Lower Bounding the Gromov–Hausdorff distance in Metric Graphs
Abstract.
Let be a compact, connected metric graph and let be a subset. If is sufficiently dense in , we show that the Gromov–Hausdorff distance matches the Hausdorff distance, namely . In a recent study, when the metric graph is the circle with circumference , the equality was established whenever . Our result for general metric graphs relaxes this hypothesis in the circle case to , and furthermore, we give an example showing that the constant is the best possible. We lower bound the Gromov–Hausdorff distance by the Hausdorff distance via a simple topological obstruction: showing a correspondence with too small distortion contradicts the connectedness of . Moreover, for two sufficiently dense subsets , we provide new lower bounds on in terms of the Hausdorff distance , which is -time computable if the subsets have at most points.
Key words and phrases:
Gromov–Hausdorff distance, Hausdorff distance, Metric graph, Convexity radius1. Introduction
The Gromov–Hausdorff distance between two abstract metric spaces and , denoted , 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 [14]). Even approximating the Gromov–Hausdorff distance by a factor of 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 -approximation scheme if and 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 and , one may:
(A)
Find sufficiently nice samples and approximating the geometry of and ,
(B)
Efficiently bound the Gromov–Hausdorff distance between the subsets and .
Indeed, when and 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 have to be in 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 and constant such that whenever . In particular if . In [2], the authors showed that no such threshold or constant exist in general. More precisely, they demonstrated examples of metric spaces and subsets 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 when is a manifold, typically with , under the assumption that for some constant . The latter inequality implies that . Our study continues that investigation and proves improved results (such as with ) when 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 with a value in another computationally simpler metric space so that and are close whenever and 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 revealing the metric disparity between and . 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 -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 is not homeomorphic to for . Proving is not homeomorphic to for relies only on connectedness: removing a single point from leaves a disconnected space, which is not the case for with . However, to prove invariance of dimension for 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 be a connected, compact metric graph with finitely many edges, and let be a subset. We show how to lower bound the Gromov–Hausdorff distance in terms of constant multiplier times the Hausdorff distance , namely . We often obtain the optimal value . Since the reverse inequality always holds, the inequality (with optimal constant ) immediately gives equality .
The simplest metric graph is a tree. Let denote the directed Hausdorff distance from the leaves of to the sample (see Section 2). We show the following. {restatable*}[Metric Trees]thmtrees Let be a compact metric tree. For any ,
-
(a)
if , then , and
-
(b)
for any , if , then .
The interpretation of part (a) of Theorem 1.1 is as follows: if the biggest gaps in the sampling are not near the leaves of , then .
The following theorem shows that some control on the set of leaf vertices of is needed.
thmcounterEx For any there exists a compact metric tree and a closed subset such that .
When is the circle of circumference , the authors of [2] showed that if , then . The question of the optimality of the density constant was subsequently raised. In this paper, we show that is suboptimal: the optimal density constant is in fact . The first result below shows that the new weaker assumption is necessary, and the second result shows it is sufficient.
[Optimality of ]thmoptimality For any , there exists a nonempty compact subset with and .
[The Circle]thmsufficient For any subsets , we have
-
(a)
, and
-
(b)
for any .
We generalize the above results for the circle and trees to general metric graphs. Let be a compact metric graph and let be the shortest edge length between non-leaf vertices (see (2)). We prove that if is dense enough relative to the smallest edge length , then the Gromov–Hausdorff distance matches the Hasudorff distance.
[Metric Graphs]thmgraphs Let be a compact, connected metric graph. Let be subsets such that if . Then
-
(a)
, and
-
(b)
for any .
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 be a metric space. For any , we let denote the metric ball of radius about . If needed for the sake of clarity, we will sometimes write instead of . For any and , we let
(1) |
denote the -thickening of in , namely, the union of open metric balls of radius centered at the points of .
The diameter of a subset is the supremum of all distances between pairs of points of . More formally, the diameter is defined as
The diameter of a compact set 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 for . Our metric graphs will always be connected and have a finite number of vertices and edges.
Definition 2.1 (Metric Graph).
A metric graph is the metric space obtained by gluing a finite collection of metric segments along their boundaries to a finite collection of vertices, and equipping the result with the quotient metric. We assume this equivalence relation results in a connected space.
The metric segments are called the edges and the equivalence classes of their endpoints in are called the vertices of . Since is connected, we have for all . Since is finite, the quotient semi-metric [5, Definition 3.1.12] is in fact a metric.111 One needs to be more careful if is infinite. Indeed, suppose one has two vertices and , and one edge of length for each with endpoints and . Then the quotient semi-metric gives , even though , and hence this semi-metric is not a metric. These subtleties disappear since we assume is finite.
For an edge , its length is the length of the corresponding metric segment. Note that the length can be different from the distance between the endpoints of according to the metric . Extending the definition of length for a continuous path , we define the length as the sum of the (full or partial) lengths of the edges that traverses.
For a vertex of a graph, its degree, denoted , is the number of edges incident to . A self-loop at contributes to the degree of . As an example, the circle can be given a metric graph structure by assigning a single vertex at the north pole, of degree , having a self-loop.
A metric graph 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 as an abstract graph. Using the above definition, we are also allowing to have single-edge cycles and multiple edges between a pair of vertices. In this paper, we only consider path-connected, compact metric graphs with finitely many edges. The metric graphs we consider are automatically geodesic (i.e., for every there is an isometric embedding such that and ) 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 , there always exists a shortest path or geodesic path in joining whose length satisfies . We define an undirected simple loop of to be a simple path that intersects itself only at the endpoints. We denote by the set of all simple loops of . Since is assumed to have finitely many edges, the set is finite. Since a metric tree is defined as a metric graph with no loops, for a metric tree .
For a metric graph with , let denote the length of the smallest non-terminal edge:
(2) |
Hausdorff distances
Let be a metric space. If and are two subsets of , then the directed Hausdorff distance from to , denoted , is defined by
Note that is not symmetric in general.
To retain symmetry, the Hausdorff distance, denoted , is defined as
In other words, the Hausdorff distance finds the infimum over all real numbers such that if we thicken by it contains and if we thicken by it contains . If no such exists then the Hausdorff distance is infinity.
Gromov–Hausdorff distances
Let be non-empty sets. A relation is called a correspondence if the following two conditions hold: for every there exists some with , and for every there exists some exists with .
If is a correspondence between and , then the distortion of is:
The Gromov–Hausdorff distance between two (arbitrary) metric spaces and to be
(3) |
where the infimum is taken over all correspondences between and [5, 12]. As shown in [10], if and are both compact metric spaces, a distortion-minimizing correspondence in the definition of always exists.
The above definition immediately provides the following simple lower bound on the Gromov–Hausdorff distance between two compact metric spaces and .
Example 2.2 (Trivial Lower Bound).
Let and be compact metric spaces. Without loss of generality, we assume that . Take an arbitrary correspondence and pairs of points and such that and . The distortion satisfies
We conclude
Let be metric spaces and be a correspondence. For a subset , we define the following subset of :
Similarly, for a subset , we define the following subset of :
Using the notation above, we observe the following fact.
Lemma 2.3 (Lifting of Connected Subsets).
Let be a subset of a length metric space and let be a correspondence with distortion . If is a connected subset, then (see (1)) is connected.
Proof.
Suppose for a contradiction that , where are disjoint, non-empty, open subsets of . Note that is a union of balls in , which are connected since is a length space. Therefore, we must have a partition of , say , such that .
Since the sets are non-empty, so are the sets. To see that the sets are open, take and a corresponding point . For any with and , we have
So, as is a length metric space. As a result, both . So each set is open.
Moreover, taking and in the above, we see for any two corresponding points that . This implies that .
So, the sets create a partition of . This is a contradiction since is connected. Therefore, is connected. ∎
3. Ratio of vs 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 between a metric graph and a subset can be bounded below by their Hausdorff distance in case is dense enough near the leaves of , i.e., . When , however, the ratio of over can become arbitrarily small as we show below. Furthermore, Theorem 1.1, a sketch of which is provided in Figure 1, suggests that 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 is the infimum of the Hausdorff distances between and in , with the infimum being over all metric spaces and all isometric embeddings and .
Proof.
Choose and let . For , let be a copy of the interval with a designated basepoint .
Define a metric tree as the wedge of intervals of lengths , i.e.,
Define a metric tree as the wedge of intervals of lengths :
In Figure 1 we have two different isometric embeddings of the same tree into . One with a small Hausdorff distance as is with each ray being shortened by , and one with a large Hausdorff distance as is with the longest ray of length being replaced with a subray of length . By the definition above,
∎
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.
Proof.
We make no distinction between a path and its image .
We first prove Part (a). Assuming that , we show that , which implies . The assumption implies that there are two points with such that the midpoint of the unique shortest path joining them is at least distance away from all points of . Consequently, for any and for any subset containing both and , the -thickening cannot be connected. In particular, cannot be connected.
If we had , we could choose and there would exist a correspondence with . Since is connected, Lemma 2.3 would imply that is connected—a contradiction. Therefore, , i.e., .
In order to show Part (b), we apply the triangle inequality, and then Part (a):
∎
We remark that if , then can be arbitrarily small with respect to 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 be a compact non-empty subset of an interval . Denote by and . Then, .
Proof.
Let be an isometric copy of centered in the interval , meaning the distance from to the closest point in is equal to the distance from to the closest point in . Then,
We represent the objects involved in Figure 2.
Since , we have . There are now two cases. If , the claim holds. Alternatively, suppose that . Then, this assumption implies that , and also that since . In virtue of the latter inequality, an application of Theorem 1.1(a) to the pair gives . Hence, the claim descends since and . ∎
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 (), the authors of [2] showed that for a compact subset with . A curious question regarding the optimality of the density constant was also posed therein [2, Question 2]. While the constant provides a sufficient Gromov–Hausdorff density of a subset to guarantee the equality of its Hausdorff and Gromov–Hausdorff distances to , it turns out this constant is sub-optimal. In this section, we show that is the optimal constant in the case of the circle. That is, (i) for a subset , the density condition implies , and (ii) there is a compact with .
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 be the circle with circumference . Let be a subset and a correspondence with . Then, .
Proof.
Since and is connected, Lemma 2.3 implies that is connected. We show that is, indeed, the whole of .
Suppose for a contradiction that is a connected, contractible subset of , i.e., a segment with endpoints . Since is a simple path, we can assume a total ordering of points on such that . See Figure 3.
For any , there is a unique antipodal point on the loop , call it . We note for later that . In light of the above notation, let us define the following two subsets and of :
and
We argue that and are non-empty, disjoint, open subsets of such that . Since is connected, this would lead to a contradiction. It is trivial to see that , since the correspondence projects surjectively onto . We now prove the other three claims.
Non-empty
The subset since for any , either or . For the same reason, .
Open
We argue that is an open subset of . Let be arbitrary. Let and be such that .
Fix small such that . We will show that for any with . We note that . Let us choose and . It suffices to show that . Without any loss of generality, we assume that . The configuration is shown in Figure 5(b). We have
Since , a unique shortest path joining exists and is contained in . The same fact is also true for , since
On the other hand, . Similarly, . Also, we have
Similarly, we have .
Now, if , then and there is nothing to show. If not, i.e., , then we arrive at a contradiction in each of the following three independent cases.
Case I
Case II
Using the same calculations, we have
This implies , again a contradiction.
Case III
We have
This implies that , again a contradiction.
Therefore, is open in . A similar argument holds to show that is open in .
Disjoint
If not, let . Then there exist and such that , as well as and such that . Choosing in the calculations carried out to show that and are open, we similarly arrive at a contradiction. Therefore, .
We have contradicted the fact that is connected. This proves that . ∎
with .
We now prove the optimality of for the circle. \sufficient
Proof.
(a) We note that has only one edge and that . It suffices to show that if , then .
Let . We can fix and correspondence such that . Lemma 5.1 implies that . As a consequence .
So, we have for any satisfying . Therefore, . This proves (a).
The proof of (b) follows from (a) and the triangle inequality. ∎
We conclude the circle case by proving (ii), i.e., is also a necessary density condition. \optimality
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 be a metric graph with and with the length of the smallest edge. Let be a subset and a correspondence with . For any , there exists a simple loop contained in .
We remind that reader that while is a subset of , its thickening is an open subset of .
Proof.
Since and is connected, Lemma 2.3 implies that is connected. We show that contains at least one simple loop. Suppose for a contradiction that is a connected, contractible subset of , i.e., a tree .
For any , there is a unique antipodal point on the loop , call it . Let and . Since , we have
(4) |
The last inequality is due to our assumption that .
Claim
Let be any point. There must exist such that for some . Now, for any , we have since from (4). We first claim that always belongs to the same connected component of —irrespective of the choice of such , , and .
Proof of Claim
To see the claim, let and such that and for . Since , note that . Since is a simple loop and is the length of the shortest edge, the unique shortest arc joining must lie on ; call it . We further note that there cannot exist with for any , since we would get ; see Figure 5(A). This would imply that the unique shortest arc joining also lies on for each —leading to the following contradiction: .
For any and , therefore, , and it lies in one of the connected components of . Moreover, the association of the connected component is continuous, i.e., for any with and , we have , implying that must also belong to the same component of as . Indeed, any geodesic that is strictly shorter than and connects points of must lay inside . Since is connected, this proves the claim that both must belong to the same component of . Otherwise, it would result in a contradiction by producing a disconnection of .
Construction of
Next, we define a (possibly non-unique) directed path on by taking an arbitrary starting point . Letting , we note from the above claim that a unique connected component of exists containing for any . Let be the vertex of that is the closest to along . If it exists, i.e., contains vertices of , it is trivially well-defined; otherwise, set for any . If is a vertex of , then we reiterate the same procedure and construct a sequence of points for . This sequence terminates at if one of the following two cases occurs: either does not contain a vertex of or already contains . Since the graph is finite and has no loops, this procedure always terminates. Then, set ; see Figure 5(B). Note that . Once the sequence is fixed, we define to be the unique directed path joining the ’s in the particular order. We assume a total ordering of points on such that ; see Figure 5.
For any , let us denote by the unique point on closest to along . Evidently, for any . It is also worth noticing that and the only vertices crosses are . In light of the above notation, let us define three subsets of in the following manner:
It is trivial to see that , since the correspondence projects surjectively onto . Also, from the choice of and , we have and . We now argue that , , and are (pairwise) disjoint, open subsets of . Since is connected, this would lead to a contradiction.
Let us consider an arbitrary point, and . Fix small such that and pick satisfying . We claim that is contained in , , or provided that so is , which would show that they are open subsets of .
First of all, since and is simple, we note that the unique shortest path connecting and is contained in . Therefore, we can deduce that . Let us choose and . We have . Consequently, if and lie on two edges of , the edges must be adjacent. Moreover, since , a unique shortest path joining exists and is contained in . The same facts are also true for , since .
On the other hand, .
Openness of (and )
Let us now assume that and let and be such that and .
We consider the following two cases:
Case I
From the definition of we have . Consequently, ; see Figure 6(A) for the configuration. Since , our first observation is that . Consequently, ; otherwise, the above claim would contradict the choice of . Moreover, due to the choice of the path and the above claim. Moreover, implying . The argument also shows why cannot be in either or in this case.
Case II :
Since and , note that . Moreover, on ; see Figure 6(B) for the configuration. In order to maintain and , we must also have with implying . The argument also shows why cannot be in either or in this case.
Therefore, is open in . Choosing , i.e., , we also conclude from the cases above that and .
A similar argument holds to show that is open in and that .
Openness of
Let us now assume that and let and be such that .
Due to the construction of and the above claim, we must have for some and that both ; see Figure 6(C). Therefore, and , which are obtained as in the previous cases, imply that .
We have contradicted the fact that is connected. This proves that cannot be contractible. Hence it must contain a simple loop . ∎
We define a lifting to be a simple loop in . Now we prove that such a lifting must be unique under slightly more restrictive conditions on the length of the smallest edge.
Lemma 6.2 (Unique of Lifting of Loops).
Let be a metric graph with smallest edge length . Let be a subset and let be a correspondence with . For any , there exists a unique simple loop contained in .
Proof.
Since , it follows from Lemma 6.1 that 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 . We know from Lemma 2.3 that is connected.
Then, there must exist two open edges such that is still connected; see Figure 7(A). Let be the midpoints of the edges. So, we have
-
•
, and
-
•
is connected and non-empty.
Let and be such that . From the triangle inequality, we get
Consequently,
Because of this and the fact that is a simple loop, the subset
must have two non-empty connected components, call them and ; see Figure 7(B).
Finally, we define the following two subsets of for :
In the rest of the proof, we now argue that are non-empty, open, and disjoint subsets of such that they partition , i.e., . This would contradict the fact that is connected.
Non-empty
To see that , it suffices to argue that cannot be fully contained in . We already know from Lemma 2.3 that is connected since is connected. Hence, it suffices to argue, without any loss of generality, that cannot be entirely contained in .
If this were the case, we would have entirely contained in since is a length space. Consequently, for any and we would have
This would imply all points of are more than distance away from , contradicting the construction of . This proves that the sets are non-empty.
Partition
The direction is evident. For the other direction, take any . Correspondingly, there exist some and with . Since , we have
Similarly, we also get . So, . Therefore, .
Open
Let . Consequently, there exist and with . In order to show is open in , we let with for sufficiently small such that .
We assume the contrary that , i.e., there exist and with . From the triangle inequality, we get . This implies that
This contradicts the choice of ’s. Therefore, . The openness of can be shown similarly.
Disjoint
Let , i.e., there exist and with . Plugging in in the proof of openness above, we again get , which contradicts the choice of ’s. Therefore, .
∎
In light of Lemma 6.1 and Lemma 6.2, we can define a well-defined lifting map via as long as . We now show that the map is a bijection.
Lemma 6.3 (Lifting of Loops is Bijective).
Let be a subset of a metric graph . Let be a correspondence with . Then, the lifting map is a bijection.
Proof.
It suffices to show that is injective. Since is finite, the Pigeonhole Principle implies that must also be surjective.
To prove injectivity, suppose for a contradiction that there are two distinct simple loops with ; see Figure 8.
Since , there must be a point such that , let and be such that for some .
Let be sufficiently small such that . We claim that . To see the claim take any and . By construction . Consequently,
Since does not contain any loops, its subset cannot contain a loop. This contradicts Lemma 6.1. Hence, is injective. ∎
Using this fact, we finally prove the following theorem for metric graphs.
Proof.
If is a tree, an even stronger result follows directly from Theorem 1.1. For the non-trivial case, therefore, we assume that .
We first prove Part (a). It suffices to show that if , then .
We fix such that and we fix a correspondence such that .
Let . Since is compact and , there exists a point such that and there exist two distinct points with such that lies on the shortest path joining them.
Depending on the position of , we now consider the following two cases to argue that :
Case I
In this case, we assume lies on a simple loop . If we had , the loop would not be entirely covered by due to . Consequently, would not be in the image of , contradicting Lemma 6.3.
Case II
In this case, belongs to an edge of that does not form a simple loop. Then, either or must also belong to since lies on the shortest path joining . Without any loss of generality, let’s assume .
Let be such that and . Let be a path joining and . Lemma 2.3 implies that there is a simple path joining in . Since does not form a simple cycle, the simple path must be itself. If we had , this would render a contradiction since both .
Therefore, it must be that for all . This concludes the proof that , as desired.
In order to show Part (b), we apply the triangle inequality, and then Part (a):
from Part (a) | ||||
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 in Theorem 1.1 optimal?
We conjecture that it is not and that 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.