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

Quantitative PL bordism

Fedor Manin  and  Shmuel Weinberger Department of Mathematics, University of California, Santa Barbara, CA, United States manin@math.ucsb.edu Department of Mathematics, University of Chicago, IL, United States manin@math.ucsb.edu
Abstract.

We study PL bordism theories from a quantitative perspective. Such theories include those of PL manifolds, ordinary homology theory, as well as various more exotic theories such as bordism of Witt spaces. In all these cases we show that a null-bordant cycle of bounded geometry and VV simplices has a filling of bounded geometry whose number of simplices is slightly superlinear in VV. This bound is similar to that found in our previous work on smooth cobordism.

1. Introduction

Suppose a compact manifold MM is nullcobordant, i.e. is the boundary of some compact manifold WW. How complicated must WW be? This is not just one question, but a whole family of questions, depending on the type of manifold that we allow MM and WW to be, and how we measure the complexity of WW.

To our knowledge, Gromov is the first to ask such questions; he discusses several variations in [Gro96, §5575\frac{5}{7}] and [Gro99]. In the case that MM and WW are smooth, he gave as a possible measure of complexity the infimal volume of a metric given a fixed bound on the local geometry (say, sectional curvatures |K|1|K|\leq 1 and injectivity radius at least 11). He suggested that the minimal complexity of a WW which fills MM should be linear in the complexity of MM; a slightly superlinear bound (O(V1+ε)O(V^{1+\varepsilon}) for any ε>0\varepsilon>0) was proved in [CDMW18].

In this paper we prove an analogous bound in the PL category. In this setting, “bounded geometry of type LL” means that every vertex is incident to at most LL simplices, and volume is measured by the total number of top-dimensional simplices.

Theorem 1.1.

Let MM be a PL kk-manifold with bounded geometry of type LL and VV total simplices. Then if MM is a PL boundary, it bounds a manifold WW with bounded geometry of type LL^{\prime} (depending on kk and LL) so that the number of simplices in WW is bounded by a function (again depending inexplicitly on kk and LL) which is O(V1+ε)O(V^{1+\varepsilon}) for every ε>0\varepsilon>0.

1.1. Discussion

The proof of the bound in [CDMW18] built on the original work of René Thom [Tho53, Tho54] relating cobordism to the homotopy groups of the Thom space of the universal bundle over a Grassmannian. This Thom space is a specific finite complex XX, and to prove the quantitative bound one must relate the Lipschitz constant of a nullhomotopic map f:SNXf:S^{N}\to X to the Lipschitz constant of its most efficient nullhomotopy. Even talking about Lipschitz constants requires a metric on XX, but for an asymptotic computation the specific metric doesn’t matter because all the reasonable (Riemannian or piecewise Riemannian) metrics are Lipschitz homotopy equivalent.

The method of proof is quite general, and although this is not pointed out in that paper, it applies to oriented and non-oriented cobordism, as well as spin and complex cobordism and any similar variant. The reason is that in all such cases Thom’s method gives the appropriate reduction to the homotopy theory of finite polyhedra, and the spaces that arise are particularly simple from the point of view of rational homotopy theory.

However, this method does not apply to PL cobordism, because the classifying spaces are not close relatives of compact Lie groups. While they are abstractly homotopy equivalent to spaces with finite skeleta, we do not know of a natural way of representing them as such. This paper studies the complexity of PL nullcobordisms, with an eye towards developing techniques that can apply in situations where the traditional methods of geometric topology (such as classifying spaces, h-principles, etc.) reduce geometric problems to the homotopy theory of infinite complexes.

For infinite complexes, one can choose metrics that are inequivalent at large scales, and different choices can lead to very different measures of complexity. For example, a finite-type model for K(,n)K(\mathbb{Z},n) and a simplicial group model with infinitely many simplices in each dimension lead to very different “quantitative cohomology theories”. A related situation arises in K-theory, whose concrete manifestation is the following. Suppose one is interested in vector bundles of dimension nn over a finite complex XX of lower dimension. The number of such bundles which admit a connection with bounded curvature KK grows as a polynomial in KK whose degree is related to the cohomology of XX. However, the naturally isomorphic set of stable vector bundles over XX can be realized by bundles (of varying dimension) with a uniform bound on curvature; if one takes curvature as the measure of complexity, then infinitely many classes are realized with bounded complexity.

Now, let us be a bit more precise. A reasonable notion of the complexity of a PL manifold MM is the number of simplices in a triangulation. We could consider the problem: given a PL kk-manifold MM with VV simplices, how many simplices are contained in the simplest WW whose boundary is MM?

It is true, but not completely obvious, that the complexity of a nullcobordism can be bounded by a recursive (i.e. computable) function of the complexity of MM. This is equivalent to the statement that the PL bordism problem is algorithmically decidable, and that follows from the deep work that is explained e.g. in [MM79, Ch. 14]. One basic step in the analysis is the structure of G/PL, which is proved using surgery theory and deduced from the Poincaré conjecture. The algorithmic nature of these ingredients seems quite difficult to unravel.

One would have been led directly to the Poincaré conjecture in thinking about this question when considering whether a candidate MM that someone hands you is actually a PL manifold. Maybe your interlocutor is playing a joke on you? You can be sure that MM is a homology manifold, but it is much harder to know that MM is a manifold—you would need to see that the links of simplices are spheres. Checking that a manifold is actually a sphere, if you think it should be, is exactly what the Poincaré conjecture is about.

Of course, one might well produce a polyhedron bounded by MM and give it to the interlocutor, who can then worry about whether what you produced is a manifold. (Fair is fair.) So, despite first impressions, it is not quite necessary to complete an analysis of the algorithmic complexity of the Poincaré conjecture to handle the bordism question.

Regardless, we do not know how to deal with the PL bordism problem for arbitrary triangulations (although we do have a program that, if it succeeds, would give a tower of exponentials bound for the rise in complexity). At the same time, there is no indication that the true answer is any more nonlinear than it is for smooth bordism.

Instead we consider the case of PL manifolds of bounded geometry. That is, we consider PL manifolds with some bound on the number of simplices incident to each vertex, and build cobordisms constrained by a similar, perhaps larger bound. This condition is similar to the smooth locally bounded geometry condition in that the amount of “local information” is bounded. In fact, there is a closer correspondence: any smooth manifold of bounded geometry has a bounded geometry triangulation at scale 1\sim 1, and any smoothable PL manifold of bounded geometry admits a smoothing with locally bounded geometry and with simplices of bounded volume; see Appendix A.

Moreover, in every dimension dd, there is an L(d)L(d) such that every PL dd-manifold is PL homeomorphic to one with bounded geometry of type L(d)L(d). (However, given a manifold with VV simplices, the number of subdivisions required to implement a PL isomorphism with such a manifold grows faster than any computable function of VV.) So bordism of PL manifolds with bounded geometry is equivalent, as far as pure topology is concerned, to bordism of PL manifolds.

As in the smooth case, our methods work in a number of settings: they extend from PL manifolds to pseudomanifolds whose singularities come from a prescribed family (see §4.1 for a precise definition), as well as to bordism homology of spaces other than a point. The full statement of our main theorem is therefore more general than Theorem 1.1:

Main Theorem.

Let \mathcal{M} be a category of PL manifolds with prescribed singularities.

For every kk, LL, and ε>0\varepsilon>0, there are constants C=C(,k,L,ε)C=C(\mathcal{M},k,L,\varepsilon) and L=L(,k,L)L^{\prime}=L^{\prime}(\mathcal{M},k,L) such that every PL nullcobordant triangulated kk-dimensional \mathcal{M}-manifold XX with VV kk-simplices and bounded geometry of type LL has a filling with at most CV1+εCV^{1+\varepsilon} simplices and bounded geometry of type LL^{\prime}. This holds for both oriented and unoriented cobordism.

More generally, for a finite simplicial complex YY, there is a constant C=C(,k,L,ε,Y)C=C(\mathcal{M},k,L,\varepsilon,Y) such that if XX is as above and f:XYf:X\to Y is a simplicial map representing 0 in the bordism homology group Ωk(Y)\Omega^{\mathcal{M}}_{k}(Y), then there is a simplicial extension f~:WY\tilde{f}:W\to Y to a filling of XX with at most CV1+εCV^{1+\varepsilon} simplices and bounded geometry of type LL^{\prime}.

Note that the constants depend in an inexplicit way on the bound on geometry. This is a point of contrast with the smooth category, where we can always change the bound on geometry by rescaling, which automatically scales the nullcobordism as well.

1.2. Overview of proof

As in [CDMW18], we follow the usual proof of the classification of manifolds up to cobordism, bounding the complexity of each step. The main additional technical challenge stems from the fact that while classifying spaces for PL cobordism theories exist by Brown representability, unlike in the smooth case, there are no nice explicit models of finite type that we can use to construct our classifying maps. We resolve this by building the models we need, but these models may not actually be homotopy equivalent to the relevant classifying spaces; they do, however, have enough topological resemblance for our purposes.

Let VV be the number of kk-simplices of the manifold XX. We proceed as follows:

  1. (1)

    Embed XX in SnS^{n}, with some control over the shape of a regular neighborhood. Here we can choose any sufficiently large nn.

  2. (2)

    This induces an φ(V)\varphi(V)-Lipschitz map from SnS^{n} to a fixed compact subspace of a kind of Thom space for \mathcal{M}-manifolds.

  3. (3)

    One constructs a ψ(φ(V))\psi(\varphi(V))-Lipschitz extension of this map to Dn+1D^{n+1}, with image in a larger fixed compact subspace of this Thom space.

  4. (4)

    From a simplicial approximation of this nullhomotopy, one can extract an \mathcal{M}-manifold embedded in Dn+1D^{n+1} which fills XX and whose number of simplices is bounded by the number of simplices in the approximation and is therefore O(ψ(φ(V))n+1)O(\psi(\varphi(V))^{n+1}).

  5. (5)

    After the previous step, the bound on the local geometry of the resulting filling WW depends on nn. We tidy up by applying local modifications to WW, making the bound on the geometry independent of nn at the expense of increasing the number of simplices by an additional multiplicative constant.

The final estimate depends on bounds on the functions φ\varphi and ψ\psi. Our results from the appendix to [CDMW18] suffice to show that

φ(V)C(k,n)V1nk(logV)2k+2,\varphi(V)\leq C(k,n)V^{\frac{1}{n-k}}(\log V)^{2k+2},

and we will show that ψ(L)\psi(L) is at most linear. This gives an overall bound of O(V1+O(1/n))O(V^{1+O(1/n)}) on the size of the resulting filling; since nn can be arbitrarily large, this gives our desired bound.

2. Examples and corollaries

We now discuss the implications of our results for particular categories of manifolds with singularities.

2.1. PL manifolds

The most obvious such category is usual PL manifolds, with no singularities. Here we remark upon a prior result of F. Costantino and D. Thurston:

Theorem ([CT08, Theorem 5.2]).

Every 33-manifold MM with a triangulation with VV 33-simplices has a filling WW which has bounded geometry and O(V2)O(V^{2}) simplices.

Note that in this theorem, MM is not required to have bounded geometry. However, in dimension 3 this doesn’t matter, and our main theorem implies a stronger result:

Corollary 2.1.

Every 33-manifold MM with a triangulation with VV 33-simplices has a filling WW which has bounded geometry and f(V)f(V) 44-simplices, where f(V)=O(V1+ε)f(V)=O(V^{1+\varepsilon}) for ε>0\varepsilon>0.

Proof.

It suffices to show that every 33-manifold MM with VV 33-simplices has a triangulation of bounded geometry with O(V)O(V) simplices. Such a triangulation can be produced as follows. This strategy was outlined by Gromov [Gro96, §5575\frac{5}{7}II].

First, take the unbounded triangulation and barycentrically subdivide once.

Now let 𝒱1\mathcal{V}_{1} be the set of vertices which come from the barycenters of edges. The star of each such vertex looks like ScZScZ (the suspension of the cone on ZZ), where ZZ is a circle with some number nn of edges. It is easy to see that ZZ can be filled with a disk DD with O(n)O(n) triangles, at most 77 of which meet at a vertex, and at most 33 of which meet at a vertex on PP. So for each vertex in 𝒱1\mathcal{V}_{1}, we replace the subcomplex ScZScZ with SDSD, getting a PL homeomorphic complex MM^{\prime}.

Now let 𝒱2\mathcal{V}_{2} be the set of vertices of MM^{\prime} that correspond to the original vertices of MM. The link of such a vertex in MM^{\prime} is a triangulation of S2S^{2} with nn faces and at most 99 faces meeting at every vertex. Such a triangulation again has a filling with bounded geometry and O(n)O(n) 33-simplices. Gromov sketches the proof in [Gro96, §5575\frac{5}{7}II′′]: after smoothing out the triangulation, one gets a Riemannian metric on S2S^{2} with curvature bounded below by some K-K. By a theorem of Alexandrov [Ale06, §XII.2], such an S2S^{2} embeds isometrically into 3\mathbb{H}^{3} with curvature K-K. In hyperbolic space, this sphere bounds a ball of volume O(n)O(n). Retriangulating that ball gives us our bounded geometry filling.

Replacing the star of each vertex in 𝒱2\mathcal{V}_{2} with a bounded geometry filling of its link gives us our bounded geometry triangulation of MM. ∎

Notice that this strategy generalizes to higher dimensions: if every bounded geometry triangulated SkS^{k} of volume VV can be filled with a bounded geometry disk of volume Fk(V)F_{k}(V), then every nn-manifold MM with a triangulation of unbounded geometry and volume VV has a triangulation of bounded geometry and volume

Gn(V)=Fn1(Fn2(F3(O(V)))).G_{n}(V)=F_{n-1}(F_{n-2}(\cdots F_{3}(O(V))\cdots)).

However, we currently have no estimates on the functions FkF_{k} for k3k\geq 3. Finding such estimates seems to be a worthwhile and difficult problem. In a future paper we will show that F7F_{7} is at least quadratic, the first nonlinear lower bound for this type of problem.

2.2. General pseudomanifolds

The next most obvious application of our theorem after PL manifolds is to pseudomanifolds with arbitrary singularities. Notice that the corresponding problem with unbounded geometry is trivial, as for any pseudomanifold one can fill it by taking the cone. On the other hand, restating our main theorem in this case gives the following new result:

Corollary 2.2.

Let YY be a finite complex. Then every simplicial kk-cycle in YY of bounded geometry of type LL and volume VV has a filling of bounded geometry of type L=L(k,L)L^{\prime}=L^{\prime}(k,L) and volume f(V)f(V), where f(V)=O(V1+ε)f(V)=O(V^{1+\varepsilon}) for every ε>0\varepsilon>0.

This fact may be useful in the study of high-dimensional expanders, random complexes, and related topics.

2.3. Other classes

There are a number of other categories of manifolds with singularities whose bordism theory is relevant in geometric topology. These include Witt spaces [Sie83, Gor84], whose bordism homology theory is closely related to KO-theory. Many other examples are discussed in [Fri15, §5.2.1].

Some such examples, including Witt spaces, have infinitely generated bordism groups. In this case, the corresponding classifying spaces do not have finite type. However, for a given bound on geometry, bordism classes are classified by a finite subspace. In other words, one can find bordism classes which require arbitrarily high local complexity, as expressed by the topology of links of simplices. Contrast this with the case of PL manifolds, in which all links have the same topology. Our main theorem still holds as written in such a setting, although for any given bound LL on the geometry, we are really only considering part of the bordism theory.

This suggests that, unlike in the case of PL manifolds, the problem of finding null-bordisms for Witt spaces without bounded geometry is fundamentally different.

On the other hand, note that if, for example, one is interested in Witt null-bordisms of non-singular PL manifolds, this can be done with finitely many topological link types, even without a bounded geometry assumption on the original manifold.

2.4. Relation to secondary invariants

One area of application of these ideas is, in principle, to the connection between “secondary” invariants and complexity. The most famous secondary invariants—the η\eta-invariants of Atiyah–Patodi–Singer [APS75, APS73, Wal99]—and their L2L^{2} analogues due to Cheeger and Gromov are linearly bounded as a function of the volume of a manifold with bounded geometry [CG85]. These invariants are defined using the signature of a nullcobordism, and indeed, this inequality was Gromov’s original motivation for suggesting that cobordisms may have linear volume. Since our volume estimates are superlinear, they do not recover the optimal asymptotics of Cheeger and Gromov.

Cha [Cha16], in dimension 3, and Lim and Weinberger [LW23], in all dimensions, have proven an inequality for the Cheeger–Gromov ρ\rho-invariant in terms of the number of simplices by a mixture of algebraic and geometric methods. The followup paper of Cha and Lim [CL] makes essential use of a non–locally finite classifying space for proving such estimates, and strongly suggests the possibility of extending the results of this paper beyond bounded geometry.

For other invariants, our results do provide a new estimate. For example, the bordism homology group Ωk(BΓ)\Omega_{k}^{\mathcal{M}}(B\Gamma), when \mathcal{M} is the class of Witt spaces, is used to define higher ρ\rho-invariants of manifolds with fundamental group Γ\Gamma [Wei99]. If BΓB\Gamma is of finite type, our results can then be used to bound these higher ρ\rho-invariants, and therefore (in certain situations) to bound the number of manifolds of bounded geometry homotopy equivalent to a given one. On the other hand, if BΓB\Gamma is not of finite type, we get an important example of a situation in which our results do not apply.

Remark.

The η\eta-invariants can also be studied in terms of a much cruder complexity measure: number of handles, rather than number of simplices. This is a very different measure of complexity: in dimension three, it is the classical Heegaard genus. Unlike number of simplices, there can be infinitely many manifolds with bounded complexity, e.g. there are infinitely many 3-manifolds with Heegaard genus 1. Gromov suggests (although we have only been able to verify this assertion for framed manifolds111This is a consequence of the fact that the main theorems of surgery theory are proved by a greedy method of improvement, special facts about quadratic forms over \mathbb{Z}, and the theorem of Kervaire–Milnor showing that the number of differentiable structures on the nn-sphere is finite.) that the number of handles necessary to produce an oriented nullcobordism of a compact manifold can be bounded linearly in the number of handles in the manifold. For manifolds with fundamental group /2\mathbb{Z}/2\mathbb{Z}, and if the cobordism is assumed to have the same fundamental group, one can use the ρ\rho-invariant to give a lower bound on the number of handles in a bordism, and therefore deduce that the number of handles in such a nullcobordism cannot be bounded at all by the number of handles in the manifold.

3. Efficient Whitney embeddings

In this section we describe a way of embedding a PL kk-manifold efficiently into n\mathbb{R}^{n}, where n2k+1n\geq 2k+1, implementing step (1) of the proof outline.

In [GG12], Gromov and Guth describe “thick” embeddings of kk-dimensional simplicial complexes in unit nn-balls, for n2k+1n\geq 2k+1. They define the thickness TT of an embedding to be the maximum value such that disjoint simplices are mapped to sets at least distance TT from each other. In [GG12, Thm. 2.1], given a complex with volume VV and bounded geometry of type LL, they construct embeddings with a lower bound on thickness whose asymptotic behavior as a function of VV is close to the theoretical upper bound.

Their construction is probabilistic: they first produce a random embedding, in which some simplices may pass too close to each other. They then bend the simplices around each other on a smaller scale to resolve these near-collisions.

Gromov and Guth’s result was strengthened in [CDMW18, Appendix A §2], using essentially the same method, to give the same bound with respect to a slightly stronger notion of bounded geometry. We give a version of this strengthened result here. Denote the link of a simplex σ\sigma by lkσ\operatorname{lk}\sigma.

Theorem 3.1 (based on [CDMW18, Appendix A, Theorem 2.1]).

Suppose that XX is a kk-dimensional simplicical complex with VV vertices and each vertex lying in at most LL simplices. Suppose that n2k+1n\geq 2k+1. Then there are constants C(n,L)C(n,L) and α(n,L)>0\alpha(n,L)>0 and a subdivision XX^{\prime} of XX which embeds linearly into the nn-dimensional Euclidean ball of radius

RC(n,L)V1nk(logV)2k+2R\leq C(n,L)V^{\frac{1}{n-k}}(\log V)^{2k+2}

such that:

  1. (i)

    The embedding has Gromov–Guth thickness 1.

  2. (ii)

    For any ii-simplex σ\sigma of XX^{\prime}, the induced embedding lkσSni1\operatorname{lk}\sigma\to S^{n-i-1} has Gromov–Guth thickness α(n,L)\alpha(n,L).

  3. (iii)

    Adjacent vertices of XX^{\prime} are mapped (n,L)\leq\ell(n,L) units apart.

  4. (iv)

    The vertices of the embedding are “snapped to a grid”: they are points in a lattice in n\mathbb{R}^{n} which depends only on LL.

Remarks.
  1. (a)

    When XX is a PL kk-manifold, an embedding is locally flat if it is locally PL homeomorphic to the standard embedding of k\mathbb{R}^{k} in n\mathbb{R}^{n}. By the Zeeman unknotting theorem, this is always true unless nk=2n-k=2.

  2. (b)

    Conditions (i) and (iii) taken together imply that each simplex is bb-bilipschitz to a standard simplex for some b=b(n,L)b=b(n,L).

  3. (c)

    Together with condition (iv), this means that the simplices in the embedding come from a finite number of isometry types depending only on nn, kk and LL.

Proof.

The cited theorem gives an embedding of XX satisfying conditions (i) and (ii). Moreover, the proof given in [CDMW18] produces a subdivision whose simplices are b(n,L)b(n,L)-bilipschitz to an equilateral simplex whose edges have length (logV)2k+2(\log V)^{2k+2}. A further subdivision gives simplices which are bilipschitz to the standard simplex at scale 11. In particular, this gives condition (iii).

To see that condition (iv) can be satisfied, first take an embedding that satisfies conditions (i)–(iii) and has Gromov–Guth thickness 1+α(n,L)1+\alpha(n,L), and move all the vertices to the nearest point of a cubic lattice of side length λ=α(n,L)/4n\lambda=\alpha(n,L)/4\sqrt{n}, extending the map linearly to the rest of the complex. Then each vertex, and therefore each point of the complex, moves by at most α(n,L)/8\alpha(n,L)/8, so the resulting complex still has Gromov–Guth thickness at least 11. Moreover, since vertices are at least 11 apart, the angular movement of each point in a link is at most α(n,L)/2\alpha(n,L)/2, and therefore the thickness of links in the resulting complex is still at least α(n,L)/2\alpha(n,L)/2. Thus we can accommodate condition (iv) by at worst halving α(n,L)\alpha(n,L). ∎

4. Simplicial Pontrjagin–Thom constructions

Since cobordism theories are cohomology theories, we can represent a cobordism class by a map to a classifying space and a nullbordism by a nullhomotopy of such a map. It is somewhat difficult to describe such classifying spaces in an explicit way, although [BRS76, §II.5] and [LR78] provide possible approaches. In this section, starting with a category \mathcal{M} of PL manifolds with prescribed singularities (in other words, of pseudomanifolds with a prescribed set of permissible links), we develop a “Pontrjagin–Thom construction”: a space Th(n,k)\operatorname{Th}\mathcal{M}(n,k) such that for nn sufficiently larger than kk, \mathcal{M}-bordism classes of spaces in \mathcal{M} correspond to homotopy classes of maps SnTh(n,k)S^{n}\to\operatorname{Th}\mathcal{M}(n,k). Note that we do not guarantee that this mapping is bijective, but merely injective: πn(Th(n,k))\pi_{n}(\operatorname{Th}\mathcal{M}(n,k)) may contain classes that are not in the image of the bordism group Ωk\Omega^{\mathcal{M}}_{k}. Therefore the spaces we construct are not explicit classifying spaces for \mathcal{M}-bordism; we fail to answer, e.g., [Gor84, §8, Question 2]. Nevertheless, this construction is sufficient for our purpose.

4.1. Classes of singularities

We first specify the necessary conditions on the category \mathcal{M} that are required for bordism to make sense. These axioms are similar to those given in [BRS76, §IV.3]. A category of manifolds-with-singularity is specified by a family {n}n\{\mathcal{L}_{n}\}_{n\in\mathbb{N}} of sets of (n1)(n-1)-dimensional simplicial complexes which constitute permissible links of vertices in a closed nn-dimensional \mathcal{M}-manifold. These sets must satisfy:

  1. (1)

    n\mathcal{L}_{n} is closed under PL isomorphism.

  2. (2)

    Each member of n\mathcal{L}_{n} is a closed \mathcal{M}-manifold (that is, its links are contained in n1\mathcal{L}_{n-1}).

  3. (3)

    If PnP\in\mathcal{L}_{n}, then SPn+1SP\in\mathcal{L}_{n+1}.

  4. (4)

    If PnP\in\mathcal{L}_{n}, then cPn+1cP\notin\mathcal{L}_{n+1}.

The last axiom means that there is a notion of \mathcal{M}-manifold with boundary and that the boundary of such an object is well-defined. Namely, an nn-dimensional \mathcal{M}-manifold with boundary is a simplicial complex whose links lie in n\mathcal{L}_{n} or cn1c\mathcal{L}_{n-1}, and its boundary is the subcomplex of points whose links lie in cn1c\mathcal{L}_{n-1}. Axiom 3 ensures that if MM is an \mathcal{M}-manifold, then M×IM\times I is a \mathcal{M}-manifold with boundary.

An additional “regularity” axiom, which is not strictly necessary, guarantees that \mathcal{M}-manifolds are pseudomanifolds:

  1. (5)

    1={S0}\mathcal{L}_{1}=\{S^{0}\}.

4.2. Thom spaces

The construction of the Thom space is straightforward: it is patched together out of pieces which encode possible local behaviors of a kk-dimensional \mathcal{M}-manifold embedded simplexwise linearly in n\mathbb{R}^{n}. To be precise, Th(n,k)\operatorname{Th}\mathcal{M}(n,k) is a simplicial complex glued out of subcomplexes KP,f,NK_{P,f,N} corresponding to triples (P,f,N)(P,f,N), where:

  • PP is a permissible link in \mathcal{M}.

  • f:cPnf:cP\to\mathbb{R}^{n} is a linear embedding of the cone on PP. Let p0p_{0} be the cone point of cPcP.

  • NN is a neighborhood of f(p0)f(p_{0}) which is equipped with a linear triangulation transverse to ff and such that f1(N)P=f^{-1}(N)\cap P=\emptyset.

  • For every kk-simplex ΔP\Delta\subset P, the subcomplex NcΔNN_{c\Delta}\subseteq N defined by

    NcΔ={simplices σNf1(σ)intst(cΔ)}N_{c\Delta}=\bigcup\{\text{simplices }\sigma\subset N\mid f^{-1}(\sigma)\subseteq\operatorname{int}\operatorname{st}(c\Delta)\}

    satisfies NcΔf(cΔ)N_{c\Delta}\cap f(c\Delta)\neq\emptyset. In other words, there is at least one simplex of NN which intersects f(cΔ)f(c\Delta) nontrivially and does not intersect f(cΔ)f(c\Delta^{\prime}) for any simplex ΔP\Delta^{\prime}\subset P which does not contain Δ\Delta. This condition can be thought of as ensuring that that NN has a nice simplicial projection to the cell complex dual to f(cP)f(cP).

Figure 1. A schematic illustration of f(cP)f(cP) (light gray with thick lines) and NN (dark gray with thin lines) for a typical triple (P,f,N)(P,f,N).

We identify two triples (P,f,N)(P,f,N) and (P,f,N)(P,f^{\prime},N^{\prime}) if ff and NN differ from ff^{\prime} and NN^{\prime} by the same translation.

Given such a triple, let NNN_{\infty}\subset N be the subcomplex consisting of simplices disjoint from f(P)f(P). Then Th(n,k)\operatorname{Th}\mathcal{M}(n,k) is glued out of complexes KP,f,NNcNK_{P,f,N}\cong N\cup cN_{\infty}. Given two such subcomplexes K(P,f,N)K_{(P,f,N)} and K(P,f,N)K_{(P^{\prime},f^{\prime},N^{\prime})}, if there are vertices vPv\in P and vPv^{\prime}\in P^{\prime} such that f(cv)=f(cv)f(cv)=f^{\prime}(cv^{\prime}) and Ncv=NcvN_{cv}=N_{cv^{\prime}}, then the subcomplexes

Kv=Ncvc(NcvN)K(P,f,N)andKv=Ncvc(NcvN)K(P,f,N)K_{v}=N_{cv}\cup c(N_{cv}\cap N_{\infty})\subset K_{(P,f,N)}\quad\text{and}\quad K_{v^{\prime}}=N_{cv^{\prime}}\cup c(N_{cv^{\prime}}\cap N_{\infty})\subset K_{(P^{\prime},f^{\prime},N^{\prime})}

are identified in Th(n,k)\operatorname{Th}\mathcal{M}(n,k). Moreover, all the cone points of all the cNcN_{\infty} are identified (this is the “point at infinity” in the Thom space).

Optionally, we can make this into a Thom space for oriented \mathcal{M}-manifolds by adding orientation data for top-dimensional simplices of cPcP to the triple (P,f,N)(P,f,N), and require that this data match in order to identify subcomplexes.

Notice that Th(n,k)\operatorname{Th}\mathcal{M}(n,k) occurs as a subcomplex in Th(n+1,k+1)\operatorname{Th}\mathcal{M}(n+1,k+1). This is because for any subcomplex K(P,f,N)K_{(P,f,N)} of Th(n,k)\operatorname{Th}\mathcal{M}(n,k), Th(n+1,k+1)\operatorname{Th}\mathcal{M}(n+1,k+1) contains a subcomplex KP,f,NK_{P^{\prime},f^{\prime},N^{\prime}} corresponding to the triple

(P=(cP×[0,1]),f:(x,t)(f(x),3t1),N=N×[0,1]).(P^{\prime}=\partial(cP\times[0,1]),f^{\prime}:(x,t)\mapsto(f(x),3t-1),N^{\prime}=N\times[0,1]).

Then K(P,f,N)K_{(P,f,N)} includes into KP,f,NK_{P^{\prime},f^{\prime},N^{\prime}} as the subcomplex KvK_{v} for v=(cone point,0)(cP×[0,1])v=(\text{cone point},0)\in\partial(cP\times[0,1]). These inclusions respect the gluings that produce the complexes as a whole. If using oriented simplices, we require that cPcP face in the positive direction in cP×[0,1]cP\times[0,1].

Finally, we must show that these complexes are indeed Thom spaces in the sense we need:

Proposition 4.1.

When n>2k+1n>2k+1, there is an injective “Pontrjagin–Thom” map

Ωkπn(Th(n+1,k+1))\Omega^{\mathcal{M}}_{k}\to\pi_{n}(\operatorname{Th}\mathcal{M}(n+1,k+1))

where Ωk\Omega^{\mathcal{M}}_{k} is the bordism group of kk-dimensional \mathcal{M}-manifolds. Moreover, Th(n,k))\operatorname{Th}\mathcal{M}(n,k)) is (nk1)(n-k-1)-connected.

Proof.

Given an \mathcal{M}-manifold MM, we can construct a PL embedding f:Mnf:M\to\mathbb{R}^{n}. We then construct a suitable triangulation of n\mathbb{R}^{n} as follows.

Lemma 4.2.

Let MM be a finite simplicial complex linearly embedded in n\mathbb{R}^{n}. There is a linear triangulation τ\tau of n\mathbb{R}^{n} which is transverse to MM and has the following property. For a vertex vMv\in M, define the subset

Nv={Δτ:Δf(M)f(intst(v))}.N_{v}=\bigcup\{\Delta\in\tau:\Delta\cap f(M)\subseteq f(\operatorname{int}\operatorname{st}(v))\}.

Then the sets NvN_{v} cover f(M)f(M).

Proof.

Let bs(M)\operatorname{bs}(M) be the barycentric subdivision of MM. It suffices for τ\tau to be such that if a simplex Δ\Delta of τ\tau intersects the star in bs(M)\operatorname{bs}(M) of a vertex vv, then τM\tau\cap M is contained in the open star of vv in MM. This is true for any sufficiently fine triangulation. ∎

Now suppose we have such a triangulation. Then for any simplex Δ\Delta of MM,

NΔ=vΔNvN_{\Delta}=\bigcap_{v\in\Delta}N_{v}

satisfies the following conditions:

  • f1(NΔ)f^{-1}(N_{\Delta}) is contained in int(st(Δ))\operatorname{int}(\operatorname{st}(\Delta)), and in fact for each vΔv\in\Delta, NΔN_{\Delta} is the maximal subcomplex of NvN_{v} for which this holds.

  • NΔf(Δ)N_{\Delta}\cap f(\Delta)\neq\emptyset (by the Knaster–Kuratowski–Mazurkiewicz lemma, since for any simplex ΣΔ\Sigma\subset\Delta, vΣNv\bigcup_{v\in\Sigma}N_{v} covers f(Σ)f(\Sigma)).

Therefore, there is a well-defined map nTh(n,k)\mathbb{R}^{n}\to\operatorname{Th}\mathcal{M}(n,k) which sends each NvN_{v} to the subcomplex K(lk(v),f|st(v),Nv)K_{(\operatorname{lk}(v),f|_{\operatorname{st}(v)},N_{v})}, and simplices disjoint from f(M)f(M) to the point at infinity. This map obviously extends to Sn=n{}S^{n}=\mathbb{R}^{n}\cup\{\infty\}. This shows that for every kk-dimensional \mathcal{M}-manifold, there is a corresponding element of πn(Th(n+1,k+1))\pi_{n}(\operatorname{Th}\mathcal{M}(n+1,k+1)).

Similarly, if MM is the boundary of a (k+1)(k+1)-dimensional \mathcal{M}-manifold WW, we can extend the embedding ff to an embedding Wn×[0,1]W\to\mathbb{R}^{n}\times[0,1] and use this embedding to construct a nullhomotopy Dn+1Th(n+1,k+1)D^{n+1}\to\operatorname{Th}\mathcal{M}(n+1,k+1) of ff. This shows that the correspondence is well-defined on bordism classes.

Conversely, given a simplicial map g:SnTh(n+1,k+1)g:S^{n}\to\operatorname{Th}\mathcal{M}(n+1,k+1), we can recover a kk-dimensional \mathcal{M}-manifold. Notice that Th(n,k)\operatorname{Th}\mathcal{M}(n,k) contains a “zero section” Th0(n,k)\operatorname{Th}_{0}\mathcal{M}(n,k), consisting of the image of ff in each K(P,f,N)K_{(P,f,N)}. We claim that M=g1(Th0(n,k))M=g^{-1}(\operatorname{Th}_{0}\mathcal{M}(n,k)) is an \mathcal{M}-manifold. Indeed, let xx be a point in Mg1(Δ)M\cap g^{-1}(\Delta), where Δ\Delta is an (nj)(n-j)-simplex of Th(n,k)\operatorname{Th}\mathcal{M}(n,k). Then it is easy to see that lk(x)\operatorname{lk}(x) is the jj-fold suspension of the link of g(x)g(x) in ΔTh0(n,k)\Delta\cap\operatorname{Th}_{0}\mathcal{M}(n,k).

Similarly, for a simplicial map Dn+1Th(n+1,k+1)D^{n+1}\to\operatorname{Th}\mathcal{M}(n+1,k+1), the preimage of the zero section is a (k+1)(k+1)-dimensional \mathcal{M}-manifold with boundary.

We have now constructed well-defined maps

Ωkπn(Th(n+1,k+1))Ωk\Omega^{\mathcal{M}}_{k}\to\pi_{n}(\operatorname{Th}\mathcal{M}(n+1,k+1))\to\Omega^{\mathcal{M}}_{k}

whose composition is the identity. Therefore the first of these maps is injective.

Now notice that for every subcomplex K(P,f,N)K_{(P,f,N)} of Th(n,k)\operatorname{Th}\mathcal{M}(n,k), the (nk1)(n-k-1)-skeleton of NN does not intersect the image of ff, that is, it is contained in NN_{\infty}. Therefore, the (nk1)(n-k-1)-skeleton of Th(n,k)\operatorname{Th}\mathcal{M}(n,k) deformation retracts to the point at infinity, and so Th(n,k)\operatorname{Th}\mathcal{M}(n,k) is (nk1)(n-k-1)-connected. ∎

4.3. Modifications for bordism homology

Here we describe the modifications to the construction of Th(n,k)\operatorname{Th}\mathcal{M}(n,k) required to encode bordism classes of maps from \mathcal{M}-manifolds to a simplicial complex YY. We define the space (ThY)(n,k)(\operatorname{Th}\wedge Y)\mathcal{M}(n,k) as follows. We use the same subcomplexes K(P,f,N)K_{(P,f,N)}, but we index them by an additional datum, a simplicial map g:cPYg:cP\to Y. Given two vertices vPv\in P and vPv^{\prime}\in P^{\prime} such that f(cv)=f(cv)f(cv)=f^{\prime}(cv^{\prime}) and Nv=NvN_{v}=N_{v^{\prime}}, we identify KvK(P,f,N,g)K_{v}\subset K_{(P,f,N,g)} and KvK(P,f,N,g)K_{v^{\prime}}\subset K_{(P^{\prime},f^{\prime},N^{\prime},g^{\prime})} if in addition g|st(cv)=gst(cv)g|_{\operatorname{st}(cv)}=g^{\prime}_{\operatorname{st}(cv^{\prime})} under the obvious identification. Once again, all the cone points of all the cNcN_{\infty} are identified.

Note that (ThY)(n,k)(\operatorname{Th}\wedge Y)\mathcal{M}(n,k) comes with a natural projection map

p:(ThY)(n,k)Th(n,k)p:(\operatorname{Th}\wedge Y)\mathcal{M}(n,k)\to\operatorname{Th}\mathcal{M}(n,k)

which is finite-to-one if YY is a finite complex.

Proposition 4.3.

When n>2k+1n>2k+1, there is an injective map

Ωk(Y)πn((ThY)(n+1,k+1))\Omega^{\mathcal{M}}_{k}(Y)\to\pi_{n}((\operatorname{Th}\wedge Y)\mathcal{M}(n+1,k+1))

where Ωk(Y)\Omega^{\mathcal{M}}_{k}(Y) is the kkth \mathcal{M}-bordism homology group of YY. Moreover, πn((ThY)(n,k))\pi_{n}((\operatorname{Th}\wedge Y)\mathcal{M}(n,k)) is (nk1)(n-k-1)-connected.

Proof.

Given a triangulated \mathcal{M}-manifold XX and a simplicial map φ:XY\varphi:X\to Y, we construct a map SnTh(n,k)S^{n}\to\operatorname{Th}\mathcal{M}(n,k) as in Proposition 4.1, taking care that the embedding MnM\hookrightarrow\mathbb{R}^{n} is linear on a subdivision τ\tau of our triangulation. To lift this map to (ThY)(n,k)(\operatorname{Th}\wedge Y)\mathcal{M}(n,k), we need consistent choices of g:clk(v)Yg:c\operatorname{lk}(v)\to Y for each vertex vv of τ\tau. We obtain this by homotoping φ\varphi to a map which is simplicial on τ\tau. We use a similar method to convert a filling of (X,φ)(X,\varphi) to a nullhomotopy. This gives a well-defined map

Ωk(Y)πn((ThY)(n+1,k+1)).\Omega^{\mathcal{M}}_{k}(Y)\to\pi_{n}((\operatorname{Th}\wedge Y)\mathcal{M}(n+1,k+1)).

It remains to show that this map is injective. We will do this once again by constructing a retraction.

Given a simplicial map g:Sn(ThY)(n+1,k+1)g:S^{n}\to(\operatorname{Th}\wedge Y)\mathcal{M}(n+1,k+1), we can construct a manifold

X=g1((ThY)0(n+1,k+1))X=g^{-1}((\operatorname{Th}\wedge Y)_{0}\mathcal{M}(n+1,k+1))

as in Proposition 4.1 after forgetting the data about YY. Now we need to build a map from this manifold to YY which, when gg is the Thom map of some φ:XY\varphi:X\to Y, agrees up to homotopy with φ\varphi.

We do this by composing the induced map i:X(ThY)(n+1,k+1)i:X\to(\operatorname{Th}\wedge Y)\mathcal{M}(n+1,k+1) with a map g~\tilde{g} from the zero section (ThY)0(n+1,k+1)(\operatorname{Th}\wedge Y)_{0}\mathcal{M}(n+1,k+1) to YY. This will be a simplicial map from a triangulation of the zero section to the one-time barycentric subdivision bs(Y)\operatorname{bs}(Y), and it is constructed by induction on dimension. For a simplex Δ\Delta of YY, denote by Δ\Delta^{\vee} the dual subcomplex to Δ\Delta in bs(Y)\operatorname{bs}(Y). For each K(P,f,N,g)K_{(P,f,N,g)} we build g~:f1(N)Y\tilde{g}:f^{-1}(N)\to Y as follows:

  • For each (k+1)(k+1)-simplex Δ\Delta of PP, set g~(NΔ)=g(Δ)\tilde{g}(N_{\Delta})=g(\Delta)^{\vee} (note that g(Δ)g(\Delta)^{\vee} is a vertex of bs(Y)\operatorname{bs}(Y)).

  • For each kk-simplex Δ\Delta of PP, extend g~\tilde{g} to NΔN_{\Delta} so that its composition with the projection to Δ\Delta^{\vee} is homotopic rel endpoints to gg. So that g~\tilde{g} is well-defined, this map should depend only on NΔN_{\Delta} and g|Δg|_{\Delta}.

  • Continue this process for lower-dimensional simplices ΔP\Delta\subset P, at all points fixing maps in such a way that they depend only on NΔN_{\Delta} and g|Δg|_{\Delta}.

If we start with (X,φ)(X,\varphi), build the corresponding map Sn(ThY)(n,k)S^{n}\to(\operatorname{Th}\wedge Y)\mathcal{M}(n,k), and then use this to construct g~i\tilde{g}\circ i as above, then by construction g~iφ\tilde{g}\circ i\simeq\varphi. Therefore, as in Proposition 4.1, we have a retraction

Ωk(Y)πn((ThY)(n+1,k+1))Ωk(Y),\Omega^{\mathcal{M}}_{k}(Y)\to\pi_{n}((\operatorname{Th}\wedge Y)\mathcal{M}(n+1,k+1))\to\Omega^{\mathcal{M}}_{k}(Y),

and therefore the first map is injective.

Finally (ThY)(n,k)(\operatorname{Th}\wedge Y)\mathcal{M}(n,k) is (nk1)(n-k-1)-connected by the same reasoning as in Proposition 4.1. ∎

5. The Thom map and its nullhomotopy

In this section, we give the details of steps (2) and (3) of our outline: first use the embedding constructed in Theorem 3.1 to build a geometrically controlled Thom map to a finite subcomplex of Th(n,k)\operatorname{Th}\mathcal{M}(n,k), and then find a controlled nullhomotopy of this map in a larger finite subcomplex of Th(n+1,k+1)\operatorname{Th}\mathcal{M}(n+1,k+1).

Lemma 5.1.

For every nn, kk, LL, and \mathcal{M}, there is a finite, (nk1)(n-k-1)-connected subcomplex T(n,k,L)Th(n,k)T_{\mathcal{M}}(n,k,L)\subset\operatorname{Th}\mathcal{M}(n,k) such that the following holds. Let XX be a PL triangulated kk-dimensional \mathcal{M}-manifold with VV kk-simplices and bounded geometry of type LL, and let n2k+1n\geq 2k+1. Then there is a PL map f:SnT(n,k,L)f:S^{n}\to T_{\mathcal{M}}(n,k,L) such that f1(Th0(n,k))f^{-1}(\operatorname{Th}_{0}\mathcal{M}(n,k)) is PL homeomorphic to XX and the Lipschitz constant of ff (with respect to the standard simplexwise linear metric on T(n,k,L)T_{\mathcal{M}}(n,k,L)) is at most

C(n,L)V1nk(logV)2k+2.C(n,L)V^{\frac{1}{n-k}}(\log V)^{2k+2}.

Moreover, if YY is a finite simplicial complex and φ:XY\varphi:X\to Y is a simplicial map, then ff lifts to a PL map

f~:Snp1T(n,k,L)(ThY)(n,k)\tilde{f}:S^{n}\to p^{-1}T_{\mathcal{M}}(n,k,L)\subset(\operatorname{Th}\wedge Y)\mathcal{M}(n,k)

such that g~f~|f~1Th0(n,k)φ\tilde{g}\circ\tilde{f}|_{\tilde{f}^{-1}\operatorname{Th}_{0}\mathcal{M}(n,k)}\simeq\varphi, where g~:(ThY)0(n,k)Y\tilde{g}:(\operatorname{Th}\wedge Y)_{0}\mathcal{M}(n,k)\to Y is the projection defined in the proof of Proposition 4.3 and p:(ThY)(n,k)Th(n,k)p:(\operatorname{Th}\wedge Y)\mathcal{M}(n,k)\to\operatorname{Th}\mathcal{M}(n,k) is the projection defined at the beginning of §4.3.

Proof.

Let Λ(L)\Lambda(L) be the lattice in n\mathbb{R}^{n} specified in condition (iv) of Theorem 3.1. Let τ\tau be a Λ(L)\Lambda(L)-invariant triangulation of n\mathbb{R}^{n} with the following properties:

  • The edge lengths are at most α(n,L)2\frac{\alpha(n,L)}{2}.

  • It is transverse to every possible simplex of an embedding satisfying conditions (i)–(iv) of Theorem 3.1. (This is possible since there are finitely many such possible simplices up to the action of Λ(L)\Lambda(L).)

Now let T(n,k,L)Th(n,k)T_{\mathcal{M}}(n,k,L)\subseteq\operatorname{Th}\mathcal{M}(n,k) be the union of subcomplexes K(P,f,N)K_{(P,f,N)} such that:

  • PP is an admissible link in n\mathcal{L}_{n}.

  • ff is an embedding of cPcP satisfying conditions (i)–(iv) of Theorem 3.1.

  • NN is the subcomplex of τ\tau consisting of simplices which intersect f(cP)f(cP) and are at distance at least α(n,L)2\frac{\alpha(n,L)}{2} from f(P)f(P).

The number of such combinations is finite, and therefore T(n,k,L)T_{\mathcal{M}}(n,k,L) is a finite complex. Moreover, it’s (nk1)(n-k-1)-connected since all the K(P,f,N)K_{(P,f,N)} are (nk1)(n-k-1)-connected and glued together along (nk1)(n-k-1)-connected subcomplexes.

Now by Theorem 3.1, there is a subdivision XX^{\prime} of XX which embeds linearly into the nn-dimensional Euclidean ball of radius

R=C(n,L)V1nk(logV)2k+2R=C(n,L)V^{\frac{1}{n-k}}(\log V)^{2k+2}

such that the embedding satisfies conditions (i)–(iv). Conditions (i) and (ii) imply that the triangulation τ\tau satisfies the conditions of Lemma 4.2. This embedding induces a map from the RR-ball to T(n,k,L)T_{\mathcal{M}}(n,k,L) which is simplicial on a slight subdivision of τ\tau (depending only on nn), and hence C(n,L)C(n,L)-Lipschitz. Since the boundary of the RR-ball is mapped to \infty, this map extends to a sphere. If we rescale this to be the unit sphere, the Lipschitz constant becomes C(n,L)RC(n,L)R.

Finally, in the case of a map φ:XY\varphi:X\to Y, we can homotope this map to a map φ:XY\varphi^{\prime}:X\to Y which is simplicial on XX^{\prime}, and then decide the lift to (ThY)(n,k)(\operatorname{Th}\wedge Y)\mathcal{M}(n,k) based on the behavior of φ\varphi^{\prime} near each vertex. ∎

Note that for any finite simplicial complex YY, since pp is finite-to-one, p1T(n,k,L)p^{-1}T_{\mathcal{M}}(n,k,L) is also a finite complex. We now embed p1T(n,k,L)p^{-1}T_{\mathcal{M}}(n,k,L) in a larger, but still finite subcomplex of (ThY)(n+1,k+1)(\operatorname{Th}\wedge Y)\mathcal{M}(n+1,k+1) in which we can nullhomotope what needs to be nullhomotoped. This construction is essentially purely formal.

Lemma 5.2.

There is a finite simplicial complex U,Y(n,k,L)p1T(n,k,L)U_{\mathcal{M},Y}(n,k,L)\supseteq p^{-1}T_{\mathcal{M}}(n,k,L) equipped with a simplicial map

ι:U(n,k,L)(ThY)(n+1,k+1)\iota:U_{\mathcal{M}}(n,k,L)\to(\operatorname{Th}\wedge Y)\mathcal{M}(n+1,k+1)

which extends the inclusion ι0:p1T(n,k,L)(ThY)(n,k)\iota_{0}:p^{-1}T_{\mathcal{M}}(n,k,L)\hookrightarrow(\operatorname{Th}\wedge Y)\mathcal{M}(n,k), such that:

  1. (i)

    U,Y(n,k,L)U_{\mathcal{M},Y}(n,k,L) is (nk1)(n-k-1)-connected.

  2. (ii)

    If a map Snp1T(n,k,L)S^{n}\to p^{-1}T_{\mathcal{M}}(n,k,L) is nullhomotopic in (ThY)(n+1,k+1)(\operatorname{Th}\wedge Y)\mathcal{M}(n+1,k+1), then it is nullhomotopic in U,Y(n,k,L)U_{\mathcal{M},Y}(n,k,L).

Proof.

We know that (ThY)(n+1,k+1)(\operatorname{Th}\wedge Y)\mathcal{M}(n+1,k+1) and p1T(n,k,L)p^{-1}T_{\mathcal{M}}(n,k,L) are both (nk1)(n-k-1)-connected. We will build U,Y(n,k,L)U_{\mathcal{M},Y}(n,k,L) by adding cells of dimension n+1n+1 and therefore not change this.

Since p1T(n,k,L)p^{-1}T_{\mathcal{M}}(n,k,L) is a simply connected finite complex, its higher homotopy groups are finitely generated. In particular, the kernel of the induced map

(ι0):πn(p1T(n,k,L))πn((ThY)(n+1,k+1))(\iota_{0})_{*}:\pi_{n}(p^{-1}T_{\mathcal{M}}(n,k,L))\to\pi_{n}((\operatorname{Th}\wedge Y)\mathcal{M}(n+1,k+1))

is finitely generated. Choose simplicial maps fi:Snp1T(n,k,L)f_{i}:S^{n}\to p^{-1}T_{\mathcal{M}}(n,k,L) representing the generators of the kernel, as well as simplicial fillings

Fi:Dn+1(ThY)(n+1,k+1).F_{i}:D^{n+1}\to(\operatorname{Th}\wedge Y)\mathcal{M}(n+1,k+1).

(In this notation, we are suppressing the simplicial structures on the sphere and disk on which these maps are defined.)

We build the simplicial complex U,Y(n,k,L)U_{\mathcal{M},Y}(n,k,L) by gluing these simplicial disks onto p1T(n,k,L)p^{-1}T_{\mathcal{M}}(n,k,L) using the maps fif_{i} as attaching maps. Then the map ι\iota is constructed by extending ι0\iota_{0} to the new cells using the maps FiF_{i}. ∎

Now put the standard simplexwise linear metric on U,Y(n,k,L)U_{\mathcal{M},Y}(n,k,L); this also restricts to a metric on p1T(n,k,L)p^{-1}T_{\mathcal{M}}(n,k,L). Then [CDMW18, Corollary 4.3] yields the following fact:

Lemma 5.3.

Suppose that n2k+2n\geq 2k+2, and let f:Snp1T(n,k,L)f:S^{n}\to p^{-1}T_{\mathcal{M}}(n,k,L) be an LL-Lipschitz map such that ι0f\iota_{0}\circ f is nullhomotopic. Then ff is nullhomotopic in U,Y(n,k,L)U_{\mathcal{M},Y}(n,k,L) via a C(L+1)C(L+1)-Lipschitz nullhomotopy, where CC is a constant which depends on \mathcal{M}, nn, kk, LL, and YY.

6. Completing the proof

We now put the earlier results together to prove the main theorem, which we restate here. This section implements steps (4) and (5) of the proof outline.

Theorem.

Let \mathcal{M} be a category of manifolds with prescribed singularities, kk and LL integers, ε>0\varepsilon>0, and YY a finite simplicial complex. Then there are constants C=C(,k,L,ε,Y)C=C(\mathcal{M},k,L,\varepsilon,Y) and L=L(,k,L)L^{\prime}=L^{\prime}(\mathcal{M},k,L) such that the following holds.

Let XX be a triangulated kk-dimensional \mathcal{M}-manifold with VV kk-simplices and bounded geometry of type LL, and let φ:XY\varphi:X\to Y be a simplicial map such that (M,φ)(M,\varphi) represents the zero class in Ωk(Y)\Omega^{\mathcal{M}}_{k}(Y). Then there is a (k+1)(k+1)-dimensional \mathcal{M}-manifold WW with W=X\partial W=X with at most CV1+εCV^{1+\varepsilon} simplices and bounded geometry of type LL^{\prime}, and a simplicial map φ~:WY\tilde{\varphi}:W\to Y extending φ\varphi.

Proof.

Let n2k+2n\geq 2k+2. The Thom construction from Proposition 4.1 tells us that the simplicial map f:Snp1T(n,k,L)f:S^{n}\to p^{-1}T_{\mathcal{M}}(n,k,L) induced by any PL embedding of XX is nullhomotopic in (ThY)(n+1,k+1)(\operatorname{Th}\wedge Y)\mathcal{M}(n+1,k+1). Combining Lemmas 5.1 and 5.3, we see that there is a map F:Dn+1U,Y(n,k,L)F:D^{n+1}\to U_{\mathcal{M},Y}(n,k,L) such that F|Sn=fF|_{S^{n}}=f and λ=LipF=O(V1/q(logV)2k+2))\lambda=\operatorname{Lip}F=O(V^{1/q}(\log V)^{2k+2})).

By a quantitative simplicial approximation theorem, see e.g. [CDMW18, Prop. 2.1], we can simplicially approximate FF on a subdivided triangulation of Dn+1D^{n+1} at scale 1/λ\sim 1/\lambda. This gives us a simplicial nullhomotopy of a simplicial approximation f~\tilde{f} of ff.

We would like to get back to a nullhomotopy of ff. Notice that f~=fσ\tilde{f}=f\circ\sigma where σ:SnSn\sigma:S^{n}\to S^{n} is a simplicial map from the subdivision to the original triangulation. Moreover, f~\tilde{f} can be constructed so that for each simplex δ\delta of the original triangulation, σ1(δ)\sigma^{-1}(\delta) is a contractible subcomplex. So we fix a triangulation of Sn×[0,1]S^{n}\times[0,1] as follows. First triangulate Sn×[0,1]S^{n}\times[0,1] using the usual product triangulation, where top-dimensional simplices take the form

[(δ0,0),(δ1,0),,(δk,0),(δk,1),(δk+1,1),,(δn,1)][(\delta_{0},0),(\delta_{1},0),\ldots,(\delta_{k},0),(\delta_{k},1),(\delta_{k+1},1),\ldots,(\delta_{n},1)]

for each 0kn0\leq k\leq n and each nn-simplex [δ0,,δn][\delta_{0},\ldots,\delta_{n}] of the subdivided triangulation of SnS^{n}. Now take the mapping cylinder of σ\sigma, which is a simplicial quotient of this triangulation. Our extra assumption about f~\tilde{f} implies that the mapping cylinder is still a triangulation of Sn×[0,1]S^{n}\times[0,1], with the subset Sn×{0}S^{n}\times\{0\} given the subdivided triangulation and Sn×{1}S^{n}\times\{1\} given the original triangulation. Then the map σprojSn\sigma\circ\text{proj}_{S^{n}} on the product triangulation induces a simplicial homotopy between ff and f~\tilde{f} on the mapping cylinder. So we get a simplicial nullhomotopy of ff by attaching this to our disk as a collar.

Now Lemma 5.2 gives us a simplicial map

ι:U,Y(n,k,L)(ThY)(n+1,k+1).\iota:U_{\mathcal{M},Y}(n,k,L)\to(\operatorname{Th}\wedge Y)\mathcal{M}(n+1,k+1).

Then by the argument of Proposition 4.1, (ιF)1(ThY)0(n,k)(\iota\circ F)^{-1}(\operatorname{Th}\wedge Y)_{0}\mathcal{M}(n,k) is an \mathcal{M}-manifold WW whose boundary is XX. Since the image of ι\iota is a finite subcomplex of (ThY)(n+1,k+1)(\operatorname{Th}\wedge Y)\mathcal{M}(n+1,k+1), the link types of WW come from a finite set depending on \mathcal{M}, nn, kk, LL, and YY. In particular, there is some L(,n,k,L,Y)L^{\prime}(\mathcal{M},n,k,L,Y) which bounds the geometry of these link types.

Moreover, there is a triangulation of WW with a bounded number of simplices per simplex of Dn+1D^{n+1}: we obtain it by pulling back from the image under ιF\iota\circ F the triangulation on which the map

g~:(ThY)0(n,k)Y\tilde{g}:(\operatorname{Th}\wedge Y)_{0}\mathcal{M}(n,k)\to Y

is defined; incidentally this gives us a simplicial map Φ=g~ιF:WY\Phi=\tilde{g}\circ\iota\circ F:W\to Y. With this triangulation, WW consists of O(V1+k+1nk(logV)(2k+2)(n+1))O\bigl{(}V^{1+\frac{k+1}{n-k}}(\log V)^{(2k+2)(n+1)}\bigr{)} simplices. Given any ε>0\varepsilon>0, for sufficiently large nn, this is O(V1+ε)O(V^{1+\varepsilon}).

At this point in the proof, LL^{\prime} depends on nn and YY as well as kk, LL, and the category. We will fix this at the expense of increasing the constant CC, by an inductive process which replaces certain subcomplexes of WW with other subcomplexes. The strategy is similar to the proof of Corollary 2.1.

We start by taking a single barycentric subdivision of WW. Let ViV_{i} be the set of vertices in the interior (outside W=M\partial W=M) of the resulting triangulation that originate as barycenters of ii-simplices. Notice that the star of a vertex vViv\in V_{i} is of the form SicP(v)S^{i}cP(v) where P(v)P(v) is the (ki1)(k-i-1)-dimensional link of the original ii-simplex. We will modify WW by replacing cP(v)cP(v) with other fillings of P(v)P(v).

We start with i=k2i=k-2. If \mathcal{M}-manifolds are pseudomanifolds, then for vVk2v\in V_{k-2}, P(v)P(v) is a disjoint union of circles. This can be filled with disks with 77 faces to a vertex, at most 33 faces to a vertex on the boundary, and linearly many faces in terms of |P(v)|\lvert P(v)\rvert; let Q(v)Q(v) be this collection of disks. For each vVk2v\in V_{k-2}, we replace Sk2cP(v)WS^{k-2}cP(v)\subset W with Sk2Q(v)S^{k-2}Q(v) to get a new \mathcal{M}-manifold W2W_{2}. Notice that although we have eliminated the vertices in Vk2V_{k-2}, the sets of vertices ViV_{i} for i<k2i<k-2 are unchanged from WW, and their stars are still of the form SicP2(v)S^{i}cP_{2}(v), though the 22-complex P2(v)P_{2}(v) is different from P(v)P(v). Since P(v)P(v) consisted of cP(w)cP(w)s for wVk2w\in V_{k-2}, meeting three to a vertex, the parts of P2(v)P_{2}(v) outside MM have at most 99 faces meeting at a vertex.

Suppose now, by induction, that we have built a complex Wi1W_{i-1} in which we have replaced the stars of vertices wVk(i1)w\in V_{k-(i-1)} with subcomplexes of the form Sk(i1)Q(w)S^{k-(i-1)}Q(w), where Q(w)Q(w) is a subcomplex of bounded geometry of type some Li1(,k,L)L^{\prime}_{i-1}(\mathcal{M},k,L). Then for each vVkiv\in V_{k-i}, its star is a complex SicPi1(v)S^{i}cP_{i-1}(v) where the geometry of the (i1)(i-1)-complex Pi1(v)P_{i-1}(v) is bounded by Li=max(L,iLi1(,k,L))L_{i}=\max(L,iL^{\prime}_{i-1}(\mathcal{M},k,L)). Moreover, by induction and since we had an original bound on the complexity of link types, the volume of Pi1(v)P_{i-1}(v) is bounded as a function of \mathcal{M}, nn, kk, LL, and YY. We can then replace cPi1(v)cP_{i-1}(v) with a complex Q(v)Q(v) (a fixed \mathcal{M}-manifold whose boundary is Pi1(v)P_{i-1}(v)) such that:

  • The geometry of Q(v)Q(v) is bounded by Li=L(,2i,i1,Li,)L^{\prime}_{i}=L^{\prime}(\mathcal{M},2i,i-1,L_{i},*).

  • The volume of Q(v)Q(v) is bounded as a function of \mathcal{M}, nn, kk, LL, and YY (simply because the set of possible Pi1(v)P_{i-1}(v)s is finite and depends only on these parameters).

This completes the inductive step to construct YY.

At each stage, we extend the map Φ\Phi over Q(v)Q(v) by mapping each interior vertex to Φ(v)\Phi(v) and extending by linearity. Since we are replacing the star of vv, this is a well-defined map.

At the end of the induction, we have replaced WW with an \mathcal{M}-manifold whose boundary is MM and whose geometry still depends on \mathcal{M}, kk, and LL, but not on nn. Moreover, we have multiplied the volume by at most a constant depending on \mathcal{M}, nn, kk, and LL. ∎

Appendix A PL versus smooth bounded geometry

Here we discuss the relationship between bounded geometry for smooth and for PL manifolds. Specifically, we show that for smooth manifolds these are closely related notions:

Theorem A.1.

Let MM be a Riemannian nn-manifold with sectional curvatures |K|1|K|\leq 1 and injectivity radius at least 11, and volume VV. Then there is a triangulation of MM with at most C1(n)VC_{1}(n)V simplices and at most C2(n)C_{2}(n) simplices incident to each vertex.

Theorem A.2.

Let MM be a smoothable PL nn-manifold equipped with a PL triangulation with at most LL simplices incident to any vertex. Then for constants C1,C2,C3C_{1},C_{2},C_{3} depending on nn and LL, every smoothing of MM has an associated Riemannian metric gg such that:

  1. (1)

    Every simplex of (M,g)(M,g) has volume at most C1C_{1}.

  2. (2)

    (M,g)(M,g) has sectional curvature |K|C2|K|\leq C_{2} and injectivity radius at least C31C_{3}^{-1}.

The tools needed to prove Theorem A.1 are given in [CMS84]; here we give an outline, referring to that paper for the more difficult parts.

Proof of Theorem A.1..

We start by fixing an ε\varepsilon-net of points {x1,,xN}\{x_{1},\ldots,x_{N}\} on MM, for some ε<1/2\varepsilon<1/2 possibly depending on nn: by definition, the ε\varepsilon-balls around the xix_{i} are disjoint and the 2ε2\varepsilon-balls cover MM. The bounded geometry condition implies that the exponential map onto each of these balls is a diffeomorphism and not too distorted. Moreover, the lower bound on curvature implies that B4ε(xi)B_{4\varepsilon}(x_{i}) intersects at most some C0(n)C_{0}(n) of the 4ε4\varepsilon-balls around the other xjx_{j}. Thus there is a (C0+1)(C_{0}+1)-coloring of the xix_{i} into sets A0,,AC0A_{0},\ldots,A_{C_{0}} so that no two 4ε4\varepsilon-balls of the same color intersect.

We can triangulate each B3ε(xi)B_{3\varepsilon}(x_{i}) with the image under the exponential map of some standard Euclidean mesh; the distortion of each simplex in each of these triangulations is bounded by some constant depending only on nn. Now [CMS84, Lemma 6.3] shows that we can take any two such local boundedly distorted meshes and connect them by a mesh which still has bounded (if somewhat worse) distortion. We apply this lemma first to interpolate between the triangulations of B3ε(xi)B_{3\varepsilon}(x_{i}) and B3ε(xj)B_{3\varepsilon}(x_{j}) for xiA0x_{i}\in A_{0} and xjA1x_{j}\in A_{1}; then to add in the 3ε3\varepsilon-balls around the points in A2A_{2}; and so on. We thus worsen the geometry of the triangulation a total of C0C_{0} times. In the end, we get a triangulation of all of MM for which the geometry of the simplices still depends only on nn. ∎

Theorem A.2 follows easily from the fact that the groups Θn\Theta_{n} of exotic spheres are all finite (except perhaps in dimension 44, where any exotic spheres are not PL spheres).

Proof of Theorem A.2..

Let (M,τ)(M,\tau) be a triangulated nn-manifold, and suppose that τ\tau has at most LL simplices adjacent to every vertex. In particular, the star of each kk-simplex can have a finite number aka_{k} of possible configurations.

We will show that there is a finite set Γ\Gamma of Riemannian metrics on the nn-simplex, depending only on nn and LL, such that every smoothing of MM has a metric which, when restricted to each simplex, is isometric to one of the metrics in Γ\Gamma. This is sufficient to prove the theorem because the constants C1C_{1}, C2C_{2}, and C3C_{3} are then obtained by maximizing over the finite set of possible local configurations.

We construct Γ\Gamma by induction, giving a process that builds all possible smoothings of MM as Riemannian manifolds with metrics on simplices chosen from Γ\Gamma. First let τ\tau^{\prime} be the two times barycentric subdivision of τ\tau, and let

M0M1Mn=MM_{0}\subset M_{1}\subset\cdots\subset M_{n}=M

be open submanifolds of MM such that each MkM_{k} includes the star of the kk-skeleton of τ\tau in τ\tau^{\prime}. Then M0M_{0} is a disjoint union of balls around the vertices of xx, and going from Mk1M_{k-1} to MkM_{k} means gluing in copies of Dk×DnkD^{k}\times D^{n-k} for each kk-simplex of MM.

Clearly M0M_{0} has a unique smoothing. For every possible link of a vertex consisting of at most LL simplices, we fix a Riemannian metric on its cone; this gives a Riemannian metric on M0M_{0}.

Now suppose we have a smoothing of Mk1M_{k-1} equipped with a Riemannian metric gg, in which each nn-simplex of τ\tau^{\prime} is isometric to one of a finite list depending on nn and LL. For every kk-simplex Δ\Delta of τ\tau, if the smoothing extends to stτ(Δ)\operatorname{st}_{\tau^{\prime}}(\Delta), then the set of such extensions is in bijection with the finite group Θk\Theta_{k} of exotic kk-spheres. We fix Riemannian metrics on all these extensions, depending on the combinatorial structure of stτ(Δ)\operatorname{st}_{\tau^{\prime}}(\Delta), the metric on stτ(Δ)\operatorname{st}_{\tau^{\prime}}(\partial\Delta), and the element of Θk\Theta_{k}. By induction, this data takes a finite set of values depending only on kk and LL.

After the nnth step, for every smoothing of MM, we have obtained a metric with these properties. ∎

Appendix B Remarks on Morse complexity

In this appendix we make some remarks about the Morse complexity, that is, the minimum number of critical points of a Morse function on a manifold MM, which we denote by MC(M)\operatorname{MC}(M). We can also discuss the relative Morse complexity of a manifold with boundary, that is, the minimum number of critical points of a Morse function which is constant on the boundary. We show that Morse complexity, thought of as a complexity measure, behaves very differently from the minimum number of simplices in a triangulation of bounded geometry.

For a simply connected manifold MM of dimension >4>4, Smale [Sma62, Theorem 6.1] showed that the homology of MM determines MC(M)\operatorname{MC}(M) in the most obvious way: the number of ii-dimensional critical points is

p(i)+q(i)+q(i1),p(i)+q(i)+q(i-1),

where p(i)p(i) is the rank of the iith homology and q(i)q(i) is the minimal number of generators for the torsion. Whitehead torsion shows that this cannot be exactly correct in the non–simply connected case, at least for relative Morse complexity: a manifold with boundary has relative Morse complexity zero if and only if it is a product N×[0,1]N\times[0,1]; but the property of being a product is obstructed by the torsion and therefore isn’t purely homotopy theoretic. On the other hand, we have the following result (which might be well known):

Theorem B.1.

If m=dim(M)>5m=dim(M)>5, and f:MMf:M^{\prime}\to M is a simple homotopy equivalence, then MC(M)=MC(M)\operatorname{MC}(M^{\prime})=\operatorname{MC}(M).

Proof.

As in the proof of the hh-cobordism theorem, we can assume that MM is given a self-indexing Morse function (i.e. a Morse function so that critical points of index kk arise before ones of index k+1k+1). We also assume that there is just one critical point of index 0 and of index mm.

We construct a Morse function on MM^{\prime} as follows. In index 2\leq 2 we mimic the 22-complex M2M_{2} associated to the Morse function of MM. That is, we put f(M2)f(M_{2}) in general position, turning it into an embedding, and then take a suitable function on its regular neighborhood. We do the same thing with the dual Morse function f-f, creating a Morse function on a disjoint submanifold.

Now what remains is a simple homotopy equivalence (which we still call ff) between two manifolds with boundary, say (A,±A)(A^{\prime},\partial_{\pm}A^{\prime}) and (A,±A)(A,\partial_{\pm}A), where the manifolds and their boundaries all have fundamental group π\pi and AA is given a Morse function (locally constant on the boundary) whose critical points all have index between 33 and m3m-3. This Morse function yields a relative handlebody structure on the manifold.

Now we perform an induction on the handles: at each step, we find a handle in AA^{\prime} whose image is the lowest handle of AA, and generate a new simple homotopy equivalence by cutting them both out. Write (H,H)A(H,\partial H)\subset A for this first handle, and Hc=AH¯H^{c}=\overline{A\setminus H}. Now we apply [Wal99, Theorem 12.1]. By this theorem, ff can be homotoped so that it is transverse to H\partial H, and so that it induces a simple homotopy equivalence of triads (A;f1H,f1Hc)(A;H,Hc)(A\textquoteright;f^{-1}H,f^{-1}H^{c})\to(A;H,H^{c}). Then f1Hf^{-1}H is a manifold homotopy equivalent to (Bm;Bk×Smk1Sk1×Bmk;Sk1×Smk1)(B^{m};B^{k}\times S^{m-k-1}\cup S^{k-1}\times B^{m-k};S^{k-1}\times S^{m-k-1}); but by simply connected surgery on triads, any such homotopy equivalence is in fact a diffeomorphism, so f1HAf^{-1}H\subset A^{\prime} is a handle.

Once one has pulled back all the handles of AA to AA^{\prime}, what lies between +A\partial_{+}A^{\prime} and this union is an ss-cobordism, and therefore a product, so we can easily extend the Morse function to all of AA^{\prime}. ∎

Remark.

If one does not first make the Morse function on MM self-indexing, it is possible to give examples where one cannot find a “corresponding” function on MM^{\prime}.

In the non–simply connected case, actually understanding this number is not so easy. If NMN\to M is a dd-fold cover, then obviously MC(N)dMC(M)\operatorname{MC}(N)\leq d\operatorname{MC}(M). One can also use Betti numbers of finite covers to give lower bounds on MC(M)\operatorname{MC}(M). These are typically hard to get information about except in characteristic 0, where one usually gets this information from L2L^{2} Morse inequalities, which describe the asymptotic behavior of finite covers.

One can ask when the obvious bound fails to be sharp, for instance:

Question.

If MM is a locally symmetric space associated to a group GG without discrete series (equivalently M=KG/ΓM=K\setminus G/\Gamma has Euler characteristic 0\neq 0), is it true that MM has a sequence of finite covers MkM_{k} of index kk such that MC(Mk)/k0\operatorname{MC}(M_{k})/k\to 0?

The optimistic conjecture that the answer is yes very generally whenever L2L^{2} Betti numbers don’t prevent this is disproved by the calculations in [AOS21]. On the other hand, a recent paper of Frączyk, Mellick and Wilkens [FMW23] shows that when GG has rank >1>1, the number of generators of the fundamental group grows sublinearly as a function of the covolume. Conversely, Avramidi and Delzant [AD23] show that the Morse complexity of a hyperbolic manifold is bounded below by an increasing function of its injectivity radius.

If one asked instead about the number of simplices in a triangulation, of course the theory of simplicial norm [Gro82, LS06] prevents the question from having a positive answer. In fact this theory shows that whenever there is a map MkMM_{k}\to M of degree kk, not necessarily a covering map, MkM_{k} must have Ω(k)\Omega(k) simplices.

Similarly, one can ask:

Question (Gromov, cf. [Gro96, §812\frac{1}{2}]).

When does a nice manifold MM, e.g. negatively curved or locally symmetric, have a sequence MkMM_{k}\to M of degree kk manifolds mapping to it, with MC(Mk)/k0\operatorname{MC}(M_{k})/k\to 0?

The following proposition shows that the answer is always for odd-dimensional manifolds.

Proposition B.2.

If dim(M)\dim(M) is odd, then MM always has a sequence of kk-fold finite branched covers (for any index kk) with uniformly bounded Morse complexity.

This is an immediate consequence of Lawson’s result [Law78] that all odd-dimensional closed manifolds have open book decompositions (in dimension 33, this actually goes back to Alexander [Ale23]). The branched covers obtained by stitching multiple copies of the book together obviously have the desired property.

Question.

Which KG/ΓK\setminus G/\Gamma have open book decompositions?

Obviously negative answers to Gromov’s question give negative answers to this question as well. Thus, for example, products of surfaces do not have open book decompositions. (In principle, this should be a calculation in an algebraic theory invented by Quinn [Qui79]; unfortunately those groups of non-symmetric quadratic forms seem very difficult to understand.) We will later explain the one method we know of proving such bounds, taken from [Gro96].

Gromov’s question can be interpreted as asking about MC-minimal representatives for the fundamental class in real homology, in analogy to the simplicial volume (see [Gro96, §812\frac{1}{2}] and [Gro09, §3.2]).

One can then ask analogous questions about integral homology or even bordism. Let XX be a finite complex. We consider the bordism group Ωd(X)\Omega_{d}(X) of smooth oriented dd-dimensional manifolds mapping to XX and ask about the rough size of its elements. In light of the proposition above, the following is not surprising:

Proposition B.3.

For any odd dd, all of Ωd(X)\Omega_{d}(X) has representatives of uniformly bounded Morse complexity.

Proof.

We will show that if f:MXf:M\to X represents an element α\alpha of bordism, then a kk-fold branched cover of MM represents kαk\alpha. The result then follows as above from the finite generation of Ωd(X)\Omega_{d}(X).

To see this fact, first suppose that the binding is empty, in other words MM fibers over S1S^{1} and the branched cover is just a kk-fold cover p:M~Mp:\tilde{M}\to M. Let h:ΣS1h:\Sigma\to S^{1} be a cobordism over S1S^{1} between kk copies of the identity map and a connected kk-fold cover. Then the pullback NN of the bundle MS1M\to S^{1} along hh, equipped with the map fh~:NXf\circ\tilde{h}:N\to X, is a bordism between kfkf and fpf\circ p.

In general, the difference of two open books with the same binding is cobordant to a manifold that fibers over the circle with fiber the union of two pages, one from each book. Applying this to ff and f-f gives us a representative of 2[f]2[f] which fibers over the circle. ∎

For even dd the answer depends only on the fundamental group of XX by the theorem below. This does not seem to be obvious even when X is simply connected, but Theorem B.1 and the connection between surgery and bordism imply:

Theorem B.4.

The Morse complexity of a class in Ωd(X)\Omega_{d}(X) is determined up to a uniformly bounded error by its “higher signature class image” in iHd4i(K(π1(X),1);)\bigoplus_{i}H_{d-4i}(K(\pi_{1}(X),1);\mathbb{Q}), determined by the numbers f(α)L(M),[M]\langle f^{*}(\alpha)\cup L(M),[M]\rangle where L(M)L(M) is the Hirzebruch LL-class and α\alpha runs over the nontrivial group cohomology classes of π1(X)\pi_{1}(X).

For example, if XX is simply connected, the result is that MC(β)=sign(β)+O(1)\operatorname{MC}(\beta)=\operatorname{sign}(\beta)+O(1), where sign\operatorname{sign} is the signature of any representative manifold for the class β\beta. That no other classes obstruct is a consequence of Winkelnkemper’s theorem [Win73] (reproved by Quinn [Qui79]) that any simply connected manifold of dimension >4>4 with signature 0 is an open book. The signature obviously obstructs because signature is a cobordism invariant and gives lower bounds for Betti numbers.

The theorem follows directly from finite generation of homology and Theorem B.1, because by a classical theorem (see e.g. [Wei23, Ch. 4]), modulo torsion, all elements whose higher signature class is 0 are represented by elements of the form MMM^{\prime}-M where MMM^{\prime}\to M is a simple homotopy equivalence.

Remark.

Not all higher signatures give rise to lower bounds on Morse complexity. For example, Gromov’s question can be rephrased as asking whether the top cohomology class does, in the case that X=MX=M. When X=TnX=T^{n}, it obviously does not, for example because tori have self-covers of arbitrary degree.

A source of higher signatures which do give lower bounds is from local systems. If LL is a local system of (perhaps skew-)symmetric or Hermitian quadratic forms on K(Γ,1)K(\Gamma,1), then for a 2k2k-manifold MM equipped with a map MK(Γ,1)M\to K(\Gamma,1) one gets an inner product pairing on Hk(M;L)H^{k}(M;L) whose signature is an invariant of bordisms in K(Γ,1)K(\Gamma,1). This signature obviously gives a lower bound on MC(M)\operatorname{MC}(M).

The signature of local systems arises in showing that signature is not multiplicative on bundles [Ati69, Mey72] and (products of) Atiyah’s examples on surfaces of genus > 1 show the following:

Proposition B.5.

Let XX be a product of surfaces of genus at least 22. Then the signature associated to the class [Σ][\Sigma] of any one of the surfaces gives a lower bound on Morse complexity. (On the other hand, the signature associated to the cup product of 1-dimensional classes from different surfaces does not.)

Proof.

First suppose XX is a single surface, and let f:M4n+2Xf:M^{4n+2}\to X be a map. Let SS be another surface of genus at least 22. According to Atiyah [Ati69, §4], bundles EBE\to B with fiber SS satisfy the equation

sign(E)=c1(T)L(B),[B],\operatorname{sign}(E)=\langle c_{1}(T)\cup L(B),[B],

where T is the vector bundle induced by the monodromy of H1(S;)H^{1}(S;\mathbb{R}).

Let p:EXp:E\to X be an SS-bundle such that EE has nonzero signature, as constructed by Atiyah. Then the above identity implies that

f[Σ]L(M),[M]=sign(fE)sign(E).\langle f^{*}[\Sigma]\cup L(M),[M]\rangle=\frac{\operatorname{sign}(f^{*}E)}{\operatorname{sign}(E)}.

Since sign(f(E))\operatorname{sign}(f^{*}(E)) gives a lower bound on rk(H2n(M))+rk(H2n+2(M))\operatorname{rk}(H^{2n}(M))+\operatorname{rk}(H^{2n+2}(M)), it also gives a lower bound on the Morse complexity.

For the case of products of surfaces, it suffices to consider products of examples of this form.

Finally, if α\alpha is the product of one-dimensional classes from different surfaces, then it is realized by a map of a torus to XX. The signature of maps to this torus does not give a lower bound on Morse complexity, as already discussed. ∎

Lusztig, in his thesis [Lus72, §5], showed that there are enough local systems on Sp(2n,)\operatorname{Sp}(2n,\mathbb{Z}) to detect all of the rational cohomology of BSp(2n,)B\operatorname{Sp}(2n,\mathbb{R}). We deduce that these higher signatures give lower bounds on Morse complexity. In particular for any manifold which represents a nontrivial cycle in this group, its fundamental class in real homology has nonzero complexity. Gromov, in [Gro96, §812\frac{1}{2}], gives other examples, including all products of even-dimensional hyperbolic manifolds.

We now connect these ideas to the problem studied in this paper.

Example B.6.

According to Browder and Livesay [BL73] there are infinitely many smooth manifolds simple homotopy equivalent to 4k+3\mathbb{RP}^{4k+3} that are cobordant over B2B\mathbb{Z}_{2}. (Since Ωi(B2)\Omega_{i}(B\mathbb{Z}_{2}) is finite for odd ii, the last condition does not need to be added explicitly, but nevertheless such cobordisms can be explicitly constructed.) Chang and Weinberger [CW03] extend this construction to any oriented closed manifold M4k+3M^{4k+3} whose fundamental group contains torsion.

By Theorem B.1, manifolds in this family all have the same Morse complexity. However, by work of Hirzebruch [Hir68], the Browder–Livesay invariant can be equivalently defined as the twice the signature of the cobordism between them minus the signature of its twofold cover (in the more general situation of Chang and Weinberger, this is replaced by an L2L^{2} signature of the cobordism). This gives a lower bound on the Morse complexity of the cobordism, and in particular shows that for non–simply connected targets one cannot always bound the Morse complexity of the smallest nullcobordism of a manifold in terms of the Morse complexity of the manifold.

Using surgery theory and products of surface groups one can give examples with torsion-free fundamental group as well. Let MM be a 55-manifold whose fundamental group π\pi is the product of three genus 22 surface groups. The product of these surfaces with the Milnor manifold constructed from kk copies of E8E_{8}, mapping to S8S^{8}, gives an element of the group L14(π)L6(π)L_{14}(\pi)\cong L_{6}(\pi). Using Wall’s realization theorem for this element of L6(π)L_{6}(\pi), one obtains a cobordism from MM to MM^{\prime}, a simple homotopy equivalent manifold, whose surgery obstruction is the given element. Now the argument is the same as above: the normal cobordism “needs” to have enough handles that a product of surface bundles over it can have large signature.

References

  • [AD23] G. Avramidi and Th. Delzant, Group rings and hyperbolic geometry, preprint, arXiv:2309.16791 [math.GT], 2023.
  • [Ale23] J. W. Alexander, A lemma on systems of knotted curves, Proc. Nat. Acad. Sci. U.S.A. 9 (1923), 93–95.
  • [Ale06] A. D. Alexandrov, Selected works. Part II: Intrinsic geometry of convex surfaces, Chapman & Hall/CRC, 2006, translated from the 1948 Russian original by S. Vakhrameyev.
  • [AOS21] G. Avramidi, B. Okun, and K. Schreve, Mod pp and torsion homology growth in nonpositive curvature, Invent. Math. 226 (2021), no. 3, 711–723.
  • [APS73] M. F. Atiyah, V. K. Patodi, and I. M. Singer, Spectral asymmetry and Riemannian geometry, Bull. London Math. Soc. 5 (1973), 229–234.
  • [APS75] by same author, Spectral asymmetry and Riemannian geometry. I, Math. Proc. Cambridge Philos. Soc. 77 (1975), 43–69.
  • [Ati69] M. F. Atiyah, The signature of fibre-bundles, Global Analysis (Papers in Honor of K. Kodaira), Univ. Tokyo Press, Tokyo, 1969, pp. 73–84.
  • [BL73] W. Browder and G. R. Livesay, Fixed point free involutions on homotopy spheres, Tohoku Math. J. (2) 25 (1973), 69–87.
  • [BRS76] S. Buoncristiano, C. P. Rourke, and B. J. Sanderson, A geometric approach to homology theory, London Mathematical Society Lecture Note Series, No. 18, Cambridge University Press, Cambridge-New York-Melbourne, 1976.
  • [CDMW18] G. R. Chambers, D. Dotterrer, F. Manin, and Sh. Weinberger, Quantitative null-cobordism, J. Amer. Math. Soc. 31 (2018), no. 4, 1165–1203, With an appendix by Manin and Weinberger.
  • [CG85] J. Cheeger and M. Gromov, On the characteristic numbers of complete manifolds of bounded curvature and finite volume, Differential geometry and complex analysis, Springer, Berlin, 1985, pp. 115–154.
  • [Cha16] J. C. Cha, A topological approach to Cheeger-Gromov universal bounds for von Neumann ρ\rho-invariants, Comm. Pure Appl. Math. 69 (2016), no. 6, 1154–1209.
  • [CL] J. C. Cha and G. Lim, Linearity of universal bounds for Cheeger–Gromov ρ\rho-invariants for (4k1)(4k-1)-manifolds, in preparation.
  • [CMS84] J. Cheeger, W. Müller, and R. Schrader, On the curvature of piecewise flat spaces, Comm. Math. Phys. 92 (1984), no. 3, 405–454.
  • [CT08] F. Costantino and D. Thurston, 3-manifolds efficiently bound 4-manifolds, J. Topol. 1 (2008), no. 3, 703–745.
  • [CW03] S. Chang and Sh. Weinberger, On invariants of Hirzebruch and Cheeger-Gromov, Geom. Topol. 7 (2003), 311–319.
  • [FMW23] M. Frączyk, S. Mellick, and A. Wilkens, Poisson–Voronoi tessellations and fixed price in higher rank, preprint, arXiv:2307.01194 [math.GT], 2023.
  • [Fri15] G. Friedman, Stratified and unstratified bordism of pseudomanifolds, Topology Appl. 194 (2015), 51–92.
  • [GG12] M. Gromov and L. Guth, Generalizations of the Kolmogorov–Barzdin embedding estimates, Duke Mathematical Journal 161 (2012), no. 13, 2549–2603.
  • [Gor84] M. Goresky, Witt space cobordism theory (after P. Siegel), Intersection cohomology (Bern, 1983), Progr. Math., vol. 50, Birkhäuser Boston, Boston, MA, 1984, pp. 209–214.
  • [Gro82] M. Gromov, Volume and bounded cohomology, Inst. Hautes Études Sci. Publ. Math. (1982), no. 56, 5–99.
  • [Gro96] by same author, Positive curvature, macroscopic dimension, spectral gaps and higher signatures, Functional analysis on the eve of the 21st century, Vol. II (New Brunswick, NJ, 1993), Progr. Math., vol. 132, Birkhäuser Boston, Boston, MA, 1996, pp. 1–213.
  • [Gro99] by same author, Quantitative homotopy theory, Prospects in mathematics (Princeton, NJ, 1996), Amer. Math. Soc., Providence, RI, 1999, pp. 45–49.
  • [Gro09] by same author, Singularities, expanders and topology of maps. I. Homology versus volume in the spaces of cycles, Geom. Funct. Anal. 19 (2009), no. 3, 743–841.
  • [Hir68] F. Hirzebruch, Involutionen auf Mannigfaltigkeiten, Proc. Conf. on Transformation Groups (New Orleans, La., 1967), Springer-Verlag New York, Inc., New York, 1968, pp. 148–166.
  • [Law78] T. Lawson, Open book decompositions for odd dimensional manifolds, Topology 17 (1978), no. 2, 189–192.
  • [LR78] N. Levitt and C. Rourke, The existence of combinatorial formulae for characteristic classes, Trans. Amer. Math. Soc. 239 (1978), 391–397.
  • [LS06] J.-F. Lafont and B. Schmidt, Simplicial volume of closed locally symmetric spaces of non-compact type, Acta Math. 197 (2006), no. 1, 129–143. MR 2285319
  • [Lus72] G. Lusztig, Novikov’s higher signature and families of elliptic operators, J. Differential Geometry 7 (1972), 229–256. MR 322889
  • [LW23] G. Lim and Sh. Weinberger, Bounds on ρ\rho-invariants and simplicial complexity of triangulated manifolds, arXiv:2301.08870 [math.GT], 2023.
  • [Mey72] W. Meyer, Die Signatur von lokalen Koeffizientensystemen und Faserbündeln, Bonn. Math. Schr. (1972), no. 53, viii+59.
  • [MM79] I. Madsen and R. J. Milgram, The classifying spaces for surgery and cobordism of manifolds, Annals of Mathematics Studies, vol. 92, Princeton University Press, Princeton, N.J.; University of Tokyo Press, Tokyo, 1979.
  • [Qui79] F. Quinn, Open book decompositions, and the bordism of automorphisms, Topology 18 (1979), no. 1, 55–73.
  • [Sie83] P. H. Siegel, Witt spaces: a geometric cycle theory for KOK{\rm O}-homology at odd primes, Amer. J. Math. 105 (1983), no. 5, 1067–1105.
  • [Sma62] S. Smale, On the structure of manifolds, Amer. J. Math. 84 (1962), 387–399.
  • [Tho53] R. Thom, Variétés différentiables cobordantes, C. R. Acad. Sci. Paris 236 (1953), 1733–1735.
  • [Tho54] by same author, Quelques propriétés globales des variétés différentiables, Comment. Math. Helv. 28 (1954), 17–86.
  • [Wal99] C. T. C. Wall, Surgery on compact manifolds, second ed., Mathematical Surveys and Monographs, vol. 69, American Mathematical Society, Providence, RI, 1999, Edited and with a foreword by A. A. Ranicki.
  • [Wei99] Sh. Weinberger, Higher ρ\rho-invariants, Tel Aviv Topology Conference: Rothenberg Festschrift (1998), Contemp. Math., vol. 231, Amer. Math. Soc., Providence, RI, 1999, pp. 315–320.
  • [Wei23] by same author, Variations on a theme of Borel: an essay on the role of the fundamental group in rigidity, Cambridge Tracts in Mathematics, vol. 213, Cambridge University Press, Cambridge, 2023.
  • [Win73] H. E. Winkelnkemper, Manifolds as open books, Bull. Amer. Math. Soc. 79 (1973), 45–51.