Skip to main content
Log in

Demystifying Latschev’s Theorem: Manifold Reconstruction from Noisy Data

  • Published:
Discrete & Computational Geometry Aims and scope Submit manuscript

Abstract

For a closed Riemannian manifold \(\mathcal {M}\) and a metric space S with a small Gromov–Hausdorff distance to it, Latschev’s theorem guarantees the existence of a sufficiently small scale \(\beta >0\) at which the Vietoris–Rips complex of S is homotopy equivalent to \(\mathcal {M}\). Despite being regarded as a stepping stone to the topological reconstruction of Riemannian manifolds from noisy data, the result is only a qualitative guarantee. Until now, it had been elusive how to quantitatively choose such a proximity scale \(\beta \) in order to provide sampling conditions for S to be homotopy equivalent to \(\mathcal {M}\). In this paper, we prove a stronger and pragmatic version of Latschev’s theorem, facilitating a simple description of \(\beta \) using the sectional curvatures and convexity radius of \(\mathcal {M}\) as the sampling parameters. Our study also delves into the topological recovery of a closed Euclidean submanifold from the Vietoris–Rips complexes of a Hausdorff close Euclidean subset. As already known for Čech complexes, we show that Vietoris–Rips complexes also provide topologically faithful reconstruction guarantees for submanifolds.

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.

We’re sorry, something doesn't seem to be working properly.

Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

References

  1. Adamaszek, M., Adams, H., Frick, F.: Metric reconstruction via optimal transport. SIAM J. Appl. Algebra Geom. 2, 597–619 (2018)

    Article  MathSciNet  Google Scholar 

  2. Adams, H., Mirth, J.: Metric thickenings of Euclidean submanifolds. Topol. Appl. 254, 69–84 (2019)

    Article  MathSciNet  Google Scholar 

  3. Alexander, S.B., Bishop, R.L.: Gauss equation and injectivity radii for subspaces in spaces of curvature bounded above. Geom. Dedicata. 117(1), 65–84 (2006)

    Article  MathSciNet  Google Scholar 

  4. Attali, D., Lieutier, A., Salinas, D.: Vietoris–Rips complexes also provide topologically correct reconstructions of sampled shapes. Comput. Geom. 46(4), 448–465 (2013). 27th Annual Symposium on Computational Geometry (SoCG 2011)

  5. Berger, M.: A Panoramic View of Riemannian Geometry. Springer, Berlin (2007)

    Google Scholar 

  6. Bishop, R.L., Crittenden, R.J.: Geometry of Manifolds, vol. 344. AMS Chelsea Publications, Providence (2001)

    Google Scholar 

  7. Chambers, E.W., de Silva, V., Erickson, J., Ghrist, R.: Vietoris-Rips complexes of planar point sets. Discrete Comput. Geom. 44(1), 75–90 (2010)

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  9. Chazal, F., Lieutier, A.: Stability and computation of topological invariants of solids in \(\mathbb{R} ^n\). Discrete Comput. Geom. 37(4), 601–617 (2007)

    Article  MathSciNet  Google Scholar 

  10. Danzer, L.: Helly’s theorem and its relatives, in convexity. In: Proc. Symp. Pure Math., vol. 7, pp. 101–180. American Mathematical Society, Providence (1963)

  11. Dekster, B.V.: An extension of Jung’s Theorem. Isr. J. Math. 50, 169–180 (1985)

    Article  MathSciNet  Google Scholar 

  12. Dekster, B.V.: The Jung Theorem in metric spaces of curvature bounded above. Proc. Am. Math. Soc. 125(8), 2425–2433 (1997)

    Article  MathSciNet  Google Scholar 

  13. Dekster, B.V.: The Jung Theorem for spherical and hyperbolic spaces. Acta Math. Hungar. 67, 315–331 (1995)

    Article  MathSciNet  Google Scholar 

  14. do Carmo, M.: Riemannian Geometry. Trans. F. Flaherty. Mathematics, Theory and Applications. Birkhäuser, Boston (1992)

  15. Fasy, B.T., Komendarczyk, R., Majhi, S., Wenk, C.: On the reconstruction of geodesic subspaces of \(\mathbb{R} ^n\). Int. J. Comput. Geom. Appl. 32(01–02), 91–117 (2022)

    Article  Google Scholar 

  16. Hausmann, J.-C.: On the Vietoris–Rips Complexes and a cohomology theory for metric spaces. In: Quinn, F. (ed.) Prospects in Topology (AM-138), Proceedings of a Conference in Honor of William Browder (AM-138), pp. 175–188. Princeton University Press, Princeton (1995)

  17. Kim, J., Shin, J., Chazal, F., Rinaldo, A., Wasserman, L.: Homotopy reconstruction via the Čech complex and the Vietoris–Rips complex. In: 36th International Symposium on Computational Geometry (SoCG 2020). Schloss Dagstuhl-Leibniz-Zentrum für Informatik (2020)

  18. Latschev, J.: Vietoris-Rips complexes of metric spaces near a closed Riemannian manifold. Arch. Math. 77(6), 522–528 (2001)

    Article  MathSciNet  Google Scholar 

  19. Lemež, B., Virk, Ž.: Finite reconstruction with selective Rips complexes. arXiv Preprint (2022). arXiv:2205.05525

  20. Majhi, S.: Vietoris-Rips complexes of metric spaces near a metric graph. J. Appl. Comput. Topol. (2023). https://doi.org/10.1007/s41468-023-00122-z

    Article  MathSciNet  Google Scholar 

  21. Munkres, J.R.: Elements of Algebraic Topology, 2nd edn. Addison-Wesley, Menlo Park (1996)

    Google Scholar 

  22. Niyogi, P., Smale, S., Weinberger, S.: Finding the homology of submanifolds with high confidence from random samples. Discret. Comput. Geom. 39(1–3), 419–441 (2008)

    Article  MathSciNet  Google Scholar 

  23. Spanier, E.H.: Algebraic Topology, vol. 55. Springer, New York (1994)

    Google Scholar 

  24. Vietoris, L.: Über den höheren zusammenhang kompakter räume und eine klasse von zusammenhangstreuen abbildungen. Math. Ann. 97, 454–472 (1927)

    Article  MathSciNet  Google Scholar 

  25. Virk, Ž: Rips complexes as nerves and a functorial Dowker-nerve diagram. Mediterr. J. Math. 18(2), 1–24 (2021)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The author would like to thank Henry Adams for providing constructive feedback on the manuscript, and for hosting the author at the University of Florida, Gainesville during the summer of 2023.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sushovan Majhi.

Additional information

Editor in Charge: Kenneth Clarkson

Publisher's Note

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

Appendix A: Additional Proofs

Appendix A: Additional Proofs

Proof of Proposition 4.3

We divide the proof into the following two cases depending on the sign of \(\kappa :=\kappa (\mathcal {M})\).

Case 1 (\(\kappa \le 0\)) From (2), we get

$$\begin{aligned} \textrm{diam}_{\mathcal {M}}\left( A\right) \ge 2\mathfrak {R}(A)\sqrt{\frac{n+1}{2n}}\ge 2\mathfrak {R}(A)\frac{1}{\sqrt{2}}\ge \frac{4}{3}\mathfrak {R}(A). \end{aligned}$$

Case 2 (\(\kappa >0\)) First note that the closed ball \(\overline{{\mathbb B}_\mathcal {M}(p,\textrm{diam}_{\mathcal {M}}\left( A\right) )}\) for any \(p\in A\) contains A entirely, implying that \(\mathfrak {R}(A)\le \textrm{diam}_{\mathcal {M}}\left( A\right) \). Since \(\textrm{diam}_{\mathcal {M}}\left( A\right) <\Delta (\mathcal {M})\le \frac{\pi }{4\sqrt{\kappa }}\), we therefore have \(\mathfrak {R}(A)\in \left[ 0,\frac{\pi }{4\sqrt{\kappa }}\right] \).

Let us now define the function \(J(r){\mathop {=}\limits ^{def}}\frac{2}{\sqrt{\kappa }}\sin ^{-1}\left( \sqrt{\frac{n+1}{2n}}\sin {\sqrt{\kappa }r}\right) \), and show that J(r)/r is a strictly decreasing function for \(r\in \left( 0,\frac{\pi }{4\sqrt{\kappa }}\right) \). We take its derivative:

$$\begin{aligned} \frac{d}{dr}(J(r)/r)=\frac{rJ'(r)-J(r)}{r^2}=\frac{f(r)}{r^2}, \end{aligned}$$
(A1)

where \(f(r){\mathop {=}\limits ^{def}}rJ'(r)-J(r)\). We observe that f(r) is strictly decreasing, since

$$\begin{aligned} f'(r)=rJ''(r)=r\frac{2\sqrt{\kappa }\sqrt{\frac{n+1}{2n}}\sin {\sqrt{\kappa }r}}{\left( 1-\frac{n+1}{2n}\sin ^2{\sqrt{\kappa }r}\right) ^{3/2}}\left( \frac{1-n}{2n}\right) <0 \end{aligned}$$

for \(r\in \left( 0,\frac{\pi }{4\sqrt{\kappa }}\right) \) and \(n\ge 2\). Then it follows from \(f(0)=0\) that \(f(r)<0\) on \(r\in \left( 0,\frac{\pi }{4\sqrt{\kappa }}\right) \). When \(n=1\), we remark that the result can be established using Case 1, since \(\mathcal {M}\) must be flat.

So, (A1) implies that the function J(r)/r is a strictly decreasing function for \(r\in \left( 0,\frac{\pi }{4\sqrt{\kappa }}\right) \). Its minimum, as a result, is attained at \(r=\frac{\pi }{4\sqrt{\kappa }}\). So, for \(r\in \left( 0,\frac{\pi }{4\sqrt{\kappa }}\right] \) we get

$$\begin{aligned} \frac{J(r)}{r}\ge \frac{2}{\sqrt{\kappa }}\sin ^{-1}\left( \sqrt{\frac{n+1}{2n}}\sin {\frac{\pi }{4}}\right) \frac{4\sqrt{\kappa }}{\pi } \ge \frac{8}{\pi }\sin ^{-1}\left( \sqrt{\frac{1}{2}}\sin {\frac{\pi }{4}}\right) =\frac{8}{\pi }\frac{\pi }{6}=\frac{4}{3}. \end{aligned}$$

On the other hand, (2) implies that \(\textrm{diam}_{\mathcal {M}}\left( A\right) \ge J(\mathfrak {R}(A))\) for \(\mathfrak {R}(A)\in \left[ 0,\frac{\pi }{4\sqrt{\kappa }}\right] \). Hence, \(\textrm{diam}_{\mathcal {M}}\left( A\right) \ge \frac{4}{3}\mathfrak {R}(A)\) as desired.

\(\square \)

Proposition 4.1

(Path-connectedness) Let \((S,d_S)\) be a metric space and \(\beta >0\) a number such that \(d_{GH}(S,\mathcal {M})<\beta \), then for any positive \(\alpha \), the geometric complex of \(\mathcal {R}_{\alpha +2\beta }(S)\) is path-connected. The result also holds if \(M\subset {\mathbb R}^d\) and \(S\subset {\mathbb R}^d\) with \(d_H(\mathcal {M},S)<\beta \).

Proof

Let \(a,b\in S\), then there exist points \(a',b'\in \mathcal {M}\) such that \((a',a),(b',b)\in \mathcal {C}\), where \(\mathcal {C}\in \mathcal {C}(\mathcal {M},S)\) is a correspondence with \(\textrm{dist}(\mathcal {C})<2\beta \). Since \(\mathcal {M}\) is assumed to be path-connected, so is \(\mathcal {R}_\alpha (\mathcal {M})\). As a result, there exists a sequence \(\{p_i'\}_{i=0}^{m+1}\subset \mathcal {M}\) forming a path in \(\mathcal {R}_\alpha (\mathcal {M})\) joining \(a'\) and \(b'\). In other words, \(p_0'=a'\), \(p_{m+1}'=b'\), and \(d_\mathcal {M}(p_i',p_{i+1}')<\alpha \) for \(0\le i\le m\). There is also a corresponding sequence \(\{x_i\}_{i=0}^{m+1}\subset S\) such that \(x_0=a\), \(x_{m+1}=b\), and \((p_i',x_i)\in \mathcal {C}\) for all i. We note that

$$\begin{aligned} d_S(x_i,x_{i+1})\le d_\mathcal {M}(p_i',p_{i+1}')+2\beta <\alpha +2\beta . \end{aligned}$$

So, the sequence \(\{x_i\}\) produces a path in \(\mathcal {R}_{\alpha +2\beta }(S)\) joining a and b. We conclude that the geometric complex of \(\mathcal {R}_{\alpha +2\beta }(S)\) is path-connected.

The Hausdorff case uses the exact same argument. \(\square \)

Proof of Proposition 4.2

In [22, Prop. 6.1], it has been shown that \(\left\Vert L_\eta \right\Vert \le 1/\tau (\mathcal {M})\) for any unit norm \(\eta \in T^\perp _p(\mathcal {M})\).

  1. (i)

    Take \(\eta =B(u,v)/\left\Vert B(u,v) \right\Vert \), to get \(\left\Vert B(u,v) \right\Vert =\big \langle \eta , B(u,v)\big \rangle =\big \langle u, L_\eta (v)\big \rangle \). Since uv are unit norm, applying the Cauchy–Schwarz inequality we get

    $$\begin{aligned} \left\Vert B(u,v) \right\Vert \le \left\Vert u \right\Vert \left\Vert L_\eta \right\Vert \left\Vert v \right\Vert =\left\Vert L_\eta \right\Vert \le \frac{1}{\tau (\mathcal {M})}. \end{aligned}$$
  2. (ii)

    Use (6), then apply the Cauchy–Schwarz inequality:

    $$\begin{aligned} \kappa _p(u,v)\le \langle B(u,u),B(v,v)\rangle \le \left\Vert B(u,u) \right\Vert \cdot \left\Vert B(v,v) \right\Vert \le \frac{1}{\tau (\mathcal {M})^2}. \end{aligned}$$

    So, \(\kappa (\mathcal {M})\le 1/\tau (\mathcal {M})^2\). Similarly from (6), we get

    $$\begin{aligned} -\kappa _p(u,v)\le \left\Vert B(u,v) \right\Vert ^2\le \frac{1}{\tau (\mathcal {M})^2}. \end{aligned}$$

    So, \(\kappa (\mathcal {M})\ge -1/\tau (\mathcal {M})^2\).

  3. (iii)

    From [5, Prop. 95], we have

    $$\begin{aligned} \rho (\mathcal {M})\ge {\left\{ \begin{array}{ll} \min \left\{ \frac{\pi }{2\sqrt{\kappa (\mathcal {M})}},\frac{1}{2}\;\textrm{inj}(\mathcal {M})\right\} , &{}\text { for }\kappa (\mathcal {M})>0\\ \frac{1}{2}\;\textrm{inj}(\mathcal {M}),&{}\text { for }\kappa (\mathcal {M})\le 0. \end{array}\right. } \end{aligned}$$

    See [5, Defn. 23] for the definition of injectivity radius \(\textrm{inj}(\mathcal {M})\). On the other hand, [3, Corr. 4] shows that \(\textrm{inj}(\mathcal {M})\ge \pi \tau (\mathcal {M})\). This, combining with (ii), implies that \(\rho (\mathcal {M})\ge \pi \tau (\mathcal {M})/2\).

  4. (iv)

    Follows directly from (1) due to (iii).

\(\square \)

Proof of Proposition 4.3

Let \(\gamma :[0,L]\rightarrow \mathcal {M}\) be an arc-length parametrized, unit speed, length-minimizing geodesic joining p and q, where \(L=d_\mathcal {M}(p,q)\). From the fundamental theorem of calculus, we have

$$\begin{aligned} \gamma (L)-\gamma (0)=\int _0^L \gamma '(t)dt, \end{aligned}$$

and

$$\begin{aligned} \gamma '(t)=\gamma '(0)+\int _0^t \gamma ''(s)ds. \end{aligned}$$

Combining, we can write

$$\begin{aligned} \gamma (L)-\gamma (0)=\int _0^L\left[ \gamma '(0)+\int _0^t \gamma ''(s)ds\right] dt =L\gamma '(0) + \int _0^L\left[ \int _0^t \gamma ''(s) ds\right] dt. \end{aligned}$$

From the triangle inequality and the fact that \(\gamma \) is unit speed,

$$\begin{aligned} \left\Vert \gamma (L)-\gamma (0) \right\Vert \ge L\left\Vert \gamma '(0) \right\Vert - \left\Vert \int _0^L\left[ \int _0^t \gamma ''(s) ds\right] dt \right\Vert \ge L-\int _0^L\left[ \int _0^t\left\Vert \gamma ''(s) \right\Vert ds\right] dt. \end{aligned}$$

We note that \(\left\Vert p-q \right\Vert =\left\Vert \gamma (L)-\gamma (0) \right\Vert \) and \(L=d_\mathcal {M}(p,q)\). Since \(\gamma \) is geodesic, \(\gamma ''(s)=B\left( \gamma '(s),\gamma '(s)\right) \) for all s. So,

$$\begin{aligned} \left\Vert p-q \right\Vert \ge L-\int _0^L\left[ \int _0^t\left\Vert B(\gamma '(s),\gamma '(s)) \right\Vert ds\right] dt. \end{aligned}$$

On the other hand, \(\left\Vert B(u,u) \right\Vert \le 1/\tau (\mathcal {M})\) from Proposition 4.2 implies that

$$\begin{aligned} \left\Vert p-q \right\Vert \ge L-\frac{1}{\tau (\mathcal {M})}\int _0^L\left[ \int _0^t ds\right] dt =L-\frac{1}{\tau (\mathcal {M})}\frac{L^2}{2} =d_\mathcal {M}(p,q)-\frac{1}{2\tau (\mathcal {M})}[d_\mathcal {M}(p,q)]^2. \end{aligned}$$

The inequality is satisfied only if \(\left\Vert p-q \right\Vert \le \tau (\mathcal {M})/2\) and

$$\begin{aligned} d_\mathcal {M}(p,q)\le \tau (\mathcal {M})-\tau (\mathcal {M})\sqrt{1-\frac{2\left\Vert p-q \right\Vert }{\tau (\mathcal {M})}}. \end{aligned}$$
(A2)

Let \(f(r){\mathop {=}\limits ^{def}}\tau (\mathcal {M})-\tau (\mathcal {M})\sqrt{1-\frac{2r}{\tau (\mathcal {M})}}\) for \(0< r<2\left( \frac{\xi -1}{\xi ^2}\right) \tau (\mathcal {M})\). We note by taking derivative that \(\frac{f(r)}{r}\) is an increasing function. The maximum is attained for \(r=2\left( \frac{\xi -1}{\xi ^2}\right) \tau (\mathcal {M})\). Therefore,

$$\begin{aligned} \frac{f(r)}{r} \le \frac{1-\frac{\sqrt{(2-\xi )^2}}{\xi }}{\frac{2(\xi -1)}{\xi ^2}} =\xi \frac{\xi -\sqrt{(2-\xi )^2}}{2(\xi -1)} =\xi \frac{\xi -(2-\xi )}{2(\xi -1)} =\xi ,\text { since }1<\xi <2. \end{aligned}$$

Hence, (A2) implies

$$\begin{aligned} d_\mathcal {M}(p,q)\le f(\left\Vert p-q \right\Vert ) \le \xi \left\Vert p-q \right\Vert ,\text { since }\left\Vert p-q \right\Vert \le 2\left( \frac{\xi -1}{\xi ^2}\right) \tau (\mathcal {M})\le \tau (\mathcal {M})/2. \end{aligned}$$

\(\square \)

Proposition 4.2

(Commuting Diagram) Let \(\mathcal {K}\) be a pure m-complex and \(\mathcal {L}\) a flag complex. Let \(f:~\mathcal {K}\xrightarrow {\quad \quad }\mathcal {L}\) and \(g:\textrm{sd}^{}\left( \mathcal {K}\right) \xrightarrow {\quad \quad }\mathcal {L}\) be simplicial maps such that

  1. (a)

    \(g(v)=f(v)\) for every vertex v of \(\mathcal {K}\),

  2. (b)

    \(f(\sigma )\cup g(\widehat{\sigma })\) is a simplex of \(\mathcal {L}\) whenever \(\sigma \) is a simplex of \(\mathcal {K}\).

Then, the following diagram commutes up to homotopy:

figure e

where h is a linear homeomorphism sending each vertex of \(\textrm{sd}^{}\left( \mathcal {K}\right) \) to the corresponding point of \(\big |\mathcal {K} \big |\).

Proof

We show that the maps \(\big |g \big |\) and \(\left( \big |f \big |\circ h\right) \) are homotopic by constructing an explicit homotopy \(H:~\big |\textrm{sd}^{}\left( \mathcal {K}\right) \big |\times [0,1]\xrightarrow {\quad \quad }\big |\mathcal {L} \big |\) with \(H(\cdot ,0)=g(\cdot )\) and \(H(\cdot ,1)=\left( \big |f \big |\circ h\right) (\cdot )\). The commutativity of the diagram then follows, since h is a homeomorphism.

The complex \(\mathcal {K}\) is taken to be m-dimensional. Without any loss of generality, we can assume that every point of \(\textrm{sd}^{}\left( \mathcal {K}\right) \) belongs to an m-simplex. Take an arbitrary \(x\in \big |\textrm{sd}^{}\left( \mathcal {K}\right) \big |\). We can write x in its barycentric coordinates as \(x=\sum _{i=0}^m\zeta _i\widehat{\sigma }_i\), where \(\sigma _m=[a_0,a_1,\ldots ,a_m]\) is an m-simplex of \(\mathcal {K}\) and \(\sigma _i=[a_0,a_1,\ldots ,a_i]\) for \(0\le i\le m\). Consider a partition

$$\begin{aligned}0=t_0<t_1<\cdots <t_{m}=1\end{aligned}$$

of [0, 1]. We first define the homotopy \(H(\cdot ,t)\) at \(t=t_i\) for \(i=0,1,\ldots ,m\), and then show that we can interpolate H continuously (using the straight-line homotopy) for any \(t\in [t_{i-1},t_i]\). For any \(0\le i\le m\), define

$$\begin{aligned} H(x,t_i)=\sum _{j=0}^i\zeta _j\left( \big |f \big |\circ h\right) (\widehat{\sigma }_j)+\sum _{j=i+1}^m\zeta _j\big |g \big |(\widehat{\sigma }_j). \end{aligned}$$
(A3)

From the above definition, we note for \(i=0\) that

$$\begin{aligned} H(x,0)=\zeta _0\left( \big |f \big |\circ h\right) (\widehat{\sigma }_0)+\sum _{j=1}^m\zeta _j\big |g \big |(\widehat{\sigma }_j)&=\zeta _0\big |g \big |(\widehat{\sigma }_0) +\sum _{j=1}^m\zeta _j\big |g \big |(\widehat{\sigma }_j) \\ {}&=\sum _{j=0}^m\zeta _j\big |g \big |(\widehat{\sigma }_j)=\big |g \big |(x). \end{aligned}$$

The second equality is due to condition (a) and fact that \(\sigma _0\) is a vertex of \(\mathcal {K}\). Also, for \(i=m\) we note that

$$\begin{aligned} H(x,1)=\sum _{j=0}^m\zeta _j\left( \big |f \big |\circ h\right) (\widehat{\sigma }_j) =\left( \big |f \big |\circ h\right) \left( \sum _{j=0}^m\zeta _j\widehat{\sigma }_j\right) =\left( \big |f \big |\circ h\right) (x). \end{aligned}$$

The second equality is due to the fact that h is a linear homeomorphism.

We now fix \(0\le i\le m\), and extend the definition of H(xt) for any \(t\in [t_{i-1},t_i]\) by using the straight-line joining \(H(x,t_{i-1})\) and \(H(x,t_{i})\). Such an extension is justified, we show now both \(H(x,t_{i-1})\) and \(H(x,t_{i})\) belong to \(\big |\eta \big |\) for some simplex \(\eta \) of \(\mathcal {L}\).

Let us take \(\eta \) to be the set \(f(\sigma _i) \cup g([\widehat{\sigma }_i,\widehat{\sigma }_{i+1},\ldots ,\widehat{\sigma }_m])\). Since g is a simplicial map, \(\eta \) can be written as \(f(\sigma _i)\cup [g(\widehat{\sigma }_i),g(\widehat{\sigma }_{i+1}),\ldots ,g(\widehat{\sigma }_m)]\). Now, note from condition (b) that \(f(\sigma _j)\cup g(\widehat{\sigma }_j)\in \mathcal {L}\) for any \(i\le j\le m\). Since f is a simplicial map, \(f(\sigma _i)\cup g(\widehat{\sigma }_j)\) is a face of \(f(\sigma _j)\cup g(\widehat{\sigma }_j)\), hence a simplex of \(\mathcal {L}\). Consequently, any pair of elements in the set \(\eta =f(\sigma _i)\cup [g(\widehat{\sigma }_i),g(\widehat{\sigma }_{i+1}),\ldots ,g(\widehat{\sigma }_m)]\) is a simplex of \(\mathcal {L}\). Since \(\mathcal {L}\) is a flag complex, it implies that \(\eta \) is a simplex of \(\mathcal {L}\).

From the definition (A3) and the fact that \(h(\widehat{\sigma }_{l})\in \big |\sigma _i \big |\) for all \(0\le l\le i\), we conclude that both \(H(x,t_{i-1})\) and \(H(x,t_{i})\) belong to \(\big |\eta \big |\). So, H(xt) is well-defined and continuous for \(t\in [t_{i-1},t_i]\). Moreover, the images agree at the endpoints as noted from (A3). Therefore, H defines the desired homotopy. \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) 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

Majhi, S. Demystifying Latschev’s Theorem: Manifold Reconstruction from Noisy Data. Discrete Comput Geom (2024). https://doi.org/10.1007/s00454-024-00655-9

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00454-024-00655-9

Keywords

Navigation