Skip to main content
Log in

Stability of 2-Parameter Persistent Homology

  • Published:
Foundations of Computational Mathematics Aims and scope Submit manuscript

Abstract

The Čech and Rips constructions of persistent homology are stable with respect to perturbations of the input data. However, neither is robust to outliers, and both can be insensitive to topological structure of high-density regions of the data. A natural solution is to consider 2-parameter persistence. This paper studies the stability of 2-parameter persistent homology: we show that several related density-sensitive constructions of bifiltrations from data satisfy stability properties accommodating the addition and removal of outliers. Specifically, we consider the multicover bifiltration, Sheehy’s subdivision bifiltrations, and the degree bifiltrations. For the multicover and subdivision bifiltrations, we get 1-Lipschitz stability results closely analogous to the standard stability results for 1-parameter persistent homology. Our results for the degree bifiltrations are weaker, but they are tight, in a sense. As an application of our theory, we prove a law of large numbers for subdivision bifiltrations of random data.

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.

Similar content being viewed by others

References

  1. C. D. Aliprantis and K. C. Border. Infinite Dimensional Analysis: A Hitchhiker’s Guide. Springer Science & Business Media, 2006.

  2. U. Bauer, M. Kerber, F. Roll, and A. Rolle. A unified view on the functorial nerve theorem and its variations. arXiv preprintarXiv:2203.03571, 2022.

  3. U. Bauer and M. Lesnick. Induced matchings and the algebraic stability of persistence barcodes. Journal of Computational Geometry, 6(2):162–191, 2015.

    MathSciNet  Google Scholar 

  4. H. B. Bjerkevik. On the stability of interval decomposable persistence modules. Discrete and Computational Geometry, 66:92–121, 2021.

  5. H. B. Bjerkevik, A. Blumberg and M. Lesnick. \(\ell ^p\)-continuity properties of multicover persistent homology. In preparation.

  6. A. J. Blumberg, I. Gal, M. A. Mandell, and M. Pancia. Robust statistics, hypothesis testing, and confidence intervals for persistent homology on metric measure spaces. Foundations of Computational Mathematics, 14(4):745–789, 2014.

    Article  MathSciNet  Google Scholar 

  7. A. J. Blumberg and M. Lesnick. Universality of the homotopy interleaving distance. Transactions of the American Mathematical Society. In press.

  8. O. Bobrowski, S. Mukherjee, J. E. Taylor, et al. Topological consistency via kernel estimation. Bernoulli, 23(1):288–328, 2017.

    Article  MathSciNet  Google Scholar 

  9. M. Botnan and W. Crawley-Boevey. Decomposition of persistence modules. Proceedings of the American Mathematical Society, 148(11):4581–4596, 2020.

    Article  MathSciNet  Google Scholar 

  10. M. B. Botnan and G. Spreemann. Approximating persistent homology in Euclidean space through collapses. Applicable Algebra in Engineering, Communication and Computing, 26(1-2):73–101, 2015.

    Article  MathSciNet  Google Scholar 

  11. P. Bubenik, V. De Silva, and J. Scott. Metrics for generalized persistence modules. Foundations of Computational Mathematics, 15(6):1501–1531, 2015.

  12. M. Buchet, F. Chazal, S. Y. Oudot, and D. R. Sheehy. Efficient and robust persistent homology for measures. Computational Geometry, 58:70–96, 2016.

    Article  MathSciNet  Google Scholar 

  13. G. Carlsson, T. Ishkhanov, V. De Silva, and A. Zomorodian. On the local behavior of spaces of natural images. International Journal of Computer Vision, 76(1):1–12, 2008.

    Article  MathSciNet  Google Scholar 

  14. G. Carlsson and A. Zomorodian. The theory of multidimensional persistence. Discrete and Computational Geometry, 42(1):71–93, 2009.

    Article  MathSciNet  Google Scholar 

  15. N. J. Cavanna, K. P. Gardner, and D. R. Sheehy. When and why the topological coverage criterion works. In Proceedings of the ACM-SIAM Symposium on Discrete Algorithms, 2017.

  16. A. Cerri, B. D. Fabio, M. Ferri, P. Frosini, and C. Landi. Betti numbers in multidimensional persistent homology are stable functions. Mathematical Methods in the Applied Sciences, 36(12):1543–1557, 2013.

    Article  MathSciNet  Google Scholar 

  17. F. Chazal, D. Cohen-Steiner, M. Glisse, L. J. Guibas, and S. Y. Oudot. Proximity of persistence modules and their diagrams. In Proceedings of the 25th Annual Symposium on Computational Geometry, SCG ’09, pages 237–246, New York, NY, USA, 2009. ACM.

  18. F. Chazal, D. Cohen-Steiner, L. J. Guibas, F. Mémoli, and S. Y. Oudot. Gromov–Hausdorff stable signatures for shapes using persistence. In Proceedings of the Symposium on Geometry Processing, SGP ’09, pages 1393–1403, Aire-la-Ville, Switzerland, Switzerland, 2009. Eurographics Association.

  19. F. Chazal, D. Cohen-Steiner, and Q. Mérigot. Geometric inference for probability measures. Foundations of Computational Mathematics, pages 1–19, 2011.

  20. F. Chazal, V. de Silva, M. Glisse, and S. Oudot. The Structure and Stability of Persistence Modules. Springer International Publishing, 2016.

    Google Scholar 

  21. F. Chazal, V. De Silva, and S. Oudot. Persistence stability for geometric complexes. Geometriae Dedicata, 173(1):193–214, 2014.

    Article  MathSciNet  Google Scholar 

  22. F. Chazal, B. Fasy, F. Lecci, B. Michel, A. Rinaldo, and L. Wasserman. Robust topological inference: Distance to a measure and kernel distance. Journal of Machine Learning Research, 18(159):1–40, 2018.

    Google Scholar 

  23. F. Chazal, L. J. Guibas, S. Y. Oudot, and P. Skraba. Scalar field analysis over point cloud data. Discrete and Computational Geometry, 46(4):743–775, May 2011.

  24. F. Chazal, L. J. Guibas, S. Y. Oudot, and P. Skraba. Persistence-based clustering in Riemannian manifolds. Journal of the ACM, 60(6), Nov. 2013. Article No. 41.

  25. F. Chazal and S. Oudot. Towards persistence-based reconstruction in Euclidean spaces. In Proceedings of the 24th Annual Symposium on Computational Geometry, pages 232–241. ACM, 2008.

  26. D. Cohen-Steiner, H. Edelsbrunner, and J. Harer. Stability of persistence diagrams. Discrete and Computational Geometry, 37(1):103–120, Jan. 2007.

  27. R. Corbet, M. Kerber, M. Lesnick, and G. Osang. Computing the multicover bifiltration. In 37th International Symposium on Computational Geometry (SoCG). Leibniz International Proceedings in Informatics (LIPIcs), 2021.

  28. W. Crawley-Boevey. Decomposition of pointwise finite-dimensional persistence modules. Journal of Algebra and Its Applications, 14(05):1550066, 2015.

    Article  MathSciNet  Google Scholar 

  29. D. J. Daley and D. Vere-Jones. An introduction to the theory of point processes. Vol. I. Probability and its Applications (New York). Springer-Verlag, second edition, 2003. Elementary theory and methods.

  30. V. De Silva, E. Munch, and A. Patel. Categorified reeb graphs. Discrete & Computational Geometry, 55(4):854–906, 2016.

    Article  MathSciNet  Google Scholar 

  31. R. M. Dudley. Real analysis and probability. CRC Press, 2018.

  32. D. Dugger and D. C. Isaksen. Topological hypercovers and 1-realizations. Mathematische Zeitschrift, 246(4):667–689, 2004.

    Article  MathSciNet  Google Scholar 

  33. H. Edelsbrunner and G. Osang. The multi-cover persistence of Euclidean balls. In 34th International Symposium on Computational Geometry (SoCG 2018). Schloss Dagstuhl-Leibniz-Zentrum fuer Informatik, 2018.

  34. H. Edelsbrunner and G. Osang. A simple algorithm for computing higher order Delaunay mosaics and \(\alpha \)-shapes. arXiv:2011.03617, 2020.

  35. H. Edelsbrunner and G. Osang. The multi-cover persistence of Euclidean balls. In 34th International Symposium on Computational Geometry (SoCG 2018), Schloss Dagstuhl-Leibniz-Zentrum fuer Informatik, 2018.

  36. P. Frosini and M. Mulazzani. Size homotopy groups for computation of natural size distances. Bulletin of the Belgian Mathematical Society-Simon Stevin, 6(3):455–464, 1999.

    Article  MathSciNet  Google Scholar 

  37. A. L. Gibbs and F. E. Su. On choosing and bounding probability metrics. International Statistical Review / Revue Internationale de Statistique, 70(3):419–435, 2002.

    Google Scholar 

  38. A. Greven, P. Pfaffelhuber, and A. Winter. Convergence in distribution of random metric measure spaces (\(\lambda \)-coalescent measure trees). Probability Theory and Related Fields, 145(1-2):285–322, 2009.

    Article  MathSciNet  Google Scholar 

  39. S. Harker, M. Kramár, R. Levanger, and K. Mischaikow. A comparison framework for interleaved persistence modules. Journal of Applied and Computational Topology, 3(1-2):85–118, 2019.

    Article  MathSciNet  Google Scholar 

  40. H. A. Harrington, N. Otter, H. Schenck, and U. Tillmann. Stratifying multiparameter persistent homology. SIAM Journal on Applied Algebra and Geometry, 3(3):439–471, 2019.

    Article  MathSciNet  Google Scholar 

  41. A. Hatcher. Algebraic topology. Cambridge Univ Press, 2002.

  42. P. S. Hirschhorn. Model categories and their localizations, volume 99. American Mathematical Society, 2009.

  43. S. Janson. On the Gromov-Prohorov distance. arXiv preprintarXiv:2005.13505, 2020.

  44. J. Jardine. Cluster graphs. Preprint, 2017.

  45. J. Jardine. Data and homotopy types. arXiv:1908.06323, 2019.

  46. J. Jardine. Stable components and layers. Canadian Mathematical Bulletin, pages 1–15, 2019.

  47. J. Jardine. Persistent homotopy theory. arXiv preprintarXiv:2002.10013, 2020.

  48. M. Lesnick. Multidimensional Interleavings and Applications to Topological Inference. PhD thesis, Stanford University, June 2012.

  49. M. Lesnick. The theory of the interleaving distance on multidimensional persistence modules. Foundations of Computational Mathematics, 15(3):613–650, 2015.

    Article  MathSciNet  Google Scholar 

  50. M. Lesnick and M. Wright. Interactive visualization of 2-D persistence modules. arXiv:1512.00180, 2015.

  51. M. Lesnick and M. Wright. Computing minimal presentations and betti numbers of 2-parameter persistent homology. SIAM Journal on Applied Algebra and Geometry, 6(2):267–298, 2022.

  52. M. Lesnick and R. Zhao. Computing the degree-Rips bifiltration. In preparation.

  53. L. McInnes and J. Healy. Accelerated hierarchical density based clustering. In 2017 IEEE International Conference on Data Mining Workshops (ICDMW), pages 33–42. IEEE, 2017.

  54. F. Mémoli. Gromov-Wasserstein distances and the metric approach to object matching. Foundations of Computational Mathematics, pages 1–71, 2011. https://doi.org/10.1007/s10208-011-9093-5.

  55. B. A. Munson and I. Volić. Cubical homotopy theory, volume 25. Cambridge University Press, 2015.

  56. G. Osang. Rhomboid tiling and order-k Delaunay mosaics. https://github.com/geoo89/rhomboidtiling, 2020.

  57. S. Y. Oudot and D. R. Sheehy. Zigzag zoology: Rips zigzags for homology inference. Foundations of Computational Mathematics, 15(5):1151–1186, 2015.

    Article  MathSciNet  Google Scholar 

  58. J. M. Phillips, B. Wang, and Y. Zheng. Geometric inference on kernel density estimates. In 31st International Symposium on Computational Geometry (SoCG 2015). Schloss Dagstuhl-Leibniz-Zentrum fuer Informatik, 2015.

  59. E. Riehl. Categorical homotopy theory. Cambridge University Press, 2014.

  60. A. Rolle and L. Scoccola. Stable and consistent density-based clustering. arXiv preprintarXiv:2005.09048, 2020.

  61. L. N. Scoccola. Locally Persistent Categories And Metric Properties Of Interleaving Distances. PhD thesis, The University of Western Ontario, 2020.

  62. D. R. Sheehy. A multicover nerve for geometric inference. In CCCG, pages 309–314, 2012.

  63. K.-T. Sturm. On the geometry of metric measure spaces. Acta Mathematica, 196(1):65–131, 2006.

    Article  MathSciNet  Google Scholar 

  64. The RIVET Developers. Rivet. https://github.com/rivetTDA/rivet, 2018-2020.

  65. C. Villani. Optimal Transport: Old and New. Springer Berlin Heidelberg, 2008.

    Google Scholar 

  66. O. Vipond. Multiparameter persistence landscapes. Journal of Machine Learning Research, 21(61):1–38, 2020.

    MathSciNet  Google Scholar 

  67. C. Webb. Decomposition of graded modules. American Mathematical Society, 94(4), 1985.

  68. A. Zomorodian and G. Carlsson. Computing persistent homology. Discrete and Computational Geometry, 33(2):249–274, 2005.

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

We thank Mike Mandell for his insights on various aspects of robustness related to this paper; René Corbet, Alex Rolle, and Don Sheehy for helpful conversations about the multicover nerve theorem; Håvard Bjerkevik for valuable discussions about the Wasserstein stability of 2-parameter persistence; Alex Tchernev for pointing out an error in Sect. 2.5 of the first version of the paper; and the anonymous reviewers for many helpful suggestions. The computations and figures of Appendix A would not have been possible without the work of Matthew Wright, Bryn Keller, Roy Zhao, and Simon Segert on the RIVET software. In particular, Zhao designed and implemented RIVET’s algorithm for computing degree-Rips bifiltrations, and Segert made critical improvements to RIVET’s visualization capabilities.

Funding

The development of RIVET was supported in part by NSF grant DMS-1606967. Blumberg was partially supported by NIH Grants 5U54CA193313 and GG010211-R01-HIV, AFOSR Grant FA9550-18-1-0415, and NSF Grant CNS 1514422.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael Lesnick.

Additional information

Communicated by Peter Bubenik.

Publisher's Note

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

Appendix A. A Computational Example of the Stability of Degree-Rips Bifiltrations

Appendix A. A Computational Example of the Stability of Degree-Rips Bifiltrations

In this section, we explore the stability of the degree-Rips bifiltration in an example, using the 2-parameter persistence software RIVET [50, 51, 64]. We consider three point clouds X, Y, and Z, shown in Fig. 1:

  • X consists of 475 points sampled uniformly from an annulus in \({\mathbb {R}}^2\) with outer radius .5 and inner radius .4.

  • \(Y=X\cup N\), where N consists of 25 points sampled uniformly from a disc of radius .4.

  • Z consists of 500 points sampled uniformly from a disc of radius .5.

We would like to consider, for each \(W\in \{X,Y,Z\}\) the homology module \(H_1({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(W))\) with coefficients in \({\mathbb {Z}}/2{\mathbb {Z}}\). However, working directly with \(H_1({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(W))\) is computationally expensive, so we instead work with an approximation H(W) of \(H_1({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(W))\); this is explained in Appendix A.1. In Appendix A.2, we illustrate the stability of degree bifiltrations in practice by using RIVET to visualize invariants of H(X), H(Y), and H(Z). Then, in Appendix A.3, we apply the stability result Proposition 3.9 (ii) to explain a small part of the similarity between H(X) and H(Y) observed in our visualizations. Recall that Proposition 3.9 (ii) applies to a nested pair of data sets; in Remark A.2, at the end of this section, we observe that Theorem 1.7 (ii), which does not assume that the data is nested, does not constrain the similarity between H(X) and H(Y).

We warn the reader that the remainder of this section is somewhat technical, in part because of the approximations involved. We invite the reader to skim the section on a first reading, focusing on understanding the figures.

Fig. 1
figure 1

The point clouds X, Y, Z

1.1 A.1 Approximations to the Degree-Rips Bifiltrations

For discussing RIVET computations, it is convenient to introduce a variant of the normalized Degree-Rips bifiltration. First, for W a finite metric space, let \(\bar{{\mathcal {R}}}(W):[0,\infty )\rightarrow \mathbf {Top}\) be the filtration defined by taking

$$\begin{aligned} \bar{{\mathcal {R}}}(W)_r=\lim _{r<s} \mathcal {R}(W)_s. \end{aligned}$$

This is precisely the variant of the Rips construction mentioned in Remark 2.3.

Now define a bifiltration

$$\begin{aligned} {{\,\mathrm{\overline{\mathcal{DR}\mathcal{}}}\,}}(W):{\mathbb {R}}^{\mathrm {op}}\times [0,\infty )\rightarrow {{\,\mathrm{\mathbf {Simp}}\,}}\end{aligned}$$

by taking \({{\,\mathrm{\overline{\mathcal{DR}\mathcal{}}}\,}}(W)_{(k,r)}\) to be the maximal subcomplex of \(\bar{{\mathcal {R}}}(W)_r\) whose vertices have degree at least \(|W|(r-1)\). This bifiltration is slightly different from the normalized degree-Rips bifiltration \({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(W)\) defined in Sect. 2.3—for one thing, they are indexed by different posets—but it’s not hard to see that these two bifiltrations are equivalent, in the sense that each determines the other in a simple way. Moreover, the restriction of \({{\,\mathrm{\overline{\mathcal{DR}\mathcal{}}}\,}}(W)\) to the poset \(J\) is \(\epsilon \)-interleaved with \({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(W)\) for any \(\epsilon >0\).

The largest simplicial complex in \({{\,\mathrm{\overline{\mathcal{DR}\mathcal{}}}\,}}(W)\) is the simplex with vertices W; denote this as S. For any simplex \(\sigma \in S\), we define the set of bigrades of appearance of \(\sigma \) to be the set of minimal elements \((k,r)\in {\mathbb {R}}^{\mathrm {op}}\times [0,\infty )\) such that \(\sigma \in {{\,\mathrm{\overline{\mathcal{DR}\mathcal{}}}\,}}(W)_{(k,r)}\). This is a finite and nonempty subset of \(J\).

To control the cost of the computations, for each \(W\in \{X,Y,Z\}\) we in fact work with a “coarsening”

$$\begin{aligned} F(W):{\mathbb {R}}^{\mathrm {op}}\times [0,\infty )\rightarrow \mathbf {Top}\end{aligned}$$

of the bifiltration \({{\,\mathrm{\overline{\mathcal{DR}\mathcal{}}}\,}}(W)\), where the bigrades of appearance of all simplices are rounded upwards so as to lie on a uniform \(100\times 100\) grid; the precise definition of this coarsening is given in [50]. The grid is chosen in a way that ensures that \({{\,\mathrm{\overline{\mathcal{DR}\mathcal{}}}\,}}(W)\) and F(W) are \((\tau ^{\frac{1}{100}},\mathrm {Id})\)-interleaved. Let \(F'(W)\) denote the restriction of F(W) to \(J\). By the generalized triangle inequality for interleavings (Remark 2.40), \({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(W)\) and \(F'(W)\) are \((\tau ^{\frac{1}{100}+\epsilon },\tau ^\epsilon )\)-interleaved for all \(\epsilon >0\).

For \(W\in \{X,Y,Z\}\), let H(W) denote the homology module \(H_1(F(W))\) with coefficients in \({\mathbb {Z}}/2{\mathbb {Z}}\). Observe that H(W) is finitely presented.

1.2 A.2 Visualization

RIVET computes and visualizes three invariants of a finitely presented biperistence module \(M:{\mathbb {R}}^{\mathrm {op}}\times [0,\infty )\rightarrow \mathbf {Top}\):

  • The Hilbert function of M, i.e., the dimension of each vector space of M.

  • The bigraded Betti numbers of M; these are functions \(\beta _i^M:{\mathbb {R}}^2\rightarrow {\mathbb {N}}\), \(i\in \{0,1,2\}\) which, roughly speaking, record the birth indices of generators, relations, and relations among the relations.

  • The fibered barcode of M [16], i.e., the map sending each affine line \(\ell \subset {\mathbb {R}}^{\mathrm {op}}\times [0,\infty )\) of non-positive slope to the barcode \({\mathcal {B}}_{M^{\ell }}\), where \(M^{\ell }\) is the restriction of M to \(\ell \). We regard \({\mathcal {B}}_{M^{\ell }}\) as a collection of intervals on the line \(\ell \).

See [50, 51] for more details about these invariants and their computation in RIVET.

Figure 2 shows RIVET’s visualization of H(X), H(Y), and H(Z), and of the barcodes \({\mathcal {B}}_{H(X)^{\ell }}\), \({\mathcal {B}}_{H(Y)^{\ell }}\), and \({\mathcal {B}}_{H(Z)^{\ell }}\) for one choice of \(\ell \). To explain the figure, first note that the x-axis is mirrored in each subfigure, relative to the usual convention, so that values decrease from left to right. In each figure, the Hilbert function is represented by grayscale shading, where the darkness is proportional to the homology dimension. The lightest non-white shade of gray shown in each figure corresponds to a value of 1. The bigraded Betti numbers are represented by translucent colored dots whose area is proportional to the value. The \(0^{\mathrm {th}}\), \(1^{\mathrm {st}}\), and \(2^{\mathrm {nd}}\) bigraded Betti numbers are shown in green, red, and yellow, respectively. In each figure, the line \(\ell \) is shown in blue, and the corresponding barcode is plotted in purple, with each interval offset perpendicularly from the line.

Note that the Hilbert function of H(X) takes value 1 on a large connected region parameter space, whose restriction to the left half-plane \(k> 0\) looks roughly like a triangle. The Hilbert function of H(Y) also takes value 1 on a substantial region in parameter space, albeit one smaller than for H(X). In contrast, the Hilbert function of H(Z) does not take the value 1 in a large region of the parameter space, and in fact almost all of the support of H(Z) lies very near (0, 0).

Fig. 2
figure 2

RIVET’s visualization of H(X), H(Y), H(Z) and the barcodes \({\mathcal {B}}_{H(X)^{\ell }}\), \({\mathcal {B}}_{H(Y)^{\ell }}\), \({\mathcal {B}}_{H(Z)^{\ell }}\), for one choice of line \(\ell \)

1.3 A.3 Stability Analysis

By appealing to Proposition 3.9 (ii), we can explain a small part of the similarity between the structures of H(X) and H(Y) observed in Fig. 2. More specifically, we will show that given H(X), Proposition 3.9 (ii) implies that the Hilbert function of H(Y) has non-trivial support on a small region of parameter space. The only property of Y we use in our analysis is that Y is a metric space of cardinality 500 containing X as a subspace.

By Proposition 3.9 (ii), \({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(X)\) and \({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(Y)\) are \((\kappa ^{|X|/|Y|},\gamma ^{\delta })\)-homotopy interleaved for any \(\delta >\tfrac{25}{500}=\tfrac{1}{20}\). Note that \(|X|/|Y|=\tfrac{475}{500}=\tfrac{19}{20}\). By the discussion of Appendix A.1, \({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(X)\), \(F'(X)\) are \((\tau ^{\frac{1}{100}+\epsilon },\tau ^\epsilon )\)-interleaved for all \(\epsilon >0\), and the same is true for \({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(Y)\), \(F'(Y)\). For \(\epsilon \ge 0\), let \(\zeta ^\epsilon \) be the forward shift given by

$$\begin{aligned} \zeta ^{\epsilon }(x,y)=\left( \frac{19x}{20}-\frac{1}{100}- \epsilon ,y+\frac{1}{100}+\epsilon \right) , \end{aligned}$$

and let \(\zeta =\zeta ^0\). A strict interleaving is also a homotopy interleaving, so by Proposition 2.38, \(F'(X)\) and \(F'(Y)\) are \((\zeta ^{\epsilon },\gamma ^{\frac{6}{100}+\epsilon })\)-homotopy interleaved for all \(\epsilon >0\). From this, one can show that in fact, \(F'(X)\) and \(F'(Y)\) are \((\zeta ,\gamma ^{\frac{6}{100}})\)-homotopy interleaved, using an argument similar to the proof of [49, Theorem 6.1].

Letting \(H'(X)\) and \(H'(Y)\) denote the respective restrictions of H(X) and H(Y) to \(J\) (i.e., \(H'(X)=H_1(F'(X))\) and \(H'(Y)=H_1(F'(Y))\)), we then have that \(H'(X)\) and \(H'(Y)\) are \((\zeta ,\gamma ^{\frac{6}{100}})\)-interleaved by Proposition 2.41. In what follows, we will show that this constrains certain vector spaces in \(H'(Y)\) to have dimension at least one.

Let \(\Delta \) denote the large triangle-like connected region in \(J\) where the Hilbert function of \(H'(X)\) takes value 1. The boundary of \(\Delta \) intersects the vertical line \(k=0\) and the horizontal line

$$\begin{aligned} r=\tfrac{7891389}{20000000}\approx .396. \end{aligned}$$

By inspecting the bigraded Betti numbers and fibered barcode of H(X), as shown in Figs. 2 and 3, it can be seen that \(\Delta \) is in fact is contained in the support of a thin indecomposable summand of \(H'(X)\). Thus, if \(a\le b\in \Delta \) (with respect to the partial order on \({\mathbb {R}}^{\mathrm {op}}\times {\mathbb {R}}\)), then \({{\,\mathrm{rank}\,}}({H'(X)}_{a,b})=1\).

Fig. 3
figure 3

RIVET’s visualization of H(X), zoomed in near (0, 0)

In particular, if \((k,r)\in \Delta \) and also

$$\begin{aligned} \gamma ^{\frac{6}{100}}\circ \zeta (k,r)=\left( \frac{19k}{20}-\frac{7}{100},3r+\frac{9}{100}\right) \in \Delta , \end{aligned}$$

then

$$\begin{aligned} {{\,\mathrm{rank}\,}}H'(X)_{(k,r),\gamma ^{\frac{6}{100}}\circ \zeta (k,r)}=1. \end{aligned}$$

Now if \((k,r)\in \Delta \), then \((\frac{19k}{20}-\frac{7}{100},3r+\frac{9}{100})\in \Delta \) if and only if \(k>c_x\) and \(r<c_y\), where

$$\begin{aligned} c_x=\tfrac{7}{95}\approx .074,\quad c_y=\left( \tfrac{2030463}{20000000}\right) \approx .102. \end{aligned}$$

From the output of RIVET, it can be seen that the set

$$\begin{aligned} \Omega =\{(k,r)\in \Delta \mid k>c_x,\ r<c_y\} \end{aligned}$$

is a small but non-empty connected subregion of \(\Delta \), containing the grades of three elements in any minimal set of generators for \(H'(X)\). \(\Omega \) is shown in Fig. 4A. Explicitly,

$$\begin{aligned} \Omega ={{\,\mathrm{Rect}\,}}(v_1)\cup {{\,\mathrm{Rect}\,}}(v_2)\cup {{\,\mathrm{Rect}\,}}(v_3), \end{aligned}$$

where

$$\begin{aligned} {{\,\mathrm{Rect}\,}}(a)&=\{(x,y)\mid a_1\ge x > c_x,\ a_2\le y < c_y\},\\ v_1&=\left( \tfrac{2657}{23750},\tfrac{1897929}{20000000}\right) \approx (.112,.095),\\ v_2&=\left( \tfrac{2183}{23750},\tfrac{1698147}{20000000}\right) \approx (.092,.085),\\ v_3&=\left( \tfrac{973}{11875},\tfrac{299673}{4000000}\right) \approx (.082,.075). \end{aligned}$$
Fig. 4
figure 4

The regions \(\Omega \) and \(\zeta (\Omega )\), shown in solid black. [Note: These regions were drawn by hand in software, and so are not as precise as if they had been drawn algorithmically. However, the imprecisions are miniscule.]

For any \((k,r)\in \Omega \), a \((\zeta ,\gamma ^{\frac{6}{100}})\)-interleaving between \(H'(X)\) and \(H'(Y)\) provides a factorization of the non-zero linear map \(H'(X)_{(k,r),(\frac{19k}{20}-\frac{7}{100},3r+\frac{9}{100})}\) through

$$\begin{aligned} H'(Y)_{\zeta (k,r)}=H'(Y)_{(\frac{19k}{20}-\frac{1}{100},r+\frac{1}{100})}. \end{aligned}$$

Therefore, the support of the Hilbert function of \(H'(Y)\) must contain \(\zeta (\Omega )\). \(\Omega \) is shown in Fig. 4B. Letting

$$\begin{aligned} d_x=\tfrac{6}{100},\quad d_y=\tfrac{2230463}{20000000}\approx .112, \end{aligned}$$

\(\zeta (\Omega )\) can be written explicitly as

$$\begin{aligned} \zeta (\Omega )={{\,\mathrm{Rect}\,}}(w_1)\cup {{\,\mathrm{Rect}\,}}(w_2)\cup {{\,\mathrm{Rect}\,}}(w_3), \end{aligned}$$

where

$$\begin{aligned} {{\,\mathrm{Rect}\,}}(a)&=\left\{ (x,y)\mid a_1\ge x > d_x ,\ a_2\le y < d_y\right\} ,\\ w_1&=\left( \tfrac{2407}{25000},\tfrac{2097929}{20000000}\right) \approx (.096,.105),\\ w_2&=\left( \tfrac{1933}{23750},\tfrac{1898147}{20000000}\right) \approx (.077,.095),\\ w_3&=\left( \tfrac{212}{3125},\tfrac{339673}{4000000}\right) \approx (.068,.085). \end{aligned}$$

Figure 4B indicates that the support of the Hilbert function of \(H'(Y)\) does indeed contain \(\zeta (\Omega )\), and in fact is much larger.

Remark A.1

We have shown that given H(X), Proposition 3.9 (ii) constrains the structure of H(Y). Using a similar argument, one can show that given \(H_1({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(X))\), Proposition 3.9 (ii) constrains the structure of \(H_1({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(Y))\). However, a similar argument also shows that given H(Y), Proposition 3.9 (ii) provides no constraint on H(X); and similarly, given \(H_1({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(Y))\), Proposition 3.9 (ii) provides no constraint on \(H_1({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(X))\).

Remark A.2

By Remark 2.16, \(d_{GPr}(X,Y)\le d_P(X,Y)\le \frac{25}{500}=\frac{1}{20}\). Starting from this observation, one can perform a stability analysis similar to the one done above, but using the weaker interleaving provided by Theorem 1.7 (ii) in place of the one provided by Proposition 3.9 (ii). It is not difficult to check that the interleaving between \(H'(X)\) and \(H'(Y)\) provided by such an analysis can be taken to be trivial. Further, using a similar argument, one can show that the interleavings between \(H_1({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(X))\) and \(H_1({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(Y))\) guaranteed to exist by Theorem 1.7 (ii) can also be taken to be trivial. Thus, the existence of this interleaving does not constrain the relationship between \(H_1({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(X))\) and \(H_1({{\,\mathrm{\mathcal{DR}\mathcal{}}\,}}(Y))\) at all. Because the shared topological signal present in X and Y is especially strong, this indicates that relative to the needs of applications, Theorem 1.7 is a rather weak result.

Remark A.3

Figure 2 makes clear that while H(X) and H(Y) have rather different global structure, they do share substantial qualitative similarities that neither module shares with H(Z). While our analysis demonstrates that Proposition 3.9 (ii) non-trivially constrains the relationship between H(X) and H(Y), most of the similarity between the two modules seen in Fig. 2 is not explained by the proposition (or, to the best of our knowledge, by any other known result). It would be valuable to develop a refinement of our stability theory which more fully explains the observed similarity.

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Blumberg, A.J., Lesnick, M. Stability of 2-Parameter Persistent Homology. Found Comput Math 24, 385–427 (2024). https://doi.org/10.1007/s10208-022-09576-6

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10208-022-09576-6

Keywords

Mathematics Subject Classification

Navigation