Skip to main content
Log in

The reflection distance between zigzag persistence modules

  • Published:
Journal of Applied and Computational Topology Aims and scope Submit manuscript

Abstract

By invoking the reflection functors introduced by Bernstein et al. (Russ Math Surv 28(2):17–32, 1973), in this paper we define a metric on the space of all zigzag modules of a given length, which we call the reflection distance. We show that the reflection distance between two given zigzag modules of the same length is an upper bound for the \(\ell ^1\)-bottleneck distance between their respective persistence diagrams.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2

Similar content being viewed by others

Notes

  1. Note that we are slightly abusing notation since \(\text {Mod}_{\tau }\) may not be a set.

  2. We warn the reader that while we are referring to isomorphism classes of interval \(\tau \) modules, in order to keep our notation simple, we will use the notation \(\mathbb {I}_\tau ([b,d])\) instead of the more correct notation \([\mathbb {I}_\tau ([b,d])]\).

  3. The infinite version of the Matching Lemma holds as well (Ore 1962) and actually generalizes the Cantor–Schröder–Bernstein (CSB) theorem slightly; the CBS theorem says that if \(f:S\rightarrow T\) and \(g:T\rightarrow S\) are injections between sets S and T then there exists a bijection \(M:S\rightarrow T\). Since injective maps can be viewed as matchings with coimages being equal to the domain of the map, the infinite Matching Lemma implies the existence of a matching \(M:S\nrightarrow T\) with \(\text {coim}(M) = \text {coim}(f) = S\) and \(\text {im}(M) = \text {coim}(g) = T\), i.e. a bijection between S and T.

References

  • Aharoni, R.: Infinite matching theory. Discrete Math. 95(1), 5–22 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  • Babichev, A., Morozov, D., Dabaghian, Y.: Robust spatial memory maps encoded by networks with transient connections. PLoS Comput Biol 14(9), e1006433 (2018). https://doi.org/10.1371/journal.pcbi.1006433

    Article  Google Scholar 

  • Bauer, U., Lesnick, M.: Induced matchings of barcodes and the algebraic stability of persistence. In: Proceedings of the 30th Annual Symposium on Computational Geometry, p. 355. ACM, New York (2014)

  • Bernstein, I.N., Gel’fand, I.M., Ponomarev, V.A.: Coxeter functors and Gabriel’s theorem. Russ. Math. Surv. 28(2), 17–32 (1973)

    Article  MathSciNet  MATH  Google Scholar 

  • Bjerkevik, H.B.: Stability of higher-dimensional interval decomposable persistence modules. ArXiv e-prints (Sept. 2016)

  • Bjerkevik, H.B., Botnan, M.B., Kerber, M.: Computing the interleaving distance is NP-hard. Preprint (2018). arXiv:1811.09165

  • Borceux, F., Rota, G.C., Doran, B., Flajolet, P., Lam, T.Y., Lutwak, E., Ismail, M.: Handbook of Categorical Algebra: Volume 1, Basic Category Theory. Cambridge Textbooks in Linguis. Cambridge University Press, Cambridge (1994)

    Book  Google Scholar 

  • Botnan, M., Lesnick, M.: Algebraic stability of zigzag persistence modules. Algebraic Geom. Topol. 18(6), 3133–3204 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  • Bubenik, P., Scott, J.A.: Categorification of persistent homology. Discrete Comput. Geom. 51(3), 600–627 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  • Bubenik, P., de Silva, V., Scott, J.: Metrics for generalized persistence modules. Found. Comput. Math. 15(6), 1501–1531 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  • Carlsson, G., De Silva, V.: Zigzag persistence. Found. Comput. Math. 10(4), 367–405 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  • Chazal, F., Cohen-Steiner, D., Glisse, M., Guibas, L.J., Oudot, S.Y.: Proximity of persistence modules and their diagrams. In: Proceedings of the 25th Annual Symposium on Computational Geometry—SCG’09 (2009)

  • Chazal, F., De Silva, V., Glisse, M., Oudot, S.: The Structure and Stability of Persistence Modules. Springer, Berlin (2016)

    Book  MATH  Google Scholar 

  • Chowdhury, S., Dai, B., Mémoli, F.: The importance of forgetting: limiting memory improves recovery of topological characteristics from neural data. PLoS ONE 13(9), e0202561 (2018). https://doi.org/10.1371/journal.pone.0202561

    Article  Google Scholar 

  • Cohen-Steiner, D., Edelsbrunner, H., Morozov, D.: Vines and vineyards by updating persistence in linear time. In: Proceedings of the 22nd Annual Symposium on Computational Geometry, pp. 119–126. ACM, New York (2006a)

  • Cohen-Steiner, D., Edelsbrunner, H., Harer, J.: Stability of persistence diagrams. Discrete Comput. Geom. 37(1), 103–120 (2006b)

    Article  MathSciNet  MATH  Google Scholar 

  • Cohen-Steiner, D., Edelsbrunner, H., Harer, J., Mileyko, Y.: Lipschitz functions have \(L_{p}\)-stable persistence. Found. Comput. Math. 10(2), 127–139 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  • Corcoran, P., Jones, C.B.: Spatio-temporal modeling of the topology of swarm behavior with persistence landscapes. In: Proceedings of the 24th ACM SIGSPATIAL International Conference on Advances in Geographic Information Systems, p. 65. ACM, New York (2016)

  • Delfinado, C.J.A., Edelsbrunner, H.: An incremental algorithm for betti numbers of simplicial complexes on the 3-sphere. Comput. Aided Geom. Des. 12(7), 771–784 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  • Edelsbrunner, H., Harer, J.: Computational Topology: An Introduction. American Mathematical Society, New York (2010)

    MATH  Google Scholar 

  • Edelsbrunner, H., Letscher, D., Zomorodian, A.: Topological persistence and simplification. In: 2000. Proceedings. 41st Annual Symposium on Foundations of Computer Science, pp. 454–463. IEEE, New York (2000)

  • Frosini, P.: A distance for similarity classes of submanifolds of a euclidean space. Bull. Aust. Math. Soc. 42(3), 407–415 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  • Frosini, P.: Measuring shapes by size functions. In: Intelligent Robots and Computer Vision X: Algorithms and Techniques, vol. 1607, pp. 122–134. International Society for Optics and Photonics, New York (1992)

  • Gabriel, P.: Unzerlegbare darstellungen I. Manuscr. Math. 6(1), 71–103 (1972)

    Article  MathSciNet  MATH  Google Scholar 

  • Kalisnik, S.: Persistent homology and duality. Ph.D. Thesis. University of Ljubljana (2013)

  • Kim, W., Memoli, F.: Stable signatures for dynamic metric spaces via zigzag persistent homology. Preprint (2017). arXiv:1712.04064

  • Lesnick, M.: The theory of the interleaving distance on multidimensional persistence modules. Found. Comput. Math. 15(3), 613–650 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  • Mac Lane, S.: Categories for the Working Mathematician. Springer, Berlin (1971)

    Book  MATH  Google Scholar 

  • Ore, O.: Theory of Graphs, vol. 38. American Mathematical Society, New York (1962)

    MATH  Google Scholar 

  • Oudot, S.Y.: Persistence Theory: From Quiver Representations to Data Analysis. American Mathematical Society, New York (2015)

    Book  MATH  Google Scholar 

  • Riehl, E.: Category Theory in Context. Dover Modern Math Originals. Dover Publications, Aurora (2017)

    MATH  Google Scholar 

  • Robins, V.: Towards computing homology from finite approximations. Topol. Proc. 24, 503–532 (1999)

    MathSciNet  MATH  Google Scholar 

  • Schweizer, B.: Cantor, Schröder, and Bernstein in orbit. Math. Mag. 73(4), 311 (2000)

    Article  MathSciNet  Google Scholar 

  • Tausz, A., Carlsson, G.: Applications of zigzag persistence to topological data analysis. Preprint (2011). arXiv:1108.3545

  • Zomorodian, A., Carlsson, G.: Computing persistent homology. Discrete Comput. Geom. 33(2), 249–274 (2005)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We thank Peter Bubenik, Mike Catanzaro, Woojin Kim, and José Perea for their useful feedback. We also thank Lauren Wickman for her careful reading of the late drafts. A.E. was supported by the NSF RTG # 1547357. F.M. was supported by NSF RI # 1422400 and NSF AF #1526513.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Facundo Mémoli.

Additional information

Publisher's Note

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

Appendices

Appendices

Limits and colimits

We assume the reader is familiar with (small) categories, functors, and natural transformations (see Mac Lane 1971; Riehl 2017 for details). We denote by \(\mathbf {Vect}\) the category of vector spaces over some fixed field \(\mathbb {F}\).

Definition A1.1

Fix a small category \(\mathcal {J}\).

  1. 1.

    A diagram of vector spaces of shape\(\mathcal {J}\) is a functor \(D:\mathcal {J}\rightarrow \mathbf Vect \).

  2. 2.

    A cone over a diagram \(D:\mathcal {J}\rightarrow \mathbf Vect \) is a vector space N together with a collection of linear transformations \(\lambda :=\{\lambda _J: N\rightarrow D(J)\ | \ J\in \text {Ob}(\mathcal {J})\}\), indexed by the objects of \(\mathcal {J}\), such that for any morphism \(f:A\rightarrow B\) in \(\text {Hom}(\mathcal {J})\) we have \(\lambda _B = D(f)\circ \lambda _A\). We denote such a cone by \((N,\lambda )\).

  3. 3.

    A cocone over a diagram \(D:\mathcal {J}\rightarrow \mathbf Vect \) is a vector space M together with a collection of linear transformations \(\gamma :=\{\gamma _J: D(J)\rightarrow M \ | \ J\in \text {Ob}(\mathcal {J})\}\), indexed by the objects of \(\mathcal {J}\), such that for any morphism \(f:A\rightarrow B\) in \(\text {Hom}(\mathcal {J})\) we have \(\gamma _A = \gamma _B\circ D(f)\). We denote such a cocone by \((M,\lambda )\).

Limits and colimits are then universal cones or cocones, respectively:

Definition A1.2

Let \(D:\mathcal {J}\rightarrow \mathbf {Vect}\) be a diagram of vector spaces.

  1. 1.

    The limit of the diagram D is a cone \((L,\phi )\) over D such that for any other cone \((N,\lambda )\) over D, there exists a unique linear transformation \(\psi :N\rightarrow L\) with \(\lambda _A = \phi _A \circ \psi \) for all \(A\in \text {Ob}(\mathcal {J})\). We denote the limit L by \(\,\lim D\).

  2. 2.

    The colimit of the diagram D is a cocone \((C,\phi )\) over D such that for any other cocone \((M,\lambda )\) over D, there exists a unique linear transformation \(\psi :M\rightarrow C\) with \(\lambda _A = \phi _A \circ \psi \) for all \(A\in \text {Ob}(\mathcal {J})\). We denote the colimit C by \(\text {colim}(D)\).

Theorem A1.1

Let \(\mathcal {J}\) be a small category and let \(D^1,D^2\) be diagrams of vector spaces. If there exists a natural transformation \(\eta :D^1\rightarrow D^2\) all of whose components are monomorphisms, then \((L^1,\eta _J\circ \lambda ^1_J)\) is a cone for \(D^2\) and the unique morphism \(\psi :L^1\rightarrow L^2\) satisfying \(\eta _J\circ \lambda ^1_J = \lambda ^2_J\circ \psi \) for all \(J\in \mathcal {J}\) is a monomorphism.

Proof

The proof can be found in more generality in Borceux et al. (1994 p. 89, Corollary 2.15.3). \(\square \)

Let \(\mathbf {2}\) denote the discrete category with 2 objects. That is, \(\text {Ob}(\mathbf {2}) = \{1,2\}\) and \(\text {Hom}(1,2) = \emptyset \).

Definition A1.3

Let XY be objects in the category \(\,\mathcal {C}\) and let \(D:\mathbf {2}\rightarrow \mathcal {C}\) be the diagram given by \(D(1) = X\) and \(D(2) = Y\).

  1. 1.

    The product of B and C, denoted \(B\times C\), is the limit of D, if it exists,

  2. 2.

    The coproduct of B and C, denoted \(B\amalg C\), is the colimit of D, if it exists.

In \(\mathbf {Vect}\), products and coproducts always exist and coincide.

Theorem A1.2

Let \(\mathcal {J}\) be a small category and let \(D^1, D^2\) be diagrams of vector spaces. Then \(\lim (D^1\times D^2)\cong \lim (D^1)\times \lim (D^2)\) Dually, \(\text {colim}(D^1\amalg D^2)\cong \text {colim}(D^1)\amalg \text {colim}(D^2)\). Here \(D^1\times D^2\) and \(D^1\amalg D^2\) denote the product and coproduct, respectively, of \(D^1\) and \(D^2\) in the functor category \(\mathcal {C}^\mathcal {J}\).

Proof

Consider the product category \(\mathbf {2}\times J\) and let \(F:\mathbf {2}\times \mathcal {J}\rightarrow \mathcal {C}\) be the functor given by \(F(i,j) = D^i(j)\) for \(i = 1,2\) and \(j\in \mathcal {J}\). By commutativity of limits,

$$\begin{aligned} \lim _\mathbf {2}\lim _\mathcal {J} F(i,j) = \lim _\mathcal {J}\lim _\mathbf {2}F(i,j) \end{aligned}$$

[see Riehl (2017 p. 111, Theorem 3.8.1)]. Now

$$\begin{aligned} \lim (D^1\times D^2)&= \lim _{\mathcal {J}}(D^1\times D^2) = \lim _{\mathcal {J}}\lim _\mathbf {2} F(i,j)\\&= \lim _{\mathbf {2}}\lim _\mathcal {J} F(i,j) = \lim _{\mathbf {2}}\lim D^i\\&= \lim (D^1)\times \lim (D^2). \end{aligned}$$

The second statement follows by a duality argument. \(\square \)

Matchings

The Matching Lemma is a useful tool for combining matchings in opposite directions into a single matching. For us, the lemma was crucial for proving our stability result, while Bjerkevik makes use of the lemma in proving his generalization of the algebraic stability theorem (Bjerkevik 2016). While it may be one of the earliest results in infinite matching theory (Aharoni 1991), we believe its applications to stability-type questions in persistence theory make it worth expounding upon here.

Lemma A2.1

(Matching lemma for finite matchings) Let S and T be sets and let \(f:S\nrightarrow T\) and \(g:T\nrightarrow S\) be finite matchings. Then there exists a matching \(M:S\nrightarrow T\) such that

  1. 1.

    \(\text {coim}(f)\subseteq \text {coim}(M)\),

  2. 2.

    \( \text {coim}(g)\subseteq \text {im}(M)\),

  3. 3.

    if \(M(s) = t\) then either \(f(s) = t\) or \(g(t) = s\).

Proof

The proof is inspired by a proof of the Cantor–Schröder–Bernstein theoremFootnote 3 given in Schweizer (2000). Let \(s\in \text {coim}(f)\), \(t\in \text {coim}(g)\), and consider sequences of the form

$$\begin{aligned} \cdots {\mathop {\longrightarrow }\limits ^{g}}f^{-1}(g^{-1}(s)){\mathop {\longrightarrow }\limits ^{f}} g^{-1}(s){\mathop {\longrightarrow }\limits ^{g}}s{\mathop {\longrightarrow }\limits ^{f}} f(s) {\mathop {\longrightarrow }\limits ^{g}} g(f(s)) {\mathop {\longrightarrow }\limits ^{f}} \cdots \end{aligned}$$

and

$$\begin{aligned} \cdots {\mathop {\longrightarrow }\limits ^{f}}g^{-1}(f^{-1}(t)){\mathop {\longrightarrow }\limits ^{g}} f^{-1}(t){\mathop {\longrightarrow }\limits ^{f}}t{\mathop {\longrightarrow }\limits ^{g}} g(t) {\mathop {\longrightarrow }\limits ^{f}} f(g(t)) {\mathop {\longrightarrow }\limits ^{g}} \cdots , \end{aligned}$$

where we allow these sequences to terminate to the right or left when undefined. We refer to these sequences as the orbits of s or t. Since f and g are finite matchings, such sequences either terminate on both the left and right, or are infinite but periodic. By injectivity of f and g, every element of \(\text {coim}(f)\sqcup \text {coim}(g)\) appears in exactly one orbit. Moreover, every orbit falls into one of the following five classes:

  1. 1.

    \(s\rightarrow t\rightarrow \cdots \rightarrow s\rightarrow y\),

  2. 2.

    \(s\rightarrow t\rightarrow \cdots \rightarrow s\rightarrow t\rightarrow x\),

  3. 3.

    \(t\rightarrow s\rightarrow \cdots \rightarrow t\rightarrow x\),

  4. 4.

    \(t\rightarrow s\rightarrow \cdots \rightarrow t\rightarrow s\rightarrow y\),

  5. 5.
    figure w

where the s’s and t’s represent elements of \(\text {coim}(f)\) and \(\text {coim}(g)\), respectively, arrows represent either f or g, and y’s and x’s represent elements of \(\text {im}(f){\setminus } \text {coim}(g)\) and \(\text {im}(g){\setminus } \text {coim}(f)\), respectively. We define a matching \(M:S\nrightarrow T\) as follows: for each \(i = 1,\ldots ,5\) let

$$\begin{aligned} S_i^{\text {co}}:= \{s\in \text {coim}(f) \ | \ s \text { appears in an orbit of type } i\} \end{aligned}$$

and

$$\begin{aligned} T_i^{\text {co}}:= \{t\in \text {coim}(g) \ | \ t \text { appears in an orbit of type } i\}. \end{aligned}$$

Then and . We partition \(S_1^{\text {co}}\) and \(T_3^{\text {co}}\) further by defining

$$\begin{aligned} S_1^{\text {co}\backslash \text {p}} : = \{s\in S_1^{\text {co}} \ | \ f(s)\in \text {coim}(g)\}, \quad S_1^{\text {p}} : = \{s\in S_1^{\text {co}} \ | \ f(s)\not \in \text {coim}(g)\},\\ T_3^{\text {co}\backslash \text {p}} : = \{t\in T_3^{\text {co}} \ | \ g(t)\in \text {coim}(f)\}, \quad \text { and } \quad T_3^{\text {p}} : = \{t\in T_3^{\text {co}}\ | \ g(t)\not \in \text {coim}(f)\}, \end{aligned}$$

so that and . The image, coimage, and mapping of M is specified by the diagram

figure x

Note that \(f(S_1^{\text {co}\backslash \text {p}}) = T_1^{\text {co}}\), \(f(S_2^{\text {co}}) = T_2^{\text {co}}\), \(g(T_3^{\text {co}\backslash \text {p}}) = S_3^{\text {co}} \), \(g(T_4^{\text {co}}) = S_4^{\text {co}}\), and \(f(S_5^{\text {co}}) = T_5^{\text {co}}\) so that M is surjective. Since f and g are injective, we see that M is a bijection between each of the parts specified and thus defines a matching. Moreover, and . Property (3) evidently holds by the definition of M. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Elchesen, A., Mémoli, F. The reflection distance between zigzag persistence modules. J Appl. and Comput. Topology 3, 185–219 (2019). https://doi.org/10.1007/s41468-019-00031-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s41468-019-00031-0

Keywords

Mathematics Subject Classification

Navigation