Hyperbolic metric space

(Redirected from Δ-hyperbolic space)

In mathematics, a hyperbolic metric space is a metric space satisfying certain metric relations (depending quantitatively on a nonnegative real number δ) between points. The definition, introduced by Mikhael Gromov, generalizes the metric properties of classical hyperbolic geometry and of trees. Hyperbolicity is a large-scale property, and is very useful to the study of certain infinite groups called Gromov-hyperbolic groups.

Definitions edit

In this paragraph we give various definitions of a  -hyperbolic space. A metric space is said to be (Gromov-) hyperbolic if it is  -hyperbolic for some  .

Definition using the Gromov product edit

Let   be a metric space. The Gromov product of two points   with respect to a third one   is defined by the formula:

 

Gromov's definition of a hyperbolic metric space is then as follows:   is  -hyperbolic if and only if all   satisfy the four-point condition

 

Note that if this condition is satisfied for all   and one fixed base point  , then it is satisfied for all with a constant  .[1] Thus the hyperbolicity condition only needs to be verified for one fixed base point; for this reason, the subscript for the base point is often dropped from the Gromov product.

Definitions using triangles edit

Up to changing   by a constant multiple, there is an equivalent geometric definition involving triangles when the metric space   is geodesic, i.e. any two points   are end points of a geodesic segment   (an isometric image of a compact subinterval   of the reals). [2][3] [4] Note that the definition via Gromov products does not require the space to be geodesic.

Let  . A geodesic triangle with vertices   is the union of three geodesic segments   (where   denotes a segment with endpoints   and  ).

 
 
 
 
 
 
The δ-slim triangle condition

If for any point   there is a point in   at distance less than   of  , and similarly for points on the other edges, and   then the triangle is said to be  -slim .

A definition of a  -hyperbolic space is then a geodesic metric space all of whose geodesic triangles are  -slim. This definition is generally credited to Eliyahu Rips.

Another definition can be given using the notion of a  -approximate center of a geodesic triangle: this is a point which is at distance at most   of any edge of the triangle (an "approximate" version of the incenter). A space is  -hyperbolic if every geodesic triangle has a  -center.

These two definitions of a  -hyperbolic space using geodesic triangles are not exactly equivalent, but there exists   such that a  -hyperbolic space in the first sense is  -hyperbolic in the second, and vice versa.[5] Thus the notion of a hyperbolic space is independent of the chosen definition.

Examples edit

 

The hyperbolic plane is hyperbolic: in fact the incircle of a geodesic triangle is the circle of largest diameter contained in the triangle and every geodesic triangle lies in the interior of an ideal triangle, all of which are isometric with incircles of diameter 2 log 3.[6] Note that in this case the Gromov product also has a simple interpretation in terms of the incircle of a geodesic triangle. In fact the quantity (A,B)C is just the hyperbolic distance p from C to either of the points of contact of the incircle with the adjacent sides: for from the diagram c = (ap) + (bp), so that p = (a + bc)/2 = (A,B)C.[7]

The Euclidean plane is not hyperbolic, for example because of the existence of homotheties.

Two "degenerate" examples of hyperbolic spaces are spaces with bounded diameter (for example finite or compact spaces) and the real line.

Metric trees and more generally real trees are the simplest interesting examples of hyperbolic spaces as they are 0-hyperbolic (i.e. all triangles are tripods).

The 1-skeleton of the triangulation by Euclidean equilateral triangles is not hyperbolic (it is in fact quasi-isometric to the Euclidean plane). A triangulation of the plane   has a hyperbolic 1-skeleton if every vertex has degree 7 or more.

The two-dimensional grid is not hyperbolic (it is quasi-isometric to the Euclidean plane). It is the Cayley graph of the fundamental group of the torus; the Cayley graphs of the fundamental groups of a surface of higher genus is hyperbolic (it is in fact quasi-isometric to the hyperbolic plane).

Hyperbolicity and curvature edit

The hyperbolic plane (and more generally any Hadamard manifolds of sectional curvature  ) is  -hyperbolic. If we scale the Riemannian metric by a factor   then the distances are multiplied by   and thus we get a space that is  -hyperbolic. Since the curvature is multiplied by   we see that in this example the more (negatively) curved the space is, the lower the hyperbolicity constant.

Similar examples are CAT spaces of negative curvature. With respect to curvature and hyperbolicity it should be noted however that while curvature is a property that is essentially local, hyperbolicity is a large-scale property which does not see local (i.e. happening in a bounded region) metric phenomena. For example, the union of an hyperbolic space with a compact space with any metric extending the original ones remains hyperbolic.

Important properties edit

Invariance under quasi-isometry edit

One way to make precise the meaning of "large scale" is to require invariance under quasi-isometry. This is true of hyperbolicity.

If a geodesic metric space   is quasi-isometric to a  -hyperbolic space   then there exists   such that   is  -hyperbolic.

The constant   depends on   and on the multiplicative and additive constants for the quasi-isometry.[8]

Approximate trees in hyperbolic spaces edit

The definition of an hyperbolic space in terms of the Gromov product can be seen as saying that the metric relations between any four points are the same as they would be in a tree, up to the additive constant  . More generally the following property shows that any finite subset of an hyperbolic space looks like a finite tree.

For any   there is a constant   such that the following holds: if   are points in a  -hyperbolic space   there is a finite tree   and an embedding   such that   for all   and
 

The constant   can be taken to be   with   and this is optimal.[9]

Exponential growth of distance and isoperimetric inequalities edit

In an hyperbolic space   we have the following property:[10]

There are   such that for all   with  , every path   joining   to   and staying at distance at least   of   has length at least  .

Informally this means that the circumference of a "circle" of radius   grows exponentially with  . This is reminiscent of the isoperimetric problem in the Euclidean plane. Here is a more specific statement to this effect.[11]

Suppose that   is a cell complex of dimension 2 such that its 1-skeleton is hyperbolic, and there exists   such that the boundary of any 2-cell contains at most   1-cells. Then there is a constant   such that for any finite subcomplex   we have
 

Here the area of a 2-complex is the number of 2-cells and the length of a 1-complex is the number of 1-cells. The statement above is a linear isoperimetric inequality; it turns out that having such an isoperimetric inequality characterises Gromov-hyperbolic spaces.[12] Linear isoperimetric inequalities were inspired by the small cancellation conditions from combinatorial group theory.

Quasiconvex subspaces edit

A subspace   of a geodesic metric space   is said to be quasiconvex if there is a constant   such that any geodesic in   between two points of   stays within distance   of  .

A quasi-convex subspace of an hyperbolic space is hyperbolic.

Asymptotic cones edit

All asymptotic cones of an hyperbolic space are real trees. This property characterises hyperbolic spaces.[13]

The boundary of a hyperbolic space edit

Generalising the construction of the ends of a simplicial tree there is a natural notion of boundary at infinity for hyperbolic spaces, which has proven very useful for analysing group actions.

In this paragraph   is a geodesic metric space which is hyperbolic.

Definition using the Gromov product edit

A sequence   is said to converge to infinity if for some (or any) point   we have that   as both   and   go to infinity. Two sequences   converging to infinity are considered equivalent when   (for some or any  ). The boundary of   is the set of equivalence classes of sequences which converge to infinity,[14] which is denoted  .

If   are two points on the boundary then their Gromov product is defined to be:

 

which is finite iff  . One can then define a topology on   using the functions  .[15] This topology on   is metrisable and there is a distinguished family of metrics defined using the Gromov product.[16]

Definition for proper spaces using rays edit

Let   be two quasi-isometric embeddings of   into   ("quasi-geodesic rays"). They are considered equivalent if and only if the function   is bounded on  . If the space   is proper then the set of all such embeddings modulo equivalence with its natural topology is homeomorphic to   as defined above.[17]

A similar realisation is to fix a basepoint and consider only quasi-geodesic rays originating from this point. In case   is geodesic and proper one can also restrict to genuine geodesic rays.

Examples edit

When   is a simplicial regular tree the boundary is just the space of ends, which is a Cantor set. Fixing a point   yields a natural distance on  : two points represented by rays   originating at   are at distance  .

When   is the unit disk, i.e. the Poincaré disk model for the hyperbolic plane, the hyperbolic metric on the disk is

 

and the Gromov boundary can be identified with the unit circle.

The boundary of  -dimensional hyperbolic space is homeomorphic to the  -dimensional sphere and the metrics are similar to the one above.

Busemann functions edit

If   is proper then its boundary is homeomorphic to the space of Busemann functions on   modulo translations.[18]

The action of isometries on the boundary and their classification edit

A quasi-isometry between two hyperbolic spaces   induces a homeomorphism between the boundaries.

In particular the group of isometries of   acts by homeomorphisms on  . This action can be used[19] to classify isometries according to their dynamical behaviour on the boundary, generalising that for trees and classical hyperbolic spaces. Let   be an isometry of  , then one of the following cases occur:

  • First case:   has a bounded orbit on   (in case   is proper this implies that   has a fixed point in  ). Then it is called an elliptic isometry.
  • Second case:   has exactly two fixed points   on   and every positive orbit   accumulates only at  . Then   is called an hyperbolic isometry.
  • Third case:   has exactly one fixed point on the boundary and all orbits accumulate at this point. Then it is called a parabolic isometry.

More examples edit

Subsets of the theory of hyperbolic groups can be used to give more examples of hyperbolic spaces, for instance the Cayley graph of a small cancellation group. It is also known that the Cayley graphs of certain models of random groups (which is in effect a randomly-generated infinite regular graph) tend to be hyperbolic very often.

It can be difficult and interesting to prove that certain spaces are hyperbolic. For example, the following hyperbolicity results have led to new phenomena being discovered for the groups acting on them.

  • The hyperbolicity of the curve complex[20] has led to new results on the mapping class group.[21]
  • Similarly, the hyperbolicity of certain graphs[22] associated to the outer automorphism group Out(Fn) has led to new results on this group.

See also edit

Notes edit

  1. ^ Coornaert, Delzant & Papadopoulos 1990, pp. 2–3
  2. ^ de la Harpe & Ghys 1990, Chapitre 2, Proposition 21.
  3. ^ Bridson & Haefliger 1999, Chapter III.H, Proposition 1.22.
  4. ^ Coornaert, Delzant & Papadopoulos 1990, pp. 6–8.
  5. ^ Bridson & Haefliger 1999, Chapter III.H, Proposition 1.17.
  6. ^ Coornaert, Delzant & Papadopoulos 1990, pp. 11–12
  7. ^ Coornaert, Delzant & Papadopoulos 1990, p. 1–2s
  8. ^ de la Harpe & Ghys 1990, Chapitre 5, Proposition 15.
  9. ^ Bowditch 2006, Chapter 6.4.
  10. ^ Bridson & Haefliger 1999, Chapter III.H, Proposition 1.25.
  11. ^ a more general statement is given in Bridson & Haefliger (1999, Chapter III.H, Proposition 2.7)
  12. ^ Bridson & Haefliger 1999, Chapter III.H, Theorem 2.9.
  13. ^ Dyubina (Erschler), Anna; Polterovich, Iosif (2001). "Explicit constructions of universal R-trees and asymptotic geometry of hyperbolic spaces". Bull. London Math. Soc. Vol. 33. pp. 727–734. MR 1853785.
  14. ^ de la Harpe & Ghys 1990, Chapitre 7, page 120.
  15. ^ de la Harpe & Ghys 1990, Chapitre 7, section 2.
  16. ^ de la Harpe & Ghys 1990, Chapitre 7, section 3.
  17. ^ de la Harpe & Ghys 1990, Chapitre 7, Proposition 4.
  18. ^ Bridson & Haefliger 1999, p. 428.
  19. ^ de la Harpe & Ghys 1990, Chapitre 8.
  20. ^ Masur, Howard A.; Minsky, Yair N. (1999). "Geometry of the complex of curves. I. Hyperbolicity". Invent. Math. Vol. 138. pp. 103–149. doi:10.1007/s002220050343. MR 1714338.
  21. ^ Dahmani, François; Guirardel, Vincent; Osin, Denis (2017). "Hyperbolically embedded subgroups and rotating families in groups acting on hyperbolic spaces". Memoirs of the American Mathematical Society. 245 (1156). arXiv:1111.7048. doi:10.1090/memo/1156.
  22. ^ Bestvina, Mladen; Feighn, Mark (2014). "Hyperbolicity of the complex of free factors". Advances in Mathematics. 256: 104–155. arXiv:1107.3308. doi:10.1016/j.aim.2014.02.001. MR 3177291.

References edit

  • Bowditch, Brian (2006), A course on geometric group theory (PDF), Mat. soc. Japan
  • Bridson, Martin R.; Haefliger, André (1999), Metric spaces of non-positive curvature, Springer
  • Coornaert, M.; Delzant, T.; Papadopoulos, A. (1990), Géométrie et théorie des groupes. Les groupes hyperboliques de Gromov, Lecture Notes in Mathematics (in French), vol. 1441, Springer-Verlag, ISBN 3-540-52977-2
  • de la Harpe, Pierre; Ghys, Etienne (1990), Sur les groupes hyperboliques d'après Mikhael Gromov (in French), Birkhäuser
  • Gromov, Mikhael (1987), "Hyperbolic groups", in Gersten, S.M. (ed.), Essays in group theory, Springer, pp. 75–264
  • Roe, John (2003), Lectures on Coarse Geometry, University Lecture Series, vol. 31, American Mathematical Society, ISBN 978-0-8218-3332-2
  • Väisälä, Jussi (2005), "Gromov hyperbolic spaces", Expositiones Mathematicae, 23 (3): 187–231, doi:10.1016/j.exmath.2005.01.010, MR 2164775.