Hostname: page-component-848d4c4894-4hhp2 Total loading time: 0 Render date: 2024-05-04T08:34:58.540Z Has data issue: false hasContentIssue false

Kernels of localities

Published online by Cambridge University Press:  02 September 2022

Valentina Grazian
Affiliation:
Department of Mathematics and Applications, University of Milano - Bicocca, Via Roberto Cozzi 55, 20125 Milano, Italy; E-mail: valentina.grazian@unimib.it
Ellen Henke
Affiliation:
Institut für Algebra, Fakultät Mathematik, Technische Universität Dresden, 01062 Dresden, Germany; E-mail: ellen.henke@tu-dresden.de

Abstract

We state a sufficient condition for a fusion system to be saturated. This is then used to investigate localities with kernels: that is, localities that are (in a particular way) extensions of groups by localities. As an application of these results, we define and study certain products in fusion systems and localities, thus giving a new method to construct saturated subsystems of fusion systems.

Type
Algebra
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2022. Published by Cambridge University Press

1 Introduction

The problem of showing that a given fusion system is saturated arises in different contexts. For example, proving that certain subsystems of fusion systems are saturated is one of the major difficulties in developing a theory of saturated fusion systems in analogy to the theory of finite groups. When studying extensions of fusion systems, it is also crucial to understand under which conditions such extensions are saturated. In this paper, we seek to take up both themes simultaneously in a systematic way. To formulate and study extension problems, it is common to work with linking systems or, more generally, with transporter systems associated to fusion systems (see, e.g., [Reference Broto, Castellana, Grodal, Levi and Oliver6, Reference Oliver and Ventura25, Reference Oliver23]). The equivalent framework of localities (introduced by Chermak [Reference Chermak7]) was recently used by Chermak and the second named author of this paper to construct saturated subsystems of fusion systems, thereby enriching the theory of fusion systems by some new concepts (compare [Reference Chermak and Henke10, Theorem C] and [Reference Henke20]).

In the present paper (apart from the appendix), we work with localities rather than transporter systems. Our first main result is, however, formulated purely in terms of fusion systems. It gives a sufficient condition for a fusion system to be saturated. The proof generalises an argument used by Oliver [Reference Oliver23] to show that the fusion systems associated to certain extensions of groups by linking systems are saturated.

Our saturation criterion serves us as an important tool for studying kernels of localities as introduced in Definition 4 below. A kernel of a locality $\mathcal {L}$ is basically a partial normal subgroup $\mathcal {N}$ such that the factor locality $\mathcal {L}/\mathcal {N}$ is a group and $\mathcal {N}$ itself supports the structure of a locality. We show in an appendix that kernels of localities correspond to ‘normal pairs of transporter systems’ (compare Definition A.3).

In Section 6, we prove some results demonstrating that the theory of kernels can be used to construct saturated subsystems of fusion systems. More precisely, we study certain products in localities that give rise to ‘sublocalities’ whose fusion systems are saturated. As a special case, if $\mathcal {F}$ is a saturated fusion system over S, one can define a notion of a product of a normal subsystem with a subgroup of a model for $N_{\mathcal {F}}(S)$ (or equivalently with a saturated subsystem of $N_{\mathcal {F}}(S)$ ). In particular, our work generalises the notion of a product of a normal subsystem with a subgroup of S, which was introduced by Aschbacher [Reference Aschbacher3]. Our results on products are also used by the second named author of this article [Reference Henke19] to construct normalisers and centralisers of subnormal subsystems of saturated fusion systems.

For the remainder of this introduction, let $\mathcal {F}$ be a fusion system over a finite p -group S.

We will adapt the terminology and notation regarding fusion systems from [Reference Aschbacher, Kessar and Oliver2, Chapter 1], except that we will write homomorphisms on the right-hand side of the argument (similarly as in [Reference Aschbacher, Kessar and Oliver2, Chapter 2]) and that we will define centric radical subgroups of $\mathcal {F}$ differently, namely as follows.

Definition 1. Define a subgroup $P\leq S$ to be centric radical in $\mathcal {F}$ if

  • P is centric: that is, $C_S(Q)\leq Q$ for every $\mathcal {F}$ -conjugate Q of P; and

  • $O_p(N_{\mathcal {F}}(Q))=Q$ for every fully $\mathcal {F}$ -normalised $\mathcal {F}$ -conjugate Q of P.

Write $\mathcal {F}^{cr}$ for the set of subgroups of S that are centric radical in $\mathcal {F}$ .

If $\mathcal {F}$ is saturated, then our notion of centric radical subgroups of $\mathcal {F}$ coincides with the usual notion (compare Lemma 2.6). However, defining centric radical subgroups as above is crucial if we want to conclude that the fusion systems of certain localities are saturated.

1.1 A saturation criterion

The following definition will be used to formulate the previously mentioned sufficient condition for a fusion system to be saturated.

Definition 2. Let $\Delta $ be a set of subgroups of S.

  • The set $\Delta $ is called $\mathcal {F}$ -closed if $\Delta $ is closed under $\mathcal {F}$ -conjugacy and overgroup-closed in S.

  • $\mathcal {F}$ is called $\Delta $ -generated if every morphism in $\mathcal {F}$ can be written as a product of restrictions of $\mathcal {F}$ -morphisms between subgroups in $\Delta $ .

  • $\mathcal {F}$ is called $\Delta $ -saturated if each $\mathcal {F}$ -conjugacy class in $\Delta $ contains a subgroup that is fully automised and receptive in $\mathcal {F}$ (as defined in [Reference Aschbacher, Kessar and Oliver2, Definition I.2.2]).

Generalising arguments used by Oliver [Reference Oliver23], we prove the following theorem.

Theorem A. Let $\mathcal {E}$ be an $\mathcal {F}$ -invariant saturated subsystem of $\mathcal {F}$ . Suppose $\mathcal {F}$ is $\Delta $ -generated and $\Delta $ -saturated for some $\mathcal {F}$ -closed set $\Delta $ of subgroups of S with $\mathcal {E}^{cr}\subseteq \Delta $ . Then $\mathcal {F}$ is saturated.

1.2 Kernels of localities

The reader is referred to Section 3 for an introduction to partial groups and localities. We say that a locality $(\mathcal {L},\Delta ,S)$ is a locality over $\mathcal {F}$ to indicate that $\mathcal {F}=\mathcal {F}_S(\mathcal {L})$ . The set $\Delta $ is called the object set of $(\mathcal {L},\Delta ,S)$ . It follows from the definition of a locality that the normaliser $N_{\mathcal {L}}(P)$ of any object $P\in \Delta $ is a subgroup of $\mathcal {L}$ and thus a finite group. We will use the following definition.

Definition 3.

  • A locality $(\mathcal {L},\Delta ,S)$ over $\mathcal {F}$ is called cr-complete if $\mathcal {F}^{cr}\subseteq \Delta $ .

  • A finite group G is said to be of characteristic p if $C_G(O_p(G))\leq O_p(G)$ .

  • A locality $(\mathcal {L},\Delta ,S)$ is called a linking locality if it is cr-complete and $N_{\mathcal {L}}(P)$ is of characteristic p for every $P\in \Delta $ .

The slightly nonstandard notion of centric radical subgroups introduced in Definition 1 ensures that the fusion system $\mathcal {F}_S(\mathcal {L})$ of a cr-complete locality $(\mathcal {L},\Delta ,S)$ is saturated (compare Proposition 3.18(c)). If $(\mathcal {L},\Delta ,S)$ is a cr-complete locality over $\mathcal {F}$ , then it gives rise to a transporter system $\mathcal {T}$ associated to $\mathcal {F}$ whose object set is $\Delta $ and thus contains the set $\mathcal {F}^{cr}$ . It follows from [Reference Oliver and Ventura25, Proposition 4.6] that the p-completed nerve of such a transporter system $\mathcal {T}$ is homotopy equivalent to the p-completed nerve of a linking system associated to $\mathcal {F}$ . The results we present next are centred around the following concept.

Definition 4. A kernel of a locality $(\mathcal {L},\Delta ,S)$ is a partial normal subgroup $\mathcal {N}$ of $\mathcal {L}$ such that $P\cap \mathcal {N}\in \Delta $ for every $P\in \Delta $ .

We show in Appendix A that kernels of localities correspond to ‘normal pairs of transporter systems’. In particular, the results presented below can be translated to results on transporter systems. The reader is referred to Definition A.3, Proposition A.4, Theorem A.7 and Remark A.8 for details.

If $\mathcal {N}$ is a kernel of a locality $(\mathcal {L},\Delta ,S)$ , then, setting

$$\begin{align*}T:=\mathcal{N}\cap S\mbox{ and }\Gamma:=\{P\cap\mathcal{N}\colon P\in\Delta\}, \end{align*}$$

it is easy to see that $(\mathcal {N},\Gamma ,T)$ is a locality (compare Lemma 5.2). We also say in this situation that $(\mathcal {N},\Gamma ,T)$ is a kernel of $(\mathcal {L},\Delta ,S)$ .

Suppose now that $(\mathcal {N},\Gamma ,T)$ is a kernel of $(\mathcal {L},\Delta ,S)$ . Observe that T is an element of $\Gamma \subseteq \Delta $ , so $N_{\mathcal {L}}(T)$ is a subgroup of $\mathcal {L}$ . It follows therefore from [Reference Chermak8, Theorem 4.3(b), Corollary 4.5] that $\mathcal {L}/\mathcal {N}\cong N_{\mathcal {L}}(T)/N_{\mathcal {N}}(T)$ is a group. Thus, $\mathcal {L}$ can be seen as an extension of the group $\mathcal {L}/\mathcal {N}$ by the locality $(\mathcal {N},\Gamma ,T)$ .

If the kernel $(\mathcal {N},\Gamma ,T)$ is cr-complete, then the following theorem implies that $\mathcal {F}_S(\mathcal {L})$ is saturated. Its proof uses Theorem A.

Theorem B. Let $(\mathcal {N},\Gamma ,T)$ be a kernel of a locality $(\mathcal {L},\Delta ,S)$ . Then $(\mathcal {L},\Delta ,S)$ is cr-complete if and only if $(\mathcal {N},\Gamma ,T)$ is cr-complete. If so, then $\mathcal {F}_T(\mathcal {N})$ is a normal subsystem of $\mathcal {F}_S(\mathcal {L})$ .

Theorem C. Let $(\mathcal {L},\Delta ,S)$ be a locality with a kernel $(\mathcal {N},\Gamma ,T)$ . Then the following conditions are equivalent:

  1. (i) $(\mathcal {L},\Delta ,S)$ is a linking locality;

  2. (ii) $(\mathcal {N},\Gamma ,T)$ is a linking locality and $N_{\mathcal {L}}(T)$ is of characteristic p;

  3. (iii) $(\mathcal {N},\Gamma ,T)$ is a linking locality and $C_{\mathcal {L}}(T)$ is of characteristic p.

We now want to consider special kinds of linking localities. The object set of any linking locality over $\mathcal {F}$ is always contained in the set $\mathcal {F}^s$ of $\mathcal {F}$ -subcentric subgroups (defined in Definition 3.19). If $\mathcal {F}$ is saturated, then the existence and uniqueness of centric linking systems imply conversely that for every $\mathcal {F}$ -closed set $\Delta $ of subgroups of $\mathcal {F}^s$ , there is an essentially unique linking locality over $\mathcal {F}$ with object set $\Delta $ . Chermak introduced an $\mathcal {F}$ -closed set $\delta (\mathcal {F})\subseteq \mathcal {F}^s$ , which by [Reference Chermak and Henke10, Lemma 7.21] can be described as the set of all subgroups of S containing an element of $F^*(\mathcal {F})^s$ (where $F^*(\mathcal {F})$ is the generalised Fitting subsystem of $\mathcal {F}$ introduced by Aschbacher [Reference Aschbacher3]). Notice that there always exists an essentially unique linking locality over $\mathcal {F}$ whose object set is the set $\delta (\mathcal {F})$ . Such a linking locality is called a regular locality.

For an arbitrary locality $(\mathcal {L},\Delta ,S)$ , there is a largest subgroup R of S with $\mathcal {L}=N_{\mathcal {L}}(R)$ . This subgroup is denoted by $O_p(\mathcal {L})$ . Setting $\tilde {\Delta }:=\{P\leq S\colon PO_p(\mathcal {L})\in \Delta \}$ , the triple $(\mathcal {L},\tilde {\Delta },S)$ is also a locality (but with a possibly larger object set). It turns out that $(\mathcal {L},\Delta ,S)$ is a linking locality if and only if $(\mathcal {L},\tilde {\Delta },S)$ is a linking locality (compare [Reference Chermak and Henke10, Lemma 3.28]). This flexibility in the choice of object sets makes it possible to formulate a result similar to Theorem C for regular localities.

Theorem D. Let $(\mathcal {L},\Delta ,S)$ be a locality with a kernel $(\mathcal {N},\Gamma ,T)$ . Set

$$\begin{align*}\tilde{\Delta}:=\{P\leq S\colon PO_p(\mathcal{L})\in\Delta\}\mbox{ and }\tilde{\Gamma}:=\{Q\leq T\colon QO_p(\mathcal{N})\in\Gamma\}.\end{align*}$$

Then the following conditions are equivalent:

  1. (i) $(\mathcal {L},\tilde {\Delta },S)$ is a regular locality;

  2. (ii) $(\mathcal {N},\tilde {\Gamma },T)$ is a regular locality and $N_{\mathcal {L}}(T)$ is of characteristic p;

  3. (iii) $(\mathcal {N},\tilde {\Gamma },T)$ is a regular locality and $C_{\mathcal {L}}(T)$ is of characteristic p.

Moreover, if these conditions hold, then $E(\mathcal {L})=E(\mathcal {N})$ .

In an unpublished preprint, Chermak defined a locality $(\mathcal {L},\Delta ,S)$ to be semiregular if (in our language) it has a kernel $(\mathcal {N},\Gamma ,T)$ that is a regular locality. He observed furthermore that a locality is semiregular if and only if it is an image of a regular locality under a projection of localities. As a consequence, images of semiregular localities under projections are semiregular. Moreover, since partial normal subgroups of regular localities form regular localities, it follows that partial normal subgroups of semiregular localities form semiregular localities. Thus, the category of semiregular localities and projections might provide a good framework to study extensions. This is one of our motivations to study kernels of localities more generally.

Remark. Extensions of partial groups and localities have already been studied by Gonzalez [Reference Gonzalez12]. He starts by giving important insights into the existence of extensions of partial groups. Basically, Gonzalez considers partial groups as simplicial sets and uses the concept of a simplicial fibre bundle. Gonzalez then states some results about extensions of localities in Section 7 of his paper. He calls a locality ‘saturated’ if it is cr-complete in our sense. Under certain conditions, it is shown that extensions of localities lead to (saturated) localities. To summarise, Gonzalez starts by defining isotypical extensions (compare [Reference Gonzalez12, Definition 7.1]) and shows that an isotypical extension of a locality $(\mathtt {L}",\Delta ",S")$ by a locality $(\mathtt {L}',\Delta ',S')$ leads to a locality $(\mathtt {T},\Delta ,S)$ (compare [Reference Gonzalez12, Proposition 7.6]). Slightly more precisely, we have $\mathtt {T}\subseteq \mathtt {L}$ for an extension $\mathtt {L}$ of the partial group $\mathtt {L}"$ by the partial group $\mathtt {L}'$ .

The situation Gonzalez studies is principally different from ours. However, in [Reference Gonzalez12, Example 7.9, Corollary 7.10], he considers a setup where $\mathtt {L}=\mathtt {T}$ (with $\mathtt {L}$ and $\mathtt {T}$ as above). In this situation, one can observe easily that $(\mathtt {L}',\Delta ',S')$ is a kernel of $(\mathtt {L},\Delta ,S)$ . Indeed, our Theorem B shows that the assumption in [Reference Gonzalez12, Corollary 7.10] that $\Delta '$ contains all $\mathcal {F}'$ -centric subgroups is redundant. It would be the subject of further research to see how far our results have other interesting applications in the context of Gonzalez’s work.

1.3 Products in regular localities and fusion systems

We now demonstrate that the theory of kernels can be used to study certain products in regular localities and thereby construct saturated subsystems of saturated fusion systems.

We study regular localities rather than arbitrary (linking) localities, mainly because every partial normal subgroup of a regular locality can be given the structure of a regular locality. To be exact, if $(\mathcal {L},\Delta ,S)$ is a regular locality and $\mathcal {N}\unlhd \mathcal {L}$ , then $\mathcal {E}:=\mathcal {F}_{S\cap \mathcal {N}}(\mathcal {N})$ is saturated and $(\mathcal {N},\delta (\mathcal {E}),S\cap \mathcal {N})$ is a regular locality.

In localities or, more generally, in partial groups, there is a natural notion of products of subsets. More precisely, if $\mathcal {L}$ is a partial group with product $\Pi \colon \mathbf {D}\rightarrow \mathcal {L}$ , then for $\mathcal {X},\mathcal {Y}\subseteq \mathcal {L}$ , we set

$$\begin{align*}\mathcal{X}\mathcal{Y}:=\{\Pi(x,y)\colon x\in\mathcal{X},\;y\in\mathcal{Y},\;(x,y)\in\mathbf {D}\}.\end{align*}$$

The product of a partial normal subgroup with another partial subgroup is only in special cases known to be a partial subgroup. For example, the product of two partial normal subgroups of a locality $(\mathcal {L},\Delta ,S)$ is a partial normal subgroup (and thus forms a regular locality if $(\mathcal {L},\Delta ,S)$ is regular). Our next theorem gives a further example of a product that is a partial subgroup and can be given the structure of a regular locality.

For the theorem below to be comprehensible, a few preliminary remarks may be useful. For any linking locality $(\mathcal {L},\Delta ,S)$ , one can define the generalised Fitting subgroup $F^*(\mathcal {L})$ as a certain partial normal subgroup of $\mathcal {L}$ (see [Reference Henke18, Definition 3]). If $(\mathcal {L},\Delta ,S)$ is a regular locality, then $F^*(\mathcal {L})$ is a kernel of $(\mathcal {L},\Delta ,S)$ , and thus $T^*:=S\cap F^*(\mathcal {L})$ is an element of $\Delta $ . In particular, $N_{\mathcal {L}}(T^*)$ forms a group of characteristic p. If $\mathcal {N}\unlhd \mathcal {L}$ is a partial normal subgroup of $\mathcal {L}$ , then $N_{\mathcal {N}}(T^*)$ is a normal subgroup of $N_{\mathcal {L}}(T^*)$ . Hence, for any subgroup H of $N_{\mathcal {L}}(T^*)$ , the product $N_{\mathcal {N}}(T^*)H$ is a subgroup of $N_{\mathcal {L}}(T^*)$ .

Theorem E. Let $(\mathcal {L},\Delta ,S)$ be a regular locality. Moreover, fix

$$\begin{align*}\mathcal{N}\unlhd\mathcal{L},\; T:=\mathcal{N}\cap S,\;\mathcal{E}:=\mathcal{F}_T(\mathcal{N}),\;T^*:=F^*(\mathcal{L})\cap S\mbox{ and }H\leq N_{\mathcal{L}}(T^*).\end{align*}$$

Then $\mathcal {N} H$ is a partial subgroup of $\mathcal {L}$ . Moreover, for every Sylow p-subgroup $S_0$ of $N_{\mathcal {N}}(T^*)H$ with $T\leq S_0$ , the following hold:

  1. (a) There exists a unique set $\Delta _0$ of subgroups of $S_0$ such that $(\mathcal {N} H,\Delta _0,S_0)$ is a cr-complete locality with kernel $(\mathcal {N},\delta (\mathcal {E}),T)$ .

  2. (b) Let $\Delta _0$ be as in (a), and set $\tilde {\Delta }_0:=\{P\leq S\colon PO_p(\mathcal {N} H)\in \Delta _0\}$ . Then $(\mathcal {N} H,\tilde {\Delta }_0,S_0)$ is a regular locality if and only if $N_{\mathcal {N}}(T^*)H$ is a group of characteristic p.

If the hypothesis of Theorem E holds and $S\cap H$ is a Sylow p-subgroup of H, then

$$\begin{align*}S_0:=T(S\cap H)\end{align*}$$

is a Sylow p-subgroup of $N_{\mathcal {N}}(T^*)H$ that is contained in S (compare Lemma 6.2). Thus, the cr-complete locality $(\mathcal {N} H,\Delta _0,S_0)$ from Theorem E(a) gives rise to a saturated subsystem $\mathcal {F}_{S_0}(\mathcal {N} H)=\mathcal {F}_{T(H\cap S)}(\mathcal {N} H)$ of $\mathcal {F}$ . This leads us to a statement that can be formulated purely in terms of fusion systems. We use here that, for every regular locality $(\mathcal {L},\Delta ,S)$ over $\mathcal {F}$ , there is by [Reference Chermak and Henke10, Theorem A] a bijection from the set of partial normal subgroups of $\mathcal {L}$ to the set of normal subsystems of $\mathcal {F}$ given by $\mathcal {N}\mapsto \mathcal {F}_{S\cap \mathcal {N}}(\mathcal {N})$ . By [Reference Chermak and Henke10, Theorem E(d)], this bijection takes $F^*(\mathcal {L})$ to $F^*(\mathcal {F})$ . In particular, $F^*(\mathcal {F})$ is a fusion system over $F^*(\mathcal {L})\cap S$ .

To formulate the result we obtain, we rely on the fact that every constrained fusion system is realised by a model: that is, by a finite group of characteristic p. Furthermore, if $\mathcal {F}$ is constrained and G is a model for $\mathcal {F}$ , then every normal subsystem of $\mathcal {F}$ is realised by a normal subgroup of G. We also use that, for every saturated fusion system $\mathcal {F}$ over S, every normal subsystem $\mathcal {E}$ of $\mathcal {F}$ and every subgroup R of S, there is a product subsystem $\mathcal {E} R$ defined (compare [Reference Aschbacher3, Chapter 8] or [Reference Henke13]).

Corollary F. Let $\mathcal {F}$ be a saturated fusion system over S and $\mathcal {E}\unlhd \mathcal {F}$ over $T\leq S$ . Let $T^*,T_0\leq S$ such that

$$\begin{align*}F^*(\mathcal{F})\mbox{ is a fusion system over }T^*\mbox{ and} \ E(\mathcal{E}) \ \mbox{is a fusion system over} \ T_0. \end{align*}$$

Then $N_{\mathcal {F}}(T^*)$ is constrained and $N_{\mathcal {E}}(T_0)\unlhd N_{\mathcal {F}}(T^*)$ . Thus we may choose a model G for $N_{\mathcal {F}}(T^*)$ and $N\unlhd G$ with $\mathcal {F}_{S\cap N}(N)=N_{\mathcal {E}}(T_0)$ . Let $H\leq G$ with $S\cap H\in \mathrm {Syl}_p(H)$ . Set

$$\begin{align*}S_0:=T(S\cap H)\mbox{ and }\mathcal{E} H:=\langle E(\mathcal{E})S_0,\mathcal{F}_{S_0}(NH)\rangle.\end{align*}$$

Then the following hold:

  1. (a) $\mathcal {E} H$ is a saturated fusion system over $S_0$ with $\mathcal {E}\unlhd \mathcal {E} H$ .

  2. (b) If $\mathcal {D}$ is a saturated subsystem of $\mathcal {F}$ with $E(\mathcal {E})\unlhd \mathcal {D}$ and $\mathcal {F}_{S_0}(N H)\subseteq \mathcal {D}$ , then $\mathcal {E} H\subseteq \mathcal {D}$ .

For a saturated fusion system $\mathcal {F}$ with $\mathcal {E}\unlhd \mathcal {F}$ , it is shown, for example, in [Reference Chermak and Henke10, Lemma 7.13(a)] that $E(\mathcal {E})\unlhd \mathcal {F}$ . Hence, it makes sense in the situation above to form the product subsystem $E(\mathcal {E})S_0$ . Moreover, part (b) of Corollary F implies the following statement: If $\mathcal {D}$ is a saturated subsystem of $\mathcal {F}$ with $\mathcal {E}\unlhd \mathcal {D}$ and $\mathcal {F}_{S_0}(N H)\subseteq \mathcal {D}$ , then $\mathcal {E} H\subseteq \mathcal {D}$ .

Organisation of the paper

We start by proving our saturation criterion (Theorem A) in Section 2. An introduction to partial groups and localities is given in Section 3. After proving some preliminary results, we study kernels of localities in Section 5. This is used in Section 6 to prove Theorem E and Corollary F as well as some more detailed results on products.

2 Proving saturation

Throughout this section, let $\mathcal {F}$ be a fusion system over S.

In this section, we prove Theorem A. The reader is referred to [Reference Aschbacher, Kessar and Oliver2, Chapter I] for an introduction to fusion systems. We will adopt the notation and terminology from there with the following two caveats: firstly, we write homomorphisms on the right-hand side of the arguments similarly as in [Reference Aschbacher, Kessar and Oliver2, Chapter II]. Secondly, we define the set $\mathcal {F}^r$ of $\mathcal {F}$ -radical subgroups differently, namely as in Definition 2.5 below.

Definition 2.1. A subgroup $P\leq S$ is said to respect $\mathcal {F}$ -saturation if there exists an element of $P^{\mathcal {F}}$ that is fully automised and receptive in $\mathcal {F}$ (as defined in [Reference Aschbacher, Kessar and Oliver2, Definition I.2.2]).

If $\Delta $ is a set of subgroups of S that is closed under $\mathcal {F}$ -conjugacy, then notice that $\mathcal {F}$ is $\Delta $ -saturated (as defined in the introduction) if and only if every element of $\Delta $ respects $\mathcal {F}$ -saturation. On the other hand, P respects $\mathcal {F}$ -saturation if $\mathcal {F}$ is $P^{\mathcal {F}}$ -saturated. Observe also that a fusion system is saturated if every subgroup of S respects $\mathcal {F}$ -saturation, or equivalently if $\mathcal {F}$ is $\Delta $ -saturated, where $\Delta $ is the set of all subgroups of S. Roberts and Shpectorov [Reference Roberts and Shpectorov26] proved the following lemma, which we will use from now on, most of the time without reference.

Lemma 2.2. Let $\mathcal {C}$ be an $\mathcal {F}$ -conjugacy class of $\mathcal {F}$ . Then $\mathcal {F}$ is $\mathcal {C}$ -saturated if and only if the following two conditions hold:

  1. (I) (Sylow axiom) Each subgroup $P\in \mathcal {C}$ that is fully $\mathcal {F}$ -normalised is also fully $\mathcal {F}$ -centralised and fully automised in $\mathcal {F}$ .

  2. (II) (Extension axiom) Each subgroup $P\in \mathcal {C}$ that is fully $\mathcal {F}$ -centralised is also receptive in $\mathcal {F}$ .

Furthermore, if $\mathcal {F}$ is $\mathcal {C}$ -saturated, then for every fully $\mathcal {F}$ -normalised $P\in \mathcal {C}$ and every $Q\in P^{\mathcal {F}}$ , there exists $\alpha \in \mathrm {Hom}_{\mathcal {F}}(N_S(Q),N_S(P))$ such that $Q\alpha =P$ .

Proof. If (I) and (II) hold, then every fully $\mathcal {F}$ -normalised subgroup $P\in \mathcal {C}$ is fully automised and receptive, and thus $\mathcal {F}$ is $\mathcal {C}$ -saturated. On the other hand, if $\mathcal {F}$ is $\mathcal {C}$ -saturated, it follows from [Reference Aschbacher, Kessar and Oliver2, Lemma I.2.6(c)] that (I) and (II) and the statement of the lemma hold.

Corollary 2.3. If $\mathcal {F}$ is saturated and $P\leq S$ is fully $\mathcal {F}$ -normalised, then for every $Q\in P^{\mathcal {F}}$ , there exists $\alpha \in \mathrm {Hom}_{\mathcal {F}}(N_S(Q),N_S(P))$ such that $Q\alpha =P$ .

Recall that a subgroup $P\leq S$ is called $\mathcal {F}$ -centric if $C_S(Q)\leq Q$ for every $Q\in P^{\mathcal {F}}$ . Write $\mathcal {F}^c$ for the set of $\mathcal {F}$ -centric subgroups of S. We will need the following lemma.

Lemma 2.4. Let $\mathcal {F}$ be a fusion system over S, and let $P\leq S$ be fully $\mathcal {F}$ -centralised. Then P is $\mathcal {F}$ -centric if and only if $C_S(P)\leq P$ .

Proof. If P is $\mathcal {F}$ -centric, then clearly $C_S(P)\leq P$ . Suppose now that $C_S(P)\leq P$ . Then for any $Q\in P^{\mathcal {F}}$ , we have $C_S(Q)\geq Z(Q)\cong Z(P)=C_S(P)$ . Hence, as P is fully $\mathcal {F}$ -centralised, we have $C_S(Q)=Z(Q)\leq Q$ for all $Q\in P^{\mathcal {F}}$ : that is, P is $\mathcal {F}$ -centric.

We recall that $O_p(\mathcal {F})$ denotes the largest subgroup of S that is normal in $\mathcal {F}$ ([Reference Aschbacher, Kessar and Oliver2, Definition I.4.3]). Moreover, if $\mathcal {F}$ is saturated and $O_p(\mathcal {F}) \in \mathcal {F}^c$ , then $\mathcal {F}$ is called constrained ([Reference Aschbacher, Kessar and Oliver2, Definition I.4.8]). The Model Theorem for constrained fusion systems [Reference Aschbacher, Kessar and Oliver2, Theorem III.5.10] guarantees that every constrained fusion system $\mathcal {F}$ over S has a model: that is, there is a finite group G such that $S \in \mathrm {Syl}_p(G)$ , $\mathcal {F}_S(G) = \mathcal {F}$ and $C_G(O_p(G)) \leq O_p(G)$ .

Definition 2.5.

  • A subgroup $P\leq S$ is called $\mathcal {F}$ -radical if there exists a fully $\mathcal {F}$ -normalised $\mathcal {F}$ -conjugate Q of P such that $O_p(N_{\mathcal {F}}(Q)) = Q$ . We denote by $\mathcal {F}^r$ the set of $\mathcal {F}$ -radical subgroups of S.

  • Set $\mathcal {F}^{cr} = \mathcal {F}^c \cap \mathcal {F}^r$ , and call the elements of $\mathcal {F}^{cr}$ the $\mathcal {F}$ -centric radical subgroups of S.

  • A subgroup $P\leq S$ is called $\mathcal {F}$ -critical if P is $\mathcal {F}$ -centric and, for every $\mathcal {F}$ -conjugate Q of P, we have

    $$\begin{align*}\mathrm{Out}_S(Q)\cap O_p(\mathrm{Out}_{\mathcal{F}}(Q))=1.\end{align*}$$

As remarked before, our definition of radical subgroups differs from the usual one given, for example, in [Reference Aschbacher, Kessar and Oliver2, Definition I.3.1]. We show in part (b) of our next lemma that, for a saturated fusion system $\mathcal {F}$ , the set $\mathcal {F}^{cr}$ equals the set of $\mathcal {F}$ -centric radical subgroups in the usual definition.

Lemma 2.6.

  1. (a) For every $R\leq S$ , the following implications hold:

    $$\begin{align*}R\in\mathcal{F}^c\mbox{ and }O_p(\mathrm{Aut}_{\mathcal{F}}(R))=\mathrm{Inn}(R)\Longrightarrow R\mbox{ is} \ \mathcal{F}\mbox{-critical} \Longrightarrow R\in\mathcal{F}^{cr}. \end{align*}$$
  2. (b) If $\mathcal {F}$ is saturated, then we have

    $$\begin{align*}\mathcal{F}^{cr}=\{R\in\mathcal{F}^c\colon O_p(\mathrm{Aut}_{\mathcal{F}}(R))=\mathrm{Inn}(R)\}=\{R\leq S\colon R\mbox{ is} \ \mathcal{F}\mbox{-critical}\}. \end{align*}$$

Proof. If $\mathrm {Inn}(R)=O_p(\mathrm {Aut}_{\mathcal {F}}(R))$ , then $\mathrm {Inn}(Q)=O_p(\mathrm {Aut}_{\mathcal {F}}(Q))=\mathrm {Aut}_S(Q)\cap O_p(\mathrm {Aut}_{\mathcal {F}}(Q))$ for every $\mathcal {F}$ -conjugate Q of R, so R is $\mathcal {F}$ -critical if in addition $R\in \mathcal {F}^c$ . This shows the first implication in (a).

Now let $R\in \mathcal {F}^c$ such that $R\not \in \mathcal {F}^r$ . If we pick a fully $\mathcal {F}$ -normalised $\mathcal {F}$ -conjugate Q of R, we have $Q<Q^*:=O_p(N_{\mathcal {F}}(Q))$ . So $\mathrm {Inn}(Q)<\mathrm {Aut}_{Q^*}(Q)$ as $Q\in \mathcal {F}^c$ . Moreover, $\mathrm {Aut}_{Q^*}(Q)$ is normal in $\mathrm {Aut}_{\mathcal {F}}(Q)$ , as by definition of $Q^*$ every element of $\mathrm {Aut}_{\mathcal {F}}(Q)$ extends to an element of $\mathrm {Aut}_{\mathcal {F}}(Q^*)$ . Hence, $\mathrm {Inn}(Q)<\mathrm {Aut}_{Q^*}(Q)\leq \mathrm {Aut}_S(Q)\cap O_p(\mathrm {Aut}_{\mathcal {F}}(Q))$ . This shows that R is not $\mathcal {F}$ -critical, so (a) holds. In particular,

$$\begin{align*}\mathcal{F}^{cr} \supseteq \{R\leq S\colon R\mbox{ is} \ \mathcal{F}\mbox{-critical }\} \supseteq \{R\in\mathcal{F}^c\colon O_p(\mathrm{Aut}_{\mathcal{F}}(R))=\mathrm{Inn}(R)\}. \end{align*}$$

For the proof of (b), it is thus sufficient to show that $O_p(\mathrm {Aut}_{\mathcal {F}}(R))=\mathrm {Inn}(R)$ for every $R\in \mathcal {F}^{cr}$ . Now fix $R\in \mathcal {F}^{cr}$ . Since the property $O_p(\mathrm {Aut}_{\mathcal {F}}(R))=\mathrm {Inn}(R)$ is preserved if R is replaced by an $\mathcal {F}$ -conjugate, we may assume without loss of generality that R is fully $\mathcal {F}$ -normalised and $R=O_p(N_{\mathcal {F}}(R))$ . Note that $N_{\mathcal {F}}(R)$ is saturated. So as $R\in \mathcal {F}^c$ , the subsystem $N_{\mathcal {F}}(R)$ is constrained. Thus, we may choose a model G for $N_{\mathcal {F}}(R)$ . Then $O_p(G)=R=O_p(N_{\mathcal {F}}(R))$ and

$$\begin{align*}\mathrm{Aut}_{\mathcal{F}}(R)\cong G/C_G(R)\cong G/Z(R). \end{align*}$$

Hence, $O_p(\mathrm {Aut}_{\mathcal {F}}(R))\cong O_p(G/Z(R))=R/Z(R)\cong \mathrm {Inn}(R)$ . Thus $O_p(\mathrm {Aut}_{\mathcal {F}}(R))= \mathrm {Inn}(R)$ . This proves that $\mathcal {F}^{cr} \subseteq \{R\in \mathcal {F}^c\colon O_p(\mathrm {Aut}_{\mathcal {F}}(R))=\mathrm {Inn}(R)\}\}$ , so (b) holds.

The $\mathcal {F}$ -critical subgroups play a crucial role in showing that a fusion system is saturated. This is made precise in the following theorem, which we restate for the reader’s convenience.

Theorem 2.7. Suppose $\Delta $ is a set of subgroups of S that is closed under $\mathcal {F}$ -conjugacy and contains every $\mathcal {F}$ -critical subgroup. If $\mathcal {F}$ is $\Delta $ -generated and $\Delta $ -saturated, then $\mathcal {F}$ is saturated.

Proof. This is a reformulation of [Reference Broto, Castellana, Grodal, Levi and Oliver5, Theorem 2.2]. The reader might also want to note that the theorem follows from Lemma 2.9 below.

Later, we will need to prove saturation in a situation where it appears impossible to apply Theorem 2.7 directly. We will therefore have a closer look at the arguments used in the proof of that theorem.

Define a partial order $\preceq $ on the set of $\mathcal {F}$ -conjugacy classes by writing $\mathcal {P}\preceq \mathcal {Q}$ if some (and thus every) element of $\mathcal {Q}$ contains an element of $\mathcal {P}$ .

Lemma 2.8. Let $\mathcal {H}$ be a set of subgroups of S closed under $\mathcal {F}$ -conjugacy such that $\mathcal {F}$ is $\mathcal {H}$ -generated and $\mathcal {H}$ -saturated. Let $\mathcal {P}$ be an $\mathcal {F}$ -conjugacy class that is maximal with respect to $\preceq $ among those $\mathcal {F}$ -conjugacy classes that are not contained in $\mathcal {H}$ .

Write $\mathcal {S}_{\geq P}\supseteq \mathcal {S}_{>P}$ for the sets of subgroups of $N_S(P)$ that contain, or properly contain, an element $P\in \mathcal {P}$ . Then the following hold for every $P\in \mathcal {P}$ that is fully $\mathcal {F}$ -normalised:

  1. (a) The subsystem $N_{\mathcal {F}}(P)$ is $\mathcal {S}_{>P}$ -generated and $\mathcal {S}_{>P}$ -saturated.

  2. (b) If $N_{\mathcal {F}}(P)$ is $\mathcal {S}_{\geq P}$ -saturated, then $\mathcal {F}$ is $\mathcal {H}\cup \mathcal {P}$ -saturated.

  3. (c) P is fully $\mathcal {F}$ -centralised. Moreover, for every $Q\in \mathcal {P}$ , there exists $\alpha \in \mathrm {Hom}_{\mathcal {F}}(N_S(Q),N_S(P))$ such that $Q\alpha =P$ .

  4. (d) If $C_S(P)\leq P$ , then P is $\mathcal {F}$ -centric, and if $\mathrm {Out}_S(P)\cap O_p(\mathrm {Out}_{\mathcal {F}}(P))=1$ , then $\mathrm {Out}_S(Q)\cap O_p(\mathrm {Out}_{\mathcal {F}}(Q))=1$ for all $Q\in \mathcal {P}$ . In particular, if $C_S(P)\leq P$ and $\mathrm {Out}_S(P)\cap O_p(\mathrm {Out}_{\mathcal {F}}(P))=1$ , then P is $\mathcal {F}$ -critical.

Proof. Basically, this follows from [Reference Broto, Castellana, Grodal, Levi and Oliver5, Lemma 2.4] and its proof. We take the opportunity to point out the following small error in the proof of part (b) of that lemma: in l.9 on p.334 of [Reference Broto, Castellana, Grodal, Levi and Oliver5] it says ‘replacing each $\varphi _i$ by $\chi _i\circ \varphi _i\circ \chi _i^{-1}\in \mathrm {Hom}_{\mathcal {F}}(\chi _i(Q_i),S)$ ’. However, it should be ‘replacing each $\varphi _i$ by $\chi _{i+1}\circ \varphi _i\circ \chi _i^{-1}\in \mathrm {Hom}_{\mathcal {F}}(\chi _i(Q_i),S)$ ’.

Let us now explain in detail how the assertion follows. By [Reference Broto, Castellana, Grodal, Levi and Oliver5, Lemma 2.4(a), (c)], $N_{\mathcal {F}}(P)$ is $\mathcal {S}_{>P}$ -saturated and (b) holds. To prove (a), one needs to argue that $N_{\mathcal {F}}(P)$ is also $\mathcal {S}_{>P}$ -generated. If $\varphi \in \mathrm {Hom}_{N_{\mathcal {F}}(P)}(A,B)$ , then $\varphi $ extends to a morphism $\hat {\varphi }\in \mathrm {Hom}_{\mathcal {F}}(AP,BP)$ that normalises P. If $A\not \leq P$ , then $AP\in \mathcal {S}_{>P}$ , and thus $\varphi $ is the restriction of a morphism between subgroups in $\mathcal {S}_{>P}$ . If $A\leq P$ , then $\hat {\varphi }\in \mathrm {Aut}_{\mathcal {F}}(P)$ , and by [Reference Broto, Castellana, Grodal, Levi and Oliver5, Lemma 2.4(b)], $\hat {\varphi }$ (and thus $\varphi $ ) is a composite of restrictions of morphisms in $N_{\mathcal {F}}(P)$ between subgroups in $\mathcal {S}_{>P}$ . Hence, $N_{\mathcal {F}}(P)$ is $\mathcal {S}_{>P}$ generated, and (a) holds.

The following property is shown in the proof of Lemma 2.4 in [Reference Broto, Castellana, Grodal, Levi and Oliver5] (see p.333, property (3)).

  1. (*) There is a subgroup $\hat {P}\in \mathcal {P}$ fully $\mathcal {F}$ -centralised such that, for all $Q\in \mathcal {P}$ , there exists a morphism $\varphi \in \mathrm {Hom}_{\mathcal {F}}(N_S(Q),N_S(\hat {P}))$ with $Q\varphi =\hat {P}$ .

In particular, there exists $\psi \in \mathrm {Hom}_{\mathcal {F}}(N_S(P),N_S(\hat {P}))$ such that $P\psi =\hat {P}$ . Since P is fully $\mathcal {F}$ -normalised, the map $\psi $ is an isomorphism. In particular, P is fully $\mathcal {F}$ -centralised as $\hat {P}$ is fully $\mathcal {F}$ -centralised. Moreover, if $Q\in \mathcal {P}$ , then by (*), there exists $\varphi \in \mathrm {Hom}_{\mathcal {F}}(N_S(Q),N_S(\hat {P}))$ with $Q\varphi =\hat {P}$ , and we have $\alpha :=\varphi \psi ^{-1}\in \mathrm {Hom}_{\mathcal {F}}(N_S(Q),N_S(P))$ , with $Q\alpha =\hat {P}\psi ^{-1}=P$ . Hence, (c) holds.

It follows from Lemma 2.4 and the first part of (c) that P is $\mathcal {F}$ -centric if $C_S(P)\leq P$ . If $Q\in \mathcal {P}$ , then the second part of (c) allows us to choose $\alpha \in \mathrm {Hom}_{\mathcal {F}}(N_S(Q),N_S(P))$ with $Q\alpha =P$ . Notice that the map $\alpha ^*\colon \mathrm {Aut}_{\mathcal {F}}(Q)\rightarrow \mathrm {Aut}_{\mathcal {F}}(P),\gamma \mapsto \alpha ^{-1}\gamma \alpha $ is a group isomorphism with $\mathrm {Inn}(Q)\alpha ^*=\mathrm {Inn}(P)$ and $\mathrm {Aut}_S(Q)\alpha ^*\leq \mathrm {Aut}_S(P)$ . So $\alpha ^*$ induces an isomorphism from $\mathrm {Out}_{\mathcal {F}}(Q)$ to $\mathrm {Out}_{\mathcal {F}}(P)$ that maps $\mathrm {Out}_S(Q)\cap O_p(\mathrm {Out}_{\mathcal {F}}(Q))$ into $\mathrm {Out}_S(P)\cap O_p(\mathrm {Out}_{\mathcal {F}}(P))$ . This implies (d).

Lemma 2.9. Suppose that $\mathcal {H}$ is one of the following two sets:

  1. (i) the set of all subgroups of S respecting $\mathcal {F}$ -saturation; or

  2. (ii) the set of all $P\leq S$ such that every subgroup of S containing an $\mathcal {F}$ -conjugate of P respects $\mathcal {F}$ -saturation.

Assume that $\mathcal {F}$ is $\mathcal {H}$ -generated. Let $\mathcal {P}$ be an $\mathcal {F}$ -conjugacy class, which is maximal with respect to $\preceq $ among the $\mathcal {F}$ -conjugacy classes not contained in $\mathcal {H}$ . Then the elements of $\mathcal {P}$ are $\mathcal {F}$ -critical.

Proof. Observe first that $\mathcal {H}$ is in either case closed under $\mathcal {F}$ -conjugacy and $\mathcal {F}$ is $\mathcal {H}$ -saturated. Moreover, note that $\mathcal {F}$ is not $\mathcal {H}\cup \mathcal {P}$ -saturated; this is clear if $\mathcal {H}$ is as in (i); if $\mathcal {H}$ is as in (ii), then note that because of the maximal choice of $\mathcal {P}$ , $\mathcal {F}$ being $\mathcal {H}\cup \mathcal {P}$ -saturated would imply that every subgroup containing an element of $\mathcal {P}$ would be either in $\mathcal {P}$ or in $\mathcal {H}$ and thus respect $\mathcal {F}$ -saturation.

Now fix $P\in \mathcal {P}$ fully $\mathcal {F}$ -normalised. By Lemma 2.8(a),(b) (using the notation introduced in that lemma), $N_{\mathcal {F}}(P)$ is $\mathcal {S}_{>P}$ generated and $\mathcal {S}_{>P}$ -saturated, but not $\mathcal {S}_{\geq P}$ -saturated as $\mathcal {F}$ is not $\mathcal {H}\cup \mathcal {P}$ -saturated. Thus, by [Reference Broto, Castellana, Grodal, Levi and Oliver5, Lemma 2.5] applied with $N_{\mathcal {F}}(P)$ in place of $\mathcal {F}$ , we have $\mathrm {Out}_S(P)\cap O_p(\mathrm {Out}_{\mathcal {F}}(P))=1$ and $P\in N_{\mathcal {F}}(P)^c$ . It now follows from Lemma 2.8(d) that P is $\mathcal {F}$ -critical.

Lemma 2.10. Let $\mathcal {E}$ be an $\mathcal {F}$ -invariant subsystem of $\mathcal {F}$ over $T\leq S$ . Let $P\leq S$ , and set $P_0:=P\cap T$ . Then the following hold:

  1. (a) If $C_S(P)\leq P$ and $\mathrm {Out}_S(P)\cap O_p(\mathrm {Out}_{\mathcal {F}}(P))=1$ , then $C_T(P_0)\leq P_0$ and $\mathrm {Out}_T(P_0)\cap O_p(\mathrm {Out}_{\mathcal {E}}(P_0))=1$ .

  2. (b) If $\mathcal {E}$ is saturated, $P_0$ is fully $\mathcal {E}$ -normalised and P is $\mathcal {F}$ -critical, then $P_0\in \mathcal {E}^{cr}$ .

Proof. Assume that $C_T(P_0)\not \leq P_0$ or $\mathrm {Out}_T(P_0)\cap O_p(\mathrm {Out}_{\mathcal {E}}(P_0))\neq 1$ . Let Q be the preimage of $\mathrm {Out}_T(P_0)\cap O_p(\mathrm {Out}_{\mathcal {E}}(P_0))$ in $N_T(P_0)$ . Note that $C_T(P_0)\leq Q$ , so our assumption implies in any case that $P_0<Q$ and thus $Q\not \leq P$ . As P normalises $P_0$ , P also normalises Q. Hence, $PQ$ is a p-group with $P<PQ$ . This yields $P<N_{PQ}(P)=PN_Q(P)$ , and thus $N_Q(P)\not \leq P$ . Note that $[P,N_Q(P)]\leq P\cap T=P_0$ . So $X:=\langle \mathrm {Aut}_Q(P)^{\mathrm {Aut}_{\mathcal {F}}(P)}\rangle $ acts trivially on $P/P_0$ . At the same time, by definition of Q, the elements of $\mathrm {Aut}_Q(P)$ , and thus the elements of X, restrict to automorphisms of $P_0$ that lie in $O_p(\mathrm {Out}_{\mathcal {E}}(P_0))\unlhd \mathrm {Aut}_{\mathcal {F}}(P_0)$ . Hence, a $p^{\prime }$ -element of X acts trivially on $P/P_0$ and on $P_0$ , and therefore it is trivial by properties of coprime action (compare [Reference Kurzweil and Stellmacher21, 8.2.2(b)] or [Reference Aschbacher, Kessar and Oliver2, Lemma A.2]). This shows that X is a p-group, so $\mathrm {Aut}_Q(P)\leq \mathrm {Aut}_S(P)\cap O_p(\mathrm {Aut}_{\mathcal {F}}(P))$ . As $N_Q(P)\not \leq P$ , this implies either $C_S(P)\not \leq P$ or $1\neq \mathrm {Out}_Q(P)\leq \mathrm {Out}_S(P)\cap O_p(\mathrm {Out}_{\mathcal {F}}(P))$ . This proves (a).

Assume now that $\mathcal {E}$ is saturated, $P_0$ is fully $\mathcal {E}$ -normalised and P is $\mathcal {F}$ -critical. The latter condition implies $C_S(P)\leq P$ and $\mathrm {Out}_S(P)\cap O_p(\mathrm {Out}_{\mathcal {F}}(P))=1$ . So by (a), $C_T(P_0)\leq P_0$ and $\mathrm {Out}_T(P_0)\cap O_p(\mathrm {Out}_{\mathcal {E}}(P_0))=1$ . As $\mathcal {E}$ is saturated and $P_0$ is fully $\mathcal {E}$ -normalised, $P_0$ is fully $\mathcal {E}$ -centralised and fully $\mathcal {E}$ -automised. Hence, $P_0$ is $\mathcal {E}$ -centric by Lemma 2.4 and $O_p(\mathrm {Out}_{\mathcal {E}}(P_0))\leq \mathrm {Out}_T(P_0)$ . The latter condition implies $O_p(\mathrm {Out}_{\mathcal {E}}(P_0))=1$ . Using Lemma 2.6(a), we conclude that $P_0\in \mathcal {E}^{cr}$ .

The basic idea in the proof of the following theorem is taken from Step 5 of the proof of [Reference Oliver23, Theorem 9]. As we argue afterwards, it easily implies Theorem A. If $P,Q\leq S$ , $\varphi \in \mathrm {Hom}_{\mathcal {F}}(P,Q)$ and $A\leq \mathrm {Aut}_{\mathcal {F}}(P)$ , note that $A^{\varphi }:=\varphi ^{-1}A\varphi \leq \mathrm {Aut}_{\mathcal {F}}(P\varphi )$ .

Theorem 2.11. Let $\mathcal {H}$ be the set of all subgroups $P\leq S$ such that every subgroup of S containing an $\mathcal {F}$ -conjugate of P respects $\mathcal {F}$ -saturation. Assume that $\mathcal {F}$ is $\mathcal {H}$ -generated. Suppose furthermore that there exists an $\mathcal {F}$ -invariant saturated subsystem $\mathcal {E}$ of $\mathcal {F}$ with $\mathcal {E}^{cr}\subseteq \mathcal {H}$ . Then $\mathcal {F}$ is saturated.

Proof. Let $T\leq S$ be such that $\mathcal {E}$ is a subsystem of $\mathcal {F}$ over T. By $\mathcal {H}_0$ , denote the set of subgroups of T that are elements of $\mathcal {H}$ . Write $\mathcal {H}^{\perp }$ for the set of subgroups of S not in $\mathcal {H}$ and $\mathcal {H}_0^{\perp }$ for the set of elements of $\mathcal {H}^{\perp }$ that are subgroups of T. If $P\leq S$ , write $P_0$ for $P\cap T$ , and similarly for subgroups of S with different names.

For the proof of the assertion, we will proceed in three steps. In the first two steps, we will argue that property (2.1) below implies the assertion, and in the third step, we will prove that (2.1) holds.

(2.1) $$ \begin{align} &\mbox{Let }P\leq S\mbox{ such that }P_0\mbox{ is an element of }\mathcal{H}_0^{\perp}\mbox{ of maximal order. If }P_0\mbox{ is not fully}\\ &\mathcal{E}\mbox{-normalised, then there exists} \ P"\in P^{\mathcal{F}} \ \mbox{such that} \ |N_T(P_0)|<|N_T(P_0")|.\nonumber \end{align} $$

Step 1: We argue that (2.1) implies the following property:

(2.2) $$ \begin{align} &\mbox{If} \ P\leq S\ \mbox{such that} \ P_0 \ \mbox{is an element of} \ \mathcal{H}_0^{\perp} \ \mbox{of maximal order, then there exists an}\\ &\mathcal{F}\mbox{-conjugate} \ Q \ \mbox{of} \ P \ \mbox{such that} \ Q_0\ \mbox{is fully} \ \mathcal{E}\mbox{-normalised.}\nonumber \end{align} $$

This can be seen as follows. If P is as in (2.2) and $P_0$ is fully $\mathcal {E}$ -normalised, then we are done. If not, then by (2.1), we can pass on to an $\mathcal {F}$ -conjugate $P"$ of P such that $|N_T(P_0)|<|N_T(P_0")|$ . Again, if $P_0"$ is fully $\mathcal {E}$ -normalised, then we are done; otherwise, we can repeat the process. Since T is finite, we will eventually end up with an $\mathcal {F}$ -conjugate Q of P such that $Q_0$ is fully $\mathcal {E}$ -normalised. So (2.1) implies (2.2).

Step 2: We show that property (2.2) implies the assertion. Assume that (2.2) holds and $\mathcal {F}$ is not saturated. Then $\mathcal {H}^{\perp }\neq \emptyset $ . By construction, $\mathcal {H}\supseteq \mathcal {H}_0$ is $\mathcal {F}$ -closed. So, for every subgroup $P\leq S$ , $P_0\in \mathcal {H}_0$ implies $P\in \mathcal {H}$ . In particular, as $\mathcal {H}^{\perp }\neq \emptyset $ , we can conclude $\mathcal {H}_0^{\perp }\neq \emptyset $ . Set

$$ \begin{align*} m &:= \max\{|Q|\colon Q\in\mathcal{H}_0^{\perp}\},\\ \mathcal{M} &:= \{P\in\mathcal{H}^{\perp}\colon |P_0|=m\}. \end{align*} $$

For any $Q\in \mathcal {H}_0^{\perp }$ with $|Q|=m$ , we have $Q=Q_0$ , and thus $Q\in \mathcal {M}$ . Hence, $\mathcal {M}\neq \emptyset $ . As T is strongly closed, $\mathcal {M}$ is closed under $\mathcal {F}$ -conjugacy. Thus we can choose an $\mathcal {F}$ -conjugacy class $\mathcal {P}\subseteq \mathcal {M}$ such that the elements of $\mathcal {P}$ are of maximal order among the elements of $\mathcal {M}$ .

We argue first that $\mathcal {P}$ is maximal with respect to $\preceq $ among the $\mathcal {F}$ -conjugacy classes contained in $\mathcal {H}^{\perp }$ . For that, let $P\in \mathcal {P}$ and $P\leq R\leq S$ . We need to show that $R=P$ or $R\in \mathcal {H}$ . Notice that $m=|P_0|\leq |R_0|$ . If $m<|R_0|$ , then by definition of m, we have $R_0\in \mathcal {H}_0\subseteq \mathcal {H}$ , and thus $R\in \mathcal {H}$ . If $m=|R_0|$ , then $R\in \mathcal {M}$ or $R\in \mathcal {H}$ . So the maximality of $|P|$ yields $R=P$ or $R\in \mathcal {H}$ . Hence, $\mathcal {P}$ is maximal with respect to $\preceq $ among the $\mathcal {F}$ -conjugacy classes contained in $\mathcal {H}^{\perp }$ . Thus Lemma 2.9 implies that the elements of $\mathcal {P}$ are $\mathcal {F}$ -critical. By (2.2), there exists $P\in \mathcal {P}$ such that $P_0$ is fully $\mathcal {E}$ -normalised. Then by Lemma 2.10(b), we have $P_0\in \mathcal {E}^{cr}$ . Since by assumption $\mathcal {E}^{cr}\subseteq \mathcal {H}$ , it follows that $P_0\in \mathcal {H}$ , so $P\in \mathcal {H}$ . This contradicts the choice of $\mathcal {P}\subseteq \mathcal {M}\subseteq \mathcal {H}^{\perp }$ . Hence, (2.2) implies that $\mathcal {F}$ is saturated.

Step 3: We complete the proof by proving (2.1). Whenever we have subgroups $Q_1\leq Q\leq S$ , we set $\mathrm {Aut}_{\mathcal {F}}(Q:Q_1):=N_{\mathrm {Aut}_{\mathcal {F}}(Q)}(Q_1)$ , $\mathrm {Aut}_S(Q:Q_1):=N_{\mathrm {Aut}_S(Q)}(Q_1)$ , $N_S(Q:Q_1):=N_S(Q)\cap N_S(Q_1)$ , and so on. Now let $P\leq S$ be such that $P_0$ is an element of $\mathcal {H}_0^{\perp }$ of maximal order that is not fully $\mathcal {E}$ -normalised. Note that $T\in \mathcal {E}^{cr}\subseteq \mathcal {H}$ and thus $P_0\neq T$ . Hence, $P_0<N_T(P_0)$ . So the maximality of $|P_0|$ yields $R:=N_T(P_0)\in \mathcal {H}_0$ . In particular, every $\mathcal {F}$ -conjugate of R respects $\mathcal {F}$ -saturation. Since $\mathcal {E}$ is saturated, there exists $\rho \in \mathrm {Hom}_{\mathcal {E}}(R,T)$ such that $P_0':=P_0\rho $ is fully $\mathcal {E}$ -normalised. As $R':=R\rho $ respects $\mathcal {F}$ -saturation, there exists $\sigma \in \mathrm {Hom}_{\mathcal {F}}(R',S)$ such that $R":=R'\sigma $ is fully automised and receptive. Set $P_0":=P_0' \sigma =P_0 \rho \sigma $ . As $R"$ is fully automised, by Sylow’s Theorem, there exists $\xi \in \mathrm {Aut}_{\mathcal {F}}(R")$ such that

$$\begin{align*}\mathrm{Aut}_{\mathcal{F}}(R":P_0")^{\xi}\cap\mathrm{Aut}_S(R")=\mathrm{Aut}_S(R": P_0" \xi)\end{align*}$$

is a Sylow p-subgroup of $\mathrm {Aut}_{\mathcal {F}}(R":P_0")^{\xi }=\mathrm {Aut}_{\mathcal {F}}(R": P_0" \xi )$ . So replacing $\sigma $ by $\sigma \xi $ , we may assume

$$\begin{align*}\mathrm{Aut}_S(R":P_0")\in\mathrm{Syl}_p(\mathrm{Aut}_{\mathcal{F}}(R":P_0")).\end{align*}$$

Notice that $\mathrm {Aut}_S(R:P_0)^{\rho \sigma }$ and $\mathrm {Aut}_S(R':P_0')^{\sigma }$ are p-subgroups of $\mathrm {Aut}_{\mathcal {F}}(R":P_0")$ . Hence, again by Sylow’s Theorem, there exist $\gamma ,\delta \in \mathrm {Aut}_{\mathcal {F}}(R":P_0")$ such that $\mathrm {Aut}_S(R:P_0)^{\rho \sigma \gamma }$ and $\mathrm {Aut}_S(R':P_0')^{\sigma \delta }$ are contained in $\mathrm {Aut}_S(R":P_0")$ . This implies $N_S(R:P_0)\leq N_{\rho \sigma \gamma }$ and $N_S(R':P_0')\leq N_{\sigma \delta }$ . Hence, as $R"$ is receptive, the map $\rho \sigma \gamma \in \mathrm {Hom}_{\mathcal {F}}(R,R")$ extends to $\varphi \in \mathrm {Hom}_{\mathcal {F}}(N_S(R:P_0),S)$ , and similarly $\sigma \delta \in \mathrm {Hom}_{\mathcal {F}}(R',R")$ extends to $\psi \in \mathrm {Hom}_{\mathcal {F}}(N_S(R':P_0'),S)$ . Notice that

$$\begin{align*}P_0 \varphi=P_0 {\rho\sigma\gamma}=P_0" \gamma=P_0"\mbox{ and } P_0' \psi= P_0' {\sigma\delta}=P_0" \delta=P_0".\end{align*}$$

Moreover, since $P_0=P\cap T$ , $R=N_T(P_0)$ and T is strongly closed, we have $P\leq N_S(P_0)=N_S(R:P_0)$ . Hence

$$\begin{align*}P":=P\varphi\end{align*}$$

is a well-defined $\mathcal {F}$ -conjugate of P, for which we have indeed that $P"\cap T=P \varphi \cap T=(P\cap T)\varphi =P_0 \varphi =P_0"$ as our notation suggests. So it only remains to show that $|N_T(P_0)|<|N_T(P_0")|$ . We will show this by arguing that

(2.3) $$ \begin{align} |N_T(P_0)|=|R|<|N_T(R':P_0')|\leq |N_T(P_0")|. \end{align} $$

Recall that $P_0$ is by assumption not fully $\mathcal {E}$ -normalised, whereas $P_0'=P_0 \rho $ is fully $\mathcal {E}$ -normalised. Therefore, we have $R'=R \rho =N_T(P_0) \rho <N_T(P_0')$ . Thus $R'<N_{N_T(P_0')}(R')=N_T(R':P_0')$ . As $|R|=|R'|$ , this shows the first inequality in (2.3). As $P_0' \psi =P_0"$ and T is strongly closed, we have $N_T(R':P_0') \psi \leq N_T(P_0")$ . Since $\psi $ is injective, this shows the second inequality in (2.3). So the proof of (2.1) is complete. As argued in Steps 1 and 2, this proves the assertion.

Proof of Theorem A.

Write $\mathcal {H}$ for the set of all subgroups $P\leq S$ such that every subgroup of S containing an $\mathcal {F}$ -conjugate of P respects $\mathcal {F}$ -saturation. Since $\Delta $ is $\mathcal {F}$ -closed and $\mathcal {F}$ is $\Delta $ -saturated, it follows that $\Delta \subseteq \mathcal {H}$ . Hence, as $\mathcal {F}$ is $\Delta $ -generated, $\mathcal {F}$ is also $\mathcal {H}$ -generated. Moreover, $\mathcal {E}^{cr}\subseteq \Delta \subseteq \mathcal {H}$ . Thus the assertion follows from Theorem 2.11 above.

3 Partial groups and localities

In this section, we introduce some basic definitions and notations that will be used in the remainder of the paper. We refer the reader to [Reference Chermak8] for a more comprehensive introduction to partial groups and localities.

3.1 Partial groups

Following the notation introduced in [Reference Chermak8], we will write $\mathbf {W}(\mathcal {L})$ for the set of words in a set $\mathcal {L}$ . The elements of $\mathcal {L}$ will be identified with the words of length one, and $\emptyset $ denotes the empty word. The concatenation of words $u_1,u_2,\dots ,u_k\in \mathbf {W}(\mathcal {L})$ will be denoted $u_1\circ u_2\circ \cdots \circ u_k$ .

Definition 3.1 [Reference Chermak8, Definition 1.1].

Suppose $\mathcal {L}$ is a nonempty set and $\mathbf {D}\subseteq \mathbf {W}(\mathcal {L})$ . Let $\Pi \colon \mathbf {D} \longrightarrow \mathcal {L}$ be a map, and let $(-)^{-1} \colon \mathcal {L} \longrightarrow \mathcal {L}$ be an involutory bijection, which we extend to a map

$$\begin{align*}(-)^{-1} \colon \mathbf {W}(\mathcal{L}) \longrightarrow \mathbf {W}(\mathcal{L}), w = (g_1, \dots, g_k) \mapsto w^{-1} = (g_k^{-1}, \dots, g_1^{-1}).\end{align*}$$

Then $\mathcal {L}$ is called a partial group with product $\Pi $ and inversion $(-)^{-1}$ if the following hold for all words $u,v,w\in \mathbf {W}(\mathcal {L})$ :

  • $\mathcal {L} \subseteq \mathbf {D}$ and

    $$\begin{align*}u \circ v \in \mathbf {D} \Longrightarrow u,v \in \mathbf {D}.\end{align*}$$
    (So in particular, $\emptyset \in \mathbf {D}$ .)
  • $\Pi $ restricts to the identity map on $\mathcal {L}$ .

  • $u \circ v \circ w \in \mathbf {D} \Longrightarrow u \circ (\Pi (v)) \circ w \in \mathbf {D}$ , and $\Pi (u \circ v \circ w) = \Pi (u \circ (\Pi (v)) \circ w)$ .

  • $w \in \mathbf {D} \Longrightarrow w^{-1} \circ w\in \mathbf {D}$ and $\Pi (w^{-1} \circ w) = \operatorname {\mathbf {1}}$ , where $\operatorname {\mathbf {1}}:=\Pi (\emptyset )$ .

For the remainder of this section, let ${\mathcal {L}}$ always be a partial group with product ${\Pi \colon \mathbf {D}\rightarrow \mathcal {L}}$ . As above, set ${\operatorname {\mathbf {1}}:=\Pi (\emptyset )}$ .

If $w=(f_1,\dots ,f_n)\in \mathbf {D}$ , then we sometimes write $f_1f_2\cdots f_n$ for $\Pi (f_1,\dots ,f_n)$ . In particular, if $(x,y)\in \mathbf {D}$ , then $xy$ denotes the product $\Pi (x,y)$ .

Notation 3.2.

  • For every $f\in \mathcal {L}$ , write $\mathbf {D}(f):=\{x\in \mathcal {L}\colon (f^{-1},x,f)\in \mathbf {D}\}$ for the set of all x such that the conjugate $x^f:=\Pi (f^{-1},x,f)$ is defined.

  • By $c_f$ , denote the conjugation map $c_f\colon \mathbf {D}(f)\rightarrow \mathcal {L},x\mapsto x^f$ .

  • For $f\in \mathcal {L}$ and $\mathcal {X}\subseteq \mathbf {D}(f)$ , set $\mathcal {X}^f:=\{x^f\colon x\in \mathcal {X}\}$ .

  • Given $\mathcal {X}\subseteq \mathcal {L}$ , set

    $$\begin{align*}N_{\mathcal{L}}(\mathcal{X}):=\{f\in\mathcal{L}\colon \mathcal{X}\subseteq\mathbf {D}(f)\mbox{ and }\mathcal{X}^f=\mathcal{X}\}\end{align*}$$
    and
    $$\begin{align*}C_{\mathcal{L}}(\mathcal{X}):=\{f\in\mathcal{L}\colon x\in\mathbf {D}(f)\mbox{ and }x^f=x\mbox{ for all }x\in\mathcal{X}\}.\end{align*}$$
  • For $\mathcal {X},\mathcal {Y}\subseteq \mathcal {L}$ , define $N_{\mathcal {Y}}(\mathcal {X})=\mathcal {Y}\cap N_{\mathcal {L}}(\mathcal {X})$ and $C_{\mathcal {Y}}(\mathcal {X})=\mathcal {Y}\cap C_{\mathcal {L}}(\mathcal {X})$ .

We call $N_{\mathcal {Y}}(\mathcal {X})$ the normaliser of $\mathcal {X}$ in $\mathcal {Y}$ and $C_{\mathcal {Y}}(\mathcal {X})$ the centraliser of $\mathcal {X}$ in $\mathcal {Y}$ .

Definition 3.3.

  • A subset $\mathcal {H}\subseteq \mathcal {L}$ is called a partial subgroup of $\mathcal {L}$ if $h^{-1}\in \mathcal {H}$ for all $h\in \mathcal {H}$ , and moreover $\Pi (w)\in \mathcal {H}$ for all $w\in \mathbf {D}\cap \mathbf {W}(\mathcal {H})$ .

  • A partial subgroup $\mathcal {H}$ of $\mathcal {L}$ is a called a subgroup of $\mathcal {L}$ if $\mathbf {W}(\mathcal {H})\subseteq \mathbf {D}(\mathcal {L})$ .

  • By a p-subgroup of $\mathcal {L}$ , we mean a subgroup S of $\mathcal {L}$ such that $|S|$ is a power of p.

  • Let $\mathcal {N}$ be a partial subgroup of $\mathcal {L}$ . Then we call $\mathcal {N}$ a partial normal subgroup of $\mathcal {L}$ (and write $\mathcal {N}\unlhd \mathcal {L}$ ) if $n^f\in \mathcal {N}$ for all $f\in \mathcal {L}$ and all $n\in \mathcal {N}\cap \mathbf {D}(f)$ .

  • A partial subgroup $\mathcal {H}$ of $\mathcal {L}$ is called a partial subnormal subgroup of $\mathcal {L}$ if there exists a series $\mathcal {H}=\mathcal {H}_0\unlhd \mathcal {H}_1\unlhd \cdots \unlhd \mathcal {H}_k=\mathcal {L}$ of partial subgroups of $\mathcal {L}$ .

If $\mathcal {H}$ is a partial subgroup of $\mathcal {L}$ , notice that $\mathcal {H}$ is itself a partial group with product $\Pi |_{\mathbf {W}(\mathcal {H})\cap \mathbf {D}}$ . If $\mathcal {H}$ is a subgroup of $\mathcal {L}$ , then $\mathcal {H}$ forms a group with binary product defined by $x\cdot y:=\Pi (x,y)$ for all $x,y\in \mathcal {H}$ . In particular, every p-subgroup of $\mathcal {L}$ forms a p-group. The following definition will be crucial in the definition of a locality.

Definition 3.4. Let $\mathcal {L}$ be a partial group, and let $\Delta $ be a collection of subgroups of $\mathcal {L}$ . Define $\mathbf {D}_{\Delta }$ to be the set of words $w=(g_1, \dots , g_k) \in \mathbf {W}(\mathcal {L})$ such that there exist $P_0, \dots ,P_k \in \Delta $ with $P_{i-1} \subseteq \mathbf {D}(g_i)$ and $P_{i-1}^{g_i} = P_i$ for all $1 \leq i \leq k$ . If such w and $P_0,\dots ,P_k$ are given, then we also say that $w\in \mathbf {D}_{\Delta }$ via $P_0,P_1,\dots ,P_k$ , or just that $w\in \mathbf {D}_{\Delta }$ via $P_0$ .

Lemma 3.5. Let $\mathcal {L}$ be a partial group and $\mathcal {N}\unlhd \mathcal {L}$ .

  1. (a) If $f\in \mathcal {L}$ and $P\subseteq \mathbf {D}(f)$ , then $(P\cap \mathcal {N})^f=P^f\cap \mathcal {N}$ .

  2. (b) Let S be a subgroup of $\mathcal {L}$ and $\Gamma $ a set of subgroups of $\mathcal {N}\cap S$ . Set

    $$\begin{align*}\Delta:=\{P\leq S\colon P\cap \mathcal{N}\in\Gamma\}.\end{align*}$$
    Then $\mathbf {D}_{\Gamma }=\mathbf {D}_{\Delta }$ .

Proof. (a) Notice that $(P\cap \mathcal {N})^f\subseteq P^f\cap \mathcal {N}$ as $\mathcal {N}\unlhd \mathcal {L}$ . Since $(c_f)^{-1}=c_{f^{-1}}$ by [Reference Chermak8, Lemma 1.6(c)], we have $(P^f)^{f^{-1}}=P$ . So, similarly, $(P^f\cap \mathcal {N})^{f^{-1}}\subseteq P\cap \mathcal {N}$ , and thus $P^f\cap \mathcal {N}\subseteq (P\cap \mathcal {N})^f$ . This shows $(P\cap \mathcal {N})^f=P^f\cap \mathcal {N}$ as required.

(b) As $\Gamma \subseteq \Delta $ , we have $\mathbf {D}_{\Gamma }\subseteq \mathbf {D}_{\Delta }$ . Conversely, if $w=(f_1,\dots ,f_n)\in \mathbf {D}_{\Delta }$ via $P_0,P_1,\dots ,P_n\in \Delta $ , then (a) yields that $w\in \mathbf {D}_{\Gamma }$ via $P_0\cap \mathcal {N},P_1\cap \mathcal {N},\dots ,P_n\cap \mathcal {N}$ . Thus $\mathbf {D}_{\Delta } \subseteq \mathbf {D}_{\Gamma }$ , implying $\mathbf {D}_{\Gamma }=\mathbf {D}_{\Delta }$ .

3.2 Localities

Definition 3.6. Let $\mathcal {L}$ be a partial group, and let S be a p-subgroup of $\mathcal {L}$ . For $f\in \mathcal {L}$ , set

$$\begin{align*}S_f:=\{x\in S\colon x\in\mathbf {D}(f)\mbox{ and }x^f\in S\}.\end{align*}$$

More generally, if $w=(f_1,\dots ,f_k)\in \mathbf {W}(\mathcal {L})$ , then write $S_w$ for the set of $s\in S$ such that there exists a sequence $s=s_0,\dots ,s_k$ of elements of S with $s_{i-1}\in \mathbf {D}(f_i)$ and $s_{i-1}^{f_i}=s_i$ for $i=1,\dots ,k$ .

Definition 3.7. Let $\mathcal {L}$ be a finite partial group with product $\Pi \colon \mathbf {D}\rightarrow \mathcal {L}$ , let S be a p-subgroup of $\mathcal {L}$ , and let $\Delta $ be a nonempty set of subgroups of S. We say that $(\mathcal {L}, \Delta , S)$ is a locality if the following hold:

  1. 1. S is maximal with respect to inclusion among the p-subgroups of $\mathcal {L}$ ;

  2. 2. $\mathbf {D} = \mathbf {D}_{\Delta }$ ;

  3. 3. $\Delta $ is closed under passing to $\mathcal {L}$ -conjugates and overgroups in S; that is, $\Delta $ is overgroup-closed in S and $P^f\in \Delta $ for all $P\in \Delta $ and $f\in \mathcal {L}$ with $P \subseteq S_f$ .

If $(\mathcal {L},\Delta ,S)$ is a locality, then $\Delta $ is also called the set of objects of $(\mathcal {L},\Delta ,S)$ .

As argued in [Reference Henke16, Remark 5.2], Definition 3.7 is equivalent to the definition of a locality given by Chermak [Reference Chermak8, Definition 2.7]. We will use this fact throughout.

For the remainder of this subsection, let $(\mathcal {L},\Delta ,S)$ be a locality.

Lemma 3.8 (Important properties of localities).

The following hold:

  1. (a) $N_{\mathcal {L}}(P)$ is a subgroup of $\mathcal {L}$ for each $P\in \Delta $ .

  2. (b) Let $P\in \Delta $ and $g\in \mathcal {L}$ with $P\subseteq S_g$ . Then $Q:=P^g\in \Delta $ , $N_{\mathcal {L}}(P)\subseteq \mathbf {D}(g)$ , and

    $$ \begin{align*} c_g\colon N_{\mathcal{L}}(P)\rightarrow N_{\mathcal{L}}(Q) \end{align*} $$
    is an isomorphism of groups.
  3. (c) Let $w=(g_1,\dots ,g_n)\in \mathbf {D}$ via $(X_0,\dots ,X_n)$ . Then

    $$ \begin{align*} c_{g_1}\circ \dots \circ c_{g_n}=c_{\Pi(w)} \end{align*} $$
    is a group isomorphism $N_{\mathcal {L}}(X_0)\rightarrow N_{\mathcal {L}}(X_n)$ .
  4. (d) $S_g\in \Delta $ for every $g\in \mathcal {L}$ . In particular, $S_g$ is a subgroup of S. Moreover, $S_g^g=S_{g^{-1}}$ , and $c_g\colon S_g\rightarrow S$ is an injective group homomorphism.

  5. (e) For every $w\in \mathbf {W}(\mathcal {L})$ , $S_w$ is a subgroup of S. Moreover, $S_w\in \Delta $ if and only if $w\in \mathbf {D}$ .

  6. (f) If $w\in \mathbf {D}$ , then $S_w\leq S_{\Pi (w)}$ .

Proof. Properties (a),(b) and (c) correspond to the statements (a),(b) and (c) in [Reference Chermak8, Lemma 2.3] except for the fact stated in (b) that $Q\in \Delta $ , which is however clearly true if one uses the definition of a locality above. Property (d) holds by [Reference Chermak8, Proposition 2.5(a), (b)], and property (e) is stated in [Reference Chermak8, Corollary 2.6]. Property (f) follows from (e) and (c).

We use from now on without further reference that $c_g\colon S_g\rightarrow S$ is an injective group homomorphism for all $g\in \mathcal {L}$ . By $\mathcal {F}_S(\mathcal {L})$ , we denote the fusion system over S generated by all maps of this form. More generally, we have the following definition.

Definition 3.9. Let $\mathcal {H}$ be a partial subgroup of $\mathcal {L}$ . Write $\mathcal {F}_{S\cap \mathcal {H}}(\mathcal {H})$ for the fusion system over $S\cap \mathcal {H}$ generated by the injective group homomorphisms of the form $c_h\colon S_h\cap \mathcal {H}\rightarrow S\cap \mathcal {H}$ with $h\in \mathcal {H}$ .

Lemma 3.10. Let $\mathcal {H}$ be a partial subgroup of $\mathcal {L}$ and $P\in \Delta $ with $P\leq S\cap \mathcal {H}$ . Set $\mathcal {D}:=\mathcal {F}_{S\cap \mathcal {H}}(\mathcal {H})$ . Then the following hold:

  1. (a) For every $\varphi \in \mathrm {Hom}_{\mathcal {D}}(P,S\cap \mathcal {H})$ , there exists $h\in \mathcal {H}$ with $P\leq S_h$ and $\varphi =c_h$ .

  2. (b) $N_{\mathcal {D}}(P)=\mathcal {F}_{N_{S\cap \mathcal {H}}(P)}(N_{\mathcal {H}}(P))$ .

Proof. For part (a), see [Reference Henke18, Lemma 3.11(b)]. Set $T:=S\cap \mathcal {H}$ . Clearly, $\mathcal {F}_{N_T(P)}(N_{\mathcal {H}}(P))\subseteq N_{\mathcal {D}}(P)$ and $N_{\mathcal {D}}(P)$ is a fusion system over $N_T(P)$ . If $\varphi \in \mathrm {Hom}_{N_{\mathcal {D}}(P)}(X,Y)$ , where $X,Y\leq N_T(P)$ is a morphism in $N_{\mathcal {D}}(P)$ , then we may assume that $P\leq X\cap Y$ and $P\varphi =P$ . As $P\in \Delta $ , we have $X,Y\in \Delta $ . Hence, by (a), there exists $h\in \mathcal {H}$ such that $X\leq S_h$ and $\varphi =c_h|_X$ . As $P\varphi =P$ , it follows $h\in N_{\mathcal {H}}(P)$ , and thus $\varphi $ is a morphism in $\mathcal {F}_{N_T(P)}(N_{\mathcal {H}}(P))$ , as required.

Definition 3.11. If $(\mathcal {L},\Delta ,S)$ is a locality, then set

$$\begin{align*}O_p(\mathcal{L}):=\bigcap\{S_w\colon w\in\mathbf {W}(\mathcal{L})\}.\end{align*}$$

Lemma 3.12. If $(\mathcal {L},\Delta ,S)$ is a locality, then $O_p(\mathcal {L})$ is the unique largest p-subgroup of $\mathcal {L}$ that is a partial normal subgroup of $\mathcal {L}$ . Moreover, a subgroup $P\leq S$ is a partial normal subgroup of $\mathcal {L}$ if and only if $N_{\mathcal {L}}(P)=\mathcal {L}$ .

Proof. This is proved as [Reference Henke18, Lemma 3.13] based on [Reference Chermak8, Lemma 2.13].

We will use the characterisation of $O_p(\mathcal {L})$ given in Lemma 3.12 most of the time from now on without further reference.

Lemma 3.13. Let $(\mathcal {L},\Delta ,S)$ be a locality, $\mathcal {N}\unlhd \mathcal {L}$ and $T:=\mathcal {N}\cap S$ . Then for every $g\in \mathcal {L}$ , there exist $n\in \mathcal {N}$ and $f\in N_{\mathcal {L}}(T)$ such that $(n,f)\in \mathbf {D}$ , $g=nf$ and $S_g=S_{(n,f)}$ .

Proof. By the Frattini Lemma [Reference Chermak8, Corollary 3.11], there exist $n\in \mathcal {N}$ and $f\in \mathcal {L}$ such that $(n,f)\in \mathbf {D}$ , $g=nf$ and f is $\uparrow $ -maximal (in the sense of [Reference Chermak8, Definition 3.6]). By [Reference Chermak8, Proposition 3.9] and then [Reference Chermak8, Lemma 3.1(a)], every $\uparrow $ -maximal element is in $N_{\mathcal {L}}(T)$ . Moreover, the Splitting Lemma [Reference Chermak8, Lemma 3.12] gives $S_g=S_{(n,f)}$ .

Definition 3.14. Let $(\mathcal {L}^+,\Delta ^+,S)$ be a locality with a partial product $\Pi ^+\colon \mathbf {D}^+\longrightarrow \mathcal {L}^+$ . Suppose that $\emptyset \neq \Delta \subseteq \Delta ^+$ such that $\Delta $ is $\mathcal {F}_S(\mathcal {L}^+)$ -closed. Set

$$\begin{align*}\mathcal{L}^+|_{\Delta}:=\{f\in\mathcal{L}^+\colon S_f\in\Delta\}.\end{align*}$$

Note that $\mathbf {D}:=\mathbf {D}_{\Delta }\subseteq \mathbf {D}^+\cap \mathbf {W}(\mathcal {L}^+|_{\Delta })$ and, by Lemma 3.8(c), $\Pi ^+(w)\in \mathcal {L}|_{\Delta }$ for all $w\in \mathbf {D}$ . We call $\mathcal {L}:=\mathcal {L}^+|_{\Delta }$ together with the partial product $\Pi ^+|_{\mathbf {D}}\colon \mathbf {D}\longrightarrow \mathcal {L}$ and the restriction of the inversion map on $\mathcal {L}^+$ to $\mathcal {L}$ the restriction of $\mathcal {L}^+$ to $\Delta $ .

With the hypothesis and notation as in the preceding definition, it turns out that the restriction of $\mathcal {L}^+$ to $\Delta $ forms a partial group and the triple $(\mathcal {L}^+|_{\Delta },\Delta ,S)$ is a locality (see [Reference Chermak7, Lemma 2.21(a)] and [Reference Henke17, Lemma 2.23(a), (c)] for details). It might be worth pointing out that in the definition of the restriction, $S_f$ and $\mathbf {D}_{\Delta }$ are a priori formed inside of $\mathcal {L}^+$ , but as argued in [Reference Henke17, Lemma 2.23(b)], it does not matter whether one forms $S_f$ and $\mathbf {D}_{\Delta }$ inside of $\mathcal {L}^+$ or inside of the partial group $\mathcal {L}^+|_{\Delta }$ .

3.3 Homomorphisms of partial groups and projections

Definition 3.15. Let $\mathcal {L}$ and $\mathcal {L}'$ be partial groups with products $\Pi \colon \mathbf {D}\rightarrow \mathcal {L}$ and $\Pi '\colon \mathbf {D}'\rightarrow \mathcal {L}'$ , respectively. Let $\alpha \colon \mathcal {L}\rightarrow \mathcal {L}',f\mapsto f\alpha $ be a map.

  • Write $\alpha ^*$ for the induced map:

    $$\begin{align*}\alpha^*\colon \mathbf {W}(\mathcal{L})\rightarrow \mathbf {W}(\mathcal{L}'), (f_1,\dots,f_n)\mapsto(f_1\alpha,\dots,f_n\alpha).\end{align*}$$
  • We call $\alpha $ a homomorphism of partial groups if $\mathbf {D}\alpha ^*\subseteq \mathbf {D}'$ and $\Pi (w)\alpha =\Pi '(w\alpha ^*)$ .

  • The map $\alpha $ is called a projection of partial groups if $\alpha $ is a homomorphism of partial groups and $\mathbf {D}\alpha ^*=\mathbf {D}'$ .

  • A bijective projection of partial groups is called an isomorphism of partial groups and an isomorphism from $\mathcal {L}$ to itself is called an automorphism of $\mathcal {L}$ . Write $\mathrm {Aut}(\mathcal {L})$ for the set of automorphisms of $\mathcal {L}$ .

  • If $\alpha \colon \mathcal {L}\rightarrow \mathcal {L}'$ is a homomorphism of partial groups and $\operatorname {\mathbf {1}}'$ denotes the identity in $\mathcal {L}'$ , then $\ker (\alpha ):=\{f\in \mathcal {L}\colon f\alpha =\operatorname {\mathbf {1}}'\}$ is called the kernel of $\alpha $ .

If $\mathcal {L}'$ is a partial group, then $\mathcal {L}'\subseteq \mathbf {D}'$ . Hence, a projection of partial groups $\mathcal {L}\rightarrow \mathcal {L}'$ is always surjective.

If $\alpha \colon \mathcal {L}\rightarrow \mathcal {L}'$ is a homomorphism of partial groups, then $\ker (\alpha )\unlhd \mathcal {L}$ by [Reference Chermak8, Lemma 1.14]. If $(\mathcal {L},\Delta ,S)$ is a locality and $\mathcal {N}\unlhd \mathcal {L}$ , then conversely, one can construct a partial group $\mathcal {L}'$ and a projection of partial groups $\mathcal {L}\rightarrow \mathcal {L}'$ with kernel $\mathcal {N}$ . Namely, define a coset of $\mathcal {N}$ in $\mathcal {L}$ to be a subset of the form

$$\begin{align*}\mathcal{N} f:=\{\Pi(n,f)\colon n\in\mathcal{N},\;f\in\mathcal{L}\mbox{ such that }(n,f)\in\mathbf {D}\}.\end{align*}$$

Call a coset maximal if it is maximal with respect to inclusion among all cosets of $\mathcal {N}$ in $\mathcal {L}$ . Write $\mathcal {L}/\mathcal {N}$ for the set of maximal cosets of $\mathcal {N}$ in $\mathcal {L}$ . By [Reference Chermak8, Proposition 3.14(d)], $\mathcal {L}/\mathcal {N}$ forms a partition of $\mathcal {L}$ . Thus there is a natural map

$$\begin{align*}\alpha\colon\mathcal{L}\rightarrow \mathcal{L}/\mathcal{N}\end{align*}$$

that sends every element $f\in \mathcal {L}$ to the unique maximal coset containing f. It turns out that the set $\mathcal {L}/\mathcal {N}$ can be (in a unique way) given the structure of a partial group such that the map $\alpha $ above is a projection of partial groups (compare [Reference Chermak8, Lemma 3.16]). We therefore call the map $\alpha $ the natural projection from $\mathcal {L}$ to $\mathcal {L}/\mathcal {N}$ . The identity element of this partial group is the maximal coset $\mathcal {N}=\mathcal {N}\operatorname {\mathbf {1}}$ . In particular, $\mathcal {N}$ is the kernel of the natural projection.

Given a locality $(\mathcal {L},\Delta ,S)$ and $\mathcal {N}\unlhd \mathcal {L}$ , we will sometimes write $\overline {\mathcal {L}}$ for the set $\mathcal {L}/\mathcal {N}$ . We mean then implicitly that we use a kind of ‘bar notation’ similar to the way it is commonly used for groups. Namely, if X is an element or subset of $\mathcal {L}$ , then $\overline {X}$ denotes the image of X under the natural projection $\alpha \colon \mathcal {L}\rightarrow \overline {\mathcal {L}}$ . Later, we will consider cases where $\mathcal {N}\cap S=\operatorname {\mathbf {1}}$ . Since $\mathcal {N}$ equals the kernel of $\alpha $ , the restriction $\alpha |_S\colon S\rightarrow \overline {S}$ is then a bijection. Thus we will be able to identify S with $\overline {S}$ .

Definition 3.16. Let $(\mathcal {L}, \Delta , S)$ and $(\mathcal {L}', \Delta ', S')$ be localities, and let $\alpha \colon \mathcal {L} \rightarrow \mathcal {L}'$ be a projection of partial groups. We say that $\alpha $ is a projection of localities from $(\mathcal {L},\Delta ,S)$ to $(\mathcal {L}',\Delta ',S')$ if the set $\Delta \alpha := \{P\alpha \mid P \in \Delta \}$ equals $\Delta '$ .

If $\alpha $ is a projection of localities from $(\mathcal {L},\Delta ,S)$ to $(\mathcal {L}',\Delta ',S')$ , then notice that $\alpha $ maps S to $S'$ , as S and $S'$ are the unique maximal elements of $\Delta $ and $\Delta '$ respectively.

3.4 Localities with special properties

The following definition was partly introduced in the introduction. Recall that a finite group G is of characteristic p if $C_G(O_p(G))\leq O_p(G)$ .

Definition 3.17. Let $(\mathcal {L},\Delta ,S)$ be a locality.

  • We call $(\mathcal {L},\Delta ,S)$ cr-complete if $\mathcal {F}_S(\mathcal {L})^{cr}\subseteq \Delta $ .

  • A locality $(\mathcal {L},\Delta ,S)$ is said to be of objective characteristic p if $N_{\mathcal {L}}(P)$ is of characteristic p for every $P \in \Delta $ .

  • A locality is called a linking locality if it is cr-complete and of objective characteristic p.

Proposition 3.18. Let $(\mathcal {L},\Delta ,S)$ be a locality.

  1. (a) $\mathcal {F}_S(\mathcal {L})$ is $\Delta $ -generated and $\Delta $ -saturated.

  2. (b) If every $\mathcal {F}_S(\mathcal {L})$ -critical subgroup is an element of $\Delta $ , then $\mathcal {F}_S(\mathcal {L})$ is saturated and $(\mathcal {L},\Delta ,S)$ is cr-complete.

  3. (c) If $(\mathcal {L},\Delta ,S)$ is cr-complete, then $\mathcal {F}_S(\mathcal {L})$ is saturated. Hence, the fusion system of a linking locality is saturated.

Proof. The fusion system $\mathcal {F}_S(\mathcal {L})$ is $\Delta $ -generated by definition and $\Delta $ -saturated by [Reference Chermak8, Lemma 2.9] and [Reference Henke18, Lemma 3.27]. Thus Theorem 2.7 implies that $\mathcal {F}_S(\mathcal {L})$ is saturated if $\Delta $ contains every $\mathcal {F}_S(\mathcal {L})$ -critical subgroup. Part (b) follows now from Lemma 2.6(b) and part (c) from Lemma 2.6(a). Alternatively, (c) is proved in [Reference Henke18, Theorem 3.26].

It is of crucial importance that cr-complete localities (and thus linking localities) lead to saturated fusion systems. This fact was first observed by Chermak. Conversely, it follows from the existence and uniqueness of centric linking systems (see [Reference Chermak7, Reference Oliver24, Reference Glauberman and Lynd11]) that there is a linking locality attached to every saturated fusion system. In this context, the following definition plays a crucial role.

Definition 3.19 [Reference Henke16, Definition 1, Lemma 3.1].

Let $\mathcal {F}$ be a saturated fusion system. A subgroup $P\leq S$ is called $\mathcal {F}$ -subcentric if $N_{\mathcal {F}}(Q)$ is constrained for every fully $\mathcal {F}$ -normalised $\mathcal {F}$ -conjugate Q of P. We denote by $\mathcal {F}^s$ the set of $\mathcal {F}$ -subcentric subgroups of S.

It can be shown that for every $\mathcal {F}$ -closed set $\Delta $ of subgroups of S with $\mathcal {F}^{cr}\subseteq \Delta \subseteq \mathcal {F}^s$ , there exists an (essentially unique) linking locality over $\mathcal {F}$ with object set $\Delta $ . Moreover, $\mathcal {F}^s$ is $\mathcal {F}$ -closed, and thus there is an (essentially unique) linking locality over $\mathcal {F}$ with object set $\mathcal {F}^s$ , which can be seen as the ‘largest’ linking locality over $\mathcal {F}$ . The reader is referred to [Reference Henke16, Theorem A] for a precise statement of these results.

3.5 Regular localities

Regular localities were introduced by Chermak [Reference Chermak9], but we will refer to the treatment of the subject in [Reference Henke18]. If $(\mathcal {L},\Delta ,S)$ is a linking locality over a saturated fusion system $\mathcal {F}$ , then there is the generalised Fitting subgroup $F^*(\mathcal {L})$ of $\mathcal {L}$ defined as a certain partial normal subgroup of $\mathcal {L}$ (compare [Reference Henke18, Definition 3]). It turns out that the set $\{P\leq S\colon P\cap F^*(\mathcal {L})\in \mathcal {F}^s\}$ depends only on $\mathcal {F}$ and not on the linking locality $(\mathcal {L},\Delta ,S)$ (see [Reference Henke18, Lemma 10.2]). Thus the following definition makes sense.

Definition 3.20. If $\mathcal {F}$ is a saturated fusion system and $(\mathcal {L},\Delta ,S)$ is a linking locality over $\mathcal {F}$ , then set

$$\begin{align*}\delta(\mathcal{F}):=\{P\leq S\colon P\cap F^*(\mathcal{L})\in\mathcal{F}^s\}. \end{align*}$$

A linking locality $(\mathcal {L},\Delta ,S)$ is called a regular locality if $\Delta =\delta (\mathcal {F})$ .

For every saturated fusion system $\mathcal {F}$ , the set $\delta (\mathcal {F})$ is $\mathcal {F}$ -closed, and thus there exists a regular locality over $\mathcal {F}$ (compare [Reference Henke18, Lemma 10.4]).

It turns out that regular localities have particularly nice properties. To describe these, suppose that $(\mathcal {L},\Delta ,S)$ is a regular locality. If $\mathcal {H}$ is a partial normal subgroup of $\mathcal {L}$ or, more generally, a partial subnormal subgroup of $\mathcal {L}$ , then by [Reference Henke18, Theorem 3, Corollary 10.19], $\mathcal {E}:=\mathcal {F}_{S\cap \mathcal {H}}(\mathcal {H})$ is saturated and $(\mathcal {H},\delta (\mathcal {E}),S\cap \mathcal {H})$ is a regular locality over $\mathcal {E}$ . This leads to a natural notion of components of $\mathcal {L}$ (see [Reference Henke18, Definition 11.1]) and to a theory of components of regular localities that mirrors results from finite group theory (see [Reference Henke18, Chapter 11]). In particular, the product $E(\mathcal {L})$ of components of $\mathcal {L}$ forms a partial normal subgroup with $F^*(\mathcal {L})=E(\mathcal {L})O_p(\mathcal {L})$ .

If $(\mathcal {L},\Delta ,S)$ is a regular locality over $\mathcal {F}$ (or, more generally, a linking locality over $\mathcal {F}$ with $\delta (\mathcal {F})\subseteq \Delta $ ), then it is shown in [Reference Chermak and Henke10, Theorem A] that the assignment $\mathcal {N}\mapsto \mathcal {F}_{S\cap \mathcal {N}}(\mathcal {N})$ defines a bijection $\Phi $ from the set of partial normal subgroups of $\mathcal {L}$ to the set of normal subsystems of $\mathcal {F}$ . This bijection sends $F^*(\mathcal {L})$ to $F^*(\mathcal {F})$ and $E(\mathcal {L})$ to $E(\mathcal {F})$ . Moreover, in the case that $(\mathcal {L},\Delta ,S)$ is regular, we have $C_{\mathcal {L}}(\mathcal {N})\unlhd \mathcal {L}$ and $\Phi (C_{\mathcal {L}}(\mathcal {N}))=C_{\mathcal {F}}(\Psi (\mathcal {N}))$ (see [Reference Henke18, Theorem 3] and [Reference Chermak and Henke10, Proposition 6.7]). We refer the reader here to [Reference Aschbacher3, p.3] or [Reference Chermak and Henke10, Definition 7.2] for the definitions of $F^*(\mathcal {F})$ and $E(\mathcal {F})$ , and to [Reference Aschbacher3, Chapter 6] or [Reference Henke15] for a definition of the centraliser $C_{\mathcal {F}}(\mathcal {E})$ of a normal subsystem $\mathcal {E}$ .

The results we just summarised are used to prove our next lemma, which collects most of the information we need in Section 6. It will be convenient to use the following notation.

Notation 3.21. If $\mathcal {F}$ is a fusion system over S and $\mathcal {E}$ is a subsystem of $\mathcal {F}$ over $T\leq S$ , then we set $\mathcal {E}\cap P:=T\cap P$ for every subgroup $P\leq S$ .

Lemma 3.22. Let $(\mathcal {L},\Delta ,S)$ be a regular locality over $\mathcal {F}$ and $\mathcal {N}\unlhd \mathcal {L}$ . Set $T:=\mathcal {N}\cap S$ , $\mathcal {E}:=\mathcal {F}_T(\mathcal {N})$ , $T^*:=F^*(\mathcal {L})\cap S$ , $T_0:=E(\mathcal {N})\cap S$ and $R:=E(C_{\mathcal {L}}(\mathcal {N}))\cap S$ . Then the following hold:

  1. (a) $N_{\mathcal {L}}(T^*)=N_{\mathcal {L}}(T_0)\cap N_{\mathcal {L}}(R)$ and $N_{\mathcal {N}}(T^*)=N_{\mathcal {N}}(T_0)$ .

  2. (b) We have $T_0=E(\mathcal {E})\cap S$ and $R=E(C_{\mathcal {F}}(\mathcal {E}))\cap S$ .

  3. (c) $N_{\mathcal {E}}(T_0)=\mathcal {F}_T(N_{\mathcal {N}}(T_0))$ .

  4. (d) $\delta (\mathcal {E})$ is closed under passing to $\mathcal {L}$ -conjugates in S.

Proof. By [Reference Henke18, Theorem 3], we have $C_{\mathcal {L}}(\mathcal {N})\unlhd \mathcal {L}$ ; moreover, $(\mathcal {N},\delta (\mathcal {E}),T)$ is a regular locality over $\mathcal {E}$ , and $C_{\mathcal {L}}(\mathcal {N})$ can also be given the structure of a regular locality. In particular, $E(\mathcal {N})$ and $E(C_{\mathcal {L}}(\mathcal {N}))$ are well-defined.

(a) By [Reference Henke18, Lemma 11.13], $E(\mathcal {N})\unlhd \mathcal {L}$ and similarly $E(C_{\mathcal {L}}(\mathcal {N}))\unlhd \mathcal {L}$ . Moreover, by [Reference Henke18, Lemma 11.16], $E(\mathcal {L})=E(\mathcal {N})E(C_{\mathcal {L}}(\mathcal {N}))$ , which by [Reference Henke14, Theorem 1] implies $T_0R=E(\mathcal {L})\cap S\leq T^*$ . In particular, $N_{\mathcal {L}}(T^*)$ normalises $T_0=T^*\cap E(\mathcal {N})$ and $R=T^*\cap E(C_{\mathcal {L}}(\mathcal {N}))$ . On the other hand, by [Reference Henke18, Lemma 11.9], we have $T^*=(E(\mathcal {L})\cap S)O_p(\mathcal {L})$ , so $N_{\mathcal {L}}(T_0)\cap N_{\mathcal {L}}(R)\leq N_{\mathcal {L}}(T^*)$ . Thus, $N_{\mathcal {L}}(T^*)=N_{\mathcal {L}}(T_0)\cap N_{\mathcal {L}}(R)$ . As $R\subseteq C_{\mathcal {L}}(\mathcal {N})$ , it follows from [Reference Henke18, Lemma 3.5] that $\mathcal {N}\subseteq C_{\mathcal {L}}(R)\subseteq N_{\mathcal {L}}(R)$ . Hence, we can conclude that $N_{\mathcal {N}}(T_0)=\mathcal {N}\cap N_{\mathcal {L}}(T_0)\cap N_{\mathcal {L}}(R)=\mathcal {N}\cap N_{\mathcal {L}}(T^*)=N_{\mathcal {N}}(T^*)$ . This proves (a).

(b) By [Reference Chermak and Henke10, Proposition 6.7], we have $\mathcal {F}_{C_S(\mathcal {N})}(C_{\mathcal {L}}(\mathcal {N}))=C_{\mathcal {F}}(\mathcal {E})$ . This means $(C_{\mathcal {L}}(\mathcal {N}),\delta (C_{\mathcal {F}}(\mathcal {E})),C_S(\mathcal {N}))$ is a regular locality over $C_{\mathcal {F}}(\mathcal {E})$ . Recall also that $(\mathcal {N},\delta (\mathcal {E}),T)$ is a regular locality over $\mathcal {E}$ . Hence, [Reference Chermak and Henke10, Theorem E(d)] yields (b).

(c) By [Reference Henke18, Corollary 11.10], $T_0\in \delta (\mathcal {E})$ . Hence, (c) follows from Lemma 3.10(b) applied with $(\mathcal {N},\delta (\mathcal {E}),T)$ in place of $(\mathcal {L},\Delta ,S)$ .

(d) The set $\delta (\mathcal {E})$ is $\mathcal {E}$ -closed by [Reference Henke18, Lemma 10.4]. Moreover, it follows from [Reference Henke18, Theorem 10.16(f)] that $\delta (\mathcal {E})$ is closed under passing to $N_{\mathcal {L}}(T)$ -conjugates. By Lemma 3.13, for every $g\in \mathcal {L}$ , there exist $n\in \mathcal {N}$ and $f\in N_{\mathcal {L}}(T)$ such that $(n,f)\in \mathbf {D}$ , $g=nf$ and $S_g=S_{(n,f)}$ . Now (d) follows using Lemma 3.8(c),(e).

4 Some additional lemmas

4.1 Lemmas on groups

Lemma 4.1. Let G be a finite group, $S\in \mathrm {Syl}_p(G)$ a Sylow p-subgroup of G and $N\mathrel {\unlhd } G$ a normal subgroup of G. If $H\leq G$ is such that $S \cap H \in \mathrm {Syl}_p(H)$ , then

$$\begin{align*}(S \cap N)(S \cap H) \in \mathrm{Syl}_p(NH).\end{align*}$$

Proof. Notice that $P:=(S \cap N)(S \cap H)\leq S \cap NH$ and $|P| = \frac {|S \cap N||S \cap H|}{|S \cap N \cap H|}$ . By assumption, $S \cap H \in \mathrm {Syl}_p(H)$ . As $N\unlhd G$ and $N\cap H\unlhd H$ , it follows that $S\cap N\in \mathrm {Syl}_p(N)$ and $S\cap N\cap H\in \mathrm {Syl}_p(N\cap H)$ . The assertion follows now from $|NH|=\frac {|N|\;|H|}{|N\cap H|}$ .

Recall that a finite group G is said to be of characteristic p if $ C_G(O_p(G)) \leq O_p(G)$ .

Lemma 4.2. Let G be a finite group of characteristic p. Then the following hold:

  1. (a) $N_G(P)$ and $ C_G(P)$ are of characteristic p for all nontrivial p-subgroups $P \leq G$ .

  2. (b) Every subnormal subgroup of G is of characteristic p.

  3. (c) If N is a normal subgroup of G of p-power index, then G is of characteristic p if and only if N is of characteristic p.

Proof. Let P be a p-subgroup of G. By [Reference Meierfrankenfeld and Stellmacher22, Lemma 1.2(a), (c)], every subnormal subgroup of G is of characteristic p, and $N_G(P)$ is of characteristic p. In particular, $ C_G(P) \mathrel {\unlhd } N_G(P)$ is of characteristic p.

Now let N be as in (c). By [Reference Meierfrankenfeld and Stellmacher22, Lemma 1.3], G is of characteristic p if and only if $O^p(G)$ is of characteristic p. Similarly, N is of characteristic p if and only if $O^p(N)$ is of characteristic p. As $O^p(N)=O^p(G)$ , this implies (c).

Lemma 4.3. Let G be a finite group, and let N be a normal subgroup of G with Sylow p-subgroup T. Then the following are equivalent:

  1. (i) G is of characteristic p.

  2. (ii) N and $N_G(T)$ are of characteristic p.

  3. (iii) N and $C_G(N)$ are of characteristic p.

Proof. It follows from Lemma 4.2 that (i) implies (ii). As N is normal in G, $C_G(N)$ is also normal in G. Notice that $C_G(N)\leq C_G(T)\leq N_G(T)$ . Thus $C_G(N)$ is a normal subgroup of $N_G(T)$ . So by Lemma 4.2(b), $C_G(N)$ has characteristic p if $N_G(T)$ has characteristic p. Hence, (ii) implies (iii). Assume now that (iii) holds. It follows from [Reference Kurzweil and Stellmacher21, 6.5.2] that every component of G is a component either of N or of $C_G(N)$ . As N and $C_G(N)$ do not have any components, it follows that $E(G)=1$ . Moreover, $[N,O_{p^{\prime }}(G)]\leq N\cap O_{p^{\prime }}(G)=O_{p^{\prime }}(N)=1$ . Thus $O_{p^{\prime }}(G)\leq C_G(N)$ , so $O_{p^{\prime }}(G)=O_{p^{\prime }}(C_G(N))=1$ . Hence, (i) holds, and the proof is complete.

4.2 Some properties of fusion systems

It will be convenient to use the following definition.

Definition 4.4. Suppose $\mathcal {F}$ and $\tilde {\mathcal {F}}$ are fusion systems over S and $\tilde {S}$ , respectively.

  • If $\alpha \colon S\rightarrow \tilde {S}$ is a group isomorphism, then $\alpha $ is said to induce an isomorphism from $\mathcal {F}$ to $\tilde {\mathcal {F}}$ if $\{\alpha ^{-1} \varphi \alpha \colon \varphi \in \mathrm {Mor}_{\mathcal {F}}(P,Q)\}=\mathrm {Mor}_{\tilde {\mathcal {F}}}(P\alpha ,Q\alpha )$ for all $P,Q\leq S$ .

  • We say that $\alpha \in \mathrm {Aut}(S)$ induces an automorphism of $\mathcal {F}$ if $\alpha $ induces an isomorphism from $\mathcal {F}$ to $\mathcal {F}$ . Write $\mathrm {Aut}(\mathcal {F})$ for the set of all $\alpha \in \mathrm {Aut}(S)$ that induce an automorphism of $\mathcal {F}$ .

Lemma 4.5. Let $\mathcal {F}$ be a fusion system over S, and let $\mathcal {E}$ be an $\mathcal {F}$ -invariant subsystem of $\mathcal {F}$ . Then $\mathcal {E}^c$ , $\mathcal {E}^r$ and $\mathcal {E}^{cr}$ are closed under $\mathcal {F}$ -conjugacy. Similarly, the set of $\mathcal {E}$ -critical subgroups is closed under $\mathcal {F}$ -conjugacy.

Proof. Let $T\leq S$ be such that $\mathcal {E}$ is a subsystem over T. Let $\Gamma $ be one of the sets $\mathcal {E}^c$ , $\mathcal {E}^r$ , $\mathcal {E}^{cr}$ or the set of $\mathcal {E}$ -critical subgroups. It follows directly from the definition of these sets that $\Gamma $ is closed under $\mathcal {E}$ -conjugacy. By the Frattini condition for $\mathcal {F}$ -invariant subsystems [Reference Aschbacher, Kessar and Oliver2, Definition I.6.1], it is thus sufficient to argue that $\Gamma $ is $\mathrm {Aut}_{\mathcal {F}}(T)$ -invariant. The definition of $\mathcal {F}$ -invariant subsystems implies, moreover, that every element of $\mathrm {Aut}_{\mathcal {F}}(T)$ induces an automorphism of $\mathcal {E}$ . Therefore, $\alpha \in \mathrm {Aut}_{\mathcal {F}}(T)$ maps $\mathcal {E}$ -conjugacy classes to $\mathcal {E}$ -conjugacy classes and fully $\mathcal {E}$ -normalised subgroups to fully $\mathcal {E}$ -normalised subgroups. Moreover, if $P\leq T$ with $C_T(P)\leq P$ , then $C_T(P\alpha )\leq P\alpha $ . So $\mathcal {E}^c$ is $\mathrm {Aut}_{\mathcal {F}}(T)$ -invariant. Notice also that $\alpha |_{N_T(Q)}$ induces an isomorphism from $N_{\mathcal {E}}(Q)$ to $N_{\mathcal {E}}(Q\alpha )$ for every $Q\leq T$ . Hence, $\alpha $ maps $\mathcal {E}$ -radical subgroups to $\mathcal {E}$ -radical subgroups. So $\mathcal {E}^r$ and $\mathcal {E}^{cr}=\mathcal {E}^c\cap \mathcal {E}^r$ are $\mathrm {Aut}_{\mathcal {F}}(T)$ -invariant. For $Q\leq T$ , the map $\mathrm {Aut}_{\mathcal {E}}(Q)\rightarrow \mathrm {Aut}_{\mathcal {E}}(Q\alpha ),\varphi \mapsto \alpha ^{-1}\varphi \alpha $ is an isomorphism that takes $\mathrm {Inn}(Q)$ to $\mathrm {Inn}(Q\alpha )$ and $\mathrm {Aut}_T(Q)$ to $\mathrm {Aut}_T(Q\alpha )$ . Hence, the set of $\mathcal {E}$ -critical subgroups is $\mathrm {Aut}_{\mathcal {F}}(T)$ invariant. This proves the assertion.

For the remainder of this subsection, let $\mathcal {F}$ be a saturated fusion system over a p -group S.

In the proofs of the following two lemmas, we cite [Reference Andersen, Oliver and Ventura1]. It should be pointed out that normal subsystems in the sense of [Reference Andersen, Oliver and Ventura1, Definition 1.18] correspond to weakly normal subsystems in the language of [Reference Aschbacher, Kessar and Oliver2] (i.e., in the language that we are using in this paper).

Lemma 4.6. Let $\mathcal {E}$ be a weakly normal subsystem of $\mathcal {F}$ over T. If $R\in \mathcal {F}^{cr}$ , then $R\cap T\in \mathcal {E}^{cr}$ .

Proof. This is [Reference Andersen, Oliver and Ventura1, Lemma 1.20(d)] but also follows easily from the results stated before. Namely, by Corollary 2.3, there exists $\alpha \in \mathrm {Hom}_{\mathcal {F}}(N_S(R\cap T),S)$ such that $(R\cap T)^{\alpha }$ is fully normalised. Since $R\leq N_S(R\cap T)$ , the subgroup $R^{\alpha }$ is well-defined. As T is strongly closed, $R^{\alpha }\cap T=(R\cap T)^{\alpha }$ . By Lemma 4.5, $\mathcal {E}^{cr}$ and $\mathcal {F}^{cr}$ are closed under $\mathcal {F}$ -conjugacy. So, replacing R by $R^{\alpha }$ , we may assume that $R\cap T$ is fully $\mathcal {F}$ -normalised. If Q is an $\mathcal {E}$ -conjugate of $R\cap T$ , then by Corollary 2.3, there is an $\mathcal {F}$ -morphism mapping $N_S(Q)$ into $N_S(R\cap T)$ and thus $N_T(Q)$ into $N_T(R\cap T)$ . Hence, $R\cap T$ is fully $\mathcal {E}$ -normalised. Now the claim follows from Lemma 2.6(b) and Lemma 2.10.

Lemma 4.7. Let $\mathcal {E}$ be an $\mathcal {F}$ -invariant (not necessarily saturated) subsystem of $\mathcal {F}$ over $T\leq S$ . Let P be an $\mathcal {E}$ -critical subgroup of T such that P is fully $\mathcal {F}$ -normalised. Then there exists $R\in \mathcal {F}^{cr}$ with $R\cap T=P$ .

Proof. This is a special case of [Reference Andersen, Oliver and Ventura1, Lemma 1.19].

We will now look at some properties of the set $\delta (\mathcal {F})$ introduced in Subsection 3.5. As shown in [Reference Chermak and Henke10, Lemma 7.21], $\delta (\mathcal {F})$ can be characterised as the set of subgroups of S containing an element of $F^*(\mathcal {F})^s$ . The reader is referred to [Reference Aschbacher3] or [Reference Chermak and Henke10, Definition 7.2] for the definition of the generalised Fitting subsystem $F^*(\mathcal {F})$ and the layer $E(\mathcal {F})$ of $\mathcal {F}$ . Recall Notation 3.21.

Lemma 4.8. Let $P\leq S$ with $O_p(\mathcal {F})\leq P$ . Then $P\in \delta (\mathcal {F})$ if and only if $E(\mathcal {F})\cap P\in E(\mathcal {F})^s$ .

Proof. We use that $F^*(\mathcal {F})=E(\mathcal {F})*\mathcal {F}_{O_p(\mathcal {F})}(O_p(\mathcal {F}))$ by [Reference Chermak and Henke10, Theorem 7.10(e)]. In particular, $F^*(\mathcal {F})\cap S=(E(\mathcal {F})\cap S)O_p(\mathcal {F})$ . As $O_p(\mathcal {F})\leq P$ , a Dedekind Argument yields thus $F^*(\mathcal {F})\cap P=( E(\mathcal {F})\cap P)O_p(\mathcal {F})$ . It is shown in [Reference Chermak and Henke10, Lemma 7.21] that $P\in \delta (\mathcal {F})$ if and only if $F^*(\mathcal {F})\cap P\in F^*(\mathcal {F})^s$ . The assertion follows now from [Reference Henke18, Lemma 2.14(g)].

Lemma 4.9. Let $\mathcal {E}$ be a normal subsystem of $\mathcal {F}$ and $O_p(\mathcal {F})\leq P\leq S$ . Then the following hold:

  1. (a) If $E(\mathcal {F})=E(\mathcal {E})$ , then $P\in \delta (\mathcal {F})$ if and only if $ \mathcal {E}\cap P\in \delta (\mathcal {E})$ .

  2. (b) Suppose $E(C_{\mathcal {F}}(\mathcal {E}))\cap S\leq P$ . Then

    $$\begin{align*}P\in\delta(\mathcal{F})\Longleftrightarrow E(\mathcal{E})\cap P\in E(\mathcal{E})^s\Longleftrightarrow \mathcal{E}\cap P\in\delta(\mathcal{E}).\end{align*}$$

Proof. By Lemma 4.8, we have $P\in \delta (\mathcal {F})$ if and only if $E(\mathcal {F})\cap P\in E(\mathcal {F})^s$ . As $O_p(\mathcal {E})\leq O_p(\mathcal {F})\leq P$ by [Reference Henke16, Lemma 2.12(b)], it follows similarly that $\mathcal {E}\cap P\in \delta (\mathcal {E})$ if and only if $E(\mathcal {E})\cap P\in E(\mathcal {E})^s$ . This implies (a). By [Reference Chermak and Henke10, Lemma 7.13(c)], $E(\mathcal {F})=E(\mathcal {E})*E(C_{\mathcal {F}}(\mathcal {E}))$ . As $R:=E(C_{\mathcal {F}}(\mathcal {E}))\cap S\leq P$ , we have $E(\mathcal {F})\cap P=(E(\mathcal {E})\cap P)R$ . Notice that $R\in E(C_{\mathcal {F}}(\mathcal {E}))^s$ . Hence, by [Reference Henke18, Lemma 2.14(g)], $E(\mathcal {F})\cap P\in E(\mathcal {F})^s$ if and only if $E(\mathcal {E})\cap P\in E(\mathcal {E})^s$ . This implies (b).

Remark 4.10. If $\mathcal {E}$ is a normal subsystem of $\mathcal {F}$ defined over $T\leq S$ and R is a subgroup of S, then there is a unique saturated subsystem $\mathcal {D}$ of $\mathcal {F}$ over $TR$ with $O^p(\mathcal {D}) = O^p(\mathcal {E})$ . It is denoted by $(\mathcal {E} R)_{\mathcal {F}}$ (or sometimes simply by $\mathcal {E} R$ ). This was first shown by Aschbacher [Reference Aschbacher3, Chapter 8]. The result was revisited in [Reference Henke13], where a concrete description is given.

We conclude this section with the following lemma, which was first proved by Aschbacher [Reference Aschbacher4, 1.3.2]. We give a new proof using localities.

Lemma 4.11. Let $\mathcal {E}$ be a normal subsystem of $\mathcal {F}$ over $T\leq S$ . Then $\mathcal {F}=\langle \mathcal {E} S, N_{\mathcal {F}}(T)\rangle $ .

Proof. By [Reference Henke16, Theorem A], there exists a linking locality $(\mathcal {L},\Delta ,S)$ over $\mathcal {F}$ with $\Delta =\mathcal {F}^s$ . Moreover, by [Reference Chermak and Henke10, Theorem A], there exists a partial normal subgroup $\mathcal {N}$ of $\mathcal {L}$ with $T=S\cap \mathcal {N}$ and $\mathcal {E}=\mathcal {F}_T(\mathcal {N})$ . Now, by [Reference Chermak and Henke10, Corollary 4.10], we have $\mathcal {E} S=\mathcal {F}_S(\mathcal {N} S)$ . As $\mathcal {F}=\mathcal {F}_S(\mathcal {L})$ is generated by maps of the form $c_g\colon S_g\rightarrow S$ with $g\in \mathcal {L}$ , it is sufficient to prove that such maps are in $\langle \mathcal {E} S,N_{\mathcal {F}}(T)\rangle $ . Let $g\in \mathcal {L}$ . By Lemma 3.13, there exist $n\in \mathcal {N}$ and $f\in N_{\mathcal {L}}(T)$ such that $(n,f)\in \mathbf {D}$ , $g=nf$ and $P:=S_g=S_{(n,f)}$ . Hence, $c_g=(c_n|_P)(c_f|_{P^n})$ . Notice that $c_n|_P\in \mathcal {F}_S(\mathcal {N} S)=\mathcal {E} S$ and that $c_f|_{P^n}$ is a morphism in $N_{\mathcal {F}}(T)$ . This shows the assertion.

5 Kernels of localities

In this section, we investigate the properties of kernels of localities. In particular, we prove Theorems B, C and D. Recall the following definition.

Definition 5.1. Let $(\mathcal {L},\Delta ,S)$ be a locality. A kernel of $\mathcal {L}$ is a partial normal subgroup $\mathcal {N}$ of $\mathcal {L}$ such that $P\cap \mathcal {N}\in \Delta $ for every $P\in \Delta $ .

Lemma 5.2. Let $\mathcal {N}$ be a kernel of a locality $(\mathcal {L},\Delta ,S)$ . Set $\mathcal {F}=\mathcal {F}_S(\mathcal {L})$ , $T:=S\cap \mathcal {N}$ , $\mathcal {E}=\mathcal {F}_T(\mathcal {N})$ and $\Gamma :=\{P\in \Delta \colon P\leq T\}=\{P\cap T\colon P\in \Delta \}$ . Then the following hold:

  1. (a) $(\mathcal {N},\Gamma ,T)$ is a locality over $\mathcal {E}$ , and $\Gamma $ is closed under $\mathcal {F}$ -conjugacy.

  2. (b) For every $f\in N_{\mathcal {L}}(T)$ , $\mathcal {L}=\mathbf {D}(f)$ , $c_f\in \mathrm {Aut}(\mathcal {L})$ and $c_f|_{\mathcal {N}}\in \mathrm {Aut}(\mathcal {N})$ .

  3. (c) $O_p(\mathcal {N})=O_p(\mathcal {L})\cap \mathcal {N}\unlhd \mathcal {L}$ .

  4. (d) The subsystem $\mathcal {E}$ is $\mathcal {F}$ -invariant.

Proof. (a,b) It follows from [Reference Henke18, Lemma 3.29(b), (c)] and the definition of $\mathcal {E}$ that $(\mathcal {N},\Gamma ,T)$ is a locality over $\mathcal {E}$ and that (b) holds. Since $\Delta $ is closed under passing to $\mathcal {L}$ -conjugates in S, $\mathcal {N}\unlhd \mathcal {L}$ and $\Gamma =\{P\in \Delta \colon P\subseteq \mathcal {N}\}$ , it follows that $\Gamma $ is closed under passing to $\mathcal {L}$ -conjugates in S. So $\Gamma $ is closed under $\mathcal {F}$ -conjugacy, and (a) holds.

(c) Notice that $O_p(\mathcal {N})$ is invariant under conjugation by elements of $N_{\mathcal {L}}(T)$ as the elements of $N_{\mathcal {L}}(T)$ induce automorphisms of $\mathcal {N}$ . Since $O_p(\mathcal {N})\unlhd \mathcal {N}$ , it follows from [Reference Chermak8, Corollary 3.13] that $O_p(\mathcal {N})\unlhd \mathcal {L}$ and thus $O_p(\mathcal {N})\leq O_p(\mathcal {L})$ . Clearly, $O_p(\mathcal {L})\cap \mathcal {N}\leq O_p(\mathcal {N})$ , so (c) holds.

(d) As $\mathrm {Aut}_{\mathcal {F}}(T)$ is generated by maps of the form $c_f|_T$ with $f\in N_{\mathcal {L}}(T)$ and these maps induce automorphisms of $\mathcal {E}$ by [Reference Henke18, Lemma 3.29(d)], it follows that the elements of $\mathrm {Aut}_{\mathcal {F}}(T)$ induce automorphisms of $\mathcal {E}$ . Now (d) is a consequence of [Reference Henke18, Lemma 3.28].

Definition 5.3. We say that $(\mathcal {N},\Gamma ,T)$ is a kernel of $(\mathcal {L},\Delta ,S)$ to indicate that $\mathcal {N}$ is a kernel of $(\mathcal {L},\Delta ,S)$ , $T=S\cap \mathcal {N}$ and $\Gamma =\{P\in \Delta \colon P\leq T\}=\{P\cap T\colon P\in \Delta \}$ .

The following lemma can be seen as a converse to Lemma 5.2(a).

Lemma 5.4. Suppose we are given a finite partial group $\mathcal {L}$ with product $\Pi \colon \mathbf {D}\rightarrow \mathcal {L}$ . Let S be a maximal p-subgroup of $\mathcal {L}$ and $\mathcal {N}\unlhd \mathcal {L}$ . Set $T:=\mathcal {N}\cap S$ , let $\Gamma $ be a set of subgroups of T, and set $\Delta :=\{P\leq S\colon P\cap T\in \Gamma \}$ . Suppose the following hold:

  • $\Gamma $ is closed under passing to overgroups and $\mathcal {L}$ -conjugates that lie in T.

  • $\mathbf {D}=\mathbf {D}_{\Gamma }$ .

Then $(\mathcal {L},\Delta ,S)$ is a locality with kernel $(\mathcal {N},\Gamma ,T)$ .

Proof. If $(\mathcal {L},\Delta ,S)$ is a locality, then clearly $(\mathcal {N},\Gamma ,T)$ is a kernel of $(\mathcal {L},\Delta ,S)$ . By Lemma 3.5(b), we have $\mathbf {D}_{\Delta }=\mathbf {D}_{\Gamma }=\mathbf {D}$ . Recall that S is a maximal p-subgroup of $\mathcal {L}$ by assumption. As $\Gamma $ is overgroup-closed in T, it follows that $\Delta $ is overgroup-closed in S. Hence, it remains to show that $\Delta $ is closed under passing to $\mathcal {L}$ -conjugates in S.

Let $P\in \Delta $ and $f\in \mathcal {L}$ with $P\subseteq \mathbf {D}(f)$ and $P^f\subseteq S$ . Observe that $Q:=P\cap \mathcal {N}\in \Gamma $ is normal in P and $Q^f\in \Gamma $ by assumption. Hence, if $x,y\in P$ , then $u:=(f^{-1},x,f,f^{-1},y,f)\in \mathbf {D}=\mathbf {D}_{\Gamma }$ via $Q^f$ . Thus, by the axioms of a partial group, $x^fy^f=\Pi (u)=\Pi (f^{-1},x,y,f)=(xy)^f$ , so $c_f\colon P\rightarrow S,x\mapsto x^f$ is a group homomorphism. Therefore, $P^f$ is a subgroup of S. By Lemma 3.5(a), $P^f\cap T=P^f\cap \mathcal {N}=Q^f\in \Gamma $ , so $P^f\in \Delta $ .

We will use from now on without further reference that, by Lemma 5.2(a), $(\mathcal {N},\Gamma ,T)$ is a locality if $(\mathcal {N},\Gamma ,T)$ is a kernel of a locality $(\mathcal {L},\Delta ,S)$ . In particular, it makes sense to say that a kernel $(\mathcal {N},\Gamma ,T)$ is cr-complete.

Proposition 5.5. Let $(\mathcal {L},\Delta ,S)$ be a locality with kernel $(\mathcal {N},\Gamma ,T)$ . Then $(\mathcal {L},\Delta ,S)$ is cr-complete if and only if $(\mathcal {N},\Gamma ,T)$ is cr-complete.

Proof. Set $\mathcal {E}:=\mathcal {F}_T(\mathcal {N})$ and $\mathcal {F}:=\mathcal {F}_S(\mathcal {L})$ . By Lemma 5.2(a),(d), $(\mathcal {N},\Gamma ,T)$ is a locality and $\mathcal {E}$ is $\mathcal {F}$ -invariant.

Assume first that $(\mathcal {L},\Delta ,S)$ is cr-complete. Then in particular, $\mathcal {F}$ is saturated by Proposition 3.18(c). We show:

(5.1) $$ \begin{align} \mbox{If} \ Q\leq T \ \mbox{is} \ \mathcal{E}\mbox{-critical, then} \ Q\in\Gamma. \end{align} $$

For the proof, let Q be an $\mathcal {E}$ -critical subgroup of T. Then there exists a fully $\mathcal {F}$ -normalised $\mathcal {F}$ -conjugate P of Q, which by Lemma 4.5 is also $\mathcal {E}$ -critical. Hence, Lemma 4.7 implies the existence of an element $R\in \mathcal {F}^{cr}$ with $R\cap T=P$ . As $\mathcal {F}^{cr}\subseteq \Delta $ and $(\mathcal {N},\Gamma ,T)$ is a kernel of $(\mathcal {L},\Delta ,S)$ , it follows that $P\in \Gamma $ . Hence, Lemma 5.2(a) yields $Q\in \Gamma $ , and thus equation (5.1) holds. Proposition 3.18(b) implies now that $\mathcal {E}$ is saturated and $(\mathcal {N},\Gamma ,T)$ is cr-complete.

To prove the other implication, assume that $(\mathcal {N},\Gamma ,T)$ is cr-complete. Then by Proposition 3.18(c), $\mathcal {E}$ is saturated. Observe also that $\mathcal {E}^{cr}\subseteq \Gamma \subseteq \Delta $ . As $\mathcal {F}$ is $\Delta $ -saturated and $\Delta $ -generated by Proposition 3.18(a), it follows now from Theorem A that $\mathcal {F}$ is saturated. In particular, $\mathcal {E}$ is weakly normal in $\mathcal {F}$ . So Lemma 4.6 gives $R\cap T\in \mathcal {E}^{cr}\subseteq \Gamma \subseteq \Delta $ for every $R\in \mathcal {F}^{cr}$ . As $\Delta $ is overgroup-closed, this implies that $(\mathcal {L},\Delta ,S)$ is cr-complete.

Lemma 5.6. Let $(\mathcal {N},\Gamma ,T)$ be a kernel of a locality $(\mathcal {L},\Delta ,S)$ . Set $\mathcal {F}:=\mathcal {F}_S(\mathcal {L})$ and $\mathcal {E}:=\mathcal {F}_T(\mathcal {N})$ . Suppose $C_{\mathcal {N}}(T)\leq T$ . Then the following hold:

  1. (a) $N_{\mathcal {N}}(T)\subseteq N_{\mathcal {N}}(TC_S(T))$ .

  2. (b) If $\mathcal {E}$ and $\mathcal {F}$ are saturated, then $\mathcal {E}$ is normal in $\mathcal {F}$ .

Proof. (a) Notice that $T\in \Gamma \subseteq \Delta $ , and thus $N_{\mathcal {L}}(T)$ is a group. The assumption $C_{\mathcal {N}}(T)\leq T$ yields that $N_{\mathcal {N}}(T)$ is of characteristic p.

Let $g\in N_{\mathcal {N}}(T)$ , and set $P:=S_g$ . Then $T\leq P$ , so $P^g=P$ by [Reference Chermak8, Lemma 3.1(b)]. Hence, $g\in K:=N_{\mathcal {N}}(P)$ . Note that $N_{\mathcal {L}}(P)\leq N_{\mathcal {L}}(T)$ , so $K=N_{N_{\mathcal {N}}(T)}(P)\leq N_{\mathcal {L}}(T)$ . Calculating inside the group $N_{\mathcal {L}}(T)$ , we have

$$\begin{align*}[K,N_{C_S(T)}(P)]\leq \mathcal{N}\cap C_{\mathcal{L}}(T)=C_{\mathcal{N}}(T)\leq T,\end{align*}$$

where the last inclusion holds by assumption. Thus $[g,N_{C_S(T)}(P)]\leq T$ , so $N_{C_S(T)}(P)\leq S_g=P$ . Therefore, we have $N_{PC_S(T)}(P)=PN_{C_S(T)}(P)=P$ . As $PC_S(T)$ is a p-group, this implies $C_S(T)\leq P$ and thus g acts on T and $C_S(T)$ . Hence, (a) holds.

(b) Notice that part (a) implies $[TC_S(T),N_{\mathcal {N}}(T)]\leq (TC_S(T))\cap \mathcal {N}=T$ . Since $\mathrm {Aut}_{\mathcal {F}}(T)$ is generated by maps of the form $c_f|_T$ , with $f\in N_{\mathcal {L}}(T)$ , the extension condition as stated in [Reference Aschbacher, Kessar and Oliver2, Definition I.6.1] follows. Now (b) follows from Lemma 5.2(b).

Lemma 5.7. A projection from a locality $(\mathcal {L}^*,\Delta ^*,S^*)$ to a locality $(\mathcal {L},\Delta ,S)$ sends every kernel of $(\mathcal {L}^*,\Delta ^*,S^*)$ to a kernel of $(\mathcal {L},\Delta ,S)$ .

Proof. Let $\varphi \colon \mathcal {L}^* \rightarrow \mathcal {L}$ be a projection of localities, and let $\mathcal {N}^*$ be a kernel of $\mathcal {L}^*$ . Then $\mathcal {N}:=(\mathcal {N}^*)\varphi $ is a partial normal subgroup of $\mathcal {L}$ by [Reference Henke20, Lemma 2.5]. Let $P \in \Delta $ . Then $P=(P^*)\varphi $ for some $P^* \in \Delta ^*$ since $\Delta ^*\varphi =\Delta $ . Since $\mathcal {N}^*$ is a kernel of $\mathcal {L}^*$ , we have $\mathcal {N}^* \cap P^*\in \Delta ^*$ . Hence

$$\begin{align*}\mathcal{N}\cap P = (\mathcal{N}^*)\varphi \cap P^*\varphi\geq (\mathcal{N}^* \cap P^*)\varphi \in \Delta^*\varphi = \Delta.\end{align*}$$

As $\Delta $ is overgroup-closed, it follows that $\mathcal {N}\cap P\in \Delta $ , and thus $\mathcal {N}$ is a kernel of $\mathcal {L}$ .

For the next lemma, the reader might want to recall Definition 3.14. By $\mathcal {E}^q$ , we denote the quasicentric subgroups of a saturated fusion system $\mathcal {E}$ ; for the definition, see [Reference Aschbacher, Kessar and Oliver2, Definition III.4.5].

Lemma 5.8. Let $(\mathcal {L},\Delta ,S)$ be a locality over $\mathcal {F}$ with cr-complete kernel $(\mathcal {N},\Gamma ,T)$ . Set $\mathcal {E}:=\mathcal {F}_T(\mathcal {N})$ . Let $\mathcal {E}^{cr}\subseteq \Gamma _0\subseteq \Gamma \cap \mathcal {E}^q$ such that $\Gamma _0$ is $\mathcal {E}$ -closed and $\mathrm {Aut}_{\mathcal {F}}(T)$ -invariant, and let $\Delta _0$ be the set of overgroups in S of the elements of $\Gamma _0$ . Set

$$\begin{align*}\mathcal{L}_0:=\mathcal{L}|_{\Delta_0},\;\mathcal{N}_0:=\mathcal{N}\cap\mathcal{L}_0\mbox{ and }\Theta=\bigcup_{P\in\Gamma_0}O_{p^{\prime}}(N_{\mathcal{N}}(P)).\end{align*}$$

Then the following hold:

  1. (a) $(\mathcal {L}_0,\Delta _0,S)$ is a cr-complete locality over $\mathcal {F}$ with cr-complete kernel $(\mathcal {N}_0,\Gamma _0,T)$ . Moreover, $\mathcal {N}_0=\mathcal {N}|_{\Gamma _0}$ and $\mathcal {F}_T(\mathcal {N}_0)=\mathcal {E}$ .

  2. (b) $\Theta $ is a partial normal subgroup of $\mathcal {L}_0$ with $\Theta \cap S=1$ . In particular, setting $\overline {\mathcal {L}}_0=\mathcal {L}_0/\Theta $ and using the ‘bar notation’ as usual, S and T are naturally isomorphic to $\overline {S}$ and $\overline {T}$ via the restriction of the natural projection map $\mathcal {L}_0\rightarrow \overline {\mathcal {L}}_0$ .

  3. (c) Identify $\overline {S}$ and $\overline {T}$ with S and T. Then the following conditions hold:

    • $(\overline {\mathcal {L}}_0,\Delta _0,S)$ is a locality over $\mathcal {F}$ .

    • $(\overline {\mathcal {N}}_0,\Gamma _0,T)$ is a kernel of $(\overline {\mathcal {L}}_0,\Delta _0,S)$ , which is a linking locality over $\mathcal {E}$ .

Proof. (a) Recall from Lemma 5.2(d) that $\mathcal {E}$ is $\mathcal {F}$ -invariant. As $\Gamma _0$ is $\mathrm {Aut}_{\mathcal {F}}(T)$ -invariant and closed under $\mathcal {E}$ -conjugacy, it follows from the Frattini condition for $\mathcal {F}$ -invariant subsystems (compare [Reference Aschbacher, Kessar and Oliver2, Definition I.6.1]) that $\Gamma _0$ is closed under $\mathcal {F}$ -conjugacy. Thus, $\Delta _0$ is $\mathcal {F}$ -closed, $\mathcal {L}_0$ is well-defined, and $(\mathcal {L}_0,\Delta _0,S)$ is a locality. It is easy to observe that $\mathcal {N}_0$ is a partial normal subgroup of $\mathcal {L}_0$ . It follows from the definition of $\Delta _0$ that $(\mathcal {N}_0,\Gamma _0,T)$ is a kernel of $(\mathcal {L}_0,\Delta _0,S)$ and $\mathcal {N}_0=\mathcal {N}|_{\Gamma _0}$ (see also Lemma 3.5). As $\mathcal {E}^{cr}\subseteq \Gamma _0$ , we can conclude from Lemma 4.6 that $\mathcal {F}^{cr}\subseteq \Delta _0\subseteq \Delta $ . In particular, $\mathcal {F}$ and $\mathcal {E}$ are saturated, by Proposition 3.18(c). Note furthermore that for all $P\in \Delta $ , $N_{\mathcal {L}}(P)=N_{\mathcal {L}_0}(P)$ , and thus, by Lemma 3.10(a), $\mathrm {Aut}_{\mathcal {F}}(P)=\mathrm {Aut}_{\mathcal {F}_S(\mathcal {L}_0)}(P)$ . Similarly, $\mathrm {Aut}_{\mathcal {E}}(Q)=\mathrm {Aut}_{\mathcal {F}_T(\mathcal {N}_0)}(Q)$ for all $Q\in \Gamma $ . Hence, it follows from Alperin’s Fusion Theorem [Reference Aschbacher, Kessar and Oliver2, Theorem I.3.6] (combined with Lemma 2.6(b)) that $\mathcal {F}=\mathcal {F}_S(\mathcal {L}_0)$ and $\mathcal {E}=\mathcal {F}_T(\mathcal {N}_0)$ . This proves (a).

(b,c) We argue first that

(5.2) $$ \begin{align} N_{\mathcal{N}}(P)/O_{p^{\prime}}(N_{\mathcal{N}}(P)) \ \mbox{ is of characteristic} \ p \ \mbox{for all} \ P\in \Gamma_0. \end{align} $$

By Lemma 3.8(b), we may reduce to the case that P is fully $\mathcal {E}$ -normalised. So let $P\in \Gamma _0$ be fully $\mathcal {E}$ -normalised. As $\Gamma _0\subseteq \mathcal {E}^q$ , [Reference Henke16, Proposition 1(c)] gives that $C_{\mathcal {N}}(P)/O_{p^{\prime }}(C_{\mathcal {N}}(P))$ is a p-group. In particular, $C_{\mathcal {N}}(P)/O_{p^{\prime }}(C_{\mathcal {N}}(P))$ is of characteristic p, and thus $C_{\mathcal {N}}(P)=C_{N_{\mathcal {N}}(P)}(P)$ is ‘almost of characteristic p’ in the sense of [Reference Henke16, Definition 2.6]. It follows therefore from [Reference Henke16, Lemma 2.9] (applied with $N_{\mathcal {N}}(P)$ in place of G) that $N_{\mathcal {N}}(P)$ is ‘almost of characteristic p’, which means $N_{\mathcal {N}}(P)/O_{p^{\prime }}(N_{\mathcal {N}}(P))$ is of characteristic p. This proves equation (5.2).

As $N_{\mathcal {N}_0}(P)=N_{\mathcal {N}}(P)$ for all $P\in \Gamma _0$ , it follows from equation (5.2) and from [Reference Henke16, Proposition 6.4] applied with $(\mathcal {N}_0,\Gamma _0,T)$ in place of $(\mathcal {L},\Delta ,S)$ that $\Theta $ is a partial normal subgroup of $\mathcal {N}_0$ , that $T\cap \Theta =1$ , that the canonical projection $\mathcal {N}_0\rightarrow \mathcal {N}_0/\Theta $ is injective on T and that $(\mathcal {N}_0/\Theta ,\Gamma _0,T)$ is a linking locality over $\mathcal {E}$ if we identify T with its image in $\mathcal {N}_0/\Theta $ . The elements of $N_{\mathcal {L}}(T)=N_{\mathcal {L}_0}(T)$ induce $\mathcal {F}$ -automorphisms of T and act thus by assumption on $\Gamma _0$ via conjugation. Hence, it follows from Lemma 3.8(b) that $\Theta $ is invariant under conjugation by elements of $N_{\mathcal {L}_0}(T)$ . Now [Reference Chermak8, Corollary 3.13] gives that $\Theta $ is a partial normal subgroups of $\mathcal {L}_0$ . Moreover, $S\cap \Theta =S\cap \mathcal {N}\cap \Theta =T\cap \Theta =1$ . This means the kernel of the natural projection $\rho \colon \mathcal {L}_0\rightarrow \overline {\mathcal {L}_0}:=\mathcal {L}_0/\Theta $ intersects trivially with S and restricts thus to isomorphisms $S\rightarrow \overline {S}$ and $T\rightarrow \overline {T}$ . Hence, (b) holds.

We now use the notation introduced in (b) and (c). By [Reference Chermak8, Theorem 4.3], $(\overline {\mathcal {L}}_0,\Delta _0,S_0)$ is a locality. It is a special case of [Reference Henke17, Lemma 2.21(b)] that $\mathcal {F}_S(\overline {\mathcal {L}}_0)=\mathcal {F}_S(\mathcal {L}_0)=\mathcal {F}$ . Using [Reference Chermak8, Lemma 3.15], one can observe that $\rho |_{\mathcal {N}_0}$ coincides with the canonical projection $\mathcal {N}_0\rightarrow \mathcal {N}_0/\Theta $ . Thus, $(\overline {\mathcal {N}}_0,\Gamma _0,T)=(\mathcal {N}_0/\Theta ,\Gamma _0,T)$ is a linking locality over $\mathcal {E}$ . Now (c) follows from Lemma 5.7.

Proposition 5.9. Let $\mathcal {F}$ be a fusion system over S with a subsystem $\mathcal {E}$ over T. Then the following conditions are equivalent:

  1. (i) There exists a locality $(\mathcal {L},\Delta ,S)$ over $\mathcal {F}$ with a cr-complete kernel $(\mathcal {N},\Gamma ,T)$ over $\mathcal {E}$ .

  2. (ii) There exists a locality $(\mathcal {L},\Delta ,S)$ over $\mathcal {F}$ with a kernel $(\mathcal {N},\Gamma ,T)$ , which is a linking locality over $\mathcal {E}$ .

If either of these two conditions holds, then $\mathcal {E}$ is a normal subsystem of $\mathcal {F}$ .

Proof. Clearly, (ii) implies (i) as every linking locality is cr-complete. Assume now that (i) holds. Notice that $\mathcal {E}^{cr}\subseteq \Gamma _0:=\Gamma \cap \mathcal {E}^c\subseteq \Gamma \cap \mathcal {E}^q$ . Moreover, $\Gamma _0$ is $\mathcal {E}$ -closed and $\mathrm {Aut}_{\mathcal {F}}(T)$ invariant as the same holds for $\Gamma $ and $\mathcal {E}^c$ (compare Lemma 4.5). Hence, (ii) follows from Lemma 5.8(c).

Assume now that (ii) holds. If $(\mathcal {N},\Gamma ,T)$ is a kernel of a locality $(\mathcal {L},\Delta ,S)$ over $\mathcal {F}$ such that $(\mathcal {N},\Gamma ,T)$ is a linking locality over $\mathcal {E}$ , then $(\mathcal {N},\Gamma ,T)$ is in particular cr-complete. Thus, $(\mathcal {L},\Delta ,S)$ is cr-complete, by Proposition 5.5. Hence, $\mathcal {E}$ and $\mathcal {F}$ are saturated, by Proposition 3.18(c). So $\mathcal {E}\unlhd \mathcal {F}$ , by Lemma 5.6(b). This proves the assertion.

Proof of Theorem B.

The statement follows from Proposition 5.5 and Proposition 5.9.

Proposition 5.10. Let $(\mathcal {L},\Delta ,S)$ be a locality with a kernel $(\mathcal {N},\Gamma ,T)$ . Then the following conditions are equivalent:

  1. (i) $(\mathcal {L},\Delta ,S)$ is of objective characteristic p.

  2. (ii) $N_{\mathcal {L}}(T)$ is of characteristic p, and $(\mathcal {N},\Gamma ,T)$ is of objective characteristic p.

  3. (iii) $ C_{\mathcal {L}}(T)$ is of characteristic p, and $(\mathcal {N},\Gamma ,T)$ is of objective characteristic p.

  4. (iv) $N_{\mathcal {L}}(P)$ is of characteristic p for every $P\in \Gamma $ .

Proof. By Lemma 5.2(a), $(\mathcal {N},\Gamma ,T)$ is a locality. Notice that $T=S\cap \mathcal {N}\in \Gamma \subseteq \Delta $ , and thus $N_{\mathcal {L}}(T)$ is a finite group with $C_{N_{\mathcal {L}}(T)}(T)=C_{\mathcal {L}}(T)$ . So properties (ii) and (iii) are equivalent by [Reference Henke16, Lemma 2.9]. Hence, it is sufficient to show that properties (i),(ii) and (iv) are equivalent.

If (i) holds, then $N_{\mathcal {L}}(P)$ is of characteristic p for every $P\in \Gamma $ . In particular, $N_{\mathcal {L}}(T)$ is of characteristic p. Moreover, by Lemma 4.2(b), $N_{\mathcal {N}}(P)\unlhd N_{\mathcal {L}}(P)$ is of characteristic p for every $P\in \Gamma $ , showing that $(\mathcal {N},\Gamma ,T)$ is of objective characteristic p. So (i) implies (ii).

To show that (ii) implies (iv), assume now that (ii) holds. Suppose furthermore that (iv) is false: that is, there exists $P\in \Gamma $ such that $G:=N_{\mathcal {L}}(P)$ is not of characteristic p. Choose such P of maximal order. By Lemma 3.8(b) and [Reference Chermak8, Lemma 2.9], we may replace P by a suitable $\mathcal {L}$ -conjugate of P and assume that $N_S(P)$ is a Sylow p-subgroup of G. As (ii) holds, we have $P\neq T$ and thus $P<N_T(P)$ . Hence, by the maximality of $|P|$ , the group $N_{\mathcal {L}}(N_T(P))$ has characteristic p. Hence, by Lemma 4.2(a), $N_G(N_T(P))=N_{N_{\mathcal {L}}(N_T(P))}(P)$ is of characteristic p. As $(\mathcal {N},\Gamma ,T)$ is of objective characteristic p, we also know that $N:=N_{\mathcal {N}}(P)\unlhd G$ is of characteristic p. Notice that $N_T(P)=N_S(P)\cap N$ is a Sylow p-subgroup of N, as $N_S(P)$ is a Sylow p-subgroup of G. So it follows from Lemma 4.3 that G is of characteristic p, contradicting our assumption. Hence, (ii) implies (iv).

Assume now that (iv) holds. If $P\in \Delta $ is arbitrary, then $P\cap \mathcal {N}\in \Gamma $ , so $N_{\mathcal {L}}(P\cap \mathcal {N})$ is of characteristic p. Observe that $N_{\mathcal {L}}(P)\subseteq N_{\mathcal {L}}(P\cap \mathcal {N})$ , and thus $N_{\mathcal {L}}(P)=N_{N_{\mathcal {L}}(P\cap \mathcal {N})}(P)$ is of characteristic p by Lemma 4.2(a). Hence, (iv) implies (i).

Proof of Theorem C.

The statement follows from Proposition 5.5 and Proposition 5.10.

Proof of Theorem D.

Set $\mathcal {F}:=\mathcal {F}_S(\mathcal {L})$ and $\mathcal {E}:=\mathcal {F}_T(\mathcal {N})$ . By [Reference Chermak and Henke10, Lemma 3.28(c)], $(\mathcal {L},\Delta ,S)$ is a linking locality if and only if $(\mathcal {L},\tilde {\Delta },S)$ is a linking locality, and similarly, $(\mathcal {N},\Gamma ,T)$ is a linking locality if and only if $(\mathcal {N},\tilde {\Gamma },T)$ is a linking locality.Footnote 1 Thus, supposing from now on that $(\mathcal {L},\Delta ,S)$ and $(\mathcal {N},\Gamma ,T)$ are linking localities, it is by Theorem C enough to show that $\tilde {\Delta }=\delta (\mathcal {F})$ if and only if $\tilde {\Gamma }=\delta (\mathcal {E})$ and that $E(\mathcal {L})=E(\mathcal {N})$ .

By Proposition 5.9, $\mathcal {E}\unlhd \mathcal {F}$ . Moreover, our assumption yields that $N_{\mathcal {L}}(T)$ is of characteristic p. Hence, by Lemma 3.10(b), $N_{\mathcal {F}}(T)=\mathcal {F}_S(N_{\mathcal {L}}(T))$ is constrained. It follows then from [Reference Chermak and Henke10, Corollary 7.18, Lemma 7.19] that $E(C_{\mathcal {F}}(\mathcal {E}))=E(N_{\mathcal {F}}(T))$ is the trivial fusion system. By [Reference Chermak and Henke10, Lemma 7.13(c)], this implies $E(\mathcal {F})=E(\mathcal {E})$ .

By [Reference Henke18, Lemma 11.13], $E(\mathcal {N})\unlhd \mathcal {L}$ . Moreover, by [Reference Chermak and Henke10, Theorem E(d)], $\mathcal {F}_{S\cap E(\mathcal {N})}(E(\mathcal {N}))=E(\mathcal {E})=E(\mathcal {F})$ , and $E(\mathcal {L})$ is the unique partial normal subgroup of $\mathcal {L}$ with $\mathcal {F}_{S\cap E(\mathcal {L})}(E(\mathcal {L}))=E(\mathcal {F})$ . Hence, $E(\mathcal {N})=E(\mathcal {L})$ .

By Lemma 3.12 and [Reference Henke16, Proposition 5], $O_p(\mathcal {F})=O_p(\mathcal {L})$ and $O_p(\mathcal {E})=O_p(\mathcal {N})$ . Moreover, $O_p(\mathcal {L})\cap T=O_p(\mathcal {L})\cap \mathcal {N}=O_p(\mathcal {N})$ by Lemma 5.2(c).

Assume first that $\tilde {\Gamma }=\delta (\mathcal {E})$ . Let $P\leq S$ with $O_p(\mathcal {L})=O_p(\mathcal {F})\leq P$ . Then $O_p(\mathcal {N})\leq P\cap T$ and thus $P\cap T\in \Gamma $ if and only if $P\cap T\in \tilde {\Gamma }$ . Hence,

$$ \begin{align*} P\in\delta(\mathcal{F}) &\Longleftrightarrow P\cap T\in\delta(\mathcal{E})\ \mbox{(by Lemma~4.9(a))}\\ &\Longleftrightarrow P\cap T\in\tilde{\Gamma} (\mbox{ as } \ \tilde{\Gamma}=\delta(\mathcal{E}))\\ &\Longleftrightarrow P\cap T\in\Gamma \mbox{ (as noted above)}\\ &\Longleftrightarrow P\in\Delta \ (\mbox{since} \ (\mathcal{N},\Gamma,T) \mbox{is a kernel}). \end{align*} $$

Since the above equivalences hold for every $P\leq S$ with $O_p(\mathcal {L})\leq P$ , it follows now from [Reference Henke18, Lemma 10.6] that $\delta (\mathcal {F})=\{P\leq S\colon PO_p(\mathcal {L})\in \delta (\mathcal {F})\}=\{P\leq S\colon PO_p(\mathcal {L})\in \Delta \}=\tilde {\Delta }$ . This proves one direction.

Assume now the other way around that $\delta (\mathcal {F})=\tilde {\Delta }$ . Let $Q\leq T$ with $O_p(\mathcal {N})\leq Q$ . Then $QO_p(\mathcal {L})\cap T=Q(O_p(\mathcal {L})\cap T)=QO_p(\mathcal {N})=Q$ . Hence,

$$ \begin{align*} Q\in\delta(\mathcal{E}) &\Longleftrightarrow QO_p(\mathcal{L})\cap T\in\delta(\mathcal{E}) \ (\mbox{as} \ Q=QO_p(\mathcal{L})\cap T)\\ &\Longleftrightarrow QO_p(\mathcal{L})\in\delta(\mathcal{F})\ (\mbox{by Lemma~4.9(a) and since} \ O_p(\mathcal{L})=O_p(\mathcal{F})) \\ &\Longleftrightarrow QO_p(\mathcal{L})\in\tilde{\Delta} \ (\mbox{ as} \ \delta(\mathcal{F})=\tilde{\Delta})\\ &\Longleftrightarrow QO_p(\mathcal{L})\in\Delta \ (\mbox{by definition of} \tilde{\Delta})\\ &\Longleftrightarrow QO_p(\mathcal{L})\cap T\in\Gamma \ (\mbox{as} \ (\mathcal{N},\Gamma,T) \ \mbox{is a kernel})\\ &\Longleftrightarrow Q\in \Gamma \ (\mbox{as} \ Q=QO_p(\mathcal{L})\cap T). \end{align*} $$

The above equivalences hold for all $Q\leq T$ with $O_p(\mathcal {N})\leq Q$ . Hence, it follows from [Reference Henke18, Lemma 10.6] that $\delta (\mathcal {E})=\{Q\leq T\colon QO_p(\mathcal {N})\in \delta (\mathcal {E})\}=\{Q\leq T\colon QO_p(\mathcal {N})\in \Gamma \}=\tilde {\Gamma }$ . This proves the assertion.

Lemma 5.11. Let $(\mathcal {L},\Delta ,S)$ be a locality with kernel $(\mathcal {N},\Gamma ,T)$ such that $(\mathcal {N},\Gamma ,T)$ is a regular locality. Set $\mathcal {E}:=\mathcal {F}_T(\mathcal {N})$ , $T_0:=E(\mathcal {N})\cap S=E(\mathcal {E})\cap S$ and $\tilde {\Delta }:=\{P\leq S\colon PO_p(\mathcal {L})\in \Delta \}$ . Then the following are equivalent:

  1. (i) $(\mathcal {L},\Delta ,S)$ is a linking locality.

  2. (ii) $(\mathcal {L},\tilde {\Delta },S)$ is a regular locality.

  3. (iii) $N_{\mathcal {L}}(T)$ is a group of characteristic p.

  4. (iv) $N_{\mathcal {L}}(T_0)$ is a group of characteristic p.

Proof. Set $\mathcal {F}:=\mathcal {F}_S(\mathcal {L})$ . By [Reference Chermak and Henke10, Theorem E(d)], $E(\mathcal {E})=\mathcal {F}_{E(\mathcal {N})\cap S}(E(\mathcal {N}))$ , and in particular $T_0:=E(\mathcal {N})\cap S=E(\mathcal {E})\cap S$ . As $(\mathcal {N},\Gamma ,T)$ is regular (and in particular a linking locality), it follows from Theorem C that (i) and (iii) are equivalent. Moreover, by [Reference Henke18, Lemma 10.6], $\Gamma =\delta (\mathcal {E})=\{Q\leq T\colon QO_p(\mathcal {N})\in \Gamma \}$ . Therefore, properties (ii) and (iii) are equivalent by Theorem D. Hence, it remains to prove that (iii) and (iv) are equivalent.

As $T_0\in E(\mathcal {E})^s\subseteq \delta (\mathcal {E})=\Gamma \subseteq \Delta $ by [Reference Chermak and Henke10, Lemma 7.22], it follows that $G:=N_{\mathcal {L}}(T_0)$ is a group with $N:=N_{\mathcal {N}}(T_0)\unlhd G$ . Indeed, since $(\mathcal {N},\Gamma ,T)$ is a regular locality, N is of characteristic p. Hence, by Lemma 4.3, G is of characteristic p if and only if $N_G(T)$ is of characteristic p. By [Reference Henke18, Lemma 11.12], every automorphism of $\mathcal {N}$ leaves $E(\mathcal {N})$ invariant, so $N_{\mathcal {L}}(T)$ acts by Lemma 5.2(b) on $E(\mathcal {N})$ via conjugation. In particular, $N_{\mathcal {L}}(T)\subseteq G$ , and thus $N_{\mathcal {L}}(T)=N_G(T)$ . This shows the equivalence of (iii) and (iv) as required.

6 Products

In this section, we prove Theorem E and Corollary F. Indeed, we state and prove here some more detailed results. The following theorem implies Theorem E. The reader might want to recall the notation introduced in Remark 4.10.

Theorem 6.1. Let $(\mathcal {L}, \Delta , S)$ be a regular locality, and let $\mathcal {N}\mathrel {\unlhd } \mathcal {L}$ be a partial normal subgroup of $\mathcal {L}$ . Set

$$\begin{align*}T^* := F^*(\mathcal{L}) \cap S, \; T_0:=E(\mathcal{N})\cap S,\; T:=S\cap\mathcal{N}\mbox{ and }\mathcal{E}:=\mathcal{F}_T(\mathcal{N}).\end{align*}$$

Let $H\leq N_{\mathcal {L}}(T^*)$ and set $\tilde {H}:=N_{\mathcal {N}}(T^*)H$ . Fix $S_0\in \mathrm {Syl}_p(\tilde {H})$ with $T\leq S_0$ . Set

$$\begin{align*}\Delta_0 := \{ P \leq S_0 \mid P \cap \mathcal{N} \in \delta(\mathcal{E})\}.\end{align*}$$

Then the following hold:

  1. (a) $\mathcal {N} H=H\mathcal {N}$ is a partial subgroup of $\mathcal {L}$ .

  2. (b) $N_{\mathcal {N}}(T_0)=N_{\mathcal {N}}(T^*)$ , and $\tilde {H}=N_{\mathcal {N} H}(T^*)=N_{\mathcal {N} H}(T_0)=N_{\mathcal {N}}(T_0)H$ .

  3. (c) $(\mathcal {N} H, \Delta _0, S_0)$ is a cr-complete locality with kernel $(\mathcal {N},\delta (\mathcal {E}),T)$ .

  4. (d) Set $\tilde {\Delta }_0:=\{P\leq S_0\colon PO_p(\mathcal {N} H)\in \Delta _0\}$ . Then the following are equivalent:

    1. $\tilde {H}$ is of characteristic p;

    2. $(\mathcal {N} H,\Delta _0,S_0)$ is a linking locality;

    3. $(\mathcal {N} H,\tilde {\Delta }_0,S_0)$ is a regular locality.

  5. (e) $\mathcal {F}_0:=\mathcal {F}_{S_0}(\mathcal {N} H)$ is saturated. Moreover, $\mathcal {E}$ and $E(\mathcal {E})$ are normal subsystems of $\mathcal {F}_0$ , $N_{\mathcal {F}_0}(T_0)=\mathcal {F}_{S_0}(\tilde {H})$ , and

    $$\begin{align*}\mathcal{F}_0=\langle(E(\mathcal{E})S_0)_{\mathcal{F}_0},\mathcal{F}_{S_0}(\tilde{H})\rangle.\end{align*}$$

Proof. Set $\mathcal {F}:= \mathcal {F}_S(\mathcal {L})$ . As usual, we write $\Pi \colon \mathbf {D}\rightarrow \mathcal {L}$ for the product on $\mathcal {L}$ .

(a) By definition of a regular locality, for every $P\leq S$ , we have $P\in \Delta $ if and only if $P\cap T^*\in \Delta $ . Thus, Lemma 3.8(e) yields that

(6.1) $$ \begin{align} \mbox{ for all } \ u\in\mathbf {W}(\mathcal{L}), \ \mbox{we have} \ u\in\mathbf {D} \ \mbox{if and only if} \ S_u\cap T^*\in\Delta. \end{align} $$

In particular, $S_f \cap T^* \in \Delta $ for all $f\in \mathcal {L}$ . If $n \in \mathcal {N}$ and $g\in \mathcal {N}_{\mathcal {L}}(T^*)$ , then we can conclude that $(g^{-1},g,n)\in \mathbf {D}$ via $S_n\cap T^*$ . Hence, by the axioms of a partial group, $gn$ is defined and $g^{-1}(gn)=\Pi (g^{-1},g,n)=n$ . Note also that $S_{gn}\cap T^*\leq S_{(g,g^{-1},gn)}$ . Using Lemma 3.8(f), we conclude that

$$\begin{align*}S_{(g,n)} \cap T^* \leq S_{gn} \cap T^* \leq S_{(g,g^{-1},gn)} \cap T^* \leq S_{(g,g^{-1}(gn))} \cap T^* \leq S_{(g,n)} \cap T^*\end{align*}$$

and so

(6.2) $$ \begin{align} S_{(g,n)} \cap T^* = S_{gn} \cap T^*\mbox{ for all } \ g\in N_{\mathcal{L}}(T^*) \ \mbox{and all} \ n\in\mathcal{N}. \end{align} $$

Let $n\in \mathcal {N}$ and $g \in H$ . We argue first that $\mathcal {N} H=H\mathcal {N}$ . Observe that $u_1:=(g,g^{-1},n,g)\in \mathbf {D}$ via $S_n\cap T^*$ and $u_2:=(g,n,g^{-1},g)\in \mathbf {D}$ via $(S_n\cap T^*)^{g^{-1}}$ . Hence, $ng=\Pi (u_1)=g(n^g)\in H\mathcal {N}$ and $gn=\Pi (u_2)=(n^{g^{-1}})g\in \mathcal {N} H$ . This shows that $\mathcal {N} H=H\mathcal {N}$ . In particular, if $f=ng\in \mathcal {N} H$ , then [Reference Chermak8, Lemma 1.4(f)] yields $(g^{-1}, n^{-1}) \in \mathbf {D}$ and $f^{-1} = g^{-1}n^{-1}\in H\mathcal {N}=\mathcal {N} H$ .

It remains thus to show that $\mathcal {N} H=H\mathcal {N}$ is closed under the partial product $\Pi \colon \mathbf {D}\rightarrow \mathcal {L}$ on $\mathcal {L}$ . Let $w = (f_1,\dots ,f_k) \in \mathbf {D} \cap \mathbf {W}(H\mathcal {N})$ . Then for every $1 \leq i \leq k $ , we have $f_i = g_in_i$ for some $g_i \in H$ and $n_i \in \mathcal {N}$ . Set $u = (g_1,n_1, \dots , g_k,n_k)$ . Then by (6.2), we get $S_u\cap T^* = S_w\cap T^*$ . Now equation (6.1) yields first $S_u\cap T^* = S_w\cap T^*\in \Delta $ and then $u \in \mathbf {D}$ . By [Reference Chermak8, Lemma 3.4], $\Pi (u)=\Pi (g_1,\dots ,g_k,n)$ for some $n\in \mathcal {N}$ . Hence, by the axioms of a partial group,

$$\begin{align*}\Pi(w) = \Pi(u) =\Pi(g_1,\dots,g_k,n)= \Pi(g_1,\dots,g_k)n \in H\mathcal{N}=\mathcal{N} H.\end{align*}$$

Thus $\mathcal {N} H$ is closed under the partial product and thus a partial subgroup. Hence, (a) holds.

(b) Lemma 3.22(a) gives $N_{\mathcal {L}}(T_0)\leq N_{\mathcal {L}}(T^*)$ and $N_{\mathcal {N}}(T_0)=N_{\mathcal {N}}(T^*)$ . It follows from the Dedekind Lemma [Reference Chermak8, Lemma 1.10] that $N_{\mathcal {N} H}(T^*)=\mathcal {N} H\cap N_{\mathcal {N} H}(T^*)=N_{\mathcal {N}}(T^*)H=\tilde {H}$ . As $H\leq N_{\mathcal {L}}(T^*)\subseteq N_{\mathcal {L}}(T_0)$ and $N_{\mathcal {L}}(T_0)$ is by [Reference Chermak8, Lemma 2.12(a)] a partial subgroup of $\mathcal {L}$ , the Dedekind Lemma yields similarly that $N_{\mathcal {N} H}(T_0)=\mathcal {N} H\cap N_{\mathcal {N} H}(T_0)=N_{\mathcal {N}}(T_0)H$ . This implies (b).

(c,d) Set $\Gamma :=\delta (\mathcal {E})$ and $R:=E(C_{\mathcal {L}}(\mathcal {N}))\cap S$ . We show first that

(6.3) $$ \begin{align} \mathbf {D} \cap \mathbf {W}(\mathcal{N} H) =\mathbf {D}_{\Gamma}\cap \mathbf {W}(\mathcal{N} H). \end{align} $$

As $R\subseteq C_{\mathcal {L}}(\mathcal {N})$ , [Reference Henke18, Lemma 3.5] yields $\mathcal {N}\subseteq C_{\mathcal {L}}(R)\subseteq N_{\mathcal {L}}(R)$ . Moreover, by Lemma 3.22(a), $H\subseteq N_{\mathcal {L}}(T^*)\subseteq N_{\mathcal {L}}(R)$ . As $N_{\mathcal {L}}(R)$ is by [Reference Chermak8, Lemma 2.12(a)] a partial subgroup of $\mathcal {L}$ , it follows that $\mathcal {N} H\subseteq N_{\mathcal {L}}(R)$ . Moreover, we use that $O_p(\mathcal {L})=O_p(\mathcal {F})$ , by [Reference Henke16, Proposition 5] and Lemma 3.12. So fixing $w=(f_1,\dots ,f_n)\in \mathbf {W}(\mathcal {N} H)$ , we have $RO_p(\mathcal {F})=RO_p(\mathcal {L})\leq S_w$ . Hence, Lemma 4.9(b) gives

$$\begin{align*}S_w\in \delta(\mathcal{F})=\Delta\Longleftrightarrow S_w\cap T\in\Gamma. \end{align*}$$

Notice that $S_w\in \mathbf {D}=\mathbf {D}_{\Delta }$ if and only if $S_w\in \Delta $ . By [Reference Henke18, Lemma 10.4], $\Gamma $ is overgroup-closed in T. Hence, $w\in \mathbf {D}_{\Gamma }$ implies $S_w\cap T\in \Gamma $ . Moreover, by Lemma 3.22(d), $\Gamma $ is closed under passing to $\mathcal {L}$ -conjugates in S. Therefore, $S_w\cap T\in \Gamma $ implies $w\in \mathbf {D}_{\Gamma }$ . So $S_w\cap T\in \Gamma $ if and only if $w\in \mathbf {D}_{\Gamma }$ . This proves equation (6.3). We show next

(6.4) $$ \begin{align} S_0\mbox{ is maximal with respect to inclusion among the} \ p\mbox{-subgroups of} \ \mathcal{N} H. \end{align} $$

Let $S_1$ be a p-subgroup of $\mathcal {N} H$ such that $S_0\leq S_1$ . Notice that $T\leq S_0\leq S_1$ , so $T_0=E(\mathcal {N})\cap S\leq E(\mathcal {N})\cap S_1$ . By [Reference Henke18, Lemma 11.13], $E(\mathcal {N})\unlhd \mathcal {L}$ . In particular, [Reference Chermak8, Lemma 3.1(c)] yields that $T_0$ is a maximal p-subgroup of $E(\mathcal {N})$ . Hence, $T_0=E(\mathcal {N})\cap S_1\unlhd S_1$ . Now (b) yields $S_1\leq N_{\mathcal {N} H}(T_0)=\tilde {H}$ . As $S_0\in \mathrm {Syl}_p(\tilde {H})$ , it follows that $S_1=S_0$ . This proves equation (6.4).

Recall that by Lemma 3.22(d), $\Gamma $ is closed under passing to $\mathcal {L}$ -conjugates in S. As T is by [Reference Chermak8, Lemma 3.1(c)] a maximal p-subgroup of $\mathcal {N}$ , we have $T=S_0\cap \mathcal {N}$ . Since the elements of $\Gamma $ are contained in T and $\mathcal {N}\unlhd \mathcal {N} H$ , it follows that $\Gamma $ is closed under passing to $\mathcal {N} H$ -conjugates in $S_0$ . So by equations (6.3) and (6.4), the hypothesis of Lemma 5.4 is fulfilled with $(\mathcal {N} H,S_0)$ in place of $(\mathcal {L},S)$ . Hence, it follows from this lemma that $(\mathcal {N} H,\Delta _0,S_0)$ is a locality with kernel $(\mathcal {N},\Gamma ,T)$ . By [Reference Henke18, Theorem 10.16(a)], $(\mathcal {N},\Gamma ,T)$ is a regular locality, and in particular it is cr-complete. Hence, $(\mathcal {N} H,\Delta _0,S_0)$ is cr-complete by Theorem B: that is, (c) holds. Part (d) follows from (b) and Lemma 5.11.

(e) By [Reference Henke18, Corollary 11.10], we have $T_0\in \delta (\mathcal {E})=\Gamma \subseteq \Delta _0$ . Hence, parts (b) and (c) together with Lemma 3.10(b) yield that $N_{\mathcal {F}_0}(T_0):=\mathcal {F}_{S_0}(N_{\mathcal {N} H}(T_0))=\mathcal {F}_{S_0}(\tilde {H})$ . Moreover, by Theorem B and Proposition 3.18(c), part (c) implies that $\mathcal {F}_0$ is saturated and $\mathcal {E}\unlhd \mathcal {F}_0$ . Hence, $E(\mathcal {E})\unlhd \mathcal {F}_0$ by [Reference Chermak and Henke10, Lemma 7.13(a)]. By Lemma 3.22(b), we have $T_0=E(\mathcal {E})\cap S$ . Hence, (e) follows from Lemma 4.11.

If the subgroup H in Theorem 6.1 has the property that $S\cap H\in \mathrm {Syl}_p(H)$ , then the following lemma says that we can choose the p-subgroup $S_0$ as a subgroup of S.

Lemma 6.2. Let $(\mathcal {L},\Delta ,S)$ be a regular locality, and set $T^*:=F^*(\mathcal {L})\cap S$ . Let $\mathcal {N}\unlhd \mathcal {L}$ , $T:=S\cap \mathcal {N}$ and $H\leq N_{\mathcal {L}}(T^*)$ such that $S\cap H\in \mathrm {Syl}_p(H)$ . Then

$$\begin{align*}S_0:=T(S\cap H)=S\cap (\mathcal{N} H)\in\mathrm{Syl}_p(N_{\mathcal{N}}(T^*)H). \end{align*}$$

Proof. Note that $N_{\mathcal {L}}(T^*)$ is a group, S is a Sylow p-subgroup of $N_{\mathcal {L}}(T^*)$ , $N_{\mathcal {N}}(T^*) \mathrel {\unlhd } N_{\mathcal {L}}(T^*)$ , and $H \leq N_{\mathcal {L}}(T^*)$ by assumption. As $S\cap H\in \mathrm {Syl}_p(H)$ , Lemma 4.1 yields that

$$\begin{align*}S_0:=(S \cap N_{\mathcal{N}}(T^*))(S \cap H)=T(S\cap H)\end{align*}$$

is a Sylow p-subgroup of $N_{\mathcal {N}}(T^*)H$ . In particular, it is a consequence of Theorem 6.1(a),(c) that $S_0$ is a maximal p-subgroup of the partial group $\mathcal {N} H$ and thus equals $S\cap (\mathcal {N} H)$ .

Assuming the hypothesis of Corollary 6.3 below, we have $\mathcal {E}\unlhd \mathcal {F}$ by [Reference Chermak and Henke10, Theorem A]. Hence, it is a consequence of [Reference Chermak and Henke10, Lemma 7.13(a)] that $E(\mathcal {F})\unlhd \mathcal {F}$ , and thus $(E(\mathcal {E})S_0)_{\mathcal {F}}$ is well-defined. This is implicitly used in the statement of part (b).

Corollary 6.3. Let $(\mathcal {L},\Delta ,S)$ be a regular locality, $T^*:=F^*(\mathcal {L})\cap S$ , $\mathcal {N}\unlhd \mathcal {L}$ , and $H\leq N_{\mathcal {L}}(T^*)$ with $S\cap H\in \mathrm {Syl}_p(H)$ . Set $T:=S\cap T$ , $\mathcal {E}:=\mathcal {F}_T(\mathcal {N})$ , $S_0:=T(S\cap H)$ , $\tilde {H}:=N_{\mathcal {N}}(T^*)H$ and

$$\begin{align*}\mathcal{E} H:=\mathcal{F}_{S_0}(\mathcal{N} H). \end{align*}$$

Then the following hold:

  1. (a) $\mathcal {E} H$ is a saturated subsystem of $\mathcal {F}$ , $\mathcal {E}\unlhd \mathcal {E} H$ and $E(\mathcal {E})\unlhd \mathcal {E} H$ .

  2. (b) $S_0\in \mathrm {Syl}_p(\tilde {H})$ and $\mathcal {E} H=\langle E(\mathcal {E})S_0,\mathcal {F}_{S_0}(\tilde {H})\rangle $ , where $E(\mathcal {E})S_0:=(E(\mathcal {E})S_0)_{\mathcal {F}}=(E(\mathcal {E})S_0)_{\mathcal {E} H}$ .

  3. (c) If $\mathcal {D}$ is a saturated subsystem of $\mathcal {F}$ such that $E(\mathcal {E})\unlhd \mathcal {D}$ and $\mathcal {F}_{S_0}(\tilde {H})\subseteq \mathcal {D}$ , then $\mathcal {E} H\subseteq \mathcal {D}$ .

Proof. By Lemma 6.2, $S_0\in \mathrm {Syl}_p(\tilde {H})$ . Therefore, setting $\Delta _0:=\{P\leq S_0\colon P\cap T\in \delta (\mathcal {E})\}$ , the hypothesis and thus the conclusion of Theorem 6.1 hold. Theorem 6.1(e) implies now that (a) holds and $\mathcal {E} H=\langle (E(\mathcal {E})S_0)_{\mathcal {E} H},\mathcal {F}_{S_0}(\tilde {H})\rangle $ . By [Reference Chermak and Henke10, Remark 2.27], $(E(\mathcal {E})S_0)_{\mathcal {E} H}=(E(\mathcal {E})S_0)_{\mathcal {F}}$ . Hence, (b) holds as well.

Now let $\mathcal {D}$ be as in (c), and suppose $\mathcal {D}$ is a subsystem over $R\leq S$ . As $\mathcal {F}_{S_0}(\tilde {H})\subseteq \mathcal {D}$ , we have in particular that $S_0\leq R$ . Using [Reference Chermak and Henke10, Remark 2.27] again, we observe that $(E(\mathcal {E})S_0)_{\mathcal {F}}=(E(\mathcal {E})S_0)_{\mathcal {D}}\subseteq \mathcal {D}$ . As $\mathcal {F}_{S_0}(\tilde {H})\subseteq \mathcal {D}$ by assumption, it follows now from (b) that $\mathcal {E} H\subseteq \mathcal {D}$ , so (c) holds.

Proof of Corollary F.

By [Reference Henke18, Lemma 10.4], there exists a regular locality $(\mathcal {L},\Delta ,S)$ over $\mathcal {F}$ . Moreover, by [Reference Chermak and Henke10, Theorem A], there exists $\mathcal {N}\unlhd \mathcal {L}$ with $T=S\cap \mathcal {N}$ and $\mathcal {E}=\mathcal {F}_T(\mathcal {N})$ . By [Reference Chermak and Henke10, Theorem E(d)], we have $T^*:=F^*(\mathcal {F})\cap S=F^*(\mathcal {L})\cap S$ . Moreover, Lemma 3.22(b) gives $T_0:=E(\mathcal {E})\cap S=E(\mathcal {N})\cap S$ . In particular, by [Reference Henke18, Lemma 10.4, Remark 10.12], we have $T^*\in \delta (\mathcal {F})=\Delta \subseteq \mathcal {F}^s$ . So $N_{\mathcal {F}}(T^*)$ is constrained, and $N_{\mathcal {L}}(T^*)$ is by Lemma 3.10(b) a model for $N_{\mathcal {F}}(T^*)$ . Using Lemma 3.22(a), one observes that $N_{\mathcal {N}}(T_0)=N_{\mathcal {N}}(T^*)\unlhd N_{\mathcal {L}}(T^*)$ . Moreover, Lemma 3.22(c) gives $N_{\mathcal {E}}(T_0)=\mathcal {F}_{T}(N_{\mathcal {N}}(T^*))$ . Hence, it follows from [Reference Aschbacher, Kessar and Oliver2, Proposition I.6.2] that $N_{\mathcal {E}}(T_0)\unlhd N_{\mathcal {F}}(T^*)$ .

Now let G be an arbitrary model for $N_{\mathcal {F}}(T^*)$ . As $N_{\mathcal {E}}(T_0)\unlhd N_{\mathcal {F}}(T^*)$ , it follows from [Reference Aschbacher, Kessar and Oliver2, Theorem II.7.5] that there exists a unique normal subgroup N of G with $S\cap N=T$ and $\mathcal {F}_T(N)=N_{\mathcal {E}}(T_0)$ . Now let $H\leq G$ with $S\cap H\in \mathrm {Syl}_p(H)$ . Set $S_0:=T(S\cap H)$ . By [Reference Aschbacher, Kessar and Oliver2, Theorem III.5.10], there exists an isomorphism $\alpha \colon G\rightarrow N_{\mathcal {L}}(T^*)$ with $\alpha |_S=\operatorname {id}_S$ . Notice that $N\alpha \unlhd N_{\mathcal {L}}(T^*)$ and $H\alpha \leq N_{\mathcal {L}}(T^*)$ with

$$\begin{align*}T=S\cap (N\alpha),\;\mathcal{F}_T(N\alpha)=\mathcal{F}_{T}(N)=N_{\mathcal{E}}(T^*), \end{align*}$$
$$\begin{align*}S\cap H=S\cap (H\alpha)\in\mathrm{Syl}_p(H\alpha)\mbox{ and }\mathcal{F}_{S_0}(N H)=\mathcal{F}_{S_0}((NH)\alpha). \end{align*}$$

By [Reference Aschbacher, Kessar and Oliver2, Theorem II.7.5], $N_{\mathcal {N}}(T^*)$ is the unique normal subgroup of $N_{\mathcal {L}}(T^*)$ realising $N_{\mathcal {E}}(T^*)$ . Hence, $N\alpha =N_{\mathcal {N}}(T^*)$ and $\mathcal {F}_{S_0}(NH)=\mathcal {F}_{S_0}(N_{\mathcal {N}}(T^*)(H\alpha ))$ . Therefore the assertion follows from Corollary 6.3 applied with $H\alpha $ in place of H.

A Normal pairs of transporter systems

It is not within the scope of this paper to construct extensions of localities, but Theorem A and our theorems on kernels provide some tools for examining the properties of existing extensions. The results in this appendix may help to compare our theorems on kernels to theorems on extensions of linking systems in the literature. Such extensions have been studied in various places, starting with [Reference Broto, Castellana, Grodal, Levi and Oliver6]. A more transparent algebraic framework is used in [Reference Oliver23, Reference Andersen, Oliver and Ventura1], where the definition of a normal pair of linking systems is crucial (see also [Reference Aschbacher, Kessar and Oliver2, Definition III.4.12]). This definition naturally generalises to a definition of a ‘normal pair of transporter systems’, which we state below. Transporter systems were defined by Oliver and Ventura [Reference Oliver and Ventura25], generalising the concept of a linking system. The goal of this appendix is to show that localities with kernels correspond to normal pairs of transporter systems (compare Proposition A.4 and Theorem A.7). We use here a correspondence between localities and transporter systems that was observed by Chermak [Reference Chermak7, Appendix A].

For any functor $\alpha \colon \mathcal {C}\rightarrow \mathcal {D}$ between categories and any objects $P,Q\in \mathcal {C}$ , we let

$$\begin{align*}\alpha_{P,Q}\colon \mathrm{Mor}_{\mathcal{C}}(P,Q)\rightarrow \mathrm{Mor}_{\mathcal{D}}(\alpha(P),\alpha(Q)) \end{align*}$$

denote the induced map between morphisms sets. Moreover, we set $\alpha _P:=\alpha _{P,P}$ .

The literature on linking systems and transporter systems is mostly written in ‘left-hand notation’. Therefore, in this appendix (unlike in the rest of the paper), we will write maps on the left-hand side of the argument and conjugate from the left. In particular, if G is a finite group and $\Delta $ is a set of subgroups of G, then $\mathcal {T}_{\Delta }(G)$ denotes the category whose object set is $\Delta $ and such that the morphism set $\mathrm {Mor}_{\mathcal {T}_{\Delta }(G)}(P,Q)$ between any two objects $P,Q\in \Delta $ is the set of all $g\in G$ with

$$\begin{align*}^g\!P:=P^{g^{-1}}=gPg^{-1}\leq Q.\end{align*}$$

A transporter system associated to a fusion system $\mathcal {F}$ over S is a category $\mathcal {T}$ whose object set is an $\mathcal {F}$ -closed collection of subgroups of S, together with functors

$$\begin{align*}\mathcal{T}_{\mathrm{Ob}(\mathcal{T})}(S)\xrightarrow{\;\;\;\delta\;\;\;}\mathcal{T}\xrightarrow{\;\;\;\pi\;\;\;}\mathcal{F}\end{align*}$$

subject to certain axioms. In particular, $\delta $ is the identity on objects and $\pi $ is the inclusion on objects, $\delta $ is injective on morphism sets, $\pi $ is surjective on morphism sets, and $\pi \circ \delta $ sends an element $g\in \mathrm {Mor}_{\mathcal {T}_{\mathrm {Ob}(\mathcal {T})}(S)}(P,Q)$ to the corresponding conjugation map $c_{g^{-1}}\colon P\rightarrow Q,x\mapsto {}^g\!x$ . The reader is referred to [Reference Oliver and Ventura25, Definition 3.1] for the precise definition. If $P,Q\in \mathrm {Ob}(\mathcal {T})$ and $\varphi \in \mathrm {Mor}_{\mathcal {T}}(P,Q)$ , we will usually write $\pi (\varphi )$ for $\pi _{P,Q}(\varphi )$ . The following definition is nonstandard.

Definition A.1. If $(\mathcal {T},\delta ,\pi )$ is a transporter system associated to a fusion system $\mathcal {F}$ , then we will say that $(\mathcal {T},\delta ,\pi )$ is a transporter system over $\mathcal {F}$ if $\mathcal {F}$ is $\mathrm {Ob}(\mathcal {T})$ -generated: that is,

$$\begin{align*}\mathcal{F}=\langle\pi(\varphi)\colon P,Q\in\mathrm{Ob}(\mathcal{T}),\;\varphi\in\mathrm{Mor}_{\mathcal{T}}(P,Q)\rangle.\end{align*}$$

If $(\mathcal {T},\delta ,\pi )$ is a linking system associated to a saturated fusion system $\mathcal {F}$ , then it follows from Alperin’s Fusion Theorem (compare [Reference Aschbacher, Kessar and Oliver2, Theorem I.3.6]) that $(\mathcal {T},\delta ,\pi )$ is a transporter system over $\mathcal {F}$ . However, in general, there can be transporter systems associated to a fusion system $\mathcal {F}$ that are not transporter systems over $\mathcal {F}$ (but just transporter systems over a subsystem of $\mathcal {F}$ ).

For a locality $(\mathcal {L},\Delta ,S)$ , there is a transporter system $\mathcal {T}_{\Delta }(\mathcal {L})$ over $\mathcal {F}_S(\mathcal {L})$ defined. The object set of the category $\mathcal {T}_{\Delta }(\mathcal {L})$ is $\Delta $ , and for $P,Q\in \Delta $ , the set $\mathrm {Mor}_{\mathcal {T}_{\Delta }(\mathcal {L})}(P,Q)$ consists of all $f\in \mathcal {L}$ with

$$\begin{align*}P\leq S_{f^{-1}}\mbox{ and }^f\!P:=P^{f^{-1}}\leq Q. \end{align*}$$

It turns out (see [Reference Chermak7, Proposition A.3]) that $(\mathcal {T}_{\Delta }(\mathcal {L}),\delta ,\pi )$ is a transporter system, where the functor $\delta \colon \mathcal {T}_{\Delta }(S)\rightarrow \mathcal {T}_{\Delta }(\mathcal {L})$ is the identity on objects and the inclusion on morphism sets, and the functor $\pi \colon \mathcal {T}_{\Delta }(\mathcal {L})\rightarrow \mathcal {F}_S(\mathcal {L})$ is the inclusion on objects and sends a morphism $f\in \mathrm {Mor}_{\mathcal {T}_{\Delta }(\mathcal {L})}(P,Q)$ to $c_f|_P\in \mathrm {Hom}_{\mathcal {F}_S(\mathcal {L})}(P,Q)$ .

To review some notation and basic results, let $(\mathcal {T},\delta ,\pi )$ be a transporter system, and fix $P,Q\in \mathrm {Ob}(\mathcal {T})$ . If $P\leq Q$ , then the morphism

$$\begin{align*}\iota_{P,Q}:=\delta_{P,Q}(1) \end{align*}$$

is regarded as an ‘inclusion map’. If $\varphi \in \mathrm {Mor}_{\mathcal {T}}(P,Q)$ and $P'\leq P$ , then set

$$\begin{align*}\varphi(P'):=\pi(\varphi)(P'). \end{align*}$$

By [Reference Oliver and Ventura25, Lemma A.6], a morphism $\varphi \in \mathrm {Mor}_{\mathcal {T}}(P,Q)$ is an isomorphism in the categorical sense if and only if $\pi (\varphi )$ is an isomorphism, which is the case if and only if $\varphi (P)=Q$ . It is shown in [Reference Oliver and Ventura25, Lemma 3.2(c)] that for all $P',Q'\in \Delta $ with $P'\leq P$ , $Q'\leq Q$ and all $\varphi \in \mathrm {Mor}_{\mathcal {T}}(P,Q)$ with $\varphi (P')\leq Q'$ , there exists a unique morphism $\varphi |_{P',Q'}\in \mathrm {Mor}_{\mathcal {T}}(P',Q')$ with

$$\begin{align*}\varphi\circ \iota_{P',P}=\iota_{Q',Q}\circ \varphi|_{P',Q'}. \end{align*}$$

The morphism $\varphi |_{P',Q'}$ is called the restriction of $\varphi $ to a morphism from $P'$ to $Q'$ .

Lemma A.2. Let $(\mathcal {T},\delta ,\pi )$ be a transporter system associated to some fusion system over a p-group S. Then the following hold:

  1. (a) Let $P"\leq P'\leq P$ and $Q"\leq Q'\leq Q$ be objects in $\mathcal {T}$ . Let $\varphi \in \mathrm {Mor}_{\mathcal {T}}(P,Q)$ with $\varphi (P')\leq Q'$ and $\varphi (P")\leq Q"$ . Then

    $$\begin{align*}\varphi|_{P",Q"}=(\varphi|_{P',Q'})|_{P",Q"}. \end{align*}$$
  2. (b) Let $P,Q,R,P',R'\in \mathrm {Ob}(\mathcal {T})$ , $\varphi \in \mathrm {Mor}_{\mathcal {T}}(P,Q)$ and $\psi \in \mathrm {Mor}_{\mathcal {T}}(Q,R)$ with $P'\leq P$ and $(\psi \circ \varphi )(P')\leq R'\leq R$ . Then

    $$\begin{align*}(\psi\circ \varphi)|_{P',R'}=\psi|_{\varphi(P'),R'}\circ \varphi|_{P',\varphi(P')}. \end{align*}$$
  3. (c) Let $x\in S$ and $\alpha =\delta _S(x)$ . For every $P\in \Delta $ , we have $\alpha (P)={}^x\!P\in \Delta $ and

    $$\begin{align*}\alpha|_{P,\alpha(P)}=\delta_{P,\alpha(P)}(x). \end{align*}$$
  4. (d) Let $P,Q,P'\in \mathrm {Ob}(\mathcal {T})$ and $\psi ,\varphi \in \mathrm {Mor}_{\mathcal {T}}(P,Q)$ such that $P'\leq P$ , $Q':=\psi (P')=\varphi (P')$ and $\varphi |_{P',Q'}=\psi |_{P',Q'}$ . Then $\varphi =\psi $ .

Proof. (a) Setting $\varphi '=\varphi |_{P',Q'}$ , it follows from the definition of the restriction that

$$\begin{align*}\iota_{Q',Q}\circ \varphi'=\varphi\circ\iota_{P',P}\mbox{ and }\iota_{Q",Q'}\circ \varphi'|_{P",Q"}=\varphi'\circ \iota_{P",P'}. \end{align*}$$

Hence,

$$ \begin{align*} \iota_{Q",Q}\circ \varphi'|_{P",Q"} &=\iota_{Q',Q}\circ \iota_{Q",Q'}\circ \varphi'|_{P",Q"}\\ &=\iota_{Q',Q}\circ \varphi'\circ \iota_{P",P'}\\ &=\varphi\circ \iota_{P',P}\circ \iota_{P",P'}\\ &=\varphi\circ \iota_{P",P}. \end{align*} $$

This implies (a).

(b) See [Reference Chermak7, Lemma A.7(b)].

(c) By axiom (B) of a transporter system as stated in [Reference Oliver and Ventura25, Definition 3.1], we have $\pi (\alpha )=c_{x^{-1}}$ , and thus $^x\!P=P^{x^{-1}}=\alpha (P)$ . As $\delta $ is a functor, it follows that

$$\begin{align*}\iota_{\alpha(P),S}\circ \delta_{P,\alpha(P)}(x)=\delta_{P,S}(1\cdot x)=\delta_{P,S}(x\cdot 1)=\delta_S(x)\circ \delta_{P,S}(1)=\alpha\circ \iota_{P,S}.\end{align*}$$

This shows (c).

(d) Assume the hypothesis of (d) and set $\varphi _0:=\varphi |_{P',Q'}=\psi |_{P',Q'}$ . It follows from the definition restrictions that

$$\begin{align*}\varphi\circ\iota_{P',P}=\iota_{Q',Q}\circ \varphi_0=\psi\circ\iota_{P',P}.\end{align*}$$

As every morphism in $\mathcal {T}$ is by [Reference Oliver and Ventura25, Lemma 3.2(d)] an epimorphism, it follows that $\varphi =\psi $ , as required.

Central to the considerations in this appendix is the following definition that generalises the definition of a normal pair of linking systems (compare [Reference Aschbacher, Kessar and Oliver2, Definition III.4.12]).

Definition A.3. Fix a pair of fusion systems $\mathcal {F}_0\subseteq \mathcal {F}$ over p-groups $S_0\leq S$ such that $\mathcal {F}_0$ is $\mathcal {F}$ -invariant. Let $\mathcal {T}_0\subseteq \mathcal {T}$ be transporter systems over $\mathcal {F}_0\subseteq \mathcal {F}$ , respectively. Then $\mathcal {T}_0$ is called normal in $\mathcal {T}$ (written as $\mathcal {T}_0\unlhd \mathcal {T}$ ) if the following hold:

  1. (i) $\mathrm {Ob}(\mathcal {T})=\{P\leq S\colon P\cap S_0\in \mathrm {Ob}(\mathcal {T}_0)\}$ ;

  2. (ii) for all $P\in \mathrm {Ob}(\mathcal {T}_0)$ and $\psi \in \mathrm {Mor}_{\mathcal {T}}(P,S_0)$ , there are morphisms $\gamma \in \mathrm {Aut}_{\mathcal {T}}(S_0)$ and $\psi _0\in \mathrm {Mor}_{\mathcal {T}_0}(P,S_0)$ such that $\psi =\gamma \circ \psi _0$ ; and

  3. (iii) for all $\gamma \in \mathrm {Aut}_{\mathcal {T}}(S_0)$ , $P,Q\in \mathrm {Ob}(\mathcal {T}_0)$ , and $\psi \in \mathrm {Mor}_{\mathcal {T}_0}(P,Q)$ ,

    $$\begin{align*}\gamma|_{Q,\gamma(Q)}\circ \psi\circ (\gamma|_{P,\gamma(P)})^{-1}\in\mathrm{Mor}_{\mathcal{T}_0}(\gamma(P),\gamma(Q)).\end{align*}$$

We also say then that $\mathcal {T}_0\unlhd \mathcal {T}$ is a normal pair of transporter systems over $\mathcal {F}_0\subseteq \mathcal {F}$ .

Saying that $\mathcal {T}_0\subseteq \mathcal {T}$ are transporter systems over $\mathcal {F}_0\subseteq \mathcal {F}$ , respectively, means here more precisely that $\mathcal {T}_0$ is a subcategory of $\mathcal {T}$ , that $(\mathcal {T},\delta ,\pi )$ is a transporter system over $\mathcal {F}$ (for appropriate functors $\delta $ and $\pi $ ) and that $(\mathcal {T}_0,\delta |_{\mathcal {T}_{\mathrm {Ob}(\mathcal {T}_0)}(S_0)},\pi |_{\mathcal {T}_0})$ is a transporter system over $\mathcal {F}_0$ .

We now prove that kernels of localities lead to normal pairs of transporter systems.

Proposition A.4. Let $(\mathcal {N},\Gamma ,T)$ be a kernel of a locality $(\mathcal {L},\Delta ,S)$ . Then $\mathcal {T}_{\Gamma }(\mathcal {N})\unlhd \mathcal {T}_{\Delta }(\mathcal {L})$ is a normal pair of transporter systems over $\mathcal {F}_T(\mathcal {N})\subseteq \mathcal {F}_S(\mathcal {L})$ .

Proof. It follows from the discussion above that $\mathcal {T}_{\Gamma }(\mathcal {N})\subseteq \mathcal {T}_{\Delta }(\mathcal {L})$ are transporter systems over $\mathcal {F}_T(\mathcal {N})\subseteq \mathcal {F}_S(\mathcal {L})$ . By Lemma 5.2(d), $\mathcal {F}_T(\mathcal {N})$ is $\mathcal {F}_S(\mathcal {L})$ -invariant. Referring to the properties (i),(ii),(iii) in Definition A.3, it follows from the definition of a kernel that (i) holds. Property (ii) is a consequence of Lemma 3.13. Property (iii) holds since $\mathcal {N}\unlhd \mathcal {L}$ .

We outline now how a locality can be constructed from a transporter system and use this afterwards to show that normal pairs of transporter systems lead to localities with kernels.

Let $(\mathcal {T},\delta ,\pi )$ be a transporter system over a fusion system $\mathcal {F}$ on S. Write $\mathrm {Iso}(\mathcal {T})$ for the set of all isomorphisms in $\mathcal {T}$ . Define a relation $\uparrow _{\mathcal {T}}$ on $\mathrm {Iso}(\mathcal {T})$ by writing

$$\begin{align*}\varphi\uparrow_{\mathcal{T}} \psi \end{align*}$$

if $\psi $ restricts to $\varphi $ . More precisely, this means $\varphi \uparrow _{\mathcal {T}}\psi $ , if $\varphi \in \mathrm {Iso}_{\mathcal {T}}(P',Q')$ and $\psi \in \mathrm {Iso}_{\mathcal {T}}(P,Q)$ for some $P,P',Q,Q'\in \mathrm {Ob}(\mathcal {T})$ with $P'\leq P$ , $Q'\leq Q$ , $\psi (P')=Q'$ and $\varphi =\psi |_{P',Q'}$ . Let then $\equiv _{\mathcal {T}}$ be the smallest equivalence relation on $\mathrm {Iso}(\mathcal {T})$ containing $\uparrow _{\mathcal {T}}$ . Write $[\varphi ]$ for the $\equiv _{\mathcal {T}}$ -equivalence class of $\varphi \in \mathrm {Iso}(\mathcal {T})$ and $\mathcal {L}(\mathcal {T})$ for the set of all equivalence classes of $\mathrm {Iso}(\mathcal {T})$ .

Remark A.5. We have $\varphi \equiv _{\mathcal {T}}\psi $ if and only if there exists a sequence $\varphi _1,\varphi _2,\cdots ,\varphi _k\in \mathrm {Iso}(\mathcal {T})$ such that, for all $i=1,2,\dots ,k-1$ ,

$$\begin{align*}\varphi_i\uparrow_{\mathcal{T}}\varphi_{i+1}\mbox{ or }\varphi_{i+1}\uparrow_{\mathcal{T}}\varphi_i. \end{align*}$$

By $\mathbf {D}$ denote the set of tuples $w=(f_1,f_2,\dots ,f_k)\in \mathbf {W}(\mathcal {L}(\mathcal {T}))$ for which there exist $\varphi _i\in f_i$ for $i=1,\dots ,k$ such that the composition $\varphi _1\circ \varphi _2\circ \cdots \circ \varphi _k$ is defined in the category $\mathcal {T}$ . Moreover, given such w and $\varphi _i$ , set

$$\begin{align*}\Pi(w):=[\varphi_1\circ\varphi_2\circ\cdots\circ \varphi_k]. \end{align*}$$

The map $\Pi \colon \mathbf {D}\longrightarrow \mathcal {L}(\mathcal {T})$ is well-defined. Together with $\Pi $ and the map

$$\begin{align*}\mathcal{L}(\mathcal{T})\longrightarrow\mathcal{L}(\mathcal{T}),[\varphi]\mapsto[\varphi^{-1}], \end{align*}$$

which is also well-defined, the set $\mathcal {L}(\mathcal {T})$ forms a partial group by [Reference Chermak7, Proposition A.9]. Moreover, the map

$$\begin{align*}S\longrightarrow\mathcal{L}(\mathcal{T}),\;x\mapsto [\delta_S(x)] \end{align*}$$

is an injective homomorphism of partial groups, and its image is a subgroup of $\mathcal {L}(\mathcal {T})$ . We will usually identify $x\in S$ with $[\delta _S(x)]\in \mathcal {L}(\mathcal {T})$ . With this identification, S is a subgroup of $\mathcal {L}(\mathcal {T})$ . Setting $\Delta :=\mathrm {Ob}(\mathcal {T})$ , it is shown in [Reference Chermak7, Proposition A.13] that $(\mathcal {L}(\mathcal {T}),\Delta ,S)$ is a locality. As we assume that $(\mathcal {T},\delta ,\pi )$ is a transporter system over $\mathcal {F}$ , it follows from [Reference Henke17, Lemma 4.4(a)] that $(\mathcal {L}(\mathcal {T}),\Delta ,S)$ is a locality over $\mathcal {F}$ . We will use these properties throughout without further reference.

Remark A.6. The partial group structure on $\mathcal {L}(\mathcal {T})$ constructed above is not exactly the same as the one constructed by Chermak [Reference Chermak7, Appendix A]. The reason is that Chermak consistently uses the ‘right-hand notation’ for maps and also for the category $\mathcal {T}$ . So if $\Pi \colon \mathbf {D}\rightarrow \mathcal {L}(\mathcal {T})$ is the partial product defined above and $\Pi '\colon \mathbf {D}'\rightarrow \mathcal {L}(\mathcal {T})$ is the product constructed by Chermak, then

$$\begin{align*}\mathbf {D}'=\{(f_n,f_{n-1},\dots,f_1)\colon (f_1,f_2,\dots,f_n)\in\mathbf {D}\} \end{align*}$$

and $\Pi '(f_n,f_{n-1},\dots ,f_1)=\Pi (f_1,f_2,\dots ,f_n)$ for all $(f_1,f_2,\dots ,f_n)\in \mathbf {D}$ . In particular, conjugation by $f\in \mathcal {L}(\mathcal {T})$ with respect to the partial group with product $\Pi '$ corresponds to conjugation by $f^{-1}$ in the partial group constructed above.

The main goal of this appendix is to prove the following theorem.

Theorem A.7. Let $\mathcal {F}_0\subseteq \mathcal {F}$ be fusion systems over $S_0\leq S$ such that $\mathcal {F}_0$ is $\mathcal {F}$ -invariant. Let $\mathcal {T}_0\unlhd \mathcal {T}$ be a normal pair of transporter systems over $\mathcal {F}_0\subseteq \mathcal {F}$ . Set $\Delta :=\mathrm {Ob}(\mathcal {T})$ , $\Gamma :=\mathrm {Ob}(\mathcal {T}_0)$ , $\mathcal {L}:=\mathcal {L}(\mathcal {T})$ and

$$\begin{align*}\mathcal{N}:=\{f\in\mathcal{L}\colon f\cap \mathrm{Iso}(\mathcal{T}_0)\neq\emptyset\}.\end{align*}$$

Then $(\mathcal {N},\Gamma ,S_0)$ is a kernel of $(\mathcal {L},\Delta ,S)$ with $\mathcal {F}_{S_0}(\mathcal {N})=\mathcal {F}_0$ . Moreover, there is an isomorphism $\mathcal {N}\rightarrow \mathcal {L}(\mathcal {T}_0)$ , which is the identity on $S_0$ , and there exists an invertible functor $\mathcal {T}_{\Gamma }(\mathcal {N})\rightarrow \mathcal {T}_0$ that is the identity on $\Gamma $ .

Remark A.8. Note that Theorem A.7 allows us to conclude results about normal pairs of transporter systems from our results about kernels of localities. Let us point out one example: Call a transporter system $(\mathcal {T},\delta ,\pi )$ over $\mathcal {F}$ cr-complete if $\mathcal {F}^{cr}\subseteq \mathrm {Ob}(\mathcal {T})$ . Consider a normal pair of transporter systems $\mathcal {T}_0\unlhd \mathcal {T}$ over $\mathcal {F}_0\subseteq \mathcal {F}$ . Then it follows from Theorem B that $\mathcal {T}_0$ is cr-complete if and only if $\mathcal {T}$ is cr-complete. Moreover, if so, then $\mathcal {F}_0$ is normal in $\mathcal {F}$ . As a particular consequence, if $\mathcal {T}_0\unlhd \mathcal {T}$ is a normal pair of linking systems over $\mathcal {F}_0\subseteq \mathcal {F}$ (as defined in [Reference Andersen, Oliver and Ventura1] and [Reference Oliver23]), then $\mathcal {F}_0$ is always normal in $\mathcal {F}$ .

To prove Theorem A.7, we assume the following hypothesis:

From now on, let $\mathcal {F}_0\subseteq \mathcal {F}$ be fusion systems over $S_0\leq S$ , respectively, such that $\mathcal {F}_0$ is $\mathcal {F}$ -invariant, and let $\mathcal {T}_0\unlhd \mathcal {T}$ be a normal pair of transporter systems over $\mathcal {F}_0\subseteq \mathcal {F}$ , respectively.

More precisely, let $(\mathcal {T},\delta ,\pi )$ be a transporter system over $\mathcal {F}$ and $(\mathcal {T},\delta |_{\mathcal {T}_{\mathrm {Ob}(\mathcal {T}_0)}(S)},\pi |_{\mathcal {T}_0})$ be a transporter system over $\mathcal {F}_0$ .

Lemma A.9. Let $P,Q\in \mathrm {Ob}(\mathcal {T}_0)$ and $\psi \in \mathrm {Iso}_{\mathcal {T}}(P,Q)$ . Then there exist $\gamma \in \mathrm {Aut}_{\mathcal {T}}(S_0)$ , $R\in \mathrm {Ob}(\mathcal {T}_0)$ and $\psi _0\in \mathrm {Iso}_{\mathcal {T}_0}(P,R)$ such that $\gamma (R)=Q$ and $\psi =\gamma |_{R,Q}\circ \psi _0$ .

Proof. Notice that $\hat {\psi }:=\iota _{Q,S_0}\circ \psi \in \mathrm {Mor}_{\mathcal {T}_0}(P,S_0)$ with $\hat {\psi }|_{P,Q}=\psi $ . Hence, by axiom (ii) of Definition A.3, there exist $\gamma \in \mathrm {Aut}_{\mathcal {T}}(S_0)$ and $\hat {\psi }_0\in \mathrm {Mor}_{\mathcal {T}_0}(P,S_0)$ such that

$$\begin{align*}\hat{\psi}=\gamma\circ \hat{\psi}_0.\end{align*}$$

Setting $R:=\hat {\psi }_0(P)$ and $\psi _0:=\hat {\psi }_0|_{P,R}\in \mathrm {Iso}_{\mathcal {T}_0}(P,R)$ , we have $\gamma (R)=Q$ and it follows from Lemma A.2(b) that $\psi =\hat {\psi }|_{P,Q}=\gamma |_{R,Q}\circ \psi _0$ .

Lemma A.10. Let $P,\overline {P},Q,\overline {Q}\in \mathrm {Ob}(\mathcal {T}_0)$ with $P\leq \overline {P}$ and $Q\leq \overline {Q}$ . Let $\psi \in \mathrm {Iso}_{\mathcal {T}}(\overline {P},\overline {Q})$ with $\psi (P)\leq Q$ and $\psi |_{P,Q}\in \mathrm {Mor}_{\mathcal {T}_0}(P,Q)$ . Then $\psi $ is a morphism in $\mathcal {T}_0$ .

Proof. As $\overline {P}$ is a p-group, P is subnormal in $\overline {P}$ . As $\mathrm {Ob}(\mathcal {T}_0)$ is overgroup-closed in $S_0$ , induction on $|\overline {P}:P|$ allows us to reduce to the case that $P\unlhd \overline {P}$ . Moreover, replacing Q by $\psi (P)$ , we may assume that $\psi (P)=Q$ , $\varphi :=\psi |_{P,Q}\in \mathrm {Iso}_{\mathcal {T}_0}(P,Q)$ and $Q=\psi (P)\unlhd \psi (\overline {P})=\overline {Q}$ . It follows from [Reference Oliver and Ventura25, Lemma 3.3] that $\varphi \circ \delta _P(x)=\delta _Q(\pi (\psi )(x))\circ \varphi $ for all $x\in \overline {P}$ . As $\pi (\psi )(\overline {P})\leq \overline {Q}$ , this implies that $\varphi \circ \delta _P(\overline {P})\circ \varphi ^{-1}\leq \delta _Q(\overline {Q})$ . Hence, by Axiom II of a transporter systems (as stated in [Reference Oliver and Ventura25, Definition 3.1]) applied to the transporter system $\mathcal {T}_0$ , there exists $\overline {\varphi }\in \mathrm {Mor}_{\mathcal {T}_0}(\overline {P},\overline {Q})$ with $\overline {\varphi }|_{P,Q}=\varphi =\psi |_{P,Q}$ . It follows now from Lemma A.2(d) that $\psi =\overline {\varphi }$ is a morphism in $\mathcal {T}_0$ .

Set now $\Delta :=\mathrm {Ob}(\mathcal {T})$ , $\Gamma :=\mathrm {Ob}(\mathcal {T}_0)$ , $\mathcal {L}:=\mathcal {L}(\mathcal {T})$ and

$$\begin{align*}\mathcal{N}:=\{f\in\mathcal{L}\colon f\cap \mathrm{Iso}(\mathcal{T}_0)\neq\emptyset\}. \end{align*}$$

As before, we write $\Pi \colon \mathbf {D}\rightarrow \mathcal {L}$ for the product on $\mathcal {L}$ .

Notation A.11. For $P\leq S$ , set $P_0:=P\cap S_0$ . Similarly, define $Q_0,P^{\prime }_0,Q^{\prime }_0$ for subgroups $Q,P',Q'\leq S$ . For $P,Q\in \Delta $ and $\varphi \in \mathrm {Iso}_{\mathcal {T}}(P,Q)$ , set moreover

$$\begin{align*}\varphi^0:=\varphi|_{P_0,Q_0}. \end{align*}$$

As $S_0$ is strongly closed, we have in the situation above that $\varphi (P_0)=\pi (\varphi )(P_0)=(\pi (\varphi )(P))\cap S_0=Q_0$ . So $\varphi ^0$ is well-defined and an element of $\mathrm {Iso}_{\mathcal {T}}(P_0,Q_0)$ . Observe also that $\varphi ^0\uparrow _{\mathcal {T}}\varphi $ and thus $[\varphi ]=[\varphi ^0]$ for all $\varphi \in \mathrm {Iso}(\mathcal {T})$ .

Lemma A.12. If $\varphi ,\psi \in \mathrm {Iso}(\mathcal {T})$ with $\psi \uparrow _{\mathcal {T}}\varphi $ , then $\psi ^0\uparrow _{\mathcal {T}}\varphi ^0$ .

Proof. Let $P'\leq P$ and $Q'\leq Q$ be objects in $\mathcal {T}$ such that $\varphi \in \mathrm {Mor}_{\mathcal {T}}(P,Q)$ , $\psi \in \mathrm {Mor}_{\mathcal {T}}(P',Q')$ , $\varphi (P')\leq Q'$ and $\varphi |_{P',Q'}=\psi $ . Applying Lemma A.2(a) twice gives then

$$\begin{align*}\varphi^0|_{P^{\prime}_0,Q^{\prime}_0}=\varphi|_{P^{\prime}_0,Q^{\prime}_0}=\psi|_{P_0',Q_0'}=\psi^0. \end{align*}$$

Hence, $\psi ^0\uparrow _{\mathcal {T}}\varphi ^0$ .

Lemma A.13. The following hold:

  1. (a) If $\varphi ,\psi \in \mathrm {Iso}(\mathcal {T}_0)$ , then $\varphi \equiv _{\mathcal {T}}\psi $ if and only if $\varphi \equiv _{\mathcal {T}_0}\psi $ .

  2. (b) Let $\varphi \in \mathrm {Iso}(\mathcal {T}_0)$ and $\psi \in \mathrm {Iso}_{\mathcal {T}}(P,Q)$ for some $P,Q\in \mathrm {Ob}(\mathcal {T}_0)$ . If $\varphi \equiv _{\mathcal {T}}\psi $ , then $\psi \in \mathrm {Iso}(\mathcal {T}_0)$ .

  3. (c) If $\varphi \in \mathrm {Iso}(\mathcal {T})$ with $[\varphi ]\in \mathcal {N}$ , then $\varphi ^0\in \mathrm {Iso}(\mathcal {T}_0)$ .

Proof. (a,b) Clearly, the relation $\uparrow _{\mathcal {T}_0}$ is contained in the relation $\uparrow _{\mathcal {T}}$ , so, if $\varphi ,\psi \in \mathrm {Iso}(\mathcal {T}_0)$ with $\varphi \equiv _{\mathcal {T}_0}\psi $ , then $\varphi \equiv _{\mathcal {T}}\psi $ .

Suppose now that $\varphi \in \mathrm {Iso}(\mathcal {T}_0)$ and $\psi \in \mathrm {Iso}_{\mathcal {T}}(P,Q)$ for some $P,Q\in \mathrm {Ob}(\mathcal {T}_0)$ . Assume $\varphi \equiv _{\mathcal {T}}\psi $ . To show (a) and (b), it remains to show that $\psi \in \mathrm {Iso}(\mathcal {T}_0)$ and $\varphi \equiv _{\mathcal {T}_0}\psi $ . By Remark A.5, there exists a series $\varphi =\varphi _1,\varphi _2,\dots ,\varphi _n=\psi \in \mathrm {Iso}(\mathcal {T})$ such that, for all $i=1,2,\dots ,n-1$ , we have $\varphi _i\uparrow _{\mathcal {T}} \varphi _{i+1}$ or $\varphi _{i+1}\uparrow _{\mathcal {T}}\varphi _i$ . Notice that $\varphi ^0=\varphi $ and $\psi ^0=\psi $ as $P,Q\in \mathrm {Ob}(\mathcal {T}_0)$ . Thus, Lemma A.12 allows us to replace $\varphi _1,\varphi _2,\dots ,\varphi _n$ by $\varphi _1^0,\varphi _2^0,\dots ,\varphi _n^0$ . Thus, we may assume that $\varphi _1,\varphi _2,\dots ,\varphi _n$ are isomorphisms in $\mathcal {T}$ between objects of $\mathcal {T}_0$ . Then Lemma A.10 implies that for all $i=1,2,\dots ,n-1$ , $\varphi _i\in \mathrm {Iso}(\mathcal {T}_0)$ if and only if $\varphi _{i+1}\in \mathrm {Iso}(\mathcal {T}_0)$ . As $\varphi _1=\varphi \in \mathrm {Iso}(\mathcal {T}_0)$ by assumption, it follows therefore inductively that $\varphi _1,\varphi _2,\dots ,\varphi _n\in \mathrm {Iso}(\mathcal {T}_0)$ . In particular, $\psi =\varphi _n\in \mathrm {Iso}(\mathcal {T}_0)$ . For $\alpha ,\beta \in \mathrm {Iso}(\mathcal {T}_0)$ it is easy to observe that $\alpha \uparrow _{\mathcal {T}}\beta $ if and only if $\alpha \uparrow _{\mathcal {T}_0}\beta $ . Hence, we also have $\varphi _i\uparrow _{\mathcal {T}_0} \varphi _{i+1}$ or $\varphi _{i+1}\uparrow _{\mathcal {T}_0}\varphi _i$ for $i=1,2,\dots ,n-1$ . This shows $\varphi \equiv _{\mathcal {T}_0}\psi $ . So (a) and (b) hold.

(c) If $\varphi \in \mathrm {Iso}(\mathcal {T})$ with $[\varphi ]\in \mathcal {N}$ , then by definition of $\mathcal {N}$ , there exists $\psi \in \mathrm {Iso}(\mathcal {T}_0)$ with $\psi \in [\varphi ]$ . Then $\varphi ^0\equiv _{\mathcal {T}}\varphi \equiv _{\mathcal {T}}\psi $ . Hence, it follows from part (b) that $\varphi ^0\in \mathrm {Iso}(\mathcal {T}_0)$ .

For the following lemma, recall that we identify $x\in S$ with $[\delta _S(x)]\in \mathcal {L}$ .

Lemma A.14.

  1. (a) $\mathcal {N}$ is a partial subgroup of $\mathcal {L}$ with $\mathcal {N}\cap S=S_0$ .

  2. (b) Set $\mathcal {L}_0:=\mathcal {L}(\mathcal {T}_0)$ , write $[\varphi ]_0$ for the $\equiv _{\mathcal {T}_0}$ -equivalence class of $\varphi \in \mathrm {Iso}(\mathcal {T}_0)$ , and identify $x\in S_0$ with $[\delta _{S_0}(x)]_0$ . Then the map

    $$\begin{align*}\theta\colon\mathcal{N}\rightarrow \mathcal{L}_0,f\mapsto f\cap \mathrm{Iso}(\mathcal{T}_0)\end{align*}$$
    is an isomorphism of localities that restricts to the identity on $S_0$ .

Proof. We prove first

(A.1) $$ \begin{align} \mathcal{N}\mbox{ is closed under inversion.} \end{align} $$

For the proof, note that every $n\in \mathcal {N}$ can be written as $n=[\varphi ]$ for some $\varphi \in \mathrm {Iso}(\mathcal {T}_0)$ . Then $\varphi ^{-1}\in \mathrm {Iso}(\mathcal {T}_0)$ , and it follows from the definition of the inversion on $\mathcal {L}$ that $n^{-1}=[\varphi ^{-1}]\in \mathcal {N}$ . This shows (A.1). We argue next that

(A.2) $$ \begin{align} \mbox{the map} \ \theta \ \mbox{ is well-defined and a bijection.} \end{align} $$

Indeed, it follows from Lemma A.13(a) that $[\varphi ]\cap \mathrm {Iso}(\mathcal {T}_0)=[\varphi ]_0$ for all $\varphi \in \mathrm {Iso}(\mathcal {T}_0)$ . Hence, $\theta $ is well-defined and surjective. Note also that $f\cap g\neq \emptyset $ for all $f,g\in \mathcal {N}$ with $\theta (f)=\theta (g)$ . As the equivalence classes of $\equiv _{\mathcal {T}}$ form a partition of $\mathrm {Iso}(\mathcal {T})$ , we can therefore conclude that the map $\theta $ is injective. So (A.2) holds. We show next

(A.3) $$ \begin{align} S_0\subseteq \mathcal{N}\cap S\mbox{ and }\theta|_{S_0}=\operatorname{id}_{S_0}. \end{align} $$

For the proof, let $x\in S_0$ . Lemma A.2(c) implies $\delta _S(x)|_{S_0}=\delta _{S_0}(x)\in \mathrm {Aut}_{\mathcal {T}_0}(S_0)\subseteq \mathrm {Iso}(\mathcal {T}_0)$ , so $x=[\delta _S(x)]=[\delta _{S_0}(x)]\in \mathcal {N}$ . Moreover, $x=[\delta _S(x)]$ gets mapped $[\delta _{S_0}(x)]_0$ , which we also identify with x (as stated in part (b)). This proves (A.3). We show next

(A.4) $$ \begin{align} \mathcal{N} \ \mbox{is a partial subgroup, and} \ \theta \ \mbox{is a homomorphism of partial groups.} \end{align} $$

For the proof, let $w:=(n_1,n_2,\dots ,n_k)\in \mathbf {W}(\mathcal {N})\cap \mathbf {D}$ . As $w\in \mathbf {D}$ , there exist $\varphi _i\in n_i$ for $i=1,2,\dots ,k$ such that the composition $\varphi _1\circ \varphi _2\circ \cdots \circ \varphi _k$ is defined in $\mathcal {T}$ . Note that then the composition $\varphi _1^0\circ \varphi _2^{\circ } \cdots \circ \varphi _k^0$ is also defined. Moreover, as $\varphi _i^0\uparrow _{\mathcal {T}}\varphi _i$ , we have $\varphi _i^0\in n_i$ for $i=1,2,\dots ,k$ . By Lemma A.13(c), $\varphi _i^0\in \mathrm {Iso}(\mathcal {T}_0)$ . So, replacing $\varphi _1,\varphi _2,\dots ,\varphi _k$ by $\varphi _1^0,\varphi _2^0,\dots ,\varphi _k^0$ , we may assume $\varphi _i\in \mathrm {Iso}(\mathcal {T}_0)$ . Then $\varphi _1\circ \varphi _2\circ \cdots \circ \varphi _k\in \mathrm {Iso}(\mathcal {T}_0)$ , and hence

$$\begin{align*}\Pi(w)=[\varphi_1\circ \varphi_2\circ\cdots\circ \varphi_k]\in\mathcal{N}. \end{align*}$$

Together with (A.1), this shows that $\mathcal {N}$ is a partial subgroup.

Note also that $\theta (n_i)=n_i\cap \mathrm {Iso}(\mathcal {T}_0)=[\varphi _i]_0$ for $i=1,2,\dots ,k$ . Moreover, the composition $\varphi _1\circ \varphi _2\circ \cdots \circ \varphi _k$ is defined in $\mathcal {T}_0$ . Hence, writing $\Pi _0\colon \mathbf {D}_0\rightarrow \mathcal {L}_0$ for the product on $\mathcal {L}_0$ (defined in the usual way), it follows that $\theta ^*(w)=(\theta (n_1),\theta (n_2),\dots ,\theta (n_k))\in \mathbf {D}_0$ and

$$\begin{align*}\Pi_0(\theta^*(w))=[\varphi_1\circ\varphi_2\circ\cdots\circ\varphi_n]_0=[\varphi_1\circ\varphi_2\circ\cdots\circ\varphi_n]\cap \mathrm{Iso}(\mathcal{T}_0)=\Pi(w)\cap\mathrm{Iso}(\mathcal{T}_0)=\theta(\Pi(w)). \end{align*}$$

This proves that $\theta $ is a homomorphism of partial groups, and (A.4) holds.

For the proofs of the next two properties, recall Definition 3.15 and the notation introduced there. We show next that

(A.5) $$ \begin{align} \theta \ \mbox{is an isomorphism of partial groups.} \end{align} $$

Writing again $\Pi _0\colon \mathbf {D}_0\rightarrow \mathcal {L}_0$ for the product on $\mathcal {L}_0$ , it is by (A.2) and (A.4) sufficient to prove that $\mathbf {D}_0\subseteq \theta ^*(\mathbf {D}\cap \mathbf {W}(\mathcal {N}))$ . For the proof of this property, let $u=(g_1,g_2,\dots ,g_n)\in \mathbf {D}_0$ , and fix $\varphi _i\in g_i$ for $i=1,2,\dots ,n$ such that the composition $\varphi _1\circ \varphi _2\circ \cdots \circ \varphi _n$ is defined in $\mathcal {T}_0$ . Note that such $\varphi _i$ exist by definition of $\mathbf {D}_0$ . Then $\varphi _i\in \mathrm {Iso}(\mathcal {T}_0)$ , and hence $f_i:=[\varphi _i]\in \mathcal {N}$ with $\theta (f_i)=g_i$ . As the composition $\varphi _1\circ \varphi _2\circ \cdots \circ \varphi _n$ is defined in $\mathcal {T}_0$ , it is also defined in $\mathcal {T}$ . Therefore, $w:=(f_1,\dots ,f_n)\in \mathbf {D}\cap \mathbf {W}(\mathcal {N})$ with $\theta ^*(w)=u$ . This shows $\mathbf {D}_0\subseteq \theta ^*(\mathbf {D}\cap \mathbf {W}(\mathcal {N}))$ , and thus (A.5).

Note that (b) holds by (A.3) and (A.5). Moreover, $S_0\subseteq S\cap \mathcal {N}$ . Also recall that $\mathcal {N}$ is a partial subgroup of $\mathcal {L}$ by (A.4) and that $(\mathcal {L}_0,S_0,\Gamma )$ is by [Reference Chermak7, Proposition A.13] a locality. In particular, $S_0$ is a maximal p-subgroup of $\mathcal {L}_0$ and thus by (b) also a maximal p-subgroup of $\mathcal {N}$ . This implies $S\cap \mathcal {N}=S_0$ .

Lemma A.15. Let $P,Q,R,P',Q',R'\in \mathrm {Ob}(\mathcal {T}_0)$ , $\varphi \in \mathrm {Iso}_{\mathcal {T}}(P,Q)$ , $\psi \in \mathrm {Iso}_{\mathcal {T}}(P',Q')$ , $\varphi _0\in \mathrm {Iso}_{\mathcal {T}_0}(P,R)$ , $\psi _0\in \mathrm {Iso}_{\mathcal {T}_0}(P',R')$ , $\alpha ,\gamma \in \mathrm {Aut}_{\mathcal {T}}(S_0)$ such that $\alpha (R)=Q$ , $\gamma (R')=Q'$ ,

$$\begin{align*}\varphi=\alpha|_{R,Q}\circ\varphi_0\mbox{ and }\psi=\gamma|_{R',Q'}\circ\psi_0. \end{align*}$$

If $\varphi \equiv _{\mathcal {T}}\psi $ , then $\gamma ^{-1}\circ \alpha \in \mathrm {Aut}_{\mathcal {T}_0}(S_0)$ .

Proof. By Remark A.5, there exists a series $\varphi =\beta _1,\beta _2,\dots ,\beta _n=\psi \in \mathrm {Iso}(\mathcal {T})$ such that $\beta _i\uparrow \beta _{i+1}$ or $\beta _{i+1}\uparrow \beta _i$ for $i=1,\dots ,n-1$ . By Lemma A.9, every $\beta _i$ can be factored as a restriction of an automorphism $\alpha _i\in \mathrm {Aut}_{\mathcal {T}}(S_0)$ with an isomorphism in $\mathcal {T}_0$ . We may assume that $\alpha _1=\alpha $ and $\alpha _n=\gamma $ . If $\alpha _{i+1}^{-1}\circ \alpha _i\in \mathrm {Aut}_{\mathcal {T}_0}(S_0)$ for $i=1,\dots ,n-1$ , then

$$\begin{align*}\gamma^{-1}\circ\alpha=\alpha_n^{-1}\circ\alpha_1=(\alpha_n^{-1}\circ \alpha_{n-1})\circ (\alpha_{n-1}^{-1}\circ \alpha_{n-2})\circ\cdots \circ (\alpha_2^{-1}\circ\alpha_1)\in\mathrm{Aut}_{\mathcal{T}_0}(S_0).\end{align*}$$

Thus, we may reduce to the case $\varphi \uparrow _{\mathcal {T}}\psi $ or $\psi \uparrow _{\mathcal {T}}\varphi $ . As $\gamma ^{-1}\circ \alpha \in \mathrm {Aut}_{\mathcal {T}_0}(S_0)$ if and only if $\alpha ^{-1}\circ \gamma =(\gamma ^{-1}\circ \alpha )^{-1}\in \mathrm {Aut}_{\mathcal {T}_0}(S_0)$ , we may indeed assume that

$$\begin{align*}\varphi\uparrow_{\mathcal{T}}\psi.\end{align*}$$

This means $P\leq P'$ , $\psi (P)=Q\leq Q'$ and $\varphi =\psi |_{P,Q}$ . Setting $X:=\psi _0(P)$ , Lemma A.2(b) gives

$$\begin{align*}\alpha|_{R,Q}\circ\varphi_0=\varphi=\psi|_{P,Q}=\gamma|_{X,Q}\circ \psi_0|_{P,X}. \end{align*}$$

As $\varphi _0$ and $\psi _0$ are morphisms in $\mathcal {T}_0$ , this implies, together with Lemma A.2(b), that

$$\begin{align*}(\gamma^{-1}\circ \alpha)|_{R,X}=(\gamma|_{X,Q})^{-1}\circ \alpha|_{R,Q}=\psi_0|_{P,X}\circ \varphi_0^{-1}\in\mathrm{Iso}_{\mathcal{T}_0}(R,X). \end{align*}$$

Hence, Lemma A.10 gives $\gamma ^{-1}\circ \alpha \in \mathrm {Aut}_{\mathcal {T}_0}(S_0)$ .

Lemma A.16. We have $\mathcal {N}\unlhd \mathcal {L}$ .

Proof. Recall that $\mathcal {N}$ is a partial subgroup of $\mathcal {L}$ by Lemma A.14(a). Let $f\in \mathcal {L}$ and $n\in \mathcal {N}$ such that $(f^{-1},n,f)\in \mathbf {D}$ . Then there exist

$$\begin{align*}\varphi^{-1}\in f^{-1},\;\rho\in n,\;\psi\in f \end{align*}$$

such that the composition $\varphi ^{-1}\circ \rho \circ \psi $ is defined. Replacing $\varphi ,\rho ,\psi $ by $\varphi ^0,\rho ^0,\psi ^0$ , we may assume that for some $P,Q,P',Q'\in \mathrm {Ob}(\mathcal {T}_0)$ , we have

$$\begin{align*}\psi\in\mathrm{Iso}_{\mathcal{T}}(Q',P'),\;\rho\in\mathrm{Iso}_{\mathcal{T}_0}(P',P),\;\varphi^{-1}\in\mathrm{Iso}_{\mathcal{T}}(P,Q). \end{align*}$$

By Lemma A.9, there exist then $R,R'\in \mathrm {Ob}(\mathcal {T}_0)$ , $\varphi _0\in \mathrm {Mor}_{\mathcal {T}_0}(P,R)$ , $\psi _0\in \mathrm {Mor}_{\mathcal {T}_0}(P',R')$ and $\alpha ,\gamma \in \mathrm {Aut}_{\mathcal {T}_0}(S_0)$ such that $\alpha (R)=Q$ , $\gamma (R')=Q'$ ,

$$\begin{align*}\varphi^{-1}=\alpha|_{R,Q}\circ\varphi_0\mbox{ and }\psi^{-1}=\gamma|_{R',Q'}\circ\psi_0. \end{align*}$$

As $f=[\psi ]$ , we have $[\psi ^{-1}]=f^{-1}=[\varphi ^{-1}]$ , so $\psi ^{-1}\equiv _{\mathcal {T}}\varphi ^{-1}$ . Thus, Lemma A.15 (applied with $(\psi ^{-1},\varphi ^{-1})$ in place of $(\varphi ,\psi )$ ) implies that $\gamma ^{-1}\circ \alpha \in \mathrm {Aut}_{\mathcal {T}_0}(S_0)$ . By definition of the product $\Pi $ on $\mathcal {L}$ , we have

$$\begin{align*}n^f=\Pi(f^{-1},n,f)=[\varphi^{-1}\circ\rho\circ\psi]. \end{align*}$$

Using Lemma A.2(b), observe that

$$ \begin{align*} \varphi^{-1}\circ\rho\circ\psi &= \alpha|_{R,Q}\circ\varphi_0\circ\rho\circ\psi_0^{-1}\circ (\gamma|_{R',Q'})^{-1}\\ &= \alpha|_{R,Q}\circ \varphi_0\circ\rho\circ\psi_0^{-1}\circ (\gamma^{-1}\circ \alpha)|_{\alpha^{-1}(Q'),R'}\circ (\alpha|_{\alpha^{-1}(Q'),Q'})^{-1}. \end{align*} $$

As $\varphi _0\circ \rho \circ \psi _0^{-1}\circ (\gamma ^{-1}\circ \alpha )|_{\alpha ^{-1}(Q'),R'}\in \mathrm {Iso}(\mathcal {T}_0)$ , it follows from axiom (iii) in Definition A.3 that $\varphi ^{-1}\circ \rho \circ \psi \in \mathrm {Iso}(\mathcal {T}_0)$ . Hence, $n^f=[\varphi ^{-1}\circ \rho \circ \psi ]\in \mathcal {N}$ . This shows the assertion.

Proof of Theorem A.7.

By Lemmas A.14(a) and A.16, $\mathcal {N}$ is a partial normal subgroup of $\mathcal {L}$ with $S\cap \mathcal {N}=S_0$ . Using that $\mathrm {Ob}(\mathcal {T}_0)\subseteq \mathrm {Ob}(\mathcal {T})$ and axiom (i) in Definition A.3 holds, we can see that $\Gamma =\{P\cap S_0\colon P\in \Delta \}\subseteq \Delta $ . Hence, $(\mathcal {N},\Gamma ,S_0)$ is a kernel of $(\mathcal {L},\Delta ,S)$ .

By Lemma A.14(b), there is an isomorphism $\theta \colon \mathcal {N}\rightarrow \mathcal {L}_0:=\mathcal {L}(\mathcal {T}_0)$ that (with appropriate identifications) restricts to the identity on $S_0$ . In particular, $\mathcal {F}_{S_0}(\mathcal {N})=\mathcal {F}_{S_0}(\mathcal {L}_0)$ , and there exists an invertible functor $\mathcal {T}_{\Gamma }(\mathcal {N})\rightarrow \mathcal {T}_{\Gamma }(\mathcal {L}_0)$ that is the identity on $\Gamma $ . We have seen in our general discussion above that $\mathcal {F}=\mathcal {F}_S(\mathcal {L})$ . So we have similarly that $\mathcal {F}_0=\mathcal {F}_{S_0}(\mathcal {L}_0)=\mathcal {F}_S(\mathcal {N})$ . It is moreover shown in [Reference Chermak7, Lemma A.15] that there is an invertible functor $\mathcal {T}_{\Gamma }(\mathcal {L}_0)\rightarrow \mathcal {T}_0$ that restricts to the identity on $\Gamma $ . This implies the assertion.

Conflicts of Interest

None.

Financial Support

The authors were partially supported by EPSRC grant no. EP/R010048/1. Moreover, they would like to thank the Isaac Newton Institute for Mathematical Sciences, Cambridge, for support and hospitality during the programme Groups, representations and applications: new perspectives, where some of the work on this paper was undertaken and supported by EPSRC grant no EP/R014604/1. Open Access Funding by the Publication Fund of the TU Dresden.

Footnotes

1 It should be noted here that the term ‘proper locality’ used in [Reference Chermak and Henke10] means exactly the same as the term ‘linking locality’.

References

Andersen, K. K. S., Oliver, B. and Ventura, J., Reduced, tame and exotic fusion systems. Proc. Lond. Math. Soc. (3), 105(2012), no. 1, 87152. http://doi.org/10.1112/plms/pdr065 CrossRefGoogle Scholar
Aschbacher, M., Kessar, R. and Oliver, B., Fusion systems in algebra and topology. London Mathematical Society Lecture Note Series, 391, Cambridge University Press, Cambridge, 2011.CrossRefGoogle Scholar
Aschbacher, M., The generalized Fitting subsystem of a fusion system. Mem. Amer. Math. Soc., 209(2011), no. 986, vi+110.Google Scholar
Aschbacher, M., On fusion systems of component type. Mem. Amer. Math. Soc., 257(2018), no. 1236, vi+182.Google Scholar
Broto, C., Castellana, N., Grodal, J., Levi, R. and Oliver, B., Subgroup families controlling $p$ -local finite groups. Proc. London Math. Soc. (3) 91(2005), no. 2, 325354. http://doi.org/10.1112/S0024611505015327 Google Scholar
Broto, C., Castellana, N., Grodal, J., Levi, R. and Oliver, B., Extensions of p-local finite groups. Trans. Amer. Math. Soc. 359(2007), no. 8, 37913858. http://doi.org/10.1090/S0002-9947-07-04225-0 Google Scholar
Chermak, A., Fusion systems and localities. Acta Math. 211(2013), no. 1, 47139. http://doi.org/10.1007/s11511-013-0099-5 Google Scholar
Chermak, A., Finite Localities I. Forum Math. Sigma 10(2022), Paper No. e43, 31. http://doi.org/10.1017/fms.2022.31 Google Scholar
Chermak, A., Finite localities III. 2016. arxiv:1610.06161v2 Google Scholar
Chermak, A. and Henke, E., Fusion systems and localities – a dictionary. Advances in Mathematics (in press) arxiv:1706.05343v2 Google Scholar
Glauberman, G. and Lynd, J., Control of fixed points and existence and uniqueness of centric linking systems. Invent. Math. 206(2016), no. 2, 441484. http://doi.org/10.1007/s00222-016-0657-5 Google Scholar
Gonzalez, A., An extension theory for partial groups and localities. 2015. arxiv:1507.04392v2 Google Scholar
Henke, E., Products in Fusion systems. J. Algebra 376(2013), 300319. http://doi.org/10.1016/j.jalgebra.2012.11.037 Google Scholar
Henke, E., Products of partial normal subgroups. Pacific J. Math. 279(2015), no. 1-2, 255268. http://doi.org/10.2140/pjm.2015.279.255 CrossRefGoogle Scholar
Henke, E., Centralizers of normal subsystems revisited. J. Algebra 511(2018), 364387. http://doi.org/10.1016/j.jalgebra.2018.05.034 CrossRefGoogle Scholar
Henke, E., Subcentric linking systems. Trans. Amer. Math. Soc. 371(2019), no. 5, 33253373. http://doi.org/10.1090/tran/7388 CrossRefGoogle Scholar
Henke, E., Extensions of homomorphisms between localities. Forum Math. Sigma 9(2021), Paper No. e63, 31. http://doi.org/10.1017/fms.2021.57 CrossRefGoogle Scholar
Henke, E., Commuting partial normal subgroups and regular localities. 2021. arxiv:2103.00955v3 Google Scholar
Henke, E., Normalizers and centralizers of subnormal subsystems of fusion systems. 2021. arxiv:2109.14038 CrossRefGoogle Scholar
Henke, E., Some products in fusion systems and localities. Preprint 2021, accepted to Proc. Amer. Math. Soc., arxiv:2107.00301v2 Google Scholar
Kurzweil, H. and Stellmacher, B., The theory of finite groups – an introduction. Universitext. Springer-Verlag, New York, 2004, xii+387. http://doi.org/10.1007/b97433 CrossRefGoogle Scholar
Meierfrankenfeld, U. and Stellmacher, B., Applications of the FF-module theorem and related results. J. Algebra 351(2012), 64106. http://doi.org/10.1016/j.jalgebra.2011.10.028 CrossRefGoogle Scholar
Oliver, B., Extensions of linking systems and fusion systems. Trans. Amer. Math. Soc. 362(2010), no. 10, 54835500. http://doi.org/10.1090/S0002-9947-2010-05022-6 CrossRefGoogle Scholar
Oliver, B., Existence and uniqueness of linking systems: Chermak’s proof via obstruction theory. Acta Math. 211(2013), no. 1, 141175. http://doi.org/10.1007/s11511-013-0100-3 CrossRefGoogle Scholar
Oliver, B. and Ventura, J., Extensions of linking systems with p-group kernel. Math. Ann. 338(2007), no. 4, 9831043. http://doi.org/10.1007/s00208-007-0104-4 CrossRefGoogle Scholar
Roberts, K. and Shpectorov, S., On the definition of saturated fusion systems. J. Group Theory 12(2009), no. 5, 679687. http://doi.org/10.1515/JGT.2009.003 CrossRefGoogle Scholar