Hostname: page-component-76fb5796d-5g6vh Total loading time: 0 Render date: 2024-04-26T18:38:07.374Z Has data issue: false hasContentIssue false

FINITE GROUPS WITH INDEPENDENT GENERATING SETS OF ONLY TWO SIZES

Published online by Cambridge University Press:  23 February 2023

ANDREA LUCCHINI
Affiliation:
Dipartimento di Matematica, ‘Tullio Levi-Civita’, University of Padova, Via Trieste 53, 35121 Padova, Italy e-mail: lucchini@math.unipd.it
PABLO SPIGA*
Affiliation:
Dipartimento di Matematica Pura e Applicata, University of Milano-Bicocca, Via Cozzi 55, 20126 Milano, Italy
Rights & Permissions [Opens in a new window]

Abstract

A generating set S for a group G is independent if the subgroup generated by $S\setminus \{s\}$ is properly contained in G for all $s \in S.$ We describe the structure of finite groups G such that there are precisely two numbers appearing as the cardinalities of independent generating sets for G.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1 Introduction

The minimal number of generators of a finite group G is denoted by $d(G).$ A generating set S for a group G is independent (sometimes called irredundant) if

$$ \begin{align*}\langle S\setminus \{s\}\rangle < G \quad\mbox{for all } s \in S.\end{align*} $$

Let $m(G)$ denote the maximal size of an independent generating set for $G.$

The finite groups with $m(G)=d(G)$ are classified by Apisa and Klopsch.

Theorem 1.1 (Apisa–Klopsch, [Reference Apisa and Klopsch1, Theorem 1.6]).

If $d(G)=m(G)$ , then G is soluble. Moreover, either

  • $G/\operatorname {\mathrm {Frat}}(G)$ is an elementary abelian p-group for some prime p; or

  • $G/\operatorname {\mathrm {Frat}}(G)=PQ,$ where P is an elementary abelian p-group and Q is a nontrivial cyclic q-group for distinct primes p and $q,$ such that Q acts by conjugation faithfully on P and P (viewed as a module for Q) is a direct sum of $m(G)-1$ isomorphic copies of one simple Q-module.

In view of this result, Apisa and Klopsch suggest a natural ‘classification problem’: given a nonnegative integer c, characterise all finite groups G which satisfy $m(G) - d(G) \leq c$ . The particular case $c=1$ has been recently highlighted by Glasby (see [Reference Glasby7, Problem 2.3]).

A nice result in universal algebra, due to Tarski and known as the Tarski irredundant basis theorem (see for example [Reference Burris and Sankappanavar3, Theorem 4.4]), implies that, for every positive integer k with $d(G)\leq k\leq m(G), G$ contains an independent generating set of cardinality  $k.$ So the condition $m(G) - d(G)= 1$ is equivalent to the fact that there are only two possible cardinalities for an independent generating set of $G.$

Let G be a finite group. We recall that the socle of G, denoted $\operatorname {\mathrm {soc}}(G)$ , is the subgroup generated by the minimal normal subgroups of G; moreover, G is said to be monolithic primitive if G has a unique minimal normal subgroup and the Frattini subgroup $\operatorname {\mathrm {Frat}}(G)$ of G is the identity.

In this paper, we prove the following two main results.

Theorem 1.2. Let G be a finite group with $\operatorname {\mathrm {Frat}}(G)=1$ and $m(G)=d(G)+1.$ If G is not soluble, then $d(G)=2$ , G is a monolithic primitive group and $G/\operatorname {\mathrm {soc}}(G)$ is cyclic of prime power order.

It was proved by Whiston and Saxl [Reference Whiston and Saxl15] that $m(\operatorname {\mathrm {PSL}}(2,p))=3$ for any prime p with p not congruent to $\pm 1$ modulo 8 or 10. In particular, as $d(S)=2$ for every nonabelian simple group, we deduce that there are infinitely many nonabelian simple groups G with $m(G)=d(G)+1$ . We also give examples of nonsimple groups G having ${m(G)=d(G)+1}$ in Section 4.

Theorem 1.3. Let G be a finite group with $\operatorname {\mathrm {Frat}}(G)=1$ and $m(G)=d(G)+1.$ If G is soluble, then one of the following occurs:

  1. (1) $G\cong V\rtimes P$ , where P is a finite noncyclic p-group and V is an irreducible P-module, which is not a p-group; in this case, $d(G)=d(P)$ ;

  2. (2) $G\cong V^t \rtimes H,$ where V is a faithful irreducible H-module, $m(H)=2$ and either $t=1$ or H is abelian; in this case, $d(G)=t+1$ ;

  3. (3) there exist two normal subgroups $N_1$ , $N_2$ such that $1 \lneq N_1 \leq N_2, N_1$ is an abelian minimal normal subgroup of $G, N_2/N_1\leq \operatorname {\mathrm {Frat}}(G/N_1)$ and $G/N_2\cong V^t\rtimes H$ , where V is an irreducible H-module and H is a nontrivial cyclic group of prime power order; in this case, $d(G)=t+1.$

In Section 4, we give examples of finite soluble groups G with $m(G)=d(G)+1$ for each of the three possibilities arising in Theorem 1.3.

2 Preliminary results

Let L be a monolithic primitive group and let A be its unique minimal normal subgroup. For each positive integer k, let $L^k$ be the k-fold direct product of L. The crown-based power of L of size k is the subgroup $L_k$ of $L^k$ defined by

$$ \begin{align*}L_k:=\{(l_1, \ldots , l_k) \in L^k \mid l_1 \equiv \cdots \equiv l_k \ {\mbox{mod}}\, A \}.\end{align*} $$

In [Reference Dalla Volta and Lucchini4], it is proved that for every finite group G, there exists a monolithic group L and a homomorphic image $L_k$ of G such that

  1. (1) $ d(L/\operatorname {\mathrm {soc}} L) < d(G) $ ; and

  2. (2) $d(L_k) =d(G).$

A group $L_{k}$ with this property is called a generating crown-based power for G.

In [Reference Dalla Volta and Lucchini4], it is explained how $d(L_{k})$ can be explicitly computed in terms of k and the structure of L. A key ingredient (when one wants to determine $d(G)$ from the behaviour of the crown-based power homomorphic images of G) is to evaluate, for each monolithic group L, the maximal k such that $L_{k}$ is a homomorphic image of G. This integer k arises from an equivalence relation among the chief factors of G. In what follows, we give some details.

Given groups G and A, we say that A is a G-group if G acts on A via automorphisms. In addition, A is irreducible if G does not stabilise any nontrivial proper subgroups of A. Two G-groups A and B are G-isomorphic if there exists a group isomorphism $\phi : A\to B$ such that $\phi (g(a))=g(\phi (a))$ for all $a\in A$ and $g\in G.$ Following [Reference Jiménez-Seral and Lafuente8], we say that two irreducible G-groups A and B are G-equivalent, denoted $A \sim _G B$ , if there is an isomorphism $\Phi : A\rtimes G \rightarrow B\rtimes G$ which restricts to a G-isomorphism $\phi \colon A \to B$ and induces the identity $G \cong AG/A \to BG/B \cong G$ , in other words, such that the following diagram commutes:

Observe that two G-isomorphic G-groups are G-equivalent, and the converse holds if A and B are abelian.

Let $A=X/Y$ be a chief factor of G. A complement U of A in G is a subgroup of G such that

$$ \begin{align*} UX=G \quad\text{and}\quad U \cap X=Y. \end{align*} $$

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 $\delta _G(A)$ of non-Frattini chief factors that are G-equivalent to A, in any chief series of G, does not depend on the particular choice of such a series.

Now, we denote by $L_G(A)$ the monolithic primitive group associated to A, that is,

$$ \begin{align*}L_G(A):= \begin{cases} A\rtimes (G/\mathbf{C}_G(A)) & \text{ if }A\text{ is abelian}, \\ G/\mathbf{C}_G(A)& \text{ otherwise}. \end{cases} \end{align*} $$

If A is a non-Frattini chief factor of G, then $L_G(A)$ is a homomorphic image of G. More precisely, there exists a normal subgroup N such that $G/N \cong L_G(A)$ and $\operatorname {\mathrm {soc}} (G/N) \sim _G~A$ . We identify $\operatorname {\mathrm {soc}}( L_G(A))$ with A, as G-groups.

Consider now all the normal subgroups N of G with the property that ${G/N \cong L_G(A)}$ and $\operatorname {\mathrm {soc}} (G/N) \sim _G A$ . The intersection $R_G(A)$ of all these subgroups has the property that $G/R_G(A)$ is isomorphic to the crown-based power $(L_G(A))_{\delta _G(A)}$ . The socle $I_G(A)/R_G(A)$ of $G/R_G(A)$ is called the A-crown of G and it is a direct product of $\delta _G(A)$ minimal normal subgroups G-equivalent to A.

Note that if L is monolithic primitive and $L_k$ is a homomorphic image of G for some $k\geq 1$ , then $L \cong L_G(A)$ for some non-Frattini chief factor A of G and $k \leq \delta _G(A)$ . Furthermore, if $(L_G(A))_k$ is a generating crown-based power, then so is $(L_G(A))_{\delta _G(A)}$ ; in this case, we say that A is a generating chief factor for G.

For an irreducible G-module M, set

$$ \begin{align*} r_G(M)&:=\dim_{\operatorname{\mathrm{End}}_G(M)}M,\\ s_G(M)&:=\dim_{\operatorname{\mathrm{End}}_G(M)} H^1(G,M), \\ t_G(M)&:=\dim_{\operatorname{\mathrm{End}}_G(M)} H^1(G/\mathbf{C}_G(M),M). \end{align*} $$

It can be seen that

$$ \begin{align*}s_G(M)=t_G(M)+\delta_G(M)\end{align*} $$

(see for example [Reference Lucchini10, 1.2]). Now, define

$$ \begin{align*}h_{G}(M):= \begin{cases} \delta_G(M)&\text{if }M\text{ is a trivial } G\textrm{-module}, \\ \displaystyle \bigg\lfloor\frac{s_G(M)-1}{r_G(M)}\bigg\rfloor+2=\bigg\lfloor\frac{\delta_G(M)+t_G(M)-1}{r_G(M)}\bigg\rfloor+2& \text{otherwise}. \end{cases} \end{align*} $$

By [Reference Aschbacher and Guralnick2, Theorem A], $t_G(M)<r_G(M)$ for any irreducible G-module M, and therefore

(2.1) $$ \begin{align}h_G(M)\leq \delta_G(M)+1. \end{align} $$

The importance of $h_G(M)$ is clarified by the following proposition.

Proposition 2.1 [Reference Detomi and Lucchini6, Proposition 2.1].

If there exists an abelian generating chief factor A of G, then $ d(G)=h_G(A)$ .

When G admits a nonabelian generating chief factor A, a relation between $\delta _G(A)$ and $d(G)$ is provided by the following result.

Proposition 2.2. If $d(G)\geq 3$ and there exists a nonabelian generating chief factor A of G, then

$$ \begin{align*}\delta_G(A)>\frac{|A|^{d(G)-1}}{2|\mathbf{C}_{\operatorname{\mathrm{Aut}} A}(L_G(A)/A)|}\geq \frac{|A|^{d(G)-2}}{2\log_2|A|}. \end{align*} $$

Proof. Suppose that $d(G)\ge 3$ and let A be a nonabelian generating chief factor of G.

For a finite group X, let $\phi _X(m)$ denote the number of ordered m-tuples $(x_1,\ldots ,x_m)$ of elements of X generating X. Define

$$ \begin{align*} L&:=L_G(A),\\ \gamma&:=|\mathbf{C}_{\operatorname{\mathrm{Aut}} A}(L/A)|,\\ \delta&:=\delta_G(A),\\ d&:=d(G). \end{align*} $$

In [Reference Dalla Volta and Lucchini4], it is proved that if $m\geq d(L),$ then

(2.2) $$ \begin{align}d(L_k)\leq m \quad {\text {if and only if}} \quad k\leq \frac{\phi_{L/A}(m)}{\phi_L(m)\gamma}. \end{align} $$

By the main result in [Reference Lucchini and Menegazzo13], $d(L)=\max (2,d(L/A))$ . Since A is a generating chief factor, from the definition, we have $d(L/A) < d(L_{\delta _G(A)})=d(G)$ . As $2 < d(G),$ it follows ${d(L)<d(G).}$ Now, by applying (2.2) with $k=\delta _G(A)$ and $m=d(G)-1,$ we deduce that

(2.3) $$ \begin{align}\delta_G(A)>\frac{\phi_{L/A}(d(G)-1)}{\phi_L(d(G)-1)\gamma}.\end{align} $$

By [Reference Detomi and Lucchini6, Corollary 1.2],

(2.4) $$ \begin{align}\frac{\phi_{L/A}(d(G)-1)}{\phi_L(d(G)-1)}\geq \frac{|A|^{d(G)-1}}{2}.\end{align} $$

Moreover, $A\cong S^n$ , where n is a positive integer and S is a nonabelian simple group. In the proof of Lemma 1 in [Reference Dalla Volta and Lucchini5], it is shown that

$$ \begin{align*}\gamma \leq n|S|^{n-1}|{\operatorname{\mathrm{Aut}}}(S)|.\end{align*} $$

Now, [Reference Kohl9] shows that $|{\operatorname {\mathrm {Out}}}(S)| \leq \log _2(|S|)$ and hence

(2.5) $$ \begin{align}\gamma\leq n|S|^{n}\log_2(|S|)\leq |S|^{n}\log_2(|S|^n)=|A|\log_2(|A|).\end{align} $$

From (2.3), (2.4) and (2.5), we obtain

$$ \begin{align*} \delta_G(A)> \frac{\phi_{L/A}(d(G)-1)}{\phi_L(d(G)-1)\gamma}\geq \frac{|A|^{d(G)-1}}{2|A|\log_2|A|}=\frac{|A|^{d(G)-2}}{2\log_2|A|}.\\[-42pt] \end{align*} $$

Recall that $m(G)$ is the largest cardinality of an independent generating set of G.

Theorem 2.3 [Reference Lucchini and Spiga14, Theorem 1.3].

Let G be a finite group. Then $m(G)\geq a+b,$ where a and b are, respectively, the number of non-Frattini and nonabelian factors in a chief series of G. Moreover, if G is soluble, then $m(G)=a.$

Corollary 2.4. Assume that G is a finite group with a unique minimal normal subgroup A. If A is nonabelian, then $m(G)\geq 3.$

Proof. Suppose first that G is simple. Let l be an element of G of order 2. Since $G=\langle l^x\mid x\in G\rangle $ , the set $\{l ^x\mid x \in G\}$ contains a minimal generating set of $G.$ Since G cannot be generated by two involutions, this minimal generating set has cardinality at least three. Thus, $m(G)\ge 3$ .

Suppose next that G is not simple. Let a and b be the number of non-Frattini and nonabelian factors in a chief series of G. As G is not simple, there exists a maximal normal subgroup N of G containing A and we have a chief series $1\unlhd N_1\unlhd \cdots \unlhd N_{t-1}\unlhd N_t=G$ with $N_1=A$ and $N_{t-1}=N.$ Then, $a\geq 2$ , $b\geq 1$ and $m(G)\geq a+b\geq 3$ by Theorem 2.3.

3 Proof of the main results

Let G be a finite group, let $d:=d(G)$ and let $m:=m(G)$ . Suppose that $m=d+1.$ Let A be a generating chief factor of G and let $\delta :=\delta _G(A)$ , $L:=L_G(A).$

3.1 A is nonabelian

First, suppose that $\delta \geq 2.$ By Theorem 2.3, $m\geq 2\delta $ and therefore $d\geq 2\delta -1\geq 3.$ By Proposition 2.2,

$$ \begin{align*}\delta> \frac{|A|^{d-2}}{2\log_2|A|}\geq \frac{ |A|^{2\delta-3}}{2\log_2|A|}\geq \frac{60^{2\delta-3}}{2\log_260},\end{align*} $$

but this is never true.

Suppose now that $\delta =1$ . In this case, by the main theorem in [Reference Lucchini and Menegazzo13], $d=d(L)=\max (2,d(L/A))=2$ and therefore $m=3.$ Since L is an epimorphic image of G, we must have $m(L)\leq 3$ . However, $m(L)\geq 3$ by Corollary 2.4. Hence, $m(L)=m=3$ and therefore it follows from [Reference Lucchini11, Lemma 11] that $G/\operatorname {\mathrm {Frat}}(G)\cong L.$ Finally, by Theorem 2.3, $m(L)=3$ implies $m(L/A)\leq 1,$ and this is possible only if $L/A$ is a cyclic p-group. This concludes the proof of Theorem 1.2.

3.2 A is abelian

It follows from Proposition 2.1 and (2.1) that

$$ \begin{align*}\delta-1 \leq m-1 = d = h_G(A)\leq \delta+1.\end{align*} $$

If $d=\delta -1,$ then $m=\delta $ and this is possible if $G/\operatorname {\mathrm {Frat}}(G)\cong A^\delta .$ However, in this case, A would be a trivial G-module and therefore $d=h_G(A)=\delta =m,$ which is a contradiction.

Now suppose that $d=\delta .$ By Theorem 2.3, G is soluble and contains only one non-Frattini chief factor which is not G-isomorphic to $A.$ If A is noncentral in G, then $G/\operatorname {\mathrm {Frat}}(G)\cong L_\delta $ and $L/A$ is a cyclic p-group. However, this implies $r_G(A)\,{=}\,1,\ t_G(A)\,{=}\,0$ and $d=h_G(A)=\delta +1,$ which is a contradiction. If A is central, then $G/\operatorname {\mathrm {Frat}}(G)\cong V\rtimes P$ , where P is a finite p-group, V is an irreducible P-module and $d(P)=d.$ In particular, we obtain item (1) in Theorem 1.3.

Finally assume $d=\delta +1.$ Notice that in this case, $L=A\rtimes H,$ where A is a faithful, nontrivial, irreducible H-module, and

$$ \begin{align*}m(H)\leq m-\delta=\delta+2-\delta=2.\end{align*} $$

In particular, by Corollary 2.4, H is soluble.

If $m(H)=2,$ then $G/\operatorname {\mathrm {Frat}}(G)\cong L_\delta .$ In particular, we obtain item (2) in Theorem 1.3.

If $m(H)=1,$ then there exist two normal subgroups $N_1$ and $N_2$ of G such that $1 \lneq N_1 \leq N_2, G/N_2\cong L_\delta , N_2/N_1\leq \operatorname {\mathrm {Frat}}(G/N_1)$ and $N_1/\operatorname {\mathrm {Frat}}(G)$ is an abelian minimal normal subgroup of $G/\operatorname {\mathrm {Frat}}(G).$ As $m(H)=1$ , H is cyclic of prime power order. In particular, we obtain item (3) in Theorem 1.3.

4 Examples for Theorems 1.2 and 1.3

4.1 Monolithic groups: examples for Theorem 1.2

Let G be monolithic primitive with nonabelian socle $N= S_1\times \cdots \times S_n$ , with $S\cong S_i$ for each $1\leq i \leq n.$ The number $\mu (G)=m(G)-m(G/N)$ has been investigated in [Reference Lucchini12]. The group G acts by conjugation on the set $\{S_1,\ldots ,S_n\}$ of the simple components of $N.$ This produces a group homomorphism $G\to \operatorname {\mathrm {Sym}}(n)$ and the image K of G under this homomorphism is a transitive subgroup of $\operatorname {\mathrm {Sym}}(n).$ Moreover, the subgroup X of $\operatorname {\mathrm {Aut}} S$ induced by the conjugation action of $\textbf {N}_G(S_1)$ on the first factor $S_1$ is an almost simple group with socle $S.$

By [Reference Lucchini12, Proposition 4], $\mu (G)\geq \mu (X)=m(X)-m(X/S).$ Assume $m(G)=3.$ Observe that by Theorems 1.1 and 1.2, $G/N$ is cyclic of prime power order. If $X=S$ , then

$$ \begin{align*} 3=m(G)=m(G/N)+\mu(G)& \geq m(G/N)+\mu(X)= m(G/N)+m(S)\\ &\geq m(G/N)+3. \end{align*} $$

This implies that $G/N=1$ and $G=S$ is a simple group. If $X\neq S$ , then $G\neq N$ and

$$ \begin{align*}3=m(G)\geq m(G/N)+\mu(G)\geq 1+\mu(X).\end{align*} $$

Moreover, $X/S$ is a nontrivial cyclic group of prime power order, so

$$ \begin{align*}m(X)=m(X/S)+\mu(X)\leq 1+\mu(X)\leq 1+2=3.\end{align*} $$

By Corollary 2.4, $m(X)=3.$

The groups

$$ \begin{align*}\mathrm{P}\Sigma\mathrm{L}_2(9), M_{10},\mathrm{Aut}(\mathrm{PSL}_2(7))\end{align*} $$

are currently the only known examples (to the best knowledge of the authors) of almost simple groups X with $X\neq \operatorname {\mathrm {soc}}(X)$ and $m(X)=3.$ We believe that there are other such examples, but our current computer codes are not efficient enough to carry out a thorough investigation.

Let $S:=\mathrm {PSL}_2(7)$ and $H:=\mathrm {Aut}(\mathrm {PSL}_2(7)),$ or let $S:=\mathrm {PSL}_2(9)$ and $H\in \{\mathrm {P}\Sigma \mathrm {L}_2(9), M_{10}\}$ . Consider the wreath product $W:=H \wr \operatorname {\mathrm {Sym}}(n)$ . Any element $w\in W$ can be written as $w=\pi (a_1,\ldots ,a_n),$ with $\pi \in \operatorname {\mathrm {Sym}}(n)$ and $a_i\in H$ for $1\leq i\leq n$ . In particular, $N=\operatorname {\mathrm {soc}}(W) =S_1\times \cdots \times S_n=\{(s_1,\ldots ,s_n)\mid s_i \in S\}$ .

Proposition 4.1. Let G be the subgroup of W generated by $N=\operatorname {\mathrm {soc}}(W)$ and $\gamma =\sigma (a,1,\ldots ,1),$ where $\sigma =(1\,2\cdots n)\in \operatorname {\mathrm {Sym}}(n)$ and $a\in H\setminus S.$ If $n=2^t$ for some positive integer $t,$ then $m(G)=3.$

In particular, this gives infinitely many examples of nonsimple, nonsoluble groups G with $m(G)=d(G)+1$ in Theorem 1.2.

Proof. Suppose that $n=2^t$ for some positive integer t. Let $r:=m(G)$ ; we aim to prove that $r=3$ .

Let $\{g_1,\ldots ,g_r\}$ be an independent generating set of $G.$ Observe that

$$ \begin{align*}\gamma^n=(a,\ldots,a)\in G\setminus N\end{align*} $$

and hence $G/N$ is cyclic of order $2^{t+1}.$ Therefore, relabelling the elements of the independent generating set if necessary, we may assume $G=\langle g_1, N\rangle $ . Hence, ${g_1=\sigma (as_1,s_2,\ldots ,s_n)}$ with $s_1,\ldots ,s_n\in S.$ Moreover, for $2\leq i\leq r,$ there exists $u_i\in \mathbb Z$ such that $g_ig_1^{u_i}\in N$ . Observe that $\{g_1,g_2g_1^{u_2},\ldots ,g_rg_1^{u_r}\}$ is still an independent generating set having cardinality r.

Let

$$ \begin{align*}m=(s_2\cdots s_n, s_3\cdots s_n,\ldots,s_{n-1}s_n,s_n,1) \in N.\end{align*} $$

Then, $Y=\{g_1^m,(g_2g_1^{u_2})^m,\ldots ,(g_rg_1^{u_r})^m\}$ is another independent generating set for G having cardinality r. We have

$$ \begin{align*}y_1:=g_1^m=\sigma(b,1,\ldots,1),\end{align*} $$

with $b=as_1\cdots s_n \in \operatorname {\mathrm {Aut}} S\setminus S,$ and for $2\leq i\leq r,$ there exist $s_{i1},\ldots s_{in}\in S$ such that

$$ \begin{align*}y_i:=(g_ig_1^{u_i})^m=(s_{i1},\ldots,s_{in}).\end{align*} $$

Let $Z:=\{b, s_{ij} \mid 2\leq i\leq r, 1\leq j\leq n\}$ and $T=\langle Z\rangle .$ Since $G=\langle y_1,\ldots ,y_t\rangle \leq T \wr \langle \sigma \rangle ,$ we must have $\operatorname {\mathrm {Aut}}(S)=T.$ However, $m(\operatorname {\mathrm {Aut}}(S))=3,$ so $\operatorname {\mathrm {Aut}}(S)=\langle b, s_{iu}, s_{jv}\rangle $ for suitable $2\leq i,j\leq r$ and $2\leq u, v\leq n.$

Let $H:=\langle y_1, y_i, y_j\rangle $ and, for $1\leq k\leq n,$ consider the projection $\pi _k: N\to S$ sending $(s_1,\ldots ,s_n)$ to $s_k.$ Notice that $\pi _1(y_1^n)=b, \pi _1((y_i)^{y_1^{1-u}})=s_{iu}, \pi _1((y_j)^{y_1^{1-v}})=s_{jv}.$ In particular, $\pi _1(H\cap N)=S$ and $H\cap N$ is a subdirect product of $N= S_1\times \cdots \times S_n$ .

Recall that a subgroup D of $N=S_1\times \cdots \times S_n$ is said to lie fully diagonally in N if each projection $\pi _i: D\to S_i$ is an isomorphism. To each pair $(\Phi ,\alpha )$ , where ${\Phi =\{B_1,\ldots ,B_c\}}$ is a partition of the set $\{1,\ldots ,n\}$ and $\alpha =(\alpha _1,\ldots ,\alpha _n)\in (\operatorname {\mathrm {Aut}} S)^n$ , we associate a direct product $\Delta (\Phi ,\alpha )=D_1\times \cdots \times D_c$ , where each factor $D_j=\{(x^{\alpha _{i_1}},\ldots , x^{\alpha _{i_d}})\mid x\in S\}$ is a full diagonal subgroup of the direct product $S_{i_1} \times \cdots \times S_{i_d}$ corresponding to the block $B_j=\{i_1,\ldots ,i_d\}$ in $\Phi .$

Since $H\cap N$ is a subdirect product of $N,$ we must have $H\cap N=\Delta (\Phi ,\alpha )$ for a suitable choice of the pair $(\Phi ,\alpha ).$ As $G=\langle H,N\rangle $ , the action by conjugation of H on $\{S_1,\ldots ,S_n\}$ is transitive and hence the partition $\{B_1,\ldots ,B_c\}$ corresponds to an imprimitive system for the permutation action of $\langle \sigma \rangle $ on $\{1,\ldots ,n\}.$ So there exist $c=2^\gamma $ and $d=2^\delta $ with $c\cdot d=n$ such that

$$ \begin{align*}B_i:=\{i,i+c,i+2c,\ldots,i+(d-1)c\} \quad\mbox{for } 1\leq i \leq c.\end{align*} $$

Notice that $y_1\in H$ normalises $\Delta (\Phi ,\alpha )$ . In particular, $y_1^c$ normalises $\Delta (\Phi ,\alpha ).$ However, $y_1^c$ normalises $L=S_1\times S_{1+c}\times \cdots \times S_{1+(d-1)c}$ and acts on L as $\pi \cdot l,$ where $\pi $ is the d-cycle $(1,1+c,\ldots ,1+(d-1)c)$ and $l=(b,1,\ldots ,1)\in L.$ In particular, $\pi \cdot l$ normalises the full diagonal subgroup $D_1$ of L. Therefore, setting $\phi _i=\alpha _{1+(i-1)c}$ , for every $s\in S$ , there exists $t\in T$ such that

$$ \begin{align*}(s^{\phi_db},s^{\phi_1},s^{\phi_2},\ldots,s^{\phi_{d-1}})= (t^{\phi_1},t^{\phi_2},t^{\phi_3 },\ldots,t^{\phi_{d}}).\end{align*} $$

It follows that

$$ \begin{align*}\begin{aligned} \phi_{d}b\phi_1^{-1}\phi_2&=\phi_1,\\ \phi_{d}b\phi_1^{-1}\phi_3&=\phi_{2},\\ \cdots\\ \phi_{d}b\phi_1^{-1}\phi_d&=\phi_{d-1}. \end{aligned}\end{align*} $$

In particular, $(\phi _1\phi _d^{-1})^d\equiv b^{d-1}$ modulo $S.$ If d is even, then $b\in \langle x^2\mid x\in \operatorname {\mathrm {Aut}}(S)\rangle =S,$ against our assumption. Thus, $d=1$ and hence $c=n$ . However, this implies that ${H\cap N=N}$ and consequently $H=G.$ Thus, $m(G)=r\leq 3$ . However, $m(G)\geq 3$ by Theorem 2.3. So we conclude that $m(G)=3.$

4.2 Soluble groups: examples for Theorem 1.3

We give three elementary examples, but with the same ideas, one can construct more complicated examples. Let $S_n$ be the symmetric group of degree n and let $C_n$ be the cyclic group of order n.

The group $G:=S_3 \times C_2^t = C_3 : C_2^{t+1}$ with $t\ge 1$ satisfies $d(G)=t+1$ and ${m(G)=t+2}$ . This gives examples of groups satisfying item (1) in Theorem 1.3.

The group $G:=S_4=K:S_3$ with K the Klein subgroup of $S_4$ and the group ${G:=(C_3^t : C_2) \times C_2}$ with $C_2$ acting on $C_3^t$ by inversion also satisfy $m(G)=d(G)+1$ . These two examples yield groups satisfying item (2) in Theorem 1.3 with $m(H)=2$ in the first case and with H abelian in the second case.

As above, let K be the Klein subgroup of $S_4$ and let $G:=K:(S_3\times C_2^{t-1})$ . This gives examples of groups satisfying item (3) in Theorem 1.3.

References

Apisa, P. and Klopsch, B., ‘A generalization of the Burnside basis theorem’, J. Algebra 400 (2014), 816.10.1016/j.jalgebra.2013.11.005CrossRefGoogle Scholar
Aschbacher, M. and Guralnick, R., ‘Some applications of the first cohomology group’, J. Algebra 90(2) (1984), 446460.10.1016/0021-8693(84)90183-2CrossRefGoogle Scholar
Burris, S. and Sankappanavar, H. P., A Course in Universal Algebra, Graduate Texts in Mathematics, 78 (Springer-Verlag, New York–Berlin, 1981).10.1007/978-1-4613-8130-3CrossRefGoogle Scholar
Dalla Volta, F. and Lucchini, A., ‘Finite groups that need more generators than any proper quotient’, J. Aust. Math. Soc. Ser. A 64 (1998), 8291.10.1017/S1446788700001312CrossRefGoogle Scholar
Dalla Volta, F. and Lucchini, A., ‘The smallest group with non-zero presentation rank’, J. Group Theory 2(2) (1999), 147155.Google Scholar
Detomi, E. and Lucchini, A., ‘Probabilistic generation of finite groups with a unique minimal normal subgroup’, J. Lond. Math. Soc. (2) 87(3) (2013), 689706.10.1112/jlms/jds076CrossRefGoogle Scholar
Glasby, S. P., ‘Classifying uniformly generated groups’, Comm. Algebra 48(1) (2020), 101104.10.1080/00927872.2019.1632333CrossRefGoogle Scholar
Jiménez-Seral, P. and Lafuente, J., ‘On complemented nonabelian chief factors of a finite group’, Israel J. Math. 106 (1998), 177188.10.1007/BF02773467CrossRefGoogle Scholar
Kohl, S., ‘A bound on the order of the outer automorphism group of a finite simple group of given order’, online note (2003), available at http://www.gap-system.org/DevelopersPages/StefanKohl/preprints/outbound.pdf.Google Scholar
Lucchini, A., ‘Generating wreath products and their augmentation ideals’, Rend. Semin. Mat. Univ. Padova 98 (1997), 6787.Google Scholar
Lucchini, A., ‘The largest size of a minimal generating set of a finite group’, Arch. Math. (Basel) 101(1) (2013), 18.10.1007/s00013-013-0527-yCrossRefGoogle Scholar
Lucchini, A., ‘Minimal generating sets of maximal size in finite monolithic groups’, Arch. Math. (Basel) 101(5) (2013), 401410.10.1007/s00013-013-0583-3CrossRefGoogle Scholar
Lucchini, A. and Menegazzo, F., ‘Generators for finite groups with a unique minimal normal subgroup’, Rend. Semin. Mat. Univ. Padova 98 (1997), 173191.Google Scholar
Lucchini, A. and Spiga, P., ‘Independent sets of generators of prime power order’, Expo. Math. 40(1) (2022), 140154.10.1016/j.exmath.2021.06.003CrossRefGoogle Scholar
Whiston, J. and Saxl, J., ‘On the maximal size of independent generating sets of ${\mathrm{PSL}}_2(q)$ ’, J. Algebra 258(2) (2002), 651657.10.1016/S0021-8693(02)00648-8CrossRefGoogle Scholar