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

Homotopy homomorphisms and the classifying space functor

R.M. Vogt rvogt@uos.de Fachbereich Mathematik/Informatik
Universität Osnabrück
Germany
D-49069 Osnabrück
(Month Day, Year; Month Day, Year)
Abstract

We show that the classifying space functor B:on𝒯opB:\mathcal{M}on\to{\mathcal{T}\!op}^{\ast} from the category of topological monoids to the category of based spaces is left adjoint to the Moore loop space functor Ω:𝒯opon\Omega^{\prime}:{\mathcal{T}\!op}^{\ast}\to\mathcal{M}on after we have localized on\mathcal{M}on with respect to all homomorphisms whose underlying maps are homotopy equivalences and 𝒯op{\mathcal{T}\!op}^{\ast} with respect to all based maps which are (not necessarily based) homotopy equivalences. It is well-known that this localization of 𝒯op{\mathcal{T}\!op}^{\ast} exists, and we show that the localization of on\mathcal{M}on is the category of monoids and homotopy classes of homotopy homomorphisms. To make this statement precise we have to modify the classical definition of a homotopy homomorphism, and we discuss the necessary changes. The adjunction is induced by an adjunction up to homotopy B:onw𝒯opw:ΩB:\mathcal{H}\mathcal{M}on^{w}\leftrightarrows{\mathcal{T}\!op}^{w}:\Omega^{\prime} between the category of well-pointed monoids and homotopy homomorphisms and the category of well-pointed spaces. This adjunction is shown to lift to diagrams. As a consequence, the well-known derived adjunction between the homotopy colimit and the constant diagram functor can also be seen to be induced by an adjunction up to homotopy before taking homotopy classes. As applications we among other things deduce a more algebraic version of the group completion theorem and show that the classifying space functor preserves homotopy colimits up to natural homotopy equivalences.

:
55P99
:
55P65,55R35,55R37,55U35
keywords:
Homotopy homomorphism, classifying space, localizations of topologically enriched categories, homotopy adjunction, homotopy colimit, group completion, Moore loop space, James Construction
\published

Month Day, Year \submittedName of Editor

\volumeyear

2012 \volumenumber1 \issuenumber1

\startpage

1

1 Introduction

Let 𝒯op\mathcal{T}\!op denote the category of kk-spaces, 𝒯op{\mathcal{T}\!op}^{\ast} the category of based kk-spaces, and 𝒯opw{\mathcal{T}\!op}^{w} the category of well-pointed kk-spaces. Recall that a space XX is a kk-space if AXA\subset X is closed iff p1(A)p^{-1}(A) is closed in CC for each map p:CXp:C\to X where CC is a compact Hausdorff space, and that a space is called well-pointed if the inclusion of the base point is a closed cofibration.

Let on\mathcal{M}on denote the category of topological monoids and continuous homomorphisms, and onw\mathcal{M}on^{w} and 𝒞on\mathcal{C}\mathcal{M}on the full subcategories of well-pointed, respectively, commutative monoids. A monoid is canonically based by its unit.

We are interested in the relationship between Milgram’s classifying space functor B:on𝒯opB:\mathcal{M}on\to{\mathcal{T}\!op}^{\ast} and the Moore loop space functor Ω:𝒯opon\Omega^{\prime}:{\mathcal{T}\!op}^{\ast}\to\mathcal{M}on (for explicit definitions see Section 4).

The related question for commutative monoids is easily answered: it is well-known that the classifying space BMBM of a commutative monoid is a commutative monoid [18], so that we have a functor B:𝒞on𝒞onB:\mathcal{C}\mathcal{M}on\to\mathcal{C}\mathcal{M}on. The usual loop space functor induces a functor Ω:𝒞on𝒞on\Omega:\mathcal{C}\mathcal{M}on\to\mathcal{C}\mathcal{M}on by defining the multiplication in ΩM\Omega M by point-wise multiplication in MM. The category 𝒞on\mathcal{C}\mathcal{M}on is enriched over 𝒯op{\mathcal{T}\!op}^{\ast} in an obvious way, and it is tensored and cotensored (for definitions see [7] or Section 3). The cotensor MKM^{K} of M𝒞onM\in\mathcal{C}\mathcal{M}on and K𝒯opK\in{\mathcal{T}\!op}^{\ast} is the function space with point-wise multiplication. It is well-known that B(M)MS1B(M)\cong M\boxtimes S^{1}, the tensor of MM and S1S^{1}. Since K-\boxtimes K is left adjoint to ()K(-)^{K} we obtain:

Proposition 1.1.

The functors

B:𝒞on𝒞on:ΩB:\mathcal{C}\mathcal{M}on\rightleftarrows\mathcal{C}\mathcal{M}on:\Omega

form a 𝒯op{\mathcal{T}\!op}^{\ast}-enriched adjoint pair.

In the non-commutative case there is no hope for a similar result. A candidate for a right adjoint of the classifying functor

B:on𝒯opB:\mathcal{M}on\to{\mathcal{T}\!op}^{\ast}

is the Moore loop space functor

Ω:𝒯opon,\Omega^{\prime}:{\mathcal{T}\!op}^{\ast}\to\mathcal{M}on,

but Ω\Omega^{\prime} does not preserve products. In fact, there is no product preserving functor

F:𝒯oponF:{\mathcal{T}\!op}^{\ast}\to\mathcal{M}on

such that F(X)Ω(X)F(X)\simeq\Omega(X) for all XX [6, Prop. 6.1].

Remark 1.2.

In [10] Fiedorowicz showed that the Moore loop space functor into a different target category is right adjoint to what he called the Moore suspension functor: Let 𝒯op[+]{\mathcal{T}\!op}^{\ast}[\mathbb{R}_{+}] be the category whose objects are based spaces XX together with a continuous map p:X+p:X\to\mathbb{R}_{+} (the non-negative real numbers) such that p1(0)=p^{-1}(0)=\ast and whose morphisms are maps over +\mathbb{R}_{+}. Then

Ω:𝒯op𝒯op[+]X(ΩX,l),\Omega^{\prime}:{\mathcal{T}\!op}^{\ast}\to{\mathcal{T}\!op}^{\ast}[\mathbb{R}_{+}]\qquad X\mapsto(\Omega^{\prime}X,l),

where ll is the length function, has this Moore suspension functor as left adjoint.

The Moore loop space funtor Ω:𝒯opon\Omega^{\prime}:{\mathcal{T}\!op}^{\ast}\to\mathcal{M}on preserves products up to natural homotopy. So one might expect it to be a right adjoint of BB after formally inverting homotopy equivalences. We will prove this in this paper.

We will have to localize our categories 𝒞\mathcal{C}, and it is a priori not clear that these localizations exist. A common procedure is to define a Quillen model structure on 𝒞\mathcal{C} such that the morphisms we want to invert are the weak equivalences in these structures. The localization then is the homotopy category Ho𝒞\mathop{\rm Ho}\nolimits\mathcal{C} associated with this model structure.

There are two standard model structures on 𝒯op\mathcal{T}\!op: The structure due to Quillen [21] whose weak equivalences are weak homotopy equivalences and whose fibrations are Serre fibrations, and the structure due to Strøm [25] whose weak equivalences are homotopy equivalences, whose fibrations are Hurewicz fibrations, and whose cofibrations are closed cofibrations.

Although mainstream homotopy theory usually works with the Quillen model structure and the proofs of our results would be considerably shorter in this context (because we could use the rich literature, in particular, the results of Fiedorowicz [10]), we choose the Strøm setting because we share D. Puppe’s point of view [20]: “Frequently a weak homotopy equivalence is considered as good as a genuine one, because for spaces having the homotopy type of a CWCW-complex there is no difference and most interesting spaces in algebraic topology are of that kind. I am not going to argue against this because I agree with it, but I do think that the methods by which we establish the genuine homotopy equivalences give some new insight into homotopy theory.” Moreover, there are spaces of interest which rarely have the homotopy type of a CWCW complex such as function spaces and spaces of foliations, which account for a growing interest in results in the Strøm setting.

So we call a based map in 𝒯op{\mathcal{T}\!op}^{\ast} a weak equivalence if it is a not necessarily based homotopy equivalence, and a homomorphism in on\mathcal{M}on a weak equivalence if the underlying map of spaces is a weak equivalence in 𝒯op{\mathcal{T}\!op}^{\ast}. Let Ho𝒯op\mathop{\rm Ho}\nolimits{\mathcal{T}\!op}^{\ast} and Hoon\mathop{\rm Ho}\nolimits\mathcal{M}on be the categories obtained from 𝒯op{\mathcal{T}\!op}^{\ast} respectively on\mathcal{M}on by formally inverting weak equivalences.

Theorem 1.3.

The categories Ho𝒯op\mathop{\rm Ho}\nolimits{\mathcal{T}\!op}^{\ast} and Hoon\mathop{\rm Ho}\nolimits\mathcal{M}on exist and the classifying space functor and the Moore loop space functor induce a derived adjoint pair

HoB:HoonHo𝒯op:HoΩ\mathop{\rm Ho}\nolimits B:\mathop{\rm Ho}\nolimits\mathcal{M}on\rightleftarrows\mathop{\rm Ho}\nolimits{\mathcal{T}\!op}^{\ast}:\mathop{\rm Ho}\nolimits\Omega^{\prime}
Remark 1.4.

This contrasts the situation in the simplicial category: The loop group functor G:𝒮𝒮ets𝒮𝒢roupsG:\mathcal{SS}ets\to\mathcal{SG}roups from simplicial sets to simplicial groups is left adjoint to the simplicial classifying space functor W¯:𝒮𝒢roups𝒮𝒮ets\overline{W}:\mathcal{SG}roups\to\mathcal{SS}ets (e.g. see [14, Lemma V.5.3]).

With our choice of weak equivalences the Strøm model structure on 𝒯op\mathcal{T}\!op lifts to 𝒯op{\mathcal{T}\!op}^{\ast} so that Ho𝒯op\mathop{\rm Ho}\nolimits{\mathcal{T}\!op}^{\ast} exists, but in contrast to the Quillen model structure, it is not known that the Strøm model lifts to on\mathcal{M}on (there is a model structure on on\mathcal{M}on whose weak equivalences are homotopy equivalences in on\mathcal{M}on rather than homotopy equivalences of underlying spaces; this follows from work of Cole [8] and Barthel and Riel [2]).

In the construction of Hoon\mathop{\rm Ho}\nolimits\mathcal{M}on in the Strøm setting homotopy homomorphisms between monoids come into play: A topological monoid can be considered as an algebra over the operad 𝒜ss\mathcal{A}ss of monoid structures or as a topologically enriched category with one object. The homotopy homomorphisms of this paper are based on the enriched category aspect and describe “functors up to coherent homotopies”. They were introduced for monoids by Sugawara in 1960 [26] and extensively studied by Fuchs in 1965 [11]. Homotopy homomophisms of 𝒜ss\mathcal{A}ss-algebras were introduced in [5], and we will indicate their relation to the ones considered in this paper in Section 2. An extension of our results to arbitrary category objects in 𝒯op\mathcal{T}\!op may be of separate interest.

If we define a semigroup to be a topological space with a continuous associative multiplication, an inspection of the definition shows that a homotopy homomorphism f:MNf:M\to N of monoids is nothing but a semigroup homomorphism W¯MN\overline{W}M\to N where W¯\overline{W} is a variant of the Boardman-Vogt WW-construction [5] (not to be confused with the functor W¯\overline{W} of Remark 1.4). If 𝒮gp\mathop{\rm\mathcal{S}gp} denotes the category of semigroups and continuous homomorphisms then W¯:𝒮gp𝒮gp\overline{W}:\mathop{\rm\mathcal{S}gp}\to\mathop{\rm\mathcal{S}gp} is a functor equipped with a natural transformation ε¯:W¯Id\overline{\varepsilon}:\overline{W}\to\mathop{\rm Id}\nolimits. The Boardman-Vogt WW-construction W:ononW:\mathcal{M}on\to\mathcal{M}on and its associated natural transformation ε:WId\varepsilon:W\to\mathop{\rm Id}\nolimits are obtained from (W¯,ε¯)(\overline{W},\ \overline{\varepsilon}) by factoring out a unit relation. In particular, for any monoid MM there is a natural projection ε(M):W¯MWM\varepsilon^{\prime}(M):\overline{W}M\to WM of semigroups such that ε(M)ε(M)=ε¯(M)\varepsilon(M)\circ\varepsilon^{\prime}(M)=\overline{\varepsilon}(M).

The lack of conditions for the unit is an indication that Sugawara’s notion of a homotopy homomorphism is not quite the correct one. So we define unitary homotopy homomorphisms from MM to NN to be monoid homomorphisms WMNWM\to N; those were studied in 1999 by Brinkmeier [4].

Composition of homotopy homomorphisms and their unitary versions is only associative up to homotopy. To obtain genuine categories of monoids and (unitary) homotopy homomorphisms we modify both notions: A homotopy homomorphisms from MM to NN will be a semigroup homomorphism W¯MW¯N\overline{W}M\to\overline{W}N and a unitary one a monoid homomorphism WMWNWM\to WN. ¿From a homotopy theoretical point of view this modification is not significant:

Proposition 1.5.

If M,NM,\ N are monoids and MM is well-pointed and G,HG,\ H are semigroups then the maps

ε(N):on(WM,WN)on(WM,N)ε¯(N):𝒮gp(W¯G,W¯H)𝒮gp(W¯G,H)\begin{array}[]{rcl}\varepsilon(N)_{\ast}:\mathcal{M}on(WM,WN)&\to&\mathcal{M}on(WM,N)\\ \overline{\varepsilon}(N)_{\ast}:\mathop{\rm\mathcal{S}gp}(\overline{W}G,\overline{W}H)&\to&\mathop{\rm\mathcal{S}gp}(\overline{W}G,H)\end{array}

are homotopy equivalences.

It is well-known that WMMWM\to M has the flavor of a cofibrant replacement of MM as known from model category theory provided MM is well-pointed (e.g. see [3], [27]). So it is no surprise that the category of well-pointed monoids and homotopy classes of unitary homotopy homomorphisms is the localization of onw\mathcal{M}on^{w} with respect to its weak equivalences. If we want to construct Hoon\mathop{\rm Ho}\nolimits\mathcal{M}on we have to relax unitary homotopy homomorphisms to homotopy unitary homotopy homomorphisms and the corresponding statement holds. We will study these various notions of homotopy homomorphisms in Section 2 in detail.

The lack of the appropriate Quillen model structure in some of our categories is made up for by their topological enrichment with nice properties. This topological enrichment allows us to prove stronger results. E.g. the restriction of Theorem 1.3 to the well-pointed case is the path-component version of the following result.

Theorem 1.6.

Let onw\mathcal{H}\mathcal{M}on^{w} be the category of well-pointed monoids and unitary homotopy homomorphisms. Then the classifying space functor and the Moore loop space functor induce an adjunction up to homotopy

onw𝒯opw.\mathcal{H}\mathcal{M}on^{w}\leftrightarrows{\mathcal{T}\!op}^{w}.

In Section 3 we will introduce the necessary notions to make this precise. There we will also recall basic facts from enriched category theory and show that topologically enriched categories with a class of weak equivalences which admit a cofibrant replacement functor can be localized. We believe that these results are of separate interest.

In Section 4 we prove Theorem 1.6 and related results and hence Theorem 1.3. In Section 5 we draw some immediate consequences of Theorem 1.3 and of the intermediate steps in the proof of Theorem 1.6.

E.g. we obtain yet another but considerably shorter proof of a strong version of the James construction.

Definition 1.7.

A Dold space is a topological space admitting a numerable cover {Uγ;γΓ}\{U_{\gamma};\ \gamma\in\Gamma\} such that each inclusion UγXU_{\gamma}\subset X is nullhomotopic.

A space of the homotopy type of a CWCW-complex is a Dold space. For more details on Dold spaces see [22].

Proposition 1.8.

(1) If XX is a well-pointed space and JXJX is the based free topological monoid on XX (the James construction), then BJXΣXBJX\simeq\Sigma X.
(2) If XX is a well-pointed path-connected Dold space, then JXΩΣXJX\simeq\Omega\Sigma X.

Part (2) was first proven in [9], shorter proofs can be found in [20] and [22].

We also obtain a new interpretation of the group completion theorem of a monoid without any additional assumptions on the multiplication.

Definition 1.9.

A topological monoid is called grouplike if it admits a continuous homotopy inversion.

A standard example of a grouplike monoid is the Moore loop space ΩX\Omega^{\prime}X of a space XX.

Theorem 1.10.

Let MM be a well-pointed topological monoid. Then there is a unitary homotopy homomorphism μM:MΩBM\mu_{M}:M\to\Omega^{\prime}BM, natural up to homotopy, having the following universal property: Given any unitary homotopy homomorphism f:MNf:M\to N into a grouplike monoid NN there is a unitary homotopy homomorphism f¯:ΩBMN\bar{f}:\Omega^{\prime}BM\to N, unique up to homotopy, such that f¯μMf\bar{f}\circ\mu_{M}\simeq f. (Here homotopy means homotopy in the category, i.e. homotopy through unitary homotopy homomorphisms.)

¿From the intermediate steps of the proof of Theorem 1.6 we obtain the following extension and strengthening of a theorem of Fuchs [11, Satz 7.7]

Proposition 1.11.

(1) If MM and NN are well-pointed monoids and NN is grouplike then

B:on(WM,WN)𝒯op(BWM,BWN)B:\mathcal{M}on(WM,WN)\to{\mathcal{T}\!op}^{\ast}(BWM,BWN)

is a homotopy equivalence.
(2) If XX is a well-pointed path-connected Dold space then WΩ:𝒯opw(X,Y)onw(WΩX,WΩY)W\Omega^{\prime}:{\mathcal{T}\!op}^{w}(X,Y)\to\mathcal{M}on^{w}(W\Omega^{\prime}X,W\Omega^{\prime}Y) is a homotopy equivalence.

The reader may object that Fuchs considers homotopy homomorphisms while Proposition 1.11 addresses unitary homotopy homomorphisms. Since Fuchs only considers well-pointed grouplike monoids and all his spaces are of the homotopy type of CWCW-complexes the two notions are linked by

Proposition 1.12.

Let MM and NN be well-pointed monoids and NN be grouplike. Then

(ε):on(WM,N)𝒮gp(W¯M,N)(\varepsilon^{\prime})^{\ast}:\mathcal{M}on(WM,N)\to\mathop{\rm\mathcal{S}gp}(\overline{W}M,N)

is a homotopy equivalence,

Section 6 deals with diagrams in topologically enriched categories \mathcal{M} with weak equivalences and a “good” cofibrant replacement functor. We first show that their localizations with respect to maps of diagrams which are objectwise weak equivalences exist. We then show that the well-known derived adjunction induced by the colimit functor and the constant diagram functor is the path-component version of an adjunction up to homotopy between the homotopy colimit functor and the constant diagram functor. We believe that this is of separate interest, too. We then show that the homotopy adjunction of Theorem 1.6 lifts to a homotopy adjunction between the corresponding categories of diagrams. In contrast to strict adjunctions this is a priori not clear, because the associated unit is natural only up to homotopy and hence does not lift to diagrams. We apply this result to prove

Theorem 1.13.

The classifying space functor B:on𝒯opB:\mathcal{M}on\to{\mathcal{T}\!op}^{\ast} preserves homotopy colimits up to natural homotopy equivalences.

The path-component versions of most of our main results are more or less known if we restrict to grouplike monoids. The paper extends these results to general monoids and shows that they arise from stronger statements. Moreover, we show that a topological enrichment with good properties can make up for the non-existence of Quillen model structures.

Acknowledgement: I want to thank P. May for pointing out possible shortcuts to Theorem 1.3 in the Quillen context and for an extended e-mail exchange on the presentation of the paper, and to M. Stelzer for clarifying discussions. I am indebted to the referee for his careful reading of the paper, for requiring a number of clarifications, for suggesting explicit improvements of a number of formulations which had been a bit opaque, and for his patience with my many typos. In particular, the organisation of the present proof of Proposition 4.13 is due to him.

2 Homotopy homomorphisms revisited

In 1960 Sugawara introduced the notion of a strongly homotopy multiplicative map between monoids, which we will call a homotopy homomorphism or hh-morphism, for short [26].

Definition 2.1.

A homotopy homomorphism, or hh-morphism f:MNf:M\to N between two monoids is a sequence of maps

fn:Mn+1×InNnf_{n}:M^{n+1}\times I^{n}\longrightarrow N\quad n\in\mathbb{N}

such that (xiM,tjI)(x_{i}\in M,t_{j}\in I)

fn(x0,t1,x1,t2,,tn,xn)={fn1(x0,t1,,xi1xi,,tn,xn) if ti=0fi1(x0,t1,,xi1)fni(xi,ti+1,,xn) if ti=1.f_{n}(x_{0},t_{1},x_{1},t_{2},\ldots,t_{n},x_{n})\\ =\left\{\begin{array}[]{ll}f_{n-1}(x_{0},t_{1},\ldots,x_{i-1}\cdot x_{i},\ldots,t_{n},x_{n})&\textrm{ if }t_{i}=0\\ f_{i-1}(x_{0},t_{1},\ldots,x_{i-1})\cdot f_{n-i}(x_{i},t_{i+1},\ldots,x_{n})&\textrm{ if }t_{i}=1.\end{array}\right.

We call f0:MNf_{0}:M\to N the underlying map of ff.

If in addition f0(eM)=eNf_{0}(e_{M})=e_{N} and

fn(x0,t1,x1,t2,,tn,xn)={fn1(x1,t2,,xn) if x0=eMfn1(x0,,xi1,max(ti,ti+1),xi+1,,xn) if xi=eMfn1(x0,t1,,xn1) if xn=eMf_{n}(x_{0},t_{1},x_{1},t_{2},\ldots,t_{n},x_{n})\\ =\left\{\begin{array}[]{ll}f_{n-1}(x_{1},t_{2},\ldots,x_{n})&\textrm{ if }x_{0}=e_{M}\\ f_{n-1}(x_{0},\ldots,x_{i-1},\max(t_{i},t_{i+1}),x_{i+1},\ldots,x_{n})&\textrm{ if }x_{i}=e_{M}\\ f_{n-1}(x_{0},t_{1},\ldots,x_{n-1})&\textrm{ if }x_{n}=e_{M}\end{array}\right.

where eMMe_{M}\in M and eNNe_{N}\in N are the units. We call ff a unitary homotopy homomorphism or uhuh-morphism, for short.

Since an hh-morphism does not pay tribute to the unit it does not seem to be the right notion for maps between monoids. E.g. if we require f0f_{0} to be a based map so that it preserves the unit we would like the path

f0(x0x1)\textstyle{f_{0}(x_{0}\cdot x_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f1(x0,t,x1)\scriptstyle{f_{1}(x_{0},t,x_{1})}f0(x0)f0(x1)\textstyle{f_{0}(x_{0})\cdot f_{0}(x_{1})}

to be the constant one, if x0x_{0} or x1x_{1} is the unit. Unitary hh-morphisms have this property. Nevertheless, in the past one usually considered hh-morphisms because the additional conditions for uhuh-morphisms make it harder to work with them.

We will later find it more convenient to work with homotopy unitary homotopy homomorphisms which preserve the unit only up to homotopy. We will introduce those at the end of this section.

The most extensive study of hh-morphisms and their induced maps on classifying spaces was done by Fuchs [11], who constructed composites of hh-morphisms, proved that composition is homotopy associative and stated that an hh-morphism f:MNf:M\to N whose underlying map is a homotopy equivalence has a homotopy inverse hh-morphism g:NMg:N\to M. In fact, he constructed g0g_{0}, g1g_{1} and the homotopies gfidg\circ f\simeq\mathop{\rm id}\nolimits and fgidf\circ g\simeq\mathop{\rm id}\nolimits in dimensions 0 and 1 in [11, p.205-p.208], but left the rest to the reader. He produced a complete proof in [12].

We handle these problems by interpreting homotopy homomorphisms as genuine homomorphisms of a “cofibrant” replacement of MM.

By a semigroup we will mean a kk-space with a continuous associative multiplication. Let 𝒮gp\mathop{\rm\mathcal{S}gp} denote the category of semigroups and continuous homomorphisms.

Constructions 2.2.

We will construct continuous functors

W¯:𝒮gp𝒮gp and W:onon\overline{W}:\mathop{\rm\mathcal{S}gp}\longrightarrow\mathop{\rm\mathcal{S}gp}\quad\textrm{ and }\quad W:\mathcal{M}on\longrightarrow\mathcal{M}on

and natural transformations

ε¯:W¯Id and ε:WId\overline{\varepsilon}:\overline{W}\longrightarrow\mathop{\rm Id}\nolimits\quad\textrm{ and }\quad\varepsilon:W\longrightarrow\mathop{\rm Id}\nolimits

as follows:

W¯M=(n=0Mn+1×In)/\overline{W}M=\left(\coprod\limits^{\infty}_{n=0}M^{n+1}\times I^{n}\right)/\sim

with the relation

(1) (x0,t1,x1,t2,,tn,xn)(x0,t1,,xi1xi,,tn,xn)(x_{0},t_{1},x_{1},t_{2},\ldots,t_{n},x_{n})\sim(x_{0},t_{1},\ldots,x_{i-1}\cdot x_{i},\ldots,t_{n},x_{n})  if ti=0t_{i}=0
and WMWM is the quotient of W¯M\overline{W}M by imposing the additional relations

(2) (x0,t1,x1,t2,,tn,xn){(x1,t2,,xn) if x0=e(x0,,xi1,max(ti,ti+1),xi+1,,xn) if xi=e(x0,t1,,xn1) if xn=e(x_{0},t_{1},x_{1},t_{2},\ldots,t_{n},x_{n})\\ {}\qquad\ \sim\left\{\begin{array}[]{ll}(x_{1},t_{2},\ldots,x_{n})&\textrm{ if }x_{0}=e\\ (x_{0},\ldots,x_{i-1},\max(t_{i},t_{i+1}),x_{i+1},\ldots,x_{n})&\textrm{ if }x_{i}=e\\ (x_{0},t_{1},\ldots,x_{n-1})&\textrm{ if }x_{n}=e\end{array}\right.
The multiplications of W¯M\overline{W}M and WMWM are given on representatives by

(x0,t1,,xk)(y0,u1,,yl)=(x0,t1,xk,1,y0,u1,,yl).(x_{0},t_{1},\ldots,x_{k})\cdot(y_{0},u_{1},\ldots,y_{l})=(x_{0},t_{1},\ldots x_{k},1,y_{0},u_{1},\ldots,y_{l}).

The natural transformations ε¯\overline{\varepsilon} and ε\varepsilon are defined by

ε¯(M),ε(M):(x0,t1,,xn)x0x1xn.\overline{\varepsilon}(M),\ \varepsilon(M):(x_{0},t_{1},\ldots,x_{n})\longmapsto x_{0}\cdot x_{1}\cdot\ldots\cdot x_{n}.

Their underlying maps have natural sections

ι¯(M),ι(M):x(x)\bar{\iota}(M),\ \iota(M):x\longmapsto(x)

which are not homomorphisms, and there is a homotopy over MM

hs:(x0,t1,x1,,tn,xn)(x0,st1,x1,,stn,xn)h_{s}:(x_{0},t_{1},x_{1},\ldots,t_{n},x_{n})\longmapsto(x_{0},s\cdot t_{1},x_{1},\ldots,s\cdot t_{n},x_{n})

from ι¯(M)ε¯(M)\overline{\iota}(M)\circ\overline{\varepsilon}(M) respectively ι(M)ε(M)\iota(M)\circ\varepsilon(M) to the identity. In particular, ε¯(M)\overline{\varepsilon}(M) and ε(M)\varepsilon(M) are shrinkable as maps.

If MM is a monoid the projection

ε(M):W¯MWM\varepsilon^{\prime}(M):\overline{W}M\to WM

is a homomorphism of semigroups satisfying

ε¯(M)=ε(M)ε(M)andε(M)ι¯(M)=ι(M).\bar{\varepsilon}(M)=\varepsilon(M)\circ\varepsilon^{\prime}(M)\qquad\textrm{and}\qquad\varepsilon^{\prime}(M)\circ\overline{\iota}(M)=\iota(M).

By inspection we see

Observation 2.3.
  1. (1)

    hh-morphisms (fn):MN(f_{n}):M\to N correspond bijectively to homomorphisms f¯:W¯MN\bar{f}:\overline{W}M\to N of semigroups, and f0=f¯ι¯(M)f_{0}=\bar{f}\circ\overline{\iota}(M)

  2. (2)

    uhuh-morphisms (fn):MN(f_{n}):M\to N correspond bijectively to homomorphisms f:WMNf:WM\to N of monoids, and f0=fι(M)f_{0}=f\circ\iota(M)

Observation 2.4.

Algebraically, W¯M\overline{W}M is a free semigroup and WMWM is a free monoid. The indecomposables are precisely those elements which have a representative (x0,t1,x1,,xn)(x_{0},t_{1},x_{1},\ldots,x_{n}) where no tit_{i} equals 11.

2.5.

The formal relation between W¯\overline{W} and WW: The forgetful functor i:on𝒮gpi:\mathcal{M}on\to\mathop{\rm\mathcal{S}gp} has a left adjoint

()+:𝒮gpon,GG+,(-)_{+}:\mathop{\rm\mathcal{S}gp}\to\mathcal{M}on,\qquad G\mapsto G_{+},

where G+=G{}G_{+}=G\sqcup\{\ast\} with \ast as unit. It follows from the definitions that the diagram

𝒮gp\textstyle{\mathop{\rm\mathcal{S}gp}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}()+\scriptstyle{(-)_{+}}W¯\scriptstyle{\overline{W}}𝒮gp\textstyle{\mathop{\rm\mathcal{S}gp}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}()+\scriptstyle{(-)_{+}}on\textstyle{\mathcal{M}on\ignorespaces\ignorespaces\ignorespaces\ignorespaces}W\scriptstyle{W}on\textstyle{\mathcal{M}on}

commutes up to natural isomorphisms in on\mathcal{M}on.

Both constructions have a universal property, which is a consequence of the following result. We give 𝒯op(X,Y){\mathcal{T}\!op}^{\ast}(X,Y) and 𝒯op(X,Y)\mathcal{T}\!op(X,Y) the kk-function space topology, obtained by turning the space of all maps from XX to YY with the compact-open topology into a kk-space. We give on(M,N)\mathcal{M}on(M,N) and 𝒮gp(M,N)\mathop{\rm\mathcal{S}gp}(M,N) the subspace topologies of the corresponding function spaces in 𝒯op{\mathcal{T}\!op}^{\ast} respectively 𝒯op\mathcal{T}\!op.

Definition 2.6.

We call a homomorphism f:MNf:M\to N in on\mathcal{M}on or 𝒮gp\mathop{\rm\mathcal{S}gp} a weak equivalence if its underlying map of spaces is a homotopy equivalence in 𝒯op\mathcal{T}\!op. (Recall that a weak equivalence in on\mathcal{M}on is a homotopy equivalence of underlying spaces in 𝒯op{\mathcal{T}\!op}^{\ast} if MM and NN are well-pointed.)

Proposition 2.7.
  1. (1)

    Let MM be a well-pointed monoid and p:XYp:X\to Y a homomorphism of monoids. Let

    p:on(WM,X)on(WM,Y)p_{\ast}:\mathcal{M}on(WM,X)\longrightarrow\mathcal{M}on(WM,Y)

    be the induced map. If pp is a fibration of underlying spaces, so is pp_{\ast}. If pp is a weak equivalence, pp_{\ast} is a homotopy equivalence.

  2. (2)

    The same holds for W¯\overline{W} and an arbitrary object MM in the category 𝒮gp\mathop{\rm\mathcal{S}gp}.

Proof.

Let p:XYp:X\to Y be a weak equivalence. By the HELP-Lemma [28] in 𝒯op\mathcal{T}\!op with the Strøm model structure [24] we have to show: Given a diagram of spaces

(A)

A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i\scriptstyle{i}f¯A\scriptstyle{\bar{f}_{A}}on(WM,X)\textstyle{\mathcal{M}on(WM,X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p_{\ast}}B\textstyle{B\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g¯\scriptstyle{\bar{g}}on(WM,Y)\textstyle{\mathcal{M}on(WM,Y)}

which commutes up to a homotopy h¯A,t:g¯ipf¯A\bar{h}_{A,t}:\bar{g}\circ i\simeq p_{\ast}\circ\bar{f}_{A}, where ii is a closed cofibration, there are extensions f¯:Bon(WM,X)\bar{f}:B\to\mathcal{M}on(WM,X) of f¯A\bar{f}_{A} and h¯t:Bon(WM,Y)\bar{h}_{t}:B\to\mathcal{M}on(WM,Y) of h¯A,t\bar{h}_{A,t} such that h¯t:g¯pf¯\bar{h}_{t}:\bar{g}\simeq p_{\ast}\circ\bar{f}.

Passing to adjoints we obtain a diagram

WM×A\textstyle{WM\times A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id×i\scriptstyle{id\times i}fA\scriptstyle{f_{A}}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}WM×B\textstyle{WM\times B\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{g}Y\textstyle{Y}

commuting up to a homotopy hA,th_{A,t}, such that each fa=fA|WM×{a}f_{a}=f_{A}|WM\times\{a\}, each gb=g|WM×{b}g_{b}=g|WM\times\{b\}, and each ha,t=hA,t|WM×{a}h_{a,t}=h_{A,t}|WM\times\{a\} is a homomorphism. We have to construct extensions f:WM×BXf:WM\times B\to X and ht:WM×BYh_{t}:WM\times B\to Y of fAf_{A} and hA,th_{A,t} such that ht:gpfh_{t}:g\simeq p\circ f and each hb,th_{b,t} and fb,bBf_{b},\ b\in B is a homomorphism.

We filter WM×BWM\times B by closed subspaces Fn×BF_{n}\times B, where FnF_{n} is the submonoid of WMWM generated by all elements having a representative (x0,t1,,tk,xk)(x_{0},t_{1},\ldots,t_{k},x_{k}) with knk\leq n. We put F1={e}F_{-1}=\{e\}. Then ff and hth_{t} are uniquely determined on F1×BF_{-1}\times B.

Suppose that ff and hth_{t} have been defined on Fn1×BF_{n-1}\times B. An element (x0,t1,,tn,xn)(x_{0},t_{1},\ldots,t_{n},x_{n}) represents an element in Fn1F_{n-1} iff one of the following conditions holds

  • some xi=ex_{i}=e  (relation 2.2.2)

  • some ti=0t_{i}=0  (relation 2.2.1)

  • some ti=1t_{i}=1  (it represents a product in Fn1F_{n-1}).

If DMn+1Mn+1DM^{n+1}\subset M^{n+1} denotes the subspace of points with some coordinate ee, then ff and hth_{t} are already defined on (DMn+1×InMn+1×In)×BMn+1×In×A(DM^{n+1}\times I^{n}\cup M^{n+1}\times\partial I^{n})\times B\cup M^{n+1}\times I^{n}\times A. The elements in (Mn+1×In)\(DMn+1×InMn+1×In)(M^{n+1}\times I^{n})\backslash(DM^{n+1}\times I^{n}\cup M^{n+1}\times\partial I^{n}) represent indecomposables of filtration nn, but not of lower filtration. Consider the diagram

(B)

(DMn+1×InMn+1×In)×BMn+1×In×A\textstyle{(DM^{n+1}\times I^{n}\cup M^{n+1}\times\partial I^{n})\times B\cup M^{n+1}\times I^{n}\times A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}j\scriptstyle{j}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}Mn+1×In×B\textstyle{M^{n+1}\times I^{n}\times B\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{g}Y\textstyle{Y}

(in abuse of notation we use gg for the composite Mn+1×In×BWM×BYM^{n+1}\times I^{n}\times B\to WM\times B\to Y). Diagram (B) commutes up to the homotopy hth_{t} and we need an extension of ff and hth_{t} to Mn+1×In×BM^{n+1}\times I^{n}\times B. These extensions exist by the HELP-Lemma, because our assumptions ensure that jj is a closed cofibration. So we have defined ff and hth_{t} for indecomposable generators (x0,t1,,tn,xn)(x_{0},t_{1},\ldots,t_{n},x_{n}) of FnF_{n}. We extend these maps to Fn×BF_{n}\times B by the conditions that each fbf_{b} and hb,t,bBh_{b,t},\ b\in B be a homomorphism using Observation 2.4.

Now suppose that pp is a fibration. By [24, Thm. 8] we need to consider a commutative diagram (A), where ii is a closed cofibration and a homotopy equivalence, and we have to find an extension f¯:Bon(WM,X)\bar{f}:B\to\mathcal{M}on(WM,X) of f¯A\bar{f}_{A} such that g¯=pf¯\bar{g}=p_{\ast}\circ\bar{f}. We proceed as above. In the inductive step we have a commutative diagram (B). Since ii is a closed cofibration and a homotopy equivalence so is jj by the pushout-product theorem for cofibrations. Hence the required extension f:Mn+1×In×BXf:M^{n+1}\times I^{n}\times B\to X exists by [24, Thm. 8].

Part (2) is proved in the same way starting with F1M=F_{-1}M=\emptyset. ∎

As an immediate consequence we obtain the

2.8.

Lifting Theorem: (1) Given homomorphisms of monoids

X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}WM\textstyle{WM\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}Y\textstyle{Y}

such that pp is a weak equivalence and MM is well-pointed, then there exists a homomorphism g:WMXg:WM\to X, unique up to homotopy in on\mathcal{M}on (i.e. a homotopy through homomorphisms), such that fpgf\simeq p\circ g in on\mathcal{M}on.
If, in addition, the underlying map of pp is a fibration there is a homomorphism g:WMXg:WM\to X, unique up to homotopy in on\mathcal{M}on, such that f=pgf=p\circ g.

(2) For W¯\overline{W} the analogous results hold in the category 𝒮gp\mathop{\rm\mathcal{S}gp}.

2.9.

By Proposition 2.7 the second one of the maps

ε(N):on(WM,WN)on(WM,N)ε¯(N):𝒮gp(W¯M,W¯N)𝒮gp(W¯M,N)\begin{array}[]{rcl}\varepsilon(N)_{\ast}:\mathcal{M}on(WM,WN)&\to&\mathcal{M}on(WM,N)\\ \overline{\varepsilon}(N)_{\ast}:\mathop{\rm\mathcal{S}gp}(\overline{W}M,\overline{W}N)&\to&\mathop{\rm\mathcal{S}gp}(\overline{W}M,N)\end{array}

is a homotopy equivalence, and the first one is a homotopy equivalence if MM is well-pointed.

To guarantee the well-pointedness condition we introduce the whiskering functor.

2.10.

The whiskering construction: We define a functor

Vt:𝒯op𝒯opwV^{t}:{\mathcal{T}\!op}^{\ast}\to{\mathcal{T}\!op}^{w}

by Vt(X,x0)=(XI)/(x01)V^{t}(X,x_{0})=(X\sqcup I)/(x_{0}\sim 1) and choose 0I0\in I as base-point of XIX_{I}. Then VtXV^{t}X is well-pointed, and the natural map q(X):VtXXq(X):V^{t}X\to X mapping II to x0x_{0} is a homotopy equivalence. Its homotopy inverse q¯(X):XVtX\bar{q}(X):X\to V^{t}X is the canonical map. If XX is well-pointed, q(X)q(X) is a based homotopy equivalence.

This functor lifts to a functor

V:ononwV:\mathcal{M}on\to\mathcal{M}on^{w}

defined by V(M)=Vt(M)V(M)=V^{t}(M) with x0x_{0} replaced by eMe_{M} with the multiplication

xy={xyMif x,yMxif xM,yIyif yM,xImax(x,y)if x,yIx\cdot y=\left\{\begin{array}[]{ll}x\cdot y\in M&\textrm{if }x,y\in M\\ x&\textrm{if }x\in M,\ y\in I\\ y&\textrm{if }y\in M,\ x\in I\\ \textrm{max}(x,y)&\textrm{if }x,y\in I\end{array}\right.

Since 0I0\in I is the unit of VMVM the monoid VMVM is well-pointed. The natural map q(M):VMMq(M):VM\to M is a weak equivalence in on\mathcal{M}on, but observe that q¯(M):XVM\bar{q}(M):X\to VM is not a homomorphism because it does not preserve the unit.

A homomorphism f:WVMNf:WVM\to N can be considered a homotopy unitary homotopy homomorphism. Strictly speaking, the underlying map of f:WVMNf:WVM\to N is

f0=fι(VM)q¯(M):MVMWVMN.f_{0}=f\circ\iota(VM)\circ\bar{q}(M):M\to VM\to WVM\to N.

We note that f0f_{0} preserves the unit eMe_{M} only up to homotopy.

By 2.9 the following change of our notations of homotopy homomorphisms is insignificant from a homotopy theoretic point of view:

Definition 2.11.

From now on a homotopy unitary homotopy homomorphism, huhhuh-morphism for short, from MM to NN is a homomorphism f:WVMWVNf:WVM\to WVN. Its underlying map is q(N)ε(VN)fι(VM)q¯(M)q(N)\circ\varepsilon(VN)\circ f\circ\iota(VM)\circ\bar{q}(M).
A unitary homotopy homomorphism, uhuh-morphism for short, from MM to NN is a homomorphism f:WMWNf:WM\to WN. Its underlying map is ε(N)fι(M)\varepsilon(N)\circ f\circ\iota(M).
A homotopy homomorphism, hh-morphism for short, from the semigroup MM to the semigroup NN is a homomorphism f:W¯MW¯Nf:\overline{W}M\to\overline{W}N. Its underlying map is ε¯(N)fι¯(M)\overline{\varepsilon}(N)\circ f\circ\overline{\iota}(M).

This solves the problem of composition, and from 2.7 we obtain

Proposition 2.12.

If f:WMWNf:WM\to WN is a uhuh-morphism from MM to NN whose underlying map is a homotopy equivalence, and MM and NN are well-pointed, then ff is a homotopy equivalence in the category on\mathcal{M}on.
If f:WVMWVNf:WVM\to WVN is a huhhuh-morphism from MM to NN, whose underlying map is a homotopy equivalence, then ff is a homotopy equivalence in the category on\mathcal{M}on.
The analogous statement in 𝒮gp\mathop{\rm\mathcal{S}gp} holds for homomorphisms W¯MW¯N\overline{W}M\to\overline{W}N.

Monoids are algebras over the operad 𝒜ss\mathcal{A}ss of monoid structures, and there is the notion of an “operadic” homotopy homomorphism defined by Boardman and Vogt in [5]. M. Klioutch compared the operadic notion with the one considered in this paper and could show [17]

Proposition 2.13.

Let MM and NN be well-pointed monoids and let H(M,N)(M,N) be the space of operadic homotopy homomorphisms from MM to NN, then there is a natural homotopy equivalence

H(M,N)on(WM,N).\textrm{H}(M,N)\simeq\mathcal{M}on(WM,N).

3 Categorical prerequisites and localizations

The functors WV:ononWV:\mathcal{M}on\to\mathcal{M}on and W¯:𝒮gp𝒮gp\overline{W}:\mathop{\rm\mathcal{S}gp}\to\mathop{\rm\mathcal{S}gp} resemble cofibrant replacement functors as known from Quillen model category theory. Unfortunately, there is no known model category structure on on\mathcal{M}on with our choice of weak equivalences. This draw-back is made up by the topological enrichment of our categories as we will see in this section.

Our categories are enriched over 𝒯op{\mathcal{T}\!op}^{\ast} or 𝒯op\mathcal{T}\!op. So we have a natural notion of homotopy. Moreover, they are tensored and cotensored. Recall that a 𝒯op{\mathcal{T}\!op}^{\ast}-enriched category \mathcal{M} is tensored and cotensored (over 𝒯op{\mathcal{T}\!op}^{\ast}) if there are functors

𝒯op×,(X,M)XM(𝒯op)op×,(X,M)MX\begin{array}[]{rclcl}{\mathcal{T}\!op}^{\ast}\times\mathcal{M}&\to&\mathcal{M},&&(X,M)\mapsto X\boxtimes M\\ ({\mathcal{T}\!op}^{\ast})^{\mathop{\rm op}\nolimits}\times\mathcal{M}&\to&\mathcal{M},&&(X,M)\mapsto M^{X}\end{array}

and natural homeomorphisms

(XM,N)𝒯op(X,(M,N))(M,NX).\mathcal{M}(X\boxtimes M,N)\cong{\mathcal{T}\!op}^{\ast}(X,\mathcal{M}(M,N))\cong\mathcal{M}(M,N^{X}).

These properties imply that for based spaces XX and YY and objects MM\in\mathcal{M} there are natural isomorphisms

(XY)MX(YM).(X\wedge Y)\boxtimes M\cong X\boxtimes(Y\boxtimes M).

The definition in the 𝒯op\mathcal{T}\!op-enriched case is similar. To distinguish between the based and the non-based case we denote the tensor over 𝒯op\mathcal{T}\!op by XMX\otimes M. The natural isomorphism in the non-based case reads

(X×Y)MX(YM).(X\times Y)\otimes M\cong X\otimes(Y\otimes M).

Forgetting base points turns a 𝒯op{\mathcal{T}\!op}^{\ast}-enriched category \mathcal{M} into a 𝒯op\mathcal{T}\!op-enriched one. If \mathcal{M} is tensored over 𝒯op{\mathcal{T}\!op}^{\ast} it is also tensored over 𝒯op\mathcal{T}\!op: we define

XM=X+MX\otimes M=X_{+}\boxtimes M

where X+=X{}X_{+}=X\sqcup\{\ast\} with the additional point as base point.

Example 3.1.

on\mathcal{M}on is 𝒯op{\mathcal{T}\!op}^{\ast}-enriched, tensored and cotensored [19, Prop. 2.10]. The cotensor MXM^{X} is the kk-function space with pointwise multiplication, XMX\boxtimes M is more complicated: as a set, it is a free product of copies MM, one copy for each xXx\in X different from the base point. By the same argument as in [19] the category 𝒮gp\mathop{\rm\mathcal{S}gp} is 𝒯op\mathcal{T}\!op-enriched and tensored and cotensored over 𝒯op\mathcal{T}\!op.

If 𝒮gp\otimes_{\mathop{\rm\mathcal{S}gp}} denotes the tensor in 𝒮gp\mathop{\rm\mathcal{S}gp} and \otimes the one over 𝒯op\mathcal{T}\!op in on\mathcal{M}on, then the universal properties of the tensor and of the adjunction of 2.5 imply that there is a natural isomorphism

(K𝒮gpG)+K(G+)(K\otimes_{\mathop{\rm\mathcal{S}gp}}G)_{+}\cong K\otimes(G_{+})

in on\mathcal{M}on for semigroups GG.

Definition 3.2.

Let \mathcal{M} be a 𝒯op\mathcal{T}\!op-enriched category. Two morphisms f,g:AXf,g:A\to X are called homotopic if there is a path in (A,X)\mathcal{M}(A,X) joining ff and gg.

Clearly, the homotopy relation is an equivalence relation preserved under composition. Passing to path components we obtain the homotopy category π\pi\mathcal{M}.

If \mathcal{M} is tensored over 𝒯op\mathcal{T}\!op it has a canonical cylinder functor MIMM\mapsto I\otimes M. The associated homotopy notion coincides with the one of Definition 3.2.

Definition 3.3.

Let \mathcal{M} be a category and 𝒲\mathscr{W} a class of morphisms in \mathcal{M}, which we will call weak equivalences. The localization of \mathcal{M} with respect to 𝒲\mathscr{W} is a category [𝒲1]\mathcal{M}[\mathscr{W}^{-1}] with ob[𝒲1]=ob\mathop{\rm ob}\nolimits\mathcal{M}[\mathscr{W}^{-1}]=\mathop{\rm ob}\nolimits\mathcal{M} and a functor γ:[𝒲1]\gamma:\mathcal{M}\to\mathcal{M}[\mathscr{W}^{-1}] such that
(1) γ\gamma is the identity on objects
(2) γ(f)\gamma(f) is an isomorphism for all f𝒲f\in\mathscr{W}
(3) if F:𝒟F:\mathcal{M}\to\mathcal{D} is a functor such that F(f)F(f) is an isomorphism for all f𝒲f\in\mathscr{W} then there exists a unique functor F¯:[𝒲1]𝒟\overline{F}:\mathcal{M}[\mathscr{W}^{-1}]\to\mathcal{D} such that F=F¯γF=\overline{F}\circ\gamma.

Proposition 3.4.

Let \mathcal{M} be a 𝒯op\mathcal{T}\!op-enriched tensored category and 𝒲\mathscr{W} a class of morphisms in \mathcal{M} such that
(1) 𝒲\mathscr{W} contains all homotopy equivalences,
(2) there is a functor Q:Q:\mathcal{M}\to\mathcal{M} and a natural transformation ε:QId\varepsilon:Q\to\mathop{\rm Id}\nolimits or a natural transformation η:IdQ\eta:\mathop{\rm Id}\nolimits\to Q taking values in 𝒲\mathscr{W} such that QfQf is a homotopy equivalence for each f𝒲f\in\mathscr{W}.
Then [𝒲1]\mathcal{M}[\mathscr{W}^{-1}] exists. Precisely, let \mathcal{HM} be the category with ob𝒞=ob\mathop{\rm ob}\nolimits\mathcal{HC}=\mathop{\rm ob}\nolimits\mathcal{M} and (M1,M2)=(QM1,QM2)\mathcal{HM}(M_{1},M_{2})=\mathcal{M}(QM_{1},QM_{2}). Then [𝒲1]=π𝒞\mathcal{M}[\mathscr{W}^{-1}]=\pi\mathcal{HC}, the quotient category obtained by passing to homotopy classes. The functor γ:[𝒲1]\gamma:\mathcal{M}\to\mathcal{M}[\mathscr{W}^{-1}] is the identity on objects and maps a morphism ff to the homotopy class of QfQf.

Proof.

The proof is essentially the same as in the case of a Quillen model category (e.g. see [13, Thm 8.3.5]). We recall the construction of the localization [𝒲1]\mathcal{M}[\mathscr{W}^{-1}] in this case. So let \mathcal{M} be a Quillen model category, let ε:CId\varepsilon:C\to\mathop{\rm Id}\nolimits respectively η;IdR\eta;\mathop{\rm Id}\nolimits\to R be a cofibrant respectively fibrant replacement functor. There are cylinder objects giving rise to the left homotopy relation.
Step 1: Using the fact that RC(X)RC(X) is fibrant and cofibrant for each object XX in \mathcal{M} one proves that left homotopy is an equivalence relation on (RC(A),RC(X))\mathcal{M}(RC(A),RC(X)) which is preserved under composition. Let π(RC(A),RC(X))\pi\mathcal{M}(RC(A),RC(X)) be the set of equivalence classes. One defines

ob[𝒲1]=oband[𝒲1](A,B)=π(RC(A),RC(B)),\mathop{\rm ob}\nolimits\mathcal{M}[\mathscr{W}^{-1}]=\mathop{\rm ob}\nolimits\mathcal{M}\quad\textrm{and}\quad\mathcal{M}[\mathscr{W}^{-1}](A,B)=\pi\mathcal{M}(RC(A),RC(B)),

and it follows that [𝒲1]\mathcal{M}[\mathscr{W}^{-1}] is a category.
Step 2: One proves that RC(f)RC(f) is a homotopy equivalence if f:AXf:A\to X is a weak equivalence. Then one defines

γ:[𝒲1]fRC(f).\gamma:\mathcal{M}\to\mathcal{M}[\mathscr{W}^{-1}]\qquad f\mapsto RC(f).

In particular, γ\gamma maps weak equivalences to isomorphisms.
Step 3: One shows that a functor F:𝒩F:\mathcal{M}\to\mathcal{N}, which maps weak equivalences to isomorphisms, maps homotopic morphisms to the same morphism.
Step 4: Given a functor F:𝒩F:\mathcal{M}\to\mathcal{N}, which maps weak equivalences to isomorphisms, then there is a unique functor F¯:[𝒲1]𝒩\bar{F}:\mathcal{M}[\mathscr{W}^{-1}]\to\mathcal{N} such that F=F¯γF=\bar{F}\circ\gamma, and F¯\bar{F} is defined on objects by F¯(X)=F(X)\bar{F}(X)=F(X) and on morphisms [f][𝒲1](A,X)[f]\in\mathcal{M}[\mathscr{W}^{-1}](A,X) by

F¯([f])=F(ε(X))(F(η(CX)))1F(f)F(η(CA))(F(ε(A)))1,\bar{F}([f])=F(\varepsilon(X))\circ(F(\eta(CX)))^{-1}\circ F(f)\circ F(\eta(CA))\circ(F(\varepsilon(A)))^{-1},

where [f][f] is the homotopy class of ff.

We now prove Proposition 3.4. We deal with the case where we have a natural transformation ε:QId\varepsilon:Q\to\mathop{\rm Id}\nolimits taking values in 𝒲\mathscr{W}.
Step 1 follows from the topological enrichment

ob[𝒲1]=oband[𝒲1](A,B)=π(Q(A),Q(B))\mathop{\rm ob}\nolimits\mathcal{M}[\mathscr{W}^{-1}]=\mathop{\rm ob}\nolimits\mathcal{M}\quad\textrm{and}\quad\mathcal{M}[\mathscr{W}^{-1}](A,B)=\pi\mathcal{M}(Q(A),Q(B))

which is a category.
Step 2 holds by Assumption 3.4.2, and we define

γ:[𝒲1]fQ(f).\gamma:\mathcal{M}\to\mathcal{M}[\mathscr{W}^{-1}]\qquad f\mapsto Q(f).

γ\gamma maps weak equivalences to isomorphisms.
For Step 3 we need the cylinder functor: the bottom and top inclusions i0id,i1id:XXIXi_{0}\otimes\mathop{\rm id}\nolimits,i_{1}\otimes\mathop{\rm id}\nolimits:X\cong\ast\otimes X\to I\otimes X into the cylinder are homotopy equivalences with the common homotopy inverse rid:IXXXr\otimes\mathop{\rm id}\nolimits:I\otimes X\to\ast\otimes X\cong X.
Step 4: Given a functor F:𝒩F:\mathcal{M}\to\mathcal{N}, which maps weak equivalences to isomorphisms, we define F¯:[𝒲1]𝒩\bar{F}:\mathcal{M}[\mathscr{W}^{-1}]\to\mathcal{N} by

F¯(X)=F(X)andF¯([f])=F(ε(X))F(f)(F(ε(A)))1\bar{F}(X)=F(X)\quad\textrm{and}\quad\bar{F}([f])=F(\varepsilon(X))\circ F(f)\circ(F(\varepsilon(A)))^{-1}

for [f][𝒲1](A,X)[f]\in\mathcal{M}[\mathscr{W}^{-1}](A,X). The rest follows like in [13, Thm 8.3.5]. ∎

Remark 3.5.

For Proposition 3.4 we do not need that the tensor XMX\otimes M exists for all topological spaces: it suffices that \mathcal{M} is tensored over the full subcategory of 𝒯op\mathcal{T}\!op consisting of a point \ast and the unit interval II.

Notation 3.6.

Following the standard convention we denote [𝒲1]\mathcal{M}[\mathscr{W}^{-1}] by Ho\mathop{\rm Ho}\nolimits\mathcal{M} if the class 𝒲\mathscr{W} has been specified.
A pair (Q,ε:QId)(Q,\ {\varepsilon:Q\to\mathop{\rm Id}\nolimits}) respectively (Q,η:IdQ)(Q,\ {\eta:\mathop{\rm Id}\nolimits\to Q}) satisfying the requirements of 3.4 will be called a cofibrant respectively fibrant replacement functor. Each 𝒯op\mathcal{T}\!op-enriched category \mathcal{M} considered in this paper will have a continuous cofibrant replacement functor, and we call the category \mathcal{H}\mathcal{M} the category of QQ-morphisms associated with \mathcal{M}.

Definition 3.7.

A functor Q:Q:\mathcal{M}\to\mathcal{M} together with a natural transformation ε:QId\varepsilon:Q\to\mathop{\rm Id}\nolimits is called a strong cofibrant replacement functor if each ε(M):Q(M)M\varepsilon(M):Q(M)\to M is a weak equivalence and p:(QA,B)(QA,C)p_{\ast}:\mathcal{M}(QA,B)\to\mathcal{M}(QA,C) is a homotopy equivalence whenever p:BCp:B\to C is a weak equivalence.

Clearly, a strong cofibrant replacement functor is a cofibrant replacement functor.

3.8.

Examples:

  1. 1.

    Let 𝒲on\mathscr{W}\subset\mathcal{M}on be the class of weak equivalences in the sense of 2.6. Then WV:ononWV:\mathcal{M}on\to\mathcal{M}on together with WVMε(VM)VMq(M)MWVM\xrightarrow{\varepsilon(VM)}VM\xrightarrow{q(M)}M is a strong cofibrant replacement functor, and the QQ-morphisms are the huhhuh-morphisms. This follows from informations in 2.2, 2.7, 2.10, and 2.12.

  2. 2.

    Let 𝒲onw\mathscr{W}\subset\mathcal{M}on^{w} be again the class of weak equivalences. Then W:onwonwW:\mathcal{M}on^{w}\to\mathcal{M}on^{w} together with ε:WId\varepsilon:W\to\mathop{\rm Id}\nolimits is a strong cofibrant replacement functor, and the QQ-morphisms are the uhuh-morphisms. The required information is obtained from 2.2, 2.7, and 2.12.

  3. 3.

    Let 𝒲𝒮gp\mathscr{W}\subset\mathop{\rm\mathcal{S}gp} be the class of weak equivalences. Then W¯:𝒮gp𝒮gp\overline{W}:\mathop{\rm\mathcal{S}gp}\to\mathop{\rm\mathcal{S}gp} together with ε¯:W¯Id\overline{\varepsilon}:\overline{W}\to\mathop{\rm Id}\nolimits is a strong cofibrant replacement functor, and the QQ-morphisms are the hh-morphisms by informations from 2.2 and 2.7.

  4. 4.

    Let 𝒲𝒯op\mathscr{W}\subset{\mathcal{T}\!op}^{\ast} be the class of based maps which are (not necessarily based) homotopy equivalences. Then Vt:𝒯op𝒯opV^{t}:{\mathcal{T}\!op}^{\ast}\to{\mathcal{T}\!op}^{\ast} together with q:VtIdq:V^{t}\to\mathop{\rm Id}\nolimits is a strong cofibrant replacement functor by the lemma below, the proof of which we leave as an exercise.

  5. 5.

    Let 𝒲𝒯opw\mathscr{W}\subset{\mathcal{T}\!op}^{w} be the class of homotopy equivalences. Then Id:𝒯opw𝒯opw\mathop{\rm Id}\nolimits:{\mathcal{T}\!op}^{w}\to{\mathcal{T}\!op}^{w} is a strong cofibrant replacement functor and each map is a QQ-morphism.

Lemma 3.9.

Let AA be a well-pointed space and p:XYp:X\to Y a map in 𝒯op{\mathcal{T}\!op}^{\ast} which is a not necessarily based homotopy equivalence. Then

p:𝒯op(A,X)𝒯op(A,Y)p_{\ast}:{\mathcal{T}\!op}^{\ast}(A,X)\to{\mathcal{T}\!op}^{\ast}(A,Y)

is a homotopy equivalence in 𝒯op\mathcal{T}\!op.

Proposition 3.10.

The localizations of the categories of 3.8 with respect to their weak equivalences exist.

Proof.

We apply 3.4 and 3.5. We have to show that our categories are tensored over the full subcategory of 𝒯op\mathcal{T}\!op consisting of a point \ast and the unit interval II, the other assumptions of 3.4 have been verified above.
We already know that on\mathcal{M}on and 𝒮gp\mathop{\rm\mathcal{S}gp} are tensored over 𝒯op\mathcal{T}\!op. The category 𝒯op{\mathcal{T}\!op}^{\ast} is tensored over itself by the smash product and hence also tensored over 𝒯op\mathcal{T}\!op. For the Examples 3.8.2 and 3.8.5 it suffices to know that for any object MM in the category the tensor IMI\otimes M is well-pointed (recall MM\ast\otimes M\cong M). This is well known for 𝒯opw{\mathcal{T}\!op}^{w} and holds for onw\mathcal{M}on^{w} by [19, Prop. 7.8]. ∎

Definition 3.11.

Let \mathcal{M} be a category and 𝒲\mathscr{W} a class of morphisms in \mathcal{M} such that [𝒲1]\mathcal{M}[\mathscr{W}^{-1}] exists. Let F:𝒟F:\mathcal{M}\to\mathcal{D} be a functor. A functor 𝐋F:[𝒲1]𝒟\mathop{\bf L}F:\mathcal{M}[\mathscr{W}^{-1}]\to\mathcal{D} together with a natural transformation τ:𝐋FγF\tau:\mathop{\bf L}F\circ\gamma\to F is called left derived functor of FF, if given any functor T:[𝒲1]𝒟T:\mathcal{M}[\mathscr{W}^{-1}]\to\mathcal{D} and natural transformation σ:TγF\sigma:T\circ\gamma\to F, there is a unique natural transformation ρ:T𝐋F\rho:T\to\mathop{\bf L}F such that σ=τ(ργ)\sigma=\tau\circ(\rho\ast\gamma).
Dually, a functor 𝐑F:[𝒲1]𝒟\mathop{\bf R}F:\mathcal{M}[\mathscr{W}^{-1}]\to\mathcal{D} together with a natural transformation μ:F𝐑Fγ\mu:F\to\mathop{\bf R}F\circ\gamma is called right derived functor of FF, if given any functor G:[𝒲1]𝒟G:\mathcal{M}[\mathscr{W}^{-1}]\to\mathcal{D} and natural transformation ν:FGγ\nu:F\to G\circ\gamma, there is a unique natural transformation ξ:𝐑FG\xi:\mathop{\bf R}F\to G such that (ξγ)μ(\xi\ast\gamma)\circ\mu.

Remark 3.12.

(1) A left or right derived functor is unique up to natural isomorphism if it exists.
(2) If F:𝒟F:\mathcal{M}\to\mathcal{D} maps weak equivalences to isomorphisms, then the induced functor F¯:[𝒲1]𝒟\overline{F}:\mathcal{M}[\mathscr{W}^{-1}]\to\mathcal{D} is the right and left derived functor of FF.

Proposition 3.13.

Let \mathcal{M} be as in Proposition 3.4, and let F:F:\mathcal{M}\to\mathcal{B} be a functor which maps homotopy equivalences to isomorphisms. Then 𝐋F:[𝒲1]\mathop{\bf L}F:\mathcal{M}[\mathscr{W}^{-1}]\to\mathcal{B} exists if \mathcal{M} has a cofibrant replacement functor, and 𝐑F:[𝒲1]\mathop{\bf R}F:\mathcal{M}[\mathscr{W}^{-1}]\to\mathcal{B} exists if \mathcal{M} has a fibrant replacement functor. In both cases the derived functor is induced by FQ:F\circ Q:\mathcal{M}\to\mathcal{B}.

Proof.

The proof is the same as in the case of a model category (e.g. see [13, 8.4.]). ∎

Let F:F:\mathcal{M}\to\mathcal{B} be a functor between 𝒯op\mathcal{T}\!op-enriched categories admitting cofibrant replacement functors Q:Q_{\mathcal{M}}:\mathcal{M}\to\mathcal{M} and Q:Q_{\mathcal{B}}:\mathcal{B}\to\mathcal{B}. Proposition 3.13 motivates the introduction of the functor

3.14.

F:F^{\mathcal{H}}:\mathcal{H}\mathcal{M}\to\mathcal{H}\mathcal{B}

defined on objects by F(X)=F(QX)F^{\mathcal{H}}(X)=F(Q_{\mathcal{M}}X) and on morphisms by

F:(QX,QY)QF(QFQX,QFQY).F^{\mathcal{H}}:\mathcal{M}(Q_{\mathcal{M}}X,Q_{\mathcal{M}}Y)\xrightarrow{Q_{\mathcal{B}}\circ F}\mathcal{B}(Q_{\mathcal{B}}FQ_{\mathcal{M}}X,Q_{\mathcal{B}}FQ_{\mathcal{M}}Y).

If FF preserves homotopy equivalences, e.g. if FF is continuous, and π:π\pi_{\mathcal{B}}:\mathcal{B}\to\pi\mathcal{B} is the canonical functor, then πF\pi_{\mathcal{B}}\circ F^{\mathcal{H}} induces the left derived functor

HoF:HoHo\mathop{\rm Ho}\nolimits F:\mathop{\rm Ho}\nolimits\mathcal{M}\to\mathop{\rm Ho}\nolimits\mathcal{B}

of πF\pi_{\mathcal{B}}\circ F. Following model category terminology, we call HoF\mathop{\rm Ho}\nolimits F the total left derived functor of FF.

One of the objectives of this paper is to show that the classifying space functor and the Moore loop space functor induce an adjoint derived pair (see Theorem 4.6 below). This is the path-component version of the more general result (Theorem 4.5 below) that

B:on\textstyle{B^{\mathcal{H}}:\mathcal{H}\mathcal{M}on\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒯op:Ω\textstyle{{\mathcal{T}\!op}^{\ast}:\Omega^{\prime\mathcal{H}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

are a homotopically adjoint pair. To make this last statement precise we need some preparations.

Definition 3.15.

Let 𝒜\mathcal{A} and \mathcal{B} be topologically enriched categories. A functor F:𝒜F:\mathcal{A}\to\mathcal{B} is called continuous if

F:𝒜(A,B)(FA,FB)F:\mathcal{A}(A,B)\longrightarrow\mathcal{B}(FA,FB)

is continuous for all AA and BB in 𝒜\mathcal{A}.

If F,G:𝒜F,G:\mathcal{A}\to\mathcal{B} are continuous functors, a collection of morphisms {α(A):FAGA;Aob𝒜}\{\alpha(A):FA\to GA;\;A\in\mathop{\rm ob}\nolimits\mathcal{A}\} is called a natural transformation up to homotopy if the diagram

𝒜(A,B)\textstyle{\mathcal{A}(A,B)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F\scriptstyle{F}G\scriptstyle{G}(FA,FB)\textstyle{\mathcal{B}(FA,FB)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α(B)\scriptstyle{\alpha(B)_{\ast}}(GA,GB)\textstyle{\mathcal{B}(GA,GB)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α(A)\scriptstyle{\alpha(A)^{\ast}}(FA,GB)\textstyle{\mathcal{B}(FA,GB)}

is homotopy commutative.

A pair of continuous functors

F:𝒜:GF:\mathcal{A}\leftrightarrows\mathcal{B}:G

is called a homotopy adjoint pair if there is a natural transformation up to homotopy

α(A,X):(FA,X)𝒜(A,GX)\alpha(A,X):\mathcal{B}(FA,X)\to\mathcal{A}(A,GX)

such that each α(A,X)\alpha(A,X) is a homotopy equivalence. The homotopy equivalences are called the homotopy adjunctions.

Just as the usual notion of adjunction is equivalently encoded by the concepts of unit and counit, Proposition 3.18 below describes how a homotopy adjunction is specified by a homotopy unit and a homotopy counit.

Observe that we have chosen a strong form of a natural transformation α:FG\alpha:F\to G up to homotopy: for each morphism f:ABf:A\to B in 𝒜\mathcal{A} we have a square

FA\textstyle{FA\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α(A)\scriptstyle{\alpha(A)}Ff\scriptstyle{Ff}GA\textstyle{GA\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gf\scriptstyle{Gf}FB\textstyle{FB\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α(B)\scriptstyle{\alpha(B)}GB\textstyle{GB}

commuting up to a homotopy H(f)H(f) which is continuous in ff.

The proofs of the following two lemmas are easy exercises.

Lemma 3.16.

Let S,T,U:𝒜S,T,U:\mathcal{A}\to\mathcal{B} be continuous functors of topologically enriched categories.

(1) Each natural transformation α:ST\alpha:S\to T is a natural transformation up to homotopy.

(2) If ε:ST\varepsilon:S\to T and η:TU\eta:T\to U are natural transformations up to homotopy, then ηε:SU\eta\circ\varepsilon:S\to U is one.

(3) Let ε:ST\varepsilon:S\to T be a natural transformation up to homotopy such that each ε(A)\varepsilon(A) is a homotopy equivalence. Choose a homotopy inverse η(A)\eta(A) of ε(A)\varepsilon(A) for each AA in 𝒜\mathcal{A}. Then the η(A)\eta(A) form a natural transformation η:TS\eta:T\to S up to homotopy. \Box

Lemma 3.17.

Let S,T,U,V:𝒜S,T,U,V:\mathcal{A}\to\mathcal{B} be continuous functors of topologically enriched categories, and let ε:ST\varepsilon:S\to T and η:UV\eta:U\to V be natural transformations up to homotopy.

(1) Let F,G:𝒜op×𝒜𝒯opF,G:\mathcal{A}^{\mathop{\rm op}\nolimits}\times\mathcal{A}\to\mathcal{T}\!op be defined by F(A,B)=𝒜(A,B)F(A,B)=\mathcal{A}(A,B) and G(A,B)=(TA,TB)G(A,B)=\mathcal{B}(TA,TB). Then

τ(A,B):𝒜(A,B)𝑇(TA,TB)\tau(A,B):\mathcal{A}(A,B)\xrightarrow{T}\mathcal{B}(TA,TB)

is a natural transformation from FF to GG.

(2) Let F,G:𝒜op×𝒜𝒯opF,G:\mathcal{A}^{\mathop{\rm op}\nolimits}\times\mathcal{A}\to\mathcal{T}\!op be defined by F(A,B)=(VA,SB)F(A,B)=\mathcal{B}(VA,SB) and G(A,B)=(UA,TB)G(A,B)=\mathcal{B}(UA,TB). Then

α(A,B):(VA,SB)ε(B)η(A)(UA,TB)\alpha(A,B):\mathcal{B}(VA,SB)\xrightarrow{\varepsilon(B)_{\ast}\circ\eta(A)^{\ast}}\mathcal{B}(UA,TB)

is a natural transformation from FF to GG up to homotopy. \Box

Proposition 3.18.

Let F:𝒜:GF:\mathcal{A}\leftrightarrows\mathcal{B}:G be a pair of continuous functors of topologically enriched categories. Suppose there are natural transformations up homotopy

μ(A):AGF(A)andη(X):FG(X)X\mu(A):A\to GF(A)\qquad\textrm{and}\qquad\eta(X):FG(X)\to X

such that

G(η(X))μ(GX)idGXandη(FA)(F(μ(A))idFA.G(\eta(X))\circ\mu(GX)\simeq\mathop{\rm id}\nolimits_{GX}\qquad\textrm{and}\qquad\eta(FA)\circ(F(\mu(A))\simeq\mathop{\rm id}\nolimits_{FA}.

Then FF and GG are a homotopy adjoint pair. (We call μ:IdGF\mu:\mathop{\rm Id}\nolimits\to GF the homotopy unit and η:FGId\eta:FG\to\mathop{\rm Id}\nolimits the homotopy counit of the resulting homotopy adjunction.)

Proof.

We define

α(A,X):(FA,X)𝐺𝒜(GFA,GX)μ(A)𝒜(A,GX)\alpha(A,X):\mathcal{B}(FA,X)\xrightarrow{G}\mathcal{A}(GFA,GX)\xrightarrow{\mu(A)^{\ast}}\mathcal{A}(A,GX)

and

β(A,X):𝒜(A,GX)𝐹(FA,FGX)η(X)(FA,X).\beta(A,X):\mathcal{A}(A,GX)\xrightarrow{F}\mathcal{B}(FA,FGX)\xrightarrow{\eta(X)_{\ast}}\mathcal{B}(FA,X).

By 3.17 both are natural transformations up to homotopy. The following diagram shows that β(A,X)α(A,X)id\beta(A,X)\circ\alpha(A,X)\simeq\mathop{\rm id}\nolimits.

𝒜(GFA,GX)\textstyle{\mathcal{A}(GFA,GX)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F\scriptstyle{F}μ(A)\scriptstyle{\mu(A)^{\ast}}𝒜(A,GX)\textstyle{\mathcal{A}(A,GX)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F\scriptstyle{F}II(FA,X)\textstyle{\mathcal{B}(FA,X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}G\scriptstyle{G}η(FA)\scriptstyle{\eta(FA)^{\ast}}I(FGFA,FGX)\textstyle{\mathcal{B}(FGFA,FGX)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(Fμ(A))\scriptstyle{(F\mu(A))^{\ast}}η(X)\scriptstyle{\eta(X)_{\ast}}(FA,FGX)\textstyle{\mathcal{B}(FA,FGX)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}η(X)\scriptstyle{\eta(X)_{\ast}}III(FGFA,X)\textstyle{\mathcal{B}(FGFA,X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(Fμ(A))\scriptstyle{(F\mu(A))^{\ast}}(FA,X)\textstyle{\mathcal{B}(FA,X)}

The squares II and III commute and square I commutes up to homotopy, and (Fμ(A))η(FA)id(F\mu(A))^{\ast}\circ\eta(FA)^{\ast}\simeq\mathop{\rm id}\nolimits by assumption.
The proof that α(A,X)β(A,X)id\alpha(A,X)\circ\beta(A,X)\simeq\mathop{\rm id}\nolimits is dual. ∎

Definition 3.19.

A homotopy adjunction F:𝒜:GF:\mathcal{A}\leftrightarrows\mathcal{B}:G is called natural if there is a natural homotopy equivalence

β(A,X):𝒜(A,GX)(FA,X)\beta(A,X):\mathcal{A}(A,GX)\to\mathcal{B}(FA,X)

and conatural if there is a natural homotopy equivalence

α(A,X):(FA,X)𝒜(A,GX)\alpha(A,X):\mathcal{B}(FA,X)\to\mathcal{A}(A,GX)

(because in this case there is a natural homotopy unit, respectively, a natural homotopy counit).

4 The classifying space and the Moore loop space functor

4.1.

The 2-sided bar construction: Let 𝒞\mathcal{C} be a small topologically enriched category, XX a 𝒞op\mathcal{C}^{op}-diagram and YY a 𝒞\mathcal{C}-diagram in 𝒯op\mathcal{T}\!op. We define a simplicial space B(X,𝒞,Y)B_{\bullet}(X,\mathcal{C},Y) by

B0(X,𝒞,Y)=A𝒞X(A)×Y(A)Bn(X,𝒞,Y)=A,B𝒞X(B)×𝒞n(A,B)×Y(A)for n>0,\begin{array}[]{rcl}B_{0}(X,\mathcal{C},Y)&=&\coprod_{A\in\mathcal{C}}X(A)\times Y(A)\\ B_{n}(X,\mathcal{C},Y)&=&\coprod_{A,B\in\mathcal{C}}X(B)\times\mathcal{C}_{n}(A,B)\times Y(A)\quad\textrm{for }n>0,\end{array}

where 𝒞n(A,B)\mathcal{C}_{n}(A,B) is the space of all composable nn-tuples of morphisms (f1,,fn)(f_{1},\ldots,f_{n}) such that source(fn)=A(f_{n})\ =\ A and target(f1)=B(f_{1})\ =\ B, with boundary and degeneracy maps given by

di(x,f1,,fn,y)=(X(f1)(x),f2,,fn,y)i=0di(x,f1,,fn,y)=(x,f1,,fifi+1,,fn,y)0<i<ndi(x,f1,,fn,y)=(x,f1,,fn1,Y(fn)(y))i=nsi(x,f1,,fn,y)=(x,f1,,fi,id,fi+1,,fn,y)0in\begin{array}[]{ll}d^{i}(x,f_{1},\ldots,f_{n},y)=(X(f_{1})(x),f_{2},\cdots,f_{n},y)&i=0\\ d^{i}(x,f_{1},\ldots,f_{n},y)=(x,f_{1},\ldots,f_{i}\circ f_{i+1},\ldots,f_{n},y)&0<i<n\\ d^{i}(x,f_{1},\ldots,f_{n},y)=(x,f_{1},\ldots,f_{n-1},Y(f_{n})(y))&i=n\\ s^{i}(x,f_{1},\ldots,f_{n},y)=(x,f_{1},\ldots,f_{i},\mathop{\rm id}\nolimits,f_{i+1},\ldots,f_{n},y)&0\leq i\leq n\end{array}

Let B(X,𝒞,Y)=|B(X,𝒞,Y)|B(X,\mathcal{C},Y)=|B_{\bullet}(X,\mathcal{C},Y)| be its topological realization.

We consider a topological monoid as a topologically enriched category with one object and define the classifying space functor

B:on𝒯opB:\mathcal{M}on\longrightarrow{\mathcal{T}\!op}^{\ast}

by BM=B(,M,)BM=B(\ast,M,\ast). Since BMBM is well-pointed if MM is, the classifying space functor is a functor of pairs

B:(on,onw)(𝒯op,𝒯opw).B:(\mathcal{M}on,\mathcal{M}on^{w})\to({\mathcal{T}\!op}^{\ast},{\mathcal{T}\!op}^{w}).
4.2.

We will also work with the variant

B~:on𝒯op\widetilde{B}:\mathcal{M}on\longrightarrow{\mathcal{T}\!op}^{\ast}

where the topological realization of B(,M,)B_{\bullet}(\ast,M,\ast) is replaced by the fat realization which disregards degeneracies. Since the fat realization does not make use of identities the functor B~\widetilde{B} extends to 𝒮gp\mathop{\rm\mathcal{S}gp}; moreover, B~G\widetilde{B}G is well-pointed for any semigroup GG so that

B~:𝒮gp𝒯opw.\widetilde{B}:\mathop{\rm\mathcal{S}gp}\to{\mathcal{T}\!op}^{w}.

By construction, there is a natural homeomorphism B~(G)B(G+)\widetilde{B}(G)\cong B(G_{+}) for semigroups GG, and the diagram

B~(M)\textstyle{\widetilde{B}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}p(M)\scriptstyle{p(M)}B(M+)\textstyle{B(M_{+})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B(κ(M))\scriptstyle{B(\kappa(M))}B(M)\textstyle{B(M)}

commutes for monoids MM, where κ:M+M\kappa:M_{+}\to M is the counit of the adjunction 2.5 and p:B~Bp:\widetilde{B}\to B is the natural projection.

It is well-known that p(M):B~(M)B(M)p(M):\widetilde{B}(M)\to B(M) and hence B(κ(M)):B(M+)B(M)B(\kappa(M)):B(M_{+})\to B(M) are homotopy equivalences if MM is well-pointed.

4.3.

The Moore path and loop space: Let XX be a (not necessarily based) space. The Moore path space of XX is the subspace Path(X)X+×+\mathop{\rm Path}(X)\subset X^{\mathbb{R}_{+}}\times\mathbb{R}_{+} consisting of all pairs (w,r)(w,r) such that w(t)=w(r)w(t)=w(r) for all trt\geq r. We call rr the length of ww and denote it by r=l(w)r=l(w).
For two paths (w1,r1)(w_{1},r_{1}) and (w2,r2)(w_{2},r_{2}) with (w1)(r1)=(w2)(0)(w_{1})(r_{1})=(w_{2})(0) we define path addition by

(w1,r1)+(w2,r2)=(w,r1+r2)(w_{1},r_{1})+(w_{2},r_{2})=(w,r_{1}+r_{2})

with

w(t)={w1(t),0tr1,w2(tr1)r1t.w(t)=\left\{\begin{array}[]{ll}w_{1}(t),&0\leq t\leq r_{1},\\ w_{2}(t-r_{1})&r_{1}\leq t.\end{array}\right.

If (X,)(X,\ast) is a based space, the Moore loop space Ω(X)Path(X)\Omega^{\prime}(X)\subset\mathop{\rm Path}(X) is the subspace of all pairs (w,r)(w,r) with (w)(r)=(w)(0)=(w)(r)=(w)(0)=\ast. Path addition defines a monoid structure on ΩX\Omega^{\prime}X with (c,0)(c,0) as unit, where c:+Xc:\mathbb{R}_{+}\to X is the constant map to \ast. The usual loop space ΩX\Omega X is embedded in Ω(X)\Omega^{\prime}(X) as a deformation retract.

It follows from [9, (11.3)] that Ω(X)\Omega^{\prime}(X) is well-pointed if XX is. Hence Ω\Omega^{\prime} defines a functor of pairs

Ω:(𝒯op,𝒯opw)(on,onw).\Omega^{\prime}:({\mathcal{T}\!op}^{\ast},{\mathcal{T}\!op}^{w})\to(\mathcal{M}on,\mathcal{M}on^{w}).

Following 3.14 we have pairs of continuous functors

B:on𝒯op:ΩB^{\mathcal{H}}:\mathcal{H}\mathcal{M}on\leftrightarrows\mathcal{H}{\mathcal{T}\!op}^{\ast}:\Omega^{\prime\mathcal{H}}

and

Bw:onw𝒯opw=𝒯opw:Ωw.B^{w{\mathcal{H}}}:\mathcal{H}\mathcal{M}on^{w}\leftrightarrows\mathcal{H}{\mathcal{T}\!op}^{w}={\mathcal{T}\!op}^{w}:\Omega^{\prime w{\mathcal{H}}}.

We shall prove

Theorem 4.4.

The functors

Bw:onw\textstyle{B^{w{\mathcal{H}}}:\mathcal{H}\mathcal{M}on^{w}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒯opw:Ωw\textstyle{{\mathcal{T}\!op}^{w}:\Omega^{\prime w{\mathcal{H}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

are a conatural homotopically adjoint pair: There is a continuous natural map

λ(WM,X):on(WM,WΩX)𝒯op(BWM,X)\lambda(WM,X):\mathcal{M}on(WM,W\Omega^{\prime}X)\longrightarrow{\mathcal{T}\!op}^{\ast}(BWM,X)

which is a homotopy equivalence.

As an immediate consequence we obtain

Theorem 4.5.

The functors

B:on\textstyle{B^{\mathcal{H}}:\mathcal{H}\mathcal{M}on\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒯op:Ω\textstyle{\mathcal{H}{\mathcal{T}\!op}^{\ast}:\Omega^{\prime\mathcal{H}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

are a conatural homotopically adjoint pair: There is a continuous natural map

λ(WVM,VtX):on(WVM,WVΩVtX)𝒯op(VtBWVM,VtX)\lambda(WVM,V^{t}X):\mathcal{M}on(WVM,WV\Omega^{\prime}V^{t}X)\longrightarrow{\mathcal{T}\!op}^{\ast}(V^{t}BWVM,V^{t}X)

which is a homotopy equivalence.

Proof.

Replacing MM by VMVM and XX by VtXV^{t}X in Proposition 4.4 we obtain a natural homotopy equivalence

on(WVM,WVΩVtX)𝒯op(BWVM,VtX).\mathcal{M}on(WVM,WV\Omega^{\prime}V^{t}X)\xrightarrow{\simeq}{\mathcal{T}\!op}^{\ast}(BWVM,V^{t}X).

Since BWVMBWVM is well-pointed the natural map q(BWVM):VtBWVMBWVMq(BWVM):V^{t}BWVM\to BWVM is a based homotopy equivalence inducing a natural homotopy equivalence

q(BWVM):𝒯op(BWVM,VtX)𝒯op(VtBWVM,VtX).q(BWVM)^{\ast}:{\mathcal{T}\!op}^{\ast}(BWVM,V^{t}X)\to{\mathcal{T}\!op}^{\ast}(V^{t}BWVM,V^{t}X).

Passing to homotopy classes (see 3.4) we obtain

Theorem 4.6.

The functors

HoB:HoonHo𝒯op:HoΩ\mathop{\rm Ho}\nolimits B:\mathop{\rm Ho}\nolimits\mathcal{M}on\leftrightarrows\mathop{\rm Ho}\nolimits{\mathcal{T}\!op}^{\ast}:\mathop{\rm Ho}\nolimits\Omega^{\prime}

are an adjoint pair. Moreover, HoB\mathop{\rm Ho}\nolimits B is the left derived of γ𝒯opB\gamma_{{\mathcal{T}\!op}^{\ast}}\circ B and HoΩ\mathop{\rm Ho}\nolimits\Omega^{\prime} the left derived of γonΩ\gamma_{\mathcal{M}on}\circ\Omega^{\prime}.

Proof.

This follows from our explicit description of the localizations and the derived functors in Section 3. ∎

The rest of this Section is devoted to the proof of Theorem 4.4. By 3.18 it suffices to construct a homotopy unit μ:IdonwΩwBw\mu:\mathop{\rm Id}\nolimits_{\mathcal{H}\mathcal{M}on^{w}}\to\Omega^{\prime w{\mathcal{H}}}B^{w{\mathcal{H}}} and a homotopy counit η:BwΩwId𝒯opw\eta:B^{w{\mathcal{H}}}\Omega^{\prime w{\mathcal{H}}}\to\mathop{\rm Id}\nolimits_{\mathcal{H}{\mathcal{T}\!op}^{w}}. Then λ(WM,X)\lambda(WM,X) is the composite

on(WM,WΩX)𝐵𝒯op(BWM,BWΩX)η(X)𝒯op(BWM,X).\mathcal{M}on(WM,W\Omega^{\prime}X)\xrightarrow{B}{\mathcal{T}\!op}^{\ast}(BWM,BW\Omega^{\prime}X)\xrightarrow{\eta(X)_{\ast}}{\mathcal{T}\!op}^{\ast}(BWM,X).
4.7.

This means, we have to construct continuous homomorphisms

μ(WM):WMWΩBWM\mu(WM):WM\longrightarrow W\Omega^{\prime}BWM

which constitute a natural transformation up to homotopy with respect to homomorphisms WMWNWM\to WN, and a natural transformation

η(X):BWΩXX,\eta(X):BW\Omega^{\prime}X\longrightarrow X,

such that
(1) WΩη(X)μ(WΩX)idWΩXW\Omega^{\prime}\eta(X)\circ\mu(W\Omega^{\prime}X)\simeq\mathop{\rm id}\nolimits_{W\Omega^{\prime}X} in onw\mathcal{M}on^{w} and
(2) η(BWM)Bμ(WM)idBWM\eta(BWM)\circ B\mu(WM)\simeq\mathop{\rm id}\nolimits_{BWM} in 𝒯opw{\mathcal{T}\!op}^{w}.
(For λ\lambda to be a natural transformation we need η\eta to be a natural transformation.)

4.8.

The homotopy counit: Let XX be a based space and let

Δn={(t0,,tn)n+1;i=0nti=1,ti0 for all i}\Delta^{n}=\{(t_{0},\ldots,t_{n})\in\mathbb{R}^{n+1};\quad\sum\limits^{n}_{i=0}t_{i}=1,\;t_{i}\geq 0\textrm{ for all }i\}

denote the standard nn-simplex. The evaluation map

ev(X):BΩX=(n0(ΩX)n×Δn)/X\mathop{\rm ev}\nolimits(X):B\Omega^{\prime}X=\left(\coprod\limits_{n\geq 0}(\Omega^{\prime}X)^{n}\times\Delta^{n}\right)/\sim\ \longrightarrow X

is defined by

ev(X)((w1,,wn)(t0,,tn))=(w1++wn)(i=1ntij=1il(wj))\mathop{\rm ev}\nolimits(X)((w_{1},\ldots,w_{n})(t_{0},\ldots,t_{n}))=(w_{1}+\ldots+w_{n})\left(\sum\limits^{n}_{i=1}t_{i}\cdot\sum\limits^{i}_{j=1}l(w_{j})\right)

where l(wj)l(w_{j}) is the length of wjw_{j}.

The homotopy counit η\eta is the natural map

η(X):BWΩX\textstyle{\eta(X):BW\Omega^{\prime}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bε(ΩX)\scriptstyle{B\varepsilon(\Omega^{\prime}X)}BΩX\textstyle{B\Omega^{\prime}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ev(X)\scriptstyle{\mathop{\rm ev}\nolimits(X)}X.\textstyle{X.}
4.9.

The homotopy unit: For a monoid MM let EMEM denote the 2-sided bar construction B(M,M,)B(M,M,\ast). Then

z(x0,x1,,xn)=(zx0,x1,,xn)z\cdot(x_{0},x_{1},\ldots,x_{n})=(z\cdot x_{0},x_{1},\ldots,x_{n})

defines a left MM-action on the simplicial space B(M,M,)B_{\bullet}(M,M,\ast) and hence on EMEM.

Let P(EM)P(EM) denote the space of Moore paths in EMEM starting at the base-point (e)(e) in the 0-skeleton MM of EMEM. The endpoint projection

P(EM)EMP(EM)\longrightarrow EM

is known to be a fibration. Moreover, it is a homotopy equivalence because P(EM)P(EM) and EMEM are contractible. Let P(EM,M)P(EM,M) be the pullback

P(EM,M)\textstyle{P(EM,M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π(M)\scriptstyle{\pi(M)}P(EM)\textstyle{P(EM)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i\scriptstyle{i}EM\textstyle{EM}

where ii is the inclusion of the 0-skeleton, i.e. P(EM,M)P(EM,M) is the space of Moore paths in EMEM starting at (e)(e) and ending in MM. Then π(M)\pi(M) is a fibration and a homotopy equivalence. We define a monoid structure \oplus in P(EM,M)P(EM,M) by

w1w2=w1+xw2w_{1}\oplus w_{2}=w_{1}+x\cdot w_{2}

where ++ is the usual path addition, xMx\in M is the endpoint of w1w_{1}, and xw2x\cdot w_{2} is the path txw2(t)t\mapsto x\cdot w_{2}(t). Then π(M):P(EM,M)M\pi(M):P(EM,M)\to M is a homomorphism and hence a weak equivalence of monoids.

Factoring out the operation of MM on EMEM we obtain a projection

EMBMEM\to BM

inducing a homomorphism

ρ(M):(P(EM,M),)(ΩBM,+).\rho^{\prime}(M):(P(EM,M),\oplus)\longrightarrow(\Omega^{\prime}BM,+).

Since we do not know whether or not (P(EM,M))(P(EM,M)) is well-pointed we apply the whiskering process to it and obtain a homomorphism

ρ(M):V(P(EM,M),)q((P(EM,M))(P(EM,M),)ρ(M)(ΩBM,+).\rho(M):V(P(EM,M),\oplus)\xrightarrow{q((P(EM,M))}(P(EM,M),\oplus)\xrightarrow{\rho^{\prime}(M)}(\Omega^{\prime}BM,+).

The homomorphism σ(M):WV(P(EM,M)M\sigma(M):WV(P(EM,M)\to M defined by

WVP(EM,M)\textstyle{WVP(EM,M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ε(VP(EM,M))\scriptstyle{\varepsilon(VP(EM,M))}σ(M)\scriptstyle{\sigma(M)}VP(EM,M)\textstyle{VP(EM,M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q(P(EM,M))\scriptstyle{q(P(EM,M))}P(EM,M)\textstyle{P(EM,M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π(M)\scriptstyle{\pi(M)}M\textstyle{M}

is a weak equivalence. All these constructions are functorial in MM and the maps between them are natural in MM. We apply them to WMWM rather than to MM; in particular σ(WM)\sigma(WM) is a homotopy equivalence in onw\mathcal{M}on^{w}.

We choose a homotopy inverse of σ(WM)\sigma(WM) in onw\mathcal{M}on^{w}

ν(WM):WMWVP(EWM,WM),\nu(WM):WM\longrightarrow WVP(EWM,WM),

which is a natural transformation up to homotopy with respect to homomorphisms WMWNWM\to WN by Lemma 3.16.

We define our homotopy unit by

μ(WM):WM\textstyle{\mu(WM):WM\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ν(WM)\scriptstyle{\nu(WM)}WVP(EWM,WM)\textstyle{WVP(EWM,WM)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Wρ(WM)\scriptstyle{W\rho(WM)}WΩBWM,\textstyle{W\Omega^{\prime}BWM,}

which is a natural transformation up to homotopy by Lemma 3.16.

Our verification of the conditions 4.7 depends on an explicit description of an hh-morphism MΩBMM\to\Omega^{\prime}BM defined by a natural homomorphism

ζ(M):W¯(M)ΩBM\zeta^{\prime}(M):\overline{W}(M)\longrightarrow\Omega^{\prime}BM

and the interplay of W¯(M)\overline{W}(M) and WMWM.

We define ζ(M)\zeta^{\prime}(M) as a composite of homomorphisms

W¯(M)\textstyle{\overline{W}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ζ(M)\scriptstyle{\zeta(M)}P(EM,M)\textstyle{P(EM,M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ(M)\scriptstyle{\rho^{\prime}(M)}ΩBM\textstyle{\Omega^{\prime}BM}

The homomorphism ζ(M)\zeta(M) maps the element represented by (x0,t1,,xn)(x_{0},t_{1},\ldots,x_{n}) to the path

v0+v1++vnv_{0}+v_{1}+\ldots+v_{n}

of length t1++tn+1t_{1}+\ldots+t_{n}+1 in the simplex (e,x0,x1,,xn)×Δn+1EM(e,x_{0},x_{1},\ldots,x_{n})\times\Delta^{n+1}\subset EM, where

vk(s)=(e,x0,,xn)×(u0,,un+1)andl(vk)=tk+1v_{k}(s)=(e,x_{0},\ldots,x_{n})\times(u_{0},\ldots,u_{n+1})\quad\textrm{and}\quad l(v_{k})=t_{k+1}

with

ur={(1s)trj=r+1k(1tj)rksr=k+10rk+2u_{r}=\left\{\begin{array}[]{ll}(1-s)\cdot t_{r}\cdot\prod^{k}_{j=r+1}(1-t_{j})&r\leq k\\ s&r=k+1\\ 0&r\geq k+2\end{array}\right.

and the conventions that t0=1t_{0}=1 and tn+1=1t_{n+1}=1.

Observe that ++ is the usual path addition of Moore paths in EMEM and not the monoid structure of P(EM,M)P(EM,M).

Example: (x0,t1,x1,t2,x2)(x_{0},t_{1},x_{1},t_{2},x_{2}) is mapped to the path v0+v1+v2v_{0}+v_{1}+v_{2} of length t1+t2+1t_{1}+t_{2}+1 given by

v1v_{1}v0v_{0}v2v_{2}eex0x1x_{0}x_{1}x0x1x2x_{0}x_{1}x_{2}x0x_{0}t1t_{1}t2t_{2}
Figure 1:
4.10.

By construction, π(M)ζ(M)=ε¯(M)\pi(M)\circ\zeta(M)=\overline{\varepsilon}(M). In particular, ζ(M):W¯MP(EM,M)\zeta(M):\overline{W}M\to P(EM,M) is a weak equivalence of semigroups.

Remark 4.11.

We will show below that ρ(M):P(EM,M)ΩBM\rho^{\prime}(M):P(EM,M)\to\Omega^{\prime}BM is a weak equivalence if MM is grouplike, so that ρ(M)ζ(M)\rho^{\prime}(M)\circ\zeta(M) is an hh-morphism which is a weak equivalence if MM is grouplike. It is well-known that such an hh-morphism exists, but to our knowledge there is no explicit description in the literature.

4.12.

Consider the following diagram

(W¯WM)+\textstyle{(\overline{W}WM)_{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ε¯(WM)+\scriptstyle{\overline{\varepsilon}(WM)^{+}}ζ(WM)+\scriptstyle{\zeta(WM)^{+}}IWM\textstyle{WM\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ν(WM)\scriptstyle{\nu(WM)}id\scriptstyle{\mathop{\rm id}\nolimits}WVP(EWM,WM)\textstyle{WVP(EWM,WM)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ1(WM)\scriptstyle{\sigma_{1}(WM)}σ(WM)\scriptstyle{\sigma(WM)}P(EWM,WM)\textstyle{P(EWM,WM)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π(WM)\scriptstyle{\pi(WM)}WM\textstyle{WM}

where σ1(WM)=q(P(EM,M))ε(VP(EM,M))\sigma_{1}(WM)=q(P(EM,M))\circ\varepsilon(VP(EM,M)) and f+:G+Mf^{+}:G_{+}\to M is the adjoint of the homomorphism f:GMf:G\to M from a semigroup into a monoid. By definition of ν(WM)\nu(WM) and σ(WM)\sigma(WM) the left lower triangle commutes up to homotopy in onw\mathcal{M}on^{w} and the right lower triangle is commutative. Since

π(WM)ζ(WM)=ε¯(WM)π(WM)σ1(WM)ν(WM)ε¯(WM)\pi(WM)\circ\zeta(WM)=\overline{\varepsilon}(WM)\simeq\pi(WM)\circ\sigma_{1}(WM)\circ\nu(WM)\circ\overline{\varepsilon}(WM)

Proposition 2.7 implies that

σ1(WM)ν(WM)ε¯(WM)ζ(WM)in 𝒮gp\sigma_{1}(WM)\circ\nu(WM)\circ\overline{\varepsilon}(WM)\simeq\zeta(WM)\qquad\textrm{in }\ \mathop{\rm\mathcal{S}gp}

which in turn is equivalent to the saying that square I commutes up to homotopy in on\mathcal{M}on.

We are now in the position to prove

Proposition 4.13.

η(BWM)Bμ(WM)idBWM\eta(BWM)\circ B\mu(WM)\simeq\mathop{\rm id}\nolimits_{BWM} in 𝒯opw{\mathcal{T}\!op}^{w}.

This result is a fairly easy consequence of

Lemma 4.14.

The diagram

B~W¯M\textstyle{\widetilde{B}\overline{W}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B~ε¯(M)\scriptstyle{\widetilde{B}\overline{\varepsilon}(M)}B~ζ(M)\scriptstyle{\widetilde{B}\zeta(M)}B~M\textstyle{\widetilde{B}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p(M)\scriptstyle{p(M)}B~P(EM,M)\textstyle{\widetilde{B}P(EM,M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B~ρ(M)\scriptstyle{\widetilde{B}\rho^{\prime}(M)}B~ΩBM\textstyle{\widetilde{B}\Omega^{\prime}BM\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p(ΩBM)\scriptstyle{p(\Omega^{\prime}BM)}BΩBM\textstyle{{B}\Omega^{\prime}BM\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ev(BM)\scriptstyle{\mathop{\rm ev}\nolimits(BM)}BM\textstyle{BM}

commutes up to homotopy.

Proof.

Let f=ev(BM)p(ΩBM)B~ρ(M)B~ζf=\mathop{\rm ev}\nolimits(BM)\circ p(\Omega^{\prime}BM)\circ\widetilde{B}\rho^{\prime}(M)\circ\widetilde{B}\zeta and let g=p(M)B~ε¯(M)g=p(M)\circ\widetilde{B}\overline{\varepsilon}(M). Let z=(z1,,zn)z=(z_{1},\ldots,z_{n}) be an element in (W¯M)n(\overline{W}M)^{n}, so that z×Δnz\times\Delta^{n} is an nn-simplex in B~W¯M\widetilde{B}\overline{W}M. If zj=(xj0,tj1,,xjrj)z_{j}=(x_{j0},t_{j1},\ldots,x_{jr_{j}}), then ff maps z×Δnz\times\Delta^{n} to the image of the path ρ(M)ζ(z1)++ρ(M)ζ(zn)\rho^{\prime}(M)\circ\zeta(z_{1})+\ldots+\rho^{\prime}(M)\circ\zeta(z_{n}) which lies in the simplex

σ=σ(z)=(x10,,x1r1,,xn0,,xnrn)×Δr1++rn+n\sigma=\sigma(z)=(x_{10},\ldots,x_{1r_{1}},\ldots,x_{n0},\ldots,x_{nr_{n}})\times\Delta^{r_{1}+\cdots+r_{n}+n}

in BMBM, while gg maps z×Δnz\times\Delta^{n} identically (modulo possible degenerations) onto the simplex

τ=τ(z)=(x10x1r1,,xn0xnrn)×Δn\tau=\tau(z)=(x_{10}\cdot\ldots\cdot x_{1r_{1}},\ldots,x_{n0}\cdot\ldots\cdot x_{nr_{n}})\times\Delta^{n}

in BMBM, which is a face of σ\sigma. So f|z×Δnf|z\times\Delta^{n} is homotopic to g|z×Δng|z\times\Delta^{n} by a linear homotopy. We call a homotopy from ff to gg admissible if it maps z×Δnz\times\Delta^{n} to σ(z)\sigma(z) throughout the homotopy.

We are going to construct an admissible homotopy H:B~W¯M×IBMH:\widetilde{B}\overline{W}M\times I\to BM from ff to gg by induction on the canonical filtration (B~W¯M)(n)(\widetilde{B}\overline{W}M)^{(n)} of B~W¯M\widetilde{B}\overline{W}M.

(B~W¯M)(0)(\widetilde{B}\overline{W}M)^{(0)} is a point, which is mapped by ff and gg to the base-point. Now suppose that we have constructed an admissible homotopy

H:(B~W¯M)(n1)×IBM.H:(\widetilde{B}\overline{W}M)^{(n-1)}\times I\to BM.

Let z×Δnz\times\Delta^{n} be an nn-simplex in B~W¯M\widetilde{B}\overline{W}M as above. We define

q(z)=q(z1,,zn)=r1++rnq(z)=q(z_{1},\ldots,z_{n})=r_{1}+\cdots+r_{n}\in\mathbb{N}

and we extend HH over (B~W¯M)(n)×I(\widetilde{B}\overline{W}M)^{(n)}\times I by induction on qq.

If q=0q=0, then z=(z1,,zn)z=(z_{1},\ldots,z_{n}) with zj=(xj0)z_{j}=(x_{j0}) for j=1,,nj=1,\ldots,n and σ(z)=τ(z)=(x10,,xn0)×Δn\sigma(z)=\tau(z)=(x_{10},\ldots,x_{n0})\times\Delta^{n}. Hence the space of all nn-simplices z(W¯M)nz\in(\overline{W}M)^{n} with q(z)=0q(z)=0 is MnM^{n}. By induction, we have to find a homotopy

h:Mn×Δn×IMn×Δnh:M^{n}\times\Delta^{n}\times I\to M^{n}\times\Delta^{n}

over MnM^{n} which is already determined on Mn×(Δn×I)M^{n}\times\partial(\Delta^{n}\times I). If bnb_{n} denotes the barycenter of Δn\Delta^{n} we map ((x1,,xn),bn,12)((x_{1},\ldots,x_{n}),b_{n},\frac{1}{2}) to ((x1,,xn),bn)((x_{1},\ldots,x_{n}),b_{n}) and cone off.

If q>0q>0 we have qq coordinates tjkIt_{jk}\in I in zz. So the space of all elements zz with q(z)=qq(z)=q is the union of spaces of the form Mn+q×IqM^{n+q}\times I^{q} which may intersect on their lower faces Mn+q×LIqM^{n+q}\times LI^{q} due to the relations, where LIq={(t1,,tq)Iq; some ti=0}LI^{q}=\{(t_{1},\ldots,t_{q})\in I^{q};\textrm{ some }t_{i}=0\}. So possible intersections are of lower filtration. We have to find a map

h:Mn+q×Iq×Δn×IMn+q×Δn+qh:M^{n+q}\times I^{q}\times\Delta^{n}\times I\to M^{n+q}\times\Delta^{n+q}

over Mn+qM^{n+q} which is already defined on

Mn+q×(LIq×Δn×IIq×(Δn×I)).M^{n+q}\times(LI^{q}\times\Delta^{n}\times I\cup I^{q}\times\partial(\Delta^{n}\times I)).

Since LIqLI^{q} is a strong deformation retract of IqI^{q}, the inclusion

LIq×Δn×IIq×(Δn×I)Iq×Δn×ILI^{q}\times\Delta^{n}\times I\cup I^{q}\times\partial(\Delta^{n}\times I)\subset I^{q}\times\Delta^{n}\times I

is an inclusion of a strong deformation retract. Hence hh exists. ∎

Proof of Proposition 4.13: Since MM is well-pointed, the projection p(M):B~MBMp(M):\widetilde{B}M\to BM is a homotopy equivalence. If h:XYh:X\to Y is a weak equivalence of semigroups, then B~h:B~XB~Y\widetilde{B}h:\widetilde{B}X\to\widetilde{B}Y is a based homotopy equivalence. Hence it suffices to show that

η(BWM)Bμ(WM)p(WM)B~ε¯(WM)p(WM)B~ε¯(WM).\eta(BWM)\circ B\mu(WM)\circ p(WM)\circ\widetilde{B}\overline{\varepsilon}(WM)\simeq p(WM)\circ\widetilde{B}\overline{\varepsilon}(WM).

Now

η(BWM)\displaystyle\eta(BWM) Bμ(WM)p(WM)B~ϵ¯(WM)\displaystyle\circ B\mu(WM)\circ p(WM)\circ\widetilde{B}\bar{\epsilon}(WM)
=\displaystyle= ev(BWM)Bε(ΩBWM)BWρ(WM)Bν(WM)p(WM)B~ε¯(WM),\displaystyle\mathop{\rm ev}\nolimits(BWM)\circ B\varepsilon(\Omega^{\prime}BWM)\circ BW\rho(WM)\circ B\nu(WM)\circ p(WM)\circ\widetilde{B}\bar{\varepsilon}(WM),
since η(BWM)=ev(BWM)Bε(ΩBWM)\eta(BWM)=\mathop{\rm ev}\nolimits(BWM)\circ B\varepsilon(\Omega^{\prime}BWM) and μ(WM)=Wρ(WM)ν(WM)\mu(WM)=W\rho(WM)\circ\nu(WM),
=ev(BWM)Bε(ΩBWM)BWρ(WM)p(WVP(EWM,WM))B~ν(WM)B~ε¯(WM),\displaystyle\begin{split}=&\mathop{\rm ev}\nolimits(BWM)\circ B\varepsilon(\Omega^{\prime}BWM)\circ BW\rho(WM)\circ p(WVP(EWM,WM))\\ &\circ\widetilde{B}\nu(WM)\circ\widetilde{B}\bar{\varepsilon}(WM),\end{split}
by naturality of pp,
=ev(BWM)Bρ(WM)p(VP(EWM,WM))B~ε(VP(EWM,WM))B~ν(WM)B~ε¯(WM),\displaystyle\begin{split}=&\mathop{\rm ev}\nolimits(BWM)\circ B\rho(WM)\circ p(VP(EWM,WM))\circ\widetilde{B}\varepsilon(VP(EWM,WM))\\ &\circ\widetilde{B}\nu(WM)\circ\widetilde{B}\bar{\varepsilon}(WM),\end{split}
by naturality of ε\varepsilon,
=ev(BWM)p(ΩBWM)B~ρ(WM)B~ε(VP(EWM,WM))B~ν(WM)B~ε¯(WM),\displaystyle\begin{split}=&\mathop{\rm ev}\nolimits(BWM)\circ p(\Omega^{\prime}BWM)\circ\widetilde{B}\rho(WM)\circ\widetilde{B}\varepsilon(VP(EWM,WM))\\ &\circ\widetilde{B}\nu(WM)\circ\widetilde{B}\bar{\varepsilon}(WM),\end{split}
by naturality of pp again,
=ev(BWM)p(ΩBWM)B~ρ(WM)B~q(P(EWM,WM))B~ε(VP(EWM,WM))B~ν(WM)B~ε¯(WM),\displaystyle\begin{split}=&\mathop{\rm ev}\nolimits(BWM)\circ p(\Omega^{\prime}BWM)\circ\widetilde{B}\rho^{\prime}(WM)\circ\widetilde{B}q(P(EWM,WM))\\ &\circ\widetilde{B}\varepsilon(VP(EWM,WM))\circ\widetilde{B}\nu(WM)\circ\widetilde{B}\bar{\varepsilon}(WM),\end{split}
by the definition of ρ(WM)\rho(WM),
=ev(BWM)p(ΩBWM)B~ρ(WM)B~σ1(WM)B~ν(WM)B~ε¯(WM),\displaystyle\begin{split}=&\mathop{\rm ev}\nolimits(BWM)\circ p(\Omega^{\prime}BWM)\circ\widetilde{B}\rho^{\prime}(WM)\circ\widetilde{B}\sigma_{1}(WM)\circ\widetilde{B}\nu(WM)\\ &\circ\widetilde{B}\bar{\varepsilon}(WM),\end{split}
by the definition of σ1(WM)\sigma_{1}(WM) from 4.12,
\displaystyle\simeq ev(BWM)p(ΩBWM)B~ρ(WM)B~ζ(WM),\displaystyle\mathop{\rm ev}\nolimits(BWM)\circ p(\Omega^{\prime}BWM)\circ\widetilde{B}\rho^{\prime}(WM)\circ\widetilde{B}\zeta(WM),
by Diagram 4.12,
\displaystyle\simeq p(WM)B~ε(WM),\displaystyle p(WM)\circ\widetilde{B}\varepsilon(WM),

by Lemma 4.14. \Box

Remark 4.15.

If we use the Quillen model structure on 𝒯op\mathcal{T}\!op rather than the Strøm structure we can construct a homotopy unit μ(WM)\mu(WM) and deduce Proposition 4.13 fairly easily from [10, Thm. 7.3] and its proof.

The proof of the first part of 4.7 needs some preparation. Let 𝒥\mathcal{J} denote the category of ordered sets [n]={0<1<<n}[n]=\{0<1<\dots<n\} and order preserving injections, and let 𝒥𝒯op0\mathcal{J}\mathcal{T}\!op_{0} denote the category of all diagrams

X:𝒥op𝒯op,[n]XnX_{\bullet}:\mathcal{J}^{\mathop{\rm op}\nolimits}\to\mathcal{T}\!op,\qquad[n]\mapsto X_{n}

such that X0X_{0} is a single point, i.e. an object in 𝒥𝒯op0\mathcal{J}\mathcal{T}\!op_{0} is a reduced simplicial space without degeneracies. Of lately, such an object is called a reduced semisimplicial space. The usual fat topological realization functor

𝒥𝒯op0𝒯op,XX\mathcal{J}\mathcal{T}\!op_{0}\to{\mathcal{T}\!op}^{\ast},\qquad X_{\bullet}\mapsto\parallel X_{\bullet}\parallel

has a right adjoint, the reduced singular functor

Sing0:𝒯op𝒥𝒯op0,Singn0(Y)=𝒯op((Δn,Δ0n),(Y,))\mathop{\rm Sing}\nolimits_{\bullet}^{0}:{\mathcal{T}\!op}^{\ast}\to\mathcal{J}\mathcal{T}\!op_{0},\qquad\mathop{\rm Sing}\nolimits_{n}^{0}(Y)=\mathcal{T}\!op((\Delta^{n},\Delta^{n}_{0}),(Y,\ast))

where Δ0n\Delta^{n}_{0} is the 0-skeleton of Δn\Delta^{n}. The unit of this adjunction

τ(X):XSing0X\tau_{\bullet}(X_{\bullet}):X_{\bullet}\to\mathop{\rm Sing}\nolimits_{\bullet}^{0}\parallel X_{\bullet}\parallel

sends xXnx\in X_{n} to the singular simplex

ΔnixkXk×ΔkX\Delta^{n}\xrightarrow{i_{x}}\coprod_{k}X_{k}\times\Delta^{k}\rightarrow\parallel X_{\bullet}\parallel

where ixi_{x} is the inclusion of the simplex {x}×Δn\{x\}\times\Delta^{n}. The counit

ev^(Y):Sing0(Y)Y\widehat{\mathop{\rm ev}\nolimits}(Y):\parallel\mathop{\rm Sing}\nolimits_{\bullet}^{0}(Y)\parallel\rightarrow Y

is induced by the evaluation maps Singn0(Y)×ΔnY\mathop{\rm Sing}\nolimits_{n}^{0}(Y)\times\Delta^{n}\to Y. The formula defining our evaluation map of 4.8 defines a natural semisimplicial map

α(Y):NΩYSing0Y\alpha_{\bullet}(Y):N_{\bullet}\Omega^{\prime}Y\rightarrow\mathop{\rm Sing}\nolimits_{\bullet}^{0}Y

where NΩYN_{\bullet}\Omega^{\prime}Y is the semisimplicial nerve of ΩY\Omega^{\prime}Y. Let v0,,vnv_{0},\ldots,v_{n} denote the vertices of Δn\Delta^{n} and let LnΔnL_{n}\subset\Delta^{n} denote the union of the 11-simplexes [vi1,vi][v_{i-1},v_{i}], i=1,,ni=1,\ldots,n. Then LnL_{n} is a strong deformation retract of Δn\Delta^{n}. The composite

(ΩY)n=Nn(ΩY)αn(Y)Singn0(Y)𝑟𝒯op((Ln,Δ0n),(Y,))=(ΩY)n,(\Omega^{\prime}Y)^{n}=N_{n}(\Omega^{\prime}Y)\xrightarrow{\alpha_{n}(Y)}\mathop{\rm Sing}\nolimits_{n}^{0}(Y)\xrightarrow{r}\mathcal{T}\!op((L_{n},\Delta^{n}_{0}),(Y,\ast))=(\Omega Y)^{n},

where rr is the restriction to LnL_{n}, is the map normalizing the loop lengths to 11. In particular, αn(Y)\alpha_{n}(Y) is a homotopy equivalence inducing a homotopy equivalence α(Y)\parallel\alpha_{\bullet}(Y)\parallel. Moreover, the diagram

4.16.
B~ΩY\textstyle{\widetilde{B}\Omega^{\prime}Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p(ΩY)\scriptstyle{p(\Omega^{\prime}Y)}α(Y)\scriptstyle{\parallel\alpha_{\bullet}(Y)\parallel}BΩY\textstyle{B\Omega^{\prime}Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ev(Y)\scriptstyle{\mathop{\rm ev}\nolimits(Y)}Sing0(Y)\textstyle{\parallel\mathop{\rm Sing}\nolimits_{\bullet}^{0}(Y)\parallel\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ev^(Y)\scriptstyle{\widehat{\mathop{\rm ev}\nolimits}(Y)}Y\textstyle{Y}

commutes.

Proposition 4.17.
  1. (1)

    If MM is a grouplike well-pointed monoid, then ρ(M):P(EM,M)ΩBM\rho^{\prime}(M):P(EM,M)\to\Omega^{\prime}BM and hence μ(WM):WMWΩBWM\mu(WM):WM\to W\Omega^{\prime}BWM are weak equivalences.

  2. (2)

    If YY is a well-pointed path-connected Dold space (see Definition 1.7), then ev(Y):BΩYY\mathop{\rm ev}\nolimits(Y):B\Omega^{\prime}Y\to Y is a based homotopy equivalence, and, hence so is

    η(Y):BWΩYBε(ΩY)BΩYev(Y)Y.\eta(Y):BW\Omega^{\prime}Y\xrightarrow{B\varepsilon(\Omega^{\prime}Y)}B\Omega^{\prime}Y\xrightarrow{\mathop{\rm ev}\nolimits(Y)}Y.
  3. (3)

    If YY is a well-pointed space, then Ωev(Y):ΩBΩYΩY\Omega^{\prime}\mathop{\rm ev}\nolimits(Y):\Omega^{\prime}B\Omega^{\prime}Y\to\Omega^{\prime}Y is a weak equivalence. Hence so is Ωη(Y):ΩBWΩYΩY\Omega^{\prime}\eta(Y):\Omega^{\prime}BW\Omega^{\prime}Y\to\Omega^{\prime}Y.

  4. (4)

    If MM is a well-pointed monoid, then Bμ(WM):BWMBWΩBWMB\mu(WM):BWM\to BW\Omega^{\prime}BWM is a homotopy equivalence.

Proof.

(1) The diagram

M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ1(NM)\scriptstyle{\tau_{1}(N_{\bullet}M)}ζι¯(M)\scriptstyle{\zeta\circ\bar{\iota}(M)}ΩB~M\textstyle{\Omega\widetilde{B}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωp(M)\scriptstyle{\Omega p(M)}ΩBM\textstyle{\Omega BM\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i(BM)\scriptstyle{i(BM)}P(EM,M)\textstyle{P(EM,M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ(M)\scriptstyle{\rho^{\prime}(M)}ΩBM\textstyle{\Omega^{\prime}BM}

commutes. Here i(X):ΩXΩXi(X):\Omega X\to\Omega^{\prime}X is the inclusion and ι¯(M):MW¯M\bar{\iota}(M):M\to\overline{W}M the section (see 2.2). It is well known that τ1(NM)\tau_{1}(N_{\bullet}M) is a homotopy equivalence if MM is grouplike (e.g. see [23]). Since p(M),i(BM)p(M),\ i(BM), and ζι¯(M)\zeta\circ\bar{\iota}(M) are homotopy equivalences in 𝒯op\mathcal{T}\!op, so is ρ(M)\rho^{\prime}(M).

(2) In the commutative diagram 4.16 the map p(ΩY)p(\Omega^{\prime}Y) is a homotopy equivalence because ΩY\Omega^{\prime}Y is well-pointed and ev^(Y)\widehat{\mathop{\rm ev}\nolimits}(Y) is a homotopy equivalence by [22, Prop. 5.6].

(3) Consider the following commutative diagram in 𝒥𝒯op0\mathcal{J}\mathcal{T}\!op_{0}

NΩY\textstyle{N_{\bullet}\Omega^{\prime}Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ(NΩY)\scriptstyle{\tau_{\bullet}(N_{\bullet}\Omega^{\prime}Y)}α(Y)\scriptstyle{\alpha_{\bullet}(Y)}Sing0B~ΩY\textstyle{\mathop{\rm Sing}\nolimits_{\bullet}^{0}\widetilde{B}\Omega^{\prime}Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Sing0α(Y)\scriptstyle{\mathop{\rm Sing}\nolimits_{\bullet}^{0}\parallel\alpha_{\bullet}(Y)\parallel}Sing0Y\textstyle{\mathop{\rm Sing}\nolimits_{\bullet}^{0}Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ(Sing0Y)\scriptstyle{\tau_{\bullet}(\mathop{\rm Sing}\nolimits_{\bullet}^{0}Y)}id\scriptstyle{\mathop{\rm id}\nolimits}Sing0Sing0Y\textstyle{\mathop{\rm Sing}\nolimits_{\bullet}^{0}\parallel\mathop{\rm Sing}\nolimits_{\bullet}^{0}Y\parallel\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Sing0ev^(Y)\scriptstyle{\mathop{\rm Sing}\nolimits_{\bullet}^{0}\widehat{\mathop{\rm ev}\nolimits}(Y)}Sing0Y\textstyle{\mathop{\rm Sing}\nolimits_{\bullet}^{0}Y}

Restricting this diagram to degree 0 we obtain a commutative diagram of spaces

ΩY\textstyle{\Omega^{\prime}Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ1(NΩY)\scriptstyle{\tau_{1}(N_{\bullet}\Omega^{\prime}Y)}α1\scriptstyle{\alpha_{1}}ΩB~ΩY\textstyle{\Omega\widetilde{B}\Omega^{\prime}Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωα(Y)\scriptstyle{\Omega\parallel\alpha_{\bullet}(Y)\parallel}Ωp(ΩY)\scriptstyle{\Omega p(\Omega^{\prime}Y)}ΩBΩY\textstyle{\Omega B\Omega^{\prime}Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωev(Y)\scriptstyle{\Omega\mathop{\rm ev}\nolimits(Y)}ΩY\textstyle{\Omega Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ1(Sing0Y)\scriptstyle{\tau_{1}(\mathop{\rm Sing}\nolimits_{\bullet}^{0}Y)}ΩSing0Y\textstyle{\Omega\parallel\mathop{\rm Sing}\nolimits_{\bullet}^{0}Y\parallel\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ωev^(Y)\scriptstyle{\Omega\widehat{\mathop{\rm ev}\nolimits}(Y)}ΩY.\textstyle{\Omega Y.}

Since ΩY\Omega^{\prime}Y is grouplike, τ1(NΩY)\tau_{1}(N_{\bullet}\Omega^{\prime}Y) is a homotopy equivalence. Since YY and hence ΩY\Omega^{\prime}Y is well-pointed, Ωp(ΩY)\Omega p(\Omega^{\prime}Y) is a homotopy equivalence. Since α1\alpha_{1} and α(Y)\parallel\alpha_{\bullet}(Y)\parallel are homotopy equivalences, τ1(Sing0Y)\tau_{1}(\mathop{\rm Sing}\nolimits_{\bullet}^{0}Y) is one. Hence so is Ωev^(Y)\Omega\widehat{\mathop{\rm ev}\nolimits}(Y) and hence also Ωev(Y)\Omega\mathop{\rm ev}\nolimits(Y), which implies the result.

(4) Since BWMBWM is a well-pointed path-connected Dold space by [22, Cor. 5.2] the statement follows from Part (2) and Proposition 4.13. ∎

Proposition 4.18.

WΩη(X)μ(WΩX)idWΩXW\Omega^{\prime}\eta(X)\circ\mu(W\Omega^{\prime}X)\simeq\mathop{\rm id}\nolimits_{W\Omega^{\prime}X} in on\mathcal{M}on.

Proof.

It follows from Proposition 4.13 and the homotopy naturality of μ\mu and η\eta that the following diagram commutes up to homotopy.

WΩX\textstyle{W\Omega^{\prime}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ(WΩX)\scriptstyle{\mu(W\Omega^{\prime}X)}μ(WΩX)\scriptstyle{\mu(W\Omega^{\prime}X)}WΩBWΩX\textstyle{W\Omega^{\prime}BW\Omega^{\prime}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}WΩη(X)\scriptstyle{W\Omega^{\prime}\eta(X)}μ(WΩBWΩX)\scriptstyle{\mu(W\Omega^{\prime}BW\Omega^{\prime}X)}WΩX\textstyle{W\Omega^{\prime}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ(WΩX)\scriptstyle{\mu(W\Omega^{\prime}X)}WΩBWΩX\textstyle{W\Omega^{\prime}BW\Omega^{\prime}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}WΩBμ(WΩX)\scriptstyle{W\Omega^{\prime}B\mu(W\Omega^{\prime}X)}id\scriptstyle{\mathop{\rm id}\nolimits}WΩBWΩBWΩX\textstyle{W\Omega^{\prime}BW\Omega^{\prime}BW\Omega^{\prime}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}WΩBWΩη(X)\scriptstyle{W\Omega^{\prime}BW\Omega^{\prime}\eta(X)}WΩη(BWΩX)\scriptstyle{W\Omega^{\prime}\eta(BW\Omega^{\prime}X)}WΩBWΩX\textstyle{W\Omega^{\prime}BW\Omega^{\prime}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}WΩη(X)\scriptstyle{W\Omega^{\prime}\eta(X)}WΩBWΩX\textstyle{W\Omega^{\prime}BW\Omega^{\prime}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}WΩη(X)\scriptstyle{W\Omega^{\prime}\eta(X)}WΩX\textstyle{W\Omega^{\prime}X}

We obtain

WΩη(X)μ(WΩX)WΩη(X)μ(WΩX)WΩη(X)μ(WΩX).W\Omega^{\prime}\eta(X)\circ\mu(W\Omega^{\prime}X)\circ W\Omega^{\prime}\eta(X)\circ\mu(W\Omega^{\prime}X)\simeq W\Omega^{\prime}\eta(X)\circ\mu(W\Omega^{\prime}X).

Since ΩX\Omega^{\prime}X is grouplike μ(WΩX)\mu(W\Omega^{\prime}X) and WΩη(X)W\Omega^{\prime}\eta(X) are weak equivalences by Proposition 4.17. By Proposition 2.12 both homomorphisms have homotopy inverses in on\mathcal{M}on so that

WΩη(X)μ(WΩX)idWΩXW\Omega^{\prime}\eta(X)\circ\mu(W\Omega^{\prime}X)\simeq\mathop{\rm id}\nolimits_{W\Omega^{\prime}X}

in on\mathcal{M}on. ∎

5 Immediate consequences

The James Construction:

The underlying space functor U:(on,onw)(𝒯op,𝒯opw)U:(\mathcal{M}on,\mathcal{M}on^{w})\to({\mathcal{T}\!op}^{\ast},{\mathcal{T}\!op}^{w}) has a left adjoint

J:(𝒯op,𝒯opw)(on,onw)J:({\mathcal{T}\!op}^{\ast},{\mathcal{T}\!op}^{w})\to(\mathcal{M}on,\mathcal{M}on^{w})

commonly called the James construction, which associates with each based space XX the free based topological monoid on XX.

Proposition 5.1.

(James [15]) For each path-connected based space there is a weak homotopy equivalence of spaces

JXΩΣX.JX\simeq\Omega\Sigma X.

D. Puppe investigated the conditions which would imply for this weak homotopy equivalence to be a genuine homotopy equivalence.

Proposition 5.2.

(Puppe [9]): If XX is a well-pointed path-connected Dold space then there is a homotopy equivalence

JXΩΣX.JX\simeq\Omega\Sigma X.

Consider the diagram of functors

onw\textstyle{\mathcal{M}on^{w}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B\scriptstyle{B}U\scriptstyle{U}𝒯opw\textstyle{{\mathcal{T}\!op}^{w}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ω\scriptstyle{\Omega^{\prime}}Ω\scriptstyle{\Omega}𝒯opw\textstyle{{\mathcal{T}\!op}^{w}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Σ\scriptstyle{\Sigma}J\scriptstyle{J}

All functors preserve weak equivalences. Hence they induce a diagram

Hoonw\textstyle{\mathop{\rm Ho}\nolimits\mathcal{M}on^{w}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}HoB\scriptstyle{\mathop{\rm Ho}\nolimits B}HoU\scriptstyle{\mathop{\rm Ho}\nolimits U}Ho𝒯opw\textstyle{\mathop{\rm Ho}\nolimits{\mathcal{T}\!op}^{w}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}HoΩ\scriptstyle{\mathop{\rm Ho}\nolimits\Omega^{\prime}}HoΩ\scriptstyle{\mathop{\rm Ho}\nolimits\Omega}Ho𝒯opw\textstyle{\mathop{\rm Ho}\nolimits{\mathcal{T}\!op}^{w}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}HoΣ\scriptstyle{\mathop{\rm Ho}\nolimits\Sigma}HoJ\scriptstyle{\mathop{\rm Ho}\nolimits J}

consisting of adjoint pairs. Since the Moore loop space functor is naturally homotopy equivalent to the usual loop space functor there is a natural transformation

τ(X):UΩ(X)Ω(X)\tau(X):U\circ\Omega^{\prime}(X)\to\Omega(X)

which is a homotopy equivalence. Hence HoΩ\mathop{\rm Ho}\nolimits\Omega and HoUHoΩ\mathop{\rm Ho}\nolimits U\circ\mathop{\rm Ho}\nolimits\Omega^{\prime} are naturally isomorphic. Since their left adjoints are unique up to natural isomorphisms this implies that HoBHoJ\mathop{\rm Ho}\nolimits B\circ\mathop{\rm Ho}\nolimits J and HoΣ\mathop{\rm Ho}\nolimits\Sigma are naturally isomorphic. We obtain

Proposition 5.3.

For each X𝒯opwX\in{\mathcal{T}\!op}^{w} there is a homotopy equivalence

BJ(X)Σ(X)BJ(X)\simeq\Sigma(X)

natural up to homotopy. \square

We obtain Puppe’s result by combining 5.3 with another well-known result:

Proposition 5.4.

If MM is a well-pointed monoid whose underlying space is a Dold space and π0(M)\pi_{0}(M) is a group, then MM is grouplike [9, (12.7)].

Proof of 5.2: If XX is a path-connected Dold space, so is JXJX. Hence JXJX is grouplike and μ(WJX):WJXWΩBWJX\mu(WJX):WJX\to W\Omega^{\prime}BWJX is a weak equivalence by 4.17, so that ΩBε(JX)ε(ΩBJX)μ(WJX)ι(JX):JXΩBJX\Omega^{\prime}B\varepsilon(JX)\circ\varepsilon(\Omega^{\prime}BJX)\circ\mu(WJX)\circ\iota(JX):JX\to\Omega^{\prime}BJX is a homotopy equivalence. We have a sequence of homotopy equivalences

JXΩBJXΩBJXΩΣX.JX\simeq\Omega^{\prime}BJX\simeq\Omega BJX\simeq\Omega\Sigma X.

Homotopical group completion: Homotopical group completion is the replacement of a monoid by a grouplike one having a universal property. We state our result for the full subcategory Hoonw\mathop{\rm Ho}\nolimits\mathcal{M}on^{w} of Hoon\mathop{\rm Ho}\nolimits\mathcal{M}on of well-pointed monoids. Since q(M):VMMq(M):VM\to M is a weak equivalence, Hoonw\mathop{\rm Ho}\nolimits\mathcal{M}on^{w} is equivalent to Hoon\mathop{\rm Ho}\nolimits\mathcal{M}on so that the corresponding statement for Hoon\mathop{\rm Ho}\nolimits\mathcal{M}on follows.

Proposition 5.5.

Let MM be a well-pointed monoid. The homotopy class of the homomorphism μ(WM):WMWΩBWM\mu(WM):WM\to W\Omega^{\prime}BWM, considered as a morphism in Hoonw(M,ΩBWM)\mathop{\rm Ho}\nolimits\mathcal{M}on^{w}(M,\Omega^{\prime}BWM), is a group completion in the following sense: Given a diagram

M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[μ(WM)]\scriptstyle{[\mu(WM)]}[g]\scriptstyle{[g]}ΩBWM\textstyle{\Omega^{\prime}BWM\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[g¯]\scriptstyle{[\overline{g}]}N\textstyle{N}

in Hoonw\mathop{\rm Ho}\nolimits\mathcal{M}on^{w} with NN grouplike, there exists a unique morphism [g¯]:ΩBWMN[\overline{g}]:\Omega^{\prime}BWM\to N making the diagram commute.

Proof.

Consider the homotopy commutative diagram in onw\mathcal{M}on^{w}

WM\textstyle{WM\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ(WM)\scriptstyle{\mu(WM)}g\scriptstyle{g}WΩBWM\textstyle{W\Omega^{\prime}BWM\ignorespaces\ignorespaces\ignorespaces\ignorespaces}WΩBg\scriptstyle{W\Omega^{\prime}Bg}WN\textstyle{WN\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ(WN)\scriptstyle{\mu(WN)}WΩBWN\textstyle{W\Omega^{\prime}BWN}

Since NN is well-pointed and grouplike μ(WN)\mu(WN) is homotopy invertible in onw\mathcal{M}on^{w} by 4.17. We choose a homotopy inverse h:WΩBWNWNh:W\Omega^{\prime}BWN\to WN and define g¯=hWΩBg\overline{g}=h\circ W\Omega^{\prime}Bg. Then g¯μ(WM)g\overline{g}\circ\mu(WM)\simeq g in onw\mathcal{M}on^{w}.

For the uniqueness of [g¯][\overline{g}] suppose there is a homomorphism g:WΩBMWNg^{\prime}:W\Omega^{\prime}BM\longrightarrow WN such that hμ(WM)gh\circ\mu(WM)\simeq g. Put j1=μ(WN)gj_{1}=\mu(WN)\circ g^{\prime} and j2=WΩBgj_{2}=W\Omega^{\prime}Bg. It suffices to show that j1j2j_{1}\simeq j_{2} in onw\mathcal{M}on^{w}. Since Bj1Bμ(WM)μ(WN)BgBj2Bμ(WM)Bj_{1}\circ B\mu(WM)\simeq\mu(WN)\circ Bg\simeq Bj_{2}\circ B\mu(WM) and Bμ(WM)B\mu(WM) is a homotopy equivalence by 4.17, we obtain Bj1Bj2Bj_{1}\simeq Bj_{2}. Since μ(WΩBM)\mu(W\Omega^{\prime}BM) and μ(WΩBN)\mu(W\Omega^{\prime}BN) are homotopy equivalences in onw\mathcal{M}on^{w} by 4.17 and μ\mu is natural up to homotopy the following diagram is homotopy commutative and establishes the result:

WΩBM\textstyle{W\Omega^{\prime}BM\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ(WΩBM)\scriptstyle{\mu(W\Omega^{\prime}BM)}jk\scriptstyle{j_{k}}WΩBWΩBM\textstyle{W\Omega^{\prime}BW\Omega^{\prime}BM\ignorespaces\ignorespaces\ignorespaces\ignorespaces}WΩBjk\scriptstyle{W\Omega^{\prime}Bj_{k}}WΩBN\textstyle{W\Omega^{\prime}BN\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ(WΩBN)\scriptstyle{\mu(W\Omega^{\prime}BN)}WΩBWΩBN\textstyle{W\Omega^{\prime}BW\Omega^{\prime}BN}

Dold spaces and grouplike monoids

For details on Dold spaces see [22]. We restrict our attention to the well-pointed case. Using the whiskering process it is easy to extend our results to the general case.

Let 𝒯opDoldw𝒯opw{\mathcal{T}\!op}^{w}_{Dold}\subset{\mathcal{T}\!op}^{w} denote the full subcategory of well-pointed path-connected Dold spaces. Since BMBM is in 𝒯opDoldw{\mathcal{T}\!op}^{w}_{Dold} for any well-pointed monoid by [22, Cor. 5.2], the classifying space functor restricts to a functor

B:onw𝒯opDoldw.B:\mathcal{M}on^{w}\to{\mathcal{T}\!op}^{w}_{Dold}.

Let ongroupwonw\mathcal{M}on^{w}_{group}\subset\mathcal{M}on^{w} denote the full subcategory of grouplike well-pointed monoids. Then Proposition 4.17 implies

Theorem 5.6.

The functors

Bw:ongroupw𝒯opDoldw:ΩwB^{w{\mathcal{H}}}:\mathcal{H}\mathcal{M}on^{w}_{group}\leftrightarrows\mathcal{H}{\mathcal{T}\!op}^{w}_{Dold}:\Omega^{\prime w{\mathcal{H}}}

define an equivalence up to homotopy of categories, i.e. the natural transformations up to homotopy μ:IdΩwBw\mu:\mathop{\rm Id}\nolimits\to\Omega^{\prime w{\mathcal{H}}}\circ B^{w{\mathcal{H}}} and η:BwΩwId\eta:B^{w{\mathcal{H}}}\circ\Omega^{\prime w{\mathcal{H}}}\to\mathop{\rm Id}\nolimits take values in homotopy equivalences. In particular,

HoB:HoongroupwHo𝒯opDoldw:HoΩ\mathop{\rm Ho}\nolimits B:\mathop{\rm Ho}\nolimits\mathcal{M}on^{w}_{group}\leftrightarrows\mathop{\rm Ho}\nolimits{\mathcal{T}\!op}^{w}_{Dold}:\mathop{\rm Ho}\nolimits\Omega^{\prime}

define an equivalence of categories. \Box

The second part is a slight extension of a well-known result (e.g. see [4, Section 4]).

The following two propositions extend and strengthen results of Fuchs [11, Satz 7.7].

The diagram

onw(WM,WN)\textstyle{\mathcal{M}on^{w}(WM,WN)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ(WN)\scriptstyle{\mu(WN)_{\ast}}B\scriptstyle{B}onw(WM,WΩBWN)\textstyle{\mathcal{M}on^{w}(WM,W\Omega^{\prime}BWN)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λ(WM,BWN)\scriptstyle{\lambda(WM,BWN)}𝒯opw(BWM,BWN)\textstyle{{\mathcal{T}\!op}^{w}(BWM,BWN)}

commutes up to homotopy because by 4.13

(λ(WM,BWN)μ(WN))(f)=η(BWN)Bμ(WN))(f)=η(BWN)Bμ(WN)BfBfcontinuously in f\begin{array}[]{rcl}(\lambda(WM,BWN)\circ\mu(WN)_{\ast})(f)&=&\eta(BWN)\circ B^{\prime}\circ\mu(WN)_{\ast})(f)\\ &=&\eta(BWN)\circ B\mu(WN)\circ Bf\\ &\simeq&Bf\quad\textrm{continuously in }f\end{array}

with B:onw(WM,WΩBWN)𝒯opw(BWM,BWΩBWN)B^{\prime}:\mathcal{M}on^{w}(WM,W\Omega^{\prime}BWN)\to{\mathcal{T}\!op}^{w}(BWM,BW\Omega^{\prime}BWN). If NN is grouplike μ(WN)\mu(WN) is a homotopy equivalence in onw\mathcal{M}on^{w}, and we obtain

Proposition 5.7.

If NN is a well-pointed grouplike monoid then

B:onw(WM,WN)𝒯opw(BWM,BWN)B:\mathcal{M}on^{w}(WM,WN)\to{\mathcal{T}\!op}^{w}(BWM,BWN)

is a homotopy equivalence. \Box

Since η(X):BWΩXX\eta(X):BW\Omega^{\prime}X\to X is a natural transformation the following diagram commutes

onw(WΩX,WΩY)\textstyle{\mathcal{M}on^{w}(W\Omega^{\prime}X,W\Omega^{\prime}Y)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B\scriptstyle{B}𝒯opw(X,Y)\textstyle{{\mathcal{T}\!op}^{w}(X,Y)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}WΩ\scriptstyle{W\Omega^{\prime}}η(X)\scriptstyle{\eta(X)^{\ast}}𝒯opw(BWΩX,BWΩY)\textstyle{{\mathcal{T}\!op}^{w}(BW\Omega^{\prime}X,BW\Omega^{\prime}Y)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}η(Y)\scriptstyle{\eta(Y)_{\ast}}𝒯opw(BWΩX,Y).\textstyle{{\mathcal{T}\!op}^{w}(BW\Omega^{\prime}X,Y).}

Since WΩYW\Omega^{\prime}Y is grouplike the map BB is a homotopy equivalence by 5.7. Since η(Y)B=λ(WΩX,Y)\eta(Y)_{\ast}\circ B=\lambda(W\Omega^{\prime}X,Y) the map η(Y):𝒯opw(BWΩX,BWΩY)𝒯opw(BWΩX,Y)\eta(Y)_{\ast}:{\mathcal{T}\!op}^{w}(BW\Omega^{\prime}X,BW\Omega^{\prime}Y)\to{\mathcal{T}\!op}^{w}(BW\Omega^{\prime}X,Y) is a homotopy equivalence. If XX is a well-pointed path-connected Dold space η(X)\eta(X) is a based homotopy equivalence by 4.17. We obtain

Proposition 5.8.

If XX is a well-pointed path-connected Dold space then WΩ:𝒯opw(X,Y)onw(WΩX,WΩY)W\Omega^{\prime}:{\mathcal{T}\!op}^{w}(X,Y)\to\mathcal{M}on^{w}(W\Omega^{\prime}X,W\Omega^{\prime}Y) is a homotopy equivalence. \Box

Homotopy homomorphisms and unitary homotopy homomorphisms

Proposition 5.9.

Let MM and NN be well-pointed monoids and NN be grouplike. Then ε(M):W¯MWM\varepsilon^{\prime}(M):\overline{W}M\to WM induces a homotopy equivalence

on(WM,N)𝒮gp(W¯M,N).\mathcal{M}on(WM,N)\to\mathop{\rm\mathcal{S}gp}(\overline{W}M,N).
Proof.

By 2.7 we may replace NN by WNWN. Since 𝒮gp(W¯M,WN)\mathop{\rm\mathcal{S}gp}(\overline{W}M,WN) is naturally homeomorphic to on(W(M+),WN)\mathcal{M}on(W(M_{+}),WN) by 2.5 it suffices to show that the counit κ(M):M+M\kappa(M):M_{+}\to M induces a homotopy equivalence

κ(M):on(WM,WN)on(W(M+),WN)\kappa(M)^{\ast}:\mathcal{M}on(WM,WN)\to\mathcal{M}on(W(M_{+}),WN)

The diagram

on(WM,WN)\textstyle{\mathcal{M}on(WM,WN)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B\scriptstyle{B}κ(M)\scriptstyle{\kappa(M)^{\ast}}𝒯op(BWM,BWN)\textstyle{{\mathcal{T}\!op}^{\ast}(BWM,BWN)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bκ(M)\scriptstyle{B\kappa(M)^{\ast}}on(W(M+),WN)\textstyle{\mathcal{M}on(W(M_{+}),WN)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B\scriptstyle{B}𝒯op(BW(M+),BWN)\textstyle{{\mathcal{T}\!op}^{\ast}(BW(M_{+}),BWN)}

commutes. By 5.7 the maps BB are homotopy equivalences, and by 4.2 the map Bκ(M)B\kappa(M)^{\ast} is a homotopy equivalence. Hence so is κ(M)\kappa(M)^{\ast}. ∎

Remark 5.10.

In general we cannot expect that ε(M):W¯MWM\varepsilon^{\prime}(M):\overline{W}M\to WM induces a homotopy equivalence. E.g. it can happen that a homomorphism W¯MN\overline{W}M\to N does not map (eM)(e_{M}) into the path-component of eNe_{N} so that there is no chance to homotop it into a homomorphism WMNWM\to N.

Proposition 5.11.

If MM is a well-pointed monoid then Wq(M):WVMWMWq(M):WVM\to WM is a homotopy equivalence in onw\mathcal{M}on^{w} by 2.12 inducing a homotopy equivalence

on(WM,N)on(WVM,N).\mathcal{M}on(WM,N)\to\mathcal{M}on(WVM,N).

6 Diagrams of monoids

We want to show that the homotopy adjunction of Theorem 4.5 lifts to diagram categories. This is not evident: since the unit of our homotopy adjunction is only natural up to homotopy it does not lift to diagrams.

Let \mathcal{M} be a cocomplete 𝒯op\mathcal{T}\!op-enriched tensored category with a class 𝒲\mathscr{W} of weak equivalences containing the homotopy equivalences. We assume that \mathcal{M} has a strong cofibrant replacement functor (QM,εM)(Q_{M},\varepsilon_{M}). We use \otimes for the tensor in \mathcal{M} and QQ for QMQ_{M} as long as there is no ambiguity.

Definition 6.1.

Let 𝒞\mathcal{C} be a small indexing category. A morphism f:D1D2f:D_{1}\to D_{2} of 𝒞\mathcal{C}-diagrams in \mathcal{M} is called a weak equivalence if it is objectwise a weak equivalence in \mathcal{M}. We denote the class of weak equivalences in 𝒞\mathcal{M}^{\mathcal{C}} by 𝒲𝒞\mathscr{W}^{\mathcal{C}}.

Our first aim is to show that 𝒞\mathcal{M}^{\mathcal{C}} admits a strong cofibrant replacement functor in order to make additional applications of Proposition 3.4. Therefore we proceed as in 2.7 and 2.2.

We define a 𝒞×𝒞op\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}-diagram B(𝒞,𝒞,𝒞)B(\mathcal{C},\mathcal{C},\mathcal{C}) in 𝒯op\mathcal{T}\!op as follows:

B(𝒞,𝒞,𝒞)(b,a)=B(𝒞(,b),𝒞,𝒞(a,))B(\mathcal{C},\mathcal{C},\mathcal{C})(b,a)=B(\mathcal{C}(-,b),\mathcal{C},\mathcal{C}(a,-))

where the right side is the 2-sided bar construction of 4.1.

The 𝒞×𝒞op\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits} structure on Bn(𝒞,𝒞,𝒞)B_{n}(\mathcal{C},\mathcal{C},\mathcal{C}) is given by

(g,h)(f0,,fn+1)=(gf0,f1,,fn,fn+1h)(g,h)\cdot(f_{0},\ldots,f_{n+1})=(g\circ f_{0},f_{1},\ldots,f_{n},f_{n+1}\circ h)

Analogously we define a 𝒞op\mathcal{C}^{\mathop{\rm op}\nolimits}-diagram B(,𝒞,𝒞)B(\ast,\mathcal{C},\mathcal{C}) in 𝒯op\mathcal{T}\!op, where \ast denotes the constant 𝒞op\mathcal{C}^{\mathop{\rm op}\nolimits}-diagram on a single point.

Lemma 6.2.

Let XX and YY be 𝒞×𝒞op\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}-diagrams in 𝒯op\mathcal{T}\!op, let p:XYp:X\to Y be a map of diagrams which is objectwise a homotopy equivalence. Then pp induces a homotopy equivalence

p:𝒯op𝒞×𝒞op(B(𝒞,𝒞,𝒞),X)𝒯op𝒞×𝒞op(B(𝒞,𝒞,𝒞),Y)p_{\ast}:\mathcal{T}\!op^{\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}}(B(\mathcal{C},\mathcal{C},\mathcal{C}),X)\to\mathcal{T}\!op^{\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}}(B(\mathcal{C},\mathcal{C},\mathcal{C}),Y)

in 𝒯op\mathcal{T}\!op.

Proof.

We apply the HELP-Lemma. So given a diagram

K\textstyle{K\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f¯K\scriptstyle{\bar{f}_{K}}i\scriptstyle{i}𝒯op𝒞×𝒞op(B(𝒞,𝒞,𝒞),X)\textstyle{\mathcal{T}\!op^{\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}}(B(\mathcal{C},\mathcal{C},\mathcal{C}),X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p_{\ast}}L\textstyle{L\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g¯\scriptstyle{\bar{g}}𝒯op𝒞×𝒞op(B(𝒞,𝒞,𝒞),Y)\textstyle{\mathcal{T}\!op^{\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}}(B(\mathcal{C},\mathcal{C},\mathcal{C}),Y)}

which commutes up to a homotopy h¯K,t:g¯ipf¯K\bar{h}_{K,t}:\bar{g}\circ i\simeq p_{\ast}\circ\bar{f}_{K}, where ii is a closed cofibration, we have to construct extensions

f¯:L𝒯op𝒞×𝒞op(B(𝒞,𝒞,𝒞),X)h¯t:L𝒯op𝒞×𝒞op(B(𝒞,𝒞,𝒞),Y)\begin{array}[]{rcl}\bar{f}:L&\rightarrow&\mathcal{T}\!op^{\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}}(B(\mathcal{C},\mathcal{C},\mathcal{C}),X)\\ \bar{h}_{t}:L&\rightarrow&\mathcal{T}\!op^{\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}}(B(\mathcal{C},\mathcal{C},\mathcal{C}),Y)\end{array}

of f¯K\bar{f}_{K} respectively h¯K,t\bar{h}_{K,t} such that h¯t:g¯pf¯\bar{h}_{t}:\bar{g}\simeq p_{\ast}\circ\bar{f}.

Taking adjoints the above diagram translates to the following diagram of 𝒞×𝒞op\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}-spaces

K×B(𝒞,𝒞,𝒞)\textstyle{K\times B(\mathcal{C},\mathcal{C},\mathcal{C})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f^{\prime}}i×id\scriptstyle{i\times\mathop{\rm id}\nolimits}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}L×B(𝒞,𝒞,𝒞)\textstyle{L\times B(\mathcal{C},\mathcal{C},\mathcal{C})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{g}Y\textstyle{Y}

which commutes up to a homotopy ht:g(i×id)pfh^{\prime}_{t}:g\circ(i\times\mathop{\rm id}\nolimits)\simeq p\circ f^{\prime} in 𝒯op𝒞×𝒞op\mathcal{T}\!op^{\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}}, and it suffices to construct extensions f:L×B(𝒞,𝒞,𝒞)Xf:L\times B(\mathcal{C},\mathcal{C},\mathcal{C})\to X of ff^{\prime} and ht:L×B(𝒞,𝒞,𝒞)Yh_{t}:L\times B(\mathcal{C},\mathcal{C},\mathcal{C})\to Y of hth^{\prime}_{t} such that ht:gpfh_{t}:g\simeq p\circ f in 𝒯op𝒞×𝒞op\mathcal{T}\!op^{\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}}.

We construct these extensions by induction on the natural filtration FnF_{n} of L×B(𝒞,𝒞,𝒞)L\times B(\mathcal{C},\mathcal{C},\mathcal{C}) induced by the realization of the simplicial set B(𝒞,𝒞,𝒞)B_{\bullet}(\mathcal{C},\mathcal{C},\mathcal{C}). We start with F0=a,b,cL×𝒞(c,b)×𝒞(a,c)F_{0}=\coprod_{a,b,c}L\times\mathcal{C}(c,b)\times\mathcal{C}(a,c). The diagram

K×{(idc,idc)}\textstyle{K\times\{(\mathop{\rm id}\nolimits_{c},\mathop{\rm id}\nolimits_{c})\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f^{\prime}}X(c,c)\textstyle{X(c,c)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}L×{(idc,idc)\textstyle{L\times\{(\mathop{\rm id}\nolimits_{c},\mathop{\rm id}\nolimits_{c})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{g}Y(c,c)\textstyle{Y(c,c)}

commutes up to a homotopy given by hth_{t}^{\prime}. Since p:X(c,c)Y(c,c)p:X(c,c)\to Y(c,c) is a homotopy equivalence and KLK\to L is a closed cofibration the required extensions exist by the HELP-Lemma. We extend ff over all of F0F_{0} by f(l,j0,j1)=X(j0,j1)f(l,id,id)f(l,j_{0},j_{1})=X(j_{0},j_{1})\circ f(l,\mathop{\rm id}\nolimits,\mathop{\rm id}\nolimits) and analogously for hth_{t}.
Now suppose that ff and hth_{t} have been defined on Fn1F_{n-1}. We obtain FnF_{n} from Fn1F_{n-1} by attaching spaces L×(j0,,jn+1)×ΔnL\times(j_{0},\ldots,j_{n+1})\times\Delta^{n} along L×(j0,,jn+1)×ΔnL\times(j_{0},\ldots,j_{n+1})\times\partial\Delta^{n}. Here the jkj_{k} are morphisms in 𝒞\mathcal{C} such that the composition

j0jn+1:acnc0bj_{0}\circ\ldots j_{n+1}:a\to c_{n}\to\ldots\ c_{0}\to b

is defined and j1,,jnj_{1},\ldots,j_{n} are not identities. Hence the extension ff and the homotopy hth_{t} are already defined on

D(L×(j0,j1,,jn,jn+1)×Δn)=K×(j0,j1,,jn,jn+1)×ΔnL×(j0,j1,,jn,jn+1)×Δn.D(L\times(j_{0},j_{1},\ldots,j_{n},j_{n+1})\times\Delta^{n})\\ {}\qquad\qquad=K\times(j_{0},j_{1},\ldots,j_{n},j_{n+1})\times\Delta^{n}\cup L\times(j_{0},j_{1},\ldots,j_{n},j_{n+1})\times\partial\Delta^{n}.

We apply the HELP-Lemma to the homotopy commutative diagram

D(L×(idc0,j1,,jn,idcn)×Δn)\textstyle{D(L\times(\mathop{\rm id}\nolimits_{c_{0}},j_{1},\ldots,j_{n},\mathop{\rm id}\nolimits_{c_{n}})\times\Delta^{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f′′\scriptstyle{f^{\prime\prime}}X(c0,cn)\textstyle{X(c_{0},c_{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}L×(idc0,j1,,jn,idcn)×Δn\textstyle{L\times(\mathop{\rm id}\nolimits_{c_{0}},j_{1},\ldots,j_{n},\mathop{\rm id}\nolimits_{c_{n}})\times\Delta^{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{g}Y(c0,cn)\textstyle{Y(c_{0},c_{n})}

where f′′f^{\prime\prime} and the commuting homotopy are given by the already defined extensions. Since p{p} is objectwise a homotopy equivalence and the inclusion

D(L×(idc0,j1,,jn,idcn)×Δn)L×(idc0,j1,,jn,idcn)×ΔnD(L\times(\mathop{\rm id}\nolimits_{c_{0}},j_{1},\ldots,j_{n},\mathop{\rm id}\nolimits_{c_{n}})\times\Delta^{n})\subset L\times(\mathop{\rm id}\nolimits_{c_{0}},j_{1},\ldots,j_{n},\mathop{\rm id}\nolimits_{c_{n}})\times\Delta^{n}

is a closed cofibration the required extensions exist. We extend our maps to maps of diagrams as in the F0F_{0}-case. ∎

Let DD be a 𝒞\mathcal{C}-diagram in \mathcal{M} and XX a 𝒞op\mathcal{C}^{\mathop{\rm op}\nolimits}-diagram in 𝒯op\mathcal{T}\!op. We define X𝒞DX\otimes_{\mathcal{C}}D to be the coequalizer in \mathcal{M} of

f mor 𝒞X(target(f))D(source(f))\textstyle{\coprod\limits_{f\in\textrm{ mor }\mathcal{C}}X(\textrm{target}(f))\otimes D(\textrm{source}(f))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{\alpha}β\scriptstyle{\beta}aob𝒞X(a)D(a)\textstyle{\coprod\limits_{a\in\rm{ob}\mathcal{C}}X(a)\otimes D(a)}

where for f:abf:a\to b in 𝒞\mathcal{C} the ff-summand X(b)D(a)X(b)\otimes D(a) is mapped as follows

α=X(f)id:X(b)D(a)X(a)D(a)β=idD(f):X(b)D(a)X(b)D(b)\begin{array}[]{rcrrcl}\alpha&=&X(f)\otimes\mathop{\rm id}\nolimits:&X(b)\otimes D(a)&\longrightarrow&X(a)\otimes D(a)\\ \beta&=&\mathop{\rm id}\nolimits\otimes D(f):&X(b)\otimes D(a)&\longrightarrow&X(b)\otimes D(b)\end{array}

We define a functor

R:𝒞𝒞,DB(𝒞,𝒞,𝒞)𝒞QDR:\mathcal{M}^{\mathcal{C}}\rightarrow\mathcal{M}^{\mathcal{C}},\qquad D\mapsto B(\mathcal{C},\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}QD

where B(𝒞,𝒞,𝒞)𝒞DB(\mathcal{C},\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}D is the 𝒞\mathcal{C}-diagram

aB(𝒞(,a),𝒞,𝒞)𝒞Da\mapsto B(\mathcal{C}(-,a),\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}D

in \mathcal{M}.

Proposition 6.3.

Let D0,D1D_{0},\ D_{1}, and D2D_{2} be 𝒞\mathcal{C} diagrams in \mathcal{M}, let p:D1D2p:D_{1}\to D_{2} be a weak equivalence in 𝒞\mathcal{M}^{\mathcal{C}} and q:A1A2q:A_{1}\to A_{2} a weak equivalence in \mathcal{M}. Then pp and qq induce homotopy equivalences

p:𝒞(RD0,D1)𝒞(RD0,D2)q:(B(,𝒞,𝒞)𝒞QD0,A1)(B(,𝒞,𝒞)𝒞QD0,A2)\begin{array}[]{rcl}p_{\ast}:\mathcal{M}^{\mathcal{C}}(RD_{0},D_{1})&\rightarrow&\mathcal{M}^{\mathcal{C}}(RD_{0},D_{2})\\ q_{\ast}:\mathcal{M}(B(\ast,\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}QD_{0},A_{1})&\rightarrow&\mathcal{M}(B(\ast,\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}QD_{0},A_{2})\end{array}

in 𝒯op\mathcal{T}\!op.

Proof.

Since 𝒞(RD0,Di)𝒯op𝒞×𝒞op(B(𝒞,𝒞,𝒞),(QD0,Di))\mathcal{M}^{\mathcal{C}}(RD_{0},D_{i})\cong\mathcal{T}\!op^{\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}}(B(\mathcal{C},\mathcal{C},\mathcal{C}),\mathcal{M}(QD_{0},D_{i})) it follows from Lemma 6.2 with X(b,a)=(QD0(a),D1(b))X(b,a)=\mathcal{M}(QD_{0}(a),D_{1}(b)) and Y(b,a)=(QD0(a),D2(b))Y(b,a)=\mathcal{M}(QD_{0}(a),D_{2}(b)) that pp_{\ast} is a homotopy equivalence.

There is a sequence of natural homeomorphisms

(B(,𝒞,𝒞)𝒞QD0,Ai)𝒯op𝒞op(B(,𝒞,𝒞),(QD0,Ai)𝒯op𝒞op(colim𝒞B(𝒞,𝒞,𝒞),(QD0,Ai)𝒯op𝒞×𝒞op(B(𝒞,𝒞,𝒞),(QD0,constAi)\begin{array}[]{rcl}\mathcal{M}(B(\ast,\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}QD_{0},A_{i})&\cong&\mathcal{T}\!op^{\mathcal{C}^{\mathop{\rm op}\nolimits}}(B(\ast,\mathcal{C},\mathcal{C}),\mathcal{M}(QD_{0},A_{i})\\ &\cong&\mathcal{T}\!op^{\mathcal{C}^{\mathop{\rm op}\nolimits}}(\mathop{\rm colim}\nolimits_{\mathcal{C}}B(\mathcal{C},\mathcal{C},\mathcal{C}),\mathcal{M}(QD_{0},A_{i})\\ &\cong&\mathcal{T}\!op^{\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}}(B(\mathcal{C},\mathcal{C},\mathcal{C}),\mathcal{M}(QD_{0},\textrm{const}A_{i})\end{array}

where constAi\textrm{const}A_{i} are the constant 𝒞\mathcal{C}-diagrams on AiA_{i}. As in the first part, it follows that qq_{\ast} is a homotopy equivalence. ∎

Let 𝒞\mathcal{C}_{\bullet} denote the 𝒞×𝒞op\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}-diagram of simplical sets sending (b,a)(b,a) to the constant simplicial set 𝒞(a,b)\mathcal{C}(a,b). The maps

δn:Bn(𝒞,𝒞,𝒞)(b,a)=Bn(𝒞(,b),𝒞,𝒞(a,))𝒞(a,b)(f0,,fn+1)f0fn+1\begin{array}[]{rrcl}\delta_{n}:&B_{n}(\mathcal{C},\mathcal{C},\mathcal{C})(b,a)=B_{n}(\mathcal{C}(-,b),\mathcal{C},\mathcal{C}(a,-))&\longrightarrow&\mathcal{C}(a,b)\\ &(f_{0},\ldots,f_{n+1})&\longmapsto&f_{0}\circ\ldots\circ f_{n+1}\end{array}

define a simplicial map B(𝒞,𝒞,𝒞)𝒞B_{\bullet}(\mathcal{C},\mathcal{C},\mathcal{C})\to\mathcal{C}_{\bullet}. Let δ:B(𝒞,𝒞,𝒞)𝒞\delta:B(\mathcal{C},\mathcal{C},\mathcal{C})\to\mathcal{C} be its realization.

Proposition 6.4.

δ(D)=δ𝒞idD:B(𝒞,𝒞,𝒞)𝒞D𝒞𝒞DD\delta(D)=\delta\otimes_{\mathcal{C}}\mathop{\rm id}\nolimits_{D}:B(\mathcal{C},\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}D\to\mathcal{C}\otimes_{\mathcal{C}}D\cong D is objectwise a homotopy equivalence in \mathcal{M} and hence a weak equivalence in 𝒞\mathcal{M}^{\mathcal{C}}.

The proposition is an immediate consequence of the following Lemma:

Lemma 6.5.

For each object b𝒞b\in\mathcal{C} the map ε:B(𝒞,𝒞,𝒞)(,b)𝒞(,b)\varepsilon:B(\mathcal{C},\mathcal{C},\mathcal{C})(-,b)\to\mathcal{C}(-,b) is a homotopy equivalence in the category 𝒯op𝒞op\mathcal{T}\!op^{\mathcal{C}^{\mathop{\rm op}\nolimits}}.

Proof.

For a𝒞a\in\mathcal{C} let 𝒳a\mathcal{X}_{a} denote the category whose objects are diagrams aj1cj0ba\xrightarrow{j_{1}}c\xrightarrow{j_{0}}b and whose morphisms from this object to aj1cj0ba\xrightarrow{j^{\prime}_{1}}c^{\prime}\xrightarrow{j^{\prime}_{0}}b are morphisms h:cch:c\to c^{\prime} in 𝒞\mathcal{C} making the diagram

c\textstyle{c\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h\scriptstyle{h}j0\scriptstyle{j_{0}}a\textstyle{a\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j1\scriptstyle{j_{1}}j1\scriptstyle{j^{\prime}_{1}}b\textstyle{b}c\textstyle{c^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j0\scriptstyle{j^{\prime}_{0}}

commute. Let 𝒞(a,b)\mathcal{C}(a,b) stand for the discrete category whose object set is 𝒞(a,b)\mathcal{C}(a,b). Then

εa:𝒳a𝒞(a,b),(aj1cj0b)(j0j1:ab)\varepsilon_{a}:\mathcal{X}_{a}\to\mathcal{C}(a,b),\qquad(a\xrightarrow{j_{1}}c\xrightarrow{j_{0}}b)\mapsto(j_{0}\circ j_{1}:a\to b)

defines a functor which has the section

sa:𝒞(a,b)𝒳a,j(a𝑗bidb).s_{a}:\mathcal{C}(a,b)\to\mathcal{X}_{a},\qquad j\mapsto(a\xrightarrow{j}b\xrightarrow{\mathop{\rm id}\nolimits}b).

There is a natural transformation τa:Id𝒳asaεa\tau_{a}:\mathop{\rm Id}\nolimits_{\mathcal{X}_{a}}\to s_{a}\circ\varepsilon_{a} defined by the diagram

c\textstyle{c\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j0\scriptstyle{j_{0}}j0\scriptstyle{j_{0}}a\textstyle{a\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j1\scriptstyle{j_{1}}j0j1\scriptstyle{j_{0}\circ j_{1}}b\textstyle{b}b\textstyle{b\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathop{\rm id}\nolimits}

So εa\varepsilon_{a} induces a homotopy equivalence of the classifying spaces. Now B(𝒳a)=B(𝒞,𝒞,𝒞(a,b))B(\mathcal{X}_{a})=B(\mathcal{C},\mathcal{C},\mathcal{C}(a,b)) and B(𝒞(a,b))=𝒞(a,b)B(\mathcal{C}(a,b))=\mathcal{C}(a,b). Moreover all data are natural with respect to a𝒞opa\in\mathcal{C}^{\mathop{\rm op}\nolimits}. Hence we obtain the required result. ∎

When we combine 6.3 and 6.4 we obtain the following corollary.

Corollary 6.6.

R:𝒞𝒞R:\mathcal{M}^{\mathcal{C}}\rightarrow\mathcal{M}^{\mathcal{C}} together with ϵ=δ𝒞εM:RId\epsilon=\delta\otimes_{\mathcal{C}}\varepsilon_{M}:R\to\mathop{\rm Id}\nolimits is a strong cofibrant replacement functor.

Let 𝒩\mathcal{N} be another cocomplete 𝒯op\mathcal{T}\!op-enriched tensored category with a class of weak equivalences containing the homotopy equivalences and a strong cofibrant replacement functor (QN,εN)(Q_{N},\varepsilon_{N}).

Theorem 6.7.

Let

F:𝒩:GF:\mathcal{M}\leftrightarrows\mathcal{N}:G

be continuous functors inducing a natural homotopy equivalence

λ(QMA,QNX):(QMA,QMGQNX)𝒩(QNFQMA,QNX)\lambda(Q_{M}A,Q_{N}X):\mathcal{M}(Q_{M}A,Q_{M}GQ_{N}X)\to\mathcal{N}(Q_{N}FQ_{M}A,Q_{N}X)

so that

F:𝒩:GF^{\mathcal{H}}:\mathcal{H}\mathcal{M}\leftrightarrows\mathcal{H}\mathcal{N}:G^{\mathcal{H}}

is a conatural adjunction up to homotopy. Then

(F𝒞):(𝒞)(𝒩𝒞):(G𝒞)(F^{\mathcal{C}})^{\mathcal{H}}:\mathcal{H}(\mathcal{M}^{\mathcal{C}})\leftrightarrows\mathcal{H}(\mathcal{N}^{\mathcal{C}}):(G^{\mathcal{C}})^{\mathcal{H}}

is an adjunction up to homotopy, and hence

Ho(F𝒞):Ho(𝒞)Ho(𝒩𝒞):Ho(G𝒞)\mathop{\rm Ho}\nolimits(F^{\mathcal{C}}):\mathop{\rm Ho}\nolimits(\mathcal{M}^{\mathcal{C}})\leftrightarrows\mathop{\rm Ho}\nolimits(\mathcal{N}^{\mathcal{C}}):\mathop{\rm Ho}\nolimits(G^{\mathcal{C}})

a genuine adjunction.

Proof.

For diagrams D:𝒞D:\mathcal{C}\to\mathcal{M} and Z:𝒞𝒩Z:\mathcal{C}\to\mathcal{N} we have a sequence of natural maps

𝒯op𝒞×𝒞op(B(𝒞,𝒞,𝒞),𝒩(QNFQMD,QNZ))\textstyle{\mathcal{T}\!op^{\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}}(B(\mathcal{C},\mathcal{C},\mathcal{C}),\mathcal{N}(Q_{N}FQ_{M}D,Q_{N}Z))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}𝒩𝒞(RNFQMD,QNZ)\textstyle{\mathcal{N}^{\mathcal{C}}(R_{N}FQ_{M}D,Q_{N}Z)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}RNFδ(QMD)\scriptstyle{R_{N}F\delta(Q_{M}D)^{\ast}}𝒯op𝒞×𝒞op(B(𝒞,𝒞,𝒞),(QMD,QMGQNZ))\textstyle{\mathcal{T}\!op^{\mathcal{C}\times\mathcal{C}^{\mathop{\rm op}\nolimits}}(B(\mathcal{C},\mathcal{C},\mathcal{C}),\mathcal{M}(Q_{M}D,Q_{M}GQ_{N}Z))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λ(QMD,QNZ)\scriptstyle{\lambda(Q_{M}D,Q_{N}Z)_{\ast}}𝒩𝒞(RNFRMD,QNZ)\textstyle{\mathcal{N}^{\mathcal{C}}(R_{N}FR_{M}D,Q_{N}Z)}𝒞(RMD,QMGQNZ)\textstyle{\mathcal{M}^{\mathcal{C}}(R_{M}D,Q_{M}GQ_{N}Z)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}𝒩𝒞(RNFRMD,RNZ)\textstyle{\mathcal{N}^{\mathcal{C}}(R_{N}FR_{M}D,R_{N}Z)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δ(QNZ)\scriptstyle{\delta(Q_{N}Z)_{\ast}}𝒞(RMD,RMGQNZ)\textstyle{\mathcal{M}^{\mathcal{C}}(R_{M}D,R_{M}GQ_{N}Z)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δ(QMGQNZ)\scriptstyle{\delta(Q_{M}GQ_{N}Z)_{\ast}}𝒞(RMD,RMGRNZ)\textstyle{\mathcal{M}^{\mathcal{C}}(R_{M}D,R_{M}GR_{N}Z)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}RMGδ(QNZ)\scriptstyle{R_{M}G\delta(Q_{N}Z)_{\ast}}

By assumption λ(QMD,QNZ)\lambda(Q_{M}D,Q_{N}Z) is a homotopy equivalence. Since δ(D)\delta(D) is objectwise a homotopy equivalence and since continuous functors preserve homotopy equivalences, RMGδ(QNZ)R_{M}G\delta(Q_{N}Z) and RNFδ(QMD)R_{N}F\delta(Q_{M}D) are homotopy equivalences in 𝒞\mathcal{M}^{\mathcal{C}} by 6.3 so that RMGδ(QNZ)R_{M}G\delta(Q_{N}Z)_{\ast} and RNFδ(QMD)R_{N}F\delta(Q_{M}D)^{\ast} are homotopy equivalences in 𝒯op\mathcal{T}\!op, and δ(QNZ)\delta(Q_{N}Z)_{\ast} and δ(QMGQNZ)\delta(Q_{M}GQ_{N}Z)_{\ast} are homotopy equivalences in 𝒯op\mathcal{T}\!op by 6.3. ∎

6.8.

Addendum: The last natural map in the proof of the theorem points in the wrong direction. So we cannot conclude that (F𝒞)(F^{\mathcal{C}})^{\mathcal{H}} and (G𝒞)(G^{\mathcal{C}})^{\mathcal{H}} are a conatural homotopy adjoint pair.

η(QNY)=λ(QMGQNY,QNY)(idQMGQNY):QNFQMGQNYQNY\eta(Q_{N}Y)=\lambda(Q_{M}GQ_{N}Y,Q_{N}Y)(\mathop{\rm id}\nolimits_{Q_{M}GQ_{N}Y}):Q_{N}FQ_{M}GQ_{N}Y\to Q_{N}Y

is natural with respect to morphisms f:QNY1QNY2f:Q_{N}Y_{1}\to Q_{N}Y_{2} in 𝒩\mathcal{N}. If η\eta extends to a natural map η(Y):QNFQMGYY\eta(Y):Q_{N}FQ_{M}GY\to Y for all Y𝒩Y\in\mathcal{N} or at least for all YY of the form Y=RNYY=R_{N}Y^{\prime} we obtain a natural map λ𝒞(RMD,RNZ)\lambda^{\mathcal{C}}(R_{M}D,R_{N}Z) defined by

𝒞(RMD,RMGRNZ)\textstyle{\mathcal{M}^{\mathcal{C}}(R_{M}D,R_{M}GR_{N}Z)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}RNF\scriptstyle{R_{N}F}λ𝒞(RMD,RNZ)\scriptstyle{\lambda^{\mathcal{C}}(R_{M}D,R_{N}Z)}𝒞(RNFRMD,RNFRMGRNZ)\textstyle{\mathcal{M}^{\mathcal{C}}(R_{N}FR_{M}D,R_{N}FR_{M}GR_{N}Z)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}η(RNZ)\scriptstyle{\eta(R_{N}Z)}𝒩𝒞(RNFRMD,RNZ)\textstyle{\mathcal{N}^{\mathcal{C}}(R_{N}FR_{M}D,R_{N}Z)}

which makes the diagram of the proof of the theorem commute so that (F𝒞)(F^{\mathcal{C}})^{\mathcal{H}} and (G𝒞)(G^{\mathcal{C}})^{\mathcal{H}} are a conatural homotopy adjoint pair.

For use in the next proposition we note

Lemma 6.9.

Let D:𝒞onwD:\mathcal{C}\to\mathcal{M}on^{w} be a diagram of well-pointed monoids. Then B(,𝒞,𝒞)𝒞DB(\ast,\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}D is a well-pointed space, and B(𝒞,𝒞,𝒞)𝒞DB(\mathcal{C},\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}D and B(𝒞,𝒞,𝒞)𝒞WDB(\mathcal{C},\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}WD are diagrams of well-pointed monoids.

Proof.

The first part holds by [19, Prop. 7.8]. The second and third statement follow by the argument used in [19, Prop. 7.8]. ∎

¿From 6.6 and 6.9 we obtain

Proposition 6.10.

With the choices of weak equivalences 𝒲\mathscr{W} as in 3.8 the functors

on𝒞on𝒞DB(𝒞,𝒞,𝒞)𝒞WVD(onw)𝒞(onw)𝒞DB(𝒞,𝒞,𝒞)𝒞WD𝒮gp𝒞𝒮gp𝒞DB(𝒞,𝒞,𝒞)𝒞W¯D(𝒯op)𝒞(𝒯op)𝒞DB(𝒞,𝒞,𝒞)+𝒞VtD(𝒯opw)𝒞(𝒯opw)𝒞DB(𝒞,𝒞,𝒞)+𝒞D\begin{array}[]{rcll}\mathcal{M}on^{\mathcal{C}}&\rightarrow&\mathcal{M}on^{\mathcal{C}}&\qquad D\mapsto B(\mathcal{C},\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}WVD\\ (\mathcal{M}on^{w})^{\mathcal{C}}&\rightarrow&(\mathcal{M}on^{w})^{\mathcal{C}}&\qquad D\mapsto B(\mathcal{C},\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}WD\\ \mathop{\rm\mathcal{S}gp}^{\mathcal{C}}&\rightarrow&\mathop{\rm\mathcal{S}gp}^{\mathcal{C}}&\qquad D\mapsto B(\mathcal{C},\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}\overline{W}D\\ ({\mathcal{T}\!op}^{\ast})^{\mathcal{C}}&\rightarrow&({\mathcal{T}\!op}^{\ast})^{\mathcal{C}}&\qquad D\mapsto B(\mathcal{C},\mathcal{C},\mathcal{C})_{+}\wedge_{\mathcal{C}}V^{t}D\\ ({\mathcal{T}\!op}^{w})^{\mathcal{C}}&\rightarrow&({\mathcal{T}\!op}^{w})^{\mathcal{C}}&\qquad D\mapsto B(\mathcal{C},\mathcal{C},\mathcal{C})_{+}\wedge_{\mathcal{C}}D\end{array}

together with the corresponding natural transformations ϵ\epsilon are strong cofibrant replacement functors with respect to the weak equivalences in 𝒲𝒞\mathscr{W}^{\mathcal{C}}. In particular, the localizations of these categories with respect to 𝒲𝒞\mathscr{W}^{\mathcal{C}} exist. (Recall that K+XK_{+}\wedge X is the tensor over 𝒯op\mathcal{T}\!op in 𝒯op{\mathcal{T}\!op}^{\ast}.)

Since Addendum 6.8 applies to our situation in Section 4 we obtain

Theorem 6.11.

The homotopy adjunctions of Theorems 4.4 and 4.5 lift to conatural homotopy adjunctions

(B𝒞):(onw)𝒞\textstyle{(B^{\mathcal{C}})^{\mathcal{H}}:\mathcal{H}(\mathcal{M}on^{w})^{\mathcal{C}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(𝒯opw)𝒞:(Ω𝒞)\textstyle{({\mathcal{T}\!op}^{w})^{\mathcal{C}}:({\Omega^{\prime}}^{\mathcal{C}})^{\mathcal{H}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

and

(B𝒞):on𝒞\textstyle{(B^{\mathcal{C}})^{\mathcal{H}}:\mathcal{H}\mathcal{M}on^{\mathcal{C}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒯op𝒞:(Ω𝒞).\textstyle{{{\mathcal{T}\!op}^{\ast}}^{\mathcal{C}}:({\Omega^{\prime}}^{\mathcal{C}})^{\mathcal{H}}.\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

There are natural adjunction homotopy equivalences

λ(RD,QZ):(onw)𝒞(RD,RΩQZ)(𝒯opw)𝒞(QBRD,QZ)λ(RVD,QVtZ):on𝒞(RVD,RVΩQVtZ)𝒯op𝒞(QVtBRVD,QVtZ)\begin{array}[]{rcl}\lambda(RD,QZ):(\mathcal{M}on^{w})^{\mathcal{C}}(RD,R\Omega^{\prime}QZ)&\rightarrow&({\mathcal{T}\!op}^{w})^{\mathcal{C}}(QBRD,QZ)\\ \lambda(RVD,QV^{t}Z):\mathcal{M}on^{\mathcal{C}}(RVD,RV\Omega^{\prime}QV^{t}Z)&\rightarrow&{{\mathcal{T}\!op}^{\ast}}^{\mathcal{C}}(QV^{t}BRVD,QV^{t}Z)\end{array}

in 𝒯op\mathcal{T}\!op, where (R,ϵ)(R,\epsilon) and (Q,ϵt)(Q,\epsilon^{t}) are the cofibrant replacement functors in (onw)𝒞(\mathcal{M}on^{w})^{\mathcal{C}} respectively (𝒯opw)𝒞({\mathcal{T}\!op}^{w})^{\mathcal{C}} of 6.10. Hence

HoB𝒞:Ho(onw)𝒞Ho(𝒯opw)𝒞:HoΩ𝒞\mathop{\rm Ho}\nolimits B^{\mathcal{C}}:\mathop{\rm Ho}\nolimits(\mathcal{M}on^{w})^{\mathcal{C}}\leftrightarrows\mathop{\rm Ho}\nolimits({\mathcal{T}\!op}^{w})^{\mathcal{C}}:\mathop{\rm Ho}\nolimits\Omega^{\prime\mathcal{C}}

and

HoB𝒞:Ho(on)𝒞Ho(𝒯op)𝒞:HoΩ𝒞\mathop{\rm Ho}\nolimits B^{\mathcal{C}}:\mathop{\rm Ho}\nolimits(\mathcal{M}on)^{\mathcal{C}}\leftrightarrows\mathop{\rm Ho}\nolimits({\mathcal{T}\!op}^{\ast})^{\mathcal{C}}:\mathop{\rm Ho}\nolimits\Omega^{\prime\mathcal{C}}

are genuine adjunctions.

Theorem 6.12.

Let \mathcal{M} be as above. Then the adjoint pair of functors

colim:𝒞:const\mathop{\rm colim}\nolimits:\mathcal{M}^{\mathcal{C}}\leftrightarrows\mathcal{M}:\textrm{const}

induces a conatural adjunction up to homotopy

colim:𝒞:const.\mathop{\rm colim}\nolimits^{\mathcal{H}}:\mathcal{H}\mathcal{M}^{\mathcal{C}}\leftrightarrows\mathcal{H}\mathcal{M}:\textrm{const}^{\mathcal{H}}.

Hence we obtain a genuine adjunction

Hocolim:Ho𝒞Ho:Hoconst\mathop{\rm Ho}\nolimits\mathop{\rm colim}\nolimits:\mathop{\rm Ho}\nolimits\mathcal{M}^{\mathcal{C}}\leftrightarrows\mathop{\rm Ho}\nolimits\mathcal{M}:\mathop{\rm Ho}\nolimits\textrm{const}
Proof.

We have the following sequence of natural homotopy equivalences and homeomorphisms from 𝒞(D,constA)=𝒞(RD,R(constQA))\mathcal{H}\mathcal{M}^{\mathcal{C}}(D,\textrm{const}^{\mathcal{H}}A)=\mathcal{M}^{\mathcal{C}}(RD,R(\textrm{const}QA)) to (colimD,A)\mathcal{H}\mathcal{M}(\mathop{\rm colim}\nolimits^{\mathcal{H}}D,A) =(Q(colimRD),QA)=\mathcal{M}(Q(\mathop{\rm colim}\nolimits RD),QA):

(1)𝒞(RD,R(constQA))ϵ(constQA)𝒞(RD,constQA)(2)(colimRD,QA)(3)εM(colimRD)(Q(colimQD),QA).\begin{array}[]{lrcl}(1)&\mathcal{M}^{\mathcal{C}}(RD,R(\textrm{const}QA))&\xrightarrow{\epsilon(\textrm{const}QA)_{\ast}}&\mathcal{M}^{\mathcal{C}}(RD,\textrm{const}QA)\\ (2)&&\stackrel{{\scriptstyle\cong}}{{\longrightarrow}}&\mathcal{M}(\mathop{\rm colim}\nolimits RD,QA)\\ (3)&&\xrightarrow{\varepsilon_{M}(\mathop{\rm colim}\nolimits RD)^{\ast}}&\mathcal{M}(Q(\mathop{\rm colim}\nolimits QD),QA).\end{array}

The first map is a homotopy equivalence by 6.3, the second one is the adjunction homeomorphism, and the third one is a homotopy equivalence, because εM(colimRD):QcolimRDcolimRD\varepsilon_{M}(\mathop{\rm colim}\nolimits RD):Q\mathop{\rm colim}\nolimits RD\to\mathop{\rm colim}\nolimits RD is a homotopy equivalence in \mathcal{M} by 6.3. ∎

Definition 6.13.

Let \mathcal{M} be a cocomplete 𝒯op\mathcal{T}\!op-enriched tensored category with a class 𝒲\mathscr{W} of weak equivalences containing the homotopy equivalences and equipped with a strong cofibrant replacement functor (Q,ε)(Q,\varepsilon). Then the homotopy colimit functor hocolim:𝒞\mathop{\rm hocolim}\nolimits:\mathcal{M}^{\mathcal{C}}\to\mathcal{M} is defined by

hocolimD=colimRD=B(,𝒞,𝒞)𝒞QD.\mathop{\rm hocolim}\nolimits D=\mathop{\rm colim}\nolimits RD=B(\ast,\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}QD.
Remark 6.14.

In the literature one often finds the homotopy colimit defined by hocolimD=B(,𝒞,𝒞)𝒞D\mathop{\rm hocolim}\nolimits D=B(\ast,\mathcal{C},\mathcal{C})\otimes_{\mathcal{C}}D (e.g. see [13, 18.1.1]). This has historical reasons because homotopy colimits were first defined in categories where all object were cofibrant.

We apply these results to on\mathcal{M}on and prove

Theorem 6.15.

The classifying space functor

B:(on,onw)(𝒯op,𝒯opw)B:(\mathcal{M}on,\mathcal{M}on^{w})\to({\mathcal{T}\!op}^{\ast},{\mathcal{T}\!op}^{w})

preserves homotopy colimits up to genuine homotopy equivalences. More precisely, for any diagram D:𝒞onD:\mathcal{C}\to\mathcal{M}on the natural map

hocolim𝒯opBDB(hocolimonD)\mathop{\rm hocolim}\nolimits^{{\mathcal{T}\!op}^{\ast}}BD\to B(\mathop{\rm hocolim}\nolimits^{\mathcal{M}on}D)

is a homotopy equivalence.

Proof.

By definition of the homotopy colimit functor it suffices to prove the well-pointed case.

Consider the diagram

(onw)𝒞\textstyle{(\mathcal{M}on^{w})^{\mathcal{C}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γ(onw)𝒞\scriptstyle{\gamma_{(\mathcal{M}on^{w})^{\mathcal{C}}}}B𝒞\scriptstyle{B^{\mathcal{C}}}Ho(onw)𝒞\textstyle{\mathop{\rm Ho}\nolimits(\mathcal{M}on^{w})^{\mathcal{C}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hocolim\scriptstyle{\mathop{\rm Ho}\nolimits\mathop{\rm colim}\nolimits}HoB𝒞\scriptstyle{\mathop{\rm Ho}\nolimits B^{\mathcal{C}}}Hoon\textstyle{\mathop{\rm Ho}\nolimits\mathcal{M}on\ignorespaces\ignorespaces\ignorespaces\ignorespaces}HoB\scriptstyle{\mathop{\rm Ho}\nolimits B}(𝒯opw)𝒞\textstyle{({\mathcal{T}\!op}^{w})^{\mathcal{C}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γ𝒯opw\scriptstyle{\gamma_{{\mathcal{T}\!op}^{w}}}Ho(𝒯op)𝒞\textstyle{\mathop{\rm Ho}\nolimits({\mathcal{T}\!op}^{\ast})^{\mathcal{C}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hocolim\scriptstyle{\mathop{\rm Ho}\nolimits\mathop{\rm colim}\nolimits}Ho𝒯opw\textstyle{\mathop{\rm Ho}\nolimits{\mathcal{T}\!op}^{w}}

and recall that Hocolim\mathop{\rm Ho}\nolimits\mathop{\rm colim}\nolimits is induced by the homotopy colimit functor. Since BB preserves weak equivalences in the well-pointed case, B𝒞B^{\mathcal{C}} induces HoB𝒞\mathop{\rm Ho}\nolimits B^{\mathcal{C}} so that the left square commutes up to natural equivalence. The right square commutes up to natural equivalence, because the corresponding square of right adjoints commutes. Hence, for any diagram DD in on\mathcal{M}on, the natural map

hocolim𝒯opwBDB(hocolimonwD)\mathop{\rm hocolim}\nolimits^{{\mathcal{T}\!op}^{w}}BD\to B(\mathop{\rm hocolim}\nolimits^{\mathcal{M}on^{w}}D)

becomes an isomorphism in Ho𝒯opw=π𝒯opw\mathop{\rm Ho}\nolimits{\mathcal{T}\!op}^{w}=\pi{\mathcal{T}\!op}^{w}. ∎

References

  • [1] G. Allaud, De-looping homotopy equivalences, Arch. Math. 23 (1972), 167 - 169.
  • [2] T. Barthel and E. Riehl, On the construction of functorial factorizations for model categories, Preprint, arXiv:1204.5427v1 [math.AT] (2012).
  • [3] C. Berger, I. Moerdijk, The Boardman-Vogt resolution of operads in monoidal model categories, Topology 45 (2006), 807-849.
  • [4] M. Brinkmeier, Strongly homotopy-commutative monoids revisited, Documenta Mathematica 5 (2000), 613-624.
  • [5] J.M. Boardman, R.M. Vogt, Homotopy invariant algebraic structures on topological spaces, Lecture Notes in Mathematics 347, Springer, 1973.
  • [6] J. M. Boardman and R. M. Vogt, Tensor products of theories, applications to infinite loop spaces, J. Pure Appl. Algebra 14 (1979), 117-129.
  • [7] F. Borceux, Handbook of categorical algebra 2, Encyclopedia of Mathematics and its Applications 51, Cambridge University Press 1994.
  • [8] M. Cole, Many homotopy categories are homotopy categories, Topol. Appl. 153 (2006), 1084-1099.
  • [9] T. tom Dieck, K.H. Kamps, D. Puppe, Homotopietheorie, Springer Lecture Notes in Mathematics 157 (1970).
  • [10] Z. Fiedorowicz, Classifying spaces of topological monoids and categories, Amer. J. Math 106 (1984), 301-350.
  • [11] M. Fuchs, Verallgemeinerte Homotopie-Homomorphismen und klassifizierende R”aume, Math. Annalen 161 (1965), 197-230.
  • [12] M. Fuchs, Homotopy equivalences in equivariant topology, Proc. Amer. Math. Soc. 58 (1976), 347-352.
  • [13] P.S. Hirschhorn, Model categories and their localizations, Mathematical Surveys and Monographs 99, Amer. Math. Soc. (2002).
  • [14] P.G. Goerss, J.F. Jardine, Simplicial homotopy theory, Progress in Mathematics 174, Birkhäuser Verlag, Basel (1999).
  • [15] I.M. James, Reduced product spaces, Ann. of Math. 62 (1955), 170-197.
  • [16] G.M. Kelly, Basic concepts of enriched category theory, London Math. Soc. Lecture Note Series 64, Cambridge University Press (1982).
  • [17] M. Klioutch, Vergleich zweier verschiedener Konzepte von Homotopiehomomorphismen von Monoiden, Diploma Thesis, University of Osnabrück (2008).
  • [18] R.J. Milgram, The bar construction and abelian HH-spaces, Illinois J. Math. 11 (1967), 242-250.
  • [19] T. Panov, N. Ray, R.M. Vogt, Colimits, Stanley-Reisner algebras, and loop spaces, Progress in Math. 215 (2003), 261-291.
  • [20] D. Puppe, Some well known weak homotopy equivalences are genuine homotopy equivalences, Symposia Mathematica 5, Istituto Nazionale di Alta Mathematica (1971), 363 - 374.
  • [21] D.G. Quillen, Homotopical algebra, Springer Lecture Notes in Math. 43 (1967).
  • [22] E. Schwamberger, R.M. Vogt, Dold spaces in homotopy theory, Algebraic & Geometric Topology (2009), 1585-1622.
  • [23] G. Segal, Categories and cohomology theories, Topology 13 (1974), 293-312.
  • [24] A. Strøm, Note on cofibrations II, Math. Scand. 22 (1968), 130-142.
  • [25] A. Strøm, The homotopy category is a homotopy category, Arch. Math. 23 (1972), 435-441.
  • [26] M. Sugawara, On the homotopy-commutativity of groups and loop spaces, Mem. Coll. Sci. Univ. Kyoto Ser. A. 33 (1960), 257-269.
  • [27] R.M. Vogt, Cofibrant operads and universal EE_{\infty}-operads, Top. Appl. 133 (2003), 69-87.
  • [28] R.M. Vogt, The HELP-Lemma and its converse in Quillen model categories, J. Homotopy and Related Structures 6 (2011), 115-118.