Skip to main content
Log in

Rigidity of Riemannian embeddings of discrete metric spaces

  • Published:
Inventiones mathematicae Aims and scope

Abstract

Let M be a complete, connected Riemannian surface and suppose that \({\mathcal {S}}\subset M\) is a discrete subset. What can we learn about M from the knowledge of all Riemannian distances between pairs of points of \({\mathcal {S}}\)? We prove that if the distances in \({\mathcal {S}}\) correspond to the distances in a 2-dimensional lattice, or more generally in an arbitrary net in \({\mathbb {R}}^2\), then M is isometric to the Euclidean plane. We thus find that Riemannian embeddings of certain discrete metric spaces are rather rigid. A corollary is that a subset of \({\mathbb {Z}}^3\) that strictly contains \({\mathbb {Z}}^2 \times \{ 0 \}\) cannot be isometrically embedded in any complete Riemannian surface.

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

References

  1. Álvarez Paiva, J.C., Berck, G.: What is wrong with the Hausdorff measure in Finsler spaces. Adv. Math. 204(2), 647–663 (2006)

    Article  MathSciNet  Google Scholar 

  2. Ballman, W., Gromov, M., Schroeder, V.: Manifolds of nonpositive curvature. In: Progress in Mathematics, vol. 61. Birkhäuser (1985)

  3. Bangert, V., Emmerich, P.: On the flatness of Riemannian cylinders without conjugate points. Commun. Anal. Geom. 19(4), 773–805 (2011)

    Article  MathSciNet  Google Scholar 

  4. Bangert, V., Emmerich, P.: Area growth and rigidity of surfaces without conjugate points. J. Differ. Geom. 94(3), 367–385 (2013)

    Article  MathSciNet  Google Scholar 

  5. Belkin, M., Niyogi, P.: Convergence of Laplacian eigenmaps. In: Neural Information Processing Systems (NIPS ’06), pp. 19–136 (2006)

  6. Burago, D., Burago, Y., Ivanov, S.: A course in metric geometry. In: Graduate Studies in Mathematics, vol. 33. American Mathematical Society, Providence, RI (2001)

  7. Burago, D., Ivanov, S.: Riemannian tori without conjugate points are flat. Geom. Funct. Anal. (GAFA) 4(3), 259–269 (1994)

    Article  MathSciNet  Google Scholar 

  8. Burago, D., Ivanov, S.: Boundary distance, lens maps and entropy of geodesic flows of Finsler metrics. Geom. Topol. 20(1), 469–490 (2016)

    Article  MathSciNet  Google Scholar 

  9. Burago, D., Ivanov, S., Kurylev, Y.: A graph discretization of the Laplace-Beltrami operator. J. Spectr. Theory 4(4), 675–714 (2014)

    Article  MathSciNet  Google Scholar 

  10. Burns, K., Knieper, G.: Rigidity of surfaces with no conjugate points. J. Differ. Geom. 34(3), 623–650 (1991)

    Article  MathSciNet  Google Scholar 

  11. Busemann, H.: The Geometry of Geodesics. Academic Press Inc., New York (1955)

    MATH  Google Scholar 

  12. Cheeger, J., Ebin, D.G.: Comparison Theorems in Riemannian Geometry, North-Holland Mathematical Library, vol. 9. Elsevier, Amsterdam (1975)

    MATH  Google Scholar 

  13. Coiffman, R., Lafon, S.: Diffusion maps. Appl. Comput. Harmon. Anal. 21(1), 5–30 (2006)

    Article  MathSciNet  Google Scholar 

  14. Croke, C.B.: Volumes of balls in manifolds without conjugate points. Int. J. Math. 3(4), 455–467 (1992)

    Article  MathSciNet  Google Scholar 

  15. Croke, C.B.: A synthetic characterization of the hemisphere. Proc. Am. Math. Soc. 136(3), 1083–1086 (2008)

    Article  MathSciNet  Google Scholar 

  16. Evans, L.C., Gangbo, W.: Differential equations methods for the Monge-Kantorovich mass transfer problem. Mem. Am. Math. Soc. 137, 653 (1999)

    MathSciNet  MATH  Google Scholar 

  17. Fefferman, C., Ivanov, S., Kurylev, Y., Lassas, M., Narayanan, H.: Reconstruction and interpolation of manifolds, I: the geometric whitney problem. Found. Comput. Math. 2019, 1–99 (2019)

    MATH  Google Scholar 

  18. Fefferman, C., Ivanov, S., Lassas M., Narayanan, H.: Reconstruction of a Riemannian manifold from noisy intrinsic distances. arXiv:1905.07182

  19. Feldman, M., McCann, R.J.: Monge’s transport problem on a Riemannian manifold. Trans. Am. Math. Soc. 354(4), 1667–1697 (2002)

    Article  MathSciNet  Google Scholar 

  20. Fujiwara, K.: Eigenvalues of Laplacians on a closed Riemannian manifold and its nets. Proc. Am. Math. Soc. 123(8), 2585–2594 (1995)

    Article  MathSciNet  Google Scholar 

  21. Hopf, E.: Closed surfaces without conjugate points. Proc. Nat. Acad. Sci. USA 34, 47–51 (1948)

    Article  MathSciNet  Google Scholar 

  22. Ivanov, S.: On two-dimensional minimal fillings. St. Petersburg Math. J. 13(1), 17–25 (2002)

    MathSciNet  MATH  Google Scholar 

  23. Klartag, B.: Needle decompositions in Riemannian geometry. Mem. Am. Math. Soc. 249, 1180 (2017)

    MathSciNet  MATH  Google Scholar 

  24. Kloeckner, B.R., Kuperberg, G.: The Cartan-Hadamard conjecture and the Little Prince. Rev. Mat. Iberoam. 35(4), 1195–1258 (2019)

    Article  MathSciNet  Google Scholar 

  25. Michel, R.: Sur la rigidité imposée par la longueur des géodésiques. French. Invent. Math. 65(1), 71–83 (1981)

    Article  Google Scholar 

  26. Morse, M., Hedlund, G.A.: Manifolds without conjugate points. Trans. Am. Math. Soc. 51, 362–386 (1942)

    Article  MathSciNet  Google Scholar 

  27. Pestov, L., Uhlmann, G.: Two dimensional compact simple Riemannian manifolds are boundary distance rigid. Ann. Math. 161(2), 1093–1110 (2005)

    Article  MathSciNet  Google Scholar 

  28. Singer, A.: From graph to manifold Laplacian: the convergence rate. Appl. Comput. Harmon. Anal. 21(1), 128–134 (2006)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The second-named author would like to thank Charles Fefferman for interesting discussions on possible Riemannian analogs of Whitney’s extension problem, and to Adrian Nachman for excellent explanations on the boundary rigidity problem and other inverse problems. Both authors thank Itai Benjamini for his interest and encouragement. Supported by a grant from the Israel Science Foundation (ISF).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Bo’az Klartag.

Additional information

Publisher's Note

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

Appendix: Continuity of area

Appendix: Continuity of area

Suppose that for any \(m \ge 1\) we are given a metric \(d_m\) on \({\mathbb {R}}^2\), such that the following hold:

  1. (i)

    For any m, the metric \(d_m\) induces the standard topology on \({\mathbb {R}}^2\).

  2. (ii)

    For any \(x,y \in {\mathbb {R}}^2\) we have \(d_m(x,y) \longrightarrow |x-y|\) as \(m \rightarrow \infty \), and the convergence is locally uniform.

  3. (iii)

    For any m, the metric space \(({\mathbb {R}}^2, d_m)\) is isometric to a complete, connected, 2-dimensional Riemannian manifold in which all geodesics are minimizing.

Write \(\text {Area}_m\) for the Riemannian area measure on \({\mathbb {R}}^2\) that corresponds to \(d_m\) under the above isometry.

Proposition 7.1

For \(D = D(0,1) =\{ x \in {\mathbb {R}}^2 \, ; \, |x| < 1 \}\) we have \(\text {Area}_m(D) \longrightarrow \pi \) as \(m \rightarrow \infty \).

We were unable to find a proof of Proposition 7.1 in the literature, even though Ivanov’s paper [22] contains a deeper result which “almost” implies this proposition. A proof of Proposition 7.1 is thus provided in this Appendix.

By a \(d_m\)-geodesic in \({\mathbb {R}}^2\) we mean a geodesic with respect to the metric \(d_m\). The \(d_m\)-length of a \(d_m\)-rectifiable curve \(\gamma \) is denoted by \(\text {Length}_m(\gamma )\). A set \(K \subseteq {\mathbb {R}}^2\) is \(d_m\)-convex if the intersection of any \(d_m\)-geodesic with K is connected. All \(d_m\)-geodesics are minimizing, and each complete \(d_m\)-geodesic divides \({\mathbb {R}}^2\) into two connected components. Each of these connected components is a \(d_m\)-convex, open set called a \(d_m\)-half-plane. The intersection of finitely many \(d_m\)-half-planes, if bounded and non-empty, is called a \(d_m\)-polygon. Note that our polygons are always open and convex. The boundary of any \(d_m\)-polygon consists of finitely many vertices and the same number of edges, and each edge is a \(d_m\)-geodesic segment.

Write \({\mathcal {G}}_m\) for the collection of all complete \(d_m\)-geodesics in \({\mathbb {R}}^2\), where we identify between two geodesics if they differ by an orientation-preserving reparametrization. Write \(\sigma _m\) for the Liouville (or étendue) measure on \({\mathcal {G}}_m\), see Kloeckner and Kuperberg [24, Section 5.2] and Álvarez-Paiva and Berck [1, Section 5]) and references therein for the basic properties of this measure, and for the formulae of Santaló and Crofton from integral geometry. The Santaló formula implies that for any open set \(U \subseteq {\mathbb {R}}^2\),

$$\begin{aligned} \text {Area}_m( U ) = \frac{1}{2 \pi } \int _{{\mathcal {G}}_m} \text {Length}_m(\gamma \cap U ) d \sigma _m(\gamma ). \end{aligned}$$

(We remark that \(\gamma \cap U\) can be disconnected, yet it is a disjoint, countable union of \(d_m\)-geodesics, and \(\text {Length}_m(\gamma \cap U)\) is the sum of the \(d_m\)-lengths of these \(d_m\)-geodesics). The Crofton formula implies that for any \(d_m\)-polygon \(P \subseteq {\mathbb {R}}^2\),

$$\begin{aligned} \text {Perimeter}_m( P ) = \frac{1}{2} \cdot \sigma _m \left( \left\{ \gamma \in {\mathcal {G}}_m \, ; \, \gamma \cap P \ne \emptyset \right\} \right) , \end{aligned}$$

where \(\text {Perimeter}_m(P) = \text {Length}_m(\partial P)\). When we discuss \(\text {Area}, \text {Length}, \text {Perimeter}\) or polygons without the subscript m we refer to the usual Euclidean geometry in \({\mathbb {R}}^2\). Write \({\mathcal {G}}\) for the collection of all lines in \({\mathbb {R}}^2\), where we identify between two lines if they differ by an orientation-preserving reparametrization. Write \(\sigma \) for the Euclidean Liouville measure on \({\mathcal {G}}\). We require the following Euclidean lemma:

Lemma 7.2

Let \({\tilde{\sigma }}\) be a Borel measure on \({\mathcal {G}}\) such that for any convex polygon \(P \subseteq D(0,2) \subseteq {\mathbb {R}}^2\),

$$\begin{aligned} \mathrm{perimeter}( P ) = \frac{1}{2} \cdot {\tilde{\sigma }} \left( \left\{ \ell \in {\mathcal {G}}\, ; \, \ell \cap P \ne \emptyset \right\} \right) . \end{aligned}$$
(74)

Then

$$\begin{aligned} \frac{1}{2 \pi } \int _{{\mathcal {G}}} \mathrm{length}(\ell \cap D ) d {\tilde{\sigma }}(\ell ) = \mathrm{area}(D) = \pi . \end{aligned}$$
(75)

Proof

For any rotation \(U \in SO(2)\), the perimeter of the rotated polygon \(U(P) \subset {\mathbb {R}}^2\) is the same as the perimeter of P. Hence formula (74) holds true with \({\tilde{\sigma }}\) replaced by \(U_* {\tilde{\sigma }}\), where by \(U_* {\tilde{\sigma }}\) we mean the push-forward of \({\tilde{\sigma }}\) under the map U acting on \({\mathcal {G}}\) by rotating lines. Moreover, since \(U(D) = D\), replacing \({\tilde{\sigma }}\) by \(U_* {\tilde{\sigma }}\) does not change the value of the integral on the left-hand side of (75).

We may thus replace the measure \({\tilde{\sigma }}\) by the average of \(U_* {\tilde{\sigma }}\) over \(U \in SO(2)\), and assume from now on that \({\tilde{\sigma }}\) is a rotationally-invariant measure on \({\mathcal {G}}\). The validity of (74) for any convex polygon \(P \subset D(0,2)\) implies its validity for all convex sets in the disc D(0, 2). Indeed, both the left-hand side and the right-hand side of (74) are monotone in the convex set P under inclusion, and convex polygons are dense in the class of all convex subsets of D(0, 2). Consequently, for any \(0< \rho < 2\),

$$\begin{aligned} 2 \pi \rho = \text {Perimeter}(D(0,\rho )) = \frac{1}{2} \cdot {\tilde{\sigma }} \left( \left\{ \ell \in {\mathcal {G}}\, ; \, \ell \cap D(0,\rho ) \ne \emptyset \right\} \right) . \end{aligned}$$
(76)

For \(\ell \in {\mathcal {G}}\) write \(r(\ell ) = \inf _{x \in \ell } |x| \in [0, \infty )\). We may reformulate (76) as follows:

$$\begin{aligned} {\tilde{\sigma }} \left( \left\{ \ell \in {\mathcal {G}}\, ; \, r(\ell )< \rho \right\} \right) = \sigma \left( \left\{ \ell \in {\mathcal {G}}\, ; \, r(\ell )< \rho \right\} \right) \qquad \text{ for } \text{ any } 0< \rho < 2. \qquad \end{aligned}$$
(77)

Since both \(\sigma \) and \({\tilde{\sigma }}\) are rotationally-invariant measures on \({\mathcal {G}}\), they are completely determined by their push-forward under the map \(\ell \mapsto r(\ell )\). From (77) we learn that \({\tilde{\sigma }}\) coincides with \(\sigma \) on the set \({\mathcal {G}}\cap r^{-1}([0,2))\). By the Satanló formula for \(\sigma \),

$$\begin{aligned} \frac{1}{2\pi } \int _{{\mathcal {G}}} \text {Length}(\ell \cap D) d {\tilde{\sigma }}(\ell ) =\frac{1}{2\pi } \int _{{\mathcal {G}}} \text {Length}(\ell \cap D) d \sigma (\ell ) = \text {Area}(D) = \pi . \end{aligned}$$

\(\square \)

When we refer to the Hausdorff metric below, we always mean the Euclidean Hausdorff metric (the Hausdorff metric is defined, e.g., in [6]). Write \((x,y) \subseteq {\mathbb {R}}^2\) for the Euclidean interval between \(x,y \in {\mathbb {R}}^2\) excluding the endpoints, and \([x,y] = (x,y) \cup \{ x,y \}\). We similarly write \([x,y]_m\) and \((x,y)_m\) for the \(d_m\)-geodesic between x and y, with and without the endpoints. We claim that for any \(x,y \in {\mathbb {R}}^2\),

$$\begin{aligned}{}[x,y]_m \xrightarrow {m \rightarrow \infty } [x,y] \end{aligned}$$
(78)

in the Hausdorff metric. Indeed, for any \(0 \le \lambda \le 1\), the point on \([x,y]_m\) whose \(d_m\)-distance from x equals \(\lambda \cdot d_m(x,y)\) must converge to the point on [xy] whose Euclidean distance from x equals \(\lambda \cdot |x-y|\). It follows from our assumptions that the convergence is uniform in \(\lambda \), and (78) follows. Moreover, it follows that the Hausdorff convergence in (78) is locally uniform in \(x,y \in {\mathbb {R}}^2\).

Write \({\overline{A}}\) for the closure of a set A. The Euclidean \(\varepsilon \)-neighborhood of a subset \(A \subseteq {\mathbb {R}}^2\) is the collection of all \(x \in {\mathbb {R}}^2\) with \(d(x,A) < \varepsilon \) where \(d(x,A) = \inf _{y \in A} |x-y|\). Given a convex polygon \(P \subseteq {\mathbb {R}}^2\), for a sufficiently large m we define \(P^{(m)} \subseteq {\mathbb {R}}^2\) to be the \(d_m\)-polygon with the same vertices as P. We need m to be sufficiently large in order to guarantee that no vertex of P is in the \(d_m\)-convex hull of the other vertices.

Lemma 7.3

Let \(P_0, P_1 \subseteq D(0,3)\) be convex polygons such that \(\overline{P_0} \subseteq P_1\). Then there exist \(m_0 \ge 1\) and \(\varepsilon > 0\) such that the following holds: For any \(m \ge m_0\) and any \(x, x',y,y' \in D(0,3)\) with \(|x -x'| < \varepsilon \) and \(|y - y'| < \varepsilon \),

$$\begin{aligned} (x,y) \cap P_0 \ne \emptyset \qquad \Longrightarrow \qquad (x',y')_m \cap P_1^{(m)} \ne \emptyset , \end{aligned}$$

and

$$\begin{aligned} (x',y')_m \cap P_0^{(m)} \ne \emptyset \qquad \Longrightarrow \qquad (x,y) \cap P_1 \ne \emptyset . \end{aligned}$$

Proof

From the Hausdorff convergence in (78) it follows that for a sufficiently large m, the closure of \(P_0^{(m)} \cup P_0\) is contained in \(P_1 \cap P_1^{(m)}\). In fact, there exist \(\delta > 0\) and \(m_1 \ge 1\) such that for \(m \ge m_1\), the Euclidean \(\delta \)-neighborhood of \(P_0^{(m)} \cup P_0\) is contained in \(P_1 \cap P_1^{(m)}\).

Set \(\varepsilon = \delta / 2\). Since the convergence in (78) is uniform in \(x,y \in D(0,3)\), there exists \(m_0 \ge m_1\) such that for any \(m \ge m_0\) and \(x',y'\in D(0,3)\), the Hausdorff distance between \((x',y')\) and \((x',y')_{m}\) is at most \(\varepsilon \). Thus for any \(m \ge m_0\) and \(x, x',y,y' \in D(0,3)\) with \(|x -x'| < \varepsilon \) and \(|y - y'| < \varepsilon \), the Hausdorff distance between (xy) and \((x',y')\) is at most \(\varepsilon \), and by the triangle inequality, the Hausdorff distance between (xy) and \((x',y')_{m}\) is at most \(2\varepsilon = \delta \).

Hence if (xy) intersects \(P_0\), then \((x',y')_m\) intersects the Euclidean \(\delta \)-neighborhood of \(P_0\), which is contained in \(P_1^{(m)}\). Similarly, if \((x',y')_m\) intersects \(P_0^{(m)}\), then (xy) intersects the \(\delta \)-neighborhood of \(P_0^{(m)}\) which is contained in \(P_1\). \(\square \)

Lemma 7.4

Let \(K \subseteq {\mathbb {R}}^2\) be a bounded, open, convex set. Then there exist \(d_m\)-polygons \(K_m^{\pm } \subseteq {\mathbb {R}}^2\) for \(m \ge 1\), real numbers \(\varepsilon _m \searrow 0\) and \(m_0 \ge 1\), such that for any \(m \ge m_0\) the following hold:

$$\begin{aligned} K_m^- \subseteq K \subseteq K_m^+, \end{aligned}$$

both boundaries \(\partial K_m^{\pm }\) are \(\varepsilon _m\)-close to \(\partial K\) in the Hausdorff metric, and the \(d_m\)-perimeters of \(K_m^{\pm }\) differ from \(\text {Perimeter}(K)\) by at most \(\varepsilon _m\).

Proof

It suffices to show that for any fixed \(\varepsilon > 0\) there exist \(m_0 \ge 1\) and \(d_m\)-polygons \(K_m^\pm \subset {\mathbb {R}}^2\), defined for any \(m \ge m_0\), such that

$$\begin{aligned} K_m^- \subseteq K \subseteq K_m^+, \end{aligned}$$

both boundaries \(\partial K_m^{\pm }\) are \(\varepsilon \)-close to \(\partial K\) in the Hausdorff metric, and the \(d_m\)-perimeters of \(K_m^{\pm }\) differ from \(\text {Perimeter}(K)\) by at most \(\varepsilon \).

Fix \(\varepsilon > 0\). We may pick finitely many points in \(\partial K\), cyclically ordered, such that when connecting each point via a segment to its two adjacent points, the result is a convex polygon whose boundary is \((\varepsilon /2)\)-close to \(\partial K\) in the Hausdorff metric. We may also require that the perimeter of this convex polygon differs from \(\text {Perimeter}(K)\) by at most \(\varepsilon /2\).

We slightly move these finitely many points inside K, and replace the segments between the points by \(d_m\)-geodesics. For a sufficiently large m, this defines a \(d_m\)-polygon \(K_m^-\). It follows from (78) that for a sufficiently large m, the boundary \(\partial K_m^-\) is \(\varepsilon \)-close to \(\partial K\) in the Hausdorff metric, the \(d_m\)-perimeter of \(K_m^-\) differs from \(\text {Perimeter}(K)\) by at most \(\varepsilon \), and \(K_m^- \subseteq K\).

We still need to construct \(K_m^+\). Approximate K by a convex polygon containing the closure of K in its interior, whose boundary is \((\varepsilon /2)\)-close to \(\partial K\) in the Hausdorff metric, and whose perimeter differs from \(\text {Perimeter}(K)\) by at most \(\varepsilon /2\). Replace the edges of this polygon by \(d_m\)-geodesics in order to form \(K_m^+\). It follows from (78) that for a sufficiently large m, the \(d_m\)-convex set \(K_m^+\) has the desired properties. \(\square \)

We apply Lemma 7.4 for the unit disc \(D \subseteq {\mathbb {R}}^2\), and obtain two \(d_m\)-polygons \(D_m^{\pm }\) with \(D_m^- \subseteq D \subseteq D_m^+\) that satisfy the conclusions of the lemma. It follows from (78) that for \(x,y \in {\mathbb {R}}^2\),

$$\begin{aligned} \text {Length}_m( (x,y)_m \cap D_m^{\pm } ) \xrightarrow {m \rightarrow \infty } \text {Length}( (x,y) \cap D ), \end{aligned}$$
(79)

and the convergence is locally uniform in \(x, y \in {\mathbb {R}}^2\). Let us fix a convex polygon \(K \subseteq {\mathbb {R}}^2\) such that \(\overline{D(0,2)} \subseteq K\) and \({\overline{K}} \subseteq D(0,3)\). We apply Lemma 7.4 and obtain \(d_m\)-polygons \(K_m = K_m^+\) for \(m \ge 1\) that approximate K. For a set \(A \subseteq {\mathbb {R}}^2\) denote

$$\begin{aligned} {\mathcal {G}}(A) = \{ \ell \in {\mathcal {G}}\, ; \, \ell \cap A \ne \emptyset \} \qquad \text {and} \qquad {\mathcal {G}}_{m}(A) = \{ \gamma \in {\mathcal {G}}_{m} \, ; \, \gamma \cap A \ne \emptyset \}. \end{aligned}$$

Definition 7.5

Define the map \(T_K :{\mathcal {G}}(K) \rightarrow \partial K \times \partial K \subset {\mathbb {R}}^2 \times {\mathbb {R}}^2\) by

$$\begin{aligned} T_K(\ell ) = (a(\ell ), b(\ell )) \in \partial K \times \partial K, \end{aligned}$$

where \(\ell \cap \partial K = \{ a(\ell ), b(\ell ) \}\) and the line \(\ell \) is oriented from the point \(a(\ell )\) towards the point \(b(\ell )\). We analogously define the map \(T_{m}: {\mathcal {G}}_m(K_m) \rightarrow \partial K_m \times \partial K_m \subset {\mathbb {R}}^2 \times {\mathbb {R}}^2\) via

$$\begin{aligned} T_m(\gamma ) = (a(\gamma ), b(\gamma )) \in \partial K_m \times \partial K_m, \end{aligned}$$

where \(\gamma \cap \partial K_m = \{ a(\gamma ), b(\gamma ) \}\) and the geodesic \(\gamma \) is oriented from \(a(\gamma )\) towards \(b(\gamma )\).

Denote by \(\mu \) the push-forward of \(\sigma |_{{\mathcal {G}}(K)}\) under the map \(T_{K}\), and denote by \(\mu _m^{}\) the push-forward of \(\sigma _m|_{{\mathcal {G}}(K_m)}\) under the map \(T_{m}\). By Lemma 7.4 and the Crofton formula,

$$\begin{aligned} \frac{1}{2} \cdot \mu _m^{}({\mathbb {R}}^2 \times {\mathbb {R}}^2)= & {} \text {Perimeter}_m(K_m) \xrightarrow {m \rightarrow \infty } \text {Perimeter}(K) \nonumber \\= & {} \frac{1}{2} \cdot \mu ({\mathbb {R}}^2 \times {\mathbb {R}}^2). \end{aligned}$$
(80)

For a convex polygon \(P \subseteq {\mathbb {R}}^2\) we write \({\mathcal {F}}(P) \subseteq \partial K \times \partial K\) for the collection of all pairs of points \(x \ne y \in \partial K\) with \((x,y) \cap P \ne \emptyset \). For a \(d_m\)-polygon P we denote by \({\mathcal {F}}_m(P) \subseteq \partial K_m \times \partial K_m\) the collection of all pairs of points \(x \ne y \in \partial K_{m}\) with \((x,y)_m \cap P \ne \emptyset \). Note that if \({\overline{P}} \subseteq K_m\) then by the Crofton formula,

$$\begin{aligned} \frac{1}{2} \cdot \mu _m^{}({\mathcal {F}}_m(P)) = \text {Perimeter}_m(P). \end{aligned}$$
(81)

For a subset \(A \subseteq {\mathbb {R}}^2 \times {\mathbb {R}}^2\) and \(\varepsilon > 0\) we write \({\mathcal {N}}_{\varepsilon }(A) \subseteq {\mathbb {R}}^2 \times {\mathbb {R}}^2\) for the Euclidean \(\varepsilon \)-neighborhood, i.e., the collection of all \((x,y) \in {\mathbb {R}}^2 \times {\mathbb {R}}^2\) for which there exists \((z,w) \in A\) with \(|x - z| < \varepsilon \) and \(|y - w| < \varepsilon \).

Lemma 7.6

Fix two convex polygons \(P_0, P_1 \subseteq {\mathbb {R}}^2\) with \( \overline{P_0} \subseteq P_1\) and \(\overline{P_1} \subseteq K\). For \(i=0,1\) abbreviate \({\mathcal {F}}_i = {\mathcal {F}}(P_i)\). Then there exists \(\varepsilon _0 > 0\) such that

$$\begin{aligned} \limsup _{m \rightarrow \infty } \mu _m^{}({\mathcal {N}}_{\varepsilon _0}({\mathcal {F}}_0)) \le \mu ({\mathcal {F}}_1) = 2 \cdot \mathrm{perimeter}(P_1). \end{aligned}$$
(82)

Furthermore, for any \(0< \varepsilon < \varepsilon _0\),

$$\begin{aligned} \liminf _{m \rightarrow \infty } \mu _m^{}({\mathcal {N}}_{\varepsilon }({\mathcal {F}}_1)) \ge \mu ({\mathcal {F}}_0) = 2 \cdot \mathrm{perimeter}(P_0). \end{aligned}$$
(83)

Proof

Recall that \(P^{(m)} \subseteq {\mathbb {R}}^2\) was defined to be the \(d_m\)-polygon with the same vertices as P, which is well-defined for a sufficiently large m. Write \({\mathcal {F}}_i^{(m)} = {\mathcal {F}}_m(P_i^{(m)})\). According to Lemma 7.3 there exist \(\varepsilon _0 > 0\) and \(m_0 \ge 1\) such that for any \(m \ge m_0\),

$$\begin{aligned}&{\mathcal {N}}_{\varepsilon _0}({\mathcal {F}}_0) \cap (\partial K_m \times \partial K_m) \subseteq {\mathcal {F}}_1^{(m)} \qquad \text{ and } \qquad \nonumber \\&{\mathcal {N}}_{\varepsilon _0}({\mathcal {F}}_{0}^{(m)}) \cap (\partial K \times \partial K) \subseteq {\mathcal {F}}_1. \end{aligned}$$
(84)

By increasing \(m_0\) if necessary, we may assume that \(\overline{P_{i}^{(m)}} \subseteq K \subseteq K_{m}\) for all \(m \ge m_0\) and \(i=0,1\). Using (84) and (81), for \(m \ge m_0\),

$$\begin{aligned} \mu _m^{}( {\mathcal {N}}_{\varepsilon _0}({\mathcal {F}}_0) ) \le \mu _m^{}( {\mathcal {F}}_1^{(m)} ) = 2 \cdot \text {Perimeter}_m(P_1^{(m)}) \xrightarrow {m \rightarrow \infty } 2 \cdot \text {Perimeter}(P_1). \end{aligned}$$

Fix \(0< \varepsilon < \varepsilon _0\). By Lemma 7.4, there exists \(m_1 \ge m_0\) such that for any \(m \ge m_1\), the Hausdorff distance between \(\partial K_m\) and \(\partial K\) is at most \(\varepsilon \). From (84) we obtain that for \(m \ge m_1\),

$$\begin{aligned} {\mathcal {F}}_0^{(m)} \subseteq {\mathcal {N}}_{\varepsilon }({\mathcal {F}}_1). \end{aligned}$$

Hence,

$$\begin{aligned} \mu _m^{}( {\mathcal {N}}_{\varepsilon }({\mathcal {F}}_1) ) \ge \mu _m^{}( {\mathcal {F}}_0^{(m)} ) = 2 \cdot \text {Perimeter}_m(P_0^{(m)}) \xrightarrow {m \rightarrow \infty } 2 \cdot \text {Perimeter}(P_0). \end{aligned}$$

This completes the proof of (82) and (83). \(\square \)

Proof of Proposition 7.1

By passing to a subsequence, we may assume that \(\text {Area}_m(D)\) converges to an element of \({\mathbb {R}}\cup \{ + \infty \}\) as \(m \rightarrow \infty \), and our goal is to prove that this limit equals \(\pi = \text {Area}(D)\).

The total mass of the measures \(\mu _m^{}\) is uniformly bounded, by (80). Lemma 7.4 implies that the support of \(\mu _m^{}\), which is contained in \(\partial K_m \times \partial K_m\), is uniformly bounded in \({\mathbb {R}}^2\). We may thus pass to a subsequence and assume that

$$\begin{aligned} \mu _m^{} \xrightarrow {m \rightarrow \infty } {\tilde{\mu }} \end{aligned}$$
(85)

weakly for some measure \({\tilde{\mu }}\). This means that for any continuous test function \(\varphi \) on \({\mathbb {R}}^2 \times {\mathbb {R}}^2\) we have \(\int \varphi d \mu _m^{} \longrightarrow \int \varphi d {\tilde{\mu }}\). The measure \({\tilde{\mu }}\) is supported on \(\partial K \times \partial K\), by Lemma 7.4.

Recall that \(\overline{D(0,2)} \subseteq K\). We claim that for any convex polygon \(P \subseteq {\mathbb {R}}^2\) with \(P \subseteq D(0,2)\),

$$\begin{aligned} {\tilde{\mu }}({\mathcal {F}}(P)) = \mu ({\mathcal {F}}(P)). \end{aligned}$$
(86)

Since \(\mu ({\mathcal {F}}(P)) = 2 \cdot \text {Perimeter}(P)\) is continuous in P and monotone in P under inclusion, and since \({\mathcal {F}}(P') \subseteq {\mathcal {F}}(P)\) when \(P' \subseteq P\), in order to prove (86) it suffices to prove the following: For any two convex polygons \(P_0, P_1 \subseteq {\mathbb {R}}^2\) with \( \overline{P_0} \subseteq P_1\) and \(\overline{P_1} \subseteq K\),

$$\begin{aligned} {\tilde{\mu }}({\mathcal {F}}(P_0)) \le \mu ({\mathcal {F}}(P_1)) \qquad \text {and} \qquad \mu ({\mathcal {F}}(P_0)) \le {\tilde{\mu }}(\overline{{\mathcal {F}}(P_1)}). \end{aligned}$$
(87)

From Lemmas 7.6 and (85), there exists \(\varepsilon _0 > 0\) with

$$\begin{aligned} {\tilde{\mu }}({\mathcal {F}}(P_0)) \le \limsup _{m \rightarrow \infty } \mu _m({\mathcal {N}}_{\varepsilon _0}({\mathcal {F}}(P_0))) \le \mu ({\mathcal {F}}(P_1)) \end{aligned}$$

and for any \(0< \varepsilon < \varepsilon _0\),

$$\begin{aligned} {\tilde{\mu }}({\mathcal {N}}_{2 \varepsilon }({\mathcal {F}}(P_1))) \ge \liminf _{m \rightarrow \infty } \mu _m( {\mathcal {N}}_{\varepsilon }({\mathcal {F}}(P_1)) ) \ge \mu ({\mathcal {F}}(P_0)). \end{aligned}$$

By letting \(\varepsilon \) tend to zero, we obtain (87), and hence (86) is proven. The map \(T_K^{-1}\) is a well-defined map from \(A = \{ (x,y) \in \partial K \times \partial K \, ; \, x \ne y \}\) to \({\mathcal {G}}\). By (86), the push-forward of \({\tilde{\mu }}|_A\) under the map \(T_K^{-1}\) is a measure \({\tilde{\sigma }}\) on \({\mathcal {G}}\) which satisfies the assumptions of Lemma 7.2. From the conclusion of Lemma 7.2,

$$\begin{aligned}&\frac{1}{2 \pi } \int _{\partial K \times \partial K} \text {Length}((x,y) \cap D ) d {\tilde{\mu }}(x,y) \nonumber \\&\quad = \frac{1}{2 \pi } \int _{A} \text {Length}((x,y) \cap D ) d {\tilde{\mu }}(x,y) = \pi . \end{aligned}$$
(88)

By the Santaló formula,

$$\begin{aligned} \text {Area}_m(D_m^{\pm }) = \frac{1}{2\pi } \int _{{\mathbb {R}}^2 \times {\mathbb {R}}^2} \text {Length}_m( (x,y)_m \cap D_m^{\pm } ) d \mu _m^{}(x,y). \end{aligned}$$

We thus deduce from (79), (85) and (88) that

$$\begin{aligned} \lim _{m \rightarrow \infty } \text {Area}_m(D_m^{\pm }) = \frac{1}{2\pi } \int _{{\mathbb {R}}^2 \times {\mathbb {R}}^2} \text {Length}( (x,y) \cap D) d {\tilde{\mu }}(x,y) = \pi . \end{aligned}$$
(89)

However, \(D_m^- \subseteq D \subseteq D_m^+\). Hence (89) implies that \(\text {Area}_m(D) \longrightarrow \pi \). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Eilat, M., Klartag, B. Rigidity of Riemannian embeddings of discrete metric spaces. Invent. math. 226, 349–391 (2021). https://doi.org/10.1007/s00222-021-01048-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00222-021-01048-y

Navigation