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

Computing persistent Stiefel-Whitney classes
of line bundles

Abstract.

We propose a definition of persistent Stiefel-Whitney classes of vector bundle filtrations. It relies on seeing vector bundles as subsets of some Euclidean spaces. The usual Čech filtration of such a subset can be endowed with a vector bundle structure, that we call a Čech bundle filtration. We show that this construction is stable and consistent. When the dataset is a finite sample of a line bundle, we implement an effective algorithm to compute its first persistent Stiefel-Whitney class. In order to use simplicial approximation techniques in practice, we develop a notion of weak simplicial approximation. As a theoretical example, we give an in-depth study of the normal bundle of the circle, which reduces to understanding the persistent cohomology of the torus knot (1,2). We illustrate our method on several datasets inspired by image analysis.

1 Introduction

The inference of relevant topological properties of data represented as point clouds in Euclidean spaces is a central challenge in Topological Data Analysis (TDA). Given a (finite) set of points XX in n\mathbb{R}^{n}, persistent homology provides a now classical and powerful tool to construct persistence diagrams whose points can be interpreted as homological features of XX at different scales.

In this work, we aim at developing a similar theoretical framework for another topological invariant: the Stiefel-Whitney classes. These classes, and more generally characteristic classes, are a powerful tool from algebraic topology, that contains additional information to the cohomology groups. For the Stiefel-Whitney classes to be defined, the input topological space has to be endowed with an additional structure: a real vector bundle. They have been widely used in differential topology, for instance in the problem of deciding orientability of manifolds, of immersing manifolds in low-dimensional spaces, or in cobordism problems (Milnor and Stasheff, 2016). Our work is motivated by introducing this tool to the TDA community.

Previous work.

To our knowledge, the problem of estimating Stiefel-Whitney classes from a point cloud observation has received little attention. In the work of Aubrey (2011), one finds an algorithm to compute the Stiefel-Whitney classes in the particular case of the tangent bundle of a Euler mod-22 space (that is, a simplicial complex for which the link of each simplex has even Euler characteristic). Close to the subject, Perea (2018) proposes a dimensionality reduction algorithm, based on the choice of a Stiefel-Whitney class, seen as a persistent cohomology class.

Recently, Scoccola and Perea (2021) developed several notions of vector bundle adapted to finite simplicial complexes, one of which is used in this paper. They propose algorithms to compute the first two Stiefel-Whitney classes, which are conceptually different than the one presented here.

Our contributions.

Just as persistent homology allows to extract homological features from filtrations of topological spaces, we propose a framework that allows to extract Stiefel-Whitney classes features from filtrations of vector bundles. It is briefly motivated here.

In general, if XX is a topological space endowed with a vector bundle ξ\xi of dimension dd, there exists a collection of cohomology classes w1(ξ),,wd(ξ)w_{1}(\xi),...,w_{d}(\xi), the Stiefel-Whitney classes, such that wi(ξ)w_{i}(\xi) is an element of the cohomology group Hi(X)H^{i}(X) over 2\mathbb{Z}_{2} for i1,di\in\llbracket 1,d\rrbracket. In order to define Stiefel-Whitney classes in a persistent-theoretic framework, we will use a convenient definition of vector bundles: defining a vector bundle over a compact space XX is equivalent (up to isomorphism of vector bundles) to defining a continuous map p:X𝒢d(m)p\colon X\rightarrow\mathcal{G}_{d}(\mathbb{R}^{m}) for mm large enough, where 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) is the Grassmann manifold of dd-planes in m\mathbb{R}^{m}. Such a map is called a classifying map for ξ\xi.

Given a classifying map p:X𝒢d(m)p\colon X\rightarrow\mathcal{G}_{d}(\mathbb{R}^{m}) of a vector bundle ξ\xi, the Stiefel-Whitney classes w1(ξ),,wd(ξ)w_{1}(\xi),...,w_{d}(\xi) can be defined by pushing forward some particular classes of the Grassmannian via the induced map in cohomology p:H(X)H(𝒢d(m))p^{*}\colon H^{*}(X)\leftarrow H^{*}(\mathcal{G}_{d}(\mathbb{R}^{m})). If wiw_{i} denotes the ithi^{\text{th}} Stiefel-Whitney class of the Grassmannian, then the ithi^{\text{th}} Stiefel-Whitney class of the vector bundle ξ\xi is

wi(ξ)=p(wi).\displaystyle w_{i}(\xi)=p^{*}(w_{i}). (1)

In order to translate these considerations in a persistent-theoretic setting, suppose that we are given a dataset of the form (X,p)(X,p), where XX is a finite subset of n\mathbb{R}^{n}, and pp is a map p:X𝒢d(m)p\colon X\rightarrow\mathcal{G}_{d}(\mathbb{R}^{m}). Denote by (Xt)t0(X^{t})_{t\geq 0} the Čech filtration of XX, that is, the collection of the tt-thickenings XtX^{t} of XX in the ambient space n\mathbb{R}^{n}. It is also known as the offset filtration of XX. In order to define some persistent Stiefel-Whitney classes, one would try to extend the map p:X𝒢d(m)p\colon X\rightarrow\mathcal{G}_{d}(\mathbb{R}^{m}) to pt:Xt𝒢d(m)p^{t}\colon X^{t}\rightarrow\mathcal{G}_{d}(\mathbb{R}^{m}). However, we did not find any interesting way to extend this map. To overcome this issue, we propose to look at the dataset in a different way. Transform the vector bundle (X,p)(X,p) into a subset of n×𝒢d(m)\mathbb{R}^{n}\times\mathcal{G}_{d}(\mathbb{R}^{m}) via

Xˇ={(x,p(x)),xX}.\displaystyle\check{X}=\left\{\left(x,p(x)\right),~{}x\in X\right\}.

The Grassmann manifold 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) can be naturally embedded in M(m)M(\mathbb{R}^{m}), the space of m×mm\times m matrices. From this viewpoint, Xˇ\check{X} can be seen as a subset of n×M(m)\mathbb{R}^{n}\times M(\mathbb{R}^{m}). Let (Xˇt)t0(\check{X}^{t})_{t\geq 0} denotes the Čech filtration of Xˇ\check{X} in the ambient space n×M(m)\mathbb{R}^{n}\times M(\mathbb{R}^{m}). A natural map pt:Xˇt𝒢d(m)p^{t}\colon\check{X}^{t}\rightarrow\mathcal{G}_{d}(\mathbb{R}^{m}) can be defined: map a point (x,A)Xˇt(x,A)\in\check{X}^{t} to the projection of AA on 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}), seen as a subset of M(m)M(\mathbb{R}^{m}):

pt:(x,A)n×M(m)proj(A,𝒢d(m)).\displaystyle p^{t}\colon(x,A)\in\mathbb{R}^{n}\times M(\mathbb{R}^{m})\longmapsto\mathrm{proj}\left(A,\mathcal{G}_{d}(\mathbb{R}^{m})\right).

The projection is well-defined if AA does not belong to the medial axis of 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}). We show that this condition can be verified in practice (Lemma 3.1). The Čech filtration of Xˇ\check{X}, endowed with the extended projection maps (pt:Xˇt𝒢d(m))t(p^{t}\colon\check{X}^{t}\rightarrow\mathcal{G}_{d}(\mathbb{R}^{m}))_{t}, is called the Čech bundle filtration. Now we can define the ithi^{\text{th}} persistent Stiefel-Whitney class as the collection of classes wi(X)=(wit(X))tw_{i}(X)=(w_{i}^{t}(X))_{t}, where wit(X)w_{i}^{t}(X) is the push-forward

wit(X)=(pt)(wi),\displaystyle w_{i}^{t}(X)=(p^{t})^{*}(w_{i}),

and where wiw_{i} is the ithi^{\text{th}} Stiefel-Whitney class of the Grassmann manifold (compare with Equation (1)). We summarize the information given by a persistent Stiefel-Whitney class in a diagram, that we call a lifebar.

The construction we propose is defined for any subset of n×M(m)\mathbb{R}^{n}\times M(\mathbb{R}^{m}). We prove that this construction is stable, a result reminiscent of the usual stability theorem of persistent homology (Corollary 3.4). We also show that the persistent Stiefel-Whitney classes are consistent estimators of Stiefel-Whitney classes (Corollary 3.7).

Moreover, we propose a concrete algorithm to compute the persistent Stiefel-Whitney classes. This algorithm is based on several ingredients, including the triangulation of projective spaces, and the simplicial approximation method. The simplicial approximation, widely used in theory, can be applied only if the simplicial complex is refined enough, a property that is attested by the star condition. However, this condition cannot be verified in practice. We circumvent this problem by introducing the weak star condition, a variant that only depends on the combinatorial structure of the simplicial complex. When the simplicial complex is fine enough, the star condition and the weak star condition turn out to be equivalent notions (Proposition 5.5).

Numerical experiments.

A Python notebook, containing a concise demonstration of our method, can be found at https://github.com/raphaeltinarrage/PersistentCharacteristicClasses/blob/master/Demo.ipynb. Another notebook, containing experiments on datasets inspired by image analysis, can be found at https://github.com/raphaeltinarrage/PersistentCharacteristicClasses/blob/master/Experiments.ipynb.

Outline.

The rest of the paper is as follows. Sect. 2 gathers usual definitions related to vector bundles, Stiefel-Whitney classes, simplicial approximation and persistent cohomology. The definitions of vector bundle filtrations and persistent Stiefel-Whitney classes are given in Sect. 3, where their stability and consistency properties are established. In Sect. 4, we propose a sketch of algorithm to compute these classes, based on simplicial approximation techniques. In Sect. 5 we give a particular attention to some technical details needed to implement this algorithm. In Sect. 6 we apply our algorithm on concrete datasets. For the clarity of the paper, the proofs of some results have been postponed to the appendices.

2 Background

2.1 Stiefel-Whitney classes

In this subsection, we define vector bundles and Stiefel-Whitney classes. The reader may refer to Milnor and Stasheff (2016) for an extended presentation. Let XX be a topological space and d1d\geq 1 an integer.

Vector bundles.

A vector bundle ξ\xi of dimension dd over XX consists of a topological space A=A(ξ)A=A(\xi), the total space, a continuous map π=π(ξ):AX\pi=\pi(\xi)\colon A\rightarrow X, the projection map, and for every xXx\in X, a structure of dd-dimensional vector space on the fiber π1({x})\pi^{-1}(\{x\}). Moreover, ξ\xi must satisfy the local triviality condition: for every xXx\in X, there exists a neighborhood UXU\subseteq X of xx and a homeomorphism h:U×dπ1(U)h\colon U\times\mathbb{R}^{d}\rightarrow\pi^{-1}(U) such that for every yUy\in U, the map zh(y,z)z\mapsto h(y,z) defines an isomorphism between the vector spaces d\mathbb{R}^{d} and π1({y})\pi^{-1}(\{y\}).

A bundle map between two vector bundles ξ\xi and η\eta with base spaces XX and YY is a continuous map f:A(ξ)A(η)f\colon A(\xi)\rightarrow A(\eta) which sends each fiber π(ξ)1({x})\pi(\xi)^{-1}(\{x\}) isomorphically into another fiber π(η)1({x})\pi(\eta)^{-1}(\{x^{\prime}\}). If such a map exists, there exist a unique map f¯\overline{f} which makes the following diagram commute:

A(ξ){A(\xi)}A(η){A(\eta)}X{X}Y{Y}f\scriptstyle{f}π(ξ)\scriptstyle{\pi(\xi)}π(η)\scriptstyle{\pi(\eta)}f¯\scriptstyle{\overline{f}}

If X=YX=Y, and if ff is a homeomorphism, we say that ff is an isomorphism of vector bundles, and that ξ\xi and η\eta are isomorphic. The trivial bundle of dimension dd over XX, denoted ϵ\epsilon, is defined with the total space A(ϵ)=X×dA(\epsilon)=X\times\mathbb{R}^{d}, with the projection map π(ϵ)\pi(\epsilon) being the projection on the first coordinate, and where each fiber is endowed with the usual vector space structure of d\mathbb{R}^{d}. A vector bundle ξ\xi over XX is said trivial if it is isomorphic to ϵ\epsilon.

Let m0m\geq 0. The Grassmann manifold 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) is a set which consists of all dd-dimensional linear subspaces of m\mathbb{R}^{m}. It can be given a smooth manifold structure. When d=1d=1, 𝒢1(m)\mathcal{G}_{1}(\mathbb{R}^{m}) corresponds to the real projective space m1()\operatorname{\mathbb{P}}_{m-1}(\mathbb{R}). In order to avoid mentioning mm, it is convenient to consider the infinite Grassmannian. The infinite Grassmann manifold 𝒢d()\mathcal{G}_{d}(\mathbb{R}^{\infty}) is the set of all dd-dimensional linear subspaces of \mathbb{R}^{\infty}, where \mathbb{R}^{\infty} is the vector space of sequences with a finite number of nonzero terms.

Let XX be a paracompact space. There exists a correspondence between the vector bundles over XX (up to isomorphism) and the continuous maps X𝒢d()X\rightarrow\mathcal{G}_{d}(\mathbb{R}^{\infty}) (up to homotopy). Such a map is called a classifying map. When XX is compact, there exist an integer m1m\geq 1 such that a classifying map factorizes through

X{X}𝒢d(m){\mathcal{G}_{d}(\mathbb{R}^{m})}𝒢d().{\mathcal{G}_{d}(\mathbb{R}^{\infty}).}

Consequently, in the rest of this paper, we shall consider that vector bundles are given as a continuous maps X𝒢d(m)X\rightarrow\mathcal{G}_{d}(\mathbb{R}^{m}) or X𝒢d()X\rightarrow\mathcal{G}_{d}(\mathbb{R}^{\infty}).

Axioms for Stiefel-Whitney classes.

To each vector bundle ξ\xi over a paracompact base space XX, one associates a sequence of cohomology classes

wi(ξ)Hi(X,2),i,\displaystyle w_{i}(\xi)\in H^{i}(X,\mathbb{Z}_{2}),~{}~{}~{}~{}~{}i\in\mathbb{N},

called the Stiefel-Whitney classes of ξ\xi. These classes satisfy:

  • Axiom 1: w0w_{0} is equal to 1H0(X,2)1\in H^{0}(X,\mathbb{Z}_{2}), and if ξ\xi is of dimension dd, then wi(ξ)=0w_{i}(\xi)=0 for i>di>d.

  • Axiom 2: if f:ξηf\colon\xi\rightarrow\eta is a bundle map, then wi(ξ)=f¯wi(η)w_{i}(\xi)=\overline{f}^{*}w_{i}(\eta), where f¯\overline{f}^{*} is the map in cohomology induced by the underlying map f¯:XY\overline{f}\colon X\rightarrow Y between base spaces.

  • Axiom 3: if ξ,η\xi,\eta are bundles over the same base space XX, then for all kk\in\mathbb{N}, wk(ξη)=i=0kwi(ξ)wki(η)w_{k}(\xi\oplus\eta)=\sum_{i=0}^{k}w_{i}(\xi)\smile w_{k-i}(\eta), where \oplus denotes the Withney sum, and \smile denotes the cup product.

  • Axiom 4: if γ11\gamma_{1}^{1} denotes the tautological bundle of the projective line 𝒢1(2)\mathcal{G}_{1}(\mathbb{R}^{2}), then w1(γ11)0w_{1}(\gamma_{1}^{1})\neq 0.

If such classes exists, then one proves that they are unique. A way to show that they actually exist relies on the cohomology of the Grassmannians.

Construction of the Stiefel-Whitney classes.

The cohomology rings of the Grassmann manifolds admit a simple description: H(𝒢d(),2)H^{*}(\mathcal{G}_{d}(\mathbb{R}^{\infty}),\mathbb{Z}_{2}) is the free abelian ring generated by dd elements w1,,wdw_{1},...,w_{d}. As a graded algebra, the degree of these elements are |w1|=1,,|wd|=d|w_{1}|=1,...,|w_{d}|=d. Hence we can write

H(𝒢d(),2)2[w1,,wd].\displaystyle H^{*}(\mathcal{G}_{d}(\mathbb{R}^{\infty}),\mathbb{Z}_{2})\cong\mathbb{Z}_{2}[w_{1},...,w_{d}].

The generators w1,,wdw_{1},...,w_{d} can be seen as the Stiefel-Whitney classes of a particular vector bundle on 𝒢d()\mathcal{G}_{d}(\mathbb{R}^{\infty}), called the tautological bundle. Now, for any vector bundle ξ\xi, define

wi(ξ)=fξ(wi),\displaystyle w_{i}(\xi)=f_{\xi}^{*}(w_{i}),

where fξ:X𝒢d()f_{\xi}\colon X\rightarrow\mathcal{G}_{d}(\mathbb{R}^{\infty}) is a classifying map for ξ\xi, and fξ:H(X)H(𝒢d())f_{\xi}^{*}\colon H^{*}(X)\leftarrow H^{*}(\mathcal{G}_{d}(\mathbb{R}^{\infty})) the induced map in cohomology. This construction yields the Stiefel-Whitney classes.

Interpretation of the Stiefel-Whitney classes.

The Stiefel-Whitney classes are invariants of isomorphism classes of vector bundles, and carry topological information. Their main interpretation is the following: the Stiefel-Whitney classes are obstructions to the existence of nowhere vanishing sections of vector bundles. Let us explain this result. A section of a vector bundle π(ξ):AX\pi(\xi)\colon A\rightarrow X is a continuous map s:XAs\colon X\rightarrow A such that s(x)π1({x})s(x)\in\pi^{-1}(\{x\}) for all xXx\in X. It is nowhere vanishing if s(x)0s(x)\neq 0 for all xXx\in X, where 0 denotes the origin of the vector space π1({x})\pi^{-1}(\{x\}). Given kk sections s1,,sks_{1},\dots,s_{k}, we say that they are independent if the family (s1(x),,sk(x))\left(s_{1}(x),\dots,s_{k}(x)\right) is free for all xXx\in X. Then the following result holds: if a vector bundle ξ\xi of dimension dd admits kk independent and nowhere vanishing sections, then the top kk Stiefel-Whitney classes wd(ξ),,wdk+1(ξ)w_{d}(\xi),\dots,w_{d-k+1}(\xi) are zero.

Another property that we will use in this paper is the following: the first Stiefel-Whitney class detects orientability. More precisely, the first Stiefel-Whitney class w1(ξ)w_{1}(\xi) is zero if and only if the vector bundle ξ\xi is orientable. In the same vein, if XX is a compact manifold and τ\tau its tangent bundle, then the manifold XX is orientable if and only if w1(τ)=0w_{1}(\tau)=0.

In this paper, we will particularly study line bundles, that is, vector bundles of dimension d=1d=1. As a consequence of being an obstruction to nowhere vanishing sections, a line bundle ξ\xi on any topological space XX is trivial if and only if w1(ξ)=0w_{1}(\xi)=0. More generally, the first Stiefel-Whitney class establishes a bijection between the isomorphism classes of line bundles over XX and its first cohomology group H1(X)H^{1}(X) over 2\mathbb{Z}_{2}. As an example, the circle 𝕊1\operatorname{\mathbb{S}}_{1} has cohomology group H1(𝕊1)=2H^{1}(\operatorname{\mathbb{S}}_{1})=\mathbb{Z}_{2}, hence admits only two isomorphism classes of line bundles. As another example, the sphere 𝕊2\operatorname{\mathbb{S}}_{2} has trivial cohomology group H1(𝕊2)=0H^{1}(\operatorname{\mathbb{S}}_{2})=0, hence only admits trivial line bundles.

2.2 Simplicial approximation

We start by defining the simplicial complexes and their topology. We then describe the technique of simplicial approximation, based on the book of Hatcher (2002).

Simplicial complexes.

A simplicial complex is a set KK such that there exists a set VV, the set of vertices, with KK a collection of finite and non-empty subsets of VV, and such that KK satisfies the following condition: for every σK\sigma\in K and every non-empty subset νσ\nu\subseteq\sigma, ν\nu is in KK. The elements of KK are called faces or simplices of the simplicial complex KK.

For every simplex σK\sigma\in K, we define its dimension dim(σ)=card(σ)1\dim(\sigma)=\mathrm{card}(\sigma)-1. The dimension of KK, denoted dim(K)\dim(K), is the maximal dimension of its simplices. For every i0i\geq 0, the ii-skeleton Ki{K}^{i} is defined as the subset of KK consisting of simplices of dimension at most ii. Note that K0{K}^{0} corresponds to the underlying vertex set VV, and that K1{K}^{1} is a graph. Given a graph GG, the corresponding clique complex is the simplicial complex whose simplices are the sets of vertices of the cliques of GG. We say that a simplicial complex KK is a flag complex if it is the clique complex of its 1-skeleton K1{K}^{1}.

Given a simplex σK\sigma\in K, its (open) star St(σ)\mathrm{St}\left(\sigma\right) is the set of all the simplices νK\nu\in K that contain σ\sigma. The open star is not a simplicial complex in general. We also define its closed star St¯(σ)\overline{\mathrm{St}}\left(\sigma\right) as the smallest simplicial subcomplex of KK which contains St(σ)\mathrm{St}\left(\sigma\right).

Geometric realizations.

For every p0p\geq 0, the standard pp-simplex Δp\Delta^{p} is the topological space defined as the convex hull of the canonical basis vectors e1,,ep+1e_{1},...,e_{p+1} of p+1\mathbb{R}^{p+1}, endowed with the subspace topology. To each simplicial complex KK is attached a geometric realization. It is a topological space, denoted |K|\left|K\right|, obtained by gluing the simplices of KK together. According to this construction, each simplex σK\sigma\in K admits a geometric realization |σ|\left|\sigma\right| which is a subset of |K|\left|K\right|. The following set is a partition of |K|\left|K\right|:

{|σ|,σK}.\left\{\left|\sigma\right|,\sigma\in K\right\}.

This allows to define the face map of KK. It is the unique map K:|K|K\mathcal{F}_{K}\colon\left|K\right|\rightarrow K that satisfies x|K(x)|x\in\left|\mathcal{F}_{K}(x)\right| for every x|K|x\in\left|K\right|.

If σ\sigma is a face of KK of dimension at least 1, the subset |σ|\left|\sigma\right| is canonically homeomorphic to the interior of the standard pp-simplex Δp\Delta^{p}, where p=dim(σ)p=\dim(\sigma). This allows to define on |K|\left|K\right| the barycentric coordinates: for every face σ=[v0,,vp]K\sigma=[v_{0},...,v_{p}]\in K, the points x|σ|x\in\left|\sigma\right| can be written as

x=i=0pλivix=\sum_{i=0}^{p}\lambda_{i}v_{i}

with λ0,,λp>0\lambda_{0},...,\lambda_{p}>0 and i=0pλi=1\sum_{i=0}^{p}\lambda_{i}=1.

If XX is any a topological space, a triangulation of XX consists of a simplicial complex KK together with a homeomorphism h:X|K|h\colon X\rightarrow\left|K\right|.

Simplicial approximation.

A simplicial map between simplicial complexes KK and LL is a map between geometric realizations g:|K||L|g\colon\left|K\right|\rightarrow\left|L\right| which sends each simplex of KK to a simplex of LL by a linear maps that sends vertices to vertices. In other words, for every σ=[v0,,vp]K\sigma=[v_{0},...,v_{p}]\in K, the subset [g(v0),,g(vp)][g(v_{0}),...,g(v_{p})] is a simplex of LL, and the map gg restricted to |σ||K|\left|\sigma\right|\subset\left|K\right| can be written in barycentric coordinates as

i=0pλivii=0pλig(vi).\sum_{i=0}^{p}\lambda_{i}v_{i}~{}\longmapsto~{}\sum_{i=0}^{p}\lambda_{i}g(v_{i}). (2)

A simplicial map g:|K||L|g\colon\left|K\right|\rightarrow\left|L\right| is uniquely determined by its restriction to the vertex sets g|K0:K0L0g_{|{K}^{0}}\colon{K}^{0}\rightarrow{L}^{0}. Reciprocally, let f:K0L0f\colon{K}^{0}\rightarrow{L}^{0} be a map between vertex sets which satisfies the following condition:

σK,f(σ)L.\forall\sigma\in K,~{}f(\sigma)\in L. (3)

Then ff induces a simplicial map via barycentric coordinates, denoted |f|:|K||L||f|\colon\left|K\right|\rightarrow\left|L\right|. In the rest of this paper, a simplicial map shall either refer to a map g:|K||L|g\colon\left|K\right|\rightarrow\left|L\right| which satisfies Equation (2), to a map f:K0L0f\colon{K}^{0}\rightarrow{L}^{0} which satisfies Equation (3), or to the induced map f:KLf\colon K\rightarrow L.

Let g:|K||L|g\colon\left|K\right|\rightarrow\left|L\right| be any continuous map. The problem of simplicial approximation consists in finding a simplicial map f:KLf\colon K\rightarrow L with geometric realization |f|:|K||L|\left|f\right|\colon\left|K\right|\rightarrow\left|L\right| homotopy equivalent to gg. A way to solve this problem is to consider the following property (Hatcher, 2002, Proof of Theorem 2C.1): we say that the map gg satisfies the star condition if for every vertex vv of KK, there exists a vertex ww of LL such that

g(|St¯(v)|)|St(w)|.\displaystyle g\left(\left|\overline{\mathrm{St}}\left(v\right)\right|\right)\subseteq\left|\mathrm{St}\left(w\right)\right|.

If this is the case, let f:K0L0f\colon{K}^{0}\rightarrow{L}^{0} be any map between vertex sets such that for every vertex vv of KK, we have g(|St¯(v)|)|St(f(v))|g\left(\left|\overline{\mathrm{St}}\left(v\right)\right|\right)\subseteq\left|\mathrm{St}\left(f(v)\right)\right|. Equivalently, ff satisfies

g(St¯(v))St(f(v)).g\left(\overline{\mathrm{St}}\left(v\right)\right)\subseteq\mathrm{St}\left(f(v)\right).

Such a map is called a simplicial approximation to gg. One shows that it is a simplicial map, and that its geometric realization |f|\left|f\right| is homotopic to gg.

In general, a map gg may not satisfy the star condition. However, there is always a way to subdivise the simplicial complex KK in order to obtain an induced map which does. We describe this construction in the following paragraph.

We point out that, for some authors, such as Munkres (1984), the star condition is defined by the property g(|St(v)|)|St(w)|g\left(\left|\mathrm{St}\left(v\right)\right|\right)\subseteq\left|\mathrm{St}\left(w\right)\right|. The defintion we used above, although harder to satisfy than this one, will be enough for our purposes.

Barycentric subdivisions.

Let Δp\Delta^{p} denote the standard pp-simplex, with vertices denoted v0,,vpv_{0},...,v_{p}. The barycentric subdivision of Δp\Delta^{p} consists in decomposing Δp\Delta^{p} into (p+1)!(p+1)! simplices of dimension pp. It is a simplicial complex, whose vertex set corresponds to the points i=0pλivi\sum_{i=0}^{p}\lambda_{i}v_{i} for which some λi\lambda_{i} are zero and the other ones are equal. Equivalently, one can see this new set of vertices as a the power set of the set of vertices of Δp\Delta^{p}.

More generally, if KK is a simplicial complex, its barycentric subdivision sub(K)\text{sub}(K) is the simplicial complex obtained by subdivising each of its faces. The set of vertices of sub(K)\text{sub}(K) can be seen as a subset of the power set of the set of vertices of KK.

If g:|K||L|g\colon\left|K\right|\rightarrow\left|L\right| is any map, there exists a canonical extended map |sub(K)||L|\left|\text{sub}(K)\right|\rightarrow\left|L\right|, still denoted gg. The simplicial approximation theorem states that for any two simplicial complexes K,LK,L with KK finite, and g:|K||L|g\colon\left|K\right|\rightarrow\left|L\right| any a continuous map, there exists n0n\geq 0 such that g:|subn(K)||L|g\colon\left|\mathrm{sub}^{n}\left(K\right)\right|\rightarrow\left|L\right| satisfies the star condition. As a consequence, such a map g:|subn(K)||L|g\colon\left|\mathrm{sub}^{n}\left(K\right)\right|\rightarrow\left|L\right| admits a simplicial approximation.

2.3 Persistent cohomology

In this subsection, we write down the definitions of persistence modules, and their associated pseudo-distances, in the context of cohomology. Compared to the standard definitions of persistent homology, the arrows go backward. Let T[0,+)T\subseteq[0,+\infty) be an interval that contains 0, let EE be a Euclidean space, and kk a field.

Persistence modules.

A persistence module over TT is a pair (𝕍,𝕧)(\mathbb{V},\mathbbm{v}) where 𝕍=(Vt)tT\mathbb{V}=(V^{t})_{t\in T} is a family of kk-vector spaces, and 𝕧=(vst)stT\mathbbm{v}=(v_{s}^{t})_{s\leq t\in T} is a family of linear maps vst:VsVtv_{s}^{t}\colon V^{s}\leftarrow V^{t} such that:

  • for every tTt\in T, vtt:VtVtv_{t}^{t}\colon V^{t}\leftarrow V^{t} is the identity map,

  • for every r,s,tTr,s,t\in T such that rstr\leq s\leq t, vrsvst=vrtv_{r}^{s}\circ v_{s}^{t}=v_{r}^{t}.

When there is no risk of confusion, we may denote a persistence module by 𝕍\mathbb{V} instead of (𝕍,𝕧)(\mathbb{V},\mathbbm{v}). Given ϵ0\epsilon\geq 0, an ϵ\epsilon-morphism between two persistence modules (𝕍,𝕧)(\mathbb{V},\mathbbm{v}) and (𝕎,𝕨)(\mathbb{W},\mathbbm{w}) is a family of linear maps (ϕt:VtWtϵ)tϵ(\phi_{t}\colon V^{t}\rightarrow W^{t-\epsilon})_{t\geq\epsilon} such that the following diagram commutes for every ϵst\epsilon\leq s\leq t:

Vs{V^{s}}Vt{V^{t}}Wsϵ{W^{s-\epsilon}}Wtϵ{W^{t-\epsilon}}ϕs\scriptstyle{\phi_{s}}vst\scriptstyle{v_{s}^{t}}ϕt\scriptstyle{\phi_{t}}wsϵtϵ\scriptstyle{w_{s-\epsilon}^{t-\epsilon}}

If ϵ=0\epsilon=0 and each ϕt\phi_{t} is an isomorphism, the family (ϕt)tT(\phi_{t})_{t\in T} is an isomorphism of persistence modules. An ϵ\epsilon-interleaving between two persistence modules (𝕍,𝕧)(\mathbb{V},\mathbbm{v}) and (𝕎,𝕨)(\mathbb{W},\mathbbm{w}) is a pair of ϵ\epsilon-morphisms (ϕt:VtWtϵ)tϵ(\phi_{t}\colon V^{t}\rightarrow W^{t-\epsilon})_{t\geq\epsilon} and (ψt:WtVtϵ)tϵ(\psi_{t}\colon W^{t}\rightarrow V^{t-\epsilon})_{t\geq\epsilon} such that the following diagrams commute for every t2ϵt\geq 2\epsilon:

Vt2ϵ{V^{t-2\epsilon}}Vt{V^{t}}Wtϵ{W^{t-\epsilon}}vt2ϵt\scriptstyle{v_{t-2\epsilon}^{t}}ϕt\scriptstyle{\phi_{t}}ψtϵ\scriptstyle{\psi_{t-\epsilon}}
Vtϵ{V^{t-\epsilon}}Wt2ϵ{W^{t-2\epsilon}}Wt{W^{t}}ϕtϵ\scriptstyle{\phi_{t-\epsilon}}wt2ϵt\scriptstyle{w_{t-2\epsilon}^{t}}ψt\scriptstyle{\psi_{t}}

The interleaving pseudo-distance between (𝕍,𝕧)(\mathbb{V},\mathbbm{v}) and (𝕎,𝕨)(\mathbb{W},\mathbbm{w}) is defined as

di(𝕍,𝕎)=inf{ϵ0,𝕍 and 𝕎 are ϵ-interleaved}.\mathrm{d}_{\mathrm{i}}\left(\mathbb{V},\mathbb{W}\right)=\inf\{\epsilon\geq 0,~{}\mathbb{V}\text{ and }\mathbb{W}\text{ are }\epsilon\text{-interleaved}\}.

Persistence barcodes.

A persistence module (𝕍,𝕧)(\mathbb{V},\mathbbm{v}) is said to be pointwise finite-dimensional if for every tTt\in T, VtV^{t} is finite-dimensional. This implies that we can define a notion of persistence barcode (Botnan and Crawley-Boevey, 2020). It comes from the algebraic decomposition of the persistence module into interval modules. Moreover, given two pointwise finite-dimensional persistence modules 𝕍,𝕎\mathbb{V},\mathbb{W} with persistence barcodes Barcode(𝕍)\mathrm{Barcode}\left(\mathbb{V}\right) and Barcode(𝕎)\mathrm{Barcode}\left(\mathbb{W}\right), the so-called isometry theorem states that

db(Barcode(𝕍),Barcode(𝕎))=di(𝕍,𝕎),\displaystyle\mathrm{d}_{\mathrm{b}}\left(\mathrm{Barcode}\left(\mathbb{V}\right),\mathrm{Barcode}\left(\mathbb{W}\right)\right)=\mathrm{d}_{\mathrm{i}}\left(\mathbb{V},\mathbb{W}\right),

where di(,)\mathrm{d}_{\mathrm{i}}\left(\cdot,\cdot\right) denotes the interleaving distance between persistence modules, and db(,)\mathrm{d}_{\mathrm{b}}\left(\cdot,\cdot\right) denotes the bottleneck distance between barcodes.

More generally, the persistence module (𝕍,𝕧)(\mathbb{V},\mathbbm{v}) is said to be qq-tame if for every s,tTs,t\in T such that s<ts<t, the map vstv_{s}^{t} is of finite rank. The qq-tameness of a persistence module ensures that we can still define a notion of persistence barcode, even though the module may not be decomposable into interval modules. Moreover, the isometry theorem still holds (Chazal et al., 2016).

Filtrations of sets and simplicial complexes.

A family of subsets 𝕏=(Xt)tT\mathbb{X}=(X^{t})_{t\in T} of EE is a filtration if it is non-decreasing for the inclusion, i.e. for any s,tTs,t\in T, if sts\leq t then XsXtX^{s}\subseteq X^{t}. Given ϵ0\epsilon\geq 0, two filtrations 𝕏=(Xt)tT\mathbb{X}=(X^{t})_{t\in T} and 𝕐=(Yt)tT\mathbb{Y}=(Y^{t})_{t\in T} of EE are ϵ\epsilon-interleaved if, for every tTt\in T, XtYt+ϵX^{t}\subseteq Y^{t+\epsilon} and YtXt+ϵY^{t}\subseteq X^{t+\epsilon}. The interleaving pseudo-distance between 𝕏\mathbb{X} and 𝕐\mathbb{Y} is defined as the infimum of such ϵ\epsilon:

di(𝕏,𝕐)=inf{ϵ,𝕏 and 𝕐 are ϵ-interleaved}.\mathrm{d}_{\mathrm{i}}\left(\mathbb{X},\mathbb{Y}\right)=\inf\{\epsilon,~{}\mathbb{X}\text{ and }\mathbb{Y}\text{ are }\epsilon\text{-interleaved}\}.

Filtrations of simplicial complexes and their interleaving distance are similarly defined: given a simplicial complex SS, a filtration of SS is a non-decreasing family 𝕊=(St)tT\operatorname{\mathbb{S}}=(S^{t})_{t\in T} of subcomplexes of SS. The interleaving pseudo-distance between two filtrations (S1t)tT(S_{1}^{t})_{t\in T} and (S2t)tT(S_{2}^{t})_{t\in T} of SS is the infimum of the ϵ0\epsilon\geq 0 such that they are ϵ\epsilon-interleaved, i.e., for any tTt\in T, we have S1tS2t+ϵS_{1}^{t}\subseteq S_{2}^{t+\epsilon} and S2tS1t+ϵS_{2}^{t}\subseteq S_{1}^{t+\epsilon}.

Relation between filtrations and persistence modules.

Applying the singular cohomology functor to a set filtration gives rise to a persistence module whose linear maps between cohomology groups are induced by the inclusion maps between sets. As a consequence, if two filtrations are ϵ\epsilon-interleaved, then their associated persistence modules are also ϵ\epsilon-interleaved, the interleaving homomorphisms being induced by the interleaving inclusion maps. As a consequence of the isometry theorem, if the modules are qq-tame, then the bottleneck distance between their persistence barcodes is upperbounded by ϵ\epsilon. The same remarks hold when applying the simplicial cohomology functor to simplicial filtrations.

2.4 Notations

We adopt the following notations:

  • II denotes a set, card(I)\mathrm{card}(I) its cardinal and Ic{I}^{c} its complement.

  • if ii and jj are intergers such that iji\leq j, i,j\llbracket i,j\rrbracket denotes the set of integers between ii and jj included.

  • n\mathbb{R}^{n} and m\mathbb{R}^{m} denote the Euclidean spaces of dimension nn and mm, EE denotes a Euclidean space.

  • M(m)M(\mathbb{R}^{m}) denotes the vector space of m×mm\times m matrices, 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) the Grassmannian of dd-subspaces of m\mathbb{R}^{m}, and 𝕊kk+1\operatorname{\mathbb{S}}_{k}\subset\mathbb{R}^{k+1} the unit kk-sphere.

  • \left\|\cdot\right\| denotes the usual Euclidean norm on n\mathbb{R}^{n}, F\left\|\cdot\right\|_{\mathrm{F}} the Frobenius norm on M(m)M(\mathbb{R}^{m}), γ\left\|\cdot\right\|_{\gamma} the norm on n×M(m)\mathbb{R}^{n}\times M(\mathbb{R}^{m}) defined as (x,A)γ2=x2+γ2AF2\left\|(x,A)\right\|_{\gamma}^{2}=\left\|x\right\|^{2}+\gamma^{2}\left\|A\right\|_{\mathrm{F}}^{2} where γ>0\gamma>0 is a parameter.

  • 𝕏=(Xt)tT\mathbb{X}=(X^{t})_{t\in T} denotes a set filtration. 𝕍[𝕏]\mathbb{V}[\mathbb{X}] denotes the corresponding persistent cohomology module. If XX is a subset of EE, then 𝕏=(Xt)tT\mathbb{X}=(X^{t})_{t\in T} denotes the Čech set filtration of XX (also called the offset filtration).

  • (𝕍,𝕧)(\mathbb{V},\mathbbm{v}) denotes a persistence module over TT, with 𝕍=(Vt)tT\mathbb{V}=(V^{t})_{t\in T} a family of vector spaces, and 𝕧=(vst:XsXt)stT\mathbbm{v}=(v_{s}^{t}\colon X^{s}\leftarrow X^{t})_{s\leq t\in T} a family of linear maps.

  • 𝒰\mathcal{U} denotes a cover of a topological space, and 𝒩(𝒰)\mathcal{N}(\mathcal{U}) its nerve. 𝕊=(St)tT\operatorname{\mathbb{S}}=(S^{t})_{t\in T} denotes a simplicial filtration.

  • If XX is a topological space, H(X)H^{*}(X) denotes its cohomology ring over 2\mathbb{Z}_{2} (the field with two elements), and Hi(X)H^{i}(X) its ithi^{\text{th}} cohomology group over 2\mathbb{Z}_{2}. If f:XYf\colon X\rightarrow Y is a continuous map, f:H(X)H(Y)f^{*}\colon H^{*}(X)\leftarrow H^{*}(Y) is the map induced in cohomology.

  • If ξ\xi is a vector bundle, wi(ξ)w_{i}(\xi) denotes its ithi^{\text{th}} Stiefel-Whitney class.

  • If AA is a subset of EE, then med(A)\mathrm{med}\left(A\right) denotes its medial axis, reach(A)\mathrm{reach}(A) its reach and dist(,A)\mathrm{dist}\left(\cdot,A\right) the distance to AA. The projection on AA is denoted proj(,A)\mathrm{proj}\left(\cdot,A\right) or projA()\mathrm{proj}_{A}\left(\cdot\right). dH(,)\mathrm{d}_{\mathrm{H}}\left(\cdot,\cdot\right) denotes the Hausdorff distance between two sets of EE.

  • If KK is a simplicial complex, Ki{K}^{i} denotes its ii-skeleton. For every vertex vK0v\in{K}^{0}, St(v)\mathrm{St}\left(v\right) and St¯(v)\overline{\mathrm{St}}\left(v\right) denote its open and closed star. The geometric realization of KK is denoted |K|\left|K\right|, and the geometric realization of a simplex σK\sigma\in K is |σ|\left|\sigma\right|. The face map is denoted K:|K|K\mathcal{F}_{K}\colon\left|K\right|\rightarrow K.

  • If f:KLf\colon K\rightarrow L is a simplicial map, |f|:|K||L|\left|f\right|\colon\left|K\right|\rightarrow\left|L\right| denotes its geometric realization. The ithi^{\text{th}} barycentric subdivision of the simplicial complex KK is denoted subi(K)\mathrm{sub}^{i}\left(K\right).

3 Persistent Stiefel-Whitney classes

3.1 Definition

Let E=nE=\mathbb{R}^{n} be a Euclidean space, and 𝕏=(Xt)tT\mathbb{X}=(X^{t})_{t\in T} a set filtration of EE. Let us denote by isti_{s}^{t} the inclusion map from XsX^{s} to XtX^{t}. In order to define persistent Stiefel-Whitney classes, we have to give such a filtration a vector bundle structure. The infinite Grassmann manifold is denoted 𝒢d()\mathcal{G}_{d}(\mathbb{R}^{\infty}).

Definition 3.1.

A vector bundle filtration of dimension dd on EE is a couple (𝕏,𝕡)(\mathbb{X},\mathbbm{p}) where 𝕏=(Xt)tT\mathbb{X}=(X^{t})_{t\in T} is a set filtration of EE and 𝕡=(pt)tT\mathbbm{p}=(p^{t})_{t\in T} a family of continuous maps pt:Xt𝒢d()p^{t}\colon X^{t}\rightarrow\mathcal{G}_{d}(\mathbb{R}^{\infty}) such that, for every s,tTs,t\in T with sts\leq t, we have ptist=psp^{t}\circ i_{s}^{t}=p^{s}. In other words, the following diagram commutes:

Xs{X^{s}}Xt{X^{t}}𝒢d(){\mathcal{G}_{d}(\mathbb{R}^{\infty})}ps\scriptstyle{p^{s}}ist\scriptstyle{i_{s}^{t}}pt\scriptstyle{p^{t}}

Note that for any mm\in\mathbb{N}, and by using the inclusion 𝒢d(m)𝒢d()\mathcal{G}_{d}(\mathbb{R}^{m})\hookrightarrow\mathcal{G}_{d}(\mathbb{R}^{\infty}), one may define a vector bundle filtration by considering maps pt:Xt𝒢d(m)p^{t}\colon X^{t}\rightarrow\mathcal{G}_{d}(\mathbb{R}^{m}).

Following Subsect. 2.1, the induced map in cohomology, (pt)(p^{t})^{*}, allows to define the Stiefel-Whitney classes of this vector bundle. Let us introduce some notations. The Stiefel-Whitney classes of 𝒢d()\mathcal{G}_{d}(\mathbb{R}^{\infty}) are denoted w1,,wdw_{1},\dots,w_{d}. The Stiefel-Whitney classes of the vector bundle (Xt,pt)(X^{t},p^{t}) are denoted w1t(𝕡),,wdt(𝕡)w_{1}^{t}(\mathbbm{p}),\dots,w_{d}^{t}(\mathbbm{p}), and can be defined as wit(𝕡)=(pt)(wi)w_{i}^{t}(\mathbbm{p})=(p^{t})^{*}(w_{i}).

(pt):H(Xt){(p^{t})^{*}\colon H^{*}\big{(}X^{t}\big{)}}H(𝒢d()){H^{*}\big{(}\mathcal{G}_{d}(\mathbb{R}^{\infty})\big{)}}w1t(𝕡){~{}~{}~{}~{}~{}~{}~{}~{}w_{1}^{t}(\mathbbm{p})}w1{w_{1}}wdt(𝕡){~{}~{}~{}~{}~{}~{}~{}~{}w_{d}^{t}(\mathbbm{p})}wd{w_{d}}{\vdots}

Let (𝕍,𝕧)(\mathbb{V},\mathbbm{v}) denote the persistence module associated to the filtration 𝕏\mathbb{X}, with 𝕍=(Vt)tT\mathbb{V}=(V^{t})_{t\in T} and 𝕧=(vst)stT\mathbbm{v}=(v_{s}^{t})_{s\leq t\in T}. Explicitly, VtV^{t} is the cohomology ring H(Xt)H^{*}(X^{t}), and vstv_{s}^{t} is the induced map H(Xs)H(Xt)H^{*}(X^{s})\leftarrow H^{*}(X^{t}). For every tTt\in T, the classes w1t(𝕡),,wdt(𝕡)w_{1}^{t}(\mathbbm{p}),\dots,w_{d}^{t}(\mathbbm{p}) belong to the vector space VtV^{t}. The persistent Stiefel-Whitney classes are defined to be the collection of such classes over tt.

Definition 3.2.

Let (𝕏,𝕡)(\mathbb{X},\mathbbm{p}) be a vector bundle filtration. The persistent Stiefel-Whitney classes of (𝕏,𝕡)(\mathbb{X},\mathbbm{p}) are the families of classes

w1(𝕡)\displaystyle w_{1}(\mathbbm{p}) =(w1t(𝕡))tT\displaystyle=\big{(}w_{1}^{t}(\mathbbm{p})\big{)}_{t\in T}
\displaystyle\vdots
wd(𝕡)\displaystyle w_{d}(\mathbbm{p}) =(wdt(𝕡))tT.\displaystyle=\big{(}w_{d}^{t}(\mathbbm{p})\big{)}_{t\in T}.

Let i1,di\in\llbracket 1,d\rrbracket, and consider a persistent Stiefel-Whitney class wi(𝕡)w_{i}(\mathbbm{p}). Note that it satisfies the following property: for all s,tTs,t\in T such that sts\leq t, wis(𝕡)=vst(wit(𝕡))w_{i}^{s}(\mathbbm{p})=v_{s}^{t}\big{(}w_{i}^{t}(\mathbbm{p})\big{)}. As a consequence, if a class wit(𝕡)w_{i}^{t}(\mathbbm{p}) is given for a tTt\in T, one obtains all the others wis(𝕡)w_{i}^{s}(\mathbbm{p}), with sts\leq t, by applying the maps vstv_{s}^{t}. In particular, if wit(𝕡)=0w_{i}^{t}(\mathbbm{p})=0, then wis(𝕡)=0w_{i}^{s}(\mathbbm{p})=0 for all sTs\in T such that sts\leq t.

Lifebar.

In order to visualize the evolution of a persistent Stiefel-Whitney class through the persistence module (𝕍,𝕧)(\mathbb{V},\mathbbm{v}), we propose the following bar representation: the lifebar of wi(𝕡)w_{i}(\mathbbm{p}) is the set

{tT,wit(𝕡)0}.\displaystyle\left\{t\in T,w_{i}^{t}(\mathbbm{p})\neq 0\right\}.

According to the last paragraph, the lifebar of a persistent class is an interval of TT, of the form [t,sup(T))[t^{\dagger},\sup(T)) or (t,sup(T))(t^{\dagger},\sup(T)), where

t=inf{tT,wit(𝕡)0},\displaystyle t^{\dagger}=\inf\left\{t\in T,w_{i}^{t}(\mathbbm{p})\neq 0\right\},

with the convention inf()=inf(T)\inf(\emptyset)=\inf(T). In order to distinguish the lifebar of a persistent Stiefel-Whitney class from the bars of the persistence barcodes, we draw the rest of the interval hatched (see Figure 1).

Refer to caption
Figure 1: Example of a lifebar of a persistent Stiefel-Whitney class with t=0.2t^{\dagger}=0{.}2 and max(T)=1\max(T)=1.

3.2 Čech bundle filtrations

In this subsection, we propose a particular construction of vector bundle filtration, called the Čech bundle filtration. We shall work in the ambient space E=n×M(m)E=\mathbb{R}^{n}\times M(\mathbb{R}^{m}). Let \left\|\cdot\right\| be the usual Euclidean norm on the space n\mathbb{R}^{n}, and F\left\|\cdot\right\|_{\mathrm{F}} the Frobenius norm on M(m)M(\mathbb{R}^{m}), the space of m×mm\times m matrices. Let γ>0\gamma>0. We endow the vector space EE with the Euclidean norm γ\left\|\cdot\right\|_{\gamma} defined for every (x,A)E(x,A)\in E as

(x,A)γ2=x2+γ2AF2.\left\|(x,A)\right\|_{\gamma}^{2}=\left\|x\right\|^{2}+\gamma^{2}\left\|A\right\|_{\mathrm{F}}^{2}. (4)

See Subsection 5.4 for a discussion about the parameter γ\gamma.

In order to define the Čech bundle filtration, we shall first study the usual embedding of the Grassmann manifold 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) into the matrix space M(m)M(\mathbb{R}^{m}).

Embedding of 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}).

We embed the Grassmannian 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) into M(m)M(\mathbb{R}^{m}) via the application which sends a dd-dimensional subspace TmT\subset\mathbb{R}^{m} to its orthogonal projection matrix PTP_{T}. We can now see 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) as a submanifold of M(m)M(\mathbb{R}^{m}). Recall that M(m)M(\mathbb{R}^{m}) is endowed with the Frobenius norm. According to this metric, 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) is included in the sphere of center 0 and radius d\sqrt{d} of M(m)M(\mathbb{R}^{m}).

In the metric space (M(m),F)(M(\mathbb{R}^{m}),\left\|\cdot\right\|_{\mathrm{F}}), consider the distance function to 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}), denoted dist(,𝒢d(m))\mathrm{dist}\left(\cdot,\mathcal{G}_{d}(\mathbb{R}^{m})\right). Let med(𝒢d(m))\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right) denote the medial axis of 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}). It consists in the points AM(m)A\in M(\mathbb{R}^{m}) which admit at least two projections on 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}):

med(𝒢d(m))={AM(m),P,P\displaystyle\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)=\{A\in M(\mathbb{R}^{m}),\exists P,P^{\prime} 𝒢d(m),PP,\displaystyle\in\mathcal{G}_{d}(\mathbb{R}^{m}),P\neq P^{\prime},
APF=APF=dist(A,𝒢d(m))}.\displaystyle\left\|A-P\right\|_{\mathrm{F}}=\left\|A-P\right\|_{\mathrm{F}}=\mathrm{dist}\left(A,\mathcal{G}_{d}(\mathbb{R}^{m})\right)\}.

On the set M(m)med(𝒢d(m))M(\mathbb{R}^{m})\setminus\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right), the projection on 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) is well-defined:

proj(,𝒢d(m)):M(m)med(𝒢d(m))\displaystyle\mathrm{proj}\left(\cdot,\mathcal{G}_{d}(\mathbb{R}^{m})\right)\colon M(\mathbb{R}^{m})\setminus\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right) 𝒢d(m)M(m)\displaystyle\longrightarrow\mathcal{G}_{d}(\mathbb{R}^{m})\subset M(\mathbb{R}^{m})
A\displaystyle A P s.t. PAF=dist(A,𝒢d(m)).\displaystyle\longmapsto P~{}\text{ s.t. }\left\|P-A\right\|_{\mathrm{F}}=\mathrm{dist}\left(A,\mathcal{G}_{d}(\mathbb{R}^{m})\right).

The following lemma describes this projection explicitly.

Lemma 3.1.

For any AM(m)A\in M(\mathbb{R}^{m}), let AsA^{s} denote the matrix As=12(A+A)A^{s}=\frac{1}{2}(A+A^{\prime}), where AA^{\prime} is the transpose of AA, and let λ1(As),,λm(As)\lambda_{1}(A^{s}),...,\lambda_{m}(A^{s}) be the eigenvalues of AsA^{s} in decreasing order. The distance from AA to med(𝒢d(m))\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right) is

dist(A,med(𝒢d(m)))=22|λd(As)λd+1(As)|.\displaystyle\mathrm{dist}\left(A,\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)\right)=\frac{\sqrt{2}}{2}\big{|}\lambda_{d}(A^{s})-\lambda_{d+1}(A^{s})\big{|}.

If this distance is positive, the projection of AA on 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) can be described as follows: consider the symmetric matrix AsA^{s}, and let As=ODOA^{s}=ODO^{\prime}, with OO an orthogonal matrix, and DD the diagonal matrix containing the eigenvalues of AsA^{s} in decreasing order. Let JdJ_{d} be the diagonal matrix whose first dd terms are 1, and the other ones are zero. We have

proj(A,𝒢d(m))=OJdO.\displaystyle\mathrm{proj}\left(A,\mathcal{G}_{d}(\mathbb{R}^{m})\right)=OJ_{d}O^{\prime}.
Proof.

Note that 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) is contained in the linear subspace SS\SS of symmetric matrices. Therefore, to project a matrix AM(m)A\in M(\mathbb{R}^{m}) onto 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}), we may project on SS\SS first. It is well known that the projection of AA onto SS\SS is the matrix As=12(A+A)A^{s}=\frac{1}{2}(A+A^{\prime}).

Suppose now that we are given a symmetric matrix BB. Let it be diagonalized as B=ODOB=ODO^{\prime} with OO an orthogonal matrix. A projection of BB onto 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) is a matrix PP which minimizes the following quantity:

minP𝒢d(E)BPF.\min_{P\in\mathcal{G}_{d}(E)}\left\|B-P\right\|_{\mathrm{F}}. (5)

By applying the transformation POPOP\mapsto O^{\prime}PO, we see that this problem is equivalent to minP𝒢d(E)DPF\min_{P\in\mathcal{G}_{d}(E)}\left\|D-P\right\|_{\mathrm{F}}. Now, let e1,,eme_{1},\cdots,e_{m} denote the canonical basis of m\mathbb{R}^{m}. We have

DPF2\displaystyle\left\|D-P\right\|_{\mathrm{F}}^{2} =DF2+PF22D,PF\displaystyle=\left\|D\right\|_{\mathrm{F}}^{2}+\left\|P\right\|_{\mathrm{F}}^{2}-2\langle D,P\rangle_{\mathrm{F}}
=DF2+PF22λiei,P(ei),\displaystyle=\left\|D\right\|_{\mathrm{F}}^{2}+\left\|P\right\|_{\mathrm{F}}^{2}-2\sum\left\langle\lambda_{i}e_{i},P(e_{i})\right\rangle,

where ,F\langle\cdot,\cdot\rangle_{\mathrm{F}} is the Frobenius inner product, and ,\left\langle\cdot,\cdot\right\rangle the usual inner product on m\mathbb{R}^{m}. Therefore, Equation (5) is a problem equivalent to

maxP𝒢d(E)λiei,P(ei).\displaystyle\max_{P\in\mathcal{G}_{d}(E)}\sum\lambda_{i}\left\langle e_{i},P(e_{i})\right\rangle.

Since PP is an orthogonal projection, we have ei,P(ei)=P(ei),P(ei)=P(ei)2\left\langle e_{i},P(e_{i})\right\rangle=\left\langle P(e_{i}),P(e_{i})\right\rangle=\left\|P(e_{i})\right\|^{2} for all i1,mi\in\llbracket 1,m\rrbracket. Moreover, d=PF2=P(ei)2d=\left\|P\right\|_{\mathrm{F}}^{2}=\sum\left\|P(e_{i})\right\|^{2}. Denoting pi=P(ei)2[0,1]p_{i}=\left\|P(e_{i})\right\|^{2}\in[0,1], we finally obtain the following alternative formulation of Equation (5):

maxp1,pm[0,1]p1++pm=dλipi.\displaystyle\max_{\begin{subarray}{c}p_{1},...p_{m}\in[0,1]\\ p_{1}+...+p_{m}=d\end{subarray}}\sum\lambda_{i}p_{i}.

Using that λ1λm\lambda_{1}\geq...\geq\lambda_{m}, we see that this maximum is attained when p1==pd=1p_{1}=...=p_{d}=1 and pd+1==pm=0p_{d+1}=...=p_{m}=0. Consequently, a minimizer of Equation (5) is P=JdP=J_{d}, where JdJ_{d} is the diagonal matrix whose first dd terms are 1, and the other ones are zero. Moreover, it is unique if λdλd+1\lambda_{d}\neq\lambda_{d+1}. As a consequence of these considerations, we obtain the following characterization: for every BM(m)B\in M(\mathbb{R}^{m}),

Bmed(𝒢d(m))λd(Bs)=λd+1(Bs).B\in\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)\iff\lambda_{d}(B^{s})=\lambda_{d+1}(B^{s}). (6)

Let us now show that for every matrix AM(m)A\in M(\mathbb{R}^{m}), we have

dist(A,med(𝒢d(m)))=22|λd(As)λd+1(As)|.\displaystyle\mathrm{dist}\left(A,\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)\right)=\frac{\sqrt{2}}{2}\big{|}\lambda_{d}(A^{s})-\lambda_{d+1}(A^{s})\big{|}.

First, remark that

dist(A,med(𝒢d(m)))=dist(As,med(𝒢d(m))).\mathrm{dist}\left(A,\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)\right)=\mathrm{dist}\left(A^{s},\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)\right). (7)

Indeed, if BB is a projection of AA on med(𝒢d(m))\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right), then BsB^{s} is still in med(𝒢d(m))\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right) according to Equation (6), and

dist(A,med(𝒢d(m)))=ABFAsBsFdist(As,med(𝒢d(m))).\displaystyle\mathrm{dist}\left(A,\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)\right)=\left\|A-B\right\|_{\mathrm{F}}\geq\left\|A^{s}-B^{s}\right\|_{\mathrm{F}}\geq\mathrm{dist}\left(A^{s},\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)\right).

Conversely, if BB is a projection of AsA^{s} on med(𝒢d(m))\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right), then B^=B+AAs\hat{B}=B+A-A^{s} is still in med(𝒢d(m))\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right), and

dist(A,med(𝒢d(m)))AB^F=AsBF=dist(As,med(𝒢d(m))).\displaystyle\mathrm{dist}\left(A,\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)\right)\leq\left\|A-\hat{B}\right\|_{\mathrm{F}}=\left\|A^{s}-B\right\|_{\mathrm{F}}=\mathrm{dist}\left(A^{s},\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)\right).

We deduce Equation (7). Now, let ASSA\in\SS and Bmed(𝒢d(m))B\in\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right). Let e1,,eme_{1},...,e_{m} be a basis of m\mathbb{R}^{m} that diagonalizes AA. Writing ABF=A(ei)B(ei)2=λi(A)eiB(ei)2\left\|A-B\right\|_{\mathrm{F}}=\sum\left\|A(e_{i})-B(e_{i})\right\|^{2}=\sum\left\|\lambda_{i}(A)e_{i}-B(e_{i})\right\|^{2}, it is clear that the closest matrix BB must satisfy B(ei)=λi(B)eiB(e_{i})=\lambda_{i}(B)e_{i}, with

  • λi(B)=λi(A)\lambda_{i}(B)=\lambda_{i}(A) for i{d,d+1}i\notin\{d,d+1\},

  • λd(B)=λd+1(B)=12(λd(A)+λd+1(A))\lambda_{d}(B)=\lambda_{d+1}(B)=\frac{1}{2}(\lambda_{d}(A)+\lambda_{d+1}(A)).

We finally compute

ABF2\displaystyle\left\|A-B\right\|_{\mathrm{F}}^{2} =λi(A)eiλi(B)ei2\displaystyle=\sum\left\|\lambda_{i}(A)e_{i}-\lambda_{i}(B)e_{i}\right\|^{2}
=|λd(A)λd(B)|2+|λd+1(A)λd+1(B)|2\displaystyle=\left\lvert\lambda_{d}(A)-\lambda_{d}(B)\right\rvert^{2}+\left\lvert\lambda_{d+1}(A)-\lambda_{d+1}(B)\right\rvert^{2}
=12|λd(A)λd+1(A)|2\displaystyle=\frac{1}{2}\left\lvert\lambda_{d}(A)-\lambda_{d+1}(A)\right\rvert^{2}

which yields the result. ∎

Observe that, as a consequence of this lemma, every point of 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) is at equal distance from med(𝒢d(m))\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right), and this distance is equal to 22\frac{\sqrt{2}}{2}. Therefore the reach of the subset 𝒢d(m)M(m)\mathcal{G}_{d}(\mathbb{R}^{m})\subset M(\mathbb{R}^{m}) is

reach(𝒢d(m))=22.\displaystyle\mathrm{reach}(\mathcal{G}_{d}(\mathbb{R}^{m}))=\frac{\sqrt{2}}{2}.

Čech bundle filtration.

Let XX be a subset of E=n×M(m)E=\mathbb{R}^{n}\times M(\mathbb{R}^{m}). Consider the usual Čech filtration 𝕏=(Xt)t0\mathbb{X}=(X^{t})_{t\geq 0}, where XtX^{t} denotes the tt-thickening of Xˇ\check{X} in the metric space (E,γ)(E,\left\|\cdot\right\|_{\gamma}). It is also known as the offset filtration. In order to give this filtration a vector bundle structure, consider the map ptp^{t} defined as the composition

Xtn×M(m){X^{t}\subset\mathbb{R}^{n}\times M(\mathbb{R}^{m})}M(m)med(𝒢d(m)){M(\mathbb{R}^{m})\setminus\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)}𝒢d(m),{\mathcal{G}_{d}(\mathbb{R}^{m}),}proj2\scriptstyle{\mathrm{proj}_{2}}proj(,𝒢d(m))\scriptstyle{\mathrm{proj}\left(\cdot,\mathcal{G}_{d}(\mathbb{R}^{m})\right)} (8)

where proj2\mathrm{proj}_{2} represents the projection on the second coordinate of n×M(m)\mathbb{R}^{n}\times M(\mathbb{R}^{m}), and proj(,𝒢d(m))\mathrm{proj}\left(\cdot,\mathcal{G}_{d}(\mathbb{R}^{m})\right) the projection on 𝒢d(m)M(m)\mathcal{G}_{d}(\mathbb{R}^{m})\subset M(\mathbb{R}^{m}). Note that ptp^{t} is well-defined only when XtX^{t} does not intersect n×med(𝒢d(m))\mathbb{R}^{n}\times\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right). The supremum of such tt’s is denoted tγmax(X)t_{\gamma}^{\mathrm{max}}\left(X\right). We have

tγmax(X)=inf{distγ(x,n×med(𝒢d(m))),xX},t_{\gamma}^{\mathrm{max}}\left(X\right)=\inf\left\{\mathrm{dist}_{\gamma}\left(x,\mathbb{R}^{n}\times\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)\right),~{}x\in X\right\}, (9)

where distγ(x,n×med(𝒢d(m)))\mathrm{dist}_{\gamma}\left(x,\mathbb{R}^{n}\times\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)\right) is the distance between the point xn×M(m)x\in\mathbb{R}^{n}\times M(\mathbb{R}^{m}) and the subspace n×med(𝒢d(m))\mathbb{R}^{n}\times\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right), with respect to the norm γ\left\|\cdot\right\|_{\gamma}. By definition of γ\left\|\cdot\right\|_{\gamma}, Equation (9) rewrites as

tγmax(X)=γinf{dist(A,med(𝒢d(m))),(y,A)X},t_{\gamma}^{\mathrm{max}}\left(X\right)=\gamma\cdot\inf\{\mathrm{dist}\left(A,\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)\right),~{}(y,A)\in X\},

where dist(A,med(𝒢d(m)))\mathrm{dist}\left(A,\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)\right) represents the distance between the matrix AA and the subset med(𝒢d(m))\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right) with respect to the Frobenius norm F\left\|\cdot\right\|_{\mathrm{F}}. Denoting tmax(X)t^{\mathrm{max}}\left(X\right) the value tγmax(X)t_{\gamma}^{\mathrm{max}}\left(X\right) for γ=1\gamma=1, we obtain

tγmax(X)=γtmax(X)andtmax(X)=inf{dist(A,med(𝒢d(m))),(y,A)X}.\displaystyle\begin{split}t_{\gamma}^{\mathrm{max}}\left(X\right)&=\gamma\cdot t^{\mathrm{max}}\left(X\right)\\ \mathrm{and}~{}~{}~{}~{}~{}~{}~{}t^{\mathrm{max}}\left(X\right)&=\inf\{\mathrm{dist}\left(A,\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)\right),~{}(y,A)\in X\}.\end{split} (10)

Note that the values tmax(X)t^{\mathrm{max}}\left(X\right) can be computed explicitly thanks to Lemma 3.1. In particular, if XX is a subset of n×𝒢d(m)\mathbb{R}^{n}\times\mathcal{G}_{d}(\mathbb{R}^{m}), then tmax(X)=22t^{\mathrm{max}}\left(X\right)=\frac{\sqrt{2}}{2}. Accordingly,

tγmax(X)=22γ.t_{\gamma}^{\mathrm{max}}\left(X\right)=\frac{\sqrt{2}}{2}\gamma. (11)
Definition 3.3.

Consider a subset XX of E=n×M(m)E=\mathbb{R}^{n}\times M(\mathbb{R}^{m}), and suppose that tmax(X)>0t^{\mathrm{max}}\left(X\right)>0. The Čech bundle filtration associated to XX in the ambient space (E,γ)(E,\left\|\cdot\right\|_{\gamma}) is the vector bundle filtration (𝕏,𝕡)(\mathbb{X},\mathbbm{p}) consisting of the Čech filtration 𝕏=(Xt)tT\mathbb{X}=(X^{t})_{t\in T}, and the maps 𝕡=(pt)tT\mathbbm{p}=(p^{t})_{t\in T} as defined in Equation (8). This vector bundle filtration is defined on the index set T=[0,tγmax(X))T=\left[0,t_{\gamma}^{\mathrm{max}}\left(X\right)\right), where tγmax(X)t_{\gamma}^{\mathrm{max}}\left(X\right) is defined in Equation (10).

The ithi^{\text{th}} persistent Stiefel-Whitney class of the Čech bundle filtration (𝕏,𝕡)(\mathbb{X},\mathbbm{p}), as in Definition 3.2, shall be denoted wi(X)w_{i}(X) instead of wi(𝕡)w_{i}(\mathbbm{p}).

Example 3.2.

Let E=2×M(2)E=\mathbb{R}^{2}\times M(\mathbb{R}^{2}). Let XX and YY be the subsets of EE defined as:

X={((cos(θ)sin(θ)),(cos(θ)2cos(θ)sin(θ)cos(θ)sin(θ)sin(θ)2)),θ[0,2π)}\displaystyle X=\bigg{\{}\bigg{(}\begin{pmatrix}\cos(\theta)\\ \sin(\theta)\end{pmatrix},\begin{pmatrix}\cos(\theta)^{2}&\cos(\theta)\sin(\theta)\\ \cos(\theta)\sin(\theta)&\sin(\theta)^{2}\end{pmatrix}\bigg{)},\theta\in[0,2\pi)\bigg{\}}
Y={((cos(θ)sin(θ)),(cos(θ2)2cos(θ2)sin(θ2)cos(θ2)sin(θ2)sin(θ2)2)),θ[0,2π)}\displaystyle Y=\bigg{\{}\bigg{(}\begin{pmatrix}\cos(\theta)\\ \sin(\theta)\end{pmatrix},\begin{pmatrix}\cos(\frac{\theta}{2})^{2}&\cos(\frac{\theta}{2})\sin(\frac{\theta}{2})\\ \cos(\frac{\theta}{2})\sin(\frac{\theta}{2})&\sin(\frac{\theta}{2})^{2}\end{pmatrix}\bigg{)},\theta\in[0,2\pi)\bigg{\}}

The set XX is to be seen as the normal bundle of the circle, and YY as the tautological bundle of the circle, also known as the Mobius strip. They are pictured in Figures 2 and 3. We have tmax(X)=tmax(Y)=22t^{\mathrm{max}}\left(X\right)=t^{\mathrm{max}}\left(Y\right)=\frac{\sqrt{2}}{2} as in Lemma 3.1. Let γ=1\gamma=1.

Refer to caption
Refer to caption
Figure 2: Representation of the sets XX and Y2×M(2)Y\subset\mathbb{R}^{2}\times M(\mathbb{R}^{2}): the black points correspond to the 2\mathbb{R}^{2}-coordinate, and the pink segments over them correspond to the orientation of the M(2)M(\mathbb{R}^{2})-coordinate.
Refer to caption
Refer to caption
Figure 3: The sets XX and Y2×M(2)Y\subset\mathbb{R}^{2}\times M(\mathbb{R}^{2}), projected in a 3-dimensional subspace of 3\mathbb{R}^{3} via Principal Component Analysis.

We now compute the persistence barcodes of the Čech filtrations of XX and YY in the ambient space EE, as represented in Figure 4.

Refer to captionRefer to caption
Refer to captionRefer to caption
Figure 4: H0H^{0} and H1H^{1} persistence barcodes of the Čech filtration of XX (left) and YY (right).

Consider the first persistent Stiefel-Whitney classes w1(X)w_{1}(X) and w1(Y)w_{1}(Y) of the corresponding Čech bundle filtrations. We compute that their lifebars are \emptyset for w1(X)w_{1}(X), and [0,tmax(Y))\left[0,t^{\mathrm{max}}\left(Y\right)\right) for w1(Y)w_{1}(Y). This is illustrated in Figure 5. One reads these bars as follows: w1t(X)w_{1}^{t}(X) is zero for every t[0,22)t\in\left[0,\frac{\sqrt{2}}{2}\right), while w1t(Y)w_{1}^{t}(Y) is nonzero.

Refer to caption
Refer to caption
Figure 5: Lifebars of the first persistent Stiefel-Whitney classes w1(X)w_{1}(X) and w1(Y)w_{1}(Y).

3.3 Stability

In this subsection we derive a straightforward stability result for persistent Stiefel-Whitney classes. We start by defining a notion of interleavings for vector bundle filtrations, in the same vein as the usual interleavings of set filtrations.

Definition 3.4.

Let ϵ0\epsilon\geq 0, and consider two vector bundle filtrations (𝕏,𝕡)(\mathbb{X},\mathbbm{p}), (𝕐,𝕢)(\mathbb{Y},\mathbbm{q}) of dimension dd on EE with respective index sets TT and UU. They are ϵ\epsilon-interleaved if the underlying filtrations 𝕏=(Xt)tT\mathbb{X}=(X^{t})_{t\in T} and 𝕐=(Yt)tU\mathbb{Y}=(Y^{t})_{t\in U} are ϵ\epsilon-interleaved, and if the following diagrams commute for every tT(Uϵ)t\in T\cap(U-\epsilon) and sU(Tϵ)s\in U\cap(T-\epsilon):

Xt{X^{t}}Yt+ϵ{Y^{t+\epsilon}}𝒢d(){\mathcal{G}_{d}(\mathbb{R}^{\infty})}pt\scriptstyle{p^{t}}qt+ϵ\scriptstyle{q^{t+\epsilon}}
Ys{Y^{s}}Xs+ϵ{X^{s+\epsilon}}𝒢d(){\mathcal{G}_{d}(\mathbb{R}^{\infty})}qs\scriptstyle{q^{s}}qs+ϵ\scriptstyle{q^{s+\epsilon}}

The following theorem shows that interleavings of vector bundle filtrations give rise to interleavings of persistence modules which respect the persistent Stiefel-Whitney classes.

Theorem 3.3.

Consider two vector bundle filtrations (𝕏,𝕡)(\mathbb{X},\mathbbm{p}), (𝕐,𝕢)(\mathbb{Y},\mathbbm{q}) of dimension dd with respective index sets TT and UU. Suppose that they are ϵ\epsilon-interleaved. Then there exists an ϵ\epsilon-interleaving (ϕ,ψ)(\phi,\psi) between their corresponding persistent cohomology modules which sends persistent Stiefel-Whitney classes on persistent Stiefel-Whitney classes. In other words, for every i1,di\in\llbracket 1,d\rrbracket, and for every t(T+ϵ)Ut\in(T+\epsilon)\cap U and sU(T+ϵ)s\in U\cap(T+\epsilon), we have

ϕt(wit(𝕡))=witϵ(𝕢)\displaystyle\phi^{t}(w_{i}^{t}(\mathbbm{p}))=w_{i}^{t-\epsilon}(\mathbbm{q})
and ψs(wis(𝕡))=wisϵ(𝕢).\displaystyle\psi^{s}(w_{i}^{s}(\mathbbm{p}))=w_{i}^{s-\epsilon}(\mathbbm{q}).
Proof.

Define (ϕ,ψ)(\phi,\psi) to be the ϵ\epsilon-interleaving between the cohomology persistence modules 𝕍(𝕏)\mathbb{V}(\mathbb{X}) and 𝕍(𝕐)\mathbb{V}(\mathbb{Y}) given by the ϵ\epsilon-interleaving between the filtrations 𝕏\mathbb{X} and 𝕐\mathbb{Y}. Explicitly, if itt+ϵi_{t}^{t+\epsilon} denotes the inclusion XtYt+ϵX^{t}\hookrightarrow Y^{t+\epsilon} and jss+ϵj_{s}^{s+\epsilon} denotes the inclusion YsXs+ϵY^{s}\hookrightarrow X^{s+\epsilon}, then ϕ=(ϕt)t(T+ϵ)U\phi=(\phi^{t})_{t\in(T+\epsilon)\cap U} is given by the induced maps in cohomology ϕt=(itϵt)\phi^{t}=(i_{t-\epsilon}^{t})^{*}, and ψ=(ψs)s(U+ϵ)T\psi=(\psi^{s})_{s\in(U+\epsilon)\cap T} is given by ψs=(jsϵs)\psi^{s}=(j_{s-\epsilon}^{s})^{*}.

Now, by fonctoriality, the diagrams of Definition 3.4 give rise to commutative diagrams in cohomology:

H(Xtϵ){H^{*}(X^{t-\epsilon})}H(Yt){H^{*}(Y^{t})}H(𝒢d()){H^{*}(\mathcal{G}_{d}(\mathbb{R}^{\infty}))}ϕt\scriptstyle{\phi^{t}}(ptϵ)\scriptstyle{(p^{t-\epsilon})^{*}}(qt)\scriptstyle{(q^{t})^{*}}
H(Ysϵ){H^{*}(Y^{s-\epsilon})}H(Xs){H^{*}(X^{s})}H(𝒢d()){H^{*}(\mathcal{G}_{d}(\mathbb{R}^{\infty}))}ψs\scriptstyle{\psi^{s}}(qsϵ)\scriptstyle{(q^{s-\epsilon})^{*}}(ps)\scriptstyle{(p^{s})^{*}}

Let i1,di\in\llbracket 1,d\rrbracket. By definition, the persistent Stiefel-Whitney classes wi(𝕡)=(wit(𝕡))tTw_{i}(\mathbbm{p})=(w_{i}^{t}(\mathbbm{p}))_{t\in T} and wi(𝕢)=(wis(𝕢))sUw_{i}(\mathbbm{q})=(w_{i}^{s}(\mathbbm{q}))_{s\in U} are equal to wit(𝕡)=(pt)(wi)w_{i}^{t}(\mathbbm{p})=(p^{t})^{*}(w_{i}) and wis(𝕢)=(qs)(wi)w_{i}^{s}(\mathbbm{q})=(q^{s})^{*}(w_{i}), where wiw_{i} is the ithi^{\text{th}} Stiefel-Whitney class of 𝒢d()\mathcal{G}_{d}(\mathbb{R}^{\infty}). The previous commutative diagrams then translates as ϕt(wit(𝕡))=witϵ(𝕢)\phi^{t}(w_{i}^{t}(\mathbbm{p}))=w_{i}^{t-\epsilon}(\mathbbm{q}) and ψs(wis(𝕡))=wisϵ(𝕢)\psi^{s}(w_{i}^{s}(\mathbbm{p}))=w_{i}^{s-\epsilon}(\mathbbm{q}), as wanted. ∎

Consider two vector bundle filtrations (𝕏,𝕡)(\mathbb{X},\mathbbm{p}), (𝕐,𝕢)(\mathbb{Y},\mathbbm{q}) such that there exists an ϵ\epsilon-interleaving (ϕ,ψ)(\phi,\psi) between their persistent cohomology modules 𝕍(𝕏)\mathbb{V}(\mathbb{X}), 𝕍(𝕐)\mathbb{V}(\mathbb{Y}) which sends persistent Stiefel-Whitney classes on persistent Stiefel-Whitney classes. Let i1,di\in\llbracket 1,d\rrbracket. Then the lifebars of their ithi^{\text{th}} persistent Stiefel-Whitney classes wi(𝕡)w_{i}(\mathbbm{p}) and wi(𝕢)w_{i}(\mathbbm{q}) are ϵ\epsilon-close in the following sense: if we denote t(𝕡)=inf{tT,wit(𝕡)0}t^{\dagger}(\mathbbm{p})=\inf\{t\in T,w_{i}^{t}(\mathbbm{p})\neq 0\} and t(𝕢)=inf{tT,wit(𝕢)0}t^{\dagger}(\mathbbm{q})=\inf\{t\in T,w_{i}^{t}(\mathbbm{q})\neq 0\}, then |t(𝕡)t(𝕢)|ϵ|t^{\dagger}(\mathbbm{p})-t^{\dagger}(\mathbbm{q})|\leq\epsilon. This can be seen from their lifebar representations, as shown in Figure 6.

Refer to caption
Refer to caption
Figure 6: Two ϵ\epsilon-close lifebars, with ϵ=0.1\epsilon=0{.}1.

Let us apply this result to the Čech bundle filtrations. Let XX and YY be two subsets of E=n×M(m)E=\mathbb{R}^{n}\times M(\mathbb{R}^{m}). Suppose that the Hausdorff distance dH(X,Y)\mathrm{d}_{\mathrm{H}}\left(X,Y\right), with respect to the norm γ\left\|\cdot\right\|_{\gamma}, is lower than ϵ\epsilon, meaning that the ϵ\epsilon-thickenings XϵX^{\epsilon} and YϵY^{\epsilon} satisfy YXϵY\subseteq X^{\epsilon} and XYϵX\subseteq Y^{\epsilon}. It is then clear that the vector bundle filtrations are ϵ\epsilon-interleaved, and we can apply Theorem 3.3 to obtain the following result.

Corollary 3.4.

If two subsets X,YEX,Y\subset E satisfy dH(X,Y)ϵ\mathrm{d}_{\mathrm{H}}\left(X,Y\right)\leq\epsilon, then there exists an ϵ\epsilon-interleaving between the persistent cohomology modules of their corresponding Čech bundle filtrations which sends persistent Stiefel-Whitney classes on persistent Stiefel-Whitney classes.

Example 3.5.

In order to illustrate Corollary 3.4, consider the sets XX^{\prime} and YY^{\prime} represented in Figure 7. They are noisy samples of the sets XX and YY defined in Example 3.2. They contain 50 points each.

Figure 8 represents the barcodes of the Čech filtrations of the sets XX^{\prime} and YY^{\prime}, together with the lifebar of the first persistent Stiefel-Whitney class of their corresponding Čech bundle filtrations. We see that they are close to the original descriptors of XX and YY (Figure 5). Experimentally, we computed that the Hausdorff distances between X,XX,X^{\prime} and Y,YY,Y^{\prime} are approximately dH(X,X)0.5\mathrm{d}_{\mathrm{H}}\left(X,X^{\prime}\right)\approx 0{.}5 and dH(Y,Y)0.4\mathrm{d}_{\mathrm{H}}\left(Y,Y^{\prime}\right)\approx 0{.}4. We observe that this is coherent with the lifebar of w1(Y)w_{1}(Y^{\prime}), which is ϵ\epsilon-close to the lifebar of w1(Y)w_{1}(Y) with ϵ0.30.4\epsilon\approx 0{.}3\leq 0{.}4.

Refer to caption
Refer to caption
Figure 7: Representation of the sets X,Y2×M(2)X^{\prime},Y^{\prime}\subset\mathbb{R}^{2}\times M(\mathbb{R}^{2}). The black points correspond to the 2\mathbb{R}^{2}-coordinate, and the pink segments over them correspond to the orientation of the M(2)M(\mathbb{R}^{2})-coordinate.
Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Figure 8: Left: H0H^{0} and H1H^{1} barcodes of XX^{\prime} and lifebar of w1(X)w_{1}(X^{\prime}). Right: same for YY^{\prime}. Only bars of length larger than 0.050.05 are represented.

3.4 Consistency

In this subsection we describe a setting where the persistent Stiefel-Whitney classes wi(X)w_{i}(X) of the Čech bundle filtration of a set XX can be seen as consistent estimators of the Stiefel-Whitney classes of some underlying vector bundle.

Let 0\mathcal{M}_{0} be a compact 𝒞3\mathcal{C}^{3}-manifold, and u:0nu\colon\mathcal{M}_{0}\rightarrow\mathcal{M}\subset\mathbb{R}^{n} an immersion. Suppose that 0\mathcal{M}_{0} is given a dd-dimensional vector bundle structure p:0𝒢d(m)p\colon\mathcal{M}_{0}\rightarrow\mathcal{G}_{d}(\mathbb{R}^{m}). Let E=n×M(m)E=\mathbb{R}^{n}\times M(\mathbb{R}^{m}), and consider the set

ˇ={(u(x0),Pp(x0)),x00}E,\check{\mathcal{M}}=\left\{\left(u(x_{0}),P_{p(x_{0})}\right),x_{0}\in\mathcal{M}_{0}\right\}\subset E, (12)

where Pp(x0)P_{p(x_{0})} denotes the orthogonal projection matrix onto the subspace p(x0)mp(x_{0})\subset\mathbb{R}^{m}. The set ˇ\check{\mathcal{M}} is called the lift of 0\mathcal{M}_{0}. Consider the lifting map defined as

uˇ:0ˇEx0(u(x0),Pp(x0)).\displaystyle\begin{split}\check{u}\colon\mathcal{M}_{0}&\longrightarrow\check{\mathcal{M}}\subset E\\ x_{0}&\longmapsto\left(u(x_{0}),P_{p(x_{0})}\right).\end{split} (13)

We make the following assumption: uˇ\check{u} is an embedding. As a consequence, ˇ\check{\mathcal{M}} is a submanifold of EE, and 0\mathcal{M}_{0} and ˇ\check{\mathcal{M}} are diffeomorphic.

The persistent cohomology of ˇ\check{\mathcal{M}} can be used to recover the cohomology of 0\mathcal{M}_{0}. To see this, select γ>0\gamma>0, and denote by reach(ˇ)\mathrm{reach}(\check{\mathcal{M}}) the reach of \mathcal{M}, where EE is endowed with the norm γ\left\|\cdot\right\|_{\gamma}. Since ˇ\check{\mathcal{M}} is a 𝒞2\mathcal{C}^{2}-submanifold, reach(ˇ)\mathrm{reach}(\check{\mathcal{M}}) is positive. Note that it depends on γ\gamma. Let V[ˇ]=(ˇt)t0V[\check{\mathcal{M}}]=(\check{\mathcal{M}}^{t})_{t\geq 0} be the Čech set filtration of ˇ\check{\mathcal{M}} in the ambient space (E,γ)(E,\left\|\cdot\right\|_{\gamma}), and let 𝕍(ˇ)\mathbb{V}(\check{\mathcal{M}}) be the corresponding persistent cohomology module. For every s,t[0,reach(ˇ))s,t\in[0,\mathrm{reach}(\check{\mathcal{M}})) such that sts\leq t, we know that the inclusion maps ist:ˇsˇti_{s}^{t}\colon\check{\mathcal{M}}^{s}\hookrightarrow\check{\mathcal{M}}^{t} are homotopy equivalences. Hence the persistence module 𝕍(ˇ)\mathbb{V}(\check{\mathcal{M}}) is constant on the interval [0,reach(ˇ))[0,\mathrm{reach}(\check{\mathcal{M}})), and is equal to the cohomology H(ˇ)=H(0)H^{*}(\check{\mathcal{M}})=H^{*}(\mathcal{M}_{0}).

Consider the Čech bundle filtration (V[ˇ],𝕡)(V[\check{\mathcal{M}}],\mathbbm{p}) of ˇ\check{\mathcal{M}}. The following theorem shows that the persistent Stiefel-Whitney classes wit(ˇ)w_{i}^{t}(\check{\mathcal{M}}) are also equal to the usual Stiefel-Whitney classes of the vector bundle (0,p)(\mathcal{M}_{0},p).

Theorem 3.6.

Let 0\mathcal{M}_{0} be a compact 𝒞3\mathcal{C}^{3}-manifold, u:0nu\colon\mathcal{M}_{0}\rightarrow\mathbb{R}^{n} an immersion and p:0𝒢d(m)p\colon\mathcal{M}_{0}\rightarrow\mathcal{G}_{d}(\mathbb{R}^{m}) a continuous map. Let ˇ\check{\mathcal{M}} be the lift of 0\mathcal{M}_{0} (Equation (12)) and uˇ\check{u} the lifting map (Equation (13)). Suppose that uu is an embedding.

Let γ>0\gamma>0 and consider the Čech bundle filtration (V[ˇ],𝕡)(V[\check{\mathcal{M}}],\mathbbm{p}) of ˇ\check{\mathcal{M}}. Its maximal filtration value is tγmax(ˇ)=22γt_{\gamma}^{\mathrm{max}}\left(\check{\mathcal{M}}\right)=\frac{\sqrt{2}}{2}\gamma. Denote by wi(𝕡)=(wit(𝕡))tTw_{i}(\mathbbm{p})=(w_{i}^{t}(\mathbbm{p}))_{t\in T} its persistent Stiefel-Whitney classes, i1,di\in\llbracket 1,d\rrbracket. Denote also by i0ti_{0}^{t} the inclusion ˇˇt\check{\mathcal{M}}\rightarrow\check{\mathcal{M}}^{t}, for t[0,reach(ˇ))t\in[0,\mathrm{reach}(\check{\mathcal{M}})).

Let t0t\geq 0 be such that t<min(reach(ˇ),tγmax(ˇ))t<\min\left(\mathrm{reach}(\check{\mathcal{M}}),t_{\gamma}^{\mathrm{max}}\left(\check{\mathcal{M}}\right)\right). Then the map i0tuˇ:0ˇti_{0}^{t}\circ\check{u}\colon\mathcal{M}_{0}\rightarrow\check{\mathcal{M}}^{t} induces an isomorphism H(0)H(ˇt)H^{*}(\mathcal{M}_{0})\leftarrow H^{*}(\check{\mathcal{M}}^{t}) which maps the ithi^{\text{th}} persistent Stiefel-Whitney class wit(𝕡)w_{i}^{t}(\mathbbm{p}) of (V[ˇ],𝕡)(V[\check{\mathcal{M}}],\mathbbm{p}) to the ithi^{\text{th}} Stiefel-Whitney class of (0,p)(\mathcal{M}_{0},p).

Proof.

Consider the following commutative diagram, defined for every t<tγmax(ˇ)t<t_{\gamma}^{\mathrm{max}}\left(\check{\mathcal{M}}\right):

0{\mathcal{M}_{0}}ˇ{\check{\mathcal{M}}}ˇt{\check{\mathcal{M}}^{t}}𝒢d(m){\mathcal{G}_{d}(\mathbb{R}^{m})}uˇ\scriptstyle{\check{u}}p\scriptstyle{p}i0t\scriptstyle{i_{0}^{t}}pt\scriptstyle{p^{t}}

We obtain a commutative diagram in cohomology:

H(0){H^{*}(\mathcal{M}_{0})}H(ˇ){H^{*}(\check{\mathcal{M}})}H(ˇt){H^{*}(\check{\mathcal{M}}^{t})}H(𝒢d(m)){H^{*}(\mathcal{G}_{d}(\mathbb{R}^{m}))}uˇ\scriptstyle{\check{u}^{*}}(i0t)\scriptstyle{(i_{0}^{t})^{*}}(pt)\scriptstyle{(p^{t})^{*}}p\scriptstyle{p^{*}}

Since t<reach(ˇ)t<\mathrm{reach}(\check{\mathcal{M}}), the map (i0t)(i_{0}^{t})^{*} is an isomorphism. So is uˇ\check{u}^{*} since uˇ\check{u} is an embedding. As a consequence, the map i0tuˇi_{0}^{t}\circ\check{u} induces an isomorphism H(0)H(ˇt)H^{*}(\mathcal{M}_{0})\simeq H^{*}(\check{\mathcal{M}}^{t}).

Let wiw_{i} denotes the ithi^{\text{th}} Stiefel-Whitney class of 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}). By definition, the ithi^{\text{th}} Stiefel-Whitney class of (0,p)(\mathcal{M}_{0},p) is p(wi)p^{*}(w_{i}), and the ithi^{\text{th}} persistent Stiefel-Whitney class of (V[ˇ],𝕡)(V[\check{\mathcal{M}}],\mathbbm{p}) is wit(𝕡)=(pt)(wi)w_{i}^{t}(\mathbbm{p})=(p^{t})^{*}(w_{i}). By commutativity of the diagram, we obtain p(wi)=(pt)(wi)p^{*}(w_{i})=(p^{t})^{*}(w_{i}) under the identification H(0)H(ˇt)H^{*}(\mathcal{M}_{0})\simeq H^{*}(\check{\mathcal{M}}^{t}). ∎

We deduce an estimation result.

Corollary 3.7.

Let XEX\subset E be any subset such that dH(X,ˇ)117reach(ˇ)\mathrm{d}_{\mathrm{H}}\left(X,\check{\mathcal{M}}\right)\leq\frac{1}{17}\mathrm{reach}(\check{\mathcal{M}}). Define ϵ=dH(X,ˇ)\epsilon=\mathrm{d}_{\mathrm{H}}\left(X,\check{\mathcal{M}}\right). Then for every t[4ϵ,reach(ˇ)3ϵ)t\in\left[4\epsilon,\mathrm{reach}(\check{\mathcal{M}})-3\epsilon\right), the composition of inclusions 0ˇXt\mathcal{M}_{0}\hookrightarrow\check{\mathcal{M}}\hookrightarrow X^{t} induces an isomorphism H(0)H(Xt)H^{*}(\mathcal{M}_{0})\leftarrow H^{*}(X^{t}) which sends the ithi^{\text{th}} persistent Stiefel-Whitney class wit(X)w_{i}^{t}(X) of the Čech bundle filtration of XX to the ithi^{\text{th}} Stiefel-Whitney class of (0,p)(\mathcal{M}_{0},p).

Proof.

This is a consequence of Theorems 3.3 and 3.6 previously stated, and Proposition 3.1 of Chazal et al. (2009). ∎

As a consequence of this corollary, on the set [4ϵ,reach(ˇ)3ϵ)[4\epsilon,\mathrm{reach}(\check{\mathcal{M}})-3\epsilon), the ithi^{\text{th}} persistent Stiefel-Whitney class of the Čech bundle filtration of XX is zero if and only if the ithi^{\text{th}} Stiefel-Whitney class of (0,p)(\mathcal{M}_{0},p) is.

Example 3.8.

In order to illustrate Corollary 3.7, consider the torus and the Klein bottle, immersed in 3\mathbb{R}^{3} as in Figure 9.

Refer to caption
Refer to caption
Figure 9: Immersion of the torus and the Klein bottle in 3\mathbb{R}^{3}.

Let them be endowed with their normal bundles. They can be seen as submanifolds ˇ,ˇ\check{\mathcal{M}},\check{\mathcal{M}}^{\prime} of 3×M(3)\mathbb{R}^{3}\times M(\mathbb{R}^{3}). We consider two samples X,XX,X^{\prime} of ˇ,ˇ\check{\mathcal{M}},\check{\mathcal{M}}^{\prime}, represented in Figure 10. They contain respectively 346 and 1489 points. We computed experimentally the Hausdorff distances dH(X,ˇ)0.6\mathrm{d}_{\mathrm{H}}\left(X,\check{\mathcal{M}}\right)\approx 0{.}6 and dH(X,ˇ)0.45\mathrm{d}_{\mathrm{H}}\left(X^{\prime},\check{\mathcal{M}}^{\prime}\right)\approx 0{.}45, with respect to the norm γ\left\|\cdot\right\|_{\gamma} where γ=1\gamma=1.

Refer to caption
Refer to caption
Figure 10: Samples XX and XX^{\prime} of ˇ\check{\mathcal{M}} and ˇ\check{\mathcal{M}}^{\prime}. The black points corresponds to the 3\mathbb{R}^{3}-coordinate, and the pink arrows over them correspond to the orientation of the M(3)M(\mathbb{R}^{3})-coordinate.

Figure 11 represents the barcodes of the persistent cohomology of XX and XX^{\prime}, and the lifebars of their first persistent Stiefel-Whitney classes w1(X)w_{1}(X) and w1(X)w_{1}(X^{\prime}). Observe that w1(X)w_{1}(X) is always zero, while w1(X)w_{1}(X^{\prime}) is nonzero for t0.3t\geq 0{.}3. This is an indication that ˇ\check{\mathcal{M}}, the underlying manifold of XX, is orientable, while ˇ\check{\mathcal{M}}^{\prime} is not. Lemma 3.9, stated below, justifies this assertion. Therefore, one interprets these lifebars as follows: XX is sampled on an orientable manifold, while XX^{\prime} is sampled on a non-orientable one.

Refer to captionRefer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to captionRefer to caption
Figure 11: Left: H0H^{0}, H1H^{1} and H2H^{2} barcodes of XX and lifebar of w1(X)w_{1}(X). Right: same for XX^{\prime}. Only bars of length larger than 0.20.2 are represented.
Lemma 3.9.

Let 0\mathcal{M}_{0}\rightarrow\mathcal{M} be an immersion of a manifold 0\mathcal{M}_{0} in a Euclidean space. Then 0\mathcal{M}_{0} is orientable if and only if the first Stiefel-Whitney class of its normal bundle is zero.

Proof.

Let τ\tau and ν\nu denote the tangent and normal bundles of 0\mathcal{M}_{0}. The Whitney sum τν\tau\oplus\nu is a trivial bundle, hence its first Stiefel-Whitney class is w1(τν)=0w_{1}(\tau\oplus\nu)=0. Using Axioms 1 and 3 of the Stiefel-Whitney classes, we obtain

w1(τν)\displaystyle w_{1}(\tau\oplus\nu) =w1(τ)w0(ν)+w0(τ)w1(ν)\displaystyle=w_{1}(\tau)\smile w_{0}(\nu)+w_{0}(\tau)\smile w_{1}(\nu)
=w1(τ)1+1w1(ν)\displaystyle=w_{1}(\tau)\smile 1+1\smile w_{1}(\nu)
=w1(τ)+w1(ν).\displaystyle=w_{1}(\tau)+w_{1}(\nu).

Therefore, w1(τ)=w1(ν)w_{1}(\tau)=w_{1}(\nu), hence w1(τ)w_{1}(\tau) is zero if and only if w1(ν)w_{1}(\nu) is zero. Besides, it is known that the first Stiefel-Whitney class of the tangent bundle of a manifold is zero if and only if the manifold is orientable. We deduce the result. ∎

4 Computation of persistent Stiefel-Whitney classes

In order to build an effective algorithm to compute the persistent Stiefel-Whitney classes, we have to find an equivalent formulation in terms of simplicial cohomology. We will make use of the well-known technique of simplicial approximation, as described in Subsect. 2.2.

4.1 Simplicial approximation to Čech bundle filtrations

Let XX be a subset of E=n×M(m)E=\mathbb{R}^{n}\times M(\mathbb{R}^{m}). Let us recall Definition 3.3: the Čech bundle filtration associated to XX is the vector bundle filtration (𝕏,𝕡)(\mathbb{X},\mathbbm{p}) whose underlying filtration is the Čech filtration 𝕏=(Xt)tT\mathbb{X}=(X^{t})_{t\in T}, with T=[0,tγmax(X))T=[0,t_{\gamma}^{\mathrm{max}}\left(X\right)), and whose maps 𝕡=(pt)tT\mathbbm{p}=(p^{t})_{t\in T} are given by the following composition, as in Equation (8):

Xt{X^{t}}M(m)med(𝒢d(m)){M(\mathbb{R}^{m})\setminus\mathrm{med}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right)}𝒢d(m).{\mathcal{G}_{d}(\mathbb{R}^{m}).}proj2\scriptstyle{\mathrm{proj}_{2}}pt\scriptstyle{p^{t}}proj(,𝒢d(m))\scriptstyle{\mathrm{proj}\left(\cdot,\mathcal{G}_{d}(\mathbb{R}^{m})\right)}

Let tTt\in T. The aim of this subsection is to describe a simplicial approximation to pt:Xt𝒢d(m)p^{t}\colon X^{t}\rightarrow\mathcal{G}_{d}(\mathbb{R}^{m}). To do so, let us fix a triangulation LL of 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}). It comes with a homeomorphism h:𝒢d(m)|L|h\colon\mathcal{G}_{d}(\mathbb{R}^{m})\rightarrow\left|L\right|. We shall now triangulate the thickenings XtX^{t} of the Čech set filtration. The thickening XtX^{t} is a subset of the metric space (E,γ)(E,\left\|\cdot\right\|_{\gamma}) which consists in a union of closed balls centered around points of XX:

Xt=xX¯γ(x,t),\displaystyle X^{t}=\bigcup_{x\in X}\overline{\mathcal{B}}_{\gamma}\left(x,t\right),

where ¯γ(x,t)\overline{\mathcal{B}}_{\gamma}\left(x,t\right) denotes the closed ball of center xx and radius tt for the norm γ\left\|\cdot\right\|_{\gamma}. Let 𝒰t\mathcal{U}^{t} denote the cover {¯γ(x,t),xX}\left\{\overline{\mathcal{B}}_{\gamma}\left(x,t\right),~{}x\in X\right\} of XtX^{t}, and let 𝒩(𝒰t)\mathcal{N}(\mathcal{U}^{t}) be its nerve. By the nerve theorem for convex closed covers (Boissonnat et al., 2018, Theorem 2.9), the simplicial complex 𝒩(𝒰t)\mathcal{N}(\mathcal{U}^{t}) is homotopy equivalent to its underlying set XtX^{t}. That is to say, there exists a continuous map gt:|𝒩(𝒰t)|Xtg^{t}\colon\left|\mathcal{N}(\mathcal{U}^{t})\right|\rightarrow X^{t} which is a homotopy equivalence.

As a consequence, in cohomological terms, the map pt:Xt𝒢d(E)p^{t}\colon X^{t}\rightarrow\mathcal{G}_{d}(E) is equivalent to the map qtq^{t} defined as qt=hptgtq^{t}=h\circ p^{t}\circ g^{t}.

Xt{X^{t}}𝒢d(m){\mathcal{G}_{d}(\mathbb{R}^{m})}|𝒩(𝒰t)|{\left|\mathcal{N}(\mathcal{U}^{t})\right|}|L|{\left|L\right|}pt\scriptstyle{p^{t}}h\scriptstyle{h}gt\scriptstyle{g^{t}}qt\scriptstyle{q^{t}} (14)

This gives a way to compute the induced map (pt):H(Xt)H(𝒢d(m))(p^{t})^{*}\colon H^{*}\left(X^{t}\right)\leftarrow H^{*}\left(\mathcal{G}_{d}(\mathbb{R}^{m})\right) algorithmically:

  • Subdivise 𝒩(𝒰t)\mathcal{N}(\mathcal{U}^{t}) until qtq^{t} satisfies the star condition.

  • Choose a simplicial approximation ftf^{t} to qtq^{t}.

  • Compute the induced map between simplicial cohomology groups (ft):H(𝒩(𝒰t))H(L)(f^{t})^{*}\colon H^{*}(\mathcal{N}(\mathcal{U}^{t}))\leftarrow H^{*}(L).

By correspondence between simplicial and singular cohomology, the map (ft)(f^{t})^{*} corresponds to (pt)(p^{t})^{*}. Hence the problem of computing (pt)(p^{t})^{*} is solved, if it were not for the following issue: in practice, the map gt:|𝒩(𝒰t)|Xtg^{t}\colon\left|\mathcal{N}(\mathcal{U}^{t})\right|\rightarrow X^{t} given by the nerve theorem is not explicit. The rest of this subsection is devoted to showing that gtg^{t} can be chosen canonically as the shadow map.

Shadow map.

We still consider XtX^{t}, the corresponding cover 𝒰t\mathcal{U}^{t} and its nerve 𝒩(𝒰t)\mathcal{N}(\mathcal{U}^{t}). The underlying vertex set of the simplicial complex 𝒩(𝒰t)\mathcal{N}(\mathcal{U}^{t}) is the set XX itself. The shadow map gt:|𝒩(𝒰t)|Xtg^{t}\colon\left|\mathcal{N}(\mathcal{U}^{t})\right|\rightarrow X^{t} is defined as follows: for every simplex σ=[x0,,xp]𝒩(𝒰t)\sigma=[x_{0},...,x_{p}]\in\mathcal{N}(\mathcal{U}^{t}) and every point i=0pλixi\sum_{i=0}^{p}\lambda_{i}x_{i} of |σ|\left|\sigma\right| written in barycentric coordinates, associate the point i=0pλixi\sum_{i=0}^{p}\lambda_{i}x_{i} of EE:

gt:i=0pλixi|σ|i=0pλixiE.\displaystyle g^{t}\colon\sum_{i=0}^{p}\lambda_{i}x_{i}\in\left|\sigma\right|\longmapsto\sum_{i=0}^{p}\lambda_{i}x_{i}\in E.

The following lemma states that this map is a homotopy equivalence. We are not aware whether the general position assumption can be removed.

Lemma 4.1.

Suppose that XX is finite and in general position. Then the shadow map gt:|𝒩(𝒰t)|Xtg^{t}\colon|\mathcal{N}(\mathcal{U}^{t})|\rightarrow X^{t} is a homotopy equivalence. Consequently, it induces an isomorphism (gt):H(|𝒩(𝒰t)|)H(Xt)(g^{t})^{*}\colon H^{*}(|\mathcal{N}(\mathcal{U}^{t})|)\leftarrow H^{*}(X^{t}).

Proof.

Recall that 𝒰t={¯γ(x,t),xX}\mathcal{U}^{t}=\left\{\overline{\mathcal{B}}_{\gamma}\left(x,t\right),x\in X\right\}. Let us consider a smaller cover. For every xXx\in X, let Vor(x)\mathrm{Vor}(x) denote the Voronoi cell of xx in the ambient metric space (E,γ)(E,\left\|\cdot\right\|_{\gamma}), and define

𝒱t={¯γ(x,t)Vor(x),xX}.\displaystyle\mathcal{V}^{t}=\left\{\overline{\mathcal{B}}_{\gamma}\left(x,t\right)\cap\mathrm{Vor}(x),x\in X\right\}.

The set 𝒱t\mathcal{V}^{t} is a cover of XtX^{t}, and its nerve 𝒩(𝒱t)\mathcal{N}(\mathcal{V}^{t}) is known as the Delaunay complex. Let ht:|𝒩(𝒱t)|Xth^{t}\colon\left|\mathcal{N}(\mathcal{V}^{t})\right|\rightarrow X^{t} denote the shadow map of 𝒩(𝒱t)\mathcal{N}(\mathcal{V}^{t}). The Delaunay complex is a subcomplex of the Čech complex, hence we can consider the following diagram:

|𝒩(𝒱t)|{|\mathcal{N}(\mathcal{V}^{t})|}|𝒩(𝒰t)|{|\mathcal{N}(\mathcal{U}^{t})|}Xt.{X^{t}.}ht\scriptstyle{h^{t}}gt\scriptstyle{g^{t}}

Now, Edelsbrunner (1993, Theorem 3.2) has proven that the shadow map ht:|𝒩(𝒱t)|Xth^{t}\colon|\mathcal{N}(\mathcal{V}^{t})|\rightarrow X^{t} is a homotopy equivalence (it is required here that XX is in general position). Moreover, we know from Bauer and Edelsbrunner (2017, Theorem 5.10) that 𝒩(𝒰t)\mathcal{N}(\mathcal{U}^{t}) collapses to 𝒩(𝒱t)\mathcal{N}(\mathcal{V}^{t}). Therefore the inclusion |𝒩(𝒱t)||𝒩(𝒰t)|\left|\mathcal{N}(\mathcal{V}^{t})\right|\hookrightarrow\left|\mathcal{N}(\mathcal{U}^{t})\right| also is a homotopy equivalence. By the 2-out-of-3 property of homotopy equivalences, we conclude that gtg^{t} is a homotopy equivalence. ∎

4.2 A sketch of algorithm

Suppose that we are given a finite set XE=n×M(m)X\subset E=\mathbb{R}^{n}\times M(\mathbb{R}^{m}). Choose d1,n1d\in\llbracket 1,n-1\rrbracket and γ>0\gamma>0. Consider the Čech bundle filtration of dimension dd of XX. Let T=[0,tγmax(X))T=\left[0,t_{\gamma}^{\mathrm{max}}\left(X\right)\right), tTt\in T and i1,di\in\llbracket 1,d\rrbracket. From the previous discussion we can infer an algorithm to solve the following problem:

Compute the persistent Stiefel-Whitney class wit(X)w_{i}^{t}(X) of the Čech bundle filtration of XX, using a cohomology computation software.

Denote:

  • 𝕏=(Xt)t0\mathbb{X}=(X^{t})_{t\geq 0} the Čech set filtration of XX,

  • 𝕊\operatorname{\mathbb{S}} the Čech simplicial filtration of XX, and gt:|St|Xtg^{t}\colon\left|S^{t}\right|\rightarrow X^{t} the shadow map,

  • LL a triangulation of 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) and h:𝒢d(m)|L|h\colon\mathcal{G}_{d}(\mathbb{R}^{m})\rightarrow\left|L\right| a homeomorphism,

  • (𝕏,𝕡)(\mathbb{X},\mathbbm{p}) the Čech bundle filtration of XX,

  • (𝕍,𝕧)(\mathbb{V},\mathbbm{v}) the persistent cohomology module of 𝕏\mathbb{X},

  • wiHi(𝒢d(m))w_{i}\in H^{i}(\mathcal{G}_{d}(\mathbb{R}^{m})) the ithi^{\text{th}} Stiefel-Whitney class of the Grassmannian.

Let tTt\in T and consider the map qtq^{t}, as defined in Equation (14):

|St|{\left|S^{t}\right|}Xt{X^{t}}𝒢d(m){\mathcal{G}_{d}(\mathbb{R}^{m})}|L|.{\left|L\right|.}gt\scriptstyle{g^{t}}qt\scriptstyle{q^{t}}pt\scriptstyle{p^{t}}h\scriptstyle{h}

We propose the following algorithm:

  • Subdivise barycentrically StS^{t} until qtq^{t} satisfies the star condition. Denote kk the number of subdivisions needed.

  • Consider a simplicial approximation ft:subk(St)Lf^{t}\colon\mathrm{sub}^{k}\left(S^{t}\right)\rightarrow L to qtq^{t}.

  • Compute the class (ft)(wi)(f^{t})^{*}(w_{i}).

The output (ft)(wi)(f^{t})^{*}(w_{i}) is equal to the persistent Stiefel-Whitney class wit(X)w_{i}^{t}(X) at time tt, seen in the simplicial cohomology group Hi(St)=Hi(subk(St))H^{i}(S^{t})=H^{i}(\mathrm{sub}^{k}\left(S^{t}\right)). In the following section, we gather some technical details needed to implement this algorithm in practice.

Note that this also gives a way to compute the lifebar of wi(X)w_{i}(X). This bar is determined by the value t=inf{tT,wi(X)0}t^{\dagger}=\inf\{t\in T,w_{i}(X)\neq 0\}. This quantity can be approximated by dichotomic search, by computing the classes wit(X)w_{i}^{t}(X) for several values of tt. We point out that, in order to compute the value tt^{\dagger}, there may exist a better algorithm than evaluating the class wit(X)w_{i}^{t}(X) several times.

Let us describe briefly such an algorithm when i=1i=1, that is, when the first persistent Stiefel-Whitney class w1(X)w_{1}(X) is to be computed. First, we remind the reader that the first cohomology group H1(𝒢d(m))H^{1}(\mathcal{G}_{d}(\mathbb{R}^{m})) of the Grassmannian is generated by one element, the first Stiefel-Whitney class w1w_{1}. For any tTt\in T, consider the map ptp^{t} as above, and the map induced in cohomology, (pt):H1(Xt)H1(𝒢d(m))(p^{t})^{*}\colon H^{1}(X^{t})\leftarrow H^{1}(\mathcal{G}_{d}(\mathbb{R}^{m})). Since w1t(X)=(pt)(w1)w_{1}^{t}(X)=(p^{t})^{*}(w_{1}), and since H1(𝒢d(m))H^{1}(\mathcal{G}_{d}(\mathbb{R}^{m})) has dimension 1, we have that w1t(X)w_{1}^{t}(X) is nonzero if and only if rank(pt)\mathrm{rank}(p^{t})^{*} is nonzero.

Next, let C(pt)C(p^{t}) denotes the mapping cone of pt:Xt𝒢d(m)p^{t}\colon X^{t}\rightarrow\mathcal{G}_{d}(\mathbb{R}^{m}). This is a usual construction in algebraic topology. In a few words, the mapping cone of a map is a topological space that contains information about the map. The mapping cone C(pt)C(p^{t}) comes with a long exact sequence

Hk(Xt)Hk+1(C(pt))Hk+1(𝒢d(m))Hk+1(Xt)...\longrightarrow H^{k}(X^{t})\longrightarrow H^{k+1}(C(p^{t}))\longrightarrow H^{k+1}(\mathcal{G}_{d}(\mathbb{R}^{m}))\longrightarrow H^{k+1}(X^{t})\longrightarrow...

from which we deduce the formula

rank(pt)=k=1+(1)k(dimHk(Xt)dimHk+1(C(pt))+dimHk+1(𝒢d(m)))\mathrm{rank}(p^{t})^{*}=\sum_{k=1}^{+\infty}(-1)^{k}\bigg{(}\dim H^{k}(X^{t})-\dim H^{k+1}(C(p^{t}))+\dim H^{k+1}(\mathcal{G}_{d}(\mathbb{R}^{m}))\bigg{)}

On the simplicial side, is not complicated to build a triangulation CtC^{t} of the mapping cone C(pt)C(p^{t}), nor to build a simplicial filtration (Ct)tT(C^{t})_{t\in T} of (C(pt))tT\left(C(p^{t})\right)_{t\in T}. It relies on finding simplicial approximations to the maps qt:StLq^{t}\colon S^{t}\rightarrow L, as in the proof of Hatcher (2002, Theorem 2C.5). We point out that, just as in the previous algorithm, we may have to apply barycentric subdivisions to StS^{t} here. Now, the previous formula translates in simplicial cohomology as

rank(qt)=k=1+(1)k(dimHk(St)dimHk+1(Ct)+dimHk+1(L)).\mathrm{rank}(q^{t})^{*}=\sum_{k=1}^{+\infty}(-1)^{k}\bigg{(}\dim H^{k}(S^{t})-\dim H^{k+1}(C^{t})+\dim H^{k+1}(L)\bigg{)}.

All these terms can be computed efficiently by applying the persistent homology algorithm to the filtrations (St)tT(S^{t})_{t\in T}, (C(pt))tT\left(C(p^{t})\right)_{t\in T} and (L)tT(L)_{t\in T}. Finally, we identify the value tt^{\dagger} as the first value of tTt\in T such that rank(qt)\mathrm{rank}(q^{t})^{*} is nonzero.

5 An algorithm when d=1d=1

Even though the last sections described a theoretical way to compute the persistent Stiefel-Whitney classes, some concrete issues are still to be discussed:

  • verifying that the star condition is satisfied,

  • the Grassmann manifold has to be triangulated,

  • in practice, the Vietoris-Rips filtration is preferred to the Čech filtration,

  • the parameter γ\gamma has to be tuned.

The following subsections will elucidate these points. Concerning the first one, we are not aware of a computational-explicit process to triangulate the Grassmann manifolds 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}), except when d=1d=1, which corresponds to the projective spaces 𝒢1(m)\mathcal{G}_{1}(\mathbb{R}^{m}). We shall then restrict to the case d=1d=1. Note that, in this case, the only nonzero Stiefel-Whitney classes are the first two (by Axiom 1 of Stiefel-Whitney classes). Since w0w_{0} is always equal to 1, the only class to estimate is w1w_{1}.

5.1 The star condition in practice

Let us get back to the context of Subsect. 2.2: K,LK,L are two simplicial complexes, KK is finite, and g:|K||L|g\colon\left|K\right|\rightarrow\left|L\right| is a continuous map. We have seen that finding a simplicial approximation to gg reduces to finding a small enough barycentric subdivision subn(K)\mathrm{sub}^{n}\left(K\right) of KK such that g:|subn(K)||L|g\colon\left|\mathrm{sub}^{n}\left(K\right)\right|\rightarrow\left|L\right| satisfies the star condition, that is, for every vertex vv of subn(K)\mathrm{sub}^{n}\left(K\right), there exists a vertex ww of LL such that

g(|St¯(v)|)|St(w)|.g\left(\left|\overline{\mathrm{St}}\left(v\right)\right|\right)\subseteq\left|\mathrm{St}\left(w\right)\right|.

In practice, one can compute the closed star St¯(v)\overline{\mathrm{St}}\left(v\right) from the finite simplicial complex subn(K)\mathrm{sub}^{n}\left(K\right). However, computing g(|St¯(v)|)g\left(\left|\overline{\mathrm{St}}\left(v\right)\right|\right) requires to evaluate gg on the infinite set |St¯(v)|\left|\overline{\mathrm{St}}\left(v\right)\right|. In order to reduce the problem to a finite number of evaluations of gg, we shall consider a related property that we call the weak star condition.

Definition 5.1.

A map g:|K||L|g\colon\left|K\right|\rightarrow\left|L\right| between geometric realizations of simplicial complexes KK and LL satisfies the weak star condition if for every vertex vv of subn(K)\mathrm{sub}^{n}\left(K\right), there exists a vertex ww of LL such that

|g(St¯(v)0)||St(w)|,\left|g\left({\overline{\mathrm{St}}\left(v\right)}^{0}\right)\right|\subseteq\left|\mathrm{St}\left(w\right)\right|,

where St¯(v)0{\overline{\mathrm{St}}\left(v\right)}^{0} denotes the 0-skeleton of St¯(v)\overline{\mathrm{St}}\left(v\right), i.e. its vertices.

Observe that the practical verification of the condition |g(St¯(v)0)||St(w)|\left|g\left({\overline{\mathrm{St}}\left(v\right)}^{0}\right)\right|\subseteq\left|\mathrm{St}\left(w\right)\right| requires only a finite number of computations. Indeed, one just has to check whether every neighbor vv^{\prime} of vv in the graph K1{K}^{1}, vv included, satisfies g(v)|St(w)|g(v^{\prime})\in\left|\mathrm{St}\left(w\right)\right|. The following lemma rephrases this condition by using the face map L:|L|L\mathcal{F}_{L}\colon\left|L\right|\rightarrow L. We remind the reader that the face map is defined by the relation xL(x)x\in\mathcal{F}_{L}(x) for all x|L|x\in\left|L\right|.

Lemma 5.1.

The map gg satisfies the weak star condition if and only if for every vertex vv of KK, there exists a vertex ww of LL such that for every neighbor vv^{\prime} of vv in K1{K}^{1}, we have

wL(g(v)).w\in\mathcal{F}_{L}(g(v^{\prime})).
Proof.

Let us show that the assertion “wL(g(v))w\in\mathcal{F}_{L}(g(v^{\prime}))” is equivalent to “g(v)|St(w)|g(v^{\prime})\in\left|\mathrm{St}\left(w\right)\right|”. Recall that the open star St(w)\mathrm{St}\left(w\right) consists of the simplices of LL that contain ww. Moreover, the geometric realization |St(w)|\left|\mathrm{St}\left(w\right)\right| is the union of the |σ|\left|\sigma\right| for σSt(w)\sigma\in\mathrm{St}\left(w\right). As a consequence, g(v)g(v^{\prime}) belongs to |St(w)|\left|\mathrm{St}\left(w\right)\right| if and only if it belongs to |σ|\left|\sigma\right| for some simplex σL\sigma\in L that contains ww. Equivalently, the face map L(g(v))\mathcal{F}_{L}(g(v^{\prime})) contains ww. ∎

Suppose that gg satisfies the weak star condition. Let f:K0L0f\colon{K}^{0}\rightarrow{L}^{0} be a map, between vertex sets, such that for every vK0v\in{K}^{0},

|g(St¯(v)0)||St(f(v))|.\left|g\left({\overline{\mathrm{St}}\left(v\right)}^{0}\right)\right|\subseteq\left|\mathrm{St}\left(f(v)\right)\right|.

According to the proof of Lemma 5.1, an equivalent formulation of this condition is: for all neighbor vv^{\prime} of vv in K1{K}^{1},

f(v)L(g(v)).f(v)\in\mathcal{F}_{L}(g(v^{\prime})). (15)

Such a map is called a weak simplicial approximation to gg. It plays a similar role as the simplicial approximations to gg.

Lemma 5.2.

If f:K0L0f\colon{K}^{0}\rightarrow{L}^{0} is a weak simplicial approximation to g:|K||L|g\colon\left|K\right|\rightarrow\left|L\right|, then ff is a simplicial map.

Proof.

Let σ=[v0,,vn]\sigma=[v_{0},...,v_{n}] be a simplex of KK. We have to show that f(σ)=[f(v0),,f(vn)]f(\sigma)=[f(v_{0}),...,f(v_{n})] is a simplex of LL. Note that each closed star St¯(vi)\overline{\mathrm{St}}\left(v_{i}\right) contains σ\sigma. Therefore each |g(St¯(vi)0)|\left|g\left({\overline{\mathrm{St}}\left(v_{i}\right)}^{0}\right)\right| contains |g(σ0)|={g(v0),,g(vn)}\left|g\left({\sigma}^{0}\right)\right|=\{g(v_{0}),...,g(v_{n})\}. Using the weak simplicial approximation property of ff, we deduce that each |St(f(vi))|\left|\mathrm{St}\left(f(v_{i})\right)\right| contains {g(v0),,g(vn)}\{g(v_{0}),...,g(v_{n})\}. Using Lemma 5.3 stated below, we obtain that [f(v0),,f(vn)][f(v_{0}),...,f(v_{n})] is a simplex of LL. ∎

Lemma 5.3 (Hatcher, 2002, Lemma 2C.2).

Let w0,,wnw_{0},...,w_{n} be vertices of a simplicial complex LL. Then i=0nSt(wi)\bigcap_{i=0}^{n}\mathrm{St}\left(w_{i}\right)\neq\emptyset if and only if [w0,,wn][w_{0},...,w_{n}] is a simplex of LL.

As one can see from the definitions, the weak star condition is weaker than the star condition. Consequently, the simplicial approximation theorem admits the following corollary.

Corollary 5.4.

Consider two simplicial complexes K,LK,L with KK finite, and let g:|K||L|g\colon\left|K\right|\rightarrow\left|L\right| be a continuous map. Then there exists n0n\geq 0 such that g:|subn(K)||L|g\colon\left|\mathrm{sub}^{n}\left(K\right)\right|\rightarrow\left|L\right| satisfies the weak star condition.

However, some weak simplicial approximations to gg may not be simplicial approximations, and may not even be homotopic to gg. Figure 12 gives such an example.

Refer to caption

KK

Refer to caption

LL

Refer to caption

g:|K||L|g\colon\left|K\right|\rightarrow\left|L\right|

Figure 12: The map gg is non-trivial but admits a weak simplicial approximation which is constant.

Fortunately, these two notions coincides under the star condition assumption:

Proposition 5.5.

Suppose that gg satisfies the star condition. Then every weak simplicial approximation to gg is a simplicial approximation.

Proof.

Let ff be a weak simplicial approximation to gg, and ff^{\prime} any simplicial approximation. Let us show that ff and ff^{\prime} are contiguous simplicial maps. Let σ=[v0,,vn]\sigma=[v_{0},...,v_{n}] be a simplex of KK. We have to show that [f(v0),,f(vn),f(v0),,f(vn)][f(v_{0}),...,f(v_{n}),f^{\prime}(v_{0}),...,f^{\prime}(v_{n})] is a simplex of LL. As we have seen in the proof of Lemma 5.2, each |g(St¯(vi)0)|\left|g\left({\overline{\mathrm{St}}\left(v_{i}\right)}^{0}\right)\right| contains {g(v0),,g(vn)}\{g(v_{0}),...,g(v_{n})\}. Therefore, by definition of weak simplicial approximations and simplicial approximations, each |St(f(vi))|\left|\mathrm{St}\left(f(v_{i})\right)\right| and |St(f(vi))|\left|\mathrm{St}\left(f^{\prime}(v_{i})\right)\right| contains {g(v0),,g(vn)}\{g(v_{0}),...,g(v_{n})\}. We conclude by applying Lemma 5.3. ∎

Remark that the proof of this proposition can be adapted to obtain the following fact: any two weak simplicial approximations are equivalent—as well as any two simplicial approximations.

Let us comment Proposition 5.5. If KK is subdivised enough, then every weak simplicial approximation to gg is homotopic to gg. We face the following problem in practice: the number of subdivisions needed by the star condition is not known. In order to work around this problem, we propose to subdivise the complex KK until it satisfies the weak star condition, and then use a weak simplicial approximation to gg. However, such a weak simplicial approximation may not be homotopic to gg, and our algorithm would output a wrong result.

To close this subsection, we state a lemma that gives a quantitative idea of the number of subdivisions needed by the star condition. We say that a Lebesgue number for an open cover 𝒰\mathcal{U} of a compact metric space XX is a positive number ϵ\epsilon such that every subset of XX with diameter less than ϵ\epsilon is included in some element of the cover 𝒰\mathcal{U}.

Lemma 5.6.

Let |K|,|L|\left|K\right|,\left|L\right| be endowed with metrics. Suppose that g:|K||L|g\colon\left|K\right|\rightarrow\left|L\right| is ll-Lipschitz with respect to these metrics. Let ϵ\epsilon be a Lebesgue number for the open cover {|St(w)|,wL}\left\{\left|\mathrm{St}\left(w\right)\right|,w\in L\right\} of |L|\left|L\right|. Let pp be the dimension of KK and DD an upper bound on the diameter of its faces. Then for n>log(Dlϵ)/log(p+1p)n>\log(\frac{Dl}{\epsilon})\big{/}\log(\frac{p+1}{p}), the map g:|subn(K)||L|g\colon\left|\mathrm{sub}^{n}\left(K\right)\right|\rightarrow\left|L\right| satisfies the star condition.

Proof.

The map gg satisfies the star condition if for every vertex vv of KK, there exists a vertex ww of LL such that g(|St¯(v)|)|St(w)|g(\left|\overline{\mathrm{St}}\left(v\right)\right|)\subseteq\left|\mathrm{St}\left(w\right)\right|. Since the cover {|St(w)|,wL}\left\{\left|\mathrm{St}\left(w\right)\right|,w\in L\right\} admits ϵ\epsilon as a Lebesgue number, it is enough for vv to satisfy the following inequality:

diam(g(|St¯(v)|))<ϵ.\mathrm{diam}\left(g(\left|\overline{\mathrm{St}}\left(v\right)\right|)\right)<\epsilon. (16)

Since gg is ll-Lipschitz, we have diam(g(|St¯(v)|))ldiam(|St¯(v)|)\mathrm{diam}\left(g\left(\left|\overline{\mathrm{St}}\left(v\right)\right|\right)\right)\leq l\cdot\mathrm{diam}\left(\left|\overline{\mathrm{St}}\left(v\right)\right|\right). Using the hypothesis diam(|St¯(v)|)D\mathrm{diam}\left(\left|\overline{\mathrm{St}}\left(v\right)\right|\right)\leq D, Equation (16) leads to the condition Dl<ϵDl<\epsilon. Now, we use the fact that a barycentric subdivision reduces the diameter of each face by a factor pp+1\frac{p}{p+1}. After nn barycentric subdivision, the last inequality rewrites (pp+1)nDl<ϵ\left(\frac{p}{p+1}\right)^{n}Dl<\epsilon. It admits n>log(Dlϵ)/log(p+1p)n>\log(\frac{Dl}{\epsilon})\big{/}\log(\frac{p+1}{p}) as a solution. ∎

5.2 Triangulating the projective spaces

As we described in Subsect. 5.1, the algorithm we propose rests on a triangulation LL of the Grassmannian 𝒢1(m)\mathcal{G}_{1}(\mathbb{R}^{m}), together with the map Lh:𝒢1(m)L\mathcal{F}_{L}\circ h\colon\mathcal{G}_{1}(\mathbb{R}^{m})\rightarrow L, where h:𝒢1(m)|L|h\colon\mathcal{G}_{1}(\mathbb{R}^{m})\rightarrow\left|L\right| is a homeomorphism and L:𝒢1(m)L\mathcal{F}_{L}\colon\mathcal{G}_{1}(\mathbb{R}^{m})\rightarrow L is the face map. In the following, we also call :=Lh\mathcal{F}:=\mathcal{F}_{L}\circ h the face map.

We shall use the triangulation of the projective space 𝒢1(m)\mathcal{G}_{1}(\mathbb{R}^{m}) by von Kühnel (1987). It uses the fact that the quotient of the sphere 𝕊m1\operatorname{\mathbb{S}}_{m-1} by the antipodal relation gives 𝒢1(m)\mathcal{G}_{1}(\mathbb{R}^{m}). Let Δm\Delta^{m} denote the standard mm-simplex, v0,,vmv_{0},...,v_{m} its vertices, and Δm\partial\Delta^{m} its boundary. The simplicial complex Δm\partial\Delta^{m} is a triangulation of the sphere 𝕊m1\mathbb{S}_{m-1}. Denote its first barycentric subdivision as sub1(Δm)\mathrm{sub}^{1}\left(\partial\Delta^{m}\right). The vertices of sub1(Δm)\mathrm{sub}^{1}\left(\partial\Delta^{m}\right) are in bijection with the non-empty proper subsets of {v0,,vm}\{v_{0},...,v_{m}\}. Consider the equivalence relation on these vertices which associates a vertex to its complement. The quotient simplicial complex under this relation, LL, is a triangulation of 𝒢1(m)\mathcal{G}_{1}(\mathbb{R}^{m}). Figures 13 and 14 represent this construction.

Refer to caption

Δ2\partial\Delta^{2}

Refer to caption

sub1(Δ2)\mathrm{sub}^{1}\left(\partial\Delta^{2}\right)

Refer to caption

Equivalence relation

Refer to caption

Quotient complex LL

Figure 13: Triangulating 𝒢1(2)\mathcal{G}_{1}(\mathbb{R}^{2}).

Let us now describe how to define the homeomorphism h:𝒢1(m)|L|h\colon\mathcal{G}_{1}(\mathbb{R}^{m})\rightarrow\left|L\right|. First, embed Δm\Delta^{m} in m+1\mathbb{R}^{m+1} via vi(0,,0,1,0,)v_{i}\mapsto(0,...,0,1,0,...), where 11 sits at the ithi^{\text{th}} coordinate. Its image lies on a mm-dimensional affine subspace PP, with origin being the barycenter of v0,,vmv_{0},...,v_{m}. Seen in PP, the vertices of Δm\Delta^{m} now belong to the sphere centered at the origin and of radius mm+1\sqrt{\frac{m}{m+1}} (see Figure 14). Let us denote this sphere as 𝕊m1\mathbb{S}_{m-1}. Next, subdivise barycentrically Δm\partial\Delta^{m} once, and project each vertex of sub1(Δm)\mathrm{sub}^{1}\left(\partial\Delta^{m}\right) on 𝕊m1\mathbb{S}_{m-1}. By taking the convex hulls of its faces, we now see |sub1(Δm)|\left|\mathrm{sub}^{1}\left(\partial\Delta^{m}\right)\right| as a subset of PP. Define an application p:𝕊m1|sub1(Δm)|p\colon\mathbb{S}_{m-1}\rightarrow\left|\mathrm{sub}^{1}\left(\partial\Delta^{m}\right)\right| as follows: for every x𝕊m1x\in\mathbb{S}_{m-1}, the image p(x)p(x) is the unique intersection point between the segment [0,x][0,x] and the set |sub1(Δm)|\left|\mathrm{sub}^{1}\left(\partial\Delta^{m}\right)\right|. The application pp can also be seen as the inverse function of the projection on 𝕊m1\operatorname{\mathbb{S}}_{m-1}, written proj𝕊m1:|sub1(Δm)|𝕊m1\text{proj}_{\operatorname{\mathbb{S}}_{m-1}}\colon\left|\mathrm{sub}^{1}\left(\partial\Delta^{m}\right)\right|\rightarrow\mathbb{S}_{m-1}. As a consequence, we can factorize p:𝕊m1|sub1(Δm)|p\colon\mathbb{S}_{m-1}\rightarrow\left|\mathrm{sub}^{1}\left(\partial\Delta^{m}\right)\right| as

h:(𝕊m1/)(|sub1(Δm)|/).\displaystyle h\colon\left(\mathbb{S}_{m-1}/{\sim}\right)\rightarrow\left(\left|\mathrm{sub}^{1}\left(\partial\Delta^{m}\right)\right|/{\sim}\right).

Using Lemma 5.7 stated below, we can identify these spaces with

h:𝒢1(m)|L|,\displaystyle h\colon\mathcal{G}_{1}(\mathbb{R}^{m})\rightarrow\left|L\right|,

giving the desired triangulation.

Refer to caption

Δ3\partial\Delta^{3} is included in 𝕊m1\mathbb{S}_{m-1}

Refer to caption

sub1(Δ3)\mathrm{sub}^{1}\left(\partial\Delta^{3}\right) and 𝕊m1\mathbb{S}_{m-1}

Refer to caption

LL

Figure 14: Triangulating 𝒢1(3)\mathcal{G}_{1}(\mathbb{R}^{3}).
Lemma 5.7.

For any vertex xsub1(Δm)x\in\mathrm{sub}^{1}\left(\partial\Delta^{m}\right), denote by |x|\left|x\right| its embedding in PP. Let |x|-\left|x\right| denote the image of |x|\left|x\right| by the antipodal relation on 𝕊m1\operatorname{\mathbb{S}}_{m-1}. Denote by yy the image of xx by the relation on sub1(Δm)\mathrm{sub}^{1}\left(\partial\Delta^{m}\right). Then y=|x|y=-\left|x\right|.

More generally, pulling back the antipodal relation onto |sub1(Δm)|\left|\mathrm{sub}^{1}\left(\partial\Delta^{m}\right)\right| via pp gives the relation we defined on sub1(Δm)\mathrm{sub}^{1}\left(\partial\Delta^{m}\right).

Proof.

Pick a vertex xx of sub1(Δm)\mathrm{sub}^{1}\left(\partial\Delta^{m}\right). It can be described as a proper subset {vi,iI}\{v_{i},i\in I\} of the vertex set (Δm)0={v0,,vm}{(\partial\Delta^{m})}^{0}=\{v_{0},...,v_{m}\}, where I0,mI\subset\llbracket 0,m\rrbracket. According to the relation on (Δm)(\partial\Delta^{m}), the vertex xx is in relation with the vertex yy described by the proper subset {vi,iIc}\{v_{i},i\in{I}^{c}\}. The point xx can be written in barycentric coordinates as 1card(I)iI|vi|\frac{1}{\mathrm{card}(I)}\sum_{i\in I}\left|v_{i}\right|. Seen in PP, |x|\left|x\right| can be written |x|=proj𝕊m1(iIvi)\left|x\right|=\mathrm{proj}_{\operatorname{\mathbb{S}}_{m-1}}\left(\sum_{i\in I}v_{i}\right). Similarly, |y|\left|y\right| can be written |y|=proj𝕊m1(iIcvi)\left|y\right|=\mathrm{proj}_{\operatorname{\mathbb{S}}_{m-1}}\left(\sum_{i\in{I}^{c}}v_{i}\right).

Now, denote by 0 the origin of the hyperplane PP, and embed the vertices v0,,vmv_{0},...,v_{m} in PP. Observe that

0=i0vi=iIvi+iIcvi.\displaystyle 0=\sum_{i\leq 0}v_{i}=\sum_{i\in I}v_{i}+\sum_{i\in{I}^{c}}v_{i}.

Hence iIvi=iIcvi-\sum_{i\in I}v_{i}=\sum_{i\in{I}^{c}}v_{i}, and we deduce that

|x|=proj𝕊m1(iIvi)=proj𝕊m1(iIcvi)=|y|.\displaystyle-\left|x\right|=\mathrm{proj}_{\operatorname{\mathbb{S}}_{m-1}}\left(-\sum_{i\in I}v_{i}\right)=\mathrm{proj}_{\operatorname{\mathbb{S}}_{m-1}}\left(\sum_{i\in{I}^{c}}v_{i}\right)=\left|y\right|.

Applying the same reasoning, one obtains the following result: for every simplex σ\sigma of sub1(Δm)\mathrm{sub}^{1}\left(\partial\Delta^{m}\right), if ν\nu denotes the image of σ\sigma by the relation on sub1(Δm)\mathrm{sub}^{1}\left(\partial\Delta^{m}\right), then the image of |σ|\left|\sigma\right| by the antipodal relation is also |ν|\left|\nu\right|. As a consequence, these two relations coincide. ∎

At a computational level, let us describe how to compute the face map :𝒢1(m)L\mathcal{F}\colon\mathcal{G}_{1}(\mathbb{R}^{m})\rightarrow L. Since \mathcal{F} can be obtained as a quotient, it is enough to compute the face map of the sphere, :𝕊m1sub1(Δm)\mathcal{F}^{\prime}\colon\mathbb{S}_{m-1}\rightarrow\mathrm{sub}^{1}\left(\partial\Delta^{m}\right), which corresponds to the homeomorphism p:𝕊m1|sub1(Δm)|p\colon\operatorname{\mathbb{S}}_{m-1}\rightarrow\left|\mathrm{sub}^{1}\left(\partial\Delta^{m}\right)\right|. It is given by the following lemma, which can be used in practice.

Lemma 5.8.

For every x𝕊m1x\in\operatorname{\mathbb{S}}_{m-1}, the image of xx by \mathcal{F}^{\prime} is equal to the intersection of all maximal faces σ=[w0,,wm]\sigma=[w_{0},...,w_{m}] of sub1(Δm)\mathrm{sub}^{1}\left(\partial\Delta^{m}\right) that satisfies the following conditions: denoting by x0x_{0} any point of the affine hyperplane spanned by {w0,,wm}\{w_{0},...,w_{m}\}, and by hh a vector orthogonal to the corresponding linear hyperplane,

  • the inner product x,h\left\langle x,h\right\rangle has the same sign as x0,h\left\langle x_{0},h\right\rangle,

  • the point x0,hx,hx\frac{\left\langle x_{0},h\right\rangle}{\left\langle x,h\right\rangle}x, which is included in the affine hyperplane spanned by {w0,,wm}\{w_{0},...,w_{m}\}, has nonnegative barycentric coordinates.

Proof.

Recall that for every x𝕊m1x\in\mathbb{S}_{m-1}, the image p(x)p(x) is defined as the unique intersection point between the segment [0,x][0,x] and the set |sub1(Δm)|\left|\mathrm{sub}^{1}\left(\partial\Delta^{m}\right)\right|. Besides, the face map (x)\mathcal{F}^{\prime}(x) is the unique simplex σsub1(Δm)\sigma\in\mathrm{sub}^{1}\left(\partial\Delta^{m}\right) such that p(x)|σ|p(x)\in\left|\sigma\right|. Equivalently, (x)\mathcal{F}^{\prime}(x) is equal to the intersection of all maximal faces σsub1(Δm)\sigma\in\mathrm{sub}^{1}\left(\partial\Delta^{m}\right) such that p(x)p(x) belongs to the closure |σ|¯\overline{\left|\sigma\right|}.

Consider any maximal face σ=[w0,,wm]\sigma=[w_{0},...,w_{m}] of sub1(Δm)\mathrm{sub}^{1}\left(\partial\Delta^{m}\right). The first condition of the lemma ensures that the segment [0,x][0,x] intersects the affine hyperplane spanned by {w0,,wm}\{w_{0},...,w_{m}\}. In this case, a computation shows that this intersection consists of the point x0,hx,hx\frac{\left\langle x_{0},h\right\rangle}{\left\langle x,h\right\rangle}x. Then, the second condition of the lemma tests whether this point belongs to the convex hull of {w0,,wk}\{w_{0},...,w_{k}\}. In conclusion, if σ\sigma satisfies these two conditions, then p(x)|σ|¯p(x)\in\overline{\left|\sigma\right|}. ∎

As a remark, let us point out that the verification of the conditions of this lemma is subject to numerical errors. In particular, the point x0,hx,hx\frac{\left\langle x_{0},h\right\rangle}{\left\langle x,h\right\rangle}x may have nonnegative coordinates, yet mathematical softwares may return (small) negative values. Consequently, the algorithm may recognize less maximal faces that satisfy these conditions, hence return a simplex that strictly contains the wanted simplex (x)\mathcal{F}^{\prime}(x). Nonetheless, such an error will not affect the output of the algorithm. Indeed, if we denote by ~\widetilde{\mathcal{F}^{\prime}} the face map computed by the algorithm, we have that (x)~(x)\mathcal{F}^{\prime}(x)\subseteq\widetilde{\mathcal{F}^{\prime}}(x) for all x𝕊m1x\in\operatorname{\mathbb{S}}_{m-1}. As a consequence of Lemma 5.1, ~\widetilde{\mathcal{F}^{\prime}} satisfies the weak star condition if \mathcal{F}^{\prime} does, and Equation (15) shows that every weak simplicial approximations for \mathcal{F}^{\prime} are weak simplicial approximations for ~\widetilde{\mathcal{F}^{\prime}}. Since every weak simplicial approximations are homotopic, we obtain that the induced maps in cohomology are equal, therefore the output of the algorithm is unchanged.

5.3 Vietoris-Rips version of the Čech bundle filtration

We still consider a subset Xn×M(m)X\subset\mathbb{R}^{n}\times M(\mathbb{R}^{m}). Denote by 𝕏\mathbb{X} the corresponding Čech set filtration, and by 𝕊=(St)t0\operatorname{\mathbb{S}}=(S^{t})_{t\geq 0} the simplicial Čech filtration. For every t0t\geq 0, let RtR^{t} be the flag complex of StS^{t}, i.e. the clique complex of the 1-skeleton (St)1{(S^{t})}^{1} of StS^{t}. It is known as the Vietoris-Rips complex of XX at time tt. The collection =(Rt)t0\mathbb{R}=(R^{t})_{t\geq 0} is called the Vietoris-Rips filtration of XX. The simplicial filtrations 𝕊\operatorname{\mathbb{S}} and \mathbb{R} are multiplicatively 2\sqrt{2}-interleaved (Bell et al., 2019, Theorem 3.1). In other words, for every t0t\geq 0, we have

StRtS2t.\displaystyle S^{t}\subseteq R^{t}\subseteq S^{\sqrt{2}t}.

Let γ>0\gamma>0 and consider the Čech bundle filtration (𝕏,𝕡)(\mathbb{X},\mathbbm{p}) of XX. Suppose that its maximal filtration value tγmax(X)t_{\gamma}^{\mathrm{max}}\left(X\right) is positive. Let ||=(|Rt|)t0\left|\mathbb{R}\right|=(\left|R^{t}\right|)_{t\geq 0} denote the geometric realization of the Vietoris-Rips filtration. We can give ||\left|\mathbb{R}\right| a vector bundle filtration structure with (p)t:|Rt|𝒢d(m)(p^{\prime})^{t}\colon\left|R^{t}\right|\rightarrow\mathcal{G}_{d}(\mathbb{R}^{m}) defined as

(p)t=p2tit,(p^{\prime})^{t}=p^{\sqrt{2}t}\circ i^{t},

where p2tp^{\sqrt{2}t} denotes the maps of the Čech bundle filtration (𝕏,𝕡)(\mathbb{X},\mathbbm{p}), and iti^{t} denotes the inclusion |Rt||S2t|\left|R^{t}\right|\hookrightarrow\left|S^{\sqrt{2}t}\right|. These maps fit in the following diagram:

|Rt|{\left|R^{t}\right|}|S2t|{\left|S^{\sqrt{2}t}\right|}X2t{X^{\sqrt{2}t}}𝒢d(m){\mathcal{G}_{d}(\mathbb{R}^{m})}it\scriptstyle{i^{t}}(p)t\scriptstyle{(p^{\prime})^{t}}p2t\scriptstyle{p^{\sqrt{2}t}}

This new vector bundle filtration is defined on the index set T=[0,12tγmax(X))T^{\prime}=\left[0,\frac{1}{\sqrt{2}}t_{\gamma}^{\mathrm{max}}\left(X\right)\right).

It is clear from the construction that the vector bundle filtrations (𝕏,𝕡)(\mathbb{X},\mathbbm{p}) and (||,𝕡)(\left|\mathbb{R}\right|,\mathbbm{p}^{\prime}) are multiplicatively 2\sqrt{2}-interleaved, with an interleaving that preserves the persistent Stiefel-Whitney classes. This property is a multiplicative equivalent of Theorem 3.3. Remember that if XX is a subset of n×𝒢d(m)\mathbb{R}^{n}\times\mathcal{G}_{d}(\mathbb{R}^{m}), then the maximal filtration value of the Čech bundle filtration on XX is tγmax(X)=22γt_{\gamma}^{\mathrm{max}}\left(X\right)=\frac{\sqrt{2}}{2}\gamma (see Equation (11)). Consequently, the maximal filtration value of its Vietoris-Rips version is 12γ\frac{1}{2}\gamma.

From an application perspective, we choose to work with the Vietoris-Rips filtration since it is easier to compute. Indeed, its construction only relies on computing pairwise distances, and finding cliques in graphs.

5.4 Choice of the parameter γ\gamma

This subsection is devoted to discussing the influence of the parameter γ>0\gamma>0. Recall that γ\gamma affects the norm γ\left\|\cdot\right\|_{\gamma} we chose on n×M(m)\mathbb{R}^{n}\times M(\mathbb{R}^{m}):

(x,A)γ2=x2+γ2AF2.\displaystyle\left\|(x,A)\right\|_{\gamma}^{2}=\left\|x\right\|^{2}+\gamma^{2}\left\|A\right\|_{\mathrm{F}}^{2}.

Let Xn×M(m)X\subset\mathbb{R}^{n}\times M(\mathbb{R}^{m}). If γ1γ2\gamma_{1}\leq\gamma_{2} are two positive real numbers, the corresponding Čech filtrations 𝕏1\mathbb{X}_{1} and 𝕏2\mathbb{X}_{2}, as well as the Čech bundle filtrations (𝕏1,𝕡1)(\mathbb{X}_{1},\mathbbm{p}_{1}) and (𝕏2,𝕡2)(\mathbb{X}_{2},\mathbbm{p}_{2}), are γ2γ1\frac{\gamma_{2}}{\gamma_{1}}-interleaved multiplicatively. This comes from the straightforward inequality

γ1γ2γ2γ1γ1.\displaystyle\|\cdot\|_{\gamma_{1}}~{}\leq~{}\|\cdot\|_{\gamma_{2}}~{}\leq~{}\frac{\gamma_{2}}{\gamma_{1}}\|\cdot\|_{\gamma_{1}}.

Note that we also have the additive inequality

(x,A)γ1(x,A)γ2(x,A)γ1+γ22γ12AF.\displaystyle\|(x,A)\|_{\gamma_{1}}~{}\leq~{}\|(x,A)\|_{\gamma_{2}}~{}\leq~{}\|(x,A)\|_{\gamma_{1}}+\sqrt{\gamma_{2}^{2}-\gamma_{1}^{2}}\left\|A\right\|_{\mathrm{F}}.

One deduces that the Čech bundle filtrations (𝕏1,𝕡1)(\mathbb{X}_{1},\mathbbm{p}_{1}) and (𝕏2,𝕡2)(\mathbb{X}_{2},\mathbbm{p}_{2}) are γ22γ12tmax(X)\sqrt{\gamma_{2}^{2}-\gamma_{1}^{2}}\cdot t^{\mathrm{max}}\left(X\right)-interleaved additively, where tmax(X)t^{\mathrm{max}}\left(X\right) is the maximal filtration value when γ=1\gamma=1. As a consequence of these interleavings, when the values γ1\gamma_{1} and γ2\gamma_{2} are close, the persistence barcodes and the lifebars of the persistent Stiefel-Whitney classes are close (see Theorem 3.3).

As a general principle, one would choose the parameter γ\gamma to be large, since it would lead to large filtrations. More precisely, if tγ1max(X)t_{\gamma_{1}}^{\mathrm{max}}\left(X\right) and tγ2max(X)t_{\gamma_{2}}^{\mathrm{max}}\left(X\right) denote respectively the maximal filtration values of (𝕏1,𝕡1)(\mathbb{X}_{1},\mathbbm{p}_{1}) and (𝕏2,𝕡2)(\mathbb{X}_{2},\mathbbm{p}_{2}), then tγ1max(X)=γ1tmax(X)t_{\gamma_{1}}^{\mathrm{max}}\left(X\right)=\gamma_{1}\cdot t^{\mathrm{max}}\left(X\right) and tγ2max(X)=γ2tmax(X)t_{\gamma_{2}}^{\mathrm{max}}\left(X\right)=\gamma_{2}\cdot t^{\mathrm{max}}\left(X\right), as in Equation (10). Moreover, we have the following inclusion:

X1tγ1max(X)X2tγ2max(X),\displaystyle X_{1}^{t_{\gamma_{1}}^{\mathrm{max}}\left(X\right)}\subseteq X_{2}^{t_{\gamma_{2}}^{\mathrm{max}}\left(X\right)},

where X1tγ1max(X)X_{1}^{t_{\gamma_{1}}^{\mathrm{max}}\left(X\right)} denotes the thickening of XX with respect to the norm γ1\|\cdot\|_{\gamma_{1}}, and X2tγ2max(X)X_{2}^{t_{\gamma_{2}}^{\mathrm{max}}\left(X\right)} with respect to γ2\|\cdot\|_{\gamma_{2}}. This inclusion can be proven from the following fact, valid for every xnx\in\mathbb{R}^{n} and AM(m)A\in M(\mathbb{R}^{m}) such that AFtmax(X)\left\|A\right\|_{\mathrm{F}}\leq t^{\mathrm{max}}\left(X\right):

(x,A)γ1tγ1max(X)(x,A)γ2tγ2max(X).\displaystyle\|(x,A)\|_{\gamma_{1}}\leq t_{\gamma_{1}}^{\mathrm{max}}\left(X\right)\implies\|(x,A)\|_{\gamma_{2}}\leq t_{\gamma_{2}}^{\mathrm{max}}\left(X\right).

Hence larger parameters γ\gamma lead to larger maximal filtration values and larger filtrations.

However, as we show in the following examples, different values of γ\gamma may result in different behaviours of the persistent Stiefel-Whitney classes. In Example 5.10, large values of γ\gamma highlight properties of the dataset that are not consistent with the underlying vector bundle, which is orientable. Notice that, so far, we always picked the value γ=1\gamma=1, for it seemed experimentally relevant with the datasets we chose.

Example 5.9.

Consider the set Y2×M(2)Y\subset\mathbb{R}^{2}\times M(\mathbb{R}^{2}) representing the Mobius strip, as in Example 3.2 of Subsect. 3.2:

Y={((cos(θ)sin(θ)),(cos(θ2)2cos(θ2)sin(θ2)cos(θ2)sin(θ2)sin(θ2)2)),θ[0,2π)}.\displaystyle Y=\bigg{\{}\bigg{(}\begin{pmatrix}\cos(\theta)\\ \sin(\theta)\end{pmatrix},\begin{pmatrix}\cos(\frac{\theta}{2})^{2}&\cos(\frac{\theta}{2})\sin(\frac{\theta}{2})\\ \cos(\frac{\theta}{2})\sin(\frac{\theta}{2})&\sin(\frac{\theta}{2})^{2}\end{pmatrix}\bigg{)},\theta\in[0,2\pi)\bigg{\}}.

As we show in Appendix A.1, YY is a circle, included in a 2-dimensional affine subspace of 2×M(2)\mathbb{R}^{2}\times M(\mathbb{R}^{2}). Its radius is 1+γ22\sqrt{1+\frac{\gamma^{2}}{2}}. As a consequence, the persistence of the Čech filtration of YY consists of two bars: one H0H^{0}-feature, the bar [0,+)[0,+\infty), and one H1H^{1}-feature, the bar [0,1+γ22)\left[0,\sqrt{1+\frac{\gamma^{2}}{2}}\right).

For any γ>0\gamma>0, the maximal filtration value of the Čech bundle filtration of YY is tγmax(Y)=22γt_{\gamma}^{\mathrm{max}}\left(Y\right)=\frac{\sqrt{2}}{2}\gamma. Moreover, the persistent Stiefel-Whitney class w1t(Y)w_{1}^{t}(Y) is nonzero all along the filtration. In this example, we see that the parameter γ\gamma does not influence the qualitative interpretation of the persistent Stiefel-Whitney class. It is always nonzero where it is defined. The following example shows a case where γ\gamma does influence the persistent Stiefel-Whitney class.

Example 5.10.

Consider the set X2×M(2)X\subset\mathbb{R}^{2}\times M(\mathbb{R}^{2}) representing the normal bundle of the circle 𝕊1\operatorname{\mathbb{S}}_{1}, as in Example 3.2:

X={((cos(θ)sin(θ)),(cos(θ)2cos(θ)sin(θ)cos(θ)sin(θ)sin(θ)2)),θ[0,2π)}.\displaystyle X=\bigg{\{}\bigg{(}\begin{pmatrix}\cos(\theta)\\ \sin(\theta)\end{pmatrix},\begin{pmatrix}\cos(\theta)^{2}&\cos(\theta)\sin(\theta)\\ \cos(\theta)\sin(\theta)&\sin(\theta)^{2}\end{pmatrix}\bigg{)},\theta\in[0,2\pi)\bigg{\}}.

As we show in Appendix A.2, XX is a subset of a 2-dimensional flat torus embedded in 2×M(2)\mathbb{R}^{2}\times M(\mathbb{R}^{2}), hence can be seen as a torus knot.

Before studying the Čech bundle filtration of XX, we discuss its Čech filtration 𝕏\mathbb{X}. Its behaviour depends on γ\gamma:

  • if γ22\gamma\leq\frac{\sqrt{2}}{2}, then XtX^{t} retracts on a circle for t[0,1)t\in[0,1), XtX^{t} retracts on a 3-sphere for t[1,1+12γ2)t\in\left[1,\sqrt{1+\frac{1}{2}\gamma^{2}}\right), and XtX^{t} retracts on a point for t1+12γ2t\geq\sqrt{1+\frac{1}{2}\gamma^{2}}.

  • if γ22\gamma\geq\frac{\sqrt{2}}{2}, then XtX^{t} retracts on a circle for t[0,1)t\in[0,1), XtX^{t} retracts on another circle for t[1,221+γ2+14γ2)t\in\left[1,\frac{\sqrt{2}}{2}\sqrt{1+\gamma^{2}+\frac{1}{4\gamma^{2}}}\right), XtX^{t} retracts on a 3-sphere for t[221+γ2+14γ2,1+12γ2)t\in\left[\frac{\sqrt{2}}{2}\sqrt{1+\gamma^{2}+\frac{1}{4\gamma^{2}}},\sqrt{1+\frac{1}{2}\gamma^{2}}\right), and XtX^{t} has the homotopy type of a point for t1+12γ2t\geq\sqrt{1+\frac{1}{2}\gamma^{2}}.

Let us interpret these facts. If γ22\gamma\leq\frac{\sqrt{2}}{2}, then the persistent cohomology of XX looks similar to the persistent cohomology of the underlying set {(cos(θ)sin(θ)),θ[0,2π)}2\left\{\begin{pmatrix}\cos(\theta)\\ \sin(\theta)\end{pmatrix},\theta\in[0,2\pi)\right\}\subset\mathbb{R}^{2}, but with a H3H^{3} cohomology feature added. Besides, if γ22\gamma\geq\frac{\sqrt{2}}{2}, a new topological feature appears in the H1H^{1}-barcode: the bar [1,21+γ2+14γ2)\left[1,\sqrt{2}\sqrt{1+\gamma^{2}+\frac{1}{4\gamma^{2}}}\right). These barcodes are depicted in Figures 15 and 16.

Let us now discuss the corresponding Čech bundle filtrations. For any γ>0\gamma>0, the maximal filtration value of the Čech bundle filtration of XX is tγmax(X)=22γt_{\gamma}^{\mathrm{max}}\left(X\right)=\frac{\sqrt{2}}{2}\gamma. We observe two behaviours: if γ22\gamma\leq\frac{\sqrt{2}}{2}, then w1t(X)w_{1}^{t}(X) is zero all along the filtration, and if γ>22\gamma>\frac{\sqrt{2}}{2}, then w1t(X)w_{1}^{t}(X) is nonzero from t=1t^{\dagger}=1. This in proven in Appendix A.2. To conclude, this persistent Stiefel-Whitney class is consistent with the underlying bundle—the normal bundle of the circle, which is trivial—only for t1t\leq 1.

Refer to captionRefer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to captionRefer to caption
Figure 15: H0H^{0}-, H1H^{1}-, H3H^{3}-barcodes and lifebar of the first persistent Stiefel-Whitney class of XX with γ=12\gamma=\frac{1}{2} (left) and γ=22\gamma=\frac{\sqrt{2}}{2} (right).
Refer to captionRefer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to captionRefer to caption
Figure 16: H0H^{0}-, H1H^{1}-, H3H^{3}-barcodes and lifebar of the first persistent Stiefel-Whitney class of XX with γ=1\gamma=1 (left) and γ=2\gamma=2 (right).

6 Numerical experiments

In this last section, we propose to apply our algorithm on synthetic datasets, inspired by image analysis. We aim at illustrating how the first persistent Stiefel-Whintey class may reveal two properties: when the datasets contain certain symmetries, and when the datasets are close to non-orientables vector bundles. A Python notebook gathering these experiments can be found at https://github.com/raphaeltinarrage/PersistentCharacteristicClasses/blob/master/Experiments.ipynb.

6.1 Datasets with symmetries

Giraffe moving forward.

Let us start with a simple dataset: it consists in a picture of a giraffe, that we translate to the right via cyclic permutations. The dataset contains of 150 images, each of size 150×300150\times 300 pixels, in RGB format. Since 150×300×3=135 000150\times 300\times 3=135\,000, the dataset can be enbedded in 135 000\mathbb{R}^{135\,000}. Some of these images can be seen in Figure 17.

Refer to caption
Refer to caption
Refer to caption
Figure 17: The dataset consists in a giraffe moving forward.

By performing a Principal Component Analysis (PCA), we project the dataset on the three principal axes. The result is a subset X={x1,,x150}X=\{x_{1},...,x_{150}\} of 3\mathbb{R}^{3}. As a last pre-processing step, we divide each point of XX by the value max{x,xX}\max\{\left\|x\right\|,~{}x\in X\}, so that XX becomes a subset of the unit ball ¯(0,1)3\overline{\mathcal{B}}\left(0,1\right)\subset\mathbb{R}^{3}. The point cloud XX is represented on Figure 18. Next to it, we give the persistence barcodes of its Rips filtration (we choose the coefficient field 2\mathbb{Z}_{2}, and represent H0H^{0} in red and H1H^{1} in green). Note that XX actually lies close to the unit sphere 𝕊2\operatorname{\mathbb{S}}_{2} of 3\mathbb{R}^{3}, and moreover that it describes a circle.

Refer to caption
Refer to captionRefer to caption
Figure 18: Left: The point cloud X3X\subset\mathbb{R}^{3}. Right: The barcode of its Rips filtration.

Let us consider on XX two 11-dimensional bundles, defined via classifying maps X𝒢1(3)M(m)X\rightarrow\mathcal{G}_{1}(\mathbb{R}^{3})\subset M(\mathbb{R}^{m}):

p:xiP(xi)p\colon x_{i}\mapsto P(x_{i})            and            p:xiP(xi+1xi1)p^{\prime}\colon x_{i}\mapsto P(x_{i+1}-x_{i-1})

where P(x)P(x) represents the orthogonal projection matrix on the 1-dimensional subspace of 3\mathbb{R}^{3} spanned by xx. The vector bundle pp is to be seen as the normal bundle of 𝕊2\operatorname{\mathbb{S}}_{2} restricted to XX, and pp^{\prime} is to be seen as the tangent bundle of the circle. These two theoretical bundles — restriction of the normal bundle of the sphere, and tangent bundle of a circle — are trivial. This follows from the general fact that 1-dimensional bundles on the sphere are trivial.

We represent on Figure 19 the vector bundles pp and pp^{\prime}, seen in 3\mathbb{R}^{3}. One observes that, while going around the circle, the lines make a complete twist. This is the same behavior as the trivial bundle of the circle, that we studied earlier (see Figure 2).

Refer to caption
Refer to caption
Figure 19: The vector bundles pp and pp^{\prime}.

Next, let γ=1\gamma=1, and consider the lifted sets Xˇ={(x,γp(x)),xX}\check{X}=\left\{\left(x,~{}\gamma p(x)\right),~{}x\in X\right\} and Xˇ={(x,γp(x)),xX}\check{X}^{\prime}=\left\{\left(x,~{}\gamma p^{\prime}(x)\right),~{}x\in X\right\}. We remind the reader that this construction has been studied in Subsection 3.4. We represent the point clouds Xˇ\check{X} and Xˇ\check{X}^{\prime} on Figure 20, projected in 3\mathbb{R}^{3} via PCA, as well as the persistence barcodes of their Rips filtration. On these two barcodes, one observes one prominent H0H^{0}-feature, and one prominent H1H^{1}-feature. Below, we plot the lifebars of their first persistent Stiefel-Whitney classes, up to the maximal filtration value 12γ\frac{1}{2}\gamma (see Subsection 5.3). Both are empty, meaning that the persistent Stiefel-Whitney classes are zero all along the filtration. This is consistent with the fact that the underlying vector bundles are trivial.

Refer to captionRefer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to captionRefer to caption
Figure 20: Left: The set Xˇ\check{X} (projected in 3\mathbb{R}^{3} via PCA), the barcodes of its Rips filtration, and the lifebar of its first persistent Stiefel-Witney class. Right: Same for Xˇ\check{X}^{\prime}. Only green bars of length larger than 0.10.1 are represented.

Giraffe moving behind two trees.

We will now study a variation of this dataset. As before, we consider a giraffe walking straight, but now with two trees on the foreground. The dataset consists of 150 images, each of size 130×300130\times 300 pixels, in RGB format. Since 130×300×3=117 000130\times 300\times 3=117\,000, the dataset can be seen as a subset of 117 000\mathbb{R}^{117\,000}. Some of the images can be seen in Figure 21.

Refer to caption
Refer to caption
Refer to caption
Figure 21: The dataset consists in a giraffe moving forward, with two trees on the foreground.

Just as before, we project the point cloud in 3\mathbb{R}^{3} via PCA, and renormalize it. The corresponding point cloud, that we denote Y={y1,,y150}Y=\{y_{1},...,y_{150}\}, and the barcode of its Rips filtration can be seen on Figure 22.

Refer to caption
Refer to captionRefer to caption
Figure 22: Left: The point cloud Y3Y\subset\mathbb{R}^{3}. Right: The barcode of its Rips filtration. Only green bars of length larger than 0.050.05 are represented.

Note that, in this collection, the images where the giraffe goes behind the trees are close to each other. Indeed, only a few pixels differ between them. As a consequence, the point cloud YY, though lying close to a circle, seems to come closer to itself at some point. This behavior translates in its persistent cohomology as two prominent H1H^{1}-features.

We now consider the vector bundles qq and qq^{\prime} defined as before:

q:yiP(yi)q\colon y_{i}\mapsto P(y_{i})            and            q:yiP(yi+1yi1)q^{\prime}\colon y_{i}\mapsto P(y_{i+1}-y_{i-1})

They are represented on Figure 23. Let γ=0.5\gamma=0.5, and consider the lifted sets Yˇ={(y,γq(y)),yY}\check{Y}=\left\{\left(y,~{}\gamma q(y)\right),~{}y\in Y\right\} and Yˇ={(y,γq(y)),yY}3×𝒢1(3)\check{Y}^{\prime}=\left\{\left(y,~{}\gamma q^{\prime}(y)\right),~{}y\in Y\right\}\subset\mathbb{R}^{3}\times\mathcal{G}_{1}(\mathbb{R}^{3}). We represent them on Figure 24, as well as the persistence barcode of their Rips filtration, and the lifebars of their first Stiefel-Whitney classes, up to their maximal filtration value. We read that the second one is empty, while the first one is not.

Refer to caption
Refer to caption
Figure 23: The vector bundles qq and qq^{\prime}.
Refer to captionRefer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to captionRefer to caption
Figure 24: Left: The set Yˇ\check{Y} (projected in 3\mathbb{R}^{3} via PCA), the barcodes of its Rips filtration, and the lifebar of its first persistent Stiefel-Witney class. Right: Same for Yˇ\check{Y}^{\prime}. Only green bars of length larger than 0.050.05 are represented.

Let us explain this phenomenon. In both cases, as we can see on Figure 23, the lines make a full twist while turning around the circle. This corresponds to a trivial bundle. However, in the case of qq, the points xx close to the almost self-intersection of YY corresponds to lines q(x)q(x) close to each other. As a consequence, in the Rips filtration of the lifted set Yˇ\check{Y}, these points will connect early. Hence the filtration behaves as if YY were composed of two loops. But, on these loops, the lines make only a half-twist. This correspond to the Mobius bundle on the circle (see Figure 2). Therefore we obtain a non-trivial Stifel-Whitney class.

The bundle qq^{\prime} does not reflect this property. This is because the points xx close to the self-intersection of YY does not correspond to lines q(x)q^{\prime}(x) that are close to each other. In the Rips filtration of the lifted set Yˇ\check{Y}^{\prime}, the two loops connect late, hence the non-orientability does not appear on its persistent Stiefel-Whitney class.

Rotating cylinders.

We now propose a dataset whose tangent bundle reflects non-orientability. Consider the union of two cylinders in 3\mathbb{R}^{3}, as represented on Figure 25. By applying rotations, we obtain a dataset of 100 pictures, in RGB format, of 500×500500\times 500 pixels.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 25: The dataset consists in different views of two cylinders in 3\mathbb{R}^{3}.

As before, we embed them into 750 000\mathbb{R}^{750\,000}, project them into 3\mathbb{R}^{3} via PCA, and renormalize them. The corresponding point cloud, denoted Z={z1,,z100}Z=\{z_{1},...,z_{100}\}, and the barcodes of its Rips filtration are represented on Figure 26.

Refer to caption
Refer to captionRefer to caption
Figure 26: Left: The point cloud Z3Z\subset\mathbb{R}^{3}. Right: The barcode of its Rips filtration. Only green bars of length larger than 0.020.02 are represented.

Define on ZZ the tangent bundle r:ziP(zi+1zi1)𝒢1(3)r\colon z_{i}\mapsto P(z_{i+1}-z_{i-1})\in\mathcal{G}_{1}(\mathbb{R}^{3}). Let γ=0.2\gamma=0.2 and consider the lifted set Zˇ={(z,γr(z)),zZ}3×𝒢1(3)\check{Z}=\left\{\left(z,\gamma r(z)\right),~{}z\in Z\right\}\subset\mathbb{R}^{3}\times\mathcal{G}_{1}(\mathbb{R}^{3}). The persistence barcodes of its Rips filtration, and the lifebar of its first persistent Stiefel-Whitney class are represented on Figure 27. We see that the lifebar is not trivial.

Refer to captionRefer to captionRefer to caption
Figure 27: The barcodes of the Rips filtration of Zˇ\check{Z}, and the lifebar of its first persistent Stiefel-Witney class. Only green bars of length larger than 0.020.02 are represented.

Here, the same phenomenon as before occurs: there are two images of the collection which are almost equal (when the cylinders are one in front of the other), resulting in a point cloud that almost intersect itself. Moreover, points zz close to the area where ZZ almost intersect itself correspond to lines r(z)r(z) that are close to each other. Consequently, the persistent homology of Zˇ\check{Z} shows two prominent loops, whose corresponding vector bundles are non-orientable.

In these two last experiments, we observed the following fact: trivial vector bundles, but whose underlying point clouds present self-similarity, or self-intersection, may result in a shortcut of the vector bundle, implying a non-trivial persistent Stiefel-Whitney class.

6.2 Datasets with intrisic (non-)orientability

We will now present three datasets which reflect some underlying theoretical orientability or non-orientability.

Gorilla on a torus.

Let us consider a picture of a gorilla, that we translate to the right and downwards via cyclic permutations (see Figure 28). Each image has size 130×120130\times 120 pixels, in RGB format. The dataset consists of 3900 images (65 vertical permutations and 60 horizontal permutations).

Note that the images behave as if the gorilla was on a torus. Indeed, the torus can be obtained from the square [0,1]2[0,1]^{2} by gluing its opposite edges. Hence the various images can be seen as translations of the original image, whose gluing pattern follows the one of the torus.

Since 130×120×3=46 800130\times 120\times 3=46\,800, the dataset can be seen as a point cloud of 46 800\mathbb{R}^{46\,800}, that we project into 4\mathbb{R}^{4} via PCA. The resulting subset is denoted X={xi,j,i1,65,j1,60}X=\left\{x_{i,j},~{}i\in\llbracket 1,65\rrbracket,j\in\llbracket 1,60\rrbracket\right\}. The first index ii correponds to translations downwards, and the second index jj corresponds to translations to the right.

Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Figure 28: The dataset consists in a gorilla moving forward and downwards on a torus.

Consider on XX the vertical tangent bundle and the horizontal tangent bundle, defined respectively as

p:xi,jP(xi1,jxi+1,j)p\colon x_{i,j}\mapsto P(x_{i-1,j}-x_{i+1,j})            and            p:xi,jP(xi,j1xi,j+1)p^{\prime}\colon x_{i,j}\mapsto P(x_{i,j-1}-x_{i,j+1})

where we recall that P(x)P(x) is the orthogonal projection matrix on the line spanned by xx. The vector bundle pp is to be seen as the vertical component of the tangent bundle of a torus, and pp^{\prime} as the horizontal one. Both are trivial bundles.

Now, let γ=0.3\gamma=0.3, and consider the corresponding lifted sets Xˇ={(x,γp(x)),xX}\check{X}=\left\{\left(x,\gamma p(x)\right),~{}x\in X\right\} and Xˇ={(x,γp(x)),xX}3×𝒢1(3)\check{X}^{\prime}=\left\{\left(x,\gamma p^{\prime}(x)\right),~{}x\in X\right\}\subset\mathbb{R}^{3}\times\mathcal{G}_{1}(\mathbb{R}^{3}). We represent on Figure 29 the barcodes of the Rips filtrations of the sets Xˇ\check{X} and Xˇ\check{X}^{\prime}, and the lifebars of their first persistent Stiefel-Whitney classes. We read that they are zero all along the filtration. This is consistent with the underlying line bundles on the torus being trivial.

Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Figure 29: Left: The barcodes of the Rips filtration of Xˇ\check{X}, and the lifebar of its first persistent Stiefel-Witney class. Right: Same for Xˇ\check{X}^{\prime}. Only green bars of length larger than 0.030.03 are represented.

Gorilla on a Klein bottle.

We will modify this dataset: we still translate the gorilla, but while translating it to the right, we inverse the part that arrives at the left (see Figure 30). It behaves as if the picture was glued to itself according to the gluing of a Klein bottle. Just as before, the dataset consists of 3900 images of size 130×120130\times 120 pixels. We embed the images into 46 800\mathbb{R}^{46\,800} and project them into 4\mathbb{R}^{4} via PCA. The resulting subset is denoted Y={yi,j,i1,65,j1,60}Y=\{y_{i,j},~{}i\in\llbracket 1,65\rrbracket,j\in\llbracket 1,60\rrbracket\}.

Again, consider the vector bundles q:yi,jP(yi1,jyi+1,j)q\colon y_{i,j}\mapsto P(y_{i-1,j}-y_{i+1,j}) and q:yi,jP(yi,j1yi,j+1)q^{\prime}\colon y_{i,j}\mapsto P(y_{i,j-1}-y_{i,j+1}). They correspond to the horizontal and vertical components of the tangent bundle of a Klein bottle. Only one of them is non-trivial. Let γ=0.3\gamma=0.3, and consider the corresponding lifted sets Yˇ\check{Y} and Yˇ\check{Y}^{\prime}. We represent on Figure 31 the barcodes of their Rips filtrations, and the lifebars of their first persistent Stiefel-Whitney classes. The first lifebar is non-empty, reflecting the non-triviality of the underlying bundle.

Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Figure 30: The dataset consists in a gorilla moving forward and downwards on a Klein bottle.
Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Figure 31: Left: The barcodes of the Rips filtration of Yˇ\check{Y}, and the lifebar of its first persistent Stiefel-Witney class. Right: Same for Yˇ\check{Y}^{\prime}. Only green bars of length larger than 0.030.03 are represented.

Gorilla on the projective plane.

We close this subsection with a last variation of the dataset: the gorilla is translated to the right, with an inversion of the left part, and translated downwards, with also an inversion of the upper part (see Figure 32). It behaves as if the image was glued to itself according to the gluing of the projective plane 𝒢1(3)\mathcal{G}_{1}(\mathbb{R}^{3}). The dataset still consists of 3900 images of size 130×120130\times 120 pixels, that we embed into 46 800\mathbb{R}^{46\,800} and project into 4\mathbb{R}^{4} via PCA. The resulting subset is denoted Z={zi,j,i1,65,j1,60}Z=\{z_{i,j},~{}i\in\llbracket 1,65\rrbracket,j\in\llbracket 1,60\rrbracket\}.

We still consider the vector bundles r:zi,jP(zi1,jzi+1,j)r\colon z_{i,j}\mapsto P(z_{i-1,j}-z_{i+1,j}) and r:zi,jP(zi,j1zi,j+1)r^{\prime}\colon z_{i,j}\mapsto P(z_{i,j-1}-z_{i,j+1}). Let γ=0.3\gamma=0.3, and consider the corresponding lifted sets Zˇ\check{Z} and Zˇ\check{Z}^{\prime}. We represent on Figure 33 the barcodes of their Rips filtrations, and the lifebars of their first persistent Stiefel-Whitney classes. Both lifebars are non-empty. We deduce that the underlying line bundles are non-trivial.

Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Figure 32: The dataset consists in a gorilla moving forward and downwards on the projective plane.
Refer to captionRefer to captionRefer to caption
Refer to captionRefer to captionRefer to caption
Figure 33: Left: The barcodes of the Rips filtration of Zˇ\check{Z}, and the lifebar of its first persistent Stiefel-Witney class. Right: Same for Zˇ\check{Z}^{\prime}. Only green bars of length larger than 0.030.03 are represented.

In these various experiments, we applied the same method: in ordered datasets, indexed by only one value ixii\mapsto x_{i}, two values (i,j)xi,j(i,j)\mapsto x_{i,j} or more, one can consider directional line bundles by approximating partial derivatives, such as 12(xi+1,jxi1,j)\frac{1}{2}(x_{i+1,j}-x_{i-1,j}). The first persistent Stiefel-Whitney classes of such bundles deliver information about the dataset in this particular direction.

7 Conclusion

In this paper we defined the persistent Stiefel-Whitney classes of vector bundle filtrations. We proved that they are stable with respect to the interleaving distance between vector bundle filtrations. We studied the particular case of Čech bundle filtrations of subsets of n×M(m)\mathbb{R}^{n}\times M(\mathbb{R}^{m}), and showed that they yield consistent estimators of the usual Stiefel-Whitney classes of some underlying vector bundle.

Moreover, when the dimension of the bundle is 1 and XX is finite, we proposed an algorithm to compute the first persistent Stiefel-Whitney class. We also described a way to compute their lifebars via mapping cones.

Our algorithm is limited to the bundles of dimension 1 since we only implemented triangulations of the Grassmannian 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) when d=1d=1. However, any other triangulation of 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}), with a computable face map, could be included in the algorithm without any modification. As far as we know, no triangulation of a Grassmannian 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) with d>1d>1 has never been given explicitely (i.e., as a list of simplices stored in a computer). This is a problem of interest, which will be adressed in a further work. A strategy could consist in using the usual CW-structures of the Grassmannians, and converting them into simplicial complexes, as done theoretically in Hatcher (2002, Theorem 2C.5). A recent result of Govc et al. (2020) gives an idea about the complexity of this problem: the number of simplices of minimal triangulations of 𝒢d(m)\mathcal{G}_{d}(\mathbb{R}^{m}) must grow exponentially in both dd and mm.

Acknowledgements.

I wish to thank Frédéric Chazal and Marc Glisse for fruitful discussions and corrections, as well as the anonymous reviewers for corrections and clarifications. I also thank Luis Scoccola for pointing out a strengthening of Lemma 4.1.

Appendix A Supplementary material for Sect. 5

A.1 Study of Example 5.9

We consider the set

X={((cos(θ)sin(θ)),(cos(θ2)2cos(θ2)sin(θ2)cos(θ2)sin(θ2)sin(θ2)2)),θ[0,2π)}.\displaystyle X=\bigg{\{}\bigg{(}\begin{pmatrix}\cos(\theta)\\ \sin(\theta)\end{pmatrix},\begin{pmatrix}\cos(\frac{\theta}{2})^{2}&\cos(\frac{\theta}{2})\sin(\frac{\theta}{2})\\ \cos(\frac{\theta}{2})\sin(\frac{\theta}{2})&\sin(\frac{\theta}{2})^{2}\end{pmatrix}\bigg{)},\theta\in[0,2\pi)\bigg{\}}.

In order to study the Čech filtration of XX, we shall apply the following affine transformation: let YY be the subset of 2×M(2)\mathbb{R}^{2}\times M(\mathbb{R}^{2}) defined as

Y={((cos(θ)sin(θ)),γ(cos(θ2)2cos(θ2)sin(θ2)cos(θ2)sin(θ2)sin(θ2)2)),θ[0,2π)}.\displaystyle Y=\bigg{\{}\bigg{(}\begin{pmatrix}\cos(\theta)\\ \sin(\theta)\end{pmatrix},\gamma\begin{pmatrix}\cos(\frac{\theta}{2})^{2}&\cos(\frac{\theta}{2})\sin(\frac{\theta}{2})\\ \cos(\frac{\theta}{2})\sin(\frac{\theta}{2})&\sin(\frac{\theta}{2})^{2}\end{pmatrix}\bigg{)},\theta\in[0,2\pi)\bigg{\}}.

and let 𝕐=(Yt)t0\mathbb{Y}=(Y^{t})_{t\geq 0} be the Čech filtration of YY in 2×M(2)\mathbb{R}^{2}\times M(\mathbb{R}^{2}) endowed with the norm (x,A)1=x2+AF2\left\|(x,A)\right\|_{1}=\sqrt{\left\|x\right\|^{2}+\left\|A\right\|_{\mathrm{F}}^{2}}. We recall that the Čech filtration of XX, denoted 𝕏=(Xt)t0\mathbb{X}=(X^{t})_{t\geq 0}, is defined with respect to the norm γ\left\|\cdot\right\|_{\gamma}. It is clear that, for every t0t\geq 0, the thickenings XtX^{t} and YtY^{t} are homeomorphic via the application

h:2×M(2)\displaystyle h\colon\mathbb{R}^{2}\times M(\mathbb{R}^{2}) 2×M(2)\displaystyle\longrightarrow\mathbb{R}^{2}\times M(\mathbb{R}^{2})
(x,A)\displaystyle(x,A) (x,γA).\displaystyle\longmapsto(x,\gamma A).

As a consequence, the persistence cohomology modules associated to 𝕏\mathbb{X} and 𝕐\mathbb{Y} are isomorphic.

Next, notice that YY is a subset of the affine subspace of dimension 2 of 2×M(2)\mathbb{R}^{2}\times M(\mathbb{R}^{2}) with origin OO and spanned by the vectors e1e_{1} and e2e_{2}, where

O\displaystyle O =((00),γ2(1001)),e1=((10),γ2(1001)),e2=((01),γ2(0110)).\displaystyle=\left(\begin{pmatrix}0\\ 0\end{pmatrix},\frac{\gamma}{2}\begin{pmatrix}1&0\\ 0&1\end{pmatrix}\right),~{}~{}~{}~{}~{}e_{1}=\left(\begin{pmatrix}1\\ 0\end{pmatrix},\frac{\gamma}{2}\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}\right),~{}~{}~{}~{}~{}e_{2}=\left(\begin{pmatrix}0\\ 1\end{pmatrix},\frac{\gamma}{2}\begin{pmatrix}0&1\\ 1&0\end{pmatrix}\right).

Indeed, using the equality

(cos(θ2)2cos(θ2)sin(θ2)cos(θ2)sin(θ2)sin(θ2)2))=12(1001)+12(cos(θ)sin(θ)sin(θ)cos(θ)),\displaystyle\begin{pmatrix}\cos(\frac{\theta}{2})^{2}&\cos(\frac{\theta}{2})\sin(\frac{\theta}{2})\\ \cos(\frac{\theta}{2})\sin(\frac{\theta}{2})&\sin(\frac{\theta}{2})^{2}\end{pmatrix}\bigg{)}=\frac{1}{2}\begin{pmatrix}1&0\\ 0&1\end{pmatrix}+\frac{1}{2}\begin{pmatrix}\cos(\theta)&\sin(\theta)\\ \sin(\theta)&-\cos(\theta)\end{pmatrix},

we obtain

Y=O+{cos(θ)e1+sin(θ)e2,θ[0,2π)}.\displaystyle Y=O+\left\{\cos(\theta)e_{1}+\sin(\theta)e_{2},~{}~{}\theta\in[0,2\pi)\right\}.

We see that YY is a circle, of radius e1=e2=1+γ22\left\|e_{1}\right\|=\left\|e_{2}\right\|=\sqrt{1+\frac{\gamma^{2}}{2}}.

Let EE denotes the affine space with origin OO and spanned by the vectors e1e_{1} and e2e_{2}. Lemma A.1, stated below, shows that the persistent cohomology of YY, seen in the ambient space 2×M(2)\mathbb{R}^{2}\times M(\mathbb{R}^{2}), is the same as the persistent cohomology of YY restricted to the subspace EE. As a consequence, YY has the same persistence as a circle of radius 1+γ22\sqrt{1+\frac{\gamma^{2}}{2}} in the plane. Hence its barcode can be described as follows:

  • one H0H^{0}-feature: the bar [0,+)[0,+\infty),

  • one H1H^{1}-feature: the bar [0,1+γ22)\left[0,\sqrt{1+\frac{\gamma^{2}}{2}}\right).

Lemma A.1.

Let YnY\subset\mathbb{R}^{n} be any subset, and define Yˇ=Y×{(0,,0)}n×m\check{Y}=Y\times\{(0,...,0)\}\subset\mathbb{R}^{n}\times\mathbb{R}^{m}. Let these spaces be endowed with the usual Euclidean norms. Then the Čech filtrations of YY and Yˇ\check{Y} yields isomorphic persistence modules.

Proof.

Let projn:n×mn\mathrm{proj}_{n}\colon\mathbb{R}^{n}\times\mathbb{R}^{m}\rightarrow\mathbb{R}^{n} be the projection on the first nn coordinates. One verifies that, for every t0t\geq 0, the map projn:YˇtYt\mathrm{proj}_{n}\colon\check{Y}^{t}\rightarrow Y^{t} is a homotopy equivalence. At cohomology level, these maps induce an isomorphism of persistence modules. ∎

Let us now study the Čech bundle filtration of YY, denoted (𝕐,𝕡)(\mathbb{Y},\mathbbm{p}). According to Equation (11), its filtration maximal value is tmax(Y)=tγmax(X)=γ2t^{\mathrm{max}}\left(Y\right)=t_{\gamma}^{\mathrm{max}}\left(X\right)=\frac{\gamma}{\sqrt{2}}. Note that γ2\frac{\gamma}{\sqrt{2}} is lower than 1+γ22\sqrt{1+\frac{\gamma^{2}}{2}}, which is the radius of the circle YY. Hence, for t<tmax(Y)t<t^{\mathrm{max}}\left(Y\right), the inclusion YYtY\hookrightarrow Y^{t} is a homotopy equivalence. Consider the following commutative diagram:

Y{Y}Yt{Y^{t}}𝒢1(2){\mathcal{G}_{1}(\mathbb{R}^{2})}p0\scriptstyle{p^{0}}pt\scriptstyle{p^{t}}

It induces the following diagram in cohomology:

H(Y){H^{*}(Y)}H(Yt){H^{*}(Y^{t})}H(𝒢1(2)){H^{*}(\mathcal{G}_{1}(\mathbb{R}^{2}))}\scriptstyle{\sim}(p0)\scriptstyle{(p^{0})^{*}}(pt)\scriptstyle{(p^{t})^{*}}

The horizontal arrow is an isomorphism. Hence the map (pt):H(Yt)H(𝒢1(2))(p^{t})^{*}\colon H^{*}(Y^{t})\leftarrow H^{*}(\mathcal{G}_{1}(\mathbb{R}^{2})) is equal to (p0)(p^{0})^{*}. We only have to understand (p0)(p^{0})^{*}.

Remark that the map p0:Y𝒢1(2)p^{0}\colon Y\rightarrow\mathcal{G}_{1}(\mathbb{R}^{2}) can be seen as the tautological bundle of the circle. It is then a standard result that (p0):H(Y)H(𝒢1(2))(p^{0})^{*}\colon H^{*}(Y)\leftarrow H^{*}(\mathcal{G}_{1}(\mathbb{R}^{2})) is nontrivial. Alternatively, p0p^{0} can be seen as a map between two circles. It is injective, hence its degree (modulo 2) is 1. We still deduce that (p0)(p^{0})^{*} is nontrivial. As a consequence, the persistent Stiefel-Whitney class w1t(X)w_{1}^{t}(X) is nonzero for every t<tmax(Y)t<t^{\mathrm{max}}\left(Y\right).

A.2 Study of Example 5.10

We consider the set

X={((cos(θ)sin(θ)),(cos(θ)2cos(θ)sin(θ)cos(θ)sin(θ)sin(θ)2)),θ[0,2π)}.\displaystyle X=\bigg{\{}\bigg{(}\begin{pmatrix}\cos(\theta)\\ \sin(\theta)\end{pmatrix},\begin{pmatrix}\cos(\theta)^{2}&\cos(\theta)\sin(\theta)\\ \cos(\theta)\sin(\theta)&\sin(\theta)^{2}\end{pmatrix}\bigg{)},\theta\in[0,2\pi)\bigg{\}}.

As we explained in the previous subsection, the Čech filtration of XX with respect to the norm γ\left\|\cdot\right\|_{\gamma} yields the same persistence as the Čech filtration of YY with respect to the norm 1\left\|\cdot\right\|_{1}, where

Y={((cos(θ)sin(θ)),γ(cos(θ)2cos(θ)sin(θ)cos(θ)sin(θ)sin(θ)2)),θ[0,2π)}.\displaystyle Y=\bigg{\{}\bigg{(}\begin{pmatrix}\cos(\theta)\\ \sin(\theta)\end{pmatrix},\gamma\begin{pmatrix}\cos(\theta)^{2}&\cos(\theta)\sin(\theta)\\ \cos(\theta)\sin(\theta)&\sin(\theta)^{2}\end{pmatrix}\bigg{)},\theta\in[0,2\pi)\bigg{\}}.

Notice that YY is a subset of the affine subspace of dimension 4 of 2×M(2)\mathbb{R}^{2}\times M(\mathbb{R}^{2}) with origin O=((00),12(1001))O=\left(\begin{pmatrix}0\\ 0\end{pmatrix},\frac{1}{2}\begin{pmatrix}1&0\\ 0&1\end{pmatrix}\right) and spanned by the vectors e1,e2,e3e_{1},e_{2},e_{3} and e4e_{4}, where

e1=\displaystyle e_{1}= ((10),(0000)),e2=((01),(0000)),\displaystyle\left(\begin{pmatrix}1\\ 0\end{pmatrix},\begin{pmatrix}0&0\\ 0&0\end{pmatrix}\right),~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}e_{2}=\left(\begin{pmatrix}0\\ 1\end{pmatrix},\begin{pmatrix}0&0\\ 0&0\end{pmatrix}\right),
e3=12\displaystyle e_{3}=\frac{1}{\sqrt{2}} ((00),(1001)),e4=12((00),(0110)).\displaystyle\left(\begin{pmatrix}0\\ 0\end{pmatrix},\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}\right),~{}~{}~{}~{}~{}~{}~{}e_{4}=\frac{1}{\sqrt{2}}\left(\begin{pmatrix}0\\ 0\end{pmatrix},\begin{pmatrix}0&1\\ 1&0\end{pmatrix}\right).

Indeed, YY can be written as

Y=O+{cos(θ)e1+sin(θ)e2+γ2cos(2θ)e3+γ2sin(2θ)e4,θ[0,2π)}.\displaystyle Y=O+\left\{\cos(\theta)e_{1}+\sin(\theta)e_{2}+\frac{\gamma}{\sqrt{2}}\cos(2\theta)e_{3}+\frac{\gamma}{\sqrt{2}}\sin(2\theta)e_{4},~{}~{}\theta\in[0,2\pi)\right\}.

This comes from the equality

(cos(θ)2cos(θ)sin(θ)cos(θ)sin(θ)sin(θ)2)=12(1001)+12(cos(2θ)sin(2θ)sin(2θ)cos(2θ)).\displaystyle\begin{pmatrix}\cos(\theta)^{2}&\cos(\theta)\sin(\theta)\\ \cos(\theta)\sin(\theta)&\sin(\theta)^{2}\end{pmatrix}=\frac{1}{2}\begin{pmatrix}1&0\\ 0&1\end{pmatrix}+\frac{1}{2}\begin{pmatrix}\cos(2\theta)&\sin(2\theta)\\ \sin(2\theta)&-\cos(2\theta)\end{pmatrix}.

Observe that YY is a torus knot, i.e. a simple closed curve included in the torus 𝕋\mathbb{T}, defined as

𝕋=O+{cos(θ)e1+sin(θ)e2+γ2cos(ν)e3+γ2sin(ν)e4,θ,ν[0,2π)}.\displaystyle\mathbb{T}=O+\left\{\cos(\theta)e_{1}+\sin(\theta)e_{2}+\frac{\gamma}{\sqrt{2}}\cos(\nu)e_{3}+\frac{\gamma}{\sqrt{2}}\sin(\nu)e_{4},~{}~{}\theta,\nu\in[0,2\pi)\right\}.

The curve YY winds one time around the first circle of the torus, and two times around the second one, as represented in Figure 34. It is known as the torus knot (1,2)(1,2).

Let EE denotes the affine subspace with origin OO and spanned by e1,e2,e3,e4e_{1},e_{2},e_{3},e_{4}. Since YY is a subset of EE, it is equivalent to study the Čech filtration of YY restricted to EE (as in Lemma A.1). We shall denote the coordinates of points xEx\in E with respect to the orthonormal basis (e1,e2,e3,e4)(e_{1},e_{2},e_{3},e_{4}). That is, a tuple (x1,x2,x3,x4)(x_{1},x_{2},x_{3},x_{4}) shall refer to the point O+x1e1+x2e2+x3e3+x4e4O+x_{1}e_{1}+x_{2}e_{2}+x_{3}e_{3}+x_{4}e_{4} of EE. Seen in EE, the set YY can be written as

Y={(cos(θ),sin(θ),γ2cos(2θ),γ2sin(2θ)),θ[0,2π)}.\displaystyle Y=\left\{\left(\cos(\theta),\sin(\theta),\frac{\gamma}{\sqrt{2}}\cos(2\theta),\frac{\gamma}{\sqrt{2}}\sin(2\theta)\right),\theta\in[0,2\pi)\right\}.

Moreover, for every θ[0,2π)\theta\in[0,2\pi), we shall denote yθ=(cos(θ),sin(θ),γ2cos(2θ),γ2sin(2θ))y_{\theta}=\left(\cos(\theta),\sin(\theta),\frac{\gamma}{\sqrt{2}}\cos(2\theta),\frac{\gamma}{\sqrt{2}}\sin(2\theta)\right).

Refer to caption
Refer to caption
Figure 34: Representations of the set YY, lying on a torus, for a small value of γ\gamma (left) and a large value of γ\gamma (right).

We now state two lemmas that will be useful in what follows.

Lemma A.2.

For every θ[0,2π)\theta\in[0,2\pi), the map θyθyθ\theta^{\prime}\mapsto\left\|y_{\theta}-y_{\theta^{\prime}}\right\| admits the following critical points:

  • θθ=0\theta^{\prime}-\theta=0 and θθ=π\theta^{\prime}-\theta=\pi if γ12\gamma\leq\frac{1}{\sqrt{2}},

  • θθ=0\theta^{\prime}-\theta=0, π\pi, arccos(12γ2)\arccos(-\frac{1}{2\gamma^{2}}) and arccos(12γ2)-\arccos(-\frac{1}{2\gamma^{2}}) if γ12\gamma\geq\frac{1}{\sqrt{2}}.

They correspond to the values

  • yθyθ=0\left\|y_{\theta}-y_{\theta^{\prime}}\right\|=0  if  θθ=0\theta^{\prime}-\theta=0,

  • yθyθ=2\left\|y_{\theta}-y_{\theta^{\prime}}\right\|=2  if  θθ=π\theta^{\prime}-\theta=\pi,

  • yθyθ=21+γ2+14γ2\left\|y_{\theta}-y_{\theta^{\prime}}\right\|=\sqrt{2}\sqrt{1+\gamma^{2}+\frac{1}{4\gamma^{2}}}  if  θθ=±arccos(12γ2)\theta^{\prime}-\theta=\pm\arccos(-\frac{1}{2\gamma^{2}}).

Moreover, we have 21+γ2+14γ22\sqrt{2}\sqrt{1+\gamma^{2}+\frac{1}{4\gamma^{2}}}\geq 2 when γ12\gamma\geq\frac{1}{\sqrt{2}}.

Proof.

Let θ,θ[0,2π)\theta,\theta^{\prime}\in[0,2\pi). One computes that

yθyθ2=4sin2(θθ2)+2γ2sin2(θθ).\left\|y_{\theta}-y_{\theta^{\prime}}\right\|^{2}=4\sin^{2}\left(\frac{\theta-\theta^{\prime}}{2}\right)+2\gamma^{2}\sin^{2}(\theta-\theta^{\prime}).

Consider the map f:x[0,2π)4sin2(x2)+2γ2sin2(x)f\colon x\in[0,2\pi)\mapsto 4\sin^{2}\left(\frac{x}{2}\right)+2\gamma^{2}\sin^{2}(x). Its derivative is

f(x)\displaystyle f^{\prime}(x) =4cos(x2)sin(x2)+4γ2cos(x)sin(x)\displaystyle=4\cos\left(\frac{x}{2}\right)\sin\left(\frac{x}{2}\right)+4\gamma^{2}\cos(x)\sin(x)
=2sin(x)(1+2γ2cos(x)).\displaystyle=2\sin(x)\left(1+2\gamma^{2}\cos(x)\right).

It vanishes when x=0x=0, x=πx=\pi, or x=±arccos(12γ2)x=\pm\arccos(-\frac{1}{2\gamma^{2}}) if γ12\gamma\geq\frac{1}{\sqrt{2}}. To conclude, a computation shows that f(0)=0f(0)=0, f(π)=4f(\pi)=4 and f(±arccos(12γ2))=2(1+γ2+14γ2)f\left(\pm\arccos\left(-\frac{1}{2\gamma^{2}}\right)\right)=2\left(1+\gamma^{2}+\frac{1}{4\gamma^{2}}\right). ∎

Lemma A.3.

For every xEx\in E such that x0x\neq 0, the map θxyθ\theta\mapsto\left\|x-y_{\theta}\right\| admits at most two local maxima and two local minima.

Proof.

Consider the map g:θ[0,2π)xyθ2g\colon\theta\in[0,2\pi)\mapsto\left\|x-y_{\theta}\right\|^{2}. It can be written as

g(θ)\displaystyle g(\theta) =x2+yθ22x,yθ\displaystyle=\left\|x\right\|^{2}+\left\|y_{\theta}\right\|^{2}-2\left\langle x,y_{\theta}\right\rangle
=x2+1+γ222x,yθ.\displaystyle=\left\|x\right\|^{2}+1+\frac{\gamma^{2}}{2}-2\left\langle x,y_{\theta}\right\rangle.

Let us show that its derivative gg^{\prime} vanishes at most four times on [0,2π)[0,2\pi), which will prove the result. Using the expression of yθy_{\theta}, we see that gg^{\prime} can be written as

g(θ)=acos(θ)+bsin(θ)+ccos(2θ)+dsin(2θ),\displaystyle g^{\prime}(\theta)=a\cos(\theta)+b\sin(\theta)+c\cos(2\theta)+d\sin(2\theta),

where a,b,c,da,b,c,d\in\mathbb{R} are not all zero. Denoting ω=cos(θ)\omega=\cos(\theta) and ξ=sin(θ)\xi=\sin(\theta), we have ξ2=1ω2\xi^{2}=1-\omega^{2}, cos(2θ)=cos2(θ)sin2(θ)=2ω21\cos(2\theta)=\cos^{2}(\theta)-\sin^{2}(\theta)=2\omega^{2}-1 and sin(2θ)=2cos(θ)sin(θ)=2ωξ\sin(2\theta)=2\cos(\theta)\sin(\theta)=2\omega\xi. Hence

g(θ)=aω+bξ+2cω2+2dωξ.\displaystyle g^{\prime}(\theta)=a\omega+b\xi+2c\omega^{2}+2d\omega\xi.

Now, if g(θ)=0g^{\prime}(\theta)=0, we get

aω+2cω2=(b+2dω)ξa\omega+2c\omega^{2}=-(b+2d\omega)\xi (17)

Squaring this equality yields (aω+2cω2)2=(b+2dω)2(1ω2)\left(a\omega+2c\omega^{2}\right)^{2}=\left(b+2d\omega\right)^{2}(1-\omega^{2}). This degree four equation, with variable ω\omega, admits at most four roots. To each of these ww, there exists a unique ξ=±1w2\xi=\pm\sqrt{1-w^{2}} that satisfies Equation (17). In other words, the corresponding θ[0,2π)\theta\in[0,2\pi) such that ω=cos(θ)\omega=\cos(\theta) is unique. We deduce that gg^{\prime} vanishes at most four times on [0,2π)[0,2\pi). ∎

Before studying the Čech filtration of YY, let us describe some geometric quantities associated to it. Using a symbolic computation software, we see that the curvature of YY is constant and equal to

ρ=1+8γ21+2γ2.\displaystyle\rho=\frac{\sqrt{1+8\gamma^{2}}}{1+2\gamma^{2}}.

In particular, we have ρ1\rho\geq 1 if γ1\gamma\leq 1, and ρ<1\rho<1 if γ>1\gamma>1. We also have an expression for the diameter of YY:

12diam(Y)={1 if γ12,121+γ2+14γ2 if γ12.\displaystyle\frac{1}{2}\mathrm{diam}\left(Y\right)=\left\{\begin{array}[]{ll}1&\mbox{ if }\gamma\leq\frac{1}{\sqrt{2}},\\ \frac{1}{\sqrt{2}}\sqrt{1+\gamma^{2}+\frac{1}{4\gamma^{2}}}&\mbox{ if }\gamma\geq\frac{1}{\sqrt{2}}.\end{array}\right.

It is a consequence of Lemma A.2. We now describe the reach of YY:

reach(Y)={1+2γ21+8γ2 if γ1,1 if γ1.\mathrm{reach}(Y)=\left\{\begin{array}[]{ll}\frac{1+2\gamma^{2}}{\sqrt{1+8\gamma^{2}}}&\mbox{ if }\gamma\leq 1,\\ 1&\mbox{ if }\gamma\geq 1.\end{array}\right. (18)

To prove this, we first define a bottleneck of YY as pair of distinct points (y,y)Y2(y,y^{\prime})\in Y^{2} such that the open ball (12(y+y),12yy)\mathcal{B}\left(\frac{1}{2}(y+y^{\prime}),\frac{1}{2}\left\|y-y^{\prime}\right\|\right) does not intersect YY. Its length is defined as 12yy\frac{1}{2}\left\|y-y^{\prime}\right\|. According to the results of Aamari et al. (2019, Theorem 3.4), the reach of YY is equal to

reach(Y)=min{1ρ,δ},\displaystyle\mathrm{reach}(Y)=\min\left\{\frac{1}{\rho},\delta\right\},

where 1ρ\frac{1}{\rho} is the inverse curvature of YY, and δ\delta is the minimal length of bottlenecks of YY. As we computed, 1ρ\frac{1}{\rho} is equal to 1+2γ21+8γ2\frac{1+2\gamma^{2}}{\sqrt{1+8\gamma^{2}}}. Besides, according to Lemma A.2, a bottleneck (yθ,yθ)(y_{\theta},y_{\theta^{\prime}}) has to satisfy θθ=π\theta^{\prime}-\theta=\pi or ±arccos(12γ2)\pm\arccos(-\frac{1}{2\gamma^{2}}). The smallest length is attained when θθ=π\theta^{\prime}-\theta=\pi, for which 12yθyθ=1\frac{1}{2}\left\|y_{\theta}-y_{\theta^{\prime}}\right\|=1. It is straightforward to verify that the pair (yθ,yθ)(y_{\theta},y_{\theta^{\prime}}) is indeed a bottleneck. Therefore we have δ=1\delta=1, and we deduce the expression of reach(Y)\mathrm{reach}(Y).

Last, the weak feature size of YY does not depend on γ\gamma and is equal to 1:

wfs(Y)=1.\mathrm{wfs}\left(Y\right)=1. (19)

We shall prove it by using the characterization of Boissonnat et al. (2018): wfs(Y)\mathrm{wfs}\left(Y\right) is the infimum of distances dist(x,Y)\mathrm{dist}\left(x,Y\right), where xEx\in E is a critical point of the distance function dYd_{Y}. In this context, xx is a critical point if it lies in the convex hull of its projections on YY. Remark that, if x0x\neq 0, then xx admits at most two projections on YY. This follows from Lemma A.3. As a consequence, if xx is a critical point, then there exists y,yYy,y^{\prime}\in Y such that xx lies in the middle of the segment [y,y][y,y^{\prime}], and the open ball (x,dist(x,Y))\mathcal{B}\left(x,\mathrm{dist}\left(x,Y\right)\right) does not intersect YY. Therefore yy^{\prime} is a critical point of yyyy^{\prime}\mapsto\left\|y-y^{\prime}\right\|, hence Lemma A.2 gives that yy2\left\|y-y^{\prime}\right\|\geq 2. We deduce the result.

We now describe the thickenings YtY^{t}. They present four different behaviours:

  • 0t<10\leq t<1: YtY^{t} is homotopy equivalent to a circle,

  • 1t<12diam(Y)1\leq t<\frac{1}{2}\mathrm{diam}\left(Y\right): YtY^{t} is homotopy equivalent to a circle,

  • 12diam(Y)t<1+γ22\frac{1}{2}\mathrm{diam}\left(Y\right)\leq t<\sqrt{1+\frac{\gamma^{2}}{2}}: YtY^{t} is homotopy equivalent to a 3-sphere,

  • t1+γ22t\geq\sqrt{1+\frac{\gamma^{2}}{2}}: YtY^{t} is homotopy equivalent to a point.

Recall that, in the case where γ12\gamma\leq\frac{1}{\sqrt{2}}, we have 12diam(Y)=1\frac{1}{2}\mathrm{diam}\left(Y\right)=1. Consequently, the interval [1,12diam(Y))\left[1,\frac{1}{2}\mathrm{diam}\left(Y\right)\right) is empty, and the second point does not appear in this case.

Study of the case 0t<10\leq t<1. For t[0,1)t\in[0,1), let us show that YtY^{t} deform retracts on YY. According to Equation (19), we have wfs(Y)=1\mathrm{wfs}\left(Y\right)=1. Moreover, Equation (18) gives that reach(Y)>0\mathrm{reach}(Y)>0. Using the results of Boissonnat et al. (2018), we deduce that YtY^{t} is isotopic to YY.

Study of the case 1t<12diam(Y)1\leq t<\frac{1}{2}\mathrm{diam}\left(Y\right). Denote zθ=(0,0,γ2cos(2θ),γ2sin(2θ))z_{\theta}=\left(0,0,\frac{\gamma}{\sqrt{2}}\cos(2\theta),\frac{\gamma}{\sqrt{2}}\sin(2\theta)\right), and define the circle Z={zθ,θ[0,π)}Z=\left\{z_{\theta},\theta\in[0,\pi)\right\}. It is repredented in Figure 35.

Refer to caption
Figure 35: Representation of the set YY (black) and the circle ZZ (red).

We claim that YtY^{t} deform retracts on ZZ. To prove so, we shall define a continuous application f:YtZf\colon Y^{t}\rightarrow Z such that, for every yYty\in Y^{t}, the segment [y,f(y)][y,f(y)] is included in YtY^{t}. This would lead to a deformation retraction of YtY^{t} onto ZZ, via

(s,y)[0,1]×Yt(1s)y+sf(y).\displaystyle(s,y)\in[0,1]\times Y^{t}\mapsto(1-s)y+sf(y).

Equivalently, we shall define an application Θ:Yt[0,π)\Theta\colon Y^{t}\rightarrow[0,\pi) such that the segment [y,zΘ(y)][y,z_{\Theta(y)}] is included in YtY^{t}.

Let yYty\in Y^{t}. According to Lemma A.3, yy admits at most two projection on YY. We start with the case where yy admits only one projection, namely yθy_{\theta} with θ[0,2π)\theta\in[0,2\pi). Let θ¯[0,π)\overline{\theta}\in[0,\pi) be the reduction of θ\theta modulo π\pi, and consider the point zθ¯z_{\overline{\theta}} of ZZ. A computation shows that the distance yθzθ¯\left\|y_{\theta}-z_{\overline{\theta}}\right\| is equal to 1. Besides, since yYty\in Y^{t}, the distance yθy\left\|y_{\theta}-y\right\| is at most tt. By convexity, the segment [y,zθ¯]\left[y,z_{\overline{\theta}}\right] is included in the ball ¯(yθ,t)\overline{\mathcal{B}}\left(y_{\theta},t\right), which is a subset of YtY^{t}. We then define Θ(y)=θ¯\Theta(y)=\overline{\theta}.

Now suppose that yy admits exactly two projection yθy_{\theta} and yθy_{\theta^{\prime}}. According to Lemma A.2, these angles must satisfy θθ=π\theta^{\prime}-\theta=\pi. Indeed, the case yθyθ=21+γ2+14γ2\left\|y_{\theta}-y_{\theta^{\prime}}\right\|=\sqrt{2}\sqrt{1+\gamma^{2}+\frac{1}{4\gamma^{2}}} does not occur since we chose t<12diam(Y)=221+γ2+14γ2t<\frac{1}{2}\mathrm{diam}\left(Y\right)=\frac{\sqrt{2}}{2}\sqrt{1+\gamma^{2}+\frac{1}{4\gamma^{2}}}. The angles θ\theta and θ\theta^{\prime} correspond to the same reduction modulo π\pi, denoted θ¯\overline{\theta}, and we also define Θ(y)=θ¯\Theta(y)=\overline{\theta}.

Study of the case t[12diam(X),1+γ22)t\in\left[\frac{1}{2}\mathrm{diam}\left(X\right),\sqrt{1+\frac{\gamma^{2}}{2}}\right). Let 𝕊3\operatorname{\mathbb{S}}_{3} denotes the unit sphere of EE. For every v=(v1,v2,v3,v4)𝕊3v=(v_{1},v_{2},v_{3},v_{4})\in\operatorname{\mathbb{S}}_{3}, we shall denote by v\langle v\rangle the linear subspace spanned by vv, and by v+\langle v\rangle_{+} the cone {λv,λ0}\{\lambda v,\lambda\geq 0\}. Moreover, we define the quantity

δ(v)=minyYdist(y,v+).\displaystyle\delta(v)=\min_{y\in Y}\mathrm{dist}\left(y,\langle v\rangle_{+}\right).

and the set

S={δ(v)v,v𝕊3}.\displaystyle S=\left\{\delta(v)v,v\in\operatorname{\mathbb{S}}_{3}\right\}.

The situation is depicted in Figure 36. We claim that SS is a subset of YtY^{t}, and that YtY^{t} deform retracts on it. This follows from the two following facts: for every v𝕊3v\in\operatorname{\mathbb{S}}_{3},

  1. 1.

    δ(v)\delta(v) is not greater than 12diam(Y)\frac{1}{2}\mathrm{diam}\left(Y\right),

  2. 2.

    v+Yt\langle v\rangle_{+}\cap Y^{t} consists of one connected component: an interval centered on δ(v)v\delta(v)v, that does not contain the point 0.

Suppose that these assertions are true. Then one defines a deformation retraction of YtY^{t} on SS by retracting each fiber v+Yt\langle v\rangle_{+}\cap Y^{t} linearly on the singleton {δ(v)v}\{\delta(v)v\}. We shall now prove the two items.

Refer to caption
Figure 36: Representation of the set YY (dashed), lying on a 3-sphere of radius 1+γ22\sqrt{1+\frac{\gamma^{2}}{2}}.

Item 1. Note that Item 1 can be reformulated as follows:

maxv𝕊3minyYdist(y,v+)12diam(Y).\max_{v\in\operatorname{\mathbb{S}}_{3}}\min_{y\in Y}\mathrm{dist}\left(y,\langle v\rangle_{+}\right)\leq\frac{1}{2}\mathrm{diam}\left(Y\right). (20)

Let us justify that the pairs (v,y)(v,y) that attain this maximum-minimum are the same as in

maxv𝕊3minyYyv.\max_{v\in\operatorname{\mathbb{S}}_{3}}\min_{y\in Y}\left\|y-v\right\|. (21)

From the definition of Y={yθ,θ[0,2π)}Y=\{y_{\theta},\theta\in[0,2\pi)\}, we see that minyYdist(y,v+)=minyYdist(y,v)\min_{y\in Y}\mathrm{dist}\left(y,\langle v\rangle_{+}\right)=\min_{y\in Y}\mathrm{dist}\left(y,\langle v\rangle\right). A vector v𝕊3v\in\operatorname{\mathbb{S}}_{3} being fixed, let us show that ydist(y,v)y\mapsto\mathrm{dist}\left(y,\langle v\rangle\right) is minimized when yvyy\mapsto\left\|v-y\right\| is. Let yYy\in Y. Since vv is a unit vector, the projection of yy on v\langle v\rangle can be written as y,vv\left\langle y,v\right\rangle v. Hence dist(y,v)2=y,vvy2\mathrm{dist}\left(y,\langle v\rangle\right)^{2}=\left\|\left\langle y,v\right\rangle v-y\right\|^{2}, and expanding this norm yields

dist(y,v)2=y2y,v2.\displaystyle\mathrm{dist}\left(y,\langle v\rangle\right)^{2}=\left\|y\right\|^{2}-\left\langle y,v\right\rangle^{2}.

Expanding the norm yv2\left\|y-v\right\|^{2} and using that y2=1+γ22\left\|y\right\|^{2}=1+\frac{\gamma^{2}}{2}, we get y,v=12(2+γ22yv2)\left\langle y,v\right\rangle=\frac{1}{2}\left(2+\frac{\gamma^{2}}{2}-\left\|y-v\right\|^{2}\right). We inject this relation in the preceding equation to obtain

dist(y,v)2=(γ2)4+γ2+14yv2(4+γ2yv2).\displaystyle\mathrm{dist}\left(y,\langle v\rangle\right)^{2}=-\left(\frac{\gamma}{2}\right)^{4}+\gamma^{2}+\frac{1}{4}\left\|y-v\right\|^{2}\left(4+\gamma^{2}-\left\|y-v\right\|^{2}\right).

Now we can deduce that ydist(y,v)2y\mapsto\mathrm{dist}\left(y,\langle v\rangle\right)^{2} is minimized when yyvy\mapsto\left\|y-v\right\| is minimized. Indeed, the map yv14yv2(4+γ2yv2)\left\|y-v\right\|\mapsto\frac{1}{4}\left\|y-v\right\|^{2}\left(4+\gamma^{2}-\left\|y-v\right\|^{2}\right) is increasing on [0,12(4+γ2)]\left[0,\frac{1}{2}(4+\gamma^{2})\right]. But yvy+v=12(4+γ2)\left\|y-v\right\|\leq\left\|y\right\|+\left\|v\right\|=\frac{1}{2}(4+\gamma^{2}).

We deduce that studying the left hand term of Equation (20) is equivalent to studying Equation (21). We shall denote by g:𝕊3g\colon\operatorname{\mathbb{S}}_{3}\rightarrow\mathbb{R} the map

g(v)=minyYyv.g(v)=\min_{y\in Y}\left\|y-v\right\|. (22)

Let v𝕊3v\in\operatorname{\mathbb{S}}_{3} that attains the maximum of gg, and let yy be a corresponding point that attains the minimum of yv\left\|y-v\right\|. The points vv and yy attains the quantity in Equation (20). In order to prove that dist(y,v)12diam(Y)\mathrm{dist}\left(y,\langle v\rangle\right)\leq\frac{1}{2}\mathrm{diam}\left(Y\right), let p(y)p(y) denotes the projection of yy on v\langle v\rangle. We shall show that there exists another point yYy^{\prime}\in Y such that p(y)p(y) is equal to 12(y+y)\frac{1}{2}(y+y^{\prime}) Consequently, we would have yp(y)=12yy12diam(Y)\left\|y-p(y)\right\|=\frac{1}{2}\left\|y^{\prime}-y\right\|\leq\frac{1}{2}\mathrm{diam}\left(Y\right), i.e.

dist(y,v)12diam(Y).\displaystyle\mathrm{dist}\left(y,\langle v\rangle\right)\leq\frac{1}{2}\mathrm{diam}\left(Y\right).

Remark the following fact: if w𝕊3w\in\operatorname{\mathbb{S}}_{3} is a unit vector such that p(y)y,w>0\left\langle p(y)-y,w\right\rangle>0, then for ϵ>0\epsilon>0 small enough, we have

dist(y,v+ϵw)>dist(y,v).\displaystyle\mathrm{dist}\left(y,\langle v+\epsilon w\rangle\right)>\mathrm{dist}\left(y,\langle v\rangle\right).

Equivalently, this statement reformulates as 0y,1v+ϵw(v+ϵw)<y,v0\leq\left\langle y,\frac{1}{\left\|v+\epsilon w\right\|}(v+\epsilon w)\right\rangle<\left\langle y,v\right\rangle. Let us show that

y,1v+ϵw(v+ϵw)=y,vϵκ+o(ϵ),\left\langle y,\frac{1}{\left\|v+\epsilon w\right\|}(v+\epsilon w)\right\rangle=\left\langle y,v\right\rangle-\epsilon\kappa+o(\epsilon), (23)

where κ=p(y)y,w>0\kappa=\left\langle p(y)-y,w\right\rangle>0, and where o(ϵ)o(\epsilon) is the little-o notation. Note that 1v+ϵw=1ϵv,w+o(ϵ)\frac{1}{\left\|v+\epsilon w\right\|}=1-\epsilon\left\langle v,w\right\rangle+o(\epsilon). We also have

1v+ϵw(v+ϵw)\displaystyle\frac{1}{\left\|v+\epsilon w\right\|}(v+\epsilon w) =v+ϵ(wv,wv)+o(ϵ).\displaystyle=v+\epsilon\left(w-\left\langle v,w\right\rangle v\right)+o(\epsilon).

Expanding the inner product in Equation (23) gives

y,1v+ϵw(v+ϵw)\displaystyle\left\langle y,\frac{1}{\left\|v+\epsilon w\right\|}(v+\epsilon w)\right\rangle =y,v+ϵ(y,wv,wy,v)+o(ϵ)\displaystyle=\left\langle y,v\right\rangle+\epsilon\bigg{(}\left\langle y,w\right\rangle-\left\langle v,w\right\rangle\left\langle y,v\right\rangle\bigg{)}+o(\epsilon)
=y,v+ϵyy,vv,w+o(ϵ)\displaystyle=\left\langle y,v\right\rangle+\epsilon\bigg{\langle}y-\left\langle y,v\right\rangle v,w\bigg{\rangle}+o(\epsilon)
=y,v+ϵyp(y),w+o(ϵ),\displaystyle=\left\langle y,v\right\rangle+\epsilon\left\langle y-p(y),w\right\rangle+o(\epsilon),

and we obtain the result.

Next, let us prove that yy is not the only point of YY that attains the minimum in Equation (22). Suppose that it is the case by contradiction. Let w𝕊3w\in\operatorname{\mathbb{S}}_{3} be a unit vector such that p(y)y,w>0\left\langle p(y)-y,w\right\rangle>0. For ϵ\epsilon small enough, let us prove that the vector v=1v+ϵw(v+ϵw)v^{\prime}=\frac{1}{\left\|v+\epsilon w\right\|}(v+\epsilon w) of 𝕊3\operatorname{\mathbb{S}}_{3} contradicts the maximality of vv. That is, let us prove that g(v)>g(v)g(v^{\prime})>g(v). Let yYy^{\prime}\in Y be a minimizer yv\left\|y^{\prime}-v^{\prime}\right\|. We have to show that yv>yv\left\|y^{\prime}-v^{\prime}\right\|>\left\|y-v\right\|. This would lead to g(v)>g(v)g(v^{\prime})>g(v), hence the contradiction.

Expanding the norm yields

vy2=vv+vy2\displaystyle\left\|v^{\prime}-y^{\prime}\right\|^{2}=\left\|v^{\prime}-v+v-y^{\prime}\right\|^{2} vv2+vy22vv,vy.\displaystyle\geq\left\|v^{\prime}-v\right\|^{2}+\left\|v-y^{\prime}\right\|^{2}-2\left\langle v^{\prime}-v,v-y^{\prime}\right\rangle.

Using vv20\left\|v^{\prime}-v\right\|^{2}\geq 0 and vy2vy2\left\|v-y^{\prime}\right\|^{2}\geq\left\|v-y\right\|^{2} by definition of yy, we obtain

vy2vy22vv,vy.\displaystyle\left\|v^{\prime}-y^{\prime}\right\|^{2}\geq\left\|v-y\right\|^{2}-2\left\langle v^{\prime}-v,v-y^{\prime}\right\rangle.

We have to show that vv,yy\left\langle v^{\prime}-v,y-y^{\prime}\right\rangle is positive for ϵ\epsilon small enough. By writing vy=vy+(yy)v-y^{\prime}=v-y+(y-y^{\prime}) we get

vv,vy\displaystyle\left\langle v^{\prime}-v,v-y^{\prime}\right\rangle =vv,vvv,y+vv,yy\displaystyle=\left\langle v^{\prime}-v,v\right\rangle-\left\langle v^{\prime}-v,y\right\rangle+\left\langle v^{\prime}-v,y-y^{\prime}\right\rangle

According to Equation (23), vv,y=ϵκ+o(ϵ)-\left\langle v^{\prime}-v,y\right\rangle=\epsilon\kappa+o(\epsilon). Besides, using vv=ϵ(wv,wv)+o(ϵ)v^{\prime}-v=\epsilon(w-\left\langle v,w\right\rangle v)+o(\epsilon), we get vv,v=o(ϵ)\left\langle v^{\prime}-v,v\right\rangle=o(\epsilon). Last, Cauchy-Schwarz inequality gives |vv,yy|vvyy|\left\langle v^{\prime}-v,y-y^{\prime}\right\rangle|\leq\left\|v^{\prime}-v\right\|\left\|y-y^{\prime}\right\|. Therefore, vv,yy=O(ϵ)yy\left\langle v^{\prime}-v,y-y^{\prime}\right\rangle=O(\epsilon)\left\|y-y^{\prime}\right\|, where O(ϵ)O(\epsilon) is the big-o notation. Gathering these three equalities, we obtain

vv,vy\displaystyle\left\langle v^{\prime}-v,v-y^{\prime}\right\rangle =o(ϵ)+ϵκ+O(ϵ)yy.\displaystyle=o(\epsilon)+\epsilon\kappa+O(\epsilon)\left\|y-y^{\prime}\right\|.

As we can read from this equation, if yy\left\|y-y^{\prime}\right\| goes to zero as ϵ\epsilon does, then vv,vy\left\langle v^{\prime}-v,v-y^{\prime}\right\rangle is positive for ϵ\epsilon small enough. Observe that vv^{\prime} goes to vv when ϵ\epsilon goes to 0. By assumption yy is the only minimizer in Equation (22). By continuity of gg, we deduce that yy^{\prime} goes to yy.

By contradiction, we deduce that there exists another point yy^{\prime} which attains the minimum in g(v)g(v). Note that it is the only other one, according to Lemma A.3. Let us show that p(y)p(y) lies in the middle of the segment [y,y][y,y^{\prime}]. Suppose that it is not the case. Then p(y)yp(y)-y is not equal to (p(y)y)-(p(y^{\prime})-y^{\prime}), where p(y)p(y^{\prime}) denotes the projection of yy^{\prime} on v\langle v\rangle. Consequently, the half-spaces {wE,p(y)y,w>0}\{w\in E,\left\langle p(y)-y,w\right\rangle>0\} and {wE,p(y)y,w>0}\{w\in E,\left\langle p(y^{\prime})-y^{\prime},w\right\rangle>0\} intersects. Let ww be any vector in the intersection. For ϵ>0\epsilon>0, denote v=1v+ϵw(1+ϵw)v^{\prime}=\frac{1}{\left\|v+\epsilon w\right\|}(1+\epsilon w). If ϵ\epsilon is small enough, the same reasoning as before shows that vv^{\prime} contradicts the maximality of vv. The situation is represented in Figure 37.

Refer to caption
Refer to caption
Figure 37: Left: Representation of the situation where yy and yy^{\prime} are minimizers of Equation (22). Right: Representation in the plane passing through the points yy, yy^{\prime} and p(y)p(y). The dashed area corresponds to the intersection of the half-spaces {wE,p(y)y,w>0}\{w\in E,\left\langle p(y)-y,w\right\rangle>0\} and {wE,p(y)y,w>0}\{w\in E,\left\langle p(y^{\prime})-y^{\prime},w\right\rangle>0\}.

Item 2. Let v𝕊3v\in\operatorname{\mathbb{S}}_{3}. The set v+Yt\langle v\rangle_{+}\cap Y^{t} can be described as

v+yY¯(y,t).\displaystyle\langle v\rangle_{+}\cap\bigcup_{y\in Y}\overline{\mathcal{B}}\left(y,t\right).

Let yYy\in Y such that v+¯(y,t)\langle v\rangle_{+}\cap\overline{\mathcal{B}}\left(y,t\right)\neq\emptyset. Denote by p(y)p(y) the projection of yy on v+\langle v\rangle_{+}. It is equal to y,vv\left\langle y,v\right\rangle v. Using Pythagoras’ theorem, we obtain that the set v+¯(y,t)\langle v\rangle_{+}\cap\overline{\mathcal{B}}\left(y,t\right) is equal to the interval

[p(y)±t2dist(y,v)2v].\displaystyle\left[p(y)\pm\sqrt{t^{2}-\mathrm{dist}\left(y,\langle v\rangle\right)^{2}}v\right].

Using the identity dist(y,v)2=yy,v2=1+γ22y,v2\mathrm{dist}\left(y,\langle v\rangle\right)^{2}=\left\|y\right\|-\left\langle y,v\right\rangle^{2}=1+\frac{\gamma^{2}}{2}-\left\langle y,v\right\rangle^{2}, we can write this interval as

[I1(y)v,I2(y)v],\displaystyle\big{[}I_{1}(y)\cdot v,~{}I_{2}(y)\cdot v\big{]},

where I1(y)=y,vy,v2(1+γ22t2)I_{1}(y)=\left\langle y,v\right\rangle-\sqrt{\left\langle y,v\right\rangle^{2}-(1+\frac{\gamma^{2}}{2}-t^{2})} and I2(y)=y,v+y,v2(1+γ22t2)I_{2}(y)=\left\langle y,v\right\rangle+\sqrt{\left\langle y,v\right\rangle^{2}-(1+\frac{\gamma^{2}}{2}-t^{2})}. Seen as functions of y,v\left\langle y,v\right\rangle, the map I1I_{1} is decreasing, and the map I2I_{2} is increasing (see Figure 38). Let yYy^{*}\in Y that minimizes dist(y,v)\mathrm{dist}\left(y,\langle v\rangle\right). Equivalently, yy^{*} maximizes y,v\langle y,v\rangle. It follows that the corresponding interval [I1(y)v,I2(y)v]\big{[}I_{1}(y^{*})\cdot v,~{}I_{2}(y^{*})\cdot v\big{]} contains all the others. We deduce that the set v+Yt\langle v\rangle_{+}\cap Y^{t} is equal to this interval.

Refer to caption
Refer to caption
Figure 38: Left: Representation of two intervals v+¯(y,t)\langle v\rangle_{+}\cup\overline{\mathcal{B}}\left(y,t\right) and v+¯(y,t)\langle v\rangle_{+}\cup\overline{\mathcal{B}}\left(y^{\prime},t\right). Right: Representation of the maps xx±x21x\mapsto x\pm\sqrt{x^{2}-1}.

Study of the case t1+12γ2t\geq\sqrt{1+\frac{1}{2}\gamma^{2}}. For every yYy\in Y, we have y=1+12γ2\left\|y\right\|=\sqrt{1+\frac{1}{2}\gamma^{2}}. Therefore, if t1+12γ2t\geq\sqrt{1+\frac{1}{2}\gamma^{2}}, then YtY^{t} is star shaped around the point 0, hence it deform retracts on it.

To close this subsection, let us study the Čech bundle filtration (𝕐,𝕡)(\mathbb{Y},\mathbbm{p}) of YY. According to Equation (11), its filtration maximal value is tmax(Y)=tγmax(X)=γ2t^{\mathrm{max}}\left(Y\right)=t_{\gamma}^{\mathrm{max}}\left(X\right)=\frac{\gamma}{\sqrt{2}}. Note that γ2\frac{\gamma}{\sqrt{2}} is lower than 12diam(Y)\frac{1}{2}\mathrm{diam}\left(Y\right). Consequently, only two cases are to be studied: t[0,1)t\in[0,1), and t[1,12diam(Y))t\in\left[1,\frac{1}{2}\mathrm{diam}\left(Y\right)\right).

The same argument as in Subsect. A.2 yields that for every t[0,1)t\in[0,1), the persistent Stiefel-Whitney class w1t(Y)w_{1}^{t}(Y) is equal to w10(Y)w_{1}^{0}(Y). Accordingly, for every t[1,12diam(Y))t\in\left[1,\frac{1}{2}\mathrm{diam}\left(Y\right)\right), the class w1t(Y)w_{1}^{t}(Y) is equal to w11(Y)w_{1}^{1}(Y). Let us show that w10(Y)w_{1}^{0}(Y) is zero, and that w11(Y)w_{1}^{1}(Y) is not.

First, remark that the map p0:Y𝒢1(2)p^{0}\colon Y\rightarrow\mathcal{G}_{1}(\mathbb{R}^{2}) can be seen as the normal bundle of the circle. Hence (p0):H(Y)H(𝒢1(2))(p^{0})^{*}\colon H^{*}(Y)\leftarrow H^{*}(\mathcal{G}_{1}(\mathbb{R}^{2})) is nontrivial, and we deduce that w10(Y)=0w_{1}^{0}(Y)=0. As a consequence, the persistent Stiefel-Whitney class w1t(X)w_{1}^{t}(X) is nonzero for every t<1t<1.

Next, consider p1:Y1𝒢1(2)p^{1}\colon Y^{1}\rightarrow\mathcal{G}_{1}(\mathbb{R}^{2}). Recall that Y1Y^{1} deform retracts on the circle

Z={(0,0,γ2cos(2θ),γ2sin(2θ)),θ[0,π)}.\displaystyle Z=\left\{\left(0,0,\frac{\gamma}{\sqrt{2}}\cos(2\theta),\frac{\gamma}{\sqrt{2}}\sin(2\theta)\right),\theta\in[0,\pi)\right\}.

Seen in 2×M(2)\mathbb{R}^{2}\times M(\mathbb{R}^{2}), we have

Z={((00),γ(cos(θ)2cos(θ)sin(θ)cos(θ)sin(θ)sin(θ)2)),θ[0,π)}.\displaystyle Z=\bigg{\{}\bigg{(}\begin{pmatrix}0\\ 0\end{pmatrix},\gamma\begin{pmatrix}\cos(\theta)^{2}&\cos(\theta)\sin(\theta)\\ \cos(\theta)\sin(\theta)&\sin(\theta)^{2}\end{pmatrix}\bigg{)},\theta\in[0,\pi)\bigg{\}}.

Notice that the map q:Z𝒢1(2)q\colon Z\rightarrow\mathcal{G}_{1}(\mathbb{R}^{2}), the projection on 𝒢1(2)\mathcal{G}_{1}(\mathbb{R}^{2}), is injective. Seen as a map between two circles, it has degree (modulo 2) equal to 1. We deduce that q:H(Z)H(𝒢1(2))q^{*}\colon H^{*}(Z)\leftarrow H^{*}(\mathcal{G}_{1}(\mathbb{R}^{2})) is nontrivial. Now, remark that the map qq factorizes through p1p^{1}:

Z{Z}Y1{Y^{1}}𝒢1(2){\mathcal{G}_{1}(\mathbb{R}^{2})}q\scriptstyle{q}p1\scriptstyle{p^{1}}

It induces the following diagram in cohomology:

H(Z){H^{*}(Z)}H(Y1){H^{*}(Y^{1})}H(𝒢1(2)){H^{*}(\mathcal{G}_{1}(\mathbb{R}^{2}))}\scriptstyle{\sim}q\scriptstyle{q^{*}}(p1)\scriptstyle{(p^{1})^{*}}

Since qq^{*} is nontrivial, this commutative diagram yields that the persistent Stiefel-Whitney class w11(Y)w_{1}^{1}(Y) is nonzero. As a consequence, the persistent Stiefel-Whitney class w1t(Y)w_{1}^{t}(Y) is nonzero for every t1t\geq 1.

References

  • Aamari et al. (2019) Aamari E, Kim J, Chazal F, Michel B, Rinaldo A, Wasserman L (2019) Estimating the Reach of a Manifold. Electronic journal of statistics
  • Aubrey (2011) Aubrey H (2011) Persistent cohomology operations. PhD thesis, Duke University
  • Bauer and Edelsbrunner (2017) Bauer U, Edelsbrunner H (2017) The Morse theory of Čech and Delaunay complexes. Transactions of the American Mathematical Society 369(5):3741–3762
  • Bell et al. (2019) Bell G, Lawson A, Martin J, Rudzinski J, Smyth C (2019) Weighted persistent homology. Involve, a Journal of Mathematics 12(5):823–837
  • Boissonnat et al. (2018) Boissonnat JD, Chazal F, Yvinec M (2018) Geometric and topological inference, vol 57. Cambridge University Press
  • Botnan and Crawley-Boevey (2020) Botnan M, Crawley-Boevey W (2020) Decomposition of persistence modules. Proceedings of the American Mathematical Society 148(11):4581–4596
  • Chazal et al. (2009) Chazal F, Cohen-Steiner D, Lieutier A (2009) A sampling theory for compact sets in Euclidean space. Discrete & Computational Geometry 41(3):461–479
  • Chazal et al. (2016) Chazal F, de Silva V, Glisse M, Oudot S (2016) The Structure and Stability of Persistence Modules. SpringerBriefs in Mathematics
  • Edelsbrunner (1993) Edelsbrunner H (1993) The union of balls and its dual shape. In: Proceedings of the ninth annual symposium on Computational geometry, pp 218–231
  • Govc et al. (2020) Govc D, Marzantowicz W, Pavešić P, et al. (2020) How many simplices are needed to triangulate a Grassmannian? Topological Methods in Nonlinear Analysis
  • Hatcher (2002) Hatcher A (2002) Algebraic Topology. Cambridge University Press
  • von Kühnel (1987) von Kühnel W (1987) Minimal triangulations of kummer varieties. In: Abhandlungen aus dem Mathematischen Seminar der Universität Hamburg, Springer, vol 57, pp 7–20
  • Milnor and Stasheff (2016) Milnor J, Stasheff JD (2016) Characteristic Classes.(AM-76), vol 76. Princeton university press
  • Munkres (1984) Munkres JR (1984) Elements of Algebraic Topology. Addison-Wesley
  • Perea (2018) Perea JA (2018) Multiscale projective coordinates via persistent cohomology of sparse filtrations. Discrete & Computational Geometry 59(1):175–225
  • Scoccola and Perea (2021) Scoccola L, Perea JA (2021) Approximate and discrete euclidean vector bundles. arXiv preprint arXiv:210407563