Skip to content
Publicly Available Published by De Gruyter January 12, 2021

New Lie products for groups and their automorphisms

  • James B. Wilson EMAIL logo
From the journal Journal of Group Theory

Abstract

We generalize the common notion of descending and ascending central series. The descending approach determines a naturally graded Lie ring and the ascending version determines a graded module for this ring. We also link derivations of these rings to the automorphisms of a group. This process uncovers new structure in 4/5 of the approximately 11.8 million groups of size at most 1000 and beyond that point pertains to at least a positive logarithmic proportion of all finite groups.

1 Introduction

Since early work by Magnus, it was known that the lower central series could be used to make a correspondence with free groups and graded Lie rings. This approach became a mainstay of nilpotent groups in part owing to Graham Higman’s promotion in lectures such as β€œLie ring methods in the theory of finite nilpotent groups”, which he delivered to the International Congress of Mathematics in 1958 [7]. Already Lazard and Mal’cev had exploited this and even more subtle connections; see [9, Chapter 3] for an excellent modern survey. Higman’s point was to popularize connections to Lie rings β€œβ€¦β€‰less precise, but much more generally applicable” [7, p. 307]. The spirit of the present article is similar.

This article points out a few ingredients of group/graded Lie ring correspondences that perhaps could have been known at the time but seem largely overlooked in the literature since. First we observe that the upper central series has an appropriate dual role, leading to Lie module upon which the associated graded Lie ring acts. Second we show that in general one can refine gradings, and the upper and lower series of a group. We call descending generalization a filter and the ascending generalization a layering. The result of such refinement is to further refine inductive studies of groups. As it stands, the common concept of nilpotence degree clusters most finite group theory into a single family of groups of nilpotence class 2. Fragmenting this huge family of groups by whatever means helps when studying these groups in detail, for example in enumeration and isomorphism studies [2, 12, 14, 13, 17]. With such applications in mind, we point out how to lift our filters and layerings of a group 𝐺 to Aut⁑(G) and link the associated graded Lie algebra of Aut⁑(G) to the graded derivation algebra of the Lie algebra for 𝐺. This serves in a way to linearize larger parts of the automorphism group of a group.

For simplicity, we first state our results for the traditional Z+ grading which later we relax to gradings by arbitrary commutative monoids. As usual, the relationship between groups and Lie rings is born of simple ingredients relying on the general observations that group commutators [x,y]=x-1⁒xy=x-1⁒y-1⁒x⁒y obey relations that mimic the distributive, alternating, and the Jacobi identities,

[x⁒y,z]=[x,z]y⁒[y,z],[x,x]=1,[[x,y-1],z]y⁒[[y,z-1],x]z⁒[[z,x-1],y]x=1.

Now consider the usual lower central series where Ξ³1=G and Ξ³i+1=[G,Ξ³i]. Thus we may define L*⁒(Ξ³)=L*⁒(γ⁒(G))=βŠ•i>0Ξ³i/Ξ³i-1 and also the following operations on L*⁒(Ξ³) (for x∈γi, set xΒ―=x⁒γi+1):

x¯∘y¯=[x,y]¯,x¯+y¯=x⁒y¯,-x¯=(x-1)¯,0=1¯.

This gives L*⁒(γ) the structure of a graded Lie ring; see [9, Chapter 3]. Perhaps known but certainly underused is the following related product. Consider the usual upper central series where ΢0=1 and ΢i+1 is the centralizer of G/΢i. We observe the upper central series can be used to define graded modules of L*⁒(γ) as follows.

Theorem 1

Given any group 𝐺, a commutative ring 𝑅, and an 𝑅-module 𝑄,

L*⁒(ΞΆ;Q)=βŠ•i>0homZ⁑(ΞΆi/ΞΆi-1,Q)

is a natural graded Lie module of the natural Lie ring L*⁒(Ξ³)βŠ—ZR.

For torsion free groups, the use of R=Q=Q is a natural choice. For finite groups, if we use R=Z and Q=Q/Z, then hom⁑(ΞΆi/ΞΆi-1,Q/Z)β‰…ΞΆi/ΞΆi-1, and for 𝑝-groups, we might prefer Q=Zpe or Zp∞. So we prove the following.

Corollary 1.1

For finite groups, the upper central series factors Li⁒(ΞΆ)=ΞΆi/ΞΆi-1 form a graded Lie module L*⁒(ΞΆ)=βŠ•i>0Li⁒(ΞΆ) for the graded Lie ring

L*⁒(Ξ³)=βŠ•i>0Li⁒(Ξ³)

of the lower central series factors Li⁒(γ)=γi/γi+1.

We will make both these results far more general below. Our further connection between central factors and Lie algebras relates to the structure of automorphism groups. In general, we expect no relationship to exist between the central series of a group and its automorphism group (consider G=Zpn and Aut⁑(G)β‰…GL⁑(d,p)). However, we argue the more appropriate Lie algebra to explore in connection to Aut⁑(G) is the graded derivation ring Der⁑(L⁒(Ξ³)). A hint of this might be recognized from the family of automorphisms πœ™ on 𝐺 that are the identity on G/ΞΆ1. Such β€œcentral” automorphisms induce a linear map G/Ξ³2β†’ΞΆ1 by Ξ³2⁒x↦x-1⁒(x⁒ϕ). The following theorem can be seen as generalizing central automorphisms but retaining the vital Lie structure. Stated just for the traditional central factors, for any groups 𝐺 and 𝐴, and a homomorphism f:Aβ†’Aut⁑(G), define

Ai={a∈A∣for all⁒j⁒and all⁒g∈γj⁒(G),g-1⁒ga∈γi+j⁒(G)},i>0.
Theorem 2

Let 𝐺 and 𝐴 be groups and f:Aβ†’Aut⁑(G) a group homomorphism. Then, for all i,j>0, [Ai,Aj]≀Ai+j and L*⁒(A)=βŠ•i>0Ai/Ai+1 is a graded Lie ring with a canonical graded Lie homomorphism into the graded derivation ring Der⁑(L*⁒(Ξ³)). Also, A/A1 is the faithful representation of 𝐴 acting on L*⁒(Ξ³) by conjugation.

Both L⁒(γ) and Der⁑(L⁒(γ)) are objects whose structure we can capture through linear algebra and related efficient algorithms as in [3]. Meanwhile, it is difficult to construct a generating set for Aut⁑(G) in most practical settings. So, in this light, Theorem 2 offers a practical approach to learn of some of the structure of Aut⁑(G) without knowing generators for Aut⁑(G).

Theorem 1 and especially its corollary appear classical to us, though we have not discovered them in the literature so far. So we give a synopsis. The mechanics came from a largely unrelated setting. We imitate Whitney’s cap-product from algebraic topology; cf. [6, Section 3.3]. The cap-product is a mixture of homology and cohomology groups. Likewise, commutation products involving the upper central series factors mix the lower and upper central series factors as follows:

∘:Li(Ξ³)Γ—Li+j(ΞΆ)↣Lj(ΞΆ).

(Here and throughout, we use ↣ to distinguish multilinear maps from linear and other functions.) To make a module, we β€œshuffle” this product. For example, we create

∘:Li(Ξ³)Γ—hom(Lj(ΞΆ),Q/Z)↣hom(Li+j(ΞΆ),Q/Z).

When the Li⁒(ΞΆ) are finite, hom⁑(Li⁒(ΞΆ),Q/Z)β‰…Li⁒(ΞΆ) leading to Corollary 1.1, but these isomorphisms are not canonical, so it is more natural to retain the coefficients.

2 Generalized ascending and descending central factors

We aim for a generalization of Theorems 1 and 2 that introduces Lie products for groups that are graded by monoids other than the positive integers. The purpose is to find new structure.

By permitting acyclic gradings, we can first profit from all the known Lie methods with standard gradings, and then add recursive tools to refine the rings and modules by increasing the grading parameter.

Throughout, we fix M=⟨M,+,0,β‰ΊβŸ© a pre-ordered commutative monoid which in particular requires that, for all s∈M, 0β‰Ίs, and for all s,t,u,v∈M, if sβ‰Ίt and uβ‰Ίv, then s+uβ‰Ίt+v. In general, we think of M=Nn with either the point-wise or the lexicographic pre-order. In a group 𝐺, commutators are [g,h]=g-1⁒gh=g-1⁒h-1⁒g⁒h, and for X,YβŠ†G,

[X,Y]=⟨[x,y]:x∈X,y∈Y⟩.

We liberally assume the usual commutator calculus such as in [16, Chapter 5].

Following [17, Section 3], a filterΟ•:Mβ†’2G on a group 𝐺 is a function on a pre-ordered commutative monoid M=⟨M,+,0,β‰ΊβŸ© into the subsets 2G of 𝐺 with the following properties. Each Ο•s is a subgroup, and for all 𝑠 and 𝑑 in 𝑀,

[Ο•s,Ο•t]≀ϕs+t and sβ‰ΊtβŸΉΟ•sβ‰₯Ο•t.

Thus the burden of explaining the structure of commutation is shared between the operation and the ordering on the monoid. For instance, it is possible to give a total order, i.e. a series of subgroups, but use indices that come from a non-cyclic monoid so that the commutation between the series terms can be far more intricate than by in usual central series.

To every filter Ο•:Mβ†’2G, there is an associated boundary filter βˆ‚β‘Ο•:Mβ†’2G given by βˆ‚sΟ•=βŸ¨Ο•s+t:tβ‰ 0⟩. Fix a commutative unital associative ring 𝑅, and define the following 𝑅-modules:

L0⁒(Ο•;R)=0,sβ‰ 0⟹Ls⁒(Ο•;R)=(Ο•s/βˆ‚s⁑ϕ)βŠ—ZR.

Every filter determines the following 𝑀-graded Lie ring [17, Theorem 3.1]:

L*⁒(Ο•;R)=βŠ•s∈MLs⁒(Ο•;R),(xΒ―sβŠ—r)∘(yΒ―tβŠ—s)=[x,y]Β―s+tβŠ—r⁒s.

Here, xΒ―s denotes a coset in Ls, etc. Set

L*⁒(Ο•)=L*⁒(Ο•;Z).

Actually, each Ls⁒(Ο•;R) is a natural R⁒[Ο•0/βˆ‚0⁑ϕ]-module so the portion at the top of a filter can be regarded as storing operators.

In a dual manner, a layeringΟ€:Mβ†’2G on a group 𝐺 is a function where each Ο€s is a subgroup, and for all 𝑠 and 𝑑 in 𝑀,

(2.1)[Ο€s,β‹‚uβ‰ 0Ο€t+u]≀πt,sβ‰ΊtβŸΉΟ€s≀πt.

Form the boundary βˆ‚β‘Ο€:Mβ†’2G in the reversed order as βˆ‚s⁑π=β‹‚tβ‰ 0Ο€s+t. So, if M=N, then βˆ‚i⁑π=Ο€i+1. If G=Ο€M=βŸ¨Ο€s:s∈M⟩, then (2.1) says simply that [G,Ο€i+1]≀πi for every 𝑖. This will be recognized as Hall’s condition for a series to be called β€œcentral”. Using general monoids, which may not be well-ordered, it is no longer possible to simply take successors to move up; hence βˆ‚s⁑π is used instead. Also, the flexibility to adjust Ο€0 and Ο€M makes it possible to define such series in settings that are non-nilpotent but still related, e.g. in automorphism groups of nilpotent groups.

These definitions determine abelian groups as follows (proved in Section 3):

L0⁒(Ο€)=0,sβ‰ 0⟹Ls⁒(Ο€)=βˆ‚s⁑π/Ο€s.

To state the main result, we need a version with coefficients. For that, we fix an 𝑅-module 𝑄 and set

Ls⁒(Ο€;Q)=homZ⁑(Ls⁒(Ο€),Q).

While filters give natural ring structures, layerings produce natural modules for these rings. A ring is always better understood with a module.

Theorem 3

For every filter ϕ:M→2G and layering π:M→2G, if

(2.2)(for alls,t∈M) [Ο•s,Ο€s+t]≀πt,

then L*⁒(Ο€;Q) is a graded module of the 𝑀-graded Lie ring L*⁒(Ο•;R).

Observe that, in the case where Ο•=Ξ³ and Ο€=ΞΆ, we know that, for all i,j>0, [Ξ³i,ΞΆi+j]≀΢j. So Theorem 1 is a consequence of Theorem 3.

The next important thrust is to push the effects of a filter into the automorphism group and thus linearize various properties of automorphisms that once required more difficult inspection. Fix groups 𝐺 and 𝐴, and a group homomorphism Aβ†’Aut⁑(G). For x∈G and a∈A, we define [x,a]=x-1⁒xa. For a filter Ο•:Mβ†’(2G)A (that is a filter of 𝐴-invariant subgroups), define the function Δ⁒ϕ:Mβ†’2A given by

Ξ”s⁒ϕ={a∈A:for all⁒t∈M,[Ο•t,a]≀ϕt+s}.
Theorem 4

For groups 𝐺 and 𝐴, a group homomorphism f:Gβ†’A, and a filter Ο•:Mβ†’(2G)A, Δ⁒ϕ:Mβ†’2A is a filter, and there is a natural graded Lie ring homomorphism from L*⁒(Δ⁒ϕ) into the derivation Lie ring Der⁑(L*⁒(Ο•)).

Note Theorem 2 follows from Theorem 4.

3 The layers Lt⁒(Ο€;QR) and the action by Lt⁒(Ο•;R)

Since filters were covered in detail along with examples in [17], we can forgo most of that and focus on the new aspects which concern the interplay with layerings. We begin by confirming the claims that Ls⁒(Ο€;R) is well-defined.

Fix a layering Ο€:Mβ†’2G. The universe being described here is between the subgroup Ο€0 and Ο€M=βŸ¨Ο€s:s∈M⟩. For every s,u, sβ‰Ίs+u, which implies that Ο€sβ‰€βˆ‚s⁑π=β‹‚uβ‰ 0Ο€s+u. Hence, for all 𝑠,

[Ο€M,Ο€s]≀[Ο€M,βˆ‚s⁑π]β‰€βˆt∈M[Ο€t,βˆ‚s⁑π]≀πs.

Thus Ο€s and βˆ‚s⁑π are normal in Ο€M and Ls⁒(Ο€)=βˆ‚s⁑π/Ο€s is central in Ο€M/βˆ‚s⁑π. In particular, Ls⁒(Ο€) is abelian. Also notice that βˆ‚β‘Ο€:Mβ†’2G, and as πœ‹ is order preserving so too is βˆ‚β‘Ο€. Finally,

[Ο€M,βˆ‚s⁑(βˆ‚β‘Ο€)]=β‹‚tβ‰ 0[Ο€M,βˆ‚s+t⁑π]=β‹‚t,uβ‰ 0[Ο€M,Ο€s+t+u]≀⋂tβ‰ 0Ο€s+t=βˆ‚s⁑π.

In particular, βˆ‚β‘Ο€ is a layering as well.

Unlike the case of filters, there is no immediate commutation product to impose on L*⁒(Ο€). Nevertheless, we recognize within these ingredients the mechanics similar to the cap-product in algebraic topology; cf. [6, Section 3.3]. We make up a product of a similar shape.

Straining our metaphors, if (2.2) holds, we say that a filter ϕ:M→2Gsifts a layering π:M→2G.

Proposition 3.1

If a filter Ο•:Mβ†’2G sifts a layering Ο€:Mβ†’2G, then for sβ‰ 0, ∘:Ls(Ο•)Γ—Ls+t(Ο€)↣Lt(Ο€) defined by

x¯∘y¯≑[x,y](modΟ€t)

is well-defined and biadditive.

Proof

Under the assumptions, πœ™ also sifts βˆ‚β‘Ο€ and βˆ‚β‘Ο• sifts πœ‹:

(3.2)[Ο•s,βˆ‚s+t⁑π]=β‹‚uβ‰ 0[Ο•s,Ο€s+t+u]≀⋂uβ‰ 0Ο€t+u=βˆ‚t⁑π,
[βˆ‚s⁑ϕ,Ο€s+t]=⟨[Ο•s+u,Ο€s+t]:uβ‰ 0βŸ©β‰€βŸ¨[Ο•s+u,Ο€s+t+u]:uβ‰ 0βŸ©β‰€Ο€s.
This means that the commutation map Ο•sΓ—βˆ‚s+tβ‘Ο€β†’βˆ‚t⁑ϕ factors through

∘:Ls(Ο•)Γ—Ls+t(Ο€)β†’Lt(Ο•).

Now we prove biadditivity of ∘. For all x,yβˆˆΟ•s and all zβˆˆβˆ‚s+t⁑π,

[[x,z],y]∈[[Ο•s,βˆ‚s+t⁑π],Ο•s]≀[βˆ‚t⁑π,Ο•s].

As sβ‰Ίs+t, [Ο•s,βˆ‚t⁑π]≀[Ο•s,βˆ‚s+t⁑π]≀πt. Hence,

(xΒ―+yΒ―)∘z¯≑[x,z]⁒[[x,z],y]⁒[y,z]≑x¯∘zΒ―+y¯∘zΒ―(modΟ€t).

Also, for xβˆˆΟ•s, y,zβˆˆβˆ‚s+t⁑π,

[[x,y],z]∈[[Ο•s,βˆ‚s+t⁑π],βˆ‚s+t⁑π]≀[βˆ‚t⁑π,Ο€M]≀ϕt.

So

x¯∘(yΒ―+zΒ―)≑[x,z]⁒[x,y]⁒[[x,y],z]≑x¯∘yΒ―+x¯∘yΒ―(modΟ€t).∎

There is of course a similar map Ls+t⁒(Ο€)Γ—Ls⁒(Ο•)↣Lt⁒(Ο€), but this is nothing more than the negative transpose of the map above since [x,y]-1=[y,x]. A more pressing concern is the awkwardness of the indices in the product – it is not graded but nearly. If our monoid has cancellation, we could pass to the associated Grothendieck group and rewrite the grading as LsΓ—Lt↣Lt-s, but still we would need to clarify at each stage that t-s∈M in order that the product be defined. We promote a remedy that makes the resulting algebra more natural at the cost of introducing coefficients.

Suppose we have an arbitrary bimap (biadditive map) *:UΓ—V↣W and an abelian group 𝑄. We create a second bimap #:UΓ—homZ⁑(W,Q)↣homZ⁑(V,Q) with the following definition:

(for allu∈Uand allf∈homZ(W,Q)) (u#f):v↦(u*v)f.

Such shuffling of products was introduced for nonassociative division rings (semifields) by Knuth and reformulated in a basis free way by Liebler [10, 11]. Though Knuth called these β€œtransposes”, we use the less overloaded term of β€œshuffle” as suggested by Uriya First [5]. Here, we use them simply to configure the objects above into familiar vocabulary of Lie modules.

Now let Q=QR be an 𝑅-module for a commutative associative unital ring 𝑅. As defined, Ls⁒(Ο€;Q)=homZ⁑(Ls⁒(Ο€),Q). When we apply a Knuth–Liebler shuffle to the products ∘:Ls(Ο•)Γ—Ls+t(Ο€)↣Lt(Ο€), we arrive at a new product

β‹…:Ls(Ο•;R)Γ—Lt(Ο€;Q)↣Ls+t(Ο€;Q),

where, for x¯∈Ls⁒(Ο•), r∈R, f:Lt⁒(Ο€)β†’Q∈Lt⁒(Ο€;Q), (xΒ―βŠ—r)β‹…f is the following map Ls+t⁒(Ο€)β†’Q:

(3.3)(xΒ―βŠ—r)β‹…f:y¯↦-(x¯∘yΒ―)⁒f⁒r.

This new product is 𝑅-bilinear and further permits us to take the direct sum over all s,t∈M to create a single graded 𝑅-bimap L*⁒(Ο•;R)Γ—L*⁒(Ο€;Q)↣L*⁒(Ο€;Q). It remains to show this makes L*⁒(Ο€;Q) into a Lie module for L*⁒(Ο•;R).

Proposition 3.4

For all x¯∈Ls⁒(Ο•), all y¯∈Lt⁒(Ο•), and all z¯∈Ls+t+u⁒(Ο€),

[xΒ―,yΒ―]∘z¯≑x¯∘(y¯∘zΒ―)-y¯∘(x¯∘zΒ―)(modΟ€u).

Proof

From the Hall–Witt identity, it follows that, for all x,y,z,

[[x,y],z]=[y-1,[x-1,z]-1]⁒[[y-1,[x-1,z]-1],x⁒y][x,[y-1,z-1]]⁒[[x,[y-1,z-1]],z⁒y].

Fix xβˆˆΟ•s, yβˆˆΟ•t, zβˆˆβˆ‚s+t+u⁑π. By (2.2) and (3.2), we note the following containments:

[[y-1,[x-1,z]-1],x⁒y]∈[[Ο•t,[Ο•s,βˆ‚s+t+uΟ€],Ο•sΟ•t]≀[[Ο•t,βˆ‚t+u⁑π],Ο•0]≀[Ο€u+0,Ο•0]≀πu,
[[x,[y-1,z-1]],z⁒y]∈[[Ο•s,[Ο•t,βˆ‚s+t+u⁑π]],βˆ‚s+t+u⁑π⁒ϕ0]≀[Ο€u,Ο€M]⁒[Ο€u,Ο•0]≀πu.
Hence, in the original product ∘, we find the following identity:

[xΒ―,yΒ―]∘z¯≑[[x,y],z](modΟ€u)≑[y-1,[x-1,z]-1]⁒[[y-1,[x-1,z]-1],x⁒y][x,[y-1,z-1]]⁒[[x,[y-1,z-1]],z⁒y]≑[y-1,[x-1,z]-1]⁒[x,[y-1,z-1]]≑x¯∘(y¯∘zΒ―)-y¯∘(x¯∘zΒ―).∎

4 Proof of Theorem 3

To complete the proof of Theorem 3, we must now consider Proposition 3.4 in the presence of the appropriate Knuth–Liebler shuffle.

Fix x¯∈Ls⁒(Ο•), y¯∈Lt⁒(Ο•), r,s∈R, and f:Lu⁒(Ο€)β†’Q∈Lu⁒(Ο€;Q). Recall 𝑅 is commutative and each Β· is 𝑅-bilinear. So, for all z¯∈Ls+t+u⁒(Ο€), following (3.3), we find the following holds:

(zΒ―)⁒((xΒ―βŠ—r)∘(yΒ―βŠ—s)β‹…f)=-((x¯∘yΒ―)∘zΒ―)⁒f⁒r⁒s=-(x¯∘(y¯∘zΒ―))⁒(f⁒s)⁒r+(y¯∘(x¯∘zΒ―))⁒(f⁒r)⁒s=(y¯∘zΒ―)⁒((xΒ―βŠ—r)β‹…(f⁒s))-(x¯∘zΒ―)⁒(yΒ―βŠ—sβ‹…(f⁒r))=-(zΒ―)⁒(yΒ―βŠ—s)β‹…((xΒ―βŠ—r)β‹…f)+(zΒ―)⁒(xΒ―βŠ—r)β‹…(yΒ―βŠ—sβ‹…f).

Hence, we have

((xΒ―βŠ—r)∘(yΒ―βŠ—s))β‹…f=(xΒ―βŠ—r)β‹…((yΒ―βŠ—s)β‹…f)-(yΒ―βŠ—s)β‹…((xΒ―βŠ—r)β‹…f).

So in L*⁒(Ο€;Q) is a graded Lie module for L*⁒(Ο•;R).∎

5 Proof of Theorem 4

Now we prove our second claim, Theorem 4, which claims that, for an 𝐴-group 𝐺 with an 𝐴-invariant filter Ο•:Mβ†’2G, there is a naturally induced filter

Δ⁒ϕ:Mβ†’2A

as follows:

Ξ”s⁒ϕ={a∈A:for all⁒t∈M,[Ο•t,a]≀ϕt+s}.

It further follows that L*⁒(Δ⁒ϕ) maps naturally into Der⁑(L*⁒(Ο•)).

Fix s∈M. Let a,bβˆˆΞ”s⁒ϕ. So, for all t∈M and all xβˆˆΟ•t,

[x,a-1⁒b]=[x-1,a]-b⁒[x,b]∈[Ο•t,a]b⁒[Ο•t,b]≀ϕt+s

(using also the fact that πœ™ is 𝐴-invariant). So Ξ”s⁒ϕ is a subgroup of 𝐴. Indeed, if a∈A, then, as Ο•t is 𝐴-invariant, [Ο•t,a]≀ϕt=Ο•t+0. So aβˆˆΞ”0⁒ϕ proving that A=Ξ”0⁒ϕ.

Now suppose s,t∈M. For each u∈M, aβˆˆΞ”s⁒ϕ, bβˆˆΞ”t⁒ϕ, xβˆˆΟ•u, we find

[x,[a,b]]=[b,[x-1,a-1]]-x⁒a⁒[a-1,[b-1,x]]-b⁒a=[[x-1,a-1],b]x⁒a⁒[[b-1,x],a-1]b⁒a∈[Ο•s+u,b]⁒[Ο•t+u,a]≀ϕs+t+u.

So [Ο•u,[Ξ”s⁒ϕ,Ξ”t⁒ϕ]]≀ϕs+t+u. Hence,

[Ξ”s⁒ϕ,Ξ”t⁒ϕ]≀Δs+t⁒ϕ.

Finally, suppose that s,t∈M and that sβ‰Ίt. Take aβˆˆΞ”t⁒ϕ. For each u∈M, s+uβ‰Ίt+u, and so Ο•s+uβ‰₯Ο•t+u. Hence, [Ο•u,a]≀ϕt+u≀ϕs+u, which proves aβˆˆΞ”s⁒ϕ. So Ξ”s⁒ϕβ‰₯Ξ”t⁒ϕ. This proves Δ⁒ϕ is a filter.

Next we show that, for each s∈M-0, Ξ”s⁒ϕ acts on each Lu⁒(Ο•)=Ο•u/βˆ‚u⁑ϕ as the identity. For taking aβˆˆΞ”s⁒ϕ, [Ο•u,a]≀ϕu+sβ‰€βˆ‚u⁑ϕ. Now define an action of Ξ”s⁒ϕ on L*⁒(Ο•) so that, for each u∈M-0 and aβˆˆΞ”s⁒ϕ, x¯⁒Da=[x,a]Β― mapping Lu⁒(Ο•)β†’Ls+u⁒(Ο•). First we observe that

[Ο•u,Ξ”s⁒ϕ]βˆˆΟ•u+s and [βˆ‚u⁑ϕ,Ξ”s⁒ϕ]β‰€βˆ‚s+u⁑ϕ

so that [y⁒x,a]=[y,a]x⁒[x,a]≑[x,a](modβˆ‚s+u⁑ϕ). So, for each 𝑒, we have Da:Lu⁒(Ο•)↦Lu+s⁒(Ο•). Next take x,yβˆˆΟ•u. As [x,a]βˆˆΟ•s+u, it follows that [[x,a],y]βˆˆΟ•s+2⁒uβ‰€βˆ‚s+u⁑ϕ. Hence,

(xΒ―+yΒ―)⁒Da≑[x⁒y,a]≑[x,a]y⁒[y,a]≑[x,a]⁒[[x,a],y]⁒[y,a]≑[x,a]⁒[y,a]

modulo βˆ‚s+u⁑ϕ. Therefore, Da is an additive endomorphism of L*⁒(Ο•).

Last we show that Da is derivation of L*⁒(Ο•). For u∈M-0 and x,yβˆˆΟ•u,

(x¯∘yΒ―)⁒Da≑[[x,y],a]≑[[x-1,a]-1,y-1]-x⁒y⁒[x,[y-1,a-1]]a⁒y(modβˆ‚s+u⁑ϕ)≑[[x-1,a]-1,y-1]-1⁒[x,[y-1,a-1]]≑x¯⁒Da∘yΒ―+x¯∘y¯⁒Da.

So Da is a derivation on L*⁒(Ο•) shifting the grading by 𝑠.∎

6 Examples

We close with some examples. In particular, we demonstrate how filters and layerings arise naturally within arbitrary groups and are not confined solely to nilpotent contexts.

Consider first the case when G=HΓ—K. Evidently,

Ξ³i⁒(HΓ—K)=Ξ³i⁒(H)Γ—Ξ³i⁒(K) and ΢j⁒(HΓ—K)=ΞΆj⁒(H)Γ—ΞΆj⁒(K).

Using an acyclic monoid, we get at the structure of 𝐻 and 𝐾 independently. For example, use M=N2 and

Ο•(a,b)=Ξ³a⁒(H)Γ—Ξ³b⁒(K),Ο€(a,b)=ΞΆa⁒(H)Γ—ΞΆb⁒(K).

It follows that πœ™ is a filter sifting the layering πœ‹. More generally, given groups β„‹ and each H∈H has a filter Ο•H:MHβ†’2H sifting a layering Ο€H:Mβ†’2H, we obtain a filter Ο•:∏H∈HMHβ†’2∏H sifting a layering Ο€:∏H∈HMHβ†’2∏H, where

Ο•m=∏H∈HΟ•m⁒(H)H,Ο€m=∏H∈HΟ€Hm⁒(H).

We call this a product filter and respectively a product layering.

6.1 Classical and semi-classical filters

Commutation products that are graded by cyclic semigroups we call classical. Suppose that 𝐺 is an arbitrary finite or pro-finite group. In particular, such groups have Sylow 𝑝-subgroups and also 𝑝-cores Op⁒(G), i.e. the intersection of all Sylow 𝑝-subgroups of 𝐺. Given a set 𝒫 of primes, the 𝒫-core OP⁒(G) is a product of the 𝑝-cores for p∈P. In the case where 𝒫 consists of all the primes, we have the usual Fitting subgroups. In any case, we can create a product filter or layering. A natural example of this is to use the exponent-𝑝 lower central series of each Op⁒(G). As a result in that example, the homogeneous components of the associated Lie products are vector spaces over Zp for the various p∈P. If it is more important that the filters remain a series, we can apply an order to the primes and then use the lexicographic order on NP. A cyclic subfilter of a highly related form was introduced by Eick–Horn called the 𝐹-series [4, Section 2].

The filters so described depend essentially on words, the commutator, 𝑝-th powers, etc. and consist of verbal and marginal subgroups. While they need not have cyclic gradings, they are all characterized by passing through the Fitting subgroup (or begin at the nilpotent residual in the case of layerings) and are then built from classical filters of the primary components of a residually nilpotent group. All such filters we refer to as semi-classical, and they serve as the natural starting point for later refinements which we explore next.

6.2 Some non-semi-classical filters

Now we turn our attention to non-semi-classical filters and explain more carefully what we are reporting in Figure 1. We know of five easily computed non-classical commutation products, three introduced in [17] and two here.

In [17, Section 4], three rings were introduced as tools to create filers (and now also layerings). There are two further rings more compatible with decomposing the composition products of layerings so we now include those as well.

Suppose we start with a known filter πœ™ sifting a layering πœ‹. We therefore have various bimaps ∘:UΓ—V↣W from the products between the homogeneous components of the Lie ring L*⁒(Ο•;R) and its module L*⁒(Ο€;Q). For any bimap, we can define the following nonassociative rings.

First we have a Lie ring of derivations

Der(∘)={(f,g;h)∈g⁒l⁒(U)Γ—g⁒l⁒(V)Γ—g⁒l⁒(W):(uf)∘v+u∘(vg)=(u∘v)h}.

Next we have three unital associative rings of left, mid, and right scalars. (Here, we use End(U)βˆ™ to denote the opposite ring, equivalently endomorphisms acting on the left.)

L(∘)={(f,g)∈End(U)βˆ™Γ—End(W)βˆ™:for alluandv,(fu)∘v=g(u∘v)},
M(∘)={(f,g)∈End(U)Γ—End(V)βˆ™:for alluandv,(uf)∘v=u∘(gv)},
R⁒(∘)={(f,g)∈End⁑(V)Γ—End⁑(W):for all⁒u⁒and⁒v,u∘(v⁒f)=(u∘v)⁒g}.
Finally, we have an associative commutative unital ring called the centroid:

Cen(∘)=Z({(f,g;h)∈End⁑(U)Γ—End⁑(V)Γ—End⁑(W):(uf)∘v=(u∘v)h=u∘(vg)}).

Here, 𝑍 means to take the center of the ring. Under very mild assumptions on ∘, the conditions on (f,g;h) already force commutativity. There is a valuable interplay between these rings and automorphisms explained in [18].

In [17], M⁒(∘) was introduced as the ring of adjoints and denoted Adj⁑(∘). The rings L⁒(∘) and R⁒(∘) did not appear. In any case, one recognizes that these rings are each determined by systems of linear equations. Thanks to algorithms of Ronyai, Giani–Miller–Traeger, Ivanyos, deGraaf, and many others, we can further compute such structure as Jacobson radicals, Levis subalgebras, Wedderburn–Mal’cev decompositions, etc.; cf. [3, 8]. Thus the action of these rings make π‘ˆ, 𝑉, and π‘Š into modules, and the structure of the rings implies characteristic structure of the modules. Furthermore, the rings are compatible with the original product ∘, e.g. ∘ is right R⁒(∘)-linear – which makes it ideal in decomposing the products Ls⁒(Ο•;R)Γ—Lt⁒(Ο€;Q)↣Ls+t⁒(Ο€;Q). So the characteristic submodules of the various π‘ˆ, 𝑉, and π‘Š can be lifted back to characteristic subgroups of our original group. See [17] for expanded details and [13] for some carefully computed examples.

6.3 Data from small groups

It had been shown in [17, Theorem 4.9] that, using the ring M⁒(∘), there is a positive proportion of all finite groups which admit a proper non-semi-classical filter refining a semi-classical filter. Although log-scale is the only option by today’s estimation techniques (see [1, Chapter 1]), it still could be the case that, in actual proportions, very few groups benefit from this process. So here we report results of an empirical study of this question.

In [13], Maglione created a series of algorithms for the computer algebra system Magma which permit the physical construction of the M⁒(∘) filter refinements described in Section 6.2. This permitted us to conduct an experiment to measure the proportions of groups where refinement is found, adapting that code to also check for refinements by the other four rings. We found at least 9,523,263 of the 11,758,814 groups (81 %) of order at most 1000 have new non-classical commutation products; Figure 1 plots this proportion. Figure 1 shows the proportion of groups of order at most 1000 which have not only a non-classical Lie product, but in fact a non-semi-classical Lie product.

Figure 1 The proportion of groups of order N≀1000N\leq 1000 with newly found characteristic non-classical commutation Lie products.
Figure 1

The proportion of groups of order N≀1000 with newly found characteristic non-classical commutation Lie products.

While certainly in some cases the newly recovered characteristic subgroups will have been knowable by other means, we hasten to mention that most of these groups (97 %) are nilpotent of class 2 and have no other standard characteristic subgroups other than the center and commutator. So we are locating structure that is genuinely new. Table 2 shows a breakdown by ring types for groups of prime power order.

Table 1

Lower bound on the proportion of pn order groups having a non-semi-classical refinement.

𝑛
𝑝4567
257 %75 %80 %97 %
360 %70 %86 %93 %
560 %68 %88 %86 %
760 %67 %88 %81 %
54160 %51 %98 %β€”
Table 2

Lower bounds on the proportions of groups of order 2n having non-semi-classical refinements by various methods.

𝑁242526272829
Total %57 %75 %80 %97 %97 %79 %
% by M⁒(∘)43 %51 %63 %81 %68 %37 %
% by Der⁑(∘)57 %75 %76 %92 %95 %79 %
% by Cen⁑(∘)57 %57 %60 %65 %31 %5.4 %

7 Closing remarks

We have emphasized groups, but the ideas generalize to nonassociative rings and also to loops by first using the work of [15] to obtain an initial associated graded ring. From there, the source for new gradings will likely benefit from the rings described in Section 6.2. There is enormous symmetry within these rings, and often a discovery by one ring can be found by another, but we have examples within the small group data above which show each is needed.

We close with some speculations as to why the proportions seen in Figure 1 are so high. First, if a group 𝐺 has a non-semi-classical refinement then so does any central extension of 𝐺. This explains some of the proportions for 𝑝-groups of class at least 3. For 𝑝-groups of class 2, we are less certain. We can take very large random polycyclic presentations of 𝑝-groups of class 2, and our methods regularly find nothing of interest. On the other hand, in another model of randomness, such as taking quotients or subgroups of upper triangular matrices, we find almost all such examples admit non-semi-classical Lie product. So the likelihood of encountering examples with interesting structure depends greatly on the origin of the groups of interest. Finally, for non-nilpotent groups, what seems a likely explanation (and occurs in random examples) is that they have relatively large Fitting subgroups, and thus the proportions for nilpotent groups have a strong influence. But this is not the whole picture as for example 99 % of the β‰ˆβ€‰1 million groups of order 768=28β‹…3 have a non-semi-classical Lie product which exceeds the proportion of nilpotent groups of lower order. Whatever the reasons, the methods above are pliable and can be used to describe other future complex interactions between subgroups. So it is not important that they remain focused on the rings given in Section 6.2.

Acknowledgements

The author wishes to thank the referee for suggested improvements, J. Maglione for discussions of the mathematics, and L. Bartholdi who pointed out Magnus’ priority to Higman in the history of associated graded Lie rings of groups.

  1. Communicated by: Evgenii I. Khukhro

References

[1] S. R. Blackburn, P. M. Neumann and G. Venkataraman, Enumeration of Finite Groups, Cambridge Tracts in Math. 173, Cambridge University, Cambridge, 2007. 10.1017/CBO9780511542756Search in Google Scholar

[2] P. A. Brooksbank, E. A. O’Brien and J. B. Wilson, Testing isomorphism of graded algebras, Trans. Amer. Math. Soc. 372 (2019), no. 11, 8067–8090. 10.1090/tran/7884Search in Google Scholar

[3] W. A. de Graaf, Lie Algebras: Theory and Algorithms, North-Holland Math. Library 56, North-Holland, Amsterdam, 2000. Search in Google Scholar

[4] B. Eick and M. Horn, The construction of finite solvable groups revisited, J. Algebra 408 (2014), 166–182. 10.1016/j.jalgebra.2013.09.028Search in Google Scholar

[5] U. First, Private Communication. Search in Google Scholar

[6] A. Hatcher, Algebraic Topology, Cambridge University, Cambridge, 2002. Search in Google Scholar

[7] G. Higman, Lie ring methods in the theory of finite nilpotent groups, Proceedings of the International Congress of Mathematicians 1958, Cambridge University, New York (1960), 307–312. Search in Google Scholar

[8] G. Ivanyos and L. RΓ³nyai, Computations in associative and Lie algebras, Some Tapas of Computer Algebra, Algorithms Comput. Math. 4, Springer, Berlin (1999), 91–120. 10.1007/978-3-662-03891-8_5Search in Google Scholar

[9] E. I. Khukhro, Nilpotent Groups and Their Automorphisms, De Gruyter Exp. Math. 8, Walter de Gruyter, Berlin, 1993. 10.1515/9783110846218Search in Google Scholar

[10] D. E. Knuth, Finite semifields and projective planes, J. Algebra 2 (1965), 182–217. 10.1016/0021-8693(65)90018-9Search in Google Scholar

[11] R. A. Liebler, On nonsingular tensors and related projective planes, Geom. Dedicata 11 (1981), no. 4, 455–464. 10.1007/BF00181205Search in Google Scholar

[12] J. Maglione, Longer nilpotent series for classical unipotent subgroups, J. Group Theory 18 (2015), no. 4, 569–585. 10.1515/jgth-2015-0008Search in Google Scholar

[13] J. Maglione, Efficient characteristic refinements for finite groups, J. Symbolic Comput. 80 (2017), no. 2, 511–520. 10.1016/j.jsc.2016.07.007Search in Google Scholar

[14] J. Maglione, Most small 𝑝-groups have an automorphism of order 2, Arch. Math. (Basel) 108 (2017), no. 3, 225–232. 10.1007/s00013-016-1005-0Search in Google Scholar

[15] J. Mostovoy, The notion of lower central series for loops, Non-Associative Algebra and its Applications, Lect. Notes Pure Appl. Math. 246, Chapman & Hall/CRC, Boca Raton (2006), 291–298. 10.1201/9781420003451.ch23Search in Google Scholar

[16] D. J. S. Robinson, A Course in the Theory of Groups, 2nd ed., Grad. Texts in Math. 80, Springer, New York, 1996. 10.1007/978-1-4419-8594-1Search in Google Scholar

[17] J. B. Wilson, More characteristic subgroups, Lie rings, and isomorphism tests for 𝑝-groups, J. Group Theory 16 (2013), no. 6, 875–897. 10.1515/jgt-2013-0026Search in Google Scholar

[18] J. B. Wilson, On automorphisms of groups, rings, and algebras, Comm. Algebra 45 (2017), no. 4, 1452–1478. 10.1080/00927872.2016.1175617Search in Google Scholar

Received: 2020-07-30
Revised: 2020-12-01
Published Online: 2021-01-12
Published in Print: 2021-07-01

Β© 2021 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 18.4.2024 from https://www.degruyter.com/document/doi/10.1515/jgth-2020-0121/html
Scroll to top button