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

Persistence Diagram Bundles:
A Multidimensional Generalization of Vineyards

Abigail Hickok
Abstract.

I introduce the concept of a persistence diagram (PD) bundle, which is the space of PDs for a fibered filtration function (a set {fp:𝒦p}pB\{f_{p}:\mathcal{K}^{p}\to\mathbb{R}\}_{p\in B} of filtrations that is parameterized by a topological space BB). Special cases include vineyards, the persistent homology transform, and fibered barcodes for multiparameter persistence modules. I prove that if BB is a smooth compact manifold, then for a generic fibered filtration function, BB is stratified such that within each stratum YBY\subseteq B, there is a single PD “template” (a list of “birth” and “death” simplices) that can be used to obtain the PD for the filtration fpf_{p} for any pYp\in Y. If BB is compact, then there are finitely many strata, so the PD bundle for a generic fibered filtration on BB is determined by the persistent homology at finitely many points in BB. I also show that not every local section can be extended to a global section (a continuous map ss from BB to the total space EE of PDs such that s(p)PD(fp)s(p)\in\text{PD}(f_{p}) for all pBp\in B). Consequently, a PD bundle is not necessarily the union of “vines” γ:BE\gamma:B\to E; this is unlike a vineyard. When there is a stratification as described above, I construct a cellular sheaf that stores sufficient data to construct sections and determine whether a given local section can be extended to a global section.

1. Introduction

In topological data analysis (TDA), our aim is to understand the global shape of a data set. Often, the data set takes the form of a collection of points in n\mathbb{R}^{n}, called a point cloud, and we hope to analyze the topology of a lower-dimensional space that the points lie on. TDA has found applications in a variety of fields, such as biology [29], neuroscience [12], and chemistry [27].

We use persistent homology (PH), a tool from algebraic topology [21]. The first step of persistent homology is to construct a filtered complex from our data; a filtered complex is a nested sequence

(1) 𝒦r0𝒦r1𝒦rn\mathcal{K}_{r_{0}}\subseteq\mathcal{K}_{r_{1}}\subseteq\cdots\subseteq\mathcal{K}_{r_{n}}\subseteq\cdots

of simplicial complexes. For example, one of the standard ways to build a filtered complex from point cloud data is to construct the Vietoris–Rips filtered complex. At filtration-parameter value rr, the Vietoris–Rips complex 𝒦r\mathcal{K}_{r} includes a simplex for every subset of points within rr of each other. In persistent homology, one studies how the topology of 𝒦r\mathcal{K}_{r} changes as the filtration parameter-value rr increases. As rr grows, new homology classes (which represent “holes” in the data) are born and old homology classes die. One way of summarizing this information is a persistence diagram: a multiset of points in the extended plane ¯2\overline{\mathbb{R}}^{2}. If there is a homology class that is born at filtration-parameter value bb and dies at filtration-parameter value dd, the persistence diagram contains the point (b,d)(b,d).

Developing new methods for analyzing how the topology of a data set changes as multiple parameters vary is a very active area of research [7]. For example, if a point cloud evolves over time (i.e., it is a dynamic metric space), then one maybe interested in using time as a second parameter, in addition to the filtration parameter rr. Common examples of time-evolving point clouds include swarming or flocking animals whose positions and/or velocities are represented by points ([15, 36, 25]). In such cases, one can obtain a filtered complex 𝒦r0t𝒦r1tKrnt\mathcal{K}_{r_{0}}^{t}\subseteq\mathcal{K}_{r_{1}}^{t}\subseteq\cdots\subseteq K_{r_{n}}^{t} at every time tt by constructing, e.g., the Vietoris–Rips filtered complex for the point cloud at time tt. It is also common to use the density of the point cloud as a parameter ([30, 10, 6]). Many other parameters can also vary in the topological analysis of point clouds or other types of data sets.

One can use a vineyard [14] to study a 1-parameter family of filtrations {𝒦r0t𝒦r1t𝒦rnt}t\{\mathcal{K}_{r_{0}}^{t}\subseteq\mathcal{K}_{r_{1}}^{t}\subseteq\cdots\mathcal{K}_{r_{n}}^{t}\}_{t\in\mathbb{R}} such as that obtained from a time-varying point cloud. At each tt\in\mathbb{R}, one can compute the PH of the filtration 𝒦r0t𝒦r1t𝒦rnt\mathcal{K}_{r_{0}}^{t}\subseteq\mathcal{K}_{r_{1}}^{t}\subseteq\cdots\subseteq\mathcal{K}_{r_{n}}^{t} and obtain a persistence diagram PD(t)(t). A vineyard is visualized as the continuously-varying “stack of PDs” {PD(t)}t\{\textnormal{PD}(t)\}_{t\in\mathbb{R}}. See Figure 1 for an illustration. As tt\in\mathbb{R} varies, the points in the PDs trace out curves (“vines”) in 3\mathbb{R}^{3}. Each vine corresponds to a homology class (i.e., one of the holes in the data), and shows how the persistence of that homology class changes with time (or, more generally, as some other parameter varies).

Refer to caption
Figure 1. An example of a vineyard. There is a persistence diagram for each time tt. Each curve is a vine in the vineyard. (This figure is a slightly modified version of a figure that appeared originally in [26].)

However, one cannot use a vineyard for a set of filtrations that is parameterized by a space that is not a subset of \mathbb{R}. For example, suppose that we have a time-varying point cloud whose dynamics depend on some system-parameter values μ1,,μm\mu_{1},\ldots,\mu_{m}\in\mathbb{R}. Many such systems exist. For example, the D’Orsogna model is a multi-agent dynamical system that models attractive and repulsive interactions between particles [11]. Each particle is represented by a point in a point cloud. In certain parameter regimes, there are interesting topological features, such as mills or double mills [31]. For each time tt\in\mathbb{R} and for each μ1,,μm\mu_{1},\ldots,\mu_{m}\in\mathbb{R}, one can obtain a filtered complex

(2) 𝒦r0t,μ1,,μm𝒦r1t,μ1,,μm𝒦rnt,μ1,,μm\mathcal{K}_{r_{0}}^{t,\mu_{1},\ldots,\mu_{m}}\subseteq\mathcal{K}_{r_{1}}^{t,\mu_{1},\ldots,\mu_{m}}\subseteq\cdots\subseteq\mathcal{K}_{r_{n}}^{t,\mu_{1},\ldots,\mu_{m}}

by constructing, e.g., the Vietoris–Rips filtered complex for the point cloud at time tt at system-parameter values μ1,,μm\mu_{1},\ldots,\mu_{m}. A parameterized set of filtered complexes like the one in (2) cannot be studied using a vineyard for the simple reason that there are too many parameters.

Such a parameterized set of filtered complexes cannot be studied using multiparameter PH [8], either. A multifiltration is a set {𝒦𝒖}𝒖n\{\mathcal{K}_{\bm{u}}\}_{\bm{u}\in\mathbb{R}^{n}} of simplicial complexes such that 𝒦𝒖𝒦𝒗\mathcal{K}_{\bm{u}}\subseteq\mathcal{K}_{\bm{v}} whenever 𝒖𝒗\bm{u}\leq\bm{v}. Multiparameter PH is the 𝔽[x1,,xn]\mathbb{F}[x_{1},\ldots,x_{n}] module obtained by applying homology (over a field 𝔽\mathbb{F}) to a multifiltration. (For more details, see reference [8].) The parameterized set of filtered complexes in (2) is not typically a multifiltration because it is not necessarily the case that 𝒦rit,μ1,,μm𝒦rit,μ1,,μm\mathcal{K}_{r_{i}}^{t,\mu_{1},\ldots,\mu_{m}}\not\subseteq\mathcal{K}_{r_{i}}^{t^{\prime},\mu_{1}^{\prime},\ldots,\mu_{m}^{\prime}} for all values of tt, tt^{\prime}, {μi}\{\mu_{i}\}, and {μi}\{\mu_{i}^{\prime}\}. Not only is there not necessarily an inclusion 𝒦rit,μ1,,μm𝒦rit,μ1,,μm\mathcal{K}_{r_{i}}^{t,\mu_{1},\ldots,\mu_{m}}\xhookrightarrow{}\mathcal{K}_{r_{i}}^{t^{\prime},\mu_{1}^{\prime},\ldots,\mu_{m}^{\prime}}, but there is not any given simplicial map 𝒦rit,μ1,,μm𝒦rit,μ1,,μm\mathcal{K}_{r_{i}}^{t,\mu_{1},\ldots,\mu_{m}}\to\mathcal{K}_{r_{i}}^{t^{\prime},\mu_{1}^{\prime},\ldots,\mu_{m}^{\prime}}. Therefore, we cannot use multiparameter PH.

In what follows, we will work with a slightly different notion of filtered complex than that of (1). A filtration function is a function f:𝒦f:\mathcal{K}\to\mathbb{R}, where 𝒦\mathcal{K} is a simplicial complex, such that every sublevel set 𝒦r:={σ𝒦f(σ)r}\mathcal{K}_{r}:=\{\sigma\in\mathcal{K}\mid f(\sigma)\leq r\} is a simplicial complex (i.e., f(τ)f(σ)f(\tau)\leq f(\sigma) if τ\tau is a face of σ\sigma). For every rsr\leq s, we have that 𝒦r𝒦s\mathcal{K}_{r}\subseteq\mathcal{K}_{s}. A simplex σ𝒦\sigma\in\mathcal{K} appears in the filtration at r=f(σ)r=f(\sigma). By setting {ri}=Im(f)\{r_{i}\}=\textnormal{Im}(f), where ri<ri+1r_{i}<r_{i+1}, we obtain a nested sequence as in (1). Conversely, given a nested sequence of simplicial complexes, the associated filtration function is f(σ)=min{riσ𝒦ri}f(\sigma)=\min\{r_{i}\mid\sigma\in\mathcal{K}_{r_{i}}\}, with 𝒦=i𝒦ri\mathcal{K}=\bigcup_{i}\mathcal{K}_{r_{i}}.

1.1. Contributions

I introduce the concept of a persistence diagram (PD) bundle, in which PH varies over an arbitrary “base space” BB. A PD bundle gives a way of studying a fibered filtration function, which is a set {fp:𝒦p𝔽}pB\{f_{p}:\mathcal{K}^{p}\to\mathbb{F}\}_{p\in B} of functions such that fpf_{p} is a filtration of a simplicial complex 𝒦p\mathcal{K}^{p}. At each pBp\in B, the sublevel sets of fpf_{p} form a filtered complex. For example, in (2), we have B=n+1B=\mathbb{R}^{n+1} and we obtain a fibered filtration function {ft,μ1,,μm:𝒦}(t,μ1,,μm)m+1\{f_{t,\mu_{1},\ldots,\mu_{m}}:\mathcal{K}\to\mathbb{R}\}_{(t,\mu_{1},\ldots,\mu_{m})\in\mathbb{R}^{m+1}} by defining ft,μ1,,μmf_{t,\mu_{1},\ldots,\mu_{m}} to be the filtration function associated with the filtered complex in (2). The associated PD bundle is the space of persistence diagrams PD(fp)(f_{p}) as they vary with pBp\in B (see Definition 3.2). In the special case in which BB is an interval in \mathbb{R}, a PD bundle is equivalent to a vineyard.

I prove that for “generic” fibered filtration functions (see Section 4.2), the base space BB can be stratified in a way that makes PD bundles tractable to compute and analyze. Theorem 4.15 says that for a “generic” fibered filtration function on a smooth compact manifold BB, the base space BB is stratified such that within each stratum, there is a single PD “template” that can be used to obtain PD(fp)PD(f_{p}) at any point pp in the stratum. Proposition 4.5 shows that all “piecewise-linear” PD bundles (see Definition 3.4) have such a stratification. The template is a list of (birth, death) simplex pairs, and the diagram PD(fp)\text{PD}(f_{p}) is obtained by evaluating fpf_{p} on each simplex. In particular, when BB is a smooth compact manifold, the number of strata is finite, so the PD bundle is determined by the PH at a finite number of points in the base space.

I show that unlike vineyards, PD bundles do not necessarily decompose into a union of “vines”. More precisely, there may not exist continuous maps γ1,,γm:BE\gamma_{1},\ldots,\gamma_{m}:B\to E such that

(3) PD(fp)=i=1mγi(p)\text{PD}(f_{p})=\bigcup_{i=1}^{m}\gamma_{i}(p)

for all pBp\in B. This is a consequence of Proposition 5.3, in which it is shown that nontrivial global sections are not guaranteed to exist. That is, given a point z0PD(fp0)z_{0}\in PD(f_{p_{0}}) for some p0Bp_{0}\in B, it may not be possible to extend p0z0p_{0}\mapsto z_{0} to a continuous map s:BE:={(p,z)zPD(fp)}s:B\to E:=\{(p,z)\mid z\in PD(f_{p})\} such that s(p)PD(fp)s(p)\in\text{PD}(f_{p}) for all pp. This behavior is a feature that gives PD bundles a richer mathematical structure than vineyards.

For any fibered filtration with a stratification as described above (see Theorem 4.15 and Proposition 4.5), I construct a “compatible cellular sheaf” (see Section 6.2) that stores the data in the PD bundle. Rather than analyzing the entire PD bundle, which consists of continuously varying PDs over the base space BB, we can analyze the cellular sheaf, which is discrete. For example, in Proposition 6.4, I prove that an extension of p0z0p_{0}\mapsto z_{0} to a global section exists if a certain associated global section of the cellular sheaf exists. A compatible cellular sheaf stores sufficient data to reconstruct the associated PD bundle and analyze its sections.

Though not the focus of this paper, I also give a simple example of vineyard instability in Appendix A.4. It is often quoted in the research literature that “vineyards are unstable”; however, this “well-known fact” has been shared only in private correspondence and, to the best of my knowledge, has never been published. The example of vineyard instability is furnished from an example in Proposition 5.3.

1.2. Related work

PD bundles are a generalization of vineyards, which were introduced in [14]. Two other important special cases of PD bundles are the fibered barcode of a multiparameter persistence module [4] and the persistent homology transform (B=SnB=S^{n}) from shape analysis [2, 32]. I discuss the special case of fibered barcodes in detail in Section 3.2.2; the base space BB is a subset of the space of lines in n\mathbb{R}^{n}. The persistent homology transform (PHT) is defined for a constructible set Mn+1M\subseteq\mathbb{R}^{n+1}. For any unit vector vSnv\in S^{n}, one defines the filtration Mrv={xMxvr}M_{r}^{v}=\{x\in M\mid x\cdot v\leq r\} (i.e., the sublevel filtration of the height function with respect to the direction vv). PHT is the map that sends vSnv\in S^{n} to the persistence diagram for the filtration {Mrv}r\{M_{r}^{v}\}_{r\in\mathbb{R}}. The significance of PHT is that it is a sufficient statistic for shapes in 2\mathbb{R}^{2} and 3\mathbb{R}^{3} [32]. Applications of PHT are numerous and include protein docking [34], barley-seed shape analysis [3], and heel-bone analysis in primates [32].

For PHT, Curry et al. [18] proved that the base space SnS^{n} is stratified such that the PHT of a shape MM is determined by the PH of {Mrv}r\{M_{r}^{v}\}_{r\in\mathbb{R}} for finitely many directions vSnv\in S^{n} (one direction vv per stratum). This is related to the stratification given by Theorem 4.15, in which I show that a “generic” PD bundle whose base space BB is a compact smooth manifold (such as SnS^{n}) is similarly stratified and thus determined by finitely many points in BB (one pBp\in B per stratum). The primary difference between the stratifications in [18] and Theorem 4.15 is that in [18], each stratum is a subset in which the order of the vertices of a triangulated shape MM (as ordered by the height function) is constant, whereas in Theorem 4.15, each stratum is a subset in which the order of the simplices (as ordered by the filtration function) is constant. The stratification result of the present paper (Theorem 4.15) applies to general PD bundles, while [18] applies only to PHT.

The stratification that we study in the present paper is used in [23] to develop an algorithm for computing “piecewise-linear” PD bundles (see Definition 3.4). The algorithm relies on the fact that for any piecewise-linear PD bundle on a compact triangulated base space BB, there are a finite number of strata, so the PD bundle is determined by the PH at a finite number of points in BB.

The existence (or nonexistence) of nontrivial global sections in PD bundles is related to the study of “monodromy” in fibered barcodes of multiparameter persistence modules [9]. Cerri et al. [9] constructed an example in which there is a path through the fibered barcode that loops around a “singularity” (a PD in the fibered barcode for which there is a point in the PD with multiplicity greater than one) and finishes in a different place than where it starts.

1.3. Organization

This paper proceeds as follows. I review background on persistent homology in Section 2. In Section 3, I give the definition of a PD bundle, with some examples, and I compare PD bundles to multiparameter PH. In Section 4, I show how to stratify the base space BB into strata in which the (birth, death) simplex pairs are constant (see Theorem 4.15 and Proposition 4.5). I discuss sections of PD bundles and the existence of monodromy in Section 5. I construct a compatible cellular sheaf in Section 6.2. I conclude and discuss possible directions for future research in Section 7. In Appendix A.4, I use the example of monodromy from Section 5 to construct an example of vineyard instability. In Appendix A.5, I provide technical details that are needed to prove Theorem 4.15.

2. Background

We begin by reviewing persistent homology (PH) and cellular sheaves. For a more thorough treatment of PH, see [19, 28], and for more on cellular sheaves, see [22, 16].

2.1. Filtrations

Consider a simplicial complex 𝒦\mathcal{K}. A filtration function on 𝒦\mathcal{K} is a real-valued function f:𝒦f:\mathcal{K}\to\mathbb{R} such that if τ𝒦\tau\in\mathcal{K} is a face of σK\sigma\in K, then f(τ)f(σ)f(\tau)\leq f(\sigma). The filtration value of a simplex σ𝒦\sigma\in\mathcal{K} is f(σ)f(\sigma). The rr-sublevel sets 𝒦r:={σ𝒦f(σ)r}\mathcal{K}_{r}:=\{\sigma\in\mathcal{K}\mid f(\sigma)\leq r\} form a filtered complex. The condition that f(τ)f(σ)f(\tau)\leq f(\sigma) if τσ\tau\subseteq\sigma guarantees that 𝒦r\mathcal{K}_{r} is a simplicial complex for all rr. For all srs\leq r, we have 𝒦s𝒦r\mathcal{K}_{s}\subseteq\mathcal{K}_{r}. The parameter rr is the filtration parameter. For an example, see Figure 2.

Refer to caption
(a) 𝒦0\mathcal{K}_{0}
Refer to caption
(b) 𝒦1\mathcal{K}_{1}
Refer to caption
(c) 𝒦2\mathcal{K}_{2}
Refer to caption
(d) 𝒦3\mathcal{K}_{3}
Refer to caption
(e) 𝒦4\mathcal{K}_{4}
Figure 2. An example of a filtration. The simplicial complex 𝒦i\mathcal{K}_{i} has the associated filtration-parameter value ii. (This figure appeared originally in [24].)

For example, suppose that X={xi}i=1MX=\{x_{i}\}_{i=1}^{M} is a point cloud. Let 𝒦\mathcal{K} be the simplicial complex that has a simplex σ\sigma with vertices {xj}jJ\{x_{j}\}_{j\in J} for all J{1,,M}J\subseteq\{1,\ldots,M\}. The Vietoris–Rips filtration function is f(σ)=12maxj,kJ{xjxk}f(\sigma)=\frac{1}{2}\max_{j,k\in J}\{\lVert x_{j}-x_{k}\rVert\}, where {xj}jJ\{x_{j}\}_{j\in J} are the vertices of σ\sigma.

2.2. Persistent homology

Let f:𝒦f:\mathcal{K}\to\mathbb{R} be a filtration function on a finite simplicial complex 𝒦\mathcal{K}, and let {Kr}r\{K_{r}\}_{r\in\mathbb{R}} be the associated filtered complex. Let r1<<rNr_{1}<\cdots<r_{N} be the filtration values of the simplices of 𝒦\mathcal{K}. These are the critical points at which simplices are added to the filtration. For all s[ri,ri+1)s\in[r_{i},r_{i+1}), we have 𝒦s=𝒦ri\mathcal{K}_{s}=\mathcal{K}_{r_{i}}.

For all iji\leq j, the inclusion ιi,j:𝒦ri𝒦rj\iota^{i,j}:\mathcal{K}_{r_{i}}\hookrightarrow\mathcal{K}_{r_{j}} induces a map ιi,j:H(𝒦ri,𝔽)H(𝒦rj,𝔽)\iota^{i,j}_{*}:H_{*}(\mathcal{K}_{r_{i}},\mathbb{F})\to H_{*}(\mathcal{K}_{r_{j}},\mathbb{F}) on homology, where 𝔽\mathbb{F} is a field that we set to /2\mathbb{Z}/2\mathbb{Z} for the rest of this paper. The qqth-persistent homology (PH) is the pair

({Hq(𝒦ri,𝔽)}1iN,{ιi,j}1ijN).\Big{(}\{H_{q}(\mathcal{K}_{r_{i}},\mathbb{F})\}_{1\leq i\leq N}\,,\{\iota_{*}^{i,j}\}_{1\leq i\leq j\leq N}\Big{)}\,.

The Fundamental Theorem of Persistent Homology yields compatible choices of bases for the vector spaces Hq(𝒦ri,𝔽)H_{q}(\mathcal{K}_{r_{i}},\mathbb{F}), which we use below in our definition of a persistence diagram.

The qqth persistence diagram PDq(f)PD_{q}(f) is a multiset of points in the extended plane ¯2\overline{\mathbb{R}}^{2} that summarizes the qqth-persistent homology. The PD contains the diagonal as well as a point for every generator. We say that a generator γHq(𝒦ri,𝔽)\gamma\in H_{q}(\mathcal{K}_{r_{i}},\mathbb{F}) is born at rir_{i} if it is not in the image of ιi,i1\iota^{i,i-1}_{*}. The homology class γ\gamma subsequently dies at rj>rir_{j}>r_{i} if ιi,j(γ)=0\iota_{*}^{i,j}(\gamma)=0 and ιi,j1(γ)0\iota_{*}^{i,j-1}(\gamma)\neq 0. If ιi,j(γ)0\iota_{*}^{i,j}(\gamma)\neq 0 for all j>ij>i, then γ\gamma never dies. For every generator, the PD contains the point (ri,rj)(r_{i},r_{j}) if the generator is born at rir_{i} and dies at rjr_{j}, or else contains the point (ri,)(r_{i},\infty) if the generator is born at rir_{i} and never dies.

2.3. Birth and death simplex pairs

Computing persistent homology can be reduced to computing the set of “birth” and “death” simplices for the generating homology classes. Informally, a birth simplex σb\sigma_{b} is a qq-simplex that creates a new qq-dimensional homology class when it is added to the filtration and a death simplex is a (q+1)(q+1)-simplex that destroys a qq-dimensional homology class when it is added to the filtration. For example, in Figure 2, the 1D PH has one generator. Its birth simplex is the 11-simplex (0,3)(0,3) and its death simplex is the 22-simplex (0,2,3)(0,2,3). For every pair (σb,σd)(\sigma_{b},\sigma_{d}) of (birth, death) simplices, the persistence diagram contains the point (f(σb),f(σd))(f(\sigma_{b}),f(\sigma_{d})). For every unpaired birth simplex σb\sigma_{b}, the persistence diagram contains the point (f(σb),)(f(\sigma_{b}),\infty).

In [20], Edelsbrunner and Harer presented an algorithm for computing the (birth, death) simplex pairs of a filtration f:𝒦f:\mathcal{K}\to\mathbb{R}. Let σ1,,σN\sigma_{1},\ldots,\sigma_{N} be the simplices of 𝒦\mathcal{K}, indexed such that i<ji<j if σi\sigma_{i} is a proper face of σj\sigma_{j}.

Definition 2.1.

The simplex order induced by ff is the strict partial order f\prec_{f} on 𝒦\mathcal{K} such that σifσj\sigma_{i}\prec_{f}\sigma_{j} if and only if f(σi)<f(σj)f(\sigma_{i})<f(\sigma_{j}).

If the simplex orders f1,f2\prec_{f_{1}},\prec_{f_{2}} induced by filtrations f1,f2f_{1},f_{2} (respectively) are the same, then f1(σi)<f1(σj)f_{1}(\sigma_{i})<f_{1}(\sigma_{j}) if and only if f2(σi)<f2(σj)f_{2}(\sigma_{i})<f_{2}(\sigma_{j}) and f1(σi)=f1(σj)f_{1}(\sigma_{i})=f_{1}(\sigma_{j}) if and only if f2(σi)=f2(σj)f_{2}(\sigma_{i})=f_{2}(\sigma_{j}).

The algorithm of [20] requires a compatible simplex indexing.

Definition 2.2.

A compatible simplex indexing is a function idx:𝒦{1,,N}\text{idx}:\mathcal{K}\to\{1,\ldots,N\} such that idx(σi)<idx(σj)\text{idx}(\sigma_{i})<\text{idx}(\sigma_{j}) if σifσj\sigma_{i}\prec_{f}\sigma_{j} or σi\sigma_{i} is a proper face of σj\sigma_{j}. Because a compatible simplex indexing may not be unique, we fix the simplex indexing induced by ff to be the unique function idxf:𝒦{1,,N}\text{idx}_{f}:\mathcal{K}\to\{1,\ldots,N\} such that idxf(σi)<idxf(σj)\text{idx}_{f}(\sigma_{i})<\text{idx}_{f}(\sigma_{j}) if either σifσj\sigma_{i}\prec_{f}\sigma_{j} or if f(σi)=f(σj)f(\sigma_{i})=f(\sigma_{j}) and i<ji<j.

The function idxf\text{idx}_{f} is a compatible simplex indexing because if σi\sigma_{i} is a proper face of σj\sigma_{j}, then i<ji<j and f(σi)f(σj)f(\sigma_{i})\leq f(\sigma_{j}). The sequence idxf1(1),,idxf1(N)\text{idx}_{f}^{-1}(1),\ldots,\text{idx}_{f}^{-1}(N) of simplices is ordered by the value of ff on each simplex, with ties broken by the order of the simplices in the sequence σ1,,σN\sigma_{1},\ldots,\sigma_{N}. The indexing idxf\text{idx}_{f} is defined such that if we define 𝒦j:={σ𝒦idxf(σ)j}\mathcal{K}^{\prime}_{j}:=\{\sigma\in\mathcal{K}\mid\text{idx}_{f}(\sigma)\leq j\}, then

𝒦1𝒦2𝒦N\mathcal{K}_{1}^{\prime}\subseteq\mathcal{K}_{2}^{\prime}\subseteq\cdots\subseteq\mathcal{K}^{\prime}_{N}

is a nested sequence of simplicial complexes and if ri=f(σj1)==f(σjk)r_{i}=f(\sigma_{j_{1}})=\cdots=f(\sigma_{j_{k}}), where j1<<jkj_{1}<\cdots<j_{k} and {ri}=Im(f)\{r_{i}\}=\text{Im}(f), with ri<ri+1r_{i}<r_{i+1}, then

𝒦ri=𝒦j1𝒦j2𝒦jk𝒦ri+1.\mathcal{K}_{r_{i}}=\mathcal{K}^{\prime}_{j_{1}}\subset\mathcal{K}^{\prime}_{j_{2}}\subset\cdots\subset\mathcal{K}_{j_{k}}^{\prime}\subset\mathcal{K}_{r_{i+1}}\,.

In other words, {𝒦j}\{\mathcal{K}^{\prime}_{j}\} is a refinement of {𝒦ri}\{\mathcal{K}_{r_{i}}\}.

The following lemma is a straightforward corollary of the work in [19], and we will rely on it repeatedly in the present paper.

Lemma 2.3 ([19]).

If f0,f1:𝒦f_{0},f_{1}:\mathcal{K}\to\mathbb{R} are two filtration functions such that f0\prec_{f_{0}} is the same as f1\prec_{f_{1}}, then idxf0=idxf1\text{idx}_{f_{0}}=\text{idx}_{f_{1}} and f1f_{1} and f2f_{2} both induce the same set of (birth, death) simplex pairs.

2.4. Updating persistent homology when the simplex indexing is updated

One of the main contributions of [14], in which vineyards were introduced, is an algorithm for updating the (birth, death) simplex pairs when the simplex indexing changes. We review the relevant details in this subsection.

Suppose that idxf0,idxf1:𝒦{1,,N}\text{idx}_{f_{0}},\text{idx}_{f_{1}}:\mathcal{K}\to\{1,\ldots,N\} are the simplex indexings that are induced by filtrations f0,f1:𝒦f_{0},f_{1}:\mathcal{K}\to\mathbb{R} (respectively), and suppose that idxf0\text{idx}_{f_{0}} and idxf1\text{idx}_{f_{1}} differ only by a transposition of a pair (σ,τ)(\sigma,\tau) of consecutive simplices. That is, idxf0(τ)=idxf0(σ)+1\text{idx}_{f_{0}}(\tau)=\text{idx}_{f_{0}}(\sigma)+1, idxf1(τ)=idxf0(σ)\text{idx}_{f_{1}}(\tau)=\text{idx}_{f_{0}}(\sigma), and idxf1(σ)=idxf0(τ)\text{idx}_{f_{1}}(\sigma)=\text{idx}_{f_{0}}(\tau). Let Sidxf0S_{\text{idx}_{f_{0}}} and Sidxf1S_{\text{idx}_{f_{1}}} be the sets of (birth, death) simplex pairs for f0f_{0} and f1f_{1}, respectively.111Recall that, by Lemma 2.3, the pairs depend only on the simplex orders idxf0\text{idx}_{f_{0}}, idxf1\text{idx}_{f_{1}}, which is why we label the sets by their associated simplex indexings. The update rule of [14] gives us bijection ϕidxf0,idxf1:Sidxf0Sidxf1\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}:S_{\text{idx}_{f_{0}}}\to S_{\text{idx}_{f_{1}}}.

We review the key properties of the bijection ϕidxf0,idxf1\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}. We write

ϕidxf0,idxf1=(ϕbidxf0,idxf1,ϕdidxf0,idxf1),\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}=(\phi_{b}^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}},\phi_{d}^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}})\,,

where ϕbidxf0,idxf1:Sidxf0𝒦\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}_{b}:S_{\text{idx}_{f_{0}}}\to\mathcal{K} maps a simplex pair (σb,σd)Sidxf0(\sigma_{b},\sigma_{d})\in S_{\text{idx}_{f_{0}}} to the birth simplex of ϕidxf0,idxf1((σb,σd))\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}((\sigma_{b},\sigma_{d})) and ϕdidxf0,idxf1:Sidxf0𝒦\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}_{d}:S_{\text{idx}_{f_{0}}}\to\mathcal{K} maps (σb,σd)Sidxf0(\sigma_{b},\sigma_{d})\in S_{\text{idx}_{f_{0}}} to the death simplex of ϕidxf0,idxf1((σb,σd))\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}((\sigma_{b},\sigma_{d})). If (σb,σd)Sidxf0(\sigma_{b},\sigma_{d})\in S_{\text{idx}_{f_{0}}} is a pair such that σb,σd{σ,τ}\sigma_{b},\sigma_{d}\not\in\{\sigma,\tau\}, then ϕidxf0,idxf1((σb,σd))=(σb,σd)\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}((\sigma_{b},\sigma_{d}))=(\sigma_{b},\sigma_{d}). If (σb1,σd1)Sidxf0(\sigma_{b}^{1},\sigma_{d}^{1})\in S_{\text{idx}_{f_{0}}} is the pair that contains σ\sigma, then let λ{b,d}\lambda\in\{b,d\} be the index such that σλ1=σ\sigma^{1}_{\lambda}=\sigma. Similarly, if (σb2,σd2)Sidxf0(\sigma_{b}^{2},\sigma_{d}^{2})\in S_{\text{idx}_{f_{0}}} is the pair that contains τ\tau, then let μ{b,d}\mu\in\{b,d\} be the index such that σμ1=τ\sigma^{1}_{\mu}=\tau. The key fact about the update rule of [14] is that ϕidxf0,idxf1\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}} is defined such that either

ϕidxf0,idxf1((σb1,σd2))\displaystyle\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}((\sigma^{1}_{b},\sigma^{2}_{d})) =(σb1,σd1),\displaystyle=(\sigma^{1}_{b},\sigma^{1}_{d})\,,
ϕidxf0,idxf1((σb2,σd2))\displaystyle\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}((\sigma^{2}_{b},\sigma^{2}_{d})) =(σb2,σd2),\displaystyle=(\sigma^{2}_{b},\sigma^{2}_{d})\,,

or

ϕλidxf0,idxf1((σb1,σd1))\displaystyle\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}_{\lambda}((\sigma^{1}_{b},\sigma^{1}_{d})) =τ,ϕλcidxf0,idxf1((σb1,σd1))=σλc1,\displaystyle=\tau\,,\qquad\phi^{\text{idx}_{f_{0}},\text{idx}_{f_{1}}}_{\lambda^{c}}((\sigma^{1}_{b},\sigma^{1}_{d}))=\sigma^{1}_{\lambda^{c}}\,,
ϕμidxf0,idxf1((σb2,σd2))\displaystyle\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}_{\mu}((\sigma^{2}_{b},\sigma^{2}_{d})) =σ,ϕμcidxf0,idxf1((σb2,σd2))=σμc1,\displaystyle=\sigma\,,\qquad\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}_{\mu^{c}}((\sigma^{2}_{b},\sigma^{2}_{d}))=\sigma^{1}_{\mu^{c}}\,,

where

λc:={b,λ=dd,λ=bμc:={b,μ=dd,μ=b.\lambda^{c}:=\begin{cases}b\,,&\lambda=d\\ d\,,&\lambda=b\end{cases}\qquad\mu^{c}:=\begin{cases}b\,,&\mu=d\\ d\,,&\mu=b\,.\end{cases}

In other words, either ϕidxf0,idxf1\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}} is the identity map or ϕidxf0,idxf1\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}} swaps σ\sigma and τ\tau in the pairs that contain them. The particular case depends on the order of f0(σb1),f0(σd1),f0(σb2),f0(σd2)f_{0}(\sigma_{b}^{1}),f_{0}(\sigma_{d}^{1}),f_{0}(\sigma_{b}^{2}),f_{0}(\sigma_{d}^{2}) (see [14] for details; they are not relevant to the present paper).

More generally, suppose that idxf0\text{idx}_{f_{0}}, idxf1\text{idx}_{f_{1}} are the simplex indexings induced by any two filtrations f0,f1f_{0},f_{1}, where idxf0\text{idx}_{f_{0}} and idxf1\text{idx}_{f_{1}} are no longer required to differ only by the transposition of two consecutive simplices. Let Sidxf0S_{\text{idx}_{f_{0}}} and Sidxf1S_{\text{idx}_{f_{1}}} be the sets of (birth, death) simplex pairs for f0f_{0} and f1f_{1}, respectively. The update rule of Cohen-Steiner et al. [14] defines a bijection ϕidxf0,idxf1:Sidxf0Sidxf1\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}:S_{\text{idx}_{f_{0}}}\to S_{\text{idx}_{f_{1}}} as follows. Every permutation can be decomposed into a sequence of transpositions that transpose consecutive elements, so there is a sequence ζ0,,ζm\zeta_{0},\ldots,\zeta_{m} of simplex indexings such that ζ0=idxf0\zeta_{0}=\text{idx}_{f_{0}}, ζm=idxf1\zeta_{m}=\text{idx}_{f_{1}}, and ζi\zeta_{i}, ζi+1\zeta_{i+1} differ only by the transposition of two consecutive simplices. Cohen-Steiner et al. [14] defined

(4) ϕidxf0,idxf1:=ϕζm1,ζmϕζ0,ζ1.\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}}:=\phi^{\zeta_{m-1},\,\zeta_{m}}\circ\cdots\circ\phi^{\zeta_{0},\,\zeta_{1}}\,.
Remark 2.4.

If idxf0\text{idx}_{f_{0}}, idxf1\text{idx}_{f_{1}} do not differ by only the transposition of two consecutive simplices, then the sequence ζ0,,ζm\zeta_{0},\ldots,\zeta_{m} is not unique. Unfortunately, the definition of ϕidxf0,idxf1\phi^{\text{idx}_{f_{0}},\,\text{idx}_{f_{1}}} does depend on the sequence ζ0,,ζm\zeta_{0},\ldots,\zeta_{m} in its definition. This is implicitly shown in Proposition 5.3.

2.5. Cellular Sheaves

A cell complex is a topological space YY with a partition into a set {Yα}αPY\{Y_{\alpha}\}_{\alpha\in P_{Y}} of subspaces (the cells of the cell complex) that satisfy the following conditions:

  1. (1)

    Every cell YαY_{\alpha} is homeomorphic to kα\mathbb{R}^{k_{\alpha}} for some kα0k_{\alpha}\geq 0. The cell YαY_{\alpha} is a kαk_{\alpha}-cell.

  2. (2)

    For every cell YαY_{\alpha}, there is a homeomorphism ϕα:BkαXα¯\phi_{\alpha}:B^{k_{\alpha}}\to\overline{X_{\alpha}}, where BkαB^{k_{\alpha}} is the closed kαk_{\alpha}-dimensional ball, such that ϕα(int(Bkα))=Xα\phi_{\alpha}(\text{int}(B^{k_{\alpha}}))=X_{\alpha}.

  3. (3)

    Axiom of the Frontier: If the intersection Yβ¯Yα\overline{Y_{\beta}}\cap Y_{\alpha} is nonempty, then YαYβ¯Y_{\alpha}\subseteq\overline{Y_{\beta}}. We say that YαY_{\alpha} is a face of YβY_{\beta}.

  4. (4)

    Locally finite: Every xXx\in X has an open neighborhood UU such that UU intersects finitely many cells.

For example, a polyhedron is a cell complex whose kk-cells are the kk-dimensional faces of the polyhedron. A graph is a another example of a cell complex; the 0-cells are the vertices and the 11-cells are the edges.

We will first review the most general definition of cellular sheaves, which uses category theory, and then we will specialize to the case of interest for the present paper, which does not require category theory. Let YY be a cell complex with cells {Yα}αPY\{Y_{\alpha}\}_{\alpha\in P_{Y}}, and let 𝒟\mathcal{D} be a category. The set PYP_{Y} is a poset with the relation αβ\alpha\leq\beta if YαYβ¯Y_{\alpha}\subseteq\overline{Y_{\beta}}.

A 𝒟\mathcal{D}-valued cellular sheaf on YY consists of

  1. (1)

    An assignment of an object (Yα)𝒟\mathcal{F}(Y_{\alpha})\in\mathcal{D} (the stalk of \mathcal{F} at YαY_{\alpha}) for every cell YαY_{\alpha} in YY, and

  2. (2)

    A morphism αβ:F(Yα)F(Yβ)\mathcal{F}_{\alpha\leq\beta}:F(Y_{\alpha})\to F(Y_{\beta}) (a restriction map) whenever YαY_{\alpha} is a face of YβY_{\beta}. The morphisms must satisfy the composition condition:

    (5) βγαβ=αγ\mathcal{F}_{\beta\leq\gamma}\circ\mathcal{F}_{\alpha\leq\beta}=\mathcal{F}_{\alpha\leq\gamma}

    whenever αβγ\alpha\leq\beta\leq\gamma.

Equivalently, a 𝒟\mathcal{D}-valued cellular sheaf on YY is a functor :PY𝒟\mathcal{F}:P_{Y}\to\mathcal{D}, where PYP_{Y} is considered as a category.

A global section of a cellular sheaf \mathcal{F} is a function

s:{Yα}αPYα(Yα)s:\{Y_{\alpha}\}_{\alpha\in P_{Y}}\to\bigcup_{\alpha}\mathcal{F}(Y_{\alpha})

such that

  1. (1)

    s(Yα)(Yα)s(Y_{\alpha})\in\mathcal{F}(Y_{\alpha}) (i.e., s(Yα)s(Y_{\alpha}) is a choice of element in the stalk at YαY_{\alpha}), and

  2. (2)

    if αβ\alpha\leq\beta, then s(Yβ)=αβ(s(Yα))s(Y_{\beta})=\mathcal{F}_{\alpha\leq\beta}(s(Y_{\alpha})) .

In what follows, we will consider a Set-valued cellular sheaf. The objects of the category Set are sets, and the morphisms between sets AA and BB are the functions from AA to BB. A Set-valued cellular sheaf on a cell complex YY consists of

  1. (1)

    A set (Yα)\mathcal{F}(Y_{\alpha}) for every cell YαY_{\alpha} in YY, and

  2. (2)

    A function αβ:(Yα)(Yβ)\mathcal{F}_{\alpha\leq\beta}:\mathcal{F}(Y_{\alpha})\to\mathcal{F}(Y_{\beta}) whenever YαY_{\alpha} is a face of YβY_{\beta}. The functions must satisfy the condition:

    βγαβ=αγ\mathcal{F}_{\beta\leq\gamma}\circ\mathcal{F}_{\alpha\leq\beta}=\mathcal{F}_{\alpha\leq\gamma}

    whenever αβγ\alpha\leq\beta\leq\gamma.

3. Definition of a Persistence Diagram Bundle

A vineyard is a 11-parameter set of persistence diagrams that is computed from a 11-parameter set of filtration functions on a simplicial complex 𝒦\mathcal{K}. We generalize a vineyard to a “persistence diagram bundle” as follows.

Definition 3.1 (Fibered filtration function).

A fibered filtration function is a set {fp:𝒦p}pB\{f_{p}:\mathcal{K}^{p}\to\mathbb{R}\}_{p\in B}, where BB is a topological space, {𝒦p}pB\{\mathcal{K}^{p}\}_{p\in B} is a set of simplicial complexes parameterized by BB, and fpf_{p} is a filtration function on 𝒦p\mathcal{K}^{p}.

When 𝒦p𝒦\mathcal{K}^{p}\equiv\mathcal{K} for all pBp\in B, we define f:𝒦×Bf:\mathcal{K}\times B\to\mathbb{R} to be the function f(σ,p):=fp(σ)f(\sigma,p):=f_{p}(\sigma). In a slight abuse of notation, we refer to f:𝒦×Bf:\mathcal{K}\times B\to\mathbb{R}, rather than to {fp:𝒦B}pB\{f_{p}:\mathcal{K}\to B\}_{p\in B}, as the fibered filtration function. For all pBp\in B, the function f(,p):𝒦f(\cdot,p):\mathcal{K}\to\mathbb{R} is a filtration of 𝒦\mathcal{K}. For several examples with 𝒦p𝒦\mathcal{K}^{p}\equiv\mathcal{K}, see Section 3.1.

Definition 3.2 (Persistence diagram bundle).

Let {fp:𝒦p}pB\{f_{p}:\mathcal{K}^{p}\to\mathbb{R}\}_{p\in B} be a fibered filtration function. The base of the bundle is BB. The qqth total space of the bundle is E:={(p,z)}pB,zPDq(fp)}E:=\{(p,z)\}\mid p\in B,z\in PD_{q}(f_{p})\}, with the subspace topology inherited from the inclusion EBׯ2E\hookrightarrow B\times\overline{\mathbb{R}}^{2}.222Technically, EE is a multiset because persistence diagrams are multisets. However, when considering EE as a topological space (which we do in Section 5 to study continuous paths in EE and sections of the PD bundle), we consider EE as a set. The qqth persistence diagram bundle is the triple (E,B,π)(E,B,\pi), where π:EB\pi:E\to B is the projection (p,z)p(p,z)\mapsto p.

For example, when BB is an interval in \mathbb{R} and 𝒦p𝒦\mathcal{K}^{p}\equiv\mathcal{K}, Definition 3.2 reduces to that of a vineyard: a 11-parameter set of PDs for a 11-parameter set of filtrations of 𝒦\mathcal{K}. As discussed in Section 1, PHT is a special case with B=SdB=S^{d}. The fibered barcode of a multiparameter persistence module is another special case; we will discuss it in Section 3.2.2.

Remark 3.3.

In Definition 3.2, we are suggestively using the language of fiber bundles. However, it is important to note that a PD bundle is not guaranteed to be a true fiber bundle. The fibers need not be homeomorphic to each other for all pBp\in B. At “singularities” (points pBp_{*}\in B at which PD(fp)\text{PD}(f_{p_{*}}) has an off-diagonal point with multiplicity), points in PD(fp)\text{PD}(f_{p}) for nearby pp may merge into each other, changing the homotopy type of the fiber. However, if f:𝒦×Bf:\mathcal{K}\times B\to\mathbb{R} is continuous and pBp\in B is not a singularity, then there is a neighborhood UBU\subseteq B and a homeomorphism ϕ:π1(U)U×PD(fp)\phi:\pi^{-1}(U)\to U\times\text{PD}(f_{p_{*}}) that preserves fibers (i.e., a local trivialization).

As a special case of fibered filtration functions, we define piecewise-linear fibered filtration functions, which are simpler to analyze.

Definition 3.4 (Piecewise-linear fibered filtration function).

Let {fp:𝒦p}pB\{f_{p}:\mathcal{K}^{p}\to\mathbb{R}\}_{p\in B} be a fibered filtration function such that 𝒦p𝒦\mathcal{K}^{p}\equiv\mathcal{K}. We define f(σ,p):=fp(σ)f(\sigma,p):=f_{p}(\sigma) for all σ𝒦\sigma\in\mathcal{K} and pBp\in B. If BB is a simplicial complex and f(σ,)f(\sigma,\cdot) is linear on each simplex of BB for all simplices σ𝒦\sigma\in\mathcal{K}, then ff is a piecewise-linear fibered filtration function. The resulting PD bundle is a piecewise-linear PD bundle.

For instance, the fibered filtration function in Example 3.6, below, is piecewise linear.

3.1. Examples

The following are concrete examples of PD bundles. We begin with the example that motivated PD bundles in Section 1.

Example 3.5.

Suppose that X(t,𝝁)={x1(t,𝝁),,xk(t,𝝁)}X(t,\bm{\mu})=\{x_{1}(t,\bm{\mu}),\ldots,x_{k}(t,\bm{\mu})\} is a point cloud that varies continuously with time tt\in\mathbb{R} and system-parameter values μ1,,μm\mu_{1},\ldots,\mu_{m}\in\mathbb{R}. We obtain a fibered filtration function f:𝒦×m+1f:\mathcal{K}\times\mathbb{R}^{m+1}\to\mathbb{R} by defining f(,(t,𝝁)):𝒦f(\cdot,(t,\bm{\mu})):\mathcal{K}\to\mathbb{R} to be the Vietoris–Rips filtration function for the point cloud X(t,𝝁)X(t,\bm{\mu}) at all (t,𝝁)m+1(t,\bm{\mu})\in\mathbb{R}^{m+1} (or any other filtration for the point cloud at each (t,𝝁)(t,\bm{\mu})). The simplicial complex 𝒦\mathcal{K} is the simplicial complex that has a simplex for every subset of points in the point cloud.

Example 3.6.

Consider a color image. Enumerate the pixels and let r(i)r(i), g(i)g(i), and b(i)b(i) denote the red, green, and blue values of the iith pixel. Triangulate each pixel to obtain a simplicial complex 𝒦\mathcal{K}. (Every pixel is split into two triangles.) Let B={(w1,w2)[0,1]20w1+w21}B=\{(w_{1},w_{2})\in[0,1]^{2}\mid 0\leq w_{1}+w_{2}\leq 1\}. For all (w1,w2)B(w_{1},w_{2})\in B, define p(i,(w1,w2))=w1r(i)+w2g(i)+(1w1w2)b(i)p(i,(w_{1},w_{2}))=w_{1}r(i)+w_{2}g(i)+(1-w_{1}-w_{2})b(i). The function p(i,(w1,w2))p(i,(w_{1},w_{2})) is a weighted average of the red, green, and blue values of the iith pixel. Define a piecewise-linear fibered filtration function f:𝒦×Bf:\mathcal{K}\times B\to\mathbb{R} as follows. For a 22-simplex σ\sigma, define f(σ,𝐰)=p(i(σ),𝐰)f(\sigma,{\bf w})=p(i(\sigma),{\bf w}), where i(σ)i(\sigma) is the pixel containing σ\sigma. For any other simplex σ𝒦\sigma\in\mathcal{K}, define f(σ,𝐰)=min{f(τ,𝐰)στ,dim(τ)=2}f(\sigma,{\bf w})=\min\{f(\tau,{\bf w})\mid\sigma\subseteq\tau,\,\dim(\tau)=2\}. At 𝐰=(1,0){\bf w}=(1,0), 𝐰=(0,1){\bf w}=(0,1), and 𝐰=(1,1){\bf w}=(1,1), the filtration function f(,𝐰)f(\cdot,{\bf w}) is the sublevel filtration by red, green, and blue pixel values, respectively. At all other 𝐰B{\bf w}\in B, the filtration function f(,𝐰)f(\cdot,{\bf w}) is the sublevel filtration by a weighted average of the red, green, and blue pixel values.

Example 3.7.

Let μ1,,μm\mu_{1},\ldots,\mu_{m}\in\mathbb{R} denote the system-parameter values of some discrete dynamical system. For given system-parameter values 𝝁m\bm{\mu}\in\mathbb{R}^{m}, let xi𝝁nx_{i}^{\bm{\mu}}\in\mathbb{R}^{n} be the solution at the iith time step and let X(𝝁)={x0𝝁,,xk𝝁}X(\bm{\mu})=\{x_{0}^{\bm{\mu}},\ldots,x_{k}^{\bm{\mu}}\} be the set of points obtained after the first kk time steps. For example, persistent homology has been used to study orbits of the linked twist map (a discrete dynamical system) [1]. We obtain a fibered filtration function f:𝒦×mf:\mathcal{K}\times\mathbb{R}^{m}\to\mathbb{R} by defining f(,𝝁):𝒦f(\cdot,\bm{\mu}):\mathcal{K}\to\mathbb{R} to be the Vietoris–Rips filtration function for the point cloud X(𝝁)X(\bm{\mu}) (or any other filtration for the point cloud at each 𝝁\bm{\mu}). The simplicial complex 𝒦\mathcal{K} has a simplex for every subset of points in the point cloud.

Example 3.8.

Suppose that X(t)={x1(t),,xk(t)}X(t)=\{x_{1}(t),\ldots,x_{k}(t)\} is a time-varying point cloud in a compact triangulable subset SnS\subseteq\mathbb{R}^{n}. Let ρh(,t)\rho_{h}(\cdot,t) be a kernel density estimator at time tt, with bandwidth parameter h>0h>0. For fixed hh and tt, we define a filtered complex by considering sublevel sets of ρh\rho_{h} as follows. Let 𝒦\mathcal{K} be a triangulation of SnS\subseteq\mathbb{R}^{n}. A vertex vv of 𝒦\mathcal{K} is included in the simplicial complex 𝒦r(t,h)\mathcal{K}_{r}^{(t,h)} if ρh(v,t)r\rho_{h}(v,t)\leq r, and a simplex of 𝒦\mathcal{K} is included in 𝒦rt,h\mathcal{K}_{r}^{t,h} if all of its vertices are in 𝒦r(t,h)\mathcal{K}_{r}^{(t,h)}. For each tt and hh, the set {𝒦r(t,h)}r\{\mathcal{K}^{(t,h)}_{r}\}_{r\in\mathbb{R}} is a filtered complex. We obtain a fibered filtration function f:𝒦×+2f:\mathcal{K}\times\mathbb{R}_{+}^{2}\to\mathbb{R} by defining f(,(t,h))f(\cdot,(t,h)) to be the filtration function associated with the filtered complex {𝒦r(t,h)}r\{\mathcal{K}_{r}^{(t,h)}\}_{r\in\mathbb{R}}.

Density sublevels of time-varying point clouds were also considered by Corcoran et al. [15], who studied a school of fish swimming in a shallow pool that was modeled as a subset of 2\mathbb{R}^{2}. However, Corcoran et al. [15] fixed a bandwidth parameter hh and a sublevel rr, and only studied how the PH changed with time (by using zigzag PH).

3.2. Comparison to multiparameter PH

Multiparameter PH was introduced in [8]; see [7] for a review. Typically, a fibered filtration function does not induce a multifiltration, but the fibered barcode of a multiparameter peristence module is an example of a PD bundle.

3.2.1. Multifiltrations

We review the definition of a multifiltration and compare it to the definition of a fibered filtration.

Definition 3.9.

A multifiltration is a set {𝒦𝒖}𝒖n\{\mathcal{K}_{\bm{u}}\}_{\bm{u}\in\mathbb{Z}^{n}} of simplicial complexes such that if 𝒖𝒗\bm{u}\leq\bm{v}, then 𝒦𝒖𝒦𝒗\mathcal{K}_{\bm{u}}\subseteq\mathcal{K}_{\bm{v}}.

The inclusion ι𝒖,𝒗:𝒦𝒖𝒦𝒗\iota^{\bm{u},\bm{v}}:\mathcal{K}_{\bm{u}}\hookrightarrow\mathcal{K}_{\bm{v}} induces a map ι𝒖,𝒗:Hq(𝒦𝒖,𝔽)Hq(𝒦𝒗,𝔽)\iota_{*}^{\bm{u},\bm{v}}:H_{q}(\mathcal{K}_{\bm{u}},\mathbb{F})\to H_{q}(\mathcal{K}_{\bm{v}},\mathbb{F}) from the qqth homology of 𝒦𝒖\mathcal{K}_{\bm{u}} to the qqth homology of 𝒦𝒗\mathcal{K}_{\bm{v}} over a field 𝔽\mathbb{F}. Given a multifiltration {𝒦𝒖}𝒖n\{\mathcal{K}_{\bm{u}}\}_{\bm{u}\in\mathbb{Z}^{n}}, the multiparameter persistence module is the graded 𝔽[x1,,xn]\mathbb{F}[x_{1},\ldots,x_{n}]-module 𝒖nHq(𝒦𝒖,𝔽)\bigoplus_{\bm{u}\in\mathbb{Z}^{n}}H_{q}(\mathcal{K}_{\bm{u}},\mathbb{F}). The action of xix_{i} on a homogeneous element γHq(𝒦𝒖,𝔽)\gamma\in H_{q}(\mathcal{K}_{\bm{u}},\mathbb{F}) is given by xiγ=ι𝒖,𝒗(γ)x_{i}\gamma=\iota_{\bm{u},\bm{v}}^{*}(\gamma), where vj=uj+δijv_{j}=u_{j}+\delta_{ij}.

Remark 3.10.

Some researchers define multifiltrations more generally as functors :𝒫𝐒𝐢𝐦𝐩\mathcal{F}:\mathcal{P}\to\bf{Simp}, where 𝒫\mathcal{P} is any poset and 𝐒𝐢𝐦𝐩\bf{Simp} is the category of simplicial complexes, with simplicial maps as morphisms. Definition 3.9 is the specific case in which 𝒫=n\mathcal{P}=\mathbb{Z}^{n} and 𝒖𝒗:𝒦𝒖𝒦𝒗\mathcal{F}_{\bm{u}\leq\bm{v}}:\mathcal{K}_{\bm{u}}\to\mathcal{K}_{\bm{v}} is an inclusion map.

To see why a fibered filtration function does not typically induce a multifiltration, consider a fibered filtration function {fp:𝒦p}pB\{f_{p}:\mathcal{K}^{p}\to\mathbb{R}\}_{p\in B} with B=nB=\mathbb{R}^{n}. Let 𝒦rp:={σ𝒦pfp(σ)r}\mathcal{K}_{r}^{p}:=\{\sigma\in\mathcal{K}^{p}\mid f_{p}(\sigma)\leq r\} denote the rr-sublevel set of fpf_{p}. It is not necessarily the case that 𝒦sp1𝒦rp2\mathcal{K}_{s}^{p_{1}}\subseteq\mathcal{K}_{r}^{p_{2}} whenever rsr\leq s and p1p2p_{1}\leq p_{2}. Moreover, there are no canonical simplicial maps 𝒦sp1𝒦rp2\mathcal{K}_{s}^{p_{1}}\to\mathcal{K}_{r}^{p_{2}}, so it is not guaranteed that {𝒦rp}(p,r)n×\{\mathcal{K}_{r}^{p}\}_{(p,r)\in\mathbb{R}^{n}\times\mathbb{R}} is a multifiltration even in the general sense of Remark 3.10. Therefore, such a set of filtered complexes cannot be analyzed using multiparameter persistent homology.

3.2.2. Fibered barcodes

Consider a multifiltration {𝒦𝒖}𝒖n\{\mathcal{K}_{\bm{u}}\}_{\bm{u}\in\mathbb{R}^{n}}. Let ¯\overline{\mathcal{L}} denote the space of lines in n\mathbb{R}^{n} with a parameterization of the form

L:n,\displaystyle L:\mathbb{R}\to\mathbb{R}^{n}\,,
L(r)=r𝒗+𝒃,𝒗[0,)n,𝒗=1,𝒃n.\displaystyle L(r)=r\bm{v}+\bm{b}\,,\qquad\bm{v}\in[0,\infty)^{n},\,\lVert\bm{v}\rVert=1,\,\bm{b}\in\mathbb{R}^{n}.

For example, when n=2n=2, the space ¯\overline{\mathcal{L}} is the space of lines in 2\mathbb{R}^{2} with non-negative slope, including vertical lines. For each line L¯L\in\overline{\mathcal{L}}, we define 𝒦rL:=𝒦L(r)\mathcal{K}^{L}_{r}:=\mathcal{K}_{L(r)}. That is, {𝒦rL}r\{\mathcal{K}^{L}_{r}\}_{r\in\mathbb{R}} is the filtered complex obtained by restricting the multifiltration {𝒦𝒖}𝒖n\{\mathcal{K}_{\bm{u}}\}_{\bm{u}\in\mathbb{R}^{n}} to the line LnL\subseteq\mathbb{R}^{n}. The set {𝒦rL}r\{\mathcal{K}^{L}_{r}\}_{r\in\mathbb{R}} is a filtered complex because L(r)iL(s)iL(r)_{i}\leq L(s)_{i} for all rsr\leq s and i{1,,n}i\in\{1,\ldots,n\}. The fibered barcode [4] is the map that sends L¯L\in\overline{\mathcal{L}} to the barcode for the persistent homology of {𝒦rL}r\{\mathcal{K}^{L}_{r}\}_{r\in\mathbb{R}}.

A fibered barcode is a PD bundle whose base space is B=¯B=\overline{\mathcal{L}}. For L¯L\in\overline{\mathcal{L}}, the filtration function is

fL:𝒦L,\displaystyle f_{L}:\mathcal{K}^{L}\to\mathbb{R}\,,
fL(σ)=min{rσ𝒦L(r)},\displaystyle f_{L}(\sigma)=\min\{r\mid\sigma\in\mathcal{K}_{L(r)}\}\,,

where 𝒦L:=r𝒦L(r)\mathcal{K}^{L}:=\bigcup_{r\in\mathbb{R}}\mathcal{K}_{L(r)}. Unlike the other examples in Section 3.1, the simplicial complex 𝒦L\mathcal{K}^{L} is not independent of L¯L\in\overline{\mathcal{L}}.

4. A Stratification of the Base Space

There are many different notions of a stratified space [35]. In the present paper, what we mean by a stratification is the following definition.

Definition 4.1.

A stratification of a topological space BB is a nested sequence

B0B1B2Bn=BB^{0}\subseteq B^{1}\subseteq B^{2}\subseteq\cdots\subseteq B^{n}=B

of closed subsets BmB^{m} such that the following hold:

  1. (1)

    For all mm, the space BmBm1B^{m}\setminus B^{m-1} is either empty or a smooth mm-dimensional submanifold of BB (where we set B1:=B^{-1}:=\emptyset). The mm-dimensional strata are the connected components of BmBm1B^{m}\setminus B^{m-1}. We denote the set of strata by 𝒴\mathcal{Y}.

  2. (2)

    The set 𝒴\mathcal{Y} of strata is locally finite: every pBp\in B has an open neighborhood UU such that UU intersects finitely many elements of 𝒴\mathcal{Y}.

  3. (3)

    The set 𝒴\mathcal{Y} of strata satisfy the Axiom of the Frontier: If YαY_{\alpha}, Yβ𝒴Y_{\beta}\in\mathcal{Y} are strata such that YβYα¯Y_{\beta}\cap\overline{Y_{\alpha}}, then YβYα¯Y_{\beta}\subseteq\overline{Y_{\alpha}}. We write that YβY_{\beta} is a face of YαY_{\alpha}.

In the present paper, BB is the base space of a fibered filtration function.

Theorem 4.15 says that for any “generic” smooth fibered filtration function (see Section 4.2), the base space BB can be stratified so that in each stratum YBY\subseteq B, the set of (birth, death) simplex pairs is constant and can be used to obtain PD(fp)\text{PD}(f_{p}) for any pYp\in Y.

4.1. Piecewise-linear fibered filtrations

As a warm-up, we first consider piecewise-linear fibered filtration functions, which will provide intuition for the general case. However, note that Proposition 4.5 below is not simply a special case of Theorem 4.15, in which we consider generic smooth fibered filtrations on smooth compact manifolds (see Section 4.2). Here, we consider all piecewise-linear fibered filtrations, rather than only generic piecewise-linear fibered filtrations.

First, we establish some notation and definitions.

Definition 4.2.

An open half-space of an affine space AA is one of the two connected components of AHA\setminus H for some hyperplane HH.

For example, an open half-space of n\mathbb{R}^{n} is a set of the form {𝒙nA𝒙>𝒃}\{\bm{x}\in\mathbb{R}^{n}\mid A\bm{x}>\bm{b}\} for some n×nn\times n matrix AA and some vector 𝒃n\bm{b}\in\mathbb{R}^{n}.

Definition 4.3.

An open polyhedron is the intersection of open half-spaces.

For example, an open polygon PP (a polygon without its faces) in 2\mathbb{R}^{2} is an open 2D polyhedron because it is the intersection of half-spaces of 2\mathbb{R}^{2}. The 1D faces (i.e., edges) of PP are 1D polyhedra because an edge is a subset of a line L2L\subseteq\mathbb{R}^{2} and the edge is the intersection of two half-spaces of LL. The 0D faces (i.e., vertices) of PP are 0D polyhedra.

We fix a simplicial complex 𝒦\mathcal{K} for the remainder of this section. For each pair (σ,τ)(\sigma,\tau) of simplices in 𝒦\mathcal{K}, we define

(6) I(σ,τ):={pBf(σ,p)=f(τ,p)}.I(\sigma,\tau):=\{p\in B\mid f(\sigma,p)=f(\tau,p)\}\,.
Lemma 4.4.

Suppose that f:𝒦×Bf:\mathcal{K}\times B\to\mathbb{R} is a continuous fibered filtration function (i.e., f(σ,):Bf(\sigma,\cdot):B\to\mathbb{R} is continuous for all simplices σ𝒦\sigma\in\mathcal{K}) and that YY is a path-connected subset of BB. If all pairs (σ,τ)(\sigma,\tau) of simplices satisfy either I(σ,τ)Y=I(\sigma,\tau)\cap Y=\emptyset or I(σ,τ)Y=YI(\sigma,\tau)\cap Y=Y, then the simplex order is constant in YY. That is, there is a strict partial order Y\prec_{Y} on 𝒦\mathcal{K} such that for all yYy\in Y, we have that f(,y)\prec_{f(\cdot,y)} is the same as Y\prec_{Y}.

Proof.

Let (σ,τ)(\sigma,\tau) be a pair of simplices. If I(σ,τ)Y=YI(\sigma,\tau)\cap Y=Y, then f(σ,p)=f(τ,p)f(\sigma,p)=f(\tau,p) for all pYp\in Y, so σf(,p)τ\sigma\not\prec_{f(\cdot,p)}\tau and τf(,p)σ\tau\not\prec_{f(\cdot,p)}\sigma for all pYp\in Y. If I(σ,τ)Y=I(\sigma,\tau)\cap Y=\emptyset, then f(σ,p)f(τ,p)f(\sigma,p)\neq f(\tau,p) for all pYp\in Y. Let p0p_{0} be a point in YY. Without loss of generality, τf(,p0)σ\tau\prec_{f(\cdot,p_{0})}\sigma. Therefore, f(σ,p0)>f(τ,p0)f(\sigma,p_{0})>f(\tau,p_{0}). To obtain a contradiction, suppose that f(σ,p1)<f(τ,p1)f(\sigma,p_{1})<f(\tau,p_{1}) for some p1Yp_{1}\in Y. Let γ:[0,1]Y\gamma:[0,1]\to Y be a continuous path from p0p_{0} to p1p_{1}, and let g(s)=f(σ,γ(s))f(τ,γ(s))g(s)=f(\sigma,\gamma(s))-f(\tau,\gamma(s)) for s[0,1]s\in[0,1]. By the Intermediate Value Theorem, there is an s[0,1]s_{*}\in[0,1] such that f(σ,γ(s))=f(τ,γ(s))f(\sigma,\gamma(s_{*}))=f(\tau,\gamma(s_{*})), but this is a contradiction. Therefore, f(σ,p)>f(τ,p)f(\sigma,p)>f(\tau,p) for all pYp\in Y, which implies that τf(,p)σ\tau\prec_{f(\cdot,p)}\sigma for all pYp\in Y. ∎

Proposition 4.5.

Let BB be a simplicial complex. If f:𝒦×Bf:\mathcal{K}\times B\to\mathbb{R} is a piecewise-linear fibered filtration function, then BB can be partitioned into disjoint polyhedra PP on which the simplex order induced by ff is constant. That is, there is a strict partial order P\prec_{P} on 𝒦\mathcal{K} such that f(,p)\prec_{f(\cdot,p)} is the same as P\prec_{P} for all pPp\in P. Consequently, the set {(σb,σd)}\{(\sigma_{b},\sigma_{d})\} of (birth, death) simplex pairs for ff is constant in each PP and for any pPp\in P, the persistence diagram PD(fp)\text{PD}(f_{p}) consists of the diagonal and the multiset {(f(σb),f(σd)}\{(f(\sigma_{b}),f(\sigma_{d})\}.

Proof.

Let Δ\Delta be an nn-dimensional simplex of the simplicial complex BB and let σ\sigma and τ\tau be distinct simplices of 𝒦\mathcal{K}. Because f(σ,)|Δf(\sigma,\cdot)|_{\Delta} and f(τ,)|Δf(\tau,\cdot)|_{\Delta} are linear, the set I(σ,τ)ΔI(\sigma,\tau)\cap\Delta is one of the following:

  1. (1)

    the intersection of an (n1)(n-1)-dimensional hyperplane with Δ\Delta ;

  2. (2)

    \emptyset ;

  3. (3)

    Δ\Delta ;

  4. (4)

    a vertex of Δ\Delta\,.

Therefore, the set Δ{I(σ,τ)Δ(I(σ,τ)Δ)Δ}σ,τ𝒦\partial\Delta\bigcup\{I(\sigma,\tau)\cap\Delta\mid\emptyset\subset(I(\sigma,\tau)\cap\Delta)\subset\Delta\}_{\sigma,\tau\in\mathcal{K}} partitions Δ\Delta into polyhedra. By Lemma 4.4, the simplex order induced by ff is constant on each polyhedron. The last statement of Proposition 4.5 follows from Lemma 2.3. ∎

For example, if BB is a triangulated surface, then the set

(7) L:=ΔBΔ{I(σ,τ)Δ(I(σ,τ)Δ)Δ}σ,τ𝒦L:=\bigcup_{\Delta\in B}\partial\Delta\cup\{I(\sigma,\tau)\cap\Delta\mid\emptyset\subset(I(\sigma,\tau)\cap\Delta)\subset\Delta\}_{\sigma,\tau\in\mathcal{K}}

partitions Δ\Delta into polyhedra such that the simplex order is constant on each polyhedron, including the 1D polyhedra (i.e., edges) and the 0D polyhedra (i.e., vertices). The polygonal subdivision induced by LL is called a line arrangement 𝒜(L)\mathcal{A}(L). For an example of such a line arrangement, see Figure 3.

Refer to caption
Figure 3. A line arrangement that represents the partition of a triangulated surface BB (the base space) into polyhedra on which the simplex order is constant.

4.2. Generic smooth fibered filtrations

We now consider generic smooth fibered filtration functions. Throughout Section 4.2, we consider a smooth fibered filtration function of the form f:𝒦×Bf:\mathcal{K}\times B\to\mathbb{R} for some nn-dimensional smooth compact manifold BB and some simplicial complex 𝒦\mathcal{K}. (A fibered filtration ff is smooth if f(σ,):Bf(\sigma,\cdot):B\to\mathbb{R} is smooth for all σ𝒦\sigma\in\mathcal{K}.) To make precise the notion of a “generic” fibered filtration function, we consider perturbations of ff of a certain form. Because the filtration value of a simplex σ\sigma must be at least as large as the filtration value of any face τ\tau at all pBp\in B, we consider only perturbations f𝒂:𝒦×Bf_{\bm{a}}:\mathcal{K}\times B\to\mathbb{R} of the form

f𝒂(σi,p):=f(σi,p)+ai,f_{\bm{a}}(\sigma_{i},p):=f(\sigma_{i},p)+a_{i}\,,

where 𝒂\bm{a} is an element of the set

(8) A:={𝒂Naiaj for all ij}A:=\{\bm{a}\in\mathbb{R}^{N}\mid a_{i}\leq a_{j}\text{ for all }i\leq j\}\

and σ1,,σN\sigma_{1},\ldots,\sigma_{N} are the simplices of 𝒦\mathcal{K}, indexed such that i<ji<j if σi\sigma_{i} is a proper face of σj\sigma_{j}. By construction, f𝒂f_{\bm{a}} is a fibered filtration function for all 𝒂A\bm{a}\in A.

For each simplex σk\sigma_{k} in 𝒦\mathcal{K}, we define the manifold

Mk:={(p,f(σk,p))pB}B×M_{k}:=\{(p,f(\sigma_{k},p))\mid p\in B\}\subseteq B\times\mathbb{R}\

and for each 𝒂A\bm{a}\in A, we define the manifold

(9) M𝒂,k:={(p,f𝒂(σk,p))pB}B×.M_{\bm{a},k}:=\{(p,f_{\bm{a}}(\sigma_{k},p))\mid p\in B\}\subseteq B\times\mathbb{R}\,.

For each pair (σi,σj)(\sigma_{i},\sigma_{j}) of simplices in 𝒦\mathcal{K}, we define I(σi,σj)I(\sigma_{i},\sigma_{j}) as in (6). The set I(σi,σj)I(\sigma_{i},\sigma_{j}) is the projection of MiMjB×M_{i}\cap M_{j}\subseteq B\times\mathbb{R} to a subset of BB. For each 𝒂A\bm{a}\in A, we define the set

I𝒂(σi,σj):={pBf𝒂(σi,p)=f𝒂(σj,p)}.I_{\bm{a}}(\sigma_{i},\sigma_{j}):=\{p\in B\mid f_{\bm{a}}(\sigma_{i},p)=f_{\bm{a}}(\sigma_{j},p)\}\,.

We also define

Em\displaystyle E^{m} :={I(σi1,σj1)I(σim,σjm)},\displaystyle:=\{I(\sigma_{i_{1}},\sigma_{j_{1}})\cap\cdots\cap I(\sigma_{i_{m}},\sigma_{j_{m}})\}\,,

which is the set of all mm-way intersections of sets I(σi,σj)I(\sigma_{i},\sigma_{j}). For all 𝒂A\bm{a}\in A, we define

(10) E𝒂m:={I𝒂(σi1,σj1)I𝒂(σim,σjm)}.E^{m}_{\bm{a}}:=\{I_{\bm{a}}(\sigma_{i_{1}},\sigma_{j_{1}})\cap\cdots\cap I_{\bm{a}}(\sigma_{i_{m}},\sigma_{j_{m}})\}\,.

Lastly, we define

(11) E𝒂m,k:={I𝒂(σi1,σj1)I𝒂(σim,σjm)ir,jrk for all r},E^{m,k}_{\bm{a}}:=\{I_{\bm{a}}(\sigma_{i_{1}},\sigma_{j_{1}})\cap\cdots\cap I_{\bm{a}}(\sigma_{i_{m}},\sigma_{j_{m}})\mid i_{r},j_{r}\leq k\text{ for all }r\}\,,

which is the set of mm-way intersections that only involve the simplices σ1,,σk\sigma_{1},\ldots,\sigma_{k}.

Remark 4.6.

There are several facts to keep in mind. First, it is not guaranteed that I𝒂(σi,σj)I_{\bm{a}}(\sigma_{i},\sigma_{j}) is homeomorphic to I(σi,σj)I(\sigma_{i},\sigma_{j}) even for arbitrarily small 𝒂\bm{a}. Additionally, the sets I𝒂(σi,σj)I_{\bm{a}}(\sigma_{i},\sigma_{j}) are not “independent” of each other; a perturbation of f(σ,)f(\sigma,\cdot) for a single simplex σ\sigma causes a perturbation of I(σ,τ)I(\sigma,\tau) for all τ𝒦\tau\in\mathcal{K}. Furthermore, not every element of EmE^{m} is an (nm)(n-m)-dimensional submanifold, even generically. For example, if I(σi2,σi1)I(\sigma_{i_{2}},\sigma_{i_{1}}) and I(σi3,σi2)I(\sigma_{i_{3}},\sigma_{i_{2}}) are (n1)(n-1)-dimensional submanifolds that intersect transversely, then I(σi3,σi1)I(σi3,σi2)I(σi2,σi1)=I(σi3,σi2)I(σi2,σi1)I(\sigma_{i_{3}},\sigma_{i_{1}})\cap I(\sigma_{i_{3}},\sigma_{i_{2}})\cap I(\sigma_{i_{2}},\sigma_{i_{1}})=I(\sigma_{i_{3}},\sigma_{i_{2}})\cap I(\sigma_{i_{2}},\sigma_{i_{1}}) is an (n2)(n-2)-dimensional submanifold, rather than an (n3)(n-3)-dimensional submanifold. Finally, r=1mI(σir,σjr)\bigcap_{r=1}^{m}I(\sigma_{i_{r}},\sigma_{j_{r}}) is not necessarily equal to the projection of r=1m(MirMjr)\bigcap_{r=1}^{m}(M_{i_{r}}\cap M_{j_{r}}) to BB. In other words, not every intersection in BB lifts to an intersection of the manifolds {Mk}k=1NB×\{M_{k}\}_{k=1}^{N}\subseteq B\times\mathbb{R}. These are the main subtleties in the proof of Theorem 4.15.

Definition 4.7.

Let SS be an element of E𝒂mE^{m}_{\bm{a}}, where E𝒂mE^{m}_{\bm{a}} is defined as in (10). The set SS is mm-reduced if it equals a set of the form I𝒂(σi1,σj1)I𝒂(σim,σjm)I_{\bm{a}}(\sigma_{i_{1}},\sigma_{j_{1}})\cap\cdots\cap I_{\bm{a}}(\sigma_{i_{m}},\sigma_{j_{m}}), where i1>i2>>imi_{1}>i_{2}>\cdots>i_{m} and ir>jri_{r}>j_{r} for all rr.

For example, if σi1\sigma_{i_{1}}, σi2\sigma_{i_{2}}, and σi3\sigma_{i_{3}} are distinct simplices, then I𝒂(σi3,σi2)I𝒂(σi2,σi1)I_{\bm{a}}(\sigma_{i_{3}},\sigma_{i_{2}})\cap I_{\bm{a}}(\sigma_{i_{2}},\sigma_{i_{1}}) is 22-reduced, but I𝒂(σi3,σi1)I𝒂(σi3,σi2)I𝒂(σi2,σi1)I_{\bm{a}}(\sigma_{i_{3}},\sigma_{i_{1}})\cap I_{\bm{a}}(\sigma_{i_{3}},\sigma_{i_{2}})\cap I_{\bm{a}}(\sigma_{i_{2}},\sigma_{i_{1}}) is not 33-reduced. We define

(12) E𝒂m¯\displaystyle\overline{E^{m}_{\bm{a}}} :={SE𝒂mS is m-reduced},\displaystyle:=\{S\in E^{m}_{\bm{a}}\mid S\text{ is }m\text{-reduced}\}\,,
(13) E𝒂m,k¯\displaystyle\overline{E^{m,k}_{\bm{a}}} :={SE𝒂m,kS is m-reduced},\displaystyle:=\{S\in E^{m,k}_{\bm{a}}\mid S\text{ is }m\text{-reduced}\}\,,

where E𝒂mE^{m}_{\bm{a}} is defined as in (10) and E𝒂m,kE^{m,k}_{\bm{a}} is defined as in (11).

Lemma 4.8.

For all m1m\geq 1, all kk, and all 𝒂A\bm{a}\in A, where AA is defined as in (8), every SE𝒂m,kS\in E^{m,k}_{\bm{a}} belongs to E𝒂m,k¯\overline{E^{m^{\prime},k}_{\bm{a}}} for some mmm^{\prime}\leq m, where E𝒂m,kE^{m,k}_{\bm{a}} and E𝒂m,k¯\overline{E^{m^{\prime},k}_{\bm{a}}} are defined as in (11) and (13), respectively.

Proof.

We prove the lemma by induction on mm. For all kk, every SE𝒂1,kS\in E^{1,k}_{\bm{a}} is 11-reduced by definition. Assume that Lemma 4.8 is true for m11m-1\geq 1, and let SS be an element of E𝒂m,kE^{m,k}_{\bm{a}}. The set SS is equal to a set of the form

I𝒂(σi1,σj1)I𝒂(σim,σjm),I_{\bm{a}}(\sigma_{i_{1}},\sigma_{j_{1}})\cap\cdots\cap I_{\bm{a}}(\sigma_{i_{m}},\sigma_{j_{m}})\,,

where ir>jri_{r}>j_{r} for all rr and ki1i2imk\geq i_{1}\geq i_{2}\geq\cdots\geq i_{m} without loss of generality. By the induction hypothesis,

I𝒂(σi2,σj2)I𝒂(σim,σjm)=I𝒂(σi2,σj2)I𝒂(σi,σj)I_{\bm{a}}(\sigma_{i_{2}},\sigma_{j_{2}})\cap\cdots\cap I_{\bm{a}}(\sigma_{i_{m}},\sigma_{j_{m}})=I_{\bm{a}}(\sigma_{i_{2}^{\prime}},\sigma_{j_{2}^{\prime}})\cap\cdots\cap I_{\bm{a}}(\sigma_{i_{\ell}^{\prime}},\sigma_{j_{\ell}^{\prime}})

for some m\ell\leq m, where ir>jri^{\prime}_{r}>j^{\prime}_{r} for all rr and ki1i2i2>i3>>ik\geq i_{1}\geq i_{2}\geq i^{\prime}_{2}>i^{\prime}_{3}>\cdots>i^{\prime}_{\ell}. If i1>i2i_{1}>i^{\prime}_{2}, then SS is an element of E,kE^{\ell,k} and we are done. Otherwise,

I𝒂(σi1,σj1)I𝒂(σi2,σj2)=I𝒂(σi1,σj1)I𝒂(σj1,σj2)I_{\bm{a}}(\sigma_{i_{1}},\sigma_{j_{1}})\cap I_{\bm{a}}(\sigma_{i_{2}^{\prime}},\sigma_{j_{2}^{\prime}})=I_{\bm{a}}(\sigma_{i_{1}},\sigma_{j_{1}})\cap I_{\bm{a}}(\sigma_{j_{1}},\sigma_{j_{2}^{\prime}})

because i1=i2i_{1}=i_{2}^{\prime}. If j1=j2j_{1}=j_{2}^{\prime}, then

S=I𝒂(σi1,σj1)I𝒂(σi3,σj3)I𝒂(σi,σj),S=I_{\bm{a}}(\sigma_{i_{1}},\sigma_{j_{1}})\cap I_{\bm{a}}(\sigma_{i_{3}^{\prime}},\sigma_{j_{3}^{\prime}})\cap\cdots\cap I_{\bm{a}}(\sigma_{i_{\ell}^{\prime}},\sigma_{j_{\ell}^{\prime}})\,,

so SS is (1)(\ell-1)-reduced. Otherwise,

S=I𝒂(σi1,σj1)I𝒂(σj1,σj2)I𝒂(σi3,σj3)I𝒂(σi,σj),S=I_{\bm{a}}(\sigma_{i_{1}},\sigma_{j_{1}})\cap I_{\bm{a}}(\sigma_{j_{1}},\sigma_{j_{2}^{\prime}})\cap I_{\bm{a}}(\sigma_{i_{3}^{\prime}},\sigma_{j_{3}^{\prime}})\cap\cdots\cap I_{\bm{a}}(\sigma_{i_{\ell}^{\prime}},\sigma_{j_{\ell}^{\prime}})\,,

where ki1>j1,j2k\geq i_{1}>j_{1},j_{2}^{\prime} and i1>iri_{1}>i_{r}^{\prime} for all r3r\geq 3. By the induction hypothesis, the set I𝒂(σj1,σj2)I𝒂(σi3,σj3)I𝒂(σi,σj)I_{\bm{a}}(\sigma_{j_{1}},\sigma_{j_{2}^{\prime}})\cap I_{\bm{a}}(\sigma_{i_{3}^{\prime}},\sigma_{j_{3}^{\prime}})\cap\cdots\cap I_{\bm{a}}(\sigma_{i_{\ell}^{\prime}},\sigma_{j_{\ell}^{\prime}}) belongs to E𝒂,k1¯\overline{E^{\ell^{\prime},k-1}_{\bm{a}}} for some 1\ell^{\prime}\leq\ell-1, so SS belongs to E𝒂+1,k¯\overline{E^{\ell^{\prime}+1,k}_{\bm{a}}}, where +1m\ell^{\prime}+1\leq\ell\leq m. ∎

Lemma 4.9.

For almost every 𝒂A\bm{a}\in A (where AA is defined as in (8)), we have that M𝒂,iM_{\bm{a},i} intersects M𝒂,jM_{\bm{a},j} transversely333When manifolds M1M_{1} and M2M_{2} intersect transversely, we will use the notation M1M2M_{1}\pitchfork M_{2}. for iji\neq j and every SE𝒂m¯S\in\overline{E^{m}_{\bm{a}}} is either \emptyset or an (nm)(n-m)-dimensional submanifold of BB for all m{1,,n}m\in\{1,\ldots,n\}, where M𝒂,iM_{\bm{a},i} is defined as in (9) and E𝒂m¯\overline{E^{m}_{\bm{a}}} is defined as in (12).

Proof.

Define gij(p):=f(σi,p)f(σj,p)g^{ij}(p):=f(\sigma_{i},p)-f(\sigma_{j},p) for all iji\neq j. For almost every 𝒂A\bm{a}\in A, the quantity ajaia_{j}-a_{i} is a regular value of gijg^{ij} by Sard’s Theorem. The set of regular values is open for all iji\neq j because gijg^{ij} is smooth and BB is compact. Therefore, there is an ϵ\epsilon^{*} such that for all iji\neq j, every y(ajai2ϵ,ajai+2ϵ)y\in(a_{j}-a_{i}-2\epsilon^{*},a_{j}-a_{i}+2\epsilon^{*}) is a regular value of gijg^{ij}.

Given an 𝒂\bm{a} and ϵ\epsilon^{*} as above, it suffices to show that for almost every ϵN\bm{\epsilon}\in\mathbb{R}^{N} with |ϵi|ϵ|\epsilon_{i}|\leq\epsilon^{*}, we have that every SE𝒂+ϵm¯S\in\overline{E^{m}_{\bm{a+\epsilon}}} is an (nm)(n-m)-dimensional submanifold of BB for all mm. For m=1m=1, every element of E𝒂+ϵ1¯\overline{E^{1}_{\bm{a+\epsilon}}} is of the form I𝒂+ϵ(σi,σj)I_{\bm{a+\epsilon}}(\sigma_{i},\sigma_{j}) for some iji\neq j. The set I𝒂+ϵ(σi,σj)I_{\bm{a+\epsilon}}(\sigma_{i},\sigma_{j}) is the (ajai+ϵjϵi)(a_{j}-a_{i}+\epsilon_{j}-\epsilon_{i})-level set of gijg^{ij}. Because (ajai+ϵjϵi)(a_{j}-a_{i}+\epsilon_{j}-\epsilon_{i}) is a regular value of gijg^{ij}, the set I𝒂+ϵ(σi,σj)I_{\bm{a+\epsilon}}(\sigma_{i},\sigma_{j}) is an (n1)(n-1)-dimensional submanifold of BB and we must have M𝒂,iM𝒂,jM_{\bm{a},i}\pitchfork M_{\bm{a},j}.

For m2m\geq 2, observe that

E𝒂m¯=E𝒂m,2¯(k=3NE𝒂m,k¯E𝒂m,k1¯),\overline{E^{m}_{\bm{a}}}=\overline{E^{m,2}_{\bm{a}}}\cup\Big{(}\bigcup_{k=3}^{N}\overline{E^{m,k}_{\bm{a}}}\setminus\overline{E^{m,k-1}_{\bm{a}}}\Big{)}\,,

where E𝒂m,k¯\overline{E^{m,k}_{\bm{a}}} is defined as in (13). We induct on k{2,,N}k\in\{2,\ldots,N\}, where NN is the number of simplices in 𝒦\mathcal{K}. When k=2k=2, we have

E𝒂+ϵm,2¯={{I𝒂+ϵ(σ1,σ2)},m=1,m2,\overline{E^{m,2}_{\bm{a+\epsilon}}}=\begin{cases}\{I_{\bm{a+\epsilon}}(\sigma_{1},\sigma_{2})\}\,,&m=1\\ \emptyset\,,&m\geq 2\,,\end{cases}

so every SE𝒂+ϵm,2¯S\in\overline{E^{m,2}_{\bm{a+\epsilon}}} is either \emptyset or an (nm)(n-m)-dimensional submanifold of BB. Now suppose that k>2k>2 and that every element of SE𝒂+ϵm,k1¯S\in\overline{E^{m,k-1}_{\bm{a+\epsilon}}} is either \emptyset or an (nm)(n-m)-dimensional submanifold for all mm. Every element SS in E𝒂+ϵm,k¯E𝒂+ϵm,k1¯\overline{E^{m,k}_{\bm{a+\epsilon}}}\setminus\overline{E^{m,k-1}_{\bm{a+\epsilon}}} is equal to a set of the form

S=I𝒂+ϵ(σk,σ)S,S=I_{\bm{a+\epsilon}}(\sigma_{k},\sigma_{\ell})\cap S^{\prime}\,,

where k1\ell\leq k-1 and SE𝒂+ϵm1,k1¯S^{\prime}\in\overline{E^{m-1,k-1}_{\bm{a+\epsilon}}}. We define the vectors ϵj:=(0,,0,ϵj,0,,0)\bm{\epsilon}^{j}:=(0,\ldots,0,\epsilon_{j},0,\ldots,0) and 𝒃j:=𝒂+ϵϵj\bm{b}^{j}:=\bm{a+\epsilon}-\bm{\epsilon}^{j}. Note that I𝒂+ϵ(σk,σ)=I𝒃+ϵk(σk,σ)I_{\bm{a}+\bm{\epsilon}}(\sigma_{k},\sigma_{\ell})=I_{\bm{b}+\bm{\epsilon}^{k}}(\sigma_{k},\sigma_{\ell}) because bik+ϵik=ai+ϵib_{i}^{k}+\epsilon_{i}^{k}=a_{i}+\epsilon_{i} for all ii, and E𝒂+ϵm1,k1¯=E𝒃km1,k1¯\overline{E^{m-1,k-1}_{\bm{a+\epsilon}}}=\overline{E^{m-1,k-1}_{\bm{b}^{k}}} because ai+ϵi=bia_{i}+\epsilon_{i}=b_{i} for all ik1i\leq k-1. Therefore, every SE𝒂+ϵm,k¯E𝒂+ϵm,k1¯S\in\overline{E^{m,k}_{\bm{a+\epsilon}}}\setminus\overline{E^{m,k-1}_{\bm{a+\epsilon}}} is equal to a set of the form

S=I𝒃k+ϵk(σk,σ)SS=I_{\bm{b}^{k}+\bm{\epsilon}^{k}}(\sigma_{k},\sigma_{\ell})\cap S^{\prime}

for some SE𝒃km1,k1¯S^{\prime}\in\overline{E^{m-1,k-1}_{\bm{b}^{k}}} and k1\ell\leq k-1. Because gkg^{\ell k} has no critical values between bbkϵkb_{\ell}-b_{k}-\epsilon_{k} and bbkb_{\ell}-b_{k}, we have that I𝒃k+ϵk(σk,σ)I_{\bm{b}^{k}+\bm{\epsilon}^{k}}(\sigma_{k},\sigma_{\ell}) is diffeomorphic to I𝒃k(σk,σ)I_{\bm{b}^{k}}(\sigma_{k},\sigma_{\ell}) for all ϵk(ϵ,ϵ)\epsilon_{k}\in(-\epsilon^{*},\epsilon^{*}). In other words, I𝒃k+ϵk(σk,σ)I_{\bm{b}^{k}+\bm{\epsilon}^{k}}(\sigma_{k},\sigma_{\ell}) is a perturbation of I𝒃k(σk,σ)I_{\bm{b}^{k}}(\sigma_{k},\sigma_{\ell}) for all ϵk(ϵ,ϵ)\epsilon_{k}\in(-\epsilon^{*},\epsilon^{*}). By Thom’s Transversality Theorem, I𝒃k+ϵk(σk,σ)I_{\bm{b}^{k}+\bm{\epsilon}^{k}}(\sigma_{k},\sigma_{\ell}) intersects every SE𝒃km1,k1S^{\prime}\in E_{\bm{b}^{k}}^{m-1,k-1} transversely for almost every ϵk(ϵ,ϵ)\epsilon_{k}\in(-\epsilon^{*},\epsilon^{*}). This shows that SS is either \emptyset or an (nm)(n-m)-dimensional submanifold of BB for almost every ϵk(ϵ,ϵ)\epsilon_{k}\in(\epsilon^{*},\epsilon^{*}). Because there are finitely many elements in E𝒂m,k¯\overline{E^{m,k}_{\bm{a}}}, we must have that every SE𝒂m,k¯S\in\overline{E^{m,k}_{\bm{a}}} is either \emptyset or an (nm)(n-m)-dimensional submanifold of BB for almost every ϵk(ϵ,ϵ)\epsilon_{k}\in(\epsilon^{*},\epsilon^{*}). Induction on kk concludes the proof. ∎

Lemma 4.10.

For all 𝒂A\bm{a}\in A, with AA defined as in (8), define

(14) B𝒂n:=BB𝒂m:=mSE𝒂n¯S for m<n.B^{n}_{\bm{a}}:=B\,\qquad B^{m}_{\bm{a}}:=\bigcup_{\ell\leq m}\bigcup_{S\in\overline{E_{\bm{a}}^{n-\ell}}}S\qquad\text{ for }m<n\,.

If 𝒂A\bm{a}\in A is such that every SE𝒂n¯S\in\overline{E_{\bm{a}}^{n-\ell}} is either \emptyset or an \ell-dimensional smooth submanifold for every {1,,n}\ell\in\{1,\ldots,n\}, where E𝒂n¯\overline{E_{\bm{a}}^{n-\ell}} is defined as in (12), then B𝒂mB𝒂m1B^{m}_{\bm{a}}\setminus B^{m-1}_{\bm{a}} is the disjoint union of smooth mm-dimensional manifolds.

Proof.

We have that

B𝒂mB𝒂m1=SE𝒂nm¯(SSE𝒂n¯m1S).B_{\bm{a}}^{m}\setminus B_{\bm{a}}^{m-1}=\bigcup_{S\in\overline{E^{n-m}_{\bm{a}}}}\Big{(}S\setminus\bigcup_{\begin{subarray}{c}S^{\prime}\in\overline{E^{n-\ell}_{\bm{a}}}\\ \ell\leq m-1\end{subarray}}S^{\prime}\Big{)}\,.

If SE𝒂n¯S^{\prime}\in\overline{E^{n-\ell}_{\bm{a}}} is a subset of SE𝒂nm¯S\in\overline{E^{n-m}_{\bm{a}}}, then SS^{\prime} is a closed subset of SS. Therefore, the set S(SE𝒂n¯m1S)S\setminus\Big{(}\bigcup_{\begin{subarray}{c}S^{\prime}\in\overline{E^{n-\ell}_{\bm{a}}}\\ \ell\leq m-1\end{subarray}}S^{\prime}\Big{)} is an open subset of the smooth manifold SS, which implies that S(SE𝒂n¯m1S)S\setminus\Big{(}\bigcup_{\begin{subarray}{c}S^{\prime}\in\overline{E^{n-\ell}_{\bm{a}}}\\ \ell\leq m-1\end{subarray}}S^{\prime}\Big{)} is a smooth manifold. If S1S_{1} and S2S_{2} are distinct elements of E𝒂nm¯\overline{E^{n-m}_{\bm{a}}}, then

(S1SE𝒂n¯m1S)(S2SE𝒂n¯m1S)=,\Big{(}S_{1}\setminus\bigcup_{\begin{subarray}{c}S^{\prime}\in\overline{E^{n-\ell}_{\bm{a}}}\\ \ell\leq m-1\end{subarray}}S^{\prime}\Big{)}\cap\Big{(}S_{2}\setminus\bigcup_{\begin{subarray}{c}S^{\prime}\in\overline{E^{n-\ell}_{\bm{a}}}\\ \ell\leq m-1\end{subarray}}S^{\prime}\Big{)}=\emptyset\,,

which completes the proof. ∎

For the remainder of Section 4.2, let {B𝒂m}m=0n\{B^{m}_{\bm{a}}\}_{m=0}^{n} be defined as in (14), and define

(15) 𝒴𝒂=m=0n𝒴𝒂m,\mathcal{Y}_{\bm{a}}=\bigcup_{m=0}^{n}\mathcal{Y}^{m}_{\bm{a}}\,,

where 𝒴𝒂m\mathcal{Y}^{m}_{\bm{a}} is the set of connected components of B𝒂mB𝒂m1B^{m}_{\bm{a}}\setminus B^{m-1}_{\bm{a}} (with B𝒂1:=B^{-1}_{\bm{a}}:=\emptyset).

Lemma 4.11.

Let AA be defined as in (8). If 𝒂A\bm{a}\in A is such that each Y𝒴𝒂Y\in\mathcal{Y}_{\bm{a}} is a manifold, then the simplex order induced by ff is constant in each YY. That is, there is a strict partial order Y\prec_{Y} on 𝒦\mathcal{K} such that f𝒂(,y)\prec_{f_{\bm{a}}(\cdot,y)} is the same as Y\prec_{Y} for all yYy\in Y.

Proof.

Let Y𝒴𝒂Y\in\mathcal{Y}_{\bm{a}}. The set YY is connected by definition. Because YY is a manifold, it is also path-connected. For each pair (σ,τ)(\sigma,\tau) of simplices, we have by construction that YI𝒂(σ,τ)Y\cap I_{\bm{a}}(\sigma,\tau) equals either \emptyset or YY. (In fact, this statement holds for all 𝒂A\bm{a}\in A and does not require YY to be a manifold.) By Lemma 4.4, the simplex order is constant in YY. ∎

Lemma 4.12.

For almost every 𝒂A\bm{a}\in A (where AA is defined as in (8)), we have that r=1mI𝒂(σir,σjr)\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) is a submanifold of BB and

(16) Tp(r=1mI𝒂(σir,σjr))=r=1mTp(I𝒂(σir,σjr))T_{p}\Big{(}\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}=\bigcap_{r=1}^{m}T_{p}\Big{(}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}

for all points pr=1mI𝒂(σir,σjr)p\in\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) and all sets {(ir,jr)}r=1m\{(i_{r},j_{r})\}_{r=1}^{m} of index pairs such that {ir,jr}{is,js}\{i_{r},j_{r}\}\neq\{i_{s},j_{s}\} if rsr\neq s.

Proof.

Because there are finitely many sets of index pairs, it suffices to fix a set {(ir,jr)}r=1m\{(i_{r},j_{r})\}_{r=1}^{m} of index pairs and show that (16) holds for all yr=1mI𝒂(σir,σjr)y\in\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) for almost every 𝒂A\bm{a}\in A. By Lemmas 4.8 and 4.9, the set r=1mI𝒂(σir,σjr)\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) is a manifold for almost every 𝒂A\bm{a}\in A. By Lemma A.2, there is a finite open cover {Uk}k=1K\{U_{k}\}_{k=1}^{K} of BB such that for each kk, there is a disjoint partition J,k={1,,m}\bigcup_{\ell}J_{\ell,k}=\{1,\ldots,m\} such that {ir,jrrJ1,k}{ir,jrrJ2,k}=\{i_{r},j_{r}\mid r\in J_{\ell_{1},k}\}\cap\{i_{r},j_{r}\mid r\in J_{\ell_{2},k}\}=\emptyset if 12\ell_{1}\neq\ell_{2} and

π(rJ,k(M𝒂,irM𝒂,jr))Uk=rJ,kI𝒂(σir,σjr)Uk\pi\Big{(}\bigcap_{r\in J_{\ell,k}}(M_{\bm{a},i_{r}}\cap M_{\bm{a},j_{r}})\Big{)}\cap U_{k}=\bigcap_{r\in J_{\ell,k}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U_{k}

for all \ell, where π\pi is the projection π:B×B\pi:B\times\mathbb{R}\to B.444Recall that an arbitrary intersection rI𝒂(σir,σjr)B\bigcap_{r}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\subseteq B does not necessarily lift to an intersection r=1m(M𝒂,iM𝒂,j)B×\bigcap_{r=1}^{m}(M_{\bm{a},i}\cap M_{\bm{a},j})\subseteq B\times\mathbb{R}. Lemma A.2 says that in local neighborhoods UBU\subseteq B, we can partition the set {1,,m}\{1,\ldots,m\} into subsets JJ_{\ell} such that the intersection rJI𝒂(σir,σjr)\bigcap_{r\in J_{\ell}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) does lift to a subset of the intersection rJ(M𝒂,iM𝒂,j)\bigcap_{r\in J_{\ell}}(M_{\bm{a},i}\cap M_{\bm{a},j}). Because the number KK of open sets is finite, it suffices to fix UkU_{k} and show that (16) holds for all pr=1mI𝒂(σir,σjr)Ukp\in\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U_{k} for almost every 𝒂A\bm{a}\in A.

By Lemma A.4, we have

Tp(r=1mI𝒂(σir,σjr))=Tp(rJ,kI𝒂(σir,σjr))=Tp(rJ,k(I𝒂(σir,σjr)))T_{p}\Big{(}\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}=T_{p}\Big{(}\bigcap_{\ell}\bigcap_{r\in J_{\ell,k}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}=\bigcap_{\ell}T_{p}\Big{(}\bigcap_{r\in J_{\ell,k}}\Big{(}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}\Big{)}

for all pr=1mI𝒂(σir,σjr)Ukp\in\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U_{k} for almost every 𝒂A\bm{a}\in A. By Lemma A.5, we have

Tp(rJ,k(I𝒂(σir,σjr)))=rJ,kTp(I𝒂(σir,σjr)\bigcap_{\ell}T_{p}\Big{(}\bigcap_{r\in J_{\ell,k}}\Big{(}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}\Big{)}=\bigcap_{\ell}\bigcap_{r\in J_{\ell,k}}T_{p}(I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})

for all pr=1mI𝒂(σir,σjr)Ukp\in\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U_{k} for almost every 𝒂A\bm{a}\in A. ∎

For any strict partial order \prec on 𝒦\mathcal{K}, we define

Z𝒂:=\displaystyle Z^{\prec}_{\bm{a}}:= {pBf𝒂(σ,p)<f𝒂(τ,p) if στ\displaystyle\{p\in B\mid f_{\bm{a}}(\sigma,p)<f_{\bm{a}}(\tau,p)\text{ if }\sigma\prec\tau
(17)  and f𝒂(σ,p)=f𝒂(τ,p) if στ and τσ}.\displaystyle\qquad\text{ and }f_{\bm{a}}(\sigma,p)=f_{\bm{a}}(\tau,p)\text{ if }\sigma\not\prec\tau\text{ and }\tau\not\prec\sigma\}\,.

That is, Z𝒂Z_{\bm{a}}^{\prec} is the subset of BB such that for all zz in Z𝒂Z_{\bm{a}}^{\prec}, the strict partial order f𝒂(,z)\prec_{f_{\bm{a}}(\cdot,z)} is the same as \prec.

Lemma 4.13.

Let AA be defined as in (8). If 𝒂A\bm{a}\in A is such that

  1. (1)

    every Y𝒴𝒂Y\in\mathcal{Y}_{\bm{a}} is a manifold, where 𝒴𝒂\mathcal{Y}_{\bm{a}} is defined as in (15),

  2. (2)

    M𝒂,iM𝒂,jM_{\bm{a},i}\pitchfork M_{\bm{a},j} for all iji\neq j, where M𝒂,iM_{\bm{a},i} is defined as in (9),

  3. (3)

    r=1mI𝒂(σir,σjr)\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) is a manifold for all sets {(ir,jr}r=1m\{(i_{r},j_{r}\}_{r=1}^{m} of index pairs, and

  4. (4)

    Tp(r=1mI𝒂(σir,σjr))=r=1mTp(I𝒂(σir,σjr))T_{p}\Big{(}\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}=\bigcap_{r=1}^{m}T_{p}\Big{(}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)} for all sets {(ir,jr}r=1m\{(i_{r},j_{r}\}_{r=1}^{m} of index pairs and all pr=1mI𝒂(σir,σjr)p\in\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}),

then 𝒴𝒂\mathcal{Y}_{\bm{a}} is locally finite.

Proof.

Let pp be a point in BB. There are finitely many strict partial orders 1,,i\prec_{1},\ldots,\prec_{i} on 𝒦\mathcal{K}. By Lemma A.6, we have that for each j{1,,i}j\in\{1,\ldots,i\}, there is a subset 𝒴𝒂j𝒴𝒂\mathcal{Y}^{\prec_{j}}_{\bm{a}}\subseteq\mathcal{Y}_{\bm{a}} such that Z𝒂j=Y𝒴𝒂jYZ^{\prec_{j}}_{\bm{a}}=\bigcup_{Y\in\mathcal{Y}^{\prec_{j}}_{\bm{a}}}Y. For each jj, the point pp has a neighborhood UjU_{j} that intersects at most one Y𝒴𝒂jY\in\mathcal{Y}^{\prec_{j}}_{\bm{a}} by Lemma A.7. Therefore j=1iUj\bigcap_{j=1}^{i}U_{j} is a neighborhood of pp that intersects at most ii elements of 𝒴𝒂\mathcal{Y}_{\bm{a}}. ∎

Lemma 4.14.

Let AA be defined as in (8). If 𝒂A\bm{a}\in A is such that

  1. (1)

    every SE𝒂n¯S\in\overline{E_{\bm{a}}^{n-\ell}} is an \ell-dimensional smooth submanifold for every {1,,n}\ell\in\{1,\ldots,n\}, where E𝒂n¯\overline{E_{\bm{a}}^{n-\ell}} is defined as in (12),

  2. (2)

    M𝒂,iM𝒂,jM_{\bm{a},i}\pitchfork M_{\bm{a},j} for all iji\neq j, where M𝒂,iM_{\bm{a},i} is defined as in (9),

  3. (3)

    i=rmI𝒂(σir,σjr)\bigcap_{i=r}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) is a manifold for all sets {(ir,jr}r=1m\{(i_{r},j_{r}\}_{r=1}^{m} of index pairs, and

  4. (4)

    Tp(i=rmI𝒂(σir,σjr))=i=1mTp(I𝒂(σir,σjr))T_{p}\Big{(}\bigcap_{i=r}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}=\bigcap_{i=1}^{m}T_{p}\Big{(}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)} for all sets {(ir,jr}r=1m\{(i_{r},j_{r}\}_{r=1}^{m} of index pairs and all pi=rmI𝒂(σir,σjr)p\in\bigcap_{i=r}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}),

then 𝒴𝒂\mathcal{Y}_{\bm{a}} satisfies the Axiom of the Frontier in Definition 4.1, where 𝒴𝒂\mathcal{Y}_{\bm{a}} is defined as in (15)

Proof.

By Lemma 4.10, each Y𝒴𝒂Y\in\mathcal{Y}_{\bm{a}} is a manifold. Let YαY_{\alpha} be an element of 𝒴𝒂\mathcal{Y}_{\bm{a}}. It suffices to show that if YβYαY_{\beta}\neq Y_{\alpha} is another element of 𝒴𝒂\mathcal{Y}_{\bm{a}} and YβYαY_{\beta}\cap\partial Y_{\alpha}\neq\emptyset, where Yα\partial Y_{\alpha} denotes the boundary of the manifold YαY_{\alpha}, then YβYαY_{\beta}\subseteq\partial Y_{\alpha}.

By Lemma 4.11, the simplex order induced by ff is constant on each YY, so there is a strict partial order α\prec_{\alpha} on 𝒦\mathcal{K} such that f𝒂(,y)\prec_{f_{\bm{a}}(\cdot,y)} is the same as α\prec_{\alpha} for all yYαy\in Y_{\alpha}. Let Z𝒂αZ^{\prec_{\alpha}}_{\bm{a}} be defined as in (4.2). By Lemma A.6, there is a subset 𝒴𝒂α𝒴𝒂\mathcal{Y}^{\prec_{\alpha}}_{\bm{a}}\subseteq\mathcal{Y}_{\bm{a}} such that Yα𝒴𝒂αY_{\alpha}\in\mathcal{Y}^{\prec_{\alpha}}_{\bm{a}} and Z𝒂α=Y𝒴𝒂αYZ^{\prec_{\alpha}}_{\bm{a}}=\bigcup_{Y\in\mathcal{Y}^{\prec_{\alpha}}_{\bm{a}}}Y. We have Z𝒂α=Y𝒴𝒂αY\partial Z^{\prec_{\alpha}}_{\bm{a}}=\bigcup_{Y\in\mathcal{Y}^{\prec_{\alpha}}_{\bm{a}}}\partial Y because 𝒴𝒂\mathcal{Y}_{\bm{a}} is locally finite by Lemma 4.13. Therefore,

(18) YβZ𝒂α=Y𝒴𝒂α(YβY).Y_{\beta}\cap\partial Z^{\prec_{\alpha}}_{\bm{a}}=\bigcup_{Y\in\mathcal{Y}^{\prec_{\alpha}}_{\bm{a}}}(Y_{\beta}\cap\partial Y)\,.

By Lemmas A.6 and A.8, we have that if Y𝒴𝒂Y\in\mathcal{Y}_{\bm{a}} intersects Z𝒂α\partial Z^{\prec_{\alpha}}_{\bm{a}}, then YZ𝒂αY\subseteq\partial Z^{\prec_{\alpha}}_{\bm{a}}. Therefore YβZ𝒂αY_{\beta}\subseteq\partial Z^{\prec_{\alpha}}_{\bm{a}} because YβZ𝒂αY_{\beta}\cap\partial Z^{\prec_{\alpha}}_{\bm{a}} contains YβYαY_{\beta}\cap\partial Y_{\alpha}\neq\emptyset. Together with (18), this shows that

(19) Yβ=Y𝒴𝒂α(YβY).Y_{\beta}=\bigcup_{Y\in\mathcal{Y}^{\prec_{\alpha}}_{\bm{a}}}(Y_{\beta}\cap\partial Y)\,.

By Lemma A.7, every point in BB has a neighborhood that intersects at most one Y𝒴𝒂αY\in\mathcal{Y}^{\prec_{\alpha}}_{\bm{a}}, so YY=\partial Y^{\prime}\cap\partial Y=\emptyset for all Y,Y𝒴𝒂αY,Y^{\prime}\in\mathcal{Y}^{\prec_{\alpha}}_{\bm{a}} such that YYY\neq Y^{\prime}. Because YβY_{\beta} is connected (by definition) and YβYαY_{\beta}\cap\partial Y_{\alpha}\neq\emptyset, we must have that YβY=Y_{\beta}\cap\partial Y=\emptyset for all Y𝒴𝒂αY\in\mathcal{Y}^{\prec_{\alpha}}_{\bm{a}} such that YYαY\neq Y_{\alpha}. By (19),

Yβ=YβYαYα.Y_{\beta}=Y_{\beta}\cap\partial Y_{\alpha}\subseteq\partial Y_{\alpha}\,.

Theorem 4.15.

Let BB be a smooth compact nn-dimensional manifold. For every 𝐚A\bm{a}\in A, define {B𝐚m}m=0n\{B^{m}_{\bm{a}}\}_{m=0}^{n} as in (14), with AA defined as in (8). For almost every 𝐚A\bm{a}\in A, we have that {B𝐚m}m=0n\{B^{m}_{\bm{a}}\}_{m=0}^{n} is a stratification of BB. In each stratum YY, the simplex order induced by f𝐚f_{\bm{a}} is constant. (In other words, there is a strict partial order Y\prec_{Y} on 𝒦\mathcal{K} such that f𝐚(,y)\prec_{f_{\bm{a}}(\cdot,y)} is the same as Y\prec_{Y} for all yYy\in Y.) Consequently, the set {(σb,σd)}\{(\sigma_{b},\sigma_{d})\} of (birth, death) simplex pairs is constant in each stratum YY and for any pYp\in Y, the persistence diagram PD(fp)PD(f_{p}) consists of the diagonal (with infinite multiplicity) and the multiset {(f(σb,p),f(σd,p))}\{(f(\sigma_{b},p),f(\sigma_{d},p))\}.

Proof.

By Lemmas 4.8, 4.9, 4.10, 4.12, 4.13, and 4.14, {B𝒂m}m=0n\{B_{\bm{a}}^{m}\}_{m=0}^{n} is a stratification of BB for almost every 𝒂A\bm{a}\in A. By Lemma 4.11, the simplex order induced by f𝒂f_{\bm{a}} is constant in each stratum Y𝒴𝒂Y\in\mathcal{Y}_{\bm{a}} whenever {B𝒂m}m=0n\{B_{\bm{a}}^{m}\}_{m=0}^{n} is a stratification of BB. The last statement of Theorem 4.15 follows from Lemma 2.3. ∎

5. Monodromy in PD Bundles

Definition 5.1 (Local section).

Let (E,B,π)(E,B,\pi) be a PD bundle. A local section is a continuous map s:UEs:U\to E, where UU is an open set in BB and πs(p)=p\pi\circ s(p)=p for all pUp\in U.

For example, consider a vineyard, in which BB is an interval II in \mathbb{R}. Let (t0,T)(t_{0},T) be an open interval in II. A local section in the vineyard is a map s:(t0,T)Es:(t_{0},T)\to E that parameterizes an open subset of one of the vines (a curve in 3\mathbb{R}^{3}).

Definition 5.2 (Global section).

Let (E,B,π)(E,B,\pi) be a PD bundle. A global section is a continuous map s:BEs:B\to E with πs(p)=p\pi\circ s(p)=p for all pBp\in B. In particular, a nontrivial global section is a global section s:BEs:B\to E such that there exists a pBp_{*}\in B for which s(p)s(p_{*}) is not on the diagonal of PD(fp)PD(f_{p_{*}}).

In a vineyard, every local section can be extended to a global section. In other words, we can trace out how the persistence of a single homology class changes over B=[t0,t1]B=[t_{0},t_{1}]\subseteq\mathbb{R}, so there are individual “vines” in the vineyard. We will show that local sections of a PD bundle cannot necessarily be extended to global sections. Consequently, a PD bundle does not necessarily have a decomposition of the form (3); if it does, then each γ\gamma is a global section.

Proposition 5.3.

There is a PD bundle (E,B,π)(E,B,\pi) for which no nontrivial global sections exist.

Proof.

The proof is constructive. Let 𝒦\mathcal{K} be the simplicial complex in Figure 4a, which has vertices 0, 11, 22, and 33. Let aa be the edge with vertices (0,1)(0,1), let bb be the edge with vertices (0,2)(0,2), let cc be the triangle with vertices (0,1,2)(0,1,2), and let dd be the triangle with vertices (0,2,3)(0,2,3).

Let f:𝒦×2f:\mathcal{K}\times\mathbb{R}^{2}\to\mathbb{R} be a continuous fibered filtration function that satisfies the following conditions:

f(c,(x,y)),f(d,(x,y))\displaystyle f(c,(x,y)),f(d,(x,y)) >f(a,(x,y)),f(b,(x,y))>0for all (x,y)2,\displaystyle>f(a,(x,y)),f(b,(x,y))>0\qquad\text{for all }(x,y)\in\mathbb{R}^{2}\,,
f(a,(x,y))\displaystyle f(a,(x,y)) >f(b,(x,y)),y>0,\displaystyle>f(b,(x,y))\,,\qquad y>0\,,
f(a,(x,y))\displaystyle f(a,(x,y)) <f(b,(x,y)),y<0,\displaystyle<f(b,(x,y))\,,\qquad y<0\,,
f(c,(x,y))\displaystyle f(c,(x,y)) >f(d,(x,y)),x>0,\displaystyle>f(d,(x,y))\,,\qquad x>0\,,
f(c,(x,y))\displaystyle f(c,(x,y)) <f(d,(x,y)),x<0,\displaystyle<f(d,(x,y))\,,\qquad x<0\,,
f(σ,(x,y))\displaystyle f(\sigma,(x,y)) =0,for all other σ, for all x,y.\displaystyle=0\,,\qquad\text{for all other }\sigma,\text{ for all }x,y\,.

The conditions on the fibered filtration function ff are illustrated in Figure 4b.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 4. (A) The simplicial complex 𝒦\mathcal{K} that is defined in the proof of Proposition 5.3. (B) The conditions on the fibered filtration f:𝒦×2f:\mathcal{K}\times\mathbb{R}^{2}\to\mathbb{R} that is defined in the proof of Proposition 5.3. (C) The (birth, death) simplex pairs in each quadrant for the 1D PH.

These conditions imply that simplices aa and bb swap their order along the xx-axis and the simplices cc and dd swap their order along the yy-axis.

In Figure 4c, we list the (birth, death) simplex pairs for the 1D PH in each quadrant. In quadrants 11, 22, and 44, the simplex pairs are (a,c)(a,c) and (b,d)(b,d). In quadrant 33, the simplex pairs are (a,d)(a,d) and (b,c)(b,c).

Let (E,2,π)(E,\mathbb{R}^{2},\pi) be the corresponding PD bundle, where E={((x,y),z)(x,y)2,zPD1(f(,(x,y)))}E=\{((x,y),z)\mid(x,y)\in\mathbb{R}^{2},z\in PD_{1}(f(\cdot,(x,y)))\} is the total space and π\pi is the projection to 2\mathbb{R}^{2}. We will show that (E,2,π)(E,\mathbb{R}^{2},\pi) has no nontrivial global sections.

If s:BEs:B\to E is a global section and s(p)s(p_{*}) is on the diagonal of PD(f(,p))\text{PD}(f(\cdot,p_{*})) for some pBp_{*}\in B, then ss is a trivial section because f(σb,p)f(σd,p)f(\sigma_{b},p)\neq f(\sigma_{d},p) for all pp for any (birth, death) simplex pairs (σb,σd)(\sigma_{b},\sigma_{d}) at pp. Therefore, if s:BEs:B\to E is a nontrivial global section, s(p)s(p) is not on the diagonal of PD(f(,p))\text{PD}(f(\cdot,p)) for any pBp\in B.

Suppose that γ:[0,1]E\gamma:[0,1]\to E is a continuous path such that γ(u)\gamma(u) is not on the diagonal of PD(f(,(x,y)))\text{PD}(f(\cdot,(x,y))) for any (x,y)2(x,y)\in\mathbb{R}^{2} and such that

(20) πγ(u)=θ(u):=(cos(2πu+π/4),sin(2πu+π/4))S1.\pi\circ\gamma(u)=\theta(u):=(\cos(2\pi u+\pi/4),\sin(2\pi u+\pi/4))\in S^{1}\,.

That is, πγ\pi\circ\gamma is a parameterization of S1S^{1} that starts in the first quadrant of 2\mathbb{R}^{2} at p0=(2/2,2/2)p_{0}=(\sqrt{2}/2,\sqrt{2}/2). The path γ\gamma is determined uniquely by its initial condition γ(0)\gamma(0). The simplex pairs in the first quadrant are (a,c)(a,c) and (b,d)(b,d), so γ(0)\gamma(0) equals either (p0,(f(a,p0),f(c,p0)))(p_{0},(f(a,p_{0}),f(c,p_{0}))) or (p0,(f(b,p0),f(d,p0)))(p_{0},(f(b,p_{0}),f(d,p_{0}))). In Figure 5, we illustrate the two possibilities for the path γ\gamma. If γ(0)=(p0,(f(a,p0),f(c,p0)))\gamma(0)=(p_{0},(f(a,p_{0}),f(c,p_{0}))), then γ(1)=(p0,(f(b,p0),f(d,p0)))\gamma(1)=(p_{0},(f(b,p_{0}),f(d,p_{0}))); if γ(0)=(p0,(f(b,p0),f(d,p0)))\gamma(0)=(p_{0},(f(b,p_{0}),f(d,p_{0}))), then γ(1)=(p0,(f(a,p0),f(c,p0)))\gamma(1)=(p_{0},(f(a,p_{0}),f(c,p_{0}))). In either case, γ(0)γ(1)\gamma(0)\neq\gamma(1).

Refer to caption
Figure 5. A visualization of the two choices for the path γ:[0,1]E\gamma:[0,1]\to E in the proof of Proposition 5.3, where EE is the total space of the PD bundle. We show 1010 fibers of the PD bundle for various points p2p\in\mathbb{R}^{2}. The first nine PDs (labeled 0 through 88) are PDs for points pS1p\in S^{1}; the kkth PD is the PD at tk=θ(uk)t_{k}=\theta(u_{k}), where uk=k/8u_{k}=k/8 and θ(u)\theta(u) is the parameterization of S1S^{1} given by (20). Note that θ(0)=(2/2,2/2)S1\theta(0)=(\sqrt{2}/2,\sqrt{2}/2)\in S^{1}. The two choices for the path γ(u)\gamma(u), which depend only on the choice of γ(u0)\gamma(u_{0}), are shown in red and blue, respectively. For each kk, the red (respectively, blue) dot in the kkth PD is equal to γ(uk)\gamma(u_{k}) when γ(u0)\gamma(u_{0}) is the red (respectively, blue) point in the 0th PD. Observe that γ(u0)γ(u8)\gamma(u_{0})\neq\gamma(u_{8}) even though p0=p8p_{0}=p_{8}. The unlabeled PD at the origin is the PD for the origin in 2\mathbb{R}^{2}, at which there is a “singularity.”

To obtain a contradiction, suppose that there were a nontrivial global section s:2Es:\mathbb{R}^{2}\to E. Let γ:[0,1]E\gamma:[0,1]\to E be the path γ=sθ\gamma=s\circ\theta, where θ\theta is the parameterization of S1S^{1} defined in (20). Then γ(0)γ(1)\gamma(0)\neq\gamma(1) because γ\gamma is a path satisfying (20), but γ(0)=s(p0)=γ(1)\gamma(0)=s(p_{0})=\gamma(1). ∎

Note that we will use the fibered filtration f:𝒦×2f:\mathcal{K}\times\mathbb{R}^{2}\to\mathbb{R} that was constructed in Proposition 5.3 as a running example throughout Section 6.2.

Remark 5.4.

Even when dim(B)=1\dim(B)=1, it is not guaranteed that a nontrivial global section exists. To see this, consider the 1D PH of the fibered filtration function above restricted to S12S^{1}\subseteq\mathbb{R}^{2}. In this example, dim(B)=1\dim(B)=1 and a nontrivial global section does not exist.

Remark 5.5.

In the example of Proposition 5.3, it was the “singularity” (the point (0,0)2(0,0)\in\mathbb{R}^{2} at which the PD had a point of multiplicity greater than one) that prevented the existence of a nontrivial global section. Restricting the PD bundle to B:=2{(0,0)}B^{\prime}:=\mathbb{R}^{2}\setminus\{(0,0)\} yields a true fiber bundle; each fiber is homeomorphic to the disjoint union of a line (the diagonal) and two points (the off-diagonal points). It is well known that fiber bundles over contractible spaces are trivial (i.e., the total space is homeomorphic to the product of the base and a fiber.) However, BB^{\prime} is not contractible, so our PD bundle restricted to BB^{\prime} is not guaranteed to be trivial. Indeed, what we showed in Proposition 5.3 is that it is not. By comparison to a vineyard,

  1. (1)

    Singularities do not occur for generic fibered filtrations f:𝒦×f:\mathcal{K}\times\mathbb{R}\to\mathbb{R}. A singularity occurs at pBp_{*}\in B when there are two (birth, death) simplex pairs (σb1,σd1)(\sigma_{b}^{1},\sigma_{d}^{1}), (σb2,σd2)(\sigma_{b}^{2},\sigma_{d}^{2}) at pp_{*} such that pI(σb1,σb2)I(σd1,σd2)p_{*}\in I(\sigma_{b}^{1},\sigma_{b}^{2})\cap I(\sigma_{d}^{1},\sigma_{d}^{2}). When dim(B)=1\dim(B)=1, the intersection I(σb1,σb2)I(σd1,σd2)I(\sigma_{b}^{1},\sigma_{b}^{2})\cap I(\sigma_{d}^{1},\sigma_{d}^{2}) is empty in the generic case, so singularities do not typically exist when dim(B)=1\dim(B)=1.

  2. (2)

    Even when singularities do occur in a vineyard, there should not be monodromy in the vineyard. As in the example above, we can remove the singularities from \mathbb{R} to obtain a disjoint union of intervals B1,,BmB_{1},\ldots,B_{m} such that when we restrict the vineyard to a BiB_{i}, we have a fiber bundle. Intervals in \mathbb{R} are contractible, so these fiber bundles must be trivial. By continuity, we can glue together the fiber bundles over each BiB_{i} to see that our PD bundle cannot have monodromy.

6. A Compatible Cellular Sheaf

For a given fibered filtration function that induces a stratification of BB as in Theorem 4.15, we construct a compatible cellular sheaf. We discuss a motivating example in Section 6.1, and give the definition in Section 6.2.

6.1. A motivating example

Again consider the example in the proof of Proposition 5.3, and also again consider the path γ:[0,1]E\gamma:[0,1]\to E that is determined uniquely by the choice of

γ(0){(p0,(f(a,p0),f(c,p0))),(p0,(f(b,p0),f(d,p0)))},\gamma(0)\in\{(p_{0},(f(a,p_{0}),f(c,p_{0}))),(p_{0},(f(b,p_{0}),f(d,p_{0})))\}\,,

where p0=(2/2,2/2)p_{0}=(\sqrt{2}/2,\sqrt{2}/2). The two possibilities for the path γ\gamma are illustrated in Figure 5. For example, if γ(0)=(p0,(f(a,p0),f(c,p0)))\gamma(0)=(p_{0},(f(a,p_{0}),f(c,p_{0}))), then

γ(u)={(θ(u),(f(a,θ(u)),f(c,θ(u)))),u[0,1/8](θ(u),(f(a,θ(u)),f(c,θ(u)))),u[1/8,3/8](θ(u),(f(b,θ(u)),f(c,θ(u)))),u[3/8,5/8](θ(u),(f(b,θ(u)),f(d,θ(u)))),u[5/8,7/8](θ(u),(f(b,θ(u)),f(d,θ(u)))),u[7/8,1],\gamma(u)=\begin{cases}\Big{(}\theta(u),(f(a,\theta(u)),f(c,\theta(u)))\Big{)}\,,&u\in[0,1/8]\\ \Big{(}\theta(u),(f(a,\theta(u)),f(c,\theta(u)))\Big{)}\,,&u\in[1/8,3/8]\\ \Big{(}\theta(u),(f(b,\theta(u)),f(c,\theta(u)))\Big{)}\,,&u\in[3/8,5/8]\\ \Big{(}\theta(u),(f(b,\theta(u)),f(d,\theta(u)))\Big{)}\,,&u\in[5/8,7/8]\\ \Big{(}\theta(u),(f(b,\theta(u)),f(d,\theta(u)))\Big{)}\,,&u\in[7/8,1]\,,\\ \end{cases}

where θ(u)\theta(u) is the parameterization of S1S^{1} given by (20). As we move through the quadrants of 2\mathbb{R}^{2}, the point in the PD that represents the pair (a,c)(a,c) in the first quadrant becomes the point that represents the pair (a,c)(a,c) in the second quadrant, which becomes the point that represents the pair (b,c)(b,c) in the third quadrant, which becomes the point that represents the pair (b,d)(b,d) in the fourth quadrant, which becomes the point that represents the pair (b,d)(b,d) in the first quadrant. One can do a similar analysis for the case in which γ(0)=(p0,(f(b,p0),f(d,p0)))\gamma(0)=(p_{0},(f(b,p_{0}),f(d,p_{0}))).

This analysis yields a bijection of the (birth, death) simplex pairs for any pair of adjacent quadrants. We illustrate the bijections in Figure 6. The bijection between the simplex pairs in a given quadrant and one of its adjacent quadrants is the same as the bijection defined by the update rule of Cohen-Steiner et al. [14] for updating the simplex pairs in a vineyard. A combinatorial perspective on Proposition 5.3 is that there is no consistent way of choosing a simplex pair in each quadrant such that if (σb,σd)(\sigma_{b},\sigma_{d}) is the (birth, death) simplex pair chosen for a given quadrant and (τb,τd)(\tau_{b},\tau_{d}) is the (birth, death) simplex pair chosen for an adjacent quadrant, then (σb,σd)(\sigma_{b},\sigma_{d}) and (τb,τd)(\tau_{b},\tau_{d}) are matched in the bijection between the two quadrants. This is because if we choose an initial simplex pair in one of the quadrants and then walk in a circle through the other quadrants, then the simplex pair at which we finish is different from the initial simplex pair. For example, if we start at (a,c)(a,c) in the first quadrant, then we finish at (b,d)(b,d) when we return to the first quadrant, and vice versa. This is a discrete way of illustrating the non-existence of a nontrivial global section.

Refer to caption
Figure 6. The (birth, death) simplex pairs in each quadrant for the 1D PH of the fibered filtration function in the proof of Proposition 5.3 (see also Figure 4). For each pair of adjacent quadrants, there is a bijection between their sets of simplex pairs; this bijection is equal to the bijection given by the update rule of Cohen-Steiner et al. [14]. The red lines connect simplex pairs that are in bijection with each other.

6.2. Definition of a compatible cellular sheaf

I generalize the discussion in Section 6.1 to fibered filtration functions of the form f:𝒦×Bf:\mathcal{K}\times B\to\mathbb{R} that have a stratification (see Definition 4.1) of BB such that in each stratum YY, the simplex order that is induced by ff is constant. (In other words, there is a strict partial order Y\prec_{Y} on 𝒦\mathcal{K} such that f(,y)\prec_{f(\cdot,y)} is the same as Y\prec_{Y} for all yYy\in Y.) Theorem 4.15 guarantees that such a stratification exists for generic fibered filtration functions, and Proposition 4.5 guarantees that such a stratification exists for all piecewise-linear fibered filtration functions. We denote the set of strata by 𝒴={Yα}αJ\mathcal{Y}=\{Y_{\alpha}\}_{\alpha\in J} for some index set JJ.

Definition 6.1.

Suppose that \mathcal{F} is a Set-valued cellular sheaf whose cell complex, stalks, and morphisms are of the following form:

  1. (1)

    The cell complex: The cell complex on which \mathcal{F} is constructed is the graph GG such that there is a vertex vαv_{\alpha} for each stratum Yα𝒴Y_{\alpha}\in\mathcal{Y} and an edge eβ,α=(vβ,vα)e_{\beta,\alpha}=(v_{\beta},v_{\alpha}) if Yβ𝒴Y_{\beta}\in\mathcal{Y} is a face of YαY_{\alpha}. The 0-cells of the cell complex are the vertices of GG and the 11-cells are the edges of GG.

  2. (2)

    The stalks: Let SαS_{\alpha} denote the set of (birth, death) simplex pairs for a stratum YαY_{\alpha}. The stalk at a 0-cell vαGv_{\alpha}\in G is (vα):=Sα\mathcal{F}(v_{\alpha}):=S_{\alpha}. For a 11-cell eβ,αGe_{\beta,\alpha}\in G, where YβY_{\beta} is a face of YαY_{\alpha}, the stalk at eβ,αe_{\beta,\alpha} is (eβ,α):=Sα\mathcal{F}(e_{\beta,\alpha}):=S_{\alpha}.

  3. (3)

    The morphisms: If Yβ𝒴Y_{\beta}\in\mathcal{Y} is a face of Yα𝒴Y_{\alpha}\in\mathcal{Y}, then the morphism vβeβ,α:(vα)(eβ,α)\mathcal{F}_{v_{\beta}\leq e_{\beta,\alpha}}:\mathcal{F}(v_{\alpha})\to\mathcal{F}(e_{\beta,\alpha}) is the identity map and the morphism vβeβ,α:(vβ)(eβ,α)\mathcal{F}_{v_{\beta}\leq e_{\beta,\alpha}}:\mathcal{F}(v_{\beta})\to\mathcal{F}(e_{\beta,\alpha}) is

    vβeβ,α:=ϕidxβ,idxα,\mathcal{F}_{v_{\beta}\leq e_{\beta,\alpha}}:=\phi^{\text{idx}_{\beta},\,\text{idx}_{\alpha}}\,,

    where ϕidxβ,idxα\phi^{\text{idx}_{\beta},\,\text{idx}_{\alpha}} is of the form in (4) and idxα:𝒦{1,,N}\text{idx}_{\alpha}:\mathcal{K}\to\{1,\ldots,N\} and idxβ:𝒦{1,,N}\text{idx}_{\beta}:\mathcal{K}\to\{1,\ldots,N\} are the simplex indexings (recall Definition 2.2) on YαY_{\alpha} and YβY_{\beta}, respectively. (Recall that by Lemma 4.11, the simplex order induced by ff is constant within YαY_{\alpha} and within YβY_{\beta}.)

Then the cellular sheaf \mathcal{F} is a compatible cellular sheaf for the fibered filtration function f:𝒦×Bf:\mathcal{K}\times B\to\mathbb{R}.

It is not guaranteed that there is a unique compatible cellular sheaf for a given fibered filtration function ff. Although the cell complex (the graph GG) is determined uniquely by ff, the stalks and morphisms are not. Recall from Definition 2.2 that the simplex indexing that is induced by ff may depend on an intrinsic indexing σ1,,σN\sigma_{1},\ldots,\sigma_{N} of the simplices in 𝒦\mathcal{K}. (The intrinsic indexing breaks ties when two simplices have the same filtration value.) For a stratum YαY_{\alpha} such that f(σ,y)=f(τ,y)f(\sigma,y)=f(\tau,y) for all yYαy\in Y_{\alpha} for some pair (σ,τ)(\sigma,\tau) of simplices, the simplex indexing idxf(,Yα)\text{idx}_{f(\cdot,Y_{\alpha})} depends on the intrinsic indexing, so SαS_{\alpha} may not be determined uniquely by ff. If SαS_{\alpha} is not determined uniquely by ff, then for any face YβY_{\beta} of YαY_{\alpha}, the stalks (vα)\mathcal{F}(v_{\alpha}) and (eβ,α)\mathcal{F}(e_{\beta,\alpha}) are not determined uniquely by ff. As discussed in Remark 2.4, a bijection ϕidxβ,idxα\phi^{\text{idx}_{\beta},\text{idx}_{\alpha}} of the form in (4) is not determined uniquely by ff if idxβ\text{idx}_{\beta} and idxα\text{idx}_{\alpha} differ by more than the transposition of two consecutive simplices. Therefore, the morphism vβeβ,α\mathcal{F}_{v_{\beta}\leq e_{\beta,\alpha}} is not necessarily determined uniquely by ff.

However, many aspects of the stalks and morphisms are determined uniquely by ff. Suppose that Yβ𝒴Y_{\beta}\in\mathcal{Y} is a face of Yα𝒴Y_{\alpha}\in\mathcal{Y}. If f(σ,y)f(τ,y)f(\sigma,y)\neq f(\tau,y) for all yy in YαY_{\alpha} and all simplices στ\sigma\neq\tau, then the simplex indexing idxf(,Yα)\text{idx}_{f(\cdot,Y_{\alpha})} is determined uniquely by ff, so the stalks (vα)\mathcal{F}(v_{\alpha}) and (eβ,α)\mathcal{F}(e_{\beta,\alpha}) are determined uniquely by ff. Theorem 4.15 guarantees that this is the generic case when YαY_{\alpha} is an nn-dimensional stratum (where n=dim(B)n=\dim(B)). There are also conditions under which a morphism is determined uniquely by ff. The morphism vαeβ,α:SαSα\mathcal{F}_{v_{\alpha}\leq e_{\beta,\alpha}}:S_{\alpha}\to S_{\alpha} must be the identity map. The morphism vαeβ,α:=ϕidxβ,idxα\mathcal{F}_{v_{\alpha}\leq e_{\beta,\alpha}}:=\phi^{\text{idx}_{\beta},\text{idx}_{\alpha}} is determined uniquely by ff when idxβ\text{idx}_{\beta} and idxα\text{idx}_{\alpha} differ by the transposition of two consecutive simplices. Theorem 4.15 guarantees that this is the generic case when YβY_{\beta} is a “top-dimensional” face of YαY_{\alpha} (i.e., when dim(Yβ)=dim(Yα)1\dim(Y_{\beta})=\dim(Y_{\alpha})-1).

Example 6.2.

Again consider a fibered filtration function f:𝒦×2f:\mathcal{K}\times\mathbb{R}^{2}\to\mathbb{R} of the form defined in Proposition 5.3, with 𝒦\mathcal{K} defined as in Figure 4a with N=11N=11 simplices. We construct a compatible cellular sheaf \mathcal{F} as follows.

  1. (1)

    The cell complex: The strata are the open quadrants Q1,,Q4Q_{1},\ldots,Q_{4}, the open half-axes A12,A23,A34,A14A_{12},A_{23},A_{34},A_{14} with Aij=(QiQj){𝟎}A_{ij}=(\partial Q_{i}\cap\partial Q_{j})\setminus\{\bm{0}\}, and the point 𝟎2\bm{0}\in\mathbb{R}^{2}. The associated graph GG (the cell complex for \mathcal{F}) has a vertex vQiv_{Q_{i}} for the iith quadrant, a vertex vAijv_{A_{ij}} for the (i,j)(i,j)th half-axis, and a vertex v𝟎v_{\bm{0}} for the point 𝟎\bm{0}. The graph GG has edges (vAij,vQi)(v_{A_{ij}},v_{Q_{i}}) and (vAij,vQj)(v_{A_{ij}},v_{Q_{j}}) for each half-axis AijA_{ij}, and it has an edge (v𝟎,v)(v_{\bm{0}},v) for every vertex vGv\in G such that vv𝟎v\neq v_{\bm{0}}.

  2. (2)

    The stalks: We index the simplices of 𝒦\mathcal{K} such that σ8=a\sigma_{8}=a, σ9=b\sigma_{9}=b, σ10=c\sigma_{10}=c, and σ11=d\sigma_{11}=d, where aa, bb, cc, dd are the simplices defined in Figure 4a. The stalk at vQ1v_{Q_{1}} is SQ1={(a,c),(b,d)}S_{Q_{1}}=\{(a,c),(b,d)\}. The vertices vQ2v_{Q_{2}} and vA12v_{A_{12}} have the same stalk {(a,c),(b,d)}\{(a,c),(b,d)\}; the vertices vQ3v_{Q_{3}}, vA23v_{A_{23}}, and v34v_{34} have the same stalk {(a,d),(b,c)}\{(a,d),(b,c)\}; and the vertices vQ4v_{Q_{4}} and vA14v_{A_{14}} have the same stalk {(b,d),(a,c)}\{(b,d),(a,c)\}. The stalks at the edges of GG are determined by the stalks at the vertices. In this example, the stalks at the vertices or edges that correspond to 2D strata are determined uniquely by ff, but the stalks at the vertices and edges that correspond to 0D or 1D strata depend on our choice of intrinsic indexing.

  3. (3)

    The morphisms: There are only three distinct nonidentity morphisms. The first two are

    vA23e(A23,Q2),v𝟎e(𝟎,Q2):{(a,c),(b,d)}\displaystyle\mathcal{F}_{v_{A_{23}}\leq e_{(A_{23},Q_{2})}},\,\mathcal{F}_{v_{\bm{0}}\leq e_{(\bm{0},Q_{2})}}:\{(a,c),(b,d)\} {(a,d),(b,c)}\displaystyle\to\{(a,d),(b,c)\}
    (a,c)\displaystyle(a,c) (b,c)\displaystyle\mapsto(b,c)
    (b,d)\displaystyle(b,d) (a,d),\displaystyle\mapsto(a,d)\,,
    vA34e(A34,Q4),v𝟎e(𝟎,Q4):{(a,d),(b,c)}\displaystyle\mathcal{F}_{v_{A_{34}}\leq e_{(A_{34},Q_{4})}},\,\mathcal{F}_{v_{\bm{0}}\leq e_{(\bm{0},Q_{4})}}:\{(a,d),(b,c)\} {(a,c),(b,d)}\displaystyle\to\{(a,c),(b,d)\}
    (a,d)\displaystyle(a,d) (a,c)\displaystyle\mapsto(a,c)
    (b,c)\displaystyle(b,c) (b,d).\displaystyle\mapsto(b,d)\,.

    The third distinct nonidentity morphism is a map

    v𝟎e(𝟎,Q1):{(a,d),(b,c)}{(a,c),(b,d)}.\mathcal{F}_{v_{\bm{0}}\leq e_{(\bm{0},Q_{1})}}:\{(a,d),(b,c)\}\to\{(a,c),(b,d)\}\,.

    As we move from 𝟎\bm{0} to Q1Q_{1}, we swap the simplex indices of aa and bb and we also swap the simplex indices of cc and dd (in the simplex indexing induced by ff). The morphism is not canonical because the bijection ϕidx𝟎,idxQ1\phi^{\text{idx}_{\bm{0}},\text{idx}_{Q_{1}}} depends on whether one first swaps aa and bb or one first swaps cc and dd. Therefore, we may define either

    v𝟎e(𝟎,Q1):(a,d)\displaystyle\mathcal{F}_{v_{\bm{0}}\leq e_{(\bm{0},Q_{1})}}:(a,d) (a,c),\displaystyle\mapsto(a,c)\,,
    (b,c)\displaystyle(b,c) (b,d)\displaystyle\mapsto(b,d)

    or

    v𝟎e(𝟎,Q1):(a,d)\displaystyle\mathcal{F}_{v_{\bm{0}}\leq e_{(\bm{0},Q_{1})}}:(a,d) (b,d),\displaystyle\mapsto(b,d)\,,
    (b,c)\displaystyle(b,c) (a,c).\displaystyle\mapsto(a,c)\,.

    Both choices results in a compatible cellular sheaf.

6.3. Sections of the cellular sheaf

Let \mathcal{F} be any compatible cellular sheaf for a fibered filtration f:𝒦×Bf:\mathcal{K}\times B\to\mathbb{R}. We write

(21) vβeβ,α=(vβeβ,αb,vβeβ,αd),\mathcal{F}_{v_{\beta}\leq e_{\beta,\alpha}}=\Big{(}\mathcal{F}^{b}_{v_{\beta}\leq e_{\beta,\alpha}},\,\mathcal{F}^{d}_{v_{\beta}\leq e_{\beta,\alpha}}\Big{)}\,,

where vβeβ,αb:SβSα\mathcal{F}^{b}_{v_{\beta}\leq e_{\beta,\alpha}}:S_{\beta}\to S_{\alpha} maps a pair (σb,σd)Sβ(\sigma_{b},\sigma_{d})\in S_{\beta} to the birth simplex of vβeβ,α((σb,σd))\mathcal{F}_{v_{\beta}\leq e_{\beta,\alpha}}((\sigma_{b},\sigma_{d})) and vβeβ,αd:SβSα\mathcal{F}^{d}_{v_{\beta}\leq e_{\beta,\alpha}}:S_{\beta}\to S_{\alpha} maps (σb,σd)Sβ(\sigma_{b},\sigma_{d})\in S_{\beta} to the death simplex of vβeβ,α((σb,σd))\mathcal{F}_{v_{\beta}\leq e_{\beta,\alpha}}((\sigma_{b},\sigma_{d})). (Recall that SαS_{\alpha}, SβS_{\beta} are the stalks at the vertices vαv_{\alpha}, vβv_{\beta} that are associated with the strata YαY_{\alpha}, YβY_{\beta}.)

In this subsection, we show that one can view sections of \mathcal{F} as sections of the associated PD bundle.

Lemma 6.3.

Let YβY_{\beta} be a face of YαY_{\alpha}. Assume that f:𝒦×Bf:\mathcal{K}\times B\to\mathbb{R} is continuous (i.e., f(σ,)f(\sigma,\cdot) is continuous for all simplices σ\sigma in 𝒦\mathcal{K}). Then for any point pYβp\in Y_{\beta} and any pair (σb,σd)(\sigma_{b},\sigma_{d}) in (vβ)\mathcal{F}(v_{\beta}), we have

(22) (f(σb,p),f(σd,p))=(f(vβeβ,αb((σb,σd)),p),f(vβeβ,αd((σb,σd)),p)),\big{(}f(\sigma_{b},p),f(\sigma_{d},p)\Big{)}=\Big{(}f(\mathcal{F}^{b}_{v_{\beta}\leq e_{\beta,\alpha}}((\sigma_{b},\sigma_{d})),p)\,,f(\mathcal{F}^{d}_{v_{\beta}\leq e_{\beta,\alpha}}((\sigma_{b},\sigma_{d})),p)\Big{)}\,,

where b\mathcal{F}^{b} and d\mathcal{F}^{d} are defined as in (21).

Proof.

If the simplex orders in YαY_{\alpha} and YβY_{\beta} differ only by a transposition of simplices (σ,τ)(\sigma,\tau) with consecutive indices in the orderings, then we must have f(σ,p)=f(τ,p)f(\sigma,p)=f(\tau,p) for all pYβp\in Y_{\beta} because ff is continuous and YβYα¯Y_{\beta}\subseteq\overline{Y_{\alpha}}. By definition, vβeβ,α\mathcal{F}_{v_{\beta}\leq e_{\beta,\alpha}} is either the identity map or the map that swaps σ\sigma and τ\tau in the pairs that contain them. In either case, (22) holds because f(σ,p)=f(τ,p)f(\sigma,p)=f(\tau,p) for all pYβp\in Y_{\beta}. Equation (22) holds in general because vβeβ,α\mathcal{F}_{v_{\beta}\leq e_{\beta,\alpha}} is defined as the composition of such maps. ∎

The following proposition says that a global section of a compatible cellular sheaf \mathcal{F} corresponds to a global section of the PD bundle.

Proposition 6.4.

Let z0z_{0} be a non-diagonal point in PDq(fp0)PD_{q}(f_{p_{0}}) for some p0Bp_{0}\in B, let (σb,σd)(\sigma_{b},\sigma_{d}) be the (birth, death) simplex pair such that z0=(f(σb,p0),f(σd,p0))z_{0}=(f(\sigma_{b},p_{0}),f(\sigma_{d},p_{0})), and let Y0Y_{0} be the stratum that contains p0p_{0}. Suppose that \mathcal{F} is a compatible cellular sheaf, and let v0v_{0} be the vertex in the graph GG (see Definition 6.1) that is associated with Y0Y_{0}. If there is a global section s¯\overline{s} of the cellular sheaf \mathcal{F} such that s¯(v0)=(σb,σd)\overline{s}(v_{0})=(\sigma_{b},\sigma_{d}), then there is a global section ss of the PD bundle such that s(p0)=z0s(p_{0})=z_{0}.

Proof.

Let s¯\overline{s} be a global section of the cellular sheaf \mathcal{F} such that s¯(v0)=(σb,σd)\overline{s}(v_{0})=(\sigma_{b},\sigma_{d}). For every stratum YαY_{\alpha}, we write

s¯(vα)=(s¯b(vα),s¯d(vα)),\overline{s}(v_{\alpha})=(\overline{s}_{b}(v_{\alpha}),\overline{s}_{d}(v_{\alpha}))\,,

where s¯b(vα)\overline{s}_{b}(v_{\alpha}) is the birth simplex of s¯(vα)\overline{s}(v_{\alpha}) and s¯d(vα)\overline{s}_{d}(v_{\alpha}) is the death simplex of s¯(vα)\overline{s}(v_{\alpha}). Let Y:B{Yα}Y:B\to\{Y_{\alpha}\} be the function that maps pBp\in B to the unique stratum YαY_{\alpha} that contains it.

We define s:BEs:B\to E to be the function

s(p):=(p,f(s¯b(vY(p)),p),f(s¯d(vY(p)),p)).s(p):=\Big{(}p,f(\overline{s}_{b}(v_{Y(p)}),p),f(\overline{s}_{d}(v_{Y(p)}),p)\Big{)}\,.

To show that s:BEs:B\to E is a global section of the PD bundle, it remains to show that it is continuous. The function s|Yαs|_{Y_{\alpha}} is continuous for all strata YαY_{\alpha} because f(σ,)f(\sigma,\cdot) is continuous for all simplices σ𝒦\sigma\in\mathcal{K}. Therefore, it suffices to show that s|Yα¯s|_{\overline{Y_{\alpha}}} is continuous on each face YβY_{\beta} of YαY_{\alpha}. Because s¯\overline{s} is a section of the cellular sheaf,

s¯(vα)=vβeβ,α(s¯(vβ)).\overline{s}(v_{\alpha})=\mathcal{F}_{v_{\beta}\leq e_{\beta,\alpha}}(\overline{s}(v_{\beta}))\,.

By Lemma 6.3,

(f(s¯b(vβ),p),f(s¯d(vβ),p))=(f(vβeβ,αb(s¯(vβ)),p),f(vβeβ,αd(s¯(vβ)),p))\Big{(}f(\overline{s}_{b}(v_{\beta}),p),f(\overline{s}_{d}(v_{\beta}),p)\Big{)}=\Big{(}f(\mathcal{F}^{b}_{v_{\beta}\leq e_{\beta,\alpha}}(\overline{s}(v_{\beta})),p),f(\mathcal{F}^{d}_{v_{\beta}\leq e_{\beta,\alpha}}(\overline{s}(v_{\beta})),p)\Big{)}

for all points pYβp\in Y_{\beta}. Therefore,

(f(s¯b(vβ),p),f(s¯d(vβ),p))=(f(s¯b(vα),p),f(s¯d(vα),p))\Big{(}f(\overline{s}_{b}(v_{\beta}),p),f(\overline{s}_{d}(v_{\beta}),p)\Big{)}=\Big{(}f(\overline{s}_{b}(v_{\alpha}),p),f(\overline{s}_{d}(v_{\alpha}),p)\Big{)}

for all pYβp\in Y_{\beta}, which completes the proof. ∎

7. Conclusions

7.1. Summary

In this paper, I introduced the concept of a persistence diagram (PD) bundle, a framework that can be used to study the persistent homology of a fibered filtration function (i.e., set of filtrations parameterized by an arbitrary “base space” BB). Special cases of PD bundles include vineyards [14], the persistent homology transform (PHT) [32], the fibered barcode of a multiparameter persistence module [4], and the barcode-decorated merge tree [17].

In Theorem 4.15, I proved that if BB is a smooth compact manifold, then for generic fibered filtrations, BB is stratified so that the simplex order is constant within each stratum. When such a stratification exists, the PD bundle is determined by the PDs at a locally finite (or finite, if BB is compact) subset of points in BB. In Proposition 4.5, I showed that every piecewise-linear PD bundle has such a stratification into polyhedra. This polyhedral stratification is utilized in [23] in an algorithm for computing piecewise-linear PD bundles.

I showed that, unlike vineyards, which PD bundles generalize, not every local section of a PD bundle can be extended to a global section (see Proposition 5.3). The implication is that PD bundles do not necessarily decompose into “vines” in the way that vineyards do (see (3)).

Lastly, I introduced a cellular sheaf that is compatible with a given PD bundle. In Proposition 6.4, I proved that one can determine whether a local section can be extended to a global section by determining whether or not there is an associated global section of a compatible cellular sheaf. A compatible cellular sheaf is a discrete mathematical data structure for summarizing the data in a PD bundle.

7.2. Discussion

For a given fibered filtration function ff with a stratification as in Theorem 4.15, I defined a compatible cellular sheaf \mathcal{F} over a graph GG. It is tempting to instead define an associated cellular sheaf directly on the stratification of BB. In particular, when ff is piecewise linear, the strata are polyhedra, so the stratification is guaranteed to be a cellular decomposition. One could certainly define stalks (Yα)\mathcal{F}(Y_{\alpha}) and functions (Yβ)(Yβ)\mathcal{F}(Y_{\beta})\to\mathcal{F}(Y_{\beta}) in the same way as in Definition 6.1. The problem is that \mathcal{F} would not necessarily satisfy the composition condition (see (5)). For instance, this issue occurs in Example 6.2 for the same reason that the morphism v𝟎e(𝟎,Q1)\mathcal{F}_{v_{\bm{0}}\leq e_{(\bm{0},Q_{1})}} in the example is not canonical (see the discussion in Example 6.2).

Additionally, I note that one could have defined a compatible cellular cosheaf rather than a sheaf.

7.3. Future research

I conclude with some questions and proposals for future work:

  • What are the conditions under which a PD bundle must have a decomposition of the form (3)?

    I conjecture that if BBB\setminus B^{*} is contractible, where BB^{*} is the set of singularities (i.e., points pBp_{*}\in B at which there is a point in PD(fp)(f_{p_{*}}) with multiplicity greater than one), then there is a decomposition of the form (3). See the discussion in Remark 5.5.

  • What algebraic or computational methods can we use to analyze global sections and to compute obstructions to the existence of global sections?

    It may help to consider the cellular-sheaf perspective from Section 6.2, which turns the question into a discrete problem that one can study computationally. One can also generalize Turner’s vineyard-module perspective [33].

  • PHT is a PD bundle over the base space B=SnB=S^{n}. Are there constructible sets Mn+1M\subseteq\mathbb{R}^{n+1} for which the associated PHT exhibits monodromy? What is the geometric interpretation (in terms of MM)?

  • Arya et al. [5] showed that the PHT of a constructible set MM can be calculated by “gluing together” the PHT of smaller, simpler subsets of MM. Can one generalize these results to all PD bundles?

  • When are PD bundles “stable”?

    PD bundles are “fiberwise stable” in the sense that if f1,f2:𝒦×Bf_{1},f_{2}:\mathcal{K}\times B\to\mathbb{R} are two fibered filtrations, then the bottleneck distance between PD(f1(,p))\text{PD}(f_{1}(\cdot,p)) and PD(f2(,p))\text{PD}(f_{2}(\cdot,p)) is bounded above by f1f2\lVert f_{1}-f_{2}\rVert_{\infty} for all pBp\in B [13]. However, this does not guarantee that the global structure of a PD bundle is stable. For example, it is well known that the structure of a vineyard is not stable (see Appendix A.4 for an example). However, vineyards are stable for generic 11-parameter filtrations; if none of the vines intersect (which is the generic case), then sufficiently small perturbations of the filtration result in small perturbations of each vine in the vine decomposition (see (3)). I expect that an analogous result holds for generic fibered filtration functions over any base space BB.

  • It will also be interesting to study real-world applications of PD bundles, such as the examples that were mentioned in Section 3.1.

Acknowledgements

I am very grateful for discussions with Andrew Blumberg, which led to the investigation of monodromy in PD bundles. I also thank Ryan Grady and Karthik Viswanathan for helpful discussions.

Appendix

A.4. Vineyard instability

For any ϵ>0\epsilon>0, we construct two 11-parameter filtration functions fϵ+f^{+}_{\epsilon}, fϵf^{-}_{\epsilon} that are ϵ\epsilon-perturbations of each other (that is, |fϵ+(σ,p)fϵ(σ,p)|<ϵ|f^{+}_{\epsilon}(\sigma,p)-f^{-}_{\epsilon}(\sigma,p)|<\epsilon for all simplices σ𝒦\sigma\in\mathcal{K} and all points pBp\in B) but such that for any bijection between the vines in the respective vineyards, not all of the matched vines are close to each other. In fact, we can define fϵ+f^{+}_{\epsilon} and fϵf^{-}_{\epsilon} so that their vines are arbitrarily far apart.

We construct our example by restricting the filtration function from Section 5 to certain paths through 2\mathbb{R}^{2}. Let 𝒦\mathcal{K} and f:𝒦×2f:\mathcal{K}\times\mathbb{R}^{2}\to\mathbb{R} be defined as in the proof of Proposition 5.3. (See Figure 4b.) Because ff is continuous, we have that for any ϵ>0\epsilon>0, there is a δ>0\delta>0 such that |f(σ,p)f(σ,𝟎)|<ϵ/2|f(\sigma,p)-f(\sigma,\bm{0})|<\epsilon/2 when p<2δ\lVert p\rVert<\sqrt{2}\delta and σ\sigma is any simplex in 𝒦\mathcal{K}. We define the paths

γϵ+(t)\displaystyle\gamma^{+}_{\epsilon}(t) :={(t,t),|t|δ(δ,δ+2t),δ<t<0(δ+2t,δ),0t<δ,\displaystyle:=\begin{cases}(t,t)\,,&|t|\geq\delta\\ (-\delta,\delta+2t)\,,&-\delta<t<0\\ \ (-\delta+2t,\delta)\,,&0\leq t<\delta\,,\end{cases}
γϵ(t)\displaystyle\gamma_{\epsilon}^{-}(t) :={(t,t),|t|δ(δ+2t,δ),δ<t<0(δ,δ+2t),0t<δ.\displaystyle:=\begin{cases}(t,t)\,,&|t|\geq\delta\\ (\delta+2t,-\delta)\,,&-\delta<t<0\\ (\delta,-\delta+2t)\,,&0\leq t<\delta\,.\end{cases}

See Figure 7 for a plot of the paths γϵ±(t)\gamma_{\epsilon}^{\pm}(t).

Refer to caption
Figure 7. The paths γϵ+(t)\gamma_{\epsilon}^{+}(t) and γϵ(t)\gamma_{\epsilon}^{-}(t).

Let fϵ±:𝒦×f^{\pm}_{\epsilon}:\mathcal{K}\times\mathbb{R}\to\mathbb{R} be the 11-parameter filtration functions defined by f±(σ,t):=f(σ,γϵ±(t))f^{\pm}(\sigma,t):=f(\sigma,\gamma_{\epsilon}^{\pm}(t)). By construction, the filtrations fϵ+f^{+}_{\epsilon} and fϵf^{-}_{\epsilon} are ϵ\epsilon-perturbations of each other.

Let V+V^{+} and VV^{-} be the vineyards for fϵ+f^{+}_{\epsilon} and fϵf^{-}_{\epsilon}, respectively, for the 11st degree PH. The vineyards V±V^{\pm} each have two vines v1±v_{1}^{\pm}, v2±v_{2}^{\pm}, which are paths vi±:3v_{i}^{\pm}:\mathbb{R}\to\mathbb{R}^{3}. The vines are

v1+(t)\displaystyle v^{+}_{1}(t) ={(f(a,γ+(t)),f(d,γ+(t))),tδ/2f(b,γ+(t)),f(d,γ+(t))),t>δ/2,\displaystyle=\begin{cases}(f(a,\gamma^{+}(t)),f(d,\gamma^{+}(t)))\,,&t\leq-\delta/2\\ f(b,\gamma^{+}(t)),f(d,\gamma^{+}(t)))\,,&t>-\delta/2\,,\end{cases}
v2+(t)\displaystyle v^{+}_{2}(t) ={(f(b,γ+(t)),f(c,γ+(t))),tδ/2f(a,γ+(t)),f(c,γ+(t))),t>δ/2,\displaystyle=\begin{cases}(f(b,\gamma^{+}(t)),f(c,\gamma^{+}(t)))\,,&t\leq-\delta/2\\ f(a,\gamma^{+}(t)),f(c,\gamma^{+}(t)))\,,&t>-\delta/2\,,\end{cases}
v1(t)\displaystyle v^{-}_{1}(t) ={(f(a,γ(t)),f(d,γ(t))),tδ/2f(a,γ(t)),f(c,γ(t))),t>δ/2,\displaystyle=\begin{cases}(f(a,\gamma^{-}(t)),f(d,\gamma^{-}(t)))\,,&t\leq-\delta/2\\ f(a,\gamma^{-}(t)),f(c,\gamma^{-}(t)))\,,&t>-\delta/2\,,\end{cases}
v2(t)\displaystyle v^{-}_{2}(t) ={(f(b,γ(t)),f(c,γ(t))),tδ/2f(b,γ(t)),f(d,γ(t))),t>δ/2.\displaystyle=\begin{cases}(f(b,\gamma^{-}(t)),f(c,\gamma^{-}(t)))\,,&t\leq-\delta/2\\ f(b,\gamma^{-}(t)),f(d,\gamma^{-}(t)))\,,&t>-\delta/2\,.\end{cases}

There is no bijection ϕ:{1,2}{1,2}\phi:\{1,2\}\to\{1,2\} such that v1+v_{1}^{+} and v2+v_{2}^{+} are close to vϕ(1)v_{\phi(1)}^{-} and vϕ(2)v_{\phi(2)}^{-}, respectively. This is because

v1+(t)v1(t)2\displaystyle\lVert v_{1}^{+}(t)-v_{1}^{-}(t)\rVert^{2} =|f(b,(t,t))f(a,(t,t))|2+|f(d,(t,t)))f(c,(t,t))|2,t>δ/2,\displaystyle=|f(b,(t,t))-f(a,(t,t))|^{2}+|f(d,(t,t)))-f(c,(t,t))|^{2}\,,\qquad t>-\delta/2\,,
v1+(t)v2(t)2\displaystyle\lVert v_{1}^{+}(t)-v_{2}^{-}(t)\rVert^{2} =|f(b,(t,t))f(a,(t,t))|2+|f(d,(t,t)))f(c,(t,t))|2,tδ/2,\displaystyle=|f(b,(t,t))-f(a,(t,t))|^{2}+|f(d,(t,t)))-f(c,(t,t))|^{2}\,,\qquad t\leq-\delta/2\,,
v2+(t)v2(t)2\displaystyle\lVert v_{2}^{+}(t)-v_{2}^{-}(t)\rVert^{2} =|f(b,(t,t))f(a,(t,t))|2+|f(d,(t,t)))f(c,(t,t))|2,t>δ/2,\displaystyle=|f(b,(t,t))-f(a,(t,t))|^{2}+|f(d,(t,t)))-f(c,(t,t))|^{2}\,,\qquad t>-\delta/2\,,
v2+(t)v1(t)2\displaystyle\lVert v_{2}^{+}(t)-v_{1}^{-}(t)\rVert^{2} =|f(b,(t,t))f(a,(t,t))|2+|f(d,(t,t)))f(c,(t,t))|2,tδ/2\displaystyle=|f(b,(t,t))-f(a,(t,t))|^{2}+|f(d,(t,t)))-f(c,(t,t))|^{2}\,,\qquad t\leq-\delta/2

and we can define ff so that |f(b,(t,t))f(a,(t,t))||f(b,(t,t))-f(a,(t,t))| and |f(d,(t,t))f(c,(t,t))||f(d,(t,t))-f(c,(t,t))| are arbitrarily large for t0t\neq 0.

A.5. Technical Details of Section 4

All notation is defined as in Section 4.

The first series of lemmas is used to prove Lemma 4.12, which shows that for almost every 𝒂A\bm{a}\in A, the tangent space of the intersection of sets I𝒂(σir,σjr)I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) is equal to the intersection of their tangent spaces.

Lemma A.1.

For almost every 𝒂A\bm{a}\in A, we have

Tp(jJM𝒂,j)=jJTp(M𝒂,j)T_{p}\Big{(}\bigcap_{j\in J}M_{\bm{a},j}\Big{)}=\bigcap_{j\in J}T_{p}(M_{\bm{a},j})

for all J{1,,N}J\subseteq\{1,\ldots,N\} and all pjJM𝒂,jp\in\bigcap_{j\in J}M_{\bm{a},j}.

Proof.

Because there are finitely many subsets of {1,,N}\{1,\ldots,N\}, it suffices to show that for a given J{1,,N}J\subseteq\{1,\ldots,N\}, we have Tp(jJM𝒂,j)=jJTp(M𝒂,j)T_{p}\Big{(}\bigcap_{j\in J}M_{\bm{a},j}\Big{)}=\bigcap_{j\in J}T_{p}(M_{\bm{a},j}) for all pjJM𝒂,jp\in\bigcap_{j\in J}M_{\bm{a},j} for almost every 𝒂A\bm{a}\in A. Let {ji}i=1k\{j_{i}\}_{i=1}^{k} be the elements of JJ, where ji<ji+1j_{i}<j_{i+1} for all ii. Because transverse intersections are generic, we have M𝒂,ji(M𝒂,j1M𝒂,ji1)M_{\bm{a},j_{i}}\pitchfork(M_{\bm{a},j_{1}}\cap\cdots\cap M_{\bm{a},j_{i-1}}) for every ii for almost every 𝒂A\bm{a}\in A. For such an 𝒂A\bm{a}\in A, we have

Tp(jJM𝒂,j)=Tp(M𝒂,jk)Tp(i=1k1M𝒂,ji)T_{p}\Big{(}\bigcap_{j\in J}M_{\bm{a},j}\Big{)}=T_{p}(M_{\bm{a},j_{k}})\cap T_{p}\Big{(}\bigcap_{i=1}^{k-1}M_{\bm{a},j_{i}}\Big{)}

because M𝒂,jk(M𝒂,j1M𝒂,jk1)M_{\bm{a},j_{k}}\pitchfork(M_{\bm{a},j_{1}}\cap\cdots\cap M_{\bm{a},j_{k-1}}). Therefore,

Tp(jJM𝒂,j)=i=1kTp(M𝒂,ji)T_{p}\Big{(}\bigcap_{j\in J}M_{\bm{a},j}\Big{)}=\bigcap_{i=1}^{k}T_{p}(M_{\bm{a},j_{i}})

by induction on ii. ∎

Lemma A.2.

Let 𝒂A\bm{a}\in A, and let {(ir,jr)}r=1m\{(i_{r},j_{r})\}_{r=1}^{m} be a set of index pairs such that {ir,jr}{is,js}\{i_{r},j_{r}\}\neq\{i_{s},j_{s}\} if rsr\neq s. If BB is a compact manifold, then there is a finite open cover {Uk}k=1K\{U_{k}\}_{k=1}^{K} and a disjoint partition J,k={1,,m}\bigcup_{\ell}J_{\ell,k}=\{1,\ldots,m\} for each kk such that

{ir,jrrJ1,k}{ir,jrrJ2,k}=\{i_{r},j_{r}\mid r\in J_{\ell_{1},k}\}\cap\{i_{r},j_{r}\mid r\in J_{\ell_{2},k}\}=\emptyset

if 12\ell_{1}\neq\ell_{2} and

π(rJ,k(M𝒂,irM𝒂,jr))Uk=rJ,kI𝒂(σir,σjr)Uk\pi\Big{(}\bigcap_{r\in J_{\ell,k}}(M_{\bm{a},i_{r}}\cap M_{\bm{a},j_{r}})\Big{)}\cap U_{k}=\bigcap_{r\in J_{\ell,k}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U_{k}

for all \ell, where π\pi is the projection π:B×B\pi:B\times\mathbb{R}\to B.

Proof.

Suppose that yr=1mI𝒂(σir,σjr)y\in\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}). Let J0:={}J^{0}_{\ell}:=\{\ell\} be an initial disjoint partition of {1,,m}\{1,\ldots,m\}. By definition, π(M𝒂,iM𝒂,j)=I𝒂(σi,σj)\pi(M_{\bm{a},i_{\ell}}\cap M_{\bm{a},j_{\ell}})=I_{\bm{a}}(\sigma_{i_{\ell}},\sigma_{j_{\ell}}). If i1=i2i_{\ell_{1}}=i_{\ell_{2}} for some 12\ell_{1}\neq\ell_{2}, then f(σj1,y)=f(σi1,y)=f(σj2,y)f(\sigma_{j_{\ell_{1}}},y)=f(\sigma_{i_{\ell_{1}}},y)=f(\sigma_{j_{\ell_{2}}},y), so yπ(M𝒂,i1M𝒂,j1M𝒂,i2M𝒂,j2)y\in\pi(M_{{\bm{a}},i_{\ell_{1}}}\cap M_{\bm{a},j_{\ell_{1}}}\cap M_{\bm{a},i_{\ell_{2}}}\cap M_{\bm{a},j_{\ell_{2}}}). We combine J10J^{0}_{\ell_{1}} and J20J^{0}_{\ell_{2}} into a single subset of the partition, and we iterate until we obtain a disjoint partition {J,y}\{J_{\ell,y}\}_{\ell} of {1,,m}\{1,\ldots,m\} such that

{ir,jrrJ1,y}{ir,jrrJ2,y}=\{i_{r},j_{r}\mid r\in J_{\ell_{1},y}\}\cap\{i_{r},j_{r}\mid r\in J_{\ell_{2},y}\}=\emptyset

if 12\ell_{1}\neq\ell_{2} and

yπ(rJ,y(M𝒂,irM𝒂,jr))y\in\pi\Big{(}\bigcap_{r\in J_{\ell,y}}(M_{\bm{a},i_{r}}\cap M_{\bm{a},j_{r}})\Big{)}

for all \ell. Therefore, for each \ell, there is a neighborhood U,yU_{\ell,y} such that

π(rJ,y(M𝒂,irM𝒂,jr))U,y=rJ,yI𝒂(σir,σjr)U,k.\pi\Big{(}\bigcap_{r\in J_{\ell,y}}(M_{\bm{a},i_{r}}\cap M_{\bm{a},j_{r}})\Big{)}\cap U_{\ell,y}=\bigcap_{r\in J_{\ell,y}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U_{\ell,k}\,.

Set Uy:=U,yU_{y}:=\bigcap_{\ell}U_{\ell,y}. Because BB is compact, there is a finite open cover {Uk}k=1K\{U_{k}\}_{k=1}^{K}, which has the desired properties by construction. ∎

The following lemma will be repeatedly used in Lemma A.4.

Lemma A.3.

Suppose that g:Bg:B\to\mathbb{R} is a smooth map and yy\in\mathbb{R} is a regular value with preimage ZyZ_{y}. If ZBZ\subseteq B is a submanifold such that ZZyZ\pitchfork Z_{y}, then yy is a regular value of g|Z:Zg|_{Z}:Z\to\mathbb{R}.

Proof.

At any zZyz\in Z_{y}, we have ker(dgz)=TzZy\ker(dg_{z})=T_{z}Z_{y}. Therefore, if zZZyz\in Z\cap Z_{y}, then TzZker(dgz)T_{z}Z\subseteq\ker(dg_{z}) only if TzZTzZyT_{z}Z\subseteq T_{z}Z_{y}. Because yy is a regular value of g:Bg:B\to\mathbb{R}, we have

dim(Tz(Zy))=dimB1,\dim(T_{z}(Z_{y}))=\dim B-1\,,

so Tz(Zy)T_{z}(Z_{y}) is a strict subset of TzBT_{z}B. Because ZZyZ\pitchfork Z_{y}, we have

TzZ+TzZy=TzB,T_{z}Z+T_{z}Z_{y}=T_{z}B\,,

so TzZT_{z}Z cannot be a subset of TzZyT_{z}Z_{y}. Therefore, TzZT_{z}Z is not a subset of ker(dgz)\ker(dg_{z}). This implies that dgz|TzZdg_{z}|_{T_{z}Z} is a surjection because dim=1\dim\mathbb{R}=1. Therefore, yy is a regular value of g|Zg|_{Z}. ∎

Lemma A.4.

Let {(ir,jr)}r=1m\{(i_{r},j_{r})\}_{r=1}^{m} be a set of index pairs such that {ir,jr}{is,js}\{i_{r},j_{r}\}\neq\{i_{s},j_{s}\} if rsr\neq s. For almost every 𝒂A\bm{a}\in A, we have that if

  1. (1)

    J={1,,m}\bigcup_{\ell}J_{\ell}=\{1,\ldots,m\} is a disjoint partition with

    {ir,jrrJ1}{ir,jrrJ2}=\{i_{r},j_{r}\mid r\in J_{\ell_{1}}\}\cap\{i_{r},j_{r}\mid r\in J_{\ell_{2}}\}=\emptyset

    for 12\ell_{1}\neq\ell_{2} and

  2. (2)

    UU is an open set in BB such that

    (23) π(rJ(M𝒂,irM𝒂,jr))U=rJI𝒂(σir,σjr)U,\pi\Big{(}\bigcap_{r\in J_{\ell}}(M_{\bm{a},i_{r}}\cap M_{\bm{a},j_{r}})\Big{)}\cap U=\bigcap_{r\in J_{\ell}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U\,,

    for all \ell, where π\pi is the projection π:B×B\pi:B\times\mathbb{R}\to B,

then

  1. (1)

    the set rJI𝒂(σir,σjr)\bigcap_{r\in J^{\prime}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) is a manifold for every J{1,,m}J^{\prime}\subseteq\{1,\ldots,m\} and

  2. (2)

    we have

    (24) Ty(rJI𝒂(σir,σjr))=Ty(rJI𝒂(σir,σjr))T_{y}\Big{(}\bigcap_{\ell}\bigcap_{r\in J_{\ell}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}=\bigcap_{\ell}T_{y}\Big{(}\bigcap_{r\in J_{\ell}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}

    for every yr=1mI𝒂(σir,σjr)Uy\in\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U.

Proof.

It suffices to show that

(25) (rJI𝒂(σir,σjr)U)(<rJI𝒂(σir,σjr)U)\Big{(}\bigcap_{r\in J_{\ell}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U\Big{)}\pitchfork\Big{(}\bigcap_{\ell^{\prime}<\ell}\bigcap_{r\in J_{\ell}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U\Big{)}

for all \ell and almost every 𝒂A\bm{a}\in A. Informally, what we show first is that at almost every 𝒂A\bm{a}\in A, perturbations of 𝒂\bm{a} produce perturbations of rJI𝒂(σir,σjr)U\bigcap_{r\in J_{\ell}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U for each \ell. Then at the end of the proof, we apply the fact that transverse intersections are generic.

By Lemmas 4.8 and 4.9, we may assume that rJI𝒂(σir,σjr)\bigcap_{r\in J^{\prime}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) is a manifold for every J{1,,m}J^{\prime}\subseteq\{1,\ldots,m\}. By (23), we may assume without loss of generality that there is a sequence k1<<kck_{1}<\cdots<k_{c} such that j1=k1j_{1}=k_{1} and ir=jr1i_{r}=j_{r-1} for all rr and jr+1=irj_{r+1}=i_{r} for all rr. In other words, we may assume that rJI𝒂(σir,σjr)\bigcap_{r\in J_{\ell}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) is of the form

I𝒂(σkc,σkc1)I𝒂(σk3,σk2)I𝒂(σk2,σk1)I_{\bm{a}}(\sigma_{k_{c}},\sigma_{k_{c-1}})\cap\cdots\cap I_{\bm{a}}(\sigma_{k_{3}},\sigma_{k_{2}})\cap I_{\bm{a}}(\sigma_{k_{2}},\sigma_{k_{1}})

for all 𝒂\bm{a}. The idea is that because the intersection lifts to an intersection of the corresponding manifolds (see (23)), we can pair up the indices however we like.

Define the function gi:Bg_{i}:B\to\mathbb{R} by

gi(p):=f(σki,p)f(σki1,p).g_{i}(p):=f(\sigma_{k_{i}},p)-f(\sigma_{k_{i-1}},p)\,.

For almost every 𝒂A\bm{a}\in A, the quantity akiaki1a_{k_{i}}-a_{k_{i-1}} is a regular value of gig_{i} for all ii, and the set of regular values is open. By the same argument as in the proof of Lemma 4.9, we have

(I𝒂(σkc,σkc1)U)(i=2c1I𝒂(σki,σki1)U)\Big{(}I_{\bm{a}}(\sigma_{k_{c}},\sigma_{k_{c-1}})\cap U\Big{)}\pitchfork\Big{(}\bigcap_{i=2}^{c-1}I_{\bm{a}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\cap U\Big{)}

for almost every 𝒂A\bm{a}\in A. Therefore, (akcakc1)(a_{k_{c}}-a_{k_{c-1}}) is a regular value of gkc|i=2c1I𝒂(σki,σki1)Ug_{k_{c}}|_{\bigcap_{i=2}^{c-1}I_{\bm{a}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\cap U} by Lemma A.3. Additionally, for ϵN\bm{\epsilon}\in\mathbb{R}^{N} such that ϵkc\epsilon_{k_{c}} and ϵkc1\epsilon_{k_{c-1}} are sufficiently small, there are no critical values between (akcakc1)(a_{k_{c}}-a_{k_{c-1}}) and (akcakc1+ϵkcϵkc1)(a_{k_{c}}-a_{k_{c-1}}+\epsilon_{k_{c}}-\epsilon_{k_{c-1}}). Because there are no critical values, the set rJI𝒂(σir,σjr)U\bigcap_{r\in J_{\ell}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U (which is the (akcakc1)(a_{k_{c}}-a_{k_{c-1}})-level set of gk|i=2c1I𝒂(σki,σki1)Ug_{k}|_{\bigcap_{i=2}^{c-1}I_{\bm{a}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\cap U}) is a submanifold of BB that is diffeomorphic to I𝒂+ϵ(σkc,σkc1)(i=2c1I𝒂(σki,σki1))UI_{\bm{a}+\bm{\epsilon}}(\sigma_{k_{c}},\sigma_{k_{c-1}})\cap\Big{(}\bigcap_{i=2}^{c-1}I_{\bm{a}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\Big{)}\cap U (which is the (akcakc1+ϵkcϵkc1)(a_{k_{c}}-a_{k_{c-1}}+\epsilon_{k_{c}}-\epsilon_{k_{c-1}})-level set of gk|i=2c1I𝒂(σki,σki1)Ug_{k}|_{\bigcap_{i=2}^{c-1}I_{\bm{a}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\cap U}), and these submanifolds are smoothly parameterized by ϵkc,ϵkc1\epsilon_{k_{c}},\epsilon_{k_{c-1}}.

Now consider any i{2,,c1}i_{*}\in\{2,\ldots,c-1\}. By induction on ii_{*}, we will show that there is a set AAA^{\prime}\subseteq A such that AAA\setminus A^{\prime} has measure zero and such that for all 𝒂A\bm{a}\in A^{\prime}, we have that

  1. (1)

    the set rJI𝒂+ϵ(σir,σjr)U\bigcap_{r\in J_{\ell}}I_{\bm{a}+\bm{\epsilon}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U is a submanifold of BB that is diffeomorphic to

    rJI𝒂(σir,σjr)U\bigcap_{r\in J_{\ell}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U

    for sufficiently small ϵN\bm{\epsilon}\in\mathbb{R}^{N}, and

  2. (2)

    these submanifolds are smoothly parameterized by ϵ\bm{\epsilon}.

Because (akiaki1)(a_{k_{i_{*}}}-a_{k_{i_{*}-1}}) is a regular value of gig_{i_{*}} and the set of regular values is open, there are no critical values between (akiaki1)(a_{k_{i_{*}}}-a_{k_{i_{*}-1}}) and (akiaki1ϵki1)(a_{k_{i_{*}}}-a_{k_{i_{*}-1}}-\epsilon_{k_{i_{*}-1}}) for sufficiently small ϵki1\epsilon_{k_{i_{*}-1}}. Therefore, for sufficiently small ϵki1\epsilon_{k_{i_{*}-1}}, the (akiaki1ϵki1)(a_{k_{i_{*}}}-a_{k_{i_{*}-1}}-\epsilon_{k_{i_{*}-1}})-level set of gig_{i_{*}} is a submanifold of BB that is diffeomorphic to the (akiaki1)(a_{k_{i_{*}}}-a_{k_{i_{*}-1}})-level set, and these submanifolds are smoothly parameterized by (sufficiently small) ϵki1\epsilon_{k_{i_{*}-1}}. Because transverse intersections are generic,

(i=i+1cI𝒂+ϵ(σki,σki1))(i=2i1I𝒂+ϵ(σki,σki1))U\Big{(}\bigcap_{i=i_{*}+1}^{c}I_{\bm{a+\epsilon}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\Big{)}\cap\Big{(}\bigcap_{i=2}^{i_{*}-1}I_{\bm{a+\epsilon}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\Big{)}\cap U

is transverse to the (akiaki1ϵki1)(a_{k_{i_{*}}}-a_{k_{i_{*}-1}}-\epsilon_{k_{i_{*}-1}})-level set of gig_{i_{*}} for almost every (sufficiently small) ϵki1\epsilon_{k_{i_{*}-1}}. Additionally, if the intersection is transverse, it is transverse for an open neighborhood of ϵki1\epsilon_{k_{i_{*}-1}}. Therefore, we can assume without loss of generality that this intersection is transverse at ϵki1=0\epsilon_{k_{i_{*}-1}}=0 (if not, we can perturb aki1a_{k_{i_{*}-1}} so that it is) and for all sufficiently small ϵki1\epsilon_{k_{i_{*}-1}}. This implies that (akiaki1ϵki1)(a_{k_{i_{*}}}-a_{k_{i_{*}-1}}-\epsilon_{k_{i_{*}-1}}) is also a regular value of gig_{i_{*}} restricted to (i=i+1cI𝒂+ϵ(σki,σki1))(i=2i1I𝒂+ϵ(σki,σki1))U\Big{(}\bigcap_{i=i_{*}+1}^{c}I_{\bm{a+\epsilon}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\Big{)}\cap\Big{(}\bigcap_{i=2}^{i_{*}-1}I_{\bm{a+\epsilon}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\Big{)}\cap U, by Lemma A.3. For sufficiently small ϵki\epsilon_{k_{i_{*}}}, there are no critical values between (akiaki1ϵki1)(a_{k_{i_{*}}}-a_{k_{i_{*}-1}}-\epsilon_{k_{i_{*}-1}}) and (akiaki1+ϵkiϵki1)(a_{k_{i_{*}}}-a_{k_{i_{*}-1}}+\epsilon_{k_{i_{*}}}-\epsilon_{k_{i_{*}-1}}). Therefore, for sufficiently small ϵki\epsilon_{k_{i_{*}}}, ϵki1\epsilon_{k_{i_{*}-1}}, we have that

(i=i+1cI𝒂+ϵ(σki,σki1))(i=2iI𝒂+ϵ(σki,σki1))U,\Big{(}\bigcap_{i=i_{*}+1}^{c}I_{\bm{a+\epsilon}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\Big{)}\cap\Big{(}\bigcap_{i=2}^{i_{*}}I_{\bm{a+\epsilon}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\Big{)}\cap U\,,

which is the (akiaki1)(a_{k_{i_{*}}}-a_{k_{i_{*}-1}})-level set of gig_{i_{*}} restricted to

(i=i+1cI𝒂+ϵ(σki,σki1))(i=2i1I𝒂+ϵ(σki,σki1))U,\Big{(}\bigcap_{i=i_{*}+1}^{c}I_{\bm{a+\epsilon}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\Big{)}\cap\Big{(}\bigcap_{i=2}^{i_{*}-1}I_{\bm{a+\epsilon}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\Big{)}\cap U\,,

is a submanifold of BB that is diffeomorphic to

(i=icI𝒂+ϵ(σki,σki1))(i=2i1I𝒂+ϵ(σki,σki1))U,\Big{(}\bigcap_{i=i_{*}}^{c}I_{\bm{a+\epsilon}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\Big{)}\cap\Big{(}\bigcap_{i=2}^{i_{*}-1}I_{\bm{a+\epsilon}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\Big{)}\cap U\,,

which is the (akiaki1+ϵkiϵki1)(a_{k_{i_{*}}}-a_{k_{i_{*}-1}}+\epsilon_{k_{i_{*}}}-\epsilon_{k_{i_{*}-1}})-level set of gig_{i_{*}} restricted to

(i=i+1cI𝒂+ϵ(σki,σki1))(i=2i1I𝒂+ϵ(σki,σki1))U.\Big{(}\bigcap_{i=i_{*}+1}^{c}I_{\bm{a+\epsilon}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\Big{)}\cap\Big{(}\bigcap_{i=2}^{i_{*}-1}I_{\bm{a+\epsilon}}(\sigma_{k_{i}},\sigma_{k_{i-1}})\Big{)}\cap U\,.

These submanifolds are smoothly parameterized by ϵki\epsilon_{k_{i_{*}}} and ϵki+1\epsilon_{k_{i_{*}+1}}. This concludes the inductive step.

Let 𝒂A\bm{a}\in A^{\prime}, where AA^{\prime} is defined as it was earlier in the proof. We showed above that for sufficiently small ϵN\bm{\epsilon}\in\mathbb{R}^{N}, the set of manifolds rI𝒂+ϵ(σir,σjr)U\bigcap_{r\in\ell}I_{\bm{a+\epsilon}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U (parameterized by ϵ\bm{\epsilon}) is a smoothly parameterized family of embeddings of rI𝒂(σir,σjr)U\bigcap_{r\in\ell}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U into UBU\subseteq B. Varying {ϵrrJ}\{\epsilon_{r}\mid r\in J_{\ell}\} (while holding ϵr\epsilon_{r} constant for rJr\not\in J_{\ell}) produces a smoothly parameterized family of embeddings of rI𝒂(σir,σjr)U\bigcap_{r\in\ell}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U while holding (<rJI𝒂(σir,σjr)U)\Big{(}\bigcap_{\ell^{\prime}<\ell}\bigcap_{r\in J_{\ell}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U\Big{)} constant. Therefore, because transverse intersections are generic,

(rJI𝒂(σir,σjr)U)(<rJI𝒂(σir,σjr)U)\big{(}\bigcap_{r\in J_{\ell}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U\Big{)}\pitchfork\Big{(}\bigcap_{\ell^{\prime}<\ell}\bigcap_{r\in J_{\ell}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U\Big{)}

for all \ell for almost every ϵ\bm{\epsilon} in a neighborhood of 𝟎N\bm{0}\in\mathbb{R}^{N}. This proves (25), which completes the proof. ∎

Lemma A.5.

Let {(ir,jr)}r=1m\{(i_{r},j_{r})\}_{r=1}^{m} be a set of index pairs such that {ir,jr}{is,js}\{i_{r},j_{r}\}\neq\{i_{s},j_{s}\} if rsr\neq s. For almost every 𝒂A\bm{a}\in A, we have that if UU is an open set in BB such that

(26) π(r=1m(M𝒂,irM𝒂,jr))U=r=1mI𝒂(σir,σjr)U,\pi\Big{(}\bigcap_{r=1}^{m}(M_{\bm{a},i_{r}}\cap M_{\bm{a},j_{r}})\Big{)}\cap U=\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U\,,

where π\pi is the projection π:B×B\pi:B\times\mathbb{R}\to B, then rJI𝒂(σir,σjr)\bigcap_{r\in J^{\prime}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) is a manifold for every J{1,,m}J^{\prime}\subseteq\{1,\ldots,m\} and

(27) Ty(I𝒂(σi1,σj1)I𝒂(σim,σjm))=r=1mTy(I𝒂(σir,σjr))T_{y}\Big{(}I_{\bm{a}}(\sigma_{i_{1}},\sigma_{j_{1}})\cap\cdots I_{\bm{a}}(\sigma_{i_{m}},\sigma_{j_{m}})\Big{)}=\bigcap_{r=1}^{m}T_{y}\Big{(}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}

for all yr=1mI𝒂(σir,σjr)Uy\in\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U.

Proof.

By Lemmas 4.8 and 4.9, rJI𝒂(σir,σjr)\bigcap_{r\in J^{\prime}}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) is a manifold for every J{1,,m}J^{\prime}\subseteq\{1,\ldots,m\} for almost every 𝒂A\bm{a}\in A. We have

Ty(I𝒂(σi1,σj1)I𝒂(σim,σjm))r=1mTy(I𝒂(σir,σjr))T_{y}\Big{(}I_{\bm{a}}(\sigma_{i_{1}},\sigma_{j_{1}})\cap\cdots\cap I_{\bm{a}}(\sigma_{i_{m}},\sigma_{j_{m}})\Big{)}\subseteq\bigcap_{r=1}^{m}T_{y}\Big{(}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}

because r=1mI𝒂(σir,σjr)I𝒂(σis,σjs)\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\subseteq I_{\bm{a}}(\sigma_{i_{s}},\sigma_{j_{s}}) for all ss.

Let 𝒗r=1mTy(I𝒂(σir,σjr))\bm{v}\in\bigcap_{r=1}^{m}T_{y}\Big{(}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}. Define π[m]:=π|r=1m(M𝒂,irM𝒂,jr)π1(U)\pi_{[m]}:=\pi|_{\bigcap_{r=1}^{m}(M_{\bm{a},i_{r}}\cap M_{\bm{a},j_{r}})\cap\pi^{-1}(U)}, and define πr:=π|M𝒂,irM𝒂,jrπ1(U)\pi_{r}:=\pi|_{M_{\bm{a},i_{r}}\cap M_{\bm{a},j_{r}}\cap\pi^{-1}(U)} for each rr. Each πr\pi_{r} is a diffeomorphism from M𝒂,irM𝒂,jrπ1(U)M_{\bm{a},i_{r}}\cap M_{\bm{a},j_{r}}\cap\pi^{-1}(U) to I𝒂(σir,σjr)UI_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U, and π[m]\pi_{[m]} is a diffeomorphism from r=1m(M𝒂,irM𝒂,jr)π1(U)\bigcap_{r=1}^{m}(M_{\bm{a},i_{r}}\cap M_{\bm{a},j_{r}})\cap\pi^{-1}(U) to r=1mI𝒂(σir,σjr)U\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\cap U. Let y~:=π[m]1\tilde{y}:=\pi^{-1}_{[m]} (which exists because π[m]1\pi^{-1}_{[m]} is a diffeomorphism), and let v~:=dπ[m]1(v)\tilde{v}:=d\pi^{-1}_{[m]}(v) (which exists because dπ[m]d\pi_{[m]} is an isomorphism). For all rr, we have y~=πr1(y)\tilde{y}=\pi^{-1}_{r}(y) and v~:=dπr1(v)\tilde{v}:=d\pi^{-1}_{r}(v). Therefore,

v~r=1mTy~(M𝒂,irM𝒂,jr).\tilde{v}\in\bigcap_{r=1}^{m}T_{\tilde{y}}\Big{(}M_{\bm{a},i_{r}}\cap M_{\bm{a},j_{r}}\Big{)}\,.

By Lemma A.1, we have Ty~(r=1m(M𝒂,irM𝒂,jr))=r=1mTy~(M𝒂,irM𝒂,jr)T_{\tilde{y}}\Big{(}\bigcap_{r=1}^{m}(M_{\bm{a},{i_{r}}}\cap M_{\bm{a},j_{r}})\Big{)}=\bigcap_{r=1}^{m}T_{\tilde{y}}(M_{\bm{a},{i_{r}}}\cap M_{\bm{a},j_{r}}) for all y~r=1m(M𝒂,irM𝒂,jr)\tilde{y}\in\bigcap_{r=1}^{m}(M_{\bm{a},{i_{r}}}\cap M_{\bm{a},j_{r}}) for almost every 𝒂A\bm{a}\in A, so

v~Ty~(r=1m(M𝒂,irM𝒂,jr))\tilde{v}\in T_{\tilde{y}}\Big{(}\bigcap_{r=1}^{m}(M_{\bm{a},i_{r}}\cap M_{\bm{a},j_{r}})\Big{)}

for almost every 𝒂A\bm{a}\in A. Therefore, v=dπ[m](v~)v=d\pi_{[m]}(\tilde{v}) is in Ty(π[m](r=1m(M𝒂,irM𝒂,jr)))=Ty(r=1mI𝒂(σir,σjr))T_{y}\Big{(}\pi_{[m]}\Big{(}\bigcap_{r=1}^{m}(M_{\bm{a},i_{r}}\cap M_{\bm{a},j_{r}})\Big{)}\Big{)}=T_{y}\Big{(}\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}, which implies that

Ty(r=1mI𝒂(σir,σjr))r=1mTy(I𝒂(σir,σjr)).T_{y}\Big{(}\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}\supseteq\bigcap_{r=1}^{m}T_{y}\Big{(}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}\,.

The following series of lemmas is used to prove Lemma 4.13, which shows that 𝒴𝒂\mathcal{Y}_{\bm{a}} is locally finite for almost every 𝒂A\bm{a}\in A, and Lemma 4.14, which shows that 𝒴𝒂\mathcal{Y}_{\bm{a}} satisfies the Axiom of the Frontier for almost every 𝒂A\bm{a}\in A. (Recall that 𝒴𝒂\mathcal{Y}_{\bm{a}} is the set of subsets of BB that is defined by (15).)

Lemma A.6.

If 𝒂A\bm{a}\in A is such that each Y𝒴𝒂Y\in\mathcal{Y}_{\bm{a}} is a manifold, then for any strict partial order \prec on 𝒦\mathcal{K}, there is a unique subset 𝒴𝒂𝒴𝒂\mathcal{Y}^{\prec}_{\bm{a}}\subseteq\mathcal{Y}_{\bm{a}} such that Z𝒂=Y𝒴𝒂YZ^{\prec}_{\bm{a}}=\bigcup_{Y\in\mathcal{Y}^{\prec}_{\bm{a}}}Y, where Z𝒂Z^{\prec}_{\bm{a}} is defined as in (4.2).

Proof.

Let Y𝒴𝒂Y\in\mathcal{Y}_{\bm{a}} and suppose that YZ𝒂Y\cap Z_{\bm{a}}^{\prec}\neq\emptyset. This implies that there is a point pYp\in Y such that f𝒂(,p)\prec_{f_{\bm{a}}(\cdot,p)} is the same as \prec. By Lemma 4.11, the simplex order induced by ff is constant in YY, so YZ𝒂Y\subseteq Z_{\bm{a}}^{\prec}. ∎

Lemma A.7.

Let \prec be a strict partial order on the simplices in 𝒦\mathcal{K}. Let 𝒂A\bm{a}\in A be such that

  1. (1)

    every Y𝒴𝒂Y\in\mathcal{Y}_{\bm{a}} is a manifold, where 𝒴𝒂\mathcal{Y}_{\bm{a}} is defined as in (15),

  2. (2)

    M𝒂,iM𝒂,jM_{\bm{a},i}\pitchfork M_{\bm{a},j} for all iji\neq j, where M𝒂,iM_{\bm{a},i} is defined as in (9),

  3. (3)

    r=1mI𝒂(σir,σjr)\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) is a manifold for all sets {(ir,jr}r=1m\{(i_{r},j_{r}\}_{r=1}^{m} of index pairs, and

  4. (4)

    we have

    (28) Tp(r=1mI𝒂(σir,σjr))=r=1mTp(I𝒂(σir,σjr))T_{p}\Big{(}\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}=\bigcap_{r=1}^{m}T_{p}\Big{(}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})\Big{)}

    for all sets {(ir,jr}r=1m\{(i_{r},j_{r}\}_{r=1}^{m} of index pairs and all pr=1mI𝒂(σir,σjr)p\in\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}).

Let 𝒴𝒂\mathcal{Y}^{\prec}_{\bm{a}} be the unique subset of 𝒴𝒂\mathcal{Y}_{\bm{a}} such that Z𝒂=Y𝒴𝒂YZ_{\bm{a}}^{\prec}=\bigcup_{Y\in\mathcal{Y}^{\prec}_{\bm{a}}}Y, which exists by Lemma A.6. Then every pBp\in B has a neighborhood that intersects at most one set Y𝒴𝒂Y\in\mathcal{Y}^{\prec}_{\bm{a}}.

Proof.

Let S(p)={(σi,σj)pI𝒂(σi,σj)}S(p)=\{(\sigma_{i},\sigma_{j})\mid p\in I_{\bm{a}}(\sigma_{i},\sigma_{j})\}. There is a neighborhood U0U_{0} of pp such that I(σi,σj)U0I(\sigma_{i},\sigma_{j})\cap U_{0}\neq\emptyset if and only if (σi,σj)S(p)(\sigma_{i},\sigma_{j})\in S(p). In a neighborhood of pp, each I(σi,σj)I(\sigma_{i},\sigma_{j}) is locally diffeomorphic (via the exponential map, for example) to Tp(I𝒂(σi,σj))T_{p}(I_{\bm{a}}(\sigma_{i},\sigma_{j})), which is an (n1)(n-1)-dimensional hyperplane. By (28), these local diffeomorphisms are compatible with each other, so there is a neighborhood UU of pp, a set {H(σi,σj)}(σi,σj)S(p)\{H(\sigma_{i},\sigma_{j})\}_{(\sigma_{i},\sigma_{j})\in S(p)} of hyperplanes in n\mathbb{R}^{n}, and a homeomorphism ϕ:U\phi:U\to\mathcal{B}, where \mathcal{B} is the open unit nn-ball, such that

ϕ((σi,σj)S(p)I𝒂(σi,σj)U)=(σi,σj)S(p)H(σi,σj)\phi\Big{(}\bigcap_{(\sigma_{i},\sigma_{j})\in S^{\prime}(p)}I_{\bm{a}}(\sigma_{i},\sigma_{j})\cap U\Big{)}=\bigcap_{(\sigma_{i},\sigma_{j})\in S^{\prime}(p)}H(\sigma_{i},\sigma_{j})\cap\mathcal{B}

for all S(p)S(p)S^{\prime}(p)\subseteq S(p). See Figure 8 for intuition, where we illustrate the neighborhood UU for a few points pBp\in B.

Refer to caption
Figure 8. For each point pip_{i}, we illustrate the idea behind the homeomorphism ϕ:Ui\phi:U_{i}\to\mathcal{B} in the proof of Lemma A.7, where UiU_{i} is a neighborhood of pip_{i} and \mathcal{B} is an open unit ball in 2\mathbb{R}^{2}. The base space is B=2B=\mathbb{R}^{2}; we show only the sets I(σ1,τ1)I(\sigma_{1},\tau_{1}) and I(σ2,τ2)I(\sigma_{2},\tau_{2}), which are curves in the plane, for some simplices σ1,σ2,τ1,τ2𝒦\sigma_{1},\sigma_{2},\tau_{1},\tau_{2}\in\mathcal{K} and some fibered filtration function f:𝒦×Bf:\mathcal{K}\times B\to\mathbb{R}.

The hyperplanes induce a stratification of \mathcal{B}, with a set 𝒴\mathcal{Y}^{\prime} of strata, such that for all Y𝒴Y^{\prime}\in\mathcal{Y}^{\prime}, we have Y=ϕ(YU)Y^{\prime}=\phi(Y\cap U) for some Y𝒴𝒂Y\in\mathcal{Y}_{\bm{a}}. Because M𝒂,iM𝒂,jM_{\bm{a},i}\pitchfork M_{\bm{a},j} for all iji\neq j, we have that H(σi,σj)\mathcal{B}\setminus H(\sigma_{i},\sigma_{j}) is the disjoint union of open sets W1(σi,σj)W_{1}(\sigma_{i},\sigma_{j}) and W2(σi,σj)W_{2}(\sigma_{i},\sigma_{j}) such that

f𝒂(σi,p)\displaystyle f_{\bm{a}}(\sigma_{i},p^{\prime}) <f𝒂(σj,p)for all pϕ1(W1(σi,σj)),\displaystyle<f_{\bm{a}}(\sigma_{j},p^{\prime})\qquad\text{for all }p^{\prime}\in\phi^{-1}(W_{1}(\sigma_{i},\sigma_{j}))\,,
f𝒂(σj,p)\displaystyle f_{\bm{a}}(\sigma_{j},p^{\prime}) <f𝒂(σi,p)for all pϕ1(W2(σi,σj))\displaystyle<f_{\bm{a}}(\sigma_{i},p^{\prime})\qquad\text{for all }p^{\prime}\in\phi^{-1}(W_{2}(\sigma_{i},\sigma_{j}))

for all (σi,σj)S(p)(\sigma_{i},\sigma_{j})\in S(p). Suppose that u1u_{1} and u2u_{2} are points in UU such that f𝒂(,u1)\prec_{f_{\bm{a}}(\cdot,u_{1})} and f𝒂(,u2)\prec_{f_{\bm{a}}(\cdot,u_{2})} are both the same as \prec. For each (σi,σj)S(p)(\sigma_{i},\sigma_{j})\in S(p), define the set

(29) V(σi,σj):={H(σi,σj),σiσj and σjσiW1(σi,σj),σiσjW2(σi,σj),σjσi.V(\sigma_{i},\sigma_{j}):=\begin{cases}H(\sigma_{i},\sigma_{j})\,,&\sigma_{i}\not\prec\sigma_{j}\text{ and }\sigma_{j}\not\prec\sigma_{i}\\ W_{1}(\sigma_{i},\sigma_{j})\,,&\sigma_{i}\prec\sigma_{j}\\ W_{2}(\sigma_{i},\sigma_{j})\,,&\sigma_{j}\prec\sigma_{i}\,.\end{cases}

We define

(30) V:=(σi,σj)S(p)V(σi,σj),V:=\bigcap_{(\sigma_{i},\sigma_{j})\in S(p)}V(\sigma_{i},\sigma_{j})\neq\emptyset\,,

which is a stratum in 𝒴\mathcal{Y}^{\prime}. Therefore, there is a Y𝒴𝒂Y\in\mathcal{Y}_{\bm{a}} such that YU=ϕ1(V)Y\cap U=\phi^{-1}(V), with u1,u2YUu_{1},u_{2}\in Y\cap U. Therefore, YY is the only element of 𝒴𝒂\mathcal{Y}^{\prec}_{\bm{a}} that UU intersects. ∎

Lemma A.8.

Let 0\prec_{0} be a strict partial order on the simplices in 𝒦\mathcal{K}, and define 𝒪\mathcal{O} to be the set of strict partial orders \prec such that

  1. (1)

    if σ0τ\sigma\prec_{0}\tau and στ\sigma\neq\tau, then either we have στ\sigma\prec\tau or we have στ\sigma\not\prec\tau and τσ\tau\not\prec\sigma,

  2. (2)

    if σ0τ\sigma\not\prec_{0}\tau and τ0σ\tau\not\prec_{0}\sigma, then στ\sigma\not\prec\tau and τσ\tau\not\prec\sigma, and

  3. (3)

    the strict partial order \prec is not the same as 0\prec_{0}.

If 𝒂A\bm{a}\in A is such that

  1. (1)

    every SE𝒂n¯S\in\overline{E_{\bm{a}}^{n-\ell}} is an \ell-dimensional smooth submanifold for every {1,,n}\ell\in\{1,\ldots,n\}, where nn is the dimension of BB,

  2. (2)

    M𝒂,iM𝒂,jM_{\bm{a},i}\pitchfork M_{\bm{a},j} for all iji\neq j,

  3. (3)

    the set r=1mI𝒂(σir,σjr)\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}}) is a manifold for all sets {(ir,jr)}r=1m\{(i_{r},j_{r})\}_{r=1}^{m} of index pairs, and

  4. (4)

    Ty(r=1mI𝒂(σir,σjr)=r=1mTy(I𝒂(σir,σjr))T_{y}(\bigcap_{r=1}^{m}I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})=\bigcap_{r=1}^{m}T_{y}(I_{\bm{a}}(\sigma_{i_{r}},\sigma_{j_{r}})) for all sets {(ir,jr)}r=1m\{(i_{r},j_{r})\}_{r=1}^{m} of index pairs,

then

Z𝒂0= in 𝒪Z𝒂.\partial Z^{\prec_{0}}_{\bm{a}}=\bigcup_{\prec\text{ in }\mathcal{O}}Z^{\prec}_{\bm{a}}\,.
Proof.

By Lemma 4.10, every Y𝒴𝒂Y\in\mathcal{Y}_{\bm{a}} is a manifold. By Lemmas A.6 and A.7, the sets Z𝒂0Z^{\prec_{0}}_{\bm{a}} and Z𝒂Z^{\prec}_{\bm{a}} (for all \prec in 𝒪\mathcal{O}) are submanifolds of BB.

Case 1: If dim(Z𝒂0)=0\dim(Z^{\prec_{0}}_{\bm{a}})=0, then we must have

Z𝒂0=I𝒂(σi1,σj1)I𝒂(σin,σjn).Z^{\prec_{0}}_{\bm{a}}=I_{\bm{a}}(\sigma_{i_{1}},\sigma_{j_{1}})\cap\cdots\cap I_{\bm{a}}(\sigma_{i_{n}},\sigma_{j_{n}})\,.

for some I𝒂(σi1,σj1)I𝒂(σin,σjn)E𝒂n¯I_{\bm{a}}(\sigma_{i_{1}},\sigma_{j_{1}})\cap\cdots\cap I_{\bm{a}}(\sigma_{i_{n}},\sigma_{j_{n}})\in\overline{E^{n}_{\bm{a}}}. If \prec is in 𝒪\mathcal{O}, then there is another pair (σin+1,σjn+1)(\sigma_{i_{n+1}},\sigma_{j_{n+1}}) of distinct simplices such that σin+1σjn+1\sigma_{i_{n+1}}\not\prec\sigma_{j_{n+1}} and σjn+1σin+1\sigma_{j_{n+1}}\not\prec\sigma_{i_{n+1}}. Therefore,

Z𝒂I𝒂(σi1,σj1)I𝒂(σin,σjn)I𝒂(σin+1,σjn+1),Z^{\prec}_{\bm{a}}\subseteq I_{\bm{a}}(\sigma_{i_{1}},\sigma_{j_{1}})\cap\cdots\cap I_{\bm{a}}(\sigma_{i_{n}},\sigma_{j_{n}})\cap I_{\bm{a}}(\sigma_{i_{n+1}},\sigma_{j_{n+1}})\,,

which is an element of E𝒂n+1¯\overline{E^{n+1}_{\bm{a}}}. By choice of 𝒂\bm{a}, every SE𝒂n+1¯S\in\overline{E^{n+1}_{\bm{a}}} is empty, so Z𝒂=Z^{\prec}_{\bm{a}}=\emptyset.

Case 2: If dim(Z𝒂0)1\dim(Z^{\prec_{0}}_{\bm{a}})\geq 1, let \prec be any strict partial order in 𝒪\mathcal{O}. Let pZ𝒂p\in Z^{\prec}_{\bm{a}}. Let S(p)={(σi,σj)pI𝒂(σi,σj)}S(p)=\{(\sigma_{i},\sigma_{j})\mid p\in I_{\bm{a}}(\sigma_{i},\sigma_{j})\}. By the same argument as in the proof of Lemma A.7, there is a neighborhood UU of pp, a set {H(σi,σj)}(σi,σj)S(p)\{H(\sigma_{i},\sigma_{j})\}_{(\sigma_{i},\sigma_{j})\in S(p)} of hyperplanes in n\mathbb{R}^{n}, and a homeomorphism ϕ:U\phi:U\to\mathcal{B}, where \mathcal{B} is the open unit nn-ball, such that

ϕ((σi,σj)S(p)I𝒂(σi,σj)U)=(σi,σj)S(p)H(σi,σj)\phi\Big{(}\bigcap_{(\sigma_{i},\sigma_{j})\in S^{\prime}(p)}I_{\bm{a}}(\sigma_{i},\sigma_{j})\cap U\Big{)}=\bigcap_{(\sigma_{i},\sigma_{j})\in S^{\prime}(p)}H(\sigma_{i},\sigma_{j})\cap\mathcal{B}

for all S(p)S(p)S^{\prime}(p)\subseteq S(p). See Figure 8.

Because M𝒂,iM𝒂,jM_{\bm{a},i}\pitchfork M_{\bm{a},j} for all iji\neq j, we have that H(σi,σj)\mathcal{B}\setminus H(\sigma_{i},\sigma_{j}) is the disjoint union of open sets W1(σi,σj)W_{1}(\sigma_{i},\sigma_{j}) and W2(σi,σj)W_{2}(\sigma_{i},\sigma_{j}) such that

f𝒂(σi,p)\displaystyle f_{\bm{a}}(\sigma_{i},p^{\prime}) <f𝒂(σj,p)for all pϕ1(W1(σi,σj)),\displaystyle<f_{\bm{a}}(\sigma_{j},p^{\prime})\qquad\text{for all }p^{\prime}\in\phi^{-1}(W_{1}(\sigma_{i},\sigma_{j}))\,,
f𝒂(σj,p)\displaystyle f_{\bm{a}}(\sigma_{j},p^{\prime}) <f𝒂(σi,p)for all pϕ1(W2(σi,σj))\displaystyle<f_{\bm{a}}(\sigma_{i},p^{\prime})\qquad\text{for all }p^{\prime}\in\phi^{-1}(W_{2}(\sigma_{i},\sigma_{j}))

for all (σi,σj)S(p)(\sigma_{i},\sigma_{j})\in S(p). For each (σi,σj)S(p)(\sigma_{i},\sigma_{j})\in S(p), define the set V(σi,σj)V(\sigma_{i},\sigma_{j}) as in (29), and define the set VV as in (30). The set ϕ1(V)\phi^{-1}(V) is a nonempty subset of UZ𝒂0U\cap Z^{\prec_{0}}_{\bm{a}}. This implies that pp is a limit point of Z𝒂0Z^{\prec_{0}}_{\bm{a}}, so Z𝒂0Z𝒂0¯Z_{\bm{a}}^{\prec_{0}}\subseteq\overline{Z^{\prec_{0}}_{\bm{a}}}. Because \prec is not the same as 0\prec_{0}, we have that Z𝒂Z𝒂0=Z^{\prec}_{\bm{a}}\cap Z^{\prec_{0}}_{\bm{a}}=\emptyset. Therefore, Z𝒂0Z𝒂0Z_{\bm{a}}^{\prec_{0}}\subseteq\partial Z^{\prec_{0}}_{\bm{a}} and

 in 𝒪Z𝒂Z𝒂0.\bigcup_{\prec\text{ in }\mathcal{O}}Z^{\prec}_{\bm{a}}\subseteq\partial Z^{\prec_{0}}_{\bm{a}}\,.

Now suppose that pp is in the complement of Z𝒂0( in 𝒪Z𝒂)Z^{\prec_{0}}_{\bm{a}}\cup\Big{(}\bigcup_{\prec\text{ in }\mathcal{O}}Z^{\prec}_{\bm{a}}\Big{)}. Because f𝒂(,p)\prec_{f_{\bm{a}}(\cdot,p)} is not the same as 0\prec_{0} or any \prec in 𝒪\mathcal{O}, there is a pair (σi,σj)(\sigma_{i},\sigma_{j}) of simplices such that f(σi,p)<f(σj,p)f(\sigma_{i},p)<f(\sigma_{j},p) and either we have σj0σi\sigma_{j}\prec_{0}\sigma_{i} or we have σjσi\sigma_{j}\not\prec\sigma_{i} and σiσj\sigma_{i}\not\prec\sigma_{j}. By continuity of ff, there is a neighborhood Uσi,σjU_{\sigma_{i},\sigma_{j}} of pp such that f𝒂(σi,p)<f𝒂(σj,p)f_{\bm{a}}(\sigma_{i},p^{\prime})<f_{\bm{a}}(\sigma_{j},p^{\prime}) for all pUσi,σjp^{\prime}\in U_{\sigma_{i},\sigma_{j}}. Therefore, Uσi,σjU_{\sigma_{i},\sigma_{j}} is in the complement of Z𝒂0Z^{\prec_{0}}_{\bm{a}}, so pp is not in Z𝒂0¯\overline{Z^{\prec_{0}}_{\bm{a}}}. This implies

Z𝒂0 in 𝒪Z𝒂,\partial Z^{\prec_{0}}_{\bm{a}}\subseteq\bigcup_{\prec\text{ in }\mathcal{O}}Z^{\prec}_{\bm{a}}\,,

which completes the proof. ∎

References

  • [1] Henry Adams, Tegan Emerson, Michael Kirby, Rachel Neville, Chris Peterson, Patrick Shipman, Sofya Chepushtanova, Eric Hanson, Francis Motta, and Lori Ziegelmeier. Persistence images: A stable vector representation of persistent homology. Journal of Machine Learning Research, 18(8):1–35, 2017.
  • [2] Pankaj K. Agarwal, Herbert Edelsbrunner, John Harer, and Yusu Wang. Extreme elevation on a 2-manifold. Discrete & Computational Geometry, 36:553–572, 2006.
  • [3] Erik J. Amézquita, Michelle Y. Quigley, Tim Ophelders, Jacob B. Landis, Daniel Koenig, Elizabeth Munch, and Daniel H. Chitwood. Measuring hidden phenotype: quantifying the shape of barley seeds using the Euler characteristic transform. in silico Plants, 4(1), 2022.
  • [4] Massimo Ferri Patrizio Frosini Claudia Landi Andrea Cerri, Barbara Di Fabio. Betti numbers in multidimensional persistent homology are stable functions. Mathematical Methods in the Applied Sciences, 36(12):1543–1557, 2013.
  • [5] Shreya Arya, Justin Curry, and Sayan Mukherjee. A sheaf-theoretic construction of shape space. arXiv:2204.09020, 2022.
  • [6] Omer Bobrowski, Sayan Mukherjee, and Jonathan E. Taylor. Topological consistency via kernel estimation. Bernoulli, 23(1):288–328, 2017.
  • [7] Magnus Bakke Botnan and Michael Lesnick. An introduction to multiparameter persistence. arxiv:2203.14289, 2022.
  • [8] Gunnar Carlsson and Afra Zomorodian. The theory of multidimensional persistence. Discrete and Computational Geometry, 42:71–93, 2007.
  • [9] Andrea Cerri, Marc Ethier, and Patrizio Frosini. A study of monodromy in the computation of multidimensional persistence. In Rocio Gonzalez-Diaz, Maria-Jose Jimenez, and Belen Medrano, editors, Discrete Geometry for Computer Imagery, volume 7749, pages 192–202, Berlin, Heidelberg, 2013. Springer.
  • [10] Frédéric Chazal, David Cohen-Steiner, and Quentin Mérigot. Geometric inference for probability measures. Foundations of Computational Mathematics, 11:733–751, 2011.
  • [11] Yao-li Chuang, Maria R. D’Orsogna, Daniel Marthaler, Andrea L. Bertozzi, and Lincoln S. Chayes. State transitions and the continuum limit for a 2D interacting, self-propelled particle system. Physica D: Nonlinear Phenomena, 232(1):33–47, 2007.
  • [12] Moo K. Chung, Peter Bubenik, and Peter T. Kim. Persistence diagrams of cortical surface data. In Jerry L. Prince, Dzung L. Pham, and Kyle J. Myers, editors, Information Processing in Medical Imaging, volume 5636, pages 386–397, Berlin, Heidelberg, 2009. Springer.
  • [13] David Cohen-Steiner, Herbert Edelsbrunner, and John Harer. Stability of persistence diagrams. Discrete and Computational Geometry, 37:103–120, 2007.
  • [14] David Cohen-Steiner, Herbert Edelsbrunner, and Dmitriy Morozov. Vines and vineyards by updating persistence in linear time. In Proceedings of the Twenty-Second Annual Symposium on Computational Geometry, SCG ’06, pages 119–126, New York, NY, USA, 2006. Association for Computing Machinery.
  • [15] Padraig Corcoran and Christopher B. Jones. Modelling topological features of swarm behaviour in space and time with persistence landscapes. IEEE Access, 5:18534–18544, 2017.
  • [16] Justin Curry. Sheaves, Cosheaves, and Applications. PhD thesis, University of Pennsylvania, 2014.
  • [17] Justin Curry, Haibin Hang, Washington Mio, Tom Needham, and Osman Berat Okutan. Decorated merge trees for persistent topology. Journal of Applied and Computational Topology, 6:371–428, 2022.
  • [18] Justin Curry, Sayan Mukherjee, and Katharine Turner. How many directions determine a shape and other sufficiency results for two topological transforms. Transactions of the American Mathematical Society, Series B, 9:1006–1043, 2022.
  • [19] Herbert Edelsbrunner and John Harer. Computational Topology: An Introduction. American Mathematical Society, Providence, RI, 2010.
  • [20] Herbert Edelsbrunner, David Letscher, and Afra Zomorodian. Topological persistence and simplification. Discrete & Computational Geometry, 28:511–533, 2022.
  • [21] Robert Ghrist. Elementary Applied Topology. Createspace, 1st edition, 2014.
  • [22] Jakob Hansen and Robert Ghrist. Toward a spectral theory of cellular sheaves. Journal of Applied. and Computational Topology, 3:315–358, 2019.
  • [23] Abigail Hickok. Computing persistence diagram bundles. arXiv:2210.06424, 2022.
  • [24] Abigail Hickok, Deanna Needell, and Mason A Porter. Analysis of spatial and spatiotemporal anomalies using persistent homology: Case studies with COVID-19 data. SIAM J. Math. Data Sci., 4(3):1116–1144, 2022.
  • [25] Woojin Kim and Facundo Mémoli. Spatiotemporal persistent homology for dynamic metric spaces. Discrete & Computational Geometry, 66:831–875, 2021.
  • [26] Yanjie Li, Dingkang Wang, Giorgio A. Ascoli, Partha Mitra, and Yusu Wang. Metrics for comparing neuronal tree shapes based on persistent homology. PLoS ONE, 12(8):e0182184, 2017.
  • [27] Shawn Martin, Aidan Thompson, Evangelos A. Coutsias, and Jean-Paul Watson. Topology of cyclo-octane energy landscape. Journal of Chemical Physics, 132(23):234115, 2010.
  • [28] Nina Otter, Mason A. Porter, Ulrike Tillmann, Peter Grindrod, and Heather A. Harrington. A roadmap for the computation of persistent homology. EPJ Data Science, 6:17, 2017.
  • [29] Raul Rabadan and Andrew J. Blumberg. Topological Data Analysis for Genomics and Evolution: Topology in Biology. Cambridge University Press, Cambridge, 2019.
  • [30] Kristian Strommen, Matthew Chantry, Joshua Dorrington, and Nina Otter. A topological perspective on weather regimes. Climate Dynamics, 2022.
  • [31] Chad M. Topaz, Lori Ziegelmeier, and Tom Halverson. Topological data analysis of biological aggregation models. PLoS ONE, 10(5):e0126383, 2015.
  • [32] Katharine Turner, Sayan Mukherjee, and Doug M. Boyer. Persistent homology transform for modeling shapes and surfaces. Information and Inference: A Journal of the IMA, 3(4):310–344, 2014.
  • [33] Katherine Turner. Representing vineyard modules. arXiv:2307.06020, 2023.
  • [34] Yusu Wang, Pankaj K. Agarwal, Paul Brown, Herbert Edelsbrunner, and Johannes Rudolph. Coarse and reliable geometric alignment for protein docking. Pacific Symposium on Biocomputing, 10:64–75, 2005.
  • [35] Schmuel Weinberger. The Topological Classification of Stratified Spaces. Chicago Lectures in Mathematics. The University of Chicago Press, 1994.
  • [36] Lu Xian, Henry Adams, Chad M. Topaz, and Lori Ziegelmeier. Capturing dynamics of time-varying data via topology. Foundations of Data Science, 4:1–36, 2022.