Startseite Finite nearfields classified, again
Artikel Öffentlich zugänglich

Finite nearfields classified, again

  • Theo Grundhöfer EMAIL logo und Christoph Hering
Veröffentlicht/Copyright: 14. Februar 2017

Abstract

We give a rather elementary and short proof for the classification of finite nearfields, using Zsigmondy primes.

All finite nearfields are Dickson nearfields, apart from seven exceptions. This was proved by Zassenhaus [27], who distinguished the two cases where the multiplicative group of a nearfield is solvable, and where it is not. For the nonsolvable case the following result of Zassenhaus [27, Satz 16] is crucial: SL(2,5) is the only finite perfect group admitting a fixed-point-free linear representation. Proofs of this remarkable result appear in the books [17, 18.6], [26, Chapter 6], [10, XII.8] and [25, pp. 335–356]; see also [28] and [2]. Shorter proofs have been found by Maierfrankenfeld [15] and Mazurov [16].

We avoid this characterization of SL(2,5) by making a different case distinction: we consider first the nearfields of dimension n3 over the prime field 𝔽p, and then those of dimension 2. In the first case we use Zsigmondy primes for pn-1 (in the proof of Theorem 1.2). This is similar to the approach in Passman [18] and in [8], but here we obtain a more elementary proof by avoiding the modular representation theory developed in Brauer’s pivotal paper [3].

1 Finite linear groups

The normality criterion (ii) in the theorem below is a weak version of Theorem 3 in Brauer [3], where only |R|=r>2n+1 is assumed; in fact, the assumption r>2n+1 suffices according to Feit–Thompson [5]. These results depend on modular representation theory. Robinson [20, p. 481] mentions that an easier proof of the normality of R is possible if |R|=r4n-1.

Theorem 1.1.

Let G be a finite subgroup of GL(n,C) and let R be a Sylow r-subgroup of G for some prime r.

  1. If R contains an element of order at least rn, then G has a nontrivial normal r-subgroup.

  2. If |R|=r4n-2, then R is normal in G.

Proof.

Up to conjugation G is a subgroup of GL(n,A), where A is the ring of all algebraic integers in a finite extension field L of ; see [9, V.11.7 (a) and V.12.5 (b)], [4, Corollary 3.8 and Corollary 3.4A] or [17, 15.11]. We choose a maximal ideal I of A with rI. Then the field A/I has characteristic r.

(i) Reduction modulo I yields a homomorphism of G into GL(n,A/I). We claim that the kernel K of this homomorphism is an r-group. Otherwise K contains an element g of prime order sr. The ring A is noetherian, hence we have m1Im={0} by Krull’s intersection theorem. For some integer m1 the ideal Im contains all entries of the matrix g-1, but Im+1 does not. Then

1=gs=(1+(g-1))s1+s(g-1)modI2m.

Thus Im+1 contains all entries of s(g-1), and hence all entries of g-1 by Bezout’s lemma. This is a contradiction. (See also [4, Theorem 3.6B].)

If K were trivial, then GL(n,A/I) would contain an element g of order rtrn. The minimal polynomial of g divides xrt-1=(x-1)rt and has degree at most nrt-1. Thus (g-1)n=0 and grt-1-1=(g-1)rt-1=0, hence g has order at most rt-1; this is a contradiction. Therefore K is a nontrivial normal r-subgroup of G. (We have solved Exercise 2 (a) in [4, p. 94].)

(ii) The following arguments evolved from the proof of Robinson [20, Theorem A]; his proof uses character theory and tensor products like VV and VVVV, which are naturally isomorphic to the endomorphism rings EndV and EndEndV, respectively.

Let V=Ln, considered as a vector space over L (or let V=n). We write the action of G on EndV by conjugation as g2(f):=gfg-1 for gG, fEndV, and the action on EndEndV as g4(ϵ):=g2ϵg2-1 for ϵEndEndV. Then G2:={g2gG}GL(EndV)EndEndV. Let H:=NGR be the normalizer of R in G.

We claim:

  1. The centralizers of G2 and H2 in EndEndV coincide.

This is the crucial step, which uses the assumptions on r. We have H2G2, hence CEndEndVG2CEndEndVH2. We assume that this inclusion is strict and aim for a contradiction. Define ψ:CEndEndVH2CEndEndVG2 by

ψ(ϵ):=|H|-1gGg4(ϵ)=tTt4(ϵ)for ϵCEndEndVH2,

where T is any left transversal of H in G, i.e. G is the disjoint union tTtH; note that (th)4(ϵ)=t4(h4(ϵ))=t4(ϵ) for hH and

ψ(ϵ)=|H|-1xG(gx)4(ϵ)=g4(ψ(ϵ))=g2ψ(ϵ)g2-1for gG.

The subring B:={xy-1xA,yAI} of L is the localization of A at I; thus B is a discrete valuation ring, and its unique maximal ideal is of the form bB with bB. The ring M:=EndB(EndBBn) is a B-module, and G2 is a subset of M as GGL(n,A)GL(n,B). The nontrivial kernel of the L-linear map ψ is described by a system of homogeneous linear equations with coefficients from AB. Thus we find a nonzero element mCMH2 with ψ(m)=0, and we may assume that mbM by dividing m by a suitable power of b.

Let R=z and let G=HdDRdH be the double coset decomposition of G with respect to (R,H), with a suitable subset DGH. Then the set T:={1}{zid0ir-1,dD} is a left transversal of H in G, because d-1zi-jdH=NGR and zi-j1 would imply dH, as |R|=r and R=zi-j. We obtain

0=ψ(m)=m+i=0r-1dDz4id4(m)=m+i=0r-1z4i(dDd4(m)).

In the next paragraph we show that i=0r-1z4i maps M into bM; this gives the contradiction mbM.

We reduce modulo b. The quotient Bn¯:=Bn/bBn is a vector space of dimension n over the field B/bB, which has characteristic r. Let z¯End(Bn¯) be the linear map induced by z. As zr=1, we have (1-z¯)r=1-z¯r=0, which implies (1-z¯)n=0 (since the minimal polynomial of 1-z¯ has degree at most n). Now z¯2=λρ-1 where λ,ρEndEnd(Bn¯) denote the left and right multiplications with z¯, respectively. Since λ and ρ commute, we obtain

(1-z¯2)2n-1=((1-λ)+(ρ-1))2n-1ρ1-2n=0,

as every summand in the binomial expansion of ((1-λ)+(ρ-1))2n-1 involves a factor equal to 0. We repeat this argument:

z¯4=λρ-1,

where now λ,ρEndEndEnd(Bn¯) denote the left and right multiplications with z¯2, respectively. Then

(1-z¯4)4n-3=((1-λ)+(ρ-1))4n-3ρ3-4n=0.

In view of r4n-2 we obtain in EndEndEnd(Bn¯) the equation

0=(1-z¯4)r-1=i=0r-1(r-1i)(-1)iz¯4i=i=0r-1z¯4i

as (r-1i)(-1)imodr. Hence i=0r-1z4iEndBM maps M into bM. This completes the proof of assertion ().

We recall the double centralizer theorem: if E is a finite-dimensional vector space over a field and X is a completely reducible subalgebra of EndE, then X=CEndE(CEndEX); see e.g. [11, Theorem 5.1, p. 258, or Theorem 4.10, p. 222]. By Maschke’s theorem, the finite groups G2 and H2 generate completely reducible L-subalgebras of EndEndV. Assertion () and the double centralizer theorem imply that G2 and H2 generate in fact the same L-subalgebra of EndEndV. As R is normal in H, the centralizer CEndVR is invariant under H2, hence also under G2. Thus G acts by conjugation on CEndVR.

Let K be the kernel of this action of G on CEndVR. Then CEndVRCEndVK. Thus we have KCEndV(CEndVK)CEndV(CEndVR), which coincides with the L-algebra generated by R by the double centralizer theorem. Since R is commutative, we infer that K is commutative. Hence the Sylow group R is characteristic in K and therefore normal in G. ∎

Let p be a prime and let g,hGL(n,p) both have order pn-1. The subring S of the matrix ring 𝔽pn×n generated by g is isomorphic to 𝔽p[x]/(μ), where μ𝔽p[x] is the minimal polynomial of g, and degμn by the Cayley–Hamilton theorem, hence |S|pn. Thus g{0}=S is a field with pn elements (and g is irreducible), hence g={xxaa0} for some field multiplication on the additive group (𝔽pn,+); compare [9, II.3.10], where irreducibility is assumed. Similarly h is described by a second field multiplication on (𝔽pn,+). These two fields are isomorphic, and every field isomorphism is an element fGL(n,p) with fg=hf. Thus any two cyclic subgroups of orderpn-1 (which are often called Singer subgroups) are conjugate inGL(n,p).

Therefore GL(n,p) contains just one conjugacy class of subgroups isomorphic to ΓL(1,pn)NGL(n,p)g, and these subgroups are embedded naturally, via 𝔽p-linear identifications of 𝔽pn with 𝔽pn.

The following special case of results of Passman [18, Theorems 2.1, 3.1] is sufficient for our purpose (the proofs in [18] rely on [3]).

Theorem 1.2.

Let n3 and let p be a prime with p5. If GGL(n,p) has order |G|=pn-1, then G is a subgroup of ΓL(1,pn), embedded naturally into GL(n,p).

Proof.

By a result of Bang (1886) and Zsigmondy (1892) there exists a Zsigmondy prime r for pn-1, i.e. a prime r dividing pn-1 but not pk-1 for 1k<n; see e.g. [7, 3.9], [18, Corollary 1.2] or [14, II.6.2] and note that these primes r are often called p-primitive prime divisors of pn-1. Then n is the multiplicative order of p modulo r; in particular, n divides r-1, hence r1modn and rn+1.

Every subgroup Cr of order r in GL(n,p) is irreducible, and the subring of the matrix ring 𝔽pn×n generated by Cr is isomorphic to the field 𝔽pn. The normalizer of Cr in GL(n,p) is the group ΓL(1,pn), embedded naturally into GL(n,p). Thus it suffices to show that G contains a normal subgroup of order r.

The Sylow r-subgroups of G are cyclic, since GL(n,p) contains cyclic (Singer) subgroups of order pn-1. Moreover, G is isomorphic to a subgroup of GL(n,) as p does not divide |G|: in fact, Serre [21, Theorem 4.25] obtains an inclusion GGL(n,p) from (a special case of) the Schur–Zassenhaus theorem, and the ring p of p-adic integers is a subring of ; see also [4, Corollary 3.8], [17, 15.11], [22, Satz 206] or [9, V.11.7a)]. Hence G contains a normal subgroup of order r by Theorem 1.1 if r2pn-1 or r4n-2.

It remains to consider the situation where all Zsigmondy primes r for pn-1 satisfy r2pn-1 and r{n+1,2n+1,3n+1}. Then Φn(p) divides the product n(n+1)(2n+1)(3n+1), where Φn is the nth cyclotomic polynomial and n is the largest prime divisor of n, see [7, 3.5]. The prime powers pn occurring here can be determined by elementary number theory, using the methods from [7, Section 3] or [18, Section 1]. In fact, this problem has been solved by Glasby, Lübeck, Niemeyer and Praeger [6], and we infer from [6, Theorem 3 and Table 5] that pn is one of the prime powers 54,56,176. (So far, we have used only that |G| is divisible by Φn(p) and not divisible by p, but not yet the equation |G|=pn-1.)

For pn=54 we have |G|=24313 and r=13. By Sylow’s theorem G has a normal subgroup of order 13.

For pn{56,176} the prime r=7 is a Zsigmondy prime for |G|, hence G is irreducible. Let s be the largest prime divisor of |G|=p6-1; thus s{31,307} and s is a Zsigmondy prime for p3-1. Since s2|G|, there exists a normal subgroup S of order s in G by Theorem 1.1 (ii). The subring of 𝔽p6×6 generated by S is normalized by the irreducible group G and therefore isomorphic to 𝔽p3. Thus GΓL(2,p3), the centralizer CGS=GGL(2,p3) is normal in G with index at most 3, and |CGS| is divisible by 7, but not by 72. As above, CGS is isomorphic to a subgroup of GL(2,). Since 742-2, we infer from Theorem 1.1 (ii) that CGS has a characteristic subgroup of order 7 (this follows also from a result of Blichfeldt, see [4, Theorem 5.5], as well as from Dickson’s list of all subgroups of PSL(2,p3), see [13, XI.2.3], [9, II.8.24] or [23, 3.6.17]). This subgroup of order 7 is normal in G. ∎

Theorem 1.2 holds also for p{2,3} except if pn=26 and G=(C7C3)×C3 acts on 𝔽26=𝔽8𝔽4𝔽2 as ΓL(1,8)×GL(1,4), with trivial action on the last direct summand 𝔽2. This can be proved as above by considering in addition the prime powers 2n with n{3,4,8,10,12,14,18,20,36} and 3n with n{4,6,18}, see [6, Theorem 3 and Table 5] and [8, Theorem 1].

Theorem 1.2 holds also for n=2 if p{5,7,11,19,23,29,59}; see [18, Theorems 2.1, 3.1] and compare Theorem 2.4 below.

2 Finite nearfields

Nearfields yield examples of two-transitive Frobenius groups, as explained in Section 3 below.

By definition, a nearfield is a set N endowed with two binary operations + and such that (N,+) and (N×,) are groups, where N×=N{0} and 0 is the additively neutral element, and such that the distributive law a(b+c)=ab+ac holds. The additive group (N,+) is always abelian (by a result of B. H. Neumann, see [25, I, Section 2]), and it is harmless to assume that 0a=0 for all aN.

A nearfield N=(N,+,) is a Dickson nearfield if there exists a skew field D=(N,+,), with the same additive group as N, and a map α:N×AutD such that ab=abα(a) for a,bN×. In this situation, the multiplicative group N× is isomorphic to a subgroup of ΓL(1,D). Thus Dickson nearfields are closely related to skew fields.

Lemma 2.1.

A finite nearfield N is a Dickson nearfield if its multiplicative group N× has a normal 2-complement.

Proof.

Let C be a normal 2-complement in N×. All Sylow subgroups of C are cyclic, see [17, 18.1] or [19, 10.5.5]. The structure of such groups C is well known, see [19, 10.1.10] or [9, IV.2.11]; in particular, the commutator group C and C/C are cyclic, hence N× is supersolvable. Let A be maximal among the abelian normal subgroups of N×. Then A coincides with its centralizer CN×(A); otherwise the supersolvable group CN×(A)/A is not trivial, hence it has a nontrivial cyclic normal subgroup, which gives a contradiction to the choice of A.

We identify A with {xaxaA}Aut(N,+) and consider a minimal A-invariant subgroup U{0} of (N,+); then U is A-irreducible. A acts regularly on U{0}, hence |A||U|-1 and the restriction AA|U to U is injective. Moreover, A|U is contained in the multiplicative group of a (skew) field by Schur’s lemma. Thus AA|U is cyclic and therefore

|N×:A|=|N×:CN×(A)||AutA||A|.

We assume that UN and derive a contradiction. We have UbU for some bN×. Since bU is A-invariant, UbU={0} and UbU is a subgroup of (N,+). Thus

|U|2|N|=1+|N×:A||A|1+|A|2,

which is a contradiction to |A||U|-1.

Thus U=N, which means that A acts irreducibly on (N,+). Hence the subring F of End(N,+) generated by A is a field by Schur’s lemma. Moreover, dimFN=1 and F is normalized by G:={xcxcN×}, hence we have GΓL(1,F) and N is a Dickson nearfield. ∎

Corollary 2.2.

Every finite nearfield N with |N|0mod2 or |N|1mod3 is a Dickson nearfield.

Proof.

If |N| is even, then N× is a normal 2-complement in N× and Lemma 2.1 applies.

If |N|1mod3, then N× contains no element of order 3. Let S be any 2-subgroup of N×. By [17, 18.1] or [19, 10.5.5], S is cyclic or a generalized quaternion group. Hence either AutS is a 2-group, or S is the quaternion group of order 8 and AutSS4 has order 24; see [17, 9.9, 9.10]. Thus S is centralized by all elements of odd order in its normalizer in N×. By a result of Frobenius, see [19, 10.3.2] or [17, 13.3 or 13.4], N× has a normal 2-complement and Lemma 2.1 applies again. ∎

The symmetric group S4 has two Schur covers (or Schur representation groups), namely 2+S4GL(2,3) and the binary octahedral group 2-S4, which is isomorphic to the normalizer SL(2,3)(i00-i) of SL(2,3) in SL(2,9). Both Schur covers have order 48, and 2-S4 contains only one involution; see [23, 3.2.21] and [10, XII.8.4], where 2-S4 is denoted by 𝔊48.

Lemma 2.3.

Let H be a group containing only one element c of order 2.

  1. If H/cA4, then HSL(2,3).

  2. If H/cS4, then H2-S4.

  3. If H/cA5, then HSL(2,5).

Proof.

The first two assertions are proved in [10, XII.8.5]; we consider the third assertion. Since SL(2,5) satisfies the assumptions, it suffices to prove that the isomorphism type of H is uniquely determined.

The group A5 has the presentation x,yx2=y3=(xy)5=1, hence H contains elements a and b with a2,b3,(ab)5c and H=a,b,c. We infer that a2=c, and we may assume that b3=1=(ab)5, multiplying b or a by c if necessary. The group H~ with the presentation

H~=x,y,zx2=z, 1=z2=y3=(xy)5,yz=zy

admits an epimorphism H~H mapping x,y,z to a,b,c in this order. The quotient H~/z has the presentation x,yx2=y3=(xy)5=1A5. Hence |H~|=2|A5|=|H| and H is isomorphic to H~. (Similar arguments work for the groups A4 and S4 with the presentations x,yx2=y3=(xy)3=1 and w,x,yw2=x2=y2=(wx)3=(xy)3=(wx)2=1.) ∎

In the situation of Lemma 2.3 we have cH, otherwise H(H/c) would contain several involutions. Thus H is a Schur cover of H/c and Lemma 2.3 follows from general results on Schur covers; see [23, 3.2.21, 3.2.22] or [9, V.25.7, V.25.12]. Moreover, every finite group H with a unique involution c is uniquely determined by H/c; see [1, Theorem 4.1].

Theorem 2.4 (Zassenhaus).

Every finite nearfield N is either a Dickson nearfield, or |N|=p2 and one of the following holds:

  1. p=5 and N×SL(2,3),

  2. p=7 and N×2-S4,

  3. p=11 and N×SL(2,3)×C5,

  4. p=11 and N×SL(2,5),

  5. p=23 and N×2-S4×C11,

  6. p=29 and N×SL(2,5)×C7,

  7. p=59 and N×SL(2,5)×C29.

In each of these seven cases, there exists precisely one isomorphism type for N.

Proof.

The group G:={xaxaN×}Aut(N,+) is isomorphic to N× and sharply transitive on N{0}. It follows that (N,+) is an elementary abelian p-group for some prime p, hence we have (N,+)=(𝔽pn,+) for some n and GAut(N,+)=GL(n,p) has order |G|=pn-1. For p2 every involution in GL(n,p) is diagonalizable, hence -id is the only involution in G.

If p{2,3}, then N is a Dickson nearfield by Corollary 2.2, hence we may assume that p5. If n=1, then N𝔽p since G=GL(1,p)𝔽p×. If n3, then N is a Dickson nearfield by Theorem 1.2.

It remains to consider the case n=2. Let C:=G𝔽p×id. If G¯:=G/CPGL(2,p) has a cyclic subgroup of index 1 or 2, then GN× has an abelian subgroup of index 1 or 2; thus N× has a normal 2-complement, and N is a Dickson nearfield by Lemma 2.1. In the remaining cases G¯ is neither cyclic nor a dihedral group. We observe that PGL(2,p)PSL(2,p2), because every element of 𝔽p is a square in 𝔽p2. Since |G| is coprime to p, we infer that G¯ is isomorphic to A4, S4 or A5 from Dickson’s enumeration of all subgroups of PSL(2,p2); see [13, XI.2.3], [9, II.8.24] or [23, 3.6.17]. We have -idC𝔽p×id, hence p+1|G¯|12(p2-1), which leads to the following cases:

  1. G¯A4 and p{5,11},

  2. G¯S4 and p{7,23},

  3. G¯A5 and p{11,19,29,59}.

In each case we have |C|=(p2-1)/|G¯|2(mod4), hence C=-id×Ct with a subgroup Ct𝔽p×id of odd order t. If p=19, then we have t=3 and A5G¯=G¯G/GC, hence GSL(2,19) contains an element g of order 3, with eigenvalues λ,λ-1𝔽19; thus Ct=λid and gλidG{id} has the eigenvalue 1, which is absurd. This eliminates the possibility p=19.

Let H:=GSL(2,p). We claim that G/C=HC/C. Both A4 and A5 have no proper normal subgroups with index dividing p-1=|GL(2,p):SL(2,p)|, hence G/C=HC/C if G¯A4,A5. For G¯S4 we use the fact that -idC is the only involution in G. This implies that every involution in G¯ is of the form gC with gG and g2=-id, in view of gC=gtC. As gC, the polynomial x2+1 is the minimal polynomial and also the characteristic polynomial of g, hence detg=1. Since S4 is generated by its involutions, this shows that G/C=HC/C in all cases. We have HC=-id and therefore

G=HC=H×CtandH/-idG¯.

By Lemma 2.3 we have one of cases (i)–(vii).

In each case there exists at most one isomorphism type for N, since PGL(2,p) contains only one conjugacy class of subgroups isomorphic to G¯. This is proved in [13, XI Appendix, Theorem 2, p. 202] and [2, 8.8, 8.14].

In fact, the group PSL(2,p) contains a subgroup isomorphic to G¯ in each case; see [21, Section 10.2], [9, II.8.13, II.8.18], [13, XI Appendix, Theorem 1, p. 201] or [23, 3.6.26]. The preimage HSL(2,p) of such a subgroup is fixed-point-free on 𝔽p2{0} since |H| is coprime to p. Hence H×Ct is fixed-point-free on 𝔽p2{0} for every subgroup Ct𝔽p×id of order t coprime to |H|. For t=12(p2-1)/|G¯| the group H×Ct is sharply transitive on 𝔽p2{0}. This implies the existence of N in each case. ∎

We remark that in cases (i), (ii), (iv) and (vi) the prime field is not contained in the center of N.

The finite Dickson nearfields can be described in detail; see [10, XII.9.7], [14, II.7], [25, IV] or [2, Section 9]. Their multiplicative groups are metacyclic (and have normal 2-complements), since this holds for the groups ΓL(1,q)𝔽q×Aut𝔽q.

3 Sharply two-transitive groups

If N is a nearfield, then

AGL(1,N):={xax+baN×,bN}=N×(N,+)

is a sharply two-transitive permutation group on the set N. Conversely, by Jordan [12, Théorème II] every finite sharply two-transitive permutation group G has a regular (elementary abelian) normal subgroup, see also [17, 8.4] or [19, 7.3.1], hence G is of the form AGL(1,N) for some finite nearfield N; cf. [10, XII.9.10]. There exist infinite sharply two-transitive permutation groups without regular normal subgroups, see e.g. [24].

In view of the equivalence between finite nearfields and finite sharply two-transitive groups, Theorem 2.4 yields the following classification of all finite sharply two-transitive groups (compare [10, XII.9]).

Corollary 3.1.

If G is a finite sharply two-transitive permutation group of degree q, then q is a prime power and up to permutation isomorphism one of the following holds:

  1. G=AGL(1,N)AΓL(1,q) for a finite Dickson nearfield N, and G is abelian-by-metacyclic.

  2. q{52,72,112,232,292,592} and G=AGL(1,N) for one of the seven nearfields N in Theorem 2.4.

Remark.

In [10, XII.9.5 Remark] the group GL(2,3) appearing in cases b) and d) should be replaced by 𝔊482-S4 as defined in [10, XII.8.4], since GL(2,3) contains several involutions.


Communicated by Linus Kramer


References

[1] L. Babai and P. J. Cameron, Automorphisms and enumeration of switching classes of tournaments, Electron. J. Combin. 7 (2000), Research Paper R38. 10.37236/1516Suche in Google Scholar

[2] H. Bender, Endliche Fastkörper und Zassenhausgruppen, Group Theory, Algebra, and Number Theory, De Gruyter, Berlin (1996), 97–143. 10.1515/9783110811957.97Suche in Google Scholar

[3] R. Brauer, On groups whose order contains a prime number to the first power. II, Amer. J. Math. 64 (1942), 421–440. 10.2307/2371694Suche in Google Scholar

[4] J. D. Dixon, The Structure of Linear Groups, Van Nostrand, London, 1971. Suche in Google Scholar

[5] W. Feit and J. G. Thompson, Groups which have a faithful representation of degree less than (p-1)/2, Pacific J. Math. 11 (1961), 1257–1262. 10.2140/pjm.1961.11.1257Suche in Google Scholar

[6] S. P. Glasby, F. Lübeck, A. C. Niemeyer and C. E. Praeger, Primitive prime divisors and the n-th cyclotomic polynomial, J. Aust. Math. Soc. 102 (2017), no. 1, 122–135. 10.1017/S1446788715000269Suche in Google Scholar

[7] C. Hering, Transitive linear groups and linear groups which contain irreducible subgroups of prime order, Geom. Dedicata 2 (1974), 425–460. 10.1007/BF00147570Suche in Google Scholar

[8] C. Hering, On the classification of finite nearfields, J. Algebra 234 (2000), 664–667. 10.1006/jabr.2000.8542Suche in Google Scholar

[9] B. Huppert, Endliche Gruppen. I, Springer, Berlin, 1967. 10.1007/978-3-642-64981-3Suche in Google Scholar

[10] B. Huppert and N. Blackburn, Finite Groups. III, Springer, Berlin, 1982. 10.1007/978-3-642-67997-1Suche in Google Scholar

[11] N. Jacobson, Basic Algebra. II, 2nd ed., W. H. Freeman, New York, 1989. Suche in Google Scholar

[12] C. Jordan, Recherches sur les substitutions, J. Math. Pures Appl. 17 (1872), 351–367. Suche in Google Scholar

[13] S. Lang, Introduction to Modular Forms, Springer, Berlin, 1976. 10.1007/978-3-642-51447-0_1Suche in Google Scholar

[14] H. Lüneburg, Translation Planes, Springer, Berlin, 1980. 10.1007/978-3-642-67412-9Suche in Google Scholar

[15] U. Maierfrankenfeld, Perfect Frobenius complements, Arch. Math. 79 (2002), 19–26. 10.1007/s00013-002-8279-0Suche in Google Scholar

[16] V. Mazurov, A new proof of Zassenhaus theorem on finite groups of fixed-point-free automorphisms, J. Algebra 263 (2003), 1–7. 10.1016/S0021-8693(03)00107-8Suche in Google Scholar

[17] D. S. Passman, Permutation Groups, W. A. Benjamin, New York, 1968. Suche in Google Scholar

[18] D. S. Passman, p-solvable doubly transitive permutation groups, Pacific J. Math. 26 (1968), 555–577. 10.2140/pjm.1968.26.555Suche in Google Scholar

[19] D. J. S. Robinson, A Course in the Theory of Groups, Springer, New York, 1982. 10.1007/978-1-4684-0128-8Suche in Google Scholar

[20] G. R. Robinson, Remarks on reduction (mod p) of finite complex linear groups, J. Algebra 83 (1983), 477–483. 10.1016/0021-8693(83)90232-6Suche in Google Scholar

[21] J.-P. Serre, Finite Groups: An Introduction, International Press, Somerville, 2016. Suche in Google Scholar

[22] A. Speiser, Die Theorie der Gruppen von endlicher Ordnung, 4th ed., Birkhäuser, Basel, 1956. 10.1007/978-3-0348-4153-5Suche in Google Scholar

[23] M. Suzuki, Group Theory I, Springer, Berlin, 1982. 10.1007/978-3-642-61804-8Suche in Google Scholar

[24] K. Tent and M. Ziegler, Sharply 2-transitive groups, Adv. Geom. 16 (2016), 131–134. 10.1515/advgeom-2015-0047Suche in Google Scholar

[25] H. Wähling, Theorie der Fastkörper, Thales, Essen, 1987. Suche in Google Scholar

[26] J. A. Wolf, Spaces of Constant Curvature, 6th ed., AMS Chelsea Publishing, Providence, 2011. 10.1090/chel/372Suche in Google Scholar

[27] H. Zassenhaus, Über endliche Fastkörper, Abh. Math. Semin. Univ. Hambg. 11 (1936), 187–220. 10.1007/BF02940723Suche in Google Scholar

[28] H. Zassenhaus, On Frobenius groups I, Results Math. 8 (1985), 132–145. 10.1007/BF03322665Suche in Google Scholar

Received: 2016-9-7
Published Online: 2017-2-14
Published in Print: 2017-7-1

© 2017 Walter de Gruyter GmbH, Berlin/Boston

Heruntergeladen am 18.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/jgth-2017-0004/html
Button zum nach oben scrollen