Higher connectivity of the Morse complex

By Nicholas A. Scoville and Matthew C. B. Zaremsky

Abstract

The Morse complex of a finite simplicial complex is the complex of all gradient vector fields on . In this paper we study higher connectivity properties of . For example, we prove that gets arbitrarily highly connected as the maximum degree of a vertex of goes to , and for a graph additionally as the number of edges goes to . We also classify precisely when is connected or simply connected. Our main tool is Bestvina–Brady Morse theory, applied to a “generalized Morse complex.”

Introduction

The Morse complex of a finite simplicial complex is the simplicial complex of all gradient vector fields on . See Section 1 for a more detailed definition. The Morse complex has several important properties. For example, two connected simplicial complexes are isomorphic if and only if their Morse complexes are isomorphic Reference CM17. Additionally, outside a few sporadic cases, for connected the group of automorphisms of is isomorphic to that of Reference LS21. The Morse complex may be viewed as a discrete analog of the space of gradient vector fields on a manifold; see, e.g., Reference PdM82.

The homotopy type of is only known for a handful of examples of , and in general it is difficult to compute. In this paper, we relax the question to just asking how highly connected is (meaning up to what bound the homotopy groups vanish). Our first main result is the following:

Theorem 2.7.

If has a vertex with degree in then is -connected.

For example this holds if . It is harder to obtain good higher connectivity bounds when the dimension of is small and vertices of have small degrees, but for certain situations we can. First we focus on the case when , i.e., is a graph . Here we are able to use Bestvina–Brady Morse theory, applied to the so called generalized Morse complex , to find higher connectivity bounds for . Let be the maximum degree of a vertex in the Hasse diagram. Our main result for graphs is:

Theorem 4.3.

The Morse complex is -connected.

Combining Theorem 4.3 with Theorem 2.7 quickly shows that, as the number of edges of goes to , becomes arbitrarily highly connected (see Corollary 4.4 for a precise statement). We conjecture that a similar result holds regardless of (Conjecture 2.8).

Our last main result is a classification of precisely which have connected and simply connected Morse complexes. Here we assume has no isolated vertices just to make the statement cleaner (isolated vertices can be deleted without affecting ).

Theorem 5.4.

Suppose has no isolated vertices. The Morse complex is connected if and only if is not an edge, and is simply connected if and only if is none of: an edge, a disjoint union of two edges, a path with three edges, a -cycle, or a -simplex.

This paper is organized as follows. In Section 1 we set up the Morse complex and generalized Morse complex . In Section 2 we prove Theorem 2.7. In Section 3 we discuss Bestvina–Brady discrete Morse theory and how to apply it to . In Section 4 we focus on the situation for graphs and prove Theorem 4.3. Finally, in Section 5 we discuss the situation for and prove Theorem 5.4.

1. The Morse complex

Let be a finite abstract simplicial complex. We will abuse notation and also write for the geometric realization of . If is a -dimensional simplex in , we may write to indicate the dimension. A primitive discrete vector field on is a pair for . A discrete vector field on is a collection of primitive discrete vector fields

such that each simplex of is in at most one pair . If the two simplices in are distinct from the two simplices in , call the primitive discrete vector fields and compatible; in particular a discrete vector field is a set of pairwise compatible primitive discrete vector fields.

The Hasse diagram of is the simple graph with a vertex for each (non-empty) simplex of and an edge between any pair of simplices such that one is a codimension- face of the other. In particular the primitive discrete vector fields on are in one-to-one correspondence with the edges of . Also, the discrete vector fields on are in one-to-one correspondence with the matchings, i.e., the collections of pairwise disjoint edges, on . We will sometimes equivocate between a discrete vector field on and its corresponding matching on .

Definition 1.1 (Generalized Morse complex).

The generalized Morse complex of is the simplicial complex whose vertices are the primitive discrete vector fields on , with a finite collection of vertices spanning a simplex whenever the primitive discrete vector fields are pairwise compatible. Said another way, the simplices of are the discrete vector fields on , with face relation given by inclusion.

Note that is a flag complex, i.e., if a finite collection of vertices pairwise span edges then they span a simplex, which makes it comparatively easy to analyze. Viewed in terms of matchings on , is precisely the matching complex of , i.e., the simplicial complex of matchings with face relation given by inclusion. Matching complexes of graphs are well studied; see Reference BGM20 for an especially extensive list of references.

The Morse complex of , introduced by Chari and Joswig in Reference CJ05, is the subcomplex of consisting of all discrete vector fields arising from a Forman discrete Morse function, or equivalently all acyclic discrete vector fields. To define all this, we need some setup, which we draw mostly from Reference Sco19, Section 2.2. Given a discrete vector field on , a -path is a sequence of simplices

such that for each , and . Such a -path is non-trivial if , and closed if . A closed non-trivial -path is called a -cycle. If there exist no -cycles, call acyclic. A -cycle is simple if , …, are pairwise distinct and , …, are pairwise distinct. We will identify -cycles up to cyclic permutation, e.g., we consider , , , …, , to be the same cycle as , , , …, , , , , and so forth.

Every acyclic discrete vector field on is the gradient vector field of a Forman discrete Morse function on . A Forman discrete Morse function on (developed by Forman in Reference For98) is a function such that for every , there is at most one with , and for every there is at most one with . The gradient vector field of is the discrete vector field whose primitive vector fields are all the with . A discrete vector field is the gradient vector field of some Forman discrete Morse function if and only if it is acyclic Reference Sco19, Theorem 2.51.

Definition 1.2 (Morse complex).

The subcomplex of consisting of all acyclic is the Morse complex of .

Note that any subset of an acyclic discrete vector field is itself acyclic, so really is a subcomplex. We should remark that the term “Morse complex” also means a certain algebraic chain complex obtained from an acyclic matching, e.g., see Reference Koz08, Definition 11.23, but in this paper “Morse complex” will always mean .

Observation 1.3 will be important later when relating and .

Observation 1.3 (1-skeleton).

The -skeleton of coincides with that of .

Proof.

Since is simplicial, fewer than three compatible primitive discrete vector fields cannot form a cycle.

Let us discuss two examples that are instructive and will be specifically relevant later.

Example 1.4.

Let be the -cycle, i.e, the cyclic graph with vertices. See Figure 1 for drawings of , , and . We see that and (this computation of agrees with Kozlov’s computation in Reference Koz99, Proposition 5.2). In particular neither nor is simply connected.

Example 1.5.

Let be the -simplex. See Figure 2 for drawings of , , and . We see that and (this computation of agrees with Chari and Joswig’s Reference CJ05, Proposition 5.1). In particular neither nor is simply connected.

It will become necessary later to consider the following generalization of and , in which certain simplices are “illegal” and cannot be used. Specifically, this will be needed in the proof of Proposition 4.2 to get an inductive argument to work.

Definition 1.6 (Relative (generalized) Morse complex).

Let be a subset of the set of simplices of . The relative generalized Morse complex is the full subcomplex of spanned by those vertices, i.e., primitive discrete vector fields , such that . The relative Morse complex is the subcomplex .

We can also phrase things using .

Definition 1.7 (Relative Hasse diagram).

The relative Hasse diagram is the induced subgraph of with vertex set given by all simplices of not in .

If we view as the matching complex of , then clearly is the matching complex of .

2. First results on higher connectivity

In this section we establish some higher connectivity bounds for the various complexes in question. In subsequent sections we will use Bestvina–Brady Morse theory to obtain more sophisticated higher connectivity bounds in certain cases. First we focus on the relative generalized Morse complex . We will use the “Belk–Forrest groundedness trick,” introduced by Belk and Forrest in Reference BF19.

Definition 2.1 (Ground, grounded).

Call a simplex in a simplicial complex an -ground if every vertex of the complex is adjacent to all but at most vertices of the simplex. The complex is -grounded if it admits a -simplex that is an -ground.

Theorem 2.2 (Groundedness trick).

Reference BF19, Theorem 4.9 Every -grounded flag complex is -connected.

Note that in Reference BF19, Theorem 4.9 the complex is assumed to be finite and are assumed to be at least , but this is not necessary: see, e.g., Reference SWZ19, Remark 4.12. Also note that in these references the bound is written , but this equals , and this form will be notationally convenient for us later.

In it is clear that every -simplex is a -ground. This is because any primitive discrete vector field only “uses” two simplices of , and so can fail to be compatible with at most two vertices of a given simplex. In particular this shows:

Observation 2.3.

If contains a -simplex then it is -connected.

Note that this only works because is a flag complex, and in particular Theorem 2.2 does not apply to .

This next result will be useful later when using Bestvina–Brady Morse theory and inductive arguments. Let be the number of edges in , and let be the maximum degree of a vertex in .

Proposition 2.4.

The complex is -connected.

Proof.

We first claim that contains a simplex of dimension . A -simplex in consists of pairwise disjoint edges in , so we need to show that admits pairwise disjoint edges. Since is a simple bipartite graph, by Kőnig’s Theorem it suffices to show that every vertex cover of has at least vertices. (Here a vertex cover is a subset of the vertex set such that every edge is incident to at least one element of .) Indeed for any graph , if is a vertex cover of then

and we have and , so this follows.

Now set , so we have shown that contains a simplex of dimension . Then is -grounded, and is flag, so Theorem 2.2 implies that is -connected, hence -connected.

Note that in Proposition 2.4 we obtain a higher connectivity bound that is better when the maximal degree is small. We can also find a better, higher connectivity bound when the maximal degree is large. We will only need this in the case (since we will not need to use Morse theory or induction later), so for simplicity we will only phrase it in that case, but one could state an analog when .

Lemma 2.5.

Suppose has a vertex that has degree in . Then is -grounded, and hence -connected.

Proof.

Let be a vertex of degree in . Let be the -simplex with each an edge incident to and each the endpoint of not equal to . We claim that is a -ground. Indeed, for we have that is not incident to , so if is an arbitrary vertex in then can intersect for at most one . This shows that is -grounded, and since it is flag Theorem 2.2 says it is -connected.

Of course the actual goal of this paper is to find higher connectivity results for . Even though is not flag, and so the groundedness trick does not apply, we can still prove the analog of Lemma 2.5 for using a more complicated argument. First let us record an easy lemma that will be important in many arguments that follow.

Lemma 2.6.

Let be a primitive discrete vector field in a discrete vector field on . Then lies in at most one simple -cycle.

Proof.

Let be the endpoint of not equal to . Assume that lies in a simple -cycle. Then must be matched in to some edge . Since cannot be matched in to more than one edge, is the unique edge with . Hence every -cycle containing also contains . Repeating this argument, we see that if lies in a simple -cycle this simple -cycle is unique.

Note that the analog of Lemma 2.6 is not true for with , since then can have more than two codimension- faces.

Recall that the star of a simplex in a simplicial complex is the subcomplex of all simplices containing along with their faces. Let us say two simplices of are joinable (in ) if they lie in a common simplex in , or equivalently if they lie in each other’s stars.

Theorem 2.7.

If has a vertex with degree in then is -connected.

Proof.

As in the proof of Lemma 2.5, let be a vertex of degree in . Let be the -simplex with each an edge incident to and each the endpoint of not equal to . Note that is acyclic, hence a simplex in , since every has as its endpoint different than . We first claim that the union of stars

is contractible. Since stars are contractible, and these stars all intersect, e.g., they all contain , it suffices by the Nerve Lemma Reference BLVŽ94, Lemma 1.2 to show that the intersection of the stars of any subcollection of the is contractible. We claim that for any face of , we have

This will prove the claim since is contractible. The reverse inclusion holds trivially, so we need to prove the forward inclusion. Let be a simplex in that is joinable in to for each vertex in . Since is flag, this implies and span a simplex in , and we need to show that is acyclic. Any simple cycle in can contain at most one , since every has as its endpoint, . Since is joinable in to each vertex of , no such cycles can exist. We conclude that lies in , so is contractible.

Now we claim that this union contains the -skeleton of . Let be a -simplex in . Since is a -grounded flag complex, with -ground by the proof of Lemma 2.5, every vertex of is adjacent (in ) to all but at most one vertex of . Since has vertices and has vertices, this implies there exists a vertex of such that every vertex of is compatible with . Since is flag, this shows lies in . As a first case, suppose lies in more than one such star, say and . If does not lie in then the discrete vector field has a cycle. Since has no cycles, this means there is a cycle in containing . This cycle necessarily contains a primitive discrete vector field of the form for some . Since can lie in at most one simple cycle in the discrete vector field , by Lemma 2.6, and since and cannot lie in a common cycle, we conclude that lies in no cycles in . In particular lies in no cycles in , so is in , which finishes this case.

For the other case, suppose only lies in one , say without loss of generality in . Then for every , has a vertex that is a primitive discrete vector field incompatible with , i.e., contains either or but not both. Since no is incident to any for , the function must be injective, so , …, are precisely the vertices of . If for any , then cannot contain a cycle. Hence suppose, without loss of generality, that . Then for each , either contains or is of the form for some -simplex . In particular no contains . Hence no cycle in can contain , which implies no cycle in can contain , which implies there are no cycles. We conclude is in , which finishes this case.

We have shown that lies in a contractible subcomplex of . Hence the inclusion induces the trivial map in all homotopy groups. We also know this map induces a surjection in all for , so is -connected.

It seems much more difficult to prove the analog of Proposition 2.4 for , but we conjecture that it holds. Let us record this here (with for simplicity):

Conjecture 2.8.

The Morse complex is -connected. In particular if then is -connected.

In the following sections we will use Bestvina–Brady Morse theory to prove this conjecture in the case when , i.e., for graphs, and for the special cases regardless of .

3. Bestvina–Brady Morse theory

An important tool we will use now is Bestvina–Brady discrete Morse theory. This is related to Forman’s discrete Morse theory, and in fact can be viewed as a generalization of it, as explained in Reference Zar. For our purposes the definition of a Bestvina–Brady discrete Morse function is as follows. (This is a special case of the situation considered in Reference Zar.)

Definition 3.1.

Let be a simplicial complex and two functions such that for any adjacent vertices we have . Extend and to maps by extending affinely to each simplex. Then we call

a Bestvina–Brady discrete Morse function provided the following holds: for any infinite sequence , , …of vertices such that for each , is adjacent to and lexicographically, the set has no lower bound in .

Note that if is finite, as it will be in our forthcoming applications, then this condition about infinite sequences holds vacuously, but for now we will continue working in full generality.

Definition 3.1 is a bit unwieldy, but we will only need the following special case:

Example 3.2.

Let be the barycentric subdivision of a simplicial complex , so the vertices of are the simplices of and adjacency in is determined by incidence in . Let be any function. Let be the function sending (viewed as a vertex of ) to (viewed as a simplex of ). If is finite dimensional and is closed and discrete (for example this holds if is finite), then is a Bestvina–Brady discrete Morse function. Indeed, adjacent vertices of have different values (hence different values), and the finite dimensionality of plus the fact that is closed and discrete ensures that the last condition of Definition 3.1 is satisfied.

Given a Bestvina–Brady discrete Morse function , we can deduce topological properties of the sublevel complexes by analyzing topological properties of the descending links of vertices. Here the sublevel complex for is the full subcomplex of spanned by vertices with . The descending link of a vertex is the space of directions out of in which decreases in the lexicographic order. More rigorously, since and are affine on simplices and not simultaneously constant on edges, the lexicographic pair achieves its maximum value on a given simplex at a unique vertex of the simplex, called its top. The descending star is the subcomplex of consisting of all simplices with top , and their faces. Then is the link of in .

The claim that an understanding of descending links leads to an understanding of sublevel complexes is made rigorous by the following Morse Lemma. This is essentially Reference BB97, Corollary 2.6, and is more precisely spelled out in this form in, e.g., Reference Zar, Corollary 1.11.

Lemma 3.3 (Morse Lemma).

Let be a Bestvina–Brady discrete Morse function on a simplicial complex . Let in . If is -connected for all vertices with then the inclusion induces an isomorphism in for all , and an epimorphism in .

Let us return to the special case from Example 3.2, so for finite dimensional, and is closed and discrete on vertices. Given a vertex in (i.e., a simplex in ), there are two types of vertex in : we can either have a face with , or a coface with . This is because goes up when passing to faces and down when passing to cofaces. Since every face of is a face of every coface of , the descending link decomposes as join

where , the descending face link, is spanned by all with , and , the descending coface link, is spanned by all with . For example if at least one of or is contractible, so is . More generally, an understanding of the topology of and yields an understanding of the topology of .

3.1. Applying Bestvina–Brady Morse theory to the relative generalized Morse complex

Now we will apply Bestvina–Brady Morse theory to . The broad strokes of this strategy are inspired by the Morse theoretic approach in Reference BFM16, Proposition 3.6 to higher connectivity properties of the matching complex of a complete graph. Let , and let be the function sending to the number of simple -cycles (since is finite, any only has finitely many simple -cycles). In particular , so if we can understand for all with , using the Bestvina–Brady discrete Morse function as in Example 3.2, then the Morse Lemma will tell us information about .

Let us inspect the descending face link.

Lemma 3.4 (Descending face link, case 1).

Let with , so is a discrete vector field on (avoiding ) with at least one -cycle. If there exists a primitive discrete vector field in that is not contained in any -cycle, then is contractible.

Proof.

Say , and say without loss of generality that is not contained in any -cycle. Now let be any vertex of , so is a simplex of with and . Then , so . Since , Reference Qui78, Section 1.5 says is contractible.

Lemma 3.5 (Descending face link, case 2).

Let with , say is a -simplex of . If every primitive discrete vector field in is contained in a -cycle, then is homeomorphic to .

Proof.

The hypothesis ensures that for every proper face , i.e., removing any part of eliminates at least one -cycle (note that removing part of cannot create new cycles, so these are in fact equivalent). Hence is homeomorphic to the boundary of (viewed as a simplex in ), so homeomorphic to .

At this point we know that the descending link of a -simplex with is either contractible or else is the join of with (so the -fold suspension of ). It remains to analyze . In Section 4 we will discuss the case when , where it turns out we can fully analyze . Then in Section 5 we will consider arbitrary , where at least we will be able to tell when is non-empty.

4. Graphs

In the special case when , i.e., is a graph, the descending coface link of those satisfying the hypotheses of Lemma 3.5 can be related to a “smaller” Morse complex (see Proposition 4.1), which allows for inductive arguments. Throughout this section denotes a finite graph, and is a subset of the set of simplices of . To us “graph” will always mean a -dimensional simplicial complex, often called a “simple graph”.

Proposition 4.1 (Modeling the descending coface link).

Let be a -simplex in such that every primitive discrete vector field in lies in a -cycle. Let be the set of simplices of used by . Then in is isomorphic to .

Proof.

Define a simplicial map as follows. A vertex of is a discrete vector field on (avoiding ) of the form for non-trivial such that any -cycle is a -cycle. In particular is acyclic, and so . Setting gives a well defined map on the level of vertices. If then , so this extends to a simplicial map . Now we have to show is bijective. It is clearly injective, since implies . It is also clear that as long as is surjective on vertices, it will be surjective. To see it is surjective on vertices, let be a vertex in , and we have to show that any -cycle is a -cycle, since then will be a vertex in . Note that for any primitive discrete vector field in , our assumptions say that lies in a -cycle. Since any -cycle is also a -cycle, Lemma 2.6 says cannot lie in any -cycles other than this one. Hence any -cycle that contains a primitive discrete vector field in must be completely contained in . Finally, note that any non-trivial -cycle cannot be fully contained in since is acyclic. We conclude that any -cycle is a -cycle.

We reiterate that the analog of Lemma 2.6 is not true for simplicial complexes of dimension greater than , so this proof does not work outside the graph case.

Proposition 4.2.

The Morse complex is -connected.

Proof.

We induct on . The base case is that is non-empty once , which is clear. Now assume . By the Morse Lemma 3.3 and Proposition 2.4, it suffices to show that for a -simplex in with , the descending link is -connected. If there exists a primitive discrete vector field in that is not contained in any -cycle, then (and hence ) is contractible by Lemma 3.4. Now assume every primitive discrete vector field in is contained in a -cycle. Then by Lemma 3.5, , and by Proposition 4.1, , where is the set of simplices used in . Since , it now suffices to show that is -connected. By induction is -connected. Note that . This is because removing edges and their endpoints and all their incident edges would normally remove at most total edges from (here we use the fact that a vertex of representing an edge of has degree at most ), but since has at least one cycle we know that we only removed at most total edges. Also, , so is -connected. Since we know by Observation 1.3. Also, if then and we are done, so we can assume . Putting all this together we compute:

and we are done.

In the special case where , we can now draw conclusions about . Let us write and , so and is the maximum degree of a vertex in the Hasse diagram (which is usually the same as the maximum degree of a vertex in , unless every vertex of has degree or ).

Theorem 4.3.

The Morse complex is -connected. In particular is connected once , simply connected once , and -connected once .

Proof.

By Proposition 4.2 is -connected, i.e., -connected.

This proves Conjecture 2.8 when . Combining this with Theorem 2.7 we can obtain a higher connectivity bound that only depends on . Let .

Corollary 4.4.

The Morse complex is -connected. In particular is connected once , simply connected once , and -connected once .

Proof.

If has no vertices of degree then is a disjoint union of edges, and so is -connected, hence -connected. Now assume has a vertex of degree . By Theorem 2.7, is -connected. If then we are done, so assume . Then , and so we are done by Theorem 4.3.

4.1. Examples

Now we discuss a couple of examples. First let us discuss an example where the homotopy type of is already known, namely when is a complete graph. This example will show that, while our results are powerful in that they apply to any , they do not necessarily yield optimal bounds.

Example 4.5.

Let be the complete graph on vertices, so . By Corollary 4.4 is -connected once , i.e., once . For example it is connected once and simply connected once . Kozlov computed the homotopy type , namely is homotopy equivalent to a wedge of spheres of dimension Reference Koz99, Theorem 3.1, so in fact is already -connected once . This shows our bounds are not always optimal.

As a remark, Kozlov also computed the homotopy type of Reference Koz99, Proposition 5.2 for the -cycle graph. Since it is easy to compare our higher connectivity bounds to the actual higher connectivity, and again we see our bounds are not optimal.

Now we discuss an example where the homotopy type of is (to the best of our knowledge) not known, namely when is complete bipartite, and compute what our results reveal.

Example 4.6.

Let be the complete bipartite graph with vertices of one type and vertices of the other type, so . By Corollary 4.4 is -connected once . For example it is connected once and simply connected once . Later in Theorem 5.4 we will see that actually it is also simply connected once , i.e., every is simply connected except , which is not even connected.

5. Higher dimensional

Now we consider arbitrary dimensional , and prove some results about higher connectivity properties of . First we observe that gets arbitrarily highly connected as goes to .

Theorem 5.1.

Suppose that contains a -simplex. Then is -connected.

Proof.

Since contains a -simplex, contains a vertex of degree . The result is now immediate from Theorem 2.7.

Our next goal is to completely classify when is connected and simply connected. To do this we will first prove that is (simply) connected if and only if is, for which we will use Bestvina–Brady Morse theory applied to , as in Section 4. The key is that, even without a full understanding of in the case, we will only need to care that is non-empty. Bestvina–Brady Morse theory is probably a more powerful tool than necessary to relate to in this way, but it makes for an elegant argument.

Lemma 5.2 (Descending link simply connected).

Assume has no isolated vertices, and is not a -simplex or a -cycle. Then for any with , either is simply connected or is connected and is non-empty. In particular, the descending link is always simply connected.

Proof.

Say is a -simplex of . The fact that implies , by Observation 1.3. We see from Lemmas 3.4 and 3.5 that is either contractible or homeomorphic to . If this is simply connected. Now we have to prove that if then is non-empty.

Say . Since we know that , , , , , , is a -cycle, so the all have the same dimension, say , and the all have dimension . First suppose . Then , so we can choose a -face and -face such that is disjoint from and . In particular is a discrete vector field, and it is clear that , so .

Now suppose , so the are vertices and the are edges. If contains an edge not equal to any then must have at least one vertex not equal to any . In this case is a discrete vector field, and it is clear that , so . Finally, suppose does not contain any edges besides the . Since isolated vertices do not contribute to the Morse complex we can assume has none, so the only options are that equals a -simplex or a -cycle, but we ruled these out.

Corollary 5.3.

We have that is connected if and only if is connected, and is simply connected if and only if is simply connected.

Proof.

First note that by Observation 1.3, so the connectivity result is true. Now we prove the simple connectivity result. If is a -simplex or a -cycle, then Examples 1.4 and 1.5 show that the result holds. Now assume is neither of these (and note that we may assume has no isolated vertices). By Lemma 5.2 the descending link of every with is simply connected. Thus by the Morse Lemma 3.3, the inclusion induces an isomorphism in .

Now we can completely classify when is connected and simply connected.

Theorem 5.4.

Suppose has no isolated vertices. The Morse complex is connected if and only if is not an edge, and is simply connected if and only if is none of: an edge, a disjoint union of two edges, a path with three edges, a -cycle, or a -simplex.

Proof.

First we prove the connectivity statement. If has a vertex with degree more than then , and hence , is connected by Lemma 2.5 and Corollary 5.3. Now assume has no vertices with degree more than , so is a disjoint union of edges. If has at least two edges (or, for trivial reasons, zero edges) then it is easy to check that is connected. If has one edge then is not connected.

Now we prove the simple connectivity statement. If has a vertex with degree more than then , and hence , is simply connected by Lemma 2.5 and Corollary 5.3. Now assume has no vertices with degree more than . Then is a disjoint union of some number of -simplices, cycle graphs, and path graphs. If has more than one connected component, and is not a disjoint union of two edges, then is a join of at least two non-empty complexes, at least one of which is connected, and so is simply connected. If is a disjoint union of two edges then is not simply connected. Now assume is connected. If it is a -simplex then is not simply connected (Example 1.5). If is an -cycle then is simply connected unless Reference Koz99, Proposition 5.2. If is a path with edges then is the matching complex of a path with edges, which is easily seen to be simply connected unless is or , so the same is true of by Corollary 5.3.

A consequence of Theorem 5.4 is that we can now verify the connectivity and simple connectivity cases of Conjecture 2.8. Let us use the notation and as before, so is the number of edges in the Hasse diagram of and is the maximum degree of a vertex in the Hasse diagram.

Corollary 5.5.

If then is connected. If then is simply connected.

Proof.

Since isolated vertices do not contribute to , , or , we can assume there are none. If is not connected then is an edge by Theorem 5.4, so . If is not simply connected then is either an edge, a disjoint union of two edges, a path with three edges, a -cycle, or a -simplex. In all these cases one can compute that .

Acknowledgments

We are grateful to the organizers of the 2019 Union College Mathematics Conference, where the idea for this project originated. We are also very grateful to the anonymous referees, for many helpful comments and in particular for suggestions that led us to proving Theorem 2.7 and its ramifications, which greatly strengthened our results.

Figures

Figure 1.

The Hasse diagram (left), the generalized Morse complex (top), and the Morse complex (bottom)

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=.7] \coordinate(z) at (0,0); \coordinate(y) at (2,0); \coordinate(x) at (4,0); \coordinate(w) at (3,1.732); \coordinate(v) at (2,3.464); \coordinate(u) at (1,1.732); \draw[line width=1] (z) -- (x) -- (v) -- (z); \filldraw(z) circle (1.5pt); \filldraw(y) circle (1.5pt); \filldraw(x) circle (1.5pt); \filldraw(w) circle (1.5pt); \filldraw(v) circle (1.5pt); \filldraw(u) circle (1.5pt); \node at ($.5*(z)+.5*(u)+(-.25,.15)$) {$a$}; \node at ($.5*(u)+.5*(v)+(-.25,.15)$) {$b$}; \node at ($.5*(v)+.5*(w)+(.25,.15)$) {$c$}; \node at ($.5*(w)+.5*(x)+(.25,.15)$) {$d$}; \node at ($.5*(x)+.5*(y)+(0,-.3)$) {$e$}; \node at ($.5*(y)+.5*(z)+(0,-.3)$) {$f$}; \begin{scope}[xshift=6cm,yshift=4cm] \coordinate(a) at (0,0); \coordinate(c) at (1.732,-1); \coordinate(e) at (0,-2); \coordinate(f) at (3.732,-1); \coordinate(d) at (5.464,0); \coordinate(b) at (5.464,-2); \filldraw[lightgray] (a) -- (c) -- (e); \filldraw[lightgray] (f) -- (d) -- (b); \draw[line width=1] (c) -- (e) -- (a) -- (c) -- (f) -- (d) -- (b) -- (f); \draw[line width=1] (a) -- (d) (e) -- (b); \filldraw(a) circle (2pt); \filldraw(b) circle (2pt); \filldraw(c) circle (2pt); \filldraw(d) circle (2pt); \filldraw(e) circle (2pt); \filldraw(f) circle (2pt); \node at ($(a)-(.3,0)$) {$a$}; \node at ($(e)-(.3,0)$) {$e$}; \node at ($(c)+(0,.3)$) {$c$}; \node at ($(f)+(0,.4)$) {$f$}; \node at ($(d)+(.3,0)$) {$d$}; \node at ($(b)+(.3,0)$) {$b$}; \end{scope} \begin{scope}[xshift=6cm,yshift=1cm] \coordinate(a) at (0,0); \coordinate(c) at (1.732,-1); \coordinate(e) at (0,-2); \coordinate(f) at (3.732,-1); \coordinate(d) at (5.464,0); \coordinate(b) at (5.464,-2); \draw[line width=1] (c) -- (e) -- (a) -- (c) -- (f) -- (d) -- (b) -- (f); \draw[line width=1] (a) -- (d) (e) -- (b); \filldraw(a) circle (2pt); \filldraw(b) circle (2pt); \filldraw(c) circle (2pt); \filldraw(d) circle (2pt); \filldraw(e) circle (2pt); \filldraw(f) circle (2pt); \node at ($(a)-(.3,0)$) {$a$}; \node at ($(e)-(.3,0)$) {$e$}; \node at ($(c)+(0,.3)$) {$c$}; \node at ($(f)+(0,.4)$) {$f$}; \node at ($(d)+(.3,0)$) {$d$}; \node at ($(b)+(.3,0)$) {$b$}; \end{scope} \end{tikzpicture}
Figure 2.

The Hasse diagram (left), the generalized Morse complex (top), and the Morse complex (bottom)

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=.8] \coordinate(z) at (0,0); \coordinate(y) at (2,0); \coordinate(x) at (4,0); \coordinate(w) at (3,1.732); \coordinate(v) at (2,3.464); \coordinate(u) at (1,1.732); \coordinate(t) at (2,1.155); \draw[line width=1] (z) -- (x) -- (v) -- (z); \draw[line width=1] (y) -- (t) -- (u) (t) -- (w); \filldraw(z) circle (1.5pt); \filldraw(y) circle (1.5pt); \filldraw(x) circle (1.5pt); \filldraw(w) circle (1.5pt); \filldraw(v) circle (1.5pt); \filldraw(u) circle (1.5pt); \filldraw(t) circle (1.5pt); \node at ($.5*(z)+.5*(u)+(-.25,.15)$) {$a$}; \node at ($.5*(u)+.5*(v)+(-.25,.15)$) {$b$}; \node at ($.5*(v)+.5*(w)+(.25,.15)$) {$c$}; \node at ($.5*(w)+.5*(x)+(.25,.15)$) {$d$}; \node at ($.5*(x)+.5*(y)+(0,-.3)$) {$e$}; \node at ($.5*(y)+.5*(z)+(0,-.3)$) {$f$}; \node at ($.5*(u)+.5*(t)+(.2,.15)$) {$g$}; \node at ($.5*(w)+.5*(t)+(-.2,.17)$) {$h$}; \node at ($.5*(y)+.5*(t)+(.2,0)$) {$i$}; \begin{scope}[xshift=6cm,yshift=5cm] \coordinate(a) at (0,0); \coordinate(c) at (1.732,-1); \coordinate(e) at (0,-2); \coordinate(f) at (3.732,-1); \coordinate(d) at (5.464,0); \coordinate(b) at (5.464,-2); \coordinate(g) at ($.5*(c)+.5*(f)+(0,.2)$); \coordinate(h) at ($.5*(e)+.5*(b)+(-1,-1)$); \coordinate(i) at ($.5*(a)+.5*(d)+(1,1)$); \filldraw[lightgray] (a) -- (c) -- (e); \draw[line width=1] (c) -- (e) -- (a) -- (c); \filldraw[lightgray] (f) -- (d) -- (b); \draw[line width=1] (f) -- (d) -- (b) -- (f); \filldraw[lightgray] (f) -- (g) -- (d); \draw[line width=1] (f) -- (g) -- (d) -- (f); \filldraw[lightgray] (f) -- (g) -- (c); \draw[line width=1] (f) -- (g) -- (c) -- (f); \filldraw[lightgray] (e) -- (g) -- (c); \draw[line width=1] (e) -- (g) -- (c) -- (e); \filldraw[lightgray] (f) -- (b) -- (h); \draw[line width=1] (f) -- (b) -- (h) -- (f); \filldraw[lightgray] (e) -- (b) -- (h); \draw[line width=1] (e) -- (b) -- (h) -- (e); \filldraw[lightgray] (e) -- (a) -- (h); \draw[line width=1] (e) -- (a) -- (h) -- (e); \filldraw[lightgray] (c) -- (a) -- (i); \draw[line width=1] (c) -- (a) -- (i) -- (c); \filldraw[lightgray] (d) -- (a) -- (i); \draw[line width=1] (d) -- (a) -- (i) -- (d); \filldraw[lightgray] (d) -- (b) -- (i); \draw[line width=1] (d) -- (b) -- (i) -- (d); \filldraw(a) circle (2pt); \filldraw(b) circle (2pt); \filldraw(c) circle (2pt); \filldraw(d) circle (2pt); \filldraw(e) circle (2pt); \filldraw(f) circle (2pt); \filldraw(g) circle (2pt); \filldraw(h) circle (2pt); \filldraw(i) circle (2pt); \node at ($(a)-(.3,0)$) {$a$}; \node at ($(e)-(.3,0)$) {$e$}; \node at ($(c)+(0,.3)$) {$c$}; \node at ($(f)+(0,-.4)$) {$f$}; \node at ($(d)+(.3,0)$) {$d$}; \node at ($(b)+(.3,0)$) {$b$}; \node at ($(g)+(0,.3)$) {$g$}; \node at ($(h)+(0,-.3)$) {$h$}; \node at ($(i)+(0,.3)$) {$i$}; \end{scope} \begin{scope}[xshift=6cm,yshift=0cm] \coordinate(a) at (0,0); \coordinate(c) at (1.732,-1); \coordinate(e) at (0,-2); \coordinate(f) at (3.732,-1); \coordinate(d) at (5.464,0); \coordinate(b) at (5.464,-2); \coordinate(g) at ($.5*(c)+.5*(f)+(0,.2)$); \coordinate(h) at ($.5*(e)+.5*(b)+(-1,-1)$); \coordinate(i) at ($.5*(a)+.5*(d)+(1,1)$); \draw[line width=1] (c) -- (e) -- (a) -- (c); \draw[line width=1] (f) -- (d) -- (b) -- (f); \filldraw[lightgray] (f) -- (g) -- (d); \draw[line width=1] (f) -- (g) -- (d) -- (f); \filldraw[lightgray] (f) -- (g) -- (c); \draw[line width=1] (f) -- (g) -- (c) -- (f); \filldraw[lightgray] (e) -- (g) -- (c); \draw[line width=1] (e) -- (g) -- (c) -- (e); \filldraw[lightgray] (f) -- (b) -- (h); \draw[line width=1] (f) -- (b) -- (h) -- (f); \filldraw[lightgray] (e) -- (b) -- (h); \draw[line width=1] (e) -- (b) -- (h) -- (e); \filldraw[lightgray] (e) -- (a) -- (h); \draw[line width=1] (e) -- (a) -- (h) -- (e); \filldraw[lightgray] (c) -- (a) -- (i); \draw[line width=1] (c) -- (a) -- (i) -- (c); \filldraw[lightgray] (d) -- (a) -- (i); \draw[line width=1] (d) -- (a) -- (i) -- (d); \filldraw[lightgray] (d) -- (b) -- (i); \draw[line width=1] (d) -- (b) -- (i) -- (d); \filldraw(a) circle (2pt); \filldraw(b) circle (2pt); \filldraw(c) circle (2pt); \filldraw(d) circle (2pt); \filldraw(e) circle (2pt); \filldraw(f) circle (2pt); \filldraw(g) circle (2pt); \filldraw(h) circle (2pt); \filldraw(i) circle (2pt); \node at ($(a)-(.3,0)$) {$a$}; \node at ($(e)-(.3,0)$) {$e$}; \node at ($(c)+(0,.3)$) {$c$}; \node at ($(f)+(0,-.4)$) {$f$}; \node at ($(d)+(.3,0)$) {$d$}; \node at ($(b)+(.3,0)$) {$b$}; \node at ($(g)+(0,.3)$) {$g$}; \node at ($(h)+(0,-.3)$) {$h$}; \node at ($(i)+(0,.3)$) {$i$}; \end{scope} \end{tikzpicture}

Mathematical Fragments

Observation 1.3 (1-skeleton).

The -skeleton of coincides with that of .

Example 1.4.

Let be the -cycle, i.e, the cyclic graph with vertices. See Figure 1 for drawings of , , and . We see that and (this computation of agrees with Kozlov’s computation in Reference Koz99, Proposition 5.2). In particular neither nor is simply connected.

Example 1.5.

Let be the -simplex. See Figure 2 for drawings of , , and . We see that and (this computation of agrees with Chari and Joswig’s Reference CJ05, Proposition 5.1). In particular neither nor is simply connected.

Theorem 2.2 (Groundedness trick).

Reference BF19, Theorem 4.9 Every -grounded flag complex is -connected.

Proposition 2.4.

The complex is -connected.

Lemma 2.5.

Suppose has a vertex that has degree in . Then is -grounded, and hence -connected.

Lemma 2.6.

Let be a primitive discrete vector field in a discrete vector field on . Then lies in at most one simple -cycle.

Theorem 2.7.

If has a vertex with degree in then is -connected.

Conjecture 2.8.

The Morse complex is -connected. In particular if then is -connected.

Definition 3.1.

Let be a simplicial complex and two functions such that for any adjacent vertices we have . Extend and to maps by extending affinely to each simplex. Then we call

a Bestvina–Brady discrete Morse function provided the following holds: for any infinite sequence , , …of vertices such that for each , is adjacent to and lexicographically, the set has no lower bound in .

Example 3.2.

Let be the barycentric subdivision of a simplicial complex , so the vertices of are the simplices of and adjacency in is determined by incidence in . Let be any function. Let be the function sending (viewed as a vertex of ) to (viewed as a simplex of ). If is finite dimensional and is closed and discrete (for example this holds if is finite), then is a Bestvina–Brady discrete Morse function. Indeed, adjacent vertices of have different values (hence different values), and the finite dimensionality of plus the fact that is closed and discrete ensures that the last condition of Definition 3.1 is satisfied.

Lemma 3.3 (Morse Lemma).

Let be a Bestvina–Brady discrete Morse function on a simplicial complex . Let in . If is -connected for all vertices with then the inclusion induces an isomorphism in for all , and an epimorphism in .

Lemma 3.4 (Descending face link, case 1).

Let with , so is a discrete vector field on (avoiding ) with at least one -cycle. If there exists a primitive discrete vector field in that is not contained in any -cycle, then is contractible.

Lemma 3.5 (Descending face link, case 2).

Let with , say is a -simplex of . If every primitive discrete vector field in is contained in a -cycle, then is homeomorphic to .

Proposition 4.1 (Modeling the descending coface link).

Let be a -simplex in such that every primitive discrete vector field in lies in a -cycle. Let be the set of simplices of used by . Then in is isomorphic to .

Proposition 4.2.

The Morse complex is -connected.

Theorem 4.3.

The Morse complex is -connected. In particular is connected once , simply connected once , and -connected once .

Corollary 4.4.

The Morse complex is -connected. In particular is connected once , simply connected once , and -connected once .

Lemma 5.2 (Descending link simply connected).

Assume has no isolated vertices, and is not a -simplex or a -cycle. Then for any with , either is simply connected or is connected and is non-empty. In particular, the descending link is always simply connected.

Corollary 5.3.

We have that is connected if and only if is connected, and is simply connected if and only if is simply connected.

Theorem 5.4.

Suppose has no isolated vertices. The Morse complex is connected if and only if is not an edge, and is simply connected if and only if is none of: an edge, a disjoint union of two edges, a path with three edges, a -cycle, or a -simplex.

References

Reference [BB97]
Mladen Bestvina and Noel Brady, Morse theory and finiteness properties of groups, Invent. Math. 129 (1997), no. 3, 445–470, DOI 10.1007/s002220050168. MR1465330,
Show rawAMSref \bib{bestvina97}{article}{ label={BB97}, author={Bestvina, Mladen}, author={Brady, Noel}, title={Morse theory and finiteness properties of groups}, journal={Invent. Math.}, volume={129}, date={1997}, number={3}, pages={445--470}, issn={0020-9910}, review={\MR {1465330}}, doi={10.1007/s002220050168}, }
Reference [BF19]
James Belk and Bradley Forrest, Rearrangement groups of fractals, Trans. Amer. Math. Soc. 372 (2019), no. 7, 4509–4552, DOI 10.1090/tran/7386. MR4009393,
Show rawAMSref \bib{belk19}{article}{ label={BF19}, author={Belk, James}, author={Forrest, Bradley}, title={Rearrangement groups of fractals}, journal={Trans. Amer. Math. Soc.}, volume={372}, date={2019}, number={7}, pages={4509--4552}, issn={0002-9947}, review={\MR {4009393}}, doi={10.1090/tran/7386}, }
Reference [BFM16]
Kai-Uwe Bux, Martin G. Fluch, Marco Marschler, Stefan Witzel, and Matthew C. B. Zaremsky, The braided Thompson’s groups are of type , J. Reine Angew. Math. 718 (2016), 59–101, DOI 10.1515/crelle-2014-0030. With an appendix by Zaremsky. MR3545879,
Show rawAMSref \bib{bux16}{article}{ label={BFM{\etalchar {+}}16}, author={Bux, Kai-Uwe}, author={Fluch, Martin G.}, author={Marschler, Marco}, author={Witzel, Stefan}, author={Zaremsky, Matthew C. B.}, title={The braided Thompson's groups are of type $\mathrm F_\infty $}, note={With an appendix by Zaremsky}, journal={J. Reine Angew. Math.}, volume={718}, date={2016}, pages={59--101}, issn={0075-4102}, review={\MR {3545879}}, doi={10.1515/crelle-2014-0030}, }
Reference [BGM20]
Margaret Bayer, Bennet Goeckner, and Marija Jelić Milutinović, Manifold matching complexes, Mathematika 66 (2020), no. 4, 973–1002, DOI 10.1112/mtk.12049. MR4143161,
Show rawAMSref \bib{bayer20}{article}{ label={BGM20}, author={Bayer, Margaret}, author={Goeckner, Bennet}, author={Milutinovi\'{c}, Marija Jeli\'{c}}, title={Manifold matching complexes}, journal={Mathematika}, volume={66}, date={2020}, number={4}, pages={973--1002}, issn={0025-5793}, review={\MR {4143161}}, doi={10.1112/mtk.12049}, }
Reference [BLVŽ94]
A. Björner, L. Lovász, S. T. Vrećica, and R. T. Živaljević, Chessboard complexes and matching complexes, J. London Math. Soc. (2) 49 (1994), no. 1, 25–39, DOI 10.1112/jlms/49.1.25. MR1253009,
Show rawAMSref \bib{bjoerner94}{article}{ label={BLV{\v {Z}}94}, author={Bj\"{o}rner, A.}, author={Lov\'{a}sz, L.}, author={Vre\'{c}ica, S. T.}, author={\v {Z}ivaljevi\'{c}, R. T.}, title={Chessboard complexes and matching complexes}, journal={J. London Math. Soc. (2)}, volume={49}, date={1994}, number={1}, pages={25--39}, issn={0024-6107}, review={\MR {1253009}}, doi={10.1112/jlms/49.1.25}, }
Reference [CJ05]
Manoj K. Chari and Michael Joswig, Complexes of discrete Morse functions, Discrete Math. 302 (2005), no. 1-3, 39–51, DOI 10.1016/j.disc.2004.07.027. MR2179635,
Show rawAMSref \bib{chari05}{article}{ label={CJ05}, author={Chari, Manoj K.}, author={Joswig, Michael}, title={Complexes of discrete Morse functions}, journal={Discrete Math.}, volume={302}, date={2005}, number={1-3}, pages={39--51}, issn={0012-365X}, review={\MR {2179635}}, doi={10.1016/j.disc.2004.07.027}, }
Reference [CM17]
Nicolas Ariel Capitelli and Elias Gabriel Minian, A simplicial complex is uniquely determined by its set of discrete Morse functions, Discrete Comput. Geom. 58 (2017), no. 1, 144–157, DOI 10.1007/s00454-017-9865-z. MR3658332,
Show rawAMSref \bib{capitelli17}{article}{ label={CM17}, author={Capitelli, Nicolas Ariel}, author={Minian, Elias Gabriel}, title={A simplicial complex is uniquely determined by its set of discrete Morse functions}, journal={Discrete Comput. Geom.}, volume={58}, date={2017}, number={1}, pages={144--157}, issn={0179-5376}, review={\MR {3658332}}, doi={10.1007/s00454-017-9865-z}, }
Reference [For98]
Robin Forman, Morse theory for cell complexes, Adv. Math. 134 (1998), no. 1, 90–145, DOI 10.1006/aima.1997.1650. MR1612391,
Show rawAMSref \bib{forman98}{article}{ label={For98}, author={Forman, Robin}, title={Morse theory for cell complexes}, journal={Adv. Math.}, volume={134}, date={1998}, number={1}, pages={90--145}, issn={0001-8708}, review={\MR {1612391}}, doi={10.1006/aima.1997.1650}, }
Reference [Koz99]
Dmitry N. Kozlov, Complexes of directed trees, J. Combin. Theory Ser. A 88 (1999), no. 1, 112–122, DOI 10.1006/jcta.1999.2984. MR1713484,
Show rawAMSref \bib{kozlov99}{article}{ label={Koz99}, author={Kozlov, Dmitry N.}, title={Complexes of directed trees}, journal={J. Combin. Theory Ser. A}, volume={88}, date={1999}, number={1}, pages={112--122}, issn={0097-3165}, review={\MR {1713484}}, doi={10.1006/jcta.1999.2984}, }
Reference [Koz08]
Dmitry Kozlov, Combinatorial algebraic topology, Algorithms and Computation in Mathematics, vol. 21, Springer, Berlin, 2008, DOI 10.1007/978-3-540-71962-5. MR2361455,
Show rawAMSref \bib{kozlov08}{book}{ label={Koz08}, author={Kozlov, Dmitry}, title={Combinatorial algebraic topology}, series={Algorithms and Computation in Mathematics}, volume={21}, publisher={Springer, Berlin}, date={2008}, pages={xx+389}, isbn={978-3-540-71961-8}, review={\MR {2361455}}, doi={10.1007/978-3-540-71962-5}, }
Reference [LS21]
Maxwell Lin and Nicholas A. Scoville, On the automorphism group of the Morse complex, Adv. in Appl. Math. 131 (2021), Paper No. 102250, 17, DOI 10.1016/j.aam.2021.102250. MR4290144,
Show rawAMSref \bib{lin21}{article}{ label={LS21}, author={Lin, Maxwell}, author={Scoville, Nicholas A.}, title={On the automorphism group of the Morse complex}, journal={Adv. in Appl. Math.}, volume={131}, date={2021}, pages={Paper No. 102250, 17}, issn={0196-8858}, review={\MR {4290144}}, doi={10.1016/j.aam.2021.102250}, }
Reference [PdM82]
Jacob Palis Jr. and Welington de Melo, Geometric theory of dynamical systems, Springer-Verlag, New York-Berlin, 1982. An introduction; Translated from the Portuguese by A. K. Manning. MR669541,
Show rawAMSref \bib{Palis82}{book}{ label={PdM82}, author={Palis, Jacob, Jr.}, author={de Melo, Welington}, title={Geometric theory of dynamical systems}, note={An introduction; Translated from the Portuguese by A. K. Manning}, publisher={Springer-Verlag, New York-Berlin}, date={1982}, pages={xii+198}, isbn={0-387-90668-1}, review={\MR {669541}}, }
Reference [Qui78]
Daniel Quillen, Homotopy properties of the poset of nontrivial -subgroups of a group, Adv. in Math. 28 (1978), no. 2, 101–128, DOI 10.1016/0001-8708(78)90058-0. MR493916,
Show rawAMSref \bib{quillen78}{article}{ label={Qui78}, author={Quillen, Daniel}, title={Homotopy properties of the poset of nontrivial $p$-subgroups of a group}, journal={Adv. in Math.}, volume={28}, date={1978}, number={2}, pages={101--128}, issn={0001-8708}, review={\MR {493916}}, doi={10.1016/0001-8708(78)90058-0}, }
Reference [Sco19]
Nicholas A. Scoville, Discrete Morse theory, Student Mathematical Library, vol. 90, American Mathematical Society, Providence, RI, 2019, DOI 10.1090/stml/090. MR3970274,
Show rawAMSref \bib{scoville19}{book}{ label={Sco19}, author={Scoville, Nicholas A.}, title={Discrete Morse theory}, series={Student Mathematical Library}, volume={90}, publisher={American Mathematical Society, Providence, RI}, date={2019}, pages={xiv+273}, isbn={978-1-4704-5298-8}, review={\MR {3970274}}, doi={10.1090/stml/090}, }
Reference [SWZ19]
Rachel Skipper, Stefan Witzel, and Matthew C. B. Zaremsky, Simple groups separated by finiteness properties, Invent. Math. 215 (2019), no. 2, 713–740, DOI 10.1007/s00222-018-0835-8. MR3910073,
Show rawAMSref \bib{skipper19}{article}{ label={SWZ19}, author={Skipper, Rachel}, author={Witzel, Stefan}, author={Zaremsky, Matthew C. B.}, title={Simple groups separated by finiteness properties}, journal={Invent. Math.}, volume={215}, date={2019}, number={2}, pages={713--740}, issn={0020-9910}, review={\MR {3910073}}, doi={10.1007/s00222-018-0835-8}, }
Reference [Zar]
Matthew C. B. Zaremsky, Bestvina–Brady discrete Morse theory and Vietoris–Rips complexes, Amer. J. Math., To appear, arXiv:1812.10976.

Article Information

MSC 2020
Primary: 55U05 (Abstract complexes in algebraic topology)
Secondary: 57Q05 (General topology of complexes)
Keywords
  • Morse complex
  • higher connectivity
  • discrete Morse theory
Author Information
Nicholas A. Scoville
Department of Mathematics and Computer Science, Ursinus College, Collegeville, Pennsylvania 19426
nscoville@ursinus.edu
ORCID
MathSciNet
Matthew C. B. Zaremsky
Department of Mathematics and Statistics, University at Albany (SUNY), Albany, New York 12222
mzaremsky@albany.edu
MathSciNet
Additional Notes

The second author was supported by grant #635763 from the Simons Foundation.

Communicated by
Patricia Hersh
Journal Information
Proceedings of the American Mathematical Society, Series B, Volume 9, Issue 14, ISSN 2330-1511, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , , , and published on .
Copyright Information
Copyright 2022 by the authors under Creative Commons Attribution 3.0 License (CC BY 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/bproc/115
  • MathSciNet Review: 4407041
  • Show rawAMSref \bib{4407041}{article}{ author={Scoville, Nicholas}, author={Zaremsky, Matthew}, title={Higher connectivity of the Morse complex}, journal={Proc. Amer. Math. Soc. Ser. B}, volume={9}, number={14}, date={2022}, pages={135-149}, issn={2330-1511}, review={4407041}, doi={10.1090/bproc/115}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.