1 Introduction

Let G be a group acting on a set Ω. Denote by G α the subgroup of G fixing the point α. G is said to be semiregular if G α =1 for each αΩ, and G is said to be regular if G is transitive and semiregular. A non-empty subset Δ of Ω is called a block for G if for each gG either Δ g=Δ or Δ gΔ=∅. Clearly, the set Ω and the singletons {α} (αΩ) are blocks for G, called the trivial blocks. Any other block is said to be non-trivial. Suppose that Δ is a non-trivial block for G. Then {Δ ggG} is the system of imprimitivity of G containing Δ. A transitive group G is primitive if G has no non-trivial blocks on Ω.

Throughout this paper, we consider undirected finite graphs without loops or multiple edges. As usual, the notation X=(V,E) denotes a graph with vertex set V and edge set E, and Aut(X) denotes its automorphism group. If two vertices u ,vV are adjacent, {u,v} denotes the edge between u and v. By X 1(v), we mean the neighborhood of a vertex v in X, consisting of vertices which are adjacent to v.

Let X=(V,E) be a graph and G≤Aut(X). Then X is said to be G-locally primitive if the vertex stabilizer G v acts primitively on X 1(v) for each vV. A graph X is said to be G-vertex-transitive or G-edge-transitive if G acts transitively on V or E, respectively. If G is replaced by Aut(X), the graph X is simply said to be vertex-transitive or edge-transitive.

An s-arc in a graph is an ordered (s+1)-tuple (v 0,v 1,…,v s−1,v s ) of vertices of the graph X such that v i−1 is adjacent to v i for 1≤is, and v i−1v i+1 for 1≤is−1. A 0-arc is a vertex and a 1-arc is also called an arc for short. A graph X is said to be (G,s)-arc-transitive if G≤Aut(X) is transitive on the set of s-arcs in X. A (G,s)-arc-transitive graph is said to be (G,s)-transitive if it is not (G,s+1)-arc-transitive. A graph X is said to be s-arc-transitive or s-transitive if the graph is (Aut(X),s)-arc-transitive or (Aut(X),s)-transitive. A graph X is G-edge-primitive if G≤Aut(X) acts primitively on the set of edges of X, and X is edge-primitive if it is Aut(X)-edge-primitive.

Weiss [9] determined all edge-primitive cubic graphs, which are the complete bipartite graph K 3,3, the Heawood graph of order 14, the Biggs–Smith cubic distance-transitive graph of order 102 and the Tutte–Coxeter graph of order 30 (also known as Tutte’s 8-cage or the Levi graph). Giudici and Li [3] systematically analyzed edge-primitive graphs via the O’Nan–Scott Theorem to determine the possible edge and vertex actions of such graphs, and determined all G-edge-primitive graphs for G an almost simple group with socle PSL(2,q), where q is a prime power and q≠2, 3. Recently, the authors [4] classified edge-primitive tetravalent graphs, which are the complete graph K 5, the co-Heawood graph of order 14 (the complement graph of the Heawood graph with respect to the complete bipartite graph K 7,7), the complete bipartite graph K 4,4, and three coset graphs defined on the almost simple groups Aut(PSL(3,3)), Aut(M12) and Aut(G 2(3)), respectively. In [6], edge-primitive 4-arc-transitive graphs are classified. In this paper, we give a classification of edge-primitive graphs of valency 5.

Theorem 1.1

Let X be an edge-primitive pentavalent graph with an edge e={u,v} and let A=Aut(X). Then X is s-transitive with s≥2, and X, s, A, A v and A e are listed in Table  1. Furthermore, such a graph X is uniquely determined by its number of vertices.

Table 1 s-transitive edge-primitive pentavalent graphs

From Theorem 1.1, we have the following corollary.

Corollary 1.2

All finite edge-primitive pentavalent graphs are 2-arc-transitive.

Remark

Let X be an edge-primitive graph with an edge e={u,v} and let A=Aut(X). Weiss classified such graphs of valency 3 in 1973. However, since then there is no much progress in this line for small valencies. In this paper, we first reduce A to an almost simple group when X has valency 5 and Theorem 1.1 follows from the classification of finite primitive groups with solvable stabilizers given in [6]. The method does not work for valency greater than 5 because A e can be non-solvable.

2 A reduction

Let X=(V,E) be a G-edge-primitive graph of valency 5 with an edge e={u,v}. Then 2|E|=5|V|, and G is a primitive permutation group on E. By [3, Lemmas 3.1 and 3.4], X is connected and G-arc-transitive. Thus, 5 | |G v |, but \(5^{2}\,{\not|}\,\,\,|G_{v}|\). In particular, X is G-locally primitive. Let N=Soc(G)=T k, the socle of G. Then T is a simple group, N is transitive on E, and hence N has at most two orbits on V. If N has two orbits on V, denote by V 1 and V 2 these orbits. In this case, X is bipartite with V 1 and V 2 as its bipartition sets.

Lemma 2.1

The socle N is a minimal normal subgroup of G and is not semiregular on V, and the graph X is N-locally primitive. If XK 5,5, T is non-abelian simple and if further k≥2, T is semiregular on V.

Proof

Let \(1\not=M\lhd G\). Suppose that M is semiregular on V. Then M v =1 and M is transitive on E, implying that M has at most two orbits on V. Thus, |V|=|M| or 2|M|. The edge-primitivity of G implies that M is transitive on E. It follows that |E| | |M| and so |E| | |V|, which is impossible because \(|E|={5|V|\over 2}\). Thus, M is not semiregular on V. Note that |X 1(v)|=5. Since MG and X is G-arc-transitive, M v is transitive on X 1(v), and hence primitive on X 1(v). Further, X is M-locally primitive. In particular, by taking M=N we see that N is not semiregular on V and X is N-locally primitive. If G has two distinct minimal normal subgroups, say N 1 and N 2, then N 1×N 2G. By taking M=N 1 or N 2, X is N 1- and N 2-locally primitive. This implies that 5 | |(N 1) v | and |(N 2) v |, forcing that 52 | |G v |, a contradiction. Thus, N is a minimal normal subgroup of G.

To prove the second part, let \(X\not=K_{5,5}\). Suppose T is abelian. Then N is abelian and hence regular on E. It follows that \(|N|=|E|=\frac{5|V|}{2}\). Recall that N is not semiregular. If N has one orbit on V then N is regular on V, a contradiction. It follows that N has two orbits on V, that is, V 1 and V 2, and for vV 1, we have \(N_{v}\not=1\). Since N is abelian, N v fixes every vertex in V 1, forcing X=K 5,5, a contradiction. Thus, N is non-abelian. To finish the proof, we further let k≥2. Suppose that T is not semiregular on V. Write N=T×L, where L=T k−1. By the minimality of N in G, L is not semiregular on V.

Assume that N is transitive on V. Since T is not semiregular on V, \(T_{w}\not=1\) for every wV, and by the minimality of N in G, \(L_{w}\not=1\). By the vertex-transitivity and locally primitivity of N, we have 5 | |T w | and 5 | |L w |. It follows that 52 | |N w |, which is impossible.

Now assume that N has two orbits on V. We may let vV 1 and uV 2. Suppose that \(5\,{\not|}\,\,\,|T_{v}|\) and \(5\,{\not|}\,\,\,|T_{u}|\). Then \(5\,{\not|}\,\,\,|T_{w}|\) and \(5\,{\not|}\,\,\,|T_{x}|\) for every wV 1 and every xV 2. Since X is N-locally primitive, T w and T x fix X 1(w) and X 1(x) pointwise, respectively. The connectivity of X implies that T u and T v fix every vertex in V. Then T u =T v =1 and hence T w =T x =1 for every wV 1 and every xV 2, contrary to the assumption that T is not semiregular on V. Thus, we may assume that 5 | |T v | (note that we cannot deduce 5 | |T v | by the locally primitivity of N when \(T_{v}\not=1\) because N has two orbits). By the minimality of N, 5 | |L u |.

Consider the orbits of L. Let {B 1,B 2,…,B m } and {C 1,C 2,…,C n } be the sets of orbits of L on V 1 and V 2, respectively. We may assume that vB 1 and uC 1. Note that B i and C i are blocks of N. Since 5 | |L u |, L u is transitive on X 1(u). Thus, X 1(u)⊆B 1, and X 1(x)⊆B 1 for every xC 1 because B 1 and C 1 are orbits of L. The edge-transitivity of N implies that if there is an edge between C i and B 1 then X 1(x)⊆B 1 for every xC i . The connectivity of X implies that m=1, that is, L is transitive on V 1. Thus, T v =T w for every vertex wV 1 because L commutes with T, which forces that X=K 5,5, a contradiction. □

Lemma 2.2

Let XK 5,5. Then G v is non-solvable and G is 2-arc-transitive.

Proof

Since G is arc-transitive, let X be (G,s)-arc-transitive for some s≥1. Note that a transitive permutation group of prime degree is either solvable or 2-transitive. To prove the lemma, we only need to show that G v is non-solvable. Suppose to the contrary that G v is solvable. Then the primitive permutation group \(G_{v}^{X_{1}(v)}\) of degree 5 is isomorphic to ℤ5, D 10 or F 20, and hence \(G_{vu}^{X_{1}(v)}\) is a 2-group (the identity subgroup is also viewed as a 2-group). Let \(G_{v}^{[1]}\) be the kernel of G v acting on X 1(v). By [6, Theorem 1.3], G v is non-solvable for s≥4. Thus, s≤3 and by [10, Theorems 4.6–4.7], \(G_{uv}^{[1]}=G_{u}^{[1]}\cap G_{v}^{[1]}=1\). Then \(G_{u}^{[1]}\times G_{v}^{[1]}= G_{u}^{[1]}G_{v}^{[1]}\lhd G_{vu}\), and \(G_{v}^{[1]}\cong G_{u}^{[1]}\cong G_{u}^{[1]}/(G_{u}^{[1]}\cap G_{v}^{[1]})\cong (G_{u}^{[1]})^{X_{1}(v)}\lhd G_{vu}^{X_{1}(v)}\), implying that \(G_{v}^{[1]}\) is a 2-group. It follows that G vu is a 2-group and hence G e is a 2-group. The maximality of G e in G implies that G e is a Sylow 2-subgroup of G. Thus, |E| is odd and so is \({1\over2}|V|\). Moreover, if N=T k with k≥2 then by Lemma 2.1, T is semiregular on V, which is impossible. Thus k=1, and G is almost simple. By [11, Theorem], if the stabilizer of an arc-transitive automorphism group of a graph with prime valency p is solvable then its order is a divisor of p(p−1)2. Thus, |G v | | 80, which forces that |G e | | 32. Since |G| is divisible by 5, by [6, Tables 14–20] we have G=PGL(2,9), M10, or PSL(2,31), which is also impossible by the Atlas [1]. □

Let XK 5,5 and s≤3. Note that \(G_{v}^{[1]}\) is a {2,3}-group and hence solvable. By Lemma 2.2, \(G_{v}^{X_{1}(v)}\) is non-solvable and so \(G_{v}^{X_{1}(v)}=\mathrm{A}_{5}\) or S5, which implies that \(G_{vu}^{X_{1}(v)}=\mathrm{A}_{4}\) or S4, respectively. If \(G_{v}^{[1]}=1\) then G v =A5 or S5. Now assume \(G_{v}^{[1]}\not=1\). Since \(G_{v}^{[1]}\lhd G_{vu}\) and \(G_{v}^{[1]}\cap G_{u}^{[1]}=1\), \(G_{v}^{[1]}\) is transitive on X 1(u)∖{v}.

We claim that G v has a normal subgroup A5 or S5 and \(G_{v}^{[1]}=\mathrm{A}_{4}\) or S4. To prove it, let \(H=\langle G_{z}^{[1]}\mid z\in X_{1}(v)\rangle\). Then HG v . Since \(G_{v}^{[1]}\cap G_{u}^{[1]}=1\), the action of \(G_{v}^{[1]}\) on X 1(u)∖{v} is non-trivial and so H has a non-trivial action on X 1(v). Since \(H^{X_{1}(v)}\unlhd G_{v}^{X_{1}(v)}=\mathrm{A}_{5}\) or S5, we have \(H^{X_{1}(v)}=\mathrm{A}_{5}\) or S5. Then \(H_{vu}^{X_{1}(v)}=\mathrm{A}_{4}\) or S4, and so H contains a non-identity element h of order 3-power such that hH vuw for some wX 1(u) with \(w\not=v\). On the other hand, it is easy to show that \([H,G_{v}^{[1]}]\leq G_{v}^{[1]}\cap G_{u}^{[1]}=1\), which implies that \(H\cap G_{v}^{[1]}\leq Z(G_{v}^{[1]})\), the center of \(G_{v}^{[1]}\). It follows that H commutes with \(G_{v}^{[1]}\) and hence h fixes X 1(u) pointwise because \(G_{v}^{[1]}\) is transitive on X 1(u)∖{v}. Thus, \(3\,|\,|G_{u}^{[1]}|\) and \(3\,|\,|G_{v}^{[1]}|\). Since \(G_{v}^{[1]}\cong G_{v}^{[1]}/G_{uv}^{[1]}\unlhd G_{vu}^{X_{1}(v)}=\mathrm{A}_{4}\) or S4, we have \(G_{v}^{[1]}=\mathrm{A}_{4}\) or S4. Furthermore, \(H\cap G_{v}^{[1]}\leq Z(G_{v}^{[1]})=1\), implying that \(G_{v}^{[1]}H=G_{v}^{[1]}\times H\) and H is faithful on X 1(v). Thus, H=A5 or S5, as claimed.

Now it is easy to see that \(|G_{v}:G_{v}^{[1]}\times H|=1\) or 2 and we have the following lemma.

Lemma 2.3

Suppose that XK 5,5 and X is connected (G,s)-transitive with s≤3. Then either

  1. (1)

    s=2, and G v =A5 or S5, or

  2. (2)

    s=3, and G v =A4×A5, S4×S5, or (A4×A5).ℤ2.

In particular, this lemma tells us that G v does not have a subnormal subgroup ℤ5.

Lemma 2.4

Suppose that XK 5,5 and X is connected (G,s)-transitive with s≤3. Then G is almost simple.

Proof

Suppose that \(1\not=M\lhd N\) is regular on E. Then X is M-edge-transitive, and hence M has at most two orbits on V. Thus, \(|M|=|E|={5|V|\over2}\) is divisible by |V| or \({1\over2}|V|\), forcing that M has exactly two orbits on V, X is bipartite, and |M v |=5. It follows that ℤ5M v N v G v , which is impossible by Lemma 2.3. Thus, N does not have a normal subgroup which is regular on E, and by O’Nan–Scott’s theorem [2, Theorem 4.1A], G is almost simple, or of product action on E.

Suppose that G is of product action on E. Then by O’Nan–Scott’s theorem, \(N_{e}=T_{e}^{k}\), \(T_{e}\not=1\), and k≥2. Since XK 5,5, by Lemma 2.1, T v =1. It follows that T e =ℤ2 and \(N_{e}=T_{e}^{k}=\mathbb{Z}_{2}^{k}\), which is impossible because 3 | |N e | (Lemma 2.3). Thus, G is an almost simple group. □

From Lemma 2.4 we find that if XK 5,5 then the group G is almost simple, and the edge stabilizer G e is a maximal subgroup, and G e =A4.ℤ2, S4.ℤ2, (A4×A4).ℤ2, (S4×S4).ℤ2, or ((A4×A4).ℤ2).ℤ2.

3 Classification

In this section, we prove Theorem 1.1. First, we introduce the so-called coset graph. Let G be a finite group, H a subgroup of G and D=D −1 a union of several double-cosets of the form HgH with gH. The coset graph X=Cos(G,H,D) of G with respect to H and D is defined to have vertex set V= [G:H], the set of the right cosets of H in G, and edge set E={{Hg,Hdg}∣gG,dD}. Then X is well defined and has valency |D|/|H|. Furthermore, X is connected if and only if D generates G. Note that G acts on V by right multiplication and so we can view G/H G as a subgroup of Aut(X), where H G is the largest normal subgroup of G contained in H. We may show that G acts transitively on the arcs of X if and only if D=HgH for some gGH (see [7, 8]). The following two examples were described in [3, Sect. 8].

Example 3.1

Let p be a prime and let G=PSL(2,p) with p≡±1, ±9 ( mod 40). Then by [3, Proposition 8.5], G has a subgroup H=A5 and one conjugacy class of maximal subgroups K=S4 such that KH=A4. Take an involution gKH. Define the pentavalent PSL(2,p)-graph as Cos(G,H,HgH). Then the PSL(2,p)-graph is edge-primitive and has automorphism group PSL(2,p). Furthermore, any connected G-edge-primitive pentavalent graph is isomorphic to the PSL(2,p)-graph.

Example 3.2

Let p be a prime and let G=PGL(2,p) with p≡±11, ±19 ( mod 40). Then by [3, Proposition 8.5], G has a subgroup H=A5 and one conjugacy class of maximal subgroups K=S4 such that KH=A4. Take an involution gKH. Define the pentavalent PGL(2,p)-graph as Cos(G,H,HgH). Then the PGL(2,p)-graph is an edge-primitive graph and has automorphism group PGL(2,p). Furthermore, any connected G-edge-primitive pentavalent graph is isomorphic to the PGL(2,p)-graph.

Now we construct an edge-primitive graph, which was given by Weiss [12].

Example 3.3

Let G=Aut(J3)=J3.ℤ2. Then by the Atlas [1], G has maximal subgroups \(H=\mathbb{Z}_{2}^{4}\rtimes \mbox{$\varGamma$L}(2,4)\) and \(K=(\mathbb{Z}_{2}^{4}\rtimes(\mathrm{A}_{4}\rtimes \mathrm{S}_{3})).\mathbb{Z}_{2}\) such that \(K\cap H=\mathbb{Z}_{2}^{4}\rtimes(\mathrm{A}_{4}\rtimes \mathrm{S}_{3})\). Define the pentavalent J3.ℤ2-graph as Cos(G,H,HgH), where gKH. Then this is a 4-transitive edge-primitive graph, and has automorphism group J3.ℤ2. Furthermore, any connected G-edge-primitive pentavalent graph is isomorphic to the J3.ℤ2-graph.

The following two edge-primitive pentavalent graphs are extracted from [5, Sect. 2].

Example 3.4

Let G=Aut(PSL(3,4))=PSL(3,4).D 12. By the Atlas [1], G has a maximal subgroup \(K=(\mathbb{Z}_{2}^{4}\rtimes(\mathrm{A}_{4}\rtimes \mathrm{S}_{3})).\mathbb{Z}_{2}\) and a subgroup \(H=\mathbb{Z}_{2}^{4}\rtimes \mbox{$\varGamma$L}(2,4)\) such that \(K\cap H=\mathbb{Z}_{2}^{4}\rtimes(\mathrm{A}_{4}\rtimes \mathrm{S}_{3})\). Define the pentavalent PSL(3,4).D 12-graph as Cos(G,H,HgH), where gKH. Then this is a 4-transitive edge-primitive graph and has automorphism group PSL(3,4).D 12. Furthermore, any connected G-edge-primitive pentavalent graph is isomorphic to the PSL(3,4).D 12-graph.

Example 3.5

Let G=Aut(PSp(4,4))=PSp(4,4).ℤ4. By the Atlas [1], G has a maximal subgroup \(K=(\mathbb{Z}_{2}^{6}\rtimes(\mathrm{A}_{4}\rtimes \mathrm{S}_{3})).\mathbb{Z}_{2}\) and a subgroup \(H=\mathbb{Z}_{2}^{6}\rtimes \mbox{$\varGamma$L}(2,4)\) such that \(K\cap H=\mathbb{Z}_{2}^{6}\rtimes(\mathrm{A}_{4}\rtimes \mathrm{S}_{3})\). Define the pentavalent PSp(4,4).ℤ4-graph as Cos(G,H,HgH), where gKH. Then this is a 5-transitive edge-primitive graph and has automorphism group PSp(4,4).ℤ4. Furthermore, any connected G-edge-primitive pentavalent graph is isomorphic to the PSp(4,4).ℤ4-graph.

Proof of Theorem 1.1

The graph X has an edge e={v,u}, and is A-edge-primitive, where A=Aut(X). Clearly, K 5,5 is 3-transitive and edge-primitive. Now assume XK 5,5. By [3, Lemmas 3.1 and 3.4], X is a connected (A,s)-transitive graph for s≥1.

Let s≤3. By Lemma 2.4, T=Soc(A) is a non-abelian simple group. Note that A e is a {2,3}-group and hence solvable. By [6, Theorem 1.1], A has a normal subgroup B which is minimal under the condition that B e =BA e is maximal in B, and the pairs (B,B e ) are given in [6, Tables 14–20]. Since \(B\unlhd A\), the edge-primitivity of A implies that B is edge-transitive and hence edge-primitive by the maximality of B e in B. Again by [3, Lemma 3.4], X is B-arc-transitive, and by Lemma 2.2, B is 2- or 3-transitive and B v is non-solvable. Clearly, Soc(B)=T. By Lemma 2.3, B e =A4.ℤ2, S4.ℤ2, (A4×A4).ℤ2, (S4×S4).ℤ2, or ((A4×A4).ℤ2).ℤ2. Checking the pairs (B,B e ) listed in [6, Tables 14–20], we have T=A6=PSL(2,9), PSL(2,p) with p a prime (p≡±1 ( mod 8) or p≡±11, ±19 ( mod 40)), or PSL(3,2). Since \(5\,{\not|}\,\,\,|\mbox{PSL}(3,2)|\), we have T=PSL(2,9) or PSL(2,p) (p≡±1 ( mod 8), or p≡±11, ±19 ( mod 40)). Since X has valency 5, by [3, Theorem 1.3], X is isomorphic to K 6, the PSL(2,p)-graph or the PGL(2,p)-graph.

Let s≥4. By [6, Theorem 1.3], X is isomorphic to the J3.ℤ2-graph, the PSL(3,4).D 12-graph or the PSp(4,4).ℤ4-graph. □