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

Extremely primitive groups and linear spaces

Melissa Lee, Gabriel Verret Department of Mathematics, University of Auckland, Auckland, New Zealand melissa.lee@auckland.ac.nz, g.verret@auckland.ac.nz
Abstract.

A finite non-regular primitive permutation group GG is extremely primitive if a point stabiliser acts primitively on each of its nontrivial orbits. Such groups have been studied for almost a century, finding various applications. The classification of extremely primitive groups was recently completed by Burness and Lee, who relied on an earlier classification of soluble extremely primitive groups by Mann, Praeger and Seress. Unfortunately, there is an inaccuracy in the latter classification. We correct this mistake, and also investigate regular linear spaces which admit groups of automorphisms that are extremely primitive on points.

1. Introduction

All groups and linear spaces in this paper are assumed to be finite. Let GG be a non-regular primitive group. We say that GG is extremely primitive if a point stabiliser GαG_{\alpha} acts primitively on each of its nontrivial orbits. Some examples of extremely primitive groups include Symn\operatorname{Sym}_{n} in its natural action and PGL2(q)\operatorname{PGL}_{2}(q) on the projective line over 𝔽q\operatorname{\mathbb{F}}_{q}. Extremely primitive groups have been studied for almost a century [13], and have arisen several other contexts, including in the constructions of the simple sporadic groups J2\mathrm{J}_{2} and HS\mathrm{HS}, as well as in the study of transitive permutation groups with a given upper bound on their subdegrees (see [14] for example).

The problem of classifying extremely primitive groups has garnered significant attention in recent years and a final classification has recently been completed following a series of papers [12, 2, 3, 4, 6, 5]. Unfortunately, an oversight was made in the classification of soluble extremely primitive groups in [12]. In this paper, the authors prove a set of necessary conditions for a soluble group to be extremely primitive [12, Lemma 3.3]. They then claim that these conditions are sufficient [12, Theorem 1.2] but this is never proved and turns out to be incorrect. This error is then reproduced in later work that relies on this result, for example [2, Theorem 1] and [6, Theorem 4].

The first purpose of this paper is to correct the classification of soluble extremely primitive groups. Recall that for a prime pp and positive integer dd, a primitive prime divisor of pd1p^{d}-1 is a prime that divides pd1p^{d}-1 but not pi1p^{i}-1 for each 1id11\leq i\leq d-1.

Theorem 1.1.

A soluble group GG is extremely primitive if and only if G=VHAΓL1(pd)G=V\rtimes H\leq\operatorname{A\Gamma L}_{1}(p^{d}) for a prime pp and d1d\geq 1, where VV is the natural vector space and CtCeHΓL1(pd)\operatorname{C}_{t}\rtimes\operatorname{C}_{e}\cong H\leq\operatorname{\Gamma L}_{1}(p^{d}), where t=|HGL1(pd)|t=|H\cap\operatorname{GL}_{1}(p^{d})| is a primitive prime divisor of pd1p^{d}-1 and either e=1e=1, or ee is a prime dividing dd and t=(pd1)/(pd/e1)t=(p^{d}-1)/(p^{d/e}-1).

This corrected version of the classification of soluble extremely primitive groups also impacts results which relied on it. For example, see Theorem 4.4 for an updated version of [2, Theorem 1].

Our next goal is to study linear spaces that admit a group of automorphisms that is extremely primitive on points. We first introduce some terminology. A linear space S=(𝒫,)S=(\mathcal{P},\mathcal{L}) is a point-line incidence structure with a set 𝒫\mathcal{P} of points and a set \mathcal{L} of lines, where each line is a subset of 𝒫\mathcal{P} such that each pair of points is contained in exactly one line. A linear space is nontrivial if it has at least two lines and if every line has at least three points. A linear space is regular if all its lines have the same size. An automorphism of SS is a permutation of 𝒫\mathcal{P} which preserves \mathcal{L}. The automorphisms of SS form its automorphism group Aut(S)\operatorname{Aut}(S). (One could consider automorphisms as acting simultaneously on points and lines but, in a nontrivial linear space, the action on points is already faithful and this viewpoint will usually be simpler for us.)

If S=(𝒫,)S=(\mathcal{P},\mathcal{L}) and S=(𝒫,)S^{\prime}=(\mathcal{P},\mathcal{L}^{\prime}) are two linear spaces with the same set of points, we say that SS^{\prime} is a refinement of SS if every line of SS^{\prime} is contained in a line of SS. (In other words, for every \ell^{\prime}\in\mathcal{L}^{\prime}, there exists \ell\in\mathcal{L} such that \ell^{\prime}\subseteq\ell.)

Theorem 1.2.

Let SS be a nontrivial regular linear space. If GAut(S)G\leq\operatorname{Aut}(S) is extremely primitive, then one of the following holds:

  1. (1)

    G=PSL2(22n)G=\operatorname{PSL}_{2}(2^{2^{n}}) with point stabiliser D2(22n+1)\operatorname{D}_{2(2^{2^{n}}+1)}, where 22n+12^{2^{n}}+1 is a Fermat prime, and SS is a refinement of 𝒮(G)\operatorname{\mathcal{LS}}(G);

  2. (2)

    GG is as in Theorem 1.1 with e2e\geq 2, and SS is a refinement of 𝒮(G)\operatorname{\mathcal{LS}}(G);

  3. (3)

    GG is as in Theorem 1.1 with e=1e=1.

In cases (1) and (2) of Theorem 1.2, there is a unique “coarsest” linear space 𝒮(G)\operatorname{\mathcal{LS}}(G) admitting GG as a group of automorphisms. These linear spaces are defined in Section 3 and some of their properties described in Section 6. It is also possible to describe the admissible refinements of 𝒮(G)\operatorname{\mathcal{LS}}(G) in a systematic manner, see Proposition 5.3 and Examples 6.6 and 6.7. Classifying the linear spaces which arise in case (3) seems much more difficult. We give some examples and discuss this further in Section 6.

2. Proof of Theorem 1.1

Let GG be a soluble extremely primitive group. As noted in the proof of [12, Lemma 3.3], GG is a soluble 3/23/2-transitive group. (Recall that a transitive permutation group GG on a set Ω\Omega is called 3/23/2-transitive if all orbits of GωG_{\omega} on Ω{ω}\Omega\setminus\{\omega\} have the same size, with this size being greater than 1.) Such groups were classified by Passman. Before stating this classification, we require the following definition. For an odd prime pp and integer d1d\geq 1, let 𝒢(pd)\mathscr{G}(p^{d}) be the subgroup of AGL2(pd)\operatorname{AGL}_{2}(p^{d}) containing all translations and whose point stabiliser consists of all diagonal and antidiagonal matrices in GL2(pd)\operatorname{GL}_{2}(p^{d}) with determinant ±1\pm 1.

Theorem 2.1 ([15, 16]).

If GG is a soluble 3/23/2-transitive group, then one of the following holds:

  1. (1)

    GG is a Frobenius group;

  2. (2)

    GAΓL1(pd)G\leq\operatorname{A\Gamma L}_{1}(p^{d}) for a prime pp and d1d\geq 1;

  3. (3)

    G=𝒢(pd)G=\mathscr{G}(p^{d}) for an odd prime pp and d1d\geq 1;

  4. (4)

    GG is one of a collection of groups of degree 32,52,72,112,1723^{2},5^{2},7^{2},11^{2},17^{2} or 343^{4}.

We now examine each of the cases of Theorem 2.1 and show that GG is extremely primitive if and only if it is as in Theorem 1.1. First, if GG is as in Theorem 2.1 (1), then the proof of [12, Lemma 3.3] implies that GG is extremely primitive if and only if it is as in Theorem 1.1 with e=1e=1. We also use GAP [8] to confirm that no group in Theorem 2.1 (4) is extremely primitive.

2.1. G=𝒢(pd)G=\mathscr{G}(p^{d})

Let G=𝒢(pd)G=\mathscr{G}(p^{d}) as in the preamble of Theorem 2.1. Let u=[0,0]u=[0,0] and v=[1,0]v=[1,0]. By definition, GuG_{u} consists of all diagonal and antidiagonal matrices in GL2(pd)\operatorname{GL}_{2}(p^{d}) with determinant ±1\pm 1. This is a group of order 4(pd1)4(p^{d}-1). Next, Guv=(1001)G_{uv}=\langle\left(\begin{smallmatrix}1&0\\ 0&-1\end{smallmatrix}\right)\rangle, so |Guv|=2|G_{uv}|=2. By Sylow’s Theorem, for GuvG_{uv} to be maximal in GuG_{u}, one must have pd1=1p^{d}-1=1, that is pd=2p^{d}=2, contradicting that fact that pp is odd. So GuvG_{uv} is not maximal in GuG_{u} and thus GG is not extremely primitive.

2.2. GAΓL1(pd)G\leq\operatorname{A\Gamma L}_{1}(p^{d})

We need the following preliminary result.

Proposition 2.2.

Let pp be a prime and d1d\geq 1, let HΓL1(pd)H\leq\operatorname{\Gamma L}_{1}(p^{d}) and let T=HGL1(pd)T=H\cap\operatorname{GL}_{1}(p^{d}). We have H=TEH=T\rtimes E with TCtT\cong\operatorname{C}_{t}, ECeE\cong\operatorname{C}_{e} and ee divides dd. Let VCpdV\cong\operatorname{C}_{p}^{d} be the natural vector space for ΓL1(pd)\operatorname{\Gamma L}_{1}(p^{d}). The following are equivalent:

  1. (1)

    TT and HH have the same orbits on VV;

  2. (2)

    the orbits of HH on V{0}V\setminus\{0\} all have the same size;

  3. (3)

    e=1e=1 or pd1p^{d}-1 divides t(pd/e1)t(p^{d/e}-1).

Proof.

Since GL1(pd)\operatorname{GL}_{1}(p^{d}) acts regularly on V{0}V\setminus\{0\}, we can identify this set with GL1(pd)\operatorname{GL}_{1}(p^{d}). Let α\alpha be a generator of GL1(pd)Cpd1\operatorname{GL}_{1}(p^{d})\cong\operatorname{C}_{p^{d}-1}. Let m=pd1tm=\frac{p^{d}-1}{t} and note that T=αmT=\langle\alpha^{m}\rangle and the orbits of TT on V{0}V\setminus\{0\} are its cosets, so of the form αiαm\alpha^{i}\langle\alpha^{m}\rangle for some ii. In particular, the orbits of TT on V{0}V\setminus\{0\} all have the same size so (1)(2)(1)\Longrightarrow(2).

Write E=fE=\langle f\rangle. Now EE acts as field automorphisms on VV and thus αf=αpd/e\alpha^{f}=\alpha^{p^{d/e}}. So

(αi+jm)f=α(i+jm)(pd/e)=αipd/e+jmpd/e.(\alpha^{i+jm})^{f}=\alpha^{(i+jm)(p^{d/e})}=\alpha^{ip^{d/e}+jmp^{d/e}}.

This calculation shows that ff preserves αiαm\alpha^{i}\langle\alpha^{m}\rangle if and only if

ipd/ei(modm).ip^{d/e}\equiv i\pmod{m}.

This always holds for i=0i=0, so ff always preserves TT. In particular, HH always has an orbit of size |T||T| on V{0}V\setminus\{0\}. On the other hand, ipd/ei(modm)ip^{d/e}\equiv i\pmod{m} holds for every ii if and only if pd/e1(modm)p^{d/e}\equiv 1\pmod{m} if and only if mm divides pd/e1p^{d/e}-1.

Now, if all the orbits of HH on V{0}V\setminus\{0\} all have the same size, they must have size |T||T| and mm must divide pd/e1p^{d/e}-1, so (2)(3)(2)\Longrightarrow(3).

Finally, if e=1e=1, then H=TH=T, and if mm divides pd/e1p^{d/e}-1, then by the above TT and HH have the same orbits on V{0}V\setminus\{0\} and thus also on VV, so (3)(1)(3)\Longrightarrow(1). ∎

Remark 2.3.

A version of Proposition 2.2 with the extra assumption that HH is pp-exceptional appears as [9, Lemma 2.7].

We are now ready to classify the extremely primitive groups in this family.

Theorem 2.4.

Let pp be a prime and d1d\geq 1, let HΓL1(pd)H\leq\operatorname{\Gamma L}_{1}(p^{d}) and let T=HGL1(pd)T=H\cap\operatorname{GL}_{1}(p^{d}). We have H=TEH=T\rtimes E with TCtT\cong\operatorname{C}_{t}, ECeE\cong\operatorname{C}_{e} and ee divides dd. Let VCpdV\cong\operatorname{C}_{p}^{d} be the natural vector space for ΓL1(pd)\operatorname{\Gamma L}_{1}(p^{d}) and let G=VHAΓL(1,pd)G=V\rtimes H\leq\operatorname{A\Gamma L}(1,p^{d}). Then GG is extremely primitive if and only if the following conditions hold:

  1. (1)

    tt is a primitive prime divisor of pd1p^{d}-1, and

  2. (2)

    e=1e=1 or pd1=t(pd/e1)p^{d}-1=t(p^{d/e}-1).

Moreover, all of the nontrivial orbits of HH have size tt.

Proof.

Note that GL1(pd)\operatorname{GL}_{1}(p^{d}) is regular on V{0}V\setminus\{0\} so TT is semiregular on V{0}V\setminus\{0\}. In particular, it acts regularly on all of its nontrivial orbits and they each have size tt. A regular group is primitive if and only if it is trivial or has prime order, so we can assume tt is 1 or prime. If tt is not a primitive prime divisor of pd1p^{d}-1, then TGL1(pb)T\leq\operatorname{GL}_{1}(p^{b}) for some proper subfield 𝔽pb𝔽pd\operatorname{\mathbb{F}}_{p^{b}}\subset\operatorname{\mathbb{F}}_{p^{d}} and Cpb<V\operatorname{C}_{p}^{b}<V is HH-invariant hence GG is not primitive. We thus assume that tt is a primitive prime divisor of pd1p^{d}-1 and show that GG is extremely primitive if and only if (2) holds. Since tt is a primitive prime divisor of pd1p^{d}-1, TT acts irreducibly on VV so VTV\rtimes T is primitive, and so is GG.

We claim that GG is extremely primitive if and only HH has the same orbits as TT. Indeed, apart from the trivial orbit of size 11, an orbit of HH must have order atat for some aa. A stabiliser of a point in such an orbit is a subgroup of order e/ae/a of Ce\operatorname{C}_{e}. This is maximal in HH if and only if a=1a=1. This proves our claim.

By Proposition 2.2, HH and TT have the same orbits if and only if e=1e=1 or pd1p^{d}-1 divides t(pd/e1)t(p^{d/e}-1). Now, tt is a primitive prime divisor of pd1p^{d}-1, so if e2e\geq 2, then tt does not divide pd/e1p^{d/e}-1 which itself divides pd1p^{d}-1. Since tt is prime, it follows that tt divides pd1pd/e1\frac{p^{d}-1}{p^{d/e}-1} and thus pd/e1p^{d/e}-1 divides pd1t\frac{p^{d}-1}{t} hence pd1=t(pd/e1)p^{d}-1=t(p^{d/e}-1). ∎

Remark 2.5.

  1. (1)

    Note that the condition pd1=t(pd/e1)p^{d}-1=t(p^{d/e}-1) is quite restrictive. For example, if d=2d=2, then it is only satisfied when p=2p=2, e=2e=2 and t=3t=3.

  2. (2)

    Moreover, this condition, together with the fact that tt is prime, implies that ee cannot be composite. Indeed, if ff is a divisor of ee, then pd/f1p^{d/f}-1 is a multiple of pd/e1p^{d/e}-1 and a divisor of pd1p^{d}-1, but pd1pd/e1\frac{p^{d}-1}{p^{d/e}-1} is a prime.

  3. (3)

    The converse of the previous remark does not hold. In other words, ee being prime is not sufficient to guarantee that pd1=t(pd/e1)p^{d}-1=t(p^{d/e}-1). The smallest counterexample is (p,d,t,e)=(5,2,3,2)(p,d,t,e)=(5,2,3,2).

3. Groups with Property ()(\star) and the linear space 𝒮(G)\operatorname{\mathcal{LS}}(G)

In this section, we define and prove some basic facts about 𝒮(G)\operatorname{\mathcal{LS}}(G), the linear space that appears in Theorem 1.2 (2).

Definition 3.1.

We say that a permutation group GG on Ω\Omega has Property ()(\star) if, for all u,v,wΩu,v,w\in\Omega with uwu\neq w,

GuvGwGuwGv.G_{uv}\leq G_{w}\Longrightarrow G_{uw}\leq G_{v}.

Let GG be a group with Property ()(\star). For u,vΩu,v\in\Omega, let Λuv={wΩGuvGw}\Lambda_{uv}=\{w\in\Omega\mid G_{uv}\leq G_{w}\} and let 𝒮(G)\operatorname{\mathcal{LS}}(G) be the point-line incidence structure having Ω\Omega as set of points and {Λuvu,vΩ,uv}\{\Lambda_{uv}\mid u,v\in\Omega,u\neq v\} as set of lines.

Proposition 3.2.

If GSym(Ω)G\leq\operatorname{Sym}(\Omega) has Property ()(\star), then 𝒮(G)\operatorname{\mathcal{LS}}(G) is a linear space with GAut(𝒮(G))G\leq\operatorname{Aut}(\operatorname{\mathcal{LS}}(G)).

Proof.

Let uu and vv be distinct points. Clearly, we have u,vΛuvu,v\in\Lambda_{uv}. We show that if wΛuvw\in\Lambda_{uv} and uwu\neq w, then Λuv=Λuw\Lambda_{uv}=\Lambda_{uw}. Since wΛuvw\in\Lambda_{uv}, we have GuvGwG_{uv}\leq G_{w} and then GuwGvG_{uw}\leq G_{v} by Property ()(\star). Let xΛuvx\in\Lambda_{uv}, then Guw=GuGuwGuGv=GuvGxG_{uw}=G_{u}\cap G_{uw}\leq G_{u}\cap G_{v}=G_{uv}\leq G_{x} so that xΛuwx\in\Lambda_{uw}. This shows that ΛuvΛuw\Lambda_{uv}\subseteq\Lambda_{uw}. Similarly, for yΛuwy\in\Lambda_{uw}, we have Guv=GuGuvGuwGyG_{uv}=G_{u}\cap G_{uv}\leq G_{uw}\leq G_{y} so yΛuvy\in\Lambda_{uv}, and Λuv=Λuw\Lambda_{uv}=\Lambda_{uw} as claimed.

Now, suppose that u,vΛabu,v\in\Lambda_{ab} for some aba\neq b and uvu\neq v. We show that Λab=Λuv\Lambda_{ab}=\Lambda_{uv}. This is clear if {u,v}={a,b}\{u,v\}=\{a,b\} so we assume this is not the case. Without loss of generality, we may assume that uavu\neq a\neq v. By the previous paragraph, we have Λau=Λab=Λav\Lambda_{au}=\Lambda_{ab}=\Lambda_{av}. So vΛauv\in\Lambda_{au} and hence, by the previous paragraph, Λau=Λuv\Lambda_{au}=\Lambda_{uv} and hence Λab=Λuv\Lambda_{ab}=\Lambda_{uv}. This shows that 𝒮(G)\operatorname{\mathcal{LS}}(G) is a linear space. The fact that GAut(𝒮(G))G\leq\operatorname{Aut}(\operatorname{\mathcal{LS}}(G)) is obvious from the definition. ∎

Recall that a flag in a linear space is a pair (u,)(u,\ell) such that uu is a point, \ell is a line and uu\in\ell.

Definition 3.3.

Let SS be a linear space and GAut(S)G\leq\operatorname{Aut}(S). We say that (S,G)(S,G) is transverse if, for every flag (u,)(u,\ell) of SS and every orbit Δ\Delta of GuG_{u}, we have |Δ|1|\ell\cap\Delta|\leq 1.

Proposition 3.4.

Let SS be a linear space and GAut(S)G\leq\operatorname{Aut}(S). If GG has Property ()(\star) and (S,G)(S,G) is transverse, then SS is a refinement of 𝒮(G)\operatorname{\mathcal{LS}}(G).

Proof.

Write S=(𝒫,)S=(\mathcal{P},\mathcal{L}), let \ell\in\mathcal{L} and let u,vu,v\in\ell, with uvu\neq v. Recall that Λuv={wΩGuvGw}\Lambda_{uv}=\{w\in\Omega\mid G_{uv}\leq G_{w}\} is the unique line of 𝒮(G)\operatorname{\mathcal{LS}}(G) containing uu and vv. We must show that Λuv\ell\subseteq\Lambda_{uv}. Suppose, by contradiction, that ww\in\ell but wΛuvw\not\in\Lambda_{uv}. Since wΛuvw\notin\Lambda_{uv}, ww is not fixed by GuvG_{uv}, so |wGuv|2|w^{G_{uv}}|\geq 2. On the other hand, GuvG_{uv} fixes uu and vv and thus the unique line of SS containing them, namely \ell. This implies that wGuvw^{G_{uv}}\subseteq\ell and thus |wGu|2|w^{G_{u}}\cap\ell|\geq 2, contradicting the hypothesis that (S,G)(S,G) is transverse. ∎

We end this section by showing exhibiting a nice family of groups with Property ()(\star). This will be useful in later sections.

Lemma 3.5.

Let pp be a prime and d1d\geq 1, let HΓL1(pd)H\leq\operatorname{\Gamma L}_{1}(p^{d}) and let VCpdV\cong\operatorname{C}_{p}^{d} be the natural vector space for ΓL1(pd)\operatorname{\Gamma L}_{1}(p^{d}). Let GG be transitive on VV with point stabiliser HH. If the equivalent conditions from Proposition 2.2 are satisfied, then GG has Property ()(\star).

Proof.

Let u,v,wu,v,w be points such that uwu\neq w and GuvGwG_{uv}\leq G_{w}. We need to show that GuwGvG_{uw}\leq G_{v}. This is clear if u=vu=v, so we may assume that uvu\neq v. We can also assume without loss of generality that u=0u=0 and thus H=GuH=G_{u}. We then have HvHwH_{v}\leq H_{w} and v,wV{0}v,w\in V\setminus\{0\}. It follows by Proposition 2.2 (2) that |Hv|=|Hw||H_{v}|=|H_{w}| hence Hw=HvGvH_{w}=H_{v}\leq G_{v}, as required. ∎

4. Proof of Theorem 1.2

We start with a few preliminary results. Recall that the rank of a transitive permutation group is the number of orbits of a point stabiliser.

Theorem 4.1 ([10, Theorem 1.1]).

Let SS be a regular linear space. If GAut(S)G\leq\operatorname{Aut}(S) is extremely primitive on 𝒫\mathcal{P}, then rank(G)3\operatorname{rank}(G)\geq 3 with equality only if SS if the affine space AGm(3)\mathrm{AG}_{m}(3) and GAΓL1(3m)G\leq\operatorname{A\Gamma L}_{1}(3^{m}).

Proposition 4.2 ([10, Lemma 2.6]).

Let SS be a linear space, let GAut(S)G\leq\operatorname{Aut}(S) be extremely primitive and let (u,)(u,\ell) be a flag of SS. If Δ\Delta is an orbit of GuG_{u}, then |Δ|{0,1,|Δ|}|\ell\cap\Delta|\in\{0,1,|\Delta|\}.

Our starting point is the following theorem.

Theorem 4.3 ([10, Corollary 1.4]).

Let SS be a nontrivial regular linear space. If GAut(S)G\leq\operatorname{Aut}(S) is extremely primitive, then one of the following holds:

  1. (1)

    GG is primitive of affine type;

  2. (2)

    GG is an almost simple exceptional group of Lie type;

  3. (3)

    G=PSL2(22n)G=\operatorname{PSL}_{2}(2^{2^{n}}) with point stabiliser D2(22n+1)\operatorname{D}_{2(2^{2^{n}}+1)}, where 22n+12^{2^{n}}+1 is a Fermat prime.

For the rest of this section, we will assume the hypothesis of Theorem 4.3. We will then deal with each case in the conclusion in turn and classify the linear spaces S=(𝒫,)S=(\mathcal{P},\mathcal{L}) that arise. First, we introduce some basic terminology and results which will often be useful. Write v=|𝒫|v=|\mathcal{P}| and b=||b=|\mathcal{L}|. Since SS is regular, all its lines have the same size, which we will denote by kk. It can be shown (see [7, Lemma 2.1], for example) that there is also a constant number of lines rr meeting each point and the following holds:

r=v1k1b=v(v1)k(k1)vk(k1)+1.r=\frac{v-1}{k-1}\qquad b=\frac{v(v-1)}{k(k-1)}\qquad v\geq k(k-1)+1.

Moreover, it follows by Fisher’s inequality that bvb\geq v and therefore rkr\geq k.

First, let GG be as in Theorem 4.3 (2), that is, an extremely primitive almost simple exceptional group of Lie type, and let HH be its point stabiliser. By [6, Theorem 1] (G,H)(G,H) is one of (G2(4),J2)(\mathrm{G}_{2}(4),\mathrm{J}_{2}) or (G2(4).2,J2.2)(\mathrm{G}_{2}(4).2,\mathrm{J}_{2}.2). In each case, the corresponding permutation group has rank 33 and it then follows by Theorem 4.1 that no regular linear space arises in this case.

Next, let GG be as in Theorem 4.3 (3). Write q=22n+1q=2^{2^{n}}+1. By [3, Proposition 5.3], the nontrivial subdegrees of GG are all qq. Since the point stabilisers are isomorphic to D2q\operatorname{D}_{2q}, it follows that, given two distinct points uu and vv, |Guv|=2|G_{uv}|=2. In particular, if uwu\neq w and GuvGwG_{uv}\leq G_{w}, then Guw=GuvG_{uw}=G_{uv}, since both groups have order 22. This shows that GG satisfies Property ()(\star). Now, let (u,)(u,\ell) be a flag of SS and Δ\Delta be an orbit of GuG_{u} that meets \ell. By Proposition 4.2, |Δ|{1,q}|\ell\cap\Delta|\in\{1,q\}. If |Δ|=q|\ell\cap\Delta|=q, then kq+1k\geq q+1 but v=|PSL2(q1):D2q|=(q1)(q2)/2v=|\operatorname{PSL}_{2}(q-1):\operatorname{D}_{2q}|=(q-1)(q-2)/2, contradicting the fact that vk(k1)+1v\geq k(k-1)+1. It follows that (S,G)(S,G) is transverse and we can apply Proposition 3.4 to conclude that SS is a refinement of 𝒮(G)\operatorname{\mathcal{LS}}(G), as in Theorem 1.2 (1).

It remains to deal with the case when GG is as in Theorem 4.3 (1), that is, an extremely primitive group of affine type. These groups were previously classified in [2, Theorem 1] but this classification relied on the incorrect [12, Theorem 1.2]. Here is then an updated classification of these groups.

Theorem 4.4.

Let pp be a prime and d1d\geq 1, let HGLd(p)H\leq\operatorname{GL}_{d}(p), let VV be the natural vector space for GLd(p)\operatorname{GL}_{d}(p) and let G=VHG=V\rtimes H. If GG is extremely primitive, then one of the following holds:

  1. (1)

    GG is as in Theorem 1.1;

  2. (2)

    p=2p=2 and H=PSLd(2)H=\operatorname{PSL}_{d}(2) with d3d\geq 3, or H=Spd(2)H={\rm Sp}_{d}(2) with d4d\geq 4;

  3. (3)

    p=2p=2 and (d,H)=(4,Alt6)(d,H)=(4,\operatorname{Alt}_{6}), (4,Alt7)(4,\operatorname{Alt}_{7}), (6,PSU3(3))(6,\operatorname{PSU}_{3}(3)) or (6,PSU3(3).2)(6,\operatorname{PSU}_{3}(3).2);

  4. (4)

    p=2p=2 and (d,H)(d,H) is one of the following:

    (10,M12)(10,M22)(10,M22.2)(11,M23)(11,M24)(22,Co3)(24,Co1)(2k,Alt2k+1)(2k,Sym2k+1)(2,Alt2+2)(2,Sym2+2)(2,Ω2±(2))(2,O2±(2))(8,PSL2(17))(8,Sp6(2))\begin{array}[]{llll}(10,{\rm M}_{12})&(10,{\rm M}_{22})&(10,{\rm M}_{22}.2)&(11,{\rm M}_{23})\\ (11,{\rm M}_{24})&(22,{\rm Co}_{3})&(24,{\rm Co}_{1})&(2k,\operatorname{Alt}_{2k+1})\\ (2k,\operatorname{Sym}_{2k+1})&(2\ell,\operatorname{Alt}_{2\ell+2})&(2\ell,\operatorname{Sym}_{2\ell+2})&(2\ell,\Omega_{2\ell}^{\pm}(2))\\ (2\ell,{\rm O}_{2\ell}^{\pm}(2))&(8,\operatorname{PSL}_{2}(17))&(8,{\rm Sp}_{6}(2))&\end{array}

    where k2k\geq 2 and 3\ell\geq 3.

In the rest of this section, we go through the cases in Theorem 4.4 and, for each case, classify regular linear spaces admitting such groups of automorphisms.

4.1. Groups arising in Theorem 4.4 (2) or (3).

The groups in Theorem 4.4 (2) are 22-transitive by [11, Lemma 2.10.5], while those in Theorem 4.4 (3) are found to be 22-transitive by direct computation. It follows by Theorem 4.1 that no regular linear space arises in this case.

4.2. Groups arising in Theorem 4.4 (4).

Adopt the notation of Theorem 4.4 (4). Let uu be a point such that H=GuH=G_{u}.

We first deal with the infinite families of groups. If H=Ω2l±(2)H=\Omega^{\pm}_{2l}(2) or O2l±(2)\mathrm{O}^{\pm}_{2l}(2), then GG has rank 3 by [11, Lemma 2.10.5] and by Theorem 4.1 no regular linear space arises in this case. Now suppose that (d,H)(d,H) is one of (2k,Alt2k+1),(2k,Sym2k+1),(2,Alt2+2)(2k,\operatorname{Alt}_{2k+1}),(2k,\operatorname{Sym}_{2k+1}),(2\ell,\operatorname{Alt}_{2\ell+2}) or (2,Sym2+2)(2\ell,\operatorname{Sym}_{2\ell+2}), and for conciseness, let n=2k+1n=2k+1 or 2+22\ell+2 as appropriate. In these cases, H=GuH=G_{u} is an alternating or symmetric group acting on the fully deleted permutation module UU. That is, the subspace V0V_{0} of Vn(2)V_{n}(2) spanned by vectors whose entries sum to 0 if nn is odd, or the quotient of V0V_{0} by the subspace spanned by the all-ones vector if nn is even; HH acts by permuting the coordinates of elements of V0V_{0}, with the corresponding induced action on UU. The orbits of HH on V0V_{0} are indexed by the weight of the vectors in it, which must be even, so HH has n/2+1\lfloor n/2\rfloor+1 orbits on V0V_{0}. When nn is even, the orbits of weight xx and nxn-x in V0V_{0} get identified in UU, so HH has n/4+1\lfloor n/4\rfloor+1 orbits on UU when nn is even. Note that the smallest nontrivial orbit has size (n2)\binom{n}{2} so, since n5n\geq 5, it follows that the size of nontrivial orbits is larger than the number of orbits. Let xx be the number of orbits of HH on UU and let Δ\Delta be a nontrivial orbit of HH on UU. We have just shown that |Δ|x|\Delta|\geq x. It follows that

(1) k(k1)v1x|Δ||Δ|2.\displaystyle k(k-1)\leq v-1\leq x|\Delta|\leq|\Delta|^{2}.

Let \ell be a line meeting uu and Δ\Delta. By Proposition 4.2, |Δ||\ell\cap\Delta| is equal to 11 or |Δ||\Delta|. If |Δ|=|Δ||\ell\cap\Delta|=|\Delta|, then k|Δ|+1k\geq|\Delta|+1, contradicting (1). We may thus assume that |Δ|=1|\ell\cap\Delta|=1. This implies that the orbit of \ell under HH has length |Δ||\Delta|. Since SS is nontrivial, there must be another nontrivial orbit Δ1\Delta_{1} of HH meeting \ell. Repeating the argument above, we find that the orbit of \ell under HH has length |Δ1||\Delta_{1}|, so |Δ|=|Δ1||\Delta|=|\Delta_{1}|. One can apply this to Δ\Delta, the smallest nontrivial orbit. As mentioned earlier, it has size (n2)\binom{n}{2} and one can check that it is the only orbit of that size, which contradicts |Δ|=|Δ1||\Delta|=|\Delta_{1}|.

It remains to consider the sporadic cases. We give the argument for (d,H)=(10,M12)(d,H)=(10,{\rm M}_{12}) here; the others are similar. In this case, v=210v=2^{10}. Recall that kk must be an integer such that 3k<v3\leq k<v, k1k-1 divides v1v-1 and k(k1)k(k-1) divides v(v1)v(v-1). We find that kk is one of 44, 1212 or 3232, and the corresponding values for rr are 341341, 9393 and 3333. Let \ell be a line through uu. If Δ\Delta is a nontrivial orbit of HH meeting \ell, then by Proposition 4.2, |Δ||\ell\cap\Delta| is equal to 11 or |Δ||\Delta|. By direct computation, we find that HH has orbit lengths 1,66,66,396,4951,66,66,396,495. If |Δ|=|Δ||\ell\cap\Delta|=|\Delta|, then k1+|Δ|67k\geq 1+|\Delta|\geq 67, a contradiction. It follows that |Δ|=1|\ell\cap\Delta|=1 and the orbit of \ell under HH has length |Δ||\Delta|. Repeating this argument for other lines through uu, we conclude that rr is a linear combination of the nontrivial orbit lengths of HH, which is a contradiction.

4.3. Groups arising in Theorem 4.4 (1).

It remains to deal with the groups arising in Theorem 4.4 (1). Let GG, pp, dd, tt, ee be as in Theorem 1.1. By Lemma 3.5, GG has Property ()(\star) and it follows by Proposition 3.2 that 𝒮(G)\operatorname{\mathcal{LS}}(G) is a linear space.

If e=1e=1, then Theorem 1.2 (3) holds. We thus assume that e2e\geq 2. Our next goal is to show that (S,G)(S,G) is transverse. Let (u,)(u,\ell) be a flag of SS and Δ\Delta be a nontrivial orbit of GuG_{u}. By Theorem 2.4, |Δ|=t|\Delta|=t. By Proposition 4.2, |Δ||\ell\cap\Delta| is equal to one of 0, 11 or tt. In view of a contradiction, we can assume that |Δ|=t|\ell\cap\Delta|=t. This implies that kt+1k\geq t+1 and, by Fisher’s inequality, rt+1r\geq t+1. Since e2e\geq 2, we have

pd1=v1=r(k1)>t2=(pd1pd/e1)2(pd1pd/21)2=(pd/2+1)2,p^{d}-1=v-1=r(k-1)>t^{2}=\left(\frac{p^{d}-1}{p^{d/e}-1}\right)^{2}\geq\left(\frac{p^{d}-1}{p^{d/2}-1}\right)^{2}=\left(p^{d/2}+1\right)^{2},

which is a contradiction. This concludes the proof that (S,G)(S,G) is transverse. We can now apply Proposition 3.4 to conclude that SS is a refinement of 𝒮(G)\operatorname{\mathcal{LS}}(G), as in Theorem 1.2 (2).

5. Refinements of line-transitive spaces

In this section, we give a construction for refinements of line-transitive linear spaces such that the line-transitive group also acts on the refined space. We also show that all such refined spaces arise in this way.

Given a line \ell of a linear space S=(𝒫,)S=(\mathcal{P},\mathcal{L}), we write 𝒫()\mathcal{P}(\ell) for the set of points of SS incident with \ell. (In most of this paper, we simply identify \ell with 𝒫()\mathcal{P}(\ell), but we avoid this in this section to reduce possible confusion.) Given GSym(Ω)G\leq\mathrm{Sym}(\Omega), we write GG_{\ell} and G[]G_{[\ell]} for the subgroup of GG preserving 𝒫()\mathcal{P}(\ell) setwise and pointwise, respectively. We also write GG_{\ell}^{\ell} for the permutation group induced by GG_{\ell} on 𝒫()\mathcal{P}(\ell).

Construction 5.1.

The input of the construction is the following:

  1. (1)

    A linear space S=(𝒫,)S=(\mathcal{P},\mathcal{L}) with GAut(S)G\leq\operatorname{Aut}(S) such that GG is transitive on \mathcal{L}, and a line \ell\in\mathcal{L}.

  2. (2)

    A linear space T=(𝒫(),𝒯)T=(\mathcal{P}(\ell),\mathcal{TL}) such that GAut(T)G_{\ell}^{\ell}\leq\operatorname{Aut}(T).

The output of the construction is an incidence structure R=(𝒫,,G,,𝒯)R=\operatorname{\mathcal{R}}(\mathcal{P},\mathcal{L},G,\ell,\mathcal{TL}). The set of points of RR is 𝒫\mathcal{P} while the set of lines of RR is {tgt𝒯,gG}\{t^{g}\mid t\in\mathcal{TL},g\in G\}.

Proposition 5.2.

Using the notation of Construction 5.1, the output RR of the construction is a linear space which is a refinement of SS and with GAut(R)G\leq\operatorname{Aut}(R).

Proof.

It is clear from the construction that GAut(R)G\leq\operatorname{Aut}(R). We first show that RR is a linear space. Let u,v𝒫u,v\in\mathcal{P}, uvu\neq v. There exists uv\ell_{uv}\in\mathcal{L} such that u,vuvu,v\in\ell_{uv}. Since GG is transitive on \mathcal{L}, there exists gGg\in G such that uvg=\ell_{uv}^{g}=\ell, so ug,vg𝒫()u^{g},v^{g}\in\mathcal{P}(\ell). Since TT is a linear space, there is t𝒯t\in\mathcal{TL} such that ug,vgtu^{g},v^{g}\in t so tg1t^{g^{-1}} is a line of RR containing uu and vv. Now, let kk be a line of RR containing ugu^{g} and vgv^{g}. By definition, there exists hGh\in G such that kh𝒯k^{h}\in\mathcal{TL}, so ugh,vgh𝒫()u^{gh},v^{gh}\in\mathcal{P}(\ell). Since \ell is the unique line of \mathcal{L} containing ugu^{g} and vgv^{g}, we have h=\ell^{h}=\ell, hence hGh\in G_{\ell} and hGAut(T)h^{\ell}\in G_{\ell}^{\ell}\leq\operatorname{Aut}(T). It follows that hh^{\ell} preserves 𝒯\mathcal{TL} and thus k𝒯k\in\mathcal{TL}. Since TT is a linear space, it follows that there is a unique line of RR containing ugu^{g} and vgv^{g} (namely kk), and the same holds for uu and vv. We have shown that u,vu,v are on a unique line in RR so RR is a linear space.

We now show that RR is a refinement of SS. Let kk be a line of RR. By definition, there exist t𝒯t\in\mathcal{TL} and gGg\in G such that k=tgk=t^{g}. By definition, tt\subseteq\ell hence k=tggk=t^{g}\subseteq\ell^{g}\in\mathcal{L}, as required. ∎

Proposition 5.3.

Let S=(𝒫,)S=(\mathcal{P},\mathcal{L}), GG and \ell be as in Construction 5.1 (1). If RR is a refinement of SS with GAut(R)G\leq\operatorname{Aut}(R), then there exists a linear space T=(𝒫(),𝒯)T=(\mathcal{P}(\ell),\mathcal{TL}) with GAut(T)G_{\ell}^{\ell}\leq\operatorname{Aut}(T) such that R=(𝒫,,G,,𝒯)R=\operatorname{\mathcal{R}}(\mathcal{P},\mathcal{L},G,\ell,\mathcal{TL}).

Proof.

Write R=(𝒫,)R=(\mathcal{P},\mathcal{RL}) and let 𝒯={kk}\mathcal{TL}=\{k\in\mathcal{RL}\mid k\subseteq\ell\}. We first show that T=(𝒫(),𝒯)T=(\mathcal{P}(\ell),\mathcal{TL}) is a linear space. Let uu and vv be distinct elements of 𝒫()\mathcal{P}(\ell). These points must be contained in a unique line of SS, which must necessarily be \ell. Moreover, they must also be contained in a unique line of RR, say kk\in\mathcal{RL}. Since RR is a refinement of SS, kk\subseteq\ell so k𝒯k\in\mathcal{TL}. If kk^{\prime} is a line in 𝒯\mathcal{TL} containing uu and vv, then by definition kk^{\prime}\in\mathcal{RL} and k=kk=k^{\prime} since RR is a linear space. This shows that TT is a linear space. By definition, GSym(𝒫())G_{\ell}^{\ell}\leq\operatorname{Sym}(\mathcal{P}(\ell)). Since GAut(R)G\leq\operatorname{Aut}(R), it follows that GAut(T)G_{\ell}^{\ell}\leq\operatorname{Aut}(T).

It remains to show that R=(𝒫,)=(𝒫,,G,,𝒯)R=(\mathcal{P},\mathcal{RL})=\operatorname{\mathcal{R}}(\mathcal{P},\mathcal{L},G,\ell,\mathcal{TL}). These have the same set of points so it remains to show that ={tgt𝒯,gG}\mathcal{RL}=\{t^{g}\mid t\in\mathcal{TL},g\in G\}. Let t𝒯t\in\mathcal{TL} and gGg\in G. By definition, tt\in\mathcal{RL} but GAut(R)G\leq\operatorname{Aut}(R), so tgt^{g}\in\mathcal{RL}. This shows that {tgt𝒯,gG}\{t^{g}\mid t\in\mathcal{TL},g\in G\}\subseteq\mathcal{RL}. For the other direction, let tt\in\mathcal{RL}. Since RR is a refinement of SS, there exists kk\in\mathcal{L} such that tkt\subseteq k and, since GG is transitive on \mathcal{L}, there exists gg such that kg=k^{g}=\ell. Now, tgkg=t^{g}\subseteq k^{g}=\ell and tgt^{g}\in\mathcal{RL} so by definition tg𝒯t^{g}\in\mathcal{TL}, as required. ∎

In light of Propositions 3.4 and 5.3, it will be important to be able to determine GG_{\ell}^{\ell}, especially in the case of 𝒮(G)\operatorname{\mathcal{LS}}(G). This is the content of our next two results, which will be useful in Section 6. (For HGH\leq G, we denote the normaliser of HH in GG by NG(H)\operatorname{N}_{G}(H).)

Lemma 5.4.

Let GSym(Ω)G\leq\mathrm{Sym}(\Omega) and u,vΩu,v\in\Omega. If ={wΩGuvGw}\ell=\{w\in\Omega\mid G_{uv}\leq G_{w}\}, then the following statements hold:

  1. (1)

    G[]=GuvG_{[\ell]}=G_{uv};

  2. (2)

    G=NG(Guv)G_{\ell}=\operatorname{N}_{G}(G_{uv});

  3. (3)

    GNG(Guv)/GuvG_{\ell}^{\ell}\cong\operatorname{N}_{G}(G_{uv})/G_{uv}.

Proof.

It follows from the definition of \ell that u,vu,v\in\ell and G[]=GuvG_{[\ell]}=G_{uv}, establishing (1). If gNG(Guv)g\in\operatorname{N}_{G}(G_{uv}) and ww\in\ell, then GuvGwG_{uv}\leq G_{w} which implies Guv=(Guv)gGwg=GwgG_{uv}=(G_{uv})^{g}\leq G_{w}^{g}=G_{w^{g}}, so wgw^{g}\in\ell. This shows NG(Guv)G\operatorname{N}_{G}(G_{uv})\leq G_{\ell}. In the other direction, if gGg\in G_{\ell}, then (G[])=G[]g=G[](G_{[\ell]})=G_{[\ell]^{g}}=G_{[\ell]} so gNG(G[])=NG(Guv)g\in\operatorname{N}_{G}(G_{[\ell]})=\operatorname{N}_{G}(G_{uv}). This completes the proof of (2). Finally, by the first isomorphism theorem, GG/G[]G_{\ell}^{\ell}\cong G_{\ell}/G_{[\ell]} and (3) follows. ∎

Recall that a permutation group is semiregular if all its point stabilisers are trivial and regular if it is also transitive.

Lemma 5.5.

Let SS be a linear space with GAut(S)G\leq\operatorname{Aut}(S) and let (u,)(u,\ell) be a flag of SS. If (S,G)(S,G) is transverse, then GuG=G[]G_{u}\cap G_{\ell}=G_{[\ell]} and in particular GG_{\ell}^{\ell} is semiregular.

Proof.

Clearly G[]GuGG_{[\ell]}\leq G_{u}\cap G_{\ell} so we must show that GuGG_{u}\cap G_{\ell} fixes every point of \ell. It clearly fixes uu, so let vv be another point of \ell. Let Δ=vGu\Delta=v^{G_{u}}. Since (S,G)(S,G) is transverse, we have Δ={v}\ell\cap\Delta=\{v\}, but Δ\ell\cap\Delta is preserved by GuGG_{u}\cap G_{\ell} so vv is fixed. The second statement follows from the first simply by unpacking the definitions involved. ∎

6. Questions and examples

Our proof of cases (1) and (2) of Theorem 1.2 both involved showing that GG has Property ()(\star), that (S,G)(S,G) is transverse and then applying Proposition 3.4. The first part of this approach still works in case (3), as we showed in Section 4.3 that GG has Property ()(\star) even in this case, but there are at least two other issues. First, in case (3), the stabiliser of two distinct points is trivial, so (G)\mathcal{L}(G) is the trivial linear space with a single line. Knowing that SS is a refinement of (G)\mathcal{L}(G) when (S,G)(S,G) is transverse therefore gives no information. The second problem is that, unlike in cases (1) and (2), (S,G)(S,G) need not be transverse, as the following examples show.

Example 6.1.

Let ={0,1,3,9}13\ell=\{0,1,3,9\}\subseteq\operatorname{\mathbb{Z}}_{13}. This is a perfect difference set so its translates form the lines of a linear space SS with point-set 13\operatorname{\mathbb{Z}}_{13}. By definition, 13Aut(S)\operatorname{\mathbb{Z}}_{13}\leq\operatorname{Aut}(S). Let α\alpha be the permutation of 13\operatorname{\mathbb{Z}}_{13} corresponding to “multiplication by 33”. Note that α=\ell^{\alpha}=\ell which implies that αAut(S)\alpha\in\operatorname{Aut}(S) hence G=13αAut(S)G=\operatorname{\mathbb{Z}}_{13}\rtimes\langle\alpha\rangle\leq\operatorname{Aut}(S). It is easy to see that GG is extremely primitive on 13\operatorname{\mathbb{Z}}_{13}. Let u=0u=0 and Δ={1,3,9}\Delta=\{1,3,9\}. Note that Gu=αG_{u}=\langle\alpha\rangle and that Δ\Delta is an orbit of GuG_{u} with |Δ|=3|\ell\cap\Delta|=3 so (S,G)(S,G) is not transverse. (Note that here v=13v=13 and k=4k=4 so SS is in fact the unique projective plane of order 33 and its automorphism group is much bigger than GG.)

Note that, in Example 6.1, GG is line-transitive (by construction). We were not able to find any other such example and wonder if any exist. More precisely:

Question 6.2.

Let SS be a nontrivial linear space with GAut(S)G\leq\operatorname{Aut}(S) such that GG that is extremely primitive on points, transitive on lines and such that (S,G)(S,G) is not transverse. Does it follow that SS is the projective plane of order 33?

As mentioned at the start of this section, in cases (1) and (2) of Theorem 1.2 (S,G)(S,G) is transverse so an example for Question 6.2 must arise from case (3), that is CpdCtGAGL(1,pd)\operatorname{C}_{p}^{d}\rtimes\operatorname{C}_{t}\cong G\leq\operatorname{AGL}(1,p^{d}) for tt a primitive prime divisor of pd1p^{d}-1. We have checked using GAP that there is no other example with fewer than 1000 points. On the other hand, there seems to be plenty of examples which are not line-transitive:

Example 6.3.

Let GG be PrimitiveGroup(25,1) in GAP. This group is generated by the following two permutations:

(2,19,6)(3,25,11)(4,7,16)(5,13,21)(8,24,9)(10,15,14)(12,17,20)(18,23,22)and\displaystyle(2,19,6)(3,25,11)(4,7,16)(5,13,21)(8,24,9)(10,15,14)(12,17,20)(18,23,22){~{}\text{a}nd}
(1,2,3,5,4)(6,7,8,10,9)(11,12,13,15,14)(16,17,18,20,19)(21,22,23,25,24).\displaystyle(1,2,3,5,4)(6,7,8,10,9)(11,12,13,15,14)(16,17,18,20,19)(21,22,23,25,24).

One can check that GG is extremely primitive and C52C3GAGL1(52)\operatorname{C}_{5}^{2}\rtimes\operatorname{C}_{3}\cong G\leq\mathrm{AGL}_{1}(5^{2}). Now, let 1={1,2,6,19}\ell_{1}=\{1,2,6,19\}, 2={1,3,11,25}\ell_{2}=\{1,3,11,25\}, =1G2G\mathcal{L}=\ell_{1}^{G}\cup\ell_{2}^{G} and S=({1,,25},)S=(\{1,\ldots,25\},\mathcal{L}). One can check that SS is a linear space with (v,k,r,b)=(25,4,8,50)(v,k,r,b)=(25,4,8,50). By construction, GAut(S)G\leq\operatorname{Aut}(S) and it turns out that GG has two orbits on lines, with representatives 1\ell_{1} and 2\ell_{2}. Note that Δi=i{1}\Delta_{i}=\ell_{i}\setminus\{1\} is an orbit of G1G_{1} with |Δii|=3|\Delta_{i}\cap\ell_{i}|=3 and again (S,G)(S,G) is not transverse. As a final remark, we note that Aut(S)\operatorname{Aut}(S) actually is isomorphic to PrimitiveGroup(25,3) and this group is not extremely primitive, nor transitive on lines.

Examples 6.1 and 6.3 show that the classification of linear spaces arising from Theorem 1.2 (3) is more complicated than in cases (1) and (2). This is not that surprising, since in case (3) GG is very “small” (its point stabiliser has prime order) so the restriction GAut(S)G\leq\operatorname{Aut}(S) is much weaker.

We now describe 𝒮(G)\operatorname{\mathcal{LS}}(G) from Theorem 1.2 (1) and (2) in more detail.

Example 6.4.

Let q=22n+1q=2^{2^{n}}+1 be a Fermat prime and G=PSL2(q1)G=\operatorname{PSL}_{2}(q-1) with point stabiliser D2q\operatorname{D}_{2q}, as in Theorem 1.2 (1). As observed in Section 4, the two-point stabilisers in GG have order 22; hence we may identify each line \ell of 𝒮(G)\mathcal{LS}(G) with the involution gg_{\ell} that fixes any two points on it. By identifying points of 𝒮(G)\mathcal{LS}(G) with their point stabilisers, we see that 𝒮(G)\mathcal{LS}(G) is the Witt-Bose-Shrikhande space W(q1)W(q-1) [1, §2.6] with parameters

(v,b,k,r)=((q1)(q2)2,q(q2),q12,q).(v,b,k,r)=\left(\frac{(q-1)(q-2)}{2},q(q-2),\frac{q-1}{2},q\right)\!.

By Lemma 5.4, G=NG(g)G_{\ell}=\operatorname{N}_{G}(\langle g_{\ell}\rangle), which is elementary abelian of order q1q-1, while GNG(g)/gG_{\ell}^{\ell}\cong\operatorname{N}_{G}(\langle g_{\ell}\rangle)/\langle g_{\ell}\rangle, an elementary abelian group of order q12\frac{q-1}{2}, which is semiregular by Lemma 5.5. Since q12=k\frac{q-1}{2}=k, it follows that GG_{\ell}^{\ell} is in fact regular. By the orbit-stabiliser theorem, the orbit of \ell under GG has size |G|/|G|=q(q2)=b|G|/|G_{\ell}|=q(q-2)=b, so GG is transitive on lines.

Example 6.5.

Let pp be a prime, d2d\geq 2, tt be a primitive prime divisor of pd1p^{d}-1 and ee be a prime dividing dd such that t=(pd1)/(pd/e1)t=(p^{d}-1)/(p^{d/e}-1). Let GAΓL1(pd)G\leq\operatorname{A\Gamma L}_{1}(p^{d}) with point stabiliser H=CtCeΓL1(pd)H=\operatorname{C}_{t}\rtimes\operatorname{C}_{e}\leq\operatorname{\Gamma L}_{1}(p^{d}), as in Theorem 1.2 (2). A two-point stabiliser in GG is conjugate to the group Ce\operatorname{C}_{e} which acts as field automorphisms on 𝔽pd\mathbb{F}_{p^{d}} and hence fixes the subfield of order pd/ep^{d/e}, so 𝒮(G)\mathcal{LS}(G) has parameters

(v,b,k,r)=(pd,pdd/et,pd/e,t).(v,b,k,r)=(p^{d},p^{d-d/e}t,p^{d/e},t).

By Lemma 5.4, G=NG(Ce)=Cpd/e×CeG_{\ell}=\operatorname{N}_{G}(\operatorname{C}_{e})=\operatorname{C}_{p}^{d/e}\times\operatorname{C}_{e}, while GNG(Ce)/CeCpd/eG_{\ell}^{\ell}\cong\operatorname{N}_{G}(\operatorname{C}_{e})/\operatorname{C}_{e}\cong\operatorname{C}_{p}^{d/e}. Since k=pd/ek=p^{d/e}, we again obtain that GG_{\ell}^{\ell} is regular. Finally, the orbit of \ell under GG has size |G|/|G|=pdte/(pd/ee)=b|G|/|G_{\ell}|=p^{d}te/(p^{d/e}e)=b, so GG is transitive on lines.

Note that in both previous examples, GG is transitive on lines of 𝒮(G)\mathcal{LS}(G) so by Proposition 5.3, all the refinements of 𝒮(G)\mathcal{LS}(G) admitting GG as a group of automorphisms arise via Construction 5.1. We now present two final examples, which give nontrivial such refinements and thus are also examples for Theorem 1.2.

Example 6.6.

Let (p,d,e,t)=(7,5,5,2801)(p,d,e,t)=(7,5,5,2801), let GG be as in Example 6.5 and let S=𝒮(G)S=\mathcal{LS}(G). If \ell is a line of SS, then ||=7|\ell|=7 and GC7G_{\ell}^{\ell}\cong\operatorname{C}_{7} is regular. If we set T=(𝒫(),𝒯)T=(\mathcal{P}(\ell),\mathcal{TL}) to be the Fano plane, then we have C7Aut(T)\operatorname{C}_{7}\leq\operatorname{Aut}(T). By Proposition 5.2, the output RR of Construction 5.1 is a linear space on 757^{5} points with lines of size 33 which is a refinement of SS and with GAut(R)G\leq\operatorname{Aut}(R). Since C7C_{7} is transitive on 𝒯\mathcal{TL}, it follows that GG is transitive on the lines of RR.

Example 6.7.

Let qq be the Fermat prime 65537=216+165537=2^{16}+1, let GG be as in Example 6.4 and let S=𝒮(G)S=\mathcal{LS}(G). If \ell is a line of SS, then ||=215|\ell|=2^{15} and GC215G_{\ell}^{\ell}\cong\operatorname{C}_{2}^{15} is regular. Let HΓL1(215)H\leq\operatorname{\Gamma L}_{1}(2^{15}) such that |HGL1(215)|=1057|H\cap\operatorname{GL}_{1}(2^{15})|=1057 and H=C1057C3H=\operatorname{C}_{1057}\rtimes\operatorname{C}_{3}. Let VC215V\cong\operatorname{C}_{2}^{15} be the natural vector space for ΓL1(215)\operatorname{\Gamma L}_{1}(2^{15}) and let A=VHAΓL1(215)A=V\rtimes H\leq\operatorname{A\Gamma L}_{1}(2^{15}). Note that Proposition 2.2 (3) is satisfied (with (p,d,e,t)=(2,15,3,1057)(p,d,e,t)=(2,15,3,1057)), hence Lemma 3.5 implies that AA has Property ()(\star) and by Proposition 3.2, T:=𝒮(A)T:=\mathcal{LS}(A) is a linear space such that C215VAAut(T)\operatorname{C}_{2}^{15}\cong V\leq A\leq\operatorname{Aut}(T). By Proposition 5.2, the output RR of Construction 5.1 is a linear space which is a refinement of SS and with GAut(R)G\leq\operatorname{Aut}(R). By similar calculations as in Example 6.5 we deduce that each line contains k:=pd/e=32k:=p^{d/e}=32 points and thus RR has parameters

(v,k,r)=(215(2161),32,qt).(v,k,r)=(2^{15}(2^{16}-1),32,qt).

Since (S,G)(S,G) is transverse, so is (R,G)(R,G). In particular, if (u,m)(u,m) is a flag of RR and uvu\neq v\in\ell, then |mGu|=|vGu|=|Gu|/|Guv|=2q/2=q|m^{G_{u}}|=|v^{G_{u}}|=|G_{u}|/|G_{uv}|=2q/2=q. It follows that GuG_{u} has r/q=t=1057r/q=t=1057 orbits on lines meeting uu. Note that GG is smaller than the number of lines of RR, so GG is not transitive on lines of RR.

Example 6.7 answers a question posed implicitly in [10] about the existence of regular linear spaces that admit an extremely primitive automorphism group with classical socle, other than the Witt-Bose-Shrikhande spaces. In [10, Lemma 3.2], Guan and Zhou show that in that case GG is as in Example 6.4, that qq is at least 6553765537 (the largest known Fermat prime) and that GuG_{u} has at least 7373 orbits on lines meeting uu, but are unable to determine if any examples actually arise. Example 6.7 gives one such example and our approach using refinements can be used to construct more.

References

  • [1] F. Buekenhout, A. Delandtsheer, and J. Doyen. Finite linear spaces with flag-transitive groups. J. Combin. Theory Ser. A, 49(2):268–293, 1988.
  • [2] T. C. Burness and M. Lee. On the classification of extremely primitive affine groups. (arXiv:2107.04302), 2021.
  • [3] T. C. Burness, C. E. Praeger, and Á. Seress. Extremely primitive classical groups. J. Pure Appl. Algebra, 216(7):1580–1610, 2012.
  • [4] T. C. Burness, C. E. Praeger, and Á. Seress. Extremely primitive sporadic and alternating groups. Bull. Lond. Math. Soc., 44(6):1147–1154, 2012.
  • [5] T. C. Burness and A. R. Thomas. A note on extremely primitive affine groups. Arch. Math. (Basel), 116(2):141–152, 2021.
  • [6] T. C. Burness and A. R. Thomas. The classification of extremely primitive groups. Int. Math. Res. Not. IMRN, (13):10148–10248, 2022.
  • [7] A. R. Camina, P. M. Neumann, and C. E. Praeger. Alternating groups acting on finite linear spaces. Proc. London Math. Soc. (3), 87(1):29–53, 2003.
  • [8] The GAP Group. GAP – Groups, Algorithms, and Programming, Version 4.9.3, 2018.
  • [9] M. Giudici, M. W. Liebeck, C. E. Praeger, J. Saxl, and P. H. Tiep. Arithmetic results on orbits of linear groups. Trans. Amer. Math. Soc., 368(4):2415–2467, 2016.
  • [10] H. Guan and S. Zhou. Extremely primitive groups and linear spaces. Czechoslovak Math. J., 66(141)(2):445–455, 2016.
  • [11] P. Kleidman and M. Liebeck. The subgroup structure of the finite classical groups, volume 129 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 1990.
  • [12] A. Mann, C. E. Praeger, and Á. Seress. Extremely primitive groups. Groups Geom. Dyn., 1(4):623–660, 2007.
  • [13] W. A. Manning. Simply transitive primitive groups. Trans. Amer. Math. Soc., 29(4):815–825, 1927.
  • [14] D. V. Pasechnik and C. E. Praeger. On transitive permutation groups with primitive subconstituents. Bull. London Math. Soc., 31(3):257–268, 1999.
  • [15] D. S. Passman. Solvable 3/2{3/2}-transitive permutation groups. J. Algebra, 7:192–207, 1967.
  • [16] D. S. Passman. Exceptional 3/23/2-transitive permutation groups. Pacific J. Math., 29:669–713, 1969.