Configuration spaces as commutative monoids
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 be the interior of a connected compact manifold with boundary. The 1-point compactification of the space of unordered configurations in may be written as
(1.1) |
the quotient formed in pointed spaces. Not-necessarily-disjoint union of unordered configurations defines a superposition product
which is associative, commutative, and unital. This gives a unital commutative monoid object in the symmetric monoidal category of -graded pointed spaces:
The goal of this note is to explain and exploit this algebraic structure.
In the following, for a pointed space we write for the -graded pointed space which consists of in grading and the point in all other gradings, and write for the free unital commutative monoid on an object .
Theorem 1.1.
There is a pushout square
of unital commutative monoids in , where is the augmentation and is induced by the diagonal inclusion . Furthermore, this square is a homotopy pushout, i.e. there is an induced equivalence
That the square is a strict pushout of unital commutative monoids is elementary: it means identifying as the quotient of the based symmetric power monoid of 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 be a vector bundle, and let
These assemble in the same way to a unital commutative monoid object . (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
of unital commutative monoids in , where is the augmentation and is induced by the diagonal inclusion . Furthermore, this square is a homotopy pushout, i.e. there is an induced equivalence
This strictly generalises Theorem 1.1, which is the case where 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 , so the conclusion of Theorem 1.2 can equivalently be stated as an equivalence
(1.2) |
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 in terms of the compactly-supported -cohomology of and its cup-product map, which in particular quickly implies homological stability. As another application we will show that the action on of the group of proper homotopy self-equivalences of 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 : this description will also follow from [Ban] as explained in [Ban23, Remark 1.1]. As such, the purpose of this paper is
-
(i)
to give a space-level implementation/interpretation of Banerjee’s ideas in a specific case, in order to popularise them among topologists, and
-
(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 be -dimensional. The space is the 1-point compactification of the -dimensional manifold
This is a vector bundle over , but is a manifold itself and is orientable if and only if the manifold is orientable and even-dimensional. To arrange this, we can take the vector bundle given by the orientation line of plus trivial line bundles. Thus by Poincaré duality we have
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 . We do not pursue this in general here, but rather focus on the case , where a complete answer is possible, and reproduces a formula of Knudsen.
2.2. Revisiting Knudsen’s formula
For an -graded pointed space we write , and similarly for chains. Write for the free graded-commutative algebra on a homologically graded vector space , i.e. , where the Koszul sign rule is implemented. If is equipped with additional -grading, then this is inherited by (but there is no Koszul sign rule associated to the -grading, only to the homological grading).
We consider . There is a map and, using the Eilenberg–Zilber maps, it extends to a map of cdga’s
which is an equivalence (since the maps are rational homology isomorphisms). Similarly, there is an equivalence of cdga’s
Furthermore, one may choose formality equivalences
i.e. chain maps inducing the identity on homology, and hence obtain equivalences
of cdga’s. In -grading 2, the map induces a map
With these choices the square
need not commute, but does commute up to homotopy in the category of cdga’s because the two chain maps induce the same map on homology, namely , so are chain homotopic. The bar construction description then gives an identification
Recall that for a free graded-commutative algebra on a homologically graded vector space (perhaps equipped with a further -grading), there is a free resolution of the trivial left -module given by equipped with the differential given by 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 , and over taking homology commutes with the formation of symmetric powers. Applying this resolution to calculate the Tor groups above gives the complex
with differential given by for , and extended by the Leibniz rule. This can be simplified as follows. If is -dimensional then the Thom isomorphism gives , where is the orientation local system of . It also gives . The involution swapping the two factors acts as on the Thom class, so because the map is a rational equivalence we find
This lets us write the complex as
(2.1) |
where the differential is dual to the map induced by cup product, so following Knudsen we can recognise this complex as the Chevelley–Eilenberg complex for the bigraded Lie algebra . Thus
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 . Let us briefly review this from the point of view taken here.
There is a canonical element , and choosing a cycle representing this element provides a map
Multiplication by this element defines a map
which under Poincaré duality gives a map ; this can be checked to be the transfer map which sums over all ways of forgetting one of the points, see [Knu17, Section 5.2] [Sta23b, Section 2.6].
Writing for the mapping cone of left multiplication by , the discussion above shows that its homology is calculated by a complex
As is connected, if we assume that then the bigraded vector spaces and both vanish in bidegrees satisfying , and hence so does the free graded-commutative algebra on them. This translates to being surjective for and an isomorphism for . For the same considerations give surjectivity for and so on (a more careful analysis gives a slope 1 range in this case too, see [Knu17, Proof of Theorem 1.3]).
2.4. The action of automorphisms on unordered configurations
Using Knudsen’s formula it is possible to mislead yourself into thinking that homeomorphisms of (or indeed pointed homotopy self-equivalences of ) act on via their action on : in other words, that such maps which act trivially on the homology of also act trivially on the homology of . 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 is orientable, and first suppose that it is odd-dimensional. Then and the analysis of Section 2.2 applied to shows that is a rational homology isomorphism. So we find:
Theorem 2.1.
If is orientable and odd-dimensional, then a pointed homotopy self-equivalence of which acts trivially on also acts trivially on .∎
The even-dimensional case is more interesting. As is assumed orientable, in this case the twisting by can be dispensed with. It is technically convenient here—for reasons of symmetric monoidality—to work in the category of simplicial -modules rather than chain complexes. We write for the tensoring of this category over simplicial sets. For a space let us abbreviate , and if it is based then let . The discussion in the previous section, ignoring the formality step and translated to simplicial -modules, shows that given the simplicial module and the map induced by the diagonal , we may form the two-sided bar construction
(2.2) |
whose bigraded homotopy groups are identified with .
A homeomorphism of , or a pointed homotopy self-equivalence of , induces an equivalence such that , meaning that the diagram of simplicial commutative rings
is commutative, which induces a self-equivalence of the two-sided bar construction (2.2). This corresponds to the induced action on .
However a weaker kind of data suffices to get an induced equivalence on two-sided bar constructions. An equivalence together with a homotopy 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 . This data suffices to obtain a self-equivalence of the two-sided bar construction (2.2), as the zig-zag
where denotes considered as a -module via .
Let be another such datum, and suppose that there is a homotopy such that the 2-cell
(2.3) |
is homotopic to . Then one may check that is homotopic to . If we let denote the set of ’s modulo the equivalence relation when there exists a homotopy having the above property, then composition of maps and pasting of homotopies makes into a group, which acts on the two-sided bar construction (2.2) in the homotopy category of simplicial -modules (and so also acts on its homotopy groups). A pointed homotopy self-equivalence of acts on the two-sided bar construction through , via elements of the special form .
We may analyse the group as follows. There is a homomorphism
to the group of homotopy classes of homotopy self-equivalences of . Using the Dold–Kan theorem the latter can be identified with the group of homotopy classes of homotopy self-equivalences of , and using a formality equivalence this is identified with the group of automorphisms of the graded vector space . Such an automorphism is in the image of precisely when it preserves the map . The kernel of consists of those such that is homotopic to the identity: by definition of the equivalence relation such an element may be written as where is obtained from and a homotopy by the 2-cell diagram (2.3). Such an is a self-homotopy of the map , so an element of . The ambiguity in when representing as comes from the choice of the homotopy , so is well-defined modulo the ambiguity coming from the self-homotopies of the identity map. In conclusion, this discussion establishes an exact sequence
Using the Dold–Kan theorem and a formality equivalence again we can identify the first map in this sequence with
and so describe by an extension
This implies the following. We continue to assume that is even-dimensional and orientable. Let denote the group of homotopy classes of pointed homotopy self-equivalences of which act as the identity on .
Theorem 2.2.
If is orientable and even-dimensional, then acts on via .∎
Example 2.3.
When is a punctured surface one has so the map has the form
which in grading 2 is the inclusion of the symplectic form and is zero otherwise. Thus the above is . Using Poincaré duality and , this can be identified with . 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 is well-based if the basepoint map is a closed cofibration: under this condition preserves weak equivalences between well-based spaces, and preserves closed cofibrations. Let us say that an -graded based space is well-based if is well-based for each .
Let us write and to ease notation, so makes into a -module.
Lemma 3.1.
and are well-based. The subspace of of those tuples which do not have distinct coordinates is well-based, and this inclusion is a closed cofibration.
Proof.
Recall that is the interior of a compact manifold with boundary . This admits a collar, showing that admits a homotopy inverse, and so the vector bundle extends to a vector bundle over , which we also call . Furthermore, choosing an inner product on this bundle we can form the closed disc bundle , and consider as lying inside it as the open disc bundle. Now is a manifold with boundary , and .
Observe that is an compact manifold pair so (is an ENR pair and hence) can be expressed as a retract of a pair of the geometric realisations of a simplicial set and a subset. We may pull back to using the retraction; let us call this . Now can be given an evident cell-structure (by induction over the relative cells of ), and is a retract of it, so is well-based. More generally, for the exterior direct sum and writing for the subcomplex where some factor lies in , there is a cell structure on for which the group acts cellularly, and so is a cell complex of which is a retract, and so is well-based. This shows that is well-based, and similar reasoning shows is.
For the second statement,
is the inclusion of a -CW-subcomplex, and so has a -equivariant open neighbourhood which equivariantly deformation retracts to it. This may be chosen to preserve the subcomplexes where some factor lies in . Thus it lifts to a -equivariant deformation retraction of an open neighbourhood of , and descends to the quotient by the subcomplexes where some factor lies in . As it is equivariant, it descends further to the -quotient. That is, it proves the claim for replaced by ; as the former data is a retract of the latter, the claim follows. ∎
Lemma 3.2.
is a flat -module, in the sense that preserves weak equivalences between left -modules whose underlying objects are well-based.
Proof.
Recall that . Define a filtration of by and
where the latter term is only taken when it makes sense: for even. This is a filtration by right -modules. One checks that the diagram
is a pushout (in and so in right -modules), where the horizontal maps are induced by the -module structure and the adjoints of the map , and the map .
We prove by induction on that is a flat -module in the indicated sense. As these properties hold for . For a left -module whose underlying object is well-based, applying to the square above gives a pushout square
(3.1) |
The map is the inclusion of the subspace of those -tuples of points in labelled by which do not have distinct coordinates, so is a closed cofibration from a well-based space by the second part of Lemma 3.1. As 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 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 is flat in the indicated sense, so is too because is an isomorphism when evaluated on , so is too. ∎
Remark 3.3.
Lemma 3.4.
The induced map is an isomorphism.
Proof.
By definition of the relative tensor product there is a coequaliser diagram
in , where is given by the -module structure on , and is induced by the augmentation . The image of is precisely the image of , whose cofibre is by definition . ∎
Proof of Theorem 1.2.
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 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 , and for let
be the open subspace of those such that no has multiplicity in all of these multisets. That is, is the configuration space of particles of different colours, having colour , which may collide except that no point of may carry points of every colour. The 1-point compactifications again have a composition product
giving a commutative monoid in -graded pointed spaces. Just as before, we can introduce labels in a vector bundle , giving and . Writing with the 1 in the th position, there is a pushout square
(A.1) |
of unital commutative monoids in , where is now induced by the inclusion . The same argument as Theorem 1.2 shows that there is an equivalence
(A.2) |
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 are -homology manifolds, being open subspaces of a product of coarse moduli spaces of the orbifolds . As before, we suppose is -dimensional and take to be given by the sum of the orientation line of plus trivial lines: then the are orientable -homology manifolds, of dimension . Again they are vector bundles over , so Poincaré duality gives
On the other hand, the bar construction formula above together with the argument of Section 2.2 identifies the multigraded vector space with
A.2.1. Odd-dimensional manifolds
As in Section 2.2 we have by the Thom isomorphism. If is odd then the permutation group acts on the Thom class via , so acts nontrivially if and trivially if . If this means that , showing that
in this case. Using Poincaré duality on both sides gives [FWW19, Theorem 1.4], except that that theorem is erroneously claimed for all . We will return to the case below.
A.2.2. Even-dimensional manifolds
If is even then acts trivially on , and using the Thom isomorphism to identify too, the Koszul complex for computing the -groups above is
The differential is induced by the map
obtained by linearly dualising the cup product map
(A.3) |
and so is trivial if (and only if) all -fold cup products of (-twisted) compactly-supported cohomology classes on vanish.
When this cup product map is trivial, so is trivial, the above just gives a formula for . Using Poincaré duality, and reindexing, to express this in terms of and we obtain an identity of multigraded vector spaces
There are stabilisation maps analogous to those constructed in Section 2.3, similarly for , and both stabilise as , just as in Section 2.3: this recovers [FWW19, Theorem 1.7]. We may take the colimit of all these stabilisations to obtain
Writing and for the Poincaré series of and respectively, this discussion identifies the homological density with the Poincaré series of . This visibly only depends on the product , 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,
Just as in the even-dimensional case, if the cup product map (A.3) is zero then one gets an explicit description of , and the homological density is given by the Poincaré series of the graded vector space . It follows from Section A.2.1 that the homological density is for , so for odd-dimensional manifolds it is not true that the homological density depends only on .
A.2.4. Euler characteristic
If the cup product map (A.3) is not zero, and either is even or is odd and , then there is instead a nontrivial differential on the multigraded vector space
of degree , whose homology is . Then one would not expect to agree with the Poincaré series of , and indeed it does not [FWW19, Remark 1.6]. However, as Euler characteristic commutes with taking homology we have the identity
in . The left-hand factor is . 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 is even-dimensional and orientable: then we can dispense with twisting by the vector bundle . We consider with its -grading reduced to an -grading via . Collapsing the complement of a small neighbourhood of a point in gives a map , inducing a map of commutative monoids
If is a left -module, it is equipped with maps and so we can define the spectrum . Using these two constructions we may therefore form the spectrum
By analogy with [Ho21, Section 7.5] we propose as a spectral form of the stable density of in . At the level of -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 -graded objects, there is an evident map from (A.1) to the analogous square for which induces a map of spectra , and this is an equivalence by (A.2) as both are identified with .
This may be simplified for as follows. The map with which the derived tensor product is formed factors over so is nullhomotopic when , and so is equivalent to as a left -module. In this situation the construction gives
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.