Skip to content
BY 4.0 license Open Access Published by De Gruyter February 18, 2020

A polynomial bound for the number of maximal systems of imprimitivity of a finite transitive permutation group

  • Andrea Lucchini ORCID logo , Mariapia Moscatiello ORCID logo and Pablo Spiga ORCID logo EMAIL logo
From the journal Forum Mathematicum

Abstract

We show that there exists a constant a such that, for every subgroup H of a finite group G, the number of maximal subgroups of G containing H is bounded above by a|G:H|3/2. In particular, a transitive permutation group of degree n has at most an3/2 maximal systems of imprimitivity. When G is soluble, generalizing a classic result of Tim Wall, we prove a much stronger bound, that is, the number of maximal subgroups of G containing H is at most |G:H|-1.

MSC 2010: 20E28; 20B15; 20F05

1 Introduction

In 1961, Tim Wall [12] has conjectured that the number of maximal subgroups of a finite group G is less than the group order |G|. Wall himself proved the conjecture under the additional hypothesis that G is soluble. The first remarkable progress towards a good understanding of Wall’s conjecture is due to Liebeck, Pyber and Shalev [9]; they proved that all, but (possibly) finitely many, simple groups satisfy Wall’s conjecture. Actually, Liebeck, Pyber and Shalev prove [9, Theorem 1.3] a polynomial version of Wall’s conjecture: there exists an absolute constant c such that every finite group G has at most c|G|3/2 maximal subgroups. Based on the conjecture of Guralnick on the dimension of certain first cohomology groups [6] and on some computer computations of Frank Lübeck, Wall’s conjecture was disproved in 2012 by the participants of an AIM workshop; see [7].

The question of Wall can be generalized in the context of finite permutation groups and this was done by Peter Cameron; see [3] (also for the motivation of this question).

Question 1.1 (Cameron [3]).

Is the number of maximal blocks of imprimitivity through a point for a transitive group G of degree n bounded above by a polynomial of degree n? Find the best bound!

To see that this question extends naturally the question of Wall, we fix some notation. Given a finite group G and a subgroup H of G, we denote by

max(H,G):=|{MM maximal subgroup of G with HM}|

the number of maximal subgroups of G containing H. Now, if Ω is the domain of a transitive permutation group G and ωΩ, then there exists a one-to-one correspondence between the maximal systems of imprimitivity of G and the maximal subgroups of G containing the point stabilizer Gω, and hence Question 1.1 asks for a polynomial upper bound for max(Gω,G) as a function of n=|G:Gω|. When n=|G|, that is, G acts regularly on itself, the question of Cameron reduces to the question of Wall and [9, Theorem 1.3] yields a positive solution in this case, with exponent 32.

The main result of this paper is a positive solution to Question 1.1.

Theorem 1.2.

There exists a constant a such that, for every finite group G and for every subgroup H of G, we have max(H,G)a|G:H|3/2. In particular, a transitive permutation group of degree n has at most an3/2 maximal systems of imprimitivity.

In the case of soluble groups, we actually obtain a much tighter bound, which extends the result of Wall [12, (8.6), p. 58] for soluble groups on his own conjecture.

Theorem 1.3.

If G is a finite soluble group and H is a proper subgroup of G, then max(H,G)|G:H|-1. In particular, a soluble transitive permutation group of degree n2 has at most n-1 maximal systems of imprimitivity.

2 Preliminaries

We start by reviewing some basic results on G-groups, on monolithic primitive groups and on crowns tailored to our proof of Theorem 1.2. For the first part we follow [5], for the second part we follow [8] and for the third part we follow [1, Chapter 1] and [5]. This section will also help for setting some notation. All groups in this paper are finite.

2.1 Monolithic primitive groups and crown-based power

Recall that an abstract group L is said to be primitive if it has a maximal subgroup with trivial core. Incidentally, given a group G and a subgroup M we denote by

coreG(M):=gGMg

the core of M in G. The soclesoc(L) of a primitive group L is either a minimal normal subgroup, or the direct product of two non-abelian minimal normal subgroups. A primitive group L is said to be monolithic if the first case occurs, that is, soc(L) is a minimal normal subgroup of L and hence (necessarily) L has a unique minimal normal subgroup.

Let L be a monolithic primitive group and let A:=soc(L). For each positive integer k, let Lk be the k-fold direct product of L. The crown-based power of L of size k is the subgroup Lk of Lk defined by

Lk:={(l1,,lk)Lkl1lk(modA)}.

Equivalently, if we denote by diag(Lk) the diagonal subgroup of Lk, then Lk=Akdiag(Lk).

For the proof of the next lemma we need some basic terminology, which we borrow from [11, Sections 4.3 and 4.4]. Let κ be a positive integer and let A be a direct product S1××Sκ, where the Si are pairwise isomorphic non-abelian simple groups. We denote by πi:ASi the natural projection onto Si. A subgroup X of A is said to be a strip if X1 and, for each i{1,,κ}, either XKer(πi)=1 or πi(X)=1. The support of the strip X is the set {i{1,,κ}πi(X)1}. The strip X is said to be full if πi(X)=Si for all i in the support of X. Two strips X and Y are disjoint if their supports are disjoint. A subgroup X of A is said to be a subdirect subgroup if πi(X)=Si for each i{1,,κ}.

Scott’s lemma (see for instance [11, Theorem 4.16]) shows (among other things) that if X is a subdirect subgroup of A, then X is a direct product of pairwise disjoint full strips of A.

Lemma 2.1.

Let Lk be a crown-based power of L of size k having non-abelian socle Nk and let H be a core-free subgroup of Lk contained in Nk. Then |Nk:H|5k.

Proof.

We argue by induction on k. If k=1, then the result is clear because Nk=N has no proper subgroups having index less then 5. Suppose that k2 and write N:=N1××Nk, where N1,,Nk are the minimal normal subgroups of Lk contained in Nk. For each i{1,,k}, we denote by πi:NkNi the natural projection onto Ni.

Suppose that there exists i{1,,k} with πi(H)<Ni. Then NiH/Ni is a core-free subgroup of Lk/NiLk-1 and is contained in Nk/Ni. Therefore, by induction, |Nk:HNi|=|Nk/Ni:HNi/Ni|5k-1. Furthermore, |HNi:H|=|Ni:HNi|5 because Ni has no proper subgroups having index less then 5. Therefore, |Nk:H|5k.

Suppose that πi(H)=Ni for every i{1,,k}. Since N is non-abelian, we may write Ni=Si,1××Si, for some pairwise isomorphic non-abelian simple groups Si,j of cardinality s. For each i{1,,k} and j{1,,}, we denote by πi,j:NkSi,j the natural projection onto Si,j. Since πi(H)=Ni, we deduce πi,j(H)=Si,j for every i{1,,k} and j{1,,}. In particular, H is a subdirect subgroup of S1,1××Sk,, and hence (by Scott’s lemma) H is a direct product of pairwise disjoint full strips. Since no Ni is contained in H, there exist two distinct indices i1,i2{1,,k} and j1,j2{1,,} such that (i1,j1) and (i2,j2) are involved in the same full strip of H. If we now consider the projection πi1,i2:NkNi1×Ni2, we obtain |Ni1×Ni2:πi1,i2(H)|s6052. The inductive hypothesis applied to Ker(πi1,i2)H yields |Ker(πi1,i2):Ker(πi1,i2)H|5k-2, and hence |Nk:H|5k. ∎

In the proofs of Theorem 1.2 and 1.3, we use without mention the following basic fact.

Lemma 2.2.

Let M be a normal subgroup of a crown-based power Lk with socle Nk. Then either MNk or NkM.

Proof.

For each i{1,,k}, we write Ni:={(n1,,nk)Nknj=1 for all j{1,,k}{i}}. In particular, N=N1××Nk.

Let M be a normal subgroup of the crown-based power Lk with socle Nk and with MNk. Let mMNk. For each i{1,,k}, since M does not centralize Ni, we deduce 1[M,Ni]MNi. As Ni is one of the minimal normal subgroups of Lk, we must have NiM. Therefore, Nk=N1××NkM. ∎

2.2 Basic facts on G-groups

Given a group G, a G-groupA is a group A together with a group homomorphism θ:GAut(A) (for simplicity, we write ag for the image of aA under the automorphism θ(g)). Given a G-group A, we have the corresponding semi-direct productAθG (or simply AG when θ is clear from the context), where the multiplication is given by

g1a1g2a2=g1g2a1g2a2

for every a1,a2A and for every g1,g2G. A G-group A is said to be irreducible if G leaves no non-identity proper normal subgroup of A invariant.

Two G-groups A and B are said to be G-isomorphic (and we write AGB) if there exists an isomorphism φ:AB such that

(ag)φ=(aφ)g

for every aA and for every gG. Similarly, we say that A and B are G-equivalent (and we write AGB) if there exist two isomorphisms φ:AB and Φ:AGBG such that the following diagram commutes:

Being “G-equivalent” is an equivalence relation among G-groups coarser than the “G-isomorphic” equivalence relation, that is, two G-isomorphic G-groups are necessarily G-equivalent. The converse is not necessarily true: for instance, if A and B are two isomorphic non-abelian simple groups and G:=A×B acts on A and on B by conjugation, then AGB and AGB. However, when A and B are abelian, the converse is true, that is, if A and B are abelian, then AGB if and only if AGB; see [8, page 178].

Let G be a group and let A:=X/Y be a chief factor of G, where X and Y are normal subgroups of G. Clearly, the action by conjugation of G endows A with the structure of a G-group and, in fact, A is an irreducible G-group. On the set of chief factors, the G-equivalence relation is easily described. Indeed, it is proved in [8, Proposition 1.4] that two chief factors A and B of G are G-equivalent if and only if either

  1. A and B are G-isomorphic, or

  2. there exists a maximal subgroup M of G such that G/coreG(M) has two minimal normal subgroups N1 and N2 that are G-isomorphic to A and B, respectively.

(The example in the previous paragraph witnesses that the second possibility does arise.) From this, it follows that, for every monolithic primitive group L and for every k, the minimal normal subgroups of the crown-based power Lk are all Lk-equivalent.

2.3 Crowns of a finite group

Let X and Y be normal subgroups of G with A=X/Y being a chief factor of G. A complementU to A in G is a subgroup U of G such that

G=UXandY=UX.

We say that A=X/Y is a Frattini chief factor if X/Y is contained in the Frattini subgroup of G/Y; this is equivalent to saying that A is abelian and there is no complement to A in G. The number δG(A) of non-Frattini chief factors G-equivalent to A in any chief series of G does not depend on the series, and hence δG(A) is a well-defined integer depending only on the chief factor A.

We denote by LA the monolithic primitive group associated to A, that is,

LA:={A(G/CG(A)) if A is abelian,G/CG(A) otherwise.

If A is a non-Frattini chief factor of G, then LA is a homomorphic image of G. More precisely, there exists a normal subgroup N of G such that

G/NLAandsoc(G/N)GA.

Consider now the collection 𝒩A of all normal subgroups N of G with G/NLA and soc(G/N)GA: the intersection

RG(A):=N𝒩AN

has the property that G/RG(A) is isomorphic to the crown-based power (LA)δG(A), that is, G/RG(A)(LA)δG(A).

The socle IG(A)/RG(A) of G/RG(A) is called the A-crown of G and it is a direct product of δG(A) minimal normal subgroups all G-equivalent to A.

We conclude this preliminary section with two technical lemmas and one of the main results from [9].

Lemma 2.3 ([1, Lemma 1.3.6]).

Let G be a finite group with trivial Frattini subgroup. There exists a chief factor A of G and a non-identity normal subgroup D of G with IG(A)=RG(A)×D.

Lemma 2.4 ([5, Proposition 11]).

Let G be a finite group with trivial Frattini subgroup, let IG(A),RG(A) and D be as in the statement of Lemma 2.3 and let K be a subgroup of G. If G=KD=KRG(A), then G=K.

Theorem 2.5 ([9, Theorem 1.4]).

There exists a constant c such that every finite group has at most cn3/2 core-free maximal subgroups of index n.

Theorem 2.5 is an improvement of [10, Corollary 2]. We warn the reader that the statement of Theorem 2.5 is slightly different from that of [9, Theorem 1.4]: to get Theorem 2.5 one should take into account [9, Theorem 1.4] and the remark following its statement.

3 Proofs of Theorems 1.2 and 1.3

In this section, we prove Theorems 1.2 and 1.3. Our proofs are inspired from some ideas developed in [4]. Moreover, our proofs have some similarities, and hence we start by deducing some general facts holding for both.

We start by defining the universal constant a. Observe that the series u=1u-3/2 converges. We write

a:=u=11u3/2.

Let c be the universal constant arising from Theorem 2.5. We define

a:=11ca1-1/23/2.

Recall that max(H,G) is the number of maximal subgroups of G containing H. For the proofs of Theorems 1.2 and 1.3 we argue by induction on |G:H|+|G|. The case |G:H|=1 for the proof of Theorem 1.2 is clear because max(H,G)=0. Similarly, the case that H is maximal in G for the proof of Theorem 1.3 is clear because max(H,G)=1. In particular, for the proof of Theorem 1.2, we suppose |G:H|>1 and, for the proof of Theorem 1.3, we suppose that H is not maximal in G.

Consider

H~:=HM<GM max. in GM.

Observe that max(H,G)=max(H~,G). In particular, when H<H~, we have |G:H~|<|G:H| and hence, by induction, we have max(H,G)=max(H~,G)a|G:H~|3/2<a|G:H|3/2. Moreover, when G is soluble, we have max(H,G)=max(H~,G)|G:H~|-1<|G:H|-1. Therefore, we may suppose H=H~, that is,

(3.1)H is an intersection of maximal subgroups of G.

Suppose that H contains a non-identity normal subgroup N of G. Since max(H,G)=max(H/N,G/N) and |G/N|<|G|, by induction, we have max(H,G)=max(H/N,G/N)a|G/N:H/N|3/2=a|G:H|3/2. Moreover, when G is soluble, we have max(H,G)=max(H/N,G/N)|G/N:H/N|-1=|G:H|-1. Therefore, we may suppose

(3.2)coreG(H)=1.

Let F be the Frattini subgroup of G. From (3.1) we have FH and hence, from (3.2), F=1. In particular, we may now apply Lemma 2.3 to the group G.

Choose I, R and D as in Lemma 2.3. By (3.1), we may write

H=X1XρY1Yσ,

where X1,,Xρ are the maximal subgroups of G not containing D and Y1,,Yσ are the maximal subgroups of G containing D. We define

X:=X1XρandY:=Y1Yσ.

Thus H=XY.

For every i{1,,ρ}, since DXi, we have G=DXi, and hence Lemma 2.4 (applied with K:=Xi) yields RXi. In particular,

(3.3)RX.

Since R=RG(A) for some chief factor A of G, Section 2.3 yields

G/RLk

for some monolithic primitive group L and for some positive integer k. We let N denote the minimal normal subgroup (a.k.a. the socle) of L. From the definition of I and R we have I/R=soc(G/R)soc(Lk)=Nk. Finally, let T:=XI. In particular,

TR=XRIR.

We have

HD=(XY)D=X(YD)=XD=X(ID)=(XI)D=TD.

It follows

|G:HD|=|G:H||HD:H|=|G:H||D:HD|=|G:H||D:TD|.

If DT, then DX, and hence DXY=H because DY. However, this is a contradiction because D1 and hence, from (3.2), DH. Therefore, DT and |D:TD|>1.

Applying our inductive hypothesis, we obtain

(3.4)σ=max(HD/D,G/D)a|G/D:HD/D|3/2=a|G:HD|3/2=a(|G:H||D:DT|)3/2a23/2|G:H|3/2.

Moreover, when G is soluble and HD is a proper subgroup of G, we obtain

(3.5)σ=max(HD/D,G/D)|G/D:HD/D|-1=|G:HD|-1=|G:H||D:DT|-1|G:H|2-1.

(Observe that, when G is soluble and G=HD, we have σ=0, and hence the inequality σ|G:H|/2-1 is valid also in this degenerate case.)

From (3.3) we deduce ρmax(HR,G). If RH, then |G:HR|<|G:H| and hence, applying our inductive hypothesis, we obtain

(3.6)ρmax(HR,G)a|G:HR|3/2=a(|G:H||HR:H|)3/2a23/2|G:H|3/2.

Moreover, when G is soluble and HR is a proper subgroup of G, we obtain

(3.7)ρmax(HR,G)|G:HR|-1=|G:H||HR:H|-1|G:H|2-1.

(As above, when G is soluble and G=HR, we have ρ=0, and hence the inequality ρ|G:H|/2-1 is valid also in this degenerate case.)

Now, from (3.4) and (3.6), we have

max(H,G)=σ+ρ2a23/2|G:H|3/2<a|G:H|3/2.

Similarly, when G is soluble, from (3.5) and (3.7) we have

max(H,G)=σ+ρ|G:H|2-1+|G:H|2-1<|G:H|-1.

In particular, for the rest of the proof, we may assume that RH. Now, (3.2) yields R=1, and hence GLk and D=I. Therefore, we may identify G with Lk and D with Nk.

Set

𝒞:={coreG(Xi)i{1,,ρ}}

and, for every C𝒞, set

C:={Xii{1,,ρ},C=coreG(Xi)}.

For the rest of our argument for proving Theorems 1.2 and 1.3, we prefer to keep the proofs separate.

Proof of Theorem 1.2.

In this proof, we distinguish two cases.

Case 1: Suppose that N is non-abelian. Since N is non-abelian, the group G=Lk has exactly k minimal normal subgroups. We denote by N1,,Nk the minimal normal subgroups of G. In particular, I=Nk=N1×N2××Nk.

First, we claim that, for every i{1,,ρ}, there exist x,y{1,,k} such that NXi for every {1,,k}{x,y}, that is, Xi contains all but possibly at most two minimal normal subgroups of G.

We argue by induction on k. The statement is clearly true when k2. Suppose then k3 and let C:=coreG(Xi). If C=1, then Xi is a maximal core-free subgroup of G, and hence the action of G on the right cosets of Xi gives rise to a faithful primitive permutation representation. Since a primitive permutation group has at most two minimal normal subgroups [2, Theorem 4.4] and since G has exactly k minimal normal subgroups, we deduce that k2, which is a contradiction. Therefore, C1.

Since N1,,Nk are the minimal normal subgroups of Lk, we deduce that there exists {1,,k} with NC. Now, the proof of the claim follows applying the inductive hypothesis to G/NLk-1 and to its maximal subgroup Xi/N.

The previous claim shows that, for every C𝒞, C contains all but possibly at most two minimal normal subgroups of Nk=I. Therefore,

|𝒞|k2.

Let C𝒞 and let MC. The reader might find it useful to see Figure 1, where we have drawn a fragment of the subgroup lattice of G relevant to our argument.

Let k be the number of minimal normal subgroups of G contained in M. In particular, IMNk. Observe that IH is contained in IM and is core-free in G. Applying Lemma 2.1 (with H replaced by IH in a crowned-based group isomorphic to Lk), we get |IM:IH|5k. As kk-2, we deduce t5k-2.

Now, M/C is a core-free maximal subgroup of G/C. From Theorem 2.5, when C=coreG(M) and z=|G:C| are fixed, we have at most cz3/2 choices for M. As t5k-2, we have z|G:H|/5k-2. Thus

ρ=C𝒞|C|C𝒞z|G:H|z|G:H|/5k-2cz3/2ck2z|G:H|z|G:H|/5k-2z3/2=ck2(|G:H|5k-2)3/2z|G:H|z|G:H|/5k-2(5k-2z|G:H|)3/2.

Therefore,

z|G:H|z|G:H|/5k-2(5k-2z|G:H|)3/2u=11u3/2=a.

Finally, it is easy to verify that k2/53(k-2)/211 for every k. Summing up,

(3.8)ρ11ca|G:H|3/2.

From (3.4), (3.8) and from the definition of a, we have

max(H,G)=σ+ρa23/2|G:H|3/2+11ca|G:H|3/2=a|G:H|3/2.
Figure 1

Subgroup lattice for G.

Case 2: Suppose that N is abelian. As N is abelian, the action of L by conjugation on N endows N with the structure of an L-module. Since L is primitive, N is irreducible. Set q:=|EndL(N)|. Now, N is a vector space over the finite field 𝔽q with q elements, and hence |N|=qk for some positive integer k.

Let C𝒞 and let MC. By Lemma 2.2, CI. Now, the action of G/C on the right cosets of M/C is a primitive permutation group with point stabilizer M/C. Observe that in this primitive action, I/C is the socle of G/C. In particular, G/C acts irreducibly as a linear group on I/C, and hence C is a maximal L-submodule of I. Since I is the direct sum of k pairwise isomorphic irreducible L-modules, we deduce that we have at most (qk-1)/(q-1) choices for C. Moreover, |G:M|=|G/C:M/C|=|N|=qk. From Theorem 2.5, when C is fixed, we have at most c|G:M|3/2=c(qk)3/2 choices for MC. This yields

(3.9)ρ|𝒞|maxC𝒞|C|qk-1q-1cq3k/2<cqk+3k/2.

As we have observed above, MI=C is an L-submodule of G. Since an intersection of L-submodules is an L-submodule, we deduce that

HI=(X1Xρ)I

is an L-submodule of G and hence HIG. Since H is core-free in G, we deduce HI=1, and hence |I|=|N|k=qkk divides |G:H|. In particular, |G:H|qkk. Therefore, from (3.9) we obtain

ρc|G:H|k+3k/2kk.

When k1 or when (k,k)(2,1), we have k+3k/2kk32. When k=1, by refining (3.9), we obtain the sharper bound ρcq3k/2c|G:H|3/2. When (k,k)=(2,1), we may again refine (3.9):

ρc(q+1)q3/2c2qq3/2=2cq5/22c|G:H|5/42c|G:H|3/2.

Summing up, in all cases we have

(3.10)ρ2c|G:H|3/2.

From (3.4) and (3.10) we have

max(H,G)=σ+ρa23/2|G:H|3/2+2c|G:H|3/2<a|G:H|3/2,

as desired. ∎

The rest of the proof of Theorem 1.3 follows the same idea as in Case 2 above, but taking in account that the whole group G is soluble.

Proof of Theorem 1.3.

Since G=Lk and I=Nk, we may write G=IK, where K is a complement of N in L. As in the proof of Theorem 1.2 for the case that N is abelian, we have that the action of L by conjugation on N endows N with the structure of an L-module. Since L is primitive, N is irreducible. Set q:=|EndL(N)|. Now, N is a vector space over the finite field 𝔽q with q elements, and hence |N|=qk for some positive integer k.

Let C𝒞 and let MC. As we have observed above (for the proof of Case 2), MI=C is a maximal L-submodule of G, HI=1 and |I|=|N|k=qkk divides |G:H|. In particular, |G:H|=qkk for some positive integer .

Since G is soluble and since M is a maximal subgroup of G supplementing I, we have M=CKx for some maximal L-submodule C of I and some xI. Arguing as in the proof of Theorem 1.2 for the case that N is abelian, we deduce that we have at most (qk-1)/(q-1) choices for C. Moreover, we have at most |I/C|=|G:M|=|N|=qk choices for x. This yields

ρqk-1q-1qk.

Now, (3.5) gives σ|G:H|/|D:DT|-1: recall that D=I=Nk and DT=DH=IH=1. Thus σ|G:H|/|D|-1=|G:H|/qkk-1=-1. Therefore,

(3.11)max(H,G)=σ+ρ-1+qk-1q-1qk.

When 2, a computation shows that the right-hand side of (3.11) is less than or equal to

qkk-1=|G:H|-1.

In particular, we may suppose that =1. In this case, |G:H|=qkk=|I| and hence G=IH=IH. Moreover, σ=0. Since H is not a maximal subgroup of G (recall the base case for our inductive argument), k2.

Assume also k=1. Since |EndL(N)|=q=|N|, we deduce that L/N is isomorphic to a subgroup of the multiplicative group of the field 𝔽q and hence |L:N| is relatively prime to q. Therefore, |G:I| is relatively prime to q and hence so is |H|. Therefore, replacing H by a suitable G-conjugate, we may suppose that K=H. Using this information, we may now refine our earlier argument bounding ρ. Let C𝒞 and let MC. Since G=IH is soluble, M is a maximal subgroup of G supplementing I and HM, we have M=CH for some maximal L-submodule C of I. We deduce that we have at most (qk-1)/(q-1) choices for C and hence we have at most (qk-1)/(q-1) choices for M. This yields

max(H,G)=σ+ρ=ρqk-1q-1qk-1=|G:H|-1,

and the result is proved in this case.

Assume k2. A computation (using =1 and k, k2) shows that the right-hand side of (3.11) is less than or equal to qkk-1=|G:H|-1. ∎


Communicated by Manfred Droste


References

[1] A. Ballester-Bolinches and L. M. Ezquerro, Classes of Finite Groups, Math. Appl. 584, Springer, Dordrecht, 2006. Search in Google Scholar

[2] P. J. Cameron, Permutation Groups, London Math. Soc. Stud. Texts 45, Cambridge University, Cambridge, 1999. 10.1017/CBO9780511623677Search in Google Scholar

[3] P. J. Cameron, https://cameroncounts.wordpress.com/2016/11/28/road-closures-and-idempotent-generated-semigroups/. Search in Google Scholar

[4] I. De Las Heras and A. Lucchini, Intersections of maximal subgroups in prosolvable groups, Comm. Algebra 47 (2019), no. 8, 3432–3441. 10.1080/00927872.2019.1566918Search in Google Scholar

[5] E. Detomi and A. Lucchini, Crowns and factorization of the probabilistic zeta function of a finite group, J. Algebra 265 (2003), no. 2, 651–668. 10.1016/S0021-8693(03)00275-8Search in Google Scholar

[6] R. M. Guralnick, The dimension of the first cohomology group, Representation Theory. II (Ottawa 1984), Lecture Notes in Math. 1178, Springer, Berlin (1986), 94–97. 10.1007/BFb0075290Search in Google Scholar

[7] R. M. Guralnick, T. Hodge, B. Parshall and L. Scott, https://aimath.org/news/wallsconjecture/wall.conjecture.pdf. Search in Google Scholar

[8] P. Jiménez-Seral and J. P. Lafuente, On complemented nonabelian chief factors of a finite group, Israel J. Math. 106 (1998), 177–188. 10.1007/BF02773467Search in Google Scholar

[9] M. W. Liebeck, L. Pyber and A. Shalev, On a conjecture of G. E. Wall, J. Algebra 317 (2007), no. 1, 184–197. 10.1016/j.jalgebra.2006.10.047Search in Google Scholar

[10] A. Mann and A. Shalev, Simple groups, maximal subgroups, and probabilistic aspects of profinite groups, Israel J. Math. 96 (1996), no. part B, 449–468. 10.1007/BF02937317Search in Google Scholar

[11] C. E. Praeger and C. Schneider, Permutation Groups and Cartesian Decompositions, London Math. Soc. Lecture Note Ser. 449, Cambridge University, Cambridge, 2018. 10.1017/9781139194006Search in Google Scholar

[12] G. E. Wall, Some applications of the Eulerian functions of a finite group, J. Aust. Math. Soc. 2 (1961/1962), 35–59. 10.1017/S1446788700026367Search in Google Scholar

Received: 2019-08-19
Published Online: 2020-02-18
Published in Print: 2020-05-01

© 2021 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 26.4.2024 from https://www.degruyter.com/document/doi/10.1515/forum-2019-0222/html
Scroll to top button