365
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Centralisers of finite groups in locally finite simple groups

Pages 4795-4803 | Received 14 Oct 2022, Accepted 17 May 2023, Published online: 31 May 2023

Abstract

We answer in the negative a question of Hartley about representations of finite groups, by constructing examples of finite simple groups with arbitrarily large representations whose endomorphism ring consists of just the scalars. We show as a consequence that there are finite simple groups of automorphisms of the locally finite simple group SL(,Fq) with trivial centralizer. The smallest of our examples is A6 with q=9.

2020 Mathematics Subject Classification:

1 Introduction

A group is said to be locally finite if every finitely generated subgroup is finite. A group is said to be linear if it has a faithful representation by matrices over a field.

In the introduction to [Citation7], and Section 3 of [Citation8], Brian Hartley mentioned the following problem.

Question 1.1.

Is it the case that in a nonlinear locally finite simple group, the centralizer of every finite subgroup is infinite?

Papers that have investigated Hartley’s Question 1.1 include Brescia and Russo [Citation5], Ersoy and Kuzucuoğlu [Citation6], Hartley and Kuzucuoğlu [Citation10], Kuzucuoğlu [Citation12–15]. In particular, it is shown in [Citation10] that if the finite subgroup is cyclic then the answer is “yes.”

A closely related problem is Problem 3.15 of Hartley [Citation9], which asks the following (with notation slightly altered to align with ours).

Question 1.2.

Let H be a finite simple group, or perhaps a finite group such that Op(H)=1. Suppose that HPSL(n,Fq)=K, where q is a power of p and CK(H)=1. Does it follow that n is bounded in terms of |H|? We are particularly interested in the case when H is also a projective special linear group over Fq. What about the algebraic group situation, that is, say, HPSLm(F¯p), KPSLn(F¯p), with CK(H)=1? Does it follow that n is bounded in terms of m?

An example of a non-linear simple locally finite group is SL(,Fq), the union of the groups SL(n,Fq) under the natural inclusions SL(n,Fq)SL(n+1,Fq) fixing the new basis vector. It is simple because (except for n=2, q=2 or 3) every proper normal subgroup of SL(n,Fq) consists of scalar multiples of the identity, and are no longer scalar in SL(n+1,Fq) unless trivial. In this group, the centralizer of a finite subgroup is obviously infinite.

We do not succeed in answering Question 1.1, but we answer a closely related question with the following theorem.

Theorem 1.3.

Let Kq be the locally finite group SL(,Fq). For every prime p, there exists a power q of p and a finite simple group of automorphisms HAut(Kq) such that the centralizer of H in Kq is trivial.

We prove Theorem 1.3 as a consequence of the following theorem, together with some information about extensions between simple modules. This theorem gives a negative answer to Question 1.2, as is shown in Section 3. We restrict our attention to finite groups and linear algebraic groups in order to address Hartley’s question, but it should be clear from the methods that other situations are also covered.

Theorem 1.4.

Let k be a field of characteristic p. Let H be a finite group, or a linear algebraic group over k. Suppose that

  1. H has non-isomorphic simple modules ST over k such that EndH(S)EndH(T)k, and dimkExtH1(S,T)3, or

  2. H has non-isomorphic simple modules R, S and T with EndH(R)EndH(S)EndH(T)k, dimkExtH1(R,T)1 and dimkExtH1(S,T)2, or

  3. H has non-isomorphic simple modules R, S1,,Sr, and T1,,Tr with endomorphism rings isomorphic to k, and such that the groups ExtH1(R,T1), ExtH1(Si,Ti), ExtH1(Si+1,Ti) and ExtH1(S1,Tr) are all non-trivial.

Then there exists a sequence of indecomposable H-modules Mm over k and embeddings MmMm+1 such that

  1. The radical (i.e., the intersection of the maximal submodules) Rad(Mm) is isomorphic to a direct sum of m copies of T (respectively of a direct sum of the Ti in case (iii)).

  2. Mm/Rad(Mm) is isomorphic to a direct sum of m copies of S in case (i) and a direct sum of one copy of R and m1 copies of S in case (ii).

  3. The endomorphism ring EndH(Mm) consists just of scalar multiples of the identity map.

The three cases of Theorem 1.4 are proved in Sections 2, 4, and 6, and examples illustrating these three cases are given in Sections 3, 5, and 7. Theorem 1.3 is proved in Section 8.

2 Three dimensional Ext1 groups Three dimensional Ext1 groups

In this section, we deal with case (i) of Theorem 1.4. Let k be a field, and let Q be the quiver with two vertices and three arrows given by the following diagram.

A representation of Q over k consists of two k-vector spaces V and W together with three k-linear maps X,Y,Z:VW. We are interested in properties of a particular family of representations Mm of kQ for m1 and embeddings MmMm+1.

For the representation Mm, the spaces V and W have dimension m, and bases v1,,vm and w1,,wm. The actions of X, Y, and Z on vi are given as follows.(2.1) XYZi=1w1002⩽i⩽m0wiwi1(2.1)

This can be pictured as follows.

Here, the actions of X, Y, and Z are described by the single, double and triple downward edges respectively. The embedding MmMm+1 takes the basis vectors vi and wi of Mm to the basis elements of Mm+1 with the same names.

Theorem 2.2.

The endomorphism ring of the kQ-module Mm is equal to k, acting as scalar multiples of the identity map.

Proof.

We have direct sum decompositions V=Ker(Y)Ker(X) and W=Im(X)Im(Y). Here, Ker(Y) is spanned by v1 and Ker(X) is spanned by v2,,vm. Similarly, Im(X) is spanned by w1 and Im(Y) is spanned by w2,,wm.

If α is an endomorphism of Mm then Xα(v1)=α(Xv1)=α(w1), so α(w1) is in the image of X. It is therefore a multiple of w1. Subtracting off a multiple of the identity map, we can suppose that α(w1)=0. Then α(v1) is in Ker(X)Ker(Y) and is hence also zero.

Suppose by induction on j that we have shown that we have α(vi)=0 and α(wi)=0 for 1i<j. Then Yα(vj)=α(Yvj)=α(wj1)=0 and Xα(vj)=α(Xvj)=0, so α(vj)Ker(Y)Ker(X) and hence α(vj)=0. Then α(wj)=α(Yvj)=Yα(vj)=0. It follows by induction that for 1⩽j⩽m we have α(vj)=0 and α(wj)=0, and hence α=0. ▪

Proof of Theorem 1.4

in case (i).

Suppose that k is a field of characteristic p and H is a finite group, or a reductive algebraic group, with non-isomorphic simple (left) modules S and T satisfying EndH(S)=EndH(T)=k and dimkExtH1(S,T)3. Then there is an H-module Δ satisfying Rad(Δ)TTT and Δ/Rad(Δ)S. Setting B=ΔT, the algebra EndH(B)op is isomorphic to the quiver algebra kQ described above. So B has a right action of kQ commuting with the left action of H. As a right kQ-module, B is a projective generator for mod-kQ. Thus B is a H-kQ-bimodule with the property that the functor BkQ:kQ-modH-mod is fully faithful and exact. It sends the simple kQ-modules to S and T and the three arrows in J(kQ) to three linearly independent elements of ExtH1(S,T). So Mm=BkQMm is a module of the following form:where the three types of lines represent the three linearly independent extension classes chosen in the construction of U. By Theorem 2.2, we have EndH(Mm)k and Mm embeds in Mm+1. Setting d=dimkS+dimkT, we have dimkMm=md. This completes the proof of Theorem 1.4. ▪

3 Examples involving three dimensional Ext1 groups Examples involving three dimensional Ext1 groups

In this section, we indicate that there are many examples of absolutely irreducible modules S and T over reductive algebraic groups and over finite simple groups H, such that ExtkH1(S,T) is at least three dimensional, and sometimes much larger. The first such examples were discovered by Scott, and already, feeding these into Case (i) of Theorem 1.4 gives negative answers to both parts of Question 1.2.

Theorem 3.1.

Suppose that n6. Then for all large enough primes p, there exist irreducible PSL(n,F¯p)-modules S and T such that dimF¯pExtPSL(n,F¯p)1(S,T)4.

And for all large enough powers q of p, the restrictions of S and T are absolutely irreducible FqPSL(n,Fq)-modules such that dimFqExtFqPSL(n,Fq)1(S,T)4.

Proof.

This is proved in Section 2 of Scott [Citation19], by computing coefficients of Kazhdan–Lusztig polynomials. The stabilisation of the dimensions of Ext1 for the finite groups to the dimension for the algebraic group comes from Theorem 2.8 of Andersen [Citation1], see Remark (iii) at the end of that section. ▪

Remark 3.2.

It is shown in Section 4 of Lübeck [Citation17] that for all large enough primes, taking S to be the trivial module F¯p, there are simple modules T for PSL(n,F¯p) withH1(PSL(n,F¯p),T)ExtPSL(n,F¯p)1(F¯p,T)having the following dimensions,n6789dimFqH1(PSL(n,F¯p),T)31646936672 and then for large enough powers q of p, the restriction of these T are irreducible modules for FqPSL(n,Fq) with these dimensions for H1(PSL(n,Fq),T). It is not known whether these dimensions are unbounded for larger values of n, though that seems very likely. In the same paper, Lübeck proves similar results for the other classical groups of types Bn, Cn, and Dn, as well as groups of types F4 and E6. These groups could therefore also be used equally well as examples in Theorem 1.3. There are also cross characteristic examples of large dimensional Ext1 for simple modules over groups of Lie type, but the Lie rank in these cases needs to be a lot larger.

The problem with the theorems above is that the phrase “large enough” is difficult to quantify. Some more explicit examples of three dimensional H1(H,T) for T an absolutely irreducible module for a finite simple group H can be found in Bray and Wilson [Citation4], for the group PSU(4,F3) over F3 with dimF3T=19 (easily verified with Magma [Citation3]), and for the group 2E6(F2) over F2 with dimF2T=1702 (not so easily verified).

4 Two dimensional Ext1 groups Two dimensional Ext1 groups

In this section, we deal with case (ii) of Theorem 1.4. The proof is very much like that of case (i). We begin with the quiver Q with three vertices and three arrows as follows.

A representation of kQ over a field k consists of three vector spaces U, V, and W together with three k-linear maps

For the kQ-module Mm, U is spanned by v1, V is spanned by v2,,vm, and W is spanned by w1,,wm. The actions of X, Y, and Z are exactly the same as in Table (2.1).

Theorem 4.1.

The endomorphism ring of the kQ-module Mm is equal to k, acting as scalar multiples of the identity map.

Proof.

The proof is the same as for Theorem 2.2. ▪

Proof of Theorem 1.4

in case (ii).

Suppose that k is a field of characteristic p, and let H be a finite group with non-isomorphic simple modules R, S and T with EndkH(R)=EndkH(S)=EndkH(T)=k, dimkExtkH1(R,T)1, and dimkExtkH1(S,T)2. Then there are kH-modules Δ1 and Δ2 withRad(Δ1)T,Δ1/Rad(Δ1)R,Rad(Δ2)TT,Δ2/Rad(Δ2)S.

Setting B=Δ1Δ2T, the algebra EndkH(B)op is isomorphic to kQ. This makes B a kH-kQ-bimodule with the property that as a right kQ-module it is a projective generator for mod-kQ. The functor BkQ is fully faithful and exact. It sends the three simple kQ-modules to R, S and T. So this time, Mm=BkQMm is a module of the following formwhose endomorphism ring is k, acting by scalar multiples of the identity. ▪

5 Examples involving two dimensional Ext1 groups Examples involving two dimensional Ext1 groups

In this section, we give examples of finite simple groups H and simple modules R, S, and T satisfying the hypotheses of Case (ii) of Theorem 1.4.

Example 5.1.

Let H be the group PSL(2,q) with q=p2, p odd, and let k=Fq. Then by Corollary 4.5 of Andersen, Jørgensen and Landrock, there are exactly two simple modules S and T, of dimensions (p12)2 and (p+12)2, such that ExtkH1(S,T) and ExtkH1(T,S) are two dimensional. All the other Ext1 groups between simple kH-modules are zero or one dimensional. In fact, the entire quiver with relations in this example may be found in Koshita [Citation11]. Since S and T are in the principal block, and are not the only simples in the principal block, we can choose a simple module R not isomorphic to S or T, so that ExtkH1(R,T)1. For example, if p=3 then PSL(2,9)A6 has simples of dimensions 3, 1, and 4 for R, S, and T.

Example 5.2.

Another example is the sporadic group M12 over F2. There are three simple modules R, S, and T in the principal block of F2M12, of dimensions 44, 1, and 10. We have dimF2ExtF2M121(R,T)=1 and dimF2ExtF2M121(S,T)=2, see Schneider [Citation18].

Example 5.3.

The following examples were computed using Magma [Citation3]. For the group PSL(3,F4) over F4, we can take for R, S, and T simple modules of dimensions 8, 9, and 1. For the group PSL(4,F3) over F3, we can take for R, S, and T simple modules of dimensions 19, 1, and 44. For the Higman–Sims group HS over F2, we can take for R, S, and T the simple modules of dimensions 56, 20, and 1.

6 One dimensional Ext1 groups One dimensional Ext1 groups

In some circumstances, our method can be modified to deal with groups where all Ext1 groups between simple modules have dimension zero or one. This technique works whenever there is an even length cycle in the Ext1 quiver with alternating directions, and with some other arrow head meeting two of its arrow heads:

The method described in Section 4 is a degenerate case of this with a cycle of length two. This will lead to Case (iii) of Theorem 1.4.

Let Q be the quiver above, with even cycle length 2r4. Then a representation of Q over k consists of vector spaces U, V1,,Vr and W1,,Wr together with k-linear maps X:UW1, Yi:ViWi, Zi+1:Vi+1Wi and Z1:V1Wr. The case r=3 is illustrated as follows.

The module Mm has basis vectors v1,,vm and w1,,wm. The space U is spanned by v1, Vi is spanned by the vj with 1<ji(modr), and Wi is spanned by the wj with 1ji(modr). The actions of the arrows are as in Table 2.1, with all Yi acting like Y and all Zi acting like Z.

Theorem 6.1.

The endomorphism ring of the kQ-module Mm is equal to k, acting as scalar multiples of the identity map.

Proof.

The proof is the same as for Theorem 2.2. ▪

Proof of Theorem 1.4

in case (iii).

We have kH-modules Δ,Δ1,,Δr withRad(Δ)T1,Δ/Rad(Δ)R,Rad(Δi)TiTi1(or T1Tr if i=1),Δi/Rad(Δi)Si.

The kH-module B is Δi=1rΔii=1rTi, and EndkH(B)kQ. This makes B a kH-kQ-bimodule with the property that as a right kQ-module it is a projective generator for mod-kQ. The functor BkQ is fully faithful and exact. It sends the simple kQ-modules to the corresponding simple kH-modules. So this time, Mm=BkQMm is a module of the following formwhere the indices are taken modulo r. ▪

7 Examples involving one dimensional Ext1 groups Examples involving one dimensional Ext1 groups

In this section, we give examples satisfying Case (iii) of Theorem 1.4.

Example 7.1.

The Ext1 quiver of the Higman–Sims group HS over a finite splitting field k of characteristic three is as follows.

This information comes from Section 2 of Waki [Citation20]. The numbers indicate the dimensions of the corresponding simple modules, and each edge denotes a single arrow in each direction. Choosing the 4-cycle 11253227481 and incoming arrow 1541253, the modules Mm take the following form.

Example 7.2.

Landrock and Michler [Citation16] showed that the Ext1 quiver of the principal block of the sporadic Janko group J1 over the splitting field F4 is as follows.

So we can take the 4-cycle 5621561201 and incoming arrow 761.

Example 7.3.

For H=PSL(4,F2) and k=F2, the Ext1 quiver was computed in Benson [Citation2] to be as follows.

So for example we can take a 4-cycle 16412011 and incoming arrow 146.

Example 7.4.

Computations using Magma [Citation3] exhibit further examples. For the group H=PSL(3,F5) and k=F5, all the Ext1 groups are zero or one dimensional, but the Ext1 quiver has a 4-cycle of the form 19618819 and an incoming arrow 356.

8 The construction

In this section, we show how to pass from the constructions of Sections 2, 4, and 6 and groups satisfying the hypotheses of Theorem 1.4 to examples proving Theorem 1.3. For convenience, we assume that H is a finite simple group, though this could probably be weakened somewhat.

Let k and Q be as in Section 2, 4, or 6. We define the kQ-module M to be the colimit of the sequence of modulesM2M3M4.

This has a basis consisting of the vi and wi with i1. Then there are submodules Mm of M such that MmMm=0, and M/(Mm+Mm) is one dimensional, spanned by the image of vi+1. Namely, we take Mm to be the submodule spanned by the basis vectors vi with im+2 and wi with im+1.

Let M and Mm be the kH-modules defined as the images of M and Mm under BkQ. Thus M is the colimit of the sequence of modulesM2M3M4.

The submodules Mm of M satisfy MmMm=0 and M/(Mm+Mm)S.

We define Gm to be the group consisting of the vector space automorphisms α of M such that α preserves and acts with determinant one on Mm, and there exists hH such that α acts in the same way as h on both M/Mm and Mm. It follows from these conditions that α(Mi)Mi for all im. Then since MmMm+1 and Mm+1Mm, we have GmGm+1. The action of H on M gives an inclusion i:HGm for each m, compatible with these inclusions.

We have group homomorphisms GmH and GmSL(Mm+1) given by the actions on Mm and on Mm+1. Since Mm and Mm+1 span M, the intersection of the kernels of the actions on these spaces is trivial, so we obtain an injective homomorphismGmH×SL(Mm+1).

This is not surjective, because the image of Gm in SL(Mm+1) stabilises Mm setwise; but it properly contains SL(Mm). However, the image contains H×1, and is therefore the direct product of H and the image in SL(Mm+1). The subgroup mapping to H×1 consists of elements acting as hH on Mm and as the identity of M/Mm. This subgroup depends on m, and is not the same as the image of i:HGm. Indeed, the composite HiGmSL(Mm+1) gives the action of H on the module Mm+1.

Lemma 8.1.

The centralizer of the image of i:HGm is trivial.

Proof.

Let z be an element of CGm(i(H)). Then z commutes with the action of H on M, and so z acts as a scalar multiple of the identity. But then z also acts on M/Mm as a multiple of the identity. The action of Gm on M/Mm factors through GmH, and Z(H) is trivial, so z=1. ▪

We define G to be the colimit of the inclusionsG2G3G4.

By Theorem 1.4 (3), we have CGm(H)Z(Gm)=1 and CG(H)Z(G)=1.

Theorem 8.2.

The group G is isomorphic to a semidirect product SL(,k)H.

Proof.

By the definition of Gm, there is a homomorphism GmH, taking an element α to the element hH that acts on M/Mm and Mm in the same way as α. The composite HiGmH is the identity. These homomorphisms are compatible with the inclusions GmGm+1, and therefore describe a homomorphism GH with kernel SL(,k)a normal complement to the inclusion i:HG. ▪

Theorem 8.3.

Given a finite simple group H satisfying condition (i), (ii) or (iii) of Theorem 1.4, there is an action of H on SL(,k) with trivial centralizer.

Proof.

This follows from Theorem 8.2 and Lemma 8.1. ▪

Proof of Theorem 1.3.

We set k=Fq in Theorem 8.3. For p odd, we can use Example 5.1, and for p=2 we can use Example 5.2 for the choice of H; there are, of course, many other possible choices. ▪

Acknowledgments

I would like to thank Alexandre Zalesskii for introducing me to these problems at breakfast at the Møller Institute in Cambridge. This was while attending the post Covid-19 resumption of the programme ‘Groups, representations and applications: new perspectives’ in 2022, at the Isaac Newton Institute, supported by EPSRC grant EP/R014604/1. My thanks also go to the Isaac Newton Institute for their support and hospitality.

References

  • Andersen, H. H. (1987). Extensions of simple modules for finite Chevalley groups. J. Algebra 111:388–403. DOI: 10.1016/0021-8693(87)90225-0.
  • Benson, D. J. (1983). The Loewy structure of the projective indecomposable modules for A8 in characteristic 2. Commun. Algebra 11(13):1395–1432.
  • Bosma, W., Cannon, J., Playoust, C. (1997). The Magma algebra system, I. The user language. J. Symbolic Comput. 24:235–265. DOI: 10.1006/jsco.1996.0125.
  • Bray, J. N., Wilson, R. A. (2008). Examples of 3-dimensional 1-cohomology for absolutely irreducible modules of finite simple groups. J. Group Theory 11(5):669–673.
  • Brescia, M., Russo, A. (2019). On centralizers of locally finite simple groups. Mediterr. J. Math. 16(5):Paper 114, 7pp. DOI: 10.1007/s00009-019-1401-3.
  • Ersoy, K., Kuzucuoğlu, M. (2012). Centralizers of subgroups in simple locally finite groups. J. Group Theory 15:9–22.
  • Hartley, B. (1990). Centralizing properties in simple locally finite groups and large finite classical groups. J. Austral. Math. Soc. 49:502–513. DOI: 10.1017/S1446788700032468.
  • Hartley, B. (1992). Finite and locally finite groups containing a small subgroup with small centralizer. In: Liebeck, M. W., Saxl, J., eds. Groups, combinatorics and geometry, Durham 1990, London Mathematical Society Lecture Note Series, Vol. 165. Cambridge: Cambridge University Press, pp. 397–402.
  • Hartley, B. (1995). Simple locally finite groups. In: Hartley, B., Seitz, G. M., Borovik, A. V., Bryant, R. M., eds. Finite and Locally Finite Groups. Berlin/New York: Springer-Verlag, pp. 1–44.
  • Hartley, B., Kuzucuoğlu, M. (1991). Centralizers of elements in locally finite simple groups. Proc. London Math. Soc. 62(2):301–324.
  • Koshita, H. (1998). Quiver and relations for SL(2,pn) in characteristic p with p odd. Commun. Algebra 26(3):681–712.
  • Kuzucuoğlu, M. (1994). Centralizers of semisimple subgroups in locally finite simple groups. Rend. Sem. Mat. Univ. Padova 92:79–90.
  • Kuzucuoğlu, M. (1997). Centralizers of abelian subroups in locally finite simple groups. Proc. Edinb. Math. Soc. 40:217–225.
  • Kuzucuoğlu, M. (2013). Centralizers in simple locally finite groups. Int. J. Group Theory 2(1):1–10.
  • Kuzucuoğlu, M. (2017). Centralizers of finite p-subgroups in simple locally finite groups. J. Siberian Fed. Univ. Math. Phys. 10(3):281–286.
  • Landrock, P., Michler, G. O. (1978). Block structure of the smallest Janko group. Math. Ann. 232:205–238. DOI: 10.1007/BF01351427.
  • Lübeck, F. (2020). Computation of Kazhdan–Lusztig polynomials and some applications to finite groups. Trans. Amer. Math. Soc. 373(4):2331–2347. DOI: 10.1090/tran/8037.
  • Schneider, G. J. A. (1993). The structure of the projective indecomposable modules of the Mathieu group M12 in characteristic 2 and 3. Arch. Math. (Basel) 60:321–326.
  • Scott, L. L. (2003). Some new examples in 1-cohomology. J. Algebra 260(1):416–425. DOI: 10.1016/S0021-8693(02)00667-1.
  • Waki, K. (1993). The projective indecomposable modules for the Higman–Sims group in characteristic 3. Commun. Algebra 21(10):3475–3487. DOI: 10.1080/00927879308824743.