Operations on stable moduli spaces
Abstract.
We construct certain operations on stable moduli spaces and use them to compare cohomology of moduli spaces of closed manifolds with tangential structure. We obtain isomorphisms in a stable range provided the -adic valuation of the Euler characteristics agree, for all primes not invertible in the coefficients for cohomology.
2010 Mathematics Subject Classification:
55P47, 55R40, 57S05, 57R15, 57R901. Introduction
An influential theorem of Harer [Har85] shows that the cohomology of the moduli stack of genus Riemann surfaces is independent of in a range of degrees called the stable range, even though there is no direct map between the moduli spaces for different genera. With rational coefficients the cohomology in the stable range is a polynomial ring, but with more general coefficients it is best described via infinite loop spaces, as shown by [Til97, MT01, MW07]. In earlier papers ([GRW14, GRW18, GRW17], see also [GRW19] for a survey) we have studied moduli spaces of higher dimensional manifolds, and in some cases have again shown that different moduli spaces have isomorphic cohomology in a range of degrees. For one can in most cases not make an integral comparison of moduli spaces of manifolds related by connected sum with copies of , at least not by an obvious generalization of the case, where a zig-zag of integral homology equivalences can be defined using manifolds with boundary. In this paper we show that a comparison is possible after all, although not with all coefficient modules. We also give examples showing that assumptions on the coefficients are necessary.
1.1. Comparing moduli spaces of closed manifolds
All manifolds in this paper will be smooth, compact, connected, and without boundary. If denotes such a manifold then there is a moduli space classifying smooth fibre bundles whose fibres are diffeomorphic to . As a model we may take , the classifying space of the diffeomorphism group of , equipped with the topology. Then for an abelian group is the group of -valued characteristic classes of such fibre bundles.
Now let and be a -manifold. The connected sum is then well defined up to (non-canonical) diffeomorphism, as admits an orientation-reversing diffeomorphism, and we write for the -fold iteration of this operation. Two manifolds and are called stably diffeomorphic if is diffeomorphic to for some . For example, any two orientable connected surfaces are stably diffeomorphic, while two non-orientable connected surfaces are stably diffeomorphic if and only if their Euler characteristic have the same parity.
In this paper we shall ask about the relationship between and when and are stably diffeomorphic. As a special case our main result will provide a canonical isomorphism
as long as these manifolds are simply-connected and of dimension , and both and are large compared with and have the same -adic valuation.
The precise statement of our main result applies more generally, and before giving it we first explain its natural setting. If is given an orientation then there is a corresponding moduli space classifying smooth fibre bundles with oriented fibres which are oriented diffeomorphic to , and a forgetful map . Then the connected sum inherits an orientation, well defined up to oriented diffeomorphism, and we say that is oriented stably diffeomorphic to provided is oriented diffeomorphic to for some . In this situation our result will also imply a canonical isomorphism , under the same hypotheses.
More generally, for a space equipped with a continuous action of a -structure on a -manifold is a -equivariant map , or, equivalently, a -equivariant map . For example, if on which acts by multiplication by the sign of the determinant, then a -structure is the same thing as an orientation: it distinguishes oriented frames from non-oriented ones. Two -structures on the same manifold are homotopic if they are homotopic through equivariant maps, and is -diffeomorphic to if there exists a diffeomorphism such that is homotopic to . The usual embedding of as the boundary of a thickened gives a trivialisation of and a -structure on extends to one on , canonically up to -diffeomorphism. For two pairs and consisting of a manifold and a -structure, we say that they are stably -diffeomorphic if is -diffeomorphic to for some .
There is a moduli space parametrising smooth fibre bundles with -dimensional fibres, and where the fibrewise tangent bundle is equipped with an equivariant map , such that all fibres of are -diffeomorphic to . Our main result is then as follows.
Theorem 1.1.
Let be as above, and let and be -structures on and such that is stably -diffeomorphic to . For an abelian group there is a canonical isomorphism
induced by a zig-zag of maps of spaces, provided
-
(i)
and and are simply connected,
-
(ii)
the integers and are both , where
-
(iii)
and are both non-zero, and for all primes which are not invertible in .
In Section 4 we give an example showing the third condition cannot be relaxed.
The main results of [GRW14, GRW18, GRW17], summarised in [GRW19], provide a map
(1.1) |
which is an isomorphism on homology in a range of degrees, when regarded as a map to the path component which it hits. Similarly there is a map
(1.2) |
which is an isomorphism on homology in a range of degrees, when regarded as a map to the path component which it hits. The definition of the codomains is recalled below. However, if then these two maps land in different path components, and the problem becomes to compare the homology of these two path components.
1.2. Operations on infinite loop spaces
The data involved in defining the common target of the maps (1.1) and (1.2) is a -equivariant fibration with domain which is cofibrant as a -space. Letting denote the Borel construction , is then the Thom spectrum of the inverse of the canonical -dimensional vector bundle over , and is its associated infinite loop space. By functoriality the group-like topological monoid of -equivariant homotopy equivalences acts on the infinite loop space , so the group-like submonoid does too. The target
of the maps (1.1) and (1.2) is the Borel construction for this action.
In order to prove Theorem 1.1 we shall construct certain operations on the space , in the case where the -space is obtained by restriction from a cofibrant -space . The space carries a canonical -dimensional vector bundle, and denotes its associated Thom spectrum; as above, by functoriality it carries an action of the monoid of -equivariant homotopy equivalences .
A key construction in this paper is a homotopy pullback diagram of infinite loop spaces, equivariant for , of the form
(1.3) |
whose bottom right corner has and all higher homotopy groups are -power torsion, and the bottom horizontal map induces a surjection on . It induces an isomorphism
(1.4) |
whose first coordinate is given by the Euler class and whose second coordinate is given by the stabilisation map. To explain this claim and its notation, first note that the -dimensional vector bundle over has an Euler class , where the coefficients are twisted by the determinant of this vector bundle, and under the Thom isomorphism this gives a class . Then is the value of this spectrum cohomology class on the Hurewicz image of an element of ; geometrically, it assigns to such an element the Euler characteristic of a manifold representing it. Similarly, the -dimensional vector bundle over has a th Stiefel–Whitney class , and under the Thom isomorphism this gives a class . Then denotes the value of this spectrum cohomology class on the Hurewicz image of .
Theorem 1.3.
For , write for the inverse image of under the map induced by the class , i.e. the union of the path components of the form under the bijection (1.4).
For any odd number there exists a self-map inducing a map
such that
-
(i)
commutes (strictly) with the action of ,
-
(ii)
is over the identity map of ,
-
(iii)
induces an isomorphism in homology with coefficients in any -module.
We shall also prove a version of Theorem 1.3 for , although it will be marginally weaker in that rather than the map being defined integrally and inducing an isomorphism with coefficients in any -module, the map will only be defined after localising the spaces involved away from .
Theorem 1.4.
In the setup of Theorem 1.3, if is even then there is a -equivariant weak equivalence of localised spaces
over the identity map of .
The proof of Theorem 1.1 will use these operations to give endomorphisms of the space which mix path-components, allowing us to compare the path components hit by the maps (1.1) and (1.2). This strategy is analogous to arguments of Bendersky–Miller [BM14] and Cantero–Palmer [CP15] for cohomology of configuration spaces. This strategy has also been used by Krannich [Kra19] to show that for an oriented manifold of dimension and an exotic sphere, in a stable range of degrees when the order of is invertible in .
Acknowledgements
SG was supported by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No. 682922), the Danish National Research Foundation through the Centre for Symmetry and Deformation (DNRF92), and the EliteForsk Prize. ORW was supported by EPSRC grant EP/M027783/1, the ERC under the European Union’s Horizon 2020 research and innovation programme (grant agreement No. 756444), and a Philip Leverhulme Prize from the Leverhulme Trust.
2. Proof of Theorem 1.1
Let be a factorisation into an -connected -equivariant cofibration and a -co-connected -equivariant fibration , and as above we write for the underlying -space of and for the underlying -equivariant map of . There is then a map
(2.1) |
which by [GRW17, Corollary 1.9] is an isomorphism on th (co)homology onto the path-component which it hits, as long as . (Note that by considering a -space rather than a -space, the tangential structure is “spherical” by the discussion after [GRW19, Definition 3.2], and so the stability range is as claimed.) Here is the stable -genus of , the largest for which there exists such that is -diffeomorphic to for some .
Let be a manifold stably -diffeomorphic to and minimising the quantity . Such a manifold has stable -genus zero and hence for large enough we must have that is -diffeomorphic to , so
It follows that (2.1) is an isomorphism on th (co)homology as long as
If is stably -diffeomorphic to then the same analysis applies, and there is a map
(2.2) |
which is an isomorphism on th (co)homology onto the path-component which it hits, as long as
By assumption we may write
for integers and all of whose prime factors are invertible in . Furthermore the two Euler characteristics have the same parity, as (de)stabilisation changes the Euler characteristic by , so if either or is even then both and are even too.
By Theorems 1.3 and 1.4, writing , (after perhaps implicitly localising away from 2) there are maps
which are -equivariant and induce isomorphisms on -homology, as is a -module. By construction these maps do not change the -component: we now analyse the components corresponding to and .
We now claim that for a suitable choice of lifting . Since these two elements of have the same Euler characteristic, it suffices to arrange that they also have the same -component. The stable -diffeomorphism from to gives a -cobordism
which is furthermore an -cobordism. We can therefore extend the -structure given by , stabilised, to a -structure on lifting the given -structure, and hence obtain a -manifold whose underlying -manifold is the stabilisation of . Now the -manifolds
(2.3) |
need not be -diffeomorphic, but must differ by an equivalence over (see [GRW17, Lemma 9.2]). However the -structure on is merely a choice of lift of along , and by re-choosing it to be we may then suppose that the manifolds (2.3) are indeed -diffeomorphic. With this choice we therefore have the desired
using the -cobordism and the fact that this cobordism theory is insensitive to stabilisation by standard ’s.
Denoting by the -orbit of , and similarly , and using the forgetful homomorphism to let the monoid act on , we therefore have a zig-zag of maps
(2.4) |
which induce isomorphisms on homology with -coefficients. The argument is completed by the following lemma.
Lemma 2.1.
The natural map is a weak equivalence.
Proof.
Working in the categories of -spaces over , or -spaces over , we have
but the natural -equivariant map has homotopy fibre so is -connected, whereas is -co-connected, so the restriction map
is an equivalence. The claim now follows by restricting to the path-components of homotopy equivalences. ∎
Remark 2.2.
This argument also gives a conclusion about homology with certain local coefficients. The maps (2.1) and (2.2) are in fact acyclic in a range of degrees [GRW17, Corollary 1.9], and the maps are acyclic with -module coefficients (as they are infinite loop maps which are isomorphisms on homology with these coefficients) so remain so after taking homotopy orbits by .
So if is a system of local coefficients on the middle space of the zig-zag (2.4), with typical fibre and having for all primes which are not invertible in , then there is also an isomorphism in a range of degrees.
3. Proof of Theorems 1.3 and 1.4
The proof of Theorem 1.3 is by an explicit construction of as a map of spectra. The main ingredient is a certain commutative diagram of spectra, which we first describe informally. It is
where is the natural map of Borel constructions. The map is homotopy equivalent to a smooth fibre bundle with fibres so we have a Becker–Gottlieb transfer , factoring as a pre-transfer composed with a map induced by the zero section of . The spectrum is defined to be the homotopy cofibre of , and both rows are cofibre sequences. It follows that the right square in the diagram is a homotopy pullback, and hence we get the homotopy pullback diagram of infinite loop spaces (1.3) mentioned in the introduction. On spectrum homology the map induces multiplication by , from which it follows that the homology and hence homotopy groups of are 2-power torsion. The space is path connected, because is, so . Thus , and the map is surjective on because is injective on .
To produce an endomorphism of satisfying part (ii) of the theorem, it therefore suffices to produce an endomorphism of over . For , we may use the map which is obviously over , at least in the homotopy category, since is the cofibre of the map . In spectrum homology, multiplies by and hence induces multiplication by on ensuring part (iii) of the theorem. Furthermore it acts by multiplication by on , so indeed sends to .
It remains to explain how to achieve part (i) of the theorem, that the continuous action of the topological monoid on the space commutes with . It is not sufficient that commutes up to homotopy with the action of individual elements of , since we want to descend to the homotopy orbit space. To give a convincing proof, it seems best to spell out a point-set model for the square (1.3).
Proof of Theorem 1.3.
As explained above, it remains to give a point-set model for the diagram (1.3) and the self-map of over , all of which commutes strictly with the action of .
We must adopt some conventions. Let us consider as lying inside using the last coordinates. Let us consider as lying inside as the subspace of vectors whose last coordinate is 0, and take to be the direct limit. To form the Borel constructions we shall take , and similarly take . The map which adds the basis vector of the first -summand as the first element of the -frame is then equivariant for the inclusion .
Then we have , which we may filter in the usual way by . Pulling back this filtration along the map , we set . There is an induced map and we shall write for the pullback of the -dimensional bundle of orthogonal complements. Then is the spectrum with th space given by the Thom space , so that
We similarly define , and hence the spectrum . There is a map
(3.1) |
given by direct sum with the 1-dimensional vector space given by the first -summand, which induces a map . The map (3.1) is -connected, but is covered by an -connected map and hence gives a -connected map . Passing to Thom spaces this gives a -connected map
These combine to define a map from to the spectrum whose st space is , and this map is a weak equivalence. This map is also -equivariant. (This weak equivalence does not come with a spectrum map in the other direction, let alone an equivariant one.)
The square (1.3) will be assembled from a square of spaces fibred over , and we first explain the constructions on fibres. Let and write for the unit sphere of and for the one-point compactification. If we shall write for the orthogonal projection. If we shall write for the nearest point in the sphere. We will describe certain explicit maps and , and explain how the composition gives rise to a model for the Becker–Gottlieb transfer for a linear sphere bundle (indeed, we will just unwrap the definition of [BG75, Section 3] in this case).
The map
is induced by the Pontryagin–Thom construction applied to the embedding . In formulas, we can take e.g.
when . The Thom space is homeomorphic to the quotient , and under this identification the map is the quotient map.
The map
is given by the zero section of the tangent bundle of . In formulas, it sends to .
If we compose these two maps and smash with , we get
Finally, we write for the map induced by collapsing to a point. Then the composition
is a continuous map of degree (by the Poincaré–Hopf theorem, see [BG75, Theorem 2.4]), depending continuously on the point . The resulting continuous map in the limit gives a map which is a model for the Becker–Gottlieb transfer of the sphere bundle over .
Now consider the diagram
where the entries in the right column are the mapping cylinders. Since induces a homeomorphism , it follows from the Puppe sequence that there is a canonical induced homeomorphism . Since has degree 2, there is a homotopy equivalence from to a mod 2 Moore space, but this is not quite sufficiently canonical for our purposes (since we get a different mod 2 Moore space for each ). We have proved that for each there is a canonical commutative diagram
(3.2) |
which is a pushout and homotopy pushout.
There is a canonical homotopy from the composition of and to the constant map. Suspending once, is canonically null homotopic. If is an integer, we may use the coordinate to form the sum of the identity map and copies of the map . We obtain a diagram
(3.3) |
which commutes up to a canonical homotopy. (The canonical nullhomotopy of each gives a homotopy from to the sum of the identity map and copies of the constant map; this is in turn canonically homotopic to the identity map.) The homotopy class of the map is determined by its degree which is , but the actual map depends in a non-trivial way on .
All spaces in the diagram “vary continuously in ”, in the sense that they are fibres over of fibre bundles over . The commutative diagram (3.2) in the category of spaces over may be pulled back along to form a diagram
(3.4) |
which is again a pushout and homotopy pushout, where is the mapping cylinder of the map given on by .
Similarly, the diagrams (3.3) assemble over to a diagram
(3.5) |
which commutes up to a canonical homotopy.
Applying to the diagram (3.4) and letting we get a model for (1.3). The monoid acts on the whole diagram (3.4), since it acts on over . This gives a weak equivalence from to the homotopy pullback in (1.3), which is also an equivariant map. The monoid also acts on the diagram (3.5), including the homotopy, and after applying and taking we obtain a self-map of which is over up to a specified homotopy. Again this self-map and the specified homotopy commutes strictly with the action of since both the map and the homotopy arose from fibrewise constructions over .
Finally, the self-map of induces an -equivariant self-map of the homotopy pullback of , and we have seen that this pullback is weakly equivalent to by an -equivariant map. ∎
Proof of Theorem 1.4.
We continue with the notation developed above. The spectrum homology of is all -torsion, so the localisation as a spectrum is contractible. However, the localised space is not contractible since it has two components. Instead, there is a spectrum map which becomes an isomorphism in homology of infinite loop spaces with coefficients in any -module. Similarly, the map
induces an isomorphism in homology with coefficients in any -module, and hence a weak equivalence of localized spaces. The spectrum map induces a self-map of commuting with the action of and whose restriction to the even-degree path components commutes with the map to . This self-map can be used in place of to produce . ∎
4. An example
In this section we will give an example to show that in Theorem 1.1 it is indeed necessary to take homology with certain primes inverted. We will take as an example the 6-manifolds given by a smooth degree hypersurface in , which we have studied in detail in [GRW19, Section 5.3]. Any unattributed claims about these manifolds may be found there. We will also consider their stabilisations
obtained by connect-sum of with copies of , which contain
copies of .
Theorem 4.1.
Let be a prime number, and suppose that . Then
The formula implies that , so the theorem may also be written
Hence the moduli spaces for the oriented stably diffeomorphic manifolds and have isomorphic if and only if , provided those -adic valuations are at most .
Proof of Theorem 4.1.
In [GRW19, Section 5.3] we computed the -cohomology of in a stable range. We will refer to details of the notation from that discussion, which differs slightly from the notation used earlier in this note.
Firstly, the -cohomology calculation goes through without significant changes for , because and have the same Moore–Postnikov -stage, and because any orientation preserving diffeomorphism of must also act trivially on . The only difference is that the formula for the -differential now involves characteristic numbers of , which can be calculated to give
Secondly, the -cohomology calculation yields an analogous -cohomology calculation for large enough primes . Specifically the spectrum is -connected, so by the Atiyah–Hirzebruch spectral sequence the Hurewicz map
is an isomorphism as long as , so as long as since we have assumed that . As is odd we have
Thus we have , with the isomorphism given by the tautological classes , and . Therefore
The submonoid of those path components which stabilise is path connected, and as the map is a -homology equivalence, since is odd, we also have that for . Thus the map is a -homology equivalence.
It remains to study the Serre spectral sequence for the fibration sequence
which in low degrees has a single differential
given by the formula above, so is given by the cokernel of this differential. The claim now follows by the identity of ideals
of , using again that is odd. ∎
References
- [BG75] J. C. Becker and D. H. Gottlieb, The transfer map and fiber bundles, Topology 14 (1975), 1–12.
- [BM14] M. Bendersky and J. Miller, Localization and homological stability of configuration spaces, Q. J. Math. 65 (2014), no. 3, 807–815.
- [CP15] F. Cantero and M. Palmer, On homological stability for configuration spaces on closed background manifolds, Doc. Math. 20 (2015), 753–805.
- [Fri17] N. Friedrich, Homological stability of automorphism groups of quadratic modules and manifolds, Doc. Math. 22 (2017), 1729–1774.
- [GRW14] S. Galatius and O. Randal-Williams, Stable moduli spaces of high-dimensional manifolds, Acta Math. 212 (2014), no. 2, 257–377.
- [GRW17] by same author, Homological stability for moduli spaces of high dimensional manifolds. II, Ann. of Math. (2) 186 (2017), no. 1, 127–204.
- [GRW18] by same author, Homological stability for moduli spaces of high dimensional manifolds. I, J. Amer. Math. Soc. 31 (2018), no. 1, 215–264.
- [GRW19] by same author, Moduli spaces of manifolds: a user’s guide, Handbook of homotopy theory, Chapman & Hall/CRC, CRC Press, Boca Raton, FL, 2019, pp. 445–487.
- [Har85] J. L. Harer, Stability of the homology of the mapping class groups of orientable surfaces, Ann. of Math. (2) 121 (1985), no. 2, 215–249.
- [Kra19] M. Krannich, On characteristic classes of exotic manifold bundles, Mathematische Annalen (2019), arXiv:1802.02609.
- [MT01] I. Madsen and U. Tillmann, The stable mapping class group and , Invent. Math. 145 (2001), no. 3, 509–544.
- [MW07] I. Madsen and M. Weiss, The stable moduli space of Riemann surfaces: Mumford’s conjecture, Ann. of Math. (2) 165 (2007), no. 3, 843–941.
- [Til97] U. Tillmann, On the homotopy of the stable mapping class group, Invent. Math. 130 (1997), no. 2, 257–275.