Home Mathematics On orders of elements of finite almost simple groups with linear or unitary socle
Article Publicly Available

On orders of elements of finite almost simple groups with linear or unitary socle

  • Mariya A. Grechkoseeva EMAIL logo
Published/Copyright: April 8, 2017

Abstract

We say that a finite almost simple G with socle S is admissible (with respect to the spectrum) if G and S have the same sets of orders of elements. Let L be a finite simple linear or unitary group of dimension at least three over a field of odd characteristic. We describe admissible almost simple groups with socle L. Also we calculate the orders of elements of the coset Lτ, where τ is the inverse-transpose automorphism of L.

1 Introduction

The spectrumω(G) of a finite group G is the set of orders of its elements, and groups with equal spectra are said to be isospectral. To solve the problem of recognition by spectrum for a given finite group G is to describe (up to isomorphism) the finite groups that are isospectral to G. If a finite group G has a non-trivial normal solvable subgroup, then there are infinitely many finite groups isospectral to G, cf. [17, Lemma 1]. This implies that a finite group G with finite set of isospectral groups must satisfy PGAutP, where P is a product of nonabelian simple groups, and the easiest situation is when G is just a nonabelian simple group. The problem of recognition of nonabelian simple groups by spectrum has been studied extensively in recent years, and as a working hypothesis of this investigation, it was conjectured that a finite group G isospectral to a “sufficiently large” simple group S must be an almost simple group with socle S, i.e. SGAutS. In 2015, this conjecture was proved with the following precise meaning of the term “sufficiently large”: S is a linear or unitary group of dimension larger than 44, or S is a symplectic or orthogonal group of dimension larger than 60, or S is one of the sporadic, alternating and exceptional groups of Lie type other than J2, A6, A10, and D43(2) (see [14] and the references therein). Thus for a vast class of simple groups, the initial problem of recognition by spectrum was reduced to a more specific problem of describing (up to isomorphism) almost simple groups with socle S that are isospectral to S, and this is the problem we address in this paper.

For brevity, we refer to a finite almost simple group with socle S that is isospectral to S as admissible for S. Clearly we are interested in non-trivial admissible groups, i.e. other than S itself. It is not hard to check that there are no non-trivial admissible groups for the alternating groups. The information collected in [7] allows to verify that the sporadic groups do not possess non-trivial admissible groups either. One of the first examples of non-trivial admissible groups was discovered by Mazurov [16]: he showed that the finite groups isospectral to PSL3(5) are exactly PSL3(5) and its extension by the graph automorphism. Later Zavarnitsine [23] provided an example demonstrating that the number of admissible groups can be arbitrarily large: PSL3(73k) has exactly k+1 admissible groups, including itself, and these are precisely extensions by field automorphisms.

Admissible groups are described for all exceptional groups of Lie type (see [27] for references), PSL2(q) (see [2]), PSL3(q) and PSU3(q) (see [23, 24]), classical groups over fields of characteristic 2 (see [11, 13, 26]), and symplectic and odd-dimensional orthogonal groups over fields of odd characteristic (see [12]). It is worth noting that for all of these groups, there is αAutS such that S0=S,α is admissible and any other admissible group is conjugate in AutS to a subgroup of S0; in other words, any admissible group is a cyclic extension of S and up to isomorphism there is a unique maximal admissible group. Below we will see that not all simple groups satisfy the latter property.

The main result of this paper is a description of admissible groups for linear and unitary groups over fields of odd characteristic (Theorems 1, 2 and 3). Also we explicitly describe spectra of some almost simple groups with linear or unitary socle (Lemmas 3.3, 4.6, and 4.7). In the rest of this section, we introduce the notation used in the theorems and then state our results.

Throughout this paper, p is a prime, F is the algebraic closure of the field of order p, H=GLn(F), with matrices acting on row vectors by right multiplication, and τ is the inverse-transpose map gg- of H. If q is a power of p, then Fq denotes the subfield of F of order q, λq denotes a fixed primitive element of Fq and φq denotes the standard Frobenius endomorphism of H of level q, i.e. the endomorphism induced by raising matrix entries to the qth power. We identify GLn(q) with CH(φq) and GUn(q) with CH(φqτ).

We write GLn+(q) instead of GLn(q) and GLn-(q) instead of GUn(q) and use a similar agreement for PGLn(q), SLn(q), and PSLn(q). For ε{+,-}, we shorten ε1 to ε in arithmetic expressions.

As usual, by (a1,,as) we denote the greatest common divisor of positive integers a1,,as, and by [a1,,as] we denote their least common multiple. If a and b are positive integers, then π(a) denotes the set of prime divisors of a, (a)b denotes the largest divisor c of a such that π(c)π(b) and (a)b denotes the number a/(a)b.

Let L=PSLnε(q) (q=pm), and define d=(n,q-ε). We write δ=δ(εq) to denote the diagonal automorphism of L induced by diag(λ,1,,1), where λ is a primitive (q-ε)th root of unity in Fq2. We denote by φ the field automorphism induced by φp. The automorphism induced by τ is denoted by the same letter. The image of αAutL in OutL is denoted by α¯.

When n3, the inverse-transpose automorphism is outer and OutL has the following presentation (we omit overbars for convenience):

δ,φ,τδd=φm=τ2=[φ,τ]=1,δφ=δp,δτ=δ-1if ε=+,
δ,φ,τδd=τ2=[φ,τ]=1,φm=τ,δφ=δp,δτ=δ-1if ε=-.

Theorem 1 is concerned with the extension by the inverse-transpose automorphism τ. A criterion of admissibility of this extension is not very short, so it seems reasonable to write it up separately.

Theorem 1.

Let L=PSLnε(q), where n3, ε{+,-} and q is odd, and let G=Lτ. Then either ω(G)=ω(L) or one of the following holds:

  1. n=pt-1+2 with t1, q-ε(mod4) and 4ptω(G)ω(L),

  2. n=2t+1 with t1, (n,q-ε)>1 and 2(q(n-1)/2-ε(n-1)/2) is an element of ω(G)ω(L),

  3. n=pt-1+1 with t1 and 2ptω(G)ω(L),

  4. n is even, (n)2(q-ε)2, qε(mod4) and qn/2+εn/2 is an element of ω(G)ω(L),

  5. n is even, (n)2>3, (n,q-ε)2>1 and 2[q(n)2-1,qn/2-(n)2+εn/2-(n)2] is an element of ω(G)ω(L).

Theorems 2 and 3 describe admissible groups in terms of admissibility of τ. As we have mentioned, the admissible groups for PSL3(q) and PSU3(q) were determined by Zavarnitsine [23, 24], so we do not consider these groups. However, we include them in the statements of the theorems for completeness. Observe that two almost simple groups with socle S are isomorphic if and only if their images in OutS are conjugate. Thus to describe admissible groups up to isomorphism, it is sufficient to describe their images in OutS up to conjugacy. We refer to αAutS as admissible if S,α is admissible.

Theorem 2.

Let L=PSLn(q), where n3, q=pm and p is an odd prime. Let d=(n,q-1), b=((q-1)/d,m)d, η=δ(d)2, ϕ=φm/(b)2, and

ψ={φm/(b)2if n4 or 12q+1,φ(m)3if n=4 or 12q+1.

Suppose that L<GAutL. Then ω(G)=ω(L) if and only if G/L is conjugate in OutL to a subgroup of α¯, where α is one of the elements specified in Table 1.

Table 1

Conditions on the automorphism α of L=PSLn(q) in Theorem 2.

Conditions on Lα=γβ=βγ
γβ
n=pt+1,(b)2>2 or (n)2<(p-1)21φm/2τη
n2u+2(m)2=2, (n)2<(p+1)21φm/2η
npt+1,τ not admissible, |ψ|>1ψ1
b oddτ admissibleψτ
npt+1, b even, pϵ(mod 4)n=ps+2u+1ψϕ,ϕiτ
nps+2u+1,(n)2(p-ϵ)2ψϕ,ϕiτ,ϕ2jτη
nps+2u+1, (n)2<(p-ϵ)2, ϵ=+ψϕ,ϕiτ,ϕ2jτη,ϕτη
nps+2u+1, (n)2<(p-ϵ)2, ϵ=-ψϕ,ϕiτ,ϕ2jτη,ϕη
t,u>0, s02i(b)2,4j(b)2

As an example of applying Theorem 2, let us consider admissible groups for PSL4(25). For this group, we have p=5, d=4, b=2, ϕ=φ, npt+1 and n=p0+21+1. Thus up to isomorphism there are two non-trivial admissible groups, namely, the extensions by φ and by φτ. Clearly they are both maximal.

Theorem 3.

Let L=PSUn(q), where n3, q=pm and p be an odd prime. Let d=(n,q+1), b=((q+1)/d,m)d and

ψ={φ2m/(b)2if n4 or 12q-1,φ2(m)3if n=4 and 12q-1.

Suppose that L<GAutL. Then ω(G)=ω(L) if and only if n-1 is not a power of p and G/L is conjugate in OutL to a subgroup of α¯, where

  1. α=ψ if τ is not admissible and |ψ|>1,

  2. α=ψτ if τ is admissible, (n)2>2 and n16,

  3. α=ψφ(m)2 if τ is admissible and either (n)22 or n12.

Returning now to the initial recognition problem, we state the following consequence of [21, Theorem 1] and the above results.

Corollary.

Let L=PSLnε(q), where n45, ε{+,-} and q is odd. A finite group is isospectral to L if and only if it is isomorphic to an almost simple group G with socle L and G/L=α¯, where α is an identity or is as specified in Theorems 2 and 3.

2 Spectra of classical groups and related number-theoretical lemmas

In this section, we collect necessary information on spectra of classical groups and related number-theoretical lemmas. Our notation for the classical groups follows that of [7]. Recall some well-known isomorphisms between classical groups (see, for example, [15, Proposition 2.9.1]). If q is odd, then

(2.1)

PSp4(q)Ω5(q),Ω6ε(q)SL4ε(q)/{±1},
PΩ6ε(q)PSL4ε(q),Ω4+(q)SL2(q)SL2(q),
PΩ4+(q)PSL2(q)×PSL2(q),Ω4-(q)PSL2(q2),
Sp2(q)SL2(q),Ω3(q)PSL2(q).

Given a prime r, we write ωr(G) to denote the set of orders of elements of G that are coprime to r. In particular, if G is a group of Lie type over a field of characteristic p, then ωp(G) is the set of orders of semisimple elements of G. By ωr~(G) we denote the difference ω(G)ωr(G).

Lemma 2.1.

Let n2, let q be a power of a prime p, ε{+,-} and let G be PGLnε(q) or PSLnε(q). Let d=1 in the first case and d=(n,q-1) in the second case. Then ω(G) consists of all divisors of the following numbers:

  1. (qn-εn)/((q-ε)d),

  2. [qn1-εn1,qn2-εn2]/(n/(n1,n2),d), where ni>0 and n1+n2=n,

  3. [qn1-εn1,,qns-εns], where s2, ni>0 and n1++ns=n,

  4. pt(qn1-εn1)/d, where t1, n1>0 and pt-1+1+n1=n,

  5. pt[qn1-εn1,,qns-εns], where t1, s2, ni>0 and pt-1+1+n1++ns=n,

  6. (n,q-1)pt/d if n=pt-1+1 for some t1.

Proof.

See [4, Corollaries 2 and 3]. ∎

Lemma 2.2.

Let n1, let q be a power of an odd prime p and let G be one of the groups Sp2n(q), PSp2n(q), and Ω2n+1(q). Let d=c=1 if G=Sp2n(q); let d=2 and c=1 if G=PSp2n(q) or G=Ω5(q),Ω3(q); and let d=c=2 if G=Ω2n+1(q) and n3. Then ω(G) consists of all divisors of the following numbers:

  1. (qn±1)/d,

  2. [qn1±1,,qns±1], where s2, ni>0 and n1++ns=n,

  3. pt(qn1±1)/c, where t1, n1>0 and pt-1+1+2n1=2n,

  4. pt[qn1±1,,qns±1], where t1, s2, ni>0 and pt-1+1+2n1++2ns=2n,

  5. 2pt/d if 2n=pt-1+1 for some t1.

Proof.

See [5, Corollaries 1, 2 and 6] for n2 and (2.1) together with Lemma 2.1 for n=1. ∎

Lemma 2.3.

Let n2, let q be a power of an odd prime p and let ε{+,-}. Then ωp(Ω2nε(q)) consists of all divisors of the following numbers:

  1. (qn-ε)/2,

  2. [qn1-κ1,,qns-κs], where s2, κi{+,-}, ni>0, n1++ns=n and κ1κ2κs=ε,

and ωp(PΩ2nε(q)) consists of all divisors of the following numbers:

  1. (qn-ε)/(4,qn-ε),

  2. [qn1-κ,qn2-εκ]/e, where κ{+,-}, n1,n2>0, n1+n2=n; e=2 when (qn1-κ)2=(qn2-εκ)2 and e=1 otherwise,

  3. [qn1-κ1,,qns-κs], where s3, κi{+,-}, ni>0, n1++ns=n and κ1κ2κs=ε.

Proof.

For n4, see [6, Theorem 6]. If n=2,3, the assertion follows from the isomorphisms given in (2.1) and Lemma 2.1. ∎

Let k3 and q an integer whose absolute value is greater than one. A prime r such that r divides qk-1 and does not divide qi-1 for any i<k is called a primitive prime divisor of qk-1 or a Zsigmondy prime. The set of primitive prime divisors of qk-1 is denoted by Rk(q), and rk(q) denotes some fixed element of Rk(q). The following result was proved by Bang [1] in 1886 and, in a more general situation, by Zsigmondy [25] in 1892 (and has been reproved many times as explained in [18, p. 34]). A modern proof of this result can be found in [19].

Lemma 2.4 (Bang–Zsigmondy).

Let q2 and k3 be integers. If q>2 or k6, then Rk(q) is not empty. In particular, if q>2 or n3, then Rk(-q) is not empty.

The two following results are well known.

Lemma 2.5.

Let q>1, k and l be positive integers, and ε{+,-}. Then:

  1. (qk-1,ql-1)=q(k,l)-1,

  2. (qk+1,ql+1) is equal to q(k,l)+1 if (k)2=(l)2 and to (2,q+1)otherwise,

  3. (qk-1,ql+1) is equal to q(k,l)+1 if (k)2>(l)2 and to (2,q+1) otherwise,

  4. ((qk-εk)/(q-ε),k)=(q-ε,k),

  5. if (k,l)=1, then (ql-εl)/(q-ε) divides (qlk-εkl)/(qk-εk) and (ql-εl)/(n,q-ε) divides (qlk-εlk)/(n,qk-εk) for any positive integer n.

Lemma 2.6.

Let q>1 and k be positive integers and ε{+,-}.

  1. If an odd prime r divides q-ε, then (qk-εk)r=(k)r(q-ε)r.

  2. If an odd prime r divides qk-εk, then r divides q(k)r-ε(k)r.

  3. If 4 divides q-ε and k is odd, then (qk-εk)2=(k)2(q-ε)2.

Lemma 2.7.

Let q be odd and let n4 be even. Then qn/2+εn/2ω(PSLnε(q)) if and only if (n)2>(q-ε)2.

Proof.

Let d=(n,q-ε). Since a=qn/2+εn/2 is divisible by a primitive divisor rn(εq), it follows from Lemma 2.1 that aω(PSLnε(q)) if and only if a divides

qn-1(q-ε)d=(qn/2+εn/2)(qn/2-εn/2)(q-ε)d;

that is, if and only if d divides (qn/2-εn/2)/(q-ε).

If n/2 is odd, by Lemma 2.6 we have that (qn/2-εn/2)/(q-ε) is odd and, therefore, it is not divisible by d, which is even. If n/2 is even and (n)2(q-ε)2, then qε(mod4) and (d)2=(n)2, and hence

(qn/2-εn/2)2(q-ε)2=(n)22<(d)2.

Also, if (n)2>(q-ε)2, then

(qn/2-εq-ε,d)=(qn/2-εq-ε,q-ε,n)=(n2,q-ε,n)=d,

where the second equality holds by Lemma 2.5 (iv). ∎

3 Extensions by field and graph-field automorphisms

In this section, we derive some formulas concerning orders of elements in extensions of PSLnε(q) by field or graph-field automorphisms. Following the lines of the proof of [24, Proposition 13], we will exploit a correspondence between σ-conjugacy classes of CK(σk) and conjugacy classes of CK(σ), where K is a connected linear algebraic group and σ is a Steinberg endomorphism of K, i.e. a surjective endomorphism with finitely many fixed points. Also we will use a slight modification of this correspondence inspired by [10, Theorem 2.1].

We begin with necessary notation and the Lang–Steinberg theorem. If K is a group and σ is an endomorphism of K, then we write Kσ to denote CK(σ).

Lemma 3.1 (Lang–Steinberg).

Let K be a connected linear algebraic group and let σ be a surjective endomorphism of K such that Kσ is finite. Then the map xx-σx from K to K is surjective.

Recall that F is the algebraic closure of the field of order p and φq is the endomorphism of GLn(F) raising matrix entries to the qth power, where q is a power of p. An endomorphism σ of a linear algebraic group K is said to be a Frobenius endomorphism if there are an identification of K with a closed subgroup of GLn(F) and a positive integer k such that σk is induced by φq. Clearly, if σ is a Frobenius endomorphism, then σ is a Steinberg endomorphism.

Lemma 3.2.

Let K be a connected linear algebraic group over F, let α be a Frobenius endomorphism of K and let τ be an automorphism of K of order 2 that commutes with α. Let k be a positive integer, l{0,1} and let β be the automorphism of Kαkτl induced by α. Given gKαkτl, choose zK such that g=z-αz and define ζ(g)=zτlz-αk. Then ζ(g)Kα, (βg)k=z-1τlζ(g)z and the map

[βg][τlζ(g)]

is a one-to-one correspondence between the Kαkτl-conjugacy classes in the coset βKαkτl and the Kα-conjugacy classes in the coset τlKα.

Proof.

Observe that αk and αkτ are Frobenius endomorphisms of K.

Let g,zK, g=z-αz and h=zτlz-αk. Define f=τlh=zτiz-αk. Then

(3.1)gαkτl=gτlz-αk+1zαkτl=z-αzzατlz-αk+1=zτlz-αkfα=f.

This if gKαkτl, then fτlKα. Furthermore,

(βg)k=βkgαk-1g=τlz-αkz=z-1fz.

Let g1=z1-αz1Kαkτl and g2=z2-αz2Kαkτl. Suppose x-1αg1x=αg2 for some xKαkτl. Then writing y=z1xz2-1, we have yα=y and

f1=z1(βg1)kz1-1=z1x(βg2)kx-1z1-1=yf2y-1.

Conversely, if f1=yf2y-1 with yKα, then x=z1-1yz2 satisfies xαkτl=x and

x-1αg1x=z2-1y-1z1αz1-αz1z1-1yz2=z2-1y-1αyz2=z2-1αz2=αg2.

It follows that the map under consideration translates conjugacy classes to conjugacy classes and is injective. Furthermore, the conjugacy class [f] does not depend on the choice of z in the equality g=z-αz.

Now let hKα. By the Lang–Steinberg theorem, there is an element zK such that h-τl=zαkτlz-1, and therefore τlh=zτlz-αk. Then (3.1) implies that g=z-αz lies in Kαkτl, with [βg] mapping to [τlh]. Thus the map is surjective, and the proof is complete. ∎

It should be noted that some special cases of Lemma 3.2 were proved in [8, Lemma 2.10] (l=0) and [10, Theorem 2.1] (l=1 and k=1).

We use Lemma 3.2 to establish the following result generalizing [24, Corollary 14].

Lemma 3.3.

Let L=PSLn(q) and U=PSUn(q), where n3 and q=pm. Let k divide m, q=q0k and β=φm/k. Then for any i, we have the following:

  1. ω(βδi(q)L)=kω(δi(q0)PSLn(q0)),

  2. if k is even, then ω(βτδi(q)L)=kω(δi(-q0)PSUn(q0)),

  3. if k is odd, then ω(βτδi(-q)U)=kω(δi(-q0)PSUn(q0)),

  4. if k is odd, then ω(βτδi(q)L)=kω(τδi(q0)PSLn(q0)),

  5. ω(βδi(-q)U)=kω(τδi(q0)PSLn(q0)).

In particular, ω(τδi(q)L)=ω(τδi(-q)U).

Proof.

Denote the center of H=GLn(F) by Z. In particular, we write |gZ| to denote the projective order of a matrix gH, that is, the order of g modulo scalars. We apply Lemma 3.2 to H with α=φpm/k and l=0. Observe that Hαk=GLn(q) and Hα=GLn(q0). Let gGLn(q) and h=ζ(g) be the element of GLn(q0) defined in Lemma 3.2. Since (βg)k is conjugate to h in H, we see that

|βgZ|=k|(βgZ)k|=k|hZ|,

and also

deth=det(gβk-1g)=(detg)q0k-1++1=(detg)(q-1)/(q0-1).

Let λ=λq and λ0=λq0. The condition gZδiL is equivalent to

detgλi,λn=λ(i,n).

Since λ(q-1)/(q0-1) is a primitive element of Fq0, it follows that

detgλ(i,n)dethλ0(i,n).

Thus (i) holds. Similarly, (ii) and (iii) follow from Lemma 3.2 with α=φpm/kτ and l=0.

Let k be odd and α=φpm/kτ. Then Hαkτ=GLn(q) and Hα=GUn(q0). Applying Lemma 3.2 with l=1, we construct from an element gGLn(q) an element hGUn(q0) such that g=z-αz and h=zτz-αk for some zH. As in the previous case, we deduce that |βτgZ|=k|τhZ|. Furthermore, we have detg=(detz)q0+1 and deth=(detz)q-1. Since (detg)q-1=1, it follows that (detz)(q0+1)(q-1)=1. Let λ=λq and λ0=λq0q0-1, and let μ be a primitive (q0+1)(q-1)st root of unity in F. Then

detgλ(i,n)detzμ(i,n),

which is equivalent to dethλ0(i,n). Hence

(3.2)ω(βτδi(q)L)=kω(τδi(-q0)PSUn(q0)).

Similarly, taking α=φpm/k and l=1, we prove (v). Applying (v) with k=1 and observing that in this case β acts on U in the same way as τ, we have

ω(τδi(q)L)=ω(τδi(-q)U).

Now this equality and (3.2) imply (iv). ∎

Note that for unitary groups, (iii) and (v) of Lemma 3.3 cover all possibilities for an element γφ: if |γ|=2k, then γ=φm/k; while if |γ|=k with k odd, then γ=φm/kτ because (m+m/k,2m)=2m/k. Thus Lemma 3.3 expresses the spectrum of an extension by a field or graph-field automorphism in terms of known spectra and the spectrum of the extension by graph automorphism, which we consider in the next section.

4 Extension by graph automorphism

This section is largely concerned with matrices, so we need to define some of them. We denote by I the identity matrix whose size is clear from the context, and by Jk the k×k unipotent Jordan block. As in the previous section, we will calculate the projective orders of matrices, that is, the orders modulo scalars.

Recall that δ¯τ¯ is a dihedral group of order 2(n,q-ε). It follows that for odd n, every τδi is conjugate to τ modulo L, and hence

(4.1)ω(τL)=ω(i=1dτδiL)=ω(τPGLn(q)).

If n is even, then for any i, we have

(4.2)ω(τL)=ω(τδ2iL)andω(τδL)=ω(τδ2i-1L).

As we saw in Lemma 3.3, the cosets τδi(q)PSLn(q) and τδi(-q)PSUn(q) have the same orders of elements, so it suffices to describe ω(τδi(q)PSLn(q)).

Since |gτ|=2|(gτ)2|=2|ggτ|, it follows that the elements of ω(τPGLn(q)) are exactly twice the projective orders of elements of

Γn(q)={ggτgGLn(q)}.

If n is even, then ω(τL) and ω(τδL) are analogously related to the projective orders of elements of

Γn(q)={ggτgGLn(q),detg(Fq×)2}

and

Γn(q)={ggτgGLn(q),detg(Fq×)2},

respectively.

A comprehensive treatment of the equation h=ggτ for a given matrix h is provided by Fulman and Guralnick in [9], and we use the terminology and some results of this paper. First of all, it is helpful to note that h=ggτ yields

(4.3)xhx-1=(xgx)(xgx)τ

for any xGLn(q), and hence h lies in Γn(q), Γn(q) or Γn(q) if and only if the whole conjugacy class [h] lies in the corresponding set. Thus we may work not with individual matrices but with conjugacy classes. Recall that the conjugacy classes of GLn(q) are parametrized by collections of partitions

{λϕϕ is a monic irreducible polynomial over Fq}

such that |λz|=0 and ϕdeg(ϕ)|λϕ|=n. In this parametrization, the collection of partitions {λϕ} corresponds to the class [h] such that the multiplicity of ϕk as an elementary divisor of h is equal to the multiplicity of parts of size k in λϕ. We denote the partition corresponding under this parametrization to a class [h] and a polynomial ϕ by λϕ(h).

A criterion for a matrix h to lie in Γn(q) was obtained by Wall in [22]. In the same paper, Wall described the conjugacy classes of finite symplectic and orthogonal groups over fields of odd characteristic, and it turns out that the matrices of Γn(q) are very similar to symplectic and orthogonal ones.

Lemma 4.1.

If q is odd, then hΓn(q) if and only if h satisfies the following:

  1. h is conjugate to h-1,

  2. all even parts of λz-1(h) have even multiplicity,

  3. all odd parts of λz+1(h) have even multiplicity.

Proof.

See [22, Theorem 2.3.1]. ∎

Lemma 4.2.

Let hGLn(q) and let q be odd. Then h is conjugate to an element of Spn(q) if and only if h satisfies the following:

  1. h is conjugate to h-1,

  2. all odd parts of λz-1(h) and λz+1(h) have even multiplicity.

Proof.

See [22, p. 36, case (B) (i)]. ∎

Lemma 4.3.

Let hGLn(q) and let q be odd. Then h is conjugate to an element of GOnε(q) for some ε if and only if h satisfies the following:

  1. h is conjugate to h-1,

  2. all even parts of λz-1(h) and λz+1(h) have even multiplicity.

Suppose that n is even and h is a unipotent element satisfying (ii). If λz-1(h) has odd parts, then h is conjugate to an element of GOn+(q) and also to an element of GOn-(q). If there are no odd parts, then h is not conjugate to an element of GOn-(q).

Suppose that n is even, |λz-1(h)|=|λz+1(h)|=0 and h satisfies (i). Then h is conjugate to an element of only one of the groups GOn+(q) and GOn-(q).

Proof.

See [22, p. 38, Case (C) (i,i’)]. ∎

Next we establish a necessary and sufficient condition for h to lie in Γn(q) or Γn(q). Let hGLn(q), let f be the characteristic polynomial of h, and suppose that f=(z-1)n1(z+1)n2f0 with (f0,z2-1)=0. Then h is conjugate to a block diagonal matrix

(4.4)diag(h1,h-1,h0)

with blocks of dimension n1, n2 and n-n1-n2, respectively, corresponding to this factorization of f. We refer to the matrix in (4.4) as the normal form of h.

Let hΓn(q). We may replace h by its normal form diag(h1,h-1,h0) with blocks of dimensions n1, n2 and n-n1-n2, respectively, and denote by V1 the subspace of Fqn spanned by the first n1 rows. Suppose that h=ggτ. Then hg=gτg=hτ. It follows that V1g is hτ-invariant and the characteristic polynomial of hτ on V1g is equal to (z-1)n1, and hence V1g=V1. The same is true for other blocks, and so g is also block diagonal with blocks g1, g-1 and g0 of the same dimensions as h1, h-1 and h0, respectively (see also [9, Lemma 8.2]). Since hi=gigiτ for i=1,-1,0 and detg=detg1detg-1detg0, it suffices to consider the following special cases: f=(z-1)n, f=(z+1)n, and (f,z2-1)=1.

Recall that for odd q, the group Ω2nε(q) is the kernel of the spinor norm

θ:SO2nε(q)Fq×/(Fq×)2.

The definition of the spinor norm in [20, pp. 163–165] implies the following way to calculate it (see also [3, Proposition 1.6.11]).

Lemma 4.4.

Let q be odd, n even, hSOnε(q) and let B be the matrix of the invariant symmetric bilinear form of SOnε(q). Suppose that det(I-h)0. Then θ(h)det((I-h)B)(mod(Fq×)2).

Lemma 4.5.

Let q be odd, n even, hΓn(q) and let f be the characteristic polynomial of h.

  1. If f=(z+1)n, then hΓn(q).

  2. Let f=(z-1)n. If λz-1(h) has no odd parts, then hΓn(q); otherwise, hΓn(q)Γn(q).

  3. Let (f,z2-1)=1. Then h is conjugate to an element of SOnε(q) for some unambiguously defined ε , and hΓn(q) if and only if -h is conjugate to an element of Ωnε(q).

Proof.

Let h=ggτ. Possibilities for g in the first two cases are found in [9, Section 8], and we use this result for calculations. For brevity, we write xy to denote that xy(mod(Fq×)2).

(i) The group GLn(q)τ can be embedded into GL2n(q), and we consider the Jordan decomposition of gτ in the latter group. So gτ=ug1τ=g1τu, where u is unipotent and (g1τ)2=-I; i.e. g1 is a skew-symmetric matrix. Since the determinant of a skew-symmetric matrix is a square and detu=1, we see that detg is also a square.

(ii) Similarly, gτ=ug1τ=g1τu, where u is unipotent and (g1τ)2=I; i.e. g1 is a symmetric matrix. Furthermore, h=u2 and the equality ug1τ=g1τu is equivalent to the condition that u preserves the bilinear form defined by g1. By Lemma 4.3, if λz-1(u)=λz-1(h) has odd parts, then u is conjugate both to an element of SOn+(q) and to an element of SOn-(q), and so we can choose g1 with any determinant, square or non-square. If λ(u) has no odd parts, then u is conjugate to an element of SOn+(q) only, and since the multiplicities of even parts are even, n is divisible by 4. In this case detg1detI.

(iii) The fact that h is conjugate to an element of SOnε(q), where ε is unambiguously defined, follows from Lemmas 4.1 and 4.3. By Lemma 4.4, we have θ(h)det(I-h)detB, where B is the matrix of the symmetric bilinear form preserved by h. Observing that θ(-I)detB, we have

θ(-h)det(I-h)=det(I-ggτ)=det(g-g)detgτdetg,

where the final equivalence holds because the determinant of the skew-symmetric matrix g-g is a square. The proof is complete. ∎

We are ready to find ω(τPGLn(q)) and ω(τPSLn(q)).

Lemma 4.6.

Let q and n3 be odd. Then

ω(τPSLn(q))=ω(τPGLn(q))=2ω(Spn-1(q)).

Proof.

The first equality was established in (4.1). To prove the second one, we need to show that the set of the projective orders of elements of Γn(q) is equal to ω(Spn-1(q)).

Let hΓn(q). Since n is odd, Lemma 4.1 implies that 1 is an eigenvalue of h. It follows that the projective order of h is equal to the ordinary order.

We show first that |h|ω(Spn-1(q)). Let diag(h1,h-1,h0) be the normal form of h as in (4.4) and denote the dimension of h1 by k. Then k is odd. By Lemmas 4.1 and 4.2, the unipotent matrix h1 is conjugate to an element of SOk(q) and the matrix diag(h-1,h0) is conjugate to an element of Spn-k(q). If k=1, then there is nothing to prove. If k>1, then |h1|ωp(SOk(q))=ωp(Spk-1(q)), and hence |h|ω(Spk(q)×Spn-k(q))ω(Spn-1(q)).

It follows from Lemmas 4.1 and 4.2 that a semisimple matrix lies in Γn(q) if and only if it is conjugate to a matrix of the form diag(1,h) with hSpn-1(q), and thus ωp(Γn(q))=ωp(Spn-1(q)).

Let aω(Spn-1(q)) and |a|p=pt>1. If a=2pt and n-1=pt-1+1 (i.e. the condition from (v) of Lemma 2.2 holds), we define h=diag(1,-Jn-1). If a=pt or n-1>pt-1+1, there is a semisimple matrix hsSpl(q), where n-1=pt-1+1+l, such that a=pt|hs|, and we define h=diag(Jpt-1+2,hs). It is easy to see that hΓn(q) and |h|=a. The proof is complete. ∎

Lemma 4.7.

Let q be odd and n4 be even. Then

ω(τPGLn(q))=2ω(PSpn(q))

and

ωp(τPSLn(q))=2ωp(PΩn+(q))2ωp(PΩn-(q)).

If n>4, then

ωp~(τPSLn(q))=2ωp~(Ωn+1(q)).

The set ωp~(τPSL4(q)) consists of all multiples of p dividing p(q±1) if p>3, and it consists of these multiples together with 18 if p=3.

Proof.

It suffices to find the sets of the projective orders of elements of Γn(q) and Γn(q). For brevity, we denote these sets by ω(PΓn(q)) and ω(PΓn(q)), respectively. Also we denote by Z the center of GLn(q).

Let hΓn(q) and diag(h1,h-1,h0) be the normal form of h. Since n is even, it follows from Lemma 4.1 that the dimension of h1 is even too. We denote this dimension by 2k and define h2=diag(h-1,h0). Then h1 is conjugate to an element of SO2k+(q) and h2 is conjugate to an element of Spn-2k(q).

Let k>0. Then |h1|ωp(SO2k+(q))=ωp(Sp2k-2(q)), and hence

|h|ω(Sp2k-2(q)×Spn-2k(q))ω(Spn-2(q)).

Using Lemmas 2.2 and 2.3, it is easy to check that

ωp(Spn-2(q))ωp(PΩn+(q))ωp(PΩn-(q))ω(PSpn(q))

and

ωp~(Spn-2(q))ωp~(Ωn+1(q))ωp~(PSpn(q)).

Thus the projective order of h lies in the required sets.

Let k=0. Then h=h2 and |hZ|ω(PSpn(q)). It follows that

(PΓn(q))ω(PSpn(q)).

Suppose, in addition, that hΓn(q). Denote the dimension of h-1 by 2l. By Lemma 4.5, we deduce that h-1Γ2l(q), and so -h0 is conjugate to an element of Ωn-2lε(q) for some ε. Observe that |hZ|=|-hZ|.

Assume that |-h-1|>1. Then

|-hZ|=|-h|=|-h-1||-h0|.

Since |-h-1|ωp(Sp2l(q))=ωp(Ω2l+1(q)), we have

|hZ|ωp~(Ω2l+1(q)×Ωn-2lε(q))ωp~(Ωn+1(q)).

Furthermore, if n=4, then either |hZ|=9 and p=3, or |hZ| lies in pΩ2ε(q) and, therefore, divides p(q-ε)/2.

If |-h-1|=1, then either l=0 and h=h0, or l>0 and |-hZ|=|-h0|. In either case, |-hZ|ω(PΩnε(q))ω(Ωn+1(q)). In particular, if n=4 and the order |-hZ| is a multiple of p, then it divides p(q±1)/2.

Now we prove the reverse containments. If h is semisimple, then hΓn(q) if and only if h is conjugate to an element of Spn(q), and hence

ωp(PSpn(q))ω(PΓn(q)).

Let aω(PSpn(q)), |a|p=pt>1. Lemma 2.2 implies that there is a semisimple matrix hsSpl(q), where n=pt-1+1+l, such that a=pt|hs|, and we define h=diag(-Jpt-1+1,-hs). It is easy to see that hΓn(q) and

|hZ|=|-hZ|=|-h|=a.

Thus

ωp~(PSpn(q)ω(PΓn(q)),

and the assertion concerning τPGLn(q) follows.

Let hΩnε(q) be semisimple and let diag(h1,h-1,h0) be the normal form of h, with h1, h-1 of dimension k and l, respectively. If k>0, then hΓn(q) by Lemma 4.5 (ii). If k=0, then applying (i) or (iii) of Lemma 4.5 according as l>0 or l=0, we deduce that -hΓn(q). In any case, |hZ|ω(PΓn(q)).

Suppose n>4 and aωp~(Ωn+1(q)), or n=4 and a=pc with c dividing (q±1)/2, or n=4, p=3 and a=9. Let |a|p=pt. By Lemmas 2.2 and 2.3, there are ε{+,-} and a semisimple matrix hsΩlε(q), where n=pt-1+1+l, such that a=pt|hs|. Defining h=diag(-Jpt-1+1,-hs), we see that hΓn(q) and |hZ|=|-hZ|=|-h|=a. The proof is complete. ∎

In contrast to the sets of the projective orders of elements of Γn(q) and Γn(q), the corresponding set for Γn(q), where n is even, is not in general closed under taking divisors. So we do not give an explicit description of the set ω(τδPSLn(q)) for even n. However, we derive some properties of this set.

Lemma 4.8.

Let q be odd and n4 even. Then ω(τδPSLn(q))ω(PSLn(q)). If k=2s, where s0, and q=q0k, then kω(τδ(q0)PSLn(q0))ω(PSUn(q)). In particular, ω(τδ(-q)PSUn(q))ω(PSUn(q)).

Proof.

Recall that ω(τδPSLn(q)) consists of the projective orders of matrices of Γn(q) multiplied by 2.

There is an element hSOnϵ(q)Ωnϵ(q) whose projective order is equal to (qn/2-ϵ)/2. By Lemma 4.5, it follows that -hΓn(q), and so we obtain that qn/2±1ω(τδPSLn(q)).

Assume that n>4 and l=(n-2)/2 is odd. Then Lemma 2.2 together with the existence of primitive divisors rl(±q) implies that

p(ql±1)ω(PSpn(q))ω(Ωn+1(q)).

Applying Lemma 4.7, we see that

(4.5)2p(ql±1)ω(τPGLn(q))ω(τPSLn(q))ω(τδPSLn(q)).

Assume now that n=4 and qϵ(mod4). The numbers p(q+1) and p(q-ϵ) lie in ω(PSp4(q)) and do not divide p(q±1)/2 (if q-1 divides (q+1)/2, then q=3), and hence

(4.6)2p(q+1),2p(q-ϵ)ω(τδPSL4(q)).

We prove the first assertion of the lemma and the second assertion for k=1 together. To this end, suppose that ω(τδPSLn(q))ω(PSLnε(q)) and as usual define d=(n,q-ε). Then qn/2+εn/2ω(PSLnε(q)). By Lemma 2.7, this is equivalent to (n)2>(q-ε)2. It follows that l=(n-2)/2 is odd and

(d)2=(q-ε)2,

and so (ql-ε)/d is odd. Then 2(ql+ε) does not divide (qn-2-1)/d and, therefore, 2p(ql+ε)ω(PSLnε(q)). If n>4, this contradicts (4.5). If n=4, then q-ε(mod4) and (4.6) implies that 2p(q+ε)ω(τδPSL4(q)) yielding a contradiction.

Suppose that k>1 and kω(τδ(q0)PSLn(q0))ω(PSUn(q)). Observe that (q+1)2=2 and choose ϵ so that q0ϵ(mod4). Let n/2 be odd. By assumption, we have a=k(q0n/2-ϵ)ω(PSUn(q)). Since a is a multiple of rn/2(ϵq0) and rn/2(ϵq0)Rn/2(q)=Rn(-q), it follows that a divides

c=qn-1(q+1)(n,q+1).

This is a contradiction because

(a)2=(qn/2-1)2>(c)2=(qn-1)24.

Thus l=(n-2)/2 is odd and by (4.5) and (4.6), we have

2p(q0l-ϵ)ω(τδPSLn(q0)).

Then a=2kp(q0l-ϵ)ω(PSUn(q)), and hence a divides p(qn-2-1)/(n,q+1). This contradicts (a)2=(qn-2-1)2.

The final assertion follows from the second one and Lemma 3.3. ∎

We close this section with a proof of Theorem 1 and one of its consequences.

Proof of Theorem 1.

Recall that L=PSLnε(q), d=(n,q-ε) and consider the difference ω(τL)ω(L). We analyze separately two cases according as n is odd or even.

Let n be odd. By Lemma 4.6, the difference under consideration is equal to 2ω(Spn-1(q))ω(L). We consider the numbers from Lemma 2.2 defining ω(Spn-1(q)) in turn.

Let a=2pt, where pt-1+1=n-1. Any element of ω(L) that is a multiple of pt divides pt(q-ε)/d. Since d is odd, 2a divides pt(q-ε)/d if and only if 4 divides q-ε. Thus 4ptω(τL)ω(L) if and only if n=pt-1+2 for some t1 and q-ε(mod4).

Let

a=pt[qn1±1,qn2±1,,qns±1],

where pt-1+1+2n1+2n2++2ns=n-1, or

a=[qn1±1,qn2±1,,qns±1],

where s2 and 2(n1+n2++ns)=n-1. Then 2a divides pt[q2n1-1,q2n2-1,,q2ns-1,q-ε] or [q2n1-1,q2n2-1,,q2ns-1,q-ε], respectively.

The remaining possibility is a=ql±1, where l=(n-1)/2. If a=ql+εl, then 2a divides

qn-1-1d=(ql+εl)ql-εld

since d is odd. Let a=ql-εl. Assume that l is not a 2-power. Then (l)2l/3, and hence l+2(l)2<n. It follows that L has an element of order

c=[ql-εl,q2(l)2-1,q-ε].

Since 2(a)2=2(ql-εl)2(q2l-1)2=(q2(l)2-1)2, we have that 2a divides c. Finally, assume that l is a 2-power. Then the 2-exponent of L is equal to (q2l-1)2 and any element of ω(L) that a multiple of (q2l-1)2 divides (q2l-1)/d. Clearly, 2a divides (q2l-1)/d if and only if d divides ql+εl, which is equivalent to d=1. Thus 2(ql-εl)ω(τL)ω(L) if and only if n-1=2t and d1.

Let n be even. By Lemma 4.7, the set ωp~(τL) is equal to 2ωp~(Ωn+1(q)) if n>4, it consists of divisors of p(q±1) if n=4 and p>3, and it consists of divisors of p(q±1) together with 18 if n=4 and p=3. We can consider the numbers from Lemma 2.2 defining ωp~(Ωn+1(q) for n6 and the numbers defining ωp~(τL) for n=4 together.

If a=pt and n=pt-1+1, then 2aω(L).

Let

a=pt[qn1±1,qn2±1,,qns±1],

where s2 and pt-1+1+2n1+2n2++2ns=n. Then we have that 2a divides pt[q2n1-1,q2n2-1,,q2ns-1]ω(L).

Let a=pt(qn1-ϵ)/2, where pt-1+1+2n1=n. If ϵ=εn1, then 2a divides pt(qn1-εn1)ω(L), while if ϵ=-εn1, then 2a divides

ptq2n1-1d=pt(qn1+εn1)qn1-εn1d.

Thus

ωp~(τL)ω(L),

whenever npt-1+1.

To handle ωp(τL)=2ω(PΩ2n+(q))2ω(PΩ2n-(q)), we consider the numbers from Lemma 2.3 that define ωp(PΩ2n±(q)).

Let

a=[qn1±1,qn2±1,,qns±1],

where s3 and 2(n1+n2++ns)=n. In this case, we have that 2a divides [q2n1-1,q2n2-1,,q2ns-1]ω(L).

Let a=(qn/2-ϵ)/(4,qn/2-ϵ). If ϵ=εn/2, then 2a divides

qn/2-εn/2ω(L).

If ϵ=-εn/2 and (4,qn/2+εn/2)=4, then 2a divides

qn-1(q-ε)d=(qn/2+εn/2)(qn/2-εn/2)(q-ε)d

because d/2 divides (n/2,q-ε)=((qn/2-εn/2)/(q-ε),q-ε). If ϵ=-εn/2 and (4,qn/2+εn/2)=2, then Lemma 2.7 implies that 2a=qn/2+εn/2 does not lie in ω(L) if and only if (n)2(q-ε)2. Assuming the last inequality, the condition (4,qn/2+εn/2)=2 is equivalent to qε(mod4).

Finally, let a=[qn1-ϵ1,qn2-ϵ2]/e, where 2(n1+n2)=n and with e=2 if (qn1-ϵ1)2=(qn2-ϵ2)2 and e=1 if (qn1-ϵ1)2(qn2-ϵ2)2. We define n0=n/(2n1,2n2). We may assume that (qn1-ϵ1)2(qn2-ϵ2)2.

If ϵ2=εn2, then 2a divides [q2n1-1,qn2-εn2]ω(L) since

2(a)22(qn1-ϵ1)2(q2n1-1)2.

Similarly, we obtain that if ϵ1=εn1 and (qn1-ϵ1)2=(qn2-ϵ2)2, then 2a divides [q2n2-1,qn1-εn1].

Let ϵ2=-εn2 and ϵ1=-εn1. If e=2, then 2a divides

[q2n1-1,q2n2-1]q-ε.

If e=1, then n1 and n2 have opposite parity, therefore, n/2 is odd. It follows that n0 is odd too, and 2a divides [q2n1-1,q2n2-1]/(n0,q-ε).

Let ϵ2=-εn2 and ϵ1=εn1. We may assume that (qn1-εn1)2>(qn2+εn2)2. If n1(n1)2, then

n1+2(n1)2+2n2<n

and 2a divides [qn1-εn1,q2n2-1,q2(n1)2-1]. If 2(qn1-εn1)2(q2n2-1)2, then 2a divides [q2n2-1,qn1-εn1]. If qn2+εn2 divides q2n1-1, then 2a divides q2n1-1 too. Finally, if (n0,q-ε)=1, then 2a divides [q2n1-1,q2n2-1].

Thus it remains to consider the case when n1=(n1)2, 2(qn1-εn1)2>(q2n2-1)2, qn2+εn2 does not divide q2n1-1 and (n0,q-ε)1. The two first conditions yield n1>(n2)2. In particular, (n/2)2=(n2)2 and n0=n/(2n2)2=(n)2. Then the third condition is equivalent to n2 not being a 2-power. Observe that

(n)2=n(2n2)2=n1+n2(n2)2=n1(n2)2+(n2)2.

It follows that (n)2 is a sum of a non-identity 2-power and an odd number greater than 1 and also ((n)2,q-ε)1. In particular, (n)2>3 and (n,q-ε)21.

Conversely, suppose that n=2tl, where t1, l5 is odd and (n,q-ε) is divisible by an odd prime r. Writing n1=(n)2=2t, n2=n/2-n1=2t-1(l-2) and a=[qn1-1,qn2+εn2], we see that

(qn1-1)2=2t-1(q2-1)2>(n2)2(q+ε)2(qn2+εn2)2

and hence 2aω(τL). Assume that 2aω(L) and let

c=[ql1-εl1,,qls-εls]f,

where l1++ls=n, be a number from Lemma 2.1 that 2a divides. Since 2(a)2=2(qn1-1)2=(q2n1-1)2, it follows that at least one of the numbers l1,,ls is a multiple of 2n1. Also r2n2(εq) divides a, and hence at least one of them is a multiple of 2n2. Observing that [2n1,2n2]=2t+1(l-2)>2tl=n, we deduce that those are different numbers, which yields s=2, l1=2n1 and l2=2n2. Then f=(n/(2n1,2n2),q-ε), and so (f)r>1. Since (r,q+ε)=(r,n1)=(r,n2)=1, it follows that

(c)r=(q-ε)r(f)r<(q-ε)r=(a)r.

This is a contradiction, therefore, 2aω(L), and the proof is complete. ∎

An interesting consequence of Theorem 1 is that whenever τ is admissible as an automorphism of PSLnε(q), it is also admissible as an automorphism of PSLnε(q1/k) for every odd k.

Lemma 4.9.

Let k be odd and q=q0k. If ω(τL)ω(L), then

ω(τPSLnε(q0))ω(PSLnε(q0)).

In particular, if ε=+, τ and β=φm/k are admissible for L, then βτ is also admissible.

Proof.

Assume that ω(τPSLnε(q0))ω(PSLnε(q0)). This implies that the numbers n and q0 satisfy the conditions of one of items (i)–(v) of Theorem 1. Since k is odd, it follows that (n,q0-ε) divides (n,q-ε) and (q-1)2=(q0-1)2, and thus the numbers n and q satisfy the same conditions. This contradicts the hypothesis that ω(τL)ω(L).

To prove the second assertion, it suffices to show that ω(βτL)ω(L). By Lemma 3.3, the proved containment and admissibility of β, we see that

ω(βτL)=kω(τPSLn(q0))kω(PSLn(q0))=ω(βL)ω(L),

and the proof is complete. ∎

5 Admissible groups

In this section, we prove Theorems 2 and 3. As we have mentioned, the theorems with n=3 were proved in [23, 24], and so we assume that n4. Throughout this section, q=pm is odd, L=PSLnε(q) and G is a group such that L<GAutL. Also, we fix the numbers d=(n,q-ε) and b=((q-ε)/d,m)d.

We begin with lemmas that hold for both linear and unitary groups. We say that a subgroup of OutL is admissible if it is the image of an admissible group.

Lemma 5.1.

If GInndiagL>L, then ω(G)ω(L). In particular, admissible groups of OutL are abelian and any non-trivial admissible subgroup of the group δ¯τ¯ is conjugate in this group to τ¯.

Proof.

By Lemma 2.1, if |GInndiagL|/|L|=i>1, then G has an element of order (qn-εn)i/(q-ε)d, which does not lie in ω(L). Thus admissible groups of OutL can be embedded into the image of the group generated by φ and γ, which is abelian. The group δ¯τ¯ is dihedral, and so all of its non-trivial subgroups that intersect trivially with δ¯ are conjugate to τ¯ or δ¯τ¯, and in the latter case we may assume that n is even. But Lemma 4.8 says that δ¯τ¯ is not admissible. ∎

Lemma 5.2.

Suppose that n5 and |G/L| is odd. Then ω(G)=ω(L) if and only if n-1 is not a p-power, G/L is conjugate in OutL to a subgroup of φ¯ and |G/L| divides ((q-ε)/d,m)d.

Proof.

If ε=+, the assertion is proved in [11, Propositions 6,7], and if ε=-, it is proved in [13, Proposition 6]. ∎

Lemma 5.3.

Let n=4 and |G/L| is odd. Then ω(G)=ω(L) if and only if 12 divides q+ε and G/L is a 3-group.

Proof.

Let ω(G)=ω(L) and rπ(G/L). Then G contains a field automorphism of order r, and hence by Lemma 3.3, ω(G) includes rω(PSL4ε(q0)), where q=q0r. It follows that rr4(q0),rr3(εq0)ω(L). Since r4(q0)R4(q), we see that r divides a=(q2+1)(q+ε)/d and, in particular, it does not divide q-ε. If r3, then r3(εq0)R3(εq), and so r divides c=(q3-ε)/d, which is a contradiction because (a,c)=1. Let r=3. Then 3 divides q+ε. Since 3p(q0-ε) is an element of ω(L), we have that 3(q0-ε) divides (q2-1)/d, and thus 4 divides q+ε.

Conversely, let 12 divide q+ε and G/L be a 3-group. We may assume that G is the extension of L by a field automorphism of order dividing (m)3. To prove that ω(G)ω(L), it suffices to check kω(PSL4ε(q0))ω(PSL4ε(q)) with q=q0k for every divisor k of (m)3. Note that 12 divides q0+ε, (q+ε)3=k(q0+ε)3 and (4,q0-ε)=2=d. Since p3, the set ω(PSL4ε(q0)) consists of divisors of (q02+1)(q0+ε)/2, (q03-ε)/2, q02-1 and p(q02-1)/2. Observing that

k(q02+1)(q0+ε)(q2+1)(q+ε)

and

k(q03-ε),k(q02-1)q2-1,

we obtain the desired containment. ∎

By Lemmas 5.2 and 5.3, it follows that up to conjugacy OutL has only one maximal admissible subgroup of odd order, and we can take this subgroup to be ψ¯, where ψφ has order (b)2 if n>4 or q1(mod12), and order (m)3 if n=4 and q-ε(mod12).

Let η=δ(d)2 and S2 be the Sylow 2-subgroup of OutL generated by η¯, φ¯(m)2 and τ¯.

Lemma 5.4.

If ω(G)=ω(L), then G/L is conjugate in OutL to a subgroup of ψ¯×S2.

Proof.

By Lemma 5.1, the group G/L is abelian, and hence is the direct product of its Hall 2-subgroup A1 and Sylow 2-subgroup A2. Since A1 is admissible, it is conjugate to a subgroup of ψ¯. Replacing G/L by a conjugate if necessary, we may assume that A1ψ¯.

Note that η¯ centralizes ψ¯. Indeed, if ε=+, then η¯ψ¯=η¯q0, where q=q0|ψ|. Since |ψ| is odd, we have

(q-1)2=(q0-1)2,

and hence (d)2 divides q0-1. If ε=-, then η¯ψ¯=η¯q0, where q2=q0|ψ|. Now (q+1)2<(q2-1)2=(q0-1)2, and again (d)2 divides q0-1. It follows that the whole group S2 centralizes ψ¯. Thus A2 is conjugate in COutL(A1) to a subgroup of S2, and the whole group G/L is conjugate to a subgroup of ψ¯×S2. ∎

The structure of the group S2 varies according to whether linear or unitary groups are under consideration, so in the rest of this section we consider the cases ε=+ and ε=- separately. We begin with the case of unitary groups, in which S2=η¯φ¯(m)2.

Lemma 5.5.

Let ε=- and 1<G/LS2. If ω(G)=ω(L), then G/L is conjugate to a subgroup of φ¯(m)2.

Proof.

We may assume that n is even. If m is odd, then S2=η¯τ¯ and by Lemma 5.1, the group G/L is conjugate to τ¯.

Let m be even. Then |η¯|=2 and S2=η¯×φ¯(m)2. Suppose Gφ¯(m)2. Then G contains φm/kη for some k>1 dividing (m)2. Let q=q0k. By Lemmas 3.3 and 4.8 together with Lemma 4.2, we have

ω(φm/kηL)=kω(τδd/2(q0)PSLn(q0))=kω(τδ(q0)PSLn(q0))ω(L),

which is a contradiction. ∎

Thus if ε=- and τ is not admissible, then S2 has no non-trivial admissible subgroups.

Lemma 5.6.

Suppose ε=-, τ is admissible, k>1 divides (m)2 and β=φm/k. Then β is admissible if and only if (n)22 or n=4,8,12. If β is admissible and γφ is admissible and has odd order, then γβ is also admissible.

Proof.

We show first that ω(βL)ω(L) if and only if (n)22 or n=4,8,12. By Lemma 3.3, the set ω(βL) is equal to kω(τPSLn(q0)), where q=q0k.

Let n be odd. Since τ is admissible and q1(mod4), Theorem 1 implies that n-2 is not a power of p. Then using Lemmas 2.2 and 2.6, it is not hard to check that kω(Spn-1(q0))ω(Spn-1(q)). Now applying Lemma 4.6 yields

kω(τPSLn(q0))=2kω(Spn-1(q0))
2ω(Spn-1(q))=ω(τL)ω(L).

Let n be even. Observe that (d)2=(q+1)2=2. By the admissibility of τ and Theorem 1, it follows that n-1 is not a p-power. Also it is not hard to verify that kω(Ωn+1(q0))ω(Ωn+1(q)) (cf. [12, Theorem 1]). Applying Lemma 4.7, we see that kωp~(τPSLn(q0))ωp~(τL) for n>4. Since k(q0±1) divides q-1, the same is true for n=4. Thus it remains to examine when 2kωp(PΩn±(q0)) is a subset of ω(L), and we consider the numbers from Lemma 2.3 in turn.

Let

a=[q0n1±1,q0n2±1,,q0ns±1],

where s3 and 2(n1+n2++ns)=n. Then 2ka divides

[q2n1-1,q2n2-1,,q2ns-1]ω(L).

Let

a=(q0n/2-ϵ)(4,q0n/2-ϵ).

If n/2 is even, then 2ka divides qn/2-1ω(L). Suppose that n/2 is odd. Then rn/2(ϵq) divides a and lies in Rn/2(q)=Rn(-q), and so 2kaω(L) if and only if 2ka divides

c=qn-1(q+1)d=(qn/2-1)(qn/2+1)(q+1)d.

Clearly, (a)2 divides (c)2. Since (d)2=2, we see that (c)2=(q-1)2/2. If (4,q0-ϵ)=4, then 2k(a)2=k(q0-ϵ)2/2=(q-1)2/2. If (4,q0-ϵ)=2, then 2k(a)2=2k(q-1)2/2.

Finally, let a=[q0n1-ϵ1,q0n2-ϵ2]/e, where 2(n1+n2)=n, and with e=2 if (qn1-ϵ1)2=(qn2-ϵ2)2 and e=1 if (qn1-ϵ1)2(qn2-ϵ2)2.

Suppose that n/2 is odd. We may assume that (qn1-ϵ1)2(qn2-ϵ2)2. If n2 is even, then 2ka divides [q2n1-1,qn2-1]. Let n2 be odd. Then n1 is even. If n1(n1)2, then 2ka divides [qn1-1,q(n1)2-1,q2n2-1]. If n1=(n1)2 and n2=1, then 2ka divides q2n1-1. If (n1)=(n1)2 and n2>1, then by the admissibility of τ, we have (n,q+1)2=1 and, therefore, (n/(2n1,2n2),q+1)=1. So [q2n1-1,q2n2-1] lies in ω(L), and it is divisible by 2ka.

Suppose that n/2 is even and n/28. We can take n1 and n2 to be odd coprime numbers larger than 1. Also we take ϵ1=+ and ϵ2=-. Then

a=[q0n1-1,q0n2+1]

and 2k(a)2=k(q02-1)2=(q2-1)2. Since a is a multiple of both rn1(q0) and rn2(-q0) and rni(±q0)R2ni(-q) for i=1,2, it follows that 2kaω(L) if and only if 2ka divides c=[q2n1-1,q2n2-1]/(n/(2n1,2n2),q+1). But we have (c)2=(q2-1)2/2<(a)2 because n/2 is even. Thus ω(βL)ω(L).

We are left with the cases n=4,8,12. If n=4, then 2a divides q02-1, and hence 2ka divides q2-1. If n=8, then a divides [q03±1,q0±1] or (q04-1)/2, and so 2ka divides q6-1 or q4-1. If n=12, then a divides [q05±1,q0±1], [q04±1,q02±1], or (q06-1)/2, and 2ka divides q10-1, q8-1, or q6-1.

We established that the condition for ω(βL) to be a subset of ω(L) depends only on n and not on q0, and thus β is admissible if and only if ω(βL)ω(L).

Next, suppose that β is admissible, γφ is admissible and has odd order l and let q0=q1l. To prove that γβ is admissible, it suffices to show the inclusion ω(γβL)ω(L). By Lemma 4.9, if we regard τ as an automorphism of PSUn(q1k), it is still be admissible, and so by the above result, it follows that

kω(τPSLn(q1))ω(PSUn(q1k)).

By Lemma 3.3, we have

ω(γβL)=lkω(τPSLn(q1))lω(PSUn(q1k))=ω(γL)ω(L),

as desired. ∎

We are now in a position to prove Theorem 3.

Proof of Theorem 3.

Suppose that ω(G)=ω(L). By Lemmas 5.4 and 5.5, it follows that G/L is conjugate to a subgroup in ψ¯×φ¯(m)2. If n-1 is a p-power, Lemmas 4.7, 5.2 and 5.3 imply that any non-trivial element of G/L is not admissible. If n-1 is not a p-power, then ψ is admissible, and applying Lemma 5.6 completes the proof. ∎

Now we assume that ε=+, and so we have S2=η¯(φ¯m2×τ¯). Given an element 1βφ(m)2, we define βϵ with ϵ{+,-} by setting β+=β and β-=βτ.

Lemma 5.7.

Let ε=+, k>1 divide (m)2 and q=q0k. Then for any ϵ{+,-}, the following hold:

  1. If n is odd or k does not divide (q-1)/d, then

    kω(PSLnϵ(q0))ω(L).
  2. If n is even, k divides (q-1)/d and n=1+pt-1, then

    kω(PSLnϵ(q0)){kpt}ω(L).
  3. If n is even, k divides (q-1)/d and n-1 is not a p-power, then

    kω(PSLnϵ(q0))ω(L).

In particular, if β=φm/k, then βϵ is admissible if and only if n is even, k divides (q-1)/d and n-1 is not a p-power. If βϵ is admissible and γφ is admissible and has odd order, then γβϵ is also admissible.

Proof.

If n is odd, then 2rn(ϵq0) is not an element of ω(L) since rn(ϵq0)Rn(q) and (qn-1)/(q-1) is odd. If n is even and k does not divide (q-1)/d, then krn-1(ϵq0)ω(L). Indeed, otherwise the fact that rn-1(ϵq0)Rn-1(q) implies that k divides (qn-1-1)/d. But (qn-1-1)2/(d)2=(q-1)2/(d)2, and thus (i) follows.

Let n be even and suppose that k divides (q-1)/d. We claim that kaω(L) for all aω(PSLnϵq0)) except a=pt for n=1+pt-1. The number a divides one of the numbers in items (i)–(v) of Lemma 2.1, and we consider these possibilities in turn.

If we have a=(q0n-1)/(q0-ϵ)(n,q0-ϵ), then ka divides qn/2-1. Similarly, if a=[q0n1-ϵn1,q0n2-ϵn2]/(n/(n1,n2),q0-ϵ], where n1+n2=n and both n1, n2 are even, then ka divides [qn1/2-1,qn2/2-1]. If a=[qn1-ϵn1,qn2-ϵn2]/(n/(n1,n2),q0-ϵ], where n1 and n2 are odd, then ka divides the element c=[qn1-1,qn2-1]/(n/(n1,n2),q-1]. Indeed, if q0-ϵ(mod4), then

k(a)2=k(c)2,

while if q0ϵ(mod4), then

k(q0-ϵ)2=(q-1)2(k)2(n)2,

which implies that (q0-ϵ)2(n)2 and k(a)2=(q-1)2/(n)2=(c)2. Also, a divides c by Lemma 2.5. Finally, if a=[q0n1-ϵn1,,q0ns-ϵns], where s3 and n1+,+ns=n, or a=pt[q0n1-ϵn1,,q0ns-ϵns], where s2 and 1+pt-1+n1++ns=n, we have that ka divides [qn1-1,,qns-1] or pt[qn1-1,,qns-1], respectively.

Let β=φm/k. By Lemma 3.3, we have ω(βϵL)=kω(PSLnϵ(q0)). Thus if βϵ is admissible, then n is even, k divides (q-1)/d and n-1 is not a p-power. Conversely, if all these three conditions are satisfied, then by the above

ω(βk1L)=(k/k1)ω(PSLn(q0k1))ω(L)

for every divisor k1 of k, and hence βϵ is admissible.

To prove the final assertion, it suffices to check ω(γβϵL)ω(L) for admissible γ and βϵ. Let |γ|=l and q0=q1l. Observing that (q-1)2=(q1k-1)2, we see that k divides (q-1)/d if and only if it divides (q1k-1)/(n,q1k-1). Thus the admissibility of βϵ yields

kω(PSLnϵ(q1))ω(PSLn(q1k)).

Applying Lemma 3.3, we have

ω(γβϵL)=lkω(PSLnϵ(q1))lω(PSLn(q1k))=ω(γL)ω(L),

and the proof is complete. ∎

Lemma 5.8.

Let ε=+, let n be even, k>1 divide (m)2, β=φm/k and q=q0k. If α=βϵηj is admissible, then α¯ is conjugate in the group S2 to either β¯ϵ or β¯ϵη¯, with q0-ϵ(mod4) and 2k(q-1)2/(d)2 in the latter case. Furthermore, if q0-ϵ(mod4) and 2k(q-1)2/(d)2, then the following statements hold:

  1. βϵη is admissible if and only if n cannot be represented as 1+pt-1+2u with t,u1 and, in addition, k=2 whenever n-1 is a p-power,

  2. If βϵη is admissible and γφ is admissible and has odd order, then γβϵη is admissible.

Proof.

Note that η¯β¯ϵ=η¯ϵq0, and so

(5.1)β¯ϵη¯=η¯-1β¯ϵη¯=β¯ϵη¯1-ϵq0.

Suppose that α=βϵηj is admissible and α¯ is not conjugate in S2 to β¯ϵ. Then (5.1) implies that (q0-ϵ,(d)2) does not divide j, or in other words

(5.2)(j)2<(q0-ϵ,d)2.

Since η=δ(d)2, applying Lemma 3.3 yields

(5.3)ω(αL)=kω(δj(d)2(ϵq0)PSLnϵ(q0)).

The index of δj(d)2(ϵq0)PSLnϵ(q0) in PGLnϵ(q0) is equal to (j(d)2,q0-ϵ). There is an element of order q0n-1-ϵ in the group PGLn(q0), and hence writing a=(q0n-1-ϵ)/(j(d)2,q-ϵ), we see that kaω(αL). Since a is a multiple of rn-1(ϵq0)Rn-1(q) and lies in ω(L), it follows that a divides c=(qn-1-1)/d. If q0ϵ(mod4), then by (5.2), we deduce that

(a)2=k(q0-ϵ)2(j)2=(q-1)2(j)2>(q-1)2(d)2=(c)2.

Thus q0-ϵ(mod4). Then (q0-ϵ,d)2=2, and so j is odd. In particular, (5.1) shows that α is conjugate to βϵη. Also, (a)2=2k(c)2=(q-1)2/(d)2, and the first assertion is proved.

Next, let α=βϵη, q0-ϵ(mod4) and 2k(q-1)2/(d)2. Observing that (q-1)2=k(q0+ϵ)2, we deduce from the last inequality that 2(d)2(q0+ϵ)2, and hence

α¯2=β¯2η¯1+ϵq0=β¯2.

Since 2k(q-1)2/(d)2, it follows from Lemma 5.7 that β2 is admissible if and only if k=2 or n-1 is not a p-power.

Suppose n=1+pt-1+2u with t,u1. Then PGLnϵ(q0) has an element of order pt(q02u-1), and so kaω(αL), where a=pt(q02u-1)/(d,q0-ϵ)2. Since ka is a multiple of kpt(q02u-1)2=pt(q2u-1)2, it lies in ω(L) only if it divides pt(q2u-1)/d, which is not the case.

To prove (i), it remains to verify that ω(αL)ω(L) whenever n cannot be represented as 1+pt-1+2u with t,u1. Since ((d)2,q0-ϵ)=(n,q0-ϵ)2=(n,q0-ϵ)/2, it follows from (5.3) that

ω(αL)=kω(2PSLnϵ(q0)PSLnϵ(q0)),

where 2PSLnϵ(q0) is the unique subgroup of PGLnϵ(q0) that contains PSLn(q0) as a subgroup of index 2. Denote ω(2PSLnϵ(q0)PSLn(q0)) by ω.

If n=pt-1+1, then any element of GLn(q) whose order is divisible by pt is conjugate to a scalar multiple of the n×n unipotent Jordan block, therefore, ptω. By Lemma 5.7, we have

k(ωω(PSLnϵ(q0)))ω(L),

and hence we are left with elements of ωω(PSLnϵ(q0)). By Lemma 2.1, it follows that this difference consists of some divisors of the following numbers:

  1. 2(q0n-1)/((q0-ϵ)(n,q0-ϵ)),

  2. [q0n1-ϵn1,qn2-ϵn2]/(n/(n1,n2),q0-ϵ), where n1,n2 are odd, n1+n2=n,

  3. 2pt(q0n1-ϵn1)/(n,q0-ϵ), where n1>0 and pt-1+1+n1=n.

Let a=2(q0n-1)/((q0-ϵ)(n,q0-ϵ)). Then

k(a)2=2k(q0n-1)2(q0-ϵ)2(q0-ϵ,n)2=(qn-1)22=(qn/2-1)2,

and so ka divides qn/2-1ω(L).

Let a=2[q0n1-ϵn1,qn2-ϵn2]/(n/(n1,n2),q0-ϵ), where n1, n2 are odd and n1+n2=n. Consider the number c=[qn1-1,qn2-1]/(n/(n1,n2),q-1) lying in ω(L). By Lemma 2.5, both numbers (qni-1)/(n/(n1,n2),q0-ϵ), i=1,2, divide c. Also, k(a)2=2k(q-1)2/d2=(c)2, and hence ka divides c.

Let a=2pt(q0n1-ϵn1)/(n,q0-ϵ), where n1>0 and pt-1+1+n1=n. Note that n1 is even and by hypothesis, (n1)2n1. Since

k(a)2=2k(q0n1-1)2(q0-ϵ)2=(qn1-1)2=(q(n1)2-1)2,

we see that ka divides pt[qn1/2-1,q(n1)2-1]ω(L).

To prove part (ii), take admissible γ and α. We may assume that n-1 is not a p-power because otherwise γ=1 by Lemmas 5.2 and 5.3. Then by the equality

α¯2=β¯2

and Lemma 5.8, to prove that γα is admissible, it suffices to check the inclusion ω(γαL)ω(L). Denote |γ| by l and let q0=q1l. Observe that

q1q0-ϵ(mod4)

and 2k divides (q1k-1)2/(n,q1k-1)2.

By Lemma 3.3, we have

ω(γαL)=lkω(δ(d)2(ϵq1)PSLnϵ(q1)).

Using the fact that (d,q1-ϵ)2=(q1-ϵ)/2 and the above results, we deduce that

kω(δ(d)2(ϵq1)PSLnϵ(q1))=kω(2PSLnϵ(q1)PSLnϵ(q1))ω(PSLn(q1k)).

Thus

ω(γαL)lω(PSLn(q1k)))=ω(γL)ω(L),

and the proof is complete. ∎

We are now ready to describe admissible 2-subgroups of OutL in the case of linear groups, and then prove Theorem 2. Recall that b=((q-ε)/d,m)d and define ϕ=φm/(b)2.

Lemma 5.9.

Let ε=+ and let 1<G/LS2=η¯(φ¯(m)2×τ¯). Then ω(G)=ω(L) if and only if G/L is cyclic and up to conjugacy in S2 is generated by the image of one of the following elements:

  1. τ if b is odd and τ is admissible,

  2. β±, where 1βϕ, if b is even and n-1 is not a p-power,

  3. βτη, where 1βϕ2, if (b)2>2 and n cannot be represented as 1+pt-1 or 1+pt-1+2u with t,u1,

  4. ϕ-κη, where pκ(mod4), if (p-κ)2>(n)2 and n cannot be represented as 1+pt-1 or 1+pt-1+2u with t,u1,

  5. φm/2τη, if (b)2>2, n=1+pt-1 for some t1 and n cannot be represented as 2+2u with u1,

  6. (φm/2)-κη, where pκ(mod4), if (m)2=2, (p-κ)2>(n)2, n=1+pt-1 for some t1 and n cannot be represented as 2+2u with u1.

Proof.

Observe that b is even if and only if all the numbers n, m and (q-1)/d are even.

Suppose that n is odd. Then S2=φ¯(m)2×τ¯. By Lemma 5.7, any non-trivial element of φ(m)2 is not admissible and thus the only potentially admissible subgroup of S2 is τ¯.

Suppose that n is even but at least one of m and (q-1)/d is odd. It follows from Lemmas 5.7 and 5.8 that every admissible subgroup of S2 is contained in η¯τ¯. Then by Lemma 5.1, it is conjugate to a subgroup of τ¯, and we are done in the case of odd b.

Suppose now that all the numbers n, m and (q-1)/d are even. In particular, (q-1)2>(d)2=(n)2 and so τ is not admissible by Theorem 4.7. Then by Lemma 5.1, we obtain that all non-trivial elements of η¯τ¯ are not admissible. It follows that every admissible subgroup of S2 is isomorphic to a subgroup of φ¯(m)2, and hence is cyclic.

Let α¯S2η¯τ¯ be admissible. We may assume that α=βϵηj, where β=φm/k for some 1<k=2l(m)2. By Lemmas 5.7 and 5.8, admissibility of α yields k(q-1)2/(d)2 or, equivalently, βϕ. Moreover, all conjugates of β¯ϵ are admissible whenever n-1 is not a p-power, and this gives the automorphisms in (ii).

It remains to consider the case when α¯ is not conjugate to β¯ϵ. In this case by Lemma 5.8, we may assume that α=βϵη,

(5.4)2k(q-1)2(d)2

and q0-ϵ(mod4), where q=q0k. Moreover, we have k=2 whenever n-1 is a p-power. Note that (q-1)2=(m)2(p-κ)2, where pκ(mod4).

Suppose that (q-1)2/(d)2(m)2 or, in other words, (p-κ)2(n)2. Then k(m)2/2, so q01(mod4) and condition (5.4) is equivalent to βϕ2. Thus (b)2>2 and α=βτη with βϕ2. In particular, if n-1 is a p-power, then α=φm/2τη.

Suppose that (q-1)2/(d)2>(m)2 or, equivalently, (p-κ)2>(n)2. Then condition (5.4) holds for all βϕ. Also, we have q01(mod4) if k<(m)2 and q0κ(mod4) if k=(m)2. Thus α=βτη with βϕ2 or α=ϕ-κη. In particular, if n-1 is a p-power, then α=φm/2τη when (b)2=(m)2>2 and α=ϕ-κη when (b)2=(m2)=2.

Conversely, let α be one of the resulting automorphisms and let the associated conditions be satisfied. Then by Lemma 5.8, α is admissible unless n cannot be represented as 1+pt-1+2u. If n=1+pt-1 and n=1+ps-1+2u, then we have pt-1=ps-1+2u or, equivalently, ps-1(pt-s-1)=2u, and so s=1 and n=2+2u. Thus we obtain the automorphisms in (iii)–(vi), and the proof is complete. ∎

Proof of Theorem 2.

Let ω(G)=ω(L). By Lemma 5.4, it follows that G/L is conjugate to a subgroup of ψ¯×S2. Admissible subgroups of ψ¯ and S2 are described in Lemmas 5.2, 5.3 and Lemma 5.9, respectively. These descriptions together with Lemmas 5.7 and 5.8 imply that the product of an admissible subgroup of ψ¯ and an admissible subgroup of S2 is also admissible, and this completes the proof. ∎


Communicated by Christopher W. Parker


Funding statement: The work is supported by Russian Science Foundation (project 14-21-00065).

References

[1] A. S. Bang, Taltheoretiske undersøgelser, Tidsskrift Math. 4 (1886), 70–80, 130–137. Search in Google Scholar

[2] R. Brandl and W. J. Shi, The characterization of PSL(2,q) by its element orders, J. Algebra 163 (1994), no. 1, 109–114. 10.1006/jabr.1994.1006Search in Google Scholar

[3] J. Bray, D. Holt and C. Roney-Dougal, The Maximal Subgroups of the Low-Dimensional Finite Classical Groups, London Math. Soc. Lecture Note Ser. 407, Cambridge University Press, Cambridge, 2013. 10.1017/CBO9781139192576Search in Google Scholar

[4] A. A. Buturlakin, Spectra of finite linear and unitary groups, Algebra Logic 47 (2008), no. 2, 91–99. 10.1007/s10469-008-9003-3Search in Google Scholar

[5] A. A. Buturlakin, Spectra of finite symplectic and orthogonal groups, Siberian Adv. Math. 21 (2011), no. 3, 176–210. 10.3103/S1055134411030035Search in Google Scholar

[6] A. A. Buturlakin and M. A. Grechkoseeva, The cyclic structure of maximal tori of the finite classical groups, Algebra Logic 46 (2007), no. 2, 73–89. 10.1007/s10469-007-0009-zSearch in Google Scholar

[7] J. H. Conway, R. T. Curtis, S. P. Norton, R. A. Parker and R. A. Wilson, Atlas of Finite Groups, Clarendon Press, Oxford, 1985. Search in Google Scholar

[8] J. Fulman and R. Guralnick, Bounds on the number and sizes of conjugacy classes in finite Chevalley groups with applications to derangements, Trans. Amer. Math. Soc. 364 (2012), no. 6, 3023–3070. 10.1090/S0002-9947-2012-05427-4Search in Google Scholar

[9] J. Fulman and R. Guralnick, Conjugacy class properties of the extension of GL(n,q) generated by the inverse transpose involution, J. Algebra 275 (2004), no. 1, 356–396. 10.1016/j.jalgebra.2003.07.004Search in Google Scholar

[10] R. Gow and C. R. Vinroot, Extending real-valued characters of finite general linear and unitary groups on elements related to regular unipotents, J. Group Theory 11 (2008), no. 3, 299–331. 10.1515/JGT.2008.017Search in Google Scholar

[11] M. A. Grechkoseeva, Recognition by spectrum for finite linear groups over fields of characteristic 2, Algebra Logic 47 (2008), no. 4, 229–241. 10.1007/s10469-008-9014-0Search in Google Scholar

[12] M. A. Grechkoseeva, On spectra of almost simple groups with symplectic or orthogonal socle, Sib. Math. J. 57 (2016), no. 4, 582–588. 10.1134/S0037446616040029Search in Google Scholar

[13] M. A. Grechkoseeva and W. J. Shi, On finite groups isospectral to finite simple unitary groups over fields of characteristic 2, Sib. Élektron. Mat. Izv. 10 (2013), 31–37. Search in Google Scholar

[14] M. A. Grechkoseeva and A. V. Vasil’ev, On the structure of finite groups isospectral to finite simple groups, J. Group Theory 18 (2015), no. 5, 741–759. 10.1515/jgth-2015-0019Search in Google Scholar

[15] P. Kleidman and M. Liebeck, The Subgroup Structure of the Finite Classical Groups, London Math. Soc. Lecture Note Ser. 129, Cambridge University Press, Cambridge, 1990. 10.1017/CBO9780511629235Search in Google Scholar

[16] V. D. Mazurov, On the set of orders of elements of a finite group, Algebra Logic 33 (1994), no. 1, 49–55. 10.1007/BF00739417Search in Google Scholar

[17] V. D. Mazurov, Recognition of finite groups by a set of orders of their elements, Algebra Logic 37 (1998), no. 6, 371–379. 10.1007/BF02671691Search in Google Scholar

[18] P. Ribenboim, The Little Book of Bigger Primes, 2nd ed., Springer, New York, 2004. Search in Google Scholar

[19] M. Roitman, On Zsigmondy primes, Proc. Amer. Math. Soc. 125 (1997), no. 7, 1913–1919. 10.1090/S0002-9939-97-03981-6Search in Google Scholar

[20] D. E. Taylor, The Geometry of the Classical Groups, Sigma Ser. Pure Math. 9, Heldermann, Lemgo, 2009. Search in Google Scholar

[21] A. V. Vasil’ev, On finite groups isospectral to simple classical groups, J. Algebra 423 (2015), 318–374. 10.1016/j.jalgebra.2014.10.013Search in Google Scholar

[22] G. E. Wall, On the conjugacy classes in the unitary, symplectic and orthogonal groups, J. Aust. Math. Soc. 3 (1963), 1–62. 10.1017/S1446788700027622Search in Google Scholar

[23] A. V. Zavarnitsine, Recognition of the simple groups L3(q) by element orders, J. Group Theory 7 (2004), no. 1, 81–97. 10.1515/jgth.2003.044Search in Google Scholar

[24] A. V. Zavarnitsine, Recognition of the simple groups U3(q) by element orders, Algebra Logic 45 (2006), no. 2, 106–116. 10.1007/s10469-006-0009-4Search in Google Scholar

[25] K. Zsigmondy, Zur Theorie der Potenzreste, Monatsh. Math. Phys. 3 (1892), 265–284. 10.1007/BF01692444Search in Google Scholar

[26] M. A. Zvezdina, Spectra of automorphic extensions of finite simple symplectic and orthogonal groups over fields of characteristic 2, Sib. Élektron. Mat. Izv. 11 (2014), 823–832. Search in Google Scholar

[27] M. A. Zvezdina, Spectra of automorphic extensions of finite simple exceptional groups of Lie type, Algebra Logic 55 (2016), 354–366. 10.1007/s10469-016-9407-4Search in Google Scholar

Received: 2016-9-7
Revised: 2017-2-6
Published Online: 2017-4-8
Published in Print: 2017-11-1

© 2017 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 7.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/jgth-2017-0010/html
Scroll to top button