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

Uniformizing Gromov hyperbolic spaces with Busemann functions

Clark Butler
Abstract.

Given a complete Gromov hyperbolic space XX that is roughly starlike from a point ω\omega in its Gromov boundary GX\partial_{G}X, we use a Busemann function based at ω\omega to construct an incomplete unbounded uniform metric space XεX_{\varepsilon} whose boundary Xε\partial X_{\varepsilon} can be canonically identified with the Gromov boundary ωX\partial_{\omega}X of XX relative to ω\omega. 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 ZZ, that there is a hyperbolic filling XX of ZZ that can be uniformized in such a way that the boundary Xε\partial X_{\varepsilon} has a biLipschitz identification with the completion Z¯\bar{Z} of ZZ. We also prove that this uniformization procedure can be done at an exponent that is often optimal in the case of CAT(1)(-1) 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 {(x,y):y>0}\{(x,y):y>0\} in 2\mathbb{R}^{2} from the hyperbolic plane 2\mathbb{H}^{2}, which is discussed in Example 1.5. The guiding example in [4], by comparison, is the relationship between 2\mathbb{H}^{2} and the Euclidean unit disk {(x,y):x2+y2<1}\{(x,y):x^{2}+y^{2}<1\}. As can be seen from these examples, the input for uniformization is a geodesic Gromov hyperbolic space XX and the output is an incomplete metric space Ω\Omega, obtained from a conformal deformation of XX, that is uniform in the sense of Definition 1.1 below. The density used for uniformizing a Gromov hyperbolic space XX in [4] is exponential in the distance to a fixed point zz of XX. 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 XX. 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 XX [1, §\S3]. Unlike in [4], we will not assume any local compactness properties on XX, 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 XX of a metric space ZZ, with a particular focus on the case in which ZZ is unbounded. When ZZ is bounded a hyperbolic filling XX of ZZ can be thought of as a Gromov hyperbolic graph whose Gromov boundary can be canonically identified with ZZ; in the case that ZZ 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 ZZ.

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 ZZ and Besov classes of functions on ZZ in the special case that ZZ 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 ZZ 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 ZZ, 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 0<εε00<\varepsilon\leq\varepsilon_{0} of exponents considered by Bonk-Heinonen-Koskela (see Theorem 1.9 below). Our applications to CAT(1)(-1) 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 (X,d)(X,d) and a curve γ:IX\gamma:I\rightarrow X, II\subset\mathbb{R} a subinterval, we write (γ)\ell(\gamma) for the length of γ\gamma measured in XX. We will follow the standard practice of using γ\gamma to denote both the parametrization of the curve and the image of the curve in XX. The curve γ\gamma is a geodesic if it is isometric as a mapping of II into XX. We say that XX is geodesic if any two points can be joined by a geodesic. We will use the following distance notation for distance from a point xx to a set EE in any metric space (X,d)(X,d),

dist(x,E)=infyEd(x,y),\mathrm{dist}(x,E)=\inf_{y\in E}d(x,y),

and in particular will write dist(x,γ)\mathrm{dist}(x,\gamma) for the distance of a point xXx\in X to (the image of) a curve γ\gamma.

We now define uniform metric spaces. We start with an incomplete metric space (Ω,d)(\Omega,d). We denote the boundary of Ω\Omega in its completion Ω¯\bar{\Omega} by Ω=Ω¯\Ω\partial\Omega=\bar{\Omega}\backslash\Omega. For xΩx\in\Omega we write dΩ(x):=dist(x,Ω)d_{\Omega}(x):=\mathrm{dist}(x,\partial\Omega) for the distance from xx to the boundary Ω\partial\Omega. For the definition below we allow I=[a,b]I=[a,b]\subset\mathbb{R} to be any closed interval, and for a curve γ:IX\gamma:I\rightarrow X we denote its endpoints by γ:=γ(a)\gamma_{-}:=\gamma(a) and γ+:=γ(b)\gamma_{+}:=\gamma(b). For such an interval II\subset\mathbb{R} we write It={sI:st}I_{\leq t}=\{s\in I:s\leq t\} and It={sI:st}I_{\geq t}=\{s\in I:s\geq t\}.

Definition 1.1.

For a constant A1A\geq 1 and a closed interval II\subset\mathbb{R}, a curve γ:IΩ\gamma:I\rightarrow\Omega is AA-uniform if

(1.1) (γ)Ad(γ,γ+),\ell(\gamma)\leq Ad(\gamma_{-},\gamma_{+}),

and if for every tIt\in I we have

(1.2) min{(γ|It),(γ|It)}AdΩ(γ(t)).\min\{\ell(\gamma|_{I\leq t}),\ell(\gamma|_{I\geq t})\}\leq Ad_{\Omega}(\gamma(t)).

We say that the metric space Ω\Omega is AA-uniform if any two points in Ω\Omega can be joined by an AA-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 AA-uniform curves minimize the distance between their endpoints up to the multiplicative constant AA. The second requirement (1.2) implies that if we cut γ\gamma at any point γ(t)\gamma(t) then at least one of the two subcurves γ|It\gamma|_{I_{\leq t}} or γ|It\gamma|_{I_{\geq t}} must have length controlled by the distance dΩ(γ(t))d_{\Omega}(\gamma(t)) of γ(t)\gamma(t) to Ω\partial\Omega. We note that it is easily verified from the definitions that the property of a curve γ\gamma being AA-uniform is independent of the choice of parametrization of γ\gamma. For the purpose of formulating our theorems it is convenient to extend the definition of AA-uniform curves to allow for arbitrary subintervals II\subset\mathbb{R} and to allow the possibility (γ)=\ell(\gamma)=\infty; as this extension is somewhat technical we refer to Definition 4.16 for the exact details.

Remark 1.2.

The definition of uniform metric spaces advanced in [4] also requires local compactness. We follow Väisälä [20] in dropping the local compactness requirement, as the output of our uniformization procedure need not be locally compact in many cases of interest.

For a continuous function ρ:X(0,)\rho:X\rightarrow(0,\infty) we write

ρ(γ)=γρ𝑑s,\ell_{\rho}(\gamma)=\int_{\gamma}\rho\,ds,

for the line integral of ρ\rho along γ\gamma. 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 ρ\rho as a density on XX. The following definition plays a key role in the statement of our main theorems.

Definition 1.3.

Let (X,d)(X,d) be a geodesic metric space and let ρ:X(0,)\rho:X\rightarrow(0,\infty) be a density on XX. The conformal deformation of XX with conformal factor ρ\rho is the metric space Xρ=(X,dρ)X_{\rho}=(X,d_{\rho}) with metric

dρ(x,y)=infρ(γ),d_{\rho}(x,y)=\inf\ell_{\rho}(\gamma),

with the infimum taken over all curves γ\gamma joining xx to yy. We say that the density ρ\rho is a Gehring-Hayman density (abbreviated as a GH-density) if there is a constant M1M\geq 1 such that for any x,yXx,y\in X and any geodesic γ\gamma joining xx to yy we have

(1.3) ρ(γ)Mdρ(x,y).\ell_{\rho}(\gamma)\leq Md_{\rho}(x,y).

We will refer to the inequality (1.3) as the GH-inequality and will sometimes refer to the constant MM as the GHGH-constant. The terminology here is inspired by the work of Gehring-Hayman [16], which shows that in a simply connected hyperbolic domain Ω\Omega 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 ρ\rho is given by the conformal change of metric relating the Euclidean metric on Ω\Omega to the hyperbolic metric. Note that if XX is a tree then the GH-inequality holds for any density ρ\rho with M=1M=1 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 XX is Gromov hyperbolic if there is a δ0\delta\geq 0 such that all geodesic triangles are δ\delta-thin, meaning that for any geodesic triangle Δ\Delta each edge of Δ\Delta is contained in a δ\delta-neighborhood of the other two edges of Δ\Delta. In this case we will also say that XX is δ\delta-hyperbolic. We write GX\partial_{G}X for the Gromov boundary of XX, to be defined in Section 2; for now we note that a geodesic ray γ:[0,)X\gamma:[0,\infty)\rightarrow X can always be identified with an equivalence class [γ]GX[\gamma]\in\partial_{G}X, but in general not every point in GX\partial_{G}X can be realized in this way.

We consider a complete geodesic δ\delta-hyperbolic space XX and a geodesic ray γ:[0,)X\gamma:[0,\infty)\rightarrow X. The Busemann function bγ:Xb_{\gamma}:X\rightarrow\mathbb{R} associated to γ\gamma is defined by the limit

(1.4) bγ(x)=limtd(γ(t),x)t.b_{\gamma}(x)=\lim_{t\rightarrow\infty}d(\gamma(t),x)-t.

Using the triangle inequality and the fact that d(γ(t),γ(0))=td(\gamma(t),\gamma(0))=t, it’s easy to check that the right side is nonincreasing in tt and bounded below by d(γ(0),x)-d(\gamma(0),x), so this limit exists. It’s also easily verified that bγb_{\gamma} is 1-Lipschitz, thus in particular is continuous. As is customary when considering Busemann functions, we will refer to any translate b=bγ+sb=b_{\gamma}+s of bγb_{\gamma} for a constant ss\in\mathbb{R} as a Busemann function as well. We write

(1.5) (X)={bγ+s:γ a geodesic ray in Xs},\mathcal{B}(X)=\{b_{\gamma}+s:\text{$\gamma$ a geodesic ray in $X$, $s\in\mathbb{R}$}\},

for the set of all Busemann functions on XX. Given a Busemann function b=bγ+sb=b_{\gamma}+s we will write ωb=[γ]GX\omega_{b}=[\gamma]\in\partial_{G}X for the point in the Gromov boundary determined by the geodesic ray γ\gamma. We will refer to ωb\omega_{b} as the basepoint of bb and say that bb is based at ωb\omega_{b}.

To state our theorems in their appropriate generality it is useful to augment the set (X)\mathcal{B}(X) with the distance functions on XX: for zXz\in X we write bz(x)=d(x,z)b_{z}(x)=d(x,z) for the distance function to zz and write

(1.6) 𝒟(X)={bz+s:zXs},\mathcal{D}(X)=\{b_{z}+s:\text{$z\in X$, $s\in\mathbb{R}$}\},

for the set of all translates of distance functions on XX. We then write ^(X)=(X)𝒟(X)\hat{\mathcal{B}}(X)=\mathcal{B}(X)\cup\mathcal{D}(X). In the case b=bz+sb=b_{z}+s we write ωb=z\omega_{b}=z and refer to ωb\omega_{b} as the basepoint of bb as well. The defining formula (1.4) for Busemann functions shows that any b(X)b\in\mathcal{B}(X) can be realized as a pointwise limit of functions bt𝒟(X)b_{t}\in\mathcal{D}(X) defined by bt(x)=d(γ(t),x)tb_{t}(x)=d(\gamma(t),x)-t.

Given b^(X)b\in\hat{\mathcal{B}}(X) and ε>0\varepsilon>0 we define a density ρε,b\rho_{\varepsilon,b} on XX by

ρε,b(x)=eεb(x).\rho_{\varepsilon,b}(x)=e^{-\varepsilon b(x)}.

We write Xε,b=(X,dε,b)X_{\varepsilon,b}=(X,d_{\varepsilon,b}) for the conformal deformation of XX with conformal factor ρε,b\rho_{\varepsilon,b}. In the theorem below we will be assuming that XX is KK-roughly starlike from the basepoint ωb\omega_{b} of bb. This is a technical condition on geodesics starting from ωb\omega_{b} 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 KK-rough starlikeness condition will be satisfied with K=12K=\frac{1}{2} in our application of Theorem 1.4 to hyperbolic fillings in Theorem 1.12.

Theorem 1.4.

Let XX be a complete geodesic δ\delta-hyperbolic space and let b^(X)b\in\hat{\mathcal{B}}(X) be given. We suppose that XX is KK-roughly starlike from ωb\omega_{b} and that ε>0\varepsilon>0 is given such that ρε,b\rho_{\varepsilon,b} is a GH-density with constant MM.

Then geodesics in XX are AA-uniform curves in Xε,bX_{\varepsilon,b}, with A=A(δ,K,ε,M)A=A(\delta,K,\varepsilon,M). Consequently Xε,bX_{\varepsilon,b} is an AA-uniform metric space. Furthermore Xε,bX_{\varepsilon,b} is bounded if and only if b𝒟(X)b\in\mathcal{D}(X).

In this statement and all subsequent ones the notation A=A(δ,K,ε,M)A=A(\delta,K,\varepsilon,M) 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 AA-uniform curves that is necessary to cover the case of an arbitrary geodesic in XX; Definition 1.1 only covers the case of geodesics defined on closed intervals. The claim that Xε,bX_{\varepsilon,b} is bounded if and only if b𝒟(X)b\in\mathcal{D}(X) does not require either the rough starlikeness hypothesis or the assumption that ρε,b\rho_{\varepsilon,b} is a GH-density, see Proposition 4.4.

We describe the motivating example for Theorem 1.4 in the case b(X)b\in\mathcal{B}(X) below.

Example 1.5.

Let 𝕌2={(x,y)2:y>0}\mathbb{U}^{2}=\{(x,y)\in\mathbb{R}^{2}:y>0\} be the upper half space in 2\mathbb{R}^{2} equipped with the Euclidean metric, which is easily seen to be a uniform metric space. Let 2\mathbb{H}^{2} denote the upper half plane model of the hyperbolic plane, which is 𝕌2\mathbb{U}^{2} equipped with the Riemannian metric ds2=dx2+dy2y2ds^{2}=\frac{dx^{2}+dy^{2}}{y^{2}}. Define γ:[0,)2\gamma:[0,\infty)\rightarrow\mathbb{H}^{2} by γ(t)=(0,et)\gamma(t)=(0,e^{t}). Then γ\gamma is a geodesic ray in 2\mathbb{H}^{2}.

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 bγ(x,y)=logyb_{\gamma}(x,y)=-\log y. Setting ε=1\varepsilon=1, the density ρ1,bγ\rho_{1,b_{\gamma}} is thus simply given by ρ1,bγ(x,y)=y\rho_{1,b_{\gamma}}(x,y)=y. Therefore the uniformized metric space 1,bγ2\mathbb{H}^{2}_{1,b_{\gamma}} is isometric to 𝕌2\mathbb{U}^{2}. We also remark that the GH-inequality (1.3) for the density ρ1,bγ\rho_{1,b_{\gamma}} can easily be verified using the standard representation of geodesics in the upper half-plane model for 2\mathbb{H}^{2} as subsegments of semicircles or vertical lines orthogonal to the horizontal line {y=0}\{y=0\} in 2\mathbb{R}^{2}.

A metric space (X,d)(X,d) is proper if its closed balls are compact. In the case b𝒟(X)b\in\mathcal{D}(X), Theorem 1.4 generalizes [4, Proposition 4.5] as it does not require XX to be proper and allows for a potentially larger range of values for the parameter ε\varepsilon; this larger range will be relevant to Theorem 1.10.

For a δ\delta-hyperbolic space XX and a point ωGX\omega\in\partial_{G}X we write ωX=GX\{ω}\partial_{\omega}X=\partial_{G}X\backslash\{\omega\} for the complement of ω\omega in the Gromov boundary of XX. We will refer to ωX\partial_{\omega}X as the Gromov boundary relative to ω\omega for reasons that will be explained prior to Proposition 2.8. We formally extend this definition to ωX\omega\in X by defining ωX=GX\partial_{\omega}X=\partial_{G}X; in this case we will still refer to ωX\partial_{\omega}X as the Gromov boundary relative to ω\omega, with the understanding that this simply coincides with the standard Gromov boundary for ωX\omega\in X. As part of the proof of Theorem 1.4, we will show that there is a canonical identification φε,b:ωbXXε,b\varphi_{\varepsilon,b}:\partial_{\omega_{b}}X\rightarrow\partial X_{\varepsilon,b} between the Gromov boundary of XX relative to ωb\omega_{b} and the boundary of Xε,bX_{\varepsilon,b} in its completion. The most important property of this identification is summarized in Theorem 1.6 below.

A function b^(X)b\in\hat{\mathcal{B}}(X) can be used to define a natural class of metrics on ωbX\partial_{\omega_{b}}X known as visual metrics based at ωb\omega_{b} (see [12, Chapter 3] as well as Section 2.3). These visual metrics have an associated parameter ε>0\varepsilon>0 and a comparison constant LL to a specific model quasi-metric θε,b\theta_{\varepsilon,b} on ωbX\partial_{\omega_{b}}X defined in (2.13). We continue to write dε,bd_{\varepsilon,b} for the canonical extension of the metric on the uniformization Xε,bX_{\varepsilon,b} to its completion X¯ε,b\bar{X}_{\varepsilon,b}.

Theorem 1.6.

Let XX be a complete geodesic δ\delta-hyperbolic space and let b^(X)b\in\hat{\mathcal{B}}(X) be such that XX is KK-roughly starlike from the basepoint ω\omega of bb. Let ε>0\varepsilon>0 be given such that ρε,b\rho_{\varepsilon,b} is a GH-density with constant MM. Then there is a canonical identification φε,b:ωXXε,b\varphi_{\varepsilon,b}:\partial_{\omega}X\rightarrow\partial X_{\varepsilon,b} under which the restriction of dε,bd_{\varepsilon,b} to Xε,b\partial X_{\varepsilon,b} defines a visual metric on ωX\partial_{\omega}X based at ω\omega with parameter ε\varepsilon and comparison constant L=L(δ,K,ε,M)L=L(\delta,K,\varepsilon,M).

For a precise description of the identification φε,b\varphi_{\varepsilon,b} we refer to the disucssion after (4.15).

Remark 1.7.

It is useful to allow some additional flexibility in the choice of function bb in Theorems 1.4 and 1.6. This flexibility will be used in Theorem 1.12. For a continuous function b:Xb:X\rightarrow\mathbb{R} and a constant κ0\kappa\geq 0 we write b^κ(X)b\in\hat{\mathcal{B}}_{\kappa}(X) if there is some b^(X)b^{\prime}\in\hat{\mathcal{B}}(X) such that |b(x)b(x)|κ|b(x)-b^{\prime}(x)|\leq\kappa for all xXx\in X; we write b𝒟κ(x)b\in\mathcal{D}_{\kappa}(x) if b𝒟(X)b^{\prime}\in\mathcal{D}(X) and bκ(x)b\in\mathcal{B}_{\kappa}(x) if b(X)b^{\prime}\in\mathcal{B}(X). We define the basepoint of bb to be the basepoint of bb^{\prime}, ωb=ωb\omega_{b}=\omega_{b^{\prime}}; while this definition may be ambiguous in the case b𝒟κ(X)b\in\mathcal{D}_{\kappa}(X), this ambiguity does not matter in the context of our theorems. Then Xε,bX_{\varepsilon,b} is eεκe^{\varepsilon\kappa}-biLipschitz to Xε,bX_{\varepsilon,b^{\prime}} via the identity map on XX and ρε,b\rho_{\varepsilon,b^{\prime}} will be a GH-density with constant e2εκMe^{2\varepsilon\kappa}M if ρε,b\rho_{\varepsilon,b} is a GH-density with constant MM. It then easily follows that Theorems 1.4 and 1.6 hold for b^κ(X)b\in\hat{\mathcal{B}}_{\kappa}(X) as well, with the uniformity parameter AA in Theorem 1.4 and the comparison constant LL in Theorem 1.6 depending additionally on κ\kappa.

Remark 1.8.

Since we do not assume that the Gromov hyperbolic space XX 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 XX is geodesic is too strong [20, Remark 3.5]. The best one can assume is that XX is a length space, i.e., that the distance between two points of XX 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 ωb\omega_{b} of bb and the assumption that ρε,b\rho_{\varepsilon,b} is a GH-density on XX. 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 ρε,b\rho_{\varepsilon,b} 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 (X,d)(X,d) be a geodesic δ\delta-hyperbolic space. There is ε0=ε0(δ)>0\varepsilon_{0}=\varepsilon_{0}(\delta)>0 depending only on δ\delta such that if a density ρ:X(0,)\rho:X\rightarrow(0,\infty) satisfies for all x,yXx,y\in X and some fixed 0<εε00<\varepsilon\leq\varepsilon_{0},

(1.7) eεd(x,y)ρ(x)ρ(y)eεd(x,y),e^{-\varepsilon d(x,y)}\leq\frac{\rho(x)}{\rho(y)}\leq e^{\varepsilon d(x,y)},

then ρ\rho is a GH-density with constant M=20M=20.

The inequality (1.7) is satisfied for ρε,b\rho_{\varepsilon,b} for any b^(X)b\in\hat{\mathcal{B}}(X) since all functions in ^(X)\hat{\mathcal{B}}(X) are 11-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 ε\varepsilon then it is always possible to assume that ρε,b\rho_{\varepsilon,b} is a GH-density with constant M=20M=20 by taking ε\varepsilon sufficiently small.

In general one wants to establish that ρε,b\rho_{\varepsilon,b} is a GH-density for as large a value of ε\varepsilon as possible, as this property is then inherited for smaller values of ε\varepsilon by Proposition 4.12. This is particularly important for applications in which there is a preferred visual metric on ωX\partial_{\omega}X, such as Theorems 1.10 and 1.12 below.

CAT(1)(-1) spaces are geodesic metric spaces in which geodesic triangles are thinner than corresponding comparison geodesic triangles in the hyperbolic plane 2\mathbb{H}^{2}. 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(1)(-1) 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 1\leq-1. A CAT(1)(-1) space is δ\delta-hyperbolic with the same hyperbolicity constant δ=δ(2)\delta=\delta(\mathbb{H}^{2}) as the hyperbolic plane, for which the optimal constant can be computed explicitly to be δ=log(1+2)\delta=\log(1+\sqrt{2}).

For a CAT(1)(-1) space XX and a function b^(X)b\in\hat{\mathcal{B}}(X) with basepoint ω\omega the model quasi-metric θ1,b\theta_{1,b} on ωX\partial_{\omega}X in fact defines a distinguished choice of visual metric on ωX\partial_{\omega}X with parameter ε=1\varepsilon=1. This metric is known as the Bourdon metric on ωX=X\partial_{\omega}X=\partial X when b𝒟(X)b\in\mathcal{D}(X) and the Hamenstädt metric on ωX\partial_{\omega}X when b(X)b\in\mathcal{B}(X). 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(1)(-1) spaces at the special value ε=1\varepsilon=1.

Theorem 1.10.

Let XX be a complete CAT(1)(-1) space and let b^(X)b\in\hat{\mathcal{B}}(X) be given with basepoint ω\omega. Then there is a universal constant M1M\geq 1 such that ρ1,b\rho_{1,b} is a GH-density with constant MM. If, furthermore, XX is KK-roughly starlike from the basepoint ω\omega of bb for some K0K\geq 0 then the conclusions of Theorems 1.4 and 1.6 hold for X1,bX_{1,b} with constants A=A(K)A=A(K) and L=L(K)L=L(K) depending only on KK. In particular the restriction of d1,bd_{1,b} to X1,b\partial X_{1,b} is LL-biLipschitz to θ1,b\theta_{1,b}.

The constant MM in Theorem 1.10 is universal in the sense that it is the same for any CAT(1)(-1) space XX and any b^(X)b\in\hat{\mathcal{B}}(X). The dependence of AA on KK can be removed when b𝒟(X)b\in\mathcal{D}(X) 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 X1,b\partial X_{1,b} of the uniformization has a canonical biLipschitz identification with the Gromov boundary ωX\partial_{\omega}X relative to ω\omega equipped with the distinguished visual metric θ1,b\theta_{1,b}.

Remark 1.11.

Theorem 1.6 produces an obstruction for ρε,b\rho_{\varepsilon,b} to be a GH-density: the Gromov boundary ωX\partial_{\omega}X must admit a visual metric based at the basepoint ω\omega of bb with parameter ε\varepsilon. In the case b=bzb=b_{z} for some zXz\in X (i.e., if bb is a distance function) then this shows in particular that εKu(X)\varepsilon\leq-K_{u}(X), where Ku(X)K_{u}(X) is the asymptotic upper curvature bound defined by Bonk and Foertsch (see [5, Theorem 1.5]). In the case of the nn-dimensional hyperbolic space n\mathbb{H}^{n} of constant negative curvature 1-1 (n2n\geq 2) we have Ku(n)=1K_{u}(\mathbb{H}^{n})=1 by [5, Proposition 1.4]. Hence ρε,b\rho_{\varepsilon,b} cannot be a GH-density for any ε>1\varepsilon>1 when b𝒟(n)b\in\mathcal{D}(\mathbb{H}^{n}). This shows in particular that the value ε=1\varepsilon=1 for ρ1,b\rho_{1,b} to be a GH-density in Theorem 1.10 is sharp in certain cases. We can in fact extend these conclusions to observe that ρε,b\rho_{\varepsilon,b} cannot be a GH-density for any ε>1\varepsilon>1 when b(n)b\in\mathcal{B}(\mathbb{H}^{n}) as well by observing that, for a fixed ωn\omega\in\partial\mathbb{H}^{n}, any visual metric based at ω\omega on ωn\partial_{\omega}\mathbb{H}^{n} with parameter ε>1\varepsilon>1 would give rise to a visual metric on n\partial\mathbb{H}^{n} with parameter ε>1\varepsilon>1 by a Möbius inversion of ωn\partial_{\omega}\mathbb{H}^{n} centered at the point ω\omega.

We will also apply our uniformization results to hyperbolic fillings of an arbitrary metric space (Z,d)(Z,d). 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 α>1\alpha>1 and τ>1\tau>1. For an r>0r>0 we say that a subset SZS\subset Z is rr-separated if for each x,ySx,y\in S we have d(x,y)rd(x,y)\geq r. Given a parameter α>1\alpha>1, we choose for each nn\in\mathbb{Z} a maximal αn\alpha^{-n}-separated subset SnS_{n} of ZZ. For nn\in\mathbb{Z} we write Vn={(z,n):zSn}V_{n}=\{(z,n):z\in S_{n}\} and set V=nVnV=\bigcup_{n\in\mathbb{Z}}V_{n}. The set VV will serve as the vertex set for XX. We define the height function h:Vh:V\rightarrow\mathbb{Z} on this vertex set by h(v)=nh(v)=n for v=(z,n)Vnv=(z,n)\in V_{n}.

We associate to each vertex v=(z,n)Vv=(z,n)\in V the ball B(v)=B(z,ταn)B(v)=B(z,\tau\alpha^{-n}) of radius ταn\tau\alpha^{-n} centered at zz. We place an edge between vertices v,wVv,w\in V if and only if their heights satisfy |h(v)h(w)|1|h(v)-h(w)|\leq 1 and their associated balls satisfy B(v)B(w)B(v)\cap B(w)\neq\emptyset. We write XX for the resulting graph and call this a hyperbolic filling of ZZ. If τ\tau is sufficiently large (see inequality (5.1)) then XX will be a connected graph by Proposition 5.5. We make XX into a geodesic metric space by declaring all edges to have unit length. We extend the height function hh to a 11-Lipschitz function h:Xh:X\rightarrow\mathbb{R} by linearly interpolating the values of hh from the vertices to the edges of XX.

As a metric space XX is δ\delta-hyperbolic with δ=δ(α,τ)\delta=\delta(\alpha,\tau) depending only on the parameters α\alpha and τ\tau. There is a distinguished point ωGX\omega\in\partial_{G}X in the Gromov boundary that can be thought of as an ideal point at infinity for ZZ. We have an identification ωXZ¯\partial_{\omega}X\cong\bar{Z} of the Gromov boundary relative to ω\omega with the completion Z¯\bar{Z} of ZZ. Under this identification the extension of the metric dd to Z¯\bar{Z} defines a visual metric on ωX\partial_{\omega}X with parameter ε=logα\varepsilon=\log\alpha. All of the results mentioned here are proved in Section 5.

We define a density ρ\rho on XX by ρ(x)=αh(x)\rho(x)=\alpha^{-h(x)} for xXx\in X. We write XρX_{\rho} for the conformal deformation of XX with conformal factor ρ\rho. By Lemma 5.12 there is a Busemann function bb based at ω\omega such that |h(x)b(x)|3|h(x)-b(x)|\leq 3 for all xXx\in X and therefore h3(X)h\in\mathcal{B}_{3}(X) in the notation of Remark 1.7. For such a Busemann function b(X)b\in\mathcal{B}(X) we have that the density ρ\rho is uniformly comparable to the density ρε,b\rho_{\varepsilon,b} with ε=logα\varepsilon=\log\alpha.

Theorem 1.12.

Let ZZ be a metric space and let XX be a hyperbolic filling of ZZ with parameters α>1\alpha>1 and τ>min{3,α/(α1)}\tau>\min\{3,\alpha/(\alpha-1)\}. Then XX is 12\frac{1}{2}-roughly starlike from ω\omega and ρ\rho is a GH-density with constant M=M(α,τ)M=M(\alpha,\tau).

Thus the conclusions of Theorems 1.4 and 1.6 hold for XρX_{\rho}. In particular we have a canonical LL-biLipschitz identification of Xρ\partial X_{\rho} and Z¯\bar{Z}, with L=L(α,τ)L=L(\alpha,\tau).

We compare our results to those of [3] in Remark 6.4. Another notable predecessor to Theorem 1.12 in the case that ZZ 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.

We are very grateful to Nageswari Shanmugalingam for providing multiple drafts of the work [3] on which a significant part of this paper is based. We also thank Tushar Das for making us aware of the results of [14] that are used to prove Theorem 1.10.

2. Hyperbolic metric spaces

2.1. Definitions

Let XX be a set and let ff, gg be real-valued functions defined on XX. For c0c\geq 0 we will write fcgf\doteq_{c}g if

|f(x)g(x)|c,|f(x)-g(x)|\leq c,

for all xXx\in X. If the exact value of the constant cc is not important or implied by context we will often just write fgf\doteq g. We will sometimes refer to the relation fgf\doteq g as a rough equality between ff and gg.

If C1C\geq 1 and ff and gg both take values in (0,)(0,\infty), we will write fCgf\asymp_{C}g if

C1g(x)f(x)Cg(x).C^{-1}g(x)\leq f(x)\leq Cg(x).

We will similarly write fgf\asymp g if the value of CC is not important or implied by context. Note that if fcgf\doteq_{c}g then efecege^{f}\asymp_{e^{c}}e^{g}, and similarly if fCgf\asymp_{C}g then logflogClogg\log f\doteq_{\log C}\log g. We will stick to a convention of using lowercase cc for additive constants and uppercase CC for multiplicative constants. When this additive constant cc is determined by other parameters δ\delta, KK, etc. under discussion we will write c=c(δ,K)c=c(\delta,K), while continuing to use the shorthand cc where it is not ambiguous (and the same for multiplicative constants CC).

For a metric space (X,d)(X,d) we write B(x,r)={y:d(x,y)<r}B(x,r)=\{y:d(x,y)<r\} for the open ball of radius r>0r>0 centered at xx. A map f:(X,d)(X,d)f:(X,d)\rightarrow(X^{\prime},d^{\prime}) between metric spaces XX and XX^{\prime} is isometric if d(f(x),f(y))=d(x,y)d^{\prime}(f(x),f(y))=d(x,y) for xx, yXy\in X. If furthermore ff is surjective then we say that it is an isometry and that XX and XX^{\prime} are isometric. For a constant c0c\geq 0 a map f:XXf:X\rightarrow X^{\prime} is defined to be cc-roughly isometric if d(f(x),f(y))cd(x,y)d^{\prime}(f(x),f(y))\doteq_{c}d(x,y). The map ff is LL-Lipschitz for a constant L0L\geq 0 if d(f(x),f(y))Ld(x,y)d^{\prime}(f(x),f(y))\leq Ld(x,y), and it is LL-biLipschitz for a constant L1L\geq 1 if d(f(x),f(y))Ld(x,y)d^{\prime}(f(x),f(y))\asymp_{L}d(x,y). As usual we will not mention the exact value of the constants if they are unimportant.

When dealing with Gromov hyperbolic spaces XX in this paper we will use the generic distance notation |xy|:=d(x,y)|xy|:=d(x,y) for the distance between xx and yy in XX, except for cases where this could cause confusion. We will often use the generic notation xyxy for a geodesic connecting two points x,yXx,y\in X, 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 Δ\Delta in XX is a collection of three points x,y,zXx,y,z\in X together with geodesics xyxy, xzxz, and yzyz joining these points, which we will refer to as the edges of Δ\Delta. We will also alternatively write xyz=Δxyz=\Delta for a geodesic triangle with vertices xx, yy and zz.

For x,y,zXx,y,z\in X the Gromov product of xx and yy based at zz is defined by

(2.1) (x|y)z=12(|xz|+|yz||xy|).(x|y)_{z}=\frac{1}{2}(|xz|+|yz|-|xy|).

We note the basepoint change inequality for x,y,z,pXx,y,z,p\in X,

(2.2) |(x|y)z(x|y)p||zp|,|(x|y)_{z}-(x|y)_{p}|\leq|zp|,

which follows from the triangle inequality.

By [17, Chapitre 2, Proposition 21] we have two key consequences of δ\delta-hyperbolicity for a metric space XX regarding Gromov products. The first is that for every x,y,z,pXx,y,z,p\in X we have

(2.3) (x|z)pmin{(x|y)p,(y|z)p}4δ.(x|z)_{p}\geq\min\{(x|y)_{p},(y|z)_{p}\}-4\delta.

We refer to (2.3) as the 4δ4\delta-inequality.

The second is that for any geodesic triangle xyzxyz in XX we have that if pxyp\in xy, qxzq\in xz are points with |xp|=|xq|(y|z)x|xp|=|xq|\leq(y|z)_{x} then |pq|4δ|pq|\leq 4\delta. Here xyxy and xzxz are referring to the corresponding geodesics in the triangle Δ\Delta. We will refer to this as the 4δ4\delta-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 44. We note that the definition using inequality (2.3) does not use the fact that XX is geodesic, and is therefore used as a definition of δ\delta-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 δ\delta-hyperbolicity (with δ\delta in place of 4δ4\delta). Wherever necessary we have multiplied the constants used in their results by 44 in order to account for this discrepancy.

Let XX be a geodesic Gromov hyperbolic space and fix pXp\in X. A sequence {xn}\{x_{n}\} converges to infinity if we have (xn|xm)p(x_{n}|x_{m})_{p}\rightarrow\infty as m,nm,n\rightarrow\infty. The Gromov boundary GX\partial_{G}X of a Gromov hyperbolic space XX is defined to be the set of all equivalence classes of sequences {xn}X\{x_{n}\}\subset X converging to infinity, with the equivalence relation {xn}{yn}\{x_{n}\}\sim\{y_{n}\} if (xn|yn)p(x_{n}|y_{n})_{p}\rightarrow\infty as nn\rightarrow\infty. Inequality (2.2) shows that these notions do not depend on the choice of basepoint pp.

A second boundary that we can associate to XX is the geodesic boundary gX\partial^{g}X, which is defined as equivalence classes of geodesic rays γ:[0,)X\gamma:[0,\infty)\rightarrow X, with two geodesic rays γ\gamma and σ\sigma being equivalent if there is a constant c0c\geq 0 such that |γ(t)σ(t)|c|\gamma(t)\sigma(t)|\leq c for t0t\geq 0. There is a natural inclusion gXGX\partial^{g}X\subseteq\partial_{G}X given by sending a geodesic ray γ\gamma to the sequence {γ(n)}n\{\gamma(n)\}_{n\in\mathbb{N}}. This inclusion need not be surjective in general. However, it is always surjective if XX is proper, meaning that closed balls in XX are compact.

For a point ωGX\omega\in\partial_{G}X and a sequence {xn}\{x_{n}\} converging to infinity we will write {xn}ω\{x_{n}\}\in\omega or xnωx_{n}\rightarrow\omega if {xn}\{x_{n}\} belongs to the equivalence class of ω\omega. For a geodesic ray γ:[a,)X\gamma:[a,\infty)\rightarrow X, aa\in\mathbb{R}, and a point ωGX\omega\in\partial_{G}X we will write γω\gamma\in\omega if {γ(n)}naω\{\gamma(n)\}_{n\geq a}\in\omega, nn\in\mathbb{N}. We will sometimes also consider geodesic rays γ:(,a]X\gamma:(-\infty,a]\rightarrow X with a reversely oriented parametrization, for which we write γω\gamma\in\omega if {γ(n)}naω\{\gamma(-n)\}_{n\geq-a}\in\omega.

For the rest of this paper we will be using the standard notation X:=GX\partial X:=\partial_{G}X for the Gromov boundary of a Gromov hyperbolic space XX. While this notation does conflict with the notation Ω=Ω¯\Ω\partial\Omega=\bar{\Omega}\backslash\Omega 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 xXx\in X and a point ξX\xi\in\partial X we will often write xξx\xi for a geodesic ray γ:[0,)X\gamma:[0,\infty)\rightarrow X with γ(0)=x\gamma(0)=x and γξ\gamma\in\xi, provided such a geodesic ray exists. Similarly, for ζ,ξX\zeta,\xi\in\partial X we will write ζξ\zeta\xi for a geodesic line γ:X\gamma:\mathbb{R}\rightarrow X with γ|(,0]ζ\gamma|_{(-\infty,0]}\in\zeta and γ|[0,)ξ\gamma|_{[0,\infty)}\in\xi, provided such a geodesic line exists. Such geodesic lines and rays always exist when XX is proper, but not necessarily in general. We extend the definition of geodesic triangles Δ\Delta in XX to allow for vertices in X\partial X: a geodesic triangle xyz=Δxyz=\Delta in XX is a collection of three points x,y,zXXx,y,z\in X\cup\partial X together with geodesics xyxy, xzxz, yzyz connecting them in the sense described above.

Remark 2.1.

It is easy to see from the definitions that there is no geodesic γ:X\gamma:\mathbb{R}\rightarrow X such that γ|[0,)\gamma|_{[0,\infty)} and γ|(,0]\gamma|_{(-\infty,0]} belong to the same equivalence class in the Gromov boundary X\partial X. Hence, for a geodesic triangle Δ\Delta, all vertices of Δ\Delta on X\partial X must be distinct.

Gromov products based at points pXp\in X can be extended to points of X\partial X by defining the Gromov product of equivalence classes ξ\xi, ζX\zeta\in\partial X based at pp to be

(ξ|ζ)p=inflim infn(xn|yn)p,(\xi|\zeta)_{p}=\inf\liminf_{n\rightarrow\infty}(x_{n}|y_{n})_{p},

with the infimum taken over all sequences {xn}ξ\{x_{n}\}\in\xi, {yn}ζ\{y_{n}\}\in\zeta. By [12, Lemma 2.2.2], if XX is δ\delta-hyperbolic then for any choices of sequences {xn}ξ\{x_{n}\}\in\xi, {yn}ζ\{y_{n}\}\in\zeta we have

(2.4) (ξ|ζ)plim infn(xn|yn)plim supn(xn|yn)p(ξ|ζ)p+8δ.(\xi|\zeta)_{p}\leq\liminf_{n\rightarrow\infty}(x_{n}|y_{n})_{p}\leq\limsup_{n\rightarrow\infty}(x_{n}|y_{n})_{p}\leq(\xi|\zeta)_{p}+8\delta.

We also have the 4δ4\delta-inequality for ξ\xi, ζ\zeta, ωX\omega\in\partial X and pXp\in X,

(2.5) (ξ|ω)pmin{(ξ|ζ)p,(ζ|ω)p}4δ.(\xi|\omega)_{p}\geq\min\{(\xi|\zeta)_{p},(\zeta|\omega)_{p}\}-4\delta.

For xXx\in X, ξX\xi\in\partial X the Gromov product is defined analogously as

(x|ξ)p=inflim infn(x|xn)p,(x|\xi)_{p}=\inf\liminf_{n\rightarrow\infty}(x|x_{n})_{p},

with the infimum taken over {xn}ξ\{x_{n}\}\in\xi, and the analogues of (2.4) and (2.5) hold as well.

We next observe that geodesic triangles Δ\Delta with vertices in XXX\cup\partial X are 10δ10\delta-thin, in the precise sense that if uΔu\in\Delta is any given point then there is a point vΔv\in\Delta satisfying |uv|10δ|uv|\leq 10\delta that does not belong to the same edge of Δ\Delta as uu. When XX is proper this can be easily deduced from the δ\delta-thin triangles property for triangles in XX by a limiting argument. Without the properness hypothesis this result can also be obtained with a larger thinness constant 200δ200\delta as a consequence of work of Väisälä [21, Theorem 6.24]; we note that he uses (2.3) as the definition of δ\delta-hyperbolicity so we have to multiply the constant he obtains by 44. 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 10δ10\delta-thinness here.

Lemma 2.2.

Let Δ\Delta be a geodesic triangle in XX with vertices in XXX\cup\partial X. Then Δ\Delta is 10δ10\delta-thin.

Proof.

Let x,y,zXXx,y,z\in X\cup\partial X be the vertices of Δ\Delta. Let uΔu\in\Delta be given. Since XX has δ\delta-thin triangles, we may assume that Δ\Delta has at least one vertex on X\partial X. We first consider the case in which Δ\Delta has exactly one vertex on X\partial X, which by relabeling we can assume is zz. We first assume that uxyu\in xy. Let {zn}xz\{z_{n}\}\subset xz and {zn}yz\{z_{n}^{\prime}\}\subset yz be sequences such that znzz_{n}\rightarrow z and znzz_{n}^{\prime}\rightarrow z. For each nn we let Δn=xyzn\Delta_{n}=xyz_{n} be the geodesic triangle sharing the edge xyxy with Δ\Delta, having a second edge be the subsegment xznxz_{n} of xzxz, and having a third edge be any choice of geodesic yznyz_{n}. Then Δn\Delta_{n} is δ\delta-thin, so we have for each nn that either dist(u,xzn)δ\mathrm{dist}(u,xz_{n})\leq\delta or dist(u,yzn)δ\mathrm{dist}(u,yz_{n})\leq\delta (or both). In the first case we are done since xznxzxz_{n}\subset xz, so we can assume that dist(u,yzn)δ\mathrm{dist}(u,yz_{n})\leq\delta. Let vnyznv_{n}\in yz_{n} be such that |uvn|δ|uv_{n}|\leq\delta. Then |vny|δ+|uy||v_{n}y|\leq\delta+|uy|.

Since both znz_{n} and znz_{n}^{\prime} converge to zz, for sufficiently large nn we will have (zn|zn)yδ+|uy|(z_{n}|z_{n}^{\prime})_{y}\geq\delta+|uy|, which implies in particular that |zny||vny||z_{n}^{\prime}y|\geq|v_{n}y|. The 4δ4\delta-tripod condition applied to yy, znz_{n}, and znz_{n}^{\prime} then implies that if wnyznw_{n}\in yz_{n}^{\prime} is the unique point such that |ywn|=|yvn||yw_{n}|=|yv_{n}| then |vnwn|4δ|v_{n}w_{n}|\leq 4\delta, from which it follows that |uwn|5δ|uw_{n}|\leq 5\delta for all sufficiently large nn. Since wnyzw_{n}\in yz this completes the proof of this case.

The other cases are uxzu\in xz and uyzu\in yz. By symmetry it suffices to prove the case uxzu\in xz. We define the sequences {zn}\{z_{n}\} and {zn}\{z_{n}^{\prime}\} and the triangle Δn\Delta_{n} as before. As in the case uxyu\in xy we can assume that dist(u,yzn)δ\mathrm{dist}(u,yz_{n})\leq\delta for all nn, as otherwise by the δ\delta-thin triangles property we have dist(u,xy)δ\mathrm{dist}(u,xy)\leq\delta and we are done. We let vnyznv_{n}\in yz_{n} be such that |uvn|δ|uv_{n}|\leq\delta, note that |vny|δ+|uy||v_{n}y|\leq\delta+|uy| as before, and choose nn large enough that (zn|zn)yδ+|uy|(z_{n}|z_{n}^{\prime})_{y}\geq\delta+|uy|. As before the 4δ4\delta-tripod condition then supplies a point wnyznw_{n}\in yz_{n}^{\prime} such that |wnvn|4δ|w_{n}v_{n}|\leq 4\delta and we conclude that dist(u,yz)5δ\mathrm{dist}(u,yz)\leq 5\delta.

We can now handle the case in which potentially two or three vertices of Δ\Delta belong to X\partial X. By symmetry it suffices to show for a point uxyu\in xy that uu is 10δ10\delta-close to either xzxz or yzyz. Let {xn}xy\{x_{n}\}\subset xy and {yn}xy\{y_{n}\}\subset xy be sequences such that xnxx_{n}\rightarrow x and ynyy_{n}\rightarrow y; if xXx\in X then we set xn=xx_{n}=x for all nn and similarly if yYy\in Y then we set yn=yy_{n}=y for all nn. Let Δn=xnynz\Delta_{n}=x_{n}y_{n}z be a geodesic triangle with one edge the subsegment xnynx_{n}y_{n} of xyxy. Then Δn\Delta_{n} has at most one vertex zz on X\partial X. We conclude from the previous case that uu is 5δ5\delta-close to either xnzx_{n}z or ynzy_{n}z. By switching the roles of xx and yy if necessary, we can then assume that there is vxnzv\in x_{n}z such that |uv|5δ|uv|\leq 5\delta. If xXx\in X then xn=xx_{n}=x and we are done. Thus we can assume that xXx\in\partial X.

Fix any point wxzw\in xz and let xnwxx_{n}^{\prime}\in wx be defined such that |wxn|=|uxn||wx_{n}^{\prime}|=|ux_{n}|. Since the geodesic rays wxwx and uxux define the same point xx of the Gromov boundary, there is a constant c0c\geq 0 such that |xnxn|c|x_{n}x_{n}^{\prime}|\leq c for all nn. We apply the previous case again to a triangle Δn=xnxnz\Delta_{n}^{\prime}=x_{n}x_{n}^{\prime}z with edges the segment xnzx_{n}^{\prime}z, the segment xnzx_{n}z, and a choice of geodesic xnxnx_{n}x_{n}^{\prime}, obtaining that vv is 5δ5\delta-close to either xnzx_{n}^{\prime}z or xnxnx_{n}x_{n}^{\prime}. If vv is 5δ5\delta-close to xnxnx_{n}x_{n}^{\prime} for all nn then

|xnu||uv|+dist(v,xnxn)+|xnxn|10δ+c,|x_{n}u|\leq|uv|+\mathrm{dist}(v,x_{n}x_{n}^{\prime})+|x_{n}x_{n}^{\prime}|\leq 10\delta+c,

contradicting that |xnu||x_{n}u|\rightarrow\infty as nn\rightarrow\infty. We conclude that vv is 5δ5\delta-close to xnzxzx_{n}^{\prime}z\subset xz for all sufficiently large nn, which implies that dist(u,xz)10δ\mathrm{dist}(u,xz)\leq 10\delta as desired. ∎

We can also now formally define rough starlikeness from points of XXX\cup\partial X. We recall that for ωX\omega\in\partial X we write ωX=X\{ω}\partial_{\omega}X=\partial X\backslash\{\omega\} for the Gromov boundary of XX relative to ω\omega. The definition is slightly different for points of XX and points of X\partial X, so we handle these two cases separately.

Definition 2.3.

Let XX be a geodesic Gromov hyperbolic space. Let zXz\in X and K0K\geq 0 be given. We say that XX is KK-roughly starlike from zz if

  1. (1)

    For each xXx\in X there is a geodesic ray γ:[0,)X\gamma:[0,\infty)\rightarrow X such that γ(0)=z\gamma(0)=z and dist(x,γ)K\mathrm{dist}(x,\gamma)\leq K.

  2. (2)

    For each ξX\xi\in\partial X there is a geodesic ray γ:[0,)X\gamma:[0,\infty)\rightarrow X such that γ(0)=z\gamma(0)=z and γξ\gamma\in\xi.

For a point ωX\omega\in\partial X we say that XX is KK-roughly starlike from ω\omega if

  1. (1)

    For each xXx\in X there is a geodesic line γ:X\gamma:\mathbb{R}\rightarrow X such that dist(x,γ)K\mathrm{dist}(x,\gamma)\leq K and γ|(,0]ω\gamma|_{(-\infty,0]}\in\omega.

  2. (2)

    For each ξωX\xi\in\partial_{\omega}X there is a geodesic line γ:X\gamma:\mathbb{R}\rightarrow X such that γ|[0,)ξ\gamma|_{[0,\infty)}\in\xi and γ|(,0]ω\gamma|_{(-\infty,0]}\in\omega.

Part (2) of Definition 2.3 implies in both cases that gX=X\partial^{g}X=\partial X, 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 ωXX\omega\in X\cup\partial X when XX is proper, since in this case any two points of XXX\cup\partial X can be joined by a geodesic. We also remark that if X\partial X consists of a single point ω\omega then XX cannot be roughly starlike from ω\omega, since no geodesic line γ:X\gamma:\mathbb{R}\rightarrow X can exist in this case. Similarly if X\partial X is empty then XX 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 ωX\omega\in\partial X. 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 γω\gamma\in\omega, that bγb_{\gamma} is a Busemann function based at ω\omega in their sense. The definition of a Busemann function given there starts with the function

(2.6) bω,p(x)=(ω|p)x(ω|x)p,b_{\omega,p}(x)=(\omega|p)_{x}-(\omega|x)_{p},

for x,pXx,p\in X and ωX\omega\in\partial X and defines a Busemann function based at ω\omega to be any function b:Xb:X\rightarrow\mathbb{R} satisfying b8δbω,p+sb\doteq_{8\delta}b_{\omega,p}+s for some pXp\in X and ss\in\mathbb{R} (recall that we are multiplying all of their constants by 44 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 ωX\omega\in\partial X, let pXp\in X, and let γω\gamma\in\omega be a geodesic ray with γ(0)=p\gamma(0)=p. Then we have bω,p24δbγb_{\omega,p}\doteq_{24\delta}b_{\gamma}.

Proof.

By [12, Example 3.1.4] we have for all xXx\in X that

bω,p(x)8δ|xp|(ω|x)p.b_{\omega,p}(x)\doteq_{8\delta}|xp|-(\omega|x)_{p}.

By inequality (2.4) we have (γ(n)|x)p8δ(ω|x)p(\gamma(n)|x)_{p}\doteq_{8\delta}(\omega|x)_{p} for nn\in\mathbb{N} sufficiently large. Since p=γ(0)p=\gamma(0) we have

(γ(n)|x)p=12(n+|xp||γ(n)x|).(\gamma(n)|x)_{p}=\frac{1}{2}(n+|xp|-|\gamma(n)x|).

Then

(2.7) |xp|2(ω|x)p16δ|γ(n)x|n.|xp|-2(\omega|x)_{p}\doteq_{16\delta}|\gamma(n)x|-n.

Since the right side converges to bγ(x)b_{\gamma}(x) as nn\rightarrow\infty, the result follows. ∎

We recall our definition of a Busemann function from (1.5): a Busemann function b(X)b\in\mathcal{B}(X) is any function b:Xb:X\rightarrow\mathbb{R} such that b=bγ+sb=b_{\gamma}+s for some geodesic ray γ\gamma in XX and some ss\in\mathbb{R}. For such a Busemann function bb we let ω=ωb=[γ]\omega=\omega_{b}=[\gamma] denote its basepoint in X\partial X. Then by Lemma 2.4 bb is a Busemann function based at ω\omega in the sense of [12, Chapter 3] as well, provided that we use a cutoff of b24δbω,p+sb\doteq_{24\delta}b_{\omega,p}+s instead of the 8δ8\delta-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 ωX\omega\in\partial X and let γ,σ:[0,)X\gamma,\sigma:[0,\infty)\rightarrow X be geodesic rays with γ,σω\gamma,\sigma\in\omega. Then there is a constant ss\in\mathbb{R} such that bσ72δbγ+sb_{\sigma}\doteq_{72\delta}b_{\gamma}+s. The constant ss depends only on the starting points γ(0)\gamma(0) and σ(0)\sigma(0) of the rays and satisfies s=0s=0 if γ(0)=σ(0)\gamma(0)=\sigma(0).

Consequently if bb is any Busemann function based at ω\omega and σω\sigma\in\omega is any geodesic ray then there is a constant ss\in\mathbb{R} such that b72δbσ+sb\doteq_{72\delta}b_{\sigma}+s.

Proof.

By [12, Lemma 3.1.2], for each p,q,xXp,q,x\in X we have

bω,p(x)24δbω,q(x)+bω,q(p).b_{\omega,p}(x)\doteq_{24\delta}b_{\omega,q}(x)+b_{\omega,q}(p).

Setting p=γ(0)p=\gamma(0), q=σ(0)q=\sigma(0), and applying Lemma 2.4 gives

bγ(x)72δbσ(x)+bω,σ(0)(γ(0)).b_{\gamma}(x)\doteq_{72\delta}b_{\sigma}(x)+b_{\omega,\sigma(0)}(\gamma(0)).

This gives the first claim of the lemma with c=bω,σ(0)(γ(0))c=b_{\omega,\sigma(0)}(\gamma(0)). The claim that s=0s=0 if γ(0)=σ(0)\gamma(0)=\sigma(0) follows from the fact that bω,p(p)=0b_{\omega,p}(p)=0 for any pXp\in X. The final claim follows immediately since for any Busemann function bb based at ω\omega there is some geodesic ray γω\gamma\in\omega such that b=bγ+sb=b_{\gamma}+s^{\prime} for some ss^{\prime}\in\mathbb{R}. ∎

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 ωX\omega\in\partial X instead of ending there. The notation (,a](-\infty,a] below should be interpreted as (,a]=(-\infty,a]=\mathbb{R} when a=a=\infty.

Lemma 2.6.

Let bb be a Busemann function on XX based at ωX\omega\in\partial X. Let a{}a\in\mathbb{R}\cup\{\infty\} and let γ:(,a]X\gamma:(-\infty,a]\rightarrow X be a geodesic with γ(t)ω\gamma(t)\rightarrow\omega as tt\rightarrow-\infty.

  1. (1)

    For any s,t(,a]s,t\in(-\infty,a] (or any s,ts,t\in\mathbb{R} in the case a=a=\infty) we have

    (2.8) b(γ(t))b(γ(s))144δts.b(\gamma(t))-b(\gamma(s))\doteq_{144\delta}t-s.
  2. (2)

    For any constant uu\in\mathbb{R} there is an arclength reparametrization γ~:(,a~]X\tilde{\gamma}:(-\infty,\tilde{a}]\rightarrow X of γ\gamma such that b(γ~(t))144δt+ub(\tilde{\gamma}(t))\doteq_{144\delta}t+u for t(,a~]t\in(-\infty,\tilde{a}].

Proof.

Let s(,a]s\in(-\infty,a] be given and let σs:[sa,)X\sigma_{s}:[s-a,\infty)\rightarrow X be defined by σs(t)=γ(st)\sigma_{s}(t)=\gamma(s-t). It’s easily checked from the definition (1.4) that bσs(σs(t))=tb_{\sigma_{s}}(\sigma_{s}(t))=-t for t[sa,)t\in[s-a,\infty). Lemma 2.5 shows that there is a constant cc\in\mathbb{R} such that b72δbσs+cb\doteq_{72\delta}b_{\sigma_{s}}+c. It follows that for any t(,a]t\in(-\infty,a],

b(γ(t))b(γ(s))144δbσs(σs(st))bσs(σs(0))=ts,b(\gamma(t))-b(\gamma(s))\doteq_{144\delta}b_{\sigma_{s}}(\sigma_{s}(s-t))-b_{\sigma_{s}}(\sigma_{s}(0))=t-s,

for t[a,)t\in[-a,\infty). This proves (1).

For the second claim we fix an s(,a)s\in(-\infty,a) and define γ~(t)=γ(tb(γ(s))+s+u)\tilde{\gamma}(t)=\gamma(t-b(\gamma(s))+s+u) for t(,a~]t\in(-\infty,\tilde{a}], a~=as+b(γ(s))u\tilde{a}=a-s+b(\gamma(s))-u (if a=a=\infty we take a~=\tilde{a}=\infty). Then by (2.8),

b(γ~(t))\displaystyle b(\tilde{\gamma}(t)) =b(γ(tb(γ(s))+s+u))\displaystyle=b(\gamma(t-b(\gamma(s))+s+u))
144δ(tb(γ(s))+s+u)s+b(γ(s))\displaystyle\doteq_{144\delta}(t-b(\gamma(s))+s+u)-s+b(\gamma(s))
=t+u.\displaystyle=t+u.

For xx, yXy\in X and b^(X)b\in\hat{\mathcal{B}}(X) the Gromov product based at bb is defined by

(x|y)b=12(b(x)+b(y)|xy|).(x|y)_{b}=\frac{1}{2}(b(x)+b(y)-|xy|).

Since bb is 11-Lipschitz we have the useful inequality

(2.9) (x|y)bmin{b(x),b(y)}.(x|y)_{b}\leq\min\{b(x),b(y)\}.

For b𝒟(X)b\in\mathcal{D}(X) this notion essentially reduces to the standard Gromov product: if b(x)=|xp|+sb(x)=|xp|+s for some pXp\in X and ss\in\mathbb{R} then (x|y)b=(x|y)p+s(x|y)_{b}=(x|y)_{p}+s. 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 b(X)b\in\mathcal{B}(X).

Let b(X)b\in\mathcal{B}(X) and let ω=ωb\omega=\omega_{b} be its basepoint. The Gromov product based at bb is extended to X\partial X by, for (ξ,ζ)(ω,ω)(\xi,\zeta)\neq(\omega,\omega),

(ξ|ζ)b=inflim infn(xn|yn)b(\xi|\zeta)_{b}=\inf\liminf_{n\rightarrow\infty}(x_{n}|y_{n})_{b}

with the infimum taken over {xn}ξ\{x_{n}\}\in\xi, {yn}ζ\{y_{n}\}\in\zeta as before, and similarly for xXx\in X and ξX\xi\in\partial X we define

(x|ξ)b=inflim infn(x|xn)b,(x|\xi)_{b}=\inf\liminf_{n\rightarrow\infty}(x|x_{n})_{b},

with the infimum taken over {xn}ξ\{x_{n}\}\in\xi. The next lemma extends the 4δ4\delta-inequality to Gromov products based at bb. It follows from [12, Lemma 3.2.4]. Recall that we have multiplied their additive constants by a total of 1212 due to the differing definition of hyperbolicity and larger cutoff in defining Busemann functions; we then round up to 600δ600\delta afterward. The corresponding additive constant in [12, Lemma 3.2.4] below is 44δ44\delta.

Lemma 2.7.

Let bb be a Busemann function based at ωX\omega\in\partial X. Then

  1. (1)

    For any ξ\xi, ζX\{ω}\zeta\in\partial X\backslash\{\omega\} and any {xn}ξ\{x_{n}\}\in\xi, {yn}ζ\{y_{n}\}\in\zeta we have

    (ξ|ζ)blim infn(xn|yn)blim supn(xn|yn)b(ξ|ζ)b+600δ,(\xi|\zeta)_{b}\leq\liminf_{n\rightarrow\infty}(x_{n}|y_{n})_{b}\leq\limsup_{n\rightarrow\infty}(x_{n}|y_{n})_{b}\leq(\xi|\zeta)_{b}+600\delta,

    and the same holds if we replace ζ\zeta with xXx\in X.

  2. (2)

    For any ξ,ζ,λXωX\xi,\zeta,\lambda\in X\cup\partial_{\omega}X we have

    (ξ|λ)bmin{(ξ|ζ)b,(ζ|λ)b}600δ.(\xi|\lambda)_{b}\geq\min\{(\xi|\zeta)_{b},(\zeta|\lambda)_{b}\}-600\delta.

Combining (1) of Lemma 2.7 with inequality (2.9) gives for all x,yXXx,y\in X\cup\partial X with (x,y)(ω,ω)(x,y)\neq(\omega,\omega),

(2.10) (x|y)bmin{b(x),b(y)}+600δ,(x|y)_{b}\leq\min\{b(x),b(y)\}+600\delta,

where we set b(ω)=b(\omega)=-\infty and b(ξ)=b(\xi)=\infty for ξωX\xi\in\partial_{\omega}X.

For a point ωgX\omega\in\partial^{g}X belonging to the geodesic boundary, a sequence {xn}\{x_{n}\} converges to infinity with respect to ω\omega if for some Busemann function bb based at ω\omega we have (xm|xn)b(x_{m}|x_{n})_{b}\rightarrow\infty as m,nm,n\rightarrow\infty. Two sequences {xn}\{x_{n}\}, {yn}\{y_{n}\} converging to infinity with respect to ω\omega are equivalent with respect to ω\omega if (xn|yn)b(x_{n}|y_{n})_{b}\rightarrow\infty. These notions do not depend on the choice of Busemann function bb based at ω\omega by Lemma 2.5. One then defines the Gromov boundary relative to ω\omega as the set of all equivalence classes of sequences converging to infinity with respect to ω\omega. We will denote this by ωX\partial_{\omega}X. As our past use of the notation ωX=X\{ω}\partial_{\omega}X=\partial X\backslash\{\omega\} suggests we have the following, which is [12, Proposition 3.4.1].

Proposition 2.8.

A sequence {xn}\{x_{n}\} converges to infinity with respect to ω\omega if and only if it converges to a point ξX\{ω}\xi\in\partial X\backslash\{\omega\}. This correspondence defines a canonical identification of ωX\partial_{\omega}X and X\{ω}\partial X\backslash\{\omega\}.

We recall that for ωX\omega\in X we will often abuse terminology and also refer to ωX=X\partial_{\omega}X=\partial X as the Gromov boundary relative to ω\omega.

2.3. Visual metrics

Let K1K\geq 1 and let ZZ be a set. A function θ:Z×Z[0,)\theta:Z\times Z\rightarrow[0,\infty) is a KK-quasi-metric if the following holds for any z,z,z′′Zz,z^{\prime},z^{\prime\prime}\in Z,

  1. (1)

    θ(z,z)=0\theta(z,z^{\prime})=0 if and only if z=zz=z^{\prime},

  2. (2)

    θ(z,z)=θ(z,z)\theta(z,z^{\prime})=\theta(z^{\prime},z),

  3. (3)

    θ(z,z′′)Kmax{θ(z,z),θ(z,z′′)}\theta(z,z^{\prime\prime})\leq K\max\{\theta(z,z^{\prime}),\theta(z^{\prime},z^{\prime\prime})\}.

By a standard construction (see [12, Lemma 2.2.5]) a KK-quasi-metric with K2K\leq 2 is always 44-biLipschitz to a metric on ZZ. Since for ε>0\varepsilon>0 we have that θε\theta^{\varepsilon} is a KεK^{\varepsilon} quasi-metric if θ\theta is a KK-quasi-metric, for any quasi-metric θ\theta we always have that θε\theta^{\varepsilon} is 44-biLipschitz to a metric dd on ZZ (by the identity map on ZZ) whenever ε\varepsilon is small enough that Kε2K^{\varepsilon}\leq 2.

Let XX be a geodesic δ\delta-hyperbolic space. For xXx\in X and ε>0\varepsilon>0 we define for ξ\xi, ζX\zeta\in\partial X,

(2.11) θε,x(ξ,ζ)=eε(ξ|ζ)x,\theta_{\varepsilon,x}(\xi,\zeta)=e^{-\varepsilon(\xi|\zeta)_{x}},

with the understanding that e=0e^{-\infty}=0. By (2.5) the function θε,x\theta_{\varepsilon,x} defines an e8δεe^{8\delta\varepsilon}-quasi-metric on X\partial X. We refer to any metric θ\theta on X\partial X that is LL-biLipschitz to θε,x\theta_{\varepsilon,x} as a visual metric on X\partial X based at xx with parameter ε\varepsilon; we call LL the comparison constant to the model quasi-metric θε,x\theta_{\varepsilon,x}. A visual metric always exists once ε\varepsilon is small enough that e8δε2e^{8\delta\varepsilon}\leq 2. We give X\partial X the topology induced by any visual metric. Equipped with a visual metric with respect to any basepoint xXx\in X and any parameter ε>0\varepsilon>0 the set X\partial X 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 xXx\in X, however the comparison constant to the quasi-metric (2.11) will depend on the basepoint. For b𝒟(X)b\in\mathcal{D}(X) of the form b(y)=d(x,y)+sb(y)=d(x,y)+s for some ss\in\mathbb{R} we then define

(2.12) θε,b(ξ,ζ)=eε(ξ|ζ)b=eεsθε,x(ξ,ζ).\theta_{\varepsilon,b}(\xi,\zeta)=e^{-\varepsilon(\xi|\zeta)_{b}}=e^{-\varepsilon s}\theta_{\varepsilon,x}(\xi,\zeta).

Let ωgX\omega\in\partial^{g}X be a point of the geodesic boundary and let b(X)b\in\mathcal{B}(X) be a Busemann function based at ω\omega. We define for ε>0\varepsilon>0 and ξ,ζωX\xi,\zeta\in\partial_{\omega}X,

(2.13) θε,b(ξ,ζ)=eε(ξ|ζ)b.\theta_{\varepsilon,b}(\xi,\zeta)=e^{-\varepsilon(\xi|\zeta)_{b}}.

Then θε,b\theta_{\varepsilon,b} defines an e600δεe^{600\delta\varepsilon}-quasi-metric on ωX\partial_{\omega}X by Lemma 2.7. A visual metric based at ω\omega with parameter ε\varepsilon is defined to be any metric θ\theta on ωX\partial_{\omega}X that is LL-biLipschitz to θε,b\theta_{\varepsilon,b}, and as before we will call LL the comparison constant to the model quasi-metric θε,b\theta_{\varepsilon,b}. Since all Busemann functions associated to ω\omega differ from each other by a constant, up to a bounded error (by Lemma 2.5), the notion of a visual metric based at ω\omega does not depend on the choice of Busemann function bb based at ω\omega. Equipped with any visual metric based at ω\omega the metric space ωX\partial_{\omega}X is complete. It is bounded if and only if ω\omega is an isolated point in X\partial X.

Remark 2.9.

For CAT(1)(-1) spaces the quasi-metric θ1,b\theta_{1,b} for b^(X)b\in\hat{\mathcal{B}}(X) with basepoint ω\omega defines a distinguished visual metric on ωX\partial_{\omega}X with parameter ε=1\varepsilon=1. This metric is known as a Bourdon metric when b𝒟(X)b\in\mathcal{D}(X) and a Hamenstädt metric when b(X)b\in\mathcal{B}(X). The basic properties of the Bourdon metric for CAT(1)(-1) spaces were established by Bourdon in [8]. The Hamenstädt metric was introduced by Hamenstädt in the setting of Hadamard manifolds with sectional curvatures 1\leq-1 in [18] through a slightly different construction. The formulation for CAT(1)(-1) spaces using Gromov products based at bb is due to Foertsch-Schroeder [15].

3. Tripod maps and Busemann functions

In this section we let XX be a geodesic δ\delta-hyperbolic space for a given parameter δ0\delta\geq 0. We will be establishing some standard claims regarding geodesic triangles in XX that have vertices on the Gromov boundary X\partial X. We will then use these claims regarding geodesic triangles in XX to evaluate Busemann functions on geodesics in XX in Proposition 3.10 and Lemma 3.12. When XX 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 δ\delta-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 Δ\Delta be a geodesic triangle in XX with vertices x,y,zXXx,y,z\in X\cup\partial X and let χ0\chi\geq 0 be given. A collection of points x^yz\hat{x}\in yz, y^xz\hat{y}\in xz, z^xy\hat{z}\in xy is χ\chi-equiradial if

diam{x^,y^,z^}=max{|x^y^|,|y^z^|,|x^z^|}χ.\mathrm{diam}\{\hat{x},\hat{y},\hat{z}\}=\max\{|\hat{x}\hat{y}|,|\hat{y}\hat{z}|,|\hat{x}\hat{z}|\}\leq\chi.

We then refer to x^\hat{x}, y^\hat{y}, z^\hat{z} as χ\chi-equiradial points for Δ\Delta.

Remark 3.2.

For x,y,zXx,y,z\in X Definition 3.1 makes sense in any geodesic metric space XX. Taking χ=δ\chi=\delta gives yet another quantitatively equivalent definition of δ\delta-hyperbolicity for XX. See [17, Chapitre 2, Proposition 21].

When x,y,zXx,y,z\in X, the 4δ4\delta-tripod condition directly provides us with a set of 4δ4\delta-equiradial points x^\hat{x}, y^\hat{y}, z^\hat{z} defined by the system of equalities |xy^|=|xz^|=(y|z)x|x\hat{y}|=|x\hat{z}|=(y|z)_{x}, |yx^|=|yz^|=(x|z)y|y\hat{x}|=|y\hat{z}|=(x|z)_{y}, and |zx^|=|zy^|=(x|y)z|z\hat{x}|=|z\hat{y}|=(x|y)_{z}. We will often refer to these points as the canonical equiradial points for Δ\Delta, since they are uniquely determined. The following definition encodes a convenient hypothesis to make on equiradial points of a geodesic triangle Δ\Delta that partially generalizes the notion of canonical equiradial points to the case that some of the vertices of Δ\Delta belong to X\partial X.

We adopt the notation convention for x,yXXx,y\in X\cup\partial X that |xy|=|xy|=\infty if xyx\neq y and one of xx or yy belongs to X\partial X and |xy|=0|xy|=0 if x=yx=y.

Definition 3.3.

Let Δ\Delta be a geodesic triangle in XX with vertices x,y,zXXx,y,z\in X\cup\partial X, let χ0\chi\geq 0 be given, and let (x^,y^,z^)(\hat{x},\hat{y},\hat{z}) be a collection of χ\chi-equiradial points for Δ\Delta. We say that this collection is calibrated if we have |x^z|=|y^z||\hat{x}z|=|\hat{y}z|, |y^x|=|z^x||\hat{y}x|=|\hat{z}x|, and |z^y|=|x^y||\hat{z}y|=|\hat{x}y|.

This condition is trivially satisfied when all vertices of Δ\Delta belong to X\partial X, since all of the subsegments involved have infinite length.

We let Υ\Upsilon be the tripod geodesic metric space composed of three copies L1L_{1}, L2L_{2}, and L3L_{3} of the closed half-line [0,)[0,\infty) identified at 0. This identification point will be denoted by oo and will be referred to as the core of the tripod Υ\Upsilon. The space Υ\Upsilon is clearly 0-hyperbolic. The Gromov boundary Υ\partial\Upsilon consists of three points ζi\zeta_{i}, i=1,2,3i=1,2,3, corresponding to the half-lines LiL_{i} thought of as geodesic rays starting from the core oo.

For a geodesic triangle Δ\Delta with a calibrated ordered triple of χ\chi-equiradial points (x^,y^,z^)(\hat{x},\hat{y},\hat{z}) as in Definitions 3.1 and 3.3, we define the associated tripod map T:ΔΥT:\Delta\rightarrow\Upsilon to be the map that sends the sides xzxz, yzyz, and xyxy isometrically into L1L3L_{1}\cup L_{3}, L2L3L_{2}\cup L_{3}, and L1L2L_{1}\cup L_{2} respectively in the unique way that satisfies T(x)L1{ζ1}T(x)\in L_{1}\cup\{\zeta_{1}\}, T(y)L2{ζ2}T(y)\in L_{2}\cup\{\zeta_{2}\},T(z)L3{ζ3}T(z)\in L_{3}\cup\{\zeta_{3}\}, and T(x^)=T(y^)=T(z^)=oT(\hat{x})=T(\hat{y})=T(\hat{z})=o. To be more precise for boundary points, when xXx\in\partial X we mean here that T(x)=ζ1T(x)=\zeta_{1}, i.e., TT maps the geodesic rays y^x\hat{y}x and z^x\hat{z}x isometrically onto L1L_{1}. A choice of ordering of the equiradial points is required to define the map TT but is not important, as changing the ordering simply corresponds to permuting the rays LiL_{i} in Υ\Upsilon while keeping the core oo fixed.

We first obtain the following direct consequence of Lemma 2.2.

Lemma 3.4.

Let Δ\Delta be a geodesic triangle with vertices x,y,zXXx,y,z\in X\cup\partial X. Then there is a calibrated 60δ60\delta-equiradial collection of points x^yz\hat{x}\in yz, y^xz\hat{y}\in xz, z^xy\hat{z}\in xy.

Proof.

If all vertices of Δ\Delta belong to XX then the canonical equiradial points give a calibrated 4δ4\delta-equiradial collection for Δ\Delta, so we can assume that at least one vertex of Δ\Delta belongs to X\partial X. Thus we can assume without loss of generality that zXz\in\partial X.

Parametrize the side xyxy by arclength as γ:IX\gamma:I\rightarrow X for an interval II\subset\mathbb{R}, oriented from xx to yy. Let ExIE_{x}\subset I be the collection of times tt such that dist(γ(t),xz)10δ\mathrm{dist}(\gamma(t),xz)\leq 10\delta and EyIE_{y}\subset I the collection of times tt such that dist(γ(t),yz)10δ\mathrm{dist}(\gamma(t),yz)\leq 10\delta. Each of the sets ExE_{x} and EyE_{y} are closed and we have ExEy=IE_{x}\cup E_{y}=I by Lemma 2.2. We claim that both ExE_{x} and EyE_{y} are always nonempty. For this we can assume without loss of generality that ExE_{x} is nonempty since ExEy=IE_{x}\cup E_{y}=I.

If Ey=E_{y}=\emptyset then Ex=IE_{x}=I. For each tIt\in I we let xtxzx_{t}\in xz be a point such that |xtγ(t)|10δ|x_{t}\gamma(t)|\leq 10\delta. For nn\in\mathbb{N} the sequence {γ(n)}\{\gamma(n)\} converges to yy, which implies that the sequence {xn}\{x_{n}\} converges to yy since these sequences are a bounded distance from one another. However any sequence of points converging to infinity in xzxz can only possibly converge to xx or zz, which is a contradiction. Thus EyE_{y} must also be nonempty.

By the connectedness of II we then conclude that ExEyE_{x}\cap E_{y}\neq\emptyset. Letting sExEys\in E_{x}\cap E_{y}, setting w:=γ(s)w:=\gamma(s), and selecting points uxzu\in xz, vyzv\in yz such that |wu|10δ|wu|\leq 10\delta and |wv|10δ|wv|\leq 10\delta, we conclude that {w,u,v}\{w,u,v\} is a 20δ20\delta-equiradial collection of points for Δ\Delta.

Lastly we need to produce a calibrated collection of equiradial points from the collection {w,u,v}\{w,u,v\}. If all vertices of Δ\Delta belong to X\partial X then the collection is trivially calibrated, so we can assume at least one vertex of Δ\Delta belongs to XX. By relabeling the vertices we can then assume that either xXx\in X and yXy\in\partial X or xXx\in X and yXy\in X. In both cases we can find uxzu^{\prime}\in xz such that |xu|=|xw||xu^{\prime}|=|xw| since |xz|=|xz|=\infty. Then

(3.1) |uu|=||xu||xu||=||xu||xw|||uw|20δ.|uu^{\prime}|=||xu|-|xu^{\prime}||=||xu|-|xw||\leq|uw|\leq 20\delta.

It follows that the collection {w,u,v}\{w,u^{\prime},v\} is 40δ40\delta-equiradial. If yXy\in\partial X then this collection is also calibrated and we are done.

If yXy\in X then we repeat this argument again by using the fact that |yz|=|yz|=\infty to find vyzv^{\prime}\in yz such that |yv|=|yw||yv^{\prime}|=|yw|. The calculation (3.1) then shows that |vv|20δ|vv^{\prime}|\leq 20\delta as well. We can then conclude that the collection {w,u,v}\{w,u^{\prime},v^{\prime}\} is calibrated and 60δ60\delta-equiradial, as desired. ∎

Our next goal will be to prove that the tripod map T:ΔΥT:\Delta\rightarrow\Upsilon 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 XX be a metric space and let x,y,zXx,y,z\in X with |xz||yz||xz|\leq|yz|. Suppose that we are given geodesics xzxz and yzyz joining xx to zz and yy to zz respectively. Let uxzu\in xz, vyzv\in yz be given points that satisfy |xu|=|yv||xu|=|yv| and let wyzw\in yz be the unique point satisfying |wz|=|uz||wz|=|uz|. Then wvzw\in vz and |wv||xy||wv|\leq|xy|.

Proof.

The point ww must belong to the subsegment vzvz of yzyz, as if wyvw\in yv and wvw\neq v then

|yz|\displaystyle|yz| =|yv|+|wz||wv|\displaystyle=|yv|+|wz|-|wv|
=|xu|+|uz||wv|\displaystyle=|xu|+|uz|-|wv|
=|xz||wv|\displaystyle=|xz|-|wv|
<|xz|,\displaystyle<|xz|,

contradicting that |yz||xz||yz|\geq|xz|. Since wvzw\in vz we then have

|yz|\displaystyle|yz| =|yv|+|vw|+|wz|\displaystyle=|yv|+|vw|+|wz|
=|xu|+|vw|+|uz|\displaystyle=|xu|+|vw|+|uz|
=|xz|+|vw|,\displaystyle=|xz|+|vw|,

which implies by the triangle inequality that |vw||xy||vw|\leq|xy|. ∎

We now apply Lemma 3.5 to the setting of a δ\delta-hyperbolic space XX.

Lemma 3.6.

Let x,yXx,y\in X, let zXXz\in X\cup\partial X, and let x¯xz\bar{x}\in xz, y¯yz\bar{y}\in yz satisfy |xx¯|=|yy¯||x\bar{x}|=|y\bar{y}|. Then we have

(3.2) |x¯y¯|3|xy|+8δ.|\bar{x}\bar{y}|\leq 3|xy|+8\delta.
Proof.

Set t=|xx¯|=|yy¯|t=|x\bar{x}|=|y\bar{y}|. If t|xy|t\leq|xy| then

|x¯y¯||x¯x|+|xy|+|y¯y|3|xy|,|\bar{x}\bar{y}|\leq|\bar{x}x|+|xy|+|\bar{y}y|\leq 3|xy|,

which verifies inequality (3.2). We can thus assume that t>|xy|t>|xy|.

We first assume that zXz\in X. We can then assume without loss of generality that |xz||yz||xz|\leq|yz|. We consider a geodesic triangle Δ=xyz\Delta=xyz with sides the given geodesics xzxz and yzyz, as well as a geodesic xyxy from xx to yy. Let wyzw\in yz be the unique point such that |wz|=|x¯z||wz|=|\bar{x}z|. Lemma 3.5 shows that wy¯zw\in\bar{y}z and |wy¯||xy||w\bar{y}|\leq|xy|.

Let xxzx^{\prime}\in xz and yyzy^{\prime}\in yz be the canonical equiradial points for Δ\Delta on these edges. These points must satisfy max{|xx|,|yy|}|xy|\max\{|x^{\prime}x|,|y^{\prime}y|\}\leq|xy| since xyxy is an edge of Δ\Delta. The assumption t>|xy|t>|xy| then implies that x¯xz\bar{x}\in x^{\prime}z and y¯yz\bar{y}\in y^{\prime}z. Thus wyzw\in y^{\prime}z. The 4δ4\delta-tripod condition then implies that |wx¯|4δ|w\bar{x}|\leq 4\delta, from which it follows that |x¯y¯||xy|+4δ|\bar{x}\bar{y}|\leq|xy|+4\delta. This proves (3.2) in this case.

We now consider the case zXz\in\partial X. For each s0s\geq 0 we define xsxzx_{s}\in xz, ysyzy_{s}\in yz to be the points such that |xxs|=s|xx_{s}|=s and |yys|=s|yy_{s}|=s. Since the geodesics xzxz and yzyz have the same endpoint zXz\in\partial X, we must have (xs|ys)x(x_{s}|y_{s})_{x}\rightarrow\infty as ss\rightarrow\infty and the same for (xs|ys)y(x_{s}|y_{s})_{y}. We choose ss large enough that (xs|ys)xt(x_{s}|y_{s})_{x}\geq t and (xs|ys)yt(x_{s}|y_{s})_{y}\geq t. We consider a geodesic triangle Δ1=xxsys\Delta_{1}=xx_{s}y_{s} with edges the subsegment xxsxx_{s} of the given geodesic xzxz as well as geodesics xsysx_{s}y_{s} and xysxy_{s}, and a triangle Δ2=xyys\Delta_{2}=xyy_{s} with edges the subsegment yysyy_{s} of the given geodesic yzyz, the edge xysxy_{s} of Δ1\Delta_{1}, and a geodesic xyxy. Then x¯xxs\bar{x}\in xx_{s} and y¯yys\bar{y}\in yy_{s} by our choice of ss.

Since (xs|ys)xt(x_{s}|y_{s})_{x}\geq t, we must have |xys|t|xy_{s}|\geq t. Therefore there is a unique point wxysw\in xy_{s} such that |xw|=|xx¯|=t|xw|=|x\bar{x}|=t. The 4δ4\delta-tripod condition applied to the triangle Δ1\Delta_{1} then implies that |x¯w|4δ|\bar{x}w|\leq 4\delta. If |xys||yys||xy_{s}|\leq|yy_{s}| then we let uyysu\in yy_{s} be the unique point such that |uys|=|wys||uy_{s}|=|wy_{s}|. By applying Lemma 3.5 we then conclude that uy¯ysu\in\bar{y}y_{s} and |uy¯||xy||u\bar{y}|\leq|xy|. Since xyxy is an edge of the triangle Δ2\Delta_{2}, the canonical equiradial points of this triangle on the edges xysxy_{s} and yysyy_{s} can be at most a distance |xy|t|xy|\leq t from the vertices xx and yy respectively. We thus conclude from the 4δ4\delta-tripod condition that |uw|4δ|uw|\leq 4\delta. Combining these inequalities together gives

(3.3) |x¯y¯||x¯w|+|wu|+|uy¯||xy|+8δ,|\bar{x}\bar{y}|\leq|\bar{x}w|+|wu|+|u\bar{y}|\leq|xy|+8\delta,

which proves (3.2). The case |xys||yys||xy_{s}|\geq|yy_{s}| is similar: we let vxysv\in xy_{s} be the point such that |vys|=|y¯ys||vy_{s}|=|\bar{y}y_{s}|, apply Lemma 3.5 to obtain |vw||xy||vw|\leq|xy| and vwysv\in wy_{s}, then apply the 4δ4\delta-tripod condition to obtain |vy¯|4δ|v\bar{y}|\leq 4\delta. 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 |x¯y¯||xy|+8δ|\bar{x}\bar{y}|\leq|xy|+8\delta when |xy|>|xx¯|=|yy¯||xy|>|x\bar{x}|=|y\bar{y}|.

We will use Lemma 3.6 to show that the tripod map associated to a collection of calibrated equiradial points for a geodesic triangle Δ\Delta is roughly isometric.

Proposition 3.8.

Let x,y,zXXx,y,z\in X\cup\partial X be given vertices of a geodesic triangle Δ\Delta in XX. Let x^yz\hat{x}\in yz, y^xz\hat{y}\in xz, z^xy\hat{z}\in xy be points such that (x^,y^,z^)(\hat{x},\hat{y},\hat{z}) is a calibrated ordered triple of χ\chi-equiradial points for Δ\Delta for a given χ0\chi\geq 0. Let T:ΔΥT:\Delta\rightarrow\Upsilon be the tripod map associated to this triple. Then TT is (6χ+16δ)(6\chi+16\delta)-roughly isometric.

In particular if (x^,y^,z^)(\hat{x},\hat{y},\hat{z}) is the calibrated 60δ60\delta-equiradial triple produced in Lemma 3.4 then TT is 400δ400\delta-roughly isometric.

Proof.

By symmetry (permuting the vertices xx, yy, and zz), to estimate |T(p)T(q)||T(p)T(q)| for p,qΔp,q\in\Delta it suffices to restrict to the case py^zp\in\hat{y}z and then consider the possible locations of qq. By construction we have |T(p)T(q)|=|pq||T(p)T(q)|=|pq| if pp and qq belong to the same edge of Δ\Delta, since the tripod map is isometric on the edges of Δ\Delta. This handles the case that qq belongs to the same edge as pp, i.e., that qxzq\in xz.

We next consider the case qyzq\in yz. Since |y^z|=|x^z||\hat{y}z|=|\hat{x}z|, we can find a point ux^zu\in\hat{x}z such that |x^u|=|y^p||\hat{x}u|=|\hat{y}p|. Then |T(p)T(q)|=|uq||T(p)T(q)|=|uq|. Applying Lemma 3.6 yields

|up|3|y^x^|+8δ3χ+8δ,|up|\leq 3|\hat{y}\hat{x}|+8\delta\leq 3\chi+8\delta,

so that

||uq||pq|||up|3χ+8δ,||uq|-|pq||\leq|up|\leq 3\chi+8\delta,

which gives the desired estimate in this case.

Lastly we must consider the case qxyq\in xy, which we subdivide into the cases qxz^q\in x\hat{z} and qz^yq\in\hat{z}y. When qxz^q\in x\hat{z} we can use the condition |xz^|=|xy^||x\hat{z}|=|x\hat{y}| to find a point vxy^v\in x\hat{y} such that |qz^|=|vy^||q\hat{z}|=|v\hat{y}|. Then |T(p)T(q)|=|vp||T(p)T(q)|=|vp|. Similarly to the previous case, Lemma 3.6 gives us the estimate |vq|3χ+8δ|vq|\leq 3\chi+8\delta which implies that

||vp||pq|||vq|3χ+8δ,||vp|-|pq||\leq|vq|\leq 3\chi+8\delta,

as desired. When qz^yq\in\hat{z}y we use the equality |z^y|=|x^y||\hat{z}y|=|\hat{x}y| to find wx^yw\in\hat{x}y such that |z^q|=|x^w||\hat{z}q|=|\hat{x}w|, and we use the equality |x^z|=|y^z||\hat{x}z|=|\hat{y}z| to find sx^zs\in\hat{x}z such that |sx^|=|py^||s\hat{x}|=|p\hat{y}|. Then |T(p)T(q)|=|sw||T(p)T(q)|=|sw|. Lemma 3.6 gives us the estimate

max{|sp|,|wq|}3χ+8δ,\max\{|sp|,|wq|\}\leq 3\chi+8\delta,

which implies by the triangle inequality that

||sw||pq||||sw||wp||+||wp||qp|||sp|+|wq|6χ+16δ.||sw|-|pq||\leq||sw|-|wp||+||wp|-|qp||\leq|sp|+|wq|\leq 6\chi+16\delta.

This completes the proof of the main claim. The final assertion follows by substituting χ=60δ\chi=60\delta 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 T:ΔΥT:\Delta\rightarrow\Upsilon. To make this more formal, for a geodesic triangle Δ\Delta in XX we will refer to a tripod map T:ΔΥT:\Delta\rightarrow\Upsilon associated to Δ\Delta as being any tripod map TT for Δ\Delta associated to an ordered triple (x^,y^,z^)(\hat{x},\hat{y},\hat{z}) of calibrated 60δ60\delta-equiradial points for Δ\Delta obtained from Lemma 3.4. We will also abuse terminology and say that x^\hat{x}, y^\hat{y} and z^\hat{z} are equiradial points for the tripod map TT (as opposed to for the triangle Δ\Delta).

3.2. Calculating Busemann functions

We recall that the Gromov boundary Υ\partial\Upsilon of the tripod Υ\Upsilon is a disjoint union of three points ζi\zeta_{i}, i=1,2,3i=1,2,3, corresponding to the geodesic rays γi:[0,)Υ\gamma_{i}:[0,\infty)\rightarrow\Upsilon that parametrize the half-lines LiL_{i} starting from oo for i=1,2,3i=1,2,3. Set bΥ:=bγ1b_{\Upsilon}:=b_{\gamma_{1}} to be the Busemann function associated to the geodesic ray γ1\gamma_{1}. A straightforward calculation shows that bΥb_{\Upsilon} is given by bΥ(s)=sb_{\Upsilon}(s)=-s for sL1s\in L_{1} and bΥ(s)=sb_{\Upsilon}(s)=s for sL2s\in L_{2} or sL3s\in L_{3}, when we consider each of these rays as identified with [0,)[0,\infty).

In this next proposition we consider a geodesic triangle Δ\Delta in XX with a distinguished vertex ωX\omega\in\partial X together with a Busemann function bb based at ω\omega. We will not keep track of exact constants in the proof of this lemma so we will not produce an explicit value for κ=κ(δ)\kappa=\kappa(\delta) below. If one does careful bookkeeping in the proof it is possible to show that κ=2000δ\kappa=2000\delta works.

Proposition 3.10.

Let Δ=ωxy\Delta=\omega xy be a geodesic triangle in XX with ωX\omega\in\partial X and x,yXωXx,y\in X\cup\partial_{\omega}X. There is a constant κ=κ(δ)\kappa=\kappa(\delta) such that the following holds: let ω^xy\hat{\omega}\in xy, x^ωy\hat{x}\in\omega y, and y^ωx\hat{y}\in\omega x be a calibrated set of 60δ60\delta-equiradial points on Δ\Delta provided by Lemma 3.4 and let bb be a Busemann function based at ω\omega. Let T:ΔΥT:\Delta\rightarrow\Upsilon be the tripod map associated to the triple (ω^,x^,y^)(\hat{\omega},\hat{x},\hat{y}). Then for each pΔp\in\Delta we have

(3.4) b(p)κbΥ(T(p))+(x|y)b.b(p)\doteq_{\kappa}b_{\Upsilon}(T(p))+(x|y)_{b}.

Consequently we have b(p)κ(x|y)bb(p)\doteq_{\kappa}(x|y)_{b} for p{ω^,x^,y^}p\in\{\hat{\omega},\hat{x},\hat{y}\} and

(3.5) (x|y)bκinfpxyb(p).(x|y)_{b}\doteq_{\kappa}\inf_{p\in xy}b(p).
Proof.

Since we will not be keeping track of the exact value of the final constant κ=κ(δ)\kappa=\kappa(\delta) in the proof, we will let \doteq denote any equality up to an additive error depending only on δ\delta. Set u=b(ω^)u=b(\hat{\omega}). Then ub(x^)u\doteq b(\hat{x}) and ub(y^)u\doteq b(\hat{y}) since bb is 1-Lipschitz. We will prove the rough equality (3.4) with uu in place of (x|y)b(x|y)_{b} and use this to deduce that u(x|y)bu\doteq(x|y)_{b}. Thus we will first show that for pΔp\in\Delta we have

(3.6) b(p)bΥ(T(p))+u.b(p)\doteq b_{\Upsilon}(T(p))+u.

We first handle the case in which pωxp\in\omega x or pωyp\in\omega y. Since the roles of xx and yy are symmetric, we can assume without loss of generality that pωxp\in\omega x. Let γ:(,a]X\gamma:(-\infty,a]\rightarrow X be an arclength parametrization of ωx\omega x with γ(0)=x^\gamma(0)=\hat{x} and γ(t)ω\gamma(t)\rightarrow\omega as tt\rightarrow-\infty. If we define s(,a]s\in(-\infty,a] such that γ(s)=p\gamma(s)=p then it follows from the construction of the tripod map that bΥ(T(γ(s))=sb_{\Upsilon}(T(\gamma(s))=s. Applying Lemma 2.6 gives

b(p)b(x^)s=bΥ(T(γ(s)),b(p)-b(\hat{x})\doteq s=b_{\Upsilon}(T(\gamma(s)),

which gives (3.6) since b(x^)ub(\hat{x})\doteq u.

The remaining case is when pxyp\in xy. By the symmetric roles of xx and yy we can assume that pxω^p\in x\hat{\omega}. As in the proof of Proposition 3.8, since |xω^|=|xy^||x\hat{\omega}|=|x\hat{y}| we can find qxy^q\in x\hat{y} such that |pω^|=|qy^||p\hat{\omega}|=|q\hat{y}|. Then by Lemma 3.6 we have

|pq|3|y^ω^|+10δ190δ.|pq|\leq 3|\hat{y}\hat{\omega}|+10\delta\leq 190\delta.

Thus b(p)b(q)b(p)\doteq b(q) since bb is 1-Lipschitz. It then follows, from the rough equality (3.6) for qωxq\in\omega x that we established above, that

b(p)b(q)|qy^|+u=|pω^|+u,b(p)\doteq b(q)\doteq|q\hat{y}|+u=|p\hat{\omega}|+u,

which gives (3.6) in this case.

We next show that u(x|y)bu\doteq(x|y)_{b}. By Lemma 2.7 we have for any sequences xnxx_{n}\rightarrow x and ynyy_{n}\rightarrow y that (xn|yn)b600δ(x|y)b(x_{n}|y_{n})_{b}\doteq_{600\delta}(x|y)_{b} for sufficiently large nn; if xXx\in X then we can just set xn=xx_{n}=x for all nn and the same goes for yy. We choose sequences {xn}\{x_{n}\} and {yn}\{y_{n}\} that belong to xyxy and consider only those nn large enough that (xn|yn)b600δ(x|y)b(x_{n}|y_{n})_{b}\doteq_{600\delta}(x|y)_{b} and xnω^xx_{n}\in\hat{\omega}x, ynω^yy_{n}\in\hat{\omega}y. Then applying (3.6),

(x|y)b\displaystyle(x|y)_{b} (xn|yn)b\displaystyle\doteq(x_{n}|y_{n})_{b}
=12(b(xn)+b(yn)|xnyn|)\displaystyle=\frac{1}{2}(b(x_{n})+b(y_{n})-|x_{n}y_{n}|)
12(|xnω^|+|ynω^|+2u|xnyn|)\displaystyle\doteq\frac{1}{2}(|x_{n}\hat{\omega}|+|y_{n}\hat{\omega}|+2u-|x_{n}y_{n}|)
=u.\displaystyle=u.

Thus we can substitute in (x|y)b(x|y)_{b} for uu in (3.6) at the cost of an additional additive constant depending only on δ\delta. The main claim (3.4) follows. The assertion that b(p)κ(x|y)bb(p)\doteq_{\kappa}(x|y)_{b} for p{ω^,x^,y^}p\in\{\hat{\omega},\hat{x},\hat{y}\} follows from (3.4) since each point of {ω^,x^,y^}\{\hat{\omega},\hat{x},\hat{y}\} has image oΥo\in\Upsilon under the tripod map TT and bΥ(o)=0b_{\Upsilon}(o)=0. The rough equality (3.5) also follows directly from (3.4) since the image of xyxy under TT is contained in L2L3L_{2}\cup L_{3} and bΥb_{\Upsilon} is nonnegative on this subset of Υ\Upsilon. ∎

Proposition 3.10 leads to the following important definition, which is useful for calculations. We recall our convention that ωX=X\partial_{\omega}X=\partial X when ωX\omega\in X.

Definition 3.11.

Let XX be a geodesic δ\delta-hyperbolic space and let b^(X)b\in\hat{\mathcal{B}}(X) be given with basepoint ω\omega. Let x,yXωXx,y\in X\cup\partial_{\omega}X and let c0c\geq 0 be a given constant. Suppose that xyxy is a geodesic joining xx to yy. We say that a parametrization γ:IX\gamma:I\rightarrow X, II\subseteq\mathbb{R}, of xyxy by arclength is cc-adapted to bb if 0I0\in I and

(3.7) b(γ(t))c|t|+(x|y)b,b(\gamma(t))\doteq_{c}|t|+(x|y)_{b},

for tIt\in I.

The inclusion of 0 in the domain of γ\gamma will be vital for our applications. When the value of cc is implied by context we will often shorten the terminology to just saying that the parametrization γ\gamma is adapted to bb.

For b^(X)b\in\hat{\mathcal{B}}(X) with basepoint ω\omega we will construct adapted parametrizations for geodesics joining any two points x,yXωXx,y\in X\cup\partial_{\omega}X under an assumption similar to the rough starlikeness hypothesis of Theorem 1.4. We emphasize that the points xx and yy in the lemma need not always be the vertices of a geodesic triangle Δ\Delta with a third vertex at ω\omega.

Lemma 3.12.

Let XX be a geodesic δ\delta-hyperbolic space and let b^(X)b\in\hat{\mathcal{B}}(X) be given with basepoint ω\omega. Let x,yXωXx,y\in X\cup\partial_{\omega}X and let xyxy be a given geodesic from xx to yy. Suppose that we are given K0K\geq 0 and points x,yXωXx^{\prime},y^{\prime}\in X\cup\partial_{\omega}X and geodesics ωx,ωy\omega x^{\prime},\omega y^{\prime} joining ω\omega to xx^{\prime} and yy^{\prime} respectively such that max{|xx|,|yy|}K\max\{|xx^{\prime}|,|yy^{\prime}|\}\leq K. Then there is a constant c=c(δ,K)c=c(\delta,K) depending only on δ\delta and KK such that there is a parametrization γ:IX\gamma:I\rightarrow X of xyxy that is cc-adapted to bb.

Proof.

We first consider the case that x=xx=x^{\prime} and y=yy=y^{\prime}, so that we can take K=0K=0. Let Δ=ωxy\Delta=\omega xy be the geodesic triangle formed by the geodesics ωx\omega x, ωy\omega y, and xyxy. Let T:ΔΥT:\Delta\rightarrow\Upsilon be a 400δ400\delta-roughly isometric tripod map associated to Δ\Delta such that T(ω)L1{ζ1}T(\omega)\in L_{1}\cup\{\zeta_{1}\}, T(x)L2{ζ2}T(x)\in L_{2}\cup\{\zeta_{2}\}, and T(y)L3{ζ3}T(y)\in L_{3}\cup\{\zeta_{3}\}, as given by Proposition 3.8. We identify the union L2L3L_{2}\cup L_{3} of geodesic rays in Υ\Upsilon with \mathbb{R} by identifying L2L_{2} with (,0](-\infty,0] and L3L_{3} with [0,)[0,\infty), sending the core oo of Υ\Upsilon to the origin in \mathbb{R}. We let I=T(xy)L2L3I=T(xy)\subset L_{2}\cup L_{3} denote the image of xyxy under TT and consider II as a subinterval II\subset\mathbb{R} under the identification of L2L3L_{2}\cup L_{3} with \mathbb{R}. Since the tripod map TT is isometric when restricted to xyxy, we can then construct an arclength parametrization γ:IX\gamma:I\rightarrow X of xyxy by inverting the restriction of TT to xyxy. By the construction of TT we have 0I0\in I since the core oo is contained in the image T(xy)T(xy) of xyxy.

When b(X)b\in\mathcal{B}(X) the rough equality (3.4) directly implies the cc-adapted condition (3.11) for γ\gamma with c=c(δ)c=c(\delta), since for zL2L3z\in L_{2}\cup L_{3} we have bΥ(z)=|zo|b_{\Upsilon}(z)=|zo|. For b𝒟(X)b\in\mathcal{D}(X) it is easy to see that it suffices to verify (3.7) for bb of the form b(x)=|xω|b(x)=|x\omega|, ωX\omega\in X. We then have to show that there is a constant c=c(δ)c=c(\delta) such that for tIt\in I we have

(3.8) d(γ(t),ω)c|t|+(x|y)ω.d(\gamma(t),\omega)\doteq_{c}|t|+(x|y)_{\omega}.

By (2.4) we can find points x¯ωx\bar{x}\in\omega x, y¯ωy\bar{y}\in\omega y such that (x¯|y¯)ωc(δ)(x|y)ω(\bar{x}|\bar{y})_{\omega}\doteq_{c(\delta)}(x|y)_{\omega} and T(x¯)L2T(\bar{x})\in L_{2}, T(y¯)L3T(\bar{y})\in L_{3}. Then (T(x¯)|T(y¯))T(ω)c(δ)(x¯|y¯)ω(T(\bar{x})|T(\bar{y}))_{T(\omega)}\doteq_{c(\delta)}(\bar{x}|\bar{y})_{\omega} since TT is c(δ)c(\delta)-roughly isometric. Since T(y¯)L2T(\bar{y})\in L_{2} and T(z¯)L3T(\bar{z})\in L_{3}, a quick calculation then shows that

(T(x¯)|T(y¯))ω=|T(ω)o|,(T(\bar{x})|T(\bar{y}))_{\omega}=|T(\omega)o|,

and therefore (x|y)ωc(δ)|T(ω)o|(x|y)_{\omega}\doteq_{c(\delta)}|T(\omega)o|. Thus for tIt\in I we have

|t|+(x|y)ωc(δ)|t|+|T(ω)o|=|T(ω)T(γ(t))|,|t|+(x|y)_{\omega}\doteq_{c(\delta)}|t|+|T(\omega)o|=|T(\omega)T(\gamma(t))|,

with the second equality following from the construction of γ\gamma. The rough equality (3.8) with c=c(δ)c=c(\delta) then directly follows from the fact that TT is c(δ)c(\delta)-roughly isometric.

We now consider the general case in which we are given points x,yXωXx^{\prime},y^{\prime}\in X\cup\partial_{\omega}X and K0K\geq 0 such that max{|xx|,|yy|}K\max\{|xx^{\prime}|,|yy^{\prime}|\}\leq K. If xx and yy both belong to ωX\partial_{\omega}X then our conventions imply that x=xx=x^{\prime} and y=yy=y^{\prime}, hence this case reduces to the case K=0K=0 considered previously.

If xx and yy both belong to XX then we apply the K=0K=0 case to the points xx^{\prime} and yy^{\prime} to obtain a cc^{\prime}-adapted parametrization σ:IX\sigma:I^{\prime}\rightarrow X of xyx^{\prime}y^{\prime} oriented from xx^{\prime} to yy^{\prime}, I=[t,t+]I^{\prime}=[t_{-}^{\prime},t_{+}^{\prime}] with c=c(δ)c^{\prime}=c^{\prime}(\delta). Since 0I0\in I^{\prime} we have t0t_{-}^{\prime}\leq 0 and t+0t_{+}^{\prime}\geq 0. Let η:IX\eta:I\rightarrow X, I=[t,t+]I=[t_{-}^{\prime},t_{+}], be the unique arclength parametrization of xyxy that is oriented from xx to yy and starts from the same time parameter tt_{-}^{\prime} as σ\sigma. The piecewise geodesic curve xxxyyyxx^{\prime}\cup x^{\prime}y^{\prime}\cup yy^{\prime} joining xx to yy can be parametrized as a 4K4K-roughly isometric map β:JX\beta:J\rightarrow X for an appropriate interval JJ\subset\mathbb{R}. By the stability of geodesics in Gromov hyperbolic spaces [12, Theorem 1.3.2] this implies that there is a constant c0=c0(δ,K)c_{0}=c_{0}(\delta,K) such that the given geodesic xyxy is contained in a c0c_{0}-neighborhood of the curve β\beta. Since the segments xxxx^{\prime} and yyyy^{\prime} of β\beta are each contained in a KK-neighborhood of σ\sigma, by increasing c0c_{0} by an amount depending only on KK we can assume that xyxy is contained in a c0c_{0}-neighborhood of σ\sigma.

Now let tIt\in I be given and let sIs\in I^{\prime} be such that |η(t)σ(s)|c0|\eta(t)\sigma(s)|\leq c_{0}. Since bb is 1-Lipschitz it follows that

b(η(t))c0b(σ(s))c|s|+(x|y)b.b(\eta(t))\doteq_{c_{0}}b(\sigma(s))\doteq_{c^{\prime}}|s|+(x|y)_{b}.

Thus it suffices to show that tc′′st\doteq_{c^{\prime\prime}}s for a constant c′′=c′′(δ,K)c^{\prime\prime}=c^{\prime\prime}(\delta,K). Since tt=|η(t)x|t-t_{-}^{\prime}=|\eta(t)x| and st=|σ(s)x|s-t_{-}^{\prime}=|\sigma(s)x^{\prime}|, we have

|ts|\displaystyle|t-s| =|(tt)(st)|\displaystyle=|(t-t_{-}^{\prime})-(s-t_{-}^{\prime})|
=||η(t)x||σ(s)x||\displaystyle=||\eta(t)x|-|\sigma(s)x^{\prime}||
||η(t)x||η(t)x||+||η(t)x||σ(s)x||\displaystyle\leq||\eta(t)x|-|\eta(t)x^{\prime}||+||\eta(t)x^{\prime}|-|\sigma(s)x^{\prime}||
|xx|+|η(t)σ(s)|\displaystyle\leq|xx^{\prime}|+|\eta(t)\sigma(s)|
K+c0,\displaystyle\leq K+c_{0},

so that we can set c′′=K+c0c^{\prime\prime}=K+c_{0}. It follows that η\eta satisfies (3.7) with constant c=c(δ,K)c=c(\delta,K) depending only on δ\delta and KK.

If 0I0\in I then η\eta gives a parametrization of xyxy that is cc-adapted to bb and we are done. We can therefore assume that 0I0\notin I which implies that t+<0t_{+}<0 since t0t_{-}^{\prime}\leq 0. We then note that |xy|2K|xy||x^{\prime}y^{\prime}|\doteq_{2K}|xy| and t+t=|xy|t_{+}^{\prime}-t_{-}^{\prime}=|x^{\prime}y^{\prime}|, t+t=|xy|t_{+}-t_{-}^{\prime}=|xy|, which implies that t+2Kt+t_{+}\doteq_{2K}t_{+}^{\prime}. Since t+0t_{+}^{\prime}\geq 0 and t+0t_{+}\leq 0, we conclude that |t+|2K|t_{+}|\leq 2K. We set I′′=[tt+,0]I^{\prime\prime}=[t_{-}^{\prime}-t_{+},0] and set γ(t)=η(t+t+)\gamma(t)=\eta(t+t_{+}) for tI′′t\in I^{\prime\prime}. Then 0I′′0\in I^{\prime\prime} by construction and this arclength parametrization γ\gamma of xyxy still satisfies (3.7) with c=c(δ,K)c=c(\delta,K) since bb is 11-Lipschitz and |t+|2K|t_{+}|\leq 2K. Thus γ\gamma gives the desired adapted parametrization.

Lastly we consider the case in which one of xx or yy belong to ωX\partial_{\omega}X, but not both. Without loss of generality we can assume that xXx\in X and yωXy\in\partial_{\omega}X. Let {yn}xy\{y_{n}\}\subset xy be the sequence of points with |xyn|=n|xy_{n}|=n for each nn\in\mathbb{N}. Let ηn:InX\eta_{n}:I_{n}\rightarrow X be the arclength parametrizations of xynxy_{n} for each nn that were constructed in the previous case, In=[sn,tn]I_{n}=[s_{n},t_{n}]. Since 0In0\in I_{n} for each nn we have sn0s_{n}\leq 0 for each nn. Since ηn(sn)=x\eta_{n}(s_{n})=x for each nn, we have from the condition that ηn\eta_{n} is cc-adapted to bb,

b(x)c|sn|+(x|y)b=sn+(x|y)bb(x)\doteq_{c}|s_{n}|+(x|y)_{b}=-s_{n}+(x|y)_{b}

with c=c(δ,K)c=c(\delta,K). It follows that smcsns_{m}\doteq_{c}s_{n} for each m,nm,n\in\mathbb{N}. Thus, by replacing ηn\eta_{n} with the parametrization γn\gamma_{n} defined by γn(t)=ηn(ts1+sn)\gamma_{n}(t)=\eta_{n}(t-s_{1}+s_{n}) on the domain In=[s1,tn+s1sn]I_{n}^{\prime}=[s_{1},t_{n}+s_{1}-s_{n}], we can assume that sn=s1:=ss_{n}=s_{1}:=s for all nn\in\mathbb{N}. Note also that, since tnt_{n}\rightarrow\infty as nn\rightarrow\infty and s0s\leq 0, we have 0In0\in I_{n} for all large enough nn. It follows that the resulting parametrization γn\gamma_{n} will be cc-adapted to bb for nn large enough that tn0t_{n}\geq 0 since bb is 1-Lipschitz, with c=c(δ,K)c=c(\delta,K).

With these modifications the parametrizations γn\gamma_{n} now have the same starting point s0s\leq 0. Since these are parametrizations of xynxy_{n} by arclength and the sequence {yn}\{y_{n}\} defines progressively longer subsegments xynxy_{n} of xyxy that exhaust xyxy, the maps γn\gamma_{n} coincide wherever their domains overlap and can therefore be used to define a parametrization γ:[s,)X\gamma:[s,\infty)\rightarrow X of xyxy that is cc-adapted to bb 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 dε,bd_{\varepsilon,b}. Section 4.2 proves the theorems in the case that b(X)b\in\mathcal{B}(X) (i.e., bb is a Busemann function). Section 4.3 then uses a special construction (Definition 4.18) to deduce the case b𝒟(X)b\in\mathcal{D}(X) from the case b(X)b\in\mathcal{B}(X). Since Theorem 1.10 follows from Theorems 1.4 and 1.6 once we’ve shown that ρ1,b\rho_{1,b} 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 (Xε,b,dε,b)(X_{\varepsilon,b},d_{\varepsilon,b}) of a geodesic δ\delta-hyperbolic space XX by the density ρε,b(x)=eεb(x)\rho_{\varepsilon,b}(x)=e^{-\varepsilon b(x)} for a given ε>0\varepsilon>0 and b^(X)b\in\hat{\mathcal{B}}(X) using the tripod maps we built in Section 3. For now we will not be assuming that ρε,b\rho_{\varepsilon,b} is a GH-density (using the terminology of Definition 1.3). Hence we can use these results to establish that ρε,b\rho_{\varepsilon,b} is a GH-density in certain important cases. To simplify notation we will drop the function bb from the notation for objects associated to the conformal deformation and write ρε:=ρε,b\rho_{\varepsilon}:=\rho_{\varepsilon,b}, Xε:=Xε,bX_{\varepsilon}:=X_{\varepsilon,b}, etc. For a curve γ:IXε\gamma:I\rightarrow X_{\varepsilon} we will write ε(γ):=ρε(γ)\ell_{\varepsilon}(\gamma):=\ell_{\rho_{\varepsilon}}(\gamma) for its length measured in the metric dεd_{\varepsilon}. We let (γ)\ell(\gamma) denote the length of γ\gamma measured in XX instead.

Remark 4.1.

Throughout the rest of this paper we will be using [4, Proposition A.7], which for a geodesic metric space XX and a continuous function ρ:X(0,)\rho:X\rightarrow(0,\infty) allows us to compute the lengths ρ(γ)\ell_{\rho}(\gamma) in the conformal deformation XρX_{\rho} of curves γ:IX\gamma:I\rightarrow X parametrized by arclength in XX as

(4.1) ρ(γ)=Iργ𝑑s,\ell_{\rho}(\gamma)=\int_{I}\rho\circ\gamma\,ds,

with dsds denoting the standard length element in \mathbb{R}.

Since bb is 11-Lipschitz we have the Harnack type inequality for x,yXx,y\in X,

(4.2) eε|xy|ρε(x)ρε(y)eε|xy|,e^{-\varepsilon|xy|}\leq\frac{\rho_{\varepsilon}(x)}{\rho_{\varepsilon}(y)}\leq e^{\varepsilon|xy|},

which made its first appearance in the statement of Theorem 1.9 earlier.

The metric spaces XεX_{\varepsilon} and XX are biLipschitz on bounded subsets of XX by inequality (4.2). A more precise estimate for this is given in the lemma below.

Lemma 4.2.

For any x,yXx,y\in X we have

ρε(x)ε1(1eε|xy|)dε(x,y)ρε(x)ε1(eε|xy|1).\rho_{\varepsilon}(x)\varepsilon^{-1}(1-e^{-\varepsilon|xy|})\leq d_{\varepsilon}(x,y)\leq\rho_{\varepsilon}(x)\varepsilon^{-1}(e^{\varepsilon|xy|}-1).
Proof.

For the upper bound we let xyxy be a geodesic joining xx to yy. Then, using (4.2),

dε(x,y)\displaystyle d_{\varepsilon}(x,y) xyρe𝑑s\displaystyle\leq\int_{xy}\rho_{e}\,ds
ρε(x)0|xy|eεt𝑑t\displaystyle\leq\rho_{\varepsilon}(x)\int_{0}^{|xy|}e^{\varepsilon t}\,dt
=ρε(x)ε1(eε|xy|1).\displaystyle=\rho_{\varepsilon}(x)\varepsilon^{-1}(e^{\varepsilon|xy|}-1).

For the lower bound we consider a rectifiable curve γ\gamma joining xx to yy in XX, which we can assume is parametrized by arclength as γ:[0,(γ)]X\gamma:[0,\ell(\gamma)]\rightarrow X with (γ)\ell(\gamma) denoting the length of γ\gamma in XX. With this parametrization γ\gamma defines a 11-Lipschitz function from [0,(γ)][0,\ell(\gamma)] to XX, so that in particular we have |xγ(t)|t|x\gamma(t)|\leq t for each t[0,(γ)]t\in[0,\ell(\gamma)]. Then by (4.2),

ε(γ)\displaystyle\ell_{\varepsilon}(\gamma) ρε(x)0(γ)eε|xγ(t)|𝑑t\displaystyle\geq\rho_{\varepsilon}(x)\int_{0}^{\ell(\gamma)}e^{-\varepsilon|x\gamma(t)|}\,dt
ρε(x)0(γ)eεt𝑑t\displaystyle\geq\rho_{\varepsilon}(x)\int_{0}^{\ell(\gamma)}e^{-\varepsilon t}\,dt
=ρε(x)ε1(1eε(γ))\displaystyle=\rho_{\varepsilon}(x)\varepsilon^{-1}(1-e^{-\varepsilon\ell(\gamma)})
ρε(x)ε1(1eε|xy|),\displaystyle\geq\rho_{\varepsilon}(x)\varepsilon^{-1}(1-e^{-\varepsilon|xy|}),

where in the final line we used that (γ)|xy|\ell(\gamma)\geq|xy|. ∎

Lemma 4.2 can be rewritten in the following useful form when |xy|1|xy|\leq 1.

Lemma 4.3.

For any x,yXx,y\in X with |xy|1|xy|\leq 1 we have

(4.3) dε(x,y)C(ε)eε(x|y)b|xy|.d_{\varepsilon}(x,y)\asymp_{C(\varepsilon)}e^{-\varepsilon(x|y)_{b}}|xy|.
Proof.

For 0t10\leq t\leq 1 we have the inequalities

1eεtεeεt,1-e^{-\varepsilon t}\geq\varepsilon e^{-\varepsilon}t,

and

eεt1εeεt,e^{\varepsilon t}-1\leq\varepsilon e^{\varepsilon}t,

as can be verified by noting that equality holds at t=0t=0 and differentiating each side. Thus for |xy|1|xy|\leq 1 the inequality of Lemma 4.2 implies that

(4.4) dε(x,y)C(ε)ρε(x)|xy|C(ε)eε(x|y)b|xy|,d_{\varepsilon}(x,y)\asymp_{C(\varepsilon)}\rho_{\varepsilon}(x)|xy|\asymp_{C(\varepsilon)}e^{-\varepsilon(x|y)_{b}}|xy|,

with the final estimate following from

ρε(x)=eεb(x)eεeε(x|y)b,\rho_{\varepsilon}(x)=e^{-\varepsilon b(x)}\asymp_{e^{\varepsilon}}e^{-\varepsilon(x|y)_{b}},

since |xy|1|xy|\leq 1 and bb is 11-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.

XεX_{\varepsilon} is bounded if and only if b𝒟(X)b\in\mathcal{D}(X).

Proof.

We first suppose that b𝒟(X)b\in\mathcal{D}(X). For xXx\in X we can then write b(x)=|xz|+sb(x)=|xz|+s for some zXz\in X and ss\in\mathbb{R}. We let xXx\in X be given and let γ\gamma be a geodesic joining zz to xx. Then

dε(x,z)\displaystyle d_{\varepsilon}(x,z) γρε𝑑t\displaystyle\leq\int_{\gamma}\rho_{\varepsilon}\,dt
=es0|xz|eεt𝑑t\displaystyle=e^{-s}\int_{0}^{|xz|}e^{-\varepsilon t}\,dt
=ε1es(1eε|xz|)\displaystyle=\varepsilon^{-1}e^{-s}(1-e^{-\varepsilon|xz|})
ε1es.\displaystyle\leq\varepsilon^{-1}e^{-s}.

It follows that XεX_{\varepsilon} is bounded with diamXε2ε1es\mathrm{diam}\,X_{\varepsilon}\leq 2\varepsilon^{-1}e^{-s}.

Now suppose that b(X)b\in\mathcal{B}(X). Then we can find a geodesic ray γ:[0,)X\gamma:[0,\infty)\rightarrow X and a constant ss\in\mathbb{R} such that b=bγ+sb=b_{\gamma}+s. For each t0t\geq 0 we apply the comparison (4.4) with x=γ(t)x=\gamma(t) and y=γ(t+1)y=\gamma(t+1) to obtain

dε(γ(t),γ(t+1))C(ε)eεb(γ(t))=eε(ts),d_{\varepsilon}(\gamma(t),\gamma(t+1))\asymp_{C(\varepsilon)}e^{-\varepsilon b(\gamma(t))}=e^{\varepsilon(t-s)},

since b(γ(t))=t+sb(\gamma(t))=-t+s. Thus as tt\rightarrow\infty we have dε(γ(t),γ(t+1))d_{\varepsilon}(\gamma(t),\gamma(t+1))\rightarrow\infty. It follows that XεX_{\varepsilon} is unbounded. ∎

We next use adapted parametrizations to estimate the length in XεX_{\varepsilon} of geodesics in XX. Below we write ω=ωb\omega=\omega_{b} for the basepoint of bb.

Lemma 4.5.

Let x,yXx,y\in X be given and let γ\gamma be a geodesic in XX joining xx to yy. Suppose that we are given K0K\geq 0 and points x,yXx^{\prime},y^{\prime}\in X and geodesics ωx,ωy\omega x^{\prime},\omega y^{\prime} joining ω\omega to xx and yy such that max{|xx|,|yy|}K\max\{|xx^{\prime}|,|yy^{\prime}|\}\leq K. Then

(4.5) ε(γ)C(δ,K,ε)eε(x|y)bmin{1,|xy|}.\ell_{\varepsilon}(\gamma)\asymp_{C(\delta,K,\varepsilon)}e^{-\varepsilon(x|y)_{b}}\min\{1,|xy|\}.
Proof.

Throughout this proof all additive constants c0c\geq 0 and C1C\geq 1 will depend only on δ\delta, KK, and ε\varepsilon; we write \doteq and \asymp for c\doteq_{c} and C\asymp_{C} respectively. We consider an arclength parametrization γ:IX\gamma:I\rightarrow X of γ\gamma that is cc-adapted to bb with c=c(δ,K)c=c(\delta,K) as constructed in Lemma 3.12. We assume that γ\gamma is oriented from xx to yy and set w=γ(0)w=\gamma(0). By (3.7) we then have b(w)(x|y)bb(w)\doteq(x|y)_{b}.

When |xy|1|xy|\leq 1 we observe that |zw||xy|1|zw|\leq|xy|\leq 1 for all zxyz\in xy. Inequality (4.2) then implies that

(4.6) ρε(z)eερε(w).\rho_{\varepsilon}(z)\asymp_{e^{\varepsilon}}\rho_{\varepsilon}(w).

By integrating the comparison (4.6) over γ\gamma we obtain

ε(γ)ρε(w)|xy|eε(x|y)b|xy|.\ell_{\varepsilon}(\gamma)\asymp\rho_{\varepsilon}(w)|xy|\asymp e^{-\varepsilon(x|y)_{b}}|xy|.

This gives the estimate (4.5) when |xy|1|xy|\leq 1.

We now suppose that |xy|1|xy|\geq 1. Let γ1:[|xw|,0]X\gamma_{1}:[-|xw|,0]\rightarrow X and γ2:[0,|yw|]X\gamma_{2}:[0,|yw|]\rightarrow X denote the parametrizations of the subsegments of γ\gamma from xx to ww and from ww to yy respectively. Then, using (3.7) and b(w)(x|y)bb(w)\doteq(x|y)_{b}, we have

ε(γ)\displaystyle\ell_{\varepsilon}(\gamma) =ε(γ1)+ε(γ2)\displaystyle=\ell_{\varepsilon}(\gamma_{1})+\ell_{\varepsilon}(\gamma_{2})
=γ1ρε𝑑t+γ2ρε𝑑t\displaystyle=\int_{\gamma_{1}}\rho_{\varepsilon}\,dt+\int_{\gamma_{2}}\rho_{\varepsilon}\,dt
eε(x|y)b(0|xw|eεt𝑑t+0|yw|eεt𝑑t)\displaystyle\asymp e^{-\varepsilon(x|y)_{b}}\left(\int_{0}^{|xw|}e^{-\varepsilon t}\,dt+\int_{0}^{|yw|}e^{-\varepsilon t}\,dt\right)
=ε1eε(x|y)b(2eε|xw|eε|yw|).\displaystyle=\varepsilon^{-1}e^{-\varepsilon(x|y)_{b}}(2-e^{-\varepsilon|xw|}-e^{-\varepsilon|yw|}).

It follows immediately that

ε(γ)Ceε(x|y)b,\ell_{\varepsilon}(\gamma)\leq Ce^{-\varepsilon(x|y)_{b}},

with C=C(δ,K,ε)C=C(\delta,K,\varepsilon). This gives the upper bound in (4.5) when |xy|1|xy|\geq 1. For the lower bound we note that since |xy|1|xy|\geq 1 and |xw|+|yw|=|xy||xw|+|yw|=|xy|, we must have min{|xw|,|yw|}12\min\{|xw|,|yw|\}\geq\frac{1}{2}. Therefore

ε1eε(x|y)b(2eε|xw|eε|yw|)ε1e(x|y)b(1eε2).\varepsilon^{-1}e^{-\varepsilon(x|y)_{b}}(2-e^{-\varepsilon|xw|}-e^{-\varepsilon|yw|})\geq\varepsilon^{-1}e^{-(x|y)_{b}}(1-e^{-\frac{\varepsilon}{2}}).

This gives the lower bound in (4.5) when |xy|1|xy|\geq 1. ∎

In connection with Lemma 4.5 it is helpful to formulate the following definition.

Definition 4.6.

Let ωXX\omega\in X\cup\partial X be given. For a constant K0K\geq 0 we say that XX is KK-roughly geodesic from ω\omega if for each xXx\in X there exists xXx^{\prime}\in X and a geodesic ωx\omega x^{\prime} joining ω\omega to xx^{\prime} such that |xx|K|xx^{\prime}|\leq K.

When XX is KK-roughly geodesic from ω\omega we can apply Lemma 4.5 freely to any geodesic γ\gamma in XX with this constant KK. If ωX\omega\in X then XX is 0-roughly geodesic from ω\omega since XX is geodesic. For the case ωX\omega\in\partial X we note that if XX is KK-roughly starlike from ω\omega then XX is clearly also KK-roughly geodesic from ω\omega. Note however that XX can be roughly geodesic from ωX\omega\in\partial X without being roughly starlike from ω\omega; this happens for instance when XX is a tree with arbitrarily long finite branches. We also remark that XX is always 0-roughly geodesic from any point of X\partial X when it is proper.

When ρε\rho_{\varepsilon} 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 XX is KK-roughly geodesic from ω\omega and that ρε\rho_{\varepsilon} is a GH-density with constant MM. Then for any x,yXx,y\in X we have

(4.7) dε(x,y)C(δ,K,ε,M)eε(x|y)bmin{1,|xy|}.d_{\varepsilon}(x,y)\asymp_{C(\delta,K,\varepsilon,M)}e^{-\varepsilon(x|y)_{b}}\min\{1,|xy|\}.

Lemma 4.5 has the following immediate corollary when combined with Lemma 4.3.

Corollary 4.8.

Suppose that XX is KK-roughly geodesic from ω\omega and that there is a constant M01M_{0}\geq 1 such that for any x,yXx,y\in X with |xy|>1|xy|>1 we have

(4.8) eε(x|y)bM0dε(x,y).e^{-\varepsilon(x|y)_{b}}\leq M_{0}d_{\varepsilon}(x,y).

Then ρε\rho_{\varepsilon} is a GH-density with constant M=M(δ,K,ε,M0)M=M(\delta,K,\varepsilon,M_{0}).

Proof.

If x,yXx,y\in X with |xy|1|xy|\leq 1 then combining Lemmas 4.3 and 4.5 establishes the GH-inequality (1.3) with M=M(δ,K,ε)M=M(\delta,K,\varepsilon). If we instead have |xy|>1|xy|>1 then the inequality (4.8) implies the GH-inequality (1.3) with M=M(δ,K,ε,M0)M=M(\delta,K,\varepsilon,M_{0}) by the estimate (4.5) and the fact that de(x,y)ε(γ)d_{e}(x,y)\leq\ell_{\varepsilon}(\gamma) for any geodesic γ\gamma joining xx to yy. ∎

We will use Corollary 4.8 to show for a CAT(1)(-1) space XX that ρ1=ρ1,b\rho_{1}=\rho_{1,b} is a GH-density for any b^(X)b\in\hat{\mathcal{B}}(X) with a universal constant M1M\geq 1. Our proof will be based on the following four point inequality for CAT(1)(-1) spaces.

Proposition 4.9.

[14, Proposition 3.3.4] Let XX be a CAT(1)(-1) space. Then for any four points x,y,z,wXx,y,z,w\in X we have

(4.9) e(x|z)we(x|y)w+e(y|z)w.e^{-(x|z)_{w}}\leq e^{-(x|y)_{w}}+e^{-(y|z)_{w}}.

Metric spaces satisfying the inequality (4.9) are called strongly hyperbolic in [14].

The inequality (4.9) can easily be improved to hold for Gromov products based at any function b^(X)b\in\hat{\mathcal{B}}(X).

Lemma 4.10.

Let XX be a CAT(1)(-1) space. Then for any x,y,zXx,y,z\in X and b^(X)b\in\hat{\mathcal{B}}(X) we have

(4.10) e(x|z)be(x|y)b+e(y|z)b.e^{-(x|z)_{b}}\leq e^{-(x|y)_{b}}+e^{-(y|z)_{b}}.
Proof.

If b𝒟(X)b\in\mathcal{D}(X) has the form b(x)=d(x,w)b(x)=d(x,w) for some wXw\in X then (4.10) is just a restatement of (4.9). The inequality for bb of the form b(x)=d(x,w)+sb(x)=d(x,w)+s for some wXw\in X and ss\in\mathbb{R} can then be obtained by multiplying the inequality (4.9) through by ese^{-s}.

Now suppose that b(X)b\in\mathcal{B}(X). Then we can find a geodesic ray γ:[0,)X\gamma:[0,\infty)\rightarrow X and an ss\in\mathbb{R} such that b=bγ+sb=b_{\gamma}+s. By multiplying the target inequality (4.10) through by ese^{s} we see that it suffices to consider the case that b=bγb=b_{\gamma}. Let x,y,zXx,y,z\in X be given. For each t[0,)t\in[0,\infty) we have from (4.9) that

e(x|z)γ(t)e(x|y)γ(t)+e(y|z)γ(t).e^{-(x|z)_{\gamma(t)}}\leq e^{-(x|y)_{\gamma(t)}}+e^{-(y|z)_{\gamma(t)}}.

After multiplying each side by ete^{t} and expanding the Gromov products we obtain that

e12(|xγ(t)|t+|yγ(t)|t|xy|)e12(|xγ(t)|t+|zγ(t)|t|xz|)+e12(|yγ(t)|t+|zγ(t)|t|yz|).e^{-\frac{1}{2}(|x\gamma(t)|-t+|y\gamma(t)|-t-|xy|)}\leq e^{-\frac{1}{2}(|x\gamma(t)|-t+|z\gamma(t)|-t-|xz|)}+e^{-\frac{1}{2}(|y\gamma(t)|-t+|z\gamma(t)|-t-|yz|)}.

Letting tt\rightarrow\infty then gives inequality (4.10). ∎

By combining Lemma 4.10 with Corollary 4.8 we can show that the density ρ1\rho_{1} for b^(X)b\in\hat{\mathcal{B}}(X) is a GH-density when XX is a complete CAT(1)(-1) space. The completeness hypothesis is only used in the case b(X)b\in\mathcal{B}(X).

Proposition 4.11.

There is a constant M1M\geq 1 such that for any complete CAT(1)(-1) space XX and any b^(X)b\in\hat{\mathcal{B}}(X) we have that ρ1=ρ1,b\rho_{1}=\rho_{1,b} is a GH-density with constant MM.

Proof.

We first observe that XX is always 0-roughly geodesic from ω\omega. For ωX\omega\in X this is trivial, while for ωX\omega\in\partial X this can be deduced from the completeness hypothesis together with the CAT(1)(-1) condition on XX [14, Proposition 4.4.4]. Thus we can apply Corollary 4.8 with K=0K=0. Let x,yXx,y\in X be given and let ωXX\omega\in X\cup\partial X denote the basepoint of bb. Since XX is δ\delta-hyperbolic with δ=δ(2)\delta=\delta(\mathbb{H}^{2}) and we are restricting to the case ε=1\varepsilon=1, it suffices by Corollary 4.8 to produce a universal constant M01M_{0}\geq 1 such that for any x,yXx,y\in X with |xy|>1|xy|>1 we have

(4.11) e(x|y)bM0d1(x,y).e^{-(x|y)_{b}}\leq M_{0}d_{1}(x,y).

We consider a rectifiable curve η:IX\eta:I\rightarrow X joining xx to yy that is parametrized by arclength (in XX) and oriented from xx to yy. We define a finite sequence of points t0,,tnt_{0},\dots,t_{n} in II inductively as follows: we set t0t_{0} to be the left endpoint of II and for each applicable k>0k>0 we set tkt_{k} to be the supremum of all points tIt\in I such that ttk1t\geq t_{k-1} and |η(s)η(tk1)|<1|\eta(s)\eta(t_{k-1})|<1 for each tk1stt_{k-1}\leq s\leq t. The finite length of η\eta in XX ensures that this sequence is finite, so that we have as a consequence that this process terminates at the right endpoint tnt_{n} of II. The assumption |xy|>1|xy|>1 implies that n2n\geq 2. By construction we then have |η(tk)η(tk+1)|=1|\eta(t_{k})\eta(t_{k+1})|=1 for 0kn20\leq k\leq n-2 and |η(tn1)η(tn)|1|\eta(t_{n-1})\eta(t_{n})|\leq 1. Since bb is 11-Lipschitz it follows that

e(x|y)bCe(η(t0)|η(tn1))b,e^{-(x|y)_{b}}\leq Ce^{-(\eta(t_{0})|\eta(t_{n-1}))_{b}},

with C=eC=e, recalling that η(t0)=x\eta(t_{0})=x. Hence to prove (4.11) it suffices to show for any rectifiable curve η\eta joining xx to yy that we have

(4.12) e(η(t0)|η(tn1))bC1(η),e^{-(\eta(t_{0})|\eta(t_{n-1}))_{b}}\leq C\ell_{1}(\eta),

for some constant C1C\geq 1. Repeated application of the inequality (4.10) based at bb gives

e(η(t0)|η(tn1))bk=0n2e(η(tk)|η(tk+1))bCk=0n2eb(η(tk)),e^{-(\eta(t_{0})|\eta(t_{n-1}))_{b}}\leq\sum_{k=0}^{n-2}e^{-(\eta(t_{k})|\eta(t_{k+1}))_{b}}\leq C\sum_{k=0}^{n-2}e^{-b(\eta(t_{k}))},

with the second inequality holding for C=eC=e since bb is 11-Lipschitz and |η(tk)η(tk+1)|=1|\eta(t_{k})\eta(t_{k+1})|=1. For 0kn20\leq k\leq n-2 we have by construction that any t[tk,tk+1]t\in[t_{k},t_{k+1}] satisfies |η(tk)η(t)|1|\eta(t_{k})\eta(t)|\leq 1. Hence the Harnack inequality (4.2) implies that

1(η|[tk,tk+1])eeb(η(tk))(η|[tk,tk+1])eb(η(tk)),\ell_{1}(\eta|_{[t_{k},t_{k+1}]})\asymp_{e}e^{-b(\eta(t_{k}))}\ell(\eta|_{[t_{k},t_{k+1}]})\geq e^{-b(\eta(t_{k}))},

since the length of η|[tk,tk+1]\eta|_{[t_{k},t_{k+1}]} in XX is at least the distance 11 between its endpoints. This shows that

k=0n2eb(η(tk))Ck=0n21(η|[tk,tk+1])C1(η),\sum_{k=0}^{n-2}e^{-b(\eta(t_{k}))}\leq C\sum_{k=0}^{n-2}\ell_{1}(\eta|_{[t_{k},t_{k+1}]})\leq C\ell_{1}(\eta),

with C=eC=e. We conclude that the desired estimate (4.12) holds, so that as a consequence ρ1\rho_{1} is a GH-density with a universal constant M1M\geq 1. ∎

By repurposing the proof of Proposition 4.11 we are also able to show that ρε\rho_{\varepsilon} being a GH-density implies that ρε\rho_{\varepsilon^{\prime}} is also a GH-density for each 0<εε0<\varepsilon^{\prime}\leq\varepsilon, with a constant independent of ε\varepsilon^{\prime}.

Proposition 4.12.

Suppose that XX is KK-roughly geodesic from ω\omega and that ρε\rho_{\varepsilon} is a GH-density with constant MM. Then there is a constant M=M(δ,K,ε,M)M^{\prime}=M^{\prime}(\delta,K,\varepsilon,M) such that ρε\rho_{\varepsilon^{\prime}} is a GH-density with constant MM^{\prime} for any 0<εε0<\varepsilon^{\prime}\leq\varepsilon.

Proof.

Let ε0=ε0(δ)\varepsilon_{0}=\varepsilon_{0}(\delta) be the threshold determined by the Harnack inequality (4.2) and Theorem 1.9 such that ρε\rho_{\varepsilon} is a GH-density with constant M=20M=20 for 0<εε00<\varepsilon\leq\varepsilon_{0}. For the purpose of proving the proposition we can then assume that ε>ε0\varepsilon>\varepsilon_{0} and ε[ε0,ε]\varepsilon^{\prime}\in[\varepsilon_{0},\varepsilon]. We will first produce a constant M0=M0(δ,K,ε,ε,M)M_{0}=M_{0}(\delta,K,\varepsilon,\varepsilon^{\prime},M) such that for any x,yXx,y\in X with |xy|>1|xy|>1 we have

(4.13) eε(x|y)bM0dε(x,y).e^{-\varepsilon^{\prime}(x|y)_{b}}\leq M_{0}d_{\varepsilon^{\prime}}(x,y).

As in the proof of Proposition 4.11, we let η:IX\eta:I\rightarrow X be a rectifiable curve joining xx to yy that is parametrized by arclength (in XX) and oriented from xx to yy. We then construct the finite sequence of points t0,,tnt_{0},\dots,t_{n} in II exactly as in the proof of Proposition 4.11, with n2n\geq 2 since |xy|>1|xy|>1. Since bb is 11-Lipschitz, we conclude as in that proof that it suffices to establish the estimate

(4.14) eε(η(t0)|η(tn1))bM1ε(η),e^{-\varepsilon^{\prime}(\eta(t_{0})|\eta(t_{n-1}))_{b}}\leq M_{1}\ell_{\varepsilon^{\prime}}(\eta),

for any rectifiable curve η\eta joining xx to yy, with M1=M1(δ,K,ε,ε,M)M_{1}=M_{1}(\delta,K,\varepsilon,\varepsilon^{\prime},M).

We set β=ε/ε\beta=\varepsilon^{\prime}/\varepsilon and observe that ε0/εβ1\varepsilon_{0}/\varepsilon\leq\beta\leq 1 by hypothesis. By the triangle inequality for dεd_{\varepsilon} we have

dε(η(t0),η(tn1))k=0n2dε(η(tk),η(tk+1)),d_{\varepsilon}(\eta(t_{0}),\eta(t_{n-1}))\leq\sum_{k=0}^{n-2}d_{\varepsilon}(\eta(t_{k}),\eta(t_{k+1})),

which implies that

dε(η(t0),η(tn1))βk=0n2dε(η(tk),η(tk+1))β,d_{\varepsilon}(\eta(t_{0}),\eta(t_{n-1}))^{\beta}\leq\sum_{k=0}^{n-2}d_{\varepsilon}(\eta(t_{k}),\eta(t_{k+1}))^{\beta},

since 0<β10<\beta\leq 1. By Lemma 4.7 we then conclude that

eε(η(t0)|η(tn1))b\displaystyle e^{-\varepsilon^{\prime}(\eta(t_{0})|\eta(t_{n-1}))_{b}} Ck=0n2eε(η(tk)|η(tk+1))b\displaystyle\leq C\sum_{k=0}^{n-2}e^{-\varepsilon^{\prime}(\eta(t_{k})|\eta(t_{k+1}))_{b}}
Ck=0n2eεb(η(tk)),\displaystyle\leq C\sum_{k=0}^{n-2}e^{-\varepsilon^{\prime}b(\eta(t_{k}))},

with C=C(δ,K,ε,ε,M)C=C(\delta,K,\varepsilon,\varepsilon^{\prime},M), where the second inequality follows from the fact that |η(tk)η(tk+1)|=1|\eta(t_{k})\eta(t_{k+1})|=1 by construction. Since |η(tk)η(t)|1|\eta(t_{k})\eta(t)|\leq 1 for each t[tk,tk+1]t\in[t_{k},t_{k+1}], we have by the Harnack inequality (4.2),

ε(η|[tk,tk+1])eεeεb(η(tk))(η|[tk,tk+1])eεb(η(tk)),\ell_{\varepsilon^{\prime}}(\eta|_{[t_{k},t_{k+1}]})\asymp_{e^{\varepsilon^{\prime}}}e^{-\varepsilon^{\prime}b(\eta(t_{k}))}\ell(\eta|_{[t_{k},t_{k+1}]})\geq e^{-\varepsilon^{\prime}b(\eta(t_{k}))},

since the length of η|[tk,tk+1]\eta|_{[t_{k},t_{k+1}]} in XX is at least the distance 11 between its endpoints. It follows that

k=0n2eεb(η(tk))C(ε)k=0n2ε(η|[tk,tk+1])C(ε)ε(η),\sum_{k=0}^{n-2}e^{-\varepsilon^{\prime}b(\eta(t_{k}))}\leq C(\varepsilon^{\prime})\sum_{k=0}^{n-2}\ell_{\varepsilon^{\prime}}(\eta|_{[t_{k},t_{k+1}]})\leq C(\varepsilon^{\prime})\ell_{\varepsilon^{\prime}}(\eta),

with C(ε)=eεC(\varepsilon^{\prime})=e^{\varepsilon^{\prime}}. 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 ε[ε0,ε]\varepsilon^{\prime}\in[\varepsilon_{0},\varepsilon] then the constants can always be chosen to depend only on ε0\varepsilon_{0} and ε\varepsilon. This shows that the GH-constant MM^{\prime} for ρε\rho_{\varepsilon^{\prime}} can be chosen such that M=M(δ,K,ε,ε0,M)M^{\prime}=M^{\prime}(\delta,K,\varepsilon,\varepsilon_{0},M). Since ε0=ε0(δ)\varepsilon_{0}=\varepsilon_{0}(\delta) we in fact obtain that M=M(δ,K,ε,M)M^{\prime}=M^{\prime}(\delta,K,\varepsilon,M), as desired. ∎

We recall that ω=ωb\omega=\omega_{b} denotes the basepoint of bb. We end this section by using Lemma 4.5 to construct a map

(4.15) φε=φε,b:ωXXε,\varphi_{\varepsilon}=\varphi_{\varepsilon,b}:\partial_{\omega}X\rightarrow\partial X_{\varepsilon},

when XX is roughly geodesic from ω\omega and ωX\partial_{\omega}X\neq\emptyset. We will be using the following corollary of the estimate (4.5) for x,yXx,y\in X,

(4.16) dε(x,y)C(δ,ε,K)eε(x|y)bmin{1,|xy|}.d_{\varepsilon}(x,y)\leq C(\delta,\varepsilon,K)e^{-\varepsilon(x|y)_{b}}\min\{1,|xy|\}.

The inequality (4.16) implies that if {xn}\{x_{n}\} is a sequence in XX that converges to infinity with respect to ω\omega (recall this means that (xm|xn)b(x_{m}|x_{n})_{b}\rightarrow\infty as m,nm,n\rightarrow\infty) then {xn}\{x_{n}\} is a Cauchy sequence in XεX_{\varepsilon}. Since XX and XεX_{\varepsilon} are biLipschitz on bounded subsets of XX by Lemma 4.2, it is easy to see that {xn}\{x_{n}\} cannot converge to a point of XεX_{\varepsilon}. It follows that XεX_{\varepsilon} is incomplete as long as ωX\partial_{\omega}X contains at least one point; the only exceptions are when either XX is bounded or ω\omega is the only point of X\partial X. Furthermore a second application of (4.16) shows that sequences {xn}\{x_{n}\} and {yn}\{y_{n}\} converging to infinity with respect to ω\omega that are equivalent with respect to ω\omega are equivalent as Cauchy sequences in XεX_{\varepsilon}, i.e., dε(xn,yn)0d_{\varepsilon}(x_{n},y_{n})\rightarrow 0 as nn\rightarrow\infty. Setting Xε=X¯ε\Xε\partial X_{\varepsilon}=\bar{X}_{\varepsilon}\backslash X_{\varepsilon} to be the complement of XεX_{\varepsilon} in its completion, we thus have a well-defined map φε:ωXXε\varphi_{\varepsilon}:\partial_{\omega}X\rightarrow\partial X_{\varepsilon} given by sending a sequence {xn}\{x_{n}\} converging to infinity with respect to ω\omega to its limit in Xε\partial X_{\varepsilon} inside of X¯ε\bar{X}_{\varepsilon}. In the next section we will show that φε\varphi_{\varepsilon} is a bijection when XX is roughly starlike from ω\omega and ρε\rho_{\varepsilon} is a GH-density. We remark that the map φε\varphi_{\varepsilon} need not be a bijection in general, see [3, Proposition 4.1].

4.2. Uniformizing by Busemann functions

For this section we let XX be a complete geodesic δ\delta-hyperbolic space. We let b(X)b\in\mathcal{B}(X) be given with basepoint ωX\omega\in\partial X and suppose that XX is KK-roughly starlike from ω\omega. For a given ε>0\varepsilon>0 we suppose ρε,b\rho_{\varepsilon,b} is a GH-density with constant MM. As in Section 4.1, to simplify notation we will drop bb from the notation for objects associated to the uniformization and write ρε:=ρε,b\rho_{\varepsilon}:=\rho_{\varepsilon,b}, Xε:=Xε,bX_{\varepsilon}:=X_{\varepsilon,b}, etc. As before, for a curve η:IXε\eta:I\rightarrow X_{\varepsilon} we will write ε(η):=ρε(η)\ell_{\varepsilon}(\eta):=\ell_{\rho_{\varepsilon}}(\eta) for its length measured in the metric dεd_{\varepsilon}. For brevity, throughout this section we write \doteq for equality up to an additive that depends only on δ\delta, KK, ε\varepsilon, and MM, and write \asymp for equality up to a multiplicative constant that depends only on those same parameters. We write c0c\geq 0 and C1C\geq 1 for additive and multiplicative constants depending only on these parameters.

The rough starlikeness of XX from ω\omega implies that ωX\partial_{\omega}X contains at least one point, so as a consequence the space XεX_{\varepsilon} is incomplete by the discussion at the conclusion of the previous section. As before we write X¯ε\bar{X}_{\varepsilon} for the completion of the uniformization XεX_{\varepsilon} and Xε=X¯ε\Xε\partial X_{\varepsilon}=\bar{X}_{\varepsilon}\backslash X_{\varepsilon} for the boundary of XεX_{\varepsilon} inside its completion. We will continue to write dεd_{\varepsilon} for the canonical extension of this metric on XεX_{\varepsilon} to the completion X¯ε\bar{X}_{\varepsilon}. We write dε(x)=dist(x,Xε)d_{\varepsilon}(x)=\mathrm{dist}(x,\partial X_{\varepsilon}) for the distance to the boundary in XεX_{\varepsilon}. We let φε:ωXXε\varphi_{\varepsilon}:\partial_{\omega}X\rightarrow\partial X_{\varepsilon} be the map constructed in (4.15) at the end of the previous section by sending a sequence {xn}\{x_{n}\} converging to infinity with respect to ω\omega to its limit in Xε\partial X_{\varepsilon} as a Cauchy sequence in XεX_{\varepsilon}. We formally extend the distance function dεd_{\varepsilon} to ωX\partial_{\omega}X by setting dε(x,y):=dε(φε(x),φε(y))d_{\varepsilon}(x,y):=d_{\varepsilon}(\varphi_{\varepsilon}(x),\varphi_{\varepsilon}(y)) for x,yXωXx,y\in X\cup\partial_{\omega}X, where we set φε(x)=x\varphi_{\varepsilon}(x)=x for xXx\in X.

Our first lemma extends the estimate (4.7) to hold for x,yXωXx,y\in X\cup\partial_{\omega}X with our formal extension of dεd_{\varepsilon} to ωX\partial_{\omega}X. We recall our convention that for x,yXXx,y\in X\cup\partial X we define |xy|=|xy|=\infty if xyx\neq y and either xXx\in\partial X or yXy\in\partial X, and define |xy|=0|xy|=0 if x=yXx=y\in\partial X.

Lemma 4.13.

For any x,yXωXx,y\in X\cup\partial_{\omega}X we have

(4.17) dε(x,y)eε(x|y)bmin{1,|xy|}.d_{\varepsilon}(x,y)\asymp e^{-\varepsilon(x|y)_{b}}\min\{1,|xy|\}.
Proof.

The case in which x,yXx,y\in X is an immediate consequence of Lemma 4.7 since XX is KK-roughly geodesic from ω\omega (because XX is KK-roughly starlike from ω\omega). We thus only need to consider the case in which at least one of the points belongs to ωX\partial_{\omega}X. Since (4.17) holds trivially when x=yx=y we can assume that xyx\neq y. We can then assume without loss of generality that xωXx\in\partial_{\omega}X. We then need to show that dε(x,y)eε(x|y)bd_{\varepsilon}(x,y)\asymp e^{-\varepsilon(x|y)_{b}}. We let {xn}\{x_{n}\} and {yn}\{y_{n}\} be sequences converging to infinity with respect to ω\omega in XX that represent the points xx and yy respectively; if yXy\in X then we instead set yn=yy_{n}=y for all nn. Our definition of the extension of dεd_{\varepsilon} to ωX\partial_{\omega}X then implies that we have limndε(xn,yn)=dε(x,y)\lim_{n\rightarrow\infty}d_{\varepsilon}(x_{n},y_{n})=d_{\varepsilon}(x,y). For nn sufficiently large we will have (x|y)b600δ(xn|yn)b(x|y)_{b}\doteq_{600\delta}(x_{n}|y_{n})_{b} by Lemma 2.7 and we will have |xnyn|1|x_{n}y_{n}|\geq 1. The comparison (4.17) then follows from the corresponding comparison for xnx_{n} and yny_{n} for sufficiently large nn. ∎

By using Lemma 4.13 we are able to show that φε\varphi_{\varepsilon} is a bijection.

Lemma 4.14.

The map φε:ωXXε\varphi_{\varepsilon}:\partial_{\omega}X\rightarrow\partial X_{\varepsilon} is a bijection.

Proof.

The injectivity of φε\varphi_{\varepsilon} follows immediately from Lemma 4.13 applied to xyωXx\neq y\in\partial_{\omega}X. Thus our focus will be on showing that φε\varphi_{\varepsilon} is surjective. Let {xn}\{x_{n}\} be a Cauchy sequence in XεX_{\varepsilon} that converges to a point zXεz\in\partial X_{\varepsilon}. We claim that the sequence {xn}\{x_{n}\} cannot belong to a bounded subset of XX. If it did then for a fixed pXp\in X there would be an r>0r>0 such that {xn}B(p,r)\{x_{n}\}\subset B(p,r) for all nn, with B(p,r)B(p,r) denoting the ball of radius rr centered at pp in XX. Lemma 4.2 shows that the metrics on XX and XεX_{\varepsilon} are biLipschitz to one another on B(p,2r)B(p,2r), which implies that {xn}\{x_{n}\} is also a Cauchy sequence in XX. Since XX is complete this Cauchy sequence must converge in XX to a point yB(p,2r)y\in B(p,2r). However this means that {xn}\{x_{n}\} also converges to yy in XεX_{\varepsilon}, contradicting that {xn}\{x_{n}\} converges to a point of Xε\partial X_{\varepsilon}.

Thus, by passing to a subsequence if necessary, we can assume that |xmxn|1|x_{m}x_{n}|\geq 1 for mnm\neq n. It then follows from Lemma 4.13 that for mnm\neq n,

dε(xm,xn)eε(xm|xn)b.d_{\varepsilon}(x_{m},x_{n})\asymp e^{-\varepsilon(x_{m}|x_{n})_{b}}.

Since dε(xn,xm)0d_{\varepsilon}(x_{n},x_{m})\rightarrow 0 as m,nm,n\rightarrow\infty, we conclude that (xm|xn)b(x_{m}|x_{n})_{b}\rightarrow\infty as m,nm,n\rightarrow\infty. Thus {xn}\{x_{n}\} converges to infinity with respect to ω\omega. Letting ξωX\xi\in\partial_{\omega}X denote the point in the Gromov boundary relative to ω\omega represented by the sequence {xn}\{x_{n}\}, the construction of φε\varphi_{\varepsilon} then shows that φε(ξ)=z\varphi_{\varepsilon}(\xi)=z. We conclude that φε\varphi_{\varepsilon} is surjective. ∎

We can now prove Theorem 1.6 in the case b(X)b\in\mathcal{B}(X). We recall the definition (2.13) of the model visual quasi-metric θε,b\theta_{\varepsilon,b} on ωX\partial_{\omega}X.

Proof of Theorem 1.6.

The fact that φε\varphi_{\varepsilon} is a bijection follows from Lemma 4.14. To complete the proof of Theorem 1.6 it then suffices to show that for any ξ,ζωX\xi,\zeta\in\partial_{\omega}X there is a constant L=L(δ,K,ε,M)L=L(\delta,K,\varepsilon,M) such that

dε(ξ,ζ)Lθε,b(ξ,ζ)=eε(ξ|ζ)b.d_{\varepsilon}(\xi,\zeta)\asymp_{L}\theta_{\varepsilon,b}(\xi,\zeta)=e^{-\varepsilon(\xi|\zeta)_{b}}.

This desired comparison then follows from Lemma 4.13. ∎

The next lemma shows that the distance to Xε\partial X_{\varepsilon} can be computed in terms of the density ρε\rho_{\varepsilon}.

Lemma 4.15.

For xXx\in X we have

(4.18) dε(x)ρε(x).d_{\varepsilon}(x)\asymp\rho_{\varepsilon}(x).
Proof.

Let xXx\in X be given. We first compute the upper bound in (4.18). By the rough-starlikeness condition we can find a geodesic line γ:X\gamma:\mathbb{R}\rightarrow X starting at ω\omega and ending at some ξX\xi\in\partial X with dist(x,γ)K\mathrm{dist}(x,\gamma)\leq K. Using Lemma 2.6 we can consider γ\gamma as parametrized by arclength with b(γ(t))144δtb(\gamma(t))\doteq_{144\delta}t for each tt\in\mathbb{R}. We let ss\in\mathbb{R} be such that |xγ(s)|K|x\gamma(s)|\leq K. We then compute,

ε(γ|[s,))\displaystyle\ell_{\varepsilon}(\gamma|_{[s,\infty)}) =seεb(γ(t))𝑑t\displaystyle=\int_{s}^{\infty}e^{-\varepsilon b(\gamma(t))}\,dt
seεt𝑑t\displaystyle\asymp\int_{s}^{\infty}e^{-\varepsilon t}\,dt
=ε1eεs.\displaystyle=\varepsilon^{-1}e^{-\varepsilon s}.

By Lemma 4.2 we then have

dε(x,ξ)\displaystyle d_{\varepsilon}(x,\xi) dε(x,γ(s))+dε(γ(s),ξ)\displaystyle\leq d_{\varepsilon}(x,\gamma(s))+d_{\varepsilon}(\gamma(s),\xi)
ε1ρε(x)(eε|xγ(s)|1)+ε(γ|[s,))\displaystyle\leq\varepsilon^{-1}\rho_{\varepsilon}(x)(e^{\varepsilon|x\gamma(s)|}-1)+\ell_{\varepsilon}(\gamma|_{[s,\infty)})
Cρε(x).\displaystyle\leq C\rho_{\varepsilon}(x).

Since dε(x)dε(x,ξ)d_{\varepsilon}(x)\leq d_{\varepsilon}(x,\xi) the upper bound follows.

For the lower bound we let ξXε\xi\in\partial X_{\varepsilon} be a given point, which we can think of as a point in ωX\partial_{\omega}X using Lemma 4.14. By rough starlikeness we can then find a geodesic line γ:X\gamma:\mathbb{R}\rightarrow X starting at ω\omega and ending at ξ\xi. For nn\in\mathbb{N} we note that |xγ(n)||x\gamma(n)|\rightarrow\infty as nn\rightarrow\infty, so we will have |xγ(n)|1|x\gamma(n)|\geq 1 for all sufficiently large nn. For sufficiently large nn we can then apply (4.17) and Lemma 2.7 to obtain

dε(x,γ(n))eε(x|γ(n))beε(x|ξ)b.d_{\varepsilon}(x,\gamma(n))\asymp e^{-\varepsilon(x|\gamma(n))_{b}}\asymp e^{-\varepsilon(x|\xi)_{b}}.

By (2.10) we have (x|ξ)bb(x)+600δ(x|\xi)_{b}\leq b(x)+600\delta. By combining this with the above we obtain that

dε(x,γ(n))C1ρε(x).d_{\varepsilon}(x,\gamma(n))\geq C^{-1}\rho_{\varepsilon}(x).

This gives the lower bound since dε(x,γ(n))dε(x,ξ)d_{\varepsilon}(x,\gamma(n))\rightarrow d_{\varepsilon}(x,\xi) as nn\rightarrow\infty. ∎

We can now complete the proof of Theorem 1.4 in the case b(X)b\in\mathcal{B}(X). Since we have already shown that XεX_{\varepsilon} is unbounded in Proposition 4.4, to prove Theorem 1.4 we only need to show that geodesics in XX are uniform curves in XεX_{\varepsilon}. For this we need to extend the definition of uniform curves to cover curves defined on arbitrary subintervals II\subset\mathbb{R}.

As in Definition 1.1, we let (Ω,d)(\Omega,d) be an incomplete metric space, set Ω=Ω¯\Ω\partial\Omega=\bar{\Omega}\backslash\Omega, and set dΩ(x)=dist(x,Ω)d_{\Omega}(x)=\mathrm{dist}(x,\partial\Omega). We consider a curve γ:IΩ\gamma:I\rightarrow\Omega defined on an arbitrary subinterval II\subset\mathbb{R}; we write t[,)t_{-}\in[-\infty,\infty) and t+(,]t_{+}\in(-\infty,\infty] for the endpoints of II. If γ\gamma has finite length, (γ)<\ell(\gamma)<\infty, then γ\gamma has well-defined endpoints γ,γ+Ω¯\gamma_{-},\gamma_{+}\in\bar{\Omega} defined by the limits γ(t)=limttγ(t)\gamma(t_{-})=\lim_{t\rightarrow t_{-}}\gamma(t) and γ(t+)=limtt+γ(t)\gamma(t_{+})=\lim_{t\rightarrow t_{+}}\gamma(t) in Ω¯\bar{\Omega}, which exist because (γ)<\ell(\gamma)<\infty.

Definition 4.16.

For a constant A1A\geq 1 and an interval II\subset\mathbb{R}, a curve γ:IΩ\gamma:I\rightarrow\Omega with (γ)<\ell(\gamma)<\infty is AA-uniform if

(4.19) (γ)Ad(γ,γ+),\ell(\gamma)\leq Ad(\gamma_{-},\gamma_{+}),

and if for every tIt\in I we have

(4.20) min{(γ|It),(γ|It)}AdΩ(γ(t)).\min\{\ell(\gamma|_{I_{\leq t}}),\ell(\gamma|_{I_{\geq t}})\}\leq Ad_{\Omega}(\gamma(t)).

If (γ)=\ell(\gamma)=\infty then we instead define γ\gamma to be AA-uniform if (4.20) holds and if d(γ(s),γ(t))d(\gamma(s),\gamma(t))\rightarrow\infty as sts\rightarrow t_{-} and tt+t\rightarrow t_{+}.

Proposition 4.17.

There is an A=A(δ,K,ε,M)A=A(\delta,K,\varepsilon,M) such that any geodesic in XX is an AA-uniform curve in XεX_{\varepsilon}. Consequently XεX_{\varepsilon} is AA-uniform.

Proof.

We first consider the case of a geodesic xyxy in XX joining two points x,yXωXx,y\in X\cup\partial_{\omega}X. Let γ:IX\gamma:I\rightarrow X be a parametrization of xyxy that is cc-adapted to bb, c=c(δ,K)c=c(\delta,K), as in Lemma 3.12; we can always find points x,yXωXx^{\prime},y^{\prime}\in X\cup\partial_{\omega}X satisfying the hypotheses of the lemma by the KK-rough starlikeness hypothesis from ω\omega. Let t[,)t_{-}\in[-\infty,\infty), t+(,]t_{+}\in(-\infty,\infty] be the endpoints of II. Let {sn},{tn}I\{s_{n}\},\{t_{n}\}\subset I be sequences such that snts_{n}\rightarrow t_{-}, tnt+t_{n}\rightarrow t_{+}, and sn<tns_{n}<t_{n} for each nn. Applying inequality (1.3) to the geodesic γ|[sn,tn]\gamma|_{[s_{n},t_{n}]} joining γ(sn)\gamma(s_{n}) to γ(tn)\gamma(t_{n}) gives

(4.21) ε(γ|[sn,tn])Mdε(γ(sn),γ(tn)).\ell_{\varepsilon}(\gamma|_{[s_{n},t_{n}]})\leq Md_{\varepsilon}(\gamma(s_{n}),\gamma(t_{n})).

Letting nn\rightarrow\infty, the left side of (4.21) converges to ε(γ)\ell_{\varepsilon}(\gamma). If xXx\in X then the sequence {γ(sn)}\{\gamma(s_{n})\} converges to xx in XX, while if xωXx\in\partial_{\omega}X then the sequence {γ(sn)}\{\gamma(s_{n})\} belongs to the equivalence class of XX with respect to ω\omega. In both cases we then have that {γ(sn)}\{\gamma(s_{n})\} converges to xx in X¯ε\bar{X}_{\varepsilon}, with the second case following from the construction of the identification φε:ωXXε\varphi_{\varepsilon}:\partial_{\omega}X\rightarrow\partial X_{\varepsilon}. The same discussion applies to the sequence {γ(tn)}\{\gamma(t_{n})\} in relation to yy. It follows that dε(γ(sn),γ(tn))dε(x,y)d_{\varepsilon}(\gamma(s_{n}),\gamma(t_{n}))\rightarrow d_{\varepsilon}(x,y) as nn\rightarrow\infty. Consequently γ\gamma has finite length in XεX_{\varepsilon} with endpoints γ=x\gamma_{-}=x and γ+=y\gamma_{+}=y. The inequality (4.19) then follows by letting nn\rightarrow\infty in (4.21).

We next verify inequality (4.20). It suffices to verify this inequality in the case that sI0s\in I_{\geq 0}, since we can deduce the case sI0s\in I_{\leq 0} from this by reversing the roles of xx and yy. We thus assume that sI0s\in I_{\geq 0}. A straightforward calculation with (3.7) gives us that

(4.22) ε(γ|Is)eε(x|y)bIseεt𝑑tε1eε(x|y)beεs.\ell_{\varepsilon}(\gamma|_{I_{\geq s}})\asymp e^{-\varepsilon(x|y)_{b}}\int_{I_{\geq s}}e^{-\varepsilon t}\,dt\leq\varepsilon^{-1}e^{-\varepsilon(x|y)_{b}}e^{-\varepsilon s}.

Since s+(x|y)bb(η(s))s+(x|y)_{b}\doteq b(\eta(s)), it then follows from (4.22) and Lemma 4.15 that

ε(γ|Is)Cρε(γ(s))Cdε(γ(s)),\ell_{\varepsilon}(\gamma|_{I_{\geq s}})\leq C\rho_{\varepsilon}(\gamma(s))\leq Cd_{\varepsilon}(\gamma(s)),

with C=C(δ,K,ε,M)C=C(\delta,K,\varepsilon,M). We conclude that γ\gamma is an AA-uniform curve in XεX_{\varepsilon} with A=A(δ,K,ε,M)A=A(\delta,K,\varepsilon,M). Since any two points x,yXx,y\in X can be joined by a geodesic xyxy in XX, this implies that the metric space XεX_{\varepsilon} is AA-uniform.

It remains to treat the case of a geodesic ωx\omega x joining the basepoint ωX\omega\in\partial X of bb to a point xXωXx\in X\cup\partial_{\omega}X. By applying (2) of Lemma 2.6 with u=0u=0 we can find an arclength parametrization γ:(,a]X\gamma:(-\infty,a]\rightarrow X of ωx\omega x, a(,]a\in(-\infty,\infty], with b(γ(t))144δtb(\gamma(t))\doteq_{144\delta}t for t(,a]t\in(-\infty,a]. For st(,a]s\leq t\in(-\infty,a] we then have, by a computation similar to the one done in Lemma 4.15,

(4.23) ε(γ|[s,t])eεseεt.\ell_{\varepsilon}(\gamma|_{[s,t]})\asymp e^{-\varepsilon s}-e^{-\varepsilon t}.

By letting ss\rightarrow-\infty we conclude that ε(γ)=\ell_{\varepsilon}(\gamma)=\infty. The comparison (4.23) together with the GH-inequality (1.3) then implies that dε(γ(s),γ(t))d_{\varepsilon}(\gamma(s),\gamma(t))\rightarrow\infty as ss\rightarrow-\infty and tat\rightarrow a. 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 s(,a]s\in(-\infty,a] that

ε(γ|Is)CeεsCρε(γ(s))Cdε(γ(s)),\ell_{\varepsilon}(\gamma|_{I_{\geq s}})\leq Ce^{-\varepsilon s}\leq C\rho_{\varepsilon}(\gamma(s))\leq Cd_{\varepsilon}(\gamma(s)),

with C=C(δ,K,ε,M)C=C(\delta,K,\varepsilon,M) in each inequality. Thus (4.20) also holds for γ\gamma with A=A(δ,K,ε,M)A=A(\delta,K,\varepsilon,M). We conclude that γ\gamma is an AA-uniform curve in XεX_{\varepsilon} 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 b𝒟(X)b\in\mathcal{D}(X) with basepoint zXz\in X. 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 (X,dX)(X,d_{X}) and a point zXz\in X we let Y=Xz0[0,)Y=X\cup_{z\sim 0}[0,\infty) be the metric space obtained by gluing the half-line [0,)[0,\infty) to XX by identifying the point zz with 0[0,)0\in[0,\infty). The metric dYd_{Y} on YY is defined by setting dY(x,y)=dX(x,y)d_{Y}(x,y)=d_{X}(x,y) for x,yXx,y\in X, dY(s,t)=|st|d_{Y}(s,t)=|s-t| for s,t[0,)s,t\in[0,\infty), and dY(x,s)=dX(x,z)+sd_{Y}(x,s)=d_{X}(x,z)+s for xXx\in X, s[0,)s\in[0,\infty). The space XX then has a canonical isometric embedding into YY. We refer to the metric space (Y,dY)(Y,d_{Y}) as the ray augmentation of XX based at zz.

The following trick will be the basis of many of the results we prove regarding the ray augmentation.

Lemma 4.19.

Let (X,dX)(X,d_{X}) be a metric space. Let (Y,dY)(Y,d_{Y}) be the ray augmentation of XX based at zXz\in X. Then for each curve γ:IY\gamma:I\rightarrow Y in YY there is a curve σ:IX\sigma:I\rightarrow X such that σ(t)=γ(t)\sigma(t)=\gamma(t) when γ(t)X\gamma(t)\in X and σ(t)=z\sigma(t)=z when γ(t)X\gamma(t)\notin X.

Proof.

Consider the retraction r:YXr:Y\rightarrow X given by setting r(x)=xr(x)=x for xXx\in X and r(t)=zr(t)=z for t[0,)t\in[0,\infty). It is easy to see that rr is 11-Lipschitz; in particular rr is continuous. For a given curve γ:IY\gamma:I\rightarrow Y the curve σ=rγ:IX\sigma=r\circ\gamma:I\rightarrow X then has the desired properties. ∎

We refer to the curve σ\sigma constructed from γ\gamma in Lemma 4.19 as the shortening of γ\gamma to XX. We apply Lemma 4.19 to conformal deformations of the ray augmentation.

Lemma 4.20.

Let (X,dX)(X,d_{X}) be a geodesic metric space and let (Y,dY)(Y,d_{Y}) be the ray augmentation of XX based at zXz\in X. Let ρ:X(0,)\rho:X\rightarrow(0,\infty) be a continuous density on XX and let ρ~:Y(0,)\tilde{\rho}:Y\rightarrow(0,\infty) be a continuous density on YY with ρ~|X=ρ\tilde{\rho}|_{X}=\rho. Then the isometric embedding XYX\rightarrow Y induces an isometric embedding XρYρ~X_{\rho}\rightarrow Y_{\tilde{\rho}} of the corresponding conformal deformations. Furthermore ρ\rho is a GH-density with constant M1M\geq 1 if and only if ρ~\tilde{\rho} is a GH-density with the same constant MM.

Proof.

We write dρd_{\rho} for the metric on XρX_{\rho} and dρ~d_{\tilde{\rho}} for the metric on Yρ~Y_{\tilde{\rho}}. Let x,yXx,y\in X be given. Let γ:IY\gamma:I\rightarrow Y be a rectifiable curve joining them and let σ:IX\sigma:I\rightarrow X be the shortening of γ\gamma to XX. Since ρ~|X=ρ\tilde{\rho}|_{X}=\rho we clearly have

σρ~𝑑s=σρ𝑑s=γ1(X)ρ𝑑sγρ~𝑑s.\int_{\sigma}\tilde{\rho}\,ds=\int_{\sigma}\rho\,ds=\int_{\gamma^{-1}(X)}\rho\,ds\leq\int_{\gamma}\tilde{\rho}\,ds.

It follows that ρ~(σ)ρ~(γ)\ell_{\tilde{\rho}}(\sigma)\leq\ell_{\tilde{\rho}}(\gamma). Thus in computing dρ~(x,y)d_{\tilde{\rho}}(x,y) it suffices to minimize ρ~(γ)\ell_{\tilde{\rho}}(\gamma) over curves γ:IX\gamma:I\rightarrow X taking values only in XX. Since for such curves we have ρ~(γ)=ρ(γ)\ell_{\tilde{\rho}}(\gamma)=\ell_{\rho}(\gamma), it immediately follows that dρ~(x,y)=dρ(x,y)d_{\tilde{\rho}}(x,y)=d_{\rho}(x,y). We conclude that the embedding XρYρ~X_{\rho}\rightarrow Y_{\tilde{\rho}} is an isometry.

It is obvious that ρ\rho is a GH-density with constant MM if ρ~\tilde{\rho} is a GH-density with constant MM, since ρ~\tilde{\rho} restricts to ρ\rho on XX and geodesics in XX are also geodesics in YY. For the converse we assume that ρ\rho is a GH-density with constant MM and let x,yYx,y\in Y be given points. We need to prove the GH-inequality (1.3) for any geodesic γ\gamma joining xx to yy in YY.

If x,yXx,y\in X then any geodesic γ\gamma joining xx to yy in YY is in fact a geodesic joining xx to yy in XX. The inequality (1.3) for ρ~\tilde{\rho} then follows from the corresponding inequality for ρ\rho. If x,y[0,)=Y\(X\{z})x,y\in[0,\infty)=Y\backslash(X\backslash\{z\}) then there is only one geodesic γ\gamma from xx to yy in YY, which is simply the interval connecting them in [0,)[0,\infty). Since any curve σ\sigma joining xx to yy in YY must contain this interval we in fact have ρ~(γ)=dρ~(x,y)\ell_{\tilde{\rho}}(\gamma)=d_{\tilde{\rho}}(x,y). Thus inequality (1.3) holds in this case as well.

The final case is that in which xXx\in X and y[0,)y\in[0,\infty). Let γ\gamma be a geodesic joining xx to yy in YY. Then we can write γ=ση\gamma=\sigma\cup\eta, where σ\sigma is a geodesic in XX joining xx to zz and η\eta is the geodesic in [0,)[0,\infty) joining 0 to yy, which is just the interval connecting these points. Given a rectifiable curve α:IY\alpha:I\rightarrow Y joining xx to yy we let β\beta be the shortening of α\alpha to XX. Then by the inequality (1.3) for ρ\rho we have ρ(σ)Mρ(β)\ell_{\rho}(\sigma)\leq M\ell_{\rho}(\beta). Since the intersection α(I)[0,)\alpha(I)\cap[0,\infty) must contain the interval η\eta joining 0 to yy, we deduce from this that

ρ~(γ)\displaystyle\ell_{\tilde{\rho}}(\gamma) =ρ~(σ)+ρ~(η)\displaystyle=\ell_{\tilde{\rho}}(\sigma)+\ell_{\tilde{\rho}}(\eta)
M(ρ~(β)+α1([0,))ρ~𝑑s)\displaystyle\leq M\left(\ell_{\tilde{\rho}}(\beta)+\int_{\alpha^{-1}([0,\infty))}\tilde{\rho}\,ds\right)
=Mρ~(α).\displaystyle=M\ell_{\tilde{\rho}}(\alpha).

Minimizing over all rectifiable curves α\alpha joining xx to yy then gives inequality (1.3). ∎

We assume now that XX is a geodesic δ\delta-hyperbolic space. We let YY be the ray augmentation of XX based at some point zXz\in X. It is an easy exercise to see that YY is also δ\delta-hyperbolic; recall that we have defined δ\delta-hyperbolicity using δ\delta-thin triangles. We also note that YY is complete if XX is complete. We will continue to use the generic distance notation |xy||xy| for the distance between x,yYx,y\in Y, noting that there is no conflict with the distance notation for XX since XX sits isometrically inside of YY.

By definition the canonical ray in YY is the geodesic ray γ:[0,)Y\gamma:[0,\infty)\rightarrow Y corresponding to the canonical parametrization of the copy of [0,)[0,\infty) that we glued onto XX. The key property of the ray augmentation is that the Busemann function bγb_{\gamma} associated to the canonical ray restricts on XX to the distance from the distinguished point zz.

Lemma 4.21.

Let γ\gamma be the canonical ray in YY. Then for xXx\in X we have bγ(x)=|xz|b_{\gamma}(x)=|xz|.

Proof.

For xXx\in X and t0t\geq 0 we have

|γ(t)x|t=t+|xz|t=|xz|,|\gamma(t)x|-t=t+|xz|-t=|xz|,

which implies upon taking tt\rightarrow\infty that bγ(x)=|xz|b_{\gamma}(x)=|xz|. ∎

Let ωY\omega\in\partial Y be the point in the Gromov boundary defined by the canonical ray. We next show that ωY=Y\{ω}\partial_{\omega}Y=\partial Y\backslash\{\omega\} identifies canonically with X\partial X and rough starlikeness from zz in XX passes over to rough starlikeness from ω\omega in YY.

Lemma 4.22.

We have ωY=X\partial_{\omega}Y=\partial X. Furthermore any visual metric on X\partial X based at zz with parameter ε>0\varepsilon>0 also defines a visual metric on ωY\partial_{\omega}Y based at ω\omega with parameter ε\varepsilon, and the converse holds as well. If XX is KK-roughly starlike from zz then YY is KK-roughly starlike from ω\omega.

Proof.

For the first assertion it suffices to show that if {xn}\{x_{n}\} is a sequence converging to infinity in YY then there is an NN\in\mathbb{N} such that for nNn\geq N the points xnx_{n} either all belong to XX or all belong to [0,)[0,\infty). Recall that {xn}\{x_{n}\} converges to infinity if we have (xn|xm)z(x_{n}|x_{m})_{z}\rightarrow\infty as m,nm,n\rightarrow\infty. If our assertion did not hold then we could find subsequences {yn}\{y_{n}\} and {zn}\{z_{n}\} of the sequence {xn}\{x_{n}\} such that {yn}X\{y_{n}\}\subset X, {zn}[0,)\{z_{n}\}\subset[0,\infty), and (yn|zn)z(y_{n}|z_{n})_{z}\rightarrow\infty as nn\rightarrow\infty. But then |ynzn|=|ynz|+|znz||y_{n}z_{n}|=|y_{n}z|+|z_{n}z| and therefore

(yn|zn)z=12(|ynz|+|znz||ynzn|)=0,(y_{n}|z_{n})_{z}=\frac{1}{2}(|y_{n}z|+|z_{n}z|-|y_{n}z_{n}|)=0,

contradicting our assumption that (yn|zn)z(y_{n}|z_{n})_{z}\rightarrow\infty. Since all sequences {xn}\{x_{n}\} converging to infinity in YY that belong exclusively to [0,)[0,\infty) must converge to ω\omega, it follows that we have a canonical identification of X\partial X with ωY\partial_{\omega}Y.

Let bγb_{\gamma} be the Busemann function associated to the canonical ray γ\gamma as in Lemma 4.21. We observe that bγ(x)=|xz|b_{\gamma}(x)=|xz| for xXx\in X implies that (x|y)bγ=(x|y)z(x|y)_{b_{\gamma}}=(x|y)_{z} for x,yXx,y\in X. Since any sequence {xn}\{x_{n}\} converging to infinity in YY that does not converge to ω\omega must eventually stay within XX, it follows that (ξ|ζ)bγ=(ξ|ζ)z(\xi|\zeta)_{b_{\gamma}}=(\xi|\zeta)_{z} for ξ,ζX\xi,\zeta\in\partial X. Thus through our identification ωY=X\partial_{\omega}Y=\partial X we have a canonical identification between visual metrics on ωY\partial_{\omega}Y based at ω\omega and visual metrics on X\partial X based at zz for any parameter ε>0\varepsilon>0.

Observe that if σ:[0,)X\sigma:[0,\infty)\rightarrow X is a geodesic ray starting at zz then the map σ~:Y\tilde{\sigma}:\mathbb{R}\rightarrow Y defined by σ~(t)=σ(t)\tilde{\sigma}(t)=\sigma(t) for t0t\geq 0 and σ~(t)=t[0,)\tilde{\sigma}(t)=-t\in[0,\infty) for t0t\leq 0 defines a geodesic line in YY that begins at ω\omega, coincides with σ\sigma inside of XX, and has the same endpoint in X=ωY\partial X=\partial_{\omega}Y as σ\sigma. This implies that if XX is KK-roughly starlike from zz for some K0K\geq 0 then YY is KK-roughly starlike from ω\omega. ∎

We can now complete the proofs of Theorems 1.4 and 1.6 by showing that they also hold for b𝒟(X)b\in\mathcal{D}(X).

Proposition 4.23.

Theorems 1.4 and 1.6 hold for b𝒟(X)b\in\mathcal{D}(X).

Proof.

Let XX be a complete geodesic δ\delta-hyperbolic space that is KK-roughly starlike from zXz\in X and let b𝒟(X)b\in\mathcal{D}(X) have the form b(x)=|xz|+sb(x)=|xz|+s for some ss\in\mathbb{R}. We assume that ε>0\varepsilon>0 is such that ρε,b\rho_{\varepsilon,b} is a GH-density with constant MM. We let Y=Xz0[0,)Y=X\cup_{z\sim 0}[0,\infty) be the ray augmentation of XX based at zz, let γ\gamma be the canonical ray in YY, and set b~=bγ+s\tilde{b}=b_{\gamma}+s. Then YY is a complete geodesic δ\delta-hyperbolic space that is KK-roughly starlike from the endpoint ωY\omega\in\partial Y of the canonical ray by Lemma 4.22. We have b~|X=b\tilde{b}|_{X}=b by Lemma 4.21. Thus by Lemma 4.20 the embedding Xε,bYε,b~X_{\varepsilon,b}\rightarrow Y_{\varepsilon,\tilde{b}} is isometric and ρε,b~\rho_{\varepsilon,\tilde{b}} is a GH-density with the same constant MM. We can then apply Theorems 1.4 and 1.6 to YY equipped with the Busemann function b~\tilde{b}.

For t[0,)Yt\in[0,\infty)\subset Y we have b~(t)=t+s\tilde{b}(t)=-t+s. Thus a straightforward calculation shows that ε(γ)=\ell_{\varepsilon}(\gamma)=\infty. It follows from this and the GH-inequality (1.3) that the only boundary points of the metric space Yε,b~Y_{\varepsilon,\tilde{b}} are the boundary points of Xε,bX_{\varepsilon,b}, i.e., Xε,b=Yε,b~\partial X_{\varepsilon,b}=\partial Y_{\varepsilon,\tilde{b}}. By applying Theorem 1.6 to YY and then using Lemma 4.22, we conclude that we have a canonical identification φε:XXε,b\varphi_{\varepsilon}:\partial X\rightarrow\partial X_{\varepsilon,b}. Since visual metrics on X\partial X based at zz with parameter ε\varepsilon correspond to visual metrics on ωY\partial_{\omega}Y based at ω\omega with the same parameter ε\varepsilon, it then follows from Theorem 1.6 that the restriction of dε,bd_{\varepsilon,b} to Xε,b\partial X_{\varepsilon,b} defines a visual metric on ωX\partial_{\omega}X with parameter ε\varepsilon and comparison constant L=L(δ,K,ε,M)L=L(\delta,K,\varepsilon,M). This completes the proof of Theorem 1.6.

For Theorem 1.4 we observe that any geodesic σ\sigma in XX is also a geodesic in YY. Since Xε,bX_{\varepsilon,b} sits isometrically inside of Yε,b~Y_{\varepsilon,\tilde{b}} and Xε,b=Yε,b~\partial X_{\varepsilon,b}=\partial Y_{\varepsilon,\tilde{b}}, by applying Theorem 1.4 to YY we conclude that σ\sigma is an AA-uniform curve in Yε,b~Y_{\varepsilon,\tilde{b}} and therefore also an AA-uniform curve in Xε,bX_{\varepsilon,b}. This completes the proof of Theorem 1.4 in this case. ∎

Remark 4.24.

The proof of Proposition 4.23 also shows that the estimates of Lemmas 4.13 and 4.15 hold for b𝒟(X)b\in\mathcal{D}(X) as well. In the notation of the proof, this is because b~|X=b\tilde{b}|_{X}=b, Xε,bX_{\varepsilon,b} isometrically embeds into Yε,b~Y_{\varepsilon,\tilde{b}}, and we have an identification of boundaries Xε,b=Yε,b~\partial X_{\varepsilon,b}=\partial Y_{\varepsilon,\tilde{b}}.

5. Hyperbolic fillings

Let (Z,d)(Z,d) be a metric space and let α,τ>1\alpha,\tau>1 be given parameters. We recall the construction of a hyperbolic filling XX of ZZ with these parameters described prior to Theorem 1.12. For each nn\in\mathbb{Z} we select a maximal αn\alpha^{-n}-separated subset SnS_{n} of ZZ. The existence of such a set is guaranteed by a standard application of Zorn’s lemma. Then for each nn\in\mathbb{Z} the balls B(z,αn)B(z,\alpha^{-n}), zSnz\in S_{n}, cover ZZ.

The vertex set of XX has the form

V=nVn,Vn={(x,n):xSn}.V=\bigcup_{n\in\mathbb{Z}}V_{n},\;\;\;V_{n}=\{(x,n):x\in S_{n}\}.

To each vertex v=(x,n)v=(x,n) we associate the dilated ball B(v)=B(x,ταn)B(v)=B(x,\tau\alpha^{-n}). We will often use vv to denote both a vertex in XX and its associated point in ZZ. We also define the height function h:Vh:V\rightarrow\mathbb{Z} by h(x,n)=nh(x,n)=n. By construction for each zZz\in Z there is a vVnv\in V_{n} such that ρ(v,z)<αn\rho(v,z)<\alpha^{-n}.

We place an edge in XX between distinct vertices vv and ww if and only if |h(v)h(w)|1|h(v)-h(w)|\leq 1 and B(v)B(w)B(v)\cap B(w)\neq\emptyset. 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 v,wv,w we write vwv\sim w if there is an edge between vv and ww. 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 α>1\alpha>1, we will need to place some constraints on the values of the parameter τ\tau based on α\alpha. We will require that

(5.1) τ>max{3,αα1}.\tau>\max\left\{3,\frac{\alpha}{\alpha-1}\right\}.

We give each connected component of XX 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 XX is actually connected and is therefore a geodesic metric space itself. For applications a standard choice of parameters satisfying (5.1) is given by α=2\alpha=2 and τ=4\tau=4.

Remark 5.1.

We do not know whether the constraint (5.1) can be relaxed while preserving the properties of XX described below. In particular we do not know whether XX is always Gromov hyperbolic or even connected for all τ>1\tau>1. However, by applying Lemma 5.2 below it is easy to see that XX is connected for any τ>1\tau>1 when ZZ is bounded. We note that one cannot take τ=1\tau=1 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 XX. We will refer to these vertical paths as vertical geodesics. We will use the generic distance notation |xy||xy| for the distance between x,yXx,y\in X. Thus for v=(x,n),w=(y,n)Vv=(x,n),w=(y,n)\in V we will denote their distance in XX by |vw||vw| and their distance in ZZ by d(v,w):=d(x,y)d(v,w):=d(x,y). Identifying an edge gg from a vertex vv to a vertex ww isometrically with [0,1][0,1], we extend the height function hh to gg by h(s)=sh(v)+(1s)h(w)h(s)=sh(v)+(1-s)h(w). Then hh defines a function h:Xh:X\rightarrow\mathbb{R} that is 11-Lipschitz on the connected components of XX.

We begin with a simple lemma.

Lemma 5.2.

Let v,wVv,w\in V with h(v)h(w)h(v)\neq h(w) and B(v)B(w)B(v)\cap B(w)\neq\emptyset. Then there is a vertical edge path from vv to ww.

Proof.

Let v=(x,m)v=(x,m), w=(y,n)w=(y,n), and let zB(v)B(w)z\in B(v)\cap B(w). We can assume without loss of generality that m<nm<n. For each integer mknm\leq k\leq n we can find a vertex vkVkv_{k}\in V_{k} with zB(vk)z\in B(v_{k}); we set vm=vv_{m}=v and vn=wv_{n}=w. Then vkvk+1v_{k}\sim v_{k+1} for each mk<nm\leq k<n by the construction of the graph XX. It follows that vv is connected to ww by a vertical edge path passing through the vertices vkv_{k}. ∎

The next lemma estimates the distance in ZZ between vertices in XX that are connected by a vertical edge path.

Lemma 5.3.

Let v,wVv,w\in V. Suppose that vv is joined to ww by a vertical edge path and h(v)h(w)h(v)\leq h(w). Then

d(v,w)2ταh(v)+1α1.d(v,w)\leq\frac{2\tau\alpha^{-h(v)+1}}{\alpha-1}.
Proof.

We first derive a sharper inequality in the case h(w)=h(v)+1h(w)=h(v)+1. Set h(v)=mh(v)=m. Let xB(v)B(w)x\in B(v)\cap B(w). Then

d(v,w)d(x,v)+d(x,w)<ταm+ταm1<2ταm.d(v,w)\leq d(x,v)+d(x,w)<\tau\alpha^{-m}+\tau\alpha^{-m-1}<2\tau\alpha^{-m}.

Now let h(v)=mh(v)=m, h(w)=nh(w)=n. For each mknm\leq k\leq n we let vkVkv_{k}\in V_{k} be the vertex at this height in the vertical edge path joining vv to ww. Then by the “h(w)=h(v)+1h(w)=h(v)+1” case we have

d(v,w)k=mn1d(vk,vk+1)2ταmk=0nm1αk2ταm+1α1,d(v,w)\leq\sum_{k=m}^{n-1}d(v_{k},v_{k+1})\leq 2\tau\alpha^{-m}\sum_{k=0}^{n-m-1}\alpha^{-k}\leq\frac{2\tau\alpha^{-m+1}}{\alpha-1},

with the final inequality following by summing the geometric series in α1\alpha^{-1}. ∎

Following the hyperbolic filling construction in [12], we define a cone point uVu\in V for a pair of vertices {v,w}V\{v,w\}\subseteq V to be a vertex that can be joined to both vv and ww by vertical geodesics and that satisfies h(u)min{h(v),h(w)}h(u)\leq\min\{h(v),h(w)\}. A branch point for {v,w}\{v,w\} is defined to be a cone point of maximal height. A branch point for {v,w}\{v,w\} always exists as long as there is at least one cone point for {v,w}\{v,w\}. When v=wv=w the vertex vv is trivially a branch point for the set {v}\{v\}.

Lemma 5.4.

Let v,wVnv,w\in V_{n} be distinct vertices with vwv\sim w. Then there is a branch point uVn1u\in V_{n-1} for the set {v,w}\{v,w\}.

Proof.

The assumptions imply that B(v)B(w)B(v)\cap B(w)\neq\emptyset. Let zB(v)B(w)z\in B(v)\cap B(w) be a point in this intersection. Since Vn1V_{n-1} is a maximal αn+1\alpha^{-n+1}-separated set in ZZ we can find uVn1u\in V_{n-1} such that d(u,z)<αn+1d(u,z)<\alpha^{-n+1}. We compute

d(v,u)d(v,z)+d(z,u)<ταn+αn+1<ταn+1,d(v,u)\leq d(v,z)+d(z,u)<\tau\alpha^{-n}+\alpha^{-n+1}<\tau\alpha^{-n+1},

by inequality (5.1), noting that the final inquality here is equivalent to

τ+α<τα,\tau+\alpha<\tau\alpha,

which is implied by (5.1). It follows that vB(u)v\in B(u) and therefore B(v)B(u)B(v)\cap B(u)\neq\emptyset. Thus vv is joined to uu by a vertical edge. Since the roles of vv and ww are symmetric, we conclude by the same calculation that B(w)B(u)B(w)\cap B(u)\neq\emptyset, i.e., ww is also joined to uu by a vertical edge. Thus uu is a cone point for {v,w}\{v,w\}. Since a cone point for a pair of distinct vertices on an adjacent level is trivially maximal, we conclude that uu is a branch point for {v,w}\{v,w\}. ∎

We can now show that the graph XX is connected.

Proposition 5.5.

For each v,wVv,w\in V there is a branch point uu for the set {v,w}\{v,w\} that satisfies

(5.2) αh(u)C(α,τ)d(v,w)+αmin{h(v),h(w)}.\alpha^{-h(u)}\asymp_{C(\alpha,\tau)}d(v,w)+\alpha^{-\min\{h(v),h(w)\}}.

Consequently the graph XX is connected.

Proof.

Let vVmv\in V_{m}, wVnw\in V_{n} be given. We can assume without loss of generality that mnm\leq n. If v=wv=w then vv is a branch point for the set {v}\{v\} and the comparison (5.2) holds trivially. We can thus assume for the rest of the proof that vwv\neq w. We let kk\in\mathbb{Z} be any integer satisfying αk>d(v,w)\alpha^{-k}>d(v,w) and kmk\leq m; note that such an integer always exists since αk\alpha^{-k}\rightarrow\infty as kk\rightarrow-\infty. Let pVkp\in V_{k} be a vertex such that d(v,p)<αkd(v,p)<\alpha^{-k} and let qVkq\in V_{k} be a vertex such that d(q,w)<αkd(q,w)<\alpha^{-k}. Then

d(p,q)d(v,p)+d(v,w)+d(w,q)<3αk<ταk,d(p,q)\leq d(v,p)+d(v,w)+d(w,q)<3\alpha^{-k}<\tau\alpha^{-k},

by (5.1). Thus qB(p)q\in B(p), so B(p)B(q)B(p)\cap B(q)\neq\emptyset. We conclude that pqp\sim q. By Lemma 5.4 we can then find a branch point xVk1x\in V_{k-1} for the set {p,q}\{p,q\}. Since B(p)B(v)B(p)\cap B(v)\neq\emptyset and B(q)B(w)B(q)\cap B(w)\neq\emptyset, Lemma 5.2 shows that pp and qq are connected to vv and ww respectively by vertical edge paths, and the requirement kmk\leq m implies that max{h(p),h(q)}min{h(v),h(w)}\max\{h(p),h(q)\}\leq\min\{h(v),h(w)\}. Since pp and qq are each connected to xx by a vertical edge, we conclude that xx is a cone point for the set {v,w}\{v,w\}.

It follows that there is a branch point uu for the set {v,w}\{v,w\}. Since uu is joined to vv and ww by vertical edge paths, the triangle inequality and Lemma 5.3 implies that

d(v,w)2max{d(v,u),d(w,u)}C(α,τ)αh(u).d(v,w)\leq 2\max\{d(v,u),d(w,u)\}\leq C(\alpha,\tau)\alpha^{-h(u)}.

Since h(u)mh(u)\leq m, we have αmαh(u)\alpha^{-m}\leq\alpha^{-h(u)} and therefore

d(v,w)+αmC(α,τ)αh(u)+αmC(α,τ)αh(u),d(v,w)+\alpha^{-m}\leq C(\alpha,\tau)\alpha^{-h(u)}+\alpha^{-m}\leq C(\alpha,\tau)\alpha^{-h(u)},

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 B(v)B(w)B(v)\cap B(w)\neq\emptyset. If h(v)=h(w)h(v)=h(w) then this implies that vwv\sim w and Lemma 5.4 implies that h(u)=h(v)1h(u)=h(v)-1. The upper bound follows immediately from this, as we then have

d(v,w)+αmin{h(v),h(w)}αh(v)=α1αh(u).d(v,w)+\alpha^{-\min\{h(v),h(w)\}}\geq\alpha^{-h(v)}=\alpha^{-1}\alpha^{-h(u)}.

If h(v)h(w)h(v)\neq h(w) then by Lemma 5.2 vv can be joined to ww by a vertical edge path. In this case vv is a branch point for the set {v,w}\{v,w\} and the inequality

d(v,w)+αh(v)αh(u),d(v,w)+\alpha^{-h(v)}\geq\alpha^{-h(u)},

holds trivially for u=vu=v.

The second case is that in which have B(v)B(w)=B(v)\cap B(w)=\emptyset. This implies in particular that we must have wB(v)w\notin B(v). Thus d(v,w)ταm>0d(v,w)\geq\tau\alpha^{-m}>0. Let kk\in\mathbb{Z} be the maximal integer such that kmk\leq m and αk>d(v,w)\alpha^{-k}>d(v,w). Then either k=mk=m or d(v,w)αk1d(v,w)\geq\alpha^{-k-1}. Since αk>d(v,w)\alpha^{-k}>d(v,w) and d(v,w)ταmd(v,w)\geq\tau\alpha^{-m}, we conclude in both cases that d(v,w)C(α,τ)αkd(v,w)\asymp_{C(\alpha,\tau)}\alpha^{-k}. Making this choice of kk in the construction of xx above, we can thus construct a cone point xx for the set {v,w}\{v,w\} with h(x)=k1h(x)=k-1 and therefore

αh(x)C(α,τ)d(v,w).\alpha^{-h(x)}\asymp_{C(\alpha,\tau)}d(v,w).

Since the branch point uu satisfies h(u)h(x)h(u)\geq h(x) it follows that

αh(u)C(α,τ)d(v,w)C(α,τ)(d(v,w)+αm).\alpha^{-h(u)}\leq C(\alpha,\tau)d(v,w)\leq C(\alpha,\tau)(d(v,w)+\alpha^{-m}).

The upper bound in (5.2) follows.

Lastly, since we can connect vv to ww through the branch point uu, it follows that vv and ww can be connected by an edge path in the graph XX. Since vv and ww were arbitrary we conclude that XX is connected. ∎

Now that we’ve shown XX is connected, the metrics we put on its connected components give it the structure of a geodesic metric space in which all edges of XX have unit length. The height function then defines a 11-Lipschitz function h:Xh:X\rightarrow\mathbb{R}. We formally define the Gromov product based at hh by, for x,yXx,y\in X,

(x|y)h=12(h(x)+h(y)|xy|).(x|y)_{h}=\frac{1}{2}(h(x)+h(y)-|xy|).

Since hh is 11-Lipschitz we have

(5.3) (x|y)hmin{h(x),h(y)}.(x|y)_{h}\leq\min\{h(x),h(y)\}.

Our next lemma gives a key relation of the Gromov product based at hh to branch points.

Lemma 5.6.

Let v,wVv,w\in V and let uu be a branch point for {v,w}\{v,w\}. Then

h(u)c(α,τ)(v|w)h,h(u)\doteq_{c(\alpha,\tau)}(v|w)_{h},

and therefore

α(v|w)hC(α,τ)d(v,w)+αmin{h(v),h(w)}.\alpha^{-(v|w)_{h}}\asymp_{C(\alpha,\tau)}d(v,w)+\alpha^{-\min\{h(v),h(w)\}}.
Proof.

Since the claims of the lemma hold trivially when v=wv=w we can assume that vwv\neq w. Proposition 5.5 gives the existence of a branch point uu for {v,w}\{v,w\} satisfying (5.2). The vertical edge path from vv to uu followed by the vertical edge path from uu to ww gives an edge path from vv to ww, which shows that

|vw||vu|+|uw|=h(v)h(u)+h(w)h(u)=h(v)+h(w)2h(u).|vw|\leq|vu|+|uw|=h(v)-h(u)+h(w)-h(u)=h(v)+h(w)-2h(u).

Rearranging this we obtain

h(u)12(h(v)+h(w)|vw|)=(v|w)h.h(u)\leq\frac{1}{2}(h(v)+h(w)-|vw|)=(v|w)_{h}.

To get a bound in the other direction, let v=v0,v1,,vk=wv=v_{0},v_{1},\dots,v_{k}=w be a sequence of vertices joined by edges that gives a geodesic γ\gamma from vv to ww. Then |vw|=k|vw|=k and k1k\geq 1 since vwv\neq w. For 1ik1\leq i\leq k we have B(vi1)B(vi)B(v_{i-1})\cap B(v_{i})\neq\emptyset and therefore, using |h(vi1)h(vi)|1|h(v_{i-1})-h(v_{i})|\leq 1,

d(vi1,vi)<2ταmin{h(vi1),h(vi)}2ταh(vi1)+1.d(v_{i-1},v_{i})<2\tau\alpha^{-\min\{h(v_{i-1}),h(v_{i})\}}\leq 2\tau\alpha^{-h(v_{i-1})+1}.

We can run the same argument viewing γ\gamma as a geodesic from ww to vv instead, setting wi=vkiw_{i}=v_{k-i} for 0ik0\leq i\leq k. We see from this that we also have

d(wi1,wi)<2ταh(wi1)+1,d(w_{i-1},w_{i})<2\tau\alpha^{-h(w_{i-1})+1},

for 1ik1\leq i\leq k. For each 1lk1\leq l\leq k we thus obtain an estimate (using h(vi1)h(v)i+1h(v_{i-1})\geq h(v)-i+1 and h(wi1)h(w)i+1h(w_{i-1})\geq h(w)-i+1),

d(v,w)\displaystyle d(v,w) i=1kd(vi1,vi)\displaystyle\leq\sum_{i=1}^{k}d(v_{i-1},v_{i})
i=1ld(vi1,vi)+i=1kl+1d(wi1,wi)\displaystyle\leq\sum_{i=1}^{l}d(v_{i-1},v_{i})+\sum_{i=1}^{k-l+1}d(w_{i-1},w_{i})
<2ταh(v)i=1lαi+2ταh(w)i=1kl+1αi\displaystyle<2\tau\alpha^{-h(v)}\sum_{i=1}^{l}\alpha^{i}+2\tau\alpha^{-h(w)}\sum_{i=1}^{k-l+1}\alpha^{i}
2ταα1(αh(v)(αl1)+αh(w)(αkl+11))\displaystyle\leq\frac{2\tau\alpha}{\alpha-1}(\alpha^{-h(v)}(\alpha^{l}-1)+\alpha^{-h(w)}(\alpha^{k-l+1}-1))
2ταα1(αlh(v)+αkl+1h(w)).\displaystyle\leq\frac{2\tau\alpha}{\alpha-1}(\alpha^{l-h(v)}+\alpha^{k-l+1-h(w)}).

We set l=12(kh(w)+h(v))l=\lceil\frac{1}{2}(k-h(w)+h(v))\rceil (the least integer greater than this quantity), observing that 1lk1\leq l\leq k since |h(v)h(w)|k|h(v)-h(w)|\leq k. This gives, after some simplification,

d(v,w)C(α,τ)α12(h(v)+h(w)k)=C(α,τ)α(v|w)h,d(v,w)\leq C(\alpha,\tau)\alpha^{-\frac{1}{2}(h(v)+h(w)-k)}=C(\alpha,\tau)\alpha^{-(v|w)_{h}},

recalling that |vw|=k|vw|=k. By Proposition 5.5 and inequality (5.3), we then have

αh(u)C(α,τ)α(v|w)h,\alpha^{-h(u)}\leq C(\alpha,\tau)\alpha^{-(v|w)_{h}},

which implies upon taking logarithms that

h(u)(v|w)hc(α,τ).h(u)\geq(v|w)_{h}-c(\alpha,\tau).

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 4δ4\delta-inequality (2.3) for our formal Gromov products based at hh.

Lemma 5.7.

Let u,v,wVu,v,w\in V. Then

(u|w)hmin{(u|v)h,(v|w)h}c(α,τ).(u|w)_{h}\geq\min\{(u|v)_{h},(v|w)_{h}\}-c(\alpha,\tau).
Proof.

Let u,v,wVu,v,w\in V be vertices. By the triangle inequality in ZZ we have

d(u,w)+αmin{h(u),h(w)}d(u,v)+αmin{h(u),h(v)}+d(v,w)+αmin{h(v),h(w)},d(u,w)+\alpha^{-\min\{h(u),h(w)\}}\leq d(u,v)+\alpha^{-\min\{h(u),h(v)\}}+d(v,w)+\alpha^{-\min\{h(v),h(w)\}},

which becomes, upon applying Lemma 5.6,

α(u|w)h\displaystyle\alpha^{-(u|w)_{h}} C(α,τ)(α(u|v)h+α(v|w)h)\displaystyle\leq C(\alpha,\tau)(\alpha^{-(u|v)_{h}}+\alpha^{-(v|w)_{h}})
C(α,τ)αmin{(u|v)h,(v|w)h}.\displaystyle\leq C(\alpha,\tau)\alpha^{-\min\{(u|v)_{h},(v|w)_{h}\}}.

Taking logarithms of each side gives the desired inequality. ∎

We can now show that XX is Gromov hyperbolic. For this we use some terminology from [12, Chapter 2]: a δ\delta-triple for δ0\delta\geq 0 is a triple (a,b,c)(a,b,c) of real numbers a,b,ca,b,c such that the two smallest numbers differ by at most δ\delta. Observe that (a,b,c)(a,b,c) is a δ\delta-triple if and only if the inequality

(5.4) cmin{a,b}δ,c\geq\min\{a,b\}-\delta,

holds for all permutations of the roles of aa, bb, and cc. We will also need the following standard claim [12, Lemma 2.1.4] which is called the Tetrahedron lemma.

Lemma 5.8.

Let d12d_{12}, d13d_{13}, d14d_{14}, d23d_{23}, d24d_{24}, d34d_{34} be six numbers such that the four triples (d23,d24,d34)(d_{23},d_{24},d_{34}), (d13,d14,d34)(d_{13},d_{14},d_{34}), (d12,d14,d24)(d_{12},d_{14},d_{24}), and (d12,d13,d23)(d_{12},d_{13},d_{23}) are δ\delta-triples. Then

(d12+d34,d13+d24,d14+d23)(d_{12}+d_{34},d_{13}+d_{24},d_{14}+d_{23})

is a 2δ2\delta-triple.

Proposition 5.9.

The space XX is δ\delta-hyperbolic with δ=δ(α,τ)\delta=\delta(\alpha,\tau).

Proof.

We will use the cross-difference triple defined in [12, Chapter 2.4]. For a quadruple of points Q=(x,y,z,u)XQ=(x,y,z,u)\in X and a fixed basepoint oXo\in X this triple is defined by

Ao(Q)=((x|y)o+(z|u)o,(x|z)o+(y|u)o,(x|u)o+(y|z)o).A_{o}(Q)=((x|y)_{o}+(z|u)_{o},(x|z)_{o}+(y|u)_{o},(x|u)_{o}+(y|z)_{o}).

The triple Ao(Q)A_{o}(Q) has the same differences among its members as the triple

Ah(Q)=((x|y)h+(z|u)h,(x|z)h+(y|u)h,(x|u)h+(y|z)h),A_{h}(Q)=((x|y)_{h}+(z|u)_{h},(x|z)_{h}+(y|u)_{h},(x|u)_{h}+(y|z)_{h}),

as a routine calculation shows for instance that

(x|y)o+(z|u)o(x|z)o(y|u)o=(x|y)h+(z|u)h(x|z)h(y|u)h,(x|y)_{o}+(z|u)_{o}-(x|z)_{o}-(y|u)_{o}=(x|y)_{h}+(z|u)_{h}-(x|z)_{h}-(y|u)_{h},

with both expressions being equal to

12(|xy||zu|+|xz|+|yu|).\frac{1}{2}(-|xy|-|zu|+|xz|+|yu|).

Similar calculations give equality for the other differences. Thus Ao(Q)A_{o}(Q) is a δ\delta-triple for a given δ0\delta\geq 0 if and only if Ah(Q)A_{h}(Q) is a δ\delta-triple.

Using Lemma 5.7 we conclude that the six numbers (x|y)h(x|y)_{h}, (z|u)h(z|u)_{h}, (x|z)h(x|z)_{h}, (y|u)h(y|u)_{h}, (x|u)h(x|u)_{h}, (y|z)h(y|z)_{h} together satisfy the hypotheses of Lemma 5.8 with parameter δ=δ(α,τ)\delta=\delta(\alpha,\tau). This implies that Ah(Q)A_{h}(Q) is a 2δ2\delta-triple and therefore that Ao(Q)A_{o}(Q) is a 2δ2\delta-triple. By [12, Proposition 2.4.1] this implies that inequality (2.3) holds for Gromov products based at oo in XX (with 2δ2\delta replacing 4δ4\delta). By [17, Chapitre 2, Proposition 21] this implies that geodesic triangles in XX are 8δ8\delta-thin, i.e., XX is 8δ8\delta-hyperbolic. ∎

We next show that any vertex in VV is part of a vertical geodesic line. We will in fact show something stronger. We let Z¯\bar{Z} denote the completion of ZZ, and continue to write dd for the canonical extension of the metric on ZZ to its completion. For r>0r>0 and a point zZ¯z\in\bar{Z} we will write B(z,r)B^{\prime}(z,r) for the open ball of radius rr centered at zz in the completion Z¯\bar{Z}.

Lemma 5.10.

Let zZ¯z\in\bar{Z}. Then there is a vertical geodesic γ:X\gamma:\mathbb{R}\rightarrow X with h(γ(t))=th(\gamma(t))=t for tt\in\mathbb{R} such that, writing γ(n)=(zn,n)\gamma(n)=(z_{n},n) for nn\in\mathbb{Z}, we have zB(zn,τ3αn)z\in B^{\prime}(z_{n},\frac{\tau}{3}\alpha^{-n}) for each nn\in\mathbb{Z}. Furthermore if v=(z,m)v=(z,m) is a given vertex of VV then we can construct γ\gamma such that γ(m)=v\gamma(m)=v.

Proof.

Since τ3>1\frac{\tau}{3}>1 by (5.1) and since for each nn\in\mathbb{Z} the balls B(y,αn)B(y,\alpha^{-n}) cover ZZ for ySny\in S_{n}, it follows from the fact that ZZ is dense in Z¯\bar{Z} that the balls B(y,τ3αn)B^{\prime}(y,\frac{\tau}{3}\alpha^{-n}) for ySny\in S_{n} cover Z¯\bar{Z}. Thus, given zZ¯z\in\bar{Z}, for each nn\in\mathbb{Z} we can find znSnz_{n}\in S_{n} such that zB(zn,τ3αn)z\in B^{\prime}(z_{n},\frac{\tau}{3}\alpha^{-n}).

Let vn=(zn,n)v_{n}=(z_{n},n) be the associated vertex in VV. We claim that for each nn\in\mathbb{Z} we have B(vn)B(vn+1)B(v_{n})\cap B(v_{n+1})\neq\emptyset. Since ZZ is dense in Z¯\bar{Z} we can find yZy\in Z such that d(y,z)<τ3αn1d(y,z)<\frac{\tau}{3}\alpha^{-n-1}. Then

d(y,zn+1)d(y,z)+d(z,zn+1)<τ3αn1+τ3αn1<ταn1,d(y,z_{n+1})\leq d(y,z)+d(z,z_{n+1})<\frac{\tau}{3}\alpha^{-n-1}+\frac{\tau}{3}\alpha^{-n-1}<\tau\alpha^{-n-1},

which implies that yB(vn+1)y\in B(v_{n+1}). A similar calculation shows that yB(vn)y\in B(v_{n}) since αn1<αn\alpha^{-n-1}<\alpha^{-n}. Thus B(vn)B(vn+1)B(v_{n})\cap B(v_{n+1})\neq\emptyset and therefore vnvn+1v_{n}\sim v_{n+1}. We can then find a vertical geodesic γ:X\gamma:\mathbb{R}\rightarrow X through the sequence of vertices {vn}n\{v_{n}\}_{n\in\mathbb{Z}}, which can be parametrized such that h(γ(t))=th(\gamma(t))=t for tt\in\mathbb{R}. Finally, if v=(z,m)v=(z,m) is a vertex of VV then we can choose zm=zz_{m}=z in our construction since we trivially have zB(z,τ3αm)z\in B^{\prime}(z,\frac{\tau}{3}\alpha^{-m}). ∎

A descending geodesic ray γ:[0,)X\gamma:[0,\infty)\rightarrow X is a vertical geodesic ray such that h(γ(t))h(\gamma(t)) is strictly decreasing as a function of tt. In this case we have h(γ(t))=h(γ(0))th(\gamma(t))=h(\gamma(0))-t for each t0t\geq 0. Similarly an ascending geodesic ray γ:[0,)X\gamma:[0,\infty)\rightarrow X is a vertical geodesic ray such that h(γ(t))h(\gamma(t)) is strictly increasing as a function of tt. In this case we instead have that h(γ(t))=h(γ(0))+th(\gamma(t))=h(\gamma(0))+t for each t0t\geq 0. A vertical geodesic γ\gamma is anchored at a point zZ¯z\in\bar{Z} if for each vertex (zm,m)(z_{m},m) belonging to γ\gamma we have zB(zm,τ3αm)z\in B^{\prime}(z_{m},\frac{\tau}{3}\alpha^{-m}); when the point zz does not need to be referenced we will just say that γ\gamma is anchored. Lemma 5.10 gives the existence of ascending and descending geodesic rays in XX anchored at any point zZ¯z\in\bar{Z}.

We will next show that all anchored descending vertical geodesic rays in XX define the same point in the Gromov boundary X\partial X.

Lemma 5.11.

Let γ\gamma, σ:[0,)X\sigma:[0,\infty)\rightarrow X be two descending geodesic rays in XX starting at vertices v=γ(0)v=\gamma(0) and w=σ(0)w=\sigma(0) of XX respectively and anchored at y,zZ¯y,z\in\bar{Z} respectively. Let kk\in\mathbb{Z} be such that kmin{h(v),h(w)}k\leq\min\{h(v),h(w)\} and τ3αk>d(y,z)\frac{\tau}{3}\alpha^{-k}>d(y,z). Let vkγVkv_{k}\in\gamma\cap V_{k}, wkσVkw_{k}\in\sigma\cap V_{k} be the vertices on these geodesics at the height kk. Then |vkwk|1|v_{k}w_{k}|\leq 1.

Proof.

By the anchoring condition we have d(vk,y)<τ3αkd(v_{k},y)<\frac{\tau}{3}\alpha^{-k} and d(wk,z)<τ3αkd(w_{k},z)<\frac{\tau}{3}\alpha^{-k}. Hence

d(vk,wk)d(vk,y)+d(y,z)+d(z,wk)<ταk.d(v_{k},w_{k})\leq d(v_{k},y)+d(y,z)+d(z,w_{k})<\tau\alpha^{-k}.

Thus wkB(vk)w_{k}\in B(v_{k}) and therefore B(vk)B(wk)B(v_{k})\cap B(w_{k})\neq\emptyset, which implies that either vk=wkv_{k}=w_{k} or vkwkv_{k}\sim w_{k}. In both cases we conclude that |vkwk|1|v_{k}w_{k}|\leq 1. ∎

The Busemann functions associated to anchored descending geodesic rays have a particularly simple form.

Lemma 5.12.

Let γ\gamma be an anchored descending geodesic ray in XX starting from a vertex vVv\in V. Then for all xXx\in X we have

(5.5) bγ(x)3h(x)h(γ(0)).b_{\gamma}(x)\doteq_{3}h(x)-h(\gamma(0)).
Proof.

Since both bγb_{\gamma} and hh are 11-Lipschitz and the edges of XX have unit length, it suffices to prove the estimate (5.5) on the vertices of XX with the constant 11 instead of 33. Let zZ¯z\in\bar{Z} be the anchoring point for γ\gamma. Let wVw\in V be an arbitrary vertex and let σ:[0,)X\sigma:[0,\infty)\rightarrow X be an anchored descending geodesic ray in XX starting at ww and anchored at the point yZy\in Z associated to ww, as constructed in Lemma 5.10. The sequences of vertices {γ(n)}n=0\{\gamma(n)\}_{n=0}^{\infty} on γ\gamma satisfies h(γ(n))=h(γ(0))nh(\gamma(n))=h(\gamma(0))-n since γ\gamma is a descending geodesic ray, and the same holds for σ\sigma. We let nn be any integer large enough that h(γ(n))h(w)h(\gamma(n))\leq h(w) and τ3αh(γ(n))>d(y,z)\frac{\tau}{3}\alpha^{-h(\gamma(n))}>d(y,z) and observe that if we define kn=h(w)h(γ(n))k_{n}=h(w)-h(\gamma(n)) then h(σ(kn))=h(γ(n))h(\sigma(k_{n}))=h(\gamma(n)). Then by Lemma 5.11 we have |γ(n)σ(kn)|1|\gamma(n)\sigma(k_{n})|\leq 1. Since σ(kn)\sigma(k_{n}) is joined to ww by a vertical geodesic of length knk_{n}, it follows immediately that

h(w)h(γ(n))|γ(n)w|h(w)h(γ(n))+1,h(w)-h(\gamma(n))\leq|\gamma(n)w|\leq h(w)-h(\gamma(n))+1,

and therefore, since h(γ(n))=h(γ(0))nh(\gamma(n))=h(\gamma(0))-n,

|γ(n)w|1h(w)+nh(γ(0)).|\gamma(n)w|\doteq_{1}h(w)+n-h(\gamma(0)).

This implies that

|γ(n)w|n1h(w)h(γ(0)).|\gamma(n)w|-n\doteq_{1}h(w)-h(\gamma(0)).

By letting nn\rightarrow\infty we conclude that

bγ(w)1h(w)h(γ(0)),b_{\gamma}(w)\doteq_{1}h(w)-h(\gamma(0)),

which gives the desired result. ∎

In particular, for an anchored descending geodesic ray γ\gamma with h(γ(0))=0h(\gamma(0))=0, Lemma 5.12 shows that bγ3hb_{\gamma}\doteq_{3}h. We fix such a descending geodesic ray γ\gamma for the remainder of this section and write b:=bγb:=b_{\gamma} for the associated Busemann function. Let ωX\omega\in\partial X be the point corresponding to the equivalence class of γ\gamma in the Gromov boundary of XX; note that Lemma 5.11 shows that all anchored descending geodesic rays belong to the equivalence class ω\omega defined by γ\gamma. Our final goal in this section is to show that the boundary ωX\partial_{\omega}X of XX relative to ω\omega can be canonically identified with the completion Z¯\bar{Z} of ZZ in such a way that the extension of the metric dd to Z¯\bar{Z} is a visual metric on ωX\partial_{\omega}X based at ω\omega with parameter logα\log\alpha.

We remark that the rough equality b3hb\doteq_{3}h implies that (x|y)b3(x|y)h(x|y)_{b}\doteq_{3}(x|y)_{h} for all x,yXx,y\in X as well, so that in particular the conclusions of Lemma 5.5 hold with bb replacing hh and (v|w)b(v|w)_{b} replacing (v|w)h(v|w)_{h} everywhere, at the cost of adding 66 to the constant c(α,τ)c(\alpha,\tau) there and multiplying the constant C(α,τ)C(\alpha,\tau) by α6\alpha^{6}. We will use this observation without further comment below.

For each point zZ¯z\in\bar{Z} we fix an ascending geodesic ray γz:[0,)X\gamma_{z}:[0,\infty)\rightarrow X anchored at zz, as given by Lemma 5.10. We define a map ψ:Z¯ωX\psi:\bar{Z}\rightarrow\partial_{\omega}X by setting ψ(z)=[γz]\psi(z)=[\gamma_{z}], i.e., ψ(z)\psi(z) is the equivalence class in ωX\partial_{\omega}X defined by the geodesic ray γz\gamma_{z}. Implicit in this definition is the fact that we cannot have [γz]=ω[\gamma_{z}]=\omega for any zZ¯z\in\bar{Z}, as it is easy to see from the fact that hh is 11-Lipschitz that ascending geodesic rays cannot be at bounded distance from descending geodesic rays. We also note that if γ\gamma is any other ascending geodesic ray anchored at zZ¯z\in\bar{Z} then we must have [γ]=[γz][\gamma]=[\gamma_{z}]: for kk\in\mathbb{Z} sufficiently large the unique vertices vkγVkv_{k}\in\gamma\cap V_{k} and wkγzVkw_{k}\in\gamma_{z}\cap V_{k} must satisfy |vkwk|1|v_{k}w_{k}|\leq 1 since zB(vk)B(wk)z\in B(v_{k})\cap B(w_{k}), hence the geodesic rays γ\gamma and γz\gamma_{z} are at a bounded distance from one another. The map ψ\psi can thus equivalently be thought of as sending zZ¯z\in\bar{Z} to the equivalence class of all ascending geodesic rays anchored at zz.

Proposition 5.13.

The map ψ:Z¯ωX\psi:\bar{Z}\rightarrow\partial_{\omega}X defines an identification of Z¯\bar{Z} with ωX\partial_{\omega}X. Under this identification the metric dd on Z¯\bar{Z} defines a visual metric on ωX\partial_{\omega}X with parameter logα\log\alpha and comparison constant depending only on α\alpha and τ\tau.

Proof.

Let x,yZx,y\in Z be given and let γx\gamma_{x} and γy\gamma_{y} be ascending geodesic rays anchored at xx and yy respectively. For nn\in\mathbb{Z} sufficiently large we let xnγxVnx_{n}\in\gamma_{x}\cap V_{n} and ynγyVny_{n}\in\gamma_{y}\cap V_{n} be the unique vertices on these rays at height nn. By Lemma 5.6 we have

(5.6) α(xn|yn)bC(α,τ)d(xn,yn)+αn.\alpha^{-(x_{n}|y_{n})_{b}}\asymp_{C(\alpha,\tau)}d(x_{n},y_{n})+\alpha^{-n}.

Since xB(xn)x\in B(x_{n}) and yB(yn)y\in B(y_{n}) we have d(x,xn)<ταnd(x,x_{n})<\tau\alpha^{-n} and d(y,yn)<ταnd(y,y_{n})<\tau\alpha^{-n}. Hence, by letting nn\rightarrow\infty in (5.6) and using Lemma 2.7, we conclude that

(5.7) α(ψ(x)|ψ(y))bC(α,τ)d(x,y).\alpha^{-(\psi(x)|\psi(y))_{b}}\asymp_{C(\alpha,\tau)}d(x,y).

It follows immediately that ψ:Z¯ωX\psi:\bar{Z}\rightarrow\partial_{\omega}X is injective. To complete the proof of the proposition it suffices to show that ψ\psi is surjective, as the estimate (5.7) then shows that the metric dd on Z¯\bar{Z} defines a visual metric on ωX\partial_{\omega}X with parameter logα\log\alpha and comparison constant depending only on α\alpha and τ\tau when we use ψ\psi to identify Z¯\bar{Z} with ωX\partial_{\omega}X.

We recall from Proposition 2.8 that ωX\partial_{\omega}X can be defined as equivalence classes of sequences {xn}\{x_{n}\} in XX such that (xm|xn)b(x_{m}|x_{n})_{b}\rightarrow\infty as m,nm,n\rightarrow\infty, with two sequences {xn}\{x_{n}\}, {yn}\{y_{n}\} being equivalent if (xn|yn)b(x_{n}|y_{n})_{b}\rightarrow\infty as nn\rightarrow\infty. Since bb is 11-Lipschitz we can always choose these sequences to consist of vertices in XX by replacing xnx_{n} with a nearest vertex vnv_{n}.

Thus let {vn}\{v_{n}\} be a sequence of vertices defining a point ξ\xi of ωX\partial_{\omega}X. Let {zn}\{z_{n}\} be the associated sequence of points in ZZ. By Lemma 5.6 we have

α(vn|vm)bC(α,τ)d(zn,zm)+αmin{b(vn),b(vm)}.\alpha^{-(v_{n}|v_{m})_{b}}\asymp_{C(\alpha,\tau)}d(z_{n},z_{m})+\alpha^{-\min\{b(v_{n}),b(v_{m})\}}.

Since (vn|vm)b(v_{n}|v_{m})_{b}\rightarrow\infty and b(vn)b(v_{n})\rightarrow\infty, it follows immediately that {zn}\{z_{n}\} is a Cauchy sequence in ZZ and therefore converges to a point zZ¯z\in\bar{Z}. We claim that ψ(z)=ξ\psi(z)=\xi.

Let γz\gamma_{z} be an ascending geodesic ray anchored at zz and let {wn}\{w_{n}\} be the sequence of vertices on γz\gamma_{z} starting from its initial point. When considered as points of ZZ this sequence of vertices must satisfy d(wn,z)0d(w_{n},z)\rightarrow 0 since γz\gamma_{z} is ascending and anchored at zz. This implies that d(wn,zn)0d(w_{n},z_{n})\rightarrow 0. Since by Lemma 5.6 we have

α(vn|wn)bC(α,τ)d(zn,wn)+αmin{b(vn),b(wn)},\alpha^{-(v_{n}|w_{n})_{b}}\asymp_{C(\alpha,\tau)}d(z_{n},w_{n})+\alpha^{-\min\{b(v_{n}),b(w_{n})\}},

we conclude that (vn|wn)b(v_{n}|w_{n})_{b}\rightarrow\infty as nn\rightarrow\infty. Hence {vn}\{v_{n}\} and {wn}\{w_{n}\} define the same point of ωX\partial_{\omega}X, i.e., ψ(z)=ξ\psi(z)=\xi. We conclude that ψ\psi 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 (Z,d)(Z,d) be a metric space and let XX be a hyperbolic filling of ZZ with parameters α>1\alpha>1 and τ>min{3,α/(α1)}\tau>\min\{3,\alpha/(\alpha-1)\} as in the previous section. We let h:Xh:X\rightarrow\mathbb{R} be the height function and set ρ(x)=αh(x)\rho(x)=\alpha^{-h(x)}. We write XρX_{\rho} for the conformal deformation of XX with conformal factor ρ\rho, dρd_{\rho} for the metric on XρX_{\rho}, and ρ\ell_{\rho} for lengths of curves measured in the metric dρd_{\rho}. Since hh is 11-Lipschitz the density ρ\rho satisfies the Harnack inequality (4.2) with ε=logα\varepsilon=\log\alpha.

In the notation of Remark 1.7, Lemma 5.12 shows that we have h3(X)h\in\mathcal{B}_{3}(X). Clearly XX is geodesic and complete, and Proposition 5.9 shows that XX is δ\delta-hyperbolic with δ=δ(α,τ)\delta=\delta(\alpha,\tau). Thus to prove Theorem 1.12 it suffices to show that XX is 12\frac{1}{2}-roughly starlike from ω\omega and that ρ\rho is a GH-density with constant M=M(α,τ)M=M(\alpha,\tau), as it then follows for a Busemann function b(X)b\in\mathcal{B}(X) with b3hb\doteq_{3}h that ρε,b\rho_{\varepsilon,b} is a GH-density with constant M=M(α,τ)M=M(\alpha,\tau) as well, where ε=logα\varepsilon=\log\alpha. The conclusions of Theorem 1.12 can then be derived from the fact that the metrics on Xε,bX_{\varepsilon,b} and XρX_{\rho} are α3\alpha^{3}-biLipschitz to one another by the identity map on XX.

We first look at rough starlikeness from the distinguished point ωX\omega\in\partial X corresponding to the equivalence class of all anchored descending geodesic rays in XX.

Lemma 6.1.

The hyperbolic filling XX is 12\frac{1}{2}-roughly starlike from ω\omega.

Proof.

Let vVv\in V be a vertex of XX with associated point zZz\in Z. Let γ:X\gamma:\mathbb{R}\rightarrow X be an ascending vertical geodesic line through vv that is anchored at zz, as constructed in Lemma 5.10, parametrized by arclength such that γ(0)=v\gamma(0)=v. We put γ¯(t)=γ(t)\bar{\gamma}(t)=\gamma(-t) for tt\in\mathbb{R}. Then γ¯|[0,)\bar{\gamma}|_{[0,\infty)} is an anchored descending geodesic ray and therefore belongs to the equivalence class ω\omega by Lemma 5.11. This shows that any vertex of XX lies on a geodesic line starting at ω\omega. Since any point in XX is within distance 12\frac{1}{2} of some vertex, condition (1) of Definition 2.3 follows.

For condition (2) we use the identification of ωX\partial_{\omega}X with Z¯\bar{Z} from Proposition 5.13. Let zZ¯z\in\bar{Z} be given and let γ:X\gamma:\mathbb{R}\rightarrow X be a vertical geodesic line anchored at zz and parametrized such that h(γ(t))=th(\gamma(t))=t, as constructed in Lemma 5.10. By the construction of the identification φ\varphi in Proposition 5.13 the geodesic ray γ|[0,)\gamma|_{[0,\infty)} then belongs to the equivalence class of zz when zz is considered as a point of ωX\partial_{\omega}X. Putting γ¯(t)=γ(t)\bar{\gamma}(t)=\gamma(-t) as above, we also have that γ¯|[0,)\bar{\gamma}|_{[0,\infty)} is a descending geodesic ray anchored at zz and therefore belongs to the equivalence class of ω\omega by Lemma 5.11. Since zZ¯z\in\bar{Z} was arbitrary, condition (2) follows. ∎

We will now show that ρ\rho is a GH-density with constant M=M(α,τ)M=M(\alpha,\tau) depending only on α\alpha and τ\tau. We will do this by estimating the distance dρd_{\rho} between points at sufficiently large scales in XX and then using Corollary 4.8.

Lemma 6.2.

Let x,yXx,y\in X with |xy|1|xy|\geq 1. Then we have

(6.1) dρ(x,y)C(α,τ)α(x|y)h.d_{\rho}(x,y)\asymp_{C(\alpha,\tau)}\alpha^{-(x|y)_{h}}.

Consequently ρ\rho is a GH-density with constant M=M(α,τ)M=M(\alpha,\tau).

Proof.

The bound on dρ(x,y)d_{\rho}(x,y) from above follows from Lemma 4.5 applied with ε=logα\varepsilon=\log\alpha, since there is a Busemann function b(X)b\in\mathcal{B}(X) such that b3hb\doteq_{3}h. Hence it suffices to establish the lower bound. Observe that for an edge gg of XX, considered as a path between its endpoints vv and ww and assuming the orientation in which h(v)h(w)h(v)\leq h(w), when h(v)=h(w)=kh(v)=h(w)=k we have

(6.2) ρ(g)=αk.\ell_{\rho}(g)=\alpha^{-k}.

On the other hand, since B(v)B(w)B(v)\cap B(w)\neq\emptyset we have d(v,w)<2ταkd(v,w)<2\tau\alpha^{-k}. It follows that

ρ(g)>C(α,τ)1d(v,w)\ell_{\rho}(g)>C(\alpha,\tau)^{-1}d(v,w)

Similarly, when h(v)=kh(v)=k and h(w)=k+1h(w)=k+1 we have

(6.3) ρ(g)=1logα(αkαk1)=1α1logααk,\ell_{\rho}(g)=\frac{1}{\log\alpha}(\alpha^{-k}-\alpha^{-k-1})=\frac{1-\alpha^{-1}}{\log\alpha}\alpha^{-k},

while B(v)B(w)B(v)\cap B(w)\neq\emptyset implies again that d(v,w)<2ταkd(v,w)<2\tau\alpha^{-k}. Thus in this case we also have

ρ(g)>C(α,τ)1d(v,w).\ell_{\rho}(g)>C(\alpha,\tau)^{-1}d(v,w).

Now let γ\gamma be a rectifiable curve joining xx to yy. Let vv be the first vertex on γ\gamma met traveling from xx to yy and let ww be the first vertex on γ\gamma met traveling from yy to xx. We first suppose that vwv\neq w. We then let σ\sigma be the subcurve of γ\gamma from vv to ww starting from this first occurrence of vv and ending at this last occurrence of ww. Let {vi}i=0l\{v_{i}\}_{i=0}^{l} be the sequence of vertices encountered along the path σ\sigma, noting that by assumption we have l1l\geq 1. Then from our calculations above we have

(6.4) ρ(σ)C(α,τ)1i=0l1d(vi,vi+1)C(α,τ)1d(v,w).\ell_{\rho}(\sigma)\geq C(\alpha,\tau)^{-1}\sum_{i=0}^{l-1}d(v_{i},v_{i+1})\geq C(\alpha,\tau)^{-1}d(v,w).

On the other hand, since vwv\neq w the curve σ\sigma must contain at least one full edge of XX with one vertex being vv and at least one full edge with one vertex being ww (these may be the same edge). Then it follows from (6.2) and (6.3) applied to those edges that

(6.5) ρ(σ)C(α,τ)1αmin{h(v),h(w)}.\ell_{\rho}(\sigma)\geq C(\alpha,\tau)^{-1}\alpha^{-\min\{h(v),h(w)\}}.

By combining (6.4) and (6.5) and then using Lemma 5.6, we conclude that

ρ(σ)C(α,τ)1α(v|w)h.\ell_{\rho}(\sigma)\geq C(\alpha,\tau)^{-1}\alpha^{-(v|w)_{h}}.

Since σ\sigma is a subcurve of γ\gamma it follows that this inequality holds for γ\gamma as well.

Now suppose that v=wv=w. Then, since |xy|1|xy|\geq 1, either the initial segment of γ\gamma from xx to vv or the final segment of γ\gamma from ww to yy has length at least 12\frac{1}{2} in XX. By reversing the roles of xx and yy if necessary we can assume that the initial segment η\eta of γ\gamma from xx to vv has length at least 12\frac{1}{2} in XX. Since η\eta is contained entirely in a single edge of XX that has vv as a vertex, it follows that

ρ(η)αh(v)1(η)12αh(v)1,\ell_{\rho}(\eta)\geq\alpha^{-h(v)-1}\ell(\eta)\geq\frac{1}{2}\alpha^{-h(v)-1},

with (η)\ell(\eta) denoting the length of η\eta in XX. Hence

ρ(γ)ρ(η)C(α)1α(v|w)h,\ell_{\rho}(\gamma)\geq\ell_{\rho}(\eta)\geq C(\alpha)^{-1}\alpha^{-(v|w)_{h}},

using that (v|w)h=h(v)(v|w)_{h}=h(v) since v=wv=w. This gives a similar lower bound on ρ(γ)\ell_{\rho}(\gamma) in this case as well. Minimizing over all rectifiable paths γ\gamma from xx to yy then gives in both cases that

dρ(x,y)C(α,τ)1α(v|w)hC(α,τ)1α(x|y)h,d_{\rho}(x,y)\geq C(\alpha,\tau)^{-1}\alpha^{-(v|w)_{h}}\geq C(\alpha,\tau)^{-1}\alpha^{-(x|y)_{h}},

with the second inequality following from the fact that hh is 11-Lipschitz and |xv|1|xv|\leq 1, |yw|1|yw|\leq 1.

To conclude that ρ\rho is a GH-density we let bb be a Busemann function on XX such that b3hb\doteq_{3}h as in Lemma 5.12. We let dε,bd_{\varepsilon,b} be the distance obtained on XX through conformal deformation with the conformal factor

ρε,b(x)=eεb(x)=αb(x),\rho_{\varepsilon,b}(x)=e^{-\varepsilon b(x)}=\alpha^{-b(x)},

where ε=logα\varepsilon=\log\alpha. Then dε,bd_{\varepsilon,b} is α3\alpha^{3}-biLipschitz to dρd_{\rho} by the identity map on XX. Consequently the comparison 6.1 holds with dε,bd_{\varepsilon,b} replacing dρd_{\rho}. We can then apply Corollary 4.8 to conclude that ρε,b\rho_{\varepsilon,b} is a GH-density with constant M=M(α,τ)M=M(\alpha,\tau), noting that XX is δ\delta-hyperbolic with δ=δ(α,τ)\delta=\delta(\alpha,\tau) and 12\frac{1}{2}-roughly starlike from ω\omega. The GH-inequality (1.3) for ρ\rho then follows immediately from the α3\alpha^{3}-biLipschitz comparison of dε,bd_{\varepsilon,b} to dρd_{\rho}. ∎

This completes the proof of Theorem 1.12 aside from the final assertion regarding the identification of Xρ\partial X_{\rho} with Z¯\bar{Z} given by the combination of Lemma 4.14 and Proposition 5.13 is biLipschitz. This is shown below.

Proposition 6.3.

The identification XρZ¯\partial X_{\rho}\cong\bar{Z} is biLipschitz with biLipschitz constant L=L(α,τ)L=L(\alpha,\tau) depending only on α\alpha and τ\tau.

Proof.

We consider ωX\partial_{\omega}X as equipped with the visual metric with parameter logα\log\alpha defined by Proposition 5.13, which coincides with the extension of the metric dd on ZZ to the completion Z¯\bar{Z} under the identification ψ:Z¯ωX\psi:\bar{Z}\rightarrow\partial_{\omega}X of that proposition. By Theorem 1.6 applied with this visual metric and ε=logα\varepsilon=\log\alpha, the identification φ:ωXXρ\varphi:\partial_{\omega}X\rightarrow\partial X_{\rho} is biLipschitz. Hence the induced identification Z¯Xε\bar{Z}\cong\partial X_{\varepsilon} given by φψ\varphi\circ\psi is also biLipschitz. Furthermore all of the parameters involved in the biLipschitz constant can be taken to depend only on α\alpha and τ\tau by the results of this section. ∎

Remark 6.4.

For kk\in\mathbb{Z} there is a canonical correspondence between hyperbolic fillings with fixed parameters α,τ>1\alpha,\tau>1 of the metric spaces (Z,d)(Z,d) and (Z,αkd)(Z,\alpha^{-k}d) given by considering αn\alpha^{-n}-separated sets in (Z,d)(Z,d) as αnk\alpha^{-n-k}-separated sets in (Z,αkd)(Z,\alpha^{-k}d). Thus when ZZ is bounded there is no harm in assuming that diamZ<1\mathrm{diam}\,Z<1 by multiplying the metric by αk\alpha^{-k} for kk sufficiently large. The hyperbolic filling XX can then be written as X=X0X0X=X_{\geq 0}\cup X_{\leq 0}, where X0=h1([0,))X_{\geq 0}=h^{-1}([0,\infty)) is the set of all points of nonnegative height and X0=h1((,0])X_{\leq 0}=h^{-1}((-\infty,0]) is the set of all points of nonpositive height. The condition diamZ<1\mathrm{diam}\,Z<1 implies that the vertex sets Vn={vn}V_{n}=\{v_{n}\} for n0n\leq 0 consist only of a single point, and in particular X0X_{\leq 0} is simply a descending geodesic ray starting from v0v_{0}. The space XX is isometric to the ray augmentation of X0X_{\geq 0} based at v0v_{0}, in the language of Definition 4.18.

The graph X0X_{\geq 0} is essentially the hyperbolic filling of ZZ 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 SmSnS_{m}\subset S_{n} for mnm\leq n. They uniformize this filling for all τ>1\tau>1 using the density ρε,v0(x)=eε|xv0|\rho_{\varepsilon,v_{0}}(x)=e^{-\varepsilon|xv_{0}|} for 0<εlogα0<\varepsilon\leq\log\alpha, for which it is easy to see that ρε,v0=ρε|X0\rho_{\varepsilon,v_{0}}=\rho_{\varepsilon}|_{X\geq 0}, where ρε(x)=eεh(x)\rho_{\varepsilon}(x)=e^{-\varepsilon h(x)} and h:Xh:X\rightarrow\mathbb{R} is the height function. When τ\tau 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 τ\tau is close to 11 it is possible to realize trees as hyperbolic fillings [3, Theorem 7.1], whereas when τ\tau satisfies (5.1) the hyperbolic filling is only a tree if ZZ 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 pp-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 CAT(1){\rm CAT}(-1)-espace. Enseign. Math. (2), 41(1-2):63–102, 1995.
  • [9] Marc Bourdon and Hervé Pajot. Cohomologie lpl_{p} 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, CAT(1){\rm CAT}(-1)-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.