Notices of the American Mathematical Society

Welcome to the current issue of the Notices of the American Mathematical Society.
With support from AMS membership, we are pleased to share the journal with the global mathematical community.


On the Geometry of Metric Spaces

Manuel Ritoré

Communicated by Notices Associate Editor Chikako Mese

Article cover

1. Introduction

The measurement of distances because of practical reasons, as the delimitation of parcels of farmland after floodings or those related to construction problems, lies behind the origins of mathematics in ancient civilizations. It was Euclid, in Proposition 20 of the first book of the Elements, who proved that the sum of two sides of a triangle is greater than the remaining one. This fact was probably used by Archimedes to state as assumption, in his work On the sphere and cylinder, that the straight line is the shortest between two points. Both achievements are now proven in basic courses on Linear Algebra.

Presently, the notion of distance on a nonempty set is formalized as a function satisfying

1.

if and only if ,

2.

for all ,

3.

for all .

These conditions imply . The second property is the symmetry one. In Gromov’s words this assumption “[…] unpleaseantly limits many applications.” Removing this symmetry condition gives rise to the concept of asymmetric distance, see Busemann’s Local metric geometry (1958). The third property is the well-known triangle inequality.

The pair , where is a nonempty set and a distance on , is called a metric space.

Metric spaces were introduced by Fréchet in 1906 in the context of functional analysis. Shortly after, a comprehensive treatment of the theory was presented by Hausdorff in the second part of his monograph Hau14 published in 1914, where the notion of topological space was introduced.

Metric spaces appear everywhere in modern mathematics. They are an invaluable tool in functional analysis, geometry of manifolds, graph theory, artificial intelligence, and many other fields. The aim of this note is to give a glance of the role of metric spaces in several fields of mathematics without any intention of being exhaustive. We focus on different notions of curvature in metric spaces, compactness results for Riemannian manifolds, and an application of metric graph theory to group theory.

2. Length Spaces

Given a subset of a Euclidean space we may consider the restriction of the Euclidean distance to the set. For instance, the distance between two antipodal points on a unit sphere is . However, this distance is not intrinsic in the sense that cannot be measured inside the subset.

Given a continuous curve on a metric space, its length can be defined as the supremum of the quantities

over all partitions , , of the interval . When is finite the curve is said to be rectifiable. Hence the length of curves, sometimes infinite, can be measured in metric spaces.

A metric space is a length space if the distance between any pair of points can be computed as

where is any rectifiable curve connecting and (i.e., such that , ). A length space is geodesic if for every pair of points , there exists a rectifiable curve connecting , such that . A geodesic in is a rectifiable curve such that for all . The interested reader is referred to Burago, Burago and Ivanov’s BBI01 for a complete presentation of length spaces.

Examples of length spaces are Riemannian, Finsler and sub-Riemannian manifolds. A Riemannian manifold is a manifold together with an scalar product on the fiber bundle (a continuously varying scalar product on each tangent space). In a Finsler manifold the scalar product is replaced by a smooth norm. A sub-Riemannian manifold is a manifold together with a nonintegrable horizontal distribution and a positive definite scalar product on . The classical example of a sub-Riemannian manifold is the Heisenberg group : the -dimensional Euclidean space together with a noncommutative product and the left-invariant vector fields

induce a left-invariant Riemannian metric making an orthonormal basis. The vectors fields generate a nonintegrable distribution that, together with the restriction of to , provide a sub-Riemannian structure on . On a connected sub-Riemannian manifold , Chow–Rashevskii theorem, see §0.4 and §1.1, §1.2 in Gromov Gro96, implies that every pair of points can be connected by a smooth horizontal curve on (i.e., everywhere tangent to the horizontal distribution).

To define a distance we consider, in the Riemannian and Finsler cases, the class of piecewise smooth curves and, in the sub-Riemannian case, the class of piecewise smooth horizontal curves. In all cases, the length of a curve can be computed as

where is the norm associated to in the Riemannian case, the Finsler norm, or the one associated to in the sub-Riemannian case. Associated to this length we can define a distance between two points as the infimum of the length of the curves joining both points. This way we obtain the Riemannian distance, the Finsler distance, or the Carnot-Carathéodory distance. Geodesics are curves of minimum distance connecting two given points.

Figure 1.

Geodesics in the Riemannian and sub-Riemannian Heisenberg group connecting two points in the same vertical line. The vertical line is a Riemannian geodesic. The helicoidal curve is a sub-Riemannian geodesic, forced to be tangent to the horizontal distribution.

Graphic without alt text

3. Metric Spaces with Bounded Curvature

Curvature in Riemannian geometry was introduced as a measure of how a Riemannian space deviates from Euclidean space, see Gromov Gro91 for an enlightening discussion. For surfaces embedded in Euclidean space, the curvature coincides with the product of the principal curvatures of the surface and it is referred to as the Gauß curvature. The classical Gauß–Bonnet theorem provides a first hint on how the curvature affects the behavior of geodesic triangles on a surface. If is a triangle delimited by geodesics making inner angles and then

Hence the sum of the inner angles of a triangle on a surface depends on the sign of the Gauß curvature: it is equal to on surfaces with , larger than on surfaces with and smaller than when .

Figure 2.

Behavior of geodesics triangles according to Gauß–Bonnet theorem.

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \coordinate(A) at (0,0); \begin{scope}[shift={(6,5.5)}] \draw(0,0) node{$K>0$}; \end{scope} \begin{scope}[shift={(6,3)}] \draw(0,0) node{$K=0$}; \end{scope} \begin{scope}[shift={(6,0.5)}] \draw(0,0) node{$K<0$}; \end{scope} \begin{scope}[shift={(1,5)}] \draw(0,0) to[out=-10, in=210] (3,0.5); \draw(0,0) to[out=90,in=235] (0.5,1.75); \draw(0.5,1.75) to[out=5,in=130] (3,0.5); \draw[fill=black] (0,0) circle (0.05); \draw[fill=black] (3,0.5) circle (0.05); \draw[fill=black] (0.5,1.75) circle (0.05); \end{scope} \begin{scope}[shift={(1,2.5)}] \draw(0,0) -- (3,0.5); \draw(0,0) -- (0.5,1.75); \draw(0.5,1.75) -- (3,0.5); \draw[fill=black] (0,0) circle (0.05); \draw[fill=black] (3,0.5) circle (0.05); \draw[fill=black] (0.5,1.75) circle (0.05); \end{scope} \begin{scope}[shift={(1,0)}] \draw(0,0) to[out=35,in=175] (3,0.5); \draw(0,0) to[out=65,in=270] (0.5,1.75); \draw(0.5,1.75) to[out=-35,in=165] (3,0.5); \draw[fill=black] (0,0) circle (0.05); \draw[fill=black] (3,0.5) circle (0.05); \draw[fill=black] (0.5,1.75) circle (0.05); \end{scope} \end{tikzpicture}

According to the historical account in the preface of Alexander, Kapovitch, and Petrunin’s monograph, see arXiv:1903.08539, the first synthetic description of curvature is due to A. Wald in a paper published in 1936, followed by a paper by Alexandrov in 1941. It is generally considered that Rauch’s comparison results for Jacobi fields in the early 1950s can be thought of as an infinitesimal triangle comparison result which was soon followed by the celebrated Toponogov’s comparison theorem in the late 1950s.

A version of triangle comparison not involving angles can be stated as follows. Assume we have a point in a Riemannian manifold and a unit-speed geodesic in a small neighborhood of . We consider a second unit-speed geodesic in the plane and a point such that , . This way we construct a comparison triangle whose vertices , , lie at the same distance as the ones of the original triangle . We define the functions , . Then we have the following.

Theorem.

Under the conditions above

(1)

If has nonnegative sectional curvatures then .

(2)

If has nonpositive curvature then .

This result is typically proven by means of Jacobi fields, see Cheeger and Ebin (1975), but perhaps the reader could find useful the following alternative argument. Assume that the sectional curvatures of are no larger than . Take , a normal neighborhood of where the exponential map is a diffeomorphism, and a geodesic parameterized by arc length. A basic result on Jacobi fields, similar to the one used in the proof of Bishop’s volume comparison, implies that the Hessian of the function , where is the distance to , satisfies

for any , , and . In Euclidean space we would have equality in 1 replacing by the Euclidean distance to . Then the functions and satisfy

Hence on and, since the function vanishes at the endpoints of the interval, we get , which implies .

Figure 3.

Illustration of triangle comparison. The one to the right lies in a Riemannian manifold . The one to the left is a comparison triangle in the Euclidean plane. If the sectional curvature of is nonnegative then . The opposite inequality holds when has nonpositive sectional curvature.

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \begin{scope}[shift={(4.5,3)}] \draw(0,1) -- (3,0); \draw[fill=black] (1,-2) circle (0.05) node[below]{$\bar{p}$}; \draw[fill=black] (0,1) circle (0.05) node[above]{$\bar{\gamma}(0)$}; \draw[fill=black] (3,0) circle (0.05) node[above]{$\bar{\gamma}(T)$}; \draw[dashed] (0,1) -- (1,-2); \draw[dashed] (3,0) -- (1,-2); \draw[fill=black] (1.35,0.55) circle (0.05) node[above]{$\bar{\gamma}(t)$}; \draw(1,-2) -- (1.35,0.55); \end{scope} \begin{scope}[shift={(.5,3)}] \draw(0,1) to[out=5, in=135] (3,0); \draw[fill=black] (1,-2) circle (0.05) node[below]{$p$}; \draw[fill=black] (0,1) circle (0.05) node[above]{$\gamma(0)$}; \draw[fill=black] (3,0) circle (0.05) node[xshift=5,yshift=12]{$\gamma(T)$}; \draw[dashed] (0,1) to[out=-90, in=130] (1,-2); \draw[dashed] (3,0) to[out=-120, in=20] (1,-2); \draw[fill=black] (1.45,0.93) circle (0.05) node[above]{$\gamma(t)$}; \draw(1,-2) to[out=90,in=250] (1.45,0.93); \end{scope} \end{tikzpicture}

This comparison can be stated also in terms of angles. Consider a small geodesic triangle in a Riemannian manifold . If has nonnegative curvature then, at every vertex, the angle made by the sides of the triangle is larger than or equal to the comparison angle at the vertex (i.e., the Euclidean angle of a comparison triangle). For nonpositive curvature we have the opposite inequality. Similar comparisons with a simply connected surface of constant curvature hold when the sectional curvature of is no smaller or no larger than .

The triangle comparison for Riemannian manifolds allows to introduce a synthetic notion of metric space with curvature bounded, above or below, in terms of properties of geodesic triangles on the metric space.

Note that angles can be defined in a metric space using the law of cosines. Given three different points in a metric space , we can build a comparison Euclidean triangle of vertices by requiring that the distances equal , , . The comparison angle is defined as the Euclidean angle , that is,

If we have two curves emanating from the same point , the angle is

when the limit exists.

The term CAT spaces is used for metric spaces with as an upper curvature bound (CAT stands for Cartan–Alexandrov–Topogonov) and the term Alexandrov spaces for spaces with lower curvature bounds. The reader should be aware that sometimes a different terminology is used depending on the direction of the bound.

The theory of CAT spaces and the one of Alexandrov spaces differ in many respects. In the case of curvature bounded above, Hadamard manifolds, complete simply connected length spaces of nonpositive curvature, play a central role. Almost all classical results for Riemannian manifolds have been extended to Hadamard manifolds, like uniqueness of geodesics connecting two given points, the validity of triangle comparison for large triangles, the fact that the squared distance function is globally concave. Essential references on CAT spaces are Bridson and Haeffliger BH99, and Ballman, Gromov, and Schroeder BGS85. As for Alexandrov spaces, there are two main points that make the theory quite different from the one of CAT spaces. The first is that triangle comparison holds for arbitrarily large triangles without additional assumptions. This is the content of Toponogov’s theorem. The second one concerns the local structure: in an Alexandrov space the Hausdorff dimension coincides with the topological dimension, which is an integer or infinite. An Alexandrov space is a manifold except on a small set of points. Examples of Alexandrov spaces are the boundaries of convex sets in Euclidean spaces. Numerous results valid for Riemannian manifolds also hold in the class of Alexandrov spaces with curvature bounded from below by . To cite just a few, an upper bound on the diameter when , a splitting theorem, Gromov–Bishop inequalities for the volume of balls. Petrunin also proved a Levy–Gromov type isoperimetric inequality (see the author’s monograph Rit23 for the Riemannian case). The reader is referred to Burago, Burago and Ivanov BBI01, the already mentioned monograph by Alexander, Kapovitch, and Petrunin, and the references cited therein, for excellent introductions to the theory of metric spaces with bounds on curvature.

4. Convergence of Metric Spaces

The introduction of curvature bounds on metric spaces might seem an abstract construction with no special interest. However, it is related to the behavior of sequences of Riemannian manifolds. One can consider the space of all Riemannian manifolds with some topology and asks whether some given subset is compact or precompact with this topology. In many cases the limit of the sequence is not a Riemannian manifold. This is one of the situations where metric spaces with bounds on curvature play a role.

A standard way of measuring the distance between two sets in a metric space is the Hausdorff distance , defined by

where for any set , is the open tubular neighborhood of of radius .

Gromov, see Gro81Gro07 introduced a way of measuring the distance between two metric spaces , nowadays known as the Gromov–Hausdorff distance, denoted by . To compute it, we take all metric spaces which contain isometric copies of and of , compute their Hausdorff distance on , and take the infimum over all possible metric spaces .

Gromov–Hausdorff distance is really a distance on the space of isometric classes of metric spaces since two isometric metric spaces have -distance equal to . To measure the Gromov–Hausdorff distance using its definition is not very practical. However, for compact metric spaces , , should we have two -nets on , and on such that

then we would have .

Important classes of Riemannian manifolds which are precompact in the Gromov–Hausdorff distance are the following:

For and , the class of -dimensional Riemannian manifolds with volume and injectivity radius .

For any and , , the class of -dimensional Riemannian manifolds with and sectional curvature . The same result holds assuming instead that the Ricci curvature is .

While the Gromov–Hausdorff distance is useful when considering compact metric spaces, it is usually a very strong convergence for unbounded metric spaces. In this case it should be replaced by the pointed Gromov–Hausdorff convergence.

Natural measures which can be considered on metric spaces are the Hausdorff measures. Given , the -dimensional Hausdorff measure associated to a set is given, up to a constant, by

where is the supremum of taken over all coverings of by sets satisfying . For a metric space there exists a constant such that for and for all . The quantity is called the Hausdorff dimension of and is denoted by .

It is also important to note that Gromov–Hausdorff limits (see next section) of Alexandrov spaces of curvature are themselves Alexandrov spaces of curvature . There are also compactness results for Alexandrov spaces. Gromov himself proved that the space composed of the Alexandov spaces with curvature , diameter and Hausdorff (topological) dimension is compact in the Gromov–Hausdorff topology.

Although the Hausdorff measure on a metric space , with , is a natural measure to consider, sometimes it is more interesting to take a different one, see Appendix 2 in CC97. This leads to the notion of metric measure space , which is nothing but a metric space together with a Borel measure on . If is a length (geodesic) space we refer to as a length (geodesic) measure space.

A sequence of metric measure spaces converges in measured Gromov–Hausdorff topology to a metric measure space if there is a sequence of measurable maps and a sequence of real numbers converging to such that the maps are -quasi-isometries:

the tubular neighborhood of radius of is , and the push-forward measures converge in weak topology to , cf. Fukaya (1987).

5. Metric Spaces with a Lower Bound on Ricci Curvature

That the curvature of a Riemannian manifold implies metric properties of triangles was the key to introduce a synthetic notion of metric spaces with bounded curvature. An analogous approach should be possible for the Ricci curvature on a Riemannian manifold.

It looks like the first ones to discuss the possibility of such generalization were Cheeger and Colding in Appendix 2 in CC97, who suggested to use lower bounds on the volume growth of sectors. See also §5.44 in Gromov Gro07. In a Riemannian manifold , the classical Bishop–Gromov theorem implies that if is a model Riemannian manifold (complete simply connected with constant sectional curvature ) of the same dimension as and satisfying , we have

for arbitrary , where is the annulus , and the corresponding annulus in , whose volume is independent of the base point. The symbol denotes the Riemannian volumes. This inequality is also valid if we consider a geodesic sector with focal point , which is nothing but a set such that, for each , there is a geodesic segment connecting and . Let . Then we have

for arbitrary .

Figure 4.

Comparison of volume of geodesic sectors.

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture} \begin{scope}[shift={(4,0)},rotate=-30] \draw[rotate=63.2] (0,0) -- (2.65,0); \draw[rotate=116.8] (0,0) --(2.65,0); \draw[ rotate=76] (0,0) -- (2.72,0); \draw[ rotate=104] (0,0) -- (2.72,0); \begin{scope}[scale=0.8,shift={(1.5,3)}] \draw(0,0) arc [start angle=0, end angle=180,x radius=1.5,y radius=0.35]; \draw[dashed] (0,0) arc [start angle=360, end angle=180,x radius=1.5,y radius=0.35]; \end{scope} \begin{scope}[scale=0.5,shift={(1.5,3)}] \draw(0,0) arc [start angle=0, end angle=180,x radius=1.5,y radius=0.35]; \draw[dashed] (0,0) arc [start angle=360, end angle=180,x radius=1.5,y radius=0.35]; \end{scope} \begin{scope}[scale=0.3,shift={(1.5,3)}] \draw(0,0) arc [start angle=0, end angle=180,x radius=1.5,y radius=0.35]; \draw[dashed] (0,0) arc [start angle=360, end angle=180,x radius=1.5,y radius=0.35]; \end{scope} \draw(0.8,0.8) node{$r_1$}; \draw(1.1,1.3) node{$r_2$}; \draw(1.6,2.2) node{$r_3$}; \end{scope} \begin{scope}[shift={(0,0)}, rotate=-30] \draw(0,0) to[out=140,in=255] (-1.055,2.7); \draw(0,0) to[out=115, in=260] (-0.4,2.98); \begin{scope}[xscale=-1,yscale=1] \draw(0,0) to[out=140,in=255] (-1.055,2.7); \draw(0,0) to[out=115, in=260] (-0.4,2.98); \end{scope} \begin{scope}[scale=1,shift={(1.05,2.7)}] \draw(0,0) arc [start angle=0, end angle=180,x radius=1.05,y radius=0.3]; \draw[dashed] (0,0) arc [start angle=360, end angle=180,x radius=1.05,y radius=0.3]; \end{scope} \begin{scope}[scale=1,shift={(1.09,1.7)}] \draw(0,0) arc [start angle=0, end angle=180,x radius=1.09,y radius=0.3]; \draw[dashed] (0,0) arc [start angle=360, end angle=180,x radius=1.09,y radius=0.3]; \end{scope} \begin{scope}[shift={(0,0.8)},scale=0.66,shift={(1.05,0)}] \draw(0,0) arc [start angle=0, end angle=180,x radius=1.05,y radius=0.3]; \draw[dashed] (0,0) arc [start angle=360, end angle=180,x radius=1.05,y radius=0.3]; \end{scope} \draw(1,0.7) node{$r_1$}; \draw(1.4,1.6) node{$r_2$}; \draw(1.35,2.65) node{$r_3$}; \end{scope} \end{tikzpicture}

Another synthetic approach to Ricci curvature mentioned in Cheeger and Colding CC97 is the use of the Laplacian of the distance function. The classical Bishop volume comparison for balls in an -dimensional manifold with follows from comparison of the mean curvatures of geodesic spheres with the same radii in and the space form of constant sectional curvature . The mean curvature of a geodesic sphere is the Laplacian of the distance function to a fixed point. So, for instance, when , this translates into the equation

at least for small radius. Note that is the mean curvature of a ball of radius in the Euclidean space . Equation 2 was shown to hold in a weak sense in Riemannian manifolds by Calabi (1958). In order for this approach to work we need the extension of the notion of Laplacian to a metric space. Observe that for a Lipschitz function on a metric measure space we can define

and the Cheeger energy

By approximation of measurable functions by Lipschitz functions, this energy can be extended to wider classes. In general, is not a quadratic form. We say that a metric measure space is infinitesimally Hilbertian if the Cheeger energy is a quadratic form in . On such spaces it is possible to define a weak notion of Laplacian, thus giving sense to inequality 2. This extension of differential calculus to metric spaces has been carried out by Gigli Gig15. See also Ambrosio’s lecture at ICM2018 Amb18. A recent monograph by Heinonen, Koskela, Shanmugalingam, and Tyson (2015) is a very good introduction to analysis on metric spaces.

Yet another generalization of the notion of Ricci curvature comes from the well-known Bochner’s formula. Recall that Bochner’s formula in a Riemannian manifold reads

If we assume and use the estimate we arrive at Bochner’s inequality

Here and, of course, can be replaced by a larger number. All the operators appearing in this formula can be defined weakly on a metric measure space assuming some extra hypotheses, thus providing a notion of Ricci curvature bounded below.

5.1. An approach based on mass transport

In the first decade of the 21st century, Lott and Villani LV09 and Sturm Stu06aStu06b introduced independently equivalent notions of metric measure spaces with Ricci curvature bounded below. These notions were based on mass transportation properties on Riemannian manifolds with Ricci curvature bounded below.

Let us introduce first a few concepts, see Villani Vil09 for an excellent exposition. Given a metric space we consider the space of probability measures on and the subset of those satisfying

for some (all) . Given , a transport plan between and is a probability measure on with marginals (i.e., the push-forward measures , by the projections , , are and ). We denote the set of transport plans between and by Plan. The Wasserstein distance on is defined by

A minimizer of this problem is called an optimal transport plan between and . The metric space inherits many properties of . The equivalence between geodesics in Wasserstein space and certain optimal transport plans was established by Lott and Villani LV09, §2.3 and Sturm Stu06a, §2.3.

Given , we define the Shannon and Rényi entropy functionals for a measure by

and

for . Here is the decomposition of as the sum of a measure absolutely continuous with respect to and a singular measure .

We say that a function is -convex if for any Wasserstein geodesic ,

Observe that -convex just means convex. Then we say that satisfies the CD property (or that has curvature ) if, for any pair , there is a Wasserstein geodesic connecting , such that is convex. We say that satisfy the CD property for if, for any , there exists a Wasserstein geodesic connecting , such that is convex for all .

Property CD is defined the same way as CD replacing convexity by -convexity of for any Wasserstein geodesic. Property CD, for , is defined by replacing -convexity by a more involved condition in which the functions , , are replaced by functions , , see Definition 5.4 in Ambrosio Amb18.

The CD property is satisfied for -dimensional Riemannian manifolds with and is stable under measured Gromov–Hausdorff convergence.

Several refinements were made later to the theory of CD spaces. The spaces satisfy locally the CD condition. Equivalence of CD and notions for nonbranching metric measure spaces has been proved by Cavalletti and Milman CM21. A geodesic space is nonbranching if the map , taking any geodesic to the pair , is injective for all . Finally RCD and RCD spaces (the R stands for Riemannian) mean that the infinitesimally Hilbertian hypothesis is added. Recently Erbar, Kuwada, and Sturm EKS15 and Ambrosio, Mondino and Savaré AMS19 obtained the validity of Bochner’s inequality in a RCD metric measure space.

Petrunin (2011) checked the compatibility of these notions with the one of curvature bounded below by proving that -dimensional Alexandrov spaces with nonnegative curvature satisfy the curvature dimension condition CD for .

Another approach to a synthetic notion of lower bound on the Ricci curvature was introduced by Ohta (2007) in terms of the metric contraction property. This property is also equivalent, on a Riemannian manifold, to having a a lower bound on the Ricci curvature. Also Alexandrov spaces satisfy this property. Roughly speaking, the measure contraction property implies that for every set with and for every , the normalized restriction of the measure to (i.e., ) can be transported in a controlled way along geodesics to the Dirac measure .

It is worth mentioning that the notion of CD and the metric contraction property do not hold in sub-Riemannian manifolds, even in the simplest case of the Heisenberg groups, as shown by Juillet (2009). Recently Milman (2021), inspired by Barilari and Rizzi (2019), has suggested a quasi-convex relaxation QCD of the CD condition for sub-Riemannian manifolds which coincides with the latter when .

6. Metric Structures on Graphs

A graph is composed of a set of vertices and a set of edges connecting two given vertices. We assume that no more than one edge connects two given vertices and that there are no loops, that is, edges connecting the same vertex. A graph can be undirected, meaning that every edge has no orientation, and in this case we denote an edge connecting the vertices by . A graph can also be directed, meaning that every edge has an initial and a final point. In this case, it is denoted by . The vertex is the origin of the edge and the endpoint. We say that are incident to (or ). Of course and are different edges in a directed graph.

Given two vertices in a graph, a path connecting and is a sequence of vertices , so that (or ) is an edge for all . A weight function assigns to each edge a positive measure of the edge. A graph is connected if any pair of vertices can be connected by a path.

We can define a distance on the set of vertices of an undirected connected graph with an arbitrary weight function by defining

where the infimum is taken over all paths connecting and . Geodesics or shortest length paths are defined as the ones that realize the distance between its extreme points. The portion of a path connecting two vertices in a geodesic is also a geodesic.

6.1. Asymmetric distances on graphs

If is a directed graph with a weight function , we can define a “distance” between two vertices simply by setting

in case there is a path connecting and . The infimum is taken over all paths connecting and . If there is no such path we define

We also define for all . This way we have defined a map such that the following properties hold

(1)

for all ,

(2)

for all .

In the triangle inequality we have used for all . In case for all or some another alternative hypothesis is assumed we also have

1’.

if .

A map satisfying 1, , and 2 is called an asymmetric distance on . The pair is called an asymmetric metric space.

Figure 5.

Geodesics or shortest length paths between vertices and in a directed graph depicted in red. The weight function here is for any .

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=1] \node[shape=circle,draw=black] (A) at (0,0) {$1$}; \node[shape=circle,draw=black] (B) at (1,0) {$2$}; \node[shape=circle,draw=black] (C) at (2,-1) {$3$}; \node[shape=circle,draw=black] (D) at (1,-1) {$4$}; \node[shape=circle,draw=black] (E) at (1,-2) {$5$}; \node[shape=circle,draw=black] (F) at (-1,-1) {$6$}; \node[shape=circle,draw=black] (G) at (-1,-2) {$7$}; \node[shape=circle,draw=black] (H) at (0,-2) {$8$}; \draw[color=red,->] (A) edge (B); \draw[color=red,->] (B) edge (C); \draw[->] (B) edge (D); \draw[->] (E) edge (D); \draw[color=red,<->] (C) edge (E); \draw[color=red,->] (A) edge (F); \draw[color=red,->] (F) edge (G); \draw[->] (G) edge (H); \draw[->] (H) edge (E); \draw[color=red,->] (G) edge[bend right=60] (E); \end{tikzpicture}

Metric structures on graphs are commonly used in computer science. For instance, the breadth-first search algorithm, an exploratory algorithm in graphs, consist essentially on the computation of the metric balls on an undirected graph, where the weight function is for any edge . Dijkstra’s algorithm computes shortest length paths between two vertices in a directed graph with positive weights using metric projections on metric balls recursively calculated, see CLRS09.

6.2. Curvature-dimension conditions on graphs

A notion of curvature-dimension curvature on graphs has been introduced by Lin and Yau LY10. See also Ollivier (2009) for a notion of Ricci curvature on Markov chains on graphs, and Erbar and Maas (2012) for finite Markov chains. For the Lin and Yau notion, consider an undirected graph with weight function . We assume that is locally finite in the sense that the number of edges connecting every vertex to others is finite. Define

and a measure of a set by

For a function the Laplacian and squared gradient are defined by

for every . In addition, for , we define the operators

According to Lin and Yaus’s paper LY10 the graph satisfies the curvature-dimension condition CD if

and the CD condition if

The CD curvature-dimension condition is the discrete counterpart of Bochner’s inequality. Lin and Yan used these notions to prove that a locally finite graph satisfies the CD curvature condition, where is the weighted degree of the graph, given by

They also obtained a lower bound of the first nonzero eigenvalue of the Laplacian on the graph in terms of the degree and the diameter of the graph.

6.3. The Cayley graph

An interesting metric structure can be introduced on groups, allowing the use of geometric arguments to obtain results on the algebraic structure. Assume we have a group with a set of generators satisfying , that is, for every we have . We define the undirected Cayley graph associated to as , where

Hence the vertices are the elements of the group and the edges incident to are of the form . We take as weight function for any edge . If is a path connecting and then there exist such that for . Hence and

Reciprocally, if we can express as a product of elements of then there is a path connecting and . Hence the associated distance is the minimum number of generators needed to express . This is called the word metric on .

Figure 6.

Unit balls for the group with set of generators , and , respectively.

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=1] \begin{scope} \draw[fill=black] (0,0) circle (0.05); \draw[fill=black] (1,0) circle (0.05); \draw[fill=black] (-1,0) circle (0.05); \draw[fill=black] (0,1) circle (0.05); \draw[fill=black] (0,-1) circle (0.05); \draw(0,0) -- (1,0); \draw(0,0) -- (-1,0); \draw(0,0) -- (0,1); \draw(0,0) -- (0,-1); \end{scope} \begin{scope}[shift={(4,0)}] \draw[fill=black] (0,0) circle (0.05); \draw[fill=black] (1,0) circle (0.05); \draw[fill=black] (-1,0) circle (0.05); \draw[fill=black] (0,1) circle (0.05); \draw[fill=black] (0,-1) circle (0.05); \draw[fill=black] (1,1) circle (0.05); \draw[fill=black] (-1,-1) circle (0.05); \draw(0,0) -- (1,0); \draw(0,0) -- (-1,0); \draw(0,0) -- (0,1); \draw(0,0) -- (0,-1); \draw(0,0) -- (1,1); \draw(0,0) -- (-1,-1); \end{scope} \end{tikzpicture}

Since left-translations are isometries of the word metric, all metric balls of a fixed radius have the same number of elements. A group has polynomial growth if there are constants and such that

The first striking application of metric theory to group theory was the following result by Gromov

Theorem (Gro81).

If a finitely generated group has polynomial growth then it contains a nilpotent subgroup of finite index.

This result is usually considered as the starting point of Geometric group theory. See also Gromov’s Hyperbolic groups (1987) and Asymptotic invariants of infinite groups (1993) for influential works on the subject.

Final Comments

The theory of metric spaces is a lively research subject in current Mathematics. The techniques developed to face the problems in this area have had a deep, unifying and simplifying influence in related areas of Mathematics. Despite the fact we have only covered in this note only a few of the geometric aspects, there is an enormous development in progress of the theory of analysis on metric spaces. The reader is referred to Semmes’s papers (2003) in these Notices for an overview of the subject.

A version of this manuscript with full references can be found on the author’s personal webpage: http://www.ugr.es/~ritore.

References

[Amb18]
Luigi Ambrosio, Calculus, heat flow and curvature-dimension bounds in metric measure spaces, Proceedings of the International Congress of Mathematicians—Rio de Janeiro 2018. Vol. I. Plenary lectures, 2018, pp. 301–340. MR3966731,
Show rawAMSref \bib{MR3966731}{inproceedings}{ author={Ambrosio, Luigi}, title={Calculus, heat flow and curvature-dimension bounds in metric measure spaces}, date={2018}, booktitle={Proceedings of the {I}nternational {C}ongress of {M}athematicians---{R}io de {J}aneiro 2018. {V}ol. {I}. {P}lenary lectures}, publisher={World Sci. Publ., Hackensack, NJ}, pages={301\ndash 340}, review={\MR {3966731}}, }
[AMS19]
Luigi Ambrosio, Andrea Mondino, and Giuseppe Savaré, Nonlinear diffusion equations and curvature conditions in metric measure spaces, Mem. Amer. Math. Soc. 262 (2019), no. 1270, v+121. MR4044464,
Show rawAMSref \bib{MR4044464}{article}{ author={Ambrosio, Luigi}, author={Mondino, Andrea}, author={Savar\'{e}, Giuseppe}, title={Nonlinear diffusion equations and curvature conditions in metric measure spaces}, date={2019}, issn={0065-9266}, journal={Mem. Amer. Math. Soc.}, volume={262}, number={1270}, pages={v+121}, url={https://doi.org/10.1090/memo/1270}, review={\MR {4044464}}, }
[BBI01]
Dmitri Burago, Yuri Burago, and Sergei Ivanov, A course in metric geometry, Graduate Studies in Mathematics, vol. 33, American Mathematical Society, Providence, RI, 2001. MR1835418,
Show rawAMSref \bib{MR1835418}{book}{ author={Burago, Dmitri}, author={Burago, Yuri}, author={Ivanov, Sergei}, title={A course in metric geometry}, series={Graduate Studies in Mathematics}, publisher={American Mathematical Society, Providence, RI}, date={2001}, volume={33}, isbn={0-8218-2129-6}, url={https://doi.org/10.1090/gsm/033}, review={\MR {1835418}}, }
[BGS85]
Werner Ballmann, Mikhael Gromov, and Viktor Schroeder, Manifolds of nonpositive curvature, Progress in Mathematics, vol. 61, Birkhäuser Boston, Inc., Boston, MA, 1985. MR823981,
Show rawAMSref \bib{MR823981}{book}{ author={Ballmann, Werner}, author={Gromov, Mikhael}, author={Schroeder, Viktor}, title={Manifolds of nonpositive curvature}, series={Progress in Mathematics}, publisher={Birkh\"{a}user Boston, Inc., Boston, MA}, date={1985}, volume={61}, isbn={0-8176-3181-X}, url={https://doi.org/10.1007/978-1-4684-9159-3}, review={\MR {823981}}, }
[BH99]
Martin R. Bridson and André Haefliger, Metric spaces of non-positive curvature, Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 319, Springer-Verlag, Berlin, 1999. MR1744486,
Show rawAMSref \bib{MR1744486}{book}{ author={Bridson, Martin~R.}, author={Haefliger, Andr\'{e}}, title={Metric spaces of non-positive curvature}, series={Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]}, publisher={Springer-Verlag, Berlin}, date={1999}, volume={319}, isbn={3-540-64324-9}, url={https://doi.org/10.1007/978-3-662-12494-9}, review={\MR {1744486}}, }
[CC97]
Jeff Cheeger and Tobias H. Colding, On the structure of spaces with Ricci curvature bounded below. I, J. Differential Geom. 46 (1997), no. 3, 406–480. MR1484888,
Show rawAMSref \bib{MR1484888}{article}{ author={Cheeger, Jeff}, author={Colding, Tobias~H.}, title={On the structure of spaces with {R}icci curvature bounded below. {I}}, date={1997}, issn={0022-040X}, journal={J. Differential Geom.}, volume={46}, number={3}, pages={406\ndash 480}, url={http://projecteuclid.org/euclid.jdg/1214459974}, review={\MR {1484888}}, }
[CLRS09]
Thomas H. Cormen, Charles E. Leiserson, Ronald L. Rivest, and Clifford Stein, Introduction to algorithms, Third, MIT Press, Cambridge, MA, 2009. MR2572804,
Show rawAMSref \bib{MR2572804}{book}{ author={Cormen, Thomas~H.}, author={Leiserson, Charles~E.}, author={Rivest, Ronald~L.}, author={Stein, Clifford}, title={Introduction to algorithms}, edition={Third}, publisher={MIT Press, Cambridge, MA}, date={2009}, isbn={978-0-262-03384-8}, review={\MR {2572804}}, }
[CM21]
Fabio Cavalletti and Emanuel Milman, The globalization theorem for the curvature-dimension condition, Invent. Math. 226 (2021), no. 1, 1–137. MR4309491,
Show rawAMSref \bib{MR4309491}{article}{ author={Cavalletti, Fabio}, author={Milman, Emanuel}, title={The globalization theorem for the curvature-dimension condition}, date={2021}, issn={0020-9910}, journal={Invent. Math.}, volume={226}, number={1}, pages={1\ndash 137}, url={https://doi.org/10.1007/s00222-021-01040-6}, review={\MR {4309491}}, }
[EKS15]
Matthias Erbar, Kazumasa Kuwada, and Karl-Theodor Sturm, On the equivalence of the entropic curvature-dimension condition and Bochner’s inequality on metric measure spaces, Invent. Math. 201 (2015), no. 3, 993–1071. MR3385639,
Show rawAMSref \bib{MR3385639}{article}{ author={Erbar, Matthias}, author={Kuwada, Kazumasa}, author={Sturm, Karl-Theodor}, title={On the equivalence of the entropic curvature-dimension condition and {B}ochner's inequality on metric measure spaces}, date={2015}, issn={0020-9910}, journal={Invent. Math.}, volume={201}, number={3}, pages={993\ndash 1071}, url={https://doi.org/10.1007/s00222-014-0563-7}, review={\MR {3385639}}, }
[Gig15]
Nicola Gigli, On the differential structure of metric measure spaces and applications, Mem. Amer. Math. Soc. 236 (2015), no. 1113, vi+91. MR3381131,
Show rawAMSref \bib{MR3381131}{article}{ author={Gigli, Nicola}, title={On the differential structure of metric measure spaces and applications}, date={2015}, issn={0065-9266}, journal={Mem. Amer. Math. Soc.}, volume={236}, number={1113}, pages={vi+91}, url={https://doi.org/10.1090/memo/1113}, review={\MR {3381131}}, }
[Gro07]
Mikhael Gromov, Metric structures for Riemannian and non-Riemannian spaces, English, Modern Birkhäuser Classics, Birkhäuser Boston, Inc., Boston, MA, 2007. Based on the 1981 French original, With appendices by M. Katz, P. Pansu and S. Semmes, Translated from the French by Sean Michael Bates. MR2307192,
Show rawAMSref \bib{MR2307192}{book}{ author={Gromov, Mikhael}, title={Metric structures for {R}iemannian and non-{R}iemannian spaces}, edition={English}, series={Modern Birkh\"{a}user Classics}, publisher={Birkh\"{a}user Boston, Inc., Boston, MA}, date={2007}, isbn={978-0-8176-4582-3; 0-8176-4582-9}, note={Based on the 1981 French original, With appendices by M. Katz, P. Pansu and S. Semmes, Translated from the French by Sean Michael Bates}, review={\MR {2307192}}, }
[Gro81]
Mikhael Gromov, Groups of polynomial growth and expanding maps, Inst. Hautes Études Sci. Publ. Math. 53 (1981), 53–73. MR623534,
Show rawAMSref \bib{MR623534}{article}{ author={Gromov, Mikhael}, title={Groups of polynomial growth and expanding maps}, date={1981}, issn={0073-8301}, journal={Inst. Hautes \'{E}tudes Sci. Publ. Math.}, number={53}, pages={53\ndash 73}, url={http://www.numdam.org/item?id=PMIHES_1981__53__53_0}, review={\MR {623534}}, }
[Gro91]
Mikhael Gromov, Sign and geometric meaning of curvature, Rend. Sem. Mat. Fis. Milano 61 (1991), 9–123 (1994). MR1297501,
Show rawAMSref \bib{MR1297501}{article}{ author={Gromov, Mikhael}, title={Sign and geometric meaning of curvature}, date={1991}, issn={0370-7377}, journal={Rend. Sem. Mat. Fis. Milano}, volume={61}, pages={9\ndash 123 (1994)}, url={https://doi.org/10.1007/BF02925201}, review={\MR {1297501}}, }
[Gro96]
Mikhael Gromov, Carnot-Carathéodory spaces seen from within, Sub-Riemannian geometry, 1996, pp. 79–323. MR1421823,
Show rawAMSref \bib{MR1421823}{incollection}{ author={Gromov, Mikhael}, title={Carnot-{C}arath\'{e}odory spaces seen from within}, date={1996}, booktitle={Sub-{R}iemannian geometry}, series={Progr. Math.}, volume={144}, publisher={Birkh\"{a}user, Basel}, pages={79\ndash 323}, review={\MR {1421823}}, }
[Hau14]
Felix Hausdorff, Grundzüge der Mengenlehre. (German), 1914.,
Show rawAMSref \bib{zbMATH02615204}{misc}{ author={Hausdorff, Felix}, title={Grundz{\"u}ge der {Mengenlehre}.}, language={German}, how={Leipzig: {Veit} \& {Comp}. viii, 476 {S}., 53 {Figuren} (1914).}, date={1914}, }
[LV09]
John Lott and Cédric Villani, Ricci curvature for metric-measure spaces via optimal transport, Ann. of Math. (2) 169 (2009), no. 3, 903–991. MR2480619,
Show rawAMSref \bib{MR2480619}{article}{ author={Lott, John}, author={Villani, C\'{e}dric}, title={Ricci curvature for metric-measure spaces via optimal transport}, date={2009}, issn={0003-486X}, journal={Ann. of Math. (2)}, volume={169}, number={3}, pages={903\ndash 991}, url={https://doi.org/10.4007/annals.2009.169.903}, review={\MR {2480619}}, }
[LY10]
Yong Lin and Shing-Tung Yau, Ricci curvature and eigenvalue estimate on locally finite graphs, Math. Res. Lett. 17 (2010), no. 2, 343–356. MR2644381,
Show rawAMSref \bib{MR2644381}{article}{ author={Lin, Yong}, author={Yau, Shing-Tung}, title={Ricci curvature and eigenvalue estimate on locally finite graphs}, date={2010}, issn={1073-2780}, journal={Math. Res. Lett.}, volume={17}, number={2}, pages={343\ndash 356}, url={https://doi.org/10.4310/MRL.2010.v17.n2.a13}, review={\MR {2644381}}, }
[Rit23]
Manuel Ritoré, Isoperimetric inequalities in Riemannian manifolds, Progress in Mathematics, Birkhauser (to appear), 2023.,
Show rawAMSref \bib{Rit22}{book}{ author={Ritoré, Manuel}, title={Isoperimetric inequalities in {R}iemannian manifolds}, publisher={Progress in Mathematics, Birkhauser (to appear)}, date={2023}, }
[Stu06a]
Karl-Theodor Sturm, On the geometry of metric measure spaces. I, Acta Math. 196 (2006), no. 1, 65–131. MR2237206,
Show rawAMSref \bib{MR2237206}{article}{ author={Sturm, Karl-Theodor}, title={On the geometry of metric measure spaces. {I}}, date={2006}, issn={0001-5962}, journal={Acta Math.}, volume={196}, number={1}, pages={65\ndash 131}, url={https://doi.org/10.1007/s11511-006-0002-8}, review={\MR {2237206}}, }
[Stu06b]
Karl-Theodor Sturm, On the geometry of metric measure spaces. II, Acta Math. 196 (2006), no. 1, 133–177. MR2237207,
Show rawAMSref \bib{MR2237207}{article}{ author={Sturm, Karl-Theodor}, title={On the geometry of metric measure spaces. {II}}, date={2006}, issn={0001-5962}, journal={Acta Math.}, volume={196}, number={1}, pages={133\ndash 177}, url={https://doi.org/10.1007/s11511-006-0003-7}, review={\MR {2237207}}, }
[Vil09]
Cédric Villani, Optimal transport, Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 338, Springer-Verlag, Berlin, 2009. Old and new. MR2459454,
Show rawAMSref \bib{MR2459454}{book}{ author={Villani, C\'{e}dric}, title={Optimal transport}, series={Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]}, publisher={Springer-Verlag, Berlin}, date={2009}, volume={338}, isbn={978-3-540-71049-3}, url={https://doi.org/10.1007/978-3-540-71050-9}, note={Old and new}, review={\MR {2459454}}, }

Credits

Opening image is courtesy of Olivier Le Moal via Getty.

Figures 1–6 and photo of Manuel Ritoré are courtesy of Manuel Ritoré.