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

Cohomology on the centric orbit category of a fusion system

George Glauberman Department of Mathematics
University of Chicago
5734 S. University Ave
Chicago, IL 60637
[email protected]
 and  Justin Lynd Department of Mathematics
University of Louisiana at Lafayette
Lafayette, LA 70504
[email protected]
Abstract.

We study here the higher derived limits of mod pp cohomology on the centric orbit category of a saturated fusion system on a finite pp-group. It is an open problem whether all such higher limits vanish. This is known in many cases, including for fusion systems realized by a finite group and for many classes of fusion systems which are not so realized. We prove that the higher limits of HjH^{j} vanish provided jp2j\leq p-2, by showing that the same is true for the contravariant part of a simple Mackey composition factor of HjH^{j} under the same conditions.

2020 Mathematics Subject Classification:
Primary 55R35, 20E25, Secondary 20J06, 20D20, 55R40
The first author was partially supported by a Simons Foundation Collaboration Grant. The second author was partially supported by NSF Grant DMS-1902152. The authors thank these organizations for their support. The authors would like to thank the Isaac Newton Institute for Mathematical Sciences, Cambridge, for support and hospitality during the program Groups, representations and applications, where work on this paper was undertaken and supported by EPSRC grant no EP/R014604/1.

1. Introduction

The purpose of this note is to show that Hj(,𝔽p)H^{j}(-,\mathbb{F}_{p}), considered as a contravariant functor on the centric orbit category of a saturated fusion system on a pp-group, has vanishing higher derived limits provided jp2j\leq p-2.

Theorem 1.1.

Fix a prime pp and a nonnegative integer jp2j\leq p-2. For each saturated fusion system \mathcal{F} on a finite pp-group,

limiHj(,𝔽p)|𝒪(c)=0{\lim}^{i}H^{j}(-,\mathbb{F}_{p})|_{\mathcal{O}(\mathcal{F}^{c})}=0

for all i1i\geq 1.

Here c\mathcal{F}^{c} denotes the full subcategory of \mathcal{F} with objects the \mathcal{F}-centric subgroups, and 𝒪(c)\mathcal{O}(\mathcal{F}^{c}) denotes the corresponding orbit category. For example, the special case j=1j=1 of Theorem 1.1 states that the functor H1(,𝔽p):𝒪(c)op𝔽p-modH^{1}(-,\mathbb{F}_{p})\colon\mathcal{O}(\mathcal{F}^{c})^{\operatorname{op}}\to\mathbb{F}_{p}\text{-}\textsf{mod} is acyclic when pp is odd. This was a case that motivated our work for reasons we outline below. But first we give context for this result, referring to [BLO03, Section 2] and [AKO11, Section III.5.6] for further details.

One way to study the classifying space of a finite group GG at a prime pp (or of a compact Lie group, or of a saturated fusion system over a pp-group) is to recognize the space as glued from classifying spaces of collections of proper subgroups. Following Dwyer’s uniform approach to such homotopy decompositions in the 1990s [Dwy97], this comes in the form of a map

hocolim𝒞FBG\operatorname{hocolim}_{\mathcal{C}}F\to BG

inducing an isomorphism in mod pp cohomology, where 𝒞\mathcal{C} is a small category and F:𝒞𝖳𝗈𝗉F\colon\mathcal{C}\to\mathsf{Top} is some functor such that for each c𝒞c\in\mathcal{C}, the composite F(c)BGF(c)\to BG identifies F(c)F(c) with the classifying space of a subgroup of GG up to homotopy. The decomposition most relevant here is the subgroup decomposition for the centric collection [Dwy97]. In this case, 𝒞=𝒪pc(G)\mathcal{C}=\mathcal{O}_{p}^{c}(G) is the full subcategory of the orbit category of GG on the pp-centric subgroups, and F(P)=B~PF(P)=\widetilde{B}P is an appropriate lifting to 𝖳𝗈𝗉\mathsf{Top} of the classifying space functor B:𝒪pc(G)𝗁𝗈𝖳𝗈𝗉B\colon\mathcal{O}_{p}^{c}(G)\to\mathsf{hoTop} to the homotopy category. For example, one can use B~P=EG×GG/P\widetilde{B}P=EG\times_{G}G/P, the Borel construction applied to the orbit G/PG/P considered as a GG-space [BS08, Section 5.2].

On the other hand, if \mathcal{F} is a saturated fusion system on a pp-group SS, there may be no finite group with Sylow pp-subgroup SS realizing the fusion in \mathcal{F} (i.e., \mathcal{F} may be exotic). Consequently there is no longer an obvious lifting of the classifying space functor. As was noticed by Broto, Levi, and Oliver [BLO03, §2], the Dwyer-Kan obstructions to such a lifting [DK92] are the same as the obstructions to the existence and uniqueness of a centric linking system \mathcal{L} associated with \mathcal{F}. By a theorem of Chermak, these obstructions vanish [Che13], and the pp-completion ||p|\mathcal{L}|_{p}^{\wedge} is then regarded as a classifying space “BB\mathcal{F}” for the fusion system. The subgroup decomposition in this context takes the form of a mod pp cohomology isomorphism hocolim𝒪(c)B~||\operatorname{hocolim}_{\mathcal{O}(\mathcal{F}^{c})}\widetilde{B}\to|\mathcal{L}|, where B~\widetilde{B} is the left homotopy Kan extension along the quotient functor π~:𝒪(c)\tilde{\pi}\colon\mathcal{L}\to\mathcal{O}(\mathcal{F}^{c}) of the constant functor \mathcal{L}\to* [AKO11, Proposition III.5.29].

The most immediate application of the subgroup decomposition for a saturated fusion system is to the computation of cohomology of the linking system. The mod pp cohomology H(||,𝔽p)H^{*}(|\mathcal{L}|,\mathbb{F}_{p}) of the linking system is the abutment of the Bousfield-Kan spectral sequence for the homotopy colimit with E2E_{2} page given by E2i,j=lim𝒪(c)iHj(,𝔽p)E_{2}^{i,j}={\lim}^{i}_{\mathcal{O}(\mathcal{F}^{c})}H^{j}(-,\mathbb{F}_{p}). Note that by a general property of pp-local functors on orbit categories, this page is bounded to the right [BLO03, Corollary 3.4]. It is an open problem whether the functor Hj(,𝔽p)H^{j}(-,\mathbb{F}_{p}) is acyclic, i.e., whether lim𝒪(c)iHj(,𝔽p)=0{\lim}^{i}_{\mathcal{O}(\mathcal{F}^{c})}H^{j}(-,\mathbb{F}_{p})=0 for all i1i\geq 1. For example, see [AO16, Problem 7.12] and [DP15, Conjecture]. Dwyer called a homology decomposition with this property “sharp”. Sharpness would immediately yield the most natural application of the subgroup decomposition, which is that the cohomology of the linking system satisfies a Cartan-Eilenberg stable elements formula:

Hj(||,𝔽p)limP𝒪(c)Hj(P,𝔽p).H^{j}(|\mathcal{L}|,\mathbb{F}_{p})\cong\lim_{P\in\mathcal{O}(\mathcal{F}^{c})}H^{j}(P,\mathbb{F}_{p}).

The stable elements formula was already shown by Broto, Levi, and Oliver [BLO03, Theorem 5.8] (along with other important consequences for combinatorial descriptions mapping spaces), but the proof relies on difficult results from homotopy theory. It would be very nice to see the stable elements formula directly via sharpness of the subgroup decomposition.

Sharpness over the centric pp-orbit category of a finite group GG was shown by Dwyer [Dwy98, §10], and over the centric orbit category of S(G)\mathcal{F}_{S}(G) by Díaz and Park [DP15, Theorem B]. Sharpness has since been established for certain families of exotic fusion systems, notably: those on pp-groups with an abelian subgroup of index pp [DP15], the smallest Benson-Solomon fusion system [HLL23], all fusion systems of characteristic pp-type/local characteristic pp (in the sense of the CFSG) [HLL23], and for the 27 exotic fusion systems on the Sylow pp-subgroup of G2(p)G_{2}(p) [GM22]. The sharpness problem has been studied from a very general point of view also by Yalçın [Yal22], who showed that sharpness for the subgroup and the normalizer decompositions are equivalent.

We follow Díaz and Park [DP15] by regarding cohomology as a Mackey functor for the fusion system, and we study the simple Mackey functors ST,VS_{T,V} occurring as composition factors of HjH^{j}. Theorem 1.1 is ultimately deduced from the following stronger result. We write k=𝔽pk=\mathbb{F}_{p} for short.

Theorem 1.2.

Fix a saturated fusion system \mathcal{F} on a pp-group SS, a subgroup TST\leq S, and a simple kOut(T)k\operatorname{Out}_{\mathcal{F}}(T)-module VV. If the simple Mackey functor ST,VS_{T,V} is a composition factor of Hj(,k)H^{j}(-,k) and jp2j\leq p-2, then the restriction ST,V|𝒪(c){S_{T,V}}^{*}|_{\mathcal{O}(\mathcal{F}^{c})} of the contravariant part of ST,VS_{T,V} has vanishing higher derived limits.

Theorem 1.2 is amenable to the standard technique of “pruning”, where one filters ST,V{S_{T,V}}^{*} by subquotient functors (not themselves contravariant parts of Mackey functors) that take the value 0 except on a single \mathcal{F}-conjugacy class of subgroups of SS. Thus, our proof of Theorem 1.1 goes by filtering HjH^{j} first as a Mackey functor completely, and then second as a coefficient system.

One reason for our looking at this problem grew out of a loose analogy with the paper [GL16], where we studied higher limits of the center functor 𝒵:𝒪(c)op(p)\mathcal{Z}_{\mathcal{F}}\colon\mathcal{O}(\mathcal{F}^{c})^{\operatorname{op}}\to\mathbb{Z}_{(p)}-mod, PZ(P)P\mapsto Z(P) in the context of Oliver’s proof [Oli13] of Chermak’s Theorem [Che13] on centric linking systems. That proof proceeds by a reduction to the case where \mathcal{F} is realizable by a finite pp-constrained group Γ\Gamma with normal centric pp-subgroup QQ. The observation of [GL16] was the relevance of finding a pp-local subgroup HH of Γ\Gamma that controls fixed points on Z(Q)Z(Q), i.e. that satisfies CZ(Q)(H)=CZ(Q)(Γ)C_{Z(Q)}(H)=C_{Z(Q)}(\Gamma), a problem that had been studied by the first author under the guise of “control of weak closure of elements”. This motivated us to look at whether techniques for “controlling transfer” in finite groups could be useful in studying the higher limits of the functor H1(,×)H^{1}(-,\mathbb{C}^{\times}) and its subfunctor H1(,𝔽p)H^{1}(-,\mathbb{F}_{p}). The issue is that it appears difficult to get reductions similar to those for the center functor in order for these techniques to be applicable. Also, general techniques for controlling transfer are known only when p5p\geq 5, so in the end what Theorem 1.1 gives in the case j=1j=1 is stronger than what those methods seemingly would have yielded even if they had been applicable.

We would like to thank Antonio Díaz for corrections and helpful suggestions on a previous version of this article.

2. Background results

2.1. Nilpotent action on group cohomology

Let kk be a commutative ring with identity. If GG is a finite group, HH is a subgroup of GG, and MM is a kGkG-module, then we use the usual notation trHG:MHMG\mathrm{tr}_{H}^{G}\colon M^{H}\to M^{G} for fixed points and the relative trace map. If pp is a rational prime which is zero in kk, then the relative trace is zero in cases where there is an element of GG outside HH but normalizing HH and acting with small nilpotence degree on MM. We state this when k=𝔽pk=\mathbb{F}_{p}, the only case we need.

Lemma 2.1.

Let GG be a finite group, pp a prime, and VV an 𝔽p[G]\mathbb{F}_{p}[G]-module. Suppose gg is an element of GG of pp-power order such that (g1)p1V=0(g-1)^{p-1}V=0. Then trHG(V)=0\mathrm{tr}_{H}^{G}(V)=0 for every subgroup HH of GG with gNG(H)Hg\in N_{G}(H)-H.

Proof.

Decompose trHG=trHgGtrHgpHgtrHHgp\mathrm{tr}_{H}^{G}=\mathrm{tr}_{H\langle g\rangle}^{G}\mathrm{tr}_{H\langle g^{p}\rangle}^{H\langle g\rangle}\mathrm{tr}_{H}^{H\langle g^{p}\rangle}. For each vtrHHgp(CV(H))v\in\mathrm{tr}_{H}^{H\langle g^{p}\rangle}(C_{V}(H)), we have

trHgpHg(v)=(1+g++gp1)v=(g1)p1v=0.\mathrm{tr}_{H\langle g^{p}\rangle}^{H\langle g\rangle}(v)=(1+g+\cdots+g^{p-1})v=(g-1)^{p-1}v=0.

In this paper, if XX and YY are two subsets of a group GG, we write [X,Y][X,Y] for the subgroup of GG generated by the set of commutators [x,y]=xyx1y1[x,y]=xyx^{-1}y^{-1} with xXx\in X and yYy\in Y. Our iterated commutators are right-associated: set [X,Y;1]=[X,Y][X,Y;1]=[X,Y], and inductively [X,Y;i]=[X,[X,Y;i1]][X,Y;i]=[X,[X,Y;i-1]] for i2i\geq 2. If GG has a left action on some module VV, the notation [X,V;i][X,V;i] should be interpreted in the semidirect product of VV by GG, in which case [X,V;i][X,V;i] is a subspace of VV, and [x,v;i]=(x1)iv[x,v;i]=(x-1)^{i}v for all xGx\in G and vVv\in V.

The techniques we have generally take advantage of situations in a finite pp-group in which some subgroup GG normalizes another subgroup PP and acts with small nilpotence degree on it, usually action which is quadratic (or trivial): [G,G,P]=1[G,G,P]=1. Then the following lemma of Miyamoto [Miy81, Lemma 2] provides a bound on the nilpotence degree of the action of GG on Hj(P,A)H^{j}(P,A) that is linear in jj when AA is finite abelian.

Lemma 2.2.

Let PP be a finite pp-group, AA a finite abelian group with trivial PP-action, and GG a finite group acting on PP and AA. Assume hh and nn are nonnegative integers such that (g1)h(g-1)^{h} acts as zero on each GG-composition factor of PP, and such that (g1)n(g-1)^{n} acts as zero on each GG-composition factor of AA. Then for all j0j\geq 0, (g1)(h1)j+n(g-1)^{(h-1)j+n} acts as zero on each GG-composition factor of Hj(P,A)H^{j}(P,A).

2.2. Mackey functors for fusion systems

The notation we use for fusion systems follows [AKO11]. We apply morphisms from right to left. Let \mathcal{F} be a saturated fusion system on a finite pp-group SS. A subgroup QQ of SS is fully \mathcal{F}-normalized (respectively, fully \mathcal{F}-centralized) if |NS(Q)||NS(Q)||N_{S}(Q)|\geq|N_{S}(Q^{\prime})| (respectively, |CS(Q)||CS(Q)||C_{S}(Q)|\geq|C_{S}(Q^{\prime})|) for each conjugate QQ^{\prime} of QQ in \mathcal{F}, i.e., for each subgroup of the form φ(Q)\varphi(Q) with φHom(Q,S)\varphi\in\operatorname{Hom}_{\mathcal{F}}(Q,S). A subgroup QQ of SS is \mathcal{F}-centric if CS(Q)QC_{S}(Q^{\prime})\leq Q^{\prime}, i.e. CS(Q)=Z(Q)C_{S}(Q^{\prime})=Z(Q^{\prime}), for each \mathcal{F}-conjugate QQ^{\prime} of QQ. The symbol c\mathcal{F}^{c} denotes the set of \mathcal{F}-centric subgroups, and also the full subcategory with the same objects. Note that a subgroup of SS is \mathcal{F}-centric if and only if it is fully centralized and contains its centralizer in SS [AKO11, Definition I.3.1]. By one of the axioms for saturation, a fully normalized subgroup PSP\leq S is also fully centralized and fully automized: AutS(P)\operatorname{Aut}_{S}(P) is a Sylow pp-subgroup of Aut(P)\operatorname{Aut}_{\mathcal{F}}(P) [AKO11, Proposition I.2.5].

For each pair of subgroups P,QSP,Q\leq S, Inn(Q)\operatorname{Inn}(Q) acts on Hom(P,Q)\operatorname{Hom}_{\mathcal{F}}(P,Q) by left composition. The orbit category 𝒪()\mathcal{O}(\mathcal{F}) of \mathcal{F} is the category with the same objects as \mathcal{F} and with morphism sets

Hom𝒪()(P,Q)=Inn(Q)\Hom(P,Q),\operatorname{Hom}_{\mathcal{O}(\mathcal{F})}(P,Q)=\operatorname{Inn}(Q)\backslash\operatorname{Hom}_{\mathcal{F}}(P,Q),

the orbits under this action. If 𝒳\mathcal{X} is a collection of subgroups of SS which is closed under \mathcal{F}-conjugacy and also closed under passing to overgroups in SS, then we abuse notation by using 𝒳\mathcal{X} also for the full subcategory of \mathcal{F} with object set 𝒳\mathcal{X}, and we write 𝒪(𝒳)\mathcal{O}(\mathcal{X}) for the corresponding orbit category. Other than the full orbit category itself, we will only need to work with the centric orbit category, the case 𝒳=c\mathcal{X}=\mathcal{F}^{c}.

Since the morphisms in a fusion system model conjugation of pp-subgroups in a finite group, it is natural that there is a notion of Mackey functor for fusion systems. We first want to recall from [DP15, Section 2] the definition of a Mackey functor in this setting in the form that is most useful later. Let kk be a commutative ring with identity. Let M=(M,M)M=(M^{*},M_{*}) be a pair of functors from 𝒪()\mathcal{O}(\mathcal{F}) to k-modk\text{-}\textsf{mod} with MM^{*} contravariant and MM_{*} covariant. Set rPQ=M([ιPQ])\mathrm{r}_{P}^{Q}=M^{*}([\iota_{P}^{Q}]), tPQ=M([ιPQ])\mathrm{t}_{P}^{Q}=M_{*}([\iota_{P}^{Q}]), and iso([φ])=M([φ])\mathrm{iso}([\varphi])=M_{*}([\varphi]) for each PQSP\leq Q\leq S and each isomorphism [φ]:Pφ(P)[\varphi]\colon P\to\varphi(P) in 𝒪()\mathcal{O}(\mathcal{F}). Then MM is a Mackey functor for \mathcal{F} if the following conditions hold [DP15, Definition 2.1, Proposition 2.2].

  1. (1)

    M(P)=defM(P)=M(P)M(P)\overset{\mathrm{def}}{=}M^{*}(P)=M_{*}(P) for each PSP\leq S,

  2. (2)

    (Isomorphism) M([φ])=defiso(φ)=M([φ]1)M_{*}([\varphi])\overset{\mathrm{def}}{=}\mathrm{iso}(\varphi)=M^{*}([\varphi]^{-1}) for each isomorphism [φ][\varphi] in 𝒪()\mathcal{O}(\mathcal{F}), and

  3. (3)

    (Mackey formula) for each P,QRSP,Q\leq R\leq S,

    rQRtPR=x[Q\R/P]tQPxQrQPxPxiso(cx|P).r_{Q}^{R}\circ t_{P}^{R}=\sum_{x\in[Q\backslash R/P]}t_{Q\cap{}^{x}P}^{Q}\circ r_{Q\cap{}^{x}P}^{{}^{x}P}\circ\mathrm{iso}(c_{x}|_{P}).

A morphism MNM\to N of Mackey functors is a family of kk-module homomorphisms ηP:M(P)N(P)\eta_{P}\colon M(P)\to N(P) such that η=(ηP)\eta=(\eta_{P}) is both a natural transformation from MNM^{*}\to N^{*} and a natural transformation MNM_{*}\to N_{*} simultaneously. A subfunctor of MM is a subfunctor of MM^{*} which is simultaneously a subfunctor of MM_{*}, and quotient functors are defined objectwise.

2.3. Simple Mackey functors

The simple objects in Mackk()\operatorname{Mack}_{k}(\mathcal{F}) are parametrized by pairs (T,V)(T,V), where TT is a subgroup of SS taken up to \mathcal{F}-conjugacy, and where VV is a simple (irreducible) kOut(T)k\operatorname{Out}_{\mathcal{F}}(T)-module taken up to isomorphism [DP15, Section 3]. When convenient we view VV as a kAut(T)k\operatorname{Aut}_{\mathcal{F}}(T)-module via inflation. The corresponding simple Mackey functor ST,VS_{T,V} has the property that ST,V(T)=VS_{T,V}(T)=V and ST,V(Q)=0S_{T,V}(Q)=0 for all subgroups QQ such that TT is not \mathcal{F}-conjugate to a subgroup of QQ.

Let TST\leq S and let VV be a simple kOut(T)k\operatorname{Out}_{\mathcal{F}}(T)-module. We give the description of the functor ST,VS_{T,V} on objects and isomorphisms in 𝒪()\mathcal{O}(\mathcal{F}) from p.153 of [DP15], since this will be important for the proof of Theorem 1.1, but interestingly the effect of ST,VS_{T,V} on nonisomorphisms is not so important for our argument. For that we refer the interested reader to the description in [DP15]. Our treatment is a little different from (but equivalent to) that in [DP15], since we need to pay somewhat closer attention to precisely how ST,V(Q)S_{T,V}(Q) decomposes as a direct sum of kOut(Q)k\operatorname{Out}_{\mathcal{F}}(Q)-modules.

The set Hom(T,Q)\operatorname{Hom}_{\mathcal{F}}(T,Q) is an Aut(Q)\operatorname{Aut}_{\mathcal{F}}(Q)-Aut(T)\operatorname{Aut}_{\mathcal{F}}(T) biset with action on either side given by composition. The orbits Hom(T,Q)/Aut(T)\operatorname{Hom}_{\mathcal{F}}(T,Q)/\operatorname{Aut}_{\mathcal{F}}(T) are in correspondence with the set of subgroups of QQ that are \mathcal{F}-conjugate to TT. Likewise the double orbits Aut(Q)\Hom(T,Q)/Aut(T)\operatorname{Aut}_{\mathcal{F}}(Q)\backslash\operatorname{Hom}_{\mathcal{F}}(T,Q)/\operatorname{Aut}_{\mathcal{F}}(T) are in correspondence with the Aut(Q)\operatorname{Aut}_{\mathcal{F}}(Q)-orbits of such subgroups. Let

AT,Q=[Aut(Q)\Hom(T,Q)/Aut(T)]A_{T,Q}=[\operatorname{Aut}_{\mathcal{F}}(Q)\backslash\operatorname{Hom}_{\mathcal{F}}(T,Q)/\operatorname{Aut}_{\mathcal{F}}(T)]

be a set of representatives for the double orbits.

For each αHom(T,Q)\alpha\in\operatorname{Hom}_{\mathcal{F}}(T,Q), we temporarily set U=α(T)U=\alpha(T) and denote (formally) by αV\alpha\otimes V the kAut(U)k\operatorname{Aut}_{\mathcal{F}}(U)-module, isomorphic to VV as a kk-module via αvv\alpha\otimes v\mapsto v, with action

φ(αv)=αα1φαv.\varphi\cdot(\alpha\otimes v)=\alpha\otimes\alpha^{-1}\varphi\alpha v.

for each φAut(U)\varphi\in\operatorname{Aut}_{\mathcal{F}}(U) and vVv\in V. Since TT acts trivially on VV, UU acts trivially on αV\alpha\otimes V. Set

(2.3) Wα=trUNQ(U)(αV),W_{\alpha}=\mathrm{tr}_{U}^{N_{Q}(U)}(\alpha\otimes V),

the image of the relative trace, where here NQ(U)N_{Q}(U) acts on the Out(U)\operatorname{Out}_{\mathcal{F}}(U)-module αV\alpha\otimes V through the composite

NQ(U)NInn(Q)(U)AutNQ(U)(U)Aut(U)N_{Q}(U)\twoheadrightarrow N_{\operatorname{Inn}(Q)}(U)\twoheadrightarrow\operatorname{Aut}_{N_{Q}(U)}(U)\to\operatorname{Aut}_{\mathcal{F}}(U)

with UCQ(U)UC_{Q}(U) acting trivially. Note WαW_{\alpha} is a kk-submodule of αV\alpha\otimes V. It has the structure of a kNAut(Q)(U)kN_{\operatorname{Aut}_{\mathcal{F}}(Q)}(U)-module on which NQ(U)N_{Q}(U) acts trivially.

Write UQU^{Q} for the QQ-conjugacy class of UU, and NAut(Q)(UQ)N_{\operatorname{Aut}_{\mathcal{F}}(Q)}(U^{Q}) for the stabilizer of this class in Aut(Q)\operatorname{Aut}_{\mathcal{F}}(Q). Since Inn(Q)\operatorname{Inn}(Q) is a normal subgroup of Aut(Q)\operatorname{Aut}_{\mathcal{F}}(Q) that acts transitively on UQU^{Q},

NAut(Q)(UQ)=NAut(Q)(U)Inn(Q).N_{\operatorname{Aut}_{\mathcal{F}}(Q)}(U^{Q})=N_{\operatorname{Aut}_{\mathcal{F}}(Q)}(U)\operatorname{Inn}(Q).

By construction, the subgroup NAut(Q)(U)Inn(Q)=NInn(Q)(U)N_{\operatorname{Aut}_{\mathcal{F}}(Q)}(U)\cap\operatorname{Inn}(Q)=N_{\operatorname{Inn}(Q)}(U) acts trivially on WαW_{\alpha}, and we may regard WαW_{\alpha} as a module for NAut(Q)(UQ)N_{\operatorname{Aut}_{\mathcal{F}}(Q)}(U^{Q}) via the composite

NAut(Q)(UQ)NAut(Q)(UQ)/Inn(Q)NAut(Q)(U)/NInn(Q)(U).N_{\operatorname{Aut}_{\mathcal{F}}(Q)}(U^{Q})\twoheadrightarrow N_{\operatorname{Aut}_{\mathcal{F}}(Q)}(U^{Q})/\operatorname{Inn}(Q)\cong N_{\operatorname{Aut}_{\mathcal{F}}(Q)}(U)/N_{\operatorname{Inn}(Q)}(U).

Set now

(2.4) ST,V(Q)α=WαNAut(Q)(UQ)Aut(Q)=φWφα,S_{T,V}(Q)_{\alpha}=W_{\alpha}\!\uparrow_{N_{\operatorname{Aut}_{\mathcal{F}}(Q)}(U^{Q})}^{\operatorname{Aut}_{\mathcal{F}}(Q)}=\bigoplus_{\varphi}W_{\varphi\alpha},

where φ\varphi runs over a set of representatives for the left cosets of NAut(Q)(UQ)N_{\operatorname{Aut}_{\mathcal{F}}(Q)}(U^{Q}) in Aut(Q)\operatorname{Aut}_{\mathcal{F}}(Q). Then ST,V(Q)αS_{T,V}(Q)_{\alpha} is a kOut(Q)k\operatorname{Out}_{\mathcal{F}}(Q)-module, that is, QQ still acts trivially. The value of ST,VS_{T,V} on the subgroup QQ is then

(2.5) ST,V(Q)=αAT,QST,V(Q)α,S_{T,V}(Q)=\bigoplus_{\alpha\in A_{T,Q}}S_{T,V}(Q)_{\alpha},

an Out(Q)\operatorname{Out}_{\mathcal{F}}(Q)-invariant direct sum decomposition.

Now let QQ^{\prime} be another subgroup of SS and β:QQ\beta\colon Q\to Q^{\prime} an isomorphism in \mathcal{F}. The bijection Hom(T,Q)βHom(T,Q)\operatorname{Hom}(T,Q)\xrightarrow{\beta\circ-}\operatorname{Hom}(T,Q^{\prime}) determines the kk-module isomorphism αVβαV\alpha\otimes V\cong\beta\alpha\otimes V intertwining the actions of NAut(Q)(α(T))N_{\operatorname{Aut}_{\mathcal{F}}(Q)}(\alpha(T)) and NAut(Q)(β(α(T)))N_{\operatorname{Aut}_{\mathcal{F}}(Q^{\prime})}(\beta(\alpha(T))) with respect to conjugation by β\beta. It induces an isomorphism of kk-modules

(2.6) WαWβαW_{\alpha}\cong W_{\beta\circ\alpha}

intertwining the actions of NAut(Q)(α(T)Q)N_{\operatorname{Aut}_{\mathcal{F}}(Q)}(\alpha(T)^{Q}) and NAut(Q)(β(α(T))Q)N_{\operatorname{Aut}_{\mathcal{F}}(Q^{\prime})}(\beta(\alpha(T))^{Q^{\prime}}). The component of iso(β)\mathrm{iso}(\beta) at α\alpha is the corresponding map of induced modules

iso(β)α:ST,V(Q)αST,V(Q)βα.\mathrm{iso}(\beta)_{\alpha}\colon S_{T,V}(Q)_{\alpha}\to S_{T,V}(Q^{\prime})_{\beta\alpha}.

and iso(β)\mathrm{iso}(\beta) is the sum of these maps.

Note that if VV is a simple 𝔽pOut(T)\mathbb{F}_{p}\operatorname{Out}_{\mathcal{F}}(T)-module, then ST,VS_{T,V} takes values in 𝔽p\mathbb{F}_{p}-mod. Further, since VV is simple, ST,VS_{T,V} is a simple functor [DP15].

2.4. Higher limits of functors on orbit categories

Let kk be a commutative (p)\mathbb{Z}_{(p)}-algebra and let MM be a contravariant functor from the orbit category of a group or a fusion system to the category of kk-modules. A common technique for computing the higher limits of MM (and especially for showing that such higher limits vanish) is to use a filtration of MM each of whose successive quotient functors is atomic, namely a functor which vanishes except on a single conjugacy class of subgroups. This method does not always work as, for example, in the case of the center functor [Oli18]. But the idea is often effective for making reductions even when it doesn’t work directly. And ultimately it is all that is needed for the proof of the main theorem here.

Let GG be a finite group and let AA be a (p)G\mathbb{Z}_{(p)}G-module. Define a functor

FA:𝒪p(G)op(p)-modF_{A}\colon\mathcal{O}_{p}(G)^{\operatorname{op}}\to\mathbb{Z}_{(p)}\text{-}\textsf{mod}

via FA(1)=AF_{A}(1)=A and FA(P)=0F_{A}(P)=0 when P1P\neq 1. The action of G=Aut𝒪p(G)(1)G=\operatorname{Aut}_{\mathcal{O}_{p}(G)}(1) on A=FA(1)A=F_{A}(1) is the given one. The higher limits of FAF_{A} are denoted Λi(G,A)\Lambda^{i}(G,A) and arise as the higher limits of atomic functors, as was first shown by Jackowski, McClure, and Oliver [AKO11, Proposition 5.20].

Proposition 2.7.

Let FF be any functor on the orbit category of a fusion system \mathcal{F} which vanishes except on the \mathcal{F}-conjugacy class of a subgroup PP. Then there is an isomorphism limFΛ(Out(P),F(P))\lim^{*}F\cong\Lambda^{*}(\operatorname{Out}_{\mathcal{F}}(P),F(P)).

Let P1,,PnP_{1},\dots,P_{n} be a set of representatives for the \mathcal{F}-conjugacy classes of subgroups of SS such that if i<ji<j, then PjP_{j} is not conjugate to a subgroup of PiP_{i}. Then one can make a filtration 0=M0M1Mn=M0=M_{0}\subset M_{1}\subseteq\cdots\subset M_{n}=M, in which MjM_{j} is the functor equal to MM on the union of the conjugacy classes PiP_{i} with iji\leq j, and 0 otherwise. Then (Mi/Mi1)(P)=M(P)(M_{i}/M_{i-1})(P)=M(P) if PP is conjugate to PiP_{i}, and it is zero otherwise, i.e. the quotient is atomic.

In general, we use the notation MQM_{Q} for the atomic subquotient functor of MM corresponding to the \mathcal{F}-conjugacy class of QQ, namely the functor with values MQ(P)=M(P)M_{Q}(P)=M(P) if PP is \mathcal{F}-conjugate to QQ, and MQ(P)=0M_{Q}(P)=0 otherwise.

The next lemma is proved using long exact sequences on higher limits corresponding to short exact sequences of functors arising out of a filtration of the above type.

Lemma 2.8 ([AKO11, Corollary 5.21(a)]).

Let MM be a contravariant functor on the centric orbit category of a fusion system \mathcal{F}. Assume that MQM_{Q} is acyclic for all QcQ\in\mathcal{F}^{c}, i.e., Λm(Out(Q),M(Q))=0\Lambda^{m}(\operatorname{Out}_{\mathcal{F}}(Q),M(Q))=0 for all m1m\geq 1. Then MM is acyclic.

The functors Λ(G,M)\Lambda^{*}(G,M) vanish in many cases. See for example Section III.5 of [AKO11] for many results along these lines. In the next lemma we state two such vanishing results.

Recall that a radical pp-chain of length mm in the finite group GG is a sequence Op(G)=P0<P1<PmO_{p}(G)=P_{0}<P_{1}<\cdots P_{m} such that Pi=Op(NG(P1,,Pi))P_{i}=O_{p}(N_{G}(P_{1},\dots,P_{i})) for each i=0,,ni=0,\dots,n, where here NG(P1,,Pi)N_{G}(P_{1},\dots,P_{i}) denotes the intersection of the normalizers in GG of the PiP_{i}.

Lemma 2.9.

Let GG be a finite group, MM a (p)G\mathbb{Z}_{(p)}G-module, and m1m\geq 1.

  1. (1)

    If Op(G)1O_{p}(G)\neq 1, then Λm(G,M)=0\Lambda^{m}(G,M)=0.

  2. (2)

    If tr1NG(P1,,Pm)(M)=0\mathrm{tr}_{1}^{N_{G}(P_{1},\dots,P_{m})}(M)=0 for each radical pp-chain 1=P0<P1<<Pm1=P_{0}<P_{1}<\cdots<P_{m} of length mm in GG, then Λm(G,M)=0\Lambda^{m}(G,M)=0.

Proof.

For (1), see [AKO11, Proposition III.5.24(b)]. Then (2) is a restatement of [AKO11, Proposition III.5.27], given (1). ∎

We want to show (Theorem 1.2) that the restriction of the contravariant part of each Mackey composition factor ST,VS_{T,V} of Hj(,𝔽p)H^{j}(-,\mathbb{F}_{p}) is acyclic when jp2j\leq p-2. For doing this, Díaz and Park show we can restrict attention to composition factors ST,VS_{T,V} with TT not \mathcal{F}-centric.

Lemma 2.10.

Let kk be a field of characteristic pp and \mathcal{F} a saturated fusion system on the finite pp-group SS. For each \mathcal{F}-centric subgroup TT and each simple kOut(T)k\operatorname{Out}_{\mathcal{F}}(T)-module VV, the restriction ST,V|𝒪(c){S_{T,V}}^{*}|_{\mathcal{O}(\mathcal{F}^{c})} of the contravariant part of ST,VS_{T,V} is acyclic.

Proof.

Proposition 3.3 of [DP15] implies that when TT is centric, ST,V|𝒪(c){S_{T,V}}^{*}|_{\mathcal{O}(\mathcal{F}^{c})} is an c\mathcal{F}^{c}-restricted Mackey functor for \mathcal{F} in the sense of Definition 2.1 of [DP15]. The lemma then follows from Theorem A there. ∎

3. Proof of Theorem 1.1

Throughout this section we fix a prime pp, and a saturated fusion system \mathcal{F} on the pp-group SS. We set k=𝔽pk=\mathbb{F}_{p} for short. For fixed j0j\geq 0, we consider Hj(,k)H^{j}(-,k) as a Mackey functor on 𝒪()\mathcal{O}(\mathcal{F}) where the contravariant structure is induced by restrictions and conjugations, and where the covariant structure is induced by transfers and conjugations (as usual).

We first fix some additional notation that we keep for the remainder of the section.

Let \mathcal{B} be the collection of all normal subgroups BB of SS such that

(3.1) CS(B)B and [B,B,S]=def[B,[B,S]]=1.\displaystyle C_{S}(B)\leq B\quad\text{ and }\quad[B,B,S]\overset{\mathrm{def}}{=}[B,[B,S]]=1.

By [Gor80, 5.3.12], each subgroup BSB\leq S maximal subject to being normal and abelian coincides with its centralizer in SS. Since [B,S]B[B,S]\leq B, we have [B,B,S][B,B]=1[B,B,S]\leq[B,B]=1. Thus, \mathcal{B} is nonempty. Further, since each member of \mathcal{B} is normal in SS, it is fully normalized, hence fully centralized by one of the saturation axioms for \mathcal{F}. This implies c\mathcal{B}\subseteq\mathcal{F}^{c}.

Definition 3.2.

Let TT be a subgroup of SS. Define 𝒬\mathcal{Q} to be the set of pairs (Q,α)(Q,\alpha) consisting of a centric subgroup QcQ\in\mathcal{F}^{c} and a morphism αHom(T,Q)\alpha\in\operatorname{Hom}_{\mathcal{F}}(T,Q) having the property that there are BB\in\mathcal{B} and an isomorphism β:QQ\beta\colon Q\to Q^{\prime} in \mathcal{F} such that Bβα(T)<Bβ(Q)B\cap\beta\alpha(T)<B\cap\beta(Q).

The proof of Theorem 1.1 is broken into two propositions. In the first one, we show that the kOut(Q)k\operatorname{Out}_{\mathcal{F}}(Q)-submodule ST,V(Q)αS_{T,V}(Q)_{\alpha} of ST,V(Q)S_{T,V}(Q) (see equation (2.4)) is 0 whenever ST,VS_{T,V} is a composition factor of Hj(,𝔽p)H^{j}(-,\mathbb{F}_{p}), (Q,α)𝒬(Q,\alpha)\in\mathcal{Q}, and jp2j\leq p-2. In the second one, we use this to show that the atomic subquotient (ST,V)Q({S_{T,V}}^{*})_{Q} is acyclic for an arbitrary \mathcal{F}-centric subgroup QQ when TT is not \mathcal{F}-centric.

Proposition 3.3.

Fix a prime pp, a saturated fusion system \mathcal{F} on a finite pp-group SS, a nonnegative integer jj, a subgroup TST\leq S, and a simple kOut(T)k\operatorname{Out}_{\mathcal{F}}(T)-module VV. If jp2j\leq p-2 and ST,VS_{T,V} is a composition factor of Hj(,k)H^{j}(-,k), then ST,V(Q)α=0S_{T,V}(Q)_{\alpha}=0 for all (Q,α)𝒬(Q,\alpha)\in\mathcal{Q}.

Proof.

Set U=α(T)U=\alpha(T) and Wα=trUNQ(U)(αV)W_{\alpha}=\mathrm{tr}_{U}^{N_{Q}(U)}(\alpha\otimes V). By (2.4), ST,V(Q)αS_{T,V}(Q)_{\alpha} takes the form

ST,V(Q)α=WαNAut(Q)(UQ)Aut(Q).S_{T,V}(Q)_{\alpha}=W_{\alpha}\!\uparrow_{N_{\operatorname{Aut}_{\mathcal{F}}(Q)}(U^{Q})}^{\operatorname{Aut}_{\mathcal{F}}(Q)}.

That is ST,V(Q)αS_{T,V}(Q)_{\alpha} is induced from WαW_{\alpha}.

Let (Q,α)𝒬(Q,\alpha)\in\mathcal{Q}. By Definition 3.2 there is an \mathcal{F}-isomorphism β:QQ\beta\colon Q\to Q^{\prime} and BB\in\mathcal{B} such that for U=β(U)U^{\prime}=\beta(U), we have BU<BQB\cap U^{\prime}<B\cap Q^{\prime}. The map β\beta induces an intertwining WαWβαW_{\alpha}\cong W_{\beta\circ\alpha}. Because of this we may as well change to lighter notation by replacing QQ by QQ^{\prime}, UU by UU^{\prime}, and α\alpha by βα\beta\circ\alpha. Thus, BU<BQB\cap U<B\cap Q and we want to show Wα=0W_{\alpha}=0.

Set BQ=BQB_{Q}=B\cap Q for short. Since BB is normal in SS, BQB_{Q} is normal in QQ and NBQ(U)N_{B_{Q}}(U) is normal in NQ(U)N_{Q}(U). Also, as BU<BQB\cap U<B_{Q}, we have U<UBQU<UB_{Q}. Hence U<NUBQ(U)=UNBQ(U)U<N_{UB_{Q}}(U)=UN_{B_{Q}}(U). Since ST,VS_{T,V} is a composition factor of Hj(,k)H^{j}(-,k), VV is a kOut(T)k\operatorname{Out}_{\mathcal{F}}(T)-composition factor of Hj(T,k)H^{j}(T,k), and αV\alpha\otimes V is an Out(U)\operatorname{Out}_{\mathcal{F}}(U)-composition factor of Hj(U,k)H^{j}(U,k). We have [NBQ(U),NBQ(U),U][B,B,S]=1[N_{B_{Q}}(U),N_{B_{Q}}(U),U]\leq[B,B,S]=1; in particular NBQ(U)N_{B_{Q}}(U) acts quadratically on every Aut(U)\operatorname{Aut}_{\mathcal{F}}(U)-composition factor of UU. The hypotheses of Lemma 2.2 thus hold with h=2h=2 and n=1n=1, and with UU, kk, and Aut(U)\operatorname{Aut}_{\mathcal{F}}(U) in the roles of PP, AA, and GG. As αV\alpha\otimes V is a composition factor of Hj(U,k)H^{j}(U,k), by that lemma we have (b1)j+1(αV)=0(b-1)^{j+1}(\alpha\otimes V)=0 for all bNBQ(U)b\in N_{B_{Q}}(U). Since composition factors are always 𝔽p\mathbb{F}_{p}-vector spaces and j+1<pj+1<p, Lemma 2.1 implies trUUNBQ(U)(αV)=0\mathrm{tr}_{U}^{UN_{B_{Q}}(U)}(\alpha\otimes V)=0. Hence

Wα=trUNQ(U)(αV)=trUNBQ(U)NQ(U)(trUUNBQ(U)(αV))=0,W_{\alpha}=\mathrm{tr}_{U}^{N_{Q}(U)}(\alpha\otimes V)=\mathrm{tr}_{UN_{B_{Q}}(U)}^{N_{Q}(U)}(\mathrm{tr}_{U}^{UN_{B_{Q}}(U)}(\alpha\otimes V))=0,

and this completes the proof. ∎

Proposition 3.4.

Fix a prime pp, a saturated fusion system \mathcal{F} on a finite pp-group SS, and a nonnegative integer jj. Let ST,VS_{T,V} be a composition factor of Hj(,k)H^{j}(-,k) such that TT is not \mathcal{F}-centric. If jp2j\leq p-2, then for all QcQ\in\mathcal{F}^{c}, the atomic subquotient functor (ST,V)Q({S_{T,V}}^{*})_{Q} is acyclic.

Proof.

Fix QcQ\in\mathcal{F}^{c}. The functor (ST,V)Q({S_{T,V}}^{*})_{Q} does not depend on QQ, but only on the \mathcal{F}-conjugacy class of QQ. By Proposition 2.7, we have

lim𝒪(c)(ST,V)QΛ(Out(Q),ST,V(Q)),{\lim}^{*}_{\mathcal{O}(\mathcal{F}^{c})}({S_{T,V}}^{*})_{Q}\cong\Lambda^{*}(\operatorname{Out}_{\mathcal{F}}(Q^{\prime}),S_{T,V}(Q^{\prime})),

for any \mathcal{F}-conjugate QQ^{\prime} of QQ. So we may assume QQ to be fully normalized in \mathcal{F}. In particular, R:=OutS(Q)R:=\operatorname{Out}_{S}(Q) is a Sylow pp-subgroup of G:=Out(Q)G:=\operatorname{Out}_{\mathcal{F}}(Q). By Lemma 2.9(1), the proposition holds if Op(G)1O_{p}(G)\neq 1, so we are reduced to Op(G)=1O_{p}(G)=1.

Adopt the notation of Section 2.3. By (2.5) and additivity of the functors Λ(G,)\Lambda^{*}(G,-), we have

(3.5) Λ(G,ST,V(Q))=αAT,QΛ(G,ST,V(Q)α).\Lambda^{*}(G,S_{T,V}(Q))=\bigoplus_{\alpha\in A_{T,Q}}\Lambda^{*}(G,S_{T,V}(Q)_{\alpha}).

Fix arbitrary BB\in\mathcal{B} and αAT,Q\alpha\in A_{T,Q}. Set BQ=BQB_{Q}=B\cap Q and U=α(T)U=\alpha(T) for short, and let X=φAut(Q)φ(U)X=\bigcap_{\varphi\in\operatorname{Aut}_{\mathcal{F}}(Q)}\varphi(U).

Assume that ST,V(Q)αS_{T,V}(Q)_{\alpha} is nonzero. By Proposition 3.3 and (2.6),

Bφ(U)=BQB\cap\varphi(U)=B_{Q}

for every choice of φAut(Q)\varphi\in\operatorname{Aut}_{\mathcal{F}}(Q). In particular, BQXBUB_{Q}\leq X\leq B\cap U.

Since TT is not \mathcal{F}-centric, UU and BQB_{Q} are not \mathcal{F}-centric, and BQ<BB_{Q}<B. Thus Q<QBQ<QB, and so Q<NQB(Q)=QNB(Q)Q<N_{QB}(Q)=QN_{B}(Q). As BB is normal in SS, NB(Q)N_{B}(Q) is normal in NS(Q)N_{S}(Q). Let CC be a normal subgroup of NS(Q)N_{S}(Q) minimal subject to BQ<CNB(Q)B_{Q}<C\leq N_{B}(Q). Since CC normalizes QQ and BB is normal in SS, we have [C,Q]BQ[C,Q]\leq B_{Q}, and hence

[C,φ(U)][C,Q]BQCXCφ(U)[C,\varphi(U)]\leq[C,Q]\leq B_{Q}\leq C\cap X\leq C\cap\varphi(U)

for each φAut(Q)\varphi\in\operatorname{Aut}_{\mathcal{F}}(Q). That is,

(3.6) CNS(φ(U)) and φ(U)NS(C)C\leq N_{S}(\varphi(U))\quad\text{ and }\quad\varphi(U)\leq N_{S}(C)

for each φAut(Q)\varphi\in\operatorname{Aut}_{\mathcal{F}}(Q).

Fix a radical pp-chain 1=Q0<Q1<<Qm1=Q_{0}<Q_{1}<\cdots<Q_{m} of GG with m1m\geq 1, and let H=NG(Q1,,Qm)H=N_{G}(Q_{1},\dots,Q_{m}) be its normalizer. Conjugating in GG in order to take QmRQ_{m}\leq R, we will show that the hypotheses of Lemma 2.9(2) hold for the conjugate chain, and then conjugating back, it holds for the one just fixed. In this way we are reduced to QmRQ_{m}\leq R.

Let φAut(Q)\varphi\in\operatorname{Aut}_{\mathcal{F}}(Q) be arbitrary. Recall that CC was chosen so that C/BQC/B_{Q} is a minimal normal subgroup of NS(Q)/BQN_{S}(Q)/B_{Q} (contained in NB(Q)/BQN_{B}(Q)/B_{Q}. As a minimal normal subgroup of a finite pp-group, C :=QC/QC/BQ\overset{{\mskip 6.0mu\leaders\hrule height=0.4pt\hfill\mskip 3.0mu}}{C}\vphantom{C}:=QC/Q\cong C/B_{Q} is therefore of order pp and contained in the center of R=OutS(Q)=NS(Q)/QR=\operatorname{Out}_{S}(Q)=N_{S}(Q)/Q, the last equality because QcQ\in\mathcal{F}^{c}. Thus, C H\overset{{\mskip 6.0mu\leaders\hrule height=0.4pt\hfill\mskip 3.0mu}}{C}\vphantom{C}\leq H. Since CBC\leq B, we have [C,φ(U)]φ(U)B[C,\varphi(U)]\leq\varphi(U)\cap B by (3.6), and so [C,C,φ(U)][C,B]=1[C,C,\varphi(U)]\leq[C,B]=1 as BB is abelian. In particular, [C,C,V1]=0[C,C,V_{1}]=0 for every Aut(φ(U))\operatorname{Aut}_{\mathcal{F}}(\varphi(U))-composition factor V1V_{1} of φ(U)\varphi(U). As φαV\varphi\alpha\otimes V is an Aut(φα(U))\operatorname{Aut}_{\mathcal{F}}(\varphi\alpha(U))-composition factor of Hj(φα(T),k)H^{j}(\varphi\alpha(T),k), we have

(c1)j+1(φαV)=0(c-1)^{j+1}(\varphi\alpha\otimes V)=0 for all cCc\in C

by Lemma 2.2 applied with Aut(φ(U))\operatorname{Aut}_{\mathcal{F}}(\varphi(U)), φ(U)\varphi(U), kk, 11, and 22 in the roles of GG, PP, AA, nn and hh. So (c1)j+1Wφα=0(c-1)^{j+1}W_{\varphi\alpha}=0 for all cCc\in C as WφαW_{\varphi\alpha} is a kk-submodule of φαV\varphi\alpha\otimes V. Since this holds for all φAut(Q)\varphi\in\operatorname{Aut}_{\mathcal{F}}(Q), we have by the direct sum decomposition (2.4) that

(c1)j+1ST,V(Q)α=0 for all cC.\text{$(c-1)^{j+1}S_{T,V}(Q)_{\alpha}=0$ for all $c\in C$}.

Lemma 2.1 now applies with HH in the role of GG, with 11 in the role of HH, and with gg a generator of C \overset{{\mskip 6.0mu\leaders\hrule height=0.4pt\hfill\mskip 3.0mu}}{C}\vphantom{C}. As j+1<pj+1<p, we have tr1H(ST,V(Q)α)=0\mathrm{tr}_{1}^{H}(S_{T,V}(Q)_{\alpha})=0 by that lemma. Therefore, Lemma 2.9(2) and (3.5) combine to give lim𝒪(c)m(ST,V)QΛm(Out(Q),ST,V(Q))=0\lim^{m}_{\mathcal{O}(\mathcal{F}^{c})}({S_{T,V}}^{*})_{Q}\cong\Lambda^{m}(\operatorname{Out}_{\mathcal{F}}(Q),S_{T,V}(Q))=0. ∎

Proof of Theorem 1.2.

Let ST,VS_{T,V} be a composition factor of Hj(,k)H^{j}(-,k) as a Mackey functor on 𝒪()\mathcal{O}(\mathcal{F}), and suppose that jp2j\leq p-2. If TT is centric, then ST,V|𝒪(c){S_{T,V}}^{*}|_{\mathcal{O}(\mathcal{F}^{c})} is acyclic by Lemma 2.10. If TT is not \mathcal{F}-centric, then Proposition 3.4 shows that the atomic functor (ST,V)Q({S_{T,V}}^{*})_{Q} is acyclic for each QcQ\in\mathcal{F}^{c}, so again ST,V|𝒪(c){S_{T,V}}^{*}|_{\mathcal{O}(\mathcal{F}^{c})} is acyclic by Lemma 2.8. ∎

Proof of Theorem 1.1.

As made explicit in the proof of [DP15, Proposition 4.3], given a filtration of Hj(,𝔽p)H^{j}(-,\mathbb{F}_{p}) whose successive quotients are simple Mackey functors for \mathcal{F}, the restrictions to 𝒪(c)\mathcal{O}(\mathcal{F}^{c}) of the contravariant parts of the members of the filtration yield a filtration for Hj(,𝔽p)|𝒪(c)H^{j}(-,\mathbb{F}_{p})|_{\mathcal{O}(\mathcal{F}^{c})}. So the theorem follows from Theorem 1.2. ∎

References

  • [AKO11] Michael Aschbacher, Radha Kessar, and Bob Oliver, Fusion systems in algebra and topology, London Mathematical Society Lecture Note Series, vol. 391, Cambridge University Press, Cambridge, 2011. MR 2848834
  • [AO16] Michael Aschbacher and Bob Oliver, Fusion systems, Bull. Amer. Math. Soc. (N.S.) 53 (2016), no. 4, 555–615. MR 3544261
  • [BLO03] Carles Broto, Ran Levi, and Bob Oliver, The homotopy theory of fusion systems, J. Amer. Math. Soc. 16 (2003), no. 4, 779–856 (electronic).
  • [BS08] David J. Benson and Stephen D. Smith, Classifying spaces of sporadic groups, Mathematical Surveys and Monographs, vol. 147, American Mathematical Society, Providence, RI, 2008. MR 2378355 (2009f:55017)
  • [Che13] Andrew Chermak, Fusion systems and localities, Acta Math. 211 (2013), no. 1, 47–139. MR 3118305
  • [DK92] W. G. Dwyer and D. M. Kan, Centric maps and realization of diagrams in the homotopy category, Proc. Amer. Math. Soc. 114 (1992), no. 2, 575–584. MR 1070515
  • [DP15] Antonio Díaz and Sejong Park, Mackey functors and sharpness for fusion systems, Homology Homotopy Appl. 17 (2015), no. 1, 147–164. MR 3338545
  • [Dwy97] W. G. Dwyer, Homology decompositions for classifying spaces of finite groups, Topology 36 (1997), no. 4, 783–804. MR 1432421 (97m:55016)
  • [Dwy98] by same author, Sharp homology decompositions for classifying spaces of finite groups, Group representations: cohomology, group actions and topology (Seattle, WA, 1996), Proc. Sympos. Pure Math., vol. 63, Amer. Math. Soc., Providence, RI, 1998, pp. 197–220. MR 1603159
  • [GL16] George Glauberman and Justin Lynd, Control of fixed points and existence and uniqueness of centric linking systems, Invent. Math. 206 (2016), no. 2, 441–484. MR 3570297
  • [GM22] Valentina Grazian and Ettore Marmo, Sharpness of saturated fusion systems on a Sylow pp-subgroup of G2(p)\mathrm{G}_{2}(p), 2022.
  • [Gor80] Daniel Gorenstein, Finite groups, second ed., Chelsea Publishing Co., New York, 1980.
  • [HLL23] Ellen Henke, Assaf Libman, and Justin Lynd, Punctured groups for exotic fusion systems, Trans. London Math. Soc. 10 (2023), no. 1, 21–99. MR 4640379
  • [JM92] Stefan Jackowski and James McClure, Homotopy decomposition of classifying spaces via elementary abelian subgroups, Topology 31 (1992), no. 1, 113–132. MR 1153240 (92k:55026)
  • [Miy81] Masahiko Miyamoto, A pp-local control of cohomology group, J. Algebra 70 (1981), no. 2, 475–479. MR 623820
  • [Oli13] Bob Oliver, Existence and uniqueness of linking systems: Chermak’s proof via obstruction theory, Acta Math. 211 (2013), no. 1, 141–175. MR 3118306
  • [Oli18] by same author, A remark on the construction of centric linking systems, Geometric and Topological Aspects of the Representation Theory of Finite Groups (J. F. Carlson, S. B. Iyengar, and J. Pevtsova, eds.), Springer Proceedings in Mathematics & Statistics, vol. 242, Springer International Publishing, 2018, pp. 327–337.
  • [Yal22] Ergün Yalçın, Higher limits over the fusion orbit category, Adv. Math. 406 (2022), Paper No. 108482, 69. MR 4438061