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

\givenname

Edoardo \surnameLanari \givennameLuis \surnameScoccola \subjectprimarymsc202018N50, 18N40 \subjectsecondarymsc202055N31, 55U10, 55U35, 62R40 \arxivreference2010.05378

Rectification of interleavings and a persistent Whitehead theorem

Edoardo Lanari Institute of Mathematics, Czech Academy of Sciences, Prague, Czech Republic    Luis Scoccola Department of Mathematics, Northeastern University
Abstract

The homotopy interleaving distance, a distance between persistent spaces, was introduced by Blumberg and Lesnick and shown to be universal, in the sense that it is the largest homotopy-invariant distance for which sublevel-set filtrations of close-by real-valued functions are close-by. There are other ways of constructing homotopy-invariant distances, but not much is known about the relationships between these choices. We show that other natural distances differ from the homotopy interleaving distance in at most a multiplicative constant, and prove versions of the persistent Whitehead theorem, a conjecture of Blumberg and Lesnick that relates morphisms that induce interleavings in persistent homotopy groups to stronger homotopy-invariant notions of interleaving.

1 Introduction

Context.

Many of the main theoretical tools of Topological Data Analysis (TDA) come in the form of stability theorems. One of the best known stability theorems, due to Cohen-Steiner, Edelsbrunner, and Harer ([7]), implies that if f,g:Xf,g:X\to\mathbb{R} are sufficiently tame functions, such as piecewise linear functions on the geometric realization of a finite simplicial complex, then

dB(Dn(f),Dn(g))fg.d_{B}(D_{n}(f),D_{n}(g))\leq\|f-g\|_{\infty}\,.

Here, Dn(f)D_{n}(f) denotes the nn-dimensional persistence diagram of ff. This consists of a multiset of points of the extended plane ¯2\overline{\mathbb{R}}^{2} that captures the isomorphism type of the nnth persistent homology of the sublevel-sets of ff, that is, of the functor 𝐑𝐕𝐞𝐜\mathbf{R}\to\mathbf{Vec} obtained by composing the sublevel-set filtration rf1(,r]:𝐑𝐓𝐨𝐩r\mapsto f^{-1}(-\infty,r]:\mathbf{R}\to\mathbf{Top} with the nnth homology functor Hn:𝐓𝐨𝐩𝐕𝐞𝐜H_{n}:\mathbf{Top}\to\mathbf{Vec}, where 𝐑\mathbf{R} denotes the poset of real numbers and 𝐕𝐞𝐜\mathbf{Vec} denotes the category of vector spaces over some fixed field. The distance dBd_{B} is the bottleneck distance, a combinatorial way of comparing persistence diagrams.

This result was later refined in [4] to the algebraic stability theorem, which says that for F,G:𝐑𝐕𝐞𝐜F,G:\mathbf{R}\to\mathbf{Vec} sufficiently tame functors, one has

dB(D(F),D(G))dI(F,G),d_{B}(D(F),D(G))\leq d_{I}(F,G),

where, as before, D(F)D(F) denotes the persistence diagram of FF, which describes the isomorphism type of FF, and dId_{I} denotes the interleaving distance, a distance between functors 𝐑C\mathbf{R}\to C for any fixed category CC, which we recall below.

Stability theorems imply that pipelines like the following, popular in TDA, are robust to perturbations of the input data and can be used for inference purposes:

DataPersistent spacesHnPersistent vector spaces𝐷Persistence diagrams\footnotesize\text{\framebox[1.1pt]{\begin{tabular}[]{c}Data\end{tabular}}}\xrightarrow{}\text{\framebox[1.1pt]{\begin{tabular}[]{c}Persistent spaces\end{tabular}}}\xrightarrow{H_{n}}\text{\framebox[1.1pt]{\begin{tabular}[]{c}Persistent vector spaces\end{tabular}}}\xrightarrow{D}\text{\framebox[1.1pt]{\begin{tabular}[]{c}Persistence diagrams\end{tabular}}}

For example, the algebraic stability theorem tells us that the last step is stable, if we endow persistent vector spaces (𝐕𝐞𝐜𝐑\mathbf{Vec}^{\mathbf{R}}) with the interleaving distance and persistence diagrams with the bottleneck distance, while functoriality implies that the second step is stable, if we also endow persistent spaces (𝐓𝐨𝐩𝐑\mathbf{Top}^{\mathbf{R}}) with the interleaving distance ([3]).

Problem statement.

Although useful in some applications, the interleaving distance on 𝐓𝐨𝐩𝐑\mathbf{Top}^{\mathbf{R}} is often too fine; for instance, it is easy to see that Vietoris–Rips and other functors S:𝐌𝐞𝐭𝐓𝐨𝐩𝐑S:\mathbf{Met}\to\mathbf{Top}^{\mathbf{R}} are not stable with respect to the Gromov–Hausdorff distance on metric spaces and the interleaving distance on 𝐓𝐨𝐩𝐑\mathbf{Top}^{\mathbf{R}}. However, when one composes these functors with a homotopy-invariant functor, such as homology Hn:𝐓𝐨𝐩𝐑𝐕𝐞𝐜𝐑H_{n}:\mathbf{Top}^{\mathbf{R}}\to\mathbf{Vec}^{\mathbf{R}}, the composite HnS:𝐌𝐞𝐭𝐕𝐞𝐜𝐑H_{n}\circ S:\mathbf{Met}\to\mathbf{Vec}^{\mathbf{R}} turns out to be stable ([5]). So, in these cases, one way to make the first step in the pipeline above stable is to force the interleaving distance on 𝐓𝐨𝐩𝐑\mathbf{Top}^{\mathbf{R}} to be homotopy-invariant ([2, Section 1.2]). For this reason, many homotopy-invariant adaptations of the interleaving distance on 𝐓𝐨𝐩𝐑\mathbf{Top}^{\mathbf{R}} and related categories have been proposed (see, e.g., [15, 2, 8]). In order to describe some of these adaptations, we recall the definition of the interleaving distance dId_{I}.

Let CC be a category. Given δ0𝐑\delta\geq 0\in\mathbf{R} and F:𝐑CF:\mathbf{R}\to C, let Fδ:𝐑CF^{\delta}:\mathbf{R}\to C be given by Fδ(r):=F(r+δ)F^{\delta}(r):=F(r+\delta), with the obvious structure morphisms. One says that F,GC𝐑F,G\in C^{\mathbf{R}} are δ\delta-interleaved if there exist natural transformations f:FGδf:F\to G^{\delta} and g:GFδg:G\to F^{\delta} such that gδf:FF2δg^{\delta}\circ f:F\to F^{2\delta} equals the natural transformation FF2δF\to F^{2\delta} given by the structure morphisms of FF, and such that fδg:GG2δf^{\delta}\circ g:G\to G^{2\delta} equals the natural transformation GG2δG\to G^{2\delta} given by the structure morphisms of GG. Then dI(F,G):=inf({δ0:F and G are δ-interleaved}{})d_{I}(F,G):=\inf\left(\{\delta\geq 0\,:\,\text{$F$ and $G$ are $\delta$-interleaved}\}\cup\{\infty\}\right).

Blumberg and Lesnick ([2]) define X,Y𝐓𝐨𝐩𝐑X,Y\in\mathbf{Top}^{\mathbf{R}} to be δ\delta-homotopy interleaved if there exist weakly equivalent persistent spaces XXX^{\prime}\simeq X and YYY^{\prime}\simeq Y such that XX^{\prime} and YY^{\prime} are δ\delta-interleaved, and use homotopy interleavings to define the homotopy interleaving distance, denoted dHId_{HI}. The homotopy interleaving distance is the (metric) quotient of the interleaving distance by the equivalence relation given by weak equivalence, in the sense that dHId_{HI} is the largest homotopy-invariant distance that is bounded above by the interleaving distance.

Instead of taking a metric quotient, one can take the categorical quotient of 𝐓𝐨𝐩𝐑\mathbf{Top}^{\mathbf{R}} by weak equivalences, and define interleavings directly in the homotopy category, similar to what is done in, e.g., [15, 8, 14]. In order to do this, one notes that the shift functors ()δ:𝐓𝐨𝐩𝐑𝐓𝐨𝐩𝐑(-)^{\delta}:\mathbf{Top}^{\mathbf{R}}\to\mathbf{Top}^{\mathbf{R}} preserve weak equivalences and thus induce functors ()δ:𝖧𝗈(𝐓𝐨𝐩𝐑)𝖧𝗈(𝐓𝐨𝐩𝐑)(-)^{\delta}:\mathsf{Ho}\left(\mathbf{Top}^{\mathbf{R}}\right)\to\mathsf{Ho}\left(\mathbf{Top}^{\mathbf{R}}\right). This lets one copy the definition of interleaving, but in the homotopy category, which gives the notions of interleaving in the homotopy category and of interleaving distance in the homotopy category, denoted dIHCd_{IHC}.

A third option, also introduced in [2], is to compare objects of 𝐓𝐨𝐩𝐑\mathbf{Top}^{\mathbf{R}} using interleavings in 𝖧𝗈(𝐓𝐨𝐩)𝐑\mathsf{Ho}\left(\mathbf{Top}\right)^{\mathbf{R}}, called homotopy commutative interleavings, which give rise to the homotopy commutative interleaving distance, denoted dHCd_{HC}.

We have described three homotopy-invariant notions of interleaving in decreasing order of coherence. On one end, homotopy interleavings can be equivalently described as homotopy coherent diagrams of spaces ([2, Section 7]). On the other end, homotopy commutative interleavings correspond to diagrams in the homotopy category of spaces. It is clear that dHIdIHCdHCd_{HI}\geq d_{IHC}\geq d_{HC}, and that any of the homotopy-invariant interleavings induce interleavings in homotopy groups.

Two questions arise: Are the three distances in some sense equivalent or are they fundamentally different? If a map induces interleavings in homotopy groups, does it follow that the map is part of one of the homotopy-invariant notions of interleaving? A conjectural answer to the second question is given in [2, Conjecture 8.6], where it is conjectured that when XX and YY are a kind of persistent CW-complex of finite dimension dd\in\mathbb{N}, if there exists a morphism between them inducing a δ\delta-interleaving in homotopy groups, then XX and YY are cδc\delta-homotopy interleaved, for a constant cc that only depends on dd.

Contributions.

Homotopy interleavings compose in any functor category of the form 𝐑m\mathcal{M}^{\mathbf{R}^{m}} for \mathcal{M} a cofibrantly generated model category (2.3). This allows us to state some of our results for any cofibrantly generated model category \mathcal{M}, or for a category of spaces SS\SS, which can be instantiated to be any of the Quillen equivalent model categories of topological spaces or simplicial sets (2.1). Our first theorem is the following rectification result.

Theorem A.

Let \mathcal{M} be a cofibrantly generated model category, let X,Y𝐑X,Y\in\mathcal{M}^{\mathbf{R}}, and let δ>0𝐑\delta>0\in\mathbf{R}. If XX and YY are δ\delta-homotopy commutative interleaved, then they are cδc\delta-homotopy interleaved for every c>2c>2.

It follows that we have 2dHCdHIdIHCdHC2d_{HC}\geq d_{HI}\geq d_{IHC}\geq d_{HC}. The above rectification result is different from many such results in homotopy theory, where a diagram of a certain shape, in the homotopy category, is lifted to a strict diagram of the same shape. The difference lies in the fact that the shape of the strict diagram we construct is different from the shape of the diagram in the homotopy category. In fact, building on the suggestion in [2] of using Toda brackets to give a lower bound for the above rectification, we show (3.12) that for =𝐓𝐨𝐩\mathcal{M}=\mathbf{Top}, if cdHCdHIc\,d_{HC}\geq d_{HI} then c3/2c\geq 3/2, so that, in particular, dHCdHId_{HC}\neq d_{HI}. This means that rectification in the usual sense is not possible in general, and thus standard results are not directly applicable. We also show that A has no analogue for multi-persistent spaces (Section 3.3).

Our second theorem relates morphisms inducing interleavings in homotopy groups to interleavings in the homotopy category. See 5.7 for the notion of persistent CW-complex and 5.2 for the notion of interleaving induced in persistent homotopy groups.

Theorem B.

Fix m1m\geq 1\in\mathbb{N} and dd\in\mathbb{N}. Let X,YSS𝐑mX,Y\in\SS^{\mathbf{R}^{m}} be (multi-)persistent spaces that are assumed to be projective cofibrant and dd-skeletal if SS=𝐬𝐒𝐞𝐭\SS=\mathbf{sSet}, or persistent CW-complexes of dimension dd if SS=𝐓𝐨𝐩\SS=\mathbf{Top}. Let δ0𝐑m\delta\geq 0\in\mathbf{R}^{m}. If there exists a morphism in the homotopy category XYδ𝖧𝗈(SS𝐑m)X\to Y^{\delta}\in\mathsf{Ho}\left(\SS^{\mathbf{R}^{m}}\right) that induces δ\delta-interleavings in all homotopy groups, then XX and YY are (4(d+1)δ)(4(d+1)\delta)-interleaved in the homotopy category.

Together, A and B give a positive answer to a version of the persistent Whitehead conjecture [2, Conjecture 8.6] (see 5.14 for a discussion and 5.15 for a statement of the conjecture).

Structure of the paper.

In Section 2, we recall and give references for the necessary background. In Section 3, we prove A, we provide a lower bound for the rectification of homotopy commutative interleavings between persistent spaces, and show that A has no analogue for multi-persistent spaces. In Section 4, we characterize projective cofibrant (multi-)persistent simplicial sets as filtered simplicial sets. In Section 5, we prove B.

Acknowledgements.

The first named author gratefully acknowledges the support of Praemium Academiae of M. Markl and RVO:67985840. The second named author thanks Dan Christensen, Rick Jardine, Mike Lesnick, and Alex Rolle for insightful conversations. We thank Alex Rolle for detailed feedback and for suggesting 3.14 to us, Mike Lesnick for suggesting improvements to the constant of A, and the anonymous referee for helpful feedback.

2 Background and conventions

The main purpose of this section is to fix notation and to provide the reader with references. This section can be referred to as needed, but we do recommend going over Section 2.2 as it contains the notions of interleaving relevant to us.

We assume that the reader is comfortable with the language of category theory. Throughout the paper, we will use the term distance to refer to any extended pseudo metric on a (possibly large) set XX, that is, to any function dX:X×X[0,]d_{X}:X\times X\to[0,\infty] that is symmetric, satisfies the triangle inequality, and is 0 on the diagonal.

2.1 Spaces and model categories

2.1.1 Spaces

We work model-independently whenever possible. This means that whenever we say space we will mean either topological space or simplicial set. Results stated for spaces will hold for both possible models. The category of spaces will be denoted by SS\SS.

For a general introduction to simplicial sets see, e.g., [9] or [12, Chapter 3]. We denote the geometric realization functor for simplicial sets by ||:𝐬𝐒𝐞𝐭𝐓𝐨𝐩|-|:\mathbf{sSet}\to\mathbf{Top}.

2.1.2 Model categories

The theory of model categories was introduced in [18]; for a modern and thorough development of this theory we recommend [12] and [11].

We recall that two objects x,yx,y\in\mathcal{M} of a model category \mathcal{M} are said to be weakly equivalent if they are isomorphic in 𝖧𝗈()\mathsf{Ho}(\mathcal{M}), which happens if and only if they are connected by a zig-zag of weak equivalences in \mathcal{M}. This is an equivalence relation, which we denote by xyx\simeq y. When there is risk of confusion, morphisms in 𝖧𝗈()\mathsf{Ho}(\mathcal{M}) will be surrounded by square brackets [f][f], to distinguish them from morphisms in \mathcal{M}.

Two of the main model structures of interest to us are the Quillen model structure on 𝐓𝐨𝐩\mathbf{Top}, the category of topological spaces ([12, Chapter 1, Section 2.4]), and the Kan–Quillen model structure on 𝐬𝐒𝐞𝐭\mathbf{sSet}, the category of simplicial sets ([12, Chapter 3]). We recall that the geometric realization functor ||:𝐬𝐒𝐞𝐭𝐓𝐨𝐩|-|:\mathbf{sSet}\to\mathbf{Top} is left adjoint to the singular functor Sing:𝐓𝐨𝐩𝐬𝐒𝐞𝐭\operatorname{Sing}:\mathbf{Top}\to\mathbf{sSet}, and that, together, they form a Quillen equivalence ([12, Chapter 1, Section 1.3], [12, Theorem 3.6.7]). For completeness, we mention that there is a subcategory 𝐓𝐨𝐩𝖢𝖦𝖶𝖧𝐓𝐨𝐩\mathbf{Top}_{\mathsf{CGWH}}\subseteq\mathbf{Top}, the category of compactly-generated weakly Hausdorff topological spaces (called compactly generated spaces in [12, Definition 2.4.21]), that is often used instead of 𝐓𝐨𝐩\mathbf{Top}. The Quillen model structure on 𝐓𝐨𝐩\mathbf{Top} restricts to a model structure on 𝐓𝐨𝐩𝖢𝖦𝖶𝖧\mathbf{Top}_{\mathsf{CGWH}}, and the inclusion 𝐓𝐨𝐩𝖢𝖦𝖶𝖧𝐓𝐨𝐩\mathbf{Top}_{\mathsf{CGWH}}\to\mathbf{Top} is part of a Quillen equivalence ([12, Theorem 2.4.25]). This model structure is, in some respects, better behaved than the Quillen model structure on topological spaces, and is in fact the model of space used in [2]. We will not concern ourselves with these subtleties since, by the observations in 2.1, there is no essential difference between using 𝐓𝐨𝐩\mathbf{Top} or 𝐓𝐨𝐩𝖢𝖦𝖶𝖧\mathbf{Top}_{\mathsf{CGWH}} when studying homotopy-invariant notions of interleaving.

We will make use of the notion of cofibrantly generated model category ([12, Chapter 2, Section 2.1]). Recall that the Kan–Quillen model structure on simplicial sets is cofibrantly generated, where a set of generating cofibrations consists of the boundary inclusions ΔnΔn\partial\Delta^{n}\hookrightarrow\Delta^{n} for n0n\geq 0 ([12, Theorem 3.6.5]). The Quillen model structure on topological spaces is also cofibrantly generated, with a set of generating cofibrations given by {Sn1Dn}n0\{S^{n-1}\hookrightarrow D^{n}\}_{n\geq 0} ([12, Theorem 2.4.19]).

We conclude by recalling the basic properties of projective model structures. Given a model category \mathcal{M} and a small category 𝐈{\mathbf{I}}, the projective model structure on the functor category 𝐈\mathcal{M}^{{\mathbf{I}}} is, when it exists, the model structure whose fibrations (respectively weak equivalences) are those which are pointwise fibrations (respectively weak equivalences) of \mathcal{M}.

The projective model structure on 𝐈\mathcal{M}^{\mathbf{I}} exists, and is cofibrantly generated, whenever \mathcal{M} is cofibrantly generated. Moreover, if {\mathcal{I}} and 𝒥{\mathcal{J}} are, respectively, generating cofibrations and generating trivial cofibration for the model structure of \mathcal{M}, then {𝐈(i,)f:i𝐈,f}\left\{{\mathbf{I}}(i,-)\odot f\ :\ i\in{\mathbf{I}},\,f\in{\mathcal{I}}\right\} and {𝐈(i,)g:i𝐈,g𝒥}\left\{{\mathbf{I}}(i,-)\odot g\ :\ i\in{\mathbf{I}},\,g\in{\mathcal{J}}\right\} are, respectively, generating cofibrations and generating trivial cofibrations for the projective model structure, where, given a functor F:𝐈𝐒𝐞𝐭F:{\mathbf{I}}\to\mathbf{Set} and an object XX\in\mathcal{M}, the functor FX:𝐈F\odot X:{\mathbf{I}}\to\mathcal{M} is defined by iaF(i)Xi\mapsto\coprod_{a\in F(i)}X ([11, Section 11.6]). For simplicity, we denote 𝐈(i,)X{\mathbf{I}}(i,-)\odot X by iXi\odot X.

We are especially interested in the projective model structure when the indexing category is a poset (P,)(P,\leq). In this case, if rPr\in P and XX\in\mathcal{M}, then rXr\odot X is the functor that takes the value XX on every srs\geq r, and has as value the initial object of \mathcal{M} when srs\nleq r. The non-trivial structure morphisms of this functor are the identity of XX.

Note that we have a functor 𝗁:𝖧𝗈(𝐈)𝖧𝗈()𝐈\mathsf{h}:\mathsf{Ho}\left(\mathcal{M}^{\mathbf{I}}\right)\to\mathsf{Ho}(\mathcal{M})^{\mathbf{I}} by the universal property of 𝖧𝗈(𝐈)\mathsf{Ho}\left(\mathcal{M}^{\mathbf{I}}\right).

2.2 Interleavings and interleavings up to homotopy

2.2.1 Strict interleavings

We denote the poset of real numbers with their standard order by 𝐑\mathbf{R}, and for mm\in\mathbb{N}, we let 𝐑m\mathbf{R}^{m} be the set of mm-tuples of real numbers with the product order. We set m¯={i:1im}\overline{m}=\{i:1\leq i\leq m\}, so that (εi)im¯(δi)im¯𝐑m(\varepsilon_{i})_{i\in\overline{m}}\leq(\delta_{i})_{i\in\overline{m}}\in\mathbf{R}^{m} if and only if εiδi\varepsilon_{i}\leq\delta_{i} for all 1im1\leq i\leq m. We denote the element (0,,0)𝐑m(0,\dots,0)\in\mathbf{R}^{m} by 0.

Fix a category CC and a natural number m1m\geq 1. An 𝒎m-persistent object of CC is any functor of the form 𝐑mC\mathbf{R}^{m}\to C. We often refer to mm-persistent objects simply as persistent objects or as multi-persistent objects when we want to stress the fact that mm is not necessarily 11. Fix persistent objects X,Y,ZC𝐑mX,Y,Z\in C^{\mathbf{R}^{m}}, r,s𝐑mr,s\in\mathbf{R}^{m}, and ε,δ0𝐑m\varepsilon,\delta\geq 0\in\mathbf{R}^{m}. We use the following conventions.

  • For f:XYf:X\to Y a natural transformation, denote the rr-component of ff by fr:X(r)Y(r)f_{r}:X(r)\to Y(r).

  • Assume rsr\leq s. The structure morphism X(r)X(s)X(r)\to X(s) will be denoted by φr,sX\varphi^{X}_{r,s}.

  • The 𝜹\delta-shift to the left of XX is the functor Xδ:𝐑mCX^{\delta}:\mathbf{R}^{m}\to C defined by Xδ(r)=X(r+δ)X^{\delta}(r)=X(r+\delta), with structure morphisms φr,sXδ:=φr+δ,s+δX\varphi^{X^{\delta}}_{r,s}:=\varphi^{X}_{r+\delta,s+\delta}. Shifting to the left gives a functor ()δ:C𝐑mC𝐑m(-)^{\delta}:C^{\mathbf{R}^{m}}\to C^{\mathbf{R}^{m}}. Dually, there is a 𝜹\delta-shift to the right functor δ():C𝐑mC𝐑m\delta\cdot(-):C^{\mathbf{R}^{m}}\to C^{\mathbf{R}^{m}} defined by mapping XX to the persistent object δX\delta\cdot X, with values given by (δX)(r)=X(rδ)(\delta\cdot X)(r)=X(r-\delta).

  • Natural transformations f:XYδf:X\to Y^{\delta} will be referred to as 𝜹\delta-morphisms, and will often be denoted by f:XδYf:X\to_{\delta}Y. Since we have natural bijections Hom(εX,Yδ)Hom(X,Yε+δ)Hom((ε+δ)X,Y)\operatorname{Hom}(\varepsilon\cdot X,Y^{\delta})\cong\operatorname{Hom}(X,Y^{\varepsilon+\delta})\cong\operatorname{Hom}((\varepsilon+\delta)\cdot X,Y), we can treat a δ\delta-morphism f:XδYf:X\to_{\delta}Y as f:XYδf:X\to Y^{\delta} or as f:δXYf:\delta\cdot X\to Y.

  • Assume εδ\varepsilon\leq\delta and let f:XεYf:X\to_{\varepsilon}Y. We can compose the rr-component of ff with φr+ε,r+δY:Y(r+ε)Y(r+δ)\varphi^{Y}_{r+\varepsilon,r+\delta}:Y(r+\varepsilon)\to Y(r+\delta), giving φr+ε,r+δYfr:X(r)Y(r+δ)\varphi^{Y}_{r+\varepsilon,r+\delta}\circ f_{r}:X(r)\to Y(r+\delta). Together, these components define the shift from ε\varepsilon to δ\delta of ff, which is a δ\delta-morphism denoted 𝖲ε,δ(f):XδY\mathsf{S}_{\varepsilon,\delta}(f):X\to_{\delta}Y.

  • Note that an ε\varepsilon-morphism f:XεYf:X\to_{\varepsilon}Y can be composed with a δ\delta-morphism g:YδZg:Y\to_{\delta}Z, yielding an (ε+δ)(\varepsilon+\delta)-morphism gεf:Xε+δYg^{\varepsilon}\circ f:X\to_{\varepsilon+\delta}Y. This composition is associative and unital, and is natural with respect to shifts of morphisms.

  • An (𝜺,𝜹)(\varepsilon,\delta)-interleaving between XX and YY consists of an ε\varepsilon-morphism f:XεYf:X\to_{\varepsilon}Y together with a δ\delta-morphism g:YδXg:Y\to_{\delta}X such that gεf=𝖲0,ε+δ(𝗂𝖽X)g^{\varepsilon}\circ f=\mathsf{S}_{0,\varepsilon+\delta}(\mathsf{id}_{X}) and fδg=𝖲0,ε+δ(𝗂𝖽Y)f^{\delta}\circ g=\mathsf{S}_{0,\varepsilon+\delta}(\mathsf{id}_{Y}). By 𝜹\delta-interleaving we mean a (δ,δ)(\delta,\delta)-interleaving.

  • If f:XεYf:X\to_{\varepsilon}Y and g:YδXg:Y\to_{\delta}X form an (ε,δ)(\varepsilon,\delta)-interleaving, we write f:XεδY:gf:X\prescript{}{{\scriptscriptstyle\delta}}{\longleftrightarrow}_{{\scriptscriptstyle\varepsilon}}\,Y:g.

Let ε1,ε2,δ1,δ20𝐑m\varepsilon_{1},\varepsilon_{2},\delta_{1},\delta_{2}\geq 0\in\mathbf{R}^{m}. Note that an (ε1,ε2)(\varepsilon_{1},\varepsilon_{2})-interleaving between XX and YY can be composed with any (δ1,δ2)(\delta_{1},\delta_{2})-interleaving between YY and ZZ, yielding an (ε1+δ1,ε2+δ2)(\varepsilon_{1}+\delta_{1},\varepsilon_{2}+\delta_{2})-interleaving. The fact that interleavings compose implies that, when m=1m=1, the formula

dI(X,Y)=inf({δ0𝐑:X and Y are δ-interleaved}{})d_{I}(X,Y)=\inf\left(\{\delta\geq 0\in\mathbf{R}:\text{$X$ and $Y$ are $\delta$-interleaved}\}\cup\{\infty\}\right)

defines an extended pseudo metric dI:𝖮𝖻𝗃(C𝐑)×𝖮𝖻𝗃(C𝐑)[0,]d_{I}:\mathsf{Obj}\left(C^{\mathbf{R}}\right)\times\mathsf{Obj}\left(C^{\mathbf{R}}\right)\to[0,\infty]. This is the interleaving distance on the class of objects of the category C𝐑C^{\mathbf{R}}. This notion of distance can be extended to objects of the functor category C𝐑mC^{\mathbf{R}^{m}} ([15]), but we will not make use of this extension.

2.2.2 Interleavings up to homotopy

If one is comparing objects of a category of functors of the form 𝐑m\mathbf{R}^{m}\to\mathcal{M}, for \mathcal{M} a model category, it makes sense to want to find a homotopy-invariant notion of interleaving. In this paper, we consider the following three homotopy-invariant relaxations of the notion of interleaving. Let \mathcal{M} be a cofibrantly generated model category and endow 𝐑m\mathcal{M}^{\mathbf{R}^{m}} with the projective model structure. Let X,Y𝐑mX,Y\in\mathcal{M}^{\mathbf{R}^{m}} and let ε,δ0𝐑m\varepsilon,\delta\geq 0\in\mathbf{R}^{m}.

  1. 1.

    Following [2], we say that XX and YY are (𝜺,𝜹)(\varepsilon,\delta)-homotopy interleaved if there exist XXX\simeq X^{\prime} and YYY\simeq Y^{\prime} such that XX^{\prime} and YY^{\prime} are (ε,δ)(\varepsilon,\delta)-interleaved.

  2. 2.

    Note that the shift functor ()δ:𝐑m𝐑m(-)^{\delta}:\mathcal{M}^{\mathbf{R}^{m}}\to\mathcal{M}^{\mathbf{R}^{m}} maps weak equivalences to weak equivalences. This implies that all the notions in Section 2.2.1 have analogues in the category 𝖧𝗈(𝐑m)\mathsf{Ho}\left(\mathcal{M}^{\mathbf{R}^{m}}\right). We say that XX and YY are (𝜺,𝜹)(\varepsilon,\delta)-interleaved in the homotopy category if they are (ε,δ)(\varepsilon,\delta)-interleaved as objects of 𝖧𝗈(𝐑m)\mathsf{Ho}\left(\mathcal{M}^{\mathbf{R}^{m}}\right).

  3. 3.

    Finally, as also done in [2], we say that XX and YY are (𝜺,𝜹)(\varepsilon,\delta)-homotopy commutative interleaved if their images 𝗁X,𝗁Y:𝐑m𝖧𝗈()\mathsf{h}X,\mathsf{h}Y:\mathbf{R}^{m}\to\mathsf{Ho}(\mathcal{M}) are (ε,δ)(\varepsilon,\delta)-interleaved.

An (ε,δ)(\varepsilon,\delta)-homotopy interleaving gives rise to an (ε,δ)(\varepsilon,\delta)-interleaving in the homotopy category, which, in turn, gives rise to an (ε,δ)(\varepsilon,\delta)-homotopy commutative interleaving.

For each of the three homotopy-invariant notions of interleaving introduced above, we have a corresponding extended pseudo metric on the collection of objects of the category 𝐑\mathcal{M}^{\mathbf{R}}. Let X,Y𝐑X,Y\in\mathcal{M}^{\mathbf{R}}. Following [2], we define the homotopy interleaving distance as

dHI(X,Y)=inf({δ0𝐑:X and Y are δ-homotopy interleaved}{}).d_{HI}(X,Y)=\inf\left(\{\delta\geq 0\in\mathbf{R}:\text{$X$ and $Y$ are $\delta$-homotopy interleaved}\}\cup\{\infty\}\right).

The fact that the homotopy interleaving distance satisfies the triangle inequality follows from 2.3. The interleaving distance in the homotopy category is

dIHC(X,Y)=inf({δ0𝐑:X and Y are δ-interleaved in the homotopy category}{}).d_{IHC}(X,Y)=\inf\left(\{\delta\geq 0\in\mathbf{R}:\text{$X$ and $Y$ are $\delta$-interleaved in the homotopy category}\}\cup\{\infty\}\right).

Again following [2], the homotopy commutative interleaving distance is defined as

dHC(X,Y)=inf({δ0𝐑:X and Y are δ-homotopy commutative interleaved}{}).d_{HC}(X,Y)=\inf\left(\{\delta\geq 0\in\mathbf{R}:\text{$X$ and $Y$ are $\delta$-homotopy commutative interleaved}\}\cup\{\infty\}\right).
Remark 2.1.

Note that if 𝒩\mathcal{M}\rightleftarrows\mathcal{N} is a Quillen equivalence between cofibrantly generated model categories, then the induced Quillen equivalence ([11, Theorem 11.6.5]) 𝐑m𝒩𝐑m\mathcal{M}^{\mathbf{R}^{m}}\rightleftarrows\mathcal{N}^{\mathbf{R}^{m}} between the projective model structures respects interleavings, in the sense that shifts commute with both the left and right adjoints. This implies that, for any of the three homotopy-invariant notions of interleaving described above, we have that two functors on one side of the adjunction are (ε,δ)(\varepsilon,\delta)-interleaved if and only if their images (along the derived adjunction) on the other side are (ε,δ)(\varepsilon,\delta)-interleaved. In particular, if m=1m=1, the two adjoints give an isometry between 𝐑\mathcal{M}^{\mathbf{R}} and 𝒩𝐑\mathcal{N}^{\mathbf{R}} independently of whether we use dHId_{HI}, dIHCd_{IHC}, or dHCd_{HC}.

2.2.3 Composability of homotopy interleavings

In this short section, we give a simplified proof of a generalization of the fact that homotopy interleavings can be composed, originally proved in [2, Section 4]. This is generalized further in [20, Theorem 4.1.4].

Lemma 2.2.

Let CC admit pullbacks. Fix m1m\geq 1\in\mathbb{N}, objects X,Y,B:𝐑mCX,Y,B:\mathbf{R}^{m}\to C, elements ε,δ0𝐑m\varepsilon,\delta\geq 0\in\mathbf{R}^{m}, an (ε,δ)(\varepsilon,\delta)-interleaving f:XεδY:gf:X\prescript{}{{\scriptscriptstyle\delta}}{\longleftrightarrow}_{{\scriptscriptstyle\varepsilon}}\,Y:g, and a map h:BYh:B\to Y. The pullback of f:XYεf:X\to Y^{\varepsilon} along hε:BεYεh^{\varepsilon}:B^{\varepsilon}\to Y^{\varepsilon}, denoted k:ABεk:A\to B^{\varepsilon}, is part of an (ε,δ)(\varepsilon,\delta)-interleaving k:AεδB:lk:A\prescript{}{{\scriptscriptstyle\delta}}{\longleftrightarrow}_{{\scriptscriptstyle\varepsilon}}\,B:l.

Proof.

We start by depicting the pullback square in the statement:

A{A}Bε{B^{\varepsilon}}X{X}Yε.{Y^{\varepsilon}.}kkffhεh^{\varepsilon}

Consider the morphisms i=𝖲0,ε+δ(𝗂𝖽B):δBBεi=\mathsf{S}_{0,\varepsilon+\delta}(\mathsf{id}_{B}):\delta\cdot B\to B^{\varepsilon} and g(δh):δBXg\circ(\delta\cdot h):\delta\cdot B\to X. Since fg(δh)=hεif\circ g\circ(\delta\cdot h)=h^{\varepsilon}\circ i, the universal property of AA gives us a map l:δBAl:\delta\cdot B\to A, or equivalently, a map l:BAδl:B\to A^{\delta}. By construction, kδl=𝖲0,ε+δ(𝗂𝖽B):BBε+δk^{\delta}\circ l=\mathsf{S}_{0,\varepsilon+\delta}(\mathsf{id}_{B}):B\to B^{\varepsilon+\delta}. To prove that lεk=𝖲0,ε+δ(𝗂𝖽A):AAε+δl^{\varepsilon}\circ k=\mathsf{S}_{0,\varepsilon+\delta}(\mathsf{id}_{A}):A\to A^{\varepsilon+\delta}, or equivalently that lεk=𝖲0,ε+δ(𝗂𝖽A):εAAδl^{\varepsilon}\circ k=\mathsf{S}_{0,\varepsilon+\delta}(\mathsf{id}_{A}):\varepsilon\cdot A\to A^{\delta}, apply the functor ()δ:C𝐑mC𝐑m(-)^{\delta}:C^{\mathbf{R}^{m}}\to C^{\mathbf{R}^{m}} to the pullback square above, and use the uniqueness part of its universal property. ∎

Proposition 2.3 (cf. [2, Section 4]).

Let \mathcal{M} be cofibrantly generated, fix m1m\geq 1, let X,Y,Z:𝐑mX,Y,Z:\mathbf{R}^{m}\to\mathcal{M}, and let ε1,ε2,δ1,δ20𝐑m\varepsilon_{1},\varepsilon_{2},\delta_{1},\delta_{2}\geq 0\in\mathbf{R}^{m}. If XX and YY are (ε1,ε2)(\varepsilon_{1},\varepsilon_{2})-homotopy interleaved and YY and ZZ are (δ1,δ2)(\delta_{1},\delta_{2})-homotopy interleaved, then XX and ZZ are (ε1+δ1,ε2+δ2)(\varepsilon_{1}+\delta_{1},\varepsilon_{2}+\delta_{2})-homotopy interleaved.

Proof.

Given interleavings Xε1ε2YX^{\prime}\prescript{}{{\scriptscriptstyle\varepsilon_{2}}}{\longleftrightarrow}_{{\scriptscriptstyle\varepsilon_{1}}}\,Y^{\prime} and Y′′δ1δ2ZY^{\prime\prime}\prescript{}{{\scriptscriptstyle\delta_{2}}}{\longleftrightarrow}_{{\scriptscriptstyle\delta_{1}}}\,Z^{\prime} with XXX\simeq X^{\prime}, YYYY^{\prime}\simeq Y\simeq Y^{\prime}, and ZZZ^{\prime}\simeq Z, we must construct an interleaving X′′ε1+δ1ε2+δ2Z′′X^{\prime\prime}\prescript{}{{\scriptscriptstyle\varepsilon_{2}+\delta_{2}}}{\longleftrightarrow}_{{\scriptscriptstyle\varepsilon_{1}+\delta_{1}}}\,Z^{\prime\prime} with X′′XX^{\prime\prime}\simeq X and Z′′ZZ^{\prime\prime}\simeq Z.

Since \mathcal{M} is cofibrantly generated, the projective model structure on 𝐑m\mathcal{M}^{\mathbf{R}^{m}} exists, and, by applying a functorial fibrant replacement \mathcal{M}\to\mathcal{M} pointwise, we get a functorial fibrant replacement 𝐑m𝐑m\mathcal{M}^{\mathbf{R}^{m}}\to\mathcal{M}^{\mathbf{R}^{m}}. By construction, the fibrant replacement 𝐑m𝐑m\mathcal{M}^{\mathbf{R}^{m}}\to\mathcal{M}^{\mathbf{R}^{m}} commutes with ()δ:𝐑m𝐑m(-)^{\delta}:\mathcal{M}^{\mathbf{R}^{m}}\to\mathcal{M}^{\mathbf{R}^{m}} so, in particular, it preserves interleavings. With this in mind, we can assume that YY^{\prime} and Y′′Y^{\prime\prime} are fibrant, which implies, and this is a general fact, that we have C𝐑mC\in\mathcal{M}^{\mathbf{R}^{m}} and trivial fibrations CYC\to Y^{\prime} and CY′′C\to Y^{\prime\prime}. Using 2.2, we can pull back the interleavings we were given along the trivial fibrations, as follows:

X′′{X^{\prime\prime}}C{C}Z′′{Z^{\prime\prime}}X{X^{\prime}}Y{Y^{\prime}}Y′′{Y^{\prime\prime}}Z.{Z^{\prime}.}ε1{\scriptscriptstyle\varepsilon_{1}}ε2{\scriptscriptstyle\varepsilon_{2}}ε1{\scriptscriptstyle\varepsilon_{1}}ε2{\scriptscriptstyle\varepsilon_{2}}δ1{\scriptscriptstyle\delta_{1}}δ2{\scriptscriptstyle\delta_{2}}δ1{\scriptscriptstyle\delta_{1}}δ2{\scriptscriptstyle\delta_{2}}

Since trivial fibrations are stable under pullback, we have that X′′XX^{\prime\prime}\simeq X and Z′′ZZ^{\prime\prime}\simeq Z, and since interleavings compose, we have that X′′X^{\prime\prime} and Z′′Z^{\prime\prime} are (ε1+δ1,ε2+δ2)(\varepsilon_{1}+\delta_{1},\varepsilon_{2}+\delta_{2})-interleaved, as required. ∎

We remark that the idea of using pullbacks to prove a triangle inequality appears in [17].

3 Interleavings in 𝐑\mathcal{M}^{\mathbf{R}} and in 𝖧𝗈()𝐑\mathsf{Ho}(\mathcal{M})^{\mathbf{R}}

This section is concerned with the rectification of homotopy commutative interleavings into homotopy interleavings. In Section 3.1, we prove A, which allows one to construct, for any c>2c>2, a cδc\delta-homotopy interleaving out of a δ\delta-homotopy commutative interleaving, when working with 11-persistent objects of any cofibrantly generated model category \mathcal{M}. We think of this result as giving a multiplicative upper bound of 22 for this rectification. In Section 3.2, we give a multiplicative lower bound of 3/23/2 for the rectification, when \mathcal{M} is the category of spaces. In Section 3.3, we show that A has no analogue for multi-persistent spaces.

3.1 Upper bound

Let 𝐙𝐑\mathbf{Z}\subseteq\mathbf{R} denote the posets of integers and real numbers respectively. The inclusion i:𝐙𝐑i:\mathbf{Z}\to\mathbf{R} induces a restriction functor i:C𝐑C𝐙i^{*}:C^{\mathbf{R}}\to C^{\mathbf{Z}} for any category CC. Given A:𝐙CA:\mathbf{Z}\to C, let i(A):𝐑Ci_{*}(A):\mathbf{R}\to C be given by AA precomposed with the functor :𝐑𝐙\lfloor-\rfloor:\mathbf{R}\to\mathbf{Z}, where r\lfloor r\rfloor is the largest integer bounded above by rr. Note that, given m0𝐙m\geq 0\in\mathbf{Z}, one has a notion of mm-interleaving between functors A,B:𝐙CA,B:\mathbf{Z}\to C, and that i:C𝐙C𝐑i_{*}:C^{\mathbf{Z}}\to C^{\mathbf{R}} preserves these interleavings.

We start with a few simplifications. Given δ>0\delta>0, let Mδ:𝐑𝐑M_{\delta}:\mathbf{R}\to\mathbf{R} be given by Mδ(r)=δ×rM_{\delta}(r)=\delta\times r. The following lemma allows us to work with integer-valued interleavings instead of δ\delta-interleavings, and its proof is immediate.

Lemma 3.1.

Let δ>0𝐑\delta>0\in\mathbf{R} and let m1𝐙m\geq 1\in\mathbf{Z}. Then X,YC𝐑X,Y\in C^{\mathbf{R}} are δ\delta-interleaved if and only if (Mδ/m)(X)(M_{\delta/m})^{*}(X) and (Mδ/m)(Y)(M_{\delta/m})^{*}(Y) are mm-interleaved.∎

The following lemma allows us to work with 𝐙\mathbf{Z}-indexed persistent objects instead of 𝐑\mathbf{R}-indexed ones. Here, by homotopy interleaving between 𝐙\mathbf{Z}-indexed functors we mean the obvious adaptation of the notion of homotopy interleaving to 𝐙\mathbf{Z}-indexed functors with values in a model category.

Lemma 3.2.

Let \mathcal{M} be cofibrantly generated. Let X,Y𝐑X,Y\in\mathcal{M}^{\mathbf{R}} and let m1𝐙m\geq 1\in\mathbf{Z}. If i(X),i(Y)𝐙i^{*}(X),i^{*}(Y)\in\mathcal{M}^{\mathbf{Z}} are mm-homotopy interleaved, then XX and YY are (m+2)(m+2)-homotopy interleaved.

Proof.

Note that XX is 11-interleaved with i(i(X))i_{*}(i^{*}(X)), as, for all r𝐑r\in\mathbf{R}, we have r1rrr+1r-1\leq\lfloor r\rfloor\leq r\leq\lfloor r\rfloor+1. Since ii_{*} preserves interleavings and weak equivalences, it is enough to show that homotopy interleavings between 𝐙\mathbf{Z}-indexed functors with values in a cofibrantly generated model category compose, which is a straightforward adaptation of 2.3 to 𝐙\mathbf{Z}-indexed functors. ∎

The next straightforward lemma gives us a special replacement of an object of the category 𝐙\mathcal{M}^{\mathbf{Z}}, with \mathcal{M} a model category, that will be useful when lifting structure from 𝖧𝗈()𝐙\mathsf{Ho}(\mathcal{M})^{\mathbf{Z}} to 𝐙\mathcal{M}^{\mathbf{Z}}.

Lemma 3.3.

Given a model category \mathcal{M} and X𝐙X\in\mathcal{M}^{\mathbf{Z}}, there exists X¯𝐙\overline{X}\in\mathcal{M}^{\mathbf{Z}} and a weak equivalence X¯X\overline{X}\to X, such that the following properties are satisfied:

  • X¯(i)\overline{X}(i) is cofibrant in \mathcal{M} for every i𝐍i\in\mathbf{N};

  • for every i0i\geq 0, the structure morphism fi:X¯(i)X¯(i+1)f_{i}:\overline{X}(i)\to\overline{X}(i+1) is a cofibration in \mathcal{M}.

Dually, we can replace Y𝐙Y\in\mathcal{M}^{\mathbf{Z}} by a pointwise fibrant Y¯\overline{Y} whose “negative” maps are fibrations.

The following lemma will allow us to lift interleavings in 𝖧𝗈()𝐙\mathsf{Ho}(\mathcal{M})^{\mathbf{Z}} to homotopy interleavings in 𝐙\mathcal{M}^{\mathbf{Z}}.

Lemma 3.4.

Let \mathcal{M} be a model category. The functor 𝗁:𝖧𝗈(𝐙)𝖧𝗈()𝐙\mathsf{h}:\mathsf{Ho}(\mathcal{M}^{\mathbf{Z}})\to\mathsf{Ho}(\mathcal{M})^{\mathbf{Z}} is essentially surjective, conservative, and full. In particular, if A,B𝐙A,B\in\mathcal{M}^{\mathbf{Z}} become isomorphic in 𝖧𝗈()𝐙\mathsf{Ho}(\mathcal{M})^{\mathbf{Z}}, then they are weakly equivalent.

Proof.

It is clear that the functor is essentially surjective and full, so we only prove the last property. Assume given X,Y𝖧𝗈(𝐙)X,Y\in\mathsf{Ho}(\mathcal{M}^{\mathbf{Z}}) together with a map f:𝗁X𝗁Yf:\mathsf{h}X\to\mathsf{h}Y. Thanks to 3.3, we can assume that XX (respectively YY) is pointwise cofibrant (respectively fibrant) in \mathcal{M}, and that all the non-negative (respectively negative) structural maps in XX (respectively YY) are cofibrations (respectively fibrations). The map ff can therefore be represented as a family {[fi]}i𝐙\{[f_{i}]\}_{i\in\mathbf{Z}} of homotopy classes of maps of \mathcal{M}. We construct a preimage of ff under 𝗁\mathsf{h} inductively, starting with a choice of representatives fif^{\prime}_{i} for the homotopy classes [fi][f_{i}]. The squares

X(1){X(-1)}X(0){X(0)}X(1){X(1)}Y(1){Y(-1)}Y(0){Y(0)}Y(1){Y(1)}f1\scriptstyle{f^{\prime}_{-1}}x1\scriptstyle{x_{-1}}f0\scriptstyle{f^{\prime}_{0}}x0\scriptstyle{x_{0}}f1\scriptstyle{f^{\prime}_{1}}y1\scriptstyle{y_{-1}}y0\scriptstyle{y_{0}}

commute up to homotopy, and since x0x_{0} and y1y_{-1} are, respectively, a cofibration and a fibration, we can deform f1f^{\prime}_{1} and f1f^{\prime}_{-1} into homotopic maps f1:X1Y1f_{1}:X_{1}\to Y_{1} and f1:X1Y1f_{-1}:X_{-1}\to Y_{-1}, which render the above squares commutative. Inductively, we can iterate this procedure to find the desired preimage of ff under 𝗁\mathsf{h}. ∎

The next result is the main rectification step involved in lifting interleavings in 𝖧𝗈()𝐙\mathsf{Ho}(\mathcal{M})^{\mathbf{Z}} to homotopy interleavings in 𝐙\mathcal{M}^{\mathbf{Z}}.

Proposition 3.5.

Let \mathcal{M} be a model category and let A,B𝐙A,B\in\mathcal{M}^{\mathbf{Z}}. Let m1𝐙m\geq 1\in\mathbf{Z}. If 𝗁A\mathsf{h}A and 𝗁B\mathsf{h}B are mm-interleaved in 𝖧𝗈()𝐙\mathsf{Ho}(\mathcal{M})^{\mathbf{Z}}, then AA and BB are 2m2m-homotopy interleaved in 𝐙\mathcal{M}^{\mathbf{Z}}.

Proof.

We start by giving the proof for the case m=1m=1, as in this case the main idea is more clear. We will use the following constructions. Let 𝖾:𝐙𝐙\mathsf{e}:\mathbf{Z}\to\mathbf{Z} be the functor that maps even numbers to themselves and an odd number nn to n1n-1. Similarly, let 𝗈:𝐙𝐙\mathsf{o}:\mathbf{Z}\to\mathbf{Z} be the functor that maps odd numbers to themselves and an even number nn to n1n-1.

Note that, for every C𝐙C\in\mathcal{M}^{\mathbf{Z}}, we have that CC is (1,0)(1,0)-interleaved with 𝖾(C)\mathsf{e}^{*}(C) and with 𝗈(C)\mathsf{o}^{*}(C), and that 𝖾(C)\mathsf{e}^{*}(C) and 𝗈(C)\mathsf{o}^{*}(C) are 11-interleaved.

Now assume given a 11-interleaving between 𝗁A\mathsf{h}A and 𝗁B\mathsf{h}B in 𝖧𝗈()𝐙\mathsf{Ho}(\mathcal{M})^{\mathbf{Z}}, that is, assume that there are morphisms fi:𝗁A(i)𝗁B(i+1)f_{i}:\mathsf{h}A(i)\to\mathsf{h}B(i+1) and gi:𝗁B(i)𝗁A(i+1)g_{i}:\mathsf{h}B(i)\to\mathsf{h}A(i+1) in 𝖧𝗈()\mathsf{Ho}(\mathcal{M}) rendering the following diagram commutative:

{\cdots}𝗁A(1){\mathsf{h}A(-1)}𝗁A(0){\mathsf{h}A(0)}𝗁A(1){\mathsf{h}A(1)}{\cdots}{\cdots}𝗁B(1){\mathsf{h}B(-1)}𝗁B(0){\mathsf{h}B(0)}𝗁B(1){\mathsf{h}B(1)}{\cdots}[β2]\,\,[\beta_{-2}]g2g_{-2}\;\;\;\;\;\;\;\;[β1][\beta_{-1}]g1g_{-1}\;\;\;\;\;\;\;\;[β0][\beta_{0}]g0g_{0}\;\;\;\;\;\;\;\;[β1][\beta_{1}]g1g_{1}\;\;\;\;\;\;\;\;[α2][\alpha_{-2}]f2f_{-2}\;\;\;\;\;\;\;\;[α1][\alpha_{-1}]f1f_{-1}\;\;\;\;\;\;\;\;[α0][\alpha_{0}]f0f_{0}\;\;\;\;\;\;\;\;\;[α1][\alpha_{1}]f1f_{1}\;\;\;\;\;\;\;\;

Consider the object C𝖧𝗈()𝐙C^{\prime}\in\mathsf{Ho}(\mathcal{M})^{\mathbf{Z}} given by one of the two diagonal zig-zags of the diagram above, namely, let

C=f2𝗁B(1)g1𝗁A(0)f0𝗁B(1)g1𝗁A(2)f2C^{\prime}=\;\;\;\cdots\xrightarrow{f_{-2}}\mathsf{h}B(-1)\xrightarrow{g_{-1}}\mathsf{h}A(0)\xrightarrow{f_{0}}\mathsf{h}B(1)\xrightarrow{g_{1}}\mathsf{h}A(2)\xrightarrow{f_{2}}\cdots

Using 3.4, construct C𝐙C\in\mathcal{M}^{\mathbf{Z}} such that 𝗁CC\mathsf{h}C\cong C^{\prime}.

Now, by construction, we have that 𝗁(𝖾(A))=𝖾(𝗁A)=𝖾(C)𝖾(𝗁C)=𝗁(𝖾(C))\mathsf{h}(\mathsf{e}^{*}(A))=\mathsf{e}^{*}(\mathsf{h}A)=\mathsf{e}^{*}(C^{\prime})\cong\mathsf{e}^{*}(\mathsf{h}C)=\mathsf{h}(\mathsf{e}^{*}(C)), so from 3.4 it follows that 𝖾(A)𝖾(C)\mathsf{e}^{*}(A)\simeq\mathsf{e}^{*}(C). Similarly, we have 𝗈(B)𝗈(C)\mathsf{o}^{*}(B)\simeq\mathsf{o}^{*}(C). Since AA is (1,0)(1,0)-interleaved with 𝖾(A)\mathsf{e}^{*}(A), 𝖾(C)\mathsf{e}^{*}(C) is 11-interleaved with 𝗈(C)\mathsf{o}^{*}(C), and 𝗈(B)\mathsf{o}^{*}(B) is (0,1)(0,1)-interleaved with BB, 2.3 implies that AA and BB are 22-homotopy interleaved, concluding the proof for the case m=1m=1.

The proof for general m1𝐙m\geq 1\in\mathbf{Z} is analogous, replacing the functor 𝖾:𝐙𝐙\mathsf{e}:\mathbf{Z}\to\mathbf{Z} with 𝖾m:𝐙𝐙\mathsf{e}_{m}:\mathbf{Z}\to\mathbf{Z} given by 𝖾m(n)=𝖾(n//m)×m\mathsf{e}_{m}(n)=\mathsf{e}(n\mathbin{/\mkern-6.0mu/}m)\times m, the functor 𝗈:𝐙𝐙\mathsf{o}:\mathbf{Z}\to\mathbf{Z} with 𝗈m:𝐙𝐙\mathsf{o}_{m}:\mathbf{Z}\to\mathbf{Z} given by 𝗈m(n)=𝗈(n//m)×m\mathsf{o}_{m}(n)=\mathsf{o}(n\mathbin{/\mkern-6.0mu/}m)\times m, and C𝖧𝗈()𝐙C^{\prime}\in\mathsf{Ho}(\mathcal{M})^{\mathbf{Z}} with

C(n)={𝗁(𝖾m(A))(n) if n//m is even 𝗁(𝗈m(B))(n) if n//m is odd,C^{\prime}(n)=\begin{cases}\mathsf{h}(\mathsf{e}_{m}^{*}(A))(n)&\text{ if $n\mathbin{/\mkern-6.0mu/}m$ is even }\\ \mathsf{h}(\mathsf{o}_{m}^{*}(B))(n)&\text{ if $n\mathbin{/\mkern-6.0mu/}m$ is odd},\end{cases}

where n//mn\mathbin{/\mkern-6.0mu/}m denotes the largest integer ll such that l×mnl\times m\leq n. ∎

We are now ready to prove the main result of this section.

Theorem A.

Let \mathcal{M} be a cofibrantly generated model category, let X,Y𝐑X,Y\in\mathcal{M}^{\mathbf{R}}, and let δ>0𝐑\delta>0\in\mathbf{R}. If XX and YY are δ\delta-homotopy commutative interleaved, then they are cδc\delta-homotopy interleaved for every c>2c>2.

Proof.

Let c>2c>2 and let m1𝐙m\geq 1\in\mathbf{Z} be large enough so that (2m+2)/mc(2m+2)/m\leq c. By 3.1, we may assume that X,Y𝐑X,Y\in\mathcal{M}^{\mathbf{R}} are mm-homotopy commutative interleaved and that we must show that they are cmcm-homotopy interleaved. Since 2m+2mc2m+2\leq mc, 3.2 reduces the problem to showing that i(X)i^{*}(X) and i(Y)i^{*}(Y) are 2m2m-homotopy interleaved in 𝐙\mathcal{M}^{\mathbf{Z}}, knowing that they are mm-homotopy commutative interleaved. 3.5 now finishes the proof. ∎

3.2 Lower bound

A implies that we have dHIcdHCd_{HI}\leq cd_{HC} as distances on 𝐑\mathcal{M}^{\mathbf{R}}, for c=2c=2 and for every cofibrantly generated model category \mathcal{M}. One could wonder if the constant c=2c=2 can be improved. In this section we show that, when =SS\mathcal{M}=\SS, we have c3/2c\geq 3/2. We do this by characterizing three-object persistent spaces which are 11-homotopy interleaved with a trivial persistent space in terms of the vanishing of a Toda bracket. The idea of using Toda brackets to prove that dHIdHCd_{HI}\neq d_{HC} is suggested in [2, Example 7.3].

The Toda bracket is an operation on composable triples of homotopy classes of pointed maps, and was originally defined to compute homotopy groups of spheres ([21]). We are interested in the use of Toda brackets as an algebraic obstruction to the rectification of diagrams. We now describe the fundamental procedure involved in the definition of Toda brackets, and the few properties that we are interested in (see, e.g., [1]).

Let SS\SS_{\bullet} denote the category of pointed spaces. For concreteness, in the arguments of this section we use SS=𝐓𝐨𝐩\SS=\mathbf{Top}. Let [3][3] denote the category freely generated by the graph \bullet\to\bullet\to\bullet\to\bullet. A diagram X𝖧𝗈(SS)[3]X\in\mathsf{Ho}(\SS_{\bullet})^{[3]}, which is given by X(0),X(1),X(2),X(3)𝖧𝗈(SS)X(0),X(1),X(2),X(3)\in\mathsf{Ho}(\SS_{\bullet}) and homotopy classes of pointed maps [f0]:X(0)X(1)[f_{0}]:X(0)\to X(1), [f1]:X(1)X(2)[f_{1}]:X(1)\to X(2), and [f2]:X(2)X(3)[f_{2}]:X(2)\to X(3), is a bracket sequence if [f1][f0][f_{1}]\circ[f_{0}] and [f2][f1][f_{2}]\circ[f_{1}] are equal to the null map, that is, to the homotopy class of the constant pointed map.

Let X𝖧𝗈(SS)[3]X^{\prime}\in\mathsf{Ho}(\SS_{\bullet})^{[3]} be a bracket sequence and let XSS[3]X\in\SS_{\bullet}^{[3]} be such that 𝗁XX\mathsf{h}X\cong X^{\prime}, which exists by 3.4. We can, and do, assume that XX takes values in CW-complexes. Consider the following diagram of pointed spaces and pointed maps:

X(0){X(0)}X(1){X(1)}{\ast}{\ast}X(2){X(2)}X(3).{X(3).}f0f_{0}f1f_{1}f2f_{2}

Since XX^{\prime} is a bracket sequence, we know that there exist (pointed) homotopies filling the squares in the diagram above. For YY a pointed space, let CYCY denote its reduced cone. Each pair of such homotopies gives us pointed maps α:CX(0)X(2)\alpha:CX(0)\to X(2) and β:CX(1)X(3)\beta:CX(1)\to X(3) such that αi=f1f0:X(0)X(2)\alpha\circ i=f_{1}\circ f_{0}:X(0)\to X(2) and βi=f2f1\beta\circ i=f_{2}\circ f_{1}, where ii is the inclusion into the cone. In particular, we have a commutative square

(3.1) X(0){X(0)}CX(1){CX(1)}CX(0){CX(0)}X(3),{X(3),}if0i\circ f_{0}iif2αf_{2}\circ\alphaβ\beta

which, by noticing that the pushout of the top and left morphisms is a model for the reduced suspension of X(0)X(0), gives us an element of [ΣX(0),X(3)][\Sigma X^{\prime}(0),X^{\prime}(3)], where [,][-,-] denotes homotopy classes of pointed maps.

Definition 3.6.

Let X𝖧𝗈(SS)[3]X\in\mathsf{Ho}(\SS_{\bullet})^{[3]} be a bracket sequence. Consider the subset of [ΣX(0),X(3)][\Sigma X(0),X(3)] consisting of all elements that can be obtained using the procedure above. This is the Toda bracket of XX. We say that the Toda bracket vanishes if it contains the null map.

It is well-known (see, e.g., [1, Section 1]) that the non-vanishing of a Toda bracket is an obstruction to the rectification of the bracket sequence, in the following sense.

Proposition 3.7.

The Toda bracket of a bracket sequence X𝖧𝗈(SS)[3]X^{\prime}\in\mathsf{Ho}(\SS_{\bullet})^{[3]} vanishes if and only if there exists XSS[3]X\in\SS_{\bullet}^{[3]} with 𝗁XX\mathsf{h}X\cong X^{\prime} and with f1f0f_{1}\circ f_{0} and f2f1f_{2}\circ f_{1} equal to the null map.∎

Although Toda brackets are defined for diagrams of pointed spaces, one can extend them to unpointed spaces, provided the spaces are simply connected. This is what we do now. A simply connected space is a non-empty, connected space whose fundamental groupoid is trivial. Let SS𝗌𝖼\SS_{\mathsf{sc}} and SS𝗌𝖼,\SS_{\mathsf{sc},\bullet} denote the categories of simply connected spaces and of pointed, simply connected spaces, respectively. We have the following well-known fact and corollary.

Lemma 3.8.

The forgetful functor U:𝖧𝗈(SS𝗌𝖼,)𝖧𝗈(SS𝗌𝖼)U:\mathsf{Ho}(\SS_{\mathsf{sc},\bullet})\to\mathsf{Ho}(\SS_{\mathsf{sc}}) is an equivalence of categories.∎

Corollary 3.9.

If X𝖧𝗈(SS𝗌𝖼,)[3]X\in\mathsf{Ho}(\SS_{\mathsf{sc},\bullet})^{[3]} is such that the composite of consecutive maps of U(X)U_{*}(X) are null-homotopic, then XX is a bracket sequence.

Let X,X𝖧𝗈(SS𝗌𝖼,)[3]X,X^{\prime}\in\mathsf{Ho}(\SS_{\mathsf{sc},\bullet})^{[3]} be such that U(X)U(X)U_{*}(X)\cong U_{*}(X^{\prime}). Then XX is a bracket sequence if and only if XX^{\prime} is; in that case, the Toda bracket of XX vanishes if and only if the Toda bracket of XX^{\prime} does. ∎

3.9 implies that, for X𝖧𝗈(SS𝗌𝖼)[3]X\in\mathsf{Ho}(\SS_{\mathsf{sc}})^{[3]}, there is a well-defined notion of XX being a bracket sequence, namely that any lift X𝖧𝗈(SS𝗌𝖼,)[3]X^{\prime}\in\mathsf{Ho}(\SS_{\mathsf{sc},\bullet})^{[3]} is a bracket sequence; in that case, we say that the Toda bracket of XX vanishes if the Toda bracket of XX^{\prime} does.

Let 𝗃:SS[3]SS𝐙\mathsf{j}:\SS^{[3]}\to\SS^{\mathbf{Z}} be given by extending XSS[3]X\in\SS^{[3]} to the right with the singleton space and to the left with the empty space. Let SS[3]\ast\in\SS^{[3]} be the constant singleton space.

Proposition 3.10.

Let XSS𝗌𝖼[3]X\in\SS_{\mathsf{sc}}^{[3]}, then

  1. 1.

    𝗁(𝗃(X))𝖧𝗈(SS)𝐙\mathsf{h}(\mathsf{j}(X))\in\mathsf{Ho}(\SS)^{\mathbf{Z}} is 11-interleaved with 𝗁(𝗃())\mathsf{h}(\mathsf{j}(\ast)) if and only if 𝗁X\mathsf{h}X is a bracket sequence;

  2. 2.

    if 𝗁(𝗃(X))𝖧𝗈(SS)𝐙\mathsf{h}(\mathsf{j}(X))\in\mathsf{Ho}(\SS)^{\mathbf{Z}} is 11-interleaved with 𝗁(𝗃())\mathsf{h}(\mathsf{j}(\ast)), then 𝗃(X)\mathsf{j}(X) is 11-homotopy interleaved with 𝗃()\mathsf{j}(\ast) if and only if the Toda bracket of 𝗁X\mathsf{h}X vanishes.

Proof.

Statement 1 follows directly from 3.9. For 2, note that if the Toda bracket of 𝗁X\mathsf{h}X vanishes, then, by 3.7, there exists XSS𝗌𝖼,[3]X^{\prime}\in\SS_{\mathsf{sc},\bullet}^{[3]} such that 𝗁X𝗁X\mathsf{h}X^{\prime}\cong\mathsf{h}X and such that the composite of consecutive maps of XX^{\prime} are null maps. In particular, 𝗃(X)\mathsf{j}(X^{\prime}) is 11-interleaved with 𝗃()\mathsf{j}(\ast), and, since 𝗁(𝗃(X))𝗁(𝗃(X))\mathsf{h}(\mathsf{j}(X^{\prime}))\cong\mathsf{h}(\mathsf{j}(X)), we have that 𝗃(X)𝗃(X)\mathsf{j}(X^{\prime})\simeq\mathsf{j}(X) by 3.4, so 𝗃(X)\mathsf{j}(X) and 𝗃()\mathsf{j}(\ast) are 11-homotopy interleaved.

For the converse of 2, assume that 𝗃(X)\mathsf{j}(X) and 𝗃()\mathsf{j}(\ast) are 11-homotopy interleaved. It follows that there exists a commutative diagram of pointed spaces and pointed maps

X(0){X^{\prime}(0)}X(1){X^{\prime}(1)}B{B}A{A}X(2){X^{\prime}(2)}X(3),{X^{\prime}(3),}f0f^{\prime}_{0}f1f^{\prime}_{1}f2f^{\prime}_{2}

with AA and BB contractible and XSS[3]X^{\prime}\in\SS_{\bullet}^{[3]} such that XXX^{\prime}\simeq X, as diagrams of unpointed spaces. It suffices to show that the Toda bracket of XX^{\prime} vanishes. For this, note that, using the diagonal morphism ABA\to B, we can find maps α:CX(0)X(2)\alpha:CX^{\prime}(0)\to X^{\prime}(2) and β:CX(1)X(3)\beta:CX^{\prime}(1)\to X^{\prime}(3) such that βCf1=f2α\beta\circ Cf^{\prime}_{1}=f^{\prime}_{2}\circ\alpha. In particular, in this case, there is a diagonal filler for the square (3.1) and thus the induced map ΣX(0)X(3)\Sigma X^{\prime}(0)\to X^{\prime}(3) is null-homotopic, as required. ∎

The following lemma is clear.

Lemma 3.11.

Let X,YC𝐙X,Y\in C^{\mathbf{Z}}. If i(X),i(Y)C𝐑i_{*}(X),i_{*}(Y)\in C^{\mathbf{R}} are rr-interleaved for some 0r<3/20\leq r<3/2, then X,YC𝐙X,Y\in C^{\mathbf{Z}} are 11-interleaved.∎

We are now ready to prove the lower bound.

Proposition 3.12.

Let =SS\mathcal{M}=\SS. If dHIcdHCd_{HI}\leq cd_{HC} then c3/2c\geq 3/2.

Proof.

By 3.11 and 3.10, it suffices to find a bracket sequence X𝖧𝗈(SS)[3]X\in\mathsf{Ho}(\SS_{\bullet})^{[3]} valued in simply connected spaces such that its Toda bracket does not vanish. Examples of this are given in [21]. A classical example, referenced in [2], is S4S4S3S3S^{4}\to S^{4}\to S^{3}\to S^{3} with the first and last maps degree 22 maps, and the middle map the suspension of the Hopf map. ∎

Remark 3.13.

3.12 implies in particular that dHIdHCd_{HI}\neq d_{HC}. As mentioned in the introduction, we know that we have dHIdIHCdHCd_{HI}\geq d_{IHC}\geq d_{HC}, so it is natural to wonder whether we have dHIdIHCd_{HI}\neq d_{IHC} or dIHCdHCd_{IHC}\neq d_{HC}, or both. We leave these as open questions.

3.3 Impossibility of rectification in higher dimensions

In this section, we show that A has no analogue for multi-persistent spaces; we thank Alex Rolle for pointing this out to us. We prove this for m=2m=2 and remark that a similar argument works for m>2m>2.

Proposition 3.14.

If m=2m=2, there is no constant c>0𝐑c>0\in\mathbf{R} such that for all δ>0𝐑m\delta>0\in\mathbf{R}^{m}, if X,YSS𝐑mX,Y\in\SS^{\mathbf{R}^{m}} are δ\delta-homotopy commutative interleaved, then they are cδc\delta-homotopy interleaved.

Let 𝐬𝐪\mathbf{sq} denote the subposet of 𝐑2\mathbf{R}^{2} spanned by {(0,0),(0,1),(1,0),(1,1)}\{(0,0),(0,1),(1,0),(1,1)\}, so that a functor 𝐬𝐪C\mathbf{sq}\to C from 𝐬𝐪\mathbf{sq} to a category CC corresponds to a commutative square in CC. We will use the following well-known fact, which says that a homotopy commutative diagram can have different, non-equivalent lifts. For a specific instance see, e.g., [10].

Lemma 3.15.

There exist A,B:𝐬𝐪SSA,B:\mathbf{sq}\to\SS such that 𝗁A𝗁B𝖧𝗈(SS)𝐬𝐪\mathsf{h}A\cong\mathsf{h}B\in\mathsf{Ho}(\SS)^{\mathbf{sq}} and such that A≄BA\not\simeq B.∎

Proof of 3.14.

Given a diagram A:𝐬𝐪SSA:\mathbf{sq}\to\SS, consider the bi-persistent space A:𝐑2SSA^{\prime}:\mathbf{R}^{2}\to\SS such that A(r,s)=A^{\prime}(r,s)=\emptyset whenever rr or ss are negative, A(r,s)=A(r,s)A^{\prime}(r,s)=A(\lfloor r\rfloor,\lfloor s\rfloor) whenever 0r,s<20\leq r,s<2, and A(r,s)A^{\prime}(r,s) is the singleton space whenever 0r,s0\leq r,s and 2max(r,s)2\leq\max(r,s). Let A,B:𝐬𝐪SSA,B:\mathbf{sq}\to\SS. Note that if (0,0)δ<(1/2,1/2)𝐑2(0,0)\leq\delta<(1/2,1/2)\in\mathbf{R}^{2} and A,BSS𝐑2A^{\prime},B^{\prime}\in\SS^{\mathbf{R}^{2}} are δ\delta-homotopy interleaved, then we have ABA\simeq B.

To prove the result, it is enough to show that there exist bi-persistent spaces X,YSS𝐑2X,Y\in\SS^{\mathbf{R}^{2}} that are 0-homotopy commutative interleaved, i.e., such that 𝗁X𝗁Y\mathsf{h}X\cong\mathsf{h}Y, which are not δ\delta-homotopy interleaved for any 0δ<(1/2,1/2)𝐑20\leq\delta<(1/2,1/2)\in\mathbf{R}^{2}. In order to do this, we can let AA and BB be as in 3.15 and take X=AX=A^{\prime} and Y=BY=B^{\prime}. ∎

4 Projective cofibrant persistent simplicial sets

The purpose of this section is to characterize projective cofibrant persistent simplicial sets as filtered simplicial sets (4.5). We work with simplicial sets indexed by an arbitrary poset (P,)(P,\leq).

Definition 4.1.

A 𝐏P-filtered simplicial set (filtered simplicial set when there is no risk of confusion) is a simplicial set XX equipped with functions βn:XnP\beta_{n}:X_{n}\to P, satisfying:

  • βn1(di(σ))βn(σ)\beta_{n-1}(d_{i}(\sigma))\leq\beta_{n}(\sigma) for every n1n\geq 1, σXn\sigma\in X_{n}, and boundary map di:XnXn1d_{i}:X_{n}\to X_{n-1};

  • βn+1(si(σ))βn(σ)\beta_{n+1}(s_{i}(\sigma))\leq\beta_{n}(\sigma) for every n0n\geq 0, σXn\sigma\in X_{n}, and degeneracy map si:XnXn+1s_{i}:X_{n}\to X_{n+1}.

When there is no risk of ambiguity, we denote the filtered simplicial set (X,β)(X,\beta) simply by XX.

Definition 4.2.

Given a filtered simplicial set (X,β)(X,\beta) define a persistent simplicial set (X,β)^𝐬𝐒𝐞𝐭P\widehat{(X,\beta)}\in\mathbf{sSet}^{P} such that for rPr\in P we have (X,β)^(r)n={σXn:βn(σ)r}\widehat{(X,\beta)}(r)_{n}=\left\{\sigma\in X_{n}:\beta_{n}(\sigma)\leq r\right\}, with faces and degeneracies given by restricting the ones of XX.

By a standard abuse of language, We say that a persistent simplicial set is a filtered simplicial set if it is isomorphic to Y^\widehat{Y}, for YY a filtered simplicial set.

The following result is a characterization of filtered simplicial sets among persistent simplicial sets by means of easily verified point-set conditions.

Lemma 4.3.

A persistent simplicial set X𝐬𝐒𝐞𝐭PX\in\mathbf{sSet}^{P} is a filtered simplicial set if and only if the following conditions are satisfied:

  1. 1.

    The structure morphism X(r)X(r)X(r)\to X(r^{\prime}) is a monomorphism for every rrr\leq r^{\prime} in PP. In particular, up to isomorphism, we may, and do, assume that X(r)X(r) is a subsimplicial set of X(r)X(r^{\prime}).

  2. 2.

    For every simplex σrPX(r)\sigma\in\bigcup_{r\in P}X(r), the set {tP:σX(t)}\{t\in P\ :\ \sigma\in X(t)\} has a minimum.

Proof.

The only non-trivial part is that if XX satisfies the two conditions in the statement then it is filtered. Set Y=rPX(r)Y=\bigcup_{r\in P}X(r), which makes sense thanks to condition 1. Given σYn\sigma\in Y_{n}, define βn(σ):=min{tP:σX(t)}\beta_{n}(\sigma):=\min\{t\in P\ :\ \sigma\in X(t)\}, which is well-defined thanks to condition 2. We then have XY^X\cong\widehat{Y}. The rest of the proof is clear. ∎

The proof of the following lemma a straightforward application of Lemma 4.3. We use the term cell attachment to indicate any pushout of a generating cofibration rΔnrΔnr\odot\partial\Delta^{n}\to r\odot\Delta^{n}.

Lemma 4.4.
  1. 1.

    A retract of a filtered simplicial set is filtered.

  2. 2.

    If the domain of a cell attachment is a filtered simplicial set, then the codomain is too.

  3. 3.

    Let ζ\zeta be a limit ordinal and let X:ζ𝐬𝐒𝐞𝐭PX_{\bullet}:\zeta\to\mathbf{sSet}^{P} be a diagram of persistent simplicial sets, where for each γ<ζ\gamma<\zeta we have that the map XγXγ+1X_{\gamma}\to X_{\gamma+1} is a cell attachment. If XγX_{\gamma} is a filtered simplicial set for every γ<ζ\gamma<\zeta, then Xζ=colimγ<ζXγX_{\zeta}=\operatorname{colim}_{\gamma<\zeta}X_{\gamma} is a filtered simplicial set.∎

The recognition principle for projective cofibrant persistent simplicial sets is now a consequence of 4.4 and the fact that the cofibrant objects in a cofibrantly generated model category are precisely the retracts of transfinite compositions of cell attachments ([12, Proposition 2.1.18(b)]).

Proposition 4.5.

A persistent simplicial set is filtered if and only if it is projective cofibrant.∎

In practice, many of the persistent spaces relevant to Topological Data Analysis are filtered simplicial sets.

Example 4.6.

The Vietoris–Rips complex associated to a metric space (X,dX)(X,d_{X}), usually defined to be a persistent simplicial complex, can be turned into a persistent simplicial set by choosing a total order on XX. It follows directly from its definition that this persistent simplicial set is filtered. Other examples of this form include the Čech complex and the filtrations of [6].

An example of a filtered multi-persistent simplicial set is the following. Given a metric space (X,dX)(X,d_{X}) together with a real-valued function f:X𝐑f:X\to\mathbf{R}, one can construct a bi-filtered simplicial set as follows. For each s𝐑s\in\mathbf{R}, consider Xs=f1(,s]X_{s}=f^{-1}(-\infty,s] and let Fs,rF_{s,r} be the Vietoris–Rips complex of XsX_{s} at scale rr.

We remark that persistent simplicial sets whose structure maps are monomorphism are not necessarily filtered. This happens in practice when the same simplex “appears at different times”, that is, when condition 2 in 4.3 is not satisfied. Examples of this include the degree-Rips bi-filtration ([16]), and Vietoris–Rips applied to the kernel density filtration of [19].

5 Interleaving in 𝖧𝗈(SS𝐑m)\mathsf{Ho}\left(\SS^{\mathbf{R}^{m}}\right) and in homotopy groups

In this section, we prove B. We start by defining the notions of persistent homotopy groups of a persistent space, and of morphism inducing an interleaving in persistent homotopy groups. The notion of persistent homotopy group we use is essentially the same as that of Jardine ([13]).

We model the nnth homotopy group πn(W,w)\pi_{n}(W,w) of a pointed space (W,w)(W,w) by the set of pointed homotopy classes of pointed maps from the nnth dimensional sphere SnS^{n} into WW.

Definition 5.1.

Let XSS𝐑mX\in\SS^{\mathbf{R}^{m}}. The persistent set π0(X):𝐑m𝐒𝐞𝐭\pi_{0}(X):\mathbf{R}^{m}\to\mathbf{Set} is defined by π0X\pi_{0}\circ X. Let n1n\geq 1, r𝐑mr\in\mathbf{R}^{m}, and xX(r)x\in X(r). The 𝐧nth persistent homotopy group of XX based at xx is the persistent group πn(X,x):𝐑m𝐆𝐫𝐩\pi_{n}(X,x):\mathbf{R}^{m}\to\mathbf{Grp} that is trivial at srs\ngeq r, and that is πn(X(s),φr,sX(x))𝐆𝐫𝐩\pi_{n}(X(s),\varphi^{X}_{r,s}(x))\in\mathbf{Grp} at srs\geq r.

Note that πn\pi_{n} is functorial for every nn\in\mathbb{N}.

Definition 5.2.

Let ε,δ0𝐑m\varepsilon,\delta\geq 0\in\mathbf{R}^{m}. Assume given a homotopy class of morphisms [f]:XYε𝖧𝗈(SS𝐑m)[f]:X^{\prime}\to Y^{\prime\varepsilon}\in\mathsf{Ho}\left(\SS^{\mathbf{R}^{m}}\right). Let XXX^{\prime}\simeq X be a cofibrant replacement, let YYY^{\prime}\simeq Y be a fibrant replacement, and let f:XεYf:X\to_{\varepsilon}Y be a representative of ff. We say that [f][f] induces an (𝛆,𝛅)(\varepsilon,\delta)-interleaving in homotopy groups if the induced map π0(f):π0(X)επ0(Y)\pi_{0}(f):\pi_{0}(X)\to_{\varepsilon}\pi_{0}(Y) is part of an (ε,δ)(\varepsilon,\delta)-interleaving of persistent sets, and if for every r𝐑mr\in\mathbf{R}^{m}, every xX(r)x\in X(r), and every n1n\geq 1\in\mathbb{N}, the induced map πn(f):πn(X,x)επn(Y,f(x))\pi_{n}(f):\pi_{n}(X,x)\to_{\varepsilon}\pi_{n}(Y,f(x)) is part of an (ε,δ)(\varepsilon,\delta)-interleaving of persistent groups.

It is clear that the definition above is independent of the choices of representatives.

A standard result in classical homotopy theory is that a fibration of Kan complexes inducing an isomorphism in all homotopy groups has the right lifting property with respect to cofibrations ([9, Theorem I.7.10]). An analogous, persistent, result (5.13), says that, for a fibration of fibrant objects inducing a δ\delta-interleaving in homotopy groups, the lift exists up to a shift, which depends on both δ\delta and on a certain “length” nn\in\mathbb{N} associated to the cofibration. To make this precise, we introduce the notion of nn-dimensional extension.

Definition 5.3.

Let A,BSS𝐑mA,B\in\SS^{\mathbf{R}^{m}} and let nn\in\mathbb{N}. A map j:ABj:A\to B is a 𝐧n-dimensional extension (of AA) if there exists a set II, a family of tuples of real numbers {ri𝐑m}iI\left\{r_{i}\in\mathbf{R}^{m}\right\}_{i\in I}, and commutative squares of the form depicted on the left below, that together give rise to the pushout square on the right below. Here, DnDn\partial D^{n}\hookrightarrow D^{n} stands for Sn1DnS^{n-1}\hookrightarrow D^{n} if SS=𝐓𝐨𝐩\ \SS=\mathbf{Top}, and for ΔnΔn\partial\Delta^{n}\hookrightarrow\Delta^{n} if SS=𝐬𝐒𝐞𝐭\ \SS=\mathbf{sSet}.

Dn{\partial D^{n}}A(ri){A(r_{i})}iIri(Dn){\coprod_{i\in I}r_{i}\odot(\partial D^{n})}A{A}Dn{D^{n}}B(ri){B(r_{i})}iri(Dn){\coprod_{i}r_{i}\odot(D^{n})}B{B}fi\scriptstyle{f_{i}}jri\scriptstyle{j_{r_{i}}}f\scriptstyle{f}j\scriptstyle{j}gi\scriptstyle{g_{i}}g\scriptstyle{g}

A single dimensional extension is an nn-dimensional extension for some nn\in\mathbb{N}.

Definition 5.4.

Let ι:AB\iota:A\to B be a projective cofibration of SS𝐑m\SS^{\mathbf{R}^{m}} and let nn\in\mathbb{N}. We say that ι\iota is an 𝐧n-cofibration if it factors as the composite of n+1n+1 maps f0,,fnf_{0},\dots,f_{n}, with fif_{i} an nin_{i}-dimensional extension for some nin_{i}\in\mathbb{N}. We say that ASS𝐑mA\in\SS^{\mathbf{R}^{m}} is nn-cofibrant if the map A\emptyset\to A is an nn-cofibration.

The next lemma, which follows directly from 4.5, gives a rich family of examples of nn-cofibrant persistent simplicial sets. Recall that a simplicial set is nn-skeletal if all its simplices in dimensions above nn are degenerate.

Lemma 5.5.

Let A𝐬𝐒𝐞𝐭𝐑mA\in\mathbf{sSet}^{\mathbf{R}^{m}} and let nn\in\mathbb{N}. If AA is projective cofibrant and pointwise nn-skeletal, then it is nn-cofibrant.∎

Example 5.6.

The Vietoris–Rips complex 𝖵𝖱(X)\mathsf{VR}(X) of a metric space XX, as defined in 4.6, is nn-cofibrant if the underlying set of XX has finite cardinality |X|=n+1|X|=n+1.

If one is interested in persistent (co)homology of some bounded degree nn, then one can restrict computations to the (n+1)(n+1)-skeleton of a Vietoris–Rips complex, which is (n+1)(n+1)-cofibrant.

A result analogous to 5.5, but for persistent topological spaces, does not hold, as cells are not necessarily attached in order of dimension. This motivates the following definition.

Definition 5.7.

Let nn\in\mathbb{N}. A persistent topological space A𝐓𝐨𝐩𝐑mA\in\mathbf{Top}^{\mathbf{R}^{m}} is an 𝐧n-dimensional persistent CW-complex if the map X\emptyset\to X can be factored as a composite of maps f0,,fnf_{0},\dots,f_{n}, with fif_{i} an ii-dimensional extension.

Example 5.8.

The geometric realization of any nn-cofibrant persistent simplicial set is an nn-dimensional persistent CW-complex.

Lemma 5.9.

Every nn-dimensional persistent CW-complex is nn-cofibrant.∎

We now make precise the notion of lifting property up to a shift.

Definition 5.10.

Let i:ABi:A\to B and p:YXp:Y\to X be morphisms in SS𝐑m\SS^{\mathbf{R}^{m}} and let δ0\delta\geq 0. We say that pp has the right 𝛅\delta-lifting property with respect to ii if for all morphisms AYA\to Y and BXB\to X making the square on the left below commute, there exists a diagonal δ\delta-morphism f:BδYf:B\to_{\delta}Y rendering the diagram commutative. Below, the diagram on the left is shorthand for the one on the right.

A{A}Y{Y}A{A}Y{Y}Yδ{Y^{\delta}}B{B}X{X}B{B}X{X}Xδ.{X^{\delta}.}iiδ{\scriptscriptstyle\delta}\;\;\,ppiipp𝖲0,δ(𝗂𝖽Y)\mathsf{S}_{0,\delta}(\mathsf{id}_{Y})pδp^{\delta}      𝖲0,δ(𝗂𝖽X)\mathsf{S}_{0,\delta}(\mathsf{id}_{X})

We now prove 5.12, an adaptation of a result of Jardine, which says that fibrations inducing interleavings in homotopy groups have a shifted right lifting property, as defined above. The main difference is that we work in the multi-persistent setting. We use simplicial notation and observe that the corresponding statement for persistent topological spaces follows from the simplicial one by using the singular complex-realization adjunction. We recall a standard, technical lemma whose proof is given within that of, e.g., [9, Theorem I.7.10].

Lemma 5.11.

Suppose given a commutative square of simplicial sets

(5.1) Δn{\partial\Delta^{n}}X{X}Δn{\Delta^{n}}Y,{Y,}α\scriptstyle{\alpha}p\scriptstyle{p}β\scriptstyle{\beta}

where pp is a Kan fibration between Kan complexes. If there is commutative diagram like the one on the left below, for which the lifting problem on the right admits a solution, then the initial square (5.1) admits a solution.

Δn{\partial\Delta^{n}}Δn×Δ1{\partial\Delta^{n}\times\Delta^{1}}X{X}Δn{\partial\Delta^{n}}X{X}Δn×Δ1{\Delta^{n}\times\Delta^{1}}Y{Y}Δn{\Delta^{n}}Y{Y}Δn{\Delta^{n}}(𝗂𝖽Δn×{1})\scriptstyle{(\mathsf{id}_{\partial\Delta^{n}}\times\{1\})}α\scriptstyle{\alpha}h\scriptstyle{h}p\scriptstyle{p}h(𝗂𝖽Δn×{0})\scriptstyle{h\circ(\mathsf{id}_{\partial\Delta^{n}}\times\{0\})}p\scriptstyle{p}g\scriptstyle{g}g(𝗂𝖽Δn×{0})\scriptstyle{g\circ(\mathsf{id}_{\Delta^{n}}\times\{0\})}(𝗂𝖽Δn×{1})\scriptstyle{(\mathsf{id}_{\Delta^{n}}\times\{1\})}β\scriptstyle{\beta}
Lemma 5.12 (cf. [13, Lemma 14]).

Let δ0\delta\geq 0, and let f:XYSS𝐑mf:X\to Y\in\SS^{\mathbf{R}^{m}} induce a (0,δ)(0,\delta)-interleaving in homotopy groups. If XX and YY are projective fibrant and ff is a projective fibration, then ff has the right 2δ2\delta-lifting property with respect to boundary inclusions rDnrDnr\odot\partial D^{n}\to r\odot D^{n}, for every r𝐑mr\in\mathbf{R}^{m} and every nn\in\mathbb{N}.

Proof.

Suppose given a commutative diagram as on the left below, which corresponds to the one on the right:

(5.2) rΔn{r\odot\partial\Delta^{n}}X{X}Δn{\partial\Delta^{n}}X(r){X(r)}rΔn{r\odot\Delta^{n}}Y{Y}Δn{\Delta^{n}}Y(r).{Y(r).}a\scriptstyle{a}p\scriptstyle{p}α\scriptstyle{\alpha}pr\scriptstyle{p_{r}}b\scriptstyle{b}β\scriptstyle{\beta}

We must find a 2δ2\delta-lift for the diagram on the right. The proof strategy is to appeal to 5.11 to simplify α\alpha, then prove that at the cost of a δ\delta-shift we can further reduce α\alpha to a constant map, and then show that the simplified lifting problem can be solved at the cost of another δ\delta-shift. So we end up with a 2δ2\delta-lift, as in the statement. We proceed by proving the claims in opposite order.

We start by showing that (5.2) can be solved up to a δ\delta-shift whenever α\alpha is constant. Let us assume that α\alpha is of the form α=\alpha=\ast for some X(r)0\ast\in X(r)_{0}. Since, then, β\beta represents an element [β]πn(Y(r),)[\beta]\in\pi_{n}(Y(r),\ast), there exists a map α:ΔnX(r+δ)\alpha^{\prime}:\Delta^{n}\to X(r+\delta) whose restriction to Δn\partial\Delta^{n} is constant on X(r)0\ast\in X(r)_{0}, and such that there is a homotopy h:βpαh:\beta\simeq p\alpha^{\prime} relative to Δn\partial\Delta^{n}. We can thus consider

Δn{\partial\Delta^{n}}(Δn×Δ1)(Δn×{0}){\left(\partial\Delta^{n}\times\Delta^{1})\cup(\Delta^{n}\times\{0\}\right)}X(r+δ){X(r+\delta)}Δn{\Delta^{n}}Δn×Δ1{\Delta^{n}\times\Delta^{1}}Y(r+δ),{Y(r+\delta),}𝗂𝖽Δn×{1}\scriptstyle{\mathsf{id}_{\partial\Delta^{n}}\times\{1\}}i\scriptstyle{i}(,α)\scriptstyle{(\ast{,}\alpha^{\prime})}pr+δ\scriptstyle{p_{r+\delta}}𝗂𝖽Δn×{1}\scriptstyle{\mathsf{id}_{\Delta^{n}}\times\{1\}}h\scriptstyle{h}H\scriptstyle{H}

where HH is a diagonal filler for the right-hand side square, which exists since the middle vertical map is a trivial cofibration of simplicial sets and pr+δp_{r+\delta} is a Kan fibration by assumption. The composite map H𝗂𝖽Δn×{1}H\circ\mathsf{id}_{\Delta^{n}}\times\{1\} is a lift for (5.2).

We now assume that α\alpha is of a specific, simplified form, and prove that, up to a δ\delta-shift, we can reduce the lifting problem (5.2) to the case in which α\alpha is constant. Let us assume that di(α)=X(r)0d_{i}(\alpha)=\ast\in X(r)_{0} for every 0<in0<i\leq n, and set α0=d0(α)\alpha_{0}=d_{0}(\alpha). We have that α0\alpha_{0} represents an element [α0]πn1(X(r),)[\alpha_{0}]\in\pi_{n-1}(X(r),\ast), with the property that p[α0]=0πn1(Y(r),)p[\alpha_{0}]=0\in\pi_{n-1}(Y(r),\ast). Since pp induces a (0,δ)(0,\delta)-interleaving in homotopy groups, we have that φr,r+δX([α0])=0πn1(X(r+δ),)\varphi^{X}_{r,r+\delta}([\alpha_{0}])=0\in\pi_{n-1}(X(r+\delta),\ast), witnessed by a homotopy h0:Δn1×Δ1X(r+δ)h_{0}:\Delta^{n-1}\times\Delta^{1}\to X(r+\delta), constant on Δn1\partial\Delta^{n-1}. If we set hi=:Δn1X(r+δ)h^{\prime}_{i}=\ast:\Delta^{n-1}\to X(r+\delta) for every 0<in10<i\leq n-1 and h0=h0h^{\prime}_{0}=h_{0}, we get a map h:Δn×Δ1X(r+δ)h^{\prime}:\partial\Delta^{n}\times\Delta^{1}\to X(r+\delta). We can now extend (φr,r+δYβ,ph):(Δn×{1})(Δn×Δ1)Y(r+δ)(\varphi^{Y}_{r,r+\delta}\circ\beta{,}ph^{\prime}):(\Delta^{n}\times\{1\})\cup(\partial\Delta^{n}\times\Delta^{1})\to Y(r+\delta) to a homotopy H:Δn×Δ1Y(r+δ)H^{\prime}:\Delta^{n}\times\Delta^{1}\to Y(r+\delta). Now observe that the following lifting problem is such that h(𝗂𝖽Δn×{0})=h^{\prime}\circ(\mathsf{id}_{\partial\Delta^{n}}\times\{0\})=*, so, thanks to 5.11, we have reduced this case to the case in which α\alpha is constant.

Δn{\partial\Delta^{n}}X(r+δ){X(r+\delta)}Δn{\Delta^{n}}Y(r+δ).{Y(r+\delta).}h1(𝗂𝖽Δn×{0})\scriptstyle{h^{\prime}_{1}\circ(\mathsf{id}_{\partial\Delta^{n}}\times\{0\})}pr+δ\scriptstyle{p_{r+\delta}}H(𝗂𝖽Δn×{0}\scriptstyle{H^{\prime}\circ(\mathsf{id}_{\Delta^{n}}\times\{0\}}

To conclude, we must show that we can reduce the original lifting problem (5.2) to one in which all but the 0th faces of α\alpha are constant on a point X(r)0\ast\in X(r)_{0}. Let K:Λ0n×Δ1Λ0nK:\Lambda^{n}_{0}\times\Delta^{1}\to\Lambda^{n}_{0} be the homotopy that contracts the simplicial horn onto its vertex 0, which determines a diagram

Λ0n{\Lambda^{n}_{0}}Λ0n×Δ1{\Lambda^{n}_{0}\times\Delta^{1}}Λ0n{\Lambda^{n}_{0}}Δn{\partial\Delta^{n}}X{X}Δ0{\Delta^{0}}𝗂𝖽Λ0n×{1}\scriptstyle{\mathsf{id}_{\Lambda^{n}_{0}}\times\{1\}}k1\scriptstyle{k_{1}}𝗂𝖽Λ0n×{0}\scriptstyle{\mathsf{id}_{\Lambda^{n}_{0}}\times\{0\}}α\scriptstyle{\alpha}α(0)\scriptstyle{\alpha(0)}

with k1=αjKk_{1}=\alpha\circ j\circ K, with j:Λ0nΔnj:\Lambda_{0}^{n}\to\partial\Delta^{n} the inclusion of the horn into the boundary. We can now extend the map (α,k1):(Δn×{1})(Λ0n×Δ1)X(r)(\alpha{,}k_{1}):\left(\partial\Delta^{n}\times\{1\}\right)\cup\left(\Lambda^{n}_{0}\times\Delta^{1}\right)\to X(r) to a homotopy k:Δn×Δ1X(r)k:\partial\Delta^{n}\times\Delta^{1}\to X(r). Similarly, we extend the map (β,pk):(Δn×{1})(Δn×Δ1)Y(r)(\beta{,}p\circ k):\left(\Delta^{n}\times\{1\}\right)\cup\left(\partial\Delta^{n}\times\Delta^{1}\right)\to Y(r) to a homotopy g:Δn×Δ1Y(r)g:\Delta^{n}\times\Delta^{1}\to Y(r). It now suffices to consider the diagram

Δn{\partial\Delta^{n}}Δn×Δ1{\partial\Delta^{n}\times\Delta^{1}}X{X}Δn×Δ1{\Delta^{n}\times\Delta^{1}}Y{Y}Δn{\Delta^{n}}(𝗂𝖽Δn×{1})\scriptstyle{(\mathsf{id}_{\partial\Delta^{n}}\times\{1\})}α\scriptstyle{\alpha}k\scriptstyle{k}p\scriptstyle{p}g\scriptstyle{g}(𝗂𝖽Δn×{1})\scriptstyle{(\mathsf{id}_{\Delta^{n}}\times\{1\})}β\scriptstyle{\beta}

observing that α:=k|Δn×{0}\alpha^{\prime}:=k_{|\partial\Delta^{n}\times\{0\}} satisfies di(α)=d_{i}(\alpha^{\prime})=\ast for 0<in0<i\leq n, and appeal to 5.11. ∎

Corollary 5.13.

Let δ0\delta\geq 0 and let f:XδYf:X\to_{\delta}Y induce a δ\delta-interleaving in homotopy groups. If XX and YY are projective fibrant and ff is a projective fibration, then ff has the right (4(n+1)δ)(4(n+1)\delta)-lifting property with respect to nn-cofibrations, for all nn\in\mathbb{N}.

Proof.

By assumption, f:XYδf:X\to Y^{\delta} induces a (0,2δ)(0,2\delta)-interleaving in all homotopy groups. Now, an nn-cofibration can be written as a composite of n+1n+1 single dimensional extensions, and any shift of a single dimensional extension is again a single dimensional extension, so it is enough to show that ff has the right 4δ4\delta-lifting property with respect to single dimensional extensions.

A single dimensional extension is the pushout of a coproduct iIri(Dn)iIriDn\coprod_{i\in I}r_{i}\odot(\partial D^{n})\to\coprod_{i\in I}r_{i}\odot D^{n}, so it is enough to show that ff has the right 4δ4\delta-lifting property with respect to coproducts of that form, which follows from 5.12 and the universal property of coproducts. ∎

We are ready to prove B.

Theorem B.

Let X,YSS𝐑mX,Y\in\SS^{\mathbf{R}^{m}} be persistent spaces that are assumed to be projective cofibrant and dd-skeletal if SS=𝐬𝐒𝐞𝐭\SS=\mathbf{sSet}, or persistent CW-complexes of dimension at most dd if SS=𝐓𝐨𝐩\SS=\mathbf{Top}. Let δ0𝐑m\delta\geq 0\in\mathbf{R}^{m}. If there exists a morphism in the homotopy category XYδ𝖧𝗈(SS𝐑m)X\to Y^{\delta}\in\mathsf{Ho}\left(\SS^{\mathbf{R}^{m}}\right) that induces δ\delta-interleavings in all homotopy groups, then XX and YY are (4(d+1)δ)(4(d+1)\delta)-interleaved in the homotopy category.

Proof.

By 5.5 and 5.9, XX and YY are dd-cofibrant. Let [f]:XYδ[f]:X\to Y^{\delta} be as in the statement. Since [f][f] is a morphism in the homotopy category, we begin by choosing a convenient representative of it. We let p:XYp:X^{\prime}\to Y^{\prime} be a projective fibration between projective fibrant objects such that there exist trivial cofibrations i:XXi:X\to X^{\prime} and j:YYj:Y\to Y^{\prime} with [p][i]=[j][f][p]\circ[i]=[j]\circ[f], in 𝖧𝗈(SS𝐑m)\mathsf{Ho}\left(\SS^{\mathbf{R}^{m}}\right).

Note that [p][p] induces a (0,2δ)(0,2\delta)-interleaving in homotopy groups, between XX^{\prime} and YδY^{\prime\delta}. Since YδY^{\delta} is dd-cofibrant, 5.13 guarantees that we can find a (4(d+1)δ)(4(d+1)\delta)-lift gg^{\prime} of pp against Y\emptyset\to Y. We can then use the fact that j:YYj:Y\to Y^{\prime} is a trivial cofibration, and that XX^{\prime} is fibrant, to construct the following lift

Y{Y}X(4d+3)δ{X^{\prime(4d+3)\delta}}Y{Y^{\prime}}gg^{\prime}jjgg

We will show that 𝖲δ,4(d+1)δ([p]):X4(d+1)δY\mathsf{S}_{\delta,4(d+1)\delta}([p]):X^{\prime}\to_{4(d+1)\delta}Y^{\prime} and 𝖲(4d+3)δ,4(d+1)δ([g]):Y4(d+1)δX\mathsf{S}_{(4d+3)\delta,4(d+1)\delta}([g]):Y^{\prime}\to_{4(d+1)\delta}X^{\prime} form a (4(d+1)δ)(4(d+1)\delta)-interleaving in the homotopy category between XX^{\prime} and YY^{\prime}.

On the one hand, note that, by construction, we have p(4d+3)δgj=p(4d+3)δg=jp^{(4d+3)\delta}\circ g\circ j=p^{(4d+3)\delta}\circ g^{\prime}=j, so, since [j][j] is an isomorphism, it follows that [p](4d+3)δ[g]=𝖲4(d+1)δ([𝗂𝖽Y])[p]^{(4d+3)\delta}\circ[g]=\mathsf{S}_{4(d+1)\delta}([\mathsf{id}_{Y^{\prime}}]), and thus that

𝖲δ,4(d+1)δ([p])4(d+1)δ𝖲(4d+3)δ,4(d+1)δ([g])=𝖲8(d+1)δ([𝗂𝖽Y]).\mathsf{S}_{\delta,4(d+1)\delta}([p])^{4(d+1)\delta}\circ\mathsf{S}_{(4d+3)\delta,4(d+1)\delta}([g])=\mathsf{S}_{8(d+1)\delta}([\mathsf{id}_{Y^{\prime}}]).

On the other hand, since XX is cofibrant and YY^{\prime} is fibrant, it follows from the previous paragraph that p4(d+1)δgδpi:XY(4d+5)δp^{4(d+1)\delta}\circ g^{\delta}\circ p\circ i:X\to Y^{\prime(4d+5)\delta} is homotopic to p4(d+1)δ𝖲0,4(d+1)δ(i)p^{4(d+1)\delta}\circ\mathsf{S}_{0,4(d+1)\delta}(i). Let H:I×XY(4d+5)δH:I\times X\to Y^{\prime(4d+5)\delta} be a homotopy between these maps, which gives the following commutative diagram

XX{X\coprod X}X4(d+1)δ{X^{\prime 4(d+1)\delta}}I×X{I\times X}Y(4d+5)δ,{Y^{\prime(4d+5)\delta},}(𝖲0,4(d+1)δ(i),gδpi)\left(\mathsf{S}_{0,4(d+1)\delta}(i){,}g^{\delta}\circ p\circ i\right)(i0,i1)(i_{0},i_{1})HHp4(d+1)δp^{4(d+1)\delta}

where the left vertical map is the inclusion of into the cylinder. We claim that, since XX is dd-cofibrant, the inclusion into the cylinder is a dd-cofibration. Indeed, a cell decomposition of this map is obtained by attaching an (n+1)(n+1)-cell for each nn-cell in the decomposition of XX. Now, by 5.13, we can find a (4(d+1)δ)(4(d+1)\delta)-lift of the diagram, which shows that

𝖲4(d+1)δ,8(n+1)δ([g]δ[p][i])=𝖲0,8(d+1)δ([i]):XX8(d+1)δ.\mathsf{S}_{4(d+1)\delta,8(n+1)\delta}([g]^{\delta}\circ[p]\circ[i])=\mathsf{S}_{0,8(d+1)\delta}([i]):X\to X^{\prime 8(d+1)\delta}.

Since the left hand side is equal to 𝖲(4d+3)δ,4(d+1)δ([g])4(d+1)δ𝖲δ,4(d+1)δ([p][i])\mathsf{S}_{(4d+3)\delta,4(d+1)\delta}([g])^{4(d+1)\delta}\circ\mathsf{S}_{\delta,4(d+1)\delta}([p]\circ[i]), and [i][i] is an isomorphism, it follows that [g]4(d+1)δ𝖲δ,4(d+1)δ([p])=𝖲8(d+1)δ([𝗂𝖽X])[g]^{4(d+1)\delta}\circ\mathsf{S}_{\delta,4(d+1)\delta}([p])=\mathsf{S}_{8(d+1)\delta}([\mathsf{id}_{X^{\prime}}]). ∎

Remark 5.14.

Together, A and B imply a version of the persistent Whitehead conjecture, which we recall as 5.15. Our result is, in a sense, stronger than the one conjectured, since B, which addresses part (i)(i) of the conjecture, applies to arbitrary multi-persistent spaces. In another respect, our result is slightly weaker, as the conjecture is stated for cofibrant, pointwise CW-complexes, which does not necessarily imply being a persistent CW-complex in our sense. We believe that this is not an issue, as many of the cofibrant, pointwise CW-complexes persistent topological spaces that appear in applications are in fact persistent CW-complexes, as they are usually the geometric realization of a filtered simplicial complex.

Conjecture 5.15 ([2, Conjecture 8.6]).

Suppose we are given connected, cofibrant X,Y:𝐑𝐓𝐨𝐩X,Y:\mathbf{R}\to\mathbf{Top}, with each X(r)X(r) and Y(r)Y(r) CW-complexes of dimension at most dd, and f:XYδf:X\to Y^{\delta} inducing a δ\delta-interleaving in all homotopy groups. Then, there is a constant cc, depending only on dd, such that

  • (i)

    ff induces a cδc\delta-interleaving in the homotopy category 𝖧𝗈(𝐓𝐨𝐩𝐑)\mathsf{Ho}\left(\mathbf{Top}^{\mathbf{R}}\right);

  • (ii)

    XX and YY are cδc\delta-homotopy interleaved.

References

  • [1] David Blanc, Mark W. Johnson and James M. Turner “A constructive approach to higher homotopy operations” In Homotopy theory: tools and applications 729, Contemp. Math. Amer. Math. Soc., Providence, RI, 2019, pp. 21–74
  • [2] Andrew J. Blumberg and Michael Lesnick “Universality of the Homotopy Interleaving Distance”, 2017 arXiv:1705.01690
  • [3] Peter Bubenik and Jonathan A. Scott “Categorification of Persistent Homology” In Discret. Comput. Geom. 51.3, 2014, pp. 600–627
  • [4] Frédéric Chazal et al. “Proximity of Persistence Modules and Their Diagrams” In Proceedings of the Twenty-Fifth Annual Symposium on Computational Geometry, SCG ’09, 2009, pp. 237–246
  • [5] Frédéric Chazal et al. “Gromov-Hausdorff Stable Signatures for Shapes Using Persistence” In Proceedings of the Symposium on Geometry Processing, SGP ’09, 2009, pp. 1393–1403
  • [6] Samir Chowdhury et al. “New families of stable simplicial filtration functors” In Topology and its Applications 279, 2020, pp. 107254
  • [7] David Cohen-Steiner, Herbert Edelsbrunner and John Harer “Stability of persistence diagrams” In Discrete Comput. Geom. 37.1, 2007, pp. 103–120
  • [8] Patrizio Frosini, Claudia Landi and Facundo Mémoli “The Persistent Homotopy Type Distance” In Homology, Homotopy and Applications 21, 2017
  • [9] Paul G. Goerss and John F. Jardine “Simplicial homotopy theory” 174, Progress in Mathematics Birkhäuser Verlag, Basel, 1999, pp. xvi+510
  • [10] Tom Goodwillie “A homotopy commutative diagram that cannot be strictified” (version: 2011-11-27), MathOverflow URL: https://mathoverflow.net/q/81962
  • [11] Philip S. Hirschhorn “Model categories and their localizations” 99, Mathematical Surveys and Monographs American Mathematical Society, Providence, RI, 2003, pp. xvi+457
  • [12] Mark Hovey “Model categories” 63, Mathematical Surveys and Monographs American Mathematical Society, Providence, RI, 1999, pp. xii+209
  • [13] J. F. Jardine “Persistent homotopy theory”, 2020 arXiv:2002.10013
  • [14] Masaki Kashiwara and Pierre Schapira “Persistent homology and microlocal sheaf theory” In Journal of Applied and Computational Topology 2.1-2 Springer ScienceBusiness Media LLC, 2018, pp. 83–113
  • [15] Michael Lesnick “Multidimensional Interleavings and Applications to Topological Inference”, 2012
  • [16] Michael Lesnick and Matthew Wright “Interactive Visualization of 2-D Persistence Modules”, 2015 arXiv:1512.00180
  • [17] Facundo Mémoli “A Distance Between Filtered Spaces Via Tripods”, 2017 arXiv:1704.03965
  • [18] Daniel G. Quillen “Homotopical algebra”, Lecture Notes in Mathematics, No. 43 Springer-Verlag, Berlin-New York, 1967, pp. iv+156 pp. (not consecutively paged)
  • [19] Alexander Rolle and Luis Scoccola “Stable and consistent density-based clustering”, 2020 arXiv:2005.09048
  • [20] Luis Scoccola “Locally Persistent Categories And Metric Properties Of Interleaving Distances”, 2020 URL: https://ir.lib.uwo.ca/etd/7119
  • [21] Hirosi Toda “Composition methods in homotopy groups of spheres”, Annals of Mathematics Studies, No. 49 Princeton University Press, Princeton, N.J., 1962, pp. v+193