Hostname: page-component-7bb8b95d7b-495rp Total loading time: 0 Render date: 2024-09-27T02:22:09.875Z Has data issue: false hasContentIssue false

GROUPS WITH FEW NONPOWER SUBGROUPS

Published online by Cambridge University Press:  10 August 2023

JIWEI ZHENG
Affiliation:
School of Mathematics and Statistics, Southwest University, Chonqing 400715, P.R. China e-mail: 1179157500@qq.com
WEI ZHOU*
Affiliation:
School of Mathematics and Statistics, Southwest University, Chonqing 400715, P.R. China
D. E. TAYLOR
Affiliation:
School of Mathematics and Statistics, The University of Sydney, New South Wales 2006, Australia e-mail: Donald.Taylor@sydney.edu.au
Rights & Permissions [Opens in a new window]

Abstract

For a group G and $m\ge 1$, let $G^m$ denote the subgroup generated by the elements $g^m$, where g runs through G. The subgroups not of the form $G^m$ are the nonpower subgroups of G. We classify the groups with at most nine nonpower subgroups.

Type
Research Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1 Introduction

For a group G and $m\ge 1$ , the power subgroup $G^m$ is the subgroup generated by the elements $g^m$ , where g runs through G. A subgroup that is not a power subgroup is a nonpower subgroup. Let $\operatorname {\mathrm {ps}}(G)$ and $\operatorname {\mathrm {nps}}(G)$ denote the number of power and nonpower subgroups of G. It is immediate that every power subgroup is a characteristic subgroup of G. But the converse is false, as illustrated by $M_{n,p}$ defined in Section 3: it has a unique maximal noncyclic subgroup, which is characteristic but not a power subgroup.

The study of nonpower subgroups was initiated by Szász [Reference Szász7] who proved that G is cyclic if and only if $\operatorname {\mathrm {nps}}(G) = 0$ . The terminology ‘nonpower subgroup’ was introduced by Zhou et al. [Reference Zhou, Shi and Duan9]. They proved that a noncyclic group G is finite if and only if $\operatorname {\mathrm {nps}}(G)$ is finite. Furthermore, if G is a finite noncyclic group, it was proved by Zhou and Ping that $\operatorname {\mathrm {nps}}(G)\ge 3$ . Therefore, from now on, we assume that all groups under consideration are finite.

For the most part, our notation follows Gorenstein [Reference Gorenstein6]. In particular, $\Phi (G)$ denotes the Frattini subgroup of G, and for subgroups H and K, $[H,K]$ is generated by the commutators $[x,y] = x^{-1}y^{-1}xy$ with $x\in H$ and $y\in K$ . For a finite p-group G, let $\Omega _{i}(G)$ be the subgroup $\langle x\in G\mid x^{p^{i}}=1 \rangle $ for $i \ge 1$ . We use [Reference Gorenstein6, Ch. 5] as a reference for standard results about p-groups.

Anabanti et al. [Reference Anabanti, Aroh, Hart and Oodo1, Reference Anabanti and Hart2] classified the groups G with $\operatorname {\mathrm {nps}}(G)\in \{3,4\}$ and showed that, for all $k> 4$ , there are infinitely many groups G with $\operatorname {\mathrm {nps}}(G) = k$ .

The following theorem extends the classification to the groups G with $\operatorname {\mathrm {nps}}(G)\le 9$ . For completeness, we include the results for $\operatorname {\mathrm {nps}}(G)\le 4$ .

Let $C_n$ denote the cyclic group of order n and let $\operatorname {\mathrm {Alt}}(n)$ and $\operatorname {\mathrm {Sym}}(n)$ denote the alternating and symmetric groups of a set of size n. See Definition 2.5 for descriptions of the other groups referred to in the following theorem.

Theorem 1.1. For $0\le k\le 9$ , a group has exactly k nonpower subgroups if and only if, up to isomorphism, it is one of the following:

  1. k = 0 a cyclic group;

  2. k = 1 no examples;

  3. k = 2 no examples;

  4. k = 3 $C_2\times C_2$ , $Q_8$ or $G_{n,3}$ for $n\ge 1$ ;

  5. k = 4 $C_3\times C_3$ ;

  6. k = 5 $C_2\times C_4$ or $G_{n,5}$ for $n\ge 1$ ;

  7. k = 6 $C_5\times C_5$ , $C_2\times C_2\times C_p$ , $Q_8\times C_p$ , where $p> 2$ is a prime, or $G_{n,3}\times C_q$ for ${n\ge 1}$ , where $q> 3$ is a prime;

  8. k = 7 $D_8$ , $\operatorname {\mathrm {Alt}}(4)$ , $C_2\times C_8$ , $Q_{16}$ , $M_{4,2}$ , $C_3\times C_9$ , $M_{3,3}$ , $G_{n,7}$ or $F_{n,7}$ for $n\ge 1$ ;

  9. k = 8 $C_7\times C_7$ or $C_3\times C_3\times C_p$ , where $p\ne 3$ is a prime;

  10. k = 9 $C_2\times C_{16}$ , $M_{5,2}$ , $C_2\times C_2\times C_{p^2}$ , $Q_8\times C_{p^2}$ , where $p> 2$ is a prime, or $G_{n,3}\times C_{q^2}$ , where $q> 3$ is a prime.

2 Preliminaries

Recall that the exponent $\exp (G)$ of a finite group G is the least positive integer e such that $g^e = 1$ for all $g\in G$ . The number of positive divisors of an integer n is denoted by $\tau (n)$ .

Lemma 2.1. The power subgroups of a finite group G are the subgroups $G^d$ , where d is a divisor of the exponent of G. Thus, $\operatorname {\mathrm {ps}}(G)\le \tau (\exp (G))$ .

Proof. Given $m\ge 1$ , we prove that $G^m = G^n$ , where $n = \gcd (m,e)$ and e is the exponent of G. To this end, we may write $n = am+be$ and $m = dn$ for some integers a, b and d. Then, for all g in G, $g^m = g^{nd}\in G^n$ and $g^n = g^{am+be} = g^{ma}\in G^m$ , from which we get $G^m = G^n$ .

Lemma 2.2 [Reference Anabanti, Aroh, Hart and Oodo1, Lemma 3].

If A and B are finite groups such that $|A|$ and $|B|$ are coprime, then $\operatorname {\mathrm {ps}}(A\times B) = \operatorname {\mathrm {ps}}(A)\operatorname {\mathrm {ps}}(B)$ and $\operatorname {\mathrm {nps}}(A\times B) = \operatorname {\mathrm {nps}}(A)s(B) + \operatorname {\mathrm {ps}}(A)\operatorname {\mathrm {nps}}(B)$ , where $s(B)$ is the number of subgroups of B.

Corollary 2.3. For any finite abelian group G, we have $\operatorname {\mathrm {ps}}(G) = \tau (\exp (G))$ .

Proof. From Lemma 2.2, it is no loss to assume that G is an abelian p-group. Then, by Lemma 2.1, it suffices to prove that, for different divisors m and n of $\exp (G)$ , we have $G^n\neq G^m$ . Let $\exp (G) = p^e$ and $e\ge i> j\ge 0$ . Then $G^{p^i} = (G^{p^{j}})^{p^{i-j}} < G^{p^j}$ .

Lemma 2.4 [Reference Zhou, Shi and Duan9, Lemma 2].

Suppose that N and H are subgroups of G such that $N\trianglelefteq G$ and $N\subseteq H$ . If $H/N$ is a nonpower subgroup of $G/N$ , then H is a nonpower subgroup of G. Therefore, $\operatorname {\mathrm {nps}}(G)\ge \operatorname {\mathrm {nps}}(G/N)$ .

Definition 2.5.

  1. (i) For $n\ge 3$ , $\langle a,b\mid a^n = b^2 = 1,\ b^{-1}ab = a^{-1} \rangle $ is a presentation for the dihedral group $D_{2n}$ of order $2n$ .

  2. (ii) For $n\ge 3$ , $\langle a,b\mid a^{2^{n-1}} = b^2 = z,\ z^2 = 1,\ b^{-1}ab = a^{-1} \rangle $ is a presentation for the generalised quaternion group $Q_{2^n}$ of order $2^n$ .

  3. (iii) For $n\ge 4$ , $\langle a,b \mid a^{2^{n-1}} = b^2 = 1,\ b^{-1}ab = a^{-1+2^{n-2}} \rangle $ is a presentation for the semidihedral group $S_{2^n}$ of order $2^n$ .

  4. (iv) For $n\ge 4$ when $p = 2$ and $n\ge 3$ when p is an odd prime, a presentation for the quasidihedral group $M_{n,p}$ of order $p^n$ is $\langle a,b \mid a^{p^{n-1}} = b^p = 1,\ b^{-1}ab = a^{1+p^{n-2}} \rangle $ . The group $M_{3,p}$ is the extraspecial group of order $p^3$ and exponent $p^2$ .

  5. (v) For an odd prime p, $\langle x,y,z \mid x^p = y^p = z^p = 1,\ [x,z] = [y,z] = 1,\ [x,y] = z \rangle $ is a presentation for the extraspecial group $M(p)$ of order $p^3$ and exponent p.

  6. (vi) For $k\ge 1$ and $n\ge 2$ , $\langle a,b\mid a^{2^n}= b^k = 1,\ a^{-1}ba = b^{-1} \rangle $ is a presentation for the group $G_{n,k}$ of order $2^n k$ . Note that $G_{1,k} = D_{2k}$ and $G_{n,2} = C_2\times C_{2^n}$ .

  7. (vii) For $n\ge 1$ and a prime $p\equiv 1\pmod 3$ , choose $i\not \equiv ~1\pmod p$ such that ${i^3\equiv 1\pmod p}$ . Then $\langle a,b\mid a^{3^n} = b^p = 1,\ a^{-1}ba = b^i \rangle $ is a presentation for the group $F_{n,p}$ of order $3^np$ .

  8. (viii) For $n\ge 1$ , $\langle a,b,c \mid a^{3^n} = b^2 = 1,\ bc = cb,\ b^a = c,\ c^a = bc \rangle $ is a presentation for the group $A_n = (C_2\times C_2)\rtimes C_{3^n}$ of order $2^2 3^n$ . When $n = 1$ , $A_1 = \operatorname {\mathrm {Alt}}(4)$ .

  9. (ix) For $n\ge 1$ and a prime p,

    $$ \begin{align*}\langle a,b,c\mid [a,b] = c,\ a^p = b^{p^n} = c^p = 1,\ [a,c] = [b,c] = 1 \rangle\end{align*} $$

    is a presentation for the group $B^1_{n,p}$ of order $p^{n+2}$ . Except for $B^1_{1,2} = D_8$ , it is nonmetacyclic (see [Reference Blackburn4, Lemma 2.5]). The quotient mod $\langle c \rangle $ is $C_p\times C_{p^n}$ and, for p odd, $B^1_{1,p}= M(p)$ .

  10. (x) For $n\ge 1$ and a prime p,

    $$ \begin{align*} \langle a,b,c\mid [a,b] = c,\ a^p = c,\ b^{p^n} = c^p = 1,\ [a,c] = [b,c] = 1 \rangle \end{align*} $$

    is a presentation for the group $B^2_{n,p}$ of order $p^{n+2}$ . It is metacyclic: $\langle a \rangle $ is a normal cyclic subgroup with cyclic quotient. The quotient mod $\langle c \rangle $ is $C_p\times C_{p^n}$ , $B^2_{1,2} = D_8$ and, for p odd, $B^2_{1,p}= M_{3,p}$ .

There are some basic facts about a $p'$ -group acting on a p-group in [Reference Gorenstein6]. For the reader’s convenience, we give these theorems as lemmas.

Lemma 2.6 [Reference Gorenstein6, Theorem 5.2.3].

If A is a $p'$ -group of automorphisms of the abelian p-group P, then $ P=C_{P}(A)\times [P,A]$ .

Lemma 2.7 [Reference Gorenstein6, Theorem 5.3.5].

If A is a $p'$ -group of automorphisms of the p-group P, then $P=CH$ , where $C=C_{P}(A)$ and $H=[P,A]$ . In particular, if $H\subseteq \Phi (P)$ , then $A=1$ .

The p-groups with a cyclic maximal subgroup are well known. It is clear that $C_{p^n}$ and $C_p\times C_{p^n}$ are the only abelian groups of this type. For the nonabelian case, we have the following lemma.

Lemma 2.8 [Reference Gorenstein6, Theorem 5.4.4].

Let P be a nonabelian p-group of order $p^n$ that contains a cyclic subgroup of index p. Then one of the following holds.

  1. (i) p is odd and P is isomorphic to $M_{n,p}$ , for $n \ge 3$ .

  2. (ii) $p = 2$ , $n = 3$ and P is isomorphic to $D_8$ or $Q_8$ .

  3. (iii) $p = 2$ , $n> 3$ and P is isomorphic to $M_{n,2}$ , $D_{2^n}$ , $Q_{2^n}$ or $S_{2^n}$ .

The p-groups in Lemma 2.8 are well studied. We collect some basic facts in the following lemmas.

Lemma 2.9 [Reference Gorenstein6, Theorem 5.4.3].

For $P=M_{n,p}$ :

  1. (i) $P'=\langle a^{p^{n-1}} \rangle $ ;

  2. (ii) $\Phi (P)=Z(P)=\langle a^{p} \rangle $ ; and

  3. (iii) $\Omega _{i}(P)= \langle a^{p^{n-i-1}},b \rangle $ is abelian of type ( $p^{i},p$ ), $1\leq i \leq n-2$ .

Lemma 2.10 [Reference Gorenstein6, Theorem 5.4.5].

Let P be a nonabelian $2$ -group of order $2^{n}$ in which $|P/P'|=4$ . Then P is isomorphic to $D_{2^n}$ , $Q_{2^n}$ or $S_{2^n}$ .

Lemma 2.11 [Reference Gorenstein6, Theorem 5.5.1].

A nonabelian p-group P of order $p^3$ is extraspecial and is isomorphic to one of the groups $M_{3,p}$ , $M(p)$ , $D_{8}$ or $Q_{8}$ .

Theorem 2.12. There is no finite p-group G such that $G/N \simeq M_{n,p}$ , where N is a central subgroup of G of order p contained in $G'$ .

Proof. From the presentation of $M_{n,p}$ in Definition 2.5, we may suppose that G has a presentation of the form

$$ \begin{align*} \langle a,b,c \mid a^{p^{n-1}} = c^i, b^p = c^j,\ c^p = 1,\ b^{-1}ab = a^{1+p^{n-2}}c^k, ac = ca,\ bc = cb \rangle, \end{align*} $$

where $N = \langle c \rangle $ , $0\le i,j,k < p$ and not all i, j, k are zero. Since $c\in Z(G)$ , it is clear that $a^p$ commutes with b and hence $z = a^{p^{n-2}}\in Z(G)$ . Therefore, $[a,b] = zc^k \in Z(G)$ and it follows from elementary properties of commutators that $G' = \langle zc^k \rangle $ , which is a contradiction.

From Lemmas 2.2, 2.5 and Theorem 2.3 in [Reference Blackburn4], we deduce the following lemma.

Lemma 2.13. For a nonabelian p-group G generated by two elements, let $R=\Phi (G')G_{3}$ , where $G_{3}=[[G,G],G]$ . Then:

  1. (i) R is the only maximal subgroup of $G'$ that is normal in G;

  2. (ii) G is metacyclic if and only if $G/R$ is metacyclic; and

  3. (iii) if the type of $G/G'$ is $(p,p^{n})$ and $G/R$ has no cyclic maximal subgroup, then $G/R$ is isomorphic to $B_{n,p}^{1}$ or $B_{n,p}^{2}$ .

Proof. (i) and (ii) are the statements of Lemma 2.2 and Theorem 2.3 in [Reference Blackburn4].

(iii) Let $H=G/R$ . Since $H'=(G/R)'=G'R/R=G'/R$ , we have $H/H'\simeq G/G'$ and $H'\subseteq Z(H)$ . Thus, we may assume that $H/H'=\langle aH',bH' \rangle $ . Then $H=\langle a,b \rangle $ and ${H'=\langle c \rangle }$ , where $c=[a,b]$ and $|c|=p$ . Thus, $c\in Z(H)$ . Since the type of $H/H'$ is $(p,p^{n})$ , we have $a^p=c^i$ , $b^{p^{n}}=c^j$ for suitable integers i, j. If $p\nmid j$ , then $\langle b \rangle $ is a cyclic maximal subgroup of $G/R$ , contrary to our assumption. Thus, $b^{p^{n}}=1$ . If $p\nmid i$ , replacing c with $c^{i}$ , we have $H\simeq B_{n,p}^{2}$ . If $p\mid i$ , then $H\simeq B_{n,p}^{1}$ .

3 A catalogue of nonpower values

The value of $\operatorname {\mathrm {nps}}(G)$ for the groups that occur in the proof of Theorem 1.1 can be computed from their presentation or from the Small Groups Database using the computer algebra system Magma [Reference Bosma, Cannon and Playoust5]. For ease of reference, we include some general formulas here.

Proposition 3.1. For an integer n and a prime p we have:

  1. (i) for $n\ge 3$ , $\operatorname {\mathrm {nps}}(D_{2^n}) = 2^n - 1$ ;

  2. (ii) for $n\ge 3$ , $\operatorname {\mathrm {nps}}(Q_{2^n}) = 2^{n-1} - 1$ ;

  3. (iii) for $n\ge 4$ , $\operatorname {\mathrm {nps}}(S_{2^n}) = 3\cdot 2^{n-2} - 1$ ;

  4. (iv) for $n\ge 3$ , $\operatorname {\mathrm {nps}}(M_{n,p}) = p(n-1) + 1$ (when $p = 2$ , assume that $n\ge 4$ );

  5. (v) $\operatorname {\mathrm {nps}}(M(p)) = p^2 + 2p + 2$ ;

  6. (vi) if $p> 2$ , then $\operatorname {\mathrm {nps}}(G_{n,p^k}) = p(p^k-1)/(p-1)$ ;

  7. (vii) if $p\equiv 1\pmod 3$ , then $\operatorname {\mathrm {nps}}(F_{n,p}) = p$ ;

  8. (viii) for $n\ge 1$ , $\operatorname {\mathrm {nps}}(A_n) = 3n + 4$ ;

  9. (ix) for $n\ge 1$ , $\operatorname {\mathrm {nps}}(B^1_{n,p})\ge 17$ except that $\operatorname {\mathrm {nps}}(B^1_{1,2}) = 7$ ;

  10. (x) for $n\ge 1$ , $\operatorname {\mathrm {nps}}(B^2_{n,p})\ge 11$ except that $\operatorname {\mathrm {nps}}(B^2_{1,2}) = \operatorname {\mathrm {nps}}(B^2_{1,3}) = 7$ .

Proof. (i), (ii) and (iii) follow from Proposition 11 and Theorems 16 and 17 of [Reference Anabanti, Aroh, Hart and Oodo1].

(iv) Suppose $G = M_{n,p}$ . From Lemma 2.9, $Z(G) = \Phi (G) = G^p = \langle a^p \rangle $ and $\Omega _i(G) = \langle a^{p^{n-i-1}},b \rangle $ . The commutator $c = [a,b] = a^{p^{n-2}}$ has order p, $[a^i,b] = c^i$ , $G' = \langle c \rangle $ and $(ba^j)^p = c^{jp(p-1)/2} a^{jp}$ .

Therefore, for $1\le i\le n-1$ , we have $G^{p^i} = \langle a^{p^i} \rangle $ and so $\operatorname {\mathrm {ps}}(G) = n$ . The maximal subgroups of G are the cyclic subgroups $\langle a \rangle $ and $\langle a^ib \rangle $ , $1\le i< p$ , and the noncyclic subgroup $\langle a^p, b \rangle $ . Therefore, the proper noncyclic subgroups of G are the abelian groups $\Omega _i(G)$ of type $(p^i,p)$ , $1\le i\le n-2$ . Thus, G has $n-1$ noncyclic subgroups each of which, except $\Omega _1(G)$ , has p maximal cyclic subgroups. There are $p+1$ cyclic subgroups in $\Omega _1(G)$ . Therefore, $s(G) = (n-1) + p(n-2) + (p+1) + 1 = p(n-1)+n+1$ and $\operatorname {\mathrm {nps}}(G) = p(n-1)+1$ .

(v) The exponent of $M(p)$ is p; therefore, it has $(p^3-1)/(p-1)$ subgroups of order p. Every subgroup of order $p^2$ is normal and hence contains the centre (of order p). Therefore, there are $(p^3-p)/(p^2-p)$ subgroups of order $p^2$ . In total there are ${p^2+2p+2}$ proper subgroups all of which are nonpower subgroups.

(vi), (vii) We only prove that $\operatorname {\mathrm {nps}}(G_{n,p^k}) = p(p^k-1)/(p-1)$ . Then $\operatorname {\mathrm {nps}}(F_{n,p})=p$ is obtained similarly.

Recall that $G_{n,p^k}=\langle a,b\mid a^{2^{n}}= b^{p^{k}} = 1,\ a^{-1}ba = b^{-1} \rangle $ . Thus, $\langle a^2 \rangle \subseteq Z(G_{n,p^k})$ . Then we have $G_{n,p^k}/\langle a^2 \rangle \simeq D_{2p^k}$ . Since the number of Sylow $2$ -subgroups of $D_{2p^k}$ is $p^k$ , the number of Sylow $2$ -subgroups of $G_{n,p^k}$ is also $p^k$ . Thus, the Sylow $2$ -subgroups of $G_{n,p^k}$ are self-normalising. For $0\leq j\leq k-1$ , $\langle b^{p^{k-j}} \rangle $ is the unique subgroup of order $p^j$ of $G_{n,p^k}$ . Let $H_{j}$ be a subgroup of $G_{n,p^k}$ and $|H_{j}|=2^{n}p^j$ . Then the Sylow $2$ -subgroups of $H_{j}$ are self-normalising. Thus, every subgroup of order $2^{n}p^j$ in $G_{n,p^k}$ contains $p^j$ Sylow $2$ -subgroups of $G_{n,p^k}$ . Thus, there are exactly $p^{k-j}$ subgroups of order $2^{n}p^j$ in $G_{n,p^k}$ and they are conjugate to each other in $G_{n,p^k}$ . The number of those subgroups is ${p(p^k-1)/(p-1)}$ . Now we prove that the other subgroups of $G_{n,p^k}$ are power subgroups. Since $\langle a^2 \rangle \subseteq Z(G_{n,p^k})$ , for $0\leq i\leq n-1$ and $0\leq s \leq k$ , $\langle a^{2^{n-i}}, b^{p^{k-s}} \rangle $ is the unique subgroup of order $2^{i}p^{s}$ in $G_{n,p^k}$ and $\langle a^{2^{n-i}}, b^{p^{k-s}} \rangle = G_{n,p^k}^{2^{n-i}p^{k-s}}$ . This completes the proof.

(viii) The power subgroups of $A_n$ are distinct except that $A_n^2 = A_n^1$ . The $2n - 1$ subgroups Q, $\langle a^{3^i} \rangle $ and $Q\langle a^{3^i} \rangle $ , where $1\le i < n$ and $Q = \langle b,c \rangle $ , are the proper nontrivial normal subgroups of $A_n$ . The other subgroups are $\langle a \rangle $ , $\langle a^{3^i}b \rangle $ , $\langle a^{3^i}c \rangle $ , $\langle a^{3^i}bc \rangle $ for $0\le i\le n$ . Thus, $A_n$ has $5(n+1)$ subgroups; therefore, $\operatorname {\mathrm {nps}}(A_n) = 3n+4$ .

(ix) There are n proper power subgroups of $B^1_{n,p}$ ; their orders are $p^i$ for $0\le i < n$ . We claim that $\operatorname {\mathrm {nps}}(B_{n,p}^1)$ is an increasing function of n and p by counting the subgroups. Since $B_{1,p}^1$ is well known, we only consider the case $n\ge 2$ . We count subgroups by considering their exponent. First, notice that $\Omega _{n-1}(B_{n,p}^1)=\langle a \rangle \times \langle b^{p} \rangle \times \langle c \rangle \simeq C_p \times C_{p^{n-1}} \times C_p$ , which implies that all the subgroups of exponent $\le p^{n-1}$ are in $\Omega _{n-1}(B_{n,p}^1)$ . Thus, the number of subgroups with exponent $\le p^{n-1}$ is $s(C_p\times C_{p^{n-1}} \times C_p)$ , which is an increasing function of n and p. Next, we consider the subgroups of exponent $p^n$ . Considering the number of elements of order $p^n$ , we see that there are $p^2$ cyclic subgroups of order $p^n$ . Let H be a subgroup of order $p^{n+1}$ with $\exp (H)=p^n$ . Then $c \in H$ and $\exp (H/\langle c \rangle )=p^n$ . Since $B_{n,p}^1/\langle c \rangle \simeq C_p\times C_{p^n}$ , we see that the number of such subgroups H is p. Therefore, the number of subgroups of exponent $p^n$ is $p^2+p+1$ . Thus, $\operatorname {\mathrm {nps}}(B_{n,p}^1)$ is an increasing function of n and p, as claimed. By direct calculation or from Magma [Reference Bosma, Cannon and Playoust5], we find that $\operatorname {\mathrm {nps}}(B^1_{2,2}) = 20$ and $\operatorname {\mathrm {nps}}(B^1_{1,3}) = \operatorname {\mathrm {nps}}(M(3)) = 17$ .

(x) There are n proper power subgroups of $B_{n,p}^2$ ; their orders are $p^i$ for $0\le i\le n-2$ and $p^n$ . Similarly, $\operatorname {\mathrm {nps}}(B^2_{n,p})$ is an increasing function of n and p. By direct calculation or from Magma [Reference Bosma, Cannon and Playoust5], we find that $\operatorname {\mathrm {nps}}(B^2_{2,2}) = 12$ , $\operatorname {\mathrm {nps}}(B^2_{2,3}) = 20$ and $\operatorname {\mathrm {nps}}(B^2_{1,5}) =\operatorname {\mathrm {nps}}(M_{3,5})= 11$ .

Remark 3.2. It can be shown that:

  1. (i) for $n\ge 2$ , $\operatorname {\mathrm {nps}}(B^1_{n,p}) = p^2(2n-1) + p(n+1) + 2$ ; and

  2. (ii) for $n\ge 2$ , $\operatorname {\mathrm {nps}}(B^2_{n,p}) = (p+1)(2 + p(n-1))$ .

Lemma 3.3. Let $G = D_{2p}\times \underbrace {C_2\times \cdots \times C_2}_n$ . For $n \ge 1$ and a prime $p> 2$ , we have $\operatorname {\mathrm {nps}}(G)\ge 3p+4$ and equality holds when $n = 1$ .

Proof. For all $n\ge 1$ , the exponent of G is $2p$ and the only proper nontrivial power subgroup is $G^2$ of order p. The group $D_{2p}\times C_2$ has $3p+7$ subgroups; therefore, G has at least $3p+4$ nonpower subgroups.

Lemma 3.4. Let $X_{n,p} = D_{2p}\times \underbrace {C_3\times \cdots \times C_3}_n$ . For $n \ge 1$ and a prime $p> 3$ , we have $\operatorname {\mathrm {nps}}(X_{n,p})=(p+3)s(C_{3}^{n})-6\geq 10$ , where $\underbrace {C_3\times \cdots \times C_3}_n$ is denoted by $C_3^n$ for short. For $n \ge 1$ , $\operatorname {\mathrm {nps}}(X_{n,3}) \geq \operatorname {\mathrm {nps}}(X_{1,3})=10$ .

Proof. By simple calculation, $s(D_{2p})=p+3$ and $\operatorname {\mathrm {ps}}(D_{2p}) = 3$ . Let $p>3$ . Then, from Lemma 2.2, $\operatorname {\mathrm {nps}}(X_{n,p})=\operatorname {\mathrm {nps}}(D_{2p})s(C_{3}^{n})+ps(D_{2p})\operatorname {\mathrm {nps}}(C_{3}^{n})= \operatorname {\mathrm {nps}}(D_{2p})s(C_{3}^{n})+ps(D_{2p})(s(C_{3}^{n})-2)=(p+3)s(C_{3}^{n})-6$ . In particular, $\operatorname {\mathrm {nps}}(X_{1,5})=10$ . For any $n\geq 1$ , $\operatorname {\mathrm {ps}}(X_{n,3})=3$ . Thus, $\operatorname {\mathrm {nps}}(X_{n,3})$ is an increasing function for n. Then a straightforward calculation shows that $\operatorname {\mathrm {nps}}(X_{1,3}) = 10$ , and this completes the proof.

Lemma 3.5 [Reference Tǎrnǎuceanu8, Theorem 3.3].

For $n_{2}\ge n_{1}\ge 1$ and a prime p, the total number of subgroups of $C_{p^{n_{1}}}\times C_{p^{n_{2}}}$ is

$$ \begin{align*} \dfrac{(n_{2}-n_{1}+1)p^{n_{1}+2}-(n_{2}-n_{1}-1)p^{n_{1}+1}-(n_{2}+n_{1}+3)p+(n_{2}+n_{1}+1)}{(p-1)^{2}}. \end{align*} $$

Lemma 3.6. For $n_{2}\ge n_{1}\ge 1$ and a prime p, the value of $\operatorname {\mathrm {nps}}(C_{p^{n_{1}}}\times C_{p^{n_{2}}})$ is

$$ \begin{align*} \frac{(n_{2}-n_{1}+1)p^{n_{1}+2}-(n_{2}-n_{1}-1)p^{n_{1}+1}-(n_{2}+n_{1}+3)p+(n_{2}+n_{1}+1)}{(p-1)^{2}}-(n_{2}+1). \end{align*} $$

Proof. This follows from Lemma 3.5 and Corollary 2.3.

Example 3.7. We have $\operatorname {\mathrm {nps}}(C_p\times C_{p^n}) = pn+1$ and $\operatorname {\mathrm {nps}}(C_{p^2}\times C_{p^2}) = (p+1)(p+2)$ .

4 The groups with at most nine nonpower subgroups

In proving Theorem 1.1, we use Theorem 1.3 of [Reference Anabanti and Hart2] and the theorems of [Reference Szász7] and [Reference Zhou, Shi and Duan9], which we summarise in the following lemma.

Lemma 4.1. For a finite group G:

  1. (1) G is cyclic if and only if $\operatorname {\mathrm {nps}}(G) = 0$ ;

  2. (2) if G is noncyclic, then $\operatorname {\mathrm {nps}}(G) \ge 3$ ;

  3. (3) if $\operatorname {\mathrm {nps}}(G) = 3$ , then G is $C_2\times C_2$ , $Q_8$ or $G_{n,3}$ for $n\ge 1$ ; and

  4. (4) if $\operatorname {\mathrm {nps}}(G) = 4$ , then G is $C_3\times C_3$ .

Let $\operatorname {\mathrm {Syl}}_p(G)$ denote the set of Sylow subgroups of G. Recall that $P\in \operatorname {\mathrm {Syl}}_p(G)$ has $|G:N_G(P)|$ conjugates and we have $|G:N_G(P)|\equiv 1\pmod p$ . Moreover, $N_G(P)$ is self-normalising; therefore, it also has $|G:N_G(P)|$ conjugates.

Lemma 4.2. Let G be a finite group and let P be a Sylow p-subgroup of G such that $P\ne N_G(P)\ne G$ . Then $\operatorname {\mathrm {nps}}(G)\le 9$ if and only if, for some $n\ge 1$ and a prime $q> 3$ :

  1. (1) $\operatorname {\mathrm {nps}}(G) = 6$ , $p = 2$ and $G\simeq G_{n,3}\times C_q$ ; or

  2. (2) $\operatorname {\mathrm {nps}}(G) = 9$ , $p = 2$ and $G\simeq G_{n,3}\times C_{q^2}$ .

Proof. Since $P\ne N_G(P)\ne G$ , both P and $N_G(P)$ have at least $p+1$ conjugates and so $2(p+1)\le \operatorname {\mathrm {nps}}(G)\le 9$ , from which we get p is 2 or 3. If $Q\in \operatorname {\mathrm {Syl}}_q(G)$ for some prime $q\ne p$ and $N_G(Q)\ne G$ , then $2(p+1)+q+1\le \operatorname {\mathrm {nps}}(G)\le 9$ , which is impossible. Therefore, for $q\ne p$ , all Sylow q-subgroups are normal. Thus, $G = NP$ , where N is a nilpotent normal subgroup such that $N\cap P = 1$ . Then $[N_N(P),P]\subseteq N\cap P = 1$ and, consequently, $N_G(P) = PC_N(P)$ . If $|G : N_G(P)|> p+1$ , then $2(2p+1) \le 9$ , which is a contradiction. Thus, $|G:N_G(P)| = p + 1$ .

If $p = 3$ , then $\operatorname {\mathrm {nps}}(G)$ is 8 or 9. Thus, for $Q\in \operatorname {\mathrm {Syl}}_2(G)$ , we have $Q\trianglelefteq G$ and ${G = N_G(P)Q}$ . If $[P,Q]\subseteq \Phi (Q)$ , it follows from Lemma 2.7 that $[P,Q] = 1$ . But then ${Q\subseteq C_G(P)\subseteq N_G(P)}$ , which is a contradiction. Thus, Q has at least three subgroups that are not normal in G. Hence, G has at least 11 nonpower subgroups, contrary to $\operatorname {\mathrm {nps}}(G)\le 9$ . Therefore, for the remainder of the proof, we take $p = 2$ .

If P is not cyclic, it follows from Lemma 4.1 that $\operatorname {\mathrm {nps}}(P)\ge 3$ . Then, in addition to the three conjugates of P and the three conjugates of $N_G(P)$ , there would be at least three nonpower subgroups H such that $H/N$ is a nonpower subgroup of $G/N\simeq P$ . In this case, $\operatorname {\mathrm {nps}}(P) = 3$ , $\operatorname {\mathrm {nps}}(G) = 9$ and P is either $C_2\times C_2$ or $Q_8$ . But $\operatorname {\mathrm {nps}}(G) = 9$ implies that the proper subgroups of P are normal in G, which is a contradiction since P is generated by its proper subgroups. Therefore, P is cyclic.

Let $R\in \operatorname {\mathrm {Syl}}_r(N)$ . Then R acts by conjugation on the three conjugates of P. If $r> 3$ , then $R\subseteq N_G(P)$ , from which we get $R\subseteq N_G(P)\cap N\subseteq C_G(P)$ . For $Q\in \operatorname {\mathrm {Syl}}_3(G)$ , we have $[R,Q]\subseteq R\cap Q = 1$ and so $R\subseteq C_G(Q)$ . Consequently, $G = PQ\times A$ , where A is a nilpotent group whose order is not divisible by two or three. From Lemma 2.2,

$$ \begin{align*} \operatorname{\mathrm{nps}}(G) = \operatorname{\mathrm{nps}}(QP)s(A) + \operatorname{\mathrm{ps}}(QP)\operatorname{\mathrm{nps}}(A). \end{align*} $$

It follows from $\operatorname {\mathrm {nps}}(G)\le 9$ that $\operatorname {\mathrm {nps}}(A) = 0$ and so A is cyclic. Furthermore, $s(A)\le 3$ ; therefore, A is either trivial or a cyclic group of order r or $r^2$ for some prime $r> 3$ .

The permutation action of G on the conjugates of P defines a homomorphism $G\to \operatorname {\mathrm {Sym}}(3)$ with kernel $K = \bigcap _{g\in G}N_G(P)^g$ . Since $|Q : C_Q(P)| = 3$ , we have $\Phi (Q)\subseteq C_Q(P)\subseteq K$ . Therefore, $M = (P\cap K)\times \Phi (Q)\times A$ is a normal subgroup of G. The group P acts on the elementary abelian group $\overline Q = Q/\Phi (Q)$ and it follows from Lemma 2.6 that $\overline Q = C_{\overline Q}(P)\times [P,\overline Q]$ . Thus, $[P,\overline Q] \simeq C_3$ and

$$ \begin{align*} G/M\simeq \operatorname{\mathrm{Sym}}(3)\times\underbrace{C_3\times\cdots\times C_3}_n. \end{align*} $$

If $n \ge 1$ , it follows from Lemmas 3.4 and 2.4 that $\operatorname {\mathrm {nps}}(G)\ge 10$ , contrary to assumption. Thus, $Q/\Phi (Q) \simeq C_3$ ; therefore, Q is cyclic. But now $Q = [P,Q]\times C_Q(P)$ ; therefore, $C_Q(P) = 1$ and $|Q| = 3$ .

Let a be a generator of P and let b a generator of Q. Then $a^{-1}ba = b^{-1}$ and so $QP\simeq G_{n,3}$ for some n. The assumption $N_G(P)\ne P$ implies that $A\ne 1$ . Thus, G is either $G_{n,3}\times C_r$ or $G_{n,3}\times C_{r^2}$ for some prime $r> 3$ .

Remark 4.3. The group $G = \operatorname {\mathrm {Sym}}(3)\times C_3$ satisfies the hypothesis of the lemma (with $p=2$ ) except that $\operatorname {\mathrm {nps}}(G) = 10$ .

Lemma 4.4. Let G be a finite group and let P be a Sylow p-subgroup of G such that $P = N_G(P)\ne G$ . Then $\operatorname {\mathrm {nps}}(G)\le 9$ if and only if, for some $n\ge 1$ , one of the following holds.

  1. (1) $\operatorname {\mathrm {nps}}(G) = 3$ , $p = 2$ and $G\simeq G_{n,3}$ .

  2. (2) $\operatorname {\mathrm {nps}}(G) = 5$ , $p = 2$ and $G\simeq G_{n,5}$ .

  3. (3) $\operatorname {\mathrm {nps}}(G) = 7$ , $p = 2$ and $G\simeq G_{n,7}$ .

  4. (4) $\operatorname {\mathrm {nps}}(G) = 7$ , $p = 3$ and $G\simeq F_{n,7}$ .

  5. (5) $\operatorname {\mathrm {nps}}(G) = 7$ , $p = 3$ and $G\simeq \operatorname {\mathrm {Alt}}(4)$ .

Proof. The Sylow subgroup P has $m = |G:N_G(P)|$ conjugates. Since $m\equiv 1\pmod p$ and $N_G(P)\ne G$ , we have $m \ge p+1$ and the conjugates of P are nonpower subgroups. The assumption $\operatorname {\mathrm {nps}}(G)\le 9$ implies that $p\in \{2,3,5,7\}$ .

If $p = 2$ , then $m\in \{3,5,7,9\}$ ; if $p = 3$ , then $m\in \{4,7\}$ ; if $p = 7$ , then $m = 8$ . However, if $p = 5$ , then $m = 6$ and $G$ is a group of twice odd order. It is an elementary fact that a group of twice odd order has a subgroup H of index two, which, in this case, contains P. Then $|H:P| = 3$ , which is impossible. Thus, in all cases m is a power of a prime q and, for $Q\in \operatorname {\mathrm {Syl}}_q(G)$ , we have $G = PQ$ and $|Q| = m$ .

The permutation action on $\operatorname {\mathrm {Syl}}_p(G)$ defines a homomorphism $G\to \operatorname {\mathrm {Sym}}(m)$ whose kernel $K = \bigcap _{g\in G}P^g$ is a proper subgroup of P.

If $N_G(Q)\ne G$ , then Q has at least $q+1$ conjugates; therefore, $m+q+1\le 9$ . In this case, either $p = 2$ and $|Q| = 3$ or $p = 3$ and $|Q| = 4$ . Furthermore, we must have $Q = N_G(Q)$ . Otherwise, both Q and $N_G(Q)$ would have at least $q+1$ conjugates. From the structure of $\operatorname {\mathrm {Sym}}(3)$ and $\operatorname {\mathrm {Sym}}(4)$ , we have $KQ\trianglelefteq G$ and, by the Frattini argument [Reference Gorenstein6, Theorem 1.3.7], $G = KN_G(Q)$ and $Q\ne N_G(Q)$ , which is a contradiction. Therefore, $Q\trianglelefteq G$ and $[K,Q]\subseteq K\cap Q = 1$ and hence $K\subseteq C_G(Q)$ .

The order of Q is either q or $q^2$ ; therefore, Q is abelian.

Case 1: $p = 2$ and $m\in \{3,5,7,9\}$ . We have $G = Q\rtimes P$ , where P is a 2-group and Q is an abelian group of order m. We treat each value of m separately.

Case 1a: $p = 2$ and $|Q| = 3$ . In this case, $|Q| = 3$ and $G/K\simeq \operatorname {\mathrm {Sym}}(3)$ . Since ${|P/K| = 2}$ , we have $\Phi (P)\subseteq K$ and so $\Phi (P)\trianglelefteq G$ . If P is not cyclic, then $G/\Phi (P)\simeq \operatorname {\mathrm {Sym}}(3)\times C_2\times \cdots \times C_2$ and it follows from Lemmas 2.4 and 3.3 that $\operatorname {\mathrm {nps}}(G)\ge 13$ , contrary to assumption. Thus, P is cyclic. If a generates P and b generates Q, then $b^a = b^{-1}$ and $G\simeq G_{n,3}$ .

Case 1b: $p = 2$ and $|Q| = 5$ . In this case, $|Q| = 5$ and $QK/K$ is a normal subgroup of $G/K$ . Therefore, $|G/K|$ is either 10 or 20. If $|G/K| = 20$ , then $G/K = \langle x, b\mid x^4 = b^5 =1, x^{-1}bx = b^2 \rangle $ . This group has 14 subgroups, 4 of which are power subgroups; therefore, there are 10 nonpower subgroups. From Lemma 2.4, $\operatorname {\mathrm {nps}}(G)\ge 10$ , contrary to our assumption. Thus, $|G/K| = 10$ . If P is not cyclic, then $G/\Phi (P)\simeq D_{10}\times C_2\times \cdots \times C_2$ and, using Lemma 3.3, we arrive at a contradiction, as in Case 1a. Thus, P is cyclic. If a generates P and b generates Q, then $b^a = b^{-1}$ and so $G\simeq G_{n,5}$ .

Case 1c: $p = 2$ and $|Q| = 7$ . In this case, $|Q| = 7$ , $G/K\simeq D_{14}$ and P has seven conjugates. If P is not cyclic, then $\operatorname {\mathrm {nps}}(P)\ge 3$ and it follows from Lemma 2.4 that $\operatorname {\mathrm {nps}}(G)\ge 10$ , contrary to assumption. Thus, P is cyclic and $G\simeq G_{n,7}$ .

Case 1d: $p = 2$ and $|Q| = 9$ . Since $\operatorname {\mathrm {nps}}(G) \le 9$ , all subgroups are normal except the Sylow 2-subgroups. In particular, if R is a subgroup of Q of order three, then  ${R\trianglelefteq G}$ ; therefore, $RP\trianglelefteq G$ . But then P is not maximal. Thus, all maximal subgroups are normal. This implies that G is nilpotent and this contradiction shows that there are no examples in this case.

Case 2: $p = 3$ . In this case, $G = Q\rtimes P$ , where P is a 3-group and $|Q| \in \{4,7\}$ .

Case 2a: $p = 3$ and $|Q| = 4$ . We must have $Q\simeq C_2\times C_2$ , otherwise $Q\simeq C_4$ and then $Q\subseteq C_G(P)$ , which contradicts the assumption that $P = N_G(P)$ .

It follows that Q has three subgroups $R_1$ , $R_2$ and $R_3$ of order two and P acts transitively on them. Then $R_1$ , $R_2$ , $R_3$ and the four conjugates of P are not normal in G, from which we get $\operatorname {\mathrm {nps}}(G)\ge 7$ . Let K be the kernel of the action of P on $\{R_1, R_2, R_3\}$ . Then $K = C_P(Q)$ , $|P : K| = 3$ and so $K\trianglelefteq G$ . Then $R_1K$ , $R_2K$ and $R_3K$ are permuted by P and, if $K\ne 1$ , we would have $\operatorname {\mathrm {nps}}(G)\ge 10$ . Therefore, $K = 1$ , $|P| = 3$ and so $G\simeq \operatorname {\mathrm {Alt}}(4)$ .

Case 2b: $p = 3$ and $|Q| = 7$ . For $Q\in \operatorname {\mathrm {Syl}}_7(G)$ , we have $|Q| = 7$ and P is cyclic; otherwise, $\operatorname {\mathrm {nps}}(P) \ge 3$ and we obtain a contradiction by applying Lemma 2.4 to $G/Q$ . The image of the homomorphism $G\to \operatorname {\mathrm {Sym}}(7)$ is the group $F_{1,7}$ of order $21$ . We may write $P = \langle a \rangle $ and $Q = \langle b \rangle $ , where $a^{3^n} = 1$ , $b^7 = 1$ and $b^a = b^2$ . Thus, $G\simeq F_{n,7}$ .

Case 3: $p = 7$ and $|Q| = 8$ . All subgroups are normal except the Sylow 7-subgroups. As in Case 1d, we see that all maximal subgroups of G are normal and so G is nilpotent, which is a contradiction.

Remark 4.5. For $n \ge 2$ , the groups $\langle x, b\mid x^{2^n} = b^5 =1, x^{-1}bx = b^2 \rangle $ have exactly 10 nonpower subgroups and their Sylow 2-subgroups are self-normalising.

Proof of Theorem 1.1.

From Lemmas 4.2 and 4.4, we may suppose that G is nilpotent. If P is a noncyclic p-group, then $\operatorname {\mathrm {nps}}(P)\ge 3$ . Suppose that $H\ne 1$ is a group whose order is not divisible by p. If $\operatorname {\mathrm {nps}}(P\times H)\le 9$ , then, from Lemma 2.2, there are two possibilities: (i) $\operatorname {\mathrm {nps}}(P) = 3$ and H is a cyclic group whose order is a prime or the square of a prime; (ii) $\operatorname {\mathrm {nps}}(P) = 4$ and H is a cyclic group of prime order.

From Lemma 4.1, if $\operatorname {\mathrm {nps}}(P) = 3$ , then G is $C_2\times C_2\times C_r$ , $C_2\times C_2\times C_{r^2}$ , $Q_8\times C_r$ or $Q_8\times C_{r^2}$ , where $r> 2$ is a prime. If $\operatorname {\mathrm {nps}}(P) = 4$ , then G is $C_3\times C_3\times C_r$ , where $r\ne 3$ is a prime.

Thus, from now on, we may suppose that G is a noncyclic p-group. Then $G/\Phi (G)$ is an elementary abelian group of order $p^d$ . The proper subgroups of $G/\Phi (G)$ are nonpower subgroups; therefore, $\operatorname {\mathrm {nps}}(G)\le 9$ implies that $d = 2$ and $p\in \{2,3,5,7\}$ .

The group G can be generated by two elements; therefore, $G/G' = C_{p^m}\times C_{p^n}$ for some $m\le n$ . It follows from Example 3.7 and Lemma 2.4 that $m = 1$ . Thus, $G/G'$ is one of $C_p\times C_p$ for $p\in \{2,3,5,7\}$ , $C_2\times C_{2^n}$ for $n\in \{2,3,4\}$ or $C_3\times C_9$ . If G is abelian, this completes the proof. From now on, we assume that $G'\ne 1$ .

Suppose that $G/G'\simeq C_2\times C_2$ . It follows from Lemma 2.10 that G is isomorphic to $D_{2^n}$ , $S_{2^n}$ or $Q_{2^n}$ . From Proposition 3.1, the only possibilities are $D_8$ , $Q_8$ and $Q_{16}$ .

Since $G'\ne 1$ , there exists $R\trianglelefteq G$ such that $|G'/R| = p$ . We shall determine the structure of $G/R$ for each choice of $G/G'$ .

Suppose that p is odd and $G/G'\simeq C_p\times C_p$ . From Lemma 2.11, $G/R$ is an extraspecial group of order $p^3$ : that is, $M_{3,p}$ or $M(p)$ . From Proposition 3.1(iv) and (v), $\operatorname {\mathrm {nps}}(M_{3,p}) = 2p+1$ and $\operatorname {\mathrm {nps}}(M(p)) = p^2 + 2p + 2$ . Thus, $p = 3$ and $G/R\simeq M_{3,3}$ . The group $M_{3,3}$ has a cyclic subgroup of order nine; therefore, it is metacyclic. It follows from Lemma 2.13 that G is metacyclic and so G has a cyclic normal subgroup that properly contains $G'$ : that is, G has a cyclic subgroup of index three. Therefore, by Lemma 2.8, $G\simeq M_{n,3}$ . (This argument is based on the MathSciNet review of [Reference Berkovich3] by Marty Isaacs.) But, from Lemma 2.9, if $M = M_{n,p}$ , then $M'$ is its unique normal subgroup of order p and $M/M' \simeq C_p\times C_{p^{n-2}}$ . Thus, $R = 1$ and $G\simeq M_{3,3}$ .

Suppose that $G/G'\simeq C_2\times C_{2^n}$ ( $n = 2,3,4$ ) or $C_3\times C_9$ . If $G/R$ has a cyclic subgroup of prime index, it follows from Lemma 2.8 that $G/R$ is isomorphic to $M_{n+2,2}$ or $M_{4,3}$ . The assumption that $\operatorname {\mathrm {nps}}(G)\le 9$ excludes $M_{6,2}$ and $M_{4,3}$ . Then, from Theorem 2.12, $R = 1$ and hence G is isomorphic to $M_{4,2}$ or $M_{5,2}$ .

We may suppose that the exponent of $G/R$ is $2^n$ or 9. Lemma 2.13 shows that $G/R$ is either $B^1_{n,p}$ or $B^2_{n,p}$ for $p\in \{2,3\}$ and $n\ge 2$ . Proposition 3.1(ix) and (x) shows that none of these groups satisfy our assumptions. This completes the proof.

Footnotes

This work was supported by the National Natural Science Foundation of China (11971391, 12071376).

References

Anabanti, C. S., Aroh, A. B., Hart, S. B. and Oodo, A. R., ‘A question of Zhou, Shi and Duan on nonpower subgroups of finite groups’, Quaest. Math. 45(6) (2022), 901910.CrossRefGoogle Scholar
Anabanti, C. S. and Hart, S. B., ‘Groups with a given number of nonpower subgroups’, Bull. Aust. Math. Soc. 106(2) (2022), 315319.CrossRefGoogle Scholar
Berkovich, Y., ‘Short proofs of some basic characterization theorems of finite $p$ -group theory’, Glas. Mat. Ser. III 41(61) (2) (2006), 239258.CrossRefGoogle Scholar
Blackburn, N., ‘On prime-power groups with two generators’, Proc. Cambridge Philos. Soc. 54 (1958), 327337.CrossRefGoogle Scholar
Bosma, W., Cannon, J. and Playoust, C., ‘The Magma algebra system I: The user language’, J. Symbolic Comput. 24 (1997), 235265.CrossRefGoogle Scholar
Gorenstein, D., Finite Groups, Harper’s Series in Mathematics (Harper and Row, New York–Evanston–London, 1968).Google Scholar
Szász, F., ‘On cyclic groups’, Fund. Math. 43 (1956), 238240.CrossRefGoogle Scholar
Tǎrnǎuceanu, M., ‘An arithmetic method of counting the subgroups of a finite abelian group’, Bull. Math. Soc. Sci. Math. Roumanie (N.S.) 53(101) (4) (2010), 373386.Google Scholar
Zhou, W., Shi, W. and Duan, Z., ‘A new criterion for finite noncyclic groups’, Comm. Algebra 34(12) (2006), 44534457.CrossRefGoogle Scholar