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

Lower Bounding the Gromov–Hausdorff distance in Metric Graphs

Henry Adams [email protected] Sushovan Majhi [email protected] Fedor Manin [email protected] Žiga Virk [email protected]  and  Nicolò Zava [email protected]
Abstract.

Let GG be a compact, connected metric graph and let XGX\subseteq G be a subset. If XX is sufficiently dense in GG, we show that the Gromov–Hausdorff distance matches the Hausdorff distance, namely dGH(G,X)=dH(G,X)d_{\mathrm{GH}}(G,X)=d_{\mathrm{H}}(G,X). In a recent study, when the metric graph is the circle G=S1G=S^{1} with circumference 2π2\pi, the equality dGH(S1,X)=dH(S1,X)d_{\mathrm{GH}}(S^{1},X)=d_{\mathrm{H}}(S^{1},X) was established whenever dGH(S1,X)<π6d_{\mathrm{GH}}(S^{1},X)<\frac{\pi}{6}. Our result for general metric graphs relaxes this hypothesis in the circle case to dGH(S1,X)<π3d_{\mathrm{GH}}(S^{1},X)<\frac{\pi}{3}, and furthermore, we give an example showing that the constant π3\frac{\pi}{3} is the best possible. We lower bound the Gromov–Hausdorff distance dGH(G,X)d_{\mathrm{GH}}(G,X) by the Hausdorff distance dH(G,X)d_{\mathrm{H}}(G,X) via a simple topological obstruction: showing a correspondence with too small distortion contradicts the connectedness of GG. Moreover, for two sufficiently dense subsets X,YGX,Y\subseteq G, we provide new lower bounds on dGH(X,Y)d_{\mathrm{GH}}(X,Y) in terms of the Hausdorff distance dH(X,Y)d_{\mathrm{H}}(X,Y), which is 𝒪(n2)\mathcal{O}(n^{2})-time computable if the subsets have at most nn points.

Key words and phrases:
Gromov–Hausdorff distance, Hausdorff distance, Metric graph, Convexity radius

1. Introduction

The Gromov–Hausdorff distance between two abstract metric spaces (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}), denoted dGH(X,Y)d_{\mathrm{GH}}(X,Y), provides a (dis)-similarity measure quantifying how far the two metric spaces are from being isometric [9, 8, 18]. In the past two decades, this distance has found applications in topological data analysis (TDA) as a theoretical framework for shape and dataset comparison [15], which motivated the study of its quantitative aspects. However, precise computations of the Gromov–Hausdorff distance between even simple spaces are mostly unknown (with certain exceptions such as the Gromov–Hausdorff distance between a line segment and a circle [11], between spheres of certain dimensions [13, 1], and between finite subsets of \mathbb{R} [14]). Even approximating the Gromov–Hausdorff distance by a factor of 33 in the case of trees with unit-edge length is known to be NP-hard ([3, 17]). A better approximation factor can only be achieved in very special cases; for example, [14] provides a 54\frac{5}{4}-approximation scheme if XX and YY are finite subsets of the real line. Developing tools and pipelines to estimate the Gromov–Hausdorff distance is a central research direction in the computational topology community.

To estimate the Gromov–Hausdorff distance between two metric spaces XX and YY, one may:
(A) Find sufficiently nice samples XXX^{\prime}\subseteq X and YYY^{\prime}\subseteq Y approximating the geometry of XX and YY,
(B) Efficiently bound the Gromov–Hausdorff distance between the subsets XX^{\prime} and YY^{\prime}.
Indeed, when XX and YY are infinite continuous objects—e.g., Riemannian manifolds, metric graphs—and one wants to use computational machinery, it is often essential to resort to finite subsets approximating the geometry of the original spaces. In this paper, we investigate both research directions (A) and (B).

We reformulate the first task (A) as follows: how dense does the sample XX^{\prime} have to be in XX to capture the geometry of the ambient space? To systematize this question, we use the Hausdorff distance as a measure of density and ask if we can provide a threshold ε\varepsilon and constant 0<C10<C\leq 1 such that dGH(X,X)CdH(X,X)d_{\mathrm{GH}}(X,X^{\prime})\geq C\cdot d_{\mathrm{H}}(X,X^{\prime}) whenever dH(X,X)<εd_{\mathrm{H}}(X,X^{\prime})<\varepsilon. In particular dGH(X,X)=dH(X,X)d_{\mathrm{GH}}(X,X^{\prime})=d_{\mathrm{H}}(X,X^{\prime}) if C=1C=1. In [2], the authors showed that no such threshold ε\varepsilon or constant CC exist in general. More precisely, they demonstrated examples of metric spaces XX and subsets XX^{\prime} thereof whose Gromov–Hausdorff distance is arbitrarily small compared to their Hausdorff distance. The authors also provided conditions for positive results. Their formulation of the problem slightly differs from that described above, but leads to stronger statements. Indeed, they provided lower bounds of the form dGH(X,X)CdH(X,X)d_{\mathrm{GH}}(X^{\prime},X)\geq C\cdot d_{\mathrm{H}}(X,X^{\prime}) when XX is a manifold, typically with 12C1\frac{1}{2}\leq C\leq 1, under the assumption that dGH(X,X)<εd_{\mathrm{GH}}(X,X^{\prime})<\varepsilon for some constant ε\varepsilon. The latter inequality implies that dH(X,X)<εd_{\mathrm{H}}(X,X^{\prime})<\varepsilon. Our study continues that investigation and proves improved results (such as with C=1C=1) when XX is a metric graph, hence relaxing the manifold assumption.

Different techniques have been proposed to tackle the second direction (B) and provide bounds on the Gromov–Hausdorff distance. Among them, a particularly fruitful approach to achieve lower bounds relies on stable metric invariants. One associates each metric space XX with a value ψ(X)\psi(X) in another computationally simpler metric space so that ψ(X)\psi(X) and ψ(Y)\psi(Y) are close whenever XX and YY are close in the Gromov–Hausdorff distance. An immediate example is the diameter; see Example 2.2. However, the diameter bound is often not informative when the metric spaces are of similar diameter but metrically very different. For further stable metric invariants, we refer the reader to [16]. Other central examples in computational topology are Vietoris–Rips persistence diagrams ([7]) and hierarchical clustering ([6]). In [19], techniques from dimension theory were applied to provide unavoidable limits to the precision of metric invariants with values in Hilbert spaces.

Our paper also provides interesting lower bounds on dGH(X,Y)d_{\mathrm{GH}}(X,Y) revealing the metric disparity between XX and YY. For example, a positive lower bound on the Gromov–Hausdorff distance is indicative of the existence of no isometry between the two metric spaces. We answer two open questions [2, Questions 2, 4] by providing optimal lower bounds on the Gromov–Hausdorff distance between the circle and a subset thereof, and by providing novel techniques lower bounding the Gromov–Hausdorff distance between a metric graph and a subset thereof. Though we generalize beyond manifolds, our generalization produces a result for the circle (the simplest 11-dimensional manifold) that is not only stronger than the result from [2], but also provably optimal.

One of the precursors of algebraic topology was Brouwer’s proof of invariance of dimension, namely that m\mathbb{R}^{m} is not homeomorphic to n\mathbb{R}^{n} for n>mn>m. Proving 1\mathbb{R}^{1} is not homeomorphic to n\mathbb{R}^{n} for n>1n>1 relies only on connectedness: removing a single point from 1\mathbb{R}^{1} leaves a disconnected space, which is not the case for n\mathbb{R}^{n} with n>1n>1. However, to prove invariance of dimension for n>m2n>m\geq 2 Brouwer relied on concepts related to the degree of a map between spheres, which is often now formalized using higher-dimensional homotopy groups or homology groups. The topological obstructions to small Gromov–Hausdorff distances in [2] relied on homological tools, namely the fundamental class of a manifold. One of the key contributions of our paper is to go “back in time” and produce (sometimes stronger) lower bounds on Gromov–Hausdorff distances using simpler topological tools: we show that small Gromov–Hausdorff distances would contradict the connectedness of a space.

1.1. Overview of Our Results

We obtain results for both research directions (A) and (B) in the case of metric graphs. Namely, we provide conditions ensuring that the Hausdorff and Gromov–Hausdorff distances between the ambient space and a sample coincide, and we derive bounds for the Gromov–Hausdorff distance between two sufficiently dense samples. Given the independent relevance of both kinds of contributions, we present many of our results with an item (a) and (b) dedicated to each.

Let GG be a connected, compact metric graph with finitely many edges, and let XGX\subseteq G be a subset. We show how to lower bound the Gromov–Hausdorff distance dGH(G,X)d_{\mathrm{GH}}(G,X) in terms of constant multiplier times the Hausdorff distance dH(G,X)d_{\mathrm{H}}(G,X), namely dGH(G,X)CdH(G,X)d_{\mathrm{GH}}(G,X)\geq C\cdot d_{\mathrm{H}}(G,X). We often obtain the optimal value C=1C=1. Since the reverse inequality dGH(G,X)dH(G,X)d_{\mathrm{GH}}(G,X)\leq d_{\mathrm{H}}(G,X) always holds, the inequality dGH(G,X)dH(G,X)d_{\mathrm{GH}}(G,X)\geq d_{\mathrm{H}}(G,X) (with optimal constant C=1C=1) immediately gives equality dGH(G,X)=dH(G,X)d_{\mathrm{GH}}(G,X)=d_{\mathrm{H}}(G,X).

The simplest metric graph is a tree. Let dH(T,X)\vec{d}_{\mathrm{H}}(\partial T,X) denote the directed Hausdorff distance from the leaves of TT to the sample XX (see Section 2). We show the following. {restatable*}[Metric Trees]thmtrees Let (T,d)(T,d) be a compact metric tree. For any X,YTX,Y\subseteq T,

  1. (a)

    if dH(T,X)>dH(T,X)d_{\mathrm{H}}(T,X)>\vec{d}_{\mathrm{H}}(\partial T,X), then dGH(T,X)=dH(T,X)d_{\mathrm{GH}}(T,X)=d_{\mathrm{H}}(T,X), and

  2. (b)

    for any εdH(T,Y)\varepsilon\geq d_{\mathrm{H}}(T,Y), if dH(X,Y)>dH(T,X)+εd_{\mathrm{H}}(X,Y)>\vec{d}_{\mathrm{H}}(\partial T,X)+\varepsilon, then dGH(X,Y)dH(X,Y)2εd_{\mathrm{GH}}(X,Y)\geq d_{\mathrm{H}}(X,Y)-2\varepsilon.

The interpretation of part (a) of Theorem 1.1 is as follows: if the biggest gaps in the sampling XTX\subseteq T are not near the leaves of TT, then dGH(T,X)=dH(T,X)d_{\mathrm{GH}}(T,X)=d_{\mathrm{H}}(T,X).

The following theorem shows that some control on the set of leaf vertices of TT is needed.

{restatable*}

thmcounterEx For any ε>0\varepsilon>0 there exists a compact metric tree TT and a closed subset XTX\subseteq T such that dGH(T,X)εdH(T,X)d_{\mathrm{GH}}(T,X)\leq\varepsilon\cdot d_{\mathrm{H}}(T,X).

When G=S1G=S^{1} is the circle of circumference 2π2\pi, the authors of [2] showed that if dGH(S1,X)<π6d_{\mathrm{GH}}(S^{1},X)<\frac{\pi}{6}, then dGH(S1,X)=dH(S1,X)d_{\mathrm{GH}}(S^{1},X)=d_{\mathrm{H}}(S^{1},X). The question of the optimality of the density constant π6\frac{\pi}{6} was subsequently raised. In this paper, we show that π6\frac{\pi}{6} is suboptimal: the optimal density constant is in fact π3\frac{\pi}{3}. The first result below shows that the new weaker assumption dGH(S1,X)<π3d_{\mathrm{GH}}(S^{1},X)<\frac{\pi}{3} is necessary, and the second result shows it is sufficient.

{restatable*}

[Optimality of π3\frac{\pi}{3}]thmoptimality For any ε(0,π6)\varepsilon\in\left(0,\frac{\pi}{6}\right), there exists a nonempty compact subset XS1X\subseteq S^{1} with dH(S1,X)=π3+εd_{\mathrm{H}}(S^{1},X)=\frac{\pi}{3}+\varepsilon and dGH(S1,X)=π3<dH(S1,X)d_{\mathrm{GH}}(S^{1},X)=\frac{\pi}{3}<d_{\mathrm{H}}(S^{1},X).

{restatable*}

[The Circle]thmsufficient For any subsets X,YS1X,Y\subseteq S^{1}, we have

  1. (a)

    dGH(S1,X)min{dH(S1,X),π3}d_{\mathrm{GH}}(S^{1},X)\geq\min\left\{d_{\mathrm{H}}(S^{1},X),\ \tfrac{\pi}{3}\right\}, and

  2. (b)

    dGH(X,Y)min{dH(X,Y)2ε,π3ε}d_{\mathrm{GH}}(X,Y)\geq\min\left\{d_{\mathrm{H}}(X,Y)-2\varepsilon,\ \tfrac{\pi}{3}-\varepsilon\right\} for any εdH(S1,Y)\varepsilon\geq d_{\mathrm{H}}(S^{1},Y).

We generalize the above results for the circle and trees to general metric graphs. Let GG be a compact metric graph and let e(G)e(G) be the shortest edge length between non-leaf vertices (see (2)). We prove that if XX is dense enough relative to the smallest edge length e(G)e(G), then the Gromov–Hausdorff distance matches the Hasudorff distance.

{restatable*}

[Metric Graphs]thmgraphs Let (G,d)(G,d) be a compact, connected metric graph. Let X,YGX,Y\subseteq G be subsets such that dH(G,X)>dH(G,X)d_{\mathrm{H}}(G,X)>\vec{d}_{\mathrm{H}}(\partial G,X) if G\partial G\neq\emptyset. Then

  1. (a)

    dGH(G,X)min{dH(G,X),112e(G)}d_{\mathrm{GH}}(G,X)\geq\min\left\{d_{\mathrm{H}}(G,X),\tfrac{1}{12}e(G)\right\}, and

  2. (b)

    dGH(X,Y)min{dH(X,Y)2ε,112e(G)ε}d_{\mathrm{GH}}(X,Y)\geq\min\left\{d_{\mathrm{H}}(X,Y)-2\varepsilon,\tfrac{1}{12}e(G)-\varepsilon\right\} for any εdH(G,Y)\varepsilon\geq d_{\mathrm{H}}(G,Y).

1.2. Organization of the Paper

We begin in Section 2 with background and notation. In Section 3, we provide an example showing that some control over the boundary of the metric graph is necessary. We begin with the case of metric trees in Section 4, proceeding next to the case of the circle in Section 5, and finally general metric graphs in Section 6.

2. Background and notation

We refer the reader to [5, 4] for more information on metric spaces, metric graphs, the Hausdorff distance, and the Gromov–Hausdorff distance.

Metric spaces

Let (X,d)(X,d) be a metric space. For any xXx\in X, we let B(x;r)={xX|d(x,x)<r}B(x;r)=\{x^{\prime}\in X~{}|~{}d(x,x^{\prime})<r\} denote the metric ball of radius rr about xx. If needed for the sake of clarity, we will sometimes write BX(x;r)B_{X}(x;r) instead of B(x;r)B(x;r). For any r0r\geq 0 and AXA\subseteq X, we let

Ar=xAB(x;r)={xGinfaAd(a,x)<r}A^{r}=\cup_{x\in A}B(x;r)=\left\{x\in G\mid\inf_{a\in A}d(a,x)<r\right\} (1)

denote the rr-thickening of AA in XX, namely, the union of open metric balls of radius rr centered at the points of AA.

The diameter of a subset AXA\subseteq X is the supremum of all distances between pairs of points of AA. More formally, the diameter is defined as

diam(A)=sup{d(x,x)x,xA}.\mathrm{diam}(A)=\sup\{d(x,x^{\prime})\mid x,x^{\prime}\in A\}.

The diameter of a compact set AA is always finite.

Metric graphs

We follow Definitions 3.1.12 (quotient metric) and 3.2.9 (metric graph) of [5] in order to define a metric graph. A metric segment is a metric space isometric to a real line segment [0,l][0,l] for l0l\geq 0. Our metric graphs will always be connected and have a finite number of vertices and edges.

Definition 2.1 (Metric Graph).

A metric graph (G,d)(G,d) is the metric space obtained by gluing a finite collection EE of metric segments along their boundaries to a finite collection VV of vertices, and equipping the result (eEe)/\left(\bigsqcup_{e\in E}e\right)/\sim with the quotient metric. We assume this equivalence relation results in a connected space.

The metric segments eEe\in E are called the edges and the equivalence classes of their endpoints in VV are called the vertices of GG. Since GG is connected, we have d(ξ,ξ)<d(\xi,\xi^{\prime})<\infty for all ξ,ξG\xi,\xi^{\prime}\in G. Since EE is finite, the quotient semi-metric [5, Definition 3.1.12] is in fact a metric.111 One needs to be more careful if EE is infinite. Indeed, suppose one has two vertices vv and vv^{\prime}, and one edge ene_{n} of length 1n\frac{1}{n} for each nn\in\mathbb{N} with endpoints vv and vv^{\prime}. Then the quotient semi-metric gives d(v,v)=infn1n=0d(v,v^{\prime})=\inf_{n\in\mathbb{N}}\frac{1}{n}=0, even though vvv\neq v^{\prime}, and hence this semi-metric is not a metric. These subtleties disappear since we assume EE is finite.

For an edge eEe\in E, its length l(e)l(e) is the length of the corresponding metric segment. Note that the length can be different from the distance between the endpoints of ee according to the metric dd. Extending the definition of length for a continuous path γ:[a,b]G\gamma:[a,b]\to G, we define the length l(γ)l(\gamma) as the sum of the (full or partial) lengths of the edges that γ\gamma traverses.

For a vertex vv of a graph, its degree, denoted deg(v)deg(v), is the number of edges incident to vv. A self-loop at vv contributes 22 to the degree of vv. As an example, the circle S1S^{1} can be given a metric graph structure by assigning a single vertex vv at the north pole, of degree 22, having a self-loop.

A metric graph GG is therefore endowed with two layers of information: a metric structure turning it into a length space [5, Definition 2.1.6] and a combinatorial structure (V,E)(V,E) as an abstract graph. Using the above definition, we are also allowing GG to have single-edge cycles and multiple edges between a pair of vertices. In this paper, we only consider path-connected, compact metric graphs (G,d)(G,d) with finitely many edges. The metric graphs we consider are automatically geodesic (i.e., for every x,yGx,y\in G there is an isometric embedding γ:[0,d(x,y)]G\gamma\colon[0,d(x,y)]\to G such that γ(0)=x\gamma(0)=x and γ(d(x,y))=y\gamma(d(x,y))=y) since they are compact (a compact length space is geodesic). This justifies why there is always at least one shortest path between two points.

For a pair of points ξ,ξG\xi,\xi^{\prime}\in G, there always exists a shortest path or geodesic path γ\gamma in GG joining ξ,ξ\xi,\xi^{\prime} whose length satisfies l(γ)=d(ξ,ξ)l(\gamma)=d(\xi,\xi^{\prime}). We define an undirected simple loop γ\mathcal{\gamma} of GG to be a simple path that intersects itself only at the endpoints. We denote by (G)\mathcal{L}(G) the set of all simple loops of GG. Since GG is assumed to have finitely many edges, the set (G)\mathcal{L}(G) is finite. Since a metric tree is defined as a metric graph with no loops, (G)=\mathcal{L}(G)=\emptyset for a metric tree GG.

For a metric graph (G,d)(G,d) with (G)\mathcal{L}(G)\neq\emptyset, let e(G)e(G) denote the length of the smallest non-terminal edge:

e(G)min{l(e)e=(u,v)E(G) such that min{deg(u),deg(v)}>1}.e(G)\coloneqq\min\{l(e)\mid e=(u,v)\in E(G)\text{ such that }\min\{deg(u),deg(v)\}>1\}. (2)

Hausdorff distances

Let (Z,d)(Z,d) be a metric space. If XX and YY are two subsets of ZZ, then the directed Hausdorff distance from XX to YY, denoted dH(X,Y)\vec{d}_{\mathrm{H}}(X,Y), is defined by

dH(X,Y)=supxXinfyYd(x,y)=inf{r0XYr}.\vec{d}_{\mathrm{H}}(X,Y)=\sup_{x\in X}\inf_{y\in Y}d(x,y)=\inf\bigg{\{}r\geq 0\mid X\subseteq Y^{r}\bigg{\}}.

Note that dH(X,Y)\vec{d}_{\mathrm{H}}(X,Y) is not symmetric in general.

To retain symmetry, the Hausdorff distance, denoted dH(X,Y)d_{\mathrm{H}}(X,Y), is defined as

dH(X,Y)=max{dH(X,Y),dH(X,Y)}.d_{\mathrm{H}}(X,Y)=\max\left\{\vec{d}_{\mathrm{H}}(X,Y),\vec{d}_{\mathrm{H}}(X,Y)\right\}.

In other words, the Hausdorff distance finds the infimum over all real numbers rr such that if we thicken YY by rr it contains XX and if we thicken XX by rr it contains YY. If no such rr exists then the Hausdorff distance is infinity.

Gromov–Hausdorff distances

Let X,YX,Y be non-empty sets. A relation X×Y\mathcal{R}\subseteq X\times Y is called a correspondence if the following two conditions hold: for every xXx\in X there exists some yYy\in Y with (x,y)(x,y)\in\mathcal{R}, and for every yYy\in Y there exists some xXx\in X exists with (x,y)(x,y)\in\mathcal{R}.

If \mathcal{R} is a correspondence between (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}), then the distortion of \mathcal{R} is:

dis()=sup(x,y),(x,y)|dX(x,x)dY(y,y)|.\mathrm{dis}(\mathcal{R})=\sup_{(x,y),(x^{\prime},y^{\prime})\in\mathcal{R}}\left|d_{X}(x,x^{\prime})-d_{Y}(y,y^{\prime})\right|.

The Gromov–Hausdorff distance between two (arbitrary) metric spaces XX and YY to be

dGH(X,Y)=12infX×Ydis(),d_{\mathrm{GH}}(X,Y)=\frac{1}{2}\inf_{\mathcal{R}\subseteq X\times Y}\mathrm{dis}(\mathcal{R}), (3)

where the infimum is taken over all correspondences \mathcal{R} between XX and YY [5, 12]. As shown in  [10], if XX and YY are both compact metric spaces, a distortion-minimizing correspondence in the definition of dGH(X,Y)d_{\mathrm{GH}}(X,Y) always exists.

The above definition immediately provides the following simple lower bound on the Gromov–Hausdorff distance between two compact metric spaces XX and YY.

Example 2.2 (Trivial Lower Bound).

Let (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}) be compact metric spaces. Without loss of generality, we assume that diam(X)diam(Y)\mathrm{diam}(X)\geq\mathrm{diam}(Y). Take an arbitrary correspondence X×Y\mathcal{R}\subseteq X\times Y and pairs of points x,xXx,x^{\prime}\in X and y,yYy,y^{\prime}\in Y such that diam(X)=dX(x,x)\mathrm{diam}(X)=d_{X}(x,x^{\prime}) and (x,y),(x,y)(x,y),(x^{\prime},y^{\prime})\in\mathcal{R}. The distortion satisfies

dis()|dX(x,x)dY(y,y)|=dX(x,x)dY(y,y)diam(X)diam(Y).\mathrm{dis}(\mathcal{R})\geq|d_{X}(x,x^{\prime})-d_{Y}(y,y^{\prime})|=d_{X}(x,x^{\prime})-d_{Y}(y,y^{\prime})\geq\mathrm{diam}(X)-\mathrm{diam}(Y).

We conclude

dGH(X,Y)=12infX×Ydis()12|diam(X)diam(Y)|.d_{\mathrm{GH}}(X,Y)=\tfrac{1}{2}\inf_{\mathcal{R}\subseteq X\times Y}\mathrm{dis}(\mathcal{R})\geq\tfrac{1}{2}|\mathrm{diam}(X)-\mathrm{diam}(Y)|.

Let Z1,Z2Z_{1},Z_{2} be metric spaces and Z1×Z2\mathcal{R}\subseteq Z_{1}\times Z_{2} be a correspondence. For a subset AZ1A\subseteq Z_{1}, we define the following subset of Z2Z_{2}:

[A]{zZ2 there is aA with (a,z)}.\mathcal{R}[A]\coloneqq\{z\in Z_{2}\mid\text{ there is }a\in A\text{ with }(a,z)\in\mathcal{R}\}.

Similarly, for a subset AZ2A\subseteq Z_{2}, we define the following subset of Z1Z_{1}:

1[A]{zZ1 there is aA with (z,a)}.\mathcal{R}^{-1}[A]\coloneqq\{z\in Z_{1}\mid\text{ there is }a\in A\text{ with }(z,a)\in\mathcal{R}\}.

Using the notation above, we observe the following fact.

Lemma 2.3 (Lifting of Connected Subsets).

Let XX be a subset of a length metric space (G,d)(G,d) and let G×X\mathcal{R}\subseteq G\times X be a correspondence with distortion dis()<2r\mathrm{dis}(\mathcal{R})<2r. If SGS\subseteq G is a connected subset, then [S]r\mathcal{R}[S]^{r} (see (1)) is connected.

Proof.

Suppose for a contradiction that [S]r=U1U2\mathcal{R}[S]^{r}=U_{1}\cup U_{2}, where U1,U2U_{1},U_{2} are disjoint, non-empty, open subsets of GG. Note that [S]r\mathcal{R}[S]^{r} is a union of balls in GG, which are connected since GG is a length space. Therefore, we must have a partition of SS, say S=S1S2S=S_{1}\cup S_{2}, such that Ui=[Si]rU_{i}=\mathcal{R}[S_{i}]^{r}.

Since the UiU_{i} sets are non-empty, so are the SiS_{i} sets. To see that the SiS_{i} sets are open, take ξSi\xi\in S_{i} and a corresponding point x[ξ]x\in\mathcal{R}[\xi]. For any ξS\xi^{\prime}\in S with d(ξ,ξ)ζ<2rdis()d(\xi,\xi^{\prime})\leq\zeta<2r-\mathrm{dis}(\mathcal{R}) and x[ξ]x^{\prime}\in\mathcal{R}[\xi^{\prime}], we have

d(x,x)d(ξ,ξ)+dis()ζ+dis()<2r.d(x,x^{\prime})\leq d(\xi,\xi^{\prime})+\mathrm{dis}(\mathcal{R})\leq\zeta+\mathrm{dis}(\mathcal{R})<2r.

So, B(x,r)B(x,r)B(x,r)\cap B(x^{\prime},r)\neq\emptyset as GG is a length metric space. As a result, both ξ,ξSi\xi,\xi^{\prime}\in S_{i}. So each SiS_{i} set is open.

Moreover, taking ζ=0\zeta=0 and ξ=ξ\xi=\xi^{\prime} in the above, we see for any two corresponding points x,x[ξ]x,x^{\prime}\in\mathcal{R}[\xi] that B(x,r)B(x,r)B(x,r)\cap B(x^{\prime},r)\neq\emptyset. This implies that S1S2=S_{1}\cap S_{2}=\emptyset.

So, the SiS_{i} sets create a partition of SS. This is a contradiction since SS is connected. Therefore, [S]r\mathcal{R}[S]^{r} is connected. ∎

3. Ratio of dGHd_{\mathrm{GH}} vs dHd_{\mathrm{H}} can be arbitrarily small for metric trees

In this section we prove Theorem 1.1, a variant of Theorem 5 of [2], now adapted to metric graphs. As we show in Theorems 1.1 and 1.1, the Gromov–Hausdorff distance dGH(T,X)d_{\mathrm{GH}}(T,X) between a metric graph TT and a subset XTX\subseteq T can be bounded below by their Hausdorff distance dH(T,X)d_{\mathrm{H}}(T,X) in case XX is dense enough near the leaves of TT, i.e., dH(T,X)>dH(T,X)d_{\mathrm{H}}(T,X)>\vec{d}_{\mathrm{H}}(\partial T,X). When dH(T,X)<dH(T,X)d_{\mathrm{H}}(T,X)<\vec{d}_{\mathrm{H}}(\partial T,X), however, the ratio of dGH(T,X)d_{\mathrm{GH}}(T,X) over dH(T,X)d_{\mathrm{H}}(T,X) can become arbitrarily small as we show below. Furthermore, Theorem 1.1, a sketch of which is provided in Figure 1, suggests that T\partial T may not map close to itself via correspondences with small distortion. We will use the definition of Gromov–Hausdorff distance presented by Gromov in [9], which is equivalent to that provided in (3) (see for example [5]), where dGH(T,X)d_{\mathrm{GH}}(T,X) is the infimum of the Hausdorff distances between F(T)F(T) and G(X)G(X) in ZZ, with the infimum being over all metric spaces ZZ and all isometric embeddings F:TZF\colon T\hookrightarrow Z and G:XZG\colon X\hookrightarrow Z.

TT
XTX^{\prime}\subseteq T
XTX\subseteq T
Figure 1. A sketch of the construction of Theorem 1.1. The metric tree TT on the left, the bold superimposed XTX^{\prime}\subseteq T in the middle (with obviously small dH(T,X)d_{\mathrm{H}}(T,X^{\prime})), and a different isometric embedding of XX^{\prime} into TT on the right indicated by XTX\subseteq T (with obviously large dH(T,X)d_{\mathrm{H}}(T,X)).
\counterEx
Proof.

Choose nn\in\mathbb{N} and let ε=1n\varepsilon=\frac{1}{n}. For t0t\geq 0, let I(t)I(t) be a copy of the interval [0,t][0,t] with a designated basepoint 0.

Define a metric tree TT as the wedge of nn intervals of lengths 1+ε,1+2ε,,1+nε=21+\varepsilon,1+2\varepsilon,\ldots,1+n\varepsilon=2, i.e.,

T=t=1nI(1+tε).T=\bigvee_{t=1}^{n}I(1+t\cdot\varepsilon).

Define a metric tree X0X_{0} as the wedge of intervals of lengths 1,1+ε,1+2ε,,1+(n1)ε=2ε1,1+\varepsilon,1+2\varepsilon,\ldots,1+(n-1)\varepsilon=2-\varepsilon:

X0=t=0n1I(1+tε).X_{0}=\bigvee_{t=0}^{n-1}I(1+t\cdot\varepsilon).

In Figure 1 we have two different isometric embeddings X,XX^{\prime},X of the same tree X0X_{0} into Z=TZ=T. One with a small Hausdorff distance dH(T,X)=εd_{\mathrm{H}}(T,X^{\prime})=\varepsilon as XX^{\prime} is TT with each ray being shortened by ε\varepsilon, and one with a large Hausdorff distance dH(T,X)=1d_{\mathrm{H}}(T,X)=1 as XX is TT with the longest ray I(2)I(2) of length 22 being replaced with a subray I(1)I(1) of length 11. By the definition above,

dGH(T,X)dH(T,X)=εdH(T,X).d_{\mathrm{GH}}(T,X)\leq d_{\mathrm{H}}(T,X^{\prime})=\varepsilon d_{\mathrm{H}}(T,X).

An analogy can be made with the construction in Theorem 5 of [2], where instead of permuting standard basis vectors in Euclidean space, we are now permuting the leaves of a metric tree.

4. Metric trees

We obtain the following optimal result for metric trees.

\trees
Proof.

We make no distinction between a path γ:[a,b]G\gamma:[a,b]\to G and its image {γ(t)t[a,b]}G\{\gamma(t)\mid t\in[a,b]\}\subseteq G.

We first prove Part (a). Assuming that dH(T,X)>dH(T,X)d_{\mathrm{H}}(T,X)>\vec{d}_{\mathrm{H}}(\partial T,X), we show that dH(T,X)dGH(T,X)d_{\mathrm{H}}(T,X)\leq d_{\mathrm{GH}}(T,X), which implies dH(T,X)=dGH(T,X)d_{\mathrm{H}}(T,X)=d_{\mathrm{GH}}(T,X). The assumption implies that there are two points x,xXx,x^{\prime}\in X with d(x,x)2dH(T,X)d(x,x^{\prime})\geq 2d_{\mathrm{H}}(T,X) such that the midpoint of the unique shortest path η\eta joining them is at least dH(T,X)d_{\mathrm{H}}(T,X) distance away from all points of XX. Consequently, for any rdH(T,X)r\leq d_{\mathrm{H}}(T,X) and for any subset AXA\subseteq X containing both xx and xx^{\prime}, the rr-thickening ArA^{r} cannot be connected. In particular, [T]r\mathcal{R}[T]^{r} cannot be connected.

If we had dGH(T,X)<dH(T,X)d_{\mathrm{GH}}(T,X)<d_{\mathrm{H}}(T,X), we could choose dGH(T,X)<rdH(T,X)d_{\mathrm{GH}}(T,X)<r\leq d_{\mathrm{H}}(T,X) and there would exist a correspondence T×X\mathcal{R}\subseteq T\times X with dis()<2r2dH(T,X)\mathrm{dis}(\mathcal{R})<2r\leq 2d_{\mathrm{H}}(T,X). Since TT is connected, Lemma 2.3 would imply that [T]r\mathcal{R}[T]^{r} is connected—a contradiction. Therefore, dGH(T,X)dH(T,X)d_{\mathrm{GH}}(T,X)\geq d_{\mathrm{H}}(T,X), i.e., dGH(T,X)=dH(T,X)d_{\mathrm{GH}}(T,X)=d_{\mathrm{H}}(T,X).

In order to show Part (b), we apply the triangle inequality, and then Part (a):

dH(X,Y)\displaystyle d_{\mathrm{H}}(X,Y) dH(T,X)+dH(T,Y)\displaystyle\leq d_{\mathrm{H}}(T,X)+d_{\mathrm{H}}(T,Y)
=max{dGH(T,X),dH(T,X)}+dH(T,Y)\displaystyle=\max\left\{d_{\mathrm{GH}}(T,X),\vec{d}_{\mathrm{H}}(\partial T,X)\right\}+d_{\mathrm{H}}(T,Y)
max{dGH(X,Y)+dGH(Y,T),dH(T,X)}+dH(T,Y)\displaystyle\leq\max\left\{d_{\mathrm{GH}}(X,Y)+d_{\mathrm{GH}}(Y,T),\vec{d}_{\mathrm{H}}(\partial T,X)\right\}+d_{\mathrm{H}}(T,Y)
max{dGH(X,Y)+dH(Y,T),dH(T,X)}+dH(T,Y)\displaystyle\leq\max\left\{d_{\mathrm{GH}}(X,Y)+d_{\mathrm{H}}(Y,T),\vec{d}_{\mathrm{H}}(\partial T,X)\right\}+d_{\mathrm{H}}(T,Y) since dGH(Y,T)dH(Y,T)\displaystyle\text{since }d_{\mathrm{GH}}(Y,T)\leq d_{\mathrm{H}}(Y,T)
=max{dGH(X,Y)+2dH(Y,T),dH(T,X)+dH(T,Y)}\displaystyle=\max\left\{d_{\mathrm{GH}}(X,Y)+2d_{\mathrm{H}}(Y,T),\vec{d}_{\mathrm{H}}(\partial T,X)+d_{\mathrm{H}}(T,Y)\right\}
max{dGH(X,Y)+2ε,dH(T,X)+ε}\displaystyle\leq\max\left\{d_{\mathrm{GH}}(X,Y)+2\varepsilon,\vec{d}_{\mathrm{H}}(\partial T,X)+\varepsilon\right\} since dH(Y,T)ε.\displaystyle\text{since }d_{\mathrm{H}}(Y,T)\leq\varepsilon.

We remark that if dH(T,X)=dH(T,X)d_{\mathrm{H}}(T,X)=\vec{d}_{\mathrm{H}}(\partial T,X), then dGH(T,X)d_{\mathrm{GH}}(T,X) can be arbitrarily small with respect to dH(T,X)d_{\mathrm{H}}(T,X) as demonstrated in Theorem 1.1.

As a particular case of Theorem 1.1, we can provide a closed formula to compute the Gromov–Hausdorff distance between a segment and a sample of it.

Corollary 4.1.

Let XX be a compact non-empty subset of an interval I=[a,b]I=[a,b]. Denote by c=minXc=\min X and d=maxXd=\max X. Then, dGH(I,X)=max{ca+bd2,dH([c,d],X)}d_{\mathrm{GH}}(I,X)=\max\left\{\frac{c-a+b-d}{2},d_{H}([c,d],X)\right\}.

Proof.

Let XIX^{\prime}\subseteq I be an isometric copy of XX centered in the interval II, meaning the distance from cc to the closest point in XX^{\prime} is equal to the distance from dd to the closest point in XX^{\prime}. Then,

dGH(I,X)=dGH(I,X)dH(I,X)=max{ca+bd2,dH([c,d],X)}.d_{\mathrm{GH}}(I,X)=d_{\mathrm{GH}}(I,X^{\prime})\leq d_{H}(I,X^{\prime})=\max\Big{\{}\tfrac{c-a+b-d}{2},d_{H}([c,d],X)\Big{\}}.

We represent the objects involved in Figure 2.

Since diam(I)diam(X)=ba(dc)\mathrm{diam}(I)-\mathrm{diam}(X)=b-a-(d-c), we have dGH(I,X)ca+bd2d_{\mathrm{GH}}(I,X)\geq\frac{c-a+b-d}{2}. There are now two cases. If dH([c,d],X)ca+bd2d_{H}([c,d],X)\leq\frac{c-a+b-d}{2}, the claim holds. Alternatively, suppose that dH([c,d],X)>ca+bd2d_{H}([c,d],X)>\frac{c-a+b-d}{2}. Then, this assumption implies that dH(I,X)=dH([c,d],X)d_{H}(I,X^{\prime})=d_{H}([c,d],X), and also that dH(I,X)>dH(I,X)d_{H}(I,X^{\prime})>\vec{d}_{H}(\partial I,X^{\prime}) since ca+bd2=dH(I,X)\tfrac{c-a+b-d}{2}=\vec{d}_{H}(\partial I,X^{\prime}). In virtue of the latter inequality, an application of Theorem 1.1(a) to the pair XIX^{\prime}\subseteq I gives dH(I,X)=dGH(I,X)d_{H}(I,X^{\prime})=d_{GH}(I,X^{\prime}). Hence, the claim descends since dH(I,X)=dH([c,d],X)d_{H}(I,X^{\prime})=d_{H}([c,d],X) and dGH(I,X)=dGH(I,X)d_{GH}(I,X^{\prime})=d_{GH}(I,X). ∎

aabbccddXXca+bd2\tfrac{c-a+b-d}{2}ca+bd2\tfrac{c-a+b-d}{2}II
aabbXX^{\prime}ca+bd2\tfrac{c-a+b-d}{2}ca+bd2\tfrac{c-a+b-d}{2}
Figure 2. A representation of the interval I=[a,b]I=[a,b] and the subsets XX and XX^{\prime} as described in Corollary 4.1 and its proof.

5. The case of the circle

Before considering general metric graphs in the following section, in this section we first understand the simplest connected metric graph which is not a tree: the circle.

In the case of the circle (G=S1G=S^{1}), the authors of [2] showed that dGH(S1,X)=dH(S1,X)d_{\mathrm{GH}}(S^{1},X)=d_{\mathrm{H}}(S^{1},X) for a compact subset XS1X\subseteq S^{1} with dGH(S1,X)<π6d_{\mathrm{GH}}(S^{1},X)<\frac{\pi}{6}. A curious question regarding the optimality of the density constant π6\frac{\pi}{6} was also posed therein [2, Question 2]. While the constant π6\frac{\pi}{6} provides a sufficient Gromov–Hausdorff density of a subset XX to guarantee the equality of its Hausdorff and Gromov–Hausdorff distances to S1S^{1}, it turns out this constant is sub-optimal. In this section, we show that π3\frac{\pi}{3} is the optimal constant in the case of the circle. That is, (i) for a subset XS1X\subseteq S^{1}, the density condition dGH(S1,X)<π3d_{\mathrm{GH}}(S^{1},X)<\frac{\pi}{3} implies dGH(S1,X)=dH(S1,X)d_{\mathrm{GH}}(S^{1},X)=d_{\mathrm{H}}(S^{1},X), and (ii) there is a compact XS1X\subseteq S^{1} with π3=dGH(S1,X)<dH(S1,X)\frac{\pi}{3}=d_{\mathrm{GH}}(S^{1},X)<d_{\mathrm{H}}(S^{1},X).

In order to prove (i), we need the following lifting lemma. Later in Lemma 6.1, we generalize the following lemma to general metric graphs.

Lemma 5.1 (Lifting of Loops for the Circle).

Let S1S^{1} be the circle with circumference 2π2\pi. Let XS1X\subseteq S^{1} be a subset and S1×X\mathcal{R}\subseteq S^{1}\times X a correspondence with 12dis()<r16e(S1)=π3\frac{1}{2}\mathrm{dis}(\mathcal{R})<r\leq\frac{1}{6}e(S^{1})=\frac{\pi}{3}. Then, [S1]r=S1\mathcal{R}[S^{1}]^{r}=S^{1}.

Proof.

Since dis()<2r\mathrm{dis}(\mathcal{R})<2r and S1S^{1} is connected, Lemma 2.3 implies that [S1]r\mathcal{R}[S^{1}]^{r} is connected. We show that [S1]r\mathcal{R}[S^{1}]^{r} is, indeed, the whole of S1S^{1}.

Suppose for a contradiction that [S1]r\mathcal{R}[S^{1}]^{r} is a connected, contractible subset of S1S^{1}, i.e., a segment ηS1\eta\subseteq S^{1} with endpoints a,ba,b. Since η\eta is a simple path, we can assume a total ordering \preceq of points on η\eta such that aba\preceq b. See Figure 3.

ξ\xiξ\xi^{\prime}ξ^\widehat{\xi}ξ^\widehat{\xi^{\prime}}
(a) The circle S1S^{1}
bbaaxxx^\widehat{x}x^\widehat{x}^{\prime}xx^{\prime}
(b) The segment η=[S1]rS1\eta=\mathcal{R}[S^{1}]^{r}\subseteq S^{1} is dashed in blue.
Figure 3.

For any ξS1\xi\in S^{1}, there is a unique antipodal point on the loop S1S^{1}, call it ξ^\widehat{\xi}. We note for later that d(ξ,ξ^)=πd(\xi,\widehat{\xi})=\pi. In light of the above notation, let us define the following two subsets AA and BB of S1S^{1}:

A{ξS1axx^b for some x[ξ],x^[ξ^]},A\coloneqq\left\{\xi\in S^{1}\mid a\preceq x\preceq\widehat{x}\preceq b\text{ for some }x\in\mathcal{R}[\xi],\widehat{x}\in\mathcal{R}[\widehat{\xi}]\right\},

and

B{ξS1ax^xb for some x[ξ],x^[ξ^]}.B\coloneqq\left\{\xi\in S^{1}\mid a\preceq\widehat{x}\preceq x\preceq b\text{ for some }x\in\mathcal{R}[\xi],\widehat{x}\in\mathcal{R}[\widehat{\xi}]\right\}.

We argue that AA and BB are non-empty, disjoint, open subsets of S1S^{1} such that AB=S1A\cup B=S^{1}. Since S1S^{1} is connected, this would lead to a contradiction. It is trivial to see that AB=S1A\cup B=S^{1}, since the correspondence \mathcal{R} projects surjectively onto S1S^{1}. We now prove the other three claims.

Non-empty

The subset AA\neq\emptyset since for any ξS1\xi\in S^{1}, either ξA\xi\in A or ξ^A\widehat{\xi}\in A. For the same reason, BB\neq\emptyset.

Open

We argue that AA is an open subset of S1S^{1}. Let ξA\xi\in A be arbitrary. Let x[ξ]x\in\mathcal{R}[\xi] and x^[ξ^]\widehat{x}\in\mathcal{R}[\widehat{\xi}] be such that axx^ba\preceq x\preceq\widehat{x}\preceq b.

Fix small ζ0\zeta\geq 0 such that dis()+ζ<2r\mathrm{dis}(\mathcal{R})+\zeta<2r. We will show that ξA\xi^{\prime}\in A for any ξS1\xi^{\prime}\in S^{1} with d(ξ,ξ)ζd(\xi,\xi^{\prime})\leq\zeta. We note that d(ξ^,ξ^)=d(ξ,ξ)ζd(\hat{\xi},\hat{\xi^{\prime}})=d(\xi,\xi^{\prime})\leq\zeta. Let us choose x[ξ]x^{\prime}\in\mathcal{R}[\xi^{\prime}] and x^[ξ^]\widehat{x^{\prime}}\in\mathcal{R}[\widehat{\xi^{\prime}}]. It suffices to show that axx^ba\preceq x^{\prime}\preceq\widehat{x^{\prime}}\preceq b. Without any loss of generality, we assume that axxba\preceq x\preceq x^{\prime}\preceq b. The configuration is shown in Figure 5(b). We have

d(x,x)dis()+d(ξ,ξ)dis()+ζ<2r<π.d(x,x^{\prime})\leq\mathrm{dis}(\mathcal{R})+d(\xi,\xi^{\prime})\leq\mathrm{dis}(\mathcal{R})+\zeta<2r<\pi.

Since d(x,x)<2rd(x,x^{\prime})<2r, a unique shortest path joining x,xx,x^{\prime} exists and is contained in η\eta. The same fact is also true for x^,x^\widehat{x},\widehat{x^{\prime}}, since

d(x^,x^)dis()+d(ξ^,ξ^)dis()+ζ<2r<π.d(\widehat{x},\widehat{x^{\prime}})\leq\mathrm{dis}(\mathcal{R})+d(\widehat{\xi},\widehat{\xi^{\prime}})\leq\mathrm{dis}(\mathcal{R})+\zeta<2r<\pi.

On the other hand, d(x,x^)d(ξ,ξ^)dis()>π2rd(x,\widehat{x})\geq d(\xi,\widehat{\xi})-\mathrm{dis}(\mathcal{R})>\pi-2r. Similarly, d(x,x^)>π2rd(x^{\prime},\widehat{x^{\prime}})>\pi-2r. Also, we have

d(x,x^)d(ξ,ξ^)dis()[d(ξ,ξ^)d(ξ^,ξ^)]dis()(πζ)dis()>π2r.d(x,\widehat{x^{\prime}})\geq d(\xi,\widehat{\xi^{\prime}})-\mathrm{dis}(\mathcal{R})\geq[d(\xi,\widehat{\xi})-d(\widehat{\xi},\widehat{\xi^{\prime}})]-\mathrm{dis}(\mathcal{R})\geq(\pi-\zeta)-\mathrm{dis}(\mathcal{R})>\pi-2r.

Similarly, we have d(x^,x)>π2rd(\widehat{x},x^{\prime})>\pi-2r.

Now, if axx^ba\preceq x^{\prime}\preceq\widehat{x^{\prime}}\preceq b, then ξA\xi^{\prime}\in A and there is nothing to show. If not, i.e., ax^xba\preceq\widehat{x^{\prime}}\preceq x^{\prime}\preceq b, then we arrive at a contradiction in each of the following three independent cases.

Case I (xx^x^x)(x\preceq\widehat{x}\preceq\widehat{x^{\prime}}\preceq x^{\prime})

The configuration is shown in Figure 5(b). In this case, we have

2r\displaystyle 2r >d(x,x)=d(x,x^)+d(x^,x^)+d(x^,x)d(x,x^)+d(x^,x^)>(π2r)+(π2r)=2π4r.\displaystyle>d(x,x^{\prime})=d(x,\widehat{x})+d(\widehat{x},\widehat{x^{\prime}})+d(\widehat{x^{\prime}},x^{\prime})\geq d(x,\widehat{x})+d(\widehat{x},\widehat{x^{\prime}})>(\pi-2r)+(\pi-2r)=2\pi-4r.

This implies r>π3r>\frac{\pi}{3}, a contradiction.

Case II (xx^x^x)(x\preceq\widehat{x^{\prime}}\preceq\widehat{x}\preceq x^{\prime})

Using the same calculations, we have

2r\displaystyle 2r >d(x,x)d(x,x^)+d(x^,x)>(π2r)+(π2r)=2π4r.\displaystyle>d(x,x^{\prime})\geq d(x,\widehat{x^{\prime}})+d(\widehat{x},x^{\prime})>(\pi-2r)+(\pi-2r)=2\pi-4r.

This implies r>π3r>\frac{\pi}{3}, again a contradiction.

Case III (xx^xx^)(x\preceq\widehat{x^{\prime}}\preceq x^{\prime}\preceq\widehat{x})

We have

2r\displaystyle 2r >d(x,x^)d(x,x^)+d(x,x^)>(π2r)+(π2r)=π4r.\displaystyle>d(x,\widehat{x})\geq d(x,\widehat{x^{\prime}})+d(x^{\prime},\widehat{x})>(\pi-2r)+(\pi-2r)=\pi-4r.

This implies that r>π3r>\frac{\pi}{3}, again a contradiction.

Therefore, AA is open in S1S^{1}. A similar argument holds to show that BB is open in S1S^{1}.

Disjoint

If not, let ξAB\xi\in A\cap B. Then there exist x[ξ]x\in\mathcal{R}[\xi] and x^[ξ^]\widehat{x}\in\mathcal{R}[\widehat{\xi}] such that axx^ba\preceq x\preceq\widehat{x}\preceq b, as well as x[ξ]x^{\prime}\in\mathcal{R}[\xi^{\prime}] and x^[ξ^]\widehat{x^{\prime}}\in\mathcal{R}[\widehat{\xi^{\prime}}] such that ax^xba\preceq\widehat{x^{\prime}}\preceq x^{\prime}\preceq b. Choosing ζ=0\zeta=0 in the calculations carried out to show that AA and BB are open, we similarly arrive at a contradiction. Therefore, AB=A\cap B=\emptyset.

We have contradicted the fact that S1S^{1} is connected. This proves that [S1]r=S1\mathcal{R}[S^{1}]^{r}=S^{1}. ∎

1[a]\mathcal{R}^{-1}[a]1[b]\mathcal{R}^{-1}[b]1[c]\mathcal{R}^{-1}[c]1[d]\mathcal{R}^{-1}[d]1[f]\mathcal{R}^{-1}[f]1[e]\mathcal{R}^{-1}[e]2π32ε\frac{2\pi}{3}-2\varepsilonπ6+ε\frac{\pi}{6}+\varepsilonπ6+ε\frac{\pi}{6}+\varepsilon
(a) Correspondence S1×X\mathcal{R}\subseteq S^{1}\times X has a distortion of 2π3\frac{2\pi}{3}. So, dGH(S1,X)π3d_{\mathrm{GH}}(S^{1},X)\leq\frac{\pi}{3}.
ddbbaaeeffcc2π3+2ε\frac{2\pi}{3}+2\varepsilonπ6ε\frac{\pi}{6}-\varepsilonπ3\frac{\pi}{3}π3\frac{\pi}{3}
(b) X={a,b,c,d,e,f}S1X=\{a,b,c,d,e,f\}\subseteq S^{1}
with dH(S1,X)=π3+εd_{\mathrm{H}}(S^{1},X)=\frac{\pi}{3}+\varepsilon.
Figure 4. The configuration of points XS1X\subseteq S^{1} is depicted to show the optimal constant for the circle.

We now prove the optimality of π3\frac{\pi}{3} for the circle. \sufficient

Proof.

(a) We note that S1S^{1} has only one edge and that e(S1)=2πe(S^{1})=2\pi. It suffices to show that if dGH(S1,X)<π3d_{\mathrm{GH}}(S^{1},X)<\frac{\pi}{3}, then dGH(S1,X)dH(S1,X)d_{\mathrm{GH}}(S^{1},X)\geq d_{\mathrm{H}}(S^{1},X).

Let dGH(S1,X)<π3d_{\mathrm{GH}}(S^{1},X)<\frac{\pi}{3}. We can fix dGH(S1,X)<rπ3d_{\mathrm{GH}}(S^{1},X)<r\leq\frac{\pi}{3} and correspondence S1×X\mathcal{R}\subseteq S^{1}\times X such that 12dis()<rπ3\frac{1}{2}\mathrm{dis}(\mathcal{R})<r\leq\frac{\pi}{3}. Lemma 5.1 implies that [S1]r=S1\mathcal{R}[S^{1}]^{r}=S^{1}. As a consequence rdH(S1,X)r\geq d_{\mathrm{H}}(S^{1},X).

So, we have rdH(S1,X)r\geq d_{\mathrm{H}}(S^{1},X) for any rr satisfying dGH(S1,X)<rπ3d_{\mathrm{GH}}(S^{1},X)<r\leq\frac{\pi}{3}. Therefore, dGH(S1,X)dH(S1,X)d_{\mathrm{GH}}(S^{1},X)\geq d_{\mathrm{H}}(S^{1},X). This proves (a).

The proof of (b) follows from (a) and the triangle inequality. ∎

We conclude the circle case by proving (ii), i.e., π3\frac{\pi}{3} is also a necessary density condition. \optimality

Proof.

We construct XX to be a six-element subset of S1S^{1}. Both XX and a particular correspondence S1×X\mathcal{R}\subseteq S^{1}\times X are shown in Figure 4. One can check that dH(S1,X)=π3+εd_{\mathrm{H}}(S^{1},X)=\frac{\pi}{3}+\varepsilon. Also, one can check that dis()=2π3\mathrm{dis}(\mathcal{R})=\frac{2\pi}{3}, giving dGH(S1,X)π3d_{\mathrm{GH}}(S^{1},X)\leq\frac{\pi}{3}. So dGH(S1,X)<dH(S1,X)d_{\mathrm{GH}}(S^{1},X)<d_{\mathrm{H}}(S^{1},X). Moreover, Theorem 1.1 implies that dGH(S1,X)=π3d_{\mathrm{GH}}(S^{1},X)=\frac{\pi}{3}. ∎

6. Metric graphs with loops

In this section, we extend our results for trees and for the circle to general metric graphs. We start by generalizing Lemma 6.1 to show the existence of lifting of loops for general metric graphs.

Lemma 6.1 (Existence of Lifting of Loops).

Let (G,d)(G,d) be a metric graph with (G)\mathcal{L}(G)\neq\emptyset and with e(G)e(G) the length of the smallest edge. Let XGX\subseteq G be a subset and G×X\mathcal{R}\subseteq G\times X a correspondence with 12dis()<r112e(G)\frac{1}{2}\mathrm{dis}(\mathcal{R})<r\leq\frac{1}{12}e(G). For any γ(G)\gamma\in\mathcal{L}(G), there exists a simple loop γ~(G)\widetilde{\gamma}\in\mathcal{L}(G) contained in [γ]r\mathcal{R}[\gamma]^{r}.

We remind that reader that while [γ]\mathcal{R}[\gamma] is a subset of XX, its thickening [γ]r\mathcal{R}[\gamma]^{r} is an open subset of GG.

Proof.

Since dis()<2r\mathrm{dis}(\mathcal{R})<2r and γ\gamma is connected, Lemma 2.3 implies that [γ]r\mathcal{R}[\gamma]^{r} is connected. We show that [γ]r\mathcal{R}[\gamma]^{r} contains at least one simple loop. Suppose for a contradiction that [γ]r\mathcal{R}[\gamma]^{r} is a connected, contractible subset of GG, i.e., a tree TGT\subseteq G.

For any ξγ\xi\in\gamma, there is a unique antipodal point on the loop γ\gamma, call it ξ^\widehat{\xi}. Let x[ξ]x\in\mathcal{R}[\xi] and x^[ξ^]\widehat{x}\in\mathcal{R}[\widehat{\xi}]. Since d(ξ,ξ^)e(G)2d(\xi,\widehat{\xi})\geq\frac{e(G)}{2}, we have

d(x,x^)d(ξ,ξ^)dis()>e(G)22r4r.d(x,\widehat{x})\geq d(\xi,\widehat{\xi})-\mathrm{dis}(\mathcal{R})>\tfrac{e(G)}{2}-2r\geq 4r. (4)

The last inequality is due to our assumption that r112e(G)r\leq\frac{1}{12}e(G).

Claim

Let tTt\in T be any point. There must exist xB(t,r)Xx\in B(t,r)\cap X such that x[ξ]x\in\mathcal{R}[\xi] for some ξγ\xi\in\gamma. Now, for any x^[ξ^]\widehat{x}\in\mathcal{R}[\widehat{\xi}], we have x^TB(t,r)\widehat{x}\in T\setminus B(t,r) since d(x,x^)>2rd(x,\widehat{x})>2r from (4). We first claim that x^\widehat{x} always belongs to the same connected component of TB(t,r)T\setminus B(t,r)—irrespective of the choice of such xB(t,r)Xx\in B(t,r)\cap X, ξ1[x]\xi\in\mathcal{R}^{-1}[x], and x^[ξ^]\widehat{x}\in\mathcal{R}[\widehat{\xi}].

Proof of Claim

To see the claim, let xiB(t,r)x_{i}\in B(t,r) and ξiγ\xi_{i}\in\gamma such that xi[ξi]x_{i}\in\mathcal{R}[\xi_{i}] and xi^[ξi^]\widehat{x_{i}}\in\mathcal{R}[\widehat{\xi_{i}}] for i=1,2i=1,2. Since d(x1,x2)<2rd(x_{1},x_{2})<2r, note that d(ξ1,ξ2)dis()+d(x1,x2)<2r+2r=4r12e(G)d(\xi_{1},\xi_{2})\leq\mathrm{dis}(\mathcal{R})+d(x_{1},x_{2})<2r+2r=4r\leq\frac{1}{2}e(G). Since γ\gamma is a simple loop and e(G)e(G) is the length of the shortest edge, the unique shortest arc joining ξ1,ξ2\xi_{1},\xi_{2} must lie on γ\gamma; call it γ\gamma^{\prime}. We further note that there cannot exist ξγ\xi\in\gamma^{\prime} with x^B(t,r)\widehat{x}\in B(t,r) for any x^[ξ^]\widehat{x}\in\mathcal{R}[\widehat{\xi}], since we would get d(ξi,ξ^)2r+d(xi,x^)<4rd(\xi_{i},\widehat{\xi})\leq 2r+d(x_{i},\widehat{x})<4r; see Figure 5(A). This would imply that the unique shortest arc joining ξi,ξ^\xi_{i},\widehat{\xi} also lies on γ\gamma for each i=1,2i=1,2—leading to the following contradiction: e(G)l(γ)=d(ξ1,ξ2)+d(ξ2,ξ^)+d(ξ^,ξ1)<12re(G)\leq l(\gamma)=d(\xi_{1},\xi_{2})+d(\xi_{2},\widehat{\xi})+d(\widehat{\xi},\xi_{1})<12r.

For any ξγ\xi\in\gamma^{\prime} and x^[ξ^]\widehat{x}\in\mathcal{R}[\widehat{\xi}], therefore, x^TB(t,r)\widehat{x}\in T\setminus B(t,r), and it lies in one of the connected components of TB(t,r)T\setminus B(t,r). Moreover, the association of the connected component is continuous, i.e., for any ξγ\xi^{\prime}\in\gamma^{\prime} with d(ξ,ξ)<2rdis()d(\xi,\xi^{\prime})<2r-\mathrm{dis}(\mathcal{R}) and x^[ξ^]\widehat{x^{\prime}}\in\mathcal{R}[\widehat{\xi^{\prime}}], we have d(x^,x^)<2rd(\widehat{x},\widehat{x^{\prime}})<2r, implying that x^\widehat{x^{\prime}} must also belong to the same component of TB(t,r)T\setminus B(t,r) as x^\widehat{x}. Indeed, any geodesic that is strictly shorter than 2r2r and connects points of [γ]\mathcal{R}[\gamma] must lay inside T=[γ]rT=\mathcal{R}[\gamma]^{r}. Since γ\gamma^{\prime} is connected, this proves the claim that both x^,x^\widehat{x},\widehat{x^{\prime}} must belong to the same component of TB(t,r)T\setminus B(t,r). Otherwise, it would result in a contradiction by producing a disconnection of γ\gamma^{\prime}.

<4r<4r<4r<4r<4r<4rξ1\xi_{1}ξ\xiξ2\xi_{2}ξ1^\widehat{\xi_{1}}ξ^\widehat{\xi}ξ2^\widehat{\xi_{2}}
(a) The loop γ(G)\gamma\in\mathcal{L}(G) is shown in gray. The point ξ\xi raising the contradiction is shown in red.
t2t_{2}t1t_{1}t3=bt_{3}=bt2^\widehat{t_{2}}t1^\widehat{t_{1}}t3^\widehat{t_{3}}a=t0a=t_{0}t0^\widehat{t_{0}}
(b) The tree T=[γ]rGT=\mathcal{R}[\gamma]^{r}\subseteq G is shown in gray and the constructed path η\eta is dashed in blue.
Figure 5.

Construction of η\eta

Next, we define a (possibly non-unique) directed path η\eta on TT by taking an arbitrary starting point aTa\in T. Letting t0at_{0}\coloneqq a, we note from the above claim that a unique connected component T0T_{0} of TB(t0,r)T\setminus B(t_{0},r) exists containing x^\widehat{x} for any xB(t0,r)x\in B(t_{0},r). Let t1t_{1} be the vertex of T0GT_{0}\subseteq G that is the closest to t0t_{0} along TT. If it exists, i.e., T0T_{0} contains vertices of GG, it is trivially well-defined; otherwise, set t1=x^t_{1}=\widehat{x} for any xB(t0,r)x\in B(t_{0},r). If t1t_{1} is a vertex of GG, then we reiterate the same procedure and construct a sequence of points tit_{i} for i=0,,ni=0,\dots,n. This sequence terminates at tnt_{n} if one of the following two cases occurs: either TnT_{n} does not contain a vertex of GG or TnT_{n} already contains tn1t_{n-1}. Since the graph is finite and TT has no loops, this procedure always terminates. Then, set btnb\coloneqq t_{n}; see Figure 5(B). Note that n1n\geq 1. Once the sequence {ti}i=0n\{t_{i}\}_{i=0}^{n} is fixed, we define η\eta to be the unique directed path joining the tit_{i}’s in the particular order. We assume a total ordering \prec of points on η\eta such that aba\prec b; see Figure 5.

For any tTt\in T, let us denote by Π(t)\Pi(t) the unique point on η\eta closest to xx along TT. Evidently, Π(t)=t\Pi(t)=t for any tηt\in\eta. It is also worth noticing that Π(Tη)={t0,,tn}\Pi(T\setminus\eta)=\{t_{0},\dots,t_{n}\} and the only vertices η\eta crosses are t1,,tn1t_{1},\dots,t_{n-1}. In light of the above notation, let us define three subsets of γ\gamma in the following manner:

A\displaystyle A {ξγd(Π(x),Π(x^))r and Π(x)Π(x^) for some x[ξ] and x^[ξ^]},\displaystyle\coloneqq\{\xi\in\gamma\mid d(\Pi(x),\Pi(\widehat{x}))\geq r\text{ and }\Pi(x)\prec\Pi(\widehat{x})\text{ for some }x\in\mathcal{R}[\xi]\text{ and }\widehat{x}\in\mathcal{R}[\widehat{\xi}]\},
B\displaystyle B {ξγd(Π(x),Π(x^))r and Π(x)Π(x^) for some x[ξ] and x^[ξ^]}, and\displaystyle\coloneqq\{\xi\in\gamma\mid d(\Pi(x),\Pi(\widehat{x}))\geq r\text{ and }\Pi(x)\succ\Pi(\widehat{x})\text{ for some }x\in\mathcal{R}[\xi]\text{ and }\widehat{x}\in\mathcal{R}[\widehat{\xi}]\},\text{ and }
C\displaystyle C {ξγd(Π(x),Π(x^))<r for some x[ξ] and x^[ξ^]}.\displaystyle\coloneqq\{\xi\in\gamma\mid d(\Pi(x),\Pi(\widehat{x}))<r\text{ for some }x\in\mathcal{R}[\xi]\text{ and }\widehat{x}\in\mathcal{R}[\widehat{\xi}]\}.

It is trivial to see that ABC=γA\cup B\cup C=\gamma, since the correspondence \mathcal{R} projects surjectively onto γ\gamma. Also, from the choice of aa and bb, we have aAa\in A and bBb\in B. We now argue that AA, BB, and CC are (pairwise) disjoint, open subsets of γ\gamma. Since γ\gamma is connected, this would lead to a contradiction.

η\etati=Π(x)t_{i}=\Pi(x)xxx^\widehat{x}ti+1=Π(x^)t_{i+1}=\Pi(\widehat{x})xx^{\prime}x^\widehat{x^{\prime}}B(Π(x),r)B(\Pi(x),r)
(a) Case I
η\etati=Π(x)t_{i}=\Pi(x)xxx^\widehat{x}ti+1=Π(x^)t_{i+1}=\Pi(\widehat{x})xx^{\prime}x^\widehat{x^{\prime}}B(Π(x),r)B(\Pi(x),r)
(b) Case II
η\etati=Π(x)t_{i}=\Pi(x)xxx^\widehat{x}ti+1t_{i+1}xx^{\prime}x^\widehat{x^{\prime}}B(Π(x),r)B(\Pi(x),r)
(c) CC
Figure 6.

Let us consider ξγ\xi\in\gamma an arbitrary point, x[ξ]x\in\mathcal{R}[\xi] and x^[ξ^]\widehat{x}\in\mathcal{R}[\widehat{\xi}]. Fix small ζ0\zeta\geq 0 such that dis()+ζ<2r\mathrm{dis}(\mathcal{R})+\zeta<2r and pick ξγ\xi^{\prime}\in\gamma satisfying d(ξ,ξ)ζd(\xi,\xi^{\prime})\leq\zeta. We claim that ξ\xi^{\prime} is contained in AA, BB, or CC provided that so is ξ\xi, which would show that they are open subsets of γ\gamma.

First of all, since d(ξ,ξ)<2rd(\xi,\xi^{\prime})<2r and γ\gamma is simple, we note that the unique shortest path connecting ξ\xi and ξ\xi^{\prime} is contained in γ\gamma. Therefore, we can deduce that d(ξ^,ξ^)=d(ξ,ξ)ζd(\hat{\xi},\hat{\xi^{\prime}})=d(\xi,\xi^{\prime})\leq\zeta. Let us choose x[ξ]x^{\prime}\in\mathcal{R}[\xi^{\prime}] and x^[ξ^]\widehat{x^{\prime}}\in\mathcal{R}[\widehat{\xi^{\prime}}]. We have d(x,x)dis()+d(ξ,ξ)dis()+ζ<2r<e(G)2d(x,x^{\prime})\leq\mathrm{dis}(\mathcal{R})+d(\xi,\xi^{\prime})\leq\mathrm{dis}(\mathcal{R})+\zeta<2r<\tfrac{e(G)}{2}. Consequently, if xx and xx^{\prime} lie on two edges of TT, the edges must be adjacent. Moreover, since d(x,x)<2rd(x,x^{\prime})<2r, a unique shortest path joining x,xx,x^{\prime} exists and is contained in TT. The same facts are also true for x^,x^\widehat{x},\widehat{x^{\prime}}, since d(x^,x^)dis()+d(ξ^,ξ^)dis()+ζ<2r<e(G)2d(\widehat{x},\widehat{x^{\prime}})\leq\mathrm{dis}(\mathcal{R})+d(\widehat{\xi},\widehat{\xi^{\prime}})\leq\mathrm{dis}(\mathcal{R})+\zeta<2r<\tfrac{e(G)}{2}.

On the other hand, d(x,x^)d(ξ,ξ^)dis()>e(G)22r4rd(x,\widehat{x})\geq d(\xi,\widehat{\xi})-\mathrm{dis}(\mathcal{R})>\frac{e(G)}{2}-2r\geq 4r.

Openness of AA (and BB)

Let us now assume that ξA\xi\in A and let x[ξ]x\in\mathcal{R}[\xi] and x^[ξ^]\widehat{x}\in\mathcal{R}[\widehat{\xi}] be such that d(Π(x),Π(x^))rd(\Pi(x),\Pi(\widehat{x}))\geq r and Π(x)Π(x^)\Pi(x)\prec\Pi(\widehat{x}).

We consider the following two cases:

Case I (xB(Π(x),r))(x\notin B(\Pi(x),r))

From the definition of AA we have Π(x^)B(Π(x),r)\Pi(\widehat{x})\notin B(\Pi(x),r). Consequently, x^B(Π(x),r)\widehat{x}\notin B(\Pi(x),r); see Figure 6(A) for the configuration. Since d(x,x)<2rd(x,x^{\prime})<2r, our first observation is that Π(x)B(Π(x),r)\Pi(x^{\prime})\in B(\Pi(x),r). Consequently, x^B(Π(x),r)\widehat{x^{\prime}}\notin B(\Pi(x),r); otherwise, the above claim would contradict the choice of ti+1t_{i+1}. Moreover, Π(x^)Π(x)\Pi(\widehat{x^{\prime}})\succ\Pi(x) due to the choice of the path η\eta and the above claim. Moreover, d(Π(x),Π(x^))rd(\Pi(x^{\prime}),\Pi(\widehat{x^{\prime}}))\geq r implying ξA\xi^{\prime}\in A. The argument also shows why ξ\xi^{\prime} cannot be in either BB or CC in this case.

Case II (xB(Π(x),r))(x\in B(\Pi(x),r)):

Since d(x,x^)>4rd(x,\widehat{x})>4r and d(x^,x^)<2rd(\widehat{x},\widehat{x^{\prime}})<2r, note that x^B(Π(x),r)\widehat{x^{\prime}}\notin B(\Pi(x),r). Moreover, Π(x^)Π(x)\Pi(\widehat{x^{\prime}})\succ\Pi(x) on η\eta ; see Figure 6(B) for the configuration. In order to maintain d(x^,x^)<2rd(\widehat{x},\widehat{x^{\prime}})<2r and d(x,x^)>4rd(x^{\prime},\widehat{x^{\prime}})>4r, we must also have Π(x^)Π(x)\Pi(\widehat{x^{\prime}})\succ\Pi(x^{\prime}) with d(Π(x),Π(x^))rd(\Pi(x^{\prime}),\Pi(\widehat{x^{\prime}}))\geq r implying ξA\xi^{\prime}\in A. The argument also shows why ξ\xi^{\prime} cannot be in either BB or CC in this case.

Therefore, AA is open in γ\gamma. Choosing ζ=0\zeta=0, i.e., ξ=ξ\xi=\xi^{\prime}, we also conclude from the cases above that AB=A\cap B=\emptyset and AC=A\cap C=\emptyset.

A similar argument holds to show that BB is open in γ\gamma and that BC=B\cap C=\emptyset.

Openness of CC

Let us now assume that ξC\xi\in C and let x[ξ]x\in\mathcal{R}[\xi] and x^[ξ^]\widehat{x}\in\mathcal{R}[\widehat{\xi}] be such that d(Π(x),Π(x^))<rd(\Pi(x),\Pi(\widehat{x}))<r.

Due to the construction of η\eta and the above claim, we must have Π(x)=Π(x^)=ti\Pi(x)=\Pi(\widehat{x})=t_{i} for some ii and that both x,x^TB(ti,r)x,\widehat{x}\in T\setminus B(t_{i},r); see Figure 6(C). Therefore, d(x,x)<2rd(x,x^{\prime})<2r and d(x^,x^)<2rd(\widehat{x},\widehat{x^{\prime}})<2r, which are obtained as in the previous cases, imply that d(Π(x),Π(x^))<rd(\Pi(x^{\prime}),\Pi(\widehat{x^{\prime}}))<r.

We have contradicted the fact that γ\gamma is connected. This proves that [γ]r\mathcal{R}[\gamma]^{r} cannot be contractible. Hence it must contain a simple loop γ~\widetilde{\gamma}. ∎

We define a lifting γ~\widetilde{\gamma} to be a simple loop in [γ]r\mathcal{R}[\gamma]^{r}. Now we prove that such a lifting γ~\widetilde{\gamma} must be unique under slightly more restrictive conditions on the length of the smallest edge.

Lemma 6.2 (Unique of Lifting of Loops).

Let (G,d)(G,d) be a metric graph with smallest edge length e(G)e(G). Let XGX\subseteq G be a subset and let G×X\mathcal{R}\subseteq G\times X be a correspondence with 12dis()<r112e(G)\frac{1}{2}\mathrm{dis}(\mathcal{R})<r\leq\frac{1}{12}e(G). For any γ(G)\gamma\in\mathcal{L}(G), there exists a unique simple loop γ~(G)\widetilde{\gamma}\in\mathcal{L}(G) contained in [γ]r\mathcal{R}[\gamma]^{r}.

Proof.

Since r18e(G)r\leq\frac{1}{8}e(G), it follows from Lemma 6.1 that [γ]r\mathcal{R}[\gamma]^{r} contains at least a simple loop. We argue that the loop is also unique.

We prove the uniqueness by contradiction. Let us assume to the contrary that there are at least two simple loops contained in [γ]r\mathcal{R}[\gamma]^{r}. We know from Lemma 2.3 that [γ]r\mathcal{R}[\gamma]^{r} is connected.

B(a,6r)B(a,6r)B(a,6r)B(a^{\prime},6r)ZZ
(a) Z[γ]rGZ\subseteq\mathcal{R}[\gamma]^{r}\subseteq G
B(ξ,2r)B(\xi,2r)B(ξ,2r)B(\xi^{\prime},2r)γ1\gamma_{1}γ2\gamma_{2}
(b) The loop γ(G)\gamma\in\mathcal{L}(G)
Figure 7.

Then, there must exist two open edges e,e[γ]re,e^{\prime}\in\mathcal{R}[\gamma]^{r} such that [γ]r{e,e}\mathcal{R}[\gamma]^{r}\setminus\{e,e^{\prime}\} is still connected; see Figure 7(A). Let a,aa,a^{\prime} be the midpoints of the edges. So, we have

  • d(a,a)e(G)12rd(a,a^{\prime})\geq e(G)\geq 12r, and

  • Z[γ]r(B(a,6r)B(a,6r))Z\coloneqq\mathcal{R}[\gamma]^{r}\setminus(B(a,6r)\cup B(a^{\prime},6r)) is connected and non-empty.

Let ξ,ξγ\xi,\xi^{\prime}\in\gamma and x[ξ],x[ξ]x\in\mathcal{R}[\xi],x^{\prime}\in\mathcal{R}[\xi^{\prime}] be such that aB(x,r),aB(x,r)a\in B(x,r),a^{\prime}\in B(x^{\prime},r). From the triangle inequality, we get

d(x,x)>d(a,a)2r12r2r=10r.d(x,x^{\prime})>d(a,a^{\prime})-2r\geq 12r-2r=10r.

Consequently,

d(ξ,ξ)d(x,x)dis()>10r2r>4r.d(\xi,\xi^{\prime})\geq d(x,x^{\prime})-\mathrm{dis}(\mathcal{R})>10r-2r>4r.

Because of this and the fact that γ\gamma is a simple loop, the subset

γγB(ξ,2r)B(ξ,2r)\gamma^{\prime}\coloneqq\gamma\setminus B(\xi,2r)\cup B(\xi^{\prime},2r)

must have two non-empty connected components, call them γ1\gamma_{1} and γ2\gamma_{2}; see Figure 7(B).

Finally, we define the following two subsets of ZZ for i=1,2i=1,2:

Zi[γi]rZ.Z_{i}\coloneqq\mathcal{R}[\gamma_{i}]^{r}\cap Z.

In the rest of the proof, we now argue that Z1,Z2Z_{1},Z_{2} are non-empty, open, and disjoint subsets of ZZ such that they partition ZZ, i.e., Z=Z1Z2Z=Z_{1}\cup Z_{2}. This would contradict the fact that ZZ is connected.

Non-empty

To see that ZiZ_{i}\neq\emptyset, it suffices to argue that [γi]r\mathcal{R}[\gamma_{i}]^{r} cannot be fully contained in B(a,6r)B(a,6r)B(a,6r)\cup B(a^{\prime},6r). We already know from Lemma 2.3 that [γi]r\mathcal{R}[\gamma_{i}]^{r} is connected since γi\gamma_{i} is connected. Hence, it suffices to argue, without any loss of generality, that [γi]r\mathcal{R}[\gamma_{i}]^{r} cannot be entirely contained in B(a,6r)B(a,6r).

If this were the case, we would have [γi]\mathcal{R}[\gamma_{i}] entirely contained in B(a,5r)B(a,5r) since GG is a length space. Consequently, for any ξiγi\xi_{i}\in\gamma_{i} and xi[ξi]x_{i}\in\mathcal{R}[\xi_{i}] we would have

d(ξi,ξ)\displaystyle d(\xi_{i},\xi^{\prime}) d(xi,x)dis()\displaystyle\geq d(x_{i},x^{\prime})-\mathrm{dis}(\mathcal{R})
d(xi,a)rdis()\displaystyle\geq d(x_{i},a^{\prime})-r-\mathrm{dis}(\mathcal{R})
(d(a,a)d(a,xi))rdis()\displaystyle\geq\big{(}d(a,a^{\prime})-d(a,x_{i})\big{)}-r-\mathrm{dis}(\mathcal{R})
>(12r5r)r2r\displaystyle>(12r-5r)-r-2r
=4r.\displaystyle=4r.

This would imply all points of γi\gamma_{i} are more than 4r4r distance away from ξ\xi^{\prime}, contradicting the construction of γi\gamma_{i}. This proves that the ZiZ_{i} sets are non-empty.

Partition

The direction Z1Z2ZZ_{1}\cup Z_{2}\subseteq Z is evident. For the other direction, take any zZz\in Z. Correspondingly, there exist some ξ1γ\xi_{1}\in\gamma and x1[ξ1]x_{1}\in\mathcal{R}[\xi_{1}] with d(x1,z)<rd(x_{1},z)<r. Since d(z,a)6rd(z,a)\geq 6r, we have

d(ξ,ξ1)d(x,x1)dis()d(a,z)2rdis()>2r.d(\xi,\xi_{1})\geq d(x,x_{1})-\mathrm{dis}(\mathcal{R})\geq d(a,z)-2r-\mathrm{dis}(\mathcal{R})>2r.

Similarly, we also get d(ξ,ξ1)>2rd(\xi^{\prime},\xi_{1})>2r. So, ξ1γ\xi_{1}\in\gamma^{\prime}. Therefore, ZZ1Z2Z\subseteq Z_{1}\cup Z_{2}.

Open

Let z1Z1z_{1}\in Z_{1}. Consequently, there exist ξ1γ1\xi_{1}\in\gamma_{1} and x1[ξ1]x_{1}\in\mathcal{R}[\xi_{1}] with x1B(z1,r)x_{1}\in B(z_{1},r). In order to show Z1Z_{1} is open in ZZ, we let z2Z1z_{2}\in Z_{1} with d(z1,z2)ηd(z_{1},z_{2})\leq\eta for sufficiently small η0\eta\geq 0 such that dis()+η<2r\mathrm{dis}(\mathcal{R})+\eta<2r.

We assume the contrary that z2Z2z_{2}\in Z_{2}, i.e., there exist ξ2γ2\xi_{2}\in\gamma_{2} and x2[ξ2]x_{2}\in\mathcal{R}[\xi_{2}] with z2B(x2,r)z_{2}\in B(x_{2},r). From the triangle inequality, we get d(x1,x2)<2r+ηd(x_{1},x_{2})<2r+\eta. This implies that

d(ξ1,ξ2)d(x1,x2)+dis()<(2r+η)+dis()<4r.d(\xi_{1},\xi_{2})\leq d(x_{1},x_{2})+\mathrm{dis}(\mathcal{R})<(2r+\eta)+\mathrm{dis}(\mathcal{R})<4r.

This contradicts the choice of γi\gamma_{i}’s. Therefore, z2Z1z_{2}\in Z_{1}. The openness of Z2Z_{2} can be shown similarly.

Disjoint

Let zZ1Z2z\in Z_{1}\cap Z_{2}, i.e., there exist ξiγi\xi_{i}\in\gamma_{i} and xi[ξi]x_{i}\in\mathcal{R}[\xi_{i}] with zB(x1,r)B(x2,r)z\in B(x_{1},r)\cap B(x_{2},r). Plugging in η=0\eta=0 in the proof of openness above, we again get d(ξ1,ξ2)<4rd(\xi_{1},\xi_{2})<4r, which contradicts the choice of γi\gamma_{i}’s. Therefore, Z1Z2=Z_{1}\cap Z_{2}=\emptyset.

In light of Lemma 6.1 and Lemma 6.2, we can define a well-defined lifting map Φ:(G)(G)\Phi\colon\mathcal{L}(G)\to\mathcal{L}(G) via γγ~\gamma\mapsto\widetilde{\gamma} as long as 12dis()<r112e(G)\frac{1}{2}\mathrm{dis}(\mathcal{R})<r\leq\frac{1}{12}e(G). We now show that the map Φ\Phi is a bijection.

Lemma 6.3 (Lifting of Loops is Bijective).

Let XX be a subset of a metric graph (G,d)(G,d). Let G×X\mathcal{R}\subseteq G\times X be a correspondence with 12dis()<r112e(G)\frac{1}{2}\mathrm{dis}(\mathcal{R})<r\leq\frac{1}{12}e(G). Then, the lifting map Φ:(G)(G)\Phi\colon\mathcal{L}(G)\to\mathcal{L}(G) is a bijection.

Proof.

It suffices to show that Φ\Phi is injective. Since (G)\mathcal{L}(G) is finite, the Pigeonhole Principle implies that Φ\Phi must also be surjective.

B(ξ,4r)B(\xi,4r)γ1\gamma_{1}γ2\gamma_{2}ξ\xi^{\prime}
(a) Z[γ]rGZ\subseteq\mathcal{R}[\gamma]^{r}\subseteq G
B(a,η)B(a,\eta)γ\gamma
(b) The loop γ(G)\gamma\in\mathcal{L}(G)
Figure 8.

To prove injectivity, suppose for a contradiction that there are two distinct simple loops γ1,γ2(G)\gamma_{1},\gamma_{2}\in\mathcal{L}(G) with Φ(γ1)=γ=Φ(γ2)\Phi(\gamma_{1})=\gamma=\Phi(\gamma_{2}); see Figure 8.

Since 8re(G)8r\leq e(G), there must be a point ξγ1\xi\in\gamma_{1} such that B(ξ,4r)γ2=B(\xi,4r)\cap\gamma_{2}=\emptyset, let and aγa\in\gamma be such that aB(x,r)a\in B(x,r) for some x[ξ]x\in\mathcal{R}[\xi].

Let ζ>0\zeta>0 be sufficiently small such that dis()+ζ<2r\mathrm{dis}(\mathcal{R})+\zeta<2r. We claim that [γ2]r[γ]rB(a,ζ)\mathcal{R}[\gamma_{2}]^{r}\subseteq\mathcal{R}[\gamma]^{r}\setminus B(a,\zeta). To see the claim take any ξγ2\xi^{\prime}\in\gamma_{2} and x[ξ]x^{\prime}\in\mathcal{R}[\xi^{\prime}]. By construction d(ξ,ξ)>4rd(\xi,\xi^{\prime})>4r. Consequently,

d(x,a)d(x,x)d(x,a)>(d(ξ,ξ)dis())r>4r(2rζ)r=r+ζ.d(x^{\prime},a)\geq d(x^{\prime},x)-d(x,a)>\left(d(\xi^{\prime},\xi)-\mathrm{dis}(\mathcal{R})\right)-r>4r-(2r-\zeta)-r=r+\zeta.

Since [γ]rB(a,ζ)\mathcal{R}[\gamma]^{r}\setminus B(a,\zeta) does not contain any loops, its subset [γ2]r\mathcal{R}[\gamma_{2}]^{r} cannot contain a loop. This contradicts Lemma 6.1. Hence, Φ\Phi is injective. ∎

Using this fact, we finally prove the following theorem for metric graphs.

\graphs
Proof.

If GG is a tree, an even stronger result follows directly from Theorem 1.1. For the non-trivial case, therefore, we assume that (G)\mathcal{L}(G)\neq\emptyset.

We first prove Part (a). It suffices to show that if dGH(G,X)<112e(G)d_{\mathrm{GH}}(G,X)<\frac{1}{12}e(G), then dGH(G,X)dH(G,X)d_{\mathrm{GH}}(G,X)\geq d_{\mathrm{H}}(G,X).

We fix r>0r>0 such that dGH(G,X)<r112e(G)d_{\mathrm{GH}}(G,X)<r\leq\frac{1}{12}e(G) and we fix a correspondence G×X\mathcal{R}\subseteq G\times X such that 12dis()<r112e(G)\frac{1}{2}\mathrm{dis}(\mathcal{R})<r\leq\frac{1}{12}e(G).

Let δ=dH(G,X)\delta=d_{\mathrm{H}}(G,X). Since GG is compact and δ<dH(G,X)\delta<\vec{d}_{\mathrm{H}}(\partial G,X), there exists a point mGm\in G such that B(m,δ)X=B(m,\delta)\cap X=\emptyset and there exist two distinct points x,xXx,x^{\prime}\in X with d(x,x)2δd(x,x^{\prime})\geq 2\delta such that mm lies on the shortest path γ\gamma joining them.

Depending on the position of mm, we now consider the following two cases to argue that r>δr>\delta:

Case I

In this case, we assume mm lies on a simple loop η(G)\eta\in\mathcal{L}(G). If we had rδr\leq\delta, the loop would not be entirely covered by [X]r\mathcal{R}[X]^{r} due to B(m,δ)X=B(m,\delta)\cap X=\emptyset. Consequently, η\eta would not be in the image of Φ\Phi, contradicting Lemma 6.3.

Case II

In this case, mm belongs to an edge ee of GG that does not form a simple loop. Then, either xx or xx^{\prime} must also belong to ee since mm lies on the shortest path joining x,xx,x^{\prime}. Without any loss of generality, let’s assume xex\in e.

Let ξ,ξG\xi,\xi^{\prime}\in G be such that x[ξ]x\in\mathcal{R}[\xi] and x[ξ]x^{\prime}\in\mathcal{R}[\xi^{\prime}]. Let α\alpha be a path joining ξ\xi and ξ\xi^{\prime}. Lemma 2.3 implies that there is a simple path joining x,xx,x^{\prime} in [α]r\mathcal{R}[\alpha]^{r}. Since ee does not form a simple cycle, the simple path must be γ\gamma itself. If we had rδr\leq\delta, this would render a contradiction since both m,xem,x\in e.

Therefore, it must be that r>δr>\delta for all dGH(G,X)<r112e(G)d_{\mathrm{GH}}(G,X)<r\leq\frac{1}{12}e(G). This concludes the proof that dGH(G,X)δd_{\mathrm{GH}}(G,X)\geq\delta, as desired.

In order to show Part (b), we apply the triangle inequality, and then Part (a):

dGH(X,Y)\displaystyle d_{\mathrm{GH}}(X,Y) dGH(G,X)dGH(G,Y)\displaystyle\geq d_{\mathrm{GH}}(G,X)-d_{\mathrm{GH}}(G,Y)
min{dH(G,X),112e(G)}dGH(G,Y)\displaystyle\geq\min\left\{d_{\mathrm{H}}(G,X),\tfrac{1}{12}e(G)\right\}-d_{\mathrm{GH}}(G,Y) from Part (a)
min{dH(G,X),112e(G)}dH(G,Y)\displaystyle\geq\min\left\{d_{\mathrm{H}}(G,X),\tfrac{1}{12}e(G)\right\}-d_{\mathrm{H}}(G,Y) since dGH(G,Y)dH(G,Y)\displaystyle\text{since }d_{\mathrm{GH}}(G,Y)\leq d_{\mathrm{H}}(G,Y)
min{dH(X,Y)dH(G,Y),112e(G)}dH(G,Y)\displaystyle\geq\min\left\{d_{\mathrm{H}}(X,Y)-d_{\mathrm{H}}(G,Y),\tfrac{1}{12}e(G)\right\}-d_{\mathrm{H}}(G,Y)
=min{dH(X,Y)2dH(G,Y),112e(G)dH(G,Y)}\displaystyle=\min\left\{d_{\mathrm{H}}(X,Y)-2d_{\mathrm{H}}(G,Y),\tfrac{1}{12}e(G)-d_{\mathrm{H}}(G,Y)\right\}
min{dH(G,X)2ε,112e(G)ε}\displaystyle\geq\min\left\{d_{\mathrm{H}}(G,X)-2\varepsilon,\tfrac{1}{12}e(G)-\varepsilon\right\} since dH(G,Y)ε.\displaystyle\text{since }d_{\mathrm{H}}(G,Y)\leq\varepsilon.

This concludes the proof. ∎

7. Conclusion and open questions

We have furthered the study of lower bounding the Gromov–Hausdorff distance of two subsets of a metric graph by their Hausdorff distance. Although we provide satisfactory results for the circle and metric graphs, the investigation also sparks several open questions and new research directions, especially for spaces beyond graphs.

We end by listing some open questions.

Question 1.

Is the density constant 112e(G)\frac{1}{12}e(G) in Theorem 1.1 optimal?

We conjecture that it is not and that 18e(G)\frac{1}{8}e(G) should suffice.

Question 2.

Do our results generalize to general length spaces? If so, under what density assumption?

A natural generalization would be to manifolds with boundary.

Question 3.

Are there classes of metric spaces (other than graphs) where the Gromov–Hausdorff distance between the space and a dense enough subset equals their Hausdorff distance?

References

  • [1] Henry Adams, Johnathan Bush, Nate Clause, Florian Frick, Mario Gómez, Michael Harrison, R. Amzi Jeffs, Evgeniya Lagoda, Sunhyuk Lim, Facundo Mémoli, Michael Moy, Nikola Sadovek, Matt Superdock, Daniel Vargas, Qingsong Wang, and Ling Zhou. Gromov–Hausdorff distances, Borsuk–Ulam theorems, and Vietoris–Rips complexes. arXiv preprint arXiv:2301.00246, 2023.
  • [2] Henry Adams, Florian Frick, Sushovan Majhi, and Nicholas McBride. Hausdorff vs Gromov–Hausdorff distances. arXiv preprint arXiv:2309.16648, 2023.
  • [3] Pankaj K. Agarwal, Kyle Fox, Abhinandan Nath, Anastasios Sidiropoulos, and Yusu Wang. Computing the Gromov–Hausdorff distance for metric trees. ACM Trans. Algorithms, 14(2), apr 2018.
  • [4] Martin R Bridson and André Haefliger. Metric spaces of non-positive curvature, volume 319. Springer Science & Business Media, 2011.
  • [5] Dmitri Burago, Yuri Burago, and Sergei Ivanov. A course in metric geometry, volume 33. American Mathematical Society, Providence, 2001.
  • [6] Gunnar Carlsson and Facundo Mémoli. Characterization, stability and convergence of hierarchical clustering methods. Journal of Machine Learning Research, 11(47):1425–1470, 2010.
  • [7] Frédéric Chazal, Vin de Silva, and Steve Oudot. Persistence stability for geometric complexes. Geometriae Dedicata, 174:193–214, 2014.
  • [8] David A Edwards. The structure of superspace. In Studies in topology, pages 121–133. Elsevier, 1975.
  • [9] Mikhael Gromov. Structures métriques pour les variétés riemanniennes. Textes Mathématiques, 1, 1981.
  • [10] Alexander Ivanov, Stavros Iliadis, and Alexey Tuzhilin. Realizations of Gromov–Hausdorff distance. arXiv preprint arXiv:1603.08850, 2016.
  • [11] Yibo Ji and Alexey A Tuzhilin. Gromov–Hausdorff distance between segment and circle. arXiv preprint arXiv:2101.05762, 2021.
  • [12] Nigel J Kalton and Mikhail I Ostrovskii. Distances between Banach spaces. Forum Math., 11:1–17, 1999.
  • [13] Sunhyuk Lim, Facundo Mémoli, and Zane Smith. The Gromov–Hausdorff distance between spheres. Accepted to appear in Geometry & Topology, 2022.
  • [14] Sushovan Majhi, Jeffrey Vitter, and Carola Wenk. Approximating Gromov–Hausdorff distance in Euclidean space. Computational Geometry, 116:102034, 2024.
  • [15] Facundo Mémoli. On the use of Gromov–Hausdorff distances for shape comparison. 2007.
  • [16] Facundo Mémoli. Some properties of Gromov–Hausdorff distances. Discrete & Computational Geometry, 48(2):416–440, 2012.
  • [17] Felix Schmiedl. Computational aspects of the Gromov-Hausdorff distance and its application in non-rigid shape matching. Discrete Comput. Geom., 57(4):854–880, 2017.
  • [18] Alexey A Tuzhilin. Who invented the Gromov–Hausdorff distance? arXiv preprint arXiv:1612.00728, 2016.
  • [19] Nicolò Zava. Coarse and bi-Lipschitz embeddability of subspaces of the gromov–Hausdorff space into Hilbert spaces. Algebr. Geom. Topol. to appear, arXiv preprint arXiv:2303.04730v2, 2024.