucas’ Theorem that we need n+1(i+1)2i+1. Assume for contradiction that (n+1)(i+1)2i+1. Then 1aiai, for all 0im1 which implies that a0=a1==am1=1. This gives i+1=1+2++2m1=n+1 that contradicts in1.

Now, for v(j)·v(i)=v(i)·v(j), we need (i+1nj)=(j+1ni). But, (i+1n+1(j+1)) is odd 1biai, for all 0im11aibi, for all 0im1(j+1n+1(i+1)) is odd.

Having established that the product is alternating we turn to the Jacobi identity . If we have basis elements e, f then, as e·e=0 and e·f=f·e, we get (e·e)·f+(e·f)·e+(f·e)·e=2(e·f)·e=0. Therefore, we only need to deal with the cases when the three basis elements are different. We will divide our analysis into few cases. Notice first, that any Jacobi relation involving one occurrence of x and two v(i)’s, for 0in1, is clearly 0. There is one remaining type of a Jacobi relation that involves x. This is the one for a triple (x,w,v(i)), where 0in1.

First, let 0in2. We have, v(i)wx+wxv(i)+xv(i)w=v(i+1)x+v(0)v(i)=(i+1n)v(i)=0, since i+1n1<n and thus n2i+1. If i=n1, then v(n1)wx+wxv(n1)+xv(n1)w=wx+v(0)v(n1)=v(0)+(nn)v(0)=0.

Let us next consider triples of the type (w,v(i),v(j)) where 0i,jn1.

Case 1:

Let i=n1 and 0jn2. Then, wv(n1)v(j)+v(n1)v(j)w+v(j)wv(n1)=wv(j)+(j+11)v(j)w+v(j+1)v(n1)=v(j+1)+(j+1)v(j+1)+(j+21)v(j+1)=2(j+2)v(j+1)=0.

Case 2:

Let 0i<jn2. Then, wv(i)v(j)+v(i)v(j)w+v(j)wv(i)=v(i+1)v(j)+(j+1ni)v(ij)w+v(j+1)v(i)=(j+1ni1)v(i(j+1))+(j+1ni)v(i(j+1))+v(i)v(j+1)=(j+1ni1)v(i(j+1))+(j+1ni)v(i(j+1))+(j+2ni)v(i(j+1))=2(j+2ni)v(i(j+1))=0, where the last equality follows from Pascal’s Rule.

Finally, consider v(i)v(j)v(k)+v(j)v(k)v(i)+v(k)v(i)v(j)=α1v(ijk)+α2v(ijk)+α3v(ijk), for 0i,j,kn1. Clearly, if all coefficients αi are even then the Jacobi identity holds. So, assume without loss of generality that α1 is odd. Then, as v(i)v(j)0 we get 1ai+bi, for i=0,,m1. Then, (i+1)+(j+1)=(a0+b0)+2(a1+b1)++2m1(am1+bm1)1+2++2m1=n+1, so i+jn1.

Case 1:

Consider the case where i+j=n1. Then, v(i)v(j)v(k)=v(0)v(k)=(k+1n)v(k0). Notice that (k+1n) is odd if and only if k=n1. Hence k=n1 and v(i)v(j)v(k)=v(0(n1))=v(0). Thus, v(i)v(j)v(k)+v(j)v(k)v(i)+v(k)v(i)v(j)=v(0)+(j+1)v(j)v(i)+(i+1)v(i)v(j)=v(0)+(j+1)v(0)+(i+1)v(0)=(n+2)v(0)=2mv(0)=0, as required.

Case 2:

Consider the case where i+jn. We want to show that if α1 is odd then exactly one of α2,α3 is odd. We shall consider each term separately. We have v(i)v(j)v(k)=(j+1n+1(i+1))v(ij)v(k)=(j+1n+1(i+1))(k+1n+1(i+jn+2))v(ijk), since i+jn=(j+1n+1(i+1))(k+12(n+1)1(i+1+j+1))v(ijk).

We assumed that α1 must be odd, hence we need both binomial coefficients to be odd. We have (j+1n+1(i+1)) is odd if and only if 1ai+bi, for all i. Let t be the smallest index such that at+bt=1. Therefore, a0==at1=b0==bt1=1.

In order for (k+12(n+1)1(i+1+j+1)) to be odd we must have that 2(n+1)1(i+1+j+1)2k+1. So, 2(n+1)1(i+1+j+1)=2(1+2+22++2m1)1((a0+b0)+2(a1+b1)++2t++2m1(am1+bm1))=2t1+2t+1(2at+1bt+1)++2m1(2am1bm1)=1+2+22++2t1+2t+1(2at+1bt+1)++2m1(2am1bm1)2c0+2c1++2t1ct1+2tct+2t+1ct+1++2m1cm1, hence, c0=c1==ct1=1 and 2ai+bi+ci, for all it+1. Notice that it follows in particular that 1ai+ci,bi+ci for all i except possibly i = t.

Notice that the assumption 2(n+1)1(i+1+j+1)2k+1 implies that 2(n+1)1(i+1+j+1)k+1. That is 2(n1)i+j+k. Then we must have that j+kn1 and i+kn1. If there is an equality, for example j+k=n1, then, by symmetry, we have already shown in Case 1 that the Jacobi identity holds. So, we may assume that j+k,i+kn. Hence, similarly as above, v(j)v(k)v(i)=(k+1n+1(j+1))(i+12(n+1)1(k+1+j+1))v(ijk) and v(k)v(i)v(j)=(i+1n+1(k+1))(j+12(n+1)1(k+1+i+1))v(ijk).

We have two cases:

  1. ct = 0: Assume without loss of generality that at = 0 and bt = 1. Then, at+ct=01 and bt+ct=1, hence (k+1n+1(j+1)) is odd and (i+1n+1(k+1)) is even. Notice that this implies that the smallest index s such that bs+cs=1 is s = t. Hence, (i+12(n+1)1(j+1+k+1)) is odd, since 2ai+bi+ci, for all it+1 and a0==at1=1. This shows that the Jacobi identity holds.

  2. ct = 1: Assume without loss of generality that at = 0 and bt = 1. Then, both at+ct,bt+ct1, hence both binomials coefficients (k+1n+1(j+1)) and (i+1n+1(k+1)) are odd. Now, similarly as in case (a), if s is the smallest index such that as+cs=1, then s = t and so (j+12(n+1)1(i+1+k+1)) is odd. It only remains to show that (i+12(n+1)1(j+1+k+1)) is even. But, bt+ct=2, so the smallest index l such that bl+cl=1 is lt+1. Then, in order for (i+12(n+1)1(j+1+k+1)) to be odd we require a0=a1==at=1, which contradicts our assumption, hence (i+12(n+1)1(j+1+k+1)) must be even and the Jacobi identity holds.□

Lemma 2.2.

The Lie algebra L has trivial center.

Proof.

Take an element of L say l=λ1x+λ0v(0)++λn1v(n1)+μw, where λ1,λ0,,λn1,μF, that lies in the center of L. Multiplying by x gives μv(0)=0, therefore μ = 0. Then, multiplying by w gives λ1v(0)+λ0v(1)++λn2v(n1)+λn1w=0 and therefore λ1==λn1=0.

Lemma 2.3.

W=Fv(0)++Fv(n1)+Fw is a simple ideal of L.

Proof.

Consider the ideal I generated by y, where y=λ0v(0)++λn1v(n1)+μw and λ0,,λn1,μF are not all zero. We first show that wI. If λ0==λn1=0, this is clear. If not, take the smallest i such that λi0, where 0in1. Taking y and mulitiplying ni times by w gives us λiw that implies that wI. Having established that wI we can multiply it by x,v(0),,v(n2) to see that v(0),,v(n1)I. Hence I = W.□

Let E=ad(x),ad(v(0)),ad(v(1)),,ad(v(n1)),ad(w)End(L). As Z(L) is trivial, E is the associative enveloping algebra of L.

Lemma 2.4.

The associative enveloping algebra E is finite-dimensional.

Proof.

This follows from the fact that dim(L)=n+2, hence dim (End (L)) = (n + 2)2, thus E must be of finite dimension.□

We will use L to construct a locally nilpotent Lie algebra over F of countably infinite dimension. This will then help us to construct a locally finite group G with a left 3-Engel element y where yG is not nilpotent. We now introduce a notation that was used in [Citation13] of modified unions of subsets of N. We let AB={AB, if AB=, otherwise

For each non-empty subset A of N we let WA be a copy of the vector space W=Fv(0)++Fv(n1)+Fw, that is WA={zA:zW} with addition zA+tA=(z+t)A. We then take the direct sum of these W*=ANWA that we turn into a Lie algebra with multiplication zA·tB=(z·t)AB when zAWA and tBWB and extend linearly on W*. The interpretation here is that z=0. Finally, we extend this to the semidirect product with Fx L*=W*Fx induced from the action zA·x=(z·x)A. Notice that L* has basis {x}{v(0)A,,v(n1)A,wA:AN} and that v(i)A·v(i)B=wA·wB=0,v(i)A·x=0,wA·x=v(0)A, for all 0i,jn1 and v(i)A·v(j)B=(j+1ni)v(ij)AB, for all 0i, jn1,v(i)A·wB=v(i+1)AB, for all 0i, jn2and v(n1)A·wB=wAB.

Lemma 2.5.

L* is locally nilpotent.

Proof.

Notice that any finitely generated subalgebra of L* is contained in some S=x,v(0)A10,,v(0)Ar0,,v(n1)A1n1,,v(n1)Atn1,wB1,,wBl.

Thus it suffices to show that S is nilpotent. Observe first that any Lie product with a repeated entry of v(i)A or wB is trivial and thus a non-trivial Lie product of the generators of S can include in total at most r+s++t+l such elements. As v(i)A·x=(wB·x)·x=0 we have (z·x)·x=0 for all zL*. Thus we see that S is nilpotent of class at most 2(r++t+l). Therefore, L* is locally nilpotent.

3. The group G

For an element yL* we denote by ad(y) the linear operator on L* induced by multiplication by y on the right. In this section we find a group G inside GL(L*) containing 1+ad(x), where 1+ad(x) is a left 3-Engel element in G, but where 1+ad(x)G is not nilpotent. The next Lemma is a preparation for this.

Lemma 3.1.

The adjoint linear operator ad(x) on L* satisfies:

  1. ad(x)2=0.

  2. ad(x)ad(y)ad(x)=0, for all yL*.

Proof.

  1. This follows from our earlier observation that (z·x)·x=0 for all zL*.

  2. Follows from wA·x·v(i)B·x=v(0)A·v(i)B·x=0,wA·x·wB·x=v(0)A·wB·x=v(1)AB·x=0.

Now let y be any of the generators x,v(i)A,wA, for 0in1. Since, ad(y)2=0 for all y it follows that (1+ad(y))2=1+2ad(y)+ad(y)2=1.

Thus, 1+ad(y) is an involution in GL(L*). Notice also that the following are elementary abelian 2-groups of countably infinite rank. V0=1+ad(v(0)A):AN,V1=1+ad(v(1)A):AN,Vn1=1+ad(v(n1)A):AN,W=1+ad(wA):AN

We will be working with the group G=1+ad(x),V0,V1,,Vn1,W.

Lemma 3.2.

The following commutator relations hold in G:

  1. [1+ad(wA),1+ad(x)]=1+ad(v(0)A).

  2. [1+ad(v(i)A),1+ad(x)]=1.

  3. [1+ad(v(i)A),1+ad(v(j)B)]=1+(j+1ni)ad(v(ij)AB).

  4. [1+ad(v(i)A),1+ad(wB)]=1+ad(v(i+1)AB), if 0in2.

  5. [1+ad(v(n1)A),1+ad(wB)]=1+ad(wAB).

Proof.

  1. We have [1+ad(wA),1+ad(x)]=(1+ad(wA))·(1+ad(x))·(1+ad(wA))·(1+ad(x))=1+ad(wA)ad(x)+ad(x)ad(wA)+ad(x)ad(wA)ad(x)=1+ad(wA·x)=1+ad(v(0)A),

    where we have used Lemma 3.1. Part (b) is proven similarly. For (c) we have [1+ad(v(i)A),1+ad(v(j)B)]=(1+ad(v(i)A))·(1+ad(v(j)B)·(1+ad(v(i)A))·(1+ad(v(j)B))=1+ad(v(i)A)ad(v(j)B)+ad(v(j)B)ad(v(i)A)=1+ad(v(i)A·v(j)B)=1+(j+1ni)ad(v(ij)AB).

Parts (d) and (e) are proved similarly.

Remark.

Notice that as L* is locally nilpotent it follows from Lemma 3.2 that G is locally nilpotent. Next proposition clarifies the structure of G.

Proposition 3.3.

We have G=1+ad(x)V0 Vn1W. In particular, for every element gG there exists an expression g=(1+ad(x))ϵ·r0  rn1·s, with ϵ{0,1},r0V0,,rn1Vn1 and sW.

Proof.

Suppose that g=g0(1+ad(x))g1 (1+ad(x))gn where g0,,gn are products of elements of the form 1+ad(v(0)A),,1+ad(v(n1)A),1+ad(wA). From Lemma 3.2 we know that (1+ad(wA))(1+ad(x))=(1+ad(x))(1+ad(wA))(1+ad(v(0)A)) and (1+ad(x)) commutes with all products of the form 1+ad(v(i)A), for 0in1. We can thus collect the (1+ad(x))’s to the left starting with the leftmost occurrence. This may introduce more elements of the form 1+ad(v(0)A), but no new elements 1+ad(x). We thus have that g=(1+ad(x))ng1  gm where gi is of the form 1+ad(v(j)A), for 0jn1 or of the form 1+ad(wA). Notice also that we can assume that n=ϵ, where ϵ{0,1}. This reduces our problem to the case when gV0,,Vn1,W. Suppose g=g1g2  gn, where the terms gi are (1+ad(v(0)A1)),,(1+ad(v(0)Ar)),,(1+ad(v(n1)B1)),,(1+ad(v(n1)Bs)),(1+ad(wC1)),,(1+ad(wCt)) in some order. By Lemma 3.2 we have that (1+ad(v(j)B))(1+ad(v(i)A))=(1+ad(v(i)A))(1+ad(v(j)B))(1+(j+1ni)ad(v(ij)AB)) and (1+ad(wB))(1+ad(v(i)A)=(1+ad(v(i)A))(1+ad(wB))(1+ad(v(i+1)AB)), if 0in2 or (1+ad(wB))(1+ad(v(n1)A))=(1+ad(v(n1)A))(1+ad(wB))(1+ad(wAB)), if i=n1. We can thus collect the terms so that g=(1+ad(v(0)A1))(1+ad(v(0)Ar))(1+ad(v(n1)B1))(1+ad(v(n1)Bs))·(1+ad(wC1))  (1+ad(wCt))·h1  hm, where hi are of the form 1+ad(v(i)D), with 0in1, or of the form 1+ad(wD), where D is a modified union of at least two sets from S={A1,,Ar,,B1,,Bs,C1,,Ct}.

Thus, g=a0a1  an1ah, where aiVi, for 0in1,aW and h is a product of elements of the form 1+ad(v(i)D), with 0in1, or of the form 1+ad(wD), where D is a modified union of at least two sets from S.

Repeating this collection process we get g=b0b1  bn1bk, where biVi, for 0in1,bW and k is a product of elements of the form 1+ad(v(i)E), where 0in1, or of the form 1+ad(wE), where E is a modified union of at least three sets from S.

Continuing in this manner we conclude that after k steps g=c0c1  cn1cf, where ciVi, for 0in1,cW and f is a product of elements of the form 1+ad(v(i)H), where 0in1, or of the form 1+ad(wH), where H is a modified union of at least k + 1 sets from S. However, any modified union of at least r++s+t+1 sets from S is trivial and thus f = 1 when k=r++s+t. This completes the proof.

Theorem 3.4.

The element 1+ad(x) is a left 3-Engel element in G such that 1+ad(x)G is not nilpotent.

Proof.

Showing that 1+ad(x) is a left 3-Engel element in G is equivalent to showing that [(1+ad(x))g,1+ad(x)] commutes with 1+ad(x) for all gG. Let g=h(1+ad(wA1)) (1+ad(wAk)) be an arbitrary element in G, where h(1+ad(x))V0  Vn1. We want to show that [(1+ad(x))g,1+ad(x),1+ad(x)]=1.

Let yL*. Then (1+ad(y))1+ad(wA)=(1+ad(wA))(1+ad(y))(1+ad(wA))=1+ad(y)+ad(y·wA).

Notice that (1+ad(x))g=(1+ad(x))(1+ad(wA1)) (1+ad(wAn)), since by Lemma 3.2 we know that 1+ad(x) commutes with all elements of the form 1+ad(v(i)B), for 0in1. Then, by induction we obtain that (1+ad(x))g=1+ad(y) where y=x+1ikv(0)Ai+1i<jkv(1)AiAj++1i(1)<i(2)<<i(n+1)kwAi(1)Ai(n+1).

Since ad(x)ad(y)ad(x)=0, the commutator of (1+ad(x))g with (1+ad(x)) is (1+ad(y))(1+ad(x))(1+ad(y))(1+ad(x))=1+ad(y)ad(x)+ad(x)ad(y)+ad(y)ad(x)ad(y).

Then, [(1+ad(x))g,1+ad(x),1+ad(x)]=[(1+ad(x))(1+ad(y))]4=(1+ad(y)ad(x)+ad(x)ad(y)+ad(y)ad(x)ad(y))2=1, using the fact that ad(x)ad(y)ad(x)=ad(y)2=ad(x)2=0.

However, the normal closure of 1+ad(x) in G is not nilpotent, as for Ai={i} we have [1+ad(wA0),1+ad(x),1+ad(wA1),1+ad(wA2),,1+ad(wAn),,1+ad(x),1+ad(wAmn+1),1+ad(wAmn+2),,1+ad(wA(m+1)n)]=1+ad(wB), where B=A0A1A(m+1)n={0,1,2,,(m+1)n}.

One might wonder if the nilpotency class of the subgroup generated by r conjugates is unbounded for the family we have constructed. That turns out not to be the case. Our next aim is to show that the subgroup generated by any r conjugates (1+ad(x))g1,,(1+ad(x))gr of 1+ad(x) in G will be nilpotent of r-bounded class.

We first work in a more general setting. For each eE and =AN, let e(A)End(L*) where v(i)B·e(A)=(v(i)·e)BAwB·e(A)=(w·e)BAx·e(A)=(x·e)A, for 0in1.

Then let E* = ⟨ad(x),e(A): eE and A⟩ . As L* is locally nilpotent, one sees readily that the elements of E* are nilpotent and thus 1+E* is a subgroup of End(L*).

Despite the fact that the normal closure of 1+ad(x) in G is not nilpotent, it turns out that the nilpotency class of the subgroup generated by any r conjugates grows linearly with respect to r. In order to see this we must first introduce some more notation. In relation with the r conjugates we let A1,A2,,Ar be any r subsets of N. For each r-tuple (i1,i2,,ir) of non-negative integers and each eE we let e(i1,,ir) =B1A1|B1|=i1BkAr|Br|=ire(B1B2Br).

Notice that e(i1,,ir)·f(j1,,jr)=(i1+j1i1)(ir+jrir)(ef)(i1+j1,,ir+jr).

Consider the r conjugates of (1+ad(x)) in G. Recall that each conjugate is of the from (1+ad(x))(1+ad(wC1)(1+ad(wCj)). Without loss of generality one can assume that each Ck is a singleton set. The following argument also works for the more general case. Let A1={1,,k1},A2={k1+1,,k2},,Ar={kr1+1,,kr} and e1=ad(v(0)),e2=ad(v(1)),,en=ad(v(n1)),en+1=ad(w).

Then we have seen (see the proof of Theorem 3.4) that (1+ad(x))(1+ad(w1))(1+ad(wk1))=1+ad(x)+e1(1,0,,0)++en+1(n+1,0,,0)(1+ad(x))(1+ad(wkr1+1))(1+ad(wkr))=1+ad(x)+e1(0,,0,1)++en+1(0,,0,n+1).

Let f(i,k)=ei(0,,0,i,0,,0),

where i is the k-th coordinate and 1in+1. Let F=f(j,k)=ej(0,,0,j,0,,0):1jn+1,1kr.

Our aim is to find an upper bound for the nilpotence class of F. For this we need to understand better the two aspects of multplying e(i1,,ir) and f(j1,,jr). These are

  • Under which conditions the binomial coefficients are non-trivial;

  • Under which conditions is the Lie product ei·ej non trivial.

The next two lemmas will help clarify these questions.

Lemma 3.5.

If i+jn+2, then (i+ji)=0, where 0i,jn+1.

Proof.

Suppose that i=α0+2α1++2m1αm1j=β0+2β1++2m1βm1.

We have that (i+ji) is odd if and only if i2i+j. The latter happens if and only if there exists no l such that αl=βl=1, where 0lm1. In particular, for the binomial coefficient to be non-zero we need i+j1+2+2m1=2m1=n+1.

Lemma 3.6.

If i+jn1 and 1i,jn, then ei·ej=0.

Proof.

As a preparation we first show that v(i)·v(j)=0 if i+jn2. To see this let i+1=a0+2a1++2m1am1j+1=b0+2b1++2m1bm1n+1(i+1)=(1a0)+2(1a1)++2m1(1am1), where 0al,bl1 for 0lm1. Then (j+1n+1(i+1)) is odd1al+bl is odd 0lm1.

In particular for (j+1n+1(i+1)) to be odd we need i+1+j+11+2++2m1=2m1=n+1. That is we need i+jn1. Thus if i+jn2 we have v(i)·v(j)=0. Having established this preliminary result we turn to ei·ej where 1i,jn. Firstly, weiej=wv(i1)v(j1)=v(i)v(j1)=(jn+1(i+1))v(i(j1)) and so it immediately follows from the result above that weiej=0, if i+j1n2, that is if i+jn1. Then, for 0kn1, we have v(k)ei·ej=v(k)v(i1)v(j1).

From our preliminary result above we know that v(k)v(i1)=0, if k+(i1)n2. We can therefore assume that k+(i1)n1 and hence k(i1)=k+(i1)(n1). Therefore, v(k)ei·ej=(in+1(k+1))v(k+(i1)(n1))v(j1)=(in+1(k+1))v(k+in)v(j1) which, by our preliminary result again, is trivial when k+in+j1n2, that is if k+i+j2n1. Given that 0kn1 we have in particular that the product is trivial when i+jn1. Hence, ei·ej=0 if i+jn1.

From these two lemmas we get the following result.

Lemma 3.7.

Let 1i,jn. If f(i,k)·f(j,s) is nonzero, then ni+jn+1.

We are now ready for establishing the linear upper bound for the nilpotence class of F.

Lemma 3.8.

F is nilpotent of a class at most 4r1.

Proof.

Notice that by Lemma 3.5 we have that f(n+1,k)·f(j,s)=0 for any 1jn+1. We thus only need to consider products of elements f(i,k) where 1in. Take any such product of even length 2u where u is going to be determined later. Suppose the product is (1) (f(i1,k1)f(j1,s1))  (f(iu,ku)f(ju,su)),(1) where 1i1,,iu,,j1,,jun and 1k1,,ku,,s1,,sur. We want to determine u so that this product becomes 0. From Lemma 3.7 we know that for this product to be non-trivial we need il+jln, for all 1lu. For 1lr, let tl be the sum of all the lth coordinates of the superfixes of the 2 u elements in the product. For this to be non-zero we need t1++tr=(i1+j1)+  +(iu+ju)nu.

If one of t1,,tr is greater than n + 1 then Lemma 3.5 implies that the product is zero. We need to find how big u has to be so that this happens. Notice that the largest value of t1,,tr is greater or equal to the mean value and this is at least nur. Therefore, it suffices that nurn+2.

This holds when u=2r. Hence, F2(2r)=F4r=0.

From this it is not difficult to obtain a linear upper bound for nilpotence class of the group generated by r conjugates of 1+ad(x). First we extend the analysis of F to the subalgebra Q=ad(x),F of E*. If no element f(n+1,k) occurs then any product of elements of Q of length 4r+1 is trivial. Here we are using the fact that ad(x)2=0, Lemma 3.8 and the fact that ad(x) commutes with elements of the form f(i,l) when 1in. Suppose therefore that at least one element f(n+1,k) is involved in a product of elements of Q. If an element of the form f(i,l) included, where 1in, we can pick f(n+1,k) and f(i,l) that are of closest distance within the product. This reduces things to the following situations: f(n+1,k)ad(x)f(i,l),f(i,l)ad(x)f(n+1,k) ,f(n+1,k)f(i,l) ,f(i,l)f(n+1,k) .

These products are trivial, as ad(x) commutes with f(i,l) and the products of f(n+1,k) and f(i,l) are trivial by Lemma 3.5. We can therefore assume that no element f(i,l) is involved, where 1in. Then we only have a product in ad(x)’s and elements of the form f(n+1,k). By Lemma 3.5 and the fact that ad(x)2=0 the product needs to alternate between ad(x) and elements of the form f(n+1,k). We will make use of the fact that ad(x)f(n+1,k)=f(n+1,k)ad(x)+f(1,k).

We have ad(x)f(n+1,k1)ad(x)f(n+1,k2)=(f(n+1,k1)ad(x)+f(1,k1))ad(x)f(n+1,k2)=f(n+1,k1)ad(x)2f(n+1,k2)+f(1,k1)ad(x)f(n+1,k2)=0+ad(x)f(1,k1)f(n+1,k2)=0, where the last equality follows from Lemma 3.5. Similarly any product of the form f(n+1,k1)ad(x)f(n+1,k2)ad(x) is 0. To conclude we have seen that Q4r+1=0. Now let H be a subgroup of G generated by any r conjugates (1+ad(x))g1,,(1+ad(x))gr of 1+ad(x) in G. Then γ4r+1(H)1+Q4r+1=1.

Thus we have the following result.

Proposition 3.9.

Let (1+ad(x))g1,,(1+ad(x))gr be any r conjugates of 1+ad(x) in G. Then H=(1+ad(x))g1,,(1+ad(x))gr is nilpotent of class at most 4r+1.

Acknowledgments

The first author is partially supported by ‘The Norton Scholarship’. We acknowledge the EPSRC (grant number 16523160) for support. Moreover, we would like to thank Marialaura Noce for drawing our attention to Lucas’ Theorem.

References

  • Baer, R. (1957). Engelsche Elemente Noetherscher Gruppen. Math. Ann. 133(3):256–270. DOI: 10.1007/BF02547953.
  • Burnside, W. (1901). On an unsettled question in the theory of discontinous groups. Q. J. Pure Appl. Math. 37:230–238.
  • Gruenberg, K. W. (1959). The Engel elements of a soluble group. Illinois J. Math. 3(2):151–169. DOI: 10.1215/ijm/1255455117.
  • Hadjievangelou, A., Noce, M., Traustason, G. (2021). Locally finite p-groups with a left 3-Engel element whose normal closure is not nilpotent. Int. J. Algebra Comput. 31(1):135–160.
  • Heineken, H. (1960). Eine Bemerkung über engelshe Elemente. Arch. Math. 11(1):321–321. DOI: 10.1007/BF01236951.
  • Ivanov, S. V. (1994). The free Burnside groups of sufficiently large exponents. Int. J. Algebra Comput. 04(01n02):1–308. DOI: 10.1142/S0218196794000026.
  • Jabara, E., Traustason, G. (2019). Left 3-Engel elements of odd order in groups. Proc. Amer. Math. Soc. 147(5):1921–1927. DOI: 10.1090/proc/14389.
  • Lysenok, I. G. (1996). Infinite Burnside groups of even period. Izv. Math. 60(3):453–654. DOI: 10.1070/IM1996v060n03ABEH000077.
  • Newell, M. (1996). On right-Engel elements of length three. Proc. R. Irish Acad. 96A(1):17–24.
  • Noce, M., Tracey, G., Traustason, G. (2020). A left 3-Engel element whose normal closure is not nilpotent. J. Pure Appl. Algebra 224(3):1092–1101. DOI: 10.1016/j.jpaa.2019.07.004.
  • Sanov, I. N. (1940). Solution of Burnside’s problem for exponent four, Leningrad State Univ. Ann. Maths. Ser. 10:166–170.
  • Tracey, G., Traustason, G. (2018). Left 3-Engel elements in groups of exponent 60. Int. J. Algebra Comput. 28(04):673–695. DOI: 10.1142/S0218196718500303.
  • Traustason, G. (1993). Engel Lie-algebras. Q. J. Math. 44(3):355–384. DOI: 10.1093/qmath/44.3.355.
  • Traustason, G. (2014). Left 3-Engel elements in groups of exponent 5. J. Algebra 414:41–71. DOI: 10.1016/j.jalgebra.2014.05.017.