Uniformizing Gromov hyperbolic spaces with Busemann functions
Abstract.
Given a complete Gromov hyperbolic space that is roughly starlike from a point in its Gromov boundary , we use a Busemann function based at to construct an incomplete unbounded uniform metric space whose boundary can be canonically identified with the Gromov boundary of relative to . This uniformization construction generalizes the procedure used to obtain the Euclidean upper half plane from the hyperbolic plane. Furthermore we show, for an arbitrary metric space , that there is a hyperbolic filling of that can be uniformized in such a way that the boundary has a biLipschitz identification with the completion of . We also prove that this uniformization procedure can be done at an exponent that is often optimal in the case of CAT spaces.
1. Introduction
The goal of this paper is to construct an unbounded analogue of the uniformizations of Gromov hyperbolic spaces built by Bonk, Heinonen and Koskela in their extensive study of a number of problems in conformal analysis [4]. The most familiar special case of our procedure is the construction of the upper half-space in from the hyperbolic plane , which is discussed in Example 1.5. The guiding example in [4], by comparison, is the relationship between and the Euclidean unit disk . As can be seen from these examples, the input for uniformization is a geodesic Gromov hyperbolic space and the output is an incomplete metric space , obtained from a conformal deformation of , that is uniform in the sense of Definition 1.1 below. The density used for uniformizing a Gromov hyperbolic space in [4] is exponential in the distance to a fixed point of . In contrast we will be using a density that is exponential in a Busemann function associated to a particular point of the Gromov boundary of . This choice of density is natural as Busemann functions are often interpreted as distance functions “from infinity” and can themselves be used to define a boundary of the space [1, 3]. Unlike in [4], we will not assume any local compactness properties on , so specializing our results back to their setting yields a small generalization of their results as well.
Our principal application of this uniformization construction will be to hyperbolic fillings of a metric space , with a particular focus on the case in which is unbounded. When is bounded a hyperbolic filling of can be thought of as a Gromov hyperbolic graph whose Gromov boundary can be canonically identified with ; in the case that is unbounded there are some additional subtleties to this notion owing to the fact that the Gromov boundary of a Gromov hyperbolic space is always bounded. We refer to the discussion prior to Theorem 1.12 for further information on this, as well as the contents of Section 5. Our use of Busemann functions in this setting is inspired by the hyperbolic filling construction of Buyalo and Schroeder [12, Chapter 6] for arbitrary metric spaces .
Our uniformization construction for hyperbolic fillings is used in a followup work [11] in order to establish a correspondence between Newton-Sobolev classes of functions on the hyperbolic filling of and Besov classes of functions on in the special case that carries a doubling measure. This is heavily inspired by work of A. Björn, J. Björn, and Shanmugalingam [3] that establishes the corresponding result in the case that is bounded. In a closely related work [10] we also generalize to our setting their results [2] on how local Poincaré inequalities transform under the uniformization in [4]. This yields some interesting new examples of uniform metric spaces satisfying Poincaré inequalities. There are a number of known variants on the correspondence between function spaces on the hyperbolic filling and function spaces on , see for instance [6], [7], [9]. Such correspondences were one of the original motivating factors in the use of hyperbolic fillings in analysis on metric spaces. For applications to trace theorems on Ahlfors regular metric spaces that demonstrate the power of these correspondences we refer to [19].
Lastly we remark that the idea of uniformizing Gromov hyperbolic spaces using Busemann functions has been developed independently by Zhou [22] for the purpose of an entirely different set of applications, including a study of Teichmueller’s displacement problem for quasi-isometries of Gromov hyperbolic spaces. The work [22] in particular gives alternative proofs of the main uniformization theorems (Theorems 1.4 and 1.6) restricted to the case of proper Gromov hyperbolic spaces and the original range of exponents considered by Bonk-Heinonen-Koskela (see Theorem 1.9 below). Our applications to CAT spaces and hyperbolic fillings require us to consider exponents outside this range however; this is a key point of departure from [22].
Stating our main theorems require some preliminary definitions. We opt to give precise definitions in the corresponding sections throughout the paper, while mostly only outlining the necessary definitions here in the introduction. For a metric space and a curve , a subinterval, we write for the length of measured in . We will follow the standard practice of using to denote both the parametrization of the curve and the image of the curve in . The curve is a geodesic if it is isometric as a mapping of into . We say that is geodesic if any two points can be joined by a geodesic. We will use the following distance notation for distance from a point to a set in any metric space ,
and in particular will write for the distance of a point to (the image of) a curve .
We now define uniform metric spaces. We start with an incomplete metric space . We denote the boundary of in its completion by . For we write for the distance from to the boundary . For the definition below we allow to be any closed interval, and for a curve we denote its endpoints by and . For such an interval we write and .
Definition 1.1.
For a constant and a closed interval , a curve is -uniform if
(1.1) |
and if for every we have
(1.2) |
We say that the metric space is -uniform if any two points in can be joined by an -uniform curve.
Many reasonable domains in Euclidean space such as the unit ball or upper half-space provide natural examples of uniform metric spaces when they are equipped with the Euclidean metric. The first requirement (1.1) implies that -uniform curves minimize the distance between their endpoints up to the multiplicative constant . The second requirement (1.2) implies that if we cut at any point then at least one of the two subcurves or must have length controlled by the distance of to . We note that it is easily verified from the definitions that the property of a curve being -uniform is independent of the choice of parametrization of . For the purpose of formulating our theorems it is convenient to extend the definition of -uniform curves to allow for arbitrary subintervals and to allow the possibility ; as this extension is somewhat technical we refer to Definition 4.16 for the exact details.
Remark 1.2.
For a continuous function we write
for the line integral of along . We refer to [4, Appendix] for a detailed discussion of line integrals in our context. We will often refer to such a positive continuous function as a density on . The following definition plays a key role in the statement of our main theorems.
Definition 1.3.
Let be a geodesic metric space and let be a density on . The conformal deformation of with conformal factor is the metric space with metric
with the infimum taken over all curves joining to . We say that the density is a Gehring-Hayman density (abbreviated as a GH-density) if there is a constant such that for any and any geodesic joining to we have
(1.3) |
We will refer to the inequality (1.3) as the GH-inequality and will sometimes refer to the constant as the -constant. The terminology here is inspired by the work of Gehring-Hayman [16], which shows that in a simply connected hyperbolic domain in the complex plane the hyperbolic geodesics minimize Euclidean length among all curves in the domain with the same end points, up to a universal multiplicative constant. Here the density is given by the conformal change of metric relating the Euclidean metric on to the hyperbolic metric. Note that if is a tree then the GH-inequality holds for any density with since any path joining two points in a tree must contain the geodesic joining those points.
We next discuss the notions we will need regarding Gromov hyperbolic spaces. Most formal definitions regarding Gromov hyperbolicity and the Gromov boundary are postponed to Section 2, as they can be found in any standard reference such as [12], [17]. A geodesic metric space is Gromov hyperbolic if there is a such that all geodesic triangles are -thin, meaning that for any geodesic triangle each edge of is contained in a -neighborhood of the other two edges of . In this case we will also say that is -hyperbolic. We write for the Gromov boundary of , to be defined in Section 2; for now we note that a geodesic ray can always be identified with an equivalence class , but in general not every point in can be realized in this way.
We consider a complete geodesic -hyperbolic space and a geodesic ray . The Busemann function associated to is defined by the limit
(1.4) |
Using the triangle inequality and the fact that , it’s easy to check that the right side is nonincreasing in and bounded below by , so this limit exists. It’s also easily verified that is 1-Lipschitz, thus in particular is continuous. As is customary when considering Busemann functions, we will refer to any translate of for a constant as a Busemann function as well. We write
(1.5) |
for the set of all Busemann functions on . Given a Busemann function we will write for the point in the Gromov boundary determined by the geodesic ray . We will refer to as the basepoint of and say that is based at .
To state our theorems in their appropriate generality it is useful to augment the set with the distance functions on : for we write for the distance function to and write
(1.6) |
for the set of all translates of distance functions on . We then write . In the case we write and refer to as the basepoint of as well. The defining formula (1.4) for Busemann functions shows that any can be realized as a pointwise limit of functions defined by .
Given and we define a density on by
We write for the conformal deformation of with conformal factor . In the theorem below we will be assuming that is -roughly starlike from the basepoint of . This is a technical condition on geodesics starting from that is described in Definition 2.3. The main purpose of this hypothesis is to rule out cases such as trees that have arbitrarily long finite branches. This -rough starlikeness condition will be satisfied with in our application of Theorem 1.4 to hyperbolic fillings in Theorem 1.12.
Theorem 1.4.
Let be a complete geodesic -hyperbolic space and let be given. We suppose that is -roughly starlike from and that is given such that is a GH-density with constant .
Then geodesics in are -uniform curves in , with . Consequently is an -uniform metric space. Furthermore is bounded if and only if .
In this statement and all subsequent ones the notation is used to indicate that a particular constant depends on the indicated parameters. We refer to Definition 4.16 for the extension of the definition of -uniform curves that is necessary to cover the case of an arbitrary geodesic in ; Definition 1.1 only covers the case of geodesics defined on closed intervals. The claim that is bounded if and only if does not require either the rough starlikeness hypothesis or the assumption that is a GH-density, see Proposition 4.4.
We describe the motivating example for Theorem 1.4 in the case below.
Example 1.5.
Let be the upper half space in equipped with the Euclidean metric, which is easily seen to be a uniform metric space. Let denote the upper half plane model of the hyperbolic plane, which is equipped with the Riemannian metric . Define by . Then is a geodesic ray in .
From explicit formulas for the hyperbolic distance in this model (see for instance [12, A.3]) it is straightforward to calculate that the associated Busemann function is given by . Setting , the density is thus simply given by . Therefore the uniformized metric space is isometric to . We also remark that the GH-inequality (1.3) for the density can easily be verified using the standard representation of geodesics in the upper half-plane model for as subsegments of semicircles or vertical lines orthogonal to the horizontal line in .
A metric space is proper if its closed balls are compact. In the case , Theorem 1.4 generalizes [4, Proposition 4.5] as it does not require to be proper and allows for a potentially larger range of values for the parameter ; this larger range will be relevant to Theorem 1.10.
For a -hyperbolic space and a point we write for the complement of in the Gromov boundary of . We will refer to as the Gromov boundary relative to for reasons that will be explained prior to Proposition 2.8. We formally extend this definition to by defining ; in this case we will still refer to as the Gromov boundary relative to , with the understanding that this simply coincides with the standard Gromov boundary for . As part of the proof of Theorem 1.4, we will show that there is a canonical identification between the Gromov boundary of relative to and the boundary of in its completion. The most important property of this identification is summarized in Theorem 1.6 below.
A function can be used to define a natural class of metrics on known as visual metrics based at (see [12, Chapter 3] as well as Section 2.3). These visual metrics have an associated parameter and a comparison constant to a specific model quasi-metric on defined in (2.13). We continue to write for the canonical extension of the metric on the uniformization to its completion .
Theorem 1.6.
Let be a complete geodesic -hyperbolic space and let be such that is -roughly starlike from the basepoint of . Let be given such that is a GH-density with constant . Then there is a canonical identification under which the restriction of to defines a visual metric on based at with parameter and comparison constant .
For a precise description of the identification we refer to the disucssion after (4.15).
Remark 1.7.
It is useful to allow some additional flexibility in the choice of function in Theorems 1.4 and 1.6. This flexibility will be used in Theorem 1.12. For a continuous function and a constant we write if there is some such that for all ; we write if and if . We define the basepoint of to be the basepoint of , ; while this definition may be ambiguous in the case , this ambiguity does not matter in the context of our theorems. Then is -biLipschitz to via the identity map on and will be a GH-density with constant if is a GH-density with constant . It then easily follows that Theorems 1.4 and 1.6 hold for as well, with the uniformity parameter in Theorem 1.4 and the comparison constant in Theorem 1.6 depending additionally on .
Remark 1.8.
Since we do not assume that the Gromov hyperbolic space in our theorems is proper, it is an interesting question whether our theorems can be applied to the “free quasiworld” considered by Väisälä [20]. The Gromov hyperbolic spaces that arise in this context are domains in Banach spaces equipped with hyperbolic metrics. However in this setting the hypothesis that is geodesic is too strong [20, Remark 3.5]. The best one can assume is that is a length space, i.e., that the distance between two points of is equal to the infimum of the lengths of all curves joining them. Thus one would need to generalize Theorems 1.4 and 1.6 to Gromov hyperbolic spaces that are not necessarily geodesic, but are still length spaces. We believe that such a generalization is possible, but since it is unnecessary for our applications we will not pursue it here.
Let’s now discuss when the hypotheses of Theorems 1.4 and 1.6 are satisfied in practice. The two key hypotheses are the rough starlikeness hypothesis from the basepoint of and the assumption that is a GH-density on . The rough starlikeness hypothesis is always easily verified in applications of interest, so as a consequence it is typically not a concern when trying to apply these theorems. Thus the main hypothesis to verify is that of being a GH-density. The most general result available regarding verifying this condition is the following theorem of Bonk-Heinonen-Koskela.
Theorem 1.9.
[4, Theorem 5.1] Let be a geodesic -hyperbolic space. There is depending only on such that if a density satisfies for all and some fixed ,
(1.7) |
then is a GH-density with constant .
The inequality (1.7) is satisfied for for any since all functions in are -Lipschitz. Theorem 1.9 builds on a number of previous works that are summarized at the beginning of [4, Chapter 5]. Thus if one is not concerned about obtaining Theorems 1.4 and 1.6 for a specific value of then it is always possible to assume that is a GH-density with constant by taking sufficiently small.
In general one wants to establish that is a GH-density for as large a value of as possible, as this property is then inherited for smaller values of by Proposition 4.12. This is particularly important for applications in which there is a preferred visual metric on , such as Theorems 1.10 and 1.12 below.
CAT spaces are geodesic metric spaces in which geodesic triangles are thinner than corresponding comparison geodesic triangles in the hyperbolic plane . We refer to [14, Definition 3.2.1] for a precise definition; since the proper definition is somewhat lengthy to state and we will only be using easily stated consequences of the CAT property, we omit a full description of the definition here. These spaces encompass many natural examples such as trees and simply connected Riemannian manifolds with sectional curvatures . A CAT space is -hyperbolic with the same hyperbolicity constant as the hyperbolic plane, for which the optimal constant can be computed explicitly to be .
For a CAT space and a function with basepoint the model quasi-metric on in fact defines a distinguished choice of visual metric on with parameter . This metric is known as the Bourdon metric on when and the Hamenstädt metric on when . For further details we refer to Remark 2.9. Our next theorem applies Theorems 1.4 and 1.6 to the special case of CAT spaces at the special value .
Theorem 1.10.
Let be a complete CAT space and let be given with basepoint . Then there is a universal constant such that is a GH-density with constant . If, furthermore, is -roughly starlike from the basepoint of for some then the conclusions of Theorems 1.4 and 1.6 hold for with constants and depending only on . In particular the restriction of to is -biLipschitz to .
The constant in Theorem 1.10 is universal in the sense that it is the same for any CAT space and any . The dependence of on can be removed when by mimicking the arguments of [4, Proposition 4.5]; this same comment applies to Theorem 1.4 as well. The conclusions of Theorem 1.10 show in particular that the boundary of the uniformization has a canonical biLipschitz identification with the Gromov boundary relative to equipped with the distinguished visual metric .
Remark 1.11.
Theorem 1.6 produces an obstruction for to be a GH-density: the Gromov boundary must admit a visual metric based at the basepoint of with parameter . In the case for some (i.e., if is a distance function) then this shows in particular that , where is the asymptotic upper curvature bound defined by Bonk and Foertsch (see [5, Theorem 1.5]). In the case of the -dimensional hyperbolic space of constant negative curvature () we have by [5, Proposition 1.4]. Hence cannot be a GH-density for any when . This shows in particular that the value for to be a GH-density in Theorem 1.10 is sharp in certain cases. We can in fact extend these conclusions to observe that cannot be a GH-density for any when as well by observing that, for a fixed , any visual metric based at on with parameter would give rise to a visual metric on with parameter by a Möbius inversion of centered at the point .
We will also apply our uniformization results to hyperbolic fillings of an arbitrary metric space . We briefly describe the construction of the hyperbolic filling here, with further details in Section 5, including proofs for the claims made here. Our construction will depend in part on two parameters and . For an we say that a subset is -separated if for each we have . Given a parameter , we choose for each a maximal -separated subset of . For we write and set . The set will serve as the vertex set for . We define the height function on this vertex set by for .
We associate to each vertex the ball of radius centered at . We place an edge between vertices if and only if their heights satisfy and their associated balls satisfy . We write for the resulting graph and call this a hyperbolic filling of . If is sufficiently large (see inequality (5.1)) then will be a connected graph by Proposition 5.5. We make into a geodesic metric space by declaring all edges to have unit length. We extend the height function to a -Lipschitz function by linearly interpolating the values of from the vertices to the edges of .
As a metric space is -hyperbolic with depending only on the parameters and . There is a distinguished point in the Gromov boundary that can be thought of as an ideal point at infinity for . We have an identification of the Gromov boundary relative to with the completion of . Under this identification the extension of the metric to defines a visual metric on with parameter . All of the results mentioned here are proved in Section 5.
We define a density on by for . We write for the conformal deformation of with conformal factor . By Lemma 5.12 there is a Busemann function based at such that for all and therefore in the notation of Remark 1.7. For such a Busemann function we have that the density is uniformly comparable to the density with .
Theorem 1.12.
Let be a metric space and let be a hyperbolic filling of with parameters and . Then is -roughly starlike from and is a GH-density with constant .
We compare our results to those of [3] in Remark 6.4. Another notable predecessor to Theorem 1.12 in the case that is compact is the work of Piaggio [13, Section 2].
We provide here an outline of the contents of the rest of the paper. In Section 2 we review several key notions in the setting of Gromov hyperbolic spaces. Section 3 establishes some basic properties of geodesic triangles in Gromov hyperbolic spaces with vertices on the Gromov boundary and gives a rough formula for evaluating certain distance functions and Busemann functions on their edges. We then use these results in Section 4 to obtain estimates for the uniformized distance and prove Theorems 1.4, 1.6, and 1.10. In Section 5 we construct the hyperbolic fillings of metric spaces that we use in Theorem 1.12 and establish their basic properties. Lastly Theorem 1.12 is proved in Section 6.
2. Hyperbolic metric spaces
2.1. Definitions
Let be a set and let , be real-valued functions defined on . For we will write if
for all . If the exact value of the constant is not important or implied by context we will often just write . We will sometimes refer to the relation as a rough equality between and .
If and and both take values in , we will write if
We will similarly write if the value of is not important or implied by context. Note that if then , and similarly if then . We will stick to a convention of using lowercase for additive constants and uppercase for multiplicative constants. When this additive constant is determined by other parameters , , etc. under discussion we will write , while continuing to use the shorthand where it is not ambiguous (and the same for multiplicative constants ).
For a metric space we write for the open ball of radius centered at . A map between metric spaces and is isometric if for , . If furthermore is surjective then we say that it is an isometry and that and are isometric. For a constant a map is defined to be -roughly isometric if . The map is -Lipschitz for a constant if , and it is -biLipschitz for a constant if . As usual we will not mention the exact value of the constants if they are unimportant.
When dealing with Gromov hyperbolic spaces in this paper we will use the generic distance notation for the distance between and in , except for cases where this could cause confusion. We will often use the generic notation for a geodesic connecting two points , even when this geodesic is not unique; in these cases there will be no ambiguity regarding the geodesic that we are referring to. A geodesic triangle in is a collection of three points together with geodesics , , and joining these points, which we will refer to as the edges of . We will also alternatively write for a geodesic triangle with vertices , and .
For the Gromov product of and based at is defined by
(2.1) |
We note the basepoint change inequality for ,
(2.2) |
which follows from the triangle inequality.
By [17, Chapitre 2, Proposition 21] we have two key consequences of -hyperbolicity for a metric space regarding Gromov products. The first is that for every we have
(2.3) |
We refer to (2.3) as the -inequality.
The second is that for any geodesic triangle in we have that if , are points with then . Here and are referring to the corresponding geodesics in the triangle . We will refer to this as the -tripod condition.
Both inequality (2.3) and the tripod condition can be taken as equivalent definitions of hyperbolicity. By [17, Chapitre 2, Proposition 21] all of these definitions are equivalent up to a factor of . We note that the definition using inequality (2.3) does not use the fact that is geodesic, and is therefore used as a definition of -hyperbolicity for general metric spaces. We will be citing several basic results from [12] in which inequality (2.3) is used as the definition of -hyperbolicity (with in place of ). Wherever necessary we have multiplied the constants used in their results by in order to account for this discrepancy.
Let be a geodesic Gromov hyperbolic space and fix . A sequence converges to infinity if we have as . The Gromov boundary of a Gromov hyperbolic space is defined to be the set of all equivalence classes of sequences converging to infinity, with the equivalence relation if as . Inequality (2.2) shows that these notions do not depend on the choice of basepoint .
A second boundary that we can associate to is the geodesic boundary , which is defined as equivalence classes of geodesic rays , with two geodesic rays and being equivalent if there is a constant such that for . There is a natural inclusion given by sending a geodesic ray to the sequence . This inclusion need not be surjective in general. However, it is always surjective if is proper, meaning that closed balls in are compact.
For a point and a sequence converging to infinity we will write or if belongs to the equivalence class of . For a geodesic ray , , and a point we will write if , . We will sometimes also consider geodesic rays with a reversely oriented parametrization, for which we write if .
For the rest of this paper we will be using the standard notation for the Gromov boundary of a Gromov hyperbolic space . While this notation does conflict with the notation introduced prior to Definition 1.1, the meaning of the notation will always be clear from context since we will never use it in the sense of Definition 1.1 in the context of Gromov hyperbolic spaces.
We now extend some notions regarding geodesic triangles to the Gromov boundary. For a point and a point we will often write for a geodesic ray with and , provided such a geodesic ray exists. Similarly, for we will write for a geodesic line with and , provided such a geodesic line exists. Such geodesic lines and rays always exist when is proper, but not necessarily in general. We extend the definition of geodesic triangles in to allow for vertices in : a geodesic triangle in is a collection of three points together with geodesics , , connecting them in the sense described above.
Remark 2.1.
It is easy to see from the definitions that there is no geodesic such that and belong to the same equivalence class in the Gromov boundary . Hence, for a geodesic triangle , all vertices of on must be distinct.
Gromov products based at points can be extended to points of by defining the Gromov product of equivalence classes , based at to be
with the infimum taken over all sequences , . By [12, Lemma 2.2.2], if is -hyperbolic then for any choices of sequences , we have
(2.4) |
We also have the -inequality for , , and ,
(2.5) |
For , the Gromov product is defined analogously as
with the infimum taken over , and the analogues of (2.4) and (2.5) hold as well.
We next observe that geodesic triangles with vertices in are -thin, in the precise sense that if is any given point then there is a point satisfying that does not belong to the same edge of as . When is proper this can be easily deduced from the -thin triangles property for triangles in by a limiting argument. Without the properness hypothesis this result can also be obtained with a larger thinness constant as a consequence of work of Väisälä [21, Theorem 6.24]; we note that he uses (2.3) as the definition of -hyperbolicity so we have to multiply the constant he obtains by . As Väisälä works in the more general context of Gromov hyperbolic spaces that are not necessarily geodesic (which greatly complicates the proofs), we prefer to give a simpler direct proof of -thinness here.
Lemma 2.2.
Let be a geodesic triangle in with vertices in . Then is -thin.
Proof.
Let be the vertices of . Let be given. Since has -thin triangles, we may assume that has at least one vertex on . We first consider the case in which has exactly one vertex on , which by relabeling we can assume is . We first assume that . Let and be sequences such that and . For each we let be the geodesic triangle sharing the edge with , having a second edge be the subsegment of , and having a third edge be any choice of geodesic . Then is -thin, so we have for each that either or (or both). In the first case we are done since , so we can assume that . Let be such that . Then .
Since both and converge to , for sufficiently large we will have , which implies in particular that . The -tripod condition applied to , , and then implies that if is the unique point such that then , from which it follows that for all sufficiently large . Since this completes the proof of this case.
The other cases are and . By symmetry it suffices to prove the case . We define the sequences and and the triangle as before. As in the case we can assume that for all , as otherwise by the -thin triangles property we have and we are done. We let be such that , note that as before, and choose large enough that . As before the -tripod condition then supplies a point such that and we conclude that .
We can now handle the case in which potentially two or three vertices of belong to . By symmetry it suffices to show for a point that is -close to either or . Let and be sequences such that and ; if then we set for all and similarly if then we set for all . Let be a geodesic triangle with one edge the subsegment of . Then has at most one vertex on . We conclude from the previous case that is -close to either or . By switching the roles of and if necessary, we can then assume that there is such that . If then and we are done. Thus we can assume that .
Fix any point and let be defined such that . Since the geodesic rays and define the same point of the Gromov boundary, there is a constant such that for all . We apply the previous case again to a triangle with edges the segment , the segment , and a choice of geodesic , obtaining that is -close to either or . If is -close to for all then
contradicting that as . We conclude that is -close to for all sufficiently large , which implies that as desired. ∎
We can also now formally define rough starlikeness from points of . We recall that for we write for the Gromov boundary of relative to . The definition is slightly different for points of and points of , so we handle these two cases separately.
Definition 2.3.
Let be a geodesic Gromov hyperbolic space. Let and be given. We say that is -roughly starlike from if
-
(1)
For each there is a geodesic ray such that and .
-
(2)
For each there is a geodesic ray such that and .
For a point we say that is -roughly starlike from if
-
(1)
For each there is a geodesic line such that and .
-
(2)
For each there is a geodesic line such that and .
Part (2) of Definition 2.3 implies in both cases that , i.e., the geodesic boundary and the Gromov boundary coincide. It will be used as a replacement for the properness hypothesis in the main theorem of [4]. We note that Property (2) of Definition 2.3 automatically holds for any when is proper, since in this case any two points of can be joined by a geodesic. We also remark that if consists of a single point then cannot be roughly starlike from , since no geodesic line can exist in this case. Similarly if is empty then cannot be roughly starlike from any of its points.
2.2. Busemann functions
In this section we closely follow [12, Chapter 3]. Throughout much of the paper we will need to work with Gromov products based at a point . These will be defined through the use of Busemann functions. In order to use the results from [12, Chapter 3] we have to show, for a geodesic ray , that is a Busemann function based at in their sense. The definition of a Busemann function given there starts with the function
(2.6) |
for and and defines a Busemann function based at to be any function satisfying for some and (recall that we are multiplying all of their constants by due to differing definitions of hyperbolicity). Note that this alternative definition (2.6) makes sense even for points in the Gromov boundary that do not belong to the geodesic boundary.
Lemma 2.4.
Let , let , and let be a geodesic ray with . Then we have .
Proof.
We recall our definition of a Busemann function from (1.5): a Busemann function is any function such that for some geodesic ray in and some . For such a Busemann function we let denote its basepoint in . Then by Lemma 2.4 is a Busemann function based at in the sense of [12, Chapter 3] as well, provided that we use a cutoff of instead of the -cutoff used there. This only has the effect of further multiplying constants by 3 in the claims of that chapter. An easy consequence of Lemma 2.4 is the following.
Lemma 2.5.
Let and let be geodesic rays with . Then there is a constant such that . The constant depends only on the starting points and of the rays and satisfies if .
Consequently if is any Busemann function based at and is any geodesic ray then there is a constant such that .
Proof.
By [12, Lemma 3.1.2], for each we have
Setting , , and applying Lemma 2.4 gives
This gives the first claim of the lemma with . The claim that if follows from the fact that for any . The final claim follows immediately since for any Busemann function based at there is some geodesic ray such that for some . ∎
We will usually use the following lemma to perform computations with Busemann functions in practice. Note that the geodesics are parametrized as starting from the basepoint instead of ending there. The notation below should be interpreted as when .
Lemma 2.6.
Let be a Busemann function on based at . Let and let be a geodesic with as .
-
(1)
For any (or any in the case ) we have
(2.8) -
(2)
For any constant there is an arclength reparametrization of such that for .
Proof.
For , and the Gromov product based at is defined by
Since is -Lipschitz we have the useful inequality
(2.9) |
For this notion essentially reduces to the standard Gromov product: if for some and then . The analogues of all the results below then follow from the discussion in the previous section. We will thus focus on the case of Busemann functions .
Let and let be its basepoint. The Gromov product based at is extended to by, for ,
with the infimum taken over , as before, and similarly for and we define
with the infimum taken over . The next lemma extends the -inequality to Gromov products based at . It follows from [12, Lemma 3.2.4]. Recall that we have multiplied their additive constants by a total of due to the differing definition of hyperbolicity and larger cutoff in defining Busemann functions; we then round up to afterward. The corresponding additive constant in [12, Lemma 3.2.4] below is .
Lemma 2.7.
Let be a Busemann function based at . Then
-
(1)
For any , and any , we have
and the same holds if we replace with .
-
(2)
For any we have
For a point belonging to the geodesic boundary, a sequence converges to infinity with respect to if for some Busemann function based at we have as . Two sequences , converging to infinity with respect to are equivalent with respect to if . These notions do not depend on the choice of Busemann function based at by Lemma 2.5. One then defines the Gromov boundary relative to as the set of all equivalence classes of sequences converging to infinity with respect to . We will denote this by . As our past use of the notation suggests we have the following, which is [12, Proposition 3.4.1].
Proposition 2.8.
A sequence converges to infinity with respect to if and only if it converges to a point . This correspondence defines a canonical identification of and .
We recall that for we will often abuse terminology and also refer to as the Gromov boundary relative to .
2.3. Visual metrics
Let and let be a set. A function is a -quasi-metric if the following holds for any ,
-
(1)
if and only if ,
-
(2)
,
-
(3)
.
By a standard construction (see [12, Lemma 2.2.5]) a -quasi-metric with is always -biLipschitz to a metric on . Since for we have that is a quasi-metric if is a -quasi-metric, for any quasi-metric we always have that is -biLipschitz to a metric on (by the identity map on ) whenever is small enough that .
Let be a geodesic -hyperbolic space. For and we define for , ,
(2.11) |
with the understanding that . By (2.5) the function defines an -quasi-metric on . We refer to any metric on that is -biLipschitz to as a visual metric on based at with parameter ; we call the comparison constant to the model quasi-metric . A visual metric always exists once is small enough that . We give the topology induced by any visual metric. Equipped with a visual metric with respect to any basepoint and any parameter the set is a complete bounded metric space. The basepoint change inequality (2.2) combined with inequality (2.4) shows that the notion of a visual metric does not actually depend on the choice of basepoint , however the comparison constant to the quasi-metric (2.11) will depend on the basepoint. For of the form for some we then define
(2.12) |
Let be a point of the geodesic boundary and let be a Busemann function based at . We define for and ,
(2.13) |
Then defines an -quasi-metric on by Lemma 2.7. A visual metric based at with parameter is defined to be any metric on that is -biLipschitz to , and as before we will call the comparison constant to the model quasi-metric . Since all Busemann functions associated to differ from each other by a constant, up to a bounded error (by Lemma 2.5), the notion of a visual metric based at does not depend on the choice of Busemann function based at . Equipped with any visual metric based at the metric space is complete. It is bounded if and only if is an isolated point in .
Remark 2.9.
For CAT spaces the quasi-metric for with basepoint defines a distinguished visual metric on with parameter . This metric is known as a Bourdon metric when and a Hamenstädt metric when . The basic properties of the Bourdon metric for CAT spaces were established by Bourdon in [8]. The Hamenstädt metric was introduced by Hamenstädt in the setting of Hadamard manifolds with sectional curvatures in [18] through a slightly different construction. The formulation for CAT spaces using Gromov products based at is due to Foertsch-Schroeder [15].
3. Tripod maps and Busemann functions
In this section we let be a geodesic -hyperbolic space for a given parameter . We will be establishing some standard claims regarding geodesic triangles in that have vertices on the Gromov boundary . We will then use these claims regarding geodesic triangles in to evaluate Busemann functions on geodesics in in Proposition 3.10 and Lemma 3.12. When is proper these claims can be obtained via limiting arguments from the corresponding claims for geodesic triangles in [17, Chapitre 2]. Without the properness hypothesis they may be obtained (with larger constants) by careful examination and specialization of the results of Väisälä on roads and biroads in -hyperbolic space [21, Section 6]. We will provide more direct proofs of these results here, as we will also need to use some particular corollaries of the proofs that cannot be found in [21]. Providing our own proofs also allows us to organize the results in a manner that is convenient for our applications.
3.1. Tripod maps
We start with a definition. The terminology is taken from [12, Chapter 2]. Compare [17, Chapitre 2, Définition 18].
Definition 3.1.
Let be a geodesic triangle in with vertices and let be given. A collection of points , , is -equiradial if
We then refer to , , as -equiradial points for .
Remark 3.2.
When , the -tripod condition directly provides us with a set of -equiradial points , , defined by the system of equalities , , and . We will often refer to these points as the canonical equiradial points for , since they are uniquely determined. The following definition encodes a convenient hypothesis to make on equiradial points of a geodesic triangle that partially generalizes the notion of canonical equiradial points to the case that some of the vertices of belong to .
We adopt the notation convention for that if and one of or belongs to and if .
Definition 3.3.
Let be a geodesic triangle in with vertices , let be given, and let be a collection of -equiradial points for . We say that this collection is calibrated if we have , , and .
This condition is trivially satisfied when all vertices of belong to , since all of the subsegments involved have infinite length.
We let be the tripod geodesic metric space composed of three copies , , and of the closed half-line identified at . This identification point will be denoted by and will be referred to as the core of the tripod . The space is clearly -hyperbolic. The Gromov boundary consists of three points , , corresponding to the half-lines thought of as geodesic rays starting from the core .
For a geodesic triangle with a calibrated ordered triple of -equiradial points as in Definitions 3.1 and 3.3, we define the associated tripod map to be the map that sends the sides , , and isometrically into , , and respectively in the unique way that satisfies , ,, and . To be more precise for boundary points, when we mean here that , i.e., maps the geodesic rays and isometrically onto . A choice of ordering of the equiradial points is required to define the map but is not important, as changing the ordering simply corresponds to permuting the rays in while keeping the core fixed.
We first obtain the following direct consequence of Lemma 2.2.
Lemma 3.4.
Let be a geodesic triangle with vertices . Then there is a calibrated -equiradial collection of points , , .
Proof.
If all vertices of belong to then the canonical equiradial points give a calibrated -equiradial collection for , so we can assume that at least one vertex of belongs to . Thus we can assume without loss of generality that .
Parametrize the side by arclength as for an interval , oriented from to . Let be the collection of times such that and the collection of times such that . Each of the sets and are closed and we have by Lemma 2.2. We claim that both and are always nonempty. For this we can assume without loss of generality that is nonempty since .
If then . For each we let be a point such that . For the sequence converges to , which implies that the sequence converges to since these sequences are a bounded distance from one another. However any sequence of points converging to infinity in can only possibly converge to or , which is a contradiction. Thus must also be nonempty.
By the connectedness of we then conclude that . Letting , setting , and selecting points , such that and , we conclude that is a -equiradial collection of points for .
Lastly we need to produce a calibrated collection of equiradial points from the collection . If all vertices of belong to then the collection is trivially calibrated, so we can assume at least one vertex of belongs to . By relabeling the vertices we can then assume that either and or and . In both cases we can find such that since . Then
(3.1) |
It follows that the collection is -equiradial. If then this collection is also calibrated and we are done.
If then we repeat this argument again by using the fact that to find such that . The calculation (3.1) then shows that as well. We can then conclude that the collection is calibrated and -equiradial, as desired. ∎
Our next goal will be to prove that the tripod map associated to the calibrated collection of equiradial points produced by Lemma 3.4 is roughly isometric. We will require the following simple lemma.
Lemma 3.5.
Let be a metric space and let with . Suppose that we are given geodesics and joining to and to respectively. Let , be given points that satisfy and let be the unique point satisfying . Then and .
Proof.
The point must belong to the subsegment of , as if and then
contradicting that . Since we then have
which implies by the triangle inequality that . ∎
We now apply Lemma 3.5 to the setting of a -hyperbolic space .
Lemma 3.6.
Let , let , and let , satisfy . Then we have
(3.2) |
Proof.
We first assume that . We can then assume without loss of generality that . We consider a geodesic triangle with sides the given geodesics and , as well as a geodesic from to . Let be the unique point such that . Lemma 3.5 shows that and .
Let and be the canonical equiradial points for on these edges. These points must satisfy since is an edge of . The assumption then implies that and . Thus . The -tripod condition then implies that , from which it follows that . This proves (3.2) in this case.
We now consider the case . For each we define , to be the points such that and . Since the geodesics and have the same endpoint , we must have as and the same for . We choose large enough that and . We consider a geodesic triangle with edges the subsegment of the given geodesic as well as geodesics and , and a triangle with edges the subsegment of the given geodesic , the edge of , and a geodesic . Then and by our choice of .
Since , we must have . Therefore there is a unique point such that . The -tripod condition applied to the triangle then implies that . If then we let be the unique point such that . By applying Lemma 3.5 we then conclude that and . Since is an edge of the triangle , the canonical equiradial points of this triangle on the edges and can be at most a distance from the vertices and respectively. We thus conclude from the -tripod condition that . Combining these inequalities together gives
(3.3) |
which proves (3.2). The case is similar: we let be the point such that , apply Lemma 3.5 to obtain and , then apply the -tripod condition to obtain . This gives inequality (3.2) through the same calculation as (3.3). ∎
Remark 3.7.
The proof of Lemma 3.6 shows that we have the sharper inequality when .
We will use Lemma 3.6 to show that the tripod map associated to a collection of calibrated equiradial points for a geodesic triangle is roughly isometric.
Proposition 3.8.
Let be given vertices of a geodesic triangle in . Let , , be points such that is a calibrated ordered triple of -equiradial points for for a given . Let be the tripod map associated to this triple. Then is -roughly isometric.
In particular if is the calibrated -equiradial triple produced in Lemma 3.4 then is -roughly isometric.
Proof.
By symmetry (permuting the vertices , , and ), to estimate for it suffices to restrict to the case and then consider the possible locations of . By construction we have if and belong to the same edge of , since the tripod map is isometric on the edges of . This handles the case that belongs to the same edge as , i.e., that .
We next consider the case . Since , we can find a point such that . Then . Applying Lemma 3.6 yields
so that
which gives the desired estimate in this case.
Lastly we must consider the case , which we subdivide into the cases and . When we can use the condition to find a point such that . Then . Similarly to the previous case, Lemma 3.6 gives us the estimate which implies that
as desired. When we use the equality to find such that , and we use the equality to find such that . Then . Lemma 3.6 gives us the estimate
which implies by the triangle inequality that
This completes the proof of the main claim. The final assertion follows by substituting and rounding up. ∎
Remark 3.9.
Throughout this paper we will often suppress the exact choice of calibrated equiradial points used to define a tripod map . To make this more formal, for a geodesic triangle in we will refer to a tripod map associated to as being any tripod map for associated to an ordered triple of calibrated -equiradial points for obtained from Lemma 3.4. We will also abuse terminology and say that , and are equiradial points for the tripod map (as opposed to for the triangle ).
3.2. Calculating Busemann functions
We recall that the Gromov boundary of the tripod is a disjoint union of three points , , corresponding to the geodesic rays that parametrize the half-lines starting from for . Set to be the Busemann function associated to the geodesic ray . A straightforward calculation shows that is given by for and for or , when we consider each of these rays as identified with .
In this next proposition we consider a geodesic triangle in with a distinguished vertex together with a Busemann function based at . We will not keep track of exact constants in the proof of this lemma so we will not produce an explicit value for below. If one does careful bookkeeping in the proof it is possible to show that works.
Proposition 3.10.
Let be a geodesic triangle in with and . There is a constant such that the following holds: let , , and be a calibrated set of -equiradial points on provided by Lemma 3.4 and let be a Busemann function based at . Let be the tripod map associated to the triple . Then for each we have
(3.4) |
Consequently we have for and
(3.5) |
Proof.
Since we will not be keeping track of the exact value of the final constant in the proof, we will let denote any equality up to an additive error depending only on . Set . Then and since is 1-Lipschitz. We will prove the rough equality (3.4) with in place of and use this to deduce that . Thus we will first show that for we have
(3.6) |
We first handle the case in which or . Since the roles of and are symmetric, we can assume without loss of generality that . Let be an arclength parametrization of with and as . If we define such that then it follows from the construction of the tripod map that . Applying Lemma 2.6 gives
which gives (3.6) since .
The remaining case is when . By the symmetric roles of and we can assume that . As in the proof of Proposition 3.8, since we can find such that . Then by Lemma 3.6 we have
Thus since is 1-Lipschitz. It then follows, from the rough equality (3.6) for that we established above, that
which gives (3.6) in this case.
We next show that . By Lemma 2.7 we have for any sequences and that for sufficiently large ; if then we can just set for all and the same goes for . We choose sequences and that belong to and consider only those large enough that and , . Then applying (3.6),
Thus we can substitute in for in (3.6) at the cost of an additional additive constant depending only on . The main claim (3.4) follows. The assertion that for follows from (3.4) since each point of has image under the tripod map and . The rough equality (3.5) also follows directly from (3.4) since the image of under is contained in and is nonnegative on this subset of . ∎
Proposition 3.10 leads to the following important definition, which is useful for calculations. We recall our convention that when .
Definition 3.11.
Let be a geodesic -hyperbolic space and let be given with basepoint . Let and let be a given constant. Suppose that is a geodesic joining to . We say that a parametrization , , of by arclength is -adapted to if and
(3.7) |
for .
The inclusion of in the domain of will be vital for our applications. When the value of is implied by context we will often shorten the terminology to just saying that the parametrization is adapted to .
For with basepoint we will construct adapted parametrizations for geodesics joining any two points under an assumption similar to the rough starlikeness hypothesis of Theorem 1.4. We emphasize that the points and in the lemma need not always be the vertices of a geodesic triangle with a third vertex at .
Lemma 3.12.
Let be a geodesic -hyperbolic space and let be given with basepoint . Let and let be a given geodesic from to . Suppose that we are given and points and geodesics joining to and respectively such that . Then there is a constant depending only on and such that there is a parametrization of that is -adapted to .
Proof.
We first consider the case that and , so that we can take . Let be the geodesic triangle formed by the geodesics , , and . Let be a -roughly isometric tripod map associated to such that , , and , as given by Proposition 3.8. We identify the union of geodesic rays in with by identifying with and with , sending the core of to the origin in . We let denote the image of under and consider as a subinterval under the identification of with . Since the tripod map is isometric when restricted to , we can then construct an arclength parametrization of by inverting the restriction of to . By the construction of we have since the core is contained in the image of .
When the rough equality (3.4) directly implies the -adapted condition (3.11) for with , since for we have . For it is easy to see that it suffices to verify (3.7) for of the form , . We then have to show that there is a constant such that for we have
(3.8) |
By (2.4) we can find points , such that and , . Then since is -roughly isometric. Since and , a quick calculation then shows that
and therefore . Thus for we have
with the second equality following from the construction of . The rough equality (3.8) with then directly follows from the fact that is -roughly isometric.
We now consider the general case in which we are given points and such that . If and both belong to then our conventions imply that and , hence this case reduces to the case considered previously.
If and both belong to then we apply the case to the points and to obtain a -adapted parametrization of oriented from to , with . Since we have and . Let , , be the unique arclength parametrization of that is oriented from to and starts from the same time parameter as . The piecewise geodesic curve joining to can be parametrized as a -roughly isometric map for an appropriate interval . By the stability of geodesics in Gromov hyperbolic spaces [12, Theorem 1.3.2] this implies that there is a constant such that the given geodesic is contained in a -neighborhood of the curve . Since the segments and of are each contained in a -neighborhood of , by increasing by an amount depending only on we can assume that is contained in a -neighborhood of .
Now let be given and let be such that . Since is 1-Lipschitz it follows that
Thus it suffices to show that for a constant . Since and , we have
so that we can set . It follows that satisfies (3.7) with constant depending only on and .
If then gives a parametrization of that is -adapted to and we are done. We can therefore assume that which implies that since . We then note that and , , which implies that . Since and , we conclude that . We set and set for . Then by construction and this arclength parametrization of still satisfies (3.7) with since is -Lipschitz and . Thus gives the desired adapted parametrization.
Lastly we consider the case in which one of or belong to , but not both. Without loss of generality we can assume that and . Let be the sequence of points with for each . Let be the arclength parametrizations of for each that were constructed in the previous case, . Since for each we have for each . Since for each , we have from the condition that is -adapted to ,
with . It follows that for each . Thus, by replacing with the parametrization defined by on the domain , we can assume that for all . Note also that, since as and , we have for all large enough . It follows that the resulting parametrization will be -adapted to for large enough that since is 1-Lipschitz, with .
With these modifications the parametrizations now have the same starting point . Since these are parametrizations of by arclength and the sequence defines progressively longer subsegments of that exhaust , the maps coincide wherever their domains overlap and can therefore be used to define a parametrization of that is -adapted to by construction. ∎
4. Uniformization
Our task in this section will be to prove Theorems 1.4, 1.6, and 1.10. Section 4.1 establishes some general estimates for the uniformized distances . Section 4.2 proves the theorems in the case that (i.e., is a Busemann function). Section 4.3 then uses a special construction (Definition 4.18) to deduce the case from the case . Since Theorem 1.10 follows from Theorems 1.4 and 1.6 once we’ve shown that is a GH-density in this case in Proposition 4.11, we will focus our efforts on proving Theorems 1.4 and 1.6 after that point.
4.1. Estimates for the uniformized distance
In this section we will derive some estimates for the conformal deformation of a geodesic -hyperbolic space by the density for a given and using the tripod maps we built in Section 3. For now we will not be assuming that is a GH-density (using the terminology of Definition 1.3). Hence we can use these results to establish that is a GH-density in certain important cases. To simplify notation we will drop the function from the notation for objects associated to the conformal deformation and write , , etc. For a curve we will write for its length measured in the metric . We let denote the length of measured in instead.
Remark 4.1.
Throughout the rest of this paper we will be using [4, Proposition A.7], which for a geodesic metric space and a continuous function allows us to compute the lengths in the conformal deformation of curves parametrized by arclength in as
(4.1) |
with denoting the standard length element in .
Since is -Lipschitz we have the Harnack type inequality for ,
(4.2) |
which made its first appearance in the statement of Theorem 1.9 earlier.
The metric spaces and are biLipschitz on bounded subsets of by inequality (4.2). A more precise estimate for this is given in the lemma below.
Lemma 4.2.
For any we have
Proof.
For the upper bound we let be a geodesic joining to . Then, using (4.2),
For the lower bound we consider a rectifiable curve joining to in , which we can assume is parametrized by arclength as with denoting the length of in . With this parametrization defines a -Lipschitz function from to , so that in particular we have for each . Then by (4.2),
where in the final line we used that . ∎
Lemma 4.2 can be rewritten in the following useful form when .
Lemma 4.3.
For any with we have
(4.3) |
Proof.
For we have the inequalities
and
as can be verified by noting that equality holds at and differentiating each side. Thus for the inequality of Lemma 4.2 implies that
(4.4) |
with the final estimate following from
since and is -Lipschitz. ∎
The comparison (4.4) in Lemma 4.3 has the following important consequence, which proves the last claim of Theorem 1.4.
Proposition 4.4.
is bounded if and only if .
Proof.
We first suppose that . For we can then write for some and . We let be given and let be a geodesic joining to . Then
It follows that is bounded with .
Now suppose that . Then we can find a geodesic ray and a constant such that . For each we apply the comparison (4.4) with and to obtain
since . Thus as we have . It follows that is unbounded. ∎
We next use adapted parametrizations to estimate the length in of geodesics in . Below we write for the basepoint of .
Lemma 4.5.
Let be given and let be a geodesic in joining to . Suppose that we are given and points and geodesics joining to and such that . Then
(4.5) |
Proof.
Throughout this proof all additive constants and will depend only on , , and ; we write and for and respectively. We consider an arclength parametrization of that is -adapted to with as constructed in Lemma 3.12. We assume that is oriented from to and set . By (3.7) we then have .
When we observe that for all . Inequality (4.2) then implies that
(4.6) |
By integrating the comparison (4.6) over we obtain
This gives the estimate (4.5) when .
We now suppose that . Let and denote the parametrizations of the subsegments of from to and from to respectively. Then, using (3.7) and , we have
It follows immediately that
with . This gives the upper bound in (4.5) when . For the lower bound we note that since and , we must have . Therefore
This gives the lower bound in (4.5) when . ∎
In connection with Lemma 4.5 it is helpful to formulate the following definition.
Definition 4.6.
Let be given. For a constant we say that is -roughly geodesic from if for each there exists and a geodesic joining to such that .
When is -roughly geodesic from we can apply Lemma 4.5 freely to any geodesic in with this constant . If then is -roughly geodesic from since is geodesic. For the case we note that if is -roughly starlike from then is clearly also -roughly geodesic from . Note however that can be roughly geodesic from without being roughly starlike from ; this happens for instance when is a tree with arbitrarily long finite branches. We also remark that is always -roughly geodesic from any point of when it is proper.
When is a GH-density we thus obtain the following corollary of Lemma 4.5 using the GH-inequality (1.3).
Lemma 4.7.
Suppose that is -roughly geodesic from and that is a GH-density with constant . Then for any we have
(4.7) |
Corollary 4.8.
Suppose that is -roughly geodesic from and that there is a constant such that for any with we have
(4.8) |
Then is a GH-density with constant .
Proof.
We will use Corollary 4.8 to show for a CAT space that is a GH-density for any with a universal constant . Our proof will be based on the following four point inequality for CAT spaces.
Proposition 4.9.
[14, Proposition 3.3.4] Let be a CAT space. Then for any four points we have
(4.9) |
The inequality (4.9) can easily be improved to hold for Gromov products based at any function .
Lemma 4.10.
Let be a CAT space. Then for any and we have
(4.10) |
Proof.
If has the form for some then (4.10) is just a restatement of (4.9). The inequality for of the form for some and can then be obtained by multiplying the inequality (4.9) through by .
Now suppose that . Then we can find a geodesic ray and an such that . By multiplying the target inequality (4.10) through by we see that it suffices to consider the case that . Let be given. For each we have from (4.9) that
After multiplying each side by and expanding the Gromov products we obtain that
Letting then gives inequality (4.10). ∎
By combining Lemma 4.10 with Corollary 4.8 we can show that the density for is a GH-density when is a complete CAT space. The completeness hypothesis is only used in the case .
Proposition 4.11.
There is a constant such that for any complete CAT space and any we have that is a GH-density with constant .
Proof.
We first observe that is always -roughly geodesic from . For this is trivial, while for this can be deduced from the completeness hypothesis together with the CAT condition on [14, Proposition 4.4.4]. Thus we can apply Corollary 4.8 with . Let be given and let denote the basepoint of . Since is -hyperbolic with and we are restricting to the case , it suffices by Corollary 4.8 to produce a universal constant such that for any with we have
(4.11) |
We consider a rectifiable curve joining to that is parametrized by arclength (in ) and oriented from to . We define a finite sequence of points in inductively as follows: we set to be the left endpoint of and for each applicable we set to be the supremum of all points such that and for each . The finite length of in ensures that this sequence is finite, so that we have as a consequence that this process terminates at the right endpoint of . The assumption implies that . By construction we then have for and . Since is -Lipschitz it follows that
with , recalling that . Hence to prove (4.11) it suffices to show for any rectifiable curve joining to that we have
(4.12) |
for some constant . Repeated application of the inequality (4.10) based at gives
with the second inequality holding for since is -Lipschitz and . For we have by construction that any satisfies . Hence the Harnack inequality (4.2) implies that
since the length of in is at least the distance between its endpoints. This shows that
with . We conclude that the desired estimate (4.12) holds, so that as a consequence is a GH-density with a universal constant . ∎
By repurposing the proof of Proposition 4.11 we are also able to show that being a GH-density implies that is also a GH-density for each , with a constant independent of .
Proposition 4.12.
Suppose that is -roughly geodesic from and that is a GH-density with constant . Then there is a constant such that is a GH-density with constant for any .
Proof.
Let be the threshold determined by the Harnack inequality (4.2) and Theorem 1.9 such that is a GH-density with constant for . For the purpose of proving the proposition we can then assume that and . We will first produce a constant such that for any with we have
(4.13) |
As in the proof of Proposition 4.11, we let be a rectifiable curve joining to that is parametrized by arclength (in ) and oriented from to . We then construct the finite sequence of points in exactly as in the proof of Proposition 4.11, with since . Since is -Lipschitz, we conclude as in that proof that it suffices to establish the estimate
(4.14) |
for any rectifiable curve joining to , with .
We set and observe that by hypothesis. By the triangle inequality for we have
which implies that
since . By Lemma 4.7 we then conclude that
with , where the second inequality follows from the fact that by construction. Since for each , we have by the Harnack inequality (4.2),
since the length of in is at least the distance between its endpoints. It follows that
with . This proves the desired inequality (4.14).
By direct inspection of the calculations in this proof, as well as in the proofs of our previous lemmas in this section, one can verify that if then the constants can always be chosen to depend only on and . This shows that the GH-constant for can be chosen such that . Since we in fact obtain that , as desired. ∎
We recall that denotes the basepoint of . We end this section by using Lemma 4.5 to construct a map
(4.15) |
when is roughly geodesic from and . We will be using the following corollary of the estimate (4.5) for ,
(4.16) |
The inequality (4.16) implies that if is a sequence in that converges to infinity with respect to (recall this means that as ) then is a Cauchy sequence in . Since and are biLipschitz on bounded subsets of by Lemma 4.2, it is easy to see that cannot converge to a point of . It follows that is incomplete as long as contains at least one point; the only exceptions are when either is bounded or is the only point of . Furthermore a second application of (4.16) shows that sequences and converging to infinity with respect to that are equivalent with respect to are equivalent as Cauchy sequences in , i.e., as . Setting to be the complement of in its completion, we thus have a well-defined map given by sending a sequence converging to infinity with respect to to its limit in inside of . In the next section we will show that is a bijection when is roughly starlike from and is a GH-density. We remark that the map need not be a bijection in general, see [3, Proposition 4.1].
4.2. Uniformizing by Busemann functions
For this section we let be a complete geodesic -hyperbolic space. We let be given with basepoint and suppose that is -roughly starlike from . For a given we suppose is a GH-density with constant . As in Section 4.1, to simplify notation we will drop from the notation for objects associated to the uniformization and write , , etc. As before, for a curve we will write for its length measured in the metric . For brevity, throughout this section we write for equality up to an additive that depends only on , , , and , and write for equality up to a multiplicative constant that depends only on those same parameters. We write and for additive and multiplicative constants depending only on these parameters.
The rough starlikeness of from implies that contains at least one point, so as a consequence the space is incomplete by the discussion at the conclusion of the previous section. As before we write for the completion of the uniformization and for the boundary of inside its completion. We will continue to write for the canonical extension of this metric on to the completion . We write for the distance to the boundary in . We let be the map constructed in (4.15) at the end of the previous section by sending a sequence converging to infinity with respect to to its limit in as a Cauchy sequence in . We formally extend the distance function to by setting for , where we set for .
Our first lemma extends the estimate (4.7) to hold for with our formal extension of to . We recall our convention that for we define if and either or , and define if .
Lemma 4.13.
For any we have
(4.17) |
Proof.
The case in which is an immediate consequence of Lemma 4.7 since is -roughly geodesic from (because is -roughly starlike from ). We thus only need to consider the case in which at least one of the points belongs to . Since (4.17) holds trivially when we can assume that . We can then assume without loss of generality that . We then need to show that . We let and be sequences converging to infinity with respect to in that represent the points and respectively; if then we instead set for all . Our definition of the extension of to then implies that we have . For sufficiently large we will have by Lemma 2.7 and we will have . The comparison (4.17) then follows from the corresponding comparison for and for sufficiently large . ∎
By using Lemma 4.13 we are able to show that is a bijection.
Lemma 4.14.
The map is a bijection.
Proof.
The injectivity of follows immediately from Lemma 4.13 applied to . Thus our focus will be on showing that is surjective. Let be a Cauchy sequence in that converges to a point . We claim that the sequence cannot belong to a bounded subset of . If it did then for a fixed there would be an such that for all , with denoting the ball of radius centered at in . Lemma 4.2 shows that the metrics on and are biLipschitz to one another on , which implies that is also a Cauchy sequence in . Since is complete this Cauchy sequence must converge in to a point . However this means that also converges to in , contradicting that converges to a point of .
Thus, by passing to a subsequence if necessary, we can assume that for . It then follows from Lemma 4.13 that for ,
Since as , we conclude that as . Thus converges to infinity with respect to . Letting denote the point in the Gromov boundary relative to represented by the sequence , the construction of then shows that . We conclude that is surjective. ∎
We can now prove Theorem 1.6 in the case . We recall the definition (2.13) of the model visual quasi-metric on .
Proof of Theorem 1.6.
The next lemma shows that the distance to can be computed in terms of the density .
Lemma 4.15.
For we have
(4.18) |
Proof.
Let be given. We first compute the upper bound in (4.18). By the rough-starlikeness condition we can find a geodesic line starting at and ending at some with . Using Lemma 2.6 we can consider as parametrized by arclength with for each . We let be such that . We then compute,
By Lemma 4.2 we then have
Since the upper bound follows.
For the lower bound we let be a given point, which we can think of as a point in using Lemma 4.14. By rough starlikeness we can then find a geodesic line starting at and ending at . For we note that as , so we will have for all sufficiently large . For sufficiently large we can then apply (4.17) and Lemma 2.7 to obtain
By (2.10) we have . By combining this with the above we obtain that
This gives the lower bound since as . ∎
We can now complete the proof of Theorem 1.4 in the case . Since we have already shown that is unbounded in Proposition 4.4, to prove Theorem 1.4 we only need to show that geodesics in are uniform curves in . For this we need to extend the definition of uniform curves to cover curves defined on arbitrary subintervals .
As in Definition 1.1, we let be an incomplete metric space, set , and set . We consider a curve defined on an arbitrary subinterval ; we write and for the endpoints of . If has finite length, , then has well-defined endpoints defined by the limits and in , which exist because .
Definition 4.16.
For a constant and an interval , a curve with is -uniform if
(4.19) |
and if for every we have
(4.20) |
If then we instead define to be -uniform if (4.20) holds and if as and .
Proposition 4.17.
There is an such that any geodesic in is an -uniform curve in . Consequently is -uniform.
Proof.
We first consider the case of a geodesic in joining two points . Let be a parametrization of that is -adapted to , , as in Lemma 3.12; we can always find points satisfying the hypotheses of the lemma by the -rough starlikeness hypothesis from . Let , be the endpoints of . Let be sequences such that , , and for each . Applying inequality (1.3) to the geodesic joining to gives
(4.21) |
Letting , the left side of (4.21) converges to . If then the sequence converges to in , while if then the sequence belongs to the equivalence class of with respect to . In both cases we then have that converges to in , with the second case following from the construction of the identification . The same discussion applies to the sequence in relation to . It follows that as . Consequently has finite length in with endpoints and . The inequality (4.19) then follows by letting in (4.21).
We next verify inequality (4.20). It suffices to verify this inequality in the case that , since we can deduce the case from this by reversing the roles of and . We thus assume that . A straightforward calculation with (3.7) gives us that
(4.22) |
Since , it then follows from (4.22) and Lemma 4.15 that
with . We conclude that is an -uniform curve in with . Since any two points can be joined by a geodesic in , this implies that the metric space is -uniform.
It remains to treat the case of a geodesic joining the basepoint of to a point . By applying (2) of Lemma 2.6 with we can find an arclength parametrization of , , with for . For we then have, by a computation similar to the one done in Lemma 4.15,
(4.23) |
By letting we conclude that . The comparison (4.23) together with the GH-inequality (1.3) then implies that as and . Thus our replacement condition in the infinite length case for (4.19) is satisfied. Combining (4.23) and Lemma 4.15 also implies for each that
with in each inequality. Thus (4.20) also holds for with . We conclude that is an -uniform curve in in this case as well. ∎
4.3. Uniformizing by distance functions
For this section we assume the same setup as in Section 4.2, with the exception that we will be assuming instead that with basepoint . We will reduce Theorems 1.4 and 1.6 in this case to the case considered in Section 4.2 using the following general construction.
Definition 4.18.
For a metric space and a point we let be the metric space obtained by gluing the half-line to by identifying the point with . The metric on is defined by setting for , for , and for , . The space then has a canonical isometric embedding into . We refer to the metric space as the ray augmentation of based at .
The following trick will be the basis of many of the results we prove regarding the ray augmentation.
Lemma 4.19.
Let be a metric space. Let be the ray augmentation of based at . Then for each curve in there is a curve such that when and when .
Proof.
Consider the retraction given by setting for and for . It is easy to see that is -Lipschitz; in particular is continuous. For a given curve the curve then has the desired properties. ∎
We refer to the curve constructed from in Lemma 4.19 as the shortening of to . We apply Lemma 4.19 to conformal deformations of the ray augmentation.
Lemma 4.20.
Let be a geodesic metric space and let be the ray augmentation of based at . Let be a continuous density on and let be a continuous density on with . Then the isometric embedding induces an isometric embedding of the corresponding conformal deformations. Furthermore is a GH-density with constant if and only if is a GH-density with the same constant .
Proof.
We write for the metric on and for the metric on . Let be given. Let be a rectifiable curve joining them and let be the shortening of to . Since we clearly have
It follows that . Thus in computing it suffices to minimize over curves taking values only in . Since for such curves we have , it immediately follows that . We conclude that the embedding is an isometry.
It is obvious that is a GH-density with constant if is a GH-density with constant , since restricts to on and geodesics in are also geodesics in . For the converse we assume that is a GH-density with constant and let be given points. We need to prove the GH-inequality (1.3) for any geodesic joining to in .
If then any geodesic joining to in is in fact a geodesic joining to in . The inequality (1.3) for then follows from the corresponding inequality for . If then there is only one geodesic from to in , which is simply the interval connecting them in . Since any curve joining to in must contain this interval we in fact have . Thus inequality (1.3) holds in this case as well.
The final case is that in which and . Let be a geodesic joining to in . Then we can write , where is a geodesic in joining to and is the geodesic in joining to , which is just the interval connecting these points. Given a rectifiable curve joining to we let be the shortening of to . Then by the inequality (1.3) for we have . Since the intersection must contain the interval joining to , we deduce from this that
Minimizing over all rectifiable curves joining to then gives inequality (1.3). ∎
We assume now that is a geodesic -hyperbolic space. We let be the ray augmentation of based at some point . It is an easy exercise to see that is also -hyperbolic; recall that we have defined -hyperbolicity using -thin triangles. We also note that is complete if is complete. We will continue to use the generic distance notation for the distance between , noting that there is no conflict with the distance notation for since sits isometrically inside of .
By definition the canonical ray in is the geodesic ray corresponding to the canonical parametrization of the copy of that we glued onto . The key property of the ray augmentation is that the Busemann function associated to the canonical ray restricts on to the distance from the distinguished point .
Lemma 4.21.
Let be the canonical ray in . Then for we have .
Proof.
For and we have
which implies upon taking that . ∎
Let be the point in the Gromov boundary defined by the canonical ray. We next show that identifies canonically with and rough starlikeness from in passes over to rough starlikeness from in .
Lemma 4.22.
We have . Furthermore any visual metric on based at with parameter also defines a visual metric on based at with parameter , and the converse holds as well. If is -roughly starlike from then is -roughly starlike from .
Proof.
For the first assertion it suffices to show that if is a sequence converging to infinity in then there is an such that for the points either all belong to or all belong to . Recall that converges to infinity if we have as . If our assertion did not hold then we could find subsequences and of the sequence such that , , and as . But then and therefore
contradicting our assumption that . Since all sequences converging to infinity in that belong exclusively to must converge to , it follows that we have a canonical identification of with .
Let be the Busemann function associated to the canonical ray as in Lemma 4.21. We observe that for implies that for . Since any sequence converging to infinity in that does not converge to must eventually stay within , it follows that for . Thus through our identification we have a canonical identification between visual metrics on based at and visual metrics on based at for any parameter .
Observe that if is a geodesic ray starting at then the map defined by for and for defines a geodesic line in that begins at , coincides with inside of , and has the same endpoint in as . This implies that if is -roughly starlike from for some then is -roughly starlike from . ∎
Proof.
Let be a complete geodesic -hyperbolic space that is -roughly starlike from and let have the form for some . We assume that is such that is a GH-density with constant . We let be the ray augmentation of based at , let be the canonical ray in , and set . Then is a complete geodesic -hyperbolic space that is -roughly starlike from the endpoint of the canonical ray by Lemma 4.22. We have by Lemma 4.21. Thus by Lemma 4.20 the embedding is isometric and is a GH-density with the same constant . We can then apply Theorems 1.4 and 1.6 to equipped with the Busemann function .
For we have . Thus a straightforward calculation shows that . It follows from this and the GH-inequality (1.3) that the only boundary points of the metric space are the boundary points of , i.e., . By applying Theorem 1.6 to and then using Lemma 4.22, we conclude that we have a canonical identification . Since visual metrics on based at with parameter correspond to visual metrics on based at with the same parameter , it then follows from Theorem 1.6 that the restriction of to defines a visual metric on with parameter and comparison constant . This completes the proof of Theorem 1.6.
5. Hyperbolic fillings
Let be a metric space and let be given parameters. We recall the construction of a hyperbolic filling of with these parameters described prior to Theorem 1.12. For each we select a maximal -separated subset of . The existence of such a set is guaranteed by a standard application of Zorn’s lemma. Then for each the balls , , cover .
The vertex set of has the form
To each vertex we associate the dilated ball . We will often use to denote both a vertex in and its associated point in . We also define the height function by . By construction for each there is a such that .
We place an edge in between distinct vertices and if and only if and . Thus there is an edge between vertices if and only if they are of the same or adjacent height and there is a nonempty intersection of their associated balls. For vertices we write if there is an edge between and . Edges between vertices of the same height are referred to as horizontal, and edges between vertices of different heights are called vertical. We say that an edge path between two vertices is vertical if it is composed exclusively of vertical edges.
While we will allow any choice of , we will need to place some constraints on the values of the parameter based on . We will require that
(5.1) |
We give each connected component of the unique geodesic metric in which all edges have unit length. The restriction (5.1) will be used in Proposition 5.5 to show that is actually connected and is therefore a geodesic metric space itself. For applications a standard choice of parameters satisfying (5.1) is given by and .
Remark 5.1.
We do not know whether the constraint (5.1) can be relaxed while preserving the properties of described below. In particular we do not know whether is always Gromov hyperbolic or even connected for all . However, by applying Lemma 5.2 below it is easy to see that is connected for any when is bounded. We note that one cannot take in the construction as it is possible for the resulting graph to fail to be Gromov hyperbolic even in the bounded case [3, Example 8.8].
Since edges can only connect vertices of the same or adjacent heights, all vertical edge paths are geodesics in . We will refer to these vertical paths as vertical geodesics. We will use the generic distance notation for the distance between . Thus for we will denote their distance in by and their distance in by . Identifying an edge from a vertex to a vertex isometrically with , we extend the height function to by . Then defines a function that is -Lipschitz on the connected components of .
We begin with a simple lemma.
Lemma 5.2.
Let with and . Then there is a vertical edge path from to .
Proof.
Let , , and let . We can assume without loss of generality that . For each integer we can find a vertex with ; we set and . Then for each by the construction of the graph . It follows that is connected to by a vertical edge path passing through the vertices . ∎
The next lemma estimates the distance in between vertices in that are connected by a vertical edge path.
Lemma 5.3.
Let . Suppose that is joined to by a vertical edge path and . Then
Proof.
We first derive a sharper inequality in the case . Set . Let . Then
Now let , . For each we let be the vertex at this height in the vertical edge path joining to . Then by the “” case we have
with the final inequality following by summing the geometric series in . ∎
Following the hyperbolic filling construction in [12], we define a cone point for a pair of vertices to be a vertex that can be joined to both and by vertical geodesics and that satisfies . A branch point for is defined to be a cone point of maximal height. A branch point for always exists as long as there is at least one cone point for . When the vertex is trivially a branch point for the set .
Lemma 5.4.
Let be distinct vertices with . Then there is a branch point for the set .
Proof.
The assumptions imply that . Let be a point in this intersection. Since is a maximal -separated set in we can find such that . We compute
by inequality (5.1), noting that the final inquality here is equivalent to
which is implied by (5.1). It follows that and therefore . Thus is joined to by a vertical edge. Since the roles of and are symmetric, we conclude by the same calculation that , i.e., is also joined to by a vertical edge. Thus is a cone point for . Since a cone point for a pair of distinct vertices on an adjacent level is trivially maximal, we conclude that is a branch point for . ∎
We can now show that the graph is connected.
Proposition 5.5.
For each there is a branch point for the set that satisfies
(5.2) |
Consequently the graph is connected.
Proof.
Let , be given. We can assume without loss of generality that . If then is a branch point for the set and the comparison (5.2) holds trivially. We can thus assume for the rest of the proof that . We let be any integer satisfying and ; note that such an integer always exists since as . Let be a vertex such that and let be a vertex such that . Then
by (5.1). Thus , so . We conclude that . By Lemma 5.4 we can then find a branch point for the set . Since and , Lemma 5.2 shows that and are connected to and respectively by vertical edge paths, and the requirement implies that . Since and are each connected to by a vertical edge, we conclude that is a cone point for the set .
It follows that there is a branch point for the set . Since is joined to and by vertical edge paths, the triangle inequality and Lemma 5.3 implies that
Since , we have and therefore
which gives the lower bound in the comparison (5.2).
For the upper bound in (5.2) we split into two cases. The first case is that in which . If then this implies that and Lemma 5.4 implies that . The upper bound follows immediately from this, as we then have
If then by Lemma 5.2 can be joined to by a vertical edge path. In this case is a branch point for the set and the inequality
holds trivially for .
The second case is that in which have . This implies in particular that we must have . Thus . Let be the maximal integer such that and . Then either or . Since and , we conclude in both cases that . Making this choice of in the construction of above, we can thus construct a cone point for the set with and therefore
Since the branch point satisfies it follows that
The upper bound in (5.2) follows.
Lastly, since we can connect to through the branch point , it follows that and can be connected by an edge path in the graph . Since and were arbitrary we conclude that is connected. ∎
Now that we’ve shown is connected, the metrics we put on its connected components give it the structure of a geodesic metric space in which all edges of have unit length. The height function then defines a -Lipschitz function . We formally define the Gromov product based at by, for ,
Since is -Lipschitz we have
(5.3) |
Our next lemma gives a key relation of the Gromov product based at to branch points.
Lemma 5.6.
Let and let be a branch point for . Then
and therefore
Proof.
Since the claims of the lemma hold trivially when we can assume that . Proposition 5.5 gives the existence of a branch point for satisfying (5.2). The vertical edge path from to followed by the vertical edge path from to gives an edge path from to , which shows that
Rearranging this we obtain
To get a bound in the other direction, let be a sequence of vertices joined by edges that gives a geodesic from to . Then and since . For we have and therefore, using ,
We can run the same argument viewing as a geodesic from to instead, setting for . We see from this that we also have
for . For each we thus obtain an estimate (using and ),
We set (the least integer greater than this quantity), observing that since . This gives, after some simplification,
recalling that . By Proposition 5.5 and inequality (5.3), we then have
which implies upon taking logarithms that
This gives the desired lower bound of the first approximate equality of the lemma. The second comparison inequality follows by using Proposition 5.5 again. ∎
We now prove an inequality similar to the -inequality (2.3) for our formal Gromov products based at .
Lemma 5.7.
Let . Then
Proof.
Let be vertices. By the triangle inequality in we have
which becomes, upon applying Lemma 5.6,
Taking logarithms of each side gives the desired inequality. ∎
We can now show that is Gromov hyperbolic. For this we use some terminology from [12, Chapter 2]: a -triple for is a triple of real numbers such that the two smallest numbers differ by at most . Observe that is a -triple if and only if the inequality
(5.4) |
holds for all permutations of the roles of , , and . We will also need the following standard claim [12, Lemma 2.1.4] which is called the Tetrahedron lemma.
Lemma 5.8.
Let , , , , , be six numbers such that the four triples , , , and are -triples. Then
is a -triple.
Proposition 5.9.
The space is -hyperbolic with .
Proof.
We will use the cross-difference triple defined in [12, Chapter 2.4]. For a quadruple of points and a fixed basepoint this triple is defined by
The triple has the same differences among its members as the triple
as a routine calculation shows for instance that
with both expressions being equal to
Similar calculations give equality for the other differences. Thus is a -triple for a given if and only if is a -triple.
Using Lemma 5.7 we conclude that the six numbers , , , , , together satisfy the hypotheses of Lemma 5.8 with parameter . This implies that is a -triple and therefore that is a -triple. By [12, Proposition 2.4.1] this implies that inequality (2.3) holds for Gromov products based at in (with replacing ). By [17, Chapitre 2, Proposition 21] this implies that geodesic triangles in are -thin, i.e., is -hyperbolic. ∎
We next show that any vertex in is part of a vertical geodesic line. We will in fact show something stronger. We let denote the completion of , and continue to write for the canonical extension of the metric on to its completion. For and a point we will write for the open ball of radius centered at in the completion .
Lemma 5.10.
Let . Then there is a vertical geodesic with for such that, writing for , we have for each . Furthermore if is a given vertex of then we can construct such that .
Proof.
Since by (5.1) and since for each the balls cover for , it follows from the fact that is dense in that the balls for cover . Thus, given , for each we can find such that .
Let be the associated vertex in . We claim that for each we have . Since is dense in we can find such that . Then
which implies that . A similar calculation shows that since . Thus and therefore . We can then find a vertical geodesic through the sequence of vertices , which can be parametrized such that for . Finally, if is a vertex of then we can choose in our construction since we trivially have . ∎
A descending geodesic ray is a vertical geodesic ray such that is strictly decreasing as a function of . In this case we have for each . Similarly an ascending geodesic ray is a vertical geodesic ray such that is strictly increasing as a function of . In this case we instead have that for each . A vertical geodesic is anchored at a point if for each vertex belonging to we have ; when the point does not need to be referenced we will just say that is anchored. Lemma 5.10 gives the existence of ascending and descending geodesic rays in anchored at any point .
We will next show that all anchored descending vertical geodesic rays in define the same point in the Gromov boundary .
Lemma 5.11.
Let , be two descending geodesic rays in starting at vertices and of respectively and anchored at respectively. Let be such that and . Let , be the vertices on these geodesics at the height . Then .
Proof.
By the anchoring condition we have and . Hence
Thus and therefore , which implies that either or . In both cases we conclude that . ∎
The Busemann functions associated to anchored descending geodesic rays have a particularly simple form.
Lemma 5.12.
Let be an anchored descending geodesic ray in starting from a vertex . Then for all we have
(5.5) |
Proof.
Since both and are -Lipschitz and the edges of have unit length, it suffices to prove the estimate (5.5) on the vertices of with the constant instead of . Let be the anchoring point for . Let be an arbitrary vertex and let be an anchored descending geodesic ray in starting at and anchored at the point associated to , as constructed in Lemma 5.10. The sequences of vertices on satisfies since is a descending geodesic ray, and the same holds for . We let be any integer large enough that and and observe that if we define then . Then by Lemma 5.11 we have . Since is joined to by a vertical geodesic of length , it follows immediately that
and therefore, since ,
This implies that
By letting we conclude that
which gives the desired result. ∎
In particular, for an anchored descending geodesic ray with , Lemma 5.12 shows that . We fix such a descending geodesic ray for the remainder of this section and write for the associated Busemann function. Let be the point corresponding to the equivalence class of in the Gromov boundary of ; note that Lemma 5.11 shows that all anchored descending geodesic rays belong to the equivalence class defined by . Our final goal in this section is to show that the boundary of relative to can be canonically identified with the completion of in such a way that the extension of the metric to is a visual metric on based at with parameter .
We remark that the rough equality implies that for all as well, so that in particular the conclusions of Lemma 5.5 hold with replacing and replacing everywhere, at the cost of adding to the constant there and multiplying the constant by . We will use this observation without further comment below.
For each point we fix an ascending geodesic ray anchored at , as given by Lemma 5.10. We define a map by setting , i.e., is the equivalence class in defined by the geodesic ray . Implicit in this definition is the fact that we cannot have for any , as it is easy to see from the fact that is -Lipschitz that ascending geodesic rays cannot be at bounded distance from descending geodesic rays. We also note that if is any other ascending geodesic ray anchored at then we must have : for sufficiently large the unique vertices and must satisfy since , hence the geodesic rays and are at a bounded distance from one another. The map can thus equivalently be thought of as sending to the equivalence class of all ascending geodesic rays anchored at .
Proposition 5.13.
The map defines an identification of with . Under this identification the metric on defines a visual metric on with parameter and comparison constant depending only on and .
Proof.
Let be given and let and be ascending geodesic rays anchored at and respectively. For sufficiently large we let and be the unique vertices on these rays at height . By Lemma 5.6 we have
(5.6) |
Since and we have and . Hence, by letting in (5.6) and using Lemma 2.7, we conclude that
(5.7) |
It follows immediately that is injective. To complete the proof of the proposition it suffices to show that is surjective, as the estimate (5.7) then shows that the metric on defines a visual metric on with parameter and comparison constant depending only on and when we use to identify with .
We recall from Proposition 2.8 that can be defined as equivalence classes of sequences in such that as , with two sequences , being equivalent if as . Since is -Lipschitz we can always choose these sequences to consist of vertices in by replacing with a nearest vertex .
Thus let be a sequence of vertices defining a point of . Let be the associated sequence of points in . By Lemma 5.6 we have
Since and , it follows immediately that is a Cauchy sequence in and therefore converges to a point . We claim that .
Let be an ascending geodesic ray anchored at and let be the sequence of vertices on starting from its initial point. When considered as points of this sequence of vertices must satisfy since is ascending and anchored at . This implies that . Since by Lemma 5.6 we have
we conclude that as . Hence and define the same point of , i.e., . We conclude that is surjective. ∎
6. Uniformizing the hyperbolic filling
This final section is devoted to proving Theorem 1.12. We retain all hypotheses and notation from the previous section. In particular we let be a metric space and let be a hyperbolic filling of with parameters and as in the previous section. We let be the height function and set . We write for the conformal deformation of with conformal factor , for the metric on , and for lengths of curves measured in the metric . Since is -Lipschitz the density satisfies the Harnack inequality (4.2) with .
In the notation of Remark 1.7, Lemma 5.12 shows that we have . Clearly is geodesic and complete, and Proposition 5.9 shows that is -hyperbolic with . Thus to prove Theorem 1.12 it suffices to show that is -roughly starlike from and that is a GH-density with constant , as it then follows for a Busemann function with that is a GH-density with constant as well, where . The conclusions of Theorem 1.12 can then be derived from the fact that the metrics on and are -biLipschitz to one another by the identity map on .
We first look at rough starlikeness from the distinguished point corresponding to the equivalence class of all anchored descending geodesic rays in .
Lemma 6.1.
The hyperbolic filling is -roughly starlike from .
Proof.
Let be a vertex of with associated point . Let be an ascending vertical geodesic line through that is anchored at , as constructed in Lemma 5.10, parametrized by arclength such that . We put for . Then is an anchored descending geodesic ray and therefore belongs to the equivalence class by Lemma 5.11. This shows that any vertex of lies on a geodesic line starting at . Since any point in is within distance of some vertex, condition (1) of Definition 2.3 follows.
For condition (2) we use the identification of with from Proposition 5.13. Let be given and let be a vertical geodesic line anchored at and parametrized such that , as constructed in Lemma 5.10. By the construction of the identification in Proposition 5.13 the geodesic ray then belongs to the equivalence class of when is considered as a point of . Putting as above, we also have that is a descending geodesic ray anchored at and therefore belongs to the equivalence class of by Lemma 5.11. Since was arbitrary, condition (2) follows. ∎
We will now show that is a GH-density with constant depending only on and . We will do this by estimating the distance between points at sufficiently large scales in and then using Corollary 4.8.
Lemma 6.2.
Let with . Then we have
(6.1) |
Consequently is a GH-density with constant .
Proof.
The bound on from above follows from Lemma 4.5 applied with , since there is a Busemann function such that . Hence it suffices to establish the lower bound. Observe that for an edge of , considered as a path between its endpoints and and assuming the orientation in which , when we have
(6.2) |
On the other hand, since we have . It follows that
Similarly, when and we have
(6.3) |
while implies again that . Thus in this case we also have
Now let be a rectifiable curve joining to . Let be the first vertex on met traveling from to and let be the first vertex on met traveling from to . We first suppose that . We then let be the subcurve of from to starting from this first occurrence of and ending at this last occurrence of . Let be the sequence of vertices encountered along the path , noting that by assumption we have . Then from our calculations above we have
(6.4) |
On the other hand, since the curve must contain at least one full edge of with one vertex being and at least one full edge with one vertex being (these may be the same edge). Then it follows from (6.2) and (6.3) applied to those edges that
(6.5) |
By combining (6.4) and (6.5) and then using Lemma 5.6, we conclude that
Since is a subcurve of it follows that this inequality holds for as well.
Now suppose that . Then, since , either the initial segment of from to or the final segment of from to has length at least in . By reversing the roles of and if necessary we can assume that the initial segment of from to has length at least in . Since is contained entirely in a single edge of that has as a vertex, it follows that
with denoting the length of in . Hence
using that since . This gives a similar lower bound on in this case as well. Minimizing over all rectifiable paths from to then gives in both cases that
with the second inequality following from the fact that is -Lipschitz and , .
To conclude that is a GH-density we let be a Busemann function on such that as in Lemma 5.12. We let be the distance obtained on through conformal deformation with the conformal factor
where . Then is -biLipschitz to by the identity map on . Consequently the comparison 6.1 holds with replacing . We can then apply Corollary 4.8 to conclude that is a GH-density with constant , noting that is -hyperbolic with and -roughly starlike from . The GH-inequality (1.3) for then follows immediately from the -biLipschitz comparison of to . ∎
This completes the proof of Theorem 1.12 aside from the final assertion regarding the identification of with given by the combination of Lemma 4.14 and Proposition 5.13 is biLipschitz. This is shown below.
Proposition 6.3.
The identification is biLipschitz with biLipschitz constant depending only on and .
Proof.
We consider as equipped with the visual metric with parameter defined by Proposition 5.13, which coincides with the extension of the metric on to the completion under the identification of that proposition. By Theorem 1.6 applied with this visual metric and , the identification is biLipschitz. Hence the induced identification given by is also biLipschitz. Furthermore all of the parameters involved in the biLipschitz constant can be taken to depend only on and by the results of this section. ∎
Remark 6.4.
For there is a canonical correspondence between hyperbolic fillings with fixed parameters of the metric spaces and given by considering -separated sets in as -separated sets in . Thus when is bounded there is no harm in assuming that by multiplying the metric by for sufficiently large. The hyperbolic filling can then be written as , where is the set of all points of nonnegative height and is the set of all points of nonpositive height. The condition implies that the vertex sets for consist only of a single point, and in particular is simply a descending geodesic ray starting from . The space is isometric to the ray augmentation of based at , in the language of Definition 4.18.
The graph is essentially the hyperbolic filling of constructed in [3], with the exceptions that they have a stricter condition for the placement of vertical edges and that they require an additional nesting condition for . They uniformize this filling for all using the density for , for which it is easy to see that , where and is the height function. When satisfies (5.1) we can use Theorem 1.12 to deduce their results from ours, up to some minor differences in the definition of the hyperbolic filling. When is close to it is possible to realize trees as hyperbolic fillings [3, Theorem 7.1], whereas when satisfies (5.1) the hyperbolic filling is only a tree if consists of a single point (by Lemma 5.4).
References
- [1] W. Ballmann, M. Gromov, and V. Schroeder. Manifolds of nonpositive curvature, volume 61 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1985.
- [2] A. Björn, J. Björn, and N. Shanmugalingam. Bounded geometry and -harmonic functions under uniformization and hyperbolization. J. Geom. Anal. To appear.
- [3] A. Björn, J. Björn, and N. Shanmugalingam. Extension and trace results for doubling metric spaces and their hyperbolic fillings. 2020. arXiv:2008.00588.
- [4] M. Bonk, J. Heinonen, and P. Koskela. Uniformizing Gromov hyperbolic spaces. Astérisque, (270):viii+99, 2001.
- [5] Mario Bonk and Thomas Foertsch. Asymptotic upper curvature bounds in coarse geometry. Math. Z., 253(4):753–785, 2006.
- [6] Mario Bonk and Eero Saksman. Sobolev spaces and hyperbolic fillings. J. Reine Angew. Math., 737:161–187, 2018.
- [7] Mario Bonk, Eero Saksman, and Tomás Soto. Triebel-Lizorkin spaces on metric spaces via hyperbolic fillings. Indiana Univ. Math. J., 67(4):1625–1663, 2018.
- [8] Marc Bourdon. Structure conforme au bord et flot géodésique d’un -espace. Enseign. Math. (2), 41(1-2):63–102, 1995.
- [9] Marc Bourdon and Hervé Pajot. Cohomologie et espaces de Besov. J. Reine Angew. Math., 558:85–108, 2003.
- [10] C. Butler. Doubling and Poincaré inequalities for uniformized measures on Gromov hyperbolic spaces. In preparation.
- [11] C. Butler. Extension and trace theorems for noncompact doubling spaces. 2020. arXiv:2009.10168.
- [12] Sergei Buyalo and Viktor Schroeder. Elements of asymptotic geometry. EMS Monographs in Mathematics. European Mathematical Society (EMS), Zürich, 2007.
- [13] Matias Carrasco Piaggio. On the conformal gauge of a compact metric space. Ann. Sci. Éc. Norm. Supér. (4), 46(3):495–548 (2013), 2013.
- [14] Tushar Das, David Simmons, and Mariusz Urbański. Geometry and dynamics in Gromov hyperbolic metric spaces, volume 218 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2017. With an emphasis on non-proper settings.
- [15] Thomas Foertsch and Viktor Schroeder. Hyperbolicity, -spaces and the Ptolemy inequality. Math. Ann., 350(2):339–356, 2011.
- [16] F. W. Gehring and W. K. Hayman. An inequality in the theory of conformal mapping. J. Math. Pures Appl. (9), 41:353–361, 1962.
- [17] É. Ghys and P. de la Harpe, editors. Sur les groupes hyperboliques d’après Mikhael Gromov, volume 83 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1990. Papers from the Swiss Seminar on Hyperbolic Groups held in Bern, 1988.
- [18] U. Hamenstädt. A new description of the Bowen-Margulis measure. Ergodic Theory Dynam. Systems, 9(3):455–464, 1989.
- [19] Eero Saksman and Tomás Soto. Traces of Besov, Triebel-Lizorkin and Sobolev spaces on metric spaces. Anal. Geom. Metr. Spaces, 5(1):98–115, 2017.
- [20] J. Väisälä. The free quasiworld. Freely quasiconformal and related maps in Banach spaces. In Quasiconformal geometry and dynamics (Lublin, 1996), volume 48 of Banach Center Publ., pages 55–118. Polish Acad. Sci. Inst. Math., Warsaw, 1999.
- [21] J. Väisälä. Gromov hyperbolic spaces. Expo. Math., 23(3):187–231, 2005.
- [22] Q. Zhou. Uniformizing gromov hyperbolic spaces and busemann functions. 2020. arXiv:2008.01399.