Skip to main content
Springer logoLink to Springer
. 2024 Sep 3;8(8):2197–2246. doi: 10.1007/s41468-024-00189-2

Metric geometry of spaces of persistence diagrams

Mauricio Che 1, Fernando Galaz-García 1,, Luis Guijarro 2, Ingrid Amaranta Membrillo Solis 3,4
PMCID: PMC11541355  PMID: 39524153

Abstract

Persistence diagrams are objects that play a central role in topological data analysis. In the present article, we investigate the local and global geometric properties of spaces of persistence diagrams. In order to do this, we construct a family of functors Dp, 1p, that assign, to each metric pair (XA), a pointed metric space Dp(X,A). Moreover, we show that D is sequentially continuous with respect to the Gromov–Hausdorff convergence of metric pairs, and we prove that Dp preserves several useful metric properties, such as completeness and separability, for p[1,), and geodesicity and non-negative curvature in the sense of Alexandrov, for p=2. For the latter case, we describe the metric of the space of directions at the empty diagram. We also show that the Fréchet mean set of a Borel probability measure on Dp(X,A), 1p, with finite second moment and compact support is non-empty. As an application of our geometric framework, we prove that the space of Euclidean persistence diagrams, Dp(R2n,Δn), 1n and 1p<, has infinite covering, Hausdorff, asymptotic, Assouad, and Assouad–Nagata dimensions.

Keywords: Alexandrov spaces, Asymptotic dimension, Metric pairs, Gromov–Hausdorff convergence, Persistence diagram, Fré chet mean set

Introduction

After first appearing in the pioneering work of Edelsbrunner et al. (2000), in recent years, persistent homology has become an important tool in the analysis of scientific datasets, covering a wide range of applications (Adcock et al. 2014; Buchet et al. 2018; Edelsbrunner and Harer 2008; Kovacev-Nikolic et al. 2016; Munch 2013; Cássio 2015; Zhu 2013) and playing a central role in topological data analysis.

In Zomorodian and Carlsson (2005), Carlsson and Zomorodian introduced persistence modules indexed by the natural numbers as the algebraic objects underlying persistent homology. The successful application of persistent homology in data analysis is, to a great extent, due to the notion of persistence diagrams. These were introduced by Cohen-Steiner, Edelsbrunner, and Harer as equivalent representations for persistence modules indexed by the positive real numbers (Cohen-Steiner et al. 2007). More precisely, a persistence diagram is a multiset of points (b,d)R¯02, where R¯02={(x,y)R¯×R¯:0x<y} and R¯=R{-,}. Persistence diagrams are objects that, in a certain sense, are easier to visualize than persistence modules. Moreover, the set of persistence diagrams supports a family of metrics, called p-Wasserstein metrics, parametrized by 1p (see Cohen-Steiner et al. 2010), with the metric corresponding to p= also known as the bottleneck distance. We will denote by Dp(R2,Δ), where Δ is the diagonal of R2, the metric space defined by the set of persistence diagrams that arise in persistent homology equipped with the p-Wasserstein metric.

Several authors have extensively studied the geometry and topology of the spaces Dp(R2,Δ). For instance, Mileyko et al. (2011) examined the completeness, separability, and compactness of subsets of the space Dp(R2,Δ) with the p-Wasserstein metric, 1p<. Turner et al. (2014) showed that D2(R2,Δ) is an (infinite dimensional) Alexandrov space with non-negative curvature. The results in Mileyko et al. (2011) imply the existence of Fréchet means for certain probability distributions on Dp(R2,Δ). For p=2, the results in Turner et al. (2014) imply the convergence of certain algorithms used to find Fréchet means of finite sets in Dp(R2,Δ). Turner (2020) studied further statistical properties, such as the median of finite sets in Dp(R2,Δ), and its relation to the mean. All these results are crucially based on the presence of the p-Wasserstein metric and, when p=2, on the Alexandrov space structure.

Our contributions

Motivated by the preceding considerations, we develop a general framework for the geometric study of generalized spaces of persistence diagrams. To the best of our knowledge, the present article is the first attempt to systematically analyze the geometric properties of such spaces. Our departure point is the existence of a family of functors Dp:MetPairMet, 1p< (resp. D:MetPairPMet), from the category MetPair of metric pairs equipped with relative Lipschitz maps into the category Met of pointed metric spaces equipped with pointed Lipschitz maps (resp. the category PMet of pointed pseudometric spaces equipped with pointed Lipschitz maps), which assign to each metric pair (XA), where X is a metric space and AX is a closed and non-empty subset, a space of persistence diagrams Dp(X,A). In particular, for (X,A)=(R2,Δ), where Δ={(x,y)R2:x=y} is the diagonal of R2, we recover the spaces Dp(R2,Δ), that arise in persistent homology, equipped with the p-Wasserstein distance. These spaces were studied in Mileyko et al. (2011), Turner et al. (2014). Bubenik and Elchesen studied such functors from an algebraic point of view in Bubenik and Elchesen (2022). Here, we disregard the algebraic structure and focus on the behavior of several basic topological, metric, and geometric properties and invariants under the functors Dp. When X=R02n={(x1,y1,,xn,yn):0xiyifori=1,,n} and A=Δn={(x1,y1,,xn,yn)R02n:xi=yifori=1,,n} for n2, the resulting spaces Dp(R02n,Δn), that we call from now on spaces of Euclidean persistence diagrams, can be considered as the parameter spaces for rectangle persistent modules. These modules arise in the context of multiparameter persistent homology and have been investigated by several authors (see, for example, Bjerkevik 2021; Botnan et al. 2022; Cochoy and Oudot 2020; Skryzalin and Carlsson 2017).

Our aim is twofold: first, to show that many basic results useful in statistical analysis on spaces of Euclidean persistence diagrams hold for the generalized persistence diagram spaces Dp(X,A), and, second, to study the intrinsic geometry of such spaces, which is of interest in its own right. As an application of our framework, we show that different notions of dimension are infinite for spaces the spaces of Euclidean diagrams Dp(R2n,Δn).

Given that each metric pair (XA) gives rise to a pointed metric space Dp(X,A), it is natural to ask whether some form of continuity holds with respect to (XA). To address this question, we introduce the Gromov–Hausdorff convergence of metric pairs, a mild generalization of the usual Gromov–Hausdorff convergence of pointed metric spaces (see Definition 3.1), and obtain our first main result.

Theorem A

The functor Dp, 1p, is sequentially continuous with respect to the Gromov–Hausdorff convergence of metric pairs if and only if p=.

One may think of Theorem A as providing formal justification for using persistence diagrams calculated by computers in applications. Since computers have finite precision, such diagrams are elements of a discrete space that approximates the ideal space of persistence diagrams in the Gromov–Hausdorff sense. Theorem A ensures that this approximation is continuous. In particular, a small perturbation of the space of parameters (XA) (which is (R2,Δ) in the Euclidean case) will result in a small perturbation in the corresponding space of persistence diagrams.

Our second main result is the invariance of several basic metric properties under Dp, 1p<, generalizing results in Mileyko et al. (2011), Turner et al. (2014) for spaces of Euclidean persistence diagrams. These properties include completeness and geodesicity (which we require of Alexandrov spaces), as well as non-negative curvature when p=2.

Theorem B

Let (X,A)MetPair and let 1p<. Then the following assertions hold:

  1. If X is complete, then Dp(X,A) is complete.

  2. If X is separable, then Dp(X,A) is separable.

  3. If X is a proper geodesic space, then Dp(X,A) is a geodesic space.

  4. If X is a proper Alexandrov space with non-negative curvature, then D2(X,A) is an Alexandrov space with non-negative curvature.

One may show that Dp(X,A) is complete if and only if X/A is complete for any p[1,] (see Che et al. (2024)). The behavior of D is substantially different to that of Dp, 1p<. Indeed, Theorem B fails for D, as shown in Che et al. (2024).

The functor Dp allows us to carry over, with some minor modifications, the Euclidean proofs of completeness and separability in Mileyko et al. (2011) to prove items (1) and (2) in Theorem B. Items (3) and  (4) generalize the corresponding Euclidean results in Turner et al. (2014), asserting that the spaces Dp(R2,Δ), 1p<, are geodesic and that D2(R2,Δ) is an Alexandrov space of non-negative curvature. Our proofs differ from those in Turner et al. (2014) and rely on a characterization of geodesics in persistence diagram spaces originally obtained by Chowdhury in Chowdhury (2019) in the Euclidean case.

Motivated by results in Turner et al. (2014) on the Alexandrov space D2(R2,Δ), we analyze the infinitesimal geometry of Alexandrov spaces arising via Theorem B  (4) at their distinguished point, the empty diagram σ. The infinitesimal structure of an Alexandrov space at a point is captured by the space of directions, which is itself a metric space and corresponds to the unit tangent sphere in the case of Riemannian manifolds.

First, we show that the space of directions Σσ at σD2(X,A) has diameter at most π/2 (Proposition 6.1). Second, we show that directions in Σσ corresponding to finite diagrams are dense in Σσ (Proposition 6.2). Finally, we use this to obtain an explicit description of the metric structure of Σσ. These results are new, even in the case of Euclidean persistence diagrams.

Theorem C

The space of persistence diagrams Σσ at σD2(X,A) has diameter at most π/2 and directions in Σσ corresponding to finite diagrams are dense in Σσ. Moreover, consider elements in Σσ given by geodesics ξσ={ξa}aσ and ξσ={ξa}aσ joining σ with σ,σD2(X,A), where ξa and ξa are geodesics in X joining points in A with a and a, respectively. Then

d2(σ,σ)d2(σ,σ)cos(ξσ,ξσ)=supϕ:ττaτd(a,A)d(a,A)cos(ξa,ξϕ(a)),

where d is the metric on X, d2 is the 2-Wasserstein metric, and ϕ ranges over all bijections between subsets τ and τ of points in σ and σ, respectively, such that ξa(0)=ξϕ(a)(0) for all aτ.

In D2(R2,Δ), persistence diagrams in a neighborhood of the empty diagram may be thought of as coming from noise. Thus, the space of directions at the empty diagram may be interpreted as directions coming from noise. Theorem C implies that the geometry at the empty diagram is singular, as directions at this point make an angle of at most π/2. In particular, the infinitesimal geometry is not Euclidean. Hence, embedding noisy sets of persistence diagrams into a Hilbert space might require large metric distortions. This might be of relevance in the vectorization of sets of persistence diagrams where noise might be present, which in turn plays a role in applying machine learning to such sets (see, for example, Bubenik 2015; Carrière and Bauer 2019; Kusano et al. 2017).

A further advantage of the metric pair framework is that it yields the existence of Fréchet mean sets for certain classes of probability measures on generalized spaces of persistence diagrams, as shown in Mileyko et al. (2011) in the Euclidean case. The elements of Fréchet mean sets (which, a priori, may be empty) are also called barycenters (see, for example, Afsari 2011) and may be interpreted as centers of mass of the given probability measure. In the case of the spaces Dp(X,A), we may interpret a finite collection of persistence diagrams as a measure μ on Dp(X,A) with finite support. An element of the corresponding Fréchet mean set then may be interpreted as an average of the diagrams determining the measure. For the spaces Dp(X,A), we have the following result.

Theorem D

Let μ be a Borel probability measure on Dp(X,A), 1p<. Then the following assertions hold:

  1. If μ has finite second moment and compact support, then the Fréchet mean set of μ is non-empty.

  2. If μ is tight and has rate of decay at infinity q>max{2,p}, then the Fréchet mean set of μ is non-empty.

The proof of this theorem follows along the lines of the proofs of the corresponding Euclidean statements in Mileyko et al. (2011), and hinges on the fact, easily shown, that the characterization of totally bounded subsets of spaces of Euclidean persistence diagrams given in Mileyko et al. (2011) also holds in the setting of metric pairs (see Proposition A.11).

As mentioned above, our constructions include, as a special case, spaces of Euclidean persistence diagrams, such as the classical space Dp(R2,Δ). Our fourth main result shows that different notions of dimension for this space are all infinite. We let Δn={(v,v)R2n:vRn}.

Theorem E

The space Dp(R2n,Δn), 1n and 1p<, has infinite covering, Hausdorff, asymptotic, Assouad, and Assouad–Nagata dimension.

It is known that every metric space of finite asymptotic dimension admits a coarse embedding into some Hilbert space (see Roe 2003, Example 11.5). The spaces Dp(R2,Δ), 2<p, do not admit such embeddings (see Bubenik and Wagner 2020, Theorem 21; Wagner 2021, Theorem 3.2). Hence, their asymptotic dimension is infinite (cf. Bubenik and Wagner 2020, Corollary 27). These observations, along with Theorem E, immediately imply the following.

Corollary F

The space Dp(R2,Δ), 1p, has infinite asymptotic dimension.

Our analysis also shows that other spaces of Euclidean persistence diagrams appearing in topological data analysis, such as Dp(R+2n,Δn) and Dp(R02n,Δn), have infinite asymptotic dimension as well (see Sect. 7 for precise definitions). One may think of these spaces as parameter spaces for persistence rectangles in multidimensional persistent homology (Bjerkevik 2021; Skryzalin and Carlsson 2017). We point out that the proof of Theorem E is based on different arguments to those in Bubenik and Wagner (2020), Wagner (2021) and provides a unified approach for all 1p<. The crucial point in our proof is the general observation that, if a metric pair (XA) contains a curve whose distance to A grows linearly, then Dp(X,A) has infinite asymptotic dimension (see Proposition 7.3). We point out that our arguments to prove Theorem E can be used to prove an analogous result for the spaces of persistence diagrams with finitely (but arbitrarily) many points equipped with the p-Wasserstein distance, 1p<, as defined, for example, in Bubenik and Elchesen (2022). Thus, all such spaces have infinite Hausdorff, covering, asymptotic, Assouad, and Assouad–Nagata dimensions (see Corollary 7.7).

Note that Theorem B (4) provides a systematic way of constructing examples of Alexandrov spaces of non-negative curvature. In particular, by Theorem E, the space D2(R2n,Δn) is an infinite-dimensional Alexandrov space. In contrast to the finite-dimensional case, where every Alexandrov space is proper, there are few results about infinite-dimensional Alexandrov spaces in the literature (see, for example, Plaut 2002, Sect. 13, and more recently, Mitsuishi 2010; Yokota 2012, 2014). Technical difficulties occur that do not arise in finite dimensions (see, for example, Halbeisen 2000), and a more thorough understanding of the infinite-dimensional case is lacking.

Related work

Different generalizations of persistence diagrams have appeared in the persistent homology literature, as well as their corresponding spaces of persistence diagrams. In Patel (2018), Patel generalizes persistent diagram invariants of persistence modules to cases where the invariants are associated to functors from a poset P to a symmetric monoidal category. In Kim and Mémoli (2021), Kim and Mémoli define the notion of rank invariant for functors with indices in an arbitrary poset, which allows defining persistence diagrams for any persistence module F over a poset regardless of whether F is interval-decomposable or not. In Divol and Lacombe (2021), Divol and Lacombe considered a persistence diagram as a discrete measure, expressing the distance between persistence diagrams as an optimal transport problem. In this context, the authors introduced Radon measures supported on the upper half plane, generalizing the notion of persistence diagrams, and studied the geometric and topological properties of spaces of Radon measures. Bubenik and Elchesen considered a functor in Bubenik and Elchesen (2022), which sends metric pairs to free commutative pointed metric monoids and studied many algebraic properties of such a functor.

In Bubenik and Hartsock (2024), which followed the first version of the present article, Bubenik and Hartsock studied topological and geometric properties of spaces of persistence diagrams and also considered the setting of pairs (XA). To address the existence of optimal matchings and geodesics, non-negative curvature in the sense of Alexandrov, and the Hausdorff and asymptotic dimension of spaces of persistence diagrams, Bubenik and Hartsock require the set AX to be distance minimizing, i.e., for all xX, there exists aA such that dist(x,A)=d(x,a). This property holds when X is proper and A is closed, which we assume in items (3) and (4) of Theorem B. The authors of Bubenik and Hartsock (2024) also show that Dp(X,A) has infinite asymptotic dimension when X is geodesic and proper, X/A is unbounded, and A is distance minimizing. The spaces of Euclidean persistence diagrams on n1 points equipped with the p-Wasserstein distance, 1p, have finite asymptotic dimension and therefore admit a coarse embedding into a Hilbert space (see Mitra and Virk 2021). On the other hand, the space of Euclidean persistence diagrams on finitely many points equipped with the bottleneck distance has infinite asymptotic dimension (Bubenik and Hartsock 2024, Corollary 27) and cannot be coarsely embedded into a Hilbert space (see Mitra and Virk 2021, Theorem 4.3). Bubenik and Hartsock have extended these results to metric pairs in Bubenik and Hartsock (2024). Carrière and Bauer have studied the Assouad dimension and bi-Lipschitz embeddings of spaces of finite persistence diagrams in Carrière and Bauer (2019). More recently, Bate et al. (2024) have shown that the space of persistence barcodes with at most m-points can be bi-Lipschitz embedded into 2. They point out that their results also hold for generalized persistence diagrams as considered in the present article whenever X=Ω¯, A=Ω, and Ω is a proper, open subset of Rn.

With respect to Fréchet means of probability measures defined on the spaces of persistence diagrams, Divol and Lacombe in Divol and Lacombe (2021) investigated the existence of such Fréchet means for probability measures defined on the space of persistence measures in (R2,Δ) equipped with the optimal partial transport, which in particular contain the spaces Dp(R2,Δ). Our results, although more particular in the hypotheses that we impose on the probability measures considered, are more general with respect to the spaces they are defined on.

Organization

Our article is organized as follows. In Sect. 2, we present the background on metric pairs, metric monoids, and Alexandrov spaces, and introduce the functor Dp. In Sect. 3, we define Gromov–Hausdorff convergence for metric pairs and prove Theorem A. The proofs of items (1) and (2) in Theorem B and of Theorem D follow, with minor modifications, along the same lines as those for the corresponding statements in the Euclidean case. For the sake of completeness, we have included a full treatment of these results in Appendix A. In Sect. 4, we analyze the geodesicity of the spaces Dp(X,A) and prove item (3) of Theorem B. In Sect. 5, we analyze the existence of lower curvature bounds for our spaces of persistence diagrams and prove item (4) of Theorem B. In Sect. 6, we make some remarks about their local structure. Finally, in Sect. 7 we specialize our constructions to the spaces of Euclidean persistence diagrams, which include the classical space of persistence diagrams, and prove Theorem E (cf. Corollary 7.5).

Preliminaries

In this section, we collect preliminary material that we will use in the rest of the article and prove some elementary results on the spaces of persistence diagrams. Our primary reference for metric geometry will be Burago et al. (2001).

Metric pairs

Let X be a set. A map d:X×X[0,) is a metric on X if d is symmetric, satisfies the triangle inequality, and is definite, i.e. d(x,y)=0 if and only if x=y. A pseudometric space is defined similarly; while keeping the other properties, and still requiring that d(x,x)=0 for all xX, we allow for points xy in X with xy and d(x,y)=0, in which case d is a pseudometric. We obtain extended metric and extended pseudometric spaces if we allow for d to take the value . Note that when d is a pseudometric, points at distance zero from each other give a partition of X, and d induces a metric in the corresponding quotient set.

Let (X,dX), (Y,dY) be two extended pseudometric spaces. A Lipschitz map f:XY with Lipschitz constant C is a map such that dY(f(x),f(x))C·dX(x,x) for all x,xX and y,yY.

Definition 2.1

Let MetPair denote the category of metric pairs, whose objects, Obj(MetPair), are pairs (XA) such that (X,dX) is a metric space and AX is closed and non-empty, and whose morphisms, Hom(MetPair), are relative Lipschitz maps, i.e. Lipschitz maps f:(X,A)(Y,B) such that f(A)B. When A is a point, we will talk about pointed metric spaces and pointed Lipschitz maps, i.e. Lipschitz maps f:(X,{x})(Y,{y}) such that f(x)=y. We will denote the category of pointed metric spaces by Met. Similarly, we define the category PMet of pointed pseudometric spaces, whose objects Obj(PMet), are pairs (D,{σ}) such that D is a pseudometric space and σ is a point in D. The morphisms of PMet are pointed Lipschitz maps.

Commutative metric monoids and spaces of persistence diagrams

Some of the definitions and results in this subsection may be found in Bubenik and Elchesen (2022), Bubenik and Elchesen (2022). For completeness, we provide full proofs of all the statements. We will denote multisets by using two curly brackets · and will usually denote persistence diagrams by Greek letters.

Let (Xd) be a metric space and fix p[1,]. We define the space (D~(X),d~p) on X as the set of countable multisets x1,x2, of elements of X equipped with the p-Wasserstein pseudometric d~p, which is given by

d~p(σ~,τ~)p=infϕ:σ~τ~xσ~d(x,ϕ(x))p 2.1

if p<, and

d~p(σ~,τ~)=infϕ:σ~τ~supxσ~d(x,ϕ(x)) 2.2

if p=, where ϕ ranges over all bijections between σ~ and τ~ in D~(X). Here, by convention, we set inf=, that is, we have d~p(σ~,τ~)= whenever σ~ and τ~ do not have the same cardinality.

The function d~p defines an extended pseudometric in D~(X), since it is clearly non-negative, symmetric, and the triangle inequality may be proved as follows: if ρ~,σ~,τ~D~(X) have the same cardinality and ϕ:ρ~σ~ and ψ:σ~τ~ are bijections, then ψϕ:ρ~τ~ is also a bijection and, if p<, then

d~p(ρ~,τ~)xρ~d(x,ψϕ(x))p1/pxρ~(d(x,ϕ(x))+d(ϕ(x),ψϕ(x)))p1/pxρ~d(x,ϕ(x))p1/p+xρ~d(ϕ(x),ψϕ(x))p1/p=xρ~d(x,ϕ(x))p1/p+yσ~d(y,ψ(y))p1/p.

Taking the infimum over bijections ϕ and ψ we get the claim. If the cardinalities of ρ~,σ~,τ~ are not the same, the inequality is trivial, since both sides or just the right-hand side would be infinite. For p= the argument is analogous and easier.

Given two multisets σ~ and τ~, we define their sum σ~+τ~ to be their disjoint union. We can make D~(X) into a commutative monoid with monoid operation given by taking sums of multisets, and with identity σ~ the empty multiset. It is easy to check that d~p is (left-)contractive, that is, d~p(σ~,τ~)d~p(ρ~+σ~,ρ~+τ~) for all σ~,τ~,ρ~D~(X).

From now on, let (X,A)MetPair. Given σ~,τ~D~(X), we write σ~Aτ~ if there exist α~,β~D~(A) such that σ~+α~=τ~+β~. It is easy to verify that A defines an equivalence relation on D~(X) such that, if α~1Aα~2 and β~1Aβ~2, then α~1+β~1Aα~2+β~2, i.e. A is a congruence relation on D~(X) (see, for example, Hungerford 1974, p. 27). We denote by D(X,A) the quotient set D~(X)/A. Given σ~D~(X), we write σ for the equivalence class of σ~ in D(X,A). Note that σ~Aτ~ if and only if σ~\A=τ~\A, that is, σ~ and τ~ share the same points with the same multiplicities outside A. The monoid operation on D~(X) induces a monoid operation on D(X,A) by defining σ+τ as the congruence class corresponding to σ~+τ~.

The function d~p on D~(X) induces a non-negative function dp:D(X,A)×D(X,A)[0,] defined by

dp(σ,τ)=infα~,β~D~(A)d~p(σ~+α~,τ~+β~). 2.3

Note that dp is also contractive, that is, dp(σ,τ)dp(ρ+σ,ρ+τ) for all σ,τ,ρD(X,A).

Definition 2.2

The space of p-persistence diagrams on the pair (XA), denoted by Dp(X,A), is the set of all σD(X,A) such that dp(σ,σ)<.

Lemma 2.3

If σ~D~(X) is a finite multiset, then σDp(X,A).

Proof

Let σ~D~(X) be a multiset of cardinality k<. Since AX is non-empty, we can pick an element aA, and so there exists a multiset ka=a,,aD~(A) of cardinality k. Therefore, there exists a bijection between the finite multisets σ~ and ka=σ~+ka, implying that dp(σ,σ)d~p(σ~,σ~+ka)<.

Lemma 2.4

The following assertions hold:

  1. If p=, then the function dp is an extended pseudometric on D(X,A) and a pseudometric on Dp(X,A).

  2. If p<, then the function dp is an extended metric on D(X,A) and a metric on Dp(X,A).

Proof

We will first show that dp, 1p, is an extended pseudometric. We will then show that, for p<, the function dp is an extended metric.

It is clear that, for all p[1,], the function dp is symmetric, non-negative, and dp(σ,σ)=0 for all σD(X,A). The triangle inequality follows from the facts that D~(X) is commutative and that d~p is contractive. More precisely, fix ρ~,σ~,τ~D~(X), and let ε>0. By the definition of dp, there exist α~,β~,γ~,δ~D~(A) such that d~p(ρ~+α~,σ~+β~)dp(ρ,σ)+ε and d~p(σ~+γ~,τ~+δ~)dp(σ,τ)+ε. Using the commutativity of D~(X), the contractivity of d~p, and the triangle inequality for d~p, we get

dp(ρ,τ)d~p(ρ~+α~+γ~,τ~+β~+δ~)d~p(ρ~+α~+γ~,σ~+β~+γ~)+d~p(σ~+β~+γ~,τ~+β~+δ~)d~p(ρ~+α~,σ~+β~)+d~p(σ~+γ~,τ~+δ~)dp(ρ,σ)+dp(σ,τ)+2ε.

Our choice of ε>0 was arbitrary, implying that dp(ρ,τ)dp(ρ,σ)+dp(σ,τ), as required. Hence, dp is an extended pseudometric on D(X,A). By the triangle inequality, dp is a pseudometric on Dp(X,A). Indeed, if σ,τDp(X,A), then dp(σ,τ)dp(σ,σ)+dp(τ,σ)<. This completes the proof of part (1).

Now, we prove part (2). Fix p< and let σ~,τ~D~(X) be multisets such that στ. It then follows that there exists a point uX\A which appears in σ~ and τ~ with different multiplicities (which includes the case when it has multiplicity 0 in one of the diagrams and positive multiplicity in the other). Without loss of generality, suppose that u appears with higher multiplicity in σ~. Now let ε1=inf{d(u,v):vτ~,vu}. Observe that ε1>0 since, otherwise, there would be a sequence of points in τ~ converging to u in X, which in turn would imply that dp(τ,σ)=. Let ε2>0 be such that d(u,a)ε2 for all aA, which exists since uX\A and X\A is open in X. We set ε=min{ε1,ε2}. Now, for any α~,β~D~(A), if ϕ:σ~+α~τ~+β~ is a bijection, then ϕ must map some copy of uσ~ to a point vτ~+β~ with d(u,v)ε, implying that d~p(σ~+α~,τ~+β~)ε. By taking the infimum over all α~,β~D~(A), it follows that dp(σ,τ)ε>0, as required. This shows that dp is an extended metric on D(X,A). The triangle inequality implies, as in part (1), that dp is a metric on Dp(X,A). This completes the proof of part (2).

For p<, the metric dp is the p-Wasserstein metric. The following example shows that, for p=, the function dp is not a metric, only a pseudometric.

Example 2.5

Let (XA) be a metric pair such that there exists a sequence {xn}nN of different points which converges to some xX\A and xxn for all nN. Then the multisets σ~=xn:nN and τ~=xn:nN{x} induce diagrams σ,τD(X,A) such that στ and d(σ,τ)=0 as can be seen considering the sequence of bijections ϕn:σ~τ~ given by

ϕn(xi)=xiifi<nxi-1ifi>nxifi=n.

Thus D(X,A) is a pseudometric space but not a metric space.

From now on, unless stated otherwise, we will only consider metric pairs (XA) where X is a metric space. Also, for the sake of simplicity, we will treat elements in Dp(X,A) as multisets, with the understanding that whenever we do so we are actually dealing with representatives of such elements in D~(X). Thus, for instance, we will consider things like xσ for σDp(X,A) or bijections ϕ:στ for σ,τDp(X,A), meaning there are representatives σ~ and τ~ and a bijection ϕ~:σ~τ~. We point out that the constructions discussed above can be carried out for extended pseudometric spaces with straightforward adjustments.

Given two metric pairs (XA) and (YB), their disjoint union is the space (XY,AB). We can form the extended pseudometric space (XY,dXY), where dXY|(X×X)=dX, dXY|(Y×Y)=dY and dXY(x,y)= for all xX and yY. The following result is an immediate consequence of the definition of the space (Dp(X,A),dp).

Proposition 2.6

If (XA) and (YB) are metric pairs, then

Dp(XY,AB)=Dp(X,A)×pDp(Y,B),

where U×pV denotes the space U×V endowed with the metric

dU×pdV((u1,v1),(u2,v2))=dU(u1,u2)p+dV(v1,v2)p1/p

if p<, and

dU×pdV((u1,v1),(u2,v2))=max{dU(u1,u2),dV(v1,v2)}

if p=.

Proof

It is clear that for any σDp(XY,AB), we can write σ=σ(X,A)+σ(Y,B) with σ(X,A)Dp(X,A) and σ(Y,B)Dp(Y,B). Therefore, given σ,τDp(XY,AB), we have

dp(σ,τ)p=dp(σ(X,A),τ(X,A))p+dp(σ(Y,B),τ(Y,B))p=dp×pdp((σ(X,A),σ(Y,B)),(τ(X,A),τ(Y,B)))p

if p<, and

dp(σ,τ)=max{dp(σ(X,A),τ(X,A)),dp(σ(Y,B),τ(Y,B))}

if p=.

Remark 2.7

Note that, if we allow (XA) and (YB) to be extended metric pairs, then the disjoint union (XY,AB) with the metric dXY defines a coproduct in the category Met¯Pair of extended metric pairs whose objects are extended metric pairs and whose morphisms are relative Lipschitz maps (cf. Definition 2.1).

Definition 2.8

Given a metric pair (XA), and a relative map f:(X,A)(Y,B) (i.e. such that f(A)B), we define a pointed map f:(Dp(X,A),σ)(Dp(Y,B),σ) as follows. Given a persistence diagram σDp(X,A), we let

f(σ)=f(x):xσ. 2.4

We now define the functor Dp, which we will study in the remaining sections.

Proposition 2.9

Consider the map Dp:(X,A)(Dp(X,A),σ).

  1. If p=, then Dp is a functor from the category MetPair of metric pairs equipped with relative Lipschitz maps to the category PMet of pointed pseudometric spaces with pointed Lipschitz maps.

  2. If p<, then Dp is a functor from the category MetPair of metric pairs equipped with relative Lipschitz maps to the category Met of pointed metric spaces with pointed Lipschitz maps.

Proof

Consider a C-Lipschitz relative map f:(X,A)(Y,B), i.e. dY(f(x),f(y))CdX(x,y) holds for all x,yX for some C>0. We will prove that the pointed map f, defined in (2.4), restricts to a C-Lipschitz map Dp(X,A)Dp(Y,B).

First, given σDp(X,A) a p-diagram, we need to prove that f(σ)Dp(Y,B). For any σ, we have

dp(f(σ),σ)p=xσdY(f(x),B)pxσdY(f(x),f(ax))pCpxσdX(x,ax)p

for any choice {ax}xσA. Since this choice is arbitrary,

dp(f(σ),σ)pCpxσdX(x,A)p=Cpdp(σ,σ)<.

Now consider two diagrams σ,σDp(X,A). Observe that, if ϕ:σσ is a bijection, then it induces a bijection fϕ:f(σ)f(σ) given by fϕ(y)=f(ϕ(x)) whenever y=f(x) for some xσ. Therefore

dp(f(σ),f(σ))pyf(σ)d(y,fϕ(y))p=xσd(f(x),f(ϕ(x)))pCpxσd(x,ϕ(x))p.

Since ϕ:σσ is an arbitrary bijection, we get that

dp(f(σ),f(σ))Cdp(σ,σ).

Thus, f:Dp(X,A)Dp(Y,B) is C-Lipschitz.

Now consider two relative Lipschitz maps f:(X,A)(Y,B) and g:(Y,B)(Z,C). Let σDp(X,A). Then

(gf)(σ)=gf(x):xσ=g(f(x):xσ)=gf(σ).

Thus, (gf)=gf.

Finally, if Id:(X,A)(X,A) is the identity map, it is clear that Id:Dp(X,A)Dp(X,A) is also the identity map. Thus, Dp defines a functor.

Remark 2.10

Note that we could have proved that Dp defines a functor on the category of metric spaces equipped with isometries or even bi-Lipschitz maps. However, Proposition 2.9 is more general.

Remark 2.11

Proposition 2.9 implies that, if (XA) is a metric pair and (g,x)g·x is an action of a group G on (XA) via relative bi-Lipschitz maps, then we get an action of G on Dp(X,A) given by

g·σ=g·a:aσ.

Observe that the Lipschitz constant of the bi-Lipschitz maps in the group action is preserved by the functor Dp. Hence, if G acts by relative isometries on (XA) (i.e., by isometries f:XX such that f(A)A) then so does the induced action on Dp(X,A).

Remark 2.12

We point out that Dp is, in fact, a functor from MetPair to CMon(Met), the category of commutative pointed metric monoids (see Bubenik and Elchesen 2022). In this case, given a map f:(X,A)(Y,B), the induced map f:Dp(X,A)Dp(Y,B) is a monoid homomorphism. Composing the functor Dp with the forgetful functor one obtains the map to MetPair. In this work we consider this last composition, since we are mainly focused on the metric properties of the spaces Dp(X,A), and leave the study of the algebraic properties of the monoids Dp(X,A) for future work.

Consider now the quotient metric space X/A, namely, the quotient space induced by the partition {{x}:xX\A}{A} endowed with the metric given by

d([x],[y])=min{d(x,y),d(x,A)+d(y,A)}

for any x,yX (cf. Munkres 2000, Ch. 2, Sect. 22 and Burago et al. 2001, Definition 3.1.12). It follows from (Bubenik and Elchesen 2022, Remark 4.14 and Lemma 4.24) that Dp(X,A) and Dp(X/A,[A]) are isometrically isomorphic. We have the following commutative diagrams of functors. For p=, graphic file with name 41468_2024_189_Figa_HTML.jpg and, for p<, graphic file with name 41468_2024_189_Figb_HTML.jpg both given by graphic file with name 41468_2024_189_Figc_HTML.jpg Observe that the map Dp(X,A)Dp(X/A,[A]) is a natural isomorphism. Therefore, diagrams (2.6) and (2.5) show that the functor Dp factors through the quotient functor Q:(X,A)(X/A,[A]) and the functor (X/A,[A])Dp(X/A,[A]) for p[1,].

Remark 2.13

Note that we also have the following commutative diagrams of functors. For p=, graphic file with name 41468_2024_189_Figd_HTML.jpg and, for p<, graphic file with name 41468_2024_189_Fige_HTML.jpg both given by graphic file with name 41468_2024_189_Figf_HTML.jpg here the categories PMet¯Pair, PMet¯, Met¯Pair and Met¯ consist of extended (pseudo)metric pairs and pointed (pseudo)metric spaces respectively.

Remark 2.14

Observe that the subspace of Dp(X,A)Dp(X/A,[A]) consisting of diagrams with finitely many points can be identified, as a set, with the infinite symmetric product of the pointed space (X/A, [A]) (see Hatcher 2002, p. 282, for the relevant definitions). These two spaces, however, might not be homeomorphic in general, as the infinite symmetric product is not metrizable unless A is open in X (see, for instance, Wofsey 2015).

Alexandrov spaces

Let X be a metric space. The length of a continuous path ξ:[a,b]X is given by

L(ξ)=supi=0n-1d(ξ(ti),ξ(ti+1)),

where the supremum is taken over all finite partitions a=t0t1tn=b of the interval [ab]. A geodesic space is a metric space X where for any x,yX there is a shortest path (or minimizing geodesic) between x,yX, i.e. a path ξ such that

d(x,y)=L(ξ). 2.5

In general, a path ξ:JX, where J is an interval, is said to be geodesic if each tJ has a neighborhood UJ such that ξ|U is a shortest path between any two of its points.

We will also consider the model spaces Mkn given by

Mκn=Sn1κ,ifκ>0,Rn,ifκ=0,Hn1-κ,ifκ<0.

Definition 2.15

A geodesic triangle pqr in X consists of three points p,q,rX and three minimizing geodesics [pq],[qr],[rp] between those points. A comparison triangle for pqr in Mk2 is a geodesic triangle ~kpqr=p~q~r~ in Mk2 such that

d(p~,q~)=d(p,q),d(q~,r~)=d(q,r),d(r~,p~)=d(r,p).

Definition 2.16

We say that X is an Alexandrov space with curvature bounded below by k if X is complete, geodesic and can be covered with open sets with the following property (cf. Fig. 1):

  • (T)
    For any geodesic triangle pqr contained in one of these open sets, any comparison triangle ~kpqr in Mk2 and any point x[qr], the corresponding point x~[q~r~] such that d(q~,x~)=d(q,x) satisfies
    d(p,x)d(p~,x~).
Fig. 1.

Fig. 1

The condition for a complete geodesic metric space X to be an Alexandrov space with curvature κ. Here, the curves [pq], [qr], [rp], [px], [p~q~], [q~r~], [r~p~], [p~x~] are geodesics, and the length of [pq] (respectively, [rp], [qx], [xr]) is equal to the length of [p~q~] (respectively, [r~p~], [q~x~], [x~r~]). Condition (T) then says that the length of [p~x~] is not greater than the length of [px]

By Toponogov’s Globalization Theorem, if X is an Alexandrov space with curvature bounded below by k, then property (T) above holds for any geodesic triangle in X (see, for example, Plaut 2002, Sect. 3.4). Well-known examples of Alexandrov spaces include complete Riemannian n-manifolds with a uniform lower sectional curvature bound, orbit spaces of such manifolds by an effective, isometric action of a compact Lie group, and, in infinite dimension, Hilbert spaces. The latter are instances of infinite-dimensional Alexandrov spaces of non-negative curvature.

The angle between two minimizing geodesics [pq], [pr] in an Alexandrov space X is defined as

qpr=limq1,r1p{q~1p~r~1:q1[pq],r1[pr]}.

Geodesics that make an angle zero determine an equivalence class called tangent direction. The set of tangent directions at a point pX is denoted by Σp. When equipped with the angle distance , the set Σp is a metric space. Note that the metric space (Σp,) may fail to be complete, as one can see by considering directions at a point in the boundary of the unit disc D in the Euclidean plane, D being an Alexandrov space of non-negative curvature. The completion of Σp, is called the space of directions of X at p, and is denoted by Σp. Note that in a complete, finite-dimensional Riemannian manifold Mn with sectional curvature uniformly bounded below, the space of directions at any point is isometric to the unit sphere in the tangent space to the manifold at the given point. For further basic results on Alexandrov geometry, we refer the reader to Burago et al. (2001), Burago et al. (1992), Plaut (2002).

We conclude this section by briefly recalling the definition of the Hausdorff dimension of a metric space (see Burago et al. 2001, Sect. 1.7, for further details). One may show that the Hausdorff dimension of an Alexandrov space is an integer or infinite (see Burago et al. 2001, Corollary 10.8.21 and Exercise 10.8.22).

Let X be a metric space and denote the diameter of a subset SX by diam(S). For any δ[0,) and any ε>0, let

Hεδ(X)=infiN(diam(Si))δ:XiNSianddiam(Si)<ε,

where {Si}iN is a countable covering of X by sets of diameter less than ε. Note that if no such covering exists, then Hεδ(X)=. The δ-dimensional Hausdorff measure of X is given by

Hδ(X)=ωδ·limε0Hεδ(X),

where ωδ>0 is a normalization constant such that, if δ is an integer n, the n-dimensional Hausdorff measure of the unit cube in n-dimensional Euclidean space Rn is 1. This is achieved by letting ωn be the Lebesgue measure of the unit ball in Rn. As its name indicates, the Hausdorff measure is a measure on the Borel σ-algebra of X. One may show that there exists 0δo such that Hδ(X)=0 for all δ>δo and Hδ(X)= for all δ<δo. We then define the Hausdorff dimension of X, denoted by dimH(X), to be δo. Thus,

dimH(A)=sup{δ:Hδ(X)>0}=sup{δ:Hδ(X)=}=inf{δ:Hδ(X)=0}=inf{δ:Hδ(X)<}.

Gromov–Hausdorff convergence and sequential continuity

In this section, we investigate the continuity of the functors Q and Dp defined in the preceding section. Since Dp, p<, takes values in Met, the category of pointed metric spaces, while D takes values in PMet, the category of pointed pseudometric spaces, we will consider each case separately. As both Q and Dp are defined on MetPair, the category of metric pairs, we will first define a notion of Gromov–Hausdorff convergence of metric pairs (XA). We do this in such a way that when A is a point, our definition implies the usual pointed Gromov–Hausdorff convergence of pointed metric spaces (see Burago et al. 2001, Definition 8.1.1 and Herron 2016; cf. Jansen 2017, Definition 2.1 for the case of proper metric spaces; see also Definition 3.4 below). After showing that Q:MetPairMet is sequentially continuous with respect to the Gromov–Hausdorff convergence of metric pairs, and that Dp:MetPairMet, p<, is not always sequentially continuous, we will prove the sequential continuity of D:MetPairPMet with respect to the Gromov–Hausdorff convergence of metric pairs on MetPair and pointed Gromov–Hausdorff convergence of pseudometric spaces on PMet.

Definition 3.1

(Gromov–Hausdorff convergence for metric pairs) A sequence {(Xi,Ai)}iN of metric pairs converges in the Gromov–Hausdorff topology to a metric pair (XA) if there exist sequences {εi}iN and {Ri}iN of positive numbers with εi0, Ri, and εi-approximations from B¯Ri(Ai) to B¯Ri(A) for each iN, i.e. maps fi:B¯Ri(Ai)X satisfying the following three conditions:

  1. |dXi(x,y)-dX(fi(x),fi(y)|εi for any x,yB¯Ri(Ai);

  2. dH(fi(Ai),A)εi, where dH stands for the Hausdorff distance in X;

  3. B¯Ri(A)B¯εi(fi(B¯Ri(Ai))).

We will denote the Gromov–Hausdorff convergence of metric pairs by (Xi,Ai)GHPair(X,A) and the pointed Gromov–Hausdorff convergence by (Xi,xi)GH(X,x).

With Definition 3.1 in hand, we now show that the functor Q is continuous, while Dp is not necessarily continuous when p<.

Proposition 3.2

The quotient functor Q:MetPairMet, given by (X,A)(X/A,[A]), is sequentially continuous with respect to the Gromov–Hausdorff convergence of metric pairs.

Proof

We will prove that, if there exist sequences {εi}iN and {Ri}iN of positive numbers with εi0, Ri and εi-approximations from B¯Ri(Ai) to B¯Ri(A), then there exist (5εi)-approximations from B¯Ri([Ai])(Xi/Ai,[Ai]) to B¯Ri([A])(X/A,[A]). For ease of notation, we will omit the subindices in the metric which indicate the corresponding metric space.

Let fi be an εi-approximation from B¯Ri(Ai) to B¯Ri(A) in the sense of Definition 3.1. Then, for any xB¯Ri(Ai), aiAi, we have

|d(x,ai)-d(fi(x),fi(ai))|εi

which implies

|d(x,Ai)-d(fi(x),fi(Ai))|εi. 3.1

Moreover, for any aiAi and aA, we have

|d(fi(x),fi(ai))-d(fi(x),a)|d(fi(ai),a)

and, since dH(fi(Ai),A)εi, this yields

|d(fi(x),fi(Ai))-d(fi(x),A)|εi. 3.2

Combining inequalities (3.1) and (3.2), we get

|d(x,Ai)-d(fi(x),A)|2εi.

Now, for each i, define f_i:B¯Ri([Ai])X/A by

f_i([x])=[fi(x)]if[x][Ai],[A]if[x]=[Ai].

We will prove that f_ is a (5εi)-approximation from B¯Ri([Ai]) to B¯Ri([A]). Indeed, consider [x],[y]B¯Ri([Ai]). Then x,yBRi(Ai) and therefore

|d([x],[y])-d(f_i([x]),f_i([y]))|=|min{d(x,y),d(x,Ai)+d(y,Ai)}-min{d(fi(x),fi(y)),d(fi(x),A)+d(fi(y),A)}||d(x,y)-d(fi(x),fi(y))|+|d(x,Ai)-d(fi(x),A)|+|d(y,Ai)-d(fi(y),A)|εi+2εi+2εi=5εi.

If [x][Ai] and [y]=[Ai], then

|d([x],[y])-d(f_i([x]),f_i([y]))|=|d(x,Ai)-d(fi(x),A)|2εi.

A similar inequality is obtained when [y][Ai] and [x]=[Ai]. When both [x]=[Ai] and [y]=[Ai], we get

|d([x],[y])-d(f_i([x]),f_i([y]))|=0.

In any case, we see that the distortion of f_i is 5εi, which is item xm(1) in Definition 3.1.

For item (2) in Definition 3.1, we simply observe that by definition of f_i they are pointed maps.

Finally, we see that for [y]BRi([A]) we have d(y,A)Ri, so given that fi is an εi-approximation from BRi(Ai) to BRi(A) there exists xBRi(Ai) such that d(y,fi(x))εi. Therefore,

d([y],f_i([x]))d(y,fi(x))εi.

Thus [y]Bεi(f_i(BRi(Ai))). This gives item (3) in Definition 3.1.

Example 3.3

(Dp:MetPairMetwith p< is not sequentially continuous) Let Xi=[-1i,1i]R and set Ai=X=A={0}. Then Dp(X,A)={σ}. Observe that for p, the space Dp(Xi,Ai) is unbounded. Indeed, if σn is the diagram that contains a single point, 1/i, with multiplicity n, then dp(σn,σ)=np/i as n.

Now, let σiDp(Xi,Ai) be the empty diagram and suppose, for the sake of contradiction, that there exist εi-approximations fi:B¯Ri(σi)Dp(X,A) for some εi0 and Ri. Then

|dp(σ,σi)-dp(fi(σ),fi(σi))|εi

for all σB¯Ri(σi). However, we have dp(fi(σ),f(σi))=dp(σ,σ)=0, implying that

dp(σ,σi)εi 3.3

for all σB¯Ri(σi). As εi0 and Ri as i, inequality (3.3) contradicts the fact that Dp(Xi,Ai) is unbounded for each i.

Finally, we turn our attention to the functor D. Recall, from Sect. 2, that D takes values in PMet, the category of pointed pseudometric spaces. Thus, to discuss the continuity of D, we must first define a notion of Gromov–Hausdorff convergence for pointed pseudometric spaces. We define this convergence in direct analogy to pointed Gromov–Hausdorff convergence of pointed metric spaces.

Definition 3.4

(Gromov–Hausdorff convergence for pointed pseudometric spaces) A sequence {(Di,σi)}iN of pointed pseudometric spaces converges in the Gromov–Hausdorff topology to a pointed pseudometric space (D,σ) if there exist sequences {εi}iN and {Ri}iN of positive numbers with εi0, Ri, and εi-approximations from B¯Ri(σi) to B¯Ri(σ) for each iN, i.e. maps fi:B¯Ri(σi)D satisfying the following three conditions:

  1. |dDi(x,y)-dD(fi(x),fi(y)|εi for any x,yB¯Ri(σi);

  2. d(fi(σi),σ)εi;

  3. B¯Ri(σ)B¯εi(fi(B¯Ri(σi))).

As for metric spaces, we will also denote the Gromov–Hausdorff convergence of pseudometric pairs by (Di,σi)GH(D,σ).

Given a pseudometric space D, we will denote by D_ the metric quotient D/, where cd if and only if dD(c,d)=0. We also denote sometimes by x_ the image of xD under the metric quotient. The following proposition shows that pointed Gromov–Hausdorff convergence of pseudometric spaces induces pointed Gromov–Hausdorff convergence of the corresponding metric quotients.

Proposition 3.5

Let {(Di,σi)}iN, (D,σ) be pointed pseudometric spaces and let πi:DiD_i, π:DD_ be the canonical metric identifications. Then the following assertions hold:

  1. If (Di,σi)GH(D,σ), then (D_i,σ_i)GH(D_,σ_).

  2. If (D_i,σ_i)GH(D_,σ_), then (Di,σi)GH(D,σ).

Proof

For each i, consider si:D_iDi such that πi(si(x))=x for all xD_i and s:D_D similarly. These maps exist due to the axiom of choice. Let fi be εi-approximations from B¯Ri(σi) to B¯Ri(σ). Define f_i:B¯Ri(σ_i)D_ as

f_i(x)=π(fi(si(x)))

for any xD_i. Then f_i is a (2εi)-approximation from B¯Ri(σ_i)) to B¯Ri(σ_). Indeed,

|d(x,y)-d(f_i(x),f_i(y))|=|d(si(x),si(y))-d(fi(si(x)),fi(si(y)))|εi.

Also

d(f_i(σ_i),σ_)=d(fi(si(σ_i)),σ)d(fi(si(σ_i)),fi(σi))+d(fi(σi),σ)d(si(σ_i),σi)+εi+d(fi(σi),σ)2εi.

Moreover, if d(x,π(σ))Ri then d(s(x),σ)Ri. Then there is some yDi with d(y,σi)Ri such that d(s(x),fi(y))εi. Therefore,

d(x,f_i(y_))=d(s(x),fi(si(y_)))d(s(x),fi(y))+d(fi(y),fi(si(y_)))εi+d(y,si(y_))+εi=2εi.

This proves item (1).

Conversely, given f_i an εi-approximation from B¯Ri(σ_i) to B¯Ri(σ_), we can define fi:B¯Ri(σi)D as

fi(x)=s(f_i(x_))

for any xDi. Then fi is an εi-approximation from B¯Ri(σi) to B¯Ri(σ). Indeed,

|d(x,y)-d(fi(x),fi(y))|=|d(x_,y_)-d(f_i(x_),f_i(y_))|εi.

Moreover

d(fi(σi),σ)=d(f_i(σ_i),σ_)εi.

Finally, if d(x_,σ)Ri then there exists yDi such that d(y_,σ_i)Ri and d(x_,f_i(y_))εi, or equivalently, d(x,fi(y))εi. This proves item (2).

In particular, if we consider the following commutative diagram graphic file with name 41468_2024_189_Figg_HTML.jpg where π:PMetMet is the canonical metric identification functor, then D is continous if and only if πD is continuous.

We will now show that, if (Xi,Ai)GHPair(X,A), then (D(Xi,Ai),σi)GH(D(X,A),σ).

Proposition 3.6

The functor (X,A)(D(X,A),σ) is sequentially continuous with respect to the Gromov–Hausdorff convergence of metric pairs.

Proof

Let (Xi,Ai)GH(X,A), Ri, εi0, and fi be εi-approximations from B¯Ri(Ai) to B¯Ri(A). We can define a map (fi):B¯Ri(σi)D(X,A) as

(fi)(σ)=fi(x):xσ\Ai.

We will prove that (fi) is a (3εi)-approximation from B¯Ri(σi) to B¯Ri(σ).

Let σ,σD(Xi,Ai). We now show that, for any bijection ϕ:σσ, there exists a bijection ϕ:(fi)(σ)(fi)(σ) such that

supxσdXi(x,ϕ(x))-supy(fi)(σ)dX(y,ϕ(y))3εi, 3.4

and, conversely, that for any bijection ϕ:(fi)(σ)(fi)(σ), there exists a bijection ϕ:σσ such that inequality (3.4) holds.

Indeed, let ϕ:σσ be a bijection, and let xσ and xσ be such that ϕ(x)=x. We set ϕ(x^)=x^, where, given any zXi, we set z^=fi(z) if zAi, and we set z^A to be a point such that dX(fi(z),z^)εi if zAi. In the latter case, such a choice is possible by item (2) in Definition 3.1. In particular, in either case we have dX(fi(z),z^)εi. Up to changing representatives of (fi)(σ) and (fi)(σ) in D(X,A), this completely defines a bijection ϕ:(fi)(σ)(fi)(σ), and we have

dXi(x,x)-dX(x^,x^)dXi(x,x)-dX(fi(x),fi(x))+dX(fi(x),fi(x))-dX(x^,fi(x))+dX(x^,fi(x))-dX(x^,x^)εi+dX(fi(x),x^)+dX(fi(x),x^)3εi

by item (1) in Definition 3.1 and the triangle inequality. Taking the supremum over all xσ yields inequality (3.4).

Conversely, let θ:(fi)(σ)(fi)(σ) be a bijection, and let y(fi)(σ) and y(fi)(σ) be such that θ(y)=y. We define a bijection θ˘:σσ by setting θ˘(y˘)=y˘, where, given any zX (viewed as an element in the multiset (fi)(σ) or (fi)(σ)), we set z˘Xi to be such that fi(z˘)=z if z is defined as fi(x) for some xXi, and such that z˘Ai and dX(fi(z˘),z)εi otherwise. In the latter case, we must have zA and hence such a choice is possible by item (2) in Definition 3.1. Similarly as above, we can then show that dXi(y˘,y˘)-dX(y,y)3εi, and hence (3.4) holds with ϕ=θ˘ and ϕ=θ.

Therefore, for any σ,σB¯Ri(σi), we have

|d(σ,σ)-d(fi(σ),fi(σ))|=infϕsupxσ{d(x,ϕ(x))}-infθsupy(fi)(σ){d(y,θ(y))}3εi.

On the other hand, by definition, we have that

d(fi(σi),σ)=d(σ,σ)=03εi.

Finally, if d(σ,σ)Ri, then d(y,A)Ri for any yσ, and since fi is an εi-approximation from B¯Ri(Ai) to B¯Ri(A), we know that there is some xyB¯Ri(Ai) such that d(y,fi(xy))εi. Hence, the diagram σ^D(Xi,Ai) given by

σ^=xy:xσ

satisfies d(σ,(fi)(σ^))εi3εi and d(σ^,σi)Ri, so we conclude that B¯Ri(σ)B¯3εi(B¯Ri(σi)).

Thus, (fi) is a 3εi-approximation from B¯Ri(σi) to B¯Ri(σ).

Proof of Theorem A

The result follows from Proposition 3.6 and Example 3.3.

Remark 3.7

Note that we have only shown that D is sequentially continuous. To show continuity, we must first introduce topologies on MetPair, Met, and PMet compatible with the definitions of Gromov–Hausdorff convergence on each of these categories. Herron has done this for Met in Herron (2016). The arguments in Herron (2016) may be generalized to MetPair and PMet, allowing to show the continuity of D. This has been carried out in Ahumada Gómez and Che (2023).

Geodesicity

In this section, we show that the functor Dp, with p[1,), preserves the property of being a geodesic space and, in the case p=2 and assuming X is a proper geodesic space, we characterize geodesics in the space D2(X,A). This section adapts the work of Chowdhury (2019) to the context of general metric pairs.

The following two lemmas are generalizations of Chowdhury (2019, Lemmas 17 and 18) and the proofs are similar. For a general metric pair (XA) where X is assumed to be proper, points in X always have a closest point in A. Here, however, as opposed to Chowdhury (2019), such a point is not necessarily unique.

Lemma 4.1

Let (X,A)MetPair. Let σ,τDp(X,A) be diagrams, ϕk:στ be a sequence of bijections such that xσd(x,ϕk(x))pdp(σ,τ)p as k. Then the following assertions hold:

  1. If xσ, yτ\A are such that limkϕk(x)=y, then there exists k0N such that ϕk(x)=y for all kk0.

  2. If xσ\A, yA are such that limkϕk(x)=y, then d(x,y)=d(x,A).

Proof

  1. Since p[1,) and τDp(X,A), there is some ε>0 such that Bε(y)τ={y}. Since ϕk(x)Bε(y)τ for sufficiently large k, the conclusion follows.

  2. For the sake of contradiction, if d(x,y)>d(x,A), then d(x,ϕk(x))>d(x,A)+2ε and d(ϕk(x),A)<ε for sufficiently large k, where ε=(d(x,y)-d(x,A))/3. This contradicts the fact that xσd(x,ϕk(x))pdp(σ,τ)p as k.

Lemma 4.2

Let (X,A)MetPair and assume X is a proper metric space. Let σ,τDp(X,A), and let ϕk:στ be a sequence of bijections such that xσd(x,ϕk(x))pdp(σ,τ)p as k. Then there exists a subsequence {ϕkl}lN and a limiting bijection ϕ such that ϕklϕ pointwise as l and xσd(x,ϕ(x))p=dp(σ,τ)p.

Proof

Since dp(σ,τ)<, for each point xσ\A the sequence {ϕk(x)}kN consists of a bounded set of points in X and at most countably many points in A. In particular, thanks to Lemma 4.1 and the fact that X is proper, and using a diagonal argument, we can assume that for each xσ\A, the sequence {ϕk(x)}kN is eventually constant equal to some point yτ\A or it is convergent to some point yA such that d(x,y)=d(x,A). In any case, we can define ϕ:σ\Aτ as

ϕ(x)=limkϕk(x).

By mapping enough points in A to all the points in τ that were not matched with points in σ\A, we get the required bijection ϕ:στ.

Corollary 4.3

(Existence of optimal bijections) Let (X,A)MetPair and assume X is a proper space, then for any σ,σDp(X,A) there exists an optimal bijection ϕ:στ, i.e. dp(σ,τ)p=xσd(x,ϕ(x))p.

Definition 4.4

A convex combination in Dp(X,A) is a path ξ:[0,1]Dp(x,A) such that there exist an optimal bijection ϕ:ξ(0)ξ(1) and a family of geodesics {ξx}xξ(0) in X such that ξx joins x with ϕ(x) for each xξ(0) and ξ(t)=ξx(t):xξ(0) for each t[0,1]. Sometimes we also write ξ=(ϕ,{ξx}xξ(0)) to indicate ξ is the convex combination with associated optimal bijection ϕ and family of geodesics {ξx}xξ(0).

With this definition in hand, the proof of geodesicity follows along the lines of Chowdhury (2019, Corollary 19).

Proposition 4.5

Let (X,A)MetPair. If X is a proper geodesic space, then Dp(X,A) is a geodesic space.

Proof

Let σ,σDp(X,A) be diagrams, ϕ:στ be an optimal bijection as in Corollary 4.3 and let ξ=(ϕ,{ξx}xσ) be some convex combination. Then ξ is a geodesic joining σ and τ. Indeed, if we consider the bijection ϕst:ξ(s)ξ(t) given by ϕst(ξxϕ(s))=ξxϕ(t), then

dp(ξ(s),ξ(t))pxξ(s)d(x,ϕst(x)p=xσd(ξxϕ(s),ξxϕ(t))p=|s-t|pxσd(x,ϕ(x))p=|s-t|pdp(σ,τ)p.

Therefore ξ is a geodesic from σ to τ.

Non-negative curvature

In this section, we prove that the functor D2 preserves non-negative curvature in the sense of Definition 2.16 (cf. Turner et al. 2014, Theorem 2.5; Chowdhury 2019, Theorems 10 and 11). On the other hand, it is known that the functor Dp does not preserve the non-negative curvature for p2 (see Turner 2020). Also, Dp does not preserve upper curvature bounds in the sense of CAT spaces for any p (cf. Turner et al. 2014, Proposition 2.4; Turner 2020, Proposition 2.4). Whether the functor D2 preserves strictly negative lower curvature bounds remains an open question. Additionally, observe that we cannot use the usual -norm in R2 to get lower curvature bounds on any space of persistence diagrams, as the following result shows.

Proposition 5.1

The space Dp(R2,Δ) is not an Alexandrov space for any p[1,] when R2 is endowed with the metric d.

Proof

For p=, the space D(R2,Δ) is only a pseudometric space, so it cannot be an Alexandrov space. Suppose now that 1p<. Consider the points x1=(0,5), x2=(0,7) and x3=(2,6), and let σi=xi for i=1,2,3. We may check that d(xi,xj)=2d(xi,Δ) for all ij, implying that for each ij there will be a geodesic ξi,j:[0,2]Dp(R2,Δ) between σi and σj such that ξi,j(t) has only one point for all t. Such geodesics are precisely paths of the form ξi,j(t)=ηi,j(t), where ηi,j:[0,2](R2,d) is a geodesic between xi and xj. But for each ij we can pick ηi,j so that ηi,j(1)=y=(1,6). This implies that, for instance, ξ1,3(t)=ξ2,3(t) for t1 but not for t<1, implying that there is a branching of geodesics at the point y, which cannot happen in an Alexandrov space.

We will use the following lemma, which does not require any curvature assumptions, to prove this section’s main result.

Lemma 5.2

Let ξ:[0,1]D2(X,A) be a geodesic. Let ϕi:ξ(1/2)ξ(i), i=0,1, be optimal bijections. Then ϕ=ϕ1ϕ0-1:ξ(0)ξ(1) is an optimal bijection and, for all xξ(1/2), x is a midpoint between ϕ0(x) and ϕ1(x).

Proof

By the triangle inequality, it is clear that

d(ϕ0(x),ϕ1(x))22d(ϕ0(x),x)2+2d(x,ϕ1(x))2

holds for all xξ(1/2). Therefore,

d2(ξ(0),ξ(1))2zξ(0)d(z,ϕ(z))2=xξ(1/2)d(ϕ0(x),ϕ1(x))2xξ(1/2)2d(ϕ0(x),x)2+2d(x,ϕ1(x))2=2xξ(1/2)d(ϕ0(x),x)2+2xξ(1/2)d(x,ϕ1(x))2=2d2(ξ(0),ξ(1/2))2+2d2(ξ(1/2),ξ(1))2=d2(ξ(0),ξ(1))2.

Thus,

d2(ξ(0),ξ(1))2=zξ(0)d(z,ϕ(z))2=xξ(1/2)d(ϕ0(x),ϕ1(x))2

and

d(ϕ0(x),ϕ1(x))2=2d(ϕ0(x),x)2+2d(x,ϕ1(x))2

for all xξ(1/2). In particular, ϕ is an optimal bijection between ξ(0) and ξ(1), and x is a midpoint between ϕ0(x) and ϕ1(x) for all xξ(1/2).

Proposition 5.3

Let (X,A)MetPair. If X is a proper Alexandrov space with non-negative curvature, then, D2(X,A) is also an Alexandrov space with non-negative curvature.

Proof

Since X is an Alexandrov space, it is complete and geodesic. Thus, by Theorem A.1, the space D2(X,A) is complete, and, since X is assumed to be proper, Proposition 4.5 implies that D2(X,A) is geodesic. Now we must show that D2(X,A) has non-negative curvature.

Let σ1,σ2,σ3D2(X,A) be diagrams and ξ:[0,1]D2(X,A) be a geodesic from σ2 to σ3. We want to show that the inequality

d2(σ1,ξ(1/2))212d2(σ1,σ2)2+12d2(σ1,σ3)2-14d2(σ2,σ3)2

holds. This inequality characterizes non-negative curvature (see, for example, Ohta 2012, Sect. 2.1).

Let ϕi:ξ(1/2)σi, i=1,2,3, be optimal bijections, and define ϕ=ϕ3ϕ2-1:σ2σ3. From the formula for the distance in D2(X,A) we observe that the following inequalities hold:

d2(σ1,ξ(1/2))2=xξ(1/2)d(x,ϕ1(x))2;d2(σ1,σ2)2xξ(1/2)d(ϕ1(x),ϕ2(x))2;d2(σ1,σ3)2xξ(1/2)d(ϕ1(x),ϕ3(x))2.

Now, since curv(X)0, we have that

d(x,ϕ1(x))212d(ϕ1(x),ϕ2(x))2+12d(ϕ1(x),ϕ3(x))2-14d(ϕ2(x),ϕ3(x))2

for all xξ(1/2). Therefore, thanks to Lemma 5.2,

d2(σ1,ξ(1/2))2=xξ(1/2)d(x,ϕ1(x))2xξ(1/2)12d(ϕ1(x),ϕ2(x))2+12d(ϕ1(x),ϕ3(x))2-14d(ϕ2(x),ϕ3(x))212d2(σ1,σ2)2+12d2(σ1,σ3)2-14d2(σ2,σ3)2.

Lemma 5.2 implies the following corollary, which one can use to give an alternative proof of Proposition 5.3 along the lines of the proof for the Euclidean case in Turner et al. (2014).

Corollary 5.4

Let (X,A)MetPair and assume X is a proper geodesic space. Then every geodesic in D2(X,A) is a convex combination.

Proof

This argument closely follows the proofs of Theorems 10 and 11 in Chowdhury (2019). We repeat some of the constructions for the convenience of the reader.

Let ξ:[0,1]D2(X,A) be a geodesic. We first claim there exists a sequence of convex combinations ξn=(ϕn,{ξx,n}xξ(0)) such that ξ(i2-n)=ξn(i2-n) for each nN and i{0,,2n}

Indeed, given nN, we define ϕn and {ξx,n}xξ(0) as follows. For each i{1,,2n-1} consider optimal bijections ϕn,i±:ξ((2i-1)2-n)ξ((2i-1±1)2-n). By Lemma 5.2,

ϕn=ϕn,2n-1+(ϕn,2n-1-)-1ϕn,1+(ϕn,1-)-1:ξ(0)ξ(1)

is an optimal bijection. Moreover, Lemma 5.2 implies that, for each xξ((2i-1)2-n), there is some geodesic joining ϕn,i-(x) with ϕn,i+(x) which has x as its midpoint. This way, starting from some point xξ(0) and following the bijections ϕn,i±, we construct a geodesic ξx,n joining x with ϕn(x).

Now, thanks to Lemma 4.2, there is a subsequence of {ϕn}nN which pointwise converges to some optimal bijection ϕ:ξ(0)ξ(1). Moreover, we can extract a further subsequence {ϕnk}kN such that, for fixed dyadic rationals l2-j and l2-j, the sequence of bijections ξ(l2-j)ξ(l2-j) induced by {ϕnk,i}kN pointwise converge as well. By Arzelà–Ascoli theorem and a applying one more diagonal argument, we may assume that for each xξ(0) the sequence {ξx,nk}kN uniformly converges to some geodesic ξx joining x with ϕ(x). By the continuity of ξ and ξ^=(ϕ,{ξx}xξ(0)) it easily follows that ξ(t)=ξ^(t) for each t[0,1].

Remark 5.5

We note that D2(X,A) cannot in general be an Alexandrov space with curvature bounded below by κ for any κ>0. To see this, let (XA) be a metric pair, where X is proper and geodesic. For i{1,2,3}, let xiX\A and let ξi:[0,1]X be a constant speed geodesic with ξi(0)A and ξi(1)=xi of minimal length, i.e. of length d(xi,A)=minaAd(xi,a); such ξi exists since X is proper and A is closed. Suppose that

d(ξi(s),ξj(t))2d(ξi(0),ξi(s))2+d(ξj(0),ξj(t))2wheneverij. 5.1

For i=1,2,3, let σi=xiD2(X,A). It follows from (5.1) that d(xi,xj)2d(xi,A)2+d(xj,A)2 for ij, and therefore d2(σi,σj)=d(xi,A)2+d(xj,A)2.

It is then easy to see that the path ηi,j:[0,1]D2(X,A), where

ηi,j(t)=ξi(1-t),ξj(t),

is a constant speed geodesic in D2(X,A) from σi to σj. But it is then easy to verify, again using (5.1), that

d2(σk,ηi,j(t))=d(xk,A)2+d(ξi(1-t),A)2+d(ξj(t),A)2,

where k{i,j}. In particular, it follows that the geodesic triangle in D2(X,A) formed by geodesics η1,2, η2,3 and η3,1 is isometric to the geodesic triangle in R3 with vertices (d(x1,A),0,0), (0,d(x2,A),0) and (0,0,d(x3,A)). It follows that D2(X,A) cannot be κ-Alexandrov for any κ>0.

The condition (5.1) is not hard to achieve: it can be achieved whenever X is a connected Riemannian manifold of dimension 2 and AX, for instance. Indeed, in that case, if |A|3 then (5.1) is satisfied for any x1,x2,x3X\A with d(xi,ai)ε/6, where a1,a2,a3A are distinct elements and ε=min{d(ai,aj):ij}. On the other hand, if |A|2 then |A|2 since X is connected of dimension 2, and so we may pick x1,x2,x3X\A in such a way that d(x1,a)=d(x2,a)=d(x3,a)=ε<d(xi,b) for any i and any bA\{a}, where aA is a fixed element. It then follows that ξi(0)=a for each i. Since dimX2, we may do this in such a way that the angle between ξi and ξj at a is >π/2 when ij; but then, as a consequence of the Rauch comparison theorem, (5.1) will be satisfied whenever ε>0 is chosen small enough.

Remark 5.6

Let X be an Alexandrov space and let KX be a convex subset, i.e. such that any geodesic joining any two points in K remains inside K (cf. Burago et al. 2001, p. 90). It is a direct consequence of the definition that K is also an Alexandrov space with the same lower curvature bound as X. In particular, if (X,A)MetPair with X an Alexandrov space of non-negative curvature, and KX is a convex subset with AK, then D2(K,A) is an Alexandrov space of non-negative curvature.

Proof of Theorem B

The result follows from Theorem A.1, Propositions A.7, 4.5, and 5.3.

Spaces of directions: the local geometry of noise

In this section we prove some metric properties of the space of directions Σσ at the empty diagram σD2(X,A) for (X,A)MetPair with X an Alexandrov space with non-negative curvature. As mentioned in the introduction, the space of directions at the empty diagram in D2(R2,Δ) could be interpreted as controlling the local geometry of small noise perturbations.

Proposition 6.1

The space of directions Σσ has diameter at most π/2

Proof

Consider σ,σD2(X,A). We can always consider a bijection ϕ:σσ such that ϕ(a)=A for every aσ different from A and ϕ-1(a)=A for every aσ different from A. Thus, by definition of the distance function d2, we have

d2(σ,σ)2aσd(a,A)2+aσd(a,A)2=d2(σ,σ)2+d2(σ,σ)2.

Therefore,

cos~σσσ=d2(σ,σ)2+d2(σ,σ)2-d2(σ,σ)22d2(σ,σ)d2(σ,σ)0,

i.e. ~σσσπ/2. This immediately implies the result.

Proposition 6.2

Directions in Σσ corresponding to diagrams with finitely many points are dense in Σσ.

Proof

Consider an arbitrary diagram σD2(X,A) and an enumeration {ai}iN of its points. We can define a sequence of finite diagrams {σn}nN given by

σn=a1,,an.

Let ξ be a minimizing geodesic joining σ with the empty diagram σ. By Corollary 5.4, we know that ξ is a convex combination, i.e. ξ=(ϕ,{ξx}xσ) for some optimal bijection ϕ:σσ and some collection of geodesics {ξx}xσ such that ξx joins xσ with ϕ(x)A. Let ξn=(ϕ|σn,{ξx}xσn) be the restricted convex combination between σn and σ. Then the inclusion is:ξn(s)ξ(s) induces a bijection between the corresponding diagrams, which in turn implies that

d2(ξn(s),ξ(s))2xξn(s)d(x,is(x))2=xσ\σnd(ξx(s),A)2=s2xσ\σnd(x,A)2=s2(d2(σ,σ)2-d2(σn,σ)2).

Thus, using the definition of the angle between geodesics in an Alexandrov space (see, for example, Burago et al. 2001, Definition 3.6.26) and the law of cosines, we get that

1cosσnσσ=lims0s2(d2(σn,σ)2+d2(σ,σ)2)-d2(ξn(s),ξ(s))22s2d2(σn,σ)d2(σ,σ)lims0s2(d2(σn,σ)2+d2(σ,σ)2-d2(σ,σ)2+d2(σn,σ)2)2s2d2(σn,σ)d2(σ,σ)=d2(σn,σ)d2(σ,σ),

and the last quotient converges to 1. Thus, σnσσ converges to 0. This way, we can conclude that the set of directions in Σσ induced by finite diagrams can approximate any geodesic direction, and since Σσ is the metric completion of that set, the result follows.

We can calculate explicitly the angle between any two directions at Σσ determined by finite diagrams, as the following result show.

Lemma 6.3

Let σ=a1,,am and σ=a1,,an be two diagrams with finitely many points, and let ξσ,ξσ:[0,1]D2(X,A) be geodesics joining σ to σ,σ, respectively, so that ξσ(t)=ξa1(t),,ξam(t) and ξσ(t)=ξa1(t),,ξan(t) for some geodesics ξai,ξaj:[0,1]X joining ξai(0)A to ai and ξaj(0)A to aj, respectively. Then

d2(σ,σ)d2(σ,σ)cos(ξσ,ξσ)=maxϕ:ττaτd(a,A)d(ϕ(a),A)cos(ξa,ξϕ(a)),

where ϕ ranges over all bijections between subsets τ and τ of points in σ and σ, respectively, such that ξa(0)=ξϕ(a)(0) for all aτ.

Proof

For each s,t(0,1], let ϕs,t:ξσ(s)ξσ(t) be a bijection realizing the distance d2(ξσ(s),ξσ(t)). Then there exists a bijection ϕs,t between subsets τ=τs,t and τ=τs,t of points in σ and σ, respectively, such that ϕs,t(ξx(s))=ξx(t) for xτ and x=ϕs,t(x)τ and such that ϕs,t matches all the other points of ξσ(s)ξσ(t) to A. Moreover, by the construction we have

d(ξa(0),ξϕ(a)(0))-sd(ξa(0),a)-td(ξϕ(a)(0),ϕ(a))d(ξa(s),ξϕ(a)(t))s2d(a,A)2+t2d(ϕ(a),A)21/2

for all aτs,t (where ϕ=ϕs,t), implying that ξa(0)=ξϕs,t(a)(0) for all aτs,t when s and t are small enough (which we will assume from now on).

Now we can compute that

d2(ξσ(s),ξσ(t))2=s2aσ\τd(a,A)2+t2aσ\τd(a,A)2+aτd(ξa(s),ξϕ(a)(t))2,

and therefore

s2d2(σ,σ)2+t2d2(σ,σ)2-d2(ξσ(s),ξσ(t))2=aτs2d(a,A)2+t2d(ϕ(a),A)2-d(ξa(s),ξϕ(a)(t))2, 6.1

where τ=τs,t and ϕ=ϕs,t. Moreover, note that since ϕs,t minimizes d2(ξσ(s),ξσ(t)), the bijection ϕ:ττ maximizes the right hand side of (6.1). It follows that

d2(σ,σ)d2(σ,σ)cos(ξσ,ξσ)=lims,t0s2d2(σ,σ)2+t2d2(σ,σ)2-d2(ξσ(s),ξσ(t))22st=lims,t0aτs,ts2d(a,A)2+t2d(ϕs,t(a),A)2-d(ξa(s),ξϕs,t(a)(t))22st=lims,t0maxϕ:ττaτs2d(a,A)2+t2d(ϕ(a),A)2-d(ξa(s),ξϕ(a)(t))22st, 6.2

where ϕ ranges over all bijections between subsets τ and τ of points in σ and σ, respectively, such that ξa(0)=ξϕ(a)(0) for all aτ. Since σ and σ each has finitely many points, there are only finitely many such bijections ϕ, allowing one to swap the limit and the maximum on the last line of (6.2). The result follows.

Proof of Theorem C

Propositions 6.1 and 6.2 correspond to the first two assertions in Theorem C. Using Lemma 6.3 and the density of the directions in Σσ corresponding to diagrams with finitely many points yields the third assertion in the theorem.

Dimension of spaces of Euclidean persistence diagrams

In this section, we analyze some aspects of the global geometry of the spaces of Euclidean persistence diagrams. We denote such spaces by Dp(R2n,Δn), 1p< and 1nN, where we let Δn={(v,v)R2n:vRn} (and for simplicity we write Δ=Δ1) and R2n is endowed with the Euclidean metric. The investigation of the geometric properties of the spaces Dp(R2,Δ), where the metric in R2 is induced by the -norm in R2, was carried out in Mileyko et al. (2011). In Turner et al. (2014), the authors showed that D2(R2,Δ), where R2 has the Euclidean metric, is an Alexandrov space of non-negative curvature.

We will also consider the sets

R02n={(x1,,xn,y1,,yn)R2n:0xiyi,i=1,,n}

and

R+2n={(x1,,xn,y1,,yn)R2n:xiyi,i=1,,n},

which are convex subsets of the Euclidean space R2n. In particular, the space D2(R02,Δ), is the classical space of persistence diagrams which arises in persistent homology, is also an Alexandrov space of non-negative curvature (cf. Remark 5.6). The interest in studying the spaces Dp(R2n,Δn) when n2 is motivated by the fact that the subspaces Dp(R02n,Δn)Dp(R2n,Δn) can be thought of as the parameter spaces of a family of n-dimensional persistence modules, namely, persistent rectangles (cf. Bjerkevik 2021, Theorem 4.3; Skryzalin and Carlsson 2017, Lemma 1).

As an application of our geometric results, we now show that the asymptotic dimension of Dp(R2n,Δn), Dp(R+2n,Δn), and Dp(R02n,Δn) is also infinite, for any 1p<. It may be feasible to also obtain these results by extending the work of Mitra and Virk in Mitra and Virk (2021), where they consider spaces of persistence diagrams with finitely many points in R2. The asymptotic dimension, introduced by Gromov in the context of finitely generated groups (see Gromov 1991), is a large scale geometric version of the covering dimension. For an introduction to this invariant, we refer the reader to Bell (2017), Bell and Dranishnikov (2008), Piotr (2012), Roe (2003).

Definition 7.1

(cf. Piotr 2012, Definition 2.2.1) Let U={Ui}iI be a cover of a metric space X. Given R>0, the R-multiplicity of U is the smallest integer n such that, for every xX, the ball B(xR) intersects at most n elements of U. The asymptotic dimension of X, which we denote by asdimX, is the smallest non-negative integer n such that, for every R>0, there exists a uniformly bounded cover U with R-multiplicity n+1. If no such integer exists, we let asdimX=.

The following lemma should be compared with (Mitra and Virk 2021, Lemma 3.2), where the authors compute the asymptotic dimension of spaces of persistence diagrams with n points.

Lemma 7.2

The asymptotic dimension of Dp([0,),{0}), 1p<, is infinite.

Proof

Consider the subspace DpN([0,),{0})Dp([0,),{0}) consisting of diagrams with N points. As a set, DpN([0,),{0}) can be identified with the quotient [0,)N/SN, where the symmetric group SN acts by permutations of coordinates. Consider two diagrams σ=a1,,aN and σ=a1,,aN in DpN([0,),{0}), where a1aN0 and a1aN0. We then claim that

dp(σ,σ)p=i=1N|ai-ai|p. 7.1

Indeed, by regarding σ and σ as atomic measures in [0,) with the same total mass, and applying the classical theory of optimal transport in dimension one, the monotone map ϕ:aiai induces an optimal bijection between σ and σ. See for example (Santambrogio 2015, Theorem 2.9). But this implies that the metric dp on DpN([0,),{0}) agrees with the quotient metric on ([0,)N,·p)/SN, where ·p denotes the p metric. This implies that the inclusion DpN([0,),{0}) into Dp([0,),{0}) is isometric.

Finally, we claim that the asymptotic dimension of DpN([0,),{0}) is N. Indeed, ([0,)N,·p) equipped with the metric is a quotient of an action of (Z/2Z)N on (RN,·p) by isometries, and DpN([0,),{0}) is a quotient of an action of SN on ([0,)N,·p) by isometries. As (RN,·p) and ([0,)N,·p) are proper, it follows by Kasprowski (2017, Theorem 1.1) that the asymptotic dimensions of (RN,·p), ([0,)N,·p) and DpN([0,),{0}) are the same. Thus the asymptotic dimension of DpN([0,),{0}) is N, as claimed. As DpN([0,),{0}) is an isometric subspace of Dp([0,),{0}) for each N, it follows that Dp([0,),{0}) has infinite asymptotic dimension, as required.

Proposition 7.3

Let (X,A)MetPair and let C1. Suppose that there exists a C-bi-Lipschitz map f:[0,)X such that f-1(A)={0} and such that dX(f(x),A)x/C for all x[0,). Then Dp(X,A), 1p<, has infinite asymptotic dimension.

Proof

Note that f induces a map of pairs f:([0,),{0})(X,A), and therefore a map

f:Dp([0,),{0})Dp(X,A).

We will show that f is a C-bi-Lipschitz equivalence onto its image. The result will then follow from Lemma 7.2. By Proposition 2.9, the map f is C-Lipschitz. Now, let ϕ0:f(σ)f(σ) be a bijection realizing the distance dp(f(σ),f(σ)), and note that ϕ0(f(x))=f(ϕ(x)) for some bijection ϕ:σσ, where f(x)=f(x) for x>0 and f(0)=A. Given any xσ, we then have

|x-ϕ(x)|C·dX(f(x),f(ϕ(x))),

since f is C-bi-Lipschitz. Furthermore, if ϕ(x)=0, then we have

|x-ϕ(x)|=xC·dX(f(x),A),

and, if x=0, we have

|x-ϕ(x)|=ϕ(x)C·dX(f(ϕ(x)),A).

It follows that

|x-ϕ(x)|C·dX(f(x),ϕ0(f(x))

in any case, and therefore

dp(σ,σ)pxσ|x-ϕ(x)|pCpyfσdX(y,ϕ0(y))p=(C·dp(fσ,fσ))p.

Hence f is C-bi-Lipschitz, as required.

Before proving the next result, we recall the definition of covering dimension.

Definition 7.4

(cf. Munkres 2000, Chapter 8) Let U={Ui}iI be an open cover of a metric space X. The order of U is the smallest number n for which each point pX belongs to at most n elements in U. The covering dimension of X is the minimum number n (if it exists) such that any finite open cover U of X has a refinement V of order n+1.

Corollary 7.5

The spaces Dp(R2n,Δn), Dp(R+2n,Δn) and Dp(R02n,Δn), for 1p<, have infinite Hausdorff, covering and asymptotic dimensions.

Proof

For each X{R2n,R+2n,R02n}, the map f:[0,)X defined by

f(x)=1n(0,,0n,x,,xn)

is an isometric (and so 2-bi-Lipshitz) embedding such that f-1(Δn)={0} and dX(f(x),Δn)=x/2. Hence, by Proposition  7.3, Dp(X,Δn) has infinite asymptotic dimension.

To see that Dp(X,Δn) has infinite covering and Hausdorff dimensions, observe that the same argument in the end the proof of Lemma 7.2, shows that the covering and Hausdorff dimensions of Dp([0,),0) is infinite. Since Dp([0,),0)Dp(X,Δn), we conclude that Dp(X,Δn) also has infinite covering and Hausdorff dimensions.

Putting the results in this section together yields the proof of our article’s last main result. Before proceeding, recall that the Assouad dimension of a metric space X, when infinite, yields an obstruction to bi-Lipschitz embedding X into a finite-dimensional Euclidean space (see Jonathan 2021 for a detailed discussion of this dimension and related results). More precisely, if X has a bi-Lipschitz embedding into some finite-dimensional Euclidean space, then X must have finite Assouad dimension (see Jonathan 2021, Ch. 13). The Assouad–Nagata dimension, which Assouad introduced in Assouad (1982), may be thought of as a variant of the asymptotic dimension (see Lang and Schlichenmaier 2005 for basic properties of this dimension).

Proof of Theorem E

The result for the Hausdorff, covering, and asymptotic dimensions follows from Corollary 7.5. Both the Hausdorff and covering dimensions are lower bounds for the Assouad dimension (see Jonathan 2021), while the asymptotic dimension is a lower bound for the Assouad–Nagata dimension (see Lang and Schlichenmaier 2005). Therefore, these dimensions are also infinite.

Remark 7.6

Recall that the Hausdorff dimension of an Alexandrov space must be either an integer or infinite (see Sect. 2). Using this fact, we can give an alternative proof that D2(R2n,Δn), n1, has infinite Hausdorff dimension. Indeed, the space D2(R2n,Δn) is not locally compact, since one can always construct sequences of points in arbitrarily small balls around Δn with no convergent subsequence (cf. Mileyko et al. 2011, Example 16). Since an Alexandrov space of finite Hausdorff dimension must be locally compact (see Burago et al. 2001, Theorem 10.8.1), the Hausdorff dimension of D2(R2n,Δn) must be infinite.

We point out that our arguments to prove Lemma 7.2, Proposition 7.3, and Corollary 7.5 can be used to prove analogous results for the spaces of persistence diagrams with finitely (but arbitrarily) many points, DpF(X,A), as defined, for example, in Bubenik and Elchesen (2022). Thus, all such spaces also have infinite Hausdorff, covering, asymptotic Assouad, and Assouad–Nagata dimensions.

Corollary 7.7

The space DpF(R2n,Δn), 1n and 1p<, has infinite covering, Hausdorff, asymptotic, Assouad, and Assouad–Nagata dimension.

Acknowledgements

Luis Guijarro and Ingrid Amaranta Membrillo Solis would like to thank the Department of Mathematical Sciences at Durham University for its hospitality while part of this paper was written. Fernando Galaz-García would like to thank Marvin Karsunky for his help at the initial stages of this project. The authors would like to thank Samir Chowdhury, Tristan Madeleine, and Motiejus Valiunas for their helpful comments on a previous version of this article. Finally, the authors would also like to thank one of the referees for their detailed comments on the manuscript and suggestions on how to improve some of our results in Sect. 7, leding to the strengthened statement of Theorem E.

Completeness, separability, and Fréchet means

Completeness and separability

In this subsection we will show that completeness and separability are both preserved by the functor Dp:MetPairMet, 1p<, as well as the existence of Fréchet means for probability measures with compact support or with rate of decay at infinity bounded below by max{2,p}. The proofs in this section follow almost verbatim the arguments in Mileyko et al. (2011, Sect. 3.1) for the properties of the classical spaces of persistence diagrams (i.e. when X=R+2 and A=Δ in our notation). We include these arguments in our more general setting for the sake of completeness. Also, we will assume throughout this section that p<, since the results do not hold in the case p= (see Che et al. 2024).

We first prove completeness of Dp(X,A). This proof will follow by putting together a series of lemmas.

Theorem A.1

Let (X,A)MetPair. If X is complete, then Dp(X,A) is complete.

Let {σn} be a Cauchy sequence in Dp(X,A). Let the multiplicity of σ, denoted by σ, be the number of points of σ outside A and, for α>0, let uα:Dp(X,A)Dp(X,A) be the function defined by

uα(σ)=xσ:d(x,A)α.

We call uα(σ) the α-upper part of σ. We define in a similar way the α-lower part of σ by letting lα:Dp(X,A)Dp(X,A) be given by

lα(σ)=xσ:d(x,A)<α.

Compare with the definition of uα and lα in Mileyko et al. (2011, Sect. 3.1). Observe that, in general, we cannot define the persistence of points in X with respect to A as usual (i.e. the difference between the coordinates of the point). However, we can take the distance to A as the notion of persistence, which in the case of the classical space of persistence diagrams (either with the norm · or ·2 in R2) is some multiple of the distance to the diagonal Δ. This will affect the computations in this and the following two sections.

Since the α-upper part of any diagram has finite multiplicity for arbitrary α, it is reasonable to consider the convergence of the α-upper parts of the diagrams σn.

Lemma A.2

Let α>0. There exist MαZ+ and δα,0<δα<α, such that, for all δ[δα,α), there exists an Nδ>0 such that uδ(σn)=Mα whenever n>Nδ.

Proof of Lemma 4.2

For δ with 0<δ<α, let Msupδ=limsupnuδ(σn) and let Minfδ=liminfnuδ(σn). Since {σn} is a Cauchy sequence, dp(σn,σ) is bounded and this directly implies that Msupδ<. Also, if δ1>δ2, then uδ1(σn)uδ2(σn) which means that Msup1Msup2 and Minf1Minf2. Therefore, the limits Msup=limδαMsupδ and Minf=limδαMinfδ exist and, moreover, there exists a δα such that Msup=Msupδ and Minf=Minfδ whenever δαδ<α. Now suppose that Minf<Msup. Fix δ(δα,α) and let ε=δ-δα>0. Let {σnk} and {σnl} be two subsequences of {σn} such that uδ(σnk)=Msup and uδα(σnl)=Minf. Since {σn} is a Cauchy sequence, there exists C>0 such that dp(σnk,σnl)<ε for all k,l>C. By assumption, uδ(σnk)>uδα(σnl), which implies that, for any bijection ϕ:σnkσnl, there is a point xσnk such that d(x,A)δ and d(ϕ(x),A)<δα. This means that d(x,ϕ(x))>ε, leading to dp(σnk,σnl)ε, which is a contradiction. We then set Mα=Msup=Minf.

Given α>0, let σnα=uδα(σn) and σn,α=lδα(σn).

Lemma A.3

For any α>0, the sequence {σnα} is a Cauchy sequence in Dp(X,A).

Proof of Lemma 4.3

Let δα be as in Lemma A.2 and let δ(δα,α). By Lemma A.2, there is an N>0 such that, for all n>N, there is no point xσn with d(x,A)[δα,δ). Let ε>0 and let ε0=min{ε,(δ-δα)/2}. If we increase N so that dp(σm,σn)<ε0 for all m,n>N, then there exists a bijection ϕ:σmσn such that

xσmd(x,ϕ(x))p1p<ε0δ-δα2,

which implies that ϕ(σmα)=σnα. Therefore,

dp(σmα,σnα)xσmαd(x,ϕ(x))p1p<ε0ε.

The following lemma shows that the sequence {σnα} converges for arbitrary α>0.

Lemma A.4

For any α>0, there is a diagram σαDp(X,A) such that limndp(σnα,σα)=0, hence σα=Mα and uα(σα)=σα. Moreover, if α1>α2, we have σα1σα2.

Proof of Lemma 4.4

Let α>0, δ(δα,α), and let N>0 be such that σnα=uδ(σn)=Mα for all n>N. Let ε(0,δ/2) and choose a subsequence {σnkα} such that n1>N and dp(σnkα,σmα)<2-kε for mnk. Let ϕk:σnkασnk+1α be a bijection realizing the p-Wasserstein distance between σnkα and σnk+1α, which, by our choice of ε, maps points outside A to points outside A. Let x1,,xMα be points outside A in σn1α and let {xk1},,{xkα} be sequences such that x1i=xi, for i=1,,Mα, and xk+1i=ϕk(xki). By the choice of our diagram subsequence, we get that each {xki} is a Cauchy sequence and we denote the corresponding limits by x^1,,x^Mα (here we are using the fact that X is complete). Let σα be the diagram whose points outside A are exactly x^1,,x^Mα, where the multiplicity of each x^i is the number of sequences whose limit is x^i.

For ε0, we choose K>0 such that, for all k>K, we have d(xki,x^i)<Mα-1/pε0/2 and dp(σnk,σm)<ε0/2, for mnk. It follows that

dp(σmα,σα)dp(σmα,σnkα)+dp(σnkα,σα)<ε0/2+ε0/2=ε0.

Hence, σα is the unique limit of σnα and does not depend on the choice of bijections ϕk, subsequences {σnkα}, or ε.

Finally, let α1>α2. Then points xσn2 such that xσn1 have d(x,A)<δα1<α1. Repeating the above argument with α=α2,N>0 such that σn1=uδ1(σn)=Mα1, and σn2=uδ2(σn)=Mα2, for n>N, where δ1(δα1,α1),δ2(δα2,α2), and ε>0 such that ε<min{δ2/2,(δ1-δ2)/2} leads to the last statement.

Lemma A.5

Let σ=α>0σα. Then σDp(X,A) and limα0dp(σα,σ)=0.

Proof of Lemma 4.5

Let α>0 and nN (big enough) such that dp(σα,σnα)<1. Then

dp(σα,σ)dp(σα,σnα)+dp(σnα,σ)1+C

for some constant C>0, since {σn} is a Cauchy sequence. Since the right-hand side of the preceding equation is independent of α, we get dp(σ,σ)1+C.

Finally, note that

dp(σα,σ)pdp(lα(σ),σ)p=xσd(x,A)<αd(x,A)p0asα0.

The last step in the proof of the completeness of Dp(X,A) is the following lemma.

Lemma A.6

For each ε>0, there exists an α0>0 such that, for all nN and α(0,α0], we have dp(σn,α,σ)<ε and, therefore, dp(σnα,σn)<ε.

Proof of Lemma 4.6

Suppose there is an ε>0 such that, for all α>0, there exists nαN with dp(σn,α,σ)ε. Let {αi}iN be a sequence of positive values monotonically decreasing to 0. Since αi0, we have nαi and we find a subsequence {σni} such that dp(σni,αi,σ)ε. Let δ(0,ε/4) and choose kN such that dp(σnk,σni)<δ, for all ik. Now, pick jk such that dp(σnk,αi,σ)<δ for all ij. This implies that

dp(σni,αi,σnk,αj)dp(σni,αi,σ)-dp(σ,σnk,αj)ε-δ>3δ.

For ij let ϕi:σniσnk be a bijection such that xσnid(x,ϕi(x))p<2δp. Then also xσni,αid(x,ϕi(x))p<2δp.

Since δαj>0, we can pick l>j such that δαj>2αi for all il. If we now take xσni,αi such that ϕi(x)σnk,αj, we see that

d(x,ϕi(x))d(x,A)-d(ϕi(x),A)δαj-αiαid(x,A),

with il. Now let ϕ^i:σni,αiσnk,αj be a bijection such that

ϕ^i(x)=ϕi(x)ifxσni,αiandϕi(x)σnk,αjAifxσni,αiandϕi(x)σnk,αj

and also points yσnk,αjwithϕi-1(y)σni,αi are getting mapped to A. Then, for il, we have

xσni,αid(x,ϕ^i(x))p=xσni,αiϕi(x)σnk,αjd(x,ϕi(x))p+xσni,αiϕi(x)σnk,αjd(x,A)p+yσnk,αjϕi-1(y)σni,αid(y,A)pxσni,αiϕi(x)σnk,αjd(x,ϕi(x))p+xσni,αiϕi(x)σnk,αjd(x,A)p+δp<2δp+δp=3δp.

Therefore, if il, we have that dp(σni,αi,σnk,αj)p<3δp, which is a contradiction.

The triangle inequality dp(σ,σn)dp(σ,σα)+dp(σα,σnα)+dp(σnα,σn) together with the aforementioned lemmas finally gives us Theorem A.1.

Let us now prove the separability of Dp(X,A).

Proposition A.7

Let (X,A)MetPair. If X is separable, then Dp(X,A) is separable.

Proof of Proposition 4.7

Let S be a countable dense subset of X and let S^Dp(X,A) be the set of persistence diagrams with finite total multiplicity and with points in S, that is,

S^={σDp(X,A):|σ|<andxSfor allxσ}.

Let σDp(X,A). Then, for each ε>0, we can find α>0 such that dp(lα(σ),σ)<ε/2. Then, we have dp(σ,uα(σ))dp(lα(σ),σ)<ε/2. Since S|uα(σ)| is dense in X|uα(σ)|, we can find σsS^ such that dp(σs,uα(σ))<ε/2. Then, dp(σ,σs)dp(σ,uα(σ))+dp(σs,uα(σ))<ε, which implies that S^ is dense.

Note that S^=m=0S^m, where S^m={σS^:|σ|=m}. Each S^m can be embedded into Sm, thus it is countable. Hence, S^ is countable.

Remark A.8

In Bubenik and Elchesen (2022), the authors consider completions of spaces of persistence diagrams in the more general context of pairs (XA) with X an extended pseudometric space (i.e. a space in which the distance between points could also be zero or infinite). They define D¯p(X,A) as the set of countable persistence diagrams such that, up to removing a finite subdiagram, have finite p-persistence. Let X/A=(X\A)A denote the quotient set obtained by collapsing A to a point and let (X,A/0) be the extended metric space obtained by identifying points with zero distance. Proposition 6.16 in Bubenik and Elchesen (2022) (cf. 5th Theorem on page 350 of Bubenik and Hartsock 2024) asserts that (D¯p(X¯,A/0),dp) is complete if and only if (X/A,d¯1) is complete, where d¯1 is the metric on on X/A given by d1(x,y)=min(d(x,y),||(d(x,A),d(y,A))||). In our context, X is a metric space in the usual sense and, therefore, (D¯p(X¯,A/0),dp)=(Dp(X¯,A),dp). Thus, one may obtain an alternative characterization of the completeness of our Dp(X,A) via Proposition 6.16 in Bubenik and Elchesen (2022).

Fréchet means

We now consider the existence of Fréchet means for probability measures on Dp(X,A). Following the arguments in Mileyko et al. (2011, Sect. 3.2), we will establish a characterization of totally bounded sets in the space of persistence diagrams Dp(X,A), 1p< (see Proposition A.11), which is the main ingredient in Mileyko et al. (2011) to prove the existence of Fréchet mean sets for probability measures with compact support. Before carrying on, we make the following elementary observation.

Proposition A.9

If (X,A)MetPair and XA, then Dp(X,A) is not totally bounded. In particular, Dp(X,A) is not compact.

Proof

Fix xX\A and consider the sequence of diagrams {σn} such that σn=nx. Then dp(σn,σ)=n1/pd(x,A), which is not bounded.

The following definition adapts Definitions 17, 18 and 20 from Mileyko et al. (2011) to our setting.

Definition A.10

Let (X,A)MetPair and let SDp(X,A).

  1. The set S is birth-death bounded if the set {xX:xσfor someσS} is bounded.

  2. The set S is off-diagonally birth-death bounded if, for all ε>0, the set uε(S) is birth-death bounded.

  3. The set S is uniform if, for all ε>0, there exists δ>0 such that dp(lδ(σ),σ)ε for all σS.

These conditions allow us to characterize totally bounded subsets of the space of diagrams, i.e. subsets SDp(X,A) such that, for each ε>0, there exists a finite collection of open balls in Dp(X,A) of radius ε whose union contains S. The proof of Proposition A.11 is a slight modification of that of Mileyko et al. (2011, Theorem 21). Observe again that our definition for the objects uα and lα differs slightly from that in Mileyko et al. (2011) due to our different definition for the persistence of points.

Recall that a metric space X is proper if it satisfies the Heine–Borel property, i.e. if every closed and bounded subset of X is compact (equivalently, if every closed ball in X is compact). Note that every proper metric space is complete.

Proposition A.11

Let (X,A)MetPair with X a proper metric space. Then, a set SDp(X,A) is totally bounded if and only if it is bounded, off-diagonally birth-death bounded, and uniform.

Proof

First, we prove the “if” statement. Assume then that SDp(X,A) is totally bounded. Then, in particular, S is bounded. Now, let ε>0, take 0<δ<ε/2, and let Bn=B(σn,δ), for n=1,,N, be a collection of balls of radius δ which cover S. For each σn we can find a ball CnX such that xCn for xσn with d(x,A)ε, and d(x,A)<ε/2 for all xσn with xCn. Let C be a ball containing C1CN. Also, we can find α>0 such that dp(lα(σn),σ)ε/4 for n=1,,N.

Let us prove that S is off-diagonally birth-death bounded. We will proceed by contradiction. Suppose that σBn and there is an xσ such that d(x,A)ε and xCε, where CεX is the ball concentric with C and with radius equal to the radius of C plus ε. Then, for any bijection ϕ:σσn, we have d(x,ϕ(x))>ε-ε/2, which implies that dp(σ,σn)ε/2. This contradicts the assumption that σBn and implies that uε(S) is birth-death bounded since, for all σuε(S) and all xσ, we have proved that xCε.

To prove that S is uniform, we also proceed by contradiction. Suppose that σBn and dp(lα/2(σ),σ)>ε. Consider a bijection ϕ:σσn and let σb and σt be maximal subdiagrams of lα/2(σ) such that d(ϕ(x),A)<α for xσb and d(ϕ(x),A)α for xσt. If dp(σb,σ)>ε/2, then

xσbd(x,ϕ(x))p1/pdp(σb,ϕ(σb))dp(σb,σ)-dp(ϕ(σb),σ)>ε2-ε4,

where ϕ(σb) denotes the subdiagram of σn which coincides with the image of σb under ϕ. Since σb and σt do not have common points outside A and lα/2(σ) is the union of σb and σt, we have

dp(lα/2(σ),σ)p=dp(σb,σ)p+dp(σt,σ)p.

Thus, if dp(σb,σ)ε/2, then

dp(σt,σ)>εp-2-pεp1/pε/2.

Note also that if xσt, then d(x,ϕ(x))α-α/2d(x,A). Therefore,

xσtd(x,ϕ(x))p1/pxσtd(x,A)p1/p=dp(σt,σ)>ε2.

Thus, for any bijection ϕ:σσn we have

xσd(x,ϕ(x))p1/p>ε4.

Therefore, dp(σ,σn)ε/4, which contradicts our assumption that σBn. Consequently,

dp(lα/2(σ),σ)ε

for all σS, which implies that S is uniform.

We now prove the “only if” statement. Assume that SDp(X,A) is bounded, off-diagonally birth-death bounded, and uniform. Given ε>0, let δ>0 be such that dp(lδ(σ),σ)<ε/2 for all σS. Take a ball CX such that, for all σS and all xuδ(σ), we have xC. Since S is bounded, we can also find a constant MN such that |uδ(σ)|M for all σS. On the other hand, since C is a bounded subset of a proper complete space, C is also totally bounded and we can find points x1,,xNC such that, for any xC, we have d(x,xn)M-1/pε/2 for some 1nN. Let σ be the diagram consisting of points xn with 1nN, each with multiplicity M and let σ1,,σL with L=(M+1)N be all subdiagrams of σ. If σS, we can find σn and a bijection ϕ:uδ(σ)σn such that

xuδ(σ)d(x,ϕ(x))p1/p<ε2.

Let ϕ¯:σσn be the extension of ϕ to σ obtained by mapping the points in lδ(σ) to A. Then,

xσd(x,ϕ¯(x))p1/p=xuδ(σ)d(x,ϕ¯(x))p+xlδ(σ)d(x,ϕ¯(x))p1/p<21p-1εε.

Therefore, dp(σ,σn)<ε and we conclude that S is totally bounded.

We now recall the definition of Fréchet mean set of probability measures on a metric space and state it in our setting.

Definition A.12

Given a Borel probability measure μ on Dp(X,A), the quantity

Var(μ)=infσDp(X,A)Fμ(σ)=Dp(X,A)dp(σ,τ)2dμ(τ)

is the Fréchet variance of μ and the Fréchet mean set of μ, denoted by F(μ) is the set of points in Dp(X,A) that realize Var(μ), i.e.

F(μ)={σDp(X,A):Fμ(σ)=Var(μ)}.

We also recall the definitions of various concepts mentioned in Theorem D.

Definition A.13

Let μ be a Borel probability measure on Dp(X,A).

  1. We say that μ has finite second moment if
    Fμ(σ)<
    for any σDp(X,A).
  2. The support of μ is the smallest closed subset S of Dp(X,A) such that μ(S)=1.

  3. We say that μ is tight if, for any ε>0, there is a compact subset SDp(X,A) such that μ(Dp(X,A)\S)<ε.

  4. We say that μ has rate of decay at infinity q if for some (and hence for all) σ0Dp(X,A) there exist C>0 and R>0 such that, for all rR,
    μ(Dp(X,A)\Br(σ0))Cr-q.

The following lemma is essential for the proof of Theorem D(1) and is an analog of Lemma 23 in Mileyko et al. (2011). We include the proof with the necessary modifications for the reader’s convenience.

Lemma A.14

Let μ be a finite Borel measure on Dp(X,A) with finite second moment and compact support SDp(X,A), and let {σn}nNDp(X,A) be a bounded sequence which is not off-diagonally birth-death bounded or uniform. Further, let C1>1 and C2>1 be bounds on S and {σn}nN, respectively, that is, dp(σ,σ)C1 for all σS and dp(σn,σ)C2 for all nN. Then there exists δ>0 (depending only on {σn}nN), a subsequence {σnk}kN, and subdiagrams σ¯nk such that

Sdp(σ¯nk,σ)2dμ(σ)Sdp(σnk,σ)2dμ(σ)-ε0μ(S),

where

ε0=(2s/2-1)(C1+C2)2-sδs,s=max{2,p}.

With these preliminaries in hand, the proof of Theorem D now follows as in the Euclidean case (see Mileyko et al. 2011, Theorems 24 and 28). We include the proof of item (1), as it is brief, and indicate the necessary steps to prove item (2), referring to Mileyko et al. (2011) for further details.

Proof of Lemma 5.6

First, consider the case when {σn}nN is not off-diagonally birth-death bounded. Fix x0X. Then there exists 0<ε<1 such that, for any C>0 and N>0, there is n>N and xσn satisfying d(x,A)ε and d(x,x0)C. Take 0<δ<ε/2 and choose C0>0 such that for all σS we have d(x,x0)C0 for xuδ(σ). Set C3=C0+C1+C2+1. Let {σnk}kN be a subsequence of {σn}nN such that each σnk contains a point x with d(x,A)ε and d(x,x0)C3, and let σ¯nk be the subdiagram of σnk obtained by removing all such points x. Take σS and let γ:σnkσ be a bijection such that

xσnkd(x,γ(x))pdp(σnk,σ)p+δp.

Note that

dp(σnk,σ)dp(σnk,σ)+dp(σ,σ)C1+C2.

Hence, for any xσnk such that d(x,x0)C3, we have

d(γ(x),x0)d(x,x0)-d(γ(x),x)C3-((C1+C2)p+δp)1/p>C0,

since we can take δ to be sufficiently small. Thus, γ(x)lδ(σ) for xσnk with d(x,x0)C3 and it follows that, for any xσnk such that d(x,A)ε and d(x,x0)C3, we have

d(x,γ(x))d(x,A)-d(γ(x),A)ε-δ>δ.

Let γ¯:σ¯nkσ be the bijection obtained from γ by pairing points γ(x) such that d(x,A)ε and d(x,x0)C3 to the diagonal. Then

xσnkd(x,γ(x))p=xσ¯nkd(x,γ(x))p+xσnk\σ¯nkd(x,γ(x))pxσ¯nkd(x,γ(x))p+δpxσ¯nkd(x,γ¯(x))p+δp A.1

which implies, after applying the inequalities in the proof of Mileyko et al. (2011, Lemma 23), that

xσnkd(x,γ(x))p2/pxσ¯nkd(x,γ¯(x))p2/p+ε0, A.2

where

ε0=(22/s-1)(C1+C2)2-sδ2,s=max{2,p}.

Therefore, after taking infimum with respect to γ and integrating with respect to μ on both sides in inequality (A.2), we obtain

Sdp(σnk,σ)2dμ(σ)Sdp(σ¯nk,σ)2dμ(σ)+ε0μ(S).

This proves the lemma in the case where {σn}nN is not off-diagonally birth-death bounded.

Suppose now that {σn}nN is not uniform. Let ε>0 be such that, for any α>0 and N>0, there exists n>N such that dp(lα(σn),σ)ε. If necessary, decrease the δ from the previous case so that 0<δ<ε/4 and choose α0 such that dp(lα0(σ),σ)δ for all σS. Take M1 and C>δ such that, for all σS, we have |uα0(σ)|M and d(x,A)C for xσ. Define f:[0,1][0,1] as f(x)=1-(1-x)p. Note that f is a continuous, monotonically increasing function and f(0)=0, f(1)=1. Set δ0=f-1(M-1C-pδp), and α1=min{δ0α0,M-1/pδ}. Let {σnk}kN be a subsequence of {σn}nN such that dp(lα1(σnk),σ)ε, k1, and let σ¯nk=uα1(σnk). Take σS and let γ:σnkσ be a bijection such that

xσnkd(x,γ(x))pdp(σnk,σ)p+δp.

Let γ~:σ¯nkσ be the bijection obtained from γ by pairing points in γ(lα1(σnk)) to the diagonal. For convenience, let

s0=σ¯nk,s1={xσnk:d(x,A)<α1,d(γ(x),A)<α0},s2={xσnk:d(x,A)<α1,d(γ(x),A)α0}.

Note that

xs2d(x,A)pMα1pδp.

Thus,

xs1d(x,A)pε-δp.

Consequently,

dp(s1,σ)-dp(γ(s1),σ)=xs1d(x,A)p1/p-xs1d(γ(x),A)p1/pε1-δεp1/p-δβδ

for any 2β(4p-1)1/p-1. Thus,

xs1d(x,γ(x))p1/pdp(s1,γ(s1))dp(s1,σ)-dp(γ(s1),σ)βδ.

Also,

xs2d(x,γ(x))pxs2(d(γ(x),A)-α1)p=xs2d(γ(x),A)p-d(γ(x),A)pfα1d(γ(x),A)xs2d(γ(x),A)p-Cpfα1α0-δp+xs2d(γ(x),A)p.

We then have

xσnkd(x,γ(x))p=xs0d(x,γ(x))p+xs1d(x,γ(x))p+xs2d(x,γ(x))pxs0d(x,γ¯(x))p+βpδp-δp+xs2d(γ(x),A)pβpδp-δp+xs0d(x,γ¯(x))p+xs1d(γ(x),A)p+xs2d(γ(x),A)pδp+xσ¯nkd(x,γ¯(x))p.

Thus, we have arrived at inequality (A.1), and we may finish the argument as in the previous case. This finishes the proof of the lemma.

Proof of Theorem D

For the proof of item (1), let SDp(X,A) be the support of μ and let {σn}nN be a sequence Dp(X,A) such that Fμ(σn)Var(μ).

We will proceed by contradiction. Suppose that {σn}nN is not bounded and let

wn=infσSdp(σn,σ).

Then {wn}nN is not bounded either. In particular,

Fμ(σn)=Sdp(σn,σ)2dμ(σ)wn2μ(S),

which is absurd. Thus, {σn}nN is bounded.

Assume now that {σn}nN is not off-diagonally birth-death bounded or it is not uniform. Then, by Lemma A.14, there exist a subsequence {σnk}kN and subdiagrams σ¯nkσnk such that

Sdp(σ¯nk,σ)2dμ(σ)Sdp(σnk,σ)2dμ(σ)-ε0μ(S).

Taking the infimum over k, we get that

Var(μ)Var(μ)-ε0μ(S),

which is a contradiction. This finishes the proof of item (1).

To prove item (2), one first proves an analog of Mileyko et al. (2011, Lemma 27), with minor modifications necessary to adapt the Euclidean proof to the general setting of Dp(X,A). This lemma then implies the result, in a similar fashion as in the proof of Mileyko et al. (2011, Theorem 28).

Author Contributions

All authors contributed equally to the manuscript.

Funding

M.C. was funded by CONACYT Doctoral Scholarship No. 769708. F.G.G. was funded in part by research grants MTM2017-85934-C3-2-P from the Ministerio de Economía y Competitividad de España (MINECO) and PID2021-124195NB-C32 from the Ministerio de Ciencia e Innovación (MICINN). L.G. was funded in part by research grants MTM2017-85934-C3-2-P from the Ministerio de Economía y Competitividad de España (MINECO), PID2021-124195NB-C32 from the Ministerio de Ciencia e Innovación (MICINN), QUAMAP - Quasiconformal Methods in Analysis and Applications (ERC grant 834728), and by ICMAT Severo Ochoa project CEX2019-000904-S (MINECO). I.A.M.S. was funded by the Leverhulme Trust (grant RPG-2019-055).

Data availability

Not applicable.

Declarations

Conflict of interest

The authors do not have any conflict of interest.

Ethical approval

Not applicable.

Footnotes

M. Che: Supported by CONACYT Doctoral Scholarship No. 769708.

F. Galaz-García, L. Guijarro: Supported in part by research grants MTM2017–85934–C3–2–P from the Ministerio de Economía y Competitividad de España (MINECO) and PID2021-124195NB-C32 from the Ministerio de Ciencia e Innovación (MICINN).

L. Guijarro: Supported in part by QUAMAP - Quasiconformal Methods in Analysis and Applications (ERC grant 834728), and by ICMAT Severo Ochoa project CEX2019-000904-S (MINECO).

I. Membrillo: Supported by the Leverhulme Trust (grant RPG-2019-055).

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References

  1. Adcock, A., Rubin, D., Carlsson, G.: Classification of hepatic lesions using the matching metric. Comput. Vis. Image Underst. 121, 36–42 (2014) [Google Scholar]
  2. Afsari, B.: Riemannian Inline graphic center of mass: existence, uniqueness, and convexity. Proc. Am. Math. Soc. 139(2), 655–673 (2011) [Google Scholar]
  3. Ahumada Gómez, A., Che, M.: Gromov-Hausdorff convergence of metric pairs and metric tuples. Differential Geom. Appl. 94, 102135 (2024) [Google Scholar]
  4. Assouad, P.: Sur la distance de Nagata. C. R. Acad. Sci. Paris Sér. I Math 294(1), 31–34 (1982) [Google Scholar]
  5. Bate, D., Garcia Pulido, A.L.: Bi-Lipschitz embeddings of the space of unordered Inline graphic-tuples with a partial transportation metric. Math. Ann. (2024). 10.1007/s00208-024-02831-x [DOI] [PMC free article] [PubMed]
  6. Bell, G.: Asymptotic Dimension, Office Hours with a Geometric Group Theorist, pp. 219–236. Princeton Univ. Press, Princeton, NJ (2017) [Google Scholar]
  7. Bell, G., Dranishnikov, A.: Asymptotic dimension. Topol. Appl. 155(12), 1265–1296 (2008) [Google Scholar]
  8. Bjerkevik, H.B.: On the stability of interval decomposable persistence modules. Discrete Comput. Geom. 66(1), 92–121 (2021) [Google Scholar]
  9. Botnan, M.B., Lebovici, V., Oudot, S.: On rectangle-decomposable 2-parameter persistence modules. Discrete Comput. Geom. 68, 1–24 (2022) [Google Scholar]
  10. Bubenik, P.: Statistical topological data analysis using persistence landscapes. J. Mach. Learn. Res. 16, 77–102 (2015) [Google Scholar]
  11. Bubenik, P., Elchesen, A.: Universality of persistence diagrams and the bottleneck and Wasserstein distances. Comput. Geom. 105/106, 101882 (2022) [Google Scholar]
  12. Bubenik, P., Elchesen, A.: Virtual persistence diagrams, signed measures, Wasserstein distances, and Banach spaces. J. Appl. Comput. Topol. 6(4), 429–474 (2022) [Google Scholar]
  13. Bubenik, P., Hartsock, I.: Topological and metric properties of spaces of generalized persistence diagrams. J. Appl. Comput. Topol. 8, 347–399 (2024) [Google Scholar]
  14. Bubenik, P., Wagner, A.: Embeddings of persistence diagrams into Hilbert spaces. J. Appl. Comput. Topol. 4(3), 339–351 (2020) [Google Scholar]
  15. Buchet, M., Hiraoka, Y., Obayashi, I.: Persistent homology and materials informatics. Nanoinformatics 75–95 (2018)
  16. Burago, Y., Gromov, M., Perel’man, G.: A. D. Aleksandrov spaces with curvatures bounded below. Uspekhi Mat. Nauk 47(2(284)), 3–51 (1992) [Google Scholar]
  17. Burago, D., Burago, Y., Ivanov, S.: A Course in Metric Geometry. Graduate Studies in Mathematics, vol. 33. American Mathematical Society, Providence, RI (2001) [Google Scholar]
  18. Carrière, M., Bauer, U.: On the metric distortion of embedding persistence diagrams into separable hilbert spaces. In: 35th International Symposium on Computational Geometry (SoCG 2019) (Dagstuhl, Germany) (Gill Barequet and Yusu Wang, eds.), Leibniz International Proceedings in Informatics (LIPIcs), vol. 129, Schloss Dagstuhl – Leibniz-Zentrum für Informatik, pp. 21:1–21:15 (2019)
  19. Cássio, M.M., de Mello, R.F.: Persistent homology for time series and spatial data clustering. Expert Syst. Appl. 42(15–16), 6026–6038 (2015) [Google Scholar]
  20. Che, M., Galaz-García, F., Guijarro, L., Membrillo Solis, I., Valiunas, M.: Basic metric geometry of the bottleneck distance. Proc. Am. Math. Soc. 152(8), 3575–3591 (2024) [Google Scholar]
  21. Chowdhury, S.: Geodesics in Persistence Diagram Space. arXiv:1905.10820 (2019)
  22. Cochoy, J., Oudot, S.: Decomposition of exact pfd persistence bimodules. Discrete Comput. Geom. 63(2), 255–293 (2020) [Google Scholar]
  23. Cohen-Steiner, D., Edelsbrunner, H., Harer, J.: Stability of persistence diagrams. Discrete Comput. Geom. 37(1), 103–120 (2007) [Google Scholar]
  24. Cohen-Steiner, D., Edelsbrunner, H., Harer, J., Mileyko, Y.: Lipschitz functions have Inline graphic-stable persistence. Found. Comput. Math. 10(2), 127–139 (2010) [Google Scholar]
  25. Divol, V., Lacombe, T.: Understanding the topology and the geometry of the space of persistence diagrams via optimal partial transport. J. Appl. Comput. Topol. 5(1), 1–53 (2021) [Google Scholar]
  26. Edelsbrunner, H., Letscher, D., Zomorodian, A.: Topological persistence and simplification. In: 41st Annual Symposium on Foundations of Computer Science, vol. 2000, pp. 454–463. IEEE Comput. Soc. Press, Los Alamitos, CA (2000)
  27. Edelsbrunner, H., Harer, J.: Persistent Homology–A Survey, Surveys on Discrete and Computational geometry. Contemp. Math., vol. 453, pp. 257–282. Amer. Math. Soc., Providence, RI (2008) [Google Scholar]
  28. Gromov, M.: Asymptotic invariants of infinite groups, Geometric group theory, vol. 2 Sussex. London Math. Soc. Lecture Note Ser., vol. 182, pp. 1–295. Cambridge Univ. Press, Cambridge (1991)
  29. Halbeisen, S.: On tangent cones of Alexandrov spaces with curvature bounded below. Manuscr. Math. 103(2), 169–182 (2000) [Google Scholar]
  30. Hatcher, A.: Algebraic Topology. Cambridge University Press, Cambridge (2002) [Google Scholar]
  31. Herron, D.A.: Gromov-Hausdorff distance for pointed metric spaces. J. Anal. 24(1), 1–38 (2016) [Google Scholar]
  32. Hungerford, T.W.: Algebra, Graduate Texts in Mathematics, vol. 73. Springer, New York-Berlin (1980). Reprint of the 1974 original
  33. Jansen, D.: Notes on Pointed Gromov–Hausdorff Convergence. arXiv:1703.09595 (2017)
  34. Jonathan, M.: Persistent Homology–A Survey. Surveys on Discrete and Computational geometry, vol. 222. Cambridge University Press, Cambridge (2021) [Google Scholar]
  35. Kasprowski, D.: The asymptotic dimension of quotients by finite groups. Proc. Am. Math. Soc. 145(6), 2383–2389 (2017) [Google Scholar]
  36. Kim, W., Mémoli, F.: Generalized persistence diagrams for persistence modules over posets. J. Appl. Comput. Topol. 5(4), 533–581 (2021) [Google Scholar]
  37. Kovacev-Nikolic, V., Bubenik, P., Nikolić, D., Heo, G.: Using persistent homology and dynamical distances to analyze protein binding. Stat. Appl. Genet. Mol. Biol. 15(1), 19–38 (2016) [DOI] [PubMed] [Google Scholar]
  38. Kusano, G., Fukumizu, K., Hiraoka, Y.: Kernel method for persistence diagrams via kernel embedding and weight factor. J. Mach. Learn. Res. 18, 189 (2017) [Google Scholar]
  39. Lang, U., Schlichenmaier, T.: Nagata dimension, quasisymmetric embeddings, and Lipschitz extensions. Int. Math. Res. Not. 58, 3625–3655 (2005) [Google Scholar]
  40. Mileyko, Y., Mukherjee, S., Harer, J.: Probability measures on the space of persistence diagrams. Inverse Probl. 27(12), 124007 (2011) [Google Scholar]
  41. Mitra, A., Virk, Ž: The space of persistence diagrams on Inline graphic points coarsely embeds into Hilbert space. Proc. Am. Math. Soc. 149(6), 2693–2703 (2021) [Google Scholar]
  42. Mitsuishi, A.: A splitting theorem for infinite dimensional Alexandrov spaces with nonnegative curvature and its applications. Geom. Dedic. 144, 101–114 (2010) [Google Scholar]
  43. Munch, E.: Applications of Persistent Homology to Time Varying Systems, Ph.D. thesis, Duke University (2013)
  44. Munkres, J.R.: Topology, 2nd edn. Prentice Hall Inc, Upper Saddle River, NJ (2000) [Google Scholar]
  45. Ohta, S.: Barycenters in Alexandrov spaces of curvature bounded below. Adv. Geom. 12(4), 571–587 (2012) [Google Scholar]
  46. Patel, A.: Generalized persistence diagrams. J. Appl. Comput. Topol. 1(3–4), 397–419 (2018) [Google Scholar]
  47. Piotr, W.: Nowak and Guoliang Yu, Large scale geometry. EMS Textbooks in Mathematics, European Mathematical Society, Zürich (2012) [Google Scholar]
  48. Plaut, C.: Metric Spaces of Curvature Inline graphic, Handbook of Geometric Topology, pp. 819–898 (2002)
  49. Roe, J.: Lectures on Coarse Geometry. University Lecture Series, vol. 31. American Mathematical Society, Providence, RI (2003) [Google Scholar]
  50. Santambrogio, F.: Optimal Transport for Applied Mathematicians. Calculus of Variations, PDEs, and Modeling, Prog. Nonlinear Differ. Equ. Appl., vol. 87. Birkhäuser/Springer, Cham (2015)
  51. Skryzalin, J., Carlsson, G.: Numeric invariants from multidimensional persistence. J. Appl. Comput. Topol. 1(1), 89–119 (2017) [Google Scholar]
  52. Turner, K.: Medians of populations of persistence diagrams. Homol. Homotopy Appl. 22(1), 255–282 (2020) [Google Scholar]
  53. Turner, K., Mileyko, Y., Mukherjee, S., Harer, J.: Fréchet means for distributions of persistence diagrams. Discrete Comput. Geom. 52(1), 44–70 (2014) [Google Scholar]
  54. Wagner, A.: Nonembeddability of persistence diagrams with Inline graphic Wasserstein metric. Proc. Am. Math. Soc. 149(6), 2673–2677 (2021) [Google Scholar]
  55. Wofsey, E.: The metrizability of symmetric products of metric spaces, MathOverflow (version: 2015-01-16) (2015)
  56. Yokota, T.: A rigidity theorem in Alexandrov spaces with lower curvature bound. Math. Ann. 353(2), 305–331 (2012) [Google Scholar]
  57. Yokota, T.: On the spread of positively curved Alexandrov spaces. Math. Z. 277(1–2), 293–304 (2014) [Google Scholar]
  58. Zhu, X.: Persistent homology: an introduction and a new text representation for natural language processing. In: Proceedings of the Twenty-Third International Joint Conference on Artificial Intelligence, pp. 1953–1959 (2013)
  59. Zomorodian, A., Carlsson, G.: Computing persistent homology. Discrete Comput. Geom. 33(2), 249–274 (2005) [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Not applicable.


Articles from Journal of Applied and Computational Topology are provided here courtesy of Springer

RESOURCES