This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the order of semiregular automorphisms of cubic vertex-transitive graphs

Marco Barbieri Dipartimento di Matematica“Felice Casorati”, University of Pavia, Via Ferrata 5, 27100 Pavia, Italy marco.barbieri07@universitadipavia.it Valentina Grazian Dipartimento di Matematica e Applicazioni, University of Milano-Bicocca, Via Cozzi 55, 20125 Milano, Italy valentina.grazian@unimib.it  and  Pablo Spiga Dipartimento di Matematica e Applicazioni, University of Milano-Bicocca, Via Cozzi 55, 20125 Milano, Italy pablo.spiga@unimib.it
Abstract.

We prove that, if Γ\Gamma is a finite connected cubic vertex-transitive graph, then either there exists a semiregular automorphism of Γ\Gamma of order at least 66, or the number of vertices of Γ\Gamma is bounded above by an absolute constant.

Key words and phrases:
Valency 3, Vertex-transitive, Semiregular
2010 Mathematics Subject Classification:
05C25, 20B25

1. Introduction

A fascinating old-standing question in the theory of group actions on graphs is the so-called Polycirculant Conjecture: non-identity 22-closed transitive permutation groups contain non-identity semiregular elements. This formulation of the conjecture was introduced by Klin [Kli98]. However, the question was previously posed independently by Marušič [Mar81, Problem 2.4] and Jordan [Jor88] in terms of graphs: vertex-transitive graphs having more than one vertex admit non-identity semiregular automorphisms.

In this paper, we focus our attention on cubic graphs. In [MS98], Marusič and Scappellato proved that, each cubic vertex-transitive graph admits a non-identity semiregular automorphism, settling the Polycirculant Conjecture for such graphs. Their proof did not take into account the order of the semiregular elements. In this direction, Cameron et al. proved in [CSS06] that, if Γ\Gamma is a cubic vertex-transitive graph, then Aut(Γ)\mathrm{Aut}(\Gamma) contains a semiregular automorphism of order at least 44. They also conjectured that, as the number of vertices of Γ\Gamma tends to infinity, the maximal order of a semiregular automorphism tends to infinity. This was proven false by the third author in [Spi14] by building a family of cubic vertex-transitive graphs where such a maximum is precisely 66. In the light of these results, it is unclear whether 66 is optimal in the sense of minimizing the maximal order of a semiregular element. Broadly speaking, we are interested in

(1.1) lim inf|VΓ|Γ cubic vertex-transitivemax{o(g)gAut(Γ),g semiregular},\displaystyle\liminf_{\begin{subarray}{c}|V\Gamma|\to\infty\\ \Gamma\textrm{ cubic vertex-transitive}\end{subarray}}\max\{o(g)\mid g\in\mathrm{Aut}(\Gamma),g\textrm{ semiregular}\},

where we denote by o(g)o(g) the order of the group element gg.

Theorem 1.1.

The value of (1.1) is 66.

Theorem 1.1 is a consequence of the following result and the main result in [Spi14].

Theorem 1.2.

Let (Γ,G)(\Gamma,G) be a pair such that Γ\Gamma is a connected cubic graph and GG is a subgroup of the automorphism group of Γ\Gamma acting vertex-transitively on VΓV\Gamma. Then either GG contains a semiregular automorphism of order at least 66 or the pair (Γ,G)(\Gamma,G) appears in Table LABEL:table:table.

There is a considerable amount of work into the proof of Theorem 1.2. Broadly speaking, the proof divides into two main cases. In the first main case, the exponent of the group GG is very small, bounded above by 55, and we use explicit knowledge on the finite groups having exponent at most 55. The second main case is concerned with graphs admitting a normal quotient which is a cycle. Here, we need to refine our knowledge on the ubiquitous Praeger-Xu graphs and on the splitting and merging operators between cubic vertex-transitive graphs and 44-valent arc-transitive graphs defined in [PSV13].

Remark 1.3.

The veracity of Theorem 1.2 for graphs with at most 1 2801\,280 vertices has been proven computationally using the database of small cubic vertex-transitive graphs in [PSV13]. Therefore, in the course of the proof of Theorem 1.2 whenever we reduce to a graph having at most 1 2801\,280 vertices we simply refer to this computation.

Table LABEL:table:table consists of six columns. In the first column, we report the number of vertices of the exceptional cubic vertex-transitive graph Γ\Gamma. In the second column, we report the order of the transitive subgroups GG of Aut(Γ)\mathrm{Aut}(\Gamma) with GG not containing semiregular elements of order at least 66: each subgroup is reported up to Aut(Γ)\mathrm{Aut}(\Gamma)-conjugacy class. In the third column, we report the cardinality of Aut(Γ)\mathrm{Aut}(\Gamma). In the forth column, when |VΓ|1 280|\mathrm{V}\Gamma|\leq 1\,280, we report the number of the graph in the database of small cubic vertex-transitive graphs in [PSV13]. In the fifth column of Table LABEL:table:table, we write the symbol when the graph is arc-transitive and the symbol {\dagger} when the graph is a split Praeger-Xu graph (see Section 2.5 for the definition of split Praeger-Xu graphs). Split Praeger-Xu graphs play an important role in our investigation and hence we are keeping track of this information in the forth column. In the sixth column, for the graphs not appearing in the database of small cubic vertex-transitive graphs, we give as much information as possible.

Table 1. Exceptional cases for Theorem 1.2
|VΓ||\mathrm{V}\Gamma| |G||G| |Aut(Γ)||\mathrm{Aut}(\Gamma)| DB /\leavevmode\hbox to11.38pt{\vbox to7.97pt{\pgfpicture\makeatletter\hbox{\hskip 0.0pt\lower 0.0pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{{}}{}{{}}{} {}{} {}{} {}{} {}\pgfsys@moveto{0.0pt}{3.98338pt}\pgfsys@lineto{2.84523pt}{0.0pt}\pgfsys@lineto{11.38092pt}{7.96661pt}\pgfsys@lineto{2.84523pt}{1.70706pt}\pgfsys@closepath\pgfsys@fill\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}\;/\;{\dagger} Comments
4 4, 4, 8, 12, 24 24 1
6 6 12 1
6, 36 24 2
8 8 16 1
8, 8, 8, 8, 16, 24, 24, 48 48 2
10 10 20 1
10 20 2
20, 60, 120 120 3
12 12, 24 24 2
24, 24 48 4 {\dagger}
16 16, 16, 32, 32, 64, 64 128 2 {\dagger}
16 32 3
16, 48 96 4
18 18, 108 216 4
36 72 5
20 20 20 2
160, 160 320 3 {\dagger}
60 120 6
120 240 7
24 24 144 2
24 48 8
24 24 9
24 48 10
24, 24 48 11 {\dagger}
30 720 1 440 8
60, 120 120 9
60 60 10
32 32 64 2
32, 32, 64, 64 128 3 {\dagger}
32, 96 192 4
36 36 72 9
40 160, 160 320 12 {\dagger}
50 100 200 7
50, 150 300 8
54 108 216 11
60 60 360 2
60, 120 120 3
60 60 4
60 120 5
60 120 6
60, 120 120 7
60 120 8
60 120 9
60 120 10
64 64, 192 384 2
64 256 4
64, 64 128 11 {\dagger}
80 80, 160 160 29
160, 160 320 31 {\dagger}
90 720 1 440 20
96 96 192 37
100 100 200 19
128 128 256 5
160 160 160 89
160 160 90
160 320 91
160 320 92
160 320 93 {\dagger}
160 320 94
180 720 77
360, 720 78
250 500 31
256 256, 768 30
360 360 720 176
360 720 177
360 720 178
360 360 179
360 720 180
360 360 181
360 720 182
360 720 183
360 720 184
360 720 185
720 1 440 268
720 1 440 270
512 512 1 024 734
810 1 620 1 620 198
1 024 1 024, 3 072 6 144 3 470
1 250 2 500 2 500 187
1 280 1 280 2 500 2 591
2 560 2 560 5 120
6 250 12 500 25 000 covers of the graph with 1 2501\,250
12 500 12 500 vertices, there are 2 graphs
31 250 62 500 125 000 covers of the graphs
62 500 125 000 with 6 2506\,250 vertices,
62 500 125 000 there are five graphs
62 500 62 500
62 500 62 500
65 610 131 220 ? cover of the graph with 810810 vertices, only one graph
2\cdot 5 4\cdot 5 7347\leq\ell\leq 34

2. Main ingredients

2.1. Permutations

A permutation on the set Ω\Omega is a derangement if it fixes no elements in Ω\Omega. A permutation is semiregular if all of its cycles have the same length. For instance, any derangement of prime order is semiregular. A permutation group GG on Ω\Omega is said to be transitive if it has a single orbit on Ω\Omega, and semiregular if the identity is the only element fixing some points. If GG is both semiregular and transitive on Ω\Omega, then GG is regular on Ω\Omega. Given a permutation group GG, and an element αΩ\alpha\in\Omega, we denote by αG\alpha^{G} the orbit of α\alpha under the action of GG.

Lemma 2.1.

Let GG be a permutation group on Ω\Omega, and let pp be a prime. If all the elements of GG of order pp are derangements, then all pp-elements of GG are semiregular.

Proof.

Let gGg\in G be an element of order pkp^{k}, for some positive integer kk. Aiming for a contradiction, assume that gg is not semiregular, that is, there exists αΩ\alpha\in\Omega such that |αg|pk1|\alpha^{\langle g\rangle}|\leq p^{k-1}. Hence gpk1g^{p^{k-1}} fixes α\alpha, which implies gpk1g^{p^{k-1}} is not a derangement, a contradiction. ∎

Lemma 2.2.

Let GG be a permutation group acting on Ω\Omega, and let pp and qq be two distinct primes. If GG has a semiregular element gg of order pp and a semiregular element hh of order qq with gh=hggh=hg, then ghgh is a semiregular element of order pqpq.

Proof.

Since gh=hggh=hg, o(gh)=pqo(gh)=pq and hence it remains to prove that ghgh is semiregular. Note that (gh)p=hp(gh)^{p}=h^{p} is semiregular, and also (gh)q=gq(gh)^{q}=g^{q} is semiregular. Therefore, each orbit of gh\langle gh\rangle has size pqpq, proving that ghgh is semiregular. ∎

2.2. Graphs

A digraph is a binary relation Γ=(VΓ,AΓ),\Gamma=(V\Gamma,A\Gamma), where AΓVΓ×VΓA\Gamma\subseteq V\Gamma\times V\Gamma. We refer to the elements of VΓV\Gamma as vertices and to the elements of AΓA\Gamma as arcs. In this paper, a graph is a finite simple undirected graph, that is, a pair Γ=(VΓ,EΓ),\Gamma=(V\Gamma,E\Gamma), where VΓV\Gamma is a set of vertices, and EΓE\Gamma is a set of unordered pairs of VΓV\Gamma, called edges. In particular, a graph can be thought of as a digraph where the binary relation is symmetric and anti-reflexive.

The valency of a vertex αVΓ\alpha\in V\Gamma is the number of edges containing α\alpha. A graph is said to be cubic when all of its vertices have valency 33. A connected graph is a cycle when all of its vertices have valency 22.

Let Γ\Gamma be a graph, and let GG be a subgroup of the automorphism group Aut(Γ)\mathrm{Aut}(\Gamma) of Γ\Gamma. If GG is transitive on VΓV\Gamma, we say that GG is vertex-transitive, similarly, if GG is transitive on AΓA\Gamma, we say that GG is arc-transitive. Moreover, Γ\Gamma is vertex- or arc-transitive when Aut(Γ)\mathrm{Aut}(\Gamma) is vertex- or arc-transitive.

Let α,βVΓ\alpha,\beta\in V\Gamma be two adjacent vertices. We denote by GαG_{\alpha} the stabilizer of the vertex α\alpha, by G{α,β}G_{\{\alpha,\beta\}} the setwise stabilizer of the edge {α,β}\{\alpha,\beta\}, by GαβG_{\alpha\beta} the pointwise stabilizer of the edge {α,β}\{\alpha,\beta\} (that is, the stabilizer of the arc (α,β)(\alpha,\beta) underlying the edge {α,β}\{\alpha,\beta\}).

Let Γ\Gamma be a graph, and let NAut(Γ)N\leq\mathrm{Aut}(\Gamma). The normal quotient Γ/N\Gamma/N is the graph whose vertices are the NN-orbits of VΓV\Gamma, and two NN-orbits αN\alpha^{N} and βN\beta^{N} are adjacent if there exists an edge {α,β}EΓ\{\alpha^{\prime},\beta^{\prime}\}\in E\Gamma such that ααN\alpha^{\prime}\in\alpha^{N} and ββN\beta^{\prime}\in\beta^{N}. Note that the valency of Γ/N\Gamma/N is at most the valency of Γ\Gamma, and that, whenever Γ\Gamma is conneted, so is Γ/N\Gamma/N. Furthermore, if the group NN is normal in some GAut(Γ)G\leq\mathrm{Aut}(\Gamma), then G/NG/N acts (possibly unfaithfully) on Γ/N\Gamma/N. If the group GG acts vertex- or arc-transitively on Γ\Gamma, then G/NG/N has the same property on Γ/N\Gamma/N.

The following result is inspired by an analogous result for 44-valent graphs in [PS21, Lemma 1.13].

Lemma 2.3.

Let Γ\Gamma be a connected cubic graph, let α\alpha be a vertex of Γ\Gamma, let GG be a vertex-transitive subgroup of Aut(Γ)\mathrm{Aut}(\Gamma) and let NN be a semiregular normal subgroup of GG. Suppose GαG_{\alpha} is a non-identity 22-group and that the normal quotient Γ/N\Gamma/N is a cycle of length r3r\geq 3, and denote by KK the kernel of the action of GG on the NN-orbits on VΓV\Gamma. Then either

  1. (1)

    GαG_{\alpha} has order 22 and |Kα|=1|K_{\alpha}|=1, or

  2. (2)

    rr is even and Gα=KαG_{\alpha}=K_{\alpha} is an elementary abelian 22-group of order at most 2r/22^{r/2}.

Proof.

Let Δ0,Δ1,,Δr1\Delta_{0},\Delta_{1},\ldots,\Delta_{r-1} be the orbits of NN in its action on VΓV\Gamma. Since Γ/N\Gamma/N is a cycle, we may assume that Δi\Delta_{i} is adjacent to Δi1\Delta_{i-1} and Δi+1\Delta_{i+1} with indices computed modulo rr. Moreover, without loss of generality, we suppose that αΔ0\alpha\in\Delta_{0}.

As GαG_{\alpha} is a non-identity 22-group, by a connectedness argument, GαG_{\alpha} induces a group of order 22 in its action on the neighbourhood of α\alpha. In particular, GαG_{\alpha} fixes a unique neighbour of α\alpha. As usual, for each βVΓ\beta\in V\Gamma, let β\beta^{\prime} be the unique neighbour of β\beta fixed by GβG_{\beta}.

Suppose that {α,α}\{\alpha,\alpha^{\prime}\} is contained in an NN-orbit. Since αΔ0\alpha\in\Delta_{0}, we deduce αΔ0\alpha^{\prime}\in\Delta_{0}. Let β\beta and γ\gamma be the other two neighbours of α\alpha. As Γ/N\Gamma/N is a cycle of length r3r\geq 3, we have βΔ1\beta\in\Delta_{1} and γΔr1\gamma\in\Delta_{r-1}. Since Aut(Γ/N)\mathrm{Aut}(\Gamma/N) is a dihedral group of order 2r2r and since GαG_{\alpha} contains an element swapping β\beta and γ\gamma, we deduce |Gα:Kα|=2|G_{\alpha}:K_{\alpha}|=2. Now, KαK_{\alpha} fixes by definition each NN-orbit and hence it fixes setwise Δ1\Delta_{1} and Δr1\Delta_{r-1}. Therefore, KαK_{\alpha} fixes β\beta and γ\gamma, because β\beta is the unique neighbour of α\alpha in Δ1\Delta_{1} and γ\gamma is the unique neighbour of α\alpha in Δr1\Delta_{r-1}. This shows that KαK_{\alpha} fixes pointwise the neighbourhood of α\alpha; now, a connectedness argument shows that Kα=1K_{\alpha}=1. In particular, part (1) is satisfied. For the rest of the argument, we suppose that {α,α}\{\alpha,\alpha^{\prime}\} is not contained in an NN-orbit.

This means that α\alpha has two neighbours in an NN-orbit, say Δ1\Delta_{1}, and only one neighbour in the other NN-orbit, say Δr1\Delta_{r-1}. (Thus αΔr1\alpha^{\prime}\in\Delta_{r-1} and β,γΔ1\beta,\gamma\in\Delta_{1}.) This implies that rr is even and, for every i{0,,r/21}i\in\{0,\ldots,r/2-1\}, each vertex in Δ2i\Delta_{2i} has two neighbours in Δ2i+1\Delta_{2i+1} and only one neighbour in Δ2i1\Delta_{2i-1}. Therefore, G/KG/K is a dihedral group of order rr when r8r\geq 8 and G/KG/K is elementary abelian of order 44 when r=4.r=4. Morever, G/KG/K acts regularly on Γ/N\Gamma/N and hence Gα=KαG_{\alpha}=K_{\alpha}. It remains to show that KαK_{\alpha} is an elementary abelian 22-group of order at most 2r2^{r}.

Since NN is normal in GG, the orbits of NN on the edge-set EΓE\Gamma form a GG- invariant partition of EΓE\Gamma. We claim that, no two edges incident to a fixed vertex of Γ\Gamma belong to the same NN-edge-orbit. We argue by contradiction and we suppose that α\alpha has two distinct neighbours vv and ww such that the edges {α,v}\{\alpha,v\} and {α,w}\{\alpha,w\} are in the same NN-edge-orbit. In particular, there exists nNn\in N with {α,v}n={α,w}\{\alpha,v\}^{n}=\{\alpha,w\}. This gives αn=α\alpha^{n}=\alpha and vn=wv^{n}=w, or αn=w\alpha^{n}=w and vn=αv^{n}=\alpha. Since there are no edges inside an NN-orbit, we cannot have αn=w\alpha^{n}=w and vn=αv^{n}=\alpha. Therefore, αn=α\alpha^{n}=\alpha and vn=wv^{n}=w. Since NN acts semiregularly on VΓV\Gamma, we have n=1n=1 and hence v=vn=wv=v^{n}=w, which is a contradiction.

Since GG is vertex-transitive, the edges between Δ2i\Delta_{2i} and Δ2i+1\Delta_{2i+1} are partitioned into precisely two NN-edge-orbits, let’s call these two orbits Θ2i\Theta_{2i} and Θ2i\Theta_{2i}^{\prime}; whereas the edges between Δ2i\Delta_{2i} and Δ2i1\Delta_{2i-1} form one NN-edge-orbit, which we call Θ2i′′\Theta_{2i}^{\prime\prime}.

An element of KK (fixing setwise the sets Δ2i\Delta_{2i} and Δ2i+1\Delta_{2i+1}) can map an edge in Θ2i\Theta_{2i} only to an edge in Θ2i\Theta_{2i} or to an edge in Θ2i\Theta_{2i}^{\prime}. On the other hand, as GαG_{\alpha} is not the identity group, for every vertex vΔ2iv\in\Delta_{2i} there is an element gGvg\in G_{v} which maps an edge of Θ2i\Theta_{2i} incident to vv to the edge of Θ2i\Theta_{2i}^{\prime} incident to vv; and this element gg is clearly an element of KK, because G/KG/K acts semiregularly on Γ/N\Gamma/N. This shows that the orbits of KK on EΓE\Gamma are precisely the sets Θ2iΘ2i,Θ2i′′\Theta_{2i}\cup\Theta_{2i}^{\prime},\Theta_{2i}^{\prime\prime}, i{0,,r/21}i\in\{0,\ldots,r/2-1\}. In other words, each orbit of the induced action of KK on the set EΓ/N={eN:eEΓ}E\Gamma/N=\{e^{N}:e\in E\Gamma\} has length at most 22. Consequently, if XX denotes the kernel of the action of KK on EΓE\Gamma, then K/XK/X embeds into Sym(2)r/2\mathrm{Sym}(2)^{r/2} and is therefore an elementary abelian 2- group of order at most 2r/22^{r/2}.

Let us now show that X=NX=N. Clearly, NXN\leq X. Let vΔ0v\in\Delta_{0}. Since NN is transitive on Δ0\Delta_{0}, it follows that X=NXvX=NX_{v}. Suppose that XvX_{v} is non-trivial and let gg be a non-trivial element of XvX_{v}. Further, let ww be a vertex which is closest to vv among all the vertices not fixed by gg, and let v=v0v1vm=wv=v_{0}\sim v_{1}\sim\cdots\sim v_{m}=w be a shortest path from vv to ww. Then vm1v_{m-1} is fixed by gg. Since gg fixes each NN-edge-orbit setwise and since every vertex of Γ\Gamma is incident to at most one edge in each NN-edge-orbit, it follows that gg fixes all the neighbours of vm1v_{m-1}, thus also vmv_{m}. This contradicts our assumptions and proves that XvX_{v} is a trivial group, and hence that X=NX=N. ∎

2.3. Praeger-Xu graphs

To introduce the infinite family of split Praeger-Xu graphs sC(r,s)\mathrm{sC}(r,s), we need two ingredients: the Praeger-Xu graphs and the splitting operation. This section is devoted to introduce the ubiquitous 44-valent Praeger-Xu graphs C(r,s)\mathrm{C}(r,s) and their automorphism group. This infinite family was originally defined in [PX89], and it was studied in detail by Gardiner, Praeger and Xu in [PX89, GP94], and more recently in [JPW19]. Here, we introduce them through their directed counterparts defined in [Pra89].

Let rr be an integer, r3r\geq 3. Then C(r,1)\vec{\mathrm{C}}(r,1) is the lexicographic product of a directed cycle of length rr with an edgeless graph on 22 vertices. In other words, VC(r,1)=r×2\mathrm{V}\vec{\mathrm{C}}(r,1)=\mathbb{Z}_{r}\times\mathbb{Z}_{2} with the out-neighbours of a vertex (x,i)(x,i) being (x+1,0)(x+1,0) and (x+1,1)(x+1,1). We will identify the (s1)(s-1)-arc

(x,ε0)(x+1,ε1)(x+s1,εs1)(x,\varepsilon_{0})\sim(x+1,\varepsilon_{1})\sim\ldots\sim(x+s-1,\varepsilon_{s-1})

with the pair (x;k)(x;k) where k=ε0ε1εs1k=\varepsilon_{0}\varepsilon_{1}\ldots\varepsilon_{s-1} is a string in 2\mathbb{Z}_{2} of length ss. For s2s\geq 2, let VC(r,s)\mathrm{V}\vec{\mathrm{C}}(r,s) be the set of all (s1)(s-1)-arcs of C(r,1)\vec{\mathrm{C}}(r,1), let hh be a string in 2\mathbb{Z}_{2} of length s1s-1 and let ε2\varepsilon\in\mathbb{Z}_{2}. The out-neighbours of (x;εh)VC(r,s)(x;\varepsilon h)\in\mathrm{V}\vec{\mathrm{C}}(r,s) are (x+1;h0)(x+1;h0) and (x+1;h1)(x+1;h1). The Praeger-Xu graph C(r,s)\mathrm{C}(r,s) is then defined as the underlying graph of C(r,s)\vec{\mathrm{C}}(r,s). We have that C(r,s)\mathrm{C}(r,s) is a connected 44-valent graph with r2sr2^{s} vertices (see [Pra89, Theorem 2.8]).

Let us now discuss the automorphisms of the graphs C(r,s)\mathrm{C}(r,s). Every automorphism of C(r,1)\vec{\mathrm{C}}(r,1) (C(r,1)\mathrm{C}(r,1), respectively) acts naturally as an automorphism of C(r,s)\vec{\mathrm{C}}(r,s) (C(r,s)\mathrm{C}(r,s), respectively) for every s2s\geq 2. For iri\in\mathbb{Z}_{r}, let τi\tau_{i} be the transposition on VC(r,1)\mathrm{V}\vec{\mathrm{C}}(r,1) swapping the vertices (i,0)(i,0) and (i,1)(i,1) while fixing every other vertex. This is clearly an automorphism of C(r,1)\vec{\mathrm{C}}(r,1), and thus also of C(r,s)\vec{\mathrm{C}}(r,s) for s2s\geq 2. Let

K:=τiir,K:=\langle\tau_{i}\mid i\in\mathbb{Z}_{r}\rangle,

and observe that KC2rK\cong C_{2}^{r}. Further, let ρ\rho and σ\sigma be the permutations on VC(r,1)\mathrm{V}\vec{\mathrm{C}}(r,1) defined by

(x,i)ρ:=(x+1,i)and(x,i)σ:=(x,i).(x,i)^{\rho}:=(x+1,i)\quad\hbox{and}\quad(x,i)^{\sigma}:=(x,-i).

Then ρ\rho is an automorphism of C(r,1)\vec{\mathrm{C}}(r,1) or order rr, and σ\sigma is an involutory automorphism of C(r,1)\mathrm{C}(r,1) (but not of C(r,1)\vec{\mathrm{C}}(r,1)). Observe that the group ρ,σ\langle\rho,\sigma\rangle normalises KK. Let

H:=Kρ,σandH+:=Kρ.H:=K\langle\rho,\sigma\rangle\quad\hbox{and}\quad H^{+}:=K\langle\rho\rangle.

Then, for every r3r\geq 3 and s1s\geq 1,

C2wrDrHAut(C(r,s))andC2wrCrH+Aut(C(r,s)).C_{2}\mathrm{wr}D_{r}\cong H\leq\mathrm{Aut}(\mathrm{C}(r,s))\quad\textup{and}\quad C_{2}\mathrm{wr}C_{r}\cong H^{+}\leq\mathrm{Aut}(\vec{\mathrm{C}}(r,s)).

Moreover, HH (H+H^{+}, respectively) acts arc-transitively on C(r,s)\mathrm{C}(r,s) (C(r,s)\vec{\mathrm{C}}(r,s), respectively) whenever 1sr11\leq s\leq r-1. With three exceptions, the groups HH and H+H^{+} are in fact the full automorphism groups of C(r,s)\mathrm{C}(r,s) and C(r,s)\vec{\mathrm{C}}(r,s), respectively.

Lemma 2.4 ([GP94, Theorem 2.13] and [Pra89, Theorem 2.8]).

The automorphism group of a directed Praeger-Xu graph is

Aut(C(r,s))=H+,\mathrm{Aut}(\vec{\mathrm{C}}(r,s))=H^{+},

and, if r4r\neq 4, the automorphism group of a Praeger-Xu graph is

Aut(C(r,s))=H.\mathrm{Aut}(\mathrm{C}(r,s))=H.

Moreover,

|Aut(C(4,1)):H|=9,|Aut(C(4,2)):H|=3|\mathrm{Aut}(\mathrm{C}(4,1)):H|=9,\quad|\mathrm{Aut}(\mathrm{C}(4,2)):H|=3
and|Aut(C(4,3)):H|=2.\hbox{and}\quad|\mathrm{Aut}(\mathrm{C}(4,3)):H|=2.

The Praeger-Xu graphs also admit the following algebraic characterization.

Lemma 2.5 ([PS21, Lemma 1.11] or [BGS22b, Lemma 3.7]).

Let Γ\Gamma be a finite connected 44-valent graph, let GG be a vertex- and edge-transitive group of automorphisms of Γ\Gamma, and let NN be a minimal normal subgroup of GG. If NN is a 22-group and Γ/N\Gamma/N is a cycle of length at least 33, then Γ\Gamma is isomorphic to a Praeger-Xu graph C(r,s)\mathrm{C}(r,s) for some positive integers r3r\leq 3 and sr1s\leq r-1.

For more details on Praeger-Xu graphs, we refer also to [JPW19, JPW22, BGS22a].

2.4. The splitting and merging operations

The operation of splitting were introduced in [PSV13, Construction 11]. Let Δ\Delta be a 44-valent graph, let 𝒞\mathcal{C} be a partition of EΔE\Delta into cycles. By applying the splitting operation to the pair (Δ,𝒞)(\Delta,\mathcal{C}), we obtain the graph, denoted by s(Δ,𝒞)\mathrm{s}(\Delta,\mathcal{C}), whose vertices are

Vs(Δ,𝒞):={(α,C)VΔ×𝒞αVC},V\mathrm{s}(\Delta,\mathcal{C}):=\{(\alpha,C)\in V\Delta\times\mathcal{C}\mid\alpha\in VC\},

and such that two vertices (α,C)(\alpha,C) and (β,D)(\beta,D) are declared adjacent if either CDC\neq D and α=β\alpha=\beta, or C=DC=D and α\alpha and β\beta are adjacent in Δ\Delta. Observe that, since Δ\Delta is 44-valent, there are precisely 22 cycles in 𝒞\mathcal{C} passing through α\alpha, thus s(Δ,𝒞)\mathrm{s}(\Delta,\mathcal{C}) is cubic and |Vs(Δ,𝒞)|=2|VΔ||V\mathrm{s}(\Delta,\mathcal{C})|=2|V\Delta|.

Notice that, for any GAut(Δ)G\leq\mathrm{Aut}(\Delta) such that its action is 𝒞\mathcal{C}-invariant, GAut(s(Δ,𝒞))G\leq\mathrm{Aut}(\mathrm{s}(\Delta,\mathcal{C})). Moreover, if GG is also arc-transitive on Δ\Delta (in particular, the action of GαG_{\alpha} on the neighbourhood of α\alpha is either the Klein four group, or the cyclic group of order 44, or the dihedral group of order 88), then GG is vertex-transitive on s(Δ,𝒞)\mathrm{s}(\Delta,\mathcal{C}). For any vertex (α,C)s(Δ,𝒞)(\alpha,C)\in\mathrm{s}(\Delta,\mathcal{C}),

G(α,C)=GαG{C},G_{(\alpha,C)}=G_{\alpha}\cap G_{\{C\}},

where G{C}G_{\{C\}} is the setwise stabilizer of the cycle CC. In particular, whenever GG is arc-transitive on Δ\Delta, as GαG_{\alpha} switches the two cycles passing through α\alpha, |Gα:G(α,C)|=2|G_{\alpha}:G_{(\alpha,C)}|=2.

Now, we introduce the tentative inverse of the splitting operator: the operation of merging (see [PSV13, Construction 7]). Let Γ\Gamma be a connected cubic graph, and let GAut(Γ)G\leq\mathrm{Aut}(\Gamma) be a vertex-transitive group such that the action of GαG_{\alpha} on the neighbourhood of α\alpha is cyclic of order 22. In particular, GαG_{\alpha} is a non-identity 22-group. Hence, GαG_{\alpha} fixes a unique neighbour of α\alpha, which we denote by α\alpha^{\prime}. Observe that (α)=α(\alpha^{\prime})^{\prime}=\alpha and Gα=GαG_{\alpha}=G_{\alpha^{\prime}}. Thus, the set :={{α,α}αVΓ}\mathcal{M}:=\{\{\alpha,\alpha^{\prime}\}\mid\alpha\in V\Gamma\} is a complete matching of Γ\Gamma, while the edges outside \mathcal{M} form a 2-factor, which we denote by \mathcal{F}. The group GG in its action on EΓE\Gamma fixes setwise both \mathcal{F} and \mathcal{M}, and acts transitively on the arcs of each of these two sets. Let Δ\Delta be the graph with vertex-set \mathcal{M} and two vertices e1,e2e_{1},e_{2}\in\mathcal{M} are declared adjacent if they are (as edges of Γ\Gamma) at distance 11 in Γ\Gamma. We may also think of Δ\Delta as being obtained by contracting all the edges in \mathcal{M}. Let 𝒞\mathcal{C} be the decomposition of EΔE\Delta into cycles given by the connected components of the the 2-factor \mathcal{F}. The merging operation applied to the pair (Γ,G)(\Gamma,G) gives as a result the pair (Δ,𝒞)(\Delta,\mathcal{C}).

Two infinite families of cubic graph have degenerate merged graphs, namely the circular and Möbius ladders. For any n3n\geq 3, a circular ladder graph is a graph isomorphic to the Cayley graph

Cay(n×2,{(0,1),(1,0),(1,0)}),\mathrm{Cay}(\mathbb{Z}_{n}\times\mathbb{Z}_{2},\{(0,1),(1,0),(-1,0)\}),

and, for any n2n\geq 2, a Möbius ladder graph is a graph isomorphic to the Cayley graph

Cay(2n,{1,1,n}).\mathrm{Cay}(\mathbb{Z}_{2n},\{1,-1,n\}).

Observe that we consider the complete graph on 44 vertices to be a Möbius ladder graph.

Lemma 2.6.

Let Λ\Lambda be a (circular or Möbius) ladder, and let GAut(Λ)G\leq\mathrm{Aut}(\Lambda) be a vertex-transitive group. Then either |VΛ|10|V\Lambda|\leq 10 or GG contains a semiregular element of order at least 66.

Lemma 2.7.

Unless Λ\Lambda is isomorphic to the skeleton of the cube or the complete graph on 44 vertices, the automorphism group of a (circular or Möbius) ladder Λ\Lambda contains NAut(Λ)N\leq\mathrm{Aut}(\Lambda), a normal cyclic subgroup of order 22, such that the normal quotient Λ/N\Lambda/N is a cycle.

Remark 2.8.

Let Γ\Gamma be a connected cubic graph that is neither a circular nor a Möbius ladder, and let GAut(Γ)G\leq\mathrm{Aut}(\Gamma) be a vertex-transitive group such that the action of GαG_{\alpha} on the neighbourhood of α\alpha is cyclic of order 22. Then [PSV13, Lemma 9 and Theorem 10] imply that the merging operator applied to the pair (Γ,G)(\Gamma,G) gives a pair (Δ,𝒞)(\Delta,\mathcal{C}) such that Δ\Delta is 44-valent, and the action of GG on Δ\Delta is faithful, arc-transitive and 𝒞\mathcal{C}-invariant. This result motivates the use of the word degenerate when referring to the circular and Möbius ladders.

In view of [PSV13, Theorem 12], the merging operator is the right-inverse of the splitting one, or, more explicitly, unless Γ\Gamma is a (circular or Möbius) ladder, splitting a pair (Δ,𝒞)(\Delta,\mathcal{C}) obtained via the merging operation on (Γ,G)(\Gamma,G) results in the starting pair. For our purposes, we need to show that the merging operator is also the left-inverse of the splitting one.

Theorem 2.9.

Let Δ\Delta be a 44-valent graph, let 𝒞\mathcal{C} be a partition of EΔE\Delta into cycles, and let GAut(Δ)G\leq\mathrm{Aut}(\Delta) be an arc-transitive and 𝒞\mathcal{C}-invariant group. Then the merging operation can be applied to the pair (s(Δ,𝒞),G)(\mathrm{s}(\Delta,\mathcal{C}),G) and it gives as a result (Δ,𝒞)(\Delta,\mathcal{C}).

Proof.

Let (α,C)(\alpha,C) be a generic vertex of s(Δ,𝒞)\mathrm{s}(\Delta,\mathcal{C}), let D𝒞D\in\mathcal{C} be the other cycle of the partition passing through α\alpha, and let β,γVΔ\beta,\gamma\in V\Delta be the neighbours of α\alpha in CC. Then, using the fact that GG is arc-transitive on CC,

(α,D)G(α,C)={(α,D)}and(β,C)G(α,C)=(γ,C)G(α,C)={(β,C),(γ,C)}.(\alpha,D)^{G_{(\alpha,C)}}=\{(\alpha,D)\}\quad\hbox{and}\quad(\beta,C)^{G_{(\alpha,C)}}=(\gamma,C)^{G_{(\alpha,C)}}=\{(\beta,C),(\gamma,C)\}.

Therefore, for any vertex (α,C)Vs(Δ,𝒞)(\alpha,C)\in V\mathrm{s}(\Delta,\mathcal{C}), G(α,C)G_{(\alpha,C)} acts on the neighbourhood of (α,C)(\alpha,C) as a cyclic group of order 22. Hence, we can apply the merging operation to the pair (s(Δ,𝒞),G)(\mathrm{s}(\Delta,\mathcal{C}),G). Furthermore, we deduce that

={{(α,C),(α,D)}αVCVD}\mathcal{M}=\{\{(\alpha,C),(\alpha,D)\}\mid\alpha\in VC\cap VD\}

is a complete matching for (s(Δ,𝒞),G)(\mathrm{s}(\Delta,\mathcal{C}),G). Thus the connected components of the resulting 22-factor =Es(Δ,𝒞)\mathcal{F}=E\mathrm{s}(\Delta,\mathcal{C})\setminus\mathcal{M} can be identified with the cycles of 𝒞\mathcal{C}. Now, consider the map defined as

θ:VΔ,{(α,C),(α,D)}α.\theta:\mathcal{M}\rightarrow V\Delta,\,\{(\alpha,C),(\alpha,D)\}\mapsto\alpha.

Since a generic vertex αVΔ\alpha\in V\Delta belongs to precisely two distinct cycles, θ\theta is bijective. Moreover, β\beta is adjacent to α\alpha in Δ\Delta if, and only if, either {(α,C),(β,C)}\{(\alpha,C),(\beta,C)\} or {(α,D),(β,D)}\{(\alpha,D),(\beta,D)\} is an edge in s(Δ,𝒞)\mathrm{s}(\Delta,\mathcal{C}). In particular, θ\theta also induces the bijection

θ^:EΔ,{(α,C),(β,C)}{α,β},\hat{\theta}:\mathcal{F}\rightarrow E\Delta,\,\{(\alpha,C),(\beta,C)\}\mapsto\{\alpha,\beta\},

which sends the connected components of \mathcal{F} into disjoint cycles of 𝒞\mathcal{C}. This shows that θ\theta is a graph isomorphism between Δ\Delta and the 44-valent graph obtained by merging the pair (s(Δ,𝒞),G)(\mathrm{s}(\Delta,\mathcal{C}),G), and that the resulting cycle partition is isomorphic to 𝒞\mathcal{C}. ∎

Corollary 2.10.

Let Δ\Delta be a 44-valent graph, let 𝒞\mathcal{C} be a partition of EΔE\Delta into cycles, and let GAut(Δ)G\leq\mathrm{Aut}(\Delta) be an arc-transitive and 𝒞\mathcal{C}-invariant group (and so GAut(s(Δ,𝒞))G\leq\mathrm{Aut}(\mathrm{s}(\Delta,\mathcal{C}))). Suppose that GAAut(s(Δ,𝒞))G\leq A\leq\mathrm{Aut}(\mathrm{s}(\Delta,\mathcal{C})) is a vertex-transitive group such that, for any vertex αVs(Δ,𝒞)\alpha\in V\mathrm{s}(\Delta,\mathcal{C}), the action of AαA_{\alpha} on the neighbourhood of α\alpha is cyclic of order 22, then AAut(Δ)A\leq\mathrm{Aut}(\Delta).

Proof.

Note that, as GG is a subgroup of AA, the actions of GG and AA on the neighbourhood of any vertex α\alpha coincide. In particular, applying the merging operation to the pair (s(Δ,𝒞),A)(\mathrm{s}(\Delta,\mathcal{C}),A) yields the same result as doing it on the pair (s(Δ,𝒞),G)(\mathrm{s}(\Delta,\mathcal{C}),G), that is, by Theorem 2.9, in both cases we obtain (Δ,𝒞)(\Delta,\mathcal{C}). The result follows by Remark 2.8. ∎

2.5. Split Praeger-Xu graphs

In this section, we bring together the information of Sections 2.3 and 2.4 to define and study the split Praeger-Xu graphs.

All the partitions of the edge set of a Praeger-Xu graph into disjoint cycles were classified in [JPW19, Section 6]. Regardless of the choice of the parameters rr and ss, there exists a decomposition into disjoint cycles of length 44 of the form

(x;0h)(x+1;h0)(x;1h)(x+1;h1)(x;0h)\sim(x+1;h0)\sim(x;1h)\sim(x+1;h1)

for some xrx\in\mathbb{Z}_{r}, and for some string hh in 2\mathbb{Z}_{2} of length s1s-1. We denote this partition by 𝒮\mathcal{S}. Moreover, observe that the only two neighbours of (x;0h)(x;0h) in the KK-orbit containing (x+1;h0)(x+1;h0) are (x+1;h1)(x+1;h1) and (x+1;h0)(x+1;h0), and similarly the only two neighbours of (x+1;h0)(x+1;h0) in the KK-orbit containing (x;0h)(x;0h) are (x;1h)(x;1h) and (x;0h)(x;0h). Therefore, 𝒮\mathcal{S} is the unique decomposition such that each cycle intersects exactly two KK-orbits.

Definition 2.11.

The split Praeger-Xu graph sC(r,s)\mathrm{sC}(r,s) is the cubic graph obtained from the pair (C(r,s),𝒮)(\mathrm{C}(r,s),\mathcal{S}) by applying the splitting operation.

Lemma 2.12.

For some positive integers r3r\geq 3 and sr1s\leq r-1, the automorphism group of the split Praeger-Xu graph is

Aut(sC(r,s))=H,\mathrm{Aut}(\mathrm{sC}(r,s))=H,

and it acts transitively on VsC(r,s)V\mathrm{sC}(r,s).

Proof.

Note that HH acts on the set of KK-orbits in VC(r,s)V\mathrm{C}(r,s), thus each automorphism of HH maps any cycle of 𝒮\mathcal{S} to a cycle intersecting exactly two KK-orbits, that is, to an element of 𝒮\mathcal{S}. Thus, HH is 𝒮\mathcal{S}-invariant, and so HAut(sC(r,s))H\leq\mathrm{Aut}(\mathrm{sC}(r,s)). We now show the opposite inclusion. Let αVsC(r,s)\alpha\in V\mathrm{sC}(r,s) be a generic vertex, aiming for a contradiction we suppose that Aut(sC(r,s))α\mathrm{Aut}(\mathrm{sC}(r,s))_{\alpha} does not act on the neighbourhood of α\alpha as a cycle of order 22. Let α,β,γ\alpha^{\prime},\beta,\gamma be the neighbours of α\alpha where α\alpha^{\prime} is fixed by the action of HαH_{\alpha}, and let δ\delta be the unique vertex at distance 11 from both β\beta and γ\gamma. Since HαAut(sC(r,s))αH_{\alpha}\leq\mathrm{Aut}(\mathrm{sC}(r,s))_{\alpha}, our hypothesis implies that there exists an element gAut(sC(r,s))αg\in\mathrm{Aut}(\mathrm{sC}(r,s))_{\alpha} such that βg=α\beta^{g}=\alpha^{\prime} and γg=γ\gamma^{g}=\gamma. This yields a contradiction because δg\delta^{g} is ill-defined: in fact there is no vertex of sC(r,s)\mathrm{sC}(r,s) at distance 11 from both γg\gamma^{g} and δg\delta^{g}. Recall that, from Lemma 2.4, if r4r\neq 4, then H=Aut(C(r,s))H=\mathrm{Aut}(\mathrm{C}(r,s)), and so, by Corollary 2.10, Aut(sC(r,s))H\mathrm{Aut}(\mathrm{sC}(r,s))\leq H. On the other hand, if r=4r=4, observe that HH is vertex-transitive on sC(r,s)\mathrm{sC}(r,s) and Aut(sC(r,s))α=Hα\mathrm{Aut}(\mathrm{sC}(r,s))_{\alpha}=H_{\alpha}, hence the equality holds by Frattini’s argument. ∎

Lemma 2.13.

Let GG be a vertex-transitive subgroup of Aut(sC(r,s))\mathrm{Aut}(\mathrm{sC}(r,s)). Then either GG contains a semiregular element of order at least 66, or (sC(r,s),G)(\mathrm{sC}(r,s),G) is one of the examples in Table LABEL:table:table marked with the symbol \dagger.

Proof.

From Lemma 2.12, we have GH=Kρ,σG\leq H=K\langle\rho,\sigma\rangle. Observe that G/GKρ,σG/G\cap K\cong\langle\rho,\sigma\rangle, otherwise GG is not transitive on the vertices of the split graph sC(r,s)\mathrm{sC}(r,s). From this, it follows that G=Vρf,σgG=V\langle\rho f,\sigma g\rangle, for some f,gKf,g\in K, where V=GKV=G\cap K. Since ρ\rho has order rr, we get that

(ρf)r=ρfρ(ρfρ)f=ρfρ(ρ2ρ1fρ)f=ρfρρ2fρf=ρfρr1fρf=fρr1fρf\begin{split}(\rho f)^{r}&=\rho f\rho\ldots(\rho f\rho)f\\ &=\rho f\rho\ldots(\rho^{2}\rho^{-1}f\rho)f\\ &=\rho f\rho\ldots\rho^{2}f^{\rho}f\\ &=\rho f\rho^{r-1}\ldots f^{\rho}f\\ &=f^{\rho^{r-1}}\ldots f^{\rho}f\end{split}

is an element of VV. Since VV is an elementary abelian 22-group, the element ρf\rho f has order either rr or 2r2r. Recalling that VKV\leq K,

(ρf)r=i=0r1τiai(\rho f)^{r}=\prod_{i=0}^{r-1}\tau_{i}^{a_{i}}

with ai{0,1}a_{i}\in\{0,1\}. Furthermore,

(ρf)rρ=ρ(fρρfρfρ)=ρ(ffρfρr2fρr1)=ρ(fρr1fρf)=ρ(ρf)r\begin{split}(\rho f)^{r}\rho&=\rho(f\rho\ldots\rho f\rho f\rho)\\ &=\rho(ff^{\rho}\ldots f^{\rho^{r-2}}f^{\rho^{r-1}})\\ &=\rho(f^{\rho^{r-1}}\ldots f^{\rho}f)\\ &=\rho(\rho f)^{r}\end{split}

thus ρ\rho centralizes (ρf)r(\rho f)^{r}. From this, and from the fact that ρ\langle\rho\rangle acts transitively on {τ0,,τr1}\{\tau_{0},\ldots,\tau_{r-1}\}, we deduce that

(ρf)r=i=0r1τia(\rho f)^{r}=\prod_{i=0}^{r-1}\tau_{i}^{a}

where aa is either 0 or 11. If a=0a=0, then ρf\rho f is a semiregular element of order rr. In particular, either r6r\geq 6, or the number of vertices of sC(r,s)\mathrm{sC}(r,s) is r2sr2^{s}, which is bounded by 525=1605\cdot 2^{5}=160, and we finish by Remark 1.3. On the other hand, if a=1a=1, ρf\rho f has order 2r2r, and it corresponds to the so-called super flip of the Praeger-Xu graph C(r,s)\mathrm{C}(r,s). Since (ρf)r(\rho f)^{r} does not fix any vertex in C(r,s)\mathrm{C}(r,s), and since the vertex-stabilizers for a split graph has index 22 in the vertex-stabilizer of the starting graph, for any vertex αVsC(r,s)\alpha\in V\mathrm{sC}(r,s), we obtain that (ρf)rGα(\rho f)^{r}\notin G_{\alpha}. Hence ρf\rho f is semiregular of order 2r62r\geq 6. ∎

To conclude this section, we show a result mimicking Lemma 2.5 for cubic graphs.

Lemma 2.14.

Let Γ\Gamma be a connected cubic vertex-transitive graph, let GAut(Γ)G\leq\mathrm{Aut}(\Gamma) be a vertex-transitive group such that the action of GαG_{\alpha} on the neighbourhood of α\alpha is cyclic of order 22, and let NN be a minimal normal subgroup of GG. If NN is a 22-group and Γ/N\Gamma/N is a cycle of length at least 33, then Γ\Gamma is isomorphic either to a circular ladder, or to a Möbius ladder, or to sC(r,s)\mathrm{sC}(r,s), for some positive integers r3r\geq 3 and sr1s\leq r-1.

Proof.

We already know by Lemma 2.7 that both ladders admit a cyclic quotient graph, thus we can suppose that Γ\Gamma is not isomorphic to a circulant ladder or to a Möbius ladder. By hypothesis, we can apply the merging operator to (Γ,G)(\Gamma,G), obtaining the pair (Δ,𝒞)(\Delta,\mathcal{C}). Since we have excluded the possibility of Γ\Gamma being a ladder, by Remark 2.8, Δ\Delta is 44-valent, and the action of GG on Δ\Delta is faithful, arc-transitive and 𝒞\mathcal{C}-invariant. Since the action of NN cannot map edges in \mathcal{M} to edges in \mathcal{F}, the quotient graph Γ/N\Gamma/N retains a partition into two disjoint sets of edges, namely /N\mathcal{M}/N and /N\mathcal{F}/N. Moreover, since \mathcal{M} is a complete matching, each edge in /N\mathcal{M}/N is adjacent to precisely two edges in /N\mathcal{F}/N, and vice versa. This implies that the edges of Δ/N\Delta/N coincide with the elements of /N\mathcal{F}/N, two of which are adjacent if they share the same neighbour in /N\mathcal{M}/N. If r6r\geq 6, then Δ/N\Delta/N is a cycle of length r/2r/2. From Lemma 2.5, we deduce that Δ\Delta is isomorphic to C(r,s)\mathrm{C}(r,s), for some positive integers r3r\geq 3 and sr1s\leq r-1. Observe that, as 𝒞\mathcal{C} coincides with the connected components of \mathcal{F}, each cycle in 𝒞\mathcal{C} intersects precisely two KK-orbits. This implies that 𝒞=𝒮\mathcal{C}=\mathcal{S}, and so [PSV13, Theorem 12] yields that Γ\Gamma is isomorphic to

s(Δ,𝒞)=s(C(r,s),𝒮)=sC(r,s).\mathrm{s}(\Delta,\mathcal{C})=\mathrm{s}(\mathrm{C}(r,s),\mathcal{S})=\mathrm{sC}(r,s).\qed

Now, suppose that r=4r=4. In this case, we have that GG is a 22-group, hence |N|=2|N|=2 and |VΓ|=8|V\Gamma|=8, and so the only possibility is for Γ\Gamma to be a (cirular or Möbius) ladder, which we already excluded.

3. Proof of Theorem 1.2

We aim to prove Theorem 1.2 by contradiction. In this section we will assume the following.

Hypothesis 3.1.

Let Γ\Gamma be a connected cubic graph, and let GAut(Γ)G\leq\mathrm{Aut}(\Gamma) such that the pair (Γ,G)(\Gamma,G) is a minimal counterexample to Theorem 1.2, first with respect to the cardinality of VΓV\Gamma, and then to the order of GG. From Remark 1.3, we have |VΓ|>1 280|V\Gamma|>1\,280. Let α\alpha be an arbitrary vertex of Γ\Gamma. Let NN be a minimal normal subgroup of GG.

Since Γ\Gamma is connected, the stabilizer GαG_{\alpha} is a {2,3}\{2,3\}-group. More generally, if Δ\Delta is a connected dd-regular graph, then no prime bigger than dd divides the order of a vertex stabilizer (this follows from an elementary connectedness argument, see for instance [Spi14, Lemma 3.1] or [MS98, Lemma 3.2]). Moreover, GG must be a {2,3,5}\{2,3,5\}-group, otherwise we can find derangements of prime order at least 77, hence semiregular elements.

Since NN is a minimal normal subgroup of GG, NN is a direct product of simple groups, any two of which are isomorphic. Clearly, NN is a {2,3,5}\{2,3,5\}-group, and NαN_{\alpha} is a {2,3}\{2,3\}-group. Thus NN is a direct product SlS^{l}, for some positive integer ll and for some simple {2,3,5}\{2,3,5\}-group SS. Using the Classification of Finite Simple Groups, we see that the collection of simple {2,3,5}\{2,3,5\}-groups consists of

C2,C3,C5,Alt(5),Alt(6),PSp(4,3),C_{2},\,C_{3},\,C_{5},\,\mathrm{Alt}(5),\,\mathrm{Alt}(6),\,\mathrm{PSp}(4,3),

see for instance [LW74].

Lemma 3.2.

Under Hypothesis 3.1, if NαN_{\alpha} is a 22-group, then NN is an elementary abelian pp-group, for some prime p{2,3,5}p\in\{2,3,5\}.

Proof.

If NN is abelian, then there is nothing to prove. Thus, suppose that N=SlN=S^{l}, where S{Alt(5),Alt(6),PSp(4,3)}S\in\{\mathrm{Alt}(5),\mathrm{Alt}(6),\mathrm{PSp}(4,3)\} and l1l\geq 1.

Assume l2l\geq 2. Let SS and TT be two distinct direct factors of NN. Then SαS_{\alpha} and TαT_{\alpha} are 22-groups, because so is NαN_{\alpha}. Thus, by Lemma 2.1, all the 33- and 55-elements of SS and TT are semiregular. Applying Lemma 2.2, we obtain that S×TS\times T, contains a semiregular element of order 1515. Thus GG contains a semiregular element of order exceeding 66, contradicting Hypothesis 3.1.

Assume l=1l=1. If N=PSp(4,3)N=\mathrm{PSp}(4,3), then Lemma 2.1 implies that the 33-elements in NN are semiregular. As PSp(4,3)\mathrm{PSp}(4,3) contains elements of order 99, GG contains a semiregular element of order 99, contradicting Hypothesis 3.1. Thus, NN is either Alt(5)\mathrm{Alt}(5) or Alt(6)\mathrm{Alt}(6).

We claim that GG is almost simple, that is, NN is the unique minimal normal subgroup of GG. Aiming for a contradiction, let MM be a minimal normal subgroup of GG distinct from NN. If Γ/M\Gamma/M is a cubic graph, then Mα=1M_{\alpha}=1, and hence each element of MM is semiregular. Since [N,M]=1[N,M]=1, by Lemma 2.2, GG contains a semiregular element of order at least 1010, against Hypothesis 3.1. On the other hand, suppose that Γ/M\Gamma/M is not cubic. Regardless of the valency of Γ/M\Gamma/M, the group that GG induces in its action on the vertices of Γ/M\Gamma/M is a subgroup of a dihedral group, hence it is a soluble group. In particular, as NN is a non-abelian simple group, NN acts trivially on the vertices of Γ/M\Gamma/M. This means that NN fixes setwise each MM-orbit. If MM is abelian, then MM acts regularly on each of its orbits. However, as NN commutes with MM and fixes each MM-orbit, this contradicts the fact that NN is non-abelian.111Recall that, if XSym(Ω)X\leq\mathrm{Sym}(\Omega) is an abelian group and XX acts regularly on Ω\Omega, then X=𝐂Sym(Ω)(X)X={\bf C}_{\mathrm{Sym}(\Omega)}(X). Therefore, MM is not abelian. In particular, there is a prime p5p\geq 5 that divides the order of MM, and the elements of MM of order pp are semiregular. As before, applying Lemma 2.2, we get that NMNM contains a semiregular element of order 3p3p, a contradiction. We conclude that NN is the unique minimal normal subgroup of GG.

Notice that Alt(5)GSym(5)\mathrm{Alt}(5)\leq G\leq\mathrm{Sym}(5) or Alt(6)GAut(Alt(6))\mathrm{Alt}(6)\leq G\leq\mathrm{Aut}(\mathrm{Alt}(6)). A computer computation in each of these cases shows that, if GAut(Γ)G\leq\mathrm{Aut}(\Gamma) has no semiregular elements of order at least 66, then |VΓ|{30,60,90,180,360}|V\Gamma|\in\{30,60,90,180,360\}, which contradicts Hypothesis 3.1. ∎

From here on, we divide the proof in five cases:

  • Gα=1G_{\alpha}=1;

  • Gα1G_{\alpha}\neq 1 and NN is transitive on VΓV\Gamma;

  • Gα1G_{\alpha}\neq 1 and NN has two orbits on VΓV\Gamma;

  • Gα1G_{\alpha}\neq 1 and Γ/N\Gamma/N is a cycle of length at least 33;

  • Gα1G_{\alpha}\neq 1 and Γ/N\Gamma/N is a cubic graph.

3.1. Gα=1G_{\alpha}=1

In this case Γ\Gamma is a Cayley graph over GG. This means that there exists an inverse-closed subset II of GG with ΓCay(G,I)\Gamma\cong\mathrm{Cay}(G,I). We recall that Cay(G,I)\mathrm{Cay}(G,I) is the graph having vertex set GG where two vertices xx and yy are declared to be adjacent if and only if yx1Iyx^{-1}\in I. Since Γ\Gamma has valency 33, we have |I|=3|I|=3. Moreover, since Γ\Gamma is connected, we have G=IG=\langle I\rangle. In particular, GG is generated by at most 33 elements. More precisely, either II consists of three involutions or II consists of an involution and an element of order greater than 22 together with its inverse.

In what follows we say that a finite group XX satisfies 𝒫\mathcal{P} if XX is generated by either three involutions, or by an involution and by an element of order greater than 22. In particular, GG satisfies 𝒫\mathcal{P}.

Since each element of GG is semiregular and since GG has no semiregular elements of order at least 66, we deduce that each element of GG has order at most 55. As customary, we let

ω(G):={o(g)gG}\omega(G):=\{o(g)\mid g\in G\}

be the spectrum of GG. Observe that

{1,2}ω(G){1,2,3,4,5}.\{1,2\}\subseteq\omega(G)\subseteq\{1,2,3,4,5\}.

Since GG is generated by at most 33 elements, we deduce from Zelmanov’s solution of the restricted Burnside problem that |G||G| is bounded above by an absolute constant. We divide the proof depending on ω(G)\omega(G).

Assume ω(G)={1,2}\omega(G)=\{1,2\}. In this case, GG is elementary abelian and, since GG is generated by at most 33 elements, we deduce |G|8|G|\leq 8, which contradicts Hypothesis 3.1.

Assume ω(G)={1,2,4}\omega(G)=\{1,2,4\}. Here, either GG is generated by an element of order 22 and an element of order 44, or GG is generated by three involutions. We resolve these two cases with a computer computation. Suppose first that GG is generated by an involution and by an element of order 44. We have constructed the free group F:=x,yF:=\langle x,y\rangle and we have constructed the set WW of words in x,yx,y of length at most 66. Then, we have constructed the finitely presented group F¯:=F|x2,{w4:wW}\bar{F}:=\langle F|x^{2},\{w^{4}:w\in W\}\rangle. We use the “bar” notation for the projection of FF onto F¯\bar{F}. Now, x¯\bar{x} has order 22 and y¯\bar{y} has order 44. Furthermore, each element of F¯\bar{F} that can be written as a word in x¯\bar{x} and y¯\bar{y} of length at most 66 has order at most 44. (The number 66 was chosen arbitrarily but large enough to guarantee to get an upper limit on the cardinality of GG.) A computer computation shows that F¯\bar{F} has order 6464 and exponent 44. This proves that the largest group of exponent 44 and generated by an involution and by an element of order 44 has order 6464. Now, GG is a quotient of F¯\bar{F} and hence |G||F¯|64|G|\leq|\bar{F}|\leq 64, which contradicts Hypothesis 3.1. Next, suppose that GG is generated by three involutions. The argument here is very similar. We have considered the free group F=x,y,z,F=\langle x,y,z\rangle, and we have considered the set WW of words in x,y,zx,y,z of length at most 66. We have verified that F|x2,y2,z2,{w4:wW}\langle F|x^{2},y^{2},z^{2},\{w^{4}:w\in W\}\rangle has order 10241024 and exponent 44. This shows that |G|1 024|G|\leq 1\,024, which contradicts Hypothesis 3.1.

Assume ω(G)={1,2,3}\omega(G)=\{1,2,3\}. The groups having spectrum {1,2,3}\{1,2,3\} are classified in [Neu37]. Routine computations in the list of groups XX classified in  [Neu37, Theorem] show that, if XX satisfies 𝒫,\mathcal{P}, then |X|18|X|\leq 18, which contradicts Hypothesis 3.1.

Assume ω(G)={1,2,5}\omega(G)=\{1,2,5\}. The groups having spectrum {1,2,5}\{1,2,5\} are classified in [New79]. As above, since GG satisfies 𝒫\mathcal{P}, we deduce from a case-by-case analysis in the groups appearing in [New79] that |G|80|G|\leq 80, which contradicts Hypothesis 3.1.

Assume ω(G)={1,2,3,4}\omega(G)=\{1,2,3,4\}. The groups having spectrum {1,2,3,4}\{1,2,3,4\} are classified in [BS91]. As above, since GG satisfies 𝒫\mathcal{P}, we deduce from a case-by-case analysis in the groups appearing in [BS91, Theorem] that |G|96|G|\leq 96, which contradicts Hypothesis 3.1.

Assume ω(G)={1,2,4,5}\omega(G)=\{1,2,4,5\}. The groups having spectrum {1,2,4,5}\{1,2,4,5\} are classified in [GM99]. This case is sligthly more involved and hence we do give more details. We have three cases to consider:

  1. (1)

    G=TDG=T\rtimes D where TT is a non-trivial elementary abelian normal 22-subgroup and DD is a non-abelian group of order 1010,

  2. (2)

    G=FTG=F\rtimes T where FF is an elementary abelian normal 55-subgroup and TT is isomorphic to a subgroup of a quaternion group of order 88,

  3. (3)

    GG contains a normal 22-subgroup TT which is nilpotent of class at most 66 such that G/TG/T is a 55-group.

Suppose that (1) holds. Clearly, DD is the dihedral group of order 1010 and TT is a module for DD over the field 𝔽2\mathbb{F}_{2} of cardinality 22. The dihedral group DD has two irreducible modules over 𝔽2\mathbb{F}_{2} up to equivalence: the trivial module and a 44-dimensional module WW. Since GG has no elements of order 1010, we deduce VWV\cong W^{\ell}, for some 1\ell\geq 1. We have verified with a computer computation that W3DW^{3}\rtimes D does not satisfy 𝒫\mathcal{P} and hence GWDG\cong W^{\ell}\rtimes D with 2\ell\leq 2. We deduce that |G|=|VΓ|{1016,10162}={160,2 560}|G|=|V\Gamma|\in\{10\cdot 16,10\cdot 16^{2}\}=\{160,2\,560\}. From Hypothesis 3.1, we have |VΓ|>1 280|V\Gamma|>1\,280 and hence GW2DG\cong W^{2}\rtimes D. We have constructed all connected cubic Cayley graphs over W2DW^{2}\rtimes D and we have found only one (up to isomorphism), therefore we obtain the example in Table LABEL:table:table.

Suppose that (2) holds. Since GG satisfies 𝒫\mathcal{P}, while the quaternion group of order 88 does not, we deduce that TT is cyclic of order 44. Thus G=FxG=F\rtimes\langle x\rangle, for some xx having order 44. As GG satisfies 𝒫\mathcal{P}, this means that G=x,yG=\langle x,y\rangle, for some involution yy. Clearly, y=fx2y=fx^{2} for some fFf\in F. As G=x,y=x,fx2=x,fG=\langle x,y\rangle=\langle x,fx^{2}\rangle=\langle x,f\rangle, we have F=f,fx,fx2,fx3F=\langle f,f^{x},f^{x^{2}},f^{x^{3}}\rangle. Since y=fx2y=fx^{2} has order 22 and xx has order 44, we deduce

1=y2=fx2fx2=ffx2,1=y^{2}=fx^{2}fx^{2}=ff^{x^{2}},

that is, fx2=f1f^{x^{2}}=f^{-1}. Now, F=f,fx,fx2,fx3=f,fx,f1,(fx)1=f,fxF=\langle f,f^{x},f^{x^{2}},f^{x^{3}}\rangle=\langle f,f^{x},f^{-1},(f^{x})^{-1}\rangle=\langle f,f^{x}\rangle. Thus |F|25|F|\leq 25 and hence |G|100|G|\leq 100, which contradicts Hypothesis 3.1.

Suppose that (3) holds. Since GG satisfies 𝒫\mathcal{P}, we deduce that G/TG/T is cyclic of order 55. Thus G=TxG=T\rtimes\langle x\rangle, for some xx having order 55. This means that G=x,yG=\langle x,y\rangle, for some involution yy. Clearly, yTy\in T. From Hypothesis 3.1, we have |G|=|VΓ|>1 280|G|=|V\Gamma|>1\,280. Let NN be a minimal normal subgroup of GG. We have NTN\leq T and NN is an irreducible 𝔽2x\mathbb{F}_{2}\langle x\rangle-module. The cyclic group of order 55 has two irreducible modules over 𝔽2\mathbb{F}_{2} up to equivalence: the trivial module and a 44-dimensional module. Since GG has no elements of order 1010, xx does not centralize NN and hence NN is the irreducible 44-dimensional module for the cyclic group of order 55. In particular, |N|=24|N|=2^{4}. Consider G¯:=G/N\bar{G}:=G/N. Now,

{1,2,5}ω(G¯)ω(G)={1,2,4,5}.\{1,2,5\}\subseteq\omega(\bar{G})\subseteq\omega(G)=\{1,2,4,5\}.

Assume ω(G¯)={1,2,5}\omega(\bar{G})=\{1,2,5\}. From the discussion above (regarding the finite groups having spectrum {1,2,5}\{1,2,5\} and satisfying 𝒫\mathcal{P}), we have |G¯|80|\bar{G}|\leq 80 and hence |G|=|G:N||N|8016=1 280|G|=|G:N||N|\leq 80\cdot 16=1\,280, which is a contradiction. Therefore, ω(G¯)={1,2,4,5}\omega(\bar{G})=\{1,2,4,5\}. Since (Γ,G)(\Gamma,G) was chosen minimal in Hypothesis 3.1, we have |G¯|1 280|\bar{G}|\leq 1\,280. Therefore (Γ/N,G¯)(\Gamma/N,\bar{G}) appears in Table LABEL:table:table. An inspection on the groups appearing in this table shows that there is only one group having spectrum {1,2,4,5}\{1,2,4,5\} and is the group of order 1 2801\,280. Thus we know precisely G¯\bar{G}. Now, the group GG is an extension of G¯\bar{G} by NN and hence it can be computed with the cohomology package in the computer algebra system magma. We have computed all the extensions EE of G¯\bar{G} via NN and we have verified that none of the extensions EE has the property that ω(E)={1,2,4,5}\omega(E)=\{1,2,4,5\} and with EE satisfying 𝒫\mathcal{P}.

Assume ω(G)={1,2,3,5}\omega(G)=\{1,2,3,5\}. The groups having spectrum {1,2,3,5}\{1,2,3,5\} are classified in [MZ99]. We deduce from [MZ99] that GA5G\cong A_{5}, which contradicts Hypothesis 3.1.

Assume ω(G)={1,2,3,4,5}\omega(G)=\{1,2,3,4,5\}. The groups having spectrum {1,2,3,4,5}\{1,2,3,4,5\} are classified in [BS91]. We deduce from [BS91, Theorem] that either GA6G\cong A_{6} or GVA5G\cong V^{\ell}\rtimes A_{5} where VV is a 44-dimensional natural module over the finite field of size 22 for A5SL2(4)A_{5}\cong\mathrm{SL}_{2}(4) and 1\ell\geq 1. The group V2A5V^{2}\rtimes A_{5} does not satisfy 𝒫\mathcal{P} (this can be verified with a computer computation). Therefore, either GA6G\cong A_{6} or GVA5G\cong V\rtimes A_{5}. Thus |G|=|VΓ|960|G|=|V\Gamma|\leq 960, which contradicts Hypothesis 3.1.

3.2. Gα1G_{\alpha}\neq 1 and NN is transitive on VΓV\Gamma.

By Hypothesis 3.1, (Γ,G)(\Gamma,G) is a minimal counterexample. This minimality and the fact that NN is transitive on VΓV\Gamma imply that G=NG=N. As NN is a minimal normal subgroup of GG, GG is simple. Thus G{Alt(5),Alt(6),PSp(4,3)}G\in\{\mathrm{Alt}(5),\mathrm{Alt}(6),\mathrm{PSp}(4,3)\}. A computer computation in each of these cases shows that, if GAut(Γ)G\leq\mathrm{Aut}(\Gamma) has no semiregular elements of order at least 66, then |VΓ|{10,20,30,60,90,180,360}|V\Gamma|\in\{10,20,30,60,90,180,360\}, which contradicts Hypothesis 3.1.

3.3. Gα1G_{\alpha}\neq 1 and NN has two orbits on VΓV\Gamma.

Suppose NN is abelian. By [PS21, Lemma 1.15], either Γ\Gamma is complete bipartite, or Γ\Gamma is a bi-Cayley graph over NN and the minimal number of generators of NN is at most 44. (Here, it is not really relevent to introduce the definition of bi-Cayley graph, however, what is really relevant is the fact that NN is generated by at most 44 elements.) Recalling that NN is a {2,3,5}\{2,3,5\}-group, it follows that |VΓ|=2|N|254=1 250|V\Gamma|=2|N|\leq 2\cdot 5^{4}=1\,250, and the equality is realized for N=C54N=C_{5}^{4}. In particular, this contradicts Hypothesis 3.1.

Suppose NN is not abelian. By Lemma 3.2, 33 divides the order of NαN_{\alpha}. A fortiori, 33 divides the order of GαG_{\alpha}, hence GG acts arc-transitively on Γ\Gamma. We can extract information on the local action of GG by consulting the amalgams in [DM80, Section 4]. In particular, with a direct inspection (on a case-by-case basis) on these amalgams, it can be verified that, for any edge {α,β}\{\alpha,\beta\} of Γ\Gamma, GG contains an element yy that swaps α\alpha and β\beta and its order is either 22 or 44. As α\alpha and β\beta belong to distinct NN-orbits, yy maps αN\alpha^{N} to βN\beta^{N}. Moreover, as NN has two orbits on VΓV\Gamma, the subgroup NyN\langle y\rangle is vertex-transitive on Γ\Gamma. Therefore, by minimality of GG, we have G=NyG=N\langle y\rangle.

Assume o(y)=2o(y)=2. Thus |G:N|=2|G:N|=2. As N=SlN=S^{l} is a minimal normal subgroup of GG, l{1,2}l\in\{1,2\}. If l=1l=1, then GG is an almost simple group whose socle is either Alt(5)\mathrm{Alt}(5), Alt(6)\mathrm{Alt}(6) or PSp(4,3)\mathrm{PSp(4,3)}. A computer computation shows that (Γ,G)(\Gamma,G) satisfies Theorem 1.2, a contradiction. If l=2l=2, then y\langle y\rangle permutes transitively the two simple direct factors of NN. Let sNs\in N be a 55-element in a simple direct factor of NN, and notice that t:=syt:=s^{y} is a 55-element in the other simple direct factor of NN. Thus [s,t]=1[s,t]=1. We claim that ysys is a semiregular element of order 1010. We get

(ys)2=ysys=tsN,(ys)^{2}=ysys=ts\in N,
(ys)5=ysysysysys=ys(ts)2yN.(ys)^{5}=ysysysysys=ys(ts)^{2}\in yN.

We have that (ys)2(ys)^{2} is a 55-element in NN, thus semiregular, and that (ys)5(ys)^{5} has order 22 and, being an element of yN=NyyN=Ny, it has no fixed points, hence it is semiregular. Therefore ysys is a semiregular element of order 1010, contradicting Hypothesis 3.1.

Assume o(y)=4o(y)=4. As |G:N|=4|G:N|=4 and NN is a minimal normal subgroup of GG, l{1,2,4}l\in\{1,2,4\}. Observe that a Sylow 33-subgroup of GαG_{\alpha} has order 33, because Γ\Gamma is cubic and GG is arc-transitive. Let xGαx\in G_{\alpha} be an element of order 33. As |G:N|=4|G:N|=4, we have xNGα=NαSlx\in N\cap G_{\alpha}=N_{\alpha}\leq S^{l}. In particular, we may write x=(s1,,sl)x=(s_{1},\ldots,s_{l}), with siSs_{i}\in S. Let κ\kappa be the number of coordinates of xx different from 11, we call κ\kappa the type of xx. Since x\langle x\rangle is a Sylow 33-subgroup of GαG_{\alpha}, from Sylow’s theorem, we deduce that each element of order 33 in GG fixing some vertex of Γ\Gamma has type κ\kappa. Let sSs\in S be an element of order 33 and let tSt\in S be an element of order 55. Suppose l=4l=4. If κ1\kappa\neq 1, then g=(s,t,1,1)g=(s,t,1,1) has order 1515 and is semiregular because g5=(s5,1,1,1)g^{5}=(s^{5},1,1,1) has order 33 but it is not of type κ\kappa. Similarly, if l=4l=4 and k=1k=1, then g=(s,s,t,1)g=(s,s,t,1) has order 1515 and is semiregular. Analogously, when l=2l=2, if κ1\kappa\neq 1, then g=(s,t)g=(s,t) has order 1515 and is semiregular. When l=2l=2, κ=1\kappa=1 and S=PSp(4,3)S=\mathrm{PSp}(4,3), the group SS contains an element rr having order 99 and hence g=(r,r)g=(r,r) is a semiregular element having order 99. Summing up, from these reductions, we may suppose that either l=1l=1, or l=2l=2 and S{Alt(5),Alt(6)}S\in\{\mathrm{Alt}(5),\mathrm{Alt}(6)\}. These cases can be dealt with a computer computation: indeed, the invaluable help of a computer shows that no counterexample to Theorem 1.2 arises.

3.4. Gα1G_{\alpha}\neq 1 and Γ/N\Gamma/N is a cycle of length r3r\geq 3.

The full automorphism group of Γ/N\Gamma/N is the dihedral group of order 2r2r. Let KK be the kernel of the action of GG on the NN-orbits. The quotient G/KG/K acts faithfully on Γ/N\Gamma/N, that is, it is a transitive subgroup of the dihedral group of order 2r2r.

We claim that

(3.1) G/K is regular in its action on the vertices of Γ/N.G/K\hbox{ is regular in its action on the vertices of }\Gamma/N.

Assume G/KG/K acts on the vertices of Γ/N\Gamma/N transitively but not regularly. In particular, G/KG/K is isomorphic to the dihedral group of order 2r2r. Thus GG has an index 22 subgroup MM such that MM is vertex-transitive and M/KM/K is isomorphic to the cyclic group of order rr. By minimality of GG, we have G=MG=M, which goes against the choice of MM. Hence G/KG/K is regular. In particular, either G/KG/K is isomorphic to the cyclic group of order rr, or rr is even and GG is isomorphic to the dihedral group of order rr. Later in this proof we resolve this ambiguity and we prove that rr is even and G/KG/K is dihedral of order rr, see (3.5).

As G/KG/K acts regularly on the vertices of Γ/N\Gamma/N, we have

1G/K=(GK)αN=GαKK.1_{G/K}=\left(\frac{G}{K}\right)_{\alpha^{N}}=\frac{G_{\alpha}K}{K}.

Therefore

(3.2) Kα=KGα=Gα.K_{\alpha}=K\cap G_{\alpha}=G_{\alpha}.

Assume GG is arc-transitive. Let β\beta be a neighbour of α\alpha and observe that αNβN\alpha^{N}\neq\beta^{N}. Since Γ\Gamma is connected, we have

G=Gα,G{α,β}=Kα,G{α,β}K,G{α,β}=KG{α,β},G=\langle G_{\alpha},G_{\{\alpha,\beta\}}\rangle=\langle K_{\alpha},G_{\{\alpha,\beta\}}\rangle\leq\langle K,G_{\{\alpha,\beta\}}\rangle=KG_{\{\alpha,\beta\}},

and hence G=KG{α,β}G=KG_{\{\alpha,\beta\}}. Recalling that KK fixes all the NN-orbits,

|G:K|=|KG{α,β}:K|=|G{α,β}:K{α,β}|=|G{α,β}:Gαβ|=2.|G:K|=|KG_{\{\alpha,\beta\}}:K|=|G_{\{\alpha,\beta\}}:K_{\{\alpha,\beta\}}|=|G_{\{\alpha,\beta\}}:G_{\alpha\beta}|=2.

Thus G/KC2G/K\cong C_{2} and r=2r=2, which is a contradiction. Therefore

GG is not arc-transitive.

This implies that GαG_{\alpha} does not act transitively on the neighbourhood of α\alpha, hence GαG_{\alpha} is a 22-group. By (3.2), we deduce Gα=KαG_{\alpha}=K_{\alpha} is a 22-group. Actually, Lemma 2.3 shows that

(3.3) Gα=Kα is an elementary abelian 2-group.G_{\alpha}=K_{\alpha}\hbox{ is an elementary abelian }2\hbox{-group}.

If NN is an elementary abelian 22-group, then, by Lemma 2.14, Γ\Gamma is either a circular ladder, or a Möbius ladder, or a split Praeger-Xu graph sC(r/2,s)\mathrm{sC}(r/2,s). Now, in the former cases, the proof follows from Lemma 2.6, while, in the latter one, we conclude by Lemma 2.13. In particular, for the rest of the proof we may suppose that NN is not an elementary abelian 22-group.

For any minimal normal subgroup MM of GG, Mα=MGαM_{\alpha}=M\cap G_{\alpha} is also a 22-group. Thus, in view of Lemma 3.2, MM is an elementary abelian pp-group, for some p{2,3,5}p\in\{2,3,5\}. This is true, in particular, for NN. Let MM be a minimal normal subgroup distinct from NN. Since [N,M]=1[N,M]=1, Lemma 2.2 gives a contradiction unless NN and MM are both pp-groups for the same prime pp. Thus,

(3.4) the socle of G is an elementary abelian p-group, for some p{3,5}.\textrm{the socle of }G\textrm{ is an elementary abelian }p\textrm{-group, for some }p\in\{3,5\}.

Before going any further, we need some extra information on the local action of GG on Γ\Gamma. Since GαG_{\alpha} is a non-identity 22-group, there exists a unique vertex αVΓ\alpha^{\prime}\in V\Gamma adjacent to α\alpha that is fixed by the action of GαG_{\alpha}. It follows that {α,α}\{\alpha,\alpha^{\prime}\} is a block of imprimitivity for the action of GG on the vertices. Hence,

GαG{α,α}and|G{α,α}:Gα|=2.G_{\alpha}\leq G_{\{\alpha,\alpha^{\prime}\}}\quad\textup{and}\quad|G_{\{\alpha,\alpha^{\prime}\}}:G_{\alpha}|=2.

We obtain that, for any βVΓ\beta\in V\Gamma, neighbour of α\alpha distinct from α\alpha^{\prime},

|G{α,α}:Gαβ|=4and|G{α,β}:Gαβ|=2.|G_{\{\alpha,\alpha^{\prime}\}}:G_{\alpha\beta}|=4\quad\textup{and}\quad|G_{\{\alpha,\beta\}}:G_{\alpha\beta}|=2.

Let {α,β,γ}\{\alpha^{\prime},\beta,\gamma\} be the neighbourhood of α\alpha.

Assume G/KG/K is cyclic of order rr. As Γ/N\Gamma/N is a cycle of length rr, this means that G/KG/K acts transitively on the vertices and on the edges of Γ/N\Gamma/N. Now, β\beta and γ\gamma are in the same KK-orbit because Kα=GαK_{\alpha}=G_{\alpha} and GαG_{\alpha} acts transitively on {β,γ}\{\beta,\gamma\}. In particular, each element in αN\alpha^{N} has two neighbours in βN\beta^{N}. As G/KG/K is transitive on edges, we reach a contradiction because each element in αN\alpha^{N} would have two neighbours in αN{\alpha^{\prime}}^{N}, contradicting the fact that α\alpha has valency 33. Thus

(3.5) r is even and G/K is dihedral of order r.r\hbox{ is even and }G/K\textrm{ is dihedral of order }r.

Recall that NN is an elementary abelian pp-group with p{3,5}p\in\{3,5\}. Thus NN is semiregular. We consider 𝐂K(N)\mathbf{C}_{K}(N). Since N𝐂K(N)N\leq{\bf C}_{K}(N) and since K=KαNK=K_{\alpha}N, we deduce 𝐂K(N)=N×Q{\bf C}_{K}(N)=N\times Q, for some subgroup QQ of KαK_{\alpha}. As KαK_{\alpha} is a 22-group, so is QQ. Therefore, QQ is characteristic in N×Q=𝐂K(N)N\times Q={\bf C}_{K}(N) and hence QGQ\unlhd G. Since GαG_{\alpha} is a core-free subgroup of GG, we get Q=1Q=1 and 𝐂K(N)=N{\bf C}_{K}(N)=N.

Since NN is a minimal normal subgroup of GG, GG acts irreducibly by conjugation on it, that is, NN is an irreducible 𝔽pG\mathbb{F}_{p}G-module. As KGK\unlhd G, by Clifford’s Theorem, NN is a completely reducible 𝔽pK\mathbb{F}_{p}K-module. As K=NGαK=NG_{\alpha} and NN is abelian, NN is a completely reducible 𝔽pGα\mathbb{F}_{p}G_{\alpha}-module. As GαG_{\alpha} is abelian, by Schur’s Lemma, GαG_{\alpha} induces on each irreducible 𝔽pGα\mathbb{F}_{p}G_{\alpha}-submodule a cyclic group action. However, since GαG_{\alpha} has exponent 22, we deduce that each irreducible 𝔽pGα\mathbb{F}_{p}G_{\alpha}-submodule has dimension 11 and GαG_{\alpha} induces on each irreducible 𝔽pGα\mathbb{F}_{p}G_{\alpha}-submodule the scalars ±1\pm 1. Therefore, GαG_{\alpha} acts on NN by conjugation as a group of diagonal matrices having eigenvalues in {±1}\{\pm 1\}. In other words, there exists a basis (n1,,ne)(n_{1},\ldots,n_{e}) of NN as a vector space over 𝔽p\mathbb{F}_{p} such that,

(3.6) for each gGα and for each ni, we have nig{ni,ni1}.\hbox{for each }g\in G_{\alpha}\hbox{ and for each }n_{i},\hbox{ we have }n_{i}^{g}\in\{n_{i},n_{i}^{-1}\}.

Furthermore, the action of GG by conjugation on NN preserves the direct product decomposition N=n1××neN=\langle n_{1}\rangle\times\cdots\times\langle n_{e}\rangle.

We claim that

(3.7) 𝐂G{α,β}(N)\displaystyle{\bf C}_{G_{\{\alpha,\beta\}}}(N) =1,\displaystyle=1,
𝐂G{α,α}(N)\displaystyle{\bf C}_{G_{\{\alpha,\alpha^{\prime}\}}}(N) =1.\displaystyle=1.

In other words, G{α,β}G_{\{\alpha,\beta\}} and G{α,α}G_{\{\alpha,\alpha^{\prime}\}} both act faithfully by conjugation on NN. Let γ{α,β}\gamma\in\{\alpha^{\prime},\beta\} and suppose, arguing by contradiction, that 𝐂G{α,γ}(N)1{\bf C}_{G_{\{\alpha,\gamma\}}}(N)\neq 1. Since 𝐂K(N)=1{\bf C}_{K}(N)=1 and |G{α,γ}:KG{α,γ}|=2|G_{\{\alpha,\gamma\}}:K\cap G_{\{\alpha,\gamma\}}|=2, we deduce 𝐂G{α,γ}(N)=x{\bf C}_{G_{\{\alpha,\gamma\}}}(N)=\langle x\rangle, where xx is an involution. Since xKx\notin K, xx acts semiregularly on Γ/N\Gamma/N and hence xx acts semiregularly on Γ\Gamma. From this and from the fact that xx centralizes NN, we deduce that GG contains semiregular elements of order 2p62p\geq 6, which contradicts Hypothesis 3.1. Thus (3.7) is proven.

Observe that (3.7) implies that an element of G{α,α}G_{\{\alpha,\alpha^{\prime}\}} or of G{α,β}G_{\{\alpha,\beta\}} is the identity if and only it its action on NN by conjugation is trivial.

We show that

(3.8) G{α,β}Gαβ contains an involution.G_{\{\alpha,\beta\}}\setminus G_{\alpha\beta}\hbox{ contains an involution.}

Let HH be the permutation group induced by G{α,α}G_{\{\alpha,\alpha^{\prime}\}} in its action on the four right cosets of GαβG_{\alpha\beta} in G{α,α}G_{\{\alpha,\alpha^{\prime}\}}. Since HH is a 22-group, HH is isomorphic to either C4C_{4}, or C2×C2C_{2}\times C_{2}, or to the dihedral group of order 88. In the first two cases, GαβG_{\alpha\beta} is a normal subgroup of both G{α,α}G_{\{\alpha,\alpha^{\prime}\}} and G{α,β}G_{\{\alpha,\beta\}}. As GαβG_{\alpha\beta} is core-free in GG and

G=G{α,α},G{α,β},G=\langle G_{\{\alpha,\alpha^{\prime}\}},G_{\{\alpha,\beta\}}\rangle,

we have that Gαβ=1G_{\alpha\beta}=1. In particular, G{α,β}G_{\{\alpha,\beta\}} is cyclic of order 22, hence it contains an involution and (3.8) follows in this case.

In the latter case, using the notation and the terminology in [Djo80], we have that the triple (G{α,α},Gαβ,G{αβ})(G_{\{\alpha,\alpha^{\prime}\}},G_{\alpha\beta},G_{\{\alpha\,\beta\}}) is a locally dihedral faithful group amalgam of type (4,2)(4,2) and GG is one of its realizations. Indeed, from the classification in [Djo80], we see that either G{α,α}GαG_{\{\alpha,\alpha^{\prime}\}}\setminus G_{\alpha} or G{α,β}GαβG_{\{\alpha,\beta\}}\setminus G_{\alpha\beta} contains an involution. If G{α,β}GαβG_{\{\alpha,\beta\}}\setminus G_{\alpha\beta} contains an involution, then (3.8) holds true also in this case. Therefore we suppose τ1G{α,α}Gα\tau_{1}\in G_{\{\alpha,\alpha^{\prime}\}}\setminus G_{\alpha} is an involution. We investigate the action by conjugation of τ1\tau_{1} on NN. By (3.1), τ1\tau_{1} is a semiregular automorphism of Γ/K\Gamma/K, because τ1K\tau_{1}\notin K. Therefore, τ1\tau_{1} is a semiregular automorphism of Γ\Gamma. Since no semiregular involution commutes with a non-identity element of NN, τ1\tau_{1} acts by conjugation on NN without fixed points, that is, for any nNn\in N, nτ1=n1n^{\tau_{1}}=n^{-1}. It follows from (3.6) that τ1\tau_{1} commutes with GαG_{\alpha} and hence G{α,α}=Gα,τ1G_{\{\alpha,\alpha^{\prime}\}}=\langle G_{\alpha},\tau_{1}\rangle is an elementary abelian 22-group. Now, as GαβG_{{\alpha\beta}} is normal in both G{α,α}G_{\{\alpha,\alpha^{\prime}\}} and G{α,β}G_{\{\alpha,\beta\}}, we can conclude, as before, that G{α,β}G_{\{\alpha,\beta\}} is cyclic of order 22, hence it contains an involution. Therefore, in any case, (3.8) holds true.

Let ee be the positive integer such that N=CpeN=C_{p}^{e}. We aim to show that

(3.9) e{1,2}.e\in\{1,2\}.

Let τ2G{α,β}Gαβ\tau_{2}\in G_{\{\alpha,\beta\}}\setminus G_{\alpha\beta} be an involution: the existence of τ2\tau_{2} is guaranteed by (3.8). Now, we look at the action by conjugation of τ2\tau_{2} on NN. Observe τ2K\tau_{2}\notin K and hence τ2\tau_{2} is a semiregular automorphism of Γ\Gamma. Therefore, arguing as in the previous paragraph (with the involution τ1\tau_{1} replaced by τ2\tau_{2}), we deduce that nτ2=n1n^{\tau_{2}}=n^{-1} for every nNn\in N. Let L:=τ2ggGL:=\langle\tau_{2}^{g}\mid g\in G\rangle. Since G/KG/K is a dihedral group and τ2\tau_{2} is an involution, we deduce that |G/K:LK/K|2|G/K:LK/K|\leq 2, that is, |G:LK|2|G:LK|\leq 2. Observe now that, for any nNn\in N, nτ2g=n1n^{\tau_{2}^{g}}=n^{-1}. Therefore, the group induced by the action by conjugation of LL on NN has order 22. This and (3.6) shows that the subgroup LKLK of GG preserves the direct sum decomposition N=n1××neN=\langle n_{1}\rangle\times\cdots\times\langle n_{e}\rangle. However, since GG acts irreducibly on NN and since |G:LK|2|G:LK|\leq 2, we finally obtain e2e\leq 2, as claimed in (3.9). Observe that from this it follows that |N|=pe{3,9,5,25}|N|=p^{e}\in\{3,9,5,25\}.

We are now ready to conclude this case. Observe that GαG_{\alpha} contains an element xx with nx=n1n^{x}=n^{-1} for every nNn\in N. This is immediate from (3.6) when e=1e=1, or when e=2e=2 and |Gα|=4|G_{\alpha}|=4. When e=2e=2 and |Gα|<4|G_{\alpha}|<4, we have |Gα|=2|G_{\alpha}|=2 and hence the non-identity element of GαG_{\alpha} acts by conjugation on NN inverting each of its elements.

Now, xx and τ2\tau_{2} both induce the same action by conjugation on NN, contradicting (3.7). This final contradiction has concluded the analysis of this case.

3.5. Gα1G_{\alpha}\neq 1 and Γ/N\Gamma/N is a cubic graph.

Under this assumption, any two distinct neighbours of α\alpha are in distinct NN-orbits, thus Nα=1N_{\alpha}=1. In particular, Lemma 3.2 gives that NN is elementary abelian. Set Γ¯:=Γ/N\bar{\Gamma}:=\Gamma/N, G¯:=G/N\bar{G}:=G/N and α¯:=αN\bar{\alpha}:=\alpha^{N}. Since |VΓ¯|<|VΓ||V\bar{\Gamma}|<|V\Gamma|, by Hypothesis 3.1 the pair (Γ¯,G¯)(\bar{\Gamma},\bar{G}) is not a counterexample to Theorem 1.2 and hence (Γ¯,G¯)(\bar{\Gamma},\bar{G}) is one of the pairs appearing in Table LABEL:table:table. Moreover, since Gα1G_{\alpha}\neq 1, we have the additional information that a vertex-stabilizer G¯α¯Gα\bar{G}_{\bar{\alpha}}\cong G_{\alpha} is not the identity.

We have resolved this case with a computer computation. Since this computer computation is quite involved, we give some details. Let (Γ¯,G¯)(\bar{\Gamma},\bar{G}) be any pair in Table LABEL:table:table, except for the last row. For each prime p{2,3,5}p\in\{2,3,5\}, we have constructed all the irreducible modules of G¯\bar{G} over the field 𝔽p\mathbb{F}_{p} having pp elements. Let VV be one of these irreducible modules. This module VV corresponds to the putative minimal normal subgroup NN of GG. We have constructed all the distinct extensions of G¯\bar{G} via VV. Let EE be one of these extensions and let π:EG¯\pi:E\to\bar{G} be the natural projection with Ker(π)=V\mathrm{Ker}(\pi)=V. This extension EE corresponds to the putative abstract group GG. For each such extension, we have computed all the subgroups HH of EE with the property that π|H\pi_{|H} is an isomorphism between HH and G¯α¯\bar{G}_{\bar{\alpha}}. This subgroup HH is our putative vertex-stabilizer GαG_{\alpha}. This computation can be performed in π1(G¯α¯)\pi^{-1}(\bar{G}_{\bar{\alpha}}). Next, we have constructed the permutation representation EpE_{p} of EE acting on the right cosets of HH in EE. This permutation group EpE_{p} is our putative permutation group GG. If EpE_{p} has semiregular elements of order at least 66, then we have discarded EE from further consideration.

For each permutation group EpE_{p} as above, we have verified, by considering the orbital graphs of EpE_{p}, whether EpE_{p} acts on a connected cubic graph. This is our putative graph Γ\Gamma. This step is by far the most expensive step in the computation.

This whole process had to be applied repeatedly starting with the pairs arising from the census of connected cubic graphs having at most 1 2801\,280 vertices.

For instance, the graphs having 65 61065\,610 vertices were found by applying this procedure starting with the graph having 810810 vertices and its transitive group of automorphisms having 1 6201\,620 elements: here the elementary abelian cover NN has cardinality 81=3481=3^{4}. Incidentally, we have found only one pair up to isomorphism. Next, by applying this procedure to this pair, we found no new examples.

We give some further details of the computation when we applied the procedure with Γ¯\bar{\Gamma} having 1 250=2541\,250=2\cdot 5^{4} vertices and with its corresponding vertex-transitive subgroup G¯\bar{G} having order 2 500=22542\,500=2^{2}\cdot 5^{4}. When we applied this procedure, we have obtained graphs having 255=6 2502\cdot 5^{5}=6\,250 vertices and admitting a group of automorphisms having 2255=12 5002^{2}\cdot 5^{5}=12\,500 elements. Actually, in this step, we have found only one pair up to isomorphism. We have repeated this procedure two more times, obtaining graphs having 256=31 2502\cdot 5^{6}=31\,250 and 257=156 2502\cdot 5^{7}=156\,250 vertices. We were not able to push this computation further. Therefore to complete the proof of Theorem 1.2, we need to show that any new pair (Γ,G)(\Gamma,G) has the property that |VΓ|=25|V\Gamma|=2\cdot 5^{\ell} and |G|=45|G|=4\cdot 5^{\ell}, with 34\ell\leq 34.

From the discussion above we may suppose that |VΓ¯|=25|V\bar{\Gamma}|=2\cdot 5^{\ell} and |G¯|=45|\bar{G}|=4\cdot 5^{\ell} with 34\ell\leq 34. Moreover, Γ¯\bar{\Gamma} is a regular cover of the graph, say Δ\Delta, having 1 2501\,250 vertices and G¯\bar{G} is a quotient of the group of automorphisms of Δ\Delta, say HH, with |H|=2 500|H|=2\,500. In particular, a Sylow 22-subgroup of G¯\bar{G} is cyclic and G¯\bar{G} has a normal Sylow 55-subgroup. (This information can be extracted from the analogous properties of HH.) Let P¯\bar{P} be a Sylow 55-subgroup of G¯\bar{G} and observe that every non-identity element of P¯\bar{P} has order 55 because every semiregular element of G¯\bar{G} has order at most 66. Let PP be the subgroup of GG with G/N=P¯G/N=\bar{P}. Assume NN is not an elementary abelian 55-group. Then NN is an elementary abelian pp-group for some p{2,3}p\in\{2,3\}. Let QQ be a Sylow 55-subgroup of PP and observe that P=NQP=N\rtimes Q. The elements in PP are semiregular and hence each element of PP has order at most 66. This implies that the elements of PP have order 11, 55 or pp. This implies that the action, by conjugation, of QQ on NN is fixed-point-free and PP is a Frobenius group with Frobenius kernel NN and Frobenius complement QQ. The structure theorem of Frobenius complements gives that QQ is cyclic and hence |Q|=5|Q|=5, which is a contradiction. This contradiction has shown that NN is an elementary abelian 55-group and hence PP is a Sylow 55-subgroup of GG. Moreover, G=PxG=P\rtimes\langle x\rangle, where x\langle x\rangle is a cyclic group of order 44. We have shown that |VΓ|=25|V\Gamma|=2\cdot 5^{\ell^{\prime}} and |G|=225|G|=2^{2}\cdot 5^{\ell^{\prime}}. Therefore, it remains to show that 34\ell^{\prime}\leq 34.

Since |Gα|=2|G_{\alpha}|=2, GαG_{\alpha} fixes a unique neighbour of α\alpha. Let us call α\alpha^{\prime} this neighbour. Now, G{α,α}G_{\{\alpha,\alpha^{\prime}\}} has order 44 because {α,α}\{\alpha,\alpha^{\prime}\} is a block of imprimitivity for the action of GG on VΓV\Gamma. Therefore, by Sylow’s theorem, we may suppose that

G{α,α}=x.G_{\{\alpha,\alpha^{\prime}\}}=\langle x\rangle.

In particular, Gα=x2G_{\alpha}=\langle x^{2}\rangle.

Let β\beta and γ\gamma be the neighbours of α\alpha with βαγ\beta\neq\alpha^{\prime}\neq\gamma. Clearly, |G{α,β}|=2|G_{\{\alpha,\beta\}}|=2 and hence, by Sylow’s theorem,

G{α,β}=(x2)y,G_{\{\alpha,\beta\}}=\langle(x^{2})^{y}\rangle,

for some yPy\in P.

Since Γ\Gamma is connected, we have

G=G{α,α},G{α,β}=x,(x2)y=x,y1yx2.G=\langle G_{\{\alpha,\alpha^{\prime}\}},G_{\{\alpha,\beta\}}\rangle=\langle x,(x^{2})^{y}\rangle=\langle x,y^{-1}y^{x^{2}}\rangle.

As PGP\unlhd G and o(x)=4o(x)=4, we deduce

P=y1yx2,(y1yx2)x,(y1yx2)x2,(y1yx2)x3.P=\langle y^{-1}y^{x^{2}},(y^{-1}y^{x^{2}})^{x},(y^{-1}y^{x^{2}})^{x^{2}},(y^{-1}y^{x^{2}})^{x^{3}}\rangle.

Now,

(y1yx2)x2=(yx2)1yx4=(yx2)1y=(y1yx2)1.(y^{-1}y^{x^{2}})^{x^{2}}=(y^{x^{2}})^{-1}y^{x^{4}}=(y^{x^{2}})^{-1}y=(y^{-1}y^{x^{2}})^{-1}.

Therefore, P=y1yx2,(y1yx2)xP=\langle y^{-1}y^{x^{2}},(y^{-1}y^{x^{2}})^{x}\rangle is a 22-generated group of exponent 55. In view of the restricted Burnside problem (see [HWW74] and [Zel91]), the order of PP is at most 5345^{34} and hence 34\ell^{\prime}\leq 34.

References

  • [BGS22a] M. Barbieri, V. Grazian, and P. Spiga. On the Cayleyness of Praeger–Xu graphs. Bulletin of the Australian Mathematical Society, 106(3):353–356, 2022.
  • [BGS22b] M. Barbieri, V. Grazian, and P. Spiga. On the number of fixed edges of automorphisms of vertex-transitive graphs of small valency. Journal of Algebraic Combinatorics, 2022.
  • [BS91] R. Brandl and W. Shi. Finite groups whose element orders are consecutive integers. Journal of Algebra, 143(2):388–400, 1991.
  • [CSS06] P. Cameron, J. Sheehan, and P. Spiga. Semiregular automorphisms of vertex-transitive cubic graphs. European Journal of Combinatorics, 27(6):924–930, aug 2006.
  • [Djo80] D. Ž. Djoković. A class of finite group-amalgams. Proceedings of the American Mathematical Society, 80:22–26, 1980.
  • [DM80] D.Ž. Djoković and G.L. Miller. Regular groups of automorphisms of cubic graphs. Journal of Combinatorial Theory, Series B, 29(2):195–230, 1980.
  • [GM99] N. D. Gupta and V. D. Mazurov. On groups with small orders of elements. Bulletin of the Australian Mathematical Society, 60(2):197–205, 1999.
  • [GP94] A. Gardiner and C. E. Praeger. A characterization of certain families of 4 symmetric graphs. European Journal of Combinatorics, 15:383–397, 1994.
  • [HWW74] G. Havas, G. E. Wall, and J. W. Wamsley. The two generator restricted Burnside group of exponent five. Bulletin of the Australian Mathematical Society, 10(3):459––470, 1974.
  • [Jor88] D. Jordan. Eine Symmetrieeigenschaft von Graphen, Graphentheorie und ihre Anwendungen (Stadt Wehlen, 1988). Dresdner Reihe Forsch, 9:17–20, 1988.
  • [JPW19] R. Jajcay, P. Potočnik, and S. Wilson. The Praeger-Xu graphs: cycle structures, maps and semitransitive orientations. Acta Mathematica Universitatis Comenianae, 88:22–26, 2019.
  • [JPW22] R. Jajcay, P. Potočnik, and S. Wilson. On the Cayleyness of Praeger-Xu graphs. Journal of Combinatorial Theory. Series B, 152:55–79, 2022.
  • [Kli98] M. Klin. On transitive permutation groups without semi-regular subgroups. ICM 1998: International Congress of Mathematicians, Berlin, 18–27 August 1998. Abstracts of short communications and poster sessions, page 279, 1998.
  • [LW74] J. S. Leon and D. B. Wales. Simple groups of order 2a3bpc2^{a}3^{b}p^{c} with cyclic Sylow pp-groups. Journal of Algebra, 29:246–254, 1974.
  • [Mar81] D. Marušič. On vertex symmetric digraphs. Discrete Mathematics, 36(1):69–81, 1981.
  • [MS98] D. Marušič and R. Scapellato. Permutation groups, vertex-transitive digraphs and semiregular automorphisms. European Journal of Combinatorics, 19(6):707–712, 1998.
  • [MZ99] V. D. Mazurov and A. Kh. Zhurtov. On recognition of the finite simple groups L2(2m)L_{2}(2^{m}) in the class of all groups. Siberian Mathematical Journal, 40(1):62–64, 1999.
  • [Neu37] B. H. Neumann. Groups Whose Elements Have Bounded Orders. The Journal of the London Mathematical Society, 12(3):195–198, 1937.
  • [New79] M. F. Newman. Groups of exponent dividing seventy. The Mathematical Scientist, 4(2):149–157, 1979.
  • [Pra89] C. E. Praeger. Highly arc transitive digraphs. European Journal of Combinatorics, 10(3):281–292, 1989.
  • [PS21] P. Potočnik and P. Spiga. On the number of fixed points of automorphisms of vertex-transitive graphs of bounded valency. Combinatorica, 41(5):703–747, 2021.
  • [PSV13] P. Potočnik, P. Spiga, and G. Verret. Cubic vertex-transitive graphs on up to 1280 vertices. Journal of Symbolic Computation, 50:465–477, 2013.
  • [PX89] C. E. Praeger and M.-Y. Xu. A characterization of a class of symmetric graphs of twice prime valency. European Journal of Combinatorics, 10:91–102, 1989.
  • [Spi14] P. Spiga. Semiregular elements in cubic vertex-transitive graphs and the restricted Burnside problem. Mathematical Proceedings of the Cambridge Philosophical Society, 157(1):45–61, 2014.
  • [Zel91] E.I. Zelmanov. Solution of the restricted Burnside problem for groups of odd exponent. Mathematics of the USSR-Izvestiya, 36:41–60, 1991.