Skip to content
Publicly Available Published by De Gruyter March 28, 2020

Isomorphism questions for metric ultraproducts of finite quasisimple groups

  • Jakob Schneider EMAIL logo
From the journal Journal of Group Theory

Abstract

New results on metric ultraproducts of finite simple groups are established. We show that the isomorphism type of a simple metric ultraproduct of groups Xni(q) (iI) for X{PGL,PSp,PGO(ε),PGU} (ε=±) along an ultrafilter 𝒰 on the index set I for which ni𝒰 determines the type X and the field size q up to the possible isomorphism of a metric ultraproduct of groups PSpni(q) and a metric ultraproduct of groups PGOni(ε)(q). This extends results of [A. Thom and J. Wilson, Metric ultraproducts of finite simple groups, Comp. Rend. Math. 352 2014, 6, 463–466].

1 Introduction

In [5, 6], Thom and Wilson discussed various properties of metric ultraproducts of finite simple groups. In particular, they asked which such ultraproducts can be isomorphic. In [5, Theorem 2.2], a metric ultraproduct of alternating groups is distinguished from a metric ultraproduct of classical groups of Lie type, where the permutation degrees resp. dimensions of the natural module tend to infinity. This is done by considering the structure of centralizers of torsion elements in these groups (see [5, Theorems 2.8 and 2.9]). In the case of a metric ultraproduct of classical groups of Lie type, in [5, Theorem 2.8], investigating the structure of such centralizers of semisimple and unipotent torsion elements, Thom and Wilson even extract the “limit characteristic” of the group. At the end of [5, Section 2], they ask which metric ultraproducts of classical groups of different types can be isomorphic.

In this note, we will give an almost complete answer to this question in the case when the field sizes are bounded. We will show that, for a metric ultraproduct of alternating or classical groups of Lie type of unbounded rank over fields of bounded size, one can extract the Lie type (up to one exception). Also one can extract the “limit field size”. Our results are summed up in Theorem 1 below. To state it, we first need to introduce some notation which we fix for the whole article.

Let =(Hi)iI be a sequence of groups where either Hi=Sni is a symmetric group or Hi=Xi(qi) is a classical group of Lie type Xi over the finite field 𝔽qi with qi elements (resp. 𝔽qi2 in the unitary case, iI). In the latter case, we let each Xi be one of GLni, Sp2mi, GO2mi±, GO2mi+1 (qi odd), or GUni for suitable mi,ni+ (iI).

Recall that a norm on a group H is a function H[0,] such that (h)=0 iff h=1H, (h)=(h-1)=(hg), and (gh)(g)+(h) for all g,hH. Call a pair (H,), where H is a group and a norm on it, a normed group. Recall that the metric ultraproduct of a sequence of normed groups (Hi,i)iI along an ultrafilter 𝒰 on the set I is defined as the quotient iIHi/N, where N:={(hi)iIiIHilim𝒰i(hi)=0} is the normal subgroup of 𝒰-null sequences.

Throughout, let G:=𝒰met be the metric ultraproduct of the groups Hi from above equipped with the normalized Hamming norm

H(σ):=|supp(σ)|/n=|{x{1,,n}x.σx}|/n

resp. the normalized rank normrk(g):=rk(1-g)/n when Hi is a symmetric resp. a classical linear group of Lie type along the ultrafilter 𝒰. Assume that the permutation degrees resp. dimensions of the natural module ni of Hi (iI) tend to infinity along 𝒰.

Note that, since 𝒰 is an ultrafilter, we may assume that each group Hi is of the same type, i.e., all groups Hi are either symmetric, linear, symplectic, orthogonal, or unitary groups. In these five distinct cases, we write S𝒰, GL𝒰, Sp𝒰, GO𝒰, or GU𝒰 for G. Also, when the field sizes qi are bounded, we may assume that qi=q is constant (iI), setting q:=lim𝒰qi. Throughout the article, set Z:=𝐙(G) to be the center of G and G¯:=G/Z.

If the groups Hi (iI) are symmetric groups, then Z=𝟏 and G¯=G. Now assume that all groups Hi are of type X(qi) (iI; i.e., they are not symmetric groups). Then G¯=G/Z=¯𝒰met is the metric ultraproduct of the groups H¯i:=Hi/𝐙(Hi) with respect to the projective rank norm

pr(h¯):=inf{rk(h)his a lift ofh¯inH},

which is defined on the general projective linear group. By the results from [3], G¯ is the unique simple quotient of G.

Similarly to the above, write PGL𝒰, PSp𝒰, PGO𝒰, or PGU𝒰 for G¯ when all the groups Hi (iI) are linear, symplectic, orthogonal, or unitary groups. Moreover, in this case, if all the fields 𝔽qi (iI) are equal to 𝔽q (or 𝔽q2 in the unitary case), we write GL𝒰(q), Sp𝒰(q), GO𝒰(q), GU𝒰(q) resp. PGL𝒰(q), PSp𝒰(q), PGO𝒰(q), PGU𝒰(q) for G resp. G¯. Write GL𝒰(k) resp. PGL𝒰(k) for the metric ultraproduct of groups GLni(k) resp. PGLni(k) over the field k (iI) along 𝒰. Write 𝐌n(k) for the matrix ring of degree n over the field k and P𝐌n(k) for the associated projective space (𝐌n(k){0})/k×. Set 𝐌n(q):=𝐌n(𝔽q) and P𝐌n(q):=P𝐌n(𝔽q). Also write 𝐌𝒰, 𝐌𝒰(q) resp. P𝐌𝒰, P𝐌𝒰(q) for the metric ultraproduct of the spaces 𝐌ni(qi), 𝐌ni(q) resp. P𝐌ni(qi), P𝐌ni(q) with respect to the metrics

drk(g,h):=rk(g-h)/n,
dpr(g¯,h¯):=inf{drk(g,h)g,hare lifts ofg¯,h¯}(iI)

so that GL𝒰𝐌𝒰, GL𝒰(q)𝐌𝒰(q), PGL𝒰P𝐌𝒰, PGL𝒰(q)P𝐌𝒰(q).

The main result of this article is now as follows.

Theorem 1.

Let G¯G¯1G¯2 with Gj=XjUj(qj), where

Xj{GL,Sp,GO,GU}(j=1,2).

Then it holds that q1=q2. Also we must have X1=X2 or {X1,X2}={Sp,GO}. Moreover, an ultraproduct X¯1U1 where the sizes qi of the finite fields Fqi (iI1) tend to infinity along U1 cannot be isomorphic to an ultraproduct X¯2U2(q).

Let us conclude this introduction by saying some words about the proof of Theorem 1. Our strategy is to compute double centralizers of semisimple torsion elements of a fixed order o+ in the above metric ultraproducts. If the sizes qi (iI) of the fields 𝔽qi are bounded, it turns out that these are always finite abelian groups. Then we consider the maximal possible exponent which such a double centralizer can have. It turns out that this data is enough to determine the limit field size q and the Lie type (up to the exception mentioned in Theorem 1). If the field sizes qi (iI) tend to infinity, a double centralizer of such a torsion element of order o>2 is always infinite. This separates this case from the former one. Throughout the article, we will make frequent use of the exposition given in [4, Subsection 3.4.2, § 1]. This paragraph in the author’s PhD thesis provides a concise presentation of the classification of the possible indecomposable blocks of elements from classical groups of Lie type stabilizing a form (which is, of course, well known). For the convenience of the reader, we will quote the facts needed from it in the subsequent section.

2 Notation and auxiliary results from [4]

In this section, we introduce the necessary notation and quote the facts from [4] which we need for the proof of Theorem 1. Keep the notation from the introduction. Subsequently, G will be one of S𝒰, GL𝒰(q), Sp𝒰(q), GO𝒰(q), or GU𝒰(q) (so that G¯ is either S𝒰, PGL𝒰(q), PSp𝒰(q), PGO𝒰(q), or PGU𝒰(q), respectively). We refer to these five distinct cases as the symmetric case, the linear case, the symplectic case, the orthogonal case, and the unitary case, respectively. The last four cases are grouped together as the classical case. The symplectic and orthogonal case are grouped together as the bilinear case.

For the rest of this section, we assume that we are in the classical case. Then set k=𝔽q in the non-unitary case and k=𝔽q2 in the unitary case. Write p:=char(k) for the characteristic of the field k.

Subsequently, assume that we are not in the linear case. Then we let σ:kk be the q-Frobenius map xxq, i.e., σ=idk in the bilinear case, whereas σ is the unique involution of k=𝔽q2 in the unitary case. For a monic polynomial

χ=a0+a1X++ak-1Xk-1+Xkk[X],

write χ=a0-σXdeg(χ)χσ(X-1) for the dual polynomial of χ. Here

χσ=a0σ+a1σX++ak-1σXk-1+Xk

is the polynomial χ with coefficients twisted by the automorphism σ. Call χ self-dual iff χ=χ. Furthermore, write fi for the sesquilinear forms stabilized by the group Hi (and Qi for the quadratic forms in the orthogonal case for p=2, iI). Note that a metric ultraproduct of groups GO2mi+1(q) for q a power of p=2 with respect to the projective rank norm where the mi tend to infinity along the chosen ultrafilter is isomorphic to such an ultraproduct of groups Sp2mi(q). Actually, GO2mi+1(q)Sp2mi(q), but the natural module of GO2mi+1(q) has one-dimensional radical rad(fi), whereas rad(fi)=0 in the second case. Hence we may and do assume that the forms fi (iI) are all non-singular and exclude groups of type GO2m+1(q) for q even. Write

F(χ):=(01000001-a0-ak-1)

for the Frobenius block associated to the monic polynomial

χ=a0+a1X++ak-1Xk-1+Xkk[X],

i.e., multiplication by X¯ in the quotient ring k[X]/(χ) represented in the basis

1,X¯,,X¯k-1.

Recall that, by the existence of the generalized Jordan normal (or primary rational canonical form) form, each linear transformation g𝐌n(k) can be written as

gχk[X]χmonic primaryF(χ)cχ(g),

where the cχ(g) are uniquely determined. A polynomial χk[X] is here called primary iff it is the power ie (e+) of an irreducible polynomial ik[X].

Here are the results needed from [4]. Subsequently, the group H is one of the groups Hi (iI), and the form f is alternating bilinear, symmetric bilinear, or σ-sesquilinear in the symplectic, orthogonal, and unitary case, respectively. Hence the sign ε of f is -1 in the symplectic case, and ε=+1 in the orthogonal and unitary case. The form Q is a compatible quadratic form in the orthogonal case for p=2, i.e., Q(u+v)=Q(u)+f(u,v)+Q(v) for all u,v. Recall that two sesquilinear forms f and f (or quadratic forms Q and Q) on a vector space V are called linearly equivalent if f(u,v)=f(u.g,v.g) (resp. Q(v)=Q(v.g)) for all u,vV and some gGL(V).

Fact 1 ([4, Fact 3.40]).

The following hold.

  1. Whenever χk[X] is (monic) self-dual and is not divided by X±1 in the bilinear case, then there exists a non-singular form f (coming from a quadratic form Q when p=2) which is preserved by F(χ) (and Q as well). The form f (and Q) is even unique up to linear equivalence.

  2. Assume now that χ=(X±1)e (which is self-dual) and we are in the bilinear case.

    1. At first, assume that p is odd. Then, when ε=(-1)e-1, there is a unique non-singular form f (up to linear equivalence) preserved by F(χ), and when ε=(-1)e, there is a unique non-singular form f (up to linear equivalence) preserved by F(χ)2, but no non-singular form is preserved by F(χ).

    2. If p=2, we use [2, Theorem 3.1] to see that if gF(χ)c, i.e., g acts χ -isotypically (see the beginning of Section 3), then any non-singular form f (and the associated quadratic form Q) restricts to all but at most constantly many blocks F(χ) to the indecomposable W(e) from this theorem (which acts as F(χ)2).

An indecomposable block of an element g from a classical group of Lie type stabilizing the sesquilinear form f with natural module V is a subspace W0 on which f is non-singular (i.e., WW=0) and which is minimal with this property. Out of the theory of indecomposables presented in [4, Subsection 3.4.2, § 1], one can also deduce the following fact which we will use in the text.

Fact 2.

Let H=GI(V,f) (resp. H=GI(V,Q) in the orthogonal case for p=2) be a classical group of Lie type stabilizing the non-singular form f (and Q). Let gH and h:UW be a partial isometry between the subspaces U,WV commuting with g. Then h extends to a full isometry hH which commutes with g.

The next result, which can also be deduced from [4, Subsection 3.4.2, § 1], describes the shape of f on a “semisimple” indecomposable block.

Fact 3.

Let H be as in Fact 2, and let gH be a semisimple element, i.e., for each Frobenius block F(χ) in the generalized Jordan normal form of g (see above), we have that χ=i1 is irreducible (or alternatively the order ord(h) of h is coprime to the characteristic p of the group H). Then the indecomposable blocks of g look as follows. Assume WV is an indecomposable block of g such that g|W has a Jordan block F(χ) acting on the subspace UW of dimension deg(χ) for χk[X] irreducible. Write Kχ:=k[X]/(χ), and let λχ:=X¯Kχ be a root of χ in Kχ. Let τ:KχKχ be the field isomorphism given by τ|k=σ and λχλχ-1. If χ is self-dual, i.e., χ=χ, we clearly have that τAut(Kχ/kσ) is an involution (or the identity in the bilinear case when χ=X±1).

  1. Assume χX±1 in the bilinear case.

    1. Assume χ is not self-dual. There exists UW such that W=UU and g acts as F(χ)F(χ) on UU. So, as k[g]-modules, UKχ and UKχ, where g acts on Kχ×KχUU=W as multiplication by (λχ,λχ). The pairing f on W is given by

      U×UKχ×Kχ(u,v)trKχ/k(uvτ),

      f|U=0, f|U=0, and extension to a sesquilinear form of sign ε (this is clearly g-invariant). In the orthogonal case for p=2, the associated quadratic form Q is uniquely determined by f and g. It is given by W=UUuvf(u,v).

    2. Assume now that χ is self-dual. Then U=W and g acts as F(χ) on U. Hence UKχ as a k[g]-module, where g acts on KχU as multiplication by λχ. In this case, the form f is given by

      U2Kχ2(u,v)trKχ/k(δuvτ),

      where δKχ is such that δτ=εδ. In the orthogonal case for p=2, the quadratic form Q is given by Uvtr(Kχ)τ/k(vvτ). Here (Kχ)τ denotes the fixed field of the involution τ.

  2. Assume now that χ=X±1 and we are in the bilinear case. Then we have g|W=idW.

    1. If we are in the symplectic case, then Wk2=𝔽q2 and f|W is the unique (up to linear equivalence) non-singular alternating form on the vector space W.

    2. Assume we are in the orthogonal case and p2. Then Wk=𝔽q and f is given by (u,v)βuv, where β𝔽q× is either 1 (the “standard square”) or a “standard non-square”.

    3. Assume we are in the orthogonal case and p=2. Then Wk2=𝔽q2 and f is the unique (up to linear equivalence) non-singular alternating form on it and Q is a quadratic form of plus type or minus type, i.e., either Q is given by k2(u,v)uv (plus type) or

      k2𝔽q2vN𝔽q2/𝔽q(v)=vq+1,

      where N𝔽q2/𝔽q:𝔽q2𝔽q is the field norm.

We will also make use of the relations presented in the next remark.

Remark 1.

When D is the block Wk2 of case (ii) (b) of Fact 3 where β=1 and E is the same block for a non-square β𝔽q×, then DDEE. Similarly, when D is the block of minus type and E the block of plus type on Wk2 in Fact 3 (ii) (c), then DDEE.

3 Description of conjugacy classes in S𝒰, GL𝒰(q), and PGL𝒰(q)

In this section, we give a description of the conjugacy classes of groups of type S𝒰 or PGL𝒰(q). We will make use of this in the subsequent sections.

We start with some definitions. Write Sn=Sym(n¯) for the symmetric group acting on the set n¯:={1,,n}. Let Ck(σ)Sn (for a permutation σSn and k+) denote the set of all k-cycles of σ, and let ck(σ):=|Ck(σ)| denote the number of k-cycles of σ. Moreover, let Ωk(σ) be the set on which all k-cycles of σ act. Set nk(σ):=|Ωk(σ)|=kck(σ) to be the cardinality of Ωk(σ). Call the permutation σSnk-isotypic when nk(σ)=n, i.e., σ has only cycles of length k. Call σ isotypic if it is k-isotypic for one number k+. Similarly, for g𝐌n(k) and a monic primary polynomial χk[X], let cχ(g) be the number of Frobenius blocks F(χ) in the generalized Jordan normal form of g (see Section 2). Let Vχ(g) be the subspace on which g acts as F(χ)cχ(g) with respect to this normal form, and set nχ(g):=dim(Vχ(g)). Clearly, Vχ(g) is not uniquely determined as it depends on the decomposition of the vector space kn associated to the chosen generalized Jordan normal form of g. However, by writing Vχ(g) we mean that these subspaces of kn are the decomposition of kn coming from such a normal form of g. Then, of course, nχ(g)=deg(χ)cχ(g). For a monic polynomial χk[X], say that g is χ-isotypic if gF(χ)c for some c+.

At first, we consider GL𝒰(q) instead of PGL𝒰(q). Subsequently, all polynomials from k[X] that occur in the text are meant to be monic polynomials.

Conjugacy classes in S𝒰 and GL𝒰(q). For an integer k+ and a polynomial ξk[X], define rk(σ):=|fix(σk)|/nresp.rξ(g):=dim(ker(ξ(g)))/n for σSn resp. g𝐌n(q). Here fix(σ):={xn¯x.σ=x} is the set of fixed points of the permutation σ. Extend this definition to S𝒰 and 𝐌𝒰(q) by setting

rk(σ):=lim𝒰rk(σi)andrξ(g):=lim𝒰rξ(gi)

for σ=(σi)¯iI resp. g=(gi)¯iI. Both expressions are well-defined since, for σ=(σi)¯iI=(τi)¯iIS𝒰, one has

(3.1)1ni||fix(σik)|-|fix(τik)||dH(σik,τik)dH(σik,σik-1τi)++dH(σiτik-1,τik)=kdH(σi,τi)𝒰0.

Similarly, if g=(gi)¯iI=(hi)¯iIGL𝒰(q) and

ξ=a0+a1X++ak-1Xk-1+Xkk[X],

we have

(3.2)1ni|dim(kerξ(gi))-dim(kerξ(hi))|drk(ξ(gi),ξ(hi))j=0kdrk(ajgij,ajhij)j=0kdrk(gij,hij)(j=0kj)drk(gi,hi)=(k+12)drk(gi,hi),

and the latter tends to zero along 𝒰. Here we used the same trick as in estimate (3.1) above to bound drk(gij,hij) by jdrk(gi,hi) (j=0,,k) in estimate (3.2). Write r(σ):=(rk(σ))k+ and r(g):=(rξ(g))ξk[X]. Now define qk(σ) for k+ via the equality

dkqd(σ)=rk(σ)for allk+.

Write q(σ):=(qk(σ))k+. Applying Möbius inversion, we obtain

qk(σ)=dkμ(k/d)rd(σ).

Alternatively, one can think of qk(σ) as the 𝒰-limit of the normalized cardinality of the support of all k-cycles in σi (iI), i.e.,

qk(σ)=lim𝒰nk(σi)/ni=klim𝒰ck(σi)/ni.

Similarly, for χ=iek[X] primary, define qχ(g) via the equality

rξ(g)=χprimarydeg(gcd{χ,ξ})deg(χ)qχ(g)

for all polynomials ξk[X]. Here gcdS is the greatest common divisor of the elements from S. Write q(g):=(qχ(g))χprimary. Alternatively, one can think of qχ(g) as the 𝒰-limit of the normalized dimensions of the (not unique) subspaces Vχ(gi) (iI), i.e.,

qχ(g)=lim𝒰nχ(gi)/ni=kχlim𝒰cχ(gi)/ni,

where kχ=deg(χ)=edeg(i). This is because, when g acts as F(χ) and ξk[X], then dim(ker(ξ(g)))=deg(gcd{χ,ξ}).

We claim that the conjugacy classes in S𝒰 resp. 𝐌𝒰(q) are in one-to-one correspondence with all tuples (qk(σ))k+ resp. (qχ(g))χprimary, where the only condition on the sequences are that

k+qk(σ)1resp.χprimaryqχ(g)1.

Here we let GL𝒰(q) act on 𝐌𝒰(q) by conjugation. The element g lies in GL𝒰(q) iff qχ(g)=0 for χ=Xe (e1). Indeed, one sees easily that r(σ) resp. r(g) is conjugacy invariant, and so is q(σ) resp. q(g) for σS𝒰 and g𝐌𝒰(q).

To see the converse, let

σ=(σi)¯iI,τ=(τi)¯iIS𝒰,
g=(gi)¯iI,h=(hi)¯iI𝐌𝒰(q)

be elements such that q(σ)=q(τ) and q(g)=q(h), respectively.

Find a sequence (Ni)iI tending to infinity along 𝒰 such that

k=1Ni|qk(σ)-qk(σi)|,k=1Ni|qk(τ)-qk(τi)|𝒰0

resp. such that

χprimarydeg(χ)Ni|qχ(g)-qχ(gi)|,χprimarydeg(χ)Ni|qχ(h)-qχ(hi)|𝒰0.

Then, by the triangle inequality,

1nik=1Ni|nk(σi)-nk(τi)|𝒰0

resp.

1niχprimarydeg(χ)Ni|dim(Vχ(gi))-dim(Vχ(hi))|𝒰0.

Hence we can conjugate a big part of k=1NiΩk(σi) equivariantly to a big part of k=1NiΩk(τi) resp. an almost full-dimensional part of

χprimarydeg(χ)NiVχ(gi)toχprimarydeg(χ)NiVχ(hi)

equivariantly (with no error in the limit; here again Vχ(gi) resp. Vχ(hi) are not unique). The remaining part of σi and τi resp. gi and hi can be modified into one big cycle resp. one big Frobenius block with no change of σ and τ resp. g and h, since Ni𝒰. Then we conjugate this cycle of σi resp. Frobenius block of gi onto the one of τi resp. hi.

The case of P𝐌𝒰(q). Let the group k×=𝔽q× act on all (monic) polynomials ξk[X] by ξ.z:=z-kξξ(zX), where kξ:=deg(ξ). Extend this action to all tuples q=(qχ)χprimary with χprimaryqχ1 via

q.z=(qχ.z)χprimary,

and denote by q¯ the orbit orbk×(q) of q under this action of k×.

Suppose that G¯=PGL𝒰(q). We claim that the conjugacy classes of elements g¯P𝐌𝒰(q) are classified by the bijection g¯G¯(qχ(g))¯χprimary, where g is any lift of g¯ in 𝐌𝒰(q) (here we exclude the tuple q where qX=1 and qχ=0 otherwise).

Indeed, the map is well-defined since any other h such that h¯=g¯P𝐌𝒰(q) is of the form zg for some zk× (as k=𝔽q is finite) so that

(qχ(h))χprimary=(qχ(zg))χprimary=(qχ.z(g))χprimary=q(g).z.

Also q is constant on conjugacy classes of 𝐌𝒰(q) (under the action of GL𝒰(q)). Conversely, if qχ(h)=qχ.z(g)=qχ(zg) for some fixed zk× and all χk[X] primary, then, from the above, we derive that the elements g and z-1h of 𝐌𝒰(q) are conjugate so that g¯ and h¯ are conjugate in P𝐌𝒰(q). This proves the claim.

Remark 2.

For G of type Sp𝒰(q), GO𝒰(q), or GU𝒰(q), the conjugacy classes of elements gG for which χprimaryqχ(g)=1 are still characterized by the tuples (qχ(g))χprimary. The only necessary additional restriction on these tuples is that qχ(g)=qχ(g) for all χk[X] primary (see the beginning of Section 2 for the definition of χ).

Indeed, assume g=(gi)¯iI,h=(hi)¯iIG, q(g)=q(h), qχ(g)=qχ(g) for all χk[X] primary, and

χprimaryqχ(g)=χprimaryqχ(h)=1.

Then g is conjugate to h.

This holds since on all but constantly many Frobenius blocks F(χ) of gi resp. hi (iI), where χ is a self-dual primary polynomial, and on all but constantly many blocks F(χ)F(χ)F(χχ), where χ is not self-dual and primary, according to Fact 1, the form fi (and in characteristic p=2 the quadratic form Qi) is uniquely determined up to linear equivalence so that we can map these blocks of gi to such blocks of hi and extend this partial map by Witt’s lemma.

Conversely, if we have a tuple (qχ)χprimary such that χprimaryqχ=1 and qχ=qχ, we can deduce again from Fact 1 that there exists an element gG such that qχ(g)=qχ for all χk[X] primary.

Recall that G¯=G/Z, where Z=𝐙(G). For a tuple q=(qχ)χprimary, let q¯ denote its orbit orbZ(q) under the action of Zk×. Then, for the elements g¯G¯ (i.e., G¯ is PSp𝒰(q), PGO𝒰(q), or PGU𝒰(q)) such that χprimaryqχ(g)=1 for one lift gG of g¯, the same characterization as for PGL𝒰(q) above holds by the same argument. Again we need to restrict the tuples q=(qχ)χprimary so that qχ=qχ for all χ primary.

However, we conjecture that the above characterization for G of type Sp𝒰(q), GO𝒰(q), or GU𝒰(q) is false if

χprimaryqχ(g)<1for an elementgG.

Remark 3.

It is easy to see that (χ.z)=χ.z for zZk×. Indeed, since zσ=z-1 for zZ, we have

(χ.z)=(z-kχ(zX))=(zk)σa0-σXk(z-k)σχσ(zσX-1)=a0-σXkχσ((zX)-1)=z-kχ(zX)=χ.z,

where χ=a0+a1X++ak-1Xk-1+Xk (recall that σAut(k) is defined in Section 2).

4 Characterization of torsion elements in S𝒰, GL𝒰(q), and PGL𝒰(q)

In this section, we characterize torsion elements in metric ultraproducts of the above type. At first, note that an invertible element in 𝐌𝒰(q), i.e., an element of GL𝒰(q), is algebraic over k=𝔽q if and only if it is torsion. Indeed, if g is torsion, then go-1=0 for some integer o1. Conversely, if g is algebraic and invertible, let χk[X] be its minimal polynomial. Setting o:=|(k[X]/(χ))×|<, one sees that go=1 as g is invertible.

Here comes the promised characterization of torsion elements.

Lemma 1.

An element σSU resp. gGLU(q) is torsion if and only if there is NZ+ such that

k=1Nqk(σ)=1resp.χ𝑝𝑟𝑖𝑚𝑎𝑟𝑦deg(χ)Nqχ(g)=1.

An element g¯PGLU(q) is torsion if and only if any lift gGLU(q) is torsion.

Proof.

Write lcmS for the least common multiple of the elements of S. Indeed, if the above two conditions are fulfilled, then writing o:=lcm{1,,N} resp. o:=lcm{|(k[X]/(χ))×|χprimary,deg(χ)N}, we have H(σio)𝒰0 resp. rk(gio)𝒰0, meaning that σo=1 resp. go=1.

Conversely, if we assume σo=1 resp. go=1, we get that H(σio)𝒰0 resp. rk(gio)𝒰0, meaning that, asymptotically, the d-cycles in σi for do support the whole set resp. all Frobenius blocks F(χ) for χXo-1 primary support the whole vector space, so taking N:=o above, we get the converse direction.

The last statement follows since the kernel of the surjective homomorphism GL𝒰(q)PGL𝒰(q) equals k×=𝔽q×, which is finite. Hence, if gGL𝒰(q) represents g¯PGL𝒰(q) and the latter is of order o<, then we have that ord(g)o(q-1)<. ∎

5 Faithful action of S𝒰 and PGL𝒰(q) on the Loeb space and the associated continuous geometry

In this section, we show that the groups S𝒰 and PGL𝒰(q) faithfully act on natural associated objects. For this purpose, we need the so-called Loeb space

L(ni)iI:=(𝒮,μ)

resp. its vector space analog, the continuous geometry

V(ni)iI:=(𝒱,dim),

which are associated naturally to the metric ultraproduct S𝒰 resp. PGL𝒰(q).

Here 𝒮 resp. 𝒱 equals iI𝒫(ni¯) resp. iISub(kni) modulo the equivalence relation (Si)iI(Ti)iI resp. (Ui)iI(Vi)iI iff μi(SiTi)𝒰0 resp. dimi(Ui+Vi)-dimi(UiVi)𝒰0, where μi resp. dimi is the normalized counting measure resp. dimension on ni¯ resp. kni (and AB denotes the symmetric difference of the sets A and B). Then one defines μ resp. dim by

μ(S)=μ((Si)¯iI):=lim𝒰μi(Si),
dim(V)=dim((Vi)¯iI):=lim𝒰dimi(Vi).

It is easy to check that both are well-defined in this way. Also the operations , resp. +, are inherited to 𝒮 resp. 𝒱 in a natural way, e.g.,

(Si)¯iI(Ti)¯iI:=(SiTi)¯iI.

Write ST resp. UV iff μ(ST)=μ(S) resp. dim(UV)=dim(U). Call a permutation of 𝒮 resp. 𝒱 an automorphism iff it preserves μ resp. dim and the relation resp. .

Then one observes that S𝒰 resp. PGL𝒰(q) is faithfully represented as group of automorphisms of (𝒮,μ) resp. (𝒱,dim).

At first, consider the case G¯=S𝒰. Indeed, assume for fixed σ=(σi)¯iIS𝒰 that S.σ=S for all S𝒮. Then take S=(Si)¯iI, where Sini¯ is taken in the following way. For each k-cycle cni¯ (k>1; here seen as a set), we pick scc and define Si by Sic={sc,sc.σi2,,sc.σi2(k/2-1)} and SiΩ1(σi)=. Then SiSi.σi= and μi(Si)1/3|supp(σi)|. This means that S is fixed by σ if and only if supp(σ):=(supp(σi))¯iI=()¯ has measure zero. But this means σ=id in the metric ultraproduct S𝒰.

Now we consider the case G¯=PGL𝒰(q). Here, similarly, assume for fixed g=(gi)¯iIGL𝒰(q) that V.g=V for all V𝒱. Then take V=(Vi)¯iI in the following way. The linear transformation gi is a direct sum of Frobenius blocks F(χ), where χk[X] runs through all (monic) primary polynomials. For each such block bkini of dimension kb>1 (here seen as a subspace) of gi select a cyclic vector vb. Then define Vi by

Vi=b,kb>1vb,vb.gi2,,vb.gi2(kb/2-1).

Then one observes that ViVi.gi=0 so that

dimi(Vi+Vi.gi)-dimi(ViVi.gi)=2dimi(Vi).

This shows that qχ(g)=0 for all χk[X] primary of degree kχ>1. But one observes that, if q(X-λ)(g),q(X-μ)(g)ε>0 for λμ elements of k, we can use the following construction. Let

ei1,,eikiVX-λ(gi)andfi1,,fikiVX-μ(gi)

such that lim𝒰ki/ni=ε. Define Vi:=eij+fijj=1,,ki (iI). Assume vViVi.gi. Then there exist numbers α1,,αki,β1,,βkik such that

v=j=1kiαj(eij+fij)=j=1kiβj(λeij+μfij).

This gives that

j=1ki(αj-βjλ)eij=j=1ki(βjμ-αj)fij

so that by disjointness of VX-λ(gi) and VX-μ(gi) both sides are zero and so, since the eij,fij (j=1,,ki) are linearly independent, we get that

αj-βjλ=βjμ-αj=0

so that, since λμ, we obtain αj=βj=0. Hence v=0 and ViVi.gi=0. But lim𝒰dim(Vi)/niε, yielding the same contradiction as above. Therefore, we must have g=λid (for λ𝔽q×, as k=𝔽q is finite) in the metric ultraproduct GL𝒰(q), i.e., PGL𝒰(q) is faithfully represented.

Remark 4.

The above statement about PGL𝒰(q) holds for any such metric ultraproduct of groups PGLni(ki) where the fields ki are not restricted with the same proof. Here the kernel of the action GL𝒰Aut(𝒱,dim) is given by 𝒰ki× (the algebraic ultraproduct of these groups).

Remark 5.

Hence, if the sequence of subsets (Si)iI resp. subspaces (Vi)iI is almost stabilized by each element of a subgroup H of G¯=S𝒰 resp. G¯=PGL𝒰(q) (or of G=GL𝒰(q)), we can restrict H to S:=(Si)¯iI resp. V:=(Vi)¯iI.

Remark 6.

For an element σ=(σi)¯iI, set

Ωk(σ):=(Ωk(σi))¯iI𝒮fork+.

Similarly, for a semisimple element g=(gi)¯iIGL𝒰(q) (see the next section for the definition of semisimple elements in such an ultraproduct), set

Vχ(g):=(Vχ(gi))¯iI𝒱forχk[X]irreducible.

Note that these definitions are independent of the chosen representatives (for the uniqueness of Vχ(g), we need that g is semisimple since then Vχ(gi)=ker(χ(gi)) is unique for a suitable representative (gi)iI of g and χk[X] irreducible).

Remark 7.

Call a sequence of subsets (Bi)iIkni a basis of V𝒱 if there is a representative (Vi)iI of V such that Bi is a basis of Vi (iI).

Remark 8.

Call V𝒱 totally singular if it has a representative (Vi)iI such that each Vi is totally singular (i.e., fi|Vi=0 for iI).

6 Centralizers in S𝒰, GL𝒰(q), Sp𝒰(q), GO𝒰(q), and GU𝒰(q)

In this section, we provide tools (Lemmas 2 and 3) to compute centralizers of certain elements from the metric ultraproducts S𝒰 and GL𝒰(q). We will use this in Section 7 to compute centralizers of elements in PGL𝒰(q). We write 𝐂(g) for the centralizer of the group element g.

Centralizers of elements in G=S𝒰, GL𝒰(q). Note that, for σ=(σi)¯iIS𝒰 resp. g=(gi)¯iIGL𝒰(q), we have 𝒰𝐂(σi)𝐂(σ) resp. 𝒰𝐂(gi)𝐂(g) (subsequently, by this notation we mean the metric ultraproduct of subgroups of the Hi (iI)). In the following lemma, we characterize when the above inclusion is actually an equality in the case of S𝒰.

Lemma 2.

An element σSU satisfies kZ+qk(σ)=1 if and only if, for each choice of a representative (σi)iI of σ, the centralizer C(σ) equals UC(σi).

Before proving Lemma 2, we turn to GL𝒰(q). An element gGL𝒰(q) is called semisimple if it has a representative (gi)iI such that each giGLni(q) is semisimple, i.e., of order prime to q.

Lemma 3.

A semisimple element gGLU(q) satisfies χ𝑝𝑟𝑖𝑚𝑎𝑟𝑦qχ(g)=1 if and only if for each choice of a representative (gi)iI of g where each gi is semisimple (iI), the centralizer C(g) equals UC(gi).

To prove Lemmas 2 and 3, we need the following auxiliary result.

Lemma 4.

The following are true.

  1. Assume σSym(n¯) is of order k and Sn¯ has normalized counting measure μ(S). Then S contains a σ -invariant subset T of measure

    μ(T)1-k(1-μ(S)).
  2. Assume gGL(V) for a k-vector space V and that the minimal polynomial of g over k has degree k. Assume UV. Then there exists a g-invariant subspace of U of codimension at most kcodim(U).

Proof.

(i) Observe that the biggest σ-invariant subset of S is equal to

T=iS.σi.

But since σk=id by assumption, we see that actually T=i=0k-1S.σi. Hence, since μ(S.σi)=μ(S) for all i, we have that μ(T)1-k(1-μ(S)).

(ii) Similarly to the above, the biggest g-invariant subspace contained in U is W=iU.gi. Now vi=0k-1U.gi means that v,,v.g-(k-1)U. But then

v.g-k=-1a0(a1v.g-(k-1)++ak-1v.g-1+v)U,

where χ=a0+a1X++ak-1Xk-1+Xk is the minimal polynomial of g. Note that a0=(-1)kdet(g)0. This shows that actually W=i=0k-1U.gi so that codim(W)kcodim(U). ∎

Remark 9.

The bounds in Lemma 4 are sharp. For example, take σ of type (kck), and set n=ckk. Take S of size n-s such that, for precisely sckk-cycles of σ, S contains k-1 elements of each of them and all elements of the remaining k-cycles. Then the set T constructed in Lemma 4 has size n-ks. In (ii), we can use a similar construction.

Now we are able to prove Lemmas 2 and 3.

Proof of Lemmas 2 and 3.

At first, we prove Lemma 2. Assume that σ=(σi)¯iI, τ=(τi)¯iIS𝒰 commute, and assume that k=1qk(σ)=1. Find a sequence (Ni)iI tending to infinity along 𝒰 such that

lim𝒰k=1Niqk(σi)=1and(Ni+12)H([σi,τi])𝒰0.

Recall that Ck(σi) denotes the set of k-cycles of the permutation σi. Call a k-cycle of σibad if it is not mapped σi-equivariantly to another k-cycle of σi by τi. Collect the set of bad k-cycles of σi in Ck(σi). For each bad k-cycle of σi, we get at least one non-fixed point of [σi,τi] so that |Ck(σi)|/niH([σi,τi]) for all k+. Hence, if we change τi such that all bad k-cycles of σi are mapped accurately for kNi and all k-cycles for k>Ni are mapped identically, we get a permutation τi such that

dH(τi,τi)1nik=1Nik|Ck(σi)|+k=Ni+1qk(σi)(k=1Nik)H([σi,τi])+k=Ni+1qk(σi)=(Ni+12)H([σi,τi])+k=Ni+1qk(σi).

By the assumption k=1qk(σ)=1, the last term in the above estimate tends to zero along 𝒰. Hence τ=(τi)¯iI=(τi)¯iI and [σi,τi]=1.

Conversely, assume that k=1qk(σ)<1. Choose the sequence (Ni)iI such that lim𝒰k=1Niqk(σi)=k=1qk(σ) and lim𝒰Ni/ni=0.

For each iI, change σi to σi such that the k-cycles of σi are the same as in σi for 1kNi and the other k-cycles of σi (k>Ni, if they exist) are grouped into one big Ki-cycle so that dH(σi,σi) is minimal possible. It is easy to see that then still dH(σi,σi)𝒰0 as Ni𝒰. Now σi eventually has precisely one Ki-cycle for Ki>Ni. Obtain σi′′ by dividing this Ki-cycle (if it exists) into two Ki/2-cycles and at most one fixed point so that dH(σi,σi′′)3/ni is minimal. Note that Ki/ni=1-k=1Niqk(σi)𝒰ε>0 so that Ki/2>Ni along 𝒰, as lim𝒰Ni/ni𝒰0 by assumption.

Now consider the restriction of the centralizers 𝐂(σi) and 𝐂(σi′′) to the support of the unique Ki-cycle of σi (which certainly both fix setwise by the previous inequality). The first group is isomorphic to the cyclic group CKi, whereas the second is isomorphic to CKi/2C2. Taking the metric ultraproducts of these groups restricted to this support (in the sense of Remark 5), we get an abelian group in the first case, and a non-abelian group in the second case. Hence, in at least one case, 𝒰𝐂(σi)𝐂(σ) or 𝒰𝐂(σi′′)𝐂(σ).

Now we prove Lemma 3. Assume that g=(gi)¯iI,h=(hi)¯iIGL𝒰(q) commute, i.e., [g,h]=id, that g and each gi (iI) is semisimple, and assume that

χirreducibleqχ(g)=1.

Recall that semisimplicity means that, for each Frobenius block F(χ) in the generalized Jordan normal form of gi, χ=i1 is irreducible. Choose the sequence (Ni)iI such that

lim𝒰χirreducibledeg(χ)Niqχ(gi)=1and(χirreducibledeg(χ)Nideg(χ))rk([gi,hi])𝒰0.

Define Ui:=ker([gi,hi]-id). Fix an irreducible polynomial χk[X] and apply Lemma 4 (ii) inside V:=Vχ(gi) to the subspace U:=UiVχ(gi) to get a gi-invariant subspace W=WiχU such that codimV(Wiχ)kχcodim(Ui), where kχ=deg(χ). Note here that Vχ(gi)=ker(χ(gi)) is unique since gi is semisimple. This large-dimensional subspace Wiχ is mapped accurately by hi, as gi commutes with hi on it. Define hi to be equal to hi on each Wiχ, and complete it on each Viχ to a map commuting with gi for deg(χ)Ni (here we use semisimplicity of gi). On Vχ(gi) with deg(χ)>Ni, set hi to be the identity. As in the proof for S𝒰 above, it follows that drk(hi,hi)𝒰0 and [gi,hi]=1.

Conversely, assume that χirreducibleqχ(g)<1. Choose the sequence (Ni)iI such that

lim𝒰χirreducibledeg(χ)NiNiqχ(gi)=χirreducibleqχ(g)andlim𝒰Ni/ni=0.

For each iI, change gi into gi such that all Frobenius blocks F(χ) for χ irreducible of degree at most Ni are left unchanged and all bigger Frobenius blocks (if there is any such block) are grouped into one big Frobenius block F(φ) of size Ki (for φ irreducible). Define gi′′ in the same way, but split the Frobenius block F(φ) (if it exists) into two or three blocks, two of which are F(ϕ) for ϕ irreducible of degree Ki/2 and, if Ki is odd, one block of size one, which is the identity. Then, as above, the centralizer of gi restricted to the large Frobenius block F(φ) of it, equals 𝐂(gi)(k[X]/(φ))×, whereas the centralizer 𝐂(gi′′) restricted to the same subspace is non-abelian (again in the sense of Remark 5). Also one sees that their metric ultraproducts are non-isomorphic, similarly to the case of permutations. The proof is complete. ∎

Remark 10.

If G is one of Sp𝒰(q), GO𝒰(q), or GU𝒰(q) and a semisimple gG is represented by (gi)iI and χirreducibleqχ(g)=1, one can adapt the above argument for GL𝒰(q) to see that still 𝐂(g)=𝒰𝐂(gi) when all gi are semisimple.

Indeed, from Fact 3, it follows that, in the space Wiχ+Wiχ (where Wiχ,Wiχ are constructed as above), we can still find a big, i.e., almost full-dimensional, gi-invariant non-singular subspace Wiχ,χ. This is because each block F(χ) must be paired with a block F(χ) (which can possibly be the same as the first one) to get an indecomposable. Now Wiχ contains almost all blocks F(χ) of gi, and Wiχ contains almost all blocks F(χ) of gi, so many of them must be paired together.

By Fact 2, we can then complete the partial map

hi|Wiχ,χ:Wiχ,χWiχ,χ.hi

to a map hiGL(Vχ(gi)+Vχ(gi)) stabilizing fi (and Qi in the orthogonal case for p=2). So we have found an element hiHi close to hi which commutes with gi. This completes the proof.

As a consequence of Lemma 1 together with Lemmas 2 and 3, and Remark 10, we get the following corollary.

Corollary 1.

If an element σSU resp. a semisimple element

gGL𝒰(q),Sp𝒰(q),GO𝒰(q),𝑜𝑟GU𝒰(q)

is torsion, then C(σ) resp. C(g) is equal to UC(σi) resp. UC(gi) for each representative (σi)iI resp. (gi)iI of σ resp. g, where we require all gi (iI) to be semisimple.

7 Centralizers in PGL𝒰(q), PSp𝒰(q), PGO𝒰(q), and PGU𝒰(q)

Now we can deduce the structure of centralizers of semisimple elements from PGL𝒰(q). By this we mean elements that lift to semisimple elements in GL𝒰(q). Let g=(gi)¯iIGL𝒰(q) be a semisimple element which maps to

g¯PGL𝒰(q)=GL𝒰(q)/k×.

All gi are assumed to be semisimple (iI).

Assume that h=(hi)¯iIGL𝒰(q) is such that [g,h]=μid for μk×. Then gh=μg so that q(g)=q(gh)=q(μg)=q(g).μ, i.e., μstabk×(q(g)). Let νstabk×(q(g))k× be a generator of this cyclic group.

It is now easy to see that the conformal centralizer

𝐂conf(g):={hGL𝒰(q)there isμk×such thatgh=μg}

is an extension 𝐂(g).stabk×(q(g))=𝐂(g).ν of 𝐂(g) by stabk×(q(g)). Hence 𝐂(g¯)=(𝐂(g).ν)/k×.

Remark 11.

The analog statement of Lemma 3 is false in PGL𝒰(q). Indeed, take a semisimple element g¯PGL𝒰(q) such that, for a lift gGL𝒰(q), the group stabk×(q(g)) is non-trivial. Choose a representative (gi)iI of gGL𝒰(q) such that qχ(gi)qξ(gi) for all χ,ξk[X] distinct irreducible and gi is semisimple (iI). Then 𝐂(gi) stabilizes each subspace Vχ(gi)=ker(χ(gi))kni. But this means that, if hC:=𝒰𝐂conf(gi), we have that gh=g so that C/k× is properly contained in 𝐂(g¯) (namely, 𝐂(g¯)/(C/k×)stabk×(q(g)), which is non-trivial).

Remark 12.

For the groups PSp𝒰(q), PGO𝒰(q), and PGU𝒰(q), the same structure for 𝐂(g¯) holds true, where Sp𝒰(q), GO𝒰(q) resp. GU𝒰(q) play the role of GL𝒰(q). The possible scalars μk× (from Z) are restricted to μ{±1} in the bilinear case, and to μq+1=1 in the unitary case.

8 Double centralizers of torsion elements

In this section, we compute the double centralizers of semisimple torsion elements of the groups G¯ of type S𝒰, PGL𝒰(q), PSp𝒰(q), PGO𝒰(q), and PGU𝒰(q). Note that, for gG a group element, 𝐂(𝐂(g))=𝐙(𝐂(g)) since g𝐂(g) so that 𝐂(𝐂(g))𝐂(g). Set 𝐂2(g):=𝐂(𝐂(g)) and 𝐂conf2(g):=𝐂conf(𝐂conf(g)) to be the double centralizer resp. double conformal centralizer of g. Here

𝐂conf(g):={hG[g,h]𝐙(G)}.

8.1 The case S𝒰

Let σ=(σi)¯iIS𝒰=G¯ be torsion of order o. Then we have koqk(σ)=1 by Lemma 1. By Corollary 1, we have that 𝐂(σ)=𝒰𝐂(σi). But 𝐂(σi) has a subgroup

koCkSym(ck(σi))

which is dense in it along 𝒰 so that C:=𝐂(σ)=𝒰koCkSym(ck(σi)).

At first, for simplicity, assume that σi is isotypic of type (kcik) (so ni=cikk). Assume that τ=(τi)¯iI𝐙(C) and τi=(aij).φiCkSym(cik). Assume that lim𝒰|supp(φi)|/cik=ε>0. Then we can conjugate φi by

ϕiSym(cik)CkSym(cik)=𝐂(σi)

such that

lim𝒰dH(φiϕi,ϕiφi)ε>0.

But this leads to the contradiction

lim𝒰dH(τiϕi,ϕiτi)ε>0.

Hence we may assume that φi=id, applying a small change to τi along 𝒰 if necessary (iI). Now assume that lim𝒰|{jaij=c}|/cik=ε(0,1). Then we find permutations ϕiSym(cik)CkSym(cik)=𝐂(σi) such that

dH(τi,τiϕi)=|{jaijaij.ϕi}|/cikmin{ε,1-ε}>0.

Hence we can assume that all aij are equal. This shows that, in this case, 𝐙(𝐂(σ)) is the metric ultraproduct 𝒰CkCk, where Ck in the ith component is generated by the element σi itself (iI).

In the general case, we obtain that

𝐙(𝐂(σ))=ko,qk(σ)>0𝒰Ckko,qk(σ)>0Ck.

This holds because σ𝐂(σ) so that, when τ𝐙(𝐂(σ)), it must commute with σ. But this implies that lim𝒰|Ωk(σi)Ωk(σi).τi|=0 so that τ must stabilize the isotypic components of σ (in the sense of Remark 5), and we can apply the above argument.

8.2 The case PGL𝒰(q), PSp𝒰(q), PGO𝒰(q), and PGU𝒰(q)

Recall from Section 2 that k=𝔽q when G is GL𝒰(q), Sp𝒰(q), or GO𝒰(q), and k=𝔽q2 when G=GU𝒰(q). Set d=1 in the first three cases and d=2 when G is unitary over 𝔽q2.

Recall that Z=k× when G=GL𝒰(q), Z={±1}k× when G=Sp𝒰(q) or G=GO𝒰(q), and Z={zk×zq+1=1}k×=𝔽q2× when G=GU𝒰(q). Also recall that, if G is not of shape GL𝒰(q), we have zσ=z-1 for zZ, where σ:kk is the identity in the bilinear case, and the q-Frobenius endomorphism xxq when G=GU𝒰(q). Let g=(gi)¯iIGGL𝒰(k) be semisimple, with gi (iI) semisimple such that g¯G¯PGL𝒰(k)=GL𝒰(k)/k× is torsion of order dividing o, i.e., there is μk× such that go=μid. This implies μZ. Then

χirreducibleXoμ(χ)qχ(g)=1

by Lemma 1. Set

P:={χk[X]χ(monic) irreducible,χXo-μ},

T:=stabZ(q(g)), Kχ:=k[X]/(χ) for χk[X] irreducible (as in Fact 3), and ciχ:=cχ(gi) (iI). Hence, similarly to the above, we have

𝐂(g)=𝒰χirreducibleXoμ(χ)qχ(g)>0𝐌ciχ(Kχ),

the centralizer being computed in 𝐌𝒰(k). Now, by Section 7, we “know” the structure of 𝐂conf(g)G. For χk[X] irreducible, consider the g-invariant subspace V:=Vχ¯(g):=ξχ¯Vξ(g)𝒱, where χ¯:=orbT(χ) is the orbit of χ under T (see Remark 6 for the definition of Vξ(g)𝒱). Set lχ:=|χ¯|, mχ:=|T|/lχ. Note that mχ=|stabT(χ)|, and so

mχ=max{mdivides|T|there existsχsuch thatχ=χ(Xm)}.

The restriction of the action of 𝐂conf(g)/Z to Vχ¯(g) is given by

((𝒰ξχ¯𝐂(g|Vξ(g))).T)/Z.

We will explain this below. In most cases, the above extension is even split.

Definition of the action of T. Let tZ be a generator of the above cyclic group T. Then the map φt, which is constructed in the following, generates the above extension.

At first, note that we can see each space Vξ(gi) as a vector space over the field Kξ=k[X]/(ξ) (ξk[X] irreducible) as g acts on it as multiplication by λξ=X¯ acts on Kξ (cf. Fact 3). With this convention, find Kξ-bases (Bξ,i)iI of each Vξ(g) (ξχ¯ for all representatives χ of orbits of the action of T on the irreducible polynomials; see Remark 7) and compatible bijections αξ1,ξ2,i:Bξ1,iBξ2,i (for iI, i.e., αξ2,ξ3,iαξ1,ξ2,i=αξ1,ξ3,i for all ξ1,ξ2,ξ3χ¯, all χ, and all iI). If G comes from groups preserving a form, we still find bijections :Bξ,iBξ,i such that b=b, the pairing fi restricted to

Kξb×Kξb=k[g]b×k[g]bk

is non-singular, the pairing fi restricted to Kξb×Kξb is zero for all bBξ,i, bBξ,i, bb, and such that commutes with the maps αξ1,ξ2,i (iI). Such bases exist by an application of Fact 3. The last condition can be fulfilled since (ξ.t)=ξ.t for all ξk[X] by Remark 3.

Now we define maps φt,i on kni by defining their restrictions

φt,i|Bξ,iKξ:Bξ,iKξBξ.t,iKξ.t.

We do this by defining a k-linear map φξ,t:KξKξ.t (which is not necessarily a homomorphism) and extending it to a map φt,i by linearity,

φt,i(bBξ,iλbb)=bBξ,iφξ,t(λb)αξ,ξ.t,i(b).

Let ϕξ,t:KξKξ.t denote the k-homomorphism such that λξtλξ.t. Here is our definition of φξ,t.

If not p>2, ε=-1, χ=χ, t=-1, and χ.t=χ, set φξ,t:=ϕξ,t. In this case, φt|T|=id, and the above extension is split.

In the opposite case,

ξ=ξ=ξ.t=ξ.(-1)for allξχ¯.

Let ηKξ be a generator of the cyclic group Kξ× and set ζ:=η(qkχ/2-1)/2, where kχ=deg(χ)=deg(ξ), so that ζqkχ/2+1=-1. Then set φξ,t:KξKξ.t to be the map xζϕξ,t(x). In this case, φξ,t2(x)=-x, so the above extension needs not to be split.

Doing this for all representatives χ of orbits of the action of T on the irreducible polynomials χk[X] with χXo-μ, this defines, up to a small error in the rank metric, a map φt,i:knikni, so set φt to be (φt,i)¯iI.

The map gt-commutes with φt. Recall from Fact 3 that gi (iI) acts on Kξ as multiplication by λξ (and so on Kξ.t as multiplication by λξ.t). Hence, by the above definition of φt, we need to verify that

(8.1)φξ,t(λξx)=tλξ.tφξ,t(x)

for all xKξ, ξχ¯.

In the first of the above cases, this is true since then φξ,t=ϕξ,t is a field isomorphism (and hence multiplicative) and ϕξ,t(λξ)=tλξ.t by definition. In the second case, the left-hand side of equation (8.1) evaluates to

φξ,t(λξx)=ζϕξ,t(λξx)=ζϕξ,t(λξ)ϕξ,t(x)=ζtλξ.tϕξ,t(x),

whereas the right-hand side is tλξ.tφξ,t(x)=tλξ.tζϕξ,t(x), so they are equal.

The action of φt preserves the forms fi (and Qi in the orthogonal case when p=2, iI). Assume G is not GL𝒰(q). Then one sees that φt preserves the forms fi (iI). We verify this in the different cases of Fact 3. Let bBξ,i (for ξχ¯), and consider the restriction of fi to U×U:=Kξb×Kξb. Recall from Fact 3 that, in this situation, τ:KξKξ is the homomorphism such that τ:λξλξ-1 and τ|h=σ. Let τ:Kξ.tKξ.t be the corresponding homomorphism such that λξ.tλξ.t-1 and τ|k=σ.

In case (i) (a) of Fact 3, we have

fi(u.t,v.t)=trKξ.t/k(φt,i(u)φt,i(v)τ)=trKξ.t/k(ϕξ,t(u)ϕξ,t(v)τ)=trKξ.t/k(ϕξ,t(uvτ))=trKξ/k(uvτ)=fi(u,v).

Here we use in the third line that τϕξ,t=ϕξ,tτ since both induce σ on the ground field k and

ϕξ,t(λξ)τ=(tλξ.t)τ=tσλξ.t-1=(tλξ.t)-1=ϕξ,t(λξ)-1=ϕξ,t(λξ-1)=ϕξ,t(λξτ).

In the orthogonal case for p=2, the forms Qi are preserved by a similar argument.

In case (i) (b) of Fact 3, let δKξ be the element used there, and let δKξ.t be the analogous element in the ξ.t-block. We may assume δ=δ=1 when ε=1. Then the same computation as in the previous case verifies that φt preserves the forms fi (and Qi). Now assume that ε=-1. If t=1, there is nothing to show, so t=-1. When χχ.t, then ξξ.t for ξχ¯, so we can assume that δ=ϕξ,t(δ). Then

fi(u.t,v.t)=trKξ.t/k(δφt,i(u)φt,i(v)τ)
=trKξ.t/k(δϕξ,t(u)ϕξ,t(v)τ)
=trKξ.t/k(δϕξ,t(uvτ))
=trKξ.t/k(ϕξ,t(δ)ϕξ,t(uvτ))
=trKξ/k(δuvτ)=fi(u,v).

So assume χ=χ.t so that ξ=ξ.t (for ξχ¯), τ=τ, and δ=δ. Then

fi(u.t,v.t)=trKξ.t/k(δφt,i(u)φt,i(v)τ)=trKξ/k(δζϕξ,t(u)ζτϕξ,t(v)τ)=trKξ/k(ζqkχ/2+1δϕξ,t(uvτ))=trKξ/k(ϕξ,t(δuvτ))=trKξ/k(δuvτ)=fi(u,v).

Here we use in the third line that ζqkχ/2+1=-1 and ϕξ,t(δ)=-δ since ϕξ,t=τ is the unique involution xxqkχ/2 in this case.

Assume now we are in case (ii). Then we may again assume that t=-1 and p>2 since |T||Z|2, and also χ=X±1χ.t=X1. Then we have φξ,t=ϕξ,t=idk so that fi is preserved according to Fact 3. So we have shown that φt preserves the forms fi (and Qi).

Now let us fix h𝐂conf2(g). We want to understand the shape of h.

Step 1: h stabilizes each Vχ(g) (χk[X] irreducible). Assume that

h𝐂conf2(g|V)𝐂conf(g|V)

does not stabilize each subspace Vξ(g) of V (ξχ¯). Write χ¯={ξ1,,ξl} and assume that Vξ1(g).h=Vξ2(g). Take

f=(M1,M2,,,)𝐂(g|V)𝐂conf(g|V),

where the jth component of f acts on Vξj(g) (j=1,,l); then

fh=h-1fh=(,M1h,,,).

Now there are three cases according to the classification presented in Fact 3. If G=GL𝒰(q), we can take M2=1Vξ2(g) and M1 far away from k×idVξ1(g). Then [f,h]=(,M1h,,,) is far away from k×idV. If G is one of Sp𝒰(q), GO𝒰(q), or GU𝒰(q), ξ1 is not self-dual and ξ2ξ1, we can do the same as before. When ξ1=ξ2 in this case, we must have M2=(M1-σ) so that

[f,h]=(,M1σM1h,,,).

Again we can choose M1𝒰GLciξ1(Kξ1) such that (M1σ)M1h is far away from Z. In the last case, ξ1=ξ1 is self-dual. Then again M1 and M2 are independent of each other, and we can choose M2=1Vξ2(g). The only restriction on M1 is that it lies in 𝒰GUciξ1(Kξ1) if ξ1X±1 or G is GU𝒰(q) (see Fact 3 (i) (b)) resp. M1𝒰Xciξ1(k) in the opposite case when ξ1=X±1, where G=X𝒰(q) (X=Sp or GO; see Fact 3 (ii)), so again we can choose M1 such that [f,h]=(,M1h,,,) is far away from Z. In all cases, we get a contradiction. This shows that h𝐂conf2(g) fixes each Vχ(g)𝒱 (χk[X] irreducible).

Assume now that h|Vχ(g)=M.α, where α corresponds to an element of

Tlχ={slχsT}

which induces a non-trivial field automorphism on Kχ and M is an ultraproduct of matrices over Kχ. This is the case since h𝐂conf(g) and by the above definition of φt.

Step 2: the automorphism α equals the identity idKχ. Then, for λKχ, we have (λid)h=λαid=(λαλ-1)λid. This implies that, for all λKχ× stabilizing the forms fi (or Qi; iI) on Vχ(g), we have λαλ-1Zk×. When G=GL𝒰(q) or χ is not self-dual, there is no restriction on λ (of course, if G is one of Sp𝒰(q), GO𝒰(q), or GU𝒰(q), then if h acts as M on Vχ(g), it must act as (M-σ) on Vχ(g)). Hence, in this case, for each λ×Kχ, there exists κλk× such that λαλ-1=κλ. However, then every vector λKχ is an eigenvector of the k-linear map α, which forces α=idKχ, since 1Kχ is fixed, a contradiction.

In the opposite case, G is one of Sp𝒰(q), GO𝒰(q), or GU𝒰(q) and χ is self-dual. Then we are in case (i) (b) and (ii) of Fact 3. Recall that τ:KχKχ is defined by τ|k=σ and τ:λχλχ-1, where λχKχ is the root of χ. Then τ2=idKχ and τ=idKχ if and only if we are in case (ii) of Fact 3. Here, if we are in case (i) (b) of Fact 3, 𝐂(g)|Vχ(g) is an ultraproduct of unitary groups over the field Kχ equipped with the involution τ. In case (ii) of that fact, 𝐂(g)|Vχ(g) is an ultraproduct of symplectic resp. orthogonal groups over Kχ=k=𝔽q. We proceed as follows. Find totally singular Kχ-subspaces

U=(Ui)¯iI,U=(Ui)¯iI,U′′=(Ui′′)¯iI𝒱ofVχ(g)

(in the sense of Remark 8) such that UU=UU′′=Vχ(g), UU′′=0, and dim(U)=dim(U)=dim(U′′)=dim(Vχ(g))/2. We may assume that

dimKχ(Ui)=dimKχ(Ui)=dimKχ(Ui′′)

and that the restrictions fi|Ui×Ui and fi|Ui×Ui′′ are non-degenerate (iI, as we may by modifying Ui, Ui, and Ui′′ a little if necessary). Then define f=(fi)¯iI, f′′=(fi′′)¯iI𝐂(g)G such that fi and fi′′ act F(φ)-isotypically on Ui and such that fi resp. fi′′ act F(φ)-isotypically on Ui resp. Ui′′ (iI) for a fixed irreducible polynomial φKχ[X] which is not self-dual with respect to τ. Then fh|Vχ(g)=zf|Vχ(g) and f′′h|Vχ(g)=z′′f′′|Vχ(g) for z,z′′Z. Note that

qφ.z-1(zf|Vχ(g))=qφ(f|Vχ(g))=1/2,
qφ.z′′-1(z′′f′′|Vχ(g))=qφ(f′′|Vχ(g))=1/2,

and φ.z-1 and φ.z′′-1 are both also not self-dual since φKχ[X] is not self-dual and z-1,z′′-1Z so that z-τ=z-σ=z and z′′-τ=z′′-σ=z′′, whence, e.g., (φ.z-1)=φ.z-1φ.z-1. Then h must stabilize the decompositions Vχ(g)=UU=UU′′, so it must stabilize U. But on the h-invariant totally isotropic subspace U, we can do the same argument as above for G=GL𝒰(q) to see that α=idKχ. Hence we have obtained that h|Vχ(g)=M𝒰𝐌ciχ(Kχ) so that h𝐂(g).

Step 3: we have that h|Vχ(g)=M=λid for λKχ. According to Fact 3, we can find Vi (iI) such that (Vi)¯iI=Vχ(g) such that either all Vi are totally singular (case (i) (a) of that fact, i.e., if χ is not self-dual; this includes the case G=GL𝒰(q)) or Hi preserves a unitary form (case (i) (b) of Fact 3) or a symplectic or orthogonal form (and a corresponding quadratic form when p=2, case (ii) of Fact 3) over Kχ on Vi (iI). Note that it follows from the classification in [4, Subsection 3.4.2, § 1] that orthogonally indecomposable blocks involving a Frobenius block of size 2 are non-central in the ambient projective linear classical group. This shows that qξ(M)=0 for all ξKχ[X] of degree 2. Assume now that there exist distinct λ,μKχ× such that qX-λ(M),qX-μ(M)ε>0. If G=GL𝒰(q) or we are in case (i) (a) of Fact 3,

F(X-λ)F(X-μ)=diag(λ,μ)GL2(Kχ)

is mapped to a non-central element in PGL2(Kχ) so that, by the assumption, since we have “many” of these blocks, h|Vχ(g) would not commute modulo scalars with all of 𝐂(g)|Vχ(g)𝒰𝐌ciχ(Kχ). In case (i) (b) of Fact 3, we use the same argument for a block of shape diag(λ,λ-τ,μ,μ-τ) acting on a four-dimensional (Kχ,τ)-unitary space. In case (ii) of Fact 3, we use the same argument with a block diag(λ,λ-1,μ,μ-1) acting on a four-dimensional symplectic or orthogonal space. In total, we get that M=λid for λKχ. If we are in case (i) (b) of Fact 3, we have the additional assumption that Nτ(λ)=1, where Nτ:Kχ(Kχ)τ is the field norm xxxτ. In case (ii) of Fact 3, we must have λ2=1.

Step 4: the precise shape of C:=𝐂conf2(g). We know now h|Vχ(g)=λχ(h)id for each irreducible χk[X], and so h commutes with all of 𝐂(g). In order that h𝐂conf2(g), we still need to check that [h,φt]=zid, where tT is the generator fixed above and zZ. Let χk[X] run through a system of representatives of the orbits of the action of T and on the irreducible polynomials (the action of is only used when G is not GL𝒰(q)). This means zh=hφt, so since h|Vχ(g)=λχ(h)id, we must have h|Vχ.t(g)=z-1φχ,t(λχ(h))idVχ.t so that h is determined on all of V=Vχ¯(g) by λχ(h). In this situation, the only condition that needs to be satisfied is that

h|Vχ(g)=h|Vχ.tlχ(g)=λχ(h)id=z-lχφlχχ,t(λχ(h))id.

Note that

φχ,tlχ:Kχ𝔽qdkχKχ𝔽qdkχ(d=1,2)

is given by xκχxqdkχ/mχ, where κχ=1 unless p>2, ε=-1, t=-1, and χ.t=χ, in which case κχ=ζ is the scalar chosen above (which depends on χ), so that the previous condition becomes

(8.2)zlχ=κχ(λχ(h))qdkχ/mχ-1.

Hence we can write C as follows. When G=GL𝒰(q), we have

(8.3)C={h=χirreducibleXoμ(χ)qχ(g)>0λχ(h)idVχ(g)|there existszZsuch thatλχ.t(h)=z-1φχ,t(λχ(h))for allχ}.

Here the condition from equation (8.3) is equivalent to equation (8.2) for χ running through a system of representatives of the action of T on the set

P:={χk[X]irreducibleχdividesXo-μandqχ(g)>0}.

For G one of Sp𝒰(q), GO𝒰(q), or GU𝒰(q), we have

(8.4)C={χirreducibleXoμ(χ)qχ(g)>0λχ(h)idVχ(g)|R},

where condition R is that there exists zZ such that λχ.t(h)=z-1φχ,t(λχ(h)) for all χP (as in the previous case) and λχ(h)λχ(h)τ=1 for all χP, where τ:KχKχ is defined as in Fact 3. If G is one of Sp𝒰(q) or GO𝒰(q) and χ=χX±1 is self-dual, this means kχ is even and λχ(h)qkχ/2+1=1. Also, in this case, if χ=X±1, it means λχ(h)2=1. If G=GU𝒰(q) and χ=χ, this means that kχ is odd (since τ needs to induce σ:xxq on k=𝔽q2) and λχ(h)qkχ+1=1.

9 Distinction of metric ultraproducts

Now we want to distinguish all (simple) metric ultraproducts G¯=X¯𝒰(q) for distinct pairs (X,q), where X{GL,Sp,GO,GU} and q is a prime power (all but PSp𝒰1(q) and PGO𝒰2(q) as mentioned in Theorem 1). For a group H, define the quantity

eH(o):=maxhH:ho=1Hexp(𝐂2(h)).

Clearly, when HL, we have

eH(o)=eL(o)for all valueso+.

Our strategy is to compute eH(o) for the groups H=G¯, where G¯=X¯𝒰(q) as above, for certain values of o to distinguish these groups (with the only exception PSp𝒰1(q)PGO𝒰2(q)?).

9.1 Computation of eG¯(o) when gcd{o,p}=gcd{o,|Z|}=1

If o is coprime to |Z| (and by semisimplicity of gG coprime to p), from the previous considerations, we can compute eG¯(o). Note that, in this situation, when go=μZ, we can replace g by g=λgG such that go=1, choosing λZ such that λo=μ-1, since the homomorphism ZZ, xxo is then bijective. So we may assume that go=1. Then

PQ:={χk[X]irreducibleχdividesXo-1}.

The case G=GL𝒰(q). From equation (8.3), we see that the bigger the group T is, for an element h𝐂conf2(g), the more restrictions are imposed to the scalars λχ(h)Kχ× (χP). Also the bigger the set P is, the “bigger” is the group 𝐂conf2(g), i.e., there are more components. Hence, to optimize the exponent of 𝐂2(g¯)=𝐂conf2(g)/Z, we choose g such that

P=Qand0<qX-1(g)qχ(g)>0for allχP{X-1}.

Namely, then T=stabZ(q(g)) must fix the polynomial X-1 so that we must have T=𝟏. Set

fq(o):=min{qe-1odividesqe-1}.

Equation (8.3) then gives

(9.1)eG¯(o)=exp(𝐂2(g¯))=exp(𝐂conf2(g)/Z)={1ifo=1,fq(o)ifo>1.

Let us demonstrate equation (9.1). The first equality in it holds by the previous argument. When o=1, we have g¯=1G¯, and so

𝐂2(g¯)=𝐙(G¯)=𝟏

so that eG¯(1)=1. Now assume o>1. For each χP, if λk¯× is a root of χ, the condition χXo-1 is equivalent to λo=1. Also Kχ=k[λ]. Let μk¯× be an element of order o with minimal polynomial ξk[X]. Then, if λ is a root of χP, we must have λo=1 and thus λ=μf for some f. Hence

Kχ=k[λ]=k[μf]k[μ]=Kξ

so that, in equation (8.3), ord(λχ(h)) divides |Kχ×| which divides |Kξ×|=fq(o). This shows that

exp(𝐂2(g¯))exp(𝐂conf2(g))lcm{|Kχ×|χdividesXo-1}=|Kξ×|=|k[μ]×|=fq(o).

To show that exp(𝐂2(g¯))=fq(o), take h𝐂conf2(g) such that λX-1(h)=1 and λξ(h) has order fq(o)=|Kξ×| in k¯×. Then, when h¯l=1G¯, we must have hlZ. But λX-1(h)l=1 so that, since qX-1(g)>0, it follows that hl=1G. Then λξ(h)l=1 so that exp(𝐂2(g¯))lord(λξ(h))=fq(o). This completes the proof.

The case G=Sp𝒰(q) or GO𝒰(q). As in the linear case, equation (8.4) shows that the optimal exponent of 𝐂2(g¯) is obtained when

P=Qand0<qX-1(g)qχ(g)>0for allχP{X-1}

so that T=𝟏. Set

(9.2)fq(o):={qe/2+1iffq(o)=qe-1,eis even, andoqe/2+1,fq(o)otherwise.

Equation (8.4) then gives

(9.3)eG¯(o)=exp(𝐂2(g¯))=exp(𝐂conf2(g)/Z)={1ifo=1,2ifo=2,fq(o)ifo>2.

We demonstrate equation (9.3). If o=1, we obtain, as in the linear case, that 𝐂2(g¯)=𝟏, and so eG¯(1)=1. If o=2, g2=1, and so P={X-1,X+1}. From equation (8.4), we see that, if h𝐂conf2(g), we have

λX-1(h)2=λX+1(h)2=1

so that h2=1. Also, defining h by λX-1(h):=1 and λX+1(h)=-1, we obtain hZ, so ordG¯(h¯)=eG¯(2)=2 (1-1 since the case p=2 does not occur due to the condition gcd{o,p}=1). Assume now that o>2. As in the linear case, for each χP, if λk¯ is a root of χ, the condition that χXo-1 is equivalent to λo-1. Choose μk¯× of order o, and let ξP be its minimal polynomial. Then, as previously, if λ is a root of χP, we have λ=μf for some f. There are two cases.

In the first case, ξ is not self-dual. This means that μ and μ-1 are not conjugate in Kξ/k. If they were conjugate, say by an automorphism α, i.e., μα=μ-1, then α=τGal(Kξ/k) needs to be the unique involution (since μμ-1 as o>2) given by xxqkξ/2; in particular, e=kξ would need to be even. Hence this case is equivalent to either e=kξ being odd or μμα=μqkξ/2+11, i.e., oqe/2+1=qkξ/2+1. This is precisely the opposite of the first case in equation (9.2). Here, for an element h𝐂conf2(g), we can choose λξ(h)Kξ×=k[μ]× arbitrarily (λξ(h) is then determined by λξ(h)). Arguing as in the linear case, we obtain exp(𝐂2(g¯))=fq(o). Indeed, for h𝐂conf2(g), as above, ord(λχ(h)) divides fq(o), and defining h such that λX-1(h)=1 and λξ(h) has order fq(o), we see that ordG¯(h¯)=eG¯(o)=fq(o).

In the opposite case, ξ is self-dual and ξX±1 as o>2. Then e=kξ needs to be even and

μμτ=μqkξ/2+1=λξ(h)λξ(h)τ=λξ(h)qkξ/2+1=1,

where τ is the involution xxqkξ/2 of Kξ𝔽qkξ from Fact 3 (i) (b). This means that oqe/2+1 and we are in the first case of equation (9.2). Note that, for each χP, the map τ restricts to an automorphism of each KχKξ of order dividing two (as all the fields are finite). Then τ|Kχ=id if and only if kχkξ/2, and τ|Kχ is the unique involution of Kχ if kξ/kχ is odd. Now, if λk¯× is a root of χ, then in the first case λ2=1, and in the second case λqkχ/2+1=1, so all χP are self-dual. Hence, if h𝐂conf2(g), for each χP, one of λχ(h)2=1 or λχ(h)qkχ/2+1=1 must hold. But 2qkξ/2+1 if p>2, and in the second case, qkχ/2+1qkξ/2+1 since kξ/kχ is then odd. Hence

exp(𝐂conf2(g))qe/2+1=qkξ/2+1=fq(o).

Defining h𝐂conf2(g) such that λX-1(h)=1 and λξ(h) has order fq(o), we see that ordG¯(h¯)=fq(o).

The case G=GU𝒰(q). Here, as well, equation (8.4) shows that the optimal exponent of 𝐂2(g¯) is obtained when P=Q and 0<qX-1(g)qχ(g)>0 for all χP{X-1} so that T=𝟏. Set

fq′′(o):={qe+1iffq2(o)=q2e-1,eis odd, andoqe+1,fq2(o)otherwise.

Equation (8.4) gives

(9.4)eG¯(o)=exp(𝐂2(g¯))=exp(𝐂conf2(g)/Z)={1ifo=1,fq′′(o)ifo>1.

Again eG¯(1)=1 is clear. If o>1, take μk¯× of order o with minimal polynomial ξ. Then the argument proceeds as in the bilinear case. But the condition that ξ is self-dual is here equivalent to μ being conjugate to μ-1 in Kξ by an automorphism τ such that τ|k=σ; xxq. This forces e=kξ to be odd and μqdkξ/2+1=1, i.e., oqe+1=qkξ+1.

9.2 Proof of Theorem 1

Set Gj:=Xj𝒰j(qj), Zj:=𝐙(Gj), and G¯j:=Gj/Zj (j=1,2). Let pj be the characteristic of the field 𝔽qj (j=1,2). Assume that G¯G¯1G¯2. We start by showing that p1=p2.

Determining the characteristic p. Choose o large enough and coprime to p1, p2, |Z1|, |Z2|. Then from equations (9.1), (9.3), and (9.4), we see that eG¯(o) is of the form q1e1±1 and q2e2±1. If we have eG¯(o)=q1e1-1=q2e2-1 or eG¯(o)=q1e1+1=q2e2+1, then q1e1=q1e2 so that p1=p2 by the uniqueness of the prime factorization. So we may assume that eG¯(o)=qe1-1=q2e2+1 for infinitely many o, and so for infinitely many pairs (e1,e2)+2. If p1p2, we get a contradiction to [1, Corollary 1.8]. Hence p1=p2=:p.

Determining qd. We can now assume qj=pej (j=1,2). Choose j{1,2}, and set X:=Xj, q:=qj, and d:=dj. Consider the quantity

f:=gcd{eG¯(o)oO},

where

O:={o+2<ocoprime top,|Z1|,|Z2|;eG¯(o)-1modulop3}

From equations (9.1), (9.3), and (9.4), it follows that, for every element oO, the number eG¯(o) is either of the form qde-1 or qe+1. But the second case is excluded by the condition that eG¯(o)-1 modulo p3. Hence qd-1 divides eG¯(o)=qde-1, and so qd-1 divides f.

For a prime r, set tr:=qr-1q-1. Then, for distinct primes r and s, we have

gcd{tr,ts}=gcd{qr-1q-1,qs-1q-1}=1q-1gcd{qr-1,qs-1}=qgcd{r,s}-1q-1=1.

Hence the numbers tr (r prime), being pairwise coprime, have arbitrarily large prime divisors. Take for r>2 a prime such that tr has a prime divisor o>p, |Z1|, |Z2|, qd-1. Then o is coprime to p, |Z1|, and |Z2| so that, by equations (9.1), (9.3), and (9.4), we have eG¯(o)fqd(o)qdr-1, as oqr-1qdr-1. Hence the number fqd(o) must be one of qdr-1 or qd-1, the latter being excluded by the condition o>qd-1, so fqd(o)=qdr-1. If X=GL, X=Sp, or X=GO, since r is odd and d=1, equations (9.1) and (9.3) show that we must have eG¯(o)=fqd(o)=qdr-1=qr-1-1 modulo p3. Hence, in this case, we have oO. If X=GU, it could be that eG¯(o)=qr+1 when oqr+1. However,

gcd{qr+1,tr}gcd{qr+1,qr-1}2,

and tr is always odd so that gcd{qr+1,tr}=1, and hence also gcd{qr+1,o}=1 as otr. This shows that here also eG¯(o)=fqd(o)=qdr-1=q2r-1-1 modulo p3. Therefore, again oO.

Applying this argument for two different primes r, say r1 and r2, which produces two different primes o, say o1 and o2, we get

f=gcd{eG¯(o)oO}gcd{eG¯(o1),eG¯(o2)}=gcd{qdr1-1,qdr2-1}=qgcd{dr1,dr2}-1=qd-1.

Altogether, we have shown that f=qd-1. Plugging in j=1,2, we obtain q1d1-1=q2d2-1 implying that q1d1=q2d2.

Now exclude all remaining possible isomorphisms but PSp𝒰1(q)PGO𝒰2(q).

Proof that PGL𝒰1(q)PSp𝒰2(q) and PGL𝒰1(q)PGO𝒰2(q). Let

G1=GL𝒰1(q)andG2=X𝒰2(q),whereX=SporGO.

Set

(9.5)o:={q2+12ifp>2,q2+1ifp=2.

Note that o>2 is coprime to p, |Z1|=q-1, and |Z2|=|{±1}|. Hence, by equation (9.3), we have eG¯(o)=eG¯2(o)=q2+1. Indeed, oq2+1q4-1. But oqf-1 for f properly dividing 4 since then oq2-1, but it is easy to see that gcd{o,q2-1}=1. This shows

fq(o)=q4-1andeG¯(o)=eG¯2(o)=fq(o)=q2+1.

But then, by G¯1G¯2, we obtain

eG¯(o)=eG¯1(o)=fq(o)=q4-1>q2+1=eG¯2(o)=eG¯(o),

a contradiction.

Proof that PSp𝒰1(q2)PGU𝒰2(q) and PGO𝒰1(q2)PGU𝒰2(q). Let

G1=X𝒰1(q2),whereX=SporGO,andG2=GU𝒰2(q).

Define o as in equation (9.5). Note that o>2 is coprime to p, |Z1|=|{±1}|, and |Z2|=q+1. Then, by equation (9.3), eG¯(o)=eG¯1(o)=fq2(o)=q2+1 (as above). But by equation (9.4), we obtain that +

eG¯=eG¯2(o)=fq′′(o)=q4-1>q2+1=eG¯1(o)=eG¯(o)

since e=2 is even, a contradiction.

Proof that PGL𝒰1(q2)PGU𝒰2(q). Let G1=GL𝒰1(q2), G2=GU𝒰2(q). Set

o:={q5+15(q+1)ifq-1modulo 5,q5+1q+1otherwise.

Note that o is coprime to p, |Z1|=q2-1, and |Z2|=q+1q2-1. Indeed,

gcd{o,q+1}gcd{q5+1q+1,q+1}=gcd{5,q+1}5.

But 5o so that gcd{o,q+1}=1. Similarly,

gcd{o,q-1}gcd{q5+1,q-1}gcd{2,q-1}2.

But o is always odd, so gcd{o,q-1}=1. We have oq10-1 so that, from equation (9.1), we obtain that eG¯(o)=eG¯1(o)=fq2(o) is either q10-1 or q2-1. But clearly q2-1<o so that we must have fq2(o)=q10-1. But equation (9.4) gives that

eG¯(o)=eG¯2(o)=fq′′(o)=q5+1<q10-1=fq(o)=eG¯1(o),

a contradiction.

Remark 13.

If qi𝒰, then double centralizers of semisimple torsion elements are infinite groups.

Remark 14.

If q is even, then PSp𝒰1(q)PGO𝒰2(q) is possible due to the isomorphism Sp2m(q)GO2m+1(q). Also it seems hard to distinguish a group PSp𝒰1(q) from a group PGO𝒰2(q) for q odd.


Communicated by John S. Wilson


Award Identifier / Grant number: 681207

Funding statement: This research was supported by ERC Consolidator Grant No. 681207.

Acknowledgements

The author wants to thank Andreas Thom for interesting discussions about the topic. The content of this article is part of the author’s PhD project [4].

References

[1] J.-H. Evertse, Linear forms in logarithms, preprint (2011), http://www.math.leidenuniv.nl/~evertse/dio2011-linforms.pdf. Search in Google Scholar

[2] S. Gonshaw, M. Liebeck and E. O’Brien, Unipotent class representatives for finite classical groups, J. Group Theory 20 (2017), no. 3, 505–525. 10.1515/jgth-2016-0047Search in Google Scholar

[3] M. Liebeck and A. Shalev, Diameters of finite simple groups: Sharp bounds and applications, Ann. of Math. (2) 154 (2001), 383–406. 10.2307/3062101Search in Google Scholar

[4] J. Schneider, On ultraproducts of compact quasisimple groups, PhD thesis, TU Dresden, 2019; to appear on http://www.qucosa.de. Search in Google Scholar

[5] A. Thom and J. Wilson, Metric ultraproducts of finite simple groups, Comp. Rend. Math. 352 (2014), no. 6, 463–466. 10.1016/j.crma.2014.03.015Search in Google Scholar

[6] A. Thom and J. Wilson, Some geometric properties of metric ultraproducts of finite simple groups, Israel J. Mathematics 227 (2018), no. 1, 113–129. 10.1007/s11856-018-1721-1Search in Google Scholar

Received: 2019-11-11
Revised: 2020-02-04
Published Online: 2020-03-28
Published in Print: 2020-09-01

© 2020 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 5.5.2024 from https://www.degruyter.com/document/doi/10.1515/jgth-2019-0165/html
Scroll to top button