1 Introduction

Throughout this section X is a simply and rectifiably connected plane set; we make no assumption about X being open or closed. Then \(\overline{X_l}\) is the metric completion of \(X_l:=(X,l)\), the metric space where l is the intrinsic (Euclidean) length distance on X.

Theorem

Suppose X is a simply and rectifiably connected plane set. Then \(\overline{X_l}\) is a Hadamard space. Moreover, \(\overline{X_l}\) is Gromov hyperbolic if and only if X does not contain Euclidean disks of arbitrarily large radii, i.e., if and only if

$$\begin{aligned} R:=\sup \left\{ r>0 \;\big |\; \exists \, \textsf{D}(x;r)\subset X\right\} <+\infty ; \end{aligned}$$

when this holds, X is 2R-hyperbolic and this is best possible.

Recall that a Hadamard metric space is a complete CAT(0) spaceFootnote 1; see Sect. 2.2.2. Our result provides a bountiful supply of easily constructed Hadamard spaces.

The special case where X is a closed bounded Jordan plane region was established in [1]; there the work of [3] is an essential ingredient. We present an elementary argument for the case where X is a simply connected plane domain; see Sect. 3.1.Footnote 2

It is easy to construct compact rectifiably connected plane sets X for which \(X_l\) is geodesic but fails to have non-positive curvature. It would be useful to have a characterization of the plane sets X for which \(\overline{X_l}\) has non-positive curvature. As Bishop mentions, in [5, Ex. 9.1.6, p. 310] the authors assert that for “any locally simply connected plane set X”, \(X_l\) has non-positive curvature. However, as Bishop comments, their discussion fails to mention certain essential details including, in particular, the existence of geodesics.

2 Preliminaries

For real numbers r and s,

$$\begin{aligned} r\wedge s:= \min \{r,s\} \quad \text {and}\quad r\vee s:= \max \{r,s\}. \end{aligned}$$

2.1 Metric Space Notation and Terminology

Throughout this section X is an arbitrary metric space with distance denoted \(|x-y|\); this is not meant to imply that X possesses any sort of linear or group structure. In this setting, all topological notions refer to the metric topology; here \({{\,\mathrm{\textsf{cl}}\,}}(A), {{\,\mathrm{\textsf{bd}}\,}}(A), {{\,\mathrm{\textsf{int}}\,}}(A)\) are the topological closure, boundary, interior (respectively) of \(A\subset X\).

Every metric space can be isometrically embedded into a complete metric space. We let \(\bar{X}\) denote the metric completion of the metric space X; thus \(\bar{X}\) is the closure of the image of X under such an isometric embedding. We call \(\partial X:=\bar{X}{\setminus }X\) the metric boundary of X.

When \(A\subset X\), there is a natural embedding \({{\bar{A}}}\hookrightarrow \bar{X}\) and \({{\,\mathrm{\textsf{bd}}\,}}(A)\subset {\partial A}\). Here if \(A\subset X\) is open and X complete, then \({\partial A}={{\,\mathrm{\textsf{bd}}\,}}(A)\), but in general \({{\bar{A}}}={\overline{{{\,\mathrm{\textsf{cl}}\,}}}}(A)\) and \({\partial A}={\overline{{{\,\mathrm{\textsf{bd}}\,}}}}(A){\setminus }A\) where \({\overline{{{\,\mathrm{\textsf{cl}}\,}}}}\) and \({\overline{{{\,\mathrm{\textsf{bd}}\,}}}}\) denote topological closure and boundary in \(\bar{X}\)

2.1.1 Paths, Arcs, Geodesics, and Length

A path in X is a continuous map \({\mathbb R}\supset I\xrightarrow {\gamma }X\) where \(I=I_\gamma \) is the parameter interval for \(\gamma \) and may be closed or open or neither and finite or infinite. The trajectory of such a path \(\gamma \) is \(|\gamma |:=\gamma (I)\) which we call a curve and often–when easily understood in context–we abuse notation and just write \(\gamma \) in place of \(|\gamma |\).

A path \(I\xrightarrow {\gamma }X\) is a geodesic if it is an isometry:

$$\begin{aligned} \forall \; s,t\in I, \quad |\gamma (s)-\gamma (t)|=|s-t| \end{aligned}$$

and X is a geodesic metric space if each pair of points can be joined by a geodesic.

When I is closed and \(I\ne {\mathbb R}\), \(\partial \gamma :=\gamma (\partial I)\) denotes the set of endpoints of \(\gamma \) and consists of one or two points depending on whether or not I is compact. For example, if \(I_\gamma =[u,v]\subset {\mathbb R}\), then \(\partial \gamma =\{\gamma (u),\gamma (v)\}\). When \(\partial \gamma =\{a,b\}\), we write \(\gamma :a\curvearrowright b\) (in X) to indicate that \(\gamma \) is a path (in X) with initial point a and terminal point b; this implies an orientation—a precedes b on \(\gamma \).

We call \(\gamma \) a compact path if its parameter interval is compact. A compact path \(\gamma \) is a loop if \(\partial \gamma \) is a single point, and then \(|\gamma |\) is often dubbed a closed curve. A loop \(\gamma :[u,v]\rightarrow X\) is a Jordan loop (aka, a simple closed curve) if \(\gamma \vert _{[u,v)}\) is injective.

An arc \(\alpha \) is an injective compact path; here \(|\alpha |\) is often called a simple curve; again, we sometimes abuse notation and call \(|\alpha |\) an arc. The interior of \(\alpha \) is .

Given points \(a,b\in |\alpha |\), there is a unique subarc \(\alpha [a,b]\) of \(\alpha \) with endpoints ab; precisely, there are unique \(u,v\in I\) with \(\alpha (u)=a\), \(\alpha (v)=b\) and \(\alpha [a,b]:=\alpha \vert _{[u,v]}\). (Again, sometimes \(\alpha [a,b]\) is this map and sometimes it denotes its trajectory.) We also use this notation for a general path \(\gamma \), but here \(\gamma [a,b]\) denotes the unique subpath of \(\gamma \) that joins ab obtained by using the last time \(\gamma \) is at a up to the first time \(\gamma \) is at b.

When \(\alpha :a\curvearrowright b\) and \(\beta :b\curvearrowright c\) are paths that join a to b and b to c respectively, \(\alpha \star \beta \) denotes the concatenationFootnote 3 of \(\alpha \) and \(\beta \); so \(\alpha \star \beta :a \curvearrowright c\). The reverse of \(\gamma \) is the path \(\tilde{\gamma }\) defined by \(\tilde{\gamma }(t):=\gamma (1-t)\) (when \(I_\gamma =[0,1]\)) and going from \(\gamma (1)\) to \(\gamma (0)\). Of course, \(|\alpha \star \beta |=|\alpha |\cup |\beta |\) and \(|\tilde{\gamma }|=|\gamma |\).

Every compact path contains an arc with the same endpoints; see [12].

The length of a compact path \([0,1]\overset{\gamma }{\rightarrow }X\) is defined in the usual way by

$$\begin{aligned} \ell (\gamma ):=\sup \left\{ \sum _{i=1}^n |\gamma (t_i)-\gamma (t_{i-1})| \,\Big \vert \, 0=t_0<t_1<\cdots <t_n=1 \right\} , \end{aligned}$$

\(\gamma \) is rectifiable when \(\ell (\gamma )<\infty \), and X is rectifiably connected provided each pair of points in X can be joined by a rectifiable path. An arbitrary path \(\gamma \) is locally rectifiable if each compact subpath of \(\gamma \) is rectifiable, and such a \(\gamma \) is rectifiable if

$$\begin{aligned} \ell (\gamma ):=\sup \bigl \{\ell (\alpha ) \,\big \vert \, \alpha \text { a compact subpath of }\gamma \big \}<+\infty . \end{aligned}$$

Rectifiable paths always have endpoints, and so have unique extensions to compact paths with the same length. Here is a precise statement; cf. [11, Thm. 3.2, p.7].

Fact 2.1

Let \({\mathbb R}\supset I\xrightarrow {\gamma }X\) be a rectifiable path with I a finite interval. Then there is a unique extension \(\bar{I}\xrightarrow {{\bar{\gamma }}}\bar{X}\) of \(\gamma \) to a compact rectifiable path \({\bar{\gamma }}\) and \(\ell ({\bar{\gamma }})=\ell (\gamma )\).

Every rectifiable path can be parametrized with respect to its arclength [11, p. 5]. When \(\gamma \) is a rectifiable path, we tacitly assume its parameter interval is \(I_\gamma =[0,\ell (\gamma )]\) unless specifically stated otherwise.

2.1.2 Intrinsic Length Distance

Every rectifiably connected metric space X admits a natural intrinsic distance, its so-called (inner) length distance given by

$$\begin{aligned} l(a,b):= \inf \bigl \{\ell (\gamma )\,\big \vert \,\hbox {} \gamma :a\curvearrowright b \hbox { a rectifiable path in}\, X\bigr \} . \end{aligned}$$

A metric space \((X,{{\,\mathrm{|\cdot |}\,}})\) is a length space provided for all points \(a,b\in X\), \(|a-b|=l(a,b)\), and we call such a \({{\,\mathrm{|\cdot |}\,}}\) a length (or intrinsic) distance function. An l-geodesic \([a,b]_l\) is a shortest path joining a and b, and any shortest path can be parametrized to be an l-geodesic.

The notation \(X_l:=(X,l)\) is convenient, and then \(\partial _l X:=\overline{X_l}{\setminus }X_l\). We note that \((\overline{X_l})_l=\overline{X_l}\), which is a consequence of the facts that the length distance \(l=l_d\) associated with a length distance d is just d, and the completion of a length distance is also a length distance.

More generally, a continuous function \(X\xrightarrow {\rho }(0,\infty )\) on a rectifiably connected metric space X induces a length distance \(d_\rho \) on X defined by

$$\begin{aligned} d_\rho (a,b):=\inf _{\gamma :a\curvearrowright b} \ell _\rho (\gamma ) \quad \text {where}\quad \ell _\rho (\gamma ):= \int _\gamma \rho \, ds \end{aligned}$$

and where the infimum is taken over all rectifiable paths \(\gamma :a\curvearrowright b\) in X. We describe this by calling \(\rho \,ds=\rho (x)|dx|\) a conformal metric on X.

There are two useful properties of length spaces that we use repeatedly. First, for any open set U in a length space X, we always have \({{\,\mathrm{\textsf{dist}}\,}}(x,{{\,\mathrm{\textsf{bd}}\,}}U)={{\,\mathrm{\textsf{dist}}\,}}(x,X{\setminus }U)\) for all points \(x\in U\). Second, \(\bar{X}\) is also a length space. In fact, for all \(x\in X, \xi \in \partial X, \varepsilon >0\) there is a path \(\gamma :x\curvearrowright \xi \) in \(X\cup \{\xi \}\) with \(\ell (\gamma )<|x-\xi |+\varepsilon \).

We utilize the fact that rectifiable arcs in \(\overline{X_l}\) can be approximated by arcs in X. Here is a precise statement.

Lemma 2.2

Let X be rectifiably connected. Suppose \(\tilde{\gamma }:{\tilde{p}}\curvearrowright {\tilde{q}}\) is a rectifiable arc in \(\overline{X_l}\). Then for each \(\varepsilon >0\), there is a rectifiable path \(\gamma :p\curvearrowright q\) in X with \(l\bigl (\gamma (t),\tilde{\gamma }(t)\bigr )<\varepsilon \) for all \(t\in I:=[0,\ell (\tilde{\gamma })]\) (where I is also the parameter interval for \(\gamma \)); thus there are rectifiable arcs \(\alpha :p\curvearrowright q\) in X and \(\tilde{\alpha }:{\tilde{p}}\curvearrowright {\tilde{q}}\) in \(X\cup \left\{ {\tilde{p}},{\tilde{q}}\right\} \) with \(\alpha ,\tilde{\alpha }\subset \textsf{N}_l(\tilde{\gamma };\varepsilon )\).

Proof Sketch

Given \(\varepsilon \in (0,\ell (\tilde{\gamma }))\), let n be the smallest positive integer with \(\ell (\tilde{\gamma })/n\le \varepsilon /10\). Put \(t_i:=(i/n)\ell (\tilde{\gamma })\) for \(0\le i\le n\). Define \(x_i:=\tilde{\gamma }(t_i)\) if \(\tilde{\gamma }(t_i)\in X\); otherwise, if \(\tilde{\gamma }(t_i)\in \partial _l X\), choose any \(x_i\in X\) with \(l\bigl (x_i,\tilde{\gamma }(t_i)\bigr )<\varepsilon /10\). Then \(l(x_{i-1},x_i)<3\varepsilon /10\) so there are rectifiable arcs \(\gamma _i:x_{i-1}\curvearrowright x_i\) in X with \(\ell (\gamma _i)<3\varepsilon /10\). Then \(\gamma :=\gamma _1\star \cdots \star \gamma _n\) has the asserted properties, where \(\gamma _i:[t_{i-1},t_i]\rightarrow X\) is parametrized proportional to arc length. \(\square \)

Here is information that we employ to construct Jordan loops inside X.

Lemma 2.3

Let X be rectifiably connected. Suppose \(\gamma _i:{\tilde{p}}\curvearrowright q_i\) (\(i=1,2\)) are rectifiable arcs in \(X\cup \left\{ {\tilde{p}}\right\} \) with \(\gamma _1\cap \gamma _2=\left\{ {\tilde{p}}\right\} \subset \overline{X_l}\). Then for each \(\varepsilon >0\), there are points \(p_i\in \gamma _i\) and a rectifiable arc \(\alpha :p_1\curvearrowright p_2\) in X with \(l(p_1,{\tilde{p}})<\varepsilon ,l(p_2,{\tilde{p}})<\varepsilon ,\ell (\alpha )<\varepsilon \) and such that \(\gamma _1^{-1}[q_1,p_1]\star \alpha \star \gamma _2[p_2,q_2]\) is a rectifiable arc \(q_1\curvearrowright q_2\) in X.

Proof

Let \(\varepsilon >0\) be given. Choose points \(a_i\in \gamma _i\) and a rectifiable arc \(\beta :a_1\curvearrowright a_2\) in X with each of \(l(a_1,{\tilde{p}}),l(a_2,{\tilde{p}}),\ell (\beta )\) less than \(\varepsilon /10\). Let \(p_1\) be the last point of \(\beta \) in \(\gamma _1\) and let \(p_2\) be the first point of \(\beta [p_1,a_2]\) in \(\gamma _2\). Then \(\alpha :=\beta [p_1,p_2]\) has the asserted properties. \(\square \)

Let \([0,1)\xrightarrow {\gamma }X\) be a path in X. If there is a point \(\xi \in \partial X\) such that \(\lim _{t\rightarrow 1^-}|\gamma (t)-\xi |=0\), then \(\xi \) is called a path accessible (metric) boundary point of X. In this situation, we define \(\gamma (1):=\xi \) and obtain a path \(\gamma :[0,1]\rightarrow X\cup \{\xi \}\subset \bar{X}\). We describe this by saying that \(\gamma \) is a path in X with terminal endpoint \(\xi \in \partial X\).

We write \(\partial ^\textrm{pa}\!X\) for the set of all path accessible boundary points of X. Restricting attention to rectifiable paths \(\gamma \) yields rectifiably accessible (metric) boundary points of X, denoted by \({\partial ^\textrm{ra}} X\). Clearly, \({\partial ^\textrm{ra}} X\subset \partial ^\textrm{pa}\!X\subset \partial X\) and each containment may be strict. We define \(X^\mathrm{{ra}}:=X\cup {\partial ^\textrm{ra}} X\).

A path in X need not be a path in \(X_l\); see [6, Ex. 3.6]. However, a rectifiable path in X is also continuous as a map into \(X_l\) and therefore a path in \(X_l\). Two rectifiable arcs in X with a common endpoint in \(\partial _l X\), say

$$\begin{aligned}{}[0,\ell (\alpha )]\xrightarrow \alpha X\cup \left\{ \xi \right\} , [0,\ell (\beta )]\xrightarrow \beta X\cup \left\{ \xi \right\} \quad \text {with }\; \alpha (0)=\xi =\beta (0)\in \partial _l X, \end{aligned}$$

are l-equivalent if and only if

$$\begin{aligned} \lim _{s\rightarrow 0^+} l\bigl (\alpha (s),\beta (s)\bigr )=0. \end{aligned}$$

There is a natural one-to-one correspondence between \(\partial _l X\) and the l-equivalence classes of such rectifiable arcs; see [6, Prop. 3.29].

The identity map \(X_l\xrightarrow {{{\,\mathrm{{\textsf{id}}}\,}}}X\) is 1-Lipschitz and so has a 1-Lipschitz extension \(\overline{X_l}\xrightarrow {\iota }\bar{X}\). In general, \(\iota =\iota _X\) need not be surjective nor injective. However, we always have \(\iota (\partial _l X)={\partial ^\textrm{ra}} X\).

We make repeated appeals to the following elementary fact; see [6, Lem. 3.17] and also Fact 2.1.

Lemma 2.4

Let \(X=(X,{{\,\mathrm{|\cdot |}\,}})\) be a rectifiably connected metric space with associated length distance space \(X_l=(X,l)\). Suppose \([0,1)\xrightarrow {\gamma } X\) is a rectifiable path in X. Then

$$\begin{aligned} \lim _{s,t\rightarrow 1^-}\ell \bigl (\gamma \vert _{[s,t]}\bigr )=0=\lim _{s,t\rightarrow 1^-} l\bigl (\gamma (s),\gamma (t)\bigr ) \end{aligned}$$

so there exist points \(z\in \partial X\) and \(\zeta \in \partial _l X\) such that

$$\begin{aligned} \lim _{t\rightarrow 1^-}|\gamma (t)-z|=0=\lim _{t\rightarrow 1^-} l\bigl (\gamma (t),\zeta \bigr ); \end{aligned}$$

therefore there are rectifiable paths

that are obtained by defining

$$\begin{aligned} {\bar{\gamma }}(t):= {\left\{ \begin{array}{ll} \gamma (t) &{} \text { for }t\in [0,1), \\ z &{} \text { for }t=1; \end{array}\right. } \quad \text {and}\quad \tilde{\gamma }(t):= {\left\{ \begin{array}{ll} \gamma (t) &{} \text { for } t\in [0,1), \\ \zeta &{} \text { for } t=1; \end{array}\right. } \end{aligned}$$

Moreover, \(z\in \partial X\) if and only if   \(\zeta \in \partial _l X\). Also, \(\ell (\tilde{\gamma })=\ell ({\bar{\gamma }})=\ell (\gamma )\).

Corollary 2.5

Suppose X is rectifiably connected and \(a,b\in X^\mathrm{{ra}}:=X\cup {\partial ^\textrm{ra}} X\). Then ab can be joined by a rectifiable path in \(X\cup \left\{ a,b\right\} \). Moreover, if \(\gamma :a\curvearrowright b\) in \(X\cup \left\{ a,b\right\} \), then there are unique points \({\tilde{a}},{\tilde{b}}\in \overline{X_l}\) and a rectifiable \(\tilde{\gamma }:{\tilde{a}}\curvearrowright {\tilde{b}}\) in \(X_l\cup \left\{ {\tilde{a}},{\tilde{b}}\right\} \) with \(\gamma =\iota {\circ }\tilde{\gamma }\) and \(\ell (\gamma )=\ell (\tilde{\gamma })\).

We define \(X^\mathrm{{ra}}\times X^\mathrm{{ra}}\xrightarrow {l^\textrm{ra}}[0,+\infty )\) by

$$\begin{aligned} l^\textrm{ra}(a,b):= \inf \bigl \{\ell (\gamma )\,\big \vert \,\hbox {} \gamma :a\curvearrowright b\, \hbox {a rectifiable path in} \; X\cup \left\{ a,b\right\} \bigr \} . \end{aligned}$$

In general, \(l^\textrm{ra}\) need not be a distance on \(X^\mathrm{{ra}}\) because the triangle inequality may fail. However, its restriction \(l^\textrm{ra}_1\) to \(X^\mathrm{{ra}}_1\times X^\mathrm{{ra}}_1\), where

$$\begin{aligned} X^\mathrm{{ra}}_1:=X\cup {\partial ^\textrm{ra}_1}X \quad \text {with}\quad {\partial ^\textrm{ra}_1}X:=\left\{ z\in {\partial ^\textrm{ra}} X \mid {{\,\mathrm{\textsf{card}}\,}}\iota ^{-1}(z)=1\right\} , \end{aligned}$$

is a distance on \(X^\mathrm{{ra}}_1\). The triangle inequality is easy to check if the intermediate point lies in X and not difficult to verify when this point lies in \({\partial ^\textrm{ra}_1}X\).

Setting

$$\begin{aligned} \partial _l^1X:=\left\{ \xi \in \partial _l X\mid \iota ^{-1}\bigl (\iota (\xi )\bigr )=\{\xi \}\right\} ,\quad X_l^1:=X_l\cup \partial _l^1 X,\quad X^\mathrm{{ra}}_l:=(X^\mathrm{{ra}}_1,l^\textrm{ra}_1) \end{aligned}$$

and \(\iota ^1:=\iota \vert _{X_l^1}\), we easily obtain the following.

Lemma 2.6

When X is rectifiably connected, \(X_l^1\xrightarrow {\iota ^1}X^\mathrm{{ra}}_l\) is an isometry.

Proof

Let \({\tilde{a}},{\tilde{b}}\in X_l^1\). Then \(a:=\iota ({\tilde{a}}),b:=\iota ({\tilde{b}})\in X^\mathrm{{ra}}_1\). Let \(\gamma :a\curvearrowright b\) be an arc in \(X\cup \left\{ a,b\right\} \). The ends of \(\gamma \) determine \({\tilde{a}},{\tilde{b}}\), so by Corollary 2.5 there is a rectifiable \(\tilde{\gamma }:{\tilde{a}}\curvearrowright {\tilde{b}}\) in \(X_l\cup \left\{ {\tilde{a}},{\tilde{b}}\right\} \) with \(\gamma =\iota {\circ }\tilde{\gamma }\) and \(\ell (\tilde{\gamma })=\ell (\gamma )\). Thus

$$\begin{aligned} l({\tilde{a}},{\tilde{b}})\le \ell (\tilde{\gamma })=\ell (\gamma ). \end{aligned}$$

Taking an infimum over all such \(\gamma \) gives

$$\begin{aligned} l({\tilde{a}},{\tilde{b}})\le l(a,b) \end{aligned}$$

and the opposite inequality holds because \(\iota \) is 1-Lipschitz. \(\square \)

2.2 CAT(0) Metric Spaces

Here our terminology and notation conforms with that in [4]; also, see [5]. We recall a few fundamental concepts, mostly copied directly from [4].

2.2.1 Geodesic and Comparison Triangles

A geodesic triangle \(\Delta \) in X consists of three points in X, say \(a,b,c\in X\), called the vertices of \(\Delta \) and three geodesics, say \(\alpha :a\curvearrowright b, \beta :b\curvearrowright c, \gamma :c\curvearrowright a\) (that we may write as [ab], [bc], [ca]) called the sides of \(\Delta \). We use the notation

$$\begin{aligned} \Delta =\Delta (\alpha ,\beta ,\gamma ) \quad \text {or}\quad \Delta =[a,b,c]:=[a,b]\star [b,c]\star [c,a] \quad \text {or}\quad \Delta =\Delta (a,b,c) \end{aligned}$$

depending on the context and the need for accuracy.

A Euclidean triangle \({\bar{\Delta }}=\Delta (\bar{a},\bar{b},\bar{c})\) in \({\mathbb C}\) is a comparison triangle for \(\Delta =\Delta (a,b,c)\) provided \(|a-b|=|\bar{a}-\bar{b}|,|b-c|=|\bar{b}-\bar{c}|,|c-a|=|\bar{c}-\bar{a}|\). We also write \({\bar{\Delta }}={\bar{\Delta }}(a,b,c)\) when a specific choice of \(\bar{a},\bar{b},\bar{c}\) is not required. A point \(\bar{x}\in [\bar{a},\bar{b}]\) is a comparison point for \(x\in [a,b]\) when \(|x-a|=|\bar{x}-\bar{a}|\). Assuming that \(b\ne a\ne c\) (so \(\bar{b}\ne \bar{a}\ne \bar{c}\)), the comparison angle of  \(\Delta \) at a is defined to be the interior Euclidean angle of \({\bar{\Delta }}\) at \(\bar{a}\) and denoted by

$$\begin{aligned} \overline{\measuredangle }_{a}(b,c):=\measuredangle ^\textsf{euc}_{\bar{a}}(\bar{b},\bar{c}). \end{aligned}$$

Assume \(a\ne p\ne b\) and let \(\alpha :p\curvearrowright a, \beta :p\curvearrowright b\) be rectifiable arcs in X parameterized by arc length. The (upper) Alexandrov angle between \(\alpha \) and \(\beta \) is defined by

$$\begin{aligned} \measuredangle _p(\alpha ,\beta ):=\limsup _{s,t\rightarrow 0^+} \overline{\measuredangle }_{p}(\alpha (s),\beta (t)); \end{aligned}$$

see [4, 1.12, p.9]. When [pa], [pb] are geodesics, \(\measuredangle _p(a,b):=\measuredangle _p([p,a],[p,b])\).

2.2.2 CAT(0) Definition

A geodesic triangle \(\Delta \) in X satisfies the CAT(0) distance inequality if and only if the distance between any two points of \(\Delta \) is not larger than the Euclidean distance between the corresponding comparison points; that is,

$$\begin{aligned} \forall \; x,y\in \Delta \;\text {and corresponding comparison points}\;\bar{x},\bar{y}\in {\bar{\Delta }}, \quad |x-y| \le |\bar{x}-\bar{y}|. \end{aligned}$$

We also say that \(\Delta \) is CAT(0)-thin when it satisfies the CAT(0) distance inequality.

A geodesic metric space is CAT(0) if and only if each of its geodesic triangles is CAT(0)-thin. A complete CAT(0) metric space is called a Hadamard space. A geodesic metric space X has non-positive curvature if and only if it is locally CAT(0), meaning that for each point \(a\in X\) there is an \(r>0\) (that can depend on a) such that the metric ball \(\textsf{B}(a;r)\) (endowed with the distance inherited from X) is CAT(0).

Of the many conditions which guarantee that a space is CAT(0), for instance, see [4, Prop. 1.7, p. 161] or [5, Thm. 4.3.5, p.116], we mention only that a geodesic metric space X is CAT(0) if and only if each of its geodesic triangles satisfies the CAT(0) vertex angle criterion. Here \(\Delta \) satisfies the CAT(0) vertex angle criterion if and only if \(\Delta \) has distinct vertices and the Alexandrov angle between any two sides of \(\Delta \) is not greater than the interior Euclidean angle between the corresponding sides of a comparison triangle for \(\Delta \); equivalently, if and only if the (Alexandrov) vertex angles of \(\Delta \) are not greater than the corresponding (Euclidean) vertex angles of a comparison triangle for \(\Delta \).

2.2.3 Triangle Tails

Let \(\Delta =[a,b,c]=[a,b]\star [b,c]\star [c,a]\) be a geodesic triangle. Suppose there are points \(b_o\in [a,b]\) and \(c_o\in [a,c]\) such that the subgeodesics \([a,b_o]\subset [a,b]\) and \([a,c_o]\subset [a,c]\) coincide: i.e., \([a,b_o]=[a,c_o]\). This common geodesic segment is a tail of \(\Delta \), and \(\Delta \) is tail-less if there are no such tails.Footnote 4

It is not difficult to verify the following. (If the lengths of two sides of an Euclidean triangle are increased by the same amount, then certain angles also increase.)

Fact 2.7

Let X be a geodesic metric space. Suppose every tail-less geodesic triangle in X satisfies the CAT(0) vertex angle criterion. Then X is CAT(0).

2.2.4 Gromov Hyperbolicity Definition

A geodesic metric space X is \(\delta \)-hyperbolic if and only if for all geodesic triangles \(\Delta \) in X, each edge of \(\Delta \) lies in the \(\delta \)-neighborhood of the union of the other two edges, and X is Gromov hyperbolic if and only if it is \(\delta \)-hyperbolic for some \(\delta \in [0,+\infty )\).

2.3 General Plane Information

We view the Euclidean plane as the complex number field \({\mathbb C}\). Everywhere \(\Omega \) is a plane domain (i.e., an open connected set), \(\Omega ^c:={\mathbb C}{\setminus }\Omega \) and \({\partial \Omega }\) denote the complement and boundary (respectively) of \(\Omega \).

The open disk of radius r centered at the point \(a\in {\mathbb C}\) is

$$\begin{aligned} \textsf{D}(a;r):=\left\{ z:|z-a|<r\right\} , \end{aligned}$$

\({\mathbb D}:=\textsf{D}(0;1)\) is the open unit disk, and the open r-neighborhood of a set \(A\subset {\mathbb C}\) is

$$\begin{aligned} \textsf{N}(A;r):= \bigcup _{a\in A} \textsf{D}(a;r) = \left\{ z:{{\,\mathrm{\textsf{dist}}\,}}(z,A)<r\right\} . \end{aligned}$$

2.3.1 Complex Analysis

The well known Riemann and Carathéodory mapping theorems assert that when \(\Omega \) is a simply connected plane domain, there is a conformal map (i.e., a holomorphic homeomorphism) \(f:{\mathbb D}\rightarrow \Omega \), and if \(\Omega \) is a Jordan domain, f extends to a homeomorphism \(\bar{\mathbb D}\rightarrow {{\bar{\Omega }}}\). So, each boundary point of a Jordan domain is path accessible from the domain.

We repeatedly use the less known fact that when \(\Omega \) is a simply connected plane domain with rectifiable boundary (e.g., if \({\partial \Omega }\) is a rectifiable Jordan loop), then each point of \({\partial \Omega }\) is rectifiably accessible from \(\Omega \); that is, \({\partial ^\textrm{ra}}\Omega ={\partial \Omega }\) and \(\Omega ^\mathrm{{ra}}={{\bar{\Omega }}}\). I am indebted to Distinguished Professor Chris Bishop for explaining this to me. It is a consequence of the fact that any Riemann map onto such a domain belongs to the Hardy class \(\textsf{H}^1\); see the “easy half” of Chris’ result in [2].

A Riemann map \({\mathbb D}\xrightarrow {f}\Omega \) provides a conformal model for the length space \(\overline{\Omega _l}\). Indeed, the conformal metric \(|f'(z)|\,|dz|\) on \({\mathbb D}\) induces the length distance

$$\begin{aligned} d_f(a,b):= & {} \inf _{\gamma :a\curvearrowright b}\int _\gamma |f'(z)|\,|dz| \end{aligned}$$

where the infimum is over all rectifiable arcs \(\gamma :a\curvearrowright b \hbox { in } {\mathbb D}\) and \({\mathbb D}_f:=({\mathbb D},d_f)\xrightarrow {f}(\Omega ,l)=:\Omega _l\) is an isometry. One can demonstrate that \(\overline{{\mathbb D}_f}={\mathbb D}\cup \partial _f{\mathbb D}\), where \(\partial _f{\mathbb D}:=\left\{ \zeta \in \partial {\mathbb D}\;\big \vert \; f\bigl ([0,\zeta )\bigr ) \text { is rectifiable}\right\} \); evidently, \(\partial _f{\mathbb D}\subset \partial {\mathbb D}_f\), and the opposite containment can be established with the help of [10, Prop. 2.14-p.29, Cor. 2.17-p.35, Thm. 4.20-p.88]. Thus \(\overline{\Omega _l}\) is isometrically equivalent to \(\overline{{\mathbb D}_f}\). With this model, the map \(\iota :\overline{\Omega _l}\rightarrow \Omega ^\mathrm{{ra}}\) can be realized as the radial limit extension \(f:\overline{{\mathbb D}_f}\rightarrow \Omega ^\mathrm{{ra}}\subset {{\bar{\Omega }}}\).

For example, we now see that a Jordan loop \(\Lambda \) in \(\overline{\Omega _l}\) corresponds to a Jordan loop in \(\overline{{\mathbb D}_f}\subset \bar{\mathbb D}\) whose interior is a simply connected domain in \({\mathbb D}\) with f image a simply connected \(D\subset \Omega \) satisfying \({\partial D}=\iota (\Lambda )\), which is a rectifiably connected loop (perhaps not Jordan) in \(\Omega ^\mathrm{{ra}}\).

3 Proofs

Here we establish the Theorem stated in the Introduction. Now X is a given simply and rectifiably connected plane set with \(\overline{X_l}\) the metric completion of the intrinsic (Euclidean) length space \(X_l\) associated with X. Also, \(\overline{X_l}\xrightarrow {\iota }\bar{X}\) is the 1-Lip extension of the identity map \(X_l\rightarrow X\).

First, we consider simply connected plane domains, then arbitrary simply and rectifiably connected plane sets.

3.1 CAT(0) Proof for X a Simply Connected Plane Domain

Assume \(X=\Omega \) is a simply connected plane domain. Evidently, \(\overline{\Omega _l}\) is a complete length metric space. We demonstrate that it is a 4-point limit of CAT(0) spaces, so by [4, Thm. 3.9, p.196] it is also CAT(0) and hence a Hadamard space.

The Riemann Mapping Theorem provides a conformal map \({\mathbb D}\xrightarrow {f}\Omega \) (i.e., a holomorphic homeomorphism). Let \((r_\nu )\) be a strictly increasing sequence in (0, 1) with \(r_\nu \nearrow 1\). For each \(\nu \in {\mathbb N}\), define

$$\begin{aligned} f_\nu (\zeta ):=f(r_\nu \zeta ), \quad \Omega _\nu :=f_\nu ({\mathbb D})=f(r_\nu {\mathbb D}), \end{aligned}$$

and let \(U_\nu :=f_\nu (r_\nu {\mathbb D})=f(r_\nu ^2{\mathbb D}),\quad \lambda _\nu :=\lambda _{\Omega _\nu }, \quad M_\nu :=\max _{{{\bar{U}}}_\nu }\lambda _\nu , \text {and} \varepsilon _\nu :=\bigl (10{{\,\mathrm{\textsf{diam}}\,}}_l(\Omega _\nu )M_\nu ^2\bigr )^{-1}\).

Note that \(U_\nu \) is compactly contained in \(\Omega _\nu \) which in turn is compactly contained in \(\Omega \), and

$$\begin{aligned} {{\,\mathrm{\textsf{diam}}\,}}_l(\Omega _\nu )={{\,\mathrm{\textsf{diam}}\,}}(\Omega _\nu ,l) \quad \text { with } l=l_\Omega . \end{aligned}$$

Also, for any plane domain D, \(\lambda _D\,ds\) denotes the Poincaré hyperbolic metric in D, i.e., \(\lambda _D\,ds\) is the maximal complete metric in D with constant Gaussian curvature \(-1\); see [8]. Note that \(\left\{ U_\nu \mid \nu \in {\mathbb N}\right\} \) is an increasing open cover of \(\Omega \) and \((M_\nu ),(\varepsilon _\nu )\) are increasing, decreasing positive sequences with \(M_\nu \nearrow +\infty ,\;\varepsilon _\nu \searrow 0\) respectively.

It is easy to check that \(\overline{\Omega _l}\) is a 4-point limit of the spaces \(({{\bar{\Omega }}}_\nu ,l_\nu )\) where \(l_\nu \) is Euclidean length distance in \(\Omega _\nu \). Since \(\Omega _\nu \) is a Jordan domain, we could appeal to Bishop’s result now, but it is easy to provide a simple alternative argument.

Consider the conformal metric \(\rho _\nu \,ds\) in \(\Omega _\nu \) where

$$\begin{aligned} \rho _\nu :=1+\varepsilon _\nu \lambda _\nu \quad (\text {and note that}\rho _\nu \le 1+\bigl (10{{\,\mathrm{\textsf{diam}}\,}}_l(\Omega _\nu )M_\nu \bigr )^{-1} \text {in } {{\bar{U}}}_\nu ). \end{aligned}$$

Let \(d_\nu \) be the length distance obtained from the metric \(\rho _\nu \,ds\) in \(\Omega _\nu \). Since \(\log \lambda _\nu \) is subharmonic and \({{\mathscr {C}}}^\infty \) smooth in \(\Omega _\nu \), so is \(\log \rho _\nu \) (see [7, 2.1, 2.2]) and therefore by classical results (e.g., see [4, Thm.1A.6, Thm. 4.1, pp. 173,193]) each space \((\Omega _\nu ,d_\nu )\) is CAT(0).

We check that \(\overline{\Omega _l}\) is a 4-point limit of the CAT(0) spaces \((\Omega _\nu ,d_\nu )\).

Let \(x_1,x_2,x_3,x_4\in \overline{\Omega _l}\) and \(\varepsilon \in (0,{{\,\mathrm{\textsf{diam}}\,}}_l\Omega _1)\) be given. Define \(z_1,z_2,z_3,z_4\in \Omega \) as follows: if \(x_i\in \Omega \), let \(z_i:=x_i\); otherwise, \(x_i\in \partial _l\Omega \), and we pick any \(z_i\in \Omega \) with \(l(z_i,x_i)<\varepsilon /10\).

Next, for all \(1\le i<j\le 4\), choose arcs \(\sigma _n^{ij}:z_i\curvearrowright z_j\) in \(\Omega \) with \(\ell (\sigma _n^{ij})\) decreasing to \(l(z_i,z_j)\). Fix N so that for all \(1\le i<j\le 4\), \(n\ge N \implies \ell (\sigma _n^{ij})<l(z_i,z_j)+\varepsilon /10\); so \(\ell (\sigma _n^{ij})<\frac{11}{10}{{\,\mathrm{\textsf{diam}}\,}}_l(\Omega _\nu )\).

For each n, \(K_n:=\bigcup _{1\le i<j\le 4}\sigma _n^{ij}\) is a compact subset of \(\Omega \), so there is an increasing sequence \((\nu _n)_{n\ge N}\) such that for each \(n\ge N\), \(M_{\nu _n}>2\varepsilon ^{-1}\) and \(K_n\subset U_{\nu _n}\subset {{\bar{U}}}_{\nu _n}\subset \Omega _{\nu _n}\). Then for each \(n\ge N\) and all \(1\le i<j\le 4\),

$$\begin{aligned} l(z_i,z_j)&\le l_{\nu _n}(z_i,z_j) \le d_{\nu _n}(z_i,z_j) \le \ell _{\rho _{\nu _n}}(\sigma _n^{ij})=\int _{\sigma _n^{ij}}\rho _{\nu _n}\,ds \\&\le \Big (1+\bigl (10{{\,\mathrm{\textsf{diam}}\,}}_l(\Omega _\nu )M_{\nu _n}\bigr )^{-1}\Bigr )\ell (\sigma _n^{ij}) \quad (\text {because}\, \sigma _{ij}^n\subset K_n\subset U_{\nu _n}) \\&\le \ell (\sigma _n^{ij}) + \frac{11}{100M_{\nu _n}}< l(z_i,z_j)+\frac{\varepsilon }{10}+\frac{11\varepsilon }{200} < l(z_i,z_j)+\frac{\varepsilon }{5} \end{aligned}$$

and so

$$\begin{aligned} l(x_i,x_j)&\le l(x_i,z_i) + l(z_i,z_j) + l(z_j,x_j) \le d_{\nu _n}(z_i,z_j)+\frac{\varepsilon }{5} \le l(z_i,z_j)+\frac{2\varepsilon }{5} \\&\le l(z_i,x_i) + l(x_i,x_j) + l(x_j,z_j) + \frac{2\varepsilon }{5} \le l(x_i,x_j) + \frac{3\varepsilon }{5}. \end{aligned}$$

Thus, for all \(n\ge N\) and \(1\le i<j\le 4\): \(z_i,z_j\in \Omega _{\nu _n}\) and \(|l(x_i,x_j)-d_{\nu _n}(z_i,z_j)|<\varepsilon \). \(\square \)

3.2 CAT(0) Proof for General Case

Let X be a simply and rectifiably connected plane set. Our primary goal here is to demonstrate that \(\overline{X_l}\) is uniquely geodesic; the CAT(0) property follows.

Since X is simply connected, whenever \(\Lambda \) is a Jordan loop in X, \({{\mathscr {D}}}(\Lambda ):=\Lambda \cup {{\,\mathrm{\textsf{Int}}\,}}(\Lambda )\subset X\). As we employ this observation again and again, it is worthwhile to review methods for constructing Jordan loops.

Given distinct points pq in X,

$$\begin{aligned} \Gamma (p,q):=\left\{ \text {all rectifiable arcs}\;\gamma :p\curvearrowright q \;\text {in} X\right\} \ne \varnothing . \end{aligned}$$

Suppose \(\beta ,\gamma \in \Gamma (p,q)\) and there is a point \(c\in \gamma {\setminus }\beta \). There are several ways to construct a Jordan loop \(\Lambda \) in X that contains an open subarc of \(\gamma \) which in turn contains c. Most simply, we move backwards, forwards along \(\gamma \) from c (towards pq respectively) and let ab be (respectively) the first points of \(\beta \cap \gamma \). Here \(\Lambda :=\gamma [a,b]\cup \beta [a,b]\) has the asserted properties with \(c\in \gamma (a,b)\).

A minor possible problem is that we do not know the order of ab along \(\beta \). To remedy this, set \(b_1:=b\) and then, move backwards along \(\gamma \) (from c to p), and let \(a_1\) be the first point of \(\beta [p,b_1]\cap \gamma \). Now \(\Lambda _1:=\gamma [a_1,b_1]\star \beta ^{-1}[b_1,a_1]\) has the asserted properties and \(p\le a<b\le q\) along both \(\beta \) and \(\gamma \). Yet another alternative is to set \(a_2:=a\), move forwards along \(\gamma \) (from c to q), let \(b_2\) be the first point of \(\beta [a_2,q]\cap \gamma \), and use \(\Lambda _2:=\gamma [a_2,b_2]\star \beta ^{-1}[b_2,a_2]\). Note that the three Jordan loops \(\Lambda ,\Lambda _1,\Lambda _2\) could all be different.

For definitiveness, we always use the first alternative construction.

We assume \({{\,\mathrm{\textsf{int}}\,}}(X)\ne \varnothing \), so

$$\begin{aligned} {{\mathscr {O}}}:=\left\{ \text {all components}\; \Omega \;\text {of} {{\,\mathrm{\textsf{int}}\,}}(X)\right\} \ne \varnothing . \end{aligned}$$

Note that even if some \(\Omega \in {{\mathscr {O}}}\) has non-rectifiably accessible boundary points, \({{\bar{\Omega }}}\subset X\) is still possible. For each \(\Omega \in {{\mathscr {O}}}\), \(\overline{\Omega _l}\xrightarrow {\iota _\Omega }{{\bar{\Omega }}}\) is the 1-Lip extension of the identity map \(\Omega _l\rightarrow \Omega \).

The following facts are useful.

  1. (3.1a)

    Rectifiable Jordan loops. Suppose \(\Lambda \) is a rectifiable Jordan loop in X. Then there is a unique \(\Omega \in {{\mathscr {O}}}\) with \({{\mathscr {D}}}(\Lambda )\subset \Omega ^\mathrm{{ra}}\cap X\), and if \(\Lambda \cap \Omega =\varnothing \), then \(\Omega ={{\,\mathrm{\textsf{Int}}\,}}(\Lambda )\).

  2. (3.1b)

    Components of  \({{\,\mathrm{\textsf{int}}\,}}(X)\). For distinct \(\Omega _1,\Omega _2\in {{\mathscr {O}}}\), \({{\,\mathrm{\textsf{card}}\,}}(\Omega ^\mathrm{{ra}}_1\cap \Omega ^\mathrm{{ra}}_2\cap X)\le 1\).

  3. (3.1c)

    Unique length boundary points. For each \(\Omega \in {{\mathscr {O}}}\), \(z\in {\partial ^\textrm{ra}}\Omega \cap X \implies {{\,\mathrm{\textsf{card}}\,}}\iota _\Omega ^{-1}(z)=1\).Footnote 5

Proof of ()

Since X is simply connected, \({{\mathscr {D}}}(\Lambda )\subset X\), so \(D:={{\,\mathrm{\textsf{Int}}\,}}(\Lambda )\subset X\) and there is an \(\Omega \in {{\mathscr {O}}}\) with \(D\subset \Omega \). Evidently, \({{\mathscr {D}}}(\Lambda )={{\bar{D}}}=D^\textrm{ra}\subset \Omega ^\mathrm{{ra}}\cap X\); see Sect. 2.3.1. Fix a point \(o\in \textsf{D}\). Given \(p\in \Omega \), let \(\alpha :o\curvearrowright p\) in \(\Omega \). If \(\Lambda \cap \Omega =\varnothing \), then \(\alpha \cap \Lambda =\varnothing \), so \(\alpha \subset D\) whence \(p\in D\) and \(\Omega =D\). \(\square \)

Proof of ()

Let ab be distinct points in \(\Omega ^\mathrm{{ra}}_1\cap \Omega ^\mathrm{{ra}}_2\cap X\) for some \(\Omega _1,\Omega _2\in {{\mathscr {O}}}\). For \(j\in \left\{ 1,2\right\} \), pick rectifiable arcs \(\alpha _j:a\curvearrowright b\) in \(\Omega _j\cup \left\{ a,b\right\} \). Since , , so we may assume that \(\alpha _1\cap \alpha _2=\left\{ a,b\right\} \). Then \(\Lambda :=\alpha _1\star \alpha _2^{-1}\) is a rectifiable Jordan loop in X, so by (3.1a) there is a unique \(\Omega \in {{\mathscr {O}}}\) with \({{\mathscr {D}}}(\Lambda )\subset \Omega ^\mathrm{{ra}}\cap X\).

Fix a point \(o\in D:={{\,\mathrm{\textsf{Int}}\,}}(\Lambda )\) and points . Then for each \(j\in \left\{ 1,2\right\} \), any \(z\in \Omega _j\) can be joined to \(z_j\) (by a rectifiable path in \(\Omega _j\)) and then to o (by a rectifiable path in \(\Omega \cup \left\{ z_j\right\} \)), so there is a rectifiable path \(z\curvearrowright o\) in \({{\,\mathrm{\textsf{int}}\,}}(X)\). It follows that \(\Omega _1=\Omega =\Omega _2\). \(\square \)

Proof of ()

Suppose \(z\in {\partial ^\textrm{ra}}\Omega \cap X\) for some \(\Omega \in {{\mathscr {O}}}\). Let \(\alpha ,\beta \) be rectifiable arcs in \(\Omega \cup \left\{ z\right\} \) both having z an endpoint. We show that \(\alpha \) and \(\beta \) determine the same point in \(\partial _l\Omega \).

Assume \([0,\ell (\alpha )]\xrightarrow {\alpha }\Omega \cup \left\{ z\right\} \) and \([0,\ell (\beta )]\xrightarrow {\beta }\Omega \cup \left\{ z\right\} \) with \(\alpha (0)=z=\beta (0)\). We verify that \(\lim _{s\rightarrow 0+} l\bigl (\alpha (s),\beta (s)\bigr )=0\).

First, suppose that for all \(\upsilon \in \bigl (0,\ell (\alpha )\wedge \ell (\beta )\bigr )\), \(\alpha \bigl ((0,\upsilon )\bigr )\cap \beta \bigl ((0,\upsilon )\bigr )\ne \varnothing \). Given such an \(\upsilon \), pick \(\sigma ,\tau \in (0,\upsilon )\) with \(\alpha (\sigma )=\beta (\tau )\). Then for any \(s\in (0,\upsilon )\),

$$\begin{aligned} l\bigl (\alpha (s),\beta (s)\bigr ) \le l\bigl (\alpha (s),\alpha (\sigma )\bigr ) + l\bigl (\alpha (\sigma ),\beta (\tau )\bigr ) + l\bigl (\beta (\tau ),\beta (s)\bigr ) \le 2\upsilon < \varepsilon \end{aligned}$$

provided \(\upsilon <2\varepsilon \).

Otherwise, we may assume \(\alpha \cap \beta =\left\{ z\right\} \). Let \(\gamma \) be a rectifiable arc in \(\Omega \) from the terminal point a of \(\alpha \) to the terminal point b of \(\beta \). Then \(\alpha \star \beta ^{-1}\) and \(\gamma \) are arcs in \(\Gamma (a,b)\) and \(z\in (\alpha \star \beta ^{-1}){\setminus }\gamma \), so there is a rectifiable Jordan loop \(\Lambda \) in X that contains an open subarc of \(\alpha \star \beta ^{-1}\) which in turn contains z. Evidently, \({{\,\mathrm{\textsf{Int}}\,}}(\Lambda )\subset \Omega \) and \({{\mathscr {D}}}(\Lambda )\subset \Omega ^\mathrm{{ra}}\cap X\), and thus \(\lim _{s\rightarrow 0+} l\bigl (\alpha (s),\beta (s)\bigr )=0\). \(\square \)

3.2.1 Entry and Exit Points

Let pq be distinct points in X and \(\Omega \in {{\mathscr {O}}}\). We say that \(\gamma \in \Gamma (p,q)\) enters \(\Omega \) if \({{\,\mathrm{\textsf{card}}\,}}(\gamma \cap \Omega ^\mathrm{{ra}})\ge 2\). We employ the following crucial facts.

  1. (3.2a)

    Both points in \(\Omega ^\mathrm{{ra}}\).    \({{\,\mathrm{\textsf{card}}\,}}\bigl (\left\{ p,q\right\} \cap \Omega ^\mathrm{{ra}}\bigr )=2 \implies \forall \;\gamma \in \Gamma (p,q), \;\gamma \subset \Omega ^\mathrm{{ra}}.\)

  2. (3.2b)

    One point in \(\Omega ^\mathrm{{ra}}\).    \({{\,\mathrm{\textsf{card}}\,}}\bigl (\left\{ p,q\right\} \cap \Omega ^\mathrm{{ra}}\bigr )=1 \implies \exists \; e:=e_\Omega \in {\partial ^\textrm{ra}}\Omega \cap X\) such that \(\forall \;\gamma \in \Gamma (p,q)\)

    $$\begin{aligned}&\left\{ p,q\right\} \cap \Omega ^\mathrm{{ra}}=\left\{ p\right\} \implies e\in \gamma ,\;\gamma [p,e]\subset \Omega ^\mathrm{{ra}}, \;\text {and}\; \gamma (e,q]\cap \Omega ^\mathrm{{ra}}=\varnothing , \\&\left\{ p,q\right\} \cap \Omega ^\mathrm{{ra}}=\left\{ q\right\} \implies e\in \gamma ,\;\gamma [e,q]\subset \Omega ^\mathrm{{ra}}, \;\text {and}\; \gamma [p,e)\cap \Omega ^\mathrm{{ra}}=\varnothing . \end{aligned}$$
  3. (3.2c)

    Neither point in \(\Omega ^\mathrm{{ra}}\).    \({{\,\mathrm{\textsf{card}}\,}}\bigl (\left\{ p,q\right\} \cap \Omega ^\mathrm{{ra}}\bigr )=0 \implies \) if some arc in \(\Gamma (p,q)\) enters \(\Omega \), then \(\exists \;a:=a_\Omega ,b:=b_\Omega \in {\partial ^\textrm{ra}}\Omega \cap X\) such that \(\forall \;\gamma \in \Gamma (p,q)\)

    $$\begin{aligned} a,b\in \gamma ,\; \gamma [a,b]\subset \Omega ^\mathrm{{ra}},\; \text { and } \bigl (\gamma [p,a)\cup \gamma (b,q]\bigr )\cap \Omega ^\mathrm{{ra}}=\varnothing . \end{aligned}$$

The points ab in (3.2c) (and e in (3.2b)) are called entry, exit points (respectively) for \(\Omega \) relative to pq. These entry, exit points depend only on pq, and \(\Omega \).

Proof of

Assume \(p,q\in \Omega ^\mathrm{{ra}}\cap X\) and let \(\gamma \in \Gamma (p,q)\). We show that \(\gamma \subset \Omega ^\mathrm{{ra}}\).

Let \(\alpha :p\curvearrowright q\) be a rectifiable arc in \(\Omega \cup \left\{ p,q\right\} \subset \Omega ^\mathrm{{ra}}\cap X\). Suppose there is a point \(o\in \gamma {\setminus }\alpha \). As discussed in the third paragraph at the beginning of this subsection, there are points \(a,b\in \alpha \cap \gamma \) such that \(p\le a<b\le q\) along both \(\alpha \) and \(\gamma \) with \(\Lambda :=\gamma [a,b]\star \alpha ^{-1}[b,a]\) a rectifiable Jordan loop in X and with \(o\in \gamma (a,b)\).

By (3.1a) there is a unique \(\Omega _o\in {{\mathscr {O}}}\) with \({{\mathscr {D}}}(\Lambda )\subset \Omega ^\mathrm{{ra}}_o\cap X\). We claim that \(D:={{\,\mathrm{\textsf{Int}}\,}}(\Lambda )\subset \Omega \), so \(\Omega =\Omega _o\) and \(o\in \Omega ^\mathrm{{ra}}\), and as o is an arbitrary point of \(\gamma {\setminus }\alpha \), \(\gamma \subset \Omega ^\mathrm{{ra}}\) as asserted.

Let \(z\in D\) and fix any point \(c\in \alpha (a,b)\). Since \(\Lambda ={\partial D}\) is rectifiable, there is a rectifiable arc \(\beta :z\curvearrowright c\) in \(D\cup \left\{ c\right\} \). As \(c\in \Omega \subset {{\,\mathrm{\textsf{int}}\,}}(X)\) and \(D={{\,\mathrm{\textsf{Int}}\,}}(\Lambda )\subset \Omega _o\subset {{\,\mathrm{\textsf{int}}\,}}(X)\), \(\beta \subset {{\,\mathrm{\textsf{int}}\,}}(X)\) and so \(z\in \beta \subset \Omega \). \(\square \)

The proof of (3.2b) is similar to, but easier than, the proof of (3.2c) and so left to the reader.

Proof of

Assume \(\gamma \in \Gamma (p,q)\) enters \(\Omega \). There are distinct points \(p_o,q_o\in \gamma \cap \Omega ^\mathrm{{ra}}\) and we label these so that \(p<p_o<q_o<q\) along \(\gamma \). According to (3.2a), \(\gamma [p_o,q_o]\subset \Omega ^\mathrm{{ra}}\). Roughly speaking, ab are the endpoints of the maximal subarc of \(\gamma \) that contains \(\gamma [p_o,q_o]\) and lies in \(\Omega ^\mathrm{{ra}}\). Some care is required because \(\Omega ^\mathrm{{ra}}\) need not be closed in \({\mathbb {C}}\) nor in X.

The set \(A{:=}\left\{ z{\in }\gamma [p,p_o]\;\big |\;z{\in }\Omega ^\mathrm{{ra}}\right\} \) is non-empty and bounded below, so it has a greatest lower bound a. Similarly, there is a least upper bound b for \(B:=\left\{ z\in \gamma [q_o,q]\;\big |\;z\in \Omega ^\mathrm{{ra}}\right\} \). Clearly \(a,b\in \gamma \cap {\partial \Omega }\), \(\bigl (\gamma [p,a)\cup \gamma (b,q]\bigr )\cap \Omega ^\mathrm{{ra}}=\varnothing \), and it is not difficult to check that \(\gamma (a,b)\subset \Omega ^\mathrm{{ra}}\).

To corroborate that \(a,b\in {\partial ^\textrm{ra}}\Omega \), we employ (3.1c) in conjunction with Lemma 2.6 as follows. Since \(\gamma (a,b)\subset \Omega ^\mathrm{{ra}}\cap X\), it lies in the image of the isometry \(\iota ^1_\Omega :\Omega _l^1\rightarrow \Omega ^\mathrm{{ra}}_l\). Thus \(\tilde{\gamma }:=(\iota ^1_\Omega )^{-1}{\circ }\gamma (a,b)\) is a rectifiable arc in \(\Omega _l^1\) and so has endpoints \({\tilde{a}},{\tilde{b}}\) that we label to have \(\iota _\Omega \) images ab. Thus \(a=\iota _\Omega ({\tilde{a}}),b=\iota _\Omega ({\tilde{b}})\in \iota _\Omega (\partial _l\Omega )={\partial ^\textrm{ra}}\Omega \).

Let \(\beta \in \Gamma (p,q)\). The path \(\gamma ^{-1}[a,p]\star \beta \star \gamma ^{-1}[q,b]\) contains a rectifiable arc \(a\curvearrowright b\) that must lie in \(\Omega ^\mathrm{{ra}}\). Since \(\bigl (\gamma [p,a)\cup \gamma (b,q]\bigr )\cap \Omega ^\mathrm{{ra}}=\varnothing \), it must be that \(a,b\in \beta \), so \(\beta [a,b]\subset \Omega ^\mathrm{{ra}}\). If, e.g., there were a point \(c\in \beta (b,q]\cap \Omega ^\mathrm{{ra}}\), then letting d be the first point of \(\beta [c,q]\) in \(\gamma [b,q]\) would give an arc \(\gamma [b,d]\star \beta ^{-1}[d,c]\), but \(b,c\in \Omega ^\mathrm{{ra}}\) would imply \(\gamma [b,d]\subset \Omega ^\mathrm{{ra}}\) violating our choice of b. Similarly \(\beta [p,a)\cap \Omega ^\mathrm{{ra}}=\varnothing \). \(\square \)

Here is a noteworthy consequence of (3.1c) and (3.2a):

$$\begin{aligned} \forall \;\Omega \in {{\mathscr {O}}}, \quad \text {there is an isometric embedding}\;\; \overline{\Omega _l}\xrightarrow {h_\Omega }\overline{X_l}\text { with } \iota _X{\circ } h_\Omega =\iota _\Omega . \end{aligned}$$
(3.3)

Proof of (3.3)

The identity map induces a 1-Lipschitz embedding which then has a 1-Lipschitz extension \(h_\Omega :\overline{\Omega _l}\rightarrow \overline{X_l}\). We explain why \({{\,\mathrm{{\textsf{id}}}\,}}_{\Omega _l}\) is an isometric embedding.

Fix \(a,b\in \Omega \) and let \(\gamma \in \Gamma (a,b)\). By (3.2a), \(\gamma \subset \Omega ^\mathrm{{ra}}\cap X\), so by (3.1c) \(\gamma \) lies in the image of the isometry \(\iota ^1_\Omega :\Omega _l^1\rightarrow \Omega ^\mathrm{{ra}}_l\); see Lemma 2.6. Thus \(\tilde{\gamma }:=(\iota ^1_\Omega )^{-1}{\circ }\gamma \) is a rectifiable arc in \(\Omega _l^1\) with \(\ell (\gamma )=\ell (\tilde{\gamma })\ge l_\Omega (a,b)\). Taking an infimum over all such arcs \(\gamma \), and using the fact that \({{\,\mathrm{{\textsf{id}}}\,}}_{\Omega _l}\) is 1-Lipschitz, we now obtain \(l_X(a,b)=l_\Omega (a,b)\). It now follows that \(h_\Omega \) is an isometric embedding.

Evidently, \(h_\Omega (z)=z\) for \(z\in \Omega \). Suppose \(\zeta \in \partial _l\Omega \). Let \([0,1)\xrightarrow {\alpha }\Omega \) be a rectifiable path that represents \(\zeta \). According to Lemma 2.4, \(\alpha \) extends to rectifiable arcs \(\tilde{\alpha }\) in \(\overline{\Omega _l}\) and \({\bar{\alpha }}=\iota _\Omega {\circ }\tilde{\alpha }\) in \(\Omega ^\mathrm{{ra}}\subset {{\bar{\Omega }}}\) with \(\zeta =\tilde{\alpha }(1)\) and \(z={\bar{\alpha }}(1)=\iota _\Omega (\zeta )\in {\partial ^\textrm{ra}}\Omega \).

But, \(\alpha ={{\,\mathrm{{\textsf{id}}}\,}}_{\Omega _l}{\circ }\alpha \) is also a rectifiable arc in X and so has extensions \({\bar{\alpha }}\) in \(X^\mathrm{{ra}}\) (i.e., \(z\in X^\mathrm{{ra}}\)) and \(\alpha _X\) in \(\overline{X_l}\). If \(z\in X\), then \(z\in \Omega ^\mathrm{{ra}}\cap X\), so \(\left\{ \zeta \right\} =\iota _\Omega ^{-1}(z)\), \(h_\Omega (\zeta )=z\), and \(\iota _X\bigl (h_\Omega (\zeta )\bigr )=\iota _X(z)=z=\iota _\Omega (\zeta )\). If \(z\in {\partial ^\textrm{ra}} X\), then \(\xi :=\alpha _X(1)\in \partial _l X\), \(z=\iota _X(\xi )\), and \(h_\Omega (\zeta )=\xi \), so \(\iota _X\bigl (h_\Omega (\zeta )\bigr )=\iota _X(\xi )=z=\iota _\Omega (\zeta )\). \(\square \)

To simplify notation, often we identify \(\overline{\Omega _l}\) with its image \(h_\Omega (\overline{\Omega _l})\subset \overline{X_l}\), but we must remember that some pointsFootnote 6 in \(\partial _l\Omega \) may lie in X (and so not in \(\partial _l X\)).

We require similar information to deal with points in \(\partial _l X\). Given distinct points pq in \(\overline{X_l}\), define

$$\begin{aligned} \Gamma _l(p,q):=\left\{ \text {all rectifiable arcs}\;\gamma :p\curvearrowright q\;\text {in} X\cup \left\{ p,q\right\} \right\} \ne \varnothing \end{aligned}$$

and

$$\begin{aligned} {\bar{\Gamma }}_l(p,q):=\left\{ \text {all rectifiable arcs}\;\gamma :p\curvearrowright q \;\text {in }\overline{X_l}\right\} \ne \varnothing . \end{aligned}$$

The arcs in \(\Gamma _l(p,q)\) are easier to work with (and all we need to compute l(pq)), but facts about \({\bar{\Gamma }}_l(p,q)\) will help us establish uniqueness of l-geodesics.

Let pq be distinct points in \(\overline{X_l}\) and \(\Omega \in {{\mathscr {O}}}\). The arcs in \({\bar{\Gamma }}(p,q)\) also have unique entry, exit points as in (3.2c); here \(\gamma \) enters \(\overline{\Omega _l}\) if \({{\,\mathrm{\textsf{card}}\,}}(\gamma \cap \overline{\Omega _l})\ge 2\). We identify \(\overline{\Omega _l}\) with its image \(h_\Omega (\overline{\Omega _l})\subset \overline{X_l}\).

  1. (3.4a)

    Both points in \(\overline{\Omega _l}\).    \({{\,\mathrm{\textsf{card}}\,}}\bigl (\left\{ p,q\right\} \cap \overline{\Omega _l}\bigr )=2 \implies \forall \;\gamma \in {\bar{\Gamma }}_l(p,q), \;\gamma \subset \overline{\Omega _l}\).

  2. (3.4b)

    One point in \(\overline{\Omega _l}\).    \({{\,\mathrm{\textsf{card}}\,}}\bigl (\left\{ p,q\right\} \cap \overline{\Omega _l}\bigr )=1 \implies \exists \; e:=e_\Omega \in {\partial ^\textrm{ra}}\Omega \cap X \text { such that } \forall \;\gamma \in {\bar{\Gamma }}_l(p,q)\)

    $$\begin{aligned}&\left\{ p,q\right\} \cap \overline{\Omega _l}=\left\{ p\right\} \implies e\in \gamma ,\; \gamma [p,e]\subset \overline{\Omega _l}, \;\text {and}\; \gamma (e,q]\cap \overline{\Omega _l}=\varnothing , \\&\left\{ p,q\right\} \cap \overline{\Omega _l}=\left\{ q\right\} \implies e\in \gamma ,\; \gamma [e,q]\subset \overline{\Omega _l}, \;\text {and}\; \gamma [p,e)\cap \overline{\Omega _l}=\varnothing . \end{aligned}$$
  3. (3.4c)

    Neither point in \(\overline{\Omega _l}\).    \({{\,\mathrm{\textsf{card}}\,}}\bigl (\left\{ p,q\right\} \cap \overline{\Omega _l}\bigr )=0 \implies \) if some arc in \({\bar{\Gamma }}_l(p,q)\) enters \(\overline{\Omega _l}\), then \(\exists \;a:=a_\Omega ,b:=b_\Omega \in {\partial ^\textrm{ra}}\Omega \cap X\) such that \(\forall \;\gamma \in {\bar{\Gamma }}_l(p,q)\)

    $$\begin{aligned} a,b\in \gamma ,\; \gamma [a,b]\subset \overline{\Omega _l}, \text { and } \bigl (\gamma [p,a)\cup \gamma (b,q]\bigr )\cap \overline{\Omega _l}=\varnothing . \end{aligned}$$

Again, we call the points ab in (3.4c) (and e in (3.4b)) entry, exit points (respectively) for \(\overline{\Omega _l}\) relative to pq. These entry, exit points depend only on pq, and \(\Omega \).

Proof of

Suppose \(p,q\in \overline{\Omega _l}\) and \(\gamma \in {\bar{\Gamma }}_l(p,q)\), but \(\gamma \not \subset \overline{\Omega _l}\). By replacing \(\gamma \) with an appropriate subarc, we may assume \(\gamma \cap \overline{\Omega _l}=\left\{ p,q\right\} \). Let \(p_1,q_1\in \gamma \) be the first pointsFootnote 7 at distance \(d:=\frac{1}{10}l(p,q)\) from pq respectively. Put \(\varepsilon :=d\wedge {{\,\mathrm{\textsf{dist}}\,}}_l(\gamma [p_1,q_1],\overline{\Omega _l})\).

Pick any points \(p_0,q_0\in \Omega \) with \(l(p_0,p)<\varepsilon , l(q_0,q)<\varepsilon \). Mimicking the proof of Lemma 2.2 gives us a rectifiable arc \(\alpha :p_0\curvearrowright q_0\) in X with \(\alpha \subset \textsf{N}_l(\gamma ;\varepsilon )\). According to (3.2a), \(\alpha \subset \Omega ^\mathrm{{ra}}\cap X\subset \overline{\Omega _l}\). Since \(l(p,q)<l(p_0,q_0)+2\varepsilon \),

$$\begin{aligned} l(p_0,q_0)\ge 10d-2\varepsilon . \end{aligned}$$

Fix a point \(a\in \alpha \) with \(l(a,p_0)=l(a,q_0)\ge \frac{1}{2}l(p_0,q_0)\ge 5d-\varepsilon \). Evidently, for all \(b\in \gamma [p_1,q_1]\),

$$\begin{aligned} a\in \alpha \subset \overline{\Omega _l}\implies l(b,a)\ge {{\,\mathrm{\textsf{dist}}\,}}_l(b,\overline{\Omega _l})\ge {{\,\mathrm{\textsf{dist}}\,}}_l(\gamma [p_1,q_1],\overline{\Omega _l})\ge \varepsilon . \end{aligned}$$

Also, if \(b\in \gamma [p,p_1]\), then

$$\begin{aligned} l(b,a)\ge l(p,a)-l(p,b)\ge l(p_0,a)-l(p_0,p)-l(p,b)\ge 4d-2\varepsilon \ge 2\varepsilon . \end{aligned}$$

Similarly, \(b\in \gamma [q_1,q]\implies l(b,a)>\varepsilon \). This contradicts \(\alpha \subset \textsf{N}_l(\gamma ;\varepsilon )\). \(\square \)

Items (3.4b) and (3.4c) now readily follow. To see that the entry and exit points lie in X, note that we can use arcs in \(X\cup \left\{ p,q\right\} \) to determine these points.

3.2.2 Stable Points

Given distinct \(p,q\in \overline{X_l}\), we call \(x\in \overline{X_l}\) a (p,q)-stable point if x lies in every \(\gamma \in {\bar{\Gamma }}_l(p,q)\). Let \(\Sigma (p,q)\) be the set of all (pq)-stable points. Evidently, pq, and all entry and exit points associated with pq belong to \(\Sigma (p,q)\). It is not difficult to see that \(\Sigma (p,q)\) is closed in \(X\cup \left\{ p,q\right\} \), ordered via any arc in \({\bar{\Gamma }}(p,q)\), and \(l(p,q)=l(p,x)+l(x,q)\) for any \(x\in \Sigma (p,q)\). To see that all arcs induce the same ordering on any \(x,y\in \Sigma (p,q)\): note that if \(\beta ,\gamma \in {\bar{\Gamma }}_l(p,q)\) with \(x<y, y<x\) along \(\gamma ,\beta \), respectively, then \(\gamma [p,x]\star \beta [x,q]\) is a path (which contains an arc) \(p\curvearrowright q\) but avoids y contradicting \(y\in \Sigma (p,q)\).

By Lemma 2.2, \(x\in \Sigma (p,q)\) provided \(x\in \gamma \) for all \(\gamma \in \Gamma _l(p,q)\). Also

$$\begin{aligned} \Sigma (p,q)=\left\{ p,q\right\} \cup \bigcup _{x<y}\Sigma (x,y) \end{aligned}$$

where the union is over all \(x,y\in X\cap \Sigma (p,q)\), and,

$$\begin{aligned} x,y\in \Sigma (p,q) \text { with } p\le x<y\le q \implies x\in \Sigma (p,y). \end{aligned}$$

Indeed, \(x,y\in \gamma \in \Gamma _l(p,q)\text { and }\exists \; x\not \in \beta \in \Gamma _l(p,y)\implies x\not \in \beta \star \gamma [y,q]\in \Gamma _l(p,q)\).

Here are two especially useful facts.

(3.5a)

Consequently,

$$\begin{aligned} \Sigma (p,q)=\left\{ p,q\right\} \implies \exists \;\Omega \in {{\mathscr {O}}}\text { such that } p,q\in \overline{\Omega _l}\subset \overline{X_l}. \end{aligned}$$
(3.5b)

Proof Sketch for (3.5)

Suppose \(\gamma \in \Gamma _l(p,q)\) and \(z\in \gamma {\setminus }\Sigma (p,q)\subset X\). Pick an arc \(\beta \in \Gamma _l(p,q)\) with \(z\not \in \beta \). Again, we can construct a Jordan loop \(\Lambda \) in X that contains an open subarc of \(\gamma \) which in turn contains z, but a wee bit of care is required. Since the ends of \(\beta ,\gamma \) both determine the same points \(p,q\in \overline{X_l}\), there are points \(p_1,p_2\) and \(q_1,q_2\) on \(\beta ,\gamma \) respectively, close in \(X_l\), and as close to pq as desired. Then \(\gamma [p_1,q_1]\cup \beta [p_2,q_2]\), together with short arcs \(p_1\curvearrowright p_2, q_1\curvearrowright q_2\) in X, forms a loop in X which contains the asserted Jordan loop \(\Lambda \); see Lemma 2.3 for details. Now (3.5b) follows from (3.1a).

To corroborate (3.5b), start with any \(\gamma \in \Gamma _l(p,q)\). By (3.5b), for each , there is an \(\Omega _z\in {{\mathscr {O}}}\) and an arc \(\alpha _z\subset \gamma \cap \Omega ^\mathrm{{ra}}_z\) with . If \(w\in \alpha _z\), then , so by (3.1b) \(\Omega _w=\Omega _z\). It now follows that there is a single \(\Omega \in {{\mathscr {O}}}\) with , but then \(\gamma \subset \overline{\Omega _l}\). \(\square \)

3.2.3 Contructing Geodesics

Let \(p,q\in \overline{X_l}\). We exhibit an l-geodesic \(p\curvearrowright q\) in \(\overline{X_l}\).

Assume \(p,q\in X\). Suppose there exists an \(\Omega \in {{\mathscr {O}}}\) with \(p,q\in \Omega ^\mathrm{{ra}}\). By (3.1c), there are unique points \({\tilde{p}},{\tilde{q}}\in \overline{\Omega _l}\) with \(p=\iota _\Omega ({\tilde{p}}),q=\iota _\Omega ({\tilde{q}})\). By Sect. 3.1, there is a unique l-geodesic \(\sigma _\Omega :{\tilde{p}}\curvearrowright {\tilde{q}}\) in \(\overline{\Omega _l}\). Then by (3.3) and its proof, \(\sigma _X:=h_\Omega {\circ }\sigma _\Omega \) is an l-geodesic in \(\overline{X_l}\) with endpoints \(h_\Omega ({\tilde{p}})=p\) and \(h_\Omega ({\tilde{q}})=q\).

Now suppose that for all \(\Omega \in {{\mathscr {O}}}\), \(\left\{ p,q\right\} \not \subset \Omega ^\mathrm{{ra}}\). We construct a path \(\sigma :p\curvearrowright q\) in \(\overline{X_l}\) that has \(\ell (\sigma )\le \ell (\gamma )\) for all \(\gamma \in \Gamma (p,q)\). Thus \(\sigma \) is a shortest path and hence an l-geodesic

If \(p\in \Omega ^\mathrm{{ra}}_p\) for some \(\Omega _p\in {{\mathscr {O}}}\), let \(e_p\in {\partial ^\textrm{ra}}\Omega _p\cap X\) be the exit point associated with \(q,\Omega _p\) as given in (3.2b) and let \(\sigma _p\) be the \(h_{\Omega _p}\) image of the l-geodesic \({\tilde{p}}\curvearrowright {\tilde{e}}_p\) in \(\overline{(\Omega _p)_l}\) where \({\tilde{p}},{\tilde{e}}_p\) are the unique points in \(\overline{(\Omega _p)_l}\) with \(p=\iota _{\Omega _p}({\tilde{p}}),e_p=\iota _{\Omega _p}({\tilde{e}}_p)\). If no such \(\Omega _p\) exists, put \(\Omega _p:=\varnothing , e_p:=p, \sigma _p:=\left\{ p\right\} \). Define \(\Omega _q,e_q,\sigma _q\) in a similar manner.

Let \(\gamma \in \Gamma (p,q)\). Note that \(\left\{ e_p,e_q\right\} \subset \gamma \). Suppose \(z\in \gamma [e_p,e_q]{\setminus }\Sigma (p,q)\). By (3.5b), there is an \(\Omega \in {{\mathscr {O}}}\) and an arc \(\alpha \subset \gamma [e_p,e_q]\cap \Omega ^\mathrm{{ra}}\) with . Thus \(\gamma \) enters \(\Omega \) and so \(z\in \gamma (a_\Omega ,b_\Omega )\) where \(a_\Omega ,b_\Omega \in {\partial ^\textrm{ra}}\Omega \cap X\) are the entry, exit points associated with \(\Omega \) as given in (3.2c).

It now follows that \(\gamma [e_p,e_q]{\setminus }\Sigma (p,q)\) is a union of countably many \(\gamma _n:=\gamma (a_n,b_n)\) where \(a_n:=a_{\Omega _n},b_n:=b_{\Omega _n}\) are the entry, exit points (given by (3.2c)) associated with the countably many \(\Omega _n\) that satisfy \({{\,\mathrm{\textsf{card}}\,}}(\gamma \cap \Omega ^\mathrm{{ra}}_n)\ge 2\) with \(\Omega _p\ne \Omega _n\ne \Omega _q\). Note that \(a_n,b_n\in {\partial ^\textrm{ra}}\Omega _n\cap X\) and these entry, exit points correspond to unique points \({\tilde{a}}_n,{\tilde{b}}_n\in \partial _l\Omega _n\subset \overline{(\Omega _n)_l}\).

For each n, let \(\sigma _n:a_n\curvearrowright b_n\) in \(\overline{X_l}\) be the \(h_{\Omega _n}\) image of the l-geodesic \({\tilde{a}}_n\curvearrowright {\tilde{b}}_n\) in \(\overline{(\Omega _n)_l}\). Replacing each of the subarcs \(\gamma [p,e_p],\gamma [e_q,q],\gamma _n\) of \(\gamma \) with \(\sigma _p,\sigma _q,\sigma _n\), respectively, we obtain an arc

$$\begin{aligned} \sigma :=\sigma _p\cup \Sigma (p,q)\cup \sigma _q\cup \bigcup _n \sigma _n:p\curvearrowright q \quad \text {in}\, \overline{X_l}. \end{aligned}$$

Since the new subarcs have lengths no larger than the replaced subarcs, \(\ell (\sigma )\le \ell (\gamma )\). Since the entry, exit points relative to pq do not depend on \(\gamma \), the construction of \(\sigma \) is independent of \(\gamma \) and \(\sigma \) is indeed an arc \(p\curvearrowright q\) in \(\overline{X_l}\) with shortest length.

Assume \(p\in X\) and \(q\in \partial _l X\). Suppose \(q\in \overline{\Omega _l}\) for some \(\Omega \in {{\mathscr {O}}}\). Assume \(p\not \in \Omega ^\mathrm{{ra}}\). By (3.4b), there is a unique exit point \(e\in {\partial ^\textrm{ra}}\Omega \cap X\) (that depends only on p), and by earlier work, there are unique l-geodesics \(\sigma _q:e\curvearrowright q\) in \(\overline{\Omega _l}\subset \overline{X_l}\) and \(\sigma _p:p\curvearrowright e\) in \(\overline{X_l}\), and we see that \(\sigma :=\sigma _p\star \sigma _q\) is a shortest arc \(p\curvearrowright q\) in \(\overline{X_l}\).

Suppose that for all \(\Omega \in {{\mathscr {O}}}\), \(q\not \in \overline{\Omega _l}\). Start with any \(\gamma \in \Gamma _l(p,q)\) and let \((z_n)\) be an increasing sequence along \(\gamma \) with \(\ell (\gamma [z_n,q])\rightarrow 0\). If \(z_n\in \Sigma (p,q)\), set \(x_n:=z_n\). If \(z_n\not \in \Sigma (p,q)\), then (3.5b) and (3.4c) provide entry, exit points \(a_n,b_n\in \gamma \cap {\partial ^\textrm{ra}}\Omega _n\) with \(z_n\in \gamma (a_n,b_n)\); here we set \(x_n:=b_n\). Thus \((x_n)\) is an increasing sequence in \(\Sigma (p,q)\) with \(l(x_n,q)\rightarrow 0\) as \(n\rightarrow +\infty \).

As \(p,x_n\in X\), there are l-geodesics \(\sigma _n:p\curvearrowright x_n\) in \(\overline{X_l}\). However, \(x_n\in \Sigma (p,x_{n+1})\), so \(\sigma _n\subset \sigma _{n+1}\).Footnote 8 Therefore, it follows that \(\sigma :=\bigcup _{n\ge 1}\sigma _n\) is a rectifiable arc in \(\overline{X_l}\) with terminal endpoint q and with \(\ell (\sigma )=l(p,q)\). Thus \(\sigma \) is a shortest arc, hence an l-geodesic in \(\overline{X_l}\).

Assume \(p,q\in \partial _l X\). If \(\Sigma (p,q)=\left\{ p,q\right\} \), then by (3.5b) \(p,q\in \overline{\Omega _l}\subset \overline{X_l}\) for some \(\Omega \in {{\mathscr {O}}}\) and hence there is an l-geodesic joining these points. Suppose there exists an \(x\in X\cap \Sigma (p,q)\). Then by a previous case there are l-geodesics \(p\curvearrowright x\) and \(x\curvearrowright q\) which paste together to give a path \(\sigma :p\curvearrowright q\) in \(\overline{X_l}\) with \(\ell (\sigma )=l(p,x)+l(x,q)=l(p,q)\).

3.2.4 Uniqueness and CAT(0)

Our penultimate task is to verify uniqueness of l-geodesics in \(\overline{X_l}\). Let pq be distinct points in \(\overline{X_l}\), let \(\sigma :p\curvearrowright q\) be the l-geodesic in \(\overline{X_l}\) constructed above, and suppose \(\psi :p\curvearrowright q\) is also an l-geodesic in \(\overline{X_l}\). Then

$$\begin{aligned} \sigma \cap \Sigma (p,q)=\Sigma (p,q)=\psi \cap \Sigma (p,q). \end{aligned}$$

Let \(z\in \sigma {\setminus }\Sigma (p,q)\).

Appealing to (3.5b) we obtain an \(\Omega \in {{\mathscr {O}}}\) and an arc \(\alpha \subset \sigma \cap \Omega ^\mathrm{{ra}}\) with . This means that \({{\,\mathrm{\textsf{card}}\,}}(\sigma \cap \overline{\Omega _l})\ge 2\), so by (3.4c) \(z\in \sigma [a,b]\subset \overline{\Omega _l}\) where \(a,b\in {\partial ^\textrm{ra}}\Omega \subset \overline{X_l}\) are the entry, exit points associated with \(p,q,\Omega \). Also by (3.4c), \(a,b\in \psi \). Since \(\psi [a,b]\subset \overline{\Omega _l}\) is a shortest arc and \(\overline{\Omega _l}\) is CAT(0) (by Sect. 3.1), it must be that \(\psi [a,b]=\sigma [a,b]\).

By symmetry it now follows that \(\sigma =\psi \).

Finally, we confirm the CAT(0) property for \(\overline{X_l}\). Let \(\Delta :=[a,b,c]_l=[a,b]_l\cup [b,c]_l\cup [c,a]_l\) be a geodesic triangle in \(\overline{X_l}\). By Fact 2.7 we may assume that \(\Delta \) is tail-less. Since l-geodesics in \(\overline{X_l}\) are unique, this means that \(\Delta \) is a rectifiable Jordan loop in \(\overline{X_l}\).

Since \([a,b]_l\cap \bigl ([b,c]_l\star [c,a]_l\bigr )=\left\{ a,b\right\} \), \(\Sigma (a,b)=\left\{ a,b\right\} \). Now (3.5b) produces an \(\Omega \in {{\mathscr {O}}}\) with \(a,b\in \overline{\Omega _l}\subset \overline{X_l}\) and so \([a,b]_l\subset \overline{\Omega _l}\). Similarly, \(\Sigma (b,c)=\left\{ b,c\right\} \) and \(\Sigma (c,a)=\left\{ c,a\right\} \), so \([b,c]_l\cup [c,a]_l\subset \overline{\Omega _l}\). Thus \(\Delta \subset \overline{\Omega _l}\) and therefore \(\Delta \) satisfies the CAT(0) vertex angle criterion.

3.3 Gromov Hyperbolicity Proof

Clearly, if X contains Euclidean disks of arbitrarily large radius, then \(\overline{X_l}\) is not Gromov hyperbolic. Suppose

$$\begin{aligned} R:=\sup \left\{ r>0 \;\big \vert \; \exists \, \textsf{D}(x;r)\subset X\right\} <+\infty . \end{aligned}$$

We show that X is 2R-hyperbolic, and that this is best possible.

Let \(\Delta =[a,b,c]_l\) be a geodesic triangle in \(\overline{X_l}\). We may assume that \(\Delta \) is tail-less; therefore, as explained immediately above, \(\Delta \subset \overline{\Omega _l}\) for some \(\Omega \in {{\mathscr {O}}}\).

There is a simply connected \(\Omega _o\subset \Omega \) with \({\partial \Omega }_o=\iota (\Delta )\); see the last paragraph of Sect. 2.3.1. Let \(D_o:=\textsf{D}(o;r)\) be a maximal open disk in \(\Omega _o\). Then \({\partial D}_o\cap {\partial \Omega }_o\) either consists of two antipodal points or has cardinality at least three. Below we verify that \({{\,\mathrm{\textsf{card}}\,}}({\partial D}_o\cap {\partial \Omega }_o)=3\).

Notice that \({{\bar{D}}}_o=D_o\cup {\partial D}_o\) isometrically embeds into \(\overline{\Omega _l}\hookrightarrow \overline{X_l}\) so \(S_l:=\partial _l D_o\) is a Euclidean circle in \(\overline{X_l}\) that bounds a Euclidean disk \(D_l\) in \(\overline{X_l}\). Note too that \(S_l\cap \Delta \subset \overline{\Omega _l}\) is isometrically equivalent to \(\partial D_o\cap {\partial \Omega }_o\); one way to see this is to use the conformal model for \(\overline{\Omega _l}\) (see the last paragraph of Sect. 2.3.1) in conjunction with [10, Prop. 2.14, Cor.2.17, pp. 29,35].

So, \({{\,\mathrm{\textsf{card}}\,}}(S_l\cap \Delta )\ge 2\); we show that \(S_l\cap \Delta =\left\{ a_o,b_o,c_o\right\} \) with \(a_o\in (b,c)_l,b_o\in (c,a)_l,c_o\in (a,b)_l\).

First, let \(E\in \left\{ [a,b]_l,[b,c]_l,[c,a]_l\right\} \) be an edge of \(\Delta \). If \(E\cap S_l\) contained two distinct points xy, then the Euclidean segment \([x,y]\ne E[x,y]\) would be an l-geodesic \(x\curvearrowright y\) in \(\overline{\Omega _l}\) which would violate unique geodesicity for \(\overline{\Omega _l}\); thus

$$\begin{aligned} {{\,\mathrm{\textsf{card}}\,}}(E\cap S_l)\le 1. \end{aligned}$$
(3.6a)

Also,

$$\begin{aligned} \exists \; p\in E\cap S_l \quad \implies \quad \forall \; q\in E{\setminus }\left\{ p\right\} , \;\; \measuredangle _p(q,o)\ge \frac{\pi }{2}. \end{aligned}$$
(3.6b)

Indeed, E is a complete convex subspace of \(\overline{\Omega _l}\), so any \(p\in E\cap S_l\) is the unique point of E nearest to o, and so (3.6b) follows from [4, Prop. II.2.4(3), p. 177]. Here \(\measuredangle _p(q,o)=\measuredangle _p(E[p,q],[p,o])\).

Thus \(2\le {{\,\mathrm{\textsf{card}}\,}}(S_l\cap \Delta )\le 3\). Suppose \(S_l\cap \Delta =\left\{ p,q\right\} \). Then \(\iota (p),\iota (q)\) are antipodal points of \({\partial D}_o\), so the Euclidean segment [pq] is the l-geodesic \(p\curvearrowright q\) in \(\overline{\Omega _l}\). Now pq are not both vertices of \(\Delta \) (otherwise (3.6a) would be violated) so we can select a vertex, say a, of \(\Delta \) so that

$$\begin{aligned} p\in [a,b]_l, \; q\in [a,c]_l, \; p\ne a\ne q \ne c \text { but maybe } p=b. \end{aligned}$$

According to (3.6b), \(\measuredangle _p(a,o)\ge \pi /2\) and \(\measuredangle _q(a,o)\ge \pi /2\). However, the comparison triangle \({\bar{\Delta }}(a,p,q)\) (in \(\mathbb {C}\)) cannot have two vertex angles that are both of size \(\pi /2\) or larger.

Thus \({{\,\mathrm{\textsf{card}}\,}}(S_l\cap \Delta )=3\). Employing (3.6a) again we see that \(S_l\cap \Delta \cap \left\{ a,b,c\right\} =\varnothing \). It now follows that there are points \(a_o\in (b,c)_l,b_o\in (c,a)_l,c_o\in (a,b)_l\) with \(S_l\cap \Delta =\left\{ a_o,b_o,c_o\right\} \).

Take any point \(z\in \Delta \), say \(z\in [a,b]_l\), or even \(z\in [a,c_o]_l\). Look at a comparison triangle \(\Delta (\bar{a},\bar{c}_o,\bar{b}_0)\) for \([a,c_o,b_o]_l\) and pick a point \(w\in [a,b_o]_l\) so that the Euclidean segment \([\bar{z},\bar{w}]\) is parallel to \([\bar{c}_o,\bar{b}_o]\). Now we see that

$$\begin{aligned} l(z,w) \le |\bar{z}-\bar{w}| \le |\bar{c}_o-\bar{b}_o| = l(c_o,b_o) = |\iota (c_o)-\iota (b_o)| \le 2r\le 2R \end{aligned}$$

and therefore \(\overline{X_l}\) is 2R-hyperbolic.

To see that this is best possible, fix \(R>0\) and consider the set

$$\begin{aligned} X:=\left\{ z\in {\mathbb C}\;\big \vert \; |{{\,\mathrm{{\text {Im}}}\,}}(z)|\le R\right\} ; \end{aligned}$$

the points \(a:=t, b:=Ri, c:=-Ri\) where \(t>2R\); and \(\Delta :=[a,b,c]=[a,b]\cup [b,c]\cup [c,a]\). Let \(2\varphi \) be the vertex angle for \(\Delta \) at a; i.e., the angle between the edges [ab] and [ac]. Pick \(y\in (0,R)\) so that \(z:=2R+iy\in [a,b]\). Then

$$\begin{aligned} \frac{t}{H} = \cos \varphi = \frac{2R}{h} \quad \text {and}\quad \frac{R}{H} = \sin \varphi = \frac{R-y}{h} \end{aligned}$$

where

$$\begin{aligned} H:=|a-b|=\sqrt{t^2+R^2} \quad \text {and}\quad h:=|z-b|=\frac{2RH}{t}. \end{aligned}$$

Now \(\dfrac{{{\,\mathrm{\textsf{dist}}\,}}(z,[a,c])}{H-h}=\sin 2\varphi =2\dfrac{tR}{H^2}\), so

$$\begin{aligned} {{\,\mathrm{\textsf{dist}}\,}}(z,[a,c])=2R\frac{t}{H}\frac{H-h}{H}= 2R\frac{t}{\sqrt{t^2+R^2}}\Bigl (1-\frac{2R}{t}\Bigr ) \rightarrow 2R \text { as } t\rightarrow +\infty . \end{aligned}$$

Since \({{\,\mathrm{\textsf{dist}}\,}}(z,[b,c])=2R\), we see that X is \(\delta \)-hyperbolic for \(\delta :=2R\) but no smaller \(\delta \) works. \(\square \)