1 Introduction and Preliminaries

In this paper, G is a finite group and p is a prime number. Let \(\mathrm{Irr}(G)\) denote the set of irreducible (complex) characters of G, \(\mathrm{Fit}(G)\) be the Fitting subgroup of G, and let Z(G) be the center of G. The solvable radical of G is the largest solvable normal subgroup of G. For a normal subgroup N of G and \(\theta \in \mathrm{Irr}(N)\), let \(I_G(\theta )\) denote the inertia group of \(\theta \) in G and let \(\mathrm{Irr}(\theta ^G)\) be the set of the irreducible constituents of the induced character \(\theta ^G\). Also, if n is a positive integer, we use \(\pi (n)\) to show the set of the prime divisors of n and by \(\pi (G)\), we mean the set of the prime divisors of |G|. For an irreducible character \(\chi \) of G, the number \(\chi ^c(1)=\frac{[G:\mathrm{ker}\chi ]}{\chi (1)}\) is called the co-degree of \(\chi \) (see [20]). Set \(\mathrm{Codeg}(G)=\{\chi ^c(1):\chi \in \mathrm{Irr}(G)\}\). In [1,2,3, 5, 6], some properties of the co-degrees of irreducible characters of finite groups have been studied.

For a set of (positive) integers \(\mathcal {A}\), the prime graph associated with \(\mathcal {A}\) is defined as a simple undirected graph whose vertices are prime divisors of the numbers in \(\mathcal {A}\) and two distinct vertices p and q are joined by an edge if and only if pq divides some number in \(\mathcal {A}\). If \(\mathcal {A}\) is the set of the irreducible character degrees, conjugacy class sizes or order elements, then many interesting results about the prime graph associated with \(\mathcal {A}\) have been obtained, for instance see [8, 9, 16].

In this paper, by the co-degree graph \(\Delta (G)\), we mean the prime graph associated with \(\mathrm{Codeg}(G)\). By [20], the vertex set of \(\Delta (G)\) is \(\pi (G)\) and if \(\Delta (G)\) is disconnected, then G is a Frobenius or a 2-Frobenius group. We note that the vertex \(p \in \pi (G)\) is an incomplete vertex in \(\Delta (G)\) when p is not adjacent to at least one other vertex; otherwise, p is called a complete vertex. By [4, Theorem A], if G has an incomplete vertex, then \(\mathrm{Fit}(G)\ne \{1\}\). As mentioned in [4], there exists a non-solvable group G of order \(2^3.3.5^2.11^2\) such that every element of \(\mathrm{Codeg}(G)\) is either \(5.11^2\) or it is co-prime to 11. This example shows the existence of non-solvable groups such that their co-degrees graphs have no complete vertices. In Theorems A and B, we are going to study such groups. Note that in GAP library of primitive groups, the group G stated in the example given above can be accessed via G=PrimitiveGroup(11\(\hat{}\) 2,57).

Theorem A

Let G be non-solvable and N be the solvable radical of G. If 2 is an incomplete vertex in \(\Delta (G)\), then G/N has the unique minimal normal subgroup S/N and \(S/N \cong \mathrm{Alt}_7\) or \(\mathrm{PSL}_2(q)\), where q is an odd prime power. In particular, a Sylow 2-subgroup of G is a generalized quaternion group and \(2 \mid |N|\).

Theorem B

Let G be non-solvable and N be the solvable radical of G. If every vertex of \(\Delta (G)\) is incomplete in \(\Delta (G)\), then \(G/N \cong \mathrm{PSL}_2(p)\) or \(\mathrm{PGL}_2(p)\), where p is an odd prime. In particular, G has the normal subgroups L and M such that \(L \le M \le N\), \(N/M \le Z(G/M)\), M is the Hall subgroup of G, \(\Delta (L)\) is complete and G/N has the unique minimal normal subgroup S/N such that \([G:S] \mid 2\), \(S/M \cong \mathrm{SL}_2(p)\) and \(S/L \cong M/L \times S/M\).

Note that in Theorem B, \(\pi (M)=\pi (N)-\{2\}\) and \([N:M]=2\).

By [14, 21], if G has an element of order m, then G admits an irreducible character whose co-degree is divisible by \(\prod _{r \in \pi (m)}r\). Recently, G. Qian in [22] has proven that if G is solvable, then for every \(x \in G\), G admits an irreducible character \(\chi \) such that o(x), the order of x, divides \(\chi ^c(1)\). Also, he has conjectured that the same result holds true for every finite group. In this paper, we prove that:

Theorem C

If every vertex of \(\Delta (G)\) is incomplete, then for every \(x \in G\), G admits an irreducible character whose co-degree is divisible by o(x).

We note that \(\Delta (G)\) is m-regular, if every vertex of G is adjacent to exactly m other vertices of \(\Delta (G)\). For a positive integer n, let \(C_n\) denote the cyclic group of order n, \(G_1\) be a Frobenius group \(C_{91} \rtimes C_6\) and \(G_2\) be a Frobenius group whose kernel is a direct product of an elementary abelian group of order \(11^2\), an elementary abelian group of order \(41^2\) and an elementary abelian group of order \(59^2\), and its complement is isomorphic to \(SL_2(5)\). Also, assume that \(G_3\) is a Frobenius group whose kernel is a direct product of an elementary abelian group of order \(41^2\), an elementary abelian group of order \(211^2\), an elementary abelian group of order \(491^2\) and an elementary abelian group of order \(631^2\), and its complement is isomorphic to \(SL_2(5) \times C_7\). Then, \(\Delta (C_5)\) and \(\Delta (D_{10})\) are 0-regular, \(\Delta (C_{15})\) and \(\Delta (G_1)\) are 1-regular, \(\Delta (C_{30})\) and \(\Delta (G_2)\) are 2-regular and also, \(\Delta (G_3)\) is 3-regular. These examples motivate us to study some finite groups with m-regular co-degree graphs in the following theorem.

Theorem D

Let \(\Delta (G)\) be m-regular.

  1. (i)

    If \(m=0\), then G is an r-group, for some prime r, or \(|\pi (G)|=2\) and G is either a Frobenius or a 2-Frobenius group. In particular, G is solvable.

  2. (ii)

    If \(m=1\), then G is solvable and \(|\pi (G)| \in \{2,4\}\). If \(|\pi (G)|=4\), then G is a Frobenius or a 2-Frobenius group.

  3. (iii)

    If \(m=2\ne |\pi (G)|-1\), then \(|\pi (G)| \le 10\). If \(|\pi (G)| \ge 6\), then G is a Frobenius or a 2-Frobenius group. Also, if G is non-solvable, then \(|\pi (G)|=6\) and G is a Frobenius group whose Frobenius kernel is abelian and its Frobenius complement has a normal subgroup of index at most 2 which is isomorphic to \(\mathrm{SL}_2(5)\).

  4. (iv)

    If \(m=3 \ne |\pi (G)|-1\) and G is non-solvable, then G has the normal subgroups K and S such \([G:S] \mid 2\) and either \(|\pi (G)|=8\) and G is a Frobenius group with Frobenius kernel K such that \(S/K \cong L \times \mathrm{SL}_2(5)\), where L is an r-group, for some prime \(r \not \in \{2,3,5\}\) or \(|\pi (G)|=6\) and \(S/K \cong \mathrm{SL}_2(t)\), where \(t\in \{5,7,17\}\).

We note that for the proof of Theorem D-(i), our argument is similar to the argument given in [24, Theorem 1.1]. However, for convenience, we have brought the proof of Theorem D-(i) here.

Let \(\sigma ^c(G)=\mathrm{max}\{|\pi (\chi ^c(1))|: \chi \in \mathrm{Irr}(G)\}\) and \(\sigma ^e(G)=\mathrm{max}\{|\pi (o(g))|: g \in G\}\). Using [15, P. 463], there exists a function C defined on the positive integers such that for every finite solvable group G, \(|\pi (G)| \le C(\sigma ^e(G))\sigma ^e(G)\). It follows from [15, P. 464] that there exists an upper bound K for C. By [17, Corollary 1.4], \(|\pi (G)|\le (K+3/2)\sigma ^c(G)\). A. Moreto in [18, Theorem B] proved that if G is solvable such that every prime p divides at most k members of \(\mathrm{Codeg}(G)\), then \(|\pi (G)| \le 24k \mathrm{log}_2k+360k\). Here, considering Theorem D leads us to prove Theorem E and propose the question after Theorem E:

Theorem E

Let G be non-solvable, and every vertex in \(\Delta (G)\) be adjacent to at most m other vertices. Then, \(m \ge 2\) and \(|\pi (G)| \le 2(m+1)\). In particular, if \(|\pi (G)|=2(m+1)\), then G is a Frobenius group and has the normal subgroups L and S such that \(L \le S\), \(|\pi (L)|=|\pi (G/L)|=m+1\), \([G:S] \mid 2\) and \(S/L \cong U\times \mathrm{SL}_2(5)\), where \(\pi (U) \cap \{2,3,5\}=\emptyset \) and \(|\pi (U)|=m-2\).

Question

Let G be solvable and every vertex in \(\Delta (G)\) be adjacent to at most m other vertices. Is there an upper bound for \(|\pi (G)|\) in terms of a function respect to m (independent of G)?

2 Proof of the Main Results

In this section, we first present a lemma that will be used frequently in this paper.

Lemma 2.1

[20, Lemma 2.1] Let N be a normal subgroup of G. Then, \(\mathrm{Codeg}(G/N) \subseteq \mathrm{Codeg}(G)\). Also, if \(\psi \in \mathrm{Irr}(N)\), then for every \(\chi \in \mathrm{Irr}(\psi ^G)\), \( \psi ^c(1) \mid \chi ^c(1)\).

Corollary 2.2

If N is a normal subgroup of G, then \(\Delta (N)\) and \(\Delta (G/N)\) are sub-graphs of \(\Delta (G)\).

Proof

The proof follows immediately from Lemma 2.1. \(\square \)

Lemma 2.3

Let \(p,q \in \pi (G)\).

  1. (i)

    [20, Theorem A] There exists a \(\chi \in \mathrm{Irr}(G)\) such that \(p \mid \chi ^c(1)\). In particular, the vertex set of \(\Delta (G)\) is \(\pi (G)\).

  2. (ii)

    [20, Theorem E] If p and q are not adjacent in \(\Delta (G)\), then G has no elements of order pq.

  3. (iii)

    [4, Theorem A] If p and q are not adjacent in \(\Delta (G)\), then \(\mathrm{Fit}(G) \ne \{1\}\). In particular, if \(\mathrm{Fit}(G)= \{1\}\), then \(\Delta (G)\) is complete.

  4. (iv)

    [20, Theorem C] If \({\Delta }(G)\) is disconnected, then \(\Delta (G)\) has exactly two connected components and G is either a Frobenius group or a 2-Frobenius group.

  5. (v)

    [20, Theorem E] The diameter of \({\Delta }(G)\) is at most 3, i.e., for every \(\Delta \subseteq \pi (G)\) with \(|\Delta | \ge 3\), there exist \(p,q \in \Delta \) such that p and q are adjacent in \(\Delta (G)\).

2.1 Proof of Theorem A

For \(r \in \pi (G)\), let \(\pi _r\) denote the set of vertices of \(\Delta (G)\) that are not adjacent to r. In order to prove Theorem A, we first prove the following lemmas:

Lemma 2.4

Let L be a solvable normal subgroup of G. If \(p,q \in \pi (L)\) are distinct, then there exists a normal subgroup N of G such that \(N \le L\), \(r \mid |N|\) and \(s \not \mid |N|\), where \(\{p,q\}=\{r,s\}\).

Proof

We complete the proof by induction on |G|. Let M be a minimal normal subgroup of G such that \(M \le L\). Since M is an elementary abelian t-group, for some prime t and \(p,q \in \pi (L)\), \(M \ne L\). If \(t \in \{p,q\}\), then the lemma follows. Otherwise, \(p,q \in \pi (L/M)\). By induction, there exists a normal subgroup N/M of G/M such that \(N/M \le L/M\), \(r \mid |N/M|\) and \(s \not \mid |N/M|\), where \(\{p,q\}=\{r,s\}\). Therefore, \(N \unlhd G\), \(N \le L\), \(r \mid |N|\) and \(s \not \mid |N|\), as wanted. \(\square \)

Lemma 2.5

Let N be the solvable radical of G and S/N be a minimal normal subgroup of G/N. If \(p\in \pi (S/N)\) such that \( \pi _p\ne \emptyset \), then a Sylow p-subgroup of G is cyclic or \(p=2\) and a Sylow 2-subgroup of G is a generalized quaternion group.

Proof

Let \(q \in \pi _p\). Since \(\mathrm{Fit}(G/N)=\{N\}\), Lemma 2.3(iii) shows that \(\Delta (G/N)\) is complete. So, Corollary 2.2 forces \(q \in \pi (N)-\pi (G/N)\). First, let \(p \in \pi (N)\). By Lemma 2.4, there exists a normal subgroup K of G such that \(K \le N\), \(r \mid |K|\) and \(t \not \mid |K|\), where \(\{r,t\}=\{p,q\}\). Let \(R \in \mathrm{Syl}_r(K)\). By the Frattini argument, \(G/K \cong N_G(R)/N_K(R)\). As \(t \not \mid |K|\), there exists \(T \in \mathrm{Syl}_t(G)\) such that \(T \le N_G(R)\). By Lemma 2.3(ii), G has no elements of order rt. Hence, \(\langle R,T\rangle \) is a Frobenius group. Therefore, either T is cyclic or \(t=2\) and T is a generalized quaternion group. So, if \(t=p\), then the lemma follows. Otherwise, \(t=q \ne 2\), because \(2 \not \in \pi _p\) and \(q \in \pi _p\). Thus, T is cyclic. Also, \(q \in \pi (N)-\pi (G/N)\). Therefore, \(T \in \mathrm{Syl}_t(N)\). By the Frattini argument, \(G/N \cong N_G(T)/N_N(T)\). Thus, \(N_G(T)\) has a chief factor which is isomorphic to S/N. Obviously, S/N is a direct product of some isomorphic non-abelian simple groups. On the other hand, \(N_G(T)/C_G(T) \lesssim \mathrm{Aut}(T)\) and \(\mathrm{Aut}(T)\) is abelian. This forces \(C_G(T)\) to have a chief factor isomorphic to S/N. Thus, \(p \in \pi (S/N) \subseteq \pi (C_G(T))\). This implies that G has an element of order pq. This is a contradiction with Lemma 2.3(ii). Next, assume that \(p \not \in \pi (N)\) and \(Q \in \mathrm{Syl}_q(N)\). By the Frattini argument, \(G/N \cong N_G(Q)/N_N(Q)\). So, there exists a \(P \in \mathrm{Syl}_p(G)\) such that \(P \le N_G(Q)\). By Lemma 2.3(ii), G has no elements of order pq. Hence, \(\langle P,Q\rangle \) is a Frobenius group. Therefore, either P is cyclic or \(p=2\) and P is a generalized quaternion group. \(\square \)

Proof of Theorem A

Since \(\pi _2 \ne \emptyset \), Lemma 2.5 forces a Sylow 2-subgroup of G to be cyclic or a generalized quaternion group. However, G is non-solvable. So, a Sylow 2-subgroup of G is a generalized quaternion group. It follows from Brauer–Suzuki theorem [7] that a Sylow 2-subgroup of G/N is a dihedral group. Hence, \(2 \mid |N|\) and G/N has the unique minimal normal subgroup S/N. By [10] and [11], \(S/N \cong \mathrm{Alt}_7\) or \(\mathrm{PSL}_2(q)\), where q is an odd prime power. Now, the proof is complete. \(\square \)

2.2 Proof of Theorem B

Here, we first prove the following interesting and useful lemmas.

Lemma 2.6

Let N be the solvable radical of G and S/N be a minimal normal subgroup of G/N. If a Sylow p-subgroup of G is cyclic, for some \(p \in \pi (N) \cap \pi (S/N)\), then S/N is a simple group and p divides the order of the Schur multiplier of S/N.

Proof

Since \(p \mid |S/N|\) and a Sylow p-subgroup of G is cyclic, the Sylow p-subgroups of G/N are cyclic. However, S/N is a minimal normal subgroup G/N. Hence, S/N is a direct product of some isomorphic non-abelian simple groups. Nevertheless, S/N is a simple group. Let \(K_1\) be the largest normal subgroup of G such that \(K_1 \le N\) and \(p \mid |N/K_1|\). Let \(K_2/K_1\) be a minimal normal subgroup of \(G/K_1\) such that \(K_2 \le N\). By our assumption, \(p \not \mid |N/K_2|\). Hence, \(K_2/K_1\) is an elementary abelian p-group. Since a Sylow p-subgroup of G is cyclic, \(K_2/K_1\) is cyclic. Consequently, \(|K_2/K_1|=p\). We claim that \(C_{N/K_1}(K_2/K_1)=K_2/K_1\). If not, then since \(p \not \mid |N/K_2|\), \(N/K_1\) has a \(p'\)-subgroup \(K_3/K_1\) such that \(C_{N/K_1}(K_2/K_1)=(K_2/K_1)(K_3/K_1)\), by Schur–Zassenhaus theorem. However, \(K_2/K_1 \le Z(C_{N/K_1}(K_2/K_1))\). Thus, \(C_{N/K_1}(K_2/K_1)=(K_2/K_1)\times (K_3/K_1)\). So, \(K_3/K_1\) is normal in \(G/K_1\), \(K_1 < K_3 \le N\) and \(p \mid |N/K_3|\). This is a contradiction with our assumption on \(K_1\). Therefore,

$$\begin{aligned} C_{N/K_1}(K_2/K_1)=K_2/K_1. \end{aligned}$$
(1)

On the other hand, \((S/K_1)/C_{S/K_1}(K_2/K_1) \lesssim \mathrm{Aut}(K_2/K_1)\) and \( \mathrm{Aut}(K_2/K_1)\) is a cyclic group. Hence, \(C_{S/K_1}(K_2/K_1)\) has a chief factor isomorphic to S/N. By (1), \(\frac{C_{S/K_1}(K_2/K_1)}{K_2/K_1} = \frac{C_{S/K_1}(K_2/K_1)}{C_{N/K_1}(K_2/K_1)} \cong \frac{C_{S/K_1}(K_2/K_1)N/K_1}{N/K_1} \unlhd \frac{S/K_1}{N/K_1} \cong S/N\). Since S/N is simple, \(\frac{C_{S/K_1}(K_2/K_1)}{K_2/K_1} \cong S/N\). Also, \(K_2/K_1 \le Z(C_{S/K_1}(K_2/K_1))\) and \(|K_2/K_1|=p\). So, \((C_{S/K_1}(K_2/K_1))' \cap K_2/K_1=\{K_1\}\) or \(K_2/K_1\). Therefore, either \((C_{S/K_1}(K_2/K_1))'\cong S/N\) and \(C_{S/K_1}(K_2/K_1)=K_2/K_1 \times (C_{S/K_1}(K_2/K_1))'\) or \((C_{S/K_1}(K_2/K_1))'=C_{S/K_1}(K_2/K_1)\). The former case is impossible, because \(p \mid |S/N|,|K_2/K_1|\) and a Sylow p-subgroup of G is cyclic. In the latter case, \(C_{S/K_1}(K_2/K_1)\) is the Schur cover of S/N. Hence, \(|K_2/K_1|=p\) divides the order of the Schur multiplier of S/N, as desired. \(\square \)

Lemma 2.7

Let N be the solvable radical of G and S/N be a minimal normal subgroup of G/N. If a Sylow 2-subgroup of G is a generalized quaternion group, then S/N is a simple group and G has a normal subgroup M such that \(M \le N\), \(2 \not \mid |M|\) and S/M is a Schur cover of S/N.

Proof

Since a Sylow 2-subgroup of G is a generalized quaternion group, Brauer–Suzuki theorem [7] shows that a Sylow 2-subgroup of G/N is a dihedral group. So, \(2 \mid |N|\) and S/N is simple. Suppose that \(N_1\) is the largest normal subgroup of G such that \(N_1 \le N\) and \(2 \not \mid |N_1|\). Then, \(G/N_1\) has the largest normal subgroup \(L/N_1\) which is a 2-group and \(L \le N\). Since G/N is non-solvable, \(4 \mid |G/N|\). Hence, \(4 \mid [G/N_1:L/N_1]\). Also, a Sylow 2-subgroup of \(G/N_1\) is a generalized quaternion group. So, \(L/N_1\) is cyclic. Thus, \(\mathrm{Aut}(L/N_1)\) is an abelian group of order \(|L/N_1|/2\). On the other hand, \(\frac{N/N_1}{C_{N/N_1}(L/N_1)} \lesssim \mathrm{Aut}(L/N_1)\). We claim that \(L=N\). If not, G/L has a normal minimal subgroup K/L such that \(K \le N\). Obviously, \(2 \not \mid |K/L|\). Therefore, \(K/N_1 \le C_{N/N_1}(L/N_1)\). By Schur–Zassenhaus theorem, there exists a \(2'\)-subgroup \(K_1/N_1\) of \(K/N_1\) such that \((K/N_1)=(L/N_1)(K_1/N_1)\). Since \(L/N_1 \le Z(C_{N/N_1}(L/N_1))\) and \(K/N_1 \le C_{N/N_1}(L/N_1)\), \((K/N_1)=(L/N_1) \times (K_1/N_1)\). Therefore, \(K_1/N_1 \) is normal in \(G/N_1\), \(N_1 <K_1 \le N\) and \(2 \not \mid |K_1|\). This is a contradiction with our assumption on \(N_1\). This forces \(L=N\). However, \(\frac{S/N_1}{C_{S/N_1}(N/N_1)} \lesssim \mathrm{Aut}(N/N_1)\) and \(\mathrm{Aut}(N/N_1)\) is abelian. Thus, \(C_{S/N_1}(N/N_1) \) is non-solvable. Also, \(N/N_1 < C_{S/N_1}(N/N_1)\) and \(\frac{C_{S/N_1}(N/N_1)}{N/N_1} \unlhd \frac{S/N_1}{N/N_1} \cong S/N\). Therefore, \(C_{S/N_1}(N/N_1)=S/N_1\). Set \(N_2/N_1=(N/N_1) \cap (S/N_1)'\). Then, \(S/N_2 =(S'N_2/N_2) \times (N/N_2)\). Consequently,

$$\begin{aligned} (P\cap S')N_2/N_2 \times N/N_2=PN_2/N_2, \end{aligned}$$
(2)

where \(P \in \mathrm{Syl}_2(S)\). Since \(P \cap S' \) is not cyclic, we get that P is a generalized quaternion group. It follows from (2) that \(N/N_2=\{N_2\}\). So, \(N/N_1 \le (S/N_1)'\). This forces \((S/N_1)'=S/N_1\). Note that \(N/N_1 \le Z(C_{S/N_1}(N/N_1))=Z(S/N_1)\), considering \(C_{S/N_1}(N/N_1)=S/N_1\). Therefore, \(S/N_1\) is a Schur cover of S/N, as desired. \(\square \)

Proof of Theorem B

Since 2 is incomplete in \(\Delta (G)\), Theorem A forces G/N to have the unique minimal normal subgroup S/N and \(S/N \cong \mathrm{Alt}_7\) or \(\mathrm{PSL}_2(q)\), where q is a power of an odd prime p. Let \(t \in \pi (S/N)-\{2\}\). By Lemma 2.5, a Sylow t-subgroup of G is cyclic. Hence, a Sylow t-subgroup of S/N is cyclic. This shows that the Sylow subgroups of S/N of odd orders are cyclic. However, a Sylow 3-subgroup of \(\mathrm{Alt}_7\) is not cyclic and a Sylow p-subgroup of \(\mathrm{PSL}_2(q)\) is cyclic if and only if \(q=p\). Therefore, \(S/N \cong \mathrm{PSL}_2(p)\). Since S/N is the unique minimal normal subgroup of G/N, \(C_{G/N}(S/N)=\{N\}\). Consequently, \(G/N \lesssim \mathrm{Aut}(S/N)=\mathrm{Aut}(\mathrm{PSL}_2(p)) \cong \mathrm{PGL}_2(p)\). So, \(G/N \cong \mathrm{PSL}_2(p)\) or \(\mathrm{PGL}_2(p)\). Therefore, \([G:S]=[G/N:S/N] \) divides 2.

By Lemma 2.5, a Sylow 2-subgroup of G is a generalized quaternion group and for every \(t \in \pi (G/N)-\{2\}\), a Sylow t-subgroup of G is cyclic. Also, the order of the Schur multiplier of \(S/N \cong \mathrm{PSL}_2(p)\) is 2. Thus, Lemmas 2.6 and 2.7 guarantee the existence of a normal Hall subgroup M of G such that \(S/M \cong \mathrm{SL}_2(p)\) and \(N/M \le Z(S/M)\) is of order 2. In particular,

$$\begin{aligned} \pi (M)=\pi (G) -\pi (G/N),~ N/M \le Z(G/M)~\mathrm{and}~[N:M]=2. \end{aligned}$$
(3)

Now, suppose that K is the smallest non-solvable normal subgroup of G. Obviously, \(K \le S\) and KM is normal in S. So, \(K/(K\cap M) \cong KM/M\) is normal in \(S/M \cong \mathrm{SL}_2(p)\). Set \(L =K \cap M\). As K is non-solvable, \(KM=S\) and \(K/L \cong \mathrm{SL}_2(p)\). Thus, \(S/L=KM/L =K/L \times M/L \cong \mathrm{SL}_2(p) \times M/L\). We claim that \(\Delta (L)\) is complete. Working towards a contradiction, suppose that there are \(r,s \in \pi (L)\) which are not adjacent in \(\Delta (L)\). Then, \(r,s \ne 2\). Without loss of generality, we can apply the same argument given in the proof of Lemma 2.5 to conclude that a Sylow r-subgroup of G is cyclic. Let T be the largest normal subgroup of G such that \(T \le L\) and \(r \mid |L/T|\). Assume that \(L_1/T\) is a minimal normal subgroup of G/T such that \(L_1 \le L\). By our assumption on T, \(L_1/T\) is an elementary abelian r-group and \(r \not \mid |L/L_1|\). Since a Sylow r-subgroup of G is cyclic, \(|L_1/T|=r\). Also, \(L_1/T\) is normal in K/T. Hence, \(\frac{K/T}{C_{K/T}(L_1/T)} \lesssim \mathrm{Aut}(L_1/T)\) and \(\mathrm{Aut}(L_1/T)\) is a cyclic group. Therefore, \(C_{K/T}(L_1/T)\) is non-solvable. So, our assumption on K forces

$$\begin{aligned} C_{K/T}(L_1/T)=K/T. \end{aligned}$$
(4)

On the other hand, the assumption on T forces \(r \not \mid |L/L_1|\). By (3), \(r \not \mid |K/L|\). Therefore, \(r \not \mid |K /L_1|\). It follows from Schur–Zassenhaus theorem that there exists a subgroup \(K_1/T\) of K/T such that \(r\not \mid |K_1/T|\) and \(K/T=(K_1/T)(L_1/T)\). By (4), \(L_1/T \le Z(K/T)\). This shows that \(K/T=(K_1/T) \times (L_1/T)\). So, \(K' \le K_1\) is a non-solvable normal subgroup of G whose order is strictly less than |K|. This is a contradiction with our assumption on K. This guarantees that \(\Delta (L)\) is complete, as desired. \(\square \)

2.3 Proof of Theorem C

In this sub-section, let p be an odd prime. In the following, we first bring some lemmas and a proposition that will be used in the proof of Theorem C.

Lemma 2.8

[22, Proposition 2.2] Let \(Z \unlhd G\), \(\lambda \in \mathrm{Irr}(Z)\) and G/Z be solvable. If \(x \in I_G(\lambda )\), then there exists a \(\chi \in \mathrm{Irr}(\lambda ^G)\) such that \(o(xZ) \chi (1) \mid [G:Z] \lambda (1)\), where o(xZ) denotes the order of xZ in G/Z.

Lemma 2.9

Let S be a normal subgroup of G such that \(|Z(G)|=2\), \([G:S]=l \mid 2\), \(S \cong \mathrm{SL}_2(p)\) and \(G/Z(G) \cong \mathrm{PSL}_2(p)\) or \(\mathrm{PGL}_2(p)\). Then:

  1. (i)

    for every \(x \in G\), \(o(x) \in \{p,lp\} \) or \(o(x) \mid l(p\pm 1)\);

  2. (ii)

    G admits the faithful irreducible characters whose degrees are \(p+1\) and \(p-1\).

Proof

The proof follows by checking the character table of \(\mathrm{SL}_2(p)\) (see [23]). \(\square \)

Now, we fix the following hypothesis:

Hypothesis \(*\). Let N, U and S be normal subgroups of G such that \(N \le U \le S\), \(U/N \le Z(G/N)\) is of order 2, \(\mathrm{gcd}(|N|,|G/N|)=1\), [G : S] divides 2, \(S/N \cong \mathrm{SL}_2(p)\) and \(G/U \cong \mathrm{PSL}_2(p)\) or \(\mathrm{PGL}_2(p)\), where p is an odd prime.

Proposition 2.10

Let N, U and S be the normal subgroups of G satisfying Hypothesis \(*\) and let \(2 \mid |C_G(N)|\). Then:

  1. (i)

    G has the unique involution z. In particular, \(z \in Z(G)\).

  2. (ii)

    If \(\psi _1,\psi _2 \in \mathrm{Irr}(G)\) such that \(2 \mid |\mathrm{ker}\psi _1|\) and \(2 \not \mid |\mathrm{ker}\psi _2|\), then \(2 \not \mid |\mathrm{ker}\psi _1\psi _2|\).

  3. (iii)

    If \(H \le G\) and \( \psi \in \mathrm{Irr}(H) \) such that \(2 \not \mid |\mathrm{ker}\psi |\), then \(2 \not \mid |\mathrm{ker} \psi ^G|\).

Proof

Since |N| is odd and \(2 \mid |C_G(N)|\), \(2 \mid |C_G(N)N/N|\). However, \(C_G(N)N/N\) is a normal subgroup of G/N. Thus, \(C_G(N)N/N\in \{G/N,S/N\}\) or \(C_G(N)N/N\) is a cyclic group of order 2. This shows that \(C_G(N)N/N\) has the unique central involution. On the other hand, \(\mathrm{gcd}(|N|,|G/N|)=1\) and \(C_G(N)N/N \cong C_G(N)/Z(N)\). Hence, \(\mathrm{gcd}(|Z(N)|,|C_G(N)/Z(N)|)=1\). By Schur–Zassenhaus theorem, \(C_G(N)\) has a subgroup H such that \(H \cong C_G(N)/Z(N) \cong C_G(N)N/N\), \(H \cap Z(N) =\{1\}\) and \(C_G(N)=H Z(N)\). So, \(C_G(N)=H \times Z(N)\). We note that \(H \cap N=\{1\}\), because \(\mathrm{gcd}(|N|,|H|)=1\). As \(H \cong HN/N=HZ(N)N/N=C_G(N)N/N \) and \(C_G(N)N/N\) has the unique central involution, we get that H has the unique central involution. Thus, \(C_G(N)\) has the unique central involution z. Since \(C_G(N)\) is a normal subgroup of G, \(\langle z\rangle \) is normal in G. Therefore, \(z \in Z(G)\). Also, a Sylow 2-subgroup of G is a generalized quaternion group. Hence, z is the unique involution of G, as desired in (i). Now, let \(\psi _1,\psi _2 \in \mathrm{Irr}(G)\) such that \(2 \mid |\mathrm{ker}\psi _1|\) and \(2 \not \mid |\mathrm{ker}\psi _2|\). Thus, \(z \in \mathrm{ker}\psi _1\) and \(z \not \in \mathrm{ker}\psi _2\). So, \(z \not \in \mathrm{ker}\psi _1\psi _2\). Since z is the unique involution of G, we get that \(2 \not \mid |\mathrm{ker}\psi _1\psi _2|\), as wanted in (ii). Finally, let \(H \le G\) and \( \psi \in \mathrm{Irr}(H) \) such that \(2 \not \mid |\mathrm{ker}\psi |\). Since \(z \in Z(G)\), \(\psi ^G(z)=[G:H] \dot{\psi }(z)\). However, \(z \not \in \mathrm{ker}\psi \). Hence, \(\dot{\psi }(z) \ne \psi (1)\), so \(\psi ^G(1) \ne \psi ^G(z)\). Consequently, \(z \not \in \mathrm{ker}\psi ^G\). As z is the unique involution of G, we have \(2 \not \mid |\mathrm{ker}\psi ^G|\). Now, the proof is complete. \(\square \)

Proposition 2.11

Let N, U and S be normal subgroups of G satisfying Hypothesis \(*\) and let M be a normal subgroup of G such that \(M \le N\). If \(\lambda \in \mathrm{Irr}(M)\) and \(x \in I_G(\lambda )\), then there exists a \(\chi \in \mathrm{Irr}(\lambda ^G)\) such that \(o(xM) \chi (1) \mid [G:M] \lambda (1)\) and if \(2 \mid |C_G(N)|\), then \(2 \not \mid |\mathrm{ker}\chi |\).

Proof

First let \(I_G(\lambda ) \ne G\). If \(I_G(\lambda )\) is non-solvable, then since \(I_G(\lambda )N/N \le G/N\), we get that \(I_S(\lambda )/(N \cap I_S(\lambda )) \cong \mathrm{SL}_2(r)\) for some prime r. So, the inductive hypothesis and Lemma 2.8 guarantee the existence of \(\psi \in \mathrm{Irr}(I_G(\lambda ))\) such that

$$\begin{aligned} o(xM) \psi (1) \mid [I_G(\lambda ):M] \lambda (1). \end{aligned}$$
(5)

Also, if \(I_G(\lambda )\) is non-solvable and \(2 \mid |C_{I_G(\lambda )}(I_N(\lambda ))|\), then \(2 \not \mid |\mathrm{ker} \psi |\). Recall that if \(2 \mid |C_G(N)|\), then G has a central involution z and hence, \(z \in C_{I_G(\lambda )}(I_N(\lambda ))\). Thus, if \(I_G(\lambda )\) is non-solvable and \(2 \mid |C_G(N)|\), then \(2 \not \mid |\mathrm{ker}\psi |\). Next, let \(I_G(\lambda )\) is solvable and \(2 \mid |C_G(N)|\). By Proposition 2.10(i), G has the unique involution z which is central in G. Thus, \(M \times \langle z\rangle \trianglelefteq G\). Let \(\mathrm{Irr}(\langle z\rangle )=\{1_{\langle z\rangle },\lambda _2\}\). Since \((\lambda \lambda _2)(1)=\lambda (1)\), \(I_G(\lambda \lambda _2)=I_G(\lambda )\) and \(\mathrm{Irr}((\lambda \lambda _2)^{I_G(\lambda )})\subseteq \mathrm{Irr}(\lambda ^{I_G(\lambda )})\), we can replace \(\lambda \) by \(\lambda \lambda _2\) in the above argument. Since \(\psi \in \mathrm{Irr}((\lambda \lambda _2)^{I_G(\lambda )})\), \(z \not \in \mathrm{ker}\psi \). Also, z is the unique involution of G. Thus, \(2 \not \mid |\mathrm{ker}\psi |\). Clifford correspondence [13, Theorem 6.11] forces \(\chi =\psi ^G \in \mathrm{Irr}(\lambda ^G)\) and \(\chi (1)=[G:I_G(\lambda )]\psi (1)\). It follows from (5) that

$$\begin{aligned} o(xM)\chi (1)= & {} o(xM)[G:I_G(\lambda )] \psi (1) \mid ([I_G(\lambda ):M] \lambda (1))[G:I_G(\lambda )] \!=\! [G:M] \lambda (1). \end{aligned}$$

Further, Proposition 2.10(iii) shows that if \(2 \mid |C_G(N)|\), then \(2\not \mid |\mathrm{ker}\chi |\), as wanted. Now, assume that \(I_G(\lambda )=G\). If \(N=M\), then since \(\mathrm{gcd}(|G/N|,|N|)=1\), we get from [13, Exercise 6.18] that there exists a \(\chi \in \mathrm{Irr}(\lambda ^G)\) such that \(\chi _N=\lambda \). By Lemma 2.9, there exists a faithful character \(\psi _2 \in \mathrm{Irr}(G/N)\) such that \(o(xN)\psi _2(1) \mid |G/N| \). It follows from [13, Corollary 6.17] and Proposition 2.10(ii) that \(\chi \psi _2 \in \mathrm{Irr}(\lambda ^G)\) and \(2 \not \mid |\mathrm{ker}\psi \psi _2|\). Therefore, \(o(xN)(\chi \psi _2)(1)=o(xN)\psi _2(1)\lambda (1) \) divides \( [G:N]\lambda (1)\), as needed. Finally, assume that \(N \ne M\) and L/M is a chief factor of G such that \(L\le N\). Then, L/M is an elementary abelian p-group for some prime p. Using [13, Exercise 6.12], \(\lambda ^L=e\theta \), where \(e^2=[L:M]\) and \(\theta \in \mathrm{Irr}(L)\) or \(\lambda ^L=\theta _1+\cdots +\theta _t\), where \(\theta _1, \ldots ,\theta _t \in \mathrm{Irr}(L)\), \(\theta _1(1)=\cdots =\theta _t(1)=\lambda (1)\) and \(t=[L:M]\). Now, rewriting the argument given in the proof of [22, Proposition 2.2] guarantees that there exists \(\psi \in \mathrm{Irr}(\lambda ^L)\) and a divisor \(t_0\) of o(xL) such that \(x^{t_0} \in I_G(\psi )\), where either \(t_0=1\) or \(\lambda ^L=\theta _1+\cdots +\theta _t\), \(t_0\) is a power of p and \(t_0 o(xM)/o(xL) \mid t=[L:M]\). By the inductive hypothesis, there exists \(\chi \in \mathrm{Irr}(\psi ^G)\) such that

$$\begin{aligned} o(x^{t_0}L)\chi (1) \mid [G:L]\psi (1) \end{aligned}$$
(6)

and if \(2 \mid |C_G(N)|\), then \(2 \not \mid |\mathrm{ker}\chi |\). If \(t_0=1\), then since \(p\psi (1)/\lambda (1) \mid |L/M|\) and \(o(xM) \mid po(xL)\), we have \(o(xM) \chi (1) \mid [G:M] \lambda (1) \), regarding (6). Otherwise, \(\psi (1)=\lambda (1)\) and \(o(x^{t_0}L)=o(xL)/t_0\). Thus, \(o(xM)=t_0o(x^{t_0}L)o(xM)/o(xL) \mid [L:M]o(x^{t_0}L)\). It follows from (6) that \(o(xM) \chi (1) \mid [G:M] \lambda (1)\), as desired. \(\square \)

Proposition 2.12

Let N, U and S be normal subgroups of G satisfying Hypothesis \(*\). Then, for every \(x \in G\), G admits an irreducible character whose co-degree is divisible by o(x).

Proof

We complete the proof by induction on |G|. Let \(\{N_1,\ldots ,N_t\}\) be the set of all minimal normal subgroups of G which are subgroups of N. Then, for every \(1 \le i \le t\), \(N_i\) is an elementary abelian \(p_i\)-group, for some prime \(p_i\). If there exists an \(1 \le i \le t\) such that \(o(xN_i)=o(x)\), then we get from induction that there exists a \(\chi \in \mathrm{Irr}(G/N_i)\) such that \(o(x)=o(xN_i) \mid \chi ^c(1)\). So, Lemma 2.1 completes the proof. Now, suppose that for every \(1 \le i \le t\), \(o(xN_i)\ne o(x)\). Set \(x_i=x^{o(x)/p_i}\). If \(x_i =1\), then \(o(x)=p_i\) and hence, Lemma 2.3(i) completes the proof. Next, we assume that \(x_i \ne 1\), for every \(1 \le i \le t\). Then, \( x_i \in N_i\) and hence, \(\{p_1,\ldots ,p_t\}\) are pair-wise distinct. Since \(x \in C_G(x_i)\), we get from [13, Theorem 6.32] that there exists a non-principal character \(\lambda _i \in \mathrm{Irr}(N_i)\) such that \(x \in I_G(\lambda _i)\). Set \(\lambda =\lambda _1 \times \cdots \times \lambda _t \) and \(M=N_1 \times \cdots \times N_t\). Then, \(M \unlhd G\) and \(\lambda \in \mathrm{Irr}(M)\). By Proposition 2.11, there exists a \(\chi \in \mathrm{Irr}(\lambda ^G)\) such that

$$\begin{aligned} o(xM)\chi (1) \mid [G:M] \lambda (1). \end{aligned}$$
(7)

Also, if \(2 \mid |C_G(N)|\), then \(2 \not \mid |\mathrm{ker}\chi |\). If \(\mathrm{ker} \chi \ne \{1\}\), then since \(\mathrm{ker} \chi \unlhd G\), there exists a minimal normal subgroup L of G such that \(L \le \mathrm{ker} \chi \). If \(L \le N\), then \(L \in \{N_1,\ldots ,N_t\}\). So, \(L=N_i\), for some \(1 \le i \le t\). Thus, \(\chi _{N_i}=\chi (1)1_{N_i}\). This is a contradiction, because \(\chi \in \mathrm{Irr}(\lambda _i^G)\). This shows that \(L \not \le N\). However, \(L \ne L \cap N \unlhd G\) and L is a minimal normal subgroup of G. Therefore, \(L \cap N=\{1\}\). So, \(L \times N \unlhd G\) and \(L \cong LN/N \unlhd G/N\). Hence, considering the normal subgroups of G/N forces \(2 \mid |Z(L)|\). Consequently, \(2 \mid |C_G(N)|\) and \(2 \mid |\mathrm{ker} \chi |\), which is a contradiction. This shows that \(\mathrm{ker} \chi = \{1\}\). On the other hand, \((p_1 \ldots p_t)o(xM)=o(x)\) and \((p_1 \ldots p_t)[G:M]=(p_1 \ldots p_t)|G|/(|N_1|\ldots |N_t|) \mid |G| \). It follows from (7) that \(o(x) \chi (1) \mid |G| \lambda (1)=|G|\). Consequently, o(x) divides \(\chi ^c(1)\), as wanted. \(\square \)

We note that in some parts of the proof of Proposition 2.12, we follow the ideas given in the proof of [22, Theorem B].

Proof of Theorem C

If G is solvable, then [22, Theorem B] completes the proof. Now, let G be non-solvable. By Theorem C, G has the normal subgroups N, U and S satisfying Hypothesis \(*\). So, the theorem follows from Proposition 2.12. \(\square \)

2.4 Proof of Theorem D

First let \(m=0\). Then, every connected component of \(\Delta (G)\) is a single point. However, \(\Delta (G)\) has at most two connected components, by Lemma 2.3(iv). It follows that \(|\pi (G)| =1\) or \(\Delta (G)\) is disconnected and \(|\pi (G)| =2\). Thus, G is solvable. Also, if \(\Delta (G)\) is disconnected, then Lemma 2.3(iv) shows that G is a Frobenius or a 2-Frobenius group.

Now, let \(m=1\). Then, every connected component of \(\Delta (G)\) is a path of length 1. However, \(\Delta (G)\) has at most two connected components, by Lemma 2.3(iv). It follows that \(|\pi (G)| =2\) or \(\Delta (G)\) is disconnected and \(|\pi (G)| =4\). Thus, if \(\Delta (G)\) is connected, then G is solvable. If \(\Delta (G)\) is disconnected, then Lemma 2.3(iv) shows that G is a Frobenius or a 2-Frobenius group. If G is a Frobenius group with the Frobenius complement H, then \(\pi (H)\) is a connected component of \(\Delta (G)\). Thus, \(|\pi (H)|=2\). Hence, H is solvable. Since the kernel of a Frobenius group is nilpotent, G is solvable. Also, the 2-Frobenius groups are solvable. Nevertheless, \(|\pi (G)| \in \{2,4\}\) and G is solvable, as wanted.

Next, assume that \(m=2\). Thus, every connected component of \(\Delta (G)\) is a cycle. By Lemma 2.3(v), diameter of \(\Delta (G)\) is at most 3. Thus, every connected component of \(\Delta (G)\) contains at most 5 vertices. However, \(\Delta (G)\) has at most two connected components, by Lemma 2.3(iv). It follows that \(|\pi (G)| \le 5\) or \(\Delta (G)\) is disconnected and \(|\pi (G)| \le 10\). In the latter case, Lemma 2.3(iv) shows that G is either a Frobenius or a 2-Frobenius group. Note that \(m < |\pi (G)|-1\) and the co-degree graph of a simple group is complete. So, 2-regularity of \(\Delta (G)\) and [19, Theorem 18.6] shows that G is solvable or G is a Frobenius group with the abelian kernel N such that G/N has a normal subgroup of index at most 2 which is isomorphic to \(\mathrm{SL}_2(5)\).

Assume that \(m = 3\) and G is non-solvable. Since \(m < |\pi (G)|-1\), every vertex of \(\Delta (G)\) is incomplete. It follows that from Theorem B that G has the normal subgroups N and S such that \(\mathrm{gcd}(|N|,|G/N|)=1\), \([G:S] \mid 2\) and \(S/N \cong \mathrm{SL}_2(p)\). However, \(\Delta (\mathrm{SL}_2(p))\) is complete, \(|\pi (\mathrm{SL}_2(p))| \ge 3\) and by Lemma 2.2, \(\Delta (S/N) \) is a sub-graph of \(\Delta (G)\). So, \(|\pi (\mathrm{SL}_2(p))| \le 4\).

If \(|\pi (\mathrm{SL}_2(p))| =4\), then for every \(r \in \pi (N)\) and \(s \in \pi (S/N)\), rs divides none of the elements of \(\mathrm{Codeg}(G)\). It follows from Lemma 2.3(ii) that \(\mathrm{SL}_2(p)\) acts fixed point freely on N. Therefore, S is a non-solvable Frobenius group with the kernel N and its complement is isomorphic to S/N. Thus, N is abelian and it follows from [19, Theorem 18.6] that \(S/N \cong \mathrm{SL}_2(5) \times L\), where L is a group whose order is a prime power. This is a contradiction, because \(S/N \cong \mathrm{SL}_2(p)\) and \(|\pi (\mathrm{SL}_2(p))|=4\).

Next, let \(|\pi (\mathrm{SL}_2(p))| = 3\). Applying [12], \(p \in \{5,7,17\}\). By Theorem B, there exists a normal subgroup K of G such that \(\Delta (K)\) is complete and \(S/K \cong N/K \times S/N\). If \(N/K \ne \{K\}\), then since \(\Delta (S/N)\) is complete, we get that \(|\pi (N/K)|=1\), \(|\pi (K)|=4\) and S/K acts fixed point freely on K. Therefore, \(S/K \cong \mathrm{SL}_2(5) \times N/K\), by [19, Theorem 18.6]. Nevertheless, \(|\pi (G)|=8\). If \(N/K=\{K\}\), then since \(\Delta (K)\) is complete and \(\Delta (G/K)\) is 2-regular, we get that \(|\pi (K)|=3\) and for every \(t\in \pi (K)\), there exists exactly one element \(r \in \pi (G/K)\) such that t and r are adjacent in \(\Delta (G)\). Therefore, \(|\pi (G)|=6\), as wanted. \(\square \)

2.5 Proof of Theorem E

If \(\Delta (G)\) has a complete vertex, then \(m=|\pi (G)|-1\) and hence, \(|\pi (G)| =(m+1) < 2(m+1)\), as wanted. Now, suppose that every vertex of \(\Delta (G)\) is incomplete. By Theorem B, G has the normal subgroups L, M and S such that \( L\le M \le S\), \([G:S] \mid 2\), \(\Delta (L)\) is complete and \(S/L \cong \mathrm{SL}_2(p) \times M/L\), for some odd prime p. So,

$$\begin{aligned} \pi (G) = \pi (S/L) \cup \pi (L). \end{aligned}$$
(8)

Let \(r \in \pi (\mathrm{SL}_2(p))\) and \(t \in \pi (S/L)\). If \(t \in \pi (\mathrm{SL}_2(p))\), then Lemma 2.3(iii) and Corollary 2.2 force r and t to be adjacent in \(\Delta (G)\). If \(t \in \pi (M/L)\), then G/L contains an element of order tr, so does G. By Lemma 2.3(ii), r and t are adjacent in \(\Delta (G)\). So, r is adjacent to at least \(|\pi (S/L)|-1\) other vertices. Therefore, \(|\pi (S/L)|-1 \le m\). Also, \(\Delta (L)\) is complete. Hence, every vertex of \(\Delta (L)\) is adjacent to at least \(|\pi (L)|-1\) other vertices. Consequently, \(|\pi (L)|-1 \le m\). It follows from (8) that \(|\pi (G)| \le |\pi (S/L)|+|\pi (L)| \le 2(m+1)\), as desired.

Now, suppose that \(|\pi (G)| =2(m+1)\). Thus, the above argument shows that \(\pi (L) \cap \pi (G/L)=\pi (L) \cap \pi (S/L) =\emptyset \) and \(|\pi (L)|=|\pi (S/L)|= m+1\). Therefore, G has a subgroup H such that \(H \cong G/L\) and \(G=HL\). However, \(\Delta (L)\) is complete. This signifies that no element of \(\pi (L)\) is adjacent to the elements of \(\pi (S/L)=\pi (H)\). Hence, Lemma 2.3(ii) shows that H acts fixed freely on L. So, [19, Theorem 18.6] completes the proof. \(\square \)