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

0002010 Mathematics Subject Classification: 55T05, 55T20, 55N25, 55P35, 58A40, 58A10
Key words and phrases. Spectral sequence, Borel construction, free loop space, diffeology. Department of Mathematical Sciences, Faculty of Science, Shinshu University, Matsumoto, Nagano 390-8621, Japan e-mail:kuri@math.shinshu-u.ac.jp

On multiplicative spectral sequences for nerves and the free loop spaces

Katsuhiko KURIBAYASHI
Abstract.

We construct a multiplicative spectral sequence converging to the cohomology algebra of the diagonal complex of a bisimplicial set with coefficients in a field. The construction provides a spectral sequence converging to the cohomology algebra of the classifying space of a category internal to the category of topological spaces. By applying the machinery to a Borel construction, we explicitly determine the mod pp cohomology algebra of the free loop space of the real projective space for each odd prime pp. This example is emphasized as an important computational case. Moreover, we represent generators in the singular de Rham cohomology algebra of the diffeological free loop space of a non-simply connected manifold MM with differential forms on the universal cover of MM via Chen’s iterated integral map.

1. Introduction

A stack is a generalization of a sheaf. More precisely, it is a weak 2-functor from a site to the category of groupoids which satisfies the gluing conditions on objects and morphisms. In particular, differentiable and topological stacks are obtained by Lie and topological groupoids, respectively, via the stakifications of prestacks associated with such groupoids; see [5, Section 2], [20] and [6, Chapter 1] for more details. In [39], loop stacks are investigated by using diffeological groupoids; see [22] and Appendix B for diffeological spaces. It is worthwhile mentioning that the study of orbifolds is developed with those Lie groupoid presentations; see [1, 35].

In [4], Behrend introduced the de Rham cohomology and the singular homology for a differentiable stack, which are those of the classifying space of a Lie groupoid presenting the stack. In particular, the cohomology is invariant under Morita equivalence. Thus, such results on stacks and groupoids motivate us to consider how to compute the cohomology of such a classifying space; see Remark 3.3.

This manuscript aims to introduce multiplicative spectral sequences computing the cohomology algebras of the classifying spaces of topological categories with coefficients in a field 𝕂{\mathbb{K}}; see Theorems 2.1 and 2.2. Although there is no application for a stack in this study, the product structure demonstrates its power in giving more additional structure to the spectral sequence and in the computations of the cohomology algebras of non-simply connected spaces. To be more precise, let LXLX denote the free loop space of a space XX, BGBG the classifying space of a finite group GG and EG×GMEG\times_{G}M the Borel construction of a GG-space MM. Then, Remark 5.11 enables one to obtain a multiplicative spectral sequence endowed with an H(LBG;𝕂)H^{*}(LBG;{\mathbb{K}})-module structure converging to the cohomology algebra of L(EG×GM)L(EG\times_{G}M). We also refer the reader to Corollaries 5.4 and 5.7, Theorem 5.2 and Proposition 5.9 each of which provides a method for computing the Borel cohomology algebra of a GG-space.

As a computational example, we explicitly determine the cohomology algebra of the free loop space of the real projective space with coefficients in /p{\mathbb{Z}}/p and {\mathbb{Q}}, where pp is an odd prime; see Theorem 4.1. To our knowledge, this result is novel. Despite the lack of nontrivial information in the rational cohomology of the even-dimensional projective space, the cohomology algebra of the free loop space can be beneficial.

A diffeological space is a generalization of a manifold. Therefore, it is crucial to consider smooth (homotopy) invariants of the generalized objects. In particular, the de Rham complex and its singular variant of a diffeological space are introduced in [43] and [29], respectively. In this manuscript, we moreover attempt to represent generators in the singular de Rham cohomology of the free loop space of a non-simply connected manifold MM by using differential forms on the universal cover of MM within the framework of diffeology; see Theorem 6.1 and subsequent comments. As a consequence, because of Theorem 4.1, we can describe generators of the singular de Rham cohomology algebra of the diffeological free loop space of nn-dimensional real projective space with the volume form on the sphere SnS^{n} via Chen’s iterated integral map; see Theorem 6.4. While the original iterated integrals due to Chen work well for the cohomology of the free loop space of a simply connected manifold, the diffeological argument above shows that we can also deal with non-simply connected manifolds in Chen’s theory for free loop spaces. That is an advantage of considering manifolds in diffeology.

An outline of this manuscript is as follows. Section 2 introduces a spectral sequence with a multiplicative structure for a bisimplicial set. The spectral sequence gives rise to those for topological categories, Borel constructions, and diffeological categories. In Section 3, we establish Theorems 2.1 and 2.2. Section 4 describes the computational example mentioned above. In Section 5, by generalizing the computations in Section 4, we present results of the cohomology algebras of Borel constructions. Moreover, we consider spectral sequences for transformation groupoids including an inertia groupoid. In Section 6, we investigate the singular de Rham cohomology of the diffeological free loop space of a non-simply connected manifold by applying results concerning Borel constructions in Section 5.

Appendix A gives a weak homotopy equivalence between a Borel construction and the free loop space of a quotient space, which is used in the computation in Section 4. In Appendix B, we briefly recall the category of diffeological spaces together with adjoint functors between the category of topological spaces. Appendix C proves that the diffeological free loop space of smooth maps from S1S^{1} is weak homotopy equivalent to the pullback of the evaluation map from a path space along the diagonal map in the category of diffeological spaces. This result is critical for proving Theorem 6.1.

2. A multiplicative spectral sequence for a bisimplicial set and its variants

We introduce multiplicative spectral sequences associated with a bisimplicial set by explicitly describing the product structure.

For a simplicial set KK, we denote by C(K;𝕂)C^{*}(K;{\mathbb{K}}) and H(K;𝕂)H^{*}(K;{\mathbb{K}}) the cochain algebra and the cohomology algebra of KK with coefficients in a field 𝕂{\mathbb{K}}, respectively. Let Sing(X)\text{Sing}_{\text{\tiny{$\bullet$}}}(X) be the singular simplicial set of a space XX. We may write Hsing(X;𝕂)H^{*}_{\text{s}ing}(X;{\mathbb{K}}) or simply H(X;𝕂)H^{*}(X;{\mathbb{K}}) for the singular cohomology algebra H(Sing(X);𝕂)H^{*}(\text{Sing}_{\text{\tiny{$\bullet$}}}(X);{\mathbb{K}}).

Let S=SS=S_{\text{\tiny{$\bullet$}}\text{\tiny{$\bullet$}}} be a bisimplicial set. Then, the vertical face maps give rise to a differential (dv)(d^{v})^{*} by the alternating sum on the graded algebra Cp,:=C(Sp;𝕂)C^{p,*}:=C^{*}(S_{p\text{\tiny{$\bullet$}}};{\mathbb{K}}) with the usual cup product \cup for any p0p\geq 0. Moreover, we have a double complex {{Cp,q}p,q0,(dh),(dv)}\{\{C^{p,q}\}_{p,q\geq 0},(d^{h})^{*},(d^{v})^{*}\}, where the differential (dh)(d^{h})^{*} is induced by the horizontal face maps of the bisimplicial set SS_{\text{\tiny{$\bullet$}}\text{\tiny{$\bullet$}}}. A product T\cup_{T} on the total complex TotC,\text{Tot}\,C^{*,*} is defined by

(2.1) ωTη=(1)qp(dp+1hdp+ph)ω(d0hdp1h)η\displaystyle\omega\cup_{T}\eta=(-1)^{qp^{\prime}}(d_{p+1}^{h}\cdots d_{p+p^{\prime}}^{h})^{*}\omega\cup(d_{0}^{h}\cdots d_{p-1}^{h})^{*}\eta

for ωCp,q\omega\in C^{p,q} and ηCp,q\eta\in C^{p^{\prime},q^{\prime}}. Observe that the differential on TotC,\text{Tot}\,C^{*,*} is given by δ(ω)=(dh)(ω)+(1)p(dv)(ω)\delta(\omega)=(d^{h})^{*}(\omega)+(-1)^{p}(d^{v})^{*}(\omega) for ωCp,q\omega\in C^{p,q}. Thus, we obtain a spectral sequence {Er,,dr}\{E_{r}^{*,*},d_{r}\} associated with the total complex.

Theorem 2.1.

The first quadrant spectral sequence {Er,,dr}\{E_{r}^{*,*},d_{r}\} with the multiplicative structure defined by (2.1) converges to H(diagS;𝕂)H^{*}(\text{\em diag}\,S_{\text{\tiny{$\bullet$}}\text{\tiny{$\bullet$}}};{\mathbb{K}}) as an algebra with

E2,H(H(S,(dv)),(dh))E_{2}^{*,*}\cong H^{*}(H^{*}(S_{\text{\tiny{$\bullet$}}\text{\tiny{$\bullet$}}},(d^{v})^{*}),(d^{h})^{*})

as a bigraded algebra, where diagS\text{\em diag}\,S_{\text{\tiny{$\bullet$}}\text{\tiny{$\bullet$}}} denotes the diagonal simplicial set of SS_{\text{\tiny{$\bullet$}}\text{\tiny{$\bullet$}}}; see, for example, [18, Chapter IV, 1]. Therefore, for a simplicial space X={Xn,i,sj}X_{\text{\tiny{$\bullet$}}}=\{X_{n},\partial_{i},s_{j}\}, one has a first quadrant spectral sequence converging to the singular cohomology algebra Hsing(X;𝕂)H^{*}_{\text{s}ing}(||X_{\text{\tiny{$\bullet$}}}||;{\mathbb{K}}) with E2,H(Hsing(X,𝕂),i(1)ii)E_{2}^{*,*}\cong H^{*}\big{(}H^{*}_{\text{s}ing}(X_{\text{\tiny{$\bullet$}}},{\mathbb{K}}),\sum_{i}(-1)^{i}\partial_{i}^{*}\big{)} as a bigraded algebra, where ||||||\ || denotes the fat geometric realization in the sense of Segal [40].

A prototype of the spectral sequence in Theorem 2.1 is one stated in [34, II 6.8. Corollary]; see also [40, Proposition (5.1)] for a generalized cohomology. The novelty here is that we explicitly provide an algebraic structure in the spectral sequence.

Let 𝒞=[C1tsC0]{\mathcal{C}}=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 9.72014pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-9.72014pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[C_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 12.95625pt\raise-7.30554pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.15279pt\hbox{$\scriptstyle{t}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 24.72014pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 12.57951pt\raise 6.65971pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.50694pt\hbox{$\scriptstyle{s}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 24.72014pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 24.72014pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{C_{0}]}$}}}}}}}\ignorespaces}}}}\ignorespaces be a category internal to 𝖳𝗈𝗉\mathsf{Top} the category of topological spaces; that is, structure maps containing the source map ss and the target map tt are continuous. We may drop the maps ss and tt in the notation of an internal category. The nerve functor gives rise to a cosimplicial cohain complex

nC(Nerven𝒞,𝕂)n\mapsto C^{*}(\text{Nerve}_{n}{\mathcal{C}},{\mathbb{K}})

and then this induces a cosimplicial abelian group nHq(Nerven𝒞,𝕂)n\mapsto H^{q}(\text{Nerve}_{n}{\mathcal{C}},{\mathbb{K}}) for any qq. In what follows, for a cosimplicial abelian group AA^{\bullet}, we denote by HΔ(A)H_{\Delta}(A^{\bullet}) the cohomology of AA^{\bullet}. Let B𝒞\text{B}{\mathcal{C}} be the classifying space, namely, B𝒞=Nerve𝒞\text{B}{\mathcal{C}}=||\text{Nerve}_{\bullet}{\mathcal{C}}||, which is the fat geometric realization of the simplicial space Nerve𝒞\text{Nerve}_{\bullet}{\mathcal{C}}; see [40]. The multiplicative spectral sequence in Theorem 2.1 is adaptable to many situations. The following results illustrate it with crucial examples.

Theorem 2.2.

(cf. [4], [34, II 6.8.Corollary, IV 4.1.Theorem], [17, Corollary 3.10])

i) Let 𝒞=[C1C0]{\mathcal{C}}=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 9.72014pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-9.72014pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[C_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 24.72014pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 24.72014pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 24.72014pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{C_{0}]}$}}}}}}}\ignorespaces}}}}\ignorespaces be a category internal to 𝖳𝗈𝗉\mathsf{Top}. Then there exists a spectral sequence {Er,,dr}\{E_{r}^{*,*},d_{r}\} converging to H(B𝒞;𝕂)H^{*}(\text{\em B}{\mathcal{C}};{\mathbb{K}}) as an algebra with

E2p,qHΔp(Hq(Nerve𝒞;𝕂)).E_{2}^{p,q}\cong H^{p}_{\Delta}(H^{q}(\text{\em Nerve}_{\text{\tiny{$\bullet$}}}{\mathcal{C}};{\mathbb{K}})).

ii) Let GG be a topological group and XX a GG-space. Then there exists a spectral sequence converging to the Borel cohomology HG(X;𝕂):=H(EG×GX;𝕂)H^{*}_{G}(X;{\mathbb{K}}):=H^{*}(EG\times_{G}X;{\mathbb{K}}) as an algebra with E2p,qHp(Hq(Nerve𝒢;𝕂)).E_{2}^{p,q}\cong H^{p}(H^{q}(\text{\em Nerve}_{\text{\tiny{$\bullet$}}}{\mathcal{G}};{\mathbb{K}})). Here 𝒢:=[G×XX]{\mathcal{G}}:=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 18.96591pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-18.96591pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[G\times X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 33.96591pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 33.96591pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 33.96591pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{X]}$}}}}}}}\ignorespaces}}}}\ignorespaces denotes the transformation groupoid associated to the GG-space XX whose source map and target map are the projection on the first factor and the action of GG on XX, respectively. In particular, one has an isomorphism

E2p,qCotorH(G)p,q(𝕂,H(X))E_{2}^{p,q}\cong\text{\em Cotor}^{p,q}_{H^{*}(G)}({\mathbb{K}},H^{*}(X))

provided H(G)H^{*}(G) and H(X)H^{*}(X) are locally finite; see Remark 5.11 for a more structure of the spectral sequence in the case where GG is a finite group.

iii) Let 𝒞=[C1C0]{\mathcal{C}}=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 9.72014pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-9.72014pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[C_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 24.72014pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 24.72014pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 24.72014pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{C_{0}]}$}}}}}}}\ignorespaces}}}}\ignorespaces be a category internal to the category 𝖣𝗂𝖿𝖿\mathsf{Diff} of diffeological spaces; that is, the sets C0C_{0} and C1C_{1} of objects and morphisms are diffeological spaces, respectively, and structure maps in 𝒞{\mathcal{C}} are smooth; see Section B. Let ADR(SD(Nerve(𝒞)))A_{DR}^{*}(S^{D}_{\text{\tiny{$\bullet$}}}(\text{\em Nerve}_{\text{\tiny{$\bullet$}}}({\mathcal{C}}))) denote the cosimplicial de Rham complex of the nerve of 𝒞{\mathcal{C}} introduced in [29, §2]; see Remark 3.2 and Appendix B. Then, there exists a spectral sequence converging to H(diagSD(Nerve(𝒞));)H^{*}(\text{\em diag}\,S^{D}_{\text{\tiny{$\bullet$}}}(\text{\em Nerve}_{\text{\tiny{$\bullet$}}}({\mathcal{C}}));{\mathbb{R}}) as an algebra with

E2p,qHΔp(HDRq(Nerve(𝒞))).E_{2}^{p,q}\cong H^{p}_{\Delta}(H_{DR}^{q}(\text{\em Nerve}_{\text{\tiny{$\bullet$}}}({\mathcal{C}}))).

Here HDR(Nerve(𝒞))H_{DR}^{*}(\text{\em Nerve}_{\text{\tiny{$\bullet$}}}({\mathcal{C}})) denotes the cohomology of the complex ADR(SD(Nerve(𝒞)))A_{DR}^{*}(S^{D}_{\text{\tiny{$\bullet$}}}(\text{\em Nerve}_{\text{\tiny{$\bullet$}}}({\mathcal{C}}))).

One might expect that the target of the spectral sequence in Theorem 2.2 iii) is replaced with the cohomology of a more familiar object. We discuss the topic in Remarks 3.1, 3.2 and 3.3.

Remark 2.3.

Let 𝒞{\mathcal{C}} be a category internal to 𝖳𝗈𝗉\mathsf{Top}. Then, by using the polynomial de Rham functor APLA_{{PL}} (see, for example, [15, II 10 (a), (b) and (c)] and Appendix B.1) instead of the singular cochain functor in Theorem 2.2 i), we have a spectral sequence converging to the rational cohomology of B𝒞\text{B}{\mathcal{C}}. In this case, an appropriate Sullivan model for each Nerven𝒞\text{Nerve}_{n}{\mathcal{C}} may be useful when computing the E2E_{2}-term as an algebra; see, for example, [15, Part II] for Sullivan models. Observe that the product T\wedge_{T} on APL(Nerve𝒞)A_{PL}(\text{Nerve}_{\text{\tiny{$\bullet$}}}{\mathcal{C}}) is of the form

(2.2) ωTη=(1)qpπ1ωπ2η\displaystyle\omega\wedge_{T}\eta=(-1)^{qp^{\prime}}\pi_{1}^{*}\omega\wedge\pi_{2}^{*}\eta

for ωAPLq(Nervep𝒞)\omega\in A_{{PL}}^{q}(\text{Nerve}_{p}{\mathcal{C}}) and ηAPLq(Nervep𝒞)\eta\in A_{{PL}}^{q^{\prime}}(\text{Nerve}_{p^{\prime}}{\mathcal{C}}), where π1\pi_{1} and π2\pi_{2} are maps assigning (f1,,fp)(f_{1},\dots,f_{p}) and (fp+1,,fp+p)(f_{p+1},\dots,f_{p+p^{\prime}}) to (f1,,fp,fp+1,,fp+p)(f_{1},\dots,f_{p},f_{p+1},\dots,f_{p+p^{\prime}}), respectively; see [4, (6)].

The spectral sequences in Theorem 2.2 i) and ii) are variants of that in Theorem 2.1. Therefore, each of them converges to the cohomology of the total complex TotC(Nerve(𝒞);𝕂)\text{Tot}\,C^{*}(\text{Nerve}_{\text{\tiny{$\bullet$}}}({\mathcal{C}});{\mathbb{K}}) as an algebra. The spectral sequence described in Theorem 2.2 iii) is also constructed by applying Theorem 2.1. Then, it converges to the cohomology algebra of the total complex TotADR(SD(Nerve(𝒞)))\text{Tot}\,A_{DR}^{*}(S^{D}_{\text{\tiny{$\bullet$}}}(\text{Nerve}_{\text{\tiny{$\bullet$}}}({\mathcal{C}}))) with the same product T\wedge_{T} as in (2.2). These targets of the convergences are isomorphic to the cohomology algebras described in Theorem 2.2. This follows from the proof of Theorem 2.2. Thus, it may be possible to reconstruct the algebra structure of the target from that of the EE_{\infty}-term in the same way as in [30, Section 7] with the formula of the product; that is, we may solve extension problems in the spectral sequences in Theorem 2.2.

Moreover, the formulae (2.1) and (2.2) enable us to explicitly consider the multiplication on the E2E_{2}-term of the spectral sequence; see the proof of Proposition 4.4 in which the cohomology algebra of the free loop space of a Borel construction is investigated for low degrees.

The spectral sequence in [17, Corollary 3.10] converging to the homology of the Borel construction EG×GMEG\times_{G}M for a GG-space MM has a differential coalgebra structure, and the condition on local finiteness for the homology groups H(G)H_{*}(G) and H(M)H_{*}(M) is not required in constructing the spectral sequence. Indeed, the torsion product of chain complexes is used in the construction. In contrast, in proving Theorem 2.2 ii), we consider the nerve of a category internal to 𝖳𝗈𝗉\mathsf{Top} and apply the Künneth theorem. Then, the local finiteness for cohomology groups H(G)H^{*}(G) and H(X)H^{*}(X) is required in our theorem. We stress that the multiplicative structure in our spectral sequence is given explicitly by (2.1) without an argument on dualizing the homology.

3. Constructions of the spectral sequences

The goal of this section is to construct the spectral sequences described in Theorems 2.1 and 2.2.

Proof of Theorem 2.1.

We observe that the decreasing filtration {FpTotC,}p0\{F^{p}\text{Tot}\,C^{*,*}\}_{p\geq 0} defined by (FpTotC,)n=i+j=n,ipCi,j(F^{p}\text{Tot}\,C^{*,*})^{n}=\bigoplus_{i+j=n,i\geq p}\,C^{i,j} provides the spectral sequence. Since the product T\cup_{T} in (2.1) preserves the filtration, it induces a multiplicative structure in the spectral sequence. Thus, for proving the first assertion, it suffices to show that the product in H(TotC,)H^{*}(\text{Tot}\,C^{*,*}) given by T\cup_{T} is compatible with the cup product on H(diagS;𝕂)H^{*}(\text{diag}\,S_{\text{\tiny{$\bullet$}}\text{\tiny{$\bullet$}}};{\mathbb{K}}) under an appropriate isomorphism between the cohomology groups. To this end, we use an argument with universal δ\delta-functors; see [45, Chapter 2].

The simplicial identities for the horizontal face maps of SS_{\text{\tiny{$\bullet$}}\text{\tiny{$\bullet$}}} enable us to deduce that the cup product T\cup_{T} on TotC,\text{Tot}\,C^{*,*} is a cochain map. A direct computation gives the fact. We show that a diagram

(3.1) H(TotC,)H(TotC,)\textstyle{H^{*}(\text{Tot}\,C^{*,*})\otimes H^{*}(\text{Tot}\,C^{*,*})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H(AW)H(AW)\scriptstyle{H(\text{AW}^{*})\otimes H(\text{AW}^{*})}\scriptstyle{\cong}T\scriptstyle{\cup_{T}}H(TotC,)\textstyle{H^{*}(\text{Tot}\,C^{*,*})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H(AW)\scriptstyle{H(\text{AW}^{*})}\scriptstyle{\cong}dualπ(diag𝕂(S))dualπ(diag𝕂(S))\textstyle{\text{dual}\,\pi_{*}(\text{diag}\,{\mathbb{K}}(S_{\text{\tiny{$\bullet$}}\text{\tiny{$\bullet$}}}))\otimes\text{dual}\,\pi_{*}(\text{diag}\,{\mathbb{K}}(S_{\text{\tiny{$\bullet$}}\text{\tiny{$\bullet$}}}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cup}dualπ(diag𝕂(S))\textstyle{\text{dual}\,\pi_{*}(\text{diag}\,{\mathbb{K}}(S_{\text{\tiny{$\bullet$}}\text{\tiny{$\bullet$}}}))}

is commutative, where \cup is the usual cup product and AW denotes the Alexander-Whitney map; see, for example, [45, 8.5.4]. The diagram

(3.2) H(dual(TotCA))H(dual(TotCA))\textstyle{H_{*}(\text{dual}\,(\text{Tot}\,CA))\otimes H_{*}(\text{dual}\,(\text{Tot}\,CA))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H(AW)H(AW)\scriptstyle{H(\text{AW}^{*})\otimes H(\text{AW}^{*})}H(dual(TotCA)dual(TotCA))\textstyle{H_{*}(\text{dual}\,(\text{Tot}\,CA)\otimes\text{dual}\,(\text{Tot}\,CA))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H(AWAW)\scriptstyle{H(\text{AW}^{*}\otimes\text{AW}^{*})}π(dualdiagA)π(dualdiagA)\textstyle{\pi_{*}(\text{dual}\,\text{diag}A)\otimes\pi_{*}(\text{dual}\,\text{diag}A)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π(dualdiagAdualdiagA))\textstyle{\pi_{*}(\text{dual}\,\text{diag}A\otimes\text{dual}\,\text{diag}A))}

is commutative, where the horizontal maps are the canonical ones, AA is the bisimplicial vector space 𝕂(S){\mathbb{K}}(S_{\text{\tiny{$\bullet$}}\text{\tiny{$\bullet$}}}) and CC denotes the double complex functor; see [45, 8.5]. Then, in order to prove the commutativity of the diagram (3.1), we show that the diagram

(3.3) H(dual(TotCA)dual(TotCA))\textstyle{H_{*}(\text{dual}\,(\text{Tot}\,CA)\otimes\text{dual}\,(\text{Tot}\,CA))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H(AWAW)\scriptstyle{H(\text{AW}^{*}\otimes\text{AW}^{*})}\scriptstyle{\cong}T\scriptstyle{\cup_{T}}H(dual(TotCA))\textstyle{H_{*}(\text{dual}\,(\text{Tot}\,CA))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H(AW)\scriptstyle{H(\text{AW}^{*})}\scriptstyle{\cong}π(dualdiagAdualdiagA))\textstyle{\pi_{*}(\text{dual}\,\text{diag}A\otimes\text{dual}\,\text{diag}A))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cup}π(dualdiag𝕂(S))\textstyle{\pi_{*}(\text{dual}\,\text{diag}\,{\mathbb{K}}(S_{\text{\tiny{$\bullet$}}\text{\tiny{$\bullet$}}}))}

is commutative. It is proved that F():=H(dual(TotC())dual(TotC()))F(\ ):=H_{*}(\text{dual}\,(\text{Tot}\,C(\ ))\otimes\text{dual}\,(\text{Tot}\,C(\ ))) and π(dualdiag())\pi_{*}(\text{dual}\,\text{diag}(\ )) are universal δ\delta-functors from the category of bisimplicial 𝕂{\mathbb{K}}-vector spaces to the opposite category of graded 𝕂{\mathbb{K}}-vector spaces. In fact, functors dual()\text{dual}(\ ) and dual()dual()\text{dual}(\ )\otimes\text{dual}(\ ) are exact and preserve projectives. Moreover, functors TotC()\text{Tot}\,C(\ ) and diag()\text{diag}(\ ) are also exact and preserve projectives; see [45, 8.5.2 and the proof of 8.5.1]. Thus, it follows that

π(dualdiag())\displaystyle\pi_{*}(\text{dual}\,\text{diag}(\ ))\!\!\! =\displaystyle= (Lπ0)(dualdiag)()=L(π0dualdiag)()and\displaystyle\!\!\!(L_{*}\pi_{0})\circ(\text{dual}\circ\text{diag})(\ )=L_{*}(\pi_{0}\circ\text{dual}\circ\text{diag})(\ )\ \ \text{and}
F()\displaystyle F(\ )\!\!\! =\displaystyle= (LH0)(dual(TotC())dual(TotC()))\displaystyle\!\!\!(L_{*}H_{0})(\text{dual}\,(\text{Tot}\,C(\ ))\otimes\text{dual}\,(\text{Tot}\,C(\ )))
=\displaystyle= (LH0)(dualdual(TotC))()\displaystyle\!\!\!(L_{*}H_{0})\circ(\text{dual}\otimes\text{dual}\circ\,(\text{Tot}\,C))(\ )
=\displaystyle= L(H0dualdualTotC))().\displaystyle\!\!\!L_{*}(H_{0}\circ\text{dual}\otimes\text{dual}\circ\text{Tot}\,C))(\ ).

The result [45, Theorem 2.4.7] allows us to deduce that F()F(\ ) and π(dualdiag())\pi_{*}(\text{dual}\,\text{diag}(\ )) are universal δ\delta-functors.

Since the map H0(AW)H_{0}(\text{AW}^{*}) is induced by the identity map on A00A_{00} and the product T\cup_{T} is nothing but the cup product, it follows that the diagram (3.3) is commutative on H0H_{0}. Therefore, the universality of the functor F()F(\ ) enables us to conclude that the diagram (3.3) is commutative. Thus, the double complex TotC,\text{Tot}\,C^{*,*} induces the multiplicative spectral sequence in the assertion.

In order to prove the latter half of the assertion, we deal with bisimplicial sets and their geometric realizations. For a simplicial space XX_{\text{\tiny{$\bullet$}}}, we see that |(|Sing(X)|)||diagSing(X)||(|\text{Sing}_{\text{\tiny{$\bullet$}}^{\prime}}(X_{\text{\tiny{$\bullet$}}})|_{\text{\tiny{$\bullet$}}^{\prime}})|\simeq|\text{diag}\,\text{Sing}_{\text{\tiny{$\bullet$}}^{\prime}}(X_{\text{\tiny{$\bullet$}}})| by the Eilenberg-Zilber theorem; see, for example, [38, Lemma, page 94] and [16, 7. Theorem].

Let YY denote the simplicial space |Sing(X)||\text{Sing}_{\text{\tiny{$\bullet$}}^{\prime}}(X_{\text{\tiny{$\bullet$}}})|_{\text{\tiny{$\bullet$}}^{\prime}} which is the geometric realization with respect to indices \text{\tiny{$\bullet$}}^{\prime}. Since YY is good in the sense that each degeneracy map is a closed cofibration, it follows that there exists a natural homotopy equivalence Y|Y|||Y||\stackrel{{\scriptstyle\simeq}}{{\longrightarrow}}|Y|; see [40, Proposition A.1. (iv)]. Moreover, the counit of the geometric realization functor gives a natural weak homotopy equivalence YnwXnY_{n}\stackrel{{\scriptstyle\simeq_{w}}}{{\longrightarrow}}X_{n} for each nn. Therefore, the equivalence induces a homology isomorphism YX||Y||\to||X||; see [14, Lemma 5.16]. It turns out that H(diagSing(X),𝕂)H(X,𝕂)H^{*}(\text{diag}\,\text{Sing}(X_{\text{\tiny{$\bullet$}}}),{\mathbb{K}})\cong H^{*}(||X||,{\mathbb{K}}) as an algebra by a natural map; see [10, Lemmas 1.2 and 1.3] for a homotopical proof of the fact. ∎

Remark 3.1.

One may expect a version of Theorem 2.1 for a simplicial diffeological space. The isomorphism in the proof of the latter half of the theorem appears to be well known. To obtain this fact, we take advantage of the Eilenberg–Zilber theorem, which is also applied in [10]. The key to proving the powerful theorem is the use of the homeomorphism |K×Δ[n]||K|×|Δ[n]||K\times\Delta[n]|\cong|K|\times|\Delta[n]| for a simplicial set KK and the standard simplicial set Δ[n]\Delta[n]. However, the diffeological realization functor ||D|\ |_{D} in the sense of Kihara [24, Remark 22.1] or Christensen and Wu [13, Proposition 4.13] does not preserve the product even if KK is the standard simplicial set in the example above. Thus, we cannot prove verbatim a diffeological version of Theorem 2.1. Indeed, it is not easy to replace the target of the spectral sequence in Theorem 2.2 iii) with a more familiar one for a general category internal to 𝖣𝗂𝖿𝖿\mathsf{Diff}. To explain this inconvenience, we have described the proof of the isomorphism.

Proof of Theorem 2.2.

The assertion i) follows from the direct application of Theorem 2.1. As for the assertion iii), we recall the result [29, Proposition 3.4] which yields that for any simplicial set KK, there exists a sequence of quasi-isomorphisms of cochain algebras between C(K;𝕂)C^{*}(K;{\mathbb{K}}) and ADR(K)A_{DR}^{*}(K). Then, by applying the first half of Theorem 2.1 to the bisimplicial set SD(Nerve(𝒞))S^{D}_{\bullet}(\text{Nerve}_{\text{\tiny{$\bullet$}}}({\mathcal{C}})), we obtain iii).

To demonstrate the assertion ii), we recall the proof of the result in [4, Equivariant homology] which describes an equivalence between the Borel cohomology and the cohomology of a groupoid.

Let 𝒢{\mathcal{G}} be the transformation groupoid [G×X\textstyle{[G\times X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}X].\textstyle{X].} The principal GG-bundle π:EG×XXG:=EG×GX\pi:EG\times X\to X_{G}:=EG\times_{G}X defined by the quotient map gives rise to the banal groupoid 𝒢~:=[(EG×X)×XG(EG×X)π1π1(EG×X)],\widetilde{\mathcal{G}}:=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 58.07648pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-58.07648pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[(EG\times X)\times_{X_{G}}(EG\times X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 59.58139pt\raise-7.30415pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.8625pt\hbox{$\scriptstyle{\pi_{1}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 73.07648pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 59.58139pt\raise 7.30415pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.8625pt\hbox{$\scriptstyle{\pi_{1}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 73.07648pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 73.07648pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{(EG\times X)]}$}}}}}}}\ignorespaces}}}}\ignorespaces\!, where πi\pi_{i} is the projection in the iith factor. Then, the groupoid 𝒢~\widetilde{\mathcal{G}} is isomorphic to a transformation groupoid of the form [G×(EG×X)\textstyle{[G\times(EG\times X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(EG×X)].\textstyle{(EG\times X)].} In fact, we have a bundle isomorphism

G×(EG×X)\textstyle{G\times(EG\times X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ\scriptstyle{\mu}\scriptstyle{\cong}pr\scriptstyle{pr}(EG×X)×XG(EG×X)\textstyle{(EG\times X)\times_{X_{G}}(EG\times X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π2\scriptstyle{\pi_{2}}EG×G\textstyle{EG\times G}

which is defined by μ(g,(e,x))=((eg1,gx),(e,x))\mu(g,(e,x))=((eg^{-1},gx),(e,x)). Here prpr denotes the projection in the second factor.

Since the quotient map π\pi is a topological submersion, it follows from [4, Lemma 32] that the edge homomorphism HG(X)=H(XG)H(TotC(Nerve𝒢~))H^{*}_{G}(X)=H^{*}(X_{G})\stackrel{{\scriptstyle\cong}}{{\to}}H(\text{Tot}\,C^{*}(\text{Nerve}_{\text{\tiny{$\bullet$}}}\widetilde{\mathcal{G}})) is an isomorphism of algebras. The projection EG×XXEG\times X\to X in the second factor induces a morphism h:𝒢~𝒢h:\widetilde{\mathcal{G}}\to{\mathcal{G}} of groupoids. Since EGEG is contractible, it follows from the spectral sequence argument that hh induces an isomorphism h:H(TotC(Nerve𝒢))H(TotC(Nerve𝒢~))h^{*}:H(\text{Tot}\,C^{*}(\text{Nerve}_{\text{\tiny{$\bullet$}}}{\mathcal{G}}))\stackrel{{\scriptstyle\cong}}{{\to}}H(\text{Tot}\,C^{*}(\text{Nerve}_{\text{\tiny{$\bullet$}}}\widetilde{\mathcal{G}})). As a consequence, we have the spectral sequence converging to HG(X)H^{*}_{G}(X) in ii). The local finiteness of H(G)H^{*}(G) and H(X)H^{*}(X) allows us to conclude that the chain complex {H(Nerven𝒢)}n\{H^{*}(\text{Nerve}_{n}\mathcal{G})\}_{n} with the horizontal differential is nothing but the cobar complex computing the cotorsion functor. We have the result. ∎

As an input datum, our spectral sequence admits the nerve of a category internal to 𝖣𝗂𝖿𝖿\mathsf{Diff}. Thus, it is expected that the machinery widely contributes to the calculation of cohomology algebras for not only topological categories but also diffeological ones; see Remark 3.2 below.

We recall the D-topology functor D:𝖣𝗂𝖿𝖿𝖳𝗈𝗉D:\mathsf{Diff}\to\mathsf{Top} from the category 𝖣𝗂𝖿𝖿\mathsf{Diff} of diffeological spaces to that of topological spaces that admits the right adjoint CC; see Appendix B. Let 𝒱D{\mathcal{V}}_{D} be the class consisting of diffeological spaces MM for each of which the identity map id:MCDMid:M\to CDM is a weak equivalence in 𝖣𝗂𝖿𝖿\mathsf{Diff}. Especially, the result [24, Theorem 11.2] implies that a CC^{\infty}-manifold in the sense of [26, Section 27] and hence each component of the nerve of a Lie groupoid 𝒢\mathcal{G} is in 𝒱D{\mathcal{V}}_{D}. In particular, paracompact manifolds modeled on Hilbert spaces and the space of smooth maps between finite-dimensional manifolds are also in 𝒱D{\mathcal{V}}_{D}; see [24, Chapter 11.4] for more details.

Remark 3.2.

Given a diffeological space XX, let SnD(X)S_{n}^{D}(X) be the set of smooth maps to XX from the standard nn-simplex Δstn\Delta^{n}_{st} endowed with the diffeology in the sense of Kihara [23, 1.2]. Then, the cosimplicial set structure on Δstn\Delta^{n}_{st} gives a simplicial set SD(X):={SnD(X)}S_{\bullet}^{D}(X):=\{S_{n}^{D}(X)\}.

It follows from [23, Proposition 3.2] that D(Δstn)D(\Delta^{n}_{st}) is the standard simplex which is a subspace of n+1{\mathbb{R}}^{n+1}. Thus, for a category 𝒞{\mathcal{C}} internal to 𝖣𝗂𝖿𝖿\mathsf{Diff}, the smoothing theorem [24, Theorem 1.7] implies that the natural map

η:SD(Nerve(𝒞))Sing(D(Nerve(𝒞)))\eta:S^{D}_{\bullet}(\text{Nerve}_{\text{\tiny{$\bullet$}}}({\mathcal{C}}))\to\text{Sing}_{\text{\tiny{$\bullet$}}}(D(\text{Nerve}_{\text{\tiny{$\bullet$}}}({\mathcal{C}})))

induced by the functor DD is a weak homotopy equivalence of simplicial sets provided each A:=Nerven(𝒞)A:=\text{Nerve}_{n}({\mathcal{C}}) is in the class 𝒱D{\mathcal{V}}_{D}; see Theorem 6.2 for a particular version of the smoothing theorem. Therefore, by [18, Chapter IV, Proposition 1.7], we see that diagSD(Nerve(𝒞))wdiag Sing(D(Nerve(𝒞)))\text{diag}\,S^{D}_{\text{\tiny{$\bullet$}}}(\text{Nerve}_{\text{\tiny{$\bullet$}}}({\mathcal{C}}))\simeq_{w}\text{diag}\!\text{\ Sing}_{\text{\tiny{$\bullet$}}}(D(\text{Nerve}_{\text{\tiny{$\bullet$}}}({\mathcal{C}}))). It turns out that the spectral sequence in Theorem 2.2 iii) converges to the cohomology algebra H(D(Nerve(𝒞));)H^{*}(||D(\text{Nerve}_{\bullet}({\mathcal{C}}))||;{\mathbb{R}}).

We apply the argument above to a transformation diffeological groupoid 𝒢:=[G×NN]{\mathcal{G}}:=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 18.99368pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-18.99368pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[G\times N\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 33.99368pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 33.99368pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 33.99368pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{N]}$}}}}}}}\ignorespaces}}}}\ignorespaces for which GG is a Lie group and NN is in 𝒱D{\mathcal{V}}_{D}. It follows from [12, Lemma 4.1] that the natural map D(X×Y)D(X)×D(Y)D(X\times Y)\to D(X)\times D(Y) is a homeomorphism if D(X)D(X) is locally compact Hausdorff. This implies that D(G×n×N)G×n×D(N)D(G^{\times n}\times N)\cong G^{\times n}\times D(N). Moreover, the functor CC is the right adjoint to DD and hence CC preserves the products. We conclude that each component Nerven(𝒢)=G×n×N\text{Nerve}_{n}(\mathcal{G})=G^{\times n}\times N of the nerve of 𝒢\mathcal{G} is in 𝒱D{\mathcal{V}}_{D}. As a consequence, the spectral sequence in Theorem 2.2 iii) converges to H(D(Nerve(𝒢));)H(Nerve(D𝒢);)H(EG×GDN;)H^{*}(||D(\text{Nerve}_{\bullet}(\mathcal{G}))||;{\mathbb{R}})\cong H^{*}(||\text{Nerve}_{\bullet}(D\mathcal{G})||;{\mathbb{R}})\cong H^{*}(EG\times_{G}DN;{\mathbb{R}}) as algebras, where D𝒢D\mathcal{G} denotes a topological groupoid of the form [G×DN\textstyle{[G\times DN\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}DN]\textstyle{DN]}; see the proof of Theorem 2.2 for the second isomorphism. Thus, we may describe the generators of H(EG×GD(N);)H^{*}(EG\times_{G}D(N);{\mathbb{R}}) with differential forms on NN, although we do not pursue such a topic in this article; see Section 6 for a related topic.

Remark 3.3.

Let 𝒢\mathcal{G} be a Lie groupoid presenting a differentiable stack 𝔛{\mathfrak{X}}. By definition, the de Rham cohomology HDR(𝔛)H_{DR}^{*}({\mathfrak{X}}) of the stack is the cohomology of the total complex of the bigraded de Rham complex A:=(Nerve(𝒢))A:=\wedge^{*}(\text{Nerve}_{\text{\tiny{$\bullet$}}}(\mathcal{G})); see [4, Definition 9]. By Proposition B.4, the factor map α:(Nerve(𝒢))ADR(SD(Nerve(𝒢))\alpha:\wedge^{*}(\text{Nerve}_{\text{\tiny{$\bullet$}}}(\mathcal{G}))\to A_{DR}^{*}(S^{D}_{\text{\tiny{$\bullet$}}}(\text{Nerve}_{\text{\tiny{$\bullet$}}}(\mathcal{G})) is a natural quasi-isomorphism.

Moreover, the result [29, Proposition 3.4] asserts that there exists a sequence of natural quasi-isomorphisms of cochain algebras between C:=C(SD(Nerve(𝒢);𝕂)C:=C^{*}(S^{D}_{\text{\tiny{$\bullet$}}}(\text{Nerve}_{\text{\tiny{$\bullet$}}}(\mathcal{G});{\mathbb{K}}) and ADR(SD(Nerve(𝒢))A_{DR}^{*}(S^{D}_{\text{\tiny{$\bullet$}}}(\text{Nerve}_{\text{\tiny{$\bullet$}}}(\mathcal{G})). Thus, we have a sequence of quasi-isomorphisms between the total complexes (TotC,T)(\text{Tot}\,C,\cup_{T}) and (TotA,T)(\text{Tot}\,A,\wedge_{T}) which preserve products; see (2.1) and (2.2) for the formulae of T\cup_{T} and T\wedge_{T}, respectively. Therefore, because Nerve(𝒢)\text{Nerve}_{\text{\tiny{$\bullet$}}}(\mathcal{G}) is a manifold, the proof of the latter half of Theorem 2.1 and the same argument as in Remark 3.2 enable us to conclude that HDR(𝔛)H(B𝒢,)H_{DR}^{*}({\mathfrak{X}})\cong H^{*}(B{\mathcal{G}},{\mathbb{R}}) as an algebra. We observe that the product T\wedge_{T} coincides with that in the total complex mentioned in [4, (6)].

4. Computational examples

The aim of this section is to compute the cohomology algebra of the free loop space of the real projective space. Let GG be a discrete group acting on a topological space MM. For gGg\in G, we define 𝒫g(M):={γ:[0,1]Mγ(1)=gγ(0)}{\mathcal{P}}_{g}(M):=\{\gamma:[0,1]\to M\mid\gamma(1)=g\gamma(0)\} which is a subspace of the space of continuous paths M[0,1]M^{[0,1]} from the interval [0,1][0,1] to MM with the compact-open topology. Moreover, we put

(4.1) 𝒫G(M):=gG(𝒫g(M)×{g}).\displaystyle{\mathcal{P}}_{G}(M):=\coprod_{g\in G}({\mathcal{P}}_{g}(M)\times\{g\}).

Then, the space 𝒫G(M){\mathcal{P}}_{G}(M) admits a GG-action defined by h(γ,g)=(γh,hgh1)h\cdot(\gamma,g)=({}_{h}\gamma,hgh^{-1}), where γh(t)=hγ(t){}_{h}\gamma(t)=h\cdot\gamma(t). For a space XX, let LXLX denote the free loop space of XX, namely, the space of continuous maps from the circle S1S^{1} to XX. Let GMpM/GG\to M\stackrel{{\scriptstyle p}}{{\to}}M/G be a principal GG-bundle. Proposition A.1 enables us to obtain a weak homotopy equivalence

(4.2) p¯:EG×G𝒫G(M)wL(M/G)\overline{p}:EG\times_{G}{\mathcal{P}}_{G}(M)\stackrel{{\scriptstyle\simeq_{w}}}{{\longrightarrow}}L(M/G)

which is induced by the projection p:MM/Gp:M\to M/G. Thus, we see that the spectral sequence in Theorem 2.2 ii) computes the cohomology algebra of a transformation groupoid of the form 𝒢=[G×𝒫G(M)𝒫G(M)]\mathcal{G}=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 28.20912pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-28.20912pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[G\times{\mathcal{P}}_{G}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 43.20912pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 43.20912pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 43.20912pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{{\mathcal{P}}_{G}(M)]}$}}}}}}}\ignorespaces}}}}\ignorespaces and hence that of L(M/G)L(M/G).

Suppose that MM is simply connected. For each gGg\in G, to compute H(𝒫g(M);𝕂)H^{*}({\mathcal{P}}_{g}(M);{\mathbb{K}}) with coefficients in a field 𝕂{\mathbb{K}}, we may use the Eilenberg–Moore spectral sequence (henceforth EMSS for short) for the pullback diagram

(4.7)

where εi\varepsilon_{i} is the evaluation map at ii for i=0,1i=0,1 and gg denotes the map induced by the action on MM with the element gg. We observe that the EMSS {Er,,dr}\{E_{r}^{*,*},d_{r}\} converges to H(𝒫g(M);𝕂)H^{*}({\mathcal{P}}_{g}(M);{\mathbb{K}}) as an algebra with

E2,TorH(M;𝕂)H(M;𝕂),(H(M;𝕂)g,H(M;𝕂))E_{2}^{*,*}\cong\text{Tor}_{H^{*}(M;{\mathbb{K}})\otimes H^{*}(M;{\mathbb{K}})}^{*,*}(H^{*}(M;{\mathbb{K}})_{g},H^{*}(M;{\mathbb{K}}))

as a bigraded algebra. Here H(M;𝕂)gH^{*}(M;{\mathbb{K}})_{g} denotes the cohomology algebra H(M;𝕂)H^{*}(M;{\mathbb{K}}) endowed with the right H(M;𝕂)H(M;𝕂)H^{*}(M;{\mathbb{K}})\otimes H^{*}(M;{\mathbb{K}})-action defined by a(λλ)=a(λg(λ))a\cdot(\lambda\otimes\lambda^{\prime})=a(\lambda g^{*}(\lambda^{\prime})) for aH(M;𝕂)ga\in H^{*}(M;{\mathbb{K}})_{g} and λ,λH(M;𝕂)\lambda,\lambda^{\prime}\in H^{*}(M;{\mathbb{K}}).

For an element hh, the GG-action on 𝒫G(M){\mathcal{P}}_{G}(M) induces the map h:𝒫g(M)𝒫hgh1(M)h_{*}:{\mathcal{P}}_{g}(M)\to{\mathcal{P}}_{hgh^{-1}}(M) which fits in the commutative diagram

(4.16)

Then, the naturality of the EMSS gives rise to a morphism of spectral sequences that is compatible with the map (h):H(𝒫hgh1(M);𝕂)H(𝒫g(M);𝕂)(h_{*})^{*}:H^{*}({\mathcal{P}}_{hgh^{-1}}(M);{\mathbb{K}})\to H^{*}({\mathcal{P}}_{g}(M);{\mathbb{K}}).

In the rest of this section, by utilizing the spectral sequence in Theorem 2.2 ii), we determine the cohomology algebra of the free loop space LPnL{\mathbb{R}}P^{n} of the real projective space with coefficients in /p{\mathbb{Z}}/p for p2p\neq 2. Moreover, we investigate the mod 22 cohomology algebra of LPnL{\mathbb{R}}P^{n} in a more general context; see Proposition 4.4.

A portion of the computations, in general, follows from a description of the cotorsion functor with the cohomology of a group; see Lemma 5.1, Theorem 5.2 and Corollary 5.4 below. However, we here compute the functor with the cobar complex to see what is happening in the E1E_{1}-term of the spectral sequence that we apply in the computation.

Theorem 4.1.

Let pp be an odd prime or 0 and mm a positive integer, then as algebras,

H(LP2m+1;/p)\displaystyle H^{*}(L{\mathbb{R}}P^{2m+1};{\mathbb{Z}}/p) \displaystyle\cong H(LS2m+1;/p)H(LS2m+1;/p)\displaystyle H^{*}(LS^{2m+1};{\mathbb{Z}}/p)\oplus H^{*}(LS^{2m+1};{\mathbb{Z}}/p)
\displaystyle\cong ((y)Γ[y¯])2and\displaystyle(\wedge(y)\otimes\Gamma[\overline{y}])^{\oplus 2}\ \ \ \text{and}
H(LP2m;/p)\displaystyle H^{*}(L{\mathbb{R}}P^{2m};{\mathbb{Z}}/p) \displaystyle\cong ((xu)Γ[w])/p,\displaystyle(\wedge(x\otimes u)\otimes\Gamma[w])\oplus\mathbb{Z}/p,

where degy=2m+1\deg y=2m+1, degy¯=2m\deg\overline{y}=2m, deg(xu)=4m1\deg(x\otimes u)=4m-1, degw=4m2\deg w=4m-2 and /0:={\mathbb{Z}}/0:={\mathbb{Q}}.

Before starting the proof of Theorem 4.1, we recall a result on a right GG action on a vector space and the right GG-coaction associated with the action. Let GG be a finite group and η:𝕂[G]VV\eta:{\mathbb{K}}[G]\otimes V\to V a GG-action on a finite-dimensional vector space VV. The left GG-action gives rise to the right GG-action φ:V𝕂[G]V\varphi:V^{\vee}\otimes{\mathbb{K}}[G]\to V^{\vee} defined by φ(fg)(v)=f(η(gv))\varphi(f\otimes g)(v)=f(\eta(g\otimes v)) for vVv\in V. Moreover, we see that the adjoint ad(φ):V𝕂[G]Vad(\varphi):V^{\vee}\to{\mathbb{K}}[G]^{\vee}\otimes V^{\vee} coincides with the dual coaction η:V𝕂[G]V\eta^{\vee}:V^{\vee}\to{\mathbb{K}}[G]^{\vee}\otimes V^{\vee}. We use the fact in the computation below without mentioning that.

In what follows, we may omit the coefficients in the cohomology groups that we deal with.

Proof of Theorem 4.1.

The antipodal action of G:=/2G:={\mathbb{Z}}/2 on the sphere SnS^{n} gives rise to the real projective space Pn=Sn/G{\mathbb{R}}P^{n}=S^{n}/G. By applying Theorem 2.2 ii) to the groupoid [G×𝒫G(Sn)\textstyle{[G\times{\mathcal{P}}_{G}(S^{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒫G(Sn)]\textstyle{{\mathcal{P}}_{G}(S^{n})]}​, we have a spectral sequence {Er,,dr}\{E_{r}^{*,*},d_{r}\} converging to H(LPn;/p)H^{*}(L{\mathbb{R}}P^{n};{\mathbb{Z}}/p) with E2,CotorH(G),(/p,H(𝒫G(Sn)))E_{2}^{*,*}\cong\text{Cotor}^{*,*}_{H^{*}(G)}({\mathbb{Z}}/p,H^{*}({\mathcal{P}}_{G}(S^{n}))) as an algebra. Since GG is abelian, it follows that the GG action on 𝒫G(Sn){\mathcal{P}}_{G}(S^{n}) is restricted to each 𝒫g(Sn){\mathcal{P}}_{g}(S^{n}) for gGg\in G. Then, we see that L(Pn)wEG×G𝒫G(Sn)=gG(EG×G𝒫g(Sn))L({\mathbb{R}}P^{n})\simeq_{w}EG\times_{G}{\mathcal{P}}_{G}(S^{n})=\coprod_{g\in G}\big{(}EG\times_{G}{\mathcal{P}}_{g}(S^{n})\big{)} and

CotorH(G),(/p,H(𝒫G(Sn)))=gGCotorH(G),(/p,H(𝒫g(Sn))).\text{Cotor}^{*,*}_{H^{*}(G)}({\mathbb{Z}}/p,H^{*}({\mathcal{P}}_{G}(S^{n})))=\oplus_{g\in G}\text{Cotor}^{*,*}_{H^{*}(G)}({\mathbb{Z}}/p,H^{*}({\mathcal{P}}_{g}(S^{n}))).

We compute the cotorsion functor with the normalized cobar complex

(/p{τ}kH(𝒫g(Sn)),k=G1+(1)k+11τ)k0,\big{(}{\mathbb{Z}}/p\{\tau^{*}\}^{\otimes k}\otimes H^{*}({\mathcal{P}}_{g}(S^{n})),\partial_{k}=\nabla_{G}\otimes 1+(-1)^{k+1}1\otimes\nabla_{\tau^{*}}\big{)}_{k\geq 0},

where τG\tau\in G denotes the nontrivial element, τ:H(𝒫g(Sn))H~0(G)H(𝒫g(Sn))=/p{τ}H(𝒫g(Sn))\nabla_{\tau^{*}}:H^{*}({\mathcal{P}}_{g}(S^{n}))\to\widetilde{H}^{0}(G)\otimes H^{*}({\mathcal{P}}_{g}(S^{n}))={\mathbb{Z}}/p\{\tau^{*}\}\otimes H^{*}({\mathcal{P}}_{g}(S^{n})) is the coaction induced by the GG-action on 𝒫g(Sn){\mathcal{P}}_{g}(S^{n}) and the projection G×𝒫g(Sn)𝒫g(Sn)G\times{\mathcal{P}}_{g}(S^{n})\to{\mathcal{P}}_{g}(S^{n}) gives rise to the map G\nabla_{G}. Observe that the complex is nothing but the E1E_{1}-term of the spectral sequence {Er,,dr}\{E_{r}^{*,*},d_{r}\}.

In what follows, we may write 𝕂{\mathbb{K}} for the underlying field /p{\mathbb{Z}}/p. We consider the EMSS associated with the fibre square (4.7) for M=SnM=S^{n}.

For the case n=2m+1n=2m+1, the action on H(S2m+1)(y)H^{*}(S^{2m+1})\cong\wedge(y) induced by the nontrivial element τ\tau is given by τ(y)=y\tau^{*}(y)=y. Therefore, the computation in [31, Theorem 2.1] allows us to conclude that the cohomology algebra H(𝒫g(Sn))H^{*}({\mathcal{P}}_{g}(S^{n})) is isomorphic to H(LS2m+1)H^{*}(LS^{2m+1}) for each gGg\in G. The naturality of the EMSS implies that the action by gg^{*} on H(LS2m+1)H^{*}(LS^{2m+1}) is trivial and then so is the coaction. This yields that E20,CotorH(G)0,(/p,H(𝒫G(Sn)))H(LS2m+1)2E_{2}^{0,*}\cong\text{Cotor}^{0,*}_{H^{*}(G)}({\mathbb{Z}}/p,H^{*}({\mathcal{P}}_{G}(S^{n})))\cong H^{*}(LS^{2m+1})^{\oplus 2} and E2p,=0E_{2}^{p,*}=0 for p>0p>0. Then, it follows that E2,E,E_{2}^{*,*}\cong E_{\infty}^{*,*} and there is no extension problem. The explicit form of H(LS2m+1)H^{*}(LS^{2m+1}) follows from [31, Theorem 2.1]. This enables us to obtain the first assertion.

We consider the case where n=2mn=2m. In order to compute the torsion functor which gives the E2E_{2}-term of the EMSS, we recall a Koszul–Tate resolution of the form

=(ΛΛ(u)Γ[w],d)εΛ0{\mathcal{F}}=(\Lambda\otimes\Lambda\otimes\wedge(u)\otimes\Gamma[w],d)\stackrel{{\scriptstyle\varepsilon}}{{\to}}\Lambda\to 0

of Λ:=H(S2m)=𝕂[x]/(x2)\Lambda:=H^{*}(S^{2m})={\mathbb{K}}[x]/(x^{2}) as a left ΛΛ\Lambda\otimes\Lambda-module, where ε\varepsilon is the multiplication on Λ\Lambda, d(ΛΛ)=0d(\Lambda\otimes\Lambda)=0, d(u)=x11xd(u)=x\otimes 1-1\otimes x, d(γr(w))=(x1+1x)uγr1(w)d(\gamma_{r}(w))=(x\otimes 1+1\otimes x)u\otimes\gamma_{r-1}(w), bidegu=(1,degx)\text{bideg}\ u=(-1,\deg x) and bidegγr(w)=r(2,2degx)\text{bideg}\ \gamma_{r}(w)=r(-2,2\deg x); see [41, Proposition 3.5] and [27, Proposition 1.1]. The complex (ΛΛΛ,1d)(\Lambda\otimes_{\Lambda\otimes\Lambda}{\mathcal{F}},1\otimes d) computes the Hochschild homology of Λ\Lambda. Using the resolution, we determine H(𝒫0(Sn))H^{*}({\mathcal{P}}_{0}(S^{n})) and H(𝒫τ(Sn))H^{*}({\mathcal{P}}_{\tau}(S^{n})) for the nontrivial element τG\tau\in G

Claim 4.2.

i) H(𝒫0(S2m)){𝕂[x]/(x2)(u)/(xu)A}{(x,u)A/(xu)A}Γ+[w]H^{*}({\mathcal{P}}_{0}(S^{2m}))\cong\{{\mathbb{K}}[x]/(x^{2})\otimes\wedge(u)/(xu)_{A}\}\oplus\{(x,u)_{A}/(xu)_{A}\}\otimes\Gamma^{+}[w] as an algebra, where (S)A(S)_{A} denotes the ideal of A:=𝕂[x]/(x2)(u)A:={\mathbb{K}}[x]/(x^{2})\otimes\wedge(u) generated by a set SS. Moreover, τ(z)=z\tau^{*}(z)=-z for zH~(𝒫0(Sn))z\in\widetilde{H}^{*}({\mathcal{P}}_{0}(S^{n})).
ii) H(𝒫τ(S2m))(xu)Γ[w]H^{*}({\mathcal{P}}_{\tau}(S^{2m}))\cong\wedge(x\otimes u)\otimes\Gamma[w] as an algebra and τ(z)=z\tau^{*}(z)=z for each element z(xu)Γ[w]z\in\wedge(x\otimes u)\otimes\Gamma[w].

It follows from Claim 4.2 i) that CotorH(G),(/p,H(𝒫0(S2m)))=/p\text{Cotor}^{*,*}_{H^{*}(G)}({\mathbb{Z}}/p,H^{*}({\mathcal{P}}_{0}(S^{2m})))={\mathbb{Z}}/p. Moreover, Claim 4.2 ii) enables us to conclude that

CotorH(G)i,(/p,H(𝒫τ(S2m))){(xu)Γ[w]for i=00for i0\displaystyle\text{Cotor}^{i,*}_{H^{*}(G)}({\mathbb{Z}}/p,H^{*}({\mathcal{P}}_{\tau}(S^{2m})))\cong\left\{\begin{array}[]{ll}\wedge(x\otimes u)\otimes\Gamma[w]&\text{for $i=0$}\\ 0&\text{for $i\neq 0$}\end{array}\right.

We have the result. ∎

Proof of Claim 4.2.

Let nn be an even integer 2m2m. i) Since 𝒫0(Sn){\mathcal{P}}_{0}(S^{n}) is nothing but the free loop space LSnLS^{n}, the result on the algebra structure follows from the results in [37, 4.1] and [31, Theorem 2.2]. In order to prove the latter assertion on the action, we represent elements in the Koszul–Tate resolution with the bar resolution 𝔹,{\mathbb{B}}^{*,*} of Λ\Lambda as a left ΛΛ\Lambda\otimes\Lambda-module. Let ψ:𝔹,\psi:{\mathbb{B}}^{*,*}\to{\mathcal{F}} be the chain map constructed in the proof of [27, Lemma 1.5], which induces an isomorphism between the torsion functors. We see that ψ(x)=x\psi(x)=x, ψ(u)=[x11x]\psi(u)=[x\otimes 1-1\otimes x] and

Ψ(1ΛΛ[x1+1xx11xx1+1xx11x]1Λ)=γr(w).\Psi(1_{\Lambda\otimes\Lambda}[x\otimes 1+1\otimes x\mid x\otimes 1-1\otimes x\mid\cdots\mid x\otimes 1+1\otimes x\mid x\otimes 1-1\otimes x]1_{\Lambda})=\gamma_{r}(w).

Since τ(x)=x\tau^{*}(x)=-x for xH(Sn)x\in H^{*}(S^{n}), it follows that τ(u)=u\tau^{*}(u)=-u and τ(γr(w))=γr(w)\tau^{*}(\gamma_{r}(w))=\gamma_{r}(w). Consider the morphism of spectral sequences induced by the diagram (4.16). Then, the naturality of the EMSS and the forms of algebra generators of H(𝒫0(Sn))H^{*}({\mathcal{P}}_{0}(S^{n})) enable us to obtain the latter half of i).

ii) To compute the cohomology algebra H(𝒫τ(Sn))H^{*}({\mathcal{P}}_{\tau}(S^{n})) for the nontrivial element τG\tau\in G, we consider the EMSS {E~r,,d~r}\{\widetilde{E}_{r}^{*,*},\widetilde{d}_{r}\} associated with the pullback diagram (4.7) converging to the cohomology H(𝒫τ(Sn))H^{*}({\mathcal{P}}_{\tau}(S^{n})). As mentioned in the proof of i), the τ\tau-action on Hn(Sn)H^{n}(S^{n}) is nothing but the multiplication by 1-1. Thus, the right ΛΛ\Lambda\otimes\Lambda-module structure on Λτ:=H(Sn)\Lambda_{\tau}:=H^{*}(S^{n}) is given by a(λλ)=aλλa\cdot(\lambda\otimes\lambda^{\prime})=-a\lambda\lambda^{\prime} for aΛτa\in\Lambda_{\tau} and λλΛΛ\lambda\otimes\lambda\in\Lambda\otimes\Lambda. This yields that

E~2,TorΛΛ(Λτ,Λ)(ΛτΛΛ,d(u)=2x,d(γr(w))=0)(xu)Γ[w]\widetilde{E}_{2}^{*,*}\cong\text{Tor}_{\Lambda\otimes\Lambda}(\Lambda_{\tau},\Lambda)\cong(\Lambda_{\tau}\otimes_{\Lambda\otimes\Lambda}{\mathcal{F}},d^{\prime}(u)=2x,d^{\prime}(\gamma_{r}(w))=0)\cong\wedge(x\otimes u)\otimes\Gamma[w]

as bigraded algebras, where bideg(xu)=(1,2n)\text{bideg}(x\otimes u)=(-1,2n). For degree reasons, we see that the EMSS collapses at the E2E_{2}-term. In fact, the possibility of a nontrivial differential appears as ds(γpf(w))=αxuγl(w)d_{s}(\gamma_{p^{f}}(w))=\alpha x\otimes u\cdot\gamma_{l}(w) for some positive integers ss, ff, ll and some α𝕂\alpha\in{\mathbb{K}}. We observe that s2s\geq 2. Comparing the total degrees of both elements in the equality, we have (2n1)pf+1=(2n1)+(2n2)l(2n-1)p^{f}+1=(2n-1)+(2n-2)l and then pf=l+1p^{f}=l+1. By comparing the filtration degree, we see that 2pf+s=12(pf1)-2p^{f}+s=-1-2(p^{f}-1) and s=1s=1, which is a contradiction.

We have to show that γpf(w)p=0\gamma_{p}^{f}(w)^{p}=0 in H(𝒫τ(Sn))H^{*}({\mathcal{P}}_{\tau}(S^{n})) for f0f\geq 0. The extension problems are solved by degree reasons. It turns out that H(𝒫τ(Sn))TotE~,TotE~,H^{*}({\mathcal{P}}_{\tau}(S^{n}))\cong\text{Tot}\widetilde{E}_{\infty}^{*,*}\cong\text{Tot}\widetilde{E}_{\infty}^{*,*} as algebras.

By using the chain map Ψ:𝔹,\Psi:{\mathbb{B}}^{*,*}\to{\mathcal{F}} as mentioned above, we see that τ(xu)=(1)(1)xu\tau^{*}(x\otimes u)=(-1)(-1)x\otimes u and τ(γr(w))=γr(w)\tau^{*}(\gamma_{r}(w))=\gamma_{r}(w). Thus, by considering again the morphism of spectral sequences induced by the diagram (4.16), we have the result on the action. ∎

Remark 4.3.

The path space 𝒫0(Sn){\mathcal{P}}_{0}(S^{n}) is nothing but the free loop space LSnLS^{n}. Then, the inclusion LSn=𝒫0(Sn)EG×𝒫0(Sn)LS^{n}={\mathcal{P}}_{0}(S^{n})\to EG\times{\mathcal{P}}_{0}(S^{n}) defines a natural map α:LSnEG×G𝒫0(Sn)\alpha:LS^{n}\to EG\times_{G}{\mathcal{P}}_{0}(S^{n}). Moreover, we have a commutative diagram

LSn\textstyle{LS^{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Lp\scriptstyle{Lp}α\scriptstyle{\alpha}LPn\textstyle{L{\mathbb{R}}P^{n}}gG(EG×G𝒫g(Sn))\textstyle{\coprod_{g\in G}(EG\times_{G}{\mathcal{P}}_{g}(S^{n}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p¯\scriptstyle{\overline{p}}w\scriptstyle{\simeq_{w}}EG×G𝒫0(Sn)\textstyle{EG\times_{G}{\mathcal{P}}_{0}(S^{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Sn\textstyle{S^{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}s\scriptstyle{s}p\scriptstyle{p}Pn,\textstyle{{\mathbb{R}}P^{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces,}s\scriptstyle{s}

where pp is the universal cover and ss denotes the map that assigns the constant loop at rr to each element rr in SnS^{n} and Pn{\mathbb{R}}P^{n}, respectively. Thus, it follows that LSnLS^{n} is weak homotopy equivalent to the path component of LPnL{\mathbb{R}}P^{n} consisting of constant loops under the equivalence (4.2). In particular, the proof of Theorem 4.1 allows us to conclude that the cohomology of the component with coefficients in /p{\mathbb{Z}}/p is isomorphic to /p{\mathbb{Z}}/p if nn is even.

We here attempt to compute the cohomology of the free loop space of a Borel construction with coefficients in /2\mathbb{Z}/2 by using the same method as above. We indeed use the multiplicative structure of the spectral sequence that we apply.

Let GG be the cyclic group /2\mathbb{Z}/2 and MM a simply-connected, mod 22 homology nn-sphere that admits a GG-action, where n2n\geq 2. Further assume that the Borel cohomology H(EG×GM;/2)H^{*}(EG\times_{G}M;{\mathbb{Z}}/2) is of finite dimension. Let L0(EG×GM)L_{0}(EG\times_{G}M) be the connected component of the free loop space L(EG×GM)L(EG\times_{G}M) containing the constant loops. Then, we have the following result.

Proposition 4.4.

As an algebra, H(L0(EG×GM);/2)Γ[y](/2[t]/(tn+1))H^{*}(L_{0}(EG\times_{G}M);{\mathbb{Z}}/2)\cong\Gamma[y]\otimes\big{(}{\mathbb{Z}}/2[t]/(t^{n+1})\big{)} for 2n3*\leq 2n-3, where degy=n1\deg y=n-1 and degt=1\deg t=1.

Proof.

To obtain the isomorphism, we compare spectral sequences given by the groupoid 𝒢1:=[G×𝒫0(EG×M)𝒫0(EG×M)]\mathcal{G}_{1}:=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 41.42911pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-41.42911pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[G\times{\mathcal{P}}_{0}(EG\times M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 56.42911pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 56.42911pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 56.42911pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{{\mathcal{P}}_{0}(EG\times M)]}$}}}}}}}\ignorespaces}}}}\ignorespaces and the translation groupoid 𝒢2:=[G×(EG×M)EG×M]\mathcal{G}_{2}:=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 37.7374pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-37.7374pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[G\times(EG\times M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 52.7374pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 52.7374pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 52.7374pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{EG\times M]}$}}}}}}}\ignorespaces}}}}\ignorespaces​, respectively. Let ev0~:𝒢1𝒢2\widetilde{ev_{0}}:\mathcal{G}_{1}\to\mathcal{G}_{2} be the morphism of groupoids induced by the evaluation map ev0:𝒫0(EG×M)EG×Mev_{0}:{\mathcal{P}}_{0}(EG\times M)\to EG\times M at 0I0\in I, where ev0ev_{0} is the evaluation map at 0. Observe that ev0ev_{0} is a GG-equivariant map. Thus, we have a commutative diagram

EG×G𝒫0(EG×M)\textstyle{EG\times_{G}{\mathcal{P}}_{0}(EG\times M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1×ev0\scriptstyle{1\times ev_{0}}p¯\scriptstyle{\overline{p}}w\scriptstyle{\simeq_{w}}L0(EG×GM)\textstyle{L_{0}(EG\times_{G}M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ev0\scriptstyle{ev_{0}}EG×G(EG×M)\textstyle{EG\times_{G}(EG\times M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}w\scriptstyle{\simeq_{w}}q\scriptstyle{q}EG×GM,\textstyle{EG\times_{G}M,}

where p¯\overline{p} is the weak equivalence described in Proposition A.1 and qq is induced by the natural projection. Let {Er,1,dr1}\{{}_{1}E_{r}^{*,*},{}_{1}d_{r}\} and {Er,2,dr2}\{{}_{2}E_{r}^{*,*},{}_{2}d_{r}\} be the spectral sequences in Theorem 2.2 ii) with coefficients in /2{\mathbb{Z}}/2 constructed by groupoids 𝒢1\mathcal{G}_{1} and 𝒢2\mathcal{G}_{2}, respectively. We observe that 𝒫0(EG×M)=L(EG×M)LM{\mathcal{P}}_{0}(EG\times M)=L(EG\times M)\simeq LM.

By the assumption on MM, we see that H(LM)/2[xn]/(xn2)Γ[y]H^{*}(LM)\cong{\mathbb{Z}}/2[x_{n}]/(x_{n}^{2})\otimes\Gamma[y] as an algebra, where degxn=n\deg x_{n}=n and degy=n1\deg y=n-1. In fact, the isomorphism for nn odd follows from [31, Theorem 2.1]. Moreover, the computation in [37, 4.1] gives the result for nn even. Indeed, we need an explicit form of the mod 22 cohomology group of ΩM\Omega M as an input datum; that is, H(ΩM;/2)Γ[y]H^{*}(\Omega M;{\mathbb{Z}}/2)\cong\Gamma[y], where degy=n1\deg y=n-1. This fact follows from the computation of the EMSS converging to H(ΩM;/2)H^{*}(\Omega M;{\mathbb{Z}}/2) with E2,TorH(M),(/2,/2)E_{2}^{,*}\cong\text{Tor}_{H^{*}(M)}^{*,*}({\mathbb{Z}}/2,{\mathbb{Z}}/2).

Since each dimension of Ai:=Hi(𝒫0(EG×M))A^{i}:=H^{i}({\mathcal{P}}_{0}(EG\times M)) is one or zero. Thus, the GG-action on AiA^{i} is trivial. This fact and the formula (2.1) allow us to deduce that

E2,1CotorH(G),(/2,A)A/2[t]/2[xn]/(xn2)Γ[y]/2[t]{}_{1}E_{2}^{*,*}\cong\text{Cotor}_{H^{*}(G)}^{*,*}({\mathbb{Z}}/2,A^{*})\cong A^{*}\otimes{\mathbb{Z}}/2[t]\cong{\mathbb{Z}}/2[x_{n}]/(x_{n}^{2})\otimes\Gamma[y]\otimes{\mathbb{Z}}/2[t]

as algebras, where bidegt=(1,0)\text{bideg}\ t=(1,0). The same argument as above is applicable to the spectral sequence {Er,2,dr2}\{{}_{2}E_{r}^{*,*},{}_{2}d_{r}\}. Then, it follows that

E2,2CotorH(G),(/2,H(M))/2[xn]/(xn2)/2[t]{}_{2}E_{2}^{*,*}\cong\text{Cotor}_{H^{*}(G)}^{*,*}({\mathbb{Z}}/2,H^{*}(M))\cong{\mathbb{Z}}/2[x_{n}]/(x_{n}^{2})\otimes{\mathbb{Z}}/2[t]

as algebras.

We consider the morphism {fr}:{Er,2,dr2}{Er,1,dr2}\{f_{r}\}:\{{}_{2}E_{r}^{*,*},{}_{2}d_{r}\}\to\{{}_{1}E_{r}^{*,*},{}_{2}d_{r}\} of spectral sequences induced by ev0~\widetilde{ev_{0}}. The construction of the spectral sequence yields that f2(xn)=xnf_{2}(x_{n})=x_{n} and f2(t)=tf_{2}(t)=t. By assumption, the vector space H(EG×GM)H^{*}(EG\times_{G}M) is of finite dimensional. Therefore, we see that dn+12(xn)=tn+1{}_{2}d_{n+1}(x_{n})=t^{n+1} and hence dn+11(xn)=tn+1{}_{1}d_{n+1}(x_{n})=t^{n+1}. To complete the proof, it remains to show that dn11(y)=0{}_{1}d_{n-1}(y)=0 but not dn1(y)=tn{}_{1}d_{n}(y)=t^{n}. Since ev0:L0(EG×GM)EG×GMev_{0}:L_{0}(EG\times_{G}M)\to EG\times_{G}M has a section, it follows that the map ev0ev_{0}^{*} induced on the cohomology is a monomorphism. This implies that dn11(y)=0{}_{1}d_{n-1}(y)=0. ∎

Remark 4.5.

In the spectral sequence {Er,1,dr1}\{{}_{1}E_{r}^{*,*},{}_{1}d_{r}\} above, there is a possibility that dn1(γ2(xn¯))=xn¯tn{}_{1}d_{n}(\gamma_{2}(\overline{x_{n}}))=\overline{x_{n}}\otimes t^{n}. In the computation above, we can use the Leray–Serre spectral sequence for a Borel fibration of the form XEG×GXBGX\to EG\times_{G}X\to BG instead of the spectral sequence in Theorem 2.2 ii). However, the same possibility of the non-trivial differential as above remains.

Remark 4.6.

We observe that Proposition 4.4 is applicable to the real projective space. Let MM be the 33-dimensional real projective space P3{\mathbb{R}}P^{3} which is regarded as the homogeneous space of S3=SU(2)S^{3}=SU(2) by the center /2{\mathbb{Z}}/2. Since MM is an H-space, it follows that L(M)Ω[e]M×M(ΩeS3ΩeS3)×ML(M)\simeq\Omega_{[e]}M\times M\simeq(\Omega_{e}S^{3}\coprod\Omega_{e}S^{3})\times M, where eS3e\in S^{3} denotes the unit; see the proof of Proposition A.1 for the second weak homotopy equivalence. Then it turns out that H(LM;/2)(Γ[s1x])2(/2[t]/(t4))H^{*}(LM;{\mathbb{Z}}/2)\cong(\Gamma[s^{-1}x])^{\oplus 2}\otimes\big{(}{\mathbb{Z}}/2[t]/(t^{4})\big{)} as an algebra.

5. A spectral sequence for a Borel construction

In the previous section, we apply the spectral sequence in Theorem 2.2 ii) to a computation of the cohomology of a transformation groupoid. In this section, the result is generalized with a description of the cotorsion functor by the cohomology of a finite group. Moreover, we give a spectral sequence for computing the cohomology of the classifying space of inertia groupoids, while there is no computational example.

We begin with a lemma relating the cotorsion functor to the cohomology of a group. Let GG be a finite group and NN a finite-dimensional left 𝕂[G]{\mathbb{K}}[G]^{\vee}-comodule. Then the module NN^{\vee} is regarded as a left 𝕂[G]{\mathbb{K}}[G]-module via the natural map :𝕂[G]NN\nabla^{\vee}:{\mathbb{K}}[G]\otimes N^{\vee}\to N^{\vee} induced by the left comodule structure \nabla on NN. Thus, by using the isomorphism N(N)N\cong(N^{\vee})^{\vee}, we consider NN a right 𝕂[G]{\mathbb{K}}[G]-module.

For a Hopf algebra AA with antipode SS, we have an injective algebra homomorphism δ:AAe=AAop\delta:A\to A^{e}=A\otimes A^{\text{op}} defined by δ(a)=a1S(a2)\delta(a)=\sum a_{1}\otimes S(a_{2}); see [46, Lemma 9.4.1].

Lemma 5.1.

Under the setup above, there are isomorphisms of vector spaces

Cotor𝕂[G](𝕂,N)HH(𝕂[G],𝕂N)Ext𝕂[G](𝕂,𝕂N)=H(G,N).\text{\em Cotor}^{*}_{{\mathbb{K}}[G]^{\vee}}({\mathbb{K}},N)\cong\text{\em HH}^{*}({\mathbb{K}}[G],{\mathbb{K}}\otimes N)\cong\text{\em Ext}^{*}_{{\mathbb{K}}[G]}({\mathbb{K}},{\mathbb{K}}\otimes N)=H^{*}(G,N).

Here the module NN in the Hochschild cohomology is regarded as a right 𝕂[G]{\mathbb{K}}[G]-module mentioned above. Moreover, the module 𝕂N{\mathbb{K}}\otimes N in the group cohomology is considered a left 𝕂[G]{\mathbb{K}}[G]-module via the monomorphism δ\delta.

Proof.

The first isomorphism follows from [2, Theorem 3.4]. As for the second isomorphism, we repeat the proof of [46, Theorem 9.4.5] verbatim. For a general Hopf algebra AA, the result [46, Lemma 9.4.2] enables us to conclude that the map f:AAeA𝕂f:A\to A^{e}\otimes_{A}{\mathbb{K}} defined by f(a)=a11f(a)=a\otimes 1\otimes 1 is an isomorphism of AeA^{e}-modules, where AeA^{e} denotes the enveloping algebra of AA. Then, it follows that ExtAen(A,T)ExtAen(AeA𝕂,T)ExtAn(𝕂,T)\text{Ext}^{n}_{A^{e}}(A,T)\cong\text{Ext}^{n}_{A^{e}}(A^{e}\otimes_{A}{\mathbb{K}},T)\cong\text{Ext}_{A}^{n}({\mathbb{K}},T) for a left AeA^{e}-module TT. We observe that the latter isomorphism is given by Eckmann-Shapiro Lemma; see, for example, [46, Lemma A.6.2]. This completes the proof. ∎

We can generalize the computation in Section 4.

Theorem 5.2.

(cf. [9, Chapter II, Theorem 19.2]) Let GG be a finite group acting on a space XX whose cohomology with coefficients in 𝕂{\mathbb{K}} is locally finite. Suppose that ch(𝕂)\text{\em ch}({\mathbb{K}}) the characteristic of 𝕂{\mathbb{K}} is coprime with |G||G|. Then as an algebra

H(EG×GX;𝕂)𝕂𝕂[G]H(X;𝕂),H^{*}(EG\times_{G}X;{\mathbb{K}})\cong{\mathbb{K}}\Box_{{\mathbb{K}}[G]^{\vee}}H^{*}(X;{\mathbb{K}}),

where the right-hand side algebra denotes the cotensor product of the trivial right 𝕂[G]{\mathbb{K}}[G]^{\vee}-comodule 𝕂{\mathbb{K}} and the left 𝕂[G]{\mathbb{K}}[G]^{\vee}-comodule H(X;𝕂)H^{*}(X;{\mathbb{K}}).

Proof.

In view of Theorem 2.2 ii), we have a spectral sequence {Er,,dr}\{E_{r}^{*,*},d_{r}\} converging to H(EG×GX)H^{*}(EG\times_{G}X) with E2,CotorH(G),(𝕂,H(X))E_{2}^{*,*}\cong\text{Cotor}_{H^{*}(G)}^{*,*}({\mathbb{K}},H^{*}(X)) as an algebra. Since H(G)=H0(G)=𝕂[G]H^{*}(G)=H^{0}(G)={\mathbb{K}}[G]^{\vee} and the characteristic of 𝕂{\mathbb{K}} is coprime with |G||G|, it follows from Lemma 5.1 that

CotorH(G)p,q(𝕂,H(X))Cotor𝕂[G]p(𝕂,Hq(X))={𝕂𝕂[G]Hq(X)for p=00for p0.\displaystyle\text{Cotor}^{p,q}_{H^{*}(G)}({\mathbb{K}},H^{*}(X))\cong\text{Cotor}_{{\mathbb{K}}[G]^{\vee}}^{p}({\mathbb{K}},H^{q}(X))=\left\{\begin{array}[]{ll}{\mathbb{K}}\Box_{{\mathbb{K}}[G]^{\vee}}H^{q}(X)&\text{for $p=0$}\\ 0&\text{for $p\neq 0$}.\end{array}\right.

We see that the coalgebra structure :H(X)H(G)H(X)\nabla:H^{*}(X)\to H^{*}(G)\otimes H^{*}(X) is a morphism of algebras. Then, the cotensor product 𝕂𝕂[G]H(X){\mathbb{K}}\Box_{{\mathbb{K}}[G]^{\vee}}H^{*}(X) is a subalgebra of H(X)H^{*}(X). It turns out that the spectral sequence of algebras collapses at the E2E_{2}-term and there is no extension problem. We have the result. ∎

Remark 5.3.

Let GG be a finite group acting on a space XX. Then, the inclusion from the unit to GG gives rise to a morphism ι:[{e}×X\textstyle{\iota:[\{e\}\times X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}X]\textstyle{X]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[G×X\textstyle{[G\times X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}X]\textstyle{X]} of the translation groupoids, whose domain is the trivial one. The naturality of the spectral sequence in Theorem 2.1 and the proof of Theorem 2.2 ii) enable us to obtain a morphism {ιr}:{Er,,dr}{Er,,dr}\{\iota^{*}_{r}\}:\{E_{r}^{*,*},d_{r}\}\to\{{}^{\backprime}E_{r}^{*,*},{}^{\backprime}d_{r}\} of spectral sequences and a commutative diagram

H(EG×GX;K)\textstyle{H^{*}(EG\times_{G}X;K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j\scriptstyle{j^{*}}E0,\textstyle{E_{\infty}^{0,*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ \ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι\scriptstyle{\iota^{*}_{\infty}}E20,\textstyle{E_{2}^{0,*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι2\scriptstyle{\iota^{*}_{2}}\scriptstyle{\cong}𝕂𝕂[G]H(X;𝕂)\textstyle{{\mathbb{K}}\Box_{{\mathbb{K}}[G]^{\vee}}H^{*}(X;{\mathbb{K}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}inclusion\scriptstyle{inclusion}H(X;𝕂)\textstyle{H^{*}(X;{\mathbb{K}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}=\scriptstyle{=}E0,\textstyle{{}^{\backprime}E_{\infty}^{0,*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}=\scriptstyle{=}E20,\textstyle{{}^{\backprime}E_{2}^{0,*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}=\scriptstyle{=}H(X;𝕂),\textstyle{H^{*}(X;{\mathbb{K}}),}

where jj is the composite XE{e}×XEG×XEG×GXX\to E\{e\}\times X\to EG\times X\to EG\times_{G}X. The proof of Theorem 5.2 shows that the upper arrows are isomorphisms if the characteristic of 𝕂{\mathbb{K}} is coprime with |G||G|. Then, the isomorphism of algebras in the theorem is regarded as the map induced by j:XEG×GXj:X\to EG\times_{G}X mentioned above.

Theorem 5.2 allows us to compute the cohomology algebra of the free loop space of a Borel construction.

Corollary 5.4.

(cf. [32, Corollary 6.5]) Let GG be a finite group acting on a space MM and 𝒫G(M){\mathcal{P}}_{G}(M) the space defined in (4.1). Suppose that H(𝒫G(M);𝕂)H^{*}({\mathcal{P}}_{G}(M);{\mathbb{K}}) is locally finite and (ch(𝕂),|G|)=1(\text{\em ch}({\mathbb{K}}),|G|)=1. Then as an algebra

H(L(EG×GM);𝕂)𝕂𝕂[G]H(𝒫G(M);𝕂).H^{*}(L(EG\times_{G}M);{\mathbb{K}})\cong{\mathbb{K}}\Box_{{\mathbb{K}}[G]^{\vee}}H^{*}({\mathcal{P}}_{G}(M);{\mathbb{K}}).
Proof.

Theorem 5.2 and Corollary A.2 imply the assertion. ∎

Remark 5.5.

Let MM be a simply-connected GG-space whose cohomology with coefficients in a field 𝕂{\mathbb{K}} is locally finite. Then, the total complex of the bar complex, which computes the E2E_{2}-term of the EMSS for the fibre square (4.7), is locally finite and hence so is H(𝒫G(M);𝕂)H^{*}({\mathcal{P}}_{G}(M);{\mathbb{K}}) if GG is finite.

Remark 5.6.

Let GMpM/GG\to M\stackrel{{\scriptstyle p}}{{\to}}M/G be a principal GG-bundle with finite fibre GG. The projection pp induces natural maps p~:𝒫G(M)L(M/G)\widetilde{p}:{\mathcal{P}}_{G}(M)\to L(M/G) and p¯:EG×G𝒫G(M)L(M/G)\overline{p}:EG\times_{G}{\mathcal{P}}_{G}(M)\to L(M/G) with p¯j=p~\overline{p}\circ j=\widetilde{p}, where jj is the map described in Remark 5.3. By virtue of Proposition A.1 and Theorem 5.2, we see that p~\widetilde{p} gives rise to an isomorphism p~:H(L(M/G);𝕂)K𝕂[G]H(𝒫G(M);𝕂)\widetilde{p}^{*}:H^{*}(L(M/G);{\mathbb{K}})\stackrel{{\scriptstyle\cong}}{{\longrightarrow}}K\Box_{{\mathbb{K}}[G]^{\vee}}H^{*}({\mathcal{P}}_{G}(M);{\mathbb{K}}) of algebras if (ch(𝕂),G)=1(\text{ch}({\mathbb{K}}),{G})=1.

Corollary 5.7.

Let GG be a finite group and MM a simply-connected GG-space whose rational cohomology is locally finite. Assume further that MM is formal and for each gGg\in G, the map g:MMg:M\to M induced by gg via the GG-action is formal in the sense of [44, Définition 2.3.3]; see also [3, Definition 2.3]. Then as an algebra

H(L(EG×GM);)[G](gGHH(H(M;),H(M;)g)),H^{*}(L(EG\times_{G}M);{\mathbb{Q}})\cong{\mathbb{Q}}\Box_{{\mathbb{Q}}[G]^{\vee}}\big{(}\oplus_{g\in G}HH_{*}(H^{*}(M;{\mathbb{Q}}),H^{*}(M;{\mathbb{Q}})_{g})\big{)},

where HH(A,H)HH_{*}(A,H) denotes the Hochschild homology of a graded commutative algebra AA with coefficients in an AeA^{e}-algebra HH.

Proof.

The assumption on the map gg implies that the map 1×g:MM×M1\times g:M\to M\times M is formal. In fact, let MAPL(M){\mathcal{M}}_{M}\stackrel{{\scriptstyle\simeq}}{{\to}}A_{PL}^{*}(M) be a minimal model for MM, where APL(M)A_{PL}(M) denotes the complex of polynomial forms on MM; see Appendix B.1. Suppose that gg is (φM)(\varphi_{M})-(φM)(\varphi_{M}^{\prime})-formal with quasi-isomorphisms φM:MH(M)\varphi_{M}:{\mathcal{M}}_{M}\stackrel{{\scriptstyle\simeq}}{{\to}}H^{*}({\mathcal{M}}_{M}) and φM:MH(M)\varphi_{M}^{\prime}:{\mathcal{M}}_{M}\stackrel{{\scriptstyle\simeq}}{{\to}}H^{*}({\mathcal{M}}_{M}). Then the map 1×g1\times g is (φMφM)(\varphi_{M}\otimes\varphi_{M}^{\prime})-(φM)(\varphi_{M}^{\prime})-formal. This fact enables us to deduce that

H(𝒫g(M);)HH(APL(M),APL(M)g)HH(H(M;),H(M;)g)H^{*}({\mathcal{P}}_{g}(M);{\mathbb{Q}})\cong HH_{*}(A_{PL}^{*}(M),A_{PL}^{*}(M)_{g}\big{)}\cong HH_{*}(H^{*}(M;{\mathbb{Q}}),H^{*}(M;{\mathbb{Q}})_{g}\big{)}

as algebras. Corollary 5.4 yields the result. ∎

We consider a spectral sequence for the inertia groupoid associated with a topological groupoid 𝒢\mathcal{G}; see [6, Chapters 11 and 16] for such a groupoid and its importance in string topology for stacks. In particular, we have a canonical morphism from the inertia groupoid of 𝒢\mathcal{G} to the loop groupoid; see [6, 12.4].

Let 𝒢=[G1G0]\mathcal{G}=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 9.72014pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-9.72014pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[G_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 24.72014pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 24.72014pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 24.72014pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{G_{0}]}$}}}}}}}\ignorespaces}}}}\ignorespaces be a topological groupoid. By definition, a space XX with a map p:XG0p:X\to G_{0} is a 𝒢\mathcal{G}-space if XX is endowed with an action G1×G0ptXXG_{1}{}^{t}\times^{p}_{G_{0}}X\to X, where G1×G0ptXG_{1}{}^{t}\!\times^{p}_{G_{0}}X denotes the pullback of pp along the target map t:G1G0t:G_{1}\to G_{0}. Then, the translation groupoid associated with the 𝒢\mathcal{G}-space XX is defined by 𝒢X:=[G1×G0ptXX]\mathcal{G}\ltimes X:=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 20.84138pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-20.84138pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[G_{1}{}^{t}\!\times^{p}_{G_{0}}X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 35.84138pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 35.84138pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 35.84138pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{X]}$}}}}}}}\ignorespaces}}}}\ignorespaces whose source and target maps are defined by the projection in the second factor and the action, respectively.

For a topological groupoid 𝒢\mathcal{G}, we define a subspace S𝒢S_{\mathcal{G}} of G1G_{1} and map p:S𝒢G0p:S_{\mathcal{G}}\to G_{0} by S𝒢:={gG1s(g)=t(g)}S_{\mathcal{G}}:=\{g\in G_{1}\mid s(g)=t(g)\} and p(g)=s(g)p(g)=s(g), respectively. Then, the conjugation on G1G_{1} gives rise to an action G1×G0ptS𝒢S𝒢G_{1}{}^{t}\!\times_{G_{0}}^{p}S_{\mathcal{G}}\to S_{\mathcal{G}}. The translation groupoid Λ𝒢:=𝒢S𝒢\Lambda\mathcal{G}:=\mathcal{G}\ltimes S_{\mathcal{G}} is called the inertia groupoid associated with 𝒢\mathcal{G}. We also define the evaluation map

ev0:Λ𝒢=[G1×G0ptS𝒢S𝒢]𝒢=[G1G0]ev_{0}:\Lambda\mathcal{G}=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 23.45079pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-23.45079pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[G_{1}{}^{t}\!\times_{G_{0}}^{p}S_{\mathcal{G}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 38.45079pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 38.45079pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 38.45079pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{S_{\mathcal{G}}]}$}}}}}}}\ignorespaces}}}}\ignorespaces\to\mathcal{G}=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 9.72014pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-9.72014pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[G_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 24.72014pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 24.72014pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 24.72014pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{G_{0}]}$}}}}}}}\ignorespaces}}}}\ignorespaces

by the projection in the first factor in morphisms and by the source map of 𝒢\mathcal{G} in the objects.

Let GG be a finite group and GM:=[G×MM]G\ltimes M:=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 19.82703pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-19.82703pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[G\times M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 34.82703pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 34.82703pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 34.82703pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{M]}$}}}}}}}\ignorespaces}}}}\ignorespaces a translation groupoid. Applying the construction above, we have the inertia groupoid Λ(GM)\Lambda(G\ltimes M). Moreover, we see that the groupid has the form

Λ(GM)=[G×gG(Mg×{g})gG(Mg×{g})],\Lambda(G\ltimes M)=\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 45.85919pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-45.85919pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[G\times\coprod_{g\in G}(M^{g}\times\{g\})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 60.85919pt\raise-2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 60.85919pt\raise 2.15277pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 60.85919pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\coprod_{g\in G}(M^{g}\times\{g\})]}$}}}}}}}\ignorespaces}}}}\ignorespaces,

where MgM^{g} denotes the space of fixed points of gg. Observe that the source map is the projection of the second factor and the conjugation gives rise to the target map tt, more precisely, t(h,(m,g))=(hm,hgh1)t(h,(m,g))=(hm,hgh^{-1}); see [6, (17.2.1)]. By virtue of Theorem 2.2 ii), we have the following result.

Proposition 5.8.

Let GMG\ltimes M be a transformation groupoid for which GG is a finite group and H(Mg)H^{*}(M^{g}) is locally finite for each gGg\in G. Then there exists a spectral sequence {Er,dr}\{E_{r},d_{r}\} converging to the cohomology of the classifying space BΛ(GM)\text{\em B}\Lambda(G\ltimes M) of the inertia groupoid Λ(GM)\Lambda(G\ltimes M), as an algebra with

E2p,qCotorH0(G)p,q(𝕂,gGH(Mg;𝕂)).E_{2}^{p,q}\cong\text{\em Cotor}^{p,q}_{H^{0}(G)}({\mathbb{K}},\oplus_{g\in G}H^{*}(M^{g};{\mathbb{K}})).

Moreover, the evaluation map gives rise to a morphism from the spectral sequence in Theorem 2.2 ii) to {Er,dr}\{E_{r},d_{r}\}. Assume further that (ch(𝕂),|G|)=1(\text{\em ch}({\mathbb{K}}),|G|)=1, then one has H(BΛ(GM);𝕂)𝕂𝕂[G](gGH(Mg;𝕂))H^{*}(B\Lambda(G\ltimes M);{\mathbb{K}})\cong{\mathbb{K}}\Box_{{\mathbb{K}}[G]^{\vee}}(\oplus_{g\in G}H^{*}(M^{g};{\mathbb{K}})) as an algebra.

Let GG be a discrete group. We consider the free loop space LBGLBG of the classifying space BG=EG/G=EG×GBG=EG/G=EG\times_{G}\ast.

Proposition 5.9.

(cf. [32, Example 6.2] and [7, Proposition 2.12.2]) Let GG be a finite group. Then as an algebra,

H(LBG;𝕂)Cotor𝕂[G](𝕂,(𝕂[G]ad)),H^{*}(LBG;{\mathbb{K}})\cong\text{\em Cotor}_{{\mathbb{K}}[G]^{\vee}}^{*}({\mathbb{K}},({\mathbb{K}}[G]_{\text{\em ad}})^{\vee}),

where 𝕂[G]ad{\mathbb{K}}[G]_{\text{\em ad}} denotes the adjoint representation of GG. Especially, if GG is abelian, then H(LBG;𝕂)H(G;𝕂)|G|H^{*}(LBG;{\mathbb{K}})\cong H^{*}(G;{\mathbb{K}})^{\oplus|G|} as an algebra.

Proof.

It follows that H(𝒫g();𝕂)=𝕂H_{*}({\mathcal{P}}_{g}(\ast);{\mathbb{K}})={\mathbb{K}} for each gGg\in G. By the definition of the GG-action on 𝒫G(){\mathcal{P}}_{G}(\ast), we see that H(𝒫G();𝕂)𝕂[G]adH_{*}({\mathcal{P}}_{G}(\ast);{\mathbb{K}})\cong{\mathbb{K}}[G]_{\text{ad}}. Then, the weak equivalence in (4.2)(\ref{eq:w}) and the spectral sequence {Er,,dr}\{E_{r}^{*,*},d_{r}\} in Theorem 2.2 ii) enable us to deduce the first assertion. Observe that E2,q=0E_{2}^{*,q}=0 for q>0q>0.

The latter half follows from the fact that the GG-action on 𝕂[G]ad{\mathbb{K}}[G]_{\text{ad}} is trivial. In particular, the explicit formula (2.1) of the multiplication on the cotorsion product yields the result on the algebra structure. ∎

Remark 5.10.

If GG is abelian, then the classifying space BGBG is an H-space. This allows us to obtain a homotopy equivalence BG×ΩBGLBGBG\times\Omega BG\stackrel{{\scriptstyle\simeq}}{{\to}}LBG induced by the product on BGBG. The latter half of Proposition 5.9 also follows from the fact.

We conclude this section with comments on a spectral sequence converging to the free loop space of a Borel construction, which is obtained by Theorem 2.2 ii).

Remark 5.11.

Let GG be a finite group with an action on a space MM and {Er,,dr}\{E_{r}^{*,*},d_{r}\} the spectral sequence in Theorem 2.2 ii) for the Borel construction EG×G𝒫G(M)EG\times_{G}{\mathcal{P}}_{G}(M). By Corollary A.2, we see that L(EG×GM)L(EG\times_{G}M) is connected to EG×G𝒫G(M)EG\times_{G}{\mathcal{P}}_{G}(M) with weak homotopy equivalences. Thus, under the same condition as in Corollary 5.4 but (ch(𝕂),|G|)=1(\text{ch}({\mathbb{K}}),|G|)=1, the spectral sequence converges to H(L(EG×GM);𝕂)H^{*}(L(EG\times_{G}M);{\mathbb{K}}) as an algebra.

The form of the E2E_{2}-term enables us to deduce that the vertical edge (qq-axis) E20,E_{2}^{0,*} is isomorphic to the cotensor product 𝕂𝕂[G]H(𝒫G(M);𝕂){\mathbb{K}}\Box_{{\mathbb{K}}[G]^{\vee}}H^{*}({\mathcal{P}}_{G}(M);{\mathbb{K}}), which is a subspace of H(𝒫G(M);𝕂)H^{*}({\mathcal{P}}_{G}(M);{\mathbb{K}}) detected by the transfer homomorphism if (ch(𝕂),|G|)=1(\text{ch}({\mathbb{K}}),|G|)=1; see [9, Chapter II, Theorem 19.2] and Lemma 5.1.

We consider the horizontal edge (pp-axis). Since the trivial map v:𝒫g(M)v:{\mathcal{P}}_{g}(M)\to\ast gives a GG-equivariant map H0(𝒫G(M);𝕂)𝕂[G]adH_{0}({\mathcal{P}}_{G}(M);{\mathbb{K}})\to{\mathbb{K}}[G]_{\text{ad}}, it follows from Proposition 5.9 that the edge E2,0E_{2}^{*,0} is a module over the cotorsion product Cotor𝕂[G](𝕂,(𝕂[G]ad))\text{Cotor}_{{\mathbb{K}}[G]^{\vee}}^{*}({\mathbb{K}},({\mathbb{K}}[G]_{\text{ad}})^{\vee}) and hence H(LBG;𝕂)H^{*}(LBG;{\mathbb{K}}) via the morphism

v:H(LBG;𝕂)Cotor𝕂[G],0(𝕂,H(𝒫G(M);𝕂))v^{*}:H^{*}(LBG;{\mathbb{K}})\to\text{Cotor}_{{\mathbb{K}}[G]^{\vee}}^{*,0}({\mathbb{K}},H^{*}({\mathcal{P}}_{G}(M);{\mathbb{K}}))

of algebras induced by vv. Observe that the vv^{*} is an isomorphism of algebras if MM is simply-connected. In fact, we see that each 𝒫g(M){\mathcal{P}}_{g}(M) is connected in that case. We remark that the underlying algebra H(LBG;𝕂)H^{*}(LBG;{\mathbb{K}}) is indeed the Hochschild cohomology of the group ring 𝕂[G]{\mathbb{K}}[G]; see [7, Proposition 2.12.2].

Each differential on Er,0E_{r}^{*,0} for r2r\geq 2 is trivial. Therefore, the multiplicative structure on the spectral sequence gives rise to an H(LBG;𝕂)H^{*}(LBG;{\mathbb{K}})-module structure on the spectral sequence; that is, each term Er,E_{r}^{*,*} admits an H(LBG;𝕂)H^{*}(LBG;{\mathbb{K}})-module structure defined by

:Hp(LBG;𝕂)Er,\textstyle{\bullet:H^{p}(LBG;{\mathbb{K}})\otimes E_{r}^{*,*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}v1\scriptstyle{v^{*}\otimes 1}E2p,0Er,\textstyle{E_{2}^{p,0}\otimes E_{r}^{*,*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pr1\scriptstyle{p_{r}\otimes 1}Erp,0Er,\textstyle{E_{r}^{p,0}\otimes E_{r}^{*,*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}m\scriptstyle{m}Er+p,,\textstyle{E_{r}^{*+p,*},}

where prp_{r} is the canonical projection and mm is the product structure on the ErE_{r}-term. Thus we see that dr(ax)=(1)padr(x)d_{r}(a\bullet x)=(-1)^{p}a\bullet d_{r}(x) for aHp(LBG;𝕂)a\in H^{p}(LBG;{\mathbb{K}}) and xEr,x\in E_{r}^{*,*}.

6. The de Rham cohomology of the diffeological free loop space of a non-simply connected manifold

Let GG be a finite group acting freely and smoothly on a manifold MM. Then, a principal GG-bundle of the form GMpM/GG\to M\stackrel{{\scriptstyle p}}{{\to}}M/G in the category of manifolds is obtained. Let 𝒫G(M){\mathcal{P}}_{G}^{\infty}(M) be the diffeological space obtained by applying the construction (4.1) in 𝖣𝗂𝖿𝖿\mathsf{Diff} the category of diffeological spaces; see Appendix B. We consider a smooth map p~:𝒫G(M)L(M/G)\widetilde{p}:{\mathcal{P}}_{G}^{\infty}(M)\to L^{\infty}(M/G) defined by p~((γ,g))=pγ\widetilde{p}((\gamma,g))=p\circ\gamma, where L(M/G)L^{\infty}(M/G) denotes the diffeological free loop space of M/GM/G; see the pullback diagram (4.7).

For a diffeological space XX, we may write HDR(X)H^{*}_{DR}(X) for the singular de Rham cohomology H(ADR(SD(X)))H^{*}(A_{DR}(S^{D}_{\bullet}(X))); see Appendix B. The purpose of this section is to prove the following theorem.

Theorem 6.1.

Under the same setting as above, suppose further that MM is simply connected. Then, the smooth map p~\widetilde{p} gives rise to a well-defined isomorphism

p~:HDR(L(M/G))[G]HDR(𝒫G(M))\widetilde{p}^{*}:H_{DR}^{*}(L^{\infty}(M/G))\stackrel{{\scriptstyle\cong}}{{\to}}{\mathbb{R}}\Box_{{\mathbb{R}}[G]^{\vee}}H^{*}_{DR}({\mathcal{P}}_{G}^{\infty}(M))

of algebras.

Each component 𝒫g(M){\mathcal{P}}_{g}^{\infty}(M) of 𝒫G(M){\mathcal{P}}_{G}^{\infty}(M) is constructed with a pullback of the form (4.7). Then, by utilizing Theorem B.5, we may represent algebra generators in the de Rham cohomology HDR(L(M/G))H^{*}_{DR}(L^{\infty}(M/G)) with differential forms on MM via Chen’s iterated integral map and the factor map, which connects the Souriau-de Rham complex and the singular de Rham complex; see Appendix B.1. Indeed, the idea is realized in Theorem 6.4.

Before proving Theorem 6.1, we recall the smoothing theorem due to Kihara for a particular case. Let MM and NN be diffeological spaces and C(M,N)C^{\infty}(M,N) the space of smooth maps from MM to NN with the functional diffeology. We recall the functor D:𝖣𝗂𝖿𝖿𝖳𝗈𝗉D:\mathsf{Diff}\to\mathsf{Top} from Appendix B. Then, we see that D(Δstn)D(\Delta^{n}_{st}) is homeomorphic to Δn\Delta^{n} the standard nn simplex which is a subspace of n+1{\mathbb{R}}^{n+1}; see [24]. Moreover, the inclusion i:D(C(M,N))C0(DM,DN)i:D(C^{\infty}(M,N))\to C^{0}(DM,DN) is continuous; see [12, Proposition 4.2]. Thus, it follows that the functor DD induces a morphism ξ:SD(C(M,N))Sing(C0(DM,DN))\xi:S^{D}_{\bullet}(C^{\infty}(M,N))\to\text{Sing}_{\bullet}(C^{0}(DM,DN)) of simplicial sets.

Theorem 6.2.

(Smoothing theorem [24, Theorems 1.1 and 1.7]) Let MM and NN be finite-dimensional manifolds. Then, the well-defined map

ξ:SD(C(M,N))Sing(C0(DM,DN))\xi:S^{D}_{\bullet}(C^{\infty}(M,N))\to\text{\em Sing}_{\bullet}(C^{0}(DM,DN))

is a weak homotopy equivalence.

For a manifold MM, we consider the composite

λ:=i(Dj):D(𝒫g(M))\textstyle{\lambda:=i\circ(Dj):D({\mathcal{P}}_{g}^{\infty}(M))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D(C([0,1],M))\textstyle{D(C^{\infty}([0,1],M))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}C0(D[0,1],DM),\textstyle{C^{0}(D[0,1],DM),}

where jj denotes the smooth inclusion 𝒫g(M)C([0,1],M){\mathcal{P}}_{g}^{\infty}(M)\to C^{\infty}([0,1],M). Since DM=MDM=M and D[0,1]D[0,1] is the subspace II of {\mathbb{R}} (see [12, Lemma 3.16]), it follows that λ:D(𝒫g(M))𝒫g(M)\lambda:D({\mathcal{P}}_{g}^{\infty}(M))\to{\mathcal{P}}_{g}(M) is continuous. Therefore, the composite ξ:=λξ:SD(𝒫g(M))Sing(𝒫g(M))\xi^{\prime}:=\lambda_{*}\circ\xi:S^{D}_{\bullet}({\mathcal{P}}_{g}^{\infty}(M))\to\text{Sing}_{\bullet}({\mathcal{P}}_{g}(M)) is a morphism of simplicial sets.

The following lemma is a key to proving Theorem 6.1.

Lemma 6.3.

Let MM be a simply-connected manifold. Then, one has a sequence of quasi-isomorphisms

APL(Sing(𝒫g(M)))\textstyle{A_{\text{PL}}(\text{\em Sing}_{\bullet}({\mathcal{P}}_{g}(M)))\otimes_{{\mathbb{Q}}}{\mathbb{R}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(ξ)\scriptstyle{(\xi^{\prime})^{*}}\scriptstyle{\simeq}APL(SD(𝒫g(M)))\textstyle{A_{\text{PL}}(S^{D}_{\bullet}({\mathcal{P}}_{g}^{\infty}(M)))\otimes_{{\mathbb{Q}}}{\mathbb{R}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ζ\scriptstyle{\zeta}\scriptstyle{\simeq}ADR(SD(𝒫g(M))).\textstyle{A_{\text{DR}}(S^{D}_{\bullet}({\mathcal{P}}_{g}^{\infty}(M))).}
Proof.

The result [29, Corollary 3.5] allows us to obtain a quasi-isomorphism ζ\zeta. We consider a commutative diagram

H(ADR(SD(𝒫g(M)))\textstyle{H^{*}(A_{DR}(S^{D}_{\bullet}({\mathcal{P}}_{g}^{\infty}(M)))}TorADR(SD(M×2))(ADR(SD(M)),ADR(SD(MI)))\textstyle{\text{Tor}_{A_{DR}(S^{D}(M^{\times 2}))}^{*}(A_{DR}(S^{D}(M)),A_{DR}(S^{D}(M^{I})))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}EM1\scriptstyle{EM_{1}}H(APL(SD(𝒫g(M))))\textstyle{H^{*}(A_{PL}(S^{D}_{\bullet}({\mathcal{P}}_{g}^{\infty}(M))))_{\mathbb{R}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H(ζ)\scriptstyle{H(\zeta)}\scriptstyle{\cong}TorAPL(SD(M×2))(APL(SD(M)),APL(SD(MI)))\textstyle{\text{Tor}_{A_{PL}(S^{D}(M^{\times 2}))}^{*}(A_{PL}(S^{D}(M)),A_{PL}(S^{D}(M^{I})))_{\mathbb{R}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}EM2\scriptstyle{EM_{2}}Tor(ζ,ζ)\scriptstyle{\text{Tor}(\zeta,\zeta)}\scriptstyle{\cong}H(APL(Sing(𝒫g(M))))\textstyle{H^{*}(A_{PL}(\text{Sing}_{\bullet}({\mathcal{P}}_{g}(M))))_{\mathbb{R}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H((ξ))\scriptstyle{H((\xi^{\prime})^{*})}TorAPL(M×2)(APL(M),APL(MI))\textstyle{\text{Tor}_{A_{PL}(M^{\times 2})}^{*}(A_{PL}(M),A_{PL}(M^{I}))_{\mathbb{R}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}EM3\scriptstyle{EM_{3}}Tor(ξ,ξ)\scriptstyle{\text{Tor}(\xi^{*},\xi^{*})}

in which the horizontal maps are induced by the Eilenberg–Moore map; see [21, 20.6]. Here, we write APL(X)A_{PL}(X) for APL(Sing(X))A_{PL}(\text{Sing}_{\bullet}(X)) in the right-hand corner and ()(\ )_{\mathbb{R}} denotes the tensor product ()(\ )\otimes_{\mathbb{Q}}{\mathbb{R}}.

By assumption, the manifold MM is simply connected. Then, the proofs of [29, Theorem 5.5] and [21, 20.6] yield that EM1EM_{1} and EM3EM_{3} are isomorphisms, respectively. Since ζ\zeta is a quasi-isomorphism, it follows that the vertical maps in the upper square are isomorphisms and then so is EM2EM_{2}. We see that MIM^{I} is smooth homotopy equivalent to MM. Therefore, Theorem 6.2 implies that Tor(ξ,ξ){\text{Tor}(\xi^{*},\xi^{*})} is an isomorphism. It turns out that (ξ)(\xi^{\prime})^{*} is a quasi-isomorphism. ∎

Proof of Theorem 6.1.

For the translation groupoid

[G×𝒫G(M)\textstyle{[G\times{\mathcal{P}}_{G}^{\infty}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}t\scriptstyle{t}s\scriptstyle{s}𝒫G(M)]\textstyle{{\mathcal{P}}_{G}^{\infty}(M)]}

in which ss and tt are defined by the projection and the action, respectively, we see that sp~=tp~s\circ\widetilde{p}=t\circ\widetilde{p}. Then the map p~:HDR(L(M/G))HDR(𝒫G(M))\widetilde{p}^{*}:H^{*}_{DR}(L^{\infty}(M/G))\to H^{*}_{DR}({\mathcal{P}}_{G}^{\infty}(M)) induced by p~\widetilde{p} factors through [G]HDR(𝒫G(M)){\mathbb{R}}\Box_{{\mathbb{R}}[G]^{\vee}}H^{*}_{DR}({\mathcal{P}}_{G}^{\infty}(M)). We show that the morphism p~\widetilde{p}^{*} of algebras in the theorem is an isomorphism. Consider a commutative diagram

H(C0(S1,M/G);)\textstyle{H^{*}(C^{0}(S^{1},M/G);{\mathbb{R}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(ξ~)\scriptstyle{(\widetilde{\xi})^{*}}\scriptstyle{\cong}HDR(C(S1,M/G))\textstyle{H_{DR}^{*}(C^{\infty}(S^{1},M/G))}H(L(M/G);)\textstyle{H^{*}(L(M/G);{\mathbb{R}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(ξ~)\scriptstyle{(\widetilde{\xi})^{*}}(q)\scriptstyle{(q^{*})^{*}}\scriptstyle{\cong}p~\scriptstyle{\widetilde{p}^{*}}\scriptstyle{\cong}HDR(L(M/G))\textstyle{H_{DR}^{*}(L^{\infty}(M/G))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(q)\scriptstyle{(q^{*})^{*}}\scriptstyle{\cong}p~\scriptstyle{\widetilde{p}^{*}}[G]H(𝒫G(M);)\textstyle{{\mathbb{R}}\Box_{{\mathbb{R}}[G]^{\vee}}H^{*}({\mathcal{P}}_{G}(M);{\mathbb{R}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H(ξ~1)\scriptstyle{H(\widetilde{\xi}_{1})}[G]HDR(𝒫G(M)),\textstyle{{\mathbb{R}}\Box_{{\mathbb{R}}[G]^{\vee}}H^{*}_{DR}({\mathcal{P}}^{\infty}_{G}(M)),}

where ξ~1\widetilde{\xi}_{1} denotes the composite of quasi-isomorphisms in Lemma 6.3 and ξ~\widetilde{\xi} is the composite of the morphism induced by ξ\xi in Theorem 6.2 and the quasi-isomorphism ζ:APL(K)ADR(K)\zeta:A_{PL}(K)\otimes_{{\mathbb{Q}}}{\mathbb{R}}\to A_{DR}(K) for a simplicial set KK in [29, Corollary 3.5]. Thus, Theorem 6.2 implies that (ξ~)(\widetilde{\xi})^{*} in the top row is an isomorphism. We observe that, by Remark 5.6, the map p~\widetilde{p}^{*} on the left-hand side is an isomorphism.

Lemma C.1 implies that map (q)(q^{*})^{*} on the right-hand side induced by the smooth map q:IS1q:I\to S^{1} in 𝖣𝗂𝖿𝖿\mathsf{Diff} is an isomorphism. A usual argument on the quotient map q:IS1q:I\to S^{1} shows that the projection induces a weak homotopy equivalence (q):C0(S1,M/G)L(M/G)(q^{*}):C^{0}(S^{1},M/G)\to L(M/G) and hence the left-hand side map (q)(q^{*})^{*} is an isomorphism. Since the map H(ξ~1)H(\widetilde{\xi}_{1}) in the lowest row is also an isomorphism, it follows that the right-hand side map p~\widetilde{p}^{*} is an isomorphism. We have the result. ∎

The following result follows from Theorems 4.1 and 6.1.

Theorem 6.4.

One has sequences

HDR(LP2m+1)\textstyle{H^{*}_{DR}(L^{\infty}{\mathbb{R}}P^{2m+1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p~\scriptstyle{\widetilde{p}^{*}}\scriptstyle{\cong}[G]HDR(𝒫G(S2m+1))\textstyle{{\mathbb{R}}\Box_{{\mathbb{R}}[G]^{\vee}}H^{*}_{DR}({\mathcal{P}}_{G}^{\infty}(S^{2m+1}))}((α𝖨𝗍(v2m+1))[α𝖨𝗍([v2m+1)])2and\textstyle{\big{(}\wedge(\alpha\circ\mathsf{It}(v_{2m+1}))\otimes{\mathbb{R}}[\alpha\circ\mathsf{It}([v_{2m+1})]\big{)}^{\oplus 2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ \ \ \text{and}}\scriptstyle{\cong}HDR(LP2m)\textstyle{H^{*}_{DR}(L^{\infty}{\mathbb{R}}P^{2m})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p~\scriptstyle{\widetilde{p}^{*}}\scriptstyle{\cong}[G]HDR(𝒫G(S2m))\textstyle{{\mathbb{R}}\Box_{{\mathbb{R}}[G]^{\vee}}H^{*}_{DR}({\mathcal{P}}_{G}^{\infty}(S^{2m}))}((α𝖨𝗍(v2m[v2m]))[α𝖨𝗍(1[v2m|v2m]))\textstyle{\big{(}\wedge(\alpha\circ\mathsf{It}(v_{2m}[v_{2m}]))\otimes{\mathbb{R}}[\alpha\circ\mathsf{It}(1[v_{2m}|v_{2m}])\big{)}\oplus{\mathbb{R}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}

of isomorphisms of algebras, where vnv_{n} denotes the volume form on HDR(Sn)H^{*}_{DR}(S^{n}), 𝖨𝗍\mathsf{It} and α\alpha are Chen’s iterated integral map and the factor map, respectively; see Appendices B.1 and B.2.

Proof.

We prove the result on the isomorphisms in the second sequence. By virtue of Theorem B.5, we see that the composite

Ω(M)1×gB¯(Ω~(M))\textstyle{\Omega(M)\otimes_{1\times g}\overline{B}(\widetilde{\Omega}(M))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝖨𝗍\scriptstyle{\mathsf{It}}Ω(𝒫g(M))\textstyle{\Omega({\mathcal{P}}_{g}^{\infty}(M))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{\alpha}ADR(SD(𝒫g(M)))\textstyle{A_{DR}(S^{D}_{\bullet}({\mathcal{P}}_{g}^{\infty}(M)))}

is a quasi-isomorphism for M=SnM=S^{n}; see Proposition B.4. It is immediate to show that 1×g1\times g is an induction. Then, Theorem B.5 is applicable to this case. Moreover, it follows from Theorem 6.1, Lemma 6.3 and the computation in Theorem 4.1 that

HDR(LP2m)H(ADR(SD(𝒫τ(S2m)))H(𝒫τ(S2m));)H^{*}_{DR}(L^{\infty}{\mathbb{R}}P^{2m})\cong H^{*}(A_{DR}(S^{D}_{\bullet}({\mathcal{P}}_{\tau}^{\infty}(S^{2m})))\oplus{\mathbb{R}}\cong H^{*}({\mathcal{P}}_{\tau}(S^{2m}));{\mathbb{R}})\oplus{\mathbb{R}}

as algebras. Observe that p~\widetilde{p}^{*} gives the first isomorphism. The same computation with the cyclic bar complex as in [28, Theorem 2.1] allows us to deduce that v2m[v2m]v_{2m}[v_{2m}] and 1[v2m|v2m]1[v_{2m}|v_{2m}] are non-exact cocycles. With the indecomposable elements of the second algebra in Theorem 4.1, we see that deg(xu)=4m1=degv2m[v2m]\deg(x\otimes u)=4m-1=\deg v_{2m}[v_{2m}] and degw=4m2=deg1[v2m|v2m]\deg w=4m-2=\deg 1[v_{2m}|v_{2m}]. Thus, we have the result. The same argument as above enables us to obtain the isomorphisms in the first sequence. ∎

Acknowledgements. The author thanks Takahito Naito, Shun Wakatsuki and Toshihiro Yamaguchi for many valuable suggestions on the first draft of this manuscript. He is also grateful to Jean-Claude Thomas and Luc Menichi for fruitful discussions on the computation in Section 4. The author would like to express his gratitude to Jim Stasheff and the referee for suggestions in revising a version of this manuscript. This work was partially supported by JSPS KAKENHI Grant Numbers JP19H05495 and JP21H00982.

Appendix A The free loop space of a global quotient

Let GMpM/GG\to M\stackrel{{\scriptstyle p}}{{\to}}M/G be a principal GG-bundle with discrete fibre GG. Under the same notations as in Section 4, we show the following proposition.

Proposition A.1.

(cf. [6, Proposition 5.9]) The map p¯:EG×G𝒫G(M)L(M/G)\overline{p}:EG\times_{G}{\mathcal{P}}_{G}(M)\to L(M/G) induced naturally by the projection p:MM/Gp:M\to M/G is a weak homotopy equivalence.

Let GG be a discrete group acting on a space MM. Let gg be an element of GG. We recall a fibre square of the form

𝒫g(M)\textstyle{{\mathcal{P}}_{g}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}qg\scriptstyle{q_{g}}M[0,1]\textstyle{M^{[0,1]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(ev0,ev1)\scriptstyle{(ev_{0},ev_{1})}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕg\scriptstyle{\phi_{g}}M×M,\textstyle{M\times M,}

where eviev_{i} denotes the evaluation map at ii for i=0,1i=0,1 and ϕg\phi_{g} is the map defined by ϕg(x)=(x,gx)\phi_{g}(x)=(x,gx) for xMx\in M. Thus, for each mMm\in M, we have a fibration

𝒫Gm(M)\textstyle{{\mathcal{P}}_{G}^{m}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒫G(M)\textstyle{{\mathcal{P}}_{G}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q:=qg\scriptstyle{q:=\coprod q_{g}}M,\textstyle{M,}

where 𝒫Gm(M)=gG𝒫gm(M){\mathcal{P}}_{G}^{m}(M)=\coprod_{g\in G}{\mathcal{P}}_{g}^{m}(M) and 𝒫gm(M):={γ:[0,1]Mγ(0)=m,γ(1)=gγ(0)=gm}.{\mathcal{P}}_{g}^{m}(M):=\{\gamma:[0,1]\to M\mid\gamma(0)=m,\gamma(1)=g\gamma(0)=gm\}.

Since the projection p2:EG×MMp_{2}:EG\times M\to M is a GG-equivariant map and a homotopy equivalence, it follows that p2~:𝒫Gm(EG×M)𝒫Gm(M)\widetilde{p_{2}}:{\mathcal{P}}_{G}^{m}(EG\times M)\to{\mathcal{P}}_{G}^{m}(M) induced by p2p_{2} is weak homotopy equivalent and hence so is p2~:EG×G𝒫G(EG×M)EG×G𝒫G(M)\widetilde{p_{2}}:EG\times_{G}{\mathcal{P}}_{G}(EG\times M)\to EG\times_{G}{\mathcal{P}}_{G}(M). Moreover, Proposition A.1 is applicable to the GG-bundle GEG×MEG×GMG\to EG\times M\to EG\times_{G}M.

Corollary A.2.

(cf. [32, Theorem 2.3]) One has weak homotopy equivalences

EG×G𝒫G(M)\textstyle{EG\times_{G}{\mathcal{P}}_{G}(M)}EG×G𝒫G(EG×M)\textstyle{EG\times_{G}{\mathcal{P}}_{G}(EG\times M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p2~\scriptstyle{\widetilde{p_{2}}}\scriptstyle{\simeq}p¯\scriptstyle{\overline{p}}\scriptstyle{\simeq}L(EG×GM).\textstyle{L(EG\times_{G}M).}

The proof of Proposition A.1 we present here differs from that of [32, Theorem 2.3] which assumes GG to be finite.

Proof of Proposition A.1.

With the same notations as above, we have a commutative diagram

𝒫Gm(M)\textstyle{{\mathcal{P}}_{G}^{m}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p¯\scriptstyle{\overline{p}}Ω[m](M/G)\textstyle{\Omega_{[m]}(M/G)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}EG×G𝒫G(M)\textstyle{EG\times_{G}{\mathcal{P}}_{G}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1×Gev0\scriptstyle{1\times_{G}ev_{0}}p¯\scriptstyle{\overline{p}}L(M/G)\textstyle{L(M/G)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ev0\scriptstyle{ev_{0}}EG×GM\textstyle{EG\times_{G}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π~\scriptstyle{\widetilde{\pi}}M/G\textstyle{M/G}

in which two vertical sequences are fibrations and π~\widetilde{\pi} is induced by the projection π:EG×MM\pi:EG\times M\to M in the second factor. Observe that the result [36, Proposition 3.2.2] yields the left-hand side fibration; see also [33, Proposition B.1]. The maps π\pi and π~\widetilde{\pi} give a morphism of fibrations from GEG×MEG×GMG\to EG\times M\to EG\times_{G}M to GMM/GG\to M\to M/G which is the identity map on the fibres. Since EGEG is contractible, it follows that π~\widetilde{\pi} is a weak homotopy equivalence. Therefore, in order to prove the result, it suffices to show that the map p¯:𝒫Gm(M)Ω[m](M/G)\overline{p}:{\mathcal{P}}_{G}^{m}(M)\to\Omega_{[m]}(M/G) is weak homotopy equivalent.

We consider the map p¯:πn(𝒫Gm(M),γ~)πn(Ω[m](M/G),γ)\overline{p}_{*}:\pi_{n}({\mathcal{P}}_{G}^{m}(M),\widetilde{\gamma})\to\pi_{n}(\Omega_{[m]}(M/G),\gamma) induced by the projection p:MM/Gp:M\to M/G for n1n\geq 1. For an element β:SnΩ[m](M/G)\beta:S^{n}\to\Omega_{[m]}(M/G) in πn(Ω[m](M/G),γ)\pi_{n}(\Omega_{[m]}(M/G),\gamma), the adjoint gives rise to a map ad(β):ΣSnMad(\beta):\Sigma S^{n}\to M, where ΣSn\Sigma S^{n} denotes the unreduced suspension of SnS^{n}. Let fβ:Sn×IMf_{\beta}:S^{n}\times I\to M be the composite of ad(β)ad(\beta) and the projection Sn×IΣSnS^{n}\times I\to\Sigma S^{n}. Observe that fβ(,t)=γ(t)f_{\beta}(*,t)=\gamma(t). Since ΣSn\Sigma S^{n} is homeomorphic to Sn+1S^{n+1}, it is immediate that (fβ)(π1(Sn×I))={0}p(π1(M))(f_{\beta})_{*}(\pi_{1}(S^{n}\times I))=\{0\}\subset p_{*}(\pi_{1}(M)). Then the lifting theorem for covering spaces allows us to obtain a lift fβ~\widetilde{f_{\beta}} of fβf_{\beta}. We see that (p¯)(ad(fβ~))=β(\overline{p})_{*}(ad(\widetilde{f_{\beta}}))=\beta. Thus, the map p¯\overline{p}_{*} is surjective.

Let H:Sn×IΩ[m](M/G)H:S^{n}\times I\to\Omega_{[m]}(M/G) be a homotopy from p¯(α0)\overline{p}(\alpha_{0}) to p¯(α1)\overline{p}(\alpha_{1}) based at γ\gamma, where α0\alpha_{0} and α1\alpha_{1} are elements in πn(𝒫Gm(M),γ~)\pi_{n}({\mathcal{P}}_{G}^{m}(M),\widetilde{\gamma}). Then, we have a lift K:Sn×I×IMK:S^{n}\times I\times I\to M of the adjoint ad(H):Sn×I×IΣSn×IM/Gad(H):S^{n}\times I\times I\to\Sigma S^{n}\times I\to M/G. It follows from the uniqueness of the lift that ad(K):Sn×I𝒫Gm(M)ad(K):S^{n}\times I\to{\mathcal{P}}_{G}^{m}(M) is a homotopy from α0\alpha_{0} to α1\alpha_{1} based at γ~\widetilde{\gamma}.

The same argument with the lifting theorem as above allows us to show the bijectivity of the map p¯:π0(𝒫Gm(M))π0(Ω[m](M/G))\overline{p}_{*}:\pi_{0}({\mathcal{P}}_{G}^{m}(M))\to\pi_{0}(\Omega_{[m]}(M/G)). This completes the proof. ∎

Appendix B Diffeological spaces

We begin by reviewing the definitions of a diffeology and a diffeological space. A good reference for the subjects is the book [22]. Additionally, we recall the definitions of Souriau–de Rham complex, the singular de Rham complex for a diffeological space and a morphism of differential graded algebras connecting the two complexes; see [29] for the details.

Definition B.1.

Let XX be a set. A set 𝒟{\mathcal{D}} of functions UXU\to X for each open subset UU in n{\mathbb{R}}^{n} and each nn\in{\mathbb{N}} is a diffeology of XX if the following three conditions hold:

  1. (1)

    Every constant map UXU\to X for all open subset UnU\subset{\mathbb{R}}^{n} is in 𝒟{\mathcal{D}};

  2. (2)

    If UXU\to X is in 𝒟{\mathcal{D}}, then for any smooth map VUV\to U from an open subset VV of m{\mathbb{R}}^{m}, the composite VUXV\to U\to X is also in 𝒟{\mathcal{D}};

  3. (3)

    If U=iUiU=\cup_{i}U_{i} is an open cover and UXU\to X is a map such that each restriction UiXU_{i}\to X is in 𝒟{\mathcal{D}}, then the map UXU\to X is in 𝒟{\mathcal{D}}.

We call an open subset of n{\mathbb{R}}^{n} a domain. A diffeological space (X,𝒟)(X,{\mathcal{D}}) consists of a set XX and a diffeology 𝒟{\mathcal{D}} of XX. An element of a diffeology 𝒟{\mathcal{D}} is called a plot of XX. Let (X,𝒟X)(X,{\mathcal{D}}^{X}) and (Y,𝒟Y)(Y,{\mathcal{D}}^{Y}) be diffeological spaces. A map XYX\to Y is smooth if for any plot p𝒟Xp\in{\mathcal{D}}^{X}, the composite fpf\circ p is in 𝒟Y{\mathcal{D}}^{Y}. All diffeological spaces and smooth maps form a category 𝖣𝗂𝖿𝖿\mathsf{Diff}. It is worthwhile mentioning that the category 𝖣𝗂𝖿𝖿\mathsf{Diff} is complete, cocomplete and cartesian closed. Moreover, the category of manifolds embeds into 𝖣𝗂𝖿𝖿\mathsf{Diff}; see also [12, Section 2].

Let {(Xi,𝒟i)}iI\{(X_{i},{\mathcal{D}}_{i})\}_{i\in I} be a family of diffeological spaces. Then, the product ΠiIXi\Pi_{i\in I}X_{i} has a diffeology 𝒟{\mathcal{D}}, called the product diffeology, defined to be the set of all maps p:UΠiIXip:U\to\Pi_{i\in I}X_{i} from a domain such that πip\pi_{i}\circ p are plots of XiX_{i} for each iIi\in I, where πi:ΠiIXiXi\pi_{i}:\Pi_{i\in I}X_{i}\to X_{i} denotes the canonical projection. Moreover, for diffeological space XX and YY, the set F:=C(X,Y)F:=C^{\infty}(X,Y) of smooth maps from XX to YY is endowed with the functional diffeology 𝒟F{\mathcal{D}}_{F} defined by 𝒟F:={p:UFUis domain andad(p):U×XYis smooth}{\mathcal{D}}_{F}:=\{p:U\to F\mid U\ \text{is domain and}\ ad(p):U\times X\to Y\ \text{is smooth}\}, where ad(p)ad(p) denotes the adjoint to pp.

The category 𝖣𝗂𝖿𝖿\mathsf{Diff} is related to 𝖳𝗈𝗉\mathsf{Top} the category of topological spaces with adjoint functors. Let XX be a topological space. Then the continuous diffeology is defined by the family of continuous maps UXU\to X from domains. This yields a functor C:𝖳𝗈𝗉𝖣𝗂𝖿𝖿C:\mathsf{Top}\to\mathsf{Diff}. For a diffeological space (M,𝒟M)(M,{\mathcal{D}}_{M}), we say that a subset AA of MM is D-open if for every plot p𝒟Mp\in{\mathcal{D}}_{M}, the inverse image p1(A)p^{-1}(A) is an open subset of the domain of pp equipped with the standard topology. The family of D-open subsets of MM defines a topology of MM. Thus, by giving the topology to each diffeological space, we have a functor D:𝖣𝗂𝖿𝖿𝖳𝗈𝗉D:\mathsf{Diff}\to\mathsf{Top} which is the left adjoint to CC; see [42, 12] for more details. The topology for a diffeological space MM is called the D-topology of MM.

For a finite-dimensional manifold MM, the set of all smooth maps from domains to MM define a diffeology 𝒟M\mathcal{D}^{M}, which is called the standard diffeology. Thus, a functor :𝖬𝖿𝖽𝖣𝗂𝖿𝖿\ell:\mathsf{Mfd}\to\mathsf{Diff} is defined by (M)=(M,𝒟M)\ell(M)=(M,\mathcal{D}_{M}), where 𝖬𝖿𝖽\mathsf{Mfd} is the category consisting of finite-dimensional manifolds and smooth maps. Observe that \ell is a fully faithful embedding. Moreover, we see that the forgetful functor UU from 𝖬𝖿𝖽\mathsf{Mfd} to 𝖳𝗈𝗉\mathsf{Top} factors through the category 𝖣𝗂𝖿𝖿\mathsf{Diff}. We summarize the categories and functors mentioned above with the diagram

𝖬𝖿𝖽\textstyle{\mathsf{Mfd}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}:fully faithful\scriptstyle{\ell:\ \text{fully faithful}}U:forgetful functor\scriptstyle{U:\ \text{forgetful functor}}\scriptstyle{\circlearrowright}𝖣𝗂𝖿𝖿\textstyle{\mathsf{Diff}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D\scriptstyle{D}𝖳𝗈𝗉.\textstyle{\mathsf{Top}.\ignorespaces\ignorespaces\ignorespaces\ignorespaces}C\scriptstyle{C}\scriptstyle{\bot}

B.1. The Souriau–de Rham complex, the simplicial de Rham complex and the factor map

We here recall the de Rham complex Ω(X)\Omega^{*}(X) of a diffeological space (X,𝒟X)(X,{\mathcal{D}}^{X}) in the sense of Souriau [43]. For an open set UU of n{\mathbb{R}}^{n}, let 𝒟X(U){\mathcal{D}}^{X}(U) be the set of plots with UU as the domains and

Λ(U)={h:U(i=1ndxi)his smooth}\Lambda^{*}(U)=\{h:U\longrightarrow\wedge^{*}(\oplus_{i=1}^{n}{\mathbb{R}}dx_{i})\mid h\ \text{is smooth}\}

the usual de Rham complex of UU. We can regard 𝒟X(){\mathcal{D}}^{X}(\ ) and Λ()\Lambda^{*}(\ ) as functors from 𝖤𝗎𝖼op\mathsf{Euc}^{\text{op}} to 𝖲𝖾𝗍𝗌\mathsf{Sets} the category of sets. A pp-form is a natural transformation from 𝒟X(){\mathcal{D}}^{X}(\ ) to Λ()\Lambda^{*}(\ ). Then, the de Rham complex Ω(X)\Omega^{*}(X) is defined by the cochain algebra consisting of pp-forms for p0p\geq 0; that is, Ω(X)\Omega^{*}(X) is the direct sum of

Ωp(X):={𝖤𝗎𝖼op𝒟XΛpω𝖲𝖾𝗍𝗌|ωis a natural transformation}\Omega^{p}(X):=\left\{\>{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 14.35835pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-14.35835pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathsf{Euc}^{\text{op}}\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{{}{{}{}{}{{{{{}}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}}}}{}{{{}{}{}{{{{{}}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}}}}}}{{}{}{}\lx@xy@spline@}{}}}}\ignorespaces{}{}{}{{}{}}\ignorespaces\ignorespaces{\hbox{\kern 23.15001pt\raise 13.75pt\hbox{\hbox{\kern 3.0pt\raise-3.075pt\hbox{$\textstyle{\scriptstyle{\mathcal{D}}^{X}}$}}}}}\ignorespaces\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{\hbox{\kern 49.35965pt\raise 5.48257pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{{}{{}{}{}{{{{{}}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}}}}{}{{{}{}{}{{{{{}}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}}}}}}{{}{}{}\lx@xy@spline@}{}}}}\ignorespaces{}{}{}{{}{}}\ignorespaces\ignorespaces{\hbox{\kern 24.20071pt\raise-13.75pt\hbox{\hbox{\kern 3.0pt\raise-2.82222pt\hbox{$\textstyle{\scriptstyle\Lambda^{p}}$}}}}}\ignorespaces\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{\hbox{\kern 55.4707pt\raise-3.00137pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 30.63751pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\hbox{{\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{\kern 0.0pt\hbox{\ignorespaces\hbox{\kern 0.0pt\raise 5.0pt\hbox{\hbox{\line@@}}}}}}\hbox{\kern-1.0pt\raise 0.0pt\hbox{\kern 0.0pt\hbox{\ignorespaces\hbox{\kern 0.0pt\raise 5.0pt\hbox{\hbox{\line@@}}}}}}}}{\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{\kern 0.0pt\hbox{\ignorespaces\hbox{\kern 0.0pt\raise 2.5pt\hbox{\hbox{\line@@}}}}}}\hbox{\kern-1.0pt\raise 0.0pt\hbox{\kern 0.0pt\hbox{\ignorespaces\hbox{\kern 0.0pt\raise 2.5pt\hbox{\hbox{\line@@}}}}}}}}{\hbox{\kern 0.0pt\hbox{\ignorespaces\hbox{\kern 0.0pt\raise-5.0pt\hbox{\lx@xy@tip{1.5}\lx@xy@tip{-1.5}}}}}}}}}}}}}{}\ignorespaces\ignorespaces{\hbox{\kern 28.61368pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-1.50694pt\hbox{$\textstyle{\scriptstyle\hskip 5.69046pt\omega}$}}}}}\ignorespaces{\hbox{\kern 49.35835pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathsf{Sets}}$}}}}}}}\ignorespaces}}}}\ignorespaces}\;\middle|\;{\omega\ \text{is a natural transformation}}\>\right\}

with the cochain algebra structure induced by that of Λ(U)\Lambda^{*}(U) pointwisely. In what follows, we may write ωp\omega_{p} for ωU(p)\omega_{U}(p) for a plot p:UXp:U\to X. The de Rham complex defined above is a generalization of the usual de Rham complex of a manifold.

Remark B.2.

Let MM be a manifold and (M)\wedge^{*}(M) the usual de Rham complex of MM. We recall the tautological map θ:(M)Ω(M)\theta:\wedge^{*}(M)\to\Omega^{*}(M) defined by θ(ω)={pω}p𝒟M,\theta(\omega)=\{p^{*}\omega\}_{p\in{{\mathcal{D}}^{M}}}, where 𝒟M{\mathcal{D}}^{M} denotes the standard diffeology of MM. Then, it follows that θ\theta is an isomorphism of cochain algebras; see [19, Section 2].

Let 𝔸n:={(x0,,xn)n+1i=0nxi=1}{\mathbb{A}}^{n}:=\{(x_{0},...,x_{n})\in{\mathbb{R}}^{n+1}\mid\sum_{i=0}^{n}x_{i}=1\} be the affine space which is diffeomorphic to the Euclidean space n{\mathbb{R}}^{n}. Let (ADR)(A_{DR}^{*})_{\bullet} be the simplicial cochain algebra defined by (ADR)n:=(𝔸n)(A^{*}_{DR})_{n}:=\wedge^{*}({\mathbb{A}}^{n}) for each n0n\geq 0. Then, for a simplicial set KK, we define the de Rham complex ADR(K)A^{*}_{DR}(K) by the set of simplicial maps from KK to (ADR)(A_{DR}^{*})_{\bullet} endowed with a differential graded algebra structure induced by that of each (ADR)n(A^{*}_{DR})_{n}; see [29, Section 2] for more details. For a general simplicial cochain algebra AA_{\bullet}, we denote by A(K)A(K) the cochain algebra

𝖲𝖾𝗍𝗌Δ𝗈𝗉(K,A):={ΔopKAω𝖲𝖾𝗍𝗌|ωis a natural transformation}\mathsf{Sets^{\Delta^{op}}}(K,A_{\bullet}):=\left\{\>{\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 10.12224pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-10.12224pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathsf{\Delta}^{\text{op}}\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{{}{{}{}{}{{{{{}}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}}}}{}{{{}{}{}{{{{{}}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}}}}}}{{}{}{}\lx@xy@spline@}{}}}}\ignorespaces{}{}{}{{}{}}\ignorespaces\ignorespaces{\hbox{\kern 22.29655pt\raise 13.75pt\hbox{\hbox{\kern 3.0pt\raise-2.39166pt\hbox{$\textstyle{\scriptstyle K}$}}}}}\ignorespaces\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{\hbox{\kern 45.1257pt\raise 5.78339pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{{}{{}{}{}{{{{{}}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}}}}{}{{{}{}{}{{{{{}}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}}}}}}{{}{}{}\lx@xy@spline@}{}}}}\ignorespaces{}{}{}{{}{}}\ignorespaces\ignorespaces{\hbox{\kern 21.89447pt\raise-13.75pt\hbox{\hbox{\kern 3.0pt\raise-1.94722pt\hbox{$\textstyle{\scriptstyle A_{\bullet}}$}}}}}\ignorespaces\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{\hbox{\kern 51.63583pt\raise-3.00137pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 28.51945pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\hbox{{\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{\kern 0.0pt\hbox{\ignorespaces\hbox{\kern 0.0pt\raise 5.0pt\hbox{\hbox{\line@@}}}}}}\hbox{\kern-1.0pt\raise 0.0pt\hbox{\kern 0.0pt\hbox{\ignorespaces\hbox{\kern 0.0pt\raise 5.0pt\hbox{\hbox{\line@@}}}}}}}}{\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{\kern 0.0pt\hbox{\ignorespaces\hbox{\kern 0.0pt\raise 2.5pt\hbox{\hbox{\line@@}}}}}}\hbox{\kern-1.0pt\raise 0.0pt\hbox{\kern 0.0pt\hbox{\ignorespaces\hbox{\kern 0.0pt\raise 2.5pt\hbox{\hbox{\line@@}}}}}}}}{\hbox{\kern 0.0pt\hbox{\ignorespaces\hbox{\kern 0.0pt\raise-5.0pt\hbox{\lx@xy@tip{1.5}\lx@xy@tip{-1.5}}}}}}}}}}}}}{}\ignorespaces\ignorespaces{\hbox{\kern 26.49562pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-1.50694pt\hbox{$\textstyle{\scriptstyle\hskip 5.69046pt\omega}$}}}}}\ignorespaces{\hbox{\kern 45.12224pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathsf{Sets}}$}}}}}}}\ignorespaces}}}}\ignorespaces}\;\middle|\;{\omega\ \text{is a natural transformation}}\>\right\}

whose cochain algebra structure is induced by that of AA_{\bullet}.

For a diffeological space XX, let SD(X)S^{D}_{\bullet}(X) be the simplicial set defined by SD(X)aff:={{σ:𝔸nXσis a C-map}}n0.S^{D}_{\bullet}(X)_{\text{aff}}:=\{\{\sigma:{\mathbb{A}}^{n}\to X\mid\sigma\ \text{is a $C^{\infty}$-map}\}\}_{n\geq 0}. Thus, we have the singular de Rham complex ADR(SD(X)aff)A_{DR}^{*}(S^{D}_{\text{\tiny{$\bullet$}}}(X)_{\text{aff}}) for a diffeological space XX. This is regarded as a diffeological variant of Sullivan’s simplicial polynomial form on a topological space. In fact, for a space XX, the polynomial-de Rham complex APL(X)A_{PL}^{*}(X) is defined by APL(X):=APL(Sing(X))A_{PL}^{*}(X):=A_{PL}(\text{Sing}_{\bullet}(X)) with the simplicial differential graded algebra (APL)(A_{PL}^{*})_{\bullet} of polynomial forms; see [8] and [15, II 10 (a), (b) and (c)].

Remark B.3.

By [23, Lemma 3.1], we have that the inclusion i:Δstn𝔸ni:\Delta^{n}_{st}\to{\mathbb{A}}^{n} is smooth; see Remark 3.2 for the notation. Moreover, the consideration at the end of [29, Section 5] yields that the chain map induced by i:SD(X)affSD(X)i^{*}:S^{D}_{\bullet}(X)_{\text{aff}}\to S^{D}_{\bullet}(X) is a quasi-isomorphism for every diffeological space XX; see also [29, Table 1, page 959]. The de Rham theorem holds for the singular de Rham cohomology; see [29, Theorem 2.4 and Corollary 2.5]. Therefore, the results on the singular de Rham cohomology H(ADR(SD(X)aff))H^{*}(A_{DR}^{*}(S^{D}_{\text{\tiny{$\bullet$}}}(X)_{\text{aff}})) in [29] hold for the cohomology H(ADR(SD(X)))H^{*}(A_{DR}^{*}(S^{D}_{\text{\tiny{$\bullet$}}}(X))).

We recall the factor map α:Ω(X)ADR(SD(X)aff)\alpha:\Omega^{*}(X)\to A_{DR}^{*}(S^{D}_{\bullet}(X)_{\text{aff}}) of cochain algebras defined by

α(ω)(σ)=σ(ω).\alpha(\omega)(\sigma)=\sigma^{*}(\omega).

We refer the reader to the result [29, Theorem 2.4] for an important role of the factor map in the de Rham theorem for diffeological spaces. In particular, we have

Proposition B.4.

([29, Theorem 2.4]) Suppose that XX is a manifold, more generally, a stratifold in the sense of Kreck [25]. Then the factor map α\alpha for XX is a quasi-isomorphism.

B.2. Chen’s iterated integral map in diffeology

Let NN be a diffeological space and ρ:I\rho:{\mathbb{R}}\to I a cut-off function with ρ(0)=0\rho(0)=0 and ρ(1)=1\rho(1)=1. Then, we call a pp-form uu on the diffeological space I×NI\times N an Ωp(N)\Omega^{p}(N)-valued function on II if for any plot ψ:UN\psi:U\to N of NN, the pp-form uρ×ψu_{\rho\times\psi} on ×U{\mathbb{R}}\times U is of type

ai1ip(t,ξ)dξi1dξip,\sum a_{i_{1}\cdots i_{p}}(t,\xi)d\xi_{i_{1}}\wedge\cdots\wedge d\xi_{i_{p}},

where (ξ1,,ξn)(\xi_{1},...,\xi_{n}) denotes the coordinates of UU we fix. For such an Ωp(N)\Omega^{p}(N)-valued function uu on II, an integration 01u𝑑tΩp(N)\int_{0}^{1}u\ dt\in\Omega^{p}(N) is defined by

(01u𝑑t)ψ=(01ai1ip(t,ξ)𝑑t)dξi1dξip.(\int_{0}^{1}u\ dt)_{\psi}=\sum(\int_{0}^{1}a_{i_{1}\cdots i_{p}}(t,\xi)\ dt)d\xi_{i_{1}}\wedge\cdots\wedge d\xi_{i_{p}}.

Each pp-form uu has the form u=dt((/t)u)+u′′u=dt\wedge((\partial/\partial t)\rfloor u)+u^{\prime\prime}, where (/t)u(\partial/\partial t)\rfloor u and u′′u^{\prime\prime} are an Ωp1(N)\Omega^{p-1}(N)-valued function and an Ωp(N)\Omega^{p}(N)-valued function on II, respectively. Let F:I×NINIF:I\times N^{I}\to N^{I} be the homotopy defined by F(t,γ)(s)=γ(ts)F(t,\gamma)(s)=\gamma(ts). The Poincaré operator F:Ω(NI)Ω(NI)\int_{F}:\Omega(N^{I})\to\Omega(N^{I}) associated with the homotopy FF is defined by Fv=01((/t)Fv)dt\int_{F}v=\int_{0}^{1}((\partial/\partial t)\rfloor F^{*}v)dt. Moreover, for forms ω1\omega_{1}, …, ωr\omega_{r} on NN, we define the iterated integral ω1ωr\int\omega_{1}\cdots\omega_{r}, which is an element in Ω(NI)\Omega^{*}(N^{I}), by ω1=Fε1ω1\int\omega_{1}=\int_{F}\varepsilon_{1}^{*}\omega_{1} and

ω1ωr=F{J(ω1ωr1)ε1ωr},\int\omega_{1}\cdots\omega_{r}=\int_{F}\{J(\int\omega_{1}\cdots\omega_{r-1})\wedge\varepsilon_{1}^{*}\omega_{r}\},

where εi\varepsilon_{i} denotes the evaluation map at ii, Ju=(1)deguuJu=(-1)^{\deg u}u and ω1ωr=1\int\omega_{1}\cdots\omega_{r}=1 if r=0r=0; see [11, Definition 1.5.1]. We observe that the Poincaré operator is of degree 1-1 and then ω1ωr\int\omega_{1}\cdots\omega_{r} is of degree 1ir(degωi1)\sum_{1\leq i\leq r}(\deg\omega_{i}-1).

With a decomposition of the form Ω~1(N)dΩ0(N)\widetilde{\Omega}^{1}(N)\oplus d\Omega^{0}(N), we have a cochain subalgebra Ω~(N)\widetilde{\Omega}(N) of Ω(N)\Omega(N) which satisfies the condition that Ω~p(N)=Ω(N)\widetilde{\Omega}^{p}(N)=\Omega(N) for p>1p>1 and Ω~0(N)=\widetilde{\Omega}^{0}(N)={\mathbb{R}}. The cochain algebra Ω~(N)\widetilde{\Omega}(N) gives rise to the normalized bar complex B(Ω(N),Ω~(N),Ω(N))B(\Omega(N),\widetilde{\Omega}(N),\Omega(N)); see [11, §4.1]. Consider the pullback diagram

(B.5)

of ε0×ε1:NIN×N\varepsilon_{0}\times\varepsilon_{1}:N^{I}\to N\times N along a smooth map f:MN×Nf:M\to N\times N. We write B¯(Ω~(N))\overline{B}(\widetilde{\Omega}(N)) for B(,Ω~(N),)B({\mathbb{R}},\widetilde{\Omega}(N),{\mathbb{R}}). Then we have a map

𝖨𝗍:Ω(M)Ω(N)Ω(N)B(Ω(N),Ω~(N),Ω(N))Ω(M)fB¯(Ω~(N))Ω(Ef)\mathsf{It}:\Omega(M)\otimes_{\Omega(N)\otimes\Omega(N)}B(\Omega(N),\widetilde{\Omega}(N),\Omega(N))\cong\Omega(M)\otimes_{f}\overline{B}(\widetilde{\Omega}(N))\to\Omega(E_{f})

defined by 𝖨𝗍(v[ω1||ωr])=pfvf~ω1ωr\mathsf{It}(v\otimes[\omega_{1}|\cdots|\omega_{r}])=p_{f}^{*}v\wedge\widetilde{f}^{*}\int\omega_{1}\cdots\omega_{r}. Observe that the domain of 𝖨𝗍\mathsf{It} gives rise to the differential on Ω(M)fB¯(Ω~(N))\Omega(M)\otimes_{f}\overline{B}(\widetilde{\Omega}(N)) by definition. Since ρ(0)=0\rho(0)=0 and ρ(1)=1\rho(1)=1 for the cut-off function ρ\rho which we use when defining the Ωp(N)\Omega^{p}(N)-valued function on II, it follows that the result [11, Lemma 1.4.1] remains valid. Then the formula of iterated integrals with respect to the differential in [11, Proposition 1.5.2] implies that 𝖨𝗍\mathsf{It} is a well-defined morphism of differential graded Ω(M)\Omega^{*}(M)-modules.

The following theorem enables us to compute the singular de Rham cohomology of a pullback diffeological space with a bar complex via Chen’s iterated integral map and the factor map mentioned above; see Remark B.3.

Theorem B.5.

([29, Theorem 5.2]) Suppose that, in the pullback diagram (B.5), the diffeological space NN is simply connected and ff is an induction; that is, pp is a plot of MM if and only if fpf\circ p is a plot of N×NN\times N. Assume further that the factor maps for NN and MM are quasi-isomorphisms and each vector space Hi(SD(N))H^{i}(S^{D}_{\bullet}(N)) is of finite dimension. Then the composite α𝖨𝗍:Ω(M)fB¯(Ω~(N))Ω(Ef)ADR(SD(Ef))\alpha\circ\mathsf{It}:\Omega^{*}(M)\otimes_{f}\overline{B}(\widetilde{\Omega}(N))\to\Omega(E_{f})\to A^{*}_{DR}(S^{D}_{\bullet}(E_{f})) is a quasi-isomorphism of Ω(M)\Omega^{*}(M)-modules.

Appendix C C(S1,M)C^{\infty}(S^{1},M) versus LML^{\infty}M

Let MM be a diffeological space. We have two free loop spaces of MM. One of them is the diffeological space C(S1,M)C^{\infty}(S^{1},M) of smooth maps from the circle S1S^{1} to MM with the functional diffeology. Another one is the diffeological space LML^{\infty}M which fits in the pullback diagram

(C.5)

in the category 𝖣𝗂𝖿𝖿\mathsf{Diff}, where I:=[0,1]I:=[0,1] is the diffeological subspace of {\mathbb{R}} the Euclidean space and Δ\Delta denotes the diagonal map. We observe that LML^{\infty}M is diffeomorphic to the diffeological subspace of MIM^{I} consisting of smooth maps γ\gamma with γ(0)=γ(1)\gamma(0)=\gamma(1).

Let q:S1q:{\mathbb{R}}\to S^{1} be the smooth map defined by q(t)=e2π1tq(t)=e^{2\pi\sqrt{-1}t}. Then, the restriction q:IS1q:I\to S^{1} is smooth. In the category 𝖳𝗈𝗉\mathsf{Top}, the continuous map q:IS1q:I\to S^{1} is regarded as a quotient map. The fact enables us to conclude that qq induces a weak homotopy equivalence q:C0(S1,M)wLMq^{*}:C^{0}(S^{1},M)\stackrel{{\scriptstyle\simeq_{w}}}{{\longrightarrow}}LM in 𝖳𝗈𝗉\mathsf{Top}. We obtain a diffeological version of the equivalence.

In order to define the weak homotopy equivalence between diffeological spaces, we first recall the smooth homotopy groups of a pointed diffeological space.

We define an equivalence relation on a diffeological space ZZ by zwz\simeq w if there exists a smooth path l:IZl:I\to Z such that l(0)=zl(0)=z and l(1)=wl(1)=w. Let SnS^{n} be the nn-sphere endowed with sub-diffeology of the manifold n+1{\mathbb{R}}^{n+1}. We use the north pole \ast as a base point of SnS^{n}. For a pointed diffeological space (X,x0)(X,x_{0}), let C((Sn,),(X,x0))C^{\infty}((S^{n},\ast),(X,x_{0})) be the diffeological subspace of the mapping space C(Sn,X)C^{\infty}(S^{n},X) consisting of smooth maps that preserve base points. Then, given a positive integer nn, the nnth smooth homotopy group πnD(X,x0)\pi_{n}^{D}(X,x_{0}) is defined by the set C((Sn,x0),(X.x0))/C^{\infty}((S^{n},x_{0}),(X.x_{0}))/\!\simeq. Moreover, we define π0D(X)\pi_{0}^{D}(X) by X/X/\!\simeq. We observe that, while the original smooth homotopy group πn(X,x0)\pi_{n}(X,x_{0}) of a pointed diffeological space (X,x0)(X,x_{0}) due to Iglesias-Zemmour [22] is defined by using an iterated loop space of XX, there is a natural bijection between the smooth homotopy set πnD(X,x0)\pi_{n}^{D}(X,x_{0}) and πn(X,x0)\pi_{n}(X,x_{0}); see [13, Theorem 3.2] for more details.

By definition, we call a smooth map f:XYf:X\to Y in 𝖣𝗂𝖿𝖿\mathsf{Diff} a weak homotopy equivalence if the induced maps f:π0D(X)π0D(Y)f_{*}:\pi_{0}^{D}(X)\to\pi_{0}^{D}(Y) and f:πnD(X,x0)πnD(Y,f(x0))f_{*}:\pi_{n}^{D}(X,x_{0})\to\pi_{n}^{D}(Y,f(x_{0})) for each nn and x0Xx_{0}\in X are bijective.

Lemma C.1.

The smooth map q:IS1q:I\to S^{1} mentioned above gives rise to a weak homotopy equivalence q:C(S1,M)wLMq^{*}:C^{\infty}(S^{1},M)\stackrel{{\scriptstyle\simeq_{w}}}{{\longrightarrow}}L^{\infty}M.

Proof.

We prove that the maps (q):π0(C(S1,M))π0(LM)(q^{*})_{*}:\pi_{0}(C^{\infty}(S^{1},M))\to\pi_{0}(L^{\infty}M) and (q):πn(C(S1,M),γ0)πn(LM,γ0q)(q^{*})_{*}:\pi_{n}(C^{\infty}(S^{1},M),\gamma_{0})\to\pi_{n}(L^{\infty}M,\gamma_{0}\circ q) induced by qq are bijective for n1n\geq 1 and each smooth loop γ0:S1M\gamma_{0}:S^{1}\to M. To this end, we first show that the adjoint

(1×q):π0({η:N×S1Mη|×S1=γ0})\displaystyle(1\times q)^{*}:\pi_{0}(\{\eta:N\times S^{1}\to M\mid\eta|_{\ast\times S^{1}}=\gamma_{0}\})
π0({η:N×IMη|×I=γ0q,η|N×{0}=η|N×{1}})\displaystyle\longrightarrow\pi_{0}(\{\eta^{\prime}:N\times I\to M\mid\eta^{\prime}|_{\ast\times I}=\gamma_{0}\circ q,\eta^{\prime}|_{N\times\{0\}}=\eta^{\prime}|_{N\times\{1\}}\})

is bijective, where N=SnN=S^{n} for n1n\geq 1. In what follows, we use the same notation for a homotopy class and its representative.

We consider a function ρ~:(12ε,1+2ε)I\widetilde{\rho}:(-1-2\varepsilon,1+2\varepsilon)\to I with ρ~(t)=0\widetilde{\rho}(t)=0 for t(12ε,1+2ε)[0,2ε)t\in(-1-2\varepsilon,-1+2\varepsilon)\cup[0,2\varepsilon) and ρ~(t)=1\widetilde{\rho}(t)=1 for t(2ε,0)(12ε,1+2ε)t\in(-2\varepsilon,0)\cup(1-2\varepsilon,1+2\varepsilon) for a sufficiently small positive number ε\varepsilon. Let (U,φU)(U,\varphi_{U}) and (V,φV)(V,\varphi_{V}) be local coordinates of S1S^{1} which satisfy the condition that φU1=q\varphi_{U}^{-1}=q, U:=φU(U)=(14ε,14+ε)U^{\prime}:=\varphi_{U}(U)=(-\frac{1}{4}-\varepsilon,\frac{1}{4}+\varepsilon), φV1=q\varphi_{V}^{-1}=q and V:=φV(V)=(14ε,34+ε)V^{\prime}:=\varphi_{V}(V)=(\frac{1}{4}-\varepsilon,\frac{3}{4}+\varepsilon).

We define a smooth path γ0~:S1M\widetilde{\gamma_{0}}:S^{1}\to M by γ0~|U:=γ0qρ~φU\widetilde{\gamma_{0}}|_{U}:=\gamma_{0}\circ q\circ\widetilde{\rho}\circ\varphi_{U} and γ0~|V:=γ0qρ~φV\widetilde{\gamma_{0}}|_{V}:=\gamma_{0}\circ q\circ\widetilde{\rho}\circ\varphi_{V}. By using the map ρ~\widetilde{\rho}, we have a smooth map ρ¯:(12ε,1+2ε)(12ε,1+2ε)\overline{\rho}:(-1-2\varepsilon,1+2\varepsilon)\to(-1-2\varepsilon,1+2\varepsilon) defined by ρ¯(t)=ρ~(t)1\overline{\rho}(t)=\widetilde{\rho}(t)-1 for t(12ε,0)t\in(-1-2\varepsilon,0) and ρ¯(t)=ρ~(t)\overline{\rho}(t)=\widetilde{\rho}(t) for t[0,1+2ε)t\in[0,1+2\varepsilon). Since the map ρ¯\overline{\rho} is smooth homotopic to the identity map on (12ε,1+2ε)(-1-2\varepsilon,1+2\varepsilon) and qρ~=qρ¯q\circ\widetilde{\rho}=q\circ\overline{\rho}, it follows that γ0\gamma_{0} is smooth homotopic to γ0~\widetilde{\gamma_{0}}. Therefore, in order to prove the bijectivity of (1×q)(1\times q)_{*}, it suffices to show the bijectivity for γ0~\widetilde{\gamma_{0}} instead of γ0\gamma_{0}. In what follows, we write γ0\gamma_{0} for γ0~\widetilde{\gamma_{0}}. Let AA and BB be the domain and codomain of the map (1×q)(1\times q)_{*}, respectively.

We show the surjectivity of (1×q)(1\times q)_{*}. Let η\eta^{\prime} be an element in BB. We consider a function ρ:(1ε,1+ε)I\rho:(-1-\varepsilon,1+\varepsilon)\to I with ρ(t)=0\rho(t)=0 for t[1ε,1+ε)[0,ε)t\in[-1-\varepsilon,-1+\varepsilon)\cup[0,\varepsilon), ρ(t)=1\rho(t)=1 for t(ε,0)[1ε,1+ε]t\in(-\varepsilon,0)\cup[1-\varepsilon,1+\varepsilon], ρ(t)=t\rho(t)=t for t[2ε,12ε]t\in[2\varepsilon,1-2\varepsilon] and ρ(t)=t+1\rho(t)=t+1 for t[1+2ε,2ε]t\in[-1+2\varepsilon,-2\varepsilon]. Observe that ρ\rho is smooth except for the point 0. We define map η:N×S1M\eta:N\times S^{1}\to M by η|N×U=η(1×ρ)(1×φU)\eta|_{N\times U}=\eta^{\prime}\circ(1\times\rho)\circ(1\times\varphi_{U}) and η|N×V=η(1×ρ)(1×φV)\eta|_{N\times V}=\eta^{\prime}\circ(1\times\rho)\circ(1\times\varphi_{V}). Since η\eta is constant in a neighborhood of zero, it follows that the map is a well-defined smooth map. Moreover, we see that η|×U=η(1×ρ)(1×φU)=γ0qρφU=γ0|U\eta|_{\ast\times U}=\eta^{\prime}\circ(1\times\rho)\circ(1\times\varphi_{U})=\gamma_{0}\circ q\circ\rho\circ\varphi_{U}=\gamma_{0}|_{U}. The last equality follows from the fact that ρ~ρ=ρ~\widetilde{\rho}\circ\rho=\widetilde{\rho} on UU^{\prime}. The same argument as above yields that η|×V=γ0|V\eta|_{\ast\times V}=\gamma_{0}|_{V}. Thus, we have η|×S1=γ0\eta|_{\ast\times S^{1}}=\gamma_{0}.

Moreover, we see that η(1×q)=η(1×ρ)\eta\circ(1\times q)=\eta^{\prime}\circ(1\times\rho). Extending ρ|[0,1]\rho|_{[0,1]}, we define a smooth map ρ:\rho^{\prime}:{\mathbb{R}}\to{\mathbb{R}} so that ρ|[1,)=1\rho^{\prime}|_{[1,\infty)}=1 and ρ|(,0]=0\rho^{\prime}|_{(-\infty,0]}=0. Define a smooth ρ^:×\widehat{\rho}:{\mathbb{R}}\times{\mathbb{R}}\to{\mathbb{R}} by ρ^(t,s)=(1s)ρ(t)+st\widehat{\rho}(t,s)=(1-s)\rho^{\prime}(t)+st. Then the restriction ρ^:I×II\widehat{\rho}:I\times I\to I gives rise to a smooth homotopy between ρ\rho and the identity on II. It follows that (1×q)(η)=η(1×ρ)η(1\times q)^{*}(\eta)=\eta^{\prime}\circ(1\times\rho)\sim\eta^{\prime} with the smooth homotopy η(1×ρ^)\eta^{\prime}\circ(1\times\widehat{\rho}). We conclude that the map (1×q)(1\times q)^{*} is surjective.

Let η0\eta_{0} and η1\eta_{1} be elements in AA with (1×q)(η0)=(1×q)(η1)(1\times q)^{*}(\eta_{0})=(1\times q)^{*}(\eta_{1}). Then, there exists a smooth path from η0(1×q)\eta_{0}\circ(1\times q) to η1(1×q)\eta_{1}\circ(1\times q). Let H:(N×I)×IMH:(N\times I)\times I\to M be the smooth homotopy which is the adjoint to the path. We define a map H~:(N×S1)×IM\widetilde{H}:(N\times S^{1})\times I\to M by H~|N×W×I=H(1×ρ×1)(1×φW×1)\widetilde{H}|_{N\times W\times I}=H\circ(1\times\rho\times 1)\circ(1\times\varphi_{W}\times 1) for W=UW=U and VV, respectively. We see that H~\widetilde{H} is a well-defined smooth map that satisfies the condition that H~(,t,s)=γ0(t)\widetilde{H}(\ast,t,s)=\gamma_{0}(t). We define a smooth map ρ′′:(1ε,1+ε)\rho^{\prime\prime}:(-1-\varepsilon,1+\varepsilon)\to{\mathbb{R}} by ρ′′|(1ε,0]=ρ|(1ε,0]1\rho^{\prime\prime}|_{(-1-\varepsilon,0]}=\rho|_{(-1-\varepsilon,0]}-1 and ρ′′|[0,1+ε)=ρ|[0,1+ε)\rho^{\prime\prime}|_{[0,1+\varepsilon)}=\rho|_{[0,1+\varepsilon)}. Then, it follows that qρ=qρ′′q\circ\rho=q\circ\rho^{\prime\prime}. Moreover, we define a smooth homotopy ρs:(1ε,1+ε)×II\rho_{s}:(-1-\varepsilon,1+\varepsilon)\times I\to I by ρs(t)=(1s)ρ′′(t)+st\rho_{s}(t)=(1-s)\rho^{\prime\prime}(t)+st. Then the map ρs\rho_{s} gives rise to smooth homotopies H~iηi\widetilde{H}_{i}\sim\eta_{i} for i=0i=0 and 11. In fact, for example, we see that

H~0|N×U\displaystyle\widetilde{H}_{0}|_{N\times U}\!\!\! =\displaystyle= H0(1×ρ)(1×φU)=η0(1×q)(1×ρ)(1×φU)\displaystyle\!\!\!H_{0}\circ(1\times\rho)\circ(1\times\varphi_{U})=\eta_{0}\circ(1\times q)\circ(1\times\rho)\circ(1\times\varphi_{U})
=\displaystyle= η0(1×π)(1×ρ)(1×φU)=η0(1×q)(1×ρ0)(1×φU)\displaystyle\!\!\!\eta_{0}\circ(1\times\pi)\circ(1\times\rho)\circ(1\times\varphi_{U})=\eta_{0}\circ(1\times q)\circ(1\times\rho_{0})\circ(1\times\varphi_{U})
\displaystyle\sim η0(1×q)(1×ρ1)(1×φU)=η0|N×U,\displaystyle\!\!\!\eta_{0}\circ(1\times q)\circ(1\times\rho_{1})\circ(1\times\varphi_{U})=\eta_{0}|_{N\times U},

where π:S1\pi:{\mathbb{R}}\to S^{1} denotes the natural smooth extension of qq. It turns out that η1=η0\eta_{1}=\eta_{0} in AA. We observe that the homotopy induced by ρs\rho_{s} fixes the path γ0\gamma_{0}.

The same argument as above enables us to prove that q:π0(C(S1,M))π0(LM)q_{*}:\pi_{0}(C^{\infty}(S^{1},M))\to\pi_{0}(L^{\infty}M) is a bijection. We have the result. ∎

References

  • [1] A. Adem, J. Leida, and Y. Ruan, Orbifolds and Stringy Topology, Cambridge Tracts in Mathematics 171, Cambridge University Press, Cambridge 2007.
  • [2] L. Abrams and C. Weibel, Cotensor products of modules, Trans. Am. Math. Soc. 354 (2002), 2173–2185.
  • [3] M. Arkowitz, Formal differential graded algebras and homomorphisms, J. Pure Appl. Algebra 51 (1988), 35-52.
  • [4] K. Behrend, Cohomology of stacks, In: Arbarello, Ellingsrud, Goettsche (Eds.), Intersection Theory and Moduli, ICTP Lecture Notes Series, vol. 19, (2004), 249–294.
  • [5] K. Behrend and P. Xu, Differentiable stacks and gerbes, J. Symplectic Geom. 9 (2011), 285–341.
  • [6] K. Behrend, G. Ginot, B. Noohi and P. Xu, String topology for stacks, Astérisque 343 (2012).
  • [7] D.J. Benson, Representations and cohomology. II: Cohomology of groups and modules, Cambridge Studies in Advanced Mathematics, 31, Cambridge University Press, 1991.
  • [8] A.K. Bousfield and V.K.A.M. Gugenheim, On PL de Rham theory and rational homotopy type, Memoirs of AMS 179(1976).
  • [9] G.E. Bredon, Sheaf theory, 2nd ed. Graduate Texts in Mathematics, 170, New York, Springer, 1997.
  • [10] D. Carchedi, On the homotopy type of higher orbifolds and Haefliger classifying spaces, Advances in Mathematics 294 (2016), 756–818.
  • [11] K.-T. Chen, Iterated path integrals, Bull. Amer. Math. Soc. 83 (1977), 831–879.
  • [12] J.D. Christensen, G. Sinnamon and E. Wu, The DD-topology for diffeological spaces, Pacific J. Math. 272 (2014), 87–110.
  • [13] J.D. Christensen and E. Wu, The homotopy theory of diffeological spaces, New York J. Math. 20 (2014), 1269–1303.
  • [14] J.L. Dupont, Curvature and characteristic classes, Lecture Notes in Math., 640, Springer, 1978.
  • [15] Y. Félix, S. Halperin and J.-C. Thomas, Rational Homotopy Theory, Graduate Texts in Mathematics 205, Springer-Verlag, 2000.
  • [16] S.I. Gelfand and Yu.I. Manin, Methods of Homological Algebra, Springer, 2002.
  • [17] V.K.A.M. Gugenheim and J.P. May, On the theory and applications of differential torsion products, Mem. Am. Math. Soc. 142 (1974).
  • [18] P.G. Goerss and J.F. Jardine, Simplicial homotopy theory, Progress in Mathematics 174. Birkhäuser, 1999.
  • [19] G. Hector, E. Macías-Virgós and E. Sanmartín-Carbón, De Rham cohomology of diffeological spaces and foliations, Indagationes math. 21 (2011), 212–220.
  • [20] J. Heinloth, Notes on differentiable stacks, Mathematisches Institut, Georg-August-Universität Göttingen Seminars: Winter Term 2004–05, (2005), 1–32.
  • [21] S. Halperin, Lectures on minimal models, Mém. Soc. Math. France (N.S.) No. 9–10 (1983).
  • [22] P. Iglesias-Zemmour, Diffeology, Mathematical Surveys and Monographs, 185, AMS, Providence, 2012.
  • [23] H. Kihara, Model category of diffeological spaces, J. Homotopy Relat. Struct. 14 (2019), 51–90.
  • [24] H. Kihara, Smooth homotopy of infinite-dimensional CC^{\infty}-manifolds, Memoirs of the American Mathematical Society 1436. Providence, RI: American Mathematical Society (AMS), 2023.
  • [25] M. Kreck, Differential Algebraic Topology, From Stratifolds to Exotic Spheres, Graduate Studies in Math.110, AMS, 2010.
  • [26] A. Kriegl and P. W. Michor, The convenient setting of global analysis, Mathematical Surveys and Monographs, 53, Providence, AMS, 1997.
  • [27] K. Kuribayashi, On the mod pp cohomology of the spaces of free loops on the Grassmann and Stiefel manifolds, J. Math. Soc. Japan 43 (1991), 331–346.
  • [28] K. Kuribayashi, On the real cohomology of spaces of free loops on manifolds, Fund. Math. 150 (1996), 173–188.
  • [29] K. Kuribayashi, Simplicial cochain algebras for diffeological spaces, Indag. Math., New Ser. 31 (2020), 934-967.
  • [30] K. Kuribayashi, M. Mimura and T. Nishimoto, Twisted tensor products related to the cohomology of the classifying spaces of loop groups, Mem. Am. Math. Soc. 849 (2006).
  • [31] K. Kuribayashi and T. Yamaguchi, The cohomology of certain free loop spaces, Fundam. Math. 154 (1997), 57–73.
  • [32] E. Lupercio, B. Uribe and M.A. Xicotencatl, Orbifold string topology, Geom. Topol. 12 (2008), 2203–2247.
  • [33] K. Matsuo, The Borel cohomology of the loop space of a homogeneous space, Topology Appl. 160 (2013), 1313–1332.
  • [34] I. Moerdijk, Classifying spaces and classifying topoi, Lecture Notes in Mathematics. 1616. Berlin: Springer-Verlag, 1995.
  • [35] I. Moerdijk and D.A. Pronk, Orbifolds, sheaves and groupoids. K-Theory 12 (1997), 3–21.
  • [36] J. Neisendorfer, Algebraic Methods in Unstable Homotopy Theory, New Mathematical Monographs, vol. 12, Cambridge University Press, Cambridge, 2010.
  • [37] B. Ndombol and J.-C. Thomas, On the cohomology algebra of free loop spaces, Topology 41 (2002), 85–106.
  • [38] D. Quillen, Higher algebraic K-theory I, Algebr. K-Theory I, Proc. Conf. Battelle Inst. 1972, Lect. Notes Math. 341 (1973), 85–147.
  • [39] D.M. Roberts and R.F. Vozzo, Smooth loop stacks of differentiable stacks and gerbes, Cah. Topol. Géom. Différ. Catég., 59 (2018), 95–141.
  • [40] G. Segal, Categories and cohomology theories, Topology 13 (1974), 293–312.
  • [41] L. Smith, On the characteristic zero cohomology of the free loop space, Amer. J. Math. 103 (1981), 887–910.
  • [42] K. Simakawa, K.Yoshida and T. Hraguchi, Homology and cohomology via enriched bifunctors, Kyushu Journal of Mathematics, 72 (2018), 239–252.
  • [43] J.-M. Souriau, Groupes différentiels, Lecture Notes in Math., 836, Springer, 1980, 91–128.
  • [44] M. Vigué-Poirrier, Réalization de morphismes donnés en cohomologie et site spectrale d’Eilenberg–Moore, Trans. Amer. Math. Soc. 265 (1981), 447–484.
  • [45] C.A. Weibel, An introduction to homological algebra, Cambridge studies in advances mathematics 38, Cambridge, 1994.
  • [46] S.J. Witherspoon, Hochschild cohomology for algebras, Graduate Studies in Mathematics 204. Providence, RI: American Mathematical Society, 2019.