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

Configuration spaces as commutative monoids

Oscar Randal-Williams [email protected] Centre for Mathematical Sciences
Wilberforce Road
Cambridge CB3 0WB
UK
Abstract.

After 1-point compactification, the collection of all unordered configuration spaces of a manifold admits a commutative multiplication by superposition of configurations. We explain a simple (derived) presentation for this commutative monoid object. Using this presentation, one can quickly deduce Knudsen’s formula for the rational cohomology of configuration spaces, prove rational homological stability, and understand how automorphisms of the manifold act on the cohomology of configuration spaces. Similar considerations reproduce the work of Farb–Wolfson–Wood on homological densities.

1. Introduction

Let M𝑀M be the interior of a connected compact manifold with boundary. The 1-point compactification of the space Cn(M)subscript𝐶𝑛𝑀C_{n}(M) of unordered configurations in M𝑀M may be written as

(1.1) Cn(M)+=[(M+)nlocus where two points coincide]𝔖n,subscript𝐶𝑛superscript𝑀subscriptdelimited-[]superscriptsuperscript𝑀𝑛locus where two points coincidesubscript𝔖𝑛C_{n}(M)^{+}=\left[\frac{(M^{+})^{\wedge n}}{\text{locus where two points coincide}}\right]_{\mathfrak{S}_{n}},

the quotient formed in pointed spaces. Not-necessarily-disjoint union of unordered configurations defines a superposition product

Cn(M)+Cn(M)+Cn+n(M)+subscript𝐶𝑛superscript𝑀subscript𝐶superscript𝑛superscript𝑀subscript𝐶𝑛superscript𝑛superscript𝑀C_{n}(M)^{+}\wedge C_{n^{\prime}}(M)^{+}\longrightarrow C_{n+n^{\prime}}(M)^{+}

which is associative, commutative, and unital. This gives a unital commutative monoid object in the symmetric monoidal category 𝖳𝗈𝗉superscriptsubscript𝖳𝗈𝗉\mathsf{Top}_{*}^{\mathbb{N}} of \mathbb{N}-graded pointed spaces:

(M):nCn(M)+.:𝑀𝑛subscript𝐶𝑛superscript𝑀\mathbb{C}(M):n\longmapsto C_{n}(M)^{+}.

The goal of this note is to explain and exploit this algebraic structure.

In the following, for a pointed space X𝑋X we write X[n]𝑋delimited-[]𝑛X[n] for the \mathbb{N}-graded pointed space which consists of X𝑋X in grading n𝑛n and the point in all other gradings, and write 𝐂𝐨𝐦(Y)𝐂𝐨𝐦𝑌\mathbf{Com}(Y) for the free unital commutative monoid on an object Y𝖳𝗈𝗉𝑌superscriptsubscript𝖳𝗈𝗉Y\in\mathsf{Top}_{*}^{\mathbb{N}}.

Theorem 1.1.

There is a pushout square

𝐂𝐨𝐦(M+[2])𝐂𝐨𝐦superscript𝑀delimited-[]2{\mathbf{Com}(M^{+}[2])}S0[0]superscript𝑆0delimited-[]0{S^{0}[0]}𝐂𝐨𝐦(M+[1])𝐂𝐨𝐦superscript𝑀delimited-[]1{\mathbf{Com}(M^{+}[1])}(M)𝑀{\mathbb{C}(M)}ϵitalic-ϵ\scriptstyle{\epsilon}ΔΔ\scriptstyle{\Delta}

of unital commutative monoids in 𝖳𝗈𝗉superscriptsubscript𝖳𝗈𝗉\mathsf{Top}_{*}^{\mathbb{N}}, where ϵitalic-ϵ\epsilon is the augmentation and ΔΔ\Delta is induced by the diagonal inclusion M+[M+M+]𝔖2=𝐂𝐨𝐦(M+[1])(2)superscript𝑀subscriptdelimited-[]superscript𝑀superscript𝑀subscript𝔖2𝐂𝐨𝐦superscript𝑀delimited-[]12M^{+}\to[M^{+}\wedge M^{+}]_{\mathfrak{S}_{2}}=\mathbf{Com}(M^{+}[1])(2). Furthermore, this square is a homotopy pushout, i.e. there is an induced equivalence

𝐂𝐨𝐦(M+[1])𝐂𝐨𝐦(M+[2])𝕃S0[0](M).subscriptsuperscripttensor-product𝕃𝐂𝐨𝐦superscript𝑀delimited-[]2𝐂𝐨𝐦superscript𝑀delimited-[]1superscript𝑆0delimited-[]0similar-to𝑀\mathbf{Com}(M^{+}[1])\otimes^{\mathbb{L}}_{\mathbf{Com}(M^{+}[2])}S^{0}[0]\overset{\sim}{\longrightarrow}\mathbb{C}(M).

That the square is a strict pushout of unital commutative monoids is elementary: it means identifying (M)𝑀\mathbb{C}(M) as the quotient of the based symmetric power monoid of M+superscript𝑀M^{+} by the ideal given by those tuples which contain a repeated element, which is a reformulation of (1.1). The content of the theorem is that the square is also a homotopy pushout, rendering it amenable to homological calculation.

More generally, let π:LM:𝜋𝐿𝑀\pi:L\to M be a vector bundle, and let

Cn(M;L)+=[(L+)nlocus where two points have the same projection in M]𝔖n.subscript𝐶𝑛superscript𝑀𝐿subscriptdelimited-[]superscriptsuperscript𝐿𝑛locus where two points have the same projection in Msubscript𝔖𝑛C_{n}(M;L)^{+}=\left[\frac{(L^{+})^{\wedge n}}{\text{locus where two points have the same projection in $M$}}\right]_{\mathfrak{S}_{n}}.

These assemble in the same way to a unital commutative monoid object (M;L)𝑀𝐿\mathbb{C}(M;L). (Of course more general spaces of labels can be implemented too, but the above suffices for us.)

Theorem 1.2.

There is a pushout square

𝐂𝐨𝐦([(LL)+]𝔖2[2])𝐂𝐨𝐦subscriptdelimited-[]superscriptdirect-sum𝐿𝐿subscript𝔖2delimited-[]2{\mathbf{Com}([(L\oplus L)^{+}]_{\mathfrak{S}_{2}}[2])}S0[0]superscript𝑆0delimited-[]0{S^{0}[0]}𝐂𝐨𝐦(L+[1])𝐂𝐨𝐦superscript𝐿delimited-[]1{\mathbf{Com}(L^{+}[1])}(M;L)𝑀𝐿{\mathbb{C}(M;L)}ϵitalic-ϵ\scriptstyle{\epsilon}ΔΔ\scriptstyle{\Delta}

of unital commutative monoids in 𝖳𝗈𝗉superscriptsubscript𝖳𝗈𝗉\mathsf{Top}_{*}^{\mathbb{N}}, where ϵitalic-ϵ\epsilon is the augmentation and ΔΔ\Delta is induced by the diagonal inclusion [(LL)+]𝔖2[L+L+]𝔖2=𝐂𝐨𝐦(L+[1])(2)subscriptdelimited-[]superscriptdirect-sum𝐿𝐿subscript𝔖2subscriptdelimited-[]superscript𝐿superscript𝐿subscript𝔖2𝐂𝐨𝐦superscript𝐿delimited-[]12[(L\oplus L)^{+}]_{\mathfrak{S}_{2}}\to[L^{+}\wedge L^{+}]_{\mathfrak{S}_{2}}=\mathbf{Com}(L^{+}[1])(2). Furthermore, this square is a homotopy pushout, i.e. there is an induced equivalence

𝐂𝐨𝐦(L+[1])𝐂𝐨𝐦([(LL)+]𝔖2[2])𝕃S0[0](M;L).subscriptsuperscripttensor-product𝕃𝐂𝐨𝐦subscriptdelimited-[]superscriptdirect-sum𝐿𝐿subscript𝔖2delimited-[]2𝐂𝐨𝐦superscript𝐿delimited-[]1superscript𝑆0delimited-[]0similar-to𝑀𝐿\mathbf{Com}(L^{+}[1])\otimes^{\mathbb{L}}_{\mathbf{Com}([(L\oplus L)^{+}]_{\mathfrak{S}_{2}}[2])}S^{0}[0]\overset{\sim}{\longrightarrow}\mathbb{C}(M;L).

This strictly generalises Theorem 1.1, which is the case where L𝐿L is the 0-dimensional vector bundle, so we shall mostly focus on Theorem 1.2 for the rest of the paper.

Recall that the derived relative tensor product may be computed by the two-sided bar construction, formed in 𝖳𝗈𝗉superscriptsubscript𝖳𝗈𝗉\mathsf{Top}_{*}^{\mathbb{N}}, so the conclusion of Theorem 1.2 can equivalently be stated as an equivalence

(1.2) B(𝐂𝐨𝐦(L+[1]),𝐂𝐨𝐦([(LL)+]𝔖2[2]),S0[0])(M;L).𝐵𝐂𝐨𝐦superscript𝐿delimited-[]1𝐂𝐨𝐦subscriptdelimited-[]superscriptdirect-sum𝐿𝐿subscript𝔖2delimited-[]2superscript𝑆0delimited-[]0similar-to𝑀𝐿B(\mathbf{Com}(L^{+}[1]),\mathbf{Com}([(L\oplus L)^{+}]_{\mathfrak{S}_{2}}[2]),S^{0}[0])\overset{\sim}{\longrightarrow}\mathbb{C}(M;L).

This formula has many applications to the homology of configuration spaces. As one application we will show how to recover Knudsen’s [Knu17] formula for H(Cn(M);)superscript𝐻subscript𝐶𝑛𝑀H^{*}(C_{n}(M);\mathbb{Q}) in terms of the compactly-supported \mathbb{Q}-cohomology of M𝑀M and its cup-product map, which in particular quickly implies homological stability. As another application we will show that the action on H(Cn(M);)superscript𝐻subscript𝐶𝑛𝑀H^{*}(C_{n}(M);\mathbb{Q}) of the group of proper homotopy self-equivalences of M𝑀M factors over a surprisingly small group. Finally, in an appendix written with Quoc P. Ho, we show how similar considerations reproduces the work of Farb–Wolfson–Wood [FWW19] on homological densities.

Context. This note is my attempt to give a topological implementation of some of the sheaf-theoretic ideas of Banerjee [Ban23] in the case of configuration spaces. The applications to the homology of configuration spaces given in Section 2 arise by taking singular chains of the equivalence (1.2) to obtain a derived tensor product description of the chains on (M;L)𝑀𝐿\mathbb{C}(M;L): this description will also follow from [Ban] as explained in [Ban23, Remark 1.1]. As such, the purpose of this paper is

  1. (i)

    to give a space-level implementation/interpretation of Banerjee’s ideas in a specific case, in order to popularise them among topologists, and

  2. (ii)

    to explain how several classical, recent, and new results about the rational homology of configuration spaces can be obtained very efficiently from (1.2) (or its chain-level analogue).

Everything I will describe has much to do with the work of Ho [Ho21, Ho20], Petersen [Pet20], Knudsen [Knu17], Getzler [Get99a, Get99b], Kallel [Kal98], Bödigheimer–Cohen–Milgram [BCM93], Segal [Seg79], and Arnol’d [Arn70].

Acknowledgements. I am grateful to Andrea Bianchi, Sadok Kallel, and the anonymous referee for their useful feedback on an earlier version of the paper. ORW was supported by the ERC under the European Union’s Horizon 2020 research and innovation programme (grant agreement No. 756444).

2. Applications

2.1. Homology of configuration spaces

Let M𝑀M be d𝑑d-dimensional. The space Cn(M;L)+subscript𝐶𝑛superscript𝑀𝐿C_{n}(M;L)^{+} is the 1-point compactification of the n(d+dim(L))𝑛𝑑dimension𝐿n\cdot(d+\dim(L))-dimensional manifold

Cn(M;L):=[Ln{(l1,,ln)|π(li)=π(lj) and ij}]𝔖n.assignsubscript𝐶𝑛𝑀𝐿subscriptdelimited-[]superscript𝐿𝑛conditional-setsubscript𝑙1subscript𝑙𝑛𝜋subscript𝑙𝑖𝜋subscript𝑙𝑗 and 𝑖𝑗subscript𝔖𝑛C_{n}(M;L):=\left[L^{n}\setminus\{(l_{1},\ldots,l_{n})\,|\,\pi(l_{i})=\pi(l_{j})\text{ and }i\neq j\}\right]_{\mathfrak{S}_{n}}.

This is a vector bundle over Cn(M)subscript𝐶𝑛𝑀C_{n}(M), but is a manifold itself and is orientable if and only if the manifold L𝐿L is orientable and even-dimensional. To arrange this, we can take the vector bundle W𝑊W given by the orientation line of M𝑀M plus (d1)𝑑1(d-1) trivial line bundles. Thus by Poincaré duality we have

H(Cn(M);𝕜)H(Cn(M;W);𝕜)H~2dn(Cn(M;W)+;𝕜).superscript𝐻subscript𝐶𝑛𝑀𝕜superscript𝐻subscript𝐶𝑛𝑀𝑊𝕜subscript~𝐻limit-from2𝑑𝑛subscript𝐶𝑛superscript𝑀𝑊𝕜H^{*}(C_{n}(M);\mathbbm{k})\cong H^{*}(C_{n}(M;W);\mathbbm{k})\cong\widetilde{H}_{2dn-*}(C_{n}(M;W)^{+};\mathbbm{k}).

In view of this, the bar construction description (1.2) can be used, in combination with the homology of free commutative monoids (see [Mil69]), to investigate H(Cn(M);𝕜)superscript𝐻subscript𝐶𝑛𝑀𝕜H^{*}(C_{n}(M);\mathbbm{k}). We do not pursue this in general here, but rather focus on the case 𝕜=𝕜\mathbbm{k}=\mathbb{Q}, where a complete answer is possible, and reproduces a formula of Knudsen.

2.2. Revisiting Knudsen’s formula

For an \mathbb{N}-graded pointed space we write Hn,i(X):=H~i(X(n))assignsubscript𝐻𝑛𝑖𝑋subscript~𝐻𝑖𝑋𝑛H_{n,i}(X):=\widetilde{H}_{i}(X(n)), and similarly for chains. Write S(V)superscript𝑆𝑉S^{*}(V) for the free graded-commutative algebra on a homologically graded vector space V𝑉V, i.e. S(V)=n0[Vn]𝔖nsuperscript𝑆𝑉subscriptdirect-sum𝑛0subscriptdelimited-[]superscript𝑉tensor-productabsent𝑛subscript𝔖𝑛S^{*}(V)=\bigoplus_{n\geq 0}[V^{\otimes n}]_{\mathfrak{S}_{n}}, where the Koszul sign rule is implemented. If V𝑉V is equipped with additional \mathbb{N}-grading, then this is inherited by S(V)superscript𝑆𝑉S^{*}(V) (but there is no Koszul sign rule associated to the \mathbb{N}-grading, only to the homological grading).

We consider (M;W)𝑀𝑊\mathbb{C}(M;W). There is a map C~(W+;)[1]C~(𝐂𝐨𝐦(W+[1]);)subscript~𝐶superscript𝑊delimited-[]1subscript~𝐶𝐂𝐨𝐦superscript𝑊delimited-[]1\widetilde{C}_{*}(W^{+};\mathbb{Q})[1]\to\widetilde{C}_{*}(\mathbf{Com}(W^{+}[1]);\mathbb{Q}) and, using the Eilenberg–Zilber maps, it extends to a map of cdga’s

S(C~(W+;)[1])C,(𝐂𝐨𝐦(W+[1]);),superscript𝑆subscript~𝐶superscript𝑊delimited-[]1subscript𝐶𝐂𝐨𝐦superscript𝑊delimited-[]1S^{*}(\widetilde{C}_{*}(W^{+};\mathbb{Q})[1])\longrightarrow{C}_{*,*}(\mathbf{Com}(W^{+}[1]);\mathbb{Q}),

which is an equivalence (since the maps [(W+)n]h𝔖n[(W+)n]𝔖nsubscriptdelimited-[]superscriptsuperscript𝑊𝑛subscript𝔖𝑛subscriptdelimited-[]superscriptsuperscript𝑊𝑛subscript𝔖𝑛[(W^{+})^{\wedge n}]_{h\mathfrak{S}_{n}}\to[(W^{+})^{\wedge n}]_{\mathfrak{S}_{n}} are rational homology isomorphisms). Similarly, there is an equivalence of cdga’s

S(C~([(WW)+]𝔖2;)[2])C,(𝐂𝐨𝐦([(WW)+]𝔖2[2]);).superscript𝑆subscript~𝐶subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2delimited-[]2subscript𝐶𝐂𝐨𝐦subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2delimited-[]2S^{*}(\widetilde{C}_{*}([(W\oplus W)^{+}]_{\mathfrak{S}_{2}};\mathbb{Q})[2])\longrightarrow{C}_{*,*}(\mathbf{Com}([(W\oplus W)^{+}]_{\mathfrak{S}_{2}}[2]);\mathbb{Q}).

Furthermore, one may choose formality equivalences

H~(W+;)subscript~𝐻superscript𝑊\displaystyle\widetilde{H}_{*}(W^{+};\mathbb{Q}) C~(W+;)absentsubscript~𝐶superscript𝑊\displaystyle\longrightarrow\widetilde{C}_{*}(W^{+};\mathbb{Q})
H~([(WW)+]𝔖2;)subscript~𝐻subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2\displaystyle\widetilde{H}_{*}([(W\oplus W)^{+}]_{\mathfrak{S}_{2}};\mathbb{Q}) C~([(WW)+]𝔖2;),absentsubscript~𝐶subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2\displaystyle\longrightarrow\widetilde{C}_{*}([(W\oplus W)^{+}]_{\mathfrak{S}_{2}};\mathbb{Q}),

i.e. chain maps inducing the identity on homology, and hence obtain equivalences

S(H~(W+;)[1])superscript𝑆subscript~𝐻superscript𝑊delimited-[]1\displaystyle S^{*}(\widetilde{H}_{*}(W^{+};\mathbb{Q})[1]) S(C~(W+;)[1])absentsuperscript𝑆subscript~𝐶superscript𝑊delimited-[]1\displaystyle\longrightarrow S^{*}(\widetilde{C}_{*}(W^{+};\mathbb{Q})[1])
S(H~([(WW)+]𝔖2;)[2])superscript𝑆subscript~𝐻subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2delimited-[]2\displaystyle S^{*}(\widetilde{H}_{*}([(W\oplus W)^{+}]_{\mathfrak{S}_{2}};\mathbb{Q})[2]) S(C~([(WW)+]𝔖2;)[2])absentsuperscript𝑆subscript~𝐶subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2delimited-[]2\displaystyle\longrightarrow S^{*}(\widetilde{C}_{*}([(W\oplus W)^{+}]_{\mathfrak{S}_{2}};\mathbb{Q})[2])

of cdga’s. In \mathbb{N}-grading 2, the map ΔΔ\Delta induces a map

δ:H~([(WW)+]𝔖2;)ΔH~([(W+)2]𝔖2;)[H~(W+;)2]𝔖2.:subscript𝛿subscript~𝐻subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2subscriptΔsubscript~𝐻subscriptdelimited-[]superscriptsuperscript𝑊2subscript𝔖2subscriptdelimited-[]subscript~𝐻superscriptsuperscript𝑊tensor-productabsent2subscript𝔖2\delta_{*}:\widetilde{H}_{*}([(W\oplus W)^{+}]_{\mathfrak{S}_{2}};\mathbb{Q})\overset{\Delta_{*}}{\longrightarrow}\widetilde{H}_{*}([(W^{+})^{\wedge 2}]_{\mathfrak{S}_{2}};\mathbb{Q})\cong[\widetilde{H}_{*}(W^{+};\mathbb{Q})^{\otimes 2}]_{\mathfrak{S}_{2}}.

With these choices the square

S(H~([(WW)+]𝔖2;)[2])superscript𝑆subscript~𝐻subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2delimited-[]2{S^{*}(\widetilde{H}_{*}([(W\oplus W)^{+}]_{\mathfrak{S}_{2}};\mathbb{Q})[2])}C,(𝐂𝐨𝐦([(WW)+]𝔖2[2]);)subscript𝐶𝐂𝐨𝐦subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2delimited-[]2{{C}_{*,*}(\mathbf{Com}([(W\oplus W)^{+}]_{\mathfrak{S}_{2}}[2]);\mathbb{Q})}S(H~(W+;)[1])superscript𝑆subscript~𝐻superscript𝑊delimited-[]1{S^{*}(\widetilde{H}_{*}(W^{+};\mathbb{Q})[1])}C,(𝐂𝐨𝐦(W+[1]);)subscript𝐶𝐂𝐨𝐦superscript𝑊delimited-[]1{{C}_{*,*}(\mathbf{Com}(W^{+}[1]);\mathbb{Q})}S(δ)superscript𝑆subscript𝛿\scriptstyle{S^{*}(\delta_{*})}similar-to\scriptstyle{\sim}Δ#subscriptΔ#\scriptstyle{\Delta_{\#}}similar-to\scriptstyle{\sim}

need not commute, but does commute up to homotopy in the category of cdga’s because the two chain maps H~([(WW)+]𝔖2;)C~([(W+)2]𝔖2;)subscript~𝐻subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2subscript~𝐶subscriptdelimited-[]superscriptsuperscript𝑊2subscript𝔖2\widetilde{H}_{*}([(W\oplus W)^{+}]_{\mathfrak{S}_{2}};\mathbb{Q})\to\widetilde{C}_{*}([(W^{+})^{\wedge 2}]_{\mathfrak{S}_{2}};\mathbb{Q}) induce the same map on homology, namely δsubscript𝛿\delta_{*}, so are chain homotopic. The bar construction description then gives an identification

TorS(H~([(WW)+]𝔖2;)[2])(S(H~(W+;)[1]),[0])H,((M;W);).superscriptsubscriptTorsuperscript𝑆subscript~𝐻subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2delimited-[]2superscript𝑆subscript~𝐻superscript𝑊delimited-[]1delimited-[]0subscript𝐻𝑀𝑊\mathrm{Tor}_{*}^{S^{*}(\widetilde{H}_{*}([(W\oplus W)^{+}]_{\mathfrak{S}_{2}};\mathbb{Q})[2])}(S^{*}(\widetilde{H}_{*}(W^{+};\mathbb{Q})[1]),\mathbb{Q}[0])\cong H_{*,*}(\mathbb{C}(M;W);\mathbb{Q}).

Recall that for a free graded-commutative algebra S(V)superscript𝑆𝑉S^{*}(V) on a homologically graded vector space V𝑉V (perhaps equipped with a further \mathbb{N}-grading), there is a free resolution of the trivial left S(V)superscript𝑆𝑉S^{*}(V)-module \mathbb{Q} given by ϵ:S(VΣV):italic-ϵsuperscript𝑆direct-sum𝑉Σ𝑉similar-to\epsilon:S^{*}(V\oplus\Sigma V)\overset{\sim}{\to}\mathbb{Q} equipped with the differential given by (Σv)=vΣ𝑣𝑣\partial(\Sigma v)=v and extended by the Leibniz rule. It is usually called the Koszul resolution. It is indeed a resolution because it is the free graded-commutative algebra on the acyclic chain complex ΣVidVΣ𝑉𝑖𝑑𝑉\Sigma V\overset{id}{\to}V, and over \mathbb{Q} taking homology commutes with the formation of symmetric powers. Applying this resolution to calculate the Tor groups above gives the complex

(S(H~(W+;)[1]ΣH~([(WW)+]𝔖2;)[2]),)superscript𝑆direct-sumsubscript~𝐻superscript𝑊delimited-[]1Σsubscript~𝐻subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2delimited-[]2\left(S^{*}\big{(}\widetilde{H}_{*}(W^{+};\mathbb{Q})[1]\oplus\Sigma\widetilde{H}_{*}([(W\oplus W)^{+}]_{\mathfrak{S}_{2}};\mathbb{Q})[2]\big{)},\partial\right)

with differential given by (Σx)=Δ(x)S2(H~(W+[1];))Σ𝑥subscriptΔ𝑥superscript𝑆2subscript~𝐻superscript𝑊delimited-[]1\partial(\Sigma x)=\Delta_{*}(x)\in S^{2}(\widetilde{H}_{*}(W^{+}[1];\mathbb{Q})) for xH~([(WW)+]/𝔖2[2];)𝑥subscript~𝐻delimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2delimited-[]2x\in\widetilde{H}_{*}([(W\oplus W)^{+}]/\mathfrak{S}_{2}[2];\mathbb{Q}), and extended by the Leibniz rule. This can be simplified as follows. If M𝑀M is d𝑑d-dimensional then the Thom isomorphism gives H~(W+;)=ΣdH~(M+;w1)subscript~𝐻superscript𝑊superscriptΣ𝑑subscript~𝐻superscript𝑀superscriptsubscript𝑤1\widetilde{H}_{*}(W^{+};\mathbb{Q})=\Sigma^{d}\widetilde{H}_{*}(M^{+};\mathbb{Q}^{w_{1}}), where w1superscriptsubscript𝑤1\mathbb{Q}^{w_{1}} is the orientation local system of M𝑀M. It also gives H~((WW)+;)=Σ2dH~(M+;)subscript~𝐻superscriptdirect-sum𝑊𝑊superscriptΣ2𝑑subscript~𝐻superscript𝑀\widetilde{H}_{*}((W\oplus W)^{+};\mathbb{Q})=\Sigma^{2d}\widetilde{H}_{*}(M^{+};\mathbb{Q}). The involution swapping the two W𝑊W factors acts as (1)dsuperscript1𝑑(-1)^{d} on the Thom class, so because the map [(WW)+]h𝔖2[(WW)+]𝔖2subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2[(W\oplus W)^{+}]_{h\mathfrak{S}_{2}}\to[(W\oplus W)^{+}]_{\mathfrak{S}_{2}} is a rational equivalence we find

H~([(WW)+]𝔖2;)={Σ2dH~(M+;)d even0d odd.subscript~𝐻subscriptdelimited-[]superscriptdirect-sum𝑊𝑊subscript𝔖2casessuperscriptΣ2𝑑subscript~𝐻superscript𝑀𝑑 even0𝑑 odd\widetilde{H}_{*}([(W\oplus W)^{+}]_{\mathfrak{S}_{2}};\mathbb{Q})=\begin{cases}\Sigma^{2d}\widetilde{H}_{*}(M^{+};\mathbb{Q})&d\text{ even}\\ 0&d\text{ odd}.\end{cases}

This lets us write the complex as

(2.1) (S(ΣdH~(M+;w1)[1]{Σ2d+1H~(M+;)d even0d odd[2]),),superscript𝑆direct-sumsuperscriptΣ𝑑subscript~𝐻superscript𝑀superscriptsubscript𝑤1delimited-[]1casessuperscriptΣ2𝑑1subscript~𝐻superscript𝑀𝑑 even0𝑑 odddelimited-[]2\left(S^{*}\big{(}\Sigma^{d}\widetilde{H}_{*}(M^{+};\mathbb{Q}^{w_{1}})[1]\oplus\begin{cases}\Sigma^{2d+1}\widetilde{H}_{*}(M^{+};\mathbb{Q})&d\text{ even}\\ 0&d\text{ odd}\end{cases}[2]\big{)},\partial\right),

where the differential is dual to the map S2(Hc(M;w1))Hc(M;)superscript𝑆2subscriptsuperscript𝐻𝑐𝑀superscriptsubscript𝑤1subscriptsuperscript𝐻𝑐𝑀S^{2}(H^{*}_{c}(M;\mathbb{Q}^{w_{1}}))\to H^{*}_{c}(M;\mathbb{Q}) induced by cup product, so following Knudsen we can recognise this complex as the Chevelley–Eilenberg complex for the bigraded Lie algebra Hc(M;Lie(Σd1w1[1]))subscriptsuperscript𝐻𝑐𝑀LiesuperscriptΣ𝑑1superscriptsubscript𝑤1delimited-[]1H^{*}_{c}(M;\mathrm{Lie}(\Sigma^{d-1}\mathbb{Q}^{w_{1}}[1])). Thus

H2nd(Cn(M);)H~(Cn(M;W)+;)HLie(Hc(M;Lie(Σd1w1[1])))(n).superscript𝐻limit-from2𝑛𝑑subscript𝐶𝑛𝑀subscript~𝐻subscript𝐶𝑛superscript𝑀𝑊subscriptsuperscript𝐻Liesubscriptsuperscript𝐻𝑐𝑀LiesuperscriptΣ𝑑1superscriptsubscript𝑤1delimited-[]1𝑛H^{2nd-*}(C_{n}(M);\mathbb{Q})\cong\widetilde{H}_{*}(C_{n}(M;W)^{+};\mathbb{Q})\cong H^{*}_{\mathrm{Lie}}(H^{*}_{c}(M;\mathrm{Lie}(\Sigma^{d-1}\mathbb{Q}^{w_{1}}[1])))(n).

After appropriate dualisations and reindexings, this agrees with Knudsen’s formula.

2.3. Homological stability

Stability for the homology of configuration spaces is by now a classical subject, with a large number of contributions by many authors: notable examples are [Arn70, Seg79, Chu12, RW13, BM14, CP15, KM15, Knu17]. In particular Knudsen has explained [Knu17, Section 5.3] how his formula implies rational (co)homological stability for the spaces Cn(M)subscript𝐶𝑛𝑀C_{n}(M). Let us briefly review this from the point of view taken here.

There is a canonical element [M]H~d(M+;w1)H~2d(W+;)delimited-[]𝑀subscript~𝐻𝑑superscript𝑀superscriptsubscript𝑤1subscript~𝐻2𝑑superscript𝑊[M]\in\widetilde{H}_{d}(M^{+};\mathbb{Q}^{w_{1}})\cong\widetilde{H}_{2d}(W^{+};\mathbb{Q}), and choosing a cycle representing this element provides a map

σ:Σ2d[1]C,(𝐂𝐨𝐦(W+[1]);)C,((M;W);).:𝜎superscriptΣ2𝑑delimited-[]1subscript𝐶𝐂𝐨𝐦superscript𝑊delimited-[]1subscript𝐶𝑀𝑊\sigma:\Sigma^{2d}\mathbb{Q}[1]\longrightarrow{C}_{*,*}(\mathbf{Com}(W^{+}[1]);\mathbb{Q})\longrightarrow{C}_{*,*}(\mathbb{C}(M;W);\mathbb{Q}).

Multiplication by this element defines a map

(σ):H~n1,2d(n1)i((M;W);)H~n,2dni((M;W);)(\sigma\cdot-)_{*}:\widetilde{H}_{n-1,2d(n-1)-i}(\mathbb{C}(M;W);\mathbb{Q})\longrightarrow\widetilde{H}_{n,2dn-i}(\mathbb{C}(M;W);\mathbb{Q})

which under Poincaré duality gives a map Hi(Cn1(M);)Hi(Cn(M);)superscript𝐻𝑖subscript𝐶𝑛1𝑀superscript𝐻𝑖subscript𝐶𝑛𝑀H^{i}(C_{n-1}(M);\mathbb{Q})\to H^{i}(C_{n}(M);\mathbb{Q}); this can be checked to be the transfer map which sums over all ways of forgetting one of the n𝑛n points, see [Knu17, Section 5.2] [Sta23b, Section 2.6].

Writing C,((M;W);)/σsubscript𝐶𝑀𝑊𝜎{C}_{*,*}(\mathbb{C}(M;W);\mathbb{Q})/\sigma for the mapping cone of left multiplication by σ𝜎\sigma, the discussion above shows that its homology is calculated by a complex

(S(ΣdH~(M+;w1)[M][1]{Σ2d+1H~(M+;)d even0d odd[2]),).superscript𝑆direct-sumsuperscriptΣ𝑑subscript~𝐻superscript𝑀superscriptsubscript𝑤1delimited-⟨⟩delimited-[]𝑀delimited-[]1casessuperscriptΣ2𝑑1subscript~𝐻superscript𝑀𝑑 even0𝑑 odddelimited-[]2\left(S^{*}\big{(}\Sigma^{d}\frac{\widetilde{H}_{*}(M^{+};\mathbb{Q}^{w_{1}})}{\langle[M]\rangle}[1]\oplus\begin{cases}\Sigma^{2d+1}\widetilde{H}_{*}(M^{+};\mathbb{Q})&d\text{ even}\\ 0&d\text{ odd}\end{cases}[2]\big{)},\partial\right).

As M𝑀M is connected, if we assume that d3𝑑3d\geq 3 then the bigraded vector spaces ΣdH~(M+;w1)[M][1]superscriptΣ𝑑subscript~𝐻superscript𝑀superscriptsubscript𝑤1delimited-⟨⟩delimited-[]𝑀delimited-[]1\Sigma^{d}\frac{\widetilde{H}_{*}(M^{+};\mathbb{Q}^{w_{1}})}{\langle[M]\rangle}[1] and Σ2d+1H~(M+;)[2]superscriptΣ2𝑑1subscript~𝐻superscript𝑀delimited-[]2\Sigma^{2d+1}\widetilde{H}_{*}(M^{+};\mathbb{Q})[2] both vanish in bidegrees (n,j)𝑛𝑗(n,j) satisfying j>(2d1)n𝑗2𝑑1𝑛j>(2d-1)n, and hence so does the free graded-commutative algebra on them. This translates to Hi(Cn1(M);)Hi(Cn(M);)superscript𝐻𝑖subscript𝐶𝑛1𝑀superscript𝐻𝑖subscript𝐶𝑛𝑀H^{i}(C_{n-1}(M);\mathbb{Q})\to H^{i}(C_{n}(M);\mathbb{Q}) being surjective for i<n𝑖𝑛i<n and an isomorphism for i<n1𝑖𝑛1i<n-1. For d=2𝑑2d=2 the same considerations give surjectivity for i<12n𝑖12𝑛i<\tfrac{1}{2}n and so on (a more careful analysis gives a slope 1 range in this case too, see [Knu17, Proof of Theorem 1.3]).

Analysing the complex (2.1) can also establish other kinds of stability results, e.g. [BY21, KMT23, Yam23].

2.4. The action of automorphisms on unordered configurations

Using Knudsen’s formula it is possible to mislead yourself into thinking that homeomorphisms of M𝑀M (or indeed pointed homotopy self-equivalences of M+superscript𝑀M^{+}) act on H(Cn(M);)subscript𝐻subscript𝐶𝑛𝑀H_{*}(C_{n}(M);\mathbb{Q}) via their action on H(M;)subscript𝐻𝑀H_{*}(M;\mathbb{Q}): in other words, that such maps which act trivially on the homology of M𝑀M also act trivially on the homology of Cn(M)subscript𝐶𝑛𝑀C_{n}(M). This is not true: in the case of surfaces see Bianchi [Bia20, Section 7], Looijenga [Loo23], and the complete analysis given by Stavrou [Sta23a].

From the point of view taken here this phenomenon can be explained as follows. For simplicity suppose that M𝑀M is orientable, and first suppose that it is odd-dimensional. Then H(Cn(M);)H~2dn(Cn(M;M×d)+;)superscript𝐻subscript𝐶𝑛𝑀subscript~𝐻limit-from2𝑑𝑛subscript𝐶𝑛superscript𝑀𝑀superscript𝑑H^{*}(C_{n}(M);\mathbb{Q})\cong\widetilde{H}_{2dn-*}(C_{n}(M;M\times\mathbb{R}^{d})^{+};\mathbb{Q}) and the analysis of Section 2.2 applied to (M;M×d)𝑀𝑀superscript𝑑\mathbb{C}(M;M\times\mathbb{R}^{d}) shows that 𝐂𝐨𝐦(SdM+[1])(M;M×d)𝐂𝐨𝐦superscript𝑆𝑑superscript𝑀delimited-[]1𝑀𝑀superscript𝑑\mathbf{Com}(S^{d}\wedge M^{+}[1])\to\mathbb{C}(M;M\times\mathbb{R}^{d}) is a rational homology isomorphism. So we find:

Theorem 2.1.

If M𝑀M is orientable and odd-dimensional, then a pointed homotopy self-equivalence of M+superscript𝑀M^{+} which acts trivially on H~(M+;)subscript~𝐻superscript𝑀\widetilde{H}_{*}(M^{+};\mathbb{Q}) also acts trivially on H(Cn(M);)superscript𝐻subscript𝐶𝑛𝑀H^{*}(C_{n}(M);\mathbb{Q}).∎

The even-dimensional case is more interesting. As M𝑀M is assumed orientable, in this case the twisting by W𝑊W can be dispensed with. It is technically convenient here—for reasons of symmetric monoidality—to work in the category of simplicial \mathbb{Q}-modules rather than chain complexes. We write direct-product-\odot- for the tensoring of this category over simplicial sets. For a space X𝑋X let us abbreviate [X]:=[Sing(X)]assigndelimited-[]𝑋delimited-[]subscriptSing𝑋\mathbb{Q}[X]:=\mathbb{Q}[\mathrm{Sing}_{\bullet}(X)], and if it is based then let ~[X]=[X]/[]~delimited-[]𝑋delimited-[]𝑋delimited-[]\widetilde{\mathbb{Q}}[X]=\mathbb{Q}[X]/\mathbb{Q}[*]. The discussion in the previous section, ignoring the formality step and translated to simplicial \mathbb{Q}-modules, shows that given the simplicial module ~[M+]~delimited-[]superscript𝑀\widetilde{\mathbb{Q}}[M^{+}] and the map δ:~[M+][~[M+]2]𝔖2:𝛿~delimited-[]superscript𝑀subscriptdelimited-[]~superscriptdelimited-[]superscript𝑀tensor-productabsent2subscript𝔖2\delta:\widetilde{\mathbb{Q}}[M^{+}]\to\big{[}\widetilde{\mathbb{Q}}[M^{+}]^{\otimes 2}\big{]}_{\mathfrak{S}_{2}} induced by the diagonal M+M+M+superscript𝑀superscript𝑀superscript𝑀M^{+}\to M^{+}\wedge M^{+}, we may form the two-sided bar construction

(2.2) B(S(~[M+][1]),S(~[M+][2]),[0])𝐵superscript𝑆~delimited-[]superscript𝑀delimited-[]1superscript𝑆~delimited-[]superscript𝑀delimited-[]2delimited-[]0B(S^{*}(\widetilde{\mathbb{Q}}[M^{+}][1]),S^{*}(\widetilde{\mathbb{Q}}[M^{+}][2]),\mathbb{Q}[0])

whose bigraded homotopy groups are identified with H~(C(M)+;)subscript~𝐻subscript𝐶superscript𝑀\widetilde{H}_{*}(C_{*}(M)^{+};\mathbb{Q}).

A homeomorphism of M𝑀M, or a pointed homotopy self-equivalence of M+superscript𝑀M^{+}, induces an equivalence ϕ:~[M+]~[M+]:italic-ϕ~delimited-[]superscript𝑀~delimited-[]superscript𝑀\phi:\widetilde{\mathbb{Q}}[M^{+}]\to\widetilde{\mathbb{Q}}[M^{+}] such that δϕ=([ϕϕ]𝔖2)δ𝛿italic-ϕsubscriptdelimited-[]tensor-productitalic-ϕitalic-ϕsubscript𝔖2𝛿\delta\circ\phi=([\phi\otimes\phi]_{\mathfrak{S}_{2}})\circ\delta, meaning that the diagram of simplicial commutative rings

S(~[M+][1])superscript𝑆~delimited-[]superscript𝑀delimited-[]1{S^{*}(\widetilde{\mathbb{Q}}[M^{+}][1])}S(~[M+][2])superscript𝑆~delimited-[]superscript𝑀delimited-[]2{S^{*}(\widetilde{\mathbb{Q}}[M^{+}][2])}[0]delimited-[]0{\mathbb{Q}[0]}S(~[M+][1])superscript𝑆~delimited-[]superscript𝑀delimited-[]1{S^{*}(\widetilde{\mathbb{Q}}[M^{+}][1])}S(~[M+][2])superscript𝑆~delimited-[]superscript𝑀delimited-[]2{S^{*}(\widetilde{\mathbb{Q}}[M^{+}][2])}[0]delimited-[]0{\mathbb{Q}[0]}S(ϕ)superscript𝑆italic-ϕ\scriptstyle{S^{*}(\phi)}S(δ)superscript𝑆𝛿\scriptstyle{S^{*}(\delta)}ϵitalic-ϵ\scriptstyle{\epsilon}S(ϕ)superscript𝑆italic-ϕ\scriptstyle{S^{*}(\phi)}id𝑖𝑑\scriptstyle{id}S(δ)superscript𝑆𝛿\scriptstyle{S^{*}(\delta)}ϵitalic-ϵ\scriptstyle{\epsilon}

is commutative, which induces a self-equivalence of the two-sided bar construction (2.2). This corresponds to the induced action on H~(C(M)+;)subscript~𝐻subscript𝐶superscript𝑀\widetilde{H}_{*}(C_{*}(M)^{+};\mathbb{Q}).

However a weaker kind of data suffices to get an induced equivalence on two-sided bar constructions. An equivalence ϕ:~[M+]~[M+]:italic-ϕ~delimited-[]superscript𝑀~delimited-[]superscript𝑀\phi:\widetilde{\mathbb{Q}}[M^{+}]\to\widetilde{\mathbb{Q}}[M^{+}] together with a homotopy h:δϕ([ϕϕ]𝔖2)δ:𝛿italic-ϕsubscriptdelimited-[]tensor-productitalic-ϕitalic-ϕsubscript𝔖2𝛿h:\delta\circ\phi\Rightarrow([\phi\otimes\phi]_{\mathfrak{S}_{2}})\circ\delta gives a diagram of simplicial commutative rings as above where the right-hand square commutes and the left-hand squares commutes up to the homotopy S(h):S(Δ1~[M+][2])S(~[M+][1]):superscript𝑆superscript𝑆direct-productsuperscriptΔ1~delimited-[]superscript𝑀delimited-[]2superscript𝑆~delimited-[]superscript𝑀delimited-[]1S^{*}(h):S^{*}(\Delta^{1}\odot\widetilde{\mathbb{Q}}[M^{+}][2])\to S^{*}(\widetilde{\mathbb{Q}}[M^{+}][1]). This data suffices to obtain a self-equivalence χ(ϕ,h)𝜒italic-ϕ\chi(\phi,h) of the two-sided bar construction (2.2), as the zig-zag

B(S(~[M+][1]),S(Δ1~[M+][2]),[0])𝐵superscript𝑆~delimited-[]superscript𝑀delimited-[]1superscript𝑆direct-productsuperscriptΔ1~delimited-[]superscript𝑀delimited-[]2delimited-[]0{B(S^{*}(\widetilde{\mathbb{Q}}[M^{+}][1]),S^{*}(\Delta^{1}\odot\widetilde{\mathbb{Q}}[M^{+}][2]),\mathbb{Q}[0])}B(S(~[M+][1]),S(~[M+][2]),[0])𝐵superscript𝑆superscript~delimited-[]superscript𝑀delimited-[]1superscript𝑆~delimited-[]superscript𝑀delimited-[]2delimited-[]0{B(S^{*}(\widetilde{\mathbb{Q}}[M^{+}][1])^{\prime},S^{*}(\widetilde{\mathbb{Q}}[M^{+}][2]),\mathbb{Q}[0])}B(S(~[M+][1]),S(~[M+][2]),[0])𝐵superscript𝑆~delimited-[]superscript𝑀delimited-[]1superscript𝑆~delimited-[]superscript𝑀delimited-[]2delimited-[]0{B(S^{*}(\widetilde{\mathbb{Q}}[M^{+}][1]),S^{*}(\widetilde{\mathbb{Q}}[M^{+}][2]),\mathbb{Q}[0])}B(S(~[M+][1]),S(~[M+][2]),[0])𝐵superscript𝑆~delimited-[]superscript𝑀delimited-[]1superscript𝑆~delimited-[]superscript𝑀delimited-[]2delimited-[]0{B(S^{*}(\widetilde{\mathbb{Q}}[M^{+}][1]),S^{*}(\widetilde{\mathbb{Q}}[M^{+}][2]),\mathbb{Q}[0])}B(id,S(d0),id)𝐵𝑖𝑑superscript𝑆subscript𝑑0𝑖𝑑\scriptstyle{B(id,S^{*}(d_{0}),id)}B(id,S(ϕ),id)𝐵𝑖𝑑superscript𝑆italic-ϕ𝑖𝑑\scriptstyle{B(id,S^{*}(\phi),id)}B(S(ϕ),S(d1),id)𝐵superscript𝑆italic-ϕsuperscript𝑆subscript𝑑1𝑖𝑑\scriptstyle{B(S^{*}(\phi),S^{*}(d_{1}),id)}χ(ϕ,h)𝜒italic-ϕ\scriptstyle{\chi(\phi,h)}

where S(~[M+][1])superscript𝑆superscript~delimited-[]superscript𝑀delimited-[]1S^{*}(\widetilde{\mathbb{Q}}[M^{+}][1])^{\prime} denotes S(~[M+][1])superscript𝑆~delimited-[]superscript𝑀delimited-[]1S^{*}(\widetilde{\mathbb{Q}}[M^{+}][1]) considered as a S(~[M+][2])superscript𝑆~delimited-[]superscript𝑀delimited-[]2S^{*}(\widetilde{\mathbb{Q}}[M^{+}][2])-module via S(ϕ)S(δ)superscript𝑆italic-ϕsuperscript𝑆𝛿S^{*}(\phi)\circ S^{*}(\delta).

Let (ϕ,h)superscriptitalic-ϕsuperscript(\phi^{\prime},h^{\prime}) be another such datum, and suppose that there is a homotopy Φ:ϕϕ:Φitalic-ϕsuperscriptitalic-ϕ\Phi:\phi\Rightarrow\phi^{\prime} such that the 2-cell

(2.3) [~[M+]2]𝔖2subscriptdelimited-[]~superscriptdelimited-[]superscript𝑀tensor-productabsent2subscript𝔖2{{[\widetilde{\mathbb{Q}}[M^{+}]^{\otimes 2}]_{\mathfrak{S}_{2}}}}~[M+]~delimited-[]superscript𝑀{\widetilde{\mathbb{Q}}[M^{+}]}[~[M+]2]𝔖2subscriptdelimited-[]~superscriptdelimited-[]superscript𝑀tensor-productabsent2subscript𝔖2{{[\widetilde{\mathbb{Q}}[M^{+}]^{\otimes 2}]_{\mathfrak{S}_{2}}}}~[M+]~delimited-[]superscript𝑀{\widetilde{\mathbb{Q}}[M^{+}]}[~[M+]2]𝔖2subscriptdelimited-[]~superscriptdelimited-[]superscript𝑀tensor-productabsent2subscript𝔖2{{[\widetilde{\mathbb{Q}}[M^{+}]^{\otimes 2}]_{\mathfrak{S}_{2}}}}~[M+]~delimited-[]superscript𝑀{\widetilde{\mathbb{Q}}[M^{+}]}[~[M+]2]𝔖2subscriptdelimited-[]~superscriptdelimited-[]superscript𝑀tensor-productabsent2subscript𝔖2{{[\widetilde{\mathbb{Q}}[M^{+}]^{\otimes 2}]_{\mathfrak{S}_{2}}}}~[M+]~delimited-[]superscript𝑀{\widetilde{\mathbb{Q}}[M^{+}]}[(ϕ)2]𝔖2subscriptdelimited-[]superscriptsuperscriptitalic-ϕtensor-productabsent2subscript𝔖2\scriptstyle{[(\phi^{\prime})^{\otimes 2}]_{\mathfrak{S}_{2}}}ϕsuperscriptitalic-ϕ\scriptstyle{\phi^{\prime}}δ𝛿\scriptstyle{\delta}[ϕ2]𝔖2subscriptdelimited-[]superscriptitalic-ϕtensor-productabsent2subscript𝔖2\scriptstyle{[\phi^{\otimes 2}]_{\mathfrak{S}_{2}}}δ𝛿\scriptstyle{\delta}ϕitalic-ϕ\scriptstyle{\phi}δ𝛿\scriptstyle{\delta}δ𝛿\scriptstyle{\delta}[Φ2]𝔖2subscriptdelimited-[]superscriptΦtensor-productabsent2subscript𝔖2{[\Phi^{\otimes 2}]_{\mathfrak{S}_{2}}\quad}ΦΦ{\Phi}h{h}Id𝐼𝑑{Id}Id𝐼𝑑{Id}

is homotopic to hsuperscripth^{\prime}. Then one may check that χ(ϕ,h)𝜒superscriptitalic-ϕsuperscript\chi(\phi^{\prime},h^{\prime}) is homotopic to χ(ϕ,h)𝜒italic-ϕ\chi(\phi,h). If we let ΓΓ\Gamma denote the set of (ϕ,h)italic-ϕ(\phi,h)’s modulo the equivalence relation (ϕ,h)(ϕ,h)similar-toitalic-ϕsuperscriptitalic-ϕsuperscript(\phi,h)\sim(\phi^{\prime},h^{\prime}) when there exists a homotopy ΦΦ\Phi having the above property, then composition of maps and pasting of homotopies makes ΓΓ\Gamma into a group, which acts on the two-sided bar construction (2.2) in the homotopy category of simplicial \mathbb{Q}-modules (and so also acts on its homotopy groups). A pointed homotopy self-equivalence of M+superscript𝑀M^{+} acts on the two-sided bar construction through ΓΓ\Gamma, via elements of the special form [(ϕ,Id)]delimited-[]italic-ϕ𝐼𝑑[(\phi,Id)].

We may analyse the group ΓΓ\Gamma as follows. There is a homomorphism

ρ:Γ:𝜌Γ\displaystyle\rho:\Gamma Aut(~[M+])absentAut~delimited-[]superscript𝑀\displaystyle\longrightarrow\mathrm{Aut}(\widetilde{\mathbb{Q}}[M^{+}])
[(ϕ,h)]delimited-[]italic-ϕ\displaystyle[(\phi,h)] [ϕ]absentdelimited-[]italic-ϕ\displaystyle\longmapsto[\phi]

to the group of homotopy classes of homotopy self-equivalences of ~[M+]~delimited-[]superscript𝑀\widetilde{\mathbb{Q}}[M^{+}]. Using the Dold–Kan theorem the latter can be identified with the group of homotopy classes of homotopy self-equivalences of C~(M+;)subscript~𝐶superscript𝑀\widetilde{C}_{*}(M^{+};\mathbb{Q}), and using a formality equivalence H~(M+;)C~(M+;)subscript~𝐻superscript𝑀similar-tosubscript~𝐶superscript𝑀\widetilde{H}_{*}(M^{+};\mathbb{Q})\overset{\sim}{\to}\widetilde{C}_{*}(M^{+};\mathbb{Q}) this is identified with the group Aut(H~(M+;))Autsubscript~𝐻superscript𝑀\mathrm{Aut}(\widetilde{H}_{*}(M^{+};\mathbb{Q})) of automorphisms of the graded vector space H~(M+;)subscript~𝐻superscript𝑀\widetilde{H}_{*}(M^{+};\mathbb{Q}). Such an automorphism is in the image of ρ𝜌\rho precisely when it preserves the map δ:H~(M+;)[H~(M+;)2]𝔖2:subscript𝛿subscript~𝐻superscript𝑀subscriptdelimited-[]subscript~𝐻superscriptsuperscript𝑀tensor-productabsent2subscript𝔖2\delta_{*}:\widetilde{H}_{*}(M^{+};\mathbb{Q})\to[\widetilde{H}_{*}(M^{+};\mathbb{Q})^{\otimes 2}]_{\mathfrak{S}_{2}}. The kernel of ρ𝜌\rho consists of those [(ϕ,h)]delimited-[]italic-ϕ[(\phi,h)] such that ϕitalic-ϕ\phi is homotopic to the identity: by definition of the equivalence relation similar-to\sim such an element may be written as [(Id,h)]delimited-[]𝐼𝑑superscript[(Id,h^{\prime})] where hsuperscripth^{\prime} is obtained from hh and a homotopy Φ:ϕid:Φitalic-ϕ𝑖𝑑\Phi:\phi\Rightarrow id by the 2-cell diagram (2.3). Such an hsuperscripth^{\prime} is a self-homotopy of the map δ𝛿\delta, so an element of π1(map(~[M+],[~[M+]2]𝔖2));δ)\pi_{1}(\mathrm{map}(\widetilde{\mathbb{Q}}[M^{+}],\big{[}\widetilde{\mathbb{Q}}[M^{+}]^{\otimes 2}\big{]}_{\mathfrak{S}_{2}}));\delta). The ambiguity in hsuperscripth^{\prime} when representing [(ϕ,h)]Ker(ρ)delimited-[]italic-ϕKer𝜌[(\phi,h)]\in\mathrm{Ker}(\rho) as [(Id,h)]delimited-[]𝐼𝑑superscript[(Id,h^{\prime})] comes from the choice of the homotopy ΦΦ\Phi, so hsuperscripth^{\prime} is well-defined modulo the ambiguity coming from the self-homotopies π1(map(~[M+],~[M+]);id)subscript𝜋1map~delimited-[]superscript𝑀~delimited-[]superscript𝑀𝑖𝑑\pi_{1}(\mathrm{map}(\widetilde{\mathbb{Q}}[M^{+}],\widetilde{\mathbb{Q}}[M^{+}]);id) of the identity map. In conclusion, this discussion establishes an exact sequence

π1(map(~[M+],~[M+]);id)subscript𝜋1map~delimited-[]superscript𝑀~delimited-[]superscript𝑀𝑖𝑑{\pi_{1}(\mathrm{map}(\widetilde{\mathbb{Q}}[M^{+}],\widetilde{\mathbb{Q}}[M^{+}]);id)}π1(map(~[M+],[~[M+]2]𝔖2);δ)subscript𝜋1map~delimited-[]superscript𝑀subscriptdelimited-[]~superscriptdelimited-[]superscript𝑀tensor-productabsent2subscript𝔖2𝛿{\pi_{1}(\mathrm{map}(\widetilde{\mathbb{Q}}[M^{+}],\big{[}\widetilde{\mathbb{Q}}[M^{+}]^{\otimes 2}\big{]}_{\mathfrak{S}_{2}});\delta)}ΓΓ{\Gamma}Aut(H~(M+;),δ)Autsubscript~𝐻superscript𝑀subscript𝛿{\mathrm{Aut}(\widetilde{H}_{*}(M^{+};\mathbb{Q}),\delta_{*})}1.1{1.}δ\scriptstyle{\delta\circ-}ρ𝜌\scriptstyle{\rho}

Using the Dold–Kan theorem and a formality equivalence again we can identify the first map in this sequence with

δ:Hom(ΣH~(M+;),H~(M+;))Hom(ΣH~(M+;),[H~(M+;)2]𝔖2),\delta_{*}\circ-:\mathrm{Hom}(\Sigma\widetilde{H}_{*}(M^{+};\mathbb{Q}),\widetilde{H}_{*}(M^{+};\mathbb{Q}))\longrightarrow\mathrm{Hom}(\Sigma\widetilde{H}_{*}(M^{+};\mathbb{Q}),[\widetilde{H}_{*}(M^{+};\mathbb{Q})^{\otimes 2}]_{\mathfrak{S}_{2}}),

and so describe ΓΓ\Gamma by an extension

1Hom(ΣH~(M+;),S2(H~(M+;))/Im(δ))ΓAut(H~(M+;),δ)1.1HomΣsubscript~𝐻superscript𝑀superscript𝑆2subscript~𝐻superscript𝑀Imsubscript𝛿ΓAutsubscript~𝐻superscript𝑀subscript𝛿11\to\mathrm{Hom}(\Sigma\widetilde{H}_{*}(M^{+};\mathbb{Q}),S^{2}(\widetilde{H}_{*}(M^{+};\mathbb{Q}))/\mathrm{Im}(\delta_{*}))\to\Gamma\to\mathrm{Aut}(\widetilde{H}_{*}(M^{+};\mathbb{Q}),\delta_{*})\to 1.

This implies the following. We continue to assume that M𝑀M is even-dimensional and orientable. Let G𝐺G denote the group of homotopy classes of pointed homotopy self-equivalences of M+superscript𝑀M^{+} which act as the identity on H~(M+;)subscript~𝐻superscript𝑀\widetilde{H}_{*}(M^{+};\mathbb{Q}).

Theorem 2.2.

If M𝑀M is orientable and even-dimensional, then G𝐺G acts on H(Cn(M);)superscript𝐻subscript𝐶𝑛𝑀{H}^{*}(C_{n}(M);\mathbb{Q}) via Hom(ΣH~(M+;),S2(H~(M+;))/Im(δ))HomΣsubscript~𝐻superscript𝑀superscript𝑆2subscript~𝐻superscript𝑀Imsubscript𝛿\mathrm{Hom}(\Sigma\widetilde{H}_{*}(M^{+};\mathbb{Q}),S^{2}(\widetilde{H}_{*}(M^{+};\mathbb{Q}))/\mathrm{Im}(\delta_{*})).∎

Example 2.3.

When M𝑀M is a punctured surface one has H~(M+;)=ΣH1(M;)Σ2subscript~𝐻superscript𝑀direct-sumΣsubscript𝐻1𝑀superscriptΣ2\widetilde{H}_{*}(M^{+};\mathbb{Q})=\Sigma H_{1}(M;\mathbb{Q})\oplus\Sigma^{2}\mathbb{Q} so the map δ:H~(M+;)S2(H~(M+;)):subscript𝛿subscript~𝐻superscript𝑀superscript𝑆2subscript~𝐻superscript𝑀\delta_{*}:\widetilde{H}_{*}(M^{+};\mathbb{Q})\to S^{2}(\widetilde{H}_{*}(M^{+};\mathbb{Q})) has the form

ΣH1(M;)Σ2Σ2Λ2(H1(M;))Σ3H1(M;)Σ4,direct-sumΣsubscript𝐻1𝑀superscriptΣ2direct-sumsuperscriptΣ2superscriptΛ2subscript𝐻1𝑀superscriptΣ3subscript𝐻1𝑀superscriptΣ4\Sigma H_{1}(M;\mathbb{Q})\oplus\Sigma^{2}\mathbb{Q}\longrightarrow\Sigma^{2}\Lambda^{2}(H_{1}(M;\mathbb{Q}))\oplus\Sigma^{3}H_{1}(M;\mathbb{Q})\oplus\Sigma^{4}\mathbb{Q},

which in grading 2 is the inclusion of the symplectic form ωΛ2(H1(M;))𝜔superscriptΛ2subscript𝐻1𝑀\omega\in\Lambda^{2}(H_{1}(M;\mathbb{Q})) and is zero otherwise. Thus the above is Hom(H1(M;),Λ2(H1(M;))/ω)H1(M;)direct-sumHomsubscript𝐻1𝑀superscriptΛ2subscript𝐻1𝑀delimited-⟨⟩𝜔subscript𝐻1𝑀\mathrm{Hom}(H_{1}(M;\mathbb{Q}),\Lambda^{2}(H_{1}(M;\mathbb{Q}))/\langle\omega\rangle)\oplus H_{1}(M;\mathbb{Q}). Using Poincaré duality and Λ2(H1(M;)){ω}Λ2(H1(M;))/ωsuperscriptΛ2subscript𝐻1𝑀direct-sum𝜔superscriptΛ2subscript𝐻1𝑀delimited-⟨⟩𝜔\Lambda^{2}(H_{1}(M;\mathbb{Q}))\cong\mathbb{Q}\{\omega\}\oplus\Lambda^{2}(H_{1}(M;\mathbb{Q}))/\langle\omega\rangle, this can be identified with Hom(H1(M;),Λ2(H1(M;)))Homsubscript𝐻1𝑀superscriptΛ2subscript𝐻1𝑀\mathrm{Hom}(H_{1}(M;\mathbb{Q}),\Lambda^{2}(H_{1}(M;\mathbb{Q}))). This is the target of the Johnson homomorphism, cf. [Sta23a].

Remark 2.4.

The results of this section should also follow from [Sta23a, Theorem 1.2] and some rational homotopy theory.

3. Proof of Theorem 1.2

Recall that X𝖳𝗈𝗉𝑋subscript𝖳𝗈𝗉X\in\mathsf{Top}_{*} is well-based if the basepoint map i:Xi:*\to X is a closed cofibration: under this condition XX\wedge- preserves weak equivalences between well-based spaces, and preserves closed cofibrations. Let us say that an \mathbb{N}-graded based space Y𝑌Y is well-based if Y(n)𝑌𝑛Y(n) is well-based for each n𝑛n\in\mathbb{N}.

Let us write :=𝐂𝐨𝐦(L+[1])assign𝐂𝐨𝐦superscript𝐿delimited-[]1\mathbb{R}:=\mathbf{Com}(L^{+}[1]) and 𝕊:=𝐂𝐨𝐦([(LL)+]𝔖2[2])assign𝕊𝐂𝐨𝐦subscriptdelimited-[]superscriptdirect-sum𝐿𝐿subscript𝔖2delimited-[]2\mathbb{S}:=\mathbf{Com}([(L\oplus L)^{+}]_{\mathfrak{S}_{2}}[2]) to ease notation, so Δ:𝕊:Δ𝕊\Delta:\mathbb{S}\to\mathbb{R} makes \mathbb{R} into a 𝕊𝕊\mathbb{S}-module.

Lemma 3.1.

𝕊𝕊\mathbb{S} and \mathbb{R} are well-based. The subspace of [(L+)p]𝔖psubscriptdelimited-[]superscriptsuperscript𝐿𝑝subscript𝔖𝑝[(L^{+})^{\wedge p}]_{\mathfrak{S}_{p}} of those tuples which do not have distinct M𝑀M coordinates is well-based, and this inclusion is a closed cofibration.

Proof.

Recall that M𝑀M is the interior of a compact manifold with boundary M¯¯𝑀\overline{M}. This admits a collar, showing that i:MM¯:𝑖𝑀¯𝑀i:M\to\overline{M} admits a homotopy inverse, and so the vector bundle LM𝐿𝑀L\to M extends to a vector bundle over M¯¯𝑀\overline{M}, which we also call L𝐿L. Furthermore, choosing an inner product on this bundle we can form the closed disc bundle D(L)M¯𝐷𝐿¯𝑀D(L)\to\overline{M}, and consider L𝐿L as lying inside it as the open disc bundle. Now D(L)𝐷𝐿D(L) is a manifold with boundary D(L)=S(L)D(L)|M¯𝐷𝐿𝑆𝐿evaluated-at𝐷𝐿¯𝑀\partial D(L)=S(L)\cup D(L)|_{\partial\overline{M}}, and L+=D(L)/D(L)superscript𝐿𝐷𝐿𝐷𝐿L^{+}=D(L)/\partial D(L).

Observe that (M¯,M¯)¯𝑀¯𝑀(\overline{M},\partial\overline{M}) is an compact manifold pair so (is an ENR pair and hence) can be expressed as a retract of a pair (|X|,|X|)subscript𝑋subscript𝑋(|X_{\bullet}|,|\partial X_{\bullet}|) of the geometric realisations of a simplicial set and a subset. We may pull L𝐿L back to |X|subscript𝑋|X_{\bullet}| using the retraction; let us call this LXsubscript𝐿𝑋L_{X}. Now D(LX)/S(LX)D(LX)||X|𝐷subscript𝐿𝑋𝑆subscript𝐿𝑋evaluated-at𝐷subscript𝐿𝑋subscript𝑋D(L_{X})/S(L_{X})\cup D(L_{X})|_{|\partial X_{\bullet}|} can be given an evident cell-structure (by induction over the relative cells of |X||X|subscript𝑋subscript𝑋|\partial X_{\bullet}|\to|X_{\bullet}|), and L+=D(L)/D(L)superscript𝐿𝐷𝐿𝐷𝐿L^{+}=D(L)/\partial D(L) is a retract of it, so is well-based. More generally, for the exterior direct sum LXp|Xp|superscriptsubscript𝐿𝑋𝑝superscriptsubscript𝑋𝑝L_{X}^{\boxplus p}\to|X_{\bullet}^{p}| and writing |Xp|superscriptsubscript𝑋𝑝\partial|X_{\bullet}^{p}| for the subcomplex where some factor lies in Xsubscript𝑋\partial X_{\bullet}, there is a cell structure on D(LXp)/S(LXp)D(LXp)||Xp|𝐷superscriptsubscript𝐿𝑋𝑝𝑆superscriptsubscript𝐿𝑋𝑝evaluated-at𝐷superscriptsubscript𝐿𝑋𝑝superscriptsubscript𝑋𝑝D(L_{X}^{\boxplus p})/S(L_{X}^{\boxplus p})\cup D(L_{X}^{\boxplus p})|_{\partial|X_{\bullet}^{p}|} for which the group 𝔖psubscript𝔖𝑝\mathfrak{S}_{p} acts cellularly, and so [D(LXp)/S(LXp)D(LXp)||Xp|]𝔖psubscriptdelimited-[]𝐷superscriptsubscript𝐿𝑋𝑝𝑆superscriptsubscript𝐿𝑋𝑝evaluated-at𝐷superscriptsubscript𝐿𝑋𝑝superscriptsubscript𝑋𝑝subscript𝔖𝑝[D(L_{X}^{\boxplus p})/S(L_{X}^{\boxplus p})\cup D(L_{X}^{\boxplus p})|_{\partial|X_{\bullet}^{p}|}]_{\mathfrak{S}_{p}} is a cell complex of which [(L+)p]𝔖psubscriptdelimited-[]superscriptsuperscript𝐿𝑝subscript𝔖𝑝[(L^{+})^{\wedge p}]_{\mathfrak{S}_{p}} is a retract, and so is well-based. This shows that \mathbb{R} is well-based, and similar reasoning shows 𝕊𝕊\mathbb{S} is.

For the second statement,

inc:F:=fat diagonal of |X|p=|fat diagonal of Xp||Xp|=|X|p:𝑖𝑛𝑐assign𝐹fat diagonal of superscriptsubscript𝑋𝑝fat diagonal of superscriptsubscript𝑋𝑝superscriptsubscript𝑋𝑝superscriptsubscript𝑋𝑝inc:F:=\text{fat diagonal of }|X_{\bullet}|^{p}=|\text{fat diagonal of }X_{\bullet}^{p}|\longrightarrow|X_{\bullet}^{p}|=|X_{\bullet}|^{p}

is the inclusion of a 𝔖psubscript𝔖𝑝\mathfrak{S}_{p}-CW-subcomplex, and so has a 𝔖psubscript𝔖𝑝\mathfrak{S}_{p}-equivariant open neighbourhood U𝑈U which equivariantly deformation retracts to it. This may be chosen to preserve the subcomplexes where some factor lies in |X|subscript𝑋|\partial X_{\bullet}|. Thus it lifts to a 𝔖psubscript𝔖𝑝\mathfrak{S}_{p}-equivariant deformation retraction of an open neighbourhood of LXp|FLXpevaluated-atsuperscriptsubscript𝐿𝑋𝑝𝐹superscriptsubscript𝐿𝑋𝑝L_{X}^{\boxplus p}|_{F}\to L_{X}^{\boxplus p}, and descends to the quotient by the subcomplexes where some factor lies in |X|subscript𝑋|\partial X_{\bullet}|. As it is equivariant, it descends further to the 𝔖psubscript𝔖𝑝\mathfrak{S}_{p}-quotient. That is, it proves the claim for (M¯,M¯,L)¯𝑀¯𝑀𝐿(\overline{M},\partial\overline{M},L) replaced by (|X|,|X|,LX)subscript𝑋subscript𝑋subscript𝐿𝑋(|X_{\bullet}|,|\partial X_{\bullet}|,L_{X}); as the former data is a retract of the latter, the claim follows. ∎

Lemma 3.2.

\mathbb{R} is a flat 𝕊𝕊\mathbb{S}-module, in the sense that 𝕊\mathbb{R}\otimes_{\mathbb{S}}- preserves weak equivalences between left 𝕊𝕊\mathbb{S}-modules whose underlying objects are well-based.

Proof.

Recall that (n)=[(L+)n]𝔖n𝑛subscriptdelimited-[]superscriptsuperscript𝐿𝑛subscript𝔖𝑛\mathbb{R}(n)=[(L^{+})^{\wedge n}]_{\mathfrak{S}_{n}}. Define a filtration of \mathbb{R} by F0=𝕊subscript𝐹0𝕊F_{0}\mathbb{R}=\mathbb{S} and

Fp(n):=Fp1(n)Im((L+)p((LL)+)(np)/2(n)),assignsubscript𝐹𝑝𝑛subscript𝐹𝑝1𝑛Imsuperscriptsuperscript𝐿𝑝superscriptsuperscriptdirect-sum𝐿𝐿𝑛𝑝2𝑛F_{p}\mathbb{R}(n):=F_{p-1}\mathbb{R}(n)\cup\mathrm{Im}\big{(}(L^{+})^{\wedge p}\wedge((L\oplus L)^{+})^{\wedge(n-p)/2}\to\mathbb{R}(n)\big{)},

where the latter term is only taken when it makes sense: for np𝑛𝑝n-p even. This is a filtration by right 𝕊𝕊\mathbb{S}-modules. One checks that the diagram

Fp2(p)[p]𝕊tensor-productsubscript𝐹𝑝2𝑝delimited-[]𝑝𝕊{F_{p-2}\mathbb{R}(p)[p]\otimes\mathbb{S}}Fp1subscript𝐹𝑝1{F_{p-1}\mathbb{R}}(p)[p]𝕊tensor-product𝑝delimited-[]𝑝𝕊{\mathbb{R}(p)[p]\otimes\mathbb{S}}Fpsubscript𝐹𝑝{F_{p}\mathbb{R}}

is a pushout (in 𝖳𝗈𝗉superscriptsubscript𝖳𝗈𝗉\mathsf{Top}_{*}^{\mathbb{N}} and so in right 𝕊𝕊\mathbb{S}-modules), where the horizontal maps are induced by the 𝕊𝕊\mathbb{S}-module structure and the adjoints of the map inc:Fp2(p)Fp1(p):𝑖𝑛𝑐subscript𝐹𝑝2𝑝subscript𝐹𝑝1𝑝inc:F_{p-2}\mathbb{R}(p)\to F_{p-1}\mathbb{R}(p), and the map id:(p)Fp(p):𝑖𝑑𝑝subscript𝐹𝑝𝑝id:\mathbb{R}(p)\to F_{p}\mathbb{R}(p).

We prove by induction on p𝑝p that Fpsubscript𝐹𝑝F_{p}\mathbb{R} is a flat 𝕊𝕊\mathbb{S}-module in the indicated sense. As F0=𝕊subscript𝐹0𝕊F_{0}\mathbb{R}=\mathbb{S} these properties hold for p=0𝑝0p=0. For 𝕄𝕄\mathbb{M} a left 𝕊𝕊\mathbb{S}-module whose underlying object is well-based, applying 𝕊𝕄-\otimes_{\mathbb{S}}\mathbb{M} to the square above gives a pushout square

(3.1) Fp2(p)[p]𝕄tensor-productsubscript𝐹𝑝2𝑝delimited-[]𝑝𝕄{F_{p-2}\mathbb{R}(p)[p]\otimes\mathbb{M}}Fp1𝕊𝕄subscripttensor-product𝕊subscript𝐹𝑝1𝕄{F_{p-1}\mathbb{R}\otimes_{\mathbb{S}}\mathbb{M}}(p)[p]𝕄tensor-product𝑝delimited-[]𝑝𝕄{\mathbb{R}(p)[p]\otimes\mathbb{M}}Fp𝕊𝕄.subscripttensor-product𝕊subscript𝐹𝑝𝕄{F_{p}\mathbb{R}\otimes_{\mathbb{S}}\mathbb{M}.}

The map Fp2(p)(p)subscript𝐹𝑝2𝑝𝑝F_{p-2}\mathbb{R}(p)\to\mathbb{R}(p) is the inclusion of the subspace of those p𝑝p-tuples of points in M𝑀M labelled by L𝐿L which do not have distinct M𝑀M coordinates, so is a closed cofibration from a well-based space by the second part of Lemma 3.1. As 𝕄𝕄\mathbb{M} is assumed well-based, the left-hand vertical map in (3.1) is a closed cofibration in each grading, and so this square is also a homotopy pushout. A weak equivalence f:𝕄𝕄:𝑓𝕄similar-tosuperscript𝕄f:\mathbb{M}\overset{\sim}{\to}\mathbb{M}^{\prime} then induces a map of homotopy pushout squares which is a weak equivalence on all but the bottom right corner, by inductive assumption, so also induces a weak equivalence on this corner.

Thus each Fpsubscript𝐹𝑝F_{p}\mathbb{R} is flat in the indicated sense, so \mathbb{R} is too because Fpsubscript𝐹𝑝F_{p}\mathbb{R}\to\mathbb{R} is an isomorphism when evaluated on n<p𝑛𝑝n<p, so Fp𝕊𝕄𝕊𝕄subscripttensor-product𝕊subscript𝐹𝑝𝕄subscripttensor-product𝕊𝕄F_{p}\mathbb{R}\otimes_{\mathbb{S}}\mathbb{M}\to\mathbb{R}\otimes_{\mathbb{S}}\mathbb{M} is too. ∎

Remark 3.3.

In the case that L𝐿L is the 00-dimensional vector bundle, the filtration stage Fp(n)subscript𝐹𝑝𝑛F_{p}\mathbb{R}(n) consists of those elements in the n𝑛nth based symmetric power [(M+)n]𝔖nsubscriptdelimited-[]superscriptsuperscript𝑀𝑛subscript𝔖𝑛[(M^{+})^{\wedge n}]_{\mathfrak{S}_{n}} containing at most p𝑝p unrepeated elements. Up to reindexing, this is the same as the filtration used by Arnol’d [Arn70] and by Segal [Seg79].

Lemma 3.4.

The induced map 𝕊S0[0](M;L)subscripttensor-product𝕊superscript𝑆0delimited-[]0𝑀𝐿\mathbb{R}\otimes_{\mathbb{S}}S^{0}[0]\to\mathbb{C}(M;L) is an isomorphism.

Proof.

By definition of the relative tensor product there is a coequaliser diagram

𝕊tensor-product𝕊{\mathbb{R}\otimes\mathbb{S}}{\mathbb{R}}𝕊S0[0]subscripttensor-product𝕊superscript𝑆0delimited-[]0{\mathbb{R}\otimes_{\mathbb{S}}S^{0}[0]}α𝛼\scriptstyle{\alpha}β𝛽\scriptstyle{\beta}

in 𝖳𝗈𝗉superscriptsubscript𝖳𝗈𝗉\mathsf{Top}_{*}^{\mathbb{N}}, where α𝛼\alpha is given by the 𝕊𝕊\mathbb{S}-module structure on \mathbb{R}, and β𝛽\beta is induced by the augmentation ϵ:𝕊S0[0]:italic-ϵ𝕊superscript𝑆0delimited-[]0\epsilon:\mathbb{S}\to S^{0}[0]. The image of ker(ϵ)(n)(n)=[(L+)n]𝔖ntensor-productkeritalic-ϵ𝑛𝑛subscriptdelimited-[]superscriptsuperscript𝐿𝑛subscript𝔖𝑛\mathbb{R}\otimes\mathrm{ker}(\epsilon)(n)\to\mathbb{R}(n)=[(L^{+})^{\wedge n}]_{\mathfrak{S}_{n}} is precisely the image of (L+)n2(LL)+[(L+)n]𝔖nsuperscriptsuperscript𝐿𝑛2superscriptdirect-sum𝐿𝐿subscriptdelimited-[]superscriptsuperscript𝐿𝑛subscript𝔖𝑛(L^{+})^{\wedge n-2}\wedge(L\oplus L)^{+}\to[(L^{+})^{\wedge n}]_{\mathfrak{S}_{n}}, whose cofibre is by definition (M;L)𝑀𝐿\mathbb{C}(M;L). ∎

Proof of Theorem 1.2.

Apply Lemma 3.2 to the weak equivalence B(𝕊,𝕊,S0[0])S0[0]𝐵𝕊𝕊superscript𝑆0delimited-[]0similar-tosuperscript𝑆0delimited-[]0B(\mathbb{S},\mathbb{S},S^{0}[0])\overset{\sim}{\to}S^{0}[0], giving an equivalence B(,𝕊,S0[0])𝕊S0[0]subscripttensor-product𝕊𝐵𝕊superscript𝑆0delimited-[]0similar-tosuperscript𝑆0delimited-[]0B(\mathbb{R},\mathbb{S},S^{0}[0])\overset{\sim}{\to}\mathbb{R}\otimes_{\mathbb{S}}S^{0}[0], and the latter is isomorphic to (M)𝑀\mathbb{C}(M) by Lemma 3.4. ∎

Remark 3.5.

It is possible to fool oneself into thinking that the above argument can be adapted to the case of ordered configuration spaces, considered in the category of symmetric sequences of pointed spaces, in order to prove a statement analogous to the equivalence (1.2) in this category. Unfortunately, that statement is false. One can verify this directly in the case M=𝑀M=* with trivial 0-dimensional Euclidean bundle, in grading 3. If there is an analogue for ordered configuration spaces, its statement must be more complicated.

Appendix A Homological densities
by Quoc P. Ho and Oscar Randal-Williams

A.1. Spaces of 0-cycles

It is easy to generalise Theorem 1.2 to the following variant of configuration spaces, called “spaces of 0-cycles” by Farb–Wolfson–Wood [FWW19]. Let m,k1𝑚𝑘1m,k\geq 1, and for n1,n2,,nmsubscript𝑛1subscript𝑛2subscript𝑛𝑚n_{1},n_{2},\ldots,n_{m}\in\mathbb{N} let

Zn1,,nmk(M)Symn1,,nm(M):=[Mn1]𝔖n1×[Mn2]𝔖n2××[Mnm]𝔖nmsubscriptsuperscript𝑍𝑘subscript𝑛1subscript𝑛𝑚𝑀subscriptSymsubscript𝑛1subscript𝑛𝑚𝑀assignsubscriptdelimited-[]superscript𝑀subscript𝑛1subscript𝔖subscript𝑛1subscriptdelimited-[]superscript𝑀subscript𝑛2subscript𝔖subscript𝑛2subscriptdelimited-[]superscript𝑀subscript𝑛𝑚subscript𝔖subscript𝑛𝑚Z^{k}_{n_{1},\ldots,n_{m}}(M)\subset\mathrm{Sym}_{n_{1},\ldots,n_{m}}(M):=[M^{n_{1}}]_{\mathfrak{S}_{n_{1}}}\times[M^{n_{2}}]_{\mathfrak{S}_{n_{2}}}\times\cdots\times[M^{n_{m}}]_{\mathfrak{S}_{n_{m}}}

be the open subspace of those ({x11,,xn11},{x12,,xn22},,{x1m,,xnmm})superscriptsubscript𝑥11superscriptsubscript𝑥subscript𝑛11superscriptsubscript𝑥12superscriptsubscript𝑥subscript𝑛22superscriptsubscript𝑥1𝑚superscriptsubscript𝑥subscript𝑛𝑚𝑚(\{x_{1}^{1},\ldots,x_{n_{1}}^{1}\},\{x_{1}^{2},\ldots,x_{n_{2}}^{2}\},\ldots,\{x_{1}^{m},\ldots,x_{n_{m}}^{m}\}) such that no xjisubscriptsuperscript𝑥𝑖𝑗x^{i}_{j} has multiplicity kabsent𝑘\geq k in all of these m𝑚m multisets. That is, Zn1,,nmk(M)subscriptsuperscript𝑍𝑘subscript𝑛1subscript𝑛𝑚𝑀Z^{k}_{n_{1},\ldots,n_{m}}(M) is the configuration space of particles of m𝑚m different colours, nisubscript𝑛𝑖n_{i} having colour i𝑖i, which may collide except that no point of M𝑀M may carry kabsent𝑘\geq k points of every colour. The 1-point compactifications Zn1,,nm(M)+subscript𝑍subscript𝑛1subscript𝑛𝑚superscript𝑀Z_{n_{1},\ldots,n_{m}}(M)^{+} again have a composition product

Zn1,,nmk(M)+Zn1,,nmk(M)+Zn1+n1,,nm+nmk(M)+,subscriptsuperscript𝑍𝑘subscript𝑛1subscript𝑛𝑚superscript𝑀subscriptsuperscript𝑍𝑘subscriptsuperscript𝑛1subscriptsuperscript𝑛𝑚superscript𝑀subscriptsuperscript𝑍𝑘subscript𝑛1subscriptsuperscript𝑛1subscript𝑛𝑚subscriptsuperscript𝑛𝑚superscript𝑀Z^{k}_{n_{1},\ldots,n_{m}}(M)^{+}\wedge Z^{k}_{n^{\prime}_{1},\ldots,n^{\prime}_{m}}(M)^{+}\longrightarrow Z^{k}_{n_{1}+n^{\prime}_{1},\ldots,n_{m}+n^{\prime}_{m}}(M)^{+},

giving a commutative monoid m,k(M)superscript𝑚𝑘𝑀\mathbb{Z}^{m,k}(M) in msuperscript𝑚\mathbb{N}^{m}-graded pointed spaces. Just as before, we can introduce labels in a vector bundle LM𝐿𝑀L\to M, giving Zn1,,nmk(M;L)subscriptsuperscript𝑍𝑘subscript𝑛1subscript𝑛𝑚𝑀𝐿Z^{k}_{n_{1},\ldots,n_{m}}(M;L) and m,k(M;L)superscript𝑚𝑘𝑀𝐿\mathbb{Z}^{m,k}(M;L). Writing 1i=(0,,0,1,0,0)msubscript1𝑖00100superscript𝑚1_{i}=(0,\ldots,0,1,0\ldots,0)\in\mathbb{N}^{m} with the 1 in the i𝑖ith position, there is a pushout square

(A.1) 𝐂𝐨𝐦([(Lmk)+]𝔖km[k,k,,k])𝐂𝐨𝐦subscriptdelimited-[]superscriptsuperscript𝐿direct-sum𝑚𝑘superscriptsubscript𝔖𝑘𝑚𝑘𝑘𝑘{\mathbf{Com}([(L^{\oplus mk})^{+}]_{\mathfrak{S}_{k}^{m}}[k,k,\ldots,k])}S0[0,,0]superscript𝑆000{S^{0}[0,\ldots,0]}𝐂𝐨𝐦(i=1mL+[1i])𝐂𝐨𝐦superscriptsubscript𝑖1𝑚superscript𝐿delimited-[]subscript1𝑖{{\displaystyle\mathbf{Com}(\bigvee_{i=1}^{m}L^{+}[1_{i}])}}m,k(M;L)superscript𝑚𝑘𝑀𝐿{\mathbb{Z}^{m,k}(M;L)}ϵitalic-ϵ\scriptstyle{\epsilon}ΔΔ\scriptstyle{\Delta}

of unital commutative monoids in 𝖳𝗈𝗉msuperscriptsubscript𝖳𝗈𝗉superscript𝑚\mathsf{Top}_{*}^{\mathbb{N}^{m}}, where ΔΔ\Delta is now induced by the inclusion [(Lmk)+]𝔖km[(L+)k]𝔖k[(L+)k]𝔖k=𝐂𝐨𝐦(i=1mL+[1i])(k,,k)subscriptdelimited-[]superscriptsuperscript𝐿direct-sum𝑚𝑘superscriptsubscript𝔖𝑘𝑚subscriptdelimited-[]superscriptsuperscript𝐿𝑘subscript𝔖𝑘subscriptdelimited-[]superscriptsuperscript𝐿𝑘subscript𝔖𝑘𝐂𝐨𝐦superscriptsubscript𝑖1𝑚superscript𝐿delimited-[]subscript1𝑖𝑘𝑘[(L^{\oplus mk})^{+}]_{\mathfrak{S}_{k}^{m}}\to[(L^{+})^{\wedge k}]_{\mathfrak{S}_{k}}\wedge\cdots\wedge[(L^{+})^{\wedge k}]_{\mathfrak{S}_{k}}=\mathbf{Com}(\bigvee_{i=1}^{m}L^{+}[1_{i}])(k,\ldots,k). The same argument as Theorem 1.2 shows that there is an equivalence

(A.2) 𝐂𝐨𝐦(i=1mL+[1i])𝐂𝐨𝐦([(Lmk)+]𝔖km[k,,k])𝕃S0[0,,0]m,k(M;L).subscriptsuperscripttensor-product𝕃𝐂𝐨𝐦subscriptdelimited-[]superscriptsuperscript𝐿direct-sum𝑚𝑘superscriptsubscript𝔖𝑘𝑚𝑘𝑘𝐂𝐨𝐦superscriptsubscript𝑖1𝑚superscript𝐿delimited-[]subscript1𝑖superscript𝑆000similar-tosuperscript𝑚𝑘𝑀𝐿\mathbf{Com}(\vee_{i=1}^{m}L^{+}[1_{i}])\otimes^{\mathbb{L}}_{\mathbf{Com}([(L^{\oplus mk})^{+}]_{\mathfrak{S}_{k}^{m}}[k,\ldots,k])}S^{0}[0,\ldots,0]\overset{\sim}{\longrightarrow}\mathbb{Z}^{m,k}(M;L).

A.2. Revisiting homological densities

This can be used to revisit the work of Farb–Wolfson–Wood [FWW19] and Ho [Ho21] on homological densities, and in particular to explain coincidences of homological densities at the level of topology rather than algebra, as proposed in [Ho21, 1.5.1].

The spaces Zn1,,nmk(M;L)subscriptsuperscript𝑍𝑘subscript𝑛1subscript𝑛𝑚𝑀𝐿Z^{k}_{n_{1},\ldots,n_{m}}(M;L) are \mathbb{Q}-homology manifolds, being open subspaces of a product of coarse moduli spaces [Ln]𝔖nsubscriptdelimited-[]superscript𝐿𝑛subscript𝔖𝑛[L^{n}]_{\mathfrak{S}_{n}} of the orbifolds Ln//𝔖nL^{n}/\!\!/\mathfrak{S}_{n}. As before, we suppose M𝑀M is d𝑑d-dimensional and take L=W𝐿𝑊L=W to be given by the sum of the orientation line of M𝑀M plus (d1)𝑑1(d-1) trivial lines: then the Zn1,,nmk(M;W)subscriptsuperscript𝑍𝑘subscript𝑛1subscript𝑛𝑚𝑀𝑊Z^{k}_{n_{1},\ldots,n_{m}}(M;W) are orientable \mathbb{Q}-homology manifolds, of dimension 2dni2𝑑subscript𝑛𝑖2d\cdot\sum n_{i}. Again they are vector bundles over Zn1,,nmk(M)subscriptsuperscript𝑍𝑘subscript𝑛1subscript𝑛𝑚𝑀Z^{k}_{n_{1},\ldots,n_{m}}(M), so Poincaré duality gives

H(Zn1,,nmk(M))H(Zn1,,nmk(M;W))H~2dni(Zn1,,nmk(M;W)+).superscript𝐻subscriptsuperscript𝑍𝑘subscript𝑛1subscript𝑛𝑚𝑀superscript𝐻subscriptsuperscript𝑍𝑘subscript𝑛1subscript𝑛𝑚𝑀𝑊subscript~𝐻limit-from2𝑑subscript𝑛𝑖subscriptsuperscript𝑍𝑘subscript𝑛1subscript𝑛𝑚superscript𝑀𝑊H^{*}(Z^{k}_{n_{1},\ldots,n_{m}}(M))\cong H^{*}(Z^{k}_{n_{1},\ldots,n_{m}}(M;W))\cong\widetilde{H}_{2d\sum n_{i}-*}(Z^{k}_{n_{1},\ldots,n_{m}}(M;W)^{+}).

On the other hand, the bar construction formula above together with the argument of Section 2.2 identifies the multigraded vector space H,(m,k(M;W))subscript𝐻superscript𝑚𝑘𝑀𝑊H_{*,*}(\mathbb{Z}^{m,k}(M;W)) with

TorS(H~([(Wmk)+]𝔖km)[k,,k])(S(i=1mH~(W+)[1i]),[0,,0]).superscriptsubscriptTorsuperscript𝑆subscript~𝐻subscriptdelimited-[]superscriptsuperscript𝑊direct-sum𝑚𝑘superscriptsubscript𝔖𝑘𝑚𝑘𝑘superscript𝑆superscriptsubscriptdirect-sum𝑖1𝑚subscript~𝐻superscript𝑊delimited-[]subscript1𝑖00\mathrm{Tor}_{*}^{S^{*}(\widetilde{H}_{*}([(W^{\oplus mk})^{+}]_{\mathfrak{S}_{k}^{m}})[k,\ldots,k])}(S^{*}(\bigoplus_{i=1}^{m}\widetilde{H}_{*}(W^{+})[1_{i}]),\mathbb{Q}[0,\ldots,0]).

A.2.1. Odd-dimensional manifolds

As in Section 2.2 we have H~([(Wmk)+]𝔖km)[ΣdmkH~(M+)]𝔖kmsubscript~𝐻subscriptdelimited-[]superscriptsuperscript𝑊direct-sum𝑚𝑘superscriptsubscript𝔖𝑘𝑚subscriptdelimited-[]superscriptΣ𝑑𝑚𝑘subscript~𝐻superscript𝑀superscriptsubscript𝔖𝑘𝑚\widetilde{H}_{*}([(W^{\oplus mk})^{+}]_{\mathfrak{S}_{k}^{m}})\cong[\Sigma^{dmk}\widetilde{H}_{*}(M^{+})]_{\mathfrak{S}_{k}^{m}} by the Thom isomorphism. If d𝑑d is odd then the permutation group 𝔖kmsuperscriptsubscript𝔖𝑘𝑚\mathfrak{S}_{k}^{m} acts on the Thom class via 𝔖km𝔖mksign×superscriptsubscript𝔖𝑘𝑚subscript𝔖𝑚𝑘𝑠𝑖𝑔𝑛superscript\mathfrak{S}_{k}^{m}\leq\mathfrak{S}_{mk}\overset{sign}{\to}\mathbb{Z}^{\times}, so acts nontrivially if k2𝑘2k\geq 2 and trivially if k=1𝑘1k=1. If k2𝑘2k\geq 2 this means that H~([(Wmk)+]𝔖km)=0subscript~𝐻subscriptdelimited-[]superscriptsuperscript𝑊direct-sum𝑚𝑘superscriptsubscript𝔖𝑘𝑚0\widetilde{H}_{*}([(W^{\oplus mk})^{+}]_{\mathfrak{S}_{k}^{m}})=0, showing that

H,(𝐂𝐨𝐦(i=1mW+[1i]))H,(m,k(M;W))subscript𝐻𝐂𝐨𝐦superscriptsubscript𝑖1𝑚superscript𝑊delimited-[]subscript1𝑖similar-tosubscript𝐻superscript𝑚𝑘𝑀𝑊H_{*,*}(\mathbf{Com}(\vee_{i=1}^{m}W^{+}[1_{i}]))\overset{\sim}{\longrightarrow}H_{*,*}(\mathbb{Z}^{m,k}(M;W))

in this case. Using Poincaré duality on both sides gives [FWW19, Theorem 1.4], except that that theorem is erroneously claimed for all k1𝑘1k\geq 1. We will return to the case k=1𝑘1k=1 below.

A.2.2. Even-dimensional manifolds

If d𝑑d is even then 𝔖kmsuperscriptsubscript𝔖𝑘𝑚\mathfrak{S}_{k}^{m} acts trivially on ΣdmkH~(M+)superscriptΣ𝑑𝑚𝑘subscript~𝐻superscript𝑀\Sigma^{dmk}\widetilde{H}_{*}(M^{+}), and using the Thom isomorphism to identify H~(W+)ΣdH~(M+)subscript~𝐻superscript𝑊superscriptΣ𝑑subscript~𝐻superscript𝑀\widetilde{H}_{*}(W^{+})\cong\Sigma^{d}\widetilde{H}_{*}(M^{+}) too, the Koszul complex for computing the TorTor\mathrm{Tor}-groups above is

(S(i=1mΣdH~(M+;w1)[1i]Σdmk+1H~(M+;(w1)mk)[k,,k]),).superscript𝑆direct-sumsuperscriptsubscriptdirect-sum𝑖1𝑚superscriptΣ𝑑subscript~𝐻superscript𝑀superscriptsubscript𝑤1delimited-[]subscript1𝑖superscriptΣ𝑑𝑚𝑘1subscript~𝐻superscript𝑀superscriptsuperscriptsubscript𝑤1tensor-productabsent𝑚𝑘𝑘𝑘(S^{*}\big{(}\bigoplus_{i=1}^{m}\Sigma^{d}\widetilde{H}_{*}(M^{+};\mathbb{Q}^{w_{1}})[1_{i}]\oplus\Sigma^{dmk+1}\widetilde{H}_{*}(M^{+};(\mathbb{Q}^{w_{1}})^{\otimes mk})[k,\ldots,k]\big{)},\partial).

The differential \partial is induced by the map

ΣdmkH~(M+;(w1)mk)Sk(ΣdH~(M+;w1))Sk(ΣdH~(M+;w1))superscriptΣ𝑑𝑚𝑘subscript~𝐻superscript𝑀superscriptsuperscriptsubscript𝑤1tensor-productabsent𝑚𝑘tensor-productsuperscript𝑆𝑘superscriptΣ𝑑subscript~𝐻superscript𝑀superscriptsubscript𝑤1superscript𝑆𝑘superscriptΣ𝑑subscript~𝐻superscript𝑀superscriptsubscript𝑤1\Sigma^{dmk}\widetilde{H}_{*}(M^{+};(\mathbb{Q}^{w_{1}})^{\otimes mk})\to S^{k}(\Sigma^{d}\widetilde{H}_{*}(M^{+};\mathbb{Q}^{w_{1}}))\otimes\cdots\otimes S^{k}(\Sigma^{d}\widetilde{H}_{*}(M^{+};\mathbb{Q}^{w_{1}}))

obtained by linearly dualising the cup product map

(A.3) Hc(M;w1)mkHc(M;(w1)mk),subscriptsuperscript𝐻𝑐superscript𝑀superscriptsubscript𝑤1tensor-productabsent𝑚𝑘subscriptsuperscript𝐻𝑐𝑀superscriptsuperscriptsubscript𝑤1tensor-productabsent𝑚𝑘H^{*}_{c}(M;\mathbb{Q}^{w_{1}})^{\otimes mk}\longrightarrow H^{*}_{c}(M;(\mathbb{Q}^{w_{1}})^{\otimes mk}),

and so is trivial if (and only if) all mk𝑚𝑘mk-fold cup products of (w1subscript𝑤1w_{1}-twisted) compactly-supported cohomology classes on M𝑀M vanish.

When this cup product map is trivial, so \partial is trivial, the above just gives a formula for H,(m,k(M;W))subscript𝐻superscript𝑚𝑘𝑀𝑊H_{*,*}(\mathbb{Z}^{m,k}(M;W)). Using Poincaré duality, and reindexing, to express this in terms of H(Zn1,,nmk(M))superscript𝐻subscriptsuperscript𝑍𝑘subscript𝑛1subscript𝑛𝑚𝑀H^{*}(Z^{k}_{n_{1},\ldots,n_{m}}(M)) and H(Symn1,,nm(M))superscript𝐻subscriptSymsubscript𝑛1subscript𝑛𝑚𝑀H^{*}(\mathrm{Sym}_{n_{1},\ldots,n_{m}}(M)) we obtain an identity of multigraded vector spaces

H(Zk(M))H(Sym(M))S(Σd(mk1)1H(M;(w1)mk1)[k,,k]).superscript𝐻subscriptsuperscript𝑍𝑘𝑀tensor-productsuperscript𝐻subscriptSym𝑀superscript𝑆superscriptΣ𝑑𝑚𝑘11superscript𝐻𝑀superscriptsuperscriptsubscript𝑤1tensor-productabsent𝑚𝑘1𝑘𝑘H^{*}(Z^{k}_{\bullet}(M))\cong H^{*}(\mathrm{Sym}_{\bullet}(M))\otimes S^{*}(\Sigma^{d(mk-1)-1}{H}^{*}(M;(\mathbb{Q}^{w_{1}})^{\otimes mk-1})[k,\ldots,k]).

There are stabilisation maps σi:H(Zn1,,nmk(M))H(Zn1,,ni+1,,nmk(M)):subscript𝜎𝑖superscript𝐻subscriptsuperscript𝑍𝑘subscript𝑛1subscript𝑛𝑚𝑀superscript𝐻subscriptsuperscript𝑍𝑘subscript𝑛1subscript𝑛𝑖1subscript𝑛𝑚𝑀\sigma_{i}:H^{*}(Z^{k}_{n_{1},\ldots,n_{m}}(M))\to H^{*}(Z^{k}_{n_{1},\ldots,n_{i}+1,\ldots,n_{m}}(M)) analogous to those constructed in Section 2.3, similarly for H(Symn1,,nm(M))superscript𝐻subscriptSymsubscript𝑛1subscript𝑛𝑚𝑀H^{*}(\mathrm{Sym}_{n_{1},\ldots,n_{m}}(M)), and both stabilise as njsubscript𝑛𝑗n_{j}\to\infty, just as in Section 2.3: this recovers [FWW19, Theorem 1.7]. We may take the colimit of all these stabilisations to obtain

H(Z,,k(M))H(Sym,,(M))S(Σd(mk1)1H(M;(w1)mk1)).superscript𝐻subscriptsuperscript𝑍𝑘𝑀tensor-productsuperscript𝐻subscriptSym𝑀superscript𝑆superscriptΣ𝑑𝑚𝑘11superscript𝐻𝑀superscriptsuperscriptsubscript𝑤1tensor-productabsent𝑚𝑘1H^{*}(Z^{k}_{\infty,\ldots,\infty}(M))\cong H^{*}(\mathrm{Sym}_{\infty,\ldots,\infty}(M))\otimes S^{*}(\Sigma^{d(mk-1)-1}{H}^{*}(M;(\mathbb{Q}^{w_{1}})^{\otimes mk-1})).

Writing PZm,k(t)subscript𝑃superscript𝑍𝑚𝑘𝑡P_{Z^{m,k}}(t) and PSymm(t)subscript𝑃𝑆𝑦superscript𝑚𝑚𝑡P_{Sym^{m}}(t) for the Poincaré series of H(Z,,k(M))superscript𝐻subscriptsuperscript𝑍𝑘𝑀H^{*}(Z^{k}_{\infty,\ldots,\infty}(M)) and H(Sym,,(M))superscript𝐻subscriptSym𝑀H^{*}(\mathrm{Sym}_{\infty,\ldots,\infty}(M)) respectively, this discussion identifies the homological density PZm,k(t)/PSymm(t)subscript𝑃superscript𝑍𝑚𝑘𝑡subscript𝑃𝑆𝑦superscript𝑚𝑚𝑡{P_{Z^{m,k}}(t)}/{P_{Sym^{m}}(t)} with the Poincaré series of S(Σd(mk1)1H(M;(w1)mk1))superscript𝑆superscriptΣ𝑑𝑚𝑘11superscript𝐻𝑀superscriptsuperscriptsubscript𝑤1tensor-productabsent𝑚𝑘1S^{*}(\Sigma^{d(mk-1)-1}{H}^{*}(M;(\mathbb{Q}^{w_{1}})^{\otimes mk-1})). This visibly only depends on the product mk𝑚𝑘mk, giving “coincidences between homological densities”: this recovers [FWW19, Theorem 1.2]; in fact it also recovers the stronger Theorem 3.6 of that paper.

A.2.3. Odd-dimensional manifolds, k=1𝑘1k=1

Just as in the even-dimensional case, if the cup product map (A.3) is zero then one gets an explicit description of H(Z(M))superscript𝐻subscript𝑍𝑀H^{*}(Z_{\bullet}(M)), and the homological density is given by the Poincaré series of the graded vector space S(Σd(m1)1H(M;(w1)m1))superscript𝑆superscriptΣ𝑑𝑚11superscript𝐻𝑀superscriptsuperscriptsubscript𝑤1tensor-productabsent𝑚1S^{*}(\Sigma^{d(m-1)-1}{H}^{*}(M;(\mathbb{Q}^{w_{1}})^{\otimes m-1})). It follows from Section A.2.1 that the homological density is 111 for k>1𝑘1k>1, so for odd-dimensional manifolds it is not true that the homological density depends only on mk𝑚𝑘mk.

A.2.4. Euler characteristic

If the cup product map (A.3) is not zero, and either d𝑑d is even or d𝑑d is odd and k=1𝑘1k=1, then there is instead a nontrivial differential on the multigraded vector space

H(Sym(M))S(Σd(mk1)1H(M;(w1)mk1)[k,,k]),tensor-productsuperscript𝐻subscriptSym𝑀superscript𝑆superscriptΣ𝑑𝑚𝑘11superscript𝐻𝑀superscriptsuperscriptsubscript𝑤1tensor-productabsent𝑚𝑘1𝑘𝑘H^{*}(\mathrm{Sym}_{\bullet}(M))\otimes S^{*}(\Sigma^{d(mk-1)-1}{H}^{*}(M;(\mathbb{Q}^{w_{1}})^{\otimes mk-1})[k,\ldots,k]),

of degree (+1,0)10(+1,0), whose homology is H(Z,,(M))superscript𝐻subscript𝑍𝑀H^{*}(Z_{*,\ldots,*}(M)). Then one would not expect PZ(t)PSym(t)subscript𝑃𝑍𝑡subscript𝑃𝑆𝑦𝑚𝑡\frac{P_{Z}(t)}{P_{Sym}(t)} to agree with the Poincaré series of S(Σd(mk1)1H(M;(w1)mk1))superscript𝑆superscriptΣ𝑑𝑚𝑘11superscript𝐻𝑀superscriptsuperscriptsubscript𝑤1tensor-productabsent𝑚𝑘1S^{*}(\Sigma^{d(mk-1)-1}{H}^{*}(M;(\mathbb{Q}^{w_{1}})^{\otimes mk-1})), and indeed it does not [FWW19, Remark 1.6]. However, as Euler characteristic commutes with taking homology we have the identity

n1,,nm0χ(Zn1,,nm(M))s1n1smnm=(i=1m(1si))χ(M)(1(s1sm)k)χ(M,(w1)mk1)subscriptsubscript𝑛1subscript𝑛𝑚0𝜒subscript𝑍subscript𝑛1subscript𝑛𝑚𝑀superscriptsubscript𝑠1subscript𝑛1superscriptsubscript𝑠𝑚subscript𝑛𝑚superscriptsuperscriptsubscriptproduct𝑖1𝑚1subscript𝑠𝑖𝜒𝑀superscript1superscriptsubscript𝑠1subscript𝑠𝑚𝑘𝜒𝑀superscriptsuperscriptsubscript𝑤1tensor-productabsent𝑚𝑘1\sum_{n_{1},\ldots,n_{m}\geq 0}\chi(Z_{n_{1},\ldots,n_{m}}(M))s_{1}^{n_{1}}\cdots s_{m}^{n_{m}}=\big{(}\prod_{i=1}^{m}(1-s_{i})\big{)}^{-\chi(M)}\cdot(1-(s_{1}\cdots s_{m})^{k})^{\chi(M,(\mathbb{Q}^{w_{1}})^{\otimes mk-1})}

in [[s1,,sm]]delimited-[]subscript𝑠1subscript𝑠𝑚\mathbb{Z}[[s_{1},\ldots,s_{m}]]. The left-hand factor is n1,,nm0χ(Symn1,,nm(M))s1n1smnmsubscriptsubscript𝑛1subscript𝑛𝑚0𝜒subscriptSymsubscript𝑛1subscript𝑛𝑚𝑀superscriptsubscript𝑠1subscript𝑛1superscriptsubscript𝑠𝑚subscript𝑛𝑚\sum_{n_{1},\ldots,n_{m}\geq 0}\chi(\mathrm{Sym}_{n_{1},\ldots,n_{m}}(M))s_{1}^{n_{1}}\cdots s_{m}^{n_{m}}. This recovers [FWW19, Theorem 1.9 1.].

A.3. Spectral densities

The construction of homological densities can be promoted to the level of spectra, addressing [Ho21, 1.5.1], as follows. Let us assume that M𝑀M is even-dimensional and orientable: then we can dispense with twisting by the vector bundle WM𝑊𝑀W\to M. We consider m,k(M)superscript𝑚𝑘𝑀\mathbb{Z}^{m,k}(M) with its msuperscript𝑚\mathbb{N}^{m}-grading reduced to an \mathbb{N}-grading via sum:m:sumsuperscript𝑚\text{sum}:\mathbb{N}^{m}\to\mathbb{N}. Collapsing the complement of a small neighbourhood of a point in M𝑀M gives a map M+Sdsuperscript𝑀superscript𝑆𝑑M^{+}\to S^{d}, inducing a map of commutative monoids

𝐂𝐨𝐦(i=1mM+[1])𝐂𝐨𝐦(Sd[1]).𝐂𝐨𝐦superscriptsubscript𝑖1𝑚superscript𝑀delimited-[]1𝐂𝐨𝐦superscript𝑆𝑑delimited-[]1\mathbf{Com}(\bigvee_{i=1}^{m}M^{+}[1])\longrightarrow\mathbf{Com}(S^{d}[1]).

If X𝑋X is a left 𝐂𝐨𝐦(Sd[1])𝐂𝐨𝐦superscript𝑆𝑑delimited-[]1\mathbf{Com}(S^{d}[1])-module, it is equipped with maps SdX(n)X(n+1)superscript𝑆𝑑𝑋𝑛𝑋𝑛1S^{d}\wedge X(n)\to X(n+1) and so we can define the spectrum X¯:=hocolimnSndΣX(n)assign¯𝑋subscripthocolim𝑛superscript𝑆𝑛𝑑superscriptΣ𝑋𝑛\overline{X}:=\operatorname*{hocolim}_{n\to\infty}S^{-nd}\wedge\Sigma^{\infty}X(n). Using these two constructions we may therefore form the spectrum

Δm,k:=𝐂𝐨𝐦(Sd[1])𝐂𝐨𝐦(i=1mM+[1])𝕃m,k(M)¯.assignsuperscriptΔ𝑚𝑘¯subscriptsuperscripttensor-product𝕃𝐂𝐨𝐦superscriptsubscript𝑖1𝑚superscript𝑀delimited-[]1𝐂𝐨𝐦superscript𝑆𝑑delimited-[]1superscript𝑚𝑘𝑀\Delta^{m,k}:=\overline{\mathbf{Com}(S^{d}[1])\otimes^{\mathbb{L}}_{\mathbf{Com}(\bigvee_{i=1}^{m}M^{+}[1])}\mathbb{Z}^{m,k}(M)}.

By analogy with [Ho21, Section 7.5] we propose Δm,ksuperscriptΔ𝑚𝑘\Delta^{m,k} as a spectral form of the stable density of Zn1,,nmk(M)subscriptsuperscript𝑍𝑘subscript𝑛1subscript𝑛𝑚𝑀Z^{k}_{n_{1},\ldots,n_{m}}(M) in Symn1,,nm(M)subscriptSymsubscript𝑛1subscript𝑛𝑚𝑀\mathrm{Sym}_{n_{1},\ldots,n_{m}}(M). At the level of \mathbb{Q}-chains it recovers the construction from the proof of Theorem 7.5.1 of [Ho21]. We can prove the spectral form of that theorem analogously: as \mathbb{N}-graded objects, there is an evident map from (A.1) to the analogous square for 1,mk(M)superscript1𝑚𝑘𝑀\mathbb{Z}^{1,mk}(M) which induces a map of spectra Δm,kΔ1,mksuperscriptΔ𝑚𝑘superscriptΔ1𝑚𝑘\Delta^{m,k}\to\Delta^{1,mk}, and this is an equivalence by (A.2) as both are identified with 𝐂𝐨𝐦(Sd[1])𝐂𝐨𝐦(M+[mk])𝕃S0[0]¯¯subscriptsuperscripttensor-product𝕃𝐂𝐨𝐦superscript𝑀delimited-[]𝑚𝑘𝐂𝐨𝐦superscript𝑆𝑑delimited-[]1superscript𝑆0delimited-[]0\overline{\mathbf{Com}(S^{d}[1])\otimes^{\mathbb{L}}_{\mathbf{Com}(M^{+}[mk])}S^{0}[0]}.

This may be simplified for mk2𝑚𝑘2mk\geq 2 as follows. The map M+[(Sd)mk]𝔖mksuperscript𝑀subscriptdelimited-[]superscriptsuperscript𝑆𝑑𝑚𝑘subscript𝔖𝑚𝑘M^{+}\to[(S^{d})^{\wedge mk}]_{\mathfrak{S}_{mk}} with which the derived tensor product is formed factors over (Sd)mksuperscriptsuperscript𝑆𝑑𝑚𝑘(S^{d})^{\wedge mk} so is nullhomotopic when mk2𝑚𝑘2mk\geq 2, and so 𝐂𝐨𝐦(Sd[1])𝐂𝐨𝐦(M+[mk])𝕃S0[0]subscriptsuperscripttensor-product𝕃𝐂𝐨𝐦superscript𝑀delimited-[]𝑚𝑘𝐂𝐨𝐦superscript𝑆𝑑delimited-[]1superscript𝑆0delimited-[]0\mathbf{Com}(S^{d}[1])\otimes^{\mathbb{L}}_{\mathbf{Com}(M^{+}[mk])}S^{0}[0] is equivalent to 𝐂𝐨𝐦(Sd[1])(S0[0]𝐂𝐨𝐦(M+[mk])𝕃S0[0])tensor-product𝐂𝐨𝐦superscript𝑆𝑑delimited-[]1subscriptsuperscripttensor-product𝕃𝐂𝐨𝐦superscript𝑀delimited-[]𝑚𝑘superscript𝑆0delimited-[]0superscript𝑆0delimited-[]0\mathbf{Com}(S^{d}[1])\otimes(S^{0}[0]\otimes^{\mathbb{L}}_{\mathbf{Com}(M^{+}[mk])}S^{0}[0]) as a left 𝐂𝐨𝐦(Sd[1])𝐂𝐨𝐦superscript𝑆𝑑delimited-[]1\mathbf{Com}(S^{d}[1])-module. In this situation the ()¯¯\overline{(-)} construction gives

Δm,ksuperscriptΔ𝑚𝑘\displaystyle\Delta^{m,k} n0SndΣ(S0[0]𝐂𝐨𝐦(M+[mk])𝕃S0[0])(n)similar-to-or-equalsabsentsubscript𝑛0superscript𝑆𝑛𝑑superscriptΣsubscriptsuperscripttensor-product𝕃𝐂𝐨𝐦superscript𝑀delimited-[]𝑚𝑘superscript𝑆0delimited-[]0superscript𝑆0delimited-[]0𝑛\displaystyle\simeq\bigvee_{n\geq 0}S^{-nd}\wedge\Sigma^{\infty}(S^{0}[0]\otimes^{\mathbb{L}}_{\mathbf{Com}(M^{+}[mk])}S^{0}[0])(n)
n0SndΣ𝐂𝐨𝐦(S1M+[mk])(n).similar-to-or-equalsabsentsubscript𝑛0superscript𝑆𝑛𝑑superscriptΣ𝐂𝐨𝐦superscript𝑆1superscript𝑀delimited-[]𝑚𝑘𝑛\displaystyle\simeq\bigvee_{n\geq 0}S^{-nd}\wedge\Sigma^{\infty}\mathbf{Com}(S^{1}\wedge M^{+}[mk])(n).

References

  • [Arn70] V. I. Arnol’d, Certain topological invariants of algebraic functions, Trudy Moskov. Mat. Obšč. 21 (1970), 27–46.
  • [Ban] O. Banerjee, Stability in cohomology via the symmetric simplicial category, in preparation.
  • [Ban23] by same author, Filtration of cohomology via symmetric semisimplicial spaces, https://arxiv.org/abs/1909.00458v3, 2023.
  • [BCM93] C.-F. Bödigheimer, F. R. Cohen, and R. J. Milgram, Truncated symmetric products and configuration spaces, Math. Z. 214 (1993), no. 2, 179–216.
  • [Bia20] A. Bianchi, Splitting of the homology of the punctured mapping class group, J. Topol. 13 (2020), no. 3, 1230–1260.
  • [BM14] M. Bendersky and J. Miller, Localization and homological stability of configuration spaces, Q. J. Math. 65 (2014), no. 3, 807–815.
  • [BY21] B. Berceanu and M. Yameen, Strong and shifted stability for the cohomology of configuration spaces, Bull. Math. Soc. Sci. Math. Roumanie (N.S.) 64(112) (2021), no. 2, 159–191.
  • [Chu12] T. Church, Homological stability for configuration spaces of manifolds, Invent. Math. 188 (2012), no. 2, 465–504.
  • [CP15] F. Cantero and M. Palmer, On homological stability for configuration spaces on closed background manifolds, Doc. Math. 20 (2015), 753–805.
  • [FWW19] B. Farb, J. Wolfson, and M. M. Wood, Coincidences between homological densities, predicted by arithmetic, Adv. Math. 352 (2019), 670–716.
  • [Get99a] E. Getzler, The homology groups of some two-step nilpotent Lie algebras associated to symplectic vector spaces, https://arxiv.org/abs/math/9903147, 1999.
  • [Get99b] by same author, Resolving mixed Hodge modules on configuration spaces, Duke Math. J. 96 (1999), no. 1, 175–203.
  • [Ho20] Q. P. Ho, Higher representation stability for ordered configuration spaces and twisted commutative factorization algebras, https://arxiv.org/abs/2004.00252, 2020.
  • [Ho21] by same author, Homological stability and densities of generalized configuration spaces, Geom. Topol. 25 (2021), no. 2, 813–912.
  • [Kal98] S. Kallel, Divisor spaces on punctured Riemann surfaces, Trans. Amer. Math. Soc. 350 (1998), no. 1, 135–164.
  • [KM15] A. Kupers and J. Miller, Improved homological stability for configuration spaces after inverting 2, Homology Homotopy Appl. 17 (2015), no. 1, 255–266.
  • [KMT23] B. Knudsen, J. Miller, and P. Tosteson, Extremal stability for configuration spaces, Mathematische Annalen 386 (2023), no. 3, 1695–1716.
  • [Knu17] B. Knudsen, Betti numbers and stability for configuration spaces via factorization homology, Algebr. Geom. Topol. 17 (2017), no. 5, 3137–3187.
  • [Loo23] E. Looijenga, Torelli group action on the configuration space of a surface, J. Topol. Anal. 15 (2023), no. 1, 215–222.
  • [Mil69] R. J. Milgram, The homology of symmetric products, Trans. Amer. Math. Soc. 138 (1969), 251–265.
  • [Pet20] D. Petersen, Cohomology of generalized configuration spaces, Compos. Math. 156 (2020), no. 2, 251–298.
  • [RW13] O. Randal-Williams, Homological stability for unordered configuration spaces, Q. J. Math. 64 (2013), no. 1, 303–326.
  • [Seg79] G. Segal, The topology of spaces of rational functions, Acta Math. 143 (1979), no. 1-2, 39–72.
  • [Sta23a] A. Stavrou, Cohomology of configuration spaces of surfaces as mapping class group representations, Trans. Amer. Math. Soc. 376 (2023), no. 4, 2821–2852.
  • [Sta23b] by same author, Homology of configuration spaces of surfaces as mapping class group representations, Ph.D. thesis, University of Cambridge, 2023.
  • [Yam23] M. Yameen, A remark on extremal stability of configuration spaces, J. Pure Appl. Algebra 227 (2023), no. 1, Paper No. 107154, 5.