Quantitative PL bordism
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 simplices has a filling of bounded geometry whose number of simplices is slightly superlinear in . This bound is similar to that found in our previous work on smooth cobordism.
1. Introduction
Suppose a compact manifold is nullcobordant, i.e. is the boundary of some compact manifold . How complicated must be? This is not just one question, but a whole family of questions, depending on the type of manifold that we allow and to be, and how we measure the complexity of .
To our knowledge, Gromov is the first to ask such questions; he discusses several variations in [Gro96, §] and [Gro99]. In the case that and 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 and injectivity radius at least ). He suggested that the minimal complexity of a which fills should be linear in the complexity of ; a slightly superlinear bound ( for any ) was proved in [CDMW18].
In this paper we prove an analogous bound in the PL category. In this setting, “bounded geometry of type ” means that every vertex is incident to at most simplices, and volume is measured by the total number of top-dimensional simplices.
Theorem 1.1.
Let be a PL -manifold with bounded geometry of type and total simplices. Then if is a PL boundary, it bounds a manifold with bounded geometry of type (depending on and ) so that the number of simplices in is bounded by a function (again depending inexplicitly on and ) which is for every .
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 , and to prove the quantitative bound one must relate the Lipschitz constant of a nullhomotopic map to the Lipschitz constant of its most efficient nullhomotopy. Even talking about Lipschitz constants requires a metric on , 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 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 over a finite complex of lower dimension. The number of such bundles which admit a connection with bounded curvature grows as a polynomial in whose degree is related to the cohomology of . However, the naturally isomorphic set of stable vector bundles over 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 is the number of simplices in a triangulation. We could consider the problem: given a PL -manifold with simplices, how many simplices are contained in the simplest whose boundary is ?
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 . 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 that someone hands you is actually a PL manifold. Maybe your interlocutor is playing a joke on you? You can be sure that is a homology manifold, but it is much harder to know that 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 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 , 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 , there is an such that every PL -manifold is PL homeomorphic to one with bounded geometry of type . (However, given a manifold with simplices, the number of subdivisions required to implement a PL isomorphism with such a manifold grows faster than any computable function of .) 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 be a category of PL manifolds with prescribed singularities.
For every , , and , there are constants and such that every PL nullcobordant triangulated -dimensional -manifold with -simplices and bounded geometry of type has a filling with at most simplices and bounded geometry of type . This holds for both oriented and unoriented cobordism.
More generally, for a finite simplicial complex , there is a constant such that if is as above and is a simplicial map representing in the bordism homology group , then there is a simplicial extension to a filling of with at most simplices and bounded geometry of type .
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 be the number of -simplices of the manifold . We proceed as follows:
-
(1)
Embed in , with some control over the shape of a regular neighborhood. Here we can choose any sufficiently large .
-
(2)
This induces an -Lipschitz map from to a fixed compact subspace of a kind of Thom space for -manifolds.
-
(3)
One constructs a -Lipschitz extension of this map to , with image in a larger fixed compact subspace of this Thom space.
-
(4)
From a simplicial approximation of this nullhomotopy, one can extract an -manifold embedded in which fills and whose number of simplices is bounded by the number of simplices in the approximation and is therefore .
-
(5)
After the previous step, the bound on the local geometry of the resulting filling depends on . We tidy up by applying local modifications to , making the bound on the geometry independent of at the expense of increasing the number of simplices by an additional multiplicative constant.
The final estimate depends on bounds on the functions and . Our results from the appendix to [CDMW18] suffice to show that
and we will show that is at most linear. This gives an overall bound of on the size of the resulting filling; since 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 -manifold with a triangulation with -simplices has a filling which has bounded geometry and simplices.
Note that in this theorem, 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 -manifold with a triangulation with -simplices has a filling which has bounded geometry and -simplices, where for .
Proof.
It suffices to show that every -manifold with -simplices has a triangulation of bounded geometry with simplices. Such a triangulation can be produced as follows. This strategy was outlined by Gromov [Gro96, §II′].
First, take the unbounded triangulation and barycentrically subdivide once.
Now let be the set of vertices which come from the barycenters of edges. The star of each such vertex looks like (the suspension of the cone on ), where is a circle with some number of edges. It is easy to see that can be filled with a disk with triangles, at most of which meet at a vertex, and at most of which meet at a vertex on . So for each vertex in , we replace the subcomplex with , getting a PL homeomorphic complex .
Now let be the set of vertices of that correspond to the original vertices of . The link of such a vertex in is a triangulation of with faces and at most faces meeting at every vertex. Such a triangulation again has a filling with bounded geometry and -simplices. Gromov sketches the proof in [Gro96, §II′′]: after smoothing out the triangulation, one gets a Riemannian metric on with curvature bounded below by some . By a theorem of Alexandrov [Ale06, §XII.2], such an embeds isometrically into with curvature . In hyperbolic space, this sphere bounds a ball of volume . Retriangulating that ball gives us our bounded geometry filling.
Replacing the star of each vertex in with a bounded geometry filling of its link gives us our bounded geometry triangulation of . ∎
Notice that this strategy generalizes to higher dimensions: if every bounded geometry triangulated of volume can be filled with a bounded geometry disk of volume , then every -manifold with a triangulation of unbounded geometry and volume has a triangulation of bounded geometry and volume
However, we currently have no estimates on the functions for . Finding such estimates seems to be a worthwhile and difficult problem. In a future paper we will show that 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 be a finite complex. Then every simplicial -cycle in of bounded geometry of type and volume has a filling of bounded geometry of type and volume , where for every .
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 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 -invariants of Atiyah–Patodi–Singer [APS75, APS73, Wal99]—and their 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 -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 , when is the class of Witt spaces, is used to define higher -invariants of manifolds with fundamental group [Wei99]. If is of finite type, our results can then be used to bound these higher -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 is not of finite type, we get an important example of a situation in which our results do not apply.
Remark.
The -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 , and the theorem of Kervaire–Milnor showing that the number of differentiable structures on the -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 , and if the cobordism is assumed to have the same fundamental group, one can use the -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 -manifold efficiently into , where , implementing step (1) of the proof outline.
In [GG12], Gromov and Guth describe “thick” embeddings of -dimensional simplicial complexes in unit -balls, for . They define the thickness of an embedding to be the maximum value such that disjoint simplices are mapped to sets at least distance from each other. In [GG12, Thm. 2.1], given a complex with volume and bounded geometry of type , they construct embeddings with a lower bound on thickness whose asymptotic behavior as a function of 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 by .
Theorem 3.1 (based on [CDMW18, Appendix A, Theorem 2.1]).
Suppose that is a -dimensional simplicical complex with vertices and each vertex lying in at most simplices. Suppose that . Then there are constants and and a subdivision of which embeds linearly into the -dimensional Euclidean ball of radius
such that:
-
(i)
The embedding has Gromov–Guth thickness 1.
-
(ii)
For any -simplex of , the induced embedding has Gromov–Guth thickness .
-
(iii)
Adjacent vertices of are mapped units apart.
-
(iv)
The vertices of the embedding are “snapped to a grid”: they are points in a lattice in which depends only on .
Remarks.
-
(a)
When is a PL -manifold, an embedding is locally flat if it is locally PL homeomorphic to the standard embedding of in . By the Zeeman unknotting theorem, this is always true unless .
-
(b)
Conditions (i) and (iii) taken together imply that each simplex is -bilipschitz to a standard simplex for some .
-
(c)
Together with condition (iv), this means that the simplices in the embedding come from a finite number of isometry types depending only on , and .
Proof.
The cited theorem gives an embedding of satisfying conditions (i) and (ii). Moreover, the proof given in [CDMW18] produces a subdivision whose simplices are -bilipschitz to an equilateral simplex whose edges have length . A further subdivision gives simplices which are bilipschitz to the standard simplex at scale . 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 , and move all the vertices to the nearest point of a cubic lattice of side length , extending the map linearly to the rest of the complex. Then each vertex, and therefore each point of the complex, moves by at most , so the resulting complex still has Gromov–Guth thickness at least . Moreover, since vertices are at least apart, the angular movement of each point in a link is at most , and therefore the thickness of links in the resulting complex is still at least . Thus we can accommodate condition (iv) by at worst halving . ∎
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 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 such that for sufficiently larger than , -bordism classes of spaces in correspond to homotopy classes of maps . Note that we do not guarantee that this mapping is bijective, but merely injective: may contain classes that are not in the image of the bordism group . Therefore the spaces we construct are not explicit classifying spaces for -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 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 of sets of -dimensional simplicial complexes which constitute permissible links of vertices in a closed -dimensional -manifold. These sets must satisfy:
-
(1)
is closed under PL isomorphism.
-
(2)
Each member of is a closed -manifold (that is, its links are contained in ).
-
(3)
If , then .
-
(4)
If , then .
The last axiom means that there is a notion of -manifold with boundary and that the boundary of such an object is well-defined. Namely, an -dimensional -manifold with boundary is a simplicial complex whose links lie in or , and its boundary is the subcomplex of points whose links lie in . Axiom 3 ensures that if is an -manifold, then is a -manifold with boundary.
An additional “regularity” axiom, which is not strictly necessary, guarantees that -manifolds are pseudomanifolds:
-
(5)
.
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 -dimensional -manifold embedded simplexwise linearly in . To be precise, is a simplicial complex glued out of subcomplexes corresponding to triples , where:
-
•
is a permissible link in .
-
•
is a linear embedding of the cone on . Let be the cone point of .
-
•
is a neighborhood of which is equipped with a linear triangulation transverse to and such that .
-
•
For every -simplex , the subcomplex defined by
satisfies . In other words, there is at least one simplex of which intersects nontrivially and does not intersect for any simplex which does not contain . This condition can be thought of as ensuring that that has a nice simplicial projection to the cell complex dual to .
We identify two triples and if and differ from and by the same translation.
Given such a triple, let be the subcomplex consisting of simplices disjoint from . Then is glued out of complexes . Given two such subcomplexes and , if there are vertices and such that and , then the subcomplexes
are identified in . Moreover, all the cone points of all the are identified (this is the “point at infinity” in the Thom space).
Optionally, we can make this into a Thom space for oriented -manifolds by adding orientation data for top-dimensional simplices of to the triple , and require that this data match in order to identify subcomplexes.
Notice that occurs as a subcomplex in . This is because for any subcomplex of , contains a subcomplex corresponding to the triple
Then includes into as the subcomplex for . These inclusions respect the gluings that produce the complexes as a whole. If using oriented simplices, we require that face in the positive direction in .
Finally, we must show that these complexes are indeed Thom spaces in the sense we need:
Proposition 4.1.
When , there is an injective “Pontrjagin–Thom” map
where is the bordism group of -dimensional -manifolds. Moreover, is -connected.
Proof.
Given an -manifold , we can construct a PL embedding . We then construct a suitable triangulation of as follows.
Lemma 4.2.
Let be a finite simplicial complex linearly embedded in . There is a linear triangulation of which is transverse to and has the following property. For a vertex , define the subset
Then the sets cover .
Proof.
Let be the barycentric subdivision of . It suffices for to be such that if a simplex of intersects the star in of a vertex , then is contained in the open star of in . This is true for any sufficiently fine triangulation. ∎
Now suppose we have such a triangulation. Then for any simplex of ,
satisfies the following conditions:
-
•
is contained in , and in fact for each , is the maximal subcomplex of for which this holds.
-
•
(by the Knaster–Kuratowski–Mazurkiewicz lemma, since for any simplex , covers ).
Therefore, there is a well-defined map which sends each to the subcomplex , and simplices disjoint from to the point at infinity. This map obviously extends to . This shows that for every -dimensional -manifold, there is a corresponding element of .
Similarly, if is the boundary of a -dimensional -manifold , we can extend the embedding to an embedding and use this embedding to construct a nullhomotopy of . This shows that the correspondence is well-defined on bordism classes.
Conversely, given a simplicial map , we can recover a -dimensional -manifold. Notice that contains a “zero section” , consisting of the image of in each . We claim that is an -manifold. Indeed, let be a point in , where is an -simplex of . Then it is easy to see that is the -fold suspension of the link of in .
Similarly, for a simplicial map , the preimage of the zero section is a -dimensional -manifold with boundary.
We have now constructed well-defined maps
whose composition is the identity. Therefore the first of these maps is injective.
Now notice that for every subcomplex of , the -skeleton of does not intersect the image of , that is, it is contained in . Therefore, the -skeleton of deformation retracts to the point at infinity, and so is -connected. ∎
4.3. Modifications for bordism homology
Here we describe the modifications to the construction of required to encode bordism classes of maps from -manifolds to a simplicial complex . We define the space as follows. We use the same subcomplexes , but we index them by an additional datum, a simplicial map . Given two vertices and such that and , we identify and if in addition under the obvious identification. Once again, all the cone points of all the are identified.
Note that comes with a natural projection map
which is finite-to-one if is a finite complex.
Proposition 4.3.
When , there is an injective map
where is the th -bordism homology group of . Moreover, is -connected.
Proof.
Given a triangulated -manifold and a simplicial map , we construct a map as in Proposition 4.1, taking care that the embedding is linear on a subdivision of our triangulation. To lift this map to , we need consistent choices of for each vertex of . We obtain this by homotoping to a map which is simplicial on . We use a similar method to convert a filling of to a nullhomotopy. This gives a well-defined map
It remains to show that this map is injective. We will do this once again by constructing a retraction.
Given a simplicial map , we can construct a manifold
as in Proposition 4.1 after forgetting the data about . Now we need to build a map from this manifold to which, when is the Thom map of some , agrees up to homotopy with .
We do this by composing the induced map with a map from the zero section to . This will be a simplicial map from a triangulation of the zero section to the one-time barycentric subdivision , and it is constructed by induction on dimension. For a simplex of , denote by the dual subcomplex to in . For each we build as follows:
-
•
For each -simplex of , set (note that is a vertex of ).
-
•
For each -simplex of , extend to so that its composition with the projection to is homotopic rel endpoints to . So that is well-defined, this map should depend only on and .
-
•
Continue this process for lower-dimensional simplices , at all points fixing maps in such a way that they depend only on and .
If we start with , build the corresponding map , and then use this to construct as above, then by construction . Therefore, as in Proposition 4.1, we have a retraction
and therefore the first map is injective.
Finally is -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 , and then find a controlled nullhomotopy of this map in a larger finite subcomplex of .
Lemma 5.1.
For every , , , and , there is a finite, -connected subcomplex such that the following holds. Let be a PL triangulated -dimensional -manifold with -simplices and bounded geometry of type , and let . Then there is a PL map such that is PL homeomorphic to and the Lipschitz constant of (with respect to the standard simplexwise linear metric on ) is at most
Moreover, if is a finite simplicial complex and is a simplicial map, then lifts to a PL map
such that , where is the projection defined in the proof of Proposition 4.3 and is the projection defined at the beginning of §4.3.
Proof.
Let be the lattice in specified in condition (iv) of Theorem 3.1. Let be a -invariant triangulation of with the following properties:
-
•
The edge lengths are at most .
-
•
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 .)
Now let be the union of subcomplexes such that:
-
•
is an admissible link in .
-
•
is an embedding of satisfying conditions (i)–(iv) of Theorem 3.1.
-
•
is the subcomplex of consisting of simplices which intersect and are at distance at least from .
The number of such combinations is finite, and therefore is a finite complex. Moreover, it’s -connected since all the are -connected and glued together along -connected subcomplexes.
Now by Theorem 3.1, there is a subdivision of which embeds linearly into the -dimensional Euclidean ball of radius
such that the embedding satisfies conditions (i)–(iv). Conditions (i) and (ii) imply that the triangulation satisfies the conditions of Lemma 4.2. This embedding induces a map from the -ball to which is simplicial on a slight subdivision of (depending only on ), and hence -Lipschitz. Since the boundary of the -ball is mapped to , this map extends to a sphere. If we rescale this to be the unit sphere, the Lipschitz constant becomes .
Finally, in the case of a map , we can homotope this map to a map which is simplicial on , and then decide the lift to based on the behavior of near each vertex. ∎
Note that for any finite simplicial complex , since is finite-to-one, is also a finite complex. We now embed in a larger, but still finite subcomplex of 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 equipped with a simplicial map
which extends the inclusion , such that:
-
(i)
is -connected.
-
(ii)
If a map is nullhomotopic in , then it is nullhomotopic in .
Proof.
We know that and are both -connected. We will build by adding cells of dimension and therefore not change this.
Since is a simply connected finite complex, its higher homotopy groups are finitely generated. In particular, the kernel of the induced map
is finitely generated. Choose simplicial maps representing the generators of the kernel, as well as simplicial fillings
(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 by gluing these simplicial disks onto using the maps as attaching maps. Then the map is constructed by extending to the new cells using the maps . ∎
Now put the standard simplexwise linear metric on ; this also restricts to a metric on . Then [CDMW18, Corollary 4.3] yields the following fact:
Lemma 5.3.
Suppose that , and let be an -Lipschitz map such that is nullhomotopic. Then is nullhomotopic in via a -Lipschitz nullhomotopy, where is a constant which depends on , , , , and .
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 be a category of manifolds with prescribed singularities, and integers, , and a finite simplicial complex. Then there are constants and such that the following holds.
Let be a triangulated -dimensional -manifold with -simplices and bounded geometry of type , and let be a simplicial map such that represents the zero class in . Then there is a -dimensional -manifold with with at most simplices and bounded geometry of type , and a simplicial map extending .
Proof.
Let . The Thom construction from Proposition 4.1 tells us that the simplicial map induced by any PL embedding of is nullhomotopic in . Combining Lemmas 5.1 and 5.3, we see that there is a map such that and .
By a quantitative simplicial approximation theorem, see e.g. [CDMW18, Prop. 2.1], we can simplicially approximate on a subdivided triangulation of at scale . This gives us a simplicial nullhomotopy of a simplicial approximation of .
We would like to get back to a nullhomotopy of . Notice that where is a simplicial map from the subdivision to the original triangulation. Moreover, can be constructed so that for each simplex of the original triangulation, is a contractible subcomplex. So we fix a triangulation of as follows. First triangulate using the usual product triangulation, where top-dimensional simplices take the form
for each and each -simplex of the subdivided triangulation of . Now take the mapping cylinder of , which is a simplicial quotient of this triangulation. Our extra assumption about implies that the mapping cylinder is still a triangulation of , with the subset given the subdivided triangulation and given the original triangulation. Then the map on the product triangulation induces a simplicial homotopy between and on the mapping cylinder. So we get a simplicial nullhomotopy of by attaching this to our disk as a collar.
Now Lemma 5.2 gives us a simplicial map
Then by the argument of Proposition 4.1, is an -manifold whose boundary is . Since the image of is a finite subcomplex of , the link types of come from a finite set depending on , , , , and . In particular, there is some which bounds the geometry of these link types.
Moreover, there is a triangulation of with a bounded number of simplices per simplex of : we obtain it by pulling back from the image under the triangulation on which the map
is defined; incidentally this gives us a simplicial map . With this triangulation, consists of simplices. Given any , for sufficiently large , this is .
At this point in the proof, depends on and as well as , , and the category. We will fix this at the expense of increasing the constant , by an inductive process which replaces certain subcomplexes of with other subcomplexes. The strategy is similar to the proof of Corollary 2.1.
We start by taking a single barycentric subdivision of . Let be the set of vertices in the interior (outside ) of the resulting triangulation that originate as barycenters of -simplices. Notice that the star of a vertex is of the form where is the -dimensional link of the original -simplex. We will modify by replacing with other fillings of .
We start with . If -manifolds are pseudomanifolds, then for , is a disjoint union of circles. This can be filled with disks with faces to a vertex, at most faces to a vertex on the boundary, and linearly many faces in terms of ; let be this collection of disks. For each , we replace with to get a new -manifold . Notice that although we have eliminated the vertices in , the sets of vertices for are unchanged from , and their stars are still of the form , though the -complex is different from . Since consisted of s for , meeting three to a vertex, the parts of outside have at most faces meeting at a vertex.
Suppose now, by induction, that we have built a complex in which we have replaced the stars of vertices with subcomplexes of the form , where is a subcomplex of bounded geometry of type some . Then for each , its star is a complex where the geometry of the -complex is bounded by . Moreover, by induction and since we had an original bound on the complexity of link types, the volume of is bounded as a function of , , , , and . We can then replace with a complex (a fixed -manifold whose boundary is ) such that:
-
•
The geometry of is bounded by .
-
•
The volume of is bounded as a function of , , , , and (simply because the set of possible s is finite and depends only on these parameters).
This completes the inductive step to construct .
At each stage, we extend the map over by mapping each interior vertex to and extending by linearity. Since we are replacing the star of , this is a well-defined map.
At the end of the induction, we have replaced with an -manifold whose boundary is and whose geometry still depends on , , and , but not on . Moreover, we have multiplied the volume by at most a constant depending on , , , and . ∎
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 be a Riemannian -manifold with sectional curvatures and injectivity radius at least , and volume . Then there is a triangulation of with at most simplices and at most simplices incident to each vertex.
Theorem A.2.
Let be a smoothable PL -manifold equipped with a PL triangulation with at most simplices incident to any vertex. Then for constants depending on and , every smoothing of has an associated Riemannian metric such that:
-
(1)
Every simplex of has volume at most .
-
(2)
has sectional curvature and injectivity radius at least .
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 -net of points on , for some possibly depending on : by definition, the -balls around the are disjoint and the -balls cover . 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 intersects at most some of the -balls around the other . Thus there is a -coloring of the into sets so that no two -balls of the same color intersect.
We can triangulate each 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 . 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 and for and ; then to add in the -balls around the points in ; and so on. We thus worsen the geometry of the triangulation a total of times. In the end, we get a triangulation of all of for which the geometry of the simplices still depends only on . ∎
Theorem A.2 follows easily from the fact that the groups of exotic spheres are all finite (except perhaps in dimension , where any exotic spheres are not PL spheres).
Proof of Theorem A.2..
Let be a triangulated -manifold, and suppose that has at most simplices adjacent to every vertex. In particular, the star of each -simplex can have a finite number of possible configurations.
We will show that there is a finite set of Riemannian metrics on the -simplex, depending only on and , such that every smoothing of has a metric which, when restricted to each simplex, is isometric to one of the metrics in . This is sufficient to prove the theorem because the constants , , and are then obtained by maximizing over the finite set of possible local configurations.
We construct by induction, giving a process that builds all possible smoothings of as Riemannian manifolds with metrics on simplices chosen from . First let be the two times barycentric subdivision of , and let
be open submanifolds of such that each includes the star of the -skeleton of in . Then is a disjoint union of balls around the vertices of , and going from to means gluing in copies of for each -simplex of .
Clearly has a unique smoothing. For every possible link of a vertex consisting of at most simplices, we fix a Riemannian metric on its cone; this gives a Riemannian metric on .
Now suppose we have a smoothing of equipped with a Riemannian metric , in which each -simplex of is isometric to one of a finite list depending on and . For every -simplex of , if the smoothing extends to , then the set of such extensions is in bijection with the finite group of exotic -spheres. We fix Riemannian metrics on all these extensions, depending on the combinatorial structure of , the metric on , and the element of . By induction, this data takes a finite set of values depending only on and .
After the th step, for every smoothing of , 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 , which we denote by . 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 of dimension , Smale [Sma62, Theorem 6.1] showed that the homology of determines in the most obvious way: the number of -dimensional critical points is
where is the rank of the th homology and 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 ; 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 , and is a simple homotopy equivalence, then .
Proof.
As in the proof of the -cobordism theorem, we can assume that is given a self-indexing Morse function (i.e. a Morse function so that critical points of index arise before ones of index ). We also assume that there is just one critical point of index and of index .
We construct a Morse function on as follows. In index we mimic the -complex associated to the Morse function of . That is, we put 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 , creating a Morse function on a disjoint submanifold.
Now what remains is a simple homotopy equivalence (which we still call ) between two manifolds with boundary, say and , where the manifolds and their boundaries all have fundamental group and is given a Morse function (locally constant on the boundary) whose critical points all have index between and . 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 whose image is the lowest handle of , and generate a new simple homotopy equivalence by cutting them both out. Write for this first handle, and . Now we apply [Wal99, Theorem 12.1]. By this theorem, can be homotoped so that it is transverse to , and so that it induces a simple homotopy equivalence of triads . Then is a manifold homotopy equivalent to ; but by simply connected surgery on triads, any such homotopy equivalence is in fact a diffeomorphism, so is a handle.
Once one has pulled back all the handles of to , what lies between and this union is an -cobordism, and therefore a product, so we can easily extend the Morse function to all of . ∎
Remark.
If one does not first make the Morse function on self-indexing, it is possible to give examples where one cannot find a “corresponding” function on .
In the non–simply connected case, actually understanding this number is not so easy. If is a -fold cover, then obviously . One can also use Betti numbers of finite covers to give lower bounds on . These are typically hard to get information about except in characteristic , where one usually gets this information from 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 is a locally symmetric space associated to a group without discrete series (equivalently has Euler characteristic ), is it true that has a sequence of finite covers of index such that ?
The optimistic conjecture that the answer is yes very generally whenever 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 has rank , 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 of degree , not necessarily a covering map, must have simplices.
Similarly, one can ask:
Question (Gromov, cf. [Gro96, §8]).
When does a nice manifold , e.g. negatively curved or locally symmetric, have a sequence of degree manifolds mapping to it, with ?
The following proposition shows that the answer is always for odd-dimensional manifolds.
Proposition B.2.
If is odd, then always has a sequence of -fold finite branched covers (for any index ) 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 , 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 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, §8] and [Gro09, §3.2]).
One can then ask analogous questions about integral homology or even bordism. Let be a finite complex. We consider the bordism group of smooth oriented -dimensional manifolds mapping to 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 , all of has representatives of uniformly bounded Morse complexity.
Proof.
We will show that if represents an element of bordism, then a -fold branched cover of represents . The result then follows as above from the finite generation of .
To see this fact, first suppose that the binding is empty, in other words fibers over and the branched cover is just a -fold cover . Let be a cobordism over between copies of the identity map and a connected -fold cover. Then the pullback of the bundle along , equipped with the map , is a bordism between and .
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 and gives us a representative of which fibers over the circle. ∎
For even the answer depends only on the fundamental group of 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 is determined up to a uniformly bounded error by its “higher signature class image” in , determined by the numbers where is the Hirzebruch -class and runs over the nontrivial group cohomology classes of .
For example, if is simply connected, the result is that , where is the signature of any representative manifold for the class . That no other classes obstruct is a consequence of Winkelnkemper’s theorem [Win73] (reproved by Quinn [Qui79]) that any simply connected manifold of dimension with signature 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 are represented by elements of the form where 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 . When , 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 is a local system of (perhaps skew-)symmetric or Hermitian quadratic forms on , then for a -manifold equipped with a map one gets an inner product pairing on whose signature is an invariant of bordisms in . This signature obviously gives a lower bound on .
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 be a product of surfaces of genus at least . Then the signature associated to the class 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 is a single surface, and let be a map. Let be another surface of genus at least . According to Atiyah [Ati69, §4], bundles with fiber satisfy the equation
where T is the vector bundle induced by the monodromy of .
Let be an -bundle such that has nonzero signature, as constructed by Atiyah. Then the above identity implies that
Since gives a lower bound on , 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 is the product of one-dimensional classes from different surfaces, then it is realized by a map of a torus to . 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 to detect all of the rational cohomology of . 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, §8], 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 that are cobordant over . (Since is finite for odd , 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 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 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 be a -manifold whose fundamental group is the product of three genus surface groups. The product of these surfaces with the Milnor manifold constructed from copies of , mapping to , gives an element of the group . Using Wall’s realization theorem for this element of , one obtains a cobordism from to , 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 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 -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 -invariants for -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 -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 -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 -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.