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

Chain duality for categories over complexes

James F. Davis, Carmen Rovi
Abstract

We show that the additive category of chain complexes parametrized by a finite simplicial complex KK forms a category with chain duality. This fact, never fully proven in the original reference (Ranicki, 1992), is fundamental for Ranicki’s algebraic formulation of the surgery exact sequence of Sullivan and Wall, and his interpretation of the surgery obstruction map as the passage from local Poincaré duality to global Poincaré duality.

Our paper also gives a new, conceptual, and geometric treatment of chain duality on KK-based chain complexes.

Dedicated to Dennis Sullivan on the occasion of his 80th birthday.

Keywords: Chain duality, assembly, controlled surgery, LL-theory, dual cell decomposition, KK-dissection.

1 Introduction

Kervaire and Milnor [KM] developed and applied the new field of surgery to classify exotic smooth structures on spheres. Browder and Novikov independently extended and relativized the theory. Sullivan in his thesis [Sullivan-thesis] investigated the obstruction theory for deforming a homotopy equivalence to a homeomorphism. In seminar notes [Sullivan-Notes] written shortly after his thesis, Sullivan’s Theorem 3 packaged this in what is now called the surgery exact sequence. (We will be ahistorical and concentrate on topological manifolds; Kervaire-Milnor concentrated on smooth manifolds and Sullivan on PL-manifolds. The extension to topological manifolds is due to the deep work of Kirby and Siebenmann [KS].). It was extended to the nonsimply-connected case and to the case of compact manifolds by Wall [Wall]. The surgery exact sequence for a closed nn-dimensional manifold XX with n5n\geq 5 is

Ln+1([π1(X)])𝕊TOP(X)𝒩TOP(X)σLn([π1X]).\dots\to L_{n+1}({\mathbb{Z}}[\pi_{1}(X)])\to\mathbb{S}^{TOP}(X)\to{\cal N}^{TOP}(X)\xrightarrow{\sigma}L_{n}({\mathbb{Z}}[\pi_{1}X]).

The object one wants to compute is the structure set 𝕊TOP(X)\mathbb{S}^{TOP}(X), first defined by Sullivan. Representatives of the structure set are given by (simple) homotopy equivalences from a closed nn-manifold to XX. Computing the structure set is the key ingredient in computing the manifold moduli set, the set of homeomorphism types of nn-manifolds homotopy equivalent to XX. The beauty of the surgery exact sequence is marred by many flaws. One is that it is an exact sequence of pointed sets. Another flaw is standard with a long exact sequence, to do computations one needs to compute the surgery obstruction map σ\sigma, including its domain, the normal invariants 𝒩TOP(X){\cal N}^{TOP}(X), and its codomain, the LL-groups. For many fundamental groups (e.g. finite groups or finitely generated abelian groups) one can compute the LL-groups algebraically. Sullivan, in his thesis, analyzed the normal invariants, using transversality to establish a bijection 𝒩TOP(X)[X,G/TOP]{\cal N}^{TOP}(X)\cong[X,G/TOP]. He also computed the homotopy groups πi(G/TOP)\pi_{i}(G/TOP) which vanish for ii odd, have order 2 for i2(mod4)i\equiv 2\pmod{4}, and are infinite cyclic for i0(mod4)i\equiv 0\pmod{4} (this follows from the generalized Poincaré conjecture). Furthermore, Sullivan analyzed the homotopy type of G/TOPG/TOP and established Sullivan periodicity, Ω4(×G/TOP)×G/TOP\Omega^{4}({\mathbb{Z}}\times G/TOP)\simeq{\mathbb{Z}}\times G/TOP. This then determines an Ω\Omega-spectrum 𝕃.\mathbb{L}_{.} and its 1-connective cover 𝕃.1\mathbb{L}_{.}\langle 1\rangle. There is a formal identification [X,G/TOP]=H0(X;𝕃.1)[X,G/TOP]=H^{0}(X;\mathbb{L}_{.}\langle 1\rangle).

The surgery exact sequence is now becoming more presentable, but it is still marred by functoriality issues: the LL-groups are covariant in XX, the normal invariants are contravariant in XX, and the structure set has no obvious variance at all. Furthermore it is only defined for manifolds, and one would like an abelian group structure. These flaws make computing the surgery obstruction map difficult. Quinn’s [Quinn] vision (largely carried out by Ranicki [bluebook], see also [KMM]) is to find a bijection between the surgery exact sequence and a long exact sequences of abelian groups defined for every space XX and fully covariant in XX. In more detail, there is the following commutative diagram

Ln+1([π1X]){L_{n+1}({\mathbb{Z}}[\pi_{1}X])}𝕊TOP(X){\mathbb{S}^{TOP}(X)}𝒩TOP(X){{\cal N}^{TOP}(X)}Ln([π1X]){L_{n}({\mathbb{Z}}[\pi_{1}X])}Ln+1([π1X]){L_{n+1}({\mathbb{Z}}[\pi_{1}X])}𝕊1n+1(X){\mathbb{S}^{\langle 1\rangle}_{n+1}(X)}Hn(X;𝕃.1){H_{n}(X;\mathbb{L}_{.}\langle 1\rangle)}Ln([π1X]){L_{n}({\mathbb{Z}}[\pi_{1}X])}Ln+1([π1X]){L_{n+1}({\mathbb{Z}}[\pi_{1}X])}𝕊n+1(X){\mathbb{S}_{n+1}(X)}Hn(X;𝕃.){H_{n}(X;\mathbb{L}_{.})}Ln([π1X]){L_{n}({\mathbb{Z}}[\pi_{1}X])}=\scriptstyle{=}\scriptstyle{\cong}\scriptstyle{\cong}=\scriptstyle{=}=\scriptstyle{=}A1\scriptstyle{A\langle 1\rangle}=\scriptstyle{=}A\scriptstyle{A}

where the vertical maps labelled \cong are bijections when XX is a closed nn-manifold and the bottom two horizontal lines are exact sequences of abelian groups, defined for any space XX. These two lines are called the 1-connective algebraic surgery exact sequence and the algebraic surgery exact sequence, respectively. The maps A1A\langle 1\rangle and AA are called assembly maps; they are defined at the spectrum level. Hence there is a long exact sequence of homotopy groups, where the algebraic structure groups are defined be to the homotopy groups of the cofiber of the assembly maps. The map AA is conjectured to be an isomorphism when X=BπX=B\pi with π\pi torsionfree.

There are myriad ways of constructing the assembly maps (the construction in [DL98] seems best for computations). Different constructions are identified via axiomatics (see [WW], also [DL98]). Ranicki’s version of assembly, needed for his approach to the above diagram, was motivated by his earlier work with Weiss [RW] viewing the assembly maps as a passage from local to global Poincaré duality. Much earlier Ranicki [ats1] reinterpreted Wall’s algebraic LL-groups as bordism groups of algebraic Poincaré complexes over the group ring [π1X]{\mathbb{Z}}[\pi_{1}X]. This is the global Poincaré duality. The local Poincaré duality comes from making a geometric degree one normal map transverse to the dual cones of XX (see Section 3 for the definition). These degree one normal maps to the cones are then assembled to give the original degree one normal map.

More precisely, Ranicki [bluebook] defined the notion of an additive category with chain duality 𝔸{\mathbb{A}}, the associated algebraic bordism category Λ(𝔸)\Lambda({\mathbb{A}}) (see his Example 3.3), and the corresponding LL-groups Ln(𝔸)L_{n}({\mathbb{A}}) (see his Definition 1.8). In his notation, the assembly map is given by establishing a map of algebraic bordism categories (see his Proposition 9.11)

Λ((,X)mod))Λ([π1X]mod)\Lambda(({\mathbb{Z}},X)\operatorname{-mod}))\to\Lambda({\mathbb{Z}}[\pi_{1}X]\operatorname{-mod})

and defining the assembly map to be the induced map on LL-groups. However, one flaw in his argument is that he never provided a proof that (,X)mod({\mathbb{Z}},X)\operatorname{-mod} is an additive category with chain duality, despite his assertion in Proposition 5.1 of [bluebook]. Our modest contribution to this saga is to provide a self-contained, conceptual, and geometric proof that (,X)mod({\mathbb{Z}},X)\operatorname{-mod} is an additive category with chain duality.

We are not the first to provide a proof of this result – one is given in Section 5 of [Spiros-Tibor]. However, we found the proof and its notation rather dense. Another account of this result is given in a recent preprint of Frank Connolly [Frank]. Although his aims are quite similar to ours, the approach is different, the reader may wish to compare.

We now outline our paper. In Section 2 we review Ranicki’s notion of an additive category with chain duality, this is an additive category with a chain duality functor satisfying a chain homotopy equivalence condition. In Section 3 we fix a finite simplicial complex KK (e.g. a triangulation of a compact manifold), and we define Ranicki’s additive categories of KK-based chain complexes. Here we need to warn the reader that we have deviated from Ranicki’s notation in [bluebook], which we found difficult to use. A comparison between our notation and Ranicki’s is given in Remark LABEL:R_notation. The two key additive categories are Ch((K)mod)\operatorname{{Ch}}({\mathbb{Z}}(K)\operatorname{-mod}) and Ch((Kop)mod)\operatorname{{Ch}}({\mathbb{Z}}(K^{\operatorname{{op}}})\operatorname{-mod}). The latter category is the one whose LL-theory gives the normal invariants, so is perhaps more important. The simplicial chain complex ΔK\Delta K gives an object of Ch((K)mod)\operatorname{{Ch}}({\mathbb{Z}}(K)\operatorname{-mod}) and the simplicial cochain complex ΔK\Delta K^{-*} gives an object of Ch((Kop)mod)\operatorname{{Ch}}({\mathbb{Z}}(K^{\operatorname{{op}}})\operatorname{-mod}). More generally, given a CW-complex XX with a KK-dissection, the cellular chains C(X)C(X) give an object of Ch((K)mod)\operatorname{{Ch}}({\mathbb{Z}}(K)\operatorname{-mod}) and given a CW-complex XX with a KopK^{\operatorname{{op}}}-dissection, the cellular chains C(X)C(X) give an object of Ch((Kop)mod)\operatorname{{Ch}}({\mathbb{Z}}(K^{\operatorname{{op}}})\operatorname{-mod}). We related this to dual cell decompositions, defined even when KK is not a manifold. In Section LABEL:sec:cat_point_of_view, we develop homological algebra necessary for our proof that these categories admit a chain duality.

Section LABEL:sec:dual_cell may be of independent interest. For a finite simplicial complex KK, we define the dual cell decomposition DKDK which is a regular CW-complex refining the simplicial structure on KK. Corollary LABEL:varepsilon_is_a_weak_eq and Remark LABEL:two-sided say that this, in some sense, gives a two-sided bar resolution for the category of posets of KK.

Finally, in Section LABEL:K-based_chain_duality we define chain duality functors (T,τ)(T,\tau) on Ch((Kop)mod)\operatorname{{Ch}}({\mathbb{Z}}(K^{\operatorname{{op}}})\operatorname{-mod}) and Ch((K)mod)\operatorname{{Ch}}({\mathbb{Z}}(K)\operatorname{-mod}) and prove our main theorem.

Theorem 1.

The following are additive categories with chain duality

{(Ch((Kop)mod),T,τ),(Ch((K)mod),T,τ).\begin{cases}(\operatorname{{Ch}}({\mathbb{Z}}(K^{\operatorname{{op}}})\operatorname{-mod}),T,\tau),\\ (\operatorname{{Ch}}({\mathbb{Z}}(K)\operatorname{-mod}),T,\tau).\end{cases}

2 Chain Duality

For a category KK, write σK\sigma\in K when σ\sigma is an object of KK and K(σ,τ)K(\sigma,\tau) for the set of morphisms from σ\sigma to τ\tau. A preadditive category is a category where all morphism sets are abelian groups and composition is bilinear. An additive category is a preadditive category which admits finite products and coproducts. An example of an additive category is the category of finitely generated free abelian groups.

Let 𝔸{\mathbb{A}} be an additive category and let Ch(𝔸)\operatorname{{Ch}}({\mathbb{A}}) be the category of finite chain complexes over 𝔸{\mathbb{A}} where finite means that Cn=0C_{n}=0 for all but a finite number of nn. Homotopy notions make sense in this category: the notions of two chain maps being chain homotopic, a chain map being a chain homotopy equivalence, two chain complexes being chain homotopy equivalent, and a chain complex being contractible. The notion of homology of a chain complex over an additive category does not make sense.

Let Ch,(𝔸)\operatorname{{Ch}}_{\bullet,\bullet}({\mathbb{A}}) be the category of finite bigraded chain complexes over 𝔸{\mathbb{A}}. There are functors

Tot\displaystyle\operatorname{{Tot}} :Ch,(𝔸)Ch(𝔸)\displaystyle:\operatorname{{Ch}}_{\bullet,\bullet}({\mathbb{A}})\to\operatorname{{Ch}}({\mathbb{A}})
Hom,\displaystyle\operatorname{{Hom}}_{\bullet,\bullet} :Ch(𝔸)op×Ch(𝔸)Ch,(mod)\displaystyle:\operatorname{{Ch}}({\mathbb{A}})^{\operatorname{{op}}}\times\operatorname{{Ch}}({\mathbb{A}})\to\operatorname{{Ch}}_{\bullet,\bullet}({\mathbb{Z}}\operatorname{-mod})

where Tot(C,)n=p+q=nCp+q\operatorname{{Tot}}(C_{\bullet,\bullet})_{n}=\bigoplus_{p+q=n}C_{p+q} and Hom(C,D)p,q=𝔸(Cp,Dq)\operatorname{{Hom}}(C,D)_{p,q}={\mathbb{A}}(C_{-p},D_{q}). (Throughout this paper, if the differentials are standard or can be easily determined, we omit them for readability). If CC and DD are finite chain complexes over an additive category 𝔸{\mathbb{A}}, then

Hom𝔸(C,D):=Tot(Hom,(C,D))\operatorname{{Hom}}_{{\mathbb{A}}}(C,D):=\operatorname{{Tot}}(\operatorname{{Hom}}_{\bullet,\bullet}(C,D))

is a chain complex of abelian groups with differentials

dHom𝔸(C,D):Hom𝔸(C,D)n\displaystyle d_{\operatorname{{Hom}}_{{\mathbb{A}}}(C,D)}:\operatorname{{Hom}}_{{\mathbb{A}}}(C,D)_{n} Hom𝔸(C,D)n1\displaystyle\to\operatorname{{Hom}}_{{\mathbb{A}}}(C,D)_{n-1}
d(f)\displaystyle d(f) =dDf+(1)n+1fdC.\displaystyle=d_{D}\circ f+(-1)^{n+1}f\circ d_{C}.

A 0-cycle is a chain map; the difference of chain maps is a boundary if and only if the chain maps are chain homotopic. In particular, there is a monomorphism of abelian groups Ch(𝔸)(C,D)Hom𝔸(C,D)0\operatorname{{Ch}}({\mathbb{A}})(C,D)\to\operatorname{{Hom}}_{{\mathbb{A}}}(C,D)_{0}.

If CC and DD are chain complexes of abelian groups, then there is a chain complex (CD,d)(C\otimes D,d_{\otimes}) with differentials

d:(CD)n\displaystyle d_{\otimes}:(C\otimes D)_{n} (CD)n1\displaystyle\to(C\otimes D)_{n-1}
d(xy)\displaystyle d_{\otimes}(x\otimes y) =dC(x)y+(1)|x|xdD(y).\displaystyle=d_{C}(x)\otimes y+(-1)^{|x|}x\otimes d_{D}(y).
Definition 2.

A chain duality functor (T,τ)(T,\tau) on an additive category 𝔸{\mathbb{A}} is an additive functor T:Ch(𝔸)Ch(𝔸)opT:\operatorname{{Ch}}({\mathbb{A}})\to\operatorname{{Ch}}({\mathbb{A}})^{\operatorname{{op}}} together with a natural chain map

τC,D:Hom𝔸(TC,D)Hom𝔸(TD,C)\tau_{C,D}:\operatorname{{Hom}}_{{\mathbb{A}}}(TC,D)\to\operatorname{{Hom}}_{{\mathbb{A}}}(TD,C)

defined for each pair of chain complexes C,DCh(𝔸)C,D\in\operatorname{{Ch}}({\mathbb{A}}) so that τ2=Id\tau^{2}=\operatorname{Id} in the sense that τD,CτC,D=Id\tau_{D,C}\circ\tau_{C,D}=\operatorname{Id}.

Remark 3.

By restricting to 0-cycles, the natural chain map τ\tau induces a natural isomorphism of abelian groups

τC,D:Ch(𝔸)(TC,D)Ch(𝔸)(TD,C).\tau_{C,D}:\operatorname{{Ch}}({\mathbb{A}})(TC,D)\to\operatorname{{Ch}}({\mathbb{A}})(TD,C).
Lemma 4.

Let (T,τ)(T,\tau) be a chain duality functor on 𝔸{\mathbb{A}}. For CCh(𝔸)C\in\operatorname{{Ch}}({\mathbb{A}}), let eC:T2CCe_{C}:T^{2}C\to C be τ(IdTC)\tau(\operatorname{Id}_{TC}). This defines a natural transformation

e:T2Id:Ch(𝔸)Ch(𝔸)e:T^{2}\to\operatorname{Id}:\operatorname{{Ch}}({\mathbb{A}})\to\operatorname{{Ch}}({\mathbb{A}})

so that for each object CCh(𝔸)C\in\operatorname{{Ch}}({\mathbb{A}}),

eTCT(eC)=IdTC:TCT3CTC.e_{TC}\circ T(e_{C})=\operatorname{Id}_{TC}:TC\to T^{3}C\to TC.
Proof.

Suppose α:TUV\alpha:TU\to V and β:VW\beta:V\to W are chain maps. Then naturality of τ\tau implies that

τ(βα)=τ(α)T(β).\tau(\beta\circ\alpha)=\tau(\alpha)\circ T(\beta).

Thus

eTCT(eC)\displaystyle e_{TC}\circ T(e_{C}) =τ(IdT2C)T(eC)\displaystyle=\tau(\operatorname{Id}_{T^{2}C})\circ T(e_{C})
=τ(eCIdT2C)\displaystyle=\tau(e_{C}\circ\operatorname{Id}_{T^{2}C})
=τ(eC)\displaystyle=\tau(e_{C})
=IdTC.\displaystyle=\operatorname{Id}_{TC}.

It is also true, conversely, that an additive functor T:Ch(𝔸)Ch(𝔸)opT:\operatorname{{Ch}}({\mathbb{A}})\to\operatorname{{Ch}}({\mathbb{A}})^{\operatorname{{op}}} and natural transformation e:T2Ide:T^{2}\to\operatorname{Id} satisfying eTCT(eC)=IdTCe_{TC}\circ T(e_{C})=\operatorname{Id}_{TC} for all CCh(𝔸)C\in\operatorname{{Ch}}({\mathbb{A}}) determines a chain duality functor (T,τ)(T,\tau) where τ(f:TCD):=eCT(f)\tau(f:TC\to D):=e_{C}\circ T(f), but we omit the proof of this fact.

Definition 5.

A chain duality on an additive category 𝔸{\mathbb{A}} is a chain duality functor (T,τ)(T,\tau) so that eC:T2CCe_{C}:T^{2}C\to C is a chain homotopy equivalence for all CCh(𝔸)C\in\operatorname{{Ch}}({\mathbb{A}}).

This is equivalent to the definition in Andrew Ranicki’s book [bluebook, Definition 1.1]. Notice that a chain duality functor does not necessarily give a chain duality, because of the extra condition that eCe_{C} is a chain homotopy equivalence. We separately defined a chain duality functor because there can be uses for the weaker notion, for example, see the thesis of Christopher Palmer [P15].

3 KK-based chain complexes

Let KK be a finite set.

Definition 6.

An abelian group MM is KK-based if it is expressed as a direct sum

M=σKM(σ).M=\bigoplus_{\sigma\in K}M(\sigma).

A morphism f:MNf:M\to N of KK-based abelian groups is simply a homomorphism of the underlying abelian groups MM and NN. Equivalently, it is a collection of homomorphisms {M(σ)N(τ))σ,τK}\{M(\sigma)\to N(\tau))\mid\sigma,\tau\in K\}.

In our exposition, we choose to work with MM being an abelian group. However, everything we say (and everything Ranicki says in [bluebook]) generalizes to the context of RR-modules where RR is a ring with involution.

When the set KK is a finite poset, we are interested in a subcategory of the KK-based abelian groups.

Definition 7.

Let KK be a finite poset.

The objects of (K)mod{\mathbb{Z}}(K)\operatorname{-mod} are the KK-based abelian groups

M=σKM(σ)M=\oplus_{\sigma\in K}M(\sigma)

where M(σ)M(\sigma) is a finitely generated free abelian group for each σK\sigma\in K. A KK-based morphism f:MNf:M\to N is a morphism in (K)mod{\mathbb{Z}}(K)\operatorname{-mod} if, for all τK\tau\in K,

f(M(τ))στN(σ).f(M(\tau))\subset\bigoplus_{\sigma\leq\tau}N(\sigma).

The slogan for morphisms is “bigger to smaller.”

Let KK be a finite simplicial complex. There is an associated poset, also called KK, whose objects are the simplices of KK and whose morphisms are inclusions: στ\sigma\leq\tau means στ\sigma\subseteq\tau. Our quintessential examples of a poset will be either KK or KopK^{\operatorname{{op}}}. Our convention will be that στ\sigma\leq\tau means σKτ\sigma\leq_{K}\tau and we will try and minimize the use of τKopσ\tau\leq_{K^{\operatorname{{op}}}}\sigma. The simplicial chain complex Δ(K)Ch((K)mod)\Delta(K)\in\operatorname{{Ch}}({\mathbb{Z}}(K)\operatorname{-mod}) illustrates the bigger-to-smaller slogan. Here Δ(K)n=σKnΔ(K)n(σ)\Delta(K)_{n}=\oplus_{\sigma\in K^{n}}\Delta(K)_{n}(\sigma), with

Δ(K)n(σ){if n=|σ|0otherwise\Delta(K)_{n}(\sigma)\cong\begin{cases}{\mathbb{Z}}&\textnormal{if }n=|\sigma|\\ 0&\textnormal{otherwise}\end{cases}

Since duality is the fundamental feature of this paper, we introduce it immediately.

Definition 8.

Let KK be a finite poset. The duality functor

:(K)mod((Kop)mod)op*:{\mathbb{Z}}(K)\operatorname{-mod}\to({\mathbb{Z}}(K^{\operatorname{{op}}})\operatorname{-mod})^{\operatorname{{op}}}

is defined on objects by

M=σKM(σ)M^{*}=\bigoplus_{\sigma\in K}M(\sigma)^{*}

where M(σ)=Hom(M(σ),)M(\sigma)^{*}=\operatorname{{Hom}}_{\mathbb{Z}}(M(\sigma),{\mathbb{Z}}). There is a natural isomorphism
E:IdE:\operatorname{Id}\Rightarrow** with EM:MME_{M}:M\to M^{**} induced by EM(σ):M(σ)M(σ)E_{M}(\sigma):M(\sigma)\to M(\sigma)^{**} given by m(ϕϕ(m))m\mapsto(\phi\mapsto\phi(m)). The duality functor and natural isomorphism extend to chain complexes

:Ch((K)mod)Ch((Kop)mod)op-*:\operatorname{{Ch}}({\mathbb{Z}}(K)\operatorname{-mod})\to\operatorname{{Ch}}({\mathbb{Z}}(K^{\operatorname{{op}}})\operatorname{-mod})^{\operatorname{{op}}}

with (C)n=Cn:=(Cn)(C^{-*})_{n}=C^{-n}:=(C_{n})^{*}.

This definition illustrates some of our notational conventions. We write CC (and not CC_{*}) to denote a chain complex. We use CC^{-*} (and not CC^{*}) so that the dual is also a chain complex, whose differential has degree minus one. There are also sign conventions on the differential; we follow the sign conventions of Dold [Dold]: the differential (C)n+1(C)n(C_{-*})_{n+1}\to(C_{-*})_{n} is given by (1)n+1(n)(-1)^{n+1}(\partial_{-n})^{*}.

The simplicial cochain complex ΔKCh((Kop)mod)\Delta K^{-*}\in\operatorname{{Ch}}({\mathbb{Z}}(K^{\operatorname{{op}}})\operatorname{-mod}) of a finite simplicial complex illustrates the KopK^{\operatorname{{op}}}-slogan “smaller-to-bigger.”

Definition 9.

Let KK be a finite simplicial complex and let XX be a finite CW-complex.

  1. 1.

    A KK-dissection of XX is a collection {X(σ)σK}\{X(\sigma)\mid\sigma\in K\} of subcomplexes of XX so that

    1. (a)

      X(σ)X(ρ)={X(σρ)if σρK otherwiseX(\sigma)\cap X(\rho)=\begin{cases}X(\sigma\cap\rho)&\text{if }\sigma\cap\rho\in K\\ \emptyset&\text{ otherwise}\end{cases}

    2. (b)

      X=σKX(σ)X=\displaystyle\bigcup_{\sigma\in K}X(\sigma).

  2. 2.

    A KopK^{\operatorname{{op}}}-dissection of XX is a collection {X(σ)σK}\{X(\sigma)\mid\sigma\in K\} of subcomplexes of XX so that

    1. (a)

      X(σ)X(ρ)={X(σρ)if σρK otherwiseX(\sigma)\cap X(\rho)=\begin{cases}X(\sigma\cup\rho)&\text{if }\sigma\cup\rho\in K\\ \emptyset&\text{ otherwise}\end{cases}

    2. (b)

      X=σKX(σ)X=\displaystyle\bigcup_{\sigma\in K}X(\sigma).

Here σρ\sigma\cup\rho is the smallest simplex of KK which contains σ\sigma and ρ\rho, if it exists. Note that in a KK-dissection στ\sigma\leq\tau implies that X(σ)X(τ)X(\sigma)\subset X(\tau), while in a KopK^{\operatorname{{op}}}-dissection, στ\sigma\leq\tau implies that X(τ)X(σ)X(\tau)\subset X(\sigma).

Remark 10.

The KopK^{\operatorname{{op}}}-dissections described here are KK-dissections in Ranicki’s terminology.

Example 11.

The geometric realization of a finite simplicial complex KK has both a KK-dissection given by the geometric realization of the simplices and a KopK^{\operatorname{{op}}}-dissection given by the dual cones of simplices. We describe the latter in order to fix notation.

Let KK^{\prime} be the barycentric subdivision of KK. The vertices of KK^{\prime} are the barycenters σ¯i{\overline{\sigma}}_{i} of the geometric realization of the simplices σiK\sigma_{i}\in K. An rr-simplex in KK^{\prime} is given by a sequence σ¯0σ¯1σ¯r{\overline{\sigma}}_{0}{\overline{\sigma}}_{1}\dots\overline{\sigma}_{r} where σi<σi+1\sigma_{i}<\sigma_{i+1}, and KK^{\prime} is a subdivision of KK (see Chapter 3, Section 3 of Spanier [S66] for the definition of a subdivision); in particular there is a PL-homeomorphism |K||K||K^{\prime}|\to|K|. For στK\sigma\leq\tau\in K, the dual cell DτσD_{\tau}\sigma is the union of the geometric realization of all simplices σ¯0σ¯1σ¯r{\overline{\sigma}}_{0}{\overline{\sigma}}_{1}\dots\overline{\sigma}_{r} of the barycentric subdivision so that σσ0<σ1<<σrτ\sigma\leq\sigma_{0}<\sigma_{1}<\dots<\sigma_{r}\leq\tau. Define the dual cone of σK\sigma\in K to be DKσ={τστ}DτσD_{K}\sigma=\cup_{\{\tau\mid\sigma\leq\tau\}}D_{\tau}\sigma. Then {DKσ}\{D_{K}\sigma\} gives a KopK^{\operatorname{{op}}}-dissection of the geometric realization of KK.

With KK a 22-simplex, Figure 1 shows a KK and KopK^{\operatorname{{op}}} dissection of the geometric realization of a 22-simplex.

\labellist\hair

2pt \pinlabel KK^{\prime} at 47 50 \pinlabelσ0¯\overline{\sigma_{0}} at 5 67 \pinlabelσ1¯\overline{\sigma_{1}} at 87 67 \pinlabelσ2¯\overline{\sigma_{2}} at 45 155 \pinlabelσ02¯\overline{\sigma_{02}} at 15 112 \pinlabelσ12¯\overline{\sigma_{12}} at 77 112 \pinlabelτ¯\overline{\tau} at 52 108 \pinlabelσ01¯\overline{\sigma_{01}} at 47 67 \pinlabeldual cells at 170 50 \pinlabelDτσ0D_{\tau}\sigma_{0} at 150 85 \pinlabelDτσ1D_{\tau}\sigma_{1} at 190 85 \pinlabelDτσ2D_{\tau}\sigma_{2} at 170 115 \pinlabelDσ0σ0D_{\sigma_{0}}\sigma_{0} at 130 67 \pinlabel KopK^{\operatorname{{op}}}-dissection at 305 110 \pinlabel DKσ0D_{K}\sigma_{0} at 285 141 \pinlabel DKσ1D_{K}\sigma_{1} at 322 141 \pinlabel DKσ2D_{K}\sigma_{2} at 303 170 \pinlabelKK-dissection at 305 -10 \pinlabelσ0\sigma_{0} at 265 03 \pinlabelσ1\sigma_{1} at 352 03 \pinlabelσ2\sigma_{2} at 309 87 \pinlabelσ02\sigma_{02} at 273 43 \pinlabelσ12\sigma_{12} at 341 43 \pinlabelσ01\sigma_{01} at 309 03 \pinlabelτ\tau at 308 35 \endlabellistRefer to caption

Figure 1: Dual cells and KK- and KopK^{\operatorname{{op}}}-dissections of a 2-simplex