1 Introduction and Main Results

Let G be a group. There are several ways to associate a graph to G (see [8] and the references therein). In this paper, we consider the intersection graph of G which is denoted by \(\Gamma (G)\). The intersection graph \(\Gamma (G)\) of a nontrivial group G is a simple and undirected graph defined as follows: the vertex set is the set of all proper non-trivial subgroups of G, and there is an edge between two distinct vertices H and K if and only if \(H\cap K\ne 1\), where 1 denotes the trivial subgroup of G. The graph \(\Gamma (G)\) has been extensively studied (see, for example, [1, 10, 11, 14, 19, 21, 22, 26]), when G is finite. Also in [12], the subgraph \(\Gamma M(G)\) whose vertices are maximal subgroups of a finitely generated group G is investigated. Intersection graphs of subsemigroups of a semigroup, submodules of a module, and ideals of a ring, were investigated in [2, 7] and [9, 15], respectively.

Let \(\Gamma \) be a simple graph. The set of vertices of every complete subgraph of \(\Gamma \) is called a clique of \(\Gamma \). The maximum size of a complete subgraph of \(\Gamma \) is called the clique number of \(\Gamma \) and it is denoted by \(\omega (\Gamma )\). For convenience, we write \(\omega (\Gamma )=0\) if \(\Gamma \) has no vertices (i. e. \(\Gamma \) is the empty graph) and \(\omega (\Gamma )=1\) if \(\Gamma \) has a non-empty vertex set with no edges (i. e. \(\Gamma \) is null).

In group theory, it is well known that the quantitative properties of some special subgroups play an important role in characterizing the solvability of groups (see [16, 17, 25]). In this paper we give a criterion for solvability of G by \(\omega (\Gamma (G))\).

Theorem 1.1

Let G be a finite group such that \(\omega (\Gamma (G))<13\). Then G is solvable.

We point out that \(\omega (\Gamma (A_5))=13\), where \(A_5\) is the alternating group on 5 letters (see the proof of Lemma 3.3), therefore, the bound in Theorem 1.1 is the best possible.

As a consequence of Theorem 1.1, we give a characterization of \(A_5\).

Corollary 1.2

Let G be a non-solvable group. Then \(\omega (\Gamma (G))=13\) if and only if \(G\cong A_5\).

Remark 1.3

For a finite group G, it is clear that \(\omega (\Gamma (G))=0\) if and only if \(G\cong C_p\), where \(C_p\) is a cyclic group of order p, for some prime number p. By Lemma 2.2 of [19], we see that \(\omega (\Gamma (G)=1\) if and only if G is isomorphic to one of the following groups:

$$\begin{aligned} C_{p^2},~~~C_{pq},~~~C_p\times C_p,~~~C_p\rtimes C_q, \end{aligned}$$

where p and q are distinct prime numbers and the last group is the semidirect product of \(C_p\) by \(C_q\).

In what follows, we determine groups G with \(2\le \omega (\Gamma (G))\le 4\).

Theorem 1.4

Let G be a finite group. Then \(\omega (\Gamma (G))=2\) if and only if one of the following cases occurs::

\(\mathrm{(a)}\) G is a cyclic group of order \(p^3\), for a prime p.

\(\mathrm{(b)}\) \(G=N\rtimes H\) is Frobenius whose kernel N is the minimal normal subgroup of G such that \(N\cong C_p\times C_p\) and \(H\cong C_q\), where p and q are primes.

Theorem 1.5

Let G be a finite group. Then \(\omega (\Gamma (G))=3\) if and only if one of the following statements holds:

\(\mathrm{(a)}\) G is cyclic and \(G\cong C_{p^4},~C_{p^2q}\) or \(C_{pqr}\), where pq and r are distinct primes.

\(\mathrm{(b)}\) \(G=N\rtimes H\) is Frobenius whose kernel N is the (unique) minimal normal subgroup of G and we have either

\(\mathrm{(1)}\) \(N\cong C_p\) and \(H\cong C_{q^2}\) or \(C_{qr}\) or

\(\mathrm{(2)}\) \(N\cong C_p\times C_p\), \(H\cong C_{q^2}\) and G does not have any subgroup of order pq, where pq and r are distinct primes.

Theorem 1.6

Let G be a finite group. Then \(\omega (\Gamma (G))=4\) if and only if one of the following holds:

\(\mathrm{(a)}\) G is abelian and \(G\cong C_{p^5},~C_2\times C_2\times C_q\) or \(C_4\times C_2\), where pq are primes and q is odd..

\(\mathrm{(b)}\) G is non-abelian and isomorphic to one of the followings:

\(\mathrm{(1)}\) \(D_8\) or \(Q_8\) or

\(\mathrm{(2)}\) \(G=N\rtimes H\) is Frobenius whose kernel N is the (unique) minimal normal subgroup of G such that \(N\cong C_p\times C_p\), \(H\cong C_{qr}\) and G does not have subgroups of orders pr and pq, where pqr are distinct primes.

Now we give some examples of groups, satisfying in Theorems 1.4, 1.5 and 1.6. Since \(A_4\cong (C_2\times C_2)\rtimes C_3\) satisfies in Theorem 1.4(b),we have \(\omega (\Gamma (A_4))=2\). In GAP-System [23], two groups

$$\begin{aligned} G_1=AllSmallGroups(42, IsAbelian,false)[1]\cong C_7\rtimes C_6 \end{aligned}$$

and

$$\begin{aligned} G_2=AllSmallGroups(52, IsAbelian,false)[2]\cong C_{13}\rtimes C_4, \end{aligned}$$

satisfy in Theorem 1.5(b)(1) and so \(\omega (\Gamma (G_1))=\omega (\Gamma (G_2)=3\). Also, for \(G_3=AllSmallGroups(11^2. 4, IsAbelian,false)[5]\cong (C_{11}\times C_{11})\rtimes C_4\), we have \(\omega (\Gamma (G_3))=3\), by Theorem 1.5(b)(2). Finally if \(G_4=AllSmallGroups(11^2. 6, IsAbelian,false)[5]\cong (C_{11}\times C_{11})\rtimes C_6\), then \(\omega (\Gamma (G_4))=4\), by Theorem 1.6(b)(2).

In this paper all groups are finite and we use the usual notation, for example \(C_n\), \(A_n\), \(S_n\), PSL(2, q) and Sz(q), respectively, denote the cyclic group of order n, the alternative group on n letters, the symmetric group on n letters, the projective special linear group of degree 2 over the finite field of size q and the Suzuki group over the field with q elements. For a group G, \(x\in G\) and \(H\le G\), we denote by \(Z(G), C_G(x), N_G(H), H^x\), the center of G, the centralizer of x in G, the normalizer of H in G and the conjugate of H in G by x, respectively. Also \(\mathcal {S}(G)\) is the set of all subgroups of G. The rest of the notation is standard and can be found mainly in [18].

2 Preliminary Results

We frequently use the following results and we state it here for the reader’s convenience.

Lemma 2.1

(see 1.3.11 of [18]) Let G be a group and H and K be subgroups of G. Then \(|G: H\cap K|\le |G:H||G:K|\), with equality if the indices |G : H| and |G : K| are coprime.

Theorem 2.2

(see 8.5.5 of [18]) If G is a finite group with a subgroup H such that \(H\cap H^x = 1\), for all \(x \in G{\setminus } H\), then \(N = (G{\setminus } (\cup _{x\in G}H^x))\cup \{1\}\) is a normal subgroup of G and \(G =N\rtimes H\).

A group G which has a proper nontrivial subgroup H, satisfying in the hypothesis of Theorem 2.2 is called a Frobenius group, H is called a Frobenius complement and N the Frobenius kernel. In this case, \(Z(G)=1\) and |H| divides \(|N|-1\), by Exercises 8.5 (2) and (6) of [18] respectively. Note that if \(1\ne x\in N\), then \(C_G(x)\le N\), by Exercise 8.5 (5) of [18].

Recall that a subgroup H of a group G is supplemented in G if there is a subgroup K of G such that \(G=HK\). Moreover if \(H\cap K=1\), then H is complemented in G by K.

Proposition 2.3

Let G be a group and N be a proper minimal normal abelian subgroup of G. If \(G=HN\), for some subgroups H of G and N is complemented in G by H, then H is a maximal subgroup of G.

Proof

Assume that there is a subgroup K of G such that \(H\lvertneqq K\). Since \(G=HN\), we have \(K=H(K\cap N)\). Therefore, \(K\cap N\ne 1\). Since N is the abelian normal subgroup of G and \(G=KN\), we have \(K\cap N\lhd G\). Thus \(K\cap N=N\), by minimality of N and so \(N<K\). Consequently \(K=G\), as required. \(\square \)

Recall that a finite group G is said to be primitive if it has a maximal subgroup M with \(Core_G(M):=\cap _{g\in G}M^g=1\). In this situation we call M a stabilizer of G. We need the following theorem of R. Baer on primitive groups.

Theorem 2.4

(see [3]) Let G be a finite primitive group with a stabilizer M. Then one of the following three statements holds:

  1. (1)

    G has a unique minimal normal subgroup N, this subgroup N is self-centralizing, and N is complemented by M in G.

  2. (2)

    G has a unique minimal normal subgroup N, this N is non-abelian, and N is supplemented by M in G.

  3. (3)

    G has exactly two minimal normal subgroups N and \(N^*\), and each of them is complemented by M in G. Also \(C_G(N) = N^*\), \(C_G(N^*) = N\) and \(N\cong N^*\cong NN^* \cap M\).

The following lemma is straightforward but is useful in the sequel.

Lemma 2.5

Let G be a finite group.

\(\mathrm{(i)}\) If \(1<H<G\), then \(\omega (\Gamma (H))< \omega (\Gamma (G))\).

\(\mathrm{(ii)}\) If \(1<N\lhd G\), then \(|\mathcal {S}(\frac{G}{N})|-1+\omega (\Gamma (N))\le \omega (\Gamma (G))\).

\(\mathrm{(iii)}\) Suppose that \(\Omega \) is a clique in \(\Gamma (G)\), and \(H\in \Omega \). If |H| is prime, then \(H\le K\), for every \(K\in \Omega \).

Proof

The proofs of (i) and (iii) are clear. For (ii), if \(\omega (\Gamma (N))=k\), for some positive integer k, then there are proper subgroups \(N_1, \ldots , N_k\) of N such that \(N_i\cap N_j\ne 1\), for each \(1\le i<j\le k\). If T is a subgroup of \(\frac{G}{N}\), then there exists a unique subgroup K of G such that \(N\le K\) and \(T=\frac{K}{N}\). Now assume that \(|\mathcal {S}(\frac{G}{N})|=l\), for some positive integer l. Then there exist subgroups \(K_1=N, K_2, \ldots , K_{l-1}, K_l=G\) containing N of G such that \(\mathcal {S}(\frac{G}{N})=\{\frac{K_1}{N}, \frac{K_2}{N}, \ldots , \frac{K_l}{N}\}\). So we have \(\{N_1, \ldots , N_k, K_1,\ldots , K_{l-1}\}\) is a clique in \(\Gamma (G)\). It follows that \(\omega (\Gamma (G))\ge k+l-1\), as required. \(\square \)

We denote by m(G) the number of maximal subgroups of a group G. In the following, we give m(G), for any p-group G, where p is a prime number.

Lemma 2.6

Let G be a finite p-group. Then \(m(G)=m(\frac{G}{\Phi (G)})=\frac{|\frac{G}{\Phi (G)}|-1}{p-1}\).

Proof

Since \(\Phi (G)\) is contained in any maximal subgroup of G, it is easy to see that \(\frac{M}{\Phi (G)}\) is a maximal subgroup of \(\frac{G}{\Phi (G)}\) whenever M is a maximal subgroup of G. On the other hand, similar to the proof of Lemma 2.5(ii), if T is a maximal subgroup of \(\frac{G}{\Phi (G)}\), then there is a unique maximal subgroup M of G such that \(T=\frac{M}{\Phi (G)}\). Thus we conclude that there exists a one-one correspondence from the set of all maximal subgroups of G to the set of all maximal subgroups of \(\frac{G}{\Phi (G)}\) and then \(m(G)=m(\frac{G}{\Phi (G)})\). By the proof of 5.3.2 of [18], we have \(\frac{G}{\Phi (G)}\) is elementary abelian. Now if \(|\frac{G}{\Phi (G)}|=p^m\), then every maximal subgroup of \(\frac{G}{\Phi (G)}\) has order \(p^{m-1}\) and so by [5, Exercise 1(d), p. 81]] we have \(m(\frac{G}{\Phi (G)})=1+p+\cdots +p^{m-1}\). It follows that \(m(\frac{G}{\Phi (G)})=\frac{|\frac{G}{\Phi (G)}|-1}{p-1}\), as desired. \(\square \)

3 A Criterion for Solvability by \(\omega (\Gamma (G))\)

In this section, we prove Theorem 1.1 and Corollary 1.2. The famous family of simple finite groups are minimal simple groups (i.e., finite non-abelian simple groups all of whose proper subgroups are solvable). The classification of minimal simple groups is given by Thompson (see Corollary 1 of [24]). As Thompson’s classification of the minimal simple groups is a very useful tool to obtain solvability criteria in the class of finite groups. In Lemmas 3.3, 3.4, 3.5, 3.6 and 3.7, we show that if G is a minimal simple group non-isomorphic to \(A_5\), then \(\omega (\Gamma (G)>13\). We need these results in the proof of Theorem 1.1 and Corollary 1.2.

In the following we give some facts about subgroups of \(G=A_5\). The proofs are elementary (for example, one may check by software GAP [23]).

Fact 1. \(|\mathcal {S}(A_5)|=59\) and if \(n_k\) is the number of subgroups of G of order k, then \(n_2=15, n_3=n_6=10, n_4=n_{12}=5\) and \(n_5=n_{10}=6\).

Fact 2. If M is a maximal subgroup of G, then \(|M|\in \{6, 10, 12\}\). Also all maximal subgroups of G of the same order are conjugate in G.

Fact 3. If \(M_1, M_2\) and \(M_3\) are maximal subgroups of G such that \(|M_1|=6, |M_2|=10\) and \(|M_3|=12\), then \(|M_i^x\cap M_j^y|\ne 1\), for every \(1\le i<j\le 3\) and \(x, y\in G\). Moreover, \(|M_2\cap M_2^a|=2\) and \(|M_3\cap M_3^b|=3\), for each \(a\in G{\setminus } M_2\) and \(b\in G{\setminus } M_3\). Therefore, \(\Lambda =\{M_1, M_2^g, M_3^g:~~g\in G\}\) is a clique in \(\Gamma (G)\) and so \(\omega (\Gamma (G))\ge 1+n_{10}+n_{12}=12\).

Fact 4. Assume that \(M_1, M_2\) and \(M_3\) are the same as in Fact 3. If \(M_i, M_i^x\) and \(M_i^y\) are distinct conjugates of \(M_i\) in G, for some \(x, y\in G\), then \(|M_i\cap M_i^x\cap M_i^y|=1\), for each i and \(\{M_1, M_1^x, M_1^y\}\) does not form a clique in \(\Gamma (G)\). Also we have \(|M_1\cap M_1^g|\le 2\), for every \(g\in G{\setminus } M_1\) and there is \(z\in G\) such that \(|M_1\cap M_1^z|=2\). Thus \(\Lambda \cup \{M_1, M_1^z\}\) is a clique, by Fact 3, and so \(\omega (\Gamma (G))\ge 13\).

We use of above facts in the proof of the following lemma.

Lemma 3.1

We have \(\omega (\Gamma (A_5))=13\).

Proof

Let \(\Omega \) be a clique in \(\Gamma (A_5)\) of maximum size. Then \(|\Omega |\ge 13\), by Fact 4. Now, we consider the following cases:

Case 1. In this case, we show that \(\Omega \) does not have any member of size 2. Assume that there is \(H_2\in \Omega \) with \(|H_2|=2\). Then other subgroups of order 2, all subgroups of order 3 and all subgroups of order 5 of \(A_5\) do not belong to \(\Omega \) and so \(|\Omega |\le 57-((n_2-1)+n_3+n_5)=27\), by Fact 1. Since the intersection of any two distinct subgroups of order 4 is trivial, there is only one subgroup \(H_4\) of order 4 containing \(H_2\). Thus, \(|\Omega |\le 27-(n_4-1)=23\). Note that if \(X\in \Omega \), then \(H_2\subseteq X\), by Lemma 2.5(iii). By Fact 4, \(\Omega \) has at most two subgroups of order 6. Also, by Facts 3, 4 and Lemma 2.5 (iii), \(\Omega \) has at most two subgroups of order 10. Since the intersection of any two distinct subgroups of order 12 has order 3 by Fact 3, \(\Omega \) has at most one subgroup of order 12. It follows that \(|\Omega |\le 23-((10-2)+(6-2)+(5-1))=7\), a contradiction.

Case 2. In this case, we show that \(\Omega \) does not have any member of size 3. Assume that there exists \(H_3\in \Omega \) with \(|H_3|=3\). Then \(\Omega \) does not have any subgroup of order 2, 4, 5 and 10 and so \(|\Omega |\le 57-(n_2+n_4+n_5+n_{10})=25\), by Fact 1. Since \(\Omega \) does not have two subgroups of order 3, we have \(|\Omega |\le 25-9=16\). Also if \(X\in \Omega \), then \(H_3\subseteq X\), by Lemma 2.5(iii). By Fact 4, \(H_3\) is contained in only one subgroup of order 6. Therefore, \(|\Omega |\le 16-9=7\), which is impossible.

Case 3. In this case, we show that \(\Omega \) does not have any member of size 4. Assume that there exists \(H_4\in \Omega \) with \(|H_4|=4\). By Case 1, \(\Omega \) does not have any subgroup of order 2. Since \(n_2=15, n_3=10, n_4=5\) and \(n_5=6\), we have \(|\Omega |\le 57-(15+10+4+6)=22\). Since every clique can contain at most two subgroups of order 6 by fact 4, we have \(|\Omega |\le 22-(n_6-2)=14\). Note that \(H_4\) is contained in at least one maximal subgroup, which is certainly of order 12. By Fact 3, there is only one subgroup K of order 12 such that \(H_4\subset K\) and hence \(|\Omega |\le 14-4=10\), which is a contradiction.

Case 4. In this case, we show that \(\Omega \) does not have any member of size 5. Assume that there exists \(H_5\in \Omega \) with \(|H_5|=5\). Since \(n_2=15, n_3=n_6=10, n_4=n_{12}=5\) and \(n_5=6\), we have \(|\Omega |\le 57-(15+ 20+10+5)=7\), a contradiction.

Case 5. In this case, we first show that \(\Omega \) has at least one member of size 6 and then prove that \(|\Omega |=13\). Suppose, for a contradiction, that \(\Omega \) does not have any member of size 6. By Cases 1–4, we conclude that \(\Omega \) does not have any subgroup of order 2, 3, 4 and 5. By Fact 1, every member of \(\Omega \) has size either 10 or 12 and so \(|\Omega |\le n_{10}+ n_{12}=11\), a contradiction. It follows that there exists \(H_6\in \Omega \) with \(|H_6|=6\). Again, by Cases 1-4, we have \(|\Omega |\le n_6+ n_{10}+ n_{12}=16\). On the other hand \(\Omega \) has at most two members of size 6 by Fact 4. Thus \(|\Omega |=13\) by Fact 4 and this completes the proof. \(\square \)

In the proofs of Lemmas 3.3, 3.4 and 3.5, we need the following result about the Sylow p-subgroups of \(PSL(2, p^n)\), where p is a prime number and n is a positive integer.

Proposition 3.2

Let \(G=PSL(2, p^n)\). Then a Sylow p-subgroup P of G is elementary abelian of order \(p^n\) and the number of Sylow p-subgroups of G is \(p^n+1\) (or equivalently, we have \(|G: N_G(P)|=p^n+1\)).

Proof

See Satz 8.2, p. 191 of [13]. \(\square \)

Lemma 3.3

For any odd prime p, we have \(\omega (\Gamma (PSL(2, 2^p))>13\).

Proof

If \(p=3\), then \(G=PSL(2,8)\) has a maximal subgroup M of index 28 in G. Since G is simple, M is not normal in G and so \(N_G(M)=M\). It follows that M has 28 conjugates in G,  say \(M_1=M, M_2,\ldots , M_{28} \). It is easily checked by GAP [23] that \(|M_i\cap M_j|=2\), for all distinct ij and so \(\omega (\Gamma (PSL(2, 2^m))\ge 28\), as required.

If \(p\ge 5\), then \(q=2^p\ge 32\). By Proposition 3.2, we have \(|G: N_G(P)|=q+1\), where \(P\in Syl_2(G)\). Since \(N_G(N_G(P))=N_G(P)\), the number of conjugates of \(N_G(P)\) is \(q+1\). It follows that \(N_G(P)\) has \(q+1\) conjugates in G. If \(N_G(P)\ne N_G(P)^g\), for some \(g\in G\), then \(|G: N_G(P)\cap N_G(P)^g|\le (q+1)^2\), by lemma 2.1. Since \(|G|=\frac{q(q^2-1)}{2}\), we have \(|N_G(P)\cap N_G(P)^g|\ne 1\). Therefore, \(\{N_G(P)^x:~x\in G\}\) is a clique in the \(\Gamma (G)\) and so \(\omega (\Gamma (G))\ge q+1>13\), as claimed. \(\square \)

Lemma 3.4

\(\omega (\Gamma (PSL(2, 3^p))>13\), for any odd prime p.

Proof

By Proposition 3.2, we have \(|G: N_G(P)|=3^p+1\), where \(P\in Syl_3(G)\). Similar to the proof of Lemma 3.3, \(\{N_G(P)^x:~x\in G\}\) is a clique in the \(\Gamma (G)\) and so \(\omega (\Gamma (G))\ge 3^p+1>13\) and this completes the proof. \(\square \)

Lemma 3.5

\(\omega (\Gamma (PSL(2, p))>13\), where \(p > 3\) is a prime with \(p^2 + 1\) divisible by 5.

Proof

If \(G=PSL(2,7)\), then G has 14 (maximal) subgroups of index 7 ( one can check it by GAP [23]), say \(M_1, M_2,\ldots , M_{14}\). Since \(|G|=168\) and \(|G:M_i\cap M_j|\le 49\), for every \(1\le i\ne j\le 14\), we have \(|M_i\cap M_j|\ne 1\), by Lemma 2.1 and so \(\{M_1, \ldots , M_{14}\}\) is a clique in the \(\Gamma (G)\) and so \(\omega (\Gamma (G))\ge 14\), as wanted.

Now if \(p\ge 13\), then by the similar argument in the proof of Lemmas 3.3 and 3.4, we see that \(\omega (\Gamma (G))\ge p+1>13\) and the proof is complete. \(\square \)

Lemma 3.6

We have \(\omega (\Gamma (PSL(3, 3))>13\).

Proof

By GAP [23], G has 26 maximal subgroups of index 13, say \(M_1, M_2,\ldots , M_{26}\). Since \(|G|=5616\) and \(|G:M_i\cap M_j|\le 169\), for every \(1\le i\ne j\le 26\), we have \(|M_i\cap M_j|\ne 1\) and so \(\omega (\Gamma (G))\ge 26\), as required. \(\square \)

Lemma 3.7

We have \(\omega (\Gamma (Sz(2^m)))>13\), where m is an odd prime.

Proof

Suppose that \(q=2^m\), \(G=Sz(q)\) and \(F\in Syl_2(G)\). Then it is well-known that \(|F|=q^2\) and Z(F) is elementary abelian of order q. It follows from Lemma 5.9 in Chapter XI of [13] that \(|C_F(g): Z(F)|=2\), for all \(g\in F{\setminus } Z(F)\) and so \(C_F(g)\) is abelian. Assume that \(\{C_F(x_1), C_F(x_2),\ldots , C_F(x_n)\}\) is the set of all proper centralizers of elements in F. Then \(F=\cup _{i=1}^nC_F(x_i)\) and \(C_F(x_i)\cap C_F(x_j)=Z(F)\), for each \(i\ne j\). Since \(|F|=q^2, |Z(F)|=q\) and \(|C_F(x_i)|=2q\), we have \(n=q-1\). Also Z(F) has \(q-1\) maximal subgroups, by Lemma 2.6 and the intersection of any pair of these maximal subgroups are non-trivial, by Lemma 2.1. If \(Z_1, \ldots , Z_{q-1}\) are maximal subgroups of Z(F), then \(\{Z_1, \ldots , Z_{q-1}, C_F(x_1), \ldots , C_F(x_{q-1})\}\) is a clique in \(\Gamma (G)\). Since \(q\ge 8\), we have \(\omega (\Gamma (Sz(2^m)))\ge 2(q-1)\ge 14\). This completes the proof. \(\square \)

Now we are ready to prove the first main result:

Proof of Theorem 1.1

Suppose, on the contrary, that there exists a non-solvable finite group G of the least possible order with \(\omega (\Gamma (G))<13\). If there exists a non-trivial proper normal subgroup N of G, then \(\omega (\Gamma (N))<13\) and \(\omega (\Gamma (\frac{G}{N}))<13\), by Lemma 2.5(i-ii). By minimality of |G|, we have \(\frac{G}{N}\) and N are solvable, which implies that G is solvable, a contradiction. Thus G is simple and so by Theorem 1 of [4] and Lemma 2.5(i), G is a minimal simple group. By Thompson’s classification of minimal simple groups [24], G is isomorphic to one of the following simple groups: \(A_5\), \(PSL(2, 2^m)\), where m is an odd prime; \(PSL(2, 3^m)\), where m is an odd prime; PSL(2, p), where \(p > 3\) is a prime with \(p^2 + 1\) divisible by 5; PSL(3, 3); or \(Sz(2^m)\), where m is an odd prime. By Lemmas 3.1, 3.3, 3.4, 3.5, 3.6 and 3.7, we have \(\omega (\Gamma (G))\ge 13\), which is impossible. \(\square \)

Proof of Corollary 1.2

If \(G\cong A_5\), then \(\omega (\Gamma (G))=13\), by Lemma 3.1. Conversely, assume that G is non-solvable with \(\omega (\Gamma (G))=13\). If M is a maximal subgroup of G, then \(\omega (\Gamma (M))<13\), by Lemma 2.5(i). So M is solvable by Theorem 1.1. Thus every maximal subgroup of G is solvable and so G is a minimal non-solvable group. Therefore, G is a minimal simple group. By the proof of Theorem 1.1 and Lemmas 3.1, 3.3, 3.4, 3.5, 3.6 and 3.7, we conclude that \(G\cong A_5\), as desired. \(\square \)

4 Groups G with \(2\le \omega (\Gamma (G))\le 4\)

In this section, we prove Theorems 1.4, 1.5 and 1.6. First, we obtain \( \omega (\Gamma (G))\), for some groups which are needed later.

Lemma 4.1

For any distinct primes pq and r and positive integer n, we have

  1. (1)

    \(\omega (\Gamma (C_{p^n}))=n-1\).

  2. (2)

    \( \omega (\Gamma (C_p\times C_p\times C_p))=p^2+p+1\).

  3. (3)

    \(\omega (\Gamma (C_{pqr}))=\omega (\Gamma (C_{p^2q}))=3\).

  4. (4)

    \(\omega (\Gamma (C_{p^2}\times C_p))= \omega (\Gamma (C_{pq}\times C_p))=p+2\).

  5. (5)

    If G is a non-abelian group of order \(p^3\), then \(\omega (\Gamma (G))=p+2\).

Proof

  1. (1)

    If \(G=C_{p^n}\), then G has a unique subgroup \(H_i\) of order \(p^i\), for each \(0\le i\le n\) and so \(H_i<H_{i+1}\), for every i. Therefore, \(\{H_1, H_2,\ldots , H_{n-1}\}\) is the unique clique in \(\Gamma (G)\) with maximum size, as required.

  2. (2)

    Suppose that \(G=C_p\times C_p\times C_p\) and \(\Omega \) is a clique in \(\Gamma (G)\). If \(H\in \Omega \), with \(|H|=p\), then, by Lemma 2.5 (iii), \(H<K\), for every \(K\in \Omega {\setminus } \{H\}\) and so \(|K|=p^2\). Since \(\frac{G}{H}\cong C_p\times C_p\) and the number of subgroups of \(C_p\times C_p\) of order p is \(p+1\), by the proof of Lemma 2.5(ii),the number of proper subgroups of G containing H is \(p+2\) and so \(|\Omega |\le p+2\). Now suppose that \(\Omega \) does not contain any subgroup of G of order p. Then \(|K|=p^2\), for every \(K\in \Omega \) and so K is a maximal subgroup of G. Since \(\Phi (G)=1\), the number of maximal subgroups of G is \(\frac{|\frac{G}{\Phi (G)}|-1}{p-1}=p^2+p+1\), by Lemma 2.6. Therefore \(|\Omega |\le p^2+p+1\). It follows from Lemma 2.1 that the intersection of any two distinct maximal subgroups of G are non-trivial. This implies that all maximal subgroups of G forms a clique with maximum size in \(\Gamma (G)\), as wanted.

  3. (3)

    The proof is clear.

  4. (4)

    Suppose that \(G=C_{p^2}\times C_p\) and \(\Omega \) is a clique in \(\Gamma (G)\). If \(H\in \Omega \) such that \(|H|=p\), then \(H<K\), for every \(K\in \Omega \setminus \{H\}\), by Lemma 2.5 (iii) and so \(|K|=p^2\). Since the number of proper subgroups of G containing H is at most \(p+2\) and so \(|\Omega |\le p+2\). On the other hand \(|\Phi (G)|=p\) and so the number of maximal subgroups of G is \(p+1\). It follows that these maximal subgroups of G together with \(\Phi (G)\) forms a clique with maximum size in \(\Gamma (G)\) and so \(\omega (\Gamma (G))=p+2\), as desired.

Suppose that \(G=C_{pq}\times C_p\) and \(\Omega \) is a clique in \(\Gamma (G)\). If \(\Omega \) does not contain any subgroup of G of prime order, then \(\Omega =\{T\}\) such that \(|T|=p^2\). If there is a subgroup H of G such that \(H\in \Omega \) and \(|H|=p\), then \(\frac{G}{H}\cong C_{pq}\) and so \(|\Omega |\le 3\). Finally assume that \(Q\in \Omega \) such that \(|Q|=q\). Then \(Q\le K\), for every \(K\in \Omega \) by Lemma 2.5 (iii). On the other hand the number of proper subgroups of G containing Q is \(p+2\) and so \(|\Omega |\le p+2\). Hence \(\omega (\Gamma (G))=p+2\), as required.

  1. (5)

    The proof is similar to the first case of (4).

\(\square \)

Lemma 4.2

Let \(G=N\rtimes H\) be a Frobenius group such that N is a minimal normal subgroup of G and H is a Frobenius complement. Also assume that pq and r are distinct primes.

  1. (1)

    If \(N\cong C_p\) and \(H\cong C_{q^2}\) or \(C_{qr}\), then \(\omega (\Gamma (G))=3\).

  2. (2)

    If \(N\cong C_p\times C_p\) and \(H\cong C_q\), then \(\omega (\Gamma (G))=2\).

  3. (3)

    If \(N\cong C_p\times C_p\) and \(H\cong C_{q^2}\) such that \(q \not \mid p-1\), then \(\omega (\Gamma (G))=3\).

  4. (4)

    If \(N\cong C_p\times C_p\) and \(H\cong C_{qr}\) such that \(q\not \mid p-1\) and \(r\not \mid p-1\), then \(\omega (\Gamma (G))=4\).

Proof

(1) Suppose that \(N\cong C_p\) and \(H\cong C_{q^2}\) and \(1<H_1<H\). Since \(G=N\rtimes H\) is a Frobenius group, \(H\cap H^x=1\) and so \(H_1\cap H_1^x=1\), for every \(1\ne x\in N\). Therefore, G has p subgroups of order q and so all of them are contained in the non-abelian subgroups \(K=NH_1\). It follows that \(K=NH_1\) is the unique subgroup of G of order pq. It is clear that \(K\cap T\ne 1\), for every non-trivial proper subgroup T of G. Thus K belongs to any clique with maximum size. Note that every clique contains at most one subgroup of order \(q^2\) and also at most one subgroup of order q. If N belongs to a clique \(\Omega \), then \(H_1^x\) and \(H^x\) do not belong to \(\Omega \), for every \( x\in N\). Thus \(\{H_1^x, H^x, K\}\) is a clique with maximum size in \(\Gamma (G)\), for every \(x\in N\), as required.

Now suppose that \(H\cong C_{qr}\). Then H contains (unique) subgroups Q and R of order q and r respectively. Since \(G=N\rtimes H\) is a Frobenius group, we have \(H\cap H^x=1\), \(Q\cap Q^x=1\) and \(R\cap R^x=1\), for every \(1\ne x\in N\). Therefore \(|Syl_q(G)|=|Syl_r(G)|=|G:N_G(H)|=|G:H|=p\). Since G is Frobenius, \(Z(G)=1\) and so NQ and NR are non-abelian. Therefore NQ and NR are the only subgroups of G of orders pq and pr, respectively. Note that the only proper subgroups of G containing properly \(Q^x\), for some \(x\in N\) are \(H^x\) and NQ, the only proper subgroups of G containing properly \(R^x\), for some \(x\in N\) are \(H^x\) and NR and also the only proper subgroups of G containing properly N are NQ and NR. Now suppose that \(\Omega \) is a clique in \(\Gamma (G)\) with maximum size.

If \(Q^x\in \Omega \), for some \(x\in N\), then \(\{Q^x, H^x, NQ\}= \Omega \). If \(R^x\in \Omega \), for some \(x\in N\), then \(\{R^x, H^x, NR\}= \Omega \). If \(N\in \Omega \), then \(\{N, NQ, NR\}= \Omega \). If none of the three cases does not occur, for \(\Omega \), then \(\Omega =\{H^x, NQ, NR\}\), for some \(x\in N\). So we conclude that \(\Omega \) has 3 members, as wanted.

(2) Since N is a minimal normal abelian subgroup of G, we conclude that \(H^x\) is a maximal subgroup of G, for all \(x\in N\) by Proposition 2.3 and so G does not have any subgroup of order pq. Therefore,

$$\begin{aligned} \mathcal {S}(G)=\{1, G, N, N_1, \ldots , N_{p+1}, H^x:~~x\in N\}, \end{aligned}$$

where \(|N_i|=p\), for each i. Since \(G=N\rtimes H\) is Frobenius, \(\{N_i, N\}\) is a clique with maximum size in \(\Gamma (G)\), for each i, as wanted.

(3) Since \(G=N\rtimes H\cong (C_p\times C_p)\rtimes C_{q^2}\) is Frobenius with complement H, the group G has \(p+1\) subgroups of order p (contained in N), say \(N_1,\ldots , N_{p+1}\) and \(p^2\) subgroups \(H_1^x\) of order q, for each \(x\in N\), where \(1<H_1<H\). Note that H is a maximal subgroup of G by Proposition 2.3. Therefore, G does have any subgroup of order \(pq^2\). Clearly, \(H_1^x\le NH_1\), for every \(x\in N\) and so \(NH_1\) is the unique non-abelian subgroup of G of order \(p^2q\). If G has a subgroup K of order pq, then K is not abelian since \(C_G(x)=N\), for \(1\ne x\in N\) by Exercise 8.5 (5) of [18]. It follows that \(q|p-1\), which is contrary to our assumption in this part. Hence the order of a non-trivial proper subgroup of G belongs to \(\{p,p^2, q, p^2q, q^2\}\) and so we have

$$\begin{aligned} \mathcal {S}(G)=\{1, G, N, N_1, \ldots , N_{p+1}, H^x, H_1^x, NH_1:~~x\in N\}. \end{aligned}$$

Since \(N\cap H^y=H\cap H^x=H_1\cap H_1^x=N_i\cap N_j=1\), for every \(y\in N\), \(x\in N{\setminus } \{1\}\) and \(i\ne j\), we have \(\{N_i, N, NH_1\}\) is a clique with maximum size in \(\Gamma (G)\), for each i. Also \(\{H_1^x, H^x, NH_1\}\) is a clique with maximum, for every \(x\in N\). This completes the proof.

(4) The proof is similar to the part (3).

\(\square \)

Proof of Theorem 1.4:

Since \(\omega (G)=2\), G contains two proper subgroups \(H_1\) and \(H_2\) such that \(H_1\cap H_2\ne 1\) and so \(\{H_1, H_2, H_1\cap H_2\}\) is a clique in \(\Gamma (G)\), which implies that either \(H_1<H_2\) or \(H_2<H_1\). Without loss of generality, assume that \(H_1<H_2\). Since \(\omega (G)=2\), we have \(|H_1|=p\) is prime, \(H_1\) is a maximal subgroup of \(H_2\) and \(H_2\) is a maximal subgroup of G. Now we claim that \(|H_2: H_1|\) and \(|G: H_2|\) are primes.

If \(H_1\lhd H_2\) and \(H_2\lhd G\), then the claim is valid. If \(H_1\) is not normal in \(H_2\), then \(N_{H_2}(H_1)=H_1\) and since \(|H_1|\) is prime, we have \(H_1\cap H_1^x=1\), for all \(x\in H_2{\setminus } H_1\). It follows from Theorem 2.2 that \(H_2\) is Frobenius and so there is a normal subgroup N such that \(H_2=N\rtimes H_1\). If |N| is not prime, then N has a proper non-trivial subgroup \(N_1\) and so \(\{N_1, N, H_2\}\) is a clique in \(\Gamma (G)\) of size 3, which is a contradiction. Therefore, \(|N|=|H_2: H_1|\) is prime. Now if \(H_2\lhd G\), then \(|G: H_2|\) is prime by maximlity of \(H_2\) in G. So assume that \(H_2\) is not normal in G. Then, by a similar argument to \(H_1\ntriangleleft H_2\), we conclude that G is Frobenius and \(|G:H_2|\) is prime. Thus \(|G|=|H_1||H_2: H_1||G: H_2|=pqr\), for some primes pq and r.

First suppose that G is abelian. If pq and r are distinct, then \(G\cong C_{pqr}\) and so by Lemma 4.1(3) we have \(\omega (\Gamma (G))=3\), a contradiction. If \(p=r\ne q\), then \(G\cong C_{p^2q}\) or \(C_p\times C_p\times C_q\), which implies that \(\omega (\Gamma (G))\ne 2\) by Lemma 4.1(3-4). If \(p=q=r\), then \(G\cong C_{p^3}, C_{p^2}\times C_p\) or \(C_p\times C_p\times C_p\). Since \(\omega (G)=2\), we have \(G\cong C_{p^3}\) by Lemma 4.1.

Now suppose that G is not abelian. If \(p=q=r\), then \(|G|=p^3\) and we have a contradiction by Lemma 4.1(5). If pq and r are distinct, then at least one of Sylow subgroups of G is normal, which implies that \(\omega (\Gamma (G))\ge 3\), a contradiction. Finally suppose that \(p=r\ne q\). If \(Syl_q(G)=\{Q\}\) and \(P\in Syl_p(G)\), then \(\{P_1, P, P_1Q\}\) is a clique in \(\Gamma (G)\), where \(P_1\) is a proper non-trivial subgroup of P, again a contradiction. Hence we have \(Syl_p(G)=\{P\}\). If \(Q\in Syl_q(G)\), then \(G=P\rtimes Q\). Since \(\omega (\Gamma (G))=2\), we see that P must be a minimal normal subgroup of G and so \(P\cong C_p\times C_p\), which implies that Q is a maximal subgroup of G Proposition 2.3. Since Q is not normal in G, we have \(N_G(Q)=Q\). It follows \(Q\cap Q^g=1\), for every \(g\in G{\setminus } Q\) and so G is a Frobenius group. This completes the proof.

The converse follows from Lemmas 4.1(1) and 4.2(2). \(\square \)

Proposition 4.3

Let G be a finite abelian group. Then \(\omega (\Gamma (G))=3\) if and only if G is isomorphic to one of the following groups:

$$\begin{aligned} C_{p^4},\quad C_{p^2q},\quad C_{pqr}, \end{aligned}$$

where pqr are distinct primes.

Proof

Assume that \(\omega (\Gamma (G))=3\). We consider two cases:

  1. (i)

    Suppose that G is a p-group, for some prime p. Since \(\omega (\Gamma (G))=3\), we have \(|G|\ge p^3\). If \(\Phi (G)=1\), then G is elementary abelian and so G has a subgroup isomorphic to \(E=C_p\times C_p\times C_p\). It follows from Lemma 4.1(2) that \(\omega (\Gamma (E))= p^2+p+1>3\), a contradiction. Hence \(\Phi (G)\ne 1.\) Moreover, the number of maximal subgroups of G is \(t=\frac{|\frac{G}{\Phi (G)}|-1}{p-1}\), by Lemma 2.6. Therefore, either \(t=1\), which follows that \(G\cong C_{p^4}\), by Lemma 4.1(1) or G has at least \(p+1\) maximal subgroups, say \(M_1, M_2,\ldots , M_{p+1}.\) By Lemma 2.1, \(\{M_1, M_2,\ldots , M_{p+1}, \Phi (G)\}\) is a clique and so \(\omega (\Gamma (G))\ge p+2\), which is impossible.

  2. (ii)

    Now suppose that |G| has at least two distinct prime divisors, p and q. Then G has two maximal subgroups \(M_1\) and \(M_2\) such that \(|G:M_1|=p\) and \(|G:M_2|=q\). Since \(\omega (\Gamma (G))=3\), we have \(M_1\cap M_2\ne 1\) and so \(\{M_1, M_2, M_1\cap M_2\}\) is a clique in \(\Gamma (G)\). Note that G has a maximal subgroup different from \(M_1\) and \(M_2\), say \(M_3\) since \(G\ne M_1\cup M_2\). We conclude that \(\{M_1, M_2, M_3\}\) is a clique in \(\Gamma (G)\). Since \(\omega (\Gamma (G))=3\), we have \(\cap _{i=1}^3M_i=1\), which follows that \(G\hookrightarrow \frac{G}{M_1}\times \frac{G}{M_2}\times \frac{G}{M_3}\). Consequently \(|G|=pqr\), where r is prime. By Lemma 4.1(3-4), we have \(G\cong C_{pqr}\) or \(C_{p^2q}\), as wanted.

The converse of the lemma follows from Lemma 4.1. \(\square \)

Proposition 4.4

Let G be a finite non-abelian group. Then \(\omega (\Gamma (G))=3\) if and only if \(G=N\rtimes H\) is Frobenius whose kernel N is the unique minimal normal subgroup of G and we have either

  1. (1)

    \(N\cong C_p\) and \(H\cong C_{q^2}\) or \(C_{qr}\) or

  2. (2)

    \(N\cong C_p\times C_p\), \(H\cong C_{q^2}\) and G does not have any subgroup of order pq, where pqr are distinct primes.

Proof

Suppose that \(\omega (\Gamma (G))=3\). Since G is not cyclic, G has at least three maximal subgroups \(M_1, M_2\) and \(M_3\). By hypothesis, \(\cap _{i=1}^3M_i=1\). If \(M_i\lhd G\), for each i, then \(|G: M_i|\) is prime, for each i and so G is abelian, a contradiction. Without loss of generality, assume that \(M:=M_1\) is not normal in G. Then \(N_G(M)=M\) and \(|G: M|\ge 3.\) Thus M has at least three conjugates \(M, M^x, M^y\), for some \(x, y\in G\). If \(Core_G(M)\ne 1\), then \(\omega (\Gamma (G))\ge 4\), which is a contradiction. So \(Core_G(M)=1\), which implies that G is a primitive group. It follows from Theorem 1.1 that G is solvable and by Theorem 2.4, G has a unique minimal normal subgroup N such that \(C_G(N)=N\) and \(G=N\rtimes M.\) Since N is a principal factor of G, we have N is elementary abelian of order \(p^n\), for some prime p, by 5.4.3 of [18]. Since \(\omega (\Gamma (G))=3\), we have \(n=1\) or 2, by Lemmas 4.1(2) and 2.5. Also note that if \(M\cap M^g\ne 1\), for some \(g\in G{\setminus } M\), then \(\{M, M^g, M\cap M^g, N(M\cap M^g)\}\) is a clique in \(\Gamma (G)\), a contradiction. Thus G is a Frobenius group.

Subcase 1. If \(n=1\), then \(G=N\rtimes M\), where \(N\cong C_p\). Since \(Aut(C_p)\) is cyclic of order \(p-1\), we have M is cyclic of order divisor of \(p-1\). It follows from Lemma 2.5(i) that \(\omega (\Gamma (M))< \omega (\Gamma (G))=3\). If \(\omega (M)=2\), then by Theorem 1.4, we have \(M\cong C_{q^3}\), for some prime q. Therefore, M has two subgroups H and K of orders q and \(q^2\) respectively and so \(\{H, K, M, NH\}\) is a clique, which is impossible. Hence \(\omega (\Gamma (M))=1\), which follows that \(M\cong C_{qr}\) or \(C_{q^2}\), by Remark 1.3, as required.

Subcase 2. If \(n=2\), then \(G=N\rtimes M\), where \(N\cong C_p\times C_p\). Therefore, |M| is not prime, by hypothesis and Theorem 1.4. We conclude that \(|\mathcal {S}(M)|\ge 3\). Since \(\omega (\Gamma (G))=3\) and \(\omega (\Gamma (N))=1\), we have \(|\mathcal {S}(M)|\le 3\), by Lemma 2.5(ii). Therefore, \(M\cong C_{q^2}\), as wanted. It remains to show that G does not have any (non-abelian) subgroup of order pq.

Suppose, for a contradiction, that G has a subgroup K of order pq. Since \(C_G(x)=N\), for every \(1\ne x\in N\), by Exercise 8.5 (5) of [18], we have K is non-abelian and so there are subgroups \(1<N_1<N\) and \(1<M_1<M\) such that \(K=N_1\rtimes M_1\). It follows that \(\{N_1, N, K, NM_1\}\) is a clique in \(\Gamma (G)\) with four elements, which is impossible.

The converse follows from Lemmas 4.2(1) and 4.2(3). \(\square \)

Proof of Theorem 1.5

The proof follows from Propositions 4.3 and 4.4. \(\square \)

Proposition 4.5

Let G be a finite abelian group. Then \(\omega (\Gamma (G))=4\) if and only if G is isomorphic to one of the follwing groups:

$$\begin{aligned} C_{p^5},~~~~~C_2\times C_2\times C_q,~~~~~C_2\times C_4, \end{aligned}$$

where p and q are primes and q is odd.

Proof

Assume that \(\omega (\Gamma (G))=4\). Then, by Remark 1.3 and Lemma 4.1, |G| has at least three prime divisors (not necessarily distinct), say pq and r. Note that the number of maximal subgroups of G is either one or at least three. So we consider three cases in terms of the number of maximal subgroups of G in the following:

Case 1. If G has at least four maximal subgroups, \(M_i\), for \(1\le i\le 4\), then \(M_i\cap M_j\ne 1\) and since \(\omega (\Gamma (G))=4\), we have \(\cap _{i=1}^4M_i=1\) and so |G| has at most four prime divisors (not necessarily distinct). Therefore, |G| divides pqrs, for some prime s.

First suppose that \(|G|=pqrs\). If all prime divisors of |G| are distinct, then G is cyclic of order pqrs and so \(\omega (\Gamma (G))>4\), a contradiction. If \(|G|=p^2qr\) such that pqr are distinct, then \(\{H, P, HQ, HR, PQ, PR\}\) is a clique in \(\Gamma (G)\), where \(|H|=p\), \(P\in Syl_p(G), Q\in Syl_q(G)\) and \(R\in Syl_r(G)\), again a contradiction. By a similar argument, we see that \(|G|\ne p^2q^2, p^3q, p^4\), where pq are distinct. So such group G does not exist.

Now suppose that \(|G|=pqr\). By the number of maximal subgroups of G, we have G is not cyclic. Also \(G\ncong C_p\times C_p\times C_p\), by Lemma 4.1(2). Hence \(G\cong C_{p^2}\times C_p\) or \(C_p\times C_p\times C_q\). It follows from Lemma 4.1(4) that \(p=2\) and q is odd. But the number of maximal subgroups \(C_{4}\times C_2\) is 3, a contradiction. Thus \(G\cong C_2\times C_2\times C_q\), as wanted.

Case 2. Suppose that G has exactly three maximal subgroups. If G is cyclic, then |G| has three distinct prime divisors and so \(G\cong C_{pqr}\). So we have a contradiction, by Lemma 4.1(3) since \(\omega (\Gamma (G))=4\). Thus G is not cyclic and then G has a Sylow t-subgroup, which is not cyclic, for some prime t. Without loss of generality, assume \(t=p\) and \(P\in Syl_p(G)\). Therefore, P (or equivalently G) has at least \(p+1\) maximal subgroups. By hypothesis in this case, we have \(p=2\) and G must be a 2-group. On the other hand, the number of maximal subgroups of G is \(|\frac{G}{\Phi ((G)}|-1\), which implies that \(\frac{G}{\Phi ((G)}\cong C_2\times C_2\). Since \(|\mathcal {S}(C_2\times C_2)|=5\) and \(\omega (\Gamma (G))=4\), we have \(\omega (\Gamma (\Phi (G)))=0\), by Lemma 2.5(ii). Thus \(|\Phi (G)|=2\) and so \(|G|=8\). Hence \(G\cong C_4\times C_2\), as desired.

Case 3.Suppose that G has exactly one maximal subgroup. Then G is cyclic of order the power of a prime p and so \(G\cong C_{p^5}\), by Lemma 4.1(1). This completes the proof.

The converse is clear, by Lemma 4.1.

\(\square \)

Proposition 4.6

Let G be a finite non-abelian group. Then \(\omega (\Gamma (G))=4\) if and only if G is one of the follwing groups:

  1. (i)

    \(G\cong D_8,~Q_8.\) or

  2. (ii)

    \(G=N\rtimes H\) is Frobenius whose kernel N is the (unique) minimal normal subgroup of G such that \(N\cong C_p\times C_p\), \(H\cong C_{qr}\) and G does not have subgroups of orders pr and pq

Proof

Assume that \(\omega (\Gamma (G))=4\). Suppose first that every maximal subgroup is normal in G. Then G is nilpotent. Since G is non-abelian, there is a Sylow p-subgroup P of G such that P is non-abelian, for some prime p. Therefore, \(\Phi (P)\ne 1\) and so \(\omega (\Gamma (P))\ge m(P)+1\ge p+1+1\), where m(P) is the number of maximal subgroups of P. Thus, by Lemma 2.5(i), we see that \(4=\omega (\Gamma (G))\ge \omega (\Gamma (P))\ge 4\), which implies that \(G=P\) and then we have \(p=2\) and \(m(P)=p+1=3\). Now if \(M_1, M_2\) and \(M_3\) are maximal subgroups of G, then \(\{\Phi (G), M_1, M_2, M_3\}\) is a clique with maximum size and so \(|\Phi (G)|=2\) and \(M_i\cap M_j=\Phi (G)\), for every \(i\ne j\le 3\). Since \(|G: M_i|=2\), for each i, we have \(|G|=8\). Thus \(G\cong D_8\) or \(Q_8\), as wanted.

Now assume that G has a maximal subgroup M, which is not normal in G. Then \(N_G(M)=M\) and \(|G:M|\ge 3\). We claim that \(Core_G(M)=1\).

Suppose, for a contradiction, that \(M_G:=Core_G(M)\ne 1\). If M has at least four conjugates, say \(M, M^x, M^y, M^z\), for some \(x, y, z\in G\), then \(\omega (\Gamma (G))\ge 5\), a contradiction. So we have \(|G:M|=3\). It follows that \(\frac{G}{M_G}\hookrightarrow S_3\). If \(\frac{G}{M_G}\cong S_3\), then by Lemma 2.5(ii), \(\omega (\Gamma (G))\ge |\mathcal {S}(S_3)|-1=5\), which is a contradiction. Since 3 divides \(|G: M_G|\), we have \(M_G=M\) and so \(M\lhd G\), a contradiction. This implies that \(M_G=1\), as claimed. Thus G is primitive. By Theorems 1.1, G is solvable and by Theorem 2.4, G has a unique minimal normal subgroup N such that \(C_G(N)=N\) and \(G=N\rtimes M\). It follows that N is elementary abelian p-group, for some prime p. If \(|N|\ge p^3\), then N contains a subgroup \(E\cong C_p\times C_p\times C_p\) and so \(\omega (\Gamma (E))>4\), by Lemma 4.1(2), which is impossible, by Lemma 2.5(i). Thus \(|N|=p\) or \(p^2\).

If \(|N|=p\), then, by Normalizer-Centralizer Theorem we have \(\frac{G}{C_G(N)}=\frac{G}{N}\hookrightarrow Aut(C_p)\cong C_{p-1}\) and so M is cyclic. It follows from Lemma 2.5(i) that \(\omega (\Gamma (M))<4\). If \(\omega (\Gamma (M))=2\) or 3, then by Theorems 1.4 and 4.3, we have \(M\cong C_{q^3},~C_{q^2r},~C_{q^4}\) or \(C_{qrs}\). Thus M has at least two proper subgroups of different orders, which are divisible by q, say \(M_1, M_2\) and so \(\{M_1, M_2, NM_1, NM_2, M\}\) is a clique in \(\Gamma (G)\), a contradiction. Also if \(\omega (\Gamma (M))=1\), then \(M\cong C_{q^2}\) or \(C_{qr}\), where qr are distinct primes. Since M is cyclic and maximal subgroup of G, we have \(M\cap M^x\lhd \langle M, M^x\rangle =G\), for any \(x\in G{\setminus } M\). Now since G has a unique minimal subgroup N, we see that \(M\cap M^x=1\), for every \(x\in G{\setminus } M\) and then G is Frobenius. Therefore, \(G=N\rtimes M\), where \(N\cong C_p\) and \(M\cong C_{q^2}\) or \(C_{qr}\), which follows \(\omega (\Gamma (G))=3\) by Theorem 4.4, another contradiction. Finally if \(\omega (\Gamma (M))= 0\), then \(|M|=q\) is prime and so \(G\cong C_p\rtimes C_q\). It follows that \(\omega (\Gamma (G))=1\), by Remark 1.3, our final contradiction.

Now suppose that \(N\cong C_p\times C_p\). Then \(\omega (\Gamma (M))<\omega (\Gamma (G))=4\), by Lemma 2.5(i). Also by Remark 1.3 and Theorem 1.4, \(\omega (\Gamma (M))\ne 0\). If \(\omega (\Gamma (M))=1\), then \(M\cong C_{q^2}, C_{qr}, C_q\times C_q \) or \(C_q\rtimes C_r\), where q and r are distinct primes. In two last cases, M (or equivalently \(\frac{G}{N}\)) has \(q+1\) non-trivial proper subgroups and so \(\{H, N, K_1, K_2, \ldots , K_{q+1}\}\) is a clique in \(\Gamma (G)\), where \(1<H<N\) and \(K_i\) is a proper subgroup of G containing N, for each i. It follows that \(\omega (\Gamma (G))\ge q+3\), a contradiction. If \(M\cong C_{q^2}\), then \(M\cap M^x\lhd \langle M, M^x\rangle =G\), for any \(1\ne x\in N\), we have \(M\cap M^x=1\), for every \(1\ne x\in N\) and then G is Frobenius. By Theorem 4.4, we have \(\omega (\Gamma (G))=3\). Hence \(M\cong C_{qr}\) and similarly G is Frobenius, as wanted.

If \(\omega (\Gamma (M))=2\) and \(\{H_1, H_2\}\) is a clique in \(\Gamma (M)\) such that \(H_1<H_2\), then \(\{H_1, H_2, M, NH_1, NH_2\}\) is a clique in \(\Gamma (G))\), which is impossible.

If \(\omega (\Gamma (M))=3\) and \(\{M_1, M_2, M_3\}\) is a clique in \(\Gamma (M))\), then

\(\{M_1, M_2, M_3, M, NM_1\}\) is a clique in \(\Gamma (G))\), a contradiction. It remains to prove that G does not have subgroups of orders pq and pr.

Suppose, for a contradiction, that G has a subgroup K of order pq. Then there are subgroups \(N_1<N\) and \(Q\in Syl_q(G)\) such that \(K=N_1\rtimes Q\). It follows that \(\{N_1, N, K, NQ, NR\}\) is a clique in \(\Gamma (G)\), where R is a Sylow r-subgroup of G. This is a contradiction with our assumption. Similarly G does not have any subgroup of order pr.

The converse of the lemma follows from Lemmas 4.1(5) and 4.2(4). \(\square \)

Proof of Theorem 1.6

The proof follows from Propositions 4.5 and 4.6. \(\square \)