Home Free subgroups of inverse limits of iterated wreath products of non-abelian finite simple groups in primitive actions
Article Publicly Available

Free subgroups of inverse limits of iterated wreath products of non-abelian finite simple groups in primitive actions

  • Felix Leinen EMAIL logo and Orazio Puglisi
Published/Copyright: February 14, 2017

Abstract

Let 𝒲={Gi1i} be a set of non-abelian finite simple groups. Set W1=G1 and choose a faithful transitive primitive W1-set Δ1. Assume that we have already constructed Wn-1 and chosen a transitive faithful primitive Wn-1-set Δn-1. The group Wn is then defined as Wn=GnwrΔn-1Wn-1. If W is the inverse limit W=lim(Wn,ρn) with respect to the natural projections ρn:WnWn-1, we prove that, for each k2, the set of k-tuples of W that freely generate a free subgroup of rank k is comeagre in Wk and its complement has Haar measure zero.

1 Introduction

In 1971, the following theorem was proved by D. B. A. Epstein [5]:

Let G be a connected, finite-dimensional nonsolvable Lie group. Then, for each k>0 and for almost all k-tuples (g1,,gk) of elements in G, the group generated by g1,,gk is free of rank k.

Here, almost all k-tuples means that the set of k-tuples in Gk, which do not generate a free group of rank k, form a set of measure zero in Gk with respect to the Haar measure on G.

Epstein’s result is the starting point of a series of papers in which the phenomenon of the so-called ubiquity of free groups is considered. Roughly speaking, it has been noticed that, for many groups G, the set G(k) of k-tuples (g1,,gk)Gk such that g1,,gk is free of rank k, is very large. Typically, the largeness of G(k) is expressed by the fact that, either the group G carries a natural topology and G(k) is co-meagre (i.e., its complement is a meagre set), or G is a measure space and the complement of G(k) is a set of measure zero. In such cases we say, that almost every k-tuple of elements in G generates a free subgroup of rank k.

The next author who investigated the ubiquity of free subgroups was J. D. Dixon in [3]. He studied the case G=Sym(). The pointwise stabilizers in G of the finite subsets of form a basis of open neighborhoods of the identity, and this gives rise to a natural topology on G. Dixon proved the strong result that, for every k2, the set

H(k)={(g1,,gk)Gkg1,,gk is free of rank kand acts highly transitively on }

is co-meagre in Gk.

In 2003 the question about abundance of free subgroups was taken up by P. M. Gartside and R. W. Knight [6] from a different point of view. For Polish groups the authors find several conditions which are equivalent to co-meagreness of G(k). Amongst others they show that, if G is non-abelian and contains a dense free subgroup, then G(k) is always co-meagre.

An interesting consequence can be drawn using a result of H. D. Macpherson in [9]. He proved that the automorphism group of any 0-categorical structure contains a dense free subgroup of countable rank. Since Aut(M) is a Polish group whenever M is a countable first order structure, it turns out that, for every 0-categorical structure M, almost every k-tuple of elements of Aut(M) generates a free group of rank k. In particular, this holds for some well-known structures like the ordered set or the random graph.

The starting point of our investigation is a paper of M. Bhattacharjee [2], in which the question about abundance of free subgroups is addressed for a class of groups consisting of inverse limits of certain wreath products. Consider the set 𝒲={(Gi,Δi)1i}, where the group Gi acts faithfully and transitively on Δi for each i. For every n, let Σn=Δn××Δ1. Now W1=G1 acts naturally on Σ1. Thus, assuming that we have already defined Wn-1 and noticed that it acts naturally on Σn-1, we let Wn=GnwrΣn-1Wn-1. Clearly Wn acts naturally on Σn. With ρn:WnWn-1 denoting the natural projection map, we can form the inverse limit W=lim(Wn,ρn). For each n there is an epimorphism σn:WWn, and the set {ker(σn)n1} can be taken as a basis of open neighborhoods of 1. Bhattacharjee proved that, with respect to this topology, W(k) is co-meagre for each k. Moreover, if all the groups Gi are finite, then W is a profinite group, so that it carries a Haar measure. In this case the complement of G(k) is a set of measure zero for all k2. Improvements have been later obtained by other authors, see [1] and [15], using different methods.

In the present paper we study groups which can be constructed in a similar fashion. Let 𝒲={Gi1i} be a set of non-abelian finite simple groups Gi. Let W1=G1 and choose a faithful transitive primitive W1-set Δ1. Assume that we have already constructed Wn-1 and chosen a transitive faithful primitive Wn-1-set Δn-1. The group Wn is then defined as Wn=GnwrΔn-1Wn-1. As above, we can form the inverse limit W=lim(Wn,ρn) with respect to the natural projections ρn:WnWn-1.

Inverse limits of wreath products of finite simple groups have been studied in several paper. J. S. Wilson [16] produced examples of hereditarily just-infinite profinite groups by means of inverse limits of wreath products of finite simple groups. This construction has been generalized recently by M. Vannacci [14]. The groups constructed by these authors are, in fact, of a quite particular nature. A characterization of (hereditarily) just infinite profinite groups is given in [12], where C. Reid finds necessary and sufficient conditions for a profinite group to be just infinite or hereditarily just infinite, in terms of the inverse system of its finite images. On the other hand (hereditarily) just infinite profinite groups form a pretty large class and very few general properties are known. Therefore specific constructions like the ones cited above, are of interest.

Inverse limits of wreath products of simple groups have been considered also by M. Quick, who showed in [10] and [11] that such inverse limits (when at each stage the wreathing action is transitive and faithful) are positively finitely generated. He gives lower bounds for the probability that two randomly chosen elements generate such a group W. The problem of finding an explicit set of generators was considered subsequently by Vannacci [13].

We study the abundance of free subgroups of the groups W and prove that, for each such group and each k2, the set W(k) is co-meagre and its complement has measure zero with respect to the Haar measure (Theorem 3.1). Our methods rely upon the ones developed in [2], but the fact that the primitive sets Δn can be chosen freely, requires a different kind of approach.

2 Non-triviality of words in iterated wreath products

In this paper, we shall consider the following classes.

  1. 𝒲1 shall be the class of all pairs (G,Γ), where G is a non-abelian finite simple group and Γ denotes some primitive faithful G-set.

  2. For every n>1, the class 𝒲n shall consist of all pairs (SwrΔG,Γ) such that

    1. S is a non-abelian finite simple group,

    2. (G,Δ)𝒲n-1, and

    3. Γ is a faithful primitive (SwrΔG)-set.

Consider the free group Fk of rank k with free generators x1,x2,,xk. We shall make use of the following notation. If w=y1ϵ1y2ϵ2ynϵn is a reduced word of length n with yi{xjj=1,,k} and ϵi=±1, we let

  1. w0=1 and wi=y1ϵ1y2ϵ2yiϵi for i=1,,n,

  2. w¯i=wi(x1,,xk)-1 if ϵi=-1 and w¯i=wi-1(x1,,xk)-1 if ϵi=1.

For k-tuples g=(g1,,gk) and h=(h1,,hk) from Gk we shall denote the k-tuple (g1h1,,gkhk) by gh. Moreover, for every pair of element a,b of a group, we shall indicate by ab the element b-1ab.

The proof of the following lemma is straightforward.

Lemma 2.1.

Let w=xi1ϵ1xi2ϵ2xinϵn be a reduced word of length n in Fk and let G be any group with a normal subgroup N. Consider (g1,,gk)Gk and (b1,,bk)Nk. Then

w(b1g1,,bkgk)=a1ϵ1anϵnw(g1,,gk),

where aj=(bij)w¯j(g1,,gk) for j=1,,n.

Lemma 2.2.

Let (G,Γ) be an element of Wn for some n2, with G=SwrΔH, where (H,Δ)Wn-1. Then one of the following holds:

  1. The action of G on Γ is the diagonal action.

  2. Γ=ΣΔ for a suitable primitive S-set Σ, and the action of G on Γ is the product action.

  3. G is isomorphic to a twisted wreath product SwrΘX with XH, and the socle SΘ of G acts regularly on Γ.

Proof.

The socle of G is the base group B=SΔ, and it acts transitively on Γ. If the action of B is regular, then G is the semidirect product of B and Gγ for any γΓ. Moreover, X=Gγ acts on the components of B in the same way as H. By [7], the group G is isomorphic to a twisted wreath product SwrΘX in this case.

We may thus assume that B is not regular, so that Bγ1. Since |Δ|>1, we can use [4, Theorem 4.6.A] to see that one of the following cases must hold: either G is of diagonal type, or G is a primitive subgroup of UwrΛSym(Λ) acting in product action and U is itself primitive with non-regular socle. In order to complete the proof, it suffices to show that U=S and Λ=Δ in the latter case.

But then, soc(U)Λ is the socle of G, so that the components of soc(U) form a block under the action of G on the components of B. This action is primitive, so that soc(U) must have just one component. Therefore U is isomorphic to a subgroup of Aut(S) containing S, so that U/S is solvable. The normal closure UG of U contains B and it is easy to see that UG/B is solvable. However, H has no non-trivial solvable normal subgroups. This shows that UG=B and U=S. ∎

Consider some (G,Γ)𝒲n. In order to give a lower bound for the size of Γ, we define the function ξ: recursively via

ξ(0)=1  and  ξ(n)=5ξ(n-1)for all n1.

Lemma 2.3.

Using the above notation,

ξ(n)|Γ|𝑎𝑛𝑑(n+12)<3ξ(n-1)-2for all n3.

Proof.

The first claim holds true for n=1, because the symmetric group of degree 4 is solvable. If (SwrΔH,Γ)𝒲n and ξ(n-1)|Δ| by induction, then Lemma 2.2 implies that |Γ|5|Δ|5ξ(n-1)=ξ(n).

The second inequality holds true for n=3. For n4, induction gives

(n+12)=12((n+1)2+(n+1))=12((n2+n)+2(n+1))
<32(n2+n)=3(n2)<3ξ(n-2)-1<3ξ(n-1)-2.

The following example shows that the above bound on the degree of G is best possible.

Example 2.4.

For each n1, there exists (Gn,Ωn)Wn such that |Ωn|=ξ(n).

Proof.

For n=1 we choose G1=Alt(5) and the natural Alt(5)-set Ω1. Suppose that we have already found (Gn-1,Ωn-1)𝒲n-1 with |Ωn-1|=ξ(n-1). Then Gn=Alt(5)wrΩn-1Gn-1 acts primitively on Ωn=Ω1Ωn-1 in product action, and the degree of this representation is ξ(n). ∎

Consider some (G,Γ)𝒲n. In the sequel,

1=KnKn-1K1K0=G

shall denote the ascending socle series in G, i.e., K-1/K=soc(G/K) for =n,,1. Each K is characteristic in G, and K-1/K is the unique minimal normal subgroup in G/K.

For k-tuples g=(g1,,gk) and h=(h1,,hk) in Gk, we shall write gh(modK) in order to express, that giK=hiK for i=1,,k.

We are now in a position to prove the main theorem of this section.

Theorem 2.5.

Consider a reduced word w of length n>0 in the free group Fk of rank k. Let (G,Γ)Wd+n for some dN. Then, for any y=(y1,,yk) in GkKdk, there exist g=(g1,,gk)Gk and γΓ such that

  1. gy(modKd),

  2. the elements γj=γwj(g) are pairwise distinct for j=0,,n.

In particular, w(g)1.

Proof.

If k>n, then at least k-n variables do not appear in the reduced form of w. We may therefore assume that kn. It is also easy to check that the result is true for n2. Let n3 and assume by induction, that the statement holds for words of length at most n-1. By Lemma 2.2 the group G=SwrΔH acts either in diagonal action or in product action, with respect to a suitable primitive representation of S, or its socle is regular. Notice that soc(G)=Kd+n-1 is the base subgroup B of the wreath product. Another fact which will be used frequently in the sequel is that, if hy(modKd) for some hHk, then bhy(modKd) for every bBk.

Our induction provides elements h=(h1,h2,,hk)Hk and δΔ such that hy(modKd) and such that the elements δi=δwi(h) are pairwise distinct for i=0,,n-1. Choose any faithful transitive S-set Ω and let G act on Ω×Δ in the natural way. The remark in the first paragraph of the proof of [2, Lemma 4] shows that we can find ε=(ω,δ)Ω×Δ and a subset Bk containing at least

|S||Δ|(k-1)|S||Δ|-1(|S|-1)=|S||Δ|k(1-1|S|)=|B|k(1-1|S|)

elements such that, whenever b, then εwi(bh)εwj(bh) for 0i<jn. In particular, wi(bh)wj(bh) for 0i<jn and b.

Consider any primitive faithful G-set Γ. We aim to show that, for each b, we can find some γΓ such that the element xij=wi(bh)wj(bh)-1 does not belong to Gγ for 0i<jn. Assume by way of contradiction that this does not hold for a certain b. It follows that Γ is the union of the sets

Γij={γγxij=γ}.

The number of these sets is (n+12). Hence there exists a pair (i,j) such that

μ=|Γ|(n+12)|Γij|.

Therefore xij fixes at least μ elements from Γ. The desired contradiction will now be derived separately in each of the three cases furnished by Lemma 2.2.

Case 1: Product action. In this case, G acts on Γ=ΣΔ, where Σ is a primitive faithful S-set. We shall try to estimate the size of the set of fixed points fixΓ(g) for a non-trivial element g=fxG (with fB, xH). Recall that g acts on Γ via (φg)(δ)=φ(δx-1)f(δx-1) for all φ:ΔΣ.

Suppose first that x-1=(δ1δ2δ) is a non-trivial cycle. If φ:ΔΣ is a fixed point for g, and if δfixΔ(x), then φ(δ)fixΣ(f(δ)). Moreover, we have φ(δi+1)f(δi+1)=φ(δi) for i=1,,-1 and φ(δ)f(δ)=φ(δ1).

Therefore, if we want to construct an element φfixΓ(g), then we should, first of all, choose φ(δ)fixΣ(f(δ)) whenever δsuppΔ(x). Then, in order to define φ on suppΔ(x), we can choose freely the value φ(δ1) (this allows |Σ| different choices), and then the value of φ on the other points δi is determined by the above relations. However, not every such choice is admissible, since the -th relation φ(δ)f(δ)=φ(δ1) might not hold. We conclude that

|fixΓ(g)||Σ|(δfixΔ(x)|fixΣ(f(δ))|).

It is now an easy matter to work out the general case: If x has r non-trivial cycles in its action on Δ, then

|fixΓ(g)||Σ|r(δfixΔ(x)|fixΣ(b(δ))|)|Σ|r+|fixΔ(x)|.

Suppose that x1. Since n3, the group H is not an alternating group. Therefore [8, Corollary 3] yields suppΔ(x)2(|Δ|-1), whence

|fixΔ(x)||Δ|-2(|Δ|-1).

Now r|Δ|-|fixΔ(x)|2. Hence

r+|fixΔ(x)||Δ|+|fixΔ(x)|2|Δ|-|Δ|+1,|fixΓ(g)||Σ||Δ|-|Δ|+1.

Suppose now that x=1 and g=fB. In this situation it is clear that

|fixΓ(g)|(|Σ|-3)|Σ||Δ|-1,

corresponding to the case when S is an alternating group in natural action, f has a single non-trivial entry and this entry is a 3-cycle.

As in the proof of [2, Lemma 3] at most one of the elements xij (say x0n) has trivial top component, so that the set

Γ0={i,j}{0,n}ΓijΓΓ0n

has size |Γ0||Σ||Δ|-(|Σ|-3)|Σ||Δ|-1=3|Σ||Δ|-1. An application of the pigeonhole principle yields that there exists a 2-set {i,j}{0,n} for which Γij has size at least

μ0=3|Σ||Δ|-1(n+12)-1.

The inequality μ0|Σ||Δ|-|Δ|+1 now implies |Σ||Δ|-2(n+12)-1 . But this contradicts Lemma 2.3.

Case 2: Diagonal action. In this case, the set Γ can be identified with the set of right cosets in B=SΔ of the diagonal subgroup D of B. As before, we try to describe fixΓ(g) for g=fxG (with fB, xH). Recall that the action of G on Γ is induced from the product action of G on B with respect to the right regular representation of S on Σ=S.

If g fixes Dt (where tB), then there exists dS such that (tf)x=d¯t, where d¯ indicates the element of D, whose entries are all equal to d. Thus,

t(δ)f(δ)=dt(δx)

for every δΔ. In particular, f(δ)=dt(δ) whenever δfixΔ(x).

We shall now use this information in order to count the elements in fixΓ(g). There are at most |S| possible choices for d. Once d has been selected, each t(δ) (δfixΔ(x)) can be chosen in |CS(d)| different ways. Let χ(x) denote the set of x-orbits c in Δ with length (c)>1. If c=(δ1δ2δ(c))χ, then, for any choice of t(δ1), the values t(δi) (i=2,,(c)) are uniquely determined by the equation t(δi)f(δi)=dt(δi+1). Moreover, also the equation

t(δ(c))f(δ(c))=dt(δ1)

must hold, and it is not difficult to see, that this happens if and only if

t(δ1)=d-(c)t(δ1)f(δ1)f(δm).

Thus t(δ1) must satisfy

(d(c))t(δ1)=f(δ1)f(δ(c)).

Hence the number of possible choices for t(δ1) is |CS(d(c))|. Altogether we conclude that

|fixΓ(g)||S||CS(d)||fixΔ(x)|cχ(x)|CS(d(c))||S|1+|fixΔx|+|χ(x)|.

If x1, then the same arguments as in Case 1 (product action) yield

|fixΓ(g)||S||Δ|-|Δ|+2.

On the other hand, if x=1 but f1, then we obtain that all values of f must lie in the conjugacy class of a single non-trivial element dS, and that we have |CS(d)| choices for each of the values t(δ) (δΔ). Moreover, if t1 and t2 have been obtained in this way, then t1t2-1CS(d)Δ. In particular, Dt1=Dt2 if and only if t1t2-1DCS(d)Δ. Therefore,

|fixΓ(g)||CS(d)||Δ|-1|S||Δ|-1.

Considering Γ0 and μ0 as in Case 1 (product action), we find that

|Γ0||S||Δ|-|S||Δ|-1=|S||Δ|-1(|S|-1)

and

μ0=|S||Δ|-1(|S|-1)(n+12)-1|S||Δ|-|Δ|+2.

This leads to the contradiction |S||Δ|-3(|S|-1)(n+12)-1 (cf. Lemma 2.3).

Case 3: The socle acts regularly. Now we have Γ=B and G=BX, where B=SΔ acts regularly on Γ and X is a point stabilizer. In order to avoid confusion, the action of sG on γΓ shall be denoted by γ.s. As before, we consider g=fxG with fB, xX and try to count the number of its fixed points. Notice that x might not belong to H. Therefore x=bh with bB, hH.

Now γfixΓ(g) if and only if γ=γ.g=γ.(fbh) if and only if γ.h-1=γfb. The latter means that

γ(δh)=γ(δ)f(δ)b(δ)for all δΔ.

If h fixes δ, then f(δ)b(δ)=1 and γ(δ) can take any value in S. If (δ1δ2δm) is an h-orbit, then γ(δi+1) is determined by γ(δi) via

γ(δi+1)=γ(δi)f(δi)b(δi)for i=1,,m-1,

but we can choose γ(δ1). With r denoting the number of non-trivial h-orbits in Δ, we obtain

|fixΓ(g)||S||fixΔh|+r|S||Δ|-|Δ|+1

as in Case 1 (product action) for elements g satisfying h1. However, any non-trivial gG either has no fixed-point in Γ or satisfies h1, because B acts regularly on Γ. In particular,

μ=|S||Δ|(n+12)|Γij||S||Δ|-|Δ|+1and|S||Δ|-1(n+12),

a contradiction to Lemma 2.3. ∎

A closer look at the proof of the above result, gives information on the number of k-tuples g which satisfy the assertion of Theorem 2.5.

Corollary 2.6.

In the notation of Theorem 2.5, let

G=SnwrΔn(Sn-1wrΔn-1((S1wrΔ1H))),

where (H,Δ1)Wd and S1,,Sn are non-abelian finite simple groups. If |Si|=si for all i and |G|=r|H|, then the number of k-tuples gGk satisfying the statement of Theorem 2.5 is at least

rki=1n(1-1si).

Proof.

Clearly, G=KdH. Suppose first that w(y)1. Then, for every bKdk, the k-tuple g=by satisfies w(g)1 too. The number of such tuples is rk. Suppose next, that w(y)=1. Following the recursive procedure in the proof of Theorem 2.5, the required elements are constructed in n steps. In the i-the step, every k-tuple constructed so far, can be extended in at least si|Δi|(1-1si) different ways. Since i=1nsi|Δi|=|Kd|, the claim follows in this case too. ∎

3 Ubiquity of free subgroups in certain profinite groups

Let 𝒮={Sii1} be a family of non-abelian finite simple groups. We define recursively a sequence of groups and permutation representations as follows:

  1. W1=S1 and Δ1 is any faithful primitive W1-set,

  2. Wn=SnwrΔn-1Wn-1 and Δn is any faithful primitive Wn-set for all n1.

Clearly, (Wn,Δn)𝒲n for all n, and there are the natural surjective homomorphisms ρn:WnWn-1. We can thus form the inverse limit W=lim(Wn,ρn), which is a profinite group. A basis of open neighborhoods of the identity in W is given by the subgroups Kn=ker(πn), where πn:WWn denotes the natural projection. We shall denote the class of groups constructed in this way by 𝒲(𝒮).

For every wFk, let

C(w)={hWkw(h)1}andC=1wFkC(w).

If h=(h1,,hk)C, then h1,,hk is free of rank k, because w(h)1 for every non-trivial wFk. Thus C=W(k). It is our aim to prove that W(k) is a rather large subset of Wk. Recall that a meagre subset of a topological space is a countable union of nowhere dense subsets (i.e., of subsets, whose closure has empty interior).

Theorem 3.1.

If S is a set of non-abelian finite simple groups, and if WW(S), then the set W(k) is dense in Wk and its complement is a meagre subset of Wk.

Proof.

Consider a non-trivial wFk. The evaluation map μw:WkW, which sends every hWk to w(h)W, is continuous. Moreover, W is a Hausdorff space. Therefore μw-1(1) is closed and C(w)=Wkμw-1(1) is open.

Let n be the length of w and choose a non-trivial element yWk. There exists d such that yKdk. An application of Theorem 2.5 gives hWk such that hy(modKd) and w(h)Kd+n. In particular, w(h)1, so that hC(w). On the other hand hKd=yKd, so that the intersection of the open neighborhood yKd of y with C(w) is non-empty. This shows that C(w) is dense in Wk. Since Wk is a compact Hausdorff space, Baire’s theorem yields that W(k) is dense, being a countable intersection of open dense subsets. Finally,

WkW(k)=Wk1wFkC(w)=1wFk(WkC(w)).

Since each C(w) is open and dense, WkC(w) is nowhere dense. This proves the claim. ∎

Theorem 3.1 has the following consequence.

Corollary 3.2.

Suppose that S is a set of non-abelian finite simple groups. Then every WW(S) contains 20 free subgroups of rank 20.

Proof.

The group W is metrizable and complete. Therefore it is a Polish space. Since W is not discrete, [6, Theorem 1] yields that W contains a dense free subgroup. The claim now follows from [6, Corollary 3]. ∎

The largeness of W(k) can also be considered from a measure-theoretic point of view, because every group in 𝒲(𝒮) is a measure space with respect to its canonical Haar measure. Since each compact group has finite volume, we may choose the normalized Haar measure and consider the group as a probability space.

The arguments contained in [2, Section 5] can be applied, without any change, in our set up. This gives the following result.

Theorem 3.3.

Let S be a set of non-abelian finite simple groups, and let WS. Then the set WkW(k) has measure zero with respect to the normalized Haar measure of W.

We finish this article with two open problems.

Questions 3.4.

Suppose that the profinite group W is formed from iterated wreath products as above, but without further restrictions to the nature of the faithful Wn-sets (i.e., we do not necessarily require that these sets are primitive).

  1. Is it still true then that WkW(k) is meagre?

  2. Is it still true then that WkW(k) has measure zero with respect to the normalized Haar measure?

In order to answer these questions, we would need a stronger form of Theorem 2.5. Namely it would be useful to know that the claim of Theorem 2.5 holds when the wreath products considered are defined without restrictions on the faithful permutation representations used at each stage. Although we have been unable to supply a proof of this stronger result, we believe that the answers to the above questions are positive.


Communicated by Andrea Lucchini


Acknowledgements

The second author wishes to thank the University of Mainz for financial support and for the warm hospitality during the period in which the research reported in this paper was carried out.

References

[1] M. Abért, Group laws and free subgroups of topological groups, Bull. Lond. Math. Soc. 37 (2005), no. 4, 525–534. 10.1112/S002460930500425XSearch in Google Scholar

[2] M. Bhattacharjee, The ubiquity of free subgroups in certain inverse limits of groups, J. Algebra 172 (1995), 134–146. 10.1006/jabr.1995.1053Search in Google Scholar

[3] J. D. Dixon, Most finitely generated permutation groups are free, Bull. Lond. Math. Soc. 22 (1990), 222–226. 10.1112/blms/22.3.222Search in Google Scholar

[4] J. D. Dixon and B. Mortimer, Permutation Groups, Grad. Texts in Math. 163, Springer, New York, 1996. 10.1007/978-1-4612-0731-3Search in Google Scholar

[5] D. B. A. Epstein, Almost all subgroups of a Lie group are free, J. Algebra 19 (1971), 261–262. 10.1016/0021-8693(71)90107-4Search in Google Scholar

[6] P. M. Gartside and R. W. Knight, Ubiquity of free subgroups, Bull. Lond. Math. Soc. 35 (2003), 624–634. 10.1112/S0024609303002194Search in Google Scholar

[7] J. Lafuente, On restricted twisted wreath products of groups, Arch. Math. 43 (1984), 208–209. 10.1007/BF01247564Search in Google Scholar

[8] M. W. Liebeck and J. Saxl, Minimal degrees of primitive permutation groups, with an application to monodromy groups of covers of Riemann surfaces, Proc. Lond. Math. Soc. (3) 63 (1991), 266–314. 10.1112/plms/s3-63.2.266Search in Google Scholar

[9] H. D. Macpherson, Groups of automorphisms of 0-categorical structures, Q. J. Math. 37 (1986), 449–465. 10.1093/qmath/37.4.449Search in Google Scholar

[10] M. Quick, Probabilistic generation of wreath products of non-abelian finite simple groups, Comm. Algebra 32 (2004), 4753–4768. 10.1081/AGB-200036751Search in Google Scholar

[11] M. Quick, Probabilistic generation of wreath products of non-abelian finite simple groups. II, Int. J. Algebra Comput. 16 (2006), 493–503. 10.1142/S0218196706003074Search in Google Scholar

[12] C. D. Reid, Inverse system characterizations of the (hereditarily) just infinite property in profinite groups, Bull. Lond. Math. Soc. 44 (2012), no. 3, 413–425. 10.1112/blms/bdr099Search in Google Scholar

[13] M. Vannacci, Finite generation of iterated wreath products in product action, Arch. Math. (Basel) 105 (2015), 205–214. 10.1007/s00013-015-0797-7Search in Google Scholar

[14] M. Vannacci, On hereditarily just infinite profinite groups obtained via iterated wreath products, J. Group Theory 19 (2016), no. 2, 233–238. 10.1515/jgth-2015-0032Search in Google Scholar

[15] J. S. Wilson, On just infinite abstract and profinite groups, New Horizons in Pro-p Groups, Progr. Math. 184, Birkhäuser, Boston (2000), 219–246. 10.1007/978-1-4612-1380-2_5Search in Google Scholar

[16] J. S. Wilson, Large hereditarily just infinite groups, J. Algebra 324 (2010), 248–255. 10.1016/j.jalgebra.2010.03.023Search in Google Scholar

Received: 2016-4-18
Revised: 2016-12-30
Published Online: 2017-2-14
Published in Print: 2017-7-1

© 2017 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 20.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/jgth-2016-0061/html
Scroll to top button