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

Projective Limits and Ultraproducts of Nonabelian Finite Groups

Nazih Nahlus, Yilong Yang
Abstract

Groups that can be approximated by finite groups have been the center of much research. This has led to the investigations of the subgroups of metric ultraproducts of finite groups. This paper attempts to study the dual problem: what are the quotients of ultraproducts of finite groups?

Since an ultraproduct is an abstract quotient of the direct product, this also led to a more general question: what are the abstract quotients of profinite groups? Certain cases were already well-studied. For example, if we have a pro-solvable group, then it is well-known that any finite abstract quotients are still solvable.

This paper studies the case of profinite groups from a surjective inverse system of non-solvable finite groups. We shall show that, for various classes of groups 𝒞\mathcal{C}, any finite abstract quotient of a pro-𝒞\mathcal{C} group is still in 𝒞\mathcal{C}. Here 𝒞\mathcal{C} can be the class of finite semisimple groups, or the class of central products of quasisimple groups, or the class of products of finite almost simple groups and finite solvable groups, or the class of finite perfect groups with bounded commutator width.

Furthermore, in the case of finite perfect groups with unbounded commutator width, we show that this is NOT the case. In fact, any finite group could be a quotient of an ultraproduct of finite perfect groups. We provide an explicit construction on how to achieve this for each finite group.

1 Introduction

1.1 Background and Main Results

Groups that can be approximated by finite groups have been the center of much research. The standard approach is to study subgroups of “limits” of finite groups. For example, in the terminology of Hold and Rees [6], sofic groups are subgroups of the metric ultraproducts of finite symmetric groups, with respect to the normalized Hamming distance (in the sense of Definition 1.6 from [14]). It has gathered the attention of many works, and it is related to Kaplansky’s Direct Finiteness Conjecture [4], Connes’ embedding conjecture [3], and many other problems. As a generalization of sofic groups, one can also consider the 𝒞\mathcal{C}-approximable groups, which are groups that can be embedded as subgroups of the metric ultraproducts of groups in a class of finite groups 𝒞\mathcal{C}. See [11] for further discussion.

This paper was motivated by the “dual problem”, i.e., given an ultraproduct of finite groups, what are the possible quotients? It turns out that such “dual” problems have their own importance. For example, motivated by gauge field theories in physics, Zilbert asks the following question in [18]:

Conjectures 1.1.

Can a compact simple Lie group be a quotient of an ultraproduct of finite groups?

The answer to Zilbert’s problem is negative, shown by [11]. Nevertheless, the questions on quotients of ultraproducts of finite groups have not been studied much.

In a parallel manner, Bergman and Nahlus have shown the following interesting results on Lie algebra in [2].

Theorem 1.2.

Let 𝒞\mathcal{C} be the class of nilpotent Lie algebras, the class of solvable Lie algebras, or the class of semi-simple Lie algebras with characteristics other than 2 or 3. If we have an ultraproduct of Lie algebra sAnA_{n} in a class 𝒞\mathcal{C}, and a surjective homomorphism f:nωAnBf:\prod_{n\to\omega}A_{n}\to B where BB is a finite dimensional Lie algebra, then BB is also in class 𝒞\mathcal{C}.

Motivated by such results, it is natural to consider the parallel situation in the case of groups. Let 𝒞\mathcal{C} be a certain class of finite groups. We ask two kinds of questions:

  1. 1.

    (Weak version) If we have an ultraproduct of groups GnG_{n} in a class 𝒞\mathcal{C}, and a surjective homomorphism f:nωGnHf:\prod_{n\to\omega}G_{n}\to H where HH is a finite group, then when is BB also in class 𝒞\mathcal{C}?

  2. 2.

    (Strong version) If we have a projective limit GG of a surjective inverse direct system of groups GnG_{n} in a class 𝒞\mathcal{C}, and a surjective homomorphism f:GHf:G\to H where HH is a finite group, then when is BB also in class 𝒞\mathcal{C}?

The strong version would immediately imply the weak version since the ultraproduct of groups is a quotient of the direct product of the same groups.

The answer to the strong version of our question is already well known in certain cases. For example, if 𝒞\mathcal{C} is the class of finite abelian groups, then obviously any projective limit would also be abelian, and hence any quotient would also be abelian. By Corollary 4.2.3 of [12], the case of pp-groups is proven. Since each nilpotent group is a product of its pp-Sylow subgroups, the case of nilpotent groups follows. Finally, the case of solvable groups is proven by Corollary 4.2.4 of [12]. For all these groups, the answer to the strong version of our question is already known to be affirmative.

In this paper, we shall prove that the strong version is true for several other cases as well.

Theorem 1.3.

Let 𝒞\mathcal{C} be the class of finite semisimple groups, or the class of finite central products of finite quasisimple groups, or the class of finite perfect groups with bounded commutator width. Then any finite quotient of a pro-𝒞\mathcal{C} group must remain in the class 𝒞\mathcal{C}.

If 𝒞\mathcal{C} is the class of finite products of finite almost simple groups, then any finite quotient of a pro-𝒞\mathcal{C} group is a direct product of almost simple groups and solvable groups.

In fact, we also showed that statements of this type are true for certain variants of the classes of groups above. Note that Theorem 1.3 is dependent on the classification of finite simple groups.

However, what if we move away from the solvable groups or groups that are close to being simple? For example, if we choose 𝒞\mathcal{C} to be the class of finite perfect groups, then even the weak version of our question is false. For example, in Theorem 2.1.11 of [5], Holt and Plesken have constructed specific examples of finite perfect groups GnG_{n} with huge centers, such that GnG_{n} must have large commutator width. Here the commutator width of a group GG refers to the least integer nn such that every element of GG is a product of at most nn commutators.

If a sequence of perfect groups has increasing commutator width, then it should be expected that their “limit” of any kind would have “infinite commutator width”, i.e., it is probably not going to be perfect after all. Then it would have a non-trivial abelian quotient, which is not in 𝒞\mathcal{C}. This is indeed the case, as shown in Proposition 1.5 in Section 3 of this paper.

Another such example is constructed by Nikolov [10], where the groups GnG_{n} are even constructed to be finite perfect linear groups of dimension 15. However, it is worth noting that in both examples by Holt and Plesken (Theorem 2.1.11 of [5]) and the one by Nikolov, nωGn\prod_{n\to\omega}G_{n} must either have non-solvable quotients, or abelian quotients. There is nothing in between.

In pursuit of this phenomenon, Holt111https://mathoverflow.net/questions/280887/subdirect-product-of-perfect-groups asked that, if GG is a subdirect product of finitely many perfect groups, must its solvable quotients be abelian? This is a “finite” version of our inquiry, and it was answered by Mayr and Ruškuc in the negative [9]. It turns out that such a quotient is not necessarily abelian, but must always be nilpotent.

In a related manner, Thiel222https://mathoverflow.net/questions/289390/inverse-limits-of-perfect-groups asked whether all groups can be achieved as a projective limit of perfect groups. An answer was provided by de Cornulier, but his construction was through an injective directed system of infinite perfect groups.

In this regard, we would establish the following result in our paper as well.

Theorem 1.4.

For any finite group GG, GG is a quotient of the projective limit of a surjective inverse directed system of finite perfect groups.

Note that Theorem 1.4 is not dependent on the classification of finite simple groups.

1.2 Outline of the paper

In Section 2, we shall establish the proof of Theorem 1.3, using mostly elementary arguments combined with the main results on ultraproducts of finite simple groups from [15]. We can break it down into the following subsections:

  • In Section 2.1, we establish properties of a class 𝒞\mathcal{C} such that any pro-𝒞\mathcal{C} is a direct product of groups in 𝒞\mathcal{C}. This shall come in handy in almost all cases below.

  • In Section 2.2, we establish the semisimple case of Theorem 1.3.

  • In Section 2.3, we establish the quasisimple case of Theorem 1.3.

  • In Section 2.4, we establish the almost simple case of Theorem 1.3.

  • In Section 2.5, we define a slightly more generalized notion of almost semi-simple groups, and prove a similar result to Theorem 1.3.

  • In Section 2.6, we establish the case of finite perfect groups with bounded commutator width for Theorem 1.3.

We shall prove Theorem 1.4 in Section 3. The construction here is a bit elaborate, so we shall provide a brief breakdown of the arguments here.

We shall show that any finite group can be realized as a quotient of Pn\prod P_{n}, a direct product of a family of finite perfect groups. First of all, if groups in the sequence PnP_{n} are perfect groups with increasing commutator width, then their direct product is not perfect, and therefore they would have a non-trivial abelian quotient, as shown in the proposition below.

Proposition 1.5.

If {Pn}n\{P_{n}\}_{n\in\mathbb{N}} is a sequence of finite perfect groups with strictly increasing commutator width, and ω\omega is any non-principal ultrafilter on \mathbb{N}, then P=nωPnP=\prod_{n\to\omega}P_{n} is not perfect.

Proof.

Since PnP_{n} has strictly increasing commutator width, each PnP_{n} must have a commutator width of at least nn. Pick gnPng_{n}\in P_{n} such that it cannot be written as the product of less than nn commutators.

If limnωgnP\lim_{n\to\omega}g_{n}\in P can be written as the product of dd commutators, then we must have {n:gn is the product of d commutators}ω\{n\in\mathbb{N}:g_{n}\text{ is the product of $d$ commutators}\}\in\omega. However, this set is finite, while ω\omega is a non-principal ultrafilter, a contradiction. Hence limnωgn\lim_{n\to\omega}g_{n} is not in the commutator subgroup of PP. In particular, PP is not perfect. ∎

We now establish the generic case. For an arbitrary finite group GG, we would construct a surjective homomorphism from a “pro-perfect” group onto GG. The homomorphism we constructed would in fact factor through an ultraproduct, as described in the statement of Theorem 1.6.

Theorem 1.6.

Given any finite group GG, there is a sequence of finite perfect groups {Pn}n\{P_{n}\}_{n\in\mathbb{N}} and an ultrafilter ω\omega on \mathbb{N} such that the ultraproduct nωPn\prod_{n\to\omega}P_{n} has a surjective homomorphism onto GG.

Since an ultraproduct is a quotient of the direct product, therefore Theorem 1.4 is simply a corollary of Theorem 1.6 above.

We first reduce the generic case to the case of a special kind of groups. In 1989, Zelmanov solved the restricted Burnside problem in [16, 17]. In particular, the following theorem is true.

Theorem 1.7.

For arbitrary positive integers d,md,m, there is a unique finite group B(d,m)B(d,m), such that any finite group of exponent mm on dd generators is isomorphic to a quotient of B(d,m)B(d,m).

So to prove Theorem 1.6, it is enough to show that for arbitrary positive integers d,md,m, the restricted Burnside group B(d,m)B(d,m) can be realized as an abstract quotient of an ultraproduct of finite perfect groups.

We shall provide an explicit construction of these groups PnP_{n}. The proof can be decomposed into the following steps:

  • In Section 3.1, we construct and discuss the restricted Burnside coproduct, defined on finite groups with exponents dividing mm. This is a crucial ingredient for Section 3.3.

  • In Section 3.2, we discuss some functorial properties of the ultraproduct and its interaction with the restricted Burnside coproduct. This is a crucial ingredient for Section 3.3.

  • In Section 3.3, we reduce the case of the restricted Burnside group B(d,m)B(d,m) to the case of the cyclic group /m\mathbb{Z}/m\mathbb{Z}. In particular, if the cyclic group /m\mathbb{Z}/m\mathbb{Z} could be realized as an abstract quotient of an ultraproduct of “nice” finite perfect groups, then so can the restricted Burnside group B(d,m)B(d,m).

  • In Section 3.4, we construct a sequence of finite perfect groups GnG_{n}.

  • In Section 3.5, we construct a sequence of finite abelian groups MnM_{n} with exponents dividing mm, and each MnM_{n} has a GnG_{n} action.

  • In Section 3.6, we show that the groups Pn=MnGnP_{n}=M_{n}\rtimes G_{n} are perfect with increasing commutator width.

  • In Section 3.7, we establish some properties on nωMn\prod_{n\to\omega}M_{n} and its “dual”.

  • In Section 3.8, we construct an explicit homomorphism ψ:nωMn/m\psi:\prod_{n\to\omega}M_{n}\to\mathbb{Z}/m\mathbb{Z}, which could be extended to a surjective homomorphism from nωPn\prod_{n\to\omega}P_{n} to /m\mathbb{Z}/m\mathbb{Z}.

1.3 Notations and Conventions

Given a perfect group GG and an element gGg\in G, we define its commutator length to be the smallest integer dd such that it can be written as a product of at most dd commutators. We define the commutator width of GG to be the smallest integer dd such that every element of GG has commutator length at most dd.

Given a group GG, for elements x,yGx,y\in G, subgroups H,KGH,K\leq G, subset SGS\subseteq G, we adopt the following notations. Here S\langle S\rangle means the subgroup generated by the subset SS.

gh\displaystyle g^{h} h1gh,\displaystyle\coloneqq h^{-1}gh,
[g,h]\displaystyle[g,h] g1h1gh,\displaystyle\coloneqq g^{-1}h^{-1}gh,
[H,K]\displaystyle[H,K] [g,h]gH and hK,\displaystyle\coloneqq\langle[g,h]\mid\text{$g\in H$ and $h\in K$}\rangle,
H\displaystyle H^{\prime} [H,H],\displaystyle\coloneqq[H,H],
Sk\displaystyle S^{\star k} {s1s2sks1,,skS}.\displaystyle\coloneqq\{s_{1}s_{2}\dots s_{k}\mid s_{1},\dots,s_{k}\in S\}.

In our paper, we adopt the convention that all group actions are from the right. Given a group GG acting on another group HH via ϕ:GAut(H)\phi:G\to\mathrm{Aut}(H), the corresponding semi-direct product is written as HGH\rtimes G. Elements of this semi-direct product group is written as hghg where hHh\in H and gGg\in G, while the action of gg on hh is written as hgh^{g}.

For hH,gGh\in H,g\in G, and subgroup KHK\leq H, we adopt the following conventions.

hg\displaystyle h^{g} ϕg(h)=g1hg,\displaystyle\coloneqq\phi_{g}(h)=g^{-1}hg,
[h,g]\displaystyle[h,g] h1hg,\displaystyle\coloneqq h^{-1}h^{g},
[K,G]\displaystyle[K,G] [h,g]hK,gG.\displaystyle\coloneqq\langle[h,g]\mid h\in K,g\in G\rangle.

Throughout this paper, all finite groups will be equipped with the discrete topology whenever needed. Let 𝒞\mathcal{C} be a class of finite groups equipped with the discrete topology. Then we say GG is a pro-𝒞\mathcal{C} group if it is the inverse direct limit of a surjective system of groups in 𝒞\mathcal{C}.

Other than the inverse direct limit, another kind of limit that concerns us is the ultraproduct. Given any index set II, a collection ω\omega of subsets of II is an ultrafilter if the followings are true:

  1. 1.

    If J1,J2ωJ_{1},J_{2}\in\omega, then J1J2ωJ_{1}\cap J_{2}\in\omega.

  2. 2.

    If J1J2J_{1}\subseteq J_{2} as subsets of II, and J1ωJ_{1}\in\omega, then J2ωJ_{2}\in\omega.

  3. 3.

    For any subset JJ of II, exactly one of JJ and its complement IJI-J is in ω\omega.

An ultrafilter is principal if the intersection of all elements of ω\omega is non-empty.

Given a sequence of groups {Gn}n\{G_{n}\}_{n\in\mathbb{N}}, and an ultrafilter ω\omega on \mathbb{N}, consider the subgroup NωN_{\omega} of the group nGn\prod_{n\in\mathbb{N}}G_{n} defined as Nω{(gn)nnGn:{gn=e}ω}N_{\omega}\coloneqq\{(g_{n})_{n\in\mathbb{N}}\in\prod_{n\in\mathbb{N}}G_{n}:\{g_{n}=e\}\in\omega\}, where ee is the identity element. This is a normal subgroup, and we denote the quotient nGn/Nω\prod_{n\in\mathbb{N}}G_{n}/N_{\omega} as nωGn\prod_{n\to\omega}G_{n}. This is the ultraproduct of the sequence {Gn}n\{G_{n}\}_{n\in\mathbb{N}} through the ultrafilter ω\omega.

For the element (gn)nnGn(g_{n})_{n\in\mathbb{N}}\in\prod_{n\in\mathbb{N}}G_{n}, its quotient image in nωGn\prod_{n\to\omega}G_{n} is denoted as limnωgn\lim_{n\to\omega}g_{n}, called the ultralimit of the sequence (gn)n(g_{n})_{n\in\mathbb{N}}.

For further discussions, see [13].

2 Projective limits of non-abelian groups

In this section, we consider projective limits of finite groups that are close to being non-abelian finite simple. Throughout this section, we use the term semisimple group to refer to a direct product of non-abelian finite simple groups.

2.1 Preliminary on Profinite groups

It is well-known that any projective limits of finite semisimple groups must be a direct product of finite simple groups. (E.g., see Lemma 8.2.3 of [12].) In this preliminary section, we try to generalize such a result to some similar classes of groups.

Let 𝒞\mathcal{C} be a class of finite groups. Let 𝒫(𝒞)\mathcal{P}(\mathcal{C}) be the class of finite products of groups in 𝒞\mathcal{C}.

Let us say that 𝒞\mathcal{C} is a central-simple class of groups if all groups in 𝒞\mathcal{C} are non-abelian, and for any homomorphism f:GHf:G\to H between groups G,H𝒞G,H\in\mathcal{C}, f(G)CH(f(G))=Hf(G)\mathrm{C}_{H}(f(G))=H implies that ff is trivial or ff is an isomorphism. Here CH(f(G))\mathrm{C}_{H}(f(G)) refers to the centralizer of f(G)f(G). We claim that the following result is true.

Proposition 2.1.

Suppose 𝒞\mathcal{C} is a central-simple class of finite groups. Then any pro-𝒫(𝒞)\mathcal{P}(\mathcal{C}) group is a direct product of groups in 𝒞\mathcal{C}.

Now let us prove Proposition 2.1. For any group GG, let us say a normal subgroup NN of GG is 𝒞\mathcal{C}-normal if G/N𝒞G/N\in\mathcal{C}, and we write N𝒞GN\trianglelefteq_{\mathcal{C}}G.

Lemma 2.2.

Suppose 𝒞\mathcal{C} is a central-simple class of finite groups. Then for any G𝒫(𝒞)G\in\mathcal{P}(\mathcal{C}), the natural homomorphism from GG to N𝒞GG/N\prod_{N\trianglelefteq_{\mathcal{C}}G}G/N is an isomorphism.

Proof.

Suppose G𝒫(𝒞)G\in\mathcal{P}(\mathcal{C}), and it is the internal direct product G=iIGiG=\prod_{i\in I}G_{i} where Gi𝒞G_{i}\in\mathcal{C} for all iIi\in I. It is enough to show that the only 𝒞\mathcal{C}-normal subgroups are kernels of the projection maps onto some factor group.

We perform induction on the number of factor groups, i.e., on the cardinality of the index set II. If II has only one element, then GG is a group in 𝒞\mathcal{C}. If NN is a 𝒞\mathcal{C}-normal subgroup of GG, then the quotient map qq from GG to G/NG/N is a surjective homomorphism between groups in 𝒞\mathcal{C}. In particular, we must have q(G)CG/N(q(G))=G/Nq(G)\mathrm{C}_{G/N}(q(G))=G/N, so qq is either trivial or an isomorphism. Since the trivial group (which is abelian) is not in 𝒞\mathcal{C}, therefore the quotient map must be an isomorphism. Hence the only 𝒞\mathcal{C}-normal subgroup of GG is the trivial subgroup, and the statement is true.

Now suppose II has more than one element. Let NN be any 𝒞\mathcal{C}-normal subgroup. Consider the quotient map q:GG/Nq:G\to G/N. For any iIi\in I, since GiG_{i} is a direct factor of GG, any element of q(Gi)q(G_{i}) must commute with any element of q(Gj)q(G_{j}) for jij\neq i. In particular, we have q(Gi)CG/N(q(Gi))=G/Nq(G_{i})\mathrm{C}_{G/N}(q(G_{i}))=G/N. So qq restricted to GiG_{i} must either be trivial or an isomorphism.

If qq is trivial on all GiG_{i}, then G/NG/N is trivial and not in 𝒞\mathcal{C}, which is impossible. Hence qq restricts to an isomorphism to some GiG_{i}. Furthermore, if qq restricted to GiG_{i} and qq restricted to GjG_{j} are both isomorphisms for some iji\neq j, then all elements in q(Gi)=G/Nq(G_{i})=G/N and in q(Gj)=G/Nq(G_{j})=G/N must commute, so G/NG/N is an abelian group and not in 𝒞\mathcal{C}, also impossible. Hence, there is a unique iIi\in I such that qq restricted to GiG_{i} is an isomorphism, and qq kills all other GjG_{j} where jij\neq i.

In particular, qq is a projection map onto a factor group GiG_{i}. So we are done. ∎

Proof of Proposition 2.1.

Let GG be a pro-𝒫(𝒞)\mathcal{P}(\mathcal{C}) group. Then it is the projective limits of G/U𝒫(𝒞)G/U\in\mathcal{P}(\mathcal{C}) for all open 𝒫(𝒞)\mathcal{P}(\mathcal{C})-normal subgroups UU.

But for each open 𝒫(𝒞)\mathcal{P}(\mathcal{C})-normal subgroup UU, by Lemma 2.2, we have an isomorphism G/UNUG/NG/U\simeq\prod_{N\in\mathcal{M}_{U}}G/N where U\mathcal{M}_{U} is the set of open 𝒞\mathcal{C}-normal subgroups containing UU. By taking inverse limits of these isomorphisms, we have an isomorphism GNG/NG\simeq\prod_{N\in\mathcal{M}}G/N where U\mathcal{M}_{U} is the set of open 𝒞\mathcal{C}-normal subgroups of GG. ∎

2.2 Semisimple groups

We start with projective limits of finite semisimple groups. Note that, as we have already noted, this must be a direct product of non-abelian finite simple groups. Such groups are also sometimes called semisimple profinite groups. Let us first describe all maximal normal subgroups of a semisimple profinite group.

Given a semisimple profinite group G=iISiG=\prod_{i\in I}S_{i} for a family of non-abelian finite simple groups SiS_{i}, for any subset JIJ\subseteq I, we define GJ:={(gi)iIGgi=e if iJ}G_{J}:=\{(g_{i})_{i\in I}\in G\mid g_{i}=e\text{ if $i\notin J$}\}, i.e., the subgroup of elements whose “support” is inside JJ. Clearly GG is the internal direct product of GJG_{J} and GIJG_{I-J}.

For any ultrafilter 𝒰\mathcal{U} on II, we define N𝒰:={(gi)iIG{iIgi=e}𝒰}N_{\mathcal{U}}:=\{(g_{i})_{i\in I}\in G\mid\{i\in I\mid g_{i}=e\}\in\mathcal{U}\}. Note that the ultraproduct of the groups SiS_{i} is by definition G/N𝒰G/N_{\mathcal{U}}.

Proposition 2.3.

If G=iISiG=\prod_{i\in I}S_{i} for a family of non-abelian finite simple groups SiS_{i}, and MM is any abstract maximal normal subgroup of GG, then N𝒰MN_{\mathcal{U}}\subseteq M for some ultrafilter 𝒰\mathcal{U} on MM.

Proof.

Note that if GG is an arbitrary product of non-abelian finite simple groups, then each factor group of GG has commutator width 11 by the classification of non-abelian finite simple groups. So GG has commutator width 11, and thus it is perfect. So G/MG/M is also a perfect group.

Since MM is maximal, G/MG/M is perfect and simple. So by Corollary 2.9 below, if G=iISiG=\prod_{i\in I}S_{i} for a family of non-abelian finite simple groups SiS_{i}, then the quotient map GG/MG\to G/M must factor through some ultraproduct i𝒰Si\prod_{i\to\mathcal{U}}S_{i} for some ultrafilter 𝒰\mathcal{U} on II. So we are done. ∎

Corollary 2.4.

If G=iISiG=\prod_{i\in I}S_{i} for a family of non-abelian finite simple groups SiS_{i}, and MM is any abstract maximal normal subgroup of GG with finite index, then M=N𝒰M=N_{\mathcal{U}} for some ultrafilter 𝒰\mathcal{U} on II, and G/MG/M is isomorphic to SiS_{i} for some ii.

Proof.

By Proposition 2.3, we know that we can find an ultrafilter 𝒰\mathcal{U} such that N𝒰MN_{\mathcal{U}}\subseteq M. But by [15], G/N𝒰G/N_{\mathcal{U}} is either isomorphic to SiS_{i} for some ii, or it cannot have any non-trivial finite dimensional unitary representation. But G/MG/M is a finite quotient of G/N𝒰G/N_{\mathcal{U}}, which must have a non-trivial finite dimensional unitary representation. Hence G/N𝒰G/N_{\mathcal{U}} is either isomorphic to SiS_{i} for some ii, and thus M=N𝒰M=N_{\mathcal{U}} and G/MG/M is isomorphic to SiS_{i} for some ii. ∎

Corollary 2.5.

If G=iISiG=\prod_{i\in I}S_{i} for a family of non-abelian finite simple groups SiS_{i}, then intersections of maximal normal subgroups of finite indices of GG is a perfect group.

Proof.

Using Corollary 2.4, let these maximal normal subgroups of finite indices be {N𝒰k}kK\{N_{\mathcal{U}_{k}}\}_{k\in K} for some index set KK and ultrafilters 𝒰k\mathcal{U}_{k} on II. For each g=(gi)iIGg=(g_{i})_{i\in I}\in G, let supp(g)={iI:gie}\mathrm{supp}(g)=\{i\in I:g_{i}\neq e\}. Then gN𝒰kg\in N_{\mathcal{U}_{k}} if and only if supp(g)𝒰k\mathrm{supp}(g)\notin\mathcal{U}_{k}. So gkKN𝒰kg\in\bigcap_{k\in K}N_{\mathcal{U}_{k}} if and only if supp(g)kK𝒰k\mathrm{supp}(g)\notin\bigcup_{k\in K}\mathcal{U}_{k}.

In particular, kKN𝒰k\bigcap_{k\in K}N_{\mathcal{U}_{k}} is generated by GJG_{J} where JkK𝒰kJ\notin\bigcup_{k\in K}\mathcal{U}_{k}. Since each GJG_{J} is also a semisimple profinite group, hence a perfect group, therefore the subgroup they generate must also be perfect. So we are done. ∎

Now we can establish our desired result.

Proposition 2.6.

If GG is a semisimple profinite group, then any finite quotient of GG is a finite semisimple group.

Proof.

Say G=iISiG=\prod_{i\in I}S_{i} for a family of non-abelian finite simple groups SiS_{i}. Suppose for contradiction that our statement is not true, and let BB be a minimal counterexample, i.e., BB is a finite quotient of GG, but it is not semisimple. Let q:GBq:G\to B be the quotient map.

If BB has a non-trivial center ZZ, then by minimality, B/ZB/Z is a finite semisimple group. Therefore, ZZ is an intersection of maximal normal subgroups of BB with finite indices. Pulling back to GG, we see that q1(Z)q^{-1}(Z) is an intersection of maximal normal subgroups of BB with finite indices. Therefore according to Corollary 2.5, it is a perfect group. Hence its image ZZ must also be a perfect group. But ZZ is also abelian, hence it must be trivial, a contradiction.

So BB must be centerless. Furthermore, if BB is the direct product of two non-trivial groups, then by minimality of BB, the two factor groups must both be semisimple groups. Hence BB itself is a semisimple group, a contradiction.

So BB must be a centerless group, and it is not the direct product of two non-trivial groups. By Corollary 2.8 below, this means that q:GBq:G\to B must factor through an ultraproduct. But by [15] again, this means that BB is either infinite, or BB is isomorphic to SiS_{i} for some ii. Neither case is possible.

To sum up, such a counterexample is impossible. ∎

Here are the necessary lemmas for the proofs above. The idea is adapted from Bergman and Nahlus [2], who proved a similar statement for Lie algebras.

Lemma 2.7 ([2, Lemma 7]).

Suppose (Ai)iI(A_{i})_{i\in I} is a family of sets, BB is a set, and f:iIAiBf:\prod_{i\in I}A_{i}\to B is a set map whose image has more than one element. Then the following are equivalent:

  1. (a)

    For every subset JIJ\subseteq I, the map ff factors either through the projection iIAiiJAi\prod_{i\in I}A_{i}\to\prod_{i\in J}A_{i}, or through the projection iIAiiIJAi\prod_{i\in I}A_{i}\to\prod_{i\in I-J}A_{i}.

  2. (b)

    The map ff factors through the natural map iIAiiIAi/𝒰\prod_{i\in I}A_{i}\to\prod_{i\in I}A_{i}/\mathcal{U}, where 𝒰\mathcal{U} is an ultrafilter on the index set II.

When these hold, the ultrafilter 𝒰\mathcal{U} is uniquely determined by the map ff.

Corollary 2.8 (Analogue of [2, Lemma 5]).

Let BB be a non-trivial group. Then the following are equivalent:

  1. (i)

    Every surjective homomorphism f:A1×A2Bf:A_{1}\times A_{2}\to B from a direct product of two groups A1,A2A_{1},A_{2} onto BB factors through the projection of A1×A2A_{1}\times A_{2} onto A1A_{1} or the projection of A1×A2A_{1}\times A_{2} onto A2A_{2}.

  2. (ii)

    There are no non-trivial normal subgroups N1,N2N_{1},N_{2} of BB such that B=N1N2B=N_{1}N_{2} and [N1,N2]={e}[N_{1},N_{2}]=\{e\}.

  3. (iii)

    The center Z(B)Z(B) of BB is trivial, and BB is not the direct product of two non-trivial groups.

  4. (iv)

    For every surjective homomorphism f:iIGiBf:\prod_{i\in I}G_{i}\to B from a direct product of a family of groups GiG_{i} over any index set II onto BB, ff factors through a quotient map from iIGi\prod_{i\in I}G_{i} onto some ultraproduct iIGi\prod_{i\in I}G_{i} over some ultrafilter 𝒰\mathcal{U} on II.

Proof.

First let us show that the negation of (i)(i), (ii)(ii), and (iii)(iii) are equivalent.

Suppose (i)(i) is false. Then let f:A1×A2Bf:A_{1}\times A_{2}\to B be a counterexample. Let N1=f(A1×{e})N_{1}=f(A_{1}\times\{e\}) and N2=f({e}×A2)N_{2}=f(\{e\}\times A_{2}). Then N1N_{1} and N2N_{2} are non-trivial. Since ff is surjective, N1N_{1} and N2N_{2} are normal subgroups of BB, and B=N1N2B=N_{1}N_{2}. Finally,

[N1,N2]=[f(A1×{e}),f({e}×A2)]=f([A1×{e},{e}×A2])={e}.[N_{1},N_{2}]=[f(A_{1}\times\{e\}),f(\{e\}\times A_{2})]=f([A_{1}\times\{e\},\{e\}\times A_{2}])=\{e\}.

So (ii)(ii) is false.

Suppose (ii)(ii) is false. Let N1,N2N_{1},N_{2} be a pair of normal subgroups of BB and a counterexample to (ii)(ii). If N1N2={e}N_{1}\cap N_{2}=\{e\}, then B=N1×N2B=N_{1}\times N_{2}, and (iii)(iii) is false. Suppose N1N2N_{1}\cap N_{2} is non-trivial. For any element aN1N2a\in N_{1}\cap N_{2}, since aN2a\in N_{2} and [N1,N2]={e}[N_{1},N_{2}]=\{e\}, aa must commute with every element of N1N_{1}. Similarly, since aN1a\in N_{1}, it must commute with every element of N2N_{2}. And since B=N1N2B=N_{1}N_{2}, we must have aZ(B)a\in Z(B). So Z(B)N1N2Z(B)\supseteq N_{1}\cap N_{2} is non-trivial. So (iii)(iii) is false.

Now suppose (iii)(iii) is false. If BB is the direct product of two non-trivial groups, then obviously (i)(i) is false. Suppose Z(B)Z(B) is non-trivial. Then let f:Z(B)×BBf:Z(B)\times B\to B be the homomorphism induced by the inclusion homomorphism of Z(B)Z(B) into BB and the identity homomorphism of BB. This provides a counterexample of (i)(i).

Now we shall show that (i)(i) and (iv)(iv) are equivalent.

(i)(i) is a special case of (iv)(iv) when I={1,2}I=\{1,2\}. So we only need to show that (i)(i) implies (iv)(iv). But this follows easily from Lemma 2.7. ∎

Corollary 2.9.

If BB is a perfect simple group, then for any surjective homomorphism f:iIGiBf:\prod_{i\in I}G_{i}\to B from a direct product of a family of groups GiG_{i} over any index set II onto BB, ff factors through a quotient map from iIGi\prod_{i\in I}G_{i} onto some ultraproduct i𝒰Gi\prod_{i\to\mathcal{U}}G_{i} over some ultrafilter 𝒰\mathcal{U} on II.

Proof.

Since BB is perfect and simple, it cannot have a non-trivial center, and it cannot be the direct product of two non-trivial subgroups. So the statement is true by Corollary 2.8. ∎

2.3 Quasisimple groups

A group is quasisimple if it is a perfect central extension of a non-abelian finite simple group. Analogously, one can define a “quasi-semisimple group” to be a perfect central extension of a finite semisimple group. It turns out that such a group is simply a central product of quasisimple groups.

Lemma 2.10.

A finite perfect central extension of a finite semisimple group is a central product of finite quasisimple groups.

Proof.

Suppose GG is a perfect central extension of a finite semisimple group SS, which is the internal direct product of non-abelian finite simple subgroups S1,,SkS_{1},\dots,S_{k}. Let the quotient map be q:GSq:G\to S, let Mi=q1(Si)M_{i}=q^{-1}(S_{i}), i.e., the pull-back in the following diagram.

Mi{M_{i}}Gi{Gi}Si{S_{i}}S{S}

Let QiQ_{i} be the unique maximal perfect normal subgroup of MiM_{i}. Now each QiQ_{i} is perfect by construction, and since MiM_{i} is finite, Mi/QiM_{i}/Q_{i} is solvable. Therefore q(Qi)=Siq(Q_{i})=S_{i}. So, we have the following diagram.

Qi{Q_{i}}Mi{M_{i}}Gi{Gi}Si{S_{i}}Si{S_{i}}S{S}

Finally, since the kernel of qq is the center of GG, we see that QiQ_{i} must be a perfect central extension of SiS_{i}, i.e., it is quasisimple. I now claim that GG is the central product of all these QiQ_{i}.

Let ZZ be the center of GG. Since q(Qi)=Siq(Q_{i})=S_{i}, we see that qq restricted to Q1QkQ_{1}\dots Q_{k} is already surjective onto SG/ZS\simeq G/Z. Hence we must have G=Q1QkZG=Q_{1}\dots Q_{k}Z.

If Q1QkQ_{1}\dots Q_{k} is a proper normal subgroup of GG, then G/(Q1Qk)G/(Q_{1}\dots Q_{k}) is non-trivial and abelian, which cannot happen since GG is perfect. So G=Q1QkG=Q_{1}\dots Q_{k}. It remains to be shown that [Qi,Qj][Q_{i},Q_{j}] is trivial for all iji\neq j.

Now QjQ_{j} acts on QiQ_{i} via conjugation, but [Qi,Qj]QiQj[Q_{i},Q_{j}]\subseteq Q_{i}\cap Q_{j} is in the kernel of qq, hence inside of the center ZZ. So QjQ_{j} acts trivially on Qi/(QiZ)SiQ_{i}/(Q_{i}\cap Z)\simeq S_{i}. But any automorphism of a finite quasisimple group must be trivial if it acts trivially on the corresponding simple group. So [Qi,Qj][Q_{i},Q_{j}] is in fact trivial.

So GG is indeed the central product of Q1,,QkQ_{1},\dots,Q_{k}. ∎

In general, a direct product of quasisimple groups might have a quotient that is not quasisimple anymore, but rather a central product of quasisimple groups, as shown in the following proposition.

Proposition 2.11.

If GG is an arbitrary direct product of finite quasisimple groups, then any finite quotient must be a central product of finite quasisimple groups.

Proof.

Say G=iIQiG=\prod_{i\in I}Q_{i} for a family of finite quasisimple groups QiQ_{i}. Let ZiZ_{i} be the center of QiQ_{i}, and set Z=iIZiZ=\prod_{i\in I}Z_{i}. Then ZZ is the center of GG, and G/ZG/Z is a semisimple profinite group.

Suppose we have a surjective homomorphism q:GBq:G\to B. Then B/q(Z)B/q(Z) is a finite quotient of the semisimple profinite group G/ZG/Z. Hence B/q(Z)B/q(Z) is semisimple. Furthermore, since ZZ is the center of GG and qq is surjective, q(Z)q(Z) is inside the center of BB. So BB is a central extension of the semisimple group B/q(Z)B/q(Z).

Finally, by classification of finite simple groups, any finite quasisimple group has commutator width at most 22. So GG has commutator width at most 22. So GG is perfect, and its quotient BB is also perfect.

So BB is a finite perfect central extension of a finite semisimple group. ∎

Now, the class of quasisimple groups is not central-simple (proper quotients of a quasisimple group might still be quasisimple). So the result above does not immediately generalize to projective limits that are not direct products.

However, by the theory of Schur multipliers, each non-abelian finite simple group has a unique maximal perfect central extension, i.e., the universal covering group for the non-abelian finite simple group. (See, e.g., Proposition 33.4 of [1].) And each quasisimple group must be a quotient of such a group.

These universal covering groups (i.e., quasisimple groups with trivial Schur multiplier) are also sometimes called superperfect quasisimple groups. Then we have the following result.

Proposition 2.12.

The class of finite superperfect quasisimple groups is central-simple.

Proof.

Suppose G,HG,H are finite superperfect quasisimple groups, and f:GHf:G\to H is any group homomorphism such that f(G)CH(f(G))=Hf(G)\mathrm{C}_{H}(f(G))=H. Since the subgroups f(G)f(G) and CH(f(G))\mathrm{C}_{H}(f(G)) both normalize f(G)f(G), therefore f(G)CH(f(G))=Hf(G)\mathrm{C}_{H}(f(G))=H implies that f(G)f(G) is normal in HH.

If f(G)Hf(G)\neq H, then note that all proper normal subgroups of a quasisimple group are central, so f(G)f(G) is abelian. But since GG is perfect, f(G)f(G) must be trivial.

If f(G)=Hf(G)=H, then GG is a central perfect extension of HH. But since HH is super perfect, it is already a maximal central perfect extension, hence ff can only be an isomorphism. ∎

Proposition 2.13.

If GG is a projective limit of finite central products of finite quasisimple groups, then any finite quotient must be a central product of finite quasisimple groups.

Proof.

Let 𝒞\mathcal{C} be the class of superperfect quasisimple groups. Then for each finite central product QQ of finite quasisimple groups {Qi}iI\{Q_{i}\}_{i\in I}, then QQ is a quotient of the direct products of these QiQ_{i} by a central subgroup. Let Q~i\widetilde{Q}_{i} be the universal perfect central extension of QiQ_{i}, then QQ is a central quotient of Q~=iIQ~i𝒫(𝒞)\widetilde{Q}=\prod_{i\in I}\widetilde{Q}_{i}\in\mathcal{P}(\mathcal{C}).

Now let GG be the projective limit of a directed family of finite central products of finite quasisimple groups {Gi}iI\{G_{i}\}_{i\in I}. Let G~i\widetilde{G}_{i} be the universal perfect central extension of GiG_{i}, so that G~i𝒫(𝒞)\widetilde{G}_{i}\in\mathcal{P}(\mathcal{C}). Let ZiZ_{i} be the kernel of the quotient map from G~i\widetilde{G}_{i} to GiG_{i}.

Note that any surjective homomorphism ϕij:GiGj\phi_{ij}:G_{i}\to G_{j} must induce a surjective homomorphism ϕ~ij:G~iG~j\widetilde{\phi}_{ij}:\widetilde{G}_{i}\to\widetilde{G}_{j}, and ϕ~ij\widetilde{\phi}_{ij} must also send the center into the center. So the directed family {Gi}iI\{G_{i}\}_{i\in I} must induce directed families {G~i}iI\{\widetilde{G}_{i}\}_{i\in I} and {Zi}iI\{Z_{i}\}_{i\in I}, and thus we have corresponding induced projective limits G~\widetilde{G} and Z~\widetilde{Z}.

Note that the projective limit on inverse systems of finite groups is an exact functor. (See, e.g., Proposition 2.2.4 of [12].) Therefore, GG is a quotient of G~\widetilde{G} by the closed subgroup ZZ. Since G~\widetilde{G} is a pro-𝒫(𝒞)\mathcal{P}(\mathcal{C}) group and 𝒞\mathcal{C} is a central-simple class, G~\widetilde{G} must be a direct product of superperfect quasisimple groups, and thus any finite quotient of G~\widetilde{G} must be a central product of finite quasisimple groups.

Since GG is a quotient of G~\widetilde{G}, the result holds for GG as well. ∎

Now, note that the class of finite central products of finite quasisimple groups is NOT closed under sub-direct product, so it is not a formation of groups.

Example 2.14.

Consider the group SL2(5)\mathrm{SL}_{2}(5) of 22 by 22 matrices over the field of order 55, whose determinants are 11. It is quasisimple over the finite simple group A5\mathrm{A}_{5}. Then SL2(5)×SL2(5)\mathrm{SL}_{2}(5)\times\mathrm{SL}_{2}(5) is a perfect central extension of A5×A5\mathrm{A}_{5}\times\mathrm{A}_{5}. Now take the diagonal subgroup of A5×A5\mathrm{A}_{5}\times\mathrm{A}_{5}, and take its pre-image in SL2(5)×SL2(5)\mathrm{SL}_{2}(5)\times\mathrm{SL}_{2}(5). This is a sub-direct product of finite quasisimple groups, but it is isomorphic to SL2(5)×(/2)\mathrm{SL}_{2}(5)\times(\mathbb{Z}/2\mathbb{Z}), NOT a central product of finite quasisimple groups.

So if one prefers to use a formation of groups, we can consider a larger class of groups, closed under sub-direct product.

Proposition 2.15.

The following three classes of groups are the same:

  1. 1.

    𝒞\mathcal{C} is the smallest class of groups containing all finite quasisimple groups which is closed under sub-direct products and quotients.

  2. 2.

    𝒞\mathcal{C} is the class of finite central extensions of finite semisimple groups.

  3. 3.

    𝒞\mathcal{C} is the class of finite central products of finite quasisimple groups and finite abelian groups.

Proof.

Let us call the three classes 𝒞1,𝒞2,𝒞3\mathcal{C}_{1},\mathcal{C}_{2},\mathcal{C}_{3}.

Since central products are quotients of the direct products, 𝒞1\mathcal{C}_{1} is closed under central products. Also note that any cyclic group ZZ can be realized as a center for some quasisimple group over PSLn(q)\mathrm{PSL}_{n}(q) for some n,qn,q. Let QQ be such a quasisimple group over a non-abelian finite simple group SS with center ZZ. Then Q2Q^{2} has a quotient S2S^{2}, and the pullback of the diagonal subgroup of S2S^{2} in Q2Q^{2} is a sub-direct product of two copies of QQ, and it is isomorphic to Q×ZQ\times Z. Hence 𝒞1\mathcal{C}_{1} must also contain all finite abelian groups. So 𝒞3𝒞1\mathcal{C}_{3}\subseteq\mathcal{C}_{1}.

Suppose G𝒞2G\in\mathcal{C}_{2}. Then we have a quotient map q:GSq:G\to S such that SS is semisimple, and the kernel of qq is the center ZZ of GG. Since SS is perfect, the unique largest perfect normal subgroup NN in GG must have q(N)=Sq(N)=S, and the restriction of qq to NN must have kernel NZN\cap Z, which is in the center of NN. So NN is a perfect central extension of SS. In particular, by Lemma 2.10, NN is a finite central product of finite quasisimple groups.

But since q(N)=Sq(N)=S and ZZ is the kernel of qq, we have G=NZG=NZ. And since ZZ is the center, [N,Z][N,Z] must be trivial. So GG is the central product of NN and the abelian group ZZ. So 𝒞2𝒞3\mathcal{C}_{2}\subseteq\mathcal{C}_{3}.

It remains to show that 𝒞1𝒞2\mathcal{C}_{1}\subseteq\mathcal{C}_{2}. Since 𝒞2\mathcal{C}_{2} contains all finite quasisimple groups, we only need to show that it is closed under sub-direct products and quotients.

The class 𝒞2\mathcal{C}_{2} is obviously closed under quotients. We now prove that sub-direct products of groups in 𝒞2\mathcal{C}_{2} are still in 𝒞2\mathcal{C}_{2}.

Suppose GG has two normal subgroups N1,N2N_{1},N_{2} such that N1N2N_{1}\cap N_{2} is trivial, and Gi=G/Ni𝒞2G_{i}=G/N_{i}\in\mathcal{C}_{2} for i=1,2i=1,2. Then GG is a finite group. Let MM be the intersection of all maximal normal subgroups of GG whose corresponding quotients are non-abelian finite simple groups. Then G/MG/M is a finite semisimple group. We need to show that MM is in the center of GG.

Consider the quotient map qi:GGiq_{i}:G\to G_{i}. Then qi(M)q_{i}(M) is in the intersection of all maximal normal subgroups of GiG_{i} whose corresponding quotients are non-abelian finite simple groups. Since Gi𝒞2G_{i}\in\mathcal{C}_{2}, qi(M)q_{i}(M) is the center of GiG_{i}. In particular, for any gGg\in G and xMx\in M, we have xgNi=gxNixgN_{i}=gxN_{i}. So, since N1N2N_{1}\cap N_{2} is trivial, we have

{xg}=(xgN1)(xgN2)=(gxN1)(gxN2)={gx}.\{xg\}=(xgN_{1})\cap(xgN_{2})=(gxN_{1})\cap(gxN_{2})=\{gx\}.

So xx is in the center of GG. So GG is a central extension of G/MG/M, and G𝒞2G\in\mathcal{C}_{2}. So we conclude that 𝒞1𝒞2\mathcal{C}_{1}\subseteq\mathcal{C}_{2}. ∎

Now we have a corresponding result on projective limits of finite central extensions of finite semisimple groups.

Proposition 2.16.

If GG is a projective limit of a family of finite central extensions of finite semisimple groups, then any finite quotient of GG is still a finite central extension of a finite semisimple group.

Proof.

Now if GG is a projective limit of a family of groups GiG_{i}, each a finite central extension of a finite semisimple group SiS_{i}. Let ZiZ_{i} be the center of GiG_{i}. Then we have a corresponding projective limit ZZ of the abelian groups ZiZ_{i}, and a projective limit SS of the semisimple groups SiS_{i}. Note that the projective limit on inverse systems of finite groups is an exact functor. (See, e.g., Proposition 2.2.4 of [12].) So ZZ is a closed central subgroup of GG and G/ZSG/Z\simeq S, a semisimple profinite group.

Suppose we have a surjective homomorphism q:GBq:G\to B where BB is finite. Then q(Z)q(Z) is a central subgroup of BB, and B/q(Z)B/q(Z) is a finite quotient of the semisimple profinite group G/ZG/Z. By Proposition 2.6, B/q(Z)B/q(Z) is a finite semisimple group. So BB is a finite central extension of a finite semisimple group. ∎

2.4 Projective limits via almost simple groups

We now briefly introduce the concept of almost simple groups.

Definition 2.17 (Definition of CR-radicals).

Given any finite group GG, we define its CR-radical CR(G)\mathrm{CR}(G) to be the largest semisimple normal subgroup. (“CR” stands for “completely reducible”.)

The CR-radical uniquely exists for all groups, and a solvable group must have a trivial CR-radical.

Definition 2.18 (Definition of Almost Simple Groups).

Given any finite group GG, we say it is almost simple if CR(G)\mathrm{CR}(G) is a non-abelian finite simple group, and GG has no non-trivial solvable normal subgroups.

Note that the above definition is equivalent to the usual definition of an almost simple group, which is a subgroup of Aut(S)\mathrm{Aut}(S) containing Inn(S)\mathrm{Inn}(S) for some non-abelian finite simple group SS.

Proposition 2.19 (Alternative Definition of Almost Simple Groups).

If GG is almost simple, then it is isomorphic to a subgroup of Aut(S)\mathrm{Aut}(S) containing Inn(S)\mathrm{Inn}(S) for some non-abelian finite simple group SS.

Proof.

Let SS be the CR-radical of GG. Then GG acts on SS by conjugation, which induces a homomorphism from GG to Aut(S)\mathrm{Aut}(S). Let NN be the kernel of this homomorphism. Assume for contradiction that it is not trivial.

Since NN is a normal subgroup of GG and CR(N)\mathrm{CR}(N) is characteristic in NN, therefore CR(N)\mathrm{CR}(N) is a semisimple normal subgroup of GG. Therefore CR(N)CR(G)=S\mathrm{CR}(N)\subseteq\mathrm{CR}(G)=S and by construction, every element of CR(N)\mathrm{CR}(N) commutes with every element of SS. Hence CR(N)\mathrm{CR}(N) is in the center of SS. But SS is a non-abelian finite simple group. So CR(N)\mathrm{CR}(N) is trivial.

Therefore, any minimal normal subgroup of NN must be a direct product of cyclic subgroups of prime order. In particular, the solvable radical MM of NN (the unique largest solvable normal subgroup of NN) is non-trivial. Since MM is characteristic in NN, it is a non-trivial solvable normal subgroup of GG, a contradiction.

So we must conclude that NN is trivial. Hence GG is isomorphic to a subgroup of Aut(S)\mathrm{Aut}(S). And since GG contains SS, the image of GG in Aut(S)\mathrm{Aut}(S) would contain Inn(S)\mathrm{Inn}(S). ∎

By the classification of finite simple groups, we also see that if GG is almost simple, then G/CR(G)G/\mathrm{CR}(G) is solvable.

Proposition 2.20.

The class of finite almost simple groups is central-simple.

Proof.

Suppose G,HG,H are finite almost simple groups, and f:GHf:G\to H is any group homomorphism such that f(G)CH(f(G))=Hf(G)\mathrm{C}_{H}(f(G))=H. Then f(G)f(G) is a normal subgroup of HH.

Suppose it is not trivial. Let SS be the unique minimal normal subgroup of HH. Then it is a finite simple group, and any non-trivial normal subgroup of HH must contain S=CR(H)S=\mathrm{CR}(H). So f(G)f(G) contains SS. However, since HH can be seen as a subgroup of Aut(S)\mathrm{Aut}(S), the centralizer of SS in HH is trivial. Hence CH(f(G))\mathrm{C}_{H}(f(G)) is trivial. So f(G)CH(f(G))=Hf(G)\mathrm{C}_{H}(f(G))=H implies that f(G)=Hf(G)=H.

Finally, note that proper quotients of an almost simple group must be solvable. But f(G)=Hf(G)=H is not solvable. So ff has a trivial kernel, and thus it is an isomorphism. ∎

As a result of Proposition 2.20, projective limits of finite products of almost simple groups must be a direct product of almost simple groups. Now, the quotient of such a profinite group might no longer be almost simple. For example, if we take a single almost simple group that is not simple itself, then it will have a solvable quotient. So we need to look at a larger class of groups.

Let 𝒞\mathcal{C} be the class of finite groups that are direct products of solvable groups and almost simple groups. Note that this class is closed under central products.

Proposition 2.21.

If G,H𝒞G,H\in\mathcal{C}, then any central product of G,HG,H is still in 𝒞\mathcal{C}.

Proof.

Note that 𝒞\mathcal{C} is obviously closed under direct products. So we only need to show that it is closed under central quotients.

Suppose G𝒞G\in\mathcal{C}, say GG is the internal direct product of almost simple subgroups G1,,GkG_{1},\dots,G_{k} and a solvable subgroup SS. Now GG acts on its normal subgroup CR(G1)CR(Gk)\mathrm{CR}(G_{1})\dots\mathrm{CR}(G_{k}) by conjugation. Since G1,,GkG_{1},\dots,G_{k} are all almost simple, the kernel of the induced homomorphism GAut(CR(G1)CR(Gk))G\to\mathrm{Aut}(\mathrm{CR}(G_{1})\dots\mathrm{CR}(G_{k})) is exactly SS. So the center ZZ of GG is contained inside of SS. Hence G/ZG1××Gk×(S/Z)𝒞G/Z\simeq G_{1}\times\dots\times G_{k}\times(S/Z)\in\mathcal{C}. ∎

In the remainder of this section we establish the following:

Proposition 2.22.

Let GG be a pro-𝒞\mathcal{C} group. Then any finite homomorphic image of GG must still be in 𝒞\mathcal{C}.

Proof.

Consider a potential counterexample. Suppose for contradiction that we have a pro-𝒞\mathcal{C} group G=iIGiG=\prod_{i\in I}G_{i}, each GiG_{i} is either finite solvable or finite almost simple, and ϕ:GB\phi:G\to B is a surjective homomorphism onto a finite group BB NOT in 𝒞\mathcal{C}. Since BB is finite, if such counterexample exists, then we can in fact let BB be a minimal counterexample.

Then by Corollary 2.30 below, BB cannot exist. Hence we are done. ∎

In the remainder of this section, we preserve the notations of G=iIGiG=\prod_{i\in I}G_{i} and ϕ:GB\phi:G\to B as in the minimal counterexample in the proof above.

For any subset JJ of II, let GJ={(gi)iI:gi=e if iJ}G_{J}=\{(g_{i})_{i\in I}:\text{$g_{i}=e$ if $i\notin J$}\}.

We use the notation CR(GJ)={(gi)iI:gi=e if iJ and giCR(Gi) if iJ}\mathrm{CR}(G_{J})=\{(g_{i})_{i\in I}:\text{$g_{i}=e$ if $i\notin J$ and $g_{i}\in\mathrm{CR}(G_{i})$ if $i\in J$}\}. Note that GJ/CR(GJ)G_{J}/\mathrm{CR}(G_{J}) is isomorphic to iJ(Gi/CR(Gi))\prod_{i\in J}(G_{i}/\mathrm{CR}(G_{i})), which is a pro-solvable group. We use CR(G)\mathrm{CR}(G) to denote CR(GI)\mathrm{CR}(G_{I}).

Lemma 2.23 (Dichotomy between complementary factors).

For any subset JJ of II, either ϕ(GJ)=B\phi(G_{J})=B or ϕ(GIJ)=B\phi(G_{I-J})=B. In particular, one is BB and the other one is inside the center of BB.

Proof.

Note that GG is the internal direct product of GJG_{J} and GIJG_{I-J}, hence B=ϕ(G)B=\phi(G) is the central product of ϕ(GJ)\phi(G_{J}) and ϕ(GIJ)\phi(G_{I-J}).

Note that ϕ(GJ)\phi(G_{J}) and ϕ(GIJ)\phi(G_{I-J}) are also finite homomorphic images of products of groups in 𝒞\mathcal{C}. So if both are proper subgroups, then by minimality of BB, both are in 𝒞\mathcal{C}. Hence by Proposition 2.21, BB as their central product is also in 𝒞\mathcal{C}. A contradiction.

So one of ϕ(GJ)\phi(G_{J}) and ϕ(GIJ)\phi(G_{I-J}) is BB. But since BB is the central product of these two groups, the other one must be in the center. ∎

In particular, let J={iI:Gi is solvable}J=\{i\in I:\text{$G_{i}$ is solvable}\}. Then BB is the central product of ϕ(GJ)\phi(G_{J}) and ϕ(GIJ)\phi(G_{I-J}). but since B𝒞B\notin\mathcal{C} while ϕ(GJ)\phi(G_{J}) must be solvable, we see that B=ϕ(GIJ)B=\phi(G_{I-J}). So we may safely throw away all solvable factor groups GiG_{i} from GG when we construct our counterexample. WLOG, from now on we may simply assume that all factor groups GiG_{i} of GG are almost simple groups.

Proposition 2.24 (Dichotomy via ultrafilter).

There is an ultrafilter ω\omega on II, such that ϕ(GJ)=B\phi(G_{J})=B if and only if JωJ\in\omega, and ϕ(GJ)\phi(G_{J}) is in the center Z(B)\mathrm{Z}(B) of BB if and only if JωJ\notin\omega.

Proof.

We define ω\omega to be the collection of all subsets JJ such that ϕ(GJ)=B\phi(G_{J})=B. Then it is enough to show that this is an ultrafilter.

We already see that for any JJ, exactly one of JJ and IJI-J is in ω\omega.

If J1J2J_{1}\subset J_{2} and J1ωJ_{1}\in\omega, then GJ1GJ2G_{J_{1}}\subseteq G_{J_{2}}. Hence B=ϕ(GJ1)ϕ(GJ2)B=\phi(G_{J_{1}})\subseteq\phi(G_{J_{2}}). So J2ωJ_{2}\in\omega as desired.

If J1,J2ωJ_{1},J_{2}\in\omega, then ϕ(GJ1)=ϕ(GJ2)=B\phi(G_{J_{1}})=\phi(G_{J_{2}})=B. Then, as a result, ϕ(GIJ1),ϕ(GIJ2)\phi(G_{I-J_{1}}),\phi(G_{I-J_{2}}) are both in the center of BB. Then

ϕ(GI(J1J2))=ϕ(G(IJ1)(IJ2))=ϕ(GIJ1GIJ2)=ϕ(GIJ1)ϕ(GIJ2).\phi(G_{I-(J_{1}\cap J_{2})})=\phi(G_{(I-J_{1})\cup(I-J_{2})})=\phi(G_{I-J_{1}}G_{I-J_{2}})=\phi(G_{I-J_{1}})\phi(G_{I-J_{2}}).

So ϕ(GI(J1J2))\phi(G_{I-(J_{1}\cap J_{2})}) is also in the center of BB. So ϕ(GJ1J2)=B\phi(G_{J_{1}\cap J_{2}})=B and J1J2ωJ_{1}\cap J_{2}\in\omega.

So ω\omega is indeed an ultrafilter. ∎

We fix this ultrafilter ω\omega from now on. We now show that, in fact, CR(B)\mathrm{CR}(B) is an ω\omega-ultraproduct of CR(Gi)\mathrm{CR}(G_{i}) for iIi\in I, and the corresponding quotient map is exactly ϕ\phi restricted to CR(G)\mathrm{CR}(G).

Corollary 2.25 (CR(B)\mathrm{CR}(B) is simple).

ϕ(CR(G))=CR(B)\phi(\mathrm{CR}(G))=\mathrm{CR}(B), and this quotient map yields CR(B)\mathrm{CR}(B) as an ω\omega-ultraproduct of CR(Gi)\mathrm{CR}(G_{i}) for iIi\in I. In particular, CR(B)\mathrm{CR}(B) is a non-abelian finite simple group isomorphic to CR(Gi)\mathrm{CR}(G_{i}) for some iIi\in I.

Proof.

Note that CR(G)\mathrm{CR}(G) is a direct product of non-abelian finite simple groups. Hence any finite homomorphic image must also be a direct product of non-abelian finite simple groups. So ϕ(CR(G))\phi(\mathrm{CR}(G)) is inside of CR(B)\mathrm{CR}(B).

In particular, CR(B)\mathrm{CR}(B) cannot be trivial. This is because B/CR(B)B/\mathrm{CR}(B) is a finite quotient of G/CR(G)G/\mathrm{CR}(G), which is solvable, but B𝒞B\notin\mathcal{C}.

Now B/ϕ(CR(G))B/\phi(\mathrm{CR}(G)) is a finite quotient of G/CR(G)G/\mathrm{CR}(G), which is a pro-solvable group. So B/ϕ(CR(G))B/\phi(\mathrm{CR}(G)) is solvable. In particular, CR(B)/ϕ(CR(G))\mathrm{CR}(B)/\phi(\mathrm{CR}(G)) is solvable, and thus it must be trivial. So ϕ(CR(G))=CR(B)\phi(\mathrm{CR}(G))=\mathrm{CR}(B).

Now for any JωJ\in\omega, ϕ(GJ)=B\phi(G_{J})=B, and by similar reasoning, we see that ϕ(CR(GJ))=CR(B)\phi(\mathrm{CR}(G_{J}))=\mathrm{CR}(B). While for JωJ\notin\omega, ϕ(CR(GJ))ϕ(GJ)\phi(\mathrm{CR}(G_{J}))\subset\phi(G_{J}) is in the center of BB. But it must also be a direct product of non-abelian finite simple groups. Hence it is trivial.

In particular, the restriction of ϕ\phi to CR(G)\mathrm{CR}(G) must in fact factor through the ω\omega-ultraproduct. Such an ultraproduct must either have no non-trivial finite quotient, or it is isomorphic to one of the factor groups of CR(G)\mathrm{CR}(G). Since CR(B)\mathrm{CR}(B) is non-trivial, then CR(G)\mathrm{CR}(G) has a non-trivial finite quotient. So in our case, ωCR(Gi)\prod_{\omega}\mathrm{CR}(G_{i}) is isomorphic to the finite simple group CR(Gi)\mathrm{CR}(G_{i}) for some iIi\in I, and CR(B)=ϕ(CR(G))\mathrm{CR}(B)=\phi(\mathrm{CR}(G)) is a non-trivial quotient of this finite simple group. Hence CR(B)\mathrm{CR}(B) is a finite simple group. ∎

Now we establish a description of the center of BB. Note that BB acts on CR(B)\mathrm{CR}(B) by conjugation, and this induces a natural homomorphism from BB to Aut(CR(B))\mathrm{Aut}(\mathrm{CR}(B)). Let NN be the kernel of this map. Then B/NB/N is by definition almost simple.

Proposition 2.26 (Description of pre-images of NN).

For any element (gi)iIG(g_{i})_{i\in I}\in G, we have ϕ((gi)iI)N\phi((g_{i})_{i\in I})\in N if and only if {iI:gi=e}ω\{i\in I:g_{i}=e\}\in\omega.

Proof.

Suppose we have an element (gi)iI(g_{i})_{i\in I} where {iI:gi=e}ω\{i\in I:g_{i}=e\}\in\omega. Then for any hiCR(Gi)h_{i}\in\mathrm{CR}(G_{i}), we have {iI:gi1hi1gihi=e}ω\{i\in I:g_{i}^{-1}h_{i}^{-1}g_{i}h_{i}=e\}\in\omega. Note that, since we picked (hi)iICR(G)(h_{i})_{i\in I}\in\mathrm{CR}(G), the element (gi1hi1gihi)iI(g_{i}^{-1}h_{i}^{-1}g_{i}h_{i})_{i\in I} is also in CR(G)\mathrm{CR}(G), and ϕ\phi restricted to CR(G)\mathrm{CR}(G) is the quotient map onto the ω\omega-ultraproduct. Hence ϕ((gi1hi1gihi)iI)=e\phi((g_{i}^{-1}h_{i}^{-1}g_{i}h_{i})_{i\in I})=e. In particular, ϕ((gi)iI)\phi((g_{i})_{i\in I}) commutes with ϕ((hi)iI)\phi((h_{i})_{i\in I}). Since ϕ(CR(G))=CR(B)\phi(\mathrm{CR}(G))=\mathrm{CR}(B), ϕ((hi)iI)\phi((h_{i})_{i\in I}) could be any element of CR(B)\mathrm{CR}(B). Therefore ϕ((gi)iI)\phi((g_{i})_{i\in I}) commutes with all elements of CR(B)\mathrm{CR}(B), and it must be in NN. So we have proven necessity.

Now we pick any element of NN, say ϕ((gi)iI)N\phi((g_{i})_{i\in I})\in N. Note that each CR(Gi)\mathrm{CR}(G_{i}) is generated by at most two elements, say hih_{i} and hih^{\prime}_{i}. Consider the elements (hi)iI(h_{i})_{i\in I} and (hi)iI(h^{\prime}_{i})_{i\in I}. Their images through ϕ\phi are in ϕ(CR(G))=CR(B)\phi(\mathrm{CR}(G))=\mathrm{CR}(B), so we know ϕ((gi)iI)\phi((g_{i})_{i\in I}) must commute with ϕ((hi)iI),ϕ((hi)iI)\phi((h_{i})_{i\in I}),\phi((h_{i})_{i\in I}). In particular, we have ϕ((gi1hi1gihi)iI)=ϕ((gi1hi1gihi)iI)=e\phi((g_{i}^{-1}h_{i}^{-1}g_{i}h_{i})_{i\in I})=\phi((g_{i}^{-1}{h^{\prime}_{i}}^{-1}g_{i}h^{\prime}_{i})_{i\in I})=e. However, the element (gi1hi1gihi)iI(g_{i}^{-1}h_{i}^{-1}g_{i}h_{i})_{i\in I} is in CR(G)\mathrm{CR}(G), and ϕ\phi restricted to this subgroup is the quotient map onto the ω\omega-ultraproduct. Therefore ϕ((gi1hi1gihi)iI)=e\phi((g_{i}^{-1}h_{i}^{-1}g_{i}h_{i})_{i\in I})=e implies that {iI:gi1hi1gihi=e}ω\{i\in I:g_{i}^{-1}h_{i}^{-1}g_{i}h_{i}=e\}\in\omega. Similarly, {iI:gi1hi1gihi=e}ω\{i\in I:g_{i}^{-1}h_{i}^{-1}g_{i}h_{i}=e\}\in\omega as well. Taking their intersection, we see that {iI:gi commutes with hi,hi}ω\{i\in I:\text{$g_{i}$ commutes with $h_{i},h^{\prime}_{i}$}\}\in\omega.

But if gig_{i} commutes with hi,hih_{i},h^{\prime}_{i}, this means gig_{i} commutes with all elements of CR(Gi)\mathrm{CR}(G_{i}). But since GiG_{i} is an almost simple group, this imples that gi=eg_{i}=e. Hence {iI:gi=e}ω\{i\in I:g_{i}=e\}\in\omega. ∎

Corollary 2.27.

NN is the center of BB.

Proof.

Let ZZ be the center of BB. Since B/NB/N is almost simple, which is centerless, we must have ZNZ\subseteq N.

Now pick any element of NN, say ϕ((gi)iI)N\phi((g_{i})_{i\in I})\in N. Then {iI:gi=e}ω\{i\in I:g_{i}=e\}\in\omega. Let J={iI:gi=e}J=\{i\in I:g_{i}=e\}, then (gi)iIGIJ(g_{i})_{i\in I}\in G_{I-J}, where IJωI-J\notin\omega. Hence ϕ((gi)iI)ϕ(GIJ)\phi((g_{i})_{i\in I})\in\phi(G_{I-J}) is in the center of BB. So NZN\subseteq Z. ∎

Now we proceed to show that NN is in fact a direct factor of BB, which would result in a contradiction to the minimality of BB, finishing our proof of Proposition 2.22.

Proposition 2.28.

The derived subgroup BB^{\prime} has a trivial intersection with NN.

Proof.

Pick any element of BB^{\prime}, say ϕ(([ai,bi])iI)\phi(([a_{i},b_{i}])_{i\in I}). Suppose this is also inside NN, then {iI:[ai,bi]=e}ω\{i\in I:[a_{i},b_{i}]=e\}\in\omega.

Let J={iI:[ai,bi]=e}J=\{i\in I:[a_{i},b_{i}]=e\}. We define ai=aia^{\prime}_{i}=a_{i} if iJi\notin J and ai=ea^{\prime}_{i}=e if iJi\in J. Similarly, we define bi=bib^{\prime}_{i}=b_{i} if iJi\notin J and bi=eb^{\prime}_{i}=e if iJi\in J. Then [ai,bi]=[ai,bi][a^{\prime}_{i},b^{\prime}_{i}]=[a_{i},b_{i}] always. So ϕ(([ai,bi])iI)=ϕ(([ai,bi])iI)\phi(([a_{i},b_{i}])_{i\in I})=\phi(([a^{\prime}_{i},b^{\prime}_{i}])_{i\in I}).

However, we clearly have (ai)iI,(bi)iIGIJ(a^{\prime}_{i})_{i\in I},(b^{\prime}_{i})_{i\in I}\in G_{I-J}. So ϕ((ai)iI),ϕ((bi)iI)\phi((a^{\prime}_{i})_{i\in I}),\phi((b^{\prime}_{i})_{i\in I}) are in the center NN. So we have

ϕ(([ai,bi])iI)=[ϕ((ai)iI),ϕ((bi)iI)]=e.\phi(([a^{\prime}_{i},b^{\prime}_{i}])_{i\in I})=[\phi((a^{\prime}_{i})_{i\in I}),\phi((b^{\prime}_{i})_{i\in I})]=e.

Hence BB^{\prime} has a trivial intersection with NN. ∎

For the moment, we use the notation Bk:={gkgB}B^{k}:=\{g^{k}\mid g\in B\} and Nk:={gkgN}N^{k}:=\{g^{k}\mid g\in N\}.

Proposition 2.29.

For any integer kk, NBBkBNkBNB^{\prime}\cap B^{k}B^{\prime}\subseteq N^{k}B^{\prime}.

Proof.

First let us show that NBBkNkBNB^{\prime}\cap B^{k}\subseteq N^{k}B^{\prime}. Suppose ϕ((xi)iIk)[ϕ((ai)iI),ϕ((bi)iI)]NBBk\phi((x_{i})_{i\in I}^{k})[\phi((a_{i})_{i\in I}),\phi((b_{i})_{i\in I})]\in NB^{\prime}\cap B^{k}, say

ϕ((xi)iIk)=[ϕ((ai)iI),ϕ((bi)iI)]ϕ((yi)iI).\phi((x_{i})_{i\in I}^{k})=[\phi((a_{i})_{i\in I}),\phi((b_{i})_{i\in I})]\phi((y_{i})_{i\in I}).

Here yi=xi[ai,bi]1y_{i}=x_{i}[a_{i},b_{i}]^{-1} and ϕ((yi)iI)N\phi((y_{i})_{i\in I})\in N. Then since ϕ((yi)iI)N\phi((y_{i})_{i\in I})\in N, the index set J={iI:yi=e}J=\{i\in I:y_{i}=e\} is an element of ω\omega.

Define ai,bi,xia^{\prime}_{i},b^{\prime}_{i},x^{\prime}_{i} to be ai,bi,xia_{i},b_{i},x_{i} respectively if iJi\notin J, and to be ee if iJi\in J. Then we have

ϕ((xi)iIk)=[ϕ((ai)iI),ϕ((bi)iI)]ϕ((yi)iI).\phi((x^{\prime}_{i})_{i\in I}^{k})=[\phi((a^{\prime}_{i})_{i\in I}),\phi((b^{\prime}_{i})_{i\in I})]\phi((y_{i})_{i\in I}).

But ϕ((ai)iI),ϕ((bi)iI)N\phi((a^{\prime}_{i})_{i\in I}),\phi((b^{\prime}_{i})_{i\in I})\in N since ω\omega-almost-all coordinates of (ai)iI(a^{\prime}_{i})_{i\in I} and (bi)iI(b^{\prime}_{i})_{i\in I} are trivial. Hence their commutator is also trivial. So ϕ((yi)iI)=ϕ((xi)iIk)\phi((y_{i})_{i\in I})=\phi((x^{\prime}_{i})_{i\in I}^{k}). We also have ϕ((xi)iI)N\phi((x^{\prime}_{i})_{i\in I})\in N since ω\omega-almost-all coordinates of (xi)iI(x^{\prime}_{i})_{i\in I} are trivial. Therefore ϕ((yi)iI)Nk\phi((y_{i})_{i\in I})\in N^{k}.

So we have

ϕ((xi)iIk)=[ϕ((ai)iI),ϕ((bi)iI)]ϕ((yi)iI)NkB.\phi((x_{i})_{i\in I}^{k})=[\phi((a_{i})_{i\in I}),\phi((b_{i})_{i\in I})]\phi((y_{i})_{i\in I})\in N^{k}B^{\prime}.

Now, suppose ghNBBkBgh\in NB^{\prime}\cap B^{k}B^{\prime} where gBkg\in B^{k} and hBh\in B^{\prime}. Then gNBh1=NBg\in NB^{\prime}h^{-1}=NB^{\prime}. So we have

gNBBkNkB.g\in NB^{\prime}\cap B^{k}\subseteq N^{k}B^{\prime}.

Hence ghNkBB=NkBgh\in N^{k}B^{\prime}B^{\prime}=N^{k}B^{\prime}. ∎

Corollary 2.30.

BB cannot exist.

Proof.

We claim that NN is a direct factor of BB. Consider NB/BNB^{\prime}/B^{\prime} in the abelian group B/BB/B^{\prime}. Then for any integer kk,

(NB/B)(B/B)k(NBBkB)/BNkB/B=(NB/B)k.(NB^{\prime}/B^{\prime})\cap(B/B^{\prime})^{k}\subseteq(NB^{\prime}\cap B^{k}B^{\prime})/B^{\prime}\subseteq N^{k}B^{\prime}/B^{\prime}=(NB^{\prime}/B^{\prime})^{k}.

By the Kulikov criteria [7], this means that NB/BNB^{\prime}/B^{\prime} is a direct summand of the abelian group B/BB/B^{\prime}. Let its complement subgroup be H/BH/B^{\prime} for some subgroup HH of BB containing BB^{\prime}. So NBH=BNB^{\prime}\cap H=B^{\prime} and (NB)H=B(NB^{\prime})H=B.

Then

NHNNBH=NB.N\cap H\subseteq N\cap NB^{\prime}\cap H=N\cap B^{\prime}.

So NHN\cap H is trivial. But we also have B=(NB)H=NHB=(NB^{\prime})H=NH. Hence BB is the internal direct product of NN and HH.

But now NN is abelian, and HB/NH\simeq B/N is almost simple. So B𝒞B\in\mathcal{C}, a contradiction. ∎

2.5 Almost semisimple groups

When studying a class of groups, it is sometimes useful if the class of groups is in fact a formation of groups.

Definition 2.31.

A class of groups is a formation if it is closed under quotients and subdirect products.

Note that the class of finite products of finite almost simple groups and finite solvable groups is not closed under subdirect product. Hence it is not a formation of groups.

Example 2.32.

Think of the symmetric group S5\mathrm{S}_{5} as the semi-direct product A5/2\mathrm{A}_{5}\rtimes\mathbb{Z}/2\mathbb{Z}, and the symmetric group S6\mathrm{S}_{6} as the semi-direct product A6/2\mathrm{A}_{6}\rtimes\mathbb{Z}/2\mathbb{Z}. Now, let /2\mathbb{Z}/2\mathbb{Z} act on both A5\mathrm{A}_{5} and A6\mathrm{A}_{6} simultaneously, then we have a group (A5×A6)/2(\mathrm{A}_{5}\times\mathrm{A}_{6})\rtimes\mathbb{Z}/2\mathbb{Z}, which is a sub-direct product of the almost simple groups S5\mathrm{S}_{5} and S6\mathrm{S}_{6}. However, the resulting group is not a direct product of almost simple groups and solvable groups.

So if we want to study a formation of groups, we need to consider a larger class of groups.

Definition 2.33.

We define a group GG to be almost semisimple if GG has no solvable normal subgroups.

These groups are natural generalizations of almost simple groups.

Proposition 2.34.

If GG is an almost semisimple group, then GG acts on its CR-radical CR(G)\mathrm{CR}(G) faithfully. In particular, we may think of GG as a subgroup of Aut(S)\mathrm{Aut}(S) containing SS for some semisimple group SS.

Proof.

Note that GG acts on its CR-radical CR(G)\mathrm{CR}(G) by conjugation. Let NN be the kernel of the induced homomorphism GAut(CR(G))G\to\mathrm{Aut}(\mathrm{CR}(G)). Suppose for contradiction that it is not trivial. Let MM be a minimal non-trivial normal subgroup of NN. Then MM is either semisimple or abelian.

If MM is semisimple, then MCR(N)M\subseteq\mathrm{CR}(N). Since CR(N)\mathrm{CR}(N) is characteristic in NN, it is also a semisimple normal subgroup of GG, so CR(N)CR(G)\mathrm{CR}(N)\subseteq\mathrm{CR}(G) and CR(N)N\mathrm{CR}(N)\subseteq N. However, since CR(G)\mathrm{CR}(G) is semisimple, therefore it acts on itself faithfully. So the intersection NCR(G)N\cap\mathrm{CR}(G) is trivial, which implies that CR(N)\mathrm{CR}(N) is trivial as well. Then CR(N)\mathrm{CR}(N) cannot contain MM, a contradiction.

Suppose MM is abelian. Then the unique maximal solvable normal subgroup of NN is non-trivial. But this subgroup is also characteristic in NN, hence it is a non-trivial solvable normal subgroup of GG, which is impossible as we require GG to be almost semisimple. Contradiction.

So NN must be trivial. ∎

Now to have a formation of groups, we also want our class to be closed under quotient. Hence we need to throw solvable groups into the mix as well.

Proposition 2.35.

If G/CR(G)G/\mathrm{CR}(G) is solvable, then GG is a subdirect product of a solvable group and an almost semisimple group.

Proof.

Let sol(G)\mathrm{sol}(G) be the solvable radical of GG (i.e., the unique largest solvable normal subgroup of GG). Then G/sol(G)G/\mathrm{sol}(G) is an almost simple group, while G/CR(G)G/\mathrm{CR}(G) is solvable. Finally, the intersection of sol(G)\mathrm{sol}(G) and CR(G)\mathrm{CR}(G) must be both CR and solvable, so it is trivial. So we are done. ∎

Now let 𝒞\mathcal{C} be the class of groups GG such that G/CR(G)G/\mathrm{CR}(G) is solvable. It is clearly closed under taking quotients.

Proposition 2.36.

𝒞\mathcal{C} is closed under subdirect product.

Proof.

Suppose GG has two normal subgroups N1,N2N_{1},N_{2} such that G/N1,G/N2𝒞G/N_{1},G/N_{2}\in\mathcal{C}. We can find normal subgroups M1,M2M_{1},M_{2} of GG that are the pre-images of CR(G/N1),CR(G/N2)\mathrm{CR}(G/N_{1}),\mathrm{CR}(G/N_{2}). Then G/M1,G/M2G/M_{1},G/M_{2} is solvable. As a result, since G/(M1M2)G/(M_{1}\cap M_{2}) is a subdirect product of solvable groups G/M1,G/M2G/M_{1},G/M_{2}, G/(M1M2)G/(M_{1}\cap M_{2}) is also solvable, and its subgroups M1/(M1M2)M_{1}/(M_{1}\cap M_{2}) and M2/(M1M2)M_{2}/(M_{1}\cap M_{2}) are also solvable.

G{G}N1(M1M2){N_{1}(M_{1}\cap M_{2})}M1{M_{1}}M2{M_{2}}N2(M1M2){N_{2}(M_{1}\cap M_{2})}N1{N_{1}}M1M2{M_{1}\cap M_{2}}N2{N_{2}}N1M2{N_{1}\cap M_{2}}N2M1{N_{2}\cap M_{1}}{e}{\{e\}}solvablesolvableCRsolvableCRsolvableCRCR

Let us now show that N1(M1M2)=M1N_{1}(M_{1}\cap M_{2})=M_{1}. To see this, note that M1/(N1(M1M2))M_{1}/(N_{1}(M_{1}\cap M_{2})) is a quotient of the solvable group M1/(M1M2)M_{1}/(M_{1}\cap M_{2}), hence it is solvable. However, it is also the quotient of M1/N1CR(G/N1)M_{1}/N_{1}\simeq\mathrm{CR}(G/N_{1}), therefore it is a CR group. Since it is both solvable and CR, it has to be trivial. So we have N1(M1M2)=M1N_{1}(M_{1}\cap M_{2})=M_{1}, and similarly N2(M1M2)=M2N_{2}(M_{1}\cap M_{2})=M_{2} as well.

So we have

(M1M2)/(N1M2)=(M1M2)/(N1(M1M2))N1(M1M2)/N1=M1/N1CR(G/N1).(M_{1}\cap M_{2})/(N_{1}\cap M_{2})=(M_{1}\cap M_{2})/(N_{1}\cap(M_{1}\cap M_{2}))\simeq N_{1}(M_{1}\cap M_{2})/N_{1}=M_{1}/N_{1}\simeq\mathrm{CR}(G/N_{1}).

So (M1M2)/(N1M2)(M_{1}\cap M_{2})/(N_{1}\cap M_{2}) and similarly (M1M2)/(M1N2)(M_{1}\cap M_{2})/(M_{1}\cap N_{2}) are both CR groups. We also see that (N1M2)(M1N2)N1N2(N_{1}\cap M_{2})\cap(M_{1}\cap N_{2})\subseteq N_{1}\cap N_{2} is trivial. Hence M1M2M_{1}\cap M_{2} is the subdirect product of CR groups (M1M2)/(N1M2)(M_{1}\cap M_{2})/(N_{1}\cap M_{2}) and (M1M2)/(M1N2)(M_{1}\cap M_{2})/(M_{1}\cap N_{2}). So M1M2M_{1}\cap M_{2} must be a CR group itself. In particular, M1M2CR(G)M_{1}\cap M_{2}\subseteq\mathrm{CR}(G).

On the other hand, since G/(M1M2)G/(M_{1}\cap M_{2}) is solvable, CR(G)M1M2\mathrm{CR}(G)\subseteq M_{1}\cap M_{2}. So we see that M1M2=CR(G)M_{1}\cap M_{2}=\mathrm{CR}(G), and G/CR(G)G/\mathrm{CR}(G) is solvable. ∎

Proposition 2.37.

If GG is a pro-𝒞\mathcal{C} group, then any finite homomorphic image of GG must still be in 𝒞\mathcal{C}.

Proof.

Suppose GG is formed via a surjective inverse directed system {Gi}iI\{G_{i}\}_{i\in I}. Let SiS_{i} be the CR-radical of GiG_{i}.

For any surjective homomorphism ϕij:GiGj\phi_{ij}:G_{i}\to G_{j}, then SiS_{i} must be mapped into SjS_{j}. Hence this induces homomorphisms SiSjS_{i}\to S_{j} and Gi/SiGj/SjG_{i}/S_{i}\to G_{j}/S_{j}. So the inverse directed system {Gi}iI\{G_{i}\}_{i\in I} would induce the corresponding inverse directed systems {Si}iI\{S_{i}\}_{i\in I} and {Gi/Si}iI\{G_{i}/S_{i}\}_{i\in I}.

Let SS be the projective limit of the system {Si}iI\{S_{i}\}_{i\in I}, then SS is a semisimple profinite group and a closed normal subgroup of GG. Note that the projective limit on inverse systems of finite groups is an exact functor. (See, e.g., Proposition 2.2.4 of [12].) As a result, G/SG/S is the projective limit of the system {Gi/Si}iI\{G_{i}/S_{i}\}_{i\in I}, and therefore it is pro-solvable.

Now suppose BB is a finite quotient of GG. Let qq be the quotient map from GG to BB. Then since SS is a semisimple profinite group, q(S)q(S) must be a semisimple normal subgroup of BB. Hence q(S)CR(B)q(S)\leq\mathrm{CR}(B). So SS is in the kernel of the quotient map from GG to B/CR(B)B/\mathrm{CR}(B). This induces a surjective homomorphism from G/SG/S to B/CR(B)B/\mathrm{CR}(B). Since G/SG/S is pro-solvable, we see that B/CR(B)B/\mathrm{CR}(B) is solvable. Hence BB is in 𝒞\mathcal{C}. ∎

2.6 Perfect groups with bounded commutator width

We devote a small section to perfect groups with bounded commutator width.

Proposition 2.38.

Let PnP_{n} be a surjective inverse direct system of perfect groups, and suppose they all have commutator width at most dd. Let GG be their projective limit. Then GG is also perfect with commutator width dd.

Proof.

Since the class of finite perfect groups with commutator width at most dd is quotient closed, therefore any continuous finite quotient of GG must be perfect with commutator width at most dd. Let XX be the set of commutators in GG, then XdX^{\star d} is mapped surjectively onto all continuous finite quotients. In particular, XdX^{\star d} is dense in GG.

However, consider the map w:G2dGw:G^{2d}\to G sending (g1,h1,,gd,hd)(g_{1},h_{1},\dots,g_{d},h_{d}) to the element [g1,h1][gd,hd][g_{1},h_{1}]\dots[g_{d},h_{d}]. Since ww must be continuous, and G2dG^{2d} is compact, therefore its image Xd=w(G2d)X^{\star d}=w(G^{2d}) is compact. Since GG is Hausdorff, XdX^{\star d} is a dense closed subset of GG. Hence Xd=GX^{\star d}=G.

So GG is perfect with commutator length at most dd. ∎

3 Perfect Groups with Unbounded Commutator Width

In this section, we shall prove Theorem 1.6 according to the layout in Section 1.2.

3.1 The restricted Burnside coproduct

For any positive integer mm, consider the category 𝒞\mathcal{C} of finite groups with exponent dividing mm. Then Theorem 1.7 showed that B(d,m)B(d,m) is the free object on dd generators in this category.

Proposition 3.1.

The category 𝒞\mathcal{C} is a reflexive subcategory of the category of finitely generated groups (i.e., the inclusion functor has a left adjoint).

Proof.

Let us create this reflexive functor to the inclusion functor. Given any finitely generated group GG, then it is the quotient of the free group Fd\mathrm{F}_{d} for some positive integer dd. Let the kernel of this quotient map q:FdGq:\mathrm{F}_{d}\to G be NGN_{G}.

According to Theorem 1.7, the free group Fd\mathrm{F}_{d} has a normal subgroup NBN_{B} such that any group homomorphism from Fd\mathrm{F}_{d} to a group in 𝒞\mathcal{C} must factor through Fd/NB\mathrm{F}_{d}/N_{B}. Let NmN_{m} be the normal subgroup generated by NGN_{G} and NBN_{B}, and we define Gm=F/NmG_{m}=F/N_{m}. Since GmG_{m} is a quotient of the restricted Burnside group B(d,m)B(d,m), therefore Gm𝒞G_{m}\in\mathcal{C} as desired.

We claim that for any H𝒞H\in\mathcal{C}, any homomorphism f:GHf:G\to H, then there is a unique induced homomorphism fm:GmHf_{m}:G_{m}\to H. This would show that the functor GGmG\mapsto G_{m} is indeed a left adjoint to the inclusion functor.

To see the claim, note that fq:FdHf\circ q:\mathrm{F}_{d}\to H must have both NGN_{G} and NBN_{B} in its kernel. So NmN_{m} is in this kernel. Hence this map factors through a unique homomorphism fm:GmHf_{m}:G_{m}\to H. ∎

Corollary 3.2.

In the category 𝒞\mathcal{C}, finite product and finite coproduct exist. And the coproduct of B(d1,m)B(d_{1},m) and B(d2,m)B(d_{2},m) is isomorphic to B(d1+d2,m)B(d_{1}+d_{2},m).

Proof.

Since the category of finitely generated groups has coproducts, therefore any reflexive subcategory has coproducts. The other statements are trivial. ∎

We define a functor d:𝒞𝒞\mathcal{B}_{d}:\mathcal{C}\to\mathcal{C} that maps each group GG in 𝒞\mathcal{C} to the coproduct of GG with itself dd-times, and maps each morphism ϕ:GH\phi:G\to H to the corresponding morphism on the dd-times iterated coproducts.

Proposition 3.3.

d\mathcal{B}_{d} preserves all colimits.

Proof.

Note that the coproduct functor :𝒞d𝒞\mathcal{F}:\mathcal{C}^{d}\to\mathcal{C} is left adjoint to the diagonal functor 𝒢:𝒞𝒞d\mathcal{G}:\mathcal{C}\to\mathcal{C}^{d}, while the diagonal functor 𝒢\mathcal{G} is left adjoint to the product functor. Therefore the functor d=𝒢\mathcal{B}_{d}=\mathcal{F}\circ\mathcal{G} is left adjoint and preserves colimits. ∎

Corollary 3.4.

d(ϕ)\mathcal{B}_{d}(\phi) is surjective if ϕ\phi is surjective.

Lemma 3.5.

Let ι1,,ιd:Gd(G)\iota_{1},\dots,\iota_{d}:G\to\mathcal{B}_{d}(G) be the coprojection morphisms. Then they are injective. Furthermore, ι1(G),,ιd(G)\iota_{1}(G),\dots,\iota_{d}(G) are independent subgroups generating d(G)\mathcal{B}_{d}(G).

Proof.

Since the category 𝒞\mathcal{C} has zero morphisms, therefore all coprojections are split monomorphisms, and the images ι1(G),,ιd(G)\iota_{1}(G),\dots,\iota_{d}(G) are independent subgroups.

Since the restricted Burnside coproduct d(G)\mathcal{B}_{d}(G) is a quotient of the free product of dd copies of GG, therefore the subgroups ι1(G),,ιd(G)\iota_{1}(G),\dots,\iota_{d}(G) generates d(G)\mathcal{B}_{d}(G). ∎

We now examine the interaction of d\mathcal{B}_{d} with group actions. In the following, MM is a finite group with exponents dividing mm, and M1,,MdM_{1},\dots,M_{d} are the images of the coprojection morphisms from MM into d(M)\mathcal{B}_{d}(M). We use e1,,ede_{1},\dots,e_{d} to denote the generators of B(d,m)=d(/m)B(d,m)=\mathcal{B}_{d}(\mathbb{Z}/m\mathbb{Z}).

Proposition 3.6.

Suppose GG is a group acting on a finite group MM, and the exponent of MM divides mm. Then GG has an induced action on d(M)\mathcal{B}_{d}(M).

If the coprojection images of MM in d(M)\mathcal{B}_{d}(M) are M1,,MdM_{1},\dots,M_{d}, then these subgroups are GG-invariant, and the GG-action on d(M)\mathcal{B}_{d}(M) restricted to each MiM_{i} is identical to the GG-action on MM.

Proof.

Suppose we have ϕ:GAut(M)\phi:G\to\mathrm{Aut}(M). As d\mathcal{B}_{d} is functorial and MM is a finite group with exponent dividing mm, we have a map d:Aut(M)Aut(d(M))\mathcal{B}_{d}:\mathrm{Aut}(M)\to\mathrm{Aut}(\mathcal{B}_{d}(M)). So dϕ\mathcal{B}_{d}\circ\phi is the induced GG-action on d(M)\mathcal{B}_{d}(M).

Let ι1,,ιd:Md(M)\iota_{1},\dots,\iota_{d}:M\to\mathcal{B}_{d}(M) be the coprojection morphisms so that Mi=ιi(M)M_{i}=\iota_{i}(M). For any ϕAut(M)\phi\in\mathrm{Aut}(M), then ιiϕ=d(ϕ)ιi\iota_{i}\circ\phi=\mathcal{B}_{d}(\phi)\circ\iota_{i} by definition of the coproduct. So

d(ϕ)(Mi)=d(ϕ)ιi(M)=ιiϕ(M)=ιi(M)=Mi.\mathcal{B}_{d}(\phi)(M_{i})=\mathcal{B}_{d}(\phi)\circ\iota_{i}(M)=\iota_{i}\circ\phi(M)=\iota_{i}(M)=M_{i}.

So MiM_{i} is d(ϕ)\mathcal{B}_{d}(\phi)-invariant, and the d(ϕ)\mathcal{B}_{d}(\phi)-action on d(M)\mathcal{B}_{d}(M) restricted to MiM_{i} is identical to the ϕ\phi-action on MM. ∎

Proposition 3.7.

Suppose GG is a group acting on a finite group MM, and the exponent of MM divides mm, and MGM\rtimes G is perfect. Then for the induced action of GG on d(M)\mathcal{B}_{d}(M), and d(M)G\mathcal{B}_{d}(M)\rtimes G is perfect.

Proof.

Now d(M)\mathcal{B}_{d}(M) is generated by GG-invariant subgroups M1,,MdM_{1},\dots,M_{d} as in Proposition 3.6. However, since GG acts on MiM_{i} the same way as its action on MM, therefore MiGM_{i}\rtimes G as a subgroup of d(M)G\mathcal{B}_{d}(M)\rtimes G is isomorphic to MGM\rtimes G, which is perfect. As these perfect subgroups MiGM_{i}\rtimes G generate the whole group, d(M)G\mathcal{B}_{d}(M)\rtimes G is perfect. ∎

Proposition 3.8.

Suppose GG is a group acting on a finite group MM, and the exponent of MM divides mm. Then for any GG-invariant homomorphism ϕ:M/m\phi:M\to\mathbb{Z}/m\mathbb{Z}, the induced homomorphism d(ϕ)\mathcal{B}_{d}(\phi) is GG-invariant.

Proof.

Since ϕ\phi is GG-invariant, for any xMx\in M and gGg\in G, we have ϕ(x)=ϕ(xg)\phi(x)=\phi(x^{g}). Let ι1,,ιd:Md(M)\iota_{1},\dots,\iota_{d}:M\to\mathcal{B}_{d}(M) be the coprojection morphisms so that Mi=ιi(M)M_{i}=\iota_{i}(M). Then

d(ϕ)ιi(xg)=ιi(ϕ(xg))=ιi(ϕ(x))=d(ϕ)ιi(x).\mathcal{B}_{d}(\phi)\circ\iota_{i}(x^{g})=\iota_{i}(\phi(x^{g}))=\iota_{i}(\phi(x))=\mathcal{B}_{d}(\phi)\circ\iota_{i}(x).

So d(ϕ)\mathcal{B}_{d}(\phi) restricted to each MiM_{i} is GG-invariant. But since these MiM_{i} generates the domain d(M)\mathcal{B}_{d}(M), therefore we see that d(ϕ)\mathcal{B}_{d}(\phi) is GG-invariant. ∎

3.2 The ultraproduct functor

The ultraproduct of groups can be thought of as a functor from Grp\mathrm{Grp}^{\mathbb{N}} to Grp\mathrm{Grp}. It has the following categorical properties:

  1. 1.

    The ultraproduct functor preserves all finite limits. (In particular, the ultraproduct of subgroups is a subgroup of the ultraproduct, and the ultraproduct of intersections is the intersection of ultraproducts.)

  2. 2.

    The ultraproduct functor preserves short exact sequences. (In particular, it preserves normal subgroups, monomorphism, epimorphism, and quotients.)

  3. 3.

    The ultraproduct functor preserves semi-direct products.

  4. 4.

    The ultraproduct of abelian groups is an abelian group.

  5. 5.

    The ultraproduct of groups of exponent dividing mm is a group of exponent dividing mm.

  6. 6.

    The ultraproduct of a sequence of the same finite group is the finite group itself.

Given any sequence of finite groups MnM_{n} of exponents dividing a fixed integer mm, we want to understand homomorphisms from nωMn\prod_{n\to\omega}M_{n} to /m\mathbb{Z}/m\mathbb{Z}. Note that many of such homomorphisms are elements of nωMn\prod_{n\to\omega}M_{n}^{*}, as shown in the example below.

Example 3.9.

One way to construct such a homomorphism is to first pick a homomorphism ϕn:Mn/m\phi_{n}:M_{n}\to\mathbb{Z}/m\mathbb{Z} for each nn, and take the ultralimit limnωϕn\lim_{n\to\omega}\phi_{n}, which is a homomorphism from nωMn\prod_{n\to\omega}M_{n} to nω/m=/m\prod_{n\to\omega}\mathbb{Z}/m\mathbb{Z}=\mathbb{Z}/m\mathbb{Z}.

If we use MnM_{n}^{*} to denote the abelian group of homomorphisms from MnM_{n} to /m\mathbb{Z}/m\mathbb{Z} (i.e., the “dual group”), then these limnωϕn\lim_{n\to\omega}\phi_{n} correspond to elements of the ultraproduct nωMn\prod_{n\to\omega}M_{n}^{*}.

Unfortunately, not all homomorphisms from nωMn\prod_{n\to\omega}M_{n} to /m\mathbb{Z}/m\mathbb{Z} are elements of nωMn\prod_{n\to\omega}M_{n}^{*}. Here is another kind of construction: we take all homomorphisms from nωMn\prod_{n\to\omega}M_{n} to /m\mathbb{Z}/m\mathbb{Z} in nωMn\prod_{n\to\omega}M_{n}^{*} as above, and take their “consensus” according to some ultrafilter 𝒰\mathcal{U} on nωMn\prod_{n\to\omega}M_{n}^{*}.

Definition 3.10.

Let 𝒰\mathcal{U} be any ultrafilter on the set nωMn\prod_{n\to\omega}M_{n}^{*}, then we define the 𝒰\mathcal{U}-consensus to be the homomorphism ϕ𝒰=limϕ𝒰ϕ\phi_{\mathcal{U}}=\lim_{\phi\to\mathcal{U}}\phi such that ϕ𝒰(x)=limϕ𝒰ϕ(x)\phi_{\mathcal{U}}(x)=\lim_{\phi\to\mathcal{U}}\phi(x) for all xnωMnx\in\prod_{n\to\omega}M_{n},

Then it is straight forward to verify that ϕ𝒰:nωMn/m\phi_{\mathcal{U}}:\prod_{n\to\omega}M_{n}\to\mathbb{Z}/m\mathbb{Z} is indeed a homomorphism.

It turns out that, when all groups MnM_{n} are abelian, then any homomorphism ϕ:nωMn/m\phi:\prod_{n\to\omega}M_{n}\to\mathbb{Z}/m\mathbb{Z} is equal to ϕ𝒰\phi_{\mathcal{U}} for some ultrafilter 𝒰\mathcal{U} on nωMn\prod_{n\to\omega}M_{n}^{*}, as in the above example.

Proposition 3.11.

Fix a sequence of finite abelian groups MnM_{n} of exponents dividing a fixed integer mm. Then any homomorphism ϕ:nωMn/m\phi:\prod_{n\to\omega}M_{n}\to\mathbb{Z}/m\mathbb{Z} is equal to ϕ𝒰\phi_{\mathcal{U}} for some ultrafilter 𝒰\mathcal{U} on nωMn\prod_{n\to\omega}M_{n}^{*}.

Proof.

This is a special case of Proposition 3.12 below, when the group actions are all trivial. ∎

In fact, we can prove something stronger. Proposition 3.12 below shows that we can in fact do this even with respect to a given group action. To do this, we need a definition of GG-invariant ultrafilter on nωMn\prod_{n\to\omega}M_{n}^{*}.

Suppose we have a finite group GnG_{n} acting on MnM_{n} for each nn. Then there is an induced action of G=nωGnG=\prod_{n\to\omega}G_{n} on nωMn\prod_{n\to\omega}M_{n}.

For each gGg\in G, let SgnωMnS_{g}\subseteq\prod_{n\to\omega}M_{n}^{*} be the collection of all gg-invariant elements of nωMn\prod_{n\to\omega}M_{n}^{*}. Then we say the ultrafilter 𝒰\mathcal{U} is GG-invariant if and only if Sg𝒰S_{g}\in\mathcal{U} for all gGg\in G. Then we have the following result.

Proposition 3.12.

Fix a sequence of finite abelian groups MnM_{n} of exponents dividing a fixed integer mm. Suppose for each nn, we have a finite group GnG_{n} acting on MnM_{n}. So we have an action of G=nωGnG=\prod_{n\to\omega}G_{n} on nωMn\prod_{n\to\omega}M_{n}.

Then any GG-invariant homomorphism ϕ:nωMn/m\phi:\prod_{n\to\omega}M_{n}\to\mathbb{Z}/m\mathbb{Z} is equal to ϕ𝒰\phi_{\mathcal{U}} for some GG-invariant ultrafilter 𝒰\mathcal{U} on nωMn\prod_{n\to\omega}M_{n}^{*}.

Proof.

For each gGg\in G, let SgnωMnS_{g}\subseteq\prod_{n\to\omega}M_{n}^{*} be the collection of all gg-invariant elements of nωMn\prod_{n\to\omega}M_{n}^{*}. For each xnωMnx\in\prod_{n\to\omega}M_{n}, let SxnωMnS_{x}\subseteq\prod_{n\to\omega}M_{n}^{*} be the collection of all ψnωMn\psi\in\prod_{n\to\omega}M_{n}^{*} such that ψ(x)=ϕ(x)\psi(x)=\phi(x).

Then 𝒰\mathcal{U} is GG-invariant if and only if it contains all SgS_{g} for all gGg\in G, and we have ϕ=ϕ𝒰\phi=\phi_{\mathcal{U}} if and only if 𝒰\mathcal{U} contains SxS_{x} for all xnωMnx\in\prod_{n\to\omega}M_{n}. So our proposition is the same as the claim that there is an ultrafilter on nωMn\prod_{n\to\omega}M_{n}^{*} containing all SgS_{g} and all SxS_{x}.

To prove this claim, we need to show that the collection of all SgS_{g} and all SxS_{x} has finite intersection property. This is done by Lemma 3.14. ∎

First, let us show that the collection of all SxS_{x} has the finite intersection property.

Lemma 3.13.

Let MnM_{n} be a sequence of finite abelian groups with exponents dividing mm. For any x1,,xknωMnx_{1},\dots,x_{k}\in\prod_{n\to\omega}M_{n} and any homomorphism ϕ:nωMn/m\phi:\prod_{n\to\omega}M_{n}\to\mathbb{Z}/m\mathbb{Z}, we can find a homomorphism ψnωMn\psi\in\prod_{n\to\omega}M_{n}^{*} such that ψ(xi)=ϕ(xi)\psi(x_{i})=\phi(x_{i}) for all ii.

Proof.

Note that finite abelian groups with exponents dividing mm are the same as /m\mathbb{Z}/m\mathbb{Z}-modules. Picking elements x1,,xknωMnx_{1},\dots,x_{k}\in\prod_{n\to\omega}M_{n} is the same as picking a /m\mathbb{Z}/m\mathbb{Z}-modules homomorphism f:(/m)knωMnf:(\mathbb{Z}/m\mathbb{Z})^{k}\to\prod_{n\to\omega}M_{n}. Say we let e1,,eke_{1},\dots,e_{k} be the standard generators of /m\mathbb{Z}/m\mathbb{Z}, so that we have f(ei)=xif(e_{i})=x_{i}.

Let xi=limnωxi,nx_{i}=\lim_{n\to\omega}x_{i,n}, then we can define /m\mathbb{Z}/m\mathbb{Z}-modules homomorphism fn:(/m)kMnf_{n}:(\mathbb{Z}/m\mathbb{Z})^{k}\to M_{n} such that fn(ei)=xi,nf_{n}(e_{i})=x_{i,n}. Then f=limnωfnf=\lim_{n\to\omega}f_{n} in the sense that f(v)=limnωfn(v)f(v)=\lim_{n\to\omega}f_{n}(v) for all v(/m)kv\in(\mathbb{Z}/m\mathbb{Z})^{k}.

Let N=Ker(f)N=\mathrm{Ker}(f). Since f(v)=0f(v)=0 if and only if vNv\in N, the sets {nfn(v)=0}\{n\in\mathbb{N}\mid f_{n}(v)=0\} for vNv\in N are inside the ultrafilter ω\omega, and the sets {nfn(v)0}\{n\in\mathbb{N}\mid f_{n}(v)\neq 0\} for v(/m)kNv\in(\mathbb{Z}/m\mathbb{Z})^{k}-N are also inside the ultrafilter ω\omega. Since (/m)k(\mathbb{Z}/m\mathbb{Z})^{k} is a finite set, we have the following result.

{nKer(fn)=N}=(vN{nfn(v)=0})(v(/m)kN{nfn(v)0})ω.\{n\in\mathbb{N}\mid\mathrm{Ker}(f_{n})=N\}=(\bigcap_{v\in N}\{n\in\mathbb{N}\mid f_{n}(v)=0\})\cap(\bigcap_{v\in(\mathbb{Z}/m\mathbb{Z})^{k}-N}\{n\in\mathbb{N}\mid f_{n}(v)\neq 0\})\in\omega.

So Ker(fn)=N\mathrm{Ker}(f_{n})=N for ω\omega-almost all nn.

(/m)k{(\mathbb{Z}/m\mathbb{Z})^{k}}(/m)k/N{(\mathbb{Z}/m\mathbb{Z})^{k}/N}Mn{M_{n}}/m{\mathbb{Z}/m\mathbb{Z}}q\scriptstyle{q}ϕf\scriptstyle{\phi\circ f}fn\scriptstyle{f_{n}}fn\scriptstyle{f^{\prime}_{n}}f\scriptstyle{f^{\prime}}ψn\scriptstyle{\psi_{n}}

Now the homomorphism ϕf:(/m)k/m\phi\circ f:(\mathbb{Z}/m\mathbb{Z})^{k}\to\mathbb{Z}/m\mathbb{Z} factors through the quotient map q:(/m)k(/m)k/Nq:(\mathbb{Z}/m\mathbb{Z})^{k}\to(\mathbb{Z}/m\mathbb{Z})^{k}/N, and induces a homomorphism f:(/m)k/N/mf^{\prime}:(\mathbb{Z}/m\mathbb{Z})^{k}/N\to\mathbb{Z}/m\mathbb{Z}. For ω\omega-almost all nn, the map fn:(/m)kMnf_{n}:(\mathbb{Z}/m\mathbb{Z})^{k}\to M_{n} also factors through qq as well, and induces an injective homomorphism fm:(/m)k/NMnf^{\prime}_{m}:(\mathbb{Z}/m\mathbb{Z})^{k}/N\to M_{n}.

Note that by Baer’s criterion (see, e.g., [8]), /m\mathbb{Z}/m\mathbb{Z} is an injective module over itself. Since for ω\omega-almost all nn, fn:(/m)k/NMnf^{\prime}_{n}:(\mathbb{Z}/m\mathbb{Z})^{k}/N\to M_{n} is injective, therefore we have an extension ψn:Mn/m\psi_{n}:M_{n}\to\mathbb{Z}/m\mathbb{Z} such that ψnfn=f\psi_{n}\circ f^{\prime}_{n}=f^{\prime}. Clearly ψnMn\psi_{n}\in M_{n}^{*}. In particular, ψnfn=ϕf\psi_{n}\circ f_{n}=\phi\circ f for ω\omega-almost all nn.

Let ψ=limnωψnnωMn\psi=\lim_{n\to\omega}\psi_{n}\in\prod_{n\to\omega}M_{n}^{*}. Let us verify that ψ\psi is what we need.

ψ(xi)=limnωψn(xi,n)=limnωψn(fn(ei))=limnωϕf(ei)=ϕf(ei)=ϕ(xi).\psi(x_{i})=\lim_{n\to\omega}\psi_{n}(x_{i,n})=\lim_{n\to\omega}\psi_{n}(f_{n}(e_{i}))=\lim_{n\to\omega}\phi\circ f(e_{i})=\phi\circ f(e_{i})=\phi(x_{i}).

So we are done. ∎

Lemma 3.14.

Let MnM_{n} be a sequence of finite abelian groups with exponents dividing mm. Let g1,n,,gk,ng_{1,n},\dots,g_{k,n} be automorphisms of MnM_{n} for each nn. Then gi=limnωgi,ng_{i}=\lim_{n\to\omega}g_{i,n} is an automorphism of nωMn\prod_{n\to\omega}M_{n} for each ii.

For any x1,,xknωMnx_{1},\dots,x_{k}\in\prod_{n\to\omega}M_{n} and any g1,,gkg_{1},\dots,g_{k}-invariant homomorphism ϕ:nωMn/m\phi:\prod_{n\to\omega}M_{n}\to\mathbb{Z}/m\mathbb{Z}, we can find a g1,,gkg_{1},\dots,g_{k}-invariant homomorphism ψnωMn\psi\in\prod_{n\to\omega}M_{n}^{*} such that ψ(xi)=ϕ(xi)\psi(x_{i})=\phi(x_{i}) for all ii.

Proof.

Suppose xi=limnωxi,nx_{i}=\lim_{n\to\omega}x_{i,n}.

For each nn, let NnN_{n} be the subgroup generated by [Mn,g1,n],,[Mn,gk,n][M_{n},g_{1,n}],\dots,[M_{n},g_{k,n}]. For simplicity of notation, let N=nωNnN=\prod_{n\to\omega}N_{n}. Let qn:MnMn/Nnq_{n}:M_{n}\to M_{n}/N_{n} be the quotient map.

Then if ϕ\phi is g1,,gkg_{1},\dots,g_{k}-invariant, then nωNnKer(ϕ)\prod_{n\to\omega}N_{n}\subseteq\mathrm{Ker}(\phi). So we have an induced homomorphism ϕ:nωMn/Nn/m\phi^{\prime}:\prod_{n\to\omega}M_{n}/N_{n}\to\mathbb{Z}/m\mathbb{Z}. By Lemma 3.13, we can find an element ψ=limnωψnnω(Mn/Nn)\psi^{\prime}=\lim_{n\to\omega}\psi^{\prime}_{n}\in\prod_{n\to\omega}(M_{n}/N_{n})^{*} such that for all ii,

ϕ(ximodN)=ψ(ximodN)=ψ(limnωqn(xi,n))=limnω(ψnqn)(xi,n).\phi^{\prime}(x_{i}\bmod N)=\psi^{\prime}(x_{i}\bmod N)=\psi^{\prime}(\lim_{n\to\omega}q_{n}(x_{i,n}))=\lim_{n\to\omega}(\psi^{\prime}_{n}\circ q_{n})(x_{i,n}).

Let ψ=limnωψnqn\psi=\lim_{n\to\omega}\psi^{\prime}_{n}\circ q_{n}, then ψ(xi)=ϕ(ximodN)=ϕ(xi)\psi(x_{i})=\phi^{\prime}(x_{i}\bmod N)=\phi(x_{i}) as desired. Furthermore, each homomorphism ψnqn\psi^{\prime}_{n}\circ q_{n} is invariant under g1,n,,gk,ng_{1,n},\dots,g_{k,n}. Thus ψ\psi is invariant under g1,,gkg_{1},\dots,g_{k}. ∎

Now we want to apply the restricted Burnside coproduct on the homomorphisms ϕnωMn\phi\in\prod_{n\to\omega}M_{n}^{*} and ϕ𝒰\phi_{\mathcal{U}} for ultrafilters 𝒰\mathcal{U} on nωMn\prod_{n\to\omega}M_{n}^{*}. Here are the constructions.

For each element limnωϕnnωMn\lim_{n\to\omega}\phi_{n}\in\prod_{n\to\omega}M_{n}^{*}, note that each ϕn:Mn/m\phi_{n}:M_{n}\to\mathbb{Z}/m\mathbb{Z} induces a homomorphism d(ϕn):d(Mn)d(/m)=B(d,m)\mathcal{B}_{d}(\phi_{n}):\mathcal{B}_{d}(M_{n})\to\mathcal{B}_{d}(\mathbb{Z}/m\mathbb{Z})=B(d,m). As a result, we make the following definition.

Definition 3.15.

Given ϕ=limnωϕnnωMn\phi=\lim_{n\to\omega}\phi_{n}\in\prod_{n\to\omega}M_{n}^{*}, we define its restricted Burnside power to be the homomorphism

d(ϕ):=limnωd(ϕn).\mathcal{B}_{d}(\phi):=\lim_{n\to\omega}\mathcal{B}_{d}(\phi_{n}).

This is a homomorphism from nωd(Mn)\prod_{n\to\omega}\mathcal{B}_{d}(M_{n}) to B(d,m)B(d,m).

Definition 3.16.

For any ultrafilter 𝒰\mathcal{U} on the set nωMn\prod_{n\to\omega}M_{n}^{*}, we may define

d(ϕ𝒰):=limϕ𝒰d(ϕ).\mathcal{B}_{d}(\phi_{\mathcal{U}}):=\lim_{\phi\to\mathcal{U}}\mathcal{B}_{d}(\phi).

such that for all xnωd(Mn)x\in\prod_{n\to\omega}\mathcal{B}_{d}(M_{n}), we have

d(ϕ𝒰)(x)=limϕ𝒰d(ϕ)(x).\mathcal{B}_{d}(\phi_{\mathcal{U}})(x)=\lim_{\phi\to\mathcal{U}}\mathcal{B}_{d}(\phi)(x).

The image limϕ𝒰d(ϕ)(x)\lim_{\phi\to\mathcal{U}}\mathcal{B}_{d}(\phi)(x) is well-defined because B(d,m)B(d,m) is a finite group.

Then under these definitions, the properties of homomorphisms from nωMn\prod_{n\to\omega}M_{n} to /m\mathbb{Z}/m\mathbb{Z} will carry over to homomorphisms from nωd(Mn)\prod_{n\to\omega}\mathcal{B}_{d}(M_{n}) to d(/m)=B(d,m)\mathcal{B}_{d}(\mathbb{Z}/m\mathbb{Z})=B(d,m). Two particularly notable properties are surjectivity and invariance under a group action.

Proposition 3.17.

Fix a sequence of finite groups MnM_{n} of exponents dividing mm. For any ultrafilter 𝒰\mathcal{U} on the set nωMn\prod_{n\to\omega}M_{n}^{*}, if ϕ𝒰\phi_{\mathcal{U}} is surjective onto /m\mathbb{Z}/m\mathbb{Z}, then d(ϕ𝒰)\mathcal{B}_{d}(\phi_{\mathcal{U}}) is also surjective onto B(d,m)B(d,m).

Proof.

Let ι:/mB(d,m)\iota:\mathbb{Z}/m\mathbb{Z}\to B(d,m) be the ii-th coprojection map. Let ιn:Mnd(Mn)\iota_{n}:M_{n}\to\mathcal{B}_{d}(M_{n}) be the corresponding ii-th coprojection map. We claim that d(ϕ𝒰)(limnωιn)=ιϕ𝒰\mathcal{B}_{d}(\phi_{\mathcal{U}})\circ(\lim_{n\to\omega}\iota_{n})=\iota\circ\phi_{\mathcal{U}}.

If the claim is true, then we can pick any xnωMnx\in\prod_{n\to\omega}M_{n} that is a pre-image of 1/m1\in\mathbb{Z}/m\mathbb{Z} under the map ϕ𝒰\phi_{\mathcal{U}}. Then d(ϕ𝒰)(limnωιn)(x)=ιϕ𝒰(x)=ι(1)\mathcal{B}_{d}(\phi_{\mathcal{U}})\circ(\lim_{n\to\omega}\iota_{n})(x)=\iota\circ\phi_{\mathcal{U}}(x)=\iota(1) is the ii-th generator of B(d,m)B(d,m). So each generator of B(d,m)B(d,m) is in the image of d(ϕ𝒰)\mathcal{B}_{d}(\phi_{\mathcal{U}}), hence d(ϕ𝒰)\mathcal{B}_{d}(\phi_{\mathcal{U}}) is surjective.

We now prove our claim. For any homomorphism ϕn:Mn/m\phi_{n}:M_{n}\to\mathbb{Z}/m\mathbb{Z}, by definition of the restricted Burnside coproduct functor, we have ιϕn=d(ϕn)ιn\iota\circ\phi_{n}=\mathcal{B}_{d}(\phi_{n})\circ\iota_{n}.

Mn{M_{n}}/m{\mathbb{Z}/m\mathbb{Z}}d(Mn){\mathcal{B}_{d}(M_{n})}B(d,m){B(d,m)}ϕn\scriptstyle{\phi_{n}}ιn\scriptstyle{\iota_{n}}ι\scriptstyle{\iota}d(ϕn)\scriptstyle{\mathcal{B}_{d}(\phi_{n})}

Taking ultralimit nωn\to\omega, we have ιϕ=d(ϕ)(limnωιn)\iota\circ\phi=\mathcal{B}_{d}(\phi)\circ(\lim_{n\to\omega}\iota_{n}) for any ϕ=limnωϕnnωMn\phi=\lim_{n\to\omega}\phi_{n}\in\prod_{n\to\omega}M_{n}^{*}.

nωMn{\prod_{n\to\omega}M_{n}}/m{\mathbb{Z}/m\mathbb{Z}}nωd(Mn){\prod_{n\to\omega}\mathcal{B}_{d}(M_{n})}B(d,m){B(d,m)}ϕ\scriptstyle{\phi}limnωιn\scriptstyle{\lim_{n\to\omega}\iota_{n}}ι\scriptstyle{\iota}d(ϕ)\scriptstyle{\mathcal{B}_{d}(\phi)}

Note that the right side of the diagram stays unchanged because the groups on the right side are all finite.

So we have the calculation

d(ϕ𝒰)(limnωιn)=limϕ𝒰[d(ϕ)(limnωιn)]=limϕ𝒰ιϕ=ι(limϕ𝒰ϕ)=ιϕ𝒰.\mathcal{B}_{d}(\phi_{\mathcal{U}})\circ(\lim_{n\to\omega}\iota_{n})=\lim_{\phi\to\mathcal{U}}[\mathcal{B}_{d}(\phi)\circ(\lim_{n\to\omega}\iota_{n})]=\lim_{\phi\to\mathcal{U}}\iota\circ\phi=\iota\circ(\lim_{\phi\to\mathcal{U}}\phi)=\iota\circ\phi_{\mathcal{U}}.

Here the ultralimit operator limϕ𝒰\lim_{\phi\to\mathcal{U}} and the homomorphism ι\iota commute because ι\iota is a homomorphism between finite groups independent of ϕ\phi. ∎

Proposition 3.18.

Fix a sequence of finite groups MnM_{n} of exponents dividing mm. Suppose we have a finite group GnG_{n} acting on MnM_{n} for each nn. So we have an action of G=nωGnG=\prod_{n\to\omega}G_{n} on nωMn\prod_{n\to\omega}M_{n}.

For any GG-invariant ultrafilter 𝒰\mathcal{U} on the set nωMn\prod_{n\to\omega}M_{n}^{*}, then both ϕ𝒰\phi_{\mathcal{U}} and d(ϕ)𝒰\mathcal{B}_{d}(\phi)_{\mathcal{U}} are GG-invariant.

Proof.

Suppose 𝒰\mathcal{U} is GG-invariant. For any gGg\in G, let SgS_{g} be the subset of gg-invariant elements of nωMn\prod_{n\to\omega}M_{n}^{*}. Then Sg𝒰S_{g}\in\mathcal{U} for all GG. So for each xnωMnx\in\prod_{n\to\omega}M_{n}, we have ϕ𝒰(xg)=limϕ𝒰ϕ(xg)=limϕ𝒰ϕ(x)=ϕ𝒰(x)\phi_{\mathcal{U}}(x^{g})=\lim_{\phi\to\mathcal{U}}\phi(x^{g})=\lim_{\phi\to\mathcal{U}}\phi(x)=\phi_{\mathcal{U}}(x).

Now for each ϕSg\phi\in S_{g}, say ϕ=limnωϕn\phi=\lim_{n\to\omega}\phi_{n} and g=limnωgng=\lim_{n\to\omega}g_{n}. Then ϕn\phi_{n} is gng_{n}-invariant for ω\omega-almost all nn. But by Proposition 3.8, this means d(ϕn)\mathcal{B}_{d}(\phi_{n}) is gng_{n}-invariant for ω\omega-almost all nn. So d(ϕ)=limnωd(ϕn)\mathcal{B}_{d}(\phi)=\lim_{n\to\omega}\mathcal{B}_{d}(\phi_{n}) is gg-invariant.

So d(ϕ)\mathcal{B}_{d}(\phi) is gg-invariant for all ϕSg\phi\in S_{g}, i.e., for 𝒰\mathcal{U}-almost all ϕ\phi. So for each xnωd(Mn)x\in\prod_{n\to\omega}\mathcal{B}_{d}(M_{n}), we have d(ϕ)𝒰(xg)=limϕ𝒰d(ϕ)(xg)=limϕ𝒰d(ϕ)(x)=d(ϕ)𝒰(x)\mathcal{B}_{d}(\phi)_{\mathcal{U}}(x^{g})=\lim_{\phi\to\mathcal{U}}\mathcal{B}_{d}(\phi)(x^{g})=\lim_{\phi\to\mathcal{U}}\mathcal{B}_{d}(\phi)(x)=\mathcal{B}_{d}(\phi)_{\mathcal{U}}(x).

Since gg is an arbitrary element of GG, both ϕ𝒰\phi_{\mathcal{U}} and d(ϕ)𝒰\mathcal{B}_{d}(\phi)_{\mathcal{U}} are GG-invariant. ∎

3.3 Reduction to the cyclic case

Recall that our goal is to realize the restricted Burnside group B(d,m)B(d,m) as a quotient of an ultraproduct of finite perfect groups. In this subsection, we shall use the fact B(d,m)=d(/m)B(d,m)=\mathcal{B}_{d}(\mathbb{Z}/m\mathbb{Z}) to reduce our goal to this: it is enough to realize the cyclic group /m\mathbb{Z}/m\mathbb{Z} as a quotient of an ultraproduct of “nice” finite perfect groups.

In particular, we shall show that it is enough to have the following lemma. We delay its proof to later subsections.

Lemma 3.19.

Pick any non-principal ultrafilter ω\omega on \mathbb{N} and any positive integer mm. Then we can find a sequence of finite groups {Gn}n\{G_{n}\}_{n\in\mathbb{N}}, each GnG_{n} acting on the corresponding finite abelian group MnM_{n} whose exponent divides mm, such that they satisfy the following conditions.

  1. 1.

    MnGnM_{n}\rtimes G_{n} is perfect for each nn.

  2. 2.

    There is a nωGn\prod_{n\to\omega}G_{n}-invariant surjective homomorphism ψ:nωMn/m\psi:\prod_{n\to\omega}M_{n}\to\mathbb{Z}/m\mathbb{Z}.

Note that the conditions here guarantee the following corollary.

Corollary 3.20.

Any cyclic group /m\mathbb{Z}/m\mathbb{Z} can be realized as an abstract quotient of an ultraproduct of finite perfect groups.

Proof.

Pick any non-principal ultrafilter ω\omega on \mathbb{N} and any positive integer mm. Let {Mn}n\{M_{n}\}_{n\in\mathbb{N}} and {Gn}n\{G_{n}\}_{n\in\mathbb{N}} and ψ\psi be the sequences of groups and the homomorphism in Proposition 3.19.

Since ψ\psi is nωGn\prod_{n\to\omega}G_{n}-invariant, by Lemma 3.21 below, it has an extension ψ~:(nωMn)(nωGn)/m\widetilde{\psi}:(\prod_{n\to\omega}M_{n})\rtimes(\prod_{n\to\omega}G_{n})\to\mathbb{Z}/m\mathbb{Z}. So ψ~\widetilde{\psi} is surjective, and as ultraproduct preserves semi-direct product, the domain of ψ~\widetilde{\psi} is canonically isomorphic to nω(MnGn)\prod_{n\to\omega}(M_{n}\rtimes G_{n}), an ultraproduct of finite perfect groups. ∎

Lemma 3.21.

Suppose we have a group GG acting on another group MM, and a GG-invariant homomorphism ψ:MB\psi:M\to B to some group BB. Then there is an extension ψ~:MGB\widetilde{\psi}:M\rtimes G\to B of ψ\psi.

Proof.

Let N=[M,G]N=[M,G]. Then since ψ\psi is GG-invariant, NKer(ψ)N\subseteq\mathrm{Ker}(\psi). So the map ψ\psi factors into a quotient map q:MM/Nq:M\to M/N and a map ψ:M/NB\psi^{\prime}:M/N\to B.

For any x,yMx,y\in M and g,hGg,h\in G, we have [x,g]h=[xh,gh]N[x,g]^{h}=[x^{h},g^{h}]\in N and [x,g]y=[xy,g][y,g]1N[x,g]^{y}=[xy,g][y,g]^{-1}\in N, so NN is a GG-invariant normal subgroup of MGM\rtimes G. So NGN\rtimes G is a subgroup of MGM\rtimes G. Furthermore, for any xN,gG,yM,hGx\in N,g\in G,y\in M,h\in G, then in the group MGM\rtimes G we have the following calculation:

(xg)y=xygy=xygyg1g=xy[y,g1]gNG.(xg)^{y}=x^{y}g^{y}=x^{y}g^{y}g^{-1}g=x^{y}[y,g^{-1}]g\in N\rtimes G.
(xg)h=xhgh=xx1xhgh=x[x,h]ghNG.(xg)^{h}=x^{h}g^{h}=xx^{-1}x^{h}g^{h}=x[x,h]g^{h}\in N\rtimes G.

Hence we see that NGN\rtimes G is a normal subgroup of MGM\rtimes G. But for each xMx\in M and gGg\in G, then [x,g]=(x1g1x)g[x,g]=(x^{-1}g^{-1}x)g is in the normal closure of GG. So NGN\rtimes G is in the normal closure of GG, and hence it is the normal closure of GG.

Let q~\widetilde{q} be the quotient homomorphism from MGM\rtimes G to (MG)/(NG)M/N(M\rtimes G)/(N\rtimes G)\simeq M/N. Then we have a commutative diagram:

N{N}M{M}M/N{M/N}B{B}NG{N\rtimes G}MG{M\rtimes G}(MG)/(NG){(M\rtimes G)/(N\rtimes G)}q\scriptstyle{q}\scriptstyle{\simeq}ψ\scriptstyle{\psi^{\prime}}q~\scriptstyle{\widetilde{q}}

So we have a map ψ~:MGB\widetilde{\psi}:M\rtimes G\to B extending ψ:MB\psi:M\to B. ∎

Assuming Lemma 3.19, which builds a homomorphism onto /m\mathbb{Z}/m\mathbb{Z}, now we show that it induces a similar construction which shall induce a homomorphism onto B(d,m)=d(/m)B(d,m)=\mathcal{B}_{d}(\mathbb{Z}/m\mathbb{Z}).

Lemma 3.22.

Pick any non-principal ultrafilter ω\omega on \mathbb{N} and any positive integers d,md,m. Let {Mn}n\{M_{n}\}_{n\in\mathbb{N}} and {Gn}n\{G_{n}\}_{n\in\mathbb{N}} and ψ\psi be the sequences of groups and the homomorphism in Lemma 3.19. Then they satisfy the following conditions:

  1. 1.

    d(Mn)Gn\mathcal{B}_{d}(M_{n})\rtimes G_{n} is perfect for each nn.

  2. 2.

    We can find a nωGn\prod_{n\to\omega}G_{n}-invariant surjective homomorphism from nωd(Mn)\prod_{n\to\omega}\mathcal{B}_{d}(M_{n}) to B(d,m)B(d,m).

Proof.

By Proposition 3.7, d(Mn)Gn\mathcal{B}_{d}(M_{n})\rtimes G_{n} is perfect for each nn.

Let G=nωGnG=\prod_{n\to\omega}G_{n}. Since ψ\psi is GG-invariant, by Proposition 3.12, we can find a GG-invariant ultrafilter 𝒰\mathcal{U} on nωMn\prod_{n\to\omega}M_{n}^{*} such that ψ=ϕ𝒰\psi=\phi_{\mathcal{U}}. Then by Proposition 3.18, d(ϕ)𝒰\mathcal{B}_{d}(\phi)_{\mathcal{U}} is also GG-invariant. Furthermore, by Proposition 3.17, since ϕ𝒰=ψ\phi_{\mathcal{U}}=\psi is surjective, d(ϕ)𝒰\mathcal{B}_{d}(\phi)_{\mathcal{U}} is also surjective. So we are done. ∎

Corollary 3.23.

For any positive integer d,md,m, the restricted Burnside group B(d,m)B(d,m) can be realized as an abstract quotient of an ultraproduct of finite perfect groups.

Proof.

Pick any non-principal ultrafilter ω\omega on \mathbb{N} and any positive integer mm. Let {Mn}n\{M_{n}\}_{n\in\mathbb{N}} and {Gn}n\{G_{n}\}_{n\in\mathbb{N}} and ψ\psi be the sequences of groups and the homomorphism in Lemma 3.19.

Let G=nωGnG=\prod_{n\to\omega}G_{n}. For any positive integer dd, by Lemma 3.22, we can then find the GG-invariant surjective homomorphism d(ϕ)𝒰:nωd(Mn)B(d,m)\mathcal{B}_{d}(\phi)_{\mathcal{U}}:\prod_{n\to\omega}\mathcal{B}_{d}(M_{n})\to B(d,m). By Lemma 3.21, d(ϕ)𝒰\mathcal{B}_{d}(\phi)_{\mathcal{U}} has an extension ψ~:nωd(Mn)GnB(d,m)\widetilde{\psi}:\prod_{n\to\omega}\mathcal{B}_{d}(M_{n})\rtimes G_{n}\to B(d,m).

Note that d(Mn)Gn\mathcal{B}_{d}(M_{n})\rtimes G_{n} is perfect for each nn, so B(d,m)B(d,m) is indeed an abstract quotient of an ultraproduct of finite perfect groups. ∎

Hence Theorem 1.4 is proven.

3.4 Constructing the group GnG_{n}

We now move on to the proof of Lemma 3.19, by an explicit construction of the groups GnG_{n} and MnM_{n} and the homomorphism ψ\psi.

Pick any two distinct prime numbers p,qp,q coprime to mm, and p5p\geq 5. For each nn\in\mathbb{N}, the alternating group Ap\mathrm{A}_{p} acts on the 𝔽q\mathbb{F}_{q}-vector space (𝔽qn)p(\mathbb{F}_{q}^{n})^{p} by permuting the pp factor spaces 𝔽qn\mathbb{F}_{q}^{n}. Let VnV_{n} be the subspace of (𝔽qn)p(\mathbb{F}_{q}^{n})^{p} of tuples (v1,,vp)(v_{1},\dots,v_{p}) with v1,,vp𝔽qnv_{1},\dots,v_{p}\in\mathbb{F}_{q}^{n} such that vi=0\sum v_{i}=0. Note that VnV_{n} is Ap\mathrm{A}_{p}-invariant.

Proposition 3.24.

VnApV_{n}\rtimes\mathrm{A}_{p} is a perfect group with commutator width at most 2.

Proof.

For any σAp\sigma\in\mathrm{A}_{p}, we have (v1,,vp)σ=(vσ(1),,vσ(p))(v_{1},\dots,v_{p})^{\sigma}=(v_{\sigma(1)},\dots,v_{\sigma(p)}) for any v1,,vp𝔽qnv_{1},\dots,v_{p}\in\mathbb{F}_{q}^{n}. Now if v1++vp=0v_{1}+\dots+v_{p}=0, then vσ(1)++vσ(p)=0v_{\sigma(1)}+\dots+v_{\sigma(p)}=0 as well. Hence VnV_{n} is an Ap\mathrm{A}_{p}-invariant subspace.

Since p5p\geq 5 and it is a prime, it must be odd. So let σ\sigma be the pp-cycle in Ap\mathrm{A}_{p} such that (v1,,vp)σ=(vp,v1,v2,,vp1)(v_{1},\dots,v_{p})^{\sigma}=(v_{p},v_{1},v_{2},\dots,v_{p-1}) for any v1,,vp𝔽qnv_{1},\dots,v_{p}\in\mathbb{F}_{q}^{n}. Note that the action by σ\sigma induces an automorphism of VnV_{n}, and if (v1,,vp)σ=(v1,,vp)(v_{1},\dots,v_{p})^{\sigma}=(v_{1},\dots,v_{p}) in VnV_{n}, then we must have vi=vjv_{i}=v_{j} for all i,ji,j. Since p,qp,q are coprime, 0=v1++vp=pvi0=v_{1}+\dots+v_{p}=pv_{i} implies that vi=0v_{i}=0 for all ii. So σ\sigma induces an automorphism with only trivial fixed point.

Now the map v[v,σ]v\mapsto[v,\sigma] is a linear automorphism of VnV_{n}, and its kernel is exactly the set of fixed points of σ\sigma, which must be trivial. So this map is bijective. So for any vVnv\in V_{n}, we can find wVnw\in V_{n} such that v=[w,σ]v=[w,\sigma]. So all elements of VnV_{n} are commutators of VnApV_{n}\rtimes\mathrm{A}_{p}.

Now for the group VnApV_{n}\rtimes\mathrm{A}_{p}, all elements in Ap\mathrm{A}_{p} are commutators since p5p\geq 5, and all elements in VnV_{n} are commutators. Therefore VnApV_{n}\rtimes\mathrm{A}_{p} is perfect with commutator width at most 2. ∎

Remark 3.25.

The bound here is tight. Suppose p=5p=5. Let σ=(12)(34)\sigma=(12)(34) and pick any vVnv\in V_{n}, and consider vσVnApv\sigma\in V_{n}\rtimes\mathrm{A}_{p}. Suppose vσ=[v1σ1,v2σ2]v\sigma=[v_{1}\sigma_{1},v_{2}\sigma_{2}], then [σ1,σ2]=σ[\sigma_{1},\sigma_{2}]=\sigma. Then σ1(5)=σ2(5)=5\sigma_{1}(5)=\sigma_{2}(5)=5 by Lemma 3.26 below.

Now [v1σ1,v2σ2]=(L1(L2I)v1+L1L2(L11I)v2)σ[v_{1}\sigma_{1},v_{2}\sigma_{2}]=(L_{1}(L_{2}-I)v_{1}+L_{1}L_{2}(L_{1}^{-1}-I)v_{2})\sigma, where LiL_{i} is the linear automorphism on VnV_{n} induced by σi\sigma_{i}. But since both L1,L2L_{1},L_{2} fix the fifth component, L1(L2I)L_{1}(L_{2}-I) and L1L2(L11I)L_{1}L_{2}(L_{1}^{-1}-I) must kill the fifth component. So we must have vσ[v1σ1,v2σ2]v\sigma\neq[v_{1}\sigma_{1},v_{2}\sigma_{2}] when vv has non-zero fifth component.

Lemma 3.26 (Remark Lemma).

If [σ1,σ2]=(12)(34)[\sigma_{1},\sigma_{2}]=(12)(34) in A5\mathrm{A}_{5}, then σ1(5)=σ2(5)=5\sigma_{1}(5)=\sigma_{2}(5)=5.

Proof.

The proof is basically a direct enumeration of all possible cycle-types of σ1\sigma_{1}.

Suppose σ1\sigma_{1} is a 55-cycle, then this implies that a product of two 5-cycles σ=σ11\sigma=\sigma_{1}^{-1} and σ=σ21σ1σ2\sigma^{\prime}=\sigma_{2}^{-1}\sigma_{1}\sigma_{2} is (12)(34)(12)(34).

Suppose σ,σ\sigma,\sigma^{\prime} are two 5-cycles such that σσ=(12)(34)\sigma\sigma^{\prime}=(12)(34). Then σ(σ(5))=5\sigma^{\prime}(\sigma(5))=5. By symmetry, WLOG, say σ(5)=1\sigma(5)=1. Since σ(σ(1))=2\sigma^{\prime}(\sigma(1))=2 and σ\sigma^{\prime} has no fixed point, we see that σ(1)2\sigma(1)\neq 2. We also know that σ(1)1\sigma(1)\neq 1 since σ\sigma has no fixed point and σ(1)5\sigma(1)\neq 5 since σ\sigma is a 5-cycle. Hence σ(1)=3\sigma(1)=3 or 44. By symmetry, WLOG say σ(1)=3\sigma(1)=3.

Now σ(σ(3))=4\sigma^{\prime}(\sigma(3))=4, so σ(3)4\sigma(3)\neq 4. But we also have σ(3)3\sigma(3)\neq 3 or 55 since σ\sigma is a 5-cycle, and σ(3)1\sigma(3)\neq 1 since σ(5)=1\sigma(5)=1. This means σ(3)=2\sigma(3)=2.

Now σ(2)2,3,5\sigma(2)\neq 2,3,5 since they are already taken by other σ\sigma-values, and σ(σ(2))=1\sigma^{\prime}(\sigma(2))=1 implies σ(2)1\sigma(2)\neq 1. So σ(2)=4\sigma(2)=4. And finally, this means σ(4)=5\sigma(4)=5. So σ=(13245)\sigma=(13245). It is then easy to see that σ=(14235)1\sigma^{\prime}=(14235)^{-1}. But now we see that (13245)(13245) and (14235)(14235) are NOT conjugate in A5\mathrm{A}_{5}. However, by definition σ=σ11\sigma=\sigma_{1}^{-1} and σ=σ21σ1σ2\sigma^{\prime}=\sigma_{2}^{-1}\sigma_{1}\sigma_{2} are conjugate in A5\mathrm{A}_{5}, a contradiction. So σ1\sigma_{1} cannot be a 5-cycle. Similarly, σ2\sigma_{2} also cannot be a 5-cycle.

Now suppose σ1=σ\sigma_{1}=\sigma is a 3-cycle. Then (12)(34)(12)(34) is the product of two 3-cycles. Note that the union of the supports of the two 3-cycles must obviously include {1,2,3,4}\{1,2,3,4\}. However, if two 3-cycles have only one element in common in their support, then their product is a 5-cycle. So the two 3-cycles must have two elements in common. So they must both have support inside {1,2,3,4}\{1,2,3,4\}. WLOG suppose the other fixed point of σ\sigma is 44. Then σ(3)=1\sigma(3)=1 or 22. WLOG suppose σ(3)=1\sigma(3)=1. Then σ=(123)\sigma=(123) and σ=(143)=(134)1\sigma^{\prime}=(143)=(134)^{-1}. So σ2=(234),(142)\sigma_{2}=(234),(142) or (13)(24)(13)(24). We see that in this case, indeed we have σ1(5)=σ2(5)=5\sigma_{1}(5)=\sigma_{2}(5)=5.

Finally, suppose neither σ1\sigma_{1} nor σ2\sigma_{2} is a 3-cycle. Then both σ=σ11\sigma=\sigma_{1}^{-1} and σ=σ21σ1σ2\sigma^{\prime}=\sigma_{2}^{-1}\sigma_{1}\sigma_{2} must be of cycle type (2,2,1)(2,2,1). Then σσ=(12)(34)=[(12)(34)]1=(σ)1σ1=σσ\sigma\sigma^{\prime}=(12)(34)=[(12)(34)]^{-1}=(\sigma^{\prime})^{-1}\sigma^{-1}=\sigma^{\prime}\sigma. Suppose σ(5)5\sigma(5)\neq 5, say WLOG σ(5)=1\sigma(5)=1. Then since σσ=(12)(34)\sigma\sigma^{\prime}=(12)(34), we see that σ(5)=1\sigma^{\prime}(5)=1 as well. So σ=(15)(ab)\sigma=(15)(ab) and σ=(15)(cd)\sigma^{\prime}=(15)(cd), where a,b,c,d{2,3,4}a,b,c,d\in\{2,3,4\}. But then σσ\sigma\sigma^{\prime} is a 3-cycle, a contradiction. So σ(5)=5\sigma(5)=5, and therefore σ1(5)=σ2(5)=5\sigma_{1}(5)=\sigma_{2}(5)=5 as desired. ∎

We shall set Gn=VnApG_{n}=V_{n}\rtimes\mathrm{A}_{p}. Let V=nωVnV=\prod_{n\to\omega}V_{n} and G=nωGnG=\prod_{n\to\omega}G_{n}.

Proposition 3.27.

G=VApG=V\rtimes\mathrm{A}_{p}.

Proof.

Note that the ultraproduct of semi-direct products is the semidirect product of ultraproducts. And since Ap\mathrm{A}_{p} is finite, we have a canonical identification nωAp=Ap\prod_{n\to\omega}\mathrm{A}_{p}=\mathrm{A}_{p}. ∎

We now establish some properties of GG to be used later. Note that 𝔽qn\mathbb{F}_{q}^{n} has a natural “inner product”. If v,w𝔽qnv,w\in\mathbb{F}_{q}^{n}, let them be v=(v1,,vn),w=(w1,,wn)v=(v_{1},\dots,v_{n}),w=(w_{1},\dots,w_{n}) for v1,,vn,w1,,wn𝔽qv_{1},\dots,v_{n},w_{1},\dots,w_{n}\in\mathbb{F}_{q}, then we define v,w=viwi\langle v,w\rangle=\sum v_{i}w_{i}. Note that this defines a non-degenerate symmetric bilinear form on 𝔽qn\mathbb{F}_{q}^{n}. This then induces a symmetric bilinear form on (𝔽qn)p(\mathbb{F}_{q}^{n})^{p} such that (v1,,vp),(w1,,wp)=vi,wi\langle(v_{1},\dots,v_{p}),(w_{1},\dots,w_{p})\rangle=\sum\langle v_{i},w_{i}\rangle for v1,,vp,w1,,wp𝔽qnv_{1},\dots,v_{p},w_{1},\dots,w_{p}\in\mathbb{F}_{q}^{n}.

Proposition 3.28.

v(σ1),w=v,wσ\langle v^{(\sigma^{-1})},w\rangle=\langle v,w^{\sigma}\rangle for all v,w(𝔽qn)pv,w\in(\mathbb{F}_{q}^{n})^{p} and σAp\sigma\in\mathrm{A}_{p}.

Proof.

(v1,,vp)(σ1),(w1,,wp)=vσ1(i)wi=viwσ(i)=(v1,,vp),(w1,,wp)σ\langle(v_{1},\dots,v_{p})^{(\sigma^{-1})},(w_{1},\dots,w_{p})\rangle=\sum v_{\sigma^{-1}(i)}w_{i}=\sum v_{i}w_{\sigma(i)}=\langle(v_{1},\dots,v_{p}),(w_{1},\dots,w_{p})^{\sigma}\rangle. ∎

Proposition 3.29.

The induced bilinear form on (𝔽qn)p(\mathbb{F}_{q}^{n})^{p} restricts to a symmetric non-degenerate bilinear form on VnV_{n}.

Proof.

We only needs to verify that it is non-degenerate. Suppose for some vVnv\in V_{n}, v,w=0\langle v,w\rangle=0 for all wVnw\in V_{n}. Then if v=(v1,,vp)v=(v_{1},\dots,v_{p}) for v1,,vp𝔽qnv_{1},\dots,v_{p}\in\mathbb{F}_{q}^{n}, pick any w𝔽qnw^{\prime}\in\mathbb{F}_{q}^{n}, then v,(w,w,0,,0)=0\langle v,(w^{\prime},-w^{\prime},0,\dots,0)\rangle=0 implies that v1,w=v2,w\langle v_{1},w^{\prime}\rangle=\langle v_{2},w^{\prime}\rangle for all w𝔽qnw^{\prime}\in\mathbb{F}_{q}^{n}. Since ,\langle-,-\rangle is non-degenerate on 𝔽qn\mathbb{F}_{q}^{n}, we have v1=v2v_{1}=v_{2}. Similarly we have vi=vjv_{i}=v_{j} for all i,ji,j. But since vVnv\in V_{n}, we have vi=0\sum v_{i}=0, so pvi=0pv_{i}=0. Finally, since p,qp,q are coprime, we have vi=0v_{i}=0 for all ii. So v=0v=0. ∎

Lemma 3.30.

The symmetric non-degenerate bilinear forms on VnV_{n} induce a symmetric non-degenerate bilinear form on the ultraproduct V=nωVnV=\prod_{n\to\omega}V_{n}.

Proof.

Taking the ultralimit of ,:Vn×Vn𝔽q\langle-,-\rangle:V_{n}\times V_{n}\to\mathbb{F}_{q}, we obtain a symmetric bilinear map ,:V×V𝔽q\langle-,-\rangle:V\times V\to\mathbb{F}_{q}. Let us verify that it is indeed non-degenerate.

Suppose limnωvn\lim_{n\to\omega}v_{n} is non-zero in VV. Then vn0v_{n}\neq 0 for ω\omega-almost all nn. For each vn0v_{n}\neq 0, since ,\langle-,-\rangle is non-degenerate, we can find wnVnw_{n}\in V_{n} such that vn,wn0\langle v_{n},w_{n}\rangle\neq 0. If vn=0v_{n}=0, we pick wnw_{n} arbitrarily. So vn,wn0\langle v_{n},w_{n}\rangle\neq 0 for ω\omega-almost all nn by construction.

Then since our bilinear form takes value in a finite set 𝔽q\mathbb{F}_{q}, we have limnωvn,limnωvn=limnωvn,wn0\langle\lim_{n\to\omega}v_{n},\lim_{n\to\omega}v_{n}\rangle=\lim_{n\to\omega}\langle v_{n},w_{n}\rangle\neq 0. So ,\langle-,-\rangle is indeed non-degenerate. ∎

We now study subspaces of VV with finite codimension, which would be useful later. We first make a definition of a “closed” subspace.

Definition 3.31.

Given a vector space VV with a symmetric non-degenerate bilinear form, we say v,wVv,w\in V are perpendicular if v,w=0\langle v,w\rangle=0.

We define the orthogonal complement of a subset SVS\subseteq V to be the set S:={vV:v,w=0 for all wS}S^{\perp}:=\{v\in V:\langle v,w\rangle=0\text{ for all $w\in S$}\}, i.e., the subspace of elements perpendicular to all elements of SS.

We say a subspace WW of VV is closed if W=SW=S^{\perp} for some subset SS.

Lemma 3.32.

If STS\subseteq T for two subsets S,TS,T in a vector space VV with a symmetric non-degenerate bilinear form, then TST^{\perp}\subseteq S^{\perp} and SSS\subseteq S^{\perp\perp}.

Proof.

If vTv\in T^{\perp}, then vv is perpendicular to all elements of TT, which includes all elements of SS. Hence vSv\in S^{\perp}.

If vSv\in S, then it is by definition perpendicular to all elements of SS^{\perp}, so sSs\in S^{\perp\perp}. ∎

Lemma 3.33.

Given a vector space VV with a symmetric non-degenerate bilinear form, then we have the following results.

  1. 1.

    A subspace WW is closed if and only if W=WW=W^{\perp\perp}.

  2. 2.

    If WW is a closed subspace with finite codimension cc, then WW^{\perp} has dimension cc.

  3. 3.

    Finite dimensional subspaces are closed.

Proof.

Suppose W=SW=S^{\perp}, then SS=WS\subseteq S^{\perp\perp}=W^{\perp}, and hence WS=WW^{\perp\perp}\subseteq S^{\perp}=W. On the other hand, WWW\subseteq W^{\perp\perp}. So we see W=WW=W^{\perp\perp}.

Now suppose WW is a subspace with finite codimension cc. Pick linearly independent vectors v1,,vcVv_{1},\dots,v_{c}\in V such that W,v1,,vcW,v_{1},\dots,v_{c} span VV. Then since the bilinear form is non-degenerate, W{v1}{vc}W^{\perp}\cap\{v_{1}\}^{\perp}\cap\dots\cap\{v_{c}\}^{\perp} must be the trivial.

Since each viv_{i} is non-zero and the bilinear form is non-degenerate, the map wv,ww\mapsto\langle v,w\rangle is a surjective linear map from VV to 𝔽q\mathbb{F}_{q} with kernel {vi}\{v_{i}\}^{\perp}, so {vi}\{v_{i}\}^{\perp} is a closed subspace of codimension 11.

Since W,v1,,vcW,v_{1},\dots,v_{c} are linearly independent, the subspaces W,{v1},,{vc}W^{\perp},\{v_{1}\}^{\perp},\dots,\{v_{c}\}^{\perp} intersect each other transversally. Since W{v1}{vc}W^{\perp}\cap\{v_{1}\}^{\perp}\cap\dots\cap\{v_{c}\}^{\perp} has dimension zero, WW^{\perp} has dimension cc.

Finally, suppose WW is finite dimensional. Say it has a basis v1,,vkv_{1},\dots,v_{k}. Then W={v1}{vk}W^{\perp}=\{v_{1}\}^{\perp}\cap\dots\cap\{v_{k}\}^{\perp}, so it is a closed subspace of codimension kk. So dimW=k\dim W^{\perp\perp}=k. Since we also have dimW=k\dim W=k and WWW\subseteq W^{\perp\perp}, we conclude that W=WW=W^{\perp\perp}, i.e., it is a closed subspace. ∎

We now go back to our discussion of the group G=VApG=V\rtimes\mathrm{A}_{p} where V=nωVnV=\prod_{n\to\omega}V_{n}.

Proposition 3.34.

If WW is a closed Ap\mathrm{A}_{p}-invariant subspace of VV of codimension cc, then for ω\omega-almost all nn, we can find closed Ap\mathrm{A}_{p}-invariant subspace WnW_{n} of VnV_{n} of codimension cc such that W=nωWnW=\prod_{n\to\omega}W_{n}.

Proof.

By Lemma 3.33, WW^{\perp} is cc-dimensional. Let v1,,vcv_{1},\dots,v_{c} be a basis for WW^{\perp}, say vi=limnωvi,nVv_{i}=\lim_{n\to\omega}v_{i,n}\in V. Let WnW_{n} be the orthogonal complement of v1,n,,vc,nv_{1,n},\dots,v_{c,n}. Since v1,,vcv_{1},\dots,v_{c} are linearly independent, and there are only finitely many possible linear combinations among cc vectors, we see that v1,n,,vc,nv_{1,n},\dots,v_{c,n} are linearly independent for ω\omega-almost all nn. So WnW_{n} has codimension cc.

Let us show that nωWn\prod_{n\to\omega}W_{n} is exactly the orthogonal complement of {v1,,vc}\{v_{1},\dots,v_{c}\}.

Fix 1ic1\leq i\leq c. For each w=limnωwnnωWnw=\lim_{n\to\omega}w_{n}\in\prod_{n\to\omega}W_{n}, then since wnWnw_{n}\in W_{n} for all nn, we see that wnvi,nw_{n}\perp v_{i,n}. So w,vi=limnωwn,vi,n=0\langle w,v_{i}\rangle=\lim_{n\to\omega}\langle w_{n},v_{i,n}\rangle=0.

Conversely, suppose w=limnωwnVw=\lim_{n\to\omega}w_{n}\in V is orthogonal to all viv_{i}. Then limnωwn,vi,n=0\lim_{n\to\omega}\langle w_{n},v_{i,n}\rangle=0 for all ii. Since there are only finitely many possible ii, therefore for ω\omega-almost all nn, wn,vi,n=0\langle w_{n},v_{i,n}\rangle=0 for all ii, and hence wnWnw_{n}\in W_{n}. So wnωWnw\in\prod_{n\to\omega}W_{n}.

In conclusion, we have seen that W=nωWnW=\prod_{n\to\omega}W_{n}.

Now we need to show that WnW_{n} is Ap\mathrm{A}_{p}-invariant for ω\omega-almost all nn. Suppose for contradiction that this is not the case. Then for ω\omega-almost all nn, WnW_{n} is NOT Ap\mathrm{A}_{p}-invariant, and we can find wnWnw_{n}\in W_{n} such that wnσnWnw_{n}^{\sigma_{n}}\notin W_{n} for some σn\sigma_{n}. So limnωwnW\lim_{n\to\omega}w_{n}\in W but limnωwnσnW\lim_{n\to\omega}w_{n}^{\sigma_{n}}\notin W.

However, since Ap\mathrm{A}_{p} is a finite set, let σ=limnωσn\sigma=\lim_{n\to\omega}\sigma_{n}. Then limnωwnσn=limnωwnσ=(limnωwn)σW\lim_{n\to\omega}w_{n}^{\sigma_{n}}=\lim_{n\to\omega}w_{n}^{\sigma}=(\lim_{n\to\omega}w_{n})^{\sigma}\in W, a contradiction. So for ω\omega-almost all nn, WnW_{n} must be an Ap\mathrm{A}_{p}-invariant subspace. ∎

Proposition 3.35.

Suppose WW is a closed subspace of VV with finite codimension. Then we can find a closed Ap\mathrm{A}^{p}-invariant subspace WWW^{\prime}\subseteq W with finite codimension in VV.

Proof.

Let v1,,vkv_{1},\dots,v_{k} be a basis for WW^{\perp}. Then the orbit 𝒪(vi)\mathcal{O}(v_{i}) for each viv_{i} under the Ap\mathrm{A}^{p} action has at most p!2\frac{p!}{2} elements. So the union of all these orbits has at most k(p!)2\frac{k(p!)}{2} elements, and they span a finite dimensional Ap\mathrm{A}_{p}-invariant subspace UU containing WW^{\perp}. So UU^{\perp} is an Ap\mathrm{A}_{p}-invariant closed subspace.

Note that UU is finite dimensional, so by Lemma 3.33, it must be a closed subspace. So UU^{\perp} has codimension dimU\dim U, which is at most k(p!)2\frac{k(p!)}{2}. ∎

Finally, here is a lemma we need for future use.

Lemma 3.36.

If WnVnW_{n}\subseteq V_{n} is a subspace of codimension cc, and n>c(p1)n>c(p-1). Then we can find wWnw\in W_{n} such that its orbit 𝒪(w)\mathcal{O}(w) in VnV_{n} under the Ap\mathrm{A}_{p} action has exactly pp elements.

Proof.

Since VnV_{n} is finite dimensional, all subspaces are closed. Let WnW_{n} be the orthogonal complement of v1,,vcVnv_{1},\dots,v_{c}\in V_{n}. Let vi=(vi,1,,vi,p)Vn(𝔽qn)pv_{i}=(v_{i,1},\dots,v_{i,p})\in V_{n}\subseteq(\mathbb{F}_{q}^{n})^{p}, where vi,j𝔽qnv_{i,j}\in\mathbb{F}_{q}^{n} and jvi,j=0\sum_{j}v_{i,j}=0 for each ii.

Now consider the span of these vi,jv_{i,j} in 𝔽qn\mathbb{F}_{q}^{n}. Since jvi,j=0\sum_{j}v_{i,j}=0 for each ii, this subspace has dimension at most c(p1)c(p-1). Let UU be the orthogonal complment of this subspace in 𝔽qn\mathbb{F}_{q}^{n}. Since n>c(p1)n>c(p-1), UU is non-trivial. Pick any non-zero uUu\in U. Consider w=(u,,u,(1p)u)Vnw=(u,\dots,u,(1-p)u)\in V_{n}. Since uvi,ju\perp v_{i,j} for all i,ji,j, we see that wviw\perp v_{i} for all ii. Hence wWnw\in W_{n}.

Now Ap\mathrm{A}_{p} acts on VnV_{n} by permuting the pp components. Since p,qp,q are coprime, we must have (1p)u=upuu(1-p)u=u-pu\neq u. Hence 𝒪(w)\mathcal{O}(w) has exactly pp elements. ∎

3.5 Constructing the /m\mathbb{Z}/m\mathbb{Z}-modules MnM_{n} and an action by GnG_{n}

We now move on to the construction of MnM_{n} in Lemma 3.19. Note that the category of finite abelian groups whose exponent divides mm is the same as the category of finite /m\mathbb{Z}/m\mathbb{Z}-modules. So we construct MnM_{n} as /m\mathbb{Z}/m\mathbb{Z}-modules. We use the same p,qp,q, and GnG_{n} as in the last section.

Consider the free /m\mathbb{Z}/m\mathbb{Z}-module (/m)q(\mathbb{Z}/m\mathbb{Z})^{q}. Let BB be the submodule of (/m)q(\mathbb{Z}/m\mathbb{Z})^{q} of tuples (t1,,tq)(t_{1},\dots,t_{q}) with t1,,tq/mt_{1},\dots,t_{q}\in\mathbb{Z}/m\mathbb{Z} such that ti=0\sum t_{i}=0. Consider the module automorphism f:BBf:B\to B such that f:(t1,,tq)(tq,t1,,tq1)f:(t_{1},\dots,t_{q})\mapsto(t_{q},t_{1},\dots,t_{q-1}).

Proposition 3.37.

BB is a free module, and ff only has the trivial fixed-point and has order qq.

Proof.

Obviously fq=idf^{q}=\mathrm{id} the identity automorphism, and if f((t1,,tq))=(t1,,tq)f((t_{1},\dots,t_{q}))=(t_{1},\dots,t_{q}) for an element (t1,,tq)B(t_{1},\dots,t_{q})\in B, then we see that all tit_{i} are the same. But since ti=0\sum t_{i}=0 and q,mq,m are coprime, we see that ti=0t_{i}=0 for all ii. Hence ff only has the trivial fixed-point.

Let us now show that BB is a free module. Set b1=(1,1,0,,0)b_{1}=(1,-1,0,\dots,0), and set bi+1=f(bi)b_{i+1}=f(b_{i}) for each i/qi\in\mathbb{Z}/q\mathbb{Z}. We claim that b1,,bq1b_{1},\dots,b_{q-1} form a basis.

For any b=(t1,,tq)Bb=(t_{1},\dots,t_{q})\in B, then since tq=t1tq1t_{q}=-t_{1}-\dots-t_{q-1}, we have

b=t1b1+(t1+t2)b2++(t1++tq1)bq1.b=t_{1}b_{1}+(t_{1}+t_{2})b_{2}+\dots+(t_{1}+\dots+t_{q-1})b_{q-1}.

So b1,,bq1b_{1},\dots,b_{q-1} are spanning.

Now suppose tibi=(0,,0)\sum t_{i}b_{i}=(0,\dots,0). Then since the first component of tibi\sum t_{i}b_{i} is zero, and only b1b_{1} contributes to this component, we see that t1=0t_{1}=0. Moving on inductively, we see that ti=0t_{i}=0 for all ii, and hence b1,,bq1b_{1},\dots,b_{q-1} are linearly independent. ∎

Corollary 3.38.

fidf-\mathrm{id} is also an automorphism of BB, where id\mathrm{id} is the identity automorphism of BB.

Let Mn=BVn{0}M_{n}=B^{V_{n}-\{0\}}, the /m\mathbb{Z}/m\mathbb{Z}-module of all set-functions from Vn{0}V_{n}-\{0\} to BB. Then clearly MnM_{n} is a free /m\mathbb{Z}/m\mathbb{Z} module.

We let VnV_{n} act on MnM_{n} such that xv(w)=fv,w(x(w))x^{v}(w)=f^{\langle v,w\rangle}(x(w)) for each xMn,vVn,wVn{0}x\in M_{n},v\in V_{n},w\in V_{n}-\{0\}. This is well defined because v,w𝔽q=/q\langle v,w\rangle\in\mathbb{F}_{q}=\mathbb{Z}/q\mathbb{Z} and ff has order qq.

We also let Ap\mathrm{A}_{p} act on MnM_{n} such that xσ(v)=x(v(σ1))x^{\sigma}(v)=x(v^{(\sigma^{-1})}) for each xMn,vVn{0},σApx\in M_{n},v\in V_{n}-\{0\},\sigma\in\mathrm{A}_{p}.

Proposition 3.39.

We have a group action of Gn=VnApG_{n}=V_{n}\ltimes\mathrm{A}_{p} on MnM_{n} induced from the group actions of VnV_{n} and Ap\mathrm{A}_{p} on MnM_{n}.

Proof.
xσ1vσ(w)=xσ1v(w(σ1))=fv,w(σ1)(xσ1(w(σ1)))=fv,w(σ1)(x(w))=fvσ,w(x(w))=x(vσ)(w).x^{\sigma^{-1}v\sigma}(w)=x^{\sigma^{-1}v}(w^{(\sigma^{-1})})=f^{\langle v,w^{(\sigma^{-1})}\rangle}(x^{\sigma^{-1}}(w^{(\sigma^{-1})}))=f^{\langle v,w^{(\sigma^{-1})}\rangle}(x(w))=f^{\langle v^{\sigma},w\rangle}(x(w))=x^{(v^{\sigma})}(w).

So the group action is well-defined. ∎

3.6 MnGnM_{n}\rtimes G_{n} is perfect

We not only want this to be perfect, but we also want MnGnM_{n}\rtimes G_{n} to have increasing commutator width. Otherwise, nωMnGn\prod_{n\to\omega}M_{n}\rtimes G_{n} will also have bounded commutator length and become a perfect group, in which case any ψ:nωMnGn/m\psi:\prod_{n\to\omega}M_{n}\rtimes G_{n}\to\mathbb{Z}/m\mathbb{Z} would be trivial.

Proposition 3.40.

The group MnGnM_{n}\rtimes G_{n} is perfect.

Proof.

We already know that Gn=VnApG_{n}=V_{n}\rtimes\mathrm{A}_{p} is perfect. So we just need to show that MnM_{n} is also in the commutator subgroup of MnGnM_{n}\rtimes G_{n}. We are going to show that Mn=[Mn,Vn]M_{n}=[M_{n},V_{n}].

Pick any vVn{0}v\in V_{n}-\{0\} and any bBb\in B, and let xvbMnx_{v\to b}\in M_{n} be the function such that xvb(v)=bx_{v\to b}(v)=b and xvb(v)=0x_{v\to b}(v^{\prime})=0 for all vvv^{\prime}\neq v. Note that such elements generate MnM_{n}. So we only need to show that xvbx_{v\to b} for all vVn{0}v\in V_{n}-\{0\} and all bBb\in B are in [Vn,Mn][V_{n},M_{n}].

Pick any wVnw\in V_{n} such that ww is NOT perpendicular to vv, say w,v=1\langle w,v\rangle=1. Then for any vvv^{\prime}\neq v, we have

[xvb,w](v)=(xvbwxvb)(v)=f(0)0=0.[x_{v\to b},w](v^{\prime})=(x_{v\to b}^{w}-x_{v\to b})(v^{\prime})=f(0)-0=0.

And we also have

[xvb,w](v)=(xvbwxvb)(v)=f(b)b=(fid)(b).[x_{v\to b},w](v)=(x_{v\to b}^{w}-x_{v\to b})(v)=f(b)-b=(f-\mathrm{id})(b).

Here id\mathrm{id} is the identity automorphism on BB. So [xvb,w]=xv(fid)(b)[x_{v\to b},w]=x_{v\to(f-\mathrm{id})(b)}.

Since fidf-\mathrm{id} is also an automorphism, we see that for any bBb^{\prime}\in B, we can find bBb\in B such that b=(fid)bb^{\prime}=(f-\mathrm{id})b, so

xvb=[xvb,w][Vn,Mn].x_{v\to b^{\prime}}=[x_{v\to b},w]\in[V_{n},M_{n}].

So we are done. ∎

We make the following useful definition.

Definition 3.41.

For each xMnx\in M_{n}, we define the null set for xx as

null(x):={0}{vVn{0}x(v)=0}.\mathrm{null}(x):=\{0\}\cup\{v\in V_{n}-\{0\}\mid x(v)=0\}.

We include zero for convenience. Under certain conditions, we want null(x)\mathrm{null}(x) to be some subspace of VnV_{n}.

Lemma 3.42.

For any σAp\sigma\in\mathrm{A}_{p} vVnv\in V_{n} and xMnx\in M_{n}, we have

[x,vσ]=\displaystyle[x,v\sigma]= [xv,σ]+[x,v],\displaystyle[x^{v},\sigma]+[x,v],
{v}\displaystyle\{v\}^{\perp}\subseteq null([x,v]).\displaystyle\mathrm{null}([x,v]).

And for each Ap\mathrm{A}_{p}-orbit SS of Vn{0}V_{n}-\{0\}, we have

wS[x,σ](w)=0.\sum_{w\in S}[x,\sigma](w)=0.
Proof.
[x,vσ]=(xv)σx=(xv)σxv+xvx=[xv,σ]+[x,v].[x,v\sigma]=(x^{v})^{\sigma}-x=(x^{v})^{\sigma}-x^{v}+x^{v}-x=[x^{v},\sigma]+[x,v].

If v,w=0\langle v,w\rangle=0, then

[x,v](w)=xv(w)x(w)=fv,w(x(w))x(w)=f0(x(w))x(w)=0.[x,v](w)=x^{v}(w)-x(w)=f^{\langle v,w\rangle}(x(w))-x(w)=f^{0}(x(w))-x(w)=0.

Finally, for each Ap\mathrm{A}_{p}-orbit SS of Vn{0}V_{n}-\{0\},

wS[x,σ](w)=(wSxσ(w))(wSx(w))=(wSx(w(σ1)))(wSx(w))=(wSx(w))(wSx(w))=0.\sum_{w\in S}[x,\sigma](w)=(\sum_{w\in S}x^{\sigma}(w))-(\sum_{w\in S}x(w))=(\sum_{w\in S}x(w^{(\sigma^{-1})}))-(\sum_{w\in S}x(w))=(\sum_{w\in S}x(w))-(\sum_{w\in S}x(w))=0.

Proposition 3.43.

MnGnM_{n}\rtimes G_{n} has a commutator width of at least n(p1)(p!)\frac{n(p-1)}{(p!)}.

Proof.

We know that Mn=[Mn,Gn]M_{n}=[M_{n},G_{n}]. We first investigate the diameter of MnM_{n} with respect to the generators [g,x][g,x] for gGn,xMng\in G_{n},x\in M_{n}.

Fix a positive integer kk. Suppose xMnx\in M_{n} has x=[x1,v1σ1]++[xk,vkσk]x=[x_{1},v_{1}\sigma_{1}]+\dots+[x_{k},v_{k}\sigma_{k}] for some σ1,,σkAp\sigma_{1},\dots,\sigma_{k}\in\mathrm{A}_{p}, v1,,vkVnv_{1},\dots,v_{k}\in V_{n} and x1,,xkMnx_{1},\dots,x_{k}\in M_{n}.

Let WnW_{n} be an Ap\mathrm{A}_{p}-invariant subspace of VnV_{n} orthogonal to all viv_{i}. Thus by Proposition 3.35, it has codimension at most kp!2k\frac{p!}{2}. If k<2n(p1)p!k<\frac{2n(p-1)}{p!}, then WnW_{n} is not trivial. Furthermore, since WnW_{n} is Ap\mathrm{A}_{p}-invariant, WnW_{n} is a collection of Ap\mathrm{A}_{p}-orbits.

By Lemma 3.42, we have

x=[x1,v1σ1]++[xk,vkσk]=[xivi,σi]+[xi,vi].x=[x_{1},v_{1}\sigma_{1}]+\dots+[x_{k},v_{k}\sigma_{k}]=\sum[x_{i}^{v_{i}},\sigma_{i}]+\sum[x_{i},v_{i}].

Now for each ii, we have Wn{vi}null([xi,vi])W_{n}\subseteq\{v_{i}\}^{\perp}\subseteq\mathrm{null}([x_{i},v_{i}]). Hence Wnnull([xi,vi])W_{n}\subseteq\mathrm{null}(\sum[x_{i},v_{i}]).

Furthermore, for each Ap\mathrm{A}_{p}-orbit SS of Vn{0}V_{n}-\{0\}, we have wS[xivi,σi](w)=0\sum_{w\in S}[x_{i}^{v_{i}},\sigma_{i}](w)=0.

Therefore, for each Ap\mathrm{A}_{p}-orbit SS inside of WnW_{n}, we have

wSi[xivi,σi](w)\displaystyle\sum_{w\in S}\sum_{i}[x_{i}^{v_{i}},\sigma_{i}](w) =0,\displaystyle=0,
wSi[xi,vi](w)\displaystyle\sum_{w\in S}\sum_{i}[x_{i},v_{i}](w) =0.\displaystyle=0.

So together, we see that wSx(w)=0\sum_{w\in S}x(w)=0.

In particular, let us pick any xMnx\in M_{n} such that wSx(w)0\sum_{w\in S}x(w)\neq 0 for each Ap\mathrm{A}_{p}-orbit SS of Vn{0}V_{n}-\{0\}. Then x=[x1,σ1v1]++[xk,σkvk]x=[x_{1},\sigma_{1}v_{1}]+\dots+[x_{k},\sigma_{k}v_{k}] implies that the corresponding WnW_{n} must be trivial. In particular, we must have k2n(p1)p!k\geq\frac{2n(p-1)}{p!}.

So the collection of [g,x][g,x] for gGn,xMng\in G_{n},x\in M_{n} generates MnM_{n} with a diameter of at least 2n(p1)p!\frac{2n(p-1)}{p!}. Thus the commutator width of our group is at least n(p1)(p!)\frac{n(p-1)}{(p!)} by Lemma 3.44 below. ∎

Lemma 3.44.

Let MM be an abelian group, and suppose we have a group GG acting on MM. If the group MGM\rtimes G has commutator width at most dd, then the elements [x,g][x,g] for all xM,gGx\in M,g\in G generate MM with a diameter of at most 2d2d.

Proof.

We write the group operation on MM multiplicatively for the sake of this argument. Suppose MGM\rtimes G has commutator width at most dd.

Then pick any xgMGxg\in M\rtimes G, it must be the product of dd commutators, say

xg=i=1d[xigi,xigi].xg=\prod_{i=1}^{d}[x_{i}g_{i},x^{\prime}_{i}g^{\prime}_{i}].

Then by taking all the gi,gig_{i},g^{\prime}_{i} to the right, for some ri,si,pi,qir_{i},s_{i},p_{i},q_{i} in GG, we have

xg=i=1d((xi1)ri((xi)1)sixipi(xi)qi)(i=1d[gi,gi]).xg=\prod_{i=1}^{d}((x_{i}^{-1})^{r_{i}}((x^{\prime}_{i})^{-1})^{s_{i}}x_{i}^{p_{i}}(x^{\prime}_{i})^{q_{i}})(\prod_{i=1}^{d}[g_{i},g^{\prime}_{i}]).

For the two sides to be equal, we must have

x=i=1d((xi1)ri((xi)1)sixipi(xi)qi).x=\prod_{i=1}^{d}((x_{i}^{-1})^{r_{i}}((x^{\prime}_{i})^{-1})^{s_{i}}x_{i}^{p_{i}}(x^{\prime}_{i})^{q_{i}}).

By replacing xi,xix_{i},x^{\prime}_{i} by xiri,(xi)six_{i}^{r_{i}},(x^{\prime}_{i})^{s_{i}}, and replacing pi,qip_{i},q_{i} by ri1pi,si1qir_{i}^{-1}p_{i},s_{i}^{-1}q_{i}, we have

x=i=1d(xi1(xi)1xipi(xi)qi).x=\prod_{i=1}^{d}(x_{i}^{-1}(x^{\prime}_{i})^{-1}x_{i}^{p_{i}}(x^{\prime}_{i})^{q_{i}}).

Note that MM is abelian. Then

x=i=1d(xi1(xi)1xipi(xi)qi)=i=1d([xi,pi][xi,qi]).x=\prod_{i=1}^{d}(x_{i}^{-1}(x^{\prime}_{i})^{-1}x_{i}^{p_{i}}(x^{\prime}_{i})^{q_{i}})=\prod_{i=1}^{d}([x_{i},p_{i}][x^{\prime}_{i},q_{i}]).

In particular, since xx is artibrary, MM must be generated by elements [x,g][x,g] for all xM,gGx\in M,g\in G with a diameter of at most 2d2d. ∎

In particular, as nn increases, we see that the commutator width of MnGnM_{n}\rtimes G_{n} is unbounded as desired.

3.7 Ultraproducts of MnM_{n} and MnM_{n}^{*}

Let V=nωVnV=\prod_{n\to\omega}V_{n}, M=nωMnM=\prod_{n\to\omega}M_{n}, and let M=nωMnM^{*}=\prod_{n\to\omega}M_{n}^{*}. We shall establish some properties of them to be used later.

Note that each MnM_{n} is the collection of set functions from Vn{0}V_{n}-\{0\} to BB. Hence given a sequence (xn)n(x_{n})_{n\in\mathbb{N}}, their ultralimit limnωxnM\lim_{n\to\omega}x_{n}\in M is a set function from nω(Vn{0})=V{0}\prod_{n\to\omega}(V_{n}-\{0\})=V-\{0\} to BB. So MM is canonically a submodule of BV{0}B^{V-\{0\}}.

Remark 3.45.

We do NOT have M=BV{0}M=B^{V-\{0\}}. Note that all MnM_{n} are finite, so the cardinality of their ultraproduct MM by a non-principal ultrafilter is that of the continuum. Similarly, note that all VnV_{n} are also finite, so the cardinality of VV is also that of the continuum. Then BV{0}B^{V-\{0\}} will have cardinality strictly larger than that of the continuum. In particular, MBV{0}M\neq B^{V-\{0\}}.

This means we can treat each xMx\in M as a set-function from V{0}V-\{0\} to BB. So we can make the following definition.

Definition 3.46.

For each xMx\in M, we define its null set as

null(x):={0}{vV{0}x(v)=0}.\mathrm{null}(x):=\{0\}\cup\{v\in V-\{0\}\mid x(v)=0\}.
Lemma 3.47.

If x=limnωxnx=\lim_{n\to\omega}x_{n}, then

null(x)=nωnull(xn)V.\mathrm{null}(x)=\prod_{n\to\omega}\mathrm{null}(x_{n})\subseteq V.
Proof.

Note that for each v=limnωvnV{0}v=\lim_{n\to\omega}v_{n}\in V-\{0\}, then x(v)=0x(v)=0 if and only if

{n:xn(vn)=0}ω.\{n\in\mathbb{N}:x_{n}(v_{n})=0\}\in\omega.

But note that xn(vn)=0x_{n}(v_{n})=0 implies that vnnull(xn)v_{n}\in\mathrm{null}(x_{n}), so we have

{n:vnnull(xn)}ω.\{n\in\mathbb{N}:v_{n}\in\mathrm{null}(x_{n})\}\in\omega.

So we have

vnωnull(xn).v\in\prod_{n\to\omega}\mathrm{null}(x_{n}).

Since this is true for all vV{0}v\in V-\{0\}, we have

null(x)nωnull(xn).\mathrm{null}(x)\subseteq\prod_{n\to\omega}\mathrm{null}(x_{n}).

The other direction is obvious. ∎

Proposition 3.48.

For any xMx\in M and vVv\in V, then

{v}null([x,v]).\{v\}^{\perp}\subseteq\mathrm{null}([x,v]).
Proof.

Say v=limnωvnv=\lim_{n\to\omega}v_{n} and x=limnωxnx=\lim_{n\to\omega}x_{n}. Then

[x,v]=xvx=limnω(xnvnxn)=limnω[xn,vn].[x,v]=x^{v}-x=\lim_{n\to\omega}(x_{n}^{v_{n}}-x_{n})=\lim_{n\to\omega}[x_{n},v_{n}].

Since {vn}null([xn,vn])\{v_{n}\}^{\perp}\subseteq\mathrm{null}([x_{n},v_{n}]) by Lemma 3.42, we see that

{v}=nω{vn}nωnull([xn,vn])=null([x,v]).\{v\}^{\perp}=\prod_{n\to\omega}\{v_{n}\}^{\perp}\subseteq\prod_{n\to\omega}\mathrm{null}([x_{n},v_{n}])=\mathrm{null}([x,v]).

Proposition 3.49.

For any xMx\in M and vVv\in V and wV{0}w\in V-\{0\}, then

xv(w)=fv,w(x(w)).x^{v}(w)=f^{\langle v,w\rangle}(x(w)).
Proof.

Say v=limnωvnv=\lim_{n\to\omega}v_{n}, w=limnωwnw=\lim_{n\to\omega}w_{n} and x=limnωxnx=\lim_{n\to\omega}x_{n}. Then

xv(w)=limnωxnvn(wn)=limnωfvn,wn(xn(wn)).x^{v}(w)=\lim_{n\to\omega}x_{n}^{v_{n}}(w_{n})=\lim_{n\to\omega}f^{\langle v_{n},w_{n}\rangle}(x_{n}(w_{n})).

But since we have

{n:vn,wn=v,w}\displaystyle\{n\in\mathbb{N}:\langle v_{n},w_{n}\rangle=\langle v,w\rangle\} ω,\displaystyle\in\omega,
{n:xn(wn)=x(w)}\displaystyle\{n\in\mathbb{N}:x_{n}(w_{n})=x(w)\} ω.\displaystyle\in\omega.

So their intersection is still in ω\omega, so we have

limnωfvn,wn(xn(wn))=fv,w(x(w)).\lim_{n\to\omega}f^{\langle v_{n},w_{n}\rangle}(x_{n}(w_{n}))=f^{\langle v,w\rangle}(x(w)).

Now we define the “support for ϕ\phi” as the dual concept of “null set for xx”.

Definition 3.50.

For each ϕnMn\phi_{n}\in M_{n}^{*}, we say a subset SS of VnV_{n} is a supporting set for ϕn\phi_{n} if for all xnMnx_{n}\in M_{n} such that null(xn)S\mathrm{null}(x_{n})\supseteq S, we have ϕn(xn)=0\phi_{n}(x_{n})=0. Define supp(ϕn)\mathrm{supp}(\phi_{n}) to be the intersection of all supporting sets for ϕn\phi_{n}.

Proposition 3.51.

supp(ϕn)\mathrm{supp}(\phi_{n}) is also a supporting set for all ϕnMn\phi_{n}\in M_{n}^{*}.

Proof.

First, the whole space VnV_{n} is a supporting set for ϕn\phi_{n}.

For any two supporting sets for ϕn\phi_{n}, say S1,S2S_{1},S_{2}, suppose S1S2null(x)S_{1}\cap S_{2}\subseteq\mathrm{null}(x) for some xMnx\in M_{n}.

Let x1Mnx_{1}\in M_{n} be the function that agrees with xx on S1S2S_{1}-S_{2}, but sends all other inputs to zero. Then S2null(x1)S_{2}\subseteq\mathrm{null}(x_{1}) and hence ϕn(x1)=0\phi_{n}(x_{1})=0.

Now we have

S1S2\displaystyle S_{1}\cap S_{2}\subseteq null(x),\displaystyle\mathrm{null}(x),
S1S2\displaystyle S_{1}\cap S_{2}\subseteq null(x1).\displaystyle\mathrm{null}(x_{1}).

Therefore we have

S1S2null(xx1).S_{1}\cap S_{2}\subseteq\mathrm{null}(x-x_{1}).

But we also see that

S1S2null(xx1).S_{1}-S_{2}\subseteq\mathrm{null}(x-x_{1}).

So we can conclude that S1null(xx1)S_{1}\subseteq\mathrm{null}(x-x_{1}) and hence ϕn(xx1)=0\phi_{n}(x-x_{1})=0.

Combining results above, we have

ϕn(x)=ϕn(xx1)+ϕn(x1)=0.\phi_{n}(x)=\phi_{n}(x-x_{1})+\phi_{n}(x_{1})=0.

So finite intersections of supporting sets are supporting sets.

Finally, since VnV_{n} is finite, therefore supp(ϕn)\mathrm{supp}(\phi_{n}) is a finite intersection of supporting sets for ϕn\phi_{n}, so it is a supporting set for ϕn\phi_{n}. ∎

Proposition 3.52.

For any vnVnv_{n}\in V_{n}, then vnsupp(ϕn)v_{n}\in\mathrm{supp}(\phi_{n}) if and only if there exists xnMnx_{n}\in M_{n} such that the followings are true

null(xn)\displaystyle\mathrm{null}(x_{n}) =Vn{vn},\displaystyle=V_{n}-\{v_{n}\},
ϕn(xn)\displaystyle\phi_{n}(x_{n}) 0.\displaystyle\neq 0.
Proof.

Suppose vnsupp(ϕn)v_{n}\in\mathrm{supp}(\phi_{n}). Then Vn{vn}V_{n}-\{v_{n}\} is NOT a supporting set for ϕn\phi_{n}. So there exists xnMnx_{n}\in M_{n} such that we have the following

null(xn)\displaystyle\mathrm{null}(x_{n}) Vn{vn},\displaystyle\supseteq V_{n}-\{v_{n}\},
ϕn(xn)\displaystyle\phi_{n}(x_{n}) 0.\displaystyle\neq 0.

In particular, ϕn(xn)0\phi_{n}(x_{n})\neq 0 implies that xn0x_{n}\neq 0. Therefore null(xn)Vn\mathrm{null}(x_{n})\neq V_{n}, and we must in fact have null(xn)=Vn{vn}\mathrm{null}(x_{n})=V_{n}-\{v_{n}\} as desired.

Conversely, suppose vnsupp(ϕn)v_{n}\notin\mathrm{supp}(\phi_{n}). Then for all xnMnx_{n}\in M_{n} with null(xn)=Vn{vn}\mathrm{null}(x_{n})=V_{n}-\{v_{n}\}, we have

supp(ϕn)null(xn).\mathrm{supp}(\phi_{n})\subseteq\mathrm{null}(x_{n}).

And thus

ϕn(xn)=0.\phi_{n}(x_{n})=0.

Now consider ϕ=limnωϕnM\phi=\lim_{n\to\omega}\phi_{n}\in M^{*}. We make the following definition.

Definition 3.53.

We say a subset SS of VV is a supporting set for ϕ\phi if ϕ(x)=0\phi(x)=0 whenever null(x)S\mathrm{null}(x)\supseteq S. Define supp(ϕ)\mathrm{supp}(\phi) to be the intersection of all supporting sets for ϕ\phi.

Proposition 3.54.

If ϕ=limnωϕn\phi=\lim_{n\to\omega}\phi_{n}, then supp(ϕ)\mathrm{supp}(\phi) is a supporting set for ϕ\phi and is equal to nωsupp(ϕn)\prod_{n\to\omega}\mathrm{supp}(\phi_{n}).

Proof.

For x=limnωxnMx=\lim_{n\to\omega}x_{n}\in M, suppose

nωsupp(ϕn)null(x)\prod_{n\to\omega}\mathrm{supp}(\phi_{n})\subseteq\mathrm{null}(x)

.

Then by Proposition 3.47, we have

nωsupp(ϕn)nωnull(xn).\prod_{n\to\omega}\mathrm{supp}(\phi_{n})\subseteq\prod_{n\to\omega}\mathrm{null}(x_{n}).

Then for ω\omega-almost all nn, we have

supp(ϕn)null(xn).\mathrm{supp}(\phi_{n})\subseteq\mathrm{null}(x_{n}).

Thus ϕn(xn)=0\phi_{n}(x_{n})=0 for ω\omega-almost all nn. So ϕ(x)=0\phi(x)=0.

So the set nωsupp(ϕn)\prod_{n\to\omega}\mathrm{supp}(\phi_{n}) is a supporting set for ϕ\phi.

Conversely, for each vnsupp(ϕn)v_{n}\in\mathrm{supp}(\phi_{n}), then there exists xnMnx_{n}\in M_{n} such that

null(xn)\displaystyle\mathrm{null}(x_{n}) Vn{vn},\displaystyle\supseteq V_{n}-\{v_{n}\},
ϕn(xn)\displaystyle\phi_{n}(x_{n}) 0.\displaystyle\neq 0.

Then for v=limnωvnnωsupp(ϕn)v=\lim_{n\to\omega}v_{n}\in\prod_{n\to\omega}\mathrm{supp}(\phi_{n}) and x=limnωxnx=\lim_{n\to\omega}x_{n}, we have

null(x)\displaystyle\mathrm{null}(x) V{v},\displaystyle\supseteq V-\{v\},
ϕ(x)\displaystyle\phi(x) 0.\displaystyle\neq 0.

So NO subsets of V{v}V-\{v\} can be a supporting set for ϕ\phi. Hence any supporting set for ϕ\phi must contain vv.

Since we started with arbitrary vnsupp(ϕn)v_{n}\in\mathrm{supp}(\phi_{n}), the statement above is true for arbitrary vnωsupp(ϕn)v\in\prod_{n\to\omega}\mathrm{supp}(\phi_{n}). As a result, nωsupp(ϕn)\prod_{n\to\omega}\mathrm{supp}(\phi_{n}) is the unique smallest supporting set for ϕ\phi. ∎

3.8 The existence of 𝒰\mathcal{U} and ψ\psi

We now construct the homomorphism ψ:nωMn/m\psi:\prod_{n\to\omega}M_{n}\to\mathbb{Z}/m\mathbb{Z} by picking the right ultrafilter 𝒰\mathcal{U} on MM^{*}, and set ψ\psi to be the 𝒰\mathcal{U}-consensus homomorphism ϕ𝒰\phi_{\mathcal{U}}. To facilitate computations, we shall first pick out bases for all the free /m\mathbb{Z}/m\mathbb{Z}-modules involved.

For BB, set b1:=(1,1,0,,0)b_{1}:=(1,-1,0,\dots,0), and set bi+1:=f(bi)b_{i+1}:=f(b_{i}) for each i/qi\in\mathbb{Z}/q\mathbb{Z}. Then we know that b1,,bq1b_{1},\dots,b_{q-1} form a basis. We fix this basis for BB for this section.

For each vVn{0}v\in V_{n}-\{0\} and each integer 1iq11\leq i\leq q-1, let xviMnx_{v\to i}\in M_{n} be the function that sends vv to bib_{i} and all other inputs to zero.

Proposition 3.55.

The elements xvix_{v\to i} for each vVnv\in V_{n} and each integer 1iq11\leq i\leq q-1 form a basis for MnM_{n}.

Proof.

Straightforward verification. ∎

We shall use the basis xvix_{v\to i} as a standard basis for MnM_{n}. For each vVn{0}v\in V_{n}-\{0\}, let ϕv:Mn/m\phi_{v}:M_{n}\to\mathbb{Z}/m\mathbb{Z} such that xv1x_{v\to 1} is sent to 11, and all other standard basis elements xwix_{w\to i} for (w,i)(v,1)(w,i)\neq(v,1) are sent to zero.

Lemma 3.56.

ϕv(xσ)=ϕvσ(x)\phi_{v}(x^{\sigma})=\phi_{v^{\sigma}}(x) for all xMnx\in M_{n}, vVn{0}v\in V_{n}-\{0\}, and σAp\sigma\in\mathrm{A}_{p}.

Proof.

Note that (xvi)(σ1)=xvσi(x_{v\to i})^{(\sigma^{-1})}=x_{v^{\sigma}\to i} by definition. So the xv1x_{v\to 1}-component of xσx^{\sigma} is the same as the xvσ1x_{v^{\sigma}\to 1}-component of xx. So ϕv(xσ)=ϕvσ(x)\phi_{v}(x^{\sigma})=\phi_{v^{\sigma}}(x). ∎

Proposition 3.57.

For any vVn{0}v\in V_{n}-\{0\}, we have

  1. 1.

    null(xvi)=Vn{v}\mathrm{null}(x_{v\to i})=V_{n}-\{v\}.

  2. 2.

    supp(ϕv)={v}\mathrm{supp}(\phi_{v})=\{v\}.

  3. 3.

    For any homomorphism ϕ:Mn/m\phi:M_{n}\to\mathbb{Z}/m\mathbb{Z}, then vsupp(ϕ)v\in\mathrm{supp}(\phi) if and only if ϕ(xvi)0\phi(x_{v\to i})\neq 0 for some ii.

Proof.

We have null(xvi)=Vn{v}\mathrm{null}(x_{v\to i})=V_{n}-\{v\} by definition.

For the second statement, note that ϕv(xv1)=1\phi_{v}(x_{v\to 1})=1, so vsupp(ϕv)v\in\mathrm{supp}(\phi_{v}). And for any xMnx\in M_{n} with {v}null(x)\{v\}\subseteq\mathrm{null}(x), then all xvix_{v\to i}-components of xx are zero. So ϕv(x)=0\phi_{v}(x)=0. So {v}\{v\} is the smallest supporting set for ϕv\phi_{v}.

For the last statement, if ϕ(xvi)0\phi(x_{v\to i})\neq 0 for some ii, note that the null set for xvix_{v\to i} is Vn{v}V_{n}-\{v\}. Then all subsets of Vn{v}V_{n}-\{v\} cannot be a supporting set for ϕ\phi. So we must have vsupp(ϕ)v\in\mathrm{supp}(\phi).

Conversely, if vsupp(ϕ)v\in\mathrm{supp}(\phi), then by Proposition 3.52, for some xMnx\in M_{n} with null(x)=Vn{v}\mathrm{null}(x)=V_{n}-\{v\}, we have ϕ(x)0\phi(x)\neq 0. But the identity null(x)=Vn{v}\mathrm{null}(x)=V_{n}-\{v\} implies that xx must be a linear combination of xv1,,xvq1x_{v\to 1},\dots,x_{v\to q-1}, and thus ϕ(xvi)0\phi(x_{v\to i})\neq 0 for some xvix_{v\to i}. ∎

Now we start to build the homomorphism ψ\psi. We want to guarantee that ψ\psi is surjective, so it must some specific element of MM to the generator 11 of /m\mathbb{Z}/m\mathbb{Z}.

We define znz_{n} to be the element in MnM_{n} such that

zn:=vVn{0}xv1.z_{n}:=\sum_{v\in V_{n}-\{0\}}x_{v\to 1}.

We then take an ultralimit and set

z:=limnznM=nωMn.z:=\lim_{n\in\infty}z_{n}\in M=\prod_{n\to\omega}M_{n}.

Our goal is to construct ψ:=ϕ𝒰\psi:=\phi_{\mathcal{U}} such that it sends zz to 11, and it is GG-invariant. Let us now reduce this to the task of finding a nice ultrafilter 𝒰\mathcal{U} on MM^{*}.

For any closed subspace WVW\subseteq V, let MWM_{W}^{*} be the collection of all ϕM\phi\in M^{*} such that supp(ϕ)W\mathrm{supp}(\phi)\subseteq W.

Let EE^{*} be the collection of all ϕM\phi\in M^{*} such that ϕ(z)=1\phi(z)=1, and ϕ(x)=ϕ(xσ)\phi(x)=\phi(x^{\sigma}) for all xMx\in M and σAp\sigma\in\mathrm{A}_{p}.

Proposition 3.58.

Suppose 𝒰\mathcal{U} is an ultrafilter on MM^{*} such that MW𝒰M_{W}^{*}\in\mathcal{U} for all closed subspace WW of finite codimension in VV, and E𝒰E^{*}\in\mathcal{U}. Then 𝒰\mathcal{U} is GG-invariant and ϕ𝒰:M/m\phi_{\mathcal{U}}:M\to\mathbb{Z}/m\mathbb{Z} is surjective and GG-invariant.

Proof.

Note that for all ϕE\phi\in E^{*}, ϕ(z)=1\phi(z)=1. Since E𝒰E^{*}\in\mathcal{U}, we conclude that ϕ𝒰(z)=1\phi_{\mathcal{U}}(z)=1. So the image of ϕ𝒰\phi_{\mathcal{U}} contains a generator of /m\mathbb{Z}/m\mathbb{Z}, and thus ϕ𝒰\phi_{\mathcal{U}} is surjective.

For any gG=VApg\in G=V\rtimes\mathrm{A}_{p}, say g=vσg=v\sigma. Let SgS_{g} be the set of all gg-invariant elements ϕM\phi\in M^{*}. Let us show that Sg𝒰S_{g}\in\mathcal{U}.

Set W={v}W=\{v\}^{\perp}, and consider any ϕEMW\phi\in E^{*}\cap M_{W}^{*}. Then since ϕ\phi is in EE^{*}, it is σ\sigma-invariant.

Since ϕ\phi is in MWM_{W}^{*}, for any xMx\in M, we have

supp(ϕ)Wnull([x,v]).\mathrm{supp}(\phi)\subseteq W\subseteq\mathrm{null}([x,v]).

Thus we have

0=ϕ([x,v])=ϕ(xvx).0=\phi([x,v])=\phi(x^{v}-x).

Therefore ϕ(x)=ϕ(xv)\phi(x)=\phi(x^{v}). Hence ϕ\phi is vv-invariant.

In conclusion, ϕ\phi is both σ\sigma-invariant and vv-invariant, therefore it is gg-invariant. So EMWSgE^{*}\cap M_{W}^{*}\subseteq S_{g}. But since EMW𝒰E^{*}\cap M_{W}^{*}\in\mathcal{U}, therefore we see that Sg𝒰S_{g}\in\mathcal{U}.

So 𝒰\mathcal{U} is GG-invariant, and consequently, ϕ𝒰\phi_{\mathcal{U}} is also GG-invariant. ∎

In particular, the proof of Lemma 3.19 depends entirely on the existence of an ultrafilter containing EE^{*} and MW𝒰M_{W}^{*}\in\mathcal{U} for all closed subspace WW of finite codimension in VV. So we need to show that EE^{*} and these MWM_{W}^{*} have the finite intersection property.

Lemma 3.59.

Let WnW_{n} be an Ap\mathrm{A}_{p}-invariant subspace of VnV_{n} of codimension cc, such that n>c(p1)n>c(p-1). Then there exists ϕMn\phi\in M_{n}^{*} such that for all xMnx\in M_{n} and σAp\sigma\in\mathrm{A}_{p}, we have

ϕ(zn)\displaystyle\phi(z_{n}) =1,\displaystyle=1,
supp(ϕ)\displaystyle\mathrm{supp}(\phi) Wn,\displaystyle\subseteq W_{n},
ϕ(x)\displaystyle\phi(x) =ϕ(xσ).\displaystyle=\phi(x^{\sigma}).
Proof.

Since n>c(p1)n>c(p-1), by Lemma 3.36, we can find wWnw\in W_{n} whose Ap\mathrm{A}_{p}-orbit 𝒪(w)\mathcal{O}(w) has exactly pp elements. Let ϕ=v𝒪(w)ϕv\phi=\sum_{v\in\mathcal{O}(w)}\phi_{v}.

Since WnW_{n} is Ap\mathrm{A}_{p}-invariant, we have 𝒪(w)Wn\mathcal{O}(w)\subseteq W_{n}, and thus

supp(ϕ)=𝒪(w)Wn.\mathrm{supp}(\phi)=\mathcal{O}(w)\subseteq W_{n}.

By Lemma 3.56, for all xMnx\in M_{n}, vVn{0}v\in V_{n}-\{0\}, and σAp\sigma\in\mathrm{A}_{p}, we have

ϕv(xσ)=ϕvσ(x).\phi_{v}(x^{\sigma})=\phi_{v^{\sigma}}(x).

So we see that for all xMnx\in M_{n} and σAp\sigma\in\mathrm{A}_{p}, we have

ϕ(x)=v𝒪(w)ϕv(x)=v𝒪(w)ϕvσ(x)=v𝒪(w)ϕv(xσ)=ϕ(xσ).\phi(x)=\sum_{v\in\mathcal{O}(w)}\phi_{v}(x)=\sum_{v\in\mathcal{O}(w)}\phi_{v^{\sigma}}(x)=\sum_{v\in\mathcal{O}(w)}\phi_{v}(x^{\sigma})=\phi(x^{\sigma}).

Finally, note that ϕv(zn)=1\phi_{v}(z_{n})=1 for all vVn{0}v\in V_{n}-\{0\}. So we have

ϕ(zn)=|𝒪(w)|=p.\phi(z_{n})=|\mathcal{O}(w)|=p.

But note that p,mp,m are coprime, so pp is invertible in /m\mathbb{Z}/m\mathbb{Z}. So by replacing ϕ\phi with a suitable multiple, we can guarantee that ϕ(zn)=1\phi(z_{n})=1. ∎

Proposition 3.60.

For any closed subspace WW of finite codimension in VV, then MWEM_{W}^{*}\cap E^{*} is non-empty.

Proof.

Note that if WW is closed in VV with finite codimension, then by Proposition 3.35, we can find WWW^{\prime}\subseteq W which is closed, Ap\mathrm{A}_{p}-invariant, and has finite codimension in VV. We also necessarily have MWMWM_{W^{\prime}}^{*}\subseteq M_{W}^{*}. So, replacing WW by the smaller subspace WW^{\prime}, we can assume that WW is also Ap\mathrm{A}_{p}-invariant.

By Proposition 3.34, since WW is Ap\mathrm{A}_{p}-invariant and closed with finite codimension in VV, therefore W=nωWnW=\prod_{n\to\omega}W_{n} where each WnW_{n} is Ap\mathrm{A}_{p}-invariant and closed with bounded codimension in VnV_{n} for all nn. Say the codimension are all at most cc. Then for all n>c(p1)n>c(p-1), we can find ϕnMn\phi_{n}\in M_{n}^{*} such that for all xMnx\in M_{n} and σAp\sigma\in\mathrm{A}_{p}, we have

ϕn(zn)\displaystyle\phi_{n}(z_{n}) =1,\displaystyle=1,
supp(ϕn)\displaystyle\mathrm{supp}(\phi_{n}) Wn,\displaystyle\subseteq W_{n},
ϕn(x)\displaystyle\phi_{n}(x) =ϕn(xσ).\displaystyle=\phi_{n}(x^{\sigma}).

Then since ω\omega is not principal, we only need to consider the cases where n>c(p1)n>c(p-1) as above. So this defines ϕ=limnωϕnM\phi=\lim_{n\to\omega}\phi_{n}\in M^{*} such that for all x=limnωxnMx=\lim_{n\to\omega}x_{n}\in M and σAp\sigma\in\mathrm{A}_{p},

ϕ(z)\displaystyle\phi(z) =limnωϕn(zn)=1,\displaystyle=\lim_{n\to\omega}\phi_{n}(z_{n})=1,
supp(ϕ)\displaystyle\mathrm{supp}(\phi) =nωsupp(ϕn)nωWn=W,\displaystyle=\prod_{n\to\omega}\mathrm{supp}(\phi_{n})\subseteq\prod_{n\to\omega}W_{n}=W,
ϕ(x)\displaystyle\phi(x) =limnωϕn(xn)=limnωϕn(xnσ)=ϕ(xσ).\displaystyle=\lim_{n\to\omega}\phi_{n}(x_{n})=\lim_{n\to\omega}\phi_{n}(x_{n}^{\sigma})=\phi(x^{\sigma}).

So ϕMWE\phi\in M_{W}^{*}\cap E^{*}. So we are done. ∎

Proposition 3.61.

There is an ultrafilter 𝒰\mathcal{U} on MM^{*} such that MW𝒰M_{W}^{*}\in\mathcal{U} for all closed subspace WW of finite codimension in VV, and E𝒰E^{*}\in\mathcal{U}.

Proof.

We only need to show that these subsets of MM^{*} have finite intersection property. Pick W1,,WkW_{1},\dots,W_{k} closed subspaces of finite codimension in VV, we want to show that E(i=1kMWi)E^{*}\cap(\bigcap_{i=1}^{k}M_{W_{i}}^{*}) is non-empty.

Let W=i=1kWiW=\bigcap_{i=1}^{k}W_{i}. This is a finite intersection of closed subspaces of finite codimensions, so it is itself closed with finite codimension.

Note that

ϕi=1kMWi if and only if\displaystyle\phi\in\bigcap_{i=1}^{k}M_{W_{i}}^{*}\text{ if and only if } supp(ϕ)Wi for all i,\displaystyle\mathrm{supp}(\phi)\subseteq W_{i}\text{ for all }i,
if and only if supp(ϕ)i=1kWi=W,\displaystyle\mathrm{supp}(\phi)\subseteq\bigcap_{i=1}^{k}W_{i}=W,
if and only if ϕMW.\displaystyle\phi\in M_{W}^{*}.

So we have i=1kMWi=MW\bigcap_{i=1}^{k}M_{W_{i}}^{*}=M_{W}^{*}.

So we only need to show that EMWE^{*}\cap M_{W}^{*} is non-empty. But since WW is closed with finite codimension, this is non-empty by Proposition 3.60. So we have the desired finite intersection property, and the desired ultrafilter exists. ∎

We now finish the proof of Lemma 3.19.

Proof of Lemma 3.19.

Let 𝒰\mathcal{U} be the ultrafilter in Proposition 3.61. Then it satisfies all the requirements of Proposition 3.58, so ϕ𝒰\phi_{\mathcal{U}} is surjective and GG-invariant.

Furthermore, by Proposition 3.40, the groups MnGnM_{n}\rtimes G_{n} are finite perfect groups as desired. ∎

References

  • Asc [00] Michael Aschbacher. Finite group theory, volume 10. Cambridge University Press, 2000.
  • BN [11] George M Bergman and Nazih Nahlus. Homomorphisms on infinite direct product algebras, especially lie algebras. Journal of Algebra, 333(1):67–104, 2011.
  • CLP [15] Valerio Capraro, Martino Lupini, and Vladimir Pestov. Introduction to sofic and hyperlinear groups and Connes’ embedding conjecture, volume 1. Springer, 2015.
  • ES [04] Gábor Elek and Endre Szabó. Sofic groups and direct finiteness. J. Algebra, 280(2):426–434, 2004.
  • HP [89] Derek F Holt and W Plesken. Perfect groups. Oxford, 1989.
  • HR [17] Derek Holt and Sarah Rees. Some closure results for 𝒞\mathcal{C}-approximable groups. Pacific Journal of Mathematics, 287(2):393–409, 2017.
  • Kul [45] L Ya Kulikov. On the theory of abelian groups of arbitrary cardinality. Mat. Sb, 16(2):129–162, 1945.
  • Lam [12] Tsit-Yuen Lam. Lectures on modules and rings, volume 189. Springer Science & Business Media, 2012.
  • MR [19] Peter Mayr and Nik Ruškuc. Generating subdirect products. Journal of the London Mathematical Society, 100(2):404–424, 2019.
  • Nik [04] Nikolay Nikolov. On the commutator width of perfect groups. Bulletin of the London Mathematical Society, 36(1):30–36, 2004.
  • NST [18] Nikolay Nikolov, Jakob Schneider, and Andreas Thom. Some remarks on finitarily approximable groups. Journal de l’École polytechnique-Mathématiques, 5:239–258, 2018.
  • RZ [00] Luis Ribes and Pavel Zalesskii. Profinite groups. In Profinite Groups, pages 19–77. Springer, 2000.
  • SB [81] Hanamantagouda P Sankappanavar and Stanley Burris. A course in universal algebra, volume 78. Citeseer, 1981.
  • Tho [12] Andreas Thom. About the metric approximation of Higman’s group. 2012.
  • Yan [16] Yilong Yang. Ultraproducts of quasirandom groups with small cosocles. Journal of Group Theory, 19(6):1137–1164, 2016.
  • Zel [90] Efim Isaakovich Zel’manov. Solution of the restricted Burnside problem for groups of odd exponent. Izvestiya Rossiiskoi Akademii Nauk. Seriya Matematicheskaya, 54(1):42–59, 1990.
  • Zel [91] Efim Isaakovich Zel’manov. A solution of the restricted Burnside problem for 2-groups. Matematicheskii Sbornik, 182(4):568–592, 1991.
  • Zil [14] Boris Zilber. Perfect infinities and finite approximation. In Infinity and truth, pages 199–223. World Scientific, 2014.